1. Introduction
The purpose of the present paper is to derive Hadamard’s variational formulae for simple eigenvalue of the Laplacian and is to derive several new inequalities concerning the first eigenvalue concerning the mixed boundary condition.
In the previous work [
9] we studied Hadamard’s variational formula for general domain deformation and extended the results [
3] on two or three-space dimensions under normal perturbations of the domain. There, we develop an abstract theory of perturbation of self-adjoint operators, refining the argument in [
6].
To be precise, let
and
be a pair of Hilbert spaces over
with compact embedding
. Henceforth,
denotes a generic constant. Let
and
for
,
, be symmetric bilinear forms satisfying
and
for some
. We take the abstract eigenvalue problem
which ensures a sequence of eigenvalues denoted by
The associated normalized eigenfunctions,
furthermore, form a complete ortho-normal system in
X, provided with the inner product induced by
. Hence it holds that
for any
and
.
The abstract theory developed in [
9] is stated as follows. First, if
holds for fixed
, we obtain
at this
t, where
are arbitrary (Theorem 8 of [
9]). Second, if there exist bilinear forms
and
for any
such that
and if it holds that
for fixed
, we obtain the existence of the unilateral derivatives
at this
t, where
are arbitrary (Theorem 12 of [
9]). If inequalities (
5) are valid to any
and it also holds that
for the specified
t, furthermore, the above unilateral derivatives satisfy
We assume, furthermore,
for some
, under the agreement of
. Assume, also that (
5) and (
7) hold for any
. Then, there exists a family of
curves denoted by
,
, made by a rearrangement of
at most countably many times in
I, where
(Theorem 3, Theorem 14 of [
9]).
Third, if we have the other bilinear forms
and
satisfying
for any
, and
for the fixed
t, then there are
for this
t, where
are arbitrary (Remark 12 of [
9]). These
, furthermore, satisfy
if inequalities (
10) are valid to any
and it holds that
for the above specified
t (Theorem 23 of [
9]). Furthermore, if (
10) and (
11) hold for any
and if it holds that (
8), then the above described family of curves
,
, are
(Theorem 24 of [
9]).
Finally, these derivatives
and
for
are characterized as the eigenvalues of associated eigenvalue problems on the
m-dimensional space,
(Theorem 12 and Theorem 15 of [
9]). Henceforth, we assume (
1), (
2), (
4), (
5), (
7), (
9), (
10), and (
11) for any
.
If
holds in (
8), for example, we obtain
and
where
,
, and
. Here,
is defined for
by
where
,
, and
is the orthogonal projection, and
Thus we obtain the following theorem because the bilinear form is non-negative definite on if .
Theorem 1.
If holds in (8), there arises that
for .
Theorem 1 implies the following result.
Harmonic convexity of the first eigenvalue, inequality (
16), was noticed by [
3] for the Dirichlet problem of Laplacian under the conformal deformations of the domain in two space dimension. Here we calculate the values
,
,
,
and the validity of (
15) under general setting of the deformations of the domain, and then turn to the dynamical and the conformal deformations. Taking preliminaries in
§2, thus, we show the results on dynamical and conformal deformations in
§3 (Theorem 5 and Theorem 6) and
§4 (Theorem 8 and Theorem 9), respectively. As applications, we show several new inequalities on the first eigenvalue of the two-dimensional problem.
2. General Deformations around
Let
be a bounded Lipschitz domain in
n-dimensional Euclidean space
for
. Suppose that its boundary
is divided into two relatively open disconnected sets
and
, satisfying
We study the eigenvalue problem of the Laplacian with mixed boundary condition,
where
and
denotes the outer unit normal vector on
. This problem takes the weak form, finding
u satisfying
defined for
and
where
This
V is a closed subspace of
under the norm
The above reduction of (
18) to (
19)-(
22) is justified via the trace operator to the boundary since
is a bounded Lipschitz domain (Theorem 2 of [
8]).
To confirm the well-posedness of (
3), we note, first, that if
, there is a coercivity of
, which means the existence of
such that
If
we replace
A by
, denoted by
. Then this
is coercive, and the eigenvalue problem
is equivalent to (
3) by
. Hence we can assume (
1), using this reduction if it is necessary.
If the bounded domain
is provided with the cone property, the inclusion
is compact by Rellich-Kondrachov’s theorem [
1]. Thus there is a sequence of eigenvalues to (
3), denoted by
The associated eigenfunctions,
, furthermore, form a complete ortho-normal system in
X, provided with the inner product induced by
:
The
j-th eigenvalue of (
3) is given by the mini-max principle
where
is the Rayleigh quotient, and
and
denote the families of all subspaces of
V with dimension and codimension
j and
, respectively.
The following well-known fact is valid without the smoothness of . The proof given in Appendix A for completeness.
Theorem 3. If is a bounded Lipschitz domain, the first eigenvalue to (3) formulated to and defined by (20)-(22) is simple.
