1 Introduction

The phenomenon of conventional superconductivity (see, e.g., [55] for a review of the physics of superconductors) is nowadays very well understood at the microscopic level thanks to the Bardeen–Cooper–Schrieffer (BCS) theory [8]: a collective behavior of the current carriers in the material is responsible for a sudden drop of the resistivity below a certain critical temperature. It is however astonishing how a phenomenological model as the Ginzburg–Landau (GL) theory [42] is capable of predicting most of the key equilibrium features of the phenomenon, in particular concerning the response of the superconducting material to an external field. When it was introduced in the ‘50s, indeed, the GL model was motivated only from purely phenomenological considerations. Only later it was shown that the GL theory emerges as an effective macroscopic model from the BCS theory suitably close to the critical temperature [37, 40, 43].

The interplay between superconductivity and strong magnetic fields is known to generate a very rich variety of physical phenomena since the pioneering works of Abrikosov [2] and St. James and De Gennes [54] in the late ‘50s/early ‘60s, who predicted the occurrence of the famous vortex lattice and of surface superconductivity, respectively, working only in the framework of the GL theory. In extreme synthesis, the response of a type-II superconducting material to the external magnetic field can vary from a perfect repulsion of the field (Meissner effect), for small fields, to a complete loss of superconductivity, for sufficiently strong ones. In between, several different phases of the material can be observed, ranging from various kinds of vortex states to configurations where the superconduction gets restricted to boundary regions. Each of these phase transitions can be associated with a critical magnetic field marking the threshold for the transition: the three major critical fields are

  • the first critical field, which separate the Meissner behavior, i.e., when the magnetic field inside the material is zero and superconductivity is unaffected, from states where the penetration of the field has occurred at least at isolated points (vortices), where superconductivity is lost;

  • the second critical field, above which the superconducting behavior gets confined at the surface of the sample (surface superconductivity);

  • the third critical field, which marks the complete loss of superconductivity.

Let us now introduce in more detail the GL theory: the free energy of the material is given by a nonlinear functional, which in the case of a superconducting infinite wire of cross section \( \Omega \subset {\mathbb {R}}^2 \) reads in suitable units

$$\begin{aligned}&{\mathcal {E}}_{\varepsilon }^{\mathrm {GL}}[\psi , \mathbf{A }]= \displaystyle \int _{\Omega } \mathrm {d}\mathbf{r }\; \bigg \{ \bigg | \left( \nabla + i \frac{\mathbf{A }}{\varepsilon ^ 2}\right) \psi \bigg |^2 \nonumber \\&\quad -\frac{1}{2b\varepsilon ^2}(2|\psi |^2-|\psi |^4) \bigg \}+\frac{1}{\varepsilon ^4}\displaystyle \int _{{\mathbb {R}}^2} \mathrm {d}\mathbf{r }\; |\text{ curl }\mathbf{A } - 1|^2, \end{aligned}$$
(1.1)

where \( \varepsilon , b > 0 \) are two parameters depending on the London penetration depth and the intensity of the applied magnetic field, which is assumed to be parallel to the wire. The function \( \psi \), a.k.a. wave function or order parameter, is complex, while \( \mathbf{A }\) is the induced magnetic potential, whose \( \mathrm {curl}\) yields the intensity of the magnetic field outside and inside the sample (measured in units \( \varepsilon ^{-2} \)). The physical meaning of the order parameter is twofold: \( |\psi |^2 \) yields the relative density of Cooper pairs and, at the same time, the phase of \( \psi \) contains the information about the stationary current flowing in the superconductor, i.e.,

$$\begin{aligned} \mathbf{j }[\psi ] : = \textstyle \frac{i}{2} \left( \psi \nabla \psi ^* - \psi ^* \nabla \psi \right) = \mathrm {Im}\left( \psi ^* \nabla \psi \right). \end{aligned}$$
(1.2)

Hence, one typically speaks of a normal state, if \( \psi = 0 \) and \( \mathbf{A }\) is such that \( \mathrm {curl}\mathbf{A }= 1 \), while the perfect superconducting state is identified by \( |\psi | = 1 \), \( \mathbf{A }= 0 \). Whenever \( |\psi | \) is non-vanishing everywhere but not identically 1, the superconductor is said to be in a mixed state. Any equilibrium state of the sample minimizes the free energy (1.1) and thus we set

$$\begin{aligned} E_{\varepsilon }^{\mathrm {GL}}:= \displaystyle \min _{(\psi ,\mathbf{A })\in {\mathscr {D}}^{\mathrm {GL}}} {\mathcal {E}}_{\varepsilon }^{\mathrm {GL}}[\psi ,\mathbf{A }], \end{aligned}$$
(1.3)

and denote by \( (\psi ^{\mathrm {GL}}, \mathbf{A }^{\mathrm {GL}}) \) any minimizing configuration, where

$$\begin{aligned} {\mathscr {D}}^{\mathrm {GL}}= \left\{ \left( \psi , \mathbf{A }\right) \in H^1(\Omega ) \times H^1_{\mathrm {loc}}({\mathbb {R}}^2; {\mathbb {R}}^2) \, \big | \, \mathrm {curl}\mathbf{A }- 1 \in L^2({\mathbb {R}}^2) \right\} . \end{aligned}$$
(1.4)

We provide some details about the above minimization and the properties of any minimizing configuration \( (\psi ^{\mathrm {GL}}, \mathbf{A }^{\mathrm {GL}}) \) in “Appendix B.1”. We also use the following convention: if we need to specify the dependence on the domain \( \Omega \), we write \( {\mathcal {E}}_{\varepsilon }^{\mathrm {GL}}[\psi , \mathbf{A }; \Omega ] \) for the functional and \( E_{\varepsilon }^{\mathrm {GL}}(\Omega ) \) for the corresponding ground state energy.

In the rest of the paper we are going to study the minimization (1.3) in the asymptotic regime

$$\begin{aligned} \varepsilon \ll 1, \end{aligned}$$
(1.5)

corresponding to an extreme type-II superconductor. Under this idealization, one can identify the mathematical counterparts of the critical values of the external magnetic field described above in terms of properties of the minimizing configuration \( (\psi ^{\mathrm {GL}}, \mathbf{A }^{\mathrm {GL}}) \) and it is also possible to precisely identify the behavior of such thresholds (see, e.g., [53] for an extensive discussion of the first phase transition). In particular, assuming that \( \Omega \) is a simply connected domain with smooth boundary \( \partial \Omega \), the second critical field associated with the transition from bulk to surface superconductivity is identified with \( b = 1 \) [33, Chpt. 10.6] and thus with a field of intensity

$$\begin{aligned} H_{\mathrm {c}2}= \frac{1}{\varepsilon ^2}, \end{aligned}$$
(1.6)

based on sharp estimates (Agmon estimates) of the decay of \( \psi ^{\mathrm {GL}}\) in the distance from the boundary (see “Appendix B.3”); the third critical field marking the transition to the normal state on the other hand corresponds to \( b = \Theta _0^{-1} > 1 \), where \( \Theta _0\simeq 0.59 \) is a universal constant, i.e., more precisely [33, Chpt. 13],

$$\begin{aligned} H_{\mathrm {c}3}= \frac{1}{\Theta _0\varepsilon ^2} + {\mathcal {O}}(1). \end{aligned}$$
(1.7)

1.1 Setting: domains with corners

In this paper we are exclusively concerned with the behavior of the superconductor for very strong magnetic fields above the second critical one, i.e., we always assume that \( h_{\mathrm {ex}}> H_{\mathrm {c}2}\), or, more concretely,

$$\begin{aligned} b > 1. \end{aligned}$$
(1.8)

The main novelty of this paper compared to other works on the GL functional above the second critical field is that we assume that \( \Omega \) is a bounded domain with a Lipschitz boundary, i.e., we allow for the presence of corners on \( \partial \Omega \) (see Fig. 1). Indeed, apart from few physics papers (see [14, 31, 52]), the GL theory on domains with corners has already been studied only in [15, 45, 46, 50], with the focus on the third critical field though, and in [19], whose results are improved in this work.

Fig. 1
figure 1

A domain \( \Omega \) with Lipschitz boundary and finitely many corners on \( \partial \Omega \)

The main reason why it is interesting to study the behavior of the GL functional in domains with corners for large magnetic fields is that for smaller fields one expects that the presence of corners does not affect the salient features of superconductivity. Indeed, the occurrence of vortices but also their uniform distribution and arrangement in regular lattices, which occur for magnetic fields below \( H_{\mathrm {c}2}\), are bulk phenomena and, as such, little influenced by the boundary regularity. On the opposite, the surface superconductivity regime, where the density of Cooper pairs is non-vanishing only at and close to the boundary, might clearly depend on the presence of singularities along \( \partial \Omega \). It is then important to know if and to what extent corners can modify the boundary behavior, even more so, considered that in physics experiments it is hardly possible to distinguish between a sample with smooth boundary and another which has corners there (see, e.g., [48, Fig. 1]).

We now specify in more detail our assumptions on the domain \( \Omega \). First, we require that it is simply connected and its boundary \( \partial \Omega \) is Lipschitz (see, e.g., [44, Def. 1.4.5.1]) and, more concretely, it is a curvilinear polygon of class \( C^{\infty } \), given by smooth pieces glued together at finitely many points, where however the curvature remains finite (no cusps). These assumptions are the same made, e.g., in [15, 19, 45].

Assumption 1

(Piecewise smooth boundary) Let \(\Omega \) be a bounded open subset of \({\mathbb {R}}^2\). We assume that \( \partial \Omega \) is a smooth curvilinear polygon, i.e., for every \( \mathbf{r }\in \partial \Omega \) there exists a neighborhood U of \( \mathbf{r }\) and a map \( {\varvec{\Phi }}: U \rightarrow {\mathbb {R}}^2 \), such that

  1. (1)

    \( {\varvec{\Phi }}\) is injective;

  2. (2)

    \( {\varvec{\Phi }}\) together with \( {\varvec{\Phi }}^{-1} \) (defined from \( {\varvec{\Phi }}(U) \)) are smooth;

  3. (3)

    the region \(\Omega \cap U \) coincides with either \( \left\{ \mathbf{r }\in \Omega \, | \, \left( {\varvec{\Phi }}(\mathbf{r }) \right)_1 < 0 \right\} \) or \( \left\{ \mathbf{r }\in \Omega \, | \, \left( {\varvec{\Phi }}(\mathbf{r }) \right)_2 < 0 \right\} \) or \( \left\{ \mathbf{r }\in \Omega \, | \, \left( {\varvec{\Phi }}(\mathbf{r }) \right)_1< 0, \left( {\varvec{\Phi }}(\mathbf{r }) \right)_2 < 0 \right\} \), where \( \left( {\varvec{\Phi }}\right)_j\) stands for the \( j-\)th component of \( {\varvec{\Phi }}\).

Assumption 2

(Boundary with corners) We assume that the set \( \Sigma : = \left\rbrace \mathbf{r }_1, \ldots , \mathbf{r }_N \right\lbrace \) of corners of \( \partial \Omega \), i.e., the points where the normal \( {\varvec{\nu }}\) does not exist, is non empty but finite and given by N points. We denote by \( \beta _j\) the angle of the \(j-\)th corner (measured towards the interior) and by \( {\mathsf {s}}_j \) its boundary coordinate.

The inward normal \( {\varvec{\nu }}\) to \( \partial \Omega \) is thus defined almost everywhere and jumps at the corners. More precisely, we can find a counterclockwise parametrization \( {\varvec{\gamma }}({\mathsf {s}}): [0, |\partial \Omega |) \rightarrow \partial \Omega \) of \( \partial \Omega \) which is smooth a.e., i.e., for \( s \ne s_j \), \( j = 1, \ldots , N \), \( s_j \) being the curvilinear coordinate of the j-th corner, and such that \( |{\varvec{\gamma }}^{\prime }({\mathsf {s}})| = 1 \). The mean curvature \( {\mathfrak {K}}({\mathsf {s}}) \) is then defined a.e. through the identity

$$\begin{aligned} {\varvec{\gamma }}^{\prime \prime }({\mathsf {s}}) = {\mathfrak {K}}({\mathsf {s}}) {\varvec{\nu }}({\mathsf {s}}). \end{aligned}$$
(1.9)

We can then introduce a convenient system of tubular coordinates in a neighborhood of the boundary (see also [33, Appendix F]) far from the corners: for any \( \left({\mathsf {s}}, {\mathsf {t}}\right) \in {\mathcal {A}}_{{\mathsf {t}}_0} \), where

with \( {\mathsf {t}}_0 \ll 1 \) small enough and c depending only on the corner opening angles, we can set

$$\begin{aligned} \mathbf{r }({\mathsf {s}}, {\mathsf {t}}) = : {\varvec{\gamma }}^{\prime }({\mathsf {s}}) + {\mathsf {t}}{\varvec{\nu }}({\mathsf {s}}), \end{aligned}$$
(1.10)

which identifies a diffeomorphism from \( {\mathcal {A}}_{t_0} \) to the region

$$\begin{aligned} \Omega _{{\mathsf {t}}_0} : = \left\{ \mathbf{r }\in \Omega \, | \, \mathrm {dist}(\mathbf{r }, \partial \Omega ) < {\mathsf {t}}_0, \mathrm {dist}(\mathbf{r }, \Sigma ) > c {\mathsf {t}}_0 \right\} , \end{aligned}$$

satisfying

$$\begin{aligned} \mathrm {dist}\left(\mathbf{r }({\mathsf {s}},{\mathsf {t}}), \partial \Omega \right) = {\mathsf {t}}. \end{aligned}$$
(1.11)

Therefore, such a map is invertible in the same region and identifies a system of local coordinates \( ({\mathsf {s}}, {\mathsf {t}}) \) there. The parameters \( {\mathsf {s}}, {\mathsf {t}}\) (with the latter defined through (1.11)) are well-posed also closer to the corners but there they do not identify a diffeomorphism and thus a set of coordinates.

1.2 Heuristics

Before entering the discussion of what is mathematically known on the phenomenon of surface superconductivity, we resume here its key features for smooth domains, neglecting errors and remainders: if \( b > 1 \), as \( \varepsilon \rightarrow 0 \),

  • the order parameter \( \psi ^{\mathrm {GL}}\) is non-vanishing only close to \( \partial \Omega \); more precisely it is exponentially small in \( \varepsilon \) at distances from the boundary much larger than \( \varepsilon \);

  • the induced magnetic field \( \mathrm {curl}\mathbf{A }^{\mathrm {GL}}\) is suitably close to the applied one, i.e., a uniform magnetic field of unit strength, and, consequently, one can find a local gauge close to \( \partial \Omega \) in which \( \mathbf{A }^{\mathrm {GL}}\) is purely tangential and \( |\mathbf{A }^{\mathrm {GL}}| = \mathrm {dist}\{\mathbf{r }, \partial \Omega \} \);

  • the modulus of \( \psi ^{\mathrm {GL}}\) is essentially independent of the tangential coordinate \( {\mathsf {s}}\) and therefore optimizes an effective one-dimensional problem where the only variable is the distance from the boundary;

  • the phase of \( \psi ^{\mathrm {GL}}\) is on the other hand constant in \( {\mathsf {t}}\) and linear in \( {\mathsf {s}}\), with rapid oscillations, or, more precisely, the current (1.2) is constant along \( {\mathsf {s}}\).

Summing up, we expect that \( \mathbf{A }^{\mathrm {GL}}\) can be locally replaced by \( - {\mathsf {t}}\mathbf{e }_{{\mathsf {s}}} \) close to \( \partial \Omega \) and

$$\begin{aligned} \left| \psi ^{\mathrm {GL}}(\mathbf{r }) \right| \simeq f\left( \textstyle \frac{{\mathsf {t}}}{\varepsilon } \right), \qquad \mathbf{j }[\psi ^{\mathrm {GL}}] \simeq \frac{|\Omega |}{\varepsilon ^2} - \frac{\alpha }{\varepsilon }, \end{aligned}$$
(1.12)

for some f positive and \( \alpha \in {\mathbb {R}}\), which leads to

$$\begin{aligned} \psi ^{\mathrm {GL}}(\mathbf{r }) \simeq f({\mathsf {t}}/\varepsilon ) e^{- \frac{i \alpha {\mathsf {s}}}{\varepsilon }} e^{ i \phi _{\varepsilon }(\mathbf{r })}, \end{aligned}$$
(1.13)

\( \phi _{\varepsilon } \) standing for the gauge transformation mentioned above. Note the scaling factors \( 1/\varepsilon \) we have extracted for later convenience, so that f and \( \alpha \) are quantities of order \( {\mathcal {O}}(1) \).

If we plug the ansatz (1.13) into the GL energy (1.1), we get

$$\begin{aligned} \frac{\left| \partial \Omega \right|}{\varepsilon } \int ^{\ell _{\varepsilon }}_0 \text{ dt } \left\{ |\partial _t f|^2 + (t+\alpha )^2 f^2 -\frac{1}{2b} (2f^2-f^4)\right\} , \end{aligned}$$
(1.14)

i.e., up to the prefactor \( |\partial \Omega |{/\varepsilon } \), a one-dimensional (1D) energy functional evaluated on f and depending on the real parameter \( \alpha \). The value \( \ell _{\varepsilon }> 0 \) is to some extent arbitrary and is chosen much larger than 1 in order to cover all the superconducting layer: we make the following explicit choice

$$\begin{aligned} \ell _{\varepsilon }: = c_1 |\log \varepsilon |, \end{aligned}$$
(1.15)

for a large constant \( c_1 \). The minimization of the 1D functional above and some variants of it w.r.t. both f and \( \alpha \) is discussed in “Appendix A”. This identifies the leading term contribution in the GL energy asymptotics \( E^{\mathrm {1D}}_{\star }/\varepsilon \), the optimal 1D profile \( f_{\star }(t) \) and the optimal phase \( \alpha _{\star }\).

The next-to-leading order term in the GL energy asymptotics is of order \( {\mathcal {O}}(1) \) and depends on the mean curvature of the boundary: one can indeed refine the 1D model problem (1.14) keeping track of \( {\mathcal {O}}(\varepsilon ) \) contributions coming from the curvature-dependent terms due to the change of coordinates \( \mathbf{r }\rightarrow ({\mathsf {s}}, {\mathsf {t}}) \). Indeed, if we define the rescaled tubular coordinates as

$$\begin{aligned} {\left\{ \begin{array}{ll} t : = {\mathsf {t}}/\varepsilon \in [0, \ell _{\varepsilon }], \\ s : = {\mathsf {s}}/\varepsilon \in \left[0, \frac{|\partial \Omega |}{\varepsilon } \right]. \end{array}\right. } \end{aligned}$$
(1.16)

we get \( \mathrm {d}\mathbf{r }= \mathrm {d}{\mathsf {t}}\mathrm {d}{\mathsf {s}}\left(1 - {\mathfrak {K}}({\mathsf {s}}){\mathsf {t}}\right) \), or, equivalently,

$$\begin{aligned} \mathrm {d}\mathbf{r }= \varepsilon ^2 \mathrm {d}t \mathrm {d}s \left(1 - \varepsilon k(s) t \right), \end{aligned}$$
(1.17)

where we have set

$$\begin{aligned} k(s) : = {\mathfrak {K}}(\varepsilon s). \end{aligned}$$
(1.18)

1.3 Summary

In this paper we continue the analysis started with [19]. The expansion (2.10) provides indeed only the leading order term in the energy asymptotics and does not capture the corner effects, that we are going to investigate. More precisely, we prove that the presence of corners modifies the \( {\mathcal {O}}(1) \) term in the expansion (2.2). We also identify the model problem which yields such a new contribution in terms of a genuine 2D model. Finally, we prove that the pointwise estimate of \( |\psi ^{\mathrm {GL}}| \) in terms of \( f_{\star }\) still holds far from the corners, precisely as in the smooth case.

After having introduced some notation in § 1.4, we define in § 2 the effective model in the corner region and state our main results about the GL energy asymptotics and the pointwise estimate of the order parameter far from corners. Further comments about the effective model and a conjecture about the possible explicit expression of the effective energy are contained in § 2.3. The well-posedness of the effective model problem is proven in detail in § 3, whereas in § 4 and § 5 we provide the energy lower bound and the rest of the arguments needed to complete the proof of our main results, respectively. The Appendix is divided in three parts: in “Appendix A” we discuss the effective 1D problems and their related properties; the GL minimization and some useful technical estimates are treated in “Appendix B”; finally, “Appendix C” recalls the salient steps of the proof of the energy asymptotics in domains with smooth boundaries, which are used to complete the proof of energy expansion.

1.4 Notation

Given their key role in the rest of the paper, we recall the definitions (1.10) and (1.16) of tubular coordinates \( ({\mathsf {s}}, {\mathsf {t}}) \) and their rescaled counterparts (st) . We stress that \( ({\mathsf {s}}, {\mathsf {t}}) \) or, equivalently, (st) provide a smooth diffeomorphism, e.g., in

$$\begin{aligned} \left\{ \mathbf{r }\in {\mathcal {A}}_{\varepsilon }\, \big | \, \mathrm {dist}(\mathbf{r }, \Sigma ) \geqslant \varepsilon |\log \varepsilon | \right\} , \end{aligned}$$

where \( \Sigma \) is the set of corner positions and

$$\begin{aligned} {\mathcal {A}}_{\varepsilon }: = \left\{ \mathbf{r }\in \Omega \, \big | \, \mathrm {dist}\left(\mathbf{r }, \partial \Omega \right) \leqslant \varepsilon \ell _{\varepsilon }\right\} , \end{aligned}$$
(1.19)

for \( \varepsilon \ll 1 \), where (see (1.15))

$$\begin{aligned} \ell _{\varepsilon }: = c_1 |\log \varepsilon | \end{aligned}$$

and \( c_1 \) is large enough constant, which is set once for all (see next (2.24)). Given a differentiable function \( \psi (\mathbf{r }) \) and a vector \( \mathbf{A }(\mathbf{r }) \), the transformations induced by the change of coordinates \( \mathbf{r }\rightarrow ({\mathsf {s}}, {\mathsf {t}}) \) are

$$\begin{aligned} \left( \nabla \psi \right) (\mathbf{r }({\mathsf {s}},{\mathsf {t}})) = \left(1 - {\mathfrak {K}}({\mathsf {s}}){\mathsf {t}}\right)^{-1} \big ( \partial _{\mathsf {s}}{\widetilde{\psi }} \big ) \mathbf{e }_{{\mathsf {s}}} + \big ( \partial _{{\mathsf {t}}} {\widetilde{\psi }} \big ) \mathbf{e }_{{\mathsf {t}}}, \end{aligned}$$
(1.20)

where we have set \( {{\widetilde{\psi }}}({\mathsf {s}},{\mathsf {t}}) : = \psi (\mathbf{r }({\mathsf {s}},{\mathsf {t}})) \) and

$$\begin{aligned} \mathbf{e }_{{\mathsf {s}}} : = {\varvec{\gamma }}^{\prime }({\mathsf {s}}), \qquad \mathbf{e }_{{\mathsf {t}}} : = {\varvec{\nu }}({\mathsf {s}}), \end{aligned}$$
(1.21)

for short. As a consequence, for any vector \( \mathbf{A }\),

$$\begin{aligned} \left( \mathrm {curl}\, \mathbf{A }\right)(\mathbf{r }({\mathsf {s}},{\mathsf {t}}))&= - \partial _{{\mathsf {t}}} \left( \mathbf{A }(\mathbf{r }({\mathsf {s}},{\mathsf {t}})) \cdot \mathbf{e }_{{\mathsf {s}}} \right) \nonumber \\&\quad + \left(1 - {\mathfrak {K}}({\mathsf {s}}){\mathsf {t}}\right)^{-1} \left[ \partial _{{\mathsf {s}}} \left( \mathbf{A }(\mathbf{r }({\mathsf {s}},{\mathsf {t}})) \cdot \mathbf{e }_{{\mathsf {t}}} \right) + {\mathfrak {K}}({\mathsf {s}})\mathbf{A }(\mathbf{r }({\mathsf {s}},{\mathsf {t}})) \cdot \mathbf{e }_{{\mathsf {s}}} \right]. \end{aligned}$$
(1.22)

We are going to make use of Landau symbols, with the following convention: given two functions f(x), g(x) , with \( g > 0 \),

  • \( f = {\mathcal {O}}(g) \), if \( \displaystyle \lim _{x \rightarrow 0^+} |f|/g \leqslant C \);

  • \( f = o(g) \), if \( \displaystyle \lim _{x \rightarrow 0^+} |f|/g = 0 \);

  • \( f \sim g \), if \( f = {\mathcal {O}}(g) \) and \( \displaystyle \lim _{x \rightarrow 0^+} |f|/g > 0 \);

  • for \( f > 0 \), \( f \ll g \) or \( f \gg g \), if \( f = o(g) \) or \( g = o(f) \), respectively;

  • for \( f > 0 \), \( f \lesssim g \) or \( f > rsim g \), if \( f = {\mathcal {O}}(g) \) or \( g = {\mathcal {O}}(f) \), respectively.

