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

Hadamard variation of eigenvalues with respect to general domain perturbations

Takashi Suzuki Center of Mathematical Modeling and Data Science,
Osaka University,
Toyonaka, 560-8531, Japan
suzuki@sigmath.es.osaka-u.ac.jp
   Takuya Tsuchiya Center of Mathematical Modeling and Data Science,
Osaka University,
Toyonaka, 560-8531, Japan
tsuchiya.takuya.plateau@kyudai.jp
Abstract.

We study Hadamard variation of eigenvalues of Laplacian with respect to general domain perturbations. We show their existence up to the second order rigorously and characterize the derivatives, using associated eigenvalue problems in finite dimensional spaces. Then smooth rearrangement of multiple eigenvalues is explicitly given. This result follows from an abstract theory, applicable to general perturbations of symmetric bilinear forms.

Key words and phrases:
eigenvalue problem, perturbation theory of linear operators, domain deformation, Hadamard’s variational formulae, Garabedian-Schiffer’s formula
     This work was promoted in RIMS program for joint research during 2019-2021. The authors thank Professors Hideyuki Azegami and Erika Ushikoshi for many detailed discussions at these occasions. The first author was supported by JSPS Grant-in-Aid for Scientific Research 19H01799. The second author was supported by JSPS Grant-in-Aid for Scientific Research 21K03372
2020 Mathematics Subject Classification:
Primary 35J25; Secondary 35R35

1. Introduction

Our purpose is to study Hadamard variation of eigenvalues of the Laplace operator with mixed boundary conditions. We characterize the first and the second derivatives, using associated finite dimensional eigenvalue problems, particularly, for multiple eigenvalues.

Hadamard’s variational formulae are used to provide effective numerical schemes for shape optimization and free boundary problems. In [12, 13] we have introduced a new variational formulation to filtration problem and applied the Hadamard variation. We have also derived Hadamard’s variational formulae for Green’s functions of the Poisson equation, extending Liouville’s volume and area formulae up to the second order [14, 15]. This paper is devoted to the eigenvalue problem concerning general perturbation of Lipschitz domains. Meanwhile we develop abstract theory applicable to other problems.

So far, C1C^{1} and analytic rearrangements of multiple eigenvalues of self-adjoint operator have been discussed, for example, pp. 44–52 of [10] and p.487 of [1]. There are, however, several comments on the complexity of the proof on C1C^{1} category, as in pp. 122–123 of [5] and p. 490 of [1]. In this paper we examine this process of rearrangement in details, using the above described characterization of the derivatives, to ensure their efficiencies in C2C^{2} categories.

Let Ω\Omega be a bounded Lipschitz domain in nn-dimensional Euclidean space n\mathbb{R}^{n} for n2n\geq 2. Suppose that its boundary Ω\partial\Omega is divided into two relatively open sets γ0\gamma_{0} and γ1\gamma_{1}, satisfying

γ0¯γ1¯=Ω,γ0¯γ1¯=.\overline{\gamma_{0}}\cup\overline{\gamma_{1}}=\partial\Omega,\quad\overline{\gamma_{0}}\cap\overline{\gamma_{1}}=\emptyset. (1)

As in [15], we thus avoid the technical difficulty described in p. 272 of [2], when γi\gamma_{i}, i=0,1i=0,1 have their boundaries on Ω\partial\Omega.

We study the eigenvalue problem of the Poisson equation with mixed boundary condition,

Δu=λuin Ω,u=0on γ0,uν=0on γ1,-\Delta u=\lambda u\ \mbox{in $\Omega$},\quad u=0\ \mbox{on $\gamma_{0}$},\quad\frac{\partial u}{\partial\nu}=0\ \mbox{on $\gamma_{1}$}, (2)

where

Δ=i=1n2xi2\Delta=\sum_{i=1}^{n}\frac{\partial^{2}}{\partial x_{i}^{2}}

is the Laplacian and ν\nu denotes the outer unit normal vector on Ω\partial\Omega. This problem takes the weak form, finding uu satisfying

uV,B(u,u)=1,A(u,v)=λB(u,v),vVu\in V,\ B(u,u)=1,\quad A(u,v)=\lambda B(u,v),\ \forall v\in V (3)

defined for

A(u,v)=Ωuvdx,B(u,v)=Ωuv𝑑xA(u,v)=\int_{\Omega}\nabla u\cdot\nabla v\ dx,\quad B(u,v)=\int_{\Omega}uv\ dx

and

V={vH1(Ω)v|γ0=0}.V=\{v\in H^{1}(\Omega)\mid\left.v\right|_{\gamma_{0}}=0\}. (4)

This VV is a closed subspace of H1(Ω)H^{1}(\Omega) under the norm

vV=v22+v22.\|v\|_{V}=\sqrt{\|\nabla v\|_{2}^{2}+\|v\|_{2}^{2}}.

To justify the above weak formulation, we confirm several fundamental facts on the Lipschitz domain [6]. First, the trace operator

vC(Ω¯)v|ΩC(Ω)v\in C^{\infty}(\overline{\Omega})\mapsto\left.v\right|_{\partial\Omega}\in C(\partial\Omega)

defined for

C(Ω¯)={vΩ¯v~C0(n),v~|Ω¯=v}C^{\infty}(\overline{\Omega})=\{v\in\overline{\Omega}\rightarrow\mathbb{R}\mid\exists\tilde{v}\in C_{0}^{\infty}(\mathbb{R}^{n}),\ \left.\tilde{v}\right|_{\overline{\Omega}}=v\}

is so extended as a bounded linear operator

vH1(Ω)v|ΩH1/2(Ω),v\in H^{1}(\Omega)\ \mapsto\ \left.v\right|_{\partial\Omega}\in H^{1/2}(\partial\Omega),

and there arises the isomorphism

vV/H01(Ω)v|γ1H1/2(γ1).v\in V/H^{1}_{0}(\Omega)\ \mapsto\left.v\right|_{\gamma_{1}}\in H^{1/2}(\gamma_{1}).

Second, if ΔvV\Delta v\in V^{\prime} is satisfied in the sense of distributions in Ω\Omega, the normal derivative of vVv\in V on γ1\gamma_{1} is defined as in

vνH1/2(γ1)=(H1/2(γ1)),\frac{\partial v}{\partial\nu}\in H^{-1/2}(\gamma_{1})=(H^{1/2}(\gamma_{1}))^{\prime},

and it holds that

φ,vνH1/2(γ1),H1/2(γ1)=(v,φ)L2(Ω)+φ,ΔvV,V,φV,\left\langle\varphi,\frac{\partial v}{\partial\nu}\right\rangle_{H^{1/2}(\gamma_{1}),H^{-1/2}(\gamma_{1})}=(\nabla v,\nabla\varphi)_{L^{2}(\Omega)}+\langle\varphi,\Delta v\rangle_{V,V^{\prime}},\quad\forall\varphi\in V,

where ,Y,Y\langle\cdot,\cdot\rangle_{Y,Y^{\prime}} denotes the paring between the Banach space YY and its dual space YY^{\prime}. See Theorem 2 of [15].

To confirm the well-posedness of (3), we note, first, that if γ0\gamma_{0}\neq\emptyset, there is coercivity of A:V×VA:V\times V\rightarrow\mathbb{R}, which means the existence of δ>0\delta>0 such that

A(v,v)δvV2,vV.A(v,v)\geq\delta\|v\|_{V}^{2},\quad\forall v\in V. (5)

If γ0=\gamma_{0}=\emptyset we replace AA by A+BA+B, denoted by A~\tilde{A}. Then this A~:V×V\tilde{A}:V\times V\rightarrow\mathbb{R} is coercive, and the eigenvalue problem

uV,B(u,u)=1,A~(u,v)=λ~B(u,v),vV,u\in V,\ B(u,u)=1,\quad\tilde{A}(u,v)=\tilde{\lambda}B(u,v),\ \forall v\in V,

is equivalent to (3) by λ~=λ+1\tilde{\lambda}=\lambda+1. Henceforth, we assume (5), using this reduction if it is necessary.

Second, we note that A:V×V𝐑A:V\times V\rightarrow{\bf R} and B:X×X𝐑B:X\times X\rightarrow{\bf R} are bounded, coercive, and symmetric bilinear forms for X=L2(Ω)X=L^{2}(\Omega). Since VXV\hookrightarrow X is compact, there is a sequence of eigenvalues to (3), denoted by

0<λ1λ2+.0<\lambda_{1}\leq\lambda_{2}\leq\cdots\rightarrow+\infty.

The associated eigenfunctions, u1,u2,u_{1},u_{2},\cdots, furthermore, form a complete ortho-normal system in XX, provided with the inner product induced by B=B(,)B=B(\cdot,\cdot):

B(ui,uj)=δij,A(uj,v)=λjB(uj,v),vV,i,j=1,2,.B(u_{i},u_{j})=\delta_{ij},\quad A(u_{j},v)=\lambda_{j}B(u_{j},v),\ \forall v\in V,\quad i,j=1,2,\cdots.

The jj-th eigenvalue of (3) is given by the mini-max principle

λj=minLjmaxvLj{0}R[v]=maxWjminvWj{0}R[v],\lambda_{j}=\min_{L_{j}}\max_{v\in L_{j}\setminus\{0\}}R[v]=\max_{W_{j}}\min_{v\in W_{j}\setminus\{0\}}R[v], (6)

where

R[v]=A(v,v)B(v,v)R[v]=\frac{A(v,v)}{B(v,v)}

is the Rayleigh quotient, and {Lj}\{L_{j}\} and {Wj}\{W_{j}\} denote the families of all subspaces of VV with dimension and codimension jj and j1j-1, respectively. See [11], for example, for these fundamental facts.

Let

Tt:ΩΩt=Tt(Ω),|t|<ε0T_{t}:\Omega\rightarrow\Omega_{t}=T_{t}(\Omega),\quad|t|<\varepsilon_{0} (7)

be a family of bi-Lipschitz homeomorphisms for ε0>0\varepsilon_{0}>0, satisfying T0=IT_{0}=I, the identity mapping. We assume that TtxT_{t}x is continuous in tt uniformly in xΩx\in\Omega, and recall the following definition used in [15].

Definition 1.

The family {Tt}\{T_{t}\} of bi-Lipschitz homeomorphisms is said to be pp-differentiable in tt for p1p\geq 1, if TtxT_{t}x is pp-times differentiable in tt for any xΩx\in\Omega and the mappings

tDTt,t(DTt)1:ΩMn(),0p\frac{\partial^{\ell}}{\partial t^{\ell}}DT_{t},\ \frac{\partial^{\ell}}{\partial t^{\ell}}(DT_{t})^{-1}:\Omega\rightarrow M_{n}(\mathbb{R}),\quad 0\leq\ell\leq p\\

are uniformly bounded in (x,t)Ω×(ε0,ε0)(x,t)\in\Omega\times(-\varepsilon_{0},\varepsilon_{0}), where DTtDT_{t} denotes the Jacobi matrix of Tt:ΩΩtT_{t}:\Omega\rightarrow\Omega_{t} and Mn()M_{n}(\mathbb{R}) stands for the set of real n×nn\times n matrices. This {Tt}\{T_{t}\} is said to be continuously pp-differentiable in tt if it is pp-differentiable and the mappings

t(ε0,ε0)tDTt,t(DTt)1L(ΩMn()),0pt\in(-\varepsilon_{0},\varepsilon_{0})\mapsto\frac{\partial^{\ell}}{\partial t^{\ell}}DT_{t},\ \frac{\partial^{\ell}}{\partial t^{\ell}}(DT_{t})^{-1}\in L^{\infty}(\Omega\rightarrow M_{n}(\mathbb{R})),\quad 0\leq\ell\leq p

are continuous.

Putting

Tt(γi)=γit,i=0,1,T_{t}(\gamma_{i})=\gamma_{it},\quad i=0,1, (8)

we introduce the other eigenvalue problem

Δu=λuin Ωt,u=0on γ0t,uν=0on γ1t,-\Delta u=\lambda u\ \mbox{in $\Omega_{t}$},\quad u=0\ \mbox{on $\gamma_{0t}$},\quad\frac{\partial u}{\partial\nu}=0\ \mbox{on $\gamma_{1t}$}, (9)

which is reduced to finding

uVt,Ωtu2𝑑x=1,Ωtuvdx=λΩtuv𝑑x,vVtu\in V_{t},\ \int_{\Omega_{t}}u^{2}\ dx=1,\quad\int_{\Omega_{t}}\nabla u\cdot\nabla v\ dx=\lambda\int_{\Omega_{t}}uv\ dx,\ \forall v\in V_{t} (10)

for

Vt={vH1(Ωt)v|γ0t=0}.V_{t}=\{v\in H^{1}(\Omega_{t})\mid\left.v\right|_{\gamma_{0t}}=0\}. (11)

Let λj(t)\lambda_{j}(t) be the jj-th eigenvalue of the eigenvalue problem (9). In Section 4 we confirm that the eigenvalue problem (10)-(11) is reduced to

uV,Bt(u,u)=1,At(u,v)=λBt(u,v),vVu\in V,\ B_{t}(u,u)=1,\quad A_{t}(u,v)=\lambda B_{t}(u,v),\ \forall v\in V (12)

by the transformation of variables y=Ttxy=T_{t}x for VH1(Ω)V\subset H^{1}(\Omega) defined by (4), where

Bt(u,v)=Ωuvat𝑑x,At(u,v)=ΩQt[u,v]at𝑑x,B_{t}(u,v)=\int_{\Omega}uva_{t}\ dx,\quad A_{t}(u,v)=\int_{\Omega}Q_{t}[\nabla u,\nabla v]a_{t}\ dx, (13)

and

at=detDTt,Qt=(DTt)1(DTt)1T.a_{t}=\det DT_{t},\quad Q_{t}=(DT_{t})^{-1}(DT_{t})^{-1T}. (14)

Several representation formulae for

λj(t)=limh01h(λj(t+h)λj(t))\lambda^{\prime}_{j}(t)=\lim_{h\rightarrow 0}\frac{1}{h}(\lambda_{j}(t+h)-\lambda_{j}(t))

and

λj′′(t)=limh01h(λj(t+h)λj(t))\lambda_{j}^{\prime\prime}(t)=\lim_{h\rightarrow 0}\frac{1}{h}(\lambda^{\prime}_{j}(t+h)-\lambda^{\prime}_{j}(t))

have been derived in [3]. Here we characterize these derivatives, using associated finite dimensional eigenvalue problems (Theorems 1 and 4). We prove also the existence of derivatives, particularly, the second derivatives of multiple eigenvalues. These results follow from the differentiability of AtA_{t} and BtB_{t} in tt, defined by

A˙t(u,v)=ddtAt(u,v),B˙t(u,v)=ddtBt(u,v),u,vV\dot{A}_{t}(u,v)=\frac{d}{dt}A_{t}(u,v),\ \dot{B}_{t}(u,v)=\frac{d}{dt}B_{t}(u,v),\quad u,v\in V (15)

and

A¨t(u,v)=d2dt2At(u,v),B¨t(u,v)=d2dt2Bt(u,v),u,vV.\ddot{A}_{t}(u,v)=\frac{d^{2}}{dt^{2}}A_{t}(u,v),\ \ddot{B}_{t}(u,v)=\frac{d^{2}}{dt^{2}}B_{t}(u,v),\quad u,v\in V. (16)

Then we show C1C^{1} and C2C^{2} rearrangements of eigenvalues, if these bilinear forms are continuous in tt (Theorems 3 and 6). We give the algorithm explicitly, as the transversal rearrangement in Definition 3. Consequently, no rearrangment is necessary to confirm C1C^{1} or C2C^{2} smoothness of eigenvalues, if their multiplicities are constant. Also elementary symmetric functions made by possible multiple eigenvalues are C1C^{1} or C2C^{2}. These properties are noticed by [7] and [8] in the real analytic category.

In [4], unilateral derivatives of the first order of the eigenvalue of the Stokes operator are calculated. It characterizes the derivatives using boundary integrals when the deformation of the domain is of normal direction. Our abstract theory reproduces and extends this result by the use of Piola transformation described in [9].

Our argument is executed in the H1H^{1} category without requiring any further elliptic regularities. Hence the Lipschitz continuity of Ω\partial\Omega is sufficient to ensure all the results on (9) under the general perturbation of domains. We recall, in this context, that numerical computations on partial differential equations are mostly executed on Lipschitz domains. There, perturbation of the domain using Lipschitz continuous vector fields is often applied. This is the method of trial domains, efficient even to the case that normal perturbation of domains, called in [12], does not work because of the presence of corners on the boundary [12, 13].

2. Summary

As is noted in the previous section, C1C^{1} and analytic categories for the smoothness of λj(t)\lambda_{j}(t) in tt have been discussed. Since each eigenvalue λ\lambda of (12) is isolated, we can reduce this problem to a finite dimensional eigenvalue problem by the Lyapunov-Schmidt reduction as in p.486 of [1]. Rellich [10] in p.45 showed in this case that the continuous differentiablity in tt of AtA_{t} and BtB_{t} implies that of λj(t)\lambda_{j}(t) under a suitable change of its order. Kato [5] in p.123 then provided an alternative proof of this C1C^{1} rearrangement.

The other category of analyticity is studied in Chapter II of [10] and p.370 of [5]. If AtA_{t} and BtB_{t} are analytic in tt, the eigenvlue problem is reduced to an algebraic equation with analytic coefficients in tt. The Puiseux expansion of λj(t)\lambda_{j}(t) at t=0t=0 is described in p.31 of [10] and Chapter 2, Section 1 of [5]. Hence this λj(t)\lambda_{j}(t) is realized as an analytic function defined on a Riemann surface.

