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

Asymptotics for slowly converging evolution equations

Beomjun Choi Department of Mathematics, POSTECH, Pohang, 37673, Republic of Korea bchoi@postech.ac.kr  and  Pei-Ken Hung School of Mathematics, University of Minnesota, Minneapolis, MN 55455, USA pkhung@umn.edu
Abstract.

We investigate slowly converging solutions for non-linear evolution equations of elliptic or parabolic type. These equations arise from the study of isolated singularities in geometric variational problems. Slowly converging solutions have previously been constructed assuming the Adams-Simon positivity condition. In this study, we identify a necessary condition for slowly converging solutions to exist, which we refer to as the Adams-Simon non-negativity condition. Additionally, we characterize the rate and direction of convergence for these solutions. Our result partially confirms Thom’s gradient conjecture in the context of infinite-dimensional problems.

1. Introduction

Analyzing the behavior of solutions near singularities is essential to understand geometric equations such as minimal surfaces, harmonic maps and mean curvature flows. Many open problems reduce to questions on the singularity formation and its asymptotics. In pioneering works of Leon Simon [Sim83, Sim85], the idea of using Łojasiewicz gradient inequality [Łoj63] from real algebraic geometry was introduced for the first time, and the uniqueness of blow-ups was shown for a class of elliptic and parabolic equations. This uniqueness shows that the solution converges to the unique tangent cone or tangent flow as it approaches a singular point or infinity.


A natural subsequent question is to investigate the rate of convergence and the next order asymptotics that describes the difference between the solution and the limit. This often serves as a crucial starting point for further analysis. For example, recent progresses on the classification of ancient solutions to geometric flows [SAS20][DH21][BCS22][CM22] and complete non-compact solutions to minimal surface [SS86] are based on higher order asymptotics and its improvement. For the singularity formation in parabolic problem, the higher order asymptotics at a singularity gives structural results on the singularity set in a neighborhood [SX22][Gan21]. As the uniqueness of blow-ups implies the second differentiability of arrival time [CM15, CM18], the higher order asymptotics and the convergence rate have a strong relation to further regularity of arrival time [KS06][Ses08].


The convergence rate and direction (i.e., the secant at the limit) are mostly understood when the solution converges at an exponential rate. This is because when the solution decays exponentially, the equation is well-approximated by its linearization. However, without the integrability of the limit, it is possible to have solutions that converge algebraically slowly. In [AS88], Adams-Simon discovered a sufficient condition (later called the Adams-Simon positivity condition or simply ASpAS_{p} condition) on the limit so that they constructed a slowly converging solution to elliptic equations of the form (1.1). Carlotto-Chodosh-Rubinstein [CCR15] found explicit examples of critical points of normalized Yamabe functional with the ASpAS_{p} condition and constructed normalized Yamabe flows converging slowly. As questioned in [CCR15, p.1533], it is of great interest to understand the general behavior of slowly converging solutions. More precisely, one can ask whether such a solution must satisfy the Adams-Simon positivity condition and whether the higher order asymptotics follows the ansatz used in the construction of [AS88] and [CCR15].


We answer this question by showing that the Adams-Simon ‘non-negativity’ condition is necessary for slowly converging solutions to exist. Moreover, when a positivity is satisfied, the convergence rate and the higher asymptotics agree with those of previously constructed examples. (See Theorem 2.3 and 2.5 for detailed statements.) We concern with the elliptic and parabolic equations of forms

(1.1) u′′mu+Σu=N1(u),u^{\prime\prime}-mu^{\prime}+\mathcal{M}_{\Sigma}u=N_{1}(u),

and

(1.2) uΣu=N2(u).u^{\prime}-\mathcal{M}_{\Sigma}u=N_{2}(u).

Here, Σ\mathcal{M}_{\Sigma} is the Euler-Lagrange operator of an analytic functional Σ\mathcal{F}_{\Sigma} as in (2.2)-(2.3), mm is a nonzero constant, and we assume N1(u)N_{1}(u) and N2(u)N_{2}(u) satisfy the structure (2.6). The presence of nonzero constant mm accounts that the equation (1.1) appears in the study of the isolated singularity or the asymptotics near the infinity for geometric variational problems. In particular, minimal surfaces, harmonic maps [Sim83, Sim85, AS88] and G2G_{2} manifolds [Che18] fall into this category. Equation (1.2) models a geometric flow which converges to a stationary solution when tt goes to ++\infty or .-\infty. For instance, the (rescaled) mean curvature flow (see Appendix B) and the harmonic map heat flow [CM22] can be described by (1.2). Our results also apply to the normalized Yamabe flow. See Remark 2.9 (3) for more details.


The main results of the paper, Theorems 2.3-2.6, can be placed in line as a generalization of Thom’s gradient conjecture which we discuss in this paragraph. Let x=x(t)nx=x(t)\in\mathbb{R}^{n} be a solution to the gradient flow x=f(x)x^{\prime}=-\nabla f(x) with an analytic potential f:nf:\mathbb{R}^{n}\to\mathbb{R} and suppose xx converges to the origin as tt\to\infty. In [Tho89], Thom conjectured that the secant x/|x|{x}/{|x|} should converge to a direction θ0𝕊n1\theta_{0}\in\mathbb{S}^{n-1} as tt\to\infty. Kurdyka-Mostowski-Parusiński [KKP00] settled this conjecture by showing a stronger result that the secant has a finite length in 𝕊n1\mathbb{S}^{n-1}. However, the precise convergence rate is not yet completed revealed. If θ0\theta_{0} does not belong to ker(2f(0))\ker\,(\nabla^{2}f(0)), one obtains that θ0\theta_{0} has to be an eigenvector of the hessian for a positive eigenvalue λ\lambda, and the solution decays with higher order asymptotics |x(t)ceλtθ0|=o(eλt)|x(t)-ce^{-\lambda t}\theta_{0}|=o(e^{-\lambda t}) for some cc\in\mathbb{R}. If θ0\theta_{0} belongs to the kernel of hessian, however, we merely know that the convergence takes place between algebraic rates tα1|x|tα2t^{-\alpha_{1}}\lesssim|x|\lesssim t^{-\alpha_{2}} for some 0<α2α110<\alpha_{2}\leq\alpha_{1}\leq 1. It is unknown if each solution has a specific convergence rate |x|tγ|x|\sim t^{-\gamma} some γ>0\gamma>0.


Let us briefly summarize the main motive and results of the paper. In the slow decaying regime, the equation becomes well-approximated by a gradient flow on the kernel of Σ\mathcal{L}_{\Sigma}, the linearization of Σ\mathcal{M}_{\Sigma} at u0u\equiv 0. The potential f:kerΣJf:\ker\mathcal{L}_{\Sigma}\approx\mathbb{R}^{J}\to\mathbb{R} is given by the pull-back of Σ\mathcal{F}_{\Sigma} through Lyapunov-Schmidt reduction (see Proposition 2.1). Near the origin, the flow dynamics shall be determined by the first non-constant homogeneous polynomial appearing in the expansion of potential ff. Let us denote this polynomial by fpf_{p}. We reserve p3p\geq 3 to denote the degree of this polynomial and we call it the order of integrability. When the gradient flow x=fp(x)x^{\prime}=-\nabla f_{p}(x) is considered, one readily finds a radial solution converging to the origin whenever there exists a critical point of f^p:=fp|𝕊J1\hat{f}_{p}:=f_{p}|_{\mathbb{S}^{J-1}} with positive critical value. The slowly decaying solutions of [AS88] and [CCR15] are obtained as perturbations of such radial ansatzes.


In Theorem 2.3 and 2.5, we prove slowly decaying solution u(ω,t)u(\omega,t) has a dichotomy on its convergence rate that either t1p2u(,t)L2t^{\frac{1}{p-2}}\|u(\cdot,t)\|_{L^{2}} converges to a positive number or diverges to infinity. In the first case, we further show u(,t)/u(,t)L2u(\cdot,t)/\|u(\cdot,t)\|_{L^{2}} converges smoothly to a critical point of f^p\hat{f}_{p}, say vkerΣv\in\ker\mathcal{L}_{\Sigma}, with positive critical value. This shows the secant u(,t)/u(,t)L2u(\cdot,t)/\|u(\cdot,t)\|_{L^{2}} has a limit and confirms Thom’s gradient conjecture for the concerned case. Note that one may interpret the theorem as the higher order asymptotics u(t)=ct1p2v+o(t1p2)u(t)=ct^{-\frac{1}{p-2}}v+o(t^{-\frac{1}{p-2}}), where cc is a constant of pp and fp(v)f_{p}(v). Next, the second alternative shows a solution possibly decays even at a slower rate than t1p2t^{-\frac{1}{p-2}}. As seen in the classical gradient flow of potential f=x14+x28f=x_{1}^{4}+x_{2}^{8} on 2\mathbb{R}^{2} (here, p=4p=4 but a solution may decay at rate t182t^{-\frac{1}{8-2}}), the second alternative can actually take place. We show this can only occur if f^p\hat{f}_{p} admits critical point(s) of zero critical value and the secant u(,t)/u(,t)L2u(\cdot,t)/\|u(\cdot,t)\|_{L^{2}} accumulates on those critical point(s) as tt\to\infty. It remains open whether the secant u(,t)/u(,t)L2u(\cdot,t)/\|u(\cdot,t)\|_{L^{2}} has a unique limit in the second case, and thereby, Thom’s conjecture holds for all slowly decaying solutions. We would like to point out that Theorem 2.5 is also novel for finite-dimensional gradient flows. For a finite-dimensional gradient flow given by x=f(x)x^{\prime}=-\nabla f(x), the theorem states that there exists an integer p3p\geq 3 which depends only on ff, such that any slowly converging solution satisfies either t1p2|x|=c(1+o(1))t^{\frac{1}{p-2}}|x|=c(1+o(1)) for some c(0,)c\in(0,\infty), or t1p2|x|t^{\frac{1}{p-2}}|x|\to\infty.


In Theorem 2.4 and 2.6, we show the higher order asymptotic behavior of fast (exponentially) decaying solutions for elliptic and parabolic equations, respectively. This type of result has been expected among researchers and has been obtained for specific problems such as [SS86][CM22]. Nevertheless, we could not find proper literature covering the general forms (1.1) and (1.2), so we provide a proof in Section 4. The elliptic equation (1.1) can be viewed as a perturbation of the second order ODE u′′mu+Σu=0u^{\prime\prime}-mu^{\prime}+\mathcal{L}_{\Sigma}u=0. Suppose Σ\mathcal{L}_{\Sigma} has an eigenvalue larger than 41m24^{-1}m^{2}, a solution might oscillate while decaying exponentially. This type of solutions has been constructed for minimal graphs over Simons cones in 4\mathbb{R}^{4} and 6\mathbb{R}^{6}. See [CP18, Remark 1.21] and [BDGG69] for more details. For this reason, the original form of Thom’s gradient conjecture is not true in the elliptic problem. Moreover, if 41m24^{-1}m^{2} is an eigenvalue of Σ\mathcal{L}_{\Sigma} and m<0m<0, a resonance might occur and result in a solution that decays at a rate te21mtte^{2^{-1}mt}. Note [HS85] considered this possibility for minimal graphs over stable (but not strictly stable) minimal cones.

This paper is organized as follows. In Section 2, we introduce the notation, condition, spectral property of linearized operator Σ\mathcal{L}_{\Sigma}, and main theorems. In Section 3, we set up the elliptic problem (1.1) as a first order ODE system on function spaces. In Section 4, Theorem 2.4 and Theorem 2.6 for exponentially decaying solutions to elliptic and parabolic equations are proved, respectively. In Section 5, we show in Proposition 5.1 that slowly converging solutions to (1.1) are governed by a finite-dimensional gradient flow with a small perturbation. The parabolic analogue, Proposition 6.1, is proved in Section 6. In Section 7, we complete the proofs for Theorems 2.3 and 2.5 by analyzing a finite-dimensional gradient flow with a small perturbation, using a version of the Łojasiewicz argument motivated by [KKP00]. Appendix A contains auxiliary tools we need in the paper. In Appendix B, we show that (rescaled) mean curvature flows can be written in the form (1.2).

2. Preliminary

Let us introduce our setting to study (1.1) and (1.2) and state the main results. Let Σ\Sigma be a closed nn-dimensional Riemannian manifold and dμd\mu be a smooth volume form on Σ\Sigma which is mutually absolutely continuous with respect to the volume form induced by the metric. Let 𝐕Σ\mathbf{V}\to\Sigma be a smooth vector bundle equipped with a smooth inner product ,\left\langle\cdot,\cdot\right\rangle. For u𝐕u\in\mathbf{V}, we write |u|2:=u,u|u|^{2}:=\left\langle u,u\right\rangle. For ωΣ\omega\in\Sigma, we denote by 𝐕ω\mathbf{V}_{\omega} the fiber over ω\omega. Let ∇̸\not{\nabla} be a connection on 𝐕\mathbf{V} which is compatible with the inner product.

We denote by L2(Σ;𝐕)L^{2}(\Sigma;\mathbf{V}) the space of L2L^{2}-sections of 𝐕\mathbf{V} with respect to dμd\mu. Namely, a section uu belongs to L2(Σ;𝐕)L^{2}(\Sigma;\mathbf{V}) provided

uL22:=Σ|u|2𝑑μ<.\displaystyle\|u\|^{2}_{L^{2}}:=\int_{\Sigma}|u|^{2}\,d\mu<\infty.

For 0\ell\in\mathbb{N}_{0}, we denote by H(Σ;𝐕)H^{\ell}(\Sigma;\mathbf{V}) the collection of sections that satisfies

uH2:=i=0Σ|∇̸iu|2𝑑μ<.\displaystyle\|u\|^{2}_{H^{\ell}}:=\sum_{i=0}^{\ell}\int_{\Sigma}\left|\not{\nabla}^{i}u\right|^{2}\,d\mu<\infty.

For <a<b-\infty<a<b\leq\infty, we define Qa,b:=Σ×[a,b)Q_{a,b}:=\Sigma\times[a,b) and equip Qa,bQ_{a,b} with the product metric. We denote by 𝐕~\widetilde{\mathbf{V}} the pull back bundle of 𝐕\mathbf{V} through the projection Qa,bΣQ_{a,b}\to\Sigma. For uCs(Qa,b;𝐕~)u\in C^{s}(Q_{a,b};\widetilde{\mathbf{V}}) and t[a,b)t\in[a,b), u(t)Cs(Σ;𝐕)u(t)\in C^{s}(\Sigma;\mathbf{V}) is defined through u(t)(ω):=u(ω,t)u(t)(\omega):=u(\omega,t). Its HH^{\ell}-norm is denoted by

uH(t):=u(t)H.\|u\|_{H^{\ell}}(t):=\|u(t)\|_{H^{\ell}}.

Also,

(2.1) uCs(t):=supωΣsupk+s|ktk∇̸u(ω,t)|.\displaystyle\|u\|_{C^{s}}(t):=\sup_{\omega\in\Sigma}\sup_{k+\ell\leq s}\left|\frac{\partial^{k}}{\partial t^{k}}\not{\nabla}^{\ell}u(\omega,t)\right|.

Here, note that the norms of time derivatives are included in the definition of Cs(t)\|\cdot\|_{C^{s}}(t). We often use uu^{\prime} as an abbreviation of ut\frac{\partial u}{\partial t} .

Our assumption on Σ\mathcal{M}_{\Sigma} is almost identical to the one in [Sim83]. The only difference is that the volume form, dμd\mu, is not necessarily the one induced from the Riemannian metric. Let Σ\mathcal{F}_{\Sigma} be a functional defined for uC1(Σ;𝐕)u\in C^{1}(\Sigma;\mathbf{V}) by

(2.2) Σ(u):=ΣF(ω,u,∇̸u)𝑑μ.\displaystyle\mathcal{F}_{\Sigma}(u):=\int_{\Sigma}F(\omega,u,\not{\nabla}u)\,d\mu.

Here the integrand FF satisfies the following:

  1. (1)

    F=F(ω,z,p)F=F(\omega,z,p) is a smooth function defined on an open set of 𝐕×(TΣ𝐕)\mathbf{V}\times(T\Sigma\otimes\mathbf{V}) that contains the zero section.

  2. (2)

    For each ωΣ\omega\in\Sigma, F(ω,,)F(\omega,\cdot,\cdot) is analytic on 𝐕ω×(TωΣ𝐕ω)\mathbf{V}_{\omega}\times(T_{\omega}\Sigma\otimes\mathbf{V}_{\omega}).

  3. (3)

    FF satisfies the Legendre-Hadamard ellipticity condition

    Dp2F(ω,0,0)[ηξ,ηξ]c|η|2|ξ|2,\displaystyle D^{2}_{p}F(\omega,0,0)[\eta\otimes\xi,\eta\otimes\xi]\geq c|\eta|^{2}|\xi|^{2},

    for c>0c>0 independent of ωΣ\omega\in\Sigma, ηTωΣ\eta\in T_{\omega}\Sigma and ξ𝐕ω\xi\in\mathbf{V}_{\omega}.

Σ\mathcal{M}_{\Sigma} is defined to be the negative Euler-Lagrange operator of Σ\mathcal{F}_{\Sigma}. Namely, for any ζC(Σ;𝐕)\zeta\in C^{\infty}(\Sigma;\mathbf{V}),

(2.3) ΣΣ(u),ζ𝑑μ=ddsΣ(u+sζ)|s=0.\displaystyle\int_{\Sigma}\left\langle\mathcal{M}_{\Sigma}(u),\zeta\right\rangle\,d\mu=-\frac{d}{ds}\mathcal{F}_{\Sigma}(u+s\zeta)|_{s=0}.

We further assume 0 is a critical point of Σ\mathcal{F}_{\Sigma}. Namely, we assume Σ(0)=0\mathcal{M}_{\Sigma}(0)=0. We denote by Σ\mathcal{L}_{\Sigma} the linearization of Σ\mathcal{M}_{\Sigma} at 0. It is clear that the difference between Σ(u)\mathcal{M}_{\Sigma}(u) and Σu\mathcal{L}_{\Sigma}u is a quadratic term of the form

(2.4) ΣuΣu=j=02cj∇̸ju,\mathcal{M}_{\Sigma}u-\mathcal{L}_{\Sigma}u=\sum_{j=0}^{2}c_{j}\cdot\not{\nabla}^{j}u,

where cj=cj(ω,u,∇̸u)c_{j}=c_{j}(\omega,u,\not{\nabla}u) are smooth with cj(ω,0,0)=0c_{j}(\omega,0,0)=0. The Legendre-Hadamard condition implies Σ\mathcal{L}_{\Sigma} is elliptic. In particular, there exists a constant C>0C>0 such that for any uC2(Σ;𝐕)u\in C^{2}(\Sigma;\mathbf{V}),

(2.5) ΣΣu,u+C|u|2dμuH12.\displaystyle\int_{\Sigma}-\left\langle\mathcal{L}_{\Sigma}u,u\right\rangle+C|u|^{2}\,d\mu\approx\|u\|^{2}_{H^{1}}.

Furthermore, Σ\mathcal{L}_{\Sigma} is self-adjoint with respect to dμd\mu. This can be seen from

ΣΣu,v𝑑μ=2s1s2Σ(s1u+s2v)|s1=s2=0.\displaystyle\int_{\Sigma}\left\langle\mathcal{L}_{\Sigma}u,v\right\rangle\,d\mu=-\frac{\partial^{2}}{\partial s_{1}\partial s_{2}}\mathcal{F}_{\Sigma}(s_{1}u+s_{2}v)\bigg{|}_{s_{1}=s_{2}=0}.

We suppose N1(u)N_{1}(u) and N2(u)N_{2}(u) in (1.1) and (1.2) are of the form

(2.6) N1(u)=a1Du+a2u+a3Σ(u),N2(u)=b1Σ(u).\begin{split}N_{1}(u)=&a_{1}\cdot Du^{\prime}+a_{2}\cdot u^{\prime}+a_{3}\cdot\mathcal{M}_{\Sigma}(u),\\ N_{2}(u)=&b_{1}\cdot\mathcal{M}_{\Sigma}(u).\end{split}

Here ai=ai(ω,u,∇̸u,u)a_{i}=a_{i}(\omega,u,\not{\nabla}u,u^{\prime}) and b1=b1(ω,u,∇̸u)b_{1}=b_{1}(\omega,u,\not{\nabla}u) are smooth with ai(ω,0,0,0)=0a_{i}(\omega,0,0,0)=0 and b1(ω,0,0)=0b_{1}(\omega,0,0)=0; D={t,∇̸}D=\left\{\frac{\partial}{\partial t},\not{\nabla}\right\}.

Let λ1λ2\lambda_{1}\geq\lambda_{2}\geq\dots be the eigenvalues of Σ\mathcal{L}_{\Sigma} and φ1,φ2,\varphi_{1},\varphi_{2},\dots be the corresponding eigensections which form a complete orthonormal basis of L2(Σ;𝐕)L^{2}(\Sigma;\mathbf{V}). We separate \mathbb{N} into four parts according to the eigenvalues.

(2.7) I1:={i:λi>41m2},I2:={i:λi=41m2},I3:={i:λi=0},I4:=(I1I2I3).\begin{split}I_{1}:=&\{i\in\mathbb{N}\,:\,\lambda_{i}>4^{-1}m^{2}\},\ I_{2}:=\{i\in\mathbb{N}\,:\,\lambda_{i}=4^{-1}m^{2}\},\\ I_{3}:=&\{i\in\mathbb{N}\,:\,\lambda_{i}=0\},\ I_{4}:=\mathbb{N}\setminus(I_{1}\cup I_{2}\cup I_{3}).\end{split}

Note that I1I_{1}, I2I_{2} and I3I_{3} are (possibly empty) finite sets and kerΣ\ker\mathcal{L}_{\Sigma} is spanned by {φi}iI3\{\varphi_{i}\}_{i\in I_{3}}. Let JJ be the cardinality of I3I_{3}, the dimension of kerΣ\ker\mathcal{L}_{\Sigma}. This implies I3={ι+1,ι+2,,ι+J}I_{3}=\{\iota+1,\iota+2,\dots,\iota+J\} for some ι0\iota\in\mathbb{N}_{0}.

We denote by ΠT\Pi^{T} and Π\Pi^{\perp} the orthogonal projection of L2(Σ;𝐕)L^{2}(\Sigma;\mathbf{V}) to kerΣ\ker\mathcal{L}_{\Sigma} and (kerΣ)\left(\ker\mathcal{L}_{\Sigma}\right)^{\perp} respectively. The following is a version of the implicit function theorem. See [Sim96, §3].

Proposition 2.1 (Lyapunov–Schmidt reduction).

Let BρB_{\rho} be the open ball of radius ρ\rho in L2(Σ;𝐕)L^{2}(\Sigma;\mathbf{V}). There exist ρ>0\rho>0 and a map

H:kerΣBρC(Σ;𝐕)(kerΣ)H:\ker\mathcal{L}_{\Sigma}\cap B_{\rho}\rightarrow C^{\infty}(\Sigma;\mathbf{V})\cap(\ker\mathcal{L}_{\Sigma})^{\perp}

such that the following statements hold. First, for any k2k\geq 2 and α(0,1)\alpha\in(0,1), HH is an analytic map from kerΣBρ\ker\mathcal{L}_{\Sigma}\cap B_{\rho} to Ck,α(Σ;𝐕)C^{k,\alpha}(\Sigma;\mathbf{V}). Second,

H(0)=0,DH(0)=0.H(0)=0,\,DH(0)=0.

Lastly,

{ΠΣ(v+H(v))=0,ΠTΣ(v+H(v))=f(v).\begin{cases}\begin{aligned} \Pi^{\perp}\mathcal{M}_{\Sigma}(v+H(v))&=0,\\ \Pi^{T}\mathcal{M}_{\Sigma}(v+H(v))&=-\nabla f(v).\end{aligned}\end{cases}

Here f:kerΣBρf:\ker\mathcal{L}_{\Sigma}\cap B_{\rho}\to\mathbb{R} is given by f(v):=Σ(v+H(v)).f(v):=\mathcal{F}_{\Sigma}(v+H(v)).

The function ff plays a crucial role in this paper and we will call it reduced functional. The reduced functional ff is real analytic with f(0)=0\nabla f(0)=0 and 2f(0)=0\nabla^{2}f(0)=0. We may view ff as an analytic function defined on an open ball in J\mathbb{R}^{J} through the identification

(2.8) (x1,x2,,xJ)j=1Jxjφι+j.\displaystyle(x_{1},x_{2},\dots,x^{J})\mapsto\sum_{j=1}^{J}x_{j}\varphi_{\iota+j}.

Let us review the integrable condition. The kernel kerΣ\ker\mathcal{L}_{\Sigma} is called integrable if for any vkerΣv\in\ker\mathcal{L}_{\Sigma}, there exists a family {vs}s(0,1)C2(Σ;𝐕)\{v_{s}\}_{s\in(0,1)}\subset C^{2}(\Sigma;\mathbf{V}) such that vs0v_{s}\to 0 in C2(Σ;𝐕)C^{2}(\Sigma;\mathbf{V}), Σ(vs)0\mathcal{M}_{\Sigma}(v_{s})\equiv 0 and lims0vs/s=v\lim_{s\to 0}v_{s}/s=v in L2(Σ;𝐕)L^{2}(\Sigma;\mathbf{V}). It is well-known [AS88, Lemma 1] that kerΣ\ker\mathcal{L}_{\Sigma} is integrable if and only if the reduced functional ff is a constant. Moreover, if integrable condition is satisfied, any decaying solution to (1.1) or (1.2) decays exponentially [AS88, CCR15]. For this reason, whenever non-exponentially decaying solution is considered, the integrable condition should necessarily fails. Namely, the reduced functional ff is not a constant function. In particular, there exists an integer p3p\geq 3 such that

(2.9) f=f(0)+jpfj,f=f(0)+\sum_{j\geq p}f_{j},

where fjf_{j} are homogeneous polynomials with degree jj and fp0f_{p}\not\equiv 0. This integer pp is called the order of integrability [CCR15].

As explained in Introduction, the gradient flow of fpf_{p} has a dominant role in the asymptotic behavior. Let f^p\hat{f}_{p} be the restriction of fpf_{p} on {wkerΣ:wL2=1}𝕊J1\{w\in\ker\mathcal{L}_{\Sigma}\,:\,\|w\|_{L^{2}}=1\}\approx\mathbb{S}^{J-1}. Consider the critical points of f^p\hat{f}_{p}:

(2.10) 𝐂:={wkerΣ:wL2=1andwis a critical point off^p}.\mathbf{C}:=\left\{w\in\ker\mathcal{L}_{\Sigma}\,:\,\|w\|_{L^{2}}=1\ \textup{and}\ w\ \textup{is a critical point of}\ \hat{f}_{p}\right\}.

If w𝐂w\in\mathbf{C} satisfies fp(w)>0f_{p}(w)>0, then one checks that

x(t)=[p(p2)fp(w)(t+c)]1p2wx(t)=[p(p-2)f_{p}(w)(t+c)]^{-\frac{1}{p-2}}w

becomes a radial solution to the flow x=fp(x)x^{\prime}=-\nabla f_{p}(x). The higher order asymptotics in Theorem 2.3 and 2.5 will be modeled on such solutions.

Definition 2.2 (Adams-Simon conditions. c.f. (4.1) in [AS88]).

We say Σ\Sigma satisfies the Adams-Simon non-negativity condition for (1.2) if there exists w𝐂w\in\mathbf{C} such that fp(w)0{f}_{p}(w)\geq 0. We say Σ\Sigma satisfies the Adams-Simon non-negativity condition for (1.1) if there exists w𝐂w\in\mathbf{C} such that m1fp(w)0m^{-1}{f}_{p}(w)\geq 0.

In both equations, the Adams-Simons positivity conditions are defined similarly by requiring that the critical values are positive.

Let us state main results concerning the asymptotic behavior and the convergence rate of decaying solutions. We begin with the elliptic equation (1.1).

Theorem 2.3 (slow decay in elliptic equation).

Let uC(Q0,,𝐕~)u\in C^{\infty}(Q_{0,\infty},\widetilde{\mathbf{V}}) be a solution to (1.1) with uC1(t)=o(1)\|u\|_{C^{1}}(t)=o(1) that does not decay exponentially as tt\to\infty. Then Σ\Sigma satisfies the Adams-Simon non-negativity condition for (1.1). Moreover, one of the following alternatives holds:

  1. (1)

    We have

    limtt1/(p2)u(t)L2=β(0,).\displaystyle\lim_{t\to\infty}t^{1/(p-2)}\|u(t)\|_{L^{2}}=\beta\in(0,\infty).

    Moreover,

    limtu(t)/u(t)L2=winC(Σ;𝐕),\displaystyle\lim_{t\to\infty}{u(t)}/{\|u(t)\|_{L^{2}}}=w\ \textup{in}\ C^{\infty}(\Sigma;\mathbf{V}),

    where w𝐂w\in\mathbf{C} with m1f^p(w)=1/(p(p2)βp2)>0m^{-1}\hat{f}_{p}(w)=1/({p(p-2)\beta^{p-2}})>0 .

  2. (2)

    We have

    limtt1/(p2)u(t)L2=.\displaystyle\lim_{t\to\infty}t^{1/(p-2)}\|u(t)\|_{L^{2}}=\infty.

    Moreover,

    limtdistCk(u(t)/u(t)L2,𝐂{w:f^p(w)=0})=0for allk.\lim_{t\to\infty}\textup{dist}_{C^{k}}\bigg{(}u(t)/\|u(t)\|_{L^{2}}\ ,\ \mathbf{C}\cap\{w:\hat{f}_{p}(w)=0\}\bigg{)}=0\ \textup{for all}\ k\in\mathbb{N}.
Theorem 2.4 (fast decay in elliptic equation).

