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

Examples of compact embedded convex λ\lambda-hypersurfaces

Qing-Ming Cheng, Junqi Lai and Guoxin Wei Qing-Ming Cheng
Department of Applied Mathematics, Faculty of Science, Fukuoka University, 814-0180, Fukuoka, Japan, [email protected]
Junqi Lai
School of Mathematical Sciences, South China Normal University, 510631, Guangzhou, China, [email protected]
Guoxin Wei
School of Mathematical Sciences, South China Normal University, 510631, Guangzhou, China, [email protected]
Abstract.

In the paper, we construct compact embedded convex λ\lambda-hypersurfaces which are diffeomorphic to a sphere and are not isometric to a standard sphere. As the special case of our result, we solve Sun’s problem (Int Math Res Not: 11818-11844, 2021). In this sense, one can not expect to have Alexandrov type theorem for λ\lambda-hypersurfaces.

footnotetext: The first author was partially supported by JSPS Grant-in-Aid for Scientific Research (B): No.16H03937 and No.22K03303. The third author was partly supported by NSFC Grant No.12171164, GDUPS (2018).

1. Introduction

A hypersurface Σnn+1\Sigma^{n}\subset\mathbb{R}^{n+1} is called a λ\lambda-hypersurface if it satisfies

(1.1) H+X,ν=λ,H+{\langle}X,\nu{\rangle}=\lambda,

where λ\lambda is a constant, XX is the position vector, ν\nu is an inward unit normal vector and HH is the mean curvature. The notation of λ\lambda-hypersurfaces were first introduced by Cheng and Wei in [5] (also see [17]). Cheng and Wei [5] proved that λ\lambda-hypersurfaces are critical points of the weighted area functional with respect to weighted volume-preserving variations. This equation of λ\lambda-hypersurfaces also arises in the study of isoperimetric problems in weighted (Gaussian) Euclidean spaces, which is a long-standing topic studied in various fields in science. λ\lambda-hypersurfaces can also be viewed as stationary solutions to the isoperimetric problem in the Gaussian space. For more information on λ\lambda-hypersurfaces, one can see [5] and [17].

Firstly, we give some examples of λ\lambda-hypersurfaces. It is well known that there are several special complete embedded solutions to (1.1):

  • hyperplanes with a distance of |λ||\lambda| from the origin,

  • sphere with radius λ+λ2+4n2\frac{-\lambda+\sqrt{\lambda^{2}+4n}}{2} centered at origin,

  • cylinders with an axis through the origin and radius λ+λ2+4(n1)2\frac{-\lambda+\sqrt{\lambda^{2}+4(n-1)}}{2}.

In [6], Cheng and Wei constructed the first nontrivial example of a λ\lambda-hypersurface which is diffeomorphic to 𝕊n1×𝕊1\mathbb{S}^{n-1}\times\mathbb{S}^{1} by using techniques similar to Angenent [2]. In [18], using a similar method to McGrath [16], Ross constructed a λ\lambda-hypersurface in 2n+2\mathbb{R}^{2n+2} which is diffeomorphic to 𝕊n×𝕊n×𝕊1\mathbb{S}^{n}\times\mathbb{S}^{n}\times\mathbb{S}^{1} and exhibits a SO(n)×SO(n)SO(n)\times SO(n) rotational symmetry. In [15], Li and Wei constructed an immersed SnS^{n} λ\lambda-hypersurface using a similar method to [9]. It is quite interesting to find other nontrivial examples of λ\lambda-hypersurfaces.

Secondly, we introduce some rigidity results about λ\lambda-hypersurfaces. If λ=0\lambda=0, X,ν+H=λ=0\langle X,\nu\rangle+H=\lambda=0, then X:Σnn+1X:\Sigma^{n}\to\mathbb{R}^{n+1} is a self-shrinker of mean curvature flow, which plays an important role for study on singularities of the mean curvature flow. Abresch and Langer [1] proved that the only 11-dimensional compact embedded self-shrinker is the circle. Huisken [14] proved any compact embedded, and mean convex (H0H\geq 0) self-shrinkers are spheres. Later, Colding and Minicozzi [7] generalized Huisken’s results to the case of complete self-shrinkers.

If λ0\lambda\neq 0, there are relatively few results about λ\lambda-hypersurfaces. In [5], Cheng and Wei proved that generalized cylinders 𝕊m×nm\mathbb{S}^{m}\times\mathbb{R}^{n-m}, 0mn0\leq m\leq n are the only complete embedded λ\lambda-hypersurfaces with polynomial volume growth in n+1\mathbb{R}^{n+1} if Hλ0H-\lambda\geq 0 and λ(f3(Hλ)S)0\lambda(f_{3}(H-\lambda)-S)\geq 0, where f3=i,j,k=1nhijhjkhkif_{3}=\sum_{i,j,k=1}^{n}h_{ij}h_{jk}h_{ki}, S=i,j=1nhij2S=\sum_{i,j=1}^{n}h_{ij}^{2}, hijh_{ij} denotes the components of the second fundamental form. This classification result generalizes the result of Huisken [14] and Colding and Minicozzi [7].

For λ>0\lambda>0, in [13], Heilman proved that convex nn-dimensional λ\lambda-hypersurfaces are generalized cylinders if λ>0\lambda>0, which generalized the rigidity result of Colding and Minicozzi [7] to λ\lambda-hypersurfaces with λ>0\lambda>0.

The case λ<0\lambda<0 is much more complicated. In [12], by using the explicit expressions of the derivatives of the principal curvatures at the non-umbilical points of the surface, Guang obtained that any strictly mean convex 22-dimensional λ\lambda-hypersurfaces are convex if λ0\lambda\leq 0. Later, by using the maximum principle, Lee [11] showed that any compact embedded and mean convex nn-dimensional λ\lambda-hypersurfaces are convex if λ0\lambda\leq 0. In [4], Chang constructed 11-dimensional λ\lambda-curves in 2\mathbb{R}^{2}, which are not circles. These surprising examples show that λ\lambda-surfaces behave very differently when λ<0\lambda<0 compared to the case λ>0\lambda>0. New techniques must be introduced to study this phenomenon. We imagine these examples may carry meaningful information in probability theory (also see [19]).

In [19], Sun developed the compactness theorem for λ\lambda-surface in 3\mathbb{R}^{3} with uniform λ\lambda and genus. As the application of the compactness theorem, he also showed a rigidity theorem for convex λ\lambda-surfaces. In the same paper, he [19] proposed the problem that constructing a compact convex λ\lambda-surface which is not a sphere (see Question 4.0.4. on page 25, [19]).