We study C2C^{2} category. To begin with, we confirm Rellich’s theorem on C1C^{1} category, the continuous differentiablity of rearranged eigenvalues. Here we show this rearrangement explicitly, to reach existence, characterization, and continuity of the second derivatives (Definition 3). In more details, first, the existence of A˙t\dot{A}_{t} and B˙t\dot{B}_{t} in (15) implies that of the first unilateral derivatives

λ˙j±(t)=limh±01h(λj(t+h)λj(t))\dot{\lambda}^{\pm}_{j}(t)=\lim_{h\rightarrow\pm 0}\frac{1}{h}(\lambda_{j}(t+h)-\lambda_{j}(t)) (17)

for each jj and tt. These derivatives, furthermore, are characterized by the other finite dimensional eigenvalue problems in accordance with the multiplicity of λj(t)\lambda_{j}(t) (Theorem 12). Second, if the above A˙t\dot{A}_{t} and B˙t\dot{B}_{t} are continuous in t, and if

λk1(t)<λk(t)λk+m1(t)<λk+m(t)\lambda_{k-1}(t)<\lambda_{k}(t)\leq\cdots\leq\lambda_{k+m-1}(t)<\lambda_{k+m}(t) (18)

holds for tI=(ε0,ε0)t\in I=(-\varepsilon_{0},\varepsilon_{0}), there are C1C^{1} curves

C~j,kjk+m1,\tilde{C}_{j},\ k\leq j\leq k+m-1,

made by at most countably many rearrangements of the C0C^{0} curves

Cj={λj(t)tI},kjk+m1.C_{j}=\{\lambda_{j}(t)\mid t\in I\},\quad k\leq j\leq k+m-1.

(Theorem 14).

In this paper, we notice two properties of the unilateral derivatives λ˙j±\dot{\lambda}_{j}^{\pm}, for the proof of this Rellich’s theorem. First, there arises that

λ˙j+(t)=λ˙2+n1j(t),j+n1\dot{\lambda}_{j}^{+}(t)=\dot{\lambda}_{2\ell+n-1-j}^{-}(t),\quad\ell\leq j\leq\ell+n-1

if

λ1(t)<λ(t)==λ+n1(t)<λ+n(t)\lambda_{\ell-1}(t)<\lambda_{\ell}(t)=\cdots=\lambda_{\ell+n-1}(t)<\lambda_{\ell+n}(t) (19)

holds for k\ell\geq k, +nm\ell+n\leq m (Theorem 12). Second, the unilateral derivative λ˙j±\dot{\lambda}_{j}^{\pm} are provided with the unilateral continuity as in

limh±0λ˙j±(t+h)=λ˙j±(t)\lim_{h\rightarrow\pm 0}\dot{\lambda}_{j}^{\pm}(t+h)=\dot{\lambda}_{j}^{\pm}(t)

if both A˙t\dot{A}_{t} and B˙t\dot{B}_{t} are continuous in tIt\in I (Theorem 13).

As for the second derivatives of eigenvalues, we assume the existence of A¨t\ddot{A}_{t} and B¨t\ddot{B}_{t} in (16) besides A˙t\dot{A}_{t} and B˙t\dot{B}_{t}. Then each j=1,2,j=1,2,\cdots admits the existence of

λ¨j±(t)=limh±02h2(λj(t+h)λj(t)hλ˙j±(t)).\ddot{\lambda}_{j}^{\pm}(t)=\lim_{h\rightarrow\pm 0}\frac{2}{h^{2}}(\lambda_{j}(t+h)-\lambda_{j}(t)-h\dot{\lambda}_{j}^{\pm}(t)). (20)

These limits are again unilateral and characterized by the other eigenvalue problem in a finite dimensional space (Theorem 15). The unilateral continuity of these λ¨j±(t)\ddot{\lambda}_{j}^{\pm}(t) is then assured under the continuity of A¨t\ddot{A}_{t} and B¨t\ddot{B}_{t} in tt, similarly. These properties induce C2C^{2} smoothness of C~j\tilde{C}_{j}, kjk+m1k\leq j\leq k+m-1, once their C1C^{1} smoothness is achieved (Theorem 24).

This paper is composed of eight sections. The results on the Hadamard variation to (9) are described in the next section. In Section 4 we use the transformation of variables to introduce an abstract setting of the problem. Section 5 is concerned on the continuity of eigenvalues and eigenspaces. We study the first derivative of eigenvalues in Section 6, and then its C1C^{1} rearrangement in Section 7. Finally, Section 8 is devoted to the second derivatives.

3. Hadamard variation

Here we state the results on the first and the second derivatives of λj=λj(t)\lambda_{j}=\lambda_{j}(t) in (12). Fix tIt\in I, and assume (19) for =k\ell=k and n=mn=m for k,m=1,2,k,m=1,2,\cdots with the convention λ0(t)=\lambda_{0}(t)=-\infty. Put

λλk(t)==λk+m1(t),\lambda\equiv\lambda_{k}(t)=\cdots=\lambda_{k+m-1}(t), (21)

and let YtλY^{\lambda}_{t}, dimYtλ=m\dim Y^{\lambda}_{t}=m, be the eigenspace corresponding to this eigenvalue λ\lambda.

Theorem 1.

Assume the above situation, and let {Tt}\{T_{t^{\prime}}\} be 11-differentiable at t=tt^{\prime}=t. Then, there exist the unilateral limits

λ˙j±=limh±01h(λj(t+h)λ),kjk+m1,\dot{\lambda}_{j}^{\pm}=\lim_{h\rightarrow\pm 0}\frac{1}{h}(\lambda_{j}(t+h)-\lambda),\quad k\leq j\leq k+m-1,

which satisfies

νjλ˙j+=λ˙2k+m1j,kjk+m1.\nu_{j}\equiv\dot{\lambda}_{j}^{+}=\dot{\lambda}_{2k+m-1-j}^{-},\quad k\leq j\leq k+m-1. (22)

This νj\nu_{j} is the qq-th eigenvalue of the matrix

Gtλ=(Etλ(ϕ~i,ϕ~j))1i,jmG^{\lambda}_{t}=\left(E^{\lambda}_{t}(\tilde{\phi}_{i},\tilde{\phi}_{j})\right)_{1\leq i,j\leq m} (23)

for q=jk+1q=j-k+1, where {ϕ~j1jm}\{\tilde{\phi}_{j}\mid 1\leq j\leq m\} is a basis of YtλY^{\lambda}_{t}, satisfying

Bt(ϕ~i,ϕ~j)=δij,1i,jmB_{t}(\tilde{\phi}_{i},\tilde{\phi}_{j})=\delta_{ij},\quad 1\leq i,j\leq m (24)

and

Etλ=A˙tλB˙tE^{\lambda}_{t}=\dot{A}_{t}-\lambda\dot{B}_{t}

for A˙t\dot{A}_{t} and B˙t\dot{B}_{t} defined by (13), (14), and (15).

Remark 1.

If ϕj1jm\langle\phi_{j}\mid 1\leq j\leq m\rangle is the other basis of YtλY^{\lambda}_{t} satisfying

Bt(ϕi,ϕj)=δij,1i,jm,B_{t}(\phi_{i},\phi_{j})=\delta_{ij},\quad 1\leq i,j\leq m, (25)

it holds that

ϕj=i=1mqjiϕ~i, 1jm\phi_{j}=\sum_{i=1}^{m}q^{i}_{j}\tilde{\phi}_{i},\ 1\leq j\leq m

with the orthogonal m×mm\times m matrix Q=(qji)Q=(q^{i}_{j}). Hence νj\nu_{j}, kjk+m1k\leq j\leq k+m-1, in (22) is determined, indpendent of the choice of ϕ~j1jm\langle\tilde{\phi}_{j}\mid 1\leq j\leq m\rangle.

Remark 2.

Under (21), it holds that

λ˙k+(t)λ˙k+m1+(t),λ˙k(t)λ˙k+m1(t).\dot{\lambda}_{k}^{+}(t)\leq\cdots\leq\dot{\lambda}_{k+m-1}^{+}(t),\quad\dot{\lambda}_{k}^{-}(t)\geq\cdots\geq\dot{\lambda}_{k+m-1}^{-}(t).

A direct consequnece of this theorem is the existence of the unilateral derivatives λ˙j±(t)\dot{\lambda}_{j}^{\pm}(t) in (17) for any tIt\in I and j=1,2,j=1,2,\cdots, if {Tt}\{T_{t}\} is 11-differentiable in II. Then we obtain the following theorem.

Theorem 2.

If {Tt}\{T_{t}\} is continuously 11-differentiable in tIt\in I, the functions λ˙j±=λ˙j±(t)\dot{\lambda}^{\pm}_{j}=\dot{\lambda}^{\pm}_{j}(t) are unilaterally continuous, so that it holds that

limh±0λ˙j±(t+h)=λ˙j±(t)\lim_{h\rightarrow\pm 0}\dot{\lambda}_{j}^{\pm}(t+h)=\dot{\lambda}_{j}^{\pm}(t)

for any tt and jj.

This fact ensures the following theorem of Rellich.

Theorem 3.

Let {Tt}\{T_{t}\} be continuously 11-differentiable in tt, and assume

λk1(t)<λk(t)λk+m1(t)<λk+m(t),tI\lambda_{k-1}(t)<\lambda_{k}(t)\leq\cdots\leq\lambda_{k+m-1}(t)<\lambda_{k+m}(t),\quad\forall t\in I

for some k,m=1,2,k,m=1,2,\cdots. Let

Cj={λj(t)tI},kjk+m1C_{j}=\{\lambda_{j}(t)\mid t\in I\},\quad k\leq j\leq k+m-1

be C0C^{0} curves. Then, there exist C1C^{1} curves denoted by C~i\tilde{C}_{i}, kik+m1k\leq i\leq k+m-1, made by a rearrangement of

{Cjkjk+m1}\{C_{j}\mid k\leq j\leq k+m-1\}

at most countably many times.

Turning to the second derivatives, we fix tIt\in I again, and assume that {Tt}\{T_{t^{\prime}}\} is twice differentiable at t=tt^{\prime}=t. Suppose (19) for =k\ell=k and n=mn=m, put λ\lambda as in (21), and let k<rk+mk\leq\ell<r\leq k+m be such that

λ˙1+<λλ˙+==λ˙r1+<λ˙r+\dot{\lambda}^{+}_{\ell-1}<\lambda^{\prime}\equiv\dot{\lambda}^{+}_{\ell}=\cdots=\dot{\lambda}^{+}_{r-1}<\dot{\lambda}^{+}_{r} (26)

in Theorem 1. To state the finite dimensonal eigenvalue problem characterizing the second derivatives of λj(t)\lambda_{j}(t^{\prime}) at t=tt^{\prime}=t for jr1\ell\leq j\leq r-1, we introduce the following definition.

Definition 2.

Let R:X=L2(Ω)YtλR:X=L^{2}(\Omega)\rightarrow Y^{\lambda}_{t} be the orthogonal projection with respect to Bt(,)B_{t}(\cdot,\cdot) and P=IRP=I-R, where I:XXI:X\rightarrow X is the identity operator. Let, furthermore, AtA_{t}, BtB_{t}, A˙t\dot{A}_{t}, B˙t\dot{B}_{t}, A¨t\ddot{A}_{t}, and B¨t\ddot{B}_{t} be as in (13), (14), (15), and (16). Then we define w=γ(u)PVw=\gamma(u)\in PV for uVu\in V by

Ct(w,v)=C˙tλ,λ(u,v),vPV,C_{t}(w,v)=-\dot{C}_{t}^{\lambda,\lambda^{\prime}}(u,v),\quad\forall v\in PV, (27)

where

Ct=AtλBt,C˙tλ,λ=A˙tλB˙tλBt.C_{t}=A_{t}-\lambda B_{t},\quad\dot{C}_{t}^{\lambda,\lambda^{\prime}}=\dot{A}_{t}-\lambda\dot{B}_{t}-\lambda^{\prime}B_{t}.

We put also

Ftλ,λ(u,v)=(A¨tλB¨t2λB˙t)(u,v)2Ct(γ(u),γ(v)),u,vV.F_{t}^{\lambda,\lambda^{\prime}}(u,v)=(\ddot{A}_{t}-\lambda\ddot{B}_{t}-2\lambda^{\prime}\dot{B}_{t})(u,v)-2C_{t}(\gamma(u),\gamma(v)),\quad u,v\in V.
Remark 3.

To confirm the unique solvability of w=γ(u)w=\gamma(u), let Q:L2(Ω)ZtkQ:L^{2}(\Omega)\rightarrow Z^{k}_{t} be the orthogonal projection with respect to Bt(,)B_{t}(\cdot,\cdot), where ZtkZ_{t}^{k} denotes the finite dimensional space of L2(Ω)L^{2}(\Omega) generated by the eigenfunctions corresponding to the eigenvalues λ1(t),,λk1(t)\lambda_{1}(t),\cdots,\lambda_{k-1}(t). Let, furthermore, V0=QVV_{0}=QV and V1=(IQ)RVV_{1}=(I-Q)RV. First, there is a unique w0V0w_{0}\in V_{0} satisfying

Ct(w0,v)=C˙tλ,λ(u,v),vV0C_{t}(w_{0},v)=-\dot{C}_{t}^{\lambda,\lambda^{\prime}}(u,v),\quad\forall v\in V_{0} (28)

because CtC_{t} is negative definite on V0×V0V_{0}\times V_{0}. Second, there is also a unique w1V1w_{1}\in V_{1} satisfying

Ct(w1,v)=C˙tλ,λ(u,v),vV1C_{t}(w_{1},v)=-\dot{C}_{t}^{\lambda,\lambda^{\prime}}(u,v),\quad\forall v\in V_{1} (29)

because CtC_{t} is positive definite on V1×V1V_{1}\times V_{1}. Then we obtain (27) for w=w0+w1w=w_{0}+w_{1}, because

Ct(w0,v)=Ct(w0,Qv),Ct(w1,v)=Ct(w1,(IQ)v),vRV,C_{t}(w_{0},v)=C_{t}(w_{0},Qv),\ C_{t}(w_{1},v)=C_{t}(w_{1},(I-Q)v),\quad\forall v\in RV,

and hence

Ct(w,v)\displaystyle C_{t}(w,v) =\displaystyle= Ct(w0,v)+Ct(w1,v)=Ct(w0,Qv)+Ct(w1,(IQ)v)\displaystyle C_{t}(w_{0},v)+C_{t}(w_{1},v)=C_{t}(w_{0},Qv)+C_{t}(w_{1},(I-Q)v)
=\displaystyle= C˙tλ,λ(u,Qv)C˙tλ,λ(u,(IQ)v)=C˙t(u,v),vRV.\displaystyle-\dot{C}_{t}^{\lambda,\lambda^{\prime}}(u,Qv)-\dot{C}_{t}^{\lambda,\lambda^{\prime}}(u,(I-Q)v)=-\dot{C}_{t}(u,v),\quad\forall v\in RV.

We thus obtain wRVw\in RV satisfying (27). If wRVw\in RV is a solution to (27), conversely, then w0=QwV0w_{0}=Qw\in V_{0} and w1=(IQ)wV1w_{1}=(I-Q)w\in V_{1} solve (28) and (29), respectively. Then there arises the uniqueness of such w=w0+w1w=w_{0}+w_{1} because these w0V0w_{0}\in V_{0} and w1V1w_{1}\in V_{1} are unique.

Recall

Ytλ=ϕ~jkjk+m1Y^{\lambda}_{t}=\langle\tilde{\phi}_{j}\mid k\leq j\leq k+m-1\rangle

with (24).

Theorem 4.

Under the above situation of (19) and (26), there exist

λj′′=limh+02h2(λj(t+h)λhλ),jr1.\lambda^{\prime\prime}_{j}=\lim_{h\rightarrow+0}\frac{2}{h^{2}}(\lambda_{j}(t+h)-\lambda-h\lambda^{\prime}),\quad\ell\leq j\leq r-1. (30)

This λj′′\lambda^{\prime\prime}_{j} is the qq-th eigenvalue of the matrix

Htλ,λ=(Ftλ,λ(ϕ~i,ϕ~j))i,jr1H^{\lambda,\lambda^{\prime}}_{t}=\left(F^{\lambda,\lambda^{\prime}}_{t}(\tilde{\phi}_{i},\tilde{\phi}_{j})\right)_{\ell\leq i,j\leq r-1} (31)

for q=j+1q=j-\ell+1. If (26) is replaced by

λ˙1(t)>λλ˙(t)==λ˙r1(t)>λ˙r(t),\dot{\lambda}^{-}_{\ell-1}(t)>\lambda^{\prime}\equiv\dot{\lambda}^{-}_{\ell}(t)=\cdots=\dot{\lambda}_{r-1}^{-}(t)>\dot{\lambda}_{r}^{-}(t),

there arises that

λj′′(t)=limh02h2(λj(t+h)λhλ),jr1.\lambda_{j}^{\prime\prime}(t)=\lim_{h\rightarrow-0}\frac{2}{h^{2}}(\lambda_{j}(t+h)-\lambda-h\lambda^{\prime}),\quad\ell\leq j\leq r-1.

As in Remark 1 on the first derivative, the above λj′′\lambda^{\prime\prime}_{j}, jr1\ell\leq j\leq r-1, are determined independent of the choice of

ϕ~jjr1.\langle\tilde{\phi}_{j}\mid\ell\leq j\leq r-1\rangle.

Theorems 1, 4 imply also the existence of the unilateral limits λ¨j±(t)\ddot{\lambda}_{j}^{\pm}(t) in (20) for any tt and jj if {Tt}\{T_{t}\} is 22-differentiable in tIt\in I, that is,

λ¨j±(t)=limh±02h2(λj(t+h)λj(t)hλ˙j±(t)).\ddot{\lambda}_{j}^{\pm}(t)=\lim_{h\rightarrow\pm 0}\frac{2}{h^{2}}(\lambda_{j}(t+h)-\lambda_{j}(t)-h\dot{\lambda}_{j}^{\pm}(t)).
Remark 4.

By Liouville’s theorem on general perturbation of domains studied in [15], the matrix GtλG_{t}^{\lambda} in (23) is represented by the surface integrals of ϕ~j\tilde{\phi}_{j}, 1jm1\leq j\leq m. This property is confirmed by [4] for the Stokes operator with Dirichlet condition under a special perturbation of domains, called the normal perturbation in [14]. Similarly, the matrix Htλ,λH_{t}^{\lambda,\lambda^{\prime}} in (31) is represented by the surface integrals of ϕ~j\tilde{\phi}_{j} and γ(ϕ~j)\gamma(\tilde{\phi}_{j}), 1jm1\leq j\leq m.

Then we obtain the following theorems.

Theorem 5.

If {Tt}\{T_{t}\} is continuously 22-differentiable in tt, then λ¨j±=λ¨j±(t)\ddot{\lambda}_{j}^{\pm}=\ddot{\lambda}_{j}^{\pm}(t) are unilaterally continuous, so that it holds that

limh±0λ¨j±(t+h)=λ¨j±(t)\lim_{h\rightarrow\pm 0}\ddot{\lambda}_{j}^{\pm}(t+h)=\ddot{\lambda}_{j}^{\pm}(t)

for any tIt\in I and j=1,2,j=1,2,\cdots.