Let uC(Q0,,𝐕~)u\in C^{\infty}(Q_{0,\infty},\widetilde{\mathbf{V}}) be a solution to (1.1) with uC1(t)=O(eεt)\|u\|_{C^{1}}(t)=O(e^{-\varepsilon t}) for some ε>0\varepsilon>0 as tt\to\infty. If uu is not the zero section, then one of the following alternatives holds:

  1. (1)

    There exists an eigenvalue λ<41m2\lambda<4^{-1}m^{2} of Σ\mathcal{L}_{\Sigma} such that

    γ+<0andlimteγ+tu(t)L2(0,),\displaystyle\gamma^{+}<0\ \textup{and}\ \lim_{t\to\infty}e^{-\gamma^{+}t}\|u(t)\|_{L^{2}}\in(0,\infty),

    or

    γ<0andlimteγtu(t)L2(0,).\displaystyle\gamma^{-}<0\ \textup{and}\ \lim_{t\to\infty}e^{-\gamma^{-}t}\|u(t)\|_{L^{2}}\in(0,\infty).

    Here γ±=21m±41m2λ.\gamma^{\pm}=2^{-1}m\pm\sqrt{4^{-1}m^{2}-\lambda}. Moreover, for some eigensection ww with Σw=λw\mathcal{L}_{\Sigma}w=\lambda w,

    limtu(t)/u(t)L2=winC(Σ;𝐕).\displaystyle\lim_{t\to\infty}{u(t)}/{\|u(t)\|_{L^{2}}}=w\ \textup{in}\ C^{\infty}(\Sigma;\mathbf{V}).
  2. (2)

    We have

    m<0andlimtt1e21mtu(t)L2(0,).\displaystyle m<0\ \textup{and}\ \lim_{t\to\infty}t^{-1}e^{-2^{-1}mt}\|u(t)\|_{L^{2}}\in(0,\infty).

    Moreover, for some eigensection ww with Σw=41m2w\mathcal{L}_{\Sigma}w=4^{-1}m^{2}w,

    limtu(t)/u(t)L2=winC(Σ;𝐕).\displaystyle\lim_{t\to\infty}{u(t)}/{\|u(t)\|_{L^{2}}}=w\ \textup{in}\ C^{\infty}(\Sigma;\mathbf{V}).
  3. (3)

    We have

    m<0andlim supte21mtu(t)L2(0,).\displaystyle m<0\ \textup{and}\ \limsup_{t\to\infty}e^{-2^{-1}mt}\|u(t)\|_{L^{2}}\in(0,\infty).

    Moreover, there exist wiw_{i}\in\mathbb{C} for iI1i\in I_{1} and cic_{i}\in\mathbb{R} for iI2i\in I_{2} such that

    limt(e21mtu(t)iI1Re(wie𝐢βit)φiiI2ciφi)=0inC(Σ;𝐕).\displaystyle\lim_{t\to\infty}\left(e^{-2^{-1}mt}u(t)-\sum_{i\in I_{1}}\textup{Re}\left(w_{i}e^{\mathbf{i}\beta_{i}t}\right)\varphi_{i}-\sum_{i\in I_{2}}c_{i}\varphi_{i}\right)=0\ \textup{in}\ C^{\infty}(\Sigma;\mathbf{V}).

    Here βi=λi41m2\beta_{i}=\sqrt{\lambda_{i}-4^{-1}m^{2}} and 𝐢=1\mathbf{i}=\sqrt{-1}.

Next, we consider decaying solutions to the parabolic equation (1.2).

Theorem 2.5 (slow decay in parabolic equation).

Let uC(Q0,,𝐕~)u\in C^{\infty}(Q_{0,\infty},\widetilde{\mathbf{V}}) be a solution to (1.2) with uHn+4(t)=o(1)\|u\|_{H^{n+4}}(t)=o(1) that does not decay exponentially as tt\to\infty. Then Σ\Sigma satisfies the Adams-Simon non-negativity condition for (1.2). Moreover, one of the following alternatives holds:

  1. (1)

    We have

    limtt1/(p2)u(t)L2=β(0,).\displaystyle\lim_{t\to\infty}t^{1/(p-2)}\|u(t)\|_{L^{2}}=\beta\in(0,\infty).

    Moreover,

    limtu(t)/u(t)L2=winC(Σ;𝐕),\displaystyle\lim_{t\to\infty}{u(t)}/{\|u(t)\|_{L^{2}}}=w\ \textup{in}\ C^{\infty}(\Sigma;\mathbf{V}),

    where wkerΣw\in\ker\mathcal{L}_{\Sigma} and ww is a critical point of f^p\hat{f}_{p} with f^p(w)=1/(p(p2)βp2)>0\hat{f}_{p}(w)=1/({p(p-2)\beta^{p-2}})>0 .

  2. (2)

    We have

    limtt1/(p2)u(t)L2=.\displaystyle\lim_{t\to\infty}t^{1/(p-2)}\|u(t)\|_{L^{2}}=\infty.

    Moreover,

    limtdistCk(u(t)/u(t)L2,𝐂{w:f^p(w)=0})=0for allk.\lim_{t\to\infty}\textup{dist}_{C^{k}}\bigg{(}u(t)/\|u(t)\|_{L^{2}}\ ,\ \mathbf{C}\cap\{w:\hat{f}_{p}(w)=0\}\bigg{)}=0\ \textup{for all}\ k\in\mathbb{N}.
Theorem 2.6 (fast decay in parabolic equation).

Let uC(Q0,,𝐕~)u\in C^{\infty}(Q_{0,\infty},\widetilde{\mathbf{V}}) be a solution to (1.2) with uHn+4(t)=O(eεt)\|u\|_{H^{n+4}}(t)=O(e^{-\varepsilon t}) for some ε>0\varepsilon>0 as tt\to\infty. Then there exists a negative eigenvalue λ\lambda of Σ\mathcal{L}_{\Sigma} such that

limteλtu(t)L2(0,).\displaystyle\lim_{t\to\infty}e^{-\lambda t}\|u(t)\|_{L^{2}}\in(0,\infty).

Moreover, for some eigensection ww with Σw=λw\mathcal{L}_{\Sigma}w=\lambda w,

limtu(t)/u(t)L2=winC(Σ;𝐕).\displaystyle\lim_{t\to\infty}{u(t)}/{\|u(t)\|_{L^{2}}}=w\ \textup{in}\ C^{\infty}(\Sigma;\mathbf{V}).

In fact, it is possible that no slowly converging solution exists even in the presence of non-integrable kernel. e.g., we refer [CCK21][CS20]. Theorem 2.3 and 2.5 provide a criterion for this non-existence result.

Corollary 2.7.

Let uC(Q0,,𝐕~)u\in C^{\infty}(Q_{0,\infty},\widetilde{\mathbf{V}}) be a solution to (1.1) with uC1(t)=o(1)\|u\|_{C^{1}}(t)=o(1) as tt\to\infty. Suppose the Adams-Simon non-negative condition for (1.1) fails. Then uu decays exponentially.

Corollary 2.8.

Let uC(Q0,,𝐕~)u\in C^{\infty}(Q_{0,\infty},\widetilde{\mathbf{V}}) be a solution to (1.2) with uHn+4(t)=o(1)\|u\|_{H^{n+4}}(t)=o(1) as tt\to\infty. Suppose the Adams-Simon non-negative condition for (1.2) fails. Then uu decays exponentially.

Remark 2.9.

We note that the main results, Theorems 2.3-2.6, can be generalized to cover other cases.

  1. (1)

    We may consider ancient solutions defined on t(,0]t\in(-\infty,0] which decay to zero when tt approaches minus infinity. For ancient solutions to the elliptic equation (1.1), we can simply perform a change of variable ttt\mapsto-t. For ancient solutions to the parabolic equation (1.2), simple changes in the proofs yield Theorems 2.5 and 2.6.

  2. (2)

    We may consider the hyperbolic equation of the form

    u′′mu+Σu=N1(u).\displaystyle-u^{\prime\prime}-mu^{\prime}+\mathcal{M}_{\Sigma}u=N_{1}(u).

    The only difference from the elliptic equation (1.1) is that, due to the lack of elliptic regularity, we need to assume uCs(t)\|u\|_{C^{s}}(t) decays to zero for all ss\in\mathbb{N}.

  3. (3)

    Due to volume normalization, the normalized Yamabe flow (NYF) involves a non-local term, the total scalar curvature, in its speed. Nevertheless, our proof for Theorem 2.5 can be modified to cover the NYF. Specifically, all parts of the proof, except for Lemma 6.2, remain the same. Lemma 6.2 can be easily established with the explicit forms of the NYF.

  4. (4)

    The main results apply for the gradient flow on n\mathbb{R}^{n} by choosing Σ\Sigma a point, dμd\mu a Dirac mass, and VV a trivial n\mathbb{R}^{n} bundle. Indeed, (1.1) and (1.2) cover wider class of ODE systems which evolve under analytic potentials.


We finish this section with two lemmas related to the Lyapunov-Schmidt reduction map HH given in Proposition 2.1. We decompose uu as follows: let uT=ΠTuu^{T}=\Pi^{T}u and define u~\tilde{u}^{\perp} by

(2.11) u=uT+H(uT)+u~.\displaystyle u=u^{T}+H(u^{T})+\tilde{u}^{\perp}.

Since uTkerΣu^{T}\in\ker\mathcal{L}_{\Sigma}, uTu^{T} can be written as a linear combination of {φι+j}1jJ\{\varphi_{\iota+j}\}_{1\leq j\leq J} as

(2.12) uT=j=1Jxjφι+j.u^{T}=\sum_{j=1}^{J}x_{j}\varphi_{\iota+j}.
Lemma 2.10.

Let ρ>0\rho>0 be the constant given in Proposition 2.1 and uC2,α(Σ;𝐕)u\in C^{2,\alpha}(\Sigma;\mathbf{V}) be a section with uTL221ρ\|u^{T}\|_{L^{2}}\leq 2^{-1}\rho. Let x=(x1,x2,,xJ)x=(x_{1},x_{2},\dots,x_{J}) be given by (2.12). Then for any α(0,1)\alpha\in(0,1), there exists a positive constant C=C(Σ,α)C=C(\mathcal{M}_{\Sigma},\alpha) such that

(2.13) |Σ(u)+f(x)Σu~|CuC2,α(Σ)u~C2,α(Σ).\left|\mathcal{M}_{\Sigma}(u)+\nabla f(x)-\mathcal{L}_{\Sigma}\tilde{u}^{\perp}\right|\leq C\|u\|_{C^{2,\alpha}(\Sigma)}\|\tilde{u}^{\perp}\|_{C^{2,\alpha}(\Sigma)}.
Proof.

In the proof we use CC to represent a positive constant that depends on Σ\mathcal{M}_{\Sigma}, α\alpha, and its value may vary from one line to another. Let ¯Σ\bar{\mathcal{L}}_{\Sigma} be the linearization of Σ\mathcal{M}_{\Sigma} at uT+H(uT)u^{T}+H(u^{T}). From Proposition 2.1 (Lyapunov-Schmidt reduction) and (2.11),

|Σ(u)+f(x)Σu~|\displaystyle\left|\mathcal{M}_{\Sigma}(u)+\nabla f(x)-\mathcal{L}_{\Sigma}\tilde{u}^{\perp}\right|
=\displaystyle= |Σ(uT+H(uT)+u~)Σ(uT+H(uT))Σu~|\displaystyle\left|\mathcal{M}_{\Sigma}(u^{T}+H(u^{T})+\tilde{u}^{\perp})-\mathcal{M}_{\Sigma}(u^{T}+H(u^{T}))-\mathcal{L}_{\Sigma}\tilde{u}^{\perp}\right|
\displaystyle\leq |¯Σu~Σu~|+|Σ(uT+H(uT)+u~)Σ(uT+H(uT))¯Σu~|.\displaystyle|\bar{\mathcal{L}}_{\Sigma}\tilde{u}^{\perp}-\mathcal{L}_{\Sigma}\tilde{u}^{\perp}|+|\mathcal{M}_{\Sigma}(u^{T}+H(u^{T})+\tilde{u}^{\perp})-\mathcal{M}_{\Sigma}(u^{T}+H(u^{T}))-\bar{\mathcal{L}}_{\Sigma}\tilde{u}^{\perp}|.

Because Σ\mathcal{M}_{\Sigma} is an analytic map from C2,α(Σ;𝐕)C^{2,\alpha}(\Sigma;\mathbf{V}) to Cα(Σ;𝐕)C^{\alpha}(\Sigma;\mathbf{V}), the above is bounded by

CuT+H(uT)C2,α(Σ)u~C2,α(Σ)+Cu~C2,α(Σ)2\displaystyle C\|u^{T}+H(u^{T})\|_{C^{2,\alpha}(\Sigma)}\|\tilde{u}^{\perp}\|_{C^{2,\alpha}(\Sigma)}+C\|\tilde{u}^{\perp}\|^{2}_{C^{2,\alpha}(\Sigma)}

From (2.12) and the analyticity of HH, there holds uT+H(uT)C2,α(Σ)C|x|CuC2,α(Σ).\|u^{T}+H(u^{T})\|_{C^{2,\alpha}(\Sigma)}\leq C|x|\leq C\|u\|_{C^{2,\alpha}(\Sigma)}. Together with (2.11), we have u~C2,α(Σ)CuC2,α(Σ).\|\tilde{u}^{\perp}\|_{C^{2,\alpha}(\Sigma)}\leq C\|u\|_{C^{2,\alpha}(\Sigma)}. Hence the assertion holds.

Lemma 2.11 (boundedness of decomposition).

Let ρ>0\rho>0 be the constant given in Proposition 2.1 and uC(Qa,b;𝐕~)u\in C^{\infty}(Q_{a,b};\widetilde{\mathbf{V}}) be a smooth section with uCs(t)M<\|u\|_{C^{s}}(t)\leq M<\infty. Then when uTL2(t)21ρ\|u^{T}\|_{L^{2}}(t)\leq 2^{-1}\rho,

uTCs(t)+H(uT)Cs(t)+u~Cs(t)CuCs(t),\displaystyle\|u^{T}\|_{C^{s}}(t)+\|H(u^{T})\|_{C^{s}}(t)+\|\tilde{u}^{\perp}\|_{C^{s}}(t)\leq C\|u\|_{C^{s}}(t),

for some constant C=C(Σ,M,s)C=C(\mathcal{M}_{\Sigma},M,s).

Proof.

In the proof we use CC to represent a positive constant that depends on Σ\mathcal{M}_{\Sigma}, MM, ss, and its value may vary from one line to another. Let x(t)=(x1(t),,xJ(t))x(t)=(x_{1}(t),\dots,x_{J}(t)) be the coefficients given by (2.12) for uT(t)u^{T}(t). Then

uTCs(t)Ck=0s|dkdtkx(t)|CuCs(t).\displaystyle\|u^{T}\|_{C^{s}}(t)\leq C\sum_{k=0}^{s}\left|\frac{d^{k}}{dt^{k}}x(t)\right|\leq C\|u\|_{C^{s}}(t).

Through (2.8), we may view HH as a map from an open ball in J\mathbb{R}^{J} to C(Σ;𝐕)C^{\infty}(\Sigma;\mathbf{V}). We abuse the notation and write H(x(t))H(x(t)) for H(uT)H(u^{T}). For 1ks1\leq k\leq s,

ktkH(x(t))=i=1kk1++ki=kkj1DiH(x(t))[dk1x(t)dtk1,,dkix(t)dtki].\displaystyle\frac{\partial^{k}}{\partial t^{k}}H(x(t))=\sum_{i=1}^{k}\sum_{\begin{subarray}{c}k_{1}+\dots+k_{i}=k\\ k_{j}\geq 1\end{subarray}}D^{i}H(x(t))\left[\frac{d^{k_{1}}x(t)}{dt^{k_{1}}},\dots,\frac{d^{k_{i}}x(t)}{dt^{k_{i}}}\right].

Let i,,α=((J)i,C,α(Σ;𝐕))\mathcal{B}_{i,\ell,\alpha}=\mathcal{L}(\left(\mathbb{R}^{J}\right)^{\otimes i},C^{\ell,\alpha}(\Sigma;\mathbf{V})) be the Banach space of bounded linear maps from (J)i\left(\mathbb{R}^{J}\right)^{\otimes i} to C,α(Σ;𝐕)C^{\ell,\alpha}(\Sigma;\mathbf{V}) equipped with the operator norm. From Proposition 2.1, DiHD^{i}H is an analytic map from Bρ(0)JB_{\rho}(0)\subset\mathbb{R}^{J} to i,,α\mathcal{B}_{i,\ell,\alpha}. In particular, the operator norm of DiHD^{i}H is bounded in B21ρ(0)B_{2^{-1}\rho}(0). Therefore, provided |x(t)|21ρ|x(t)|\leq 2^{-1}\rho,

|∇̸[DiH(x(t))[dk1x(t)dtk1,,dkix(t)dtki]]|C|dk1x(t)dtk1||dkix(t)dtki|CuCs(t).\displaystyle\left|\not{\nabla}^{\ell}\left[D^{i}H(x(t))\left[\frac{d^{k_{1}}x(t)}{dt^{k_{1}}},\dots,\frac{d^{k_{i}}x(t)}{dt^{k_{i}}}\right]\right]\right|\leq C\left|\frac{d^{k_{1}}x(t)}{dt^{k_{1}}}\right|\dots\left|\frac{d^{k_{i}}x(t)}{dt^{k_{i}}}\right|\leq C\|u\|_{C^{s}}(t).

We used the assumption uCs(t)M\|u\|_{C^{s}}(t)\leq M in the second inequality. This ensures H(uT)Cs(t)CuCs(t).\|H(u^{T})\|_{C^{s}}(t)\leq C\|u\|_{C^{s}}(t). Then because of (2.11), u~Cs(t)CuCs(t)\|\tilde{u}^{\perp}\|_{C^{s}}(t)\leq C\|u\|_{C^{s}}(t) holds. ∎

3. First order ODE system

In this section, we transform (1.1) into a first order ODE system. Let uC2(Q0,,𝐕~)u\in C^{2}(Q_{0,\infty},\widetilde{\mathbf{V}}) be a solution to (1.1). By setting E1(u)=N1(u)Σu+Σu{E}_{1}(u)=N_{1}(u)-\mathcal{M}_{\Sigma}u+\mathcal{L}_{\Sigma}u, we may rewirte (1.1) as

(3.1) u′′mu+Σu=E1(u).\displaystyle u^{\prime\prime}-mu^{\prime}+\mathcal{L}_{\Sigma}u=E_{1}(u).

From (2.4) and (2.6), the error term E1(u)E_{1}(u) has the structure

(3.2) E1(u)=a1Du(t)+a2u(t)+j=02a4,j∇̸ju,{E}_{1}(u)=a_{1}\cdot Du^{\prime}(t)+a_{2}\cdot u^{\prime}(t)+\sum_{j=0}^{2}a_{4,j}\cdot\not{\nabla}^{j}u,

where a4,j=a4,j(ω,u,∇̸u,u)a_{4,j}=a_{4,j}(\omega,u,\not{\nabla}u,u^{\prime}) are smooth with a4,j(ω,0,0,0)=0a_{4,j}(\omega,0,0,0)=0. We aim to vectorize (3.1) and view it as a first order ODE.

Let us begin to set up some notion.

Definition 3.1.

Let 𝐋\mathbf{L} be an operator from H2(Σ;𝐕)×H1(Σ;𝐕)H^{2}(\Sigma;\mathbf{V})\times H^{1}(\Sigma;\mathbf{V}) to H1(Σ;𝐕)×L2(Σ;𝐕)H^{1}(\Sigma;\mathbf{V})\times L^{2}(\Sigma;\mathbf{V}) given by

𝐋(v,w):=(21mv+w,Σv+41m2v+21mw).\displaystyle\mathbf{L}(v,w):=\left(2^{-1}mv+w,-\mathcal{L}_{\Sigma}v+4^{-1}m^{2}v+2^{-1}mw\right).
Definition 3.2.

For a section uC2(Q0,,𝐕~)u\in C^{2}(Q_{0,\infty},\widetilde{\mathbf{V}}), define

q(u):=(u,u21mu),(u):=(0,E1(u)).\displaystyle q(u):=(u,u^{\prime}-2^{-1}mu),\ \mathcal{E}(u):=(0,E_{1}(u)).

If there is no confusion, we often omit the argument and write qq, \mathcal{E} to denote q(u)q(u) and (u)\mathcal{E}(u), respectively. Moreover, q(t)q(t) and (t)\mathcal{E}(t) denote the restriction of q(u)q(u) and (u)\mathcal{E}(u) on {t}×Σ\{t\}\times\Sigma, respectively.

The equation (3.1) can be rewritten as

(3.3) q=𝐋q+.\displaystyle q^{\prime}=\mathbf{L}q+\mathcal{E}.

We now identify the eigenvalues and eigensections of 𝐋\mathbf{L}. For ii\in\mathbb{N}, let

γi±:=21m±41m2λi.\gamma^{\pm}_{i}:=2^{-1}m\pm\sqrt{4^{-1}m^{2}-\lambda_{i}}.

Recall that \mathbb{N} is divided into i=14Ii\cup_{i=1}^{4}I_{i} in (2.7). The eigenvalues γi±\gamma^{\pm}_{i} are real and different from 21m2^{-1}m if and only if iI3I4i\in I_{3}\cup I_{4}. For iI3I4i\in I_{3}\cup I_{4}, let

(3.4) ψi±:=(mm2γi±φi,m2φi).\psi_{i}^{\pm}:=\bigg{(}\frac{m}{m-2\gamma^{\pm}_{i}}\varphi_{i},-\frac{m}{2}\varphi_{i}\bigg{)}.

From Σφ=λiφi\mathcal{L}_{\Sigma}\varphi=\lambda_{i}\varphi_{i}, one can check that

(3.5) 𝐋ψi±=γi±ψi±.\mathbf{L}\psi^{\pm}_{i}=\gamma^{\pm}_{i}\psi^{\pm}_{i}.

For iI1i\in I_{1}, γi±\gamma^{\pm}_{i} are not real with the imaginary part βi=λi41m2>0\beta_{i}=\sqrt{\lambda_{i}-4^{-1}m^{2}}>0. Set

ψi,1:=(0,21/2mφi),ψi,2:=(21/2mβi1φi,0).\displaystyle\psi_{i,1}:=(0,2^{-1/2}m\varphi_{i}),\ \psi_{i,2}:=(2^{-1/2}m\beta_{i}^{-1}\varphi_{i},0).

Then

(3.6) 𝐋ψi,1=21mψi,1+βiψi,2,𝐋ψi,2=21mψi,2βiψi,1.\mathbf{L}\psi_{i,1}=2^{-1}m\psi_{i,1}+\beta_{i}\psi_{i,2},\ \mathbf{L}\psi_{i,2}=2^{-1}m\psi_{i,2}-\beta_{i}\psi_{i,1}.

For iI2i\in I_{2}, let

ψi,3:=(0,21/2mφi),ψi,4:=(21/2mφi,0).\displaystyle\psi_{i,3}:=(0,2^{-1/2}m\varphi_{i}),\ \psi_{i,4}:=(2^{-1/2}m\varphi_{i},0).

Then

(3.7) 𝐋ψi,3=21mψi,3+ψi,4,𝐋ψi,4=21mψi,4.\mathbf{L}\psi_{i,3}=2^{-1}m\psi_{i,3}+\psi_{i,4},\ \mathbf{L}\psi_{i,4}=2^{-1}m\psi_{i,4}.

Next, we introduce a bilinear form GG. Let G:(H1(Σ;𝐕)×H0(Σ;𝐕))2G:(H^{1}(\Sigma;\mathbf{V})\times H^{0}(\Sigma;\mathbf{V}))^{2}\to\mathbb{R} be defined by

G((v1,w1);(v2,w2)):=\displaystyle G((v_{1},w_{1});(v_{2},w_{2})):= 2m2Σ(Σv141m2v1,v2+w1,w2)𝑑μ\displaystyle 2m^{-2}\int_{\Sigma}\big{(}-\left\langle\mathcal{L}_{\Sigma}v_{1}-4^{-1}m^{2}v_{1},v_{2}\right\rangle+\left\langle w_{1},w_{2}\right\rangle\big{)}\,d\mu
+2m2iI12βi2Σv1,φi𝑑μΣv2,φi𝑑μ+2m2iI2Σv1,φi𝑑μΣv2,φi𝑑μ.\displaystyle+2m^{-2}\sum_{i\in I_{1}}2\beta^{2}_{i}\int_{\Sigma}\left\langle v_{1},\varphi_{i}\right\rangle d\mu\int_{\Sigma}\left\langle v_{2},\varphi_{i}\right\rangle d\mu+2m^{-2}\sum_{i\in I_{2}}\int_{\Sigma}\left\langle v_{1},\varphi_{i}\right\rangle d\mu\int_{\Sigma}\left\langle v_{2},\varphi_{i}\right\rangle d\mu.

We denote (v,w)G2:=G((v,w);(v,w))\|(v,w)\|^{2}_{G}:=G((v,w);(v,w)). The positive-definiteness of GG will be justified in Lemma 3.3 below. We write 𝐋\mathbf{L}^{\dagger} for the adjoint operator of 𝐋\mathbf{L} with respect to GG. Namely,

G(𝐋(v1,w1);(v2,w2))=G((v1,w1);𝐋(v2,w2)).\displaystyle G(\mathbf{L}(v_{1},w_{1});(v_{2},w_{2}))=G((v_{1},w_{1});\mathbf{L}^{\dagger}(v_{2},w_{2})).

We define the collection of vectors

(3.8) :={ψi,1,ψi,2}iI1{ψi,3,ψi,4}iI2{ψi+,ψi}iI3I4.\mathcal{B}:=\{\psi_{i,1},\psi_{i,2}\}_{i\in I_{1}}\cup\{\psi_{i,3},\psi_{i,4}\}_{i\in I_{2}}\cup\{\psi^{+}_{i},\psi^{-}_{i}\}_{i\in I_{3}\cup I_{4}}.
Lemma 3.3 (spectral decomposition of 𝐋\mathbf{L}).
  1. (1)

    The bilinear form GG is equivalent to the standard inner product. Namely,

    (3.9) (v,w)G(v,w)H1×H0.\|(v,w)\|_{G}\approx\|(v,w)\|_{H^{1}\times H^{0}}.
  2. (2)

    The collection \mathcal{B} forms a complete GG-orthonormal basis.

  3. (3)

    For iI3I4i\in I_{3}\cup I_{4},

    (3.10) 𝐋ψi±=γi±ψi±.\displaystyle\mathbf{L}^{\dagger}\psi^{\pm}_{i}=\gamma^{\pm}_{i}\psi^{\pm}_{i}.

    For iI1i\in I_{1},

    (3.11) 𝐋ψi,1=21mψi,1βiψi,2,𝐋ψi,2=21mψi,2+βiψi,1.\displaystyle\mathbf{L}^{\dagger}\psi_{i,1}=2^{-1}m\psi_{i,1}-\beta_{i}\psi_{i,2},\ \mathbf{L}^{\dagger}\psi_{i,2}=2^{-1}m\psi_{i,2}+\beta_{i}\psi_{i,1}.

    For iI2i\in I_{2},

    (3.12) 𝐋ψi,3=21mψi,3,𝐋ψi,4=21mψi,4+ψi,3.\displaystyle\mathbf{L}^{\dagger}\psi_{i,3}=2^{-1}m\psi_{i,3},\ \mathbf{L}^{\dagger}\psi_{i,4}=2^{-1}m\psi_{i,4}+\psi_{i,3}.
Proof.

We start to prove (3.9). By expressing vv as v=i=1aiφiv=\sum_{i=1}^{\infty}a_{i}\varphi_{i},

G((v,0);(v,0))=2m2(iI2|41mλi|ai2+iI2ai2)ci=1ai2=cvL22,\displaystyle G((v,0);(v,0))=2m^{-2}\left(\sum_{i\notin I_{2}}\left|4^{-1}m-\lambda_{i}\right|a_{i}^{2}+\sum_{i\in I_{2}}a_{i}^{2}\right)\geq c\sum_{i=1}^{\infty}a_{i}^{2}=c\|v\|^{2}_{L^{2}},

for some c=c(Σ,m)>0c=c(\mathcal{M}_{\Sigma},m)>0. In view of (2.5), this implies (3.9). It is straightforward to check that vectors in \mathcal{B} are GG-orthonormal. Suppose (v,w)H1(Σ;𝐕)×H0(Σ;𝐕)(v,w)\in H^{1}(\Sigma;\mathbf{V})\times H^{0}(\Sigma;\mathbf{V}) is GG-orthogonal to every vector in \mathcal{B}. Note that for all jj\in\mathbb{N}, (0,φj)(0,\varphi_{j}) and (φj,0)(\varphi_{j},0) lie in the linear span of \mathcal{B}. This implies that

0=21m2G((v,w);(0,φj))=Σw,φj𝑑μ,\displaystyle 0=2^{-1}m^{2}G((v,w);(0,\varphi_{j}))=\int_{\Sigma}\left\langle w,\varphi_{j}\right\rangle\,d\mu,

and that

0=21m2G((v,w);(φj,0))=CjΣv,φj𝑑μ.\displaystyle 0=2^{-1}m^{2}G((v,w);(\varphi_{j},0))=C_{j}\int_{\Sigma}\left\langle v,\varphi_{j}\right\rangle\,d\mu.

Here Cj=1C_{j}=1 for jI2j\in I_{2} and Cj=|41mλj|C_{j}=\left|4^{-1}m-\lambda_{j}\right| for jI2j\notin I_{2}. Therefore, v=w=0v=w=0 and \mathcal{B} is complete. Lastly, (3.10)-(3.12) follow from (3.5)-(3.7) and the GG-orthogonality of \mathcal{B}. ∎

The next lemma shows that 𝐋\mathbf{L} behaves like ∇̸\not{\nabla}.

