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

Asymptotic convergence of evolving hypersurfaces

Carlo Mantegazza Carlo Mantegazza
Dipartimento di Matematica e Applicazioni, Università di Napoli Federico II, Via Cintia, Monte S. Angelo 80126 Napoli, Italy
c.mantegazza@sns.it
 and  Marco Pozzetta Marco Pozzetta
Dipartimento di Matematica e Applicazioni, Università di Napoli Federico II, Via Cintia, Monte S. Angelo 80126 Napoli, Italy
marco.pozzetta@unina.it
(Date: August 6, 2025)
Abstract.

If ψ:Mnn+1\psi:M^{n}\to\mathbb{R}^{n+1} is a smooth immersed closed hypersurface, we consider the functional

m(ψ)=M1+|mν|2dμ,\mathcal{F}_{m}(\psi)=\int_{M}1+|\nabla^{m}\nu|^{2}\,d\mu,

where ν\nu is a local unit normal vector along ψ\psi, \nabla is the Levi–Civita connection of the Riemannian manifold (M,g)(M,g), with gg the pull–back metric induced by the immersion and μ\mu the associated volume measure. We prove that if m>n/2m>\lfloor n/2\rfloor then the unique globally defined smooth solution to the L2L^{2}–gradient flow of m\mathcal{F}_{m}, for every initial hypersurface, smoothly converges asymptotically to a critical point of m\mathcal{F}_{m}, up to diffeomorphisms. The proof is based on the application of a Łojasiewicz–Simon gradient inequality for the functional m\mathcal{F}_{m}.

Key words and phrases:
Geometric flows, Łojasiewicz–Simon gradient inequality, smooth convergence
2010 Mathematics Subject Classification:
53E40, 35R01, 46N20

1. Introduction

We consider a closed connected differentiable manifold MnM^{n} of dimension n1n\geq 1 and ψ:Mnn+1\psi:M^{n}\to\mathbb{R}^{n+1} a smooth immersion of MnM^{n} in the Euclidean space n+1\mathbb{R}^{n+1}. We shall usually omit the superscript nn denoting the dimension of MM. For such an immersion, we always assume that MM is endowed with the metric tensor g=ψ,n+1g=\psi^{*}\langle\cdot,\cdot\rangle_{\mathbb{R}^{n+1}}, induced by the immersion ψ\psi that we also sometimes simply denote as ,\langle\cdot,\cdot\rangle. The Levi–Civita connection of the Riemannian manifold (M,g)(M,g) is denoted by \nabla and the associated volume measure by μ\mu. Then, for mm\in\mathbb{N} with m1m\geq 1, we consider the functional

m(ψ)M1+|mν|2dμ\mathcal{F}_{m}(\psi)\coloneqq\int_{M}1+|\nabla^{m}\nu|^{2}\,d\mu

on the smooth immersions ψ:Mnn+1\psi:M^{n}\to\mathbb{R}^{n+1}, where ν\nu is a (locally defined) unit normal vector field along ψ\psi. Let us specify that in the above definition, if ν=ναeα\nu=\nu^{\alpha}e_{\alpha} and eαe_{\alpha} is the standard basis in n+1\mathbb{R}^{n+1}, we mean |mν|2α=1n+1|mνα|2|\nabla^{m}\nu|^{2}\coloneqq\sum_{\alpha=1}^{n+1}|\nabla^{m}\nu^{\alpha}|^{2}. Notice that m\mathcal{F}_{m} is independent of the local choice of the unit normal ν\nu and it is well-defined without further hypotheses on MM. However, by the discussion below in Remark 1.2, we shall always assume without loss of generality that MM is orientable and that a global choice of unit normal field ν\nu is understood.

We observe that in case n=m=1n=m=1, we recognize the well-known elastic energy of closed curves ψ:𝕊12\psi:\mathbb{S}^{1}\to\mathbb{R}^{2}, i.e., 1(ψ)=𝕊11+|k|2ds\mathcal{F}_{1}(\psi)=\int_{\mathbb{S}^{1}}1+|k|^{2}ds, where kk is the curvature of ψ\psi. If instead n=2n=2, m=1m=1, then one has |ν|2=|A|2|\nabla\nu|^{2}=|{\mathrm{A}}|^{2}, where A{\mathrm{A}} is the second fundamental form (see (1.1)), and then 1\mathcal{F}_{1} yields the sum of the area and of (an equivalent form of) the Willmore energy of an immersed surface ψ:M23\psi:M^{2}\to\mathbb{R}^{3}. Let us also notice that if instead n=m=2n=m=2, then 2(ψ)=M1+|A|2+𝔭0(A,A,A,A)dμ\mathcal{F}_{2}(\psi)=\int_{M}1+|\nabla{\mathrm{A}}|^{2}+\mathfrak{p}_{0}({\mathrm{A}},{\mathrm{A}},{\mathrm{A}},{\mathrm{A}})d\mu for any given immersion ψ:M23\psi:M^{2}\to\mathbb{R}^{3}, where 𝔭0\mathfrak{p}_{0} is some polynomial as in (1.3) below.

By the formula for the first variation of m\mathcal{F}_{m} (see Theorem 2.1), one can prove that the associated L2L^{2}–gradient flow is defined by an evolution equation

φt(p,t)=Em(φt)(p)νt(p)\frac{\partial\varphi}{\partial t}(p,t)=-{\mathrm{E}}_{m}(\varphi_{t})(p)\nu_{t}(p)

for a smooth map φ:M×[0,T)n+1\varphi:M\times[0,T)\to\mathbb{R}^{n+1} (where φt=φ(,t):Mn+1\varphi_{t}=\varphi(\cdot,t):M\to\mathbb{R}^{n+1} describes the moving hypersurface and νt\nu_{t} is its unit normal vector field), which turns out to be a (quasilinear and degenerate) parabolic system of PDEs.
If m>n/2m>\lfloor n/2\rfloor, the study carried out in [17] shows that for every initial smooth immersed hypersurface φ0:Mn+1\varphi_{0}:M\to\mathbb{R}^{n+1}, there exists a unique smooth solution φt\varphi_{t} with initial datum φ0\varphi_{0}, defined for all positive times; moreover, φt\varphi_{t} sub–converges to a critical point φ:Mn+1\varphi_{\infty}:M\to\mathbb{R}^{n+1} of the functional m\mathcal{F}_{m}, that is, such that Em(φ)=0{\mathrm{E}}_{m}(\varphi_{\infty})=0 (see Theorem 4.3). By sub–convergence we mean that for some sequence of times tj+t_{j}\to+\infty, the sequence φtj\varphi_{t_{j}} smoothly converges to φ\varphi_{\infty}, up to diffeomorphisms and translations in n+1\mathbb{R}^{n+1}. More precisely, there exist a sequence of smooth diffeomorphisms σj:MM\sigma_{j}:M\to M and a sequence of points pjn+1p_{j}\in\mathbb{R}^{n+1} such that the sequence of immersions φtjσjpj\varphi_{t_{j}}\circ\sigma_{j}-p_{j} converge to φ\varphi_{\infty} in Ck(M)C^{k}(M), for any kk\in\mathbb{N}. From such a sub–convergence result it is anyway not possible to immediately deduce that the flow fully converges, i.e., that there exists the full limit of φt\varphi_{t} as t+t\to+\infty in Ck(M)C^{k}(M) for any kk (up to diffeomorphisms). Actually, the sub–convergence of the flow does not even guarantee that the limits of the flow along different diverging sequences of times coincide. Moreover, as the evolution equations involved here are of order greater or equal than four with respect to the parametrization, it is not even possible to conclude that the flow stays in a compact set of n+1\mathbb{R}^{n+1} for all times by means of comparison arguments, as maximum principles are not applicable.

In this work we address this issue, that is, we prove that the gradient flow of m\mathcal{F}_{m} does actually converge, for any initial hypersurface. Our main result is the following theorem.

Theorem 1.1.

Let φ0:Mnn+1\varphi_{0}:M^{n}\to\mathbb{R}^{n+1} be a smooth immersion of a closed hypersurface and let m>n/2m>\lfloor n/2\rfloor. Then the unique smooth solution φ:M×[0,+)n+1\varphi:M\times[0,+\infty)\to\mathbb{R}^{n+1} to the evolution problem

{φt=Em(φt)νtφ(0,)=φ0\begin{cases}\frac{\partial\varphi}{\partial t}=-{\mathrm{E}}_{m}(\varphi_{t})\nu_{t}\\ \varphi(0,\cdot)=\varphi_{0}\end{cases}

converges in Ck(M)C^{k}(M) to a smooth critical point φ:Mn+1\varphi_{\infty}:M\to\mathbb{R}^{n+1} of m\mathcal{F}_{m} as t+t\to+\infty, for every kk\in\mathbb{N} up to diffeomorphisms of MM; more precisely, there exists a one-parameter family of diffeomorphisms σt:MM\sigma_{t}:M\to M such that the flow φtσt\varphi_{t}\circ\sigma_{t} converges in Ck(M)C^{k}(M) to a smooth critical point φ\varphi_{\infty} of m\mathcal{F}_{m} as t+t\to+\infty, for every kk\in\mathbb{N}.
In particular, there exists a compact set Kn+1K\subseteq\mathbb{R}^{n+1} such that Mt=φt(M)KM_{t}=\varphi_{t}(M)\subseteq K for any t0t\geq 0.

We remark that the assumption m>n/2m>\lfloor n/2\rfloor is sharp, in fact if mn/2m\leq\lfloor n/2\rfloor then one gets flows that may develop singularities in finite time.

A relevant motivation for the study of the gradient flow of the functionals m\mathcal{F}_{m} goes back to Ennio De Giorgi. In one of his last papers, he conjectured that any compact nn–dimensional hypersurface in n+1\mathbb{R}^{n+1}, evolving by the gradient flow of certain functionals depending on sufficiently high derivatives of the curvature does not develop singularities during the flow (see [10] and [11, Section 5, Conjecture 2] for an English translation, see also [17, Section 9]). This result was central in his program to approximate singular geometric flows, as the mean curvature flow, with sequences of smooth ones (see [17, Sec. 9] and [2] for a result in this direction). The functionals m\mathcal{F}_{m} are strictly related to the ones proposed by De Giorgi since, roughly speaking, the derivative of the normal field yields the curvature of MM (see (1.1)). Though not exactly the same, the energies m\mathcal{F}_{m} can then play the same role in the approximation process he suggested and the analysis of the asymptotic behavior of their gradient flow is another step in understanding such process.

The main tool in the proof of Theorem 1.1 is a Łojasiewicz–Simon gradient inequality for the functional m\mathcal{F}_{m} (see Corollary 4.2). Such an estimate bounds a less–than–1/21/2 power of the difference in “energy” (the value of the functional) between a critical point and a point sufficiently close to it in terms of a suitable norm of the first variation of the functional. The use of this kind of inequalities in the study of the convergence of parabolic equations of gradient–type goes back to Łojasiewicz [15, 16] and to the seminal paper of Simon [23]. More recently, useful sufficient hypotheses implying a Łojasiewicz–Simon gradient inequality have been derived in [4] (see also [13] for several recent generalizations). Building on the abstract tools developed in [4], a first recent application of the inequality to get convergence of an extrinsic geometric flow is contained in [5], where the authors investigate the Willmore flow of surfaces in neighborhoods of critical points. In the last years and in the context of higher order geometric gradient flows, the Łojasiewicz–Simon inequality appeared as a tool for “promoting” the sub–convergence of a flow to its full convergence. As applications of this method we mention [9], in which it is proved the full convergence of the elastic flow of open clamped curves, and [20], in which the sub–convergence of the pp–elastic flow of closed curves on Riemannian manifolds is shown to imply the full convergence. The analysis in [20] led to a further simplification and deeper understanding of the method, which is exposed in [19]. In this work we essentially generalize the strategy employed in [19] to the gradient flows of the functionals m\mathcal{F}_{m}. Moreover we tried to keep most of the arguments as general as possible, in order that the method could be possibly applied also to other geometric gradient flows, also in the context of extrinsic geometric flows in higher codimension possibly in Riemannian manifolds.

