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

Translating Annuli for Mean Curvature Flow

D. Hoffman David Hoffman
Department of Mathematics
Stanford University
Stanford, CA 94305, USA
E-mail address:dhoffman@stanford.edu
F. Martín Francisco Martín
Departamento de Geometría y Topología
Instituto de Matemáticas de Granada (IMAG)
Universidad de Granada
18071 Granada, Spain
E-mail address:fmartin@ugr.es
 and  B. White Brian White
Department of Mathematics
Stanford University
Stanford, CA 94305, USA
E-mail address:bcwhite@stanford.edu
(Date: 25 August, 2023. Revised 24 July, 2024)
Abstract.

We construct a family 𝒜\mathscr{A} of complete, properly embedded, annular translators MM such that MM lies in a slab and is invariant under reflections in the vertical coordinate planes. For each MM in 𝒜\mathscr{A}, MM is asymptotic as zz\to-\infty to four vertical planes {y=±b}\{y=\pm b\} and {y=±B}\{y=\pm B\} where 0<bB<0<b\leq B<\infty. We call bb and BB the inner width and the (outer) width of MM. We show that for each bπ/2b\geq\pi/2 and each s>0s>0, there is an M𝒜M\in\mathscr{A} with inner width bb and with necksize ss. (We also show that there are no translators with inner width <π/2<\pi/2 having the properties of the examples we construct.)

Key words and phrases:
mean curvature flow, translators.
2010 Mathematics Subject Classification:
Primary 53E10, 53C21, 53C42
F. Martín was partially supported by the MICINN grant PID2020-116126-I00, by the IMAG–Maria de Maeztu grant CEX2020-001105-M / AEI / 10.13039/501100011033 and by the Regional Government of Andalusia and ERDEF grant P20-01391. B. White was partially supported by grants from the Simons Foundation (#396369) and from the National Science Foundation (DMS 1404282, DMS 1711293).
The authors would like to thank Leonor Ferrer for helpful suggestions. They would also like to thank the reviewer for the careful reading and helpful suggestions.

1. Introduction

A (normalized) translator is a surface MM in 𝐑3\mathbf{R}^{3} such that

tMt𝐞3t\mapsto M-t\mathbf{e}_{3}

is a mean-curvature flow. This is equivalent to the condition that the mean curvature vector at each point of MM is equal to (𝐞3)(-\mathbf{e}_{3})^{\perp}.

A vertical plane is a translator, and it is complete. There are also graphical translators, i.e., complete, translating surfaces that are graphs z=z(x,y)z=z(x,y) over open subsets of 𝐑2\mathbf{R}^{2}. They have been completely classified. In particular, every graphical translator is a grim reaper surface (tilted or untilted), a Δ\Delta-wing, or a bowl soliton. See Figure 1. These surfaces are described in Section 2. The graphical examples are all simply connected.

Refer to caption
Refer to caption
Refer to caption
Figure 1. From left to right: Two grim reaper surfaces (b=π/2b=\pi/2 in yellow and b>π/2b>\pi/2 in blue), a bowl soliton, and a Δ\Delta-wing.

The next simplest examples are the rotationally-invariant translating annuli (also known as translating catenoids): for every R>0R>0, there is a complete translating annulus MM that is rotationally invariant about ZZ (the zz-axis) and whose distance from ZZ is RR. Furthermore, it is unique up to a vertical translation. See [CSS].

Refer to caption
Figure 2. A translating catenoid of revolution.

It is natural to wonder whether there are other complete, translating annuli. That is, are there complete translating annuli that are not surfaces of revolution? In this paper, we show that there is a large family of such annuli.

Gluing is a powerful tool that has been successfully used in many geometric problems, including construction of translators [del-pino, nguyen09, nguyen13, nguyen-survey, smith21]. It would be natural to try to use gluing to construct non-rotational translating annuli by connecting two Δ\Delta-wings (or two untilted grim reaper surfaces), one slightly above the other, by a small catenoidal neck. (Grim reaper surfaces and Δ\Delta-wings are shown in Figure 1 and discussed in Section 2.) We do in fact construct surfaces that fit that description (Figure 4, left). However, much to our surprise, our method (a continuity method) also produced other surfaces, “uncapped annuloids”, that exhibit strikingly different behavior (Figure 4, right). We believe that such uncapped annuloids could not arise from any gluing method.

We define an annuloid to be a complete, properly embedded translator MM such that

  1. (1)

    MM is an annulus.

  2. (2)

    MM lies in a slab {|y|B}\{|y|\leq B^{\prime}\}.

  3. (3)

    MM is symmetric with respect to the vertical coordinate planes.

  4. (4)

    MM is disjoint from the zz-axis, ZZ.

  5. (5)

    M+(0,0,z)M+(0,0,z) converges smoothly as zz\to\infty to four planes {y=±b}\{y=\pm b\} and {y=±B}\{y=\pm B\} for some 0bB<0\leq b\leq B<\infty.

  6. (6)

    M(0,0,z)M-(0,0,z) converges as zz\to\infty to the empty set.

We define the width B(M)B(M) of MM to be the number BB. (One can prove that BB is also the smallest BB^{\prime} such that (2) holds; see Corollary B.8.) We define the inner width b(M)b(M) of MM to be the number bb.

In this paper, we prove existence of a large family of annuloids. In particular, we prove

Theorem 1.1.

For each bπ/2b\geq\pi/2 and for each 0<s<0<s<\infty, there exists an annuloid with inner width bb and with necksize ss.

The restriction bπ/2b\geq\pi/2 is not arbitrary. All of our examples have “finite type” (as defined in Section 5.) There are no finite-type annuloids with inner width <π/2<\pi/2. See Theorem B.10.

The annuloids in Theorem 1.1 are obtained as follows. Using a path-lifting argument, we prove an analogous existence theorem for compact translating annuli bounded by pairs of nested rectangles. As the lengths of the rectangles tend to infinity, the rectangles converge to four parallel lines. We prove that suitable vertical translates of the compact annuli converge to annuloids (without boundary).

To make Theorem 1.1 precise, one needs to define necksize. There are various natural definitions, such as: the length of the shortest homotopically nontrivial curve in MM, or the radius of the smallest ball containing a nontrivial curve, or the distance from the zz-axis, ZZ, to the surface. Our existence result is true for any of those definitions. For small necks, the different notions of necksize are essentially equivalent. (See, for example, Corollary 16.4.) However, the following definition turns out to be most convenient notion of necksize. In particular, for large necks, it is more suitable than the other notions.

Definition 1.2.

If MM is a surface, we let x(M)x(M) be the distance from ZZ to M{y=0}M\cap\{y=0\}.

One can think of x(M)x(M) as measuring the size of the neck in the xx-direction. See Figure 3.

Conditions (4), (5), and (6) in the definition of annuloids imply that if MM is an annuloid, then x(M)>0x(M)>0. We do not know whether there is a unique annuloid with a given inner width and a given necksize. However, the family of annuloids we construct behaves, in some ways, like a nice 22-parameter family, as indicated by the following theorem. (See also Theorems 7.3, 21.2, 17.1, and 17.2.)

Refer to caption
Figure 3. The distance x(M)x(M) for a translating annulus. This annulus is bounded by a pair of nested convex curves.
Theorem 1.3.

There is a family 𝒜\mathscr{A} of annuloids in 𝐑3\mathbf{R}^{3} such that the map M𝒜B(M)M\in\mathscr{A}\mapsto B(M) is continuous, and such that the map

Φ:𝒜[π/2,)×(0,),\displaystyle\Phi:\mathscr{A}\to[\pi/2,\infty)\times(0,\infty),
Φ(M)=(b(M),x(M))\displaystyle\Phi(M)=(b(M),x(M))

is continuous, proper, and surjective.

For bπ/2b\geq\pi/2, let

𝒜(b):={M𝒜:b(M)=b}.\mathscr{A}(b):=\{M\in\mathscr{A}:b(M)=b\}.

There is a closed, connected subset =(b)\mathscr{F}=\mathscr{F}(b) of 𝒜(b)\mathscr{A}(b) such that Mx(M)M\mapsto x(M) is a proper, surjective map from \mathscr{F} onto (0,)(0,\infty).

The behavior as necksize tends to 0 or to infinity is described by the following theorem. (See Theorems 15.6 and 16.1.)

Theorem 1.4.

Suppose that Mn𝒜M_{n}\in\mathscr{A} and that b(Mn)b<b(M_{n})\to b<\infty.

  1. (1)

    If x(Mn)0x(M_{n})\to 0, then MnM_{n} converges with multiplicity 22 to the translating graph

    fb:𝐑×(b,b)𝐑f_{b}:\mathbf{R}\times(-b,b)\to\mathbf{R}

    with fb(0,0)=0f_{b}(0,0)=0 and Dfb(0,0)=0Df_{b}(0,0)=0. The convergence is smooth away from the origin. After suitable rescaling, the surfaces converge to a catenoid.

  2. (2)

    If x(Mn)x(M_{n})\to\infty, then Mn(0,0,ζn)M_{n}-(0,0,\zeta_{n}) converges smoothly to a pair of grim reaper surfaces, one over 𝐑×(b,b+π)\mathbf{R}\times(b,b+\pi) and the other over 𝐑×((b+π),b)\mathbf{R}\times(-(b+\pi),-b). Here

    ζn:=max{z:(0,y,z)Mn}.\zeta_{n}:=\max\{z:(0,y,z)\in M_{n}\}.

By the classification of graphical translators (Theorem 2.1), the graph of fbf_{b} in Assertion (1) is an untilted grim reaper surface if b=π/2b=\pi/2 and a Δ\Delta-wing if b>π/2b>\pi/2.

It is natural to ask what happens if we fix necksize and let the inner width tend to infinity.

Conjecture 1.5.

Suppose that MnM_{n} are annuloids in 𝒜\mathscr{A} such that x(Mn)=Rx(M_{n})=R and b(Mn)b(M_{n})\to\infty. Then the MnM_{n} converge to the rotationally symmetric translating annulus whose neck is a horizontal circle of radius RR.

In Conjecture 1.5, if R>0R>0 is small, then, by Theorem 1.4, MnM_{n} resembles two copies of graph(fbn)\operatorname{graph}(f_{b_{n}}) joined by a neck. The fact that graph(fbn)\operatorname{graph}(f_{b_{n}}) converges smoothly to a rotationally symmetric surface (a bowl soliton) makes Conjecture 1.5 plausible.

Behavior away from the zz-axis

Let M𝒜M\in\mathscr{A}. We know that as zz\to-\infty, the surface looks like 44 planes, and as zz\to\infty, it looks like the empty set.

What does it look like as xx\to\infty? (The symmetric behavior occurs as xx\to-\infty.) We show that M{x>x(M)}M\cap\{x>x(M)\} has two connected components, MupperM^{\textnormal{upper}} and MlowerM^{\textnormal{lower}}. For large xx, MlowerM^{\textnormal{lower}} looks like a downward-tilted grim reaper surface over 𝐑×(b,b)\mathbf{R}\times(-b,b).

Refer to caption
Refer to caption
Figure 4. A capped example (left) and an uncapped one (right).

Also, for large xx, MupperM^{\textnormal{upper}} looks like a tilted grim reaper surface GG over 𝐑×(B,B)\mathbf{R}\times(-B,B). But (assuming B>π/2B>\pi/2) there are two kinds of MM:

  1. (1)

    For some MM, GG is a downward-tilted grim reaper surface over 𝐑×(B,B)\mathbf{R}\times(-B,B). In this case, we say that MM is capped.

  2. (2)

    For other MM, GG is an upward-tilted grim reaper surface over 𝐑×(B,B)\mathbf{R}\times(-B,B). In this case, we say that MM is uncapped.

Equivalently (assuming B>π/2B>\pi/2), MM is capped if and only z()|Mz(\cdot)|M is bounded, and MM is uncapped if and only if z()|Mz(\cdot)|M is unbounded. See §22.

If B(M)>b(M)B(M)>b(M), then MM is uncapped (by Theorem 14.1(4)). However, the converse is not true: there are examples of uncapped MM with b(M)=B(M)b(M)=B(M). See Corollary 22.7.

The following theorem describes the relation between necksize and capping (see Theorem 22.5) when b>π/2b>\pi/2.

Theorem 1.6.

Suppose b>π/2b>\pi/2. There are constants 0<cd<0<c\leq d<\infty (depending on bb) with the following properties:

  1. (1)

    If M𝒜(b)M\in\mathscr{A}(b) and x(M)<cx(M)<c, then MM is capped (and thefore b(M)=B(M)b(M)=B(M)).

  2. (2)

    If M𝒜(b)M\in\mathscr{A}(b) and x(M)dx(M)\geq d, then b(M)<B(M)b(M)<B(M) (and therefore MM is uncapped).

In the case b=π/2b=\pi/2, there is a d<d<\infty for which (2) holds. But (in that case) we do not know whether B(M)=π/2B(M)=\pi/2 for all sufficiently small neck sizes. Indeed, we do not know if there are any annuloids with B(M)=π/2B(M)=\pi/2. We do know that for every ϵ>0\epsilon>0, there is a c=c(ϵ)c=c(\epsilon) such that if M𝒜(π/2)M\in\mathscr{A}(\pi/2) and if x(M)<cx(M)<c, then B(M)<π/2+ϵB(M)<\pi/2+\epsilon. See Theorems 16.1(5) and 22.5.

Prongs

Consider an M𝒜(b)M\in\mathscr{A}(b) with x(M)x(M) very large. Then (according to Theorem 1.4), near ZZ (the zz-axis), MM looks like a pair of grim reaper surfaces. What does it look like away from ZZ? In particular, what does it look like in the region where xx(M)x\sim x(M)? That is perhaps the most interesting region, since x(M)x(M) is the value of tt for which the topology of M{|x|t}M\cap\{|x|\leq t\} undergoes a change.

To answer that question, consider a sequence Mn𝒜(b)M_{n}\in\mathscr{A}(b) with x(Mn)x(M_{n})\to\infty. We show that the surfaces

M(x(Mn),0,0)M-(x(M_{n}),0,0)

converge smoothly, perhaps after passing to a subsequence, to a simply connected translator Σ\Sigma that we call a prong.

Refer to caption
Refer to caption
Figure 5. Two different views of a prong.

As zz\to\infty, Σ+(0,0,z)\Sigma+(0,0,z) converges to the planes {y=±b}\{y=\pm b\} and {y=±B}\{y=\pm B\}, where B=b+πB=b+\pi.

The portion of Σ\Sigma with {x>0}\{x>0\} behaves just like an uncapped annuloid. In particular, Σ{x>0}\Sigma\cap\{x>0\} has two components, Σupper\Sigma^{\textnormal{upper}} and Σlower\Sigma^{\textnormal{lower}}. Where xx is large, Σupper\Sigma^{\textnormal{upper}} looks like an upward-tilted grim reaper surface over 𝐑×(bπ,b+π)\mathbf{R}\times(-b-\pi,b+\pi) and Σlower\Sigma^{\textnormal{lower}} looks like a downward-tilted grim reaper surface over 𝐑×(b,b)\mathbf{R}\times(-b,b).

Where xx is very negative, Σ\Sigma looks like a pair of untilted grim reaper surfaces over 𝐑×(bπ,b)\mathbf{R}\times(-b-\pi,-b) and 𝐑×(b,b+π)\mathbf{R}\times(b,b+\pi).

The prong Σ\Sigma projects diffeomorphically (under (x,y,z)(y,z)(x,y,z)\mapsto(y,z)) onto an open subset of the yzyz-plane. Thus it is the sideways graph of a function x=x(y,z)x=x(y,z).

Prongs are discussed in Section 15. The x=x(y,z)x=x(y,z) sideways graph property is proved in Section 23.

Singularity Models

All of the examples of complete translators produced in this paper have finite topology and finite entropy, and thus could conceivably arise by blowing up (at a singularity) the mean curvature flow of an initially smooth, closed surface. If the closed surface is embedded, then, by recent work of Bamler and Kleiner [bamler-kleiner], the bowl soliton is the only translator (other than a vertical plane) that can arise as a blow-up. Whether any of the new examples in this paper can arise as blow-ups of immersed surfaces is an open question that seems very challenging.

The paper [annuloid-survey] discusses how the construction of annuloids is related to the way Δ\Delta-wings arise as limits of compact translating graphs.

2. Graphical Translators

As mentioned in the introduction, graphical translators have been completely classified. In this section, we describe the classification. See also Figure 1. For more information about graphical translators, see the survey article [himw-survey].

There is a unique entire translating graph, the bowl soliton. It is a surface of revolution. Here, and below, ‘unique’ means unique up to translations in 𝐑3\mathbf{R}^{3} and rotation about a vertical axis.

Every other properly embedded translating graph is defined on a strip of width 2bπ2b\geq\pi. We may take the strip to be 𝐑×(b,b)\mathbf{R}\times(-b,b). As zz\rightarrow-\infty, the surface is asymptotic to the planes {y=±b}\{y=\pm b\}. Up to the Euclidean motions mentioned above, here is a complete list of the graphical translators defined on strips.

  1. (1)

    For b=π/2b=\pi/2, the untilted grim reaper surface is the graph

    u:𝐑×(π/2,π/2)𝐑,\displaystyle u:\mathbf{R}\times(-\pi/2,\pi/2)\rightarrow\mathbf{R},
    u(x,y)=log(cosy).\displaystyle u(x,y)=\log(\cos y).

    The grim reaper surface uu is ruled by horizontal lines. It is symmetric with respect to reflection in the vertical coordinate planes.

  2. (2)

    For b>π/2b>\pi/2, the tilted grim reaper surface

    ub:𝐑×(b,b)𝐑,u_{b}:\mathbf{R}\times(-b,b)\rightarrow\mathbf{R},

    is produced from the untilted grim reaper surface by dilation followed by rotation around the yy-axis. Let σb=2bπ\sigma_{b}=\frac{2b}{\pi} and s(b)=σb21s(b)=\sqrt{\sigma_{b}^{2}-1}. Then

    ub(x,y)=σb2log(cos(y/σb))+s(b)x.u_{b}(x,y)=\sigma_{b}^{2}\log(\cos(y/\sigma_{b}))+s(b)x.

    The angle of rotation is θb=arctans(b)\theta_{b}=\arctan s(b), and the dilation is by a factor of σb\sigma_{b}. The tilted grim reapers are ruled by horizontal lines making an angle of θb\theta_{b} with the horizontal plane. Note that ub(x,y)ub(x,y)u_{b}(x,y)\equiv u_{b}(x,-y). Note also that ubu_{b} is the unique translator on 𝐑×(b,b)\mathbf{R}\times(-b,b) such that

    ub(0,0)=0andubxs(b).u_{b}(0,0)=0\,\,\mbox{and}\,\,\frac{\partial u_{b}}{\partial x}\equiv s(b).

    When b=π/2b=\pi/2, uπ/2u_{\pi/2} is the untilted grim reaper surface. We let wb(x,y)=ub(x,y)w_{b}(x,y)=u_{b}(-x,y). Thus wbw_{b} is the unique translator on 𝐑×(b,b)\mathbf{R}\times(-b,b) such that

    wb(0,0)=0andwbxs(b).w_{b}(0,0)=0\quad\text{and}\quad\frac{\partial w_{b}}{\partial x}\equiv-s(b).

    The grim reaper surfaces (tilted and untilted) are the complete translating graphs that are intrinsically flat.

  3. (3)

    For each bπ/2b\geq\pi/2, there is a unique graphical translator

    fb:𝐑×(b,b)𝐑f_{b}:\mathbf{R}\times(-b,b)\to\mathbf{R}

    such that

    fb(0,0)=0andDfb(0,0)=0.f_{b}(0,0)=0\quad\text{and}\quad Df_{b}(0,0)=0.

    When b=π/2b=\pi/2, fπ/2f_{\pi/2} is the untilted grim reaper surface. When b>π/2b>\pi/2, D2fbD^{2}f_{b} is everywhere negative definite. In this case, graph(fb)\operatorname{graph}(f_{b}) is called a Δ\Delta-wing. It is invariant under reflection in the vertical coordinate planes: fb(x,y)=fb(x,y)=fb(x,y)f_{b}(x,-y)=f_{b}(x,y)=f_{b}(-x,y). The function fbf_{b} attains its maximum value 0 at the origin. The Δ\Delta-wings have strictly positive curvature, as does the bowl soliton.

Theorem 2.1.

[graphs]*Theorem 7.1 Every complete graphical translator is a grim reaper surface, a Δ\Delta-wing, or a bowl soliton.

The Δ\Delta-wings and grim reaper surfaces are related as follows. The function

fb(x+t,y)fb(t,0)f_{b}(x+t,y)-f_{b}(t,0)

converges to ub(x,y)u_{b}(x,y) as tt\to-\infty and to wb(x,y)w_{b}(x,y) as t+t\to+\infty.

As bb\rightarrow\infty, the tilted grim reaper surfaces converge to the vertical plane {x=0}\{x=0\} and the Δ\Delta-wings converge to the bowl soliton.

3. Bounds on Area and on Curvature

The surfaces constructed in this paper will all be either compact translators bounded by a pair of nested convex curves in a horizontal plane or limits of such compact translators. The following theorem gives area and curvature bounds that hold for all such surfaces. Here ZZ is the zz-axis, and ρZ\rho_{Z} is rotation by π\pi around ZZ.

The following lemma for translators is an analog of the convex hull property for minimal surfaces in Euclidean space.

Lemma 3.1.

Suppose MM is a compact translator. If 𝐯\mathbf{v} is a horizontal vector and if F𝐯(p):=𝐯pF_{\mathbf{v}}(p):=\mathbf{v}\cdot p, then

minMF𝐯\displaystyle\min_{M}F_{\mathbf{v}} =minMF𝐯,\displaystyle=\min_{\partial M}F_{\mathbf{v}},
maxMF𝐯\displaystyle\max_{M}F_{\mathbf{v}} =maxMF𝐯.\displaystyle=\max_{\partial M}F_{\mathbf{v}}.

If z=ϕ(x,y)z=\phi(x,y) is the equation of a bowl soliton and if Φ(x,y,z)=zϕ(x,y)\Phi(x,y,z)=z-\phi(x,y), then

minMΦ\displaystyle\min_{M}\Phi =minMΦ,\displaystyle=\min_{\partial M}\Phi,
maxMΦ\displaystyle\max_{M}\Phi =maxMΦ.\displaystyle=\max_{\partial M}\Phi.
Proof.

The lemma follows immediately from the maximum principle. ∎

Corollary 3.2.

Suppose that MiM_{i} is a sequence of compact translators.

  1. (1)

    If the Mi\partial M_{i} lie in a bounded subset of 𝐑3\mathbf{R}^{3}, then the MiM_{i} lie in a bounded subset of 𝐑3\mathbf{R}^{3}.

  2. (2)

    If λi\lambda_{i}\to\infty and if the dilated boundaries λi(Mi)\lambda_{i}(\partial M_{i}) lie in a bounded subset of 𝐑3\mathbf{R}^{3}, then the surfaces λiMi\lambda_{i}M_{i} lie in a bounded subdset of 𝐑3\mathbf{R}^{3}.

Proposition 3.3.

Suppose that MiM_{i} is a sequence of simply connected translators.

  1. (1)

    If MiM_{i} converges smoothly to an embedded surface MM, then MM is simply connected.

  2. (2)

    If λi\lambda_{i}\to\infty and if λiMi\lambda_{i}M_{i} converges to an embedded surface MM, then MM is simply connected.

Proof.

To prove (1), let CC be a simple closed curve in MM. Then CC is the smooth limit of simple closed curves CiC_{i} in MiM_{i}. Since MiM_{i} is simply connected, CiC_{i} bounds a disk DiD_{i} in MiM_{i}. By Corollary 3.2, the DiD_{i} lie in a bounded subset of 𝐑3\mathbf{R}^{3}. Thus the DiD_{i} converge smoothly to a disk DD in MM with D=C\partial D=C. The proof of (2) is essentially the same. ∎

Theorem 3.4.

There are finite constants c1c_{1} and c2c_{2} with the following properties. Suppose that MM is a compact, embedded translator bounded by a pair of convex curves in horizontal planes. More generally, suppose that MM is a smooth limit of such surfaces. For pMp\in M and r0r\geq 0, let

M(p,r):={qM:distM(q,p)r}M(p,r):=\{q\in M:\operatorname{dist}_{M}(q,p)\leq r\}

and let R(M,p)R(M,p) be the infimum of r>0r>0 such that M(p,r)M(p,r) contains a homotopically non-trivial curve in MM. Then

  1. (1)

    area(M𝐁(p,r))c1r2\operatorname{area}(M\cap\mathbf{B}(p,r))\leq c_{1}r^{2} for all balls 𝐁(p,r)\mathbf{B}(p,r).

  2. (2)

    The 2nd fundamental form satisfies

    |A(M,p)|min{1,R(M,p),distM(p,M)}c2,|A(M,p)|\,\min\{1,R(M,p),\operatorname{dist}_{M}(p,\partial M)\}\leq c_{2},

    where distM\operatorname{dist}_{M} denotes geodesic distance in MM.

  3. (3)

    If, in addition, MM is an annulus disjoint from ZZ and invariant under ρZ\rho_{Z},

    R(M,p)\displaystyle R(M,p) dist(p,Z), and\displaystyle\geq\operatorname{dist}(p,Z)\text{, and}
    R(M,p)\displaystyle R(M,p) x(M):=dist(Z,M{y=0}),\displaystyle\geq x(M):=\operatorname{dist}(Z,M\cap\{y=0\}),

    and thus

    |A(M,p)|min{1,dist(p,ZM)}\displaystyle|A(M,p)|\,\min\{1,\operatorname{dist}(p,Z\cup\partial M)\} c2,\displaystyle\leq c_{2},
    |A(M,p)|min{1,x(M),distM(p,M)}\displaystyle|A(M,p)|\,\min\{1,x(M),\operatorname{dist}_{M}(p,\partial M)\} c2.\displaystyle\leq c_{2}.
Proof.

It suffices to prove Assertions (1) and (2) for compact MM, since the general case follows trivially.

The existence of c1c_{1} is proved in [white-entropy].

Suppose that Assertion (2) fails. Then there exist compact examples MiM_{i}, points piMip_{i}\in M_{i}, and radii rir_{i} such that

0<ri<min{1,R(Mi,pi),dist(pi,Mi)}0<r_{i}<\min\{1,R(M_{i},p_{i}),\operatorname{dist}(p_{i},\partial M_{i})\}

and

|A(Mi,pi)|ri|A(M_{i},p_{i})|\,r_{i}\to\infty

Note that Mi(pi,ri)M_{i}(p_{i},r_{i}) is a compact subset of MiMiM_{i}\setminus\partial M_{i}. Let qiq_{i} be a point in Mi(pi,ri)M_{i}(p_{i},r_{i}) that maximizes

fi(q):=|A(Mi,q)|(ridist(q,pi)).f_{i}(q):=|A(M_{i},q)|\,(r_{i}-\operatorname{dist}(q,p_{i})).

Then fi(qi)f_{i}(q_{i})\to\infty since fi(pi)f_{i}(p_{i})\to\infty.

Let ρi=ridist(pi,qi)\rho_{i}=r_{i}-\operatorname{dist}(p_{i},q_{i}). Then

|A(Mi,)|2|A(Mi,qi)||A(M_{i},\cdot)|\leq 2|A(M_{i},q_{i})|

on Mi(qi,ρi/2)M_{i}(q_{i},\rho_{i}/2).

Now dilate MiqiM_{i}-q_{i} by |A(Mi,qi)||A(M_{i},q_{i})| to get MiM_{i}^{\prime}. After passing to a subsequence, the MiM_{i}^{\prime} converge smoothly to a MM^{\prime} such that MM^{\prime} is a smoothly embedded minimal surface and

|A(M,0)|\displaystyle|A(M^{\prime},0)| =1,\displaystyle=1,
dist(0,M)\displaystyle\operatorname{dist}(0,\partial M^{\prime}) =,\displaystyle=\infty,
R(M,0)\displaystyle R(M^{\prime},0) =.\displaystyle=\infty.

Thus MM^{\prime} is complete, properly embedded (it is the limit of embedded surfaces) and simply connected (by Proposition 3.3). By Assertion (1) of this theorem, MM^{\prime} has quadratic area growth. Thus it is a plane, a contradiction, because |A(M,0)|=1|A(M^{\prime},0)|=1. Hence Assertion (2) is proved.

To prove Assertion (3), let r>R(M,p)r>R(M,p). Then there is a simple closed curve CC in M(p,r)M(p,r) that is homotopically nontrivial in MM. By elementary topology (see Lemma 3.5), CC is homotopically nontrivial in 𝐑3Z\mathbf{R}^{3}\setminus Z. Let q+q^{+} be a point in C{y=0,x>0}C\cap\{y=0,\,x>0\} and qq^{-} be a point in C{y=0,x<0}C\cap\{y=0,\,x<0\}. Then x(q+)x(M)x(q^{+})\geq x(M) and x(q)x(M)x(q^{-})\leq-x(M), so

2x(M)\displaystyle 2x(M) |q+q|\displaystyle\leq|q^{+}-q^{-}|
distM(q+,p)+distM(p,q)\displaystyle\leq\operatorname{dist}_{M}(q^{+},p)+\operatorname{dist}_{M}(p,q^{-})
2r.\displaystyle\leq 2r.

Thus,

(1) x(M)r.x(M)\leq r.

Since

CM(p,r)𝐁(p,r)¯,C\subset M(p,r)\subset\overline{\mathbf{B}(p,r)},

and since CC is homotopically nontrivial in 𝐑3𝐙\mathbf{R}^{3}\setminus\mathbf{Z}, it follows that 𝐁(p,r)Z\mathbf{B}(p,r)\cap Z is nonempty, so

(2) dist(p,Z)r.\operatorname{dist}(p,Z)\leq r.

The inequalities (1) and (2) hold for all r>R(M,p)r>R(M,p). Hence they hold for r=R(M,p)r=R(M,p). ∎

Lemma 3.5 (Topology Lemma).

Suppose that MM is an annulus in 𝐑3Z\mathbf{R}^{3}\setminus Z that is invariant under ρZ\rho_{Z}. Then the inclusion of MM into 𝐑3Z\mathbf{R}^{3}\setminus Z induces a monomorphism of first homology.

If, in addition, the annulus MM is a translator and if VV is a vertical plane, then MVM\cap V does not contain a closed curve.

Proof.

Since H1(M;𝐙)=H1(𝐑3Z;𝐙)=𝐙H_{1}(M;\mathbf{Z})=H_{1}(\mathbf{R}^{3}\setminus Z;\mathbf{Z})=\mathbf{Z}, It suffices to prove that MM contains a closed curve that is homotopically nontrivial in 𝐑3Z\mathbf{R}^{3}\setminus Z. Let pMp\in M and γ\gamma be a oriented curve in MM from pp to ρZp\rho_{Z}p. Then

γ𝑑θ\int_{\gamma}\,d\theta

is an odd multiple of π\pi, where

dθ:=xdyydxx2+y2.d\theta:=\frac{x\,dy-y\,dx}{x^{2}+y^{2}}.

If C=γρZγC=\gamma\cup\rho_{Z}\gamma, then

C𝑑θ=γ𝑑θ+ρZ𝑑θ=2γ𝑑θ.\int_{C}d\theta=\int_{\gamma}d\theta+\int_{\rho_{Z}}d\theta=2\int_{\gamma}d\theta.

Thus Cθ\int_{C}\theta is an odd multiple of 2π2\pi, and therefore CC is homotopically nontrivial in 𝐑3𝐙\mathbf{R}^{3}\setminus\mathbf{Z}.

To prove the second statement, note that a closed curve in VV disjoint from ZZ is homotopically trivial in 𝐑3Z\mathbf{R}^{3}\setminus Z. (Indeed, it is homotopically trivial in VV if ZVZ\not\subset V and in VZV\setminus Z if ZVZ\subset V.) If MVM\cap V contained a closed curve SS, it would be homotopically trival in 𝐑Z\mathbf{R}\setminus Z, and therefore SS would bound a disk in MM. By the maximum principle, that disk would lie in VV, which is impossible. (By unique continuation, all of MM would lie in VV.) ∎

4. Morse-Radó Theory

Let MM be a translator. There are a number of standard foliations \mathscr{F} of 𝐑3\mathbf{R}^{3} or of open subsets WW of 𝐑3\mathbf{R}^{3} by translators. For example, \mathscr{F} could be a family of parallel vertical planes. Bounding the number of points of tangency of MM with the leaves of \mathscr{F} is the subject of Morse-Radó Theory, a powerful tool that plays a major role in this paper. In this section, we recall the basic facts in Morse-Radó Theory. More details about the results included in this section can be found in [morse-rado].

Recall [ilmanen_1994] that M𝐑3M\subset\mathbf{R}^{3} is a translator if and only if it is minimal with respect to the translator metric

(3) g=ez(dx2+dy2+dz2).g=e^{-z}(dx^{2}+dy^{2}+dz^{2}).
Definition 4.1.

Let MM be an embedded or immersed minimal surface in a Riemannian 33-manifold NN. Let \mathscr{F} be a foliation of an open subset WW of NN by minimal surfaces. Let M^\hat{M} be the union of the components of MWM\cap W that are not contained in leaves of \mathscr{F}. A critical point of MM with respect to \mathscr{F} is an interior point pp of M^\hat{M} at which MM is tangent to the leaf of \mathscr{F} through pp. The multiplicity of the critical point is the order of contact of MM and the leaf. We let

𝖭(|M)\mathsf{N}(\mathscr{F}|M)

be the total number of interior critical points, counting multiplicity.