In this paper, motivated by [9, 10, 15, 19], we construct nontrivial embedded λ\lambda-hypersurfaces which are diffeomorphic to 𝕊n\mathbb{S}^{n} and they are not isometric to a standard sphere. As the special case of our result, that is, n=2n=2, we solve Sun’s problem. In fact, we obtain the following theorem:

Theorem 1.1.

For n2n\geq 2 and 2n+2<λ<0-\frac{2}{\sqrt{n+2}}<\lambda<0, there exists an embedded convex λ\lambda-hypersurface Σnn+1\Sigma^{n}\subset\mathbb{R}^{n+1} which is diffeomorphic to 𝕊n\mathbb{S}^{n} and is not isometric to a standard sphere.

Remark 1.1.

It is well-known that for compact embedded hypersurfaces with constant mean curvature in n+1\mathbb{R}^{n+1}, Alexandrov theorem holds, that is, a compact embedded hypersurface with constant mean curvature in n+1\mathbb{R}^{n+1} is isometric to a round sphere. But for λ\lambda-hypersurfaces, one can not expect to have Alexandrov type theorem for λ\lambda-hypersurfaces according to the above theorem 1.1.

Remark 1.2.

For self-shrinkers, there is a well-known conjecture asserts that the round sphere should be the only embedded self-shrinker which is diffeomorphic to a sphere. Brendle [3] proved the above conjecture for 22-dimension self-shrinker. For the higher dimensional self-shrinker, the conjecture is still open. But for λ\lambda-hypersurfaces, we can construct compact embedded λ\lambda-hypersurface which is diffeomorphic to a sphere and is not isometric to a round sphere.

Remark 1.3.

For 23<λ<0-\frac{2}{\sqrt{3}}<\lambda<0, Chang [4] proved that there exists a compact embedded λ\lambda-curve with 22-symmetry in 2\mathbb{R}^{2}, which is not a circle.

2. Preliminaries

Let SO(n)SO(n) denote the special orthogonal group and act on n+1={(x,y):x,yn}\mathbb{R}^{n+1}=\left\{(x,y):x\in\mathbb{R},y\in\mathbb{R}^{n}\right\} in the usual way, then we can identify the space of orbits n+1/SO(n)\mathbb{R}^{n+1}/SO(n) with the half plane ={(x,r)2:x,r0}\mathbb{H}=\left\{(x,r)\in\mathbb{R}^{2}:x\in\mathbb{R},r\geq 0\right\} under the projection (see [18])

Π(x,y)=(x,|y|)=(x,r).\Pi(x,y)=(x,|y|)=(x,r).

If a hypersurface Σ\Sigma is invariant under the action SO(n)SO(n), then the projection Π(Σ)\Pi(\Sigma) will give us a profile curve in the half plane, which can be parametrized by Euclidean arc length and write as γ(s)=(x(s),r(s))\gamma(s)=(x(s),r(s)). Conversely, if we have a curve γ(s)=(x(s),r(s)),s(a,b)\gamma(s)=(x(s),r(s)),\ \ s\in(a,b) parametrized by Euclidean arc length in the half plane, then we can reconstruct the hypersurface by

X:(a,b)×Sn1(1)n+1,\displaystyle{X:(a,b)\times S^{n-1}(1)\hookrightarrow\mathbb{R}^{n+1}},
(2.1) (s,α)(x(s),r(s)α).\displaystyle(s,\alpha)\mapsto(x(s),r(s)\alpha).

Let

(2.2) ν=(r˙,x˙α),\nu=(-\dot{r},\dot{x}\,\alpha),

where the dot denotes taking derivative with respect to arc length ss. A direct calculation shows that ν\nu is an inward unit normal vector for the hypersurface. Then we can calculate that the principal curvatures of hypersurface (see [6, 8]):

(2.3) κi\displaystyle\kappa_{i} =x˙r,i=1, 2,,n1,\displaystyle=-\frac{\dot{x}}{r},\ \ \ \ i=1,\,2,\,\dots,\,n-1,
κn\displaystyle\kappa_{n} =x˙r¨x¨r˙.\displaystyle=\dot{x}\,\ddot{r}-\ddot{x}\,\dot{r}.

Hence the mean curvature vector equals to

(2.4) H=Hν,hereH=i=1nκi=x˙r¨x¨r˙(n1)x˙r,\overrightarrow{H}=H\nu,\ \ \ \ {\mbox{h}ere}\ \ H=\sum_{i=1}^{n}\kappa_{i}=\dot{x}\,\ddot{r}-\ddot{x}\,\dot{r}-(n-1)\frac{\dot{x}}{r},

and then, by (2.1), (2.2) and (2.4), equation (1.1) reduces to (see also [6, 10, 18])

(2.5) x˙r¨x¨r˙=(n1rr)x˙+xr˙+λ,\dot{x}\,\ddot{r}-\ddot{x}\,\dot{r}=(\frac{n-1}{r}-r)\dot{x}+x\,\dot{r}+\lambda,

where (x˙)2+(r˙)2=1(\dot{x})^{2}+(\dot{r})^{2}=1. Let θ(s)\theta(s) denote the angles between the tangent vectors of the profile curve and xx-axis, then (2.5) can be written as the following system of differential equation:

(2.6) {x˙=cosθ,r˙=sinθ,θ˙=(n1rr)cosθ+xsinθ+λ.\left\{\begin{aligned} \dot{x}&=\cos\theta,\\ \dot{r}&=\sin\theta,\\ \dot{\theta}&=(\frac{n-1}{r}-r)\cos\theta+x\,\sin\theta+\lambda.\end{aligned}\right.

Let PP denote the projection from ×\mathbb{H}\times\mathbb{R} to \mathbb{H}. Obviously, if γ(s)\gamma(s) is a solution of (2.6), then the curve P(γ(s))P(\gamma(s)) will generate a λ\lambda-hypersurface by (2.1). If we can, by this way, find a curve that starts and ends on the xx-axis and is perpendicular to the xx-axis at both ends, then we obtain an embedded λ\lambda-hypersurface, which is diffeomorphic to a sphere. Hence, Theorem 1.1 will be proved. By some symmetry of (2.6), a curve that starts perpendicularly on the xx-axis and ends perpendicularly on the rr-axis will also solve the problem. This paper’s main goal is to find the latter.
Letting (x0,r0,θ0)(x_{0},r_{0},\theta_{0}) be a point in {(x,r,θ):x,r>0,θ}\{(x,r,\theta):x\in\mathbb{R},r>0,\theta\in\mathbb{R}\}, by an existence and uniqueness theorem of the solutions for first order ordinary differential equations, there is a unique solution Γ(x0,r0,θ0)(s)\Gamma(x_{0},r_{0},\theta_{0})(s) to (2.6) satisfying initial conditions Γ(x0,r0,θ0)(0)=(x0,r0,θ0)\Gamma(x_{0},r_{0},\theta_{0})(0)=(x_{0},r_{0},\theta_{0}). Moreover, the solution depends smoothly on the initial conditions. Note that (2.6) has a singularity at r=0r=0 due to the term 1r\frac{1}{r}. But there is also an existence and uniqueness theorem for (2.6) with initial conditions at singularity and θ0=π/2\theta_{0}=\pi/2. And the previously mentioned smooth dependence can also extend to singularity smoothly (see [9]).

