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

Helical magnetic fields and semi-classical asymptotics of the lowest eigenvalue

B. Helffer, A. Kachmar
Abstract

We study the 3D Neuman magnetic Laplacian in the presence of a semi-classical parameter and a non-uniform magnetic field with constant intensity. We determine a sharp two term asymptotics for the lowest eigenvalue, where the second term involves a quantity related to the magnetic field and the geometry of the domain. In the special case of the unit ball and a helical magnetic field, the concentration takes place on two symmetric points of the unit sphere.

1 Main results

Let Ω3\Omega\subset\mathbb{R}^{3} be an open and bounded set with a smooth boundary Ω\partial\Omega. Let us consider a smooth magnetic field 𝐁:Ω¯3\mathbf{B}:\overline{\Omega}\to\mathbb{R}^{3} (so 𝐁\mathbf{B} should be closed) which will always be assumed to satisfy

xΩ,|𝐁(x)|=b\forall\,x\in\Omega,\quad|\mathbf{B}(x)|=b (1.1)

where b>0b>0 is a constant. Without loss of generality, we assume from now on that b=1b=1. Let 𝐀(x){\bf A}(x) be a magnetic potential such that

curl𝐀=𝐁.{\rm curl}\,{\bf A}=\mathbf{B}\;. (1.2)

We are interested in the analysis of the lowest eigenvalue λ1(𝐀,h)\lambda_{1}({\bf A},h) of the Neumann realization of the Schrödinger operator in Ω\Omega with magnetic field

P𝐀h:=Δh,𝐀=j=13(hDxj+Aj(x))2.P_{\bf A}^{h}:=\Delta_{h,{\bf A}}=\sum_{j=1}^{3}(hD_{x_{j}}+A_{j}(x))^{2}\;. (1.3)

We introduce the following assumptions.

Assumption 1.1 (C1).

The set of boundary points where 𝐁\mathbf{B} is tangent to Ω\partial\Omega, i.e.

Γ:={xΩ|𝐁𝐍(x)=0},\displaystyle\Gamma:=\{x\in\partial\Omega\,\big{|}\,\mathbf{B}\cdot\mathbf{N}(x)=0\}, (1.4)

is a regular submanifold of Ω\partial\Omega :

κn,𝐁(x):=|dT(𝐁𝐍)(x)|0,xΓ.\kappa_{n,\mathbf{B}}(x):=|d^{T}(\mathbf{B}\cdot\mathbf{N})(x)|\neq 0\;,\;\forall x\in\Gamma\;. (1.5)

Here dTd^{T} is the differential defined on functions on Ω\partial\Omega and 𝐍(x)\mathbf{N}(x) is the unit inward normal of Ω\Omega.

Assumption 1.2 (C2).

The set of points where 𝐁\mathbf{B} is tangent to Γ\Gamma is finite.

These assumptions are rather generic and for instance satisfied for ellipsoids, when 𝐁\mathbf{B} is constant. When |𝐁||\mathbf{B}| is constant, the above assumptions hold for the sphere with a helical magnetic field (see Sec. 3).

Let us introduce the constant γ^0,𝐁\widehat{\gamma}_{0,\mathbf{B}} involving the “magnetic curvature” in (1.5), which is defined by

γ^0,𝐁:=infxΓγ~0,𝐁(x),\widehat{\gamma}_{0,\mathbf{B}}:=\inf_{x\in\Gamma}\widetilde{\gamma}_{0,\mathbf{B}}(x), (1.6)

where

γ~0,𝐁(x):=22/3ν^0δ01/3|κn,𝐁(x)|2/3(1(1δ0)|𝐓(x)𝐁(x)|2)1/3.\widetilde{\gamma}_{0,\mathbf{B}}(x):=2^{-2/3}\widehat{\nu}_{0}\delta_{0}^{1/3}|\kappa_{n,\mathbf{B}}(x)|^{2/3}\Big{(}1-(1-\delta_{0})|\mathbf{T}(x)\cdot\mathbf{B}(x)|^{2}\Big{)}^{1/3}\;. (1.7)

Here 𝐓(x)\mathbf{T}(x) is the oriented, unit tangent vector to Γ\Gamma at the point xx, δ0]0,1[\delta_{0}\in]0,1[ and ν^0>0\widehat{\nu}_{0}>0 are spectral quantities relative to the De Gennes and Montgomery operators which will be introduced in (4.2) and (4.5).

When 𝐁\mathbf{B} is constant, the following two-term asymptotics of λ1(B)\lambda_{1}(B) has been established by Helffer-Morame [HelMo4] and Pan [Pan3].

Theorem 1.3.

Let us assume that 𝐁\mathbf{B} is constant. Then, if Ω\Omega and 𝐁\mathbf{B} satisfy (C1)-(C2), there exists η>0\eta>0 such that the lowest eigenvalue λ1N(𝐀,h)\lambda_{1}^{N}({\bf A},h) satisfies as h0h\rightarrow 0

λ1N(𝐀,h)=Θ0h+γ^0,𝐁h43+𝒪(h43+η).\lambda_{1}^{N}({\bf A},h)=\Theta_{0}h+\widehat{\gamma}_{0,\mathbf{B}}h^{\frac{4}{3}}+{\mathcal{O}}(h^{\frac{4}{3}+\eta})\,. (1.8)

The aim of this paper is to prove that Theorem 1.3 also holds under the weaker assumption that |𝐁||\mathbf{B}| is constant.

Theorem 1.4.

Under the assumptions (C1)-(C2), if |𝐁||\mathbf{B}| is constant, then the asymptotics in (1.8) holds for the lowest eigenvalue λ1N(𝐀,h)\lambda_{1}^{N}({\bf A},h).

An interesting example of a non-constant magnetic field but with a constant intensity is the helical magnetic field occurring in the theory of liquid crystals. Up to the action of an orthogonal matrix, it can be expressed as follows [Pan6]

𝐁=curl𝐧τ=τ𝐧τ,𝐧τ=(1τcos(τx3),1τsin(τx3),0).\mathbf{B}={\rm curl}\,\mathbf{n}_{\tau}=-\tau\mathbf{n}_{\tau},\quad\mathbf{n}_{\tau}=\Big{(}\frac{1}{\tau}\cos(\tau x_{3}),\frac{1}{\tau}\sin(\tau x_{3}),0\Big{)}. (1.9)

Here τ>0\tau>0 is a given constant. In this situation (𝐁=τ𝐧τ\mathbf{B}=-\tau\mathbf{n}_{\tau}), [Pan6] derived an upper bound on the eigenvalue λ1N(𝐀,h)\lambda_{1}^{N}({\bf A},h), which is consistent with Theorem 1.4. Our contribution is valid for a more general class of magnetic fields with constant intensity and also determines the asymptotically matching lower bound of the lowest eigenvalue.

Discussion and applications

The inspection of the eigenvalue λ1N(𝐀,h)\lambda_{1}^{N}({\bf A},h) is vital in understanding the transition between superconducting and normal states in the Ginzburg-Landau model [FoHe2]. In this context, the magnetic field is typically constant. Accurate estimates of the lowest eigenvalue λ1N(𝐀,h)\lambda_{1}^{N}({\bf A},h) under constant magnetic fields [HelMo3, HelMo4] led to a precise understanding of the transition between superconducting and normal states [FoHe1, FS].

Non-homogeneous magnetic fields with constant intensity are encountered in the Landau–de Gennes theory of liquid crystals, which is the analog of the Ginzburg-Landau theory of superconductivity. Here a transition between smectic and nematic phases occurs. Our main result, Theorem 1.4, yields an accurate estimate of the lowest eigenvalue λ1N(𝐀,h)\lambda_{1}^{N}({\bf A},h) for magnetic fields with constant intensity, and by analogy with [FoHe1], we expect it to yield a precise description of the transition between surface smectic and nematic states (see [Pan2]).

At the threshold of the phase transition, both superconductive and smectic states nucleate on the surface of the domain (near the curve Γ\Gamma introduced in (3.7)). The paper [Pan5] contains a nice discussion of this interesting analogy. The analysis of 3D surface superconductivity is the subject of the papers [Pan3, FKP, FMP], while surface smectics are rigorously studied in [HePa2, FKPa]. It would be interesting to complete this analysis by providing more accurate estimates at the threshold, where the linear analysis (such as the one in this paper) becomes handy.

The analysis in this paper concerns the lowest eigenvalue. In the presence of a constant magnetic field, and a “single well” assumption (i.e. the minimum in (1.6) is non-degenerate and attained at a unique point), accurate estimates of the low-lying eigenvalues were obtained recently in [HR]. In our setting of a non-homogeneous magnetic field, the example of the ball under the helical magnetic field suggests the presence of multiple wells (see Remark 3.5).

The interaction between magnetic fields and 3D domains is interesting in other situations. In particular, for the Robin problem, we observe pure magnetic wells on the surface of the domain [HKR], and in the case of a constant magnetic field, strong diamgnetism does not hold for the ball [Mi].

Organization and outline of the proof

The proof of Theorem 1.4 is split into two parts. In the first part, we establish a lower bound of the lowest eigenvalue, by comparing the quadratic form via a simpler form related to a new model operator. Comparing with the constant magnetic field in [HelMo4], we prove that the model operator in our setting is a perturbation of the one considered in [HelMo4].

The second part of the proof is devoted to an upper bound of the lowest eigenvalue, already studied for 𝐁\mathbf{B} in (1.9) [Pan6], but we revisit it since our formulation is not the same as [Pan6]. The upper bound follows after computing the quadratic form of a suitable trial state, having the same structure as the constant magnetic field case in [HelMo4, Pan3]. However, there are additional terms in the computations due to the varying magnetic field, which require a careful handling.

The model operator takes into consideration two phenomena. First, after decomposing our domain into small cells and working in a small cell near the domain’s boundary, we have to express the integrals in a flat geometry, which requires a careful expansion of the Riemannian metric in particular. This part is essentially the same as for the constant magnetic field case in [HelMo4].

Then, we have to express the magnetic potential in adapted coordinates, in each small cell, and apply a Taylor expansion and a gauge transformation to obtain a “normal” form, i.e. a simpler effective magnetic potential. In this part, we deviate from the constant magnetic field situation and find additional terms in the effective magnetic potential. Interestingly, we can still show that the analysis with this magnetic potential is somehow independent of those additional terms and treat the new model as a perturbation of the model with a constant magnetic field.

The paper is organized as follows. In Section 2 we introduce the adapted coordinates in a small “boundary” cell. In Section 3, we analyze the case of the unit ball with the “helical” magnetic field occurring in liquid crystals and verify that Assumptions 1.1 and 1.2 hold. Interestingly, after computing the energy in (1.6), we notice that this example shows a phenomenon of multiple “wells” induced by the “magnetic” geometry.

In Section 4, we review two standard 1D operators that we need in defining the quantities appearing in (1.6) and the statement in Theorem 1.4. Then, in Section 5, we introduce a new model, specific to our case of a varying magnetic field with a constant intensity, and analyze it through a perturbation argument.

With the model in Section 5, we can adjust the proof in [HelMo4] and prove Theorem 1.4. The first step is to localize the ground states near the boundary, which is the content of Section 6. Then, the approximation of the quadratic form and the magnetic potential are the subject of Section 7, which allows us, in the subsequent Section 8, to obtain a lower bound on the lowest eigenvalue.

Finally, Section 9 is devoted to the computation of the energy of a trial state, which yields an upper bound of the lowest eigenvalue, and thereby completes the proof of Theorem 1.4.

2 Adapted coordinates

We recall a rather standard choice of coordinates in the neighborhood of Γ\Gamma.

2.1 Description of the coordinates

Let g0g_{0} be the Riemannian metric on 3\mathbb{R}^{3}, which induces a Riemanian metric GG on Ω\partial\Omega. Given two vector fields 𝐗,𝐘\mathbf{X},\mathbf{Y} of 3\mathbb{R}^{3}, we denote by

𝐗𝐘=𝐗,𝐘:=g0(𝐗,𝐘).\mathbf{X}\cdot\mathbf{Y}=\langle\mathbf{X},\mathbf{Y}\rangle:=g_{0}(\mathbf{X},\mathbf{Y})\,. (2.1)

Consider a direct frame (𝐕(x),𝐓(x),𝐍(x))xΓ(\mathbf{V}(x),\mathbf{T}(x),\mathbf{N}(x))_{x\in\Gamma} along Γ\Gamma such that

  • 𝐓(x)\mathbf{T}(x) is an oriented unit tangent vector of Γ\Gamma ;

  • 𝐕(x):=𝐓(x)×𝐍(x)\mathbf{V}(x):=\mathbf{T}(x)\times\mathbf{N}(x), hence determining an oriented normal to the curve Γ\Gamma in the tangent space to Ω\partial\Omega.

For mΓm\in\Gamma, let Λm\Lambda_{m} be the geodesic that passes through mm and is normal to Γ\Gamma. Let x0Γx_{0}\in\Gamma. In some neighborhood 𝒩x0Ω¯\mathcal{N}_{x_{0}}\subset\overline{\Omega} of x0x_{0}, we can introduce new coordinates (r,s,t)(r,s,t) as follows:

  • For x𝒩x0x\in\mathcal{N}_{x_{0}}, p(x)Ωp(x)\in\partial\Omega is defined by dist(x,p(x))=t(x):=dist(x,Ω){\rm dist}(x,p(x))=t(x):={\rm dist}(x,\partial\Omega);

  • For x𝒩x0x\in\mathcal{N}_{x_{0}}, γ(x)Γ\gamma(x)\in\Gamma is defined by distΩ(p(x),γ(x))=distΩ(p(x),Γ){\rm dist}_{\partial\Omega}(p(x),\gamma(x))={\rm\;dist\;}_{\partial\Omega}(p(x),\Gamma), where distΩ{\rm dist}_{\partial\Omega} denotes the (geodesic) distance in Ω\partial\Omega ;

  • Γ\Gamma is parameterized by arc-length ss so that s=s0s=s_{0} defines x0x_{0}, and for x𝒩x0x\in\mathcal{N}_{x_{0}}, s=s(x)s=s(x) defines γ(x)\gamma(x) ;

  • For x𝒩x0x\in\mathcal{N}_{x_{0}}, the geodesic Λp(x)\Lambda_{p(x)} passing through p(x)p(x) is parameterized by arclength rr, so that r=0r=0 defines γ(x)\gamma(x) and r=r(x)r=r(x) defines p(x)p(x).

In this way, we observe that Φx0\Phi_{x_{0}}

𝒩x0xΦx0(x):=(r(x),s(x),t(x))××+\mathcal{N}_{x_{0}}\ni x\mapsto\Phi_{x_{0}}(x):=(r(x),s(x),t(x))\in\mathbb{R}\times\mathbb{R}\times\mathbb{R}_{+} (2.2)

is a local diffeomorphism. Thus, we can pick a sufficiently small ϵ0>0\epsilon_{0}>0 such that

(r,s,t)(ϵ0,ϵ0)×(ϵ0+s0,s0+ϵ0)×(0,ϵ0)x=Φx01(r,s,t)(r,s,t)\in(-\epsilon_{0},\epsilon_{0})\times(-\epsilon_{0}+s_{0},s_{0}+\epsilon_{0})\times(0,\epsilon_{0})\to x=\Phi_{x_{0}}^{-1}(r,s,t) (2.3)

is a diffeomorphism, whose image is a neighborhood of x0Γx_{0}\in\Gamma parameterized by (r,s,t)(r,s,t). Within these coordinates, t=0t=0 means that we are on Ω\partial\Omega, and r=t=0r=t=0 means we are on the curve Γ\Gamma. We can then compute

|dT(𝐁𝐍)(x)|=|r(𝐁𝐍)|r=0|(xΓ).|d^{T}(\mathbf{B}\cdot\mathbf{N})(x)|=|\partial_{r}(\mathbf{B}\cdot\mathbf{N})|_{r=0}|\quad(x\in\Gamma)\,. (2.4)

It is convenient to express the magnetic field along Γ\Gamma as follows

𝐁(x)=sinθ𝐓(x)+cosθ𝐕(x)(x=Φx01(0,s,0)Γ),\mathbf{B}(x)=\sin\theta\,\mathbf{T}(x)+\cos\theta\,\mathbf{V}(x)\quad\big{(}x=\Phi_{x_{0}}^{-1}(0,s,0)\in\Gamma\big{)}, (2.5)

where θ=θ(s)[π2,π2]\theta=\theta(s)\in[-\frac{\pi}{2},\frac{\pi}{2}] is the angle defined by

θ=arcsin(𝐁(x)𝐓(x)).\theta=\arcsin\big{(}\mathbf{B}(x)\cdot\mathbf{T}(x)\big{)}. (2.6)

2.2 The metric in the new coordinates

Let us consider an arbitrary point x0Γx_{0}\in\Gamma and a neighborhood 𝒩x0Ω¯\mathcal{N}_{x_{0}}\subset\overline{\Omega} of x0x_{0} such that the adapted coordinates introduced in (2.2) and (2.3) are valid. Modulo a translation, we can center the coordinates at x0x_{0} so that (r=0,s=0,t=0)(r=0,s=0,t=0) are the coordinates of x0x_{0} in the new frame. In the sequel, we follow closely the presentation of [HelMo4, Sec. 8] mainly following the first chapter of [DHKW] (see also the volume two of Spivak’s book [Sp]).

We label the new coordinates as follows

(y1,y2,y3)=(r,s,t),(y_{1},y_{2},y_{3})=(r,s,t)\,, (2.7)

and the Riemanian metric g0g_{0} becomes [HelMo4, Eq. (8.26)]

g0=dy3dy3+1i,j2[Gij2y3Kij+y32Lij]dyidyjg_{0}=dy_{3}\otimes dy_{3}+\sum_{1\leq i,j\leq 2}\big{[}G_{ij}-2y_{3}K_{ij}+y_{3}^{2}L_{ij}\big{]}dy_{i}\otimes dy_{j} (2.8)

where:

  • G:=1i,j2GijdyidyjG:=\sum\limits_{1\leq i,j\leq 2}G_{ij}dy_{i}\otimes dy_{j} is the first fundamental form on Ω\partial\Omega ;

  • K:=1i,j2KijdyidyjK:=\sum\limits_{1\leq i,j\leq 2}K_{ij}dy_{i}\otimes dy_{j} is the second fundamental form on Ω\partial\Omega ;

  • L:=1i,j2LijdyidyjL:=\sum\limits_{1\leq i,j\leq 2}L_{ij}dy_{i}\otimes dy_{j} is the third fundamental form on Ω\partial\Omega .

The matrix gg of the metric g0g_{0} takes the form

g:=(gij)1i,j3=(g11g120g21g220001)g:=(g_{ij})_{1\leq i,j\leq 3}=\left(\begin{array}[]{lll}g_{11}&g_{12}&0\\ g_{21}&g_{22}&0\\ 0&0&1\end{array}\right) (2.9)

whose inverse is

g1=(gij)1i,j3=(g11g120g21g220001).g^{-1}=(g^{ij})_{1\leq i,j\leq 3}=\left(\begin{array}[]{lll}g^{11}&g^{12}&0\\ g^{21}&g^{22}&0\\ 0&0&1\end{array}\right)\,. (2.10)

We will express these matrices in a more pleasant form involving, in particular, the curvatures on the boundary. To that end, let sγ(s)s\mapsto\gamma(s) be an arc-length parameterization of Γ\Gamma near x0x_{0}, so that |γ˙(s)|=1|\dot{\gamma}(s)|=1, γ(0)=x0\gamma(0)=x_{0} and 𝐓(γ(s))=γ˙(s)\mathbf{T}(\gamma(s))=\dot{\gamma}(s). We can introduce the geodesic and normal curvatures at γ(s)\gamma(s), κg(γ(s))\kappa_{g}(\gamma(s)) and κn(γ(s))\kappa_{n}(\gamma(s)), as follows

γ¨(s)=κg(γ(s))𝐕(γ(s))+κn(γ(s))𝐍(γ(s)).\ddot{\gamma}(s)=-\kappa_{g}(\gamma(s))\mathbf{V}(\gamma(s))+\kappa_{n}(\gamma(s))\mathbf{N}(\gamma(s))\,. (2.11)

The choice of our coordinates (r,s)(r,s) ensures that the metric GG is diagonal on Ω\partial\Omega [HelMo4, Lem. 8.2]

G=drdr+α(r,s)dsds,G=dr\otimes dr+\alpha(r,s)ds\otimes ds, (2.12)

with

α(r,s)=12κg(γ(s))r+𝒪(r2),α(0,s)=1,\alpha(r,s)=1-2\kappa_{g}\big{(}\gamma(s)\big{)}r+{\cal O}(r^{2})\;,\quad\alpha(0,s)=1\,, (2.13)

and

αs(0,s)=0.\frac{\partial\alpha}{\partial s}(0,s)=0\;. (2.14)

Then, with (2.7), we have for the determinant of the matrix of gg (see [HelMo4, Eq. (8.29) & (8.30)]),

|g|=α(r,s)2t[α(r,s)K11(r,s)+K22(r,s)]+t2ε3(r,s,t),|g|=\alpha(r,s)-2t\big{[}\alpha(r,s)K_{11}(r,s)+K_{22}(r,s)\big{]}+t^{2}\varepsilon_{3}(r,s,t)\,, (2.15)

and

(gij)1i,j2=(100α1(r,s))+2t(K11(r,s)α1K12(r,s)α1K21(r,s)α2K22(r,s))+t2R,(g^{ij})_{1\leq i,j\leq 2}=\left(\begin{array}[]{ll}1&0\\ 0&\alpha^{-1}(r,s)\end{array}\right)+2t\left(\begin{array}[]{ll}K_{11}(r,s)&\alpha^{-1}K_{12}(r,s)\\ \alpha^{-1}K_{21}(r,s)&\alpha^{-2}K_{22}(r,s)\end{array}\right)+t^{2}R\,, (2.16)

where ε3\varepsilon_{3} and RR are smooth functions.

2.3 The operator and quadratic form

We continue to work in the setting of Subsection 2.2 . We introduce the following neighborhood of x0x_{0}

Vx0=Φx01(V~x0),V_{x_{0}}=\Phi_{x_{0}}^{-1}(\tilde{V}_{x_{0}})\,, (2.17)

where (recall (2.7))

V~x0={(y1,y2,y3)(ϵ0,ϵ0)×(ϵ0,ϵ0)×(0,ϵ0)}.\tilde{V}_{x_{0}}=\{(y_{1},y_{2},y_{3})\in(-\epsilon_{0},\epsilon_{0})\times(-\epsilon_{0},\epsilon_{0})\times(0,\epsilon_{0})\}\,. (2.18)

Given a function u:Vx0u:V_{x_{0}}\to\mathbb{C}, we assign to it the function u~:Vx0\tilde{u}:V_{x_{0}}\to\mathbb{C} defined by

u~(y1,y2,y3)=u(Φx01(y1,y2,y3)).\tilde{u}(y_{1},y_{2},y_{3})=u(\Phi_{x_{0}}^{-1}(y_{1},y_{2},y_{3}))\,. (2.19)

By the considerations in Subsection 2.2 on the Riemanian metric, if uL2(Vx0,dx)u\in L^{2}(V_{x_{0}},dx), then u~L2(V~x0,|g|1/2dy)\tilde{u}\in L^{2}(\tilde{V}_{x_{0}},|g|^{1/2}dy) and

Vx0|u(x)|2𝑑x=V~x0|u~(y)|2|g|1/2𝑑y.\int_{V_{x_{0}}}|u(x)|^{2}dx=\int_{\tilde{V}_{x_{0}}}|\tilde{u}(y)|^{2}|g|^{1/2}dy\,. (2.20)

Moreover, assuming uu supported in Vx0V_{x_{0}}, we have the quadratic form formula [HelMo4, Eq. (8.27)]

q𝐀h(u)\displaystyle q_{{\bf A}}^{h}(u) :=Vx0|(hi𝐀)u|2𝑑x\displaystyle:=\int_{V_{x_{0}}}|(h\nabla-i{\bf A})u|^{2}dx (2.21)
=V~x0[|(hDy3A~3)u~|2+1i,j2gij(hDyiA~i)u~(hDyjA~j)u~¯]|g|1/2𝑑y\displaystyle=\int_{\tilde{V}_{x_{0}}}\Big{[}|(hD_{y_{3}}-\tilde{A}_{3})\tilde{u}|^{2}+\sum_{1\leq i,j\leq 2}g^{ij}(hD_{y_{i}}-\tilde{A}_{i})\tilde{u}\cdot\overline{(hD_{y_{j}}-\tilde{A}_{j})\tilde{u}}\Big{]}|\,g|^{1/2}dy

where the new magnetic potential 𝐀~=(A~1,A~2,A~3)\tilde{\bf A}=(\tilde{A}_{1},\tilde{A}_{2},\tilde{A}_{3}) is assigned to 𝐀=(A1,A2,A3){\bf A}=(A_{1},A_{2},A_{3}) by the relation

A1dx1+A2dx2+A3dx3=A~1dy1+A~2dy2+A~3dy3,A_{1}dx_{1}+A_{2}dx_{2}+A_{3}dx_{3}=\tilde{A}_{1}dy_{1}+\tilde{A}_{2}dy_{2}+\tilde{A}_{3}dy_{3}\,, (2.22)

and after performing a (local) gauge transformation, we may assume that

A~3=0.\tilde{A}_{3}=0\,. (2.23)

The operator P𝐀hP_{{\bf A}}^{h} in (1.3) can be expressed in the new coordinates as follows [HelMo4, Eq. (8.28)]

P𝐀h=(hDy3A~3)2+h2i|g|1y3|g|(hDy3A~3)+|g|1/21i,j2(hDyjA~j)|g|1/2gij(hDyiA~i).P_{{\bf A}}^{h}=(hD_{y_{3}}-\tilde{A}_{3})^{2}+\frac{h}{2i}|g|^{-1}\frac{\partial}{\partial y_{3}}|g|(hD_{y_{3}}-\tilde{A}_{3})\\ +|g|^{-1/2}\sum_{1\leq i,j\leq 2}(hD_{y_{j}}-\tilde{A}_{j})|g|^{1/2}g^{ij}(hD_{y_{i}}-\tilde{A}_{i})\,. (2.24)

3 Helical magnetic fields

3.1 Preliminaries

Let τ>0\tau>0 and consider the magnetic potential

𝐀(x)=𝐧τ(x):=(1τcos(τx3),1τsin(τx3),0),{\bf A}(x)=\mathbf{n}_{\tau}(x):=\Big{(}\frac{1}{\tau}\cos(\tau x_{3}),\frac{1}{\tau}\sin(\tau x_{3}),0\Big{)}\,, (3.1)

which generates the magnetic field

𝐁(x)=curl𝐀(x)=τ𝐀(x)\mathbf{B}(x)={\rm curl}\,{\bf A}(x)=-\tau{\bf A}(x) (3.2)

with constant intensity

|𝐁(x)|=1.|\mathbf{B}(x)|=1\,. (3.3)

We will verify that Assumptions C1-C2 hold for this particular magnetic field in the case where Ω\Omega is the unit ball. In particular, with in mind that γ^0,𝐁\widehat{\gamma}_{0,\mathbf{B}} and γ~0,𝐁\widetilde{\gamma}_{0,\mathbf{B}} are introduced in (1.6) and (1.7) respectively and that δ0]0,1[\delta_{0}\in]0,1[ and ν^0>0\widehat{\nu}_{0}>0 will be introduced in (4.2) and in (4.5) (there is no need in this subsection to know more about them) we will find that

γ^0,𝐁=22/3ν^0δ01/3C(τ,δ0),\widehat{\gamma}_{0,\mathbf{B}}=2^{-2/3}\widehat{\nu}_{0}\delta_{0}^{1/3}C(\tau,\delta_{0})\,, (3.4)

and for ττ0\tau\leq\tau_{0}, the equality,

{xΓ|γ~0,𝐁(x)=22/3ν^0δ01/3}={(0,±1,0)},\{x\in\Gamma~{}|~{}\widetilde{\gamma}_{0,\mathbf{B}}(x)=2^{-2/3}\widehat{\nu}_{0}\delta_{0}^{1/3}\}=\{(0,\pm 1,0)\}\,, (3.5)

where τ0\tau_{0} is a constant and C(τ,δ0)C(\tau,\delta_{0}) is explicitly computed (see Proposition 3.4).

