Lower bounds for the first eigenvalue of the Laplacian with zero magnetic field in planar domains
Abstract
We study the Laplacian with zero magnetic field acting on complex functions of a planar domain , with magnetic Neumann boundary conditions. If is simply connected then the spectrum reduces to the spectrum of the usual Neumann Laplacian; therefore we focus on multiply connected domains bounded by convex curves and prove lower bounds for its ground state depending on the geometry and the topology of . Besides the area, the perimeter and the diameter, the geometric invariants which play a crucial role in the estimates are the the fluxes of the potential one-form around the inner holes and the distance between the boundary components of the domain; more precisely, the ratio between its minimal and maximal width. Then, we give a lower bound for doubly connected domains which is sharp in terms of this ratio, and a general lower bound for domains with an arbitrary number of holes. When the inner holes shrink to points, we obtain as a corollary a lower bound for the first eigenvalue of the so-called Aharonov-Bohm operators with an arbitrary number of poles.
Classification AMS : 58J50, 35P15
Keywords: Magnetic Laplacian, spectrum, lowest eigenvalue, planar domains
Acknowledgments: Research partially supported by INDAM and GNSAGA of Italy
1 Introduction
1.1 Definitions and state of the art
Let be a bounded, open, connected domain with smooth boundary in a Riemannian manifold and let be a smooth real one-form on , (the potential one-form). Define a connection on the space of complex-valued functions as follows:
for all vector fields on , where is the Levi-Civita connection of . The magnetic Laplacian with potential is the operator acting on :
In this gives explicitly, in the usual notation:
where is the dual vector field of , the vector potential. The two-form is the magnetic field; dually, in dimension , is the vector field .
Scope of this paper is to discuss the spectrum of for planar domains. Hence in what follows we take .
The spectrum of the magnetic Laplacian has been studied extensively for Dirichlet boundary conditions ( on ), and we denote by the first eigenvalue. First we remark that, thanks to the diamagnetic inequality, one always has:
and in particular . For planar domains and constant magnetic field (that is, and constant), a Faber-Krahn inequality holds, in the sense that the first eigenvalue of a planar domain is minimized by that of the disk of the same area (see [5]). Estimates for sums of eigenvalues can be found in [9].
However in this paper we deal with magnetic Neumann boundary conditions, that is we impose on the boundary, where is the inner unit normal to . It is known that then admits a discrete spectrum
diverging to . The first eigenvalue has the following variational characterization:
(1) |
For computing lower bounds the diamagnetic inequality is of no use; in fact it gives:
because is simply the first eigenvalue of the usual Neumann Laplacian, which is zero (the associated eigenspace being spanned by the constant functions). There are fewer estimates in this regard; let us first discuss the case of a constant magnetic field on planar domains. The paper [4] gives a lower bound of in terms of the inradius of , and of course . Asymptotic expansions as are obtained in [7]. We also mention the paper [6] which investigates the validity of a reverse Faber-Krahn inequality for constant magnetic field , that is: is it true that is always bounded above by that of a disk with equal volume ? It is proved there that this inequality is true when is either sufficiently small or sufficiently large, but the general case is still open in the simply connected case.
In this paper we prove three lower bounds for the first eigenvalue of planar domains under Neumann conditions, when the magnetic field is identically zero. Since this will be the only boundary condition we consider, from now on we will simply write instead of .
Let us first clarify the circumstances under which the first eigenvalue might be positive even if the magnetic potential is a closed one-form on . This is intimately related to a phenomenon in quantum mechanics predicted in 1959 and known as Aharonov-Bohm effect, which has also experimental evidence: a particle travelling a region in the plane might be affected by the magnetic field even if this is identically zero on its path. In fact what the particle ”feels” is not the magnetic field but, rather, the magnetic potential , provided that is closed but not exact, and that the flux of around the pole may assume non-integer values (see below for the precise condition).
Let us be more precise. From the definition we see that, if , the spectrum of coincides with the spectrum of the usual Laplacian under Neumann boundary conditions. The same is true when is an exact one-form, by the well-known gauge invariance of the magnetic Laplacian. This fundamental property states that the spectrum of is the same as the spectrum of , for any , which follows from the identity:
showing that and are unitarily equivalent.
