The Structure of Translating Surfaces with Finite Total Curvature
Abstract.
In this paper, we prove that any mean curvature flow translator with finite total curvature and one end must be a plane. We also prove that if the translator has multiple ends, they are asymptotic to a plane containing the direction of translation and can be written as graphs over . Finally, we determine that the ends of are strongly asymptotic to and obtain quantitative estimates for their asymptotic behavior.
1. Introduction
A proper, embedded hypersurface is a mean curvature flow translator if there is a unit vector such that satisfies the following equation:
(1.1) |
where is the mean curvature of at a point , and is the component of perpendicular to the tangent plane to at . This is equivalent to saying that the family of surfaces satisfies the mean cuvature flow (MCF) equation,
(1.2) |
where is the vector in corresponding to the position of the point . 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 be a complete embedded MCF translator with one end, and finite total curvature
(1.3) |
Then, is a plane parallel to , the direction of motion.
We also prove
Theorem 1.2.
Let be a complete embedded translator with finite total curvature, as in Theorem 1.1. Outside of some ball , the ends of may be written as the graphs of functions over a fixed plane . Furthermore, each decays at a rate for any as . In particular, each decays radially at an exponential rate in any sector in that excludes the ray , , where 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 , 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 , 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 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 . Thus, inside any annulus , any end can be written as a graph over a plane with gradient bounded by . If we fix an annulus we can take a blow-down subsequence that can be written in as a sequence of graphs over a single plane with gradient bounded by where . 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 , the quantity , is uniformly bounded when taken over all centers in and radii . 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, , then the “in between” annular region can be written as a graph over the blow-down limit plane . This is shown by examining two cases: (1) that has the same orientation as a graph over as , and (2) that these two annular regions have different orientations as graphs over . 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 . In the second case, we integrate the geodesic curvature over the boundary of 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 , . Our main tool is the maximum principle: first we use an exponential barrier to bound the growth of the functions in the upper half-plane as . In particular, this decay implies the uniform boundedness of the 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 be a unit vector in the direction of motion of the translating graphs of and , which are defined on an open set . Then, satisfies an equation of the form
where has ellipticity constants and depending only on the upper bounds for the gradients .
Proof.
The , , both satisfy the quasilinear elliptic equation
(2.1) |
Set
We follow the proofs of Gilbarg and Trudinger in [GT, Theorem 10.7] and Colding and Minicozzi in [CM, Lemma 1.26]. Let , and . Let
By the Fundamental Theorem of calculus and the chain rule, satisfies
All that remains to show is the uniform ellipticity of . We show this by demonstrating the positive definiteness of the matrix differential . If is a unit vector and , then
Thus, is a positive definite matrix whose ellipticity constants depend on the upper bounds of . ∎
Corollary 2.2.
Let be an open connected neighborhood of the origin. If satisfy the translator equation (2.1) with respect to the direction with and , then .
In addition to this maximum principle, we also obtain a weak maximum principle for the gradients of translators that are graphs over planes containing , the direction of motion.
Lemma 2.3.
If , is a bounded domain in , and satisfies (2.1), then the partial derivatives and achieve their maxima and minima on the boundary.
Proof.
We set without loss of generality. The equation (2.1) becomes
(2.2) |
We differentiate this equation with respect to to obtain
Let . We calculate
This implies that satisfies a differential equation of the form , where the differential operator has coefficients given by
We calculate
Similarly
Since is assumed to be smooth and is bounded, the operator is uniformly elliptic. In particular, satisfies the hypotheses of [GT, Theorem 3.1] and by applying this theorem, we obtain the weak maximum/minimum principle for . ∎
We now prove a result about the pointwise decay of the second fundamental form and its derivatives. To obtain this result, we use Ecker’s -Regularity theorem for the mean curvature flow of surfaces.
Theorem 2.4.
[E1, Ecker Regularity Theorem] There exist constants and such that for any solution of mean curvature flow, any , and the inequality
(2.3) |
implies the mean value estimate
(2.4) |
Lemma 2.5.
For a translator with finite total curvature, if is distance from the origin, then
Furthermore, in the positive half-space defined by ,
Proof.