The inward normal of Ω={x3||x|<1}\Omega=\{x\in\mathbb{R}^{3}~{}|~{}|x|<1\} along Ω\partial\Omega is

𝐍(x)=x(|x|=1).\mathbf{N}(x)=-x\quad(|x|=1)\,. (3.6)

The restriction of the magnetic field 𝐁\mathbf{B} to the boundary is then tangent to Ω\partial\Omega on the following set

Γ={xΩ|x𝐀(x)=0}.\Gamma=\{x\in\partial\Omega~{}|~{}x\cdot{\bf A}(x)=0\}\,. (3.7)

3.2 Γ\Gamma is a regular curve

For |x|=1|x|=1, the equation x𝐀(x)=0x\cdot{\bf A}(x)=0 reads as follows

x1cos(τx3)+x2sin(τx3)=0.x_{1}\cos(\tau x_{3})+x_{2}\sin(\tau x_{3})=0\,. (3.8)
Proposition 3.1.

The set Γ\Gamma introduced in (3.7) is a CC^{\infty} regular curve.

Proof.

The proof follows by constructing an atlas on Γ\Gamma,

{(𝐜i,U:=(1,1)),1i4}\{\big{(}\mathbf{c}_{i},U:=(-1,1)\big{)},~{}1\leq i\leq 4\}

which turns Γ\Gamma to a CC^{\infty} regular curve.

Let us introduce the charts (𝐜1,U)(\mathbf{c}_{1},U) and (𝐜2,U)(\mathbf{c}_{2},U) which cover Γ{(0,0,±1)}\Gamma\setminus\{(0,0,\pm 1)\}. These charts are obtained by expressing x1x_{1} and x2x_{2} in (3.8) in terms of x3(1,1)x_{3}\in(-1,1), provided that (x1,x2,x3)(0,0,±1)(x_{1},x_{2},x_{3})\not=(0,0,\pm 1). We write for α]π,π]\alpha\in]-\pi,\pi]

x1=1x32cosα,x2=1x32sinα.x_{1}=\sqrt{1-x_{3}^{2}}\,\cos\alpha,\quad x_{2}=\sqrt{1-x_{3}^{2}}\,\sin\alpha\,. (3.9)

Then (3.8) becomes, for x32<1x_{3}^{2}<1,

cos(τx3α)=0\cos(\tau x_{3}-\alpha)=0 (3.10)

which in turn yields

α=τx3π2+kπ,k.\alpha=\tau x_{3}-\frac{\pi}{2}+k\pi,\quad k\in\mathbb{Z}\,.

In this way, we get two branches of Γ\Gamma parameterized by x3x_{3} and defined as follows

x3(1,1)𝐜1(x3):=(x1=1x32sin(τx3)x2=1x32cos(τx3)x3)x_{3}\in(-1,1)\mapsto\mathbf{c}_{1}(x_{3}):=\left(\begin{array}[]{c}x_{1}=\sqrt{1-x_{3}^{2}}\,\sin(\tau x_{3})\\ x_{2}=-\sqrt{1-x_{3}^{2}}\,\cos(\tau x_{3})\\ x_{3}\end{array}\right)

and

x3(1,1)𝐜2(x3):=(x1=1x32sin(τx3)x2=1x32cos(τx3)x3).x_{3}\in(-1,1)\mapsto\mathbf{c}_{2}(x_{3}):=\left(\begin{array}[]{c}x_{1}=-\sqrt{1-x_{3}^{2}}\sin(\tau x_{3})\\ x_{2}=\sqrt{1-x_{3}^{2}}\cos(\tau x_{3})\\ x_{3}\end{array}\right)\,.

Both of the foregoing branches represent regular curves. Furthermore, 𝐜1\mathbf{c}_{1} and 𝐜2\mathbf{c}_{2} can be extended by continuity to the interval [1,1][-1,1], yielding a continuous representation of all Γ\Gamma.

Now we introduce the charts (𝐜3,U)(\mathbf{c}_{3},U) and 𝐜4,U)\mathbf{c}_{4},U) that cover the points (0,0,±1)(0,0,\pm 1). In a neighborhood of (x1,x2,x3)=(0,0,±1)(x_{1},x_{2},x_{3})=(0,0,\pm 1), we parameterize a branch of Γ\Gamma with respect to ρ:=x12+x22\rho:=\sqrt{x_{1}^{2}+x_{2}^{2}} as follows

x1=ρcosα,x2=ρsinα,x3=1ρ2.x_{1}=\rho\cos\alpha\,,~{}x_{2}=\rho\sin\alpha\,,~{}x_{3}=\sqrt{1-\rho^{2}}\,.

With this in hand, (3.10) continues to hold for x30x_{3}\not=0 and we can write again α=τx3π2+kπ\alpha=\tau x_{3}-\frac{\pi}{2}+k\pi for some kk\in\mathbb{Z}. Consequently, we get two regular branches of Γ\Gamma defined as follows

ρ(1,1)𝐜3(ρ):=(x1=ρsin(τ1ρ2)x2=ρcos(τ1ρ2)x3=1ρ2)\rho\in(-1,1)\mapsto\mathbf{c}_{3}(\rho):=\left(\begin{array}[]{rl}x_{1}&=\rho\sin(\tau\sqrt{1-\rho^{2}})\\ x_{2}&=-\rho\cos(\tau\sqrt{1-\rho^{2}})\\ x_{3}&=\sqrt{1-\rho^{2}}\end{array}\right)

and

ρ(1,1)𝐜4(ρ):=(x1=ρsin(τ1ρ2)x2=ρcos(τ1ρ2)x3=1ρ2).\rho\in(-1,1)\mapsto\mathbf{c}_{4}(\rho):=\left(\begin{array}[]{rl}x_{1}&=-\rho\sin(\tau\sqrt{1-\rho^{2}})\\ x_{2}&=\rho\cos(\tau\sqrt{1-\rho^{2}})\\ x_{3}&=\sqrt{1-\rho^{2}}\end{array}\right)\,.

 

3.3 Explicit formulas in adapted coordinates

Note that 𝐜:=𝐜1\mathbf{c}:=\mathbf{c}_{1} and 𝐜2\mathbf{c}_{2} parameterize all of Γ{(0,0,±1)}\Gamma\setminus\{(0,0,\pm 1)\}. By symmetry considerations, we will compute, on 𝐜((1,1))\mathbf{c}\big{(}\,(-1,1)\,\big{)} only,

|dT(𝐁𝐍)|=τ|dT(𝐀𝐍)|and|𝐁𝐓|=τ|𝐀𝐓|.|d^{T}(\mathbf{B}\cdot\mathbf{N})|=\tau|d^{T}(\mathbf{A}\cdot\mathbf{N})|\quad{\rm and}\quad|\mathbf{B}\cdot\mathbf{T}|=\tau|\mathbf{A}\cdot\mathbf{T}|\,. (3.11)

First we note that 𝐍=x\mathbf{N}=-x on Ω\partial\Omega and introduce the arc-length parameter

s(x3)=0x3|𝐜(x~3)|𝑑x~3s(x_{3})=\int_{0}^{x_{3}}|\mathbf{c}^{\prime}(\tilde{x}_{3})|d\tilde{x}_{3}

of x3𝐜(x3)x_{3}\mapsto\mathbf{c}(x_{3}), which satisfies

s(x3)=|𝐜(x3)|=1+τ2(1x32)21x32.s^{\prime}(x_{3})=|\mathbf{c}^{\prime}(x_{3})|=\sqrt{\frac{1+\tau^{2}(1-x_{3}^{2})^{2}}{1-x_{3}^{2}}}\,. (3.12)

Clearly, x3(1,1)x_{3}\in(-1,1) can be expressed in terms of the arc-length parameter as x3=x3(s)x_{3}=x_{3}(s) with

m(x3):=dx3ds(s(x3))=1x321+τ2(1x32)2.m(x_{3}):=\frac{dx_{3}}{ds}(s(x_{3}))=\sqrt{\frac{1-x_{3}^{2}}{1+\tau^{2}(1-x_{3}^{2})^{2}}}\,. (3.13)

The arc-length parameterization is now given by

γ(s):=𝐜(x3(s)),\gamma(s):=\mathbf{c}(x_{3}(s))\,, (3.14)

and consequently, with 𝐜=𝐜1\mathbf{c}=\mathbf{c}_{1}, we have

𝐍(γ(s))=γ(s)=(1x32sin(τx3)1x32cos(τx3)x3)withx3=x3(s),\mathbf{N}(\gamma(s))=-\mathbf{\gamma}(s)=\left(\begin{array}[]{c}-\sqrt{1-x_{3}^{2}}\sin(\tau x_{3})\\ \sqrt{1-x_{3}^{2}}\cos(\tau x_{3})\\ -x_{3}\end{array}\right)~{}{\rm with~{}}x_{3}=x_{3}(s)\,, (3.15)

and

𝐓(γ(s))\displaystyle\mathbf{T}(\gamma(s)) =ddsγ(s)=(T1T2T3)\displaystyle=\frac{d}{ds}\gamma(s)=\left(\begin{array}[]{c}T_{1}\\ T_{2}\\ T_{3}\end{array}\right)
=m(x3)(x3sin(τx3)1x32+τ1x32cos(τx3)x3cos(τx3)1x32+τ1x32sin(τx3)1).\displaystyle=m(x_{3})\left(\begin{array}[]{c}-\frac{x_{3}\sin(\tau x_{3})}{\sqrt{1-x_{3}^{2}}}+\tau\sqrt{1-x_{3}^{2}}\cos(\tau x_{3})\\ \frac{x_{3}\cos(\tau x_{3})}{\sqrt{1-x_{3}^{2}}}+\tau\sqrt{1-x_{3}^{2}}\sin(\tau x_{3})\\ 1\end{array}\right)\,.

We also introduce the normal vector to Γ\Gamma on γ(s)\gamma(s),

𝐕(γ(s))\displaystyle\mathbf{V}(\gamma(s)) =𝐓(γ(s))×𝐍(γ(s))=(V1V2V3)\displaystyle=\mathbf{T}(\gamma(s))\times\mathbf{N}(\gamma(s))=\left(\begin{array}[]{c}V_{1}\\ V_{2}\\ V_{3}\end{array}\right)
=m(x3)(x32cos(τx3)1x32τx31x32sin(τx3)1x32cos(τx3)x32sin(τx3)1x32+τx31x32cos(τx3)1x32sin(τx3)τ(1x32)).\displaystyle=m(x_{3})\left(\begin{array}[]{c}-\frac{x_{3}^{2}\cos(\tau x_{3})}{\sqrt{1-x_{3}^{2}}}-\tau x_{3}\sqrt{1-x_{3}^{2}}\sin(\tau x_{3})-\sqrt{1-x_{3}^{2}}\cos(\tau x_{3})\\ -\frac{x_{3}^{2}\sin(\tau x_{3})}{\sqrt{1-x_{3}^{2}}}+\tau x_{3}\sqrt{1-x_{3}^{2}}\cos(\tau x_{3})-\sqrt{1-x_{3}^{2}}\sin(\tau x_{3})\\ \tau(1-x_{3}^{2})\end{array}\right)\,.

We are now ready to prove that our magnetic field 𝐁\mathbf{B} verifies the condition (C2) appearing in Assumption 1.2

Proposition 3.2.

Let 𝐁\mathbf{B} be the magnetic field introduced in (3.2). For all xΓx\in\Gamma, we have

|𝐁(x)𝐓(x)|=τ(1x32)1+τ2(1x32)2.|\mathbf{B}(x)\cdot\mathbf{T}(x)|=\frac{\tau(1-x_{3}^{2})}{\sqrt{1+\tau^{2}(1-x_{3}^{2})^{2}}}\,.

In particular, 𝐁\mathbf{B} satisfies the condition (C2)\rm(C2).

Proof.

It is straightforward to compute

|𝐀(x)𝐓(x)|=1τ(|cos(τx3)T1+sin(τx3)T2|)=1x321+τ2(1x32)2,|\mathbf{A}(x)\cdot\mathbf{T}(x)|=\frac{1}{\tau}(|\cos(\tau x_{3})T_{1}+\sin(\tau x_{3})T_{2}|)=\frac{1-x_{3}^{2}}{\sqrt{1+\tau^{2}(1-x_{3}^{2})^{2}}}\,, (3.16)

which holds for all 1x3<1-1\leq x_{3}<1 and x=𝐜(x3)x=\mathbf{c}(x_{3}). Similarly, we can compute |𝐀(x)𝐓(x)||\mathbf{A}(x)\cdot\mathbf{T}(x)| for all x=𝐜2(x3)Γx=\mathbf{c}_{2}(x_{3})\in\Gamma, and get that (3.16) holds globally on Γ\Gamma, since Γ\Gamma is a regular curve. Finally, 𝐁(x)\mathbf{B}(x) is orthogonal to 𝐓(x)\mathbf{T}(x) if and only if x32=1x_{3}^{2}=1, thereby (C2) holds.  

Our next task is to show that our magnetic field satisfies the condition (C1) in Assumption 1.1.

Proposition 3.3.

Let 𝐁\mathbf{B} be the magnetic field introduced in (3.2). For all xΓx\in\Gamma, we have

κn,𝐁(x)=1+τ2(1x3(s)2)2.\kappa_{n,{\mathbf{B}}}(x)=\sqrt{1+\tau^{2}(1-x_{3}(s)^{2})^{2}}\,. (3.17)

In particular, 𝐁\mathbf{B} satisfies the condition (C1)\rm(C1).

Proof.

By Proposition 3.1, Γ\Gamma is a regular curve. So all we need to verify that 𝐁\mathbf{B} satisfies (C1)\rm(C1), is to derive (3.17) and observe that it yields κn,𝐁(x)0\kappa_{n,{\mathbf{B}}}(x)\not=0 everywhere on Γ\Gamma.

Consider x=𝐜1(x3)x=\mathbf{c}_{1}(x_{3}) with x3=x3(s)x_{3}=x_{3}(s), i.e. x=γ(s)x=\gamma(s). At the point γ(s)\gamma(s), the geodesic Λγ(s)\Lambda_{\gamma(s)} normal to the curve Γ\Gamma is the great circle (of center 0 and radius 11) in the (𝐕(γ(s)),𝐍(γ(s))(\mathbf{V}(\gamma(s)),\mathbf{N}(\gamma(s)) plane. A point P=P(r,s)P=P(r,s) on Λγ(s)\Lambda_{\gamma(s)} can be described by the corresponding vector 𝐩(r,s)=OP\mathbf{p}(r,s)=\vbox{\halign{#\cr\tiny\rightarrowfill\cr\nointerlineskip\vskip 1.0pt\cr$OP\mskip 2.0mu$\cr}} as follows

𝐩(r,s)=cosr𝐍(γ(s))sinr𝐕(γ(s)),\mathbf{p}(r,s)=-\cos r\,\mathbf{N}(\gamma(s))-\sin r\,\mathbf{V}(\gamma(s))\,,

where rr is the angle between 𝐩\mathbf{p} and 𝐍-\mathbf{N}; hence rr is an arc-length length parameter of Λγ(s)\Lambda_{\gamma(s)}, and for r=0r=0, p(r,s)=γ(s)p(r,s)=\gamma(s).

Refer to caption
Figure 1: The curve Γ\Gamma and the geodesic Λγ(s)\Lambda_{\gamma(s)} passing through γ(s)\gamma(s).

Now, we can introduce the coordinates (r,s,t)(r,s,t) in a neighborhood of γ(s0)\gamma(s_{0}) as follows (see Fig. 1)

x(r,s,t)=(cosr+t)𝐍(γ(s))sinr𝐕(γ(s)).x(r,s,t)=-(\cos r+t)\mathbf{N}(\gamma(s))-\sin r\mathbf{V}(\gamma(s))\,. (3.18)

For x=γ(s)x=\gamma(s), we would like to compute κn,𝐁(x)=|dT(𝐁𝐍)|\kappa_{n,\mathbf{B}}(x)=|d^{T}(\mathbf{B}\cdot\mathbf{N})|. We will show that κn,𝐁(x)=r(𝐁𝐍)|r=t=0\kappa_{n,\mathbf{B}}(x)=\partial_{r}(\mathbf{B}\cdot\mathbf{N})|_{r=t=0} and end up with the computation of |r(𝐁𝐍)|r=t=0.\big{|}\partial_{r}(\mathbf{B}\cdot\mathbf{N})|_{r=t=0}\,.

Notice that, by (3.15), we have

x3(r,s,t)=(cosr+t)𝐍3(γ(s))sinr𝐕3(γ(s))=(cosr+t)x3(s)sinrm(x3(s))τ(1x3(s)2),\begin{array}[]{ll}x_{3}(r,s,t)&=-(\cos r+t)\mathbf{N}_{3}(\gamma(s))-\sin r\mathbf{V}_{3}(\gamma(s))\\ &=(\cos r+t)x_{3}(s)-\sin r\,m(x_{3}(s))\tau(1-x_{3}(s)^{2})\,,\end{array}

and we observe that by (3.18),

xr|r=t=0=V(γ(s)).\frac{\partial x}{\partial r}\,\big{|}_{r=t=0}=-V(\gamma(s))\,. (3.19)

In particular we have

x3r|r=t=0=τ(1x32(s))m(x3(s)).\frac{\partial x_{3}}{\partial r}\,\big{|}_{r=t=0}=-\tau(1-x_{3}^{2}(s))m(x_{3}(s))\,.

Now, using (3.13) and (3.15), we get from (3.1) that

𝐀r𝐍|r=t=0=τ(1x32)21+τ2(1x32)2.\frac{\partial{\bf A}}{\partial r}\cdot\mathbf{N}\,\big{|}_{r=t=0}=-\frac{\tau(1-x_{3}^{2})^{2}}{\sqrt{1+\tau^{2}(1-x_{3}^{2})^{2}}}\,. (3.20)

Moreover, by (3.19) we have

r𝐍(x(r,s,t))|r=t=0=𝐕(γ(s))\frac{\partial}{\partial r}\mathbf{N}(x(r,s,t))\,\big{|}_{r=t=0}=\mathbf{V}(\gamma(s))

and

𝐀r𝐍(x(r,s,t))|r=t=0\displaystyle\mathbf{A}\cdot\frac{\partial}{\partial r}\mathbf{N}(x(r,s,t))\,\big{|}_{r=t=0} =1τcos(τx3(s))V1+1τsin(τx3(s))V2\displaystyle=\frac{1}{\tau}\cos(\tau x_{3}(s))V_{1}+\frac{1}{\tau}\sin(\tau x_{3}(s))V_{2} (3.21)
=1τ1+τ2(1x32)2.\displaystyle=-\frac{1}{\tau\sqrt{1+\tau^{2}(1-x_{3}^{2})^{2}}}\,.

Summing up, we deduce from (3.20) and (3.21) that

|r(𝐀𝐍)|r=t=0|=1τ1+τ2(1x32)2.\big{|}\partial_{r}(\mathbf{A}\cdot\mathbf{N})\,|_{r=t=0}\big{|}=\frac{1}{\tau}\sqrt{1+\tau^{2}(1-x_{3}^{2})^{2}}\,. (3.22)

We also observe that s(𝐀𝐍)|r=t=0=0\partial_{s}(\mathbf{A}\cdot\mathbf{N})\,|_{r=t=0}=0 and we get

|dT(𝐀𝐍)|(γ(s))=1τ1+τ2(1x32)2,|d^{T}(\mathbf{A}\cdot\mathbf{N})|(\gamma(s))=\frac{1}{\tau}\sqrt{1+\tau^{2}(1-x_{3}^{2})^{2}}\,, (3.23)

on each branch (including the end points). Inserting this into (3.11), we get the identity in (3.17).  

We return to the function in (1.7) and can give its expression in coordinates. We deduce from (3.16) and (3.17):

γ~0,𝐁(x)=22/3ν^0δ01/3(1+τ2(1x32)2)1/3(1(1δ0)τ(1x32)1+τ2(1x32)2)1/3\widetilde{\gamma}_{0,\mathbf{B}}(x)=2^{-2/3}\widehat{\nu}_{0}\delta_{0}^{1/3}\left(1+\tau^{2}(1-x_{3}^{2})^{2}\right)^{1/3}\left(1-(1-\delta_{0})\frac{\tau(1-x_{3}^{2})}{\sqrt{1+\tau^{2}(1-x_{3}^{2})^{2}}}\right)^{1/3}

for all x=(±1x32sin(τx3),1x32cos(τx3),x3)x=(\pm\sqrt{1-x_{3}^{2}}\sin(\tau x_{3}),\mp\sqrt{1-x_{3}^{2}}\cos(\tau x_{3}),x_{3}) with 1x31-1\leq x_{3}\leq 1.

Consequently, we can compute the quantity appearing in the two terms asymptotics by computing infxΓγ~0,𝐁(x)\inf_{x\in\Gamma}\widetilde{\gamma}_{0,\mathbf{B}}(x) and determining where the infimum is attained.

Proposition 3.4.

Let

τ0=12(1δ0+δ0(1δ0)1)1/2.\tau_{0}=\frac{1}{\sqrt{2}}\left(\frac{1}{\sqrt{\delta_{0}+\delta_{0}(1-\delta_{0})}}-1\right)^{1/2}\,.

The following holds:

  1. 1.

    If 0<ττ00<\tau\leq\tau_{0}, then

    infxΓγ~0,𝐁(x)=22/3ν^0δ01/3(1+τ2)1/3(1(1δ0)τ1/3(1+τ2)1/6)=γ~0,𝐁(0,±1,0).\inf_{x\in\Gamma}\widetilde{\gamma}_{0,\mathbf{B}}(x)=2^{-2/3}\widehat{\nu}_{0}\delta_{0}^{1/3}(1+\tau^{2})^{1/3}\Big{(}1-(1-\delta_{0})\frac{\tau^{1/3}}{(1+\tau^{2})^{1/6}}\Big{)}=\widetilde{\gamma}_{0,\mathbf{B}}(0,\pm 1,0)\,.
  2. 2.

    If τ>τ0\tau>\tau_{0}, then

    infxΓγ~0,𝐁(x)=22/3ν^0δ01/3(1+τ02)1/3(1(1δ0)τ01/3(1+τ02)1/6),\inf_{x\in\Gamma}\widetilde{\gamma}_{0,\mathbf{B}}(x)=2^{-2/3}\widehat{\nu}_{0}\delta_{0}^{1/3}(1+\tau_{0}^{2})^{1/3}\Big{(}1-(1-\delta_{0})\frac{\tau_{0}^{1/3}}{(1+\tau_{0}^{2})^{1/6}}\Big{)}\,,

    and the minimum is attained on the points

    (±τ0τsinτ1τ0τ,τ0τcosτ1τ0τ,1τ0τ)\Big{(}\,\pm\sqrt{\frac{\tau_{0}}{\tau}}\sin\tau\sqrt{1-\frac{\tau_{0}}{\tau}},\mp\sqrt{\frac{\tau_{0}}{\tau}}\cos\tau\sqrt{1-\frac{\tau_{0}}{\tau}},\sqrt{1-\frac{\tau_{0}}{\tau}}\,\Big{)}

    and

    (±τ0τsinτ1τ0τ,±τ0τcosτ1τ0τ,1τ0τ).\Big{(}\,\pm\sqrt{\frac{\tau_{0}}{\tau}}\sin\tau\sqrt{1-\frac{\tau_{0}}{\tau}},\pm\sqrt{\frac{\tau_{0}}{\tau}}\cos\tau\sqrt{1-\frac{\tau_{0}}{\tau}},-\sqrt{1-\frac{\tau_{0}}{\tau}}\,\Big{)}\,.
Remark 3.5.