On the other hand, if the magnetic field is non-zero, then is strictly positive. One could then ask if has to vanish whenever the magnetic field is zero, that is, whenever is a closed one-form.
To that end, let be a closed curve in (a loop). The quantity:
is called the flux of across (we assume that is travelled once, and we will not specify the orientation of the loop; this will not affect any of the statements, definitions or results which we will prove in this paper).
It turns out that
if and only if is closed and the cohomology class of is an integer, that is, the flux of around any loop is an integer.
This was first observed by Shigekawa [12] for closed manifolds, and then proved in [8] for manifolds with boundary. This remarkable feature of the magnetic Laplacian shows its deep relation with the topology of the underlying manifold . In this paper we will focus precisely on the situation where the potential one form is closed, and we will then give two lower bounds for the first eigenvalue .
Let us then recall a few previous results when the magnetic field is assumed to vanish. A lower bound for a general Riemannian cylinder (i.e. the surface endowed with a Riemannian metric) and zero magnetic field has been given in [3], and is somewhat the inspiration of this work: one of two main results here is in fact to improve such bound when is a doubly connected planar domain.
Directly related to the Aharonov-Bohm effect, we mention the papers [1] and [10] which investigate the behavior of the spectrum of a domain with a pole when the pole approaches the boundary, for Dirichlet boundary conditions. We remark here that the pole is a distinguished point and the potential is the harmonic one-form:
which has flux across any closed curve enclosing , giving rise to a magnetic field which is a Dirac distribution concentrated at the pole (therefore, the magnetic field indeed vanishes on ). The magnetic Laplacian acting on is often called an Aharonov-Bohm operator. One could think to a domain with a pole as a doubly connected domain for which the inner boundary curve shrinks to a point.
We will in fact give a lower bound for the first eigenvalue of Aharonov-Bohm operators with many poles, and Neumann boundary conditions (see Theorem 3).
The Aharonov-Bohm operators play an interesting role in the study of minimal partitions, see chapter 8 of [2].
For Neumann boundary conditions, we mention the paper [8], where the authors study the multiplicity and the nodal sets corresponding to the ground state for non-simply connected planar domains with harmonic potential. For doubly connected domains, it is shown that is maximal precisely when is congruent to modulo integers (this fact is no longer true when there are more than two holes). The proof relies on a delicate argument involving the nodal line of a first eigenfunction and the conclusion does not follow from a specific comparison argument, or from an explicit lower bound.
The focus of this paper is on lower bounds for multiply connected planar domains and zero magnetic field defined by the closed potential form . By what we have just said, it is clear that estimating the first eigenvalue is a trivial problem if is simply connected, because then any closed one-form is automatically exact, and therefore by gauge invariance. Therefore, we restrict our study to domains with holes, with .
In this paper we will prove: an improved lower bound for doubly connected domains; a general lower bound for multiply connected domains with an arbitrary number of convex holes; a lower bound for a general convex domain with an arbitrary number of punctures. Let us describe these results in detail.
1.2 A lower bound for doubly connected domains
Let us start from doubly connected domains () hence domains of type:
where and are open and smooth. We assume and convex. Let be the flux of the closed potential around the inner boundary curve : by Shigekawa result, the lower bound is simply zero when is an integer. Then, to hope for a positive lower bound, we need to measure how much is far from being an integer, and the natural invariant will then be:
The second important ingredient for our lower bounds is the ratio between the minimal width and the maximal width of . To be more precise, let us say that the line segment is an orthogonal ray if it hits the inner boundary orthogonally. By definition, the minimal width (resp. maximal width ) of is the minimal (resp. maximal ) length of an orthogonal ray contained in :

Note that the ratio is invariant under homotheties, and reaches its largest value whenever the boundary components are parallel curves.
In Theorem 2 of [3] we prove the lower bound:
(2) |
We insist on the fact that if is bounded below away from zero we get a positive, uniform lower bound even if tends to zero. Think for example to a concentric annulus of radii and ; then and as the lower bound will approach , a positive number, which is just the first eigenvalue of the unit circle.
This means that (for fixed perimeter) in order to get small, the ratio (and not just ) has to be small.
