Helical magnetic fields and semi-classical asymptotics of the lowest eigenvalue
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 be an open and bounded set with a smooth boundary . Let us consider a smooth magnetic field (so should be closed) which will always be assumed to satisfy
(1.1) |
where is a constant. Without loss of generality, we assume from now on that . Let be a magnetic potential such that
(1.2) |
We are interested in the analysis of the lowest eigenvalue of the Neumann realization of the Schrödinger operator in with magnetic field
(1.3) |
We introduce the following assumptions.
Assumption 1.1 (C1).
The set of boundary points where is tangent to , i.e.
(1.4) |
is a regular submanifold of :
(1.5) |
Here is the differential defined on functions on and is the unit inward normal of .
Assumption 1.2 (C2).
The set of points where is tangent to is finite.
These assumptions are rather generic and for instance satisfied for ellipsoids, when is constant. When is constant, the above assumptions hold for the sphere with a helical magnetic field (see Sec. 3).
Let us introduce the constant involving the “magnetic curvature” in (1.5), which is defined by
(1.6) |
where
(1.7) |
Here is the oriented, unit tangent vector to at the point , and are spectral quantities relative to the De Gennes and Montgomery operators which will be introduced in (4.2) and (4.5).
When is constant, the following two-term asymptotics of has been established by Helffer-Morame [HelMo4] and Pan [Pan3].
Theorem 1.3.
Let us assume that is constant. Then, if and satisfy (C1)-(C2), there exists such that the lowest eigenvalue satisfies as
(1.8) |
The aim of this paper is to prove that Theorem 1.3 also holds under the weaker assumption that is constant.
Theorem 1.4.
Under the assumptions (C1)-(C2), if is constant, then the asymptotics in (1.8) holds for the lowest eigenvalue .
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]
(1.9) |
Here is a given constant. In this situation (), [Pan6] derived an upper bound on the eigenvalue , 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 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 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 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 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).
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 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.
2 Adapted coordinates
We recall a rather standard choice of coordinates in the neighborhood of .
2.1 Description of the coordinates
Let be the Riemannian metric on , which induces a Riemanian metric on . Given two vector fields of , we denote by
(2.1) |
Consider a direct frame along such that
-
•
is an oriented unit tangent vector of ;
-
•
, hence determining an oriented normal to the curve in the tangent space to .
For , let be the geodesic that passes through and is normal to . Let . In some neighborhood of , we can introduce new coordinates as follows:
-
•
For , is defined by ;
-
•
For , is defined by , where denotes the (geodesic) distance in ;
-
•
is parameterized by arc-length so that defines , and for , defines ;
-
•
For , the geodesic passing through is parameterized by arclength , so that defines and defines .
In this way, we observe that
(2.2) |
is a local diffeomorphism. Thus, we can pick a sufficiently small such that
(2.3) |
is a diffeomorphism, whose image is a neighborhood of parameterized by . Within these coordinates, means that we are on , and means we are on the curve . We can then compute
(2.4) |
It is convenient to express the magnetic field along as follows
(2.5) |
where is the angle defined by
(2.6) |
2.2 The metric in the new coordinates
Let us consider an arbitrary point and a neighborhood of such that the adapted coordinates introduced in (2.2) and (2.3) are valid. Modulo a translation, we can center the coordinates at so that are the coordinates of 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
(2.7) |
and the Riemanian metric becomes [HelMo4, Eq. (8.26)]
(2.8) |
where:
-
•
is the first fundamental form on ;
-
•
is the second fundamental form on ;
-
•
is the third fundamental form on .