Of course the case of interest in this paper is when NN is 𝐑3\mathbf{R}^{3} with the translator metric.

The first important fact about 𝖭(|M)\mathsf{N}(\mathscr{F}|M) is that it depends lower semicontinuously on \mathscr{F} and on MM:

Theorem 4.2 ([morse-rado], Corollary 40).

Suppose that gig_{i} are Riemannian metrics on a 33-manifold NN that converge smoothly to a Riemannian metric gg. Suppose that MiM_{i} are gig_{i}-minimal surfaces that converges smoothly to a gg-minimal surface MM. Suppose i\mathscr{F}_{i} are gig_{i}-minimal foliations of open subsets WiW_{i} of NN such that the leaves of i\mathscr{F}_{i} converge smoothly to the leaves of a gg-minimal foliation \mathscr{F} of an open subset WW of NN. Then

𝖭(|M)lim inf𝖭(i|Mi).\mathsf{N}(\mathscr{F}|M)\leq\liminf\mathsf{N}(\mathscr{F}_{i}|M_{i}).

In particular, if pp is a critical point of (,M)(\mathscr{F},M), then pp is a limit of critical points pip_{i} of (i,Mi)(\mathscr{F}_{i},M_{i}).

A minimal foliation function on NN is a continuous function FF from an open subset of WW of NN to an open interval I𝐑I\subset\mathbf{R} such that for each tIt\in I,

  1. F1(t)F^{-1}(t) is a minimal surface, and

  2. F1(t)F^{-1}(t) is in the closures of {F>t}\{F>t\} and of {F<t}\{F<t\}.

If FF is a minimal foliation function on NN, we let 𝖭(F|M)=𝖭(|M)\mathsf{N}(F|M)=\mathsf{N}(\mathscr{F}|M), where \mathscr{F} is the foliation whose leaves are the level sets of FF.

In the following theorem, if SS is a set, then |S||S| denotes the number of elements in the set.

Theorem 4.3 ([morse-rado], Theorem 4).

Let F:WNIF:W\subset N\to I be a minimal foliation function on an open subset WW of a Riemannian 33-manifold NN. Let MM be a minimal surface in NN homeomorphic to a 22-manifold-with-boundary. Suppose that F|MWF|M\cap W is proper, that the set QQ of local minima of F|MF|\partial M is finite, that MWM\cap W has finite genus, and that (M){F<t}(\partial M)\cap\{F<t\} is empty for some tIt\in I. Then

𝖭(F|M)|Q||A|χ(MW),\mathsf{N}(F|M)\leq|Q|-|A|-\chi(M\cap W),

where AA is the set of local maxima or minima of F|MF|\partial M that are not local maxima or minima of F|MF|M, and where χ()\chi(\cdot) denotes Euler Characteristic.

Equivalently,

𝖭(F|M)|S||T|χ(MW),\mathsf{N}(F|M)\leq|S|-|T|-\chi(M\cap W),

where SS is the set of local minima of F|MF|\partial M that are also local minima of F|MF|M, and where TT is the set of local maxima of F|MF|\partial M that are not local maxima of F|MF|M.

Remark 4.4.

In practice, one sometimes encounters FF and MM that satisfy all but one of the hypotheses of Theorem 4.3, namely the hypothesis that the set of local minima of F|MF|\partial M is finite. In particular, that hypothesis will fail if FF is constant on one or more arcs of M\partial M. One can handle such examples as follows. Suppose FF is not constant on any connected component of M\partial M. Let M~\tilde{M} be obtained from MM by identifying each arc of M\partial M on which FF is constant to a point. Let F~\tilde{F} be the function on M~\tilde{M} corresponding to FF on MM. If the set Q~\tilde{Q} of local minima of F~|M~\tilde{F}|\partial\tilde{M} is finite, then

𝖭(F|M)\displaystyle\mathsf{N}(F|M) |Q~||A~|χ(MW),\displaystyle\leq|\tilde{Q}|-|\tilde{A}|-\chi(M\cap W),

where A~\tilde{A} is the set of local minima and local maxima of F~|M~\tilde{F}|\partial\tilde{M} that are not local minima or local maxima of F~|M\tilde{F}|M. Equivalently,

𝖭(F|M)|S~||T~|χ(MW),\mathsf{N}(F|M)\leq|\tilde{S}|-|\tilde{T}|-\chi(M\cap W),

where S~\tilde{S} is the set of local minima of F~|M~\tilde{F}|\partial\tilde{M} that are also local minima of F~|M~\tilde{F}|\tilde{M}, and T~\tilde{T} is the set of local maxima of F~|M~\tilde{F}|\partial\tilde{M} that are not local maxima of F~|M~\tilde{F}|\tilde{M}.

We now describe the main examples of gg-minimal foliation functions that we will use (where gg is the translator metric). First, if 𝐯\mathbf{v} is a horizontal unit vector in 𝐑3\mathbf{R}^{3}, then the function

(4) F𝐯:𝐑3𝐑,\displaystyle F_{\mathbf{v}}:\mathbf{R}^{3}\to\mathbf{R},
F𝐯(p)=𝐯p\displaystyle F_{\mathbf{v}}(p)=\mathbf{v}\cdot p

is a gg-minimal foliation function. Second, suppose that UU is 𝐑2\mathbf{R}^{2} or an open strip in 𝐑2\mathbf{R}^{2} and that h:U𝐑h:U\to\mathbf{R} is a function whose graph is a complete translator. Then

(5) H:U×𝐑𝐑,\displaystyle H:U\times\mathbf{R}\to\mathbf{R},
H(x,y,z)=zh(x,y)\displaystyle H(x,y,z)=z-h(x,y)

is a gg-minimal foliation function.

The complete translators MM that we construct in this paper all have the following properties (see Theorem 7.3):

  1. (i)

    For each horizontal unit vector 𝐯\mathbf{v}, 𝖭(F𝐯|M)2\mathsf{N}(F_{\mathbf{v}}|M)\leq 2.

  2. (ii)

    If HH is as in (5), then 𝖭(H|M)8\mathsf{N}(H|M)\leq 8.

(It is not hard to show that 𝖭(H|M)4\mathsf{N}(H|M)\leq 4 in (ii), but we do not need that fact.)

In the next section, we present a few useful general facts about translators that satisfy bounds such as (i) and (ii).

5. Translators of Finite Type

As mentioned in Section 4, translators can be seen as minimal surfaces in 𝐑3\mathbf{R}^{3} endowed with the metric g=ez(dx2+dy2+dz2).g={\rm e}^{-z}(dx^{2}+dy^{2}+dz^{2}). For minimal surfaces in Euclidean 33-space, there is a special class of surfaces that have very interesting properties not shared by general minimal surfaces: the surfaces with finite total curvature. In particular, for each such surface MM, there is a positive integer kk such that

𝖭(|M)k,\mathsf{N}(\mathscr{F}|M)\leq k,

for any foliation \mathscr{F} of 𝐑3\mathbf{R}^{3} by minimal surfaces (i.e., for any foliation by parallel planes). This motivates our definition of translator of finite type, which plays a similar role in this setting.

Definition 5.1.

We say that a translator MM in 𝐑3\mathbf{R}^{3} is of finite type provided there are finite numbers cc, KK, and kk such that

  1. (1)

    area(M𝐁)cr2\operatorname{area}(M\cap\mathbf{B})\leq cr^{2} for every ball 𝐁\mathbf{B} of radius rr.

  2. (2)

    For every pMp\in M,

    |A(M,p)|min{1,dist(p,M)}K.|A(M,p)|\,\min\{1,\operatorname{dist}(p,\partial M)\}\leq K.
  3. (3)

    𝖭(F𝐯|M)k\mathsf{N}(F_{\mathbf{v}}|M)\leq k for each horizontal unit vector 𝐯\mathbf{v}.

  4. (4)

    𝖭(H|M)k\mathsf{N}(H|M)\leq k for every function H(x,y,z)=zh(x,y)H(x,y,z)=z-h(x,y) whose level sets are grim reaper surfaces (tilted or untilted).

We do not require any regularity at M\partial M.

(Actually, condition (3) is redundant: it is implied by condition (4). See Remark 5.10.)

This paper studies translators that are compact minimal annuli bounded by pairs of nested rectangles, as well as translating annuli without boundary obtained as limits of such compact examples. All such translators are of finite type. (See Theorems 6.4, 6.6, and 7.3.)

Remark 5.2.

Condition (1) in Definition 5.1 is equivalent (by a simple calculation) to the condition that MM have finite entropy. See [white21]*Theorem 9.1.

Remark 5.3.

Let \mathscr{H} be the collection of all functions HH as in condition (4) of Definition 5.1. By lower semicontinuity (Theorem 4.2), to prove 𝖭(H|M)k\mathsf{N}(H|M)\leq k for all HH\in\mathscr{H}, it suffices to prove it for a dense set of HH\in\mathscr{H}. Likewise, to prove 𝖭(F𝐯|M)k\mathsf{N}(F_{\mathbf{v}}|M)\leq k for all horizontal unit vectors 𝐯\mathbf{v}, it suffices to prove it for a dense set of such 𝐯\mathbf{v}.

We are mainly interested in translators without boundary. For a surface without boundary, dist(p,M)=\operatorname{dist}(p,\partial M)=\infty, so the curvature bound in (2) of Definition 5.1 simplifies to

|A(M,p)|K|A(M,p)|\leq K

for all pMp\in M.

Translators of finite type have a nice compactness property. If MnM_{n} is a sequence of such surfaces satisfying (1)–(4) in Definition 5.1, with cc, kk, and KK independent of nn, then, after passing to a subsequence, the MnM_{n} will converge smoothly to a limit translator MM satisfying the bounds (1) and (2). (The convergence is smooth away from M\partial M.) By lower semicontinuity (Theorem 4.2), MM will also satisfy the bounds (3) and (4).

Theorem 5.4.

Suppose that M𝐑3M\subset\mathbf{R}^{3} is a translator of finite type. If pip_{i} is a divergent sequence and if dist(pi,M)\operatorname{dist}(p_{i},\partial M)\to\infty, then, after passing to a subsequence, MpiM-p_{i} converges smoothly to a limit surface MM^{\prime}. Furthermore, any such limit MM^{\prime} is a union of vertical planes and translating graphs (grim reaper surfaces, Δ\Delta-wings, and bowl solitons).

If the pip_{i} lie in a vertical plane containing ZZ, then MM^{\prime} is a union of vertical planes and grim reaper surfaces.

If the pi=(0,0,zi)p_{i}=(0,0,z_{i}) are in ZZ, then MM^{\prime} is a union of vertical planes.

Of course, for a surface without boundary, the condition dist(pi,M)\operatorname{dist}(p_{i},\partial M)\to\infty is vacuously true.

Proof.

We may assume that MM is connected. We may also assume that MM is not a vertical plane (as the theorem is trivial in that case). The curvature and area bounds imply smooth subsequential convergence to a limit translator MM^{\prime} that is properly immersed and without boundary. Let 𝐯\mathbf{v} be a horizontal unit vector. By hypothesis, F𝐯|MF_{\mathbf{v}}|M has a finite set VV of critical points. Since pip_{i} diverges in 𝐑3\mathbf{R}^{3}, the surfaces

Mi:=(MV)piM_{i}:=(M\setminus V)-p_{i}

converge smoothly to the same limit surface MM^{\prime}. By lower semicontinuity,

𝖭(F𝐯|M)lim inf𝖭(F𝐯|Mi).\mathsf{N}(F_{\mathbf{v}}|M^{\prime})\leq\liminf\mathsf{N}(F_{\mathbf{v}}|M_{i}).

But

𝖭(F𝐯|Mi)=𝖭(F𝐯|(MV))=0.\mathsf{N}(F_{\mathbf{v}}|M_{i})=\mathsf{N}(F_{\mathbf{v}}|(M\setminus V))=0.

Thus

(6) 𝖭(F𝐯|M)=0.\mathsf{N}(F_{\mathbf{v}}|M^{\prime})=0.

Let Σ\Sigma be a component of MM^{\prime} that is not a vertical plane. By (6) (which holds for all horizontal unit vectors 𝐯\mathbf{v}), the surface Σ\Sigma has no points at which the tangent plane is vertical. According to [spruck-xiao]*Corollary 1.2, any complete connected translator with no vertical tangent planes is a translating graph. By Theorem 2.1, a translating graph is a grim reaper surface, a Δ\Delta-wing, or a bowl soliton.

Now suppose that the pip_{i} lie in a vertical plane containing ZZ. By rotating, we may assume that the plane is the plane {y=0}\{y=0\}, and thus that pi=(xi,0,zi)p_{i}=(x_{i},0,z_{i}).

For c𝐑c\in\mathbf{R}, let

hc(x,y)=log(cos(yc))h_{c}(x,y)=\log(\cos(y-c))

be the untilted grim reaper surface over 𝐑×(cπ/2,c+π/2)\mathbf{R}\times(c-\pi/2,c+\pi/2), and let

Hc:𝐑×(cπ/2,c+π/2)×𝐑𝐑,\displaystyle H_{c}:\mathbf{R}\times(c-\pi/2,c+\pi/2)\times\mathbf{R}\to\mathbf{R},
Hc(x,y,z)=zhc(x,y).\displaystyle H_{c}(x,y,z)=z-h_{c}(x,y).

Note that Hc|MH_{c}|M has a finite set CC of critical points. Then Mi:=(MiC)piM_{i}^{\prime}:=(M_{i}\setminus C)-p_{i} converges to MM^{\prime} and 𝖭(Hc|Mi)=0\mathsf{N}(H_{c}|M_{i}^{\prime})=0, so

(7) 𝖭(Hc|M)=0.\mathsf{N}(H_{c}|M^{\prime})=0.

Now suppose that Σ\Sigma is a bowl soliton or a Δ\Delta-wing. Then z()|Σz(\cdot)|\Sigma attains its maximum at a single point p:=(x¯,y¯,z¯)p:=(\bar{x},\bar{y},\bar{z}). Note that pp is a critical point of Hy¯|ΣH_{\bar{y}}|\Sigma. Therefore Σ\Sigma cannot be a component of MM^{\prime} by (7).

Now suppose that pi=(0,0,zi)Zp_{i}=(0,0,z_{i})\in Z. We must show that if Σ\Sigma is a translating graph, then Σ\Sigma is not a component of MM^{\prime}. We have already proved it when Σ\Sigma is a Δ\Delta-wing or bowl soliton. Thus suppose that Σ\Sigma is a grim reaper surface. By rotating, we can assume that Σ\Sigma is a grim reaper surface over I×𝐑I\times\mathbf{R} for some interval II. Note that H0|ΣH_{0}|\Sigma has a critical point. Thus, by (7), Σ\Sigma cannot be a component of MM^{\prime}. ∎

Corollary 5.5.

If MM is a complete translator of finite type that lies in a slab {|y|B}\{|y|\leq B\} and if pi=(xi,yi,zi)p_{i}=(x_{i},y_{i},z_{i}) is a sequence of points in MM with (xi,yi)(x_{i},y_{i}) bounded and with |zi||z_{i}|\to\infty, then, after passing to a subsequence, ν(M,pi)\nu(M,p_{i}) converges to 𝐞2\mathbf{e}_{2} or to 𝐞2-\mathbf{e}_{2}.

Proof.

By Theorem 5.4, M(0,0,zi)M-(0,0,z_{i}) converges smoothly (after passing to a subsequence) to a union of vertical planes. Since those planes are contained in the slab {|y|B}\{|y|\leq B\}, they are all normal to 𝐞2\mathbf{e}_{2}. The assertion follows immediately. ∎

Theorem 5.6.

Suppose MM is a translator. Let WW be the set of horizontal unit vectors 𝐯\mathbf{v} such that 𝖭(F𝐯|M)>0\mathsf{N}(F_{\mathbf{v}}|M)>0. Then WW is an open subset of the equator EE.

Now suppose that MM has finite type, lies in a slab {|y|<B}\{|y|<B\}, and has no boundary. Then

  1. (1)

    If 𝐯W\mathbf{v}\in\partial W, then 𝐯=±𝐞2\mathbf{v}=\pm\mathbf{e}_{2}.

  2. (2)

    If 𝖭(x()|M)=0\mathsf{N}(x(\cdot)|M)=0, then each component of MM is either a plane parallel to {y=0}\{y=0\}, or a Δ\Delta-wing, or grim reaper surface.

Proof.

The openness of WW is because 𝖭(F𝐯|M)\mathsf{N}(F_{\mathbf{v}}|M) is a lower-semicontinuous function of 𝐯\mathbf{v} (Theorem 4.2.) To prove Assertion (1), suppose 𝐯W\mathbf{v}\in\partial W. Then 𝐯W\mathbf{v}\notin W but there exists a sequence of points 𝐯nW\mathbf{v}_{n}\in W converging to 𝐯\mathbf{v}. Let pnMp_{n}\in M with ν(M,pn)=𝐯n\nu(M,p_{n})=\mathbf{v}_{n}. Then, after passing to a subsequence, Mn=MpnM_{n}^{\prime}=M-p_{n} converges to a translator MM^{\prime}. Since 𝐯W\mathbf{v}\notin W, 𝖭(F𝐯|Mn)=𝖭(F𝐯|M)=0\mathsf{N}(F_{\mathbf{v}}|M_{n}^{\prime})=\mathsf{N}(F_{\mathbf{v}}|M)=0 and therefore 𝖭(F𝐯|M)=0\mathsf{N}(F_{\mathbf{v}}|M^{\prime})=0 by lower-semicontinuity, But ν(M,0)=𝐯\nu(M^{\prime},0)=\mathbf{v}, so the component of MM^{\prime} containing 0 is a plane. (Otherwise 0 would be a critical point of F𝐯F_{\mathbf{v}} with positive multiplicity.) Since MM^{\prime} lies in the slab {|y|2B}\{|y|\leq 2B\}, 𝐯=±𝐞2\mathbf{v}=\pm\mathbf{e}_{2}. This completes the proof of Assertion (1).

Note that 𝖭(F𝐯|M)=𝖭(F𝐯|M)\mathsf{N}(F_{\mathbf{v}}|M)=\mathsf{N}(F_{-\mathbf{v}}|M), so WW is invariant under 𝐯𝐯\mathbf{v}\mapsto-\mathbf{v}. Thus, by Assertion (1), W\partial W is either the empty set or {𝐞2,𝐞2}\{\mathbf{e}_{2},-\mathbf{e}_{2}\}. Hence, WW is one of the following: \emptyset, EE, or E{𝐞2,𝐞2}E\setminus\{\mathbf{e}_{2},-\mathbf{e}_{2}\}. If 𝖭(x()|M)=0\mathsf{N}(x(\cdot)|M)=0, i.e., if 𝖭(F𝐞1|M)=0\mathsf{N}(F_{\mathbf{e}_{1}}|M)=0, then 𝐞1W\mathbf{e}_{1}\notin W so WW is empty. By the Spruck-Xiao Theorem [spruck-xiao]*Corollary 1.2, MM consists of vertical planes and graphs. By Theorem 2.1, every translating graph is a Δ\Delta-wing, a grim reaper surface, or bowl soliton. Since MM lies in a vertical slab, it cannot contain a bowl soliton. ∎

Theorem 5.7.

Suppose that MM is a complete translator of finite type contained in a slab {|y|B}\{|y|\leq B\}. Let 𝐯\mathbf{v} be a horizontal unit vector that is not ±𝐞2\pm\mathbf{e}_{2}. For t𝐑t\in\mathbf{R} and I𝐑I\subset\mathbf{R}, let M(t)=MF𝐯1(t)M(t)=M\cap F_{\mathbf{v}}^{-1}(t) and M(I)=MF𝐯1(I)M(I)=M\cap F_{\mathbf{v}}^{-1}(I). Let II be a connected subset of 𝐑\mathbf{R} that does not include any critical values of F𝐯|MF_{\mathbf{v}}|M, and let t0It_{0}\in I. Then

M(I)M(I)

is diffeomorphic to

I×M(t0).I\times M(t_{0}).

Indeed, there is a diffeomorphism of the form

(8) Φ:pM(I)(F𝐯(p),ϕ(p))I×M(t0).\Phi:p\in M(I)\mapsto(F_{\mathbf{v}}(p),\phi(p))\in I\times M(t_{0}).
Proof.

To simplify notation, we write FF in place of F𝐯F_{\mathbf{v}}. Consider the tangent vector field

V:=(F|M)|(FM)|2V:=\frac{\nabla(F|M)}{|\nabla(F|M)|^{2}}

on M(I)M(I).

Claim 1.

If JJ is a compact interval contained in II, then |V()||V(\cdot)| is bounded above on M(J)M(J):

cJ:=supM(J)|V()|<.c_{J}:=\sup_{M(J)}|V(\cdot)|<\infty.

To prove the claim, suppose it fails for some JJ. Then there is a sequence pi=(xi,yi,zi)p_{i}=(x_{i},y_{i},z_{i}) in M(J)M(J) such that |V(pi)||V(p_{i})|\to\infty. Therefore

(9) (F|M)(pi)0.\nabla(F|M)(p_{i})\to 0.

Since II contains no critical values of FF, the sequence pip_{i} diverges. Note that xix_{i} and yiy_{i} are bounded, and thus that |zi||z_{i}|\to\infty. By Corollary 5.5, ν(M,pi)\nu(M,p_{i}) converges (after passing to a subsequence) to 𝐞2\mathbf{e}_{2} or to 𝐞2-\mathbf{e}_{2}. We may choose the orientation on MM so that ν(M,pi)\nu(M,p_{i}) converges to 𝐞2\mathbf{e}_{2}. Thus

(F|M)(pi)\displaystyle\nabla(F|M)(p_{i}) =𝐯(𝐯ν(M,pi))ν(M,pi)\displaystyle=\mathbf{v}-(\mathbf{v}\cdot\nu(M,p_{i}))\nu(M,p_{i})
𝐯(𝐯𝐞2)𝐞2,\displaystyle\to\mathbf{v}-(\mathbf{v}\cdot\mathbf{e}_{2})\mathbf{e}_{2},

which is nonzero since 𝐯±𝐞2\mathbf{v}\neq\pm\mathbf{e}_{2}. But that contradicts (9). Thus we have proved Claim 1.

For pM(I)p\in M(I), let tIqp(t)t\in I\mapsto q_{p}(t) be the solution to the following initial value problem. The ODE is qp(t)=V(qp(t))q_{p}^{\prime}(t)=V(q_{p}(t)), and the initial condition is that qp(t)=pq_{p}(t)=p at time t=F(p)t=F(p).

Note that

(d/dt)F(qp(t))=FV=1,(d/dt)F(q_{p}(t))=\nabla F\cdot V=1,

so

F(qp(t))t+cF(q_{p}(t))\equiv t+c

for some constant cc. Putting t=F(p)t=F(p), we see that F(p)=F(p)+cF(p)=F(p)+c, so c=0c=0. Thus

F(qp(t))t.F(q_{p}(t))\equiv t.

That is, at each time tt, the point qp(t)q_{p}(t) is in the level set M(t)M(t).

The bound on |V()||V(\cdot)| in Claim 1 imply that during a compact time interval JIJ\subset I, qp(t)q_{p}(t) traces out a curve of length at most cJ|J|c_{J}|J|. Thus the solution tqp(t)t\mapsto q_{p}(t) exists for the entire interval tIt\in I.

Now we define the diffeomorphism Φ\Phi in (8) by letting

ϕ(p)=qp(t0).\phi(p)=q_{p}(t_{0}).

The inverse Ψ\Psi of Φ\Phi is given by

Ψ:M(t0)×IM(I),\displaystyle\Psi:M(t_{0})\times I\to M(I),
Ψ(p,t)=qp(t).\displaystyle\Psi(p,t)=q_{p}(t).

Theorem 5.8.

Suppose that MM is a translator of finite type and that MM has no boundary. Suppose also that MM lies in a vertical slab {|y|B}\{|y|\leq B\} and that MM is invariant under (x,y,z)(x,y,z)(x,y,z)\mapsto(-x,y,z). Then as zz\to\infty (or as zz\to-\infty), the surfaces M+(0,0,z)M+(0,0,z) converge smoothly to a finite union of vertical planes.

Note that we get convergence, not just subsequential convergence. The limit as zz\to\infty will, in general, be different from the limit as zz\to-\infty. For example, if MM is a complete translating graph in a slab, then M+(0,0,z)M+(0,0,z) converges as zz\to\infty to a pair of parallel planes, and M+(0,0,z)M+(0,0,z) converges as zz\to-\infty to the empty set.

Proof.

Note that MM intersects the plane {x=0}\{x=0\} orthogonally, and thus

Γ:=M{x=0}\Gamma:=M\cap\{x=0\}

is a smooth curve. Let 𝐯=𝐞2\mathbf{v}=\mathbf{e}_{2}. Any critical point of F𝐯|ΓF_{\mathbf{v}}|\Gamma is also a critical point of F𝐯|MF_{\mathbf{v}}|M. Thus F𝐯|ΓF_{\mathbf{v}}|\Gamma has only finitely many critical points. Consequently, if t[0,)γ(t)=(0,y(t),z(t))t\in[0,\infty)\mapsto\gamma(t)=(0,y(t),z(t)) is a parametrization of an end of Γ\Gamma, then y(t)=F𝐯(γ(t))y(t)=F_{\mathbf{v}}(\gamma(t)) is eventually mononotic and thus has a well-defined limit as tt\to\infty.

It follows that Γ+(0,0,z)\Gamma+(0,0,z) converges as zz\to\infty to a union iLi\cup_{i}L_{i} of vertical lines that are contained in the plane {x=0}\{x=0\} and in the slab {|y|B}\{|y|\leq B\}. Let PiP_{i} be the plane in the slab {|y|B}\{|y|\leq B\} such that Pi{x=0}=LiP_{i}\cap\{x=0\}=L_{i}.

Now let MM^{\prime} be any subsequential limit of M+(0,0,z)M+(0,0,z) as zz\to\infty. By Theorem 5.4, MM^{\prime} is a union of vertical planes. We know that

(10) M{x=0}=iLi.M^{\prime}\cap\{x=0\}=\cup_{i}L_{i}.

It follows that

(11) M=iPi.M^{\prime}=\cup_{i}P_{i}.

Since the limit (11) does not depend on choice of subsequence, we get convergence (and not just subsequential convergence) of M+(0,0,z)M+(0,0,z) to MM^{\prime}.

The same proof works for convergence of M+(0,0,z)M+(0,0,z) as zz\to-\infty. ∎

Theorem 5.9.

Let MM be a connected translator of finite type. Assume that MM is invariant under reflection in the plane {y=0}\{y=0\}, and that MM is not contained in {y=0}\{y=0\}. Then

  1. (1)

    Γ:=(MM){y=0}\Gamma:=(M\setminus\partial M)\cap\{y=0\} is a smooth 11-manifold.

  2. (2)

    If

    t[0,)γ(t)=(x(t),0,z(t))t\in[0,\infty)\mapsto\gamma(t)=(x(t),0,z(t))

    is an arclength parametrization of an end of Γ\Gamma, then γ(t)\gamma^{\prime}(t) converges as tt\to\infty to a limit 𝐮=(x(),0,z())\mathbf{u}=(x^{\prime}(\infty),0,z^{\prime}(\infty)).

Furthermore, if MM lies in the slab {|y|B}\{|y|\leq B\} and if dist(γ(t),M)\operatorname{dist}(\gamma(t),\partial M)\to\infty, then

  1. (3)

    Mγ(t)M-\gamma(t) converges subsequentially to a limit MM^{\prime}, and the component Σ\Sigma of MM^{\prime} containing the origin is the grim reaper surface that contains the line L:={s𝐮:s𝐑}L:=\{s\mathbf{u}:s\in\mathbf{R}\} and that is symmetric about the plane {y=0}\{y=0\}.

  2. (4)

    |z()|s(B)|x()||z^{\prime}(\infty)|\leq s(B)\,|x^{\prime}(\infty)|.

  3. (5)

    x(t)x(t) tends to \infty (if x()>0x^{\prime}(\infty)>0) or to -\infty (if x()<0x^{\prime}(\infty)<0).

Proof of Theorem 5.9.

Because MM intersects the plane {y=0}\{y=0\} orthogonally, Γ\Gamma is a smooth 11-manifold.

Let m𝐑m\in\mathbf{R}. Let βπ/2\beta\geq\pi/2 be such that there is a grim reaper function

h:𝐑×(β,β)𝐑h:\mathbf{R}\times(-\beta,\beta)\to\mathbf{R}

with xhm\frac{\partial}{\partial x}h\equiv m. Then

𝖭(H|M)k<\mathsf{N}(H|M)\leq k<\infty

where H(x,y,z):=zh(x,y)H(x,y,z):=z-h(x,y) (as in Definition 5) and where kk is as in Definition 5.1. Each point on Γ\Gamma where the tangent line is parallel to the line {z=mx,y=0}\{z=mx,\,y=0\} is a critical point of H|MH|M. Hence there are at most kk such points.

It follows immediately that γ(t)\gamma^{\prime}(t) converges to a limit 𝐮\mathbf{u}.

Now suppose that MM lies in the slab {|y|B}\{|y|\leq B\}. By Theorem 5.4, Mγ(t)M-\gamma(t) converges smoothly, perhaps after passing to subsequence, to a limit MM^{\prime} consisting of vertical planes and grim reaper surfaces. The component Σ\Sigma containing the origin is orthogonal to the plane {y=0}\{y=0\}. Thus if it were a vertical plane, it would be the plane {x=0}\{x=0\}, which is impossible since Σ\Sigma is contained in the slab {|y|B}\{|y|\leq B\}. Thus Σ\Sigma is a grim reaper surface. Now Σ\Sigma contains the line L:={t𝐮:t𝐑}L:=\{t\mathbf{u}:t\in\mathbf{R}\} and intersects the plane {y=0}\{y=0\} orthogonally along LL. Thus it is the unique grim reaper surface with those properties.

Since the grim reaper surface Σ\Sigma is contained in the slab {|y|B}\{|y|\leq B\}, the slope of the line LL is at most s(B)s(B) in absolute value. Thus Assertion (4) holds.

By Assertion (4), x()0x^{\prime}(\infty)\neq 0. Thus Assertion (5) holds. ∎

Remark 5.10.

As mentioned earlier, the condition (3) on F𝐯F_{\mathbf{v}} in the definition of finite type is redundant: it is implied by condition (4), as we now explain. By rotating, it suffices to consider the case 𝐯=𝐞1\mathbf{v}=\mathbf{e}_{1}. For bπ/2b\geq\pi/2, let

hb:𝐑×(b,b)𝐑h_{b}:\mathbf{R}\times(-b,b)\to\mathbf{R}

be the grim reaper surface with hb(0,0)=0h_{b}(0,0)=0 and xhb0\frac{\partial}{\partial x}h_{b}\geq 0. Let b\mathscr{F}_{b} be the foliation of the slab {|y|<b}\{|y|<b\} by surfaces hb=constanth_{b}=\textnormal{constant}, i.e., by the level sets of Hb(x,y,z):=zhb(x,y)H_{b}(x,y,z):=z-h_{b}(x,y). Since we are assuming condition (4) Definition 5.1,

𝖭(b|M)=𝖭(Hb|M)k.\mathsf{N}(\mathscr{F}_{b}|M)=\mathsf{N}(H_{b}|M)\leq k.

As bb\to\infty, the foliation b\mathscr{F}_{b} converges to the foliation \mathscr{F} of 𝐑3\mathbf{R}^{3} by the level sets of F𝐞1F_{\mathbf{e}_{1}}. Thus, by lower semicontinuity,

𝖭(F𝐞1|M)=𝖭(|M)lim infb𝖭(b|M)k.\mathsf{N}(F_{\mathbf{e}_{1}}|M)=\mathsf{N}(\mathscr{F}|M)\leq\liminf_{b\to\infty}\mathsf{N}(\mathscr{F}_{b}|M)\leq k.

6. The Space \mathscr{R} of Annuli with Rectangular Boundaries

Definition 6.1.

We define 𝒞\mathscr{C} to be the space of compact, embedded, translating annuli MM such that

  1. (1)

    The boundary of MM is a pair of disjoint, nested, convex closed curves in the plane {z=0}\{z=0\}:

  2. (2)

    MM is invariant under reflection in the planes {x=0}\{x=0\} and {y=0}\{y=0\}.

  3. (3)

    MM is disjoint from the zz-axis.

We let innerM\partial_{\textnormal{inner}}M and outerM\partial_{\textnormal{outer}}M denote the inner and outer components of M\partial M. We define a(M)a(M), b(M)b(M), A(M)A(M), and B(M)B(M) to be the positive numbers such that

(a(M),0,0)\displaystyle(a(M),0,0) innerM,\displaystyle\in\partial_{\textnormal{inner}}M,
(A(M),0,0)\displaystyle(A(M),0,0) outerM,\displaystyle\in\partial_{\textnormal{outer}}M,
(0,b(M),0)\displaystyle(0,b(M),0) innerM,\displaystyle\in\partial_{\textnormal{inner}}M,
(0,B(M),0)\displaystyle(0,B(M),0) outerM.\displaystyle\in\partial_{\textnormal{outer}}M.

See Figures 3 and 6.

Refer to caption
Figure 6. innerM\partial_{\textnormal{inner}}M and outerM.\partial_{\textnormal{outer}}M.

Condition (3) in Definition 6.1 is redundant, but we include it here for convenience.

Definition 6.2.

