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

\recdate

20XX1199\revdate20XX199

Liouville’s formulae and Hadamard variation with respect to general domain perturbations

Takashi Suzuki Center of Mathematical Modeling and Data Science,
Osaka University,
Toyonaka, 560-8531, Japan
[email protected]
   Takuya Tsuchiya Graduate School of Science and Engineering,
Ehime University,
Matsuyama 790-8577, Japan
[email protected]
Abstract.

We study Hadamard variations with respect to general domain perturbations, particularly for the Neumann boundary condition. They are derived from new Liouville’s formulae concerning the transformation of volume and area integrals. Then, relations to several geometric quantities are discussed; differential forms and the second fundamental form on the boundary.

Key words and phrases:
the Green function, domain perturbations, the Hadamard variation, Liouville’s formulae, the Neumann boundary condition
2020 Mathematics Subject Classification:
Primary 35J25; Secondary 35R35

1. Introduction

Our purpose is to establish Liouville’s formulae on volume and area integrals and derive Hadamard variations with respect to general domain perturbations for the Neumann boundary condition.

Let Ω\Omega be a bounded domain in dd-dimensional Euclidean space d\mathbb{R}^{d}, d2d\geq 2. If its boundary Ω\partial\Omega is represented as graphs of Lipschitz functions, Ω\Omega is called a Lipschitz domain. For the detailed definition of Lipschitz domains, see [8] and the references there in. If Ω\Omega is a Lipschitz domain, the smoothness of Ω\partial\Omega is denoted by C0,1C^{0,1}. We suppose that Ω\partial\Omega is divided into two non-overlapped closed sets γ0\gamma^{0} and γ1\gamma^{1} satisfying

γ0γ1=Ω,γ0γ1=.\displaystyle\gamma^{0}\cup\gamma^{1}=\partial\Omega,\quad\gamma^{0}\cap\gamma^{1}=\emptyset. (1)

This assumption yields that γi\gamma^{i}, i=0,1i=0,1, do not have their boundaries in Ω\partial\Omega.

We study the Poisson problem with the mixed boundary condition:

Δz=fin Ω,z=φon γ0,z𝝂=ψon γ1,\displaystyle-\Delta z=f\ \mbox{in $\Omega$},\quad z=\varphi\ \mbox{on $\gamma^{0}$},\quad\frac{\partial z}{\partial\boldsymbol{\nu}}=\psi\ \mbox{on $\gamma^{1}$}, (2)

where Δ=i=1d2xi2\Delta=\sum_{i=1}^{d}\frac{\partial^{2}}{\partial x_{i}^{2}} stands for the Laplacian and 𝝂\boldsymbol{\nu} is the unit outer normal vector on Ω\partial\Omega. The standard theory of elliptic PDE tells us that if Ω\partial\Omega is C2C^{2}, or C1,1C^{1,1} more weakly, which means that it is Lipschitz continuous up to the first derivatives, and fL2(Ω)f\in L^{2}(\Omega), φH2(Ω)\varphi\in H^{2}(\Omega), and ψH1(Ω)\psi\in H^{1}(\Omega), equation (2) admits a unique solution zH2(Ω)z\in H^{2}(\Omega). (See Section 2.1 below for the case of Lipschitz domains.)

This solution admits the representation

z(y)=ΩN(x,y)f(x)𝑑xγ0φ(x)N𝝂x(x,y)𝑑sx+γ1N(x,y)ψ(x)𝑑sxz(y)=\int_{\Omega}N(x,y)f(x)\ dx-\int_{\gamma^{0}}\varphi(x)\frac{\partial N}{\partial\boldsymbol{\nu}_{x}}(x,y)\ ds_{x}+\int_{\gamma^{1}}N(x,y)\psi(x)\ ds_{x} (3)

for yΩy\in\Omega, where dsxds_{x} is the surface element and N(x,y)N(x,y) is the Green’s function. Thus, given yΩy\in\Omega, if Γ(x)\Gamma(x) denotes the fundamental solution of Δ\Delta:

Γ(x)={12πlog|x|,d=2,1(d2)ωd|x|2d,d3,\Gamma(x)=\left\{\begin{array}[]{ll}-\frac{1}{2\pi}\log\left|x\right|,&d=2,\\ \frac{1}{(d-2)\omega_{d}}|x|^{2-d},&d\geq 3,\end{array}\right. (4)

where ωd\omega_{d} is the volume of the unit ball in d\mathbb{R}^{d}, and if u=u(x)u=u(x) is the solution to

Δu=0in Ω,u=Γ(y)on γ0,u𝝂=𝝂Γ(y)on γ1,\Delta u=0\ \mbox{in $\Omega$},\quad u=-\Gamma(\cdot-y)\ \mbox{on $\gamma^{0}$},\quad\frac{\partial u}{\partial\boldsymbol{\nu}}=-\frac{\partial\;}{\partial\boldsymbol{\nu}}\Gamma(\cdot-y)\ \mbox{on $\gamma^{1}$}, (5)

this N(x,y)N(x,y) is given by

N(x,y)=Γ(xy)+u(x).N(x,y)=\Gamma(x-y)+u(x). (6)

We take a family of domain perturbations parametrized by tt, |t|1|t|\ll 1, which is denoted by {Tt}\{T_{t}\} (the exact definition will be given in Section 2.3 below). Suppose that the boundaries γi\gamma^{i}, i=0,1i=0,1, are mapped onto γti\gamma_{t}^{i}, i=0,1i=0,1, respectively, by TtT_{t}:

Tt(γi)=γti,i=0,1.T_{t}(\gamma^{i})=\gamma^{i}_{t},\quad i=0,1.

Then the Green’s function on Ωt\Omega_{t} is defined by

N(x,y,t)=Γ(xy)+u(x,t),N(x,y,t)=\Gamma(x-y)+u(x,t), (7)

where u=u(x,t)u=u(x,t) is the solution to

Δu(,t)=0in Ωt,u(,t)=Γ(y)on γt0,uν(,t)=νΓ(y)on γt1.\Delta u(\cdot,t)=0\ \mbox{in $\Omega_{t}$},\quad u(\cdot,t)=-\Gamma(\cdot-y)\ \mbox{on $\gamma_{t}^{0}$},\quad\frac{\partial u}{\partial\nu}(\cdot,t)=-\frac{\partial}{\partial\nu}\Gamma(\cdot-y)\ \mbox{on $\gamma_{t}^{1}$}. (8)

By this definition, it holds that Ω0=Ω\Omega_{0}=\Omega, N(x,y,0)=N(x,y)N(x,y,0)=N(x,y), and u(x,0)=u(x)u(x,0)=u(x).

Given x,yΩx,y\in\Omega, we have x,yΩtx,y\in\Omega_{t} for |t|1|t|\ll 1. Then the Hadamard variation of the Green’s function N(x,y)N(x,y) is defined by

δN(x,y)=Nt(x,y,t)|t=0=ut(x,t)|t=0,(x,y)Ω×Ω.\delta N(x,y)=\frac{\partial N}{\partial t}(x,y,t)\Big{|}_{t=0}=\frac{\partial u}{\partial t}(x,t)\Big{|}_{t=0},\quad(x,y)\in\Omega\times\Omega. (9)

The second variation is defined similarly:

δ2N(x,y)=2Nt2(x,y,t)|t=0=2ut2(x,t)|t=0,(x,y)Ω×Ω.\delta^{2}N(x,y)=\frac{\partial^{2}N}{\partial t^{2}}(x,y,t)\Big{|}_{t=0}=\frac{\partial^{2}u}{\partial t^{2}}(x,t)\Big{|}_{t=0},\quad(x,y)\in\Omega\times\Omega. (10)

These notions are classical, but to clarify the meaning of these derivatives, including their existence, is one of our main aims. The other is to prescribe a class of domains which admits these limits. The Lipschitz domain is a main target, from the viewpoint of numerical computations in engineering, which are often concerned on polygons on the plane. Hence we are taking the applications to free boundary problems [6] or shape optimizations [1] in mind. These problems in engineering induce the third motivation of ours, study on the general domain perturbation of the domain. Thus we continue our previous work [8] on the Dirichlet boundary condition.

So far, the domain perturbation has been often introduced by the deformation of Ω\partial\Omega as in

Ωt:x+tδρ(x)νx,xΩ,\partial\Omega_{t}:x+t\cdot\delta\rho(x)\nu_{x},\ x\in\partial\Omega, (11)

where δρ(x)\delta\rho(x) is a smooth function of xΩx\in\partial\Omega. This method, which may be called the normal perturbation, does not always work for the general Lipschitz domain, for example, if Ω\partial\Omega has a corner. The dynamical perturbation introduced by [8] may fit more the Lipschitz domain. It is given by Ttx=X(t)T_{t}x=X(t) for the solution X=X(t)X=X(t) to

dXdt=v(X),X|t=0=xΩ¯,\frac{dX}{dt}=v(X),\quad\left.X\right|_{t=0}=x\in\overline{\Omega}, (12)

where v=v(x)v=v(x) is a Lipschitz continuous vector field defined on an open neighbourhood of Ω¯\overline{\Omega}.

In [8] we have studied the first and the second Hadamard variations under the general perturbation of Lipschitz domains for the case γ1=\gamma^{1}=\emptyset, that is, for the Dirichlet boundary condition. This paper is devoted to the general case of γ1\gamma^{1}\neq\emptyset. Hence it is concerned on the Neumann boundary condition essentially, and extends the classical result of [4] on (11) for d=2,3d=2,3. Meanwhile we carefully examine the regularity of the domain admitting these variations.

Our strategy is a systematic use of Liouville’s formulae concerning the variation of volume and area integrals under the tranformation of variables. Actually, we have derived Liouville’s first volume and area formulae in the general form to treat the Dirichlet boundary condition in [8]. Here we formulate the second formulae of these integrals to study the Neumann boundary condition. These formulate are concerned on the second derivatives and are to be used for numerical computations adapting Newton’s method. From the viewpoint of pure mathematics, on the other hand, a role of the second fundamental form of the boundary in the second Hadamard variation is clarified for general domain perturbations. The other topic is the derivation of the first and the second area formulae via the transformation of differential forms.

This paper is composed of four sections and two appendices. Taking preliminaries in §\S2, we show Liouville’s volume and area formulae in §\S3. Then §\S4 is devoted to the Hadamard variation with respect to general domain perturbations. The first appendix, §\SA, is devoted to the derivation of Liouville’s area formulae via differential forms. In the second appendix, §\SB, we show a form of Liouville’s second area formula represented by the second fundamental form of Ω\partial\Omega.

Several formulae on Hadamard variations of eigenvalues noticed by [4], such as the harmonic concavity of the first eigenvalue, will be generalized in our forthcoming paper, with rigorous proof of the existence of these variations.

2. Preliminaries

2.1. Poisson equation on Lipschitz domains

The fundamental property of the Lipschitz domain is the following fact [2, Theorem 3, p.127].

Theorem 1.

If Ωd\Omega\subset\mathbb{R}^{d} is a Lipschitz domain, then the set of functions C(Ω¯)C^{\infty}(\overline{\Omega}) is dense in W1,p(Ω)W^{1,p}(\Omega) for 1p<1\leq p<\infty, where

C(Ω¯)={v:Ω¯v~C0(d),v~|Ω¯=v}.C^{\infty}(\overline{\Omega})=\{v:\overline{\Omega}\rightarrow\mathbb{R}\mid\exists\tilde{v}\in C_{0}^{\infty}(\mathbb{R}^{d}),\ \left.\tilde{v}\right|_{\overline{\Omega}}=v\}. (13)

This theorem ensures the validity of the trace operator to Ω\partial\Omega ([2, Theorem 1, p.133]), and then the unique solvability of (2) for the Lipschitz domain Ω\Omega holds as in the case of γ1=\gamma^{1}=\emptyset. Hence we have the following theorem similarly to [8]. Note that the spaces C0,1(Ω)C^{0,1}(\partial\Omega) and Hθ(Ω)H^{\theta}(\partial\Omega) for 0<θ<10<\theta<1 are well-defined by the local chart because Ω\partial\Omega is Lipschitz continuous.

Theorem 2.

If Ωd\Omega\subset\mathbb{R}^{d} is a bounded Lipschitz domain satisfying (1), there arise the following facts: First, the trace operator vC(Ω¯)v|ΩC0,1(Ω)v\in C^{\infty}(\overline{\Omega})\mapsto\left.v\right|_{\partial\Omega}\in C^{0,1}(\partial\Omega) is extended as

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

Then there arise the isomorphisms

vH1(Ω)/H01(Ω)v|ΩH1/2(Ω)v\in H^{1}(\Omega)/H^{1}_{0}(\Omega)\ \mapsto\ \left.v\right|_{\partial\Omega}\in H^{1/2}(\partial\Omega)

and

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})

for V={vH1(Ω)v|γ0=0}V=\{v\in H^{1}(\Omega)\mid\left.v\right|_{\gamma^{0}}=0\}. Second, the normal derivative of vH1(Ω)v\in H^{1}(\Omega) on Ω\partial\Omega is defined in the sense of

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

provided that ΔvH1(Ω)\Delta v\in H^{1}(\Omega)^{\prime}, where Δ\Delta is taken in the sense of distributions in Ω\Omega. Hence there arises that

φ,vνH1/2(Ω),H1/2(Ω)=(v,φ)L2(Ω)+φ,ΔvH1(Ω),H1(Ω)\left\langle\varphi,\frac{\partial v}{\partial\nu}\right\rangle_{H^{1/2}(\partial\Omega),H^{-1/2}(\partial\Omega)}=(\nabla v,\nabla\varphi)_{L^{2}(\Omega)}+\langle\varphi,\Delta v\rangle_{H^{1}(\Omega),H^{1}(\Omega)^{\prime}}

for any φH01(Ω)\varphi\in H^{1}_{0}(\Omega). Similarly, if ΔvV\Delta v\in V^{\prime}, the normal derivative of vVv\in V on γ1\gamma^{1} is defined in the sense of

v𝝂H1/2(γ1),\frac{\partial v}{\partial\boldsymbol{\nu}}\in H^{-1/2}(\gamma^{1}),

and it holds that

φ,vνH1/2(γ1),H1/2(γ1)=(v,φ)L2(Ω)+φ,ΔvV,V\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}}

for any φV\varphi\in V. Finally, given fVf\in V^{\prime}, φH1/2(γ0)\varphi\in H^{1/2}(\gamma^{0}), and ψH1/2(γ1)\psi\in H^{-1/2}(\gamma^{1}), there is a unique solution zH1(Ω)z\in H^{1}(\Omega) to (2). Hence this zz satisfies

z|γ0=φ\left.z\right|_{\gamma^{0}}=\varphi

and

(z,ζ)=(f,ζ)+ζ,ψV,V,ζV.(\nabla z,\nabla\zeta)=(f,\zeta)+\langle\zeta,\psi\rangle_{V,V^{\prime}},\quad\forall\zeta\in V.

The Green’s function N(x,y)N(x,y) of (2) is now defined by (5)-(6). It satisfies

N(,y)V,ΔN(,y)=δ(y),N𝝂(,y)|γ1=0,yΩ,N(\cdot,y)\in V,\ -\Delta N(\cdot,y)=\delta(\cdot-y),\ \left.\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y)\right|_{\gamma^{1}}=0,\quad\forall y\in\Omega,

where δ(x)\delta(x) stands for the delta function and N𝝂(,y)\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y) on γ1\gamma^{1} is taken as an element in H1/2(γ1)H^{-1/2}(\gamma^{1}). Hence the solution zH1(Ω)z\in H^{1}(\Omega) to (2) for φH1/2(γ0)\varphi\in H^{1/2}(\gamma^{0}) and ψH1/2(γ1)\psi\in H^{-1/2}(\gamma^{1}) admits the representation

z(y)=(N(,y),f)φ,N𝝂(,y)γ0+N(,y),ψγ1,yΩ.z(y)=(N(\cdot,y),f)-\left\langle\varphi,\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y)\right\rangle_{\gamma^{0}}+\langle N(\cdot,y),\psi\rangle_{\gamma^{1}},\quad\forall y\in\Omega. (14)

Here and henceforth, (,)(\ ,\ ) and ,γi\langle\ ,\ \rangle_{\gamma^{i}}, i=0,1i=0,1, denote the inner product in L2(Ω)L^{2}(\Omega) and the paring between H1/2(γi)H^{1/2}(\gamma^{i}) and H1/2(γi)H^{-1/2}(\gamma^{i}), respectively.

We note that the above H1H^{1} theory to (2) is valid even if σγ0γ1\sigma\equiv\gamma^{0}\cap\gamma^{1}\neq\emptyset in (1), provided that the (d1)(d-1) dimensional Hausdorff measure of σ\sigma vanishes.

2.2. Differentiations on Ω\partial\Omega

We continue to suppose that Ωd\Omega\subset\mathbb{R}^{d} is a bounded Lipschitz domain satisfying (1). It follows from Rademacher’s theorem that the tangent space Tx(Ω)T_{x}(\partial\Omega) and the unit outer normal vector 𝝂\boldsymbol{\nu} exist for almost every xΩx\in\partial\Omega. At such xΩx\in\partial\Omega, we take the orthonormal moving frame

{𝐬1,,𝐬d1,𝝂}\displaystyle\{\mathbf{s}_{1},\cdots,\mathbf{s}_{d-1},\boldsymbol{\nu}\}

with positive orientation, where {𝐬1,,𝐬d1}\{\mathbf{s}_{1},\cdots,\mathbf{s}_{d-1}\} is an orthonormal frame (with positive orientation) of the tangent space Tx(Ω)T_{x}(\partial\Omega). Equalities concerning the derivatives of Lipschitz functions below are valid almost everywhere, although it is not mentioned each time.

Let γ=γi\gamma=\gamma^{i}, i=0,1i=0,1, and =x\nabla=\nabla_{x}. If γ\gamma is C1,1C^{1,1}, the above 𝐬1,,𝐬d1,𝝂\mathbf{s}_{1},\cdots,\mathbf{s}_{d-1},\boldsymbol{\nu} are C0,1(γ)C^{0,1}(\gamma). If γ\gamma is C2,1C^{2,1}, which means that it is Lipschitz continuous up to the derivatives of the second order, these vectors 𝐬1,,𝐬d1,𝝂\mathbf{s}_{1},\cdots,\mathbf{s}_{d-1},\boldsymbol{\nu} are in C1,1(γ)C^{1,1}(\gamma). In this case, if ff is a C1,1C^{1,1} function in a neighbourhood of γ\gamma, then we obtain ([8, Lemma 10, Corollary 11]),

2f\displaystyle\nabla^{2}f =\displaystyle= i=1d1(𝐬i)Tfsi+(𝝂)Tfν+i,j=1d1𝐬i𝐬j2fsisj\displaystyle\sum_{i=1}^{d-1}(\nabla\mathbf{s}_{i})^{T}\frac{\partial f}{\partial s_{i}}+(\nabla\boldsymbol{\nu})^{T}\frac{\partial f}{\partial\nu}+\sum_{i,j=1}^{d-1}\mathbf{s}_{i}\otimes\mathbf{s}_{j}\frac{\partial^{2}f}{\partial s_{i}\partial s_{j}} (15)
+i=1d1(𝐬i𝝂+𝝂𝐬i)2fsiν+𝝂𝝂2fν2,\displaystyle+\sum_{i=1}^{d-1}(\mathbf{s}_{i}\otimes\boldsymbol{\nu}+\boldsymbol{\nu}\otimes\mathbf{s}_{i})\frac{\partial^{2}f}{\partial s_{i}\partial\nu}+\boldsymbol{\nu}\otimes\boldsymbol{\nu}\frac{\partial^{2}f}{\partial\nu^{2}},

and

Δf=(𝝂)fν+2fν2+i=1d12fsi2,𝝂=i=1d1κi\Delta f=(\nabla\cdot\boldsymbol{\nu})\frac{\partial f}{\partial\nu}+\frac{\partial^{2}f}{\partial\nu^{2}}+\sum_{i=1}^{d-1}\frac{\partial^{2}f}{\partial s_{i}^{2}},\quad\nabla\cdot\boldsymbol{\nu}=\sum_{i=1}^{d-1}\kappa_{i} (16)

on Γ\Gamma, where 2f\nabla^{2}f is the Hesse matrix of ff,

(𝝂)T\displaystyle(\nabla\boldsymbol{\nu})^{T} =\displaystyle= i=1d1κi𝐬i𝐬i+i=1d1ji(𝝂si𝐬j)𝐬j𝐬i,\displaystyle\sum_{i=1}^{d-1}\kappa_{i}\mathbf{s}_{i}\otimes\mathbf{s}_{i}+\sum_{i=1}^{d-1}\sum_{j\neq i}\left(\frac{\partial\boldsymbol{\nu}}{\partial s_{i}}\cdot\mathbf{s}_{j}\right)\mathbf{s}_{j}\otimes\mathbf{s}_{i},
(𝐬j)T\displaystyle(\nabla\mathbf{s}_{j})^{T} =\displaystyle= κj𝝂𝐬jij(𝝂si𝐬j)𝝂𝐬i,\displaystyle-\kappa_{j}\boldsymbol{\nu}\otimes\mathbf{s}_{j}-\sum_{i\neq j}\left(\frac{\partial\boldsymbol{\nu}}{\partial s_{i}}\cdot\mathbf{s}_{j}\right)\boldsymbol{\nu}\otimes\mathbf{s}_{i}, (17)

and κi\kappa_{i} is the sectional curvature of γ\gamma along 𝐬j\mathbf{s}_{j}, j=1,,d1j=1,\cdots,d-1.

In the general case of the bounded Lipschitz domain Ω\Omega, each γi\gamma^{i}, i=0,1i=0,1, forms a Lipschitz manifold without boundaries. Then the Stokes theorem ensures

γi𝑑ω=0,i=1,2,\int_{\gamma^{i}}d\omega=0,\quad i=1,2, (18)

where ω\omega is a Lipschitz continuous differential form of order d2d-2 and dωd\omega is its exterior derivative.

Let FF and gg be Lipschitz continuous functions on γ1\gamma^{1}, and HH be C1,1C^{1,1} in a neighbourbood of γ1\gamma^{1}. Let, furthermore, ω1(F,H)\omega_{1}(F,H) and ω2(F,g,H)\omega_{2}(F,g,H) be differential forms with order d2d-2 defined by

ω1(F,H)=i=1d1(1)i+1FHd𝐬1dsi^d𝐬d1\displaystyle\omega_{1}(F,H)=\sum_{i=1}^{d-1}(-1)^{i+1}FH\,d\mathbf{s}_{1}\wedge\cdots\wedge\widehat{ds_{i}}\wedge\cdots\wedge d\mathbf{s}_{d-1}
ω2(F,g,H)=i=1d1(1)i+1FgH𝐬id𝐬1dsi^d𝐬d1,\displaystyle\omega_{2}(F,g,H)=\sum_{i=1}^{d-1}(-1)^{i+1}Fg\frac{\partial H}{\partial\mathbf{s}_{i}}\,d\mathbf{s}_{1}\wedge\cdots\wedge\widehat{ds_{i}}\wedge\cdots\wedge d\mathbf{s}_{d-1},

where d𝐬i^\widehat{d\mathbf{s}_{i}} means excluding of d𝐬id\mathbf{s}_{i} and 𝐬i\frac{\partial\;}{\partial\mathbf{s}_{i}} is the directional derivative along 𝐬i\mathbf{s}_{i}. Using (18), we obtain

dω1(F,H), 1γ1=dω2(F,g,H), 1γ1=0,\langle d\omega_{1}(F,H),\,1\rangle_{\gamma^{1}}=\langle d\omega_{2}(F,g,H),\,1\rangle_{\gamma^{1}}=0,

and hence the following lemma. Here, the tangential gradient τ\nabla_{\tau} on Ω\partial\Omega is defined by

τ=i=1d1𝐬i,\nabla_{\tau}=\sum_{i=1}^{d-1}\frac{\partial\;}{\partial\mathbf{s}_{i}}, (19)

which is independent of the choice on the orthonormal coordinate {𝐬1,,𝐬d1}\{\mathbf{s}_{1},\cdots,\mathbf{s}_{d-1}\}.

Lemma 3.

It holds that

i=1N1F𝐬i,Hγ1\displaystyle\sum_{i=1}^{N-1}\left\langle\frac{\partial F}{\partial\mathbf{s}_{i}},\,H\right\rangle_{\gamma^{1}} =\displaystyle= i=1N1F,H𝐬iγ1,\displaystyle-\sum_{i=1}^{N-1}\left\langle F,\,\frac{\partial H}{\partial\mathbf{s}_{i}}\right\rangle_{\gamma^{1}}, (20)
τF,gτHγ1\displaystyle\langle\nabla_{\tau}F,\,g\nabla_{\tau}H\rangle_{\gamma^{1}} =\displaystyle= F,i=1N1𝐬i(gH𝐬i)γ1.\displaystyle-\left\langle F,\,\sum_{i=1}^{N-1}\frac{\partial\;}{\partial\mathbf{s}_{i}}\left(g\frac{\partial H}{\partial\mathbf{s}_{i}}\right)\right\rangle_{\gamma^{1}}. (21)

2.3. Domain perturbations

Given the bounded Lipschitz domain Ωd\Omega\subset{\mathbb{R}}^{d}, let

Tt:ΩΩt=Tt(Ω),|t|1,T0=IT_{t}:\Omega\rightarrow\Omega_{t}=T_{t}(\Omega),\ |t|\ll 1,\quad{T}_{0}=I (22)

be a family of bi-Lipschitz homeomorphisms.

Definition 4.

The family of deformations {Tt}\{T_{t}\} of Ω\Omega in (22) is said to be differentiable if TtxT_{t}x is continuously differentiable in tt for every xΩx\in\Omega and the mappings

tDTt,t(DTt)1:Ωd\frac{\partial}{\partial t}DT_{t},\ \frac{\partial}{\partial t}(DT_{t})^{-1}:\Omega\rightarrow\mathbb{R}^{d}

are uniformly bounded, where DTtDT_{t} denotes the Jacobi matrix of Tt:ΩΩtT_{t}:\Omega\rightarrow\Omega_{t}. It is said to be twice differentiable if TtxT_{t}x is continuously differentiable twice in tt for every xΩx\in\Omega and the mappings

2t2DTt,2t2(DTt)1:Ωd\frac{\partial^{2}}{\partial t^{2}}DT_{t},\ \frac{\partial^{2}}{\partial t^{2}}(DT_{t})^{-1}:\Omega\rightarrow\mathbb{R}^{d}

are uniformly bounded.