Lemma 3.4 (equivalence of 𝐋\mathbf{L} and angular derivative).

For each 0\ell\in\mathbb{N}_{0},

(3.13) j=0𝐋j(v,w)G(v,w)H+1×H.\sum_{j=0}^{\ell}\|\mathbf{L}^{j}(v,w)\|_{G}\approx\|(v,w)\|_{H^{\ell+1}\times H^{\ell}}.
Proof.

We use an induction argument. The assertion for =0\ell=0 follows directly from (3.9). Now we assume (3.13) holds for \ell and prove it for +1\ell+1. From the induction hypothesis,

(3.14) j=0+1𝐋j(v,w)G(v,w)H1×H0+𝐋(v,w)H×H1\displaystyle\sum_{j=0}^{\ell+1}\|\mathbf{L}^{j}(v,w)\|_{G}\approx\|(v,w)\|_{H^{1}\times H^{0}}+\|\mathbf{L}(v,w)\|_{H^{\ell}\times H^{\ell-1}}

In view of Definition 3.1, 𝐋(v,w)H×H1C(v,w)H+1×H\|\mathbf{L}(v,w)\|_{H^{\ell}\times H^{\ell-1}}\leq C\|(v,w)\|_{H^{\ell+1}\times H^{\ell}}. Therefore,

(3.15) j=0+1𝐋j(v,w)GC(v,w)H+1×H.\sum_{j=0}^{\ell+1}\|\mathbf{L}^{j}(v,w)\|_{G}\leq C\|(v,w)\|_{H^{\ell+1}\times H^{\ell}}.

To obtain the inequality in the other direction, we use

wH\displaystyle\|w\|_{H^{\ell}} C21mv+wH+CvH,\displaystyle\leq C\|2^{-1}mv+w\|_{H^{\ell}}+C\|v\|_{H^{\ell}},
vH+1\displaystyle\|v\|_{H^{\ell+1}} CΣv+41m2v+21mwH1+C(v,w)H×H1.\displaystyle\leq C\|-\mathcal{L}_{\Sigma}v+4^{-1}m^{2}v+2^{-1}mw\|_{H^{\ell-1}}+C\|(v,w)\|_{H^{\ell}\times H^{\ell-1}}.

Combining Definition 3.1, (3.14) and the induction hypothesis, we obtain

(3.16) (v,w)H+1×HCj=0+1𝐋j(v,w)G.\|(v,w)\|_{H^{\ell+1}\times H^{\ell}}\leq C\sum_{j=0}^{\ell+1}\|\mathbf{L}^{j}(v,w)\|_{G}.

The assertion then follows from (3.15) and (3.16). ∎

Definition 3.5.

Fix k,0k,\ell\in\mathbb{N}_{0}. For a section uCk++1(Q0,,𝐕~)u\in C^{k+\ell+1}(Q_{0,\infty},\widetilde{\mathbf{V}}), we define

q(k,)(u):=tk𝐋q(u),(k,)(u):=tk𝐋(u).q^{(k,\ell)}(u):={\partial_{t}^{k}}\mathbf{L}^{\ell}q(u),\ \mathcal{E}^{(k,\ell)}(u):={\partial_{t}^{k}}\mathbf{L}^{\ell}\mathcal{E}(u).

Here q(u)q(u) and (u)\mathcal{E}(u) are given in Definition 3.2. We often abbreviate them to q(k,)q^{(k,\ell)}, (k,)\mathcal{E}^{(k,\ell)} and denote by q(k,)(t)q^{(k,\ell)}(t), (k,)(t)\mathcal{E}^{(k,\ell)}(t) the restriction of q(k,)(u)q^{(k,\ell)}(u), (k,)(u)\mathcal{E}^{(k,\ell)}(u) on {t}×Σ\{t\}\times\Sigma, respectively.

Corollary 3.6.

Fix s0s\in\mathbb{N}_{0}. Then for all uCs+2(Q0,,𝐕~)u\in C^{s+2}(Q_{0,\infty},\widetilde{\mathbf{V}}) and t(0,)t\in(0,\infty),

(3.17) k+sq(k,)(u)G(t)k+s+1tk∇̸uL2(t).\sum_{k+\ell\leq s}\|q^{(k,\ell)}(u)\|_{G}(t)\approx\sum_{k+\ell\leq s+1}\|\partial_{t}^{k}\not{\nabla}^{\ell}u\|_{L^{2}}(t).

Moreover, suppose uu satisfies uCs+2(t)=o(1)\|u\|_{C^{s+2}}(t)=o(1). Then

(3.18) k+s(k,)(u)G(t)=o(1)k+sq(k,)(u)G(t)\sum_{k+\ell\leq s}\|\mathcal{E}^{(k,\ell)}(u)\|_{G}(t)=o(1)\sum_{k+\ell\leq s}\|q^{(k,\ell)}(u)\|_{G}(t)
Proof.

From Definition 3.2 and Lemma 3.4 (equivalence of 𝐋\mathbf{L} and angular derivative),

k+sq(k,)(u)G(t)\displaystyle\sum_{k+\ell\leq s}\|q^{(k,\ell)}(u)\|_{G}(t) k=0s(tku,tk+1u21mtku)Hsk+1×Hsk(t)\displaystyle\approx\sum_{k=0}^{s}\|(\partial_{t}^{k}u,\partial_{t}^{k+1}u-2^{-1}m\partial_{t}^{k}u)\|_{H^{s-k+1}\times H^{s-k}}(t)
k+s+1tk∇̸uL2(t).\displaystyle\approx\sum_{k+\ell\leq s+1}\|\partial_{t}^{k}\not{\nabla}^{\ell}u\|_{L^{2}}(t).

This gives (3.17). Similarly, from Lemma 3.4 and Definition 3.2,

k+m(k,)(u)G(t)\displaystyle\sum_{k+\ell\leq m}\|\mathcal{E}^{(k,\ell)}(u)\|_{G}(t) k=0stkE1(u)Hsk(t).\displaystyle\approx\sum_{k=0}^{s}\|\partial_{t}^{k}E_{1}(u)\|_{H^{s-k}}(t).

From (3.2) and the assumption,

k=0stkE1(u)Hsk(t)=o(1)k+s+1tk∇̸uL2(t).\displaystyle\sum_{k=0}^{s}\|\partial_{t}^{k}E_{1}(u)\|_{H^{s-k}}(t)=o(1)\sum_{k+\ell\leq s+1}\|\partial_{t}^{k}\not{\nabla}^{\ell}u\|_{L^{2}}(t).

Then (3.18) follows from (3.17). ∎

Let us project the equation (3.3) onto vectors in \mathcal{B}. Let

(3.19) ξi,1(t):=G(q(t),ψi,1),ξi,2(t):=G(q(t),ψi,2)foriI1,ξi,3(t):=G(q(t),ψi,3),ξi,4(t):=G(q(t),ψi,4)foriI2,ξi±(t):=G(q(t),ψi±)foriI3I4.\begin{split}&\xi_{i,1}(t):=G(q(t),\psi_{i,1}),\ \xi_{i,2}(t):=G(q(t),\psi_{i,2})\ \textup{for}\ i\in I_{1},\\ &\xi_{i,3}(t):=G(q(t),\psi_{i,3}),\ \xi_{i,4}(t):=G(q(t),\psi_{i,4})\ \textup{for}\ i\in I_{2},\\ &\xi^{\pm}_{i}(t):=G(q(t),\psi^{\pm}_{i})\ \textup{for}\ i\in I_{3}\cup I_{4}.\end{split}

Also, let

(3.20) i,1(t):=G((t),ψi,1),i,2(t):=G((t),ψi,2)foriI1,i,3(t):=G((t),ψi,3),i,4(t):=G((t),ψi,4)foriI2,i±(t):=G((t),ψi±)foriI3I4.\begin{split}&\mathcal{E}_{i,1}(t):=G(\mathcal{E}(t),\psi_{i,1}),\ \mathcal{E}_{i,2}(t):=G(\mathcal{E}(t),\psi_{i,2})\ \textup{for}\ i\in I_{1},\\ &\mathcal{E}_{i,3}(t):=G(\mathcal{E}(t),\psi_{i,3}),\ \mathcal{E}_{i,4}(t):=G(\mathcal{E}(t),\psi_{i,4})\ \textup{for}\ i\in I_{2},\\ &\mathcal{E}^{\pm}_{i}(t):=G(\mathcal{E}(t),\psi^{\pm}_{i})\ \textup{for}\ i\in I_{3}\cup I_{4}.\end{split}

We can then rewrite (3.3) as follows. For iI1i\in I_{1},

(3.21) ddtξi,121mξi,1+βiξi,2=i,1,ddtξi,221mξi,2βiξi,1=i,2.\begin{split}&\frac{d}{dt}\xi_{i,1}-2^{-1}m\xi_{i,1}+\beta_{i}\xi_{i,2}=\mathcal{E}_{i,1},\\ &\frac{d}{dt}\xi_{i,2}-2^{-1}m\xi_{i,2}-\beta_{i}\xi_{i,1}=\mathcal{E}_{i,2}.\end{split}

For iI2i\in I_{2},

(3.22) ddtξi,321mξi,3=i,3,ddtξi,421mξi,4ξi,3=i,4.\begin{split}&\frac{d}{dt}\xi_{i,3}-2^{-1}m\xi_{i,3}=\mathcal{E}_{i,3},\\ &\frac{d}{dt}\xi_{i,4}-2^{-1}m\xi_{i,4}-\xi_{i,3}=\mathcal{E}_{i,4}.\end{split}

For iI3I4i\in I_{3}\cup I_{4},

(3.23) ddtξi±γi±ξi=i±.\frac{d}{dt}\xi^{\pm}_{i}-\gamma^{\pm}_{i}\xi_{i}=\mathcal{E}^{\pm}_{i}.

4. Fast decaying solutions

In this section, we consider solutions to (1.1) that decay exponentially and prove Theorem 2.4. The proof for Theorem 2.6 is simpler so we will omit it. We assume throughout this section that m<0m<0 and I1,I2I_{1},I_{2} are non-empty. The proof can be generalized easily to other cases. We begin with a unique continuation property at infinity. Though we closely follow the argument in [Str20], we include the proof for readers’ convenience.

Proposition 4.1.

Let uC(Q0,,𝐕~)u\in C^{\infty}(Q_{0,\infty},\widetilde{\mathbf{V}}) be a solution to (1.1) that satisfies uC1(t)=O(eγt)\|u\|_{C^{1}}(t)=O(e^{\gamma t}) as tt\to\infty for all γ<0\gamma<0. Then u0u\equiv 0.

Proof.

By the elliptic regularity, Lemma A.3, uCs(t)=O(eγt)\|u\|_{C^{s}}(t)=O(e^{\gamma t}) for all ss\in\mathbb{N} and γ<0\gamma<0. Let q=q(u)q=q(u) be given in Definition 3.2. Take γ<21m1\gamma<2^{-1}m-1. Let Πγq\Pi_{\gamma}q be the projection of qq onto the eigenspace of 𝐋\mathbf{L} whose eigenvalues are less than or equal to γ\gamma. Namely, in terms of the coefficients introduced in (3.19),

(Πγq)(t)=i:γiγξi(t)ψi.\displaystyle(\Pi_{\gamma}q)(t)=\sum_{i:\gamma^{-}_{i}\leq\gamma}\xi_{i}^{-}(t)\psi_{i}^{-}.

We claim that for any t2t1t_{2}\geq t_{1},

(4.1) eγ(t2t1)q(t2)GΠγq(t1)G+t1eγ(st1)(s)G𝑑s.\displaystyle e^{-\gamma(t_{2}-t_{1})}\|q(t_{2})\|_{G}\leq\|\Pi_{\gamma}q(t_{1})\|_{G}+\int_{t_{1}}^{\infty}e^{-\gamma(s-t_{1})}\|\mathcal{E}(s)\|_{G}\,ds.

Suppose (4.1) is true at the moment. From (3.2), there exists a uniform constant C<C<\infty such that if uC1(t)1\|u\|_{C^{1}}(t)\leq 1, then there holds

(t)GCq(t)GuC2(t).\displaystyle\|\mathcal{E}(t)\|_{G}\leq C\|q(t)\|_{G}\|u\|_{C^{2}}(t).

Define Mγ(t)=supτteγ(τt)q(τ)GM_{\gamma}(t)=\sup_{\tau\geq t}e^{-\gamma(\tau-t)}\|q(\tau)\|_{G}. Suppose t1t_{1} is large enough such that uC1(s)1\|u\|_{C^{1}}(s)\leq 1 for st1s\geq t_{1}. Then from (4.1),

Mγ(t1)Πγq(t1)G+Ct1uC2(s)𝑑sMγ(t1).\displaystyle M_{\gamma}(t_{1})\leq\|\Pi_{\gamma}q(t_{1})\|_{G}+C\int_{t_{1}}^{\infty}\|u\|_{C^{2}}(s)\,ds\cdot M_{\gamma}(t_{1}).

From the exponential decay assumption we may choose large t1t_{1} so that Ct1uC2(s)𝑑s1/2C\int_{t_{1}}^{\infty}\|u\|_{C^{2}}(s)ds\leq 1/2. Hence

(4.2) Mγ(t1)2Πγq(t1)G.\displaystyle M_{\gamma}(t_{1})\leq 2\|\Pi_{\gamma}q(t_{1})\|_{G}.

It is clear that Mγ(t1)M_{\gamma}(t_{1}) is non-increasing in γ\gamma. Therefore,

Mγ(t1)lim supγMγ(t1)2lim supγΠγq(t1)G=0.\displaystyle M_{\gamma}(t_{1})\leq\limsup_{\gamma^{\prime}\to-\infty}M_{\gamma^{\prime}}(t_{1})\leq 2\limsup_{\gamma^{\prime}\to-\infty}\|\Pi_{\gamma^{\prime}}q(t_{1})\|_{G}=0.

We conclude q(t)=0q(t)=0 (and thus u(t)=0u(t)=0) for tt1t\geq t_{1}. Next, we show q(t)=0q(t)=0 upto t=0t=0. Suppose on the contrary t0=inf{t1:q(t)=0 for tt1}>0t_{0}=\inf\{t_{1}\,:\,q(t)=0\text{ for }t\geq t_{1}\}>0. By the smoothness of u(x,t)u(x,t), we may find a small ϵ>0\epsilon>0 such that uC1(t)1\|u\|_{C^{1}}(t)\leq 1 for t[t0ϵ,t0]t\in[t_{0}-\epsilon,t_{0}] and

Ct0ϵuC2(s)𝑑s=Ct0ϵt0uC2(s)𝑑s12,C\int_{t_{0}-\epsilon}^{\infty}\|u\|_{C^{2}}(s)\,ds=C\int_{t_{0}-\epsilon}^{t_{0}}\|u\|_{C^{2}}(s)\,ds\leq\frac{1}{2},

which implies that (4.2) holds for t1=t0ϵt_{1}=t_{0}-\epsilon. This gives a contradiction and proves the statement.


It remains to show (4.1). Define non-negative functions X±(t)X_{\pm}(t) by

X+2(t)=\displaystyle X^{2}_{+}(t)= iI1|ξi,1(t)|2+|ξi,2(t)|2+iI2|ξi,3(t)|2+|ξi,4(t)|2+|ξi+(t)|2+i:γi>γ|ξi(t)|2,\displaystyle\sum_{i\in I_{1}}|\xi_{i,1}(t)|^{2}+|\xi_{i,2}(t)|^{2}+\sum_{i\in I_{2}}|\xi_{i,3}(t)|^{2}+|\xi_{i,4}(t)|^{2}+\sum|\xi^{+}_{i}(t)|^{2}+\sum_{i:\gamma^{-}_{i}>\gamma}|\xi^{-}_{i}(t)|^{2},
X2(t)=\displaystyle X^{2}_{-}(t)= i:γiγ|ξi(t)|2.\displaystyle\sum_{i:\gamma^{-}_{i}\leq\gamma}|\xi^{-}_{i}(t)|^{2}.

From (3.21)-(3.23), X+(t)X+(t)X_{+}(t)X_{+}^{\prime}(t) equals,

21m(iI1|ξi,1(t)|2+|ξi,2(t)|2+iI2|ξi,3(t)|2+|ξi,4(t)|2)+γi+|ξi+(t)|2+i:γi>γγi|ξi(t)|2\displaystyle 2^{-1}m\left(\sum_{i\in I_{1}}|\xi_{i,1}(t)|^{2}+|\xi_{i,2}(t)|^{2}+\sum_{i\in I_{2}}|\xi_{i,3}(t)|^{2}+|\xi_{i,4}(t)|^{2}\right)+\sum\gamma^{+}_{i}|\xi^{+}_{i}(t)|^{2}+\sum_{i:\gamma^{-}_{i}>\gamma}\gamma^{-}_{i}|\xi^{-}_{i}(t)|^{2}
+iI2ξi,3(t)ξi,4(t)+ξi+(t)i+(t)+i:γi>γξ+(t)+(t)+iI1ξi,1(t)i,1(t)+ξi,1(t)i,1(t)\displaystyle+\sum_{i\in I_{2}}\xi_{i,3}(t)\xi_{i,4}(t)+\sum\xi^{+}_{i}(t)\mathcal{E}^{+}_{i}(t)+\sum_{i:\gamma^{-}_{i}>\gamma}\xi^{+}_{-}(t)\mathcal{E}^{+}_{-}(t)+\sum_{i\in I_{1}}\xi_{i,1}(t)\mathcal{E}_{i,1}(t)+\xi_{i,1}(t)\mathcal{E}_{i,1}(t)
+iI2ξi,3(t)i,3(t)+ξi,4(t)i,4(t).\displaystyle+\sum_{i\in I_{2}}\xi_{i,3}(t)\mathcal{E}_{i,3}(t)+\xi_{i,4}(t)\mathcal{E}_{i,4}(t).

From |iI2ξi,3(t)ξi,4(t)|21iI2|ξi,3(t)|2+|ξi,4(t)|2\left|\sum_{i\in I_{2}}\xi_{i,3}(t)\xi_{i,4}(t)\right|\leq 2^{-1}\sum_{i\in I_{2}}|\xi_{i,3}(t)|^{2}+|\xi_{i,4}(t)|^{2} and γ<21m1\gamma<2^{-1}m-1,

(4.3) X+(t)γX+(t)+Y+(t).\displaystyle X_{+}^{\prime}(t)\geq\gamma X_{+}^{\prime}(t)+Y_{+}(t).

Here Y+(t)Y_{+}(t) is given by

X+(t)Y+(t)=\displaystyle X_{+}(t)Y_{+}(t)= ξi+(t)i+(t)+i:γi>γξ+(t)+(t)+iI1ξi,1(t)i,1(t)+ξi,1(t)i,1(t)\displaystyle\sum\xi^{+}_{i}(t)\mathcal{E}^{+}_{i}(t)+\sum_{i:\gamma^{-}_{i}>\gamma}\xi^{+}_{-}(t)\mathcal{E}^{+}_{-}(t)+\sum_{i\in I_{1}}\xi_{i,1}(t)\mathcal{E}_{i,1}(t)+\xi_{i,1}(t)\mathcal{E}_{i,1}(t)
+iI2ξi,3(t)i,3(t)+ξi,4(t)i,4(t).\displaystyle+\sum_{i\in I_{2}}\xi_{i,3}(t)\mathcal{E}_{i,3}(t)+\xi_{i,4}(t)\mathcal{E}_{i,4}(t).

Similarly, we have

(4.4) X(t)γX(t)+Y(t),\displaystyle X_{-}^{\prime}(t)\leq\gamma X_{-}^{\prime}(t)+Y_{-}(t),

where Y(t)Y_{-}(t) is given by X(t)Y(t)=i:γiγξi(t)i(t)X_{-}(t)Y_{-}(t)=\sum_{i:\gamma^{-}_{i}\leq\gamma}\xi^{-}_{i}(t)\mathcal{E}^{-}_{i}(t). Fix t2t1t_{2}\geq t_{1}. By integrating (4.4) from t1t_{1} to t2t_{2}, we obtain

(4.5) eγ(t2t1)X(t2)X(t1)+t1t2eγ(st1)Y(s)𝑑sX(t1)+t1eγ(st1)|Y(s)|𝑑s.\begin{split}e^{-\gamma(t_{2}-t_{1})}X_{-}(t_{2})\leq&X_{-}(t_{1})+\int_{t_{1}}^{t_{2}}e^{-\gamma(s-t_{1})}Y_{-}(s)\,ds\\ \leq&X_{-}(t_{1})+\int_{t_{1}}^{\infty}e^{-\gamma(s-t_{1})}|Y_{-}(s)|\,ds.\end{split}

Take any t3t2t_{3}\geq t_{2}. By integrating (4.3) from t2t_{2} to t3t_{3}, we obtain

eγ(t2t1)X+(t2)\displaystyle e^{-\gamma(t_{2}-t_{1})}X_{+}(t_{2})\leq eγ(t3t1)X+(t3)t2t3eγ(st1)Y+(s)𝑑s\displaystyle e^{-\gamma(t_{3}-t_{1})}X_{+}(t_{3})-\int_{t_{2}}^{t_{3}}e^{-\gamma(s-t_{1})}Y_{+}(s)\,ds
\displaystyle\leq eγ(t3t1)X+(t3)+t1eγ(st1)|Y+(s)|𝑑s\displaystyle e^{-\gamma(t_{3}-t_{1})}X_{+}(t_{3})+\int_{t_{1}}^{\infty}e^{-\gamma(s-t_{1})}|Y_{+}(s)|\,ds

By the decay assumption, eγ(t3t1)X+(t3)e^{-\gamma(t_{3}-t_{1})}X_{+}(t_{3}) goes to zero when t3t_{3} goes to infinity. Hence

(4.6) eγ(t2t1)X+(t2)t1eγ(st1)|Y+(s)|𝑑s.\displaystyle e^{-\gamma(t_{2}-t_{1})}X_{+}(t_{2})\leq\int_{t_{1}}^{\infty}e^{-\gamma(s-t_{1})}|Y_{+}(s)|\,ds.

Note that X+2(t)+X2(t)=q(t)G2X_{+}^{2}(t)+X_{-}^{2}(t)=\|q(t)\|^{2}_{G}, X(t)=Πγq(t)GX_{-}(t)=\|\Pi_{\gamma}q(t)\|_{G} and Y+2(t)+Y2(t)(t)G2Y_{+}^{2}(t)+Y_{-}^{2}(t)\leq\|\mathcal{E}(t)\|^{2}_{G}. Then (4.1) follows from (4.5) and (4.6). ∎

Remark 4.2.

A similar argument applies for the parabolic equation. The only difference is that to control the error term E2(u)E_{2}(u) in (6.6), one needs to differentiate the equation. See Lemma 6.2.

Let uu be an exponentially decaying solution to (1.1). Namely, uC1(t)=O(e2ε0t)\|u\|_{C^{1}}(t)=O(e^{-2\varepsilon_{0}t}) for some ε0>0\varepsilon_{0}>0. We further assume uu is not identically zero. In view of Proposition 4.1, the set

Λ:={γ<0:uC1(t)=O(eγt)}\displaystyle\Lambda:=\{\gamma<0\,:\,\|u\|_{C^{1}}(t)=O(e^{\gamma t})\}

has an infimum <γ<0-\infty<\gamma_{*}<0. From the elliptic regularity, Lemma A.3, uCs(t)=O(e(γ+ε)t)\|u\|_{C^{s}}(t)=O(e^{(\gamma_{*}+\varepsilon)t}) for all ss\in\mathbb{N} and ε>0\varepsilon>0. Let {ξi,1(t),ξi,2(t)}iI1\{\xi_{i,1}(t),\xi_{i,2}(t)\}_{i\in I_{1}}, {ξi,3(t),ξi,4(t)}iI2\{\xi_{i,3}(t),\xi_{i,4}(t)\}_{i\in I_{2}}, {ξi±(t)}iI3I4\{\xi^{\pm}_{i}(t)\}_{i\in I_{3}\cup I_{4}} be the coefficients defined in (3.19). Let {i,1(t),i,2(t)}iI1\{\mathcal{E}_{i,1}(t),\mathcal{E}_{i,2}(t)\}_{i\in I_{1}}, {i,3(t),i,4(t)}iI2\{\mathcal{E}_{i,3}(t),\mathcal{E}_{i,4}(t)\}_{i\in I_{2}}, {i±(t)}iI3I4\{\mathcal{E}^{\pm}_{i}(t)\}_{i\in I_{3}\cup I_{4}} be given by (3.20). The quadratic nature of E1(u)E_{1}(u) (see (3.2)), in particular, implies

(4.7) (0,E1(u))G2(t)=iI1(i,1(t))2+(i,2(t))2+iI2(i,3(t))2+(i,4(t))2+iI3I4(i+(t))2+(i(t))2=O(e2(γε0)t).\begin{split}\|(0,E_{1}(u))\|^{2}_{G}(t)=\sum_{i\in I_{1}}(\mathcal{E}_{i,1}(t))^{2}+(\mathcal{E}_{i,2}(t))^{2}+\sum_{i\in I_{2}}(\mathcal{E}_{i,3}(t))^{2}+(\mathcal{E}_{i,4}(t))^{2}+\sum_{i\in I_{3}\cup I_{4}}(\mathcal{E}_{i}^{+}(t))^{2}+(\mathcal{E}_{i}^{-}(t))^{2}\\ =O(e^{2(\gamma_{*}-\varepsilon_{0})t}).\end{split}

Here we used |E1(u)|Ce2ε0te(γ+ε0)t=Ce(γε0)t|E_{1}(u)|\leq Ce^{-2\varepsilon_{0}t}e^{(\gamma^{*}+\varepsilon_{0})t}=Ce^{(\gamma^{*}-\varepsilon_{0})t}.

Lemma 4.3.

There holds γ{γi+,γi}iI3I4{21m}\gamma_{*}\in\{\gamma^{+}_{i},\gamma^{-}_{i}\}_{i\in I_{3}\cup I_{4}}\cup\{2^{-1}m\}.

Proof.

Suppose the assertion fails. This implies there exists ε1(0,ε0)\varepsilon_{1}\in(0,\varepsilon_{0}) such that there is no element of {γi+,γi}iI3I4{21m}\{\gamma^{+}_{i},\gamma^{-}_{i}\}_{i\in I_{3}\cup I_{4}}\cup\{2^{-1}m\} in the interval [γ2ε1,γ+2ε1][\gamma_{*}-2\varepsilon_{1},\gamma_{*}+2\varepsilon_{1}]. We show this leads to a contradiction for the case 21m<γ2^{-1}m<\gamma_{*}. The argument for the case γ<21m\gamma_{*}<2^{-1}m is similar. Define an non-negative function X+(t)X_{+}(t) by

X+2(t)=i:γi+>γ|ξi+(t)|2+i:γi>γ|ξi(t)|2.\displaystyle X^{2}_{+}(t)=\sum_{i:\gamma^{+}_{i}>\gamma_{*}}|\xi_{i}^{+}(t)|^{2}+\sum_{i:\gamma^{-}_{i}>\gamma_{*}}|\xi_{i}^{-}(t)|^{2}.

From (3.23),

X+(t)X+(t)=\displaystyle X_{+}(t)X_{+}^{\prime}(t)= i:γi+>γγi+|ξi+(t)|2+i:γi>γγi|ξi(t)|2+i:γi+>γξi+(t)i+(t)+i:γi>γξi(t)i(t)\displaystyle\sum_{i:\gamma^{+}_{i}>\gamma_{*}}\gamma^{+}_{i}|\xi_{i}^{+}(t)|^{2}+\sum_{i:\gamma^{-}_{i}>\gamma_{*}}\gamma^{-}_{i}|\xi_{i}^{-}(t)|^{2}+\sum_{i:\gamma^{+}_{i}>\gamma_{*}}\xi_{i}^{+}(t)\mathcal{E}^{+}_{i}(t)+\sum_{i:\gamma^{-}_{i}>\gamma_{*}}\xi_{i}^{-}(t)\mathcal{E}^{-}_{i}(t)
\displaystyle\geq (γ+ε1)X+2(t)+X+(t)Y+(t).\displaystyle(\gamma_{*}+\varepsilon_{1})X^{2}_{+}(t)+X_{+}(t)Y_{+}(t).

Here Y+(t)Y_{+}(t) is given by

X+(t)Y+(t)=i:γi+>γξi+(t)i+(t)+i:γi>γξi(t)i(t).\displaystyle X_{+}(t)Y_{+}(t)=\sum_{i:\gamma^{+}_{i}>\gamma_{*}}\xi_{i}^{+}(t)\mathcal{E}^{+}_{i}(t)+\sum_{i:\gamma^{-}_{i}>\gamma_{*}}\xi_{i}^{-}(t)\mathcal{E}^{-}_{i}(t).

From the Cauchy-Schwarz inequality and (4.7), |Y+(t)|=O(e(γε0)t)|Y_{+}(t)|=O(e^{(\gamma_{*}-\varepsilon_{0})t}). Using

limte(γ+ε1)tX+(t)=0,\lim_{t\to\infty}e^{-(\gamma_{*}+\varepsilon_{1})t}X_{+}(t)=0,

we can integrate

ddt(e(γ+ε1)tX+(t))e(γ+ε1)tY+(t)\frac{d}{dt}\left(e^{-(\gamma_{*}+\varepsilon_{1})t}X_{+}(t)\right)\geq e^{-(\gamma_{*}+\varepsilon_{1})t}Y_{+}(t)

from tt to \infty to obtain

(4.8) X+(t)e(γ+ε1)tte(γ+ε1)τ|Y+(τ)|𝑑τ=O(e(γε0)t).\displaystyle X_{+}(t)\leq e^{(\gamma_{*}+\varepsilon_{1})t}\int_{t}^{\infty}e^{-(\gamma_{*}+\varepsilon_{1})\tau}|Y_{+}(\tau)|\,d\tau=O(e^{(\gamma_{*}-\varepsilon_{0})t}).

