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

11institutetext: Department of Mathematics, Stanford University, Stanford, CA 94305, USA
11email: dhoffman@stanford.edu
22institutetext: Department of Mathematics, E. T. H. Zürich, Rämistrasse 101, 8092 Zürich, Switzerland
22email: tom.ilmanen@math.ethz.ch
33institutetext: Departamento de Geometría y Topología, Universidad de Granada, 18071 Granada, Spain
33email: fmartin@ugr.es
44institutetext: Department of Mathematics, Stanford University, Stanford, CA 94305, USA
44email: bcwhite@stanford.edu

Notes on translating solitons for Mean Curvature Flow

David Hoffman 11    Tom Ilmanen 22    Francisco Martín F. Martín was partially supported by the MINECO/FEDER grant MTM2017-89677-P and by the Leverhulme Trust grant IN-2016-019.33    Brian White B. White was partially supported by grants from the Simons Foundation (#396369) and from the National Science Foundation (DMS 1404282, DMS 1711293).44
Abstract

These notes provide an introduction to translating solitons for the mean curvature flow in 𝐑3\mathbf{R}^{3}. In particular, we describe a full classification of the translators that are complete graphs over domains in 𝐑2\mathbf{R}^{2}.

keywords:
Mean curvature flow, singularities, monotonicity formula, area estimates, comparison principle.

1 Introduction

Mean curvature flow is an exciting area of mathematical research. It is situated at the crossroads of several scientific disciplines: geometric analysis, geometric measure theory, partial differential equations, differential topology, mathematical physics, image processing, computer-aided design, among others. In these notes, we give a brief introduction to mean curvature flow and we describe recent progress on translators, an important class of solutions to mean curvature flow.

In physics, diffusion is a process which equilibrates spatial variations in concentration. If we consider a initial concentration u0u_{0} on a domain Ω𝐑2\Omega\subseteq\mathbf{R}^{2} and seek solutions of the linear heat equation tuΔu=0,\frac{\partial}{\partial t}u-\operatorname{\Delta}u=0, with initial data u0u_{0} and natural boundary conditions on Ω\partial\Omega, we obtain smoothed concentrations {ut}t>0\{u_{t}\}_{t>0}.

Refer to caption
Figure 1: A surface moving by mean curvature.

If we are interested in the smoothing of perturbed surface geometries, it make sense to use analogous strategies. The geometrical counterpart of the Euclidean Laplace operator Δ\operatorname{\Delta} on a smooth surface M2𝐑3M^{2}\subset\mathbf{R}^{3} (or more generally, a hypersurface Mn𝐑n+1M^{n}\subset\mathbf{R}^{n+1}) is the Laplace-Beltrami operator, which we will denote as ΔM\operatorname{\Delta}_{M}. Thus, we obtain the geometric diffusion equation

t𝐱=ΔMt𝐱,\frac{\partial}{\partial t}\;\mathbf{x}=\operatorname{\Delta}_{M_{t}}\mathbf{x}, (1.1)

for the coordinates x of the corresponding family of surfaces {Mt}t[0,T).\{M_{t}\}_{t\in[0,T)}.

A classical formula (see [hildebrandt], for instance) says that, given a hypersurface in Euclidean space, one has:

ΔMt𝐱=H,\operatorname{\Delta}_{M_{t}}\mathbf{x}=\vec{H},

where H\vec{H} represents the mean curvature vector. This means that (1.1) can be written as:

t𝐱(p,t)=H(p,t).\frac{\partial}{\partial t}\;\mathbf{x}(p,t)=\vec{H}(p,t). (1.2)

Using techniques of parabolic PDE’s it is possible to deduce the existence and uniqueness of the mean curvature flow for a small time period in the case of compact manifolds (for details see [regularityTheoryMCF], [LecturesNotesOnMCF], among others).

Theorem 1.1.

Given a compact, immersed hypersurface MM in 𝐑n+1\mathbf{R}^{n+1} then there exists a unique mean curvature flow defined on an interval [0,T)[0,T) with initial surface MM.

The mean curvature is known to be the first variation of the area functional MM𝑑μM\mapsto\int_{M}d\mu (see [colding-minicozzi, meeks-perez].) We will obtain for the Area(Ω(t))\operatorname{Area}(\Omega(t)) of a relatively compact Ω(t)Mt\Omega(t)\subset M_{t} that

ddt(Area(Ω(t))=Ω(t)|H|2dμt.\frac{d}{dt}\,\left(\operatorname{Area}(\Omega(t)\right)=-\int_{\Omega(t)}|\vec{H}|^{2}d\mu_{t}.

In other words, we get that the mean curvature flow is the corresponding gradient flow for the area functional:

Remark 1.2.

The mean curvature flow is the flow of steepest descent of surface area.

Moreover, we also have a nice maximum principle for this particular diffusion equation.

Theorem 1.3 (Maximum/comparison principle).

If two compact immersed hypersurfaces of 𝐑n+1\mathbf{R}^{n+1} are initially disjoint, they remain so. Furthermore, compact embedded hypersurfaces remain embedded.

Similarly, convexity and mean convexity are preserved:

  • If the initial hypersurface MM is convex (i.e., all the geodesic curvatures are positive, or equivalently MM bounds a convex region of 𝐑n+1\mathbf{R}^{n+1}), then MtM_{t} is convex, for any tt.

  • If MM is mean convex (H>0H>0), then MtM_{t} is also mean convex, for any tt.

Moreover, mean curvature flow has a property which is similar to the eventual simplicity for the solutions of the heat equation. This result was proved by Huisken and asserts:

Theorem 1.4 ([huisken90]).

Convex, embedded, compact hypersurfaces converge to points p𝐑n+1.p\in\mathbf{R}^{n+1}. After rescaling to keep the area constant, they converge smoothly to round spheres.

As a consequence of the above theorems we have.

Corollary 1.5 (Existence of singularities in finite time).

Let MM be a compact hypersurface in 𝐑n+1\mathbf{R}^{n+1}. If MtM_{t} represents its evolution by the mean curvature flow, then MtM_{t} must develop singularities in finite time. Moreover, if we denote this maximal time as TmaxT_{\max}, then

2nTmax(diam𝐑n+1(M))2.2\,n\,T_{\max}\leq\left(\mbox{\rm diam}_{\mathbf{R}^{n+1}}(M)\right)^{2}.
Proof 1.6.

Since MM is compact, it is contained an open ball B(p,ρ)B(p,\rho). So, MM must develop a singularity before the flow of 𝐒pn\mathbf{S}^{n}_{p} collapses at the point pp, as otherwise we would contradict the previous theorem. The upper bound of TmaxT_{\max} is just a consequence of the collapsing time of a sphere.

A natural question is: What can we say when MM is not compact? In this case, we can have long-time existence. A trivial example is the case of a complete, properly embedded minimal hypersurface MM in 𝐑n+1\mathbf{R}^{n+1}. Under the mean curvature flow, MM remains stationary, so the flow exists for all time. If we are looking for non-stationary examples, then we can consider the following example:

Example 1.7 (grim reapers).

Consider the euclidean product M=Γ×𝐑n1M=\Gamma\times\mathbf{R}^{n-1}, where Γ\Gamma is the grim reaper in 𝐑2\mathbf{R}^{2} represented by the immersion

f:(π/2,π/2)𝐑2f:(-\pi/2,\pi/2)\to\mathbf{R}^{2} (1.3)
f(x)=(x,logcosx).f(x)=(x,\log\cos x).