The matrix of the metric takes the form
(2.9) |
whose inverse is
(2.10) |
We will express these matrices in a more pleasant form involving, in particular, the curvatures on the boundary. To that end, let be an arc-length parameterization of near , so that , and . We can introduce the geodesic and normal curvatures at , and , as follows
(2.11) |
The choice of our coordinates ensures that the metric is diagonal on [HelMo4, Lem. 8.2]
(2.12) |
with
(2.13) |
and
(2.14) |
Then, with (2.7), we have for the determinant of the matrix of (see [HelMo4, Eq. (8.29) & (8.30)]),
(2.15) |
and
(2.16) |
where and 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
(2.17) |
where (recall (2.7))
(2.18) |
Given a function , we assign to it the function defined by
(2.19) |
By the considerations in Subsection 2.2 on the Riemanian metric, if , then and
(2.20) |
Moreover, assuming supported in , we have the quadratic form formula [HelMo4, Eq. (8.27)]
(2.21) | ||||
where the new magnetic potential is assigned to by the relation
(2.22) |
and after performing a (local) gauge transformation, we may assume that
(2.23) |
The operator in (1.3) can be expressed in the new coordinates as follows [HelMo4, Eq. (8.28)]
(2.24) |
3 Helical magnetic fields
3.1 Preliminaries
Let and consider the magnetic potential
(3.1) |
which generates the magnetic field
(3.2) |
with constant intensity
(3.3) |
We will verify that Assumptions C1-C2 hold for this particular magnetic field in the case where is the unit ball. In particular, with in mind that and are introduced in (1.6) and (1.7) respectively and that and 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
(3.4) |
and for , the equality,
(3.5) |
where is a constant and is explicitly computed (see Proposition 3.4).
The inward normal of along is
(3.6) |
The restriction of the magnetic field to the boundary is then tangent to on the following set
(3.7) |
3.2 is a regular curve
For , the equation reads as follows
(3.8) |
Proposition 3.1.
The set introduced in (3.7) is a regular curve.
Proof.
The proof follows by constructing an atlas on ,
which turns to a regular curve.
Let us introduce the charts and which cover . These charts are obtained by expressing and in (3.8) in terms of , provided that . We write for
(3.9) |
Then (3.8) becomes, for ,
(3.10) |
which in turn yields
In this way, we get two branches of parameterized by and defined as follows
and
Both of the foregoing branches represent regular curves. Furthermore, and can be extended by continuity to the interval , yielding a continuous representation of all .
Now we introduce the charts and that cover the points . In a neighborhood of , we parameterize a branch of with respect to as follows
With this in hand, (3.10) continues to hold for and we can write again for some . Consequently, we get two regular branches of defined as follows
and
3.3 Explicit formulas in adapted coordinates
Note that and parameterize all of . By symmetry considerations, we will compute, on only,
(3.11) |
First we note that on and introduce the arc-length parameter
of , which satisfies
(3.12) |
Clearly, can be expressed in terms of the arc-length parameter as with
(3.13) |
The arc-length parameterization is now given by
(3.14) |
and consequently, with , we have
(3.15) |
and
We also introduce the normal vector to on ,
We are now ready to prove that our magnetic field verifies the condition (C2) appearing in Assumption 1.2
Proposition 3.2.
Let be the magnetic field introduced in (3.2). For all , we have
In particular, satisfies the condition .
Proof.
It is straightforward to compute
(3.16) |
which holds for all and . Similarly, we can compute for all , and get that (3.16) holds globally on , since is a regular curve. Finally, is orthogonal to if and only if , 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 be the magnetic field introduced in (3.2). For all , we have
(3.17) |
In particular, satisfies the condition .
Proof.
By Proposition 3.1, is a regular curve. So all we need to verify that satisfies , is to derive (3.17) and observe that it yields everywhere on .
Consider with , i.e. . At the point , the geodesic normal to the curve is the great circle (of center and radius ) in the plane. A point on can be described by the corresponding vector as follows
where is the angle between and ; hence is an arc-length length parameter of , and for , .