In the broad framework of geometric flows, Łojasiewicz–Simon gradient inequalities found many other notable applications. For example, the study of singularities of mean curvature flow can be reconducted to the study of the smooth convergence of a suitable extrinsic geometric flow. Smooth convergence of such flow has been proved exploiting Łojasiewicz–Simon inequalities in some relevant particular cases in [22, 8, 6]. Let us also mention [7], where a classification of ancient solutions to a family of geometric flows in Riemannian manifolds is derived. Łojasiewicz–Simon inequalities have been employed also in the context of intrinsic geometric flows. We refer, for instance, to the study of the rate of convergence of Yamabe flows in [3], or to the deep investigation on the Yang–Mills flow contained in [12] (see also references therein).

Notation and geometry of submanifolds

Let MnM^{n} be closed, connected, and orientable. Let ψ:Mn+1\psi:M\to\mathbb{R}^{n+1} be a smooth immersion of MM and let ν\nu be global unit normal field on MM along ψ\psi.

Remark 1.2.

In case MM is not orientable, given an initial immersion φ0:Mn+1\varphi_{0}:M\to\mathbb{R}^{n+1}, we can consider the canonical two–fold cover π:M~M\pi:\widetilde{M}\to M, where M~\widetilde{M} is orientable and the initial immersion φ~0=φ0π\widetilde{\varphi}_{0}=\varphi_{0}\circ\pi. By uniqueness of the flow φt\varphi_{t} starting at φ0\varphi_{0} (Theorem 4.3), it follows that the flow φ~t\widetilde{\varphi}_{t} starting at φ~0\widetilde{\varphi}_{0} is just φ~t=φtπ\widetilde{\varphi}_{t}=\varphi_{t}\circ\pi. Therefore, if we prove that φ~t\widetilde{\varphi}_{t} smoothly converges, then the same holds for the flow φt\varphi_{t}. Hence, also in this case Theorem 1.1 holds.

As the metric gg is obtained pulling it back with ψ\psi, in local coordinates {xi}\{x_{i}\} on MM, we have

gij(x)=ψ(x)xi,ψ(x)xjg_{ij}(x)=\left\langle\frac{\partial{{\psi}}(x)}{\partial x_{i}},\frac{\partial{{\psi}}(x)}{\partial x_{j}}\right\rangle

and the canonical volume measure induced by the metric gg is given in local coordinates by μ=det(gij)n\mu=\sqrt{\det(g_{ij})\,}\,{\mathcal{L}}^{n} where n{\mathcal{L}}^{n} is the standard Lebesgue measure on n\mathbb{R}^{n}.
The induced covariant derivative on (M,g)(M,g) of a tangent vector field XX is given by

jXi=xjXi+ΓjkiXk\nabla_{j}X^{i}=\frac{\partial\,}{\partial x_{j}}X^{i}+\Gamma^{i}_{jk}X^{k}

(in the whole paper we will adopt the Einstein convention of summation over repeated indices) where the Christoffel symbols Γjki\Gamma^{i}_{jk} are expressed by the formula

Γjki=12gil(xjgkl+xkgjlxlgjk).\Gamma^{i}_{jk}=\frac{1}{2}g^{il}\left(\frac{\partial\,}{\partial x_{j}}g_{kl}+\frac{\partial\,}{\partial x_{k}}g_{jl}-\frac{\partial\,}{\partial x_{l}}g_{jk}\right)\,.

We will write i\partial_{i} for the coordinates derivatives, opposite to the covariant ones i\nabla_{i}. With kT\nabla^{k}T we will mean the kk–th iterated covariant derivative of a tensor TT. If ff is a smooth function on a smooth immersed hypersurface, the symbol f\nabla f denotes its gradient and 2f\nabla^{2}f its Hessian, whose trace is the Laplacian Δf\Delta f.

The second fundamental form A{\mathrm{A}} of the immersion ψ\psi is the bilinear symmetric form acting on any pair of tangent vector fields X,YX,Y to the hypersurface as

A(X,Y)=Xn+1Y,ν,{\mathrm{A}}(X,Y)=-\bigl{\langle}\nabla^{\mathbb{R}^{n+1}}_{X}Y,\nu\rangle,

given a (global, since we assumed MM orientable) choice of the unit normal vector ν\nu (we will usually identify TMTM with dψ(TM)n+1d\psi(TM)\subseteq\mathbb{R}^{n+1} and in this formula the field YY is extended locally around ψ(M)\psi(M) in n+1\mathbb{R}^{n+1}). Hence A{\mathrm{A}} is defined up to a sign, that is, up to the choice of ν\nu, while Aν{\mathrm{A}}\nu is independent of the choice of ν\nu. In local coordinates, the components hijh_{ij} of A{\mathrm{A}} are given by

hij(x)=ν(x),2ψ(x)xixj.h_{ij}(x)=-\left\langle\nu(x),\frac{\partial^{2}{{\psi}}(x)}{\partial x_{i}\partial x_{j}}\right\rangle\,.

We recall that the following Gauss–Weingarten relations hold

ij2ψ=Γijkkψhijν,iν=hijgjkkψ.\partial^{2}_{ij}\psi=\Gamma^{k}_{ij}\partial_{k}\psi-h_{ij}\nu,\qquad\partial_{i}\nu=h_{ij}g^{jk}\partial_{k}\psi. (1.1)

The mean curvature H{\mathrm{H}} of ψ\psi is the trace of A{\mathrm{A}}, that is

H(x)=gij(x)hij(x).{\mathrm{H}}(x)=g^{ij}(x)h_{ij}(x).

By means of the Gauss equation, the Riemann tensor can be expressed via the second fundamental form, in local coordinates, as follows

Rijkl=\displaystyle{\mathrm{R}}_{ijkl}\,= hikhjlhilhjk.\displaystyle\,h_{ik}h_{jl}-h_{il}h_{jk}\,.

Hence, the formulae for the interchange of covariant derivatives become

ijXsjiXs=RijklgksXl=RijlsXl=(hikhjlhilhjk)gksXl,\nabla_{i}\nabla_{j}X^{s}-\nabla_{j}\nabla_{i}X^{s}={\mathrm{R}}_{ijkl}g^{ks}X^{l}={\mathrm{R}}_{ijl}^{s}X^{l}=\left(h_{ik}h_{jl}-h_{il}h_{jk}\right)g^{ks}X^{l}\,,
ijωkjiωk=Rijklglsωs=Rijksωs=(hikhjlhilhjk)glsωs,\nabla_{i}\nabla_{j}\omega_{k}-\nabla_{j}\nabla_{i}\omega_{k}={\mathrm{R}}_{ijkl}g^{ls}\omega_{s}={\mathrm{R}}_{ijk}^{s}\omega_{s}=\left(h_{ik}h_{jl}-h_{il}h_{jk}\right)g^{ls}\omega_{s}\,, (1.2)

where we recall that by ijXs\nabla_{i}\nabla_{j}X^{s} we mean the ss-th component of the field (2X)(i,j)(\nabla^{2}X)(\partial_{i},\partial_{j}).

Abusing a little the notation, if T1,,TNT_{1},...,T_{N} is a finite family of tensors, we denote by

k=1NTkT1TN\circledast_{k=1}^{N}T_{k}\coloneqq T_{1}*\ldots*T_{N}

a generic contraction of some indices of the tensors T1,,TNT_{1},...,T_{N} using the coefficients gijg_{ij} or gijg^{ij}. We will also denote

𝔭s(T1,,TN)i1++iN=sCi1,,iNi1T1iNTN,\mathfrak{p}_{s}(T_{1},\ldots,T_{N})\coloneqq\sum_{i_{1}+\ldots+i_{N}=s}C_{i_{1},\ldots,i_{N}}\nabla^{i_{1}}T_{1}*\ldots*\nabla^{i_{N}}T_{N}, (1.3)

for some constants Ci1,,iNC_{i_{1},\ldots,i_{N}}\in\mathbb{R}. Notice that in every additive term of 𝔭s(T1,,TN)\mathfrak{p}_{s}(T_{1},\ldots,T_{N}) each tensor appears exactly once (there are no repetitions).
We will use instead the symbol 𝔮s(T1,,TN)\mathfrak{q}^{s}(T_{1},\ldots,T_{N}) for “polynomials” of the form

𝔮s(T1,,TN)[i1=1M1i1T1iN=1MNiNTN],\mathfrak{q}^{s}(T_{1},\ldots,T_{N})\coloneqq\sum\left[\circledast_{i_{1}=1}^{M_{1}}\nabla^{i_{1}}T_{1}\,\ldots\,\circledast_{i_{N}=1}^{M_{N}}\nabla^{i_{N}}T_{N}\right],

with Mj1M_{j}\geq 1 for any j=1,,Nj=1,...,N and with

s=i1=1M1(i1+1)++iN=1MN(iN+1).s=\sum_{i_{1}=1}^{M_{1}}(i_{1}+1)+\ldots+\sum_{i_{N}=1}^{M_{N}}(i_{N}+1).

Hence, repetitions are allowed in 𝔮s\mathfrak{q}^{s} and in every additive term there must be present every argument of 𝔮s\mathfrak{q}^{s}.

We notice that, by the above relations, the Riemann tensor of the hypersurface can be written as R=AA{\mathrm{R}}={\mathrm{A}}*{\mathrm{A}}, exploiting the above notation.

2. Preliminary computations

Let us recall the first variation formula for the functional m\mathcal{F}_{m}.

Theorem 2.1 ([17, Theorem 3.7]).

Let φt:Mnn+1\varphi_{t}:M^{n}\to\mathbb{R}^{n+1} be a smooth family of immersions smoothly depending on t(ε,ε)t\in(-\varepsilon,\varepsilon) and Xt=tφtX_{t}=\partial_{t}\varphi_{t}. Then, for every t(ε,ε)t\in(-\varepsilon,\varepsilon), there holds

ddtm(φt)=MEm(φt)ν,Xt𝑑μt,\frac{d}{dt}\mathcal{F}_{m}(\varphi_{t})=\int_{M}{\mathrm{E}}_{m}(\varphi_{t})\langle\nu,X_{t}\rangle\,d\mu_{t},

with

Em(φt)=2(1)mΔmH+𝔮2m+1(ν,A)+H,{\mathrm{E}}_{m}(\varphi_{t})=2(-1)^{m}\Delta^{m}{\mathrm{H}}+\mathfrak{q}^{2m+1}(\nabla\nu,{\mathrm{A}})+{\mathrm{H}},

where all the quantities are relative to the hypersurface φt\varphi_{t}.