When the profile curve can be written in the form (x,u(x))(x,u(x)), by (2.6), the function u(x)u(x) satisfies the differential equation

(2.7) u′′1+(u)2=xuu+n1u+λ1+(u)2.\frac{u^{\prime\prime}}{1+(u^{\prime})^{2}}=x\,u^{\prime}-u+\frac{n-1}{u}+\lambda\sqrt{1+(u^{\prime})^{2}}.

When the profile curve can be written in the form (f(r),r)(f(r),r), by (2.6), the function f(r)f(r) satisfies the differential equation

(2.8) f′′1+(f)2=(rn1r)ffλ1+(f)2.\frac{f^{\prime\prime}}{1+(f^{\prime})^{2}}=(r-\frac{n-1}{r})f^{\prime}-f-\lambda\sqrt{1+(f^{\prime})^{2}}.

Next we consider equations (2.6) in polar coordinates, when the profile curve can be written in the form ρ=ρ(ϕ)\rho=\rho(\phi), where ρ=x2+r2\rho=\sqrt{x^{2}\,+\,r^{2}} and ϕ=arctan(r/x)\phi\,=\,\arctan(r/x), the function ρ(ϕ)\rho(\phi) satisfies the differential equation

(2.9) ρ′′=1ρ{ρ2+(ρ2+ρ2)[nρ2(n 1)ρρcotϕλρ2+ρ2]}.\rho^{\prime\prime}\,=\,\frac{1}{\rho}\left\{\rho^{\prime 2}\,+\,(\rho^{2}\,+\,\rho^{\prime 2})\left[n\,-\,\rho^{2}\,-\,(n\,-\,1)\frac{\rho^{\prime}}{\rho}\cot\phi\,-\,\lambda\,\sqrt{\rho^{2}\,+\,\rho^{\prime 2}}\right]\right\}.

Note that (2.8) has a solution f=λf=-\lambda, which corresponds to hyperplane, (2.9) has a solution ρ=λ+λ2+ 4n2\rho\,=\,\frac{-\lambda\,+\,\sqrt{\lambda^{2}\,+\,4\,n}}{2}, which corresponds to the round sphere.
We conclude this section with a lemma about the solutions of (2.8), this lemma can be found in [15].

Lemma 2.1 ([15]).

Let ff be a solution of (2.8) with f(0)>λf(0)>-\lambda and f(0)=0f^{\prime}(0)=0. Then we have f′′<0f^{\prime\prime}<0 and there exists a point r<r_{*}<\infty so that limrrf(r)=\lim_{r\rightarrow r_{*}}f^{\prime}(r)=-\infty and limrrf(r)>\lim_{r\rightarrow r_{*}}f(r)>-\infty, i.e., ff blows-up at rr_{*}.

3. Behavior near the known solutions

In order to solve the problem, we need to study the behavior near two known solutions first, and use continuity argument to find the curve that we want to have.

3.1. Behavior near the plane

Let fϵ(r)f_{\epsilon}(r) be the solution of (2.8) with fϵ(0)=λ+ϵf_{\epsilon}(0)=-\lambda+\epsilon and fϵ(0)=0f_{\epsilon}^{\prime}(0)=0 where denotes ddr\frac{{\rm d}}{{\rm d}r}. For ϵ>0\epsilon>0, by Lemma 2.1, let rϵr_{*}^{\epsilon} denote the point where fϵf_{\epsilon} blows-up, and let xϵ=fϵ(rϵ)x_{*}^{\epsilon}=f_{\epsilon}(r_{*}^{\epsilon}). Let hϵ(r)=fϵ(r)+λh_{\epsilon}(r)=f_{\epsilon}(r)+\lambda, we may write hϵ(r)h_{\epsilon}(r) to h(r)h(r), fϵ(r)f_{\epsilon}(r) to f(r)f(r), rϵr_{*}^{\epsilon} to rr_{*} and xϵx_{*}^{\epsilon} to xx_{*} when there is no ambiguity. We introduce the following proposition which gives a description of the behavior near the plane f0=λf_{0}=-\lambda.

Proposition 3.1.

For any fixed n2n\geq 2, min{23n+428n,25n630n}λ<0-\min\left\{\frac{23n+4}{28\sqrt{n}},\frac{25n-6}{30\sqrt{n}}\right\}\leq\lambda<0, there exists ϵ¯>0\bar{\epsilon}>0 so that

rϵlog1πϵ,30n+4log1πϵλxϵ<λr_{*}^{\epsilon}\geq\sqrt{\log\frac{1}{\sqrt{\pi}\epsilon}},\ \ \ \ -\frac{30n+4}{\sqrt{\log\frac{1}{\sqrt{\pi}\epsilon}}}-\lambda\leq x_{*}^{\epsilon}<-\lambda

for ϵ(0,ϵ¯]\epsilon\in(0,\bar{\epsilon}].

To prove the proposition, we need several information on fϵ(r)f_{\epsilon}(r) under small ϵ\epsilon.

Lemma 3.1.

If λ0\lambda\leq 0, 0<ϵ1π0<\epsilon\leq\frac{1}{\sqrt{\pi}}, then r>log1πϵr_{*}>\sqrt{\log\frac{1}{\sqrt{\pi}\epsilon}}.

Proof.

Since f(0)=0f^{\prime}(0)=0, f′′(r)<0f^{\prime\prime}(r)<0, and limrrf(r)=\lim_{r\rightarrow r_{*}}f^{\prime}(r)=-\infty, there exists r(0,r)r^{\prime}\in(0,r_{*}) so that f(r)=1f^{\prime}(r^{\prime})=-1. Then h(r)h^{\prime}(r^{\prime}) also equals to 1-1. For r(0,r)r\in(0,r^{\prime}), we have

ddr(er2h(r))\displaystyle\dfrac{{\rm d}}{{\rm d}r}(e^{-r^{2}}h^{\prime}(r)) =er2h′′(r)2rer2h(r)\displaystyle=e^{-r^{2}}h^{\prime\prime}(r)-2re^{-r^{2}}h^{\prime}(r)
21+h(r)2er2h′′(r)2rer2h(r)\displaystyle\geq\dfrac{2}{1+h^{\prime}(r)^{2}}e^{-r^{2}}h^{\prime\prime}(r)-2re^{-r^{2}}h^{\prime}(r)
=2er2[(rn1r)h(r)h(r)+λ(11+h(r)2)]2rer2h(r)\displaystyle=2e^{-r^{2}}\left[(r-\dfrac{n-1}{r})h^{\prime}(r)-h(r)+\lambda(1-\sqrt{1+h^{\prime}(r)^{2}})\right]-2re^{-r^{2}}h^{\prime}(r)
2er2h(r),\displaystyle\geq-2e^{-r^{2}}h(r),

where we have used that λ0\lambda\leq 0 in the last inequality. Integrating from 0tor0\ {\rm to}\ r^{\prime},

e(r)220rer2h(r)dr2ϵ0rer2drπϵ.-e^{-(r^{\prime})^{2}}\geq-2\int_{0}^{r^{\prime}}e^{-r^{2}}h(r){\rm d}r\geq-2\epsilon\int_{0}^{r^{\prime}}e^{-r^{2}}{\rm d}r\geq-\sqrt{\pi}\epsilon.