We also commit a little abuse of the notation by using the symbols \( {\mathcal {O}}( \, ) \) and \( o( \, ) \) inside an inequality to mean a precise direction of the estimate. As usual, \( {\mathcal {O}}( \, ) \) and \( o( \, ) \) stand for quantities whose sign is not known. In case of functions of two or more variables, we point out the parameter, whose asymptotics we are considering, by adding a label, e.g., \( o_x(\,) \) or \( {\mathcal {O}}_x( \, ) \) is meant to stress that we are estimating the behavior of the function as \( x \rightarrow 0^+ \). Finally, we say that a quantity is \( {\mathcal {O}}(\varepsilon ^{\infty }) \), as \( \varepsilon \rightarrow 0^+ \), if it is smaller than any power of \( \varepsilon \), i.e., it is exponentially small in \( \varepsilon \). We will also use the following convention: \( {\mathcal {O}}(\varepsilon ^a |\log \varepsilon | ^{\infty }) \), \( a > 0 \), stands for a quantity which is bounded by \( \varepsilon ^a |\log \varepsilon |^b \) for some large but finite power \( b > 0 \), which is however not relevant since the \( |\log \varepsilon |\)-factor is always dominated by \( \varepsilon \)-powers.

2 Main results

2.1 State of the art

We briefly review here the most recent and relevant results on surface superconductivity, which are related to the analysis carried on in this paper (see [16] for a more detailed review). After the series of works [17, 23,24,25], following [49], where the problem was first investigated, and [4, 34], the phenomenon of 2D surface superconductivity in domains with smooth boundaries is well understood: combining [23, Thm. 1] with [24, Lemma 2.1] (see also [17]), one gets that, whenever

$$\begin{aligned} 1< b < \Theta _0^{-1}, \end{aligned}$$
(2.1)

the GL energy asymptotics is given by

$$\begin{aligned} E_{\varepsilon }^{\mathrm {GL}}= \frac{|\partial \Omega | E^{\mathrm {1D}}_{\star }}{\varepsilon } - 2 \pi E_{\mathrm {corr}}+ o(1), \end{aligned}$$
(2.2)

where

$$\begin{aligned} E^{\mathrm {1D}}_{\star }: = \inf _{\alpha \in {\mathbb {R}}} \inf _{f \in {\mathscr {D}}^{\mathrm {1D}}} \int ^{+\infty }_0 \text{ dt } \left\{ |\partial _t f|^2 + (t+\alpha )^2 f^2 -\frac{1}{2b} (2f^2-f^4)\right\} , \end{aligned}$$
(2.3)

and

$$\begin{aligned} E_{\mathrm {corr}}: = \int _{0} ^{+\infty } \mathrm {d}t \, t\left\{ \left| \partial _t f_{\star }\right|^2 + f_0 ^2 \left( -\alpha _{\star }(t+\alpha _{\star }) -\frac{1}{b} + \frac{1}{2b} f_{\star }^2\right) \right\} = \frac{1}{3} f_{\star }^2(0) \alpha _{\star }- E^{\mathrm {1D}}_{\star },\nonumber \\ \end{aligned}$$
(2.4)

\( \alpha _{\star }, f_{\star }\) being a pair of minimizers of (2.3) (see “Appendix A.1”). Note that (2.2) can also be rewritten as

$$\begin{aligned} {E_{\varepsilon }^{\mathrm {GL}}= \int _0^{\frac{|\partial \Omega |}{\varepsilon }} \mathrm {d}s \, E^{\mathrm {1D}}_{k(s)} + {\mathcal {O}}(\varepsilon |\log \varepsilon |^{\infty }),} \end{aligned}$$

with a more precise remainder term and where \( E^{\mathrm {1D}}_k \) is defined in (A.6) in “Appendix A.2”. Expanding further \( E ^{\mathrm{1D}}_{k}\), the next-to-leading order correction in (2.2) can be shown to be

$$\begin{aligned} -E_{\mathrm {corr}}\int _{\partial \Omega } \mathrm {d}{\mathsf {s}}\, {\mathfrak {K}}({\mathsf {s}}) + o(1) = -2\pi E_{\mathrm {corr}}+ o(1), \end{aligned}$$
(2.5)

by the Gauss–Bonnet theorem, because \( \Omega \) is flat and the Euler characteristic is equal to 1. Moreover, in [17] the quantity \( E_{\mathrm {corr}}\) is numerically evaluated and it is shown that it is positive, which has some important consequences on the distribution of superconductivity near the boundary: regions with larger curvature attract Cooper pairs, which concentrate more there (to first order), although to leading order superconductivity is uniform at the boundary.

Indeed, a consequence of (2.3) is that [23, Thm. 1] the density \( |\psi ^{\mathrm {GL}}|^2 \) is \( L^2\)-close to the reference density \( f_{\star }\). Such an estimate can in fact be strengthened in two directions:

  • in [23, Thm. 2] it is proven that there exists a boundary layer \( {\mathcal {A}}_{\mathrm {bl}} \subset \left\rbrace \mathbf{r }| \mathrm {dist}(\mathbf{r }, \partial \Omega ) \leqslant \varepsilon |\log \varepsilon | \right\lbrace \), containing the bulk of superconductivity, where Pan’s conjecture holds true, i.e.,

    $$\begin{aligned} \left\Vert \left| \psi ^{\mathrm {GL}}(\, \cdot \,) \right| - f_{\star }(\mathrm {dist}(\, \cdot \,,\partial \Omega )/\varepsilon ) \right\Vert_{L^{\infty }({\mathcal {A}}_{\mathrm {bl}})} = o(1); \end{aligned}$$
    (2.6)
  • the approximation of \( |\psi ^{\mathrm {GL}}| \) in terms of \( f_{\star }\) holds also locally [24, Thm 1.1] and one can explicitly derive the asymptotics of the density of superconductivity (in fact, the \( L^4 \) norm of \( \psi ^{\mathrm {GL}}\)) in any reasonable subdomain contained in \( \Omega \).

It is expected that a regime of surface superconductivity with similar features occurs also for genuine 3D samples but so far only partial results are available [38, 39]. In particular, in [38, Thm 1.1] (see also [36]) it is shown that such a regime does exist and the leading order term in the energy asymptotics can be identified, although in terms of a rather implicit effective problem. In [39, Thm. 1.5] it is then proven that, when the magnetic field is parallel to the 3D boundary, the effective model is still given by the 1D functional (2.3) above.

One of the major differences for samples with non-smooth boundary is that one expects [15, 45,46,47, 50] a shift of the third critical field, provided there is at least one corner with acute opening angle \( 0< \beta < \pi \): more precisely, the transition to the normal state should occur [15, Thm. 1.4] for applied fields larger than

$$\begin{aligned} H_{\mathrm {c}3}= \frac{1}{\mu (\beta ) \varepsilon ^2} \end{aligned}$$
(2.7)

where

$$\begin{aligned} \mu (\beta ) : = \inf \mathrm {spec}_{L^2(W_\beta )} \left( - \left(\nabla + \textstyle \frac{1}{2} i \mathbf{x }^{\perp } \right)^2 \right), \end{aligned}$$
(2.8)

is the ground state energy of a Schrödinger operator with uniform magnetic field in an infinite sector \( W_{\beta } \) of angle \( \beta \). The above result is however conditioned to the inequality \( \mu (\beta ) < \Theta _0\) (see also [51, Chpt. 8.2] and references therein), which is known to be true for \( 0< \beta < \pi /2 + \epsilon \) [11, 30, 46] but is expected to hold in the whole interval \( \beta \in (0, \pi ) \), based on numerical simulations (see, e.g., [11, 12, 30]).

As the applied field gets closer to (2.7) from below, the order parameter concentrates around the corner of smallest opening angle and becomes smaller and smaller everywhere else. Hence, one can speak of a corner superconductivity regime occurring before the transition to the normal state. On the other hand, in [19], we proved that, if \( 1< b < \Theta _0\), superconductivity is still uniform along the boundary (although only in \( L^2 \) sense), leading to the conjectured existence of another critical field

$$\begin{aligned} H_{\mathrm {corner}}= \frac{1}{\Theta _0\varepsilon ^2}, \end{aligned}$$
(2.9)

which marks the transition from surface to corner concentration. Indeed, if \( 1< b < \Theta _0\), then [19, Thm 1.1]

$$\begin{aligned} E_{\varepsilon }^{\mathrm {GL}}= \frac{|\partial \Omega | E^{\mathrm {1D}}_{\star }}{\varepsilon } + {\mathcal {O}}(|\log \varepsilon |^2), \end{aligned}$$
(2.10)

and, more importantly,

$$\begin{aligned} \left\Vert \left|\psi ^{\mathrm {GL}}( \, \cdot \,) \right|^2 - f_{\star }^2\left( \mathrm {dist}( \, \cdot \,, \partial \Omega )/\varepsilon \right) \right\Vert_{L^2(\Omega )} = {\mathcal {O}}(\varepsilon |\log \varepsilon |), \end{aligned}$$
(2.11)

which implies, to leading order, uniform distribution of superconductivity along the boundary layer.

The result of [15] has also been recently improved in [45], where the presence of several corners is taken into account and shown that, under the same unproven assumption, one can identify several critical fields associated to the concentration of the order parameter close to the respective corner. We also stress that, as noted in [7, Rmk. 1.9] (see also [5, 6]), the behavior of superconductivity in presence of corners is expected to be recovered in the case of magnetic steps, i.e., for applied magnetic fields with a jump singularity along a curve.

2.2 GL energy and density asymptotics

Before stating our main results, we have to define the effective problem near the corners. Here we only provide a sketchy definition and in the next § 2.3 we comment further about its well-posedness and heuristic meaning. The model problem is given by first minimizing the GL functional with given magnetic potential and unit magnetic field in large wedge-like domain (see Fig. 2), and then subtracting the surface energy of the outer boundary of the wedge. The wedge domain is supposed to describe the rectified and rescaled area close to each corner, where the only relevant parameter is the opening angle \( \beta _j \).

Fig. 2
figure 2

The region \( \Gamma _{\beta }(L,\ell )\), where \( \beta \) is the opening angle \( {\widehat{AVB}} \), \( L = |{\overline{AV}}| = |{\overline{VB}}| \) and \( \ell = |{\overline{AC}}| = |{\overline{EB}}| \)

We thus define the corner energy as

$$\begin{aligned} \boxed { E_{\mathrm {corner},\beta }: = \displaystyle \lim _{\ell \rightarrow +\infty } \lim _{L \rightarrow + \infty } \left( - 2 L E ^{\mathrm{1D}}_{0}(\ell ) + \inf _{\psi \in {\mathscr {D}}_{\star }(\Gamma _{\beta }(L,\ell ))} {\mathcal {E}}^{\mathrm {GL}}_{1}\left[\psi , \mathbf{F }; \Gamma _{\beta }(L,\ell )\right] \right), }\nonumber \\ \end{aligned}$$
(2.12)

where \( E ^{\mathrm{1D}}_{0}(\ell ) \) is a 1D effective energy analogous to (2.3), which is explicitly given by (see “Appendix A.3” for further details)

$$\begin{aligned} E ^{\mathrm{1D}}_{0}(\ell ) : = \inf _{\alpha \in {\mathbb {R}}} \inf _{f \in H^1([0,\ell ])} {\mathcal {E}}^{\mathrm {1D}}_{0,\alpha }[f], \end{aligned}$$
(2.13)

with

$$\begin{aligned} {\mathcal {E}}^{\mathrm {1D}}_{0,\alpha }[f] : = \int ^{\ell }_0 \text{ dt } \left\{ |\partial _t f|^2 + (t+\alpha )^2 f^2 -\frac{1}{2b} (2f^2-f^4)\right\} , \end{aligned}$$
(2.14)

and we denote by \( \alpha _0 \in {\mathbb {R}}\), \( f_{0}\in H^1([0,\ell ]) \) a corresponding minimizing pair, i.e., \( E ^{\mathrm{1D}}_{0}(\ell ) = {\mathcal {E}}^{\mathrm {1D}}_{0, \alpha _0}[f_{0}] \). The minimization domain is

$$\begin{aligned} {\mathscr {D}}_{\star }(\Gamma _{\beta }(L,\ell )) : = \left\{ \psi \in H^1(\Gamma _{\beta }(L,\ell )) \, \big | \, \left. \psi \right|_{\partial \Gamma _{\mathrm {bd}}\cup \partial \Gamma _{\mathrm {in}}} = \psi _{\star } \right\} . \end{aligned}$$
(2.15)

where the boundaries are identified in Fig. 2 by the segments \( \partial \Gamma _{\mathrm {bd}}= {\overline{AC}} \cup {\overline{EB}} \) and \( \partial \Gamma _{\mathrm {in}}= {\overline{CD}} \cup {\overline{DE}} \) and

$$\begin{aligned} \psi _{\star }(\mathbf{r }(s,t)) : = f_0(t) \exp \left\{ -i \alpha _0 s - \textstyle \frac{1}{2} i st \right\} , \qquad {\text{ for } |s| \geqslant \textstyle \frac{\ell }{\tan (\beta /2)}}. \end{aligned}$$
(2.16)

Here, we have denoted by (st) a set of tubular coordinates similar to (1.10), yielding a smooth parametrization of \( \Gamma _{\beta }(L,\ell )\cap \left\rbrace \mathrm {dist}(\mathbf{r }, V) \geqslant c \ell \right\lbrace \). Any function in \( {\mathscr {D}}_{\star }(\Gamma _{\beta }(L,\ell )) \) has thus to satisfy Dirichlet non-zero boundary conditions on \( \partial \Gamma _{\mathrm {in}}\) and \( \partial \Gamma _{\mathrm {bd}}\) in trace sense, whose role is going to be explained in next § 2.3. Finally, the magnetic potential \( \mathbf{F }\) is fixed and equals

$$\begin{aligned} \mathbf{F }(\mathbf{r }) : = \textstyle \frac{1}{2} \left( - y, x \right) = : \textstyle \frac{1}{2} \mathbf{r }^{\perp }, \end{aligned}$$
(2.17)

in a coordinate system chosenFootnote 1 as in Fig. 3. We also point out that the existence of the limit in (2.12) is not trivial at all and, in fact, it will be the main content of Proposition 2.2. Furthermore, the GL functional in the second term on the r.h.s. of (2.12) is independent of \( \varepsilon \), but still contains the parameter \( b \in (1, \Theta _0^{-1}) \).

Fig. 3
figure 3

Cartesian coordinates for the corner domain

The main result we prove in this paper is about the GL energy asymptotics as \( \varepsilon \rightarrow 0 \), i.e., we derive the expansion of \( E_{\varepsilon }^{\mathrm {GL}}\) up to correction of order o(1) . Compared with the case of domains with smooth boundary, some new terms of order \( {\mathcal {O}}(1) \) appear: each corner indeed contributes to the energy by \( E_{\mathrm {corner}, \beta _j} \), \( \beta _j \) being the corresponding opening angle.

Theorem 2.1

(GL energy asymptotics) Let \(\Omega \subset {\mathbb {R}}^2\) be any bounded simply connected domain satisfying Assumptions 1 and 2. Then, for any fixed

$$\begin{aligned} 1< b < \Theta _0^{-1}, \end{aligned}$$
(2.18)

as \(\varepsilon \rightarrow 0\), it holds

$$\begin{aligned} \boxed {E_{\varepsilon }^{\mathrm {GL}}= \displaystyle \frac{|\partial \Omega | E^{\mathrm {1D}}_{\star }}{\varepsilon } - E_{\mathrm {corr}}\int _{0}^{|\partial \Omega |} \mathrm {d}{\mathsf {s}}\, {\mathfrak {K}}({\mathsf {s}})+ \sum _{j = 1}^N E_{\mathrm {corner}, \beta _j} + o(1). } \end{aligned}$$
(2.19)

Remark 2.1

(Critical field \( H_{\mathrm {c}2}\)) In smooth domains the regime of surface superconductivity corresponds to the parameter interval \( 1< b < \Theta _0^{-1} \), namely the second critical field is \( b = 1 \), while the third one is precisely \( b= \Theta _0^{-1} \). This is motivated by the results in [24, 25] in combination with [32, 35], where it is proven that, for \( b < 1 \), there is still superconductivity in the bulk, while, for \( b > \Theta _0^{-1} \), the normal state is a global minimizer of the GL functional, respectively. The condition \( b > 1 \) is expected to be sharp also for domains with corners and, consequently, we expect that the second critical field is given by

$$\begin{aligned} H_{\mathrm {c}2}= \frac{1}{\varepsilon ^2}, \end{aligned}$$
(2.20)

The value \( 1/\varepsilon ^2 \) can actually be taken as a definition of the second critical field, but, as for smooth domains, it would be necessary to show that, for \( b \leqslant 1 \), there is still superconductivity in the bulk. This has not yet been proven in case of samples with corners, but, based on the results proven in [35], it is highly expected.

Remark 2.2

(Critical field \( H_{\mathrm {corner}}\)) The result proven in Theorem 2.1 substantiates even more than [19] the conjecture about the appearance of an additional critical field

$$\begin{aligned} H_{\mathrm {corner}}= \frac{1}{\Theta _0\varepsilon ^2}, \end{aligned}$$
(2.21)

when corners are present along the sample boundary. Indeed, combining (2.19) and, more importantly, next Proposition 2.1, with [15, Thm. 1.6] (see also [45, Thm. 1.2]), which states the exponential decay of \( \psi ^{\mathrm {GL}}\) in the distance from \( \Sigma \) (still, based on the unproved conjecture on the linear model), one concludes that superconductivity is uniform along the boundary layer until the threshold \( b = \Theta _0^{-1} \) is crossed and, then, concentrates close to the corners with smallest opening angles. More precisely, assuming that all the angles \( \beta _j \) are acute and different, one can identify [45, Rmk. 1.4] a sequence of N critical fields

$$\begin{aligned} H_{\mathrm {corner}}= H_{\mathrm {corner},0} \leqslant H_{\mathrm {corner},1} \leqslant \ldots \leqslant H_{\mathrm {corner},N-1} \leqslant H_{\mathrm {corner},N} = H_{\mathrm {c}3}, \end{aligned}$$
(2.22)

with

$$\begin{aligned} H_{\mathrm {corner},j} = \frac{1}{\mu (\beta _j) \varepsilon ^2}, \qquad \text{ for } 1 \leqslant j \leqslant N, \end{aligned}$$
(2.23)

so that, in between \( H_{\mathrm {corner},j-1} \) and \( H_{\mathrm {corner},j} \), the material is superconducting only close to the \( j-\)th corner \( \mathbf{r }_j \). Let us stress that all these results are conditioned by the request \( \mu (\beta _j) < \Theta _0\) for all the corners, which is expected to hold true (but not proven) for any acute angle \( 0< \beta _j < \pi \).

Once the energy asymptotics is obtained, it is natural to ask whether one can extract information about the behavior of the order parameter, which would then give access to the physically relevant quantities, as the density of Cooper pairs. As already proven in [19, Thm. 1.1], the distribution of superconductivity along the boundary layer is uniform to leading order (see (2.11)). Note that such an estimate goes along with the exponential decay proven in (B.16), which implies that \( \psi ^{\mathrm {GL}}= o(1)\) at distance much larger than \( \varepsilon \) from the boundary \( \partial \Omega \): we can indeed restrict out attention to the boundary layer \( {\mathcal {A}}_{\varepsilon }\) defined in (1.19), since

$$\begin{aligned} \psi ^{\mathrm {GL}}(\mathbf{r }) = {\mathcal {O}}\left( \varepsilon ^{c_1 \cdot c(b) + 1} \right), \qquad \text{ in } \Omega \setminus {\mathcal {A}}_{\varepsilon }, \end{aligned}$$
(2.24)

and by taking \( c_1 \) large, the above quantity can be made arbitrarily small. We thus denote it as \( {\mathcal {O}}(\varepsilon ^{\infty }) \), to stress that it is an arbitrarily large power of \( \varepsilon \).

Remark 2.3

(Refined \( L^2 \) estimate)

An almost direct consequence of the energy asymptotics (2.19) is an improvement of the bound (2.11): setting

$$\begin{aligned} \Omega _{\mathrm {smooth}} : = \left\{ \mathbf{r }\in \Omega | \mathrm {dist}(\mathbf{r }, \Sigma ) \geqslant c_2 \varepsilon |\log \varepsilon | \right\} , \end{aligned}$$
(2.25)

for some large enough constant \( c_2 > 0 \), one has

$$\begin{aligned} \left\Vert \left| \psi ^{\mathrm {GL}}(\, \cdot \,) \right|^2 - {f_{\star }}^2\left( \mathrm {dist}(\, \cdot \,, \partial \Omega )/\varepsilon \right) \right\Vert_{L^2(\Omega _{\mathrm {smooth}} )} = o(\varepsilon |\log \varepsilon |). \end{aligned}$$
(2.26)

The estimates (2.11) or (2.26) do not exclude however the presence of vortices or region with very little superconductivity left close to the boundary and, therefore, one would like to prove a bound in a stronger norm, e.g., in \( L^{\infty } \), which is stated in the next Proposition 2.1.

Proposition 2.1

(GL order parameter) Under the same assumptions of Theorem 2.1,

$$\begin{aligned} \boxed {\left\Vert |\psi ^{\mathrm {GL}}(\mathbf{r })| - {f_{\star }}(0) \right\Vert_{L^{\infty }(\partial \Omega _{\mathrm {smooth}}\cap \partial \Omega ) } = o(1). } \end{aligned}$$
(2.27)

Remark 2.4

(Uniform distribution of superconductivity) The estimate (2.27) can in fact be extended to the boundary layer of points \( \mathbf{r }\) such that \( \mathrm {dist}(\mathbf{r }, \partial \Omega _{\mathrm {smooth}}) \leqslant \varepsilon \sqrt{|\log \varepsilon |} \), in the very same way as the analogous result in [23, Thm. 2]. An important consequence is the uniformity of superconductivity in \( {\mathcal {A}}_{\varepsilon }\), where one has

$$\begin{aligned} \left|\psi ^{\mathrm {GL}}(\, \cdot \,) \right| \sim f_{\star }\left( \mathrm {dist}(\, \cdot \,, \partial \Omega )/ \varepsilon \right), \end{aligned}$$
(2.28)

not only in weak sense, as proven in [19], but also pointwise. Strictly speaking, the corner regions are excluded, but, on the one hand, their overall area is \( {\mathcal {O}}( N \varepsilon ^2 |\log \varepsilon |^2) \), i.e., much smaller than \( |{\mathcal {A}}_{\varepsilon }| \), and, on the other, we do expect the minimizer of the corner problem to be close to \( f_{\star }\) almost everywhere but very close to the corner. An interested reader might wonder whether it is possible to show that \( \psi ^{\mathrm {GL}}\) is close to such an effective minimizer in the corner region, but this presumably requires to get some more information about the effective problem (2.12) as well as extract a more precise estimate of the remainders in (2.19).

Remark 2.5

(Current along \( \partial \Omega \)) An important consequence of (2.6) in smooth domains is the non-vanishing of \( \psi ^{\mathrm {GL}}\) close to the boundary, because of the strict positivity of \( f_{\star }\), and thus surface superconductivity is robust w.r.t. the inclusion of the applied magnetic field. In addition, this allows to estimate the current (1.2) along the boundary or, equivalently, the total winding number of \( \psi ^{\mathrm {GL}}\) on \( \partial \Omega \) [23, Thm. 3]:

$$\begin{aligned} \deg \left( \psi ^{\mathrm {GL}}, \partial \Omega \right) = \frac{|\Omega |}{\varepsilon ^2} - \frac{\alpha _{\star }}{\varepsilon }(1 + o(1)). \end{aligned}$$
(2.29)

Such a behavior is similar (although physically different) to the ultrafast rotation regime for angular velocities larger than the third critical one of rotating Bose-Einstein condensates, when vortices are expelled from the boundary region [18, 27] (see also [21, 22, 28] for further results on rotating condensates). In presence of corners, (2.27) guarantees the non-vanishing of \( \psi ^{\mathrm {GL}}\) only far from the corners and prevents us to estimate the current on \( \partial \Omega \). Indeed, the pointwise estimate of the gradient (B.13) allows a variation of order 1 of \( \psi ^{\mathrm {GL}}\) on a scale \( \varepsilon \), which is much smaller than the tangential length of the corner region, thus implying that \( \psi ^{\mathrm {GL}}\) may a priori vanish there.

2.3 Corner effective energy

We now give more details about the corner effective problem. Let us start by identifying more precisely the corner region depicted in Fig. 2. It is meant as a suitable stretching and rescaling (on a scale \( \varepsilon \)) of a local area around any corner of \( \Omega \) of tangential and normal lengths both of order \( \varepsilon |\log \varepsilon | \), as \( \varepsilon \rightarrow 0 \). For later convenience, however, we consider a region where the tangential length L along the angle is different from the normal length \( \ell \). Let then \( \Gamma _{\beta }(L,\ell )\) be a triangle-like region as in Fig. 2, where \( \beta \) is the opening angle at the vertex V and side lengths \( L, \ell > 0 \). In order to reproduce the shape of Fig. 2, we always assume that

$$\begin{aligned} \ell \leqslant \tan \left( \textstyle \frac{\beta }{2}\right) \, L. \end{aligned}$$
(2.30)

We recall the definition of the boundaries \( \partial \Gamma _{\mathrm {in}}, \partial \Gamma _{\mathrm {bd}}\) provided in § 2 and denote by \( \partial \Gamma _{\mathrm {out}}\) the outer boundary \( {\overline{AVB}} \), so that \( \partial \Gamma _{\beta }(L,\ell )= \partial \Gamma _{\mathrm {out}}\cup \partial \Gamma _{\mathrm {in}}\cup \partial \Gamma _{\mathrm {bd}}\).