Now, we can introduce the coordinates in a neighborhood of as follows (see Fig. 1)
(3.18) |
For , we would like to compute . We will show that and end up with the computation of
Notice that, by (3.15), we have
and we observe that by (3.18),
(3.19) |
In particular we have
Now, using (3.13) and (3.15), we get from (3.1) that
(3.20) |
Moreover, by (3.19) we have
and
(3.21) | ||||
Summing up, we deduce from (3.20) and (3.21) that
(3.22) |
We also observe that and we get
(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):
for all with .
Consequently, we can compute the quantity appearing in the two terms asymptotics by computing and determining where the infimum is attained.
Proposition 3.4.
Let
The following holds:
-
1.
If , then
-
2.
If , then
and the minimum is attained on the points
and
Remark 3.5.
In the case where is the unit ball and the magnetic field is constant, , we have and is constant on . Proposition 3.4 shows a quite different phenomenon when only the intensity of is constant, . In fact, is no more constant along and may have two symmetric minimum points, , 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 and . Then
where
We have to minimize on . Notice that
and the equation has a unique positive solution, which is the solution of
This solution is given by
and observe that for and for . Then, for ,
while for ,
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 , we consider the harmonic oscillator on :
(4.1) |
with Neumann boundary condition at . We denote by its lowest eigenvalue. admits a unique minimum at a point which in addition is non-degenerate. This leads to introduce the spectral constants, and :
(4.2) |
where .
Moreover and that .
is called the de Gennes constant.
If denotes the positive and normalized ground state of ,
(4.3) |
which amounts to saying, via the Feynman-Hellmann formula, that . We also introduce the regularized resolvent as follows
(4.4) |
4.2 The Montgomery model
Here we refer to [HelMo1] and [PanKw]. In Theorem 1.3, the constant is related to the Montgomery model [Mon] whose spectral analysis has a long story including recently (see [HeLe] and references therein). For , we introduce, in , the operator
and denote its lowest eigenvalue by . Then
(4.5) |
where is the unique minimum of , which has been later shown to be non degenerate [HeKo]. Finally, the normalized positive ground state of belongs to the Schwartz space and is an even function.
5 Model operator for non-uniform magnetic fields
Given real parameters and , we consider the operator
(5.1) |
on (actually in a neighborhood of ). Let us fix a positive constant . We assume that
(5.2) |
We note, when , we recover the model studied in [HelMo4, Sec. 11]. Our aim is to compare this situation with that when . 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 is sufficiently small.
Let us look at this model more carefully. We proceed essentially like in the case . We do the following scaling
(5.3) |
After division by , this leads to (forgetting the hats)
(5.4) |
on .
Hence we have
Unlike the case where , we can no more perform a partial Fourier transform in the -variable. But we can rewrite this operator as in the following lemma.
Lemma 5.1.
It holds,
(5.5) |
where
(5.6) |
Proof of Lemma 5.1.
When , this is the operator studied in [HelMo4], modulo a Fourier transformation with respect to the variable. Let us recall the following important result [HelMo4, Lem. 13.4] corresponding to the case .
Proposition 5.2 (Helffer-Morame).
For any , and , there exist positive constants and such that, for all , , and , we have, for any
(5.8) |
where
and is the operator introduced in (5.1).
Remark 5.3.
The underlying estimate in Proposition 5.2 is in fact
We can not directly compare and but this can be done by introducing a small perturbation of whose spectrum is just lifted. To achieve this goal we introduce for
where we have modified the coefficient of by . Heuristically this leads to a maximal shift of the bottom of the spectrum by . More precisely, we show by a slight variation of the argument in [HelMo4, Lem. 13.3]
Proposition 5.4.
For all , for any , and , there exist positive constants and such that, for all , , and , we have, for any
(5.9) |
Note that the estimate in Proposition 5.2 holds without constraint on the
support of the function in . This will not be the case for .
We now compare and when
and satisfies (5.2).
Let us fix
(5.10) |
The estimates below hold uniformly with respect to , and satisfying (5.2).