The dynamical perturbation (12) is once and twice differentiable if vC0,1(Ω~;d)v\in C^{0,1}(\tilde{\Omega};\mathbb{R}^{d}) and vC1,1(Ω~;d)v\in C^{1,1}(\tilde{\Omega};\mathbb{R}^{d}), respectively, where Ω~\tilde{\Omega} is an open neighbourhood of Ω¯\overline{\Omega}. If {Tt}\{T_{t}\} is, say, twice differentiable, the vector fields SS and RR defined by

S=Ttt|t=0,R=2Ttt2|t=0S=\frac{\partial T_{t}}{\partial t}\bigg{|}_{t=0},\quad R=\frac{\partial^{2}T_{t}}{\partial t^{2}}\bigg{|}_{t=0} (23)

are Lipschitz continuous on Ω¯\overline{\Omega}. Then the family {Tt}\{T_{t}\} admits the Taylor expansion in C0,1(Ω¯)C^{0,1}(\overline{\Omega}),

Tt=I+tS+12t2R+o(t2),t0,T_{t}=I+tS+\frac{1}{2}t^{2}R+o(t^{2}),\quad t\rightarrow 0, (24)

if it is twice differentiable, where II denotes the identity mapping. Then we put

δρ=Ttt|t=0𝝂=S𝝂,δ2ρ=2Ttt2|t=0𝝂=R𝝂,\delta\rho=\frac{\partial T_{t}}{\partial t}\bigg{|}_{t=0}\cdot\boldsymbol{\nu}=S\cdot\boldsymbol{\nu},\quad\delta^{2}\rho=\frac{\partial^{2}T_{t}}{\partial t^{2}}\bigg{|}_{t=0}\cdot\boldsymbol{\nu}=R\cdot\boldsymbol{\nu}, (25)

recalling that 𝝂\boldsymbol{\nu} is the unit outer normal vector on Ω\partial\Omega.

For the normal perturbation, it holds that

S=ρ𝝂on Ω,R=0,S=\rho\boldsymbol{\nu}\ \mbox{on $\partial\Omega$},\quad R=0, (26)

and therefore, δρ\delta\rho in (25) is consistent to that in (11). If {Tt}\{T_{t}\} is a dynamical perturbation defined by (12), then we obtain

S=v,R=(v)vS=v,\quad R=(v\cdot\nabla)v (27)

and hence

δρ=v𝝂,δ2ρ=[(v)v]𝝂.\delta\rho=v\cdot\boldsymbol{\nu},\quad\delta^{2}\rho=[(v\cdot\nabla)v]\cdot\boldsymbol{\nu}.

Careful treatments of the domain perturbation are necessary if γ0γ1\gamma^{0}\cap\gamma^{1}\neq\emptyset is admitted in (1), such as the non-pealing-off condition used in [7]. This case is left in a future study.

2.4. Jacobi matrix and its derivatives

Given the d×dd\times d matrices A=(aij)A=(a_{ij}) and B=(bij)B=(b_{ij}), their inner product A:BA:B and the associated Frobenius norm AF\|A\|_{F} are defined by

A:B=i,j=1daijbij,AF2=A:A=i,j=1daij2.\displaystyle A:B=\sum_{i,j=1}^{d}a_{ij}b_{ij},\quad\|A\|_{F}^{2}=A:A=\sum_{i,j=1}^{d}a_{ij}^{2}.

The Jacobi matrix DTtDT_{t} of the bi-Lipschitz homeomorphism TtT_{t} is defined by

DTt=(Ttixj)i,j=1,,d,\displaystyle DT_{t}=\left(\frac{\partial T_{t}^{i}}{\partial x_{j}}\right)_{i,j=1,\cdots,d},

where

Tt=(Tt1,,Ttd)T.T_{t}=\left(T_{t}^{1},\cdots,T_{t}^{d}\right)^{T}.

We continue to suppose that Ωd\Omega\subset\mathbb{R}^{d} is a bounded Lipschitz domain and {Tt}\{T_{t}\}, |t|1|t|\ll 1, is a family of bi-Lipschitz deformations of Ω\Omega.

Lemma 5.

It holds that

limt0DTt=I,limt0t(DTt)=DS,limt02t2(DTt)=DR\lim_{t\to 0}DT_{t}=I,\quad\lim_{t\to 0}\frac{\partial\;}{\partial t}(DT_{t})=DS,\quad\lim_{t\to 0}\frac{\partial^{2}}{\partial t^{2}}(DT_{t})=DR (28)

uniformly on Ω\Omega if {Tt}\{T_{t}\} is differentiable, and

limt0t(DTt)1=DS,limt02t2(DTt)1=2(DS)2DR\lim_{t\to 0}\frac{\partial\;}{\partial t}(DT_{t})^{-1}=-DS,\quad\lim_{t\to 0}\frac{\partial^{2}\;}{\partial t^{2}}(DT_{t})^{-1}=2(DS)^{2}-DR (29)

uniformly on Ω\Omega if {Tt}\{T_{t}\} is twice differentiable, furthermore. Here, SS and RR are vector fields on Ω\Omega defined by (24), and DSDS and DRDR are the Jacobi matrices of SS and RR, respectively.

Proof.

Convergences below are uniform on Ω\Omega. The limits in (28) are obvious by (24). For (29), we differentiate

I=(DTt)(DTt)1I=(DT_{t})(DT_{t})^{-1} (30)

once and twice with respect to tt, to obtain

O\displaystyle O =\displaystyle= (t(DTt))(DTt)1+(DTt)t(DTt)1,\displaystyle\left(\frac{\partial\;}{\partial t}(DT_{t})\right)(DT_{t})^{-1}+(DT_{t})\frac{\partial\;}{\partial t}(DT_{t})^{-1},
O\displaystyle O =\displaystyle= (2t2(DTt))(DTt)1+2t(DTt)t(DTt)1+(DTt)2t2(DTt)1.\displaystyle\left(\frac{\partial^{2}\;}{\partial t^{2}}(DT_{t})\right)(DT_{t})^{-1}+2\frac{\partial\;}{\partial t}(DT_{t})\frac{\partial\;}{\partial t}(DT_{t})^{-1}+(DT_{t})\frac{\partial^{2}\;}{\partial t^{2}}(DT_{t})^{-1}.

With t0t\to 0, there arises that

O=DS+t(DTt)1|t=0,O=DR+2(DS)(DS)+2t2(DTt)1|t=0O=DS+\left.\frac{\partial\;}{\partial t}(DT_{t})^{-1}\right|_{t=0},\quad O=DR+2(DS)(-DS)+\left.\frac{\partial^{2}\;}{\partial t^{2}}(DT_{t})^{-1}\right|_{t=0}

by

limt0(DTt)1=I\lim_{t\to 0}(DT_{t})^{-1}=I

derived from (30). Then (29) follows. ∎

The following lemma is given in [8, Theorem 6].

Lemma 6.

It holds that

tdetDTt|t=0=S\left.\frac{\partial}{\partial t}\,\mathrm{det}\,DT_{t}\right|_{t=0}=\nabla\cdot S (31)

uniformly on Ω\Omega if {Tt}\{T_{t}\} is differentiable, and

2t2detDTt|t=0=R+(S)2(DS)T:DS\left.\frac{\partial^{2}}{\partial t^{2}}\,\mathrm{det}\,DT_{t}\right|_{t=0}=\nabla\cdot R+(\nabla\cdot S)^{2}-(DS)^{T}:DS (32)

uniformly on Ω\Omega if {Tt}\{T_{t}\} is twice differentiable.

3. Liouville’s formulae

3.1. First formulae

This section is devoted to several Liouville’s formulae concerning volume and area integrals under general perturbation of Lipchitz domains. They are used to derive Hadamard variations associated with the Neumann boundary condition in the following section. The formulae given below are concerned on general domain perturbations. They are new, and have their own interests.

We continue to suppose that Ωd\Omega\subset\mathbb{R}^{d} is a bounded Lipschitz domain and {Tt}\{T_{t}\}, |t|<ε|t|<\varepsilon, is a family of deformations of Ω\Omega. We write ,Ω\langle\ ,\ \rangle_{\partial\Omega} either for the paring between H1/2(Ω)H^{1/2}(\partial\Omega) and H1/2(Ω)H^{-1/2}(\partial\Omega), or, for the inner product in L2(Ω)L^{2}(\partial\Omega). Differentiations in tt of the volume and area integrals below are taken in the classical sense, unless otherwise stated.

The first volume formula follows from (31) and a transformation of variables as in [8, Theorem 1]. Note that δρ\delta\rho defined by (25) is Lipschitz continuous on Ω\partial\Omega. Let Qd+1Q\subset\mathbb{R}^{d+1} be the non-cylindrical domain defined by

Q=|t|<εΩt×{t}.Q=\bigcup_{|t|<\varepsilon}\Omega_{t}\times\{t\}. (33)
Theorem 7 (first volume formula).

If Ωd\Omega\subset\mathbb{R}^{d} is a bounded Lipschitz domain, {Tt}\{T_{t}\} is differentiable, cC0,1(Q¯)c\in C^{0,1}(\overline{Q}), and ctc_{t} is continuous on Q¯\overline{Q}, it holds that

ddtΩtc𝑑x|t=0=Ωc˙𝑑x+c0,δρΩ,\frac{d\;}{dt}\left.\int_{\Omega_{t}}c\ dx\right|_{t=0}=\int_{\Omega}\dot{c}\ dx+\langle c_{0},\,\delta\rho\rangle_{\partial\Omega},

where c0=c(,0)c_{0}=c(\cdot,0) and c˙=ct(,0)\dot{c}=c_{t}(\cdot,0).

Remark 8.

Without the last requirement of ctC(Q¯)c_{t}\in C(\overline{Q}), we have

dIdt=Ωtct𝑑x+Ωtc(Tttνt)𝑑st\frac{dI}{dt}=\int_{\Omega_{t}}c_{t}\ dx+\int_{\partial\Omega_{t}}c\ \left(\frac{\partial T_{t}}{\partial t}\cdot\nu_{t}\right)\ ds_{t} (34)

in the sense of distributions in tt, where dstds_{t} denotes the area element of Ωt\partial\Omega_{t} and

I=Ωtc𝑑x.I=\int_{\Omega_{t}}c\ dx. (35)

In other words, the mapping tI(t)t\mapsto I(t) is locally absolutely continuous, and equality (34) holds for almost every tt. Then the above condition ctC(Q¯)c_{t}\in C(\overline{Q}) is used to take the initial trace at t=0t=0 in (34).

Remark 9.

In [8] the above c=c(x,t)c=c(x,t) is required to be extended outside Q¯\overline{Q}. This extension is always possible within the category of Lipschitz continuous functions because Qd+1Q\subset\mathbb{R}^{d+1} is a Lipschitz domain by the dynamical extension described later. Hence if cC0,1(Q¯)c\in C^{0,1}(\overline{Q}), there is c~C0,1(Q^)\tilde{c}\in C^{0,1}(\hat{Q}) such that c~|Q=c\left.\tilde{c}\right|_{Q}=c, where Q^\hat{Q} is an open set containing Q¯\overline{Q}.

The following form with less regularity of c=c(x,t)c=c(x,t) is used later for computations of the Hadamard variation. Note that Theorem 7 is a direct consequence of this theorem.

Theorem 10.

Let Ωd\Omega\subset\mathbb{R}^{d} be a bounded Lipschitz domain and {Tt}\{T_{t}\} be differentiable. Given cL1(Q)c\in L^{1}(Q) with cL1(Q;d)\nabla c\in L^{1}(Q;\mathbb{R}^{d}), put

b(x,t)=c(Ttx,t),β(x,t)=c(Ttx,t)Ttt(x),b(x,t)=c(T_{t}x,t),\quad\beta(x,t)=\nabla c(T_{t}x,t)\cdot\frac{\partial T_{t}}{\partial t}(x), (36)

and assume bC1(ε,ε;W1,1(Ω))b\in C^{1}(-\varepsilon,\varepsilon;W^{1,1}(\Omega)). Then it holds that

α(x,t)ct(Ttx,t)L1(Q~),Q~=Ω×(ε,ε).\alpha(x,t)\equiv c_{t}(T_{t}x,t)\in L^{1}(\tilde{Q}),\quad\tilde{Q}=\Omega\times(-\varepsilon,\varepsilon).

Assume, furthermore, βC0(ε,ε;L1(Ω))\beta\in C^{0}(-\varepsilon,\varepsilon;L^{1}(\Omega)). Then it holds that αC0(ε,ε;L1(Ω))\alpha\in C^{0}(-\varepsilon,\varepsilon;L^{1}(\Omega)), IC1(ε,ε)I\in C^{1}(-\varepsilon,\varepsilon) for I=I(t)I=I(t) defined by (35), and

dIdt|t=0=Ωc˙+(c0S)dx,\left.\frac{dI}{dt}\right|_{t=0}=\int_{\Omega}\dot{c}+\nabla\cdot(c_{0}S)\ dx, (37)

where c0=c(,0)c_{0}=c(\cdot,0), c˙=ct(,0)\dot{c}=c_{t}(\cdot,0),

Proof.

Since cL1(Q)c\in L^{1}(Q), cL1(Q;d)\nabla c\in L^{1}(Q;\mathbb{R}^{d}), and bC1(ε,ε;W1,1(Ω))b\in C^{1}(-\varepsilon,\varepsilon;W^{1,1}(\Omega)), we obtain βL1(Q~)\beta\in L^{1}(\tilde{Q}) and

bt=α+βin Q~=Ω×(ε,ε)b_{t}=\alpha+\beta\quad\mbox{in $\tilde{Q}=\Omega\times(-\varepsilon,\varepsilon)$} (38)

in the sense of distributions, and hence αL1(Q~)\alpha\in L^{1}(\tilde{Q}). If βC0(ε,ε;L1(Ω))\beta\in C^{0}(-\varepsilon,\varepsilon;L^{1}(\Omega)), furthermore, it holds that αC0(ε,ε;L1(Ω))\alpha\in C^{0}(-\varepsilon,\varepsilon;L^{1}(\Omega)), and therefore, the volume integral on the right-hand of (37) is definite by

ct(Ttx,t),c(Ttx,t)Ttt(x)C0(ε,ε;L1(Ω)).c_{t}(T_{t}x,t),\ \nabla c(T_{t}x,t)\cdot\frac{\partial T_{t}}{\partial t}(x)\ \in C^{0}(-\varepsilon,\varepsilon;L^{1}(\Omega)).

Since

Ωtc𝑑x=Ωb(x,t)detDTt(x)𝑑x\int_{\Omega_{t}}c\ dx=\int_{\Omega}b(x,t)\ \mbox{det}\ DT_{t}(x)\ dx (39)

and {Tt}\{T_{t}\} is differentiable, we have

dIdt\displaystyle\frac{dI}{dt} =\displaystyle= Ωbt+btdetDTtdx=Ωα+β+btdetDTtdx\displaystyle\int_{\Omega}b_{t}+b\frac{\partial}{\partial t}\det DT_{t}\ dx=\int_{\Omega}\alpha+\beta+b\frac{\partial}{\partial t}\det DT_{t}\ dx
=\displaystyle= Ωct(Ttx,t)+c(Ttx,t)Ttt(x)+c(Ttx,t)tdetDTtxdx\displaystyle\int_{\Omega}c_{t}(T_{t}x,t)+\nabla c(T_{t}x,t)\cdot\frac{\partial T_{t}}{\partial t}(x)+c(T_{t}x,t)\frac{\partial}{\partial t}\det DT_{t}x\ dx

with the continuity of the right-hand side in tt. Then the result follows with t=0t=0,

dIdt|t=0=Ωc˙+(c0)S+c0Sdx,\left.\frac{dI}{dt}\right|_{t=0}=\int_{\Omega}\dot{c}+(\nabla c_{0})\cdot S+c_{0}\nabla\cdot S\ dx,

from (31). ∎

Remark 11.

The differentiabilities of bb and cc in tt stand for those of the Lagrange and the Euler ones, respectively. Later in the study of Hadamard variations, we use the fact that the differentiability of the former implies that of the latter by the above theorem.

To formulate Liouville’s area formula, we define a dd-dimensional Lipschitz-manifold in d+1\mathbb{R}^{d+1} by

Γ=|t|<εΩt×{t}.\Gamma=\bigcup_{|t|<\varepsilon}\partial\Omega_{t}\times\{t\}. (40)

Here we employ the method of dynamical extension for the proof of the following lemma.

Lemma 12.

Each fC0,1(Γ)f\in C^{0,1}(\Gamma) is extended as an element in C0,1(Γ~)C^{0,1}(\tilde{\Gamma}), where Γ~\tilde{\Gamma} is an open neighbourhood of Γ\Gamma^{\prime} in d+1\mathbb{R}^{d+1} for

Γ=|t|<εΩt×{t},0<ε<ε.\Gamma^{\prime}=\bigcup_{|t|<\varepsilon^{\prime}}\partial\Omega_{t}\times\{t\},\quad 0<\varepsilon^{\prime}<\varepsilon. (41)
Proof.

We take a smooth vector field v=v(z)v=v(z), z=(x,t)z=(x,t), in d+1\mathbb{R}^{d+1}. Then, we let Z=Z(z,s)Z=Z(z,s) to be the solution to

dZds=v(Z),Z|s=0=zΓ.\frac{dZ}{ds}=v(Z),\quad\left.Z\right|_{s=0}=z\in\Gamma.

There is 0<δ010<\delta_{0}\ll 1, such that

Γ~={Z(z,s)zΓ,|s|<δ0}\tilde{\Gamma}=\{Z(z,s)\mid z\in\Gamma,\ |s|<\delta_{0}\}

forms an open neighbourhood of Γ\Gamma^{\prime}. Furthermore, by the uniqueness of the solution of the ordinary differential equation, the orbits, {𝒪zzΓ}\{{\cal O}_{z}\mid z\in\Gamma\} with 𝒪z={Z(z,s)|s|<δ0}{\mathcal{O}}_{z}=\{Z(z,s)\mid|s|<\delta_{0}\}, do not intersect each other, and form a tubular neighourhood of Γ\Gamma. Given fC0,1(Γ)f\in C^{0,1}(\Gamma), now we put

f~(Z)=f(z),Z=Z(z,s)Γ~,\tilde{f}(Z)=f(z),\quad Z=Z(z,s)\in\tilde{\Gamma},

to obtain f~C0,1(Γ~)\tilde{f}\in C^{0,1}(\tilde{\Gamma}). ∎

Remark 13.

Given 𝐚C0,1(Γ;d)\mathbf{a}\in C^{0,1}(\Gamma;\mathbb{R}^{d}), we thus obtain its extension, denoted by the same symbol, 𝐚C0,1(Γ~;d)\mathbf{a}\in C^{0,1}(\tilde{\Gamma};\mathbb{R}^{d}), which assures 𝐚tL(Γ~;d)\mathbf{a}_{t}\in L^{\infty}(\tilde{\Gamma};\mathbb{R}^{d}). This property, however, does not imply 𝐚t(,t)L(Ωt;d)\mathbf{a}_{t}(\cdot,t)\in L^{\infty}(\partial\Omega_{t};\mathbb{R}^{d}), |t|<ε|t|<\varepsilon, because of the discrepancy of the dimensions of Γ~\tilde{\Gamma} and Ωt\partial\Omega_{t}, the boundary of Ωt\Omega_{t}. Note that 𝐚(,t)L(Ωt)\nabla\cdot\mathbf{a}(\cdot,t)\in L^{\infty}(\partial\Omega_{t}) holds for each tt because 𝐚C0,1(Γ;d)\mathbf{a}\in C^{0,1}(\Gamma;\mathbb{R}^{d}) implies 𝐚(,t)C0,1(Ωt;d)\mathbf{a}(\cdot,t)\in C^{0,1}(\partial\Omega_{t};\mathbb{R}^{d}). Hence we require for both 𝝂𝐚t\boldsymbol{\nu}\cdot\mathbf{a}_{t} and 𝐚\nabla\cdot\mathbf{a} to be continuous on Γ\Gamma, besides 𝐚C0,1(Γ)\mathbf{a}\in C^{0,1}(\Gamma), in Lemma 14 below.

Recalling that dstds_{t} denotes the area element of Ωt\partial\Omega_{t}, we put ds=ds0ds=ds_{0}. The outer unit normal vector 𝝂=𝝂(,t)\boldsymbol{\nu}=\boldsymbol{\nu}(\cdot,t) is defined almost everywhere on Ωt\partial\Omega_{t} for any tt, because Ωt\Omega_{t} is a Lipschitz domain.

Lemma 14.

If Ωd\Omega\subset\mathbb{R}^{d} is a bounded Lipschitz domain, {Tt}\{T_{t}\} is differentiable, 𝐚C0,1(Γ;d)\mathbf{a}\in C^{0,1}(\Gamma;\mathbb{R}^{d}), and both 𝛎𝐚t\boldsymbol{\nu}\cdot\mathbf{a}_{t} and 𝐚\nabla\cdot\mathbf{a} are continuous on Γ\Gamma, it holds that

ddtΩt𝝂𝐚𝑑st|t=0=Ω(𝝂𝐚t)(,0)𝑑s+𝐚(,0),δρΩ.\frac{d\;}{dt}\left.\int_{\partial\Omega_{t}}\boldsymbol{\nu}\cdot\mathbf{a}\ ds_{t}\right|_{t=0}=\int_{\partial\Omega}(\boldsymbol{\nu}\cdot\mathbf{a}_{t})(\cdot,0)\,ds+\langle\nabla\cdot\mathbf{a}(\cdot,0),\delta\rho\rangle_{\partial\Omega}.
Proof.

This lemma is reduced to an identity, valid to any 𝐚W1,1(Γ)\mathbf{a}\in W^{1,1}(\Gamma), that is,

ddtΩt𝝂𝐚𝑑st=Ωt(𝝂𝐚t+(𝐚)Ttt𝝂)𝑑st,\frac{d\;}{dt}\int_{\partial\Omega_{t}}\boldsymbol{\nu}\cdot\mathbf{a}\ ds_{t}=\int_{\partial\Omega_{t}}\biggl{(}\boldsymbol{\nu}\cdot\mathbf{a}_{t}+(\nabla\cdot\mathbf{a})\frac{\partial T_{t}}{\partial t}\cdot\boldsymbol{\nu}\biggr{)}\ ds_{t}, (42)

in the sense of distributions in tt. Hence in (42), the boundary integral of the left-hand side is locally absolutely continuous in t(ε,ε)t\in(-\varepsilon,\varepsilon), and equality holds for almost all tt. Here, the required regularity ensures the Lipschitz continuity of 𝝂𝐚t\boldsymbol{\nu}\cdot\mathbf{a}_{t}, 𝐚\nabla\cdot\mathbf{a}, and Ttt𝝂\frac{\partial T_{t}}{\partial t}\cdot\boldsymbol{\nu} on Γ\Gamma, and also the differentiability in tt of the bi-Lipschitz diffeomorphism Tt:ΩΩtT_{t}:\partial\Omega\rightarrow\partial\Omega_{t}. We thus obtain the continuity of the right-hand side of (42) in tt, and hence the conclusion.

By Lemma 12 there is an extension of 𝐚\mathbf{a}, denoted by the same symbol, such that 𝐚C0,1(Γ~)\mathbf{a}\in C^{0,1}(\tilde{\Gamma}), Then C(Γ~¯)C^{\infty}(\overline{\tilde{\Gamma}}) is dense in W1,1(Γ~)W^{1,1}(\tilde{\Gamma}) by Theorem 1. We may assume, therefore, 𝐚C(Γ¯)\mathbf{a}\in C^{\infty}(\overline{\Gamma}) to verify (42), recalling the notion (13). In this case this differentiation in tt is to be valid in the classical sense, and the equality is to hold for all tt. Below we describe the proof of this fact just for t=0t=0 to make the description simple. Hence we show the lemma for 𝐚=𝐚(x,t)\mathbf{a}=\mathbf{a}(x,t) smooth in Γ~\tilde{\Gamma}, the open neighbourhood of Γ\Gamma^{\prime} in d+1\mathbb{R}^{d+1} defined by (41).

For this purpose, we extend 𝐚\mathbf{a} to a smooth vector field 𝐚~\tilde{\mathbf{a}} on Q¯\overline{Q^{\prime}} for

Q=|t|<εΩ×{t},Q^{\prime}=\bigcup_{|t|<\varepsilon^{\prime}}\Omega\times\{t\},

that is, 𝐚~=𝐚φ\tilde{\mathbf{a}}=\mathbf{a}\varphi, where 0φ=φ(x,t)10\leq\varphi=\varphi(x,t)\leq 1 is a smooth function supported in Γ~\tilde{\Gamma} and is equal to 11 near Γ\Gamma^{\prime}. Thus the proof of this lemma is reduced for smooth 𝐚\mathbf{a}, say 𝐚C1,1(Q¯,d)\mathbf{a}\in C^{1,1}(\overline{Q^{\prime}},\mathbb{R}^{d}).

Now we use the transformation of variables,

𝐛(x,t)=𝐚(y,t),y=Ttx.\mathbf{b}(x,t)=\mathbf{a}(y,t),\quad y=T_{t}x. (43)

It holds that Dx𝐛(x,t)=Dy𝐚(y,t)DTt(x)D_{x}\mathbf{b}(x,t)=D_{y}\mathbf{a}(y,t)DT_{t}(x) and hence

Dy𝐚(y,t)=Dx𝐛(x,t)(DTt(x))1,D_{y}\mathbf{a}(y,t)=D_{x}\mathbf{b}(x,t)(DT_{t}(x))^{-1},

where Dx𝐛D_{x}\mathbf{b} and Dy𝐚D_{y}\mathbf{a} denote the Jacobi matrices of 𝐛\mathbf{b} and 𝐚\mathbf{a} with respect to xx and yy, respectively. Then Green’s formula implies

Ωt(𝝂𝐚)(y,t)𝑑st\displaystyle\int_{\partial\Omega_{t}}(\boldsymbol{\nu}\cdot\mathbf{a})(y,t)\ ds_{t} =\displaystyle= Ωty𝐚(y,t)𝑑y=Ωttr[Dy𝐚(y,t)]𝑑y\displaystyle\int_{\Omega_{t}}\nabla_{y}\cdot\mathbf{a}(y,t)\ dy=\int_{\Omega_{t}}\mathrm{tr}\left[D_{y}\mathbf{a}(y,t)\right]dy (44)
=\displaystyle= Ωtr[Dx𝐛(x,t)(DxTt(x))1]det(DTt)(x)𝑑x,\displaystyle\int_{\Omega}\mathrm{tr}\left[D_{x}\mathbf{b}(x,t)(D_{x}T_{t}(x))^{-1}\right]\mathrm{det}(DT_{t})(x)\ dx,

where trX\mathrm{tr}\,X denotes the trace of the matrix XX. Note that the right-hand side of (44) is differentiable in tt by the above reduction of 𝐚C1,1(Q¯,d)\mathbf{a}\in C^{1,1}(\overline{Q},\mathbb{R}^{d}).