This is a consequence of Ecker’s -regularity theorem, Theorem 2.4. By the finite total curvature condition, there is a ball centered at the origin such that . We pick any , and take . Suppose that and take the ball . Consider the mean curvature flow solution defined on the interval . Notice that never intersects for any time . Thus,
Theorem 2.4, Ecker’s regularity theorem, yields
Thus, , where is distance to the origin (which is comparable to ). This proves the first statement of the Lemma.
Now we prove the improved curvature decay in the upper half-space of points in with positive component. Suppose that and that . Consider the ball and the MCF solution defined on the interval . Since , never intersects . Thus, by Theorem 2.4,
Choosing , we see that . Since , where is , we conclude that . ∎
Corollary 2.6.
If is the normal vector to a translator moving in the direction at the point , then
where is the distance between and the origin.
Proof.
Therefore,
∎
Before we can begin the proof of the main theorems, we must show that the embedding of into 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 be a complete surface with finite total curvature. Then there exists a constant depending only on , such that
for all and for all , where is the intrinsic ball of radius .
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 .
Theorem 2.8 (Huber).
If is a complete surface of finite total curvature immersed in , then is conformally equivalent to a compact Riemann surface with finitely many points deleted.
In particular, we note that Huber’s theorem implies that must have finitely many ends and bounded genus. By Huber’s theorem, we have a bijective conformal map, , from the annulus to a given end of . Following [MS], we consider the family of rescalings
According to [MS], there exists a sequence with the following properties:
-
(i)
There exists a plane such that the Gauss map converges to the constant in on the punctured complex plane.
-
(ii)
The maps converge uniformly on compact subsets of to a conformal mapping which satisfies .
If we consider as a holomorphic function , it can be seen that is of the form
However, if , the graph of the map must have an self-intersection, since the are converging uniformly as maps into . This contradicts the assumption of embeddedness, and thus must be zero. Müller and Šverák define the multiplicity of an end by the integer . Therefore, in our situation each end of 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 . 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 , consider the conformal parametrization given by Huber’s theorem. For each there exists such that the following statement holds. If is a square such that and if are neighboring vertices of , then
where is a constant.
By [MS, Theorem 4.2.1], for as in Proposition 2.9, approaches uniformly as . Taking piecewise--paths, we can now compare intrinsic and extrinsic distances on . Thus, if , there exists a constant such that
In particular, if , contains the connected component of containing .
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 is a complete, embedded surface with finite total curvature and is a compact domain in , there exists a constant that bounds the following area ratios:
(2.5) |
Proof.
Müller and Šverák prove in [MS, Corollary 4.2.5] that for a fixed point , the intrinsic and extrinsic distance functions coincide in the limit. That is,
(2.6) |
Thus, for very large radii , the intrinsic balls are comparable to extrinsic balls in the following sense:
Let denote the area of the geodesic ball . A result from [Sh] cited in the proof of [MS, Corollary 4.2.5] tells us that
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
such that
By comparison in the limit as , we can see that this property holds for all , and not only for points contained in .
We have a family of lower semi-continuous functions defined on
Take the supremum of this family with respect to the index . 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 . 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 . By Proposition 2.10, we know that the area ratios are bounded for balls centered in this region. We now consider points . We know that if and is the connected component of containing , then there exists a uniform constant such that
and thus,
For a general point , we consider the set , which is the nearest point projection to the ends of the surface . Then,
The theorem of Hartman, Theorem 2.7, implies that there is a constant depending on and such that
Combining the above bound with Proposition 2.10 proves a uniform bound on area ratios.
Corollary 2.11.
There exists such that
3. Blowing Down
In this section, we prove that each blow-down sequence of has a subsequential -limit.
Lemma 3.1.
Let be a translating end with finite total curvature. For every sequence there is a subsequence such that the -limit of the rescalings in is a plane that is parallel to .
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 ).
For each , there is a constant such that if is a smooth 2-manifold properly embedded in and
then there are pairwise disjoint closed disks in such that
(3.1) |
and for any such that intersects transversally and , we have
where each is an embedded disk. Furthermore, for each , there is a 2-plane , a simply-connected domain , disjoint closed balls , and a function
such that
(3.2) |
and
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 , which we call “pimples.” We recall that Simon states on p. 289 of [Si] that each the boundary of each disk is a graphical curve contained in . In particular, this tells us a) that the can be taken to be disjoint (they cannot be connected by a pimple , as these are all topological disks) and b) that if we assume , then the connected component of containing is contained in a neighborhood of the cone
Lemma 3.3.