If, ignoring parametrization, we let MtM_{t} be the result of flowing MM by mean curvature flow for time tt, then Mt=Mten+1M_{t}=M-t\,e_{n+1}, where again {e1,,en+1}\{\operatorname{e}_{1},\ldots,\operatorname{e}_{n+1}\} represents the canonical basis of 𝐑n+1\mathbf{R}^{n+1}. In other words, MM moves by vertical translations. By definition, we say that MM is a translating soliton in the direction of en+1-\,\mbox{\rm e}_{n+1}. More generally, any translator in the direction of en+1-\,\mbox{e}_{n+1} which is a Riemannian product of a planar curve and an euclidean space 𝐑n1\mathbf{R}^{n-1} can be obtained from this example by a suitable combination of a rotation and a dilation (see [himw] for further details.) We refer to these translating hypersurfaces as nn-dimesional grim reapers, or simply grim reapers if the nn is clear from the context.

Refer to caption
Figure 2: A grim reaper

2 Some Remarks About Singularities

Throughout this section, we consider a fixed compact initial hypersurface MM. Consider the maximal time T=TMT=T_{M} such that a smooth solution of the MCF F:M×[0,T)𝐑n+1F:M\times[0,T)\to\mathbf{R}^{n+1} as in Theorem 1.1 exists. Then the embedding vector FF is uniformly bounded according to Theorem 1.3. It follows that some spatial derivatives of the embedding FtF_{t} have to become unbounded as tTt\nearrow T. Otherwise, we could apply Arzelà-Ascoli Theorem and obtain a smooth limit hypersurface, MTM_{T}, such that MtM_{t} converges smoothly to MTM_{T} as tTt\nearrow T. This is impossible because, in such a case, we could apply Theorem 1.1 to restart the flow. In this way, we could extend the flow smoothly all the way up to T+εT+\varepsilon, for some ε>0\varepsilon>0 small enough, contradicting the maximality of TT. In fact, Huisken [huisken84, Theorem 8.1] showed that the second spatial derivative (i.e, the norm of the second fundamental form) blows up as tT.t\to T.

We would like to say more about the “blowing-up” of the norm of AA, as tT.t\nearrow T. The evolution equation for |A|2|A|^{2} is

t|A|2=Δ|A|22|A|2+2|A|4.\frac{\partial}{\partial t}|A|^{2}=\operatorname{\Delta}|A|^{2}-2\big{|}\nabla A\big{|}^{2}+2|A|^{4}.

Define

|A|max2(t)maxMt|A|2(,t).|A|^{2}_{\text{max}}(t)\coloneqq\max_{M_{t}}|A|^{2}(\cdot,t).

Using Hamilton’s trick (see [LecturesNotesOnMCF]) we deduce that |A|max2|A|^{2}_{\max} is locally Lipschitz and that

ddt|A|max2(t0)=t|A|2(p0,t0),\frac{d}{dt}\,|A|^{2}_{\max}(t_{0})=\frac{\partial}{\partial t}\,|A|^{2}(p_{0},t_{0}),

where p0p_{0} is any point where |A|2(,t0)|A|^{2}(\cdot,t_{0}) reaches its maximum. Thus, using the above expression, we have

ddt|A|max2(t0)\displaystyle\frac{d}{dt}\,|A|^{2}_{\max}(t_{0}) =t|A|2(p0,t0)\displaystyle=\frac{\partial}{\partial t}\,|A|^{2}(p_{0},t_{0})
=Δ|A|2(p0,t0)2|A(p0,t0)|2+2|A|4(p0,t0).\displaystyle=\operatorname{\Delta}|A|^{2}(p_{0},t_{0})-2\big{|}\nabla A(p_{0},t_{0})\big{|}^{2}+2|A|^{4}(p_{0},t_{0}).

It is well known that the Hessian of |A||A| is negative semi-definite at any maximum. In particular the Laplacian of |A||A| at these points is non-positive. Hence,

ddt|A|max2(t0)2|A|4(p0,t0)2|A|max4(t0).\frac{d}{dt}\,|A|^{2}_{\max}(t_{0})\leq 2|A|^{4}(p_{0},t_{0})\leq 2|A|_{\text{max}}^{4}(t_{0}).

Notice that |A|max2|A|_{\text{max}}^{2} is always positive, since otherwise at some instant tt we would have |A(,t)|0|A(\cdot,t)|\equiv 0 along MtM_{t}, which would imply that MtM_{t} is totally geodesic and therefore a hyperplane in 𝐑n+1\mathbf{R}^{n+1}, contrary to the fact that the initial surface was compact.

So, one can prove that 1/|A|max21/|A|_{\text{max}}^{2} is locally Lipschitz. Then the previous inequality allows us to deduce that:

ddt(1|A|max2)2a.e. in t[0,T).-\frac{d}{dt}\left(\frac{1}{|A|_{\text{max}}^{2}}\right)\leq 2\quad\mbox{\rm a.e. in $t\in[0,T)$.}

Integrating (respect to time) in any sub-interval [t,s][0,T)[t,s]\subset[0,T) we get

1|A|max2(t)1|A|max2(s)2(st).\frac{1}{|A|_{\text{max}}^{2}(t)}-\frac{1}{|A|_{\text{max}}^{2}(s)}\leq 2(s-t).

As |A(,t)||A(\cdot,t)| is not bounded as to tends to TT, there exists a time sequence siTs_{i}\nearrow T such that

|A|max2(si)+.|A|_{\text{max}}^{2}(s_{i})\to+\infty.

Substituting s=sis=s_{i} in the above inequality and taking the limit as ii\to\infty, we get

1|A|max2(t)2(Tt).\frac{1}{|A|_{\text{max}}^{2}(t)}\leq 2(T-t).

We collect all this information in the next proposition.

Proposition 2.1.

Consider the mean curvature flow for a compact initial hypersurface MM. If TT is the maximal time of existence, then the following lower bound holds

|A|max(t)12(Tt)|A|_{\text{max}}(t)\geq\frac{1}{\sqrt{2(T-t)}}

for all t[0,T)t\in[0,T).

In particular,

limtT|A|max(t)=+.\lim_{t\to T}|A|_{\text{max}}(t)=+\infty.
Definition 2.2.

When this happens we say that TT is singular time for the mean curvature flow.

So we have the following improved version of Theorem 1.1:

Theorem 2.3.

Given a compact, immersed hypersurface MM in 𝐑n+1\mathbf{R}^{n+1} then there exists a unique mean curvature flow defined on a maximal interval [0,Tmax)[0,T_{\text{max}}).
Moreover, TmaxT_{\text{max}} is finite and

|A|max(t)12(Tmaxt)|A|_{\text{max}}(t)\geq\frac{1}{\sqrt{2(T_{\text{max}}-t)}}

for each t[0,Tmax)t\in[0,T_{\text{max}}).

Remark 2.4.

From the above proposition, we deduce the following estimate for the maximal time of existence of flow:

Tmax12|A|max2(0).T_{\text{max}}\geq\frac{1}{2|A|_{\text{max}}^{2}(0)}.
Definition 2.5.