In the case where Ω=B(0,1)\Omega=B(0,1) is the unit ball and the magnetic field is constant, 𝐁=(0,0,1)\mathbf{B}=(0,0,1), we have Γ={x12+x22=1,x3=0}\Gamma=\{x_{1}^{2}+x_{2}^{2}=1,~{}x_{3}=0\} and γ~0,𝐁(x)\widetilde{\gamma}_{0,\mathbf{B}}(x) is constant on Γ\Gamma. Proposition 3.4 shows a quite different phenomenon when only the intensity of 𝐁\mathbf{B} is constant, |𝐁|=1|\mathbf{B}|=1. In fact, γ~0,𝐁(x)\widetilde{\gamma}_{0,\mathbf{B}}(x) is no more constant along Γ\Gamma and may have two symmetric minimum points, (0,±1,0)(0,\pm 1,0), which is the signature of an interesting double well tunnel effect [HeSj] related to the magnetic geometry of the problem.

Proof of Proposition 3.4.

Let us introduce v=τ(1x32)[0,τ]v=\tau(1-x_{3}^{2})\in[0,\tau] and μ0=1δ0(0,1)\mu_{0}=1-\delta_{0}\in(0,1). Then

γ~0,𝐁(x)=22/3ν^0δ01/3(f(v))1/3\widetilde{\gamma}_{0,\mathbf{B}}(x)=2^{-2/3}\widehat{\nu}_{0}\delta_{0}^{1/3}\big{(}f(v)\big{)}^{1/3}

where

f(v)=1+v2μ0v1+v2.f(v)=1+v^{2}-\mu_{0}v\sqrt{1+v^{2}}\,.

We have to minimize f(v)f(v) on [0,τ][0,\tau]. Notice that

f(v)=2vμ01+2v21+v2,f^{\prime}(v)=2v-\mu_{0}\frac{1+2v^{2}}{\sqrt{1+v^{2}}}\,,

and the equation f(v)=0f^{\prime}(v)=0 has a unique positive solution, which is the solution of

v4+v2=μ024(1μ02).v^{4}+v^{2}=\frac{\mu_{0}^{2}}{4(1-\mu_{0}^{2})}\,.

This solution is given by

τ0=12μ01+1μ021μ02\tau_{0}=\frac{1}{\sqrt{2}}\frac{\mu_{0}}{\sqrt{1+\sqrt{1-\mu_{0}^{2}}}\sqrt{1-\mu_{0}^{2}}}

and observe that f(v)<0f^{\prime}(v)<0 for 0<v<τ00<v<\tau_{0} and f(v)>0f^{\prime}(v)>0 for v>τ0v>\tau_{0}. Then, for ττ0\tau\leq\tau_{0}\,,

minv[0,τ]f(v)=f(τ),\min_{v\in[0,\tau]}f(v)=f(\tau)\,,

while for τ>τ0\tau>\tau_{0}\,,

minv[0,τ]f(v)=f(τ0).\min_{v\in[0,\tau]}f(v)=f(\tau_{0})\,.

 

4 1D Models

The aim of this section is to recall the now standard properties of two important models.

4.1 The de Gennes model

We refer to [DaHe, HelMo2] for the proof of these now standard properties which are presented below. For ξ\xi\in\mathbb{R}, we consider the harmonic oscillator on +\mathbb{R}_{+}:

H(ξ):=Dt2+(tξ)2,H(\xi):=D_{t}^{2}+(t-\xi)^{2}\,, (4.1)

with Neumann boundary condition at 0. We denote by μ(ξ)\mu(\xi) its lowest eigenvalue. ξμ(ξ)\xi\mapsto\mu(\xi) admits a unique minimum at a point ξ0\xi_{0} which in addition is non-degenerate. This leads to introduce the spectral constants, Θ0\Theta_{0} and δ0\delta_{0}:

Θ0=infξμ(ξ)=μ(ξ0),δ0=μ′′(ξ0),\Theta_{0}=\inf_{\xi\in\mathbb{R}}\mu(\xi)=\mu(\xi_{0}),\quad\delta_{0}=\mu^{\prime\prime}(\xi_{0})\,, (4.2)

where ξ0=Θ0\xi_{0}=\sqrt{\Theta_{0}}.
Moreover 12<Θ0<1\frac{1}{2}<\Theta_{0}<1 and that 0<δ0<10<\delta_{0}<1. Θ0\Theta_{0} is called the de Gennes constant.
If φ0L2(+)\varphi_{0}\in L^{2}(\mathbb{R}_{+}) denotes the positive and normalized ground state of H(ξ0)H(\xi_{0}),

+(tξ0)|φ0(t)|2𝑑t=0,\int_{\mathbb{R}_{+}}(t-\xi_{0})|\varphi_{0}(t)|^{2}dt=0\,, (4.3)

which amounts to saying, via the Feynman-Hellmann formula, that μ(ξ0)=0\mu^{\prime}(\xi_{0})=0. We also introduce the regularized resolvent 0(L2(+))\mathcal{R}_{0}\in\mathcal{L}(L^{2}(\mathbb{R}_{+})) as follows

0u={(H(ξ0)Θ0)1uifuφ00ifuφ0.\mathcal{R}_{0}u=\begin{cases}\big{(}H(\xi_{0})-\Theta_{0}\big{)}^{-1}u&{\rm if}~{}u\perp\varphi_{0}\\ 0&{\rm if}~{}u\parallel\varphi_{0}\end{cases}\,. (4.4)

4.2 The Montgomery model

Here we refer to [HelMo1] and [PanKw]. In Theorem 1.3, the constant ν^0>0\widehat{\nu}_{0}>0 is related to the Montgomery model [Mon] whose spectral analysis has a long story including recently (see [HeLe] and references therein). For ρ\rho\in\mathbb{R}, we introduce, in L2()L^{2}(\mathbb{R}), the operator

S(ρ)=Dr2+(r2ρ)2,S(\rho)=D_{r}^{2}+(r^{2}-\rho)^{2}\,,

and denote its lowest eigenvalue by μMon(ρ)\mu^{\rm Mon}(\rho). Then

ν^0:=infρμMon(ρ)=μMon(ρ0),\widehat{\nu}_{0}:=\inf_{\rho\in\mathbb{R}}\mu^{\rm Mon}(\rho)=\mu^{\rm Mon}(\rho_{0})\,, (4.5)

where ρ0\rho_{0}\in\mathbb{R} is the unique minimum of μMon\mu^{\rm Mon}, which has been later shown to be non degenerate [HeKo]. Finally, the normalized positive ground state ψ0L2()\psi_{0}\in L^{2}(\mathbb{R}) of S(ρ0)S(\rho_{0}) belongs to the Schwartz space S()S(\mathbb{R}) and is an even function.

5 Model operator for non-uniform magnetic fields

Given real parameters η,ζ,γ\eta,\zeta,\gamma and θ\theta, we consider the operator

P0;γ,θh,η,ζ:=(hDrsinθtcosθ(ηs+ζr)t)2+(hDs+cosθtsinθ(ηs+ζr)t+γr22)2+h2Dt2,\begin{array}[]{ll}P^{h,\eta,\zeta}_{0;\gamma,\theta}&:=(hD_{r}-\sin\theta\,t-\cos\theta\,(\eta s+\zeta r)t)^{2}\\ &\quad+(hD_{s}+\cos\theta\,t-\sin\theta\,(\eta s+\zeta r)\,t+\gamma\frac{r^{2}}{2})^{2}\\ &\quad+h^{2}D_{t}^{2}\;,\end{array} (5.1)

on 2×+\mathbb{R}^{2}\times\mathbb{R}^{+} (actually in a neighborhood of (0,0,0)(0,0,0)). Let us fix a positive constant MM. We assume that

η,ζ,γ[M,M].\eta,\zeta,\gamma\in[-M,M]\,. (5.2)

We note, when η=ζ=0\eta=\zeta=0, we recover the model studied in [HelMo4, Sec. 11]. Our aim is to compare this situation with that when η=ζ=0\eta=\zeta=0. Our main result on this model is Proposition 5.5 below, which is useful in our derivation of the lower bound matching with the asymptotics in Theorem 1.4. The lower bound in this proposition is uniform with respect to the various parameters appearing in (5.1) provided (5.2) holds and hh is sufficiently small.

Let us look at this model more carefully. We proceed essentially like in the case η=ζ=0\eta=\zeta=0. We do the following scaling

r=h13r^,s=h13s^,t=h12t^.r=h^{\frac{1}{3}}\hat{r}\;,\;s=h^{\frac{1}{3}}\hat{s}\;,\;t=h^{\frac{1}{2}}\hat{t}\;. (5.3)

After division by hh, this leads to (forgetting the hats)

P1;γ,θh,η,ζ:=(h16Drsinθth13cosθt(ηs+ζr))2+(h16Ds+cosθt+h16γr22h13sinθt(ηs+ζr))2+Dt2\begin{array}[]{l}P_{1;\gamma,\theta}^{h,\eta,\zeta}:=\left(h^{\frac{1}{6}}D_{r}-\sin\theta\,t-h^{\frac{1}{3}}\cos\theta\,t(\eta s+\zeta r)\right)^{2}\\ \quad\quad+\left(h^{\frac{1}{6}}D_{s}+\cos\theta\,t+h^{\frac{1}{6}}\gamma\frac{r^{2}}{2}-h^{\frac{1}{3}}\sin\theta\,t(\eta s+\zeta r)\right)^{2}+D_{t}^{2}\end{array} (5.4)

on ××+\mathbb{R}\times\mathbb{R}\times\mathbb{R}^{+}.
Hence we have

σ(P0;γ,θh,η,ζ)=hσ(P1;γ,θh,η,ζ).\sigma(P_{0;\gamma,\theta}^{h,\eta,\zeta})=h\,\sigma(P_{1;\gamma,\theta}^{h,\eta,\zeta}).

Unlike the case where η=ζ=0\eta=\zeta=0, we can no more perform a partial Fourier transform in the ss-variable. But we can rewrite this operator as in the following lemma.

Lemma 5.1.

It holds,

P1;γ,θh,η,ζ=Dt2+(th16L1,γ,θ)2+h13(L2,γ,θh,η,ζ)2,P_{1;\gamma,\theta}^{h,\eta,\zeta}=D_{t}^{2}+(t-h^{\frac{1}{6}}L_{1,\gamma,\theta})^{2}+h^{\frac{1}{3}}(L_{2,\gamma,\theta}^{h,\eta,\zeta})^{2}\;, (5.5)

where

L1;γ,θ=sinθDrcosθ(γ2r2+Ds),L2;γ,θh,η,ζ:=cosθDr+sinθ(γ2r2+Ds)h16(ζr+ηs)t.\begin{array}[]{ll}L_{1;\gamma,\theta}&=\sin\theta D_{r}-\cos\theta\left(\frac{\gamma}{2}r^{2}+D_{s}\right)\;,\\ L^{h,\eta,\zeta}_{2;\gamma,\theta}&:=\cos\theta D_{r}+\sin\theta\left(\frac{\gamma}{2}r^{2}+D_{s}\right)-h^{\frac{1}{6}}(\zeta r+\eta s)t\,.\end{array} (5.6)

Note that to compare with the case considered in [HelMo4] (η=ζ=0\eta=\zeta=0) we can write

L2;γ,θh,η,ζ=L2;γ,θh16(ζr+ηs)t,L_{2;\gamma,\theta}^{h,\eta,\zeta}=L_{2;\gamma,\theta}-h^{\frac{1}{6}}(\zeta r+\eta s)t\,, (5.7)

where L2;γ,θ:=L2;γ,θ0,0,0L_{2;\gamma,\theta}:=L_{2;\gamma,\theta}^{0,0,0}.

Proof of Lemma 5.1.

Let P1;γ,θh:=P1;γ,θh,0,0P_{1;\gamma,\theta}^{h}:=P_{1;\gamma,\theta}^{h,0,0}. Then (see [HelMo4, Eq. (11.4)])

P1;γ,θh=Dt2+(th16L1,γ,θ)2+h13(L2,γ,θ)2.P_{1;\gamma,\theta}^{h}=D_{t}^{2}+(t-h^{\frac{1}{6}}L_{1,\gamma,\theta})^{2}+h^{\frac{1}{3}}(L_{2,\gamma,\theta})^{2}\,.

With p=(ηs+ζr)tp=(\eta s+\zeta r)t, we have

P1;γ,θh,η,ζ=P1;γ,θh+h13[2(h16p)L2;γ,θh16(cosθ(Drp)+sinθ(Dsp))+(h16p)2].P_{1;\gamma,\theta}^{h,\eta,\zeta}=P_{1;\gamma,\theta}^{h}+h^{\frac{1}{3}}\big{[}-2(h^{\frac{1}{6}}p)L_{2;\gamma,\theta}-h^{\frac{1}{6}}\big{(}\cos\theta(D_{r}p)+\sin\theta(D_{s}p)\big{)}+(h^{\frac{1}{6}}p)^{2}\big{]}\,.

Finally, we observe by (5.7),

(L2;γ,θh,η,ζ)2=(L2,γ,θ)22(h16p)L2;γ,θh16(cosθ(Drp)+sinθ(Dsp))+(h16p)2.(L_{2;\gamma,\theta}^{h,\eta,\zeta})^{2}=(L_{2,\gamma,\theta})^{2}-2(h^{\frac{1}{6}}p)L_{2;\gamma,\theta}-h^{\frac{1}{6}}\big{(}\cos\theta(D_{r}p)+\sin\theta(D_{s}p)\big{)}+(h^{\frac{1}{6}}p)^{2}\,.

 

When η=ζ=0\eta=\zeta=0, this is the operator studied in [HelMo4], modulo a Fourier transformation with respect to the ss variable. Let us recall the following important result [HelMo4, Lem. 13.4] corresponding to the case (η,ζ)=(0,0)(\eta,\zeta)=(0,0).

Proposition 5.2 (Helffer-Morame).

For any C0>0C_{0}>0, δ]0,13[\delta\in]0,\frac{1}{3}[ and M>0M>0, there exist positive constants CC and h0h_{0} such that, for all θ\theta\in\mathbb{R}, |γ|M|\gamma|\leq M, and h]0,h0]h\in]0,h_{0}], we have, for any uC0(]C0hδ,C0hδ[××+¯)u\in C_{0}^{\infty}(]-C_{0}h^{\delta},C_{0}h^{\delta}[\times\mathbb{R}\times\overline{\mathbb{R}_{+}})

P0;γ,θh,0,0u,u(hΘ0+h43cconj(γ,θ)C(h118+hδ+1312))u2,\langle P_{0;\gamma,\theta}^{h,0,0}u\,,\,u\rangle\geq\big{(}h\Theta_{0}+h^{\frac{4}{3}}c^{\rm conj}(\gamma,\theta)-C(h^{\frac{11}{8}}+h^{\delta+\frac{13}{12}})\big{)}\;\|u\|^{2}\;, (5.8)

where

cconj(γ,θ):=(12)23δ013|γ|23(δ0sin2θ+cos2θ)13ν^0,c^{\rm conj}(\gamma,\theta):=\Big{(}\frac{1}{2}\Big{)}^{\frac{2}{3}}\delta_{0}^{\frac{1}{3}}|\gamma|^{\frac{2}{3}}(\delta_{0}\sin^{2}\theta+\cos^{2}\theta)^{\frac{1}{3}}\,{\hat{\nu}_{0}},

and P0;γ,θh,0,0P^{h,0,0}_{0;\gamma,\theta} is the operator introduced in (5.1).

Remark 5.3.

The underlying estimate in Proposition 5.2 is in fact

P1;γ,θh,0,0u,u(Θ0+h13cconj(γ,θ)C(h38+hδ+112))u2.\langle P_{1;\gamma,\theta}^{h,0,0}u\,,\,u\rangle\geq\big{(}\Theta_{0}+h^{\frac{1}{3}}c^{\rm conj}(\gamma,\theta)-C(h^{\frac{3}{8}}+h^{\delta+\frac{1}{12}})\big{)}\;\|u\|^{2}.

We can not directly compare P1;γ,θh,η,ζP_{1;\gamma,\theta}^{h,\eta,\zeta} and P1;γ,θh,0,0P_{1;\gamma,\theta}^{h,0,0} but this can be done by introducing a small perturbation of P1;γ,θh,0,0P_{1;\gamma,\theta}^{h,0,0} whose spectrum is just lifted. To achieve this goal we introduce for τ>0\tau>0

P1;γ,θ,τh:=Dt2+(th16L1,γ,θ)2+(1hτ)h13(L2,γ,θ)2,P_{1;\gamma,\theta,\tau}^{h}:=D_{t}^{2}+(t-h^{\frac{1}{6}}L_{1,\gamma,\theta})^{2}+(1-h^{\tau})h^{\frac{1}{3}}(L_{2,\gamma,\theta})^{2},

where we have modified the coefficient of (L2,γ,θ)2(L_{2,\gamma,\theta})^{2} by ϵ=h1/3+τ\epsilon=h^{1/3+\tau}. Heuristically this leads to a maximal shift of the bottom of the spectrum by 𝒪(h1/3+τ)\mathcal{O}(h^{1/3+\tau}). More precisely, we show by a slight variation of the argument in [HelMo4, Lem. 13.3]

Proposition 5.4.

For all τ]0,1[\tau\in\,]0,1[, for any C0>0C_{0}>0, δ]0,13[\delta\in]0,\frac{1}{3}[ and M>0M>0, there exist positive constants CC and h0h_{0} such that, for all θ\theta\in\mathbb{R}, |γ|M|\gamma|\leq M, and h]0,h0]h\in]0,h_{0}], we have, for any uC0(]C0hδ,C0hδ[××+¯)u\in C_{0}^{\infty}(]-C_{0}h^{\delta},C_{0}h^{\delta}[\times\mathbb{R}\times\overline{\mathbb{R}_{+}})

P1,γ,θ,τhu,u(Θ0+h13cconj(γ,θ)C(hτ+13+h38+hδ+112))u2.\langle P_{1,\gamma,\theta,\tau}^{h}u\,,\,u\rangle\geq\big{(}\Theta_{0}+h^{\frac{1}{3}}c^{\rm conj}(\gamma,\theta)-C(h^{\tau+\frac{1}{3}}+h^{\frac{3}{8}}+h^{\delta+\frac{1}{12}})\big{)}\;\|u\|^{2}\,. (5.9)

Note that the estimate in Proposition 5.2 holds without constraint on the support of the function in ss. This will not be the case for (η,ζ)0(\eta,\zeta)\not=0.
We now compare P1;γ,θh,η,ζu,u\langle P_{1;\gamma,\theta}^{h,\eta,\zeta}u,u\rangle and P1;γ,θ,τhu,u\langle P_{1;\gamma,\theta,\tau}^{h}u,u\rangle when

uC0(]C0hδ13,C0hδ13[×]C0hδ13,C0hδ13[×+¯).u\in C_{0}^{\infty}\big{(}\,]-C_{0}h^{\delta-\frac{1}{3}},C_{0}h^{\delta-\frac{1}{3}}[\times\,]-C_{0}h^{\delta-\frac{1}{3}},C_{0}h^{\delta-\frac{1}{3}}[\times\overline{\mathbb{R}_{+}}\,\big{)}.

and η,ζ\eta,\zeta satisfies (5.2).

Let us fix

δ]14,13[ and τ]0,16[.\delta\in\,]\frac{1}{4},\frac{1}{3}[\mbox{ and }\tau\in]0,\frac{1}{6}[\,. (5.10)

The estimates below hold uniformly with respect to uu, θ\theta\in\mathbb{R} and η,ζ,γ\eta,\zeta,\gamma satisfying (5.2).
Comparing L2,γ,θh,η,ζL_{2,\gamma,\theta}^{h,\eta,\zeta} and L2,γ,θL_{2,\gamma,\theta} in (5.7), we find111We use 2abεa2+ε1b22ab\leq\varepsilon a^{2}+\varepsilon^{-1}b^{2} with ε=hτ\varepsilon=h^{\tau}, a=L2,γ,θh,η,ζua=\|L_{2,\gamma,\theta}^{h,\eta,\zeta}u\| and b=L2,γ,θu2b=\|L_{2,\gamma,\theta}u\|^{2}., for all τ>0\tau>0,

(L2,γ,θh,η,ζ)2u,u=L2,γ,θh,η,ζu2(1hτ)L2,γ,θu2+(1hτ)(L2,γ,θh,η,ζL2,γ,θ)u2.\langle(L_{2,\gamma,\theta}^{h,\eta,\zeta})^{2}u,u\rangle=\|L_{2,\gamma,\theta}^{h,\eta,\zeta}u\|^{2}\geq(1-h^{\tau})\|L_{2,\gamma,\theta}u\|^{2}+(1-h^{-\tau})\|(L_{2,\gamma,\theta}^{h,\eta,\zeta}-L_{2,\gamma,\theta})u\|^{2}\,.

Consequently,

P1;γ,θh,η,ζu,u(Dt2+(th16L1;γ,θ)2+(1hτ)h13(L2;γ,θ)2)u,uh13τ(L2;γ,θh,η,ζL2;γ,θ)u2.\langle P_{1;\gamma,\theta}^{h,\eta,\zeta}u,u\rangle\geq\langle\big{(}D_{t}^{2}+(t-h^{\frac{1}{6}}L_{1;\gamma,\theta})^{2}+(1-h^{\tau})h^{\frac{1}{3}}(L_{2;\gamma,\theta})^{2}\big{)}u,u\rangle\\ -h^{\frac{1}{3}-\tau}||(L_{2;\gamma,\theta}^{h,\eta,\zeta}-L_{2;\gamma,\theta})u||^{2}\;. (5.11)

This implies (see (5.7) and the condition on the support of uu),

P1;γ,θh,η,ζu,uP1;γ,θ,τhu,uC(η2+ζ2)h2δτtu2,\langle P_{1;\gamma,\theta}^{h,\eta,\zeta}u,u\rangle\geq\langle P_{1;\gamma,\theta,\tau}^{h}u,u\rangle-C(\eta^{2}+\zeta^{2})h^{2\delta-\tau}||tu||^{2}\;, (5.12)

where we used (see (5.7))

L2;γ,θh,η,ζL2;γ,θ=h1/6t𝒪((|s|+|r|))=t𝒪(hδ16)L_{2;\gamma,\theta}^{h,\eta,\zeta}-L_{2;\gamma,\theta}=h^{1/6}t\,\mathcal{O}((|s|+|r|))=t\,\mathcal{O}(h^{\delta-\frac{1}{6}})

in the support of uu.
By (5.9) and (5.12) we have

P1;γ,θh,η,ζu,u(Θ0+cconj(γ,θ)h1/3C(h38+hδ+112+hτ))u2C(η2+ζ2)h2δτtu2.\langle P^{h,\eta,\zeta}_{1;\gamma,\theta}u,u\rangle\geq\big{(}\Theta_{0}+c^{\rm conj}(\gamma,\theta)h^{1/3}-C(h^{\frac{3}{8}}+h^{\delta+\frac{1}{12}}+h^{\tau})\big{)}\|u\|^{2}\\ -C(\eta^{2}+\zeta^{2})h^{2\delta-\tau}\|tu\|^{2}\;. (5.13)

Note that by (5.10) we have

h38+hδ+112+hτ+13+h2δτ=𝒪(h13+ς),h^{\frac{3}{8}}+h^{\delta+\frac{1}{12}}+h^{\tau+\frac{1}{3}}+h^{2\delta-\tau}=\mathcal{O}(h^{\frac{1}{3}+\varsigma})\,,

for some ς=ς(δ,τ)>0\varsigma=\varsigma(\delta,\tau)>0.
Consequently, there exist CC, ς>0\varsigma>0 and h0h_{0} such that, h]0,h0]\forall h\in]0,h_{0}],

P1;γ,θ,τh,η,ζu,u(Θ0+cconj(γ,θ)h1/3Ch13+ς)u2Ch13+ςtu2,\langle P_{1;\gamma,\theta,\tau}^{h,\eta,\zeta}u,u\rangle\geq\big{(}\Theta_{0}+c^{\rm conj}(\gamma,\theta)h^{1/3}-Ch^{\frac{1}{3}+\varsigma}\big{)}\|u\|^{2}-Ch^{\frac{1}{3}+\varsigma}\|tu\|^{2}\,, (5.14)

for any uC0(]Chδ13,Chδ13[2×+¯)u\in C_{0}^{\infty}(\,]-Ch^{\delta-\frac{1}{3}},Ch^{\delta-\frac{1}{3}}[^{2}\times\overline{\mathbb{R}_{+}}\,).
By coming back to the initial coordinates, we get the following generalization of Proposition 5.2.

Proposition 5.5.