Let be given, and let be a complete embedded translator, and select a point such that
Let be the connected component of containing (where is as in Theorem 3.2) and consider the collection of pimples . Then there exists such that is the graph of a function with gradient bounded by , defined on a plane parallel to the direction of motion, .
Proof.
Away from the pimples , the component is the graph of a function with small gradient over a plane, which we denote . By Corollary 2.6, . Since
we can take to be a plane that is parallel to . Let be the standard orthogonal projection map. Now, consider the following cone:
By the graphical decomposition lemma, is contained in a neighborhood of with radius :
Let denote the disk centered at of radius in the plane . This boundedness of ensures that for sufficiently small , the intersections are compact and do not intersect the boundary for all . By the diameter bounds on the pimples and the transversality theorem, there is a number such that and intersects transversely. By Theorem 3.2, this intersection is a single graphical curve over that bounds a disk, .
We now show that the surface given by is a graph over . 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 be an open set with boundary and let be simple closed curves each of which are graphs over distinct components of with bounded slope. Further assume that for any point , the tangent plane to is well defined and does not contain the vertical normal vector . Then any translator contained in with must be graphical over .
Proof.
Let be spanned by unit vectors and let the normal direction be . Given the plane , is divided into the portions above and below the plane. We reflect below the plane to obtain a new translator below the plane. Note that because is contained in , reflection in preserves the translator equation. We decrease until we encounter the first point where (a) contains the vector or (b) and have an interior point of contact. Let the critical height be called . Suppose that (a) occurs–this means that the tangent planes and coincide and we may apply the maximum principle, Corollary 2.2, to show that must have a reflection symmetry through . This contradicts the graphicality of the boundary unless coincides with .
If is not one-to-one (i.e. 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 has a reflection symmetry through . This is a contradiction, and thus neither case (a) nor case (b) occurs. Because is a disjoint union of graphs with bounded slope over disjoint planar curves, there is no boundary point of contact. Thus, the projection of to is one-to-one and is a graph. ∎
If we let and , then clearly satisfies the conditions of Theorem 3.4 and is a graph over . Furthermore, we know from the graphical decomposition lemma, Theorem 3.2, that . By Lemma 2.3, the weak maximum principle for derivatives, can be written as the graph of a function defined on with . This completes the proof of the lemma. ∎
Given some , we can find such that
and decomposes into disjoint annular ends. Given one of these ends, , we wish to glue a disk with small total curvature to the inner boundary. The resulting surface will be a topological disk without boundary to which we may apply Simon’s Lemma (Theorem 3.2). To this end, for each , we find a cylindrical curve with special properties that will allow us to carry out the gluing construction.
Lemma 3.5.
Given and as above, there exists a plane such that , a radius such that , and a graphical curve over such that
Proof.
By Corollary 2.6, we may choose large enough that in . Let . The norm of the tangential gradient is bounded by 1. The coarea formula says that
Note that by Hölder’s inequality and the assumed area bounds
There exists some such that
(3.3) | ||||
By embeddedness, we may assume that intersects transversely and is a closed curve. From the inequality (3.3), the normal vector of has very small oscillation on the curve for sufficiently small . Thus, the curve is contained in a small graphical annular region with gradient bounded by . The radius around the graphical neighborhood of each point in is extended to by applying Lemma 3.3. Approximately, this tells us that is very close (on the order of ) to the cross section of some translate of with the ball .
Ultimately, what we obtain from this argument is that the union of these neighborhoods of contains the graph of a function defined on a planar annulus with width and . Note that we may choose large enough that , where is orthogonal projection to .
Assume that and let for all . The coarea formula gives
Thus, there exists such that
where we have used the fact that . Furthermore, by construction, . Thus,
(3.4) | ||||
(3.5) |
Note that is an absolute constant that does not depend on or . Letting the curve equal completes the proof of the lemma. ∎
In the proof of Lemma 3.5, is a graph over and is contained in the graph of a function , defined over an annulus such that . We now cut out the interior region in bounded by , leaving only an annular end with total curvature bounded by . 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 to replace the excised region.
Lemma 3.6.