Sharpness in terms of . In [3] we showed that if is small then the first eigenvalue could indeed be small. We then looked for an example which could show that the dependance on is sharp, and we could not find it. Rather, in Examples 14 and 15 in [3], we constructed examples of domains such that is bounded below, say by , is bounded above, goes to zero and goes to zero proportionally to , for any non-integral flux. Therefore, if one could replace by the linear factor in (2), one would obtain a sharp inequality (with respect to ). See Figure 2 below for the example which shows sharpness.
This is in fact possible, and the theorem which follows should be regarded as the first main theorem of this paper.
Theorem 1.
Let be an annulus in the plane, with and convex with piecewise-smooth boundary. Let be a closed one-form with flux around the inner hole . Then:
a) One has the lower bound:
where and are, respectively, the minimum and maximum width of , and is the diameter of .
b) If the outer boundary is smooth, and if is less than the injectivity radius of the normal exponential map of , then we have the simpler lower bound:
(3) |
Note that, modulo a factor of , b) is formally identical to (2) with replacing .
We observe that there is no positive constant such that
for all doubly convex annuli in the plane (otherwise, the lower bound would be independent on the inner hole, and this is impossible). This means that Theorem 1 is not a trivial consequence of (2).
In fact, the proof of Theorem 1 uses a suitable partition of into overlapping annuli for which is, so to speak, as small as possible (see Section 2 below, and in particular Figure 3 for an example). Recall the -interior ball condition:
given , there is a ball of radius tangent to at and entirely contained in .
Here and for further applications, we say that the injectivity radius of is if satisfies the -interior ball condition for any . If is smooth, its injectivity radius is positive.
Finally we picture below the family of domains which realize sharpness. is the difference between two rectangles with parallel sides, with boundaries being units apart. Hence and is uniformly bounded above by .

1.3 A general lower bound for multiply connected domains
Now let be an -holed planar domain, which we write as follows:
(4) |
where the inner holes are smooth, open and disjoint. We furthermore assume that are convex. Note that:
We will call the inner boundary of . The minimal and maximal widths of are defined as in the case , namely is the minimal length of a line segment contained in and hitting the inner boundary orthogonally, and the maximal length of such line segments is by definition the maximal width .
It is clear that we could replace by the diameter of , and by the invariant:
In this section we give a lower bound of when has an arbitrary number of convex holes.
Here is the estimate.
Theorem 2.
Let be an n-holed planar domain, where are smooth, open and convex. Let be a closed potential having flux around the -th inner boundary curve , for , and let Then we have:
(5) |
where and are, respectively, the minimal and maximal width of .
1.4 A lower bound for Aharonov-Bohm operators with many poles
The power in the previous estimate is probably not sharp; it appears to be there for technical reasons. By shrinking the inner boundary curves to points we obtain an estimate in terms of , which has an interesting interpretation in terms of Aharonov-Bohm operators with many poles.
Precisely, we fix a convex domain and choose points inside it, say . Consider the punctured domain . Given a closed one-form , we define:
where is the -neighborhood of (it obviously consists of a finite set of disks of radius ). It is not our scope in this paper to investigate the convergence in terms of ; however, what we are looking at could be interpreted as the first eigenvalue of a Aharonov-Bohm operator with poles and Neumann boundary conditions. The proof of the theorem in the previous section simplifies, to give a general lower bound in terms of the distance between the poles, and the distance of each pole to the boundary. To that end, define:
Of course could be conveniently bounded above by the diameter of . Let be as usual a closed one-form having flux around the pole . Then we have:
Theorem 3.
Let be a convex domain and a finite set of poles. For the punctured domain we have the bound:
where , and is the flux of the closed potential around .
In a forthcoming paper, we will give upper bounds for the Laplacian with zero magnetic field on multiply connected planar domains, which are closely related to the topology (number of holes) of the domain.
2 Proof of Theorem 1
The proof depends on a suitable way to partition our domain . We first remark the simple fact that the first eigenvalue of a domain is controlled from below by the smallest first eigenvalue of the subdomains of a partition of (Proposition 4). Then, we need to extend inequality (2) to piecewise-smooth boundaries, see Section 2.2. In Section 2.3 we state our main geometric facts, Lemma 6 and Lemma 7, and then we prove Theorem 1 (see Section 2.4). Finally, in Section 2.5, we define the partition and we prove Lemma 6 and Lemma 7.