The effective energy in the corner region is given by a suitably rescaled GL energy with fixed magnetic potential (2.17). The effective variational model is then

$$\begin{aligned} E_{\mathrm {corner},\beta }(L,\ell ): = - 2 L E ^{\mathrm{1D}}_{0}(\ell ) + \inf _{\psi \in {\mathscr {D}}_{\star }(\Gamma _{\beta }(L,\ell ))} {\mathcal {E}}^{\mathrm {GL}}_{1}\left[\psi , \mathbf{F }; \Gamma _{\beta }(L,\ell )\right], \end{aligned}$$
(2.31)

where \( {\mathscr {D}}_{\star }(\Gamma _{\beta }(L,\ell )) \) is defined in (2.15). The heuristics behind the choice (2.31) is that in the surface superconductivity regime each portion of the boundary of the sample yields a (leading order) energy contribution proportional to \( E^{\mathrm {1D}}_{\star }\) times its length, which equals \( E^{\mathrm {1D}}_{\star }|\partial \Gamma _{\mathrm {out}}| = 2 L E^{\mathrm {1D}}_{\star }\) in the case of \( \Gamma _{\beta }(L,\ell )\). Indeed, the boundaries \( \partial \Gamma _{\mathrm {in}}\) and \( \partial \Gamma _{\mathrm {bd}}\) are not expected to give any energy contribution. More precisely, \( \partial \Gamma _{\mathrm {in}}\) is immersed in the bulk, where the order parameter is exponentially small in \( \ell \) and it could have been removed from the outset by consider a solid wedge; similarly, \( \partial \Gamma _{\mathrm {bd}}\) is a fictitious boundary, whose role is to separate the corner region from the rest. Mathematically, the non-zero Dirichlet conditions on \( \partial \Gamma _{\mathrm {in}}\) and \( \partial \Gamma _{\mathrm {bd}}\) in the minimization domain \( {\mathscr {D}}_{\star }\) guarantee that those portions of the boundary do not contribute to surface superconductivity.

Once the boundary energy \( 2 L E^{\mathrm {1D}}_{\star }\) has been subtracted, what remains is precisely the additional energy due to the presence of the corner. Such an energy is indeed of purely geometric nature and is generated by the constraint on the boundary \( \partial \Gamma _{\mathrm {out}}\): in order to reproduce the correct energy along \( \partial \Gamma _{\mathrm {out}}\), the minimizer must behave like the model order parameter \( f_{\star }(t) e^{-i \alpha _{\star }s} \) in a layer of width \( {\mathcal {O}}(1) \) around \( \partial \Gamma _{\mathrm {out}}\), but the coordinate s has a jump on the bisectrix of the domain and thus such a behavior is allowed only close far from the corner. The modulus of the minimizer \( f_{\star }(t) \) is in fact well defined and continuous everywhere, since it depends on the normal coordinate which is continuous as well. Hence, in order to glue together the two model profiles, any minimizer must accommodate a non-trivial phase factor, which must be genuinely 2D, because no 1D function can adjust the jump of \( - i \alpha _{\star }s \) along the bisectrix. Unfortunately, the explicit expression of such a phase remains unknown, expect in certain specific cases (for almost flat angles, see [20]).

The GL energy functional appearing in (2.31) is gauge invariant but we have chosen to work in a prescribed local gauge, i.e., we have made an explicit choice of the vector potential \( \mathbf{F }\), generating a unit magnetic field. In this respect the GL energy in (2.31) is similar to the effective functional studied in [15, Eq. (1.11)], although both the parameter regime and the domain are slightly different. Such a difference reflects indeed the different behavior of the minimizer: in the present setting it decays in the distance from the outer boundary, whereas in [15], the decay is in the distance from the corner.

Recalling that L and \( \ell \) are obtained via a rescaling from the tangential and normal length of the corner region and thus, in the original problem in \( \Omega \), are actually of order \( |\log \varepsilon | \gg 1 \), we have to study the limit \( L, \ell \rightarrow + \infty \) of (2.31).

Proposition 2.2

(Corner energy) Let \( \left\rbrace \ell _n \right\lbrace _{n \in {\mathbb {N}}} \), \( \left\rbrace L_n \right\lbrace _{n \in {\mathbb {N}}} \) be two monotone sequences with \( \ell _n, L_n \rightarrow + \infty \), as \( n \rightarrow + \infty \), and \( \beta \in (0, 2\pi ) \), such that \( 1 \ll \ell _n \leqslant \tan \left( \beta /2 \right) \, L_n \leqslant C \ell _n^{a} \) for some \( a > 0 \). Then, for any \( 1< b < \Theta _0^{-1} \), the limit

$$\begin{aligned} \lim _{n \rightarrow + \infty } E_{\mathrm {corner},\beta }(L_n,\ell _n) = : E_{\mathrm {corner},\beta }\end{aligned}$$
(2.32)

exists, it is finite and independent of the chosen sequences.

As stated in Proposition 2.2 (see also Proposition 3.4), the corner energy \( E_{\mathrm {corner},\beta }\) is bounded for any \( \beta \in (0,2\pi ) \), although we have no information on its sign. In fact, it might as well be zero. In a companion paper [20] however we prove that, when \( \beta \) is close to \( \pi \), this is not the case (see also below).

Once the well-posedness of the model problem has been proven, it is then natural to ask whether one can derive the explicit dependence of \( E_{\mathrm {corner},\beta }\) on the angle \( \beta \). So far we have not found such an expression but, based on some heuristic arguments, we formulate below an unproven conjecture, which is inspired again by the Gauss–Bonnet theorem. Indeed, the first order correction to the GL energy asymptotics in smooth domains reads equivalently

$$\begin{aligned} -E_{\mathrm {corr}}\int _{0}^{|\partial \Omega |} \mathrm {d}{\mathsf {s}}\, {\mathfrak {K}}({\mathsf {s}})= -2 \pi E_{\mathrm {corr}}. \end{aligned}$$
(2.33)

In presence of corner singularities on \( \partial \Omega \), the Gauss–Bonnet theorem has to be modified to take into account the corners: the only correction is that the integral of the curvature must now be performed over the smooth part of \( \partial \Omega \) and each corner yields a contribution proportional to its opening angle

$$\begin{aligned} \int _{\partial \Omega _{\mathrm {smooth}}} \mathrm {d}{\mathsf {s}}\, {\mathfrak {K}}({\mathsf {s}})+ \sum _{j =1}^N (\pi - \beta _j) = 2 \pi . \end{aligned}$$

Therefore, one can think of the above identity as if each corner contributes to the mean curvature with a Dirac mass multiplied by \( \pi - \beta _j \) and the integral is meant in distributional sense, i.e., formally replacing the curvature \( {\mathfrak {K}}({\mathsf {s}}) \) with

$$\begin{aligned} {\mathfrak {K}}({\mathsf {s}}) + \sum _{j =1}^N (\pi - \beta _j) \delta ({\mathsf {s}}_j), \end{aligned}$$

which, if substituted on the r.h.s. of (2.33), yields

$$\begin{aligned} -E_{\mathrm {corr}}\int _{\partial \Omega _{\mathrm {smooth}}} \mathrm {d}{\mathsf {s}}\, {\mathfrak {K}}({\mathsf {s}})- E_{\mathrm {corr}} \sum _{j =1}^N (\pi - \beta _j). \end{aligned}$$

After a direct comparison with the asymptotics proven in Theorem 2.1, i.e.,

$$\begin{aligned} -E_{\mathrm {corr}}\int _{\partial \Omega _{\mathrm {smooth}}} \mathrm {d}{\mathsf {s}}\, {\mathfrak {K}}({\mathsf {s}})+ \sum _{j =1}^N E_{\mathrm {corner}, \beta _j} \end{aligned}$$

it is then very natural to state the conjecture below. Note that, if true, the conjecture would imply that the next-to-leading order term in the GL energy expansion would always be given by \( - 2\pi E_{\mathrm {corr}} \), irrespective of the presence of corners.

Conjecture 1

(Corner energy) For any \( 1< b < \Theta _0^{-1} \) and \( \beta \in (0,2\pi ) \), one has

$$\begin{aligned} E_{\mathrm {corner},\beta }= -(\pi - \beta ) E_{\mathrm {corr}}. \end{aligned}$$
(2.34)

Remark 2.6

(Acute/obtuse angles) In the linear case, i.e., for a magnetic Schrödinger operator with uniform magnetic field in an infinite wedge, it is expected [15, Rmk. 1.1] and numerically verified [1, 13] that the ground state energy changes for acute or obtuse angles: for the former it is a strictly increasing function of the angle, which equals \( \Theta _0\) for flat angles, while it is believed to remain constantly equal to \( \Theta _0\) for any obtuse angle. On the opposite, in the nonlinear case, the above Conjecture would provide the same expression for acute and obtuse angles.

As already anticipated, we prove in [20] that in a wedge with opening angle \( \pi - \delta \), \( 0 < \delta \ll 1 \), the corner energy is given by

$$\begin{aligned} E_{\mathrm {corner},\beta }= -\delta E_{\mathrm {corr}}+ {\mathcal {O}}(\delta ^{4/3}|\log \delta |) + {\mathcal {O}}(\ell ^{-\infty }), \end{aligned}$$
(2.35)

i.e., it coincides to leading order in \( \delta \) with the conjectured expression. Furthermore, this also shows that the corner energy \( E_{\mathrm {corner},\beta }\) is non-trivial, at least for angles close to the flat one.

3 Corner effective problems

This section is mainly devoted to the proof of Proposition 2.2, i.e., the existence of the limit defining the effective energy contribution of each corner, and the discussion of the properties of such a limit. For later convenience, we also study another minimization problem in \( \Gamma _{\beta }\) with different boundary conditions and show that it asymptotically provides the same effective energy (Proposition 3.5).

3.1 Surface superconductivity in a finite strip

We start by studying a simple minimization problem in a finite strip. Similar problems have already been studied in [4, 23, 49], taking into account the limit of an infinite strip. Here, instead, the focus is more on boundary conditions and their effect on the ground state energy. We are going to apply the corresponding obtained results to the minimization in (2.12) to derive Proposition 3.5.

After a local gauge transformation and blow-up on a scale \( \varepsilon \), the leading contribution to the GL energy of a portion of the boundary layer of \( \Omega \) of tangential length \( \varepsilon L \) and normal length \( \varepsilon \ell \), suitably far from any corner, is (see, e.g., [23, Lemmas 2 & 4])

$$\begin{aligned} {\mathcal {G}}\left[\psi ; R(L,\ell )\right] : = \int _0^{L} \mathrm {d}s \int _{0}^{\ell } \mathrm {d}t \, \left\{ \left| \partial _t \psi \right|^2 + \left| \left(\partial _s - i t\right) \psi \right|^2 - \frac{1}{2b} \left( 2 |\psi |^2 - |\psi |^4 \right) \right\} , \end{aligned}$$
(3.1)

where \( L,\ell > 0 \), \( b \in (1, \Theta _0^{-1}) \) and \(R(L,\ell )\) stands for the rectangle

$$\begin{aligned} R(L,\ell ): = \left[ 0, L \right] \times \left[ 0, \ell \right], \qquad \text{ with } \ell \gg 1. \end{aligned}$$
(3.2)

We study two simple minimization problems associated to the energy (3.1). First, we set

$$\begin{aligned} E_{\mathrm {D}}(R(L,\ell )) : = \inf _{\psi \in {\mathscr {D}}_{\mathrm {D}}(R(L,\ell ))} {\mathcal {G}}\left[\psi ; R(L,\ell )\right], \end{aligned}$$
(3.3)

and denote by \( \psi _{\mathrm {D}}\) any corresponding minimizer. The minimization domain is given by

$$\begin{aligned}&{\mathscr {D}}_{\mathrm {D}}(R(L,\ell )) : = \left\{ \psi \in H^1(R(L,\ell )) \, \big | \, \psi (0,t) = f_{0}(t), \psi (L,t) = f_{0}(t) e^{-i \alpha _0 L}, \right. \nonumber \\&\qquad \left. \psi (s,\ell ) = f_0(\ell ) e^{-i \alpha _0 s} \right\} , \end{aligned}$$
(3.4)

where the boundary conditions are meant in trace \( H^{1/2}\)-sense and we recall that \( f_{0}, \alpha _0 \) is a minimizing pair (see also “Appendix A.2”) of (2.13). The label \( \mathrm {D} \) stands for the Dirichlet-type conditions at \( s = 0 \) and \( s = L \). The heuristic meaning of such conditions is the following:

  • on the boundary between the surface and the bulk region, i.e., for \( t = \ell \), the order parameter is exponentially small and the same holds true for \( f_{0}(\ell ) \), so the contribution of the boundary conditions there is expected to be negligible; for this reason we could as well have set \( \psi = 0 \) at \( t = \ell \), but this would make the analysis more complicated;

  • at the normal boundaries \( s = 0 \) or \( s = L \), the order parameter is set equal to the ideal minimizer (see § 1.2);

  • no condition is set on the boundary \( t = 0 \), which is meant to coincide with a blow-up of a portion of \( \partial \Omega \): this is crucial to capture the key features of surface superconductivity and leads to Neumann conditions along the line \( t = 0 \).

By setting \( \psi : = \chi + f_0(t) e^{-i\alpha _0s} \), one can reduce the variational problem (3.3) to the minimization of a functional of \( \chi \) with zero Dirichlet conditions on the boundaries \( s = 0, L \) and \( t = \ell \). This easily allows to deduce (see, e.g., [41, Chapt. 4]) the existence of a minimizer, its smoothness and the fact that any minimizer solves

$$\begin{aligned} - \left( \nabla - i t \mathbf{e }_s \right)^2 \psi = \frac{1}{b} \left(1 - \left|\psi \right|^2 \right) \psi . \end{aligned}$$
(3.5)

The alternative version of (3.4) is provided by a modification of the energy: we define

$$\begin{aligned} {\widetilde{{\mathcal {G}}}}\left[\psi ; R(L,\ell )\right] : = {\mathcal {G}}[\psi ; R(L,\ell )] - \int _0^{\ell }\mathrm {d}t\, \frac{F_{0}(t)}{f_{0}^2(t)} j_t\left[\psi \right] \bigg \vert _{s=0}^{s=L}, \end{aligned}$$
(3.6)

where \( F_{0}\) is the potential function (see also “Appendix A.3”)

$$\begin{aligned} F_{0}(t) : = 2\int _0^t\mathrm {d}\eta \, \left(\eta + \alpha _0 \right) f_{0}^2(\eta ), \end{aligned}$$
(3.7)

and \( j_t \) is the normal component of the current \( \mathbf{j }[\psi ] \) given in (1.2), i.e., \( j_t[\psi ] = \frac{i}{2} \left( \psi \partial _t \psi ^* - \psi ^* \partial _t \psi \right)\). The boundary terms appearing in (3.6) are non-trivial only if the phase of \( \psi \) varies along the normal to the boundary, which is obviously not the case for the reference function \( f_{0}(t) e^{- i\alpha _0 s} \). The reason why such terms have been added to the energy will become clear later (see the proof of Proposition 3.1 and in particular (3.22)). The minimization of (3.6) is performed on a domain without constraints on the boundaries \( s = 0 \) and \( s = L \), i.e., we set

$$\begin{aligned} E_{\mathrm {N}}(R(L,\ell )) : = \inf _{\psi \in {\mathscr {D}}_{\mathrm {N}}(R(L,\ell ))} {\widetilde{{\mathcal {G}}}}\left[\psi ; R(L,\ell )\right], \end{aligned}$$
(3.8)

where

$$\begin{aligned} {\mathscr {D}}_{\mathrm {N}}(R(L,\ell )) : = \left\{ \psi \in H^1(R(L,\ell )) \, \big | \, \psi (s,\ell ) = f_0(\ell ) e^{-i\alpha _0 s} \right\} \end{aligned}$$
(3.9)

and we denote by \( \psi _{\mathrm {N}}\) any corresponding minimizer, which enjoys the same properties as \( \psi _{\mathrm {D}}\), except for conditions of magnetic Neumann-type at \( s = 0 \) and \( s = L\), i.e.,

$$\begin{aligned} \left. \left[ \left( \partial _s + i \alpha _0 \right) \psi _{\mathrm {N}}- i \frac{F_0(t)}{f_0^2(t)} \left( \partial _t \psi _{\mathrm {N}}\right) \right] \right|_{s = 0, L} = 0. \end{aligned}$$
(3.10)

The surface superconductivity behavior occurs for \( 1< b < \Theta _0^{-1} \) and is characterized by the emergence of the 1D effective model (2.13) or, equivalently, (2.3).

Proposition 3.1

(GL energies on a finite strip) For any \( 1< b < \Theta _0^{-1} \) and \( L > 0 \), as \( \ell \rightarrow \infty \),

$$\begin{aligned} E_{\mathrm {D}/\mathrm {N}}(R(L,\ell )) = L \left( E ^{\mathrm{1D}}_{0}(\ell ) + {\mathcal {O}}(\ell ^{-\infty })\right) = L \left( E^{\mathrm {1D}}_{\star }+ {\mathcal {O}}(\ell ^{-\infty })) \right). \end{aligned}$$
(3.11)

Remark 3.1

(Boundary conditions) The boundary condition \( \psi _{\mathrm {N}}(s,\ell ) = f_0(\ell ) e^{-i\alpha _0 s} \) is needed for the asymptotics (3.11) to hold true. The reason is that otherwise one would get an additional energy contribution from the boundary \( t = \ell \), i.e., the energy would be twice the value appearing in (3.11). Indeed, without the condition at \( t = \ell \), exploiting the gauge invariance of (3.1) and replacing \( \psi , - t \mathbf{e }_s \) with \( \psi ^* e^{i \ell s}, -(\ell - t) \mathbf{e }_s \), one can exchange the boundaries \( t = 0 \) and \( t = \ell \), leaving the energy unaffected.

Proof

We first observe that the last estimate is in fact stated in Lemma A.2 in “Appendix A.3”. The rest of the statement is actually proven by showing separately that the first estimate holds true for both \( E_{\mathrm {D}}\) and \( E_{\mathrm {N}}\).

Let us first consider \( E_{\mathrm {D}}(R(L,\ell )) \). For the upper bound, we test \( {\mathcal {G}}\) on the trial state \( f_{0}(t) e^{-i \alpha _0 s} \), which immediately yields \( E_{\mathrm {D}}(R(L,\ell )) \leqslant L E ^{\mathrm{1D}}_{0}(\ell ) \). For the corresponding lower bound we use the same energy splitting used, e.g., in [24], i.e., we set

$$\begin{aligned} \psi _{\mathrm {D}}(s,t) = : f_{0}(t) e^{-i\alpha _0s} u(s,t), \end{aligned}$$
(3.12)

which, via an integration by parts and the variational equation (A.7), leads to

$$\begin{aligned} E_{\mathrm {D}}(R(L,\ell )) = L E ^{\mathrm{1D}}_{0}+ {\mathcal {E}}_0\left[u;R(L,\ell )\right], \end{aligned}$$
(3.13)

where

$$\begin{aligned} {\mathcal {E}}_0\left[u;R(L,\ell )\right]:=\int _0^{L}\mathrm {d}s\int _0^{\ell }\mathrm {d}t\, f_{0}^2 \left\{ \left|\nabla _{s,t} u \right|^2 - 2(t+\alpha _0)j_s[u] +\frac{f_{0}^2}{2b}(1-|u|^2)^2\right\} ,\nonumber \\ \end{aligned}$$
(3.14)

and \( j_s[\psi ] \) is the tangential component of (1.2), i.e., explicitly \( j_s[\psi ] = \textstyle \frac{i}{2} \left( \psi \partial _s \psi ^* - \psi ^* \partial _s \psi \right) \). We stress that the decoupling does not generate any boundary term because \(f_{0}^{\prime } \) vanishes both at \( t = 0 \) and \( t = \ell \) by (A.17): the only non-trivial computation is the following integration by parts

$$\begin{aligned} \int _0^{L}\mathrm {d}s\int _0^{\ell }\mathrm {d}t\, \left[ |u|^2 {f_0^{\prime }}^2 + f_0 f_0^{\prime } \partial _t \left| u \right|^2 \right]= - \int _0^{L}\mathrm {d}s\int _0^{\ell }\mathrm {d}t\, |u|^2 f_0 f_0^{\prime \prime }, \end{aligned}$$

where \( f_0^{\prime \prime } \) can then be replaced via the variational Eq. (A.7).

The key ingredient to bound from below \( {\mathcal {E}}_0[u] \) is the pointwise positivity of the cost function (see (A.21) and (A.22) in “Appendix A.3”)

$$\begin{aligned} K_0 : = f_{0}^2 + F_{0}, \end{aligned}$$
(3.15)

in \( I_{{\bar{\ell }}} = [0,{{\bar{\ell }}}] \) given by (A.23) (recall that \( {\bar{\ell }} = \ell + {\mathcal {O}}(1) \) by (A.24)). Indeed, we integrate by parts twice:

$$\begin{aligned}&-2 \int _0^{L}\mathrm {d}s\int _0^{\ell } \mathrm {d}t \,(t+\alpha _0) f_0^2(t) j_s[u] \nonumber \\&\quad = - \int _0^{L}\mathrm {d}s\int _0^{\ell } \mathrm {d}t \, F_{0}^{\prime }(t) j_s[u] = \int _0^{L}\mathrm {d}s\int _0^{\ell } \mathrm {d}t \, F_{0}(t) \partial _t j_s[u] \nonumber \\&\quad = 2 \int _0^{L}\mathrm {d}s\int _0^{\ell } \mathrm {d}t\, F_{0}(t) \, \mathrm {Im}\left( \partial _t u^* \partial _s u \right) + \int _0^{\ell }\mathrm {d}t\, F_{0}(t)j_t[u]\bigg \vert _{s=0}^{s=L}, \end{aligned}$$
(3.16)

where the boundary terms of the first integration by parts vanish, because \( F_{0}(0) = F_{0}(\ell ) = 0 \), and the last terms vanish as well, since, due to boundary conditions, \( u(0,t) = u(L,t) = 1 \) and thus \( j_t[u] = 0 \) there.

Using (A.25) and the simple bound \( 2 |\mathrm {Im}(a b) | \leqslant |a|^2 + |b|^2 \), one then obtains as in [23, Eq. (4.38)] (see also [23, Sect. 2.3 & Proof of Prop. 4.2])

$$\begin{aligned}&{\mathcal {E}}_0\left[u; R(L,\ell )\right] \geqslant \int _0^{L}\mathrm {d}s\int _0^{{{\bar{\ell }}}} \mathrm {d}t\,\left\{ K_0(t) \left( |\partial _s u|^2+|\partial _t u|^2 \right) + \frac{1}{2b}f_0^4 (1-|u|^2)^2 \right\} \nonumber \\&\quad + \int _0^{L}\mathrm {d}s\int _{{\bar{\ell }}}^{\ell } \mathrm {d}t\, \left\{ f_0^2 \left| \nabla u \right|^2 + 2 F_{0}(t) \, \mathrm {Im}\left( \partial _t u^* \partial _s u \right) \right\} \nonumber \\&\quad \geqslant \int _0^{L}\mathrm {d}s\int _{{\bar{\ell }}}^{\ell } \mathrm {d}t\, \left\{ f_0^2 \left| \nabla u \right|^2 + 2 F_{0}(t) \, \mathrm {Im}\left( \partial _t u^* \partial _s u \right) \right\} , \end{aligned}$$
(3.17)

by (A.22) and the positivity of the last term on the r.h.s. of the first line. It thus remains to estimate the quantity on the r.h.s. of (3.17) above, which can be done by integrating by parts back:

$$\begin{aligned}&\int _0^{L}\mathrm {d}s\int _{{\bar{\ell }}}^{\ell } \mathrm {d}t\, \left\{ f_0^2 \left| \nabla u \right|^2 + 2 F_{0}(t) \, \mathrm {Im}\left( \partial _t u^* \partial _s u \right) \right\} \nonumber \\&\quad = \int _0^{L}\mathrm {d}s\int _{{\bar{\ell }}}^{\ell } \mathrm {d}t \, \left\{ f_0^2 \left| \nabla u \right|^2 - 2 (t+\alpha _0) j_s[f_0 u] \right\} - 2 F_0({\bar{\ell }}) \int _0^{L}\mathrm {d}s \, j_s[u]\bigg \vert _{t = {{\bar{\ell }}}}. \end{aligned}$$
(3.18)

Now, exploiting (A.11) and the fact that \( {{\bar{\ell }}} = \ell + {\mathcal {O}}(1) \), we deduce that

$$\begin{aligned} f_{0}(t) = {\mathcal {O}}(\ell ^{-\infty }), \qquad f_{0}^{\prime }(t) = {\mathcal {O}}(\ell ^{-\infty }), \qquad \text{ for } \text{ any } t \geqslant {{\bar{\ell }}}. \end{aligned}$$
(3.19)

Hence, \( \left| \nabla \psi _{\mathrm {D}}\right| = f_0 \left| \nabla u \right| + {\mathcal {O}}(\ell ^{-\infty })\) in \( I_{\ell }\setminus I_{{\bar{\ell }}}\). Now, since \( F_0(\ell ) = 0 \), \( F_0({{\bar{\ell }}}) \leqslant C \ell f^2_0({{\bar{\ell }}}) \), we can bound the boundary term (last term in (3.18)) by

$$\begin{aligned} C L \sup _{s \in [0,L]} \left| \psi _{\mathrm {D}}(s,{\bar{\ell }}) \right| \left| \nabla \psi _{\mathrm {D}}(s,{{\bar{\ell }}}) \right| = L {\mathcal {O}}(\ell ^{-\infty }), \end{aligned}$$

thanks to (B.38) and the bound \( \left\Vert \nabla \psi _{\mathrm {D}}\right\Vert_{\infty } \leqslant C \) on the gradient of \( \psi _{\mathrm {D}}\) (see (B.13)). For the same reason, the first term on the r.h.s. of (3.18) can be bounded from below via Cauchy inequality and (B.17) by

$$\begin{aligned} - C \int _0^{L}\mathrm {d}s\int _{{\bar{\ell }}}^{\ell } \mathrm {d}t\, \left(t + \alpha _0 \right)^2 \left| \psi _{\mathrm {D}}\right|^2 = {\mathcal {O}}(L \ell ^{-\infty }), \end{aligned}$$

which finally yields,

$$\begin{aligned} \int _0^{L}\mathrm {d}s\int _{{\bar{\ell }}}^{\ell } \mathrm {d}t\, \left\{ f_0^2 \left| \nabla u \right|^2 + 2 F_{0}(t) \, \mathrm {Im}\left( \partial _t u^* \partial _s u \right) \right\} = {\mathcal {O}}(L \ell ^{-\infty }), \end{aligned}$$
(3.20)

and thus the statement.