Let

X2(t)=i:γi+<γ|ξi+(t)|2+i:γi<γ|ξi(t)|2+iI1|ξi,1(t)|2+|ξi,2(t)|2+iI2|ξi,3(t)|2+ε12|ξi,4(t)|2.\displaystyle X^{2}_{-}(t)=\sum_{i:\gamma^{+}_{i}<\gamma_{*}}|\xi_{i}^{+}(t)|^{2}+\sum_{i:\gamma^{-}_{i}<\gamma_{*}}|\xi_{i}^{-}(t)|^{2}+\sum_{i\in I_{1}}|\xi_{i,1}(t)|^{2}+|\xi_{i,2}(t)|^{2}+\sum_{i\in I_{2}}|\xi_{i,3}(t)|^{2}+\varepsilon_{1}^{2}|\xi_{i,4}(t)|^{2}.

From (3.21)-(3.23), X(t)X(t)X_{-}(t)X_{-}^{\prime}(t) equals

i:γi+<γγi+|ξi+(t)|2+i:γi<γγi|ξi(t)|2+21m(iI1|ξi,1(t)|2+|ξi,2(t)|2+iI2|ξi,3(t)|2+ε12|ξi,4(t)|2)\displaystyle\sum_{i:\gamma^{+}_{i}<\gamma_{*}}\gamma^{+}_{i}|\xi_{i}^{+}(t)|^{2}+\sum_{i:\gamma^{-}_{i}<\gamma_{*}}\gamma^{-}_{i}|\xi_{i}^{-}(t)|^{2}+2^{-1}m\left(\sum_{i\in I_{1}}|\xi_{i,1}(t)|^{2}+|\xi_{i,2}(t)|^{2}+\sum_{i\in I_{2}}|\xi_{i,3}(t)|^{2}+\varepsilon_{1}^{2}|\xi_{i,4}(t)|^{2}\right)
+\displaystyle+ ε12iI2ξi,3(t)ξi,4(t)+i:γi+<γξi+(t)i+(t)+i:γi<γξi(t)i(t)+iI1ξi,1(t)i,1(t)+ξi,2(t)i,2(t)\displaystyle\varepsilon_{1}^{2}\sum_{i\in I_{2}}\xi_{i,3}(t)\xi_{i,4}(t)+\sum_{i:\gamma^{+}_{i}<\gamma_{*}}\xi_{i}^{+}(t)\mathcal{E}^{+}_{i}(t)+\sum_{i:\gamma^{-}_{i}<\gamma_{*}}\xi_{i}^{-}(t)\mathcal{E}^{-}_{i}(t)+\sum_{i\in I_{1}}\xi_{i,1}(t)\mathcal{E}_{i,1}(t)+\xi_{i,2}(t)\mathcal{E}_{i,2}(t)
+\displaystyle+ iI2ξi,3(t)i,3(t)+ε12ξi,4(t)i,4(t).\displaystyle\sum_{i\in I_{2}}\xi_{i,3}(t)\mathcal{E}_{i,3}(t)+\varepsilon_{1}^{2}\xi_{i,4}(t)\mathcal{E}_{i,4}(t).

From ε12iI2|ξi,3(t)ξi,4(t)|21ε1iI2|ξi,3(t)|2+ε12|ξi,4(t)|2\varepsilon_{1}^{2}\sum_{i\in I_{2}}|\xi_{i,3}(t)\xi_{i,4}(t)|\leq 2^{-1}\varepsilon_{1}\sum_{i\in I_{2}}|\xi_{i,3}(t)|^{2}+\varepsilon^{2}_{1}|\xi_{i,4}(t)|^{2},

X(t)X(t)(γε1)X2(t)+X(t)Y(t).\displaystyle X_{-}(t)X_{-}^{\prime}(t)\leq(\gamma_{*}-\varepsilon_{1})X^{2}_{-}(t)+X_{-}(t)Y_{-}(t).

Here Y(t)Y_{-}(t) is given by

X(t)Y(t)=\displaystyle X_{-}(t)Y_{-}(t)= i:γi+<γξi+(t)i+(t)+i:γi<γξi(t)i(t)+iI1ξi,1(t)i,1(t)+ξi,2(t)i,2(t)\displaystyle\sum_{i:\gamma^{+}_{i}<\gamma_{*}}\xi_{i}^{+}(t)\mathcal{E}^{+}_{i}(t)+\sum_{i:\gamma^{-}_{i}<\gamma_{*}}\xi_{i}^{-}(t)\mathcal{E}^{-}_{i}(t)+\sum_{i\in I_{1}}\xi_{i,1}(t)\mathcal{E}_{i,1}(t)+\xi_{i,2}(t)\mathcal{E}_{i,2}(t)
+iI2ξi,3(t)i,3(t)+ε12ξi,4(t)i,4(t).\displaystyle+\sum_{i\in I_{2}}\xi_{i,3}(t)\mathcal{E}_{i,3}(t)+\varepsilon_{1}^{2}\xi_{i,4}(t)\mathcal{E}_{i,4}(t).

From the Cauchy-Schwarz inequality and (4.7), |Y(t)|=O(e(γε0)t)|Y_{-}(t)|=O(e^{(\gamma_{*}-\varepsilon_{0})t}). Integrating

ddt(e(γε1)tX(t))e(γε1)tY(t)\frac{d}{dt}\left(e^{-(\gamma_{*}-\varepsilon_{1})t}X_{-}(t)\right)\geq e^{-(\gamma_{*}-\varepsilon_{1})t}Y_{-}(t)

from 0 to tt, we obtain

(4.9) X(t)e(γε1)t(X(0)+0e(γε1)τ|Y(τ)|𝑑τ)=O(e(γε1)t).\displaystyle X_{-}(t)\leq e^{(\gamma_{*}-\varepsilon_{1})t}\left(X_{-}(0)+\int_{0}^{\infty}e^{-(\gamma_{*}-\varepsilon_{1})\tau}|Y_{-}(\tau)|d\tau\right)=O(e^{(\gamma_{*}-\varepsilon_{1})t}).

Combining (4.8) and (4.9), q(t)G=O(e(γε1)t)\|q(t)\|_{G}=O(e^{(\gamma_{*}-\varepsilon_{1})t}). From the elliptic regularity, Lemma A.3, we then have uC1(t)=O(e(γε1)t)\|u\|_{C^{1}}(t)=O(e^{(\gamma_{*}-\varepsilon_{1})t}). This contradicts to the definition of γ\gamma_{*}. ∎

Lemma 4.4.

Suppose γ=21m\gamma_{*}=2^{-1}m. Then there exists wiw_{i}\in\mathbb{C} for iI1i\in I_{1} and ci,3,ci,4c_{i,3},c_{i,4}\in\mathbb{R} for iI2i\in I_{2} such that the following holds. For

q^(t):=q(t)eγt(iI1Re(wie𝐢βit)ψi,1+Im(wie𝐢βit)ψi,2+iI2ci,3ψi,3+(tci,3+ci,4)ψi,4)\displaystyle\hat{q}(t):=q(t)-e^{\gamma_{*}t}\bigg{(}\sum_{i\in I_{1}}\textup{Re}\big{(}w_{i}e^{\mathbf{i}\beta_{i}t}\big{)}\psi_{i,1}+\textup{Im}\big{(}w_{i}e^{\mathbf{i}\beta_{i}t}\big{)}\psi_{i,2}+\sum_{i\in I_{2}}c_{i,3}\psi_{i,3}+(tc_{i,3}+c_{i,4})\psi_{i,4}\bigg{)}

there exists ε>0\varepsilon>0 such that q^(t)G=O(e(γε)t)\|\hat{q}(t)\|_{G}=O(e^{(\gamma_{*}-\varepsilon)t}).

Proof.

Define non-negative functions X±(t)X_{\pm}(t) by

X+2(t)\displaystyle X^{2}_{+}(t) =i:γi+>γ(ξi+(t))2+i:γi>γ(ξi(t))2,\displaystyle=\sum_{i:\gamma^{+}_{i}>\gamma_{*}}(\xi_{i}^{+}(t))^{2}+\sum_{i:\gamma^{-}_{i}>\gamma_{*}}(\xi_{i}^{-}(t))^{2},
X2(t)\displaystyle X^{2}_{-}(t) =i:γi+<γ(ξi+(t))2+i:γi<γ(ξi(t))2.\displaystyle=\sum_{i:\gamma^{+}_{i}<\gamma_{*}}(\xi_{i}^{+}(t))^{2}+\sum_{i:\gamma^{-}_{i}<\gamma_{*}}(\xi_{i}^{-}(t))^{2}.

By shrinking the value of ε0\varepsilon_{0} if necessary, we may assume ε0<|21mγi±|\varepsilon_{0}<|2^{-1}m-\gamma^{\pm}_{i}| for all iI3I4i\in I_{3}\cup I_{4}. An argument similar to the one in the proof of Lemma 4.3 shows X±(t)=O(e(γε0)t)X_{\pm}(t)=O(e^{(\gamma_{*}-\varepsilon_{0})t}). From (3.21), we derive for iI1i\in I_{1}

ddt(e(γ+𝐢βi)t(ξi,1(t)+𝐢ξi,2(t)))=e(γ+𝐢βi)t(i,1(t)+𝐢i,2(t)).\displaystyle\frac{d}{dt}\left(e^{-(\gamma_{*}+\mathbf{i}\beta_{i})t}\left(\xi_{i,1}(t)+\mathbf{i}\xi_{i,2}(t)\right)\right)=e^{-(\gamma_{*}+\mathbf{i}\beta_{i})t}\left(\mathcal{E}_{i,1}(t)+\mathbf{i}\mathcal{E}_{i,2}(t)\right).

Integrating the above from 0 to tt and using (4.7) yield

ξi,1(t)+𝐢ξi,2(t)=wie(γ+𝐢βi)t+O(e(γε0)t)\displaystyle\xi_{i,1}(t)+\mathbf{i}\xi_{i,2}(t)=w_{i}e^{(\gamma_{*}+\mathbf{i}\beta_{i})t}+O(e^{(\gamma_{*}-\varepsilon_{0})t})

for some wiw_{i}\in\mathbb{C}. A similar argument applying to ddt(eγtξi,3(t))=eγti,3(t)\frac{d}{dt}\left(e^{-\gamma_{*}t}\xi_{i,3}(t)\right)=e^{-\gamma_{*}t}\mathcal{E}_{i,3}(t) gives

ξi,3(t)=ci,3eγt+O(e(γε0)t)\xi_{i,3}(t)=c_{i,3}e^{\gamma_{*}t}+O(e^{(\gamma_{*}-\varepsilon_{0})t})

for some ci,3c_{i,3}\in\mathbb{R}. Lastly, from

ddt(eγtξi,4(t))=\displaystyle\frac{d}{dt}\left(e^{-\gamma_{*}t}\xi_{i,4}(t)\right)= eγtξ3,i(t)+eγti,4(t)=ci,3+O(eε0t),\displaystyle e^{-\gamma_{*}t}\xi_{3,i}(t)+e^{-\gamma_{*}t}\mathcal{E}_{i,4}(t)=c_{i,3}+O(e^{-\varepsilon_{0}t}),

we derive ξ4,i=(tci,3+ci,4)eγt+O(e(γε0)t)\xi_{4,i}=(tc_{i,3}+c_{i,4})e^{\gamma_{*}t}+O(e^{(\gamma_{*}-\varepsilon_{0})t}) for some ci,4c_{i,4}\in\mathbb{R}. ∎

For γ=γi+\gamma_{*}=\gamma^{+}_{i} or γi\gamma^{-}_{i}, an analogous result holds. We omit the proof because it is similar to and simpler than the one for Lemma 4.4.

Lemma 4.5.

Suppose γ=γi+\gamma_{*}=\gamma^{+}_{i} for some iI3I4i\in I_{3}\cup I_{4} and let NN be the multiplicity of λi\lambda_{i}. We may assume λi=λi+1==λi+N1\lambda_{i}=\lambda_{i+1}=\dots=\lambda_{i+N-1}. Then there exists a1,a2,,aNa_{1},a_{2},\dots,a_{N}\in\mathbb{R} and ε>0\varepsilon>0 such that

q(t)eγtj=1Najψi+j1+G=O(e(γε)t).\displaystyle\|q(t)-e^{\gamma_{*}t}\sum_{j=1}^{N}a_{j}\psi^{+}_{i+j-1}\|_{G}=O(e^{(\gamma_{*}-\varepsilon)t}).

Similar result holds when γ=γi\gamma_{*}=\gamma^{-}_{i} for some iI3I4i\in I_{3}\cup I_{4}

We are ready to prove Theorem 2.4.

Proof of Theorem 2.4.

Suppose γ=γi+\gamma_{*}=\gamma^{+}_{i} or γi\gamma^{-}_{i}. From Lemma 4.5 and Corollary 3.6, there exists vv with Σv=λiv\mathcal{L}_{\Sigma}v=\lambda_{i}v and ε>0\varepsilon>0 such that u(t)eγtvH1=O(e(γε)t)\|u(t)-e^{\gamma_{*}t}v\|_{H^{1}}=O(e^{(\gamma_{*}-\varepsilon)t}). Let w=v/vL2w=v/\|v\|_{L^{2}}. Then

limtu(t)/u(t)L2=winH1(Σ,𝐕).\displaystyle\lim_{t\to\infty}u(t)/\|u(t)\|_{L^{2}}=w\ \textup{in}\ H^{1}(\Sigma,\mathbf{V}).

We may upgrade the convergence to C(Σ,𝐕)C^{\infty}(\Sigma,\mathbf{V}) by taking 𝐋\mathbf{L} derivatives to (3.3). As a result, case (1) in Theorem 2.4 holds.

Suppose γ=21m\gamma_{*}=2^{-1}m. From Lemma 4.4 and Corollary 3.6, there exists wiw_{i}\in\mathbb{C} for iI1i\in I_{1} and ci,3,ci,4I2c_{i,3},c_{i,4}\in I_{2} such that

u(t)eγtiI1Re(wie𝐢βit)φieγtiI2(tci,3+ci,4)φiH1=O(e(γε)t).\displaystyle\left\|u(t)-e^{\gamma_{*}t}\sum_{i\in I_{1}}\textup{Re}\left(w_{i}e^{\mathbf{i}\beta_{i}t}\right)\varphi_{i}-e^{\gamma_{*}t}\sum_{i\in I_{2}}(tc_{i,3}+c_{i,4})\varphi_{i}\right\|_{H^{1}}=O(e^{(\gamma_{*}-\varepsilon)t}).

Then case (2) or case (3) in Theorem 2.4 holds depending on whether ci,30c_{i,3}\equiv 0 or not. ∎

5. Slowly decaying solutions to elliptic equation

In this section, we show that if a solution to (1.1) decays slowly, then the neutral mode, the projection of q(u)q(u) onto the 0-eigenspace of 𝐋\mathbf{L}, dominates the solution. Moreover, the neutral mode evolves by a gradient flow up to a small error. That is the content of Proposition 5.1.

Let uC(Q0,,𝐕~)u\in C^{\infty}(Q_{0,\infty},\widetilde{\mathbf{V}}) be a solution to (1.1) with uC1(t)=o(1)\|u\|_{C^{1}}(t)=o(1) as tt\to\infty. From the elliptic regularity, Lemma A.3, uCs(t)=o(1)\|u\|_{C^{s}}(t)=o(1) for all ss\in\mathbb{N}. We further assume that uu does not decay exponentially. Namely, for any ε>0\varepsilon>0,

(5.1) lim supteεtuC1(t)=.\limsup_{t\to\infty}e^{\varepsilon t}\|u\|_{C^{1}}(t)=\infty.

Recall that we rewrote (1.1) as an ODE system (3.21)-(3.23). For brevity, we assume throughout this section that I2=I_{2}=\varnothing. With notational changes, the proof can be readily extended to cover the case where I2I_{2}\neq\varnothing. Since I2=,I_{2}=\varnothing, the ODE system consists of (3.21) and (3.23). It is convenient to relabel the coefficients {ξi±}iI3I4\{\xi^{\pm}_{i}\}_{i\in I_{3}\cup I_{4}} in (3.23). For {ψi±}iI4\{\psi^{\pm}_{i}\}_{i\in I_{4}}, we set

{Ψi}i={ψi+|iI4andγi+>0}{ψi|iI4andγi>0},\displaystyle\{\Psi_{i}\}_{i\in\mathbb{N}}=\{\psi^{+}_{i}\,|\,i\in I_{4}\ \textup{and}\ \gamma^{+}_{i}>0\}\cup\{\psi^{-}_{i}\,|\,i\in I_{4}\ \textup{and}\ \gamma^{-}_{i}>0\},
{Ψi}i={ψi+|iI4andγi+<0}{ψi|iI4andγi<0},\displaystyle\{\Psi_{i}\}_{i\in-\mathbb{N}}=\{\psi^{+}_{i}\,|\,i\in I_{4}\ \textup{and}\ \gamma^{+}_{i}<0\}\cup\{\psi^{-}_{i}\,|\,i\in I_{4}\ \textup{and}\ \gamma^{-}_{i}<0\},

and define Γi\Gamma_{i}, a relabelling of γi±\gamma^{\pm}_{i} for iI4i\in I_{4}, by

(5.2) 𝐋Ψi=𝐋Ψi=ΓiΨi.\displaystyle\mathbf{L}\Psi_{i}=\mathbf{L}^{\dagger}\Psi_{i}=\Gamma_{i}\Psi_{i}.

Recall that I3={ι+1,,ι+J}I_{3}=\{\iota+1,\dots,\iota+J\}. If m>0m>0, we set for 1jJ1\leq j\leq J Υj=ψι+j\Upsilon_{j}=\psi^{-}_{\iota+j} and Υ¯j=ψι+j+\overline{\Upsilon}_{j}=\psi_{\iota+j}^{+}. If m<0m<0, we set Υj=ψι+j+\Upsilon_{j}=\psi^{+}_{\iota+j} and Υ¯j=ψι+j\overline{\Upsilon}_{j}=\psi_{\iota+j}^{-}. This arrangement ensures that

(5.3) 𝐋Υj=𝐋Υj=0,𝐋Υ¯j=𝐋Υ¯j=mΥ¯j.\begin{split}\mathbf{L}\Upsilon_{j}=\mathbf{L}^{\dagger}\Upsilon_{j}=0,\ \mathbf{L}\overline{\Upsilon}_{j}=\mathbf{L}^{\dagger}\overline{\Upsilon}_{j}=m\overline{\Upsilon}_{j}.\end{split}

Let

(5.4) ξi(t):=G(q(t),Ψi)fori{0},zj(t):=G(q(t),Υj),z¯j(t):=G(q(t),Υ¯j)for 1jJ,\begin{split}&\xi_{i}(t):=G(q(t),\Psi_{i})\ \textup{for}\ i\in\mathbb{Z}\setminus\{0\},\\ &z_{j}(t):=G(q(t),\Upsilon_{j}),\ \bar{z}_{j}(t):=G(q(t),\overline{\Upsilon}_{j})\ \textup{for}\ 1\leq j\leq J,\end{split}

and

(5.5) i(t):=G((t),Ψi)fori{0},𝒲j(t):=G((t),Υj),𝒲¯j(t):=G((t),Υ¯j)for 1jJ.\begin{split}&\mathcal{E}_{i}(t):=G(\mathcal{E}(t),\Psi_{i})\ \textup{for}\ i\in\mathbb{Z}\setminus\{0\},\\ &\mathcal{W}_{j}(t):=G(\mathcal{E}(t),\Upsilon_{j}),\ \overline{\mathcal{W}}_{j}(t):=G(\mathcal{E}(t),\overline{\Upsilon}_{j})\ \textup{for}\ 1\leq j\leq J.\end{split}

We rewrite (3.23) as (5.6) and (5.7) below. For i{0}i\in\mathbb{Z}\setminus\{0\},

(5.6) ddtξiΓiξi=i,\frac{d}{dt}\xi_{i}-\Gamma_{i}\xi_{i}=\mathcal{E}_{i},

and for 1jJ1\leq j\leq J,

(5.7) ddtzj=𝒲j,ddtz¯jmz¯j=𝒲¯j.\begin{split}&\frac{d}{dt}z_{j}=\mathcal{W}_{j},\ \frac{d}{dt}\bar{z}_{j}-m\bar{z}_{j}=\overline{\mathcal{W}}_{j}.\end{split}

We denote z(t):=(z1(t),zJ(t))z(t):=(z_{1}(t),\dots z_{J}(t)) and z¯(t):=(z¯1(t),z¯J(t))\bar{z}(t):=(\bar{z}_{1}(t),\dots\bar{z}_{J}(t)), and use |z(t)||z(t)| and |z¯(t)||\bar{z}(t)| to denote their Euclidean norms respectively. Recall the reduced functional ff is introduced in Proposition 2.1. The goal of this section is to prove the following proposition.

Proposition 5.1.

For a slowly decaying solution uu to (1.1) as described above, there holds

(5.8) |z¯(t)|2+iI1(|ξi,1(t)|2+|ξi,2(t)|2)+i0|ξi(t)|2=o(1)|z(t)|2.\displaystyle|\bar{z}(t)|^{2}+\sum_{i\in I_{1}}\left(|\xi_{i,1}(t)|^{2}+|\xi_{i,2}(t)|^{2}\right)+\sum_{i\neq 0}|\xi_{i}(t)|^{2}=o(1)|z(t)|^{2}.

Moreover, for all ε>0\varepsilon>0 there exists a positive constant C=C(u,m,Σ,N1,ε)C=C(u,m,\mathcal{M}_{\Sigma},N_{1},\varepsilon) such that

(5.9) |z(t)+m1f(z(t))|C|z(t)|pε/2.\displaystyle|z^{\prime}(t)+m^{-1}\nabla f(z(t))|\leq C|z(t)|^{p-\varepsilon/2}.

The remainder of this section is dedicated to proving Proposition 5.1, which involves two main parts. In Subsection 5.1, we demonstrate (as shown in Corollary 5.3) that any CsC^{s} norm of uu can be bounded by |z(t)||z(t)|. In Subsection 5.2, we obtain an enhanced decay rate in Lemma 5.11 through the decomposition (2.11). Proposition 5.1 is then a simple consequence of Lemma 5.11.

5.1. Bounding uCs\|u\|_{C^{s}}

To estimate CsC^{s} norms of uu, we need higher-derivative versions of the ODE system. Fix k,0k,\ell\in\mathbb{N}_{0}. Recall that q(k,)q^{(k,\ell)} and (k,)\mathcal{E}^{(k,\ell)} are given in Definition 3.5. Let

(5.10) ξi,1(k,)(t):=G(q(k,)(t),ψi,1),ξi,2(k,)(t):=G(q(k,)(t),ψi,2)foriI1,ξi(k,)(t):=G(q(k,)(t),Ψi)fori{0},zj(k,)(t):=G(q(k,)(t),Υj),z¯j(k,)(t):=G(q(k,)(t),Υ¯j)for 1jJ.\begin{split}&\xi^{(k,\ell)}_{i,1}(t):=G(q^{(k,\ell)}(t),\psi_{i,1}),\ \xi^{(k,\ell)}_{i,2}(t):=G(q^{(k,\ell)}(t),\psi_{i,2})\ \textup{for}\ i\in I_{1},\\ &\xi^{(k,\ell)}_{i}(t):=G(q^{(k,\ell)}(t),\Psi_{i})\ \textup{for}\ i\in\mathbb{Z}\setminus\{0\},\\ &z^{(k,\ell)}_{j}(t):=G(q^{(k,\ell)}(t),\Upsilon_{j}),\ \bar{z}^{(k,\ell)}_{j}(t):=G(q^{(k,\ell)}(t),\overline{\Upsilon}_{j})\ \textup{for}\ 1\leq j\leq J.\end{split}

Also, let

(5.11) i,1(k,)(t):=G((k,)(t),ψi,1),i,2(k,)(t):=G((k,)(t),ψi,2)foriI1,i(k,)(t):=G((k,)(t),Ψi)fori{0},𝒲j(k,)(t):=G((k,)(t),Υj),𝒲¯j(k,)(t):=G((k,)(t),Υ¯j)for 1jJ.\begin{split}&\mathcal{E}^{(k,\ell)}_{i,1}(t):=G(\mathcal{E}^{(k,\ell)}(t),\psi_{i,1}),\ \mathcal{E}^{(k,\ell)}_{i,2}(t):=G(\mathcal{E}^{(k,\ell)}(t),\psi_{i,2})\ \textup{for}\ i\in I_{1},\\ &\mathcal{E}^{(k,\ell)}_{i}(t):=G(\mathcal{E}^{(k,\ell)}(t),\Psi_{i})\ \textup{for}\ i\in\mathbb{Z}\setminus\{0\},\\ &\mathcal{W}^{(k,\ell)}_{j}(t):=G(\mathcal{E}^{(k,\ell)}(t),\Upsilon_{j}),\ \overline{\mathcal{W}}^{(k,\ell)}_{j}(t):=G(\mathcal{E}^{(k,\ell)}(t),\overline{\Upsilon}_{j})\ \textup{for}\ 1\leq j\leq J.\end{split}

Then for iI1i\in I_{1},

(5.12) ddtξi,1(k,)21mξi,1(k,)+βiξi,2(k,)=i,1(k,),ddtξi,2(k,)21mξi,2(k,)βiξi,1(k,)=i,2(k,),\begin{split}&\frac{d}{dt}\xi^{(k,\ell)}_{i,1}-2^{-1}m\xi^{(k,\ell)}_{i,1}+\beta_{i}\xi^{(k,\ell)}_{i,2}=\mathcal{E}^{(k,\ell)}_{i,1},\\ &\frac{d}{dt}\xi^{(k,\ell)}_{i,2}-2^{-1}m\xi^{(k,\ell)}_{i,2}-\beta_{i}\xi^{(k,\ell)}_{i,1}=\mathcal{E}^{(k,\ell)}_{i,2},\end{split}

for i{0}i\in\mathbb{Z}\setminus\{0\},

(5.13) ddtξi(k,)Γiξi(k,)=i(k,),\frac{d}{dt}\xi^{(k,\ell)}_{i}-\Gamma_{i}\xi^{(k,\ell)}_{i}=\mathcal{E}^{(k,\ell)}_{i},

and for 1jJ1\leq j\leq J,

(5.14) ddtzj(k,)=𝒲j(k,),ddtz¯j(k,)mz¯j(k,)=𝒲¯j(k,).\begin{split}&\frac{d}{dt}z^{(k,\ell)}_{j}=\mathcal{W}^{(k,\ell)}_{j},\ \frac{d}{dt}\bar{z}^{(k,\ell)}_{j}-m\bar{z}^{(k,\ell)}_{j}=\overline{\mathcal{W}}^{(k,\ell)}_{j}.\end{split}

In the next lemma, we use the Merle-Zaag ODE lemma [MZ98] (see Lemma A.1) to show that |z(t)||z(t)| dominates the other coefficients, thereby obtaining a stronger version of (5.8).

Lemma 5.2 (dominance of neutral mode).

For any s0s\in\mathbb{N}_{0},

k+s[|z(k,)(t)|2+|z¯(k,)(t)|2+iI1(|ξi,1(k,)(t)|2+|ξi,2(k,)(t)|2)+\displaystyle\sum_{k+\ell\leq s}\left[|z^{(k,\ell)}(t)|^{2}+|\bar{z}^{(k,\ell)}(t)|^{2}+\sum_{i\in I_{1}}\left(|\xi^{(k,\ell)}_{i,1}(t)|^{2}+|\xi^{(k,\ell)}_{i,2}(t)|^{2}\right)+\right. i0|ξi(k,)(t)|2]\displaystyle\left.\sum_{i\neq 0}|\xi^{(k,\ell)}_{i}(t)|^{2}\right]
=(1+o(1))|z(t)|2.\displaystyle=(1+o(1))|z(t)|^{2}.
Proof.

Fix s0s\in\mathbb{N}_{0}. We give the proof for the case m>0m>0. The argument for m<0m<0 is similar. Define three non-negative functions X+(t)X_{+}(t), X0(t)X_{0}(t) and X(t)X_{-}(t) by

X02(t)=\displaystyle X^{2}_{0}(t)= k+s1jJ|zj(k,)(t)|2,X2(t)=k+si|ξi(k,)(t)|2,\displaystyle\sum_{k+\ell\leq s}\sum_{1\leq j\leq J}|z^{(k,\ell)}_{j}(t)|^{2},\ X^{2}_{-}(t)=\sum_{k+\ell\leq s}\sum_{i\in-\mathbb{N}}|\xi^{(k,\ell)}_{i}(t)|^{2},

and

X+2(t)=k+s(iI1(|ξi,1(k,)(t)|2+|ξi,2(k,)(t)|2)+i|ξi(k,)(t)|2+1jJ|z¯j(k,)(t)|2).\displaystyle X^{2}_{+}(t)=\sum_{k+\ell\leq s}\left(\sum_{i\in I_{1}}\left(|\xi^{(k,\ell)}_{i,1}(t)|^{2}+|\xi^{(k,\ell)}_{i,2}(t)|^{2}\right)+\sum_{i\in\mathbb{N}}|\xi^{(k,\ell)}_{i}(t)|^{2}+\sum_{1\leq j\leq J}|\bar{z}^{(k,\ell)}_{j}(t)|^{2}\right).