We define \mathscr{R} to be the space of M𝒞M\in\mathscr{C} such that

  1. (1)

    innerM\partial_{\textnormal{inner}}M are outerM\partial_{\textnormal{outer}}M are rectangles whose sides are parallel to the coordinate axes, and

  2. (2)

    MM is the limit of a sequence of tranlators Mn𝒞M_{n}\in\mathscr{C} such that for each nn, innerMn\partial_{\textnormal{inner}}M_{n} and outerMn\partial_{\textnormal{outer}}M_{n} are smooth with nowhere vanishing curvature.

(Condition (2) might be redundant: perhaps every M𝒞M\in\mathscr{C} with property (1) also has property (2).)

When discussing a surface MM in \mathscr{R}, we will often write aa, bb, AA, and BB in place of the more cumbersome a(M)a(M), b(M)b(M), A(M)A(M), and B(M)B(M). Likewise, if MnM_{n} is a sequence of surfaces in \mathscr{R}, we will often write ana_{n}, bnb_{n}, AnA_{n}, and BnB_{n} in place of a(Mn)a(M_{n}), b(Mn)b(M_{n}), A(Mn)A(M_{n}), and B(Mn)B(M_{n}). For MM\in\mathscr{R}, the boundary is completely determined by the four numbers a:=a(M)a:=a(M), b:=b(M)b:=b(M), A:=A(M)A:=A(M), and B:=B(M)B:=B(M):

innerM\displaystyle\partial_{\textnormal{inner}}M =([a,a]×[b,b]),\displaystyle=\partial([-a,a]\times[-b,b]),
outerM\displaystyle\partial_{\textnormal{outer}}M =([A,A]×[B,B]).\displaystyle=\partial([-A,A]\times[-B,B]).
Theorem 6.3.

Suppose M𝒞M\in\mathscr{C}. Let U𝐑2U\subset\mathbf{R}^{2} be an open strip of width π\pi and let h:U𝐑h:U\to\mathbf{R} be an untilted grim reaper. Let

H:U×𝐑𝐑,\displaystyle H:U\times\mathbf{R}\to\mathbf{R},
H(x,y,z)=zh(x,y).\displaystyle H(x,y,z)=z-h(x,y).

Then

𝖭(H|M)4.\mathsf{N}(H|M)\leq 4.
Proof.

There are at most countably many lines that contain a segment in M\partial M. It suffices to prove the Theorem for strips UU that are not parallel to any of those lines; the general case follows by lower semicontinuity.

Let LL be the line in UU that bisects UU. Let CC be a closed convex curve in the plane {z=0}\{z=0\}. If LL passes through the interior of CC, then H|CH|C has exactly two local minima, namely the two points in LCL\cap C. Otherwise, LL has at most one local minimum, namely the point in CC closest to LL (if that point is in UU.)

Thus |S|4|S|\leq 4 in the formula

𝖭(H|M)|S||T|χ(M(U×𝐑)).\mathsf{N}(H|M)\leq|S|-|T|-\chi(M\cap(U\times\mathbf{R})).

Also, since MM is an annulus, χ(M(U×𝐑))0\chi(M\cap(U\times\mathbf{R}))\geq 0. (If this is not clear, see Lemma 6.5 below.) Thus 𝖭(H|M)40=4\mathsf{N}(H|M)\leq 4-0=4. ∎

Theorem 6.4.

Suppose MM\in\mathscr{R}. Let h:U𝐑h:U\to\mathbf{R} be a complete translator. (Thus hh is a grim reaper surface, a Δ\Delta-wing, or a bowl soliton.) Let

H:U×𝐑𝐑,\displaystyle H:U\times\mathbf{R}\to\mathbf{R},
H(x,y,z)=zh(x,y).\displaystyle H(x,y,z)=z-h(x,y).

Then

𝖭(H|M)8.\mathsf{N}(H|M)\leq 8.
Proof.

In case UU is a strip, we may assume that UU is not parallel to either coordinate axis; the general case then follows by lower semicontinuity (Theorem 4.2).

Note that the restriction of hh to any line segment not parallel to UU is strictly concave. Thus hh has at most one local maximum on each of the eight edges of M\partial M, so h|Mh|\partial M has at most 88 local maxima. The local maxima of h|Mh|\partial M are the local minima of H|MH|\partial M, so |S|8|S|\leq 8 in the formula

𝖭(H|M)|S||T|χ(M(U×𝐑))\mathsf{N}(H|M)\leq|S|-|T|-\chi(M\cap(U\times\mathbf{R}))

from Theorem 4.3. Also, since MM is an annulus, χ(M(U×𝐑))0\chi(M\cap(U\times\mathbf{R}))\geq 0. (If this is not clear, see Lemma 6.5 below.) Thus 𝖭(H|M)80=8\mathsf{N}(H|M)\leq 8-0=8. ∎

Lemma 6.5.

Suppose that MM is a translator and that U𝐑2U\subset\mathbf{R}^{2} is a convex set. Then the inclusion of M(U×𝐑)M\cap(U\times\mathbf{R}) into MM induces a monomorphism of first homology. In particular, if MM is an annulus, then at most one component of M(U×𝐑)M\cap(U\times\mathbf{R}) is an annulus, and the non-annular components are disks. If M𝒞M\in\mathscr{C} and if 0U0\notin U, then each component of M(U×𝐑)M\cap(U\times\mathbf{R}) is a disk.

Proof.

Let CC be a closed embedded curve in M(U×𝐑)M\cap(U\times\mathbf{R}) that is homologically trivial in MM. Then CC bounds a 22-chain in MM. Let Σ\Sigma be the support of that 22-chain. By the strong maximum principle, for each horizontal unit vector 𝐯\mathbf{v}, the maximum of F𝐯|ΣF_{\mathbf{v}}|\Sigma occurs on CC. It follows that Σ\Sigma lies in U×𝐑U\times\mathbf{R}, and thus that CC is homologically trivial in M(U×𝐑)M\cap(U\times\mathbf{R}).

Now suppose that M𝒞M\in\mathscr{C} and that 0U0\notin U. Each closed curve in M×(U×𝐑)M\times(U\times\mathbf{R}) is homotopically trivial in 𝐑3𝐙\mathbf{R}^{3}\setminus\mathbf{Z}, and therefore homotopically trivial in MM by Lemma 3.5. Hence it is homotopically trivial in M(U×𝐑)M\cap(U\times\mathbf{R}). Thus each component of M×(U×𝐑)M\times(U\times\mathbf{R}) is a disk. ∎

Theorem 6.6.

Suppose that M𝒞M\in\mathscr{C} and that 𝐯\mathbf{v} is a horizontal unit vector. Then

  1. (1)

    𝖭(F𝐯|M)\mathsf{N}(F_{\mathbf{v}}|M) is either 0 or 22. If 𝖭(F𝐯|M)=2\mathsf{N}(F_{\mathbf{v}}|M)=2, there are two critical points, each with multiplicity 11 (i.e., with Gauss curvature <0<0.)

  2. (2)

    There is at most one interior point of MM at which ν(p)=𝐯\nu(p)=\mathbf{v}. If 𝐯=±𝐞1\mathbf{v}=\pm\mathbf{e}_{1} then y(p)=0y(p)=0, and if 𝐯=±𝐞2\mathbf{v}=\pm\mathbf{e}_{2}, then x(p)=0x(p)=0.

Proof.

First we claim that

𝖭(F𝐯|M)2.\mathsf{N}(F_{\mathbf{v}}|M)\leq 2.

Except for a countable set 𝒱\mathscr{V} of unit vectors 𝐯\mathbf{v}, the function F𝐯|MF_{\mathbf{v}}|\partial M has exactly two local minima, so by the counting formula (Theorem 4.3),

(12) 𝖭(F𝐯|M)20χ(M)=2\mathsf{N}(F_{\mathbf{v}}|M)\leq 2-0-\chi(M)=2

if 𝐯𝒱\mathbf{v}\notin\mathscr{V}. By lower-semicontinuity (Theorem 4.2), (12) holds for all horizontal unit vectors 𝐯\mathbf{v}.

If F𝐯|MF_{\mathbf{v}}|M has an interior critical point pp, then ρZp\rho_{Z}p is also an interior critical point, so each has multiplicity 11. Since pp has multiplicity one, the order of contact at pp between MM and the plane {F𝐯=F𝐯(p)}\{F_{\mathbf{v}}=F_{\mathbf{v}}(p)\} is 22. Thus the Gauss curvature of MM at pp is nonzero. Since the mean curvature at pp is (𝐞3)=0(-\mathbf{e}_{3})^{\perp}=0, the Gauss curvature is negative. Thus we have proved Assertion (1).

To prove Assertion (2), note that the interior critical points of F𝐯|MF_{\mathbf{v}}|M are the points where ν=±𝐯\nu=\pm\mathbf{v}. If ν(p)=𝐯\nu(p)=\mathbf{v}, then ν(ρZp)=𝐯\nu(\rho_{Z}p)=-\mathbf{v}, and, by Assertion (1), there are no other points where ν=±𝐯\nu=\pm\mathbf{v}. If ν(p)=𝐞1\nu(p)=\mathbf{e}_{1} (or if ν(p)=𝐞1\nu(p)=-\mathbf{e}_{1}), then y(p)=0y(p)=0, since otherwise the image of pp under (x,y,z)(x,y,z)(x,y,z)\mapsto(x,-y,z) would be another point at which ν=𝐞1\nu=\mathbf{e}_{1} (or ν=𝐞1\nu=-\mathbf{e}_{1}). The analogous argument applies when ν(p)=±𝐞2\nu(p)=\pm\mathbf{e}_{2}. ∎

Corollary 6.7.

Suppose that Mn𝒞M_{n}\in\mathscr{C} and that Mn:=MnpnM_{n}^{\prime}:=M_{n}-p_{n} converges to a limit MM. Let M~\tilde{M} be the set of points pMMp\in M\setminus\partial M such that the convergence is smooth at pp and such that the component of M~\tilde{M} containing pp is not contained in a vertical plane. Suppose 𝐯\mathbf{v} is a horizontal unit vector. Then there is at most one point qM~q\in\tilde{M} such that ν(M~,q)=𝐯\nu(\tilde{M},q)=\mathbf{v}. Furthermore, the Gauss curvature of MM at such a point qq is <0<0, and if the vertical plane V𝐯V_{\mathbf{v}} containing 𝐯\mathbf{v} is a plane of symmetry of MM, then qVq\in V.

Proof.

Let Mn+={pMn:ν(p)𝐯>0}M_{n}^{+}=\{p\in M_{n}:\nu(p)\cdot\mathbf{v}>0\} and M~+={pM~:ν(p)𝐯>0}\tilde{M}^{+}=\{p\in\tilde{M}:\nu(p)\cdot\mathbf{v}>0\}. By Theorem 6.6, 𝖭(F𝐯|Mn+)1\mathsf{N}(F_{\mathbf{v}}|M_{n}^{+})\leq 1, so (by lower semicontinuity) 𝖭(F𝐯|M~+)1\mathsf{N}(F_{\mathbf{v}}|\tilde{M}^{+})\leq 1. Thus there is at most one point in M~\tilde{M} at which ν=𝐯\nu=\mathbf{v}, and if qq is such a point, it is a critical point of multiplicity 11 and thus the Gauss curvature at qq is negative.

Note that if V𝐯V_{\mathbf{v}} is a plane of symmetry of MM, then qV𝐯q\in V_{\mathbf{v}} since otherwise the image of qq under reflection in V𝐯V_{\mathbf{v}} would be a second point at which ν=𝐯\nu=\mathbf{v}. ∎

7. The Space 𝒜\mathscr{A} of Annuloids

We will show the following by a path-lifting argument:

Theorem 7.1.

Suppose that bπ/2b\geq\pi/2 and that 0<x^<a0<\hat{x}<a. Then there exists a translator MM\in\mathscr{R} (not necessarily unique) such that

a(M)\displaystyle a(M) =a,\displaystyle=a,
b(M)\displaystyle b(M) =b,\displaystyle=b,
x(M)\displaystyle x(M) =x^.\displaystyle=\hat{x}.

We postpone the proof to §1821; see Lemma 21.1. However, interested readers may skip to §18 after finishing this section (§7); the proof of Theorem 7.1 (i.e., of Lemma 21.1) does not depend on anything in the intervening sections. In any case, there is no risk of circularity in postponing that proof. In the intervening sections, we prove various properties of complete MM that arise as limits of MnM_{n}\in\mathscr{R}, but nowhere do we assume existence in those proofs. (If existence did not hold, then the assertions would be vacuously true.)

We now wish to see what happens to the surface MM in Theorem 7.1 if we fix bb and x^\hat{x} and let aa\to\infty.

Definition 7.2.

If M𝒞M\in\mathscr{C}, we let z(M)z(M) be the largest value of zz such that (x(M),0,z)M(x(M),0,z)\in M.

(In fact, there is only one zz for which (x(M),0,z)(x(M),0,z) is in MM; see Corollary 8.3.)

Theorem 7.3.

Let MiM_{i} be a sequence in \mathscr{R} such that

bi:=b(Mi)b[π/2,),\displaystyle b_{i}:=b(M_{i})\to b\in[\pi/2,\infty),
x(Mi)x^(0,),\displaystyle x(M_{i})\to\hat{x}\in(0,\infty),
ai:=a(Mi).\displaystyle a_{i}:=a(M_{i})\to\infty.

Then, after passing to a subsequence, the surfaces

Mi:=Mi(0,0,z(Mi))M_{i}^{\prime}:=M_{i}-(0,0,z(M_{i}))

converge smoothly to a complete translator MM. Furthermore,

  1. (1)

    MM is symmetric with respect to the planes {x=0}\{x=0\} and {y=0}\{y=0\}.

  2. (2)

    x(M)=x^x(M)=\hat{x} and (x^,0,0)M(\hat{x},0,0)\in M.

  3. (3)

    area(M𝐁(p,r))c1r2\operatorname{area}(M\cap\mathbf{B}(p,r))\leq c_{1}r^{2} for every p𝐑3p\in\mathbf{R}^{3} and r0r\geq 0.

  4. (4)

    |A(M,)|min{1,x(M)}c2|A(M,\cdot)|\min\{1,x(M)\}\leq c_{2}.

  5. (5)

    |A(M,p)|min{1,dist(p,Z)}c2|A(M,p)|\,\min\{1,\operatorname{dist}(p,Z)\}\leq c_{2}.

  6. (6)

    𝖭(F𝐯|M)2\mathsf{N}(F_{\mathbf{v}}|M)\leq 2 for every horizontal unit vector 𝐯\mathbf{v}.

  7. (7)

    𝖭(H|M)8\mathsf{N}(H|M)\leq 8 for every function H(x,y,z)=zh(x,y)H(x,y,z)=z-h(x,y) such that the graph of hh is a complete translator.

  8. (8)

    The connected component of MM containing (x^,0,0)(\hat{x},0,0) is an annulus.

Here c1c_{1} and c2c_{2} are the constants in Theorem 3.4.

Concerning Assertion (8), we will show later (Theorem 10.6) that MM is connected and therefore that it is an annulus.

Proof.

First, we claim that

(13) z(Mi).z(M_{i})\to\infty.

To see this, let fa,b:[a,a]×[b,b]𝐑f_{a,b}:[-a,a]\times[-b,b]\to\mathbf{R} be the translator with boundary values 0. By the maximum principle,

Mi([ai,ai]×[bi,bi]×𝐑)M_{i}\cap([-a_{i},a_{i}]\times[-b_{i},b_{i}]\times\mathbf{R})

lies in {zuai,bi(x,y)}\{z\geq u_{a_{i},b_{i}}(x,y)\}. As ii\to\infty, fai,bi(0,0)f_{a_{i},b_{i}}(0,0)\to\infty and

fai,bi(x,y)fai,bi(0,0)f_{a_{i},b_{i}}(x,y)-f_{a_{i},b_{i}}(0,0)

converges smoothly to a translator

fb:𝐑×(b,b)𝐑.f_{b}:\mathbf{R}\times(-b,b)\to\mathbf{R}.

(See [graphs]*Theorem 4.1). Hence if KK is a compact subset of 𝐑×(b,b)\mathbf{R}\times(-b,b), then

minKfai,bi.\min_{K}f_{a_{i},b_{i}}\to\infty.

Thus (13) holds.

By (13),

(14) dist(0,Mi).\operatorname{dist}(0,\partial M_{i}^{\prime})\to\infty.

Consequently, the curvature and area bounds in Theorem 3.4 give smooth convergence (after passing to a subsequence) of MiM_{i}^{\prime} to a limit translator MM. By (14), MM has no boundary. From the construction, (x^,0,0)M(\hat{x},0,0)\in M. Also, MM is disjoint from the strip (x^,x^)×{0}×𝐑(-\hat{x},\hat{x})\times\{0\}\times\mathbf{R}. Thus x(M)=x^x(M)=\hat{x}. All the assertions other than Assertion (8) follow trivially from the corresponding properties of the MiM_{i}^{\prime}. See Theorems  4.2 and 3.4.

Let Σ\Sigma be the connected component of MM containing (x^,0,0)(\hat{x},0,0). Let f:𝐑2𝐑f:\mathbf{R}^{2}\to\mathbf{R} be the bowl solition. Then, for any λ\lambda, each component of

(*) Mn{λ<z<f(x,y)+λ}M_{n}^{\prime}\cap\{-\lambda<z<f(x,y)+\lambda\}

is a disk or an annulus.

(If that is not clear, note that if one or more closed curves in (*7) bound a region KK in MnM_{n}^{\prime}, then f|Kf|K must attain its maximum on K\partial K and z()|Kz(\cdot)|K must attain its minimum on K\partial K, and thus KK is contained in (*7). Hence the inclusion of (*7) into MM induces a monomorphism of first homology.)

By smooth convergence, each component of

M{λ<z<f(x,y)+λ}M\cap\{-\lambda<z<f(x,y)+\lambda\}

is a disk or an annulus. Letting λ\lambda\to\infty, we see that Σ\Sigma is a disk or an annulus.

Thus it suffices to show that Σ\Sigma is not simply connected. No complete translator is contained in a half-slab of the form {|y|B}{x<a}\{|y|\leq B\}\cap\{x<a\}. (See Corollary A.3.) Thus Σ{x=0}\Sigma\cap\{x=0\} is nonempty. Let CC be a shortest path in Σ{x0}\Sigma\cap\{x\geq 0\} from (x(M),0,0)(x(M),0,0) to a point in Σ{x=0}\Sigma\cap\{x=0\}. Extending by reflection in {x=0}\{x=0\} and then in {y=0}\{y=0\} gives a simple closed curve CC^{\prime} in MM that winds once around the zz-axis. Since CC^{\prime} is homologically nontrivial in 𝐑3Z\mathbf{R}^{3}\setminus Z, it is homologically nontrivial in Σ\Sigma. Thus Σ\Sigma is an annulus. ∎

Definition 7.4.

Let π/2bBb+π\pi/2\leq b\leq B\leq b+\pi and x^(0,)\hat{x}\in(0,\infty). We define 𝒜(b,B,x^)\mathscr{A}(b,B,\hat{x}) to be the space of all limits of sequences

Mi(0,0,z(Mi))M_{i}-(0,0,z(M_{i}))

such that

Mi,\displaystyle M_{i}\in\mathscr{R},
a(Mi),\displaystyle a(M_{i})\to\infty,
b(Mi)b,\displaystyle b(M_{i})\to b,
B(Mi)B,\displaystyle B(M_{i})\to B,
x(Mi)x^.\displaystyle x(M_{i})\to\hat{x}.

We let

𝒜(b,x^)\displaystyle\mathscr{A}(b,\hat{x}) :=B𝒜(b,B,x^),\displaystyle:=\bigcup_{B}\mathscr{A}(b,B,\hat{x}),
𝒜(b)\displaystyle\mathscr{A}(b) :=x^𝒜(b,x^)\displaystyle:=\bigcup_{\hat{x}}\mathscr{A}(b,\hat{x})
𝒜\displaystyle\mathscr{A} :=b𝒜(b)\displaystyle:=\bigcup_{b}\mathscr{A}(b)
=b,B,x^𝒜(b,B,x^).\displaystyle=\bigcup_{b,B,\hat{x}}\mathscr{A}(b,B,\hat{x}).
Remark 7.5.

Let bπ/2b\geq\pi/2 and x^(0,)\hat{x}\in(0,\infty). By Theorems 7.1 and 7.3, 𝒜(b,x^)\mathscr{A}(b,\hat{x}) is nonempty, and the surfaces in 𝒜(b,x^)\mathscr{A}(b,\hat{x}) have the properties listed in Theorem 7.3. In particular, the surfaces in 𝒜\mathscr{A} are translators of finite type (as defined in Section 5).

For surfaces in MM in \mathscr{R} or in 𝒜\mathscr{A}, there is a curvature bound that gives more information than the bounds in Theorem 3.4 and 7.3 when x(M)x(M) is close to 0:

Theorem 7.6.

Suppose that MM\in\mathscr{R} or M𝒜M\in\mathscr{A}. If 𝐯\mathbf{v} is a horizontal unit vector, let

δ(M,𝐯,p)=min{|pq|:qM,ν(M,q)=𝐯}.\delta(M,\mathbf{v},p)=\min\{|p-q|:q\in M,\,\nu(M,q)=\mathbf{v}\}.

(Here δ(M,𝐯,p)=\delta(M,\mathbf{v},p)=\infty if there is no such point qq.) Then

|A(M,p)|min{1,δ(M,𝐯,p),dist(p,M)}C.|A(M,p)|\,\min\{1,\delta(M,\mathbf{v},p),\operatorname{dist}(p,\partial M)\}\leq C.
Proof.

It suffices to prove the theorem for M𝒞M\in\mathscr{C} with smooth boundaries, as any surface in \mathscr{R} is a limit of such surfaces.

Thus, if the theorem is false, there is a sequence Mn𝒞M_{n}\in\mathscr{C} with smooth boundaries such that

suppMn|A(Mn,p)|min{1,δ(Mn,𝐯,p),dist(p,Mn)}.\sup_{p\in M_{n}}|A(M_{n},p)|\,\min\{1,\delta(M_{n},\mathbf{v},p),\operatorname{dist}(p,\partial M_{n})\}\to\infty.

For each nn, the supremum is attained at some point pnp_{n}.

Translate MnM_{n} by pn-p_{n} and then dilate by |A(Mn,pn)||A(M_{n},p_{n})| to get MnM_{n}^{\prime}. Thus |A(Mn,0|=1|A(M_{n}^{\prime},0|=1 and

(15) δ(Mn,𝐯,0),\displaystyle\delta(M_{n}^{\prime},\mathbf{v},0)\to\infty,
(16) dist(0,M).\displaystyle\operatorname{dist}(0,\partial M^{\prime})\to\infty.

Then MnM_{n}^{\prime} converges smoothly (after passing to a subsequence) to a complete, embedded minimal surface MM^{\prime} with |A(M,0)|=1|A(M^{\prime},0)|=1. Since MM^{\prime} has genus 0 and quadratic area growth, it is a catenoid.

If the axis of MM^{\prime} is not parallet to 𝐯\mathbf{v}, then there is a point qMq\in M^{\prime} such that ν(M,q)=𝐯\nu(M^{\prime},q)=\mathbf{v}. The curvature of MM^{\prime} at qq is non zero, so qq is the limit of points qnMnq_{n}\in M_{n}^{\prime} such that ν(Mn,qn)=𝐯\nu(M_{n}^{\prime},q_{n}^{\prime})=\mathbf{v}. But that is excluded by (15).

Thus the axis of MM^{\prime} is parallel to 𝐯\mathbf{v}, so the plane 𝐯\mathbf{v}^{\perp} intersects MM^{\prime} transversely in a circle. Consequently, for large nn, 𝐯Mn\mathbf{v}^{\perp}\cap M_{n}^{\prime} contains a simple closed curve. But that is impossible by Lemma 3.5. ∎

Remark 7.7.

Suppose that MnM_{n} is a sequence of surfaces in \mathscr{R} or in 𝒜\mathscr{A}, that Mn=MnqnM_{n}^{\prime}=M_{n}-q_{n} converges to a limit MM, and that the convergence is not smooth at a point pMMp\in M\setminus\partial M. By Theorem 7.6, the two points in MnM_{n}^{\prime} where ν=±𝐞1\nu=\pm\mathbf{e}_{1} both have to converge to pp. Hence there can be at most one point in MMM\setminus\partial M where the convergence MnMM_{n}^{\prime}\to M is not smooth.

Theorem 7.8.

Suppose that MnM_{n}\in\mathscr{R} or Mn𝒜M_{n}\in\mathscr{A}, and that the MnM_{n} lie in a slab {|y|λ}\{|y|\leq\lambda\}. Let pn±p_{n}^{\pm} be the points where ν(Mn,pn±)=±𝐞1\nu(M_{n},p_{n}^{\pm})=\pm\mathbf{e}_{1}. Suppose that qnq_{n} are points in the plane {y=0}\{y=0\} such that

dist(qn,{pn+,pn}Mn).\operatorname{dist}(q_{n},\{p_{n}^{+},p_{n}^{-}\}\cup\partial M_{n})\to\infty.

Then, after passing to a subsequence, Mn=MnqnM_{n}^{\prime}=M_{n}-q_{n} converges smoothly to a limit MM^{\prime}, each component of which is a plane parallel to {y=0}\{y=0\}, or a Δ\Delta-wing, or a grim reaper surface.

Now suppose also that each qn=(xn,0,zn)q_{n}=(x_{n},0,z_{n}) is in MnM_{n}, and that xn0x_{n}\geq 0. Then the component Σ\Sigma of MM^{\prime} containing 0 is either a Δ\Delta-wing or a grim reaper surface. Furthermore, in this case,

xnx(Mn).x_{n}-x(M_{n})\to\infty.
Proof.

The curvature bound in Theorem 7.6 implies that we get smooth subsequential convergence, and that the limit surface has bounded principal curvatures. For each bounded open set UU, 𝖭(x()|MnU)=0\mathsf{N}(x(\cdot)|M_{n}^{\prime}\cap U)=0 for all sufficiently large nn, so 𝖭(x()|MU)=0\mathsf{N}(x(\cdot)|M^{\prime}\cap U)=0. Since UU is arbitrary, 𝖭(x()|M)=0\mathsf{N}(x(\cdot)|M^{\prime})=0. By Theorem 5.6, the components of MM^{\prime} are planes parallel to {y=0}\{y=0\}, Δ\Delta-wings, and grim reaper surfaces.

Now suppose also that each qnq_{n} is in Mn{y=0}M_{n}\cap\{y=0\} and that xn0x_{n}\geq 0. Then xnx(Mn)x_{n}\geq x(M_{n}) be definition of x(Mn)x(M_{n}). The component Σ\Sigma of MM^{\prime} containing 0 is perpendicular to {y=0}\{y=0\} at 0 and is contained in the slab {|y|λ}\{|y|\leq\lambda\}, so it cannot be a plane. Hence Σ\Sigma must be a Δ\Delta-wing or a grim reaper surface. In particular Σ\Sigma is the graph of a function

w:𝐑×(β,β)𝐑w:\mathbf{R}\times(-\beta,\beta)\to\mathbf{R}

for some β\beta.

Now the component Γn\Gamma_{n} of Mn{y=0}M_{n}^{\prime}\cap\{y=0\} containing 0 converges to Σ{y=0}\Sigma\cap\{y=0\} as nn\to\infty.

Note that the minimum value of x()x(\cdot) on Γn\Gamma_{n} is x(Mn)xnx(M_{n})-x_{n}. If x(Mn)xnx(M_{n})-x_{n} were bounded below by α\alpha, then x()x(\cdot) would be bounded below by α\alpha on Σ{y=0}\Sigma\cap\{y=0\} (that is, on {(x,0,w(x,0)):x𝐑}\{(x,0,w(x,0)):x\in\mathbf{R}\}), which is not the case. Thus x(Mn)xnx(M_{n})-x_{n}\to-\infty, so xnx(Mn)x_{n}-x(M_{n})\to\infty. ∎

What are the next steps?

According to Remark 7.5, for each bπ/2b\geq\pi/2 and x^(0,)\hat{x}\in(0,\infty), there is a complete smooth translator MM in 𝒜(b,x^)\mathscr{A}(b,\hat{x}). We know that it has an annular component (Theorem 7.3), but it is not obvious that MM is connected. Also, although we have examples for every x^(0,)\hat{x}\in(0,\infty) and bπ/2b\geq\pi/2, it is not obvious that these examples are distinct. We will eventually prove that 𝒜(b,x^)\mathscr{A}(b,\hat{x}) and 𝒜(b,x^)\mathscr{A}(b^{\prime},\hat{x}^{\prime}) are disjoint unless b=bb=b^{\prime} and x^=x^\hat{x}=\hat{x}^{\prime}. But a priori there is the possibility that 𝒜(b,x^)=𝒜(π/2,x^)\mathscr{A}(b,\hat{x})=\mathscr{A}(\pi/2,\hat{x}) for all bπ/2b\geq\pi/2. (We do already know that x(M)=x^x(M)=\hat{x} for M𝒜(b,x^)M\in\mathscr{A}(b,\hat{x}). Thus if x^x^\hat{x}\neq\hat{x}^{\prime}, then 𝒜(b,x^)\mathscr{A}(b,\hat{x}) and 𝒜(b,x^)\mathscr{A}(b^{\prime},\hat{x}^{\prime}) are disjoint.)

Our goal now is to prove that every surface MM in 𝒜(b,B,x^)\mathscr{A}(b,B,\hat{x}) has the properties described in Theorem 1.1. In particular, we wish to show that

  1. (1)

    MM is connected (and therefore is an annulus).

  2. (2)

    M+(0,0,z)M+(0,0,z) converges as zz\to-\infty to the empty set.

  3. (3)

    There are nonnegative numbers b(M)b(M) and B(M)B(M) with b(M)B(M)b(M)\leq B(M) such that M+(0,0,z)M+(0,0,z) converges smoothly at zz\to\infty to the planes {y=±b(M)}\{y=\pm b(M)\} and {y=±B(M)}\{y=\pm B(M)\}.

  4. (4)

    B(M)=BB(M)=B.

  5. (5)

    b(M)=bb(M)=b.

We will show (1), (2), and (3) in the next three sections; see Theorems 10.5 and 10.6.

Assertions (4) and (5) seem to be much more subtle. We deduce them from other properties of MM, which we describe and prove in Sections 11, 12, and 13. Assertions (4) and (5) are proved in Theorem 13.5.

Remark 7.9 (Entropy of annuloids).

Assertion 3 above implies that the entropy of MM is 4. This is a consequence of Corollary 8.5 in [GMM].

8. The Slice M{y=0}M\cap\{y=0\}.

Let M𝒜M\in\mathscr{A}. To analyze MM, it is helpful to analyze the slice M{y=0}M\cap\{y=0\}. To analyze that slice, it suffices by symmetry to analyze

Γ:=M{y=0}{x>0}.\Gamma:=M\cap\{y=0\}\cap\{x>0\}.
Theorem 8.1.

Let M𝒜M\in\mathscr{A}. Then Γ\Gamma is the union of two graphs

{(x,0,ϕupper(x)):x[x(M),)}\{(x,0,\phi^{\textnormal{upper}}(x)):x\in[x(M),\infty)\}

and

{(x,0,ϕlower(x)):x[x(M),)}\{(x,0,\phi^{\textnormal{lower}}(x)):x\in[x(M),\infty)\}

where ϕupper\phi^{\textnormal{upper}} and ϕlower\phi^{\textnormal{lower}} are continuous on [x(M),)[x(M),\infty), smooth on (x(M),)(x(M),\infty), and where

  1. (1)

    ϕlower(x(M))=ϕupper(x(M))=0\phi^{\textnormal{lower}}(x(M))=\phi^{\textnormal{upper}}(x(M))=0.

  2. (2)

    ϕlower(x)<ϕupper(x)\phi^{\textnormal{lower}}(x)<\phi^{\textnormal{upper}}(x) for all x>x(M)x>x(M).

  3. (3)

    The limits

    (ϕupper)():=limx(ϕupper)(x)(\phi^{\textnormal{upper}})^{\prime}(\infty):=\lim_{x\to\infty}(\phi^{\textnormal{upper}})^{\prime}(x)

    and

    (ϕlower)():=limx(ϕlower)(x)(\phi^{\textnormal{lower}})^{\prime}(\infty):=\lim_{x\to\infty}(\phi^{\textnormal{lower}})^{\prime}(x)

    exist and are finite.

Later (Theorem 14.1) we will prove that (ϕlower)()=s(b(M))(\phi^{\textnormal{lower}})^{\prime}(\infty)=-s(b(M)) and that (ϕupper)()=±S(B(M))(\phi^{\textnormal{upper}})^{\prime}(\infty)=\pm S(B(M)), where s(β)s(\beta) is the slope uβ/x\partial u_{\beta}/\partial x of the grim reaper surface uβ:𝐑×(β,β)𝐑u_{\beta}:\mathbf{R}\times(-\beta,\beta)\to\mathbf{R} in §2.

Proof.

Since {y=0}\{y=0\} intersects MM orthogonally, Γ\Gamma is a smooth 11-manifold.

By Theorem 7.3 (6) (see also Remark 7.5),

2𝖭(F𝐞1|M).2\geq\mathsf{N}(F_{\mathbf{e}_{1}}|M).

Now each critical point (x,0,z)(x,0,z) of F𝐞1|ΓF_{\mathbf{e}_{1}}|\Gamma is a critical point of F𝐞1|MF_{\mathbf{e}_{1}}|M, as is its mirror image (x,0,z)(-x,0,z). Thus F𝐞1|ΓF_{\mathbf{e}_{1}}|\Gamma has at most one critical point. Since (x(M),0,0)(x(M),0,0) is a critical point, F𝐞1|ΓF_{\mathbf{e}_{1}}|\Gamma has exactly one critical point. (Note that the point (x(M),0,0)(x(M),0,0) is in MM by Definition 7.4.) Thus no component of Γ\Gamma is a closed curve. (We also know from Lemma 3.5 that no component of Γ\Gamma is a closed curve.)