Theorem 6.

If {Tt}\{T_{t}\} is continuously 22-differentiable in tt, the C1C^{1} curves C~j\tilde{C}_{j}, kjk+m1k\leq j\leq k+m-1 in Theorem 3 are C2C^{2}.

Although the above theorems are to be extended to p3p\geq 3 of Definition 1, we restrict ourselves to p=1,2p=1,2 in this paper. Yet these results on p=2p=2 are efficient to examine the harmonic concavity of λj(t)\lambda_{j}(t) in tt studied by [3] for the first eigenvalue to (2) with γ1=\gamma_{1}=\emptyset. We emphasize also that the treatment of the second derivatives is rather different from that of the first ones. See Remark 10 in Section 8.

4. Reduction to the abstract theory

For the moment, we fix tt and treat the bi-Lipschitz homeomorphism T=Tt:ΩΩt=TtΩT=T_{t}:\Omega\rightarrow\Omega_{t}=T_{t}\Omega. Let

Ω~=Ωt,f(y)=g(x),y=Tx,\tilde{\Omega}=\Omega_{t},\quad f(y)=g(x),\ y=Tx,

and confirm the chain rule for this transformation of variables, that is,

g=fDT,dy=(detDT)dx\nabla g=\nabla f\ DT,\quad dy=(\det DT)dx

for g=(gx1,,gxn)\nabla g=(\frac{\partial g}{\partial x_{1}},\cdots,\frac{\partial g}{\partial x_{n}}), f=(fy1,,fyn)\nabla f=(\frac{\partial f}{\partial y_{1}},\cdots,\frac{\partial f}{\partial y_{n}}), and

DT=(y1x1y1xnynx1ynxn).DT=\left(\begin{array}[]{ccc}\frac{\partial y_{1}}{\partial x_{1}}&\cdots&\frac{\partial y_{1}}{\partial x_{n}}\\ \cdot&&\cdot\\ \frac{\partial y_{n}}{\partial x_{1}}&\cdots&\frac{\partial y_{n}}{\partial x_{n}}\end{array}\right).

Putting γ~i=γit=Tγi\tilde{\gamma}_{i}=\gamma_{it}=T\gamma_{i} for i=0,1i=0,1, we take the eigenvalue problem

Δu=λuin Ω~,u=0on γ~0,uν=0on γ~1,-\Delta u=\lambda u\ \mbox{in $\tilde{\Omega}$},\quad u=0\ \mbox{on $\tilde{\gamma}_{0}$},\quad\frac{\partial u}{\partial\nu}=0\ \mbox{on $\tilde{\gamma}_{1}$}, (32)

that is, (9) for T=TtT=T_{t}. Let

V~={vH1(Ω~)v|γ~0=0},\tilde{V}=\{v\in H^{1}(\tilde{\Omega})\mid\left.v\right|_{\tilde{\gamma}_{0}}=0\},

and introduce the weak form of (32),

uV~,A~(u,v)=λB~(u,v),vV~,u\in\tilde{V},\ \tilde{A}(u,v)=\lambda\tilde{B}(u,v),\quad\forall v\in\tilde{V}, (33)

where

A~(u,v)=Ω~yuyvdy,B~(u,v)=Ω~uv𝑑y.\tilde{A}(u,v)=\int_{\tilde{\Omega}}\nabla_{y}u\cdot\nabla_{y}v\ dy,\quad\tilde{B}(u,v)=\int_{\tilde{\Omega}}uv\ dy. (34)

Given ϕV\phi\in V, put

ψ(y)=ϕ(x),y=Tx.\psi(y)=\phi(x),\quad y=Tx.

Then it holds that

ϕVψV~\phi\in V\ \Leftrightarrow\ \psi\in\tilde{V}

for VH1(Ω)V\subset H^{1}(\Omega) defined by (4). Writing

U(x)=u(y),V(x)=v(y),y=Tx,U(x)=u(y),\ V(x)=v(y),\quad y=Tx, (35)

we obtain

yuyv\displaystyle\nabla_{y}u\cdot\nabla_{y}v =\displaystyle= [xU(DT)1][xV(DT)1]\displaystyle[\nabla_{x}U(DT)^{-1}]\cdot[\nabla_{x}V(DT)^{-1}]
=\displaystyle= (xU)(DT)1(DT)1T(xV)T\displaystyle(\nabla_{x}U)(DT)^{-1}(DT)^{-1T}(\nabla_{x}V)^{T}
=\displaystyle= (xU)Q(xV)T=Q[xU,xV]\displaystyle(\nabla_{x}U)Q(\nabla_{x}V)^{T}=Q[\nabla_{x}U,\nabla_{x}V]

for Q=(DT)1(DT)1TQ=(DT)^{-1}(DT)^{-1T}, where FTF^{T} denotes the transpose of the matrix FF. Then, (33) means

ΩQ[xU,xV]a𝑑x=λΩUVa𝑑x\int_{\Omega}Q[\nabla_{x}U,\nabla_{x}V]a\ dx=\lambda\int_{\Omega}UVa\ dx

for a=detDTa=\det DT. The condition of normalization

Ω~u2𝑑y=1\int_{\tilde{\Omega}}u^{2}\ dy=1

is also transformed into

ΩU2a𝑑x=1.\int_{\Omega}U^{2}a\ dx=1.

Under the family {Tt}\{T_{t}\} of homeomorphisms, therefore, the weak form (10) of (9), is equivalent to (12) for VH1(Ω)V\subset H^{1}(\Omega) defined by (4) and BtB_{t}, AtA_{t} defined by (13)-(14). Here we confirm the following lemma.

Lemma 7.

The jj-th eigenvalue of (9) is equal to that of (12).

Proof.

For the moment, let λ~j(t)\tilde{\lambda}_{j}(t) be the jj-th eigenvalues of (9) and let λj(t)\lambda_{j}(t) be that of (12) for (13) and (14). By the mini-max principle (6), it holds that

λ~j(t)=minL~jmaxvL~j{0}R~t[v]=maxW~jminvW~j{0}R~t[v],\tilde{\lambda}_{j}(t)=\min_{\tilde{L}_{j}}\max_{v\in\tilde{L}_{j}\setminus\{0\}}\tilde{R}_{t}[v]=\max_{\tilde{W}_{j}}\min_{v\in\tilde{W}_{j}\setminus\{0\}}\tilde{R}_{t}[v], (36)

where VtH1(Ωt)V_{t}\subset H^{1}(\Omega_{t}) is defined by (11),

R~t[v]=A~t(v,v)B~t(v,v),A~t(u,v)=Ωtuvdx,B~t(u,v)=Ωtuv𝑑x,\tilde{R}_{t}[v]=\frac{\tilde{A}_{t}(v,v)}{\tilde{B}_{t}(v,v)},\quad\tilde{A}_{t}(u,v)=\int_{\Omega_{t}}\nabla u\cdot\nabla v\ dx,\quad\tilde{B}_{t}(u,v)=\int_{\Omega_{t}}uv\ dx,

and {L~j}\{\tilde{L}_{j}\} and {W~j}\{\tilde{W}_{j}\} are the families of all subspaces of VtV_{t} with dimension and codimension jj and j1j-1, respectively.

It holds also that

λj(t)=minLjmaxvLj{0}Rt[v]=maxWjminvWj{0}Rt[v],\lambda_{j}(t)=\min_{L_{j}}\max_{v\in L_{j}\setminus\{0\}}R_{t}[v]=\max_{W_{j}}\min_{v\in W_{j}\setminus\{0\}}R_{t}[v], (37)

for

Rt[v]=At(v,v)Bt(v,v),R_{t}[v]=\frac{A_{t}(v,v)}{B_{t}(v,v)}, (38)

and {Lj}\{L_{j}\} and {Wj}\{W_{j}\} denote the families of all subspaces of VV with dimension and codimension jj and j1j-1, respectively.

If the set LL is a jj-dimensional subspace of VV there is ϕL\phi_{\ell}\in L, 1j1\leq\ell\leq j, such that

Ωϕϕ𝑑x=δ\int_{\Omega}\phi_{\ell}\phi_{\ell^{\prime}}dx=\delta_{\ell\ell^{\prime}}

and

ϕ==1jcϕ,c=Ωϕϕ𝑑x\phi=\sum_{\ell=1}^{j}c_{\ell}\phi_{\ell},\quad c_{\ell}=\int_{\Omega}\phi\phi_{\ell}\ dx

for any ϕL\phi\in L, which implies

ψ==1jcψ\psi=\sum_{\ell=1}^{j}c_{\ell}\psi_{\ell}

for ψ=ϕTt1\psi=\phi\circ T_{t}^{-1} and ψ=ϕTt1\psi_{\ell}=\phi\circ T_{t}^{-1}. Hence we obtain dimL~tdimL\mbox{dim}\ \tilde{L}_{t}\leq\mbox{dim}\ L for

L~t={ϕTt1ϕL}.\tilde{L}_{t}=\{\phi\circ T_{t}^{-1}\mid\phi\in L\}.

The reverse inequality follows similarly, and hence it holds that

dimL~t=dimL=j.\mbox{dim}\ \tilde{L}_{t}=\mbox{dim}\ L=j.

Since Tt:ΩΩtT_{t}:\Omega\rightarrow\Omega_{t} is a bi-Lipschitz homeomorphism, furthermore, LVL\subset V if and only if L~tVt\tilde{L}_{t}\subset V_{t}. We thus obtain

λ~j(t)=λj(t)\tilde{\lambda}_{j}(t)=\lambda_{j}(t) (39)

by (36)-(37). ∎

We are ready to develop an abstract theory, writing L2L^{2} norm in X=L2(Ω)X=L^{2}(\Omega) as ||X|\ \cdot\ |_{X}. With VV in (4), we recall that V\|\ \cdot\ \|_{V} denotes the norm in VV and that the inclusion VXV\hookrightarrow X is compact. It holds also that

|v|XKvV,vV|v|_{X}\leq K\|v\|_{V},\quad v\in V (40)

for K=1K=1.

Henceforth, CC denotes the generic positive constant. The above At:V×VA_{t}:V\times V\rightarrow\mathbb{R} and Bt:X×XB_{t}:X\times X\rightarrow\mathbb{R} for tIt\in I are symmetric bilinear forms, satisfying

|At(u,v)|CuVvV,At(u,u)δuV2,u,vV|A_{t}(u,v)|\leq C\|u\|_{V}\|v\|_{V},\ A_{t}(u,u)\geq\delta\|u\|_{V}^{2},\quad u,v\in V (41)

and

|Bt(u,v)|C|u|X|v|X,Bt(u,u)δ|u|X2,u,vX|B_{t}(u,v)|\leq C|u|_{X}|v|_{X},\ B_{t}(u,u)\geq\delta|u|_{X}^{2},\quad u,v\in X (42)

for some δ>0\delta>0. Then the eigenvalues of (12) are denoted by

0<λ1(t)λ2(t)+.0<\lambda_{1}(t)\leq\lambda_{2}(t)\leq\cdots\rightarrow+\infty.

The weak and the strong convergences of {uj}Y\{u_{j}\}\subset Y to uYu\in Y for Y=XY=X or Y=VY=V are, furthermore, indicated by

wlimjuj=uin Y\mbox{w}\mathchar 45\relax\lim_{j\rightarrow\infty}u_{j}=u\ \mbox{in $Y$}

and

slimjuj=uin Y,\mbox{s}\mathchar 45\relax\lim_{j\rightarrow\infty}u_{j}=u\ \mbox{in $Y$},

respectively.

5. Continuity of eigenvalues and eigenspaces

Let tIt\in I be fixed. We begin with the following theorem valid under the abstract setting in the previous section.

Theorem 8.

The conditions

limh0supuV,vV1|At+h(u,v)At(u,v)|=0\displaystyle\lim_{h\rightarrow 0}\sup_{\|u\|_{V},\|v\|_{V}\leq 1}|A_{t+h}(u,v)-A_{t}(u,v)|=0
limh0sup|u|X,|v|X1|Bt+h(u,v)Bt(u,v)|=0,\displaystyle\lim_{h\rightarrow 0}\sup_{|u|_{X},|v|_{X}\leq 1}|B_{t+h}(u,v)-B_{t}(u,v)|=0, (43)

imply

limh0λj(t+h)=λj(t)\lim_{h\rightarrow 0}\lambda_{j}(t+h)=\lambda_{j}(t) (44)

for any j=1,2,j=1,2,\cdots.

Proof.

We note that the jj-th eigenvalue of (9) is given by the mini-max principle as in (37), for the Rayleigh quotient Rt[v]R_{t}[v] defined by (38).

Given tt, let

α(h)=supuV,vV1|(At+hAt)(u,v)|\displaystyle\alpha(h)=\sup_{\|u\|_{V},\|v\|_{V}\leq 1}|(A_{t+h}-A_{t})(u,v)|
β(h)=sup|u|X,|v|X1|(Bt+hBt)(u,v)|.\displaystyle\beta(h)=\sup_{|u|_{X},|v|_{X}\leq 1}|(B_{t+h}-B_{t})(u,v)|. (45)

We obtain

(At+hAt)(v,v)α(h)vV2α(h)δAt(v,v)(A_{t+h}-A_{t})(v,v)\geq-\alpha(h)\|v\|_{V}^{2}\geq-\frac{\alpha(h)}{\delta}A_{t}(v,v)

and

(At+hAt)(v,v)α(h)vV2α(h)δAt(v,v)(A_{t+h}-A_{t})(v,v)\leq\alpha(h)\|v\|_{V}^{2}\leq\frac{\alpha(h)}{\delta}A_{t}(v,v)

by (41) and (43), which implies

(1δ1α(h))At(v,v)At+h(v,v)(1+δ1α(h))At(v,v).(1-\delta^{-1}\alpha(h))A_{t}(v,v)\leq A_{t+h}(v,v)\leq(1+\delta^{-1}\alpha(h))A_{t}(v,v).

Similarly, there arises that

(1δ1β(h))Bt(v,v)Bt+h(v,v)(1+δ1β(h))Bt(v,v)(1-\delta^{-1}\beta(h))B_{t}(v,v)\leq B_{t+h}(v,v)\leq(1+\delta^{-1}\beta(h))B_{t}(v,v)

for any vVv\in V.

Then it follows that

(1o(1))Rt[v]Rt+h[v](1+o(1))Rt[v](1-o(1))R_{t}[v]\leq R_{t+h}[v]\leq(1+o(1))R_{t}[v]

uniformly in vV{0}v\in V\setminus\{0\}, and hence

(1o(1))λj(t)λj(t+h)(1+o(1))λj(t)(1-o(1))\lambda_{j}(t)\leq\lambda_{j}(t+h)\leq(1+o(1))\lambda_{j}(t)

by (43). Thus we obtain (44). ∎

Let uj(t)Vu_{j}(t)\in V be the eigenfunction of (12) corresponding to the eigenvalue λj(t)\lambda_{j}(t):

Bt(uj(t),uj(t))=δjj,At(uj(t),v)=λj(t)Bt(uj(t),v),vV.B_{t}(u_{j}(t),u_{j^{\prime}}(t))=\delta_{jj^{\prime}},\quad A_{t}(u_{j}(t),v)=\lambda_{j}(t)B_{t}(u_{j}(t),v),\ \forall v\in V. (46)

Fix tt, assume (19), and define λ\lambda by (21). Although this multiplicity mm is not stable under the perturbation of tt, we obtain the following theorem concerning the continuity of eigenspaces with respect to tt.

Let

Ytλ=uj(t)kjk+m1Y_{t}^{\lambda}=\langle u_{j}(t)\mid k\leq j\leq k+m-1\rangle (47)

be the subspace of XX generated by the above eigenfunctions uj(t)u_{j}(t) for kjk+m1k\leq j\leq k+m-1.

Lemma 9.

Under the above situation, any h0h_{\ell}\rightarrow 0 admits a subsequence, denoted by the same symbol, such that the limits

slimuj(t+h)=ϕjYtλin V,kjk+m1\mbox{s}\mathchar 45\relax\lim_{\ell\rightarrow\infty}u_{j}(t+h_{\ell})=\phi_{j}\in Y_{t}^{\lambda}\ \mbox{in $V$},\quad k\leq j\leq k+m-1 (48)

exist. In particular it holds that

Bt(ϕj,ϕj)=δjj,kj,jk+m1.B_{t}(\phi_{j},\phi_{j^{\prime}})=\delta_{jj^{\prime}},\quad k\leq j,j^{\prime}\leq k+m-1. (49)
Proof.

We have

limh0λj(t+h)=λj(t),kjk+m1\lim_{h\rightarrow 0}\lambda_{j}(t+h)=\lambda_{j}(t),\quad k\leq j\leq k+m-1 (50)

by Theorem 8. It holds also that

uj(t)VC,kjk+m1,|t|<ε0\|u_{j}(t^{\prime})\|_{V}\leq C,\quad k\leq j\leq k+m-1,\ |t^{\prime}|<\varepsilon_{0} (51)

by (40)-(42) and (46):

δuj(t)V2At(uj(t),uj(t))=λj(t)Bt(uj(t),uj(t))=λj(t).\delta\|u_{j}(t^{\prime})\|_{V}^{2}\leq A_{t^{\prime}}(u_{j}(t^{\prime}),u_{j}(t^{\prime}))=\lambda_{j}(t^{\prime})B_{t}(u_{j}(t^{\prime}),u_{j}(t^{\prime}))=\lambda_{j}(t^{\prime}).

Given h0h_{\ell}\rightarrow 0, therefore, we have a subsequence, denoted by the same symbol, which admits the weak limits

wlimuj(t+h)=ϕjin V,kjk+m1,\mbox{w}\mathchar 45\relax\lim_{\ell\rightarrow\infty}u_{j}(t+h_{\ell})=\phi_{j}\ \mbox{in $V$},\quad k\leq j\leq k+m-1, (52)

for some ϕjV\phi_{j}\in V. From (46) it follows that

At(ϕj,v)=λj(t)Bt(ϕj,v),vVA_{t}(\phi_{j},v)=\lambda_{j}(t)B_{t}(\phi_{j},v),\quad\forall v\in V

and hence

ϕjYtλ,kjk+m1.\phi_{j}\in Y_{t}^{\lambda},\quad k\leq j\leq k+m-1.