Coming back to the Lipschitz bounded domain
, we introduce its deformation as follows. Let
be a family of bi-Lipschitz homeomorphisms. We assume that
is continuous in
t uniformly in
, and continue to use the following definition as in [
9].
Definition 1.
The family of bi-Lipschitz homeomorphisms is said to be p-differentiable in t for , if is p-times differentiable in t for any and the mappings
are uniformly bounded in , where denotes the Jacobi matrix of and stands for the set of real matrices. This is furthermore said to be continuously p-differentiable in t if it is p-differentiable and the mappings
are continuous.
Putting
in (
18), we introduce the other eigenvalue problem
which is reduced to finding
for
Let
be the
j-th eigenvalue of the eigenvalue problem (
27). Then Lemma 7 of [
9] ensures that the eigenvalue problem (
28)-(
29) is reduced to
by the transformation of variables
, where
for
V and
X defined by (
22), and
Recall that
is a bounded Lipshitz domain, and let
,
be a family of twice continuously differentiable bi-Lispchitz transformations. Then the abstract theory described in
§1 is applicable with
and
Henceforth, we write
for the symmetric matrix
, where
denotes the transpose of vectors or matrices. The unit matrix is denoted by
, and
for
and
. Let, furthermore,
be the identity mapping.
Theorem 4.
uniformly on Ω, where are Lipschitz continuous vector fields. Then, if in (8) for , it holds that
where , , , and are defined by (14) with , and .
Proof: First, since
uniformly on
, it holds that
uniformly on
, which implies
uniformly on
(Lemma 5 and Lemma 6 of [
8]).
Second, we have
using the standard
inner product
, and therefore, it follows that
and
By these equalities, finally, we obtain
and
and hence the conclusion.
3. Dynamical Deformations
We continue to suppose that
is a bounded Lipschitz domain. Here we study the dynamical deformation of domains introduced by [
7].
To this end, we take a Lipschitz continuous vector field defined on a neighbourhood of
, denoted by
. Then the transformation
is made by
, where
,
, is the solution to
Then we have the group property,
and therefore, the formulae derived for
are shifted to
.
If
v is a
vector field, furthermore, this
is twice continuously differentiable, and it holds that (
33) with
because of
Then we obtain the following lemma.
Lemma 1.
Under the above assumption, it holds that
Proof: Writing
, we obtain
and the proof is complete. □
Here we take two categories that the vector fields are solenoidal and gradient. In the first category, we assume
everywhere. Then it holds that
Theorem 5.
If , it holds that and
in Theorem 4.
Proof: We recall (
34). By the assumption we obtain
and
by Lemma 1, which implies
, and (
36) by Theorem 4. □
Remark 1.
in short. Then, by the proof of Lemma 1 we obtain
similarly. Hence it holds that
In the second category that the vector field is a gradient of a scalar field we obtain the following theorem.
Theorem 6.
If for the scalar field μ, it holds that
Proof: In this case, we obtain
and
It also holds that
because
is valid to the vector field
a.
Since
is valid by Lemma 1, and hence
Next, we obtain
and hence
Finally, we divide
as in
for
Equality (
39) now implies
for
, and hence
We thus end up with the result by combining these equalities. □
Remark 2.
Theorem 5 and Theorem 6 are consistent if with
In this case it holds that
Then we can reproduce the result of Theorem 5 from (38), as in
cannot have a definite sign because of (41). Second, if
we obtain by (13), which implies
regardless of the form of ϕ. For inequality (45) to hold, however, it is necessary that
by (42). This inequality implies because
for with . In other words, several possibilities can arise to the monotonicity or convexity of the first eigenvalue under the dynamical perturbation , provided with the property (35).
4. Conformal Deformations
Here we sayt that
is conformal if
holds with a scalar field
. It follows that
, and therefore, the matrix
is orthogonal:
.
Then we assume, furthermore, that
for simplicity. It follows that
and hence
, which results in
and hence
and
Then we can show the following theorem.
Theorem 7.
Let , , and be conformal at t. Assume, furthermore, that
Proof: Since
we obtain
Then it holds that
and
in Theorem 1 for
. Then the result follows from Theorem 2. □
A consequence of Theorem 7 is the following isometric inequality.
Theorem 8.
Let be a simply-connected bounded Lipschitz domain and be a univalent bi-Lipschitz homeomorphism, where . Let, furthermore,
Then it holds that
where and are the first eigenvalue of (18):
respectively with , , and furthermore, is the first eigenfunction of (51) such that
Remark 3.
Either or occurs if is simply-connected and (17), and in the former case it follows that and the above theorem is trivial. This assumption (17), however, is used in [9] just to formulate (18) as in (3) with (20)-(21), and (22). Hence we can exclude this assumption if we formulate (50) as in (19) with (20)-(21) for V standing for the closure of in , where if and only if there is , a smooth extension of v in a neighborhood of , such that
Then we re-formulate (51) as in (30) with (31) for , using the bi-Lipshichits homeomorphism . Under this agreement of (50) and (51), we exclude the assumption (17) in the above theorem.