By Theorem 5.9, x()x(\cdot) tends to \infty or -\infty on each end of Γ\Gamma. Since x>0x>0 on Γ\Gamma, we see that xx\to\infty on each end of Γ\Gamma. Thus each component of Γ\Gamma has at least one local minimum of F𝐞1|ΓF_{\mathbf{e}_{1}}|\Gamma. Since F𝐞1|ΓF_{\mathbf{e}_{1}}|\Gamma has only one critical point, the curve Γ\Gamma has only one component. (It is nonempty since it contains the point (x(M),0,0)(x(M),0,0).)

Since F𝐞1|ΓF_{\mathbf{e}_{1}}|\Gamma has its minimum at (x(M),0,0)(x(M),0,0) and has no other critical points, and since xx\to\infty on each end of Γ\Gamma, we see that Γ\Gamma is the union of two graphs satisfying (1) and (2). By Theorem 5.9, it also satisfies (3). ∎

Theorem 8.1 has an analog for surfaces in \mathscr{R}, or more generally, for surfaces in 𝒞\mathscr{C}:

Theorem 8.2.

Suppose that M𝒞M\in\mathscr{C}. Write a=a(M)a=a(M) and A=A(M)A=A(M). If x(M)<ax(M)<a, then M{y=0}{x>0}M\cap\{y=0\}\cap\{x>0\} is the union of

{(x,0,ϕlower(x)):x(M)xa}\{(x,0,\phi^{\textnormal{lower}}(x)):x(M)\leq x\leq a\}

and

{(x,0,ϕupper(x)):x(M)xA},\{(x,0,\phi^{\textnormal{upper}}(x)):x(M)\leq x\leq A\},

where

  1. (1)

    ϕupper\phi^{\textnormal{upper}} and ϕlower\phi^{\textnormal{lower}} are continuous and are smooth for x>x(M)x>x(M).

  2. (2)

    ϕlower(x(M))=ϕupper(x(M))=z(M)\phi^{\textnormal{lower}}(x(M))=\phi^{\textnormal{upper}}(x(M))=z(M) and ϕlower(x)<ϕupper(x)\phi^{\textnormal{lower}}(x)<\phi^{\textnormal{upper}}(x) for x>x(M)x>x(M).

  3. (3)

    ϕlower(a)=ϕupper(A)=0\phi^{\textnormal{lower}}(a)=\phi^{\textnormal{upper}}(A)=0.

If x(M)=ax(M)=a, then M{y=0}{x>0}M\cap\{y=0\}\cap\{x>0\} is

{(x,0,ϕ(x)):axA}.\{(x,0,\phi(x)):a\leq x\leq A\}.

for a function ϕ\phi that is continuous on [a,A][a,A] and smooth on (a,A)(a,A).

We omit the proof, since it is very similar to the proof of Theorem 8.1 (but simpler because behavior as xx\to\infty does not arise.)

Of course, the same theorem holds with the roles of xx and yy reversed, that is, for M{x=0}{y>0}M\cap\{x=0\}\cap\{y>0\}.

Corollary 8.3.

If M𝒞M\in\mathscr{C}, then there is a unique zz such that (x(M),0,z)M(x(M),0,z)\in M.

Recall (Definition 7.2) that z(M)z(M) is equal to the zz in Corollary 8.3.

Corollary 8.4.

Suppose M𝒞M\in\mathscr{C}. For each xx, the vertical line {(x,0)}×𝐑\{(x,0)\}\times\mathbf{R} intersects MM in at most 22 points. For each yy, the vertical line {(0,y)}×𝐑\{(0,y)\}\times\mathbf{R} intersects MM in at most 22 points. For each y(b,B]y\in(b,B] ,the vertical line {(0,y)×𝐑}\{(0,y)\times\mathbf{R}\} interects MM in exactly one point.

Theorem 8.5.

Suppose that MM is a surface in 𝒞\mathscr{C} or in 𝒜\mathscr{A}. If M𝒞M\in\mathscr{C}, suppose also that x(M)<a(M)x(M)<a(M). Let MlowerM^{\textnormal{lower}} be the component of M{x>x(M)}M\cap\{x>x(M)\} containing

graph(ϕlower){x>x(M)}\operatorname{graph}(\phi^{\textnormal{lower}})\cap\{x>x(M)\}

and let MupperM^{\textnormal{upper}} be the component of M{x>x(M)}M\cap\{x>x(M)\} containing

graph(ϕupper){x>x(M)}.\operatorname{graph}(\phi^{\textnormal{upper}})\cap\{x>x(M)\}.

Then MlowerMupperM^{\textnormal{lower}}\neq M^{\textnormal{upper}}.

Furthermore, if M𝒞M\in\mathscr{C}, then

(17) M{x>x(M)}=MlowerMupper.M\cap\{x>x(M)\}=M^{\textnormal{lower}}\cup M^{\textnormal{upper}}.

We will show later (Theorem 10.5) that (17) also holds for M𝒜M\in\mathscr{A}. (Knowing (17) for MM\in\mathscr{R} does not immediately imply it for M𝒜M\in\mathscr{A} because, in general, a limit of connected sets need not be connected.)

Proof.

It suffices to prove that MlowerMupperM^{\textnormal{lower}}\neq M^{\textnormal{upper}} for M𝒞M\in\mathscr{C}; the result for M𝒜M\in\mathscr{A} follows easily. Suppose, contrary to the theorem, that Mupper=MlowerM^{\textnormal{upper}}=M^{\textnormal{lower}}. Then there is a path CC in M{x>x(M)}M\cap\{x>x(M)\} joining a point in graph(ϕlower)\operatorname{graph}(\phi^{\textnormal{lower}}) to a point in graph(ϕupper)\operatorname{graph}(\phi^{\textnormal{upper}}). We can choose the path so that CC is embedded and so that C{y=0}=CC\cap\{y=0\}=\partial C. Let C0C_{0} be the arc in M{y=0}M\cap\{y=0\} joining the endpoints of CC. Since CC0C\cup C_{0} is homologically trivial in 𝐑3Z\mathbf{R}^{3}\setminus Z, it is homologically trivial in MM. (See Lemma 3.5.) Thus, since MM is an annulus, CC0C\cup C_{0} bounds a simply-connected region DD in MM. Now the union 𝒟\mathscr{D} of DD and its image under (x,y,z)(x,y,z)(x,y,z)\mapsto(x,-y,z) will be region whose boundary lies in {x>x(M)}\{x>x(M)\}. By the maximum principle,

min𝒟x()=min𝒟x().\min_{\mathscr{D}}x(\cdot)=\min_{\partial\mathscr{D}}x(\cdot).

The left-side is x(M)\leq x(M) since (x(M),0,z(M))𝒟(x(M),0,z(M))\in\mathscr{D}, whereas the right-side is >x(M)>x(M), a contradiction. Thus MupperMlowerM^{\textnormal{upper}}\neq M^{\textnormal{lower}}.

Now we show (assuming M𝒞M\in\mathscr{C}) that (17) holds. Note that MupperM^{\textnormal{upper}} contains outerM{x>x(M)}\partial_{\textnormal{outer}}M\cap\{x>x(M)\} and MlowerM^{\textnormal{lower}} contains innerM{x>x(M)}\partial_{\textnormal{inner}}M\cap\{x>x(M)\}. If there were another component Σ\Sigma of M{x>x(M)}M\cap\{x>x(M)\}, it would have no boundary in {x>x(M)}\{x>x(M)\}. But then x()|Σx(\cdot)|\Sigma would attain its maximum at an interior point, violating the strong maximum principle. Thus there is no such component, so (17) holds. ∎

9. A Slope Bound

For bπ/2b\geq\pi/2, let

fb:𝐑×(b,b)𝐑f_{b}:\mathbf{R}\times(-b,b)\to\mathbf{R}

be the translator such that f(0,0)=0f(0,0)=0 and Df(0,0)=0Df(0,0)=0. Thus fbf_{b} is an untilted grim reaper if b=π/2b=\pi/2 and a Δ\Delta-wing if b>π/2b>\pi/2.

The goal of this section is to prove

Theorem 9.1.

If MM\in\mathscr{R}, then

(18) (ϕlower)(x)<xfb(x,0)0(\phi^{\textnormal{lower}})^{\prime}(x)<\frac{\partial}{\partial x}f_{b}(x,0)\leq 0

for all x[x(M),a]x\in[x(M),a], where b=b(M)b=b(M).

If M𝒜(b)M\in\mathscr{A}(b), then the inequality (18) holds for all xx(M)x\geq x(M).

In particular, ϕlower\phi^{\textnormal{lower}} is a strictly decreasing function.

(In Theorem 9.1, we require that MM\in\mathscr{R} and not merely that M𝒞M\in\mathscr{C}. In fact, the theorem is false if we replace \mathscr{R} by 𝒞\mathscr{C}.)

Corollary 9.2.

If M𝒜(b)M\in\mathscr{A}(b), then

limx(ϕlower)(x)s(b).\lim_{x\to\infty}(\phi^{\textnormal{lower}})^{\prime}(x)\leq-s(b).
Proof of Corollary 9.2.

Let xx\to\infty in (18). (The limit of (ϕlower)(x)(\phi^{\textnormal{lower}})^{\prime}(x) exists by Theorem 8.1). ∎

Later (Theorem 14.1) we will show that equality holds in Corollary 9.2.

Proof of Theorem 9.1.

It suffices to prove the assertion for MM\in\mathscr{R}, as the result for M𝒜(b)M\in\mathscr{A}(b) follows trivially.

For α,β>0\alpha,\beta>0, let

fα,β:[α,α]×[β,β]𝐑f_{\alpha,\beta}:[-\alpha,\alpha]\times[-\beta,\beta]\to\mathbf{R}

be the translator with boundary values 0.

Now suppose that MM\in\mathscr{R}.

Let

α:[,][a,A]\displaystyle\alpha:[-\infty,\infty]\to[a,A]
β:[,][b,B]\displaystyle\beta:[-\infty,\infty]\to[b,B]

be surjections that are continuous and strictly increasing. Let UU be the region in the upper halfspace {z>0}\{z>0\} between the graphs of fa,bf_{a,b} and fA,Bf_{A,B}. By the maximum principle, the interior of MM lies in UU.

For t[,]t\in[-\infty,\infty], let ft=fα(t),β(t)f^{t}=f_{\alpha(t),\beta(t)}.

Note that the graphs of ftf^{t} with t𝐑t\in\mathbf{R} form a foliation \mathscr{F} of UU. Let T:U𝐑T:U\to\mathbf{R} be the function such that

T(p)=tfor pgraph(ft).T(p)=t\quad\text{for $p\in\operatorname{graph}(f^{t}).$}

Let M=MMM^{\prime}=M\setminus\partial M. Since T|MT|M^{\prime} is proper and since MM^{\prime} includes no boundary points,

𝖭(T|M)χ(M)=0\mathsf{N}(T|M^{\prime})\leq\chi(M^{\prime})=0

by Theorem 4.3. That is, MM^{\prime} is everywhere transverse to the foliation \mathscr{F}.

For pUp\in U, note that graph(fT(p))\operatorname{graph}(f^{T(p)}) is the leaf that passes through pp. For x[x(M),a)x\in[x(M),a), let

u(x)=xfT(x,0)(ϕlower)(x),where T=T(x,0,ϕlower(x)).u(x)=\partial_{x}f^{T}(x,0)-(\phi^{\textnormal{lower}})^{\prime}(x),\quad\text{where $T=T(x,0,\phi^{\textnormal{lower}}(x))$}.

Now MM and graph(fT)\operatorname{graph}(f^{T}) are not tangent at p=(x,0,ϕlower(x))p=(x,0,\phi^{\textnormal{lower}}(x)). Since 𝐞2\mathbf{e}_{2} is tangent to MM and to graph(fT)\operatorname{graph}(f^{T}) at pp, we see that M{y=0}M\cap\{y=0\} and graph(fT){y=0}\operatorname{graph}(f^{T})\cap\{y=0\} are not tangent at pp. In other words, u(x)0u(x)\neq 0. We have shown that u()u(\cdot) never vanishes on [x(M),a)[x(M),a). Now u(x)=u(x)=\infty for x=x(M)x=x(M). Thus u(x)>0u(x)>0 for all x[x(M),a)x\in[x(M),a). Consequently, u(x)0u(x)\geq 0 for all x[x(M),a]x\in[x(M),a]. Thus if x[x(M),a]x\in[x(M),a] and if T=T(x,0,ϕlower(x))T=T(x,0,\phi^{\textnormal{lower}}(x)), then

(ϕlower)(x)\displaystyle(\phi^{\textnormal{lower}})^{\prime}(x) xfT(x,0)\displaystyle\leq\partial_{x}f^{T}(x,0)
=xfα(T),β(T)(x,0)\displaystyle=\partial_{x}f_{\alpha(T),\beta(T)}(x,0)
<xfb(x,0),\displaystyle<\partial_{x}f_{b}(x,0),

where the last inequality is by Lemma 9.3 below. This completes the proof of Theorem 9.1. ∎

Lemma 9.3.

If π/2bβ\pi/2\leq b\leq\beta, then

xfα,β(x,0)<xfb(x,0)\frac{\partial}{\partial x}f_{\alpha,\beta}(x,0)<\frac{\partial}{\partial x}f_{b}(x,0)

for all x(0,α]x\in(0,\alpha].

To prove the lemma, it suffices to prove that

(i) xfα,β(x,0)xfb(x,0) for each x(0,α),\text{$\partial_{x}f_{\alpha,\beta}(x,0)\neq\partial_{x}f_{b}(x,0)$ for each $x\in(0,\alpha)$},

and that

(ii) xfα,β(α,0)<xfb(α,0)\partial_{x}f_{\alpha,\beta}(\alpha,0)<\partial_{x}f_{b}(\alpha,0).
Proof of (i).

Let Σ=graph(fα,β)\Sigma=\operatorname{graph}(f_{\alpha,\beta}).

Define FbF_{b} on the slab W:={|y|<b}W:=\{|y|<b\} by

Fb(x,y,z)=zfb(x,y).F_{b}(x,y,z)=z-f_{b}(x,y).

Now Fb|ΣF_{b}|\partial\Sigma has exactly two critical points, namely (α,0,0)(\alpha,0,0) and (α,0,0)(-\alpha,0,0). Thus

𝖭(Fb|Σ)\displaystyle\mathsf{N}(F_{b}|\Sigma) 2χ(Σ{|y|<β}\displaystyle\leq 2-\chi(\Sigma\cap\{|y|<\beta\}
=1\displaystyle=1

Thus Fb|ΣF_{b}|\Sigma has at most one critical point. Equivalently, the function

w:=fα,βfbw:=f_{\alpha,\beta}-f_{b}

has at most one critical point in the rectangle R:=(α,α)×(β,β)R:=(-\alpha,\alpha)\times(-\beta,\beta). Since (0,0)(0,0) is a critical point, it is the only one.

Thus if x(0,α)x\in(0,\alpha), then Dw(x,0)0Dw(x,0)\neq 0. Now yw(x,0)=0\partial_{y}w(x,0)=0 since w(x,y)w(x,y)w(x,y)\equiv w(x,-y). Thus

xw(x,0)0.\partial_{x}w(x,0)\neq 0.

This completes the proof of (i). ∎

Proof of (ii).

Let u=fbfα,βu=f_{b}-f_{\alpha,\beta} on [α,α]×(β,β)[-\alpha,\alpha]\times(-\beta,\beta). Now uu is proper and bounded above, so it attains a maximum. By the maximum principle, the maximum is attained at the boundary, that is at a point (±α,y)(\pm\alpha,y). By symmetry, it is attained at (α,y)(\alpha,y) for some yy. In fact, the maximum of u(α,y)u(\alpha,y) occurs at y=0y=0.

Thus we have shown

maxu=u(A,0).\max u=u(A,0).

By the strong maximum principle, ux(A,0)>0u_{x}(A,0)>0, i.e., that xfb(A,0)>xfα,β(A,0)\partial_{x}f_{b}(A,0)>\partial_{x}f_{\alpha,\beta}(A,0). This completes the proof of (ii), and thus the proof of Lemma 9.3. ∎

10. Sideways Graphicality

We now show that large portions of M𝒞M\in\mathscr{C} and M𝒜M\in\mathscr{A} are graphical sideways in that they can be expressed as y=y(x,z)y=y(x,z).

In this section, we let

hc(x,y)=log(cos(xc))for xIc:=(cπ/2,c+π/2),\displaystyle h_{c}(x,y)=\log(\cos(x-c))\qquad\text{for $x\in I_{c}:=(c-\pi/2,c+\pi/2)$},
Hc(x,y,z)=zhc(x,y)for (x,y,z)Wc:=Ic×𝐑2.\displaystyle H_{c}(x,y,z)=z-h_{c}(x,y)\qquad\text{for $(x,y,z)\in W_{c}:=I_{c}\times\mathbf{R}^{2}$}.

Thus the graph of hch_{c} is an untilted grim reaper over the strip Ic×𝐑I_{c}\times\mathbf{R}.

Proposition 10.1.

Suppose that MM is a surface in 𝒞\mathscr{C} or in 𝒜\mathscr{A}, and that

cc^:=x(M)+π/2.c\geq\hat{c}:=x(M)+\pi/2.

Then

𝖭(Hc|M{y0})=0.\mathsf{N}(H_{c}|M\cap\{y\neq 0\})=0.
Proof.

By lower semicontinuity (Theorem 4.2), it suffices to prove it for M𝒞M\in\mathscr{C}. By Theorem 8.5, M{x>x(M)}=MupperMlowerM\cap\{x>x(M)\}=M^{\textnormal{upper}}\cup M^{\textnormal{lower}}, so it suffices to prove that

𝖭(Hc|Σ{y0})=0\mathsf{N}(H_{c}|\Sigma\cap\{y\neq 0\})=0

for Σ=Mupper\Sigma=M^{\textnormal{upper}} and for Σ=Mlower\Sigma=M^{\textnormal{lower}}. The proofs for the two cases are essentially the same, so we give the proof for MlowerM^{\textnormal{lower}}.

We may assume that c<a(M)+(π/2)c<a(M)+(\pi/2), as otherwise MlowerWcM^{\textnormal{lower}}\cap W_{c} is empty.

If c<a(M)c<a(M), then Hc|MlowerH_{c}|\partial M^{\textnormal{lower}} has exactly two local minima, namely the two points on {x=c}innerM\{x=c\}\cap\partial_{\textnormal{inner}}M.

If c[a(M),a(M)+π)c\in[a(M),a(M)+\pi), then Hc|MlowerH_{c}|\partial M^{\textnormal{lower}} has exactly one local minimum, namely {x=c}innerM\{x=c\}\cap\partial_{\textnormal{inner}}M, which is either a point or an interval. (If it is an interval, we identify that interval to a point as in Remark 4.4.) Thus

𝖭(Hc|Mlower)|S||T|χ(MlowerW)201=1.\mathsf{N}(H_{c}|M^{\textnormal{lower}})\leq|S|-|T|-\chi(M^{\textnormal{lower}}\cap W)\leq 2-0-1=1.

(Note that χ1\chi\geq 1 by Lemma 6.5.) Consequently,

1\displaystyle 1 𝖭(Hc|Mlower)\displaystyle\geq\mathsf{N}(H_{c}|M^{\textnormal{lower}})
𝖭(Hc|Mlower{y0})\displaystyle\geq\mathsf{N}(H_{c}|M^{\textnormal{lower}}\cap\{y\neq 0\})
=2𝖭(Hc|Mlower{y>0}).\displaystyle=2\mathsf{N}(H_{c}|M^{\textnormal{lower}}\cap\{y>0\}).

It follows immediately that 𝖭(Hc|Mlower{y0})=0\mathsf{N}(H_{c}|M^{\textnormal{lower}}\cap\{y\neq 0\})=0. ∎

Corollary 10.2.

Suppose that p=(x0,y0,z0)p=(x_{0},y_{0},z_{0}) is an interior point of MM and that x0c^:=x(M)+πx_{0}\geq\hat{c}:=x(M)+\pi. Then 𝐞2Tan(M,p)\mathbf{e}_{2}\in\operatorname{Tan}(M,p) if and only if y0=0y_{0}=0.

Proof.

The “if” assertion follows from the fact that {y=0}\{y=0\} is a plane of symmetry of MM. Suppose that 𝐞2Tan(M,p)\mathbf{e}_{2}\in\operatorname{Tan}(M,p). Then there is a grim reaper surface GG that is tangent to MM at pp and that contains the line {p+s𝐞2:s𝐑}\{p+s\mathbf{e}_{2}:s\in\mathbf{R}\}. Note that GG is the level set {Hc=Hc(p)}\{H_{c}=H_{c}(p)\} for some HcH_{c} as in Proposition 10.1. Now pp is a critical point of Hc|MH_{c}|M, so y0=0y_{0}=0 by Proposition 10.1. ∎

Theorem 10.3.

Suppose M𝒞M\in\mathscr{C} or M𝒜M\in\mathscr{A}. Let c^=x(M)+π\hat{c}=x(M)+\pi. If M𝒜M\in\mathscr{A}, let me know

Ωin(M)\displaystyle\Omega^{\textnormal{in}}(M) ={(x,z):xc^ and zϕlower(x)},\displaystyle=\{(x,z):\text{$x\geq\hat{c}$ and $z\leq\phi^{\textnormal{lower}}(x)$}\},
Ωout(M)\displaystyle\Omega^{\textnormal{out}}(M) ={(x,z):xc^ and zϕupper(x)}.\displaystyle=\{(x,z):\text{$x\geq\hat{c}$ and $z\leq\phi^{\textnormal{upper}}(x)$}\}.

If M𝒞M\in\mathscr{C}, let

Ωin(M)\displaystyle\Omega^{\textnormal{in}}(M) ={(x,z):x[c^,a) and 0zϕlower(x)},\displaystyle=\{(x,z):\text{$x\in[\hat{c},a)$ and $0\leq z\leq\phi^{\textnormal{lower}}(x)$}\},
Ωout(M)\displaystyle\Omega^{\textnormal{out}}(M) ={(x,z):x[c^,A) and 0zϕupper(x)}.\displaystyle=\{(x,z):\text{$x\in[\hat{c},A)$ and $0\leq z\leq\phi^{\textnormal{upper}}(x)$}\}.

There exist nonnegative, continuous functions yinnery^{\textnormal{inner}} on Ωin\Omega^{\textnormal{in}} and youtery^{\textnormal{outer}} on Ωout\Omega^{\textnormal{out}} whose graphs y=yinner(x,z)y=y^{\textnormal{inner}}(x,z) and y=youter(x,z)y=y^{\textnormal{outer}}(x,z) are contained in MM. Furthermore, M{xc^}{y0}M\cap\{x\geq\hat{c}\}\cap\{y\geq 0\} has exactly two components, namely graph(yinner)\operatorname{graph}(y^{\textnormal{inner}}) and graph(youter)\operatorname{graph}(y^{\textnormal{outer}}) if M𝒜M\in\mathscr{A}, or

graph(yinner)(innerM{x=a})\operatorname{graph}(y^{\textnormal{inner}})\cup(\partial_{\textnormal{inner}}M\cap\{x=a\})

and

graph(youter)(outerM{x=A})\operatorname{graph}(y^{\textnormal{outer}})\cup(\partial_{\textnormal{outer}}M\cap\{x=A\})

if M𝒞M\in\mathscr{C}.

The functions yinnery^{\textnormal{inner}} and youtery^{\textnormal{outer}} are strictly positive except where z=ϕlower(x)z=\phi^{\textnormal{lower}}(x) (for yinnery^{\textnormal{inner}}) and z=ϕupper(x)z=\phi^{\textnormal{upper}}(x) (for youtery^{\textnormal{outer}}). The functions are smooth on the interior of their domains.

Proof.

It suffices to prove it for M𝒞M\in\mathscr{C}, since the M𝒜M\in\mathscr{A} case follows by taking limits.

For each x0[c^,a)x_{0}\in[\hat{c},a), the height function z()z(\cdot) on the smooth curve Mlower{x=x0}{y0}M^{\textnormal{lower}}\cap\{x=x_{0}\}\cap\{y\geq 0\} has only one critical point, namely the endpoint (x0,0,ϕlower(x0))(x_{0},0,\phi^{\textnormal{lower}}(x_{0})). Thus z()z(\cdot) maps the curve homeomorphically onto [0,ϕlower(x)][0,\phi^{\textnormal{lower}}(x)]. Hence there is a continuous, nonnegative function yinnery^{\textnormal{inner}} on Ωin\Omega^{\textnormal{in}} such that yinner(x,z)Mlowery^{\textnormal{inner}}(x,z)\in M^{\textnormal{lower}} for each (x,z)Ωin(x,z)\in\Omega^{\textnormal{in}}. The graph {(x,yinner(0,z),z):(x,z)Ωin}\{(x,y^{\textnormal{inner}}(0,z),z):(x,z)\in\Omega^{\textnormal{in}}\} of yinnery^{\textnormal{inner}} is a smooth manifold (namely Mlower{c^x<a}M^{\textnormal{lower}}\cap\{\hat{c}\leq x<a\}) and 𝐞2\mathbf{e}_{2} is not tangent to that manifold except where y=ϕlower(x)y=\phi^{\textnormal{lower}}(x). Thus yinnery^{\textnormal{inner}} is a smooth function except where y=ϕlower(x)y=\phi^{\textnormal{lower}}(x). This completes the proof in the case of yinnery^{\textnormal{inner}}.

The same proof works for youtery^{\textnormal{outer}}. ∎

Corollary 10.4.

Suppose pM{xc^}{y>0}p\in M\cap\{x\geq\hat{c}\}\cap\{y>0\}. Then

ν(M,p)𝐞2>0\displaystyle\nu(M,p)\cdot\mathbf{e}_{2}>0 if pMupper,\displaystyle\qquad\text{if $p\in M^{\textnormal{upper}}$},
ν(M,p)𝐞2<0\displaystyle\nu(M,p)\cdot\mathbf{e}_{2}<0 if pMlower.\displaystyle\qquad\text{if $p\in M^{\textnormal{lower}}$}.
Theorem 10.5.

Suppose M𝒜M\in\mathscr{A}. Then

  1. (1)

    MM is connected.

  2. (2)

    M{x>x(M)}M\cap\{x>x(M)\} has exactly two connected components, MlowerM^{\textnormal{lower}} and MupperM^{\textnormal{upper}}.

Proof.

Let MM^{\prime} be the connected component of MM containing

M{y=0}{x>0}.M\cap\{y=0\}\cap\{x>0\}.

By Theorem 10.3, MMM\setminus M^{\prime} is contained in the half-slab {|y|B}{x<x(M)+π}\{|y|\leq B\}\cap\{x<x(M)+\pi\}. But there are no nonempty translators (without boundary) in such a half-slab (Corollary A.3), so MMM\setminus M^{\prime} is empty and therefore MM is connected. Thus we have proved Assertion (1).

Now let

Σ=(M{x>x(M)})(MupperMlower).\Sigma=(M\cap\{x>x(M)\})\setminus(M^{\textnormal{upper}}\cup M^{\textnormal{lower}}).

Suppose, contrary to Assertion (2), that Σ\Sigma is nonempty. Then

(19) x(M)<supΣx().x(M)<\sup_{\Sigma}x(\cdot).

Now Σ\partial\Sigma is nonempty by Assertion (1), and x()x(M)x(\cdot)\equiv x(M) on Σ\partial\Sigma, so

(20) supΣx()=x(M).\sup_{\partial\Sigma}x(\cdot)=x(M).

By Theorem 10.3,

supΣx()x(M)+π.\sup_{\Sigma}x(\cdot)\leq x(M)+\pi.

Thus, by a version of the maximum principle in slabs (Corollary A.3),

supΣx()\displaystyle\sup_{\Sigma}x(\cdot) =supΣx()\displaystyle=\sup_{\partial\Sigma}x(\cdot)
=x(M)\displaystyle=x(M)

by (20). But this contradicts (19). Hence Assertion (2) holds. ∎

Theorem 10.6.

Suppose M𝒜M\in\mathscr{A}. Then

  1. (1)

    MM is an annulus.

  2. (2)

    M(0,0,z)M-(0,0,z) converges as zz\to\infty to the empty set.

  3. (3)

    There are numbers b(M)b(M) and B(M)B(M) with 0b(M)B(M)0\leq b(M)\leq B(M) such that M+(0,0,z)M+(0,0,z) converges smoothly as zz\to\infty to the planes {y=±b(M)}\{y=\pm b(M)\} a and {y=±B(M)}\{y=\pm B(M)\}.

  4. (4)

    For each xc^:=x(M)+πx\geq\hat{c}:=x(M)+\pi,

    limzyinner(x,z)\displaystyle\lim_{z\to-\infty}y^{\textnormal{inner}}(x,z) =b(M),\displaystyle=b(M),
    limzyouter(x,z)\displaystyle\lim_{z\to-\infty}y^{\textnormal{outer}}(x,z) =B(M).\displaystyle=B(M).
Proof.

Since MM is connected and has an annular component (Theorem 7.3), it is an annulus.

By Theorem 5.8, M(0,0,ζ)M-(0,0,\zeta) converges smoothly as ζ\zeta\to\infty to a limit MM_{\infty} that is a finite union of planes in the slab {|y|B}\{|y|\leq B\}. By Theorem 10.3, MM_{\infty} lies in the region {|y|x(M)+π}\{|y|\leq x(M)+\pi\}. Thus MM_{\infty} is empty.

By Theorem 5.8, M+(0,0,ζ)M+(0,0,\zeta) converges smoothly as ζ\zeta\to\infty to a finite union MM_{-\infty} of planes. From Theorem 10.3, we see that the number of planes is four (counting multiplicity). Thus the planes have the form {y=±b(M)}\{y=\pm b(M)\} and {y=±B(M)}\{y=\pm B(M)\} for some 0b(M)B(M)0\leq b(M)\leq B(M). It follows trivially that yinner(x,z)b(M)y^{\textnormal{inner}}(x,z)\to b(M) and youter(x,z)B(M)y^{\textnormal{outer}}(x,z)\to B(M) for every xc^x\geq\hat{c}. ∎

Theorem 10.7.

Suppose that MM is a surface in 𝒞\mathscr{C} (see Definition 6.1) with smooth, uniformly convex boundary curves, or that MM is a surface in \mathscr{R}. Then the surface S:=M{z>0}{y(b+B)/2}S:=M\cap\{z>0\}\cap\{y\geq(b+B)/2\} projects diffeomorphically onto its image in the xzxz-plane, and MM is disjoint from the open region in {z>0}\{z>0\} between SS and its image under reflection in the plane {y=(b+B)/2}\{y=(b+B)/2\}.

Proof.

For MM with smooth boundary, the proof is by the standard standard Alexandrov moving planes argument, so we omit it. The theorem for MM\in\mathscr{R} follows immediately by taking limits. ∎

Corollary 10.8.

Suppose MM\in\mathscr{R}. Then

yinner+youterb+By^{\textnormal{inner}}+y^{\textnormal{outer}}\leq b+B

at all points of domain(yinner)\operatorname{domain}(y^{\textnormal{inner}}).

Theorem 10.9.

Suppose that M𝒞M\in\mathscr{C} with smooth, strictly convex boundaries, or that MM\in\mathscr{R}. Then Mlower{z>0}M^{\textnormal{lower}}\cap\{z>0\} projects diffeomorphically onto its image in the yzyz plane. In particular, in the domain of yinnery^{\textnormal{inner}}, for each zz, yinner(x,z)y^{\textnormal{inner}}(x,z) is a decreasing function of xx.

Proof.

First consider the case that M𝒞M\in\mathscr{C} with smooth, strictly convex boundaries. The proof in this case is the following standard Alexandrov argument. For t(x(M),a(M)]t\in(x(M),a(M)], let Σ(t)\Sigma(t) be the image of Mlower{xt}M^{\textnormal{lower}}\cap\{x\geq t\} under reflection in the plane {x=t}\{x=t\}. Let 𝒯\mathscr{T} be the set of t[x(M),a(M)]t\in[x(M),a(M)] such that Σ(t)\Sigma(t) is disjoint from M{x<t}M\cap\{x<t\} and such that ν𝐞1<0\nu\cdot\mathbf{e}_{1}<0 at all points of Σ(t)\Sigma(t). Now 𝒯\mathscr{T} is nonempty since t𝒯t\in\mathscr{T} for all tt sufficiently close to a(M)a(M). Clearly 𝒯\mathscr{T} is an open subset of (x(M),a(M))(x(M),a(M)). By the strong maximum principle, it is also a relatively closed subset. The statement for \mathscr{R} follows by taking the limit. ∎

11. Vertical Graphicality

In this section, we show that if MM\in\mathscr{R}, then a large portion of MM can be written as a graph z=uM(x,y)z=u_{M}(x,y).

Definition 11.1 (Waist).

If MM is a translator and if M~\tilde{M} is the set of non-flat components of MM (i.e., the set of components that are not contained in vertical planes), then waist(M)\operatorname{waist}(M) is the set of points in M~\tilde{M} at which the tangent plane is vertical.