Let TT be the maximal time of existence of the mean curvature flow. If there is a constant C>1C>1 such that

|A|max(t)C2(Tt),|A|_{\text{max}}(t)\leq\frac{C}{\sqrt{2(T-t)}},

then we say that the flow develops a Type I singularity at instant TT. Otherwise, that is, if

lim suptT|A|max(t)(Tt)=+,\limsup_{t\to T}|A|_{\text{max}}(t)\,\sqrt{(T-t)}=+\infty,

we say that is a Type II singularity.

We conclude this brief section by pointing out that there have been substantial breakthroughs in the study and understanding of the singularities of Type I, whereas Type II singularities have been much more difficult to study. This seems reasonable since, according to the above definition and the results we have seen, the singularities of Type I are those for which one has the best possible control of blow-up of the second fundamental form.

3 Translators

A standard example of Type II singularity is given by a loop pinching off to a cusp (see Figure 3). S. Angenent [An1] proved, in the case of convex planar curves, that singularities of this kind are asymptotic (after rescaling) to the above mentioned grim reaper curve (1.3), which moves set-wise by translation. In this case, up to inner diffeomorphisms of the curve, it can be seen as a solution of the curve shortening flow which evolves by translations and is defined for all time. In this paper we are interested in this type of solitons, which we will call translating solitons (or translators) from now on. Summarizing this information, we make the following definition:

Definition 3.1 (Translator).

A translator is a hypersurface MM in 𝐑n+1\mathbf{R}^{n+1} such that

tMt𝐞n+1t\mapsto M-t\,\mathbf{e}_{n+1}

is a mean curvature flow, i.e., such that normal component of the velocity at each point is equal to the mean curvature at that point:

H=𝐞n+1.\vec{H}=-\mathbf{e}_{n+1}^{\perp}. (3.1)
Refer to caption
Figure 3:

The cylinder over a grim-reaper curve, i.e. the hypersurface in 𝐑n+1\mathbf{R}^{n+1} parametrized by 𝒢:(π2,π2)×𝐑n1𝐑n+1\mathscr{G}:\left(-\tfrac{\pi}{2},\tfrac{\pi}{2}\right)\times\mathbf{R}^{n-1}\longrightarrow\mathbf{R}^{n+1} given by

𝒢(x1,,xn)=(x1,,xn,logcosx1),\mathscr{G}(x_{1},\ldots,x_{n})=(x_{1},\ldots,x_{n},-\log\cos x_{1}),

is a translating soliton. It appears as limit of sequences of parabolic rescaled solutions of mean curvature flows of immersed mean convex hypersurfaces. For example, we can take product of the loop pinching off to a cusp times 𝐑n1\mathbf{R}^{n-1}. We can produce others examples of solitons just by scaling and rotating the grim reaper. In this way, we obtain a 11-parameter family of translating solitons parametrized by 𝒢θ:(π2cos(θ),π2cos(θ))×𝐑n1𝐑n+1\mathscr{G}_{\theta}:\left(-\tfrac{\pi}{2\cos(\theta)},\tfrac{\pi}{2\cos(\theta)}\right)\times\mathbf{R}^{n-1}\longrightarrow\mathbf{R}^{n+1}

𝒢θ(x1,,xn)=(x1,,xn,sec2(θ)logcos(x1cos(θ))tan(θ)xn),\mathscr{G}_{\theta}(x_{1},\ldots,x_{n})=(x_{1},\ldots,x_{n},\sec^{2}(\theta)\log\cos(x_{1}\cos(\theta))-\tan(\theta)x_{n}), (3.2)

where θ[0,π/2).\theta\in[0,\pi/2). Notice that the limit of the family FθF_{\theta}, as θ\theta tends to π/2\pi/2, is a hyperplane parallel to 𝐞n+1\mathbf{e}_{n+1}.

Refer to caption
Figure 4: The regular grim reaper in 𝐑3\mathbf{R}^{3} and the tilted grim reaper for θ=π/6\theta=\pi/6.

3.1 Variational approach

Ilmanen [ilmanen] noticed that a translating soliton MM in 𝐑n+1\mathbf{R}^{n+1} can be seen as a minimal surface for the weighted volume functional

𝒜f[M]=Mef𝑑μ\mathscr{A}_{f}[M]=\int_{M}e^{-f}\,d\mu

where ff represents the Euclidean height function, that is, the restriction of the last coordinate xn+1x_{n+1} to MM. We have the following

Proposition 3.2 (Ilmanen).

Let MnM^{n} be a translating soliton in 𝐑n+1\mathbf{R}^{n+1} and let NN be its unit normal. Then the translator equation

H=𝐞n+1,NH=\langle\mathbf{e}_{n+1},N\rangle (3.3)

on the relatively compact domain ΩM\Omega\subset M is the Euler-Lagrange equation of the functional

𝒜f[Ω]=volf(Ω)=Ωef𝑑μ.\mathscr{A}_{f}[\Omega]={\rm vol}_{f}(\Omega)=\int_{\Omega}e^{-f}\,d\mu. (3.4)

Moreover, the second variation formula for normal variations is given by

δ2𝒜f[Ω](u,u)=MefuLfu𝑑μ,uC0(Ω),\delta^{2}\mathscr{A}_{f}[\Omega]\cdot(u,u)=-\int_{M}e^{-f}\,u\,L_{f}u\,d\mu,\quad u\in C^{\infty}_{0}(\Omega), (3.5)

where the stability operator LfL_{f} is defined by

Lfu=Δfu+|A|2uL_{f}u=\operatorname{\Delta}^{f}u+|A|^{2}\,u (3.6)

where Δf\operatorname{\Delta}^{f} is the drift Laplacian given by

Δf=Δf,=Δ𝐞n+1,.\operatorname{\Delta}^{f}=\operatorname{\Delta}-\langle\nabla f,\nabla\,\cdot\,\rangle=\operatorname{\Delta}-\langle\mathbf{e}_{n+1},\nabla\,\cdot\,\rangle. (3.7)

Proof. Given ε>0\varepsilon>0, let Ψ:(ε,ε)×M𝐑n+1\Psi:(-\varepsilon,\varepsilon)\times M\to\mathbf{R}^{n+1} be a variation of MM compactly supported in ΩM\Omega\subset M with Ψ(0,)=Id\Psi(0,\,\cdot\,)=\mbox{Id} and normal variational vector field

Ψs|s=0=uN+T\frac{\partial\Psi}{\partial s}\Big{|}_{s=0}=uN+T

for some function uC0(Ω)u\in C^{\infty}_{0}(\Omega) and a tangent vector field TΓ(TM)T\in\Gamma(TM). Here, NN denotes a local unit normal vector field along MM. Then

dds|s=0volf[Ψs(Ω)]=Ωef(𝐞n+1,NH)u𝑑μ+Ωdiv(efT)𝑑μ.\displaystyle\frac{{\rm d}}{{\rm d}s}\Big{|}_{s=0}{\rm vol}_{f}[\Psi_{s}(\Omega)]=\int_{\Omega}e^{-f}\,(\langle\mathbf{e}_{n+1},N\rangle-H)u\,d\mu+\int_{\Omega}{\rm div}(e^{-f}T)\,d\mu.