The next lemma states the evolution formulae for the geometric quantities that we need in the computation of the second variation of the functional m\mathcal{F}_{m}.

Lemma 2.2.

Let φt:Mnn+1\varphi_{t}:M^{n}\to\mathbb{R}^{n+1} be a smooth family of immersions smoothly depending on t(ε,ε)t\in(-\varepsilon,\varepsilon) and φ=φ0\varphi=\varphi_{0}. Let X=tφt|t=0X=\partial_{t}\varphi_{t}|_{t=0} and assume that XX is a normal vector field along φ\varphi. Then, we have

tgij|t=0=2ν,Xhij,\partial_{t}g_{ij}|_{t=0}=2\langle\nu,X\rangle h_{ij}\,,
tgij|t=0=2ν,Xgikgjlhkl,\partial_{t}g^{ij}|_{t=0}=-2\langle\nu,X\rangle g^{ik}g^{jl}h_{kl}\,,
tν|t=0=ν,X,\partial_{t}\nu|_{t=0}=-\nabla\langle\nu,X\rangle\,,
tΓijk|t=0=Aν,X+Aν,X,\partial_{t}\Gamma^{k}_{ij}\bigr{|}_{t=0}=\nabla{\mathrm{A}}*\langle\nu,X\rangle+{\mathrm{A}}*\nabla\langle\nu,X\rangle\,,
thij|t=0=ij2ν,X+ν,Xhij2,\partial_{t}h_{ij}|_{t=0}=-\nabla^{2}_{ij}\langle\nu,X\rangle+\langle\nu,X\rangle h^{2}_{ij}\,, (2.1)
tH|t=0=Δν,Xν,X|A|2,\partial_{t}{\mathrm{H}}|_{t=0}=-\Delta\langle\nu,X\rangle-\langle\nu,X\rangle|{\mathrm{A}}|^{2}\,, (2.2)
tΔmf|t=0Δmtf|t=0=𝔭2m(f0,A,ν,X),\partial_{t}\Delta^{m}f|_{t=0}-\Delta^{m}\partial_{t}f|_{t=0}=\mathfrak{p}_{2m}(f_{0},{\mathrm{A}},\langle\nu,X\rangle)\,, (2.3)

for any smooth function fC(M×(ε,ε))f\in C^{\infty}(M\times(-\varepsilon,\varepsilon)) and mm\in\mathbb{N} with m1m\geq 1, where f0=f(,0)f_{0}=f(\cdot,0) and

t𝔮2m+1(ν,A)|t=0=𝔮2m+3(ν,X,ν,A)+ν,X𝔮2m+2(ν,A).\partial_{t}\mathfrak{q}^{2m+1}(\nabla\nu,{\mathrm{A}})|_{t=0}=\mathfrak{q}^{2m+3}(\langle\nu,X\rangle,\nabla\nu,{\mathrm{A}})+\langle\nu,X\rangle\mathfrak{q}^{2m+2}(\nabla\nu,{\mathrm{A}})\,. (2.4)
Proof.

The first four formulae are computed explicitly at page 150 of [17].

By means of the Gauss–Weingarten relations (1.1), setting X=βνX=\beta\nu, hence ν,X=β\langle\nu,X\rangle=\beta, we compute

thij|t=0=\displaystyle\partial_{t}h_{ij}|_{t=0}= tν,ij2φt|t=0\displaystyle\,-\partial_{t}\langle\nu,\partial^{2}_{ij}\varphi_{t}\rangle|_{t=0}
=\displaystyle= ν,ij2(βν)+β,ij2φ\displaystyle\,-\langle\nu,\partial^{2}_{ij}(\beta\nu)\rangle+\langle\nabla\beta,\partial^{2}_{ij}\varphi\rangle
=\displaystyle\,= ij2ββν,i(hjlglkkφ)+lβglssφ,Γijkkφhijν\displaystyle\,-\partial^{2}_{ij}\beta-\beta\langle\nu,\partial_{i}(h_{jl}g^{lk}\partial_{k}\varphi)\rangle+\langle\partial_{l}\beta g^{ls}\partial_{s}\varphi,\Gamma_{ij}^{k}\partial_{k}\varphi-h_{ij}\nu\rangle
=\displaystyle\,= ij2ββν,hjlglkik2φ+kβΓijk\displaystyle\,-\partial^{2}_{ij}\beta-\beta\langle\nu,h_{jl}g^{lk}\partial^{2}_{ik}\varphi\rangle+\partial_{k}\beta\Gamma_{ij}^{k}
=\displaystyle\,= ij2β+βhikgklhlj\displaystyle\,-\nabla^{2}_{ij}\beta+\beta h_{ik}g^{kl}h_{lj}

that is, thij|t=0=ij2ν,X+ν,Xhij2\partial_{t}h_{ij}|_{t=0}=-\nabla^{2}_{ij}\langle\nu,X\rangle+\langle\nu,X\rangle h^{2}_{ij}, hence it follows

tH|t=0=t(gijhij)|t=0=2ν,X|A|2Δν,X+ν,X|A|2=Δν,Xν,X|A|2.\partial_{t}{\mathrm{H}}|_{t=0}=\partial_{t}(g^{ij}h_{ij})|_{t=0}=-2\langle\nu,X\rangle|{\mathrm{A}}|^{2}-\Delta\langle\nu,X\rangle+\langle\nu,X\rangle|{\mathrm{A}}|^{2}=-\Delta\langle\nu,X\rangle-\langle\nu,X\rangle|{\mathrm{A}}|^{2}. (2.5)

We now deal with equation (2.3) arguing by induction on m1m\geq 1. Using the previous evolution formulae, for m=1m=1 we compute

tΔf|t=0=t(gij(ij2fΓijkkf))|t=0=2ν,Xgikgjlhklij2f0+Δt|t=0fgij(Aν,X+Aν,X)kf0,\begin{split}\partial_{t}\Delta f|_{t=0}&=\partial_{t}(g^{ij}(\partial^{2}_{ij}f-\Gamma^{k}_{ij}\partial_{k}f))|_{t=0}\\ &=-2\langle\nu,X\rangle g^{ik}g^{jl}h_{kl}\nabla^{2}_{ij}f_{0}+\Delta\partial_{t}|_{t=0}f-g^{ij}(\nabla{\mathrm{A}}*\langle\nu,X\rangle+{\mathrm{A}}*\nabla\langle\nu,X\rangle)\partial_{k}f_{0},\end{split}

and the claim follows. Now for m+11m+1\geq 1, by induction we get

tΔm+1f|t=0=Δ(tΔmf)|t=0+𝔭2(Δmf0,A,ν,X)=Δ(Δmtf|t=0+𝔭2m(f0,A,ν,X))+𝔭2m+2(f0,A,ν,X).\begin{split}\partial_{t}\Delta^{m+1}f|_{t=0}&=\Delta(\partial_{t}\Delta^{m}f)|_{t=0}+\mathfrak{p}_{2}(\Delta^{m}f_{0},{\mathrm{A}},\langle\nu,X\rangle)\\ &=\Delta\left(\Delta^{m}\partial_{t}f|_{t=0}+\mathfrak{p}_{2m}(f_{0},{\mathrm{A}},\langle\nu,X\rangle)\right)+\mathfrak{p}_{2m+2}(f_{0},{\mathrm{A}},\langle\nu,X\rangle).\end{split}

Finally, in order to show equation (2.4), we need to differentiate a generic term of the form

k=1Nikνl=1MjlA,\circledast_{k=1}^{N}\nabla^{i_{k}}\nabla\nu\circledast_{l=1}^{M}\nabla^{j_{l}}{\mathrm{A}},

with k=1N(ik+1)+l=1M(jl+1)=2m+1\sum_{k=1}^{N}(i_{k}+1)+\sum_{l=1}^{M}(j_{l}+1)=2m+1.
For any component να\nu^{\alpha} of ν\nu we can apply [17, Proposition 3.6] in order to get

t(ikνα)|t=0=ik+1αν,X+𝔭ik(ν,X,ν,A),\partial_{t}(\nabla^{i_{k}}\nabla\nu^{\alpha})|_{t=0}=-\nabla^{i_{k}+1}\nabla^{\alpha}\langle\nu,X\rangle+\mathfrak{p}_{i_{k}}(\langle\nu,X\rangle,\nabla\nu,{\mathrm{A}}),

where αν,X\nabla^{\alpha}\langle\nu,X\rangle denotes the α\alpha–th component in n+1\mathbb{R}^{n+1} of the gradient ν,X\nabla\langle\nu,X\rangle. Also, by [17, Lemma 3.5] and formula (2.1), we have

t(jlA)|t=0=jl(2ν,X+ν,XAA)+𝔭jl(A,A,ν,X)=jl+2ν,X+𝔭jl(A,A,ν,X).\begin{split}\partial_{t}(\nabla^{j_{l}}{\mathrm{A}})|_{t=0}&=\nabla^{j_{l}}(-\nabla^{2}\langle\nu,X\rangle+\langle\nu,X\rangle{\mathrm{A}}*{\mathrm{A}})+\mathfrak{p}_{j_{l}}({\mathrm{A}},{\mathrm{A}},\langle\nu,X\rangle)\\ &=-\nabla^{j_{l}+2}\langle\nu,X\rangle+\mathfrak{p}_{j_{l}}({\mathrm{A}},{\mathrm{A}},\langle\nu,X\rangle).\end{split}

Therefore, using these formulae and the ones above for the derivative of the metric gijg_{ij} and its inverse gijg^{ij}, formula (2.4) follows. ∎

We can now compute the second variation of m\mathcal{F}_{m}.

Theorem 2.3.

Let φt:Mnn+1\varphi_{t}:M^{n}\to\mathbb{R}^{n+1} be a smooth family of immersions smoothly depending on t(ε,ε)t\in(-\varepsilon,\varepsilon). Denote φ=φ0\varphi=\varphi_{0} and assume that φ\varphi is a critical point for m\mathcal{F}_{m}, i.e., Em(φ)=0{\mathrm{E}}_{m}(\varphi)=0. Let X=tφt|t=0X=\partial_{t}\varphi_{t}|_{t=0} and assume that XX is normal along φ\varphi. Then

d2dt2m(φt)|t=0=M(2(1)m+1Δm+1ν,X+Ω(ν,X))ν,X𝑑μ,\frac{d^{2}}{dt^{2}}\mathcal{F}_{m}(\varphi_{t})\bigg{|}_{t=0}=\int_{M}\left(2(-1)^{m+1}\Delta^{m+1}\langle\nu,X\rangle+\Omega(\langle\nu,X\rangle)\right)\langle\nu,X\rangle\,d\mu,

where Ω(ν,X)\Omega(\langle\nu,X\rangle) is linear in ν,X\langle\nu,X\rangle and depends on its covariant derivatives of order 2m2m at most.

Proof.