2.1 A simple lemma
We say that the family of open subdomains is a partition of , if . Thus, the members of the partition might overlap and some of the intersections could have positive measure. If furthemore is empty for all then we say that the partition is disjoint. We observe the following standard fact whose proof is easy:
Proposition 4.
Let be a partition of the domain . Let be any closed potential. Then, there is an index such that
(6) |
If the partition is disjoint, then:
(7) |
Proof.
We start proving (6). Let be an eigenfunction associated to . We use it as test-function for and obtain, for all :
(8) |
Now
where the index is chosen so that is maximum among all . Then:
That is: which is the assertion.
For the proof of (7), let . From (8) we have, for all :
We now sum over and obtain As is a first eigenfunction the left-hand side is precisely , and the inequality follows.
∎
2.2 Convex annuli with piecewise-smooth boundary
From now on will be an annulus in the plane with boundary components which we assume convex and piecewise-smooth. We will write where and are open, convex, with piecewise smooth boundary. In that case and .
Let be a point of where is not smooth ( will then be called a vertex). The normal cone of at is the set
Then is the closed exterior wedge bounded by the normal lines to the two smooth curves concurring at , its boundary is the broken line depicted in the figure below. Call its angle at .
We remark the obvious fact that .
We now define the minimum and maximum width in the piecewise-smooth case. These are defined in (9) and depicted in the Figure 3 below.

For a unit vector applied in and pointing inside we let denote the ray exiting in the direction , and let be the intersection of with . We define:
(9) |
We notice that at a smooth point the cone at degenerates to the normal segment at . Hence at a smooth point one has
We now define
(10) |
and will be called the minimum width and, respectively, the maximum width of . We remark that when the two boundary components are smooth and parallel then and the ratio assumes its largest possible value, which is .
As a first step in the proof of Theorem 1, we extend the inequality (2) to the piecewise-smooth case.
Theorem 5.
Let be an annulus in the plane whose boundary components are convex and piecewise smooth. Let and be the invariants defined in (10). Then for any closed potential having flux around the inner boundary curve one has the lower bound:
where is the length of the outer boundary.
Proof.
First, admits an exhaustion by convex annuli with -boundary, say . By that we mean:
a) where and are convex and have -smooth boundary;
b) and so that ;
c) and in particular .
To construct we round off corners at distance to each of the vertices of ; to construct we just take the convex domain bounded by the -neighborhood of .
Let be an eigenfunction associated to ; by restriction we obtain a test-function for , hence by the min-max principle:
Let be the functional:
We can apply (2) and obtain because has smooth boundary; then, for all :
We now pass to the limit as on both sides; as (as we can easily see from the definitions in (10)), we obtain the assertion:
∎
2.3 Preparatory results
In this section we state the two main technical lemmas; the partition of the annulus will be defined in Section 2.5.
So let be an annulus as above and let . We consider the distance functions:
where and . Fix a parameter . As is convex, with piecewise-smooth boundary, it is well-known that the equidistants are -smooth curves. We say that the parameter is regular if the equidistant is a piecewise-smooth curve. Following Appendix 2 in [11], we know that the set of regular parameters has full measure in ; as a consequence, there exists a sequence of regular parameters as .
By using an obvious limiting procedure, from now on we take
and can assume that it is a regular parameter, so that is a piecewise-smooth curve.
Lemma 6.
Let be an annulus in the plane with and convex with piecewise-smooth boundary, and let (which, by assumption, is a regular parameter). Then admits a partition into (overlapping) subdomains with the following properties.
a) is an annulus bounded by two convex piecewise-smooth curves, that is, with and convex, and contains (see figure in Section 2.5 below).
b) The number of annuli in the partition can be taken so that:
We estimate the ratio of each piece as follows.
Lemma 7.
Let be the partition in the previous lemma. For all one has the following facts.
a) and .
b) The following estimate holds:
(11) |
where is the diameter of .
c) If is less than the injectivity radius of , then the following simpler lower bound holds for all :
The proof of Lemma 6 and Lemma 7 involves rather simple geometric constructions, but there are some delicate points to take care of, and will be done in Section 2.5. In fact, these two lemmas make it possible to write as a union of subset such that the ratio is bounded below, which make the proof of Theorem 1 quite easy, as follows.