Hence, we obtain r>rlog1πϵr_{*}>r^{\prime}\geq\sqrt{\log\frac{1}{\sqrt{\pi}\epsilon}} for λ0\lambda\leq 0, 0<ϵ1π0<\epsilon\leq\frac{1}{\sqrt{\pi}}. ∎

Lemma 3.2.

For any fixed n2n\geq 2, λ\lambda\in\mathbb{R}, there exists ϵ1>0\epsilon_{1}>0 so that fϵ(r)=λf_{\epsilon}(r)=-\lambda ((i.e., hϵ(r)=0h_{\epsilon}(r)=0)) has a solution in [n,2n][\sqrt{n},\sqrt{2n}] for |ϵ|ϵ1|\epsilon|\leq\epsilon_{1}.

Proof.

We adopt an argument from the appendix B of [10]. In order to understand the behavior of fϵ(r)f_{\epsilon}(r) when ϵ\epsilon is close to 0, we study the linearization of the rotational λ\lambda-hypersurface differential equation (2.8) near the plane f0(r)=λf_{0}(r)=-\lambda. We define ww by

w(r)=ddϵ|ϵ=0fϵ(r).w(r)=\dfrac{{\rm d}}{{\rm d}\epsilon}\bigg{|}_{\epsilon=0}f_{\epsilon}(r).

Since fϵ(r)f_{\epsilon}(r) satisfies equation

fϵ′′1+(fϵ)2=(rn1r)fϵfϵλ1+(fϵ)2,\dfrac{f_{\epsilon}^{\prime\prime}}{1+(f_{\epsilon}^{\prime})^{2}}=(r-\dfrac{n-1}{r})f_{\epsilon}^{\prime}-f_{\epsilon}-\lambda\sqrt{1+(f_{\epsilon}^{\prime})^{2}},

by differentiating the above equation with respect to ϵ\epsilon and letting ϵ=0\epsilon=0, we obtain a differential equation for ww:

(3.1) w′′=(rn1r)ww,w^{\prime\prime}=(r-\dfrac{n-1}{r})w^{\prime}-w,

with w(0)=1w(0)=1 and w(0)=0w^{\prime}(0)=0, where we have used f0(r)=0f_{0}^{\prime}(r)=0. We will show that w(n)>0w(\sqrt{n})>0 and w(2n)<0w(\sqrt{2n})<0 which lead to the lemma.
Defining ξ=r2\xi=r^{2}, the (3.1) becomes into the following differential equation:

(3.2) 4ξd2wdξ2=2(ξn)dwdξw4\xi\dfrac{{\rm d}^{2}w}{{\rm d}\xi^{2}}=2(\xi-n)\dfrac{{\rm d}w}{{\rm d}\xi}-w

with the initial conditions at ξ=0\xi=0:

w(0)=1,dwdξ(0)=12n.w(0)=1,\ \ \ \dfrac{{\rm d}w}{{\rm d}\xi}(0)=-\dfrac{1}{2n}.

The equation (3.2) is one of the classical differential equations, namely a confluent hypergeometric equation. Up to a dilation of the argument ξ\xi, the solutions ww used here are called Kummer functions. Taking derivatives of (3.2), we have the following second order differential equation:

(3.3) 4ξd3wdξ3=2(ξn2)d2wdξ2+dwdξ4\xi\dfrac{{\rm d}^{3}w}{{\rm d}\xi^{3}}=2(\xi-n-2)\dfrac{{\rm d}^{2}w}{{\rm d}\xi^{2}}+\dfrac{{\rm d}w}{{\rm d}\xi}

with the initial conditions at ξ=0\xi=0:

dwdξ(0)=12n,d2wdξ2(0)=14n(n+2).\dfrac{{\rm d}w}{{\rm d}\xi}(0)=-\dfrac{1}{2n},\ \ \ \dfrac{{\rm d}^{2}w}{{\rm d}\xi^{2}}(0)=-\dfrac{1}{4n(n+2)}.

We also note that the differential equation (3.3) for dwdξ\frac{{\rm d}w}{{\rm d}\xi} satisfies a maximum principle, which yields that d2wdξ2<0\frac{{\rm d}^{2}w}{{\rm d}\xi^{2}}<0 and then dwdξdwdξ(0)<0\frac{{\rm d}w}{{\rm d}\xi}\leq\frac{{\rm d}w}{{\rm d}\xi}(0)<0. Hence ww is strictly concave and strictly decreasing on [0,)[0,\infty). Combining this fact with w|ξ=0=1w|_{\xi=0}=1 and dwdξ(0)=12n\frac{{\rm d}w}{{\rm d}\xi}(0)=-\frac{1}{2n} we obtain w|ξ=2n<0w|_{\xi=2n}<0. Let ξ=n\xi=n in (3.2), we get w|ξ=n>0w|_{\xi=n}>0. Now we have proved that w|ξ=n>0w|_{\xi=n}>0 and w|ξ=2n<0w|_{\xi=2n}<0, i.e., w|r=n>0w|_{r=\sqrt{n}}>0 and w|r=2n<0w|_{r=\sqrt{2n}}<0. ∎

Lemma 3.3.

Suppose n2n\geq 2, 25n630nλ<0-\frac{25n-6}{30\sqrt{n}}\leq\lambda<0. If there exists a point r0[n,)r_{0}\in[\sqrt{n},\infty) so that h(r0)=0h(r_{0})=0, we have

h(r)>30nrh(r)h(r)>\dfrac{30n}{r}h^{\prime}(r)

for r[r0,r)r\in[r_{0},r_{*}).

Proof.