Comparing and in (5.7), we find111We use with , and ., for all ,
Consequently,
(5.11) |
This implies (see (5.7) and the condition on the support of ),
(5.12) |
where we used (see (5.7))
in the support of .
By (5.9) and (5.12) we have
(5.13) |
Note that by (5.10) we have
for some .
Consequently, there exist , and such that, ,
(5.14) |
for any .
By coming back to the initial coordinates, we get the following generalization of Proposition 5.2.
Proposition 5.5.
Let and be given. There exist positive constants , , and , such that, for all , and , we have, for any ,
(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 in (1.3) are localized on the boundary near the curve where the magnetic field is tangent to the boundary . The localization is related with the analysis of a family of model operators in the half-space [LuPa5].
Consider and the Neumann realization in of the operator,
where .
More precisely, is self-adjoint in with the following domain
We denote by
(6.1) |
Proposition 6.1.
The following properties hold for the lowest eigenvalue of :
-
•
For all , .
-
•
is monotone increasing and .
-
•
.
-
•
As , .
Here we recall that and are introduced in (4.2).
Let us return to the magnetic field in (1.1). Recall that, for , satisfies , and it is uniquely defined when is sufficiently close to the boundary. For all , we introduce by
(6.2) |
Hence implies that is tangent to at , in other words that belongs to (see (1.4)). Now we recall the following lower bound related to the operator established in [HelMo4, Thm. 4.3]:
Proposition 6.2.
If additionally , we have for some positive constant the stronger lower bound
Combining the lower bound in Proposition 6.2 with the following leading term expansion of the lowest eigenvalue (see [HelMo4, Thm. 4.4])
(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 , there exists a positive constant such that, if is a normalized bound state of with eigenvalue , then as ,
(6.4) |
Furthermore, there exist constants such that, as ,
where
(6.5) |
and is the geodesic distance on .
Hence we have two levels of localization, first a strong one near and then an additional but weaker one near . 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 be a normalized ground state of .
- 1.
- 2.
7 Estimating the quadratic form
7.1 A comparison estimate
We fix and satisfying
(7.1) |
We also fix , , and introduce for the set
(7.2) |
where are introduced in (2.2) and, since ,
(7.3) |
For simplicity, we omit most of the time the reference to and .
Let be the magnetic potential associated with via (2.22), with (see (2.7)). We introduce the following magnetic potential
(7.4) |
which is the quadratic Taylor expansion of at . We introduce the quadratic form associated with the magnetic potential as follows
where
(7.5) |
and (see (2.11))
The next lemma compares the quadratic forms and introduced in (2.21). The errors that will arise are controlled by the following energy
(7.6) |
where . Notice that,
(7.7) | ||||
Lemma 7.1.
There exist constants such that, for all and satisfying , we have
(7.8) |
Proof.
Let us recall two useful estimates whose proof does not require that the magnetic field is constant (see [HelMo4, Lem. 10.1]):
(7.9) |
and
(7.10) |
In the sequel we use the notation in the following manner
(7.11) |
Since we have assumed (7.1), we have
We can now estimate the error terms appearing in (7.9) and (7.10). We deduce from (7.7) (a) that
where we write instead of for the sake of simplicity.
7.2 Normal form
Recall that we have fixed an arbitrary point and denoted its coordinates, in the -frame, by . Let us also recall that the magnetic field can be expressed by (2.5).
Performing an appropriate gauge transformation on the set introduced in (7.5), will yield a convenient normal form of the magnetic potential introduced in (7.4).
Lemma 7.2.
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 and , where is the geodesic curvature introduced in (2.11). Note that we do not try at the moment to explicitly compute and in the general case. We plan indeed to show that the result on the lowest eigenvalue is independent of and .
Proof of Lemma 7.2.
Our goal is to determine the Taylor expansion up to
order of the magnetic field vector and corresponding magnetic
field -form in the variables , the Taylor expansion being
computed at and . Up to a translation, we assume that .
