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

The Structure of Translating Surfaces with Finite Total Curvature

Ilyas Khan Ilyas Khan
University of Wisconsin–Madison
480 Lincoln Drive
Madison, WI 53706, USA
ikhan4@wisc.edu
Abstract.

In this paper, we prove that any mean curvature flow translator Σ23\Sigma^{2}\subset\mathbb{R}^{3} with finite total curvature and one end must be a plane. We also prove that if the translator Σ\Sigma has multiple ends, they are asymptotic to a plane Π\Pi containing the direction of translation and can be written as graphs over Π\Pi. Finally, we determine that the ends of Σ\Sigma are strongly asymptotic to Π\Pi and obtain quantitative estimates for their asymptotic behavior.

1. Introduction

A proper, embedded hypersurface Σ23\Sigma^{2}\subset\mathbb{R}^{3} is a mean curvature flow translator if there is a unit vector VV such that Σ\Sigma satisfies the following equation:

(1.1) H=V,\vec{H}=V^{\perp},

where H\vec{H} is the mean curvature of Σ\Sigma at a point pp, and VV^{\perp} is the component of VV perpendicular to the tangent plane to Σ\Sigma at pp. This is equivalent to saying that the family of surfaces {Σ+tV:t(,)}\{\Sigma+tV\;:\;t\in(-\infty,\infty)\} satisfies the mean cuvature flow (MCF) equation,

(1.2) (t𝐱)=H(\partial_{t}\mathbf{x})^{\perp}=\vec{H}

where 𝐱(p)\mathbf{x}(p) is the vector in 3\mathbb{R}^{3} corresponding to the position of the point pΣp\in\Sigma. Translating solitons, or translators, are important singularity models for Type II singularities of the MCF and are in general highly significant examples of ancient, immortal and eternal solutions of MCF.

In this paper, we prove the following structure theorems for embedded translators with finite total curvature.

Theorem 1.1.

Let Σ23\Sigma^{2}\hookrightarrow\mathbb{R}^{3} be a complete embedded MCF translator with one end, and finite total curvature

(1.3) Σ|A|2<.\int_{\Sigma}|A|^{2}<\infty.

Then, Σ\Sigma is a plane parallel to VV, the direction of motion.

We also prove

Theorem 1.2.

Let Σ\Sigma be a complete embedded translator with finite total curvature, as in Theorem 1.1. Outside of some ball BRB_{R}, the ends of Σ\Sigma may be written as the graphs of functions uiu_{i} over a fixed plane Π\Pi. Furthermore, each uiu_{i} decays at a rate o(r14(1+δ))o\big{(}r^{-\frac{1}{4(1+\delta)}}\big{)} for any δ>0\delta>0 as rr\rightarrow\infty. In particular, each uiu_{i} decays radially at an exponential rate in any sector in PP that excludes the ray tV-tV, t(r0,)t\in(r_{0},\infty), where r0r_{0} is any non-negative number.

These results can be seen as addressing translator versions of the analogous classical questions for minimal surfaces. We now briefly summarize the proof.

The main tools used to prove Theorem 1.1 are a strong maximum principle for translators, a weak maximum principle for the gradient of a graphical translator, and a lemma due to Leon Simon in [Si] which gives an approximate graphical decomposition of surfaces with small total curvature. We begin by cutting out a large “high curvature” region of Σ\Sigma, leaving only disjoint annular ends with small total curvature. We choose one of these ends, and paste in a disk with small total curvature bounded by the total curvature of the annular end. We then apply Simon’s lemma to this new disk of small total curvature ϵ2\epsilon^{2}, which is a translator outside of a small fixed radius. The lemma tells us that, away from some “pimples” of small diameter, the surface is “mostly” a graph with gradient bounded by Cϵ1/6C\epsilon^{1/6} over some plane. Away from a neighborhood of the center, we use the strong maximum principle and Schoen’s method of moving planes to show that the pimples must be graphs. Then, using the weak maximum principle for gradients, we show that these graphical pimples must also have small gradient bounded by Cϵ1/6C\epsilon^{1/6}. Thus, inside any annulus BηBCηϵ1/2B_{\eta}\setminus B_{C\eta\epsilon^{1/2}}, any end can be written as a graph over a plane with gradient bounded by Cϵ1/6C\epsilon^{1/6}. If we fix an annulus BRBR1B_{R}\setminus B_{R^{-1}} we can take a blow-down subsequence λiΣ\lambda_{i}\Sigma that can be written in BRBR1B_{R}\setminus B_{R^{-1}} as a sequence of graphs over a single plane Π\Pi with gradient bounded by Cϵi1/6C\epsilon_{i}^{1/6} where ϵi0\epsilon_{i}\rightarrow 0. Then, if we assume we have only one end, we can re-apply the moving planes method and the maximum principle to the blow-down sequence to extract Theorem 1.1 as a consequence.

Note that the application of Simon’s lemma requires that the density of Σ\Sigma, the quantity r22(Br(x)Σ)r^{-2}\mathcal{H}^{2}(B_{r}(x)\cap\Sigma), is uniformly bounded when taken over all centers xx in 3\mathbb{R}^{3} and radii r>0r\in\mathbb{R}_{>0}. In Corollary 2.11, we prove the somewhat stronger result that this is true for all surfaces with finite total curvature.

In order to prove Theorem 1.2, we must prove that if we take two points in a convergent blow-down sequence, λi>λi+1\lambda_{i}>\lambda_{i+1}, then the “in between” annular region Σ(BR1λi+11BRλi1)\Sigma\cap(B_{R^{-1}\lambda^{-1}_{i+1}}\setminus B_{R\lambda^{-1}_{i}}) can be written as a graph over the blow-down limit plane Π\Pi. This is shown by examining two cases: (1) that Σ(BRλi+11BR1λi+11)\Sigma\cap(B_{R\lambda^{-1}_{i+1}}\setminus B_{R^{-1}\lambda_{i+1}^{-1}}) has the same orientation as a graph over Π\Pi as Σ(BRλi1BR1λi1)\Sigma\cap(B_{R\lambda^{-1}_{i}}\setminus B_{R^{-1}\lambda_{i}^{-1}}), and (2) that these two annular regions have different orientations as graphs over Π\Pi. In the first case, geometric considerations allow us to apply Schoen’s method of moving planes and the strong maximum principle to obtain graphicality over Π\Pi. In the second case, we integrate the geodesic curvature over the boundary of Σ(BR1λi+11BRλi1)\Sigma\cap(B_{R^{-1}\lambda^{-1}_{i+1}}\setminus B_{R\lambda_{i}^{-1}}) and use the Gauss-Bonnet theorem to show that the surface must have total curvature uniformly bounded below by a constant. This contradicts the fact that the end has arbitrarily small total curvature outside a large radius of our choice. Then, application of the maximum principle and moving along the blow-down sequence allows us to conclude that each end may be written as a graph with sublinear growth over a single plane.

In the final section, we prove the strong asymptotics of the ends Σi=graph(ui)\Sigma_{i}=\textrm{graph}(u_{i}), i=1,,Mi=1,\ldots,M. Our main tool is the maximum principle: first we use an exponential barrier to bound the growth of the functions uiu_{i} in the upper half-plane as x1x_{1}\rightarrow\infty. In particular, this decay implies the uniform boundedness of the uiu_{i} on the entire domain. This in turn allows us to obtain a radially decaying barrier based on the modified Bessel function of the second type. This gives the stated asymptotics in Theorem 1.2 and completes the proof.

2. Preliminaries

First, we prove a strong maximum principle for translators.

Lemma 2.1.

Let V=(v1,v2,v3)V=(v_{1},v_{2},v_{3}) be a unit vector in the direction of motion of the translating graphs of u1u_{1} and u2u_{2}, which are defined on an open set Ω2\Omega\subset\mathbb{R}^{2}. Then, v=u2u1v=u_{2}-u_{1} satisfies an equation of the form

div(aijDjv)+biDiv=0\operatorname{div}(a_{ij}D_{j}v)+b_{i}D_{i}v=0

where aija_{ij} has ellipticity constants Λ\Lambda and λ\lambda depending only on the upper bounds for the gradients |Dui||Du_{i}|.

Proof.

The uαu_{\alpha}, α=1,2\alpha=1,2, both satisfy the quasilinear elliptic equation

(2.1) div(Duα(1+|Duα|2)1/2)=v1D1uαv2D2uα+v3(1+|Duα|2)1/2.\operatorname{div}\bigg{(}\frac{Du_{\alpha}}{(1+|Du_{\alpha}|^{2})^{1/2}}\bigg{)}=\frac{-v_{1}D_{1}u_{\alpha}-v_{2}D_{2}u_{\alpha}+v_{3}}{(1+|Du_{\alpha}|^{2})^{1/2}}.

Set

A(p)=p(1+|p|2)1/2B(p)=V(p,1)(1+|p|2)1/2.A(p)=\frac{p}{(1+|p|^{2})^{1/2}}\;\;\;B(p)=\frac{V\cdot(-p,1)}{(1+|p|^{2})^{1/2}}.

We follow the proofs of Gilbarg and Trudinger in [GT, Theorem 10.7] and Colding and Minicozzi in [CM, Lemma 1.26]. Let v=u2u1v=u_{2}-u_{1}, and ut=tu2+(1t)u1u_{t}=tu_{2}+(1-t)u_{1}. Let

aij(x)=01DpjAi(Dut)𝑑t and bi(x)=01DpiB(Dut)𝑑t.a_{ij}(x)=\int_{0}^{1}D_{p_{j}}A^{i}(Du_{t})dt\;\;\text{ and }b_{i}(x)=\int_{0}^{1}D_{p_{i}}B(Du_{t})dt.

By the Fundamental Theorem of calculus and the chain rule, vv satisfies

div(aijDjv)+biDiv=0.\operatorname{div}(a_{ij}D_{j}v)+b_{i}D_{i}v=0.

All that remains to show is the uniform ellipticity of aija_{ij}. We show this by demonstrating the positive definiteness of the matrix differential DADA. If ν\nu is a unit vector and p2p\in\mathbb{R}^{2}, then

DA(p)ν=ν(1+|p|2)1/2p,ν(1+|p|2)3/2pDA(p)\nu=\frac{\nu}{(1+|p|^{2})^{1/2}}-\frac{\langle p,\nu\rangle}{(1+|p|^{2})^{3/2}}p
(1+|p|2)3/2ν,DA(p)ν\displaystyle(1+|p|^{2})^{3/2}\langle\nu,DA(p)\nu\rangle =(1+|p|2)p,ν2\displaystyle=(1+|p|^{2})-\langle p,\nu\rangle^{2}
(1+|p|2)|p|2=1.\displaystyle\geq(1+|p|^{2})-|p|^{2}=1.

Thus, aij=01DA(ut)𝑑ta_{ij}=\int_{0}^{1}DA(u_{t})dt is a positive definite matrix whose ellipticity constants depend on the upper bounds of |Duα||Du_{\alpha}|. ∎

Corollary 2.2.

Let Ω2\Omega\subset\mathbb{R}^{2} be an open connected neighborhood of the origin. If u1,u2:Ωu_{1},u_{2}:\Omega\rightarrow\mathbb{R} satisfy the translator equation (2.1) with respect to the direction V=(v1,v2,v3)V=(v_{1},v_{2},v_{3}) with u1u2u_{1}\leq u_{2} and u1(0)=u2(0)u_{1}(0)=u_{2}(0), then u1u2u_{1}\equiv u_{2}.

Proof.

Immediate from Lemma 2.1 and the strong maximum principle proved in [GT, Lemma 3.5]. ∎

In addition to this maximum principle, we also obtain a weak maximum principle for the gradients of translators that are graphs over planes containing VV, the direction of motion.

Lemma 2.3.

If V2×{0}3V\in\mathbb{R}^{2}\times\{0\}\subset\mathbb{R}^{3}, Ω\Omega is a bounded domain in 2×{0}\mathbb{R}^{2}\times\{0\}, and u:Ω2u:\Omega\subset\mathbb{R}^{2}\rightarrow\mathbb{R} satisfies (2.1), then the partial derivatives D1uD_{1}u and D2uD_{2}u achieve their maxima and minima on the boundary.

Proof.

We set V=(1,0,0)V=(1,0,0) without loss of generality. The equation (2.1) becomes

(2.2) div(Du(1+|Du|2)1/2)=D1u(1+|Du|2)1/2.\operatorname{div}\bigg{(}\frac{Du}{(1+|Du|^{2})^{1/2}}\bigg{)}=\frac{-D_{1}u}{(1+|Du|^{2})^{1/2}}.

We differentiate this equation with respect to xix_{i} to obtain

Didiv(Du(1+|Du|2)1/2)=D1(Diu)(1+|Du|2)1/2+D1u(1+|Du|2)3/2DuD(Diu)D_{i}\operatorname{div}\bigg{(}\frac{Du}{(1+|Du|^{2})^{1/2}}\bigg{)}=\frac{-D_{1}(D_{i}u)}{(1+|Du|^{2})^{1/2}}+\frac{D_{1}u}{(1+|Du|^{2})^{3/2}}Du\cdot D(D_{i}u)
Di(Δu(1+|Du|2)1/2DjuDkuDkju(1+|Du|2)3/2)=(D1(Diu))(1+|Du|2)1/2+D1u(DuD(Diu))(1+|Du|2)3/2D_{i}\bigg{(}\frac{\Delta u}{(1+|Du|^{2})^{1/2}}-\frac{D_{j}uD_{k}uD_{kj}u}{(1+|Du|^{2})^{3/2}}\bigg{)}=\frac{-(D_{1}(D_{i}u))}{(1+|Du|^{2})^{1/2}}+\frac{D_{1}u(Du\cdot D(D_{i}u))}{(1+|Du|^{2})^{3/2}}

Let v=Diuv=D_{i}u. We calculate

(1+|Du|2)ΔvDjuDkuDkjv=2DjvDkuDjku(D1v)(1+|Du|2)+D1u(DuDv)+Δu(DvDu)3DjuDkuDkju(DuDv)1+|Du|2(1+|Du|^{2})\Delta v-D_{j}uD_{k}uD_{kj}v\\ =2D_{j}vD_{k}uD_{jk}u-(D_{1}v)(1+|Du|^{2})+D_{1}u(Du\cdot Dv)\\ +\Delta u(Dv\cdot Du)-\frac{3D_{j}uD_{k}uD_{kj}u(Du\cdot Dv)}{1+|Du|^{2}}

This implies that vv satisfies a differential equation of the form Lv=0Lv=0, where the differential operator L=ajkDjk+bjDjL=a_{jk}D_{jk}+b_{j}D_{j} has coefficients given by

ajk=(1+|Du|2)δjkDjuDkua_{jk}=(1+|Du|^{2})\delta_{jk}-D_{j}uD_{k}u
bj=(1+|Du|2)δ1j2DkuDjku(D1u+Δu3DkuDluDklu1+|Du|2)Djub_{j}=(1+|Du|^{2})\delta_{1j}-2D_{k}uD_{jk}u-\bigg{(}D_{1}u+\Delta u-\frac{3D_{k}uD_{l}uD_{kl}u}{1+|Du|^{2}}\bigg{)}D_{j}u

We calculate

ajkξjξk=(1+|Du|2)|ξ|2(Duξ)2(1+|Du|2)|ξ|2|Du|2|ξ|2|ξ|2.a_{jk}\xi_{j}\xi_{k}=(1+|Du|^{2})|\xi|^{2}-(Du\cdot\xi)^{2}\geq(1+|Du|^{2})|\xi|^{2}-|Du|^{2}|\xi|^{2}\geq|\xi|^{2}.