The proof for the modified functional (3.6) is very similar. The upper bound is obtained by evaluating the energy on the trial state \( f_0(t) e^{-i\alpha _0 s} \): notice that the phase of such a function is independent of t, then the normal component \( j_t \) of its current is identically zero and therefore the boundary terms in \( {\widetilde{{\mathcal {G}}}} \) do not yield any additional contribution. The final outcome is the very same bound \( E_{\mathrm {N}}(R(L,\ell )) \leqslant L E ^{\mathrm{1D}}_{0}(\ell ) \) as before.

One can then apply the splitting technique, setting (for a different u than before)

$$\begin{aligned} \psi _{\mathrm {N}}(s,t) = : f_0(t) e^{-i\alpha _0 s} u(s,t), \end{aligned}$$
(3.21)

to get the identity \( E_{\mathrm {N}}(R(L,\ell )) = L E ^{\mathrm{1D}}_{0}+ {\widetilde{{\mathcal {E}}}}_0[u;R(L,\ell )] \), where

$$\begin{aligned} {\widetilde{{\mathcal {E}}}}_0[u;R(L,\ell )] := {\mathcal {E}}_0[u; R(L,\ell )] - \int _0^{\ell }\mathrm {d}t\, F_{0}(t)j_t[u]\bigg \vert _{s=0}^{s=L}. \end{aligned}$$
(3.22)

The proof of the lower bound is then completely analogous to the one above: the only nontrivial observation is that the first integration by parts in (3.16) generates the same outcome, because of the vanishing of \( F_0 \) at the boundaries, and the last terms in (3.16) are exactly compensated by the boundary terms in the functional (3.22), so that they sum up to zero. Actually, this was the main reason to add those terms to (3.6) in first place. The lower bound then follows from the positivity of \( K_0 \), exactly as above. \(\square \)

A straightforward adaptation of the above arguments leads to the following result on a modified problem with twisted boundary conditions, which is going to play a role later.

Proposition 3.2

(GL energy with twisted boundary conditions) Let \( \varkappa \in [0,2\pi ) \), \( 1< b < \Theta _0^{-1} \) and \( L > 0 \). Let also

$$\begin{aligned}&E_{\mathrm {D},\varkappa }(R(L,\ell )) : = \inf _{\psi \in {\mathscr {D}}_{\mathrm {D},\varkappa }(R(L,\ell ))} {\mathcal {G}}[\psi ; R(L,\ell )], \end{aligned}$$
(3.23)
$$\begin{aligned}&{\mathscr {D}}_{\mathrm {D},\varkappa }(R(L,\ell )) : = \left\{ \psi \in {\mathscr {D}}_{\mathrm {N}}(R(L,\ell )) \, \big | \, \psi (0,t) = f_0(t) e^{i \varkappa }, \psi (L,t) = f_0(t) e^{-i\alpha _0L} \right\} .\nonumber \\ \end{aligned}$$
(3.24)

Then, as \( \ell \rightarrow \infty \),

$$\begin{aligned} E ^{\mathrm{1D}}_{0}(\ell ) L + {\mathcal {O}}(L \ell ^{-\infty }) \leqslant E_{\mathrm {D},\varkappa }(R(L,\ell )) \leqslant E ^{\mathrm{1D}}_{0}(\ell ) L + \frac{C}{L}. \end{aligned}$$
(3.25)

Proof

The lower bound is obtained via the splitting technique and the positivity of the cost function as discussed in the proof of Proposition 3.1. For the upper bound it is sufficient to test the functional on the trial state

$$\begin{aligned} f_0(t) e^{- i \alpha _0s} e^{i \frac{\varkappa (L - s)}{L}}, \end{aligned}$$

and recall the optimality of the phase \( \alpha _0\) yielding (A.8). \(\square \)

We conclude this section with a result which will be used later in the paper. In extreme synthesis it states that, if one has an a priori upper bound on \( {\mathcal {E}}_0[u,R(L,\ell )] \), then it is possible to extract some useful information on the corresponding order parameter \( \psi (s,t) \) and show for instance that it is pointwise close to \( f_0(t) e^{- i \alpha _0 s} \) up to a smooth phase factor.

Proposition 3.3

(Order parameter estimates) Let \( \psi \) be a solution of (3.5) in the strip \( R(L,\ell )\), with \( \ell \geqslant t_0 > 0 \) and \( L > 0 \), satisfying the boundary conditions in (3.9) and (3.10), and let u be defined as in (3.21). Let also \( {\widetilde{{\mathcal {E}}}}_0[u; R(L,\ell )] \) be the functional defined in (3.22) in the strip \( R(L,\ell )\) and assume that

$$\begin{aligned} {\widetilde{{\mathcal {E}}}}_0[u;R(L,\ell )] \leqslant {\mathfrak {e}}\leqslant 1, \end{aligned}$$
(3.26)

for some \( {\mathfrak {e}}> 0 \). Then, if \( 1< b < \Theta _0^{-1} \),

$$\begin{aligned} \left\Vert f_0^2 \nabla u \right\Vert_{L^2(R(L,\ell ))}^2 \leqslant C {\mathfrak {e}}+ {\mathcal {O}}(L \ell ^{-\infty }). \end{aligned}$$
(3.27)

Moreover, for any \( 0 < T \leqslant {{\bar{\ell }}} \), there exists a finite constant \( C > 0 \), such that

$$\begin{aligned}&\left| \left|\psi (s,t)\right| - f_0(t) \right| \leqslant \frac{C{\mathfrak {e}}^{1/4} + {\mathcal {O}}(L\ell ^{-\infty })}{\sqrt{\min _{[0,T]} f_0}}, \qquad \text{ for } \text{ any } (s,t) \in R_{L,T}, \end{aligned}$$
(3.28)
$$\begin{aligned}&\left. \left| \partial _s \int _0^\ell \mathrm {d}t \, \left| \psi \right|^2 \right| \right|_{s = L} \leqslant C \left\{ {\mathfrak {e}}+ \sqrt{{\mathfrak {e}}L }+ \frac{1}{L} \left[ \frac{{\mathfrak {e}}^{1/4} + {\mathcal {O}}(L\ell ^{-\infty })}{\sqrt{\min _{[0,T]} f_0}} + e^{-c(b)T} \right] + L \right\} .\nonumber \\ \end{aligned}$$
(3.29)

Proof

Applying elliptic regularity theory to the equation satisfied by \( \psi \) one can prove as in Lemma B.1 that

$$\begin{aligned} \left\Vert \nabla \psi \right\Vert_{L^{\infty }(R(L,\ell ))} \leqslant C, \end{aligned}$$
(3.30)

Furthermore, \( \psi \) satisfies the Agmon estimates (B.17) and (B.38).

The key estimate is then the positivity of the cost function \( K_0 \) in \( I_{{\bar{\ell }}}\), as well as the lower bound given by (A.27), i.e.,

$$\begin{aligned} K_0(t) = f_0^2(t) + F_0(t) \geqslant c_b f_0^4(t), \qquad \text{ for } \text{ any } t \in I_{{\bar{\ell }}}. \end{aligned}$$
(3.31)

Indeed, by acting as in the proof of Proposition 3.1, one immediately gets

$$\begin{aligned} {\widetilde{{\mathcal {E}}}}_0[u, R(L,\ell )]&\geqslant \int _{R(L,{{\bar{\ell }}})} \mathrm {d}s \mathrm {d}t \, K_0(t) \left|\nabla u \right|^2 {+ \frac{1}{2b} \int _{R(L, \ell )} \mathrm {d}s \mathrm {d}t \, f_0^4(t) \left(1 - \left|u\right|^2 \right)^2} \nonumber \\&\quad + {\mathcal {O}}(L \ell ^{-\infty }). \end{aligned}$$
(3.32)

Plugging in (3.31) above, one obtains (3.27) and

$$\begin{aligned} \int _{R(L, \ell )} \mathrm {d}s \mathrm {d}t \, f_0^4(t) \left(1 - \left|u\right|^2 \right)^2 \leqslant C {\mathfrak {e}}+ {\mathcal {O}}(L\ell ^{-\infty }). \end{aligned}$$
(3.33)

We now address (3.28): the starting point is provided by (3.33), which essentially implies that |u| is approximately constant and equal to 1. The idea of proof goes back to [10] and it has been used several times since then (see, e.g., [27]). Fix \( 0 < T \leqslant {{\bar{\ell }}} \) and assume by contradiction that there was a point \( (s_0,t_0) \in R_{L,T} \), where

$$\begin{aligned} \left| \left|\psi (s_0,t_0)\right| - f_0(t_0) \right| \geqslant \frac{c \, {\bar{{\mathfrak {e}}}}^{1/4}}{\sqrt{\min _{[0,T]} f_0}}, \end{aligned}$$
(3.34)

for suitable \( c > 0 \) and \( {\bar{{\mathfrak {e}}}} \geqslant {\mathfrak {e}}\) to be adjusted later. Then, by (3.30) and the analogous bound for \( |f_0^{\prime }(t)| \) (see (A.11)), we deduce that there would exist also a ball of radius \( \varrho : = c' {\bar{{\mathfrak {e}}}}^{1/4}/\sqrt{f_0(t_0)} \) centered in \( (s_0,t_0) \), with \( c' \) a constant proportional to c and depending only on the a priori bounds on the gradients, so that

$$\begin{aligned} \left| \left|\psi (s,t)\right| - f_0(t) \right| \geqslant \frac{1}{2}\frac{c \, {\bar{{\mathfrak {e}}}}^{1/4}}{\sqrt{\min _{[0,T]} f_0}}, \qquad \text{ in } {\mathcal {B}}_{\varrho }(s_0,t_0) \cap R_{L,T}. \end{aligned}$$

Furthermore, we can also assume that at least one quarter of the ball is contained inside \( R_{L,T} \). Hence,

$$\begin{aligned}&\int _{R(\ell ,T)} \mathrm {d}s \mathrm {d}t \, f_0^4(t) \left(1 - \left|u\right|^2 \right)^2 = \int _{R(\ell ,T)} \mathrm {d}s \mathrm {d}t \, \left(f_0^2(t) - \left|\psi \right|^2 \right)^2 \\&\quad \geqslant \frac{\pi c^2}{16} \varrho ^2 {\bar{{\mathfrak {e}}}}^{1/2} \min _{[0,T]} f_0 \geqslant C c^4 {\bar{{\mathfrak {e}}}} \end{aligned}$$

where C is a positive constant independent of c. Therefore, by taking c large enough and \( {\bar{{\mathfrak {e}}}} = {\mathfrak {e}}+ {\mathcal {O}}(L \ell ^{-\infty }) \), we would get a contradiction with (3.33), which completes the proof.

In order to finally get (3.29), we can restrict the integration to the interval \( t \in [0, {{\bar{\ell }}}] \), since the rest is exponentially small. We then compute

$$\begin{aligned} \partial _s \int _0^{{{\bar{\ell }}}} \mathrm {d}t \, \left| \psi (s,t) \right|^2 \bigg |_{s = L} = \int _0^{L} \mathrm {d}s \int _0^{{{\bar{\ell }}}} \mathrm {d}t \, \left[ \chi (s) \partial _s^2 \left| \psi (s,t) \right|^2 + \chi ^{\prime }(s) \partial _s \left| \psi (s,t) \right|^2 \right],\nonumber \\ \end{aligned}$$
(3.35)

for any smooth \( \chi \) such that \( \chi (L) = 1 \). Taking \( \chi (s) = s/L \), we get

$$\begin{aligned}&\bigg | \int _0^{\delta } \mathrm {d}s \int _0^{{{\bar{\ell }}}} \mathrm {d}t \, \chi ^{\prime }(s) \partial _s \left| \psi (s,t) \right|^2 \bigg | \leqslant \frac{1}{L} \int _0^{{\bar{\ell }}} \mathrm {d}t \, \left| \left| \psi (\delta ,t) \right|^2 - \left| \psi (0,t) \right|^2 \right| \\&\quad \leqslant \frac{C}{L} \left[ \frac{{\mathfrak {e}}^{1/4} + {\mathcal {O}}(L\ell ^{-\infty })}{\sqrt{\min _{[0,T]} f_0}} + e^{-c(b)T} \right], \end{aligned}$$

by (3.28) and (B.38). For the first term on the r.h.s. of (3.35), we extend the integration in t to \( \ell \): using Neumann boundary conditions at \( t = \ell \), one gets

$$\begin{aligned} \int _{{\bar{\ell }}}^{\ell } \mathrm {d}t \, \partial _s^2 \left| \psi \right|^2 = \int _{{\bar{\ell }}}^{\ell } \mathrm {d}t \, \Delta \left| \psi \right|^2 + \left. \partial _t \left| \psi \right|^2 \right|_{t = {{\bar{\ell }}}} = \int _{{\bar{\ell }}}^{\ell } \mathrm {d}t \, \Delta \left| \psi \right|^2 + {\mathcal {O}}(\ell ^{-\infty }), \end{aligned}$$

by (B.38) and the pointwise bound on the gradient of \( \psi \). Hence, exploiting the Neumann conditions also at \( t = 0 \) and (3.5), we obtain

$$\begin{aligned}&\int _{0}^{{{\bar{\ell }}}} \mathrm {d}t \, \partial _s^2 \left| \psi \right|^2 = \int _{0}^{\ell } \mathrm {d}t \, \Delta \left| \psi \right|^2 + {\mathcal {O}}(\ell ^{-\infty })= \int _{0}^{\ell } \mathrm {d}t \, \left[ 2 \mathrm {Re}\left( \psi ^* \Delta \psi \right) + 2 \left| \nabla \psi \right|^2 \right] + {\mathcal {O}}(\ell ^{-\infty })\\&\quad = 2 \int _{0}^{\ell } \mathrm {d}t \, \left[ \left| {\left( \nabla - i t \mathbf{e }_s \right)} \psi \right|^2 - \textstyle \frac{1}{b} \left(1 - \left| \psi \right|^2 \right) \left| \psi \right|^2 \right] + {\mathcal {O}}(\ell ^{-\infty }), \end{aligned}$$

which yields, after integration in s,

$$\begin{aligned}&\int _0^{L} \mathrm {d}s \int _0^{\ell } \mathrm {d}t \, \chi (s) \partial _s^2 \left| \psi (s,t) \right|^2 \nonumber \\&\quad = 2 \int _{R(L,\ell )} \mathrm {d}s \mathrm {d}t \, \chi (s) \left[ \left| {\left( \nabla - i t \mathbf{e }_s \right)} \psi \right|^2 - \textstyle \frac{1}{b} \left(1 - \left| \psi \right|^2 \right) \left| \psi \right|^2\right] + {\mathcal {O}}(L \ell ^{-\infty }). \end{aligned}$$
(3.36)

In order to estimate the quantity on the r.h.s. of the expression above we observe that

$$\begin{aligned}&\int _{R(L,\ell )} \mathrm {d}s \mathrm {d}t \, \chi (s) \left[ \left| {\left( \nabla - i t \mathbf{e }_s \right)} \psi \right|^2 - \textstyle \frac{1}{b} \left(1 - \left| \psi \right|^2 \right) \left| \psi \right|^2\right] \\&\quad \leqslant {\widetilde{{\mathcal {E}}}}_0[u] +\frac{1}{2b}\int _{R(L,\ell )} \mathrm {d}s \mathrm {d}t \left| f_0^4 - \left| \psi \right|^4 \right|+ 2 \int _{R(L,\ell )} \mathrm {d}s \mathrm {d}t \, \left(\chi (s) - 1 \right) (t + \alpha _0) f_0^2 j_s[u] \\&\quad \leqslant C \left[ {\mathfrak {e}}+ \sqrt{{\mathfrak {e}}L}+ L \right], \end{aligned}$$

by (3.30) and (3.33). Altogether we get (3.29). \(\square \)

3.2 Properties of \(E_{\mathrm {corner},\beta }(L,\ell )\)

In the present and following Sections, we study the effective model introduced in (2.12) and specifically prove the existence of the limit as well as its boundedness. The key properties we are going to use in the proof of Proposition 2.2 are:

  • change of gauge to replace the magnetic potential \( \mathbf{F }\) with \( \mathbf{a }\simeq - t \mathbf{e }_s \) (Lemma 3.1);

  • uniform boundedness of \( E_{\mathrm {corner},\beta }(L,\ell ) \) and existence of the limit \( L, \ell \rightarrow + \infty \) over suitable subsequences (Proposition 3.4);

  • further properties of the effective model and, in particular, its dependence on the boundary conditions (§ 3.4).

We recall the corner energy defined in (2.31) and set

$$\begin{aligned}&{\mathcal {G}}_{\mathbf{F }}[\psi ] : = {\mathcal {E}}^{\mathrm {GL}}_{1}\left[\psi , \mathbf{F }; \Gamma _{\beta }(L,\ell )\right]; \end{aligned}$$
(3.37)
$$\begin{aligned}&E_{\beta }(L,\ell ) : = \inf _{\psi \in {\mathscr {D}}_{\star }(\Gamma _{\beta }(L,\ell ))} {\mathcal {G}}_{\mathbf{F }}[\psi ], \end{aligned}$$
(3.38)

where both the energy \( {\mathcal {E}}^{\mathrm {GL}}_{1} \), the minimization domain and the corner region are introduced in § 2. Any corresponding minimizer is denoted by \( \psi _{\beta }\). Before proceeding further, we introduce an auxiliary problem in \( \Gamma _{\beta }(L,\ell )\), modified by the addition of analogous boundary terms as in (3.6). Such a problem will appear in the proof of the main theorem. We set

$$\begin{aligned}&{\widetilde{{\mathcal {G}}}}_{\mathbf{F }}[\psi ] : = {\mathcal {E}}^{\mathrm {GL}}_{1}\left[\psi , \mathbf{F }; \Gamma _{\beta }(L,\ell )\right] - \int _0^{\ell }\mathrm {d}t\, \frac{F_0(t)}{f_0^2(t)} j_t\left[\psi (\mathbf{r }(s,t)) e^{\frac{i}{2} st} \right] \bigg \vert _{s=-L}^{s=L}; \end{aligned}$$
(3.39)
$$\begin{aligned}&{\widetilde{E}}_{\beta }(L,\ell ) : = \inf _{\psi \in {\widetilde{{\mathscr {D}}}}_{\star }(\Gamma _{\beta }(L,\ell ))} {\widetilde{{\mathcal {G}}}}_{{\mathbf{F }}}[\psi ], \end{aligned}$$
(3.40)

where

$$\begin{aligned} {\widetilde{{\mathscr {D}}}}_{\star }(\Gamma _{\beta }(L,\ell )) : = \left\{ \psi \in H^1(\Gamma _{\beta }(L,\ell )) \, \big | \, \left. \psi \right|_{\partial \Gamma _{\mathrm {in}}} = \psi _{\star } \right\} , \end{aligned}$$
(3.41)

and \( \psi _\star \) is defined in (2.16). Note that the boundary terms are slightly different than the ones considered in (3.6), which is due to the presence of an additional phase in \( \psi _{\star } \) compared to \( f_0 e^{-i \alpha _0s} \), due to the different choice of the magnetic potential.

In the next Lemma 3.1, we show that the vector potential \( \mathbf{F }\) can be replaced with \( \mathbf{a }\), such that far from the corners

$$\begin{aligned} \mathbf{a }(\mathbf{r }(s,t)) \simeq - t \mathbf{e }_s, \end{aligned}$$
(3.42)

in boundary coordinates (st) . It is not difficult to figure out that there exists no smooth gauge transformation implementing the above change globally in \( \Gamma _{\beta }(L,\ell )\), in particular close to the bisectrix. More precisely, we define the wedge-domain \( \Gamma _{\beta }(L,\ell )\setminus {\widetilde{\Gamma }}_{\beta }(L,\ell )\) (as depicted in Fig. 4) through

$$\begin{aligned} \Gamma _{\beta }(L,\ell )\setminus {\widetilde{\Gamma }}_{\beta }(L,\ell ): = \left\{ \mathbf{r }\in \Gamma _{\beta }(L,\ell )\, \big | \, \textstyle \frac{1}{2} \beta - \textstyle \frac{1}{\ell ^3} \leqslant \vartheta \leqslant \textstyle \frac{1}{2} \beta + \textstyle \frac{1}{\ell ^3} \right\} , \end{aligned}$$
(3.43)

in polar coordinates \( (\varrho ,\vartheta ) \in [0,\ell ] \times [0,\beta ] \). Hence, we obviously have

$$\begin{aligned} \left| \Gamma _{\beta }(L,\ell )\setminus {\widetilde{\Gamma }}_{\beta }(L,\ell )\right| = {\mathcal {O}}(\ell ^{-1}). \end{aligned}$$
(3.44)

The potential \( \mathbf{a }\) is thus such that there exists a gauge phase \( \phi _{\mathbf{F }} \in H^1(\Gamma _{\beta }(L,\ell )) \) so that

$$\begin{aligned} \mathbf{a }= \mathbf{F }+ \nabla \phi _{\mathbf{F }}, \end{aligned}$$
(3.45)

and

$$\begin{aligned} \mathbf{a }= - t \mathbf{e }_s, \qquad \text{ in } {\widetilde{\Gamma }}_{\beta }(L,\ell ). \end{aligned}$$
(3.46)

As already explained, because of the jump of \( \mathbf{e }_s \) along the bisectrix of the sector, one can not set \( \mathbf{a }= - t \mathbf{e }_s \) everywhere. However, we require that

$$\begin{aligned} \mathbf{a }= {\mathcal {O}}(\ell ^4), \qquad \text{ in } \Gamma _{\beta }(L,\ell ), \end{aligned}$$
(3.47)

which is in fact a constraint only in \( \Gamma _{\beta }(L,\ell )\setminus {\widetilde{\Gamma }}_{\beta }(L,\ell )\). In next Lemma 3.1 we investigate the existence of such a phase \( \phi _{\mathbf{F }} \). Note that

$$\begin{aligned} \mathrm {curl}\, \mathbf{a }= \mathrm {curl}( - t \mathbf{e }_s ) = 1, \qquad \text{ in } {\Gamma _{\beta }(L,\ell )}, \end{aligned}$$
(3.48)

thanks to (1.22) and the gauge invariance of the \( \mathrm {curl}\).

Fig. 4
figure 4

The region \( \Gamma _{\beta }(L,\ell )\setminus {\widetilde{\Gamma }}_{\beta }(L,\ell )\) (shaded area)

Lemma 3.1

(Gauge choice) For any \( L, \ell > 0 \) satisfying (2.30) and so that

$$\begin{aligned} \frac{\left|\Gamma _{\beta }(L,\ell )\right|}{2\pi \left| \partial \Gamma _{\beta }(L,\ell )\right|} \in {\mathbb {Z}}, \end{aligned}$$
(3.49)

there exists a vector potential \( \mathbf{a }\in L^{\infty }(\Gamma _{\beta }(L,\ell )) \) and a phase \( \phi _{\mathbf{F }} \in H^1(\Gamma _{\beta }(L,\ell )) \) satisfying (3.45), (3.46) and (3.47), such that

$$\begin{aligned} \inf _{\psi \in {\mathscr {D}}_{\star }(\Gamma _{\beta }(L,\ell ))} {\mathcal {G}}_{\mathbf{F }}[\psi ]= & {} \inf _{\psi \in {\mathscr {D}}_{\mathrm {D}}(\Gamma _{\beta }(L,\ell ))} {\mathcal {G}}_{\mathbf{a }}[\psi ], \end{aligned}$$
(3.50)
$$\begin{aligned} \inf _{\psi \in {\widetilde{{\mathscr {D}}}}_{\star }(\Gamma _{\beta }(L,\ell ))} {\widetilde{{\mathcal {G}}}}_{{\mathbf{F }}}[\psi ]= & {} \inf _{\psi \in {\mathscr {D}}_{\mathrm {N}}(\Gamma _{\beta }(L,\ell ))} {\widetilde{{\mathcal {G}}}}_{\mathbf{a }}[\psi ], \end{aligned}$$
(3.51)

where

$$\begin{aligned} {\mathscr {D}}_{\mathrm {D}}(\Gamma _{\beta }(L,\ell ))&: =&\left\{ \psi \in H^1(\Gamma _{\beta }(L,\ell )) \, \big | \, \left. \psi \right|_{\partial \Gamma _{\mathrm {bd}}\cup \partial \Gamma _{\mathrm {in}}} = \psi _{0} \right\} , \end{aligned}$$
(3.52)
$$\begin{aligned} {\mathscr {D}}_{\mathrm {N}}(\Gamma _{\beta }(L,\ell ))&: =&\left\{ \psi \in H^1(\Gamma _{\beta }(L,\ell )) \, \big | \, \left. \psi \right|_{\partial \Gamma _{\mathrm {in}}} = \psi _{0} \right\} , \end{aligned}$$
(3.53)
$$\begin{aligned} \psi _0(s,t)&: =&f_0(t) e^{- i \alpha _0 s}. \end{aligned}$$
(3.54)

Remark 3.2

(Constraint on \( L, \ell \)) The condition (3.49) reads \( L - \frac{2 \ell }{\tan \beta } = c(\beta ,L \ell ) {\mathbb {Z}} \) where \( c(\beta ,\ell ) \) is uniformly bounded as \( L,\ell \rightarrow + \infty \). More precisely

$$\begin{aligned} c(\beta ,L,\ell ) \xrightarrow [L \rightarrow +\infty ]{} 0, \qquad c(\beta ,L,\ell ) \xrightarrow [\ell \rightarrow + \infty ]{} c(\beta ), \end{aligned}$$

uniformly in the other parameters. Hence, given generic \( \ell , L \rightarrow + \infty \), it suffices to replace L with \( L + {\mathcal {O}}(1) \) to enforce (3.49).

