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

ancient finite entropy flows by powers of curvature in 2\mathbb{R}^{2}

Kyeongsu Choi School of Mathematics, Korea Institute for Advanced Study, 85 Hoegiro, Dongdaemun-gu, Seoul 02455, Republic of Korea. [email protected]  and  Liming Sun Department of Mathematics, University of British Columbia, Vancouver, BC, V6T 1Z2, CA. [email protected]
Abstract.

We show the existence of non-homothetic ancient flows by powers of curvature embedded in 2\mathbb{R}^{2} whose entropy is finite. We determine the Morse indices and kernels of the linearized operator of shrinkers to the flows, and construct ancient flows by using unstable eigenfunctions of the linearized operator.

2010 Mathematics Subject Classification:
Primary 53C44, 53A04; Secondary 35K55

1. Introduction

Given α>0\alpha>0, the α\alpha-curve shortening flow (α\alpha-CSF) is a family of complete convex curves Γt\Gamma_{t} embedded in 2\mathbb{R}^{2} which evolves by the α\alpha-power-of-curvature. Namely, the position vector 𝐗(,t)\mathbf{X}(\cdot,t) of Γt\Gamma_{t} satisfies

𝐗t(p,t)=κα(p,t)𝐍(p,t),\displaystyle\frac{\partial\mathbf{X}}{\partial t}(p,t)=\kappa^{\alpha}(p,t)\mathbf{N}(p,t), (1.1)

where κ\kappa is the curvature and 𝐍\mathbf{N} is inward pointing unit normal vector of Γt\Gamma_{t}.


We say that a flow Γt\Gamma_{t} is ancient if it exists for t(,T)t\in(-\infty,T) for some T{+}T\in\mathbb{R}\cup\{+\infty\}. Geometric flows satisfy parabolic equations so that there are in general only a few number of ancient flows. For example, Wang [26] showed that a closed convex embedded ancient curve shortening flow (CSF)111Curve shortening flow means the α\alpha-CSF with α=1\alpha=1. sweeping the entire plane is a shrinking circle, and Daskalopoulos-Hamilton-Sesum [22] showed that a closed convex embedded ancient CSF is a shrinking circle or an Angenent oval.222It looks like a shortening paper clip sweeping a slab. See also Bourni, Langford, and Tinaglia [8] for the classification of non-compact ones.

Ancient flows have been intensively studied in the mean curvature flow, a higher dimensional version of the CSF. In particular, ancient mean curvature flows are useful to investigate singularities. See [6, 7, 11, 12, 19, 20, 16] (c.f. Ricci flow [10, 5, 13]).

The α\alpha-CSF is a fully nonlinear flow, which behaves like the α\alpha-Gauss curvature flow in many aspects. In particular, if α=13\alpha=\frac{1}{3} (α=1n+2\alpha=\frac{1}{n+2} in higher dimensions) then a α\alpha-CSF remains a α\alpha-CSF under any affine transform (of determinant one) of the ambient space. The affine normal flow (13\frac{1}{3}-CSF) have been widely studied due to its beauty from affine geometry. For example, Chen [15] showed that an ancient closed convex affine normal flow must be a shrinking ellipse (see an alternate proof by Ivaki [23]). See also [24] for higher dimensions.


Andrews, Guan, and Ni [3] introduced an important notion of entropy for α\alpha-CSF. We recall that the support function uz0u_{z_{0}} with respect to z02z_{0}\in\mathbb{R}^{2} is

uz0(θ):=maxzΩ(cosθ,sinθ),zz0,\displaystyle u_{z_{0}}(\theta):=\max_{z\in\Omega}\langle(\cos\theta,\sin\theta),z-z_{0}\rangle,

and the entropy α(Ω)\mathcal{E}_{\alpha}(\Omega) of a bounded convex region Ω2\Omega\subset\mathbb{R}^{2} and its boundary Ω\partial\Omega is defined by

α(Ω)=α(Ω)=supz0Ωα(Ω,z0),\mathcal{E}_{\alpha}(\partial\Omega)=\mathcal{E}_{\alpha}(\Omega)=\sup_{z_{0}\in\Omega}\mathcal{E}_{\alpha}(\Omega,z_{0}), (1.2)

where α(Ω,z0)\mathcal{E}_{\alpha}(\Omega,z_{0}) is

α(Ω,z0)={αα1log(𝕊1uz011α(θ)𝑑θ)12log|Ω|πif α1,𝕊1loguz0(θ)𝑑θ12log|Ω|πif α=1.\displaystyle\mathcal{E}_{\alpha}(\Omega,z_{0})=\begin{cases}\frac{\alpha}{\alpha-1}\log\left(\fint_{\mathbb{S}^{1}}u_{z_{0}}^{1-\frac{1}{\alpha}}(\theta)d\theta\right)-\frac{1}{2}\log\frac{|\Omega|}{\pi}&\text{if }\alpha\neq 1,\\ \fint_{\mathbb{S}^{1}}\log u_{z_{0}}(\theta)d\theta-\frac{1}{2}\log\frac{|\Omega|}{\pi}&\text{if }\alpha=1.\end{cases} (1.3)

Here |Ω||\Omega| denotes the area of it.

In [3], they showed that the entropy α(Γt)\mathcal{E}_{\alpha}(\Gamma_{t}) of the α\alpha-CSF decreases with respect to tt. Hence, we say that an ancient α\alpha-CSF has finite entropy if

limtα(Γt)<+.\lim_{t\to-\infty}\mathcal{E}_{\alpha}(\Gamma_{t})<+\infty. (1.4)

Clearly, self-shrinking ancient solutions has finite entropy, since the entropy does not change under homothetic transformation. However, every non-homothetic ancient α\alpha-CSF discovered in previous researches including [4] and [9] do not have finite entropy. See also [17] for a higher dimensional analogue. Indeed, the entropy of every non-homothetic ancient α\alpha-CSF with α(23,1]\alpha\in(\frac{2}{3},1] must diverge by [22] and [9]. In this paper, we present families of non-homothetic closed ancient α\alpha-CSFs which converge to a self-shrinker333If Γt=(t)1α+1Γ1\Gamma_{t}=(-t)^{\frac{1}{\alpha+1}}\Gamma_{-1} is the α\alpha-CSF, then we call Γ1\Gamma_{-1} a self-shrinker or a shrinker. as tt\to-\infty after rescaling. Then, their entropy is less than that of the limiting shrinker, namely the ancient flows have the finite entropy. See Theorem 3.2.


To construct ancient flows asymptotic to a self-shrinking ancient flow, we first recall the classification result of self-shrinkers.

Theorem 1.1 (Andrews [2]).

If α[18,+)\{1/3}\alpha\in[\frac{1}{8},+\infty)\backslash\{1/3\}, then the shrinker of (1.1) is a circle (denote it as Γαc\Gamma_{\alpha}^{c}). If α=13\alpha=\frac{1}{3}, then a shrinker is an ellipse. If α(0,18)\alpha\in(0,\frac{1}{8}), then a shrinker is a circle or a curve Γαk{\Gamma_{\alpha}^{k}} with kk-fold symmetry, where 3k3\leq k\in\mathbb{N} with k<1+1/αk<\sqrt{1+1/\alpha}. The curves Γαk{\Gamma_{\alpha}^{k}} depend smoothly on α<1k21\alpha<\frac{1}{k^{2}-1} and converge to regular kk-sided poloygons as α0\alpha\searrow 0 and to circles as α1k21\alpha\nearrow\frac{1}{k^{2}-1}. See Table 1 and Figure 1 for illustrations.

α\alpha Γαc and Γαk\Gamma_{\alpha}^{c}\text{ and }{\Gamma_{\alpha}^{k}}
[18,+)\13[\frac{1}{8},+\infty)\backslash\frac{1}{3} [Uncaptioned image]
[115,18)[\frac{1}{15},\frac{1}{8}) [Uncaptioned image][Uncaptioned image]
[124,115)[\frac{1}{24},\frac{1}{15}) [Uncaptioned image][Uncaptioned image][Uncaptioned image]
\cdots \cdots
Table 1. Enumeration of shrinkers for different α\alpha.
Refer to caption
Refer to caption
Refer to caption
Figure 1. The shape of Γαk{\Gamma_{\alpha}^{k}} (normalized by (1.7)) when k=3k=3, α=19,116,1100\alpha=\frac{1}{9},\frac{1}{16},\frac{1}{100} from left to right.

To fix the asymptotic self-shrinking ancient flow, we consider the normalized flow Γ¯τ\bar{\Gamma}_{\tau} defined by

𝑿¯(p,τ)=(1+α)1α+1eτ𝑿(p,e(1+α)τ),\bar{\boldsymbol{X}}(p,\tau)=(1+\alpha)^{-\frac{1}{\alpha+1}}e^{\tau}\boldsymbol{X}(p,-e^{-(1+\alpha)\tau}), (1.5)

By Proposition 2.1 the support function u¯(θ,τ)\bar{u}(\theta,\tau) of 𝑿¯\bar{\boldsymbol{X}} with respect to the origin satisfies

u¯τ=(u¯θθ+u¯)α+u¯.\displaystyle\bar{u}_{\tau}=-(\bar{u}_{\theta\theta}+\bar{u})^{\alpha}+\bar{u}. (1.6)

Hence, the support function hh of a self-shrinker Γ\Gamma with respect to the origin satisfies

hθθ+h=h1/α,\displaystyle h_{\theta\theta}+h=h^{-1/\alpha}, (1.7)

and thus the difference v=u¯hv=\bar{u}-h satisfies

vτ=(hθθ+h+vθθ+v)α+(h+v):=Γ(v)+EΓ(v).\displaystyle v_{\tau}=-{(h_{\theta\theta}+h+v_{\theta\theta}+v)^{-\alpha}}+(h+v):=\mathcal{L}_{\Gamma}(v)+E_{\Gamma}(v). (1.8)

Here Γ\mathcal{L}_{\Gamma} is the linearization of the above equation at v=0v=0

Γ(v):=αh1+1α(vθθ+v)+v\displaystyle\mathcal{L}_{\Gamma}(v):={\alpha h^{1+\frac{1}{\alpha}}}(v_{\theta\theta}+v)+v (1.9)

and

|EΓ(v)|C|vθθ+v|2,|E_{\Gamma}(v)|\leq C|v_{\theta\theta}+v|^{2}, (1.10)

for small enough vθθ+vv_{\theta\theta}+v. See Proposition 3.3 for details.

It is easy to see that the Jacobi operator Γ\mathcal{L}_{\Gamma} is a self-adjoint operator on the space Lh2(𝕊1)={f:𝕊1f2h11/α<}L_{h}^{2}(\mathbb{S}^{1})=\{f:\int_{\mathbb{S}^{1}}f^{2}h^{-1-1/\alpha}<\infty\}, and thus it has a sequence of eigenvalues and eigenfunctions which form the basis of Lh2(𝕊1)L_{h}^{2}(\mathbb{S}^{1}). We are able to characterize its kernel and Morse index444The dimension of negative space of -\mathcal{L}. as follows.

Theorem 1.2 (cf. Proposition 2.2 and Theorem 2.4).

Suppose 0<α130<\alpha\neq\frac{1}{3}.

  1. (1)

    The Morse index of Γ¯αk\mathcal{L}_{\bar{\Gamma}_{\alpha}^{k}} is 2k12k-1, and kerΓ¯αk=span{hθ}\ker\mathcal{L}_{\bar{\Gamma}_{\alpha}^{k}}=span\{h_{\theta}\}, where hh is the support function of Γ¯αk\bar{\Gamma}_{\alpha}^{k}.

  2. (2)

    The Morse index of Γ¯αc\mathcal{L}_{\bar{\Gamma}_{\alpha}^{c}} is 21+1/α12\lceil\sqrt{1+1/\alpha}\rceil-1.555x\lceil x\rceil denotes least integer greater than or equal to xx. If α=1k21\alpha=\frac{1}{k^{2}-1}, then kerΓ¯αc=span{coskθ,sinkθ}\ker\mathcal{L}_{\bar{\Gamma}_{\alpha}^{c}}=span\{\cos k\theta,\sin k\theta\}. Otherwise kerΓ¯αc=\ker\mathcal{L}_{\bar{\Gamma}_{\alpha}^{c}}=\emptyset.

The center manifold theory in functional analysis provides the existence of an II-parameter family of ancient solutions to a class of fully nonlinear parabolic equations, where II is the Morse index. See Lunardi [25, Chapter 9]. However, using the contraction mapping method, we can show the existence of such ancient solutions and even including sharp asymptotic behaviors of the solutions with layer structures. See Choi and Mantoulidis [18] and Caffarelli-Hardt-Simon [14] for quasilinear parabolic and elliptic PDEs. Here comes the second main theorem of our paper.

Theorem 1.3 (cf. Theorem 3.2).

Let α13\alpha\neq\frac{1}{3} and λ1λI<0\lambda_{1}\leq\cdots\leq\lambda_{I}<0 denote the negative eigenvalues of Γ\mathcal{L}_{\Gamma} where Γ=Γ¯αk\Gamma=\bar{\Gamma}_{\alpha}^{k} or Γ¯αc\bar{\Gamma}_{\alpha}^{c} and II is the Morse index. There exists β(0,1)\beta\in(0,1), ε0>0\varepsilon_{0}>0 and an injective continuous map 𝒮:Bε0(0)(I3)C2,β(𝕊1×(,1])\mathcal{S}:B_{\varepsilon_{0}}(0)(\subset\mathbb{R}^{I-3})\to C^{2,\beta}(\mathbb{S}^{1}\times(-\infty,-1]) such that for each 𝐚=(a1,,aI3)I3\boldsymbol{a}=(a_{1},\cdots,a_{I-3})\in\mathbb{R}^{I-3} the image v=𝒮(𝐚)v=\mathcal{S}(\boldsymbol{a}) is an ancient solution to (1.8). Moreover, if  3<kI3<k\leq I and 𝐚,𝐛I3\boldsymbol{a},\boldsymbol{b}\in\mathbb{R}^{I-3} satisfy ak3bk30a_{k-3}-b_{k-3}\neq 0 and ajbj=0a_{j}-b_{j}=0 for all j>k3j>k-3, then 𝒮\mathcal{S} satisfies

𝒮(𝒂)(θ,τ)𝒮(𝒃)(θ,τ)=(ak3bk3)eλkτφk(θ)+o(eλkτ)\mathcal{S}(\boldsymbol{a})(\theta,\tau)-\mathcal{S}(\boldsymbol{b})(\theta,\tau)=(a_{k-3}-b_{k-3})e^{-\lambda_{k}\tau}\varphi_{k}(\theta)+o(e^{-\lambda_{k}\tau}) (1.11)

when λk1<λk\lambda_{k-1}<\lambda_{k}, and

𝒮(𝒂)(θ,τ)𝒮(𝒃)(θ,τ)=eλkτi=k1k(ai3bi3)φi(θ)+o(eλkτ)\mathcal{S}(\boldsymbol{a})(\theta,\tau)-\mathcal{S}(\boldsymbol{b})(\theta,\tau)\\ =e^{-\lambda_{k}\tau}\sum_{i=k-1}^{k}(a_{i-3}-b_{i-3})\varphi_{i}(\theta)+o(e^{-\lambda_{k}\tau}) (1.12)

when λk1=λk\lambda_{k-1}=\lambda_{k}, where φi\varphi_{i} are eigenfunctions of Γ\mathcal{L}_{\Gamma} with the eigenvalue λi\lambda_{i} and φi,φjLh2=δij\langle\varphi_{i},\varphi_{j}\rangle_{L^{2}_{h}}=\delta_{ij}. In particular, 𝒮(𝟎)(θ,τ)=h(θ)\mathcal{S}(\boldsymbol{0})(\theta,\tau)=h(\theta) corresponds to the shrinker.