Similarly

ajkξjξk(1+2|Du|2)|ξ|2.a_{jk}\xi_{j}\xi_{k}\leq(1+2|Du|^{2})|\xi|^{2}.

Since uu is assumed to be smooth and Ω\Omega is bounded, the operator LL is uniformly elliptic. In particular, LL satisfies the hypotheses of [GT, Theorem 3.1] and by applying this theorem, we obtain the weak maximum/minimum principle for vv. ∎

We now prove a result about the pointwise decay of the second fundamental form |A||A| and its derivatives. To obtain this result, we use Ecker’s ϵ\epsilon-Regularity theorem for the mean curvature flow of surfaces.

Theorem 2.4.

[E1, Ecker Regularity Theorem] There exist constants ϵ0>0\epsilon_{0}>0 and c0>0c_{0}>0 such that for any solution (Mt)t[0,T)(M_{t})_{t\in[0,T)} of mean curvature flow, any x03x_{0}\in\mathbb{R}^{3}, and ρ(0,T)\rho\in(0,\sqrt{T}) the inequality

(2.3) sup[Tρ2,T)MtBρ(x0)|A|2ϵ0\sup_{[T-\rho^{2},T)}\int_{M_{t}\cap B_{\rho}(x_{0})}|A|^{2}\leq\epsilon_{0}

implies the mean value estimate

(2.4) sup[T(ρ/2)2,T)supMtBρ/2(x0)|A|2c0ρ4Tρ2TMtBρ(x0)|A|2.\sup_{[T-(\rho/2)^{2},T)}\sup_{M_{t}\cap B_{\rho/2}(x_{0})}|A|^{2}\leq c_{0}\rho^{-4}\int_{T-\rho^{2}}^{T}\int_{M_{t}\cap B_{\rho}(x_{0})}|A|^{2}.
Lemma 2.5.

For a translator Σ\Sigma with finite total curvature, if ϱ>1\varrho>1 is distance from the origin, then

|A|=O(ϱ1/2).|A|=O(\varrho^{-1/2}).

Furthermore, in the positive half-space defined by VV,

|A|=O(ϱ1).|A|=O(\varrho^{-1}).
Proof.

This is a consequence of Ecker’s ϵ\epsilon-regularity theorem, Theorem 2.4. By the finite total curvature condition, there is a ball BRB_{R} centered at the origin such that ΣBR|A|2<ϵ0\int_{\Sigma\setminus B_{R}}|A|^{2}<\epsilon_{0}. We pick any x03B2Rx_{0}\in\mathbb{R}^{3}\setminus B_{2R}, and take 2ρ2=dist3(x0,B2R)2\rho^{2}=\textrm{dist}_{\mathbb{R}^{3}}(x_{0},B_{2R}). Suppose that x0Σx_{0}\in\Sigma and take the ball Bρ(x0)B_{\rho}(x_{0}). Consider the mean curvature flow solution Σt=Σ+tV\Sigma_{t}=\Sigma+tV defined on the interval t[3ρ2/4,ρ2/4]t\in[-3\rho^{2}/4,\rho^{2}/4]. Notice that BRΣ+tVB_{R}\cap\Sigma+tV never intersects Bρ(x0)B_{\rho}(x_{0}) for any time t[3ρ2/4,ρ2/4]t\in[-3\rho^{2}/4,\rho^{2}/4]. Thus,

sup[3ρ2/4,ρ2/4)ΣtBρ(x0)|A|2ϵ0.\sup_{[-3\rho^{2}/4,\rho^{2}/4)}\int_{\Sigma_{t}\cap B_{\rho}(x_{0})}|A|^{2}\leq\epsilon_{0}.

Theorem 2.4, Ecker’s regularity theorem, yields

sup[0,ρ2/4)supΣtBρ/2(x0)|A|2\displaystyle\sup_{[0,\rho^{2}/4)}\sup_{\Sigma_{t}\cap B_{\rho/2}(x_{0})}|A|^{2} c0ρ43(ρ/2)2(ρ/2)2ΣtBρ(x0)|A|2\displaystyle\leq c_{0}\rho^{-4}\int_{-3(\rho/2)^{2}}^{(\rho/2)^{2}}\int_{\Sigma_{t}\cap B_{\rho}(x_{0})}|A|^{2}
c0ρ2ϵ0\displaystyle\leq c_{0}\rho^{-2}\epsilon_{0}

Thus, |A|2(x0)Cϱ1|A|^{2}(x_{0})\leq C\varrho^{-1}, where ϱ>0\varrho>0 is distance to the origin (which is comparable to 2ρ2=dist3(x0,B2R)2\rho^{2}=\textrm{dist}_{\mathbb{R}^{3}}(x_{0},B_{2R})). This proves the first statement of the Lemma.

Now we prove the improved curvature decay in the upper half-space 3\mathbb{H}\subset\mathbb{R}^{3} of points in 3\mathbb{R}^{3} with positive VV component. Suppose that x0Σx_{0}\in\mathbb{H}\cap\Sigma and that ρ=dist3(x0,BR)\rho=\textrm{dist}_{\mathbb{R}^{3}}(x_{0},B_{R}). Consider the ball Bρ(x0)B_{\rho}(x_{0}) and the MCF solution Σt=Σ+tV\Sigma_{t}=\Sigma+tV defined on the interval t[ρ2,0]t\in[\rho^{2},0]. Since x0V>0x_{0}\cdot V>0, ΣtBR\Sigma_{t}\cap B_{R} never intersects Bρ(x0)B_{\rho}(x_{0}). Thus, by Theorem 2.4,

sup[ρ2/4,0]supΣtBρ/2(x0)|A|2c0ρ2ϵ0.\sup_{[-\rho^{2}/4,0]}\sup_{\Sigma_{t}\cap B_{\rho/2}(x_{0})}|A|^{2}\leq c_{0}\rho^{-2}\epsilon_{0}.

Choosing t=0t=0, we see that |A|2(x0)Cρ2|A|^{2}(x_{0})\leq C\rho^{-2}. Since ρϱ\rho\approx\varrho, where ϱ>0\varrho>0 is |x0||x_{0}|, we conclude that |A|=O(ϱ1)|A|=O(\varrho^{-1}). ∎

Corollary 2.6.

If 𝐧(p)\mathbf{n}(p) is the normal vector to a translator Σ\Sigma moving in the direction VV at the point pp, then

𝐧(p)V=O(ϱ1/2),\mathbf{n}(p)\cdot V=O(\varrho^{-1/2}),

where ϱ\varrho is the distance between pp and the origin.

Proof.
|H|22|A|2,H(p)=𝐧(p)V.|H|^{2}\leq 2|A|^{2},\;\;H(p)=\mathbf{n}(p)\cdot V.

Therefore,

𝐧(p)V=O(ϱ1/2).\mathbf{n}(p)\cdot V=O(\varrho^{-1/2}).

Before we can begin the proof of the main theorems, we must show that the embedding of Σ\Sigma into 3\mathbb{R}^{3} is proper and that the area ratios, centered at any point, are uniformly bounded. We state the following theorem of Hartman which bounds the intrinsic area growth.

Theorem 2.7.

[LT, Proposition 1.3] Let Σ\Sigma be a complete surface with finite total curvature. Then there exists a constant C1>0C_{1}>0 depending only on Σ|K|𝑑A\int_{\Sigma}|K|dA, such that

Area(BrΣ(x))C1r2,\textrm{Area}(B^{\Sigma}_{r}(x))\leq C_{1}r^{2},

for all xΣx\in\Sigma and for all r>0r>0, where BrΣ(x)B^{\Sigma}_{r}(x) is the intrinsic ball of radius rr.

We will extend this statement to the analogous statement for extrinsic balls by comparing intrinsic and extrinsic distances. The following theorem due to Huber gives important information about the topology and geometry of Σ\Sigma.

Theorem 2.8 (Huber).

If Σ\Sigma is a complete surface of finite total curvature immersed in n\mathbb{R}^{n}, then Σ\Sigma is conformally equivalent to a compact Riemann surface with finitely many points deleted.

In particular, we note that Huber’s theorem implies that Σ\Sigma must have finitely many ends and bounded genus. By Huber’s theorem, we have a bijective conformal map, f:ΩΣf:\Omega^{*}\hookrightarrow\Sigma, from the annulus Ω:=B1(0)\Omega^{*}:=\mathbb{C}\setminus B_{1}(0) to a given end of Σ\Sigma. Following [MS], we consider the family of rescalings

fϵ(z)=ϵα+1[f(z/ϵ)f(2/ϵ)].f_{\epsilon}(z)=\epsilon^{\alpha+1}[f(z/\epsilon)-f(2/\epsilon)].

According to [MS], there exists a sequence ϵk0\epsilon_{k}\rightarrow 0 with the following properties:

  • (i)

    There exists a plane LGr(2,3)L\in\textrm{Gr}(2,3) such that the Gauss map converges to the constant LL in Wloc1,2W^{1,2}_{loc} on the punctured complex plane.

  • (ii)

    The maps fϵkf_{\epsilon_{k}} converge uniformly on compact subsets of {0}\mathbb{C}\setminus\{0\} to a conformal mapping f0:{0}L3f_{0}:\mathbb{C}\setminus\{0\}\rightarrow L\subset\mathbb{R}^{3} which satisfies |xif0(z)|=C|z|α|\partial_{x_{i}}f_{0}(z)|=C|z|^{\alpha}.

If we consider f0f_{0} as a holomorphic function f0:{0}f_{0}:\mathbb{C}\setminus\{0\}\rightarrow\mathbb{C}, it can be seen that f0f_{0} is of the form

f0(z)=Czα+1+D.f_{0}(z)=Cz^{\alpha+1}+D.

However, if α>0\alpha>0, the graph of the map ff must have an self-intersection, since the fϵkf_{\epsilon_{k}} are converging uniformly as maps into 3\mathbb{R}^{3}. This contradicts the assumption of embeddedness, and thus α\alpha must be zero. Müller and Šverák define the multiplicity of an end by the integer m=α+1m=\alpha+1. Therefore, in our situation each end of Σ\Sigma has multiplicity one.

Now we state a result of Müller and Šverák that allows us to relate extrinsic distances to distances in the parameter space Ω\Omega^{*}. Given that each end has unit multiplicity, we simplify the statement of the theorem.

Proposition 2.9.

[MS, Corollary 4.2.10] Given an end of Σ\Sigma, consider the conformal parametrization f:ΩΣ3f:\Omega^{*}\rightarrow\Sigma\hookrightarrow\mathbb{R}^{3} given by Huber’s theorem. For each ϵ>0\epsilon>0 there exists R>0R>0 such that the following statement holds. If QΩQ\subset\Omega^{*} is a square such that Q{z:|z|R}=Q\cap\{z\in\mathbb{C}\;:\;|z|\leq R\}=\emptyset and if ξ1,ξ2Ω\xi_{1},\xi_{2}\in\Omega^{*} are neighboring vertices of QQ, then

eλ(1ϵ)|ξ1ξ2||f(ξ1)f(ξ2)|eλ(1+ϵ)|ξ1ξ2|,e^{\lambda}(1-\epsilon)|\xi_{1}-\xi_{2}|\leq|f(\xi_{1})-f(\xi_{2})|\leq e^{\lambda}(1+\epsilon)|\xi_{1}-\xi_{2}|,

where λ\lambda is a constant.

By [MS, Theorem 4.2.1], for ξ1,ξ2Ω\xi_{1},\xi_{2}\in\Omega^{*} as in Proposition 2.9, dΣ(f(ξ1),f(ξ2))d_{\Sigma}(f(\xi_{1}),f(\xi_{2})) approaches eλ|ξ1ξ2|e^{\lambda}|\xi_{1}-\xi_{2}| uniformly as RR\rightarrow\infty. Taking piecewise-C1C^{1}-paths, we can now compare intrinsic and extrinsic distances on Σ\Sigma. Thus, if x1,x2Σ3x_{1},x_{2}\in\Sigma\subset\mathbb{R}^{3}, there exists a constant CC such that

C1dΣ(x1,x2)|x1x2|CdΣ(x1,x2).C^{-1}d_{\Sigma}(x_{1},x_{2})\leq|x_{1}-x_{2}|\leq Cd_{\Sigma}(x_{1},x_{2}).

In particular, if x0Σx_{0}\in\Sigma, BCrΣ(x0)B^{\Sigma}_{Cr}(x_{0}) contains the connected component of Br(x0)B_{r}(x_{0}) containing x0x_{0}.

We must control the number of connected components in order to use this comparison to bound area ratios. We begin by showing that area ratios are uniformly bounded with respect to centers in any compact set.

Proposition 2.10.

If Σ\Sigma is a complete, embedded surface with finite total curvature and KK is a compact domain in 3\mathbb{R}^{3}, there exists a constant β>0\beta>0 that bounds the following area ratios:

(2.5) supxKsupR>02(ΣBR(x))R2<β.\sup_{x\in K}\sup_{R>0}\frac{\mathcal{H}^{2}(\Sigma\cap B_{R}(x))}{R^{2}}<\beta.
Proof.

Müller and Šverák prove in [MS, Corollary 4.2.5] that for a fixed point x0Σx_{0}\in\Sigma, the intrinsic and extrinsic distance functions coincide in the limit. That is,

(2.6) limdistΣ(x,x0)distΣ(x,x0)|xx0|=1.\lim_{\textrm{dist}_{\Sigma}(x,x_{0})\rightarrow\infty}\frac{\textrm{dist}_{\Sigma}(x,x_{0})}{|x-x_{0}|}=1.

Thus, for very large radii RR, the intrinsic balls are comparable to extrinsic balls in the following sense:

BRΣ(x0)BRΣB2RΣ(x0).B^{\Sigma}_{R}(x_{0})\subset B_{R}\cap\Sigma\subset B^{\Sigma}_{2R}(x_{0}).