By Theorem 2.1 we have

d2dt2m(φt)|t=0=ddtMEm(φt)ν,tφt𝑑μt|t=0=M[tEm(φt)]|t=0ν,Xdμ,\begin{split}\frac{d^{2}}{dt^{2}}\mathcal{F}_{m}(\varphi_{t})\bigg{|}_{t=0}&=\frac{d}{dt}\int_{M}{\mathrm{E}}_{m}(\varphi_{t})\langle\nu,\partial_{t}\varphi_{t}\rangle\,d\mu_{t}\bigg{|}_{t=0}=\int_{M}\left[\frac{\partial}{\partial t}{\mathrm{E}}_{m}(\varphi_{t})\right]\bigg{|}_{t=0}\langle\nu,X\rangle\,d\mu,\end{split}

as Em(φ)=0{\mathrm{E}}_{m}(\varphi)=0. Using the explicit expression for Em(φt){\mathrm{E}}_{m}(\varphi_{t}) (Theorem 2.1), applying formula (2.3) with f=Hf={\mathrm{H}} and equations (2.2), (2.4), we get

ddtEm(φt)|t=0\displaystyle\frac{d}{dt}{\mathrm{E}}_{m}(\varphi_{t})\bigg{|}_{t=0} =2(1)m+1Δm(Δν,X+ν,X|A|2)+𝔭2m(H,A,ν,X)\displaystyle=2(-1)^{m+1}\Delta^{m}(\Delta\langle\nu,X\rangle+\langle\nu,X\rangle|{\mathrm{A}}|^{2})+\mathfrak{p}_{2m}({\mathrm{H}},{\mathrm{A}},\langle\nu,X\rangle)
+𝔮2m+3(ν,X,ν,A)+ν,X𝔮2m+2(ν,A)(Δν,X+ν,X|A|2)\displaystyle\indent+\mathfrak{q}^{2m+3}(\langle\nu,X\rangle,\nabla\nu,{\mathrm{A}})+\langle\nu,X\rangle\mathfrak{q}^{2m+2}(\nabla\nu,{\mathrm{A}})-(\Delta\langle\nu,X\rangle+\langle\nu,X\rangle|{\mathrm{A}}|^{2})
=2(1)m+1Δm+1ν,X+2(1)m+1Δm(ν,X|A|2)\displaystyle=2(-1)^{m+1}\Delta^{m+1}\langle\nu,X\rangle+2(-1)^{m+1}\Delta^{m}(\langle\nu,X\rangle|A|^{2})
+𝔮2m+3(ν,X,ν,A)+ν,X𝔮2m+2(ν,A)(Δν,X+ν,X|A|2).\displaystyle\indent+\mathfrak{q}^{2m+3}(\langle\nu,X\rangle,\nabla\nu,{\mathrm{A}})+\langle\nu,X\rangle\mathfrak{q}^{2m+2}(\nabla\nu,{\mathrm{A}})-(\Delta\langle\nu,X\rangle+\langle\nu,X\rangle|{\mathrm{A}}|^{2}).

Hence, the thesis follows by observing that a generic monomial in 𝔮2m+3(ν,X,ν,A)\mathfrak{q}^{2m+3}(\langle\nu,X\rangle,\nabla\nu,{\mathrm{A}}) is of the form

k=1Nikν,Xl=1Mjlνs=1PrsA,\circledast_{k=1}^{N}\nabla^{i_{k}}\langle\nu,X\rangle\circledast_{l=1}^{M}\nabla^{j_{l}}\nabla\nu\circledast_{s=1}^{P}\nabla^{r_{s}}{\mathrm{A}},

with

k=1N(ik+1)+l=1M(jl+1)+s=1P(rs+1)=2m+3,\sum_{k=1}^{N}(i_{k}+1)+\sum_{l=1}^{M}(j_{l}+1)+\sum_{s=1}^{P}(r_{s}+1)=2m+3,

and N,M,P1N,M,P\geq 1 and then ik2mi_{k}\leq 2m for any kk. ∎

It follows that, by polarization, we can define the bilinear form

(δ2m)φ(f1,f2)ddsddtm(φ+sf1ν+tf2ν)|t=0|s=0=M(2(1)m+1Δm+1f1+Ω(f1))f2𝑑μ,\begin{split}(\delta^{2}\mathcal{F}_{m})_{\varphi}(f_{1},f_{2})&\coloneqq\frac{d}{ds}\frac{d}{dt}\mathcal{F}_{m}(\varphi+sf_{1}\nu+tf_{2}\nu)\bigg{|}_{t=0}\bigg{|}_{s=0}\\ &\>=\int_{M}\left(2(-1)^{m+1}\Delta^{m+1}f_{1}+\Omega(f_{1})\right)f_{2}\,d\mu\,,\end{split} (2.6)

for any pair of smooth functions f1,f2:Mf_{1},f_{2}:M\to\mathbb{R} and Ω\Omega is as in Theorem 2.3.

3. Analysis of the second variation

Suppose that φ:Mn+1\varphi:M\to\mathbb{R}^{n+1} is a smooth critical point of m\mathcal{F}_{m}, i.e., Em(φ)=0{\mathrm{E}}_{m}(\varphi)=0. The formula for the second variation given above shows that (δ2m)φ(f1,f2)(\delta^{2}\mathcal{F}_{m})_{\varphi}(f_{1},f_{2}) is well-defined for f1W2m+2,2(M,g)f_{1}\in W^{2m+2,2}(M,g) and f2L2(μ)f_{2}\in L^{2}(\mu). This means that

(δ2m)φ(f,)L2(μ),(\delta^{2}\mathcal{F}_{m})_{\varphi}(f,\cdot)\in L^{2}(\mu)^{\star},

for any fW2m+2,2(M,g)f\in W^{2m+2,2}(M,g) and it is well-defined the map

W2m+2,2(M,g)f(δ2m)φ(f,)L2(μ).W^{2m+2,2}(M,g)\ni f\mapsto(\delta^{2}\mathcal{F}_{m})_{\varphi}(f,\cdot)\in L^{2}(\mu)^{\star}.

We are going to exploit the theory of Fredholm operators between Banach spaces. For definitions and results on the subject we refer the reader to [14, Section 19.1]. We recall that if T:V1V2T:V_{1}\to V_{2} is a Fredholm operator between Banach spaces, its index is defined to be the integer number

indexTdimkerTdimcokerT.{\rm index}\,T\coloneqq\dim\ker T-\dim\,{\rm coker}\,T.

where dim\dim denotes the dimension of a finite dimensional vector space.

Proposition 3.1.

Let φ:Mn+1\varphi:M\to\mathbb{R}^{n+1} be a smooth critical point of m\mathcal{F}_{m}, i.e., Em(φ)=0{\mathrm{E}}_{m}(\varphi)=0. Then the second variation functional

(δ2m)φ:W2m+2,2(M,g)L2(μ)(\delta^{2}\mathcal{F}_{m})_{\varphi}:W^{2m+2,2}(M,g)\to L^{2}(\mu)^{\star}

is a Fredholm operator of index zero.

In order to prove Proposition 3.1 we need the following commutation rule.

Lemma 3.2.

Let φ:Mnn+1\varphi:M^{n}\to\mathbb{R}^{n+1} be a smooth immersion and let TT be a tensor defined on MM. Assume MM is endowed with the pull-back metric gg induced by φ\varphi. Then

ΔlTΔlT=𝔭2l1(A,A,T),\nabla\Delta^{l}T-\Delta^{l}\nabla T=\mathfrak{p}_{2l-1}({\mathrm{A}},{\mathrm{A}},T),

for any ll\in\mathbb{N} with l1l\geq 1.

Proof.

As we need to prove a pointwise identity, we can take a local coordinate frame E1,,EnE_{1},...,E_{n} which is orthonormal at a given point pp (that is, Ei,Ej=δij\langle E_{i},E_{j}\rangle=\delta_{ij}) and iEj=0\nabla_{i}E_{j}=0 at pp. In this way we can compute

(ΔT)(Ek)=(2(T)(Ei,Ei))(Ek)=(i(iT)iEiT)(Ek)=(i(iT))(Ek)=i((iT)(Ek))(iT)(iEk)=i(2T(Ei,Ek))2T(Ei,iEk)=i(2T(Ei,Ek)).\begin{split}(\Delta\nabla T)(E_{k})&=(\nabla^{2}(\nabla T)(E_{i},E_{i}))(E_{k})\\ &=(\nabla_{i}(\nabla_{i}\nabla T)-\nabla_{\nabla_{i}E_{i}}\nabla T)(E_{k})\\ &=(\nabla_{i}(\nabla_{i}\nabla T))(E_{k})\\ &=\nabla_{i}((\nabla_{i}\nabla T)(E_{k}))-(\nabla_{i}\nabla T)(\nabla_{i}E_{k})\\ &=\nabla_{i}(\nabla^{2}T(E_{i},E_{k}))-\nabla^{2}T(E_{i},\nabla_{i}E_{k})\\ &=\nabla_{i}(\nabla^{2}T(E_{i},E_{k})).\end{split}

at the point pp. On the other hand, using that for any tensor SS we have the commutation rule

(2S)(Ej,El)=(2S)(El,Ej)+RS(\nabla^{2}S)(E_{j},E_{l})=(\nabla^{2}S)(E_{l},E_{j})+{\mathrm{R}}*S

for any jj and ll, we obtain

(ΔT)(Ek)=k(trace2T)=tracek2T=(k(2T))(Ei,Ei)=(3T)(Ek,Ei,Ei)=(3T)(Ei,Ek,Ei)+RT=(i(2T))(Ek,Ei)+RT=i(2T(Ek,Ei))(2T)(iEk,Ei)(2T)(Ek,iEi)+RT=i(2T(Ei,Ek)+RT)+RT=(ΔT)(Ek)+(RT)+RT=(ΔT)(Ek)+𝔭1(A,A,T),\begin{split}(\nabla\Delta T)(E_{k})&=\nabla_{k}({\rm trace\,}\nabla^{2}T)\\ &={\rm trace\,}\nabla_{k}\nabla^{2}T\\ &=(\nabla_{k}(\nabla^{2}T))(E_{i},E_{i})\\ &=(\nabla^{3}T)(E_{k},E_{i},E_{i})\\ &=(\nabla^{3}T)(E_{i},E_{k},E_{i})+{\mathrm{R}}*\nabla T\\ &=(\nabla_{i}(\nabla^{2}T))(E_{k},E_{i})+{\mathrm{R}}*\nabla T\\ &=\nabla_{i}(\nabla^{2}T(E_{k},E_{i}))-(\nabla^{2}T)(\nabla_{i}E_{k},E_{i})-(\nabla^{2}T)(E_{k},\nabla_{i}E_{i})+{\mathrm{R}}*\nabla T\\ &=\nabla_{i}(\nabla^{2}T(E_{i},E_{k})+{\mathrm{R}}*T)+{\mathrm{R}}*\nabla T\\ &=(\Delta\nabla T)(E_{k})+\nabla({\mathrm{R}}*T)+{\mathrm{R}}*\nabla T\\ &=(\Delta\nabla T)(E_{k})+\mathfrak{p}_{1}({\mathrm{A}},{\mathrm{A}},T),\end{split}

where we used that R=AA{\mathrm{R}}={\mathrm{A}}*{\mathrm{A}}, by Gauss equations. Hence, the thesis is proved for l=1l=1. Letting now l+11l+1\geq 1, by induction we obtain

ΔΔlT=ΔΔlT+𝔭1(A,A,ΔlT)=Δ(ΔlT+𝔭2l1(A,A,T))+𝔭2l+1(A,A,T),\begin{split}\nabla\Delta\Delta^{l}T&=\Delta\nabla\Delta^{l}T+\mathfrak{p}_{1}({\mathrm{A}},{\mathrm{A}},\Delta^{l}T)=\Delta(\Delta^{l}\nabla T+\mathfrak{p}_{2l-1}({\mathrm{A}},{\mathrm{A}},T))+\mathfrak{p}_{2l+1}({\mathrm{A}},{\mathrm{A}},T),\end{split}

and the thesis follows. ∎

We are now ready to prove Proposition 3.1. A relevant property about Fredholm operators that we are going to use is the following. If T:V1V2T:V_{1}\to V_{2} is a Fredholm operator between Banach spaces and K:V1V2K:V_{1}\to V_{2} is a compact operator, then T+KT+K is Fredholm and index(T+K)=indexT{\rm index}(T+K)={\rm index}\,T (see [14, Corollary 19.1.8]).