Remark 1.4.

Notice that the first three eigenfunctions of Γ\mathcal{L}_{\Gamma} are h,cosθ,sinθh,\cos\theta,\sin\theta by Proposition 2.2, which accounts for dilations and transitions of the non-rescaled α\alpha-CSF. See Proposition 3.4. Therefore, we consider (I3)(I-3)-parameter family of ancient solutions rather than II-parameter.

Moreover, if Γ=Γαc\Gamma=\Gamma_{\alpha}^{c}, then rotations accounts for 11-parameter. Namely, Theorem 1.3 provides a (I4)(I-4)-parameter family ancient flows converging to a round shrinking circle up to rigid motions and dilations.

In short, given 1k21α<1(k1)21\frac{1}{k^{2}-1}\leq\alpha<\frac{1}{(k-1)^{2}-1} with 3k3\leq k\in\mathbb{N}, by Theorem 1.3 there exist, up to rigid motions and dilations, a (2k5)(2k-5)-parameter family of closed convex ancient α\alpha-CSFs converging to a round shrinking circle and a (2m3)(2m-3)-parameter family of closed convex ancient α\alpha-CSFs converging to a shrinking mm-fold symmetric curve for each integer 3m<k3\leq m<k.

In an following paper, the authors will classify ancient finite entropy α\alpha-CSFs, and show that the solutions in Theorem 1.3 are the all solutions up to transitions and dilations with exhibiting the layer structure (1.11)-(1.12).

An outline of our paper is in order. In Section 2, we devote to studying the spectrum of the linear operator \mathcal{L}. In Section 3, we construct ancient solutions converge with finite entropy by contraction mapping theorem.

Acknowledgements. The authors are grateful to Christos Mantoulidis for fruitful discussion, and also thankful to Shibing Chen, Beomjun Choi, John Loftin, and Mohammad N. Ivaki for their comments and suggestions. K. Choi is supported by KIAS Individual Grant MG078901.

2. Spectra of linearized operators

We begin by deriving the evolution equation of the support function u¯\bar{u} of the normalized flow Γ¯τ\bar{\Gamma}_{\tau} given by (1.5).

Proposition 2.1.

Let Γ¯τ\bar{\Gamma}_{\tau} be a normalized α\alpha-CSF satisfying (1.5). The support function u¯\bar{u} of Γ¯τ\bar{\Gamma}_{\tau} satisfies

u¯τ=κ¯α+u¯=(u¯θθ+u¯)α+u¯.\displaystyle\bar{u}_{\tau}=-\bar{\kappa}^{\alpha}+\bar{u}=-(\bar{u}_{\theta\theta}+\bar{u})^{-\alpha}+\bar{u}. (2.1)
Proof.

By using (1.5), we have

κ¯=(1+α)11+αeτκ,\displaystyle\bar{\kappa}=(1+\alpha)^{\frac{1}{1+\alpha}}e^{-\tau}\kappa, (2.2)

and thus

τ𝑿¯=(1+α)α1+αeατtX+X¯=κ¯α𝐍+𝑿¯.\displaystyle\partial_{\tau}\bar{\boldsymbol{X}}=(1+\alpha)^{\frac{\alpha}{1+\alpha}}e^{-\alpha\tau}\partial_{t}X+\bar{X}={\bar{\kappa}^{\alpha}}\mathbf{N}+\bar{\boldsymbol{X}}. (2.3)

Therefore, u¯=𝑿¯,𝐍\bar{u}=\langle\bar{\boldsymbol{X}},\mathbf{N}\rangle and u¯θθ+u¯=κ¯1\bar{u}_{\theta\theta}+\bar{u}=\bar{\kappa}^{-1} yield the desired evolution equation. ∎

We are interested in normalized ancient flows Γ¯τ\bar{\Gamma}_{\tau} converging to a shrinker Γ=Γ¯αk\Gamma=\bar{\Gamma}_{\alpha}^{k} or Γ¯αc\bar{\Gamma}_{\alpha}^{c} as τ\tau\to-\infty. Namely, the difference v=u¯hv=\bar{u}-h converges to zero, where hh is the suppose function of Γ\Gamma satisfying (1.7). Moreover, the evolution equation (1.8) has the linearized operator \mathcal{L} given by (1.9).

Γ=αh1+1α(θ2+1)+1.\mathcal{L}_{\Gamma}=\alpha h^{1+\frac{1}{\alpha}}(\partial_{\theta}^{2}+1)+1. (2.4)

here hh is the support function of Γ\Gamma. We shall abbreviate Γ\mathcal{L}_{\Gamma} as \mathcal{L} whenever there is no confusion.

We introduce the space Lh2(𝕊1)=L2(𝕊1,h11/αdθ)L^{2}_{h}(\mathbb{S}^{1})=L^{2}(\mathbb{S}^{1},h^{-1-1/\alpha}d\theta) with norm fh2=𝕊1f2h11/α||f||_{h}^{2}=\int_{\mathbb{S}^{1}}f^{2}h^{-1-1/\alpha}. It is equipped with the inner product

(f,g)h=𝕊1fgh11α𝑑θ.\displaystyle(f,g)_{h}=\int_{\mathbb{S}^{1}}fgh^{-1-\frac{1}{\alpha}}d\theta. (2.5)

Since h>0h>0 on 𝕊1\mathbb{S}^{1} and (1.7), this norm is equivalent to the standard L2L^{2} norm.

It is easy to see that \mathcal{L} is a self-adjoint operator on Lh2L^{2}_{h}. Since \mathcal{L} is an elliptic operator on a compact space, thus -\mathcal{L} has a sequence of eigenvalues λ1λ2\lambda_{1}\leq\lambda_{2}\leq\cdots. We remind that an eigenfunction φLh2(𝕊1)\varphi\in L_{h}^{2}(\mathbb{S}^{1}) and the corresponding eigenvalue λ\lambda\in\mathbb{R} satisfy

αh1+1α(φθθ+φ)+(λ+1)φ=0, on 𝕊1.\displaystyle{\alpha h^{1+\frac{1}{\alpha}}}(\varphi_{\theta\theta}+\varphi)+(\lambda+1)\varphi=0,\quad\text{ on }\quad\mathbb{S}^{1}. (2.6)

Moreover, there exists a sequence of the pairs (λi,φi)(\lambda_{i},\varphi_{i}) of eigenvalues and eigenfunctions such that λiλi+1\lambda_{i}\leq\lambda_{i+1}, limiλi=+\lim_{i\to\infty}\lambda_{i}=+\infty, (φi,φj)h=δij(\varphi_{i},\varphi_{j})_{h}=\delta_{ij}, and span{φ1,φ2,}=Lh2(𝕊1)\text{span}\{\varphi_{1},\varphi_{2},\cdots\}=L_{h}^{2}(\mathbb{S}^{1}).

In this section, we will study eigenfunctions with negative or zero eigenvalues of -\mathcal{L}.

Proposition 2.2.

There are some known eigenvalues for -\mathcal{L}.

  1. (1)

    λ=1α\lambda=-1-\alpha is an eigenvalue with the eigenfunction φ=h\varphi=h. Since hh is always positive, λ=1α\lambda=-1-\alpha is the lowest eigenvalue.

  2. (2)

    λ=1\lambda=-1 is an eigenvalue with the eigenfunctions φ=sinθ,cosθ\varphi=\sin\theta,\cos\theta.

  3. (3)

    Γ¯αc=α(θ2+1)+1\mathcal{L}_{\bar{\Gamma}_{\alpha}^{c}}=\alpha(\partial^{2}_{\theta}+1)+1 has eigenvalues

    λ1=α1,λ2l=λ2l+1=α(l21)1,l1\displaystyle\lambda_{1}=-\alpha-1,\quad\lambda_{2l}=\lambda_{2l+1}=\alpha(l^{2}-1)-1,\quad l\geq 1 (2.7)

    with the eigenfunctions cos(lθ)\cos(l\theta) and sin(lθ)\sin(l\theta). Notice that -\mathcal{L} has an eigenvalue λ=0\lambda=0 only when α=1/(l21)\alpha=1/({l^{2}-1}) for some l2l\geq 2.

  4. (4)

    Γ¯αk\mathcal{L}_{\bar{\Gamma}_{\alpha}^{k}} has zero eigenvalue λ=0\lambda=0 with eigenfunction φ=hθ\varphi=h_{\theta}. More importantly, λ=0\lambda=0 is simple.

Proof.

(1), (2), (3) are easy to verify. For (4), it is obtained by differentiating (1.7) with respect to θ\theta, which gives hθh_{\theta} satisfies (2.6) when λ=0\lambda=0. Indeed, hθh_{\theta} arises from rotations of Γ¯αk{\bar{\Gamma}_{\alpha}^{k}}. Andrews [2, Lemma 7.3] shows that the eigenspace of λ=0\lambda=0 has dimension ONE, which is span{hθ}span\{h_{\theta}\}. ∎

In Proposition 2.2, we characterize all eigenfunctions of Γ¯αc-\mathcal{L}_{\bar{\Gamma}_{\alpha}^{c}} and neutral eigenfunctions of Γ¯αk-\mathcal{L}_{\bar{\Gamma}_{\alpha}^{k}}. Thus, we will focus on Γ=Γ¯αk\Gamma=\bar{\Gamma}_{\alpha}^{k} and consider negative eigenvalues of Γ¯αk-\mathcal{L}_{\bar{\Gamma}_{\alpha}^{k}}. We shall simply write =Γ¯αk\mathcal{L}=\mathcal{L}_{\bar{\Gamma}_{\alpha}^{k}} for the rest of this section.

The following lemma is equivalent to [1, Lemma 5] whose proof needs Brunn-Minkowski inequality there. We give a direct proof here.

Lemma 2.3.

There is NO eigenvalue of -\mathcal{L} in (1α,1)(-1-\alpha,-1).

Proof.

Suppose φ\varphi is an eigenfunction of -\mathcal{L} satisfying (φ,h)h=0(\varphi,h)_{h}=0 and (2.6). Then there exists cc such that φ~=φch\tilde{\varphi}=\varphi-ch satisfy 𝕊1φ~=0\int_{\mathbb{S}^{1}}\tilde{\varphi}=0. Then 𝕊1φ~2|φ~θ|20\int_{\mathbb{S}^{1}}\tilde{\varphi}^{2}-|\tilde{\varphi}_{\theta}|^{2}\leq 0. Multiplying (2.6) by φ~h11α\tilde{\varphi}\,h^{-1-\frac{1}{\alpha}} and integrating over 𝕊1\mathbb{S}^{1} give

α𝕊1(φθθ+φ)φ~+(λ+1)𝕊1φφ~h11α=0.\displaystyle\alpha\int_{\mathbb{S}^{1}}(\varphi_{\theta\theta}+\varphi)\tilde{\varphi}+(\lambda+1)\int_{\mathbb{S}^{1}}\varphi\tilde{\varphi}\,h^{-1-\frac{1}{\alpha}}=0. (2.8)

Let us simplify the left-hand side. First, using the fact 𝕊1φh1α=0\int_{\mathbb{S}^{1}}\varphi\,h^{-\frac{1}{\alpha}}=0, we have

𝕊1φφ~h11α=𝕊1φ2h11α=(φ,φ)h>0.\int_{\mathbb{S}^{1}}\varphi\tilde{\varphi}h^{-1-\frac{1}{\alpha}}=\int_{\mathbb{S}^{1}}\varphi^{2}h^{-1-\frac{1}{\alpha}}=(\varphi,\varphi)_{h}>0.

Second,

𝕊1(φθθ+φ)φ~=\displaystyle\int_{\mathbb{S}^{1}}(\varphi_{\theta\theta}+\varphi)\tilde{\varphi}= 𝕊1[φ~θθ+φ~]φ~+c𝕊1[hθθ+h]φ~\displaystyle\int_{\mathbb{S}^{1}}[\tilde{\varphi}_{\theta\theta}+\tilde{\varphi}]\tilde{\varphi}+c\int_{\mathbb{S}^{1}}[h_{\theta\theta}+h]\tilde{\varphi}
=\displaystyle= 𝕊1(φ~2φ~θ2)+c𝕊1h1α(φch)c2𝕊1h11α0,\displaystyle\int_{\mathbb{S}^{1}}(\tilde{\varphi}^{2}-\tilde{\varphi}_{\theta}^{2})+c\int_{\mathbb{S}^{1}}h^{-\frac{1}{\alpha}}(\varphi-ch)\leq-c^{2}\int_{\mathbb{S}^{1}}h^{-1-\frac{1}{\alpha}}\leq 0,

where in the second equality we used (1.7).

If λ(1α,1)\lambda\in(-1-\alpha,-1) is an eigenvalue of -\mathcal{L}, inserting the above two inequalities into the LHS of (2.8), one could find out the LHS <0<0. Contradiction. ∎

It follows from Proposition 2.2 and Lemma 2.3 that -\mathcal{L} has eigenvalues

λ1<λ2=λ3<λ4\displaystyle\lambda_{1}<\lambda_{2}=\lambda_{3}<\lambda_{4}\leq\cdots (2.9)

where λ1=1α\lambda_{1}=-1-\alpha, λ2=λ3=1\lambda_{2}=\lambda_{3}=-1.