For M𝒞M\in\mathscr{C}, then no component of MM is flat, and therefore waist(M)\operatorname{waist}(M) is the preimage of the equator

E:={(x,y,z)SS2:z=0}E:=\{(x,y,z)\in\SS^{2}:z=0\}

under the Gauss map. If pp is in the waist, then the Gauss Curvature of MM at pp is negative (by Theorem 6.6), so ν\nu maps a neighborhood of pp in MM diffeomorphically onto an open set in SS2\SS^{2}. Thus waist(M)\operatorname{waist}(M) is a smooth curve, and ν\nu is a smooth immersion from waist(M)\operatorname{waist}(M) onto its image. By Theorem 6.6, the map

ν|waist(M):waist(M)E\nu|_{\operatorname{waist}(M)}:\operatorname{waist}(M)\longrightarrow E

is one-to-one, so ν\nu is a diffeomorphism from waist(M)\operatorname{waist}(M) onto an open subset of EE.

Theorem 11.2.

Suppose that MM is a surface in 𝒞\mathscr{C} with smooth boundary curves, or a surface in \mathscr{R}. Suppose that pp is a point in MMM\setminus\partial M where Tan(M,p)\operatorname{Tan}(M,p) is vertical. If x(p)0x(p)\neq 0, then ν𝐞1\nu\cdot\mathbf{e}_{1} and x(p)x(p) have opposite signs, and if y(p)0y(p)\neq 0, then ν𝐞2\nu\cdot\mathbf{e}_{2} and y(p)y(p) have opposite signs.

(Theorem 11.2 is true, with essentially the same proof for any M𝒞M\in\mathscr{C}, but the lack of regularity of M\partial M makes the proof a bit more involved.)

Proof.

If y(p)=0y(p)=0, then pp must be one of the points (±x(M),0,z(M))(\pm x(M),0,z(M)), and the assertion holds at those points. Likewise, it holds if x(p)=0x(p)=0.

Thus we may assume that neither x(p)x(p) nor y(p)y(p) is 0. By symmetry, it suffices to consider the case when x(p)>0x(p)>0 and y(p)>0y(p)>0. Let JJ be the connected component of waist(M){x>0,y>0}\operatorname{waist}(M)\cap\{x>0,\,y>0\} containing pp.

Let qq be an endpoint of JJ. We claim that

(*) ν(q)𝐞i0 for i=1,2.\text{$\nu(q)\cdot\mathbf{e}_{i}\leq 0$ for $i=1,2$}.

To prove (*11), note that if qq is an endpoint of JJ, then one of the following must hold:

  1. (1)

    qq is a point in M\partial M where Tan(M,q)\operatorname{Tan}(M,q) is vertical, or

  2. (2)

    qq is a point in MMM\setminus\partial M with y(q)=0y(q)=0, or

  3. (3)

    qq is a point in MMM\setminus\partial M with x(q)=0x(q)=0.

In case (1), qq is in innerM\partial_{\textnormal{inner}}M since, by the maximum principle, Tan(M,)\operatorname{Tan}(M,\cdot) is never vertical on outerM\partial_{\textnormal{outer}}M. Thus (*11) holds in this case.

In case (2), by Theorem 8.1, x=x(M)x=x(M) and q=(x(M),0,z(M))q=(x(M),0,z(M)), so ν(q)=𝐞1\nu(q)=-\mathbf{e}_{1} and therefore (*11) holds.

Likewise, in case (3), ν(q)=𝐞2\nu(q)=-\mathbf{e}_{2}, so (*11) holds.

Now ν\nu is a homeomorphism of J¯\overline{J} onto its image. Note that if (x,y,z)J(x,y,z)\in J, then ν(x,y,z)\nu(x,y,z) cannot be ±𝐞1\pm\mathbf{e}_{1}, since if it were, ν(x,y,z)\nu(x,y,z) and ν(x,y,z)\nu(x,-y,z) would be equal, contrary to Theorem 6.6. For the same reason, ν(x,y,z)\nu(x,y,z) cannot be ±𝐞2\pm\mathbf{e}_{2}. Thus ν(J)\nu(J) lies in one of the four components of

E{𝐞1,𝐞2,𝐞1,𝐞2}.E\setminus\{\mathbf{e}_{1},\mathbf{e}_{2},-\mathbf{e}_{1},-\mathbf{e}_{2}\}.

By (*11), the endpoints of ν(J¯)\nu(\overline{J}) lie in {vE:v𝐞10,v𝐞20}\{v\in E:v\cdot\mathbf{e}_{1}\leq 0,\,v\cdot\mathbf{e}_{2}\leq 0\}. Thus JJ lies in {vE:v𝐞1<0,v𝐞2<0}\{v\in E:v\cdot\mathbf{e}_{1}<0,\,v\cdot\mathbf{e}_{2}<0\}. ∎

Definition 11.3.

Suppose MM is in 𝒞\mathscr{C} or 𝒜\mathscr{A}. We define a function u=uMu=u_{M} as follows. The domain of uMu_{M} is the set of (x,y)(x,y) such that {z:(x,y,z)M}\{z:(x,y,z)\in M\} is nonempty and has a greatest element ζ\zeta and such that the tangent plane to MM at (x,y,ζ)(x,y,\zeta) is not vertical. For (x,y)domain(uM)(x,y)\in\operatorname{domain}(u_{M}), we let

uM(x,y)=max{z:(x,y,z)M}.u_{M}(x,y)=\max\{z:(x,y,z)\in M\}.

If MM is smooth and outerM\partial_{\textnormal{outer}}M is smooth, then uMu_{M} is smooth. For MM\in\mathscr{R}, uMu_{M} is C1C^{1}, and it is smooth except at the corners (±A,±B)(\pm A,\pm B).

Recall that for M𝒞M\in\mathscr{C} or M𝒜M\in\mathscr{A},

x(M):=min{x>0:(x,0,z)M for some z}.x(M):=\min\{x>0:\text{$(x,0,z)\in M$ for some $z$}\}.

Likewise, we let

(21) y(M):=min{y>0:(0,y,z)M for some z}.y(M):=\min\{y>0:\text{$(0,y,z)\in M$ for some $z$}\}.
Theorem 11.4.

For MM\in\mathscr{R},

[c^,A]×[B,B]domain(uM)[\hat{c},A]\times[-B,B]\subset\operatorname{domain}(u_{M})

and

yuM(x,y)<0on [c^,A)×(0,B],\frac{\partial}{\partial y}u_{M}(x,y)<0\quad\text{on $[\hat{c},A)\times(0,B]$},

where c^:=x(M)+π\hat{c}:=x(M)+\pi.

Likewise, if B>y(M)+πB>y(M)+\pi, then

[A,A]×[y(M)+π,B]domain(uM),[-A,A]\times[y(M)+\pi,B]\subset\operatorname{domain}(u_{M}),

and

xuM(x,y)<0on (0,A]×[y(M)+π,B).\frac{\partial}{\partial x}u_{M}(x,y)<0\quad\text{on $(0,A]\times[y(M)+\pi,B)$.}
Proof.

Let Σ=Mupper{xc^}\Sigma=M^{\textnormal{upper}}\cap\{x\geq\hat{c}\}. By Corollary 10.4,

ν𝐞2>0\nu\cdot\mathbf{e}_{2}>0 on Σ{y>0}\Sigma\cap\{y>0\}.

If pM{y>0}p\in M\cap\{y>0\} and ν𝐞3=0\nu\cdot\mathbf{e}_{3}=0, then ν𝐞2<0\nu\cdot\mathbf{e}_{2}<0 by Theorem 11.2, and thus pΣp\notin\Sigma. Hence ν𝐞3\nu\cdot\mathbf{e}_{3} is never 0 on Σ{y>0}\Sigma\cap\{y>0\}. By Theorem 8.2, ν𝐞3>0\nu\cdot\mathbf{e}_{3}>0 on Σ{y=0}\Sigma\cap\{y=0\}. Thus ν𝐞3>0\nu\cdot\mathbf{e}_{3}>0 on Σ{y>0}\Sigma\cap\{y>0\} since Σ{y>0}\Sigma\cap\{y>0\} is connected.

Note that on Ωout\Omega^{\textnormal{out}},

youterz=ν𝐞3ν𝐞2<0,\frac{\partial y^{\textnormal{outer}}}{\partial z}=-\frac{\nu\cdot\mathbf{e}_{3}}{\nu\cdot\mathbf{e}_{2}}<0,

where youter(x,z)y^{\textnormal{outer}}(x,z) is as in Theorem 10.3. Thus for each x^[c^,A)\hat{x}\in[\hat{c},A), youter(x^,z)y^{\textnormal{outer}}(\hat{x},z) is a strictly decreasing function of zz. Then, we have that zz is a strictly decreasing function of yy on Σ{x=x^}\Sigma\cap\{x=\hat{x}\}. Notice that youter(x^,0)=By^{\textnormal{outer}}(\hat{x},0)=B and youter(x^,ϕupper(x^))=0y^{\textnormal{outer}}(\hat{x},\phi^{\textnormal{upper}}(\hat{x}))=0. Since this holds for each x^[c^,A)\hat{x}\in[\hat{c},A), we see that

Σ{y0}\Sigma\cap\{y\geq 0\}

is the graph of a function on [c^,A]×[0,B][\hat{c},A]\times[0,B]. By symmetry, Σ\Sigma is the graph of a function uu on [c^,A]×[B,B][\hat{c},A]\times[-B,B]. By Theorem 10.3 and Theorem 10.5, (M{xc^})Σ(M\cap\{x\geq\hat{c}\})\setminus\Sigma lies below the graph of uu. Thus u=uMu=u_{M}.

Furthermore,

uMy=ν𝐞2ν𝐞3<0on [c^,A)×(0,B).\frac{\partial u_{M}}{\partial y}=-\frac{\nu\cdot\mathbf{e}_{2}}{\nu\cdot\mathbf{e}_{3}}<0\quad\text{on $[\hat{c},A)\times(0,B)$}.

The last assertion of the theorem follows by reversing the roles of xx and yy. ∎

12. The Inner and Outer Portions of an Annuloid

In this section, we describe a way of dividing MM\in\mathscr{R} (or M𝒜M\in\mathscr{A}) into an inner and an outer portion, and we show that the outer portion is well-behaved. In particular, we show that ν𝐞3\nu\cdot\mathbf{e}_{3} is everywhere positive on the outer portion, and we get good control on the slopes of the tangent planes in the outer portion. This control is the key to showing that if M𝒜(b,B,x^)M\in\mathscr{A}(b,B,\hat{x}), then B(M)=BB(M)=B. See Theorem 12.8.

The idea is the following. For points in the outer portion that are not in the plane {y=0}\{y=0\}, the unit normal should point away from that plane, whereas on the inner portion, the unit normal should point toward the plane. Thus, except on the plane {y=0}\{y=0\}, the outer portion should be the points where (ν𝐞2)/y>0(\nu\cdot\mathbf{e}_{2})/y>0, and the inner portion should be the points where (ν𝐞2)/y<0(\nu\cdot\mathbf{e}_{2})/y<0. What about on M{y=0}M\cap\{y=0\}? The function (ν𝐞2)/y(\nu\cdot\mathbf{e}_{2})/y extends to a smooth function fMf_{M} on all of MM, so we extend the notion of outer and inner to all of MM by defining a point to be outer or inner according to whether fMf_{M} is >0>0 or <0<0.

Definition 12.1.

Suppose that MM is an oriented surface in 𝐑3\mathbf{R}^{3} such that reflection in the plane {y=0}\{y=0\} is an orientation-reversing isometry of MM. We let fM:M𝐑f_{M}:M\to\mathbf{R} be the smooth function such that

fM=ν𝐞2yon M{y0}.f_{M}=\frac{\nu\cdot\mathbf{e}_{2}}{y}\quad\text{on $M\cap\{y\neq 0\}$.}

We let

MouterY:={pM:fM(p)>0},\displaystyle M^{Y}_{\textnormal{outer}}:=\{p\in M:f_{M}(p)>0\},
MinnerY:={pM:fM(p)<0}.\displaystyle M^{Y}_{\textnormal{inner}}:=\{p\in M:f_{M}(p)<0\}.

(The function fMf_{M} exists because for any smooth function uu on MM with u(x,y,z)u(x,y,z)u(x,y,z)\equiv-u(x,-y,z), the function u(x,y,z)/yu(x,y,z)/y on M{y0}M\cap\{y\neq 0\} extends smoothly to MM.)

Note that

(22) fM(p)=𝐞2(ν𝐞2)=(𝐞2ν)𝐞2 for pM{y=0}.\text{$f_{M}(p)=\nabla_{\mathbf{e}_{2}}(\nu\cdot\mathbf{e}_{2})=(\nabla_{\mathbf{e}_{2}}\nu)\cdot\mathbf{e}_{2}$ for $p\in M\cap\{y=0\}$}.

If pM{y=0}p\in M\cap\{y=0\}, then 𝐞2\mathbf{e}_{2} is one of the principal directions of MM at pp. Thus, by (22),

(23) If pM{y=0}p\in M\cap\{y=0\} and if the Gauss curvature is 0\neq 0, then fM(p)0f_{M}(p)\neq 0.

For MM\in\mathscr{R}, note that fMf_{M} has the following behavior on M\partial M:

(24) fM>0\displaystyle f_{M}>0 on (A,A)×{±B}×{0},\displaystyle\quad\text{on $(-A,A)\times\{\pm B\}\times\{0\}$},
fM<0\displaystyle f_{M}<0 on (a,a)×{±b}×{0},\displaystyle\quad\text{on $(-a,a)\times\{\pm b\}\times\{0\}$},
fM=0\displaystyle f_{M}=0  on {±a}×[b,b]×{0},\displaystyle\quad\text{ on $\{\pm a\}\times[-b,b]\times\{0\}$},
fM=0\displaystyle f_{M}=0  on {±A}×[B,B]×{0}.\displaystyle\quad\text{ on $\{\pm A\}\times[-B,B]\times\{0\}$}.

The following is trivial, but useful:

Proposition 12.2.

Suppose that MM is in \mathscr{R} or in 𝒜\mathscr{A}. Suppose LL is a line parallel to the yy-axis. Let pp be the point in LML\cap M that maximizes y(p)y(p). Then ν(M,p)𝐞20\nu(M,p)\cdot\mathbf{e}_{2}\geq 0.

In particular, if MM and LL cross transversely at pp, then pp is in MouterYM^{Y}_{\textnormal{outer}}.

Proof.

It suffices to prove it for MM\in\mathscr{R}; the other case follows by taking limits. Let KK be the closure of the bounded component of {z0}M\{z\geq 0\}\setminus M. By definition, ν(M,)\nu(M,\cdot) is the unit normal vector field to MM that points out of KK. Proposition 12.2 follows immediately. ∎

Lemma 12.3.

If MM\in\mathscr{R}, then

  1. (1)

    the surface Mupper{x(x(M)+π,A)}M^{\textnormal{upper}}\cap\{x\in(x(M)+\pi,A)\} is contained in MouterYM^{Y}_{\textnormal{outer}}, and

  2. (2)

    the surface Mlower{x(x(M)+π,a)}M^{\textnormal{lower}}\cap\{x\in(x(M)+\pi,a)\} is contained in MinnerYM^{Y}_{\textnormal{inner}}.

Proof.

Let MM^{\prime} be the surface in Assertion (1). By Theorem 10.3, 𝐞2ν\mathbf{e}_{2}\cdot\nu is >0>0 on all the points on M{y>0}M^{\prime}\cap\{y>0\}. Thus those points are in MouterYM^{Y}_{\textnormal{outer}}. By symmetry, M{y<0}M^{\prime}\cap\{y<0\} is also contained in MouterYM^{Y}_{\textnormal{outer}}. Now ν𝐞2\nu\cdot\mathbf{e}_{2} is a Jacobi field that is >0>0 in M{y>0}M^{\prime}\cap\{y>0\} and that vanishes along M{y=0}M^{\prime}\cap\{y=0\}. Thus, by the boundary maximum principle [hopf],

𝐞2(ν𝐞2)>0\nabla_{\mathbf{e}_{2}}(\nu\cdot\mathbf{e}_{2})>0 on M{y=0}M^{\prime}\cap\{y=0\}.

Thus M{y=0}M^{\prime}\cap\{y=0\} is contained in MouterYM^{Y}_{\textnormal{outer}} (by (22)).

The proof of Assertion (2) is essentially the same. ∎

Definition 12.4.

Given λ(0,)\lambda\in(0,\infty), let (λ)\mathscr{R}(\lambda) be the collection of MM\in\mathscr{R} such that

  1. (1)

    MM lies in the slab {|y|λ}\{|y|\leq\lambda\}.

  2. (2)

    a(M)x(M)+2πa(M)\geq x(M)+2\pi.

The following theorem shows that the tangent planes to M(λ)M\in\mathscr{R}(\lambda) at points where fM=0f_{M}=0 have uniformly bounded slopes.

Theorem 12.5.

There is an η=η(λ)<\eta=\eta(\lambda)<\infty with the following property. If pM(λ)p\in M\in\mathscr{R}(\lambda) and if fM(p)=0f_{M}(p)=0, then

|ν(M,p)𝐞3|1η.|\nu(M,p)\cdot\mathbf{e}_{3}|^{-1}\leq\eta.
Proof.

First we give a slope bound on the points of the boundary where fM=0f_{M}=0, i.e., on the edges {±A}×[B,B]×{0}\{\pm A\}\times[-B,B]\times\{0\} and {±a}×[b,b]×{0}\{\pm a\}\times[-b,b]\times\{0\}. Let u:[0,2π)×[λ,λ][0,)u:[0,2\pi)\times[-\lambda,\lambda]\to[0,\infty) be the translator that is \infty on the edge

{2π}×[λ,λ]\{2\pi\}\times[-\lambda,\lambda]

and that is 0 on the other three sides of the boundary. (For existence of the function uu, see Theorem A.1.)

Let s=maxy[λ,λ](/x)u(0,y)s=\max_{y\in[-\lambda,\lambda]}(\partial/\partial x)u(0,y). (By the boundary maximum principle, s<s<\infty.) By the maximum principle,

Mupper{x[A,A+π)}M^{\textnormal{upper}}\cap\{x\in[-A,-A+\pi)\}

lies under the graph of (x,y)u(x+A,y)(x,y)\mapsto u(x+A,y). Thus the slope of the tangent plane to MM at each point of of {A}×[B,B]×{0}\{-A\}\times[-B,B]\times\{0\} is bounded above by ss. By symmetry, the same bound holds on {A}×[B,B]×{0}\{A\}\times[-B,B]\times\{0\}. The same argument gives the same upper bound on the edges {±a}×[b,b]×{0}\{\pm a\}\times[-b,b]\times\{0\}.

Thus we have shown

(25) |ν𝐞3|1(1+s2)1/2on (M){fM=0}.|\nu\cdot\mathbf{e}_{3}|^{-1}\leq(1+s^{2})^{1/2}\quad\text{on $(\partial M)\cap\{f_{M}=0\}$.}

Now suppose that Theorem 12.5 is false. Then there exist Mn(λ)M_{n}\in\mathscr{R}(\lambda) and pnMnp_{n}\in M_{n} such that

fMn(pn)=0f_{M_{n}}(p_{n})=0

and such that

|ν(Mn,pn)𝐞3|0.|\nu(M_{n},p_{n})\cdot\mathbf{e}_{3}|\to 0.

Since ν(Mn,pn)𝐞2=0\nu(M_{n},p_{n})\cdot\mathbf{e}_{2}=0, it follows that (after passing to a subsequence)

(26) ν(Mn,pn)σ𝐞1\nu(M_{n},p_{n})\to\sigma\mathbf{e}_{1}, where σ\sigma is 11 or 1-1.

By (25), we may suppose that pnp_{n} is in the interior of MnM_{n} for all nn. By symmetry, we can assume that x(pn)0x(p_{n})\geq 0 for all nn. By Lemma 12.3, we see that x(pn)x(Mn)+πx(p_{n})\leq x(M_{n})+\pi, and thus

0x(pn)anπ0\leq x(p_{n})\leq a_{n}-\pi

for all nn.

By passing to a subsequence, we can assume that

  1. Case 1:

    dist(pn,Mn)\operatorname{dist}(p_{n},\partial M_{n}) is bounded away from 0 and dist(pn,Z)\operatorname{dist}(p_{n},Z) is bounded away from 0, or

  2. Case 2:

    dist(pn,Mn)\operatorname{dist}(p_{n},\partial M_{n}) is bounded away from 0 and δn:=dist(pn,Z)0\delta_{n}:=\operatorname{dist}(p_{n},Z)\to 0, or

  3. Case 3:

    dist(pn,Mn)0\operatorname{dist}(p_{n},\partial M_{n})\to 0 (and therefore z(pn)0z(p_{n})\to 0.)

In case 1, we let

Mn=Mn(x(pn),0,z(pn))M_{n}^{\prime}=M_{n}-(x(p_{n}),0,z(p_{n}))

and

pn=pn(x(pn),0,z(pn))=(0,y(pn),0).p_{n}^{\prime}=p_{n}-(x(p_{n}),0,z(p_{n}))=(0,y(p_{n}),0).

By passing to a subsequence, we can assume (by the fundamental compactness theorem for minimal surfaces of locally bounded area and genus [white18]) that the MnM_{n}^{\prime} converge to a smooth embedded limit MM^{\prime} (possibly with multiplicity), and that pnp_{n}^{\prime} converges to a limit p=(0,y(p),0)p^{\prime}=(0,y(p^{\prime}),0). By the curvature estimates in Theorem 3.4, the convergence is smooth in a neighborhood of pp^{\prime}. Thus,

(27) fM(p)=limfMn(pn)=limfMn(pn)=0f_{M^{\prime}}(p^{\prime})=\lim f_{M_{n}^{\prime}}(p_{n}^{\prime})=\lim f_{M_{n}}(p_{n})=0

and ν(M,p)=𝐞1\nu(M^{\prime},p^{\prime})=\mathbf{e}_{1} by (26). The component of MM^{\prime} containing pp^{\prime} cannot be contained in a plane, since if were, that component would be the plane {x=0}\{x=0\}, which is impossible since MM^{\prime} is contained in the slab {|y|λ}\{|y|\leq\lambda\}. By Corollary 6.7 (with 𝐯=𝐞1\mathbf{v}=\mathbf{e}_{1}), y(qn)=0y(q_{n})=0 and the Gauss curvature of MnM_{n} at qnq_{n} is nonzero. Thus

fM(p)0f_{M^{\prime}}(p^{\prime})\neq 0

by (23), contrary to (27). The contradiction proves the Theorem in Case 1.

In Case 2, let qnq_{n} be the point in ZZ closest to pnp_{n}. Let M~n=(Mnqn)/δn\tilde{M}_{n}=(M_{n}-q_{n})/\delta_{n} and p~n:=(pnqn)/δn\tilde{p}_{n}:=(p_{n}-q_{n})/\delta_{n}. By (26) and Lemma 12.10 (at the end of this section), M~n\tilde{M}_{n} converges smoothly to the standard catenoid M~\tilde{M} whose waist is the unit circle in the the plane {z=0}\{z=0\} and p~n\tilde{p}_{n} converges to the point p~=(1,0,0)\tilde{p}=(1,0,0). Now fMn(pn)=0f_{M_{n}}(p_{n})=0, so fM~n(p~n)=0f_{\tilde{M}_{n}}(\tilde{p}_{n})=0 and therefore, letting nn\to\infty,

fM~(p~)=0.f_{\tilde{M}}(\tilde{p})=0.

But the Gauss curvature of M~\tilde{M} at p~\tilde{p} is nonzero, so fM~(p~)0f_{\tilde{M}}(\tilde{p})\neq 0 by (23). a contradiction. Thus Case 2 cannot occur.

Now we turn to Case 3: dist(pn,Mn)0\operatorname{dist}(p_{n},\partial M_{n})\to 0. The boundary of Mn\partial M_{n} lies outside

(an,an)×(bn,bn)×𝐑,(-a_{n},a_{n})\times(-b_{n},b_{n})\times\mathbf{R},

so if ρn\rho_{n} is the minimum distance from one of the points (±x(Mn),0,z(Mn))(\pm x(M_{n}),0,z(M_{n})) to Mn\partial M_{n}, then

ρn\displaystyle\rho_{n} min{anx(Mn),bn}\displaystyle\geq\min\{a_{n}-x(M_{n}),b_{n}\}
min{2π,π/2}\displaystyle\geq\min\{2\pi,\pi/2\}
=π/2.\displaystyle=\pi/2.

Thus if 0<r<π/20<r<\pi/2, then (for all sufficiently large nn),

(28) 𝐁(pn,r) contains no critical points of x()|Mn.\text{$\mathbf{B}(p_{n},r)$ contains no critical points of $x(\cdot)|M_{n}$}.

Now let Mn=(Mnpn)/z(pn)M_{n}^{\prime}=(M_{n}-p_{n})/z(p_{n}). By passing to a subsequence, we can assume that MnM_{n}^{\prime} converges smoothly in {z>1}\{z>-1\} to a surface MM^{\prime} that is minimal with respect to the Euclidean metric. Note that MM^{\prime} is contained in {z1}\{z\geq-1\} and that the boundary of MM^{\prime} is (if nonempty) a set of lines in {z=1}\{z=-1\} parallel the xx-axis. Hence the function x()x(\cdot) (i.e., the function F𝐞1F_{\mathbf{e}_{1}}) is not constant on any component of MM.

Thus by Theorem 4.2, the critical point pp^{\prime} of x()|Mx(\cdot)|M^{\prime} is a limit of critical points of x()|Mnx(\cdot)|M_{n}^{\prime}. But that is impossible by (28). ∎

Theorem 12.6.

Let M(λ)M\in\mathscr{R}(\lambda). Then:

  1. (1)

    The tangent plane is never vertical at any point of MouterY¯\overline{M^{Y}_{\textnormal{outer}}}.

  2. (2)

    MouterYM^{Y}_{\textnormal{outer}} is connected.

  3. (3)

    ν𝐞3>0\nu\cdot\mathbf{e}_{3}>0 at all points of MouterY¯\overline{M^{Y}_{\textnormal{outer}}}.

  4. (4)

    At all points (x,y,z)MouterY¯(x,y,z)\in\overline{M^{Y}_{\textnormal{outer}}},

    0(B|y|)ηB(ν𝐞3),0\leq(B-|y|)\leq\eta B(\nu\cdot\mathbf{e}_{3}),

    where η=η(λ)\eta=\eta(\lambda) is given by Theorem 12.5.

  5. (5)

    MouterY{xx(M)+π}M^{Y}_{\textnormal{outer}}\cap\{x\geq x(M)+\pi\} is the graph of a function

    u:[x(M)+π,A(M)]×[B,B]u:[x(M)+\pi,A(M)]\times[-B,B]

    such that

    (B|y|)1+|Du|2ηB,(B-|y|)\sqrt{1+|Du|^{2}}\leq\eta B,

The statement of Theorem 12.6 is somewhat involved. However, for the remainder of the paper, we only need two consequences of the theorem: Assertion (5) of the theorem, and Theorem 12.8 below.

Proof.

By Theorem 11.2, 𝐞2ν\mathbf{e}_{2}\cdot\nu and y()y(\cdot) have opposite signs on waist(M){y0}\operatorname{waist}(M)\cap\{y\neq 0\}. Hence fM(p)<0f_{M}(p)<0 on waist(M){y0}\operatorname{waist}(M)\cap\{y\neq 0\}. By Theorem 12.5, fM(p)f_{M}(p) is never equal to 0 on waist(M)\operatorname{waist}(M). Thus fM<0f_{M}<0 at all points of waist(M)\operatorname{waist}(M), so waist(M)MinnerY\operatorname{waist}(M)\subset M^{Y}_{\textnormal{inner}}. This is Assertion (1).

Claim 2.

Suppose WW is a connected component of M{fM0}M\cap\{f_{M}\neq 0\} such that ν𝐞2=0\nu\cdot\mathbf{e}_{2}=0 at all points of W\partial W. Then there must be points in WW with ν𝐞3=0\nu\cdot\mathbf{e}_{3}=0.

Proof of claim.

Note that ν𝐞2\nu\cdot\mathbf{e}_{2} is a Jacobi field on W¯\overline{W} that vanishes at the boundary. Thus W¯\overline{W} is not strictly stable. Now ν𝐞3\nu\cdot\mathbf{e}_{3} is a Jacobi field that never vanishes on W\partial W (by Theorem 12.5). Thus if it never vanished on WW, then it would never vanish on W¯\overline{W}, and thus W¯\overline{W} would be strictly stable, a contradiction. Thus we have proved the claim. ∎

Recall (see (24)) that

(29) (M){fM>0}=E+E,(\partial M)\cap\{f_{M}>0\}=\pazocal{E}^{+}\cup\pazocal{E}^{-},

where

E±=(A,A)×{B±}×{0}.\pazocal{E}^{\pm}=(-A,A)\times\{B^{\pm}\}\times\{0\}.

Let WW be the connected component of MouterYM^{Y}_{\textnormal{outer}} that contains

(*) Mupper{x(x(M)+π,A)}.M^{\textnormal{upper}}\cap\{x\in(x(M)+\pi,A)\}.

(The set (*12) is in MouterYM^{Y}_{\textnormal{outer}} by Lemma 12.3.) Thus WW contains both of the edges E±\pazocal{E}^{\pm}, since it contains the portions of those edges with x>x(M)+πx>x(M)+\pi. If WW^{\prime} were another connected component of MouterYM^{Y}_{\textnormal{outer}}, it would not contain any of the edges E±\pazocal{E}^{\pm}, and thus ν𝐞2\nu\cdot\mathbf{e}_{2} would vanish everywhere on W\partial W^{\prime} (by (29)). Thus by Claim 2, ν𝐞3\nu\cdot\mathbf{e}_{3} would vanish at some points of WW^{\prime}. But by Assertion (1), ν𝐞3\nu\cdot\mathbf{e}_{3} does not vanish at any point of W¯\overline{W^{\prime}}. The contradiction proves that MouterYM^{Y}_{\textnormal{outer}} is connected, which is Assertion (2).

Now MouterYM^{Y}_{\textnormal{outer}} is connected, ν𝐞3\nu\cdot\mathbf{e}_{3} never vanishes on MouterY¯\overline{M^{Y}_{\textnormal{outer}}}, and ν𝐞3>0\nu\cdot\mathbf{e}_{3}>0 on E+MouterY\pazocal{E}^{+}\subset M^{Y}_{\textnormal{outer}}. Thus ν𝐞3>0\nu\cdot\mathbf{e}_{3}>0 at all points of MouterY¯\overline{M^{Y}_{\textnormal{outer}}}, which is Assertion (3).

By Spruck-Xiao [spruck-xiao]*Lemma 5.7 (see also [graphs]*Theorem 2.6), the maximum of (ν𝐞3)1(B|y|)(\nu\cdot\mathbf{e}_{3})^{-1}(B-|y|) on MouterY¯\overline{M^{Y}_{\textnormal{outer}}} occurs at a point in the boundary, and thus a point where fM=0f_{M}=0. Thus Assertion (4) follows from Theorem 12.5.

Assertion (5) follows immediately from Assertion (4) and Theorem 11.4. ∎

Definition 12.7.

Let (b,B)\mathscr{L}(b,B) be the set of all limits of sequences Mn=MnqnM_{n}^{\prime}=M_{n}-q_{n} where MnM_{n}\in\mathscr{R} and qn=(xn,0,zn){y=0}q_{n}=(x_{n},0,z_{n})\in\{y=0\}, and where

a(Mn)x(Mn),\displaystyle a(M_{n})-x(M_{n})\to\infty,
bn:=bn(M) converges to b,\displaystyle\text{$b_{n}:=b_{n}(M)$ converges to $b$},
Bn:=Bn(M) converges to B,and\displaystyle\text{$B_{n}:=B_{n}(M)$ converges to $B$},\,\text{and}
dist(qn,Mn).\displaystyle\operatorname{dist}(q_{n},\partial M_{n})\to\infty.

Note that 𝒜(b,B,x^)(b,B)\mathscr{A}(b,B,\hat{x})\subset\mathscr{L}(b,B) for every x^\hat{x}, and that any limit of surfaces in (b,B)\mathscr{L}(b,B) is also in (b,B)\mathscr{L}(b,B).

Theorem 12.8.

Let λ>B\lambda>B. If M(b,B)M\in\mathscr{L}(b,B) and if pMouterp\in M^{\textnormal{outer}}, then

(30) (B|y(p)|)η(λ)B(ν(M,p)𝐞3).(B-|y(p)|)\leq\eta(\lambda)B(\nu(M,p)\cdot\mathbf{e}_{3}).

Now suppose M(b,B)M\in\mathscr{L}(b,B) is a non-empty union of planes. Then MouterM^{\textnormal{outer}} contains the planes {y=±B}\{y=\pm B\}, possibly with multiplicity, and the planes of MinnerM^{\textnormal{inner}} are contained in the slab {|y|b}\{|y|\leq b\}. If B>bB>b, then the plane {y=B}\{y=B\} occurs with multiplicity 11.

Proof.

Let λ>B\lambda>B. Note that Mn(λ)M_{n}\in\mathscr{R}(\lambda) for all sufficiently large nn. We may assume that Mn(λ)M_{n}\in\mathscr{R}(\lambda) for all nn. Thus for p(Mn)outerp\in(M_{n}^{\prime})^{\textnormal{outer}}, we have the estimate (Theorem 12.5)

(Bn|y(p)|)ηBn(ν(Mn,p)𝐞3).(B_{n}-|y(p)|)\leq\eta B_{n}(\nu(M_{n}^{\prime},p)\cdot\mathbf{e}_{3}).

Passing to the limit gives (30).