Proof of Proposition 3.1.

For f1W2m+2,2(M,g)f_{1}\in W^{2m+2,2}(M,g) the functional (δ2m)φ(f1,)(\delta^{2}\mathcal{F}_{m})_{\varphi}(f_{1},\cdot) is given by

(δ2m)φ(f1,f2)=(f1),f2L2(μ),(\delta^{2}\mathcal{F}_{m})_{\varphi}(f_{1},f_{2})=\langle\mathcal{L}(f_{1}),f_{2}\rangle_{L^{2}(\mu)},

where :W2m+2,2(M,g)L2(μ)\mathcal{L}:W^{2m+2,2}(M,g)\to L^{2}(\mu) is

(f)=2(1)m+1Δm+1f+Ω(f),\mathcal{L}(f)=2(-1)^{m+1}\Delta^{m+1}f+\Omega(f),

and Ω\Omega is as in Theorem 2.3, hence Ω\Omega is a compact operator. Therefore

(δ2m)φ:W2m+2,2(M,g)L2(μ)(\delta^{2}\mathcal{F}_{m})_{\varphi}:W^{2m+2,2}(M,g)\to L^{2}(\mu)^{\star}

is Fredholm of index zero if and only if the same holds for :W2m+2,2(M,g)L2(μ)\mathcal{L}:W^{2m+2,2}(M,g)\to L^{2}(\mu).
We then claim that the operator

Cid+2(1)m+1Δm+1:W2m+2,2(M,g)L2(μ)C{\mathrm{id}}+2(-1)^{m+1}\Delta^{m+1}:W^{2m+2,2}(M,g)\to L^{2}(\mu)

is invertible for C>0C>0 sufficiently large, thus it is Fredholm of index zero. As the inclusion id:W2m+2,2(M,g)L2(μ){\mathrm{id}}:W^{2m+2,2}(M,g)\to L^{2}(\mu) is compact, this eventually implies that 2(1)m+1Δm+1:W2m+2,2(M,g)L2(μ)2(-1)^{m+1}\Delta^{m+1}:W^{2m+2,2}(M,g)\to L^{2}(\mu) is Fredholm of index zero.
The injectivity of the above operator immediately follows, suppose indeed that we have Cf+2(1)m+1Δm+1f=0Cf+2(-1)^{m+1}\Delta^{m+1}f=0, if m=2k+1m=2k+1, multiplying by ff and integrating, we get

CMf2𝑑μ=2MfΔ2(k+1)f𝑑μ=2M(Δk+1f)2𝑑μ,C\int_{M}f^{2}\,d\mu=-2\int_{M}f\Delta^{2(k+1)}f\,d\mu=-2\int_{M}(\Delta^{k+1}f)^{2}\,d\mu,

then f=0f=0. If instead m=2km=2k, multiplying by ff and integrating we get

CMf2𝑑μ=2MfΔ2k+1f𝑑μ=2M|Δkf|2𝑑μ,C\int_{M}f^{2}\,d\mu=2\int_{M}f\Delta^{2k+1}f\,d\mu=-2\int_{M}|\nabla\Delta^{k}f|^{2}\,d\mu,

then f=0f=0 as well.
About the surjectivity, given hL2(μ)h\in L^{2}(\mu) we aim at finding fW2m+2,2(M,g)f\in W^{2m+2,2}(M,g) such that Cf+2(1)m+1Δm+1f=hCf+2(-1)^{m+1}\Delta^{m+1}f=h. We shall minimize the functional

Am:Wm+1,2(M,g)A_{m}:W^{m+1,2}(M,g)\to\mathbb{R}

defined by

Am(f){M[C2f2+(Δk+1f)2fh]𝑑μ if m=2k+1,M[C2f2+|Δkf|2fh]𝑑μ if m=2k.A_{m}(f)\coloneqq\begin{cases}\int_{M}\bigl{[}\frac{C}{2}f^{2}+(\Delta^{k+1}f)^{2}-fh\bigr{]}\,d\mu&\mbox{ if }m=2k+1,\\ \int_{M}\bigl{[}\frac{C}{2}f^{2}+|\nabla\Delta^{k}f|^{2}-fh\bigr{]}\,d\mu&\mbox{ if }m=2k.\end{cases}

We can prove that AmA_{m} is coercive on Wm+1,2(M,g)W^{m+1,2}(M,g), up to choosing C>0C>0 sufficiently large (depending on mm and the geometry of (M,g)(M,g)).
We first consider the case m=2k+1m=2k+1. Integrating by parts in the integral M(Δk+1f)2𝑑μ\int_{M}(\Delta^{k+1}f)^{2}\,d\mu, that is, using the divergence theorem and applying the commutation rule of Lemma 3.2 we get

M(Δk+1f)2𝑑μ=MΔkf,Δk+1fdμ=M[Δkf,Δk+1f+Δkf𝔭2(k+1)1(A,A,f)+Δk+1f𝔭2k1(A,A,f)]dμ=M[Δkf,Δk+1f+𝔭4k+2(A,A,f,f)]dμ=M[(1)k+1k+1f,Δk+1k+1f+𝔭4k+2(A,A,f,f)]dμ=M[|2k+2f|2+𝔭4k+2(A,A,f,f)]dμ=M[|m+1f|2+𝔭2m(A,A,f,f)]dμ.\begin{split}\int_{M}(\Delta^{k+1}f)^{2}\,d\mu=\int_{M}&-\langle\nabla\Delta^{k}f,\nabla\Delta^{k+1}f\rangle\,d\mu\\ =\int_{M}&\bigl{[}-\langle\Delta^{k}\nabla f,\Delta^{k+1}\nabla f\rangle+\nabla\Delta^{k}f*\mathfrak{p}_{2(k+1)-1}({\mathrm{A}},{\mathrm{A}},f)\\ &\,\,+\nabla\Delta^{k+1}f*\mathfrak{p}_{2k-1}({\mathrm{A}},{\mathrm{A}},f)\bigr{]}\,d\mu\\ =\int_{M}&\bigl{[}-\langle\Delta^{k}\nabla f,\Delta^{k+1}\nabla f\rangle+\mathfrak{p}_{4k+2}({\mathrm{A}},{\mathrm{A}},f,f)\bigr{]}\,d\mu\\ =\int_{M}&\bigl{[}(-1)^{k+1}\langle\nabla^{k+1}f,\Delta^{k+1}\nabla^{k+1}f\rangle+\mathfrak{p}_{4k+2}({\mathrm{A}},{\mathrm{A}},f,f)\bigr{]}\,d\mu\\ =\int_{M}&\bigl{[}|\nabla^{2k+2}f|^{2}+\mathfrak{p}_{4k+2}({\mathrm{A}},{\mathrm{A}},f,f)\bigr{]}\,d\mu\\ =\int_{M}&\bigl{[}|\nabla^{m+1}f|^{2}+\mathfrak{p}_{2m}({\mathrm{A}},{\mathrm{A}},f,f)\bigr{]}\,d\mu.\end{split}

Moreover, by definition of 𝔭s\mathfrak{p}_{s}, we can apply the divergence theorem on the integral M𝔭2m(A,A,f,f)𝑑μ\int_{M}\mathfrak{p}_{2m}({\mathrm{A}},{\mathrm{A}},f,f)\,d\mu in the above expression so that in the polynomial there appear derivatives of ff of order mm at most.
We recall that for any covariant tensor TT there holds the general inequality (see [1, Chapter 3, Section 7.6])

lTL2(μ)Cl,mm+1TL2(μ)lm+1TL2(μ)m+1lm+1εm+1TL2(μ)+Cl,m(ε)TL2(μ),\|\nabla^{l}T\|_{L^{2}(\mu)}\leq C_{l,m}\|\nabla^{m+1}T\|_{L^{2}(\mu)}^{\frac{l}{m+1}}\|T\|_{L^{2}(\mu)}^{\frac{m+1-l}{m+1}}\leq\varepsilon\|\nabla^{m+1}T\|_{L^{2}(\mu)}+C_{l,m}(\varepsilon)\|T\|_{L^{2}(\mu)}, (3.1)

for any lml\leq m and ε>0\varepsilon>0. Therefore we can estimate

M|𝔭2m(A,A,f,f)|𝑑μCm(A2)M|l1f||l2f|𝑑μ,\int_{M}|\mathfrak{p}_{2m}({\mathrm{A}},{\mathrm{A}},f,f)|\,d\mu\leq C_{m}(\|{\mathrm{A}}\|_{\infty}^{2})\sum\int_{M}|\nabla^{l_{1}}f||\nabla^{l_{2}}f|\,d\mu,

where l1,l2ml_{1},l_{2}\leq m and then

M|𝔭2m(A,A,f,f)|𝑑μεCm(A2)m+1fL2(μ)2+Cm(A2,ε)fL2(μ)2.\int_{M}|\mathfrak{p}_{2m}({\mathrm{A}},{\mathrm{A}},f,f)|\,d\mu\leq\varepsilon C_{m}(\|{\mathrm{A}}\|_{\infty}^{2})\|\nabla^{m+1}f\|^{2}_{L^{2}(\mu)}+C_{m}(\|{\mathrm{A}}\|_{\infty}^{2},\varepsilon)\|f\|^{2}_{L^{2}(\mu)}.

Therefore, taking εCm(A2)<1/2\varepsilon C_{m}(\|{\mathrm{A}}\|_{\infty}^{2})<1/2 and C=C(m,A2)C=C(m,\|{\mathrm{A}}\|_{\infty}^{2}) sufficiently large, we estimate

Am(f)C¯M[f2+|m+1f|2h2]𝑑μ,A_{m}(f)\geq\overline{C}\int_{M}\bigl{[}f^{2}+|\nabla^{m+1}f|^{2}-h^{2}\bigr{]}\,d\mu,

that by inequality (3.1) implies that AmA_{m} is coercive on Wm+1,2(M,g)W^{m+1,2}(M,g). Analogously, one can prove the coercivity of AmA_{m} also in the case m=2km=2k.
It follows that there exists a function FWm+1,2(M,g)F\in W^{m+1,2}(M,g) solving

M[CFf+2Δk+1FΔk+1f]𝑑μ=Mfh𝑑μfWm+1,2(M,g)\int_{M}\bigl{[}CFf+2\Delta^{k+1}F\Delta^{k+1}f\bigr{]}\,d\mu=\int_{M}fh\,d\mu\qquad\forall\,f\in W^{m+1,2}(M,g)

if m=2k+1m=2k+1, or

M[CFf+2ΔkF,Δkf]𝑑μ=Mfh𝑑μfWm+1,2(M,g)\int_{M}\bigl{[}CFf+2\langle\nabla\Delta^{k}F,\nabla\Delta^{k}f\rangle\bigr{]}\,d\mu=\int_{M}fh\,d\mu\qquad\forall\,f\in W^{m+1,2}(M,g)

if m=2km=2k. In any case, FF is a weak solution to an elliptic equation with constant coefficients and datum hL2(μ)h\in L^{2}(\mu) (in the sense of [1, Point (d), Page 85]). Therefore, the standard regularity theory for distributional solutions applies (see [1, Theorem, Page 85]), hence FF belongs to W2m+2,2(M,g)W^{2m+2,2}(M,g). Integrating by parts, we then get that FF solves CF+2(1)m+1Δm+1F=hCF+2(-1)^{m+1}\Delta^{m+1}F=h, as required. ∎

4. Convergence

Suppose that φ:Mn+1\varphi:M\to\mathbb{R}^{n+1} is a smooth critical point of m\mathcal{F}_{m}, that is, Em(φ)=0{\mathrm{E}}_{m}(\varphi)=0. Then for ρ0>0\rho_{0}>0 suitably small, it is well-defined the functional m:Bρ0(0)W2m+2,2(M,g)\mathcal{E}_{m}:B_{\rho_{0}}(0)\subseteq W^{2m+2,2}(M,g)\to\mathbb{R} given by

m(f)m(φ+fν).\mathcal{E}_{m}(f)\coloneqq\mathcal{F}_{m}(\varphi+f\nu).