Proof

The two different minimization problem can be treated in the same way. It suffices to prove the existence of the gauge phase \( \phi _{\mathbf{F }} \) and, in order to recover (3.46), we set

$$\begin{aligned} \phi _{\mathbf{F }}(s,t) : = - \textstyle \frac{1}{2} s t, \qquad \text{ in } {\widetilde{\Gamma }}_{\beta }(L,\ell ). \end{aligned}$$
(3.55)

Note that such a phase is actually the same gauge phase used in [33, Appendix F] or in [24, Eq. (4.7)] with vector potential set equal to \( \mathbf{F }\) and recovers the additional phase factor in the boundary terms in (3.39). Such a phase is in \( H^1({\widetilde{\Gamma }}_{\beta }(L,\ell )) \) but its definition can not be extended to the whole \( \Gamma _{\beta }(L,\ell )\). We can however continue \( \phi _{\mathbf{F }} \) arbitrarily in \( \Gamma _{\beta }(L,\ell )\setminus {\widetilde{\Gamma }}_{\beta }(L,\ell )\), just requiring continuity through the boundary of the region. There are infinitely many ways of doing that and at least one such that the bound (3.47) is satisfied (e.g., a linear interpolation).

In order to complete the proof, we need to show that \( \psi e^{i \phi _\mathbf{F }} \) is still a single-valued function. It is not difficult to see [33, Appendix F] that, to this purpose, one has to correct (3.55) by \( \varpi s \), where

$$\begin{aligned} \varpi&= \frac{1}{2\pi \left| \partial \Gamma _{\beta }(L,\ell )\right|} \int _{\Gamma _{\beta }(L,\ell )} \mathrm {d}\mathbf{r }\, \mathrm {curl}\mathbf{F }- \left\lfloor \frac{1}{2\pi \left| \partial \Gamma _{\beta }(L,\ell )\right|} \int _{\Gamma _{\beta }(L,\ell )} \mathrm {d}\mathbf{r }\, \mathrm {curl}\mathbf{F }\right\rfloor \\&= \frac{\left| \Gamma _{\beta }(L,\ell )\right|}{2\pi \left| \partial \Gamma _{\beta }(L,\ell )\right|} - \left\lfloor \frac{\left| \Gamma _{\beta }(L,\ell )\right|}{2\pi \left| \partial \Gamma _{\beta }(L,\ell )\right|} \right\rfloor , \end{aligned}$$

where \( \left\lfloor \, \cdot \, \right\rfloor \) stands for the integer part. However, by the assumption (3.49), \( \varpi = 0 \) and no additional phase is needed. \(\square \)

From now on we are to going to study only the minimization on the r.h.s. of (3.50) in Lemma 3.1, with the vector potential \( \mathbf{a }\) satisfying (3.45), (3.46) and (3.47). In order to guarantee that (3.49) is satisfied, however, we restrict the analysis to suitable monotone sequences \( \left\rbrace \ell _n \right\lbrace _{n \in {\mathbb {N}}} \), \( \left\rbrace L_n \right\lbrace _{n \in {\mathbb {N}}} \), such that

$$\begin{aligned} \ell _n, L_n \xrightarrow [n \rightarrow +\infty ]{} + \infty , \qquad (2.30) \text{ and } (3.49) \text{ hold }, \end{aligned}$$
(3.56)

and consider \( E_{\mathrm {corner},\beta }(L_n,\ell _n) \) in the following. More precisely, we are going to study the quantity

$$\begin{aligned} E_{\beta }(L_n,\ell _n) : = \inf _{\psi \in {\mathscr {D}}_{\mathrm {D}}(\Gamma _{\beta }(L_n,\ell _n))} {\mathcal {G}}_{\mathbf{a }}[\psi ; \Gamma _{\beta }(L_n,\ell _n))]. \end{aligned}$$
(3.57)

Any minimizer of (3.57) is denoted by \( \psi _n\), i.e.,

$$\begin{aligned} E_{\beta }(L_n,\ell _n) : = {\mathcal {G}}_{\mathbf{a }}\left[\psi _n; \Gamma _{\beta }(L_n,\ell _n))\right]. \end{aligned}$$
(3.58)

The existence of such a minimizer follows by standard arguments as well as the fact that any \( \psi _n\) solves the variational equation

$$\begin{aligned} {\left\{ \begin{array}{ll} - \left( \nabla + i \mathbf{a }\right)^2 \psi _n= \textstyle \frac{1}{b}(1-|\psi _n|^2)\psi _n, &{} \text{ in } \Gamma _{\beta }(L,\ell ), \\ \psi _n= f_0(t(\mathbf{r })) e^{-i \alpha _0 s(\mathbf{r })}, &{} \text{ on } \partial \Gamma _{\mathrm {bd}}\cup \partial \Gamma _{\mathrm {in}}, \\ \mathbf{n }\cdot \left(\nabla + i \mathbf{a }\right) \psi _n= 0, &{} \text{ on } \partial \Gamma _{\mathrm {out}}. \end{array}\right. } \end{aligned}$$
(3.59)

Note that the equation above coincides with (3.5) far from the vertex, where boundary coordinates are well posed and \( \mathbf{a }= - t \mathbf{e }_s \). We can thus apply to \( \psi _n\) the results in Lemmas B.1, B.3, B.5 and B.6.

3.3 Boundedness and existence of the limit

We start by proving the uniform boundedness of \( E_{\mathrm {corner},\beta }(L,\ell )\) as a function of \( \ell , L \).

Proposition 3.4

(Boundedness of \( E_{\mathrm {corner},\beta }(L_n,\ell _n)\)) Let \( \left\rbrace \ell _n \right\lbrace _{n \in {\mathbb {N}}}, \left\rbrace L_n \right\lbrace _{n \in {\mathbb {N}}} \) satisfy (3.56). Then, for any \( 1< b < \Theta _0^{-1} \), there exists a finite constant \( C < + \infty \) independent of n, such that

$$\begin{aligned} \left| E_{\mathrm {corner},\beta }(L_n,\ell _n)\right| \leqslant C. \end{aligned}$$
(3.60)

Proof

We first discuss the boundedness from below, which is the most difficult property to prove, and show that

$$\begin{aligned} E_{\mathrm {corner},\beta }(L_n,\ell _n) \geqslant - C, \end{aligned}$$
(3.61)

for some finite \( 0< C < +\infty \). The key tool is a suitable partition of unity, which isolates the region where we want to retain the energy and allow us to discard the rest. We thus consider two smooth positive functions \( \chi \) and \( \eta \), such that \(\chi ^2 + \eta ^2=1\) and whose supports are described, e.g., in Fig. 5: we assume that \( \eta \equiv 1 \) inside the shaded area, while \( \chi \equiv 1 \) in the white area.

Fig. 5
figure 5

The partition of unity \( \chi \), \(\eta \)

The dashed regions is where the supports of the two functions overlap. We choose the angle \( {\widehat{CVD}} \) equal to \( \beta /2 \) for concreteness but any angle of order 1 would work. The distance of the points A and B from the vertex V is also taken of order 1. Furthermore, the width of the transition regions can also be taken in such a way that

$$\begin{aligned} |\nabla \chi | = {\mathcal {O}}(1), \qquad |\nabla \eta | = {\mathcal {O}}(1). \end{aligned}$$
(3.62)

The rationale behind the choice of the partition of unity is that the energy contribution coming from the support of \( \chi \) reconstructs the leading term \( 2 L E ^{\mathrm{1D}}_{0}(\ell ) \), up to an \( {\mathcal {O}}(1) \) error, while the rest provides a correction of order \( {\mathcal {O}}(1) \). Therefore, the support of \( \chi \) must contain the outer boundary \( \partial \Gamma _{\mathrm {out}}\) up to \( {\mathcal {O}}(1) \) regions and the magnetic potential must be equal to \( - t \mathbf{e }_s \) there. Hence, the area close to the bisectrix is included in the support \( \eta \), because there the magnetic potential is unknown.

The key ingredient of the proof is then the IMS formula [29, Thm. 3.2], which yields

$$\begin{aligned}&E_{\beta }(L_n,\ell _n) = {\mathcal {G}}\left[\chi \psi _n\right] + {\mathcal {G}}\left[\eta \psi _n\right] - \int _{ \Gamma _{\beta }(L,\ell )}\mathrm {d}\mathbf{r }\, |\nabla \chi |^2|\psi _n|^2 -\int _{ \Gamma _{\beta }(L,\ell )} \mathrm {d}\mathbf{r }\, |\nabla \eta |^2|\psi _n|^2 \nonumber \\&\quad = {\mathcal {G}}\left[\chi \psi _n\right] + {\mathcal {G}}\left[\eta \psi _n\right] + {\mathcal {O}}(1), \end{aligned}$$
(3.63)

where we have exploited the decay (B.35) to bound the contributions on the supports of \( \nabla \chi , \nabla \eta \). We now claim that there exists a finite constant independent of n so that

$$\begin{aligned} {\mathcal {G}}_{\mathbf{a }}\left[\eta \psi _n\right]\geqslant & {} -C, \end{aligned}$$
(3.64)
$$\begin{aligned} {\mathcal {G}}_{\mathbf{a }}\left[\chi \psi _n\right]= & {} 2 L_n E ^{\mathrm{1D}}_{0}+ {\mathcal {O}}(1), \end{aligned}$$
(3.65)

which combined with (3.63) yields (3.60).

Let us first consider the first estimate above: dropping from the energy all the positive terms, we get

$$\begin{aligned} {\mathcal {G}}_{\mathbf{a }}\left[\eta \psi _n\right] \geqslant - C \int _{\mathrm {supp}(\eta )} \mathrm {d}\mathbf{r }\, \left| \psi _n(\mathbf{r }) \right|^2 \geqslant - C, \end{aligned}$$
(3.66)

by the decay of \( \psi _n\) as above. To complete the proof it remains only to deal with (3.65): since \( \mathrm {supp}(\chi ) \) is actually composed of two disconnected sets, denoted by \( T_- \) (on the right of Fig. 5) and \( T_+ \), we can use boundary coordinates in both regions \( T_{\pm } \). We can then apply the splitting technique described in the proof of Proposition 3.1 and set

$$\begin{aligned} \chi (s,t) \psi _n(\mathbf{r }(s,t)) = : {\left\{ \begin{array}{ll} f_0(t) e^{-i \alpha _0 s} u_-(s,t), &{} \text{ in } T_-, \\ f_0(t) e^{-i \alpha _0 s} u_+(s,t), &{} \text{ in } T_+. \end{array}\right. } \end{aligned}$$
(3.67)

The same computation which leads to (3.13) yields now (recall (3.14))

$$\begin{aligned} {\mathcal {G}}_{\mathbf{a }}\left[\chi \psi _n\right] = - \frac{1}{b} \int _{T_- \cup T_+} \mathrm {d}s \mathrm {d}t \, f_0^4(t) + {\mathcal {E}}_0[u_-;T_-] + {\mathcal {E}}_0[u_+;T_+]. \end{aligned}$$
(3.68)

Finally, as long as \( 1< b < \Theta _0^{-1} \), one can prove that the energies \( {\mathcal {E}}_0[u_-;T_-] \) and \( {\mathcal {E}}_0[u_+;T_+] \) are both positive, exactly as in (3.17), leading to

$$\begin{aligned} {\mathcal {G}}_{\mathbf{a }}\left[\chi \psi _n\right] \geqslant - \frac{1}{b} \int _{T_- } \mathrm {d}s \mathrm {d}t \, f_0^4(t) - \frac{1}{b} \int _{T_+ } \mathrm {d}s \mathrm {d}t \, f_0^4(t). \end{aligned}$$
(3.69)

The last step is the estimate of the two integrals on the r.h.s. of (3.69) above: the identity (A.9) and the exponential decay (A.10) (both with \( k = 0 \)) imply

$$\begin{aligned} -\frac{1}{b} \int _{T_\pm }\mathrm {d}s\mathrm {d}t \, f_0^4(t) \geqslant L_n E ^{\mathrm{1D}}_{0}- C \int _0^{\ell }\mathrm {d}t \, t e^{-2(t+\alpha _0)^2} + {\mathcal {O}}(1) \geqslant L_n E ^{\mathrm{1D}}_{0}+ {\mathcal {O}}(1), \end{aligned}$$

which together with (3.69) completes the lower bound proof.

The opposite side of the inequality (3.61) can be proven by simply using \( \chi f_0(t) e^{-i \alpha _0 s} \) as a trial state (more precisely, setting \( u_{\pm } = 1 \) in (3.67)). We omit the calculations, since they are totally analogous to the ones above. \(\square \)

We are now in position to prove the first important result of this section.

Proof of Proposition 2.2

The first important observation is that \( E_{\mathrm {corner},\beta }(L,\ell ) \) is a monotone non-increasing function of L and as such it admits a limit. Indeed, for any \( L_a < L_b \), one can easily construct a trial state for the energy in \( \Gamma _{\beta }(L_b, \ell ) \) by extending the minimizer in \( \Gamma _{\beta }(L_a, \ell ) \) and setting the trial state equal to \( f_0(t) e^{-i \alpha _0 s} \) where the minimizer is not defined. The outcome of the trivial computation is the inequality \( E_{\mathrm {corner},\beta }(L_a, \ell ) \leqslant E_{\mathrm {corner},\beta }(L_b,\ell ) \).

Let \( \left\rbrace \ell _n \right\lbrace _{n \in {\mathbb {N}}}, \left\rbrace L_n \right\lbrace _{n \in {\mathbb {N}}} \) be two monotone subsequences such that \( \lim _{n \rightarrow +\infty } \ell _n = \lim _{n \rightarrow +\infty } L_n = + \infty \) and (3.56) is satisfied (see Remark 3.2). By the monotonicity in L of the energy and its boundedness, we know that for any \( \varepsilon > 0 \) and any given \( {\bar{n}} \in {\mathbb {N}}\), there exists \( {\bar{n}}_2({\bar{n}}) \in {\mathbb {N}}\), such that

$$\begin{aligned} \left| E_{\mathrm {corner},\beta }(L_n, \ell _{{\bar{n}}}) - E_{\mathrm {corner},\beta }(L_m, \ell _{{\bar{n}}}) \right| < \textstyle \frac{1}{3} \varepsilon , \end{aligned}$$
(3.70)

for any \( n,m > {\bar{n}}_2 \).

Furthermore, by the exponential decay of the minimizer and its derivatives (B.34), one gets

$$\begin{aligned} \left| E_{\mathrm {corner},\beta }(L_n, \ell _n) - E_{\mathrm {corner},\beta }(L_n, \ell _m) \right| \leqslant C L_n e^{- c \min \left\rbrace \ell _n, \ell _m \right\lbrace }. \end{aligned}$$

Hence, if the sequences satisfy the condition

$$\begin{aligned} L_n \leqslant C \ell _n^a, \qquad \text{ for } \text{ some } a > 0, \end{aligned}$$
(3.71)

we can conclude that there exists \( {\bar{n}}_1 \in {\mathbb {N}}\), such that

$$\begin{aligned} \left| E_{\mathrm {corner},\beta }(L_n, \ell _n) - E_{\mathrm {corner},\beta }(L_n, \ell _m) \right| < \textstyle \frac{1}{3} \varepsilon \end{aligned}$$
(3.72)

for \( n,m > {\bar{n}}_1 \).

In conclusion, we can estimate

$$\begin{aligned}&\left| E_{\mathrm {corner},\beta }(L_n, \ell _n) - E_{\mathrm {corner},\beta }(L_m, \ell _m) \right| \leqslant \left| E_{\mathrm {corner},\beta }(L_n, \ell _n) - E_{\mathrm {corner},\beta }(L_n, \ell _{{\bar{n}}_1+1}) \right| \nonumber \\&\quad + \left| E_{\mathrm {corner},\beta }(L_n, \ell _{{\bar{n}}_1+1}) - E_{\mathrm {corner},\beta }(L_m, \ell _{{\bar{n}}_1+1}) \right| \nonumber \\&\quad + \left| E_{\mathrm {corner},\beta }(L_m, \ell _{{\bar{n}}_1+1}) - E_{\mathrm {corner},\beta }(L_m, \ell _m) \right| < \varepsilon \end{aligned}$$
(3.73)

for any \( n, m > \max \left\rbrace {\bar{n}}_1, {\bar{n}}_2(\bar{n_1}+1) \right\lbrace \), so that the sequence is Cauchy and the limit exists. The independence of the chosen subsequences relies on the uniqueness of the limit, while the uniform boundedness has been proven in Proposition 3.4. \(\square \)

3.4 Neumann and Dirichlet problems in \( \Gamma _{\beta }(L,\ell )\)

We are going to study the Neumann problem (3.40) on the monotone subsequences \( \left\rbrace \ell _n \right\lbrace _{n \in {\mathbb {N}}} \), \( \left\rbrace L_n \right\lbrace _{n \in {\mathbb {N}}} \) introduced in the previous § 3.2, i.e., such that (3.56) holds. Our main goal here is to show that, as in the case of the strip, the Dirichlet and Neumann energies coincide asymptotically as \( n \rightarrow + \infty \). This is going to play a key role in the proof of our main result, since it implies the identity

$$\begin{aligned} E_{\mathrm {corner},\beta }&= \lim _{n \rightarrow +\infty } \left( - 2 L E ^{\mathrm{1D}}_{0}(\ell _n) + E_{\beta }(L_n,\ell _n) \right) \nonumber \\&= \lim _{n \rightarrow +\infty } \left( - 2 L_n E ^{\mathrm{1D}}_{0}(\ell _n) + {\widetilde{E}}_{\beta }(L_n,\ell _n) \right). \end{aligned}$$
(3.74)

Before proving the result we need however a technical lemma on a variational problem with twisted boundary conditions, whose proof is postponed at the end of the section. Let then \( \varkappa \in [0,2\pi ) \) as in Proposition 3.2 and set

$$\begin{aligned}&E_{\beta ,\varkappa }(L,\ell ) : = \inf _{\psi \in {\mathscr {D}}_{\mathrm {D},\varkappa }(\Gamma _{\beta }(L,\ell ))} {\widetilde{{\mathcal {G}}}}_{\mathbf{a }}[\psi ], \end{aligned}$$
(3.75)
$$\begin{aligned}&{\mathscr {D}}_{\mathrm {D},\varkappa }(\Gamma _{\beta }(L,\ell )) : = \left\{ \psi \in H^1(\Gamma _{\beta }(L,\ell )) \, \Big | \, \left. \psi \right|_{\partial \Gamma _{\mathrm {in}}\cup \{ s = - L \}} = \psi _{0}, \left. \psi \right|_{s = L} = \psi _0 e^{i \varkappa } \right\} .\nonumber \\ \end{aligned}$$
(3.76)

Lemma 3.2

Let \( \left\rbrace \ell _n \right\lbrace _{n \in {\mathbb {N}}} \), \( \left\rbrace L_n \right\lbrace _{n \in {\mathbb {N}}} \) be two monotone subsequences such that (3.56) holds. Then,

$$\begin{aligned} E_{\beta }(L_n,\ell _n) = E_{\beta ,\varkappa }(L_n,\ell _n) + o_n(1). \end{aligned}$$
(3.77)

Proposition 3.5

(Dirichlet and Neumann energies) Let \( \left\rbrace \ell _n \right\lbrace _{n \in {\mathbb {N}}} \), \( \left\rbrace L_n \right\lbrace _{n \in {\mathbb {N}}} \) be two monotone subsequences such that (3.56) holds. Then, for any \( 1< b < \Theta _0^{-1} \),

$$\begin{aligned} E_{\beta }(L_n,\ell _n) - {\widetilde{E}}_{\beta }(L_n,\ell _n) = o_{n}(1). \end{aligned}$$
(3.78)

Proof

In view of the vanishing of the boundary terms in the functional \( {\widetilde{{\mathcal {G}}}}_{\mathbf{a }}[\psi ] \) on any \( \psi \) belonging to \( {\mathscr {D}}_{\mathrm {D}}(\Gamma _{\beta }(L_n,\ell _n)) \) (see also the proof of Proposition 3.1) and the trivial inclusion \( {\mathscr {D}}_{\mathrm {D}}(\Gamma _{\beta }(L_n,\ell _n)) \subset {\mathscr {D}}_{\mathrm {N}}(\Gamma _{\beta }(L_n,\ell _n)) \), we deduce the inequality

$$\begin{aligned} {\widetilde{E}}_{\beta }(L_n,\ell _n) \leqslant E_{\beta }(L_n,\ell _n). \end{aligned}$$
(3.79)

Hence, we only have to prove the opposite inequality, i.e.,

$$\begin{aligned} E_\beta (L_n,\ell _n) \leqslant {\widetilde{E}}_\beta (L_n,\ell _n) + o_{n}(1). \end{aligned}$$
(3.80)

Preliminarily, we observe that the quantity \( {\widetilde{E}}_{\mathrm {corner},\beta }(L_n,\ell _n) \) admits a limit, which is independent of the chosen sequences, exactly as \( E_{\mathrm {corner},\beta }(L_n,\ell _n) \). The argument to prove it is the same as in the proof of Proposition 2.2; therefore we spell in detail only the estimates showing that \( {\widetilde{E}}_{\mathrm {corner},\beta }(L,\ell ) \) is monotone in L for fixed \( \ell \), up to an exponentially small error term: let \( L_a < L_b \), then we have

$$\begin{aligned} {\widetilde{E}}_{\beta }(L_b,\ell ) = {\widetilde{{\mathcal {G}}}}_{\mathbf{a }}\left[ {\widetilde{\psi }}_{L_b,\ell }; \Gamma _{\beta }(L_a,\ell ) \right] + {\widetilde{{\mathcal {G}}}}_{\mathbf{a }}\left[ {\widetilde{\psi }}_{L_b,\ell }; R_{\pm } \right] \geqslant {\widetilde{E}}_{\beta }(L_a,\ell ) + {\widetilde{{\mathcal {G}}}}_{\mathbf{a }}\left[ {\widetilde{\psi }}_{L_b,\ell }; R_{\pm } \right]{,}\nonumber \\ \end{aligned}$$
(3.81)

where \( R_\pm \) are the rectangular regions \( [L_a, L_b] \times [0, \ell ] \) and \( [-L_b, -L_a] \times [ 0, \ell ] \), respectively. Applying Proposition 3.1, we get

$$\begin{aligned} {\widetilde{{\mathcal {G}}}}_{\mathbf{a }}\left[ {\widetilde{\psi }}_{L_b,\ell }; R_{\pm }(L_b-L_a,\ell ) \right] \geqslant (L_b - L_a) E ^{\mathrm{1D}}_{0}(\ell ) + {\mathcal {O}}\left((L_b-L_a) \ell ^{-\infty }\right), \end{aligned}$$

which, plugged into (3.81), yields

$$\begin{aligned} {\widetilde{E}}_{\mathrm {corner},\beta }(L_b,\ell ) \geqslant {\widetilde{E}}_{\mathrm {corner},\beta }(L_a,\ell ) + {\mathcal {O}}\left((L_b-L_a) \ell ^{-\infty }\right). \end{aligned}$$
(3.82)

Let \( \left\rbrace \delta _n \right\lbrace _{n \in {\mathbb {N}}} \) be such that \( 0 \leqslant \delta _n \leqslant 1 \) and the pair of sequences \( \left\rbrace L_n - \delta _n \right\lbrace _{n \in {\mathbb {N}}} \), \( \left\rbrace \ell '_n \right\lbrace _{n \in {\mathbb {N}}} \) satisfies the same conditions (3.56) as \( \left\rbrace L_n \right\lbrace \), \( \left\rbrace \ell _n \right\lbrace \). Note that we have also \( \ell '_n = \ell _n + {\mathcal {O}}(\delta _n) \), because of (3.49) (see also Remark 3.2). We denote by \( {\widetilde{\psi }}_n \) and \( {\widetilde{\psi }}_{n,\delta _n} \) for short any energy minimizer in \( \Gamma _{\beta }(L_n,\ell _n)\) and \( \Gamma _{\beta }(L_n-\delta _n,\ell '_n) \), respectively. The splitting technique used to derive (3.13), yields (recall (3.12), (3.22) and (3.67))

$$\begin{aligned}&{\widetilde{E}}_{\beta }(L_n, \ell _n) \nonumber \\&= {\widetilde{{\mathcal {G}}}}_{\mathbf{a }} \big [{\widetilde{\psi }}_{n}; \Gamma _{\beta }(L_n-\delta _n,\ell _n^{\prime }) \big ] + 2 E ^{\mathrm{1D}}_{0}(\ell ^{\prime }_n) \delta _n + {\widetilde{{\mathcal {E}}}}_0\left[{u_-}; R_-\right] + {\widetilde{{\mathcal {E}}}}_0\left[{u_+}; R_+\right] + {\mathcal {O}}(\ell _n^{-\infty }) \nonumber \\&\geqslant {\widetilde{E}}_{\beta }(L_n-\delta _n,\ell ^{\prime }_n) + 2 E ^{\mathrm{1D}}_{0}(\ell ^{\prime }_n) \delta _n + {\widetilde{{\mathcal {E}}}}_0\left[{u_-}; R_-\right] + {\widetilde{{\mathcal {E}}}}_0\left[{u_+}; R_+\right] + {\mathcal {O}}(\ell _n^{-\infty }), \end{aligned}$$
(3.83)

where \( R_- = [-L_n,-L_n+\delta _n]\times [0,\ell _n^{\prime }] \) and \( R_+ = [L_n-\delta _n,L_n]\times [0,\ell _n^{\prime }] \) and \( u_{\pm } \) are defined as in (3.12). Hence, we get that (recall that \( {\widetilde{{\mathcal {E}}}}_0[u] \geqslant {\mathcal {O}}(\ell ^{-\infty })\) if \( 1< b < \Theta _0^{-1} \))

$$\begin{aligned} {\widetilde{{\mathcal {E}}}}_0\left[{u_{\pm }}; R_{\pm } \right] \leqslant {\widetilde{E}}_{\mathrm {corner},\beta }(L_n,\ell _n) - {\widetilde{E}}_{\mathrm {corner},\beta }(L_n - \delta _n,\ell '_n) + {\mathcal {O}}(\ell _n^{-\infty }) = : {\mathfrak {e}}_n = o_n(1),\nonumber \\ \end{aligned}$$
(3.84)

for any \( \delta _n \leqslant 1 \), since the two quantities \( {\widetilde{E}}_{\mathrm {corner},\beta }(L_n,\ell _n) \), \( {\widetilde{E}}_{\mathrm {corner},\beta }(L_n - \delta _n,\ell '_n) \) admit the same limit, as proven above.