In fact, we have

t{tr[Dx𝐛(,t)(DxTt)1]det(DTt)}\displaystyle\frac{\partial}{\partial t}\left\{\mathrm{tr}\left[D_{x}\mathbf{b}(\cdot,t)(D_{x}T_{t})^{-1}\right]\mathrm{det}(DT_{t})\right\}
=tr[Dx𝐛(,t)(DxTt)1]tdet(DTt)+tr[Dx𝐛(,t)(DxTt)1](det(DTt))t.\displaystyle\quad=\mathrm{tr}\left[D_{x}\mathbf{b}(\cdot,t)(D_{x}T_{t})^{-1}\right]_{t}\mathrm{det}(DT_{t})+\mathrm{tr}\left[D_{x}\mathbf{b}(\cdot,t)(D_{x}T_{t})^{-1}\right](\mathrm{det}(DT_{t}))_{t}. (45)

Then Lemmas 5 and 6 imply

Dx𝐛(x,t)Dx𝐚(x,0),(det(DTt))t(x)(S)(x)D_{x}\mathbf{b}(x,t)\to D_{x}\mathbf{a}(x,0),\quad(\mathrm{det}(DT_{t}))_{t}(x)\to(\nabla\cdot S)(x)

and

Dx𝐛t(x,t)(DTt(x))1+Dx𝐛(x,t)t(DTt(x))1\displaystyle D_{x}\mathbf{b}_{t}(x,t)(DT_{t}(x))^{-1}+D_{x}\mathbf{b}(x,t)\frac{\partial\;}{\partial t}(DT_{t}(x))^{-1}
Dx𝐛t(x,t)|t=0Dx𝐚(x,0)(DS)(x)\displaystyle\quad\to\ D_{x}\mathbf{b}_{t}(x,t)\bigm{|}_{t=0}-D_{x}\mathbf{a}(x,0)(DS)(x)

as t0t\to 0, uniformly in xΩ¯x\in\overline{\Omega}. Here, since

Dx𝐛t(x,t)\displaystyle D_{x}\mathbf{b}_{t}(x,t)
=Dx(Dy𝐚(Ttx,t)Ttt(x)+𝐚t(Ttx,t))\displaystyle\quad=D_{x}\left(D_{y}\mathbf{a}(T_{t}x,t)\frac{\partial T_{t}}{\partial t}(x)+\mathbf{a}_{t}(T_{t}x,t)\right)
=(Dy2𝐚(Ttx,t)DTt(x))Ttt(x)+Dy𝐚(Ttx,t)t(DTt(x))+Dy𝐚t(Ttx,t)DTt(x)\displaystyle\quad=\left(D_{y}^{2}\mathbf{a}(T_{t}x,t)DT_{t}(x)\right)\frac{\partial T_{t}}{\partial t}(x)+D_{y}\mathbf{a}(T_{t}x,t)\frac{\partial\;}{\partial t}(DT_{t}(x))+D_{y}\mathbf{a}_{t}(T_{t}x,t)DT_{t}(x)

it holds that

Dx𝐛t(x,t)|t=0=Dx2𝐚(x,0)S(x)+Dx𝐚(x,0)DS(x)+Dx𝐚t(x,0),D_{x}\mathbf{b}_{t}(x,t)\bigm{|}_{t=0}=D_{x}^{2}\mathbf{a}(x,0)S(x)+D_{x}\mathbf{a}(x,0)DS(x)+D_{x}\mathbf{a}_{t}(x,0), (46)

where Dx2𝐚D_{x}^{2}\mathbf{a} is the third-order tensor consisting of the second derivative of the components of 𝐚\mathbf{a}. Hence there arises that

Dx2𝐚(,0)S=(k2aixjxk(,0)Sk)i,j=1,,dD_{x}^{2}\mathbf{a}(\cdot,0)S=\left(\sum_{k}\frac{\partial^{2}a^{i}}{\partial x_{j}\partial x_{k}}(\cdot,0)S^{k}\right)_{i,j=1,\cdots,d}

for 𝐚=(ai)\mathbf{a}=(a^{i}) and S=(Si)S=(S^{i}).

Gathering these observations, we obtain

ddtΩt(𝝂𝐚)(,t)𝑑st|t=0\displaystyle\frac{d\;}{dt}\int_{\partial\Omega_{t}}(\boldsymbol{\nu}\cdot\mathbf{a})(\cdot,t)\ ds_{t}\biggr{|}_{t=0}
=ddtΩtr[Dx𝐛(x,t)(Dx(Ttx))1]detDTt(x)𝑑x|t=0\displaystyle\quad=\frac{d\;}{dt}\left.\int_{\Omega}\mathrm{tr}\left[D_{x}\mathbf{b}(x,t)(D_{x}(T_{t}x))^{-1}\right]\mathrm{det}\ DT_{t}(x)\ dx\right|_{t=0}
=Ωtr[Dx2𝐚(,0)S+Dx𝐚(,0)DS+Dx𝐚t(,0)]+tr[Dx𝐚(,0)(DS)]\displaystyle\quad=\int_{\Omega}\mathrm{tr}[D_{x}^{2}\mathbf{a}(\cdot,0)S+D_{x}\mathbf{a}(\cdot,0)DS+D_{x}\mathbf{a}_{t}(\cdot,0)]+\mathrm{tr}[D_{x}\mathbf{a}(\cdot,0)(-DS)]
+tr[Dx𝐚(,0)](S)dx\displaystyle\qquad\qquad\qquad+\mathrm{tr}[D_{x}\mathbf{a}(\cdot,0)](\nabla\cdot S)\ dx
=Ωtr[Dx2𝐚(,0)S]+(x𝐚)(,0)(S)+(x𝐚t)(,0)dx\displaystyle\quad=\int_{\Omega}\mathrm{tr}[D_{x}^{2}\mathbf{a}(\cdot,0)S]+(\nabla_{x}\cdot\mathbf{a})(\cdot,0)(\nabla\cdot S)+(\nabla_{x}\cdot\mathbf{a}_{t})(\cdot,0)\ dx
=Ωx[(x𝐚)(,0))S]+(x𝐚t)(,0)dx\displaystyle\quad=\int_{\Omega}\nabla_{x}\cdot[(\nabla_{x}\cdot\mathbf{a})(\cdot,0))S]+(\nabla_{x}\cdot\mathbf{a}_{t})(\cdot,0)\ dx
=Ω(𝝂𝐚t)(,0)+(𝐚)(,0)(S𝝂)ds,\displaystyle\quad=\int_{\partial\Omega}(\boldsymbol{\nu}\cdot\mathbf{a}_{t})(\cdot,0)+(\nabla\cdot\mathbf{a})(\cdot,0)(S\cdot\boldsymbol{\nu})\ ds,

and hence the conclusion. ∎

Remark 15.

Lemma 14 ensures an extension of [8, Lemma 13]. Here, it is not necessary to assume c=𝐚c=\nabla\cdot\mathbf{a} with a vector field 𝐚\mathbf{a}, C2C^{2} in a neighbourhood of Γ\Gamma. The outer unit normal vector 𝝂\boldsymbol{\nu} in the following theorem is in

𝝂C0,1(Γ,Sd1),Sd1={ζd|ζ|=1}\boldsymbol{\nu}\in C^{0,1}(\Gamma,S^{d-1}),\quad S^{d-1}=\{\zeta\in\mathbb{R}^{d}\ \mid|\zeta|=1\}

from the assumption. Then we obtain 𝝂(,t)C0,1(Ωt,Sd1)\boldsymbol{\nu}(\cdot,t)\in C^{0,1}(\partial\Omega_{t},S^{d-1}) and hence (𝝂)(,t)L(Ωt)(\nabla\cdot\boldsymbol{\nu})(\cdot,t)\in L^{\infty}(\partial\Omega_{t}) for each tt, which is equal to the mean curvature of Ωt\partial\Omega_{t}.

Theorem 16 (first area formula).

If Ωd\Omega\subset\mathbb{R}^{d} is a bounded C1,1C^{1,1} domain, {Tt}\{T_{t}\} is differentiable, cC0,1(Γ)c\in C^{0,1}(\Gamma), and both ctc_{t} and c𝛎\frac{\partial c}{\partial\boldsymbol{\nu}} are continuous on Γ\Gamma, it holds that

ddtΩtc𝑑st|t=0=Ωc˙𝑑s+(𝝂)c0+c0𝝂,δρΩ,\frac{d\;}{dt}\left.\int_{\partial\Omega_{t}}c\ ds_{t}\right|_{t=0}=\int_{\partial\Omega}\dot{c}\ ds+\left\langle(\nabla\cdot\boldsymbol{\nu})c_{0}+\frac{\partial c_{0}}{\partial\boldsymbol{\nu}},\delta\rho\right\rangle_{\partial\Omega},

where c0=c(,0)c_{0}=c(\cdot,0) and c˙=ct(,0)\dot{c}=c_{t}(\cdot,0).

Proof.

Lemma 12 ensures an extension of 𝝂C0,1(Γ;Sd1)\boldsymbol{\nu}\in C^{0,1}(\Gamma;S^{d-1}) to

𝝂C0,1(Γ~;Sd1),\boldsymbol{\nu}\in C^{0,1}(\tilde{\Gamma};S^{d-1}),

where Γ~\tilde{\Gamma} is an open neighbourhood of Γ\Gamma^{\prime} in d+1\mathbb{R}^{d+1}. Then we apply Lemma 14 to 𝐚=𝝂cC0,1(Γ)\mathbf{a}=\boldsymbol{\nu}c\in C^{0,1}(\Gamma).

In fact, it holds that

𝝂𝝂t=0on Γ\boldsymbol{\nu}\cdot\boldsymbol{\nu}_{t}=0\quad\mbox{on $\Gamma$} (47)

by |𝝂|2=1|\boldsymbol{\nu}|^{2}=1 in Γ~\tilde{\Gamma}, and therefore, 𝝂𝐚t=ct\boldsymbol{\nu}\cdot\mathbf{a}_{t}=c_{t} is continuous on Γ\Gamma from the assumption. Then we obtain the result by

𝐚=(𝝂)c+𝝂con Γ.\nabla\cdot\mathbf{a}=(\nabla\cdot\boldsymbol{\nu})c+\boldsymbol{\nu}\cdot\nabla c\quad\mbox{on $\Gamma$}.

3.2. Second formulae

We turn to Liouville’s second formulae. In the following formula on the volume integral, the last term vanishes for the dynamical perturbation {Tt}\{T_{t}\} defined by (12), because of

δ2ρ[(S)S]𝝂=(R[SS])𝝂\delta^{2}\rho-[(S\cdot\nabla)S]\cdot\boldsymbol{\nu}=(R-[S\cdot\nabla S])\cdot\boldsymbol{\nu}

and (27).

Theorem 17 (second volume formula).

If Ωd\Omega\subset\mathbb{R}^{d} is a bounded Lipschitz domain, {Tt}\{T_{t}\} is twice differentiable, cC1,1(Q¯)c\in C^{1,1}(\overline{Q}), and cttc_{tt} is continuous on Q¯\overline{Q}, it holds that

d2dt2Ωtc𝑑x|t=0=Ωc¨𝑑x+2c˙+c0S,δρΩ+c0,δ2ρ[(S)S]𝝂Ω,\frac{d^{2}\;}{dt^{2}}\left.\int_{\Omega_{t}}c\ dx\right|_{t=0}=\int_{\Omega}\ddot{c}\ dx+\langle 2\dot{c}+\nabla\cdot c_{0}S,\delta\rho\rangle_{\partial\Omega}+\langle c_{0},\delta^{2}\rho-[(S\cdot\nabla)S]\cdot\boldsymbol{\nu}\rangle_{\partial\Omega},

where c0=c(,0)c_{0}=c(\cdot,0), c˙=ct(,0)\dot{c}=c_{t}(\cdot,0), and c¨=ctt(,0)\ddot{c}=c_{tt}(\cdot,0).

Proof.

The proof is reduced to the case of cC(Q¯)c\in C^{\infty}(\overline{Q}) as in Lemma 14. Then we differentiate the right-hand side of

Ωtc𝑑x=Ωc(Ttx,t)det(DTt(x))𝑑x\int_{\Omega_{t}}c\ dx=\int_{\Omega}c(T_{t}x,t)\,\mathrm{det}(DT_{t}(x))\ dx

twice in tt. It follows that

d2dt2Ωtcdx=Ω[ctt(Ttx,t)+2xct(Ttx,t)tTtx+[x2c(Ttx,t)]((tTtx)2)\displaystyle\frac{d^{2}}{dt^{2}}\int_{\Omega_{t}}c\ dx=\int_{\Omega}\Big{[}c_{tt}(T_{t}x,t)+2\nabla_{x}c_{t}(T_{t}x,t)\cdot\frac{\partial\;}{\partial t}T_{t}x+[\nabla^{2}_{x}c(T_{t}x,t)]\left(\left(\frac{\partial\;}{\partial t}T_{t}x\right)^{2}\right)
+xc(Ttx,t)2t2Ttx]det(DTt(x))+c(Ttx,t)2t2detDTt(x)\displaystyle\quad+\nabla_{x}c(T_{t}x,t)\cdot\frac{\partial^{2}\;}{\partial t^{2}}T_{t}x\Big{]}\mbox{det}(DT_{t}(x))+c(T_{t}x,t)\frac{\partial^{2}\;}{\partial t^{2}}\mathrm{det}DT_{t}(x)
+2[ct(Ttx,t)+xc(Ttx,t)tTtx]tdetDTt(x)dx.\displaystyle\quad+2\Big{[}c_{t}(T_{t}x,t)+\nabla_{x}c(T_{t}x,t)\cdot\frac{\partial\;}{\partial t}T_{t}x\Big{]}\frac{\partial\;}{\partial t}\mathrm{det}DT_{t}(x)\ dx.

Letting t0t\to 0, then we obtain

d2dt2Ωtc𝑑x|t=0=Ω[c¨+2c˙S+c0R+(2c0)(S2)]\displaystyle\left.\frac{d^{2}}{dt^{2}}\int_{\Omega_{t}}c\ dx\right|_{t=0}=\int_{\Omega}\left[\ddot{c}+2\nabla\dot{c}\cdot S+\nabla c_{0}\cdot R+(\nabla^{2}c_{0})(S^{2})\right]
+2(c˙+c0S)(S)+c0(R+(S)2DST:DS)dx\displaystyle\quad+2(\dot{c}+\nabla c_{0}\cdot S)(\nabla\cdot S)+c_{0}(\nabla\cdot R+(\nabla\cdot S)^{2}-DS^{T}:DS)\ dx

by Lemma 6. Since the divergence formula implies

Ωc0R+c0(R)dx=c0,R𝝂Ω=c0,δ2ρΩ\displaystyle\int_{\Omega}\nabla c_{0}\cdot R+c_{0}(\nabla\cdot R)\ dx=\langle c_{0},R\cdot\boldsymbol{\nu}\rangle_{\partial\Omega}=\langle c_{0},\delta^{2}\rho\rangle_{\partial\Omega}
Ωc˙S+c˙(S)dx=c˙,S𝝂Ω=c˙,δρΩ,\displaystyle\int_{\Omega}\nabla\dot{c}\cdot S+\dot{c}(\nabla\cdot S)\ dx=\langle\dot{c},S\cdot\boldsymbol{\nu}\rangle_{\partial\Omega}=\langle\dot{c},\delta\rho\rangle_{\partial\Omega},

it follows that

d2dt2Ωtc𝑑x|t=0=Ωc¨𝑑x+2c˙,δρΩ+c0,δ2ρΩ+X+Y+Z\left.\frac{d^{2}\;}{dt^{2}}\int_{\Omega_{t}}c\ dx\right|_{t=0}=\int_{\Omega}\ddot{c}\ dx+2\langle\dot{c},\delta\rho\rangle_{\partial\Omega}+\langle c_{0},\delta^{2}\rho\rangle_{\partial\Omega}+X+Y+Z

for

X=Ω[2c0](S2)𝑑x\displaystyle X=\int_{\Omega}[\nabla^{2}c_{0}](S^{2})dx
Y=2Ω((c0)S)(S)𝑑x\displaystyle Y=2\int_{\Omega}((\nabla c_{0})\cdot S)(\nabla\cdot S)\ dx
Z=Ωc0((S)2DST:DS)dx.\displaystyle Z=\int_{\Omega}c_{0}((\nabla\cdot S)^{2}-DS^{T}:DS)\ dx.

We simplify the terms XX, YY, and ZZ furthermore.

Let S=(S1,,SN)TS=(S^{1},\cdots,S^{N})^{T} and

cj=cxj,cij=2cxixj,Sji=Sixj,c_{j}=\frac{\partial c}{\partial x_{j}},\quad c_{ij}=\frac{\partial^{2}c}{\partial x_{i}\partial x_{j}},\quad S^{i}_{j}=\frac{\partial S^{i}}{\partial x_{j}},

for simplicitly. First, the divergence formula implies

X+Y\displaystyle X+Y =\displaystyle= i,jΩcijSiSj+2ciSiSjjdx\displaystyle\sum_{i,j}\int_{\Omega}c_{ij}S^{i}S^{j}+2c_{i}S^{i}S_{j}^{j}\ dx
=\displaystyle= cS,S𝝂Ω+i,jΩciSjiSj+ciSiSjjdx\displaystyle\langle\nabla c\cdot S,S\cdot\boldsymbol{\nu}\rangle_{\partial\Omega}+\sum_{i,j}\int_{\Omega}-c_{i}S^{i}_{j}S^{j}+c_{i}S^{i}S_{j}^{j}\ dx

Second, if SC1,1(Q¯)S\in C^{1,1}(\overline{Q}) we obtain

Z\displaystyle Z =\displaystyle= i,jΩc(SiiSjjSijSji)𝑑x\displaystyle\sum_{i,j}\int_{\Omega}c(S_{i}^{i}S_{j}^{j}-S_{i}^{j}S_{j}^{i})\ dx
=\displaystyle= i,jΩc(SiSjjSjSji)νi𝑑s+i,jΩ(cSjj)iSi+(cSji)iSjdx\displaystyle\sum_{i,j}\int_{\partial\Omega}c(S^{i}S^{j}_{j}-S^{j}S^{i}_{j})\nu^{i}\ ds+\sum_{i,j}\int_{\Omega}-(cS^{j}_{j})_{i}S^{i}+(cS^{i}_{j})_{i}S^{j}\ dx
=\displaystyle= c(S),S𝝂Ωc,𝝂[(S)S]Ω+i,jΩciSiSjj+ciSjSjidx,\displaystyle\langle c(\nabla\cdot S),S\cdot\boldsymbol{\nu}\rangle_{\partial\Omega}-\langle c,\boldsymbol{\nu}\cdot[(S\cdot\nabla)S]\rangle_{\partial\Omega}+\sum_{i,j}\int_{\Omega}-c_{i}S^{i}S_{j}^{j}+c_{i}S^{j}S_{j}^{i}\ dx,

for 𝝂=(ν1,,νd)T\boldsymbol{\nu}=(\nu^{1},\cdots,\nu^{d})^{T}, similarly, by

i,jΩcSiSijj𝑑x=i,jΩcSjSiji𝑑x.\sum_{i,j}\int_{\Omega}cS^{i}S^{j}_{ij}\ dx=\sum_{i,j}\int_{\Omega}cS^{j}S^{i}_{ij}\ dx.

Hence it holds that

Z=c(S),S𝝂Ωc,𝝂[(S)S]Ω+i,jΩciSiSjj+ciSjSjidxZ=\langle c(\nabla\cdot S),S\cdot\boldsymbol{\nu}\rangle_{\partial\Omega}-\langle c,\boldsymbol{\nu}\cdot[(S\cdot\nabla)S]\rangle_{\partial\Omega}+\sum_{i,j}\int_{\Omega}-c_{i}S^{i}S_{j}^{j}+c_{i}S^{j}S_{j}^{i}\ dx

for general SC0,1(Q¯)S\in C^{0,1}(\overline{Q}).

Adding these equations, we have

X+Y+Z\displaystyle X+Y+Z =\displaystyle= cS,S𝝂Ω+c(S),S𝝂Ωc,𝝂(S)SΩ\displaystyle\langle\nabla c\cdot S,S\cdot\boldsymbol{\nu}\rangle_{\partial\Omega}+\langle c(\nabla\cdot S),S\cdot\boldsymbol{\nu}\rangle_{\partial\Omega}-\langle c,\,\boldsymbol{\nu}\cdot(S\cdot\nabla)S\rangle_{\partial\Omega}
=\displaystyle= (cS),S𝝂Ωc,𝝂(S)SΩ.\displaystyle\langle\nabla\cdot(c\,S),S\cdot\boldsymbol{\nu}\rangle_{\partial\Omega}-\langle c,\,\boldsymbol{\nu}\cdot(S\cdot\nabla)S\rangle_{\partial\Omega}.

Gathering all equations above, the proof is complete. ∎

The following form with less regularity of cc is applicable to the Hadamard variation as in Theorem 10. The proof is the same and is omitted.

Theorem 18.

Let Ωd\Omega\subset\mathbb{R}^{d} be a bounded Lipschitz domain and {Tt}\{T_{t}\} be twice differentiable. Given c=c(x,t)L1(Q)c=c(x,t)\in L^{1}(Q) with cL1(Q;d)\nabla c\in L^{1}(Q;\mathbb{R}^{d}), suppose

bC2(ε,ε;W1,1(Ω)),βC0(ε,ε;L1(Ω))b\in C^{2}(-\varepsilon,\varepsilon;W^{1,1}(\Omega)),\quad\beta\in C^{0}(-\varepsilon,\varepsilon;L^{1}(\Omega))

for b=b(x,t)b=b(x,t) and β=β(x,t)\beta=\beta(x,t) defined by (36). Then it holds that αC0(ε,ε;L1(Ω))\alpha\in C^{0}(-\varepsilon,\varepsilon;L^{1}(\Omega)), IC2(ε,ε)I\in C^{2}(-\varepsilon,\varepsilon), and

d2dt2I=ΩbttdetDTt+b2t2detDTt+2(α+β)tdetDTtdx\frac{d^{2}}{dt^{2}}I=\int_{\Omega}b_{tt}\ \mbox{det}\ DT_{t}+b\ \frac{\partial^{2}}{\partial t^{2}}\mbox{det}\ DT_{t}+2(\alpha+\beta)\ \frac{\partial}{\partial t}\mbox{det}\ DT_{t}\ dx

for I=I(t)I=I(t) defined by (35).

Liouville’s second area formula is derived from the following lemma.

Lemma 19.

If Ω\Omega is C1,1C^{1,1}, {Tt}\{T_{t}\} is twice differentiable, 𝐚C1,1(Γ;d)\mathbf{a}\in C^{1,1}(\Gamma;\mathbb{R}^{d}), and both 𝛎𝐚tt\boldsymbol{\nu}\cdot\mathbf{a}_{tt} and 𝐚t\nabla\cdot\mathbf{a}_{t} are continuous on Γ\Gamma, it holds that

d2dt2Ωt𝝂𝐚𝑑st|t=0=Ω𝝂𝐚tt(,0)𝑑s+2𝐚t(,0)+[(𝐚(,0))S],δρΩ\displaystyle\frac{d^{2}\;}{dt^{2}}\left.\int_{\partial\Omega_{t}}\boldsymbol{\nu}\cdot\mathbf{a}\ ds_{t}\right|_{t=0}=\int_{\partial\Omega}\boldsymbol{\nu}\cdot\mathbf{a}_{tt}(\cdot,0)\ ds+\langle 2\nabla\cdot\mathbf{a}_{t}(\cdot,0)+\nabla\cdot\left[(\nabla\cdot\mathbf{a}(\cdot,0))S\right],\delta\rho\rangle_{\partial\Omega}
+𝐚(,0),(R(S)S)𝝂Ω.\displaystyle\quad+\langle\nabla\cdot\mathbf{a}(\cdot,0),(R-(S\cdot\nabla)S)\cdot\boldsymbol{\nu}\rangle_{\partial\Omega}.
Proof.

Again, the proof is reduced to the case of 𝐚C(Γ;d)\mathbf{a}\in C^{\infty}(\Gamma;\mathbb{R}^{d}). Then we use the transformation (43),

𝐛(x,t)=𝐚(y,t),y=Ttx,\mathbf{b}(x,t)=\mathbf{a}(y,t),\quad y=T_{t}x,

to reach (44):

Ωt𝝂𝐚(y,t)𝑑sy=Ωtr[Dx𝐛(x,t)(DxTt(x))1]det(DTt)𝑑x.\int_{\partial\Omega_{t}}\boldsymbol{\nu}\cdot\mathbf{a}(y,t)ds_{y}=\int_{\Omega}\mathrm{tr}\left[D_{x}\mathbf{b}(x,t)(D_{x}T_{t}(x))^{-1}\right]\mathrm{det}(DT_{t})dx.