From (5.12), (5.13) and (5.14), we compute

X+X+=\displaystyle X_{+}X_{+}^{\prime}= k+s(21miI1(|ξi,1(k,)|2+|ξi,2(k,)|2)+iΓi|ξi(k,)|2+m1jJ|z¯j(k,)|2)\displaystyle\sum_{k+\ell\leq s}\left(2^{-1}m\sum_{i\in I_{1}}\left(|\xi^{(k,\ell)}_{i,1}|^{2}+|\xi^{(k,\ell)}_{i,2}|^{2}\right)+\sum_{i\in\mathbb{N}}\Gamma_{i}|\xi^{(k,\ell)}_{i}|^{2}+m\sum_{1\leq j\leq J}|\bar{z}^{(k,\ell)}_{j}|^{2}\right)
+k+s(iI1(ξi,1(k,)i,1(k,)+ξi,2(k,)i,2(k,))+iξi(k,)i(k,)+1jJz¯j(k,)𝒲¯j(k,)).\displaystyle+\sum_{k+\ell\leq s}\left(\sum_{i\in I_{1}}\left(\xi^{(k,\ell)}_{i,1}\mathcal{E}^{(k,\ell)}_{i,1}+\xi^{(k,\ell)}_{i,2}\mathcal{E}^{(k,\ell)}_{i,2}\right)+\sum_{i\in\mathbb{N}}\xi^{(k,\ell)}_{i}\mathcal{E}^{(k,\ell)}_{i}+\sum_{1\leq j\leq J}\bar{z}^{(k,\ell)}_{j}\overline{\mathcal{W}}^{(k,\ell)}_{j}\right).

Denote the terms on the second line by X+(t)Y+(t)X_{+}(t)Y_{+}(t) and let bb be the minimum among 21m2^{-1}m and |Γi|,i{0}|\Gamma_{i}|,\ i\in\mathbb{Z}\setminus\{0\}. We have

X+bX+Y+.\displaystyle X_{+}^{\prime}-bX_{+}\geq Y_{+}.

Similarly, define Y0(t)Y_{0}(t) and Y(t)Y_{-}(t) by

X0(t)Y0(t)=k+s1jJzj(k,)(t)𝒲j(k,)(t),\displaystyle X_{0}(t)Y_{0}(t)=\sum_{k+\ell\leq s}\sum_{1\leq j\leq J}z^{(k,\ell)}_{j}(t)\mathcal{W}^{(k,\ell)}_{j}(t),

and

X(t)Y(t)=k+siξi(k,)(t)i(k,)(t).\displaystyle X_{-}(t)Y_{-}(t)=\sum_{k+\ell\leq s}\sum_{i\in-\mathbb{N}}\xi^{(k,\ell)}_{i}(t)\mathcal{E}^{(k,\ell)}_{i}(t).

It holds that

(5.15) X0=Y0,X+bXY.\displaystyle X_{0}^{\prime}=Y_{0},\quad X_{-}^{\prime}+bX_{-}\leq Y_{-}.

We now compare |X+(t)|2+|X0(t)|2+|X(t)|2|X_{+}(t)|^{2}+|X_{0}(t)|^{2}+|X_{-}(t)|^{2} and |Y+(t)|2+|Y0(t)|2+|Y(t)|2|Y_{+}(t)|^{2}+|Y_{0}(t)|^{2}+|Y_{-}(t)|^{2}.

|X+(t)|2+|X0(t)|2+|X(t)|2\displaystyle|X_{+}(t)|^{2}+|X_{0}(t)|^{2}+|X_{-}(t)|^{2}
=k+s[|z(k,)(t)|2+|z¯(k,)(t)|2+iI1(|ξi,1(t)|2+|ξi,2(t)|2)+i0|ξi(t)|2]\displaystyle=\sum_{k+\ell\leq s}\left[|z^{(k,\ell)}(t)|^{2}+|\bar{z}^{(k,\ell)}(t)|^{2}+\sum_{i\in I_{1}}\left(|\xi_{i,1}(t)|^{2}+|\xi_{i,2}(t)|^{2}\right)+\right.\left.\sum_{i\neq 0}|\xi_{i}(t)|^{2}\right]
=k+sq(k,)(t)G2.\displaystyle=\sum_{k+\ell\leq s}\|q^{(k,\ell)}(t)\|^{2}_{G}.

From the Cauchy-Schwarz inequality and (5.11),

|Y+(t)|2+|Y0(t)|2+|Y(t)|2\displaystyle|Y_{+}(t)|^{2}+|Y_{0}(t)|^{2}+|Y_{-}(t)|^{2}
k+s(iI1(|i,1(k,)(t)|2+|i,2(k,)(t)|2)+i{0}|i(k,)(t)|2+1jJ(|𝒲j(t)|2+|𝒲¯j(t)|2))\displaystyle\leq\sum_{k+\ell\leq s}\left(\sum_{i\in I_{1}}\left(|\mathcal{E}^{(k,\ell)}_{i,1}(t)|^{2}+|\mathcal{E}^{(k,\ell)}_{i,2}(t)|^{2}\right)+\sum_{i\in\mathbb{Z}\setminus\{0\}}|\mathcal{E}^{(k,\ell)}_{i}(t)|^{2}+\sum_{1\leq j\leq J}\left(|\mathcal{W}_{j}(t)|^{2}+|\overline{\mathcal{W}}_{j}(t)|^{2}\right)\right)
=k+s(k,)(t)G2.\displaystyle=\sum_{k+\ell\leq s}\|\mathcal{E}^{(k,\ell)}(t)\|^{2}_{G}.

From (3.18) in Corollary 3.6, |Y+(t)|2+|Y0(t)|2+|Y(t)|2=o(1)(|X+(t)|2+|X0(t)|2+|X(t)|2).|Y_{+}(t)|^{2}+|Y_{0}(t)|^{2}+|Y_{-}(t)|^{2}=o(1)\big{(}|X_{+}(t)|^{2}+|X_{0}(t)|^{2}+|X_{-}(t)|^{2}\big{)}. We can then apply the ODE lemma, Lemma A.1. In view of (A.4), the slow decay assumption (5.1) rules out the possibility that X(t)X_{-}(t) dominates. Hence

(5.16) |X+(t)|2+|X0(t)|2+|X(t)|2=(1+o(1))|X0(t)|2.|X_{+}(t)|^{2}+|X_{0}(t)|^{2}+|X_{-}(t)|^{2}=(1+o(1))|X_{0}(t)|^{2}.

It remains to show that

(5.17) |X0(t)|2=(1+o(1))|z(t)|2.|X_{0}(t)|^{2}=(1+o(1))|z(t)|^{2}.

For 1\ell\geq 1,

zj(k,)=G(𝐋qk,1,Υj)=G(qk,1,𝐋Υj)=0.\displaystyle z^{(k,\ell)}_{j}=G(\mathbf{L}q^{k,\ell-1},\Upsilon_{j})=G(q^{k,\ell-1},\mathbf{L}^{\dagger}\Upsilon_{j})=0.

For k1k\geq 1,

zj(k,)=ddtzj(k1,)=𝒲j(k1,).\displaystyle z^{(k,\ell)}_{j}=\frac{d}{dt}z^{(k-1,\ell)}_{j}=\mathcal{W}^{(k-1,\ell)}_{j}.

These imply

|X0(t)|2\displaystyle|X_{0}(t)|^{2} |z(t)|2+ks1|𝒲(k,0)(t)|2\displaystyle\leq|z(t)|^{2}+\sum_{k\leq s-1}|\mathcal{W}^{(k,0)}(t)|^{2}
|z(t)|2+o(1)(|X+(t)|2+|X0(t)|2+|X(t)|2)\displaystyle\leq|z(t)|^{2}+o(1)\left(|X_{+}(t)|^{2}+|X_{0}(t)|^{2}+|X_{-}(t)|^{2}\right)
|z(t)|2+o(1)|X0(t)|2.\displaystyle\leq|z(t)|^{2}+o(1)|X_{0}(t)|^{2}.

We used (5.16) in the last inequality. Therefore (5.17) holds and the proof is finished. ∎

In view of Corollary 3.6 and Lemma 5.2, for any s0s\in\mathbb{N}_{0},

k+stk∇̸uL2(t)=O(1)|z(t)|.\displaystyle\sum_{k+\ell\leq s}\|\partial_{t}^{k}\not{\nabla}^{\ell}u\|_{L^{2}}(t)=O(1)|z(t)|.

Applying the Sobolev embedding on Σ\Sigma, we can bound any CsC^{s} norm of uu.

Corollary 5.3 (control on higher derivatives).

For any ss\in\mathbb{N}, there exists a positive constant C=C(u,s)C=C(u,s) such that uCs(t)C|z(t)|.\|u\|_{C^{s}}(t)\leq C|z(t)|.

Lemma 5.4.

There hold the following statements:

  1. (1)

    |z(t)|C1t1|z(t)|\geq C^{-1}t^{-1} for t1t\geq 1, where C=C(u)C=C(u).

  2. (2)

    For any ε>0\varepsilon>0 there exists t0=t0(u,ε)t_{0}=t_{0}(u,\varepsilon), such that on [t0,)[t_{0},\infty), eεt|z(t)|e^{-\varepsilon t}|z(t)| non-increasing and eεt|z(t)|e^{\varepsilon t}|z(t)| non-decreasing.

Proof.

The proof directly follows from |z(t)|C|z(t)|2|z^{\prime}(t)|\leq C|z(t)|^{2} and limt|z(t)|=0\lim_{t\to\infty}|z(t)|=0. ∎

5.2. Enhanced decay rate

Recall that the decomposition of uu in (2.11)

(5.18) u=uT+H(uT)+u~.u=u^{T}+H(u^{T})+\tilde{u}^{\perp}.

Let us introduce an auxiliary quantity

(5.19) Q(t):=|z(t)|p+|z(t)||z¯(t)|+|z(t)|uC2(t)+|z(t)|u~C3(t).\displaystyle Q(t):=|z(t)|^{p}+|z(t)||\bar{z}(t)|+|z(t)|\left\|u^{\prime}\right\|_{C^{2}}(t)+|z(t)|\|\tilde{u}^{\perp}\|_{C^{3}}(t).
Lemma 5.5.

There exists a positive constant C=C(u,Σ,N1)C=C(u,\mathcal{M}_{\Sigma},N_{1}) such that

(5.20) |Σ(u)+f(z(t))Σu~|CQ(t),\left|\mathcal{M}_{\Sigma}(u)+\nabla f(z(t))-\mathcal{L}_{\Sigma}\tilde{u}^{\perp}\right|\leq CQ(t),

and

(5.21) |N1(u)|CQ(t).|N_{1}(u)|\leq CQ(t).
Proof.

In the proof we use CC to represent a positive constant that depends on uu, Σ\mathcal{M}_{\Sigma}, N1N_{1} in (2.6), and its value may vary from one line to another. Note that the coefficients of uTkerΣu^{T}\in\ker\mathcal{L}_{\Sigma} are given by zj(t)z¯j(t)z_{j}(t)-\bar{z}_{j}(t).

uT(t)=j=1J(zj(t)z¯j(t))φι+j.\displaystyle u^{T}(t)=\sum_{j=1}^{J}(z_{j}(t)-\bar{z}_{j}(t))\varphi_{\iota+j}.

Fix α=1/2\alpha=1/2. From Lemma 2.10,

|Σ(u)+f(z(t))Σu~|Cu(t)C2,α(Σ)u~(t)C2,α(Σ)+C|f(z(t)z¯(t))f(z(t))|.\displaystyle\left|\mathcal{M}_{\Sigma}(u)+\nabla f(z(t))-\mathcal{L}_{\Sigma}\tilde{u}^{\perp}\right|\leq C\|u(t)\|_{C^{2,\alpha}(\Sigma)}\|\tilde{u}^{\perp}(t)\|_{C^{2,\alpha}(\Sigma)}+C|\nabla f(z(t)-\bar{z}(t))-\nabla f(z(t))|.

Clearly u(t)C2,α(Σ)CuC3(t)\|u(t)\|_{C^{2,\alpha}(\Sigma)}\leq C\|u\|_{C^{3}}(t) and u~(t)C2,α(Σ)Cu~C3(t)\|\tilde{u}^{\perp}(t)\|_{C^{2,\alpha}(\Sigma)}\leq C\|\tilde{u}^{\perp}\|_{C^{3}}(t). From Corollary 5.3, uC3(t)C|z(t)|\|u\|_{C^{3}}(t)\leq C|z(t)|. Therefore,

u(t)C2,α(Σ)u~(t)C2,α(Σ)CQ(t).\displaystyle\|u(t)\|_{C^{2,\alpha}(\Sigma)}\|\tilde{u}^{\perp}(t)\|_{C^{2,\alpha}(\Sigma)}\leq CQ(t).

Because ff(0)f-f(0) vanishes at the origin of degree pp, |2f(x)|C|x|p2|\nabla^{2}f(x)|\leq C|x|^{p-2} near the origin. Together with |z¯(t)|C|z(t)||\bar{z}(t)|\leq C|z(t)|,

|f(z(t)z¯(t))f(z(t))|CQ(t).\displaystyle|\nabla f(z(t)-\bar{z}(t))-\nabla f(z(t))|\leq CQ(t).

Hence (5.20) holds. To show (5.21), recall that N1(u)=a1Du+a2u+a3Σ(u)N_{1}(u)=a_{1}\cdot Du^{\prime}+a_{2}\cdot u^{\prime}+a_{3}\cdot\mathcal{M}_{\Sigma}(u), where ai=ai(ω,u,∇̸u,u)a_{i}=a_{i}(\omega,u,\not{\nabla}u,u^{\prime}) are smooth with ai(ω,0,0,0)=0a_{i}(\omega,0,0,0)=0. From Corollary 5.3, |ai|C|z(t)||a_{i}|\leq C|z(t)|. Hence

|a1Du+a2u|C|z(t)|(|Du|+|u|)C|z(t)|uC1(t)CQ(t).\displaystyle|a_{1}\cdot Du^{\prime}+a_{2}\cdot u^{\prime}|\leq C|z(t)|(|Du^{\prime}|+|u^{\prime}|)\leq C|z(t)|\|u^{\prime}\|_{C^{1}}(t)\leq CQ(t).

Together with (5.20), (5.21) follows. ∎

Lemma 5.6.

There exists a positive constant C=C(u,m,Σ,N1)C=C(u,m,\mathcal{M}_{\Sigma},N_{1}) such that

(5.22) |z(t)+m1f(z(t))|CQ(t),\displaystyle|z^{\prime}(t)+m^{-1}\nabla f(z(t))|\leq CQ(t),

and

(5.23) |z¯(t)mz¯(t)+m1f(z(t))|CQ(t).\displaystyle|\bar{z}^{\prime}(t)-m\bar{z}(t)+m^{-1}\nabla f(z(t))|\leq CQ(t).
Proof.

In view of (5.7), it suffices to show

|𝒲j(t)+m1fxj(z(t))|CQ(t)and|𝒲¯j(t)+m1fxj(z(t))|CQ(t).\displaystyle\left|\mathcal{W}_{j}(t)+m^{-1}\frac{\partial f}{\partial x^{j}}(z(t))\right|\leq CQ(t)\ \textup{and}\ \left|\overline{\mathcal{W}}_{j}(t)+m^{-1}\frac{\partial f}{\partial x^{j}}(z(t))\right|\leq CQ(t).

From (5.5) and (u)=(0,E1(u))\mathcal{E}(u)=(0,E_{1}(u)), we actually have 𝒲j(t)=𝒲¯j(t)\mathcal{W}_{j}(t)=\overline{\mathcal{W}}_{j}(t). Recall that

𝒲j=G((u);Υj)=G((0,E1(u));(φι+j,21mφι+j))=m1ΣE1(u)φι+j𝑑μ,\displaystyle\mathcal{W}_{j}=G(\mathcal{E}(u);\Upsilon_{j})=G((0,E_{1}(u));(\varphi_{\iota+j},-2^{-1}m\varphi_{\iota+j}))=-m^{-1}\int_{\Sigma}E_{1}(u)\varphi_{\iota+j}\,d\mu,

where E1(u)=N1(u)Σ(u)+Σu{E}_{1}(u)=N_{1}(u)-\mathcal{M}_{\Sigma}(u)+\mathcal{L}_{\Sigma}u. Because φι+jkerΣ\varphi_{\iota+j}\in\ker\mathcal{L}_{\Sigma},

𝒲j=m1Σ(N1(u)Σ(u))φι+j𝑑μ,\displaystyle\mathcal{W}_{j}=-m^{-1}\int_{\Sigma}(N_{1}(u)-\mathcal{M}_{\Sigma}(u))\varphi_{\iota+j}\,d\mu,

Then the assertion follows from (5.20) and (5.21). ∎

Corollary 5.7.

Suppose |Q(t)|M|z(t)|q\left|Q(t)\right|\leq M|z(t)|^{q} for some M,q>0M,q>0. Then there exists a positive constant C=C(u,m,Σ,N1,M,q)C=C(u,m,\mathcal{M}_{\Sigma},N_{1},M,q) such that

|z(t)|C|z(t)|min(q,p1)and|z¯(t)|C|z(t)|min(q,p1).\left|{z}^{\prime}(t)\right|\leq C|z(t)|^{\min(q,p-1)}\ \textup{and}\ \left|\bar{z}(t)\right|\leq C|z(t)|^{\min(q,p-1)}.
Proof.

Because ff(0)f-f(0) vanishes at the origin of degree pp, |f(z(t))|C|z(t)|p1|\nabla f(z(t))|\leq C|z(t)|^{p-1}. Hence the bound for z(t)z^{\prime}(t) follows from (5.22). The the bound for z¯(t)\bar{z}(t) can be obtained by integrating (5.23). ∎

We vectorize u~\tilde{u}^{\perp} and perform the projection. Set q~:=q(u~)\tilde{q}:=q(\tilde{u}^{\perp}) and

ξ~i,1(t):=G(q~(t),ψi,1),ξ~i,2(t):=G(q~(t),ψi,2)foriI1,ξ~i(t):=G(q~(t),Ψi)fori{0},\begin{split}&\tilde{\xi}_{i,1}(t):=G(\tilde{q}(t),\psi_{i,1}),\ \tilde{\xi}_{i,2}(t):=G(\tilde{q}(t),\psi_{i,2})\ \textup{for}\ i\in I_{1},\\ &\tilde{\xi}_{i}(t):=G(\tilde{q}(t),\Psi_{i})\ \textup{for}\ i\in\mathbb{Z}\setminus\{0\},\end{split}

Note that because u~\tilde{u}^{\perp} is orthogonal to kerΣ\ker\mathcal{L}_{\Sigma}, those coefficients completely characterize q~\tilde{q}. The projections of higher order derivatives of q~\tilde{q}, namely q~(k,)=tk𝐋q~\tilde{q}^{(k,\ell)}={\partial_{t}^{k}}\mathbf{L}^{\ell}\tilde{q}, are defined similarly. In the lemma below, we show that {ξ~i,1,ξ~i,2}iI1\{\tilde{\xi}_{i,1},\tilde{\xi}_{i,2}\}_{i\in I_{1}}, {ξ~i}i{0}\{\tilde{\xi}_{i}\}_{i\in\mathbb{Z}\setminus\{0\}} and the higher order coefficients are bounded by |z(t)||z(t)|.

Lemma 5.8 (control on higher derivatives of q~\tilde{q}).

For any k,0k,\ell\in\mathbb{N}_{0}, there exists a positive constant C=C(u,m,+k)C=C(u,m,\ell+k) such that

iI1(|ξ~i,1(k,)(t)|2+|ξ~i,2(k,)(t)|2)+i0|ξ~i(k,)(t)|2C|z(t)|2.\displaystyle\sum_{i\in I_{1}}\left(|\tilde{\xi}^{(k,\ell)}_{i,1}(t)|^{2}+|\tilde{\xi}^{(k,\ell)}_{i,2}(t)|^{2}\right)+\sum_{i\neq 0}|\tilde{\xi}^{(k,\ell)}_{i}(t)|^{2}\leq C|z(t)|^{2}.
Proof.

Let s=+ks=\ell+k.

iI1(|ξ~i,1(k,)(t)|2+|ξ~i,2(k,)(t)|2)+i0|ξ~i(k,)(t)|2=q~(k,)G2(t)Cu~Cs+12(t).\displaystyle\sum_{i\in I_{1}}\left(|\tilde{\xi}^{(k,\ell)}_{i,1}(t)|^{2}+|\tilde{\xi}^{(k,\ell)}_{i,2}(t)|^{2}\right)+\sum_{i\neq 0}|\tilde{\xi}^{(k,\ell)}_{i}(t)|^{2}=\|\tilde{q}^{(k,\ell)}\|^{2}_{G}(t)\leq C\|\tilde{u}^{\perp}\|^{2}_{C^{s+1}}(t).

The assertion then follows from Lemma 2.11 (boundedness of decomposition) and Corollary 5.3 (control on higher derivatives). ∎

Let us define

(5.24) E~(u):=(u~)′′m(u~)+Σu~, and ~:=(0,E~(u)).\displaystyle\tilde{E}(u):=\left(\tilde{u}^{\perp}\right)^{\prime\prime}-m\left(\tilde{u}^{\perp}\right)^{\prime}+\mathcal{L}_{\Sigma}\tilde{u}^{\perp},\text{ and }\tilde{\mathcal{E}}:=(0,\tilde{E}(u)).

Set {~i,1,~i,2}iI1\{\tilde{\mathcal{E}}_{i,1},\tilde{\mathcal{E}}_{i,2}\}_{i\in I_{1}} and {~i}i{0}\{\tilde{\mathcal{E}}_{i}\}_{i\in\mathbb{Z}\setminus\{0\}} be the coefficients of ~\tilde{\mathcal{E}}. Namely,

(5.25) ~i,1(t):=G(~(t),ψi,1),~i,2(t):=G(~(t),ψi,2)foriI1,~i(t):=G(~(t),Ψi)fori{0}.\begin{split}&\tilde{\mathcal{E}}_{i,1}(t):=G(\tilde{\mathcal{E}}(t),\psi_{i,1}),\ \tilde{\mathcal{E}}_{i,2}(t):=G(\tilde{\mathcal{E}}(t),\psi_{i,2})\ \textup{for}\ i\in I_{1},\\ &\tilde{\mathcal{E}}_{i}(t):=G(\tilde{\mathcal{E}}(t),\Psi_{i})\ \textup{for}\ i\in\mathbb{Z}\setminus\{0\}.\end{split}

Then we have, for iI1i\in I_{1},

(5.26) ddtξ~i,121mξ~i,1+βiξ~i,2=~i,1,ddtξ~i,221mξ~i,2βiξ~i,1=~i,2,\begin{split}&\frac{d}{dt}\tilde{\xi}_{i,1}-2^{-1}m\tilde{\xi}_{i,1}+\beta_{i}\tilde{\xi}_{i,2}=\tilde{\mathcal{E}}_{i,1},\\ &\frac{d}{dt}\tilde{\xi}_{i,2}-2^{-1}m\tilde{\xi}_{i,2}-\beta_{i}\tilde{\xi}_{i,1}=\tilde{\mathcal{E}}_{i,2},\end{split}

for i{0}i\in\mathbb{Z}\setminus\{0\},

(5.27) ddtξ~iΓiξ~i=~i.\frac{d}{dt}\tilde{\xi}_{i}-\Gamma_{i}\tilde{\xi}_{i}=\tilde{\mathcal{E}}_{i}.
Lemma 5.9.

There exists a positive constant C=C(u,m,Σ,N1)C=C(u,m,\mathcal{M}_{\Sigma},N_{1}) such that

E~(u)L2(t)CQ(t).\displaystyle\|\tilde{E}(u)\|_{L^{2}}(t)\leq CQ(t).
Proof.

In the proof we use CC to represent a positive constant that depends on uu, mm, Σ\mathcal{M}_{\Sigma}, N1N_{1} in (2.6), and its value may vary from one line to another.

We compute

E~(u)\displaystyle\tilde{E}(u) =Π[E(u)(H(uT))′′+m(H(uT))ΣH(uT)]\displaystyle=\Pi^{\perp}\bigg{[}E(u)-\left(H(u^{T})\right)^{\prime\prime}+m\left(H(u^{T})\right)^{\prime}-\mathcal{L}_{\Sigma}H(u^{T})\bigg{]}
=Π[Σ(u)+Σu+N1(u)(H(uT))′′+m(H(uT))ΣH(uT)]\displaystyle=\Pi^{\perp}\bigg{[}-\mathcal{M}_{\Sigma}(u)+\mathcal{L}_{\Sigma}u+N_{1}(u)-\left(H(u^{T})\right)^{\prime\prime}+m\left(H(u^{T})\right)^{\prime}-\mathcal{L}_{\Sigma}H(u^{T})\bigg{]}
=:I+II,\displaystyle=:\textsc{I+II},

where

I =ΠΣ(u)+Σu~+ΠN1(u),II=(H(uT))′′+m(H(uT)).\displaystyle=-\Pi^{\perp}\mathcal{M}_{\Sigma}(u)+\mathcal{L}_{\Sigma}\tilde{u}^{\perp}+\Pi^{\perp}N_{1}(u),\ \textsc{II}=-\left(H(u^{T})\right)^{\prime\prime}+m\left(H(u^{T})\right)^{\prime}.

From (5.20) and (5.21), we have IL2(t)CQ(t).\|\textsc{I}\|_{L^{2}}(t)\leq CQ(t). From a direct computation,

II=D2H(zz¯)[(zz¯),(zz¯)]DH(zz¯)[(zz¯)′′]+mDH(zz¯)[(zz¯)].\displaystyle\textsc{II}=-D^{2}H(z-\bar{z})[(z-\bar{z})^{\prime},(z-\bar{z})^{\prime}]-DH(z-\bar{z})[(z-\bar{z})^{\prime\prime}]+mDH(z-\bar{z})[(z-\bar{z})^{\prime}].

Since DH(0)=0DH(0)=0, the above is bounded by

C|(zz¯)|2+C|zz¯||(zz¯)′′|+C|zz¯||(zz¯)|CuTC2(t)tuTC1(t).C|(z-\bar{z})^{\prime}|^{2}+C|z-\bar{z}||(z-\bar{z})^{\prime\prime}|+C|z-\bar{z}||(z-\bar{z})^{\prime}|\leq C\|u^{T}\|_{C^{2}}(t)\|\partial_{t}u^{T}\|_{C^{1}}(t).

Together with Corollary 5.3, we get

IIL2(t)C|z(t)|uC1(t)CQ(t).\|\textsc{II}\|_{L^{2}}(t)\leq C|z(t)|\|u^{\prime}\|_{C^{1}}(t)\leq CQ(t).

Corollary 5.10.

Suppose Q(t)M|z(t)|qQ(t)\leq M|z(t)|^{q} for some M,q>0M,q>0. Then there exists a positive constant C=C(u,m,Σ,N1,M,q)C=C(u,m,\mathcal{M}_{\Sigma},N_{1},M,q) such that

iI1(|ξ~i,1(t)|2+|ξ~i,2(t)|2)+i{0}|ξ~i(t)|2C|z(t)|2q.\displaystyle\sum_{i\in I_{1}}\left(|\tilde{\xi}_{i,1}(t)|^{2}+|\tilde{\xi}_{i,2}(t)|^{2}\right)+\sum_{i\in\mathbb{Z}\setminus\{0\}}|\tilde{\xi}_{i}(t)|^{2}\leq C|z(t)|^{2q}.
Proof.

In the proof we use CC to represent a positive constant that depends on uu, mm, Σ\mathcal{M}_{\Sigma}, N1N_{1} in (2.6), MM, qq, and its value may vary from one line to another. We shall prove

(5.28) i|ξ~i(t)|2C|z(t)|2q,\displaystyle\sum_{i\in-\mathbb{N}}|\tilde{\xi}_{i}(t)|^{2}\leq C|z(t)|^{2q},

by directly solving evolution equations. The other terms can be treated similarly. For ii\in-\mathbb{N}, from (5.27),

ξ~i(t)=eΓitξ~(0)+0teΓi(tτ)~i(τ)𝑑τ.\displaystyle\tilde{\xi}_{i}(t)=e^{\Gamma_{i}t}\tilde{\xi}(0)+\int_{0}^{t}e^{\Gamma_{i}(t-\tau)}\tilde{\mathcal{E}}_{i}(\tau)\,d\tau.

For any sequence {ai}i\{a_{i}\}_{i\in-\mathbb{N}} with i|ai|2=1\sum_{i\in-\mathbb{N}}|a_{i}|^{2}=1, we compute

|iξ~i(t)ai|\displaystyle\left|\sum_{i\in-\mathbb{N}}\tilde{\xi}_{i}(t)a_{i}\right|\leq i|eΓitξ~i(0)ai|+i0teΓi(tτ)|~i(τ)ai|𝑑τ\displaystyle\sum_{i\in-\mathbb{N}}\left|e^{\Gamma_{i}t}\tilde{\xi}_{i}(0)a_{i}\right|+\sum_{i\in-\mathbb{N}}\int_{0}^{t}e^{\Gamma_{i}(t-\tau)}\left|\tilde{\mathcal{E}}_{i}(\tau)a_{i}\right|\,d\tau
\displaystyle\leq ebti|ξ~i(0)ai|+i0teb(tτ)|~i(τ)ai|𝑑τ.\displaystyle e^{-bt}\sum_{i\in-\mathbb{N}}\left|\tilde{\xi}_{i}(0)a_{i}\right|+\sum_{i\in-\mathbb{N}}\int_{0}^{t}e^{-b(t-\tau)}\left|\tilde{\mathcal{E}}_{i}(\tau)a_{i}\right|\,d\tau.

From (1) in Lemma 5.4, ebti|ξ~i(0)ai|C|z(t)|qe^{-bt}\sum_{i\in-\mathbb{N}}\left|\tilde{\xi}_{i}(0)a_{i}\right|\leq C|z(t)|^{q}. From (5.24), (5.25), Lemma 5.9 and the assumption,

iI1(|~i,1(t)|2+|~i,2(t)|2)+i{0}|~i(t)|2=~G2=2m2E~(u)L22(t)C|z(t)|q.\displaystyle\sum_{i\in I_{1}}\left(|\tilde{\mathcal{E}}_{i,1}(t)|^{2}+|\tilde{\mathcal{E}}_{i,2}(t)|^{2}\right)+\sum_{i\in\mathbb{Z}\setminus\{0\}}|\tilde{\mathcal{E}}_{i}(t)|^{2}=\|\tilde{\mathcal{E}}\|^{2}_{G}=2m^{-2}\|\tilde{E}(u)\|^{2}_{L^{2}}(t)\leq C|z(t)|^{q}.