Theorem 2.4.

Suppose k3k\geq 3, α(0,1/(k21))\alpha\in(0,{1}/({k^{2}-1})). The negative eigenspace of Γ¯αk-\mathcal{L}_{\bar{\Gamma}_{\alpha}^{k}} has dimension 2k12k-1. In particular,

  1. (1)

    If kk is odd, every negative eigenvalue except λ1\lambda_{1} has the eigenspace of dimension two. If kk is even, every negative eigenvalue except λ1,λk,λk+1\lambda_{1},\lambda_{k},\lambda_{k+1} has the eigenspace of dimension two. In both cases, any eigenfunction of λ2l\lambda_{2l} and λ2l+1\lambda_{2l+1}, 1lk11\leq l\leq k-1, have 2l2l zeros.

  2. (2)

    Furthermore, λ2k=0\lambda_{2k}=0 and λ2k+1>0\lambda_{2k+1}>0 are simple, namely

    λ2k1<λ2k(=0)<λ2k+1<λ2k+2.\lambda_{2k-1}<\lambda_{2k}(=0)<\lambda_{2k+1}<\lambda_{2k+2}\leq\cdots.

    In addition, both φ2k\varphi_{2k} and φ2k+1\varphi_{2k+1} have 2k2k nodal sets.

Easily one can see the dimension of eigenspace of each eigenvalue is at most 22. This is because (2.6) is a second order ODE and it has at most two linearly independent solutions.

For a function φ\varphi, the term zeros (or nodal sets) refers to the set {θ:φ(θ)=0}\{\theta:\varphi(\theta)=0\}. The term nodal domain refers to the connected components of the complement of the nodal sets.

Lemma 2.5.

Each eigenfunction φ\varphi satisfies φ2+|φ|2>0\varphi^{2}+|\varphi^{\prime}|^{2}>0. Any eigenfunctions φ\varphi, except φspan{h}\varphi\in span\{h\}, has even number of zeros and even number of nodal domains.

Proof.

If φ(θ0)=φ(θ0)=0\varphi(\theta_{0})=\varphi^{\prime}(\theta_{0})=0 at some point θ0𝕊1\theta_{0}\in\mathbb{S}^{1}, then we have φ=0\varphi=0 on 𝕊1\mathbb{S}^{1} by the uniqueness of solution to the second order ODE. Hence, φ2+|φ|2>0\varphi^{2}+|\varphi^{\prime}|^{2}>0 everywhere. Therefore, if φ(θ0)=0\varphi(\theta_{0})=0 at some θ0𝕊1\theta_{0}\in\mathbb{S}^{1} then φ(θ0+ϵ)φ(θ0ϵ)<0\varphi(\theta_{0}+\epsilon)\varphi(\theta_{0}-\epsilon)<0 for small enough ϵ\epsilon. Namely, φ\varphi change its signs at zeros. Hence, φ\varphi has even number of zeros and thus it has even number of nodal domains. ∎

Lemma 2.6.

Suppose that λ\lambda has a two dimensional eigenspace. Then, its eigenfunctions have the same number of nodal sets.

Proof.

This follows from the Sturm separation theorem. For reader’s convenience, we give a proof. Suppose span{φ,ψ}span\{\varphi,\psi\} are the eigenspace of λ\lambda, where φ,ψ\varphi,\psi are linearly independent. Then the Wronskian of φ\varphi and ψ\psi is not zero for any θ\theta, that is

W[φ,ψ](θ)=|φ(θ)ψ(θ)φ(θ)ψ(θ)|0.\displaystyle W[\varphi,\psi](\theta)=\begin{vmatrix}\varphi(\theta)&\psi(\theta)\\ \varphi^{\prime}(\theta)&\psi^{\prime}(\theta)\end{vmatrix}\neq 0. (2.10)

If ψ\psi and φ\varphi has different number of nodal sets, by Lemma 2.5, then without loss of generality one nodal set φ\varphi is strictly contained in one nodal set of ψ\psi. That is, there exists θ1\theta_{1} and θ2\theta_{2} such that

φ(θ1)=φ(θ2)=0,φ(θ1)φ(θ2)<0,ψ(θ1)ψ(θ2)>0.\varphi(\theta_{1})=\varphi(\theta_{2})=0,\quad\varphi^{\prime}(\theta_{1})\varphi^{\prime}(\theta_{2})<0,\quad\psi(\theta_{1})\psi(\theta_{2})>0.

However, we will get

W(θ1)W(θ2)=φ(θ1)ψ(θ1)φ(θ2)ψ(θ2)<0.W(\theta_{1})W(\theta_{2})=\varphi^{\prime}(\theta_{1})\psi(\theta_{1})\varphi^{\prime}(\theta_{2})\psi(\theta_{2})<0.

This is not possible, because WW does not change sign. ∎

Because the above lemma, we are eligible to say the number of nodal sets corresponding to an eigenvalue λ\lambda.

Lemma 2.7.

Suppose an eigenvalue λi\lambda_{i} has a two dimensional eigenspace, then for any λj>λi\lambda_{j}>\lambda_{i}, the number of nodal sets of the eigenfunctions corresponding to λj\lambda_{j} is greater than that of λi\lambda_{i}. Similarly, if λj<λi\lambda_{j}<\lambda_{i}, then the number of nodal sets corresponding to λj\lambda_{j} is less than that of λi\lambda_{i}.

Proof.

Suppose the the eigenspace of λi\lambda_{i} is span{ψi,ψ~i}span\{\psi_{i},\tilde{\psi}_{i}\} and take any eigenfunction φj\varphi_{j} corresponding to λj\lambda_{j}. Assume φj(θ0)=0\varphi_{j}(\theta_{0})=0 for some θ0\theta_{0}, One can find ω[0,2π]\omega\in[0,2\pi] such that (cosω)ψi(θ0)+(sinω)ψ~i(θ0)=0(\cos\omega)\psi_{i}(\theta_{0})+(\sin\omega)\tilde{\psi}_{i}(\theta_{0})=0. Notice φi=(cosω)ψi+(sinω)ψ~i\varphi_{i}=(\cos\omega)\psi_{i}+(\sin\omega)\tilde{\psi}_{i} is an eigenfunction of λi\lambda_{i}. Since φi(θ0)=φj(θ0)=0\varphi_{i}(\theta_{0})=\varphi_{j}(\theta_{0})=0, the conclusions follow from the Sturm-Picone comparison theorem. For reader’s convenience, we sketch the proof of the case λj>λi\lambda_{j}>\lambda_{i}. The other case is similar. It suffices to prove that there is at least a zero of φj\varphi_{j} which lies strictly between any two consecutive zeros of φi\varphi_{i}. Assume that φi\varphi_{i} has two consecutive zeros aa and bb, and φj\varphi_{j} has no zero in (a,b)(a,b). Without loss of generality, we assume φi>0\varphi_{i}>0 and φj>0\varphi_{j}>0 in (a,b)(a,b).

Denote the Wronskian W(θ)=φiφjφiφjW(\theta)=\varphi_{i}\varphi_{j}^{\prime}-\varphi_{i}^{\prime}\varphi_{j}, then we can directly calculate W=(λiλj)1αh11/αφjφiW^{\prime}=(\lambda_{i}-\lambda_{j})\frac{1}{\alpha}h^{-1-1/\alpha}\varphi_{j}\varphi_{i}. We have the following Picone’s identity

(φiφjW)=1αh11α(λiλj)φi2(Wφj)2\displaystyle\left(\frac{\varphi_{i}}{\varphi_{j}}W\right)^{\prime}=\frac{1}{\alpha}h^{-1-\frac{1}{\alpha}}(\lambda_{i}-\lambda_{j})\varphi_{i}^{2}-\left(\frac{W}{\varphi_{j}}\right)^{2} (2.11)

wherever φj0\varphi_{j}\neq 0.

If φj(a)>0\varphi_{j}(a)>0 and φj(b)>0\varphi_{j}(b)>0, then we integrate (2.11) from aa to bb.

φiφjW|ab<0.\frac{\varphi_{i}}{\varphi_{j}}W\Big{|}_{a}^{b}<0.

This contradicts to φi(a)=φi(b)=0\varphi_{i}(a)=\varphi_{i}(b)=0.

If φj(a)=0\varphi_{j}(a)=0 and φj(b)>0\varphi_{j}(b)>0, then φj(a)>0\varphi_{j}^{\prime}(a)>0. Integrating (2.11) from aϵa-\epsilon to bb for ϵ>0\epsilon>0 small enough, we obtain

φiφjW|aϵb<0\frac{\varphi_{i}}{\varphi_{j}}W\Big{|}_{a-\epsilon}^{b}<0

However, we know φi(aϵ)>0\varphi_{i}(a-\epsilon)>0, φj(aϵ)>0\varphi_{j}(a-\epsilon)>0 and W(aϵ)<0W(a-\epsilon)<0 for sufficiently small ϵ>0\epsilon>0. Namely, the above quantity is positive. Hence it is an obvious contradiction.

The case of φj(a)>0\varphi_{j}(a)>0 with φj(b)=0\varphi_{j}(b)=0 and φj(a)=φj(b)=0\varphi_{j}(a)=\varphi_{j}(b)=0 can be ruled out in the same manner. ∎

Recall that the support function hh of Γ¯kα\bar{\Gamma}_{k}^{\alpha} is kk-fold symmetric. It has 2k2k critical points. By rotating Γ¯kα\bar{\Gamma}_{k}^{\alpha}, we may assume

h(nπ/k)=0,h^{\prime}(n\pi/k)=0, (2.12)

for all nn\in\mathbb{Z}. Then hh has even reflection symmetry with respect to nπ/kn\pi/k for any nn\in\mathbb{Z}, namely h(θ)=h(2nπ/kθ)h(\theta)=h(2n\pi/k-\theta) for any nn\in\mathbb{Z}.

Lemma 2.8.

Suppose that the eigenspace of λiλ1\lambda_{i}\neq\lambda_{1} is span{φi}span\{\varphi_{i}\}, namely λi\lambda_{i} is simple. Then, either φi\varphi_{i} has at least 2k2k zeros, or φi\varphi_{i} has exactly kk zeros and kk must be even. In the second case, zeros of φi\varphi_{i} are {2nπ/k:n}\{2n\pi/k:n\in\mathbb{Z}\} or {(2n+1)π/k:n}\{(2n+1)\pi/k:n\in\mathbb{Z}\} modulo 2π2\pi, where hh satisfies 2.12.

Proof.

By the symmetry of hh, φi(2nπ/kθ)\varphi_{i}(2n\pi/k-\theta) is an eigenfunction of λi\lambda_{i}. Since the eigenspace of λi\lambda_{i} has dimension one, we must have φi(θ)=cφi(2nπ/kθ)\varphi_{i}(\theta)=c\varphi_{i}(2n\pi/k-\theta) for some c(n)0c(n)\neq 0 and any θ𝕊1\theta\in\mathbb{S}^{1}. Replacing θ\theta by 2nπ/kθ2n\pi/k-\theta, one gets φi(2nπ/kθ)=c(n)φi(θ)\varphi_{i}(2n\pi/k-\theta)=c(n)\varphi_{i}(\theta). Thus, c(n)c(n) must be 11 or 1-1.

If c(n)=1c(n)=1, then φi\varphi_{i} has even reflection symmetry with respect to nπ/kn\pi/k. Then, we have φi(nπ/k)=0\varphi_{i}^{\prime}(n\pi/k)=0, and because Lemma 2.5, we also φi(nπ/k)0\varphi_{i}(n\pi/k)\neq 0 for such nn. If c(n)=1c(n)=-1, then φi\varphi_{i} has an odd reflection symmetry with respect to nπ/kn\pi/k. Obviously, we have φi(nπ/k)=0\varphi_{i}(n\pi/k)=0 for such nn. Conversely, if φi(nπ/k)0\varphi_{i}(n\pi/k)\neq 0 then φi\varphi_{i} has even reflection symmetry with respect to nπ/kn\pi/k. If φi(nπ/k)=0\varphi_{i}(n\pi/k)=0, then φi\varphi_{i} has odd reflection symmetry with respect to nπ/kn\pi/k.


Now, we consider φi\varphi_{i} on [0,nπ/k][0,n\pi/k]. We divide it into three cases.

First, if φi\varphi_{i} has a zero in (0,nπ/k)(0,n\pi/k), then the reflect symmetries of φ\varphi guarantees at least 2k2k zeros in 𝕊1\mathbb{S}^{1}.

Second, if φi(0)=φi(nπ/k)=0\varphi_{i}(0)=\varphi_{i}(n\pi/k)=0 and has no zero inside (0,nπ/k)(0,n\pi/k), then after the reflection symmetries of φ\varphi guarantees at least 2k2k zeros in 𝕊1\mathbb{S}^{1}.

Last, φi\varphi_{i} has only one zero on the endpoint of [0,nπ/k][0,n\pi/k], say φi(0)=0\varphi_{i}(0)=0 and φi0\varphi_{i}\neq 0 for (0,nπ/k](0,n\pi/k]. In this case, the previous paragraph shows that φi\varphi_{i} has even reflection symmetry with respect to (2n+1)π/k(2n+1)\pi/k and has odd reflection symmetry with respect to 2nπ/k2n\pi/k for any nn. Moreover, kk must be even since φi\varphi_{i} has even number of nodal sets. Counting the zeros of φi\varphi_{i}, we find it is kk in this case. ∎

Now we can prove the first part of Theorem 2.4.

Proof of Theorem 2.4 Part (1).

We will use the induction to prove that λ2l=λ2l+1\lambda_{2l}=\lambda_{2l+1} for any 1l<k1\leq l<k, except that λkλk+1\lambda_{k}\leq\lambda_{k+1} when kk is even. In any case, eigenfunctions corresponds to λ2l\lambda_{2l} and λ2l+1\lambda_{2l+1} have 2l2l nodal domains. If l=1l=1, then λ2=λ3\lambda_{2}=\lambda_{3} by Proposition 2.2. Any eigenfunction in span{cosθ,sinθ}span\{\cos\theta,\sin\theta\} has 2 nodal domains. Suppose the induction is complete for any ll such that 2l+1k12l+1\leq k-1, that is

λ1<λ2=λ3<<λ2l=λ2l+1<0\lambda_{1}<\lambda_{2}=\lambda_{3}<\cdots<\lambda_{2l}=\lambda_{2l+1}<0

and those eigenfunctions of λ2j=λ2j+1\lambda_{2j}=\lambda_{2j+1} have 2j2j nodal domains for jlj\leq l.