Let Φ(r)=130nrh(r)h(r)\Phi(r)=\frac{1}{30n}rh(r)-h^{\prime}(r). We want to show that Φ(r)>0\Phi(r)>0 for r[r0,r)r\in[r_{0},r_{*}). If r=r0r=r_{0}, then Φ(r0)=h(r0)>0\Phi(r_{0})=-h^{\prime}(r_{0})>0. If r>r0r>r_{0}, by a direct computation, we obtain

Φ(r)\displaystyle\Phi(r) =130nrh(r)h(r)\displaystyle=\frac{1}{30n}rh(r)-h^{\prime}(r)
=130nrr0rh(ξ)dξh(r)\displaystyle=\dfrac{1}{30n}r\int_{r_{0}}^{r}h^{\prime}(\xi){\rm d}\xi-h^{\prime}(r)
>130nr(rr0)h(r)h(r)\displaystyle>\dfrac{1}{30n}r(r-r_{0})h^{\prime}(r)-h^{\prime}(r)
=130nh(r)(r2r0r30n)\displaystyle=\dfrac{1}{30n}h^{\prime}(r)(r^{2}-r_{0}r-30n)
130nh(r)(r2nr30n),\displaystyle\geq\dfrac{1}{30n}h^{\prime}(r)(r^{2}-\sqrt{n}r-30n),

that is, Φ(r)>0\Phi(r)>0 for r6nr\leq 6\sqrt{n}.
Suppose Φ(r)=0\Phi(r)=0 for some r[r0,r)r\in[r_{0},r_{*}). Then r>6nr>6\sqrt{n} and there exists a point r¯(6n,r)\bar{r}\in(6\sqrt{n},r_{*}) so that Φ(r¯)=0\Phi(\bar{r})=0 and Φ(r)>0\Phi(r)>0 for r[r0,r¯)r\in[r_{0},\bar{r}). This implies that Φ(r¯)0\Phi^{\prime}(\bar{r})\leq 0 and 130nr¯h(r¯)=h(r¯)\dfrac{1}{30n}\bar{r}h(\bar{r})=h^{\prime}(\bar{r}). On the other hand, since

Φ(r¯)\displaystyle\Phi^{\prime}(\bar{r}) =130nh(r¯)+130nr¯h(r¯)h′′(r¯)\displaystyle=\dfrac{1}{30n}h(\bar{r})+\dfrac{1}{30n}\bar{r}h^{\prime}(\bar{r})-h^{\prime\prime}(\bar{r})
130nh(r¯)+130nr¯h(r¯)h′′(r¯)1+h(r¯)2\displaystyle\geq\dfrac{1}{30n}h(\bar{r})+\dfrac{1}{30n}\bar{r}h^{\prime}(\bar{r})-\dfrac{h^{\prime\prime}(\bar{r})}{1+h^{\prime}(\bar{r})^{2}}
=130nh(r¯)+130nr¯h(r¯)[(r¯n1r¯)h(r¯)f(r¯)λ1+h(r¯)2]\displaystyle=\dfrac{1}{30n}h(\bar{r})+\dfrac{1}{30n}\bar{r}h^{\prime}(\bar{r})-\left[(\bar{r}-\dfrac{n-1}{\bar{r}})h^{\prime}(\bar{r})-f(\bar{r})-\lambda\sqrt{1+h^{\prime}(\bar{r})^{2}}\right]
=130nh(r¯)+130nr¯h(r¯)[(r¯n1r¯)h(r¯)h(r¯)+λ(11+h(r¯)2)]\displaystyle=\dfrac{1}{30n}h(\bar{r})+\dfrac{1}{30n}\bar{r}h^{\prime}(\bar{r})-\left[(\bar{r}-\dfrac{n-1}{\bar{r}})h^{\prime}(\bar{r})-h(\bar{r})+\lambda(1-\sqrt{1+h^{\prime}(\bar{r})^{2}})\right]
130nh(r¯)+130nr¯h(r¯)[(r¯n1r¯)h(r¯)h(r¯)+λh(r¯)]\displaystyle\geq\dfrac{1}{30n}h(\bar{r})+\dfrac{1}{30n}\bar{r}h^{\prime}(\bar{r})-\left[(\bar{r}-\dfrac{n-1}{\bar{r}})h^{\prime}(\bar{r})-h(\bar{r})+\lambda h^{\prime}(\bar{r})\right]
=130nh(r¯)+1900n2r¯2h(r¯)[130nr¯(r¯n1r¯+λ)h(r¯)h(r¯)]\displaystyle=\dfrac{1}{30n}h(\bar{r})+\dfrac{1}{900n^{2}}\bar{r}^{2}h(\bar{r})-\left[\dfrac{1}{30n}\bar{r}(\bar{r}-\dfrac{n-1}{\bar{r}}+\lambda)h(\bar{r})-h(\bar{r})\right]
=1900n2h(r¯)[(130n)r¯230nλr¯+930n2]\displaystyle=\dfrac{1}{900n^{2}}h(\bar{r})\left[(1-30n)\bar{r}^{2}-30n\lambda\bar{r}+930n^{2}\right]
>0,\displaystyle>0,

where we have used that n2n\geq 2, 25n630nλ<0-\frac{25n-6}{30\sqrt{n}}\leq\lambda<0 and r¯>6n\bar{r}>6\sqrt{n} in the last inequality, this contradicts Φ(r¯)0\Phi^{\prime}(\bar{r})\leq 0. Hence, we complete the proof. ∎

Proof of Proposition 3.1. The first part of the proposition 3.1 has been shown in the lemma 3.1. We know that rlog1πϵr^{\prime}\geq\sqrt{\log\frac{1}{\sqrt{\pi}\epsilon}} for 0<ϵ1π0<\epsilon\leq\frac{1}{\sqrt{\pi}} from the lemma 3.1. In particular, rlog1πϵ7nr^{\prime}\geq\sqrt{\log\frac{1}{\sqrt{\pi}\epsilon}}\geq 7\sqrt{n} for 0<ϵ1πe49n0<\epsilon\leq\frac{1}{\sqrt{\pi}e^{49n}}. Choosing ϵ¯=min{ϵ1,1πe49n}\bar{\epsilon}=\min\left\{\epsilon_{1},\frac{1}{\sqrt{\pi}e^{49n}}\right\} and assuming 0<ϵϵ¯0<\epsilon\leq\bar{\epsilon}, by the lemma 3.2 and lemma 3.3, we have h(r)<h(r)<h(r0)=0h(r_{*})<h(r^{\prime})<h(r_{0})=0 (that is, x<λx_{*}<-\lambda) and

h(r)>30nr30nlog1πϵ.h(r^{\prime})>-\dfrac{30n}{r^{\prime}}\geq-\dfrac{30n}{\sqrt{\log\frac{1}{\sqrt{\pi}\epsilon}}}.