Since VXV\hookrightarrow X is compact, the weak convergence (52) implies the strong convergence

slimuj(t+h)=ϕjin X,\mbox{s}\mathchar 45\relax\lim_{\ell\rightarrow\infty}u_{j}(t+h_{\ell})=\phi_{j}\ \mbox{in $X$}, (53)

and hence (49). Now we improve this weak convergence (52) to the strong convergence (48) in VV, using (43).

For this purpose, we put

v=uj(t+h)ϕjv=u_{j}(t+h_{\ell})-\phi_{j}

and recall (21) and (50). Then there arises that

δvV2\displaystyle\delta\|v\|_{V}^{2} \displaystyle\leq At+h(uj(t+h)ϕj,v)\displaystyle A_{t+h_{\ell}}(u_{j}(t+h_{\ell})-\phi_{j},v)
=\displaystyle= At+h(uj(t+h),v)(At+hAt)(ϕj,v)At(ϕj,v)\displaystyle A_{t+h_{\ell}}(u_{j}(t+h_{\ell}),v)-(A_{t+h_{\ell}}-A_{t})(\phi_{j},v)-A_{t}(\phi_{j},v)
=\displaystyle= λj(t+h)Bt+h(uj(t+h),v)(At+hAt)(ϕj,v)λj(t)Bt(ϕj,v)\displaystyle\lambda_{j}(t+h_{\ell})B_{t+h_{\ell}}(u_{j}(t+h_{\ell}),v)-(A_{t+h_{\ell}}-A_{t})(\phi_{j},v)-\lambda_{j}(t)B_{t}(\phi_{j},v)
=\displaystyle= (λj(t+h)λj(t))Bt+h(uj(t+h),v)\displaystyle(\lambda_{j}(t+h_{\ell})-\lambda_{j}(t))B_{t+h_{\ell}}(u_{j}(t+h_{\ell}),v)
+λj(t)(Bt+hBt)(uj(t+h),v)+λj(t)Bt(uj(t+h)ϕj,v)\displaystyle+\lambda_{j}(t)(B_{t+h_{\ell}}-B_{t})(u_{j}(t+h_{\ell}),v)+\lambda_{j}(t)B_{t}(u_{j}(t+h_{\ell})-\phi_{j},v)
(At+hAt)(ϕj,v)\displaystyle-(A_{t+h_{\ell}}-A_{t})(\phi_{j},v)
\displaystyle\leq C|λj(t+h)λj(t)||uj(t+h)|XKvV\displaystyle C|\lambda_{j}(t+h_{\ell})-\lambda_{j}(t)|\cdot|u_{j}(t+h_{\ell})|_{X}\cdot K\|v\|_{V}
+λj(t)β(h)|uj(t+h)|XKvV\displaystyle+\lambda_{j}(t)\cdot\beta(h_{\ell})\cdot|u_{j}(t+h_{\ell})|_{X}\cdot K\|v\|_{V}
+λj(t)C|uj(t+h)ϕj|XKvV+α(h)ϕjVvV,\displaystyle+\lambda_{j}(t)\cdot C|u_{j}(t+h_{\ell})-\phi_{j}|_{X}\cdot K\|v\|_{V}+\alpha(h_{\ell})\|\phi_{j}\|_{V}\cdot\|v\|_{V},

and hence

vV\displaystyle\|v\|_{V} \displaystyle\leq C{|λj(t+h)λj(t)|+β(h)+|uj(t+h)ϕj|X+α(h)}\displaystyle C\{|\lambda_{j}(t+h_{\ell})-\lambda_{j}(t)|+\beta(h_{\ell})+|u_{j}(t+h_{\ell})-\phi_{j}|_{X}+\alpha(h_{\ell})\}
=\displaystyle= o(1)\displaystyle o(1)

by (43) and (53). ∎

Remark 5.

If m=1m=1, the eigenfunction uj(t)u_{j}(t) in (58) is uniquely determined by (46) up to the multiplication of ±1\pm 1, which implies ϕj=±uj(t)\phi_{j}=\pm u_{j}(t). In the other case of m2m\geq 2, the eigenfunction which attains (39) does not satisfy this property. Hence we have

ϕj=i=1mqjiui(t)\phi_{j}=\sum_{i=1}^{m}q^{i}_{j}u_{i}(t)

for Q=(qji)Q=(q^{i}_{j}) satisfying QT=Q1Q^{T}=Q^{-1} in Theorem 9. In other words, the eigenfunction uj(t)u_{j}(t) corresponding to λj(t)\lambda_{j}(t) has more varieties than ±1\pm 1 multiplication, although the eigenspace YtλY_{t}^{\lambda} is determined. By this ambiguity the limits ϕj\phi_{j} in (48) depend on the sequence h0h_{\ell}\rightarrow 0, which makes the argument below to be complicated.

6. First derivatives

If {Tt}\{T_{t}\} is 11-differentiable in the setting of Section 1, we can put

A˙t(u,v)=ΩQ˙t[u,v]at+Qt[u,v]a˙tdx,u,vV\dot{A}_{t}(u,v)=\int_{\Omega}\dot{Q}_{t}[\nabla u,\nabla v]a_{t}+Q_{t}[\nabla u,\nabla v]\dot{a}_{t}\ dx,\quad u,v\in V

and

B˙t(u,v)=Ωuva˙t𝑑x,u,vX.\dot{B}_{t}(u,v)=\int_{\Omega}uv\dot{a}_{t}\ dx,\quad u,v\in X.

These A˙t:V×V\dot{A}_{t}:V\times V\rightarrow\mathbb{R} and B˙t:X×X\dot{B}_{t}:X\times X\rightarrow\mathbb{R} in (13) are bilinear forms satisfying

|A˙t(u,v)|CuVvV,u,vV\displaystyle|\dot{A}_{t}(u,v)|\leq C\|u\|_{V}\|v\|_{V},\quad u,v\in V
|B˙t(u,v)|C|u|X|v|X,u,vX\displaystyle|\dot{B}_{t}(u,v)|\leq C|u|_{X}|v|_{X},\quad u,v\in X (54)

and

limh01hsupuV,vV1|(At+hAthA˙t)(u,v)|=0\displaystyle\lim_{h\rightarrow 0}\frac{1}{h}\sup_{\|u\|_{V},\|v\|_{V}\leq 1}\left|\left(A_{t+h}-A_{t}-h\dot{A}_{t}\right)(u,v)\right|=0
limh01hsup|u|X,|v|X1|(Bt+hBthB˙t)(u,v)|=0.\displaystyle\lim_{h\rightarrow 0}\frac{1}{h}\sup_{|u|_{X},|v|_{X}\leq 1}\left|\left(B_{t+h}-B_{t}-h\dot{B}_{t}\right)(u,v)\right|=0. (55)

Hence Theorem 1 in Section 4 is reduced to the following abstract theorem. In this theorem, the assumption made by Theorem 8 is valid, and therefore, there arises that (44).

Theorem 10.

Let X,VX,V be Hilbert spaces over \mathbb{R}, with compact embedding VXV\hookrightarrow X. Let At:V×VA_{t}:V\times V\rightarrow\mathbb{R} and Bt:X×XB_{t}:X\times X\rightarrow\mathbb{R} be symmetric bilinear forms satisfying (41)-(42) for any tIt\in I. Given tt, assume the existence of the bilinear forms A˙t:V×V\dot{A}_{t}:V\times V\rightarrow\mathbb{R} and B˙t:X×X\dot{B}_{t}:X\times X\rightarrow\mathbb{R} such that (54)-(55). Assume, finally, (19) with =k\ell=k and k=mk=m for λj(t)\lambda_{j}(t), kk+m1k\leq k+m-1, defined by (37)-(38). Then the conclusion of Theorem 1 holds.

Before proceeding to the rigorous proof, we develop a formal argument, writing (12) as

utV,Bt(ut,ut)=1,At(ut,v)=λtBt(ut,v),vV.u_{t}\in V,\ B_{t}(u_{t},u_{t})=1,\quad A_{t}(u_{t},v)=\lambda_{t}B_{t}(u_{t},v),\ v\in V. (56)

First, taking a formal differentiation in tt in this equality, we obtain

A˙t(ut,v)+At(u˙t,v)=λ˙tBt(ut,v)+λtB˙t(ut,v)+λtBt(u˙t,v),vV.\dot{A}_{t}(u_{t},v)+A_{t}(\dot{u}_{t},v)=\dot{\lambda}_{t}B_{t}(u_{t},v)+\lambda_{t}\dot{B}_{t}(u_{t},v)+\lambda_{t}B_{t}(\dot{u}_{t},v),\quad\forall v\in V. (57)

Putting v=utv=u_{t}, we obtain

A˙t(ut,ut)+At(u˙t,ut)=λ˙t+λtB˙t(ut,ut)+λtBt(u˙t,ut)\dot{A}_{t}(u_{t},u_{t})+A_{t}(\dot{u}_{t},u_{t})=\dot{\lambda}_{t}+\lambda_{t}\dot{B}_{t}(u_{t},u_{t})+\lambda_{t}B_{t}(\dot{u}_{t},u_{t}) (58)

by

Bt(ut,ut)=1.B_{t}(u_{t},u_{t})=1. (59)

To eliminate u˙t\dot{u}_{t} in (58), second, we use (59) to deduce

B˙t(ut,ut)+2Bt(u˙t,ut)=0.\dot{B}_{t}(u_{t},u_{t})+2B_{t}(\dot{u}_{t},u_{t})=0. (60)

From

λt=λtBt(ut,ut)=At(ut,ut)\lambda_{t}=\lambda_{t}B_{t}(u_{t},u_{t})=A_{t}(u_{t},u_{t})

it is derived also that

λ˙t=A˙t(ut,ut)+2At(u˙t,ut)\dot{\lambda}_{t}=\dot{A}_{t}(u_{t},u_{t})+2A_{t}(\dot{u}_{t},u_{t}) (61)

and then (58) is replaced by

A˙t(ut,ut)+12λ˙t12A˙t(ut,ut)=λ˙t+λtB˙t(ut,ut)12λtB˙t(ut,ut),\dot{A}_{t}(u_{t},u_{t})+\frac{1}{2}\dot{\lambda}_{t}-\frac{1}{2}\dot{A}_{t}(u_{t},u_{t})=\dot{\lambda}_{t}+\lambda_{t}\dot{B}_{t}(u_{t},u_{t})-\frac{1}{2}\lambda_{t}\dot{B}_{t}(u_{t},u_{t}),

or,

λ˙t=A˙t(ut,ut)λtB˙t(ut,ut).\dot{\lambda}_{t}=\dot{A}_{t}(u_{t},u_{t})-\lambda_{t}\dot{B}_{t}(u_{t},u_{t}). (62)

As is noticed in Remark 5 in Section 5, if the eigenspace

Ytλ=uj(t)kjk+m1Y_{t}^{\lambda}=\langle u_{j}(t)\mid k\leq j\leq k+m-1\rangle

corresponding to the eigenvalue λ=λj(t)\lambda=\lambda_{j}(t), kjk+m1k\leq j\leq k+m-1, to (12), is one-dimensional as in m=1m=1, the eigenfunction utu_{t} in (56) is unique up to the multiplication of ±1\pm 1, and this ambiguity is canceled in (62). This property is valid even if m2m\geq 2 as in Remark 1.

Turning to the rigorous proof, we use the following lemma, recalling uj(t)Vu_{j}(t)\in V, YtλY^{\lambda}_{t}, and ϕj\phi_{j} in (46), (47), and (48), respectively.

Lemma 11.

Under the assumption of Theorem 10, any h0h_{\ell}\rightarrow 0 admits a subsequence, denoted by the same symbol, such that

lim1h{λj(t+h)λj(t)}=A˙t(ϕj,ϕj)λj(t)B˙t(ϕj,ϕj)\lim_{\ell\rightarrow\infty}\frac{1}{h_{\ell}}\{\lambda_{j}(t+h_{\ell})-\lambda_{j}(t)\}=\dot{A}_{t}(\phi_{j},\phi_{j})-\lambda_{j}(t)\dot{B}_{t}(\phi_{j},\phi_{j}) (63)

for kjk+m1k\leq j\leq k+m-1. It holds also that

A˙t(ϕj,ϕj)λj(t)B˙t(ϕj,ϕj)=0,kjjk+m1.\dot{A}_{t}(\phi_{j},\phi_{j^{\prime}})-\lambda_{j}(t)\dot{B}_{t}(\phi_{j},\phi_{j^{\prime}})=0,\quad k\leq j\neq j^{\prime}\leq k+m-1. (64)
Proof.

Let kj,jk+m1k\leq j,j^{\prime}\leq k+m-1. Since

At+h(uj(t+h)ϕj,uj(t+h)ϕj)=At+h(uj(t+h),uj(t+h))\displaystyle A_{t+h}(u_{j}(t+h)-\phi_{j},u_{j^{\prime}}(t+h)-\phi_{j^{\prime}})=A_{t+h}(u_{j}(t+h),u_{j^{\prime}}(t+h))
At+h(uj(t+h),ϕj)At+h(ϕj,uj(t+h))+At+h(ϕj,ϕj)\displaystyle\quad-A_{t+h}(u_{j}(t+h),\phi_{j^{\prime}})-A_{t+h}(\phi_{j},u_{j^{\prime}}(t+h))+A_{t+h}(\phi_{j},\phi_{j^{\prime}})

and

At(uj(t+h)ϕj,uj(t+h)ϕj)=At(uj(t+h),uj(t+h))\displaystyle A_{t}(u_{j}(t+h)-\phi_{j},u_{j^{\prime}}(t+h)-\phi_{j^{\prime}})=A_{t}(u_{j}(t+h),u_{j^{\prime}}(t+h))
At(uj(t+h),ϕj)At(ϕj,uj(t+h))+At(ϕj,ϕj),\displaystyle\quad-A_{t}(u_{j}(t+h),\phi_{j^{\prime}})-A_{t}(\phi_{j},u_{j^{\prime}}(t+h))+A_{t}(\phi_{j},\phi_{j^{\prime}}),

it holds that

h(At+hAt)(uj(t+h)ϕjh,uj(t+h)ϕjh)\displaystyle h(A_{t+h}-A_{t})\left(\frac{u_{j}(t+h)-\phi_{j}}{h},\frac{u_{j^{\prime}}(t+h)-\phi_{j^{\prime}}}{h}\right)
=1h(At+hAt)(uj(t+h)ϕj,uj(t+h)ϕj)\displaystyle\quad=\frac{1}{h}(A_{t+h}-A_{t})(u_{j}(t+h)-\phi_{j},u_{j^{\prime}}(t+h)-\phi_{j^{\prime}})
=1h(At+hAt)(uj(t+h),uj(t+h))+1h(At+hAt)(ϕj,ϕj)\displaystyle\quad=\frac{1}{h}(A_{t+h}-A_{t})(u_{j}(t+h),u_{j^{\prime}}(t+h))+\frac{1}{h}(A_{t+h}-A_{t})(\phi_{j},\phi_{j^{\prime}})
1h(At+hAt)(uj(t+h),ϕj)1h(At+hAt)(ϕj,uj(t+h)).\displaystyle\qquad-\frac{1}{h}(A_{t+h}-A_{t})(u_{j}(t+h),\phi_{j^{\prime}})-\frac{1}{h}(A_{t+h}-A_{t})(\phi_{j},u_{j^{\prime}}(t+h)). (65)

In (65), first, we obtain

1h(At+hAt)(uj(t+h),uj(t+h))=A˙t(uj(t+h),uj(t+h))+o(1)\frac{1}{h}(A_{t+h}-A_{t})(u_{j}(t+h),u_{j^{\prime}}(t+h))=\dot{A}_{t}(u_{j}(t+h),u_{j^{\prime}}(t+h))+o(1)

as h0h\rightarrow 0 by (51) and (55). Then (48) implies

lim1h(At+hAt)(uj(t+h),uj(t+h))\displaystyle\lim_{\ell\rightarrow\infty}\frac{1}{h_{\ell}}(A_{t+h_{\ell}}-A_{t})(u_{j}(t+h_{\ell}),u_{j^{\prime}}(t+h_{\ell}))
=limA˙t(uj(t+h),uj(t+h))=A˙t(ϕj,ϕj)\displaystyle\quad=\lim_{\ell\rightarrow\infty}\dot{A}_{t}(u_{j}(t+h_{\ell}),u_{j^{\prime}}(t+h_{\ell}))=\dot{A}_{t}(\phi_{j},\phi_{j^{\prime}}) (66)

and also

lim1h(At+hAt)(ϕj,ϕj)=A˙t(ϕj,ϕj).\lim_{\ell\rightarrow\infty}\frac{1}{h_{\ell}}(A_{t+h_{\ell}}-A_{t})(\phi_{j},\phi_{j^{\prime}})=\dot{A}_{t}(\phi_{j},\phi_{j^{\prime}}). (67)