Let C0,M>0C_{0},M>0 and δ]14,13[\delta\in]\frac{1}{4},\frac{1}{3}[ be given. There exist positive constants CC, h0h_{0}, and ς\varsigma, such that, for all h]0,h0]h\in]0,h_{0}], θ\theta\in\mathbb{R} and γ,η,ζ[M,M]\gamma,\eta,\zeta\in[-M,M], we have, for any uC0(]C0hδ,C0hδ[ 2×+¯)u\in C_{0}^{\infty}(]-C_{0}h^{\delta},C_{0}h^{\delta}[^{\,2}\,\times\,\overline{\mathbb{R}^{+}}),

P0,γ,θh,η,ζu,u(hΘ0+h43cconj(γ,θ)Ch43+ς)u2Ch13+ςtu2.\langle P_{0,\gamma,\theta}^{h,\eta,\zeta}u\,,\,u\rangle\geq\big{(}h\Theta_{0}+h^{\frac{4}{3}}c^{\rm conj}(\gamma,\theta)-Ch^{\frac{4}{3}+\varsigma}\big{)}\|u\|^{2}-Ch^{\frac{1}{3}+\varsigma}\|tu\|^{2}. (5.15)

Note here that the last term will be small when considering localized states satisfying (6.6).

6 Localization of bound states

We recall that the bound states of the operator P𝐀hP^{h}_{\bf A} in (1.3) are localized on the boundary near the curve where the magnetic field is tangent to the boundary Ω\partial\Omega. The localization is related with the analysis of a family of model operators in the half-space [LuPa5].

Consider +3:={(x1,x2,x3)3|x1>0}\mathbb{R}_{+}^{3}:=\{(x_{1},x_{2},x_{3})\in\mathbb{R}^{3}\,|\,x_{1}>0\} and the Neumann realization in +3\mathbb{R}_{+}^{3} of the operator,

H(ν)=Dx12+Dx22+(Dx3+x1cosνx2sinν)2,H(\nu)=D_{x_{1}}^{2}+D_{x_{2}}^{2}+(D_{x_{3}}+x_{1}\cos\nu-x_{2}\sin\nu)^{2}\,,

where ν[π2,π2]\nu\in[-\frac{\pi}{2},\frac{\pi}{2}].
More precisely, H(ν)H(\nu) is self-adjoint in L2(+3)L^{2}(\mathbb{R}_{+}^{3}) with the following domain

Dom(H(ν))={uL2(+3)|H(ν)uL2(+3),x1u|x1=0=0}.{\rm Dom}\big{(}H(\nu)\big{)}=\{u\in L^{2}(\mathcal{\mathbb{R}}^{3}_{+})\,|\,H(\nu)u\in L^{2}(\mathcal{\mathbb{R}}^{3}_{+}),~{}\partial_{x_{1}}u|_{x_{1}=0}=0\}\,.

We denote by

σ(ν)=infν[π2,π2]spec(H(ν)).\sigma(\nu)=\inf_{\nu\in[-\frac{\pi}{2},\frac{\pi}{2}]}{\rm spec}\,\big{(}H(\nu)\big{)}\,. (6.1)

We gather some properties of the lowest eigenvalue σ(ν)\sigma(\nu) (see [LuPa5], [HelMo2], and [HelMo4, Sec. 3.3]):

Proposition 6.1.

The following properties hold for the lowest eigenvalue σ(ν)\sigma(\nu) of H(ν)H(\nu):

  • For all ν[π2,π2]\nu\in[-\frac{\pi}{2},\frac{\pi}{2}], σ(ν)=σ(ν)\sigma(-\nu)=\sigma(\nu).

  • [0,π2]νσ(ν)[0,\frac{\pi}{2}]\ni\nu\mapsto\sigma(\nu) is monotone increasing and σ(0)=Θ0\sigma(0)=\Theta_{0}.

  • σ(ν)Θ0cos2ν+sin2ν\sigma(\nu)\geq\Theta_{0}\cos^{2}\nu+\sin^{2}\nu.

  • As ν0\nu\to 0, σ(ν)=Θ0+δ0|ν|+𝒪(ν2)\sigma(\nu)=\Theta_{0}+\sqrt{\delta_{0}}\,|\nu|+\mathcal{O}(\nu^{2}).

Here we recall that Θ0\Theta_{0} and δ0\delta_{0} are introduced in (4.2).

Let us return to the magnetic field in (1.1). Recall that, for xΩx\in\Omega, p(x)Ωp(x)\in\partial\Omega satisfies dist(x,Ω)=dist(x,p(x)){\rm dist}(x,\partial\Omega)={\rm dist}(x,p(x)), and it is uniquely defined when xx is sufficiently close to the boundary. For all xΩ¯x\in\overline{\Omega}, we introduce ν(x)[π2,π2]\nu(x)\in[-\frac{\pi}{2},\frac{\pi}{2}] by

(𝐁𝐍)(p(x))=sinν(x).(\mathbf{B}\cdot\mathbf{N})(p(x))=\sin\nu(x)\,. (6.2)

Hence ν(x)=0\nu(x)=0 implies that 𝐁(p(x))\mathbf{B}(p(x)) is tangent to Ω\partial\Omega at p(x)p(x), in other words that xx belongs to Γ\Gamma (see (1.4)). Now we recall the following lower bound related to the operator P𝐀hP_{\bf A}^{h} established in [HelMo4, Thm. 4.3]:

Proposition 6.2.

Under Assumption (1.1), there exist constants C,h0>0C,h_{0}>0 such that, for all h(0,h0]h\in(0,h_{0}] and uH1(Ω)u\in H^{1}(\Omega), we have

Ω|(hi𝐀)u|2𝑑xΩ(hWh(x)Ch5/4)|u(x)|2𝑑x,\int_{\Omega}|(h\nabla-i{\bf A})u|^{2}dx\geq\int_{\Omega}(hW_{h}(x)-Ch^{5/4})|u(x)|^{2}dx\,,

where

Wh(x)={1ifdist(x,Ω)2h3/8σ(ν(x))ifdist(x,Ω)2h3/8.W_{h}(x)=\begin{cases}1&{\rm if~{}}{\rm dist}(x,\partial\Omega)\geq 2h^{3/8}\\ \,\sigma(\nu(x))&{\rm if~{}}{\rm dist}(x,\partial\Omega)\leq 2h^{3/8}\end{cases}\,.

If additionally uH01(Ω)u\in H^{1}_{0}(\Omega), we have for some positive constant C0C_{0} the stronger lower bound

Ω|(hi𝐀)u|2𝑑x(hC0h5/4)Ω|u|2𝑑x.\int_{\Omega}|(h\nabla-i{\bf A})u|^{2}dx\geq(h-C_{0}h^{5/4})\int_{\Omega}|u|^{2}\,dx\,.

Combining the lower bound in Proposition 6.2 with the following leading term expansion of the lowest eigenvalue (see [HelMo4, Thm. 4.4])

λ1N(𝐀,h)=Θ0h+o(h),\lambda_{1}^{N}({\bf A},h)=\Theta_{0}h+o(h)\,, (6.3)

we get decay estimates for the ground states. Let us recall these localization estimates (see [FoHe2, Sec. 9.4] for details).

Proposition 6.3.

Given M>0M>0, there exists a positive constant α\alpha such that, if uhu_{h} is a normalized bound state of 𝒫h\mathcal{P}_{h} with eigenvalue λ(h)Mh\lambda(h)\leq Mh, then as h0+h\to 0_{+},

Ω(|uh(x)|2+h1|(hi𝐀)uh|2)exp(αdist(x,Ω)h1/2)𝑑x=𝒪(1).\int_{\Omega}\Big{(}|u_{h}(x)|^{2}+h^{-1}|(h\nabla-i{\bf A})u_{h}|^{2}\Big{)}\exp\Big{(}\frac{\alpha\,{\rm dist}(x,\partial\Omega)}{h^{1/2}}\Big{)}dx=\mathcal{O}(1)\,. (6.4)

Furthermore, there exist constants α1,ϵ0>0\alpha_{1},\epsilon_{0}>0 such that, as h0+h\to 0_{+},

{dist(x,Ω)<ϵ0}(|uh(x)|2+h1|(hi𝐀)uh|2)exp(α1dΓ(x)h1/4)𝑑x=𝒪(1),\int_{\{{\rm dist}(x,\partial\Omega)<\epsilon_{0}\}}\Big{(}|u_{h}(x)|^{2}+h^{-1}|(h\nabla-i{\bf A})u_{h}|^{2}\Big{)}\exp\Big{(}\frac{\alpha_{1}\,d_{\Gamma}(x)}{h^{1/4}}\Big{)}dx=\mathcal{O}(1)\,,

where

dΓ(x)=distΩ(p(x),Γ),d_{\Gamma}(x)={\rm dist}_{\partial\Omega}(p(x),\Gamma)\,, (6.5)

and distΩ{\rm dist}_{\partial\Omega} is the geodesic distance on Ω\partial\Omega.

Hence we have two levels of localization, first a strong one near Ω\partial\Omega and then an additional but weaker one near Γ\Gamma. Along the proof of Theorem 1.4, we will only use (6.4) and generalizations/consequences of it, as explained in the below remark.

Remark 6.4 (Applications of Proposition 6.3).

Let uhu_{h} be a normalized ground state of P𝐀hP_{\bf A}^{h}.

  1. 1.

    By (6.3), the hypothesis in Proposition 6.3 holds, hence the ground state uhu_{h} satisfies (6.4) and (6.5).

  2. 2.

    Pick an arbitrary point x0Γx_{0}\in\Gamma. In the coordinates introduced in (2.7), where t(x)=dist(x,Ω)t(x)={\rm dist}(x,\partial\Omega), r(x)=dΓ(x)r(x)=d_{\Gamma}(x) and uh(x)=u~h(r,s,t)u_{h}(x)=\tilde{u}_{h}(r,s,t) (see (2.19)), we deduce from (6.4) the following weaker, but quite useful estimates. For any n0n\geq 0,

    V~0tn|u~h|2𝑑s𝑑r𝑑t=𝒪(hn/2),\int_{\tilde{V}_{0}}t^{n}|\tilde{u}_{h}|^{2}dsdrdt=\mathcal{O}(h^{n/2}), (6.6)

    and

    V~0tn|(hr,s,t𝐀~)u~h|2𝑑r𝑑s𝑑t=𝒪(h1+n2),\int_{\tilde{V}_{0}}t^{n}|(h\nabla_{r,s,t}-\tilde{\bf A})\tilde{u}_{h}|^{2}drdsdt=\mathcal{O}(h^{1+\frac{n}{2}})\,,\\ (6.7)

    where V~0:=V~x0\tilde{V}_{0}:=\tilde{V}_{x_{0}} and 𝐀~\tilde{\bf A} are introduced in (2.18) and (2.22) respectively.

7 Estimating the quadratic form

7.1 A comparison estimate

We fix δ\delta and ϵ2\epsilon_{2} satisfying

518<δ<13 and 0<ϵ2<1.\frac{5}{18}<\delta<\frac{1}{3}\mbox{ and }0<\epsilon_{2}<1\,. (7.1)

We also fix R0>0R_{0}>0, h0>0h_{0}>0, x0Γx_{0}\in\Gamma and introduce for h(0,h0]h\in(0,h_{0}] the set

Qh(x0,R0,δ,ϵ2)={xΩ:|r(x)r0|R0hδ,|s(x)s0|R0hδ,0<t(x)<ϵ2}Q_{h}(x_{0},R_{0},\delta,\epsilon_{2})\\ =\{x\in\Omega\,:\,|r(x)-r_{0}|\leq R_{0}h^{\delta},~{}|s(x)-s_{0}|\leq R_{0}h^{\delta},~{}0<t(x)<\epsilon_{2}\} (7.2)

where (r(x),s(x),t(x))(r(x),s(x),t(x)) are introduced in (2.2) and, since x0Γx_{0}\in\Gamma,

y0:=(r0,s0,t0):=(r(x0),s(x0),t(x0))=(0,s(x0),0).y^{0}:=(r_{0},s_{0},t_{0}):=(r(x_{0}),s(x_{0}),t(x_{0}))=(0,s(x_{0}),0)\,. (7.3)

For simplicity, we omit most of the time the reference to δ\delta and ϵ2\epsilon_{2}.

Let 𝐀~=(A~1(2),A~2(2),A~3(2))\tilde{\bf A}=(\tilde{A}_{1}^{(2)},\tilde{A}_{2}^{(2)},\tilde{A}_{3}^{(2)}) be the magnetic potential associated with 𝐀{\bf A} via (2.22), with y=(y1,y2,y3)=(r,s,t)y=(y_{1},y_{2},y_{3})=(r,s,t) (see (2.7)). We introduce the following magnetic potential

𝐀~(2)(y)=|β|2β𝐀~yβ(y0)(yy0)ββ!,\tilde{\bf A}^{(2)}(y)=\sum_{|\beta|\leq 2}\frac{\partial^{\beta}\tilde{\bf A}}{\partial y^{\beta}}(y^{0})\frac{(y-y^{0})^{\beta}}{\beta!}, (7.4)

which is the quadratic Taylor expansion of 𝐀~\tilde{\bf A} at y0y^{0}. We introduce the quadratic form associated with the magnetic potential 𝐀~(2)\tilde{\bf A}^{(2)} as follows

q𝐀~(2)h(u)=Q~h(x0,R0)(1rκg(x0))(|(hDtA~3(2))u|2+(1+2rκg(x0))|(hDsA~2(2))u|2+|(hDrA~1(2))u|2)drdsdt,q_{\tilde{\bf A}^{(2)}}^{h}(u)=\int_{\tilde{Q}_{h}(x_{0},R_{0})}(1-r\kappa_{g}(x_{0}))\Big{(}|(hD_{t}-\tilde{A}_{3}^{(2)})u|^{2}\\ +(1+2r\kappa_{g}(x_{0}))|(hD_{s}-\tilde{A}_{2}^{(2)})u|^{2}+|(hD_{r}-\tilde{A}_{1}^{(2)})u|^{2}\Big{)}drdsdt\,,

where

Q~h(x0,R0,δ,ϵ2)={(r,s,t):max(|r|,|ss0|)<R0hδ,0<t<ϵ2},\tilde{Q}_{h}(x_{0},R_{0},\delta,\epsilon_{2})=\{(r,s,t)~{}:~{}\max(|r|,|s-s_{0}|)<R_{0}h^{\delta},~{}0<t<\epsilon_{2}\}\,, (7.5)

and (see (2.11))

κg(x0) is the geodesic curvature of Γ at x0.\kappa_{g}(x_{0})\mbox{ is the geodesic curvature of }\Gamma\mbox{ at }x_{0}\,.

The next lemma compares the quadratic forms uq𝐀~(2)h(u)u\mapsto q_{\tilde{\bf A}^{(2)}}^{h}(u) and uq𝐀h(u)u\mapsto q_{\bf A}^{h}(u) introduced in (2.21). The errors that will arise are controlled by the following energy

Mh(u)=n=06hn/2Ωt(x)n(|u|2+h1|(hi𝐀)u|2)𝑑x,M_{h}(u)=\sum_{n=0}^{6}h^{-n/2}\int_{\Omega}t(x)^{n}\Big{(}|u|^{2}+h^{-1}|(h\nabla-i{\bf A})u|^{2}\Big{)}dx\,, (7.6)

where t(x)=dist(x,Ω)t(x)={\rm dist}(x,\partial\Omega). Notice that,

(a)Ω|u|2𝑑x\displaystyle{\rm(a)}\quad\int_{\Omega}|u|^{2}dx Mh(u),\displaystyle\leq M_{h}(u)\,, (7.7)
(b)Ω|(hi𝐀)u|2𝑑x\displaystyle{\rm(b)}\quad\int_{\Omega}|(h\nabla-i{\bf A})u|^{2}dx Mh(u)h,\displaystyle\leq M_{h}(u)h\,,
(c)Ωt(x)n(|u|2+h1|(hi𝐀)u|2)𝑑x\displaystyle{\rm(c)}\quad\int_{\Omega}t(x)^{n}\big{(}|u|^{2}+h^{-1}|(h\nabla-i{\bf A})u|^{2}\big{)}dx Mh(u)hn/2(1n6).\displaystyle\leq M_{h}(u)h^{n/2}\quad(1\leq n\leq 6)\,.
Lemma 7.1.

There exist constants C,h0,ς0>0C,h_{0},\varsigma_{0}>0 such that, for all h(0,h0]h\in(0,h_{0}] and uH1(Ω)u\in H^{1}(\Omega) satisfying suppuQh(x0,R0){\rm supp\,}\,u\subset Q_{h}(x_{0},R_{0}), we have

(1Ch2δ)q𝐀~(2)h(u)CMh(u)h43+ς0q𝐀h(u)(1+Ch2δ)q𝐀~(2)h(u)+CMh(u)h43+ς0.(1-C\,h^{2\delta})q_{\tilde{\bf A}^{(2)}}^{h}(u)-CM_{h}(u)h^{\frac{4}{3}+\varsigma_{0}}\leq q_{\bf A}^{h}(u)\leq(1+C\,h^{2\delta})q_{\tilde{\bf A}^{(2)}}^{h}(u)+CM_{h}(u)h^{\frac{4}{3}+\varsigma_{0}}\,. (7.8)
Proof.

Let us recall two useful estimates whose proof does not require that the magnetic field curl𝐀{\rm curl}\,{\bf A} is constant (see [HelMo4, Lem. 10.1]):

q𝐀h(u)(1Ch2δ)q𝐀~(2)h(u)Ct1/2(hDx𝐀)u2C(q𝐀~(2)h(u))1/2(h3δ+h2δt+hδt2+t3)uC(h3δ+h2δt+hδt2+t3)u2,q_{\bf A}^{h}(u)\geq(1-Ch^{2\delta})q_{\tilde{\bf A}^{(2)}}^{h}(u)-C\|t^{1/2}(hD_{x}-{\bf A})u\|^{2}\\ -C\big{(}q_{\tilde{\bf A}^{(2)}}^{h}(u)\big{)}^{1/2}\|(h^{3\delta}+h^{2\delta}t+h^{\delta}t^{2}+t^{3})u\|\\ -C\|(h^{3\delta}+h^{2\delta}t+h^{\delta}t^{2}+t^{3})u\|^{2}\,, (7.9)

and

q𝐀h(u)(1+Ch2δ)q𝐀~(2)h(u)+Ct1/2(hDx𝐀)u2+C(q𝐀~(2)h(u))1/2(h3δ+h2δt+hδt2+t3)u+C(h3δ+h2δt+hδt2+t3)u2.q_{\bf A}^{h}(u)\leq(1+Ch^{2\delta})q_{\tilde{\bf A}^{(2)}}^{h}(u)+C\|t^{1/2}(hD_{x}-{\bf A})u\|^{2}\\ +C\big{(}q_{\tilde{\bf A}^{(2)}}^{h}(u)\big{)}^{1/2}\|(h^{3\delta}+h^{2\delta}t+h^{\delta}t^{2}+t^{3})u\|\\ +C\|(h^{3\delta}+h^{2\delta}t+h^{\delta}t^{2}+t^{3})u\|^{2}\,. (7.10)

In the sequel we use the notation 𝒪(chhρ+)\mathcal{O}(c_{h}h^{\rho+}) in the following manner

fh=𝒪(chhρ+) if and only if ϵ>0 s.t. fh=𝒪(chhρ+ϵ).f_{h}=\mathcal{O}(c_{h}h^{\rho+})\mbox{ if and only if }\exists\,\epsilon>0\mbox{ s.t. }f_{h}=\mathcal{O}(c_{h}h^{\rho+\epsilon}). (7.11)

Since we have assumed (7.1), we have

min(6δ, 2δ+1, 3δ+12, 22δ)>43.\min(6\delta\,,\,2\delta+1\,,\,3\delta+\frac{1}{2}\,,\,2-2\delta)>\frac{4}{3}\,.

We can now estimate the error terms appearing in (7.9) and (7.10). We deduce from (7.7) (a) that

h3δu2=𝒪(Mhh6δ)=𝒪(Mhh43+),\|h^{3\delta}u\|^{2}=\mathcal{O}(M_{h}h^{6\delta})=\mathcal{O}(M_{h}h^{\frac{4}{3}+})\,,

where we write MhM_{h} instead of Mh(u)M_{h}(u) for the sake of simplicity.

Using again (7.7) with n=1n=1, n=2n=2, n=4n=4 and n=6n=6, we get

t1/2(hDx𝐀)u2=𝒪(Mhh54),h2δtu2=𝒪(Mhh4δ+1),hδt2u2=𝒪(Mhh2δ+2),t3u2=𝒪(Mhh3).\begin{array}[]{rl}\|t^{1/2}(hD_{x}-{\bf A})u\|^{2}&=\mathcal{O}(M_{h}h^{\frac{5}{4}}),\\ \|h^{2\delta}tu\|^{2}&=\mathcal{O}(M_{h}h^{4\delta+1})\,,\\ \|h^{\delta}t^{2}u\|^{2}&=\mathcal{O}(M_{h}h^{2\delta+2})\,,\\ \|t^{3}u\|^{2}&=\mathcal{O}(M_{h}h^{3})\,.\end{array}

Consequently,

t1/2(hDx𝐀)u2+(h3δ+h2δt+hδt2+t3)u2=𝒪(Mhh43+).\|t^{1/2}(hD_{x}-{\bf A})u\|^{2}+\|(h^{3\delta}+h^{2\delta}t+h^{\delta}t^{2}+t^{3})u\|^{2}=\mathcal{O}(M_{h}h^{\frac{4}{3}+})\,. (7.12)

Notice that |𝐀~𝐀~(2)|=𝒪(h3δ)+𝒪(t3)|\tilde{\bf A}-\tilde{\bf A}^{(2)}|=\mathcal{O}(h^{3\delta})+\mathcal{O}(t^{3}) in Q~h(x0,R0)\tilde{Q}_{h}(x_{0},R_{0}). By the triangle inequality and (2.21)

q𝐀~(2)h(u)C(q𝐀h(u)+t3u2).q_{\tilde{\bf A}^{(2)}}^{h}(u)\leq C\,\big{(}q_{{\bf A}}^{h}(u)+\|t^{3}u\|^{2}\big{)}\,.

So by using (7.7) we get

q𝐀~(2)h(u)=𝒪(Mhh).q_{\tilde{\bf A}^{(2)}}^{h}(u)=\mathcal{O}(M_{h}h)\,.

Consequently, the foregoing estimate and (7.12) yield,

(q𝐀~(2)h(u))1/2(h3δ+h2δt+hδt2+t3)u=𝒪(Mhh).\big{(}q_{\tilde{\bf A}^{(2)}}^{h}(u)\big{)}^{1/2}\|(h^{3\delta}+h^{2\delta}t+h^{\delta}t^{2}+t^{3})u\|=\mathcal{O}(M_{h}h)\,.

This finishes the proof of (7.8).  

7.2 Normal form

Recall that we have fixed an arbitrary point x0Γx_{0}\in\Gamma and denoted its coordinates, in the (r,s,t)(r,s,t)-frame, by (0,s0,0)(0,s_{0},0). Let us also recall that the magnetic field 𝐁(x0)\mathbf{B}(x_{0}) can be expressed by (2.5).

Performing an appropriate gauge transformation on the set Q~h(x0,R0)\tilde{Q}_{h}(x_{0},R_{0}) introduced in (7.5), will yield a convenient normal form of the magnetic potential 𝐀~(2)\tilde{{\bf A}}^{(2)} introduced in (7.4).

Lemma 7.2.

There exist positive constants CC and C^\widehat{C}, and for all x0Γx_{0}\in\Gamma, there exist κˇ,ζ[C^,C^]\check{\kappa},\zeta\in[-\widehat{C},\widehat{C}] and a smooth function pˇ\check{p} on a neighborhood of Q~h(x0,R0,δ,ϵ2)\tilde{Q}_{h}(x_{0},R_{0},\delta,\epsilon_{2}), such that,

|𝐀~(2)(r,s,t)𝐀00(r,s,t)+pˇ(r,s,t)|C(r3+t2+|ss0|3),|\tilde{\bf A}^{(2)}(r,s,t)-{\bf A}^{00}(r,s,t)+\nabla\check{p}(r,s,t)|\leq C\,\big{(}r^{3}+t^{2}+|s-s_{0}|^{3}\big{)}\,,

where

𝐀00(r,s,t)=(ta1(r,s),ta2(r,s)+12κn,𝐁(x0)r2,0),{\bf A}^{00}(r,s,t)=\Big{(}ta_{1}(r,s),ta_{2}(r,s)+\frac{1}{2}\kappa_{n,\mathbf{B}}(x_{0})r^{2},0\Big{)},

κn,𝐁(x0)\kappa_{n,\mathbf{B}}(x_{0}) is introduced in (1.5), and

a1(r,s)\displaystyle a_{1}(r,s) =sinθ(s0)+(ζr+κˇ(ss0))cosθ(s0),\displaystyle=\sin\theta(s_{0})+\big{(}\zeta r+\check{\kappa}(s-s_{0})\big{)}\cos\theta(s_{0})\,,
a2(r,s)\displaystyle a_{2}(r,s) =cosθ(s0)+rκg(x0)cosθ(s0)+(ζr+κˇ(ss0))sinθ(s0).\displaystyle=-\cos\theta(s_{0})+r\kappa_{g}(x_{0})\cos\theta(s_{0})+(\zeta r+\check{\kappa}(s-s_{0}))\sin\theta(s_{0})\,.

Here θ(s0)\theta(s_{0}) is the angle introduced in (2.6) with x=x0x=x_{0}.