[Si, Lemma 2.2] Let be smooth embedded, , a plane containing , for some open neighborhood of , and
Also, let satisfy
(3.6) |
Then
(3.7) |
where , is the second fundamental form of , and is 1-dimensional Hausdorff measure (i.e. arc-length measure) on ; is a fixed constant independent of , .
Note that the solution 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 in , a bounded Lipschitz domain in , and is continuous in , then there is a constant that depends only on the Lipschitz character of and independent of the diameter of such that
Since , Theorem 3.7 implies that , and thus there exists a uniform constant not depending on or such that
Combining this estimate with (3.7) and Lemma 3.5, we obtain the desired bound
where is independent of and .
We attach the graph of to the end of bounded by . Simply joining these surfaces along results in a surface , which does not satisfy the hypotheses of Theorem 3.2: a surface is needed. To improve the regularity, we approximate the piecewise surface in the -Sobolev norm by smooth functions.
Lemma 3.8.
Given as in Lemma 3.5 and a connected component of , we can find a smooth topological disk such that outside the ball .
Proof.
Take a small tubular neighborhood of in . On the outside of , is equal to the graph of over the plane . On the inside of , is equal to the graph of over . Let be the function over an annular domain containing such that .
First, we confirm that is in . Since , we need only find weak second derivatives. Notice that is smooth away from . We define away from the measure zero set :
Now, consider a test function .
Since , this is enough to show that the weak second derivatives we defined are valid and . Note that we used (3.6) to get rid of the last term in the second line.
We approximate in by smooth functions. Let be a cutoff function that is uniformly equal to 1 in a neighborhood of and compactly supported in . Let be the regularization of , with chosen so that is also compactly supported in and
Then the function
is equal to in a uniform neighborhood of the boundary and has . Since is uniformly bounded, we have
Since matches in an open neighborhood of , the surface obtained by joining them, , is smooth and equal to outside . ∎
Choose and let be a sufficiently large radius. We now prove the following lemma for . We assume that .
Lemma 3.9.
Away from the ball , the surface is the graph of a function with gradient bounded by over a plane parallel to .
Proof.
We apply Simon’s Lemma, Theorem 3.2, to . Lemma 3.3 ensures that (where ) is an embedded, closed graphical curve bounding a connected topological disk. Away from pimples, Simon’s Lemma tells us that is the graph of a function defined on a domain in with gradient bounded by . Let be orthogonal projection and let be as in Lemma 3.8. Let be a disk such that . By the diameter bounds (3.1), there exists and such that and do not intersect the pimples and intersect 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 with gradient bounded by . 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 , where , a sequence decreasing to zero , and a family of rescalings , with . Let be obtained for by cutting out a high curvature region and pasting in a disk as above. Note that in the inequality
the constant depends only on , and is independent of . Observe also that outside of some ball , for any .
Fix and . By Lemma 3.9, for sufficiently large values of , can be written as a graph of a function over a plane parallel to with -norm bounded by in . Un-fixing , we see that by the compactness of the collection of planes , there exists a subsequence such that converges to a fixed plane inside in the -norm.
Now, unfixing , we take a further diagonal subsequence such that the rescalings of our designated end, , converge to in the 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 (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 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 be a blow-down sequence converging to . Fix . Consider the sequence of annuli . Each of these can be written as the graph of a function over with gradient , where . Consider the curve and its projection to an embedded Jordan curve in , which we denote . We take to be the open set in bounded by and apply the method of moving planes, Theorem 3.4 to determine that is a graph over . Then we apply the weak maximum principle, Theorem 2.3, to determine that this graph has gradient . As , the sequence must be a sequence of graphical disks over with increasingly small gradient. Thus, 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 . 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 , the -limit in of any blow-down sequence , is a unique plane parallel to . Consequently, the ends of can be written as graphs over outside of some ball .
Proof.
Let and be the modified surface from the previous section: a complete embedded topological disk, equal to the translator outside a ball and with total curvature less than . Consider a blow-down sequence , with that converges to a plane , and take . Let be a large fixed radius, and let be large enough that and can be written as graphs of functions and over annuli in with gradient bounded by . Let us also define the curves and which will denote the inner boundary and the outer boundary respectively.