Let A(r)A(r) denote the area of the geodesic ball BrΣ(x0)B_{r}^{\Sigma}(x_{0}). A result from [Sh] cited in the proof of [MS, Corollary 4.2.5] tells us that

limrA(r)r2=2πχΣΣK=total number of ends.\lim_{r\rightarrow\infty}\frac{A(r)}{r^{2}}=2\pi\chi_{\Sigma}-\int_{\Sigma}K=\textrm{total number of ends}.

By the finite total curvature property and Huber’s theorem, the right hand side is bounded by a fixed constant. Since intrinsic and extrinsic balls are comparable, there exists some

β>2πχΣΣK\beta>2\pi\chi_{\Sigma}-\int_{\Sigma}K

such that

2(ΣBR(x0))R2<β.\frac{\mathcal{H}^{2}(\Sigma\cap B_{R}(x_{0}))}{R^{2}}<\beta.

By comparison in the limit as RR\rightarrow\infty, we can see that this property holds for all x0Kx_{0}\in K, and not only for points contained in Σ\Sigma.

We have a family of lower semi-continuous functions defined on xKx\in K

{2(ΣBR(x))R2}R[0,).\bigg{\{}\frac{\mathcal{H}^{2}(\Sigma\cap B_{R}(x))}{R^{2}}\bigg{\}}_{R\in[0,\infty)}.

Take the supremum of this family with respect to the index RR. This is again a lower semi-continuous function with finite values, since the limiting area ratio for every center is a fixed finite number. We apply the extreme value theorem for lower semi-continuous functions to this new function on the compact set KK. This proves the proposition. ∎

A neighborhood of each end can be parametrized as in Proposition 2.9. By Huber’s theorem, the complement of these neighborhoods is contained inside some compact set K3K\subset\mathbb{R}^{3}. By Proposition 2.10, we know that the area ratios are bounded for balls centered in this region. We now consider points x3Kx\in\mathbb{R}^{3}\setminus K. We know that if xΣKx\in\Sigma\setminus K and Cx(Br(x)Σ)C_{x}(B_{r}(x)\cap\Sigma) is the connected component of Br(x)ΣB_{r}(x)\cap\Sigma containing xx, then there exists a uniform constant C2>0C_{2}>0 such that

Cx(Br(x)Σ)BCrΣ(x),C_{x}(B_{r}(x)\cap\Sigma)\subset B_{Cr}^{\Sigma}(x),

and thus,

2(Cx(Br(x)Σ))πr2C22Area(BC2rΣ(x))π(C2r)2.\frac{\mathcal{H}^{2}(C_{x}(B_{r}(x)\cap\Sigma))}{\pi r^{2}}\leq C_{2}^{2}\frac{\textrm{Area}(B_{C_{2}r}^{\Sigma}(x))}{\pi(C_{2}r)^{2}}.

For a general point x3Kx\in\mathbb{R}^{3}\setminus K, we consider the set {x1,,xd}\{x_{1},\ldots,x_{d}\}, which is the nearest point projection to the dd ends of the surface Σ\Sigma. Then,

Br(x)Σi=1dBC2rΣ(xi).B_{r}(x)\cap\Sigma\subset\bigcup_{i=1}^{d}B^{\Sigma}_{C_{2}r}(x_{i}).

The theorem of Hartman, Theorem 2.7, implies that there is a constant C3>0C_{3}>0 depending on C1C_{1} and C2C_{2} such that

2(Br(x)Σ)πr2dC3.\frac{\mathcal{H}^{2}(B_{r}(x)\cap\Sigma)}{\pi r^{2}}\leq dC_{3}.

Combining the above bound with Proposition 2.10 proves a uniform bound on area ratios.

Corollary 2.11.

There exists D>0D>0 such that

supx3supR>02(ΣBR(x))R2<D.\sup_{x\in\mathbb{R}^{3}}\sup_{R>0}\frac{\mathcal{H}^{2}(\Sigma\cap B_{R}(x))}{R^{2}}<D.

3. Blowing Down

In this section, we prove that each blow-down sequence of Σ\Sigma has a subsequential Cloc1C^{1}_{loc}-limit.

Lemma 3.1.

Let Σ\Sigma be a translating end with finite total curvature. For every sequence λi0\lambda_{i}\rightarrow 0 there is a subsequence λij0{\lambda_{i}}_{j}\rightarrow 0 such that the Cloc1C^{1}_{loc}-limit of the rescalings λijΣ{\lambda_{i}}_{j}\Sigma in 3{0}\mathbb{R}^{3}\setminus\{0\} is a plane Π\Pi that is parallel to VV.

The central tool that we will use to prove this lemma is the following graphical decomposition lemma due to L. Simon, [Si, Lemma 2.1]. We restate the version of the lemma given in [I] here for convenience.

Theorem 3.2 (Simon’s Lemma on |A|2|A|^{2}).

For each β>0\beta>0, there is a constant ϵ0=ϵ0(β)\epsilon_{0}=\epsilon_{0}(\beta) such that if Σ\Sigma is a smooth 2-manifold properly embedded in BR3B_{R}\subset\mathbb{R}^{3} and

ΣBR|A|2ϵ2ϵ02,2(ΣBR)βR2,\int_{\Sigma\cap B_{R}}|A|^{2}\leq\epsilon^{2}\leq\epsilon_{0}^{2},\;\;\;\mathcal{H}^{2}(\Sigma\cap B_{R})\leq\beta R^{2},

then there are pairwise disjoint closed disks P¯1,,P¯N\bar{P}_{1},\ldots,\bar{P}_{N} in ΣBR\Sigma\cap B_{R} such that

(3.1) mdiam(Pm)C(β)ϵ1/2R\sum_{m}\operatorname{diam}(P_{m})\leq C(\beta)\epsilon^{1/2}R

and for any S[R/4,R/2]S\in[R/4,R/2] such that Σ\Sigma intersects BS\partial B_{S} transversally and BSmP¯m=\partial B_{S}\cap\bigcup_{m}\bar{P}_{m}=\emptyset, we have

ΣBS=i=1lDi\Sigma\cap B_{S}=\cup_{i=1}^{l}D_{i}

where each DiD_{i} is an embedded disk. Furthermore, for each DiD_{i}, there is a 2-plane Li3L_{i}\subset\mathbb{R}^{3}, a simply-connected domain ΩiLi\Omega_{i}\subset L_{i}, disjoint closed balls B¯i,pΩi\bar{B}_{i,p}\subset\Omega_{i}, p=1,,pip=1,\ldots,p_{i} and a function

ui:ΩiB¯i,p3u_{i}:\Omega_{i}\setminus\cup\bar{B}_{i,p}\rightarrow\mathbb{R}^{3}

such that

(3.2) sup|uiR|+|Dui|C(β)ϵ1/6\sup\bigg{|}\frac{u_{i}}{R}\bigg{|}+|Du_{i}|\leq C(\beta)\epsilon^{1/6}

and

DimP¯m=graph(ui|ΩiB¯i,p).D_{i}\setminus\cup_{m}\bar{P}_{m}=\operatorname{graph}(u_{i}|\Omega_{i}\setminus\cup\bar{B}_{i,p}).

Roughly speaking, this lemma tells us that a surface of small total curvature contained inside a ball can be expressed as the union of disks that are graphical away from a small pathological set of discs {Pi}\{P_{i}\}, which we call “pimples.” We recall that Simon states on p. 289 of [Si] that each the boundary of each disk DiD_{i} is a graphical curve contained in BSΣ\partial B_{S}\cap\Sigma. In particular, this tells us a) that the DiD_{i} can be taken to be disjoint (they cannot be connected by a pimple PmP_{m}, as these are all topological disks) and b) that if we assume 0Σ30\in\Sigma\subset\mathbb{R}^{3}, then the connected component of ΣBS(0)\Sigma\cap B_{S}(0) containing 0 is contained in a neighborhood of the cone

Xϵ(Π,0)={y3:dist(y,Π)Cϵ16|π(y)|}.X_{\epsilon}(\Pi,0)=\{y\in\mathbb{R}^{3}:\textrm{dist}(y,\Pi)\leq C\epsilon^{\frac{1}{6}}|\pi(y)|\}.
Lemma 3.3.

Let ϵ,R>0\epsilon,R>0 be given, and let Σ\Sigma be a complete embedded translator, and select a point xΣx\in\Sigma such that

ΣBR|A|2ϵ2ϵ02,2(ΣBR(x))βπR2, and |x|>Cϵ1/3.\int_{\Sigma\cap B_{R}}|A|^{2}\leq\epsilon^{2}\leq\epsilon_{0}^{2},\;\;\;\mathcal{H}^{2}(\Sigma\cap B_{R}(x))\leq\beta\pi R^{2},\text{ and }|x|>C\epsilon^{-1/3}.

Let Σx\Sigma_{x} be the connected component of ΣBS(x)\Sigma\cap B_{S}(x) containing xx (where SS is as in Theorem 3.2) and consider the collection of pimples P¯mΣ\bar{P}_{m}\subset\Sigma. Then there exists S[R/8,R/16]S^{\prime}\in[R/8,R/16] such that ΣBS\Sigma\cap B_{S^{\prime}} is the graph of a function with gradient bounded by Cϵ16C\epsilon^{\frac{1}{6}}, defined on a plane Π\Pi parallel to the direction of motion, VV.

Proof.

Away from the pimples {Pm}\{P_{m}\}, the component Σx\Sigma_{x} is the graph of a function with small gradient over a plane, which we denote LxL_{x}. By Corollary 2.6, 𝐧(x)V=O(|x|1/2)\mathbf{n}(x)\cdot V=O(|x|^{-1/2}). Since

|x|1/2<C1ϵ1/6,|x|^{-1/2}<C^{-1}\epsilon^{1/6},

we can take LxL_{x} to be a plane Π\Pi that is parallel to VV. Let π:3Π\pi:\mathbb{R}^{3}\rightarrow\Pi be the standard orthogonal projection map. Now, consider the following cone:

Xϵ(Π,x)={y3:dist(yx,Π)Cϵ16|π(yx)|}.X_{\epsilon}(\Pi,x)=\{y\in\mathbb{R}^{3}:\textrm{dist}(y-x,\Pi)\leq C\epsilon^{\frac{1}{6}}|\pi(y-x)|\}.

By the graphical decomposition lemma, Σx\Sigma_{x} is contained in a neighborhood of Xϵ(Π,x)X_{\epsilon}(\Pi,x) with radius Cϵ1/2RC\epsilon^{1/2}R:

𝒩Cϵ1/2R(Xϵ(Π,x))={y+tν:yXϵ(Π,x),t[0,Cϵ1/2R),νS2}.\mathcal{N}_{C\epsilon^{1/2}R}(X_{\epsilon}(\Pi,x))=\{y+t\nu:y\in X_{\epsilon}(\Pi,x),\;t\in[0,C\epsilon^{1/2}R),\;\nu\in S^{2}\}.

Let DsD_{s} denote the disk centered at xx of radius ss in the plane Π\Pi. This boundedness of Σx\Sigma_{x} ensures that for sufficiently small ϵ>0\epsilon>0, the intersections π1(Ds)Σx\pi^{-1}(\partial D_{s})\cap\Sigma_{x} are compact and do not intersect the boundary Σx\partial\Sigma_{x} for all s[R/4,R/8]s\in[R/4,R/8]. By the diameter bounds on the pimples and the transversality theorem, there is a number S[R/4,R/8]S^{\prime}\in[R/4,R/8] such that π1(DS){Pm}=\pi^{-1}(\partial D_{S^{\prime}})\cap\{P_{m}\}=\emptyset and π1(DS)\pi^{-1}(\partial D_{S^{\prime}}) intersects Σx\Sigma_{x} transversely. By Theorem 3.2, this intersection is a single graphical curve η\eta over DS\partial D_{S^{\prime}} that bounds a disk, π1(DS)Σ\pi^{-1}(D_{S^{\prime}})\cap\Sigma.

We now show that the surface given by π1(DS)Σ\pi^{-1}(D_{S^{\prime}})\cap\Sigma is a graph over Π\Pi. In order to prove this, we adapt the Alexandrov moving plane method of Schoen for minimal surfaces ([CM, Theorem 1.29] and [Sc, Theorem 1]). This has been previously done for translators in [MSHS].

Theorem 3.4 (Method of Moving Planes).

Let ΩΠ\Omega\subset\Pi be an open set with C1C^{1} boundary and let {σi}3\{\sigma_{i}\}\subset\mathbb{R}^{3} be simple closed curves each of which are graphs over distinct components of Ω\partial\Omega with bounded slope. Further assume that for any point xσiΣx\in\sigma_{i}\subset\Sigma, the tangent plane to Σ\Sigma is well defined and does not contain the vertical normal vector e3e_{3}. Then any translator Σ3\Sigma\subset\mathbb{R}^{3} contained in Ω×\Omega\times\mathbb{R} with Σ=iσi\partial\Sigma=\cup_{i}\sigma_{i} must be graphical over Ω\Omega.

Proof.

Let Π\Pi be spanned by unit vectors e1,e2e_{1},e_{2} and let the normal direction be e3e_{3}. Given the plane {x3=t}\{x_{3}=t\}, Σ\Sigma is divided into the portions Σt+\Sigma^{+}_{t} above and Σt\Sigma^{-}_{t} below the plane. We reflect Σt+\Sigma^{+}_{t} below the plane to obtain a new translator Σ~t+\tilde{\Sigma}^{+}_{t} below the plane. Note that because VV is contained in Π\Pi, reflection in Π\Pi preserves the translator equation. We decrease tt until we encounter the first point pp where (a)TpΣT_{p}\Sigma contains the vector e3e_{3} or (b) Σ~t+\tilde{\Sigma}^{+}_{t} and Σt\Sigma^{-}_{t} have an interior point of contact. Let the critical height tt be called t0t_{0}. Suppose that (a) occurs–this means that the tangent planes TpΣ~t0+T_{p}\tilde{\Sigma}^{+}_{t_{0}} and TpΣt0T_{p}\Sigma^{-}_{t_{0}} coincide and we may apply the maximum principle, Corollary 2.2, to show that Σ\Sigma must have a reflection symmetry through {x3=t0}\{x_{3}=t_{0}\}. This contradicts the graphicality of the boundary unless Σ\Sigma coincides with {x3=t0}\{x_{3}=t_{0}\}.