This lemma is an extension of Lemma 9.1 in [HelMo4] to the case when the magnetic field is not necessarily constant. In the constant magnetic field case we have ζ=0\zeta=0 and κˇ=κg(x0)\check{\kappa}=\kappa_{g}(x_{0}), where κg\kappa_{g} is the geodesic curvature introduced in (2.11). Note that we do not try at the moment to explicitly compute κˇ\check{\kappa} and ζ\zeta in the general case. We plan indeed to show that the result on the lowest eigenvalue is independent of κˇ\check{\kappa} and ζ\zeta.

Proof of Lemma 7.2.

Our goal is to determine the Taylor expansion up to order 11 of the magnetic field vector and corresponding magnetic field 22-form in the variables (r,s,t)(r,s,t), the Taylor expansion being computed at t=r=0t=r=0 and s=s0s=s_{0}. Up to a translation, we assume that s0=0s_{0}=0.
Writing the magnetic vector field in (1.2) as

𝐁=b~1r+b~2s+b~3t,\mathbf{B}=\widetilde{b}_{1}\partial_{r}+\widetilde{b}_{2}\partial_{s}+\widetilde{b}_{3}\partial_{t}\;, (7.13)

the Taylor expansion of order 11 at (0,0,0)(0,0,0) takes the form

b~1(r,s,t)=cosθ+γ1r+δ1s+σ1t+𝒪(r2+s2+t2),b~2(r,s,t)=sinθ+γ2r+δ2s+σ2t+𝒪(r2+s2+t2),b~3(r,s,t)=γ3r+σ3t+𝒪(r2+s2+t2).\begin{array}[]{ll}\widetilde{b}_{1}(r,s,t)&=\cos\theta+\gamma_{1}r+\delta_{1}s+\sigma_{1}t+\mathcal{O}(r^{2}+s^{2}+t^{2})\;,\\ \widetilde{b}_{2}(r,s,t)&=\sin\theta+\gamma_{2}r+\delta_{2}s+\sigma_{2}t+\mathcal{O}(r^{2}+s^{2}+t^{2})\;,\\ \widetilde{b}_{3}(r,s,t)&=\gamma_{3}r+\sigma_{3}t+\mathcal{O}(r^{2}+s^{2}+t^{2})\;.\end{array} (7.14)

where θ=θ(s0)\theta=\theta(s_{0}) and where we used (2.4)-(2.5). Here we have used that by definition of the coordinate rr, the function (r,s)b~3(r,s,0)(r,s)\mapsto\widetilde{b}_{3}(r,s,0) vanishes exactly at order 11 on r=0r=0. Note that γ3\gamma_{3} is κn,𝐁(x0)\kappa_{n,\mathbf{B}}(x_{0}), introduced in (1.5).

We now express that on t=0t=0 the norm of 𝐁\mathbf{B} should be one. In fact

|𝐁|2=1i,j1gijb~ib~j+b~32|\mathbf{B}|^{2}=\sum\limits_{1\leq i,j\leq 1}g_{ij}\tilde{b}_{i}\tilde{b}_{j}+\tilde{b}_{3}^{2} (7.15)

where the coefficients gijg_{ij} can be computed by (2.8), (2.9) and (2.12).
For t=0t=0, this reads

(b~1(r,s,0)2+α(r,s)(b~2(r,s,0))2+(b~3(r,s,0))2=1,(\widetilde{b}_{1}(r,s,0)^{2}+\alpha(r,s)(\widetilde{b}_{2}(r,s,0))^{2}+(\widetilde{b}_{3}(r,s,0))^{2}=1\;, (7.16)

where α(r,s)\alpha(r,s) is introduced in (2.12) and satisfies (2.13). We expand the last formula around t=r=s=0t=r=s=0. This leads, by taking t=0t=0 and considering the coefficients of rr and ss, to the two identities

γ1cosθ+γ2sinθκg(x0)sin2θ=0,\gamma_{1}\cos\theta+\gamma_{2}\sin\theta-\kappa_{g}(x_{0})\sin^{2}\theta=0\;,

and

δ1cosθ+δ2sinθ=0.\delta_{1}\cos\theta+\delta_{2}\sin\theta=0\;.

So it is natural to introduce the new parameters κ^\hat{\kappa} and ζ\zeta as follows

κˇ=δ1sinθ+δ2cosθ,ζ=γ1sinθ+(γ2κg(x0)sinθ)cosθ.\check{\kappa}=-\delta_{1}\sin\theta+\delta_{2}\cos\theta,\quad\zeta=-\gamma_{1}\sin\theta+(\gamma_{2}-\kappa_{g}(x_{0})\sin\theta)\cos\theta\,. (7.17)

So we observe that

δ1=κˇsinθ,δ2=κˇcosθ,\delta_{1}=-\check{\kappa}\sin\theta\;,\delta_{2}=\check{\kappa}\cos\theta\,,

and

γ1=ζsinθ,γ2=ζcosθ+κg(x0)sinθ.\gamma_{1}=-\zeta\sin\theta\;,\;\gamma_{2}=\zeta\cos\theta+\kappa_{g}(x_{0})\,\sin\theta\;.

Hence our “normal” form becomes

b~j(r,s,t)=b~j0(r,s,t)+𝒪(r2+s2+t2)\widetilde{b}_{j}(r,s,t)=\widetilde{b}_{j}^{0}(r,s,t)+{\cal O}(r^{2}+s^{2}+t^{2})

with

b~10(r,s,t)=cosθ(ζr+κˇs)sinθ+σ1t,b~2(r,s,t)=sinθ+(ζr+κˇs)cosθ+κg(x0)rsinθ+σ2t,b~3(r,s,t)γ3r+σ3t,\begin{array}[]{ll}\widetilde{b}_{1}^{0}(r,s,t)&=\cos\theta-(\zeta\,r+\check{\kappa}s)\sin\theta+\sigma_{1}t\;,\\ \widetilde{b}_{2}(r,s,t)&=\sin\theta+(\zeta\,r+\check{\kappa}\,s)\cos\theta+\kappa_{g}(x_{0})r\sin\theta+\sigma_{2}t\;,\\ \widetilde{b}_{3}(r,s,t)&\equiv\gamma_{3}\,r+\sigma_{3}t\;,\end{array} (7.18)

with

γ3=κn,𝐁(x0)=r𝐁|N.\gamma_{3}={\kappa_{n,{\bf B}}(x_{0})}=\partial_{r}\langle\mathbf{B}\,|\,N\rangle\;. (7.19)

Now consider 𝐁~=curl(r,s,t)𝐀~\widetilde{\mathbf{B}}={\rm curl}\,_{(r,s,t)}\tilde{\bf A}. We have 𝐁~=|g|1/2(b~1,b~2,b~3)\widetilde{\mathbf{B}}=|g|^{1/2}(\tilde{b}_{1},\tilde{b}_{2},\tilde{b}_{3}) (see [HelMo4, Eq. (5.13)]), where gg is introduced in (2.9). So we obtain by (2.15),

𝐁~ij(r,s,t)=𝐁~ij0(r,s,t)+𝒪(r2+s2+t2)\widetilde{\mathbf{B}}_{ij}(r,s,t)=\widetilde{\mathbf{B}}_{ij}^{0}(r,s,t)+{\cal O}(r^{2}+s^{2}+t^{2})

with

𝐁~230(r,s,t)=(1κg(x0)r)cosθ(ζr+κˇs)sinθ+σ1t,𝐁~310(r,s,t)=sinθ+(ζr+κˇs)cosθ+σ2t,𝐁~120(r,s,t)γ3r+σ3t.\begin{array}[]{ll}\widetilde{\mathbf{B}}_{23}^{0}(r,s,t)&=(1-\kappa_{g}(x_{0})r)\cos\theta-(\zeta\,r+\check{\kappa}s)\sin\theta+\sigma_{1}t\;,\\ \widetilde{\mathbf{B}}_{31}^{0}(r,s,t)&=\sin\theta+(\zeta\,r+\check{\kappa}\,s)\cos\theta+\sigma_{2}t\;,\\ \widetilde{\mathbf{B}}_{12}^{0}(r,s,t)&\equiv\gamma_{3}\,r+\sigma_{3}t\;.\end{array} (7.20)

Notice that the condition div(r,s,t)𝐁~=0{\rm div}_{(r,s,t)}\widetilde{\mathbf{B}}=0 reads (at r=t=0r=t=0 and s=0s=0) as follows

σ3=(κg(x0)κˇ)cosθ+ζsinθ.\sigma_{3}=\big{(}\kappa_{g}(x_{0})-\check{\kappa}\big{)}\cos\theta+\zeta\sin\theta\,.

We have now to choose a suitable corresponding magnetic potential to 𝐁~0\widetilde{\mathbf{B}}^{0}. We find

𝐀~00(r,s,t)=(A~100A~200A~300)=(ta1(r,s)+σ22t2ta2(r,s)+12γ3r2σ12t20)=𝐀00(r,s,t)+𝒪(t2),\tilde{{\bf A}}^{00}(r,s,t)=\left(\begin{array}[]{c}\tilde{A}_{1}^{00}\\ \tilde{A}_{2}^{00}\\ \tilde{A}_{3}^{00}\end{array}\right)=\left(\begin{array}[]{c}ta_{1}(r,s)+\frac{\sigma_{2}}{2}t^{2}\vskip 6.0pt plus 2.0pt minus 2.0pt\\ ta_{2}(r,s)+\frac{1}{2}\gamma_{3}r^{2}-\frac{\sigma_{1}}{2}t^{2}\vskip 6.0pt plus 2.0pt minus 2.0pt\\ 0\end{array}\right)={\bf A}^{00}(r,s,t)+\mathcal{O}(t^{2})\,, (7.21)

with

a1(r,s)=sinθ+(ζr+κˇs)cosθ,a_{1}(r,s)=\sin\theta+(\zeta r+\check{\kappa}s)\cos\theta\;, (7.22)
a2(r,s)=(1κg(x0)r)cosθ+(ζr+κˇs)sinθ.a_{2}(r,s)=-(1-\kappa_{g}(x_{0})r)\cos\theta+(\zeta r+\check{\kappa}s)\sin\theta\;. (7.23)

Moreover curl𝐀~(2)=𝐁~0{\rm curl\,}\tilde{\bf A}^{(2)}=\widetilde{\mathbf{B}}^{0} in the simply connected domain Q~h(x0,R0,δ,ϵ2)\tilde{Q}_{h}(x_{0},R_{0},\delta,\epsilon_{2}), so we can find a function pˇ\check{p} such that 𝐀~(2)=𝐀~00pˇ\tilde{\bf A}^{(2)}=\tilde{\bf A}^{00}-\nabla\check{p}.

Finally, γj(s):=b~jr(0,s,0)\gamma_{j}(s):=\frac{\partial\tilde{b}_{j}}{\partial r}(0,s,0) and δj(s):=b~js(0,s,0)\delta_{j}(s):=\frac{\partial\tilde{b}_{j}}{\partial s}(0,s,0) are bounded functions. Setting Mj=sup(|γj(s)|+|δj(s)|)M_{j}=\sup\big{(}|\gamma_{j}(s)|+|\delta_{j}(s)|\big{)} and M=max(M1,M2)M=\max(M_{1},M_{2}), we get from (7.17) that

|κˇ|2M and |ζ|2M+κg.|\check{\kappa}|\leq 2M\mbox{ and }|\zeta|\leq 2M+\|\kappa_{g}\|_{\infty}\,.

 

7.3 A second comparison estimate

We use the magnetic potential in Lemma 7.2 to approximate the quadratic form, as we did in Lemma 7.1. In particular, we approximate the metric by a flat one. Let us introduce the quadratic form corresponding to the magnetic potential in Lemma 7.1 (see [HelMo4, Lem. 10.2]):

q𝐀00h(v)=Q~h(x0,R0)(|hDtv|2+(1+2rκg(x0))|(hDsA200)v|2+|(hDrA100)v|2)drdsdt,q_{{\bf A}^{00}}^{h}(v)=\int_{\tilde{Q}_{h}(x_{0},R_{0})}\Big{(}|hD_{t}v|^{2}+(1+2r\kappa_{g}(x_{0}))|(hD_{s}-A^{00}_{2})v|^{2}\\ +|(hD_{r}-A_{1}^{00})v|^{2}\Big{)}drdsdt, (7.24)

where vH1(Q~h(x0,R0))v\in H^{1}(\tilde{Q}_{h}(x_{0},R_{0})) and Q~h(x0,R0)=Q~h(x0,R0,δ,ϵ2)\tilde{Q}_{h}(x_{0},R_{0})=\tilde{Q}_{h}(x_{0},R_{0},\delta,\epsilon_{2}) is the set introduced in (7.5).

We can obtain a further approximation of the quadratic form for functions obeying the conditions in (7.7).

Lemma 7.3 (Helffer-Morame).

There exist positive constants C,h0,ς0C,h_{0},\varsigma_{0} such that, for all h(0,h0]h\in(0,h_{0}] and uH1(Ω)u\in H^{1}(\Omega) s.t. suppuQh(x0,R0,δ,ϵ2){\rm supp\,}\,u\subset Q_{h}(x_{0},R_{0},\delta,\epsilon_{2}), we have

q𝐀00h(u~)CMh(u)h43+ς0q𝐀h(u)q𝐀00h(u~)+CMh(u)h43+ς0,q_{{\bf A}^{00}}^{h}(\tilde{u})-CM_{h}(u)h^{\frac{4}{3}+\varsigma_{0}}\leq q_{\bf A}^{h}(u)\leq q_{{\bf A}^{00}}^{h}(\tilde{u})+CM_{h}(u)h^{\frac{4}{3}+\varsigma_{0}}\,,

where Mh(u)M_{h}(u) is introduced in (7.6) and

u~=(1rκg(x0))1/2ueipˇ/h.\tilde{u}=(1-r\kappa_{g}(x_{0}))^{1/2}u\,e^{-i\check{p}/h}.
Proof.

We have the following two estimates from [HelMo4, Lem. 10.2] (whose proof does not require that the magnetic field curl𝐀{\rm curl}\,{\bf A} is constant)

q𝐀h(u)q𝐀00h(u~)Ct1/2(hDx𝐀)u2C(q𝐀00h(u~))1/2(h3δ+h+h2δt+t2)uC(h3δ+h+h2δt+t2)u2,q_{\bf A}^{h}(u)\geq q_{{\bf A}^{00}}^{h}(\tilde{u})-C\|t^{1/2}(hD_{x}-{\bf A})u\|^{2}\\ -C\big{(}q_{{\bf A}^{00}}^{h}(\tilde{u})\big{)}^{1/2}\|(h^{3\delta}+h+h^{2\delta}t+t^{2})u\|\\ -C\|(h^{3\delta}+h+h^{2\delta}t+t^{2})u\|^{2}\,,

and

q𝐀h(u)q𝐀00h(u~)+Ct1/2(hDx𝐀)u2+C(q𝐀00h(u~))1/2(h3δ+h+h2δt+t2)u+C(h3δ+h+h2δt+t2)u2.q_{\bf A}^{h}(u)\leq q_{{\bf A}^{00}}^{h}(\tilde{u})+C\|t^{1/2}(hD_{x}-{\bf A})u\|^{2}\\ +C\big{(}q_{{\bf A}^{00}}^{h}(\tilde{u})\big{)}^{1/2}\|(h^{3\delta}+h+h^{2\delta}t+t^{2})u\|\\ +C\|(h^{3\delta}+h+h^{2\delta}t+t^{2})u\|^{2}\,.

We can then estimate the remainder terms, using (7.7), as we did in the proof of Lemma 7.1. The only term that was not present satisfies

t2u2Mh(u)h2,\|t^{2}u\|^{2}\leq M_{h}(u)h^{2}\,,

where we used (7.7) (c) with n=4n=4.  

7.4 An estimate away from the curve Γ\Gamma

Let us now look at the quadratic form, q𝐀h(u)q_{\bf A}^{h}(u), when uu is supported away from Γ\Gamma. We start with a rough lower bound.

Lemma 7.4.

Given c>0c>0, ϵ2(0,1)\epsilon_{2}\in(0,1) and ρ(0,14)\rho\in(0,\frac{1}{4}), there exist positive constants h0,c~h_{0},\tilde{c} such that, if uH1(Ω)u\in H^{1}(\Omega) satisfies

suppu{xΩ:dist(x,Ω)<ϵ2,dΓ(x)chρ},{\rm supp\,}\,u\subset\{x\in\Omega~{}:~{}{\rm dist}(x,\partial\Omega)<\epsilon_{2}\,,~{}d_{\Gamma}(x)\geq c\,h^{\rho}\},

where dΓ(x)=distΩ(p(x),Γ)d_{\Gamma}(x)={\rm dist}_{\partial\Omega}(p(x),\Gamma) is introduced in (6.5), then

q𝐀h(u)(Θ0+c~hρ)hΩ|u|2𝑑x.q_{\bf A}^{h}(u)\geq(\Theta_{0}+\tilde{c}\,h^{\rho})h\int_{\Omega}|u|^{2}dx\,.
Proof.

If we verify that, for a given constant c>0c>0,

dΓ(x)chρc>0,|ν(x)|chρ,d_{\Gamma}(x)\geq ch^{\rho}\implies\exists\,c^{\prime}>0,~{}|\nu(x)|\geq c^{\prime}h^{\rho}\,, (7.25)

then the proof follows from Proposition 6.2, by using that h5/4=o(h1+ρ)h^{5/4}=o(h^{1+\rho}) and the lower bound from Proposition 6.1,

σ(ν)Θ0+δ02|ν|,\sigma(\nu)\geq\Theta_{0}+\frac{\sqrt{\delta_{0}}}{2}|\nu|\,,

in a neighborhood of 0.

Let us denote by m=minxΓκn,𝐁(x)m_{*}=\min_{x\in\Gamma}\kappa_{n,\mathbf{B}}(x), then m>0m_{*}>0 by Assumption 1.1, and (7.25) holds with c=mc/2c^{\prime}=m_{*}c/2. In fact, if |ν(x)|chρ|\nu(x)|\leq c^{\prime}h^{\rho}, we get by (6.2)

|𝐁𝐍(p(x))|chρ,|\mathbf{B}\cdot\mathbf{N}(p(x))|\leq c^{\prime}h^{\rho}\,,

and it follows from (2.5) that (recall that dΓ(x)=|r|d_{\Gamma}(x)=|r|, see Sec. 2)

mdΓ(x)chρ=mc2hρ.m_{*}d_{\Gamma}(x)\leq c^{\prime}h^{\rho}=m_{*}\frac{c}{2}h^{\rho}\,.

 

The next proposition is an improvement of Proposition 7.4 since it allows for the support of uu to be closer to the curve Γ\Gamma.

Proposition 7.5.

Given c>0c>0, ϵ2(0,1)\epsilon_{2}\in(0,1) and δ[14,13)\delta\in[\frac{1}{4},\frac{1}{3}), there exist positive constants h0,c,C,ς0h_{0},c_{*},C,\varsigma_{0} such that, if uH1(Ω)u\in H^{1}(\Omega) satisfies

suppu{xΩ:dist(x,Ω)<ϵ2,dΓ(x)chδ},{\rm supp\,}\,u\subset\{x\in\Omega~{}:~{}{\rm dist}(x,\partial\Omega)<\epsilon_{2},~{}d_{\Gamma}(x)\geq c\,h^{\delta}\}, (7.26)

where dΓ(x)=distΩ(p(x),Γ)d_{\Gamma}(x)={\rm dist}_{\partial\Omega}(p(x),\Gamma) is introduced in (6.5), then

q𝐀h(u)(Θ0+chδ)hΩ|u|2𝑑xCMh(u)h43+ς0,q_{\bf A}^{h}(u)\geq(\Theta_{0}+c_{*}h^{\delta})h\int_{\Omega}|u|^{2}dx-CM_{h}(u)h^{\frac{4}{3}+\varsigma_{0}}\,,

where Mh(u)M_{h}(u) is introduced in (7.6).

Proof.

Step 1. Let us fix constants c,R0>0c,R_{0}>0, ϵ2(0,1)\epsilon_{2}\in(0,1), δ[14,13)\delta\in[\frac{1}{4},\frac{1}{3}) and ρ(0,14)\rho\in(0,\frac{1}{4}). We assume that suppuQh(x0,R0,δ,ϵ2){\rm supp\,}\,u\subset Q_{h}(x_{0}^{*},R_{0},\delta,\epsilon_{2}) where x0Ωx_{0}^{*}\in\partial\Omega with boundary coordinates (r0,s0,t0=0)(r_{0},s_{0},t_{0}=0) satisfies (for hh small enough) chδ|r0|=dΓ(x0)2chρc\,h^{\delta}\leq|r_{0}|=d_{\Gamma}(x_{0}^{*})\leq 2c\,h^{\rho} and Qh(x0,R0,δ,ϵ2)Q_{h}(x_{0}^{*},R_{0},\delta,\epsilon_{2}) is introduced in (7.2).

We denote by Q~h(x0)=Q~h(x0,R0,δ,ϵ2)\tilde{Q}_{h}(x_{0}^{*})=\tilde{Q}_{h}(x_{0}^{*},R_{0},\delta,\epsilon_{2}) the neighborhood associated with Qh(x0,R0,δ,ϵ2)Q_{h}(x_{0}^{*},R_{0},\delta,\epsilon_{2}) by (7.5). By a translation, we may assume that s0=0s_{0}=0.

Consider the magnetic potential 𝐀~(2)\tilde{\bf A}^{(2)} introduced in (7.4). We modify the coordinates (r,s,t)(r,s,t) so that, locally near (r0,0,0)(r_{0},0,0), the metric GG in (2.12) is diagonal222We consider the curve Γh\Gamma_{h} defined by sΦx01(r0,s,0)s\mapsto\Phi_{x_{0}}^{-1}(r_{0},s,0), where x0=γ(x0)x_{0}=\gamma(x_{0}^{*}) and Φx0\Phi_{x_{0}} is the coordinate transformation introduced in (2.2). We parameterization Γh\Gamma_{h} by arc-length sγh(s)s\mapsto\gamma_{h}(s) and define the adapted coordinates by considering the normal geodesic to Γh\Gamma_{h} passing through x0x_{0}^{*}. with

α(r0,s)=1andαr(r0,s)=2κg(γ(s))+𝒪(hρ).\alpha(r_{0},s)=1\quad{\rm and}\quad\frac{\partial\alpha}{\partial r}(r_{0},s)=-2\kappa_{g}(\gamma(s))+\mathcal{O}(h^{\rho})\,. (7.27)

By Taylor’s formula

α(r,s)=12κg(γ(s))(rr0)+𝒪(hρ(rr0))+𝒪((rr0)2).\alpha(r,s)=1-2\kappa_{g}(\gamma(s))(r-r_{0})+\mathcal{O}(h^{\rho}(r-r_{0}))+\mathcal{O}((r-r_{0})^{2})\,.

In Q~h(x0)\tilde{Q}_{h}(x_{0}^{*}), we write

|κg(γ(s))κg(x0)|Chδ,|\kappa_{g}(\gamma(s))-\kappa_{g}(x_{0}^{*})|\leq Ch^{\delta},
α(r,s)=12κg(x0)(rr0)Chδ+ρ,\begin{aligned} \alpha(r,s)=1-2\kappa_{g}(x_{0}^{*})(r-r_{0})-Ch^{\delta+\rho}\end{aligned}\,,

and

hDy𝐀~=(hDy𝐀~(2))(𝐀~𝐀~(2)).hD_{y}-\tilde{\bf A}=(hD_{y}-\tilde{\bf A}^{(2)})-(\tilde{\bf A}-\tilde{\bf A}^{(2)})\,.

So we get, as in Lemma 7.1, the existence of C,ς0>0C^{\prime},\varsigma_{0}>0 such that

q𝐀h(u)(1Chδ+ρ)q𝐀~(2)h,x0(u)CMh(u)h43+ς0,q_{\bf A}^{h}(u)\geq(1-Ch^{\delta+\rho})q_{\tilde{\bf A}^{(2)}}^{h,x_{0}^{*}}(u)-C^{\prime}M_{h}(u)h^{\frac{4}{3}+\varsigma_{0}}\,,

where

q𝐀~(2)h,x0(u)=Q~h(x0)(1(rr0)κg(x0))(|(hDtA~3(2))u|2+(1+2(rr0)κg(x0))|(hDsA~2(2))u|2+|(hDrA~1(2))u|2)drdsdt.q_{\tilde{\bf A}^{(2)}}^{h,x_{0}^{*}}(u)=\int_{\tilde{Q}_{h}(x_{0}^{*})}(1-(r-r_{0})\kappa_{g}(x_{0}^{*}))\Big{(}|(hD_{t}-\tilde{A}_{3}^{(2)})u|^{2}\\ +(1+2(r-r_{0})\kappa_{g}(x_{0}^{*}))|(hD_{s}-\tilde{A}_{2}^{(2)})u|^{2}+|(hD_{r}-\tilde{A}_{1}^{(2)})u|^{2}\Big{)}drdsdt\,.

Performing a change of variables

(r,s)((rr0)cosωssinω,(rr0)sinω+scosω)(r,s)\mapsto\big{(}(r-r_{0})\cos\omega-s\sin\omega,(r-r_{0})\sin\omega+s\cos\omega\big{)}

which amounts to a rotation in the (r,s)(r,s)-plane (centered at (r0,0)(r_{0},0)), we may assume that the second component of 𝐁~=curl(r,s,t)𝐀~=(B~23,B~31,B~12)\widetilde{\mathbf{B}}={\rm curl}_{(r,s,t)}\tilde{\bf A}=(\tilde{B}_{23},\tilde{B}_{31},\tilde{B}_{12}) vanishes at (r0,0,0)(r_{0},0,0), by choosing ω\omega so that

B~31(x0)cosω+B~23(x0)sinω=0.\tilde{B}_{31}(x_{0}^{*})\cos\omega+\tilde{B}_{23}(x_{0}^{*})\sin\omega=0\,.