Writing the magnetic vector field in (1.2) as
(7.13) |
the Taylor expansion of order at takes the form
(7.14) |
where and where we used (2.4)-(2.5).
Here we have used that by definition of the coordinate , the
function vanishes exactly at
order on . Note that is , introduced in (1.5).
We now express that on the norm of should be one. In fact
(7.15) |
where the coefficients can be computed by (2.8), (2.9) and (2.12).
For , this reads
(7.16) |
where is introduced in (2.12) and satisfies (2.13). We expand the last formula around . This leads, by taking and considering the coefficients of and , to the two identities
and
So it is natural to introduce the new parameters and as follows
(7.17) |
So we observe that
and
Hence our “normal” form becomes
with
(7.18) |
with
(7.19) |
Now consider . We have (see [HelMo4, Eq. (5.13)]), where is introduced in (2.9). So we obtain by (2.15),
with
(7.20) |
Notice that the condition reads (at and ) as follows
We have now to choose a suitable corresponding magnetic potential to . We find
(7.21) |
with
(7.22) |
(7.23) |
Moreover in the simply connected domain , so we can find a function such that .
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]):
(7.24) |
where and 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 such that, for all and s.t. , we have
where is introduced in (7.6) and
Proof.
7.4 An estimate away from the curve
Let us now look at the quadratic form, , when is supported away from . We start with a rough lower bound.
Lemma 7.4.
Given , and , there exist positive constants such that, if satisfies
where is introduced in (6.5), then
Proof.
The next proposition is an improvement of Proposition 7.4 since it allows for the support of to be closer to the curve .
Proposition 7.5.
Proof.
Step 1. Let us fix constants , , and . We assume that where with boundary coordinates satisfies (for small enough) and is introduced in (7.2).
We denote by the neighborhood associated with by (7.5). By a translation, we may assume that .
Consider the magnetic potential introduced in (7.4). We modify the coordinates so that, locally near , the metric in (2.12) is diagonal222We consider the curve defined by , where and is the coordinate transformation introduced in (2.2). We parameterization by arc-length and define the adapted coordinates by considering the normal geodesic to passing through . with
(7.27) |
By Taylor’s formula
In , we write
and
So we get, as in Lemma 7.1, the existence of such that
where
Performing a change of variables
which amounts to a rotation in the -plane (centered at ), we may assume that the second component of vanishes at , by choosing so that
At the same time, this rotation leaves and the measure invariant. Then performing a gauge transformation (see [HelMo4, Sec. 16.3]), we may assume that
where
Here
and are constants.
Similarly to the proof of Lemma 7.1, by writing
and
we get
Thus we are left with finding a lower bound of .
Note that,
since and by (7.27), the metric satisfies on , we have by (7.15), .
Moreover, since vanishes linearly on , there exist and such that
The previous estimates yield a lower bound of by comparing with a model operator (after rescaling the variables , and ). In fact, by [HelMo4, Lemma 16.1], there exists such that,
Note that, we can use Lemma 16.1 of [HelMo4] under our assumptions on the support of .
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 -dependent partition of unity on such that
If satisfies (7.26), then
where
where in the last step we used that and .
8 Lower bound
8.1 Another model
The model in (5.1) corresponds to the quadratic form in (7.24) when . However, when , 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 (see (8.22)).
8.1.1 A new model quadratic form
Let us fix a boundary point and denote the model quadratic form near by
(8.1) |
where is given in (7.24), and is the set introduced in (7.5). Furthermore, we assume that that the metric is flat at and the coordinates of in the frame are , after performing a translation with respect to the variable.
Following the proof of [HelMo4, Lem. 15.1], we are led to the analysis of the model quadratic form (see Lemma 8.1)
(8.2) |
where
(8.3) | ||||
and, with the angle defined by (2.6), we introduce the following functions
(8.4) |
We will consider the form on the following class of functions
(8.5) |
where
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 and , there exists such that, for any and ,
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
(8.6) |
where we used that on the support of , which follows by (8.4).