It follows from Courant nodal domain theorem [21, VI.6], φ2(l+1)\varphi_{2(l+1)} has at most 2(l+1)2(l+1) nodal domains. Because 2(l+1)2(k1)2(l+1)\leq 2(k-1), Lemma 2.8 implies that λ2(l+1)\lambda_{2(l+1)} will be repeated unless kk is even and 2(l+1)=k2(l+1)=k.


Let us first consider the case that λ2(l+1)\lambda_{2(l+1)} is repeated. We can continue the induction. The dimension of the eigenspace of each eigenvalue is at most 2, thus λ2l+3=λ2(l+1)>λ2l+1=λ2l\lambda_{2l+3}=\lambda_{2(l+1)}>\lambda_{2l+1}=\lambda_{2l}. Now we only need to prove φ2l+2\varphi_{2l+2} and φ2l+3\varphi_{2l+3} has 2l+22l+2 nodal domains. First, they have the same number of nodal sets by Lemma 2.6, while Lemma 2.7 implies that they must have at least 2l+22l+2 nodal sets. Combining the previous upper bound on the number of nodal sets, our induction for l+1l+1 is complete.


If kk is even and λ2(l+1)=λk\lambda_{2(l+1)}=\lambda_{k} is simple, then Lemma 2.8 says φk\varphi_{k} has kk zeros. Now, we consider λk+1\lambda_{k+1}. By the Courant nodal domain theorem and Lemma 2.5, any eigenfunction associated to λk+1\lambda_{k+1} also has kk zeros. Therefore, the second part of Lemma 2.7 implies that λk+1\lambda_{k+1} is also simple. Then, we have λ2l+2<λ2l+3<λ2l+4\lambda_{2l+2}<\lambda_{2l+3}<\lambda_{2l+4}. By Lemma 2.8, λ2l+4\lambda_{2l+4} is repeated, and its eigenfunctions has 2l+4=k+22l+4=k+2 zeros. The induction on l+1l+1 and l+2l+2 is complete. Now, since 2(l+2)>k2(l+2)>k in this case, Lemma 2.8 shows the rest negative eigenvalue are all repeated, therefore it can be continued as the previous case.

Putting everything together, we have the following relations

λ1<λ2=λ3<<λ2l=λ2l+1<<λ2k2=λ2k1\lambda_{1}<\lambda_{2}=\lambda_{3}<\cdots<\lambda_{2l}=\lambda_{2l+1}<\cdots<\lambda_{2k-2}=\lambda_{2k-1}

except that when kk is even, the relation λk=λk+1\lambda_{k}=\lambda_{k+1} will be replaced by λkλk+1\lambda_{k}\leq\lambda_{k+1}. All these eigenvalues are negative, because λ=0\lambda=0 has the eigenfunction hθh_{\theta} with 2k2k nodal sets and φ2k2\varphi_{2k-2} has 2k22k-2 nodal sets. So, Lemma 2.7 says λ2k2<0\lambda_{2k-2}<0. ∎

Next, in order to show Part (2) of Theorem 2.4, we consider the following eigenvalue problems in the Hilbert space H1([0,π/k])H^{1}([0,\pi/k]) with various boundary conditions

ψ′′+ψ=1αh11α(μ+1)ψ,ψ(0)=ψ(π/k)=0,\displaystyle\psi^{\prime\prime}+\psi=-\frac{1}{\alpha}h^{-1-\frac{1}{\alpha}}(\mu+1)\psi,\quad\psi(0)=\psi(\pi/k)=0, (DD)
ψ′′+ψ=1αh11α(μ+1)ψ,ψ(0)=ψ(π/k)=0,\displaystyle\psi^{\prime\prime}+\psi=-\frac{1}{\alpha}h^{-1-\frac{1}{\alpha}}(\mu+1)\psi,\quad\psi(0)=\psi^{\prime}(\pi/k)=0, (DN)
ψ′′+ψ=1αh11α(μ+1)ψ,ψ(0)=ψ(π/k)=0,\displaystyle\psi^{\prime\prime}+\psi=-\frac{1}{\alpha}h^{-1-\frac{1}{\alpha}}(\mu+1)\psi,\quad\psi^{\prime}(0)=\psi(\pi/k)=0, (ND)
ψ′′+ψ=1αh11α(μ+1)ψ,ψ(0)=ψ(π/k)=0.\displaystyle\psi^{\prime\prime}+\psi=-\frac{1}{\alpha}h^{-1-\frac{1}{\alpha}}(\mu+1)\psi,\quad\psi^{\prime}(0)=\psi^{\prime}(\pi/k)=0. (NN)

Here we also assume hh satisfies (2.12). According to the Sturm-Liouville theory, the eigenvalues μiAB\mu_{i}^{AB} for the problems (AB)(AB) where A,B=DA,B=D or NN satisfy,

μ1AB<μ2AB<μ3AB<\mu_{1}^{AB}<\mu_{2}^{AB}<\mu_{3}^{AB}<\cdots

and the eigenfunction ψiAB\psi_{i}^{AB} corresponding to μiAB\mu_{i}^{AB} have i1i-1 zeros.

For (NN), it is easy to know hH1([0,π/k])h\in H^{1}([0,\pi/k]) and it is an eigenfunction to the first eigenvalue μ1NN=1α\mu_{1}^{NN}=-1-\alpha.

Proposition 2.9.

For 0<α<1/(k21)0<\alpha<{1}/({k^{2}-1}), we have μ2NN>0\mu_{2}^{NN}>0 of (NN).

Proof.

We shall write μ2=μ2NN\mu_{2}=\mu_{2}^{NN} for short within this proposition.

First, μ2\mu_{2} can not be equal to zero. Otherwise, we will have an eigenfunction ψ2\psi_{2} on [0,π/k][0,\pi/k] such that

ψ2′′+ψ2=1αh11αψ2,ψ2(0)=ψ2(π/k)=0.\psi_{2}^{\prime\prime}+\psi_{2}=-\frac{1}{\alpha}h^{-1-\frac{1}{\alpha}}\psi_{2},\quad\psi_{2}^{\prime}(0)=\psi_{2}^{\prime}(\pi/k)=0.

By reflecting ψ2\psi_{2} about nπ/kn\pi/k evenly for any nn, we can extend ψ2\psi_{2} to a smooth function defined on 𝕊1\mathbb{S}^{1}. This contradicts to the fact that the eigenspace of \mathcal{L} for λ=0\lambda=0 is one dimensional, because ψ2span{hθ}\psi_{2}\not\in span\{h_{\theta}\}.

Second, suppose μ2<0\mu_{2}<0 and ψ2\psi_{2} is an eigenfunction. In the following Lemma 2.10, we get a function η(θ)\eta(\theta) with

η′′+η=1αh11αη\eta^{\prime\prime}+\eta=-\frac{1}{\alpha}h^{-1-\frac{1}{\alpha}}\eta

with η(0)>0\eta(0)>0 and η(0)=0\eta^{\prime}(0)=0, η(π/k)<0\eta(\pi/k)<0, η(π/k)<0\eta^{\prime}(\pi/k)<0. See Figure 2 for illustration.

We claim that η\eta has only one zero in (0,π/k)(0,\pi/k).

In fact, on the contrary assume η(θ0)=η(θ1)=0\eta(\theta_{0})=\eta(\theta_{1})=0 for 0<θ0<θ1<π/k0<\theta_{0}<\theta_{1}<\pi/k. One can define a new function η~(θ)\tilde{\eta}(\theta) such that it equals η(θ)\eta(\theta) if θ[θ0,θ1]\theta\in[\theta_{0},\theta_{1}] and zero elsewhere. Then obviously η~H1([0,π/k])\tilde{\eta}\in H^{1}([0,\pi/k]) and

0π/kη~θ2η~2=0π/k1αh11/αη~2\displaystyle\int_{0}^{\pi/k}\tilde{\eta}_{\theta}^{2}-\tilde{\eta}^{2}=\int_{0}^{\pi/k}\frac{1}{\alpha}h^{-1-1/\alpha}\tilde{\eta}^{2} (2.13)

Recall the following variational characterization of μ1DD\mu_{1}^{DD},

μ1DD=infuH1([0,π/k]),u0{0π/kuθ2u20π/ku21αh11/α1|u(0)=u(π/k)=0}.\mu_{1}^{DD}=\inf_{u\in H^{1}([0,\pi/k]),u\not\equiv 0}\left\{\left.\frac{\int_{0}^{\pi/k}u_{\theta}^{2}-u^{2}}{\int_{0}^{\pi/k}u^{2}\frac{1}{\alpha}h^{-1-1/\alpha}}-1\right|u(0)=u(\pi/k)=0\right\}.

The infimum is achieved by the first eigenfunction of (DD). Using (3) in Proposition 2.2, we have μ1DD=0\mu_{1}^{DD}=0 and eigenfunction ψ1DD=hθ\psi_{1}^{DD}=h_{\theta}, because hθh_{\theta} does not change sign in (0,π/k)(0,\pi/k). However, (2.13) implies μ1DD<0\mu_{1}^{DD}<0. Contradiction. The claim is proved.

Let aa be the only zero of η\eta in (0,π/k)(0,\pi/k) and bb be that of ψ2\psi_{2}. Without loss of generality, we assume ψ2(θ)>0\psi_{2}(\theta)>0 when θ(0,b)\theta\in(0,b). Otherwise one can work on ψ2-\psi_{2}. Define W[ψ2,η]=ψ2ηψ2ηW[\psi_{2},\eta]=\psi_{2}\eta^{\prime}-\psi_{2}^{\prime}\eta. Then, we have W(0)=0W(0)=0, W(π/k)>0W(\pi/k)>0 and

W(θ)=1αh11αμ2ψ2η.W^{\prime}(\theta)=\frac{1}{\alpha}h^{-1-\frac{1}{\alpha}}\mu_{2}\psi_{2}\eta.

If aba\geq b, then W(b)0W(b)\geq 0 while W<0W^{\prime}<0 in (0,b)(0,b). This is impossible because of W(0)=0W(0)=0.

If a<ba<b, then W(b)<0W(b)<0, W<0W^{\prime}<0 on (b,π/k)(b,\pi/k). This contradicts to W(π/k)>0W(\pi/k)>0. Therefore μ2\mu_{2} can not be negative. ∎

Refer to caption
Figure 2. The graph of η\eta with η(0)=1\eta(0)=1 when α=1/16\alpha=1/16 and k=3k=3.
Lemma 2.10.

Let k3k\geq 3 and 0<α<1/(k21)0<\alpha<1/(k^{2}-1). There exists a smooth function η\eta on [0,2π][0,2\pi] satisfying

η′′+η+1αh11/αη=0\displaystyle\eta^{\prime\prime}+\eta+\frac{1}{\alpha}h^{-1-1/\alpha}\eta=0 (2.14)

and η(0)>0\eta(0)>0, η(0)=0\eta^{\prime}(0)=0, η(π/k)<0\eta(\pi/k)<0 and η(π/k)<0\eta^{\prime}(\pi/k)<0.

Proof.

We will use some notations in [2, Lemma 7.2]. Consider the function U(α,r,θ)U(\alpha,r,\theta) defined by

Uθθ+U\displaystyle U_{\theta\theta}+U =U1α\displaystyle=U^{-\frac{1}{\alpha}}
Uθ(α,r,0)=0,\displaystyle U_{\theta}(\alpha,r,0)=0,\quad Uθ(α,r,Θ(α,r))=0\displaystyle U_{\theta}(\alpha,r,\Theta(\alpha,r))=0
Uθ(α,r,θ)\displaystyle U_{\theta}(\alpha,r,\theta) <0,0<θ<Θ(α,r)\displaystyle<0,\quad 0<\theta<\Theta(\alpha,r)
U(α,r,0)\displaystyle U(\alpha,r,0) =rU(α,r,Θ(α,r))\displaystyle=rU(\alpha,r,\Theta(\alpha,r))

where Θ=Θ(α,r)\Theta=\Theta(\alpha,r) is the period function defined in [2, Definition 2.1]. Moreover, one can find

U(α,r,Θ(α,r))\displaystyle U(\alpha,r,\Theta(\alpha,r)) =(2α1α1r11/αr21)αα+1,\displaystyle=\left(\frac{2\alpha}{1-\alpha}\cdot\frac{1-r^{1-1/\alpha}}{r^{2}-1}\right)^{\frac{\alpha}{\alpha+1}}, (2.15)
U(α,r,0)\displaystyle U(\alpha,r,0) =rU(α,r,Θ(α,r)).\displaystyle=rU(\alpha,r,\Theta(\alpha,r)). (2.16)

We will omit dependence on α\alpha of UU and Θ\Theta in what follows. It follows from [2] that for each α(0,1/(k21))\alpha\in(0,1/(k^{2}-1)), there exists a unique r1r^{*}\geq 1 such that Θ(r)=π/k\Theta(r_{*})=\pi/k. The support function hh is given by h(θ)=U(r,θ)h(\theta)=U(r_{*},\theta). Define η(θ)=ddrU(r,θ)|r=r\eta(\theta)=\frac{d}{dr}U(r,\theta)|_{r=r_{*}}. Then obviously η\eta satisfies (2.14). Since Uθ(r,0)=0U_{\theta}(r,0)=0 for any rr, we have ηθ(0)=0\eta_{\theta}(0)=0. Note that (2.16) implies

U(r,0)=r(2α1α1r11/αr21)αα+1U(r,0)=r\left(\frac{2\alpha}{1-\alpha}\cdot\frac{1-r^{1-1/\alpha}}{r^{2}-1}\right)^{\frac{\alpha}{\alpha+1}}

Differentiating with respect to rr implies η(0)>0\eta(0)>0.

Since Uθ(r,Θ(r))=0U_{\theta}(r,\Theta(r))=0, Differentiating with respect to rr gives

ηθ(Θ(r))+Uθθ(r,Θ(r))ddrΘ(r)=0.\displaystyle\eta_{\theta}(\Theta(r))+U_{\theta\theta}(r,\Theta(r))\frac{d}{dr}\Theta(r)=0. (2.17)

Note that d/drΘ(r)>0d/dr\,\Theta(r)>0 if α(0,1/3)\alpha\in(0,1/3) by [2]. Also Uθθ(r,π/k)0U_{\theta\theta}(r_{*},\pi/k)\geq 0, because U(r,θ)U(r_{*},\theta) attains the minimum at r=rr=r_{*}. Here Uθθ(r,π/k)U_{\theta\theta}(r_{*},\pi/k) can not be 0, otherwise combined with Uθ(r,π/k)=0U_{\theta}(r_{*},\pi/k)=0, one gets UU is a constant. Inserting r=rr=r_{*} to the above equation, one can see ηθ(π/k)<0\eta_{\theta}(\pi/k)<0.