Proof of Theorem 8: Let
be the conformal mapping on
D, and define
and
by (
48) for (
46) with
. We obtain
in (
46) and hence it follows that
and hence
Thus we obtain
for
defined by (
24) with
and
and
, which implies
or
by
and (
12) with (
48). Then we obtain the result by (
53). □
Remark 4.
If , inequality (49) is reduced to
Since is univalent, it holds that
and therefore, if and only if
This equality implies and in particular,
Inequality (
55) for
is proven in Appendix B. We conclude this section with an analogous result to (
49).
Theorem 9.
Under the assumption of the previous theorem it holds that
where for the first eigenfunction to (50) such that
Proof: Similarly to (
54), we obtain
which implies the result by
and
5. Proof of Theorem 3
Although using the Harnack inequality is standard (c.f. Theorem 8.38 of [
4]), here we follow the argument by [
5]. In fact, by (
24) it holds that
Since
is compact, there is
in
which attains
. Then it holds that
We have
for
(Definition 6.7 and Theorem A.1 of [
2]), and therefore,
We thus obtain
because
is a Lipschitz domain. Hence we may assume
in
.
Equality (
56) implies
in the sense of distributions, and therefore, we obtain
by the elliptic regularity. Then the strong maximum principle implies
in
.
Now we use the following lemma derived from (
57).
Lemma 2.
Then it follows that and
Proof: We obtain
by (
57) and hence
because
is a Lipschitz domain. It is obvious that
while
follows from (
57). □
We are ready to give the following proof.
Proof of Theorem 3: It sufficies to show
for any first eigenfunction
. To this end, we take
and put
for
. The desired equality
is thus reduced to
or,
Letting
, we obtain
by Lemma 2, while
C is non-negative definite on
. Hence it holds that
, and hence
are the other first eigenfunctions. We thus obtain
, in particular.
Given
, we obtain
by
. Dividing both sides by
h and making
, we thus obtain
and hence
. It holds that
, and hence (
60). □
6. Proof of (55) for
If , we have , and the result is a direct consequence of the following theorem.
Lemma 3.
If for and is holomorphic in D, it holds that
Proof of Lemma 3: Writing
, we get
Since
on
, it holds that
because
is homeomorphic. Then we obatin
7. Numerical Experiments
In this section, we present numerical experiments to confirm the theoretical results obtained.
Let
and
be the unit disk. Following (
52), we define
by
for
. In this case, the images of the transformation are as follows.
Figure 1.
The images of the transformation of D. The value of t are from (first row, leftmost) to (second row, rightmost) with increments.
Figure 1.
The images of the transformation of D. The value of t are from (first row, leftmost) to (second row, rightmost) with increments.
Then, the eigenvalues of
are computed by the piecewise linear finite element method. By the calculus in the proof of Theorem 8 and Theorem 7, we have
although
is not univalent on
D. The computed profiles of the first eigenvalue and its reciprocal given in
Figure 2 are clearly consistent with Theorem 7.
Next, we define
to confirm the statement of Theorem 8. Note that
is univalent on
D. Then, the images of
and
are given in
Figure 3.
The first five eigenvalues of (
61) on
D and
are given in
Table 1. Because
in this case, the numbers on
are consistent with the inequality (
55). Although Theorem 8 and Remark 4 claim only on the first eigenvalue
, we observe similar inequalities hold for
(
).
References
- R.A. Adams and J.J.F. Fournier, Sobolev Spaces, second edition, Academic Press, Amsterdam, 2025.
- D. Kinderlehrer and G. Stampacchia, An Introduction of Variational Inequalities and their Applications, Academic Press, New York, 1980.
- P.R. Garabedian and M. Schiffer, Convexity of domain functionals, J. Ann. Math. 2 (1952-53) 281–368. [CrossRef]
- D. Gilbarg and N.S. Trudinger, Elliptic Partial Differential Equations of Second Order, second edition, Springer, Berlin, 1983.
- M. Otani and T. Teshima, On the first eigenvalue of some quasilinear elliptic equations, Proc. Japan Acd. 64A (1988) 8-10. [CrossRef]
- F. Rellich, Perturbation Theory of Eigenvalue Problems, Lecture Notes, New York Univ. 1953.
- T. Suzuki and T. Tsuchiya, First and second Hadamard variational formulae of the Green function for general domain perturbations, J. Math. Soc. Japan, 68 (2016) 1389–1419. [CrossRef]
- T. Suzuki and T. Tsuchiya, Liouville’s formulae and Hadamard variation with respect to general domain perturbations, J. Math. Soc. Japan 75 (2023) 983–1024. [CrossRef]
- T. Suzuki and T. Tsuchiya, Hadamard variation of eigenvalues with respect to general domain perturbations, J. Math. Soc. Japan, accepted for publication. https://arxiv.org/abs/2309.00273.
|
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2024 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).