The advantage of the above definition is that the functional m\mathcal{E}_{m} is now defined on an open set of a Banach space and we can then look at first and second variation functionals in the classical sense of functional analysis. More precisely, by Theorem 2.1 we have

(δm)f1(f2)ddtm(f1+tf2)|t=0=MEm(φ+f1ν)ν1,νf2𝑑μ1,(\delta\mathcal{E}_{m})_{f_{1}}(f_{2})\coloneqq\frac{d}{dt}\mathcal{E}_{m}(f_{1}+tf_{2})\Big{|}_{t=0}=\int_{M}{\mathrm{E}}_{m}(\varphi+f_{1}\nu)\langle\nu_{1},\nu\rangle\,f_{2}\,d\mu_{1},

where ν\nu (resp. ν1\nu_{1}) is a unit normal vector along φ\varphi (resp. φ+f1ν\varphi+f_{1}\nu) and μ1\mu_{1} is the volume measure induced by φ+f1ν\varphi+f_{1}\nu. In this way we see that

δm:Bρ0(0)W2m+2,2(M,g)L2(μ).\delta\mathcal{E}_{m}:B_{\rho_{0}}(0)\subseteq W^{2m+2,2}(M,g)\to L^{2}(\mu)^{\star}.

Analogously, by Theorem 2.3 and formula (2.6) the second variation of m\mathcal{E}_{m} evaluated at 0Bρ0(0)0\in B_{\rho_{0}}(0) is given by

(δ2m)0(f1,f2)=M(2(1)m+1Δm+1f1+Ω(f1))f2𝑑μ,(\delta^{2}\mathcal{E}_{m})_{0}(f_{1},f_{2})=\int_{M}\left(2(-1)^{m+1}\Delta^{m+1}f_{1}+\Omega(f_{1})\right)f_{2}\,d\mu,

for Ω\Omega as in Theorem 2.3, so that

(δ2m)0:W2m+2,2(M,g)L2(μ),(\delta^{2}\mathcal{E}_{m})_{0}:W^{2m+2,2}(M,g)\to L^{2}(\mu)^{\star},

and it is a Fredholm operator of index zero by Proposition 3.1.

In this setting we can apply the following abstract result stating sufficient conditions implying a Łojasiewicz–Simon gradient inequality.

Proposition 4.1 ([20, Corollary 2.6]).

Let E:Bρ0(0)VE:B_{\rho_{0}}(0)\subseteq V\to\mathbb{R} be an analytic map, where VV is a Banach space. Suppose that 0 is a critical point for EE, i.e., δE0=0\delta E_{0}=0. Assume that there exists a Banach space ZZ such that VZV\hookrightarrow Z, the first variation δE:Bρ0(0)Z\delta E:B_{\rho_{0}}(0)\to Z^{\star} is ZZ^{\star}–valued and analytic and the second variation δ2E0:VZ\delta^{2}E_{0}:V\to Z^{\star} evaluated at 0 is ZZ^{\star}–valued and Fredholm of index zero.
Then there exist constants C,θ>0C,\theta>0 and α(0,1/2]\alpha\in(0,1/2] such that

|E(f)E(0)|1αCδEfZ,|E(f)-E(0)|^{1-\alpha}\leq C\|\delta E_{f}\|_{Z^{\star}},

for every fBθ(0)Vf\in B_{\theta}(0)\subseteq V.

The above functional analytic result is a corollary of the useful theory developed in [4] and it has been also proved in [21] independently.

Applying Proposition 4.1 to the functional m\mathcal{E}_{m} we obtain the following corollary.

Corollary 4.2.

Let φ:Mn+1\varphi:M\to\mathbb{R}^{n+1} be a smooth critical point of m\mathcal{F}_{m}, i.e., Em(φ)=0{\mathrm{E}}_{m}(\varphi)=0. Let ρ0>0\rho_{0}>0 such that m:Bρ0(0)W2m+2,2(M,g)\mathcal{E}_{m}:B_{\rho_{0}}(0)\subseteq W^{2m+2,2}(M,g)\to\mathbb{R} is well-defined.
Then, there exist constants C>0,θ(0,ρ0]C>0,\theta\in(0,\rho_{0}] and α(0,1/2]\alpha\in(0,1/2] such that

|m(φ+fν)m(φ)|1αC(δm)fL2(μ),|\mathcal{F}_{m}(\varphi+f\nu)-\mathcal{F}_{m}(\varphi)|^{1-\alpha}\leq C\|(\delta\mathcal{E}_{m})_{f}\|_{L^{2}(\mu)^{\star}},

for every fBθ(0)W2m+2,2(M,g)f\in B_{\theta}(0)\subseteq W^{2m+2,2}(M,g).

Proof.

We want to apply Proposition 4.1 with V=W2m+2,2(M,g)V=W^{2m+2,2}(M,g) and Z=L2(μ)Z=L^{2}(\mu). By Proposition 3.1 and the discussion at the beginning of the section, we just need to check that m\mathcal{E}_{m} and δm\delta\mathcal{E}_{m} are analytic as maps between Banach spaces.

We can rewrite

m(f)=M1+α=1n+1mνfα,mνfαdμf\mathcal{E}_{m}(f)=\int_{M}1+\sum_{\alpha=1}^{n+1}\langle\nabla^{m}\nu^{\alpha}_{f},\nabla^{m}\nu^{\alpha}_{f}\rangle\,d\mu_{f}

where νf\nu_{f} is a unit normal along φ+fν\varphi+f\nu and μf\mu_{f} is the volume measure induced by φ+fν\varphi+f\nu. If ψ:Mn+1\psi:M\to\mathbb{R}^{n+1} is any immersion, we have that a unit normal along ψ\psi is νψ=1ψnψ|1ψnψ|\nu_{\psi}=\star\frac{\partial_{1}\psi\wedge\ldots\wedge\partial_{n}\psi}{|\partial_{1}\psi\wedge\ldots\wedge\partial_{n}\psi|}, where \star denotes the Euclidean Hodge star operator. As ψ\psi is an immersion, we see that ψνψ\psi\mapsto\nu_{\psi} is analytic. It follows that fνff\mapsto\nu_{f} is analytic as well. As the metric tensor induced by an immersion ψ:Mn+1\psi:M\to\mathbb{R}^{n+1} has components gij=iψ,jψg_{ij}=\langle\partial_{i}\psi,\partial_{j}\psi\rangle, we get that the metric tensor of φ+fν\varphi+f\nu depends analytically on ff and then it is analytic the dependence of μf\mu_{f} and of Christoffel symbols (and thus of the connection) on ff. Then the integrand in the definition of m\mathcal{E}_{m} is just a sum of compositions and multiplications of functions which are analytic in ff. Finally, integration is linear on L1(μ)L^{1}(\mu), then fm(f)f\mapsto\mathcal{E}_{m}(f)\in\mathbb{R} is analytic for fBρ0(0)W2m+2,2(M,g)f\in B_{\rho_{0}}(0)\subseteq W^{2m+2,2}(M,g).

By the very same arguments, one can check that also f(δm)ff\mapsto(\delta\mathcal{E}_{m})_{f} is analytic. Hence, all the hypotheses of Proposition 4.1 are satisfied and the thesis follows. ∎

The starting point for proving the smooth convergence of the gradient flow of m\mathcal{F}_{m} is the following sub–convergence theorem.

Theorem 4.3 ([17, Theorem 7.8, Theorem 8.2]).

Let φ0:Mnn+1\varphi_{0}:M^{n}\to\mathbb{R}^{n+1} be a smooth immersion and let m>n/2m>\lfloor n/2\rfloor. Then there exists a unique smooth solution φ:M×[0,+)n+1\varphi:M\times[0,+\infty)\to\mathbb{R}^{n+1} to the evolution equation

{tφ=Em(φt)νt,φ(,0)=φ0,\begin{cases}\partial_{t}\varphi=-{\mathrm{E}}_{m}(\varphi_{t})\nu_{t},\\ \varphi(\cdot,0)=\varphi_{0},\end{cases}

where νt\nu_{t} denotes a unit normal vector field along φtφ(,t)\varphi_{t}\coloneqq\varphi(\cdot,t). Moreover, the solution satisfies the estimates

kAtL(M,gt)C(k,n,φ0),\|\nabla^{k}{\mathrm{A}}_{t}\|_{L^{\infty}(M,g_{t})}\leq C(k,n,\varphi_{0}), (4.1)

for any t[0,+)t\in[0,+\infty), where At{\mathrm{A}}_{t} and gtg_{t} are the second fundamental form and the metric of φt\varphi_{t} respectively and there exists a smooth critical point φ:Mn+1\varphi_{\infty}:M\to\mathbb{R}^{n+1} of m\mathcal{F}_{m}, a sequence of times tj+t_{j}\to+\infty and a sequence of points pjn+1p_{j}\in\mathbb{R}^{n+1} such that

φtjσjpjφCk(M)j+0,\|\varphi_{t_{j}}\circ\sigma_{j}-p_{j}-\varphi_{\infty}\|_{C^{k}(M)}\xrightarrow[j\to+\infty]{}0,

for any kk\in\mathbb{N}, where σj\sigma_{j} is a sequence of diffeomorphisms of MM.

We need a preliminary lemma.

Lemma 4.4.

Let φ0,φt,φ,σj,tj,pj\varphi_{0},\varphi_{t},\varphi_{\infty},\sigma_{j},t_{j},p_{j} be as in Theorem 4.3. Then, for any ε>0\varepsilon>0 there is jεj_{\varepsilon}\in\mathbb{N} such that for any jjεj\geq j_{\varepsilon} there exists δj>0\delta_{j}>0 such that the immersion φtpj\varphi_{t}-p_{j} coincides with φ+ftν\varphi_{\infty}+f_{t}\nu_{\infty} up to diffeomorphism, where ν\nu_{\infty} is a unit normal vector along φ\varphi_{\infty}, for some “height” functions ftC(M)f_{t}\in C^{\infty}(M) smoothly depending on t[tj,tj+δj)t\in[t_{j},t_{j}+\delta_{j}). Moreover,

ftW2m+2,2(M,g)ε,\|f_{t}\|_{W^{2m+2,2}(M,g_{\infty})}\leq\varepsilon,

for any t[tj,tj+δj)t\in[t_{j},t_{j}+\delta_{j}).