Now suppose that MM is a nonempty union of planes. For large nn, the yy-axis, YY, intersects MnM_{n}^{\prime} transversely in k>0k>0 points, where kk is the number of planes. By Proposition 12.2, the point in YMnY\cap M_{n}^{\prime} for which yy is greatest is in (Mn)outer(M_{n}^{\prime})^{\textnormal{outer}}. Thus (passing to the limit) MouterM^{\textnormal{outer}} is nonempty.

Since MM consists of vertical planes, ν𝐞30\nu\cdot\mathbf{e}_{3}\equiv 0 on MM, and thus |y|B|y|\equiv B on MouterM^{\textnormal{outer}} by (30) Hence MouterM^{\textnormal{outer}} consists of the planes {y=±B}\{y=\pm B\}.

If B=bB=b, then trivially MinnerM^{\textnormal{inner}} is contained in {|y|b}\{|y|\leq b\}. Thus suppose that B>bB>b. Then, by Theorem 10.7 (applied to the MnM_{n}^{\prime}), the planes {y=B}\{y=B\} and {y=B}\{y=-B\} each occur with multiplicity 11 in MM, and the remainder of MM lies in the slab {|y|b}\{|y|\leq b\}. ∎

Corollary 12.9.

Suppose that M(b,B)M\in\mathscr{L}(b,B), that Σ\Sigma is a Δ\Delta-wing or grim reaper surface definied on the strip 𝐑×(β,β)\mathbf{R}\times(-\beta,\beta), and that Σ\Sigma is contained in MinnerM^{\textnormal{inner}}. Then βb\beta\leq b.

Proof.

Let MM^{\prime} be a subsequential limit of M+(0,0,z)M+(0,0,z) as zz\to\infty. By Theorem 5.4, such a limit exists and is a union of planes. Note that Σ+(0,0,z)\Sigma+(0,0,z) converges to the planes {y=±β}\{y=\pm\beta\}, and thus those planes are in the inner portion of MM^{\prime}. Thus βb\beta\leq b by Theorem 12.8. ∎

We conclude this section with the Lemma that was used in the proof of Theorem 12.5.

Lemma 12.10.

Suppose that MiM_{i} are surfaces in 𝒞\mathscr{C} or 𝒜\mathscr{A} that lie in a slab {|y|Λ}\{|y|\leq\Lambda\}. Suppose also that piMip_{i}\in M_{i}, that dist(pi,Z)0\operatorname{dist}(p_{i},Z)\to 0 and that ν(Mi,pi)𝐯\nu(M_{i},p_{i})\to\mathbf{v}, where 𝐯\mathbf{v} is a horizontal vector not equal to 𝐞2\mathbf{e}_{2} or 𝐞2-\mathbf{e}_{2}. Let qiq_{i} be the point in ZZ closest to pip_{i}. Then Mi:=(Miqi)/dist(pi,Z)M_{i}^{\prime}:=(M_{i}-q_{i})/\operatorname{dist}(p_{i},Z) converges smoothly and with multiplicity 11 to the catenoid whose waist is the unit circle in the plane {z=0}\{z=0\}.

(The lemma is also true when 𝐯=±𝐞2\mathbf{v}=\pm\mathbf{e}_{2}, but a different proof is required. We only need the case when 𝐯=±𝐞1\mathbf{v}=\pm\mathbf{e}_{1}.)

Proof.

By the compactness theorem [white18]*Theorem 1.1 for minimal surfaces, we can assume (by passing to a subsequence) that MiqiM_{i}-q_{i} converges to a limit MM, where MMM\setminus\partial M is smooth and embedded (possibly with multiplicity), and where the convergence is smooth away from the boundary except at a locally finite subset SS of MMM\setminus\partial M. (Indeed, by Remark 7.7, SS contains at most one point.)

The symmetries imply that Tan(M,0)\operatorname{Tan}(M,0) is one of the coordinate planes. We claim that Tan(M,0)\operatorname{Tan}(M,0) cannot be {x=0}\{x=0\}. For suppose it were. Since MM is embedded and invariant under (x,y,z)(x,y,z)(x,y,z)\mapsto(-x,y,z), if follows that MM and {x=0}\{x=0\} coincide in a neighborhood of 0. But then by unique continuation, all of {x=0}\{x=0\} would lie in MM, which is impossible since MM is contained in the slab {|y|Λ}\{|y|\leq\Lambda\}.

Thus Tan(M,0)\operatorname{Tan}(M,0) is either {y=0}\{y=0\} or {z=0}\{z=0\}. Consequently, since 𝐯{±𝐞2,±𝐞3}\mathbf{v}\notin\{\pm\mathbf{e}_{2},\pm\mathbf{e}_{3}\}, 𝐯\mathbf{v} is not perpendicular to Tan(M,0)\operatorname{Tan}(M,0).

By passing to a subsequence, we can assume that MiM_{i}^{\prime} converges to a limit MM^{\prime} and that

pi=(piqi)/dist(pi,Z)=(piqi)/|piqi|p_{i}^{\prime}=(p_{i}-q_{i})/\operatorname{dist}(p_{i},Z)=(p_{i}-q_{i})/|p_{i}-q_{i}|

converges to a point pp^{\prime}. The convergence is smooth at pp^{\prime} (and at p-p^{\prime}), so

(31) ν(M,p)=𝐯,ν(M,p)=𝐯.\nu(M^{\prime},p^{\prime})=\mathbf{v},\quad\nu(M^{\prime},-p^{\prime})=-\mathbf{v}.

By [white18]*Theorems 2.2 and 2.3, MM^{\prime} is one of the following:

  1. (1)

    A multiplicity 11 plane.

  2. (2)

    Two or more planes (counting multiplicity) parallel to Tan(M,0)\operatorname{Tan}(M,0).

  3. (3)

    A complete, embedded, non-flat minimal surface of finite total curvature whose ends are parallel to Tan(M,0)\operatorname{Tan}(M,0). In this case, the convergence is smooth and with multiplicity 11.

If MM^{\prime} consists of planes, then (by (31)) there are at least two planes, counting multiplicity, and thus those planes are parallel to Tan(M,0)\operatorname{Tan}(M,0). In that case, by (31), 𝐯\mathbf{v} would be perpendicular to Tan(M,0)\operatorname{Tan}(M,0). But we have already seen that 𝐯\mathbf{v} is not perpendicular to Tan(M,0)\operatorname{Tan}(M,0). Thus MM is a complete, nonflat, embedded, minimal surface of finite total curvature. Since MM has genus 0, it is a catenoid with ends parallel to Tan(M,0)\operatorname{Tan}(M,0), and therefore parallel to {z=0}\{z=0\} or to {y=0}\{y=0\}. If the ends were parallel to {y=0}\{y=0\}, then (by symmetry) MM^{\prime} would intersect ZZ orthogonally in a pair of points. But that is impossible since the MiM_{i}^{\prime} are disjoint from ZZ.

Hence the ends of MM^{\prime} are parallel to {z=0}\{z=0\} (and therefore Tan(M,0)\operatorname{Tan}(M,0) is horizontal.) ∎

Remark 12.11.

In the proof of Lemma 12.10, we showed that the MiM_{i} converged (after passing to a subsequence) to a smooth limit MM with Tan(M,0)\operatorname{Tan}(M,0) parallel to the ends of MM^{\prime}, i.e., with Tan(M,0)\operatorname{Tan}(M,0) horizontal. Of course, the convergence of MiM_{i} to MM is not smooth at 0, and thus the multiplicity of the component of MM containing 0 is 2\geq 2. In fact, since every vertical line in the plane {y=0}\{y=0\} intersects MiM_{i} at most twice (by Theorems 8.1 and 8.2), that multiplicity is exactly 22.

13. Behavior as zz\to-\infty

Suppose M𝒜(b,B,x^)M\in\mathscr{A}(b,B,\hat{x}). Recall from §10 that Ωin(M)\Omega^{\textnormal{in}}(M) consists of the region of the (x,z)(x,z)-plane bounded above by the curve z=ϕlower(x)z=\phi^{\textnormal{lower}}(x), and on the left by the vertical line x=c^:=x(M)+πx=\hat{c}:=x(M)+\pi. In this section we show (Theorem 13.4) that if (x,z)Ωin(M)(x,z)\in\Omega^{\textnormal{in}}(M) and if (x,z)(x,z) is far from the top edge of Ωin(M)\Omega^{\textnormal{in}}(M), then yinner(x,z)y^{\textnormal{inner}}(x,z) and youter(x,z)y^{\textnormal{outer}}(x,z) are very near bb and BB, respectively. It follows that b(M)=bb(M)=b and B(M)=BB(M)=B.

Theorem 13.4 is a direct consequence of the analogous result (Lemma 13.3) for MM\in\mathscr{R}.

Definition 13.1.

For bπ/2b\geq\pi/2, let

wb:𝐑×(b,b)𝐑w_{b}:\mathbf{R}\times(-b,b)\to\mathbf{R}

be the grim reaper surface with wb(0,0)=0w_{b}(0,0)=0 and

wbxs(b).\frac{\partial w_{b}}{\partial x}\equiv-s(b).

Let

yb:{(x,z):zs(b)x}[0,b)y^{b}:\{(x,z):z\leq-s(b)x\}\longrightarrow[0,b)

be the function such that (x,y,z)graph(wb)(x,y,z)\in\operatorname{graph}(w_{b}) if and only if y=±yb(x,z)y=\pm y^{b}(x,z).

Lemma 13.2.

Suppose that MnM_{n}\in\mathscr{R}, that bn:=b(Mn)b_{n}:=b(M_{n}) converges to bb, and that an:=a(Mn)a_{n}:=a(M_{n}) tends to infinity. Let pn=(x(Mn),0,z(Mn))p_{n}=(x(M_{n}),0,z(M_{n})) be the point where ν(Mn,pn)=𝐞1\nu(M_{n},p_{n})=-\mathbf{e}_{1}. Let xn[x(Mn),An]x_{n}\in[x(M_{n}),A_{n}] and qn:=(xn,0,ϕnlower(xn))q_{n}:=(x_{n},0,\phi_{n}^{\textnormal{lower}}(x_{n})) be such that

dist(qn,{pn}Mn).\displaystyle\operatorname{dist}(q_{n},\{p_{n}\}\cup\partial M_{n})\to\infty.

Then

Σn:=Mnlowerqn\Sigma_{n}:=M^{\textnormal{lower}}_{n}-q_{n}

converges smoothly to the grim reaper surface graph(wb)\operatorname{graph}(w_{b}).

Proof.

Let Mn=MnqnM_{n}^{\prime}=M_{n}-q_{n}. Theorem 7.8 gives smooth subsequential convergence of MnM_{n}^{\prime} to a limit MM^{\prime}. (Since xnx(Mn)x_{n}\geq x(M_{n}), the distance from qnq_{n} to (x(Mn),0,z(Mn))(-x(M_{n}),0,z(M_{n})) is greater than the distance from qnq_{n} to pnp_{n}.) Let Σ\Sigma be the component of MM^{\prime} containing 0. By Theorem 7.8,

xnx(Mn)x_{n}-x(M_{n})\to\infty

and Σ\Sigma is a Δ\Delta-wing or grim reaper surface, and thus the graph of a function

w:𝐑×(β,β)𝐑w:\mathbf{R}\times(-\beta,\beta)\to\mathbf{R}

with w(0,0)=0w(0,0)=0, for some βπ/2\beta\geq\pi/2. Note that Σ=graph(w)\Sigma=\operatorname{graph}(w) is the limit of MnlowerqnM^{\textnormal{lower}}_{n}-q_{n}.

Let x𝐑x\in\mathbf{R}. By Theorem 9.1,

(32) (ϕnlower)(xn+x)xfbn(xn+x,0).(\phi_{n}^{\textnormal{lower}})^{\prime}(x_{n}+x)\leq\frac{\partial}{\partial x}f_{b_{n}}(x_{n}+x,0).

As nn\to\infty, the left side of (32) converges to xw(t,0)\frac{\partial}{\partial x}w(t,0). Since xnx(Mn)x_{n}-x(M_{n})\to\infty, xnx_{n}\to\infty. Thus the right side of (32) converges to s(b)-s(b). Hence

xw(x,0)s(b)\frac{\partial}{\partial x}w(x,0)\leq-s(b)

for all xx. Consequently Σ=graph(wβ)\Sigma=\operatorname{graph}(w_{\beta}) for some βb\beta\geq b. On the other hand, βb\beta\leq b by Corollary 12.9. ∎

Recall that for MM\in\mathscr{R},

Ωin(M):={(x,z)[c^,a)×𝐑:0zϕlower(x)},\Omega^{\textnormal{in}}(M):=\{(x,z)\in[\hat{c},a)\times\mathbf{R}:0\leq z\leq\phi^{\textnormal{lower}}(x)\},

where c^=x(M)+π\hat{c}=x(M)+\pi and a=a(M)a=a(M). The following lemma says that for MM\in\mathscr{R}, if (x,z)Ωin(x,z)\in\Omega^{\textnormal{in}} is far from the upper and lower edges of Ωin\Omega^{\textnormal{in}}, then yinner(y,z)y^{\textnormal{inner}}(y,z) is close to b(M)b(M) and youter(y,z)y^{\textnormal{outer}}(y,z) is close to B(M)B(M). The analogous assertion for M𝒜(b,B,x^)M\in\mathscr{A}(b,B,\hat{x}) (which is the main result of this section) follows readily.

Lemma 13.3.

For every pair of positive numbers ϵ\epsilon and λ\lambda, there is an RR with the following property. Suppose that M(λ)M\in\mathscr{R}(\lambda) with b=b(M)b=b(M). Then

yinner(x,z)[bϵ,b+ϵ]y^{\textnormal{inner}}(x,z)\in[b-\epsilon,b+\epsilon] and youter(x,z)[Bϵ,B]y^{\textnormal{outer}}(x,z)\in[B-\epsilon,B]

for all (x,z)Ωin(M)(x,z)\in\Omega^{\textnormal{in}}(M) such that

z[R,ϕlower(x)R],z\in[R,\phi^{\textnormal{lower}}(x)-R],
Proof.

Suppose it is false. Then there exist Mn(λ)M_{n}\in\mathscr{R}(\lambda) and (xn,zn)Ωin(Mn)(x_{n},z_{n})\in\Omega^{\textnormal{in}}(M_{n}) such that

(33) nznϕnlower(xn)nn\leq z_{n}\leq\phi_{n}^{\textnormal{lower}}(x_{n})-n

and such that

(34) yinner(xn,zn)[bnϵ,bn+ϵ] or youter(xn,zn)[Bnϵ,Bn].\text{$y^{\textnormal{inner}}(x_{n},z_{n})\notin[b_{n}-\epsilon,b_{n}+\epsilon]$ or $y^{\textnormal{outer}}(x_{n},z_{n})\notin[B_{n}-\epsilon,B_{n}]$}.

By passing to subsequence, we can assume that bnb_{n} and BnB_{n} converge:

(35) bnb,\displaystyle b_{n}\to b,
BnB.\displaystyle B_{n}\to B.

We will prove Lemma 13.3 by showing that (33) and (35) imply (contrary to (34)) that

limnyninner(xn,zn)\displaystyle\lim_{n}y^{\textnormal{inner}}_{n}(x_{n},z_{n}) =b,and\displaystyle=b,\,\text{and}
limnynouter(xn,zn)\displaystyle\lim_{n}y^{\textnormal{outer}}_{n}(x_{n},z_{n}) =B.\displaystyle=B.

By Theorem 7.8, we can assume (passing to a subsequence) that Mn:=Mn(xn,0,zn)M_{n}^{\prime}:=M_{n}-(x_{n},0,z_{n}) converges smoothly to a limit MM^{\prime}, each component of which is one of the following: a plane parallel to {y=0}\{y=0\}, or a Δ\Delta-wing, or a grim reaper surface.

Note that if x>0x>0 and if Y+(x,0,z)Y+(x,0,z) intersects MM^{\prime} transversely, then it does so in exactly four points (counting multiplicity).

Thus MM^{\prime} is the union of 44 planes (counting multiplicity). By Theorem 12.8, the planes are {y=±B}\{y=\pm B\} and {y=b}\{y=b^{*}\} for some b[0,b]b^{*}\in[0,b].

Thus limnynouter(xn,zn)=B\lim_{n}y^{\textnormal{outer}}_{n}(x_{n},z_{n})=B and limyninner(xn,zn)=b\lim y^{\textnormal{inner}}_{n}(x_{n},z_{n})=b^{*}, so it remains only to show that bbb^{*}\geq b.

Fix (for the moment) a k(0,)k\in(0,\infty). For n>kn>k, using (33) and the fact that ϕnlower\phi_{n}^{\textnormal{lower}} is decreasing (Theorem 9.1), there is a xn>xnx_{n}^{\prime}>x_{n} such that

ϕnlower(xn)=zn+k.\phi_{n}^{\textnormal{lower}}(x_{n}^{\prime})=z_{n}+k.

Let

qn=(xn,0,ϕnlower(xn))=(xn,0,zn+k).q_{n}=(x_{n}^{\prime},0,\phi_{n}^{\textnormal{lower}}(x_{n}^{\prime}))=(x_{n}^{\prime},0,z_{n}+k).

Now

dist(qn,Mn)z(qn)=zn+kzn.\operatorname{dist}(q_{n},\partial M_{n})\geq z(q_{n})=z_{n}+k\geq z_{n}\to\infty.

Also,

|pnqn|\displaystyle|p_{n}-q_{n}| z(pn)z(qn)\displaystyle\geq z(p_{n})-z(q_{n})
=ϕnlower(x(Mn))(zn+k)\displaystyle=\phi_{n}^{\textnormal{lower}}(x(M_{n}))-(z_{n}+k)
ϕnlower(xn)znk\displaystyle\geq\phi_{n}^{\textnormal{lower}}(x_{n})-z_{n}-k
nk.\displaystyle\geq n-k\to\infty.

By Lemma 13.2, MnlowerqnM^{\textnormal{lower}}_{n}-q_{n} converges smoothly to the grim reaper surface graph(wb)\operatorname{graph}(w_{b}). Thus for every t>0t>0,

limnyninner(xn,0,ϕnlower(xn)t)=yb(0,t).\lim_{n}y^{\textnormal{inner}}_{n}(x_{n}^{\prime},0,\phi_{n}^{\textnormal{lower}}(x_{n}^{\prime})-t)=y_{b}(0,-t).

In particular, for t=kt=k, ϕnlower(xn)k=zn\phi_{n}^{\textnormal{lower}}(x_{n}^{\prime})-k=z_{n}, so

limnyninner(xn,zn)=yb(0,t).\lim_{n}y^{\textnormal{inner}}_{n}(x_{n}^{\prime},z_{n})=y^{b}(0,-t).

Since yninner(xn,zn)yninner(xn,zn)y^{\textnormal{inner}}_{n}(x_{n},z_{n})\geq y^{\textnormal{inner}}_{n}(x_{n}^{\prime},z_{n}) (by Theorem 10.9),

b=limnyninner(xn,zn)yb(0,t).b^{*}=\lim_{n}y^{\textnormal{inner}}_{n}(x_{n},z_{n})\geq y^{b}(0,-t).

This holds for every t(0,)t\in(0,\infty). Thus letting tt\to\infty gives

bb.b^{*}\geq b.

Theorem 13.4.

For every pair of positive numbers ϵ\epsilon and λ\lambda, there is an RR with the following property. Suppose that M𝒜(b,B,x^)M\in\mathscr{A}(b,B,\hat{x}). If B<λB<\lambda, then

yinner(x,z)[bϵ,b+ϵ]y^{\textnormal{inner}}(x,z)\in[b-\epsilon,b+\epsilon] and youter(x,z)[Bϵ,B]y^{\textnormal{outer}}(x,z)\in[B-\epsilon,B]

for all (x,z)Ωin(M)(x,z)\in\Omega^{\textnormal{in}}(M) such that

z<ϕlower(x)Rz<\phi^{\textnormal{lower}}(x)-R
Proof.

Let MnM_{n}\in\mathscr{R} be such that: bn:=b(Mn)b_{n}:=b(M_{n}), Bn:=B(Mn)B_{n}:=B(M_{n}), x(Mn)x(M_{n}), and a(Mn)a(M_{n}) converge to bb, BB, x^\hat{x}, and \infty, and such that M~n=Mn(0,0,z(Mn))\tilde{M}_{n}=M_{n}-(0,0,z(M_{n})) converges smoothly to MM. We may assume that Bn<λB_{n}<\lambda and that an>x(Mn)+2πa_{n}>x(M_{n})+2\pi for all nn.

By the smooth convergence,

  1. (i)

    ϕnlower()z(Mn)\phi_{n}^{\textnormal{lower}}(\cdot)-z(M_{n}) converges uniformly on compact subsets of (x(M),)(x(M),\infty) to ϕlower()\phi^{\textnormal{lower}}(\cdot), and

  2. (ii)

    yninner(x,z+z(Mn))y^{\textnormal{inner}}_{n}(x,z+z(M_{n})) and ynouter(x,z+z(Mn))y^{\textnormal{outer}}_{n}(x,z+z(M_{n})) converge uniformly on compact subsets of {x>x(M),z<ϕlower(x)}\{x>x(M),\,z<\phi^{\textnormal{lower}}(x)\} to yinner(x,z)y^{\textnormal{inner}}(x,z) and youter(x,z)y^{\textnormal{outer}}(x,z).

Let R=R(ϵ,λ)R=R(\epsilon,\lambda) be as in Lemma 13.3. Suppose first that x>x(M)+πx>x(M)+\pi and that z<ϕlower(x)Rz<\phi^{\textnormal{lower}}(x)-R.

Then for all sufficiently large nn,

z<ϕnlower(x)z(Mn)Rz<\phi_{n}^{\textnormal{lower}}(x)-z(M_{n})-R

or, equivalently,

z+z(Mn)<ϕnlowerR.z+z(M_{n})<\phi_{n}^{\textnormal{lower}}-R.

Also, since z(Mn)z(M_{n})\to\infty,

R<z+z(Mn)R<z+z(M_{n})

for all sufficiently large nn. Thus, for such nn,

yninner(x,z+z(Mn))[bϵ,b+ϵ]y^{\textnormal{inner}}_{n}(x,z+z(M_{n}))\in[b-\epsilon,b+\epsilon] and ynouter(x,z+z(Mn))[Bϵ,B]y^{\textnormal{outer}}_{n}(x,z+z(M_{n}))\in[B-\epsilon,B]

Then, letting nn\to\infty, we have (by (ii))

(36) yinner(x,z)[bϵ,b+ϵ] and youter(x,z)[Bϵ,B].\text{$y^{\textnormal{inner}}(x,z)\in[b-\epsilon,b+\epsilon]$ and $y^{\textnormal{outer}}(x,z)\in[B-\epsilon,B]$}.

We have shown that (36) holds for all (x,z)(x,z) with x>x(M)x>x(M) and z<ϕlower(x)Rz<\phi^{\textnormal{lower}}(x)-R. By continuity, (36) also holds for (x,z)(x,z) with xx(M)x\geq x(M) and zϕlower(x)Rz\leq\phi^{\textnormal{lower}}(x)-R. ∎

Theorem 13.5.

Suppose M𝒜(b,B,x^)M\in\mathscr{A}(b,B,\hat{x}). Then b(M)=bb(M)=b and B(M)=BB(M)=B.

Proof.

Let xx(M)+πx\geq x(M)+\pi. Then

limzyinner(x,z)=b,and\displaystyle\lim_{z\to-\infty}y^{\textnormal{inner}}(x,z)=b,\,\text{and}
limzyouter(x,z)=B.\displaystyle\lim_{z\to-\infty}y^{\textnormal{outer}}(x,z)=B.

by Theorem 13.4. Hence b(M)=bb(M)=b and B(M)=BB(M)=B. (See Theorem 10.6 if this is not clear.) ∎

14. Behavior as x±x\to\pm\infty

In this section we study the asymptotic behavior of the wings, MlowerM^{\textnormal{lower}} and MupperM^{\textnormal{upper}}, of M𝒜(b,B,x^)M\in\mathscr{A}(b,B,\hat{x}), as x.x\to\infty. In particular, we prove that if we take limit of Mlower(x,0,ϕlower(x))M^{\textnormal{lower}}-(x,0,\phi^{\textnormal{lower}}(x)), as xx\to\infty, then this limit is the grim reaper surface of slope s(b)-s(b). On the other hand, Mupper(x,0,ϕupper(x))M^{\textnormal{upper}}-(x,0,\phi^{\textnormal{upper}}(x)) converges, as xx\to\infty, to a grim reaper surface of slope s=±s(B)s=\pm s(B). If B>bB>b, then s=s(B).s=s(B).

Theorem 14.1.

Suppose that M𝒜(b,B,x^)M\in\mathscr{A}(b,B,\hat{x}). Let λ>B\lambda>B. Then

  1. (1)

    limx(ϕlower)(x)=s(b)\displaystyle\lim_{x\to\infty}(\phi^{\textnormal{lower}})^{\prime}(x)=-s(b), and Mlower(x,0,ϕlower(x))M^{\textnormal{lower}}-(x,0,\phi^{\textnormal{lower}}(x)) converges to the grim reaper surface graph(wb)\operatorname{graph}(w_{b}). (See Definition 13.1.)

  2. (2)

    The surface Mupper{xx(M)+π}M^{\textnormal{upper}}\cap\{x\geq x(M)+\pi\} is the graph of a function

    u:[x(M)+π,)×(B,B)𝐑u:[x(M)+\pi,\infty)\times(-B,B)\to\mathbf{R}

    that satisfies the gradient bound

    (B|y|)1+|Du|2ηB,(B-|y|)\sqrt{1+|Du|^{2}}\leq\eta B,

    for a constant η=η(λ)\eta=\eta(\lambda).

  3. (3)

    The limit L:=limx(ϕupper)(x)\displaystyle L:=\lim_{x\to\infty}(\phi^{\textnormal{upper}})^{\prime}(x) exists, and L=±s(B)L=\pm s(B). Moreover, Mupper(x,0,ϕupper(x))M^{\textnormal{upper}}-(x,0,\phi^{\textnormal{upper}}(x)) converges smoothly as xx\to\infty to the grim reaper surface graph(w)\operatorname{graph}(w) such that

    w:𝐑×(B,B)𝐑,\displaystyle w:\mathbf{R}\times(-B,B)\to\mathbf{R},
    w(0,0)=0,\displaystyle w(0,0)=0,
    wxL.\displaystyle\frac{\partial w}{\partial x}\equiv L.

    In particular,

    limxxu(x,y)=L\lim_{x\to\infty}\frac{\partial}{\partial x}u(x,y)=L

    for each y(B,B)y\in(-B,B).

  4. (4)

    If B>bB>b, then L=s(B)L=s(B).

Proof.

Let βπ/2\beta\geq\pi/2 be the number such that

(37) limx(ϕlower)(x)=s(β).\lim_{x\to\infty}(\phi^{\textnormal{lower}})^{\prime}(x)=-s(\beta).

(The limit exists by Theorem 8.1.) By Corollary 9.2, the limit in (37) is at most s(b)-s(b), and therefore

(38) βb.\beta\geq b.

Let xnx_{n}\to\infty and

Mn:=M(xn,0,ϕlower(xn))M_{n}^{\prime}:=M-(x_{n},0,\phi^{\textnormal{lower}}(x_{n}))

By passing to a subsequence, we can assume that MnM_{n}^{\prime} converges smoothly to a limit MM^{\prime}. Let Σ\Sigma be the component of MM^{\prime} containing 0. By (37),

Σ{y=0}={(x,0,z):z=s(β)x},\Sigma\cap\{y=0\}=\{(x,0,z):z=-s(\beta)x\},

Thus Σ\Sigma is the grim reaper surface wβw_{\beta}. (If this is not clear, see Theorem 5.4.) Note that Σ\Sigma is in (M)inner(M^{\prime})^{\textnormal{inner}}. Thus

βb.\beta\leq b.

by Corollary 12.9. This (together with (38)) completes the proof of Assertion (1).

Assertion (2) follows immediately from the corresponding assertion in Theorem 12.6 for MM\in\mathscr{R}.

By Theorem 8.1, the limit L=limx(ϕupper)(x)\displaystyle L=\lim_{x\to\infty}(\phi^{\textnormal{upper}})^{\prime}(x) exists and is finite. Suppose xnx_{n}\to\infty. By the gradient bound in Assertion (2) and Arzela-Ascoli, the function

u(x+xn,y)u(xn,0)u(x+x_{n},y)-u(x_{n},0)

converges smoothly, perhaps after passing to a subsequence, to a function

w:𝐑×(B,B)𝐑.w:\mathbf{R}\times(-B,B)\to\mathbf{R}.

Since w(x,0)=Lxw(x,0)=Lx, we see that ww is the grim reaper surface with

xwL,\frac{\partial}{\partial x}w\equiv L,

and thus that L=±s(B)L=\pm s(B). Since the limit is independent of the choice of subsequence, we get convergence and not just subsequential convergence.

To prove Assertion (4), suppose that b<Bb<B. Then s(b)<s(B)s(b)<s(B), so

(39) s(B)<s(b)<s(B).-s(B)<-s(b)<s(B).

Now ϕupper(x)>ϕlower(x)\phi^{\textnormal{upper}}(x)>\phi^{\textnormal{lower}}(x) for all x>x(M)x>x(M), so

L:=limx(ϕupper)(x)limx(ϕlower)(x)=s(b).L:=\lim_{x\to\infty}(\phi^{\textnormal{upper}})^{\prime}(x)\geq\lim_{x\to\infty}(\phi^{\textnormal{lower}})^{\prime}(x)=-s(b).

Since L=±s(B)L=\pm s(B), we see from (39) that L=s(B)L=s(B). ∎

We will show later (Theorem 22.5) that there are examples for which b=Bb=B and other examples for which b<Bb<B. The following theorem gives a condition guaranteeing that b=Bb=B:

Theorem 14.2.

Suppose that M𝒜M\in\mathscr{A} and that B:=B(M)>y(M)+πB:=B(M)>y(M)+\pi. Then b(M)=B(M)b(M)=B(M) and

limx(ϕupper)(x)=s(B).\lim_{x\to\infty}(\phi^{\textnormal{upper}})^{\prime}(x)=-s(B).
Proof.

By Theorem 11.4,

𝐑×[y(M)+π,B)domain(uM)\mathbf{R}\times[y(M)+\pi,B)\subset\operatorname{domain}(u_{M})

and

xuM(x,y)0on [0,)×[y(M+π,B).\frac{\partial}{\partial x}u_{M}(x,y)\leq 0\quad\text{on $[0,\infty)\times[y(M+\pi,B)$.}

In particular, for each y[y(M)+π,B)y\in[y(M)+\pi,B),

lim supxxuM(x,y)0.\limsup_{x\to\infty}\frac{\partial}{\partial x}u_{M}(x,y)\leq 0.

In particular, since uM(x,0)=ϕupper(x)u_{M}(x,0)=\phi^{\textnormal{upper}}(x) for xx(M)x\geq x(M) and since limx(ϕupper)(x)=±s(B)\lim_{x\to\infty}(\phi^{\textnormal{upper}})^{\prime}(x)=\pm s(B), we see that

limx(ϕupper)(x)=s(B).\lim_{x\to\infty}(\phi^{\textnormal{upper}})^{\prime}(x)=-s(B).

On the other hand, if b<Bb<B, then

limxxu(x,y)=s(B)>0\lim_{x\to\infty}\frac{\partial}{\partial x}u(x,y)=s(B)>0

by Theorem 14.1. Thus b=Bb=B. ∎

15. Large necks and Prongs

In this section, we analyze the behavior of Mn𝒜M_{n}\in\mathscr{A} as the necksize x(Mn)x(M_{n}) tends to \infty. We begin by examining the behavior of Mn𝒜M_{n}\in\mathscr{A} near the point (x(Mn),0,0)(x(M_{n}),0,0). Specifically, we analyze the limit (which we call a prong) of Mn(x(Mn),0,0)M_{n}-(x(M_{n}),0,0) as x(Mn)x(M_{n})\to\infty. Then, in Theorem 15.6, we describe the behavior of MnM_{n} at bounded distances from the plane {x=0}\{x=0\}: in particular, we show that suitable vertical translates of the MnM_{n} converge to a pair of untilted grim reaper surfaces over strips 𝐑×(b,b+π)\mathbf{R}\times(b,b+\pi) and 𝐑×((b+π),b)\mathbf{R}\times(-(b+\pi),-b).

Theorem 15.1.

Suppose that Mn𝒜M_{n}\in\mathscr{A} are annuloids with bn:=b(Mn)b<b_{n}:=b(M_{n})\to b<\infty and with x(Mn)x(M_{n})\to\infty. Then, perhaps after passing to a subsequence, Bn:=B(Mn)B_{n}:=B(M_{n}) converges to a limit BB, and the surfaces

Mn(x(Mn),0,0)M_{n}-(x(M_{n}),0,0)

converge smoothly to a limit MM.

Theorem 15.1 follows immediately from the curvature and area bounds in Theorem 3.4.

Definition 15.2.

We define a prong to be any surface MM obtained as in Theorem 15.1.

Theorem 15.3 gives some basic properties of prongs, including the behavior of a prong as zz tends to \infty or -\infty, and the fact that B=b+πB=b+\pi in Theorem 15.1. Theorem 15.5 describes the behavior as xx tends to \infty or to -\infty. Another interesting property of a prong, namely that it is a sideways graph x=x(y,z)x=x(y,z) over a region in the yzyz-plane, will be proved in Section 23; see Theorem 23.6.

Theorem 15.3.