On the other hand, it follows from (2.16) that

U(r,Θ(r))=(2α1α1r11/αr21)αα+1U(r,\Theta(r))=\left(\frac{2\alpha}{1-\alpha}\cdot\frac{1-r^{1-1/\alpha}}{r^{2}-1}\right)^{\frac{\alpha}{\alpha+1}}

Taking the derivative with respect to rr of the above equation reveals ddrU(r,Θ(r))<0\frac{d}{dr}U(r,\Theta(r))<0. Therefore η(π/k)=ddrU(r,Θ(r))|r=r<0\eta(\pi/k)=\frac{d}{dr}U(r,\Theta(r))|_{r=r_{*}}<0. ∎

Proof of Theorem 2.4 Part (2).

Since λ2k1\lambda_{2k-1} has a two dimensional eigenspace with 2k22k-2 nodal sets, Lemma 2.7 implies λ2k1<λ2k\lambda_{2k-1}<\lambda_{2k}. The Courant nodal domain and Lemma 2.5 imply that the eigenfunctions assosicated to λ2k\lambda_{2k} must have 2k2k nodal sets. We need to show λ2k=0\lambda_{2k}=0.

Towards a contradiction, suppose that λ2k<0\lambda_{2k}<0. hθh_{\theta} is an eigenfunction corresponding to the eigenvalue 0 and it also has 2k2k nodal sets. Thus, Lemma 2.7 implies that λ2k\lambda_{2k} must be simple. Therefore, Lemma 2.8 says that the eigenfunction φ2k\varphi_{2k} is even-reflection-symmetric with respect to nπ/kn\pi/k for all nn\in\mathbb{N} and it has exactly 2k2k nodal sets. Thus, the restriction of φ2k\varphi_{2k} on [0,π/k][0,\pi/k] is a Neumann eigenfunction to (NN). Since φ2k\varphi_{2k} changes its sign exactly once in [0,π/k][0,\pi/k], we have λ2kμ2NN\lambda_{2k}\geq\mu_{2}^{NN} which is the second Neumann eigenvalue. However, Proposition 2.9 says μ2NN>0\mu_{2}^{NN}>0.

Since λ2k1<0\lambda_{2k-1}<0 has 2k22k-2 nodal domains and 0 is an eigenvalue having 2k2k nodal domains, we have λ2k=0\lambda_{2k}=0. λ2k\lambda_{2k} is simple by Proposition 2.2, and thus we will have the next eigenvalue λ2k+1>0\lambda_{2k+1}>0. The nodal sets of eigenfunction associated to λ2k+1\lambda_{2k+1} is 2k2k by Courant nodal domain theorem. Then the second part of Lemma 2.7 implies that λ2k+1\lambda_{2k+1} also have to be simple. ∎

We completed the proof the Theorem 2.4. From now on, we discuss about why we may not have λk=λk+1\lambda_{k}=\lambda_{k+1} when kk is even. These two eigenvalues are related to the (DN)\eqref{eq:DN-eig} and (ND)\eqref{eq:ND-eig}.

Lemma 2.11.

We have μ1DN<0\mu_{1}^{DN}<0 and μ1ND<0\mu_{1}^{ND}<0.

Proof.

Suppose ψDN\psi^{DN} is an eigenfunction corresponding to μ1DN\mu_{1}^{DN}. Then make an even reflection of ψDN\psi^{DN} with respect to π/k\pi/k. We get ψDN\psi^{DN} is an eigenfunction of

ψ′′+ψ=1αh11α(μDN+1)ψ,ψ(0)=ψ(2π/k)=0\psi^{\prime\prime}+\psi=-\frac{1}{\alpha}h^{-1-\frac{1}{\alpha}}(\mu^{DN}+1)\psi,\quad\psi(0)=\psi(2\pi/k)=0

Since ψDN\psi^{DN} does not change sign in [0,2π][0,2\pi], it must be the first eigenfunction for the above problem. Note that the reflection of hθh_{\theta} also makes an eigenfunction corresponds to 0 for the above problem. We must have μ1DN<0\mu^{DN}_{1}<0.

The fact of (ND) can be proved through even reflection with respect to θ=0\theta=0. ∎

Remark 2.12.

If λk\lambda_{k} is simple, then Lemma 2.8 says φk\varphi_{k} will have kk zeros. Moreover, Lemma 2.8 indicates that the restriction of φk\varphi_{k} on [0,π/k][0,\pi/k] will give us a first eigenfunction of (DN) or (ND) corresponds to μ1DN\mu_{1}^{DN} or μ1ND\mu_{1}^{ND}. In fact λk=min{μ1DN,μ1ND}\lambda_{k}=\min\{\mu_{1}^{DN},\mu_{1}^{ND}\}. For the same reason, λk+1=max{μ1DN,μ1ND}\lambda_{k+1}=\max\{\mu_{1}^{DN},\mu_{1}^{ND}\}. A priori we do not know μ1DN=μ1ND\mu_{1}^{DN}=\mu_{1}^{ND}.

3. Construction of ancient solutions

In this section, we construct ancient solutions converging to a shrinker Γ\Gamma after rescaling by using the Morse index II we characterized in Section 2. Let us denote the Morse index of Γ\mathcal{L}_{\Gamma} by I(Γ)I(\mathcal{L}_{\Gamma}) . In the section 2, we showed I(Γ¯αk)=2k1I(\mathcal{L}_{\bar{\Gamma}_{\alpha}^{k}})=2k-1 and I(Γ¯αc)=21+1/α1I(\mathcal{L}_{\bar{\Gamma}_{\alpha}^{c}})=2\lceil 1+1/\alpha\rceil-1. Again, we shall simply suppress the notation to II and \mathcal{L}. One should interpret the following for each case Γ=Γ¯αk\Gamma=\bar{\Gamma}_{\alpha}^{k} or Γ¯αc\bar{\Gamma}_{\alpha}^{c} respectively.

We begin by considering the inhomogeneous linear PDE

τv=v+EΓ(v).\partial_{\tau}v=\mathcal{L}v+E_{\Gamma}(v).

Fix β(0,1)\beta\in(0,1), and for any f:𝕊1×f:\mathbb{S}^{1}\times\mathbb{R}_{-}\to\mathbb{R} we define the seminorm

|f(τ)|𝒞β=sup(θi,ti)𝕊1×(τ1,τ){|f(θ1,t1)f(θ2,t2)||θ1θ2|β+|t1t2|β/2:(θ1,t1)(θ2,t2)}.|f(\tau)|_{\mathcal{C}^{\beta}}=\sup_{(\theta_{i},t_{i})\in\mathbb{S}^{1}\times(\tau-1,\tau)}\left\{\frac{|f(\theta_{1},t_{1})-f(\theta_{2},t_{2})|}{|\theta_{1}-\theta_{2}|^{\beta}+|t_{1}-t_{2}|^{\beta/2}}:(\theta_{1},t_{1})\neq(\theta_{2},t_{2})\right\}.

We use the special symbol 𝒞\mathcal{C} to denote the parabolic norm in what follows. Notice that we write τ\tau explicitly in |f(τ)|𝒞β|f(\tau)|_{\mathcal{C}^{\beta}} to indicate that the parabolic norm is taken on 𝕊1×(τ1,τ)\mathbb{S}^{1}\times(\tau-1,\tau). For l0l\geq 0, define the norm

f(τ)𝒞l,β:=i+2jlsup𝕊1×(τ1,τ)|θitjf|+i+2j=l|θitjf|𝒞β.\displaystyle||f(\tau)||_{\mathcal{C}^{l,\beta}}:=\sum_{i+2j\leq l}\sup_{\mathbb{S}^{1}\times(\tau-1,\tau)}|\partial_{\theta}^{i}\partial_{t}^{j}f|+\sum_{i+2j=l}|\partial_{\theta}^{i}\partial_{t}^{j}f|_{\mathcal{C}^{\beta}}. (3.1)

For some δ>0\delta>0, define the norm

f𝒞l,β,δ:=supτ0{eδτ||f||𝒞2,β(𝕊1×(τ1,τ))}.\displaystyle||f||_{\mathcal{C}^{l,\beta,\delta}}:=\sup_{\tau\leq 0}\{e^{-\delta\tau}||f||_{\mathcal{C}^{2,\beta}(\mathbb{S}^{1}\times(\tau-1,\tau))}\}. (3.2)

Suppose XδX^{\delta} is the Banach space equipped with the norm f𝒞l,β,δ<||f||_{\mathcal{C}^{l,\beta,\delta}}<\infty.

We fix once and for all an Lh2L^{2}_{h} orthonormal sequence of eigenfunctions φj\varphi_{j} of -\mathcal{L} such that φj=λjφj\mathcal{L}\varphi_{j}=\lambda_{j}\varphi_{j} and (φj,φj)h=1(\varphi_{j},\varphi_{j})_{h}=1. Define vj=(v,φj)hv_{j}=(v,\varphi_{j})_{h} and Pjv=(v,φj)hφjP_{j}v=(v,\varphi_{j})_{h}\varphi_{j}. We also define Pj=i=1jPiP_{\leq j}=\sum_{i=1}^{j}P_{i} and

P=j=0IPj,P+={j:λj>0}Pj,P0={j:λj=0}Pj.\displaystyle P_{-}=\sum_{j=0}^{I}P_{j},\quad P_{+}=\sum_{\{j:\lambda_{j}>0\}}P_{j},\quad P_{0}=\sum_{\{j:\lambda_{j}=0\}}P_{j}. (3.3)

In addition, we define

fL2,δ=supτ0{eδτ||f(,τ)||h}.||f||_{L^{2,\delta}}=\sup_{\tau\leq 0}\{e^{-\delta\tau}||f(\cdot,\tau)||_{h}\}.

For the rest of this section, we will always choose δ\delta as some positive constant different from λj-\lambda_{j} for any jj. Denote J={j:λj<δ}{1,,I}J=\{j:\lambda_{j}<-\delta\}\subset\{1,\cdots,I\}. For example, J=J=\emptyset if δ>λ1=1α\delta>\lambda_{1}=-1-\alpha.

Lemma 3.1.

Fix any 0<δ{λj}j=10<\delta\not\in\{-\lambda_{j}\}_{j=1}^{\infty} and recall the operator \mathcal{L} in (1.9). If fL2,δ<||f||_{L^{2,\delta}}<\infty, then the equation

τuu=f(θ,τ), on 𝕊1×\displaystyle\partial_{\tau}u-\mathcal{L}u=f(\theta,\tau),\quad\text{ on }\quad\mathbb{S}^{1}\times\mathbb{R}_{-}

has a unique solution uu satisfying uL2,δ<||u||_{L^{2,\delta}}<\infty and Pj(u(,0))=0P_{j}(u(\cdot,0))=0 for jJj\in J. Furthermore, there exists C=C(α,β,δ)C=C(\alpha,\beta,\delta) such that uL2,δCfL2,δ||u||_{L^{2,\delta}}\leq C||f||_{L^{2,\delta}} and u𝒞2,β,δCf𝒞0,β,δ||u||_{\mathcal{C}^{2,\beta,\delta}}\leq C||f||_{\mathcal{C}^{0,\beta,\delta}} hold.

Proof.

Recall that {φi}\{\varphi_{i}\} is an orthonormal basis of Lh2(𝕊1)L^{2}_{h}(\mathbb{S}^{1}) with respect to (,)h(,)_{h}. It suffices to solve

τui+λiui=fi,𝕊1×,\displaystyle\partial_{\tau}u_{i}+\lambda_{i}u_{i}=f_{i},\quad\mathbb{S}^{1}\times\mathbb{R}_{-},

where ui=(u,φi)hu_{i}=(u,\varphi_{i})_{h} and fi=(f,φi)hf_{i}=(f,\varphi_{i})_{h}. Denote

uj(τ)\displaystyle u_{j}(\tau) :=τ0eλj(sτ)fj(s)𝑑s,jJ,\displaystyle:=\int_{\tau}^{0}e^{\lambda_{j}(s-\tau)}f_{j}(s)ds,\quad j\in J,
uj(τ)\displaystyle u_{j}(\tau) :=τeλj(sτ)fj(s)𝑑s,jJc=+\J.\displaystyle:=\int_{-\infty}^{\tau}e^{\lambda_{j}(s-\tau)}f_{j}(s)ds,\quad j\in J^{c}=\mathbb{Z}_{+}\backslash J.

Notice the integral on the RHS is well-defined because |fj(s)|f(,s)hfL2,δeδs|f_{j}(s)|\leq||f(\cdot,s)||_{h}\leq||f||_{L^{2,\delta}}e^{\delta s}. Define u(,τ)=j=1uj(τ)φj()u(\cdot,\tau)=\sum_{j=1}^{\infty}u_{j}(\tau)\varphi_{j}(\cdot). It is easy to see Pj(u(,0))=0P_{j}(u(\cdot,0))=0 for any jJj\in J.

Choose δ\delta^{\prime} and δ′′\delta^{\prime\prime} satisfying maxjJ{λj}<δ<δ<δ′′<minjJc{λj}\max_{j\in J}\{\lambda_{j}\}<-\delta^{\prime}<-\delta<-\delta^{\prime\prime}<\min_{j\in J^{c}}\{\lambda_{j}\}. Note that for jJj\in J

uj2(τ)τ0e2(λj+δ)(sτ)𝑑sτ0e2δ(sτ)|fj|2𝑑sCτ0e2δ(sτ)|fj|2𝑑s\displaystyle u_{j}^{2}(\tau)\leq\int_{\tau}^{0}e^{2(\lambda_{j}+\delta^{\prime})(s-\tau)}ds\int_{\tau}^{0}e^{-2\delta^{\prime}(s-\tau)}|f_{j}|^{2}ds\leq C\int_{\tau}^{0}e^{-2\delta^{\prime}(s-\tau)}|f_{j}|^{2}ds

and for jJcj\in J^{c}

uj2(τ)τe2(λj+δ′′)(sτ)𝑑sτe2δ′′(sτ)|fj|2𝑑sCτe2δ′′(sτ)|fj|2𝑑s.\displaystyle u_{j}^{2}(\tau)\leq\int_{-\infty}^{\tau}e^{2(\lambda_{j}+\delta^{\prime\prime})(s-\tau)}ds\int_{-\infty}^{\tau}e^{-2\delta^{\prime\prime}(s-\tau)}|f_{j}|^{2}ds\leq C\int_{-\infty}^{\tau}e^{-2\delta^{\prime\prime}(s-\tau)}|f_{j}|^{2}ds.