Differentiating the right-hand side twice, here we obtain

2t2tr[Dx𝐛(x,t)(DxTt(x))1]det(DTt(x))=tr[Dx𝐛tt(x,t)(DTt(x))1]det(DTt(x))\displaystyle\frac{\partial^{2}\;}{\partial t^{2}}\ \mathrm{tr}\bigl{[}D_{x}\mathbf{b}(x,t)(D_{x}T_{t}(x))^{-1}\bigr{]}\mathrm{det}(DT_{t}(x))=\mathrm{tr}\left[D_{x}\mathbf{b}_{tt}(x,t)(DT_{t}(x))^{-1}\right]\mathrm{det}(DT_{t}(x))
+tr[Dx𝐛(x,t)[(DTt(x))1]tt]det(DTt(x))+tr[Dx𝐛(x,t)(DTt)1][det(DTt(x))]tt\displaystyle\quad+\mathrm{tr}\left[D_{x}\mathbf{b}(x,t)[(DT_{t}(x))^{-1}]_{tt}\right]\mathrm{det}(DT_{t}(x))+\mathrm{tr}\left[D_{x}\mathbf{b}(x,t)(DT_{t})^{-1}\right][\mathrm{det}(DT_{t}(x))]_{tt}
+2tr[Dx𝐛t(x,t)[(DTt(x))1]t]det(DTt)+tr[Dx𝐛t(x,t)(DTt(x))1][det(DTt(x))]t\displaystyle\quad+2\mathrm{tr}\left[D_{x}\mathbf{b}_{t}(x,t)[(DT_{t}(x))^{-1}]_{t}\right]\mathrm{det}(DT_{t})+\mathrm{tr}\left[D_{x}\mathbf{b}_{t}(x,t)(DT_{t}(x))^{-1}\right][\mathrm{det}(DT_{t}(x))]_{t}
+2tr[Dx𝐛(x,t)[(DTt(x))1]t][det(DTt(x))]t\displaystyle\quad+2\mathrm{tr}\left[D_{x}\mathbf{b}(x,t)[(DT_{t}(x))^{-1}]_{t}\right][\mathrm{det}(DT_{t}(x))]_{t}

Since

Dx𝐛(x,t)=Dy𝐚(Ttx,t)(DTt(x)),D_{x}\mathbf{b}(x,t)=D_{y}\mathbf{a}(T_{t}x,t)(DT_{t}(x)),

there arises that

Dx𝐛tt(x,t)=(Dy3𝐚(Ttx,t)DTt(x))(Ttt(x))2+2(Dy2𝐚(Ttx,t)(DTt(x))t)Ttt(x)\displaystyle D_{x}\mathbf{b}_{tt}(x,t)=\left(D_{y}^{3}\mathbf{a}(T_{t}x,t)DT_{t}(x)\right)\left(\frac{\partial T_{t}}{\partial t}(x)\right)^{2}+2\left(D_{y}^{2}\mathbf{a}(T_{t}x,t)(DT_{t}(x))_{t}\right)\frac{\partial T_{t}}{\partial t}(x)
+2(Dy2𝐚t(Ttx,t)DTt(x))Ttt(x)+2Dy𝐚t(Ttx,t)(DTt(x))t\displaystyle\quad+2\left(D_{y}^{2}\mathbf{a}_{t}(T_{t}x,t)DT_{t}(x)\right)\frac{\partial T_{t}}{\partial t}(x)+2D_{y}\mathbf{a}_{t}(T_{t}x,t)(DT_{t}(x))_{t}
+(Dy2𝐚(Ttx,t)DTt(x))2Ttt2(x)+Dy𝐚(Ttx,t)(DTt(x))tt+Dy𝐚tt(Ttx,t)(DTt(x)),\displaystyle\quad+\left(D_{y}^{2}\mathbf{a}(T_{t}x,t)DT_{t}(x)\right)\frac{\partial^{2}T_{t}}{\partial t^{2}}(x)+D_{y}\mathbf{a}(T_{t}x,t)(DT_{t}(x))_{tt}+D_{y}\mathbf{a}_{tt}(T_{t}x,t)(DT_{t}(x)),

and hence

Dx𝐛tt(,t)Dx3𝐚(,0)S2+2(Dx2𝐚(,0)DS)S+2Dx2𝐚t(,0)S\displaystyle D_{x}\mathbf{b}_{tt}(\cdot,t)\to D_{x}^{3}\mathbf{a}(\cdot,0)S^{2}+2\left(D_{x}^{2}\mathbf{a}(\cdot,0)DS\right)S+2D_{x}^{2}\mathbf{a}_{t}(\cdot,0)S
+2Dx𝐚t(,0)DS+Dx2𝐚(,0)R+Dx𝐚(,0)DR+Dx𝐚tt(,0)\displaystyle\quad+2D_{x}\mathbf{a}_{t}(\cdot,0)DS+D_{x}^{2}\mathbf{a}(\cdot,0)R+D_{x}\mathbf{a}(\cdot,0)DR+D_{x}\mathbf{a}_{tt}(\cdot,0)

as t0t\to 0, uniformly on Ω¯\overline{\Omega}, where Dx3𝐚D_{x}^{3}\mathbf{a} denotes the fourth-order tensor which consists of the third derivatives of the elements of 𝐚\mathbf{a}. Thus it follows that

2t2tr[Dx𝐛(,t)(DTt)1]det(DTt)\displaystyle\frac{\partial^{2}\;}{\partial t^{2}}\ \mathrm{tr}\left[D_{x}\mathbf{b}(\cdot,t)(DT_{t})^{-1}\right]\mathrm{det}(DT_{t})
tr[Dx3𝐚(,0)S2+2(Dx2𝐚(,0)DS)S+2Dx2𝐚t(,0)S]\displaystyle\quad\rightarrow\ \mathrm{tr}\left[D_{x}^{3}\mathbf{a}(\cdot,0)S^{2}+2(D_{x}^{2}\mathbf{a}(\cdot,0)DS)S+2D_{x}^{2}\mathbf{a}_{t}(\cdot,0)S\right]
+tr[2Dx𝐚t(,0)DS+Dx2𝐚(,0)R+Dx𝐚(,0)DR]\displaystyle\qquad+\mathrm{tr}\left[2D_{x}\mathbf{a}_{t}(\cdot,0)DS+D_{x}^{2}\mathbf{a}(\cdot,0)R+D_{x}\mathbf{a}(\cdot,0)DR\right]
+tr[Dx𝐚tt(,0)+Dx𝐚(,0)(2(DS)2DR)]\displaystyle\qquad+\mathrm{tr}\left[D_{x}\mathbf{a}_{tt}(\cdot,0)+D_{x}\mathbf{a}(\cdot,0)(2(DS)^{2}-DR)\right]
+tr[Dx𝐚(,0)](R+(S)2DST:DS)\displaystyle\qquad+\mathrm{tr}\left[D_{x}\mathbf{a}(\cdot,0)\right](\nabla\cdot R+(\nabla\cdot S)^{2}-DS^{T}:DS)
+2tr[(Dx2𝐚(,0)S+Dx𝐚(,0)DS+Dx𝐚t(x,0))(DS)]\displaystyle\qquad+2\mathrm{tr}\left[(D_{x}^{2}\mathbf{a}(\cdot,0)S+D_{x}\mathbf{a}(\cdot,0)DS+D_{x}\mathbf{a}_{t}(x,0))(-DS)\right]
+2tr[Dx2𝐚(,0)S+Dx𝐚(,0)DS+Dx𝐚t(,0)](S)\displaystyle\qquad+2\mathrm{tr}\left[D_{x}^{2}\mathbf{a}(\cdot,0)S+D_{x}\mathbf{a}(\cdot,0)DS+D_{x}\mathbf{a}_{t}(\cdot,0)\right](\nabla\cdot S)
+2tr[Dx𝐚(,0)(DS)](S)\displaystyle\qquad+2\mathrm{tr}\left[D_{x}\mathbf{a}(\cdot,0)(-DS)\right](\nabla\cdot S)
=tr[Dx3𝐚(,0)S2+2Dx2𝐚t(,0)S+Dx2𝐚(,0)R+Dx𝐚tt(,0)]\displaystyle\quad=\mathrm{tr}\left[D_{x}^{3}\mathbf{a}(\cdot,0)S^{2}+2D_{x}^{2}\mathbf{a}_{t}(\cdot,0)S+D_{x}^{2}\mathbf{a}(\cdot,0)R+D_{x}\mathbf{a}_{tt}(\cdot,0)\right]
+tr[Dx𝐚(,0)](R+(S)2DST:DS)\displaystyle\qquad+\mathrm{tr}\left[D_{x}\mathbf{a}(\cdot,0)\right](\nabla\cdot R+(\nabla\cdot S)^{2}-DS^{T}:DS)
+2tr[Dx2𝐚(,0)S+2Dx𝐚t(,0)](S)\displaystyle\qquad+2\mathrm{tr}\left[D_{x}^{2}\mathbf{a}(\cdot,0)S+2D_{x}\mathbf{a}_{t}(\cdot,0)\right](\nabla\cdot S)

as t0t\rightarrow 0, uniformly on Ω¯\overline{\Omega}. Hence we obtain

2t2Ωtr[Dx𝐛(,t)(DTt)1]det(DTt)𝑑x|t=0\displaystyle\left.\frac{\partial^{2}\;}{\partial t^{2}}\int_{\Omega}\mathrm{tr}\left[D_{x}\mathbf{b}(\cdot,t)(DT_{t})^{-1}\right]\mathrm{det}(DT_{t})\ dx\right|_{t=0}
=Ωx𝐚tt(,0)+[(𝐚(,0))R]+2[(𝐚(,0))S]\displaystyle\quad=\int_{\Omega}\nabla_{x}\cdot\mathbf{a}_{tt}(\cdot,0)+\nabla\cdot[(\nabla\cdot\mathbf{a}(\cdot,0))R]+2\nabla\cdot[(\nabla\cdot\mathbf{a}(\cdot,0))S]
+tr[Dx3𝐚(,0)S2]+2tr[Dx𝐚t(,0)](S)\displaystyle\qquad+\mathrm{tr}\left[D_{x}^{3}\mathbf{a}(\cdot,0)S^{2}\right]+2\mathrm{tr}\left[D_{x}\mathbf{a}_{t}(\cdot,0)\right](\nabla\cdot S)
+tr[Dx𝐚(,0)]((S)2DST:DS)dx,\displaystyle\qquad+\mathrm{tr}\left[D_{x}\mathbf{a}(\cdot,0)\right]((\nabla\cdot S)^{2}-DS^{T}:DS)\ dx,

using

tr[Dx2𝐚(,0)R]+tr[Dx𝐚(,0)](R)=[(𝐚(,0))R]\displaystyle\mathrm{tr}\left[D_{x}^{2}\mathbf{a}(\cdot,0)R\right]+\mathrm{tr}\left[D_{x}\mathbf{a}(\cdot,0)\right](\nabla\cdot R)=\nabla\cdot\left[(\nabla\cdot\mathbf{a}(\cdot,0))R\right]
2tr[Dx2𝐚t(,0)S]+tr[Dx𝐚t(,0)](S)=2[(𝐚t(,0))S].\displaystyle 2\mathrm{tr}\left[D_{x}^{2}\mathbf{a}_{t}(\cdot,0)S\right]+\mathrm{tr}\left[D_{x}\mathbf{a}_{t}(\cdot,0)\right](\nabla\cdot S)=2\nabla\cdot\left[(\nabla\cdot\mathbf{a}_{t}(\cdot,0))S\right].

We simplify the the last three terms further by the divergence formula. Wright

𝐚(,0)=(a1,,ad)T,S=(S1,,Sd)T,𝝂=(ν1,,νd)T,\mathbf{a}(\cdot,0)=(a^{1},\cdots,a^{d})^{T},\quad S=(S^{1},\cdots,S^{d})^{T},\quad\boldsymbol{\nu}=(\nu^{1},\cdots,\nu^{d})^{T},

and

fi=fxi,fij=2fxixj,fijk=3fxixjxkf_{i}=\frac{\partial f}{\partial x_{i}},\quad f_{ij}=\frac{\partial^{2}f}{\partial x_{i}\partial x_{j}},\quad f_{ijk}=\frac{\partial^{3}f}{\partial x_{i}\partial x_{j}\partial x_{k}}

for simplicity. Then it follows that

X:\displaystyle X: =\displaystyle= Ωtr[Dx3𝐚(,0)S2]𝑑x=i,p,qΩaipqiSpSq\displaystyle\int_{\Omega}\mathrm{tr}\left[D_{x}^{3}\mathbf{a}(\cdot,0)S^{2}\right]dx=\sum_{i,p,q}\int_{\Omega}a^{i}_{ipq}S^{p}S^{q}
=\displaystyle= i,p,qΩaipi(SqpSq+SpSqq)+i,p,qaipiSp,SqνqΩ,\displaystyle-\sum_{i,p,q}\int_{\Omega}a^{i}_{ip}(S_{q}^{p}S^{q}+S^{p}S_{q}^{q})+\sum_{i,p,q}\langle a^{i}_{ip}S^{p},S^{q}\nu^{q}\rangle_{\partial\Omega},
Y:=2Ωtr[Dx2𝐚(,0)S](S)𝑑x=2i,p,qΩaipiSpSqq,Y:=2\int_{\Omega}\mathrm{tr}\left[D_{x}^{2}\mathbf{a}(\cdot,0)S\right](\nabla\cdot S)\ dx=2\sum_{i,p,q}\int_{\Omega}a^{i}_{ip}S^{p}S_{q}^{q},

and

Z:\displaystyle Z: =\displaystyle= Ωtr[Dx𝐚(,0)]((S)2DST:DS)dx\displaystyle\int_{\Omega}\mathrm{tr}\left[D_{x}\mathbf{a}(\cdot,0)\right]((\nabla\cdot S)^{2}-DS^{T}:DS)\ dx
=\displaystyle= i,p,qΩai(SppSqqSxpqSxqp)𝑑x\displaystyle\sum_{i,p,q}\int_{\Omega}a^{i}(S_{p}^{p}S_{q}^{q}-S_{x_{p}}^{q}S_{x_{q}}^{p})\ dx
=\displaystyle= i,p,qΩ(aipiSpSqq+aiSpSpqq)+(aipiSqSqp+aiiSqSpqp)dx\displaystyle\sum_{i,p,q}\int_{\Omega}-(a^{i}_{ip}S^{p}S_{q}^{q}+a^{i}S^{p}S_{pq}^{q})+(a^{i}_{ip}S^{q}S_{q}^{p}+a^{i}_{i}S^{q}S_{pq}^{p})\ dx
+i,p,q{aiiSqq,SpνpΩaiiSq,SqpνpΩ}\displaystyle+\sum_{i,p,q}\left\{\langle a^{i}_{i}S_{q}^{q},S^{p}\nu^{p}\rangle_{\partial\Omega}-\langle a^{i}_{i}S^{q},S_{q}^{p}\nu^{p}\rangle_{\partial\Omega}\right\}

if SC1,1(Ω¯)S\in C^{1,1}(\overline{\Omega}). Adding these equalities, we obtain

X+Y+Z\displaystyle X+Y+Z =\displaystyle= i,p,q{aipiSp,SqνqΩ+aiSqq,SpνpΩaiiSq,SqpνpΩ}\displaystyle\sum_{i,p,q}\left\{\langle a^{i}_{ip}S^{p},S^{q}\nu^{q}\rangle_{\partial\Omega}+\langle a^{i}S_{q}^{q},S^{p}\nu^{p}\rangle_{\partial\Omega}-\langle a^{i}_{i}S^{q},S_{q}^{p}\nu^{p}\rangle_{\partial\Omega}\right\}
=\displaystyle= [(𝐚(,0))S],S𝝂Ω𝐚(,0),[(S)S]𝝂Ω,\displaystyle\left\langle\nabla\cdot[(\nabla\cdot\mathbf{a}(\cdot,0))S],S\cdot\boldsymbol{\nu}\right\rangle_{\partial\Omega}-\left\langle\nabla\cdot\mathbf{a}(\cdot,0),[(S\cdot\nabla)S]\cdot\boldsymbol{\nu}\right\rangle_{\partial\Omega},

which is valid for SC0,1(Ω¯)S\in C^{0,1}(\overline{\Omega}).

Gathering all equations, we obtain the result by the divergence theorem. ∎

If Ωd\Omega\subset\mathbb{R}^{d} is a bounded C1,1C^{1,1} domain and {Tt}\{T_{t}\} is differentiable, there arises that 𝝂C0,1(Γ;Sd1)\boldsymbol{\nu}\in C^{0,1}(\Gamma;S^{d-1}). We have also 𝐬iC0,1(Γ;Sd1)\mathbf{s}_{i}\in C^{0,1}(\Gamma;S^{d-1}), 1id11\leq i\leq d-1, such that

{𝐬1(,t),,𝐬d1(,t),𝝂(,t)}\{\mathbf{s}_{1}(\cdot,t),\cdots,\mathbf{s}_{d-1}(\cdot,t),\boldsymbol{\nu}(\cdot,t)\}

forms a frame of Ωt\partial\Omega_{t} for each tt. The following lemma ensures again (47) for the case that Ω\Omega is C1,1C^{1,1}.

Lemma 20.

If Ω\Omega is C1,1C^{1,1} and {Tt}\{T_{t}\} is differentiable, it holds that

𝝂t=i=1d1[𝐬i(Ttt𝝂)]𝐬i,a.e. on Γ.\boldsymbol{\nu}_{t}=-\sum_{i=1}^{d-1}\left[\frac{\partial}{\partial\mathbf{s}_{i}}\left(\frac{\partial T_{t}}{\partial t}\cdot\boldsymbol{\nu}\right)\right]\mathbf{s}_{i},\quad\mbox{a.e. on $\Gamma$}.
Proof.

We may fix x0Ωx_{0}\in\Omega and assume that {𝐬1,,𝐬d1,𝝂}\{\mathbf{s}_{1},\cdots,\mathbf{s}_{d-1},\boldsymbol{\nu}\} are differentiable at (x,t)=(x0,0)(x,t)=(x_{0},0) to show the desired equality at this (x0,0)(x_{0},0). Write 𝝂(t)=𝝂(x0,t)\boldsymbol{\nu}(t)=\boldsymbol{\nu}(x_{0},t), 𝐬i=𝐬(x0,0)\mathbf{s}_{i}=\mathbf{s}(x_{0},0), 1id11\leq i\leq d-1, and 𝝂=𝝂(x0,0)\boldsymbol{\nu}=\boldsymbol{\nu}(x_{0},0). We take the exponential mapping aroud x0x_{0}:

ξ1𝐬1++ξd1𝐬d1Tx0(Ω)x(𝝃)Ω,𝝃=(ξ1,,ξd1)d1.\xi_{1}\mathbf{s}_{1}+\cdots+\xi_{d-1}\mathbf{s}_{d-1}\in T_{x_{0}}(\partial\Omega)\ \mapsto\ x(\boldsymbol{\xi})\in\partial\Omega,\quad\boldsymbol{\xi}=(\xi_{1},\cdots,\xi_{d-1})\in\mathbb{R}^{d-1}. (48)

This mapping is defined for |𝝃|1|\boldsymbol{\xi}|\ll 1, and satisfies x(0)=x0x(0)=x_{0}. Furthermore, it is a local C1,1C^{1,1} diffeomorphism, and there arises that

xξi|𝝃=𝟎=𝐬i.\frac{\partial x}{\partial\xi_{i}}\biggm{|}_{\boldsymbol{\xi}=\mathbf{0}}=\mathbf{s}_{i}.

The perturbed boundary Ωt\partial\Omega_{t} around x0(t)=Ttx0x_{0}(t)=T_{t}x_{0} is thus parametrized by 𝝃\boldsymbol{\xi} as Tt(x(𝝃))T_{t}(x(\boldsymbol{\xi})), and furthermore, the tangent space Tx0(t)(Ωt)T_{x_{0}(t)}(\partial\Omega_{t}) is spanned by

{𝐬~1(t),,𝐬~d1(t)},𝐬~i(t)=ξiTt(x(𝝃))|𝝃=0,i=1,,d1,\{\tilde{\mathbf{s}}_{1}(t),\cdots,\tilde{\mathbf{s}}_{d-1}(t)\},\quad\tilde{\mathbf{s}}_{i}(t)=\left.\frac{\partial\;}{\partial\xi_{i}}T_{t}(x(\boldsymbol{\xi}))\right|_{\boldsymbol{\xi}=0},\quad i=1,\cdots,d-1,

although {𝐬~1(t),,𝐬~d1(t),𝝂(t)}\{\tilde{\mathbf{s}}_{1}(t),\cdots,\tilde{\mathbf{s}}_{d-1}(t),\boldsymbol{\nu}(t)\} does not necessarily form a frame at x0(t)=Ttx0Ωtx_{0}(t)=T_{t}x_{0}\in\partial\Omega_{t}.

Since Ω\Omega is C1,1C^{1,1} and {Tt}\{T_{t}\} is differentiable, these vectors are Lipschitz continuous in tt, and it follows that

𝐬i(t)𝝂(t)=0,𝝂(t)𝝂(t)=0,|t|1.\mathbf{s}_{i}(t)\cdot\boldsymbol{\nu}(t)=0,\quad\boldsymbol{\nu}(t)\cdot\boldsymbol{\nu}(t)=0,\quad|t|\ll 1.

Then we obtain

𝝂t|t=0𝐬i+𝝂𝐬it|t=0=𝝂t|t=0𝝂=0,\left.\boldsymbol{\nu}_{t}\right|_{t=0}\cdot\mathbf{s}_{i}+\boldsymbol{\nu}\cdot\left.\frac{\partial\mathbf{s}_{i}}{\partial t}\right|_{t=0}=\left.\boldsymbol{\nu}_{t}\right|_{t=0}\cdot\;\boldsymbol{\nu}=0, (49)

and furthermore,

𝐬it|t=0=2tξiTt(x(𝝃))|t=0,𝝃=𝟎=xξiS(x(ξ)|ξ=0=S𝐬i=𝐬iTtt|t=0\left.\frac{\partial\mathbf{s}_{i}}{\partial t}\right|_{t=0}=\left.\frac{\partial^{2}\;}{\partial t\partial\xi_{i}}T_{t}(x(\boldsymbol{\xi}))\right|_{t=0,\ \boldsymbol{\xi}=\mathbf{0}}=\left.\frac{\partial x}{\partial\xi_{i}}S(x(\xi)\right|_{\xi=0}=\nabla S\cdot\mathbf{s}_{i}=\frac{\partial}{\partial\mathbf{s}_{i}}\left.\frac{\partial T_{t}}{\partial t}\right|_{t=0}

by (24).

Hence it follows that

𝝂t=i=1d1(𝝂𝐬it)𝐬i=i=1d1[𝐬i(Ttt𝝂)]𝐬i\boldsymbol{\nu}_{t}=-\sum_{i=1}^{d-1}\left(\boldsymbol{\nu}\cdot\frac{\partial\mathbf{s}_{i}}{\partial t}\right)\mathbf{s}_{i}=-\sum_{i=1}^{d-1}\left[\frac{\partial}{\partial\mathbf{s}_{i}}\left(\frac{\partial T_{t}}{\partial t}\cdot\boldsymbol{\nu}\right)\right]\mathbf{s}_{i}

at (x,t)=(x0,0)(x,t)=(x_{0},0). ∎

Theorem 21 (second area formula).

If Ω\Omega is C2,1C^{2,1}, {Tt}\{T_{t}\} is twice differentiable, cC1,1(Γ)c\in C^{1,1}(\Gamma), and cttc_{tt} is continuous on Γ\Gamma, it holds that

d2dt2Ωtc𝑑st|t=0=Ωc¨𝑑sc0,|τδρ|2Ω\displaystyle\left.\frac{d^{2}\;}{dt^{2}}\int_{\partial\Omega_{t}}c\ ds_{t}\right|_{t=0}=\int_{\partial\Omega}\ddot{c}\ ds-\left\langle c_{0},|\nabla_{\tau}\delta\rho|^{2}\right\rangle_{\partial\Omega}
+2(Δτδρ)c0+2(𝝂)c˙+[((𝝂)c0+c0𝝂)S],δρΩτ2c0,(δρ)2Ω\displaystyle\quad+\left\langle-2(\Delta_{\tau}\delta\rho)c_{0}+2(\nabla\cdot\boldsymbol{\nu})\dot{c}+\nabla\cdot\left[((\nabla\cdot\boldsymbol{\nu})c_{0}+\frac{\partial c_{0}}{\partial\boldsymbol{\nu}})S\right],\delta\rho\right\rangle_{\partial\Omega}-\langle\nabla_{\tau}^{2}c_{0},(\delta\rho)^{2}\rangle_{\partial\Omega}
+(𝝂)c0+c0𝝂,δ2ρ((S)S)𝝂Ω,\displaystyle\quad+\left\langle(\nabla\cdot\boldsymbol{\nu})c_{0}+\frac{\partial c_{0}}{\partial\boldsymbol{\nu}},\delta^{2}\rho-((S\cdot\nabla)S)\cdot\boldsymbol{\nu}\right\rangle_{\partial\Omega},

where c0=c(,0)c_{0}=c(\cdot,0), c˙=ct(,0)\dot{c}=c_{t}(\cdot,0), c¨=ctt(,0)\ddot{c}=c_{tt}(\cdot,0), and

Δτ=ττ,τ=(s1,,sd1)T.\Delta_{\tau}=\nabla_{\tau}\cdot\nabla_{\tau},\quad\nabla_{\tau}=\left(\frac{\partial}{\partial s_{1}},\cdots,\frac{\partial}{\partial s_{d-1}}\right)^{T}.
Proof.

By the assumption it holds that 𝝂C1,1(Γ;Sd1)\boldsymbol{\nu}\in C^{1,1}(\Gamma;S^{d-1}). Then we apply Lemma 19 to 𝐚=𝝂c\mathbf{a}=\boldsymbol{\nu}c. First, it follows that

𝐚tt=𝝂ttc+2𝝂tct+𝝂ctt,𝐚t=𝝂tc+𝝂ct.\mathbf{a}_{tt}=\boldsymbol{\nu}_{tt}c+2\boldsymbol{\nu}_{t}c_{t}+\boldsymbol{\nu}c_{tt},\quad\mathbf{a}_{t}=\boldsymbol{\nu}_{t}c+\boldsymbol{\nu}c_{t}.

Second, |𝝂|2=1|\boldsymbol{\nu}|^{2}=1 in Γ~\tilde{\Gamma} implies

𝝂𝝂tt=|𝝂t|2a.e. on Γ\boldsymbol{\nu}\cdot\boldsymbol{\nu}_{tt}=-|\boldsymbol{\nu}_{t}|^{2}\quad\mbox{a.e. on $\Gamma$}

as well as (47) everywhere. Then we obtain

𝝂𝐚tt=|𝜶|2c+ctt\boldsymbol{\nu}\cdot\mathbf{a}_{tt}=-|\boldsymbol{\alpha}|^{2}c+c_{tt}

almost everywhere with the continuity of its right-hand side on Γ\Gamma, where

𝜶=(Tt𝐬1𝝂,,Tt𝐬d1𝝂)T.\boldsymbol{\alpha}=\left(\frac{\partial T_{t}}{\partial\mathbf{s}_{1}}\cdot\boldsymbol{\nu},\cdots,\frac{\partial T_{t}}{\partial\mathbf{s}_{d-1}}\cdot\boldsymbol{\nu}\right)^{T}.