If Σ\Sigma is not one-to-one (i.e. t0t_{0} exists) and case (a) does not occur, then case (b) must occur. If there is a first interior point of contact, we may apply Corollary 2.2 directly to show that Σ\Sigma has a reflection symmetry through {x3=t0}\{x_{3}=t_{0}\}. This is a contradiction, and thus neither case (a) nor case (b) occurs. Because Σ\partial\Sigma is a disjoint union of graphs with bounded slope over disjoint planar curves, there is no boundary point of contact. Thus, the projection of Σ\Sigma to Π\Pi is one-to-one and Σ\Sigma is a graph. ∎

If we let Ω=DSΠ\Omega=D_{S^{\prime}}\subset\Pi and σ=π1(DS)Σ\sigma=\pi^{-1}(\partial D_{S^{\prime}})\cap\Sigma, then π1(DS)Σ\pi^{-1}(D_{S^{\prime}})\cap\Sigma clearly satisfies the conditions of Theorem 3.4 and is a graph over DSD_{S^{\prime}}. Furthermore, we know from the graphical decomposition lemma, Theorem 3.2, that |Du|η|Cϵ16|Du|_{\eta}|\leq C\epsilon^{\frac{1}{6}}. By Lemma 2.3, the weak maximum principle for derivatives, π1(DS)Σ\pi^{-1}(D_{S^{\prime}})\cap\Sigma can be written as the graph of a function uu defined on DSΠD_{S^{\prime}}\subset\Pi with |Du|Cϵ16|Du|\leq C\epsilon^{\frac{1}{6}}. This completes the proof of the lemma. ∎

Given some ϵ>0\epsilon>0, we can find Rϵ>0R_{\epsilon}>0 such that

ΣBRϵ|A|2𝑑2<ϵ2,\int_{\Sigma\setminus B_{R_{\epsilon}}}|A|^{2}d\mathcal{H}^{2}<\epsilon^{2},

and ΣBRϵ\Sigma\setminus B_{R_{\epsilon}} decomposes into disjoint annular ends. Given one of these ends, Σi\Sigma_{i}, we wish to glue a disk with small total curvature to the inner boundary. The resulting surface Σi\Sigma_{i}^{\prime} will be a topological disk without boundary to which we may apply Simon’s Lemma (Theorem 3.2). To this end, for each RϵR_{\epsilon}, we find a cylindrical curve with special properties that will allow us to carry out the gluing construction.

Lemma 3.5.

Given ϵ>0\epsilon>0 and Rϵ>0R_{\epsilon}>0 as above, there exists a plane Π\Pi such that VΠV\in\Pi, a radius ρ>0\rho>0 such that π(ΣBRϵ)DρΠ\pi(\Sigma\cap B_{R_{\epsilon}})\subset D_{\rho}\subset\Pi, and a graphical curve ΓΣ\Gamma\subset\Sigma over DρΠ\partial D_{\rho}\subset\Pi such that

Γ|A|2𝑑1Cϵ2ρ.\int_{\Gamma}|A|^{2}d\mathcal{H}^{1}\leq\frac{C\epsilon^{2}}{\rho}.
Proof.

By Corollary 2.6, we may choose RϵR_{\epsilon} large enough that 𝐧V<Cϵ16\mathbf{n}\cdot V<C\epsilon^{\frac{1}{6}} in ΣBRϵ\Sigma\setminus B_{R_{\epsilon}}. Let f(x)=|x|f(x)=|x|. The norm of the tangential gradient |Σf||\nabla^{\Sigma}f| is bounded by 1. The coarea formula says that

2Rϵ3RϵΣ{f=r}|A|𝑑1𝑑r\displaystyle\int_{2R_{\epsilon}}^{3R_{\epsilon}}\int_{\Sigma\cap\{f=r\}}|A|d\mathcal{H}^{1}dr =Σ{2Rϵ<f<3Rϵ}|A||Σf|𝑑2\displaystyle=\int_{\Sigma\cap\{2R_{\epsilon}<f<3R_{\epsilon}\}}|A||\nabla^{\Sigma}f|d\mathcal{H}^{2}

Note that by Hölder’s inequality and the assumed area bounds

Σ{2Rϵ<f<3Rϵ}|A|𝑑2CRϵ(Σ{2Rϵ<f<3Rϵ}|A|2𝑑2)12\int_{\Sigma\cap\{2R_{\epsilon}<f<3R_{\epsilon}\}}|A|d\mathcal{H}^{2}\leq CR_{\epsilon}\bigg{(}\int_{\Sigma\cap\{2R_{\epsilon}<f<3R_{\epsilon}\}}|A|^{2}d\mathcal{H}^{2}\bigg{)}^{\frac{1}{2}}

There exists some R(2Rϵ,3Rϵ)R\in(2R_{\epsilon},3R_{\epsilon}) such that

(3.3) Σ{f=R}|A|𝑑1\displaystyle\int_{\Sigma\cap\{f=R\}}|A|d\mathcal{H}^{1} 1RϵΣ{2Rϵ<f<3Rϵ}|A|𝑑2\displaystyle\leq\frac{1}{R_{\epsilon}}\int_{\Sigma\cap\{2R_{\epsilon}<f<3R_{\epsilon}\}}|A|d\mathcal{H}^{2}
C(Σ{2Rϵ<f<3Rϵ}|A|2𝑑2)12\displaystyle\leq C\bigg{(}\int_{\Sigma\cap\{2R_{\epsilon}<f<3R_{\epsilon}\}}|A|^{2}d\mathcal{H}^{2}\bigg{)}^{\frac{1}{2}}
Cϵ.\displaystyle\leq C\epsilon.

By embeddedness, we may assume that BR\partial B_{R} intersects Σ\Sigma transversely and γ=BRΣ\gamma=\partial B_{R}\cap\Sigma is a closed curve. From the inequality (3.3), the normal vector of Σ\Sigma has very small oscillation on the curve γ\gamma for sufficiently small ϵ\epsilon. Thus, the curve γ\gamma is contained in a small graphical annular region with gradient bounded by Cϵ1/6C\epsilon^{1/6}. The radius around the graphical neighborhood of each point in γ\gamma is extended to R/8R/8 by applying Lemma 3.3. Approximately, this tells us that γ\gamma is very close (on the order of Cϵ1/6RC\epsilon^{1/6}R) to the cross section of some translate of Π\Pi with the ball BρB_{\rho}.

Ultimately, what we obtain from this argument is that the union of these neighborhoods of γ\gamma contains the graph of a function uu defined on a planar annulus Ω=DR/16+αDR/32+α\Omega=D_{R/16+\alpha}\setminus D_{R/32+\alpha} with width R/32R/32 and α>0\alpha>0. Note that we may choose R>RϵR>R_{\epsilon} large enough that π(ΣBRϵ)DR/32Π\pi(\Sigma\cap B_{R_{\epsilon}})\subset D_{R/32}\subset\Pi, where π\pi is orthogonal projection to Π\Pi.

Assume that 0Π0\in\Pi and let g(x)=|π(x)|g(x)=|\pi(x)| for all x3x\in\mathbb{R}^{3}. The coarea formula gives

R/32+αR/16+αΣ{g=r}|A|2𝑑1𝑑r=graph(u)|A|2|Σg|𝑑2.\displaystyle\int_{R/32+\alpha}^{R/16+\alpha}\int_{\Sigma\cap\{g=r\}}|A|^{2}d\mathcal{H}^{1}dr=\int_{\textrm{graph}(u)}|A|^{2}|\nabla^{\Sigma}g|d\mathcal{H}^{2}.

Thus, there exists ρ(R/32+α,R/16+α)\rho\in(R/32+\alpha,R/16+\alpha) such that

Σ{g=ρ}|A|2𝑑132Rgraph(u)|A|2𝑑2,\int_{\Sigma\cap\{g=\rho\}}|A|^{2}d\mathcal{H}^{1}\leq\frac{32}{R}\int_{\textrm{graph}(u)}|A|^{2}d\mathcal{H}^{2},

where we have used the fact that |Σg|1|\nabla^{\Sigma}g|\leq 1. Furthermore, by construction, ρ17R/16\rho\leq 17R/16. Thus,

(3.4) Σ{g=ρ}|A|2𝑑1\displaystyle\int_{\Sigma\cap\{g=\rho\}}|A|^{2}d\mathcal{H}^{1} Cρgraph(u)|A|2𝑑2\displaystyle\leq\frac{C}{\rho}\int_{\textrm{graph}(u)}|A|^{2}d\mathcal{H}^{2}
(3.5) Cϵ2ρ.\displaystyle\leq\frac{C\epsilon^{2}}{\rho}.

Note that C>0C>0 is an absolute constant that does not depend on R,Rϵ,R,R_{\epsilon}, or ϵ\epsilon. Letting the curve {g=ρ}\{g=\rho\} equal Γ\Gamma completes the proof of the lemma. ∎

In the proof of Lemma 3.5, Γ\Gamma is a graph over DρΠ\partial D_{\rho}\subset\Pi and is contained in the graph of a function uu, defined over an annulus Ω\Omega such that DρΩ\partial D_{\rho}\subset\Omega. We now cut out the interior region in Σ\Sigma bounded by Γ\Gamma, leaving only an annular end with total curvature bounded by ϵ2\epsilon^{2}. We will appeal to the following lemma proved by L. Simon in [Si] to find a candidate disk with small total curvature and boundary equal to Γ\Gamma to replace the excised region.

Lemma 3.6.

[Si, Lemma 2.2] Let Σ23\Sigma^{2}\subset\mathbb{R}^{3} be smooth embedded, ξn\xi\in\mathbb{R}^{n}, LL a plane containing ξ\xi, uC(U)u\in C^{\infty}(U) for some open (L)(L-)neighborhood UU of LBρ(ξ)L\cap\partial B_{\rho}(\xi), and

graph uΣ,|Du|1.\textrm{graph }u\subset\Sigma,\;\;\;\;|Du|\leq 1.

Also, let wC(LB¯ρ(ξ))w\in C^{\infty}(L\cap\bar{B}_{\rho}(\xi)) satisfy

(3.6) {Δ2w=0on LBρ(ξ)w=u,Dw=Duon LBρ(ξ).\left\{\begin{array}[]{lr}\Delta^{2}w=0&\text{on }L\cap B_{\rho}(\xi)\\ w=u,Dw=Du&\text{on }L\cap\partial B_{\rho}(\xi).\end{array}\right.

Then

(3.7) LBρ(ξ)|D2w|2CρΓ|A|2𝑑1,\int_{L\cap B_{\rho}(\xi)}|D^{2}w|^{2}\leq C\rho\int_{\Gamma}|A|^{2}d\mathcal{H}^{1},

where Γ=graph(u|LBρ(ξ))\Gamma=\textrm{graph}(u|L\cap\partial B_{\rho}(\xi)), AA is the second fundamental form of Σ\Sigma, and 1\mathcal{H}^{1} is 1-dimensional Hausdorff measure (i.e. arc-length measure) on Γ\Gamma; CC is a fixed constant independent of Σ\Sigma, ρ\rho.

Note that the solution ww exists and is unique [Si], and that there is the following maximum principle for the biharmonic equation from [PV].

Theorem 3.7 (Maximum Principle for the Biharmonic Equation).

If Δ2u=0\Delta^{2}u=0 in DD, a bounded Lipschitz domain in n\mathbb{R}^{n}, and |Du||Du| is continuous in D¯\bar{D}, then there is a constant CC that depends only on the Lipschitz character of DD and independent of the diameter of DD such that

DuL(D)CDuL(D).||Du||_{L^{\infty}(D)}\leq C||Du||_{L^{\infty}(\partial D)}.

Since DuL(Bρ(ξ))Cϵ1/6||Du||_{L^{\infty}(\partial B_{\rho}(\xi))}\leq C\epsilon^{1/6}, Theorem 3.7 implies that DuL(Bρ(ξ))Cϵ1/6||Du||_{L^{\infty}(B_{\rho}(\xi))}\leq C\epsilon^{1/6}, and thus there exists a uniform constant C>0C>0 not depending on RR or ϵ\epsilon such that

graph w|A|2𝑑2CLBρ|D2w|2.\int_{\textrm{graph }w}|A|^{2}d\mathcal{H}^{2}\leq C\int_{L\cap B_{\rho}}|D^{2}w|^{2}.

Combining this estimate with (3.7) and Lemma 3.5, we obtain the desired bound

graph w|A|2𝑑2Cϵ2,\int_{\textrm{graph }w}|A|^{2}d\mathcal{H}^{2}\leq C\epsilon^{2},

where CC is independent of RR and ϵ\epsilon.

We attach the graph of ww to the end of Σ\Sigma bounded by Γ\Gamma. Simply joining these surfaces along Γ\Gamma results in a C1C^{1} surface SS, which does not satisfy the hypotheses of Theorem 3.2: a C2C^{2} surface is needed. To improve the regularity, we approximate the piecewise surface in the H2H^{2}-Sobolev norm by smooth functions.

Lemma 3.8.

Given R>0R>0 as in Lemma 3.5 and a connected component of ΣRϵ\Sigma\setminus R_{\epsilon}, we can find a smooth topological disk Σ\Sigma^{\prime} such that Σ=Σ\Sigma^{\prime}=\Sigma outside the ball B2RB_{2R}.

Proof.

Take a small tubular neighborhood 𝒯(Γ)\mathcal{T}(\Gamma) of Γ\Gamma in SS. On the outside of Γ\Gamma, 𝒯(Γ)\mathcal{T}(\Gamma) is equal to the graph of uu over the plane Π\Pi. On the inside of Γ\Gamma, 𝒯(Γ)\mathcal{T}(\Gamma) is equal to the graph of ww over Π\Pi. Let vv be the C1C^{1} function over an annular domain VΠV\subset\Pi containing Dρ\partial D_{\rho} such that graph v=𝒯(Γ)\textrm{graph }v=\mathcal{T}(\Gamma).

First, we confirm that vv is in W2,2(V)W^{2,2}(V). Since vC1(V)v\in C^{1}(V), we need only find weak second derivatives. Notice that vv is smooth away from DρV\partial D_{\rho}\subset V. We define DijvD_{ij}v away from the measure zero set Dρ\partial D_{\rho}:

Dijv(x)={Diju:xVB¯ρDijw:xVBρD_{ij}v(x)=\left\{\begin{array}[]{lr}D_{ij}u&:x\in V\setminus\overline{B}_{\rho}\\ D_{ij}w&:x\in V\cap B_{\rho}\end{array}\right.

Now, consider a test function φCc(V)\varphi\in C_{c}^{\infty}(V).

VDijvφ𝑑x\displaystyle\int_{V}D_{ij}v\varphi dx =VB¯ρDijuφ𝑑x+VBρDijwφ𝑑x\displaystyle=\int_{V\setminus\overline{B}_{\rho}}D_{ij}u\varphi dx+\int_{V\cap B_{\rho}}D_{ij}w\varphi dx
=VB¯ρDiuDjφ𝑑xVBρDiwDjφ𝑑x+Bρ(DiuDiw)νj𝑑x\displaystyle=-\int_{V\setminus\overline{B}_{\rho}}D_{i}uD_{j}\varphi dx-\int_{V\cap B_{\rho}}D_{i}wD_{j}\varphi dx+\int_{\partial B_{\rho}}(D_{i}u-D_{i}w)\nu_{j}dx
=VDivDjφ𝑑x.\displaystyle=-\int_{V}D_{i}vD_{j}\varphi dx.