Hence, stationary immersions for variations fixing the boundary of Ω\Omega are characterized by the scalar soliton equation

H𝐞n+1,N=0 on ΩM,H-\,\langle\mathbf{e}_{n+1},N\rangle=0\,\,\,\mbox{ on }\,\,\,\Omega\subset\subset M,

which yields (3.3). Now we compute the second variation formula. At a stationary immersion we have

d2ds2|s=0volf[Ψs(Ω)]=Mefdds|s=0(𝐞n+1,NsHs)udμ.\displaystyle\frac{{\rm d}^{2}}{{\rm d}s^{2}}\Big{|}_{s=0}{\rm vol}_{f}[\Psi_{s}(\Omega)]=\int_{M}e^{-f}\,\frac{d}{ds}\Big{|}_{s=0}(\langle\mathbf{e}_{n+1},N_{s}\rangle-H_{s})u\,d\mu.

Using the fact that

¯sN=u𝒲T,\bar{\nabla}_{\partial_{s}}N=-\nabla u-\mathscr{W}T, (3.8)

(where 𝒲\mathscr{W} means the Weingarten map) then we compute

dds|s=0𝐞n+1,N=¯s𝐞n+1,N+𝐞n+1,¯sN=𝐞n+1,u𝒲T.\frac{{\rm d}}{{\rm d}s}\Big{|}_{s=0}\langle\mathbf{e}_{n+1},N\rangle=\langle\bar{\nabla}_{\partial_{s}}\mathbf{e}_{n+1},N\rangle+\langle\mathbf{e}_{n+1},\bar{\nabla}_{\partial_{s}}N\rangle=\langle\mathbf{e}_{n+1},-\nabla u-\mathscr{W}T\rangle. (3.9)

Since

dds|s=0H=Δu+|A|2u+TH\frac{{\rm d}}{{\rm d}s}\Big{|}_{s=0}H=\operatorname{\Delta}u+|A|^{2}\,u+\mathscr{L}_{T}H (3.10)

and η=𝐞n+1\nabla\eta=\mathbf{e}_{n+1}^{\top}, we obtain for normal variations (when T=0T=0)

dds(H𝐞n+1,N)\displaystyle\frac{{\rm d}}{{\rm d}s}\big{(}H-\langle\mathbf{e}_{n+1},N\rangle\big{)} =Δu𝐞n+1,u+|A|2u\displaystyle=\operatorname{\Delta}u-\langle\mathbf{e}_{n+1},\nabla u\rangle+|A|^{2}\,u
=Δfu+|A|2u.\displaystyle=\operatorname{\Delta}^{f}u+|A|^{2}u.

This finishes the proof of the proposition. \square

The previous result has important consequences. It means that a hypersurface M𝐑n+1M\subset\mathbf{R}^{n+1} is a translator if and only if it is minimal with respect to the Riemannian metric

gij(x1,,xn+1)=exp(2nxn+1)δij.g_{ij}(x_{1},\ldots,x_{n+1})=\exp\left(-\frac{2}{n}\,x_{n+1}\right)\delta_{ij}.

Although the metric gijg_{ij} is not complete (notice that the length of vertical half-lines in the direction of 𝐞n+1\mathbf{e}_{n+1} is finite), we can apply all the local results of the theory of minimal hypersurfaces in Riemannian manifolds. Thus we can freely use curvature estimates and compactness theorems from minimal surface theory; cf. [white-intro, Chapter 3]. In particular, if MM is a graphical translator, then (since vertical translates of it are also gg-minimal) 𝐞3,ν\left<\mathbf{e}_{3},\nu\right> is a nowhere vanishing Jacobi field, so MM is a stable gg-minimal surface. It follows that any sequence MiM_{i} of complete translating graphs in 𝐑3\mathbf{R}^{3} has a subsequence that converges smoothly to a translator MM. Also, if a translator MM is the graph of a function u:Ω𝐑u:\Omega\to\mathbf{R}, then MM and its vertical translates from a gg-minimal foliation of Ω×𝐑\Omega\times\mathbf{R}, from which it follows that MM is gg-area minimizing in Ω×𝐑\Omega\times\mathbf{R}, and thus that if KΩ×𝐑K\subset\Omega\times\mathbf{R} is compact, then the gg-area of MKM\cap K is at most 1/21/2 of the gg-area of K\partial K. Hence, if we consider sequences of translators that are manifolds-with-boundary, then the area bounds described above together with standard compactness theorems for minimal surfaces (such as those in [white-curvature-estimates, white-controlling]) give smooth, subsequential convergence, including at the boundary. This has been a crucial tool in [mpss, himw, families-1, families-2] (The local area bounds and bounded topology mean that the only boundary singularities that could arise would be boundary branch points. In the situations that occur in these papers, obvious barriers preclude boundary branch points.)

The situation for higher dimensional translating graphs is more subtle; (see [himw, Appendix A] and [nino]).

4 Examples of translators

Besides the grim reapers that we have already described, the last decades have witnessed the appearance of numerous examples of translators. Clutterbuck, Schnürer and Schulze [CSS] (see also [Altschuler-Wu]) proved that there exists an entire graphical translator in 𝐑n+1\mathbf{R}^{n+1} which is rotationally symmetric, strictly convex with translating velocity 𝐞n+1-\mathbf{e}_{n+1}. This example is known as the translating paraboloid or bowl soliton. Moreover, they classified all the translating solitons of revolution, giving a one-parameter family {Wλn}λ>0\{W^{n}_{\lambda}\}_{\lambda>0} of rotationally invariant cylinders called translating catenoids. The parameter λ\lambda controls the size of the neck of each translating soliton. The limit, as λ0\lambda\to 0, of WλnW^{n}_{\lambda} consists of two superimposed copies of the bowl soliton with a singular point at the axis of symmetry. Furthermore, all these hypersurfaces have the following asymptotic expansion as r approaches infinity:

r22(n1)logr+O(r1),\frac{r^{2}}{2(n-1)}-\log r+O(r^{-1}),

where rr is the distance function in 𝐑n\mathbf{R}^{n}.

Refer to caption
Refer to caption
Figure 5: The bowl soliton in 𝐑3\mathbf{R}^{3} and the translating catenoid for λ=2.\lambda=2.

These rotationally symmetric translating catenoids can be seen as the desingularization of two paraboloids connected by a small neck of some radius.

Recall that the Costa-Hoffman-Meeks surfaces can be regarded as desingularizations of a plane and catenoid: a sequence of Costa-Hoffman-Meeks surfaces with genus tending to infinity converges (if suitably scaled) to the union of a catenoid and the plane through its waist. This suggests that one try to construct translators by desingularizing the union of a translating catenoid and a bowl solition. Dávila, del Pino, and Nguyen  [DDPN] were able to do that (for large genus) by glueing methods, replacing the circle of intersection by a surface similar to the singly periodic Scherk minimal surface (see also [smith].) Previously, Nguyen in [Nguyen09],[Nguyen13] and [Nguyen15] had used similar techniques to desingularize the intersection of a grim reaper and a plane. In this way she obtained a complete periodic embedded translator of infinite genus, that she called a translating trident.

Once this abundance of translating solitons is guaranteed, there arises the need to classify them. One of the first classification results was given by X.-J. Wang in [wang]. He characterized the bowl soliton as the only convex translating soliton which is an entire graph.

Very recently, J. Spruck and L. Xiao [spruck-xiao] have proved that a complete translating soliton which is graph over a domain in 𝐑2\mathbf{R}^{2} must be convex (see Section 6 below.) So, combining both results we have:

Theorem 4.1.