By the Cauchy-Schwarz inequality,

i0teb(tτ)|~i(τ)ai|𝑑τC0teb(tτ)|z(τ)|q𝑑τ.\displaystyle\sum_{i\in-\mathbb{N}}\int_{0}^{t}e^{-b(t-\tau)}\left|\tilde{\mathcal{E}}_{i}(\tau)a_{i}\right|\,d\tau\leq C\int_{0}^{t}e^{-b(t-\tau)}|z(\tau)|^{q}\,d\tau.

Let t0=t0(u,21q1b)t_{0}=t_{0}(u,2^{-1}q^{-1}b) be the constant given in (2) of Lemma 5.4. Then

0teb(tτ)|z(τ)|q𝑑τ=\displaystyle\int_{0}^{t}e^{-b(t-\tau)}|z(\tau)|^{q}\,d\tau= 0t0eb(tτ)|z(τ)|q𝑑τ+t0te21b(tτ)(e21q1b(tτ)|z(τ)|)q𝑑τ\displaystyle\int_{0}^{t_{0}}e^{-b(t-\tau)}|z(\tau)|^{q}\,d\tau+\int_{t_{0}}^{t}e^{-2^{-1}b(t-\tau)}\left(e^{-2^{-1}q^{-1}b(t-\tau)}|z(\tau)|\right)^{q}\,d\tau
C\displaystyle\leq C ebtmaxτ[0,)|z(τ)|q+C|z(t)|qC|z(t)|q.\displaystyle e^{-bt}\max_{\tau\in[0,\infty)}|z(\tau)|^{q}+C|z(t)|^{q}\leq C|z(t)|^{q}.

In short, we deduce |iξ~i(t)ai|C|z(t)|q\left|\sum_{i\in-\mathbb{N}}\tilde{\xi}_{i}(t)a_{i}\right|\leq C|z(t)|^{q} for all sequence {ai}\{a_{i}\} with i|ai|2=1\sum_{i\in-\mathbb{N}}|a_{i}|^{2}=1. This implies (5.28).

Lemma 5.11 (improvement in decay).

For 1rp11\leq r\leq p-1, suppose the following holds.

(5.29) |z(t)|2+|z¯(t)|2+iI1(|ξ~i,1(t)|2+|ξ~i,2(t)|2)+i0|ξ~i(t)|2M|z(t)|2r.\displaystyle\left|z^{\prime}(t)\right|^{2}+|\bar{z}(t)|^{2}+\sum_{i\in I_{1}}\left(|\tilde{\xi}_{i,1}(t)|^{2}+|\tilde{\xi}_{i,2}(t)|^{2}\right)+\sum_{i\neq 0}|\tilde{\xi}_{i}(t)|^{2}\leq M|z(t)|^{2r}.

Then for any ε(0,1)\varepsilon\in(0,1), there exists a positive constant C=C(u,m,Σ,N1,M,r,ε)C=C(u,m,\mathcal{M}_{\Sigma},N_{1},M,r,\varepsilon) such that

(5.30) |z(t)|2+|z¯(t)|2+iI1(|ξ~i,1(t)|2+|ξ~i,2(t)|2)+i0|ξ~i(t)|2C|z(t)|min{2r+2ε,2p2}.\displaystyle\left|z^{\prime}(t)\right|^{2}+|\bar{z}(t)|^{2}+\sum_{i\in I_{1}}\left(|\tilde{\xi}_{i,1}(t)|^{2}+|\tilde{\xi}_{i,2}(t)|^{2}\right)+\sum_{i\neq 0}|\tilde{\xi}_{i}(t)|^{2}\leq C|z(t)|^{\min\{2r+2-\varepsilon,2p-2\}}.

Moreover,

(5.31) |z(t)+m1f(z(t))|C|z(t)|r+1ε/2.\displaystyle\left|z^{\prime}(t)+m^{-1}\nabla f(z(t))\right|\leq C|z(t)|^{r+1-\varepsilon/2}.
Proof.

Fix ε(0,1)\varepsilon\in(0,1). In the proof we use CC to represent a positive constant that depends on uu, mm, Σ\mathcal{M}_{\Sigma}, N1N_{1} in (2.6), MM, rr, ε\varepsilon, and its value may vary from one line to another. Using the assumption (5.29), Lemma 5.2 (dominance of neutral mode) and Lemma A.2 (interpolation), we have for all k,0k,\ell\in\mathbb{N}_{0}, |ddtz(k,)(t)|2+|z¯(k,)(t)|2Ck,|z(t)|2rε.\left|\frac{d}{dt}z^{(k,\ell)}(t)\right|^{2}+|\bar{z}^{(k,\ell)}(t)|^{2}\leq C_{k,\ell}|z(t)|^{2r-\varepsilon}. Similarly, applying Lemma A.2 with Lemma 5.8,

(5.32) iI1(|ξ~i,1(k,)(t)|2+|ξ~i,2(k,)(t)|2)+i0|ξ~i(k,)(t)|2Ck,|z(t)|2rε.\displaystyle\sum_{i\in I_{1}}\left(|\tilde{\xi}^{(k,\ell)}_{i,1}(t)|^{2}+|\tilde{\xi}^{(k,\ell)}_{i,2}(t)|^{2}\right)+\sum_{i\neq 0}|\tilde{\xi}^{(k,\ell)}_{i}(t)|^{2}\leq C_{k,\ell}|z(t)|^{2r-\varepsilon}.

By the Sobolev embedding, (uT)Cs(t)+u~Cs(t)Cs|z(t)|rε/2\|(u^{T})^{\prime}\|_{C^{s}}(t)+\|\tilde{u}^{\perp}\|_{C^{s}}(t)\leq C_{s}|z(t)|^{r-\varepsilon/2}. From (5.18) and (5.32), we also have uCs(t)Cs|z(t)|rε/2\|u^{\prime}\|_{C^{s}}(t)\leq C_{s}|z(t)|^{r-\varepsilon/2}. In view of the definition of Q(t)Q(t) in (5.19),

Q(t)C|z(t)|min(p,r+1ε/2)=C|z(t)|r+1ε/2.Q(t)\leq C|z(t)|^{\min(p,r+1-\varepsilon/2)}=C|z(t)|^{r+1-\varepsilon/2}.

From Corollary 5.10,

(5.33) iI1(|ξ~i,1(t)|2+|ξ~i,2(t)|2)+i0|ξ~i(t)|2C|z(t)|2r+2ε.\displaystyle\sum_{i\in I_{1}}\left(|\tilde{\xi}_{i,1}(t)|^{2}+|\tilde{\xi}_{i,2}(t)|^{2}\right)+\sum_{i\neq 0}|\tilde{\xi}_{i}(t)|^{2}\leq C|z(t)|^{2r+2-\varepsilon}.

From Corollary 5.7, (5.31) and

(5.34) |z¯(t)|2C|z(t)|min{2r+2ε,2p2}.|\bar{z}(t)|^{2}\leq C|z(t)|^{\min\{2r+2-\varepsilon,2p-2\}}.

Note that (5.31) implies

(5.35) |z(t)|2C|z(t)|min{2r+2ε,2p2}.\displaystyle|z^{\prime}(t)|^{2}\leq C|z(t)|^{\min\{2r+2-\varepsilon,2p-2\}}.

Combining (5.33), (5.35) and (5.34) yields (5.30). ∎

Proof of Proposition 5.1.

The bound (5.8) is a weaker form of Lemma 5.2 (dominance of neutral mode). From Lemma 5.2 (dominance of neutral mode) and Lemma 5.8 (control on higher derivatives of q~\tilde{q}), the assumption (5.29) in Lemma 5.11 (improvement in decay) holds for r=1r=1. By iterating Lemma 5.11, (5.29) holds for r=p1r=p-1. Then the (5.9) follows from (5.31). ∎

6. Slowly decaying solutions to parabolic equation

In this section, we consider the slowly decaying solutions to the parabolic equation (1.2). The main goal is to prove Proposition 6.1, which is analogous to Proposition 5.1.

Let uC(Q0,,𝐕~)u\in C^{\infty}(Q_{0,\infty},\widetilde{\mathbf{V}}) be a solution to (1.2) with uHn+4(t)=o(1)\|u\|_{H^{n+4}}(t)=o(1) as tt\to\infty. From Lemma A.4 (parabolic regularity), uCs(t)=o(1)\|u\|_{C^{s}}(t)=o(1) for all ss\in\mathbb{N}. We further assume that uu does not decay exponentially. Namely, for any ε>0\varepsilon>0,

(6.1) lim supteεtuC1(t)=.\limsup_{t\to\infty}e^{\varepsilon t}\|u\|_{C^{1}}(t)=\infty.

We project uu onto the eigensections φi\varphi_{i}. Set

(6.2) ξi(t):=Σu,φi𝑑μ.\displaystyle\xi_{i}(t):=\int_{\Sigma}\left\langle u,\varphi_{i}\right\rangle\,d\mu.

Recall that {i:λi=0}={ι+1,ι+2,,ι+J}\{i\in\mathbb{N}\,:\,\lambda_{i}=0\}=\{\iota+1,\iota+2,\dots,\iota+J\}. The neutral mode will play a special role and we denote it by

(6.3) xj(t):=ξι+j(t).\displaystyle x_{j}(t):=\xi_{\iota+j}(t).
Proposition 6.1.

For a slowly decaying solution uu to (1.2) as described above, there holds

(6.4) i:λi0|ξi(t)|2=o(1)|x(t)|2.\displaystyle\sum_{i:\lambda_{i}\neq 0}|\xi_{i}(t)|^{2}=o(1)|x(t)|^{2}.

Moreover, for all ε>0\varepsilon>0, there exists a positive constant C=C(u,Σ,N2,ε)C=C(u,\mathcal{M}_{\Sigma},N_{2},\varepsilon) such that

(6.5) |x(t)+f(x(t))|C|x(t)|pε/2.\displaystyle|x^{\prime}(t)+\nabla f(x(t))|\leq C|x(t)|^{p-\varepsilon/2}.

The proof of Proposition 6.1 follows a similar structure to the one of Proposition 5.1. The first part is to show (in Corollary 6.4) that |x(t)||x(t)| dominates any CsC^{s} norm of uu.


Let E2(u)=N2(u)+Σ(u)ΣuE_{2}(u)=N_{2}(u)+\mathcal{M}_{\Sigma}(u)-\mathcal{L}_{\Sigma}u. Then (1.2) becomes

(6.6) uΣu=E2(u).u^{\prime}-\mathcal{L}_{\Sigma}u=E_{2}(u).

From (2.4) and (2.6), the error term E2(u)E_{2}(u) has the structure

(6.7) E2(u)=j=02b2,j∇̸ju,\displaystyle E_{2}(u)=\sum_{j=0}^{2}b_{2,j}\cdot\not{\nabla}^{j}u,

where b2,j=b2,j(ω,u,∇̸u)b_{2,j}=b_{2,j}(\omega,u,\not{\nabla}u) are smooth with b2,j(ω,0,0)=0b_{2,j}(\omega,0,0)=0. With uCs(t)=o(1)\|u\|_{C^{s}}(t)=o(1), the quadratic nature of E2E_{2} allows us to bound higher derivatives of E2E_{2} in terms of higher derivatives of uu.

Lemma 6.2.

For any s2s\geq 2,

2k+2stk∇̸E2(u)L2(t)=o(1)2k+2stk∇̸uL2(t).\displaystyle\sum_{2k+\ell\leq 2s}\|\partial^{k}_{t}\not{\nabla}^{\ell}E_{2}(u)\|_{L^{2}}(t)=o(1)\sum_{2k+\ell\leq 2s}\|\partial^{k}_{t}\not{\nabla}^{\ell}u\|_{L^{2}}(t).
Proof.

We present the proof for

2k+2stk∇̸(b2,2∇̸2u)L2(t)=o(1)2k+2stk∇̸uL2(t).\displaystyle\sum_{2k+\ell\leq 2s}\left\|\partial^{k}_{t}\not{\nabla}^{\ell}\left(b_{2,2}\cdot\not{\nabla}^{2}u\right)\right\|_{L^{2}}(t)=o(1)\sum_{2k+\ell\leq 2s}\|\partial^{k}_{t}\not{\nabla}^{\ell}u\|_{L^{2}}(t).

The other terms can be treated similarly. For simplicity, we write b(ω,u,∇̸u)b(\omega,u,\not{\nabla}u) for b2,2(ω,u,∇̸u)b_{2,2}(\omega,u,\not{\nabla}u). For m0,m1,m20m_{0},m_{1},m_{2}\in\mathbb{N}_{0}, we write b(m0,m1,m2)b^{(m_{0},m_{1},m_{2})} for the partial derivative of bb of order (m0,m1,m2)(m_{0},m_{1},m_{2}). Fix k0,0k_{0},\ell_{0} with 2k0+02s2k_{0}+\ell_{0}\leq 2s. Then the terms in the expansion of tk0∇̸0(b∇̸2u)\partial^{k_{0}}_{t}\not{\nabla}^{\ell_{0}}\left(b\cdot\not{\nabla}^{2}u\right) are of the form

b(m0,m1,m2)(tk1∇̸1utk2∇̸2utkN∇̸Nu),\displaystyle b^{(m_{0},m_{1},m_{2})}\cdot\left(\partial^{k_{1}}_{t}\not{\nabla}^{\ell_{1}}u\ast\partial^{k_{2}}_{t}\not{\nabla}^{\ell_{2}}u\ast\dots\partial^{k_{N}}_{t}\not{\nabla}^{\ell_{N}}u\right),

where N=m1+m2+1N=m_{1}+m_{2}+1, i=1Nki=k0\sum_{i=1}^{N}k_{i}=k_{0} and i=1Ni=0+m2+2\sum_{i=1}^{N}\ell_{i}=\ell_{0}+m_{2}+2. It suffices to show the pointwise bound

b(m0,m1,m2)((tk1∇̸1u)(tk2∇̸2u)(tkN∇̸Nu))=o(1)2k+2s|tk∇̸u|.b^{(m_{0},m_{1},m_{2})}\cdot\left(\big{(}\partial^{k_{1}}_{t}\not{\nabla}^{\ell_{1}}u\big{)}\cdot\big{(}\partial^{k_{2}}_{t}\not{\nabla}^{\ell_{2}}u\big{)}\cdot\ldots\cdot\big{(}\partial^{k_{N}}_{t}\not{\nabla}^{\ell_{N}}u\big{)}\right)=o(1)\sum_{2k+\ell\leq 2s}|\partial^{k}_{t}\not{\nabla}^{\ell}u|.

We first discuss the case m1=m2=0m_{1}=m_{2}=0. The above becomes b(m0,0,0)tk0∇̸0m0+2ub^{(m_{0},0,0)}\cdot\partial^{k_{0}}_{t}\not{\nabla}^{\ell_{0}-m_{0}+2}u. From

|b(m0,0,0)|C(|u|+|∇̸u|)and|tk0∇̸0m0+2u|=o(1),|b^{(m_{0},0,0)}|\leq C(|u|+|\not{\nabla}u|)\ \textup{and}\ |\partial^{k_{0}}_{t}\not{\nabla}^{\ell_{0}-m_{0}+2}u|=o(1),

the assertion holds.

Next, we consider the case m1+m21m_{1}+m_{2}\geq 1. It suffices to show mini(2ki+i)2s\min_{i}(2k_{i}+\ell_{i})\leq 2s as the other terms can be bounded by o(1)o(1). Suppose this fails. In other words, 2ki+i2s+12k_{i}+\ell_{i}\geq 2s+1 for all 1iN1\leq i\leq N. Then

2Ns+Ni=1N(2ki+i)=2k0+0+m2+22s+2N.\displaystyle 2Ns+N\leq\sum_{i=1}^{N}(2k_{i}+\ell_{i})=2k_{0}+\ell_{0}+m_{2}+2\leq 2s+2N.

In view of N2N\geq 2 and s2s\geq 2, this is a contradiction. ∎

Now we project (6.6) onto the eigensections φi\varphi_{i}. Let

ξi(k,)(t):=ΣΣtku,φi𝑑μ,i(k,)(t):=ΣΣtkE2(u),φi𝑑μ\displaystyle\xi^{(k,\ell)}_{i}(t):=\int_{\Sigma}\left\langle\mathcal{L}_{\Sigma}^{\ell}\partial^{k}_{t}u,\varphi_{i}\right\rangle\,d\mu,\ \mathcal{E}^{(k,\ell)}_{i}(t):=\int_{\Sigma}\left\langle\mathcal{L}_{\Sigma}^{\ell}\partial^{k}_{t}E_{2}(u),\varphi_{i}\right\rangle\,d\mu

Then (6.6) becomes

(6.8) ddtξi(k,)(t)λiξi(k,)(t)=i(k,)(t).\displaystyle\frac{d}{dt}\xi^{(k,\ell)}_{i}(t)-\lambda_{i}\xi^{(k,\ell)}_{i}(t)=\mathcal{E}^{(k,\ell)}_{i}(t).

For the neutral mode, we denote

xj(k,)(t):=ξι+j(k,)(t),𝒲j(k,)(t):=ι+j(k,)(t)\displaystyle x^{(k,\ell)}_{j}(t):=\xi^{(k,\ell)}_{\iota+j}(t),\ \mathcal{W}^{(k,\ell)}_{j}(t):=\mathcal{E}^{(k,\ell)}_{\iota+j}(t)
Lemma 6.3.

For any s2s\geq 2,

k+si=1|ξi(k,)(t)|2=(1+o(1))|x(t)|2.\displaystyle\sum_{k+\ell\leq s}\sum_{i=1}^{\infty}|\xi^{(k,\ell)}_{i}(t)|^{2}=(1+o(1))|x(t)|^{2}.
Proof.

Fix s2s\geq 2. Define three non-negative functions X+(t)X_{+}(t), X0(t)X_{0}(t) and X(t)X_{-}(t) by

X02(t)=\displaystyle X^{2}_{0}(t)= k+s1jJ|xj(k,)(t)|2,X±2(t)=k+si:±λi>0|ξi(k,)(t)|2.\displaystyle\sum_{k+\ell\leq s}\sum_{1\leq j\leq J}|x^{(k,\ell)}_{j}(t)|^{2},\ X^{2}_{\pm}(t)=\sum_{k+\ell\leq s}\sum_{i:\pm\lambda_{i}>0}|\xi^{(k,\ell)}_{i}(t)|^{2}.

We note that these coefficients are grouped together according to the sign of the eigenvalues. From (6.8), we compute

X+X+=\displaystyle X_{+}X_{+}^{\prime}= k+si:λi>0(λi|ξi(k,)(t)|2+ξi(k,)(t)i(k,)(t))\displaystyle\sum_{k+\ell\leq s}\sum_{i:\lambda_{i}>0}\left(\lambda_{i}|\xi^{(k,\ell)}_{i}(t)|^{2}+\xi^{(k,\ell)}_{i}(t)\mathcal{E}^{(k,\ell)}_{i}(t)\right)
=\displaystyle= k+si:λi>0(λi|ξi(k,)(t)|2)+X+Y+,\displaystyle\sum_{k+\ell\leq s}\sum_{i:\lambda_{i}>0}\left(\lambda_{i}|\xi^{(k,\ell)}_{i}(t)|^{2}\right)+X_{+}Y_{+},

here X+(t)Y+(t)X_{+}(t)Y_{+}(t) is defined by the last equality. Let b=min{|λi|:λi0}b=\min\{|\lambda_{i}|\,:\,\lambda_{i}\neq 0\}. We have

X+bX+Y+.\displaystyle X_{+}^{\prime}-bX_{+}\geq Y_{+}.

Similarly, define Y0(t)Y_{0}(t) and Y(t)Y_{-}(t) by

X0(t)Y0(t)=k+s1jJxj(k,)(t)𝒲j(k,)(t),\displaystyle X_{0}(t)Y_{0}(t)=\sum_{k+\ell\leq s}\sum_{1\leq j\leq J}x^{(k,\ell)}_{j}(t)\mathcal{W}^{(k,\ell)}_{j}(t),

and

X(t)Y(t)=k+si:λi<0ξi(k,)(t)i(k,)(t).\displaystyle X_{-}(t)Y_{-}(t)=\sum_{k+\ell\leq s}\sum_{i:\lambda_{i}<0}\xi^{(k,\ell)}_{i}(t)\mathcal{E}^{(k,\ell)}_{i}(t).

It holds that

X0=Y0,X+bXY.\displaystyle X_{0}^{\prime}=Y_{0},\quad X_{-}^{\prime}+bX_{-}\leq Y_{-}.

We now compare |X+(t)|2+|X0(t)|2+|X(t)|2|X_{+}(t)|^{2}+|X_{0}(t)|^{2}+|X_{-}(t)|^{2} and |Y+(t)|2+|Y0(t)|2+|Y(t)|2|Y_{+}(t)|^{2}+|Y_{0}(t)|^{2}+|Y_{-}(t)|^{2}.

|X+(t)|2+|X0(t)|2+|X(t)|2=k+sΣtkuL22(t).\displaystyle|X_{+}(t)|^{2}+|X_{0}(t)|^{2}+|X_{-}(t)|^{2}=\sum_{k+\ell\leq s}\|\mathcal{L}_{\Sigma}^{\ell}\partial^{k}_{t}u\|_{L^{2}}^{2}(t).

From the Cauchy-Schwarz inequality,

|Y+(t)|2+|Y0(t)|2+|Y(t)|2\displaystyle|Y_{+}(t)|^{2}+|Y_{0}(t)|^{2}+|Y_{-}(t)|^{2} k+si=1|ik,(t)|2=k+sΣtkE2(u)L22(t).\displaystyle\leq\sum_{k+\ell\leq s}\sum_{i=1}^{\infty}|\mathcal{E}^{k,\ell}_{i}(t)|^{2}=\sum_{k+\ell\leq s}\|\mathcal{L}_{\Sigma}^{\ell}\partial^{k}_{t}E_{2}(u)\|_{L^{2}}^{2}(t).

From Lemma 6.2,

k+sΣtkE2(u)L22(t)=o(1)k+sΣtkuL22(t).\displaystyle\sum_{k+\ell\leq s}\|\mathcal{L}_{\Sigma}^{\ell}\partial^{k}_{t}E_{2}(u)\|_{L^{2}}^{2}(t)=o(1)\sum_{k+\ell\leq s}\|\mathcal{L}_{\Sigma}^{\ell}\partial^{k}_{t}u\|_{L^{2}}^{2}(t).

Therefore, |Y+(t)|2+|Y0(t)|2+|Y(t)|2=o(1)(|X+(t)|2+|X0(t)|2+|X(t)|2).|Y_{+}(t)|^{2}+|Y_{0}(t)|^{2}+|Y_{-}(t)|^{2}=o(1)\big{(}|X_{+}(t)|^{2}+|X_{0}(t)|^{2}+|X_{-}(t)|^{2}\big{)}. We can then apply the ODE lemma, Lemma A.1. The slow decay assumption (6.1) rules out the possibility that X(t)X_{-}(t) dominates. Hence

(6.9) |X+(t)|2+|X0(t)|2+|X(t)|2=(1+o(1))|X0(t)|2.|X_{+}(t)|^{2}+|X_{0}(t)|^{2}+|X_{-}(t)|^{2}=(1+o(1))|X_{0}(t)|^{2}.

It remains to show that

(6.10) |X0(t)|2=(1+o(1))|x(t)|2.|X_{0}(t)|^{2}=(1+o(1))|x(t)|^{2}.

For 1\ell\geq 1,

xj(k,)=ΣΣtku,φι+j𝑑μ=Σtku,Σφι+j𝑑μ=0.\displaystyle x^{(k,\ell)}_{j}=\int_{\Sigma}\left\langle\mathcal{L}_{\Sigma}^{\ell}\partial^{k}_{t}u,\varphi_{\iota+j}\right\rangle\,d\mu=\int_{\Sigma}\left\langle\partial^{k}_{t}u,\mathcal{L}_{\Sigma}^{\ell}\varphi_{\iota+j}\right\rangle\,d\mu=0.

For k1k\geq 1,

xj(k,)=ddtzj(k1,)=𝒲j(k1,).\displaystyle x^{(k,\ell)}_{j}=\frac{d}{dt}z^{(k-1,\ell)}_{j}=\mathcal{W}^{(k-1,\ell)}_{j}.

These imply

|X0(t)|2\displaystyle|X_{0}(t)|^{2} |x(t)|2+ks1|𝒲(k,0)(t)|2\displaystyle\leq|x(t)|^{2}+\sum_{k\leq s-1}|\mathcal{W}^{(k,0)}(t)|^{2}
|x(t)|2+o(1)(|X+(t)|2+|X0(t)|2+|X(t)|2)\displaystyle\leq|x(t)|^{2}+o(1)\left(|X_{+}(t)|^{2}+|X_{0}(t)|^{2}+|X_{-}(t)|^{2}\right)
|x(t)|2+o(1)|X0(t)|2.\displaystyle\leq|x(t)|^{2}+o(1)|X_{0}(t)|^{2}.

We used (6.9) in the last inequality. Therefore (6.10) holds and the proof is finished. ∎

Corollary 6.4.

For any ss\in\mathbb{N}, there exists a positive constant C=C(u,s)C=C(u,s) such that uCs(t)C|x(t)|.\|u\|_{C^{s}}(t)\leq C|x(t)|.

The rest of the arguments are simpler than ones in the elliptic case. We omit the details and only provide the main steps. Their elliptic counterparts can be found in Subsection 5.2. Let

Q(t)=|x(t)|p+|x(t)|uC2(t)+|x(t)|u~C3(t).Q(t)=|x(t)|^{p}+|x(t)|\|u^{\prime}\|_{C^{2}}(t)+|x(t)|\|\tilde{u}^{\perp}\|_{C^{3}}(t).
Lemma 6.5 (c.f. Lemma 5.5).

There exists a positive constant C=C(u,Σ,N2)C=C(u,\mathcal{M}_{\Sigma},N_{2}) such that

|Σ(u)+f(x(t))Σu~|+|N2(u)|CQ(t).|\mathcal{M}_{\Sigma}(u)+\nabla f(x(t))-\mathcal{L}_{\Sigma}\tilde{u}^{\perp}|+|N_{2}(u)|\leq CQ(t).
Lemma 6.6 (c.f. Lemmas 5.6 and 5.9).

There exists a positive constant C=C(u,Σ,N2)C=C(u,\mathcal{M}_{\Sigma},N_{2}) such that

|x(t)+f(x(t))|+(u~)Σu~L2(t)CQ(t).|x^{\prime}(t)+\nabla f(x(t))|+\left\|\left(\tilde{u}^{\perp}\right)^{\prime}-\mathcal{L}_{\Sigma}\tilde{u}^{\perp}\right\|_{L^{2}}(t)\leq CQ(t).
Corollary 6.7 (c.f. Corollaries 5.7 and 5.10).

Suppose Q(t)M|x(t)|qQ(t)\leq M|x(t)|^{q} for some M,q>0M,q>0. Then there exists a positive constant C=C(u,Σ,N2,M,q)C=C(u,\mathcal{M}_{\Sigma},N_{2},M,q) such that |x(t)|C|x(t)|min(q,p1)|x^{\prime}(t)|\leq C|x(t)|^{\min(q,p-1)} and u~L2(t)C|x(t)|q\|\tilde{u}^{\perp}\|_{L^{2}}(t)\leq C|x(t)|^{q}.

Lemma 6.8 (c.f. Lemma 5.11).

For 1rp11\leq r\leq p-1, suppose the following holds.

|x(t)|+u~L2(t)M|x(t)|r.\displaystyle\left|x^{\prime}(t)\right|+\|\tilde{u}^{\perp}\|_{L^{2}}(t)\leq M|x(t)|^{r}.

Then for any ε(0,1)\varepsilon\in(0,1), there exists a positive constant C=C(u,Σ,N2,M,r,ε)C=C(u,\mathcal{M}_{\Sigma},N_{2},M,r,\varepsilon) such that

|x(t)|+u~L2(t)C|x(t)|min{r+1ε/2,p1}.\displaystyle\left|x^{\prime}(t)\right|+\|\tilde{u}^{\perp}\|_{L^{2}}(t)\leq C|x(t)|^{\min\{r+1-\varepsilon/2,p-1\}}.

Moreover,

|x(t)+f(x(t))|C|x(t)|r+1ε/2.\displaystyle\left|x^{\prime}(t)+\nabla f(x(t))\right|\leq C|x(t)|^{r+1-\varepsilon/2}.

With Lemma 6.8, Proposition 6.1 holds through a simple iteration argument.

7. Gradient flow

In this section, we study the gradient flow on Euclidean space with a perturbative vector field: let z(t)z(t) be a curve on J\mathbb{R}^{J} which satisfies

(7.1) z(t)+f(z(t))=G(t),z^{\prime}(t)+\nabla f(z(t))=G(t),

where f:Jf:\mathbb{R}^{J}\to\mathbb{R} is an analytic potential function and G(t)G(t) is a smooth vector field. Let ff satisfy the assumptions:

(7.2) f(0)=0,f(0)=0,2f(0)=0.f(0)=0,\quad\nabla f(0)=0,\quad\nabla^{2}f(0)=0.