Second, it holds that

At+h(uj(t+h),ϕj)=λj(t+h)Bt+h(uj(t+h),ϕj)\displaystyle A_{t+h}(u_{j}(t+h),\phi_{j^{\prime}})=\lambda_{j}(t+h)B_{t+h}(u_{j}(t+h),\phi_{j^{\prime}})
At(uj(t+h),ϕj)=λj(t)Bt(uj(t+h),ϕj)\displaystyle A_{t}(u_{j}(t+h),\phi_{j^{\prime}})=\lambda_{j}(t)B_{t}(u_{j}(t+h),\phi_{j^{\prime}})

by ϕjYtλ\phi_{j^{\prime}}\in{Y_{t}^{\lambda}}, which implies

1h(At+hAt)(uj(t+h),ϕj)\displaystyle\frac{1}{h_{\ell}}(A_{t+h_{\ell}}-A_{t})(u_{j}(t+h_{\ell}),\phi_{j^{\prime}})
=1h(λj(t+h)Bt+h(uj(t+h),ϕj)λj(t)Bt(uj(t+h),ϕj))\displaystyle\quad=\frac{1}{h_{\ell}}(\lambda_{j}(t+h_{\ell})B_{t+h_{\ell}}(u_{j}(t+h_{\ell}),\phi_{j^{\prime}})-\lambda_{j}(t)B_{t}(u_{j}(t+h_{\ell}),\phi_{j^{\prime}}))
=1h(λj(t+h)λj(t))Bt+h(uj(t+h),ϕj)\displaystyle\quad=\frac{1}{h_{\ell}}(\lambda_{j}(t+h_{\ell})-\lambda_{j}(t))B_{t+h_{\ell}}(u_{j}(t+h_{\ell}),\phi_{j^{\prime}})
+1hλj(t)(Bt+hBt)(uj(t+h),ϕj)\displaystyle\qquad+\frac{1}{h_{\ell}}\lambda_{j}(t)(B_{t+h_{\ell}}-B_{t})(u_{j}(t+h_{\ell}),\phi_{j^{\prime}})
=1h(λj(t+h)λj(t))(δjj+o(1))+λj(t)B˙t(ϕj,ϕj)+o(1)\displaystyle\quad=\frac{1}{h_{\ell}}(\lambda_{j}(t+h_{\ell})-\lambda_{j}(t))(\delta_{jj^{\prime}}+o(1))+\lambda_{j}(t)\dot{B}_{t}(\phi_{j},\phi_{j^{\prime}})+o(1) (68)

by

Bt+h(uj(t+h),ϕj)\displaystyle B_{t+h_{\ell}}(u_{j}(t+h_{\ell}),\phi_{j^{\prime}}) =\displaystyle= Bt(uj(t+h),ϕj)+o(1)\displaystyle B_{t}(u_{j}(t+h_{\ell}),\phi_{j^{\prime}})+o(1)
=\displaystyle= Bt(ϕj,ϕj)+o(1)=δjj+o(1)\displaystyle B_{t}(\phi_{j},\phi_{j^{\prime}})+o(1)=\delta_{jj^{\prime}}+o(1)

and (55). Similarly, it follows that

1h(At+hAt)(ϕj,uj(t+h))\displaystyle\frac{1}{h_{\ell}}(A_{t+h_{\ell}}-A_{t})(\phi_{j},u_{j^{\prime}}(t+h_{\ell}))
=1h(λj(t+h)λj(t))(δjj+o(1))+λj(t)B˙t(ϕj,ϕj)+o(1).\displaystyle\quad=\frac{1}{h_{\ell}}(\lambda_{j}(t+h_{\ell})-\lambda_{j}(t))(\delta_{jj^{\prime}}+o(1))+\lambda_{j}(t)\dot{B}_{t}(\phi_{j},\phi_{j^{\prime}})+o(1). (69)

Finally, we obtain

|h(At+hAt)(uj(t+h)ϕjh,uj(t+h)ϕjh)|\displaystyle\left|h_{\ell}(A_{t+h_{\ell}}-A_{t})\left(\frac{u_{j}(t+h_{\ell})-\phi_{j}}{h_{\ell}},\frac{u_{j^{\prime}}(t+h_{\ell})-\phi_{j^{\prime}}}{h_{\ell}}\right)\right|
Ch2uj(t+h)ϕjhVuj(t+h)ϕjhV\displaystyle\quad\leq Ch_{\ell}^{2}\left\|\frac{u_{j}(t+h_{\ell})-\phi_{j}}{h_{\ell}}\right\|_{V}\cdot\left\|\frac{u_{j^{\prime}}(t+h_{\ell})-\phi_{j^{\prime}}}{h_{\ell}}\right\|_{V}
=Cuj(t+h)ϕjVuj(t+h)ϕjV=o(1)\displaystyle\quad=C\|u_{j}(t+h_{\ell})-\phi_{j}\|_{V}\cdot\|u_{j^{\prime}}(t+h_{\ell})-\phi_{j^{\prime}}\|_{V}=o(1) (70)

by (48) and (55). Then equalities (63)-(64) follow from (66)-(70) as

0\displaystyle 0 =\displaystyle= 2A˙t(ϕj,ϕj)2h(λj(t+h)λj(t))(δjj+o(1))\displaystyle 2\dot{A}_{t}(\phi_{j},\phi_{j^{\prime}})-\frac{2}{h_{\ell}}(\lambda_{j}(t+h_{\ell})-\lambda_{j}(t))(\delta_{jj^{\prime}}+o(1))
2λj(t)B˙t(ϕj,ϕj)+o(1),.\displaystyle-2\lambda_{j}(t)\dot{B}_{t}(\phi_{j},\phi_{j^{\prime}})+o(1),\quad\ell\rightarrow\infty.

. ∎

Below we confirm that the process of taking subsequence in the previous lemma is not necessary, if h0h_{\ell}\rightarrow 0 is unilateral as in h+0h_{\ell}\rightarrow+0 or h0h_{\ell}\rightarrow-0. Theorem 1 is thus reduced to the following theorem.

Theorem 12.

Under the assumption of Theorem 10, the unilateral limits

λ˙j±(t)=limh±01h{λj(t+h)λj(t)}\dot{\lambda}_{j}^{\pm}(t)=\lim_{h\rightarrow\pm 0}\frac{1}{h}\{\lambda_{j}(t+h)-\lambda_{j}(t)\} (71)

exist, and it holds that

λ˙j+(t)=μjk+1,λ˙j(t)=μk+mj,kjk+m1.\dot{\lambda}^{+}_{j}(t)=\mu_{j-k+1},\ \dot{\lambda}_{j}^{-}(t)=\mu_{k+m-j},\quad k\leq j\leq k+m-1. (72)

Here, μq\mu_{q}, 1qm1\leq q\leq m, is the qq-th eigenvalue of

uYtλ,Etλ(u,v)=μBt(u,v),vYtλ,u\in Y_{t}^{\lambda},\quad E^{\lambda}_{t}(u,v)=\mu B_{t}(u,v),\ \forall v\in Y_{t}^{\lambda}, (73)

where YtλY_{t}^{\lambda} is the mm-dimensional eigenspace of (12) corresponding to the eigenvalue λ\lambda of (21) defined by (47), and

Etλ=A˙tλB˙t.E^{\lambda}_{t}=\dot{A}_{t}-\lambda\dot{B}_{t}. (74)

In particular, it holds that

λ˙j+(t)=λ˙2k+m1j(t),kjk+m1.\dot{\lambda}_{j}^{+}(t)=\dot{\lambda}_{2k+m-1-j}^{-}(t),\quad k\leq j\leq k+m-1.
Proof.

Since YtλY_{t}^{\lambda} is mm-dimensional, the eigenvalue problem (73) admits mm-eigenvalues denoted by

μ1μm.\mu_{1}\leq\cdots\leq\mu_{m}.

By Lemma 11, on the other hand, any h0h_{\ell}\rightarrow 0 takes a subsequence, denoted by the same symbol, satisfying (63)-(64) for some

ϕjYtλ,kjk+m1,\phi_{j}\in Y_{t}^{\lambda},\quad k\leq j\leq k+m-1,

with (49).

This lemma ensures also the existence of

μ~j=lim1h(λj(t+h)λj(t)),\tilde{\mu}_{j}=\lim_{\ell\rightarrow\infty}\frac{1}{h_{\ell}}(\lambda_{j}(t+h_{\ell})-\lambda_{j}(t)), (75)

and the equalities

Etλ(ϕj,ϕj)=δjjμ~j,kj,jk+m1.E^{\lambda}_{t}(\phi_{j},\phi_{j^{\prime}})=\delta_{jj^{\prime}}\tilde{\mu}_{j},\quad k\leq j,j^{\prime}\leq k+m-1. (76)

We thus obtain

ϕjYtλ,Bt(ϕj,ϕj)=1,Etλ(ϕj,v)=μ~jBt(ϕj,v),vYtλ,\phi_{j}\in{Y_{t}^{\lambda}},\quad B_{t}(\phi_{j},\phi_{j})=1,\quad E^{\lambda}_{t}(\phi_{j},v)=\tilde{\mu}_{j}B_{t}(\phi_{j},v),\ \forall v\in Y_{t}^{\lambda},

and therefore, μ=μ~j\mu=\tilde{\mu}_{j} is an eigenvalue of (73).

If h+0h_{\ell}\rightarrow+0, there arises that

μ~kμ~k+m1\tilde{\mu}_{k}\leq\cdots\leq\tilde{\mu}_{k+m-1}

by

λk(t+h)λk+m1(t+h),\lambda_{k}(t+h)\leq\cdots\leq\lambda_{k+m-1}(t+h),

and hence

μ~j=μjk+1,kjk+m1.\tilde{\mu}_{j}=\mu_{j-k+1},\quad k\leq j\leq k+m-1.

Then we obtain the result because the value μ~j\tilde{\mu}_{j} in (75) is independent of the sequence h+0h_{\ell}\rightarrow+0.

In the other case of h0h_{\ell}\rightarrow-0, we obtain

μ~j=μk+mj,kjk+m1,\tilde{\mu}_{j}=\mu_{k+m-j},\quad k\leq j\leq k+m-1,

and the result follows similarly. ∎

Theorem 2 is reduced to the following abstract theorem.

Theorem 13.

Let the assumption of Theorem 10 hold in II. Fix tIt\in I, and assume

limh0supuV,vV1|A˙t+h(u,v)A˙t(u,v)|=0\displaystyle\lim_{h\rightarrow 0}\sup_{\|u\|_{V},\|v\|_{V}\leq 1}\left|\dot{A}_{t+h}(u,v)-\dot{A}_{t}(u,v)\right|=0
limh0sup|u|X,|v|X1|B˙t+h(u,v)B˙t(u,v)|=0.\displaystyle\lim_{h\rightarrow 0}\sup_{|u|_{X},|v|_{X}\leq 1}\left|\dot{B}_{t+h}(u,v)-\dot{B}_{t}(u,v)\right|=0. (77)

Then, it follows that

limh±0λ˙j±(t+h)=λ˙j±(t).\lim_{h\rightarrow\pm 0}\dot{\lambda}_{j}^{\pm}(t+h)=\dot{\lambda}_{j}^{\pm}(t).
Proof.

Assume (19)-(21) and take kjk+m1k\leq j\leq k+m-1. Since the assumption of Theorem 10 holds in II, any tIt^{\prime}\in I admits uj(t)Vu_{j}(t^{\prime})\in V such that

Bt(uj(t),uj(t))=1,At(uj(t),uj(t))=λj(t)Bt(uj(t),uj(t))B_{t^{\prime}}(u_{j}(t^{\prime}),u_{j}(t^{\prime}))=1,\ A_{t^{\prime}}(u_{j}(t^{\prime}),u_{j}(t^{\prime}))=\lambda_{j}(t^{\prime})B_{t^{\prime}}(u_{j}(t^{\prime}),u_{j}(t^{\prime})) (78)

and

λ˙j+(t)=A˙t(uj(t),uj(t))λj(t)B˙t(uj(t),uj(t)).\dot{\lambda}_{j}^{+}(t^{\prime})=\dot{A}_{t^{\prime}}(u_{j}(t^{\prime}),u_{j}(t^{\prime}))-\lambda_{j}(t^{\prime})\dot{B}_{t^{\prime}}(u_{j}(t^{\prime}),u_{j}(t^{\prime})). (79)

by Lemma 11 and Theorem 12.

Given tt in this theorem, and take h+0h_{\ell}\rightarrow+0 and uj(t)u_{j}(t^{\prime}) in (78) for t=t+ht^{\prime}=t+h_{\ell}. Hence there is a subsequence denoted by the same symbol such that (48) with ϕjV\phi_{j}\in V. Then it holds that

λ˙j+(t)=A˙t(ϕj,ϕj)λj(t)B˙t(ϕj,ϕj).\dot{\lambda}_{j}^{+}(t)=\dot{A}_{t}(\phi_{j},\phi_{j})-\lambda_{j}(t)\dot{B}_{t}(\phi_{j},\phi_{j}).

We thus obtain

limλ˙j+(t+h)\displaystyle\lim_{\ell\rightarrow\infty}\dot{\lambda}_{j}^{+}(t+h_{\ell}) =\displaystyle= lim{A˙t+h(uj(t+h),uj(t+h))\displaystyle\lim_{\ell\rightarrow\infty}\{\dot{A}_{t+h_{\ell}}(u_{j}(t+h_{\ell}),u_{j}(t+h_{\ell}))
λj(t+h)B˙t+h(uj(t+h),uj(t+h))}\displaystyle-\lambda_{j}(t+h_{\ell})\dot{B}_{t+h_{\ell}}(u_{j}(t+h_{\ell}),u_{j}(t+h_{\ell}))\}
=\displaystyle= A˙t(ϕj,ϕj)λj(t)B˙t(ϕj,ϕj)=λ˙j+(t)\displaystyle\dot{A}_{t}(\phi_{j},\phi_{j})-\lambda_{j}(t)\dot{B}_{t}(\phi_{j},\phi_{j})=\dot{\lambda}_{j}^{+}(t)

by (79), and hence

limh+0λ˙j+(t+h)=λ˙j+(t)\lim_{h\rightarrow+0}\dot{\lambda}_{j}^{+}(t+h)=\dot{\lambda}_{j}^{+}(t)

because h+0h_{\ell}\rightarrow+0 is arbitrary.

The proof of

limh0λ˙j(t+h)=λ˙j(t)\lim_{h\rightarrow-0}\dot{\lambda}_{j}^{-}(t+h)=\dot{\lambda}_{j}^{-}(t)

is similar. ∎

Remark 6.

The limits (75) exist for any jj under the conditions (41), (42), (54), and (55). If these conditions are satisfied for any t(ε0,ε0)t\in(-\varepsilon_{0},\varepsilon_{0}), the limits (71) are unilaterally locally uniform in tIt\in I. In fact, if not, there are, for example, tkt0t_{k}\downarrow t_{0}\in (ε0,ε0)(-\varepsilon_{0},\varepsilon_{0}), δ>0\delta>0, and h0h_{\ell}\rightarrow 0, such that

|1h(λj(tk+h)λj(tk))λ˙j+(tk)|δ.\left|\frac{1}{h_{\ell}}(\lambda_{j}(t_{k}+h_{\ell})-\lambda_{j}(t_{k}))-\dot{\lambda}_{j}^{+}(t_{k})\right|\geq\delta.

Then we obtain

|1h(λj(t0+h)λj(t0))λ˙j+(t0)|δ,\left|\frac{1}{h_{\ell}}(\lambda_{j}(t_{0}+h_{\ell})-\lambda_{j}(t_{0}))-\dot{\lambda}_{j}^{+}(t_{0})\right|\geq\delta,

a contradiction with \ell\rightarrow\infty.

7. Rearrangement of eigenvalues

For simplicity we introduce the following notations to prove Theorem 3. Recall I=(ε0,ε0)I=(-\varepsilon_{0},\varepsilon_{0}), and let fjC0(I)f_{j}\in C^{0}(I), 1jm1\leq j\leq m, satisfy

f1(t)fm(t),tI.f_{1}(t)\leq\cdots\leq f_{m}(t),\quad t\in I. (80)

Assume the existence of the unilateral limits

f˙j±(t)=limh±01h(fj(t+h)fj(t))\dot{f}_{j}^{\pm}(t)=\lim_{h\rightarrow\pm 0}\frac{1}{h}(f_{j}(t+h)-f_{j}(t)) (81)

and

limh±0f˙j(t+h)=f˙j±(t)\lim_{h\rightarrow\pm 0}\dot{f}_{j}(t+h)=\dot{f}_{j}^{\pm}(t) (82)

for any jj and tt. Assume, finally,

f˙j+(t)=f˙2k+nj1(t),kjk+n1,\dot{f}_{j}^{+}(t)=\dot{f}_{2k+n-j-1}^{-}(t),\quad k\leq j\leq k+n-1, (83)

provided that

fk1(t)<fk(t)==fk+n1(t)<fk+n(t),f_{k-1}(t)<f_{k}(t)=\cdots=f_{k+n-1}(t)<f_{k+n}(t), (84)

where 1nm1\leq n\leq m, 1kmn+11\leq k\leq m-n+1, and tIt\in I. In (84) we understand

f0(t)=,fm+1(t)=+.f_{0}(t)=-\infty,\quad f_{m+1}(t)=+\infty.

We call

K=Kk,n(t)={k,,k+n1}K=K_{k,n}(t)=\{k,\cdots,k+n-1\}

the nn-cluster at tt with entry kk if (84) arises, and also

p(K)=max{jkjk+[n2]1,f˙j+(t)<f˙2k+n1j+(t)}k+1p(K)=\max\{j\mid k\leq j\leq k+[\frac{n}{2}]-1,\ \dot{f}_{j}^{+}(t)<\dot{f}_{2k+n-1-j}^{+}(t)\}-k+1

its pp-value. Here, we understand p(K)=0p(K)=0 if

f˙k+(t)=f˙k+n1+(t),\dot{f}_{k}^{+}(t)=\dot{f}_{k+n-1}^{+}(t),

noting

kijk+n1f˙i+(t)f˙j+(t).k\leq i\leq j\leq k+n-1\quad\Rightarrow\quad\dot{f}_{i}^{+}(t)\leq\dot{f}_{j}^{+}(t).

We construct a rearrangement of C0C^{0}-curves

Cj={fj(t)tI},1jm,C_{j}=\{f_{j}(t)\mid t\in I\},\quad 1\leq j\leq m,

denoted by C~j\tilde{C}_{j}, 1jm1\leq j\leq m, so that are C1C^{1} in tIt\in I. This rearrangement is done only on

I1={tIthere exists a cluster K at t such that p(K)1.}.I_{1}=\{t\in I\mid\mbox{there exists a cluster $K$ at $t$ such that $p(K)\geq 1$}.\}.

To introduce this rearrangement, we note the following facts in advance. First, given 2nm2\leq n\leq m, let

I1n={tIthere exists an n-cluster K at t such that p(K)1}.I_{1}^{n}=\{t\in I\mid\mbox{there exists an $n$-cluster $K$ at $t$ such that $p(K)\geq 1$}\}.

If tI1nt\in I_{1}^{n} and K=Kk,n(t)K=K_{k,n}(t) satisfies p(K)1p(K)\geq 1, it holds that

fk+n1(t)>fk(t),0<|tt|1.f_{k+n-1}(t^{\prime})>f_{k}(t^{\prime}),\quad 0<|t^{\prime}-t|\ll 1. (85)

Hence this tt is an isolated point of I1nI_{1}^{n}. In particular, each I1nI_{1}^{n}, 2nm2\leq n\leq m, is at most countable, and hence so is I1I_{1} by

I1=n=2mI1n.I_{1}=\bigcup_{n=2}^{m}I_{1}^{n}.

Second, given tI1t\in I_{1} and 1jm1\leq j\leq m, if jK=Kk,n(t)j\in K=K_{k,n}(t) holds for some KK in p(K)1p(K)\geq 1, this KK is unique.

Definition 3.