Combining the above two inequalities and using |fj(s)|fL2,δeδs|f_{j}(s)|\leq||f||_{L^{2,\delta}}e^{\delta s}, one obtains

u(,τ)h2=juj2(τ)\displaystyle||u(\cdot,\tau)||_{h}^{2}=\sum_{j}u_{j}^{2}(\tau)\leq Cτ0e2δ(sτ)|fj|2𝑑s+Cτe2δ′′(sτ)|fj|2𝑑s\displaystyle C\int_{\tau}^{0}e^{-2\delta^{\prime}(s-\tau)}|f_{j}|^{2}ds+C\int_{-\infty}^{\tau}e^{-2\delta^{\prime\prime}(s-\tau)}|f_{j}|^{2}ds
\displaystyle\leq CfL2,δ2e2δτ.\displaystyle C||f||_{L^{2,\delta}}^{2}e^{2\delta\tau}.

Therefore uL2,δCfL2,δ||u||_{L^{2,\delta}}\leq C||f||_{L^{2,\delta}}.

Let’s establish 𝒞2,β,δ\mathcal{C}^{2,\beta,\delta} bounds. By the interior parabolic Schauder estimates (for instance, see [18, (C.6)]), we have that for any τ0\tau\leq 0,

u𝒞2,β(𝕊1×(τ1,τ))C(uL2(𝕊1×(τ2,τ))+f𝒞0,β(𝕊1×(τ2,τ))).||u||_{\mathcal{C}^{2,\beta}(\mathbb{S}^{1}\times(\tau-1,\tau))}\leq C\left(||u||_{L^{2}(\mathbb{S}^{1}\times(\tau-2,\tau))}+||f||_{\mathcal{C}^{0,\beta}(\mathbb{S}^{1}\times(\tau-2,\tau))}\right).

Multiplying by eδτe^{\delta\tau} and taking the supremum over τ0\tau\leq 0 yield

u𝒞2,β,δC(uL2,δ+f𝒞0,β,δ)C(fL2,δ+f𝒞0,β,δ)Cf𝒞0,β,δ.\displaystyle||u||_{\mathcal{C}^{2,\beta,\delta}}\leq C(||u||_{L^{2,\delta}}+||f||_{\mathcal{C}^{0,\beta,\delta}})\leq C(||f||_{L^{2,\delta}}+||f||_{\mathcal{C}^{0,\beta,\delta}})\leq C||f||_{\mathcal{C}^{0,\beta,\delta}}.

We shall use contraction mapping theorem and the above lemma repeatedly to construct ancient solutions. Let us introduce some necessary notations. For any 𝒂=(a1,,aI)I\boldsymbol{a}=(a_{1},\cdots,a_{I})\in\mathbb{R}^{I}, denote |𝒂|=(i=1Iai2)12|\boldsymbol{a}|=(\sum_{i=1}^{I}a_{i}^{2})^{\frac{1}{2}}. We introduce auxiliary operators which maps any integer set J{1,2,,I}J\subset\{1,2,\cdots,I\} to functions,

ιJ:I\displaystyle\iota^{J}:\mathbb{R}^{I} Lh2×,\displaystyle\to L_{h}^{2}\times\mathbb{R}_{-},
ιJ(𝒂)\displaystyle\iota^{J}(\boldsymbol{a}) :=jJajeλjτφj.\displaystyle:=\sum_{j\in J}a_{j}e^{-\lambda_{j}\tau}\varphi_{j}.

Denote L=λ1/λIL=\lfloor\lambda_{1}/\lambda_{I}\rfloor. For each l=1,,Ll=1,\cdots,L, define

J(l)={m:(l+1)λI<λmlλI}.J^{(l)}=\{m:(l+1)\lambda_{I}<\lambda_{m}\leq l\lambda_{I}\}.

Then l=1LJ(l)\cup_{l=1}^{L}J^{(l)} is a partition of {1,,I}\{1,\cdots,I\} according to the negative eigenvalues of Γ\mathcal{L}_{\Gamma}. Choose δl\delta_{l} satisfying

max{λj:jJ(l+1)}(l+1)λI<δl<min{λj:jJ(l)}.\max\{\lambda_{j}:j\in J^{(l+1)}\}\leq(l+1)\lambda_{I}<-\delta_{l}<\min\{\lambda_{j}:j\in J^{(l)}\}.

Write X(l)=XδlX^{(l)}=X^{\delta_{l}}, ι(l)=ιJ(l)\iota^{(l)}=\iota^{J^{(l)}}, and P(l)=j:λj<δlPjP^{(l)}=\sum_{j:\lambda_{j}<-\delta_{l}}P_{j} for simplicity. In what follows, we will use the symbol \lesssim for inequalities that hold up to multiplicative constants that may depend on α,h\alpha,h.

Here is the main result of this section

Theorem 3.2.

Let L=λ1/λIL=\lfloor\lambda_{1}/\lambda_{I}\rfloor666x\lfloor x\rfloor means the greatest integer less than or equal to xx. There exists some ε0>0\varepsilon_{0}>0 satisfying the following significance. Given 𝐚=(a1,,aI)I\boldsymbol{a}=(a_{1},\cdots,a_{I})\in\mathbb{R}^{I} with |𝐚|<ε0|\boldsymbol{a}|<\varepsilon_{0}, there exist a set of functions {v(l)}l=1L\{v^{(l)}\}_{l=1}^{L} uniquely determined and depending continuously on 𝐚\boldsymbol{a} such that for each l=1,,Ll=1,\cdots,L, we have v(l)ι(l)(𝐚)X(l)v^{(l)}-\iota^{(l)}(\boldsymbol{a})\in X^{(l)}, P(l)(v(l)ι(l)(𝐚))(,0)=0P^{(l)}(v^{(l)}-\iota^{(l)}(\boldsymbol{a}))(\cdot,0)=0, and j=1lv(j)\sum_{j=1}^{l}v^{(j)} is an ancient solution of (1.8) for (,0](-\infty,0]. More importantly

limτeλmτ(v(l)(,τ),φm)h=am,mJ(l).\displaystyle\lim_{\tau\to-\infty}e^{\lambda_{m}\tau}(v^{(l)}(\cdot,\tau),\varphi_{m})_{h}=a_{m},\quad m\in J^{(l)}. (3.4)

Let us first prove a proposition which will be needed in the proof of Theorem 3.2.

Proposition 3.3.

There exists some constants C=C(α,h)C=C(\alpha,h) and ϵ=ϵ(α,h)>0\epsilon=\epsilon(\alpha,h)>0 such that if |vθθ+v|ϵ|v_{\theta\theta}+v|\leq\epsilon then

|EΓ(v)|\displaystyle|E_{\Gamma}(v)| C|vθθ+v|2,\displaystyle\leq C|v_{\theta\theta}+v|^{2}, (3.5)
|EΓ(v)(τ)|𝒞β\displaystyle|E_{\Gamma}(v)(\tau)|_{\mathcal{C}^{\beta}} C|(vθθ+v)(τ)|𝒞β|(vθθ+v)(τ)|𝒞0.\displaystyle\leq C|(v_{\theta\theta}+v)(\tau)|_{\mathcal{C}^{\beta}}|(v_{\theta\theta}+v)(\tau)|_{\mathcal{C}^{0}}. (3.6)

Moreover, if u,vu,v satisfy |uθθ+u|+|vθθ+v|ϵ|u_{\theta\theta}+u|+|v_{\theta\theta}+v|\leq\epsilon then

|EΓ(u)(τ)EΓ(v)(τ)|𝒞β\displaystyle|E_{\Gamma}(u)(\tau)-E_{\Gamma}(v)(\tau)|_{\mathcal{C}^{\beta}}
\displaystyle\leq C|((uv)θθ+uv)(τ)|𝒞β[|(uθθ+u)(τ)|𝒞0+|(vθθ+v)(τ)|𝒞0]\displaystyle C|((u-v)_{\theta\theta}+u-v)(\tau)|_{\mathcal{C}^{\beta}}\left[|(u_{\theta\theta}+u)(\tau)|_{\mathcal{C}^{0}}+|(v_{\theta\theta}+v)(\tau)|_{\mathcal{C}^{0}}\right]
Proof.

By the definition EΓE_{\Gamma} in (1.8), we obtain

EΓ(v)=h(1+h1α(vθθ+v))α+hαh1+1α(vθθ+v).E_{\Gamma}(v)=-h{\left(1+h^{\frac{1}{\alpha}}(v_{\theta\theta}+v)\right)^{-\alpha}}+h-\alpha h^{1+\frac{1}{\alpha}}(v_{\theta\theta}+v).

Using the Taylor expansion of (1+x)α(1+x)^{-\alpha}, it is easy to know |(1+x)α1+αx|C(α)x2|(1+x)^{-\alpha}-1+\alpha x|\leq C(\alpha)x^{2} whenever |x|<1/2|x|<1/2. Therefore we have

|EΓ(v)|C(α,h)|vθθ+v|2|E_{\Gamma}(v)|\leq C(\alpha,h)|v_{\theta\theta}+v|^{2}

whenever |vθθ+v|<12h1/α|v_{\theta\theta}+v|<\frac{1}{2}h^{-1/\alpha}.

Our second conclusion follows from the following observations

|(1+x)α+αx(1+y)ααy|C(α)(xy)2,|x|+|y|<12|(1+x)^{-\alpha}+\alpha x-(1+y)^{-\alpha}-\alpha y|\leq C(\alpha)(x-y)^{2},\quad|x|+|y|<\frac{1}{2}

and consequently for any t1,t2[τ1,τ]t_{1},t_{2}\in[\tau-1,\tau]

|EΓ(v)(θ1,t1)EΓ(v)(θ2,t2)|\displaystyle|E_{\Gamma}(v)(\theta_{1},t_{1})-E_{\Gamma}(v)(\theta_{2},t_{2})|
\displaystyle\leq C(α,h)|(vθθ+v)(θ1,t1)(vθθ+v)(θ2,t2)|2\displaystyle C(\alpha,h)|(v_{\theta\theta}+v)(\theta_{1},t_{1})-(v_{\theta\theta}+v)(\theta_{2},t_{2})|^{2}
\displaystyle\leq C(α,h)(|θ1θ2|β+|t1t2|β/2)|(vθθ+v)(τ)|𝒞β|(vθθ+v)(τ)|𝒞0.\displaystyle C(\alpha,h)(|\theta_{1}-\theta_{2}|^{\beta}+|t_{1}-t_{2}|^{\beta/2})|(v_{\theta\theta}+v)(\tau)|_{\mathcal{C}^{\beta}}|(v_{\theta\theta}+v)(\tau)|_{\mathcal{C}^{0}}.

The estimates of |EΓ(u)(τ)EΓ(v)(τ)|𝒞β|E_{\Gamma}(u)(\tau)-E_{\Gamma}(v)(\tau)|_{\mathcal{C}^{\beta}} can be proved similarly. ∎

Now we can prove the main theorem of this section.

Proof of Theorem 3.2.

We shall find all v(l)v^{(l)} by the induction. First, we notice that ι(1)\iota^{(1)} is an ancient solution to the linear equation τv=v\partial_{\tau}v=\mathcal{L}v. Therefore, to find v(1)v^{(1)}, we assume v(1)=w+ι(1)(𝒂)v^{(1)}=w+\iota^{(1)}(\boldsymbol{a}) for some ww to be determined. Then (1.8) is equivalent to

τw=w+E(w+ι(1)(𝒂))\displaystyle\partial_{\tau}w=\mathcal{L}w+E(w+\iota^{{(1)}}(\boldsymbol{a})) (3.7)

Here and in the following, we shall write E(v)=EΓ(v)E(v)=E_{\Gamma}(v) and 𝔯[v]=vθθ+v\mathfrak{r}[v]=v_{\theta\theta}+v for short.

Claim 1.

There exists small ε0\varepsilon_{0} such that if wX(1)+|𝐚|<ε0||w||_{X^{(1)}}+|\boldsymbol{a}|<\varepsilon_{0} then

E(w+ι(1)(𝒂))𝒞0,β,δ1\displaystyle||E(w+\iota^{(1)}(\boldsymbol{a}))||_{\mathcal{C}^{0,\beta,\delta_{1}}} wX(1)2+|𝒂|2,\displaystyle\lesssim||w||_{X^{(1)}}^{2}+|\boldsymbol{a}|^{2}, (3.8)
E(w1+ι(1)(𝒂))E(w2+ι(1)(𝒂))𝒞0,β,δ1\displaystyle||E(w_{1}+\iota^{(1)}(\boldsymbol{a}))-E(w_{2}+\iota^{(1)}(\boldsymbol{a}))||_{\mathcal{C}^{0,\beta,\delta_{1}}} ε0w1w2X(1).\displaystyle\lesssim\varepsilon_{0}||w_{1}-w_{2}||_{X^{(1)}}. (3.9)

In fact, one can easily derive from (3.5) and (3.6) that there exists ε0>0\varepsilon_{0}>0 and some constant C(α,β,h)C(\alpha,\beta,h) such that

E(v)(τ)𝒞0,βC𝔯[v](τ)𝒞0,β2\displaystyle||E(v)(\tau)||_{\mathcal{C}^{0,\beta}}\leq C||\mathfrak{r}[v](\tau)||_{\mathcal{C}^{0,\beta}}^{2}

provided 𝔯[v](τ)𝒞0<ε0||\mathfrak{r}[v](\tau)||_{\mathcal{C}^{0}}<\varepsilon_{0}. Furthermore

||E(v1)(τ)\displaystyle||E(v_{1})(\tau)- E(v2)(τ)||𝒞0,β\displaystyle E(v_{2})(\tau)||_{\mathcal{C}^{0,\beta}} (3.10)
\displaystyle\leq C(𝔯[v1](τ)𝒞0,β+𝔯[v2](τ)𝒞0,β)𝔯[v1v2](τ)𝒞0,β\displaystyle C(||\mathfrak{r}[v_{1}]\left(\tau)||_{\mathcal{C}^{0,\beta}}+||\mathfrak{r}[v_{2}](\tau)||_{\mathcal{C}^{0,\beta}}\right)||\mathfrak{r}[v_{1}-v_{2}](\tau)||_{\mathcal{C}^{0,\beta}}

provided 𝔯[v1](τ)𝒞0+𝔯[v2](τ)𝒞0<ε0||\mathfrak{r}[v_{1}](\tau)||_{\mathcal{C}^{0}}+||\mathfrak{r}[v_{2}](\tau)||_{\mathcal{C}^{0}}<\varepsilon_{0}. Therefore

E(w+ι(1)(𝒂))(τ)𝒞0,β\displaystyle||E(w+\iota^{(1)}(\boldsymbol{a}))(\tau)||_{\mathcal{C}^{0,\beta}}\lesssim 𝔯[w](τ)+𝔯[ι(1)(𝒂)](τ)𝒞0,β2\displaystyle||\mathfrak{r}[w](\tau)+\mathfrak{r}[\iota^{(1)}(\boldsymbol{a})](\tau)||^{2}_{\mathcal{C}^{0,\beta}}
\displaystyle\lesssim w(τ)𝒞2,β2+|𝒂|2e2λIτ.\displaystyle||w(\tau)||^{2}_{\mathcal{C}^{2,\beta}}+|\boldsymbol{a}|^{2}e^{-2\lambda_{I}\tau}.