If we project the space curves and to via the orthogonal projection map , we obtain simple closed planar curves and . Equivalently, and . Thus, and bound a large topological annulus in . Let be the normal vector to . Consider the solid cylinder . Suppose that . Then, is contained inside , and since the boundary curves and are graphs, we can apply the moving planes method, Theorem 3.4, to find that is a graph over .
Suppose that , i.e. the surface turns back and intersects the cylinder of small radius. Consider the domain in the plane such that and is bounded by the Jordan curve . We separate this situation into two cases. These cases are illustrated in Figure A and Figure A in Appendix A.
Case 1: is non-graphical or a multigraph over in either of the connected components of . Without loss of generality, let us assume this occurs in the upper half space . In this case, we may use a modified method of moving planes. Consider the surface (a topological disk) contained in the cylinder with boundary equal to . If are coordinates in and indicates height in , we take the plane parallel to and the associated surfaces and as in the proof of Theorem 3.4.
Let be the minimum height of in the upper half space . Note that
where we may assume without loss of generality that is sufficiently large that the second inequality holds. The first inequality follows from repeated applications of Lemma 3.9, which tells us that once leaves the ball , it will never re-enter.
If a first point of interior tangency occurs, it must occur for . This means that it occurs outside of and around this point of contact and 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 . However, is a graph over , so there is no point of contact outside on the boundary. Thus, the orthogonal projection to must be one-to-one, so Case 1 cannot occur.
Case 2: is a single-valued graph over . Note that the plane is spanned by two orthogonal vectors and and is normal to the coordinate vector . Let us take the cross section of with respect to the plane spanned by and passing through the origin. Since the normal part of has norm by Corollary 2.6, the intersection with this plane is transverse for sufficiently large values of . 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 . Without loss of generality, take the component that intersects on the left. This curve is an embedded curve with boundary in and , which we henceforth denote by .
We now describe some properties of the curve . Let and be the unique intersection points of with and : these are the endpoints of . Let be the length of . We parametrize by arclength such that
By assumption, crosses the line exactly once. We wish to use this condition to determine whether is close to 1 or -1. First note that
is much longer than
We then consider the two cases (1) is on the left of vertical axis and (2) is on the right of this axis. In case (1), must cross an even number of times, which cannot happen. Thus, (2) must hold. If on , integration tells us that must be on the left of , which is impossible. Thus, in all situations. In particular, .
Let and be parametrized by
and
respectively. The two surfaces are oriented by the respective 2-vectors
Since these two surfaces are graphical annuli with very small gradient, their orientation 2-vectors are close to
respectively. Note that at all points , is “almost” a tangent vector to perpendicular to , by Corollary 2.6. Since , we can deduce that
for any and . Equivalently, the normal vectors to the two annuli determine opposite orientations with respect to a fixed frame on . The crucial consequence of this fact is that the geodesic curvatures along curves “nearby” and will have the same sign and be approximately equal to . 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 and in and respectively such that
and such that
We consider the annular subset of bounded by the curves and , which we call . We apply the Gauss-Bonnet theorem:
where is the geodesic curvature, is equal to the number of connected components, is the genus, and is equal to the number of boundary components. Substituting and , we obtain
Claim 4.2.
Proof.
We apply similar reasoning to a proof on p.294-5 of [Si]. We parametrize and as follows.
(4.1) |
and
(4.2) |
where and is identified with . We calculate the integral
Differentiating (4.1) and (4.2), and combining with the above calculation, we obtain that the total curvature vector can be writted in the following form:
where is a vector field on such that and is the inward pointing unit normal to with respect to the surface . Since
for any and , the geodesic curvatures and have the same sign. We calculate
This concludes the proof of the claim. ∎
Claim 4.2 and the Gauss-Bonnet Theorem imply that
for some small . Given the assumption of small total curvature, this is contradiction. Thus, Case 2 does not occur and must be the graph of a function over . Since on and , the weak maximum principle implies that on .
Now, consider two blow down sequences and , with and has blow-down limit equal to the plane . For any sufficiently large , up to a subsequence of , we can find . The annular set is contained in the larger annulus , which must be a graph over with gradient bounded by . Now that can be written as a graph with gradient bounded by over , we see that in fact the blow-down limit of must be the plane as well.
In fact, this tells us that the entire end can be written as the graph of of a function over with and . That each end is a graph over the same plane 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 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 be an end of a translator that translates with unit speed in the -direction and has finite total curvature. There exists a function whose graph represents and whose magnitude decays uniformly at a rate for any positive as .