We also observe that :
(8.7) |
and
(8.8) |
Later on, we will choose and in a convenient way (see Remark 8.3).
8.1.2 Linearizing change of variable
In order to reduce to the case and eliminate the slightly variable coefficients of and 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 and (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 and appearing in (5.1).
Let us follow, what this change of variable was doing. We introduce
(8.9) |
Let us make the change of variables with
(8.10) |
where is the angle defined by (2.6).
The map 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 onto another neighborhood of .
Then, for small enough, is transformed by to the set satisfying :
(8.11) |
Let us write
(8.12) |
We can express the functions in terms of the variables, by using (8.10). In fact, we introduce , and observe that
Then we return back to the variables, by using (8.10). Noticing that, as ,
(8.13) |
we get
(8.14) | ||||
Let us now control the measure in the change of variable. By an easy computation, we get :
By using (8.13), satisfies
(8.15) |
where is defined in (8.10).
Similarly to Lemma 8.1 we get also that one can go from the control of to the control of the new quadratic form333We express and (see (8.3)) in terms of the variables introduced in (8.10) and neglect the terms of order .
(8.16) |
with
and
(8.17) | ||||
where .
Lemma 8.2.
For any , there exists such that, for any ,
where is associated with by the transformation .
By a unitary transformation, and after control of a commutator, we can reduce to a flat measure ( instead of ) and obtain the new quadratic form defined as follows
(8.18) |
with associated to by . In fact, we have [HelMo4, Eq. (15.29)]
(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
and on . The proof
is moreover uniform with respect to these parameters.
As a consequence, if was the transformation introduced in (8.9), the inverse (for )
, more explicitly the transformation
will bring
us (in the new variables ) to the initial model with
replaced by , and replaced by .
This can also be done by explicit computations.
Doing the transformations backwards, we are led to a magnetic Laplacian computed with a trivial metric but with a new magnetic potential
(8.20) |
and
(8.21) |
So the new model is not as simple as in the uniform magnetic field case (where ) but it is the model in (5.1), which we have studied in the previous section with
(8.22) |
In fact, since is supported in , we have,
(8.23) |
where is the operator in (5.1).
Remark 8.3.
We will choose in such a manner that . This choice is possible when satisfies .
8.1.3 Conclusion
We can now write a lower bound for the quadratic form in (7.24), assuming that and is the set introduced in (7.5). Let and . Collecting Lemmas 8.1, 8.2, (8.19), (8.23) and Proposition 5.5, we get the existence of positive constants and , such that
(8.24) |
where is introduced in Proposition 5.2 with the angle in (2.6).
8.2 The general case
We return now to the proof of the asymptotics of the lowest eigenvalue, , of the operator in (1.3). Under Assumptions (C1)-(C2), we will prove the following lower bound:
(8.25) |
for some constant , where is introduced in (1.6).
Let be a normalized ground state of , i.e.
Consider and the following neighborhood of the curve ,
(8.26) |
In terms of the coordinates introduced in Sec. 2.1,
Let be a smooth function such that
and
We introduce the function
(8.27) |
By Proposition 6.3, the eigenfunction is exponentially small outside , since by our choice of we have and . So we have
(8.28) |
Consider now a partition of unity of
and introduce the following functions
(8.29) |
We can decompose the quadratic form as follows
(8.30) |
where
(8.31) |
Let be a fixed constant that we will choose later to be sufficiently large. We will estimate the energy when the support of is near the curve , or away from , independently. So we introduce the sets of indices
(8.32) |
By Proposition 7.5,
(8.33) |
where is introduced in (7.6). Notice that
Since and , Proposition 6.3 together with (6.6) and (6.7) yield
Consequently, we infer from (8.33),
(8.34) |
For , we estimate by collecting (8.24) and the estimates in Lemma 7.1 and 7.3. We start by picking and , so that
where is introduced in (7.2). Eventually, we find
for some constant , where
and denote the coordinates of in the -frame (see Sec. 2 and Eq. 2.3). Note that we used Proposition 6.3 to control the term appearing in (8.24); in fact .