It holds also that

𝐚t\displaystyle\nabla\cdot\mathbf{a}_{t} =\displaystyle= [(𝝂)c+𝝂c]t=[(𝝂)c]t+𝝂tc+𝝂ct\displaystyle[(\nabla\cdot\boldsymbol{\nu})c+\boldsymbol{\nu}\cdot\nabla c]_{t}=[(\nabla\cdot\boldsymbol{\nu})c]_{t}+\boldsymbol{\nu}_{t}\cdot\nabla c+\boldsymbol{\nu}\cdot\nabla c_{t}
=\displaystyle= [(𝝂)c]t+i=1d1(S𝐬i𝝂)c𝐬i+ct𝝂\displaystyle[(\nabla\cdot\boldsymbol{\nu})c]_{t}+\sum_{i=1}^{d-1}\left(\frac{\partial S}{\partial\mathbf{s}_{i}}\cdot\boldsymbol{\nu}\right)\frac{\partial c}{\partial\mathbf{s}_{i}}+\frac{\partial c_{t}}{\partial\boldsymbol{\nu}}
=\displaystyle= [(𝝂)c]t+𝜶τc+ct𝝂\displaystyle[(\nabla\cdot\boldsymbol{\nu})c]_{t}+\boldsymbol{\alpha}\cdot\nabla_{\tau}c+\frac{\partial c_{t}}{\partial\boldsymbol{\nu}}

almost everywhere with its right-hand side on Γ\Gamma. Hence Lemma 19 is applicable.

Since Ω\Omega is C2,1C^{2,1} and {Tt}\{T_{t}\} is twice differentiable, 𝝂t\boldsymbol{\nu}_{t} in Lemma 20 is Lipschitz continuous on Γ\Gamma, and it holds that

𝝂t=i=1d12𝐬i2(Ttt𝝂).\nabla\cdot\boldsymbol{\nu}_{t}=-\sum_{i=1}^{d-1}\frac{\partial^{2}}{\partial\mathbf{s}_{i}^{2}}\left(\frac{\partial T_{t}}{\partial t}\cdot\boldsymbol{\nu}\right).

We thus obtain

[(𝝂)c]t\displaystyle\left.[(\nabla\cdot\boldsymbol{\nu})c\right]_{t} =\displaystyle= (𝝂t)c+(𝝂)ct\displaystyle(\nabla\cdot\boldsymbol{\nu}_{t})c+(\nabla\cdot\boldsymbol{\nu})c_{t}
=\displaystyle= (ΔτTtt)c+(𝝂)ct.\displaystyle-\left(\Delta_{\tau}\frac{\partial T_{t}}{\partial t}\right)c+(\nabla\cdot\boldsymbol{\nu})c_{t}.

Then we obtain the result by

𝐚=(𝝂)c+c𝝂,|α|2|t=0=i=1d1(δρsi)2=|τδρ|2,\nabla\cdot\mathbf{a}=(\nabla\cdot\boldsymbol{\nu})c+\frac{\partial c}{\partial\boldsymbol{\nu}},\quad\left.\left|\alpha\right|^{2}\right|_{t=0}=\sum_{i=1}^{d-1}\left(\frac{\partial\delta\rho}{\partial s_{i}}\right)^{2}=|\nabla_{\tau}\delta\rho|^{2},

and

2ατc,Tt𝝂tΩ|t=0=τc0,τ(δρ)2=τ2c0,(δρ)22\left.\left\langle\alpha\cdot\nabla_{\tau}c,\frac{\partial T_{t}\cdot\boldsymbol{\nu}}{\partial t}\right\rangle_{\partial\Omega}\right|_{t=0}=\langle\nabla_{\tau}c_{0},\nabla_{\tau}(\delta\rho)^{2}\rangle=-\langle\nabla_{\tau}^{2}c_{0},(\delta\rho)^{2}\rangle

derived from

𝜶τc=i=1d1Tt𝝂sic𝐬i=τTt𝝂tτc.\boldsymbol{\alpha}\cdot\nabla_{\tau}c=\sum_{i=1}^{d-1}\frac{\partial T_{t}\cdot\boldsymbol{\nu}}{\partial s_{i}}\frac{\partial c}{\partial\mathbf{s}_{i}}=\nabla_{\tau}\frac{\partial T_{t}\cdot\boldsymbol{\nu}}{\partial t}\cdot\nabla_{\tau}c.

4. Hadamard variation of the Green’s function

4.1. First variation

The existence of Hadamard variation (9) on the Green’s function is assured by the method of [8]. Recall that N(x,y,t)N(x,y,t) and N(x,y)N(x,y) denote the Green’s function on Ωt\Omega_{t} and Ω\Omega defined by (4), (7), (8) and (4), (5), (6), using utH1(Ωt)u_{t}\in H^{1}(\Omega_{t}) and uH1(Ω)u\in H^{1}(\Omega), respectively.

Theorem 22.

Let Ωd\Omega\subset\mathbb{R}^{d} be a bounded Lipschitz domain and fix yΩy\in\Omega. Let u=u(,t)H1(Ωt)u=u(\cdot,t)\in H^{1}(\Omega_{t}) be the solution to (8) in Theorem 2. Then, if {Tt}\{T_{t}\} is differentiable, it holds that

vC1(ε,ε;H1(Ω)),v(x,t)=u(Ttx,t).v\in C^{1}(-\varepsilon,\varepsilon;H^{1}(\Omega)),\quad v(x,t)=u(T_{t}x,t). (50)

In particular, the first Hadamard variation in (9),

δN(,y)=ut(,t)|t=0,\delta N(\cdot,y)=\left.\frac{\partial u}{\partial t}(\cdot,t)\right|_{t=0},

exists in the sense of distributions in Ω\Omega. There arises that

ΔδN(,y)=Δu˙=0in Ω,\Delta\delta N(\cdot,y)=\Delta\dot{u}=0\quad\mbox{in $\Omega$}, (51)

and furthermore, δN(,y)L2(Ω)\delta N(\cdot,y)\in L^{2}(\Omega), more precisely,

δN(x,y)=vt(x,t)|t=0(S)u0(x)\delta N(x,y)=\left.\frac{\partial v}{\partial t}(x,t)\right|_{t=0}-(S\cdot\nabla)u_{0}(x) (52)

for SC0,1(Ω¯,d)S\in C^{0,1}(\overline{\Omega},\mathbb{R}^{d}) defined by (23) and u0=u(,0)H1(Ω)u_{0}=u(\cdot,0)\in H^{1}(\Omega).

Proof.

To show (50), we take ψC0(d)\psi\in C_{0}^{\infty}(\mathbb{R}^{d}), 0ψ10\leq\psi\leq 1, such that ψ=0\psi=0 and ψ=1\psi=1 in B(y,r)B(y,r) and dB(y,2r)\mathbb{R}^{d}\setminus B(y,2r), respectively, for 0<r10<r\ll 1. Then Γ~=Γ(y)φ\tilde{\Gamma}=\Gamma(\cdot-y)\varphi is independent of tt, and w=u+Γ~w=u+\tilde{\Gamma} satisfies

Δw=hin Ωt,w=0on γt0,w𝝂=0on γt1-\Delta w=h\ \mbox{in $\Omega_{t}$},\quad w=0\ \mbox{on $\gamma_{t}^{0}$},\quad\frac{\partial w}{\partial\boldsymbol{\nu}}=0\ \mbox{on $\gamma_{t}^{1}$} (53)

for h=ΔΓ~h=-\Delta\tilde{\Gamma}. The Poisson problem (53) takes the weak form

wVt,Ωtwφdx=Ωthφ𝑑x,φVt,w\in V_{t},\quad\int_{\Omega_{t}}\nabla w\cdot\nabla\varphi\ dx=\int_{\Omega_{t}}h\varphi\ dx,\ \forall\varphi\in V_{t},

where Vt={vH1(Ωt)v|γt1=0}V_{t}=\{v\in H^{1}(\Omega_{t})\mid\left.v\right|_{\gamma_{t}^{1}}=0\}. Then we obtain

zC1(ε,ε;H1(Ω)),z(x,t)=w(Ttx,t)z\in C^{1}(-\varepsilon,\varepsilon;H^{1}(\Omega)),\quad z(x,t)=w(T_{t}x,t)

and hence (50) as in [8, Theorem 16], because {Tt}\{T_{t}\} is differentiable.

Having (50), the rest part of this theorem follows as in [8]

If Ω\Omega is C1,1C^{1,1}, furthermore, we have u0H2(Ω)u_{0}\in H^{2}(\Omega) by the elliptic regularity, recalling (1). Then it holds that δN(,y)H1(Ω)\delta N(\cdot,y)\in H^{1}(\Omega) by (50)-(52), and then H1H^{1} theory is applicable to (5) for u=δN(,y)u=\delta N(\cdot,y). We have also N(,y)H2(Ω)N(\cdot,y)\in H^{2}(\Omega), which implies

N(,y)|ΩH1/2(Ω).\left.\nabla N(\cdot,y)\right|_{\partial\Omega}\in H^{1/2}(\partial\Omega). (54)

Thus we have the well-definedness of the right-hand side of the desired identity in the following theorem.

Theorem 23 (first variational formula).

If Ω\Omega is C1,1C^{1,1} and {Tt}\{T_{t}\} is differentiable, it holds that

δN(x,y)=δρN𝝂(,x),N𝝂(,y)γ0δρτN(,x),τN(,y)γ1,x,yΩ,\delta N(x,y)=\left\langle\delta\rho\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,x),\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y)\right\rangle_{\gamma^{0}}-\left\langle\delta\rho\nabla_{\tau}N(\cdot,x),\nabla_{\tau}N(\cdot,y)\right\rangle_{\gamma^{1}},\quad x,y\in\Omega,

where δρ=S𝛎\delta\rho=S\cdot\boldsymbol{\nu} and τ\nabla_{\tau} is the tangential gradient on Ω\partial\Omega defined by (19).

Proof.

Continue to fix yΩy\in\Omega. We have readily confirmed N(,y)H2(Ω)N(\cdot,y)\in H^{2}(\Omega) and u˙=δN(,y)H1(Ω)\dot{u}=\delta N(\cdot,y)\in H^{1}(\Omega). Now we show that u˙=δN(,y)H1(Ω)\dot{u}=\delta N(\cdot,y)\in H^{1}(\Omega) solves

Δu˙=0in Ω,u˙=δρN𝝂(,y)on γ0,u˙𝝂=τ(δρτN(,y))on γ1,\Delta\dot{u}=0\ \mbox{in $\Omega$},\quad\dot{u}=-\delta\rho\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y)\ \mbox{on $\gamma^{0}$},\quad\frac{\partial\dot{u}}{\partial\boldsymbol{\nu}}=\nabla_{\tau}\cdot(\delta\rho\nabla_{\tau}N(\cdot,y))\ \mbox{on $\gamma^{1}$}, (55)

where

τ(δρτN(,y))=i=1d1𝐬i(δρN𝐬iN(,y)).\nabla_{\tau}\cdot(\delta\rho\nabla_{\tau}N(\cdot,y))=\sum_{i=1}^{d-1}\frac{\partial}{\partial\mathbf{s}_{i}}\left(\delta\rho\frac{\partial N}{\partial\mathbf{s}_{i}}N(\cdot,y)\right).

Once (55) is shown, Theorem 2 is applicable to this Poisson equation, because (54) implies

δρN𝝂(,y)H1/2(γ0),τ(δρτN(,y))H1/2(γ1)\delta\rho\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y)\in H^{1/2}(\gamma^{0}),\quad\nabla_{\tau}\cdot(\delta\rho\nabla_{\tau}N(\cdot,y))\in H^{-1/2}(\gamma^{1}) (56)

by δρC0,1(Ω)\delta\rho\in C^{0,1}(\partial\Omega). Then the desired equality follows from the representation formula (14) of the solution u˙=u˙(x)\dot{u}=\dot{u}(x) to (55), because N(x,y)=N(y,x)N(x,y)=N(y,x).

Since the boundary condition of v=δN(,y)v=\delta N(\cdot,y) on γ0\gamma^{0} in (55) is assured by the result in [8] on the Dirichlet boundary condition, we have only to confirm the boundary condition on γ1\gamma^{1} in (56). To this end, we take an open neighbourhood of γ1\gamma^{1}, denoted by Ω~\tilde{\Omega}, satisfying Ω~γ0=\tilde{\Omega}\cap\gamma^{0}=\emptyset and yΩ~y\not\in\tilde{\Omega}.

Let φC0(Ω~)\varphi\in C_{0}^{\infty}(\tilde{\Omega}). Then, for |t|1|t|\ll 1 it holds that

ΩtxN(x,y,t)φ(x)𝑑x=0\int_{\Omega_{t}}\nabla_{x}N(x,y,t)\cdot\nabla\varphi(x)\ dx=0 (57)

by

ΔN(,y,t)=0in Ωt{y},N𝝂(,y,t)=0on γt1,\Delta N(\cdot,y,t)=0\ \mbox{in $\Omega_{t}\setminus\{y\}$},\quad\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y,t)=0\ \mbox{on $\gamma_{t}^{1}$},

where the normal derivative of N(,y,t)N(\cdot,y,t) on γt1\gamma_{t}^{1} belongs to H1/2(γt1)H^{-1/2}(\gamma_{t}^{1}).

We apply Liouville’s first volume formula, Theorem 10, to (57) for

c(x,t)=xN(x,y,t)φ(x),c(x,t)=\nabla_{x}N(x,y,t)\cdot\nabla\varphi(x), (58)

satisfying cL(Q)c\in L^{\infty}(Q) and cL(Q;d)\nabla c\in L^{\infty}(Q;\mathbb{R}^{d}). In fact, we obtain

bC1(ε,ε;L2(Ω)),βC0(ε,ε;L2(Ω))b\in C^{1}(-\varepsilon,\varepsilon;L^{2}(\Omega)),\quad\beta\in C^{0}(-\varepsilon,\varepsilon;L^{2}(\Omega))

for b(x,t)=c(Ttx,t)b(x,t)=c(T_{t}x,t) and β(x,t)=xc(Ttx,t)Ttt(x)\beta(x,t)=\nabla_{x}c(T_{t}x,t)\cdot\frac{\partial T_{t}}{\partial t}(x) by Theorem 22. Then it follows that

αC0(ε,ε;L2(Ω))\alpha\in C^{0}(-\varepsilon,\varepsilon;L^{2}(\Omega))

for α(x,t)=ct(Ttx,x)\alpha(x,t)=c_{t}(T_{t}x,x) from (38).

Hence there holds that

0\displaystyle 0 =\displaystyle= ddtΩtxN(x,y,t)φ(x)𝑑x|t=0\displaystyle\left.\frac{d\;}{dt}\int_{\Omega_{t}}\nabla_{x}N(x,y,t)\cdot\nabla\varphi(x)\ dx\right|_{t=0}
=\displaystyle= Ωc˙+(c0S)dx\displaystyle\int_{\Omega}\dot{c}+\nabla\cdot(c_{0}S)\ dx
=\displaystyle= Ωu˙(x)φ(x)+x(xN(x,y)φ(x)Ttt(x))dx\displaystyle\int_{\Omega}\nabla\dot{u}(x)\cdot\nabla\varphi(x)+\nabla_{x}\cdot\left(\nabla_{x}N(x,y)\cdot\nabla\varphi(x)\frac{\partial T_{t}}{\partial t}(x)\right)\ dx
=\displaystyle= φ,u˙𝝂γ1+γ1δρxN(,y)φds\displaystyle\left\langle\varphi,\frac{\partial\dot{u}}{\partial\boldsymbol{\nu}}\right\rangle_{\gamma^{1}}+\int_{\gamma^{1}}\delta\rho\nabla_{x}N(\cdot,y)\cdot\nabla\varphi\ ds
=\displaystyle= φ,u˙𝝂γ1+γ1δρτN(,y)τφds\displaystyle\left\langle\varphi,\frac{\partial\dot{u}}{\partial\boldsymbol{\nu}}\right\rangle_{\gamma^{1}}+\int_{\gamma^{1}}\delta\rho\nabla_{\tau}N(\cdot,y)\cdot\nabla_{\tau}\varphi\ ds
=\displaystyle= φ,u˙𝝂γ1φ,τ(δρτN(,y))γ1\displaystyle\left\langle\varphi,\frac{\partial\dot{u}}{\partial\boldsymbol{\nu}}\right\rangle_{\gamma^{1}}-\langle\varphi,\,\nabla_{\tau}(\delta\rho\nabla_{\tau}N(\cdot,y))\rangle_{\gamma^{1}}

by φC0(Ω~)\varphi\in C_{0}^{\infty}(\tilde{\Omega}) and (51). Then we obtain

u˙𝝂=τ(δρτN(,y))on γ1\frac{\partial\dot{u}}{\partial\boldsymbol{\nu}}=\nabla_{\tau}\cdot(\delta\rho\nabla_{\tau}N(\cdot,y))\quad\mbox{on $\gamma^{1}$}

as an element in H1/2(γ1)H^{-1/2}(\gamma^{1}), because φC0(Ω~)\varphi\in C_{0}^{\infty}(\tilde{\Omega}) is arbitrary. ∎

4.2. Second variation

We begin with the existence of the second variation δ2N\delta^{2}N in (10) as in Theorem 22.

Theorem 24.

Let Ωd\Omega\subset\mathbb{R}^{d} be a bounded Lipschitz domain, {Tt}\{T_{t}\} be twice differentiable, and uH1(Ωt)u\in H^{1}(\Omega_{t}) be the solution to (8) for yΩy\in\Omega. Then it holds that

vC2(ε,ε;H1(Ω)),v(x,t)=u(Ttx,t).v\in C^{2}(-\varepsilon,\varepsilon;H^{1}(\Omega)),\quad v(x,t)=u(T_{t}x,t). (59)

In particular, u¨=δ2N(,y)\ddot{u}=\delta^{2}N(\cdot,y) in (10) exists in the sense of distributions in Ω\Omega, and there arises that

Δu¨=0in Ω.\Delta\ddot{u}=0\quad\mbox{in $\Omega$}.

If Ω\Omega is C1,1C^{1,1} and C2,1C^{2,1}, furthermore, this u¨\ddot{u} belongs to L2(Ω)L^{2}(\Omega) and H1(Ω)H^{1}(\Omega), respectively.

Proof.

All the results except for the regularity of u¨\ddot{u} follow from the weak form (53) as in [8]. There arises also that

2vt2|t=0=u¨+2Su˙+Ru+[2u]SSH1(Ω).\left.\frac{\partial^{2}v}{\partial t^{2}}\right|_{t=0}=\ddot{u}+2S\cdot\nabla\dot{u}+R\cdot\nabla u+[\nabla^{2}u]S\cdot S\in H^{1}(\Omega). (60)

If Ω\Omega is C1,1C^{1,1} we have uH2(Ω)u\in H^{2}(\Omega) and hence u˙H1(Ω)\dot{u}\in H^{1}(\Omega) by (52). Then u¨L2(Ω)\ddot{u}\in L^{2}(\Omega) follows from (60) and S,RC0,1(Ω¯)S,R\in C^{0,1}(\overline{\Omega}). If Ω\Omega is C2,1C^{2,1} there arises uH3(Ω)u\in H^{3}(\Omega) and hence u˙H2(Ω)\dot{u}\in H^{2}(\Omega) by (55) and (6). Then (60) implies u¨H1(Ω)\ddot{u}\in H^{1}(\Omega), similarly. ∎

Lemma 25.

If Ω\Omega is C2,1C^{2,1} and {Tt}\{T_{t}\} is twice differentiable, u¨=δ2N(,y)H1(Ω)\ddot{u}=\delta^{2}N(\cdot,y)\in H^{1}(\Omega) satisfies

Δu¨=0in Ω,u¨=gon γ0,u¨𝝂=hon γ1\Delta\ddot{u}=0\ \mbox{in $\Omega$},\quad\ddot{u}=g\ \mbox{on $\gamma^{0}$},\quad\frac{\partial\ddot{u}}{\partial\boldsymbol{\nu}}=h\quad\mbox{on $\gamma^{1}$} (61)

for

g=χN𝝂(,y)+2δρu˙𝝂\displaystyle g=-\chi\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y)+2\delta\rho\frac{\partial\dot{u}}{\partial\boldsymbol{\nu}}
h=τ(σNτ(,y))+2τ(δρτu˙),\displaystyle h=\nabla_{\tau}\cdot(\sigma N_{\tau}(\cdot,y))+2\nabla_{\tau}\cdot(\delta\rho\nabla_{\tau}\dot{u}), (62)

where δρ=S𝛎\delta\rho=S\cdot\boldsymbol{\nu} and

χ=δ2ρ+δρδρ𝝂+(𝝂)[S,S](δρ)2𝝂(S)δρ\displaystyle\chi=\delta^{2}\rho+\delta\rho\frac{\partial\delta\rho}{\partial\boldsymbol{\nu}}+(\nabla\boldsymbol{\nu})[S,S]-(\delta\rho)^{2}\nabla\cdot\boldsymbol{\nu}-(S\cdot\nabla)\delta\rho
σ=δ2ρ2(Sττ)δρ+(𝝂)[S,S],\displaystyle\sigma=\delta^{2}\rho-2(S_{\tau}\cdot\nabla_{\tau})\delta\rho+(\nabla\boldsymbol{\nu})[S,S], (63)

where Sτ=S(δρ)𝛎S_{\tau}=S-(\delta\rho)\boldsymbol{\nu}.

Proof.

We have readily obtained

u=N(,y)Γ(y)H3(Ω),u˙=δN(,y)H2(Ω),u¨=δ2N(,y)H1(Ω)u=N(\cdot,y)-\Gamma(\cdot-y)\in H^{3}(\Omega),\ \dot{u}=\delta N(\cdot,y)\in H^{2}(\Omega),\ \ddot{u}=\delta^{2}N(\cdot,y)\in H^{1}(\Omega) (64)

if Ω\Omega is C2,1C^{2,1} and {Tt}\{T_{t}\} is twice differentiable. By the same reason there arises that

χ,σC0,1(Ω)\chi,\sigma\in C^{0,1}(\partial\Omega)

for χ\chi and σ\sigma in (63). Hence it follows that

gH1/2(γ0),hH1/2(γ1)g\in H^{1/2}(\gamma^{0}),\quad h\in H^{-1/2}(\gamma^{1})

for gg and hh defined by (62).

We have confirmed Δu¨=0\Delta\ddot{u}=0 in Ω\Omega in the previous theorem. It is also shown that

u¨=g=χN𝝂(,y)+2δρu˙𝝂on γ0\ddot{u}=g=-\chi\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y)+2\delta\rho\frac{\partial\dot{u}}{\partial\boldsymbol{\nu}}\quad\mbox{on $\gamma^{0}$}

for

χ=(R(S)S)𝝂(δρ)2(𝝂)(S)δρ+(δρ)2𝝂\chi=(R-(S\cdot\nabla)S)\cdot\boldsymbol{\nu}-(\delta\rho)^{2}(\nabla\cdot\boldsymbol{\nu})-(S\cdot\nabla)\delta\rho+\frac{\partial(\delta\rho)^{2}}{\partial\boldsymbol{\nu}} (65)

by [8]. Then there arises the first equality of (63) by R𝝂=δ2ρR\cdot\boldsymbol{\nu}=\delta^{2}\rho,

[(S)S]𝝂\displaystyle[(S\cdot\nabla)S]\cdot\boldsymbol{\nu} =\displaystyle= (S)(S𝝂)[(S)𝝂]S\displaystyle(S\cdot\nabla)(S\cdot\boldsymbol{\nu})-[(S\cdot\nabla)\boldsymbol{\nu}]\cdot S (66)
=\displaystyle= (Sττ)δρ+δρδρνS[(S)𝝂],\displaystyle(S_{\tau}\cdot\nabla_{\tau})\delta\rho+\delta\rho\frac{\partial\delta\rho}{\partial\nu}-S\cdot[(S\cdot\nabla)\boldsymbol{\nu}],

and

S[(S)𝝂]=(𝝂)[S,S].S\cdot[(S\cdot\nabla)\boldsymbol{\nu}]=(\nabla\boldsymbol{\nu})[S,S]. (67)

It thus suffices to ensure

u¨𝝂=hon γ1.\frac{\partial\ddot{u}}{\partial\boldsymbol{\nu}}=h\ \mbox{on $\gamma^{1}$}. (68)

For this purpose we use the open neighbourhood Ω~\tilde{\Omega} of γ1\gamma^{1} in the proof of Theorem 23 satisfying Ω~γ0=\tilde{\Omega}\cap\gamma^{0}=\emptyset and yΩ~y\not\in\tilde{\Omega}. Taking φC0(Ω~)\varphi\in C_{0}^{\infty}(\tilde{\Omega}), we have bC2(ε,ε;L2(Ω))b\in C^{2}(-\varepsilon,\varepsilon;L^{2}(\Omega)) for b=b(x,t)b=b(x,t) defined by (58) and (36), from the proof of [8, Theorem 16]. Hence Theorem 18 is applicable.

Since Ω\Omega is C1,1C^{1,1} it holds that 2cL(Q;d×d)\nabla^{2}c\in L^{\infty}(Q;\mathbb{R}^{d}\times\mathbb{R}^{d}). Then we obtain

btt=γ+2δ+μ+σb_{tt}=\gamma+2\delta+\mu+\sigma

in the sense of distributions in Q~=Ω×(ε,ε)\tilde{Q}=\Omega\times(-\varepsilon,\varepsilon), where

γ(x,t)=ctt(Ttx,t),δ(x,t)=βt(x,t)c(Ttx,t)2Ttt2(x)\displaystyle\gamma(x,t)=c_{tt}(T_{t}x,t),\qquad\qquad\quad\ \delta(x,t)=\beta_{t}(x,t)-\nabla c(T_{t}x,t)\cdot\frac{\partial^{2}T_{t}}{\partial t^{2}}(x)
σ(x,t)=c(Ttx,t)2Ttt2(x),μ(x,t)=2c(Ttx,t)[Ttt(x),Ttt(x)].\displaystyle\sigma(x,t)=\nabla c(T_{t}x,t)\cdot\frac{\partial^{2}T_{t}}{\partial t^{2}}(x),\quad\mu(x,t)=\nabla^{2}c(T_{t}x,t)\left[\frac{\partial T_{t}}{\partial t}(x),\frac{\partial T_{t}}{\partial t}(x)\right].