The bowl soliton is the only translator that is an entire graph over 𝐑2\mathbf{R}^{2}.

Using the Alexandrov method of moving hyperplanes, Martín, Savas-Halilaj, and Smoczyk [mss] showed that the bowl soliton is the only translator (not assumed to be graphical) that has one end and is CC^{\infty}-asymptotic to a bowl soliton. Hershkovits [Her18] improved this by showing uniqueness of the bowl soliton among (not necessarily graphical) translators that have one cylindrical end (and no other ends). Haslhofer [Has15] proved a related result in higher dimensions: he showed that any translator in 𝐑n+1\mathbf{R}^{n+1} that is noncollapsed and uniformly 22-convex must be the nn-dimensional bowl soliton. At this point, we would like to mention the recent classification result of Brendle and Choi [brendle]. They prove that the rotationally symmetric bowl soliton is the only noncompact ancient solution of mean curvature flow in 𝐑3\mathbf{R}^{3} which is strictly convex and noncollapsed.

Martín, Savas-Halilaj and Smoczyk also obtained one of the first characterizations of the family of tilted grim reapers:

Theorem 4.2.

[mss] Let MM be a connected translating soliton in 𝐑n+1\mathbf{R}^{n+1}, n2n\geq 2, such that the function |A|2H2|A|^{2}H^{-2} has a local maximum in {xM:H(x)0}\{x\in M:H(x)\neq 0\}. Then MM is a tilted grim reaper.

5 Graphical translators

If a translator MM is the graph of function u:Ω𝐑n𝐑u:\Omega\subset\mathbf{R}^{n}\to\mathbf{R}, we will say that MM is a translating graph; in that case, we also refer to the function uu as a translator, and we say that uu is complete if its graph is a complete submanifold of 𝐑n+1\mathbf{R}^{n+1}. Thus u:Ω𝐑n𝐑u:\Omega\subset\mathbf{R}^{n}\to\mathbf{R} is a translator if and only if it solves the translator equation (the nonparametric form of (3.1)):

Di(Diu1+|Du|2)=11+|Du|2.D_{i}\left(\frac{D_{i}u}{\sqrt{1+|Du|^{2}}}\right)=-\frac{1}{\sqrt{1+|Du|^{2}}}. (5.1)

The equation can also be written as

(1+|Du|2)ΔuDiuDjuDiju+|Du|2+1=0.(1+|Du|^{2})\operatorname{\Delta}u-D_{i}u\,D_{j}u\,D_{ij}u+|Du|^{2}+1=0. (5.2)

In a recent preprint, we classify all complete translating graphs in 𝐑3\mathbf{R}^{3}. In two other papers [families-1, families-2], we construct new families of complete, properly embedded (non-graphical) translators: a two-parameter family of translating annuli, examples that resemble Scherk’s minimal surfaces, and examples that resemble helicoids. In [families-2], we also construct several new families of complete translators that are obtained as limits of the Scherk-type translators mentioned above. They include a 11-parameter family of single periodic surfaces called Scherkenoids (see Fig. 6) and a simply-connected translator called the pitchfork translator (see Fig. 7). The pitchfork translator resembles Nguyen’s translating tridents [Nguyen09] (see also [families-1.5]): like the tridents, it is asymptotic to a plane as zz\to\infty and to three parallel planes as zz\to-\infty. However, the pitchfork has genus 0, whereas the tridents have infinite genus.

Refer to caption
Figure 6: A Scherkenoid is a singly periodic translator. As zz\to\infty, it is asymptotic to a plane, and as zz\to-\infty, it is asymptotic to an infinite family of parallel planes. There is a one-parameter family of such Scherkenoids, the parameter being the angle between the upper plane and the lower planes.
Refer to caption
Figure 7: The pitchfork translator is a simply connected translator that is asympotic to a plane as zz\to\infty and to three parallel planes as zz\to-\infty.

As a consequence of Theorem 4.2, we have that every translator 𝐑3\mathbf{R}^{3} with zero Gauss curvature is a grim reaper surface, a tilted grim reaper surface, or a vertical plane.

In addition to the examples described in the previous section, Ilmanen (in unpublished work) proved that for each 0<k<1/20<k<1/2, there is a translator u:Ω𝐑u:\Omega\to\mathbf{R} with the following properties: u(x,y)u(x,y)u(x,y)u(x,y)\equiv u(-x,y)\equiv u(x,-y), uu attains its maximum at (0,0)Ω(0,0)\in\Omega, and

D2u(0,0)=[k00(1k)].D^{2}u(0,0)=\begin{bmatrix}-k&0\\ 0&-(1-k)\end{bmatrix}.

The domain Ω\Omega is either a strip 𝐑×(b,b)\mathbf{R}\times(-b,b) or 𝐑2\mathbf{R}^{2}. He referred to these examples as Δ\operatorname{\Delta}-wings. As k0k\to 0, he showed that the examples converge to the grim reaper surface. Uniqueness (for a given kk) was not known. It was also not known which strips 𝐑×(b,b)\mathbf{R}\times(-b,b) occur as domains of such examples. The main result in [himw] is the following:

Theorem 5.1.

For every b>π/2b>\pi/2, there is (up to translation) a unique complete, strictly convex translator ub:𝐑×(b,b)𝐑.u^{b}:\mathbf{R}\times(-b,b)\to\mathbf{R}. Up to isometries of 𝐑2\mathbf{R}^{2}, the only other complete translating graphs in 𝐑3\mathbf{R}^{3} are the grim reaper surface, the tilted grim reaper surfaces, and the bowl soliton.

Although the paper [himw] is primarily about translators in 𝐑3\mathbf{R}^{3}, the last sections extend Ilmanen’s original proof to get Δ\operatorname{\Delta}-wings in 𝐑n+1\mathbf{R}^{n+1} that have prescribed principal curvatures at the origin. For n3n\geq 3, the examples include entire graphs that are not rotationally invariant. At the end of the paper, we modify the construction to produce a family of Δ\operatorname{\Delta}-wings in 𝐑n+2\mathbf{R}^{n+2} over any given slab of width >π>\pi. See [wang] for a different construction of some higher dimensional graphical translators.

Refer to caption
Figure 8: The Δ\operatorname{\Delta}-wing of width 2π.\sqrt{2}\,\pi. As y±,y\to\pm\infty, this Δ\operatorname{\Delta}-wing is asymptotic to the tilted grim reapers 𝒢π4\mathscr{G}_{-\frac{\pi}{4}} and 𝒢π4\mathscr{G}_{\frac{\pi}{4}}, respectively.

6 The Spruck-Xiao Convexity Theorem

One of the fundamental results in the recent development of soliton theory has been the paper by Spruck and Xiao [spruck-xiao], where they proved that complete graphical translators (or, more generally, complete translators of positive mean curvature) are convex. The ideas contained in this paper are really inspiring and we would like to provide a slightly simplified exposition of their proof.

At any non-umbilic point, we let κ1>κ2\kappa_{1}>\kappa_{2} be the principal curvatures and H=κ1+κ2>0H=\kappa_{1}+\kappa_{2}>0 be the mean curvature. We let v1v_{1} and v2v_{2} be the principal direction unit vector fields, so

κiA(vi,vi) and A(v1,v2)0.\text{$\kappa_{i}\equiv A(v_{i},v_{i})$ and $A(v_{1},v_{2})\equiv 0$}.