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 matching with the asymptotics in Theorem 1.4. More precisely, under Assumptions (C1)-(C2), we will prove that:
(9.1) |
for some constant , where 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
9.1.2 Structure of the quasi-mode
Consider two positive constants and such that . Let be a smooth even function, valued in , equal to on and supported in . We set
(9.4) |
where , so that is normalized as follows,
Our quasi-mode, , is supported in the set introduced in (7.2) and is of the form
(9.5) |
where is the function from Lemma 7.2 and the function is of the form
(9.6) |
where is given by (4.2), and are introduced in (9.3).
The choice of and will be specified later444 is defined in (9.2). For the definition of , see (9.11), (9.12) and (9.13). so that, for some constants , we have [HelMo4, Eq. (12.8)]
(9.7) |
Here arises while computing the quadratic form of the quasi-mode in (9.5). It is defined as follows [HelMo4, Eq. (12.9)],
(9.8) |
where
(9.9) | ||||
Notice that, by our normalization of , we have
(9.10) |
9.1.3 Definition of the auxiliary objects
Let us recall the definition of the function and the parameter given in [HelMo4, Sec. 12]. The function depends on and is selected in the following form (see [HelMo4, Eq. (12.14)])
(9.11) |
where
The function is selected as in [HelMo4, Eq. (12.28)]:
(9.12) |
In the sequel, we skip the hats from the notation. The function is defined as follows555 For the convenience of the reader, we will recall the heuristics behind the construction of in Subsection 9.1.4. [HelMo4, Eq. (12.22)]
(9.13) |
where is the positive normalized ground state of the harmonic oscillator in (4.1),
and is the regularized resolvent introduced in (4.4). Notice that and are Schwartz functions (i.e. in , see [FH06, Appendix A]). The definition of involves the differential operator
(9.14) |
and a function defined via the ground state of the Montgomery model in (4.5) and the following phase function
where
and the constant introduced in (4.2). We define now the function as follows
where
(9.15) |
and we choose (see (4.5))
(9.16) |
We conclude by mentioning some estimates which follow easily from the definitions of and in (9.11) and (9.12):
(9.17) | ||||
9.1.4 Heuristics on the construction of .
Starting from the definition of the function in (9.12), the quadratic form in (9.7) becomes (after neglecting error terms in the magnetic potential)
where
is introduced in (9.14) and
The construction of is based on minimizing
which amounts to finding the lowest eigenvalue of the operator
Writing
it is natural to search for in the form in (9.13) and satisfying
in the following sense (after taking the coefficients of to be , for )
Eventually, this leads to , , and as in (9.13).
9.2 Energy estimates
Actually, is bounded from above by modulo error terms, where and 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 appearing in (9.20), which would equals 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 in (9.6), we have
where
and
(9.18) | ||||
Since the function is even, we have
and
Moreover, we have the estimates
and
Notice that we used (9.17) and also that in the support of . Consequently, we get
(9.19) | ||||
Let us now reduce the computations to the potentials and in (9.9) which amount to and with . A straightforward computation yields,
(9.20) |
where
and by (9.17)
So, we end up with estimating
Notice that
9.3 Conclusion
We insert this into Lemma 7.3 with given in (9.5). Notice that satisfies (7.7) with . So by Lemma 7.3 and (9.10), we get for some
Comparing (9.10) and (9.5), we get by (2.20),
(9.22) |
Applying the min-max principle, and noticing that for , 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, n∘2, 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 . 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 (General case). mparc 02-145 (2002).
- [HelMo4] B. Helffer and A. Morame. Magnetic bottles for the Neumann problem: Curvature effect in the case of dimension (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 -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 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).