2.4 Proof of Theorem 1
We use the partition of Lemma 6. Let be a closed potential having flux around the inner boundary curve ; then, has the same flux around the inner component of , by Lemma 6a. By Proposition 4a there exists such that
(12) |
By Theorem 5 applied to we see:
By b) of Lemma 7 we see:
(13) |
This, together with the inequality , gives:
We insert this inequality in (12) and use the inequality (see Lemma 6b) to conclude that
This proves part a) of Theorem 1.
If is less than the injectivity radius of we proceed as before, using the lower bound proved in Lemma 7c. We arrive easily at the inequality:
which is Theorem 1b).
2.5 The partition of and the proof of Lemma 6
We start by showing the partition on a particular example, see Figure 4 below. The initial domain is a triangle minus a disk and . We draw the first three pieces and then the last one, which is and which coincides with the -neighborhood of the exterior boundary (this is always the case). Note that the pieces overlap, hence the partition is not disjoint.

.
We now proceed to construct the partition in general. Let then be convex annulus as above, and consider the distance functions:
where and .
At step 1, we let , and . That is, is simply the subset of at distance less than to .
At step 2, we let and and define .
At the arbitrary step , we let and , and define:
Observe that for any choice of positive numbers the sets and are convex. Therefore, both and are convex; moreover and then is an annulus. This proves part a) of Lemma 6.
For b) we first prove the following fact:
Fact. Let (the smallest integer greater than or equal to ). Then and . In particular, and then, starting from , the sequence stabilizes: .
For the proof we first observe that, from the definition of , we have ; then, if we fix we have by definition hence . To show that it is enough to show:
(14) |
In fact, if not, there would be a point such that and . Let be a point at minimum distance to , and prolong the segment from to till it hits at the point . It is clear that then . By definition of we then have:
which contradicts the definition of . Hence (14) holds.
We now prove part b) of Lemma 6. Observe that and (by the definition of ) . Since for all we see that, for all :
which gives the assertion.
2.6 Proof of Lemma 7
We now study the typical piece in the partition. Observe that
where
(some of these boundary pieces may be empty). As the equidistants are smooth, and the equidistant is piecewise smooth, we see that and are both piecewise smooth, hence
is piecewise smooth.
The inner boundary is , it is piecewise smooth with vertices in the set
Now we have to estimate the ratio for a fixed . Recall the function
defined in (9). First notice that the regular parts
are parallel, at distance to each other. Hence
(15) |
at all points . Similarly, the regular sets
are parallel at distance and on .
Therefore, it only remains to control the ratio at the vertices of , which are finite, say .
Each break point gives rise to a corresponding wedge . In Figure 5 we enlarge the domain relative to the partition of Figure 4 and we show its set of wedges. In other words, every annulus is made up of strips of constant width and wedges, and we need to control only at the wedges.

Typically, the ratio is small when there are small angles; nevertheless, this ratio is controlled from below by the diameter and the volume of the outer domain, as we will see in the next section.
As is a convex subset of , we see that . Now it is clear from the construction that for all ; moreover, the inequality is attained. Therefore
for all . This proves part a) of Lemma 7.
2.7 End of proof of Lemma 7
The estimate on the wedges of the generic piece will be a consequence of Lemma 8 below.
We recall that the cut-locus of is the closure of the set of all points which can be joined to by at least two minimizing segments; moreover, the injectivity radius of is the minimum distance of to the cut-locus. If is smooth, its injectivity radius is positive. Finally the distance function is smooth outside the cut-locus.
Then, we fix a piece and recall that . For simplicity we write and recall that, by definition, we have .
Let be the vertices of . For in this set, write , where are the arcs concurring at . Note that either is an equidistant to , that is, is a subset of , and in that case we say that is parallel to , or is an equidistant to , that is, is a subset of , and in that case we say that is parallel to . There are two possibilities:
Type 1. The vertex where is parallel to and is parallel to ;
Type 2. and are both parallel to .