At the same time, this rotation leaves |𝐁||\mathbf{B}| and the measure drdsdrds invariant. Then performing a gauge transformation (see [HelMo4, Sec. 16.3]), we may assume that

𝐀~(2)(r,s,t)=𝐀~(2,0)(r,s,t)+𝒪(|rr0|t+|s|t+t2)\tilde{\bf A}^{(2)}(r,s,t)=\tilde{\bf A}^{(2,0)}(r,s,t)+\mathcal{O}(|r-r_{0}|t+|s|t+t^{2})

where

𝐀~(2,0)(r,s,t):=(c~10s2B~23(0)t+B~12(0)(rr0)+c~20(rr0)20).\tilde{\bf A}^{(2,0)}(r,s,t):=\left(\begin{array}[]{l}\tilde{c}_{1}^{0}\,s^{2}\\ \tilde{B}_{23}^{(0)}t+\tilde{B}_{12}^{(0)}(r-r_{0})+\tilde{c}_{2}^{0}\,(r-r_{0})^{2}\\ 0\end{array}\right)\,.

Here

𝐁~(0):=𝐁~(r0,0,0)=(B~23(0),B~31(0)=0,B~12(0))\widetilde{\mathbf{B}}^{(0)}:=\widetilde{\mathbf{B}}(r_{0},0,0)=(\tilde{B}_{23}^{(0)},\tilde{B}_{31}^{(0)}=0,\tilde{B}_{12}^{(0)})

and c~10,c~20\tilde{c}_{1}^{0},\tilde{c}_{2}^{0} are constants.
Similarly to the proof of Lemma 7.1, by writing

hDy𝐀~(2)=hDy𝐀~(2,0)(𝐀~(2)𝐀~(2,0))hD_{y}-\tilde{\bf A}^{(2)}=hD_{y}-\tilde{\bf A}^{(2,0)}-(\tilde{\bf A}^{(2)}-\tilde{\bf A}^{(2,0)})

and

(hDy𝐀~(2,0))u(hDy𝐀~)u+(𝐀~𝐀~(2))u+(𝐀~(2)𝐀~(2,0))u,\|(hD_{y}-\tilde{\bf A}^{(2,0)})u\|\leq\|(hD_{y}-\tilde{\bf A})u\|+\|(\tilde{\bf A}-\tilde{\bf A}^{(2)})u\|+\|(\tilde{\bf A}^{(2)}-\tilde{\bf A}^{(2,0)})u\|\,,

we get

q𝐀~(2)h,x0(u)(1Ch2δ)q𝐀~(2,0)h,x0(u)C′′Mh(u)h43+ς0.q_{\tilde{\bf A}^{(2)}}^{h,x_{0}^{*}}(u)\geq(1-Ch^{2\delta})q_{\tilde{\bf A}^{(2,0)}}^{h,x_{0}^{*}}(u)-C^{\prime\prime}M_{h}(u)h^{\frac{4}{3}+\varsigma_{0}}\,.

Thus we are left with finding a lower bound of q𝐀~(2,0)h,x0(u)q_{\tilde{\bf A}^{(2,0)}}^{h,x_{0}^{*}}(u).
Note that, since |𝐁|=1|\mathbf{B}|=1 and by (7.27), the metric satisfies |g|=1|g|=1 on x0x_{0}^{*}, we have by (7.15), |B~23(0)|2+|B~12(0)|2=1|\tilde{B}_{23}^{(0)}|^{2}+|\tilde{B}_{12}^{(0)}|^{2}=1.
Moreover, since 𝐁𝐍\mathbf{B}\cdot\mathbf{N} vanishes linearly on Γ={r=0}\Gamma=\{r=0\}, there exist C1>0C_{1}>0 and C2>0C_{2}>0 such that

1C1|r0||B~12(0)|2C2|r0|,||B~23(0)|1|C2r02,|c~10|+|c~20|C2.\frac{1}{C_{1}}|r_{0}|\leq|\tilde{B}_{12}^{(0)}|^{2}\leq C_{2}\,|r_{0}|\,,\quad|\,|\tilde{B}_{23}^{(0)}|-1\,|\leq C_{2}r_{0}^{2},\quad|\tilde{c}_{1}^{0}|+|\tilde{c}_{2}^{0}|\leq C_{2}\,.

The previous estimates yield a lower bound of q𝐀~(2,0)h(u)q_{\tilde{\bf A}^{(2,0)}}^{h}(u) by comparing with a model operator (after rescaling the variables r~=h1/3(rr0)\tilde{r}=h^{1/3}(r-r_{0}), s~=h1/3s\tilde{s}=h^{1/3}s and t~=h1/2t\tilde{t}=h^{1/2}t). In fact, by [HelMo4, Lemma 16.1], there exists c1>0c_{1}>0 such that,

q𝐀~(2,0)h,x0(u)(Θ0+c1|r0|)hΩ|u|2𝑑x.q_{\tilde{\bf A}^{(2,0)}}^{h,x_{0}^{*}}(u)\geq(\Theta_{0}+c_{1}|r_{0}|)h\int_{\Omega}|u|^{2}dx\,.

Note that, we can use Lemma 16.1 of [HelMo4] under our assumptions on the support of uu.
Step 2. We can reduce to the setting of Step 1 and Lemma 7.4 by means of a partition of unity. In fact, consider an hh-dependent partition of unity χ12+χ22=1\chi_{1}^{2}+\chi_{2}^{2}=1 on {dist(x,Ω)<ϵ2}\{{\rm dist}(x,\partial\Omega)<\epsilon_{2}\} such that

suppχ1{dΓ(x)c2hρ},suppχ2{dΓ(x)chρ},i=12|χi|2=𝒪(h2ρ).{\rm supp\,}\chi_{1}\subset\{d_{\Gamma}(x)\geq\frac{c}{2}h^{\rho}\},\quad{\rm supp\,}\chi_{2}\subset\{d_{\Gamma}(x)\leq ch^{\rho}\},\quad\sum_{i=1}^{2}|\nabla\chi_{i}|^{2}=\mathcal{O}(h^{-2\rho})\,.

If uH1(Ω)u\in H^{1}(\Omega) satisfies (7.26), then

q𝐀h(u)=i=12(q𝐀h(χiu)h2|χi|u2),q_{\bf A}^{h}(u)=\sum_{i=1}^{2}\Big{(}q_{\bf A}^{h}(\chi_{i}u)-h^{2}\|\,|\nabla\chi_{i}|u\,\|^{2}\Big{)}\,,

where

q𝐀h(χ1u)(Θ0+c~hρ)hΩ|χ1u|2𝑑xbyProposition7.4,\displaystyle q_{\bf A}^{h}(\chi_{1}u)\geq(\Theta_{0}+\tilde{c}\,h^{\rho})h\int_{\Omega}|\chi_{1}u|^{2}dx\quad{\rm by~{}Proposition~{}\ref{lem:10.1*f}}\,,
q𝐀h(χ2u)(1Chδ+ρ)(Θ0+c1hδ)hΩ|χ2u|2𝑑xMh(u)h43+ς0byStep1,\displaystyle q_{\bf A}^{h}(\chi_{2}u)\geq(1-Ch^{\delta+\rho})(\Theta_{0}+c_{1}h^{\delta})h\int_{\Omega}|\chi_{2}u|^{2}dx-M_{h}(u)h^{\frac{4}{3}+\varsigma_{0}}\quad{\rm by~{}Step~{}1}\,,
i=12h2|χi|u2=𝒪(h22ρ)=o(h1+δ),\displaystyle\sum_{i=1}^{2}h^{2}\|\,|\nabla\chi_{i}|u\,\|^{2}=\mathcal{O}(h^{2-2\rho})=o(h^{1+\delta})\,,

where in the last step we used that 0<ρ<140<\rho<\frac{1}{4} and 14<δ<13\frac{1}{4}<\delta<\frac{1}{3}.  

8 Lower bound

8.1 Another model

The model in (5.1) corresponds to the quadratic form in (7.24) when κg(x0)=0\kappa_{g}(x_{0})=0. However, when κg(x0)0\kappa_{g}(x_{0})\not=0, the situation is similar to [HelMo4, Sec. 15]. The model compatible with (7.24) can still be reduced to the one in (5.1) with appropriate choices of the parameters η,ζ,γ\eta,\zeta,\gamma (see (8.22)).

8.1.1 A new model quadratic form

Let us fix a boundary point x0Γx_{0}\in\Gamma and denote the model quadratic form near x0x_{0} by

uqm(u):=q𝐀00(u)u\mapsto q_{m}(u):=q_{{\bf A}^{00}}(u) (8.1)

where q𝐀00q_{{\bf A}^{00}} is given in (7.24), uH1(Q~h(x0,R0))u\in H^{1}(\tilde{Q}_{h}(x_{0},R_{0})) and Q~h(x0,R0)=Q~h(x0,R0,δ,ϵ2)\tilde{Q}_{h}(x_{0},R_{0})=\tilde{Q}_{h}(x_{0},R_{0},\delta,\epsilon_{2}) is the set introduced in (7.5). Furthermore, we assume that that the metric is flat at x0x_{0} and the coordinates of x0x_{0} in the (r,s,t)(r,s,t) frame are (0,s0=0,0)(0,s_{0}=0,0), after performing a translation with respect to the ss variable.

Following the proof of [HelMo4, Lem. 15.1], we are led to the analysis of the model quadratic form (see Lemma 8.1)

qm,0h(u)=Q~h(x0,R0)(h2|Dtu|2+|tuL1hu|2+|L2hu|2)𝑑r𝑑s𝑑t,q^{h}_{m,0}(u)=\int_{\tilde{Q}_{h}(x_{0},R_{0})}\Big{(}h^{2}|D_{t}u|^{2}+|tu-L^{h}_{1}u|^{2}+|L^{h}_{2}u|^{2}\Big{)}\;drdsdt\;, (8.2)

where

L1h\displaystyle L^{h}_{1} =a1hDr+a20hDs12cosθκn,𝐁(x0)r2,\displaystyle=a_{1}\,hD_{r}+a^{0}_{2}\,hD_{s}-\frac{1}{2}\,\cos\theta\,\kappa_{n,\bf B}(x_{0})\,r^{2}\;, (8.3)
L2h\displaystyle L^{h}_{2} =a21hDr+a11hDs+12sinθκn,𝐁(x0)r2,\displaystyle=a^{1}_{2}\,hD_{r}+a^{1}_{1}\,hD_{s}+\frac{1}{2}\,\sin\theta\,\kappa_{n,\bf B}(x_{0})\,r^{2}\;,

and, with θ=θ(s0)\theta=\theta(s_{0}) the angle defined by (2.6), we introduce the following functions

a1(r,s)=sinθ+cosθ(ζr+κˇs),a2(r,s)=cosθ+κg(x0)cosθr+sinθ(ζr+κˇs),a20(r,s)=cosθκg(x0)cosθr+sinθ(ζr+κˇs),a21(r,s)=cosθsinθ(ζr+κˇs),a11(r,s)=sinθ+sinθκg(x0)r+cosθ(ζr+κˇs),α(r)=1+2κg(x0)r.\begin{array}[]{ll}a_{1}(r,s)&=\sin\theta+\cos\theta(\zeta r+\check{\kappa}s)\;,\\ a_{2}(r,s)&=-\cos\theta+\kappa_{g}(x_{0})\cos\theta r+\sin\theta(\zeta r+\check{\kappa}s)\;,\\ a_{2}^{0}(r,s)&=-\cos\theta-\kappa_{g}(x_{0})\cos\theta r+\sin\theta(\zeta r+\check{\kappa}s)\;,\\ a_{2}^{1}(r,s)&=\cos\theta-\sin\theta(\zeta r+\check{\kappa}s)\;,\\ a_{1}^{1}(r,s)&=\sin\theta+\sin\theta\kappa_{g}(x_{0})r+\cos\theta(\zeta r+\check{\kappa}s)\;,\\ \alpha(r)&=1+2\kappa_{g}(x_{0})r\;.\end{array} (8.4)

We will consider the form qm,0q_{m,0} on the following class of functions

𝒟0={uH1(Ωh):u|(𝒬h)×]0,hδ[=0,u|𝒬h×{hδ}=0}\mathcal{D}_{0}=\{u\in H^{1}(\Omega^{h})~{}:~{}u|_{(\partial\mathcal{Q}^{h})\times\,]0,h^{\delta}[}=0,~{}u|_{\mathcal{Q}^{h}\times\{h^{\delta}\}}=0\} (8.5)

where

Ωh=𝒬h×]0,hδ[,𝒬h=]R0hδ,R0hδ[2.\Omega_{h}=\mathcal{Q}^{h}\times\,]0,h^{\delta}[\,,\quad\mathcal{Q}^{h}=]-R_{0}h^{\delta},R_{0}h^{\delta}[^{2}\,.

The precise relation between the model quadratic forms in (8.1) and (8.2) is given in the following lemma.

Lemma 8.1.

For any δ(518,13)\delta\in(\frac{5}{18},\frac{1}{3}) and τ1>0\tau_{1}>0, there exists C>0C>0 such that, for any u𝒟0u\in\mathcal{D}_{0} and h(0,1)h\in(0,1),

(1+Ch2δ)qmh(u)(1Chτ1)qm,0h(u)C((h2δ+hτ1)tu2+h6δτ1u2).(1+Ch^{2\delta})q^{h}_{m}(u)\geq(1-Ch^{\tau_{1}})q^{h}_{m,0}(u)-C\big{(}\|(h^{2\delta}+h^{\tau_{1}})tu\|^{2}+h^{6\delta-\tau_{1}}\|u\|^{2}\big{)}\,.
Proof.

The proof follows that of Lemma 15.1 in [HelMo4] with some adjustments in the formulas (15.9), (15.16) and (15.17) in [HelMo4].
We have indeed

|1(a1)2α(a2)2|Ch2δ,|1-(a_{1})^{2}-\alpha(a_{2})^{2}|\leq Ch^{2\delta}\;, (8.6)

where we used that α(r)1/2=1+κg(x0)r+𝒪(h2δ)\alpha(r)^{1/2}=1+\kappa_{g}(x_{0})r+\mathcal{O}(h^{2\delta}) on the support of uu, which follows by (8.4).

We also observe that :

|αa2a20|+|α1/2a2+a21|+|α1/2a1a11|C(r2+s2)|\alpha a_{2}-a^{0}_{2}|+|\alpha^{1/2}a_{2}+a^{1}_{2}|+|\alpha^{1/2}a_{1}-a^{1}_{1}|\leq C(r^{2}+s^{2}) (8.7)

and

|α1/2a1sinθ|+|αa2+cosθ|C(r2+s2)1/2.|\alpha^{1/2}a_{1}-\sin\theta|+|\alpha a_{2}+\cos\theta|\leq C(r^{2}+s^{2})^{1/2}\;. (8.8)

 

Later on, we will choose δ\delta and τ1\tau_{1} in a convenient way (see Remark 8.3).

8.1.2 Linearizing change of variable

In order to reduce to the case κg=0\kappa_{g}=0 and eliminate the slightly variable coefficients of DrD_{r} and DsD_{s} in (7.24), we argue as [HelMo4a, Sec. 15.2] by performing a change of variables. The argument does not work in our case in the same way as [HelMo4a, Sec. 15.2], but it leads to the fact that for our lower bound the only relevant parameters are η:=κˇκg\eta:=\check{\kappa}-\kappa_{g} and ζ\zeta (see (7.24)).

The below computations are essentially the same as in [HelMo4, Sec. 15.2] but we have to do them carefully in order to capture the correct η\eta and ζ\zeta appearing in (5.1).

Let us follow, what this change of variable was doing. We introduce

κ:=κg(x0).\kappa:=\kappa_{g}(x_{0}). (8.9)

Let us make the change of variables (r,s)=Φκ(p,q)(r,s)=\Phi_{\kappa}(p,q) with

r=sinθp+cosθqκ2[cosθp+sinθq]2,s=cosθp+sinθqκ2[sin(2θ)(p2q2)+2cos(2θ)pq],\begin{array}[]{l}r=\sin\theta\,p+\cos\theta\,q-\frac{\kappa}{2}[-\cos\theta\,p+\sin\theta\,q]^{2}\;,\\ s=-\cos\theta\,p+\sin\theta\,q-\frac{\kappa}{2}[\sin(2\theta)\,(p^{2}-q^{2})+2\cos(2\theta)\,pq]\;,\end{array} (8.10)

where θ=θ(s0)\theta=\theta(s_{0}) is the angle defined by (2.6).

The map Φκ\Phi_{\kappa} is a perturbation of a rotation and, by the local inversion theorem, it is easily seen as a local diffeomorphism sending a fixed neighborhood of (0,0)(0,0) onto another neighborhood of (0,0)(0,0).

Then, for hh small enough, 𝒬h:=]R0hδ,R0hδ[2\mathcal{Q}^{h}:=]-R_{0}h^{\delta},R_{0}h^{\delta}[^{2} is transformed by Φκ1\Phi_{\kappa}^{-1} to the set 𝒬0h\mathcal{Q}^{h}_{0} satisfying :

𝒬0h=Φκ1(𝒬h)]R0hδ,R0hδ[×]R0hδ,R0hδ[.\mathcal{Q}^{h}_{0}=\Phi_{\kappa}^{-1}(\mathcal{Q}^{h})\;\subset\,\;]-R_{0}^{\prime}h^{\delta}\;,\;R_{0}^{\prime}h^{\delta}[\;\times\;]-R_{0}^{\prime}h^{\delta}\;,\;R_{0}^{\prime}h^{\delta}[\;. (8.11)

Let us write

Dp=c11Dr+c12Ds,Dq=c21Dr+c22Ds,D_{p}=c_{11}D_{r}+c_{12}D_{s},\ \ \ D_{q}=c_{21}D_{r}+c_{22}D_{s}\;, (8.12)

We can express the functions cijc_{ij} in terms of the (p,q)(p,q) variables, by using (8.10). In fact, we introduce cij(r,s)=cˇij(p,q)c_{ij}(r,s)=\check{c}_{ij}(p,q), and observe that

cˇ11(p,q)\displaystyle\check{c}_{11}(p,q) =rp=sinθ+κcosθ(cosθp+sinθq);\displaystyle=\frac{\partial r}{\partial p}=\sin\theta+\kappa\cos\theta\,(-\cos\theta\,p+\sin\theta\,q)\;;
cˇ12(p,q)\displaystyle\check{c}_{12}(p,q) =sp=cosθκ(sin(2θ)p+cos(2θ)q);\displaystyle=\frac{\partial s}{\partial p}=-\cos\theta-\kappa\,(\sin(2\theta)\,p+\cos(2\theta)\,q)\;;
cˇ21(p,q)\displaystyle\check{c}_{21}(p,q) =rq=cosθκsinθ(cosθp+sinθq);\displaystyle=\frac{\partial r}{\partial q}=\cos\theta-\kappa\sin\theta\,(-\cos\theta\,p+\sin\theta\,q)\;;
cˇ22(p,q)\displaystyle\check{c}_{22}(p,q) =sq=sinθκ(sin(2θ)q+cos(2θ)p).\displaystyle=\frac{\partial s}{\partial q}=\sin\theta-\kappa\,(-\sin(2\theta)\,q+\cos(2\theta)\,p)\;.

Then we return back to the (r,s)(r,s) variables, by using (8.10). Noticing that, as (p,q)(0,0)(p,q)\to(0,0),

r=sinθp+cosθq+𝒪(p2+q2),s=cosθp+sinθq+𝒪(p2+q2),r=\sin\theta\,p+\cos\theta\,q+{\cal O}(p^{2}+q^{2})\,,\quad s=-\cos\theta\,p+\sin\theta\,q+\mathcal{O}(p^{2}+q^{2})\,, (8.13)

we get

c11(r,s)\displaystyle c_{11}(r,s) =sinθ+κcosθs+𝒪(r2+s2);\displaystyle=\sin\theta+\kappa\cos\theta\,s+{\cal O}(r^{2}+s^{2})\;; (8.14)
c12(r,s)\displaystyle c_{12}(r,s) =cosθκ(cosθrsinθs)+𝒪(r2+s2);\displaystyle=-\cos\theta-\kappa(\cos\theta\,r-\sin\theta\,s)+{\cal O}(r^{2}+s^{2})\;;
c21(r,s)\displaystyle c_{21}(r,s) =cosθκsinθs+𝒪(r2+s2);\displaystyle=\cos\theta-\kappa\sin\theta\,s+{\cal O}(r^{2}+s^{2})\;;
c22(r,s)\displaystyle c_{22}(r,s) =sinθ+κ(sinθr+cosθs)+𝒪(r2+s2).\displaystyle=\sin\theta+\kappa(\sin\theta\,r+\cos\theta\,s)+{\cal O}(r^{2}+s^{2})\;.

Let us now control the measure in the change of variable. By an easy computation, we get :

drds=αˇ1dpdq,αˇ1(p,q)=1+κ(sinθp+cosθq)+𝒪(p2+q2).dr\,ds=\check{\alpha}_{1}dpdq\,,\quad\check{\alpha}_{1}(p,q)=1+\kappa(\sin\theta\,p+\cos\theta\,q)+\mathcal{O}(p^{2}+q^{2})\,.

By using (8.13), α1(r,s)=αˇ1(p,q)\alpha_{1}(r,s)=\check{\alpha}_{1}(p,q) satisfies

|α11κr|C(r2+s2),|\alpha_{1}-1-\kappa r|\leq C(r^{2}+s^{2})\;, (8.15)

where r=r(p,q)r=r(p,q) is defined in (8.10).


Similarly to Lemma 8.1 we get also that one can go from the control of qm,0h(u)q_{m,0}^{h}(u) to the control of the new quadratic form333We express L1hL_{1}^{h} and L2hL_{2}^{h} (see (8.3)) in terms of the (p,q)(p,q) variables introduced in (8.10) and neglect the terms of order 𝒪(r2+s2)=𝒪(p2+q2)\mathcal{O}(r^{2}+s^{2})=\mathcal{O}(p^{2}+q^{2}).

qm,1h(u)=Ω0h(h2|Dtu|2+|tuM1hu|2+|M2hu|2)αˇ1𝑑p𝑑q𝑑t,q^{h}_{m,1}(u)=\int_{\Omega^{h}_{0}}\Big{(}h^{2}|D_{t}u|^{2}+|tu-M^{h}_{1}u|^{2}+|M^{h}_{2}u|^{2}\Big{)}\check{\alpha}_{1}\,dpdqdt\;, (8.16)

with

Ω0h:=𝒬0h×]0,hδ[,\Omega^{h}_{0}:=\mathcal{Q}^{h}_{0}\times]0,h^{\delta}[\;,

and

M1h\displaystyle M^{h}_{1} =hDp+h((κˇκ)s+ζr)Dq12cosθκn,𝐁(x0)(sinθp+cosθq)2,\displaystyle=hD_{p}+h\left((\check{\kappa}-\kappa)s+\zeta r\right)D_{q}-\frac{1}{2}\,\cos\theta\,\kappa_{n,\bf B}(x_{0})(\sin\theta\,p+\cos\theta\,q)^{2}\;, (8.17)
M2h\displaystyle M^{h}_{2} =hDqh((κˇκ)s+ζr)Dp+12sinθκn,𝐁(x0)(sinθp+cosθq)2,\displaystyle=hD_{q}-h\left((\check{\kappa}-\kappa)s+\zeta r\right)D_{p}+\frac{1}{2}\,\sin\theta\kappa_{n,\bf B}(x_{0})(\sin\theta\,p\;+\;\cos\theta q)^{2}\;,

where (r,s)=(sinθp+cosθq,cosθp+sinθq)(r,s)=(\sin\theta p+\cos\theta q,-\cos\theta p+\sin\theta q).

More precisely, we have the following comparison lemma (see Lemma 8.1 and [HelMo4, Lem. 15.4]).

Lemma 8.2.

For any τ1>0\tau_{1}>0, there exists C>0C>0 such that, for any u𝒟0u\in\mathcal{D}_{0},

(1+Ch2δ)qm,0h(u)(1Chτ1)qm,1h(u~)C((h2δ+hτ1)tu2+h6δτ1u2),(1+Ch^{2\delta})q^{h}_{m,0}(u)\geq(1-Ch^{\tau_{1}})q^{h}_{m,1}(\tilde{u})-C\big{(}\|(h^{2\delta}+h^{\tau_{1}})tu\|^{2}+h^{6\delta-\tau_{1}}\|u\|^{2}\big{)}\,,

where u~=uΦκ1\tilde{u}=u\circ\Phi_{\kappa}^{-1} is associated with uu by the transformation Φκ\Phi_{\kappa}.