By (59) we have γ,δ,σC(ε,ε;L2(Ω))\gamma,\delta,\sigma\in C(-\varepsilon,\varepsilon;L^{2}(\Omega)), which implies μC(ε,ε;L2(Ω))\mu\in C(-\varepsilon,\varepsilon;L^{2}(\Omega)).

Put t=0t=0 in the conclusion of Lemma 18 and apply (64). Then we obtain

0\displaystyle 0 =\displaystyle= d2dt2ΩtN(,y,t)φdx|t=0\displaystyle\left.\frac{d^{2}\;}{dt^{2}}\int_{\Omega_{t}}\nabla N(\cdot,y,t)\cdot\nabla\varphi\ dx\right|_{t=0}
=\displaystyle= Ωδ2N(,y)φdx+2φ,δρδN(,y)Ω+[(xN(,y)φ)S],δρΩ\displaystyle\int_{\Omega}\nabla\delta^{2}N(\cdot,y)\cdot\nabla\varphi\ dx+2\langle\nabla\varphi,\,\delta\rho\nabla\delta N(\cdot,y)\rangle_{\partial\Omega}+\left\langle\nabla\cdot\left[(\nabla_{x}N(\cdot,y)\cdot\nabla\varphi)S\right],\,\delta\rho\right\rangle_{\partial\Omega}
+φ,[(R(S)S)𝝂]N(,y)Ω\displaystyle+\left\langle\nabla\varphi,\,[(R-(S\cdot\nabla)S)\cdot\boldsymbol{\nu}]\nabla N(\cdot,y)\right\rangle_{\partial\Omega}

by the proof of Theorem 17.

We examine each term on the right-hand side, recalling φC0(Ω~)\varphi\in C_{0}^{\infty}(\tilde{\Omega}). First, it follows that

Ωδ2N(,y)φdx=φ,𝝂δ2N(,y)γ1\int_{\Omega}\nabla\delta^{2}N(\cdot,y)\cdot\nabla\varphi\ dx=\left\langle\varphi,\,\frac{\partial\;}{\partial\boldsymbol{\nu}}\delta^{2}N(\cdot,y)\right\rangle_{\gamma^{1}}

from Δu¨=0\Delta\ddot{u}=0. Second, we have

φ,N(,y)(R(S)S)𝝂Ω\displaystyle\langle\nabla\varphi,\,\nabla N(\cdot,y)(R-(S\cdot\nabla)S)\cdot\boldsymbol{\nu}\rangle_{\partial\Omega} =\displaystyle= τφ,[(R(S)S)𝝂]τN(,y)γ1\displaystyle\langle\nabla_{\tau}\varphi,[(R-(S\cdot\nabla)S)\cdot\boldsymbol{\nu}]\nabla_{\tau}N(\cdot,y)\rangle_{\gamma^{1}}
=\displaystyle= φ,τ([(R(S)S)𝝂]τN(,y))γ1.\displaystyle-\langle\varphi,\,\nabla_{\tau}\cdot([(R-(S\cdot\nabla)S)\cdot\boldsymbol{\nu}]\nabla_{\tau}N(\cdot,y))\,\rangle_{\gamma^{1}}.

by (21) for F=φF=\varphi, g=(R(S)S)𝝂g=(R-(S\cdot\nabla)S)\cdot\boldsymbol{\nu}, and H=N(,y)H=N(\cdot,y).

Third, there arises that

[(N(,y)φ)S],δρΩ\displaystyle\left\langle\nabla\cdot[(\nabla N(\cdot,y)\cdot\nabla\varphi)S],\,\delta\rho\right\rangle_{\partial\Omega}
=[i=1d1𝐬i([S𝐬i]N(,y)φ)+𝝂([S𝝂]N(,y)φ)],S𝝂γ1\displaystyle\quad=\left\langle\left[\sum_{i=1}^{d-1}\frac{\partial\;}{\partial\mathbf{s}_{i}}([S\cdot\mathbf{s}_{i}]\nabla N(\cdot,y)\cdot\nabla\varphi)+\frac{\partial\;}{\partial\boldsymbol{\nu}}([S\cdot\boldsymbol{\nu}]\nabla N(\cdot,y)\cdot\nabla\varphi)\right],\,S\cdot\boldsymbol{\nu}\right\rangle_{\gamma^{1}}
=N(,y)φ,(S𝝂)(S𝝂)𝝂i=1d1(S𝐬i)(S𝝂)𝐬iγ1\displaystyle\quad=\left\langle\nabla N(\cdot,y)\cdot\nabla\varphi,\,(S\cdot\boldsymbol{\nu})\frac{\partial(S\cdot\boldsymbol{\nu})}{\partial\boldsymbol{\nu}}-\sum_{i=1}^{d-1}(S\cdot\mathbf{s}_{i})\frac{\partial(S\cdot\boldsymbol{\nu})}{\partial\mathbf{s}_{i}}\right\rangle_{\gamma^{1}}
+𝝂(N(,y)φ),(S𝝂)2γ1\displaystyle\qquad+\left\langle\frac{\partial\;}{\partial\boldsymbol{\nu}}(\nabla N(\cdot,y)\cdot\nabla\varphi),\,(S\cdot\boldsymbol{\nu})^{2}\right\rangle_{\gamma^{1}}
=τN(,y)τφ,δρδρ𝝂(Sτ)δργ1+i=1d12N𝝂2φ𝝂+N𝐬i2φ𝐬i𝝂,(δρ)2γ1\displaystyle\quad=\left\langle\nabla_{\tau}N(\cdot,y)\cdot\nabla_{\tau}\varphi,\,\delta\rho\frac{\partial\delta\rho}{\partial\boldsymbol{\nu}}-(S\cdot\nabla_{\tau})\delta\rho\right\rangle_{\gamma^{1}}+\sum_{i=1}^{d-1}\left\langle\frac{\partial^{2}N}{\partial\boldsymbol{\nu}^{2}}\frac{\partial\varphi}{\partial\boldsymbol{\nu}}+\frac{\partial N}{\partial\mathbf{s}_{i}}\frac{\partial^{2}\varphi}{\partial\mathbf{s}_{i}\partial\boldsymbol{\nu}},(\delta\rho)^{2}\right\rangle_{\gamma^{1}}

by (20) for F=N(,y)φF=N(\cdot,y)\cdot\nabla\varphi and H=S𝝂H=S\cdot\boldsymbol{\nu}. Here we have

τN(,y)τφ,δρδρ𝝂(Sτ)δργ1\displaystyle\left\langle\nabla_{\tau}N(\cdot,y)\cdot\nabla_{\tau}\varphi,\,\delta\rho\frac{\partial\delta\rho}{\partial\boldsymbol{\nu}}-(S\cdot\nabla_{\tau})\delta\rho\right\rangle_{\gamma^{1}}
=φ,τ[(δρδρ𝝂(Sτ)δρ)τN(,y)]γ1\displaystyle\quad=-\left\langle\varphi,\,\nabla_{\tau}\cdot[(\delta\rho\frac{\partial\delta\rho}{\partial\boldsymbol{\nu}}-(S\cdot\nabla_{\tau})\delta\rho)\nabla_{\tau}N(\cdot,y)]\right\rangle_{\gamma^{1}}

and also

i=1d1(2N𝝂2φ𝝂+N𝐬i2φ𝐬i𝝂),(δρ)2γ1\displaystyle\left\langle\sum_{i=1}^{d-1}\left(\frac{\partial^{2}N}{\partial\boldsymbol{\nu}^{2}}\frac{\partial\varphi}{\partial\boldsymbol{\nu}}+\frac{\partial N}{\partial\mathbf{s}_{i}}\frac{\partial^{2}\varphi}{\partial\mathbf{s}_{i}\partial\boldsymbol{\nu}}\right),\,(\delta\rho)^{2}\right\rangle_{\gamma^{1}}
=φ𝝂,(δρ)2(τ)2N(,y)γ1φν,τ(δρ)2τN(,y)\displaystyle=-\left\langle\frac{\partial\varphi}{\partial\boldsymbol{\nu}},\,(\delta\rho)^{2}(\nabla_{\tau})^{2}N(\cdot,y)\right\rangle_{\gamma^{1}}-\left\langle\frac{\partial\varphi}{\partial\nu},\nabla_{\tau}(\delta\rho)^{2}\cdot\nabla_{\tau}N(\cdot,y)\right\rangle
=2φ𝝂,(δρ)2(τ)2N(,y)γ1φ𝝂,τ(δρ)2τN(,y)γ1\displaystyle=-2\left\langle\frac{\partial\varphi}{\partial\boldsymbol{\nu}},(\delta\rho)^{2}(\nabla_{\tau})^{2}N(\cdot,y)\right\rangle_{\gamma^{1}}-\left\langle\frac{\partial\varphi}{\partial\boldsymbol{\nu}},\,\nabla_{\tau}(\delta\rho)^{2}\cdot\nabla_{\tau}N(\cdot,y)\right\rangle_{\gamma^{1}}

by

N𝝂(,y)=2N𝐬i𝝂(,y)=0,i=1d12N𝐬i2(,y)+2N𝝂2(,y)=0on γ1.\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y)=\frac{\partial^{2}N}{\partial\mathbf{s}_{i}\partial\boldsymbol{\nu}}(\cdot,y)=0,\ \sum_{i=1}^{d-1}\frac{\partial^{2}N}{\partial\mathbf{s}_{i}^{2}}(\cdot,y)+\frac{\partial^{2}N}{\partial\boldsymbol{\nu}^{2}}(\cdot,y)=0\quad\mbox{on $\gamma^{1}$}.

Finally, we notice (55) to deduce

2φ,δρδN(,y)γ1\displaystyle 2\langle\nabla\varphi,\,\delta\rho\nabla\delta N(\cdot,y)\rangle_{\gamma^{1}}
=2φ𝝂,δρδN𝝂(,y)γ1+2τφ,δρτδN(,y)γ1\displaystyle\quad=2\left\langle\frac{\partial\varphi}{\partial\boldsymbol{\nu}},\,\delta\rho\frac{\partial\delta N}{\partial\boldsymbol{\nu}}(\cdot,y)\right\rangle_{\gamma^{1}}+2\left\langle\nabla_{\tau}\varphi,\delta\rho\nabla_{\tau}\delta N(\cdot,y)\right\rangle_{\gamma^{1}}
=2φ𝝂,δρτ(δρ)τN(,y)γ1+2φ𝝂,(δρ)2τ2N(,y)γ1\displaystyle\quad=2\left\langle\frac{\partial\varphi}{\partial\boldsymbol{\nu}},\,\delta\rho\nabla_{\tau}(\delta\rho)\cdot\nabla_{\tau}N(\cdot,y)\right\rangle_{\gamma^{1}}+2\left\langle\frac{\partial\varphi}{\partial\boldsymbol{\nu}},\,(\delta\rho)^{2}\nabla_{\tau}^{2}N(\cdot,y)\right\rangle_{\gamma^{1}}
2φ,τ(δρτδN(,y))γ1.\displaystyle\qquad-2\left\langle\varphi,\,\nabla_{\tau}\cdot(\delta\rho\nabla_{\tau}\delta N(\cdot,y))\right\rangle_{\gamma^{1}}.

Gathering these equalities, we obtain

0=φ,𝝂δ2N(,y)γ1φ,τ[(δρδρ𝝂(Sτ)δρ)τN(,y)]γ1\displaystyle 0=\left\langle\varphi,\,\frac{\partial\;}{\partial\boldsymbol{\nu}}\delta^{2}N(\cdot,y)\right\rangle_{\gamma^{1}}-\left\langle\varphi,\,\nabla_{\tau}\cdot\left[\left(\delta\rho\frac{\partial\delta\rho}{\partial\boldsymbol{\nu}}-(S\cdot\nabla_{\tau})\delta\rho\right)\nabla_{\tau}N(\cdot,y)\right]\right\rangle_{\gamma^{1}}
φ,τ([(R(S)S)𝝂]τN(,y))γ12φ,τ(δρτδN(,y))γ1\displaystyle\quad-\langle\varphi,\nabla_{\tau}\cdot([(R-(S\cdot\nabla)S)\cdot\boldsymbol{\nu}]\nabla_{\tau}N(\cdot,y))\rangle_{\gamma^{1}}-2\langle\varphi,\nabla_{\tau}\cdot(\delta\rho\nabla_{\tau}\delta N(\cdot,y))\rangle_{\gamma^{1}}

and hence the result because φC0(Ω~)\varphi\in C_{0}^{\infty}(\tilde{\Omega}) is arbitrary. In fact, we obtain (62) for

σ=δρδρ𝝂(Sττ)δρ+(R(S)S)𝝂\sigma=\delta\rho\frac{\partial\delta\rho}{\partial\boldsymbol{\nu}}-(S_{\tau}\cdot\nabla_{\tau})\delta\rho+(R-(S\cdot\nabla)S)\cdot\boldsymbol{\nu} (69)

and then, the second equality of (63) follows from R𝝂=δ2ρR\cdot\boldsymbol{\nu}=\delta^{2}\rho, (66), and (67). ∎

Theorem 26 (second variational formula).

If Ω\Omega is C2,1C^{2,1} and {Tt}\{T_{t}\} is twice differentiable, it holds that

δ2N(x,y)=2(δN(,x),δN(,y))+χN𝝂(,x),N𝝂(,y)γ0\displaystyle\delta^{2}N(x,y)=-2(\nabla\delta N(\cdot,x),\nabla\delta N(\cdot,y))+\left\langle\chi\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,x),\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y)\right\rangle_{\gamma^{0}}
στN(,x),τN(,y)γ1\displaystyle\quad-\left\langle\sigma\nabla_{\tau}N(\cdot,x),\nabla_{\tau}N(\cdot,y)\right\rangle_{\gamma^{1}}

for x,yΩx,y\in\Omega, where (,)(\ ,\ ) denotes the inner product in L2(Ω)L^{2}(\Omega).

Proof.

Form Lemma 25 and the representation formula (14), it follows that

δ2N(x,y)=u¨(x)=g,N𝝂(,x)γ0+N(,x),hγ1\delta^{2}N(x,y)=\ddot{u}(x)=-\left\langle g,\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,x)\right\rangle_{\gamma^{0}}+\langle N(\cdot,x),h\rangle_{\gamma^{1}} (70)

for gg, hh defined by (62)-(63), where x,yΩx,y\in\Omega.

By (55), we obtain

0\displaystyle 0 =\displaystyle= δN𝝂(,x),δN(,y)+δρN𝝂(,y)γ0\displaystyle\left\langle\frac{\partial\delta N}{\partial\boldsymbol{\nu}}(\cdot,x),\,\delta N(\cdot,y)+\delta\rho\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y)\right\rangle_{\gamma^{0}}
+δN(,y),δN𝝂(,x)τ(δρτN(,x))γ1\displaystyle+\left\langle\delta N(\cdot,y),\,\frac{\partial\delta N}{\partial\boldsymbol{\nu}}(\cdot,x)-\nabla_{\tau}\cdot(\delta\rho\nabla_{\tau}N(\cdot,x))\right\rangle_{\gamma^{1}}
=\displaystyle= δN(,y),δN𝝂(,x)Ω+δN𝝂(,x),δρN𝝂(,y)γ0\displaystyle\left\langle\delta N(\cdot,y),\,\frac{\partial\delta N}{\partial\boldsymbol{\nu}}(\cdot,x)\right\rangle_{\partial\Omega}+\left\langle\frac{\partial\delta N}{\partial\boldsymbol{\nu}}(\cdot,x),\delta\rho\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y)\right\rangle_{\gamma^{0}}
+τδN(,y),δρτN(,x)γ1\displaystyle+\left\langle\nabla_{\tau}\delta N(\cdot,y),\delta\rho\nabla_{\tau}N(\cdot,x)\right\rangle_{\gamma^{1}}
=\displaystyle= (δN(,y),δN(,x))+δN𝝂(,x),δρN𝝂(,y)γ0\displaystyle(\nabla\delta N(\cdot,y),\nabla\delta N(\cdot,x))+\left\langle\frac{\partial\delta N}{\partial\boldsymbol{\nu}}(\cdot,x),\delta\rho\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y)\right\rangle_{\gamma^{0}}
N(,x),τ(δρτδN(,y))γ1\displaystyle-\langle N(\cdot,x),\nabla_{\tau}\cdot(\delta\rho\nabla_{\tau}\delta N(\cdot,y))\rangle_{\gamma^{1}}

for xΩx\in\Omega, and hence

(δN(,x),δN(,y))\displaystyle(\nabla\delta N(\cdot,x),\nabla\delta N(\cdot,y)) =\displaystyle= δρN𝝂(,y),δN𝝂(,x)γ0\displaystyle-\left\langle\delta\rho\frac{\partial N}{\partial\boldsymbol{\nu}}(\cdot,y),\frac{\partial\delta N}{\partial\boldsymbol{\nu}}(\cdot,x)\right\rangle_{\gamma^{0}} (71)
+N(,x),τ(δρτδN(,y))γ1.\displaystyle+\langle N(\cdot,x),\nabla_{\tau}\cdot(\delta\rho\nabla_{\tau}\delta N(\cdot,y))\rangle_{\gamma^{1}}.

Then the result follows from (70)-(71) as

δ2N(x,y)χNν(,x),Nν(,y)γ0+στN(,x),τN(,y)γ1\displaystyle\delta^{2}N(x,y)-\left\langle\chi\frac{\partial N}{\partial\nu}(\cdot,x),\frac{\partial N}{\partial\nu}(\cdot,y)\right\rangle_{\gamma^{0}}+\langle\sigma\nabla_{\tau}N(\cdot,x),\nabla_{\tau}N(\cdot,y)\rangle_{\gamma^{1}}
=2δρNν(,x),Nν(,y)γ0+2N(,x),τ(δρτδN(,y)γ1\displaystyle\quad=-2\left\langle\delta\rho\frac{\partial N}{\partial\nu}(\cdot,x),\frac{\partial N}{\partial\nu}(\cdot,y)\right\rangle_{\gamma^{0}}+2\langle N(\cdot,x),\nabla_{\tau}(\delta\rho\nabla_{\tau}\delta N(\cdot,y)\rangle_{\gamma^{1}}
=2(δN(,x),δN(,y)).\displaystyle\quad=-2(\nabla\delta N(\cdot,x),\nabla\delta N(\cdot,y)).

Appendix A Liouville’s area formula and differential forms

There is an alternative argument for the proof of Lemmas 14 and 19 using differential forms. Here we assume that Ωd\Omega\subset\mathbb{R}^{d} is a bounded Lipschitz domain and suppose that the family of domain perturbations {Tt}\{T_{t}\} is twice differentiable.

Put D=ΩtD=\Omega_{t}, and let Λp(D)\Lambda^{p}(D), 1pd1\leq p\leq d, be the set of pp-forms on DD. The outer derivative and wedge product are denoted by dd and \wedge, respectively. Given

α=iαidxi,β=iβidxiΛ1(D),x=(xi),\alpha=\sum_{i}\alpha^{i}dx^{i},\ \beta=\sum_{i}\beta^{i}dx^{i}\in\Lambda^{1}(D),\quad x=(x^{i}),

let

(α,β)=iαiβi.(\alpha,\beta)=\sum_{i}\alpha^{i}\beta^{i}.

Given

λ=α1αp,μ=β1βpLp(D),\lambda=\alpha^{1}\wedge\cdots\wedge\alpha^{p},\ \mu=\beta^{1}\wedge\cdots\wedge\beta^{p}\in L^{p}(D),

we put

(λ,μ)=det((αi,βj))ij,(\lambda,\mu)=\mbox{det}\ ((\alpha^{i},\beta^{j}))_{ij},

which is independent of the choice of αi,βiΛ1(D)\alpha^{i},\beta^{i}\in\Lambda^{1}(D), 1ip1\leq i\leq p, to represent α,βΛp(D)\alpha,\beta\in\Lambda^{p}(D). Then the Hodge operator :Λp(D)Λdp(D)\ast:\Lambda^{p}(D)\rightarrow\Lambda^{d-p}(D) is defined by

ωτ=(ω,τ)dx1dxd,ωΛp(D),τΛdp(D).\omega\wedge\tau=(\ast\omega,\tau)dx^{1}\wedge\cdots\wedge dx^{d},\quad\omega\in\Lambda^{p}(D),\ \tau\in\Lambda^{d-p}(D).

It holds that

(dxj1dxjp)=sgnσdxjp+1dxjd,\ast(dx^{j_{1}}\wedge\cdots\wedge dx^{j_{p}})=\mbox{sgn}\ \sigma\cdot dx^{j_{p+1}}\wedge\cdots\wedge dx^{j_{d}},

where σ:(1,,d)(j1,,jd)\sigma:(1,\cdots,d)\mapsto(j_{1},\cdots,j_{d}). By this definition, there arises [3] that

𝝂ds=(dx1,,dxn),\boldsymbol{\nu}ds=(\ast dx_{1},\cdots,\ast dx_{n}),

where 𝝂=(νi)\boldsymbol{\nu}=(\nu^{i}) and dsds denote the outer unit normal and area element on D\partial D, respectively. Then we obtain

DC𝝂𝑑s=iDCi𝑑xi\int_{\partial D}C\cdot\boldsymbol{\nu}\ ds=\sum_{i}\int_{\partial D}C^{i}\ast dx^{i}

for C=(Ci)C=(C^{i}).

Let y=Ttxy=T_{t}x be the transformation of variables, and 𝐚=(ai)C1,1(Γ)\mathbf{a}=(a^{i})\in C^{1,1}(\Gamma) for Γd+1\Gamma\subset\mathbb{R}^{d+1} defined by (40). Then we obtain

Ωt𝝂𝐚𝑑st=iΩtai𝑑yi.\int_{\partial\Omega_{t}}\boldsymbol{\nu}\cdot\mathbf{a}\ ds_{t}=\sum_{i}\int_{\partial\Omega_{t}}a^{i}\ast dy^{i}. (72)

In (24) we have

yi=xi+tSi(x)+t22Ri(x)+o(t2),t0, 1idy^{i}=x^{i}+tS^{i}(x)+\frac{t^{2}}{2}R^{i}(x)+o(t^{2}),\quad t\rightarrow 0,\ 1\leq i\leq d

in C0,1(Ω¯)C^{0,1}(\overline{\Omega}) and hence

yji=δij+tSji+t22Rji+o(t2),t0, 1i,jdy^{i}_{j}=\delta_{ij}+tS^{i}_{j}+\frac{t^{2}}{2}R^{i}_{j}+o(t^{2}),\quad t\rightarrow 0,\ 1\leq i,j\leq d

in L(Ω)L^{\infty}(\Omega), where yji=yixjy^{i}_{j}=\frac{\partial y^{i}}{\partial x^{j}}, Sji=SixjS^{i}_{j}=\frac{\partial S^{i}}{\partial x^{j}}, Rji=RixjR^{i}_{j}=\frac{\partial R^{i}}{\partial x^{j}}, and so forth. Thus it holds that

dyi=jyjidxj=j(δij+tSji+t22Rji)dxj+o(t2).dy^{i}=\sum_{j}y^{i}_{j}dx^{j}=\sum_{j}\left(\delta_{ij}+tS^{i}_{j}+\frac{t^{2}}{2}R^{i}_{j}\right)dx^{j}+o(t^{2}).

Pull back the vector field 𝐚(y,t)\mathbf{a}(y,t) on Ωt\partial\Omega_{t} to that on Ω\partial\Omega by TtT_{t}: 𝐚(Tt(x),t)\mathbf{a}(T_{t}(x),t). Using

ati(y,t)=ait(y,t),atti(y,t)=2ait2(y,t),a_{t}^{i}(y,t)=\frac{\partial a^{i}}{\partial t}(y,t),\quad a_{tt}^{i}(y,t)=\frac{\partial^{2}a^{i}}{\partial t^{2}}(y,t),

we obtain

tai(Tt(x),t)\displaystyle\frac{\partial\;}{\partial t}a^{i}(T_{t}(x),t) =\displaystyle= xai(Tt(x),t)Tt(x)t+ati(Tt(x),t),\displaystyle\nabla_{x}a^{i}(T_{t}(x),t)\frac{\partial T_{t}(x)}{\partial t}+a_{t}^{i}(T_{t}(x),t),
2t2ai(Tt(x),t)\displaystyle\frac{\partial^{2}\;}{\partial t^{2}}a^{i}(T_{t}(x),t) =\displaystyle= x2ai(Tt(x),t)[Tt(x)t,Tt(x)t]+2xati(Tt(x),t)Tt(x)t\displaystyle\nabla_{x}^{2}a^{i}(T_{t}(x),t)\left[\frac{\partial T_{t}(x)}{\partial t},\frac{\partial T_{t}(x)}{\partial t}\right]+2\nabla_{x}a_{t}^{i}(T_{t}(x),t)\frac{\partial T_{t}(x)}{\partial t}
+xai(Tt(x),t)2Tt(x)t2+atti(Tt(x),t),\displaystyle+\nabla_{x}a^{i}(T_{t}(x),t)\frac{\partial^{2}T_{t}(x)}{\partial t^{2}}+a_{tt}^{i}(T_{t}(x),t),

and therefore,

ai(Tt(x),t)\displaystyle a^{i}(T_{t}(x),t) =\displaystyle= a0i+t(Sa0i+a˙0i)+t22(Ra0i+(2a0i)[S,S]\displaystyle a_{0}^{i}+t(S\cdot\nabla a_{0}^{i}+\dot{a}_{0}^{i})+\frac{t^{2}}{2}(R\cdot\nabla a_{0}^{i}+(\nabla^{2}a_{0}^{i})[S,S] (73)
+2a˙0iS+a¨0i)+o(t2),\displaystyle+2\nabla\dot{a}_{0}^{i}S+\ddot{a}_{0}^{i})+o(t^{2}),

where

a0i=ai(,0),a˙0i=ati(,0),a¨0i=atti(,0).a_{0}^{i}=a^{i}(\cdot,0),\quad\dot{a}_{0}^{i}=a_{t}^{i}(\cdot,0),\quad\ddot{a}_{0}^{i}=a_{tt}^{i}(\cdot,0).