Now, we claim that (3.84) implies that, up to a phase, \( {\widetilde{\psi }}_{n} \) is pointwise close to \( f_0(t) e^{-i\alpha _0 s} \) in the region \( R_- \cup R_+ \) and, in particular, along the boundary \( \partial \Gamma _{\mathrm {bd}}\). Indeed, applying Proposition 3.3 to the functionals \( {\widetilde{{\mathcal {E}}}}_0\left[u_{\pm }; R_{\pm }\right] \) (with \( \delta _n \) in place of L), we get that

$$\begin{aligned} \int _{- L_n}^{-L_n + \delta _n} \mathrm {d}s \int _{0}^{{\bar{\ell }}_n} \mathrm {d}t \, f_0^4 \left| \nabla u_- \right|^2 + \int _{L_n - \delta _n}^{L_n} \mathrm {d}s \int _{0}^{{\bar{\ell }}_n} \mathrm {d}t \, f_0^4 \left| \nabla u_+ \right|^2 \leqslant C {\mathfrak {e}}_n + {\mathcal {O}}(\ell _n^{-\infty }).\nonumber \\ \end{aligned}$$
(3.85)

Furthermore, fixing some \( 0 < T_n \leqslant {{\bar{\ell }}}_n \), then, for any \( t \in [0,T_n] \) and any \(s \in [L_n- \delta _n, L_n] \) or \( s \in [ -L_n,-L_n+\delta _n] \),

$$\begin{aligned}&\left| \big |{\widetilde{\psi }}_{n} (\mathbf{r }(s,t))\big | - f_0(t) \right| \leqslant \frac{C{\mathfrak {e}}_n^{1/4} + {\mathcal {O}}(\ell _n^{-\infty })}{\sqrt{\min _{[0,T_n]} f_0}}, \end{aligned}$$
(3.86)
$$\begin{aligned}&\left. \left| \partial _s \int _0^{\ell _n} \mathrm {d}t \, \big | {\widetilde{\psi }}_{n} \big |^2 \right| \right|_{s = \pm L_n} \leqslant C \left\{ \sqrt{{\mathfrak {e}}_n} + \frac{1}{\delta _n} \left[ \frac{{{\mathfrak {e}}_n}^{1/4} + {\mathcal {O}}(\delta _n \ell _n^{-\infty })}{\sqrt{\min _{[0,T_n]} f_0}} + e^{-c(b)T_n} \right] + \delta _n \right\} .\nonumber \\ \end{aligned}$$
(3.87)

In order to simplify the discussion, let us assume that the errors \( {\mathcal {O}}(\ell _n^{-\infty })\) appearing on the r.h.s. of (3.85) and (3.86) are much smaller than \( {\mathfrak {e}}_n \), since, if this is not the case, i.e., \( {\mathfrak {e}}_n \) is exponentially small in \( \ell _n \), then the argument is actually much simpler. Then, if we pick \( T_n \) in such a way that

$$\begin{aligned} f_0(T_n) = {\mathfrak {e}}_n^{1/12} = o_n(1), \end{aligned}$$
(3.88)

if the r.h.s. is larger than \( f_0({{\bar{\ell }}}_n) \), or \( T_n = {{\bar{\ell }}}_n \) otherwise, then

$$\begin{aligned}&\left\Vert \nabla u_{\pm } \right\Vert_{L^2({\widetilde{R}}_{\pm })}^2 \leqslant C {\mathfrak {e}}_n^{2/3} = o_n(1), \end{aligned}$$
(3.89)
$$\begin{aligned}&\left\Vert 1 - |u_{\pm }| \right\Vert_{L^{\infty }({\widetilde{R}}_{\pm })} \leqslant C {\mathfrak {e}}_n^{1/8} = o_n(1), \qquad \big \Vert {\widetilde{\psi }}_n \big \Vert _{L^{\infty }(R_{\pm } \setminus {\widetilde{R}}_{\pm })} \leqslant C e^{-\frac{1}{2} c(b) T_n}, \end{aligned}$$
(3.90)

where \( {\widetilde{R}}_+ : = [L_n - \delta _n, L_n] \times [0, T_n] \) and we used (B.35). Note that, by the pointwise lower bound on \( f_0 \) stated in (A.10), we find that

$$\begin{aligned} C \sqrt{\left| \log {\mathfrak {e}}_n \right|} \geqslant T_n \geqslant 2 \sqrt{\left| \log {\mathfrak {e}}_n \right|} (1 + o_n(1)) \gg 1. \end{aligned}$$
(3.91)

Now, we claim that (3.89) and (3.90) imply that \( u_{\pm } \) is close in \( L^2\) sense to a constant phase factor \( e^{i \varkappa _{\pm }} \), \( \varkappa _{\pm } \in {\mathbb {R}}\), or, equivalently, \( {\widetilde{\psi }}_n \simeq f_0(t) e^{-i(\alpha _0s - \varkappa _{\pm })} \) in \( {\widetilde{R}}^{\pm } \). By applying the Poincaré inequality

$$\begin{aligned} \int _{{\widetilde{R}}_+} \mathrm {d}s \mathrm {d}t \, \left| h - \langle h \rangle \right|^2 \leqslant C \int _{{\widetilde{R}}_+} \mathrm {d}s \mathrm {d}t \, \left\{ T_n^2 \left| \partial _t h \right|^2 + \delta _n^2 \left| \partial _s h \right|^2 \right\} , \end{aligned}$$

where \( \langle h \rangle \) is the average of h over \( {\widetilde{R}}_+ \), to \( h = u_+/|u_+| \), which is well posed since \( u_+ \) does not vanish in \( {\widetilde{R}}_+ \) by (3.90), we obtain that there exists \( \varkappa _+ \in [0,2\pi ) \) such that

$$\begin{aligned} \left\Vert \frac{u_+}{|u_+|} - e^{i \varkappa _+} \right\Vert_{L^2({\widetilde{R}}_+)}^2 \leqslant C T_n^2 {\mathfrak {e}}_n^{2/3} = o_n(1), \end{aligned}$$

thanks to (3.91). This in turn yields the desired estimate via (3.90):

$$\begin{aligned} \left\Vert u_+ - e^{i \varkappa _+} \right\Vert_{L^2({\widetilde{R}}_+)}^2 \leqslant C \left[ T_n^2 {\mathfrak {e}}_n^{2/3} + \delta _n T_n {\mathfrak {e}}_n^{1/4} \right] = o_n(1). \end{aligned}$$
(3.92)

The idea is now to exploit the information collected above to construct a trial state and prove an upper bound on \( E_{\beta }(L_n,\ell _n) \) in terms of \( {\widetilde{E}}_{\beta }(L_n-\delta _n,\ell _n) \) via Lemma 3.2: we set \( \psi _{\mathrm {trial}}(\mathbf{r }) := {\widetilde{\psi }}_n(\mathbf{r }) \) close to the corner, while sufficiently far from it,

$$\begin{aligned} \psi _{\mathrm {trial}}(\mathbf{r }) : = {\widetilde{\psi }}_n(\mathbf{r }) + \eta (s(\mathbf{r })) \left( f_0(t(\mathbf{r })) e^{-i\alpha _0 s(\mathbf{r })} e^{i\varkappa _{\pm }} - {\widetilde{\psi }}_n(\mathbf{r }) \right), \end{aligned}$$
(3.93)

where the phases \( \varkappa _{\pm } \) are the constants appearing in (3.92). The function \( \eta \) is smooth and satisfies \( \eta (\pm L_n) = 1 \), \( \mathrm {supp}(\eta ) \subset [-L_n, -L_n + \delta _n] \cup [L_n - \delta _n, L_n] \times [0, \ell _n] \) and \( |\nabla \eta | = {\mathcal {O}}(\delta _n^{-1}) \). Obviously, the trial state \( \psi _{\mathrm {trial}}\) does not belong to \( {\mathscr {D}}_{\mathrm {D}}\) but \( e^{- i \varkappa _-} \psi _{\mathrm {trial}}\in {\mathscr {D}}_{\mathrm {D},\varkappa }\) (recall (3.76)) with \( \varkappa = \varkappa _+ - \varkappa _- \). Hence, using Lemma 3.2, we can estimate

$$\begin{aligned}&E_{\beta }(L_n,\ell _n) \leqslant E_{\beta , \varkappa _+ - \varkappa _-} (L_n,\ell _n) + o_n(1) \leqslant {\mathcal {G}}_{\mathbf{a }}[\psi _{\mathrm {trial}};\Gamma _{\beta }(L_n,\ell _n)]+ o_n(1) \nonumber \\&\quad \leqslant {\widetilde{E}}_{\beta }(L_n-\delta _n,\ell _n) +\left. \int _0^{\ell _n}\mathrm {d}t\, \frac{F_0(t)}{f_0^2(t)} j_t \big [{\widetilde{\psi }}_{n} \big ] \right|_{s=-L_n-\delta _n}^{s=L_n-\delta _n} + {\mathcal {G}}_{\mathbf{a }}\left[\psi _{\mathrm {trial}}; R_+ \cup R_-\right]+ o_n(1). \end{aligned}$$
(3.94)

Let us first consider the boundary terms at \( L_n \), since the ones at \( - L_n \) are perfectly equivalent: thanks to the boundary conditions (3.10) satisfied by \( {\widetilde{\psi }}_{n} \), we get

$$\begin{aligned} \left. \int _0^{\ell _n}\mathrm {d}t\, \frac{F_0(t)}{f_0^2(t)} j_t \big [{\widetilde{\psi }}_{n} \big ] \right|_{s=L_n-\delta _n} = - \left. \frac{1}{2} \int _0^{\ell _n}\mathrm {d}t\, \partial _s \big | {\widetilde{\psi }}_{n} \big |^2 \right|_{s=L_n} + \int _{L_n - \delta _n}^{L_n} \int _0^{\ell _n}\mathrm {d}t\, \frac{F_0(t)}{f_0^2(t)} \partial _s j_t \big [{\widetilde{\psi }}_{n} \big ]. \end{aligned}$$

Integrating by parts as in (3.16) and using the Agmon bound provided by Lemma B.5 as well as the inequalities (A.22) and (A.25) , one can show that the second term on the r.h.s. of the expression above is bounded by

$$\begin{aligned}&\left| \int _{L_n - \delta _n}^{L_n} \int _0^{\ell _n}\mathrm {d}t\, \frac{F_0(t)}{f_0^2(t)} \partial _s j_t \big [{\widetilde{\psi }}_{n} \big ] \right| \nonumber \\&\quad \leqslant 2 \int _{L_n - \delta _n}^{L_n} \int _0^{T_n}\mathrm {d}t\, \big | \partial _t {\widetilde{\psi }}_n \big | \big | \partial _s {\widetilde{\psi }}_n \big | + 2 e^{-c(b) T_n} \int _{R_+ \setminus {\widetilde{R}}_+} \mathrm {d}s \mathrm {d}t \, e^{c(b) t} \big | \nabla {\widetilde{\psi }}_n \big |^2 \nonumber \\&\quad \leqslant C \left[ \left\Vert f_0 \nabla u \right\Vert_{L^2({\widetilde{R}}_+)}^2 + e^{-c(b) T_n} \right] \leqslant C \left( {\mathfrak {e}}_n^{5/6} + e^{-c(b) T_n} \right), \end{aligned}$$
(3.95)

while the first one can be estimated via (3.87), so obtaining

$$\begin{aligned} \left| \left. \int _0^{\ell _n}\mathrm {d}t\, \frac{F_0(t)}{f_0^2(t)} j_t \big [{\widetilde{\psi }}_{n} \big ] \right|_{s=L_n-\delta _n} \right| \leqslant C \left[ \sqrt{{\mathfrak {e}}_n} + \frac{{{\mathfrak {e}}_n}^{5/24}+ e^{-c(b)T_n}}{\delta _n} + \delta _n \right]. \end{aligned}$$
(3.96)

We now focus on the energy contributions of the regions \( R_{\pm } \) (third term on the r.h.s. of (3.94)): For simplicity, we are going to consider only the energy in the region \( R_+ \), since the corresponding one in \( R_- \) can be bounded in the very same way. We have

$$\begin{aligned}&{\mathcal {G}}_{\mathbf{a }}\left[\psi _{\mathrm {trial}}; R_+ \right] \leqslant 10 \left\Vert \left( \nabla - i t \mathbf{e }_s \right) {\widetilde{\psi }}_n \right\Vert_{L^2(R_+)}^2 + 4 \left\Vert \nabla \eta \left(f_0 e^{-i\alpha _0 s + i\varkappa } - {\widetilde{\psi }}_n \right) \right\Vert_{L^2(R_+)}^2 \nonumber \\&\quad + 8 \left\Vert \left( \nabla - i t \mathbf{e }_s \right) f_0 e^{-i\alpha _0 s + i \varkappa } \right\Vert_{L^2(R_+)}^2 + \frac{1}{b} \left( \big \Vert {\widetilde{\psi }}\big \Vert _{L^4(R_+)}^4 + \left\Vert f_0 \right\Vert_{L^4(R_+)}^4 \right) \nonumber \\&\quad \leqslant \frac{C}{\delta _n^{2}} \left\Vert f_0 e^{-i\alpha _0 s + i\varkappa } - {\widetilde{\psi }}_n \right\Vert_{L^2(R_+)}^2 + C \left( {\mathfrak {e}}_n^{5/6} + e^{-c(b) T_n} \right) + {\mathcal {O}}(\delta _n) \end{aligned}$$
(3.97)

thanks to (A.10) and (B.34) and where the first term on the r.h.s. has been bounded by Cauchy inequality, exploiting (3.85), (B.34), (B.35) and the splitting technique:

$$\begin{aligned}&\left\Vert \left( \nabla - i t \mathbf{e }_s \right) {\widetilde{\psi }}_n \right\Vert_{L^2(R_+)}^2 \leqslant \int _{R_+} \mathrm {d}s \mathrm {d}t \, f_0^2 \left\{ \left| \nabla u_+ \right|^2 - 2(t+\alpha _0) j_s[u_+] + \textstyle \frac{1}{b} \left(1 - f_0^2 \right) |u_+|^2 \right\} \nonumber \\&\quad \leqslant C \left[ \left\Vert f_0 \nabla u \right\Vert_{L^2({\widetilde{R}}_+)}^2 + e^{-c(b) T_n} \right] + {\mathcal {O}}(\delta _n) \leqslant C \left( {\mathfrak {e}}_n^{5/6} + e^{-c(b) T_n} \right) + {\mathcal {O}}(\delta _n). \end{aligned}$$
(3.98)

We now exploit (3.90) and (3.92) to deduce that

$$\begin{aligned}&\left\Vert f_0 e^{-i\alpha _0 s + i\varkappa } - {\widetilde{\psi }}_n \right\Vert_{L^2(R_+)}^2 \leqslant \int _{{\widetilde{R}}_+} \mathrm {d}s \mathrm {d}t \, f_0^2 \left| u - e^{i \varkappa } \right|^2 + C \delta _n e^{- \frac{1}{2} c(b) T_n} \nonumber \\&\quad \leqslant C \left[ T_n^2 {\mathfrak {e}}_n^{2/3} + \delta _n T_n {\mathfrak {e}}_n^{1/4} + e^{- \frac{1}{2} c(b) T_n }\right]. \end{aligned}$$
(3.99)

Putting together (3.94) with (3.96), (3.97) and (3.99), we finally get

$$\begin{aligned}&E_{\beta }(L_n,\ell _n) \leqslant {\widetilde{E}}_{\beta }(L_n-\delta _n,\ell _n) \nonumber \\&\quad +C \left\{ \sqrt{{\mathfrak {e}}_n} + \frac{{{\mathfrak {e}}_n}^{5/24}+ e^{-c(b)T_n}}{\delta _n} + \frac{T_n^2 {\mathfrak {e}}_n^{2/3} + \delta _n T_n {\mathfrak {e}}_n^{1/4} + e^{- \frac{1}{2} c(b) T_n }}{\delta _n^2} + \delta _n \right\} \nonumber \\&\quad \leqslant {\widetilde{E}}_{\beta }(L_n-\delta _n,\ell _n) + C \left[ \frac{T_n^2 {\mathfrak {e}}_n^{2/3} + \delta _n T_n {\mathfrak {e}}_n^{1/4} + e^{- \frac{1}{2} c(b) T_n }}{\delta _n^2} + \delta _n \right] + o_n(1) \nonumber \\&\quad \leqslant {\widetilde{E}}_{\beta }(L_n-\delta _n,\ell _n) + C \left[ \max \left\{ T_n^{2/3} {\mathfrak {e}}_n^{2/9}, T_n^2 {\mathfrak {e}}_n^{1/8} \right\} + e^{- \frac{1}{10} c(b) T_n } \right] + o_n(1) \nonumber \\&\quad = {\widetilde{E}}_{\beta }(L_n,\ell _n) + o_n(1), \end{aligned}$$
(3.100)

where we have optimized over \(\delta _n \) by taking \( \delta _n = \max \{ T_n^{2/3} {\mathfrak {e}}_n^{2/9}, T_n^2 {\mathfrak {e}}_n^{1/8} \} + e^{- \frac{1}{5} c(b) T_n} \) and used that \( T_n = {\mathcal {O}}(\sqrt{|\log {\mathfrak {e}}_n|}) \). \(\square \)

Proof of Lemma 3.2

We first observe that the existence of the limit as \( n \rightarrow + \infty \) of \( E_{\beta ,\varkappa }(L_n,\ell _n) - 2 E ^{\mathrm{1D}}_{0}(\ell _n) L_n \) can be shown as in the proof of Proposition 2.2. Hence for any \( 1 \ll \delta _n \ll \min \{\ell _n, L_n\} \), we have

$$\begin{aligned} E_{\beta ,\varkappa }(L_n,\ell _n) - E_{\beta ,\varkappa }(L_n-\delta _n,\ell _n) + 2 E ^{\mathrm{1D}}_{0}(\ell _n) \delta _n = o_n(1). \end{aligned}$$

By a trivial testing of the functional, exploiting the above estimate as well as Propositions 3.1 and 3.2, one gets

$$\begin{aligned}&E_{\beta }(L_n,\ell _n) \leqslant E_{\beta ,\varkappa }(L_n-\delta _n,\ell _n) + E_{\mathrm {D},\varkappa }(R(\delta _n,\ell _n)) \nonumber \\&\quad \leqslant E_{\beta ,\varkappa }(L_n-\delta _n,\ell _n) + 2 E ^{\mathrm{1D}}_{0}(\ell _n) \delta _n + {\mathcal {O}}(\delta _n^{-1}) + o_n(1) = E_{\beta ,\varkappa }(L_n,\ell _n) + o_n(1). \end{aligned}$$
(3.101)

The proof of the opposite inequality is identical. \(\square \)

4 Proof of the energy lower bound

In this Section we prove the lower bound to the GL energy which in combination with the upper bound proven in Proposition 5.1, stated in next Section, will provide the proof of Theorem 2.1.

Proposition 4.1

(GL energy lower bound) Let \(\Omega \subset {\mathbb {R}}^2\) be any bounded simply connected domain satisfying Assumptions 1 and 2. Then, for any fixed \( 1< b < \Theta _0^{-1} \), as \(\varepsilon \rightarrow 0\), it holds

$$\begin{aligned} E^{\mathrm {GL}}\geqslant \displaystyle \frac{|\partial \Omega | E ^{\mathrm{1D}}_{0}}{\varepsilon } - E_{\mathrm {corr}}\int _{0}^{|\partial \Omega |} \mathrm {d}{\mathsf {s}}\, {\mathfrak {K}}({\mathsf {s}})+ \sum _{j = 1}^N E_{\mathrm {corner}, \beta _j} + o(1). \end{aligned}$$
(4.1)

We recall the definition of the superconducting boundary layer

$$\begin{aligned} {\mathcal {A}}_{\varepsilon }: = \left\{ \mathbf{r }\in \Omega \, \big | \, \mathrm {dist}\left(\mathbf{r }, \partial \Omega \right) \leqslant \varepsilon \ell _{\varepsilon }\right\} , \end{aligned}$$

with (see (1.15)) \( \ell _{\varepsilon }= c_1 |\log \varepsilon | \), for a large constant \( c_1 \). The smooth part of the boundary layer is defined as

$$\begin{aligned} {\mathcal {A}}_{\mathrm {cut},\varepsilon }: = \left\{ \mathbf{r }\in {\mathcal {A}}_{\varepsilon }\, \big | \left| s(\mathbf{r }) - {\mathsf {s}}_j \right| \geqslant \varepsilon L_{\varepsilon }\right\} , \end{aligned}$$
(4.2)

where \( {\mathsf {s}}_j \) is the coordinate along \( \partial \Omega \) of the j-th corner and

$$\begin{aligned} L_{\varepsilon }= c_2(\varepsilon ) |\log \varepsilon |, \end{aligned}$$
(4.3)

for some

$$\begin{aligned} \textstyle \frac{c_1}{\tan \left(\beta /2\right)} \leqslant c_2(\varepsilon ) \leqslant C , \end{aligned}$$
(4.4)

so that (3.56) holds. The corner regions are denoted by \( \Gamma _{j,\varepsilon }\), \( j \in \left\rbrace 1, \ldots , N \right\lbrace \), and coincide with the complement of \( {\mathcal {A}}_{\mathrm {cut},\varepsilon }\):

$$\begin{aligned} \Gamma _{j,\varepsilon }: = \left( {\mathcal {A}}_{\varepsilon }\setminus {\mathcal {A}}_{\mathrm {cut},\varepsilon }\right) \cap \left\{ \mathbf{r }\in {\mathcal {A}}_{\varepsilon }\, | \, \mathrm {dist}(\mathbf{r }, \mathbf{r }_j) \leqslant C \varepsilon L_{\varepsilon }\right\} . \end{aligned}$$
(4.5)

In \( {\mathcal {A}}_{\mathrm {cut},\varepsilon }\), one can use the tubular coordinates \( ({\mathsf {s}}, {\mathsf {t}}) \) defined in (1.10) as well as their rescaled counterparts given in (1.16). We denote by \( {\mathcal {A}}\) the rescaling of the boundary layer \( {\mathcal {A}}_{\varepsilon }\). Similarly, the set obtained via rescaling of the domain \( {\mathcal {A}}_{\mathrm {cut},\varepsilon }\) is denoted by \( {\mathcal {A}}_{\mathrm {cut}}\), i.e., with a little abuse of notation,

$$\begin{aligned} {\mathcal {A}}_{\mathrm {cut}}: = \left( \left[0, s_{1} - L_{\varepsilon }\right] \cup \left[ s_1 + L_{\varepsilon }, s_2 - L_{\varepsilon }\right] \cup \cdots \cup \left[ s_N + L_{\varepsilon }, \textstyle \frac{|\partial \Omega |}{\varepsilon } \right] \right) \times [0, c_1 |\log \varepsilon |],\nonumber \\ \end{aligned}$$
(4.6)

while \( \Gamma _j \) stands for the rescaling of the domain \( \Gamma _{j,\varepsilon }\), i.e., \( \Gamma _j : =\left\rbrace \mathbf{r }' \in {\mathbb {R}}^2 \, \big | \, \mathbf{r }_j + \varepsilon \mathbf{r }' \in \Gamma _{j,\varepsilon }\right\lbrace \).