Since vC1v\in C^{1}, this is enough to show that the weak second derivatives we defined are valid and vW2,2v\in W^{2,2}. Note that we used (3.6) to get rid of the last term in the second line.

We approximate vv in W2,2(V)W^{2,2}(V) by smooth functions. Let ψ\psi be a cutoff function that is uniformly equal to 1 in a neighborhood of Dρ\partial D_{\rho} and compactly supported in VV. Let (ψv)h(\psi v)_{h} be the regularization of ψv\psi v, with h>0h>0 chosen so that (ψv)h(\psi v)_{h} is also compactly supported in VV and

(ψv)hψvW2,2(V)<ϵ2.||(\psi v)_{h}-\psi v||_{W^{2,2}(V)}<\epsilon^{2}.

Then the function

v~=(1ψ)v+(ψv)h\tilde{v}=(1-\psi)v+(\psi v)_{h}

is equal to vv in a uniform neighborhood of the boundary V\partial V and has D2v~L2(V)<Cϵ2||D^{2}\tilde{v}||_{L^{2}(V)}<C\epsilon^{2}. Since |Dv~||D\tilde{v}| is uniformly bounded, we have

graph v~|A|2CV|D2v~|2<Cϵ2.\int_{\textrm{graph }\tilde{v}}|A|^{2}\leq C\int_{V}|D^{2}\tilde{v}|^{2}<C\epsilon^{2}.

Since graph v~\textrm{graph }\tilde{v} matches Σ\Sigma in an open neighborhood of V\partial V, the surface obtained by joining them, Σ\Sigma^{\prime}, is smooth and equal to Σ\Sigma outside B2RB_{2R}. ∎

Choose ϵ>0\epsilon>0 and let Q>0Q>0 be a sufficiently large radius. We now prove the following lemma for Σ\Sigma^{\prime}. We assume that 0Σ0\in\Sigma^{\prime}.

Lemma 3.9.

Away from the ball BCϵ1/2Q(0)B_{C\epsilon^{1/2}Q}(0), the surface ΣBQ/8(0)\Sigma^{\prime}\cap B_{Q/8}(0) is the graph of a function uu with gradient bounded by Cϵ1/6C\epsilon^{1/6} over a plane Π\Pi parallel to VV.

Proof.

We apply Simon’s Lemma, Theorem 3.2, to Σ\Sigma^{\prime}. Lemma 3.3 ensures that BSΣ\partial B_{S}\cap\Sigma (where S[Q/4,Q/2]S\in[Q/4,Q/2]) is an embedded, closed graphical curve bounding a connected topological disk. Away from pimples, Simon’s Lemma tells us that ΣBS\Sigma^{\prime}\cap B_{S} is the graph of a function uu defined on a domain in Π\Pi with gradient bounded by Cϵ1/6C\epsilon^{1/6}. Let π:3Π\pi:\mathbb{R}^{3}\rightarrow\Pi be orthogonal projection and let R>0R>0 be as in Lemma 3.8. Let DrΠD_{r}\subset\Pi be a disk such that π(ΣB2R)Dr\pi(\Sigma^{\prime}\cap B_{2R})\subset D_{r}. By the diameter bounds (3.1), there exists s1[Q/4Cϵ1/2Q,Q/4]s_{1}\in[Q/4-C\epsilon^{1/2}Q,Q/4] and s2[r,r+Cϵ1/2Q]s_{2}\in[r,r+C\epsilon^{1/2}Q] such that π1(Bs1)\pi^{-1}(\partial B_{s_{1}}) and π1(Bs2)\pi^{-1}(\partial B_{s_{2}}) do not intersect the pimples and intersect ΣBS\Sigma^{\prime}\cap B_{S} transversely. We apply the Schoen/Alexandrov reflection method, Theorem 3.4, and the maximum principle, Lemma 2.3, as we did in Lemma 3.3 to show that the pimples lying between these cylinders are indeed graphical over Π\Pi with gradient bounded by Cϵ1/6C\epsilon^{1/6}. This completes the proof. ∎

We can now prove the main result of this section, Lemma 3.1.

Proof of Lemma 3.1.

Let us choose a sequence of annuli BmB1/mB_{m}\setminus B_{1/m}, where mm\in\mathbb{N}, a sequence decreasing to zero ϵj0\epsilon_{j}\rightarrow 0, and a family of rescalings λiΣ\lambda_{i}\Sigma^{\prime}, with λi0\lambda_{i}\rightarrow 0. Let Σj\Sigma_{j}^{\prime} be obtained for ϵ=ϵj\epsilon=\epsilon_{j} by cutting out a high curvature region and pasting in a disk as above. Note that in the inequality

Σj|A|2Cϵj2,\int_{\Sigma^{\prime}_{j}}|A|^{2}\leq C\epsilon_{j}^{2},

the constant CC depends only on β\beta, and is independent of ϵj\epsilon_{j}. Observe also that outside of some ball Σj=Σk\Sigma_{j}^{\prime}=\Sigma_{k}^{\prime}, for any j,kj,k\in\mathbb{N}.

Fix mm and jj. By Lemma 3.9, for sufficiently large values of ii, λiΣ\lambda_{i}\Sigma^{\prime} can be written as a graph of a function over a plane Πi\Pi_{i} parallel to VV with C1C^{1}-norm bounded by Cϵj1/6C\epsilon_{j}^{1/6} in BmB1/mB_{m}\setminus B_{1/m}. Un-fixing jj, we see that by the compactness of the collection of planes {ΠGr2,3:VΠ}\{\Pi\in\textrm{Gr}_{2,3}:V\in\Pi\}, there exists a subsequence λik0\lambda_{i_{k}}\rightarrow 0 such that λikΣ\lambda_{i_{k}}\Sigma^{\prime} converges to a fixed plane Π\Pi inside BmB1/mB_{m}\setminus B_{1/m} in the C1C^{1}-norm.

Now, unfixing mm, we take a further diagonal subsequence such that the rescalings of our designated end, λikΣ\lambda_{i_{k}}\Sigma, converge to Π{0}\Pi\setminus\{0\} in the Cloc1C^{1}_{loc} topology. This concludes the proof of Lemma 3.1. ∎

Note that this blow-down limit is subsequential, and thus not necessarily unique. The problem arises from the fact that in Lemma 3.9, graphicality cannot be extended to a neighborhood of the origin with diameter proportional to the graphical radius. That is, the surface may be twisting and turning inside the ball BCϵ1/6QB_{C\epsilon^{1/6}Q} (in the notation of Lemma 3.9.) Thus, we must argue further in §4 to show that this asymptotic tangent plane is indeed unique. However, in the case that the translator Σ\Sigma has only one end, this possibility does not present a serious difficulty: Theorem 1.1 is a quick corollary of Lemma 3.1.

Proof of Theorem 1.1.

Let {λiΣ}\{\lambda_{i}\Sigma\} be a blow-down sequence converging to Π\Pi. Fix mm\in\mathbb{N}. Consider the sequence of annuli ΣBλi1mBλi1m1\Sigma\cap B_{\lambda_{i}^{-1}m}\setminus B_{\lambda_{i}^{-1}m^{-1}}. Each of these can be written as the graph of a function uiu_{i} over PP with gradient |Dui|<Cϵi1/6|Du_{i}|<C\epsilon_{i}^{1/6}, where ϵi0\epsilon_{i}\rightarrow 0. Consider the curve ΣBλi1m\Sigma\cap\partial B_{\lambda_{i}^{-1}m} and its projection to an embedded Jordan curve in Π\Pi, which we denote σi\sigma_{i}. We take Ω\Omega to be the open set in Π\Pi bounded by σi\sigma_{i} and apply the method of moving planes, Theorem 3.4 to determine that ΣBλi1,\Sigma\cap B_{\lambda_{i}^{-1},} is a graph over PP. Then we apply the weak maximum principle, Theorem 2.3, to determine that this graph has gradient |Du|<Cϵi1/6|Du|<C\epsilon_{i}^{1/6}. As m,im,i\rightarrow\infty, the sequence ΣBλi1m\Sigma\cap B_{\lambda_{i}^{-1}m} must be a sequence of graphical disks over Π\Pi with increasingly small gradient. Thus, Σ\Sigma must be a plane. ∎

4. Uniqueness of the Tangent Plane

In this section, we establish the uniqueness of the asymptotic tangent plane and the graphicality of the ends of Σ\Sigma. To prove this, we show that given two annuli in a convergent blow-down sequence, the “interstitial space” between them must in fact be graphical with small gradient over the limit plane.

Proposition 4.1.

Given an end of the translator Σ\Sigma, the Cloc1C^{1}_{loc}-limit in 3{0}\mathbb{R}^{3}\setminus\{0\} of any blow-down sequence {λiΣ}\{\lambda_{i}\Sigma\}, λi0\lambda_{i}\rightarrow 0 is a unique plane Π\Pi parallel to VV. Consequently, the ends of Σ\Sigma can be written as graphs over Π\Pi outside of some ball BR03B_{R_{0}}\subset\mathbb{R}^{3}.

Proof.

Let ϵ>0\epsilon>0 and Σ\Sigma^{\prime} be the modified surface from the previous section: a complete embedded topological disk, equal to the translator Σ\Sigma outside a ball BRB_{R} and with total curvature less than ϵ2\epsilon^{2}. Consider a blow-down sequence {λiΣ}\{\lambda_{i}\Sigma^{\prime}\}, with λi0\lambda_{i}\rightarrow 0 that converges to a plane Π\Pi, and take λj>λj+1\lambda_{j}>\lambda_{j+1}. Let R>0R>0 be a large fixed radius, and let jj be large enough that ΣBλj1RBλj1R1\Sigma^{\prime}\cap B_{\lambda_{j}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}} and ΣBλj+11RBλj+11R1\Sigma^{\prime}\cap B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j+1}^{-1}R^{-1}} can be written as graphs of functions uju_{j} and uj+1u_{j+1} over annuli in Π\Pi with gradient bounded by Cϵ1/6C\epsilon^{1/6}. Let us also define the curves γj\gamma_{j} and γj+1\gamma_{j+1} which will denote the inner boundary ΣBλj1R1\Sigma^{\prime}\cap\partial B_{\lambda_{j}^{-1}R^{-1}} and the outer boundary ΣBλj+11R\Sigma^{\prime}\cap\partial B_{\lambda^{-1}_{j+1}R} respectively.

If we project the space curves γj\gamma_{j} and γj+1\gamma_{j+1} to Π\Pi via the orthogonal projection map π:3Π\pi:\mathbb{R}^{3}\rightarrow\Pi, we obtain simple closed planar curves σj=π(γj)\sigma_{j}=\pi(\gamma_{j}) and σj+1=π(γj+1)\sigma_{j+1}=\pi(\gamma_{j+1}). Equivalently, uj(σj)=γju_{j}(\sigma_{j})=\gamma_{j} and uj+1(σj+1)=γj+1u_{j+1}(\sigma_{j+1})=\gamma_{j+1}. Thus, σj\sigma_{j} and σj+1\sigma_{j+1} bound a large topological annulus Ω\Omega in Π\Pi. Let 𝐧\mathbf{n} be the normal vector to Π\Pi. Consider the solid cylinder Ω×𝐧\Omega\times\mathbf{n}\mathbb{R}. Suppose that (ΣBλj+11RBλj1R1)(σj×𝐧)=γj(\Sigma^{\prime}\cap B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}})\cap(\sigma_{j}\times\mathbf{n}\mathbb{R})=\gamma_{j}. Then, Σ(Bλj+11RBλj1R1)\Sigma^{\prime}\cap(B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}}) is contained inside Ω×𝐧\Omega\times\mathbf{n}\mathbb{R}, and since the boundary curves γj\gamma_{j} and γj+1\gamma_{j+1} are graphs, we can apply the moving planes method, Theorem 3.4, to find that Σ(Bλj+11RBλj1R1)\Sigma^{\prime}\cap(B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}}) is a graph over Ω\Omega.

Suppose that (ΣBλj+11R)(σj×𝐧)γj(\Sigma^{\prime}\cap B_{\lambda_{j+1}^{-1}R})\cap(\sigma_{j}\times\mathbf{n}\mathbb{R})\supsetneq\gamma_{j}, i.e. the surface turns back and intersects the cylinder of small radius. Consider the domain 𝒟=int(σj)\mathcal{D}=\mathrm{int}(\sigma_{j}) in the plane Π\Pi such that π(ΣBR)𝒟\pi(\Sigma^{\prime}\cap B_{R})\subset\mathcal{D} and D𝒟D\mathcal{D} is bounded by the Jordan curve σj\sigma_{j}. We separate this situation into two cases. These cases are illustrated in Figure A and Figure A in Appendix A.

Case 1: Λ:=(ΣBλj+11RBλj1R1)(𝒟×𝐧)\Lambda:=(\Sigma^{\prime}\cap B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}})\cap(\mathcal{D}\times\mathbf{n}\mathbb{R}) is non-graphical or a multigraph over DRD_{R} in either of the connected components of (𝒟×𝐧)Bλj1R1(\mathcal{D}\times\mathbf{n}\mathbb{R})\setminus B_{\lambda_{j}^{-1}R^{-1}}. Without loss of generality, let us assume this occurs in the upper half space Π×0\Pi\times\mathbb{R}_{\geq 0}. In this case, we may use a modified method of moving planes. Consider the surface ΣBλj+11R\Sigma^{\prime}\cap B_{\lambda_{j+1}^{-1}R} (a topological disk) contained in the cylinder int(σj+1)×𝐧\textrm{int}(\sigma_{j+1})\times\mathbf{n}\mathbb{R} with boundary equal to γj+1\gamma_{j+1}. If (x1,x2)(x_{1},x_{2}) are coordinates in Π\Pi and x3x_{3} indicates height in 𝐧\mathbf{n}\mathbb{R}, we take the plane {x3=t}\{x_{3}=t\} parallel to Π\Pi and the associated surfaces Σ~t+\tilde{\Sigma^{\prime}}_{t}^{+} and Σt{\Sigma^{\prime}}_{t}^{-} as in the proof of Theorem 3.4.

Let hh be the minimum height of Λ\Lambda in the upper half space x3>0x_{3}>0. Note that

h>λj1R12>R,h>\frac{\lambda_{j}^{-1}R^{-1}}{2}>R,

where we may assume without loss of generality that λj1\lambda_{j}^{-1} is sufficiently large that the second inequality holds. The first inequality follows from repeated applications of Lemma 3.9, which tells us that once Σ\Sigma^{\prime} leaves the ball Bλj1R1B_{\lambda_{j}^{-1}R^{-1}}, it will never re-enter.