Note that the second type corresponds to the situation where the vertex is a point of the cut-locus of . The situation where and are both parallel to does not occur, because then would belong to the cut locus of ; however the cut-locus of a convex domain is always contained in the interior of the domain; as is outside this is impossible.
For the partition in the example and its piece (see Figure 5), the vertices and are of type 1, while the vertex is of type 2.
Lemma 8.
a) If is of type 1 then the interior angle of at is larger than or equal to , hence the angle of the wedge at is at most . Consequently,
b) If is of type 2, then is in the cut-locus of and one has:
c) If is less than the injectivity radius of then the estimate in a) will hold at all vertices of the decomposition.
Since the lower bound in b) is always weaker than that in a), we have b) at all vertices of . It is clear that Lemma 8 completes the proof of Lemma 7.
Proof of Lemma 8a) If is of type 1 then is not on the cut-locus of , hence exists and is a well-defined unit vector in a neighborhood of . Note that points in the direction where the distance to increases (obviously an analogous observation holds for ). Now observe that the angle of the wedge is the angle between the vectors and (see Figure 6).

Hence, it is enough to show that the quantity
is non-positive. Assume on the contrary that . We let denote the segment which minimizes distance from to (parametrized by arc-length); hence . We let be the function which measures distance from to , so that:
Now . In particular,
As , this means that for small one has , but this impossible because , and all points of are, by definition, at distance at least to .
Hence and the angle of the wedge at is at most . Now, the wedge is contained in the polygon with vertices as in the picture, hence because the angle at is at most , the angles at and are , and .
Proof of b). Let be a vertex of type 2: then, the two arcs concurring at are parallel to , and belongs to the cut locus of . The boundary of the wedge is made of two distinct segment of the same length minimizing distance to . Then, is contained in a wedge of the last member of the partition, that is, . As depends only on , we could as well estimate the ratio by estimating the corresponding ratio for the wedges of , which will allow to express in terms of the geometry of , hence, in terms of the geometry of .
The relevant picture is shown below (see Figure 7), in which we evidence such an edge (dark shadowed in the picture): it has its vertex in the point of ; we let be a point such that . We omit to draw the inner boundary as it will play no role in the proof.
Let be the triangle with dotted boundary, with a vertex in and such that . As is the exterior angle at a vertex of the piecewise-smooth curve , we see that .


Each of the angles and is less than . Consider the circle with center and radius . If is inside this circle then which is impossible. Then is outside the circle; and are, each, less then the corresponding angles at the vertex obtained by projecting on the circle. It is clear that each of these two angles is less than .
Let be the angle at the vertex . Then
(16) |
and similarly If or then and we are finished because .
Hence we can assume from now on .
Lemma 9.
In the above notation we have
Proof.
We first remark that . In fact, we have and so that On the other hand, the same argument applies to , that is, . Therefore, as we conclude The same bounds are satisfied by .
As is contained in the union of two parallelograms with sides and we see:
(17) |
We set ; we also let be the convex polygon with vertices and (see Figure 8).

We have disjoint unions:
We will show that
and the assertion will follow by summing up the two inequalities. Now the first inequality is obvious, because . By the bounds for and we see that and are both, at least, . Then:
Combining the two estimates we see that as asserted. ∎
End of proof of Lemma 7. Refer to Figure 8. We can assume that . We let be the length of the segment joining and (which meets the side opposite to orthogonally, by definition), so that:
The assumptions give , so that Using the lower bound for proved before, we have and then, from (16):
the last inequality holding because evidently . This proves Lemma 8b and, with it, Lemma 7 is completely proved.
2.8 Example showing sharpness
This example is taken from [3], we repeat it below for the sake of clarity. Its scope is to show that the inequality of Theorem 1 is sharp in .
We take to be the rectangle , and consider the doubly connected domain:
We refer to the picture in the Introduction. We let be any closed -form. As a direct consequence of the gauge invariance of the magnetic Laplacian, it is proved in [3] that, for any planar domain one has:
(18) |
where is any closed, simply connected subdomain of , and where denotes the first eigenvalue of the usual Laplacian with Neumann boundary conditions on and with Dirichlet boundary conditions on .