The curves C~j\tilde{C}_{j}, 1jm1\leq j\leq m, are called the transversal rearrangement of CjC_{j}, 1jm1\leq j\leq m, if the following operations are done.

  1. (1)

    If tI1t\in I_{1}, 1jm1\leq j\leq m, and jKj\in K hold for K=Kk,n(t)K=K_{k,n}(t) with p(K)1p(K)\geq 1, the curve CjC_{j} for kjk+p1k\leq j\leq k+p-1 and knpjkn1k-n-p\leq j\leq k-n-1 on the right, is connected to C2k+nj1C_{2k+n-j-1} on the left at tt, where p=p(K)p=p(K).

  2. (2)

    No rearrangements to CjC_{j}, 1jm1\leq j\leq m, are done otherwise.

The curves C~j\tilde{C}_{j}, 1jm1\leq j\leq m, are uniquely constructed from CjC_{j}, 1jm1\leq j\leq m, by this transversal rearrangement. From the results in the previous section, Theorem 3 is a consequence of the following theorem.

Theorem 14.

Under the above situation, the C0C^{0} curves C~j\tilde{C}_{j}, 1jm1\leq j\leq m, made by the transversal rearrangement of CjC_{j}, 1jm1\leq j\leq m, are C1C^{1}.

Proof.

This theorem is obvious if m=1m=1. Now we show it by an induction on mm, assuming the assertion up to m1m-1.

Take t0II1¯t_{0}\in I\setminus\overline{I_{1}}, and make the transversal rearrangement of CjC_{j}, 1jq1\leq j\leq q, toward left and right diections. Let tt_{\ell}, =1,2,\ell=1,2,\cdots, be the successive points of I1I_{1} in the left direction:

tI1,t1>t,(t,t1)I1=,=1,2,.t_{\ell}\in I_{1},\ t_{\ell-1}>t_{\ell},\ (t_{\ell},t_{\ell-1})\cap I_{1}=\emptyset,\quad\ell=1,2,\cdots.

If {t}\{t_{\ell}\} is finite, these CjC_{j}’s are successfully rearranged to C1C^{1} curves on (ε0,t0)(-\varepsilon_{0},t_{0}). If not, there is

t=limt[ε0,t0].t_{\ast}=\lim_{\ell\rightarrow\infty}t_{\ell}\in[-\varepsilon_{0},t_{0}].

Then the case t=ε0t_{\ast}=-\varepsilon_{0} ensures the same conclusion.

Letting t>ε0t_{\ast}>-\varepsilon_{0}, we show that {C~j1jm}\{\tilde{C}_{j}\mid 1\leq j\leq m\} are C1C^{1} curves on (tδ,t+δ)(t_{\ast}-\delta,t_{\ast}+\delta) for 0<δ10<\delta\ll 1. Once this fact is proven, we can repeat this process up to t=ε0t=-\varepsilon_{0} by a covering argument. Turning to the right direction, we conclude that C~j\tilde{C}_{j}, 1jm1\leq j\leq m, are C1C^{1} curves on I=(ε0,ε0)I=(-\varepsilon_{0},\varepsilon_{0}).

To this end we distinguish the cases tI1t_{\ast}\in I_{1} and tI1t_{\ast}\not\in I_{1}.

If tI1t_{\ast}\in I_{1}, first, the above assertion follows from the assumption of induction. In fact, each K=Kk,n(t)K=K_{k,n}(t_{\ast}) with p(K)1p(K)\geq 1 admits (85), while

f~j(t)={fj(t),0<tt1f2k+nj1(t),0<tt1,j=k,j=k+n1\tilde{f}_{j}(t)=\left\{\begin{array}[]{ll}f_{j}(t),&0<t-t_{\ast}\ll 1\\ f_{2k+n-j-1}(t),&0<t_{\ast}-t\ll 1\end{array}\right.,\quad j=k,\ j=k+n-1

are C1C^{1} around t=tt=t_{\ast}. Then we apply the assumption of induction to CjC_{j}, k+1jk+n2k+1\leq j\leq k+n-2, to get nn-C1C^{1} curves in (tδ,t+δ)(t_{\ast}-\delta,t_{\ast}+\delta) made by CjC_{j}, kjk+n1k\leq j\leq k+n-1. Operating this process to any K=Kk,n(t)K=K_{k,n}(t_{\ast}) with p(K)1p(K)\geq 1 at t=tt=t_{\ast}, we get C1C^{1} curves C~j\tilde{C}_{j}, 1jm1\leq j\leq m, in (tδ,t+δ)(t_{\ast}-\delta,t_{\ast}+\delta) by this transversal rearrangement of CjC_{j}, 1jm1\leq j\leq m, at t=tt=t_{\ast}.

If tI1t_{\ast}\not\in I_{1}, second, we take the cluster decomposition of {1,,m}\{1,\cdots,m\} at t=tt=t_{\ast}, that is,

1=k1<k1+n1=k2<<ks1+ns1=ks<ks+ns=m1=k_{1}<k_{1}+n_{1}=k_{2}<\cdots<k_{s-1}+n_{s-1}=k_{s}<k_{s}+n_{s}=m

satisfying

r=1sKkr,nr(t)={1,,m}.\bigcup_{r=1}^{s}K_{k_{r},n_{r}}(t_{\ast})=\{1,\cdots,m\}.

Since

p(Kkr,nr(t))=0,1rsp(K_{k_{r},n_{r}}(t_{\ast}))=0,\quad 1\leq r\leq s

holds by the assumption, there are ara_{r}, 1rs1\leq r\leq s, such that

f˙j+(t)=ar,jKkr,nr.\dot{f}_{j}^{+}(t_{\ast})=a_{r},\quad\forall j\in K_{k_{r},n_{r}}.

We obtain, on the other hand,

fkr1(t)<fkr(t),|tt|1, 1rs+1f_{k_{r-1}}(t)<f_{k_{r}}(t),\ |t-t_{\ast}|\ll 1,\ 1\leq r\leq s+1

under the agreement

fk0(t)=,fks+1(t)=+,f_{k_{0}}(t)=-\infty,\quad f_{k_{s+1}}(t)=+\infty,

and hence C~j\tilde{C}_{j}, 1jm1\leq j\leq m, are C1C^{1} on [t,t+δ)[t_{\ast},t_{\ast}+\delta) for 0<δ10<\delta\ll 1. If tt_{\ast} is not a right accumulating point of I1I_{1}, therefore, these C~j\tilde{C}_{j}, 1jm1\leq j\leq m, made by the transversal rearrangement of CjC_{j}, 1jm1\leq j\leq m, are C1C^{1} on (tδ,t+δ)(t_{\ast}-\delta,t_{\ast}+\delta).

In the other case that tt_{\ast} is a right accumulating point of I1I_{1}, these C~j\tilde{C}_{j}, 1jm1\leq j\leq m, are C1C^{1} on (tδ,t](t_{\ast}-\delta,t_{\ast}], similarly. Then it holds that

f˙j(t)=ar,jKkr,nr=Kkr,nr(t)\dot{f}_{j}^{-}(t_{\ast})=a_{r},\quad\forall j\in K_{k_{r},n_{r}}=K_{k_{r},n_{r}}(t_{\ast})

by p(Kkr,nr)=0p(K_{k_{r},n_{r}})=0, and hence C~j\tilde{C}_{j}, 1jm1\leq j\leq m, are C1C^{1} on (tδ,t+δ)(t_{\ast}-\delta,t_{\ast}+\delta). ∎

8. Second derivatives

If Tt:ΩΩtT_{t}:\Omega\rightarrow\Omega_{t} is 22-differentiable, we have the other bilinear forms A¨t:V×V\ddot{A}_{t}:V\times V\rightarrow\mathbb{R} and B¨t:X×X\ddot{B}_{t}:X\times X\rightarrow\mathbb{R} satisfying

|A¨t(u,v)|CuVvV,u,vV\displaystyle|\ddot{A}_{t}(u,v)|\leq C\|u\|_{V}\|v\|_{V},\quad u,v\in V
|B¨t(u,v)|C|u|X|v|X,u,vX\displaystyle|\ddot{B}_{t}(u,v)|\leq C|u|_{X}|v|_{X},\quad\ u,v\in X (86)

uniformly in tt and

limh01h2supuV,vV1|(At+hAthA˙th22A¨t)(u,v)|=0\displaystyle\lim_{h\rightarrow 0}\frac{1}{h^{2}}\sup_{\|u\|_{V},\|v\|_{V}\leq 1}\left|\left(A_{t+h}-A_{t}-h\dot{A}_{t}-\frac{h^{2}}{2}\ddot{A}_{t}\right)(u,v)\right|=0
limh01h2sup|u|X,|v|X1|(Bt+hBthB˙th22B¨t)(u,v)|=0\displaystyle\lim_{h\rightarrow 0}\frac{1}{h^{2}}\sup_{|u|_{X},|v|_{X}\leq 1}\left|\left(B_{t+h}-B_{t}-h\dot{B}_{t}-\frac{h^{2}}{2}\ddot{B}_{t}\right)(u,v)\right|=0 (87)

for each tt. Hence Theorem 4 is reduced to the following abstract theorem.

Theorem 15.

In Theorem 10, assume, futhermore, (86)-(87). Then the conclusion of Theorem 4 holds.

For the moment we develop a formal argument as in the first derivative. Assuming (62), first, we deduce

λ¨t\displaystyle\ddot{\lambda}_{t} =\displaystyle= A¨t(ut,ut)+2A˙t(u˙t,ut)λ˙tB˙t(ut,ut)λtB¨t(ut,ut)2λtB˙t(u˙t,ut)\displaystyle\ddot{A}_{t}(u_{t},u_{t})+2\dot{A}_{t}(\dot{u}_{t},u_{t})-\dot{\lambda}_{t}\dot{B}_{t}(u_{t},u_{t})-\lambda_{t}\ddot{B}_{t}(u_{t},u_{t})-2\lambda_{t}\dot{B}_{t}(\dot{u}_{t},u_{t}) (88)
=\displaystyle= 2(A˙tλtB˙t)(u˙t,ut)+Dt(ut,ut)\displaystyle 2(\dot{A}_{t}-\lambda_{t}\dot{B}_{t})(\dot{u}_{t},u_{t})+D_{t}(u_{t},u_{t})

for

Dt(u,v)=A¨t(u,v)λ˙tB˙t(u,v)λtB¨t(u,v),u,vV.D_{t}(u,v)=\ddot{A}_{t}(u,v)-\dot{\lambda}_{t}\dot{B}_{t}(u,v)-\lambda_{t}\ddot{B}_{t}(u,v),\quad u,v\in V. (89)

Putting v=u˙tv=\dot{u}_{t} in (57), second, we reach

(A˙tλtB˙t)(ut,u˙t)=(AtλtBt)(u˙t,u˙t)+λ˙tBt(ut,u˙t).(\dot{A}_{t}-\lambda_{t}\dot{B}_{t})(u_{t},\dot{u}_{t})=-(A_{t}-\lambda_{t}B_{t})(\dot{u}_{t},\dot{u}_{t})+\dot{\lambda}_{t}B_{t}(u_{t},\dot{u}_{t}). (90)

Then, (60), (88), and (90) imply

λ¨t=2(AtλtBt)(z,z)+2λ˙tBt(ut,z)+Dt(ut,ut)\displaystyle\ddot{\lambda}_{t}=-2(A_{t}-\lambda_{t}B_{t})(z_{\ast},z_{\ast})+2\dot{\lambda}_{t}B_{t}(u_{t},z_{\ast})+D_{t}(u_{t},u_{t})
=2(AtλtBt)(z,z)λ˙tB˙t(ut,ut)+Dt(ut,ut)\displaystyle=-2(A_{t}-\lambda_{t}B_{t})(z_{\ast},z_{\ast})-\dot{\lambda}_{t}\dot{B}_{t}(u_{t},u_{t})+D_{t}(u_{t},u_{t})
=2(AtλtBt)(z,z)+A¨t(ut,ut)2λ˙tB˙t(ut,ut)λtB¨t(ut,ut)\displaystyle=-2(A_{t}-\lambda_{t}B_{t})(z_{\ast},z_{\ast})+\ddot{A}_{t}(u_{t},u_{t})-2\dot{\lambda}_{t}\dot{B}_{t}(u_{t},u_{t})-\lambda_{t}\ddot{B}_{t}(u_{t},u_{t}) (91)

for z=u˙tz_{\ast}=\dot{u}_{t}.

We observe that u˙tV\dot{u}_{t}\in V is not uniquely determined by (57), which is derive formally also. It has, more precisely, the ambiguity of addition of an element in YtλY_{t}^{\lambda}. This ambiguity, however, cancels in (91) by equality (92) below.

We now develop a rigorous argument valid even to (19). Define λ\lambda by (21) and recall Ytλ=uj(t)kjk+m1Y_{t}^{\lambda}=\langle u_{j}(t)\mid k\leq j\leq k+m-1\rangle for uj(t)u_{j}(t) satisfying (46). Let, also,

Ctj=Atλj(t)Bt,kjk+m1,tI.C_{t^{\prime}}^{j}=A_{t^{\prime}}-\lambda_{j}(t^{\prime})B_{t^{\prime}},\quad k\leq j\leq k+m-1,\ t^{\prime}\in I.

Then we obtain

Ctj=AtλBtCtC_{t}^{j}=A_{t}-\lambda B_{t}\equiv C_{t}

and

Ct(u,v)=0,(u,v)Ytλ×V.C_{t}(u,v)=0,\quad\forall(u,v)\in Y_{t}^{\lambda}\times V. (92)

By Lemma 11, given h0h_{\ell}\rightarrow 0, which may not be of definite sign, we have a subsequence, denoted by the same symbol, satisfying (48),

slimuj(t+h)=ϕjYtλin V,kjk+m1.\mbox{s}\mathchar 45\relax\lim_{\ell\rightarrow\infty}u_{j}(t+h_{\ell})=\phi_{j}\in Y_{t}^{\lambda}\ \mbox{in $V$},\quad k\leq j\leq k+m-1. (93)

There exists

λ˙j=lim1h(λj(t+h)λj(t)),kjk+m1\dot{\lambda}_{j}^{\ast}=\lim_{\ell\rightarrow\infty}\frac{1}{h_{\ell}}(\lambda_{j}(t+h_{\ell})-\lambda_{j}(t)),\quad k\leq j\leq k+m-1 (94)

passing to a subsequence, with the equality

λ˙jδjj=A˙t(ϕj,ϕj)λB˙t(ϕj,ϕj),kj,jk+m1.\dot{\lambda}_{j}^{\ast}\delta_{jj^{\prime}}=\dot{A}_{t}(\phi_{j},\phi_{j^{\prime}})-\lambda\dot{B}_{t}(\phi_{j},\phi_{j^{\prime}}),\quad k\leq j,j^{\prime}\leq k+m-1.
Remark 7.

If we take the subsequence of h0h_{\ell}\rightarrow 0 to be unilateral, the limit λ˙j\dot{\lambda}_{j}^{\ast} in (94) exists and is either λ˙j+(t)\dot{\lambda}_{j}^{+}(t) or λ˙j(t)\dot{\lambda}_{j}^{-}(t).

Let

C˙tj=A˙tλ˙jBtλB˙t.\dot{C}_{t}^{\ast j}=\dot{A}_{t}-\dot{\lambda}_{j}^{\ast}B_{t}-\lambda\dot{B}_{t}.

It holds that

limsupuV1,vV1|1h(Ct+hjCt)(u,v)C˙tj(u,v)|=0,kjk+m1\lim_{\ell\rightarrow\infty}\sup_{\|u\|_{V}\leq 1,\|v\|_{V}\leq 1}\left|\frac{1}{h_{\ell}}(C_{t+h_{\ell}}^{j}-C_{t})(u,v)-\dot{C}_{t}^{\ast j}(u,v)\right|=0,\quad k\leq j\leq k+m-1

by (55), and also

C˙tj(u,v)=0,u,vYtλ,kjk+m1\dot{C}_{t}^{\ast j}(u,v)=0,\quad\forall u,v\in Y_{t}^{\lambda},\ k\leq j\leq k+m-1 (95)

by Lemma 11. Let

zj=1h(uj(t+h)ϕj).z_{\ell}^{j}=\frac{1}{h_{\ell}}(u_{j}(t+h_{\ell})-\phi_{j}).
Lemma 16.

It holds that

limCt(zj,v)=C˙tj(ϕj,v),vV.\lim_{\ell\rightarrow\infty}C_{t}(z_{\ell}^{j},v)=-\dot{C}_{t}^{\ast j}(\phi_{j},v),\quad\forall v\in V. (96)
Proof.

Given vVv\in V, we obtain

Ct(zj,v)\displaystyle C_{t}(z_{\ell}^{j},v) =\displaystyle= 1hCt(uj(t+h)ϕj,v)=1hCt(uj(t+h),v)\displaystyle\frac{1}{h_{\ell}}C_{t}(u_{j}(t+h_{\ell})-\phi_{j},v)=\frac{1}{h_{\ell}}C_{t}(u_{j}(t+h_{\ell}),v)
=\displaystyle= 1h(Ct+hjCt)(uj(t+h),v)=C˙tj(uj(t+h),v)+o(1)\displaystyle-\frac{1}{h_{\ell}}(C_{t+h_{\ell}}^{j}-C_{t})(u_{j}(t+h_{\ell}),v)=-\dot{C}_{t}^{\ast j}(u_{j}(t+h_{\ell}),v)+o(1)

by (51) and (92). Then (96) follows from (48). ∎

Recall that

R:XYtλ=uj(t)kjk+m1R:X\rightarrow Y_{t}^{\lambda}=\langle u_{j}(t)\mid k\leq j\leq k+m-1\rangle

is the orthogonal projection with respect to Bt(,)B_{t}(\cdot,\cdot), and P=IRP=I-R. There is a unique zjPVz_{\ast}^{j}\in PV satisfying

Ct(zj,v)=C˙tj(ϕj,v),vPV.C_{t}(z_{\ast}^{j},v)=-\dot{C}_{t}^{\ast j}(\phi_{j},v),\quad\forall v\in PV. (97)
Remark 8.

Equality (97) ensures

γλ˙j(ϕj)=zj,kjk+m1\gamma_{\dot{\lambda}_{j}^{\ast}}(\phi_{j})=z_{\ast}^{j},\quad k\leq j\leq k+m-1 (98)

under the notation of Definition 2.

Lemma 17.