By a unitary transformation, and after control of a commutator, we can reduce to a flat measure (dpdqdpdq instead of αˇ1dpdq\check{\alpha}_{1}dpdq) and obtain the new quadratic form defined as follows

qm,2h(v)=Ω0h[h2|Dtv|2+|tvM1hv|2+|M2hv|2]𝑑p𝑑q𝑑t,q^{h}_{m,2}(v)=\int_{\Omega^{h}_{0}}[h^{2}|D_{t}v|^{2}+|tv-M^{h}_{1}v|^{2}+|M^{h}_{2}v|^{2}]\;dpdqdt\;, (8.18)

with vv associated to uu by v=αˇ11/2u~v=\check{\alpha}_{1}^{1/2}\tilde{u}. In fact, we have [HelMo4, Eq. (15.29)]

(1+Ch1/2)qm,1(u~)+Ch3/2u2qm,2h(v).(1+Ch^{1/2})q_{m,1}(\tilde{u})+Ch^{3/2}\|u\|^{2}\geq q^{h}_{m,2}(v)\,. (8.19)

Let us consider the new model associated with the quadratic form in (8.18). We first observe that the result depends only on κˇκ\check{\kappa}-\kappa and on ζ\zeta. The proof is moreover uniform with respect to these parameters. As a consequence, if Φ=Φκ\Phi=\Phi_{\kappa} was the transformation introduced in (8.9), the inverse (for κ=0\kappa=0) Φ01\Phi_{0}^{-1}, more explicitly the transformation (p,q)(r~=sinθp+cosθq,s~=cosθp+sinθq)(p,q)\mapsto(\tilde{r}=\sin\theta p+\cos\theta q\;,\;\tilde{s}=-\cos\theta p+\sin\theta q) will bring us (in the new variables (r~,s~,t)(\tilde{r},\tilde{s},t)) to the initial model with κg\kappa_{g} replaced by 0, and κˇ\check{\kappa} replaced by κˇκg(x0)\check{\kappa}-\kappa_{g}(x_{0}). This can also be done by explicit computations.

Doing the transformations backwards, we are led to a magnetic Laplacian computed with a trivial metric κg=0\kappa_{g}=0 but with a new magnetic potential

a1(r,s)new=sinθ+cosθ(ζr+(κˇκg(x0))s),a_{1}(r,s)^{\rm new}=\sin\theta+\cos\theta(\zeta r+(\check{\kappa}-\kappa_{g}(x_{0}))s)\;, (8.20)

and

a2(r,s)new=cosθ+sinθ(ζr+(κˇκg(x0))s).a_{2}(r,s)^{\rm new}=-\cos\theta+\sin\theta(\zeta r+(\check{\kappa}-\kappa_{g}(x_{0}))s)\;. (8.21)

So the new model is not as simple as in the uniform magnetic field case (where κˇ=κg\check{\kappa}=\kappa_{g}) but it is the model in (5.1), which we have studied in the previous section with

η=κˇκg(x0),γ=κn,𝐁(x0).\eta=\check{\kappa}-\kappa_{g}(x_{0}),\quad\gamma=\kappa_{n,\mathbf{B}}(x_{0})\;. (8.22)

In fact, since vv is supported in Ω0h\Omega_{0}^{h}, we have,

qm,2h(v)=P0;γ,θh,η,ζv,v,q_{m,2}^{h}(v)=\langle P_{0;\gamma,\theta}^{h,\eta,\zeta}v,v\rangle, (8.23)

where P0;γ,θh,η,ζP_{0;\gamma,\theta}^{h,\eta,\zeta} is the operator in (5.1).

Remark 8.3.

We will choose τ1\tau_{1} in such a manner that 13<τ1<6δ43\frac{1}{3}<\tau_{1}<6\delta-\frac{4}{3}. This choice is possible when δ\delta satisfies 518<δ<13\frac{5}{18}<\delta<\frac{1}{3}.

8.1.3 Conclusion

We can now write a lower bound for the quadratic form q𝐀00h(u)q^{h}_{{\bf A}^{00}}(u) in (7.24), assuming that uH1(Q~h(x0,R0))u\in H^{1}(\tilde{Q}_{h}(x_{0},R_{0})) and Q~h(x0,R0)\tilde{Q}_{h}(x_{0},R_{0}) is the set introduced in (7.5). Let 518<δ<13\frac{5}{18}<\delta<\frac{1}{3} and 13<τ1<6δ43\frac{1}{3}<\tau_{1}<6\delta-\frac{4}{3}. Collecting Lemmas 8.1, 8.2, (8.19), (8.23) and Proposition 5.5, we get the existence of positive constants CC and ς0\varsigma_{0}, such that

q𝐀00h(u)(hΘ0+h43cconj(θ,κn,𝐁(x0))Ch43+ς0)u2Ch13+ς0tu2C(h2δ+hτ1)tu2q^{h}_{{\bf A}^{00}}(u)\geq\big{(}h\Theta_{0}+h^{\frac{4}{3}}c^{\rm conj}\big{(}\theta,\kappa_{n,\mathbf{B}}(x_{0})\big{)}-Ch^{\frac{4}{3}+\varsigma_{0}}\big{)}\|u\|^{2}\\ -Ch^{\frac{1}{3}+\varsigma_{0}}\|tu\|^{2}-C\|(h^{2\delta}+h^{\tau_{1}})tu\|^{2} (8.24)

where cconj(γ,θ)c^{\rm conj}(\gamma,\theta) is introduced in Proposition 5.2 with θ=θ(s0)\theta=\theta(s_{0}) the angle in (2.6).

8.2 The general case

We return now to the proof of the asymptotics of the lowest eigenvalue, λ1N(𝐀,h)\lambda_{1}^{N}({\bf A},h), of the operator P𝐀hP_{\bf A}^{h} in (1.3). Under Assumptions (C1)-(C2), we will prove the following lower bound:

λ1N(𝐀,h)Θ0h+γ^0,𝐁h43+𝒪(h43+η),\lambda_{1}^{N}({\bf A},h)\geq\Theta_{0}h+\widehat{\gamma}_{0,\mathbf{B}}h^{\frac{4}{3}}+{\mathcal{O}}(h^{\frac{4}{3}+\eta_{*}}), (8.25)

for some constant η>0\eta_{*}>0, where γ^0,𝐁\widehat{\gamma}_{0,\mathbf{B}} is introduced in (1.6).

Let uhu_{h} be a normalized ground state of P𝐀hP_{\bf A}^{h} , i.e.

λ1N(𝐀,h)=q𝐀h(uh)=(hi𝐀)uh2.\lambda_{1}^{N}({\bf A},h)=q_{\bf A}^{h}(u_{h})=\|(h\nabla-i{\bf A})u_{h}\|^{2}\,.

Consider 518<δ<13\frac{5}{18}<\delta<\frac{1}{3} and the following neighborhood of the curve Γ\Gamma,

Γδh={xΩ:dist(x,Ω)<hδ,distΩ(x,Γ)<hδ/2}.\Gamma_{\delta}^{h}=\{x\in\Omega~{}:~{}{\rm dist}(x,\partial\Omega)<h^{\delta},~{}{\rm dist}_{\partial\Omega}(x,\Gamma)<h^{\delta/2}\}\,. (8.26)

In terms of the (r,s,t)(r,s,t) coordinates introduced in Sec. 2.1,

Γδh={0<t<hδ,hδ/2<r<hδ/2}.\Gamma_{\delta}^{h}=\{0<t<h^{\delta},~{}h^{\delta/2}<r<h^{\delta/2}\}\,.

Let χhCc(Γδh;[0,1])\chi_{h}\in C_{c}^{\infty}(\Gamma_{\delta}^{h};[0,1]) be a smooth function such that

χh=1onΓδ,0h={xΩ:dist(x,Ω)<12hδ,distΩ(x,Γ)<12hδ/2}\chi_{h}=1~{}{\rm on~{}}\Gamma_{\delta,0}^{h}=\{x\in\Omega~{}:~{}{\rm dist}(x,\partial\Omega)<\frac{1}{2}h^{\delta},~{}{\rm dist}_{\partial\Omega}(x,\Gamma)<\frac{1}{2}h^{\delta/2}\}

and

|χh|=𝒪(hδ/2).|\nabla\chi_{h}|=\mathcal{O}(h^{-\delta/2}).

We introduce the function

wh=χhuh.w_{h}=\chi_{h}u_{h}\,. (8.27)

By Proposition 6.3, the eigenfunction uhu_{h} is exponentially small outside Γδh\Gamma_{\delta}^{h}, since by our choice of δ\delta we have hδ/2h1/4h^{\delta/2}\gg h^{1/4} and hδh1/2h^{\delta}\gg h^{1/2}. So we have

λ1N(𝐀,h)=q𝐀h(uh)=q𝐀h(wh)+𝒪(h),uh=wh+𝒪(h).\lambda_{1}^{N}({\bf A},h)=q_{\bf A}^{h}(u_{h})=q_{\bf A}^{h}(w_{h})+\mathcal{O}(h^{\infty}),\quad\|u_{h}\|=\|w_{h}\|+\mathcal{O}(h^{\infty})\,. (8.28)

Consider now a partition of unity of 3\mathbb{R}^{3}

j3|χj|2=1,j3|χj|2<,suppχjj+[1,1]3,\sum_{j\in\mathbb{Z}^{3}}|\chi_{j}|^{2}=1,\quad\sum_{j\in\mathbb{Z}^{3}}|\nabla\chi_{j}|^{2}<\infty,\quad{\rm supp\,}\chi_{j}\subset j+[-1,1]^{3}\,,

and introduce the following functions

wh,j=χj,δ(x)wh(x),χj,δ(x)=χj(hδx).w_{h,j}=\chi_{j,\delta}(x)w_{h}(x),\quad\chi_{j,\delta}(x)=\chi_{j}(h^{-\delta}x)\,. (8.29)

We can decompose the quadratic form q𝐀h(wh)q_{\bf A}^{h}(w_{h}) as follows

q𝐀h(wh)=j𝒥hq𝐀h(wh,j)+𝒪(h22δ),q_{\bf A}^{h}(w_{h})=\sum_{j\in\mathcal{J}_{h}}q_{\bf A}^{h}(w_{h,j})+\mathcal{O}(h^{2-2\delta}), (8.30)

where

𝒥h={j3:suppχj,δΩ}.\mathcal{J}_{h}=\{j\in\mathbb{Z}^{3}~{}:~{}{\rm supp\,}\chi_{j,\delta}\cap\Omega\not=\emptyset\}\,. (8.31)

Let C1>0C_{1}>0 be a fixed constant that we will choose later to be sufficiently large. We will estimate the energy q𝐀h(wh,j)q_{\bf A}^{h}(w_{h,j}) when the support of wh,jw_{h,j} is near the curve Γ\Gamma, or away from Γ\Gamma, independently. So we introduce the sets of indices

𝒥h1={j𝒥h:dist(suppχγ,δ,Γ)C1hδ}𝒥h2={j𝒥h:dist(suppχγ,δ,Γ)C1hδ}.\begin{aligned} \mathcal{J}_{h}^{1}&=\{j\in\mathcal{J}_{h}~{}:~{}{\rm dist}\big{(}{\rm supp\,}\chi_{\gamma,\delta},\Gamma\big{)}\leq C_{1}h^{\delta}\}\\ \mathcal{J}_{h}^{2}&=\{j\in\mathcal{J}_{h}~{}:~{}{\rm dist}\big{(}{\rm supp\,}\chi_{\gamma,\delta},\Gamma\big{)}\geq C_{1}h^{\delta}\}\end{aligned}\,. (8.32)

By Proposition 7.5,

j𝒥h2q𝐀h(wh,j)j𝒥h2((Θ0h+ch1+δ)wh,j2Ch43+ς0Mh(wj,h),\sum_{j\in\mathcal{J}_{h}^{2}}q_{\bf A}^{h}(w_{h,j})\geq\sum_{j\in\mathcal{J}_{h}^{2}}\Big{(}(\Theta_{0}h+c_{*}h^{1+\delta})\|w_{h,j}\|^{2}-Ch^{\frac{4}{3}+\varsigma_{0}}M_{h}(w_{j,h}\Big{)}\,, (8.33)

where Mh(wj,h)M_{h}(w_{j,h}) is introduced in (7.6). Notice that

Mh(wj,h)n=06hn/2Ωt(x)n(|χj,huh|2+2h1|χj,h(hi𝐀)uh|2+2h|(χhχj,h)|2|uh|2)dx.M_{h}(w_{j,h})\leq\sum_{n=0}^{6}h^{-n/2}\int_{\Omega}t(x)^{n}\Big{(}|\chi_{j,h}u_{h}|^{2}+2h^{-1}|\chi_{j,h}(h\nabla-i{\bf A})u_{h}|^{2}\\ +2h|\nabla(\chi_{h}\chi_{j,h})|^{2}|u_{h}|^{2}\Big{)}dx\,.

Since |χj,h|21\sum|\chi_{j,h}|^{2}\leq 1 and |(χj,hχh)|2=𝒪(h2δ)\sum|\nabla(\chi_{j,h}\chi_{h})|^{2}=\mathcal{O}(h^{-2\delta}), Proposition 6.3 together with (6.6) and (6.7) yield

j𝒥hMh(wj,h)=𝒪(1).\sum_{j\in\mathcal{J}_{h}}M_{h}(w_{j,h})=\mathcal{O}(1)\,.

Consequently, we infer from (8.33),

j𝒥h2q𝐀h(wh,j)(Θ0h+ch1+δ)(j𝒥h2wh,j2)Ch43+ς0.\sum_{j\in\mathcal{J}_{h}^{2}}q_{\bf A}^{h}(w_{h,j})\geq(\Theta_{0}h+c_{*}h^{1+\delta})\Big{(}\sum_{j\in\mathcal{J}_{h}^{2}}\|w_{h,j}\|^{2}\Big{)}-C^{\prime}h^{\frac{4}{3}+\varsigma_{0}}\,. (8.34)

For j𝒥h1j\in\mathcal{J}_{h}^{1}, we estimate q𝐀h(wh,j)q_{\bf A}^{h}(w_{h,j}) by collecting (8.24) and the estimates in Lemma 7.1 and 7.3. We start by picking R0>0R_{0}>0 and x0jΓx_{0}^{j}\in\Gamma, so that

suppwh,jQh(x0j){\rm supp\,}\,w_{h,j}\subset Q_{h}(x_{0}^{j})

where Qh(x0j)Q_{h}(x_{0}^{j}) is introduced in (7.2). Eventually, we find

j𝒥h1q𝐀h(wh,j)j𝒥h1(Θ0h+h4/3cconj(θj,κn,𝐁(xj0)))wh,j2Ch43+ς,\sum_{j\in\mathcal{J}_{h}^{1}}q_{\bf A}^{h}(w_{h,j})\geq\sum_{j\in\mathcal{J}_{h}^{1}}(\Theta_{0}h+h^{4/3}c^{\rm conj}(\theta_{j},\kappa_{n,\mathbf{B}}(x_{j}^{0})))\|w_{h,j}\|^{2}-Ch^{\frac{4}{3}+\varsigma_{*}}\,,

for some constant ς>0\varsigma_{*}>0, where

θj=θ(s0j)\theta_{j}=\theta(s_{0}^{j})

and (0,s0j,0)(0,s_{0}^{j},0) denote the coordinates of x0jx_{0}^{j} in the (r,s,t)(r,s,t)-frame (see Sec. 2 and Eq. 2.3). Note that we used Proposition 6.3 to control the term j𝒥h1twh,j2\sum_{j\in\mathcal{J}_{h}^{1}}\|tw_{h,j}\|^{2} appearing in (8.24); in fact j𝒥h1twh,j2=𝒪(h)\sum_{j\in\mathcal{J}_{h}^{1}}\|tw_{h,j}\|^{2}=\mathcal{O}(h).

Since cconj(θj,κn,𝐁(xj0))c^{\rm conj}(\theta_{j},\kappa_{n,\mathbf{B}}(x_{j}^{0})) is bounded from below by γ^0,𝐁\widehat{\gamma}_{0,\mathbf{B}} (see (1.6)), we get

j𝒥h1q𝐀h(wh,j)(Θ0h+γ^0,𝐁h4/3)j𝒥h1wh,j2Ch43+ς.\sum_{j\in\mathcal{J}_{h}^{1}}q_{\bf A}^{h}(w_{h,j})\geq(\Theta_{0}h+\widehat{\gamma}_{0,\mathbf{B}}h^{4/3})\sum_{j\in\mathcal{J}_{h}^{1}}\|w_{h,j}\|^{2}-Ch^{\frac{4}{3}+\varsigma_{*}}\,. (8.35)

Inserting (8.34) and (8.35) into (8.30), and using (8.28), we deduce the lower bound in (8.25), since 518<δ<13\frac{5}{18}<\delta<\frac{1}{3}.

9 Upper bound

Fortunately, the same quasi-mode constructed in [HelMo4, Sec. 12] (see also [Pan6] for a different formulation) yields an upper bound of the lowest eigenvalue λ1(𝐀,h)\lambda_{1}({\bf A},h) matching with the asymptotics in Theorem 1.4. More precisely, under Assumptions (C1)-(C2), we will prove that:

λ1N(𝐀,h)Θ0h+γ^0,𝐁h43+𝒪(h43+η),\lambda_{1}^{N}({\bf A},h)\leq\Theta_{0}h+\widehat{\gamma}_{0,\mathbf{B}}h^{\frac{4}{3}}+{\mathcal{O}}(h^{\frac{4}{3}+\eta^{*}}), (9.1)

for some constant η>0\eta^{*}>0, where γ^0,𝐁\widehat{\gamma}_{0,\mathbf{B}} is introduced in (1.6).

However, while computing the energy of the quasi-mode, we observe additional terms (not present in [HelMo4]) due to the non-homogeneity of the magnetic field. These terms are treated in Sec. 9.2.

9.1 The quasi-mode

The construction of the quasi-mode in [HelMo4] is quite lengthy and involves many auxiliary functions related to the de Gennes and Montgomery models (see (4.1) and (4.5)). We present here the definition of the quasi-mode along with a useful result from [HelMo4, Sec. 12].

9.1.1 Geometry and normal form

Select a point x0Ωx_{0}\in\partial\Omega such that the function in (1.7) satisfies

γ~0,𝐁(x0)=γ^0,𝐁.\widetilde{\gamma}_{0,\mathbf{B}}(x_{0})=\widehat{\gamma}_{0,\mathbf{B}}\,.

Let us assume that the coordinates of x0x_{0} in the (r,s,t)(r,s,t)-frame are (0,s0=0,t0)(0,s_{0}=0,t_{0}). The normal form of the effective magnetic potential in Lemma 7.2 now becomes

𝐀00=(A100A200A300)=(tsinθ+t(ζr+κˇs)cosθtcosθ+rtκcos+t(ζr+κˇs)sinθ+12γr20),{\bf A}^{00}=\left(\begin{array}[]{c}A^{00}_{1}\\ A^{00}_{2}\\ A^{00}_{3}\end{array}\right)=\left(\begin{array}[]{c}t\sin\theta+t(\zeta r+\check{\kappa}s)\cos\theta\\ -t\cos\theta+rt\kappa\cos+t(\zeta r+\check{\kappa}s)\sin\theta+\frac{1}{2}\gamma r^{2}\\ 0\end{array}\right)\,, (9.2)

where

θ=θ(s0),κ=κg(s0),γ=κn,𝐁(x0).\theta=\theta(s_{0}),\quad\kappa=\kappa_{g}(s_{0}),\quad\gamma=\kappa_{n,\mathbf{B}}(x_{0})\,. (9.3)

9.1.2 Structure of the quasi-mode

Consider two positive constants C0C_{0} and δ\delta such that 518<δ<13\frac{5}{18}<\delta<\frac{1}{3}. Let χ\chi be a smooth even function, valued in [0,1][0,1], equal to 11 on [14,14][-\frac{1}{4},\frac{1}{4}] and supported in [12,12][-\frac{1}{2},\frac{1}{2}]. We set

χh(s)=c1hδ/2χ(C01hδs),\chi_{h}(s)=c_{1}h^{-\delta/2}\chi(C_{0}^{-1}h^{-\delta}s)\,, (9.4)

where c1=C01/2(χ(σ)2𝑑σ)1/2c_{1}=C_{0}^{-1/2}\left(\int_{\mathbb{R}}\chi(\sigma)^{2}d\sigma\right)^{1/2}, so that χh\chi_{h} is normalized as follows,

|χh(s)|2𝑑s=1.\int_{\mathbb{R}}|\chi_{h}(s)|^{2}ds=1\,.

Our quasi-mode, uu, is supported in the set Qh(x0,R0,δ,ϵ2)Q_{h}(x_{0},R_{0},\delta,\epsilon_{2}) introduced in (7.2) and is of the form

u=eipˇ/h(1rκ)1/2u~u=e^{i\check{p}/h}(1-r\kappa)^{-1/2}\tilde{u} (9.5)

where (r,s,t)pˇ(r,s,t)(r,s,t)\mapsto\check{p}(r,s,t) is the function from Lemma 7.2 and the function u~\tilde{u} is of the form

u~(r,s,t)=exp(iργsh1/3)exp(irsinθscosθh1/2ξ0)χh(s)v(r,t),\tilde{u}(r,s,t)=\exp\left(-\frac{i\rho\gamma s}{h^{1/3}}\right)\exp\left(i\frac{r\sin\theta-s\cos\theta}{h^{1/2}}\xi_{0}\right)\chi_{h}(s)v(r,t)\,, (9.6)

where ξ0=Θ0\xi_{0}=\sqrt{\Theta_{0}} is given by (4.2), θ\theta and κ\kappa are introduced in (9.3).
The choice of ρ\rho and vv will be specified later444 ρ\rho is defined in (9.2). For the definition of vv, see (9.11), (9.12) and (9.13). so that, for some constants C,ς>0C,\varsigma_{*}>0, we have [HelMo4, Eq. (12.8)]

qM00h(v)(Θ0h+γ^0,𝐁h43+Ch43+ς)vL2(×+)2.q^{h}_{M^{00}}(v)\leq\Big{(}\Theta_{0}h+\widehat{\gamma}_{0,\mathbf{B}}h^{\frac{4}{3}}+Ch^{\frac{4}{3}+\varsigma_{*}}\Big{)}\|v\|_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}^{2}\,. (9.7)

Here qM00(v)q_{M^{00}}(v) arises while computing the quadratic form of the quasi-mode in (9.5). It is defined as follows [HelMo4, Eq. (12.9)],

qM00h(v)=×+(|(hDrM100v|2+|M200v|2+|hDtv|2)drdt,q^{h}_{M^{00}}(v)=\int_{\mathbb{R}\times\mathbb{R}_{+}}\Big{(}|(hD_{r}-M_{1}^{00}v|^{2}+|M_{2}^{00}v|^{2}+|hD_{t}v|^{2}\Big{)}drdt\,, (9.8)

where

M100(r,t)\displaystyle M_{1}^{00}(r,t) =sinθ(th1/2ξ0)\displaystyle=\sin\theta(t-h^{1/2}\xi_{0}) (9.9)
M200(r,t)\displaystyle M_{2}^{00}(r,t) =(1+2κr)1/2(cosθ(th1/2ξ0)+κcosθrtbγ2(r2h2/3ρ)).\displaystyle=(1+2\kappa r)^{1/2}\Big{(}-\cos\theta(t-h^{1/2}\xi_{0})+\kappa\cos\theta rt-b\frac{\gamma}{2}(r^{2}-h^{2/3}\rho)\Big{)}\,.

Notice that, by our normalization of χh\chi_{h}, we have

2×+|u~(r,s,t)|2𝑑r𝑑s𝑑t=vL2(×+)2.\int_{\mathbb{R}^{2}\times\mathbb{R}_{+}}|\tilde{u}(r,s,t)|^{2}drdsdt=\|v\|^{2}_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}\,. (9.10)

9.1.3 Definition of the auxiliary objects

Let us recall the definition of the function vv and the parameter ρ\rho given in [HelMo4, Sec. 12]. The function vv depends on hh and is selected in the following form (see [HelMo4, Eq. (12.14)])

v(r,t)=h5/12v0(r^,t^)v(r,t)=h^{-5/12}v_{0}(\hat{r},\hat{t}) (9.11)

where

(r^,t^)=(h1/3r,h1/2t).(\hat{r},\hat{t})=(h^{-1/3}r,h^{-1/2}t)\,.