We will extend this estimate for h(r)h(r^{\prime}) to an estimate for h(r)=x+λh(r_{*})=x_{*}+\lambda. For rrr\geq r^{\prime}, we have

h′′(r)\displaystyle h^{\prime\prime}(r) <h(r)2h′′(r)1+h(r)2\displaystyle<h^{\prime}(r)^{2}\dfrac{h^{\prime\prime}(r)}{1+h^{\prime}(r)^{2}}
=h(r)2[(rn1r)h(r)h(r)+λ(11+h(r)2)]\displaystyle=h^{\prime}(r)^{2}\left[(r-\dfrac{n-1}{r})h^{\prime}(r)-h(r)+\lambda(1-\sqrt{1+h^{\prime}(r)^{2}})\right]
=h(r)2[(rn1r)h(r)h(r)+λh(r)]\displaystyle=h^{\prime}(r)^{2}\left[(r-\dfrac{n-1}{r})h^{\prime}(r)-h(r)+\lambda h^{\prime}(r)\right]
<h(r)3(r31n1r+λ)\displaystyle<h^{\prime}(r)^{3}(r-\dfrac{31n-1}{r}+\lambda)
14rh(r)3,\displaystyle\leq\dfrac{1}{4}rh^{\prime}(r)^{3},

where we have used that n2,r7nn\geq 2,\ r\geq 7\sqrt{n} and 23n+428nλ<0-\frac{23n+4}{28\sqrt{n}}\leq\lambda<0 in the last inequality.
Integrating the previous inequality from rr to rr_{*}, implies

h(r)24r2r2h^{\prime}(r)^{2}\leq\dfrac{4}{r_{*}^{2}-r^{2}}

for rrr\geq r^{\prime}. Since h(r)<0h^{\prime}(r)<0, we have

(3.4) h(r)2r2r21r+r2rrh^{\prime}(r)\geq-\dfrac{2}{\sqrt{r_{*}^{2}-r^{2}}}\geq-\dfrac{1}{\sqrt{r_{*}+r^{\prime}}}\dfrac{2}{\sqrt{r_{*}-r}}

for r[r,r)r\in[r^{\prime},r_{*}). At rr^{\prime}, this tells us that

rrr+r2r+r.-\dfrac{\sqrt{r_{*}-r^{\prime}}}{\sqrt{r_{*}+r^{\prime}}}\geq-\dfrac{2}{r_{*}+r^{\prime}}.

Finally, integrating (3.4) from rr^{\prime} to rr_{*}, we have

h(r)h(r)4r+rrr,h(r_{*})-h(r^{\prime})\geq-\dfrac{4}{\sqrt{r_{*}+r^{\prime}}}\sqrt{r_{*}-r^{\prime}},

and therefore

h(r)\displaystyle h(r_{*}) h(r)4r+rrr\displaystyle\geq h(r^{\prime})-\dfrac{4}{\sqrt{r_{*}+r^{\prime}}}\sqrt{r_{*}-r^{\prime}}
30nr8r+r\displaystyle\geq-\dfrac{30n}{r^{\prime}}-\dfrac{8}{r_{*}+r^{\prime}}
30n+4r\displaystyle\geq-\dfrac{30n+4}{r^{\prime}}
30n+4log1πϵ.\displaystyle\geq-\frac{30n+4}{\sqrt{\log\frac{1}{\sqrt{\pi}\epsilon}}}.

Hence x30n+4log1πϵλx_{*}\geq-\frac{30n+4}{\sqrt{\log\frac{1}{\sqrt{\pi}\epsilon}}}-\lambda, this competes the proof. ∎

3.2. Behavior near the round sphere

As in the proof of the lemma 3.2, we also make use of the method from the appendix B of [10]. Recall equation (2.9), if we can write the profile curve in the form ρ=ρ(ϕ)\rho=\rho(\phi) in polar coordinates, where ρ=x2+r2\rho=\sqrt{x^{2}\,+\,r^{2}} and ϕ=arctan(r/x)\phi\,=\,\arctan(r/x), then (2.9) tells us that ρ=ρ(ϕ)\rho=\rho(\phi) satisfies

(2.8) ρ′′=1ρ{ρ2+(ρ2+ρ2)[nρ2(n 1)ρρcotϕλρ2+ρ2]}.\rho^{\prime\prime}\,=\,\frac{1}{\rho}\left\{\rho^{\prime 2}\,+\,(\rho^{2}\,+\,\rho^{\prime 2})\left[n\,-\,\rho^{2}\,-\,(n\,-\,1)\frac{\rho^{\prime}}{\rho}\cot\phi\,-\,\lambda\,\sqrt{\rho^{2}\,+\,\rho^{\prime 2}}\right]\right\}.

This ordinary differential equation has a constant solution

ρ=λ+λ2+ 4n2\rho\,=\,\frac{-\lambda\,+\,\sqrt{\lambda^{2}\,+\,4\,n}}{2}

which corresponds to the round sphere. We note that this equation has a singularity when ϕ=0\phi=0 due to the cotϕ\cot\phi term. Let ρ(ϕ,ϵ)\rho(\phi,\epsilon) be the solution to (2.8) with initial conditions ρ(0,ϵ)=λ+λ2+4n2+ϵ\rho(0,\epsilon)=\frac{-\lambda+\sqrt{\lambda^{2}+4\,n}}{2}+\epsilon and dρdϕ(0,ϵ)=0\frac{{\rm d}\rho}{{\rm d}\phi}(0,\epsilon)=0. As is Drugan and Kleene said in [10], for λ=0\lambda=0, the solution ρ(ϕ,ϵ)\rho(\phi,\epsilon) is smooth when ϕ[0,π/2]\phi\in[0,\pi/2] and ϵ\epsilon is close to 0. This is also true for λ\lambda\in\mathbb{R}.
In order to understand the behavior of ρ(ϕ,ϵ)\rho(\phi,\epsilon) when ϵ\epsilon is close to 0, we study the linearization of the rotational self-shrinker differential equation near the round sphere ρ(ϕ,0)=λ+λ2+4n2\rho(\phi,0)=\frac{-\lambda+\sqrt{\lambda^{2}+4\,n}}{2}. We define ww by

w(ϕ)=ddϵ|ϵ=0ρ(ϕ,ϵ).w(\phi)=\frac{\rm d}{{\rm d}\epsilon}\bigg{|}_{\epsilon=0}\rho(\phi,\epsilon).