Recall our definition of the norm (3.2), multiplying the above inequality by eδ1τe^{\delta_{1}\tau} and noticing δ1>2λI-\delta_{1}>2\lambda_{I}, one can get (3.8) holds. Moreover

E(w1+ι(1)(𝒂))E(w2+ι(1)(𝒂))𝒞0,β(𝕊1×(τ1,τ))\displaystyle||E(w_{1}+\iota^{(1)}(\boldsymbol{a}))-E(w_{2}+\iota^{(1)}(\boldsymbol{a}))||_{\mathcal{C}^{0,\beta}(\mathbb{S}^{1}\times(\tau-1,\tau))}
\displaystyle\lesssim (𝔯[w1](τ)𝒞0,β+𝔯[w2](τ)𝒞0,β+|𝒂|eλIτ)(𝔯[w1]𝔯[w2])(τ)𝒞0,β\displaystyle(||\mathfrak{r}[w_{1}](\tau)||_{\mathcal{C}^{0,\beta}}+||\mathfrak{r}[w_{2}](\tau)||_{\mathcal{C}^{0,\beta}}+|\boldsymbol{a}|e^{-\lambda_{I}\tau})||(\mathfrak{r}[w_{1}]-\mathfrak{r}[w_{2}])(\tau)||_{\mathcal{C}^{0,\beta}}

which implies (3.9) holds. Thus, the claim is proved.

Define a map S:{fX(1):fX(1)<ε0}X(1)S:\{f\in X^{(1)}:||f||_{X^{(1)}}<\varepsilon_{0}\}\to X^{(1)} by S(w)=uS(w)=u where uu is the solution of

τuu=E(w+ι(1)(𝒂))on 𝕊1×\partial_{\tau}u-\mathcal{L}u=E(w+\iota^{(1)}(\boldsymbol{a}))\quad\text{on }\quad\mathbb{S}^{1}\times\mathbb{R}_{-}

with P(1)(u(,0))=0P^{(1)}(u(\cdot,0))=0. By Lemma 3.1, such uX(1)u\in X^{(1)} is unique, so S(w)S(w) is well-defined. Moreover, Lemma 3.1 says ,

uX(1)wX(1)2+|𝒂|2ε02+|𝒂|2\quad||u||_{X^{(1)}}\lesssim||w||_{X^{(1)}}^{2}+|\boldsymbol{a}|^{2}\leq\varepsilon_{0}^{2}+|\boldsymbol{a}|^{2}

and

S(w1)S(w2)X(1)(ε0+|𝒂|)w1w2X(1).||S(w_{1})-S(w_{2})||_{X^{(1)}}\lesssim(\varepsilon_{0}+|\boldsymbol{a}|)||w_{1}-w_{2}||_{X^{(1)}}.

Choosing ε0\varepsilon_{0} small enough, SS will be a contraction mapping on {fX(1):fX(1)<ε0}\{f\in X^{(1)}:||f||_{X^{(1)}}<\varepsilon_{0}\}. Therefore it has a fixed point ww which solves (3.7). Since weδ1τw\lesssim e^{\delta_{1}\tau}, we have

limτeλmτ(v(1),φm)h=limτeλmτ(ι(1)(𝒂),φm)h=am,mJ(1).\lim_{\tau\to-\infty}e^{\lambda_{m}\tau}(v^{(1)},\varphi_{m})_{h}=\lim_{\tau\to-\infty}e^{\lambda_{m}\tau}(\iota^{(1)}(\boldsymbol{a}),\varphi_{m})_{h}=a_{m},\quad m\in J^{(1)}.

Therefore, we have found v(1)v^{(1)} and (3.4) is true for l=1l=1.

Suppose that we have found v1v^{1} up to v(l)v^{(l)}, and (3.4) is established up to ll by the induction. If J(l+1)=J^{(l+1)}=\emptyset, let v(l+1)=0v^{(l+1)}=0. Obviously the theorem still holds for such v(l+1)v^{(l+1)}. If J(l+1)J^{(l+1)}\neq\emptyset, then we can find v(l+1)v^{(l+1)} by the following process. Let v(l+1)=w+ι(l+1)(𝒂)v^{(l+1)}=w+\iota^{(l+1)}(\boldsymbol{a}). Since we require j=1l+1v(j)\sum_{j=1}^{l+1}v^{(j)} is an ancient solution of (1.8), it suffices to find wX(l+1)w\in X^{(l+1)} such that

τw=w+E(l+1)(w)\displaystyle\partial_{\tau}w=\mathcal{L}w+E^{(l+1)}(w) (3.11)

where

E(l+1)(w)=E(w+ι(l+1)(𝒂)+j=1lv(j))E(j=1lv(j)).\displaystyle E^{(l+1)}(w)=E\left(w+\iota^{(l+1)}(\boldsymbol{a})+\sum_{j=1}^{l}v^{(j)}\right)-E\left(\sum_{j=1}^{l}v^{(j)}\right). (3.12)
Claim 2.

There exists ε0\varepsilon_{0} small such that if wX(l+1)+|𝐚|<ε0||w||_{X^{(l+1)}}+|\boldsymbol{a}|<\varepsilon_{0} then

E(l+1)(w)𝒞0,β,δl+1\displaystyle||E^{(l+1)}(w)||_{\mathcal{C}^{0,\beta,\delta_{l+1}}} wX(l+1)2+|𝒂|2,\displaystyle\lesssim||w||^{2}_{X^{(l+1)}}+|\boldsymbol{a}|^{2}, (3.13)
E(l+1)(w1)E(l+1)(w2)𝒞0,β,δl+1\displaystyle||E^{(l+1)}(w_{1})-E^{(l+1)}(w_{2})||_{\mathcal{C}^{0,\beta,\delta_{l+1}}} ε0w1w2X(l+1).\displaystyle\lesssim\varepsilon_{0}||w_{1}-w_{2}||_{X^{(l+1)}}. (3.14)

In fact, using (3.10), for 𝔯[w1]\mathfrak{r}[w_{1}] and 𝔯[w2]\mathfrak{r}[w_{2}] small

||[E(w1)\displaystyle||[E(w_{1}) E(w2)](τ)||𝒞0,β\displaystyle-E(w_{2})](\tau)||_{\mathcal{C}^{0,\beta}}
\displaystyle\lesssim (𝔯[w1](τ)𝒞0,β+𝔯[w2](τ)𝒞0,β)[𝔯[w1]𝔯[w2]](τ)𝒞0,β.\displaystyle(||\mathfrak{r}[w_{1}](\tau)||_{\mathcal{C}^{0,\beta}}+||\mathfrak{r}[w_{2}](\tau)||_{\mathcal{C}^{0,\beta}})||[\mathfrak{r}[w_{1}]-\mathfrak{r}[w_{2}]](\tau)||_{\mathcal{C}^{0,\beta}}.

This implies

E(l+1)(w)(τ)𝒞0,β\displaystyle||E^{(l+1)}(w)(\tau)||_{\mathcal{C}^{0,\beta}}\lesssim (w(τ)𝒞2,β+|𝒂|eλIτ)(w(τ)𝒞2,β+|𝒂|e(l+1)λIτ)\displaystyle(||w(\tau)||_{\mathcal{C}^{2,\beta}}+|\boldsymbol{a}|e^{-\lambda_{I}\tau})(||w(\tau)||_{\mathcal{C}^{2,\beta}}+|\boldsymbol{a}|e^{-(l+1)\lambda_{I}\tau})
\displaystyle\lesssim e(l+2)λIτ(w𝒞2,β,λI+|𝒂|)(w𝒞2,β,(l+1)λI+|𝒂|).\displaystyle e^{-(l+2)\lambda_{I}\tau}(||w||_{\mathcal{C}^{2,\beta,-\lambda_{I}}}+|\boldsymbol{a}|)(||w||_{\mathcal{C}^{2,\beta,-(l+1)\lambda_{I}}}+|\boldsymbol{a}|).

Recalling (3.2) and (l+2)λI<δl+1<(l+1)λI(l+2)\lambda_{I}<-\delta_{l+1}<(l+1)\lambda_{I}, one can see that (3.13) holds. The proof of (3.14) can be derived similarly.

Define a map S:{fX(l+1):fX(l+1)<ε0}X(l+1)S:\{f\in X^{(l+1)}:||f||_{X^{(l+1)}}<\varepsilon_{0}\}\to X^{(l+1)} by S(w)=uS(w)=u where uu is the unique solution of

τuu=E(l+1)(w)on 𝕊1×\partial_{\tau}u-\mathcal{L}u=E^{(l+1)}(w)\quad\text{on }\quad\mathbb{S}^{1}\times\mathbb{R}_{-}

with P(l+1)(u(,0))=0P^{(l+1)}(u(\cdot,0))=0. Taking ε0\varepsilon_{0} small enough, SS is a contraction mapping. Note that (3.4) is satisfied for mJ(l+1)m\in J^{(l+1)}. The existence of v(l+1)v^{(l+1)} is established.

Finally, the uniqueness and continuity of v(l)v^{(l)} also follow from the contraction mapping theorem. ∎

According to Theorem 3.2, we can have II-parameter family of ancient solutions, but only I3I-3 of them are important, because the first three of them depends on the time and space center we choose in the renormalization (1.5). That is, one can always just shift the non-rescaled ancient solution by time and space to edit the first three parameters.

Proposition 3.4.

Let Γt\Gamma_{t} be an α\alpha-CSF asymptotic to Γ\Gamma after rescaling where Γ=Γ¯αk\Gamma=\bar{\Gamma}_{\alpha}^{k} or Γ¯αc\bar{\Gamma}_{\alpha}^{c}, and let u¯\bar{u} denote the support function of the rescaled flow Γ¯τ\bar{\Gamma}_{\tau}. Then, given B=(b1,b2,b3)3B=(b_{1},b_{2},b_{3})\in\mathbb{R}^{3} the ancient flow

ΓtB=Γt+b1+(b2,b3)2\Gamma^{B}_{t}=\Gamma_{t+b_{1}}+(b_{2},b_{3})\subset\mathbb{R}^{2} (3.15)

satisfies

u¯B(θ,τ)u¯(θ,τ)\displaystyle\bar{u}^{B}(\theta,\tau)-\bar{u}(\theta,\tau)
=(1+α)1α+1[b2cosθ+b3sinθ]eτ+b11+αe(1+α)τh+o(e(1+α)τ)\displaystyle=(1+\alpha)^{-\frac{1}{\alpha+1}}[b_{2}\cos\theta+b_{3}\sin\theta]e^{\tau}+\frac{b_{1}}{1+\alpha}e^{(1+\alpha)\tau}h+o(e^{(1+\alpha)\tau})

where u¯B\bar{u}^{B} denotes the support function of the rescaled flow Γ¯τB\bar{\Gamma}_{\tau}^{B}. Consequently if Γ¯τ\bar{\Gamma}_{\tau} is constructed from 𝐚\boldsymbol{a}, then Γ¯τB\bar{\Gamma}_{\tau}^{B} is constructed from using

𝒂+(b11+α,(1+α)1α+1b2,(1+α)1α+1b3,0,,0).\boldsymbol{a}+\left(\frac{b_{1}}{1+\alpha},(1+\alpha)^{-\frac{1}{\alpha+1}}b_{2},(1+\alpha)^{-\frac{1}{\alpha+1}}b_{3},0,\cdots,0\right).
Proof.

Our assumption implies

uB(θ,t)=u(θ,t+b1)+b2cosθ+b3sinθu^{B}(\theta,t)=u(\theta,t+b_{1})+b_{2}\cos\theta+b_{3}\sin\theta

where uBu^{B} and uu are the support functions of ΓtB\Gamma_{t}^{B} and Γt\Gamma_{t}. It follows from (1.5) that

u¯B(θ,τ)=\displaystyle\bar{u}^{B}(\theta,\tau)= (1+α)1α+1eτuB(θ,e(1+α)τ)\displaystyle(1+\alpha)^{-\frac{1}{\alpha+1}}e^{\tau}u^{B}(\theta,-e^{-(1+\alpha)\tau})
=\displaystyle= (1+α)1α+1eτ[u(θ,e(1+α)τ+b1)+b2cosθ+b3sinθ]\displaystyle(1+\alpha)^{-\frac{1}{\alpha+1}}e^{\tau}\left[u(\theta,-e^{-(1+\alpha)\tau}+b_{1})+b_{2}\cos\theta+b_{3}\sin\theta\right]
=\displaystyle= eττ1u¯(θ,τ1)+(1+α)1α+1[b2cosθ+b3sinθ]eτ\displaystyle e^{\tau-\tau_{1}}\bar{u}(\theta,\tau_{1})+(1+\alpha)^{-\frac{1}{\alpha+1}}[b_{2}\cos\theta+b_{3}\sin\theta]e^{\tau} (3.16)

where

τ1=11+αlog(e(1+α)τb1).\tau_{1}=\frac{-1}{1+\alpha}\log(e^{-(1+\alpha)\tau}-b_{1}).