The function v0v_{0} is selected as in [HelMo4, Eq. (12.28)]:

v0(r^,t^)=χ(C01hδ+13r^)χ(C01hδ+12t^)wh(r^,t^),v_{0}(\hat{r},\hat{t})=\chi(C_{0}^{-1}h^{-\delta+\frac{1}{3}}\hat{r})\chi(C_{0}^{-1}h^{-\delta+\frac{1}{2}}\hat{t})w_{h}(\hat{r},\hat{t}), (9.12)

In the sequel, we skip the hats from the notation. The function whw_{h} is defined as follows555 For the convenience of the reader, we will recall the heuristics behind the construction of whw_{h} in Subsection 9.1.4. [HelMo4, Eq. (12.22)]

wh(r,t)=φ0(t)ψ(r)+h1/6φ1(t)L10(r,Dr)ψ(r)+h1/3φ2(t)(L10(r,Dr))2ψ(r),w_{h}(r,t)=\varphi_{0}(t)\psi(r)+h^{1/6}\varphi_{1}(t)L_{1}^{0}(r,D_{r})\psi(r)+h^{1/3}\varphi_{2}(t)\big{(}L_{1}^{0}(r,D_{r})\big{)}^{2}\psi(r)\,, (9.13)

where φ0\varphi_{0} is the positive normalized ground state of the harmonic oscillator in (4.1),

φ1(t)=20((tξ0)φ0),φ2(t)=20((tξ0)φ1(tξ0)φ1,φ0φ0)\varphi_{1}(t)=2\mathcal{R}_{0}\big{(}(t-\xi_{0})\varphi_{0}\big{)},\quad\varphi_{2}(t)=2\mathcal{R}_{0}\big{(}(t-\xi_{0})\varphi_{1}-\langle(t-\xi_{0})\varphi_{1},\varphi_{0}\rangle\varphi_{0}\big{)}

and 0\mathcal{R}_{0} is the regularized resolvent introduced in (4.4). Notice that φ0,φ1\varphi_{0},\varphi_{1} and φ2\varphi_{2} are Schwartz functions (i.e. in 𝒮(+)\mathcal{S}(\mathbb{R}_{+}), see [FH06, Appendix A]). The definition of whw_{h} involves the differential operator

L10(r,Dr)=sinθDr12cosθγ(r2ρ)L_{1}^{0}(r,D_{r})=\sin\theta D_{r}-\frac{1}{2}\cos\theta\gamma(r^{2}-\rho) (9.14)

and a function ψ𝒮()\psi\in\mathcal{S}(\mathbb{R}) defined via the ground state ψ0\psi_{0} of the Montgomery model in (4.5) and the following phase function

φ(r)=γα(θ)(r36+ρr2),\varphi(r)=\gamma\alpha(\theta)\Big{(}\frac{r^{3}}{6}+\frac{\rho r}{2}\Big{)}\,,

where

α(θ)=sinθcosθ(1δ0)δ0sin2θ+cos2θ,\alpha(\theta)=\frac{\sin\theta\cos\theta(1-\delta_{0})}{\delta_{0}\sin^{2}\theta+\cos^{2}\theta}\,,

and δ0\delta_{0} the constant introduced in (4.2). We define now the function ψ(r)\psi(r) as follows

ψ(r)=(cd)1/12exp(iφ(r))ψ0((cd)1/6r)\psi(r)=\Big{(}\frac{c}{d}\Big{)}^{-1/12}\exp\big{(}i\varphi(r)\big{)}\,\psi_{0}\Big{(}\Big{(}\frac{c}{d}\Big{)}^{-1/6}r\Big{)}

where

c=cos2θ+δ0sin2θ,d=δ02γ2δ0sin2θ+cos2θ,c=\cos^{2}\theta+\delta_{0}\sin^{2}\theta,\quad d=\frac{\delta_{0}^{2}\gamma^{2}}{\delta_{0}\sin^{2}\theta+\cos^{2}\theta}\,, (9.15)

and we choose (see (4.5))

ρ=(cd)1/3ρ0.\rho=\Big{(}\frac{c}{d}\Big{)}^{1/3}\rho_{0}. (9.16)

We conclude by mentioning some estimates which follow easily from the definitions of vv and v0v_{0} in (9.11) and (9.12):

vL2(×+)2\displaystyle\|v\|_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}^{2} =1+𝒪(h1/6),\displaystyle=1+\mathcal{O}(h^{1/6})\,, (9.17)
×+rktn|v|2𝑑r𝑑t\displaystyle\int_{\mathbb{R}\times\mathbb{R}_{+}}r^{k}t^{n}|v|^{2}drdt =𝒪(hk3+n2)(k,n0),\displaystyle=\mathcal{O}(h^{\frac{k}{3}+\frac{n}{2}})\quad(k,n\geq 0)\,,
×+|hDrv|2𝑑r𝑑t\displaystyle\int_{\mathbb{R}\times\mathbb{R}_{+}}|hD_{r}v|^{2}drdt =𝒪(h5/3),×+|hDtv|2=𝒪(h).\displaystyle=\mathcal{O}(h^{5/3}),\quad\int_{\mathbb{R}\times\mathbb{R}_{+}}|hD_{t}v|^{2}=\mathcal{O}(h)\,.

9.1.4 Heuristics on the construction of whw_{h}.

Starting from the definition of the function vv in (9.12), the quadratic form in (9.7) becomes (after neglecting error terms in the magnetic potential)

qM00h(v)hq~h(wh)q^{h}_{M^{00}}(v)\approx h\tilde{q}^{h}(w_{h})

where

q~h(wh):=+2(|Dtwh|2+|(tξ0h1/6L10(r,Dr))wh|2+h1/3|L20(h,Dr)wh|2)𝑑r𝑑t,\tilde{q}^{h}(w_{h}):=\int_{\mathbb{R}^{2}_{+}}\Big{(}|D_{t}w_{h}|^{2}+\big{|}\big{(}t-\xi_{0}-h^{1/6}L_{1}^{0}(r,D_{r})\big{)}w_{h}\big{|}^{2}+h^{1/3}|L_{2}^{0}(h,D_{r})w_{h}|^{2}\Big{)}drdt\,,

L10(r,Dr)L_{1}^{0}(r,D_{r}) is introduced in (9.14) and

L20=cosθDr+12sinθγ(r2ρ).L_{2}^{0}=\cos\theta D_{r}+\frac{1}{2}\sin\theta\gamma(r^{2}-\rho)\,.

The construction of whw_{h} is based on minimizing

(+(|Dtwh|2+|(tξ0h1/6L10(r,Dr))wh|2)𝑑t)𝑑r,\int_{\mathbb{R}}\left(\int_{\mathbb{R}_{+}}\Big{(}|D_{t}w_{h}|^{2}+\big{|}\big{(}t-\xi_{0}-h^{1/6}L_{1}^{0}(r,D_{r})\big{)}w_{h}\big{|}^{2}\Big{)}dt\right)dr\,,

which amounts to finding the lowest eigenvalue of the operator

𝒯h:=Dt2+(tξ0h1/6L10(r,Dr))2.\mathcal{T}_{h}:=D_{t}^{2}+\big{(}t-\xi_{0}-h^{1/6}L_{1}^{0}(r,D_{r})\big{)}^{2}\,.

Writing

𝒯h=Dt2+(tξ0)22h1/6(tξ0)2L10(r,Dr)+h1/3(L10(r,Dr))2,\mathcal{T}_{h}=D_{t}^{2}+(t-\xi_{0})^{2}-2h^{1/6}(t-\xi_{0})^{2}L_{1}^{0}(r,D_{r})+h^{1/3}\big{(}L_{1}^{0}(r,D_{r})\big{)}^{2}\,,

it is natural to search for whw_{h} in the form in (9.13) and satisfying

𝒯hwh(μ0+μ1h1/6L10(r,Dr)+μ21/3(L10(r,Dr))2)wh0\mathcal{T}_{h}w_{h}-\left(\mu_{0}+\mu_{1}h^{1/6}L_{1}^{0}(r,D_{r})+\mu_{2}^{1/3}\big{(}L_{1}^{0}(r,D_{r})\big{)}^{2}\right)w_{h}\approx 0

in the following sense (after taking the coefficients of hi/6h^{i/6} to be 0, for i=0,1,2i=0,1,2)

(Dt2+(tξ0)2μ0)φ0\displaystyle\big{(}D_{t}^{2}+(t-\xi_{0})^{2}-\mu_{0}\big{)}\varphi_{0} =0\displaystyle=0
(Dt2+(tξ0)2μ0)φ1\displaystyle\big{(}D_{t}^{2}+(t-\xi_{0})^{2}-\mu_{0}\big{)}\varphi_{1} =μ1φ0\displaystyle=\mu_{1}\varphi_{0}
(Dt2+(tξ0)2μ0)φ2\displaystyle\big{(}D_{t}^{2}+(t-\xi_{0})^{2}-\mu_{0}\big{)}\varphi_{2} =μ2φ0+μ1φ1\displaystyle=\mu_{2}\varphi_{0}+\mu_{1}\varphi_{1}

Eventually, this leads to μ0=Θ0\mu_{0}=\Theta_{0}, μ1=0\mu_{1}=0, μ2=12μ′′(ξ0)\mu_{2}=\frac{1}{2}\mu^{\prime\prime}(\xi_{0}) and φ0,φ1,φ2\varphi_{0},\varphi_{1},\varphi_{2} as in (9.13).

9.2 Energy estimates

We will estimate the following energy arising from Lemma 7.3:

q𝐀00h(u~)=2×+(|(hDrA100)u~|2+(1+2κr)|(hDsA200)u~|2+|hDtu~|2)𝑑r𝑑s𝑑t,q^{h}_{{\bf A}^{00}}(\tilde{u})=\int_{\mathbb{R}^{2}\times\mathbb{R}_{+}}\Big{(}|(hD_{r}-A_{1}^{00})\tilde{u}|^{2}+(1+2\kappa r)|(hD_{s}-A_{2}^{00})\tilde{u}|^{2}+|hD_{t}\tilde{u}|^{2}\Big{)}drdsdt\,,

where A100,A200A_{1}^{00},A_{2}^{00} are introduced in (9.2).

Actually, q𝐀00h(u~)q^{h}_{{\bf A}^{00}}(\tilde{u}) is bounded from above by qM00(v)q_{M^{00}}(v) modulo error terms, where qM00(v)q_{M^{00}}(v) and vv are introduced in (9.8) and (9.6) respectively. Due to the non-homogeneity of the magnetic field, the error terms involve a quantity666 This is A(v)+B(v)A(v)+B(v) appearing in (9.20), which would equals 0 if the magnetic field were constant. introduced in (9.20) whose control has to be done carefully.

Due to the phase terms in the definition of u~\tilde{u} in (9.6), we have

q𝐀00h(u~)=2×+(|hDtu~|2+|(hDsA2,new00)u~|2+|(hDrA1,new00)u~|2)𝑑r𝑑s𝑑tq^{h}_{{\bf A}^{00}}(\tilde{u})=\int_{\mathbb{R}^{2}\times\mathbb{R}_{+}}\Big{(}|hD_{t}\tilde{u}|^{2}+|(hD_{s}-A_{2,\rm new}^{00})\tilde{u}|^{2}+|(hD_{r}-A_{1,\rm new}^{00})\tilde{u}|^{2}\Big{)}drdsdt

where

(A1,new00A2,new00)=(M1,ζ00M2,ζ00)+(κˇcosθstκˇsinθst),\left(\begin{array}[]{c}A_{1,\rm new}^{00}\vskip 6.0pt plus 2.0pt minus 2.0pt\\ A_{2,\rm new}^{00}\end{array}\right)=\left(\begin{array}[]{c}M_{1,\zeta}^{00}\vskip 6.0pt plus 2.0pt minus 2.0pt\\ M_{2,\zeta}^{00}\end{array}\right)+\left(\begin{array}[]{c}\check{\kappa}\cos\theta st\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \check{\kappa}\sin\theta st\end{array}\right)\,,

and

M1,ζ00(r,t)\displaystyle M_{1,\zeta}^{00}(r,t) =sinθ(th1/2ξ0)+ζcosθrt\displaystyle=\sin\theta(t-h^{1/2}\xi_{0})+\zeta\cos\theta rt (9.18)
M2,ζ00(r,t)\displaystyle M_{2,\zeta}^{00}(r,t) =(1+2κr)1/2(cosθ(th1/2ξ0)\displaystyle=(1+2\kappa r)^{1/2}\Big{(}-\cos\theta(t-h^{1/2}\xi_{0})
+(κcosθ+ζsinθ)rtγ2(r2h2/3ρ)).\displaystyle\qquad+(\kappa\cos\theta+\zeta\sin\theta)rt-\frac{\gamma}{2}(r^{2}-h^{2/3}\rho)\Big{)}\,.

Since the function sχh(s)s\mapsto\chi_{h}(s) is even, we have

(hDsκˇsinθst)u,M2,ζ00uL2(2×+)=0\langle(hD_{s}-\check{\kappa}\sin\theta st)u,M_{2,\zeta}^{00}u\rangle_{L^{2}(\mathbb{R}^{2}\times\mathbb{R}_{+})}=0

and

κˇcosθstu,(hDrM1,ζ00)uL2(2×+)=0.\langle\check{\kappa}\cos\theta st\,u,(hD_{r}-M_{1,\zeta}^{00})u\rangle_{L^{2}(\mathbb{R}^{2}\times\mathbb{R}_{+})}=0\,.

Moreover, we have the estimates

(hDsκˇsinθst)u~L2(2×+)\displaystyle\|(hD_{s}-\check{\kappa}\sin\theta st)\tilde{u}\|_{L^{2}(\mathbb{R}^{2}\times\mathbb{R}_{+})} C×+(h22δ+h2δt2)|v|2𝑑r𝑑t\displaystyle\leq C\int_{\mathbb{R}\times\mathbb{R}_{+}}(h^{2-2\delta}+h^{2\delta}t^{2})|v|^{2}drdt
=𝒪(h22δ+h2δ+1),\displaystyle=\mathcal{O}(h^{2-2\delta}+h^{2\delta+1})\,,

and

κˇcosθstuL2(2×+)Ch2δ×+t2|v|2𝑑r𝑑t=𝒪(h2δ+1).\|\check{\kappa}\cos\theta st\,u\|_{L^{2}(\mathbb{R}^{2}\times\mathbb{R}_{+})}\leq Ch^{2\delta}\int_{\mathbb{R}\times\mathbb{R}_{+}}t^{2}|v|^{2}drdt=\mathcal{O}(h^{2\delta+1}).

Notice that we used (9.17) and also that |s|C0hδ|s|\leq C_{0}h^{\delta} in the support of u~\tilde{u}. Consequently, we get

q𝐀00h(u~)\displaystyle q^{h}_{{\bf A}^{00}}(\tilde{u}) ×+(|(hDrM1,ζ00)v|2+|M2,ζ00v|2+|hDtv|2)𝑑r𝑑t\displaystyle\leq\int_{\mathbb{R}\times\mathbb{R}_{+}}\Big{(}|(hD_{r}-M_{1,\zeta}^{00})v|^{2}+|M_{2,\zeta}^{00}v|^{2}+|hD_{t}v|^{2}\Big{)}drdt (9.19)
+𝒪(h22δ+h2δ+1).\displaystyle\qquad+\mathcal{O}\big{(}h^{2-2\delta}+h^{2\delta+1}\big{)}\,.

Let us now reduce the computations to the potentials M100M_{1}^{00} and M200M_{2}^{00} in (9.9) which amount to M1,ζ00M_{1,\zeta}^{00} and M2,ζ00M_{2,\zeta}^{00} with ζ=0\zeta=0. A straightforward computation yields,

(hDrM1,ζ00)vL2(×+)2+M2,ζ00vL2(×+)2=(hDrM100)vL2(×+)2+M200vL2(×+)2+ζ(A(v)+B(v)),\|(hD_{r}-M_{1,\zeta}^{00})v\|^{2}_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}+\|M_{2,\zeta}^{00}v\|^{2}_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}\\ =\|(hD_{r}-M_{1}^{00})v\|^{2}_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}+\|M_{2}^{00}v\|^{2}_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}+\zeta\big{(}A(v)+B(v)\big{)}, (9.20)

where

A(v)\displaystyle A(v) :=ζcos2θrtvL2(×+)22cosθRe(hDrM100)v,rtvL2(×+)\displaystyle:=\zeta\cos^{2}\theta\,\|rtv\|^{2}_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}-2\cos\theta\,{\rm Re}\langle(hD_{r}-M_{1}^{00})v\,,\,rtv\rangle_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}
B(v)\displaystyle B(v) :=ζsin2θrtvL2(×+)2+2sinθReM200v,rtvL2(×+)\displaystyle:=\zeta\sin^{2}\theta\,\|rtv\|^{2}_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}+2\sin\theta\,{\rm Re}\langle M_{2}^{00}v\,,\,rtv\rangle_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}

and by (9.17)

rtvL2(×+)2=𝒪(h5/3),hDrv,rtvL2(×+)=𝒪(h5/3).\|rtv\|^{2}_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}=\mathcal{O}(h^{5/3}),\quad\langle hD_{r}v\,,\,rtv\rangle_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}=\mathcal{O}(h^{5/3})\,.

So, we end up with estimating

F(v):=(cosθM100+sinθM200)v,rtvL2(×+).F(v):=\langle(\cos\theta M_{1}^{00}+\sin\theta M_{2}^{00})v\,,\,rtv\rangle_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}\,.

Notice that

cosθM100(r,t)+sinθM200(r,t)=cosθsinθ(1(1+2κr)2)(th1/2ξ0)+(1+2κr)2cosθ(κcosθrtγ2(r2h2/3ρ)).\cos\theta M_{1}^{00}(r,t)+\sin\theta M_{2}^{00}(r,t)=\cos\theta\sin\theta\big{(}1-(1+2\kappa r)^{2}\big{)}(t-h^{1/2}\xi_{0})\\ +(1+2\kappa r)^{2}\cos\theta\Big{(}\kappa\cos\theta rt-\frac{\gamma}{2}\Big{(}r^{2}-h^{2/3}\rho\Big{)}\Big{)}\,.

By expanding

(1+2κr)1/2=1+κr+𝒪(r2)(r0),(1+2\kappa r)^{1/2}=1+\kappa r+\mathcal{O}(r^{2})\quad(r\to 0)\,,

we observe that, for |r|r0|r|\leq r_{0} and r0r_{0} sufficiently small,

|cosθM100(r,t)+sinθM200(r,t)|C(r2+t2+h2/3),|\cos\theta M_{1}^{00}(r,t)+\sin\theta M_{2}^{00}(r,t)|\leq C(r^{2}+t^{2}+h^{2/3})\,,

so we get by (9.17) and the Cauchy-Schwarz inequality that

F(v)=𝒪(h3/2).F(v)=\mathcal{O}(h^{3/2})\,.

Therefore, A(v)+B(v)=𝒪(h3/2)A(v)+B(v)=\mathcal{O}(h^{3/2}) and we deduce from (9.20) and (9.19) that

q𝐀00h(u~)qM00h(v)+𝒪(h22δ+h2δ+1+h3/2),q^{h}_{{\bf A}^{00}}(\tilde{u})\leq q^{h}_{M^{00}}(v)+\mathcal{O}\big{(}h^{2-2\delta}+h^{2\delta+1}+h^{3/2}\big{)}\,, (9.21)

where qM00h(v)q^{h}_{M^{00}}(v) is introduced in (9.8).

9.3 Conclusion

Collecting (9.21) and (9.7), we get

q𝐀00h(u~)(Θ0h+γ^0,𝐁h43+Ch43+η)vL2(×+)2+Rh(v)q^{h}_{{\bf A}^{00}}(\tilde{u})\leq(\Theta_{0}h+\widehat{\gamma}_{0,\mathbf{B}}h^{\frac{4}{3}}+Ch^{\frac{4}{3}+\eta}\Big{)}\|v\|_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}^{2}+R_{h}(v)

where

Rh(v)=𝒪(h22δ+h2δ+1+h3/2)=𝒪(h43+η^)R_{h}(v)=\mathcal{O}\big{(}h^{2-2\delta}+h^{2\delta+1}+h^{3/2}\big{)}=\mathcal{O}(h^{\frac{4}{3}+\hat{\eta}})

for some η^>0\hat{\eta}>0, thanks to the condition 518<δ<13\frac{5}{18}<\delta<\frac{1}{3}.

We insert this into Lemma 7.3 with uu given in (9.5). Notice that uu satisfies (7.7) with Mh(u)=𝒪(1)M_{h}(u)=\mathcal{O}(1). So by Lemma 7.3 and (9.10), we get for some η>0\eta_{*}>0

q𝐀h(u)(Θ0h+γ^0,𝐁h43+Ch43+η)vL2(×+)2.q^{h}_{{\bf A}}(u)\leq(\Theta_{0}h+\widehat{\gamma}_{0,\mathbf{B}}h^{\frac{4}{3}}+Ch^{\frac{4}{3}+\eta_{*}}\Big{)}\|v\|_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}^{2}\,.

Comparing (9.10) and (9.5), we get by (2.20),

u2=(1+𝒪(h2δ))vL2(×+)2.\|u\|^{2}=\big{(}1+\mathcal{O}(h^{2\delta})\big{)}\|v\|^{2}_{L^{2}(\mathbb{R}\times\mathbb{R}_{+})}\,. (9.22)

Applying the min-max principle, and noticing that 1+2δ>431+2\delta>\frac{4}{3} for 518<δ<13\frac{5}{18}<\delta<\frac{1}{3}, we finish the proof of (9.1).

Acknowledgments

Preliminary discussions of the first author on this problem with Xingbin Pan more than twelve years ago are acknowledged. This work was initiated while the second author visited the Laboratoire de Mathématiques Jean Leray (LMJL) at Nantes Université in 2021. The authors would like to thank the support from the Fédération de recherche Mathématiques des Pays de Loire and Nantes Université.

References

  • [DaHe] M. Dauge, B. Helffer : Eigenvalues variation I, Neumann problem for Sturm-Liouville operators. Journal of Differential Equations, Vol. 104, n2, august 1993, p. 243-262.
  • [DHKW] U. Dierkes, S. Hildebrandt, A. Kuster, O. Wohlrab : Minimal Surfaces I, Springer-Verlag, Berlin (1992).
  • [FoHe1] S. Fournais, B. Helffer : On the Ginzburg-Landau critical field in three dimensions. Commun. Pure Appl. Anal. 62(2), 215–241 (2009).
  • [FoHe2] S. Fournais, B. Helffer : Spectral methods in surface superconductivity. Progress in Non Linear PDE’s 77 (Birkhäuser) (2010).
  • [FKPa] S. Fournais, A. Kachmar, X.B. Pan. Existence of surface smectic states of liquid crystals. J. Funct. Anal. 274(3), 900–958 (2018).
  • [FKP] S. Fournais, A. Kachmar, M. Persson. The ground state energy of the three dimensional Ginzburg-Landau functional. II: Surface regime. J. Math. Pures Appl. 99(3), 343–374 (2013).
  • [FMP] S. Fournais, J.P. Miqueu, X.B. Pan. Concentration behavior and lattice structure of 3D surface superconductivity in the half space. Math. Phys. Anal. Geom. 22(2), Paper No. 12, 33 p. (2019).
  • [FS] S. Fournais, M.P. Sundqvist. Lack of diamagnetism and the Little–Parks effect. Commun. Math. Phys. 337(1), 191–224 (2015).
  • [HKR] B. Helffer, A. Kachmar, N. Raymond. Magnetic confinement for the 3D Robin Laplacian. To appear in Pure and Applied Functional Analysis, arXiv:2005.03845 (2020).
  • [HeKo] B. Helffer and Yu. A. Kordyukov. Spectral gaps for periodic Schrödinger operators with hypersurface magnetic wells, In “Mathematical results in quantum mechanics”, Proceedings of the QMath10 Conference Moieciu, Romania 10 - 15 September 2007, World Sci. Publ., Singapore, 2008.
  • [HeLe] B. Helffer and M. Léautaud. On critical points of the eigenvalues of the Montgomery family of quartic oscillators. Work in Progress.
  • [HelMo1] B. Helffer and A. Mohamed. Semiclassical analysis for the ground state energy of a Schrödinger operator with magnetic wells. Journal of Functional Analysis 138 (1) (1996), p. 40-81.
  • [HelMo2] B. Helffer and A. Morame. Magnetic bottles in connection with superconductivity. J. Functional Analysis 185 (2) (2001), p. 604-680. (See Erratum available at http://mahery.math.u-psud.fr/ helffer/erratum164.pdf, 2005).
  • [HelMo3] B. Helffer and A. Morame. Magnetic bottles for the Neumann problem: The case of dimension 33. Proc. Indian Acad. Sci. (Math.Sci.) 112 (1) (2002), p. 71-84.
  • [HelMo4a] B. Helffer and A. Morame. Magnetic bottles for the Neumann problem: Curvature effect in the case of dimension 33 (General case). mp_\_arc 02-145 (2002).
  • [HelMo4] B. Helffer and A. Morame. Magnetic bottles for the Neumann problem: Curvature effect in the case of dimension 33 (General case). Annales de l’ENS 37 (2004), p. 105-170.
  • [HePa2] B. Helffer, X-B. Pan : Reduced Landau-de Gennes functional and surface smectic state of liquid crystals. JFA 255 (11), p. 3008-3069 (2008).
  • [HeSj] B. Helffer, J. Sjöstrand : Multiple wells in the semiclassical limit I.
    Comm. in PDE 9(4), p. 337-408, (1984).
  • [HR] F. Hérau, N. Raymond. Semi-classical gaps of the 3D Neumann Laplacian with constant magnetic field. arXiv:2203.06048v1 (March 2022).
  • [LuPa5] K. Lu, X-B. Pan : Surface nucleation of superconductivity in 33-dimension. J. of Differential Equations 168(2) (2000), p. 386-452.
  • [Mi] G. Miranda. Non–monotonicity for the 3D magnetic Robin Laplacian. arXiv:2202.01001 (2022).
  • [Mon] R. Montgomery : Hearing the zero locus of a magnetic field. Comm. Math. Physics 168 (1995), p. 651-675.
  • [Pan2] X. -B. Pan. Landau-de Gennes model of liquid crystals and critical wave number. Comm. Math. Phys. 239 (1-2) (2003), p. 343-382.
  • [Pan3] X. -B. Pan. Surface superconductivity in 33 dimensions. Trans. AMS. 356 (10) (2004), p. 3899-3937.
  • [Pan5] X. -B. Pan. Analogies between superconductors and liquid crystals: nucleation and critical fields. Advanced Studies in Pure Mathematics (ASPM) 47-2 (2007). Asymptotic Analysis and Singularities. p. 479-518.
  • [Pan6] X. B. Pan. An eigenvalue variation problem of magnetic Schrödinger operator in three dimensions. Disc. Contin. Dyn. Systems, Ser. A, special issue dedicated to the 70th birthday of Prof. P. Bates, Vol. 24, No.3 (2009), p. 933-978.
  • [PanKw] X. -B. Pan. and K.H. Kwek. Schrödinger operators with non-degenerately vanishing magnetic fields in bounded domains. Transactions of the AMS 354 (10) (2002), p. 4201-4227.
  • [Sp] M.  Spivak. Differential Geometry, Publish or Perish Inc., Houston (1975).