Note uv1\nabla_{u}v_{1} is perpendicular to v1v_{1}. Thus

v1v1=α1v2,v2v1=α2v2\nabla_{v_{1}}v_{1}=\alpha_{1}v_{2},\qquad\nabla_{v_{2}}v_{1}=\alpha_{2}v_{2} (6.1)

for some functions α1\alpha_{1} and α2\alpha_{2}. Since 0=vi(v1v2)=(viv1)v2+v1(viv2)0=\nabla_{v_{i}}(v_{1}\cdot v_{2})=(\nabla_{v_{i}}v_{1})\cdot v_{2}+v_{1}\cdot(\nabla_{v_{i}}v_{2}), we see that

v1v2=α1v1,v2v2=α2v1.\nabla_{v_{1}}v_{2}=-\alpha_{1}v_{1},\qquad\nabla_{v_{2}}v_{2}=-\alpha_{2}v_{1}. (6.2)

Thus

uκ1\displaystyle\nabla_{u}\kappa_{1} =uA(v1,v1)\displaystyle=\nabla_{u}A(v_{1},v_{1})
=(uA)(v1,v1)+2A(uv1,v1).\displaystyle=(\nabla_{u}A)(v_{1},v_{1})+2A(\nabla_{u}v_{1},v_{1}).

But uv1\nabla_{u}v_{1} is perpendicular to v1v_{1}, so A(uv1,v1)0A(\nabla_{u}v_{1},v_{1})\equiv 0. Thus

uκ1=(uA)(v1,v1).\nabla_{u}\kappa_{1}=(\nabla_{u}A)(v_{1},v_{1}). (6.3)

In particular,

iκj=hjj,i,\nabla_{i}\kappa_{j}=h_{jj,i}, (6.4)

where hij=A(vi,vj)h_{ij}=A(v_{i},v_{j}) and hij,k=(vkA)(vi,vj)h_{ij,k}=(\nabla_{v_{k}}A)(v_{i},v_{j}). From A(v1,v2)0A(v_{1},v_{2})\equiv 0, we see that

h12,i\displaystyle h_{12,i} =(i)A(v1,v2)\displaystyle=(\nabla_{i})A(v_{1},v_{2})
=i(A(v1,v2))A(iv1,v2)A(v1,iv2)\displaystyle=\nabla_{i}(A(v_{1},v_{2}))-A(\nabla_{i}v_{1},v_{2})-A(v_{1},\nabla_{i}v_{2})
=0αih22+αih11\displaystyle=0-\alpha_{i}h_{22}+\alpha_{i}h_{11}
=(κ1κ2)αi\displaystyle=(\kappa_{1}-\kappa_{2})\alpha_{i}

by (6.1) and (6.2). Thus

αi=h12,iκ1κ2.\alpha_{i}=\frac{h_{12,i}}{\kappa_{1}-\kappa_{2}}. (6.5)

Also, if we let ui=viu_{i}=v_{i} at a particular point and extend by parallel transport on radial geodesics, we have

Δκ1\displaystyle\operatorname{\Delta}\kappa_{1} =uiui(A(v1,v1))\displaystyle=\nabla_{u_{i}}\nabla_{u_{i}}(A(v_{1},v_{1}))
=ui((uiA)(v1,v1))\displaystyle=\nabla_{u_{i}}((\nabla_{u_{i}}A)(v_{1},v_{1}))
=(ΔA)(v1,v1)+2(uiA)(uiv1,v1).\displaystyle=(\operatorname{\Delta}A)(v_{1},v_{1})+2(\nabla_{u_{i}}A)(\nabla_{u_{i}}v_{1},v_{1}).

Thus

Δκ1\displaystyle\operatorname{\Delta}\kappa_{1} =(ΔA)(v1,v1)+2(iA)(iv1,v1)\displaystyle=(\operatorname{\Delta}A)(v_{1},v_{1})+2(\nabla_{i}A)(\nabla_{i}v_{1},v_{1})
=(ΔA)(v1,v1)+2(iA)(αiv2,v1)\displaystyle=(\operatorname{\Delta}A)(v_{1},v_{1})+2(\nabla_{i}A)(\alpha_{i}v_{2},v_{1})
=(ΔA)(v1,v1)+2αih12,i\displaystyle=(\operatorname{\Delta}A)(v_{1},v_{1})+2\alpha_{i}h_{12,i}
=(ΔA)(v1,v1)+2(h12,1)2+(h12,2)2κ1κ2\displaystyle=(\operatorname{\Delta}A)(v_{1},v_{1})+2\frac{(h_{12,1})^{2}+(h_{12,2})^{2}}{\kappa_{1}-\kappa_{2}}
=(ΔA)(v1,v1)+2Q2κ1κ2,\displaystyle=(\operatorname{\Delta}A)(v_{1},v_{1})+\frac{2Q^{2}}{\kappa_{1}-\kappa_{2}},

where

Q2:=(h12,1)2+(h12,2)2=(h11,2)2+(h22,1)2.Q^{2}:=(h_{12,1})^{2}+(h_{12,2})^{2}=(h_{11,2})^{2}+(h_{22,1})^{2}. (6.6)

(The second equality follows from the Codazzi equations.)

Now suppose the surface is a translator. Then, we have that (see [spruck-xiao, Lemma 2.1] or [mss, Lemma 2.1]):

ΔAe3TA+|A|2A=0.\operatorname{\Delta}A-\nabla_{e_{3}^{T}}A+|A|^{2}\,A=0.

Hence,

(ΔA)(v1,v1)\displaystyle(\operatorname{\Delta}A)(v_{1},v_{1}) =|A|2A(v1,v1)+(e3TA)(v1,v1)\displaystyle=-|A|^{2}A(v_{1},v_{1})+(\nabla_{e_{3}^{T}}A)(v_{1},v_{1})
=|A|2κ1+e3Tκ1\displaystyle=-|A|^{2}\kappa_{1}+\nabla_{e_{3}^{T}}\kappa_{1}

by (6.3). Thus

Δfκ1=|A|2κ1+2Q2κ1κ2.\operatorname{\Delta}^{f}\kappa_{1}=-|A|^{2}\kappa_{1}+\frac{2Q^{2}}{\kappa_{1}-\kappa_{2}}. (6.7)

Recall that the drift Laplacian (see (3.7)) is given by

Δf:=Δf,=Δe3,.\operatorname{\Delta}^{f}:=\operatorname{\Delta}-\langle\nabla f,\nabla\cdot\rangle=\operatorname{\Delta}-\langle e_{3},\nabla\cdot\rangle.

Likewise,

Δfκ2=|A|2κ22Q2κ1κ2.\operatorname{\Delta}^{f}\kappa_{2}=-|A|^{2}\kappa_{2}-\frac{2Q^{2}}{\kappa_{1}-\kappa_{2}}.

Adding these gives

ΔfH=|A|2H.\operatorname{\Delta}^{f}H=-|A|^{2}H. (6.8)

From (6.7) and (6.8),

κ1ΔfHHΔfκ1=2HQ2κ1κ2.\kappa_{1}\operatorname{\Delta}^{f}H-H\operatorname{\Delta}^{f}\kappa_{1}=\frac{-2HQ^{2}}{\kappa_{1}-\kappa_{2}}.