We have, on the other hand,

dzi=(1)i1dz1dzi^dzd*dz^{i}=(-1)^{i-1}dz^{1}\wedge\cdots\wedge\widehat{dz^{i}}\wedge\cdots\wedge dz^{d}

for z=yz=y or z=xz=x, recalling that dzi^\widehat{dz^{i}} indicates the exclusion of dzidz^{i}. It holds also that

dyi=j=1dyjidxj,dy^{i}=\sum_{j=1}^{d}y_{j}^{i}dx^{j},

and therefore,

dyi\displaystyle*dy^{i} =\displaystyle= dy1dyi^dyd\displaystyle dy^{1}\wedge\cdots\wedge\widehat{dy^{i}}\wedge\cdots\wedge dy^{d}
=\displaystyle= (1)i1j=1ddet(ypq)p=1,,d,pjq=1,,d,qidx1dxj^dxd\displaystyle(-1)^{i-1}\sum_{j=1}^{d}\det(y_{p}^{q})_{\begin{subarray}{c}p=1,\cdots,d,p\neq j\\ q=1,\cdots,d,q\neq i\end{subarray}}dx^{1}\wedge\cdots\wedge\widehat{dx^{j}}\wedge\cdots\wedge dx^{d}
=\displaystyle= (1)i1j=1d(1)j1det(ypq)p=1,,d,pjq=1,,d,qidxj.\displaystyle(-1)^{i-1}\sum_{j=1}^{d}(-1)^{j-1}\det(y_{p}^{q})_{\begin{subarray}{c}p=1,\cdots,d,p\neq j\\ q=1,\cdots,d,q\neq i\end{subarray}}*\!dx^{j}.

First, if i=ji=j, we have

det(ypq)p,q=1,,dp,qi=det(1+tS11+t22R11tS21+t22R21tSd1+t22Rd1tS12+t22R121+tS22+t22R22tSd2+t22Rd2tS1d+t22R1dtS2d+t22R2d1+tSdd+t22Rdd)+o(t2),\det(y_{p}^{q})_{\begin{subarray}{c}p,q=1,\cdots,d\\ p,q\neq i\end{subarray}}=\det\begin{pmatrix}1+tS_{1}^{1}+\frac{t^{2}}{2}R_{1}^{1}&tS_{2}^{1}+\frac{t^{2}}{2}R_{2}^{1}&\cdots&tS_{d}^{1}+\frac{t^{2}}{2}R_{d}^{1}\\ tS_{1}^{2}+\frac{t^{2}}{2}R_{1}^{2}&1+tS_{2}^{2}+\frac{t^{2}}{2}R_{2}^{2}&\cdots&\ tS_{d}^{2}+\frac{t^{2}}{2}R_{d}^{2}\\ \vdots&\vdots&&\vdots\\ tS_{1}^{d}+\frac{t^{2}}{2}R_{1}^{d}&tS_{2}^{d}+\frac{t^{2}}{2}R_{2}^{d}&\cdots&1+tS_{d}^{d}+\frac{t^{2}}{2}R_{d}^{d}\end{pmatrix}+o(t^{2}),

where the entries in the form of tSiq+t22RiqtS_{i}^{q}+\frac{t^{2}}{2}R_{i}^{q}, tSpi+t22RpitS_{p}^{i}+\frac{t^{2}}{2}R_{p}^{i}, or 1+tSii+t22Rii1+tS_{i}^{i}+\frac{t^{2}}{2}R_{i}^{i} are not included in this matrix. Hence it follows that

det(ypq)p,q=1,,dp,qi=1+tkiSkk+t22kiRkk+t2j,kij<k(SjjSkkSkjSjk)+o(t2).\det(y_{p}^{q})_{\begin{subarray}{c}p,q=1,\cdots,d\\ p,q\neq i\end{subarray}}=1+t\sum_{k\neq i}S_{k}^{k}+\frac{t^{2}}{2}\sum_{k\neq i}R_{k}^{k}+t^{2}\sum_{\begin{subarray}{c}j,k\neq i\\ j<k\end{subarray}}(S_{j}^{j}S_{k}^{k}-S_{k}^{j}S_{j}^{k})+o(t^{2}).

Second, if iji\neq j we have

det(ypq)p=1,,d,pjq=1,,d,qi=(1)ji+1(tSij+t22Rij+t2pi,j(SijSppSipSpj))+o(t2),\det(y_{p}^{q})_{\begin{subarray}{c}p=1,\cdots,d,p\neq j\\ q=1,\cdots,d,q\neq i\end{subarray}}=(-1)^{j-i+1}\left(tS_{i}^{j}+\frac{t^{2}}{2}R_{i}^{j}+t^{2}\sum_{p\neq i,j}(S_{i}^{j}S_{p}^{p}-S_{i}^{p}S_{p}^{j})\right)+o(t^{2}), (74)

or

(1)i+jdet(ypq)p=1,,d,pjq=1,,d,qi=tSijt22Rij+t2pi,j(SipSpjSppSij)+o(t2).(-1)^{i+j}\det(y_{p}^{q})_{\begin{subarray}{c}p=1,\cdots,d,p\neq j\\ q=1,\cdots,d,q\neq i\end{subarray}}=-tS_{i}^{j}-\frac{t^{2}}{2}R_{i}^{j}+t^{2}\sum_{p\neq i,j}(S_{i}^{p}S_{p}^{j}-S_{p}^{p}S_{i}^{j})+o(t^{2}).

Equality (74) is obtained by an expansion of the determinant, and the proof is left for the reader. We thus end up with

dyi\displaystyle*dy^{i} =\displaystyle= dxi+tjiSjjdxitjiSijdxj+t22(jiRjj)dxit22jiRijdxj\displaystyle*dx^{i}+t\sum_{j\neq i}S_{j}^{j}*\!dx^{i}-t\sum_{j\neq i}S_{i}^{j}*\!dx^{j}+\frac{t^{2}}{2}\left(\sum_{j\neq i}R_{j}^{j}\right)*\!dx^{i}-\frac{t^{2}}{2}\sum_{j\neq i}R_{i}^{j}*\!dx^{j} (75)
+t2j,kij<k(SjjSkkSkjSjk)dxi+t2jipi,j(SipSpjSppSij)dxj+o(t2).\displaystyle+t^{2}\sum_{\begin{subarray}{c}j,k\neq i\\ j<k\end{subarray}}(S_{j}^{j}S_{k}^{k}-S_{k}^{j}S_{j}^{k})*\!dx^{i}+t^{2}\sum_{j\neq i}\sum_{p\neq i,j}(S_{i}^{p}S_{p}^{j}-S_{p}^{p}S_{i}^{j})*\!dx^{j}+o(t^{2}).

Writing (73) as

ai(y,t)=a0i+tXi+t22Yi+o(t2),a^{i}(y,t)=a_{0}^{i}+tX^{i}+\frac{t^{2}}{2}Y^{i}+o(t^{2}),

with

Xi=Sa0i+a˙0i,Yi=Ra0i+(2a0i)[S,S]+2a˙0iS+a¨0i,X^{i}=S\cdot\nabla a_{0}^{i}+\dot{a}_{0}^{i},\quad Y^{i}=R\cdot\nabla a_{0}^{i}+(\nabla^{2}a_{0}^{i})[S,S]+2\nabla\dot{a}_{0}^{i}\cdot S+\ddot{a}_{0}^{i},

we obtain

ai(y,t)dyi\displaystyle a^{i}(y,t)*\!dy^{i} =\displaystyle= a0idxi+t(a0ijiSjj+Sa0i+a˙0i)dxita0ijiSijdxj\displaystyle a_{0}^{i}*\!dx^{i}+t\left(a_{0}^{i}\sum_{j\neq i}S_{j}^{j}+S\cdot\nabla a_{0}^{i}+\dot{a}_{0}^{i}\right)*\!dx^{i}-ta_{0}^{i}\sum_{j\neq i}S_{i}^{j}*\!dx^{j}
+t22a0i(jiRjj)dxit22a0ijiRijdxj\displaystyle+\frac{t^{2}}{2}a_{0}^{i}\left(\sum_{j\neq i}R_{j}^{j}\right)*\!dx^{i}-\frac{t^{2}}{2}a_{0}^{i}\sum_{j\neq i}R_{i}^{j}*\!dx^{j}
+t2a0ij,kij<k(SjjSkkSkjSjk)dxi+t2a0ijipi,j(SipSpjSppSij)dxj\displaystyle+t^{2}a_{0}^{i}\sum_{\begin{subarray}{c}j,k\neq i\\ j<k\end{subarray}}(S_{j}^{j}S_{k}^{k}-S_{k}^{j}S_{j}^{k})*\!dx^{i}+t^{2}a_{0}^{i}\sum_{j\neq i}\sum_{p\neq i,j}(S_{i}^{p}S_{p}^{j}-S_{p}^{p}S_{i}^{j})*\!dx^{j}
+t22Yidxi+t2Xi(jiSjj)dxit2XijiSijdxj+o(t2)\displaystyle+\frac{t^{2}}{2}Y^{i}*\!dx^{i}+t^{2}X^{i}\left(\sum_{j\neq i}S_{j}^{j}\right)*\!dx^{i}-t^{2}X^{i}\sum_{j\neq i}S_{i}^{j}*\!dx^{j}+o(t^{2})

by (75). Then (72) implies

Ωt𝐚(y,t)𝝂𝑑sy\displaystyle\int_{\partial\Omega_{t}}\mathbf{a}(y,t)\cdot\boldsymbol{\nu}ds_{y} =\displaystyle= i=1dΩtai(y,t)𝑑yi\displaystyle\sum_{i=1}^{d}\int_{\partial\Omega_{t}}a^{i}(y,t)*\!dy^{i} (76)
=\displaystyle= i=1dΩa0idxi+tΩI++t22ΩII+o(t2)\displaystyle\sum_{i=1}^{d}\int_{\partial\Omega}a_{0}^{i}*\!dx^{i}+t\int_{\partial\Omega}I++\frac{t^{2}}{2}\int_{\partial\Omega}II+o(t^{2})

with

I=i=1dji[(a0iSjja0jSji)+Sa0i+a˙0i]dxiI=\sum_{i=1}^{d}\sum_{j\neq i}\left[\left(a_{0}^{i}S_{j}^{j}-a_{0}^{j}S_{j}^{i}\right)+S\cdot\nabla a_{0}^{i}+\dot{a}_{0}^{i}\right]*\!dx^{i} (77)

and

II\displaystyle II =\displaystyle= i=1da0i(jiRjj)dxii=1da0ijiRijdxj\displaystyle\sum_{i=1}^{d}a_{0}^{i}\left(\sum_{j\neq i}R_{j}^{j}\right)*\!dx^{i}-\sum_{i=1}^{d}a_{0}^{i}\sum_{j\neq i}R_{i}^{j}*\!dx^{j} (78)
+2i=1da0ij,kij<k(SjjSkkSkjSjk)dxi+2i=1da0ijipi,j(SipSpjSppSij)dxj\displaystyle+2\sum_{i=1}^{d}a_{0}^{i}\sum_{\begin{subarray}{c}j,k\neq i\\ j<k\end{subarray}}\left(S_{j}^{j}S_{k}^{k}-S_{k}^{j}S_{j}^{k}\right)*\!dx^{i}+2\sum_{i=1}^{d}a_{0}^{i}\sum_{j\neq i}\sum_{p\neq i,j}\left(S_{i}^{p}S_{p}^{j}-S_{p}^{p}S_{i}^{j}\right)*\!dx^{j}
+i=1d(Ra0i+(2a0i)[S,S]+2a˙0iS+a¨0i)dxi\displaystyle+\sum_{i=1}^{d}(R\cdot\nabla a_{0}^{i}+(\nabla^{2}a_{0}^{i})[S,S]+2\nabla\dot{a}_{0}^{i}\cdot S+\ddot{a}_{0}^{i})*\!dx^{i}
+2i=1d(Sa0i+a˙0i)(jiSjj)dxi2i=1d(Sa0i+a˙0i)jiSijdxj.\displaystyle+2\sum_{i=1}^{d}(S\cdot\nabla a_{0}^{i}+\dot{a}_{0}^{i})\left(\sum_{j\neq i}S_{j}^{j}\right)*\!dx^{i}-2\sum_{i=1}^{d}(S\cdot\nabla a_{0}^{i}+\dot{a}_{0}^{i})\sum_{j\neq i}S_{i}^{j}*\!dx^{j}.

Here, we obtain

I\displaystyle I =\displaystyle= i=1d[j=1d(a0iSjja0jSji)+j=1dSja0,jij=1da0,jjSi+j=1da0,jjSi]dxi+i=1da˙0idxi\displaystyle\sum_{i=1}^{d}\left[\sum_{j=1}^{d}(a_{0}^{i}S_{j}^{j}-a_{0}^{j}S_{j}^{i})+\sum_{j=1}^{d}S^{j}a_{0,j}^{i}-\sum_{j=1}^{d}a_{0,j}^{j}S^{i}+\sum_{j=1}^{d}a_{0,j}^{j}S^{i}\right]*\!dx^{i}+\sum_{i=1}^{d}\dot{a}_{0}^{i}*\!dx^{i}
=\displaystyle= i=1d[j=1d(a0iSj)jj=1d(a0jSi)j+j=1da0,jjSi]dxi+i=1da˙0idxi\displaystyle\sum_{i=1}^{d}\left[\sum_{j=1}^{d}(a_{0}^{i}S^{j})_{j}-\sum_{j=1}^{d}(a_{0}^{j}S^{i})_{j}+\sum_{j=1}^{d}a_{0,j}^{j}S^{i}\right]*\!dx^{i}+\sum_{i=1}^{d}\dot{a}_{0}^{i}*\!dx^{i}
=\displaystyle= i=1d[(𝐚)Si+a˙0i]dxi+θ1\displaystyle\sum_{i=1}^{d}\left[(\nabla\cdot\mathbf{a})S^{i}+\dot{a}_{0}^{i}\right]*\!dx^{i}+\theta_{1}

for

θ1=i=1d[j=1d(a0iSj)jj=1d(a0jSi)j]dxi.\theta_{1}=\sum_{i=1}^{d}\left[\sum_{j=1}^{d}(a_{0}^{i}S^{j})_{j}-\sum_{j=1}^{d}(a_{0}^{j}S^{i})_{j}\right]*\!dx^{i}.

Since dxidxi=dx1dxddx^{i}\wedge\ast dx^{i}=dx^{1}\wedge\cdots\wedge dx^{d}, it holds that

dθ1=i,j=1d[(a0iSj)ij(a0jSi)ij]dx1dxd=0,d\theta_{1}=\sum_{i,j=1}^{d}\left[(a_{0}^{i}S^{j})_{ij}-(a_{0}^{j}S^{i})_{ij}\right]dx^{1}\wedge\cdots\wedge dx^{d}=0,

and hence

Ωθ1=Ω𝑑θ1=0\int_{\partial\Omega}\theta_{1}=\int_{\Omega}d\theta_{1}=0

by the Stokes theorem. We thus obtain

ΩI=i=1dΩ[(𝐚)Si+a˙0i]𝑑xi=Ω[(𝐚)S𝝂+𝐚˙0𝝂]𝑑s.\int_{\partial\Omega}I=\sum_{i=1}^{d}\int_{\partial\Omega}\left[(\nabla\cdot\mathbf{a})S^{i}+\dot{a}_{0}^{i}\right]*\!dx^{i}=\int_{\partial\Omega}\left[(\nabla\cdot\mathbf{a})S\cdot\boldsymbol{\nu}+\dot{\mathbf{a}}_{0}\cdot\boldsymbol{\nu}\right]ds. (79)

We divide IIII into four terms, involving 𝐚¨0\ddot{\mathbf{a}}_{0}, {R,𝐚0}\{R,\mathbf{a}_{0}\}, {S,𝐚˙0}\{S,\dot{\mathbf{a}}_{0}\}, and {S,𝐚0}\{S,\mathbf{a}_{0}\}, denoted by II1II_{1}, II2II_{2}, II3II_{3}, and II4II_{4}, respectively. First, we have

II1=i=1da¨0idxi,II_{1}=\sum_{i=1}^{d}\ddot{a}_{0}^{i}*\!dx^{i},

and hence

ΩII1=Ω𝐚¨0𝝂𝑑s.\int_{\partial\Omega}II_{1}=\int_{\partial\Omega}\ddot{\mathbf{a}}_{0}\cdot\boldsymbol{\nu}ds.

Second, there arises that

II2\displaystyle II_{2} =\displaystyle= i=1da0i(jiRjj)dxii=1da0ijiRijdxj+i=1dRa0idxj\displaystyle\sum_{i=1}^{d}a_{0}^{i}\left(\sum_{j\neq i}R_{j}^{j}\right)*\!dx^{i}-\sum_{i=1}^{d}a_{0}^{i}\sum_{j\neq i}R_{i}^{j}*\!dx^{j}+\sum_{i=1}^{d}R\cdot\nabla a_{0}^{i}*\!dx^{j}
=\displaystyle= i=1dji[(a0iRjja0jRji)+Ra0i]dxi\displaystyle\sum_{i=1}^{d}\sum_{j\neq i}\left[(a_{0}^{i}R_{j}^{j}-a_{0}^{j}R_{j}^{i})+R\cdot\nabla a_{0}^{i}\right]*\!dx^{i}
=\displaystyle= i=1d[j=1d(a0iRjja0jRji)+j=1dRja0,jij=1da0,jjRi+j=1da0,jjRi]dxi\displaystyle\sum_{i=1}^{d}\left[\sum_{j=1}^{d}(a_{0}^{i}R_{j}^{j}-a_{0}^{j}R_{j}^{i})+\sum_{j=1}^{d}R^{j}a_{0,j}^{i}-\sum_{j=1}^{d}a_{0,j}^{j}R^{i}+\sum_{j=1}^{d}a_{0,j}^{j}R^{i}\right]*\!dx^{i}
=\displaystyle= i=1d[j=1d(a0iRj)jj=1d(a0jRi)j+j=1da0,jjRi]dxi\displaystyle\sum_{i=1}^{d}\left[\sum_{j=1}^{d}(a_{0}^{i}R^{j})_{j}-\sum_{j=1}^{d}(a_{0}^{j}R^{i})_{j}+\sum_{j=1}^{d}a_{0,j}^{j}R^{i}\right]*\!dx^{i}
=\displaystyle= i=1d(𝐚)Ridxi+θ2\displaystyle\sum_{i=1}^{d}(\nabla\cdot\mathbf{a})R^{i}*\!dx^{i}+\theta_{2}

with

θ2=i=1d[j=1d(a0iRj)jj=1d(a0jRi)j]dxi,\theta_{2}=\sum_{i=1}^{d}\left[\sum_{j=1}^{d}(a_{0}^{i}R^{j})_{j}-\sum_{j=1}^{d}(a_{0}^{j}R^{i})_{j}\right]*\!dx^{i},

Then we obtain dθ2=0d\theta_{2}=0, and hence

Ωθ2=0,\int_{\partial\Omega}\theta_{2}=0,

similarly. It thus follows that

ΩII2=Ω(𝐚)(R𝝂)𝑑s.\int_{\partial\Omega}II_{2}=\int_{\partial\Omega}(\nabla\cdot\mathbf{a})\left(R\cdot\boldsymbol{\nu}\right)ds.

Third, we have

II3\displaystyle II_{3} =\displaystyle= 2i=1da˙0i(jiSjj)dxi2i=1da˙0ijiSijdxj+2i=1dSa˙0idxj\displaystyle 2\sum_{i=1}^{d}\dot{a}_{0}^{i}\left(\sum_{j\neq i}S_{j}^{j}\right)*\!dx^{i}-2\sum_{i=1}^{d}\dot{a}_{0}^{i}\sum_{j\neq i}S_{i}^{j}*\!dx^{j}+2\sum_{i=1}^{d}S\cdot\nabla\dot{a}_{0}^{i}*\!dx^{j}
=\displaystyle= 2i=1dji[(a˙0iSjja˙0jSji)+Sa˙0i]dxi\displaystyle 2\sum_{i=1}^{d}\sum_{j\neq i}\left[(\dot{a}_{0}^{i}S_{j}^{j}-\dot{a}_{0}^{j}S_{j}^{i})+S\cdot\nabla\dot{a}_{0}^{i}\right]*\!dx^{i}
=\displaystyle= 2i=1d[j=1d(a˙0iSjja˙0jSji)+j=1dSja˙0,jij=1da˙0,jjSi+j=1da˙0,jjSi]dxi\displaystyle 2\sum_{i=1}^{d}\left[\sum_{j=1}^{d}(\dot{a}_{0}^{i}S_{j}^{j}-\dot{a}_{0}^{j}S_{j}^{i})+\sum_{j=1}^{d}S^{j}\dot{a}_{0,j}^{i}-\sum_{j=1}^{d}\dot{a}_{0,j}^{j}S^{i}+\sum_{j=1}^{d}\dot{a}_{0,j}^{j}S^{i}\right]*\!dx^{i}
=\displaystyle= 2i=1d[j=1d(a˙0iSj)jj=1d(a˙0jSi)j+j=1da˙0,jjSi]dxi\displaystyle 2\sum_{i=1}^{d}\left[\sum_{j=1}^{d}(\dot{a}_{0}^{i}S^{j})_{j}-\sum_{j=1}^{d}(\dot{a}_{0}^{j}S^{i})_{j}+\sum_{j=1}^{d}\dot{a}_{0,j}^{j}S^{i}\right]*\!dx^{i}
=\displaystyle= 2i=1d(𝐚˙)Sidxi+2θ3\displaystyle 2\sum_{i=1}^{d}(\nabla\cdot\dot{\mathbf{a}})S^{i}*\!dx^{i}+2\theta_{3}

for

θ3=i=1d[j=1d(a˙0iSj)jj=1d(a˙0jSi)j]dxi.\theta_{3}=\sum_{i=1}^{d}\left[\sum_{j=1}^{d}(\dot{a}_{0}^{i}S^{j})_{j}-\sum_{j=1}^{d}(\dot{a}_{0}^{j}S^{i})_{j}\right]*\!dx^{i}.

Then we obtain dθ3=0d\theta_{3}=0, and hence

Ωθ3=0,\int_{\partial\Omega}\theta_{3}=0,

similarly. It thus follows that

ΩII3=2Ω(𝐚˙)(S𝝂)𝑑s.\int_{\partial\Omega}II_{3}=2\int_{\partial\Omega}(\nabla\cdot\dot{\mathbf{a}})\left(S\cdot\boldsymbol{\nu}\right)ds.

Finally, we treat

II4=2i=1da0ij,kij<k(SjjSkkSkjSjk)dxi+2i=1da0ijipi,j(SipSpjSppSij)dxj\displaystyle II_{4}=2\sum_{i=1}^{d}a_{0}^{i}\sum_{\begin{subarray}{c}j,k\neq i\\ j<k\end{subarray}}(S_{j}^{j}S_{k}^{k}-S_{k}^{j}S_{j}^{k})*\!dx^{i}+2\sum_{i=1}^{d}a_{0}^{i}\sum_{j\neq i}\sum_{p\neq i,j}(S_{i}^{p}S_{p}^{j}-S_{p}^{p}S_{i}^{j})*\!dx^{j}
+i=1d(2a0i)[S,S]dxi+2i=1d(Sa0i)(jiSjj)dxi2i=1d(Sa0i)jiSijdxj\displaystyle+\sum_{i=1}^{d}(\nabla^{2}a_{0}^{i})[S,S]*\!dx^{i}+2\sum_{i=1}^{d}(S\cdot\nabla a_{0}^{i})\left(\sum_{j\neq i}S_{j}^{j}\right)*\!dx^{i}-2\sum_{i=1}^{d}(S\cdot\nabla a_{0}^{i})\sum_{j\neq i}S_{i}^{j}*\!dx^{j}

Our goal is to show

ΩII4\displaystyle\int_{\partial\Omega}II_{4} =\displaystyle= i=1dΩ[(𝐚0)S]Si𝑑xii=1dΩ(𝐚0)(S)Si𝑑xi\displaystyle\sum_{i=1}^{d}\int_{\partial\Omega}\nabla\cdot[(\nabla\cdot\mathbf{a}_{0})S]S^{i}*\!dx^{i}-\sum_{i=1}^{d}\int_{\partial\Omega}(\nabla\cdot\mathbf{a}_{0})(S\cdot\nabla)S^{i}*\!dx^{i} (80)
=\displaystyle= Ω[(𝐚)S](S𝝂)(𝐚)[(S)S]𝝂ds.\displaystyle\int_{\partial\Omega}\nabla\cdot[(\nabla\cdot\mathbf{a})S](S\cdot\boldsymbol{\nu})-(\nabla\cdot\mathbf{a})[(S\cdot\nabla)S]\cdot\boldsymbol{\nu}\ ds.

For this purpose we note the equalities

[(𝐚0)S]Sidxi(𝐚0)(S)Sidxi\displaystyle\nabla\cdot[(\nabla\cdot\mathbf{a}_{0})S]S^{i}*\!dx^{i}-(\nabla\cdot\mathbf{a}_{0})(S\cdot\nabla)S^{i}*\!dx^{i}
=j,ka0,jkkSjSidxi+j,ka0,kkSjjSidxij,ka0,kkSjiSjdxi\displaystyle\quad=\sum_{j,k}a_{0,jk}^{k}S^{j}S^{i}*\!dx^{i}+\sum_{j,k}a_{0,k}^{k}S_{j}^{j}S^{i}*\!dx^{i}-\sum_{j,k}a_{0,k}^{k}S_{j}^{i}S^{j}*\!dx^{i}

and

2a0i[S,S]=j,ka0,jkiSjSk,Sa0i=ka0,kiSk.\nabla^{2}a_{0}^{i}[S,S]=\sum_{j,k}a_{0,jk}^{i}S^{j}S^{k},\quad S\cdot\nabla a_{0}^{i}=\sum_{k}a_{0,k}^{i}S^{k}.

We write also

2ia0ij,kij<k(SjjSkkSkjSjk)dxi=2ia0iAidxi,\displaystyle 2\sum_{i}a_{0}^{i}\sum_{\begin{subarray}{c}j,k\neq i\\ j<k\end{subarray}}(S_{j}^{j}S_{k}^{k}-S_{k}^{j}S_{j}^{k})*\!dx^{i}=2\sum_{i}a_{0}^{i}A^{i}*\!dx^{i},
2ia0ijipi,j(SipSpjSppSij)dxj=2ia0ijiBijdxj,\displaystyle 2\sum_{i}a_{0}^{i}\sum_{j\neq i}\sum_{p\neq i,j}(S_{i}^{p}S_{p}^{j}-S_{p}^{p}S_{i}^{j})*\!dx^{j}=2\sum_{i}a_{0}^{i}\sum_{j\neq i}B^{ij}*\!dx^{j},

using

Ai\displaystyle A^{i} =\displaystyle= j,kij<k(SjjSkkSkjSjk)=12jiki(SjjSkkSkjSjk),\displaystyle\sum_{\begin{subarray}{c}j,k\neq i\\ j<k\end{subarray}}(S_{j}^{j}S_{k}^{k}-S_{k}^{j}S_{j}^{k})=\frac{1}{2}\sum_{j\neq i}\sum_{k\neq i}(S_{j}^{j}S_{k}^{k}-S_{k}^{j}S_{j}^{k}),
Bij\displaystyle B^{ij} =\displaystyle= ki,j(SikSkjSkkSij)=ki(SikSkjSkkSij).\displaystyle\sum_{k\neq i,j}(S_{i}^{k}S_{k}^{j}-S_{k}^{k}S_{i}^{j})=\sum_{k\neq i}(S_{i}^{k}S_{k}^{j}-S_{k}^{k}S_{i}^{j}).