Proof.
To prove this statement, we apply a barrier argument to the -derivative of , . We set . By the proof of Lemma 2.3, satisfies a differential equation of the form , where the differential operator has coefficients given by
We now calculate .
Thus,
We now choose our barrier function, . We calculate :
We calculate the coefficients and here:
We now rely on the fact that as in order to estimate the lower order terms for large values of .
for sufficiently large. Let be such that the previous inequality holds for . Let be such that on the line . Note that such a must exist, since as the radius approaches infinity, and is in particular uniformly bounded. Now for some , define the function . Since there are no order zero terms, for . Thus,
Consider the semicircular domains parametrized by defined by the intersection of the balls , of radius and centered around , and the upper half-plane . Since as the radius approaches infinity, for each , there exists some such that on the boundary . The weak maximum principle [GT, Theorem 3.1] tells us that on . Sending and then , we see that in the upper half-plane.
This in turn implies that on . Integrating along rays in the direction of motion, we can conclude that approaches a limit along each ray parallel to the direction of motion. Fix real numbers and and suppose that
Since , there exists a radius such that on ,
Thus, for sufficiently large values of , only if . This is a contradiction, so . Now, fix , for any . The following formula holds:
By our bound ,
and
Since is arbitrary, this shows that are upper and lower barriers for , respectively. In particular, decays uniformly along rays parallel to . ∎
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 not pointing in the direction , the magnitude of the graph decays exponentially.
Proof.
Immediate from comparison to tilted grim hyperplanes. ∎
Corollary 5.3.
Each end is represented by the graph of a function 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 must have distance less than from another asymptotic plane of .
Proof.
If not, we could place a grim hyperplane between these two asymptotic planes, and obtain an interior point of contact with . ∎
In order to improve on Proposition 5.1 and its consequences, we introduce another barrier,
Here denotes the modified Bessel function of the second kind, with parameter , i.e. the solution to the modified Bessel’s equation,
The key point is that the decay and uniform boundedness of established in Proposition 5.1 allow us to find domains in which acts as a barrier for .
Proposition 5.5.
Proof.
We first apply the linear operator to .
We evaluate each summand separately.
by the definition of the modified Bessel function . We now evaluate
Now we calculate the final term.
Combining these results, we obtain
The asymptotic expansion at infinity of is as follows:
Thus, for sufficiently large , is negative. Thus,
We now consider the operator defined in (2.1), where we set .
where and
We know that has the following asymptotic expansion
Consequently, we know that for sufficiently large , the partial derivatives are bounded above by a multiple of . Thus,
as . We conclude that . Consider the family of functions
where are fixed constants which depends on the upper bound of the function in the annulus . A quick calculation yields
This expression is clearly negative for .
Finally, we calculate , where satisfies the translator equation (2.1), i.e. .
Now that we know that , we may apply the following quasilinear maximum principle:
Theorem 5.6.
[GT, Theorem 10.1] Let satisfy in , on , where
-
(i)
is locally uniformly elliptic with respect to either or ;
-
(ii)
the coefficients are independent of ;
-
(iii)
the coefficient is non-increasing in for each ;
-
(iv)
the coefficients are continuously differentiable with respect to the variables in .
It then follows that in .
It is clear that for sufficiently large , conditions (i)-(iv) are satisfied by and on the annulus . Consider the large square with corners and . Let . Set so that and set . By Proposition 5.1, Corollary 5.2, and Corollary 5.3, on for sufficiently large and some arbitrary small . On the edge from to , and . The other three edges lie in a sector not containing , so by Corollary 5.2, for sufficiently large values of , on these edges. Thus, given an , for sufficiently large , on .
Having chosen sufficiently large, on the domains , we have , so we may apply [GT, Theorem 10.1] and obtain that on . Sending , we see that on . Since is arbitrary, we send to zero and conclude that .
We finish the proof of the proposition by observing that all of the above arguments apply when we select and is an arbitrary constant. ∎
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 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).
A. Illustrations for the Proof of Proposition 4.1
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/b3c40054-c544-4ba7-801d-5d75adbc5afc/x1.png)
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/b3c40054-c544-4ba7-801d-5d75adbc5afc/x2.png)