Thus

Δf(Hκ1)\displaystyle\operatorname{\Delta}^{f}\left(\frac{H}{\kappa_{1}}\right) =κ1ΔfHHΔfκ1κ122κ1κ1(Hκ1)\displaystyle=\frac{\kappa_{1}\operatorname{\Delta}^{f}H-H\operatorname{\Delta}^{f}\kappa_{1}}{\kappa_{1}{}^{2}}-2\,\frac{\nabla\kappa_{1}}{\kappa_{1}}\cdot\nabla\left(\frac{H}{\kappa_{1}}\right) (6.9)
=2HQ2(κ1κ2)(κ1)22κ1κ1(Hκ1),\displaystyle=\frac{-2HQ^{2}}{(\kappa_{1}-\kappa_{2})(\kappa_{1})^{2}}-2\,\frac{\nabla\kappa_{1}}{\kappa_{1}}\cdot\nabla\left(\frac{H}{\kappa_{1}}\right),

so

Δf(Hκ1)+2κ1κ1(Hκ1)0.\operatorname{\Delta}^{f}\left(\frac{H}{\kappa_{1}}\right)+2\,\frac{\nabla\kappa_{1}}{\kappa_{1}}\cdot\nabla\left(\frac{H}{\kappa_{1}}\right)\leq 0. (6.10)
Theorem 6.1 (Spruck-Xiao).

Let M𝐑3M\subset\mathbf{R}^{3} be a complete translator with H>0H>0. Then MM is convex.

Proof 6.2.

Suppose the theorem is false. Then

η:=infHκ1=inf(1+κ2κ1)[0,1).\eta:=\inf\frac{H}{\kappa_{1}}=\inf\left(1+\frac{\kappa_{2}}{\kappa_{1}}\right)\in[0,1). (6.11)

Note that the set of points where H/κ1<2H/\kappa_{1}<2 contains no umbilic points, which implies that H/κ1H/\kappa_{1} is smooth on that set and that we can apply the formulas in this section (§6), which were derived assuming that κ1>κ2\kappa_{1}>\kappa_{2}.

Step 1: The infimum η\eta is not attained. For suppose it is attained at some point. Then by (6.10) and the strong minimum principle (see [gilbarg-trudinger, Theorem 3.5] or [evans-book, Chapter 6, Theorem 3]), H/κ1H/\kappa_{1} is constant on MM. Therefore κ2/κ1=H/κ11\kappa_{2}/\kappa_{1}=H/\kappa_{1}-1 is constant on MM. Since H>0H>0 and since H/κ1H/\kappa_{1} is constant, we see from (6.9) that Q0Q\equiv 0, i.e., (see (6.6)) h12,1h12,20h_{12,1}\equiv h_{12,2}\equiv 0. Hence by (6.5), (6.1) and (6.2), the frame {v1,v2}\{v_{1},v_{2}\} is parallel, so κ1κ20\kappa_{1}\kappa_{2}\equiv 0, contadicting (6.11). Thus the infimum is not attained.

Step 2: If pnMp_{n}\in M is a sequence with H(pn)/κ1(pn)ηH(p_{n})/\kappa_{1}(p_{n})\to\eta, then (after passing to a subsequence) MpnM-p_{n} converges smoothly to a limit by the curvature estimates mentioned at the end of Section 3. We claim that the limit must be a vertical plane. For suppose not. Then (after passing to a subsequence) MpnM-p_{n} converges smoothly to a complete translator MM^{\prime} with H>0H>0. By the smooth convergence, H/κ1H/\kappa_{1} attains its minimum value on MM^{\prime} at the origin, and that minimum value is η\eta, contradicting Step 1.

Step 3: Now we apply the Omori-Yau maximum principle (see Theorem 6.3 below) to get a sequence pnMp_{n}\in M such that

Hκ1=1+κ1κ2\displaystyle\frac{H}{\kappa_{1}}=1+\frac{\kappa_{1}}{\kappa_{2}} η,\displaystyle\to\eta, (6.12)
(Hκ1)\displaystyle\nabla\left(\frac{H}{\kappa_{1}}\right) 0,\displaystyle\to 0, (6.13)
Δ(Hκ1)\displaystyle\operatorname{\Delta}\left(\frac{H}{\kappa_{1}}\right) δ[0,].\displaystyle\to\delta\in[0,\infty]. (6.14)

From (6.13) and (6.14), we see that

Δf(Hκ1)δ[0,].\operatorname{\Delta}^{f}\left(\frac{H}{\kappa_{1}}\right)\to\delta\in[0,\infty]. (6.15)

By Step 2, we can assume that MpnM-p_{n} converges smoothly to a vertical plane. For the rest of the proof, any statement that some quantity tends to a limit refers only to the quantity at the points pnp_{n}.

Since AA is a quadratic form with eigenvalues κ1\kappa_{1} and κ2\kappa_{2}, A/κ1A/\kappa_{1} is a a quadratic form with eigenvalues 11 and κ2/κ1=H/κ11\kappa_{2}/\kappa_{1}=H/\kappa_{1}-1. Thus (by passing to a subsequence) we can assume that A/κ1A/\kappa_{1} (at pnp_{n}) converges to a quadratic form with eigenvalues 11 and η1\eta-1. (Note that the eigenvalue η1\eta-1 is negative by Hypothesis (6.11).)

Recall that

H=A(𝐞3T,).\nabla H=A(\mathbf{e}_{3}^{T},\cdot).

(See, for example, [mss, Lemma 2.1].) Since 𝐞3T𝐞3\mathbf{e}_{3}^{T}\to\mathbf{e}_{3}, we see that H/κ1\nabla H/\kappa_{1} (at pnp_{n}) converges to a nonzero vector NN:

Hκ1N0.\frac{\nabla H}{\kappa_{1}}\to N\neq 0. (6.16)

Now

(Hκ1)=Hκ1Hκ1κ1κ1.\nabla\left(\frac{H}{\kappa_{1}}\right)=\frac{\nabla H}{\kappa_{1}}-\frac{H}{\kappa_{1}}\frac{\nabla\kappa_{1}}{\kappa_{1}}. (6.17)

By Omori-Yau (see (6.13)), this tends to 0, so

Hκ1κ1κ1N,\frac{H}{\kappa_{1}}\frac{\nabla\kappa_{1}}{\kappa_{1}}\to N, (6.18)

or, equivalently (see (6.4)),

Hκ1h11,iκ1Ni(i=1,2),\frac{H}{\kappa_{1}}\frac{h_{11,i}}{\kappa_{1}}\to N_{i}\qquad(i=1,2), (6.19)

where Ni=NviN_{i}=N\cdot v_{i}.

We can rewrite (6.17) as

(Hκ1)\displaystyle\nabla\left(\frac{H}{\kappa_{1}}\right) =κ1+κ2κ1Hκ1κ1κ1\displaystyle=\frac{\nabla\kappa_{1}+\nabla\kappa_{2}}{\kappa_{1}}-\frac{H}{\kappa_{1}}\frac{\nabla\kappa_{1}}{\kappa_{1}}
=(1Hκ1)(κ1κ1)+κ2κ1.\displaystyle=\left(1-\frac{H}{\kappa_{1}}\right)\left(\frac{\nabla\kappa_{1}}{\kappa_{1}}\right)+\frac{\nabla\kappa_{2}}{\kappa_{1}}.