We thus obtain

II4=i[(𝐚0)S]Sidxii(𝐚0)(S)Sidxi+θ4+θ5+θ6,II_{4}=\sum_{i}\nabla\cdot[(\nabla\cdot\mathbf{a}_{0})S]S^{i}*\!dx^{i}-\sum_{i}(\nabla\cdot\mathbf{a}_{0})(S\cdot\nabla)S^{i}*\!dx^{i}+\theta_{4}+\theta_{5}+\theta_{6},

for

θ4=ij,ka0,jkiSjSkdxiij,ka0,jkkSiSjdxiij,ka0,kkSiSjjdxi,\theta_{4}=\sum_{i}\sum_{j,k}a_{0,jk}^{i}S^{j}S^{k}*\!dx^{i}-\sum_{i}\sum_{j,k}a_{0,jk}^{k}S^{i}S^{j}*\!dx^{i}-\sum_{i}\sum_{j,k}a_{0,k}^{k}S^{i}S_{j}^{j}*\!dx^{i},
θ5\displaystyle\theta_{5} =\displaystyle= ij,ka0,kkSjSjidxi+2ij,k,jia0,kiSkSjjdxi2ij,k,jia0,kiSkSijdxj\displaystyle\sum_{i}\sum_{j,k}a_{0,k}^{k}S^{j}S_{j}^{i}*\!dx^{i}+2\sum_{i}\sum_{j,k,j\neq i}a_{0,k}^{i}S^{k}S_{j}^{j}*\!dx^{i}-2\sum_{i}\sum_{j,k,j\neq i}a_{0,k}^{i}S^{k}S_{i}^{j}*\!dx^{j}
=\displaystyle= ij,ka0,kkSjSjidxi+2ij,ka0,kiSkSjjdxi2ij,ka0,kjSkSjidxi,\displaystyle\sum_{i}\sum_{j,k}a_{0,k}^{k}S^{j}S_{j}^{i}*\!dx^{i}+2\sum_{i}\sum_{j,k}a_{0,k}^{i}S^{k}S_{j}^{j}*\!dx^{i}-2\sum_{i}\sum_{j,k}a_{0,k}^{j}S^{k}S_{j}^{i}*\!dx^{i},

and

θ6=2ia0iAidxi+2ia0ijiBijdxj.\theta_{6}=2\sum_{i}a_{0}^{i}A^{i}*\!dx^{i}+2\sum_{i}a_{0}^{i}\sum_{j\neq i}B^{ij}*\!dx^{j}.

Thus, equality (80) is reduced to

Ω(θ4+θ5+θ6)=Ωd(θ4+θ5+θ6)=0,\int_{\partial\Omega}(\theta_{4}+\theta_{5}+\theta_{6})=\int_{\Omega}d(\theta_{4}+\theta_{5}+\theta_{6})=0,

which follows from

d(θ4+θ5+θ6)=0.d(\theta_{4}+\theta_{5}+\theta_{6})=0. (81)

In fact, we have

d(θ4+θ5+θ6)=Xdx1dxdd(\theta_{4}+\theta_{5}+\theta_{6})=X\ dx^{1}\wedge\cdots\wedge dx^{d}

with

X\displaystyle X =\displaystyle= i,j,ka0,ijkiSjSk+i,j,ka0,jki(SjSk)ii,j,ka0,ijkkSiSji,j,ka0,jkk(SiSj)i\displaystyle\sum_{i,j,k}a_{0,ijk}^{i}S^{j}S^{k}+\sum_{i,j,k}a_{0,jk}^{i}(S^{j}S^{k})_{i}-\sum_{i,j,k}a_{0,ijk}^{k}S^{i}S^{j}-\sum_{i,j,k}a_{0,jk}^{k}(S^{i}S^{j})_{i}
i,j,ka0,ikkSiSjji,j,ka0,kk(SiSjj)i\displaystyle-\sum_{i,j,k}a_{0,ik}^{k}S^{i}S_{j}^{j}-\sum_{i,j,k}a_{0,k}^{k}(S^{i}S_{j}^{j})_{i}
+i,j,ka0,ikkSjSji+i,j,ka0,kk(SjSji)i+2i,j,ka0,ikiSkSjj+2i,j,ka0,ki(SkSjj)i\displaystyle+\sum_{i,j,k}a_{0,ik}^{k}S^{j}S_{j}^{i}+\sum_{i,j,k}a_{0,k}^{k}(S^{j}S_{j}^{i})_{i}+2\sum_{i,j,k}a_{0,ik}^{i}S^{k}S_{j}^{j}+2\sum_{i,j,k}a_{0,k}^{i}(S^{k}S_{j}^{j})_{i}
2i,j,ka0,ikjSkSji2i,j,ka0,kj(SkSji)i\displaystyle-2\sum_{i,j,k}a_{0,ik}^{j}S^{k}S_{j}^{i}-2\sum_{i,j,k}a_{0,k}^{j}(S^{k}S_{j}^{i})_{i}
+2ia0,iiAi+2ia0iAii+2ijia0,jiBij+2ia0ijiBjij\displaystyle+2\sum_{i}a_{0,i}^{i}A_{i}+2\sum_{i}a_{0}^{i}A_{i}^{i}+2\sum_{i}\sum_{j\neq i}a_{0,j}^{i}B^{ij}+2\sum_{i}a_{0}^{i}\sum_{j\neq i}B_{j}^{ij}
\displaystyle\equiv X1+X2X3X4X5X6\displaystyle X_{1}+X_{2}-X_{3}-X_{4}-X_{5}-X_{6}
+X7+X8+2X9+2X102X112X12\displaystyle+X_{7}+X_{8}+2X_{9}+2X_{10}-2X_{11}-2X_{12}
+2X13+2X14+2X15+2X16.\displaystyle+2X_{13}+2X_{14}+2X_{15}+2X_{16}.

First, noticing the terms involving the third order derivatives of a0ia_{0}^{i}, we realize

X1X3=0.X_{1}-X_{3}=0.

Second, the terms involving the zero-th order derivatives of a0ia_{0}^{i} also cancel as

X14+X16=ia0iFiX_{14}+X_{16}=\sum_{i}a_{0}^{i}F_{i}

with

Fi\displaystyle F_{i} =\displaystyle= Aii+jiBjij\displaystyle A_{i}^{i}+\sum_{j\neq i}B_{j}^{ij}
=\displaystyle= 12j,ki(SjjSkkSkjSjk)i+j,ki(SikSkjSkkSij)j\displaystyle\frac{1}{2}\sum_{j,k\neq i}(S_{j}^{j}S_{k}^{k}-S_{k}^{j}S_{j}^{k})_{i}+\sum_{j,k\neq i}(S_{i}^{k}S_{k}^{j}-S_{k}^{k}S_{i}^{j})_{j}
=\displaystyle= 12j,ki(SijjSkk+SjjSkikSikjSjkSkjSijk)+j,ki(SijkSkj+SikSjkjSkjkSijSkkSijj)\displaystyle\frac{1}{2}\sum_{j,k\neq i}(S_{ij}^{j}S_{k}^{k}+S_{j}^{j}S_{ki}^{k}-S_{ik}^{j}S_{j}^{k}-S_{k}^{j}S_{ij}^{k})+\sum_{j,k\neq i}(S_{ij}^{k}S_{k}^{j}+S_{i}^{k}S_{jk}^{j}-S_{kj}^{k}S_{i}^{j}-S_{k}^{k}S_{ij}^{j})
=\displaystyle= 12(X17+X18X19X20)+X21+X22X23X24.\displaystyle\frac{1}{2}(X_{17}+X_{18}-X_{19}-X_{20})+X_{21}+X_{22}-X_{23}-X_{24}.

Since

X17=X24,X20=X21,X22=X23,X_{17}=X_{24},\quad X_{20}=X_{21},\quad X_{22}=X_{23},

we obtain

Fi=12j,ki(SijjSkk+SjjSkikSikjSjk+SkjSijk)=0.F_{i}=\frac{1}{2}\sum_{j,k\neq i}(-S_{ij}^{j}S_{k}^{k}+S_{j}^{j}S_{ki}^{k}-S_{ik}^{j}S_{j}^{k}+S_{k}^{j}S_{ij}^{k})=0.

Third, the terms involving the second order derivatives of a0ia_{0}^{i} cancel as

X2X4X5+X7+2X92X11\displaystyle X_{2}-X_{4}-X_{5}+X_{7}+2X_{9}-2X_{11}
=i,j,ka0,jkiSijSk+i,j,ka0,jkiSjSik+i,j,ka0,ikkSjSji+2i,j,ka0,ikiSkSjj\displaystyle\quad=\sum_{i,j,k}a_{0,jk}^{i}S_{i}^{j}S^{k}+\sum_{i,j,k}a_{0,jk}^{i}S^{j}S_{i}^{k}+\sum_{i,j,k}a_{0,ik}^{k}S^{j}S_{j}^{i}+2\sum_{i,j,k}a_{0,ik}^{i}S^{k}S_{j}^{j}
i,j,ka0,jkkSjSiii,j,ka0,jkkSiSiji,j,ka0,ikkSiSjj2i,j,ka0,ikjSkSji\displaystyle\qquad-\sum_{i,j,k}a_{0,jk}^{k}S^{j}S_{i}^{i}-\sum_{i,j,k}a_{0,jk}^{k}S^{i}S_{i}^{j}-\sum_{i,j,k}a_{0,ik}^{k}S^{i}S_{j}^{j}-2\sum_{i,j,k}a_{0,ik}^{j}S^{k}S_{j}^{i}
=X25+X26+X27+2X28X29X30X312X32=0\displaystyle\quad=X_{25}+X_{26}+X_{27}+2X_{28}-X_{29}-X_{30}-X_{31}-2X_{32}=0

by

X25=X26=X32,X27=X30,X28=X29=X31.X_{25}=X_{26}=X_{32},\quad X_{27}=X_{30},\quad X_{28}=X_{29}=X_{31}.

Finally, for the terms involving the first order derivatives of a0ia_{0}^{i} we obtain

Y=X6+X8+2X102X12+2X13\displaystyle Y=-X_{6}+X_{8}+2X_{10}-2X_{12}+2X_{13}
=i,j,ka0,kk(SiSjj)i+i,j,ka0,kk(SjSji)i+2i,j,ka0,ki(SkSjj)i\displaystyle\quad=-\sum_{i,j,k}a_{0,k}^{k}(S^{i}S_{j}^{j})_{i}+\sum_{i,j,k}a_{0,k}^{k}(S^{j}S_{j}^{i})_{i}+2\sum_{i,j,k}a_{0,k}^{i}(S^{k}S_{j}^{j})_{i}
2i,j,ka0,kj(SkSji)i+2ia0,iiAi+2ia0,jijiBij\displaystyle\qquad-2\sum_{i,j,k}a_{0,k}^{j}(S^{k}S_{j}^{i})_{i}+2\sum_{i}a_{0,i}^{i}A_{i}+2\sum_{i}a_{0,j}^{i}\sum_{j\neq i}B^{ij}
=i,j,ka0,kkSiiSjji,j,ka0,kkSiSijj+i,j,ka0,kkSijSji+i,j,ka0,kkSjSiji\displaystyle\quad=-\sum_{i,j,k}a_{0,k}^{k}S_{i}^{i}S_{j}^{j}-\sum_{i,j,k}a_{0,k}^{k}S^{i}S_{ij}^{j}+\sum_{i,j,k}a_{0,k}^{k}S_{i}^{j}S_{j}^{i}+\sum_{i,j,k}a_{0,k}^{k}S^{j}S_{ij}^{i}
+2i,j,ka0,kiSikSjj+2i,j,ka0,kiSkSijj2i,j,ka0,kjSikSji2i,j,ka0,kjSkSiji\displaystyle\qquad+2\sum_{i,j,k}a_{0,k}^{i}S_{i}^{k}S_{j}^{j}+2\sum_{i,j,k}a_{0,k}^{i}S^{k}S_{ij}^{j}-2\sum_{i,j,k}a_{0,k}^{j}S_{i}^{k}S_{j}^{i}-2\sum_{i,j,k}a_{0,k}^{j}S^{k}S_{ij}^{i}
+2ia0,iiAi+2ijia0,jiBij\displaystyle\qquad+2\sum_{i}a_{0,i}^{i}A_{i}+2\sum_{i}\sum_{j\neq i}a_{0,j}^{i}B^{ij}
=X33X34+X35+X36\displaystyle=-X_{33}-X_{34}+X_{35}+X_{36}
+2X37+2X382X392X40+2X41+2X42.\displaystyle\quad+2X_{37}+2X_{38}-2X_{39}-2X_{40}+2X_{41}+2X_{42}.

Then it holds that

X33X34+X35+X36=ka0,kk(i,jk(SiiSjjSijSji)+i,jSijSjii,jSiiSjj)\displaystyle-X_{33}-X_{34}+X_{35}+X_{36}=\sum_{k}a_{0,k}^{k}\left(\sum_{i,j\neq k}(S_{i}^{i}S_{j}^{j}-S_{i}^{j}S_{j}^{i})+\sum_{i,j}S_{i}^{j}S_{j}^{i}-\sum_{i,j}S_{i}^{i}S_{j}^{j}\right)
=2j,ka0,kk(SkjSjkSkkSjj),\displaystyle\quad=2\sum_{j,k}a_{0,k}^{k}(S_{k}^{j}S_{j}^{k}-S_{k}^{k}S_{j}^{j}),
2X372X39=2i,j,ka0,ji(SijSkkSkjSik)+2ijia0,jik(SikSkjSkkSij)\displaystyle 2X_{37}-2X_{39}=2\sum_{i,j,k}a_{0,j}^{i}(S_{i}^{j}S_{k}^{k}-S_{k}^{j}S_{i}^{k})+2\sum_{i}\sum_{j\neq i}a_{0,j}^{i}\sum_{k}(S_{i}^{k}S_{k}^{j}-S_{k}^{k}S_{i}^{j})
2X382X40=2ijia0,jik(SikSkjSkkSij)\displaystyle 2X_{38}-2X_{40}=2\sum_{i}\sum_{j\neq i}a_{0,j}^{i}\sum_{k}(S_{i}^{k}S_{k}^{j}-S_{k}^{k}S_{i}^{j})
2X41+2X42=0,\displaystyle 2X_{41}+2X_{42}=0,

which implies

Y=2j,ka0,kk(SkjSjkSkkSjj)+2i,ka0,ii(SiiSkkSkiSik)=0.Y=2\sum_{j,k}a_{0,k}^{k}(S_{k}^{j}S_{j}^{k}-S_{k}^{k}S_{j}^{j})+2\sum_{i,k}a_{0,i}^{i}(S_{i}^{i}S_{k}^{k}-S_{k}^{i}S_{i}^{k})=0.

Hence we obtain (81).

We thus end up with

Ωt𝐚(y,t)𝝂𝑑sy\displaystyle\int_{\partial\Omega_{t}}\mathbf{a}(y,t)\cdot\boldsymbol{\nu}ds_{y} =\displaystyle= i=1dΩtai(y,t)𝑑yi\displaystyle\sum_{i=1}^{d}\int_{\partial\Omega_{t}}a^{i}(y,t)*\!dy^{i}
=\displaystyle= Ω𝐚0𝝂𝑑s+tΩ[(𝐚)S𝝂+𝐚˙0𝝂]𝑑s\displaystyle\int_{\partial\Omega}\mathbf{a}_{0}\cdot\boldsymbol{\nu}ds+t\int_{\partial\Omega}[(\nabla\cdot\mathbf{a})S\cdot\boldsymbol{\nu}+\dot{\mathbf{a}}_{0}\cdot\boldsymbol{\nu}]ds
+t22Ω[([(𝐚0)S]+2𝐚˙0)S𝝂+𝐚¨0𝝂]\displaystyle+\frac{t^{2}}{2}\int_{\partial\Omega}[(\nabla\cdot[(\nabla\cdot\mathbf{a}_{0})S]+2\nabla\cdot\dot{\mathbf{a}}_{0})S\cdot\boldsymbol{\nu}+\ddot{\mathbf{a}}_{0}\cdot\boldsymbol{\nu}]
+(R(S)S)𝝂ds+o(t2),\displaystyle\quad+(R-(S\cdot\nabla)S)\cdot\boldsymbol{\nu}\ ds+o(t^{2}),

and hence reach the formulae in Lemmas 14 and 19.

Appendix B Second fundamental form on Ω\partial\Omega

In (63) if {Tt}\{T_{t}\} is the normal perturbation (11) we have

(ν)[S,S]=δρνν=0(\nabla\nu)[S,S]=\delta\rho\frac{\partial\nu}{\partial\nu}=0

by (26) and hence

χ=(δρ)2ν,σ=0.\chi=-(\delta\rho)^{2}\nabla\cdot\nu,\quad\sigma=0.

If {Tt}\{T_{t}\} is the dynamical perturbation using (12), it follows that

χ=(δρ)2ν+(vννvτ)δρ,σ=(vννvτ)δρ\chi=-(\delta\rho)^{2}\nabla\cdot\nu+\left(v\cdot\nu\frac{\partial}{\partial\nu}-v\cdot\nabla_{\tau}\right)\delta\rho,\quad\sigma=\left(v\cdot\nu\frac{\partial}{\partial\nu}-v\cdot\nabla_{\tau}\right)\delta\rho

from (65), (69), and (27). In [4] the second Hadamard variation for d=3d=3 is described in accordance with the second fundamental form of Ω\partial\Omega. It is concerned on the normal perturbation (11), where ρ=S𝝂\rho=S\cdot\boldsymbol{\nu} and R=0R=0

These values χ,σ\chi,\sigma used in Theorem 26 are actually associated with the second fundamental form on Ω\partial\Omega, defined by

[ξ,η]=νξη=(𝝂)[ξ,η],ξ,ηd.{\cal B}[\xi,\eta]=-\frac{\partial\nu}{\partial\xi}\cdot\eta=-(\nabla\boldsymbol{\nu})[\xi,\eta],\quad\xi,\eta\in\mathbb{R}^{d}.

The argument relies on the following formula. Recall

δρ=S𝝂,tr=𝝂,Sτ=(Sτ)τ.\delta\rho=S\cdot\boldsymbol{\nu},\quad\mbox{tr}\ {\cal B}=-\nabla\cdot\boldsymbol{\nu},\quad S_{\tau}=(S\cdot\tau)\tau.
Lemma 27.

[5, (3.1.1.8), p.137] It holds that

(S)δρ[(S)S]𝝂\displaystyle(\nabla\cdot S)\delta\rho-[(S\cdot\nabla)S]\cdot\boldsymbol{\nu}
=τ(δρSτ)(tr)(δρ)2(Sτ,Sτ)2(Sττ)δρ.\displaystyle\quad=\nabla_{\tau}\cdot(\delta\rho S_{\tau})-(\mbox{tr}\ {\cal B})(\delta\rho)^{2}-{\cal B}(S_{\tau},S_{\tau})-2(S_{\tau}\cdot\nabla_{\tau})\delta\rho. (82)
Lemma 28.

It holds that

[(SS)]ν+(δρν+Sττ)δρ\displaystyle[(S\cdot\nabla S)]\cdot\nu+\left(-\delta\rho\frac{\partial}{\partial\nu}+S_{\tau}\cdot\nabla_{\tau}\right)\delta\rho
=(ν)(δρ)2+(Sτ,Sτ)+2(Sττ)δρ\displaystyle\quad=-(\nabla\cdot\nu)(\delta\rho)^{2}+{\mathcal{B}}(S_{\tau},S_{\tau})+2(S_{\tau}\cdot\nabla_{\tau})\delta\rho
Proof.

The result follows from a direct computation. First, Lemma 27 implies

(Sττ)δρ+(Sτ,Sτ)+(S)δρ\displaystyle(S_{\tau}\cdot\nabla_{\tau})\delta\rho+{\cal B}(S_{\tau},S_{\tau})+(\nabla\cdot S)\delta\rho
=[(S)S]𝝂+τ(δρSτ)+(𝝂)(δρ)2(Sττ)δρ\displaystyle\quad=[(S\cdot\nabla)S]\cdot\boldsymbol{\nu}+\nabla_{\tau}\cdot(\delta\rho S_{\tau})+(\nabla\cdot\boldsymbol{\nu})(\delta\rho)^{2}-(S_{\tau}\cdot\nabla_{\tau})\delta\rho
=[(S)S]ν+(ν)(δρ)2+δρτSτ\displaystyle\quad=[(S\cdot\nabla)S]\cdot\nu+(\nabla\cdot\nu)(\delta\rho)^{2}+\delta\rho\nabla_{\tau}\cdot S_{\tau}

and hence

[(S)S]𝝂+(𝝂)(δρ)2+(S)δρ\displaystyle[(S\cdot\nabla)S]\cdot\boldsymbol{\nu}+(\nabla\cdot\boldsymbol{\nu})(\delta\rho)^{2}+(S\cdot\nabla)\delta\rho
=δρτSτ+(Sττ)δρ+(Sτ,Sτ)+(S)δρ+(S)δρ\displaystyle\quad=-\delta\rho\nabla_{\tau}\cdot S_{\tau}+(S_{\tau}\cdot\nabla_{\tau})\delta\rho+{\mathcal{B}}(S_{\tau},S_{\tau})+(\nabla\cdot S)\delta\rho+(S\cdot\nabla)\delta\rho
=(Sττ)δρ+(Sτ,Sτ)δρτSτ+(Sδρ).\displaystyle\quad=(S_{\tau}\cdot\nabla_{\tau})\delta\rho+{\mathcal{B}}(S_{\tau},S_{\tau})-\delta\rho\nabla_{\tau}\cdot S_{\tau}+\nabla\cdot(S\delta\rho). (83)

Then we use

(S)δρ=(δρν+Sττ)δρ(S\cdot\nabla)\delta\rho=\left(\delta\rho\frac{\partial}{\partial\nu}+S_{\tau}\cdot\nabla_{\tau}\right)\delta\rho

and

(Sδρ)\displaystyle\nabla\cdot(S\delta\rho) =\displaystyle= τ(Sτδρ)+ν(δρ)2\displaystyle\nabla_{\tau}(S_{\tau}\delta\rho)+\frac{\partial}{\partial\nu}(\delta\rho)^{2}
=\displaystyle= δρτSτ+(Sττ)δρ+2δρδρν,\displaystyle\delta\rho\nabla_{\tau}\cdot S_{\tau}+(S_{\tau}\cdot\nabla_{\tau})\delta\rho+2\delta\rho\frac{\partial\delta\rho}{\partial\nu},

which implies

[(S)S]ν+(ν)(δρ)2=(Sτ,Sτ)+(Sττ+δρν)δρ[(S\cdot\nabla)S]\cdot\nu+(\nabla\cdot\nu)(\delta\rho)^{2}={\mathcal{B}}(S_{\tau},S_{\tau})+\left(S_{\tau}\cdot\nabla_{\tau}+\delta\rho\frac{\partial}{\partial\nu}\right)\delta\rho

by (83). Then the result follows. ∎

Theorem 29.

It holds that

χ=δ2ρ2(Sττ)δρ(Sτ,Sτ)(δρ)2(𝝂)\displaystyle\chi=\delta^{2}\rho-2(S_{\tau}\cdot\nabla_{\tau})\delta\rho-{\cal B}(S_{\tau},S_{\tau})-(\delta\rho)^{2}(\nabla\cdot\boldsymbol{\nu})
σ=δ2ρ2(Sττ)δρ(Sτ,Sτ)\displaystyle\sigma=\delta^{2}\rho-2(S_{\tau}\cdot\nabla_{\tau})\delta\rho-{\cal B}(S_{\tau},S_{\tau})

in Theorem 26.

Proof.

Equalities (65) and (69) imply

χ=[R(S)S]ν(δρ)2ν+(δρν(Sττ))δρ\chi=[R-(S\cdot\nabla)S]\cdot\nu-(\delta\rho)^{2}\nabla\cdot\nu+\left(\delta\rho\frac{\partial}{\partial\nu}-(S_{\tau}\cdot\nabla_{\tau})\right)\delta\rho

and

σ=[R(S)S]ν+(δρ𝝂(Sττ))δρ,\sigma=[R-(S\cdot\nabla)S]\cdot\nu+\left(\delta\rho\frac{\partial}{\partial\boldsymbol{\nu}}-(S_{\tau}\cdot\nabla_{\tau})\right)\delta\rho,

respectively. Then we obtain the result by (82).

References

  • [1] H. Azegami, Shape Optimization Problems, Springer, Singapore, 2020.
  • [2] L.C. Evans and R.F. Gariepy, Measure Theory and Fine Properties of Functions, CRC Press, Boca Raton, 1992.
  • [3] H. Flanders, Differential Forms with Applications to the Physical Sciences, Academic Press, New York, 1963.
  • [4] P.R. Garabedian and M. Schiffer, Convexity of domain functionals, J. Ann. Math. 2 (1952-53) 281–368.
  • [5] P. Grisvard, Elliptic Problems in Nonsmooth Domains, Pitman, Boston, 1985.
  • [6] T. Suzuki, T. Tsuchiya, Convergence analysis of trial free boundary methods for the two-dimensional filtration problem, Numer. Math. 100 (2005) 537–564.
  • [7] T. Suzuki, T. Tsuchiya, Weak formulation of Hadamard variation applied to the filtration problem, Japan. J. Indus. Appl. Math. 28 (2011) 327–350
  • [8] 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.

Acknowledgements. This work was promoted in Research Institute for Mathematical Sciences (RIMS) Joint Research Program during 2019–2021. (RIMS is an International Joint Usage/Research Center located in Kyoto University.) 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.