Suppose that MM is a prong, and that MnM_{n}, bb, and BB are as in Theorem 15.1. Let λ>B\lambda>B, then

  1. (1)

    M{y=0}M\cap\{y=0\} is the union of two graphs

    {(x,0,ϕupper(x)):x0}and{(x,0,ϕlower(x)):x0},\{(x,0,\phi^{\textnormal{upper}}(x)):x\geq 0\}\quad\text{and}\quad\{(x,0,\phi^{\textnormal{lower}}(x)):x\geq 0\},

    where ϕupper,ϕlower:[0,)𝐑\phi^{\textnormal{upper}},\phi^{\textnormal{lower}}:[0,\infty)\to\mathbf{R} and

    ϕlower(0)\displaystyle\phi^{\textnormal{lower}}(0) =ϕupper(0)=0,and\displaystyle=\phi^{\textnormal{upper}}(0)=0,\,\text{and}
    ϕlower(x)\displaystyle\phi^{\textnormal{lower}}(x) <ϕupper(x)for x>0.\displaystyle<\phi^{\textnormal{upper}}(x)\quad\text{for $x>0$}.

    We let MupperM^{\textnormal{upper}} and MlowerM^{\textnormal{lower}} be the components of M{x>0}M\cap\{x>0\} containing {(x,0,ϕupper(x)):x>0}\{(x,0,\phi^{\textnormal{upper}}(x)):x>0\} and {(x,0,ϕlower(x)):x>0}\{(x,0,\phi^{\textnormal{lower}}(x)):x>0\}.

  2. (2)

    There are functions

    yinner:Ωin[0,B),\displaystyle y^{\textnormal{inner}}:\Omega^{\textnormal{in}}\to[0,B),
    youter:Ωout[0,B),\displaystyle y^{\textnormal{outer}}:\Omega^{\textnormal{out}}\to[0,B),

    where

    Ωin\displaystyle\Omega^{\textnormal{in}} :={(x,z):xπ,zϕlower(x)},\displaystyle:=\{(x,z):x\geq\pi,\,z\leq\phi^{\textnormal{lower}}(x)\},
    Ωout\displaystyle\Omega^{\textnormal{out}} :={(x,z):xπ,zϕupper(x)},\displaystyle:=\{(x,z):x\geq\pi,\,z\leq\phi^{\textnormal{upper}}(x)\},

    such that

    Mupper{xπ}{y0}\displaystyle M^{\textnormal{upper}}\cap\{x\geq\pi\}\cap\{y\geq 0\} ={(x,youter(x,z),z):(x,z)Ωout},\displaystyle=\{(x,y^{\textnormal{outer}}(x,z),z):(x,z)\in\Omega^{\textnormal{out}}\},
    Mlower{xπ}{y0}\displaystyle M^{\textnormal{lower}}\cap\{x\geq\pi\}\cap\{y\geq 0\} ={(x,yinner(x,z),z):(x,z)Ωin}.\displaystyle=\{(x,y^{\textnormal{inner}}(x,z),z):(x,z)\in\Omega^{\textnormal{in}}\}.
  3. (3)

    MM is connected, and M{x>0}=MupperMlowerM\cap\{x>0\}=M^{\textnormal{upper}}\cup M^{\textnormal{lower}}.

  4. (4)

    For each zz, yinner(x,z)y^{\textnormal{inner}}(x,z) is a strictly decreasing function of xx.

  5. (5)

    Mupper{xπ}M^{\textnormal{upper}}\cap\{x\geq\pi\} is the graph of a function u(x,y)u(x,y)

    u:[π,)×(B,B)𝐑,u:[\pi,\infty)\times(-B,B)\longrightarrow\mathbf{R},

    and

    (B|y|)1+|Du|2ηB,(B-|y|)\sqrt{1+|Du|^{2}}\leq\eta B,

    for a constant η=η(λ)\eta=\eta(\lambda).

  6. (6)

    As zz\to\infty, M(0,0,z)M-(0,0,z) converges to the empty set.

  7. (7)

    As zz\to\infty, M+(0,0,z)M+(0,0,z) converges smoothly to the planes {y=±b}\{y=\pm b\} and {y=±B}\{y=\pm B\}.

  8. (8)

    limx(ϕlower)(x)=s(b)\displaystyle\lim_{x\to\infty}(\phi^{\textnormal{lower}})^{\prime}(x)=-s(b), and as xx\to\infty,

    Mlower(x,0,ϕlower(x))M^{\textnormal{lower}}-(x,0,\phi^{\textnormal{lower}}(x))

    converges to the grim reaper surface wb:𝐑×(b,b)𝐑w_{b}:\mathbf{R}\times(-b,b)\to\mathbf{R} with w(0,0)=0w(0,0)=0 and w/xs(b)\partial w/\partial x\equiv-s(b).

  9. (9)

    B=b+πB=b+\pi.

Proof.

Let Mn𝒜M_{n}\in\mathscr{A} be such that Mn(x(Mn),0,0)M_{n}-(x(M_{n}),0,0) converges smoothly to MM.

The proof of Assertion (1) is almost identical to the proof of Theorem 8.1.

Assertion (2) follows from Theorem 10.3.

The proof of Assertion (3) is the same as the proof of Theorem 10.5.

Concerning Assertion (4), the fact that yinner(x,z)y^{\textnormal{inner}}(x,z) is a decreasing function of the xx follows from the corresponding property of the MnM_{n} (Theorem 10.9). Strict inequality follows, for example, by the strong maximum principle. (Specifically, ν𝐞1\nu\cdot\mathbf{e}_{1} is a Jacobi field that is nonnegative on MlowerM^{\textnormal{lower}}. If it vanished at a point of MlowerM^{\textnormal{lower}}, it would vanish identically on MlowerM^{\textnormal{lower}} by the strong maximum principle, and then on all of MM by unique continuation, which is impossible since ν𝐞1=1\nu\cdot\mathbf{e}_{1}=-1 at the origin.)

Now let λ>B\lambda>B. Then λ>Bn\lambda>B_{n} for all sufficiently large nn. We may assume that λ>Bn\lambda>B_{n} for all nn. Assertion (5) follows from the corresponding property of the MnM_{n} in Theorem 14.1.

To prove Assertion (6), note (by Theorem 5.4) that every sequence ziz_{i}\to\infty has a subsequence zi(n)z_{i(n)} such that M(0,0,zi(n))M-(0,0,z_{i(n)}) converges a limit MM^{\prime} consisting of a union of planes parallel to the plane {y=0}\{y=0\}. By Assertions (2) and (3), MM^{\prime} is contained in {|x|π}\{|x|\leq\pi\}. Thus MM^{\prime} is the empty set.

By Theorem 13.4, for each ϵ\epsilon, there is an R=R(ϵ,λ)R=R(\epsilon,\lambda) such that, for each nn,

yninner(x,z)[bnϵ,bn+ϵ]y^{\textnormal{inner}}_{n}(x,z)\in[b_{n}-\epsilon,b_{n}+\epsilon] and ynouter(x,z)[Bnϵ,Bn]y^{\textnormal{outer}}_{n}(x,z)\in[B_{n}-\epsilon,B_{n}]

provided xx(Mn)+πx\geq x(M_{n})+\pi and zϕnlower(x)Rz\leq\phi_{n}^{\textnormal{lower}}(x)-R.

Passing to the limit, we see that

yinner(x,z)[bϵ,b+ϵ] and youter(x,z)[Bϵ,B],\text{$y^{\textnormal{inner}}(x,z)\in[b-\epsilon,b+\epsilon]$ and $y^{\textnormal{outer}}(x,z)\in[B-\epsilon,B]$},

provided xπx\geq\pi and yϕlower(x)Ry\leq\phi^{\textnormal{lower}}(x)-R.

Thus for each x0x\geq 0,

(40) limzyinner(x,z)=b,limzyouter(x,z)=B.\begin{gathered}\lim_{z\to-\infty}y^{\textnormal{inner}}(x,z)=b,\\ \lim_{z\to-\infty}y^{\textnormal{outer}}(x,z)=B.\end{gathered}

We know (by Theorem 5.4) that if ζn\zeta_{n}\to\infty, then M+(0,0,ζn)M+(0,0,\zeta_{n}) converges (after passing to a subsequence) to a limit consisting of planes. By (40), those planes are {y=±b}\{y=\pm b\} and {y=±B}\{y=\pm B\}. Since this is independent of the sequence and choice of subsequence, in fact M+(0,0,ζ)M+(0,0,\zeta) converges as ζ\zeta\to\infty to those planes.

The proof of Assertion (8) is identical to the proof of the corresponding assertion for M𝒜M\in\mathscr{A} in Theorem 14.1.

To prove that B=b+πB=b+\pi (Assertion (9)), note that MM is disjoint from {y=0}{x<0}\{y=0\}\cap\{x<0\}. Let Σ\Sigma be the portion of M{x<0}M\cap\{x<0\} in the halfspace {y>0}\{y>0\}.

Then Σ\Sigma lies in the half-slab {0<yB}{x<0}\{0<y\leq B\}\cap\{x<0\}, the boundary of Σ\Sigma lies in {x=0}\{x=0\}, and

supΣz()<\sup_{\Sigma}z(\cdot)<\infty

(This last inequality follows from Assertion (6).) Also Σ+(0,0,z)\Sigma+(0,0,z) converges smoothly as zz\to\infty to the halfplanes {y=b}{x0}\{y=b\}\cap\{x\leq 0\} and {y=B}{x0}\{y=B\}\cap\{x\leq 0\}.

Hence, by a general theorem (Theorem B.7) about translators in half-slabs, Bb+πB\geq b+\pi. On the other hand, Bn<bn+πB_{n}<b_{n}+\pi, so Bb+πB\leq b+\pi. Thus B=b+πB=b+\pi. ∎

Remark 15.4 (Entropy of prongs).

Taking into account (7) in Theorem 15.3, we can use Corollary 8.5 in [GMM] to deduce that a prong has entropy 44.

The following theorem describes the behavior of the prong MM in Theorem 15.3 as xx\to-\infty.

Theorem 15.5.

Let MM be a prong as in Theorem 15.3, and let

ψ(t):=maxM{x=t}z().\psi(t):=\max_{M\cap\{x=t\}}z(\cdot).

Then

M(x,0,ψ(x))M-(x,0,\psi(x))

converges smoothy at xx\to-\infty to a pair of (untilted) grim reaper surfaces, one over 𝐑×(b,b+π)\mathbf{R}\times(b,b+\pi) and the other over 𝐑×((b+π),b)\mathbf{R}\times(-(b+\pi),-b).

Proof.

Let MM^{\prime} be a subsequential limit of M(x,0,ψ(x))M-(x,0,\psi(x)) and let Σ\Sigma be a component of MM^{\prime}. By Theorem B.7(i) (applied to each of the two components of M{x<0}M\cap\{x<0\}),

(41) MM^{\prime} is contained in the two slabs given by {b|y|b+π}\{b\leq|y|\leq b+\pi\}.

Let Σ\Sigma be a component of MM^{\prime}. By symmetry, it suffices to consider the case when Σ\Sigma is contained in the slab {byb+π}\{b\leq y\leq b+\pi\}. By Theorem 5.4, Σ\Sigma is a graph or a vertical plane. By construction,

(42) maxM{x=0}z()=0.\max_{M^{\prime}\cap\{x=0\}}z(\cdot)=0.

Thus Σ\Sigma is a graph. By (41), it is an untilted grim reaper surface.

We have shown that each component of MM^{\prime} in the halfspace {y>0}\{y>0\} is a grim reaper surface over the strip {b<y<b+π}\{b<y<b+\pi\}.

Let kk be the number of grim reaper surfaces in {y>0}\{y>0\}. Then M+(0,0,z)M^{\prime}+(0,0,z) converges as zz\to\infty to the planes {y=±b}\{y=\pm b\} and {y=±(b+π)}\{y=\pm(b+\pi)\}, each with multiplicity kk. By Theorem 12.8, the plane {y=b+π}\{y=b+\pi\} occurs with multiplicity 11. Hence k=1k=1.

Note that (41) and (42) determine the two grim reaper surfaces. Thus the limit MM^{\prime} is independent of choice of subsequence, so we get convergence and not merely subsequential convergence. ∎

Theorem 15.6.

Suppose that Mn𝒜M_{n}\in\mathscr{A}, that bn:=b(Mn)b<b_{n}:=b(M_{n})\to b<\infty, Bn:=B(Mn)BB_{n}:=B(M_{n})\to B, and x(Mn)+.x(M_{n})\to+\infty. Then y(Mn)by(M_{n})\to b and B=b+πB=b+\pi. Furthermore, if ζn=maxM{x=0}z()\zeta_{n}=\max_{M\cap\{x=0\}}z(\cdot), then

Mn(0,0,ζn)M_{n}-(0,0,\zeta_{n})

converges smoothly to a pair of grim reaper surfaces over 𝐑×(b,b+π)\mathbf{R}\times(b,b+\pi) and 𝐑×((b+π),b)\mathbf{R}\times(-(b+\pi),-b).

Proof.

That B=b+πB=b+\pi was proved in Theorem 15.3(7).

Thus, for large nn, Bn>bnB_{n}>b_{n}, and therefore

y(Mn)Bnπy(M_{n})\geq B_{n}-\pi

by Theorem 14.2. Thus

lim infy(Mn)Bπ=b.\liminf y(M_{n})\geq B-\pi=b.

On the other hand, lim supy(Mn)b\limsup y(M_{n})\leq b trivially. Thus y(Mn)by(M_{n})\to b.

Let MM be a subsequential limit of Mn(0,0,ζn)M_{n}-(0,0,\zeta_{n}). Note that MM is also a subsequential limit of

Mn:=Mn{|x|<x(Mn)}(0,0,ζn)M_{n}^{\prime}:=M_{n}\cap\{|x|<x(M_{n})\}-(0,0,\zeta_{n})

and that 𝖭(x()|Mn)=0\mathsf{N}(x(\cdot)|M_{n}^{\prime})=0. Thus 𝖭(x()|M)=0\mathsf{N}(x(\cdot)|M)=0. Let Σ\Sigma be a component of MM. By Theorem 5.6, Σ\Sigma is a plane parallel to {y=0}\{y=0\}, or a Δ\Delta-wing, or a grim reaper surface. Since maxM{x=0}z()=0\max_{M\cap\{x=0\}}z(\cdot)=0, Σ\Sigma cannot be a plane. Thus Σ\Sigma is a Δ\Delta-wing or a grim reaper surface.

Since MnM_{n} is disjoint from {x=0,|y|<y(Mn)}\{x=0,\,|y|<y(M_{n})\}, we see (by Letting nn\to\infty) that MM is disjoint from {x=0,|y|<b}\{x=0,\,|y|<b\}. Thus Σ{x=0}\Sigma\cap\{x=0\} lies in {b|y|B}\{b\leq|y|\leq B\}. Since B=b+πB=b+\pi, Σ\Sigma is an untilted grim reaper surface.

We have shown that MM^{\prime} is a union of untilted grim reaper surface. Exactly as in the proof of Theorem 15.5, MM^{\prime} consists of exactly 22 grim reaper surfaces, and we get convergence, not merely subsequential convergence. ∎

16. Small Necks

In this section, we show that if its neck is very small, then an annuloid in 𝒜\mathscr{A} resembles two Δ\Delta-wings joined by a small catenoidal neck.

Theorem 16.1.

Suppose that Mn𝒜M_{n}\in\mathscr{A}, that bn:=b(Mn)b<b_{n}:=b(M_{n})\to b<\infty, and that x(Mn)0x(M_{n})\to 0. Then

  1. (1)

    Mn/x(Mn)M_{n}/x(M_{n}) converges to the catenoid whose waist is the unit circle in {z=0}\{z=0\}.

  2. (2)

    There is an n0n_{0} such that if nn0n\geq n_{0}, then waist(Mn)\operatorname{waist}(M_{n}), the set of points pp where ν(Mn,p)\nu(M_{n},p) is horizontal, is a smooth closed curve that is contained in 𝐁(0,2x(M))\mathbf{B}(0,2x(M)) and that projects diffeomorphically to Kn\partial K_{n}, where KnK_{n} is a compact, strictly convex region in 𝐑×(bn,bn)\mathbf{R}\times(-b_{n},b_{n}).

  3. (3)

    If nn0n\geq n_{0}, then one of the components of Mnwaist(Mn)M_{n}\setminus\operatorname{waist}(M_{n}) is a graph over

    (𝐑×(Bn,Bn))Kn,(\mathbf{R}\times(-B_{n},B_{n}))\setminus K_{n},

    and the other is a graph over

    (𝐑×(bn,bn)Kn.(\mathbf{R}\times(-b_{n},b_{n})\setminus K_{n}.
  4. (4)

    MnM_{n} converges to graph(fb)\operatorname{graph}(f_{b}), where fb:𝐑×(b,b)𝐑f_{b}:\mathbf{R}\times(-b,b)\to\mathbf{R} is the translator such that fb(0,0)=0f_{b}(0,0)=0 and Dfb(0,0)=0Df_{b}(0,0)=0. The convergence is smooth with multiplicity 22 away from the origin.

  5. (5)

    Bn:=B(Mn)B_{n}:=B(M_{n}) converges to bb.

Later (Theorem 22.5) we will show that Bn=bnB_{n}=b_{n} for all sufficiently large nn, provided b>π/2b>\pi/2. (We do not know whether “provided b=π/2b=\pi/2” is necessary.)

Proof.

Assertion (1) was proved in Lemma 12.10.

If nn is large, then, by Assertion (1), MnM_{n} contains a curve CnC_{n} that is a slight perturbation of the circle {z=0}𝐁(0,x(Mn))\{z=0\}\cap\partial\mathbf{B}(0,x(M_{n})) such that Tan(Mn,)\operatorname{Tan}(M_{n},\cdot) is vertical at each point of CnC_{n}. It follows that ν(Mn,)\nu(M_{n},\cdot) maps CnC_{n} diffeomorphically onto the equator. By Corollary 6.7, CnC_{n} contains all the points of waist(Mn)\operatorname{waist}(M_{n}). Thus Assertion (2) holds.

Assertion (3) follows from Assertion (2) by Lemma 16.2 below.

Note that ν𝐞3\nu\cdot\mathbf{e}_{3} is >0>0 on one component Mn+M_{n}^{+} of components of MnΓnM_{n}\setminus\Gamma_{n} and that ν𝐞3\nu\cdot\mathbf{e}_{3} is <0<0 on the other component, MnM_{n}^{-}. Each component is stable since it is a graph.

By passing to a subsequence, we can assume that Mn+M_{n}^{+} and MnM_{n}^{-} converge as sets to limits M+M^{+} and MM^{-}, both containing the origin, and that BnB_{n} converges to a limit BB. By stability, the convergence is smooth away from the origin. By Remark 12.11, M:=M+MM^{\prime}:=M^{+}\cup M^{-} is smooth and embedded (possibly with multiplicity) and Tan(M,0)\operatorname{Tan}(M^{\prime},0) is horizontal. Let Σ+\Sigma^{+} and Σ\Sigma^{-} be the components of M+M^{+} and MM^{-} containing the origin. Since MM is smooth and embedded, Σ+\Sigma^{+} and Σ\Sigma^{-} coincide near 0. Thus, by unique continuation, Σ+=Σ\Sigma^{+}=\Sigma^{-}.

Now ν𝐞3\nu\cdot\mathbf{e}_{3} is a non-negative Jacobi field on Σ+\Sigma^{+} that is >0>0 at 0. Hence ν𝐞3>0\nu\cdot\mathbf{e}_{3}>0 everywhere on Σ+\Sigma^{+}, and therefore Σ+\Sigma^{+} is graph(fβ)\operatorname{graph}(f_{\beta}) for some β\beta, where

fβ:𝐑×(β,β)𝐑f_{\beta}:\mathbf{R}\times(-\beta,\beta)\to\mathbf{R}

is the translator with fβ(0,0)=0f_{\beta}(0,0)=0 and Dfβ(0,0)=0Df_{\beta}(0,0)=0. Since Σ=Σ+\Sigma^{-}=\Sigma^{+}, we see that Σ=Σ+=graph(fβ)\Sigma^{-}=\Sigma^{+}=\operatorname{graph}(f_{\beta}). Since MnM_{n}^{-} lies in {|y|bn}\{|y|\leq b_{n}\}, we see that Σ\Sigma^{-} lies in {|y|b}\{|y|\leq b\}. Thus

(43) βb.\beta\leq b.

Note that M+(0,0,z)M^{\prime}+(0,0,z) converges as zz\to\infty to a limit M′′M^{\prime\prime} consisting of the planes {y=±β}\{y=\pm\beta\}, each with multiplicity 22. Now M′′(b,B)M^{\prime\prime}\in\mathscr{L}(b,B), so by Theorem 12.8, the plane {y=B}\{y=B\} is in M′′M^{\prime\prime}. Thus (by (43)) β=b=B\beta=b=B. ∎

Lemma 16.2.

Suppose that M𝒜M\in\mathscr{A}, and that the waist of MM is a smooth closed curve CC that projects diffeomorphically onto the boundary of a compact, strictly convex set KK in 𝐑×(b,b)\mathbf{R}\times(-b,b), where b=b(M)b=b(M). Let M+M^{+} be the set of points of MM where ν𝐞3>0\nu\cdot\mathbf{e}_{3}>0 and MM^{-} be the set of points of MM where ν𝐞3<0\nu\cdot\mathbf{e}_{3}<0. Then M+M^{+} projects diffeomorphically onto

(𝐑2×(B,B))K,(\mathbf{R}^{2}\times(-B,B))\setminus K,

and MM^{-} projects diffeorphically onto

(𝐑2×(b,b))K,(\mathbf{R}^{2}\times(-b,b))\setminus K,

where b=b(M)b=b(M) and B=B(M)B=B(M).

Proof.

We give the proof for MM^{-}. (The proof for M+M^{+} is the same.) Let

π:M𝐑2,\displaystyle\pi:M^{-}\to\mathbf{R}^{2},
π(x,y,z)=(x,y).\displaystyle\pi(x,y,z)=(x,y).

Let Ω\Omega be a connected component of 𝐑2({y=±b}K)\mathbf{R}^{2}\setminus(\{y=\pm b\}\cup\partial K). Note that if QQ is a compact subset of Ω\Omega, then π1(Q)\pi^{-1}(Q) is compact. Then (since ν𝐞3<0\nu\cdot\mathbf{e}_{3}<0 everwhere in MM^{-}), for qΩq\in\Omega, the number dΩd_{\Omega} of points in π1(q)\pi^{-1}(q) is independent of qq.

If Ω\Omega is the interior of KK, then dΩ=0d_{\Omega}=0 since π1(0,0)=ZM\pi^{-1}(0,0)=Z\cap M^{-} is empty. if Ω\Omega is the region {y>b}\{y>b\} or the region {y<b}\{y<-b\}, then dΩ=0d_{\Omega}=0 since MM lies in the slab {|y|B}\{|y|\leq B\}. Finally, if Ω\Omega is the region

𝐑2K,\mathbf{R}^{2}\setminus K,

then dΩ=1d_{\Omega}=1, since for x>x(M)x>x(M), π1(x,0)\pi^{-1}(x,0) has exactly one point, the point (x,0,ϕlower(x))(x,0,\phi^{\textnormal{lower}}(x)). ∎

Remark 16.3.

Whether every M𝒜M\in\mathscr{A} satisfies the hypothesis (and therefore also the conclusion) of Lemma 16.2 is an interesting open question. By Theorem 16.1, the hypothesis is satisfied when the necksize x(M)x(M) is sufficiently small.

The following corollary shows that several notions of necksize are very nearly the same when x(M)x(M) is small.

Corollary 16.4.

Let MnM_{n} be as in Theorem 16.1. Then

limny(Mn)x(Mn)=1\lim_{n\to\infty}\frac{y(M_{n})}{x(M_{n})}=1

and

limnLn2πx(Mn)=1\lim_{n\to\infty}\frac{L_{n}}{2\pi x(M_{n})}=1

where LnL_{n} is the length of the shortest homotopically nontrivial closed curve in MnM_{n}.

17. Continuity and Properness

In this section, we show that b(M)b(M), B(M)B(M), x(M)x(M), and y(M)y(M) depend continuously on M𝒜M\in\mathscr{A}. We also show that the map M𝒜(b(M),x(M))M\in\mathscr{A}\mapsto(b(M),x(M)) is proper.

Theorem 17.1.

Suppose that Mn𝒜M_{n}\in\mathscr{A}, and that bn:=b(Mn)b_{n}:=b(M_{n}) and xn:=x(Mn)x_{n}:=x(M_{n}) converge to finite limits bb and x^\hat{x}. Then, after passing to a subsequence, MnM_{n} converges to an MM such that b(M)=bb(M)=b, x(M)=x^x(M)=\hat{x}, and B(M)=limnB(Mn)B(M)=\lim_{n}B(M_{n}).

In particular, the map

Φ:𝒜[π/2,)×(0,)\displaystyle\Phi:\mathscr{A}\to[\pi/2,\infty)\times(0,\infty)
Φ(M)=(b(M),x(M))\displaystyle\Phi(M)=(b(M),x(M))

is proper and surjective.

Proof.

After passing to a subsequence, we can assume that the Bn:=B(Mn)B_{n}:=B(M_{n}) converge to a limit B[b,b+π]B\in[b,b+\pi] and (by Theorem C.3) that the MnM_{n} converge as sets to a limit MM:

(44) δn(Mn,M)0,\delta_{n}(M_{n},M)\to 0,

where δ(,)\delta(\cdot,\cdot) is the metric on closed subsets of 𝐑3\mathbf{R}^{3} defined in Appendix C. Now Mn𝒜(bn,Bn,xn)M_{n}\in\mathscr{A}(b_{n},B_{n},x_{n}) (by Theorem 13.5), so, by definition of 𝒜(bn,Bn,xn)\mathscr{A}(b_{n},B_{n},x_{n}), there exists a surface M~n\tilde{M}_{n}\in\mathscr{R} such that

(45) |b(M~n)bn|<2n,|B(M~n)Bn|<2n,|x(M~n)xn|<2n,a(M~n)>n,δ(Mn,Mn)<2n,\begin{gathered}|b(\tilde{M}_{n})-b_{n}|<2^{-n},\\ |B(\tilde{M}_{n})-B_{n}|<2^{-n},\\ |x(\tilde{M}_{n})-x_{n}|<2^{-n},\\ a(\tilde{M}_{n})>n,\\ \delta(M_{n}^{\prime},M_{n})<2^{-n},\end{gathered}

where Mn:=M~n(0,0,z(M~n))M_{n}^{\prime}:=\tilde{M}_{n}-(0,0,z(\tilde{M}_{n})).

By (44) and (45), δ(Mn,M)0\delta(M_{n}^{\prime},M)\to 0. The curvature and area bounds imply that the convergence is smooth. Again, by (45), b(M~n)bb(\tilde{M}_{n})\to b and x(M~n)x^x(\tilde{M}_{n})\to\hat{x}. Thus M𝒜(b,B,x^)M\in\mathscr{A}(b,B,\hat{x}) (by definition of 𝒜(b,B,x^)\mathscr{A}(b,B,\hat{x})), and therefore b(M)=bb(M)=b, B(M)=BB(M)=B, and x(M)=x^x(M)=\hat{x}. ∎

Theorem 17.2.

The maps Mb(M)M\mapsto b(M), MB(M)M\mapsto B(M), and Mx(M)M\mapsto x(M) are continuous maps from 𝒜\mathscr{A} to 𝐑\mathbf{R}.

Proof.

Suppose that Mn𝒜M_{n}\in\mathscr{A} converges to M𝒜M\in\mathscr{A}. Trivially x(Mn)x(M)x(M_{n})\to x(M). Let bn=b(Mn)b_{n}=b(M_{n}), b=b(M)b=b(M), Bn=B(Mn)B_{n}=B(M_{n}), and B=B(M)B=B(M). By passing to a subsequence, we can assume that the bnb_{n} and BnB_{n} converge to a limits bb^{\prime} and BB^{\prime} in [π/2,][\pi/2,\infty].

First, we claim that bbb^{\prime}\leq b. For if not, let b<β<bb<\beta<b^{\prime}. For all sufficiently large nn, bn>βb_{n}>\beta. For such nn,

(ϕnlower)(x)xfbn(x,0)<xfβ(x,0).(\phi_{n}^{\textnormal{lower}})^{\prime}(x)\leq\frac{\partial}{\partial x}f_{b_{n}}(x,0)<\frac{\partial x}{\partial f}_{\beta}(x,0).

Letting nn\to\infty gives

(ϕlower)(x)xfβ(x,0)(\phi^{\textnormal{lower}})^{\prime}(x)\leq\frac{\partial}{\partial x}f_{\beta}(x,0)

and therefore (letting xx\to\infty), s(b)s(β)-s(b)\leq-s(\beta), so bβb\geq\beta, a contradiction.

Since Bn[bn,bn+π)B_{n}\in[b_{n},b_{n}+\pi), we see that B[b,b+π]B^{\prime}\in[b^{\prime},b^{\prime}+\pi].

By the Properness Theorem 17.1, we can assume, after passing to a subsequence, that MnM_{n} converges to a limit MM^{\prime} with b(M)=bb(M^{\prime})=b and B(M)=BB(M^{\prime})=B. Since MnM_{n} also converges to MM, we see that M=MM^{\prime}=M and thus b(M)=b(M)=bb(M)=b(M^{\prime})=b and B(M)=B(M)=BB(M^{\prime})=B(M)=B. ∎

Continuity of the map M𝒜y(M)M\in\mathscr{A}\mapsto y(M) requires a different argument.

Lemma 17.3.

Suppose Mn𝒜M_{n}\in\mathscr{A} converges to M𝒜M\in\mathscr{A}, and suppose that ζn\zeta_{n}\to\infty. Then Mn:=Mn+(0,0,ζn)M_{n}^{\prime}:=M_{n}+(0,0,\zeta_{n}) converges smoothly to the planes {y=±b}\{y=\pm b\} and {y=±B}\{y=\pm B\}, where b=b(M)b=b(M) and B=B(M)B=B(M).

Proof.

After passing to a subsequence, we can assume that MnM_{n}^{\prime} converges to a limit MM^{\prime} consisting of planes, Δ\Delta-wings, and grim reaper surfaces. Let x>x(M)+πx>x(M)+\pi and z𝐑z\in\mathbf{R}. By the smooth convergence, x>x(Mn)+πx>x(M_{n})+\pi for large nn and

ϕnlower(x(Mn)+x))ϕnlower(x(M)+x).\phi_{n}^{\textnormal{lower}}(x(M_{n})+x))\to\phi_{n}^{\textnormal{lower}}(x(M)+x).

Thus

(zζn)ϕnlower(x(Mn)+x).(z-\zeta_{n})-\phi_{n}^{\textnormal{lower}}(x(M_{n})+x)\to-\infty.

Hence by Theorem 13.4,

|yninner(x,zζn)bn|0|y^{\textnormal{inner}}_{n}(x,z-\zeta_{n})-b_{n}|\to 0

where bn=b(Mn)b_{n}=b(M_{n}). Since bnbb_{n}\to b (by Theorem 13.5),

yninner(x,zζn)b.y^{\textnormal{inner}}_{n}(x,z-\zeta_{n})\to b.

This holds for all (x,z)(x,z) with x>x(M)x>x(M). Hence MM^{\prime} coincides with the planes {y=±b}\{y=\pm b\} and {y=±B}\{y=\pm B\} in the halfspace {x>x(M)}\{x>x(M)\}. Since MM^{\prime} consists of planes, Δ\Delta-wings, and grim reaper surfaces, in fact MM^{\prime} consists of the planes {y=±b}\{y=\pm b\} and {y=±B}\{y=\pm B\}. ∎

Theorem 17.4.

The map M𝒜y(M)M\in\mathscr{A}\mapsto y(M) is continuous.

Proof.

Suppose that Mn𝒜M_{n}\in\mathscr{A} converges to M𝒜M\in\mathscr{A}. By Theorem 17.2, Bn:=B(Mn)B_{n}:=B(M_{n}) and bn:=b(Mn)b_{n}:=b(M_{n}) converge to B=B(M)B=B(M) and b=b(M)b=b(M). By passing to a subsequence, we can assume that y(Mn)y(M_{n}) converges.

Let pM{y=0}p\in M\cap\{y=0\}. Then there exist pnM{x=0}p_{n}\in M\cap\{x=0\} such that pnpp_{n}\to p. Now

y(Mn)|y(pn)||y(p)|,y(M_{n})\leq|y(p_{n})|\to|y(p)|,

so

limy(Mn)|y(p)|.\lim y(M_{n})\leq|y(p)|.

Taking the infimum over pM{x=0}p\in M\cap\{x=0\} gives

limy(Mn)y(M).\lim y(M_{n})\leq y(M).

Thus to prove the theorem, it suffices to prove the reverse inequality.

Choose pn=(0,yn,zn)Mnp_{n}=(0,y_{n},z_{n})\in M_{n} with

(46) 0|yn|<y(Mn)+(1/n).0\leq|y_{n}|<y(M_{n})+(1/n).

By symmetry, we can assume yn0y_{n}\geq 0. We can also assume that yny_{n} converges to a limit yy_{\infty}. Thus

ylimny(Mn)y_{\infty}\leq\lim_{n}y(M_{n})

by squeeze.

Case 1: znz_{n} is bounded below. Then (passing to a subsequence) we can assume that pnp_{n} converges to a point p=(0,y,z)Mp=(0,y_{\infty},z_{\infty})\in M. Thus

y(M)y(p)=ylimny(Mn).y(M)\leq y(p)=y_{\infty}\leq\lim_{n}y(M_{n}).