Proof.

Fixed θ>0\theta>0 and k>2m+2k>2m+2, by Theorem 4.3 there is jθj_{\theta} such that for any jjθj\geq j_{\theta} we have

φtσjpjφCk(M)<θ,\|\varphi_{t}\circ\sigma_{j}-p_{j}-\varphi_{\infty}\|_{C^{k}(M)}<\theta, (4.2)

for every t[tj,tj+δj)t\in[t_{j},t_{j}+\delta_{j}), for some δj>0\delta_{j}>0.
Let us assume that φ\varphi_{\infty} is an embedding. The general statement analogously follows by recalling that immersions are local embeddings. So for jθj_{\theta} large enough, φt\varphi_{t} is an embedding as well for every t[tj,tj+δj)t\in[t_{j},t_{j}+\delta_{j}). Moreover there exists Un+1U\subseteq\mathbb{R}^{n+1} open set containing Nφ(M)N\coloneqq\varphi_{\infty}(M) such that it is well-defined the projection map π:UN\pi:U\to N as

π(p)=p12n+1dN2(p),\pi(p)=p-\frac{1}{2}\nabla^{\mathbb{R}^{n+1}}d_{N}^{2}(p),

where dNd_{N} is the distance function from NN. The vector 12n+1dN2(p)\tfrac{1}{2}\nabla^{\mathbb{R}^{n+1}}d_{N}^{2}(p) is orthogonal to NN at π(p)\pi(p), π\pi is smooth on UU and for jθj_{\theta} sufficiently large we have that (φtσj(M)pj)U(\varphi_{t}\circ\sigma_{j}(M)-p_{j})\subseteq U for every t[tj,tj+δj)t\in[t_{j},t_{j}+\delta_{j}) (for a proof of these facts see [18, Proposition 4.2]).
Hence, for xMx\in M the “height” function ft(x)f_{t}(x) is uniquely determined by the identity

φtσj(x)pj=π(φtσj(x)pj)+ft(x)ν(φ1π(φtσj(x)pj)),\varphi_{t}\circ\sigma_{j}(x)-p_{j}=\pi(\varphi_{t}\circ\sigma_{j}(x)-p_{j})+f_{t}(x)\nu_{\infty}(\varphi_{\infty}^{-1}\circ\pi\circ(\varphi_{t}\circ\sigma_{j}(x)-p_{j})),

that is,

ft(x)=φtσj(x)pjπ(φtσj(x)pj),ν(φ1π(φtσj(x)pj)).f_{t}(x)=\langle\varphi_{t}\circ\sigma_{j}(x)-p_{j}-\pi(\varphi_{t}\circ\sigma_{j}(x)-p_{j}),\nu_{\infty}(\varphi_{\infty}^{-1}\circ\pi\circ(\varphi_{t}\circ\sigma_{j}(x)-p_{j}))\rangle. (4.3)

Then, the map (x,t)ft(x)(x,t)\mapsto f_{t}(x) is smooth on M×[tj,tj+δj)M\times[t_{j},t_{j}+\delta_{j}) and ftW2m+2,2(M,g)0\|f_{t}\|_{W^{2m+2,2}(M,g_{\infty})}\to 0 as θ0\theta\to 0, by inequality (4.2) and the fact that k>2m+2k>2m+2.
Hence, for the chosen ε>0\varepsilon>0, taking a suitable θ>0\theta>0 we have the estimate in the statement of the lemma. ∎

We are now ready for proving our main result. The proof of Theorem 1.1 is essentially a generalization of the strategy employed in [19] to show the smooth convergence of the elastic flow of closed curves in n\mathbb{R}^{n}.

Proof of Theorem 1.1..

Let φ0,φt,φ,σj,tj,pj\varphi_{0},\varphi_{t},\varphi_{\infty},\sigma_{j},t_{j},p_{j} be as in Theorem 4.3. Fixed k>2m+2k>2m+2 and chosen ε>0\varepsilon>0 smaller than the constant θ\theta given by Corollary 4.2, relative to the critical point φ\varphi_{\infty}, by Theorem 4.3 and Lemma 4.4, there exists jεj_{\varepsilon}\in\mathbb{N} such that for every jjεj\geq j_{\varepsilon} we have

φtσjpjφCk(M)<ε,\|\varphi_{t}\circ\sigma_{j}-p_{j}-\varphi_{\infty}\|_{C^{k}(M)}<\varepsilon, (4.4)

for every t[tj,tj+δj)t\in[t_{j},t_{j}+\delta_{j}) with some δj>0\delta_{j}>0, moreover, φtσjpj\varphi_{t}\circ\sigma_{j}-p_{j} coincides with φ+ftν\varphi_{\infty}+f_{t}\nu_{\infty}, up to diffeomorphism, for the functions ftf_{t} given by Lemma 4.4 (we recall that ftC(M)f_{t}\in C^{\infty}(M) depends on jj), satisfying

ftW2m+2,2(M,g)<ε<θ,\|f_{t}\|_{W^{2m+2,2}(M,g_{\infty})}<\varepsilon<\theta, (4.5)

for every t[tj,tj+δj)t\in[t_{j},t_{j}+\delta_{j}).
We claim that it is possible to choose ε>0\varepsilon>0 small enough such that for any fixed jjεj\geq j_{\varepsilon}, the hypersurfaces φtσjpj\varphi_{t}\circ\sigma_{j}-p_{j} coincide with φ+ftν\varphi_{\infty}+f_{t}\nu_{\infty} (up to diffeomorphism) for some smooth functions ftf_{t} with ftW2m+2,2(M,g)<θ\|f_{t}\|_{W^{2m+2,2}(M,g_{\infty})}<\theta for any t[tj,+)t\in[t_{j},+\infty).
We define

H(t)|m(φt)m(φ)|α,H(t)\coloneqq|\mathcal{F}_{m}(\varphi_{t})-\mathcal{F}_{m}(\varphi_{\infty})|^{\alpha},

where α(0,1/2]\alpha\in(0,1/2] is as in Corollary 4.2 applied to the critical point φ\varphi_{\infty} and, without loss of generality, we can clearly assume that H(t)>0H(t)>0 for any tt. As m(φt)=m(φ+ftν)\mathcal{F}_{m}(\varphi_{t})=\mathcal{F}_{m}(\varphi_{\infty}+f_{t}\nu_{\infty}), by Corollary 4.2 we have

H(t)1ααC(δm)ftL2(μ)=C(M|Em(φ+ftν)νt,ν|2detgtdμ)1/2C(φ,θ)(M|Em(φ+ftν)|2𝑑μt)1/2=C(φ,θ)Em(φt)L2(μ),\begin{split}H(t)^{\frac{1-\alpha}{\alpha}}&\leq C\|(\delta\mathcal{E}_{m})_{f_{t}}\|_{L^{2}(\mu_{\infty})^{\star}}\\ &=C\Bigl{(}\int_{M}|{\mathrm{E}}_{m}(\varphi_{\infty}+f_{t}\nu_{\infty})\langle\nu_{t},\nu_{\infty}\rangle|^{2}\det g_{t}\,d\mu_{\infty}\Bigr{)}^{1/2}\\ &\leq C(\varphi_{\infty},\theta)\Bigl{(}\int_{M}|{\mathrm{E}}_{m}(\varphi_{\infty}+f_{t}\nu_{\infty})|^{2}\,d\mu_{t}\Bigr{)}^{1/2}\\ &=C(\varphi_{\infty},\theta)\|{\mathrm{E}}_{m}(\varphi_{t})\|_{L^{2}(\mu)},\end{split}

where νt\nu_{t}, gtg_{t}, μt=detgtdμ\mu_{t}=\det g_{t}\,d\mu_{\infty} are the unit normal, metric tensor and volume measure on φ+ftν\varphi_{\infty}+f_{t}\nu_{\infty} and we estimated detgtC(φ,θ)\sqrt{\det g_{t}}\leq C(\varphi_{\infty},\theta), for any ttjt\geq t_{j} such that

ftW2m+2,2(M,g)<θ.\|f_{t}\|_{W^{2m+2,2}(M,g_{\infty})}<\theta.

Differentiating HH and using the above inequality, we obtain

tH(t)=αHα1αtm(φt)=αHα1αMEm(φt),tφt𝑑μ=αHα1αM|Em(φt)||tφt|𝑑μ=αHα1αEm(φt)L2(μ)tφtL2(μ)αC(φ,θ)tφtL2(μ),\begin{split}\partial_{t}H(t)&=\alpha H^{\frac{\alpha-1}{\alpha}}\partial_{t}\mathcal{F}_{m}(\varphi_{t})\\ &=\alpha H^{\frac{\alpha-1}{\alpha}}\int_{M}\langle{\mathrm{E}}_{m}(\varphi_{t}),\partial_{t}^{\perp}\varphi_{t}\rangle\,d\mu\\ &=-\alpha H^{\frac{\alpha-1}{\alpha}}\int_{M}|{\mathrm{E}}_{m}(\varphi_{t})||\partial_{t}^{\perp}\varphi_{t}|\,d\mu\\ &=-\alpha H^{\frac{\alpha-1}{\alpha}}\|{\mathrm{E}}_{m}(\varphi_{t})\|_{L^{2}(\mu)}\|\partial_{t}^{\perp}\varphi_{t}\|_{L^{2}(\mu)}\\ &\leq-\alpha C(\varphi_{\infty},\theta)\|\partial_{t}^{\perp}\varphi_{t}\|_{L^{2}(\mu)},\end{split}

for any ttjt\geq t_{j} such that ftW2m+2,2(M,g)<θ\|f_{t}\|_{W^{2m+2,2}(M,g_{\infty})}<\theta. For such times, possibly choosing a smaller ε\varepsilon, we can assume that |νtν|<1/2|\nu_{t}-\nu_{\infty}|<1/2. Letting φ~tφ+ftν\widetilde{\varphi}_{t}\coloneqq\varphi_{\infty}+f_{t}\nu_{\infty} we thus get

|tφ~t|=|tφ~t,νtνt|=|tφ~t,ννt+tφ~t,νtννt||tφ~t,ν|12|tφ~t|=12|tφ~t|\begin{split}|\partial_{t}^{\perp}\widetilde{\varphi}_{t}|&=|\langle\partial_{t}\widetilde{\varphi}_{t},\nu_{t}\rangle\nu_{t}|\\ &=|\langle\partial_{t}\widetilde{\varphi}_{t},\nu_{\infty}\rangle\nu_{t}+\langle\partial_{t}\widetilde{\varphi}_{t},\nu_{t}-\nu_{\infty}\rangle\nu_{t}|\\ &\geq|\langle\partial_{t}\widetilde{\varphi}_{t},\nu_{\infty}\rangle|-\frac{1}{2}|\partial_{t}\widetilde{\varphi}_{t}|\\ &=\frac{1}{2}|\partial_{t}\widetilde{\varphi}_{t}|\end{split}

and the above estimate becomes

tH(t)αC(φ,θ)tφtL2(μ)=αC(φ,θ)tφ~tL2(μt)αC(φ,θ)tφ~tL2(μt)\partial_{t}H(t)\leq-\alpha C(\varphi_{\infty},\theta)\|\partial_{t}^{\perp}\varphi_{t}\|_{L^{2}(\mu)}=-\alpha C(\varphi_{\infty},\theta)\|\partial_{t}^{\perp}\widetilde{\varphi}_{t}\|_{L^{2}(\mu_{t})}\leq-\alpha C(\varphi_{\infty},\theta)\|\partial_{t}\widetilde{\varphi}_{t}\|_{L^{2}(\mu_{t})}

for any ttjt\geq t_{j} such that ftW2m+2,2(M,g)<θ\|f_{t}\|_{W^{2m+2,2}(M,g_{\infty})}<\theta. Integrating the above differential inequality and estimating detgtC(φ,θ)>0\sqrt{\det g_{t}}\geq C(\varphi_{\infty},\theta)>0, we obtain