Multiply by H/κ1H/\kappa_{1}:

Hκ1(Hκ1)=(1Hκ1)(Hκ1κ1κ1)+Hκ1κ2κ1\frac{H}{\kappa_{1}}\nabla\left(\frac{H}{\kappa_{1}}\right)=\left(1-\frac{H}{\kappa_{1}}\right)\left(\frac{H}{\kappa_{1}}\frac{\nabla\kappa_{1}}{\kappa_{1}}\right)+\frac{H}{\kappa_{1}}\frac{\nabla\kappa_{2}}{\kappa_{1}}

By Omori-Yau (see (6.13)), this tends to 0, so (using (6.18)),

Hκ1κ2κ1(η1)N,\frac{H}{\kappa_{1}}\frac{\nabla\kappa_{2}}{\kappa_{1}}\to(\eta-1)N, (6.20)

or, equivalently (by (6.4)),

Hκ1h22,iκ1(η1)Ni(i=1,2).\frac{H}{\kappa_{1}}\frac{h_{22,i}}{\kappa_{1}}\to(\eta-1)N_{i}\qquad(i=1,2). (6.21)

Combining (6.19) with i=2i=2 and (6.21) with i=1i=1 gives

(Hκ1)2(Qκ1)2(N2)2+(η1)2(N1)2:=λ2>0.\left(\frac{H}{\kappa_{1}}\right)^{2}\left(\frac{Q}{\kappa_{1}}\right)^{2}\to(N_{2})^{2}+(\eta-1)^{2}(N_{1})^{2}:=\lambda^{2}>0. (6.22)

Note that λ2>0\lambda^{2}>0 because N0N\neq 0 and η<1\eta<1 by Hypothesis (6.11).

Now multiply (6.9) by H/κ1H/\kappa_{1}:

(Hκ1)Δf(Hκ1)=21κ2κ1(Hκ1)2(Qκ1)22(Hκ1κ1κ1)(Hκ1).\left(\frac{H}{\kappa_{1}}\right)\operatorname{\Delta}^{f}\left(\frac{H}{\kappa_{1}}\right)=\frac{-2}{1-\frac{\kappa_{2}}{\kappa_{1}}}\left(\frac{H}{\kappa_{1}}\right)^{2}\left(\frac{Q}{\kappa_{1}}\right)^{2}-2\left(\frac{H}{\kappa_{1}}\frac{\nabla\kappa_{1}}{\kappa_{1}}\right)\cdot\nabla\left(\frac{H}{\kappa_{1}}\right).

Using (6.12) and (6.15) for the left side, (6.12) and (6.22) for the first term on the right, and (6.18) and (6.13) for the second term, we can let nn\to\infty to get:

ηδ22ηλ2+0,\eta\,\delta\leq\frac{-2}{2-\eta}\lambda^{2}+0,

a contradiction (since η\eta and δ\delta are nonnegative).

We used the Omori-Yau Theorem (see, for example, [Alias-et-al]):

Theorem 6.3 (Omori-Yau Theorem).

Let MM be a complete Riemannian manifold with Ricci curvature bounded below. Let f:M𝐑f:M\to\mathbf{R} be a smooth function that is bounded below. Then there is a sequence pnp_{n} in MM such that

f(pn)infMf,\displaystyle f(p_{n})\to\inf_{M}f, (6.23)
f(pn)0,\displaystyle\nabla f(p_{n})\to 0, (6.24)
lim infΔf(pn)0.\displaystyle\liminf\operatorname{\Delta}f(p_{n})\geq 0. (6.25)

The theorem remains true if we replace the assumption that ff is smooth by the assumption that ff is smooth on {f<α}\{f<\alpha\} for some α>infMf\alpha>\inf_{M}f.

To see the last assertion, let ϕ:𝐑𝐑\phi:\mathbf{R}\to\mathbf{R} be a smooth, monotonic function such that ϕ(t)=0\phi(t)=0 for tαt\geq\alpha and such that ϕ1\phi\equiv 1 on an open interval containing infMf\inf_{M}f. Then ϕf\phi\circ f is smooth, so the Omori-Yau Theorem holds for ϕf\phi\circ f, from which it follows immediately that the Omori-Yau Theorem also holds for ff.

In our case, the function H/κ1H/\kappa_{1} is smooth except at umbilic points. At such points, H/κ1=2H/\kappa_{1}=2. Since we assumed that the infimum was <1<1, we could invoke the Omori-Yau Theorem.

7 Characterization of translating graphs in 𝐑3\mathbf{R}^{3}

As we mentioned before, the authors of these notes have obtained the complete classification of the complete graphical translators in Euclidean 33-space.

Recall that by translator we mean a smooth function u:Ω𝐑u:\Omega\rightarrow\mathbf{R} such that M=Graph(u)M=\operatorname{Graph}(u) is a translator. Then uu must be solution of the equation:

(1+uy2)uxx2uxuyuxy+(1+ux2)uyy+ux2+uy2+1=0.(1+u_{y}^{2})\,u_{xx}-2u_{x}\,u_{y}\,u_{xy}+(1+u_{x}^{2})\,u_{yy}+u_{x}^{2}+u_{y}^{2}+1=0. (7.1)

If we impose that MM is complete, then we will say that uu is a complete translator. In this setting Shahriyari [shari] proved in her thesis the following

Theorem 7.1 (Shahriyari).

If MM is complete, then Ω\Omega must be a strip, a halfspace, or all of 𝐑2\mathbf{R}^{2}.

In [wang], X. J. Wang proved that the only entire convex translating graph is the bowl soliton, and that there are no complete translating graphs defined over halfplanes. Thus by the Spruck-Xiao Convexity Theorem, the bowl soliton is the only complete translating graph defined over a plane or halfplane.

Hence, it remained to classify the translators u:Ω𝐑u:\Omega\to\mathbf{R} whose domains are strips. Our main new contributions in this line are:

  1. 1.

    For each b>π/2b>\pi/2, we prove ([himw][Theorem 5.7]) existence and uniqueness (up to translation) of a complete translator ub:𝐑×(b,b)𝐑u^{b}:\mathbf{R}\times(-b,b)\to\mathbf{R} that is not a tilted grim reaper. We call ubu^{b} the Δ\operatorname{\Delta}-wing of width 2b2b.

  2. 2.

    We give a simpler proof (see [himw][Theorem 6.7]) that there are no complete graphical translators in 𝐑3\mathbf{R}^{3} defined over halfplanes in 𝐑2\mathbf{R}^{2}.

Furthermore, there are no complete translating graphs defined over strips of width <π<\pi (see [spruck-xiao, bourni-et-al]), and the grim reaper surface is the only translating graph over a strip of width π\pi (see [himw]). Consequently, we have a classification: every complete, translating graph in 𝐑3\mathbf{R}^{3} is one of the following: a grim reaper surface or tilted grim reaper surface, a Δ\operatorname{\Delta}-wing, or the bowl soliton.

We remark that Bourni, Langford, and Tinaglia gave a different proof of the existence (but not uniqueness) of the Δ\operatorname{\Delta}-wings in (1) [bourni-et-al].

{bibsection}

[References]

  • \ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry
\ProcessBibTeXEntry