For some positive integer p3p\geq 3, ff has an expansion at zero

f=jpfj,f=\sum_{j\geq p}f_{j},

where fjf_{j} is homogeneous polynomial of degree jj and fp0f_{p}\not\equiv 0. The restriction of fpf_{p} on 𝕊J1\mathbb{S}^{J-1} is denoted by f^p\hat{f}_{p}. We write ∇̸\not{\nabla} for the standard connection on 𝕊J1\mathbb{S}^{J-1}. Consider the critical points and critical values of f^p\hat{f}_{p} as

𝐂:={θ𝕊J1:∇̸f^p(θ)=0},𝐃:={f^p(θ):θ𝐂}.\displaystyle\mathbf{C}:=\left\{\theta\in\mathbb{S}^{J-1}\,:\,\not{\nabla}\hat{f}_{p}(\theta)=0\right\},\ \mathbf{D}:=\left\{\hat{f}_{p}(\theta)\,:\,\theta\in\mathbf{C}\right\}.

We further assume the perturbative vector field GG has a bound

(7.3) |G(t)|C|z(t)|pε,|G(t)|\leq C|z(t)|^{p-\varepsilon},

for some uniform constants ε(0,21)\varepsilon\in(0,2^{-1}) and C<C<\infty. The main theorem of this section concerns the secant direction when the solution converges to the origin.

Theorem 7.1.

Suppose limtz(t)=0\lim_{t\to\infty}z(t)=0. Then |z(t)|pf(z(t))|z(t)|^{-p}f(z(t)) converges to a non-negative critical value α0𝐃\alpha_{0}\in\mathbf{D}. Moreover, one of the following alternatives holds:

  1. (1)

    Suppose α0>0\alpha_{0}>0. Then the secant z(t)/|z(t)|z(t)/|z(t)| converges to a critical point θ𝐂\theta^{*}\in\mathbf{C} with f^p(θ)=α0\hat{f}_{p}(\theta^{*})=\alpha_{0}. Moreover, limtt1/(p2)|z(t)|=(α0p(p2))1/(p2).\lim_{t\to\infty}t^{{1}/(p-2)}|z(t)|=(\alpha_{0}p(p-2))^{-1/(p-2)}.

  2. (2)

    Suppose α0=0\alpha_{0}=0. Then limtdist(z(t)/|z(t)|,𝐂0)=0\lim_{t\to\infty}\mathrm{dist}(z(t)/|z(t)|,\mathbf{C}_{0})=0. Here 𝐂0\mathbf{C}_{0} is some connected component of 𝐂{θ:f^p(θ)=0}.\mathbf{C}\cap\left\{\theta:\hat{f}_{p}(\theta)=0\right\}. Moreover, limtt1/(p2)|z(t)|=\lim_{t\to\infty}t^{1/(p-2)}|z(t)|=\infty.

Remark 7.2.

The existence of non-negative critical value of 𝕊J1\mathbb{S}^{J-1} is a necessary condition for the flow z(t)z(t) to converge the the origin. Moreover, if the points in 𝐂0\mathbf{C}_{0} are isolated, then (2) implies the unique secant limit direction as well.

We are ready to prove Theorems 2.3 and 2.5.

Proof of Theorem 2.3.

From Proposition 5.1, Theorem 7.1 applies with the potential function being m1fm^{-1}f. Suppose case (1) in Theorem 7.1 occurs. This ensures z(t)/|z(t)|z(t)/|z(t)| converges to θ\theta^{*}, a critical point of f^p\hat{f}_{p} with m1f^p(θ)=α0>0m^{-1}\hat{f}_{p}(\theta^{*})=\alpha_{0}>0. Moreover,

limtt1/(p2)|z(t)|=(α0p(p2))1/(p2).\lim_{t\to\infty}t^{1/(p-2)}|z(t)|=\big{(}\alpha_{0}p(p-2)\big{)}^{-1/(p-2)}.

This implies

limtt1/(p2)u(t)L2=(α0p(p2))1/(p2).\lim_{t\to\infty}t^{1/(p-2)}\|u(t)\|_{L^{2}}=\big{(}\alpha_{0}p(p-2)\big{)}^{-1/(p-2)}.

Let wkerΣw\in\ker\mathcal{L}_{\Sigma} be the section corresponding to θ\theta^{*} through (2.8). Clearly u(t)/u(t)L2u(t)/\|u(t)\|_{L^{2}} converges to ww in L2(Σ;𝐕)L^{2}(\Sigma;\mathbf{V}). From Corollary 5.3, for any kk\in\mathbb{N}, u(t)/u(t)L2u(t)/\|u(t)\|_{L^{2}} is uniformly bounded in Ck(Σ;𝐕)C^{k}(\Sigma;\mathbf{V}). Therefore u(t)/u(t)L2u(t)/\|u(t)\|_{L^{2}} converges to ww in C(Σ;𝐕)C^{\infty}(\Sigma;\mathbf{V}). We then obtain case (1) in Theorem 2.3. The other possibility, case (2) in Theorem 7.1, leads to case (2) in Theorem 2.4. ∎

Proof of Theorem 2.5.

Theorem 2.5 can be obtained through replacing Proposition 5.1 in the above proof by Proposition 6.1. ∎

The rest of the section is devoted to proving Theorem 7.1. Let us write the problem in terms of the polar coordinates z=rθz=r\theta where r=|z|r=|z| and θ=z/|z|𝕊J1\theta=z/|z|\in\mathbb{S}^{J-1}. Because fpf_{p} is homogeneous of degree pp, fp(z)=rpf^p(θ)f_{p}(z)=r^{p}\hat{f}_{p}(\theta). We compute the gradient fp=prp1f^pr+rp2∇̸f^p\nabla f_{p}=pr^{p-1}\hat{f}_{p}\tfrac{\partial}{\partial r}+r^{p-2}\not{\nabla}\hat{f}_{p}. Let G(t)G^{\perp}(t) and GT(t)G^{T}(t) be the radial and tangential parts of G(t)G(t). Then equation (7.1) can be decomposed into the radial and tangential parts:

(7.4) {r(t)=rp1(t)(pf^p(θ(t))+R(t)),θ(t)=rp2(t)(∇̸f^p(θ(t))+RT(t)).\displaystyle\begin{cases}\begin{aligned} &r^{\prime}(t)=-r^{p-1}(t)\left(p\hat{f}_{p}(\theta(t))+R^{\perp}(t)\right),\\ &\theta^{\prime}(t)=-r^{p-2}(t)\left(\not{\nabla}\hat{f}_{p}(\theta(t))+R^{T}(t)\right).\end{aligned}\end{cases}

Here

R(t)=\displaystyle R^{\perp}(t)= rp+1(t)G(t)+jp+1jrjp(t)f^j,\displaystyle r^{-p+1}(t)G^{\perp}(t)+\sum_{j\geq p+1}jr^{j-p}(t)\hat{f}_{j},
RT(t)=\displaystyle R^{T}(t)= rp+1(t)GT(t)+jp+1rjp(t)∇̸f^j.\displaystyle r^{-p+1}(t)G^{T}(t)+\sum_{j\geq p+1}r^{j-p}(t)\not{\nabla}\hat{f}_{j}.

From (7.3), there exists a constant A>0A>0 such that

(7.5) |R(t)|+|RT(t)|Ar1ε(t).\displaystyle|R^{\perp}(t)|+|R^{T}(t)|\leq Ar^{1-\varepsilon}(t).
Lemma 7.3.

lim inft|∇̸f^p(θ(t))|=0.\liminf_{t\to\infty}|\not{\nabla}\hat{f}_{p}(\theta(t))|=0.

Proof.

It is convenient to work with σ(t)=r2p(t)\sigma(t)=r^{2-p}(t). The equation (7.4) becomes

(7.6) {σ(t)=(p2)(pf^p(θ(t))R(t))θ(t)=σ1(t)(∇̸f^p(θ(t))+RT(t)).\displaystyle\begin{cases}\begin{aligned} &\sigma^{\prime}(t)=(p-2)\left(p\hat{f}_{p}(\theta(t))-R^{\perp}(t)\right)\\ &\theta^{\prime}(t)=-\sigma^{-1}(t)\left(\not{\nabla}\hat{f}_{p}(\theta(t))+R^{T}(t)\right).\end{aligned}\end{cases}

The assumption limtr(t)=0\lim_{t\to\infty}r(t)=0 becomes limtσ(t)=\lim_{t\to\infty}\sigma(t)=\infty. Suppose lim inft|∇̸f^p(θ(t))|=c0>0.\liminf_{t\to\infty}|\not{\nabla}\hat{f}_{p}(\theta(t))|=c_{0}>0. From (7.5), there exists t0>0t_{0}>0 such that |RT(t)|41c0|R^{T}(t)|\leq 4^{-1}c_{0} and |∇̸f^p(θ(t))|21c0|\not{\nabla}\hat{f}_{p}(\theta(t))|\geq 2^{-1}c_{0} for all tt0t\geq t_{0}. This implies for all tt0t\geq t_{0},

ddtf^p(θ(t))=∇̸f^p(θ(t))θ(t)c024σ(t).\displaystyle\frac{d}{dt}\hat{f}_{p}(\theta(t))=\not{\nabla}\hat{f}_{p}(\theta(t))\cdot\theta^{\prime}(t)\leq-\frac{c_{0}^{2}}{4\sigma(t)}.

From the first equation in (7.6), we infer that σ\sigma grows at most linearly. As a result,

f^p(θ(t))f^p(θ(t0))+t0tc024σ(τ)dτ as t,\hat{f}_{p}(\theta(t))\leq\hat{f}_{p}(\theta(t_{0}))+\int_{t_{0}}^{t}-\frac{c_{0}^{2}}{4\sigma(\tau)}d\tau\to-\infty\text{ as }t\to\infty,

which is a contradiction. ∎

By Lemma 7.3, there exists a sequence tit_{i} goes to infinity such that |∇̸f^p(θ(ti))||\not{\nabla}\hat{f}_{p}(\theta(t_{i}))| goes to zero. Since 𝕊J1\mathbb{S}^{J-1} is compact, there exists a subsequence (still denoted by tit_{i}) such that θ(ti)\theta(t_{i}) converges to a critical point θ𝐂.\theta^{*}\in\mathbf{C}. Let α0=f^p(θ)\alpha_{0}=\hat{f}_{p}(\theta^{*}).

Lemma 7.4.

We have α00\alpha_{0}\geq 0 and limtf^p(θ(t))=α0\lim_{t\to\infty}\hat{f}_{p}(\theta(t))=\alpha_{0}.

Proof.

Suppose α0<0\alpha_{0}<0. From (7.5) and the first equation of (7.4), r(t)r(t) is strictly increasing for tt large. This contradicts the assumption limtz(t)=0\lim_{t\to\infty}z(t)=0.

To show the second assertion, let us assume f^p\hat{f}_{p} is not a constant as otherwise the assertion is trivial. By the Łojasiewicz gradient inequality [Łoj63], for each critical point θ0𝐂\theta_{0}\in\mathbf{C}, there exists a neighborhood UU such that for all θ𝐂U\theta\in\mathbf{C}\cap U, f^p(θ)=f^p(θ0)\hat{f}_{p}(\theta)=\hat{f}_{p}(\theta_{0}). Therefore, 𝐃\mathbf{D} is a finite set. Because f^p\hat{f}_{p} is not a constant, 𝐃\mathbf{D} has at least two distinct elements. Fix small ϵ0>0\epsilon_{0}>0 such that the critical values are separated at least by 6ε06\varepsilon_{0}. That is,

infα1,α2𝐃,α1α2|α1α2|6ϵ0.\inf_{\alpha_{1},\alpha_{2}\in\mathbf{D},\ \alpha_{1}\neq\alpha_{2}}|\alpha_{1}-\alpha_{2}|\geq 6\epsilon_{0}.

For every ϵ(0,ϵ0)\epsilon\in(0,\epsilon_{0}), let

δ(ε):=inf{|∇̸f^p(θ)|:|f(θ)α|εfor allα𝐃}.\displaystyle\delta(\varepsilon):=\inf\left\{|\not{\nabla}\hat{f}_{p}(\theta)|\,:\,|f(\theta)-\alpha|\geq\varepsilon\ \textup{for all}\ \alpha\in\mathbf{D}\right\}.

It is clear that δ(ε)>0\delta(\varepsilon)>0. Now assume limtf^p(θ(t))α0\lim_{t\to\infty}\hat{f}_{p}(\theta(t))\neq\alpha_{0}. There exist ϵ(0,ϵ0)\epsilon\in(0,\epsilon_{0}) and tit_{i}\to\infty such that |f^p(θ(ti))α0|2ε.|\hat{f}_{p}(\theta(t_{i}))-\alpha_{0}|\geq 2\varepsilon. Since f^p(θ(t))\hat{f}_{p}(\theta(t)) is a continuous and differentiable function of tt, there exist t~i\tilde{t}_{i}\to\infty such that

|f^p(θ(t~i))α0|=ε and ddtf^p(θ(t~i))0.|\hat{f}_{p}(\theta(\tilde{t}_{i}))-\alpha_{0}|=\varepsilon\quad\text{ and }\quad\frac{d}{dt}\hat{f}_{p}(\theta(\tilde{t}_{i}))\geq 0.

From (7.5), for ii large enough, |RT(t~i)|21δ(ε)|R^{T}(\tilde{t}_{i})|\leq 2^{-1}\delta(\varepsilon). Then we compute from (7.6)

(7.7) ddtf^p(θ(t~i))\displaystyle\frac{d}{dt}\hat{f}_{p}(\theta(\tilde{t}_{i})) =σ1(t~i)∇̸f^p(θ(t~i))(∇̸f^p(θ(t~i))+RT(t~i))\displaystyle=-\sigma^{-1}(\tilde{t}_{i})\not{\nabla}\hat{f}_{p}(\theta(\tilde{t}_{i}))\cdot\left(\not{\nabla}\hat{f}_{p}(\theta(\tilde{t}_{i}))+R^{T}(\tilde{t}_{i})\right)
21σ1(t~i)δ2(ε)<0.\displaystyle\leq-2^{-1}\sigma^{-1}(\tilde{t}_{i})\delta^{2}(\varepsilon)<0.

This is a contradiction. ∎

Lemma 7.5.

Let 𝐂0\mathbf{C}_{0} be the component of 𝐂\mathbf{C} containing θ\theta^{*}. Then

limtdist(θ(t),𝐂0)=0.\lim_{t\to\infty}\mathrm{dist}(\theta(t),\mathbf{C}_{0})=0.
Proof.

Let NδN_{\delta} denote the (open) δ\delta-neighborhood of 𝐂0\mathbf{C}_{0}. If the assertion is false, there is a small δ>0\delta>0 such that for any t00t_{0}\geq 0, there exists t1t0t_{1}\geq t_{0} such that θ(t1)N2δ\theta(t_{1})\notin N_{2\delta}. By taking δ>0\delta>0 small enough, we may assume N3δ𝐂=𝐂0N_{3\delta}\cap\mathbf{C}=\mathbf{C}_{0}. Set 𝐀=cl(N2δNδ)\mathbf{A}=\mathrm{cl}(N_{2\delta}\setminus N_{\delta}). Because θ\theta^{*} is a limit point of θ(t)\theta(t), there exist a sequence of strictly increasing times t1<t2<t_{1}<t_{2}<\ldots such that θ(t)𝐀\theta(t)\in\mathbf{A} for t[t2k1,t2k]t\in[t_{2k-1},t_{2k}], dist(θ(t2k1),𝐂0)=2δ\mathrm{dist}(\theta(t_{2k-1}),\mathbf{C}_{0})=2\delta, and dist(θ(t2k),𝐂0)=δ\mathrm{dist}(\theta(t_{2k}),\mathbf{C}_{0})=\delta for all positive integers kk. Since 𝐀\mathbf{A} is a closed set with 𝐀𝐂=\mathbf{A}\cap\mathbf{C}=\varnothing, there exists c>0c>0 such that |∇̸f^p(θ)|c|\not{\nabla}\hat{f}_{p}(\theta)|\geq c for all θ𝐀\theta\in\mathbf{A}. From (7.5), |RT(t)|21c|R^{T}(t)|\leq 2^{-1}c for tt¯t\geq\bar{t} some large t¯\bar{t}. This implies for any t[t2k1,t2k]t\in[t_{2k-1},t_{2k}] with t2k1t¯t_{2k-1}\geq\bar{t},

ddtf^p(θ(t))=\displaystyle-\frac{d}{dt}\hat{f}_{p}(\theta(t))= σ1(t)∇̸f^p(θ(t))(∇̸f^p(θ(t))+RT(t))21σ1(t)|∇̸f^p(θ(t))|2\displaystyle\sigma^{-1}(t)\not{\nabla}\hat{f}_{p}(\theta(t))\cdot\left(\not{\nabla}\hat{f}_{p}(\theta(t))+R^{T}(t)\right)\geq 2^{-1}\sigma^{-1}(t)|\not{\nabla}\hat{f}_{p}(\theta(t))|^{2}
\displaystyle\geq 31|∇̸f^p(θ(t))||θ(t)|31c|θ(t)|.\displaystyle 3^{-1}|\not{\nabla}\hat{f}_{p}(\theta(t))||\theta^{\prime}(t)|\geq 3^{-1}c|\theta^{\prime}(t)|.

Therefore,

f^p(θ(t2k))f^p(θ(t2k1))31ct2k1t2k|θ(t)|𝑑t31cδ.\displaystyle\hat{f}_{p}(\theta(t_{2k}))-\hat{f}_{p}(\theta(t_{2k-1}))\leq-3^{-1}c\int_{t_{2k-1}}^{t_{2k}}|\theta^{\prime}(t)|\,dt\leq-3^{-1}c\delta.

In particular, f^p(θ(t))\hat{f}_{p}(\theta(t)) does not converge. This contradicts to Lemma 7.4.

Proposition 7.6.

If α0>0\alpha_{0}>0, then limtθ(t)=θ\lim_{t\to\infty}\theta(t)=\theta^{*}. Moreover, the length of curve θ(t)\theta(t) on 𝕊J1\mathbb{S}^{J-1} is finite.

Proof.

Recall α0=f^p(θ).\alpha_{0}=\hat{f}_{p}(\theta^{*}). Consider the control function

g(t):=r(t)+(f^p(θ(t))f^p(θ)).g(t):=r(t)+(\hat{f}_{p}(\theta(t))-\hat{f}_{p}(\theta^{*})).

Our goal is to show the following. There exist t00t_{0}\geq 0, c0>0c_{0}>0, ρ0(0,1)\rho_{0}\in(0,1) and a neighborhood U0U_{0} of θ\theta^{*} such that g(t)0g^{\prime}(t)\leq 0, for tt0t\geq t_{0}, and moreover,

(7.8) g(t)c0|g(t)|ρ0|θ(t)| if tt0 and θ(t)U0.g^{\prime}(t)\leq-c_{0}|g(t)|^{\rho_{0}}|\theta^{\prime}(t)|\quad\text{ if }\ t\geq t_{0}\text{ and }\theta(t)\in U_{0}.

Let us first show this implies the proposition. First, since g(t)g(t) decreases monotonically to zero, g(t)0g(t)\geq 0 for tt0t\geq t_{0}. Suppose θ(t)U0\theta(t)\in U_{0} for t[t1,t2]t\in[t_{1},t_{2}] with t1t0t_{1}\geq t_{0}. By integrating (7.8),

t1t2|θ(t)|𝑑t|g(t1)|1ρ0|g(t2)|1ρ0c0(1ρ0)|g(t1)|1ρ0c0(1ρ0).\int_{t_{1}}^{t_{2}}|\theta^{\prime}(t)|dt\leq\frac{|g(t_{1})|^{1-\rho_{0}}-|g(t_{2})|^{1-\rho_{0}}}{c_{0}(1-\rho_{0})}\leq\frac{|g(t_{1})|^{1-\rho_{0}}}{c_{0}(1-\rho_{0})}.

This shows that if t1t0t_{1}\geq t_{0}, g(t1)g(t_{1}) is sufficiently small and θ(t1)\theta(t_{1}) is sufficiently close to θ\theta^{*} (we may find such t1t_{1} by Lemma 7.4 and limiθ(ti)=θ\lim_{i\to\infty}\theta(t_{i})=\theta^{*}), then θ(t)\theta(t) remains inside of U0U_{0} for all tt1t\geq t_{1} and the length of the curve θ(t)\theta(t) for t[t1,)t\in[t_{1},\infty) is finite. In particular, θ(t)\theta(t) converges as tt\to\infty and this proves the proposition.

It remains to prove g(t)0g^{\prime}(t)\leq 0 and (7.8). Let AA be the constant in (7.5). Fix t00t_{0}\geq 0 large so that for tt0t\geq t_{0},

(7.9) pf^p(θ(t))+R(t)21pf^p(θ),\displaystyle p\hat{f}_{p}(\theta(t))+R^{\perp}(t)\geq 2^{-1}p\hat{f}_{p}(\theta^{*}),
(7.10) r12ε(t)(160)1A2pf^p(θ).\displaystyle r^{1-2\varepsilon}(t)\leq(160)^{-1}A^{-2}p\hat{f}_{p}(\theta^{*}).

This is possible because of Lemma 7.4 and f^p(θ)>0\hat{f}_{p}(\theta^{*})>0. Observe from (7.4),

(7.11) g(t)=rp1(t)(pf^p(θ)+R)rp2(t)∇̸f^p(θ(t))(∇̸f^p(θ(t))+RT(t)).g^{\prime}(t)=-r^{p-1}(t)\left(p\hat{f}_{p}(\theta^{*})+R^{\perp}\right)-r^{p-2}(t)\not{\nabla}\hat{f}_{p}(\theta(t))\cdot\left(\not{\nabla}\hat{f}_{p}(\theta(t))+R^{T}(t)\right).

For tt0t\geq t_{0},

g(t)\displaystyle g^{\prime}(t)\leq 21prp1(t)f^p(θ)rp2(t)∇̸f^p(θ(t))(∇̸f^p(θ(t))+RT(t))by (7.9)\displaystyle-2^{-1}pr^{p-1}(t)\hat{f}_{p}(\theta^{*})-r^{p-2}(t)\not{\nabla}\hat{f}_{p}(\theta(t))\cdot\left(\not{\nabla}\hat{f}_{p}(\theta(t))+R^{T}(t)\right)\qquad\text{by \eqref{eq:t0_1}}
\displaystyle\leq 21prp1(t)f^p(θ)+41rp2(t)|RT(t)|2by Cauchy-Schwarz\displaystyle-2^{-1}pr^{p-1}(t)\hat{f}_{p}(\theta^{*})+4^{-1}r^{p-2}(t)|R^{T}(t)|^{2}\qquad\text{by Cauchy-Schwarz}
\displaystyle\leq 21prp1(t)f^p(θ)+41A2rp2ε(t)0by (7.5).\displaystyle-2^{-1}pr^{p-1}(t)\hat{f}_{p}(\theta^{*})+4^{-1}A^{2}r^{p-2\varepsilon}(t)\leq 0\qquad\text{by \eqref{eq:Rbound}}.

This shows g(t)0g^{\prime}(t)\leq 0.

Next, applying the Łojasiewicz gradient inequality [Łoj63] to f^p\hat{f}_{p} by viewing it as an analytic function |x|pfp(x)|x|^{-p}f_{p}(x) on J{0}\mathbb{R}^{J}\setminus\{0\}, we obtain a neighborhood U0U_{0} of θ\theta^{*} in 𝕊J1\mathbb{S}^{J-1}, ρ1(0,1)\rho_{1}\in(0,1) and c1>0c_{1}>0 such that if θU0\theta\in U_{0} then

(7.12) |∇̸f^p(θ)|c1|f^p(θ)f^p(θ)|ρ1.|\not{\nabla}\hat{f}_{p}(\theta)|\geq c_{1}|\hat{f}_{p}(\theta)-\hat{f}_{p}(\theta^{*})|^{\rho_{1}}.

Now we suppose θ(t)U0\theta(t)\in U_{0} for some tt0t\geq t_{0} and show that (7.8) holds. We divide the discussion into two cases depending on the term we utilize from the right hand side of (7.11).

Case 1. Suppose |∇̸f^p(θ(t))+RT(t)|4Ar1ϵ(t)|\not{\nabla}\hat{f}_{p}(\theta(t))+R^{T}(t)|\geq 4Ar^{1-\epsilon}(t). This implies

(7.13) |∇̸f^p(θ(t))|3Ar1ϵ(t)|\not{\nabla}\hat{f}_{p}(\theta(t))|\geq 3Ar^{1-\epsilon}(t)

and

(7.14) |∇̸f^p(θ(t))||RT(t)|21|∇̸f^p(θ(t))+RT(t)|=21r2p(t)|θ(t)|.\displaystyle|\not{\nabla}\hat{f}_{p}(\theta(t))|-|R^{T}(t)|\geq 2^{-1}|\not{\nabla}\hat{f}_{p}(\theta(t))+R^{T}(t)|=2^{-1}r^{2-p}(t)|\theta^{\prime}(t)|.

From (7.12) and (7.13),

(7.15) |∇̸f^p(θ(t))|21(c1|f^p(θ(t))f^p(θ)|ρ1+3Ar1ϵ(t)).\displaystyle|\not{\nabla}\hat{f}_{p}(\theta(t))|\geq 2^{-1}\left(c_{1}|\hat{f}_{p}(\theta(t))-\hat{f}_{p}(\theta^{*})|^{\rho_{1}}+3Ar^{1-\epsilon}(t)\right).

Combining (7.14) and (7.15), we derive

g(t)\displaystyle g^{\prime}(t)\leq rp2(t)|∇̸f^p(θ(t))|(|∇̸f^p(θ(t))||RT(t)|)\displaystyle-r^{p-2}(t)|\not{\nabla}\hat{f}_{p}(\theta(t))|\left(|\not{\nabla}\hat{f}_{p}(\theta(t))|-|R^{T}(t)|\right)
\displaystyle\leq 41(c1|f^p(θ(t))f^p(θ)|ρ1+3Ar1ϵ(t))|θ(t)|.\displaystyle-4^{-1}\left(c_{1}|\hat{f}_{p}(\theta(t))-\hat{f}_{p}(\theta^{*})|^{\rho_{1}}+3Ar^{1-\epsilon}(t)\right)|\theta^{\prime}(t)|.

As a result, for ρ2=max(ρ1,1ε)(0,1)\rho_{2}=\max(\rho_{1},1-\varepsilon)\in(0,1) and some c2>0c_{2}>0,

g(t)c2|g(t)|ρ2|θ(t)|.g^{\prime}(t)\leq-c_{2}|g(t)|^{\rho_{2}}|\theta^{\prime}(t)|.

Case 2. Suppose |∇̸f^p(θ(t))+RT(θ(t))|<4Ar1ε(t)|\not{\nabla}\hat{f}_{p}(\theta(t))+R^{T}(\theta(t))|<4Ar^{1-\varepsilon}(t). Then |θ(t)|4Arp1ε(t).|\theta^{\prime}(t)|\leq 4Ar^{p-1-\varepsilon}(t). This implies

21prp1(t)f^p(θ)81A1pf^p(θ)rε(t)|θ(t)|=4c3rε(t)|θ(t)|.\displaystyle-2^{-1}pr^{p-1}(t)\hat{f}_{p}(\theta^{*})\leq-8^{-1}A^{-1}p\hat{f}_{p}(\theta^{*})r^{\varepsilon}(t)|\theta^{\prime}(t)|=-4c_{3}r^{\varepsilon}(t)|\theta^{\prime}(t)|.

Here c3=(32)1A1pf^p(θ)c_{3}=(32)^{-1}A^{-1}p\hat{f}_{p}(\theta^{*}) is defined by the last equality. From (7.10),

(7.16) |∇̸f^p(θ(t))|5Ar1ϵ(t)c3rϵ(t).|\not{\nabla}\hat{f}_{p}(\theta(t))|\leq 5Ar^{1-{\epsilon}}(t)\leq c_{3}r^{{\epsilon}}(t).

Hence

g(t)21prp1(t)f^p(θ)+∇̸f^p(θ(t))θ(t)3c3rε(t)|θ(t)|.\displaystyle g^{\prime}(t)\leq-2^{-1}pr^{p-1}(t)\hat{f}_{p}(\theta^{*})+\not{\nabla}\hat{f}_{p}(\theta(t))\cdot\theta^{\prime}(t)\leq-3c_{3}r^{\varepsilon}(t)|\theta^{\prime}(t)|.

Moreover, from (7.16) and (7.12),

3c3rε(t)2c3rε(t)+|∇̸f^p(θ(t))|2c3rε(t)+c1|f^p(θ(t))f^p(θ)|ρ1c4|g(t)|ρ3\displaystyle 3c_{3}r^{\varepsilon}(t)\geq 2c_{3}r^{\varepsilon}(t)+|\not{\nabla}\hat{f}_{p}(\theta(t))|\geq 2c_{3}r^{\varepsilon}(t)+c_{1}|\hat{f}_{p}(\theta(t))-\hat{f}_{p}(\theta^{*})|^{\rho_{1}}\geq c_{4}|g(t)|^{\rho_{3}}

for some c4>0c_{4}>0 and ρ3=max(ε,ρ1)(0,1)\rho_{3}=\max(\varepsilon,\rho_{1})\in(0,1). The inequality (7.8) is obtained and this finishes the proof. ∎

Proof of Theorem 7.1.

By Lemma 7.4, there exists a critical point θ𝐂\theta^{*}\in\mathbf{C} such that

limt|z|pf(z)=f^p(θ)0.\displaystyle\lim_{t\to\infty}|z|^{-p}f(z)=\hat{f}_{p}(\theta^{*})\geq 0.

As previously, we denote α0=f^p(θ)\alpha_{0}=\hat{f}_{p}(\theta^{*}). If α0>0\alpha_{0}>0, then the convergence of z(t)/|z(t)|z(t)/|z(t)| is an immediate consequence of Proposition 7.6. Moreover, by solving the first line in (7.4) with known asymptotics, we obtain

limtt1/(p2)|z(t)|=(α0p(p2))1/(p2).\lim_{t\to\infty}t^{1/(p-2)}|z(t)|=(\alpha_{0}p(p-2))^{1/(p-2)}.