Then ww satisfies the (singular) linear differential equation:

(3.5) w′′=(n 1)cotϕwλ+λ2+ 4n2λ2+ 4nww^{\prime\prime}=-(n\,-\,1)\cot\phi\,w^{\prime}\,-\,\frac{-\lambda\,+\,\sqrt{\lambda^{2}\,+\,4\,n}}{2}\,\sqrt{\lambda^{2}\,+\,4n}\,w

with w(0)=1w(0)=1 and w(0)=0w^{\prime}(0)=0. Letting A(λ)=λ+λ2+ 4n2λ2+ 4nA(\lambda)=\frac{-\lambda\,+\,\sqrt{\lambda^{2}\,+\,4\,n}}{2}\,\sqrt{\lambda^{2}\,+\,4n}, we may write A(λ)A(\lambda) for AA when there is no ambiguity. Note that the sign of AA is positive for λ\lambda\in\mathbb{R}. In the following lemma, we claim that w(π/2)<0w(\pi/2)<0 and w(π/2)<0w^{\prime}(\pi/2)<0 which yields that, for ϵ<0\epsilon<0 and ϵ\epsilon closed to 0, r>λ+λ2+ 4n2r>\frac{-\lambda\,+\,\sqrt{\lambda^{2}\,+\,4\,n}}{2} and π/2<θ<π\pi/2<\theta<\pi at the first point where the profile curve meets rr-axis.

Lemma 3.4.

Let 2n+2<λ0-\frac{2}{\sqrt{n+2}}<\lambda\leq 0, ww be the solution to (3.5) with w(0)=1w(0)=1 and w(0)=0w^{\prime}(0)=0. Then w(π/2)<0w(\pi/2)<0 and w(π/2)<0w^{\prime}(\pi/2)<0.

Proof.

By making use of the substitution ξ=cosϕ\xi=\cos\phi, (3.5) becomes into the following Legendre type differential equation:

(3.6) (1ξ2)d2wdξ2=nξdwdξAw(1-\xi^{2})\frac{{\rm d}^{2}w}{{\rm d}\xi^{2}}=n\,\xi\,\frac{{\rm d}w}{{\rm d}\xi}-A\,w

with the initial conditions at ξ=1\xi=1:

w(1)=1,dwdξ(1)=An.w(1)=1,\ \ \ \frac{{\rm d}w}{{\rm d}\xi}(1)=\frac{A}{n}.

To prove the lemma, we need to show that w=w(ξ)w=w(\xi) satisfies w(0)<0w(0)<0 and dwdξ(0)>0\frac{{\rm d}w}{{\rm d}\xi}(0)>0.
Taking derivatives of (3.6), we have the following second order differential equations:

(3.7) (1ξ2)d3wdξ3=(n+2)ξd2wdξ2+(nA)dwdξ,\displaystyle(1-\xi^{2})\frac{{\rm d}^{3}w}{{\rm d}\xi^{3}}=(n+2)\,\xi\,\frac{{\rm d}^{2}w}{{\rm d}\xi^{2}}+(n-A)\,\frac{{\rm d}w}{{\rm d}\xi},
(3.8) (1ξ2)d4wdξ4=(n+4)ξd3wdξ3+(2n+2A)d2wdξ2.\displaystyle(1-\xi^{2})\frac{{\rm d}^{4}w}{{\rm d}\xi^{4}}=(n+4)\,\xi\,\frac{{\rm d}^{3}w}{{\rm d}\xi^{3}}+(2n+2-A)\,\frac{{\rm d}^{2}w}{{\rm d}\xi^{2}}.

It follows from (3.7) and (3.8) that

d2wdξ2(1)=(nA)An(n+2),d3wdξ3(1)=(2n+2A)(nA)An(n+2)(n+4).\frac{{\rm d}^{2}w}{{\rm d}\xi^{2}}(1)=-\frac{(n-A)A}{n(n+2)},\ \ \ \frac{{\rm d}^{3}w}{{\rm d}\xi^{3}}(1)=\frac{(2n+2-A)(n-A)A}{n(n+2)(n+4)}.

By a direct calculation, we get that nA=12λ(λ2+4nλ)n<0n-A=\frac{1}{2}\lambda(\sqrt{\lambda^{2}+4n}-\lambda)-n<0 for λ0\lambda\leq 0, and we also get 2n+2A=12(λ24λλ2+4n)2n+2-A=-\frac{1}{2}(\lambda^{2}-4-\lambda\sqrt{\lambda^{2}+4n}). Then one can obtain λ>2n+2\lambda>-\frac{2}{\sqrt{n+2}} by solving inequality 2n+2A>02n+2-A>0. Summing up, we have nA<0n-A<0 and 2n+2A>02n+2-A>0 for 2n+2<λ0-\frac{2}{\sqrt{n+2}}<\lambda\leq 0, which tell us that d2wdξ2(1)>0\frac{{\rm d}^{2}w}{{\rm d}\xi^{2}}(1)>0 and d3wdξ3(1)<0\frac{{\rm d}^{3}w}{{\rm d}\xi^{3}}(1)<0. We also note that the differential equation (3.8) for d2wdξ2\frac{{\rm d}^{2}w}{{\rm d}\xi^{2}} satisfies a maximum principle since 2n+2A>02n+2-A>0.
Using the same method as in [10], we can obtain d2wdξ2(0)>0\frac{{\rm d}^{2}w}{{\rm d}\xi^{2}}(0)>0 and d3wdξ3(0)<0\frac{{\rm d}^{3}w}{{\rm d}\xi^{3}}(0)<0, which follows from (3.6) and (3.7) that w(0)<0w(0)<0 and dwdξ(0)>0\frac{{\rm d}w}{{\rm d}\xi}(0)>0. Regarding w=w(ϕ)w=w(\phi) as a function of ϕ\phi, this says that w(π/2)<0w(\pi/2)<0 and dwdξ(π/2)<0\frac{{\rm d}w}{{\rm d}\xi}(\pi/2)<0, which proves the lemma. ∎