φ~τ2φ~τ1L2(μ)=τ1τ2tφ~tdtL2(μ)C(φ,θ)τ1τ2tφ~tL2(μt)𝑑tC(α,φ,θ)(H(τ1)H(τ2))C(α,φ,θ)(m(φτ1)m(φ))α\begin{split}\|\widetilde{\varphi}_{\tau_{2}}-\widetilde{\varphi}_{\tau_{1}}\|_{L^{2}(\mu_{\infty})}&=\left\|\int_{\tau_{1}}^{\tau_{2}}\partial_{t}\widetilde{\varphi}_{t}\,dt\,\,\right\|_{L^{2}(\mu_{\infty})}\\ &\leq C(\varphi_{\infty},\theta)\int_{\tau_{1}}^{\tau_{2}}\left\|\partial_{t}\widetilde{\varphi}_{t}\right\|_{L^{2}(\mu_{t})}\,dt\\ &\leq C(\alpha,\varphi_{\infty},\theta)(H(\tau_{1})-H(\tau_{2}))\\ &\leq C(\alpha,\varphi_{\infty},\theta)(\mathcal{F}_{m}(\varphi_{\tau_{1}})-\mathcal{F}_{m}(\varphi_{\infty}))^{\alpha}\end{split}

then, since possibly choosing a larger jεj_{\varepsilon} we can assume that m(φtjε)m(φ)ε1/α\mathcal{F}_{m}(\varphi_{t_{j_{\varepsilon}}})-\mathcal{F}_{m}(\varphi_{\infty})\leq\varepsilon^{1/\alpha}, we see that

φ~τ2φ~τ1L2(μ)C(α,φ,θ)ε\|\widetilde{\varphi}_{\tau_{2}}-\widetilde{\varphi}_{\tau_{1}}\|_{L^{2}(\mu_{\infty})}\leq C(\alpha,\varphi_{\infty},\theta)\varepsilon (4.6)

for any τ2τ1tj\tau_{2}\geq\tau_{1}\geq t_{j} such that ftW2m+2,2(M,g)<θ\|f_{t}\|_{W^{2m+2,2}(M,g_{\infty})}<\theta on t[tj,τ2]t\in[t_{j},\tau_{2}]. Finally, since ftL2(μ)=φ~tφL2(μ)\|f_{t}\|_{L^{2}(\mu_{\infty})}=\|\widetilde{\varphi}_{t}-\varphi_{\infty}\|_{L^{2}(\mu_{\infty})}, we get

ftL2(μ)φ~tφ~tjL2(μ)+φ~tjφL2(μ)C(α,φ,θ)ε\|f_{t}\|_{L^{2}(\mu_{\infty})}\leq\|\widetilde{\varphi}_{t}-\widetilde{\varphi}_{t_{j}}\|_{L^{2}(\mu_{\infty})}+\|\widetilde{\varphi}_{t_{j}}-\varphi_{\infty}\|_{L^{2}(\mu_{\infty})}\leq C(\alpha,\varphi_{\infty},\theta)\varepsilon (4.7)

for any ttjt\geq t_{j} such that ftW2m+2,2(M,g)<θ\|f_{t}\|_{W^{2m+2,2}(M,g_{\infty})}<\theta.
Since m>n/2m>\lfloor n/2\rfloor, estimate (4.5) implies that the hypersurfaces φ~t\widetilde{\varphi}_{t} are represented as graph on φ\varphi_{\infty} by means of functions ftf_{t} with uniformly equibounded gradients (such bound clearly depends on ε\varepsilon and goes to zero with it). Also, the inequalities (4.1) clearly hold also for the second fundamental form of the hypersurfaces φtσj\varphi_{t}\circ\sigma_{j} and φ~t\widetilde{\varphi}_{t}, since they coincide with φt\varphi_{t} up to diffeomorphism (and translation). These facts imply uniform estimates on the “height” functions ftf_{t} in Wr,(M,g)W^{r,\infty}(M,g_{\infty}); namely, for any rr\in\mathbb{N} we have

ftWr,(M,g)C(r,n,φ0,φ),\|f_{t}\|_{W^{r,\infty}(M,g_{\infty})}\leq C(r,n,\varphi_{0},\varphi_{\infty}), (4.8)

for any t[tj,tj+δj)t\in[t_{j},t_{j}+\delta_{j}) (a tedious but straightforward way to see this is to differentiate formula (4.3) and use Gauss–Weingarten relations (1.1), taking into account that the closeness in W2m+2,2(M,g)W^{2m+2,2}(M,g_{\infty}) implies that the metric tensor and the Christoffel symbols of the covariant derivative of φ~t\widetilde{\varphi}_{t} are mutually “comparable” with the ones relative to φ\varphi_{\infty}). Hence, if r>2m+2r>2m+2 and ε>0\varepsilon>0 is small enough, combining estimates (4.7) and (4.8), the interpolation inequalities (3.1) imply that

ftW2m+2,2(M,g)<θ,\|f_{t}\|_{W^{2m+2,2}(M,g_{\infty})}<\theta,

for any t[tj,tj+δj)t\in[t_{j},t_{j}+\delta_{j}). By a maximality argument, it clearly follows that we can take δj=+\delta_{j}=+\infty, for every jjεj\geq j_{\varepsilon}. Hence, the estimate (4.6), which then holds for any ttjt\geq t_{j}, implies that the flow φ~t\widetilde{\varphi}_{t} satisfies the Cauchy criterion for convergence in L2(μ)L^{2}(\mu_{\infty}), hence φ~t\widetilde{\varphi}_{t} converges in L2(μ)L^{2}(\mu_{\infty}), as t+t\to+\infty. Interpolating as before by means of inequalities (4.8), the same holds for φ~t\widetilde{\varphi}_{t} in Wr,2(M,g)W^{r,2}(M,g_{\infty}), for any rr\in\mathbb{N} and, by Sobolev embeddings, we thus deduce that there exists the limit limt+φ~t\lim_{t\to+\infty}\widetilde{\varphi}_{t} in Cr(M)C^{r}(M) for any rr\in\mathbb{N}. Therefore, the same conclusion holds for the original flow φt\varphi_{t}, up to diffeomorphism. ∎

References

  • [1] T. Aubin, Some nonlinear problems in Riemannian geometry, Springer Monographs in Mathematics, Springer–Verlag, Berlin, 1998.
  • [2] G. Bellettini, C. Mantegazza, and M. Novaga, Singular perturbations of mean curvature flow, J. Differential Geom. 75 (2007), no. 3, 403–431.
  • [3] A. Carlotto, O. Chodosh, and Y. Rubinstein, Slowly converging Yamabe flows, Geom. Topol. 19 (2015), no. 3, 1523–1568.
  • [4] R. Chill, On the Łojasiewicz–Simon gradient inequality, J. Funct. Anal. 201 (2003), no. 2, 572–601.
  • [5] R. Chill, E. Fašangová, and R. Schätzle, Willmore blowups are never compact, Duke Math. J. 147 (2009), no. 2, 345–376.
  • [6] O. Chodosh and F. Schulze, Uniqueness of asymptotically conical tangent flows, ArXiv Preprint Server – arXiv:1901.06369. To appear: Duke Math. J., 2020.
  • [7] K. Choi and C. Mantoulidis, Ancient gradient flows of elliptic functionals and Morse index, ArXiv Preprint Server – arXiv:1902.07697. To appear: Amer. J. Math., 2020.
  • [8] T. H. Colding and W. P. Minicozzi II, Uniqueness of blowups and Łojasiewicz inequalities, Ann. of Math. 182 (2015), no. 1, 221–285.
  • [9] A. Dall’Acqua, P. Pozzi, and A. Spener, The Łojasiewicz–Simon gradient inequality for open elastic curves, J. Differential Equations 261 (2016), no. 3, 2168–2209.
  • [10] E. De Giorgi, Congetture riguardanti alcuni problemi di evoluzione. A celebration of John F. Nash, Jr., Duke Math. J. 81 (1996), no. 2, 255–268.
  • [11] E. De Giorgi, Congetture riguardanti alcuni problemi di evoluzione. A celebration of John F. Nash, Jr. (English translation), CvGmt Preprint Server – Scuola Normale Superiore di Pisa, http:/​​/cvgmt.sns.it, 1996.
  • [12] P. M. N. Feehan, Global existence and convergence of solutions to gradient systems and applications to Yang–Mills gradient flow, ArXiv Preprint Server – arXiv:1409.1525, 2016.
  • [13] P. M. N. Feehan and M. Maridakis, Łojasiewicz–Simon gradient inequalities for analytic and Morse–Bott functions on Banach spaces, J. Reine Angew. Math. 765 (2020), 35–67.
  • [14] L. Hörmander, The analysis of linear partial differential operators. III, Classics in Mathematics, Springer, Berlin, 2007, –differential operators, Reprint of the 1994 edition.
  • [15] S. Łojasiewicz, Une propriété topologique des sous–ensembles analytiques réels, Les Équations aux Dérivées Partielles (Paris, 1962), Éditions du Centre National de la Recherche Scientifique, Paris, 1963, pp. 87–89.
  • [16] by same author, Sur les trajectoires du gradient d’une fonction analytique, Seminari di Geometria (1982/83), Università degli Studi di Bologna (1984), 115–117.
  • [17] C. Mantegazza, Smooth geometric evolutions of hypersurfaces, Geom. Funct. Anal. 12 (2002), no. 1, 138–182.
  • [18] C. Mantegazza and A. C. Mennucci, Hamilton–Jacobi equations and distance functions on Riemannian manifolds, Appl. Math. Opt. 47 (2003), no. 1, 1–25.
  • [19] C. Mantegazza and M. Pozzetta, The Łojasiewicz–Simon inequality for the elastic flow, Calc. Var. 60 (2021), no. 56.
  • [20] M. Pozzetta, Convergence of elastic flows of curves into manifolds, ArXiv Preprint Server – arXiv:2007.00582. To appear: Nonlinear Anal., 2021.
  • [21] F. Rupp, On the Lojasiewicz–Simon gradient inequality on submanifolds, J. Funct. Anal. 279 (2020), no. 8, 1–32.
  • [22] F. Schulze, Uniqueness of compact tangent flows in Mean Curvature Flow, J. Reine Angew. Math. 690 (2014), 163–172.
  • [23] L. Simon, Asymptotics for a class of nonlinear evolution equations, with applications to geometric problems, Ann. of Math. (2) 118 (1983), no. 3, 525–571.