It holds that

wlimPzj=zjin V,kjk+m1.\mbox{w}\mathchar 45\relax\lim_{\ell\rightarrow\infty}Pz_{\ell}^{j}=z_{\ast}^{j}\quad\mbox{in $V$},\quad k\leq j\leq k+m-1. (99)
Proof.

Lemma 16 ensures that {Pzj}\{Pz_{\ell}^{j}\} converges weakly in PVPV, and hence is bounded there:

PzjVC.\|Pz_{\ell}^{j}\|_{V}\leq C.

Then, passing to a subsequence denoted by the same symbol, there is z~jPV\tilde{z}_{j}\in PV such that

wlimPzj=z~j,\mbox{w}\mathchar 45\relax\lim_{\ell\rightarrow\infty}Pz_{\ell}^{j}=\tilde{z}_{j},

which satisfies

Ct(z~j,v)=C˙tj(ϕj,v),vVC_{t}(\tilde{z}_{j},v)=-\dot{C}_{t}^{\ast j}(\phi_{j},v),\quad\forall v\in V (100)

by Lemma 16. Since such z~jPV\tilde{z}_{j}\in PV is unique, we obtain the result with z~j=zj\tilde{z}_{j}=z_{\ast}^{j}. ∎

Remark 9.

Since (100) holds with z~j=zj\tilde{z}_{j}=z^{j}_{\ast}, this zjPVz_{j}^{\ast}\in PV defined by (97) satisfies

Ct(zj,v)=C˙tj(ϕj,v),vV.C_{t}(z_{\ast}^{j},v)=-\dot{C}_{t}^{\ast j}(\phi_{j},v),\quad\forall v\in V.
Remark 10.

Generally, the inequality

zjVC\|z_{\ell}^{j}\|_{V}\leq C

is not expected to hold, which causes the other difficulty in later arguments. In fact, if m=1m=1 and ϕj=uj(t)\phi_{j}=u_{j}(t), for example, this property means

|Bt(uj(t),whj)|C,whj=1h(uj(t+h)uj(t)).|B_{t}(u_{j}(t),w_{h_{\ell}}^{j})|\leq C,\quad w_{h}^{j}=\frac{1}{h}(u_{j}(t+h)-u_{j}(t)).

In the formal argument, we have actually (60), which, however, does not assure the actual convergence

limh0Bt(uj(t),whj)=12B˙t(uj(t),uj(t)).\lim_{h\rightarrow 0}B_{t}(u_{j}(t),w_{h}^{j})=-\frac{1}{2}\dot{B}_{t}(u_{j}(t),u_{j}(t)). (101)

In fact, the equality

1=Bt+h(uj(t+h),uj(t+h))=Bt(uj(t),uj(t))1=B_{t+h}(u_{j}(t+h),u_{j}(t+h))=B_{t}(u_{j}(t),u_{j}(t))

just implies

0\displaystyle 0 =\displaystyle= 1h{Bt+h(uj(t+h),uj(t+h))Bt(uj(t),uj(t))}\displaystyle\frac{1}{h}\{B_{t+h}(u_{j}(t+h),u_{j}(t+h))-B_{t}(u_{j}(t),u_{j}(t))\}
=\displaystyle= 1h(Bt+hBt)(uj(t+h),uj(t+h))\displaystyle\frac{1}{h}(B_{t+h}-B_{t})(u_{j}(t+h),u_{j}(t+h))
+1h{Bt(uj(t+h),uj(t+h))Bt(uj(t),uj(t))}\displaystyle+\frac{1}{h}\{B_{t}(u_{j}(t+h),u_{j}(t+h))-B_{t}(u_{j}(t),u_{j}(t))\}
=\displaystyle= B˙t(uj(t),uj(t))+o(1)+1hBt(uj(t+h)+uj(t),uj(t+h)uj(t))\displaystyle\dot{B}_{t}(u_{j}(t),u_{j}(t))+o(1)+\frac{1}{h}B_{t}(u_{j}(t+h)+u_{j}(t),u_{j}(t+h)-u_{j}(t))
=\displaystyle= B˙t(uj(t),uj(t))+2Bt(uj(t+h)+uj(t)2,whj)+o(1)\displaystyle\dot{B}_{t}(u_{j}(t),u_{j}(t))+2B_{t}(\frac{u_{j}(t+h)+u_{j}(t)}{2},w_{h}^{j})+o(1)

and hence

limh0Bt(uj(t+h)+uj(t)2,whj)=12B˙t(uj(t),uj(t))\lim_{h\rightarrow 0}B_{t}(\frac{u_{j}(t+h)+u_{j}(t)}{2},w_{h}^{j})=-\frac{1}{2}\dot{B}_{t}(u_{j}(t),u_{j}(t)) (102)

differently from (101). Here, the condition whjX=O(1)\|w_{h}^{j}\|_{X}=O(1) is necessary to conclude

Bt(uj(t+h),whj)=Bt(uj(t),whj)+o(1)B_{t}(u_{j}(t+h),w_{h}^{j})=B_{t}(u_{j}(t),w_{h}^{j})+o(1)

in the left-hand side of (102) from ϕj=uj(t)\phi_{j}=u_{j}(t) in (48). Our purpose, however, was to assure whjV=O(1)\|w_{h}^{j}\|_{V}=O(1), which is reduced to whjX=O(1)\|w_{h}^{j}\|_{X}=O(1) by PhwhjV=O(1)\|P_{h}w_{h}^{j}\|_{V}=O(1). This is a circular reasoning.

Lemma 18.

It holds that

slimPzj=zjin V,kjk+m1.\mbox{s}\mathchar 45\relax\lim_{\ell\rightarrow\infty}Pz_{\ell}^{j}=z_{\ast}^{j}\quad\mbox{in $V$},\qquad k\leq j\leq k+m-1. (103)
Proof.

Since VXV\hookrightarrow X is compact, we have

slimPzj=zjin X\mbox{s}\mathchar 45\relax\lim_{\ell\rightarrow\infty}Pz_{\ell}^{j}=z_{\ast}^{j}\quad\mbox{in $X$} (104)

in the previous lemma. Then we obtain

δPzjzjV2\displaystyle\delta\|Pz_{\ell}^{j}-z_{\ast}^{j}\|_{V}^{2} \displaystyle\leq At(Pzjzj,Pzjzj)=Ct(Pzjzj,Pzjzj)+o(1)\displaystyle A_{t}(Pz_{\ell}^{j}-z_{\ast}^{j},Pz_{\ell}^{j}-z_{\ast}^{j})=C_{t}(Pz_{\ell}^{j}-z_{\ast}^{j},Pz_{\ell}^{j}-z_{\ast}^{j})+o(1)
=\displaystyle= Ct(Pzj,Pzjzj)+o(1)=Ct(zj,Pzjzj)+o(1)\displaystyle C_{t}(Pz_{\ell}^{j},Pz_{\ell}^{j}-z_{\ast}^{j})+o(1)=C_{t}(z_{\ell}^{j},Pz_{\ell}^{j}-z_{\ast}^{j})+o(1)

by (99) and (104).

Since ϕjYtλ\phi_{j}\in Y_{t}^{\lambda} it holds that

Ct(zj,Pzjzj)\displaystyle C_{t}(z_{\ell}^{j},Pz_{\ell}^{j}-z_{\ast}^{j}) =\displaystyle= 1hCt(uj(t+h),Pzjzj)\displaystyle\frac{1}{h_{\ell}}C_{t}(u_{j}(t+h_{\ell}),Pz_{\ell}^{j}-z_{\ast}^{j})
=\displaystyle= 1h(CtCt+hj)(uj(t+h),Pzjzj)\displaystyle\frac{1}{h_{\ell}}(C_{t}-C_{t+h_{\ell}}^{j})(u_{j}(t+h_{\ell}),Pz_{\ell}^{j}-z_{\ast}^{j})
=\displaystyle= C˙tj(uj(t+h),Pzjzj)+o(1)\displaystyle-\dot{C}_{t}^{\ast j}(u_{j}(t+h_{\ell}),Pz_{\ell}^{j}-z_{\ast}^{j})+o(1)
=\displaystyle= C˙tj(ϕj,Pzjzj)+o(1)=o(1)\displaystyle-\dot{C}_{t}^{\ast j}(\phi_{j},Pz_{\ell}^{j}-z_{\ast}^{j})+o(1)=o(1)

by (48) and (99), because

vVC˙tj(ϕj,v)v\in V\ \mapsto\ \dot{C}_{t}^{\ast j}(\phi_{j},v)\in\mathbb{R}

is a bounded linear mapping. Then the result follows as

limPzjzjV=0.\lim_{\ell\rightarrow\infty}\|Pz_{\ell}^{j}-z_{\ast}^{j}\|_{V}=0.

Remark 11.

The limit zjz_{\ast}^{j} in (103) depends on the sequence h0h_{\ell}\rightarrow 0 because it is prescribed by (93) and (98). The limit λ¨j\ddot{\lambda}_{j}^{\ast} in the following lemma depends also h0h_{\ell}\rightarrow 0, but this ambiguity is canceled unilaterally. Hence these limits are uniquely determined as either h0h\downarrow 0 or h0h\uparrow 0, because they are characterized by an eigenvalue problem on a finite dimsnsional space, similarly to the first derivative of λj(t)\lambda_{j}(t). See Theorem 22 below.

Lemma 19.

There exists

λ¨jlim2h2(λj(t+h)λj(t)hλ˙j),kjk+1m\ddot{\lambda}_{j}^{\ast}\equiv\lim_{\ell\rightarrow\infty}\frac{2}{h_{\ell}^{2}}(\lambda_{j}(t+h_{\ell})-\lambda_{j}(t)-h_{\ell}\dot{\lambda}_{j}^{\ast}),\quad k\leq j\leq k+1-m (105)

with

λ¨j=(A¨tλB¨t2λ˙jB˙t)(ϕj,ϕj)2Ct(zj,zj).\ddot{\lambda}_{j}^{\ast}=(\ddot{A}_{t}-\lambda\ddot{B}_{t}-2\dot{\lambda}_{j}^{\ast}\dot{B}_{t})(\phi_{j},\phi_{j})-2C_{t}(z_{\ast}^{j},z_{\ast}^{j}).
Proof.

By (95) and Lemma 18, we have

Ct(zj,zj)\displaystyle C_{t}(z_{\ell}^{j},z_{\ell}^{j}) =\displaystyle= Ct(Pzj,Pzj)=Ct(zj,zj)+o(1)\displaystyle C_{t}(Pz_{\ell}^{j},Pz_{\ell}^{j})=C_{t}(z_{\ast}^{j},z_{\ast}^{j})+o(1) (106)
=\displaystyle= Ct(Pzj,zj)+o(1)=Ct(zj,zj)+o(1).\displaystyle C_{t}(Pz_{\ell}^{j},z_{\ast}^{j})+o(1)=C_{t}(z_{\ell}^{j},z_{\ast}^{j})+o(1).

It holds that

Ct(zj,zj)\displaystyle C_{t}(z_{\ell}^{j},z_{\ast}^{j}) =\displaystyle= 1hCt(uj(t+h)ϕj,zj)=1hCt(uj(t+h),zj)\displaystyle\frac{1}{h_{\ell}}C_{t}(u_{j}(t+h_{\ell})-\phi_{j},z_{\ast}^{j})=\frac{1}{h_{\ell}}C_{t}(u_{j}(t+h_{\ell}),z_{\ast}^{j})
=\displaystyle= 1h(CtCt+hj)(uj(t+h),zj)=C˙tj(uj(t+h),zj)+o(1)\displaystyle\frac{1}{h_{\ell}}(C_{t}-C_{t+h_{\ell}}^{j})(u_{j}(t+h_{\ell}),z_{\ast}^{j})=-\dot{C}_{t}^{\ast j}(u_{j}(t+h_{\ell}),z_{\ast}^{j})+o(1)
=\displaystyle= C˙tj(ϕj,zj)+o(1)\displaystyle-\dot{C}_{t}^{\ast j}(\phi_{j},z_{\ast}^{j})+o(1)

by (92) and ϕjYtλ\phi_{j}\in Y_{t}^{\lambda}, which implies

Ct(zj,zj)=C˙tj(ϕj,zj)+o(1)C_{t}(z_{\ast}^{j},z_{\ast}^{j})=-\dot{C}_{t}^{\ast j}(\phi_{j},z_{\ast}^{j})+o(1) (107)

by (106). It holds also that

C˙tj(ϕj,zj)\displaystyle\dot{C}_{t}^{\ast j}(\phi_{j},z_{\ast}^{j}) =\displaystyle= 1hC˙tj(ϕj,P(uj(t+h)ϕj))+o(1)\displaystyle\frac{1}{h_{\ell}}\dot{C}_{t}^{\ast j}(\phi_{j},P(u_{j}(t+h_{\ell})-\phi_{j}))+o(1) (108)
=\displaystyle= 1hC˙tj(ϕj,uj(t+h)ϕj)+o(1)\displaystyle\frac{1}{h_{\ell}}\dot{C}_{t}^{\ast j}(\phi_{j},u_{j}(t+h_{\ell})-\phi_{j})+o(1)
=\displaystyle= 1hC˙tj(ϕj,uj(t+h))+o(1)\displaystyle\frac{1}{h_{\ell}}\dot{C}_{t}^{\ast j}(\phi_{j},u_{j}(t+h_{\ell}))+o(1)

by (95) and ϕjYtλ\phi_{j}\in Y_{t}^{\lambda}.

Here, we use the asymptotics

Ct+hj(ϕj,uj(t+h))=Ctj(ϕj,uj(t+h))+hC˙tj(ϕj,uj(t+h))\displaystyle C_{t+h_{\ell}}^{j}(\phi_{j},u_{j}(t+h_{\ell}))=C_{t}^{j}(\phi_{j},u_{j}(t+h_{\ell}))+h_{\ell}\dot{C}_{t}^{\ast j}(\phi_{j},u_{j}(t+h_{\ell}))
+12h2C¨t,j(ϕj,uj(t+h))+o(h2)\displaystyle+\frac{1}{2}h_{\ell}^{2}\ddot{C}_{t,\ell}^{\ast j}(\phi_{j},u_{j}(t+h_{\ell}))+o(h_{\ell}^{2})

for

C¨t,j=A¨tλB¨t2λ˙jB˙t2h2(λj(t+h)λj(t)hλ˙j)Bt,\ddot{C}_{t,\ell}^{\ast j}=\ddot{A}_{t}-\lambda\ddot{B}_{t}-2\dot{\lambda}_{j}^{\ast}\dot{B}_{t}-\frac{2}{h_{\ell}^{2}}(\lambda_{j}(t+h_{\ell})-\lambda_{j}(t)-h\dot{\lambda}_{j}^{\ast})B_{t}, (109)

derived from (55) and (87). Since

Ct+hj(ϕj,uj(t+h))=Ctj(ϕj,uj(t+h))=0C_{t+h_{\ell}}^{j}(\phi_{j},u_{j}(t+h_{\ell}))=C_{t}^{j}(\phi_{j},u_{j}(t+h_{\ell}))=0

holds by (92), we obtain

Ctj(zj,zj)\displaystyle C_{t}^{j}(z_{\ast}^{j},z_{\ast}^{j}) =\displaystyle= C˙tj(ϕj,zj)+o(1)=1hC˙tj(ϕj,uj(t+h))+o(1)\displaystyle-\dot{C}_{t}^{\ast j}(\phi_{j},z_{\ast}^{j})+o(1)=-\frac{1}{h_{\ell}}\dot{C}_{t}^{\ast j}(\phi_{j},u_{j}(t+h_{\ell}))+o(1)
=\displaystyle= 12C¨t,j(ϕj,uj(t+h))+o(1)=12C¨t,j(ϕj,ϕj)+o(1)\displaystyle\frac{1}{2}\ddot{C}_{t,\ell}^{\ast j}(\phi_{j},u_{j}(t+h_{\ell}))+o(1)=\frac{1}{2}\ddot{C}_{t,\ell}^{\ast j}(\phi_{j},\phi_{j})+o(1)
=\displaystyle= 12(A¨tλB¨t2λ˙jB˙t)(ϕj,ϕj)1h2(λj(t+h)λj(t)hλ˙j)+o(1)\displaystyle\frac{1}{2}(\ddot{A}_{t}-\lambda\ddot{B}_{t}-2\dot{\lambda}_{j}^{\ast}\dot{B}_{t})(\phi_{j},\phi_{j})-\frac{1}{h_{\ell}^{2}}(\lambda_{j}(t+h_{\ell})-\lambda_{j}(t)-h_{\ell}\dot{\lambda}_{j}^{\ast})+o(1)

by (107)-(108) and Bt(ϕj,ϕj)=1B_{t}(\phi_{j},\phi_{j})=1. Then it follows that

lim2h2(λj(t+h)λj(t)hλ˙j)=(A¨tλB¨t2λ˙jB˙t)(ϕj,ϕj)2Ct(zj,zj),\lim_{\ell\rightarrow\infty}\frac{2}{h_{\ell}^{2}}(\lambda_{j}(t+h_{\ell})-\lambda_{j}(t)-h_{\ell}\dot{\lambda}_{j}^{\ast})\\ =(\ddot{A}_{t}-\lambda\ddot{B}_{t}-2\dot{\lambda}_{j}^{\ast}\dot{B}_{t})(\phi_{j},\phi_{j})-2C_{t}(z_{\ast}^{j},z_{\ast}^{j}),

and the proof is complete. ∎

Lemma 20.

If λ˙λ˙j=λ˙j\dot{\lambda}_{\ast}\equiv\dot{\lambda}_{j}^{\ast}=\dot{\lambda}_{j^{\prime}}^{\ast} arises for some kjjk+m1k\leq j\neq j^{\prime}\leq k+m-1, then it holds that

(A¨tλB¨t2λ˙B˙t)(ϕj,ϕj)=2Ct(zj,zj).(\ddot{A}_{t}-\lambda\ddot{B}_{t}-2\dot{\lambda}_{\ast}\dot{B}_{t})(\phi_{j},\phi_{j^{\prime}})=2C_{t}(z_{\ast}^{j},z_{\ast}^{j^{\prime}}).
Proof.