4. Proof of the theorem

Proof of Theorem 1.1. Let S[x]S[x] denote the solution to (2.6) with initial conditions S[x](0)=(x,0,π/2)S[x](0)=(x,0,\pi/2). According to the lemma 2.1 we know that, for x>λx>-\lambda, the first component of P(S[x])P(S[x]) written as a graph over the rr-axis is concave down and decreasing before it blows-up at the point Bx=(fx+λ(rx+λ),rx+λ)=(xx+λ,rx+λ)B_{x}=(f_{x+\lambda}(r_{*}^{x+\lambda}),r_{*}^{x+\lambda})=(x_{*}^{x+\lambda},r_{*}^{x+\lambda}). The proposition 3.1 tells us that when x>λx>-\lambda and xx closed to λ-\lambda, BxB_{x} is in the first quadrant. From the lemma 3.4 we know that when λ<x<λ+λ2+ 4n2-\lambda<x<\frac{-\lambda\,+\,\sqrt{\lambda^{2}\,+\,4\,n}}{2} and xx closed to λ+λ2+ 4n2\frac{-\lambda\,+\,\sqrt{\lambda^{2}\,+\,4\,n}}{2}, BxB_{x} is in the second quadrant. Since the mapping (λ,),xBx(-\lambda,\infty)\rightarrow\mathbb{H},x\mapsto B_{x} is continuous, the previous results imply that there exists x^(λ,λ+λ2+ 4n2)\hat{x}\in(-\lambda,\frac{-\lambda\,+\,\sqrt{\lambda^{2}\,+\,4\,n}}{2}) such that Bx^B_{\hat{x}} lies in rr-axis. Suppose P(S[x^](s^))=Bx^P(S[\hat{x}](\hat{s}))=B_{\hat{x}}, then the curve [0,2s^],sP(S[x^](s))[0,2\hat{s}]\rightarrow\mathbb{H},\,s\mapsto P(S[\hat{x}](s)) will generate an embedded λ\lambda-hypersurface by (2.1), which is diffeomorphic to a sphere 𝕊n\mathbb{S}^{n} and is not isometric to the sphere Sn(λ+λ2+4n2)S^{n}(\frac{-\lambda+\sqrt{\lambda^{2}+4n}}{2}). Moreover, by (2.3) and lemma 2.1, all the principal curvatures of the hypersurface we construct is positive. In other words, the hypersurface we construct is convex. ∎

Remark 4.1.

According to some sketches, we know, for n2n\geq 2 and some λ2n+2\lambda\leq-\frac{2}{\sqrt{n+2}}, there exists an embedded λ\lambda-hypersurface Σnn+1\Sigma^{n}\subset\mathbb{R}^{n+1} which is diffeomorphic to 𝕊n\mathbb{S}^{n} and is not Sn(λ+λ2+4n2)S^{n}(\frac{-\lambda+\sqrt{\lambda^{2}+4n}}{2}) (see Figure 4.2, Figure 4.2, Figure 4.4 and Figure 4.4).

Refer to caption
Figure 4.1. n=2,λ=1,x^1.31n=2,\,\lambda=-1,\,\hat{x}\approx 1.31.
Refer to caption
Figure 4.2. n=3,λ=0.9,x^1.11n=3,\,\lambda=-0.9,\,\hat{x}\approx 1.11.
Refer to caption
Figure 4.3.  The graph of the λ\lambda-hypersurface generated by the profile curve in Figure 4.1
Refer to caption
Figure 4.4.  The graph of half of the λ\lambda-hypersurface generated by the profile curve in Figure 4.1
Remark 4.2.

In the proof of Theorem 1.1, we can not assure that the point x^\hat{x} that we find in (λ,λ+λ2+ 4n2)(-\lambda,\frac{-\lambda\,+\,\sqrt{\lambda^{2}\,+\,4\,n}}{2}) is unique. According to some sketches, for some nn and λ<0\lambda<0 (for example, n=3n=3 and λ=1\lambda=-1 ), there may be more than one points x^1\hat{x}_{1}, x^2,\hat{x}_{2},\cdots in (λ,λ+λ2+ 4n2)(-\lambda,\frac{-\lambda\,+\,\sqrt{\lambda^{2}\,+\,4\,n}}{2}) so that

P(S[x^1](s)),P(S[x^2](s)),P(S[\hat{x}_{1}](s)),P(S[\hat{x}_{2}](s)),\cdots

will generate embedded λ\lambda-hypersurfaces by (2.1) (see Figure 4.6 and Figure 4.6).

Refer to caption
Figure 4.5.  n=3,λ=1,x^1.57n=3,\,\lambda=-1,\,\hat{x}\approx 1.57.
Refer to caption
Figure 4.6.  n=3,λ=1,x^1.69n=3,\,\lambda=-1,\,\hat{x}\approx 1.69.

References

  • [1] U. Abresch and J. Langer, The normalized curve shortening flow and homothetic solutions, J. Differential Geom. 23 (1986), 175šC196.
  • [2] S. B. Angenent, Shrinking doughnuts Birkhaüser, Boston-Basel-Berlin, 7, 21-38, 1992.
  • [3] S. Brendle, Embedded self-similar shrinkers of genus 0, Ann. of Math. 183 (2016), 715-728.
  • [4] J. -E. Chang, 1-dimensional solutions of the λ\lambda-self shrinkers, Geom. Dedicata 189 (2017), 97-112.
  • [5] Q. -M. Cheng and G. Wei, Complete λ\lambda-hypersurfaces of weighted volume-preserving mean curvature flow, Cal. Var. Partial Differential Equations 57 (2018), 32.
  • [6] Q. -M. Cheng and G. Wei, Examples of compact λ\lambda-hypersurfaces in Euclidean spaces, Sci. China Math. 64 (2021), 155-166.
  • [7] T. H. Colding and W. P. Minicozzi II, Generic mean curvature flow I; generic singularities, Ann. of Math. 175 (2012), 755-833.
  • [8] M. do Carmo and M. Dajczer, Hypersurfaces in space of constant curvature, Trans. Amer. Math. Soc. 277 (1983), 685-709.
  • [9] G. Drugan, An immersed S2S^{2} self-shrinker. Trans. Amer. Math. Soc. 367 (2015), 3139-3159.
  • [10] G. Drugan and S. J. Kleene, Immersed self-shrinkers, Trans. Amer. Math. Soc. 369 (2017), 7213-7250.
  • [11] T. -K. Lee, Convexity of λ\lambda-hypersurfaces, Proc. Amer. Math. Soc. 150 (2022), 1735-1744.
  • [12] Q. Guang, A note on mean convex λ\lambda-surfaces in R3R^{3}, Proc. Amer. Math. Soc. 149 (2021), 1259-1266.
  • [13] S. Heilman, Symmetric convex sets with minimal Gaussian surface area, Amer. J. Math. 143 (2021), 53-94.
  • [14] G. Huisken, Asymptotic behavior for singularities of the mean curvature flow, J. Differential Geom. 31 (1990), 285šC299.
  • [15] Z. Li and G. Wei, An immersed SnS^{n} λ\lambda-hypersurface, J. Geom. Anal. 33 (2023), Paper No. 288, 29 pp.
  • [16] P. McGrath, Closed mean curvature self-shrinking surfaces of generalized rotational type, arXiv: 1507.00681. 2015.
  • [17] M. McGonagle and J. Ross, The hyperplane is the only stable, smooth solution to the isoperimetric problem in Gaussian space, Geom. Dedicata 178 (2015), 277-296.
  • [18] J. Ross, On the existence of a closed, embedded, rotational λ\lambda-hypersurface, J. Geom. 110 (2019), 1-12.
  • [19] A. Sun, Compactness and rigidity of λ\lambda-surfaces, Int. Math. Res. Not. IMRN (2021), 11818-11844.