Given our choice of , we remove from it the rectangle to get the simply connected subdomain called . We estimate by taking the test-function as follows:
It is readily seen that , while (note that ). Therefore:
Given (18) we conclude that:
(19) |
3 Proof of Theorem 2
Let be an -holed domain, which we write: with smooth, open and convex.
From now on we denote . The idea is to use a suitable partition of by annuli whose boundary is either a piece of or is an equidistant curve from two interior boundary curves; each is an annulus with piecewise-smooth exterior boundary which is star-shaped with respect to . We can then apply a theorem in [3] and obtain the uniform bound, valid for all :
(21) |
As the bound holds for all subdomains of a disjoint partition it holds a fortiori for , thanks to Proposition 4.
3.1 The partition of
We start by giving in Figure 9 below the picture of the partition when has three holes. The inner boundary of each piece is made of equidistant sets from two suitable holes.

Here is the construction. For each we consider the non-empty open set:
If we set:
(22) |
we see that we can write
It is clear that . We remark that is the equidistant set from and :
(23) |
We have the following general fact.
Lemma 10.
Let and be disjoint smooth convex domains. Then the equidistant set as above is a smooth curve.
Proof.
Let and let be the distance function to , . The convexity of implies that is smooth on the complement of , so that is smooth on . Let , so that is the zero set of . One has and it is enough to show that has no critical points on . Let be a point in and let be the line segment which minimizes the distance from to . One has: . The corresponding minimizing segment from to is then . If then and, as , the two minimizing segments would have the same foot , which would then belong to both and : but this impossible because and are disjoint.
Hence, on one has which proves smoothness. ∎
As are disjoint we see that and then we can introduce the annulus
that is:
The family gives rise to a disjoint partition of , as the next lemma shows.
Lemma 11.
The following properties hold:
a) .
b) For one has that and is a smooth curve (eventually empty).
c) is an annulus with smooth inner boundary and piecewise smooth outer boundary . Moreover:
where is contained in the equidistant curve from and .
Note that actually . The proof of the lemma is clear from the definitions.
3.2 Estimate of
As the partition of Lemma 11 is disjoint, from Proposition 4 we have:
Therefore, in this section, we estimate the first eigenvalue of the generic member of the partition. To that end, recall a relevant theorem from [3]. Let be an annulus with inner boundary curve , where is smooth and convex. For and , consider the segment where is the unit normal to oriented outside . Let be the first intersection of with the outer boundary curve , and let be the angle between and the outer normal to at . We set:
We recall that is said to be starlike with respect to if, for any , the segment minimizing distance from to is entirely contained in .
We also set:
which are called, respectively, the minimum and maximum width of .
The estimate in Theorem 2 of [3] says that:
(24) |
We will apply (24) to each annulus in the above partition of . We start from:
Lemma 12.
Let and let be a piece in the partition defined above. Then:
a) is an annulus which is starlike with respect to , and moreover:
b) One has the estimate:
Lemma 12 allows to prove Theorem 2 as follows. We apply (24) to and get:
To make the lower bound independent on , it is enough to observe that and . Then we get:
which is the final step of the proof.
Then, it remains to prove Lemma 12.
3.3 Proof of Lemma 12a
It is enough to prove it for . We first prove that is star shaped with respect to . Let and let be the segment starting at and minimizing distance to : let be the foot of .
Note that, as , we must have for all .
We prove that is entirely contained in . Assume by contradiction that there is such that . Then for some , and there exists with . But then:
that is, and this means that , which contradicts the assumption. Hence is star shaped.
We now estimate , and for convenience we refer to the picture below, Figure 10.

Let and draw the segment where is the unit normal vector to pointing outside . It hits at the point . If we proceed as in [3] (because is convex) and get
If (as in the picture) then for some ; we can assume that . Let be the point in such that ; observe that where is the normal to at pointing away from . Observe that is the unit vector in the direction of , and that is tangent to at . If is the angle between and then we see that , that is
Consider the triangle with vertices ; it is isosceles on the basis , (whose length is denoted ); its height is part of the tangent line to the equidistant at . One sees that
hence
Now by definition of ; as the segment joining and is entirely contained in we see that . Hence
as asserted.