As in the previous lemma we obtain

Ct(zj,zj)\displaystyle C_{t}(z_{\ast}^{j},z_{\ast}^{j^{\prime}}) =\displaystyle= Ct(Pzj,zj)+o(1)=Ct(zj,zj)+o(1)\displaystyle C_{t}(Pz_{\ell}^{j},z_{\ast}^{j^{\prime}})+o(1)=C_{t}(z_{\ell}^{j},z_{\ast}^{j^{\prime}})+o(1)
=\displaystyle= 1hCt(uj(t+h)ϕj,zj)+o(1)=1hCt(uj(t+h),zj)+o(1)\displaystyle\frac{1}{h_{\ell}}C_{t}(u_{j}(t+h_{\ell})-\phi_{j},z_{\ast}^{j^{\prime}})+o(1)=\frac{1}{h_{\ell}}C_{t}(u_{j}(t+h_{\ell}),z_{\ast}^{j^{\prime}})+o(1)
=\displaystyle= 1h(CtCt+h)(uj(t+h),zj)+o(1)\displaystyle\frac{1}{h_{\ell}}(C_{t}-C_{t+h_{\ell}})(u_{j}(t+h),z_{\ast}^{j^{\prime}})+o(1)
=\displaystyle= C˙tj(uj(t+h),zj)+o(1)=C˙tj(ϕj,zj)+o(1)\displaystyle-\dot{C}_{t}^{\ast j}(u_{j}(t+h_{\ell}),z_{\ast}^{j^{\prime}})+o(1)=-\dot{C}_{t\ast}^{j}(\phi_{j},z_{\ast}^{j^{\prime}})+o(1)
=\displaystyle= 1hC˙tj(ϕj,P(uj(t+h)ϕj))+o(1)\displaystyle-\frac{1}{h_{\ell}}\dot{C}_{t}^{\ast j}(\phi_{j},P(u_{j^{\prime}}(t+h_{\ell})-\phi_{j^{\prime}}))+o(1)
=\displaystyle= 1hC˙tj(ϕj,uj(t+h)ϕj)+o(1)\displaystyle-\frac{1}{h_{\ell}}\dot{C}_{t}^{\ast j}(\phi_{j},u_{j^{\prime}}(t+h_{\ell})-\phi_{j^{\prime}})+o(1)
=\displaystyle= 1hC˙tj(ϕj,uj(t+h))+o(1)\displaystyle-\frac{1}{h_{\ell}}\dot{C}_{t}^{\ast j}(\phi_{j},u_{j^{\prime}}(t+h_{\ell}))+o(1)

by ϕj,ϕjYtλ\phi_{j},\phi_{j^{\prime}}\in Y_{t}^{\lambda}. Then it holds that

Ct+hj(ϕj,uj(t+h))=Ct(ϕj,uj(t+h))+hC˙tj(ϕj,uj(t+h))\displaystyle C_{t+h_{\ell}}^{j}(\phi_{j},u_{j^{\prime}}(t+h_{\ell}))=C_{t}(\phi_{j},u_{j^{\prime}}(t+h_{\ell}))+h_{\ell}\dot{C}_{t}^{\ast j}(\phi_{j},u_{j^{\prime}}(t+h_{\ell}))
+12h2C¨t,j(ϕj,uj(t+h))+o(h2)\displaystyle\quad+\frac{1}{2}h_{\ell}^{2}\ddot{C}_{t,\ell}^{\ast j}(\phi_{j},u_{j^{\prime}}(t+h_{\ell}))+o(h_{\ell}^{2})

with

Ct(ϕj,uj(t+h))=0C_{t}(\phi_{j},u_{j^{\prime}}(t+h_{\ell}))=0

and

Ct+hj(ϕj,uj(t+h))\displaystyle C^{j}_{t+h_{\ell}}(\phi_{j},u_{j^{\prime}}(t+h_{\ell})) =\displaystyle= (Ct+hjCt+hj)(ϕj,uj(t+h))\displaystyle(C_{t+h_{\ell}}^{j}-C_{t+h_{\ell}}^{j^{\prime}})(\phi_{j},u_{j^{\prime}}(t+h_{\ell}))
=\displaystyle= (λj(t+h)λj(t+h))Bt+h(ϕj,uj(t+h))\displaystyle(\lambda_{j^{\prime}}(t+h_{\ell})-\lambda_{j}(t+h_{\ell}))B_{t+h_{\ell}}(\phi_{j},u_{j^{\prime}}(t+h_{\ell}))

by (92), to conclude

Ct(zj,zj)\displaystyle C_{t}(z_{\ast}^{j},z_{\ast}^{j^{\prime}}) =\displaystyle= 1h2(λj(t+h)λj(t+h))Bt+h(ϕj,uj(t+h))\displaystyle\frac{1}{h_{\ell}^{2}}(\lambda_{j}(t+h_{\ell})-\lambda_{j^{\prime}}(t+h_{\ell}))B_{t+h_{\ell}}(\phi_{j},u_{j^{\prime}}(t+h_{\ell})) (110)
+12C¨t,j(ϕj,uj(t+h))+o(1).\displaystyle+\frac{1}{2}\ddot{C}_{t,\ell}^{\ast j}(\phi_{j},u_{j^{\prime}}(t+h_{\ell}))+o(1).

Then we use

λj(t+h)=λ+hλ˙j+h22λ¨j+o(h2)\displaystyle\lambda_{j}(t+h_{\ell})=\lambda+h_{\ell}\dot{\lambda}_{j}^{\ast}+\frac{h_{\ell}^{2}}{2}\ddot{\lambda}_{j}^{\ast}+o(h_{\ell}^{2})
λj(t+h)=λ+hλ˙j+h22λ¨j+o(h2)\displaystyle\lambda_{j^{\prime}}(t+h_{\ell})=\lambda+h_{\ell}\dot{\lambda}_{j^{\prime}}^{\ast}+\frac{h_{\ell}^{2}}{2}\ddot{\lambda}_{j^{\prime}}^{\ast}+o(h_{\ell}^{2})

with λ˙j=λ˙j\dot{\lambda}_{j}^{\ast}=\dot{\lambda}_{j^{\prime}}^{\ast}, to deduce

lim1h2(λj(t+h)λj(t+h))Bt+h(ϕj,uj(t+h))\displaystyle\lim_{\ell\rightarrow\infty}\frac{1}{h_{\ell}^{2}}(\lambda_{j}(t+h_{\ell})-\lambda_{j^{\prime}}(t+h_{\ell}))B_{t+h_{\ell}}(\phi_{j},u_{j^{\prime}}(t+h_{\ell}))
=12(λ¨j+λ¨j)Bt(ϕj,ϕj)=0.\displaystyle\quad=\frac{1}{2}(\ddot{\lambda}_{j}^{\ast}+\ddot{\lambda}_{j^{\prime}}^{\ast})B_{t}(\phi_{j},\phi_{j^{\prime}})=0.

Then the result follows from (109), (110), and the previous lemma. ∎

Recall Ftλ,λF^{\lambda,\lambda^{\prime}}_{t} in Definition 2.

Lemma 21.

Define μ~kμ~k+m1\tilde{\mu}_{k}\leq\cdots\leq\tilde{\mu}_{k+m-1} by

{μ~jkjk+m1}={λ˙jkjk+m1},\{\tilde{\mu}_{j}\mid k\leq j\leq k+m-1\}=\{\dot{\lambda}_{j}^{\ast}\mid k\leq j\leq k+m-1\},

and assume k<rk+mk\leq\ell<r\leq k+m be such that

μ~1<μ~μ~==μ~r1<μ~r.\tilde{\mu}_{\ell-1}<\tilde{\mu}\equiv\tilde{\mu}_{\ell}=\cdots=\tilde{\mu}_{r-1}<\tilde{\mu}_{r}.

under the agreement of

μ~k1=,μ~k+m=+.\tilde{\mu}_{k-1}=-\infty,\quad\tilde{\mu}_{k+m}=+\infty.

Then, σ=λ¨j\sigma=\ddot{\lambda}_{j}^{\ast}, jr1\ell\leq j\leq r-1, is an eigenvalue of

uYλ,t,r,Ftλ,μ~(u,v)=σBt(u,v),vYλ,t,r,u\in Y_{\lambda,t}^{\ell,r},\quad F^{\lambda,\tilde{\mu}}_{t}(u,v)=\sigma B_{t}(u,v),\ \forall v\in Y_{\lambda,t}^{\ell,r},

for

Yλ,t,r=uj(t)jr1Ytλ.Y_{\lambda,t}^{\ell,r}=\langle u_{j}(t)\mid\ell\leq j\leq r-1\rangle\subset Y^{\lambda}_{t}.
Proof.

Lemmas 19 and 20 imply

Ftλ,μ~(ϕj,ϕj)=δjjλ¨j,j,jr1,F^{\lambda,\tilde{\mu}}_{t}(\phi_{j},\phi_{j^{\prime}})=\delta_{jj^{\prime}}\ddot{\lambda}_{j}^{\ast},\quad\ell\leq j,j^{\prime}\leq r-1,

and hence the result follows from

Yλ,t,r=ϕjjr1,Bt(ϕj,ϕj)=δjj.Y_{\lambda,t}^{\ell,r}=\langle\phi_{j}\mid\ell\leq j\leq r-1\rangle,\quad B_{t}(\phi_{j},\phi_{j^{\prime}})=\delta_{jj^{\prime}}.

Theorem 15 is now reduced to the following theorem.

Theorem 22.

Fix tIt\in I, and assume (19) for =k\ell=k and n=mn=m. Put (21) and let k<rk+mk\leq\ell<r\leq k+m be such that

λ˙1+(t)<λλ˙+(t)==λ˙r1+(t)<λ˙r+(t).\dot{\lambda}^{+}_{\ell-1}(t)<\lambda^{\prime}\equiv\dot{\lambda}^{+}_{\ell}(t)=\cdots=\dot{\lambda}_{r-1}^{+}(t)<\dot{\lambda}_{r}^{+}(t). (111)

Then there exists

λj′′(t)=limh+02h2(λj(t+h)λhλ),jr1.\lambda_{j}^{\prime\prime}(t)=\lim_{h\rightarrow+0}\frac{2}{h^{2}}(\lambda_{j}(t+h)-\lambda-h\lambda^{\prime}),\quad\ell\leq j\leq r-1. (112)

It holds, furthermore, that

λj′′(t)=σj+1,jr1,\lambda_{j}^{\prime\prime}(t)=\sigma_{j-\ell+1},\quad\ell\leq j\leq r-1, (113)

where σq\sigma_{q}, 1qr1\leq q\leq r-\ell, denotes the qq-th eigenvalue of

uYλ,t,r,Ftλ,λ(u,v)=σBt(u,v),vYλ,t,r.u\in Y_{\lambda,t}^{\ell,r},\qquad F_{t}^{\lambda,\lambda^{\prime}}(u,v)=\sigma B_{t}(u,v),\ \forall v\in Y_{\lambda,t}^{\ell,r}.

If

λ˙1(t)>λλ˙(t)==λ˙r1(t)>λ˙r(t),\dot{\lambda}^{-}_{\ell-1}(t)>\lambda^{\prime}\equiv\dot{\lambda}^{-}_{\ell}(t)=\cdots=\dot{\lambda}_{r-1}^{-}(t)>\dot{\lambda}_{r}^{-}(t),

there arises that

λj′′(t)=limh02h2(λj(t+h)λhλ),jr1\lambda_{j}^{\prime\prime}(t)=\lim_{h\rightarrow-0}\frac{2}{h^{2}}(\lambda_{j}(t+h)-\lambda-h\lambda^{\prime}),\quad\ell\leq j\leq r-1

with (113).

Proof.

In the previous lemma, we obtain

λ¨λ¨r1.\ddot{\lambda}_{\ell}^{\ast}\leq\cdots\leq\ddot{\lambda}_{r-1}^{\ast}.

Hence the result follows similarly to Theorem 12. ∎

Remark 12.

By the above theorem and Remark 7, the limits (20),

λ¨j±(t)=limh±02h2(λj(t+h)λj(t)hλ˙j±(t))\ddot{\lambda}_{j}^{\pm}(t)=\lim_{h\rightarrow\pm 0}\frac{2}{h^{2}}(\lambda_{j}(t+h)-\lambda_{j}(t)-h\dot{\lambda}_{j}^{\pm}(t))

exist for any jj, provided that the conditions (41), (42), (54), (55), (86), and (87) hold. If these conditions hold for any tt locally uniformly in II, these limits are locally uniform in tIt\in I.

Theorem 5 is reduced to the following abstract theorem. The proof is similar to that of Theorem 13.

Theorem 23.

Let the assumption of Theorem 15 hold for any tt. Fix tIt\in I, and assume, furthermore,

limh0supuV,vV1|A¨t+h(u,v)A¨t(u,v)|=0\displaystyle\lim_{h\rightarrow 0}\sup_{\|u\|_{V},\|v\|_{V}\leq 1}\left|\ddot{A}_{t+h}(u,v)-\ddot{A}_{t}(u,v)\right|=0
limh0sup|u|X,|v|X1|B¨t+h(u,v)B¨t(u,v)|=0.\displaystyle\lim_{h\rightarrow 0}\sup_{|u|_{X},|v|_{X}\leq 1}\left|\ddot{B}_{t+h}(u,v)-\ddot{B}_{t}(u,v)\right|=0. (114)

Then it holds that

limh±0λ¨j±(t+h)=λ¨j±(t).\lim_{h\rightarrow\pm 0}\ddot{\lambda}_{j}^{\pm}(t+h)=\ddot{\lambda}_{j}^{\pm}(t).

Finally, Theorem 6 is reduced to the following abstract theorem.

Theorem 24.

If (114) is valid to any tt in the previous theorem, C1C^{1} curves C~j\tilde{C}_{j}, 1jm1\leq j\leq m, in Theorem 3 are C2C^{2}.

Proof.

Define λ~j(t)\tilde{\lambda}_{j}(t) by

C~j={λ~j(t)tI}, 1jm.\tilde{C}_{j}=\{\tilde{\lambda}_{j}(t)\mid t\in I\},\ 1\leq j\leq m.

From the proof of Theorem 14, Theorems 22 and 23 guarantee the existence of

λ~j′′(t)=limh01h2(λ~j(t+h)λ~j(t)hλ~j(t))\tilde{\lambda}_{j}^{\prime\prime}(t)=\lim_{h\rightarrow 0}\frac{1}{h^{2}}(\tilde{\lambda}_{j}(t+h)-\tilde{\lambda}_{j}(t)-h\tilde{\lambda}_{j}^{\prime}(t)) (115)

together with its continuity in tt,

limh±0λ~j′′(t+h)=λ~j′′(t)\lim_{h\rightarrow\pm 0}\tilde{\lambda}_{j}^{\prime\prime}(t+h)=\tilde{\lambda}_{j}^{\prime\prime}(t)

for any tt and jj. This convergence (115), furthermore, is locally uniform in tIt\in I by Remark 12.

Then it follows that

λ~j(t+h)=λ~j(t)+hλ~j(t)+h22λ~j′′(t)+o(h2)\displaystyle\tilde{\lambda}_{j}(t+h)=\tilde{\lambda}_{j}(t)+h\tilde{\lambda}_{j}^{\prime}(t)+\frac{h^{2}}{2}\tilde{\lambda}_{j}^{\prime\prime}(t)+o(h^{2})
λ~j(t)=λ~j(t+h)hλ~j(t+h)+h22λ~j′′(t+h)+o(h2),\displaystyle\tilde{\lambda}_{j}(t)=\tilde{\lambda}_{j}(t+h)-h\tilde{\lambda}_{j}^{\prime}(t+h)+\frac{h^{2}}{2}\tilde{\lambda}_{j}^{\prime\prime}(t+h)+o(h^{2}),

as h0h\rightarrow 0, which implies

limh01h(λ~j(t+h)λ~j(t))=limh012(λ~j′′(t+h)+λ~j′′(t))=λ~j′′(t)\lim_{h\rightarrow 0}\frac{1}{h}(\tilde{\lambda}^{\prime}_{j}(t+h)-\tilde{\lambda}^{\prime}_{j}(t))=\lim_{h\rightarrow 0}\frac{1}{2}(\tilde{\lambda}^{\prime\prime}_{j}(t+h)+\tilde{\lambda}^{\prime\prime}_{j}(t))=\tilde{\lambda}^{\prime\prime}_{j}(t)

for any tIt\in I. Hence these C~j\tilde{C}_{j}’s are C2C^{2}. ∎

Acknowledgements. The authors thank the referees for valuable comments on the original manuscprit.

References

  • [1] S.-N. Chow and J.K. Hale, Methods of Bifurcation Theory, Springer-Verlag, New York, 1982.
  • [2] H. Fujita, N. Saito, and T. Suzuki, Operator Theory and Numerical Methods, Elsevier, Amsterdam, 2001.
  • [3] P.R. Garabedian and M. Schiffer, Convexity of domain functionals, J. Ann. Math. 2 (1952-53) 281–368.
  • [4] S. Jimbo and E. Ushikoshi, Hadamard variational formula for the multiple eigenvalues of the Stokes operator with the Dirichlet bundary conditions, Far East J. Math. Sci. 98 (2015) 713-739.
  • [5] T. Kato, Perturbation Theory for Linear Operators, second edition, Springer-Verlag, Berlin, 1976.
  • [6] A. Kufner, O. John, and S. Fučik, Function Spaces, Academia, Prague, 1977.
  • [7] P.D. Lamberti and M.L. De Cristoforis, An analyticity result for the dependence of multiple eigenvalues and eigenspaces of the Laplace operator upon perturbation of the domain, Glasgow Math. J. 44 (2002) 29-43.
  • [8] P.D. Lamberti and M.L. De Cristoforis, A real analyticity result for symmetric functions of the eigenvalues of a domain dpendent Dirichlet prbolem for the Laplace operator, J. Nonl. Conv. Anal. 5 (2004) 18-42.
  • [9] J.E. Marsden and T.J.R. Hughes, Mathematical Foundations of Elasticity, Dover, New York 1994.
  • [10] F. Rellich, Perturbation Theory of Eigenvalue Problems, Lecture Notes, New York Univ. 1953.
  • [11] T. Suzuki, Applied Analysis, Mathematics for Science, Mathematics for Science, Technology, Engineering, third edition, World Scientific, London, 2022.
  • [12] T. Suzuki, T. Tsuchiya, Convergence analysis of trial free boundary methods for the two-dimensional filtration problem, Numer. Math. 100 (2005) 537–564.
  • [13] T. Suzuki, T. Tsuchiya, Weak formulation of Hadamard variation applied to the filtration problem, Japan. J. Indus. Appl. Math. 28 (2011) 327–350.
  • [14] 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.
  • [15] 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.