As τ\tau\to-\infty, we have τ1=τ+(1+α)1b1e(1+α)τ+o(e(1+α)τ)\tau_{1}=\tau+(1+\alpha)^{-1}b_{1}e^{(1+\alpha)\tau}+o(e^{(1+\alpha)\tau}). Consequently eττ1=1+(1+α)1b1e(1+α)τ+o(e(1+α)τ)e^{\tau-\tau_{1}}=1+(1+\alpha)^{-1}b_{1}e^{(1+\alpha)\tau}+o(e^{(1+\alpha)\tau}).

Since Γ¯τ\bar{\Gamma}_{\tau} converges to some self-shrinker Γ\Gamma with support function hh, u¯(θ,τ)h\bar{u}(\theta,\tau)\to h as τ\tau\to-\infty. Plugging the above information of τ1\tau_{1} to (3.16), one gets the conclusion. ∎

Proof of Theorem 1.3.

It follows from Theorem 3.2 that there exists a map

𝒮:Bε0(I)\displaystyle\mathcal{S}:B_{\varepsilon_{0}}(\subset\mathbb{R}^{I}) C2,β(𝕊1×(,0])\displaystyle\to C^{2,\beta}(\mathbb{S}^{1}\times(-\infty,0])
𝒂=(a1,,aI)\displaystyle\boldsymbol{a}=(a_{1},\cdots,a_{I}) j=1Lv(j)\displaystyle\mapsto\sum_{j=1}^{L}v^{(j)}

such that 𝒮(𝒂)=j=1Lv(j)\mathcal{S}(\boldsymbol{a})=\sum_{j=1}^{L}v^{(j)} is an ancient solution of (1.8).

Suppose that 𝒂,𝒃I\boldsymbol{a},\boldsymbol{b}\in\mathbb{R}^{I} satisfy akbk0a_{k}-b_{k}\neq 0 and aibi=0a_{i}-b_{i}=0 for all i>ki>k. Assume λkJ(l+1)\lambda_{k}\in J^{(l+1)} for some ll. Then for any j{1,,l}j\in\{1,\cdots,l\}, careful tracking the proof of Theorem 3.2 shows that v𝒂(j)=v𝒃(j)v^{(j)}_{\boldsymbol{a}}=v^{(j)}_{\boldsymbol{b}}, because of the uniqueness of them obtained through the contraction mapping theorem. In the step to find v𝒂(l+1)=ι(l+1)(𝒂)+w𝒂v^{(l+1)}_{\boldsymbol{a}}=\iota^{(l+1)}(\boldsymbol{a})+w_{\boldsymbol{a}} and v𝒂(l+1)=ι(l+1)(𝒃)+w𝒃v^{(l+1)}_{\boldsymbol{a}}=\iota^{(l+1)}(\boldsymbol{b})+w_{\boldsymbol{b}}, we recall that w𝒂,w𝒃X(l+1)w_{\boldsymbol{a}},w_{\boldsymbol{b}}\in X^{(l+1)}. Therefore

𝒮(𝒂)(θ,τ)𝒮(𝒃)(θ,τ)\displaystyle\mathcal{S}(\boldsymbol{a})(\theta,\tau)-\mathcal{S}(\boldsymbol{b})(\theta,\tau) =v𝒂(l+1)v𝒃(l+1)+O(eδl+1τ)\displaystyle=v^{(l+1)}_{\boldsymbol{a}}-v^{(l+1)}_{\boldsymbol{b}}+O(e^{-\delta_{l+1}\tau})
=(akbk)eλkτφk(θ)+O(eλk1τ)+O(eδl+1τ)\displaystyle=(a_{k}-b_{k})e^{-\lambda_{k}\tau}\varphi_{k}(\theta)+O(e^{-\lambda_{k-1}\tau})+O(e^{-\delta_{l+1}\tau})
=(akbk)eλkτφk(θ)+o(eλkτ)\displaystyle=(a_{k}-b_{k})e^{-\lambda_{k}\tau}\varphi_{k}(\theta)+o(e^{-\lambda_{k}\tau})

when λk1<λk\lambda_{k-1}<\lambda_{k}, and

𝒮(𝒂)(θ,τ)𝒮(𝒃)(θ,τ)=eλkτi=k1k(aibi)φi(θ)+o(eλkτ)\displaystyle\mathcal{S}(\boldsymbol{a})(\theta,\tau)-\mathcal{S}(\boldsymbol{b})(\theta,\tau)=e^{-\lambda_{k}\tau}\sum_{i=k-1}^{k}(a_{i}-b_{i})\varphi_{i}(\theta)+o(e^{-\lambda_{k}\tau})

when λk1=λk\lambda_{k-1}=\lambda_{k}, where φi\varphi_{i} are eigenfunctions of Γ\mathcal{L}_{\Gamma} with the eigenvalue λi\lambda_{i} and φi,φjLh2=δij\langle\varphi_{i},\varphi_{j}\rangle_{L^{2}_{h}}=\delta_{ij}.

Proposition 3.4 says it suffices to have the map for 𝒂=(0,0,0,a3,,aI)\boldsymbol{a}=(0,0,0,a_{3},\cdots,a_{I}), because the other ancient solutions can be generated by these ones from a different choice of time and space center in the renormalization. Removing the first three zeros of 𝒂\boldsymbol{a} and reindexing each components, we abuse the notation by still denoting 𝒂=(a1,,aI3)\boldsymbol{a}=(a_{1},\cdots,a_{I-3}). Therefore we have a map 𝒮:Bϵ0(0)I3C2,β(𝕊1×(,0])\mathcal{S}:B_{\epsilon_{0}}(0)\subset\mathbb{R}^{I-3}\to C^{2,\beta}(\mathbb{S}^{1}\times(-\infty,0]) with the desired property. ∎

Notice that in Theorem 3.2 there is a restriction |𝒂|<ε0|\boldsymbol{a}|<\varepsilon_{0}. It is possible to get around this by translating in the τ\tau (which is equivalent to parabolic scaling in the corresponding non-rescaled ancient solution). However, one has to pay the price that these ancient solutions may not live up to τ=0\tau=0.

Theorem 3.5.

There exists a continuous map 𝒮\mathcal{S} (define in (3.18)) which maps any 𝐚I\boldsymbol{a}\in\mathbb{R}^{I} to 𝒞2,β,λI(𝕊1×(,T(𝐚))\mathcal{C}^{2,\beta,-\lambda_{I}}(\mathbb{S}^{1}\times(-\infty,T({\boldsymbol{a}})) such that 𝒮(a)\mathcal{S}(a) is an ancient solution of (1.8) on (,T(𝐚)](-\infty,T({\boldsymbol{a}})], where T(𝐚)T(\boldsymbol{a}) is defined in (3.17). Moreover there is a unique decomposition 𝒮(a)=l=1Lv(l)\mathcal{S}(a)=\sum_{l=1}^{L}v^{(l)} with

limτeλmτ(v(l)(,τ),φm)h=am,m satisfying λm((l+1)λI,lλI]\lim_{\tau\to-\infty}e^{\lambda_{m}\tau}(v^{(l)}(\cdot,\tau),\varphi_{m})_{h}=a_{m},\quad\forall\,m\text{ satisfying }\lambda_{m}\in((l+1)\lambda_{I},l\lambda_{I}]

for any l=1,,Ll=1,\cdots,L.

Proof.

For any 𝒂I\boldsymbol{a}\in\mathbb{R}^{I}, let

T(𝒂)=12(1+α)max{logε0|𝒂|2,0}\displaystyle T(\boldsymbol{a})=\frac{1}{2(1+\alpha)}\max\{\log\frac{\varepsilon_{0}}{|\boldsymbol{a}|^{2}},0\} (3.17)

then i=1Ie2λmTam2<ε0\sum_{i=1}^{I}e^{-2\lambda_{m}T}a_{m}^{2}<\varepsilon_{0}. Then by the previous theorem, one can find {v(l)}l=1L\{v^{(l)}\}_{l=1}^{L} such that l=1Lv(l)\sum_{l=1}^{L}v^{(l)} is an ancient solution of (1.8) on 𝕊1×(,0]\mathbb{S}^{1}\times(-\infty,0] and

limτeλmτ(v(l)(,τ),φm)h=eλmTam,mJ(l).\lim_{\tau\to-\infty}e^{\lambda_{m}\tau}(v^{(l)}(\cdot,\tau),\varphi_{m})_{h}=e^{-\lambda_{m}T}a_{m},\quad m\in J^{(l)}.

So we define a map 𝒮\mathcal{S} by translating v(l)v^{(l)}

𝒮(a)(,τ)=l=1Lv(l)(,τT).\displaystyle\mathcal{S}(a)(\cdot,\tau)=\sum_{l=1}^{L}v^{(l)}(\cdot,\tau-T). (3.18)

One can easily verify that 𝒮(a)\mathcal{S}(a) is an ancient solution of (1.8) on 𝕊1×(,T]\mathbb{S}^{1}\times(-\infty,-T] and

limτeλmτ(v(l)(,τT),φm)h=am,mJ(l).\lim_{\tau\to-\infty}e^{\lambda_{m}\tau}(v^{(l)}(\cdot,\tau-T),\varphi_{m})_{h}=a_{m},\quad m\in J^{(l)}.

References

  • Andrews [2002] Ben Andrews. Non-convergence and instability in the asymptotic behaviour of curves evolving by curvature. Communications in Analysis and Geometry, 10(2):409–449, 2002.
  • Andrews [2003] Ben Andrews. Classification of limiting shapes for isotropic curve flows. Journal of the American mathematical society, 16(2):443–459, 2003.
  • Andrews et al. [2016] Ben Andrews, Pengfei Guan, and Lei Ni. Flow by powers of the gauss curvature. Advances in Mathematics, 299:174–201, 2016.
  • Angenent [1992] Sigurd Angenent. Shrinking doughnuts. In Nonlinear diffusion equations and their equilibrium states, 3, pages 21–38. Springer, 1992.
  • Angenent et al. [2019a] Sigurd Angenent, Simon Brendle, Panagiota Daskalopoulos, and Natasa Sesum. Unique asymptotics of compact ancient solutions to three-dimensional ricci flow. arXiv preprint arXiv:1911.00091, 2019a.
  • Angenent et al. [2019b] Sigurd Angenent, Panagiota Daskalopoulos, and Natasa Sesum. Unique asymptotics of ancient convex mean curvature flow solutions. Journal of Differential Geometry, 111(3):381–455, 2019b.
  • Angenent et al. [2018] Sigurd B Angenent, Panagiota Daskalopoulos, and Natasa Sesum. Uniqueness of two-convex closed ancient solutions to the mean curvature flow. arXiv preprint arXiv:1804.07230, 2018.
  • Bourni et al. [2019] Theodora Bourni, Mat Langford, and Giuseppe Tinaglia. Convex ancient solutions to curve shortening flow. arXiv preprint arXiv:1903.02022, 2019.
  • Bourni et al. [2020] Theodora Bourni, Julie Clutterbuck, Xuan Hien Nguyen, Alina Stancu, Guofang Wei, and Valentina-Mira Wheeler. Ancient solutions for flow by powers of the curvature in 2\mathbb{R}^{2}. arXiv preprint arXiv:2005.07642, 2020.
  • Brendle [2018] Simon Brendle. Ancient solutions to the ricci flow in dimension 3. arXiv preprint arXiv:1811.02559, 2018.
  • Brendle and Choi [2018] Simon Brendle and Kyeongsu Choi. Uniqueness of convex ancient solutions to mean curvature flow in higher dimensions. arXiv preprint arXiv:1804.00018, 2018.
  • Brendle and Choi [2019] Simon Brendle and Kyeongsu Choi. Uniqueness of convex ancient solutions to mean curvature flow in 3\mathbb{R}^{3}. Inventiones mathematicae, 217(1):35–76, 2019.
  • Brendle et al. [2020] Simon Brendle, Panagiota Daskalopulos, and Natasa Sesum. Uniqueness of compact ancient solutions to three-dimensional ricci flow. arXiv preprint arXiv:2002.12240, 2020.
  • Caffarelli et al. [1984] Luis Caffarelli, Robert Hardt, and Leon Simon. Minimal surfaces with isolated singularities. Manuscripta Mathematica, 48(1):1432–1785, 1984.
  • Chen [2015] Shibing Chen. Classifying convex compact ancient solutions to the affine curve shortening flow. The Journal of Geometric Analysis, 25(2):1075–1079, 2015.
  • Chodosh et al. [2020] Otis Chodosh, Kyeongsu Choi, Christos Mantoulidis, and Felix Schulze. Mean curvature flow with generic initial data. arXiv preprint arXiv:2003.14344, 2020.
  • Choi et al. [2020] Beomjun Choi, Kyeongsu Choi, and Panagiota Daskalopoulos. Uniqueness of ancient solutions to gauss curvature flow asymptotic to a cylinder. arXiv preprint arXiv:2004.11754, 2020.
  • Choi and Mantoulidis [2019] Kyeongsu Choi and Christos Mantoulidis. Ancient gradient flows of elliptic functionals. arXiv preprint arXiv:1902.07697, 2019.
  • Choi et al. [2018] Kyeongsu Choi, Robert Haslhofer, and Or Hershkovits. Ancient low entropy flows, mean convex neighborhoods, and uniqueness. arXiv preprint arXiv:1810.08467, 2018.
  • Choi et al. [2019] Kyeongsu Choi, Robert Haslhofer, Or Hershkovits, and Brian White. Ancient asymptotically cylindrical flows and applications. arXiv preprint arXiv:1910.00639, 2019.
  • Courant and Hilbert [1953] R. Courant and D. Hilbert. Methods of Mathematical Physics. Number v. 1. Wiley, 1953.
  • Daskalopoulos et al. [2010] Panagiota Daskalopoulos, Richard Hamilton, and Natasa Sesum. Classification of compact ancient solutions to the curve shortening flow. J. Differential Geom., 84(3):455–464, 03 2010.
  • Ivaki [2016] Mohammad N Ivaki. Classification of compact convex ancient solutions of the planar affine normal flow. The Journal of Geometric Analysis, 26(1):663–671, 2016.
  • Loftin and Tsui [2008] John Loftin and Mao-Pei Tsui. Ancient solutions of the affine normal flow. Journal of Differential Geometry, 78(1):113–162, 2008.
  • Lunardi [2012] Alessandra Lunardi. Analytic semigroups and optimal regularity in parabolic problems. Springer Science & Business Media, 2012.
  • Wang [2011] Xu-Jia Wang. Convex solutions to the mean curvature flow. Annals of mathematics, pages 1185–1239, 2011.