If a first point of interior tangency occurs, it must occur for t>ht>h. This means that it occurs outside of BRB_{R} and around this point of contact Σ~t+\tilde{\Sigma^{\prime}}_{t}^{+} and Σt{\Sigma^{\prime}}_{t}^{-} must satisfy the translator equation. Thus, the strong maximum principle applies. This is a contradiction, so the first point of contact must occur on the boundary curve γj+1\gamma_{j+1}. However, γj+1\gamma_{j+1} is a graph over Π\Pi, so there is no point of contact outside {x3=t}\{x_{3}=t\} on the boundary. Thus, the orthogonal projection π(Λ)\pi(\Lambda) to Π\Pi must be one-to-one, so Case 1 cannot occur.

Case 2: Λ:=(ΣBλj+11RBλj1R1)(𝒟×𝐧)\Lambda:=(\Sigma^{\prime}\cap B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}})\cap(\mathcal{D}\times\mathbf{n}\mathbb{R}) is a single-valued graph over UU. Note that the plane Π\Pi is spanned by two orthogonal vectors x1=V\partial_{x_{1}}=V and x2\partial_{x_{2}} and is normal to the coordinate vector x3\partial_{x_{3}}. Let us take the cross section of ΣBλj+11RBλj1R1\Sigma^{\prime}\cap B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}} with respect to the plane spanned by x2\partial_{x_{2}} and x3\partial_{x_{3}} passing through the origin. Since the normal part of VV has norm |V|=O(ϱ1/2)|V^{\perp}|=O(\varrho^{-1/2}) by Corollary 2.6, the intersection with this plane is transverse for sufficiently large values of λj1\lambda_{j}^{-1}. This intersection can be seen to have two connected components which can be distinguished by whether they intersect the inner boundary curve on the left or the right of the line (x2,x3){0}×(x_{2},x_{3})\in\{0\}\times\mathbb{R}. Without loss of generality, take the component that intersects γj\gamma_{j} on the left. This curve is an embedded curve with boundary in γj\gamma_{j} and γj+1\gamma_{j+1}, which we henceforth denote by η\eta.

We now describe some properties of the curve η\eta. Let pjp_{j} and pj+1p_{j+1} be the unique intersection points of η\eta with γj\gamma_{j} and γj+1\gamma_{j+1}: these are the endpoints of η\eta. Let \ell be the length of η\eta. We parametrize η\eta by arclength such that

η(0)=pj,η()=pj+1, and η(0),x2=1+Cϵ1/6.\eta(0)=p_{j},\;\eta(\ell)=p_{j+1},\text{ and }\langle\eta^{\prime}(0),\partial_{x_{2}}\rangle=-1+C\epsilon^{1/6}.

By assumption, η\eta crosses the line {0}×\{0\}\times\mathbb{R} exactly once. We wish to use this condition to determine whether η(),x2\langle\eta^{\prime}(\ell),\partial_{x_{2}}\rangle is close to 1 or -1. First note that

ηBλj+11RBλj+11R1\eta\cap B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j+1}^{-1}R^{-1}}

is much longer than

ηBλj1R1Bλj1R1.\eta\cap B_{\lambda_{j}^{-1}R^{-1}}\setminus B_{\lambda_{j}^{-1}R^{-1}}.

We then consider the two cases (1) pj+1p_{j+1} is on the left of vertical axis {0}×\{0\}\times\mathbb{R} and (2) pj+1p_{j+1} is on the right of this axis. In case (1), η\eta must cross {0}×\{0\}\times\mathbb{R} an even number of times, which cannot happen. Thus, (2) must hold. If η(s),x21\langle\eta^{\prime}(s),\partial_{x_{2}}\rangle\approx-1 on ηBλj+11RBλj+11R1\eta\cap B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j+1}^{-1}R^{-1}}, integration tells us that pj+1p_{j+1} must be on the left of {0}×\{0\}\times\mathbb{R}, which is impossible. Thus, η(),x21\langle\eta^{\prime}(\ell),\partial_{x_{2}}\rangle\approx 1 in all situations. In particular, η(),η(0)1\langle\eta^{\prime}(\ell),\eta^{\prime}(0)\rangle\approx-1.

Let ΣBλj1RBλj1R1\Sigma^{\prime}\cap B_{\lambda_{j}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}} and ΣBλj+11RBλj+11R1\Sigma^{\prime}\cap B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j+1}^{-1}R^{-1}} be parametrized by

Xj(x1,x2)=(x1,x2,uj(x1,x2))X_{j}(x_{1},x_{2})=(x_{1},x_{2},u_{j}(x_{1},x_{2}))

and

Xj+1(x1,x2)=(x1,x2,uj+1(x1,x2))X_{j+1}(x_{1},x_{2})=(x_{1},x_{2},u_{j+1}(x_{1},x_{2}))

respectively. The two surfaces are oriented by the respective 2-vectors

x1Xjx2Xj and x1Xj+1x2Xj+1.\partial_{x_{1}}X_{j}\wedge\partial_{x_{2}}X_{j}\text{ and }\partial_{x_{1}}X_{j+1}\wedge\partial_{x_{2}}X_{j+1}.

Since these two surfaces are graphical annuli with very small gradient, their orientation 2-vectors are close to

Vη(0) and Vη(),V\wedge\eta^{\prime}(0)\text{ and }V\wedge\eta^{\prime}(\ell),

respectively. Note that at all points η(s)\eta(s), VV is “almost” a tangent vector to Σ\Sigma perpendicular to η(s)\eta^{\prime}(s), by Corollary 2.6. Since η(),η(0)1\langle\eta^{\prime}(\ell),\eta^{\prime}(0)\rangle\approx-1, we can deduce that

(x1Xjx2Xj)(p),(x1Xj+1x2Xj+1)(q)<0,(\langle\partial_{x_{1}}X_{j}\wedge\partial_{x_{2}}X_{j})(p),(\partial_{x_{1}}X_{j+1}\wedge\partial_{x_{2}}X_{j+1})(q)\rangle<0,

for any pΣBλj1RBλj1R1p\in\Sigma^{\prime}\cap B_{\lambda_{j}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}} and qΣBλj+11RBλj+11R1q\in\Sigma^{\prime}\cap B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j+1}^{-1}R^{-1}}. Equivalently, the normal vectors to the two annuli determine opposite orientations with respect to a fixed frame on 3\mathbb{R}^{3}. The crucial consequence of this fact is that the geodesic curvatures along curves “nearby” γj\gamma_{j} and γj+1\gamma_{j+1} will have the same sign and be approximately equal to 2π2\pi. We demonstrate this precisely in the remainder of the proof.

Next, we show that this change in orientation causes the total curvature to be large. By the same reasoning as in (3.4), we can find curves Γj\Gamma_{j} and Γj+1\Gamma_{j+1} in Σ(Bλj1RBλj1R1)\Sigma^{\prime}\cap(B_{\lambda_{j}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}}) and Σ(Bλj+11RBλj+11R1)\Sigma^{\prime}\cap(B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j+1}^{-1}R^{-1}}) respectively such that

π(Γj)=Dρj,π(Γj+1)=Dρj+1 for some ρj,ρj+1>0,\pi(\Gamma_{j})=\partial D_{\rho_{j}},\;\pi(\Gamma_{j+1})=\partial D_{\rho_{j+1}}\text{ for some }\rho_{j},\rho_{j+1}>0,

and such that

Γj|A|2𝑑12ϵ2ρj and Γj+1|A|2𝑑12ϵ2ρj+1.\int_{\Gamma_{j}}|A|^{2}d\mathcal{H}^{1}\leq\frac{2\epsilon^{2}}{\rho_{j}}\text{ and }\int_{\Gamma_{j+1}}|A|^{2}d\mathcal{H}^{1}\leq\frac{2\epsilon^{2}}{\rho_{j+1}}.

We consider the annular subset of Σ(Bλj+11RBλj1R1)\Sigma^{\prime}\cap(B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}}) bounded by the curves Γj\Gamma_{j} and Γj+1\Gamma_{j+1}, which we call S(Γj,Γj+1)S(\Gamma_{j},\Gamma_{j+1}). We apply the Gauss-Bonnet theorem:

S(Γj,Γj+1)K=2π(2M12gM0)S(Γj,Γj+1)κg,\int_{S(\Gamma_{j},\Gamma_{j+1})}K=2\pi(2M_{1}-2g-M_{0})-\int_{\partial S(\Gamma_{j},\Gamma_{j+1})}\kappa_{g},

where κg\kappa_{g} is the geodesic curvature, M1M_{1} is equal to the number of connected components, gg is the genus, and M0M_{0} is equal to the number of boundary components. Substituting M1=1,M0=2M_{1}=1,M_{0}=2 and g=0g=0, we obtain

Σ(Bλj+11RBλj1R1)K=ΓjκgΓj+1κg.\int_{\Sigma^{\prime}\cap(B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}})}K=-\int_{\Gamma_{j}}\kappa_{g}-\int_{\Gamma_{j+1}}\kappa_{g}.
Claim 4.2.
Γjκg,Γj+1κg2π\int_{\Gamma_{j}}\kappa_{g}\;\;,\;\;\int_{\Gamma_{j+1}}\kappa_{g}\approx 2\pi
Proof.

We apply similar reasoning to a proof on p.294-5 of [Si]. We parametrize Γj\Gamma_{j} and Γj+1\Gamma_{j+1} as follows.

(4.1) Γj(θ)=ρjeiθ+uj(ρjeiθ)𝐧,\Gamma_{j}(\theta)=\rho_{j}e^{i\theta}+u_{j}(\rho_{j}e^{i\theta})\mathbf{n},

and

(4.2) Γj+1(θ)=ρj+1eiθ+uj(ρj+1eiθ)𝐧,\Gamma_{j+1}(\theta)=\rho_{j+1}e^{i\theta}+u_{j}(\rho_{j+1}e^{i\theta})\mathbf{n},

where θ[0,2π)\theta\in[0,2\pi) and Π\Pi is identified with \mathbb{C}. We calculate the integral

02π|2u(ieiθ,ieiθ)||Γ(θ)|𝑑θ\displaystyle\int_{0}^{2\pi}|\nabla^{2}u(ie^{i\theta},ie^{i\theta})||\Gamma^{\prime}(\theta)|d\theta =Γ|2u(ieiθ,ieiθ)|𝑑1\displaystyle=\int_{\Gamma}|\nabla^{2}u(ie^{i\theta},ie^{i\theta})|d\mathcal{H}^{1}
(length(Γ))1/2(Γ|2u|2𝑑1)1/2\displaystyle\leq(\textrm{length}(\Gamma))^{1/2}\bigg{(}\int_{\Gamma}|\nabla^{2}u|^{2}d\mathcal{H}^{1}\bigg{)}^{1/2}
ρ1/2(2ϵ2ρ)1/22ϵ.\displaystyle\leq\rho^{1/2}\bigg{(}\frac{2\epsilon^{2}}{\rho}\bigg{)}^{1/2}\leq 2\epsilon.

Differentiating (4.1) and (4.2), and combining with the above calculation, we obtain that the total curvature vector κ\vec{\kappa} can be writted in the following form:

κ=ρ1ν+E,\vec{\kappa}=\rho^{-1}\nu+E,

where EE is a vector field on Γ\Gamma such that Γ|E|<Cϵ1/6\int_{\Gamma}|E|<C\epsilon^{1/6} and ν\nu is the inward pointing unit normal to Γ\Gamma with respect to the surface S(Γj,Γj+1)S(\Gamma_{j},\Gamma_{j+1}). Since

(x1Xjx2Xj)(p),(x1Xj+1x2Xj+1)(q)<0,(\langle\partial_{x_{1}}X_{j}\wedge\partial_{x_{2}}X_{j})(p),(\partial_{x_{1}}X_{j+1}\wedge\partial_{x_{2}}X_{j+1})(q)\rangle<0,

for any pΣBλj1RBλj1R1p\in\Sigma^{\prime}\cap B_{\lambda_{j}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}} and qΣBλj+11RBλj+11R1q\in\Sigma^{\prime}\cap B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j+1}^{-1}R^{-1}}, the geodesic curvatures (κj)g(\kappa_{j})_{g} and (κj+1)g(\kappa_{j+1})_{g} have the same sign. We calculate

Γκg\displaystyle\int_{\Gamma}\kappa_{g} =Γκν=02π(ρ+Eν)|ρdθ\displaystyle=\int_{\Gamma}\vec{\kappa}\cdot\nu=\int_{0}^{2\pi}(\rho+E\cdot\nu)|\rho d\theta
=02πρρ1𝑑θ+ΓEν\displaystyle=\int_{0}^{2\pi}\rho\rho^{-1}d\theta+\int_{\Gamma}E\cdot\nu
|Γκg2π|Cϵ1/6.\implies\bigg{|}\int_{\Gamma}\kappa_{g}-2\pi\bigg{|}\leq C\epsilon^{1/6}.

This concludes the proof of the claim. ∎

Claim 4.2 and the Gauss-Bonnet Theorem imply that

32Σ(Bλj+11RBλj1R1)|A|24πδ\frac{3}{2}\int_{\Sigma^{\prime}\cap(B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}})}|A|^{2}\geq 4\pi-\delta

for some small δ>0\delta>0. Given the assumption of small total curvature, this is contradiction. Thus, Case 2 does not occur and Σ(Bλj+11RBλj1R1)\Sigma^{\prime}\cap(B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}}) must be the graph of a function wjw_{j} over ΩΠ\Omega\subset\Pi. Since |Dwj|<Cϵ1/6|Dw_{j}|<C\epsilon^{1/6} on γj\gamma_{j} and γj+1\gamma_{j+1}, the weak maximum principle implies that |Dwj|<Cϵ1/6|Dw_{j}|<C\epsilon^{1/6} on Ω\Omega.

Now, consider two blow down sequences {μiΣ}\{\mu_{i}\Sigma\} and {λjΣ}\{\lambda_{j}\Sigma\}, with λj,μi0\lambda_{j},\mu_{i}\rightarrow 0 and {λjΣ}\{\lambda_{j}\Sigma\} has blow-down limit equal to the plane Π\Pi. For any sufficiently large μi\mu_{i}, up to a subsequence of {λj}\{\lambda_{j}\}, we can find λj>μi>λj+1\lambda_{j}>\mu_{i}>\lambda_{j+1}. The annular set Σ(Bμi1RBμi1R1)\Sigma^{\prime}\cap(B_{\mu_{i}^{-1}R}\setminus B_{\mu_{i}^{-1}R^{-1}}) is contained in the larger annulus Σ(Bλj+11RBλj1R1)\Sigma^{\prime}\cap(B_{\lambda_{j+1}^{-1}R}\setminus B_{\lambda_{j}^{-1}R^{-1}}), which must be a graph over Π\Pi with gradient bounded by Cϵj1/6C\epsilon_{j}^{1/6}. Now that Σ(Bμi1RBμi1R1)\Sigma^{\prime}\cap(B_{\mu_{i}^{-1}R}\setminus B_{\mu_{i}^{-1}R^{-1}}) can be written as a graph with gradient bounded by Cϵj1/6C\epsilon_{j}^{1/6} over Π\Pi, we see that in fact the blow-down limit of {μiΣ}\{\mu_{i}\Sigma\} must be the plane Π\Pi as well.