Suppose α0=0\alpha_{0}=0. Let 𝐂0\mathbf{C}_{0} be the connected component of 𝐂\mathbf{C} which contains θ\theta^{*}. Since there are finitely many critical values (see the proof of Lemma 7.4) and f^p(𝐂0)\hat{f}_{p}(\mathbf{C}_{0}) is connected, f^p(𝐂0)={0}\hat{f}_{p}(\mathbf{C}_{0})=\{0\}. In particular, 𝐂0\mathbf{C}_{0} is a connected component of 𝐂{θ:f^p(θ)=0}\mathbf{C}\cap\{\theta\,:\,\hat{f}_{p}(\theta)=0\}. The convergence of dist(z(t)/|z(t)|,𝐂0)\mathrm{dist}(z(t)/|z(t)|,\mathbf{C}_{0}) follows by Lemma 7.5 and the limit of t1/(p2)|z(t)|t^{1/(p-2)}|z(t)| follows by a similar argument.

Appendix A Tools

The following ODE lemma was proved in [MZ98].

Lemma A.1 (Merle-Zaag ODE lemma).

Let X+X_{+}, X0X_{0}, X:[0,)[0,)X_{-}:[0,\infty)\to[0,\infty) be absolutely continuous functions such that X+(t)+X0(t)+X(t)>0X_{+}(t)+X_{0}(t)+X_{-}(t)>0 for all t0t\geq 0 and lim inftX+(t)=0.\liminf_{t\to\infty}X_{+}(t)=0. Suppose there exist b>0b>0 and functions Y+Y_{+}, Y0Y_{0}, YY_{-} with

|Y+(t)|2+|Y0(t)|2+|Y(t)|2=o(1)(|X+(t)|2+|X0(t)|2+|X(t)|2)\displaystyle|Y_{+}(t)|^{2}+|Y_{0}(t)|^{2}+|Y_{-}(t)|^{2}=o(1)(|X_{+}(t)|^{2}+|X_{0}(t)|^{2}+|X_{-}(t)|^{2})

such that

(A.1) X+(t)bX+(t)Y+(t),|X0(t)|Y0(t),X(t)+bX(t)Y(t).\begin{array}[]{ccc}X_{+}^{\prime}(t)-bX_{+}(t)&\geq&Y_{+}(t),\\ |X_{0}^{\prime}(t)|&\leq&Y_{0}(t),\\ X_{-}^{\prime}(t)+bX_{-}(t)&\leq&Y_{-}(t).\end{array}

Then one of the following holds: either

(A.2) X+(t)+X(t)=o(1)X0(t),\displaystyle X_{+}(t)+X_{-}(t)=o(1)X_{0}(t),

or

(A.3) X+(t)+X0(t)=o(1)X(t).\displaystyle X_{+}(t)+X_{0}(t)=o(1)X_{-}(t).

Moreover, suppose (A.3) holds true. Then for all ε>0\varepsilon>0,

(A.4) lim supte(bε)t(X+(t)+X0(t)+X(t))=0.\displaystyle\limsup_{t\to\infty}e^{(b-\varepsilon)t}(X_{+}(t)+X_{0}(t)+X_{-}(t))=0.
Lemma A.2 (interpolation).

Suppose a function of single variable g(t)g(t) has bounds

supt[1,1]|djdtjg(t)|Ck\sup_{t\in[-1,1]}\left|\frac{d^{j}}{dt^{j}}g(t)\right|\leq C_{k}

for j=0,1,,kj=0,1,\ldots,k and

supt[1,1]|g(t)|C0.\sup_{t\in[-1,1]}|g(t)|\leq C_{0}.

Then, for every 0<k0\leq\ell<k, there holds a bound

|ddtg(0)|βC0αCk1α\left|\frac{d^{\ell}}{dt^{\ell}}g(0)\right|\leq\beta\cdot C_{0}^{\alpha}C_{k}^{1-\alpha}

for some α=α(,k)>0\alpha=\alpha(\ell,k)>0 and β=β(,k)>0\beta=\beta(\ell,k)>0. Moreover, α(,k)\alpha(\ell,k) converges to 11 as kk\to\infty while \ell is fixed.

Proof.

We prove the inequality with β1\beta\equiv 1 for functions which are compactly supported in the interior of (1,1)(-1,1). This is sufficient as one may multiply a general function by a cut-off and then apply the result.

By an argument which uses an induction and the integration by parts, we obtain for every <k\ell<k

tgL2gL21ktkgL2k.\|\partial^{\ell}_{t}g\|_{L^{2}}\leq\|g\|_{L^{2}}^{1-\frac{\ell}{k}}\|\partial^{k}_{t}g\|_{L^{2}}^{\frac{\ell}{k}}.
supt[1,1]|tg|t+1gL1t+1gL2gL21+1ktkgL2+1kC0αCk1α,\sup_{t\in[-1,1]}|\partial^{\ell}_{t}g|\leq\|\partial^{\ell+1}_{t}g\|_{L^{1}}\leq\|\partial^{\ell+1}_{t}g\|_{L^{2}}\leq\|g\|_{L^{2}}^{1-\frac{\ell+1}{k}}\|\partial^{k}_{t}g\|_{L^{2}}^{\frac{\ell+1}{k}}\leq C_{0}^{\alpha}C_{k}^{1-\alpha},

where α=1+1k\alpha=1-\frac{\ell+1}{k}. ∎

In the next lemma, we record the elliptic regularity in [Sim83, (1.13)].

Lemma A.3 (elliptic regularity).

There exists ρ0>0\rho_{0}>0 small depending on Σ\mathcal{M}_{\Sigma} such that the following holds. Let uu be a solution to (1.1) for t[T2,T+2]t\in[T-2,T+2]. Assume uC1(t)ρ0\|u\|_{C^{1}}(t)\leq\rho_{0} for t[T2,T+2]t\in[T-2,T+2]. Then for any ss\in\mathbb{N}, there exists a positive constant C=C(m,Σ,N1,s)C=C(m,\mathcal{M}_{\Sigma},N_{1},s) such that

supt[T1,T+1]uCs2(t)CT2T+2uL22(t)𝑑t.\displaystyle\sup_{t\in[T-1,T+1]}\|u\|^{2}_{C^{s}}(t)\leq C\int_{T-2}^{T+2}\|u\|_{L^{2}}^{2}(t)\,dt.

The parabolic regularity below is a minor extension of the one presented in [Sim83]. We include the proof for readers’ convenience.

Lemma A.4 (parabolic regularity).

Let uC(Q0,,𝐕~)u\in C^{\infty}(Q_{0,\infty},\widetilde{\mathbf{V}}) be a solution to (1.2). Assume limtuHn+4(t)=0\lim_{t\to\infty}\|u\|_{H^{n+4}}(t)=0. Then for all ss\in\mathbb{N}, limtuCs(t)=0\lim_{t\to\infty}\|u\|_{C^{s}}(t)=0.

Proof.

We claim that that for n+4\ell\geq n+4,

(A.5) limtuH(t)=0 implieslimtuH+1(t)=0.\lim_{t\to\infty}\|u\|_{H^{\ell}}(t)=0\ \textup{ implies}\ \lim_{t\to\infty}\|u\|_{H^{\ell+1}}(t)=0.

Suppose for a moment (A.5) holds true. Applying the Sobolev embedding on Σ\Sigma, u(t)u(t), as functions on Σ\Sigma, converges smoothly to zero. In view of (6.6), u(t)u^{\prime}(t) and higher order time derivatives also converge smoothly to zero.

We then turn to proving (A.5). From (6.6) and (2.5), we have the following energy estimate. For any \ell\in\mathbb{N}, there exists δ>0\delta_{\ell}>0 such that for t2>t1>0t_{2}>t_{1}>0 with t2t1δt_{2}-t_{1}\leq\delta_{\ell}, there holds that

(A.6) supt[t1,t2]uH(t)+t1t2uH+1(t)𝑑tCuH(t1)+Ct1t2E2(u)H1(t)𝑑t.\displaystyle\sup_{t\in[t_{1},t_{2}]}\|u\|_{H^{\ell}}(t)+\int_{t_{1}}^{t_{2}}\|u\|_{H^{\ell+1}}(t)\,dt\leq C_{\ell}\|u\|_{H^{\ell}}(t_{1})+C_{\ell}\int_{t_{1}}^{t_{2}}\|E_{2}(u)\|_{H^{\ell-1}}(t)\,dt.

We omit the derivation and instead refer readers to equation (4.3) in [Sim83]. Assume limtuH(t)=0\lim_{t\to\infty}\|u\|_{H^{\ell}}(t)=0. We claim that

(A.7) E2(u)H1(t)=o(1)uH+1(t)andE2(u)H(t)=o(1)uH+2(t).\|E_{2}(u)\|_{H^{\ell-1}}(t)=o(1)\|u\|_{H^{\ell+1}}(t)\ \textup{and}\ \|E_{2}(u)\|_{H^{\ell}}(t)=o(1)\|u\|_{H^{\ell+2}}(t).

Suppose for a moment (A.7) holds true. Fix δ=min(δ,21δ+1)\delta=\min(\delta_{\ell},2^{-1}\delta_{\ell+1}). Applying (A.6) and absorbing the E2(u)E_{2}(u) terms using (A.7), for t1t_{1} large enough,

t1t1+δuH+1(t)𝑑tCuH(t1),supt[t1,t1+2δ]uH+1(t)C+1uH+1(t1).\displaystyle\int_{t_{1}}^{t_{1}+\delta}\|u\|_{H^{\ell+1}}(t)\,dt\leq C_{\ell}\|u\|_{H^{\ell}}(t_{1}),\ \sup_{t\in[t_{1},t_{1}+2\delta]}\|u\|_{H^{\ell+1}}(t)\leq C_{\ell+1}\|u\|_{H^{\ell+1}}(t_{1}).

Then (A.5) follows.

It remains to prove (A.7). Recall that (6.7) E2(u)=j=02b2,j∇̸juE_{2}(u)=\sum_{j=0}^{2}b_{2,j}\cdot\not{\nabla}^{j}u, where b2,j=b2,j(ω,u,∇̸u)b_{2,j}=b_{2,j}(\omega,u,\not{\nabla}u) are smooth with b2,j(ω,0,0)=0b_{2,j}(\omega,0,0)=0. We present the proof for

(A.8) b2,2∇̸2uH(t)=o(1)uH+2(t).\displaystyle\left\|b_{2,2}\cdot\not{\nabla}^{2}u\right\|_{H^{\ell}}(t)=o(1)\|u\|_{H^{\ell+2}}(t).

The other terms can be treated similarly. For simplicity, we write b(ω,u,∇̸u)b(\omega,u,\not{\nabla}u) for b2,2(ω,u,∇̸u)b_{2,2}(\omega,u,\not{\nabla}u). For m0,m1,m20m_{0},m_{1},m_{2}\in\mathbb{N}_{0}, we write b(m0,m1,m2)b^{(m_{0},m_{1},m_{2})} for the partial derivative of bb of order (m0,m1,m2)(m_{0},m_{1},m_{2}). Fix 0\ell_{0}\leq\ell. Then the terms in the expansion of ∇̸0(b∇̸2u)\not{\nabla}^{\ell_{0}}\left(b\cdot\not{\nabla}^{2}u\right) are of the form

b(m0,m1,m2)(∇̸1u∇̸2u∇̸Nu),\displaystyle b^{(m_{0},m_{1},m_{2})}\cdot\left(\not{\nabla}^{\ell_{1}}u\ast\not{\nabla}^{\ell_{2}}u\ast\dots\not{\nabla}^{\ell_{N}}u\right),

where N=m1+m2+1N=m_{1}+m_{2}+1 and m0+i=1Ni=0+m2+2m_{0}+\sum_{i=1}^{N}\ell_{i}=\ell_{0}+m_{2}+2. Furthermore, i1\ell_{i}\geq 1 for all 1iN1\leq i\leq N, and there exist at least m2+1m_{2}+1 values of ii for which i2\ell_{i}\geq 2. We first discuss the case m1=m2=0m_{1}=m_{2}=0. Then the above becomes b(m0,0,0)∇̸0m0+2ub^{(m_{0},0,0)}\cdot\not{\nabla}^{\ell_{0}-m_{0}+2}u. From the Sobolev embedding,

|b(m0,0,0)|C(|u|+|∇̸u|)CuHn+4(t)=o(1).\displaystyle|b^{(m_{0},0,0)}|\leq C(|u|+|\not{\nabla}u|)\leq C\|u\|_{H^{n+4}}(t)=o(1).

Hence

b(m0,0,0)∇̸0m0+2uL2(t)=o(1)uH+2(t).\left\|b^{(m_{0},0,0)}\cdot\not{\nabla}^{\ell_{0}-m_{0}+2}u\right\|_{L^{2}}(t)=o(1)\left\|u\right\|_{H^{\ell+2}}(t).

Next, we consider the case m1+m21m_{1}+m_{2}\geq 1. It is then straightforward to check that the maximum of {i}1iN\{\ell_{i}\}_{1\leq i\leq N} is at most +2\ell+2 and other elements are at most +32=\lfloor\frac{\ell+3}{2}\rfloor=\ell_{*}. From n+4\ell\geq n+4 and the Sobolev embedding, ∇̸u(t)L(Σ)CuH(t)=o(1).\left\|\not{\nabla}^{\ell_{*}}u(t)\right\|_{L^{\infty}(\Sigma)}\leq C\left\|u\right\|_{H^{\ell}}(t)=o(1). This implies

b(m0,m1,m2)(∇̸1u∇̸2u∇̸Nu)L2(t)=o(1)uH+2(t),\displaystyle\left\|b^{(m_{0},m_{1},m_{2})}\cdot\left(\not{\nabla}^{\ell_{1}}u\ast\not{\nabla}^{\ell_{2}}u\ast\dots\not{\nabla}^{\ell_{N}}u\right)\right\|_{L^{2}}(t)=o(1)\left\|u\right\|_{H^{\ell+2}}(t),

and finishes the proof of (A.8). ∎

Appendix B Mean curvature flow

In this section, we show that (rescaled) mean curvature flows close to a stationary solution can be described by (1.2).

We begin with the mean curvature flow (MCF). Let Σ\Sigma be an embedded closed nn-dimensional submanifold in an n+kn+k-dimensional analytic ambient space (M,g¯)(M,\bar{g}). Let dμd\mu be the volume form induced from the metric on Σ\Sigma. Let 𝐕\mathbf{V} be the normal bundle of Σ\Sigma equipped with the inner product and the connection induced from MM. For ε>0\varepsilon>0, let 𝒰ε={u𝐕:|u|<ε}\mathcal{U}_{\varepsilon}=\{u\in\mathbf{V}:|u|<\varepsilon\} and let Nε(Σ)N_{\varepsilon}(\Sigma) be the ε\varepsilon-tubular neighborhood of Σ\Sigma in MM. For ε\varepsilon small enough, 𝒰ε\mathcal{U}_{\varepsilon} and Nε(Σ)N_{\varepsilon}(\Sigma) can be identified through the exponential map:

(B.1) exp:𝒰εNε(Σ).\displaystyle\exp:\mathcal{U}_{\varepsilon}\to N_{\varepsilon}(\Sigma).

Consider a local coordinate chart of 𝐕\mathbf{V} as {xi,yA}\{x^{i},y^{A}\}, 1in1\leq i\leq n, 1Ak1\leq A\leq k, where xix^{i} are coordinates of Σ\Sigma and yAy^{A} are fiber coordinates. We write (g0)ij(g_{0})_{ij} and (h0)AB(h_{0})_{AB} for the metric on Σ\Sigma and the inner product on 𝐕\mathbf{V}, respectively. Under the identification (B.1), the ambient metric g¯\bar{g} is of the form

g¯=gijdxidxj+hABdyAdyB+ciA(dxidyA+dyAdxi).\displaystyle\bar{g}=g_{ij}dx^{i}dx^{j}+h_{AB}dy^{A}dy^{B}+c_{iA}(dx^{i}dy^{A}+dy^{A}dx^{i}).

When yA=0y^{A}=0, gij=(g0)ijg_{ij}=(g_{0})_{ij}, hAB=(h0)ABh_{AB}=(h_{0})_{AB} and ciA=0c_{iA}=0. In particular, the induced volume form of Σ\Sigma is given by

dμ=detg0dx1dxn.\displaystyle d\mu=\sqrt{\det g_{0}}dx^{1}\wedge\dots\wedge dx^{n}.

Let uu be a section of 𝐕\mathbf{V} which is contained in 𝒰ε\mathcal{U}_{\varepsilon}. From (B.1), uu can be identified as a submanifold in Nε(Σ)N_{\varepsilon}(\Sigma) and we denote this submanifold by graph(u)\textup{graph}(u). The induced metric on graph(u)\textup{graph}(u) is

(gu)ij=gij+uAxiuBxjhAB+uAxicjA+uBxjciB.\displaystyle(g_{u})_{ij}=g_{ij}+\frac{\partial u^{A}}{\partial x^{i}}\frac{\partial u^{B}}{\partial x^{j}}h_{AB}+\frac{\partial u^{A}}{\partial x^{i}}c_{jA}+\frac{\partial u^{B}}{\partial x^{j}}c_{iB}.

Let (gu)ij(g_{u})^{ij} be the inverse metric of (gu)ij(g_{u})_{ij}. It is straightforward to check that

eA:=yA(gu)ij(ciA+uBxihBA)(xj+uCxjyC), 1Ak\displaystyle e_{A}:=\frac{\partial}{\partial y^{A}}-(g_{u})^{ij}\left(c_{iA}+\frac{\partial u^{B}}{\partial x^{i}}h_{BA}\right)\left(\frac{\partial}{\partial x^{j}}+\frac{\partial u^{C}}{\partial x^{j}}\frac{\partial}{\partial y^{C}}\right),\ 1\leq A\leq k

form a basis for the normal space of graph(u)(u). Let (hu)AB=g¯(eA,eB)(h_{u})_{AB}=\bar{g}(e_{A},e_{B}) and (hu)AB(h_{u})^{AB} be its inverse metric.

Now we compare the mean curvature vector and the Euler-Lagrange operator of the area functional. The (local) area functional Σ(u)\mathcal{F}_{\Sigma}(u) is given by

Σ(u)=detgu𝑑x1dxn.\displaystyle\mathcal{F}_{\Sigma}(u)=\int\sqrt{\det g_{u}}\,dx^{1}\wedge\dots\wedge dx^{n}.

Denote by H\vec{H} the mean curvature vector of graph(u)\textup{graph}(u). Let ξ\xi be a smooth section of 𝐕\mathbf{V}. From the first variational formula,

ddsΣ(u+sξ)|s=0=g¯(H,ξAyA)detgu𝑑x1dxn.\displaystyle\frac{d}{ds}\mathcal{F}_{\Sigma}(u+s\xi)\big{|}_{s=0}=-\int\bar{g}\left(\vec{H},\xi^{A}\frac{\partial}{\partial y^{A}}\right)\sqrt{\det g_{u}}\,dx^{1}\wedge\dots\wedge dx^{n}.

Comparing the above with

ddsΣ(u+sξ)|s=0=h0(Σ(u),ξ)detg0𝑑x1dxn,\displaystyle\frac{d}{ds}\mathcal{F}_{\Sigma}(u+s\xi)\big{|}_{s=0}=-\int h_{0}\left(\mathcal{M}_{\Sigma}(u),\xi\right)\sqrt{\det g_{0}}\,dx^{1}\wedge\dots\wedge dx^{n},

we have

(B.2) Σ(u)A=g¯(H,yB)(h0)BAdetgudetg0.\mathcal{M}_{\Sigma}(u)^{A}=\bar{g}\left(\vec{H},\frac{\partial}{\partial y^{B}}\right)(h_{0})^{BA}\sqrt{\frac{\det g_{u}}{\det g_{0}}}.

In particular, Σ(0)=0\mathcal{M}_{\Sigma}(0)=0 if Σ\Sigma is a minimal submanifold.

Next, we compute the non-parametric form of the MCF. In this case, the speed ηA=uAt\eta^{A}=\frac{\partial u^{A}}{\partial t} is determined by ηAg¯(yA,eB)=g¯(H,eB)\eta^{A}\bar{g}(\frac{\partial}{\partial y^{A}},e_{B})=\bar{g}(\vec{H},e_{B}) for all 1Bk1\leq B\leq k. From g¯(yA,eB)=(hu)AB\bar{g}(\frac{\partial}{\partial y^{A}},e_{B})=(h_{u})_{AB} and g¯(H,eB)=g¯(H,yB)\bar{g}(\vec{H},e_{B})=\bar{g}(\vec{H},\frac{\partial}{\partial y^{B}}), we have ηA=g¯(H,yB)(hu)BA\eta^{A}=\bar{g}\left(\vec{H},\frac{\partial}{\partial y^{B}}\right)(h_{u})^{BA}. Hence the non-parametric form of the MCF is given by

(B.3) uAt=g¯(H,yB)(hu)BA.\displaystyle\frac{\partial u^{A}}{\partial t}=\bar{g}\left(\vec{H},\frac{\partial}{\partial y^{B}}\right)(h_{u})^{BA}.

Using (B.2), (B.3) is equivalent to

(B.4) uAtΣ(u)A=(detg0detgu(hu)AB(h0)BCδCA)Σ(u)C.\displaystyle\frac{\partial u^{A}}{\partial t}-\mathcal{M}_{\Sigma}(u)^{A}=\left(\sqrt{\frac{\det g_{0}}{\det g_{u}}}(h_{u})^{AB}(h_{0})_{BC}-\delta_{C}^{A}\right)\cdot\mathcal{M}_{\Sigma}(u)^{C}.

In view of (2.6), (B.4) is of the form (1.2).

Next, we consider the rescaled MCF close to a shrinker. Let Σ\Sigma be an nn-dimensional embedded closed submanifold in n+k\mathbb{R}^{n+k}. We continue to use notation introduced above. For a section uu of the normal bundle 𝐕\mathbf{V}, we write X(u)X(u) for the position vector. In particular, X(0)X(0) stands for the position vector of Σ\Sigma. Let dμd\mu be the volume form with a Gaussian weight. Namely,

dμ=e|X(0)|2/4detg0dx1dxn.\displaystyle d\mu=e^{-|X(0)|^{2}/4}\sqrt{\det g_{0}}dx^{1}\wedge\dots\wedge dx^{n}.

For a general section uu, the (local) Gaussian area functional is

Σ(u)=e|X(u)|2/4detgu𝑑x1dxn.\displaystyle\mathcal{F}_{\Sigma}(u)=\int e^{-|X(u)|^{2}/4}\sqrt{\det g_{u}}dx^{1}\wedge\dots\wedge dx^{n}.

Let ξ\xi be a smooth section of 𝐕\mathbf{V}. From the first variational formula,

ddsΣ(u+sξ)|s=0=H+X(u)2,ξAyAe|X(u)|2/4detgu𝑑x1dxn,\displaystyle\frac{d}{ds}\mathcal{F}_{\Sigma}(u+s\xi)\big{|}_{s=0}=-\int\left\langle\vec{H}+\frac{X^{\perp}(u)}{2},\xi^{A}\frac{\partial}{\partial y^{A}}\right\rangle e^{-|X(u)|^{2}/4}\sqrt{\det g_{u}}\,dx^{1}\wedge\dots\wedge dx^{n},

where X(u)X^{\perp}(u) stands for the normal part of X(u)X(u). A similar argument as above yields

Σ(u)A=H+X(u)2,yB(h0)BAe|X(u)|2/4+|X(0)|2/4detgudetg0.\displaystyle\mathcal{M}_{\Sigma}(u)^{A}=\left\langle\vec{H}+\frac{X^{\perp}(u)}{2},\frac{\partial}{\partial y^{B}}\right\rangle(h_{0})^{BA}e^{-|X(u)|^{2}/4+|X(0)|^{2}/4}\sqrt{\frac{\det g_{u}}{\det g_{0}}}.

Moreover, the non-parametric form the the rescaled MCF is given by

uAtΣ(u)A=(e|X(u)|2/4|X(0)|2/4detg0detgu(hu)AB(h0)BCδCA)Σ(u)C.\displaystyle\frac{\partial u^{A}}{\partial t}-\mathcal{M}_{\Sigma}(u)^{A}=\left(e^{|X(u)|^{2}/4-|X(0)|^{2}/4}\sqrt{\frac{\det g_{0}}{\det g_{u}}}(h_{u})^{AB}(h_{0})_{BC}-\delta_{C}^{A}\right)\cdot\mathcal{M}_{\Sigma}(u)^{C}.

It is then clear that the above is of the form (1.2).

Acknowledgement

The first author has been partially supported by National Research Foundation of Korea grant No. 2022R1C1C1013511 and POSTECH Basic Science Research Institute grant No. 2021R1A6A1A10042944.

References

  • [AS88] D. Adams and L. Simon. Rates of asymptotic convergence near isolated singularities of geometric extrema. Indiana Univ. Math. J., 37(2):225–254, 1988.
  • [BCS22] W. Du R. Haslhofer B. Choi, P. Daskalopoulos and N. Sesum. Classification of bubble-sheet ovals in 4\mathbb{R}^{4}. arXiv preprint arXiv:2209.04931, 2022.
  • [BDGG69] E. Bombieri, E. De Giorgi, and E. Giusti. Minimal cones and the Bernstein problem. Invent. Math., 7:243–268, 1969.
  • [CCK21] B. Choi, K. Choi, and S. Kim. Translating surfaces under flows by sub-affine-critical powers of gauss curvature. arXiv preprint arXiv:2104.13186, 2021.
  • [CCR15] A. Carlotto, O. Chodosh, and Y.A. Rubinstein. Slowly converging Yamabe flows. Geom. Topol., 19(3):1523–1568, 2015.
  • [Che18] G. Chen. Rate of asymptotic convergence near an isolated singularity of G2G_{2} manifold. J. Geom. Anal., 28:3139–3170, 2018.
  • [CM15] T.H. Colding and W.P. Minicozzi. Uniqueness of blowups and łojasiewicz inequalities. Ann. of Math., 182(1):221–285, 2015.
  • [CM18] T.H. Colding and W.P. Minicozzi. Regularity of the level set flow. Comm. Pure Appl. Math., 71(4):814–824, 2018.
  • [CM22] K. Choi and C. Mantoulidis. Ancient gradient flows of elliptic functionals and Morse index. Amer. J. Math., 144(2):541–573, 2022.
  • [CP18] X. Cabré and G. Poggesi. Stable solutions to some elliptic problems: minimal cones, the Allen-Cahn equation, and blow-up solutions. In Geometry of PDEs and related problems, volume 2220 of Lecture Notes in Math., pages 1–45. Springer, Cham, 2018.
  • [CS20] K. Choi and L. Sun. Classification of ancient flows by sub-affine-critical powers of curvature in classification of ancient flows by sub-affine-critical powers of curvature in 2\mathbb{R}^{2}. arXiv preprint arXiv:2012.01727, 2020.
  • [DH21] W. Du and R. Haslhofer. Hearing the shape of ancient noncollapsed flows in 4\mathbb{R}^{4}. arXiv preprint arXiv:2107.04443 (to appear in CPAM), 2021.
  • [Gan21] Z. Gang. On the mean convexity of a space-and-time neighborhood of generic singularities formed by mean curvature flow. J. Geom. Anal., 31(10):9819–9890, 2021.
  • [HS85] R. Hardt and L. Simon. Area minimizing hypersurfaces with isolated singularities. J. Reine Angew. Math., 362:102–129, 1985.
  • [KKP00] T. Mostowski K. Kurdyka and A. Parusinski. Proof of the gradient conjecture of r. thom. Annals of Mathematics, 152(3):763–792, 2000.
  • [KS06] R.V. Kohn and S. Serfaty. A deterministic-control-based approach to motion by curvature. Comm. Pure Appl. Math., 59(3):344–407, 2006.
  • [Łoj63] S. Łojasiewicz. Une propriété topologique des sous-ensembles analytiques réels. In Les Équations aux Dérivées Partielles (Paris, 1962), pages 87–89. Éditions du Centre National de la Recherche Scientifique (CNRS), Paris, 1963.
  • [MZ98] F. Merle and H. Zaag. Optimal estimates for blowup rate and behavior for nonlinear heat equations. Comm. Pure Appl. Math., 51(2):139–196, 1998.
  • [SAS20] P. Daskalopoulos S. Angenent and N. Sesum. Uniqueness of two-convex closed ancient solutions to the mean curvature flow. Ann. of Math., 192(2):353–436, 2020.
  • [Ses08] N. Sesum. Rate of convergence of the mean curvature flow. Comm. Pure Appl. Math., 61(4):464–485, 2008.
  • [Sim83] L. Simon. Asymptotics for a class of nonlinear evolution equations, with applications to geometric problems. Ann. of Math. (2), 118(3):525–571, 1983.
  • [Sim85] L. Simon. Isolated singularities of extrema of geometric variational problems. In Harmonic mappings and minimal immersions (Montecatini, 1984), volume 1161 of Lecture Notes in Math., pages 206–277. Springer, Berlin, 1985.
  • [Sim96] L. Simon. Theorems on regularity and singularity of energy minimizing maps. Lectures in Mathematics ETH Zürich. Birkhäuser Verlag, Basel, 1996. Based on lecture notes by Norbert Hungerbühler.
  • [SS86] L. Simon and B. Solomon. Minimal hypersurfaces asymptotic to quadratic cones in 𝐑n+1{\bf R}^{n+1}. Invent. Math., 86(3):535–551, 1986.
  • [Str20] N. Strehlke. A unique continuation property for the level set equation. Int. Math. Res. Not., (16):4843–4851, 2020.
  • [SX22] A. Sun and J. Xue. Generic mean curvature flows with cylindrical singularities. arXiv preprint arXiv:2210.00419, 2022.
  • [Tho89] R. Thom. Problèmes rencontrés dans mon parcours mathématique: un bilan. Inst. Hautes Études Sci. Publ. Math., (70):199–214 (1990), 1989.