3.4 Proof of Lemma 12b
Recall that the typical piece of the decomposition is . We need to estimate ; this is a bit more difficult now because is no longer convex (there are circumstances under which each is actually convex - for example, when all holes are disks of the same radius - and we will discuss this case in the next section, to obtain a simpler final estimate).
Set for concreteness. We apply Green formula to the function . Note that is the curvature at of the equidistant to through ; as is convex one has on the complement of . By Green formula:
where is the inner unit normal. We let denote the -neighborhood of , so that by the definition of . Since :
By co-area formula:
because for all (we are integrating the opposite of the curvature of a closed curve, and we always obtain ). Therefore:
(25) |
On the other hand . Hence:
The first piece is . On the outer boundary we see that:
where is as in the proof of part a), and the inequality then follows from part a). Then:
and given (25) we obtain , that is:
which gives the assertion.
4 Proof of Theorem 3
About the partition of the previous section for domains with holes, we remark that if the inner holes are disks of the same radius , then the equidistant set between any pair of them is simply a straight line, and therefore each is a line segment; moreover the subdomains are all convex: see Figure 11 which illustrates the partition when all holes shrink to a point.

We can directly apply (2) to each and obtain:
As is convex, we have and therefore we arrive at the following estimate.
Theorem 13.
Let with convex and being disjoint disks of center, respectively, and common radius . Then:
with .
References
- [1] Laura Abatangelo, Veronica Felli, Benedetta Noris, and Manon Nys. Sharp boundary behavior of eigenvalues for Aharonov-Bohm operators with varying poles. J. Funct. Anal., 273(7):2428–2487, 2017.
- [2] Virginie Bonnaillie-Noël and Bernard Helffer. Nodal and spectral minimal partitions—the state of the art in 2016. In Shape optimization and spectral theory, pages 353–397. De Gruyter Open, Warsaw, 2017.
- [3] Bruno Colbois and Alessandro Savo. Lower bounds for the first eigenvalue of the magnetic Laplacian. J. Funct. Anal., 274(10):2818–2845, 2018.
- [4] Tomas Ekholm, Hynek Kovarik, and Fabian Portmann. Estimates for the lowest eigenvalue of magnetic Laplacians. J. Math. Anal. Appl., 439(1):330–346, 2016.
- [5] L. Erdös. Rayleigh-type isoperimetric inequality with a homogeneous magnetic field. Calc. Var. Partial Differential Equations, 4:283–292, 1996.
- [6] S. Fournais and B. Helffer. Inequalities for the lowest magnetic Neumann eigenvalue. Lett. Math. Phys., 109(7):1683–1700, 2019.
- [7] Soeren Fournais and Bernard Helffer. Accurate eigenvalue asymptotics for the magnetic Neumann Laplacian. Ann. Inst. Fourier (Grenoble), 56(1):1–67, 2006.
- [8] B. Helffer, M. Hoffmann-Ostenhof, T. Hoffmann-Ostenhof, and M. P. Owen. Nodal sets for groundstates of Schrödinger operators with zero magnetic field in non-simply connected domains. Comm. Math. Phys., 202:629–649, 1999.
- [9] R. S. Laugesen and B. A. Siudeja. Magnetic spectral bounds on starlike plane domains. ESAIM Control Optim. Calc. Var., 21(3):670–689, 2015.
- [10] Benedetta Noris, Manon Nys, and Susanna Terracini. On the Aharonov-Bohm operators with varying poles: the boundary behavior of eigenvalues. Comm. Math. Phys., 339(3):1101–1146, 2015.
- [11] Alessandro Savo. Lower bounds for the nodal length of eigenfunctions of the Laplacian. Ann. Global Anal. Geom., 19(2):133–151, 2001.
- [12] Ichiro Shigekawa. Eigenvalue problems for the Schrödinger operator with the magnetic field on a compact Riemannian manifold. J. Funct. Anal., 75(1):92–127, 1987.
Bruno Colbois
Université de Neuchâtel, Institut de Mathématiques
Rue Emile Argand 11
CH-2000, Neuchâtel, Suisse
Alessandro Savo
Dipartimento SBAI, Sezione di Matematica,
Sapienza Università di Roma
Via Antonio Scarpa 16
00161 Roma, Italy