In fact, this tells us that the entire end can be written as the graph of of a function uu over Π\Pi with |u|=o(ϱ)|u|=o(\varrho) and |Du|=o(1)|Du|=o(1). That each end is a graph over the same plane Π\Pi is a consequence of embeddedness. ∎

5. Strong Asymptotics of the Ends

In this section, we obtain upper bounds for the decay of an end of Σ\Sigma to its asymptotic plane. We begin by using a barrier argument to obtain unidirectional exponential decay of the ends. This is stated precisely in the following proposition:

Proposition 5.1.

Let Σ0\Sigma_{0} be an end of a translator Σ3\Sigma\subset\mathbb{R}^{3} that translates with unit speed in the x1x_{1}-direction and has finite total curvature. There exists a function u:2BRu:\mathbb{R}^{2}\setminus B_{R}\rightarrow\mathbb{R} whose graph represents Σ0\Sigma_{0} and whose magnitude decays uniformly at a rate O(eαx1)O(e^{-\alpha x_{1}}) for any positive α<1/2\alpha<1/2 as x1x_{1}\rightarrow\infty.

Proof.

To prove this statement, we apply a barrier argument to the x1x_{1}-derivative of uu, D1uD_{1}u. We set v:=D1uv:=D_{1}u. By the proof of Lemma 2.3, v=D1uv=D_{1}u satisfies a differential equation of the form Lv=0Lv=0, where the differential operator L=ajkDjk+bjDjL=a_{jk}D_{jk}+b_{j}D_{j} has coefficients given by

ajk=(1+|Du|2)δjkDjuDkua_{jk}=(1+|Du|^{2})\delta_{jk}-D_{j}uD_{k}u
bj=(1+|Du|2)δ1j2DkuDjku(Δu+D1u3DkuDluDklu1+|Du|2)Dju.b_{j}=(1+|Du|^{2})\delta_{1j}-2D_{k}uD_{jk}u-\bigg{(}\Delta u+D_{1}u-\frac{3D_{k}uD_{l}uD_{kl}u}{1+|Du|^{2}}\bigg{)}D_{j}u.

We now calculate L(v2)L(v^{2}).

L(v2)\displaystyle L(v^{2}) =aijDij2(v2)+bjDj(v2)\displaystyle=a_{ij}D^{2}_{ij}(v^{2})+b_{j}D_{j}(v^{2})
=aij(Di(2vDjv))+2vbjDjv\displaystyle=a_{ij}(D_{i}(2vD_{j}v))+2vb_{j}D_{j}v
=2aij(DivDjv)+2vaijDij2v+2vbjDjv\displaystyle=2a_{ij}(D_{i}vD_{j}v)+2va_{ij}D_{ij}^{2}v+2vb_{j}D_{j}v
=2aij(DivDjv)\displaystyle=2a_{ij}(D_{i}vD_{j}v)
=2(1+|Du|2)|Dv|2+2DiuDju(DivDjv)\displaystyle=2(1+|Du|^{2})|Dv|^{2}+2D_{i}uD_{j}u(D_{i}vD_{j}v)
=2(1+|Du|2)|Dv|2+2(Du,Dv)2.\displaystyle=2(1+|Du|^{2})|Dv|^{2}+2(\langle Du,Dv\rangle)^{2}.

Thus,

L(v2)\displaystyle L(v^{2}) 2(1+|Du|2)|Dv|2\displaystyle\geq 2(1+|Du|^{2})|Dv|^{2}
0.\displaystyle\geq 0.

We now choose our barrier function, ex12e^{-\frac{x_{1}}{2}}. We calculate L(ex12)L\big{(}e^{-\frac{x_{1}}{2}}\big{)}:

L(ex12)\displaystyle L\big{(}e^{-\frac{x_{1}}{2}}\big{)} =a11D112(ex12)+b1D1(ex12).\displaystyle=a_{11}D_{11}^{2}\big{(}e^{-\frac{x_{1}}{2}}\big{)}+b_{1}D_{1}\big{(}e^{-\frac{x_{1}}{2}}\big{)}.

We calculate the coefficients a11a_{11} and b1b_{1} here:

a11=(1+|Du|2)|D1u|2=1+|D2u|2a_{11}=(1+|Du|^{2})-|D_{1}u|^{2}=1+|D_{2}u|^{2}
b1=(1+|Du|2)|D1u|22DkuD1ku(Δu3D2u(Du,Du)1+|Du|2)D1ub_{1}=(1+|Du|^{2})-|D_{1}u|^{2}-2D_{k}uD_{1k}u-\bigg{(}\Delta u-\frac{3D^{2}u(Du,Du)}{1+|Du|^{2}}\bigg{)}D_{1}u

We now rely on the fact that DuCk=o(1)||Du||_{C^{k}}=o(1) as ρ\rho\rightarrow\infty in order to estimate the lower order terms for large values of x1x_{1}.

L(ex12)\displaystyle L\big{(}e^{-\frac{x_{1}}{2}}\big{)} =a11ex124b1ex122\displaystyle=a_{11}\frac{e^{-\frac{x_{1}}{2}}}{4}-b_{1}\frac{e^{-\frac{x_{1}}{2}}}{2}
=(1+|D2u|2)(ex124ex122)\displaystyle=(1+|D_{2}u|^{2})\bigg{(}\frac{e^{-\frac{x_{1}}{2}}}{4}-\frac{e^{-\frac{x_{1}}{2}}}{2}\bigg{)}
+ex122(2DkuD1ku+D1uΔuD1u3D2u(Du,Du)1+|Du|2)\displaystyle\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;+\frac{e^{-\frac{x_{1}}{2}}}{2}\bigg{(}2D_{k}uD_{1k}u+D_{1}u\Delta u-D_{1}u\frac{3D^{2}u(Du,Du)}{1+|Du|^{2}}\bigg{)}
=ex124+ex12o(1)\displaystyle=-\frac{e^{-\frac{x_{1}}{2}}}{4}+e^{-\frac{x_{1}}{2}}o(1)
ex128<0,\displaystyle\leq-\frac{e^{-\frac{x_{1}}{2}}}{8}<0,

for x1x_{1} sufficiently large. Let M>0M>0 be such that the previous inequality holds for x1Mx_{1}\geq M. Let CC be such that CeM2>D1uCe^{\frac{-M}{2}}>D_{1}u on the line {x1=M}\{x_{1}=M\}. Note that such a CC must exist, since |Du|=o(1)|Du|=o(1) as the radius approaches infinity, and is in particular uniformly bounded. Now for some ϵ>0\epsilon>0, define the function ρϵ=Cex12+ϵ\rho_{\epsilon}=Ce^{-\frac{x_{1}}{2}}+\epsilon. Since there are no order zero terms, Lρϵ<0L\rho_{\epsilon}<0 for x1Mx_{1}\geq M. Thus,

L(ρϵv2)<0.L(\rho_{\epsilon}-v^{2})<0.

Consider the semicircular domains 𝒟S2\mathcal{D}_{S}\subset\mathbb{R}^{2} parametrized by S>0S>0 defined by the intersection of the balls DS(p)D_{S}(p), of radius SS and centered around p=(x1,x2)=(M,0)p=(x_{1},x_{2})=(M,0), and the upper half-plane {x1>M}\{x_{1}>M\}. Since |Du|=o(1)|Du|=o(1) as the radius approaches infinity, for each ϵ>0\epsilon>0, there exists some S>0S>0 such that ρϵ>v2\rho_{\epsilon}>v^{2} on the boundary 𝒟S\partial\mathcal{D}_{S}. The weak maximum principle [GT, Theorem 3.1] tells us that ρϵ>v2\rho_{\epsilon}>v^{2} on 𝒟S\mathcal{D}_{S}. Sending SS\rightarrow\infty and then ϵ0\epsilon\rightarrow 0, we see that v2<Cϵx12v^{2}<C\epsilon^{-\frac{x_{1}}{2}} in the upper half-plane.

This in turn implies that |D1u|Cex14|D_{1}u|\leq Ce^{-\frac{x_{1}}{4}} on {x1M}\{x_{1}\geq M\}. Integrating along rays in the direction of motion, we can conclude that uu approaches a limit along each ray parallel to the direction of motion. Fix real numbers yy and y^\hat{y} and suppose that

limx1u(x1,y)=K,limx1u(x1,y^)=K^.\lim_{x_{1}\rightarrow\infty}u(x_{1},y)=K,\;\;\lim_{x_{1}\rightarrow\infty}u(x_{1},\hat{y})=\hat{K}.

Since |Du|=o(1)|Du|=o(1), there exists a radius RR such that on 2BR\mathbb{R}^{2}\setminus B_{R},

|Du|<|KK^|2|yy^|.|Du|<\frac{|K-\hat{K}|}{2|y-\hat{y}|}.

Thus, for sufficiently large values of x1x_{1}, 2|u(x1,y)u(x1,y^)|<|KK^|2|u(x_{1},y)-u(x_{1},\hat{y})|<|K-\hat{K}| only if |KK^|>0|K-\hat{K}|>0. This is a contradiction, so K=K^K=\hat{K}. Now, fix x2=yx_{2}=y, for any yy\in\mathbb{R}. The following formula holds:

x1D1u(ξ,y)𝑑ξ=Ku(x1,y).\int_{x_{1}}^{\infty}D_{1}u(\xi,y)d\xi=K-u(x_{1},y).

By our bound |D1u|Cex14|D_{1}u|\leq Ce^{-\frac{x_{1}}{4}},

u(x1,y)K+x1Ceξ4𝑑ξK+Cex14,u(x_{1},y)\leq K+\int_{x_{1}}^{\infty}Ce^{-\frac{\xi}{4}}d\xi\leq K+Ce^{-\frac{x_{1}}{4}},

and

u(x1,y)Kx1Ceξ4𝑑ξKCex14.u(x_{1},y)\geq K-\int_{x_{1}}^{\infty}Ce^{-\frac{\xi}{4}}d\xi\geq K-Ce^{-\frac{x_{1}}{4}}.

Since yy\in\mathbb{R} is arbitrary, this shows that K±Cex14K\pm Ce^{-\frac{x_{1}}{4}} are upper and lower barriers for uu, respectively. In particular, uu decays uniformly along rays parallel to x1\nabla x_{1}. ∎

Proposition 5.1 allows us to use grim hyperplanes and tilted grim hyperplanes as barriers for the ends. We obtain a number of immediate corollaries from this observation.

Corollary 5.2.

Along any ray in the coordinate plane PP not pointing in the direction V-V, the magnitude of the graph uu decays exponentially.

Proof.

Immediate from comparison to tilted grim hyperplanes. ∎

Corollary 5.3.

Each end is represented by the graph of a function uu that is uniformly bounded.

Proof.

This follows from comparison to grim hyperplanes above and below the top and bottom asymptotic planes, respectively. ∎

Corollary 5.4.

Any asymptotic plane of the translator Σ\Sigma must have distance less than π\pi from another asymptotic plane of Σ\Sigma.

Proof.

If not, we could place a grim hyperplane between these two asymptotic planes, and obtain an interior point of contact with Σ\Sigma. ∎

In order to improve on Proposition 5.1 and its consequences, we introduce another barrier,

ϕ=r1/4ex1/2K0(r/2).\phi=r^{1/4}e^{-x_{1}/2}K_{0}(r/2).

Here K0(x)K_{0}(x) denotes the modified Bessel function of the second kind, with parameter 0, i.e. the solution to the modified Bessel’s equation,

x2d2ydx2+xdydxx2y=0.x^{2}\frac{d^{2}y}{dx^{2}}+x\frac{dy}{dx}-x^{2}y=0.

The key point is that the decay and uniform boundedness of uu established in Proposition 5.1 allow us to find domains in which ϕ\phi acts as a barrier for u2u^{2}.

Proposition 5.5.

Let Σ0\Sigma_{0} and u:2BRu:\mathbb{R}^{2}\setminus B_{R}\rightarrow\mathbb{R} be as in Proposition 5.1.

|u|Cr1/αex1/4K0(r/2),|u|\leq Cr^{1/\alpha}e^{-x_{1}/4}\sqrt{K_{0}(r/2)},

where α>4\alpha>4. In particular, u(x)=O(r14(1+δ))u(x)=O\big{(}r^{-\frac{1}{4(1+\delta)}}\big{)}, for δ>0\delta>0 as rr\rightarrow\infty.

Proof.

We first apply the linear operator =Δ+D1\mathcal{L}=\Delta+D_{1} to ϕ\phi.

Δϕ+D1ϕ=ex1/2K0(r/2)(r1/4)+r1/4(ex1/2K0(r/2))+2ex1/2K0(r/2),r1/4.\Delta\phi+D_{1}\phi\\ =e^{-x_{1}/2}K_{0}(r/2)\mathcal{L}(r^{1/4})+r^{1/4}\mathcal{L}\big{(}e^{-x_{1}/2}K_{0}(r/2)\big{)}+2\langle\nabla e^{-x_{1}/2}K_{0}(r/2),\nabla r^{1/4}\rangle.

We evaluate each summand separately.