Fig. 6
figure 6

A typical corner region \( \Gamma _{j,\varepsilon }\) (or, after rescaling, \( \Gamma _j \)) before the rectification

Before proceeding further, we summarize the main steps of the proof of the lower bound. We are going to treat the smooth part of the layer and the corner regions differently. In order to extract the \( {\mathcal {O}}(1) \) contributions to the energy, it is indeed necessary to retain in the smooth part of the layer the terms depending on the boundary curvature. The same precision is not needed close to the corners. There, however, the procedure is more involved, since we have to reconstruct the model problem discussed in § 2.3.

  • The first step is the replacement of the magnetic vector potential (§ 4.1). The idea is to replace \(\mathbf{A }^{\mathrm {GL}}\) with \(-{\mathsf {t}}\mathbf{e }_{{\mathsf {s}}} {+ \frac{1}{2} {\mathfrak {K}}({\mathsf {s}}){\mathsf {t}}^2 + \varepsilon \delta _{\varepsilon }} \) far enough from the corners by means of a suitable gauge change. Close to the corners, on the other hand, we replace \( \mathbf{A }^{\mathrm {GL}}\) with \( \mathbf{F }\) (Lemma 4.3) by means of a priori bounds of the difference between \( \mathbf{A }^{\mathrm {GL}}\) and \( \mathbf{F }\) (see “Appendix B.2”);

  • The second step is the rectification of the corner regions (§ 4.2): via a suitable diffeomorphism, we map the corner region as in Fig. 6 onto a domain with the same shape as \( \Gamma _{\beta }(L,\ell )\) in Fig. 2; this allows us to reduce the lower bound to the corner effective problem introduced in (2.12);

  • The third step is simply the completion of the lower bound (§ 4.3), where we just glue together the lower bounds near the corners with the one in the smooth part of the domain discussed in “Appendix C”.

4.1 Replacement of the magnetic field

In \( {\mathcal {A}}_{\mathrm {cut},\varepsilon }\) we aim at bounding from below the GL energy by the reduced energy functional \( {\mathcal {G}}_{\varepsilon }[\psi ; {\mathcal {A}}_{\mathrm {cut}}] \), where

$$\begin{aligned} {\mathcal {G}}_{\varepsilon }[\psi ; {\mathcal {A}}_{\mathrm {cut}}]&: = \int _{{\mathcal {A}}_{\mathrm {cut}}} \mathrm {d}s \mathrm {d}t \, \left(1 - \varepsilon k(s) t \right) \left\{ \left| \partial _t \psi \right|^2 + \textstyle \frac{1}{(1 - \varepsilon k(s) t)^2} \left| \left( \partial _s + i a_{\varepsilon }(s,t) \right) \psi \right|^2 \right. \nonumber \\&\quad \left. - \textstyle \frac{1}{2 b} \left( 2 |\psi |^2 - |\psi |^4 \right) \right\} . \end{aligned}$$
(4.7)
$$\begin{aligned} \mathbf{a }_{\varepsilon }(s,t)&: = a_{\varepsilon }(s,t) \mathbf{e }_{s}, \qquad a_{\varepsilon }(s,t) = - t + \textstyle \frac{1}{2} k(s) t^2 + \varepsilon \delta _{\varepsilon }, \end{aligned}$$
(4.8)

and \( \psi (s,t) = \psi ^{\mathrm {GL}}(\mathbf{r }(s,t)) e^{-i\phi _\varepsilon (\mathbf{r }(s,t))} \), with \( \phi _{\varepsilon } \) a suitable gauge phase (see (4.10) below).

The replacement procedure by means of a local gauge choice is well described in [33, Appendix F] for smooth domains and, in more details, in [24, Sect. 5.1]. A similar discussion is extended in presence of corners at the boundary in [19, Sect. 2.4], where however the energy of the corner regions is dropped.

Lemma 4.1

(Replacement of the magnetic potential in \( {\mathcal {A}}_{\mathrm {cut},\varepsilon }\)) Under the assumptions of Proposition 4.1, there exists \( \phi _{\varepsilon } \in C^{\infty }(\Omega ) \) such that, setting \( \psi (s,t) : = \psi ^{\mathrm {GL}}(\mathbf{r }(s,t)) e^{i \phi _{\varepsilon }(\mathbf{r }(s,t))} \), we get, as \( \varepsilon \rightarrow 0 \),

$$\begin{aligned} {\mathcal {E}}_{\varepsilon }^{\mathrm {GL}}\left[\psi ^{\mathrm {GL}}, \mathbf{A }^{\mathrm {GL}}; {\mathcal {A}}_{\mathrm {cut},\varepsilon }\right] \geqslant {\mathcal {G}}_{\varepsilon }[\psi ; {\mathcal {A}}_{\mathrm {cut}}] + {\mathcal {O}}(\varepsilon ^2 |\log \varepsilon |^2). \end{aligned}$$
(4.9)

Proof

As described above there are three operations, which are performed simultaneously, to get (4.9):

  • change to boundary tubular coordinates \( ({\mathsf {s}},{\mathsf {t}}) \);

  • extraction of a suitable gauge phase to replace \( \mathbf{A }^{\mathrm {GL}}\) with \( - {\mathsf {t}}\mathbf{e }_{{\mathsf {s}}} {+ \frac{1}{2} {\mathfrak {K}}({\mathsf {s}}){\mathsf {t}}^2 + \varepsilon \delta _{\varepsilon }} \);

  • rescaling of all the lengths (e.g., via (1.16)).

As anticipated, the above procedures have been already discussed in the literature, therefore we omit the details for the sake of brevity. We only provide the expression of the gauge phase for later convenience

$$\begin{aligned} \phi _\varepsilon (s,t) := -\frac{1}{\varepsilon }\int _0^{t} \, \mathrm {d}\eta \, \mathbf{A }^{\mathrm {GL}}(\mathbf{r }(\varepsilon s, \varepsilon \eta ))\cdot {\varvec{\nu }}(\varepsilon s) - \frac{1}{\varepsilon }\int _0^{s} \, \mathrm {d}\xi \, \mathbf{A }^{\mathrm {GL}}(\mathbf{r }(\varepsilon \xi ,0))\cdot {\varvec{\gamma }}^\prime (\varepsilon \xi ) + {\mathcal {O}}(\varepsilon ) s.\nonumber \\ \end{aligned}$$
(4.10)

\(\square \)

In the corner regions, on the opposite, it suffices to use a priori bounds on the solutions of the GL equations to substitute \( \mathbf{A }^{\mathrm {GL}}\) with \( \mathbf{F }\) (recall (2.17)). Before doing that, we need however a preparatory lemma:

Lemma 4.2

For any \( j {= 1, \ldots , N} \), as \( \varepsilon \rightarrow 0 \),

$$\begin{aligned} \left\Vert \left( \nabla + i \textstyle \frac{\mathbf{A }^{\mathrm {GL}}}{\varepsilon ^2} \right) \psi ^{\mathrm {GL}}\right\Vert_{L^2(\Gamma _{j,\varepsilon })} = {\mathcal {O}}(|\log \varepsilon |). \end{aligned}$$
(4.11)

Proof

The idea is to exploit the variational equation for \( \psi ^{\mathrm {GL}}\) in (B.6), to compute

$$\begin{aligned}&\left\Vert \left( \nabla + i \textstyle \frac{\mathbf{A }^{\mathrm {GL}}}{\varepsilon ^2} \right) \psi ^{\mathrm {GL}}\right\Vert^2_{L^2(\Gamma _{j,\varepsilon })} = \frac{1}{\varepsilon ^2} \int _{\Gamma _{j,\varepsilon }} \mathrm {d}\mathbf{r }\, \left( 1 - \left| \psi ^{\mathrm {GL}}\right|^2 \right) \left| \psi ^{\mathrm {GL}}\right|^2 \nonumber \\&\quad + \int _{\partial \Gamma _{j,\varepsilon }\setminus \partial \Omega } \mathrm {d}x \, {\psi ^{\mathrm {GL}}}^* {\varvec{\nu }}\cdot \left( \nabla + i \textstyle \frac{\mathbf{A }^{\mathrm {GL}}}{\varepsilon ^2} \right) \psi ^{\mathrm {GL}}= {\mathcal {O}}(|\log \varepsilon |^2), \end{aligned}$$
(4.12)

by the bounds (B.10), (B.13) and the boundary conditions on \( \psi ^{\mathrm {GL}}\) (recall (B.6)). \(\square \)

We can now perform the vector potential replacement:

Lemma 4.3

(Replacement of the magnetic field in \( \Gamma _{j,\varepsilon }\)) For any \( j = 1, \ldots , N \), there exists \( \psi _j \in H^1({\mathcal {A}}_{\varepsilon }) \), so that, as \( \varepsilon \rightarrow 0 \),

$$\begin{aligned} {\mathcal {E}}_{\varepsilon }^{\mathrm {GL}}\left[\psi ^{\mathrm {GL}}, \mathbf{A }^{\mathrm {GL}}; \Gamma _{j,\varepsilon }\right] \geqslant {\mathcal {E}}^{\mathrm {GL}}_1\left[\psi _j, \mathbf{F }; \Gamma _j \right] + {\mathcal {O}}(\varepsilon ^{3/5}). \end{aligned}$$
(4.13)

Remark 4.1

(Kinetic energy in the corner regions) Combining Lemma 4.2 with Lemma 4.3, one can easily deduce that (4.11) holds true with \( (\psi _j, \mathbf{F }) \) in place of \(( \psi ^{\mathrm {GL}}, \mathbf{A }^{\mathrm {GL}}) \). More precisely, let \( {\psi _j} \) as in (4.13), then

$$\begin{aligned} \left\Vert \left( \nabla + i \mathbf{F }\right) \psi _j \right\Vert_{L^2(\Gamma _j)} = {\mathcal {O}}(|\log \varepsilon |). \end{aligned}$$
(4.14)

Proof

A straightforward computation yields

$$\begin{aligned}&\left\Vert \left( \nabla + i \textstyle \frac{\mathbf{A }^{\mathrm {GL}}}{\varepsilon ^2} \right) \psi ^{\mathrm {GL}}\right\Vert^2_{L^2(\Gamma _{j,\varepsilon })} - \left\Vert \left( \nabla + i \textstyle \frac{\mathbf{F }}{\varepsilon ^2} \right) \psi ^{\mathrm {GL}}\right\Vert^2_{L^2(\Gamma _{j,\varepsilon })} \nonumber \\&\quad = - 2 \mathrm {Im}\int _{\Gamma _{j,\varepsilon }} \mathrm {d}\mathbf{r }\, \left[ \left( \nabla + i \textstyle \frac{\mathbf{A }^{\mathrm {GL}}}{\varepsilon ^2} \right) \psi ^{\mathrm {GL}}\right]^* \left( \textstyle \frac{\mathbf{A }^{\mathrm {GL}}}{\varepsilon ^2} - \textstyle \frac{\mathbf{F }}{\varepsilon ^2} \right) \psi ^{\mathrm {GL}}- \int _{\Gamma _{j,\varepsilon }} \mathrm {d}\mathbf{r }\, \left| \textstyle \frac{\mathbf{A }^{\mathrm {GL}}}{\varepsilon ^2} - \textstyle \frac{\mathbf{F }}{\varepsilon ^2} \right|^2 \left| \psi ^{\mathrm {GL}}\right|^2 \nonumber \\&\quad \geqslant - \delta \left\Vert \left( \nabla + i \textstyle \frac{\mathbf{A }^{\mathrm {GL}}}{\varepsilon ^2} \right) \psi ^{\mathrm {GL}}\right\Vert^2_{L^2(\Gamma _{j,\varepsilon })} - \textstyle \frac{1}{\varepsilon ^4} \left( \textstyle \frac{1}{\delta } + 1\right) \left\Vert \mathbf{A }^{\mathrm {GL}}- \mathbf{F }\right\Vert^2_{L^2(\Gamma _{j,\varepsilon })} \nonumber \\&\quad \geqslant - C \delta |\log \varepsilon |^2 - \textstyle \frac{1}{\varepsilon ^4} \left(\textstyle \frac{1}{\delta } + 1\right) \left\Vert \mathbf{A }^{\mathrm {GL}}- \mathbf{F }\right\Vert^2_{L^{2p}(\Omega )} \left| \Gamma _{j,\varepsilon }\right|^{1 - \frac{1}{p}}, \end{aligned}$$
(4.15)

for any \( p \in [2, \infty ) \), where we have used Lemma 4.2. Plugging now (B.11), which reads for \( p' \in [2, +\infty ) \)

$$\begin{aligned} \left\Vert \mathbf{A }^{\mathrm {GL}}- \mathbf{F }\right\Vert_{L^{p'}(\Omega )} = {\mathcal {O}}(\varepsilon ^{7/4}), \end{aligned}$$
(4.16)

thanks to the Agmon decay (B.16), we get

$$\begin{aligned}&\left\Vert \left( \nabla + i \textstyle \frac{\mathbf{A }^{\mathrm {GL}}}{\varepsilon ^2} \right) \psi ^{\mathrm {GL}}\right\Vert^2_{L^2(\Gamma _{j,\varepsilon })} - \left\Vert \left( \nabla + i \textstyle \frac{\mathbf{F }}{\varepsilon ^2} \right) \psi ^{\mathrm {GL}}\right\Vert^2_{L^2(\Gamma _{j,\varepsilon })}\geqslant - C \left[ \delta |\log \varepsilon |^2 \right. \nonumber \\&\quad \left. - \textstyle \frac{1}{\sqrt{\varepsilon }} \left(\textstyle \frac{1}{\delta } + 1\right) \left(\varepsilon ^2 |\log \varepsilon |^2 \right)^{1 - \frac{1}{p}} \right] \geqslant - C \varepsilon ^{3/4 - 1/p} |\log \varepsilon |^2 \end{aligned}$$
(4.17)

after an optimization over \( \delta \) (i.e., taking \( \delta = \varepsilon ^{3/4-1/p} |\log \varepsilon |^{{-1/p}} \)). The proof is then completed by exploiting the invariance of the energy under combined translations and global gauge change. \(\square \)

4.2 Rectification of the corner regions

We aim at estimating from below the energy close to the corners (first term on the r.h.s. of (4.13)) by the minimal energy of the model problem introduced in (2.12) and discussed in § 3. To this purpose there are two difficulties to overcome. First, one needs to force the boundary conditions in the minimization of the corner energies appearing in (4.13), since an unconstrained minimization would lead to unwanted contribution from the normal boundaries at \( s_j \pm L_{\varepsilon }\). Such a problem will however be solved by exploiting Proposition 3.5.

The second issue is less trivial: the model problem is indeed defined on a domain whose boundary is straight, while typically the boundaries of the domains \( \Gamma _j \) have a non-trivial curvature, as in Fig. 6. Of course, being the corner regions rather small and the curvature bounded, the corrections induced by this adjustment are of lower order (Lemma 4.4).

Let us introduce some notation: we are going to denote by \( \Gamma _{j, \mathrm {rect}}\) the corner region of opening angle \( \beta _j \) with straight sides, longitudinal length \( L_{\varepsilon }\) and normal width \( \ell _{\varepsilon }(1 + o(1)) \). We do not require the inner boundaries of \( \Gamma _{j, \mathrm {rect}}\) to be straight since the exponential decay of any GL minimizer makes such a boundary irrelevant. We also choose the coordinates in such a way that the corner coincides with the origin. Hence, except for the inner boundaries, \( \Gamma _{j, \mathrm {rect}}\) coincide with the region described in Fig. 2, up to a rotation:

$$\begin{aligned} \Gamma _{j, \mathrm {rect}}\simeq {\mathcal {R}} \Gamma _{\beta _j}(L_{\varepsilon },\ell _{\varepsilon }), \end{aligned}$$
(4.18)

with \( {\mathcal {R}} \) a rotation around the axis perpendicular to the plane passing through the corner.

Lemma 4.4

(Rectification of the corner) Let \( \psi _j \) be the \( H^1 \) function in (4.13). Then, there exists a diffeomorphism \( \mathbf{R }: \Gamma _j \rightarrow \Gamma _{j, \mathrm {rect}}\), so that, setting \( {{\widetilde{\psi }}}_j(\mathbf{R }) : = \psi _j(\mathbf{r }(\mathbf{R })) \),

$$\begin{aligned} {\mathcal {E}}^{\mathrm {GL}}_1\left[\psi _j, \mathbf{F }; \Gamma _j \right] = {\mathcal {E}}^{\mathrm {GL}}_1\big [{{\widetilde{\psi }}}_j, \mathbf{F }; \Gamma _{\beta _j}(L_{\varepsilon },\ell _{\varepsilon })\big ] + {\mathcal {O}}(\varepsilon |\log \varepsilon | ^{\infty }). \end{aligned}$$
(4.19)

Proof

We want to map the region \( \Gamma _j \) onto \( \Gamma _{j, \mathrm {rect}}\) via a suitable diffeomorphism and exploit the fact that, thanks to the boundedness of curvature and the size of the region, such a map is suitably close to the identity. A similar a trick has already been used, e.g., in [15]. Indeed, there exists a smooth map \( \mathbf{R }(\mathbf{r }) : \Gamma _j \rightarrow \Gamma _{j, \mathrm {rect}}\) which is one-to-one, such that \( \mathbf{R }(0) = 0 \) and

$$\begin{aligned} R_j(\mathbf{r }) = r_j \left( 1 + {\mathcal {O}}(\varepsilon |\log \varepsilon |) \right), \qquad \partial _j R_k(\mathbf{r }) = \delta _{jk} + {\mathcal {O}}(\varepsilon |\log \varepsilon | ^{\infty }). \end{aligned}$$
(4.20)

Using such a map, we get

$$\begin{aligned} {\mathcal {E}}^{\mathrm {GL}}_1\left[\psi _j, \mathbf{F }; \Gamma _j \right] = \int _{\Gamma _{j, \mathrm {rect}}} \mathrm {d}\mathbf{R }\left\{ \left| \left( J \nabla _{\mathbf{R }} + i {{\widetilde{\mathbf{F }}}}\right) {{\widetilde{\psi }}}_j \right|^2 - \textstyle \frac{1}{2b} \left( 2 \big | {{\widetilde{\psi }}}_j \big |^2 - \big | {{\widetilde{\psi }}}_j \big |^4 \right) \right\} , \end{aligned}$$
(4.21)

where J is the jacobian matrix associated to the change of coordinates \( \mathbf{r }\rightarrow \mathbf{R }\) and

$$\begin{aligned} {{\widetilde{\psi }}}_j(\mathbf{R }) : = \psi _j(\mathbf{r }(\mathbf{R })), \qquad {{\widetilde{\mathbf{F }}}}(\mathbf{R }) : = \mathbf{F }(\mathbf{r }(\mathbf{R })). \end{aligned}$$
(4.22)

By (4.20),

$$\begin{aligned} {{\widetilde{\mathbf{F }}}}(\mathbf{R }) = \textstyle \frac{1}{2} \mathbf{R }^{\perp } \left(1 + {\mathcal {O}}(\varepsilon |\log \varepsilon |) \right). \end{aligned}$$
(4.23)

Therefore, we can estimate from below the r.h.s. of (4.21) exactly as in (4.15), using (4.14) in Remark 4.1, to get

$$\begin{aligned}&{\mathcal {E}}^{\mathrm {GL}}_1\left[\psi _j, \mathbf{F }; \Gamma _j \right] \geqslant \int _{\Gamma _{j, \mathrm {rect}}} \mathrm {d}\mathbf{r }\left\{ \left| \left( \nabla + i \mathbf{F }\right) {{\widetilde{\psi }}}_j \right|^2 - \textstyle \frac{1}{2b} \left( 2 \big | {{\widetilde{\psi }}}_j \big |^2 - \big | {{\widetilde{\psi }}}_j \big |^4 \right) \right\} \nonumber \\&\quad - C \left[ \delta |\log \varepsilon |^2 + \left(1 + \textstyle \frac{1}{\delta }\right) \varepsilon ^2 |\log \varepsilon |^4 \right] \geqslant {\mathcal {E}}^{\mathrm {GL}}_1\big [{{\widetilde{\psi }}}_j, \mathbf{F }; \Gamma _{j, \mathrm {rect}}\big ] + {\mathcal {O}}(\varepsilon |\log \varepsilon | ^{\infty }). \end{aligned}$$
(4.24)

The last step is then the replacement of the region \( \Gamma _{j, \mathrm {rect}}\) with \( \Gamma _{\beta _j}(L_{\varepsilon },\ell _{\varepsilon })\), which can be done exploiting the rotational invariance of the GL functional, and the exponential decay of \( \psi ^{\mathrm {GL}}\) given by (B.16) (and thus of \( \psi \)). \(\square \)

4.3 Completion of the lower bound

We are now in position to complete the proof of the lower bound.

Proof of Proposition 4.1

Combining the results proven in Lemma 4.1, Lemmas 4.3 and 4.4, we get

$$\begin{aligned} E^{\mathrm {GL}}\geqslant {\mathcal {G}}_{\varepsilon }[\psi ; {\mathcal {A}}_{\mathrm {cut}}] + \sum _{j = 1}^N {\mathcal {E}}^{\mathrm {GL}}_1\big [{{\widetilde{\psi }}}_j, \mathbf{F }; \Gamma _{\beta _j}(L_{\varepsilon },\ell _{\varepsilon })\big ] {+ {\mathcal {O}}(\varepsilon ^{3/5})}. \end{aligned}$$
(4.25)

At this stage the energy contributions of the smooth part of the domain and its complement have been completely decoupled, so we can bound them from below separately. In fact, the lower bound to \( {\mathcal {G}}_{\varepsilon }[\psi ; {\mathcal {A}}_{\mathrm {cut}}] \) can be simply taken from [24, Prop. 4]:

$$\begin{aligned}&E^{\mathrm {GL}}\geqslant \displaystyle \frac{|\partial \Omega | E ^{\mathrm{1D}}_{0}}{\varepsilon } - 2 L_{\varepsilon }E ^{\mathrm{1D}}_{0}- E_{\mathrm {corr}}\int _{0}^{|\partial \Omega |} \mathrm {d}{\mathsf {s}}\, {\mathfrak {K}}({\mathsf {s}})\nonumber \\&\quad + \sum _{j = 1}^N \left[ {\mathcal {E}}^{\mathrm {GL}}_1\big [{{\widetilde{\psi }}}_j, \mathbf{F }; \Gamma _{\beta _j}(L_{\varepsilon },\ell _{\varepsilon })\big ] - \int _0^{c_0|\log \varepsilon |}\mathrm {d}t\, \left.\frac{F_0(t)}{f_0^2(t)} j_t \left[\psi (s,t)\right] \right|_{s=s_j - L_\varepsilon }^{s= s_j +L_\varepsilon } \right] {+ {\mathcal {O}}(\varepsilon ^{3/5})}. \end{aligned}$$
(4.26)

Now, recalling the definitions of \( \psi \) in Lemma 4.1 and \( {\widetilde{\psi }}_j \) in Lemma 4.4, we claim that

$$\begin{aligned} {\mathcal {E}}^{\mathrm {GL}}_1\big [{{\widetilde{\psi }}}_j, \mathbf{F }; \Gamma _{\beta _j}(L_{\varepsilon },\ell _{\varepsilon })\big ] - \int _0^{c_0|\log \varepsilon |}\mathrm {d}t\, \left.\frac{F_0(t)}{f_0^2(t)} j_t \left[\psi (s,t)\right] \right|_{s=s_j - L_\varepsilon }^{s= s_j +L_\varepsilon } = {\widetilde{{\mathcal {G}}}}_{\mathbf{F }}\big [{\widetilde{\psi }}_j \big ] + {\mathcal {O}}({\varepsilon ^{3/5}}),\nonumber \\ \end{aligned}$$
(4.27)

where the functional \( {\widetilde{{\mathcal {G}}}}_{\mathbf{F }} \) is defined in (3.39).