(ex1/2K0(r/2))\displaystyle\mathcal{L}\big{(}e^{-x_{1}/2}K_{0}(r/2)\big{)} =14ex1/2K0(r/2)+ex1/2ΔK0(r/2)\displaystyle=\frac{1}{4}e^{-x_{1}/2}K_{0}(r/2)+e^{-x_{1}/2}\Delta K_{0}(r/2)
ex1/2x1,K0(r/2)+ex1/2D1K0(r/2)12ex1/2K0(r/2)\displaystyle\;\;\;\;\;-\langle e^{-x_{1}/2}\nabla x_{1},\nabla K_{0}(r/2)\rangle+e^{-x_{1}/2}D_{1}K_{0}(r/2)-\frac{1}{2}e^{-x_{1}/2}K_{0}(r/2)
=ex1/2ΔK0(r/2)ex1/24K0(r/2)\displaystyle=e^{-x_{1}/2}\Delta K_{0}(r/2)-\frac{e^{-x_{1}/2}}{4}K_{0}(r/2)
=ex1/2(2r2K0(r/2)+1rrK0(r/2)14K0(r/2))\displaystyle=e^{-x_{1}/2}\bigg{(}\frac{\partial^{2}}{\partial r^{2}}K_{0}(r/2)+\frac{1}{r}\frac{\partial}{\partial r}K_{0}(r/2)-\frac{1}{4}K_{0}(r/2)\bigg{)}
=ex1/2(142K0r2(r/2)+12rK0r(r/2)14K0(r/2))\displaystyle=e^{-x_{1}/2}\bigg{(}\frac{1}{4}\frac{\partial^{2}K_{0}}{\partial r^{2}}(r/2)+\frac{1}{2r}\frac{\partial K_{0}}{\partial r}(r/2)-\frac{1}{4}K_{0}(r/2)\bigg{)}
=ex1/2r2((r2)22K0r2(r/2)+r2K0r(r/2)(r2)2K0(r/2))\displaystyle=\frac{e^{-x_{1}/2}}{r^{2}}\bigg{(}\bigg{(}\frac{r}{2}\bigg{)}^{2}\frac{\partial^{2}K_{0}}{\partial r^{2}}(r/2)+\frac{r}{2}\frac{\partial K_{0}}{\partial r}(r/2)-\bigg{(}\frac{r}{2}\bigg{)}^{2}K_{0}(r/2)\bigg{)}
=0,\displaystyle=0,

by the definition of the modified Bessel function K0K_{0}. We now evaluate

r1/4\displaystyle\mathcal{L}r^{1/4} =2r2r1/4+1rrr1/4+D1r1/4\displaystyle=\frac{\partial^{2}}{\partial r^{2}}r^{1/4}+\frac{1}{r}\frac{\partial}{\partial r}r^{1/4}+D_{1}r^{1/4}
=316r7/4+14r7/4+(e1er)4r3/4\displaystyle=-\frac{3}{16}r^{-7/4}+\frac{1}{4}r^{-7/4}+\frac{(e_{1}\cdot e_{r})}{4}r^{-3/4}
=116r7/4+(e1er)4r3/4\displaystyle=\frac{1}{16}r^{-7/4}+\frac{(e_{1}\cdot e_{r})}{4}r^{-3/4}

Now we calculate the final term.

2r1/4,ex1/2K0(r/2)\displaystyle 2\langle\nabla r^{1/4},\nabla e^{-x_{1}/2}K_{0}(r/2)\rangle =r3/4ex1/24K0r(r/2)r3/4ex1/24K0(r/2)(e1er)\displaystyle=\frac{r^{-3/4}e^{-x_{1}/2}}{4}\frac{\partial K_{0}}{\partial r}(r/2)-\frac{r^{-3/4}e^{-x_{1}/2}}{4}K_{0}(r/2)(e_{1}\cdot e_{r})

Combining these results, we obtain

Δϕ+D1ϕ\displaystyle\Delta\phi+D_{1}\phi =0+ex1/2K0(r/2)(116r7/4+(e1er)4r3/4)\displaystyle=0+e^{-x_{1}/2}K_{0}(r/2)\bigg{(}\frac{1}{16}r^{-7/4}+\frac{(e_{1}\cdot e_{r})}{4}r^{-3/4}\bigg{)}
+r3/4ex1/24K0r(r/2)r3/4ex1/24K0(r/2)(e1er)\displaystyle\;\;\;\;\;\;\;+\frac{r^{-3/4}e^{-x_{1}/2}}{4}\frac{\partial K_{0}}{\partial r}(r/2)-\frac{r^{-3/4}e^{-x_{1}/2}}{4}K_{0}(r/2)(e_{1}\cdot e_{r})
=r3/4ex1/24K0r(r/2)+o(|r3/4ex1/24K0r(r/2)|).\displaystyle=\frac{r^{-3/4}e^{-x_{1}/2}}{4}\frac{\partial K_{0}}{\partial r}(r/2)+o\bigg{(}\bigg{|}\frac{r^{-3/4}e^{-x_{1}/2}}{4}\frac{\partial K_{0}}{\partial r}(r/2)\bigg{|}\bigg{)}.

The asymptotic expansion at infinity of K0(x)K_{0}^{\prime}(x) is as follows:

K0(x)π2xex(1+38x+O(1x2)).K_{0}^{\prime}(x)\sim-\sqrt{\frac{\pi}{2x}}e^{-x}\bigg{(}1+\frac{3}{8x}+O\bigg{(}\frac{1}{x^{2}}\bigg{)}\bigg{)}.

Thus, for sufficiently large rr, K0(r/2)K_{0}^{\prime}(r/2) is negative. Thus,

ϕ<0 for r sufficiently large.\mathcal{L}\phi<0\;\;\;\;\text{ for $r$ sufficiently large.}

We now consider the operator QQ defined in (2.1), where we set (v1,v2,v3)=(1,0,0)(v_{1},v_{2},v_{3})=(1,0,0).

Qϕ\displaystyle Q\phi =(1+|Dϕ|2)ϕD2ϕ(Dϕ,Dϕ)\displaystyle=(1+|D\phi|^{2})\mathcal{L}\phi-D^{2}\phi(D\phi,D\phi)
=aijDij2ϕ+bjDjϕ\displaystyle=a_{ij}D^{2}_{ij}\phi+b_{j}D_{j}\phi

where (x,u,Du)=(x,z,p)(x,u,Du)=(x,z,p) and

aij=1+|p|2δij+pipja_{ij}=1+|p|^{2}\delta_{ij}+p_{i}p_{j}
bj=(1+|p|2)δ1jb_{j}=(1+|p|^{2})\delta_{1j}

We know that ϕ\phi has the following asymptotic expansion

ϕ(x1,x2)=r1/4ex1/2K0(r/2)Cr1/4ex1r2(118r+O(1r2))\phi(x_{1},x_{2})=r^{1/4}e^{-x_{1}/2}K_{0}(r/2)\sim Cr^{-1/4}e^{\frac{-x_{1}-r}{2}}\bigg{(}1-\frac{1}{8r}+O\bigg{(}\frac{1}{r^{2}}\bigg{)}\bigg{)}

Consequently, we know that for sufficiently large rr, the partial derivatives DiϕD_{i}\phi are bounded above by a multiple of r5/4r^{-5/4}. Thus,

|D2ϕ(Dϕ,Dϕ)|=O(r5/2ex1/2K0(r/2))=o(ϕ)|D^{2}\phi(D\phi,D\phi)|=O(r^{-5/2}e^{-x_{1}/2}K_{0}(r/2))=o(\mathcal{L}\phi)

as rr\rightarrow\infty. We conclude that Qϕ<0Q\phi<0. Consider the family of functions

ϕϵ,N=C1ϕ+C2exp((x1+N)/2)+ϵ,\phi_{\epsilon,N}=C_{1}\phi+C_{2}\exp(-(x_{1}+N)/2)+\epsilon,

where C1,C2>0C_{1},C_{2}>0 are fixed constants which depends on the upper bound of the function uu in the annulus 2BR0\mathbb{R}^{2}\setminus B_{R_{0}}. A quick calculation yields

Q(exp((x1+N)/2))=exp((x1+N)/2)4exp((x1+N)/2)2exp((x1+N))4(exp((x1+N)/2)4).Q(\exp(-(x_{1}+N)/2))=\frac{\exp(-(x_{1}+N)/2)}{4}-\frac{\exp(-(x_{1}+N)/2)}{2}\\ -\frac{\exp(-(x_{1}+N))}{4}\bigg{(}\frac{\exp(-(x_{1}+N)/2)}{4}\bigg{)}.

This expression is clearly negative for x1Nx_{1}\geq-N.

Finally, we calculate Q(u2)Q(u^{2}), where uu satisfies the translator equation (2.1), i.e. Qu=0Qu=0.

Q(u2)\displaystyle Q(u^{2}) =aijDij2(u2)+bjDj(u2)\displaystyle=a_{ij}D^{2}_{ij}(u^{2})+b_{j}D_{j}(u^{2})
=aij(Di(2uDju))+2ubjDju\displaystyle=a_{ij}(D_{i}(2uD_{j}u))+2ub_{j}D_{j}u
=2aij(DiuDju)+2uaijDij2u+2ubjDju\displaystyle=2a_{ij}(D_{i}uD_{j}u)+2ua_{ij}D_{ij}^{2}u+2ub_{j}D_{j}u
=2aij(DiuDju)\displaystyle=2a_{ij}(D_{i}uD_{j}u)
=2(1+|Du|2)|Du|2+2DiuDju(DiuDju)\displaystyle=2(1+|Du|^{2})|Du|^{2}+2D_{i}uD_{j}u(D_{i}uD_{j}u)
=2(1+|Du|2)|Du|2+2|Du|4\displaystyle=2(1+|Du|^{2})|Du|^{2}+2|Du|^{4}
0.\displaystyle\geq 0.

Now that we know that Q(ϕϵ,Nu2)0Q(\phi_{\epsilon,N}-u^{2})\leq 0, we may apply the following quasilinear maximum principle:

Theorem 5.6.

[GT, Theorem 10.1] Let u,vC0(Ω¯)C2(Ω)u,v\in C^{0}(\bar{\Omega})\cap C^{2}(\Omega) satisfy QuQvQu\geq Qv in Ω\Omega, uvu\leq v on Ω\partial\Omega, where

  • (i)

    QQ is locally uniformly elliptic with respect to either uu or vv;

  • (ii)

    the coefficients aija_{ij} are independent of zz;

  • (iii)

    the coefficient b(x,p)=bj(x,p)pjb(x,p)=b_{j}(x,p)p_{j} is non-increasing in zz for each (x,p)Ω×n(x,p)\in\Omega\times\mathbb{R}^{n};

  • (iv)

    the coefficients aij,ba_{ij},b are continuously differentiable with respect to the pp variables in Ω××n\Omega\times\mathbb{R}\times\mathbb{R}^{n}.

It then follows that uvu\leq v in Ω\Omega.

It is clear that for sufficiently large R0R_{0}, conditions (i)-(iv) are satisfied by u2u^{2} and ϕ\phi on the annulus 2BR0\mathbb{R}^{2}\setminus B_{R_{0}}. Consider the large square SNS_{N} with corners (N,N),(N,N),(N,N),(N,N)(N,N),(N,-N),(-N,N),(-N,-N) and NR0>0N\geq R_{0}>0. Let ΩN=SNBR0\Omega_{N}=S_{N}\setminus B_{R_{0}}. Set C1>0C_{1}>0 so that C1ϕ|BR0u2L(2BR0)C_{1}\phi|_{\partial B_{R_{0}}}\geq||u^{2}||_{L^{\infty}(\mathbb{R}^{2}\setminus B_{R_{0}})} and set C2>u2L(2BR0)C_{2}>||u^{2}||_{L^{\infty}(\mathbb{R}^{2}\setminus B_{R_{0}})}. By Proposition 5.1, Corollary 5.2, and Corollary 5.3, ϕϵ,N>u2\phi_{\epsilon,N}>u^{2} on ΩN\partial\Omega_{N} for sufficiently large NN and some arbitrary small ϵ>0\epsilon>0. On the edge from (N,N)(N,-N) to (N,N)(-N,-N), x1Nx_{1}\equiv-N and ϕϵ,N>C2>u2\phi_{\epsilon,N}>C_{2}>u^{2}. The other three edges lie in a sector not containing V-V, so by Corollary 5.2, for sufficiently large values of NN, u2<ϵu^{2}<\epsilon on these edges. Thus, given an ϵ\epsilon, for sufficiently large NN, ϕϵ,N>u2\phi_{\epsilon,N}>u^{2} on ΩN\partial\Omega_{N}.

Having chosen R0>0R_{0}>0 sufficiently large, on the domains ΩN\Omega_{N}, we have Q(ϕϵ,N)<0Q(\phi_{\epsilon,N})<0, so we may apply [GT, Theorem 10.1] and obtain that ϕϵ,Nu20\phi_{\epsilon,N}-u^{2}\geq 0 on ΩN\Omega_{N}. Sending NN\rightarrow\infty, we see that C1ϕ+ϵu2C_{1}\phi+\epsilon\geq u^{2} on 2BR0\mathbb{R}^{2}\setminus B_{R_{0}}. Since ϵ>0\epsilon>0 is arbitrary, we send ϵ\epsilon to zero and conclude that C1ϕu2C_{1}\phi\geq u^{2}.

We finish the proof of the proposition by observing that all of the above arguments apply when we select ϕ=r2/αex1/2K0(r/2)\phi=r^{2/\alpha}e^{-x_{1}/2}K_{0}(r/2) and α(4,)\alpha\in(4,\infty) is an arbitrary constant. ∎

Proof of Theorem 1.2.

The theorem immediately follows from Proposition 4.1 and Proposition 5.5. ∎

Acknowledgments

I would like to thank Bing Wang for suggesting this problem. I would also like to thank Bing Wang and Sigurd Angenent for many helpful conversations and their careful feedback on the drafts of this paper. I am also very grateful to Sigurd Angenent for his invaluable guidance in the writing of §5, especially for explaining the importance of the Bessel functions and sharing his overall wisdom around asymptotic formulae. Lastly, I’d like to thank Professor Yuxiang Li for pointing out that conditions on the area ratios could be dropped and suggesting the proof of Proposition 2.10.

References

  • [CM] T. H. Colding and W. P. Minicozzi II, A Course in Minimal Surfaces, Graduate Studies in Mathematics, 121, American Mathematical Society, (2011).
  • [E1] K. Ecker, On regularity for mean curvature flow of hypersurfaces, Calc. Var. 3, 107-126 (1995).
  • [GT] D. Gilbarg and N. Trudinger, Elliptic Partial Differential Equations of Second Order, Springer-Verlag, 3rd Ed., (2001).
  • [I] T. Ilmanen, Singularities of mean curvature flow of surfaces, http://www.math.ethz.ch/~ilmanen/papers/sing.ps
  • [LT] P. Li and L.-F. Tam, Complete surfaces with finite total curvature, J. Diff. Geom. 33, No. 1, 139-168 (1991).
  • [MS] S. Müller and V. Šverák, On surfaces of finite total curvature, J. Diff. Geom. 43, No. 2, 229-258 (1995).
  • [MSHS] F. Martin, A. Savas-Halilaj, K. Smoczyk, On the topology of translating solitons of the mean curvature flow, Calc. Var. Partial Differ. Equ. 54, No. 3, 2853-2882 (2015).
  • [PV] J. Pipher and G Verchota, A maximum principle for biharmonic functions in Lipschitz and C1C^{1} domains, Comment. Math. Helv. 68, no. 1, 385-414 (1993).
  • [Sc] R. Schoen, Uniqueness, symmetry, and embeddedness of minimal surfaces, J. Diff. Geom. 18, No. 4, 791-809 (1983).
  • [Sh] K. Shiohama, Total curvatures and minimal areas of complete open surfaces, Proc. Amer. Math. Soc. 94, 310-316 (1985).
  • [Si] L. Simon, Existence of surfaces minimizing the Willmore functional, Comm. Anal. Geom. 1, 281-326 (1993).
\enddoc@text

A. Illustrations for the Proof of Proposition 4.1

Figure 1. Case 1
[Uncaptioned image]
Figure 2. Case 2
[Uncaptioned image]