We note that, because of the rigid translation and rotation, the boundaries \( \partial \Gamma _{\beta _j, \mathrm {bd}} \) coincides with the portion of the lines \( s = s_j \pm L_{\varepsilon }\) in \( {\mathcal {A}} \). Therefore, in order to replace \( \psi \) with \( {{\widetilde{\psi }}}_j \) in the boundary terms in (4.27), we need to estimate the contribution of the gauge phase as well as the effect of the rectification. Next, we observe that

$$\begin{aligned} \left. j_t[\psi ] \right|_{s = s_j \pm L_{\varepsilon }}&= \left. j_t\left[ \psi _j((\mathbf{r }(\varepsilon s, \varepsilon t)) - \mathbf{r }_j)/\varepsilon ) e^{-i\phi _\varepsilon (s,t)} e^{ - i \frac{ \mathbf{F }(\mathbf{r }_j) \cdot \mathbf{r }'}{\varepsilon } } \right] \right|_{s = s_j \pm L_{\varepsilon }} \nonumber \\&= \left. \left[ j_t\left[ \psi _j((\mathbf{r }(\varepsilon s, \varepsilon t)) - \mathbf{r }_j)/\varepsilon ) \right] - \partial _t \left( \phi _\varepsilon (s,t) + \textstyle \frac{1}{\varepsilon } \mathbf{F }(\mathbf{r }_j) \cdot \mathbf{r }' \right) \right] \right|_{s = s_j \pm L_{\varepsilon }}. \end{aligned}$$
(4.28)

Recalling (4.22), we immediately see that the first term on the r.h.s. of the expression above in fact reconstructs the boundary terms in \( {\widetilde{{\mathcal {G}}}}_{\mathbf{F }} \). Moreover, by direct computation

$$\begin{aligned} - \partial _t \left( \phi _\varepsilon + \textstyle \frac{1}{\varepsilon } \mathbf{F }(\mathbf{r }_j) \cdot \mathbf{r }' \right) = \textstyle \frac{1}{\varepsilon } \left( \mathbf{A }^{\mathrm {GL}}(\mathbf{r }) - \mathbf{F }(\mathbf{r }_j) \right) \cdot \mathbf{e }_t = \textstyle \frac{1}{\varepsilon } \left( \mathbf{A }^{\mathrm {GL}}(\mathbf{r }) - \mathbf{F }(\mathbf{r }) \right) \cdot \mathbf{e }_t + \frac{1}{2} {\mathbf{r }^{\prime }}^{\perp } \cdot \mathbf{e }_t, \end{aligned}$$

where we have used the change of coordinates \( \mathbf{r }= \mathbf{r }_j + \varepsilon \mathbf{r }^{\prime } \). Furthermore, the properties of the diffeomorphism discussed in the proof of Lemma 4.4 imply that

$$\begin{aligned} \textstyle \frac{1}{2} {\mathbf{r }^{\prime }}^{\perp } \cdot \mathbf{e }_t = \textstyle \frac{1}{2} s + {\mathcal {O}}(\varepsilon |\log \varepsilon | ^{\infty }). \end{aligned}$$

Hence, each boundary term can be rewritten

$$\begin{aligned}&\int _0^{c_0|\log \varepsilon |}\mathrm {d}t\, \left. \frac{F_0(t)}{f_0^2(t)} j_t \left[\psi (s,t) \right] \right|_{s=s_j - L_\varepsilon }^{s= s_j +L_\varepsilon } = \int _0^{c_0|\log \varepsilon |}\mathrm {d}t\, \left. \frac{F_0(t)}{f_0^2(t)} \left[ j_t \left[{{\widetilde{\psi }}}(\mathbf{r }(s,t))\right] + \textstyle \frac{1}{2} s \right] \right|_{s=- L_\varepsilon }^{s= +L_\varepsilon } \nonumber \\&\quad - \frac{1}{\varepsilon } \int _0^{c_0|\log \varepsilon |}\mathrm {d}t\, \left. \frac{F_0(t)}{f_0^2(t)} \left( \mathbf{A }^{\mathrm {GL}}(\mathbf{r }(s,t)) - \mathbf{F }(\mathbf{r }(s,t)) \right) \cdot \mathbf{e }_t \right|_{s=s_j - L_\varepsilon }^{s= s_j +L_\varepsilon } + {\mathcal {O}}(\varepsilon |\log \varepsilon | ^{\infty }) \end{aligned}$$
(4.29)

and it only remains to bound the last term. This can be done exploiting once more (B.11): setting \( g(\mathbf{r }) : = F_0(t(\mathbf{r }))/f_0^2(t(\mathbf{r })) \) for short and using the vanishing of \( {F_0} \) at \( t = 0, c_1 |\log \varepsilon | \), we get

$$\begin{aligned}&\frac{1}{\varepsilon ^3} \left| \int _{\Gamma _{j,\varepsilon }} \mathrm {d}\mathbf{r }\, \nabla \cdot \left[ g \left( \mathbf{A }^{\mathrm {GL}}- \mathbf{F }\right) \right] \right| = \frac{1}{\varepsilon ^3} \left| \int _{\Gamma _{j,\varepsilon }} \mathrm {d}\mathbf{r }\, \left( \nabla g \right) \cdot \left( \mathbf{A }^{\mathrm {GL}}- \mathbf{F }\right) \right| \nonumber \\&\quad \leqslant \frac{C |\log \varepsilon |^5}{\varepsilon ^3} \left\Vert \mathbf{A }^{\mathrm {GL}}- \mathbf{F }\right\Vert_{L^1(\Gamma _{j,\varepsilon })}, \end{aligned}$$
(4.30)

where we have used that both \( \mathbf{A }^{\mathrm {GL}}\) and \( \mathbf{F }\) are divergence free and the estimate

$$\begin{aligned} \left| \nabla g(\mathbf{r }) \right| = \left| \partial _t \frac{F_0(t)}{f_0^2(t)} \right| \leqslant 2 \left| t + \alpha _0 \right| + \left| \frac{F_0(t) f_0^{\prime }(t)}{f_0^3(t)} \right| \leqslant C |\log \varepsilon |^5, \end{aligned}$$
(4.31)

by (A.3), (A.11) and the simple bound (A.25). On the other hand, by (4.16),

$$\begin{aligned} \left\Vert \mathbf{A }^{\mathrm {GL}}- \mathbf{F }\right\Vert_{L^1(\Gamma _{j,\varepsilon })} \leqslant \left\Vert \mathbf{A }^{\mathrm {GL}}- \mathbf{F }\right\Vert_{L^p(\Gamma _{j,\varepsilon })} \left| \Gamma _{j,\varepsilon }\right|^{1 - \frac{1}{p}} \leqslant C \varepsilon ^{7/4} \left( \varepsilon ^2 |\log \varepsilon | \right)^{1 - \frac{1}{p}}, \end{aligned}$$
(4.32)

for \( p\in [2, +\infty ) \), which implies (4.27) via (4.30).

Putting together (4.26) with (4.27) and observing that, by the Agmon decay, we can impose the boundary condition \( {\widetilde{\psi }} = \psi _0 \) along the interior boundary \( \partial \Gamma _{\mathrm {in}}\) up to \( {\mathcal {O}}(\varepsilon ^{\infty }) \) errors, we thus get

$$\begin{aligned}&E^{\mathrm {GL}}\geqslant \displaystyle \frac{|\partial \Omega | E ^{\mathrm{1D}}_{0}}{\varepsilon } - 2 L_{\varepsilon }E ^{\mathrm{1D}}_{0}- E_{\mathrm {corr}}\int _{0}^{|\partial \Omega |} \mathrm {d}{\mathsf {s}}\, {\mathfrak {K}}({\mathsf {s}})\nonumber \\&\quad + \sum _{j \in \Sigma } \inf _{\psi \in {{\widetilde{{\mathscr {D}}}}}_{\star }(\Gamma _{\beta _j}(L_{\varepsilon },\ell _{\varepsilon }))} {\widetilde{{\mathcal {G}}}}_{{\mathbf{F }}}\left[ {{\widetilde{\psi }}}, \Gamma _{\beta _j}(L_{\varepsilon },\ell _{\varepsilon }) \right] + {\mathcal {O}}(\varepsilon ^{3/5}) \nonumber \\&\quad = \displaystyle \frac{|\partial \Omega | E ^{\mathrm{1D}}_{0}}{\varepsilon } - E_{\mathrm {corr}}\int _{0}^{|\partial \Omega |} \mathrm {d}{\mathsf {s}}\, {\mathfrak {K}}({\mathsf {s}})+ \sum _{j \in \Sigma } \left( {\widetilde{E}}_{\beta _j}(L_{\varepsilon },\ell _{\varepsilon }) - 2 L_{\varepsilon }E ^{\mathrm{1D}}_{0}\right) + {\mathcal {O}}(\varepsilon ^{3/5}). \end{aligned}$$
(4.33)

The final step of the proof is the application of Proposition 3.5, which yields (see (3.74))

$$\begin{aligned} {\widetilde{E}}_{\beta _j}(L_{\varepsilon },\ell _{\varepsilon }) - 2 L_{\varepsilon }E ^{\mathrm{1D}}_{0}= E_{\mathrm {corner},\beta _j} + o(1), \end{aligned}$$
(4.34)

and thus the result. \(\square \)

5 Other proofs

We complete in this Section the proofs of the results proven in the paper, i.e., specifically, we prove the energy upper bound matching the lower bound proven in Proposition 4.1. Finally, we show how the energy asymptotics can be used to deduce a pointwise estimate of the order parameter.

5.1 Upper bound and energy asymptotics

We state the main result of this section in the following Proposition 5.1. Note that Propositions 4.1 and 5.1 together completes the proof of Theorem 2.1, with the simple exception of the replacement of \( E ^{\mathrm{1D}}_{0}\) with \( E^{\mathrm {1D}}_{\star }\), which can be done up to remainders of order \( {\mathcal {O}}(\varepsilon ^{\infty }) \) by Lemma A.2.

Proposition 5.1

(GL energy upper bound) Let \(\Omega \subset {\mathbb {R}}^2\) be any bounded simply connected domain satisfying Assumptions 1 and 2. Then, for any fixed \( 1< b < \Theta _0^{-1} \), as \(\varepsilon \rightarrow 0\), it holds

$$\begin{aligned} E^{\mathrm {GL}}\leqslant \displaystyle \frac{|\partial \Omega | E ^{\mathrm{1D}}_{0}}{\varepsilon } - E_{\mathrm {corr}}\int _{0}^{|\partial \Omega |} \mathrm {d}{\mathsf {s}}\, {\mathfrak {K}}({\mathsf {s}})+ \sum _{j = 1}^N E_{\mathrm {corner}, \beta _j} + o(1). \end{aligned}$$
(5.1)

Proof

As usual the upper bound is obtained by testing the GL functional on a suitable trial state. As a vector potential, we pick \( \mathbf{F } = \frac{1}{2} (-y,x) \). The order parameter on the other hand is much more involved: the idea is to recover the trial state given in [24, Eq. (4.14)] far from the corners and glue it to the minimizers of the effective energies in every corner. To retain the curvature corrections, as in [23] (see also “Appendix C”), we decompose the smooth part of the layer into cells of tangential length of order \( \varepsilon \). The order parameter is constructed in such a way that its modulus is close to \( f_{k_n}(\varepsilon {\mathsf {t}}) \) (see “Appendix A.2”) in each cell, \( k_n \) being the average curvature, and to the modulus of the corner minimizer \( \psi _{\beta _j}(\mathbf{r }) \) in the \(j-\)th corner region, respectively. The phase of \( \psi _{\mathrm {trial}}\) on the other hand is given by a gauge phase analogous to (4.10), but defined in terms of the vector potential \( \mathbf{F } \), plus the optimal phase \( \exp \{- i \alpha _{k_n} {\mathsf {s}}/ \varepsilon \} \) in each cell. An additional phase is then added to patch together such factors. Explicitly, we set

$$\begin{aligned} \psi _{\mathrm {trial}}(\mathbf{r }) : = \chi \left( t(\mathbf{r }) \right) \cdot {\left\{ \begin{array}{ll} g(s(\mathbf{r }),t(\mathbf{r })) e^{- i S(s(\mathbf{r }))} e^{i \phi _{\mathrm {trial}}(s(\mathbf{r }),t(\mathbf{r }))}, &{} \text{ for } \mathbf{r }\in \widetilde{{\mathcal {A}}}_{\mathrm {cut},\varepsilon }, \\ {{\widetilde{\psi }}}_{\beta _j,\varkappa _j}\left({\mathcal {R}}^{-1}(\mathbf{r }-\mathbf{r }_j)/\varepsilon \right), &{} \text{ for } \mathbf{r }\in \Gamma _{j,\varepsilon }, \end{array}\right. } \end{aligned}$$
(5.2)

where \( {\mathcal {R}} \) is the rotation defined in (4.18) and \( \varkappa _j \) a suitable phase factor, given in (5.5) below. Moreover, for some \( a > 0 \),

$$\begin{aligned} \widetilde{{\mathcal {A}}}_{\mathrm {cut},\varepsilon }: = \left\{ \mathbf{r }\in {\mathcal {A}}_{\mathrm {cut},\varepsilon }\, \big | \, \left| s(\mathbf{r }) - s_j \right| \geqslant L_{\varepsilon }+ \varepsilon ^{a}, \forall j {= 1, \ldots , N} \right\} , \end{aligned}$$
(5.3)

is a subdomain of \( {\mathcal {A}}_{\mathrm {cut},\varepsilon }\) where boundary coordinates are well defined. In \( {\mathcal {A}}_{\mathrm {cut},\varepsilon }\setminus \widetilde{{\mathcal {A}}}_{\mathrm {cut},\varepsilon }\), we take care of the transition from the smooth part of the layer to the corner region: for any given \( j {= 1, \ldots , N} \), we concretely set

$$\begin{aligned} \psi _{\mathrm {trial}}(\mathbf{r }) : = \zeta (s) f_0 \left(t\right) e^{- i S(s_j \pm L_{\varepsilon })} +\left(1 - \zeta (s) \right) f_{k_{\pm }}\left(t\right) e^{- i S(s_j \pm L_{\varepsilon }\pm \varepsilon ^{a})}, \end{aligned}$$
(5.4)

for any \( |s(\mathbf{r }) - s_j \pm L_{\varepsilon }| \leqslant \varepsilon ^{a} \) and where we have denoted \( k_{\pm } \) the average curvature in the cells \( {\mathscr {C}}_{\pm } \) adjacent to the j-th corner region. The smooth cut-off function \( \zeta \) is chosen in such a way that it is positive and

$$\begin{aligned} \zeta (s_j \pm L_{\varepsilon }) = 1, \qquad \zeta (s_j \pm L_{\varepsilon }\pm \varepsilon ^{a}) = 0, \qquad |\zeta ^{\prime }| \leqslant C \varepsilon ^{-a} . \end{aligned}$$

Furthermore,

  • the smooth cut-off function \( \chi (t) \) equals 1 for any \( t \leqslant c_1 |\log \varepsilon | \) and vanishes for \( t \geqslant |\log \varepsilon |^2 \), so that its gradient is bounded by \( {\mathcal {O}}(|\log \varepsilon |^{-2}) \);

  • the gauge phase \( \phi _{\mathrm {trial}}\) is given by (4.10) with \( \mathbf{F }\) in place of \( \mathbf{A }^{\mathrm {GL}}\); to leading order such a phase equals \( - \frac{1}{2} st \), so recovering part of the phase of \( \psi _{\star } \) (recall (2.16));

  • the function g(st) is taken directly from [24, Eq. (4.15)] (see also [24, Eq. (4.18)]): it equals \( f_{k_n} \) in \( {\mathscr {C}}_n \), up to a smaller correction \( \chi _n \), which allows the continuous transition from \( f_{k_n} \) to \( f_{k_{n+1}} \);

  • similarly, the phase S(s) is given by [24, Eqs. (4.20) & (4.21)]: to leading order \( S(s) = - i \alpha _{k_n} s \) in \( {\mathscr {C}}_n \), but, as for the density, one needs to add a higher order correction taking into account the jump from \( \alpha _{k_n} \) to \( \alpha _{k_{n+1}} \);

  • \( {{\widetilde{\psi }}}_{\beta _j,\varkappa _j}(\mathbf{r }') \) is close to the minimizer of the effective energy (3.75) in \( \Gamma _{\beta _j}(L_{\varepsilon },\ell _{\varepsilon })\), where \( \varkappa _j : = \varkappa _{+,j} - \varkappa _{-,j} \) and

    $$\begin{aligned} \varkappa _{\pm , j} : = - S(s_j \pm L_{\varepsilon }) + \alpha _0 (s_j \pm L_{\varepsilon }) = - \int _0^{s_j \pm L_{\varepsilon }} \mathrm {d}\xi \, \left( \alpha _{k(\xi )} - \alpha _0\right) + {\mathcal {O}}(|\log \varepsilon | ^{\infty }).\nonumber \\ \end{aligned}$$
    (5.5)

    Applying the rectification procedure described in Lemma 4.4, we set

    $$\begin{aligned} {{\widetilde{\psi }}}_{\beta _j,\varkappa _j}(\mathbf{r }') = \psi _{\beta _j,\varkappa _j}(\mathbf{R }(\mathbf{r })), \end{aligned}$$
    (5.6)

    where \( \mathbf{R }\) is the diffeomorphism of Lemma 4.4 and \( \psi _{\beta ,\varkappa } \) any minimizer of (3.75).

We now sketch the main steps in the computation of the energy of \( \left( \psi _{\mathrm {trial}}, \mathbf{F } \right) \), which were already discussed elsewhere and focus afterwards on the new estimates:

  • since \( \mathrm {curl}\mathbf{F } = 1 \) in \( \Omega \), the last term in the GL energy functional (1.1) vanishes;

  • the integration can be restricted to \( {\mathcal {A}}_{\varepsilon }\), where the cut-off function \( \chi \) is 1 and all the rest of the energy can be discarded thanks to the exponential decay of the modulus of \( \psi _{\mathrm {trial}}\) as well as its derivatives (inherited from \( f_0\) and \( f_k \), see (A.10));

  • the gauge phase \( \phi _{\mathrm {trial}}\) allows to replace \( \mathbf{F }\) with \( \left( - t + \frac{1}{2} \varepsilon k(s) t^2 \right) \mathbf{e }_s \) in \( \widetilde{{\mathcal {A}}}_{\mathrm {cut},\varepsilon }\), as in Lemma 4.1 up to an error of order \( {\mathcal {O}}(\varepsilon |\log \varepsilon | ^{\infty }) \);

  • the energy bound in \( \widetilde{{\mathcal {A}}}_{\mathrm {cut},\varepsilon }\) is taken from [24] and stated in Proposition C.1:

    $$\begin{aligned} {\mathcal {E}}^{\mathrm {GL}}\left[ \psi _{\mathrm {trial}}, \mathbf{F }; {\mathcal {A}}_{\mathrm {cut},\varepsilon }\right] \leqslant \displaystyle \frac{|\partial \Omega | E ^{\mathrm{1D}}_{0}}{\varepsilon } - 2 L_{\varepsilon }N E ^{\mathrm{1D}}_{0}- \varepsilon E_{\mathrm {corr}}\int _{0}^{\frac{|\partial \Omega |}{\varepsilon }} \mathrm {d}s \, k(s) + o(1). \end{aligned}$$

Given the discussion above, it remains to compute the energy of \( \psi _{\mathrm {trial}}\) in the region \( {\mathcal {A}}_{\mathrm {cut},\varepsilon }\setminus \widetilde{{\mathcal {A}}}_{\mathrm {cut},\varepsilon }\) as well as the energy contributions of all the corner regions \( \Gamma _{j,\varepsilon }\). Let us start by considering the latter: getting rid of the diffeomorphism up to small errors, close to each corner we recover \( E_{\beta _j,\varkappa _j}(L_{\varepsilon },\ell _{\varepsilon }) = E_{\beta _j}(L_{\varepsilon },\ell _{\varepsilon }) + o(1) \) by Lemma 3.2. Summing up, we get

$$\begin{aligned} {\sum _{j = 1}^N} E_{\beta _j}(L_{\varepsilon },\ell _{\varepsilon }) + o(1) = {\sum _{j = 1}^N} E_{\mathrm {corner}, \beta _j} + 2 N E ^{\mathrm{1D}}_{0}L_{\varepsilon }+ o(1), \end{aligned}$$
(5.7)

where we have exploited the existence of the limit proven in Proposition 2.2.

Finally, let us consider the energy in \( {\mathcal {A}}_{\mathrm {cut},\varepsilon }\setminus \widetilde{{\mathcal {A}}}_{\mathrm {cut},\varepsilon }\) and restrict ourselves to the interval \( [s_j - L_{\varepsilon }- \varepsilon ^a, s_j - L_{\varepsilon }] \): the area of the region is of order \( \varepsilon ^{2+a} |\log \varepsilon | \) and we can thus discard all the terms involving \( f_0 \), \( f_{k_{\pm }} \) and their derivatives there up to errors of order \( {\mathcal {O}}(\varepsilon ^a|\log \varepsilon | ^{\infty }) \) by (A.12) and (A.10). The only non-trivial term to estimate is thus the kinetic energy of the cut-off function \( \zeta \): by grouping together the terms in a convenient way, one has to bound at the boundary \( s_j - L_{\varepsilon }\) the quantity

$$\begin{aligned}&C \int _{s_j - L_{\varepsilon }- \varepsilon ^a}^{s_j -L_{\varepsilon }} \mathrm {d}s \int _0^{\ell _{\varepsilon }} \mathrm {d}t \, \left|\zeta '(s)\right|^2 \left| f_0 \left(t\right) e^{-iS(s_j - L_{\varepsilon })} - f_{k_{-}}\left(t\right) e^{- i S(s_j - L_{\varepsilon }- \varepsilon ^{a})} \right|^2 \nonumber \\&\leqslant C \varepsilon ^{-2a} \int _{s_j {-}L_{\varepsilon }{-}\varepsilon ^a}^{s_j {-}L_{\varepsilon }} \mathrm {d}s \int _0^{\ell _{\varepsilon }} \mathrm {d}t \, \left[ \left| f_0 \left(t\right) - f_{k_{-}}\left(t\right) \right|^2 + f_0^2(t) \left| e^{- i S(s_j - L_{\varepsilon })} - e^{- i S(s_j - L_{\varepsilon }- \varepsilon ^{a})} \right|^2 \right] \nonumber \\&\quad \leqslant C \left[ \varepsilon ^{1-a} + \varepsilon ^a |\log \varepsilon | ^{\infty }\right], \end{aligned}$$
(5.8)

by the identity

$$\begin{aligned} S(s) = \alpha _0 s + \int _0^{s} \mathrm {d}\xi \, \left( \alpha _{k(\xi )} - \alpha _0\right) + {\mathcal {O}}(|\log \varepsilon | ^{\infty }). \end{aligned}$$
(5.9)

Putting together all the energy contributions, we get (5.1). \(\square \)

5.2 Order parameter

As proven in [19], any minimizing \( \psi ^{\mathrm {GL}}\) is such that its modulus is suitably close in \( L^2({\mathcal {A}}_{\varepsilon }) \) to the 1D profile \( f_0(\mathrm {dist}(\, \cdot \,, \partial \Omega )/\varepsilon ) \). The presence of the corners affects the estimate only at the precision one can approximate \( |\psi ^{\mathrm {GL}}| \) with \( f_0 \), since the result in [19, Thm. 1.1] is proven by neglecting the corner regions. The improved energy asymptotics of Theorem 2.1 obviously suggests that such an estimate can in fact be strengthened. Indeed, we prove here that a pointwise estimate of the difference \( |\psi ^{\mathrm {GL}}| - f_0 \) holds true in the smooth part of the boundary layer.

Proof of Proposition 2.1

The starting point is the combination of the energy upper bound (5.1), with the stronger lower bound which can be obtained by combining the arguments of the proof of Proposition 4.1 with Lemma C.1: in each cell contained in the smooth part of the boundary layer, we can retain the positive contribution appearing on the r.h.s. of (C.13). The final outcome is the estimate

$$\begin{aligned} \sum _{{n} = 1}^{M_{\varepsilon }} \int _{{\mathscr {C}}_n}\mathrm {d}s\mathrm {d}t\, (1-\varepsilon k_nt)f_n^4(1-|u_n|^2)^2 = o(1). \end{aligned}$$
(5.10)

A direct consequence is the estimate stated in (2.26) in Remark 2.3. Furthermore, (5.10) is the key ingredient of a typical argument (first used in [10]) to deduce a pointwise estimate of \( |u_n| \) and thus \( |\psi ^{\mathrm {GL}}| = f_0 |u_n| \) (see [24, Proof of Thm. 2, Step 2] but also the proof of (3.28) in Proposition 3.3).

Instead of providing all the details, we comment only on the needed adaptations. First of all, one has to select a subdomain of \( {\mathscr {C}}_n\), where a suitable lower bound on the density \( f_n \) holds true. In our case, we can restrict the analysis to the layer \( \left\rbrace \mathrm {dist}(\mathbf{r }, \partial \Omega ) \leqslant c \varepsilon \right\lbrace \), where \( f_n \) is bounded from below by a positive constant independent of \( \varepsilon \). As a consequence, the argument of [24, Proof of Thm. 2, Step 1] leads to

$$\begin{aligned} \left| \nabla \left|u_n \right| \right| \leqslant C, \qquad \text{ in } \left\{ \mathrm {dist}(\mathbf{r }, \partial \Omega ) \leqslant c \varepsilon \right\} \cap {\mathscr {C}}_n. \end{aligned}$$
(5.11)

With such a bound at disposal, the aforementioned argument applies straightforwardly in the boundary region \( \left\rbrace \mathrm {dist}(\mathbf{r }, \partial \Omega ) \leqslant c \varepsilon \right\lbrace \), leading to the pointwise estimate

$$\begin{aligned} \left| \left| u_n \right| - 1 \right| = o(1), \qquad \text{ in } \left\{ \mathrm {dist}(\mathbf{r }, \partial \Omega ) \leqslant c \varepsilon \right\} \cap {\mathscr {C}}_n, \end{aligned}$$
(5.12)

which immediately yields (2.27). \(\square \)