Case 2: znz_{n} is not bounded below. Then, passing to a subsequence, we can assume that znz_{n}\to-\infty.

By Lemma 17.3, Mn:=Mn+(0,0,zn)M_{n}^{\prime}:=M_{n}+(0,0,z_{n}) converges smoothly to the planes {y=±b}\{y=\pm b\} and {y=±B}\{y=\pm B\}. Hence yy_{\infty} is bb or BB. In particular, byb\leq y_{\infty}. Trivially, y(M)by(M)\leq b. Thus

y(M)bylimy(Mn).y(M)\leq b\leq y_{\infty}\leq\lim y(M_{n}).

18. Gap Theorems

In the next two sections, we prove results that will be used to establish the existence of compact translating annuli bounded by pairs of nested rectangles (The existence is proved in Section 19 using a path-lifting argument). In this section, we establish conditions guaranteeing that a pair of curves does not bound a connected translator.

Theorem 18.1.

Suppose that II is an infinite open strip in 𝐑2\mathbf{R}^{2} of width π\pi. Suppose that MM is a properly immersed translator in 𝐑3{z0}\mathbf{R}^{3}\cap\{z\geq 0\} with no boundary in the slab S:=I×𝐑S:=I\times\mathbf{R}. Then MM lies in the complement of SS.

Proof.

We may assume that the strip II is 𝐑×(π/2,π/2)\mathbf{R}\times(-\pi/2,\pi/2). For α,β>0\alpha,\beta>0, let

fα,β:[α,α]×[β,β]𝐑f_{\alpha,\beta}:[-\alpha,\alpha]\times[-\beta,\beta]\to\mathbf{R}

be the unique graphical translator [graphs]*§3 with boundary values 0. By the maximum principle, if β<π/2\beta<\pi/2, then the graph of fα,βf_{\alpha,\beta} lies below MM. That is, for (x,y,z)M(x,y,z)\in M with (x,y)[α,α]×[β,β](x,y)\in[-\alpha,\alpha]\times[-\beta,\beta],

zfα,β(x,y).z\geq f_{\alpha,\beta}(x,y).

Letting βπ/2\beta\to\pi/2, we see that

zfα,π/2(x,y).z\geq f_{\alpha,\pi/2}(x,y).

As α\alpha\to\infty, fα,π/2(x,y)fα,π/2(0,0)f_{\alpha,\pi/2}(x,y)-f_{\alpha,\pi/2}(0,0) converges smoothly to the untilted grim reaper surface fπ/2(x,y):=log(cosy)f_{\pi/2}(x,y):=\log(\cos y). (See the proof of [graphs]*Theorem 4.1.) Thus fα,π/2f_{\alpha,\pi/2} converges to \infty uniformly on compact subsets of 𝐑×(π/2,π/2)\mathbf{R}\times(-\pi/2,\pi/2). The result follows immediately. ∎

Remark 18.2.

Note this does not require any regularity of MM: MM can be any closed set that satisfies the maximum principle. (In the language of [white16], one could state the result as follows: Suppose Σ\Sigma is a (2,0)(2,0) set (with respect to the Ilmanen metric) in I×𝐑I\times\mathbf{R} that lies in {z0}\{z\geq 0\}. Then Σ\Sigma is empty.)

Corollary 18.3.

If MM is a translator in {z0}\{z\geq 0\}, then MM lies in C×[0,)C\times[0,\infty), where CC is the convex hull of the projection of M\partial M to the horizontal plane.

Lemma 18.4.

Let M𝐑3{z0}M\subset\mathbf{R}^{3}\cap\{z\geq 0\} be a translator such that M\partial M lies in a compact set KK. Then MM is bounded. In particular, if KK lies below a bowl soliton QQ, then MM also lies below QQ.

Proof.

Let R=max{|p|:pM}R=\max\{|p|\;:\;p\in\partial M\}. By Corollary 18.3,

M{(x,y,z):x2+y2R2,z0}.M\subset\{(x,y,z)\;:\;x^{2}+y^{2}\leq R^{2},\,z\geq 0\}.

By [CSS], there exists a complete translating annulus Σ\Sigma such that Σ\Sigma is rotationally invariant about the zz-axis and such that

minpΣdist(p,Z)=R.\min_{p\in\Sigma}\operatorname{dist}(p,Z)=R.

By translating Σ\Sigma vertically, we can assume that the neck of Σ\Sigma is at a height h>0h>0 and that Σ\Sigma is disjoint from {(x,y,z):x2+y2+z2(2R)2}\{(x,y,z):x^{2}+y^{2}+z^{2}\leq(2R)^{2}\}. Let 𝐯\mathbf{v} be a horizontal unit vector. By the maximum principle, Σ+s𝐯\Sigma+s\mathbf{v} is disjoint from MM for 0sR.0\leq s\leq R. It follows that M is disjoint from the plane {z=h}\{z=h\}, and thus that MM lies in

{x2+y2R2}{0zh}.\{x^{2}+y^{2}\leq R^{2}\}\cap\{0\leq z\leq h\}.

The last statement ( “In particular…”) follows immediately from the maximum principle. (If MM did not lie below Q,Q, then there would be a largest s>0s>0 such that MM has a nonempty intersection with Q+s𝐞3.Q+s\mathbf{e}_{3}. For that s, the strong maximum principle would be violated at the point of contact of MM and Q+s𝐞3.Q+s\mathbf{e}_{3}.) ∎

Theorem 18.5.

Let UnUnU_{n}\subset U_{n}^{\prime} be open convex regions in 𝐑2\mathbf{R}^{2} such that UnU_{n} converges to a bounded open convex set UU and such that UnU_{n}^{\prime} converges to an infinite strip UU^{\prime}. Suppose that

min{|pq|:pU,qU}π.\min\{|p-q|:p\in\partial U,q\in\partial U^{\prime}\}\geq\pi.

Then for all sufficiently large nn, there is no connected translator in {z0}\{z\geq 0\} whose boundary is Sn:=((Un)(Un))×{0}S_{n}:=((\partial U_{n})\cup(\partial U_{n}^{\prime}))\times\{0\}.

Proof.

Suppose the result is false. Then (after passing to a subsequence) each SnS_{n} bounds a connected translator MnM_{n} in {z0}\{z\geq 0\}. Passing to a further subsequence, we can assume that the MnM_{n} converge as sets to a closed set MM in 𝐑3\mathbf{R}^{3}. Note that UUU^{\prime}\setminus U contains two parallel infinite strips I1I_{1} and I2I_{2} each of width π\pi. Thus by Theorem 18.1 (and Remark 18.2), MM is disjoint from I1×𝐑I_{1}\times\mathbf{R} and I2×𝐑I_{2}\times\mathbf{R}. Thus MM is the union of three connected components, where one component, MM^{*}, has boundary (U)×{0}(\partial U)\times\{0\}, and where each of the other two components is bounded by one of the straight lines in (U)×{0}(\partial U^{\prime})\times\{0\}.

By Lemma 18.4, MM^{*} is compact, a contradiction.

(If the contradiction is not clear, let KK be a compact set such that MM^{*} is in the interior of KK and such that MMM\setminus M^{*} is disjoint from KK. For all sufficiently large nn, MnM_{n} contains a point in K\partial K, and therefore MKM\cap\partial K is nonempty, a contradiction.) ∎

19. Limits when innerM\partial_{\textnormal{inner}}M approaches outerM\partial_{\textnormal{outer}}M

Lemma 19.1.

Let CC be a smooth convex curve in {z=0}\{z=0\} such that at each point, the radius of curvature is R>0\geq R>0. If 0<r<R0<r<R, then the set

T(C,r)={p𝐑3:dist(p,C)r}T(C,r)=\{p\in\mathbf{R}^{3}:\operatorname{dist}(p,C)\leq r\}

is a solid torus whose boundary is foliated by circles of radius rr. At each point of the boundary, the mean curvature with respect to the inward pointing normal ν\nu is greater than or equal to

1r1Rr.\frac{1}{r}-\frac{1}{R-r}.

The proof is a standard and straightforward, so we omit it. (If 𝐮\mathbf{u} and 𝐯2\mathbf{v}_{2} are orthornomal vector fields tangent to T(C,r)\partial T(C,r) with 𝐮\mathbf{u} tangent to the circles, then 𝐮𝐮ν=1/r\nabla_{\mathbf{u}}\mathbf{u}\cdot\nu=1/r and 𝐯𝐯ν1/(Rr)\nabla_{\mathbf{v}}\mathbf{v}\cdot\nu\geq-1/(R-r).)

If rR/3r\leq R/3, then Rr2rR-r\geq 2r, so the mean curvature is 1/r1/(2r)=1/(2r)\geq 1/r-1/(2r)=1/(2r). Hence we have:

Corollary 19.2.

Let h0h\geq 0. In Lemma 19.1, if rmin{R/3,1/(2h)}r\leq\min\{R/3,1/(2h)\}, then the mean curvature of T(C,r)T(C,r) with respect to the inward unit normal is everywhere h\geq h. In particular, if

rmin{R3,12},r\leq\min\left\{\frac{R}{3},\frac{1}{2}\right\},

then T(C,r)T(C,r) is mean-convex with respect to the translator metric.

Theorem 19.3.

Suppose R>0R>0, and let

r=12min{R/3,1/2}.r=\frac{1}{2}\min\{R/3,1/2\}.

Suppose AA is an annulus bounded by a pair of smooth, nested, convex curves in {z=0}\{z=0\}. Suppose that the inner curve, C1C_{1}, has radius of curvature everywhere R\geq R, and that AA lies in T(C1,r)T(C_{1},r). Then there is exactly one translator MM in T(C1,r)T(C_{1},r) with boundary A\partial A, and it is a graph over AA.

Proof.

Let CC be the set of points pp in the the unbounded component of {z=0}C1\{z=0\}\setminus C_{1} such that dist(p,C1)=r\operatorname{dist}(p,C_{1})=r. Then CC is a convex curve whose radius of curvature is everywhere R\geq R.

Claim 3.

Suppose MM is a translator in T(C1,r)T(C_{1},r) with boundary A\partial A. Then MM is a graph over AA.

To prove the claim, note that Since MM lies in T(C1,r)T(C_{1},r), it also lies in T(C,2r)T(C,2r). By Corollary 19.2, T(C,ρ)T(C,\rho) is gg-mean-convex for every ρ2r\rho\leq 2r. For ρ[r,2r]\rho\in[r,2r], AA is contained in T(C,ρ)T(C,\rho). Thus, by the maximum principle,

(47) MT(C,r).M\subset T(C,r).

Note that the projection of T(C,r)T(C,r) to {z=0}\{z=0\} (which is also T(C,r){z=0}T(C,r)\cap\{z=0\}) is an annulus AA^{\prime} whose inner boundary is C1C_{1}. Thus MM lies in A×𝐑A^{\prime}\times\mathbf{R}. Let DD be the convex planar region bounded by C2C_{2}, the outer boundary curve of AA. By the maximum principal, MM lies in D×𝐑D\times\mathbf{R} and MMM\setminus\partial M lies in the interior of D×𝐑D\times\mathbf{R}. Thus, by (47), MM lies in A×𝐑A\times\mathbf{R} and MMM\setminus\partial M lies in the interior of A×𝐑A\times\mathbf{R}.

It follows that MM is a graph over AA. For if not, there would be a t>0t>0 such that M(M+t𝐞3)M\cap(M+t\mathbf{e}_{3}) is nonempty, and, at the largest such tt, we would get a violation of the maximum principle. This completes the proof of the claim.

It follows from the claim that there is at most one translator in T(C1,r)T(C_{1},r) with boundary A\partial A.

Thus to complete the proof of the theorem, it suffices to prove that there is at least one such translator. For that, we can let MM be a gg-area-minimizing surface (flat chain mod 22 or integral current) in T(C1,r){z0}T(C_{1},r)\cap\{z\geq 0\} with boundary A\partial A. ∎

Lemma 19.4 (Dichotomy Lemma).

Let Mi𝒞M_{i}\in\mathscr{C} and suppose that the two components of Mi\partial M_{i} both converge to the same convex curve Γ\Gamma. Let DD be the translating graph bounded by Γ\Gamma. Then, after passing to a subsequence, one of the following occurs:

  1. (1)

    maxpMidist(p,Γ)0\max_{p\in M_{i}}\operatorname{dist}(p,\Gamma)\to 0, in which case x(Mi)x(Γ)x(M_{i})\to x(\Gamma), or

  2. (2)

    MiM_{i} converges to DD. Away from ΓZ\Gamma\cup Z, the convergence is smooth with some positive integer multiplicity. In this case, x(Mi)0x(M_{i})\to 0.

Proof.

Let DiD_{i} and DiD_{i}^{\prime} be the translating graphs bounded by the inner and outer components of Mi\partial M_{i}. By the maximum principle (consider vertical translates of DiD_{i} and of DiD_{i}^{\prime}), MiM_{i} lies in the closed region of {z0}\{z\geq 0\} between DiD_{i} and DiD_{i}^{\prime}.

Thus, after passing to a subsequence, the MiM_{i} converge as sets to a closed subset MM of DD.

By the curvature and area bounds, away from ZΓZ\cup\Gamma, the convergence is smooth with some integer multiplicity k0k\geq 0.

If k=0k=0, then ΓMΓ(ZD)\Gamma\subset M\subset\Gamma\cup(Z\cap D). Since each MiM_{i} is connected, MM is connected and therefore M=ΓM=\Gamma. Thus

maxpMidist(p,Γ)0\max_{p\in M_{i}}\operatorname{dist}(p,\Gamma)\to 0

and x(Mi)x(Γ)x(M_{i})\to x(\Gamma).

Now suppose that k>0k>0. Let ϵ>0\epsilon>0 be small and let

U={p𝐑3:dist(p,ZD)>ϵ}.U=\{p\in\mathbf{R}^{3}:\operatorname{dist}(p,Z\cap D)>\epsilon\}.

Then for large ii, MUM\cap U consists of kk components, each of which converges smoothly as ii\to\infty to DUD\cap U. Hence x(Mi)0x(M_{i})\to 0. ∎

The space 𝒫\mathscr{P}

Let 𝒫\mathscr{P} be the space of curves Γ\Gamma in {z=0}\{z=0\} such that Γ\Gamma consists of a pair of disjoint, convex, simple closed curves, each of which is symmetric about the xx and yy axes. As in Definition 6.1, 𝒞\mathscr{C} is the space of embedding translating annuli MM such that M𝒫\partial M\in\mathscr{P}, MM is invariant under reflection in the planes {x=0}\{x=0\} and {y=0}\{y=0\}, and MZ=M\cap Z=\emptyset.

Lemma 19.5.

Suppose that Mi𝒞M_{i}\in\mathscr{C} and that MiΓ𝒫\partial M_{i}\to\Gamma\in\mathscr{P}. Then, after passing to a subsequence, the MiM_{i} converge to a limit M𝒞M\in\mathscr{C}, and the convergence is smooth in {z>0}\{z>0\}. Furthermore, if Mi\partial M_{i} and Γ\Gamma are C2,αC^{2,\alpha} and if the Mi\partial M_{i} converge to Γ\Gamma in C2,αC^{2,\alpha}, then the MiM_{i} converge to MM in C2,αC^{2,\alpha}.

The first statement of the lemma is the assertion that the boundary map :MM\partial:M\mapsto\partial M is a continuous, proper map from 𝒞\mathscr{C} to 𝒫\mathscr{P}.

Proof.

Recall that the MiM_{i} are minimal surfaces with respect to the translator metric. By the standard compactness theory (e.g., [white18]*Theorem 1.1) and the curvature bounds in Theorem 3.4(3), we can assume that the MiM_{i} converge to a limit MM, where MΓM\setminus\Gamma is smooth and embedded (possibly with multiplicity) and where the convergence of MiM_{i} to MM is smooth in compact subsets of 𝐑3(ΓS)\mathbf{R}^{3}\setminus(\Gamma\cup S), for some locally finite subset SS of ZZ. By Corollary 3.2, the MiM_{i} (and therefore also MM) all lie some compact subset of 𝐑3\mathbf{R}^{3}. Let ϵi\epsilon_{i} be the length of the shortest closed geodesic γi\gamma_{i} in MiM_{i}. It suffices to show that ϵi\epsilon_{i} is bounded away from 0, as then the MiM_{i} converge subsequentially to an embedded minimal annulus M𝒞M\in\mathscr{C} (by the curvature bound in Theorem 3.4(2), or by standard Douglas-Rado theory.) Suppose to the contrary that ϵi0\epsilon_{i}\to 0. Note that γi\gamma_{i} converges to a point pp in MM. One component of MiγiM_{i}\setminus\gamma_{i} converges to a disk DoutD_{\textnormal{out}} in MM with boundary Γout\Gamma_{\textnormal{out}} and the other component converges to a disk DinD_{\textnormal{in}} in MM with boundary Γin\Gamma_{\textnormal{in}}; the convergence is smooth away from {p}Γ\{p\}\cup\partial\Gamma. A priori, DinD_{\textnormal{in}} and DoutD_{\textnormal{out}} might not be smooth at {p}\{p\}. But since MM is smooth, it follows that DinD_{\textnormal{in}} and DoutD_{\textnormal{out}} are smooth at pp. But then their contact at pp violates the strong maximum principle. This completes the proof of the first Assertion.

The second assertion follows immediately from Theorem 3 of [white-curv]. Choose r>0r>0 large enough that the ball B(0,r)B(0,r) of radius rr about 0 in 𝐑3\mathbf{R}^{3} contains all the MiM_{i}. Let

N=B(0,2r)¯{z0}.N=\overline{B(0,2r)}\cap\{z\geq 0\}.

Assertion (4) of [white-curv]*Theorem 3 includes the hypothesis that N\partial N is smooth and strictly convex with respect to the Riemannian metric. (Strict convexity of N\partial N at a point means that the principal curvatures with respect to the inward pointing unit normal are positive. Equivalently, it means that mean curvature vector points into NN, and that the product of the principal curvatures is positive.) However, the proof of that theorem only uses smoothness and strict convexity of N\partial N in some open subset of N\partial N containing Γ\Gamma. Indeed, smoothness and strict mean convexity of N\partial N near Γ\Gamma suffice.

Actually, [white-curv]*Theorem 3 only asserts convergence of MiM_{i} to MM in C2,βC^{2,\beta} for β<α\beta<\alpha. But convergence in C2,αC^{2,\alpha} then follows by standard Schauder estimates. ∎

20. Path Lifting

Throughout this section, we shall consider one-parameter families of closed, convex, symmetric sets in the plane {z=0}\{z=0\}

t[0,1]\displaystyle t\in[0,1] \displaystyle\longmapsto Cin(t)\displaystyle C_{\rm in}(t)
t[0,1]\displaystyle t\in[0,1] \displaystyle\longmapsto Cout(t)\displaystyle C_{\rm out}(t)

satisfying:

  1. (a)

    Cin(0)=Cout(0),C_{\rm in}(0)=C_{\rm out}(0),

  2. (b)

    Cin(t)C_{\rm in}(t) is contained in the interior of Cout(t)C_{\rm out}(t), for all t(0,1],t\in(0,1],

  3. (c)

    Cout(t)C_{\rm out}(t) is contained in the interior of Cout(t)C_{\rm out}(t^{\prime}) for all 0t<t10\leq t<t^{\prime}\leq 1.

  4. (d)

    Cin(t)C_{\rm in}(t) is compact with nonempty interior, for all t[0,1],t\in[0,1],

  5. (e)

    Cout(1)C_{\rm out}(1) is a strip 𝐑×[d,d]×{0}\mathbf{R}\times[-d,d]\times\{0\} such that the distance from Cout(1)\partial C_{\rm out}(1) to Cin(1)C_{\rm in}(1) is π.\geq\pi.

We define:

Γin(t)\displaystyle\Gamma_{\rm in}(t) :=\displaystyle:= Cin(t),\displaystyle\partial C_{\rm in}(t),
Γout(t)\displaystyle\Gamma_{\rm out}(t) :=\displaystyle:= Cout(t),\displaystyle\partial C_{\rm out}(t),
Γ(t)\displaystyle\Gamma(t) :=\displaystyle:= Γin(t)Γout(t).\displaystyle\Gamma_{\rm in}(t)\sqcup\Gamma_{\rm out}(t).

We also define

𝒞Γ(t)\displaystyle\mathscr{C}_{\Gamma(t)} ={M𝒞:M=Γ(t)},\displaystyle=\{M\in\mathscr{C}:\partial M=\Gamma(t)\},
𝒞Γ()\displaystyle\mathscr{C}_{\Gamma(\cdot)} =t(0,1)𝒞Γ(t).\displaystyle=\cup_{t\in(0,1)}\mathscr{C}_{\Gamma(t)}.

and we define a map

π:𝒞Γ()(0,1)\pi:\mathscr{C}_{\Gamma(\cdot)}\to(0,1)

by letting π(M)\pi(M) be the unique t(0,1)t\in(0,1) such that

M=Γ(t).\partial M=\Gamma(t).

(Uniqueness of tt follows from (c).)

Theorem 18.5 implies the following:

Lemma 20.1.

For tt sufficiently close to 11, Γ(t)\Gamma(t) bounds no connected translator in the halfspace {z0}.\{z\geq 0\}.

Refer to caption
Figure 7. 𝒞Γ()\mathscr{C}_{\Gamma(\cdot)} is homeomorphic to a 11-manifold without boundary.

Let Γ=Γin(0)=Γout(0)\Gamma=\Gamma_{\rm in}(0)=\Gamma_{\rm out}(0) and let DD be the unique graphical translator with boundary Γ\Gamma. The main result of this section asserts the existence of a connected family, {\mathscr{F}^{\prime}}, of compact translating annuli, each of which has boundary Γ(t)\Gamma(t) for some t(0,1)t\in(0,1). They have the further property that, when the elements are considered as subsets of 𝐑3\mathbf{R}^{3}, the closure of \mathscr{F}^{\prime} is compact and equal to :={Γ,D}\mathscr{F}:=\mathscr{F}^{\prime}\cup\{\Gamma,D\}.

Theorem 20.2.

Let Γ=Γin(0)=Γout(0).\Gamma=\Gamma_{\rm in}(0)=\Gamma_{\rm out}(0). Let (x(Γ),0,0)(x(\Gamma),0,0) be the point of Γ\Gamma in the positive xx-axis.

  1. (1)

    If 0<x^<x(Γ)0<\hat{x}<x(\Gamma), then there exists a surface M𝒞Γ()M\in\mathscr{C}_{\Gamma(\cdot)} such that x(M)=x^x(M)=\hat{x}.

  2. (2)

    If II is a closed interval in (0,x(Γ))(0,x(\Gamma)), then there is a compact, connected subset 𝒢I\mathscr{G}_{I} of 𝒞Γ()\mathscr{C}_{\Gamma(\cdot)} such that

    {x(M):M𝒢I}=I.\{x(M):M\in\mathscr{G}_{I}\}=I.

(We are not asserting uniqueness of the MM in Assertion (1) or of the 𝒢I\mathscr{G}_{I} in Assertion (2).)

A special case

Fix an integer k2k\geq 2 and an α(0,1)\alpha\in(0,1). If MM and NN are smooth manifolds, recall that Ck,α+(M,N)C^{k,\alpha+}(M,N) is the closure in Ck,αC^{k,\alpha} of C(M,N)C^{\infty}(M,N). Let 𝒫~\tilde{\mathscr{P}} be the set of Γ𝒫\Gamma\in\mathscr{P} such that Γin\Gamma_{\textnormal{in}} and Γout\Gamma_{\textnormal{out}} are Ck,α+C^{k,\alpha+} curves with nowhere vanishing curvature. Let 𝒞~\tilde{\mathscr{C}} be the set of M𝒞M\in\mathscr{C} such that M𝒫~\partial M\in\tilde{\mathscr{P}}.

First, we will prove Theorem 20.2 under some additional hypotheses. Then we will deduce the general theorem from the special case. The additional hypotheses are:

  1. (h1).

    Each Γ(t)\Gamma(t) is in 𝒫~\tilde{\mathscr{P}}.

  2. (h2).

    The map tΓ(t)t\mapsto\Gamma(t) is a a smooth embedding of [0,1)[0,1) into 𝒫~\tilde{\mathscr{P}}.

  3. (h3).

    𝒞Γ()\mathscr{C}_{\Gamma(\cdot)} is homeomorphic to a 11-manifold without boundary. (The 11-manifold need not be connected.) Thus each connected component of 𝒞Γ()\mathscr{C}_{\Gamma(\cdot)} is homeomorphic to a circle or to 𝐑\mathbf{R} (see Figure 7.)

By [white87]*Theorem 3.3(8) and §1.5 (see also Remark 20.3), 𝒞~\tilde{\mathscr{C}} and 𝒫~\tilde{\mathscr{P}} are smooth, separable Banach manifolds and

(*) :M𝒞~M𝒫~\partial:M\in\tilde{\mathscr{C}}\mapsto\partial M\in\tilde{\mathscr{P}}

is a smooth map of Fredholm index 0. By the Sard-Smale Theorem, a generic map satisfying (h1) and (h2) is transverse to the map (*A special case), and therefore also satisfies (h3).

Proof of Theorem 20.2 assuming (h1)–(h3).

By Theorem 19.3, there is a δ(0,1)\delta\in(0,1) such that if t(0,δ]t\in(0,\delta], then Γ(t)\Gamma(t) bounds a translating graph Σ(t)\Sigma(t). By the maximum principle, Σ(t)\Sigma(t) is the unique graphical translator with boundary Γ(t)\Gamma(t):

(48) If M𝒞Γ(t) and MΣ(t), then M is not graphical.\text{If $M\in\mathscr{C}_{\Gamma(t)}$ and $M\neq\Sigma(t)$, then $M$ is not graphical}.

The uniqueness implies that Σ(t)\Sigma(t) depends smoothly on tt for t(0,δ]t\in(0,\delta].

Let E={Σ(t):0<tδ}\pazocal{E}=\{\Sigma(t):0<t\leq\delta\} and let 𝒢\mathscr{G} be the connected component of 𝒞Γ()\mathscr{C}_{\Gamma(\cdot)} containing E\pazocal{E}.

As t0t\to 0, the Σ(t)\Sigma(t) converge (as sets) to the curve Γ\Gamma. Thus E\pazocal{E} is one end of 𝒢\mathscr{G}. Let Mi𝒢M_{i}\in\mathscr{G} be a sequence that diverges to the other end. Let ti=π(Mi)t_{i}=\pi(M_{i}) (i.e, let Mi=Γ(ti)\partial M_{i}=\Gamma(t_{i})).

By Lemma 20.1, the sequence tit_{i} is bounded above by some T<1T<1. By properness of the map :𝒞𝒫\partial:\mathscr{C}\to\mathscr{P} (see Lemma 19.5), ti0t_{i}\to 0.

Thus for large ii, ti(0,δ)t_{i}\in(0,\delta), and, since MiEM_{i}\notin\pazocal{E}, it follows that MiM_{i} is not graphical (by (48)). Thus x(Mi)0x(M_{i})\to 0 by Lemma 19.4.

Let s𝐑M(s)s\in\mathbf{R}\mapsto M(s) be a parametrization of 𝒢\mathscr{G} such that M(s)EM(s)\in\pazocal{E} if and only if s<0s<0.

We have shown that x(M(s))0x(M(s))\to 0 as ss\to-\infty and that x(M(s))x(Γ)x(M(s))\to x(\Gamma) as ss\to\infty. Thus x(M())x(M(\cdot)) takes every value in (0,x(Γ))(0,x(\Gamma)), so Asssertion (1) holds.

Now let I=[d1,d2]I=[d_{1},d_{2}] be a compact subinterval of (0,x(Γ))(0,x(\Gamma)). Let s1s_{1} be the largest ss for which x(M(s))=d1x(M(s))=d_{1}. Now let s2s_{2} be the smallest ss1s\leq s_{1} for which x(M(s))=d2x(M(s))=d_{2}. Then

𝒢I:={M(s):s1ss2}\mathscr{G}_{I}:=\{M(s):s_{1}\leq s\leq s_{2}\}

is a compact, connected subset of 𝒞Γ()\mathscr{C}_{\Gamma(\cdot)} such that

{x(M):M𝒢I}=[d1,d2].\{x(M):M\in\mathscr{G}_{I}\}=[d_{1},d_{2}].

This completes the proof of Theorem 20.2 assuming the extra hypotheses (h1)–(h3). ∎

The general case

Now we prove Theorem 20.2 without assuming extra hypotheses:

Proof of Theorem 20.2.

The curves Γ(t)\Gamma(t) need not be smooth. However, we can smooth those curves to get, for n𝐍n\in\mathbf{N}, a family

t[0,1]Γn(t)t\in[0,1]\mapsto\Gamma^{n}(t)

that satisfies the hypotheses of Theorem 20.2 and also the addition hypotheses (h1) and (h2). We can do the smoothing in such a way that:

(49) If tn(0,1]t_{n}\in(0,1] converges to t(0,1]t\in(0,1], then Γn(tn)\Gamma^{n}(t_{n}) converges to Γ(t)\Gamma(t)

As mentioned at the beginning of the proof of Theorem 20.2, the condition (h3) is generic. Thus by making a small, generic perturbation of tΓn(t)t\mapsto\Gamma^{n}(t), we can assume that it also satisfies the hypothesis (h3). Since the perturbations can be arbitrarily small, we can do them in such a way that (49) still holds for the perturbed families.

Claim 4.

Suppose that Mn𝒞Γn(tn)M_{n}\in\mathscr{C}_{\Gamma^{n}(t_{n})} and that x(Mn)x(M_{n}) converges to a limit x^\hat{x} with 0<x^<x(Γ)0<\hat{x}<x(\Gamma). Then there exist i(n)i(n)\to\infty such that ti(n)t_{i(n)} converges to a limit t(0,1)t\in(0,1) and such that Mi(n)M_{i(n)} converges to a limit M𝒞M\in\mathscr{C} with M=Γ(t)\partial M=\Gamma(t).

Likewise if there is a subsequence Mi(n)𝒞Γi(n)(ti(n))M_{i(n)}\in\mathscr{C}_{\Gamma^{i(n)}}(t_{i(n)}) with x(Mi(n))x(M_{i(n)}) bounded away from 0, then, after passing to a further subsequence, we get convergence to limits t(0,)t\in(0,\infty) and M𝒞Γ(t)M\in\mathscr{C}_{\Gamma(t)}.

To prove the claim, we can assume, by passing to a subsequence, that the tnt_{n} converge to a limit t[0,1]t\in[0,1]. By Lemma 19.4, t>0t>0. By Theorem 18.5, t<1t<1. Hence Mn=Γn(tn)\partial M_{n}=\Gamma^{n}(t_{n}) converges to Γ(t)\Gamma(t). By properness (Lemma 19.5), the MnM_{n} converge (after passing to a further subsequence) to a limit M𝒞M\in\mathscr{C} with M=Γ(t)\partial M=\Gamma(t). This completes the proof of the claim.

Now let x^(0,x(Γ))\hat{x}\in(0,x(\Gamma)). Then x^(0,x(Γn))\hat{x}\in(0,x(\Gamma^{n})) for all sufficiently large nn. Thus (for such nn) there exists an tn(0,1)t_{n}\in(0,1) and an Mn𝒞M_{n}\in\mathscr{C} such that Mn=Γn(tn)\partial M_{n}=\Gamma^{n}(t_{n}) and such that x(Mn)=x^x(M_{n})=\hat{x}. By the claim, a subsequence of the MnM_{n} will converge to an M𝒞ΓM\in\mathscr{C}_{\Gamma} with x(M)=x^x(M)=\hat{x}. This proves Assertion (1) of Theorem 20.2.

Now let I=[d,d]I=[d,d^{\prime}] be a compact subinterval of (0,x(Γ))(0,x(\Gamma)). By passing to a subsequence, we can assume that d<x(Γn)d^{\prime}<x(\Gamma^{n}) for all nn. Since Theorem 20.2 holds for Γn()\Gamma^{n}(\cdot), there is a compact, connected set 𝒢nI\mathscr{G}^{I}_{n} of 𝒞Γn()\mathscr{C}_{\Gamma^{n}(\cdot)} such that

{x(M):M𝒢nI}=[d,d].\{x(M):M\in\mathscr{G}^{I}_{n}\}=[d,d^{\prime}].

By Theorem C.2, a subsequence of the 𝒢nI\mathscr{G}^{I}_{n} converges to a compacted, connected subset 𝒢I\mathscr{G}^{I} of 𝐊(𝐑3)\mathbf{K}(\mathbf{R}^{3}), the space of all closed subsets of 𝐑3\mathbf{R}^{3}. To simplify notation, we assume that the original sequence 𝒢nI\mathscr{G}^{I}_{n} converges to 𝒢I\mathscr{G}^{I}.

Let M𝒢IM\in\mathscr{G}^{I}. Then there exist Mn𝒢nIM_{n}\in\mathscr{G}^{I}_{n} such that the MnM_{n} converge as sets to MM. By Claim 4, M𝒞Γ()M\in\mathscr{C}_{\Gamma(\cdot)} and the convergence is smooth away from the boundary, so

x(M)=limx(Mn).x(M)=\lim x(M_{n}).

Thus 𝒢I𝒞Γ()\mathscr{G}^{I}\subset\mathscr{C}_{\Gamma(\cdot)} and

(50) {x(M):M𝒢I}[d,d].\{x(M):M\in\mathscr{G}^{I}\}\subset[d,d^{\prime}].

Now let x^[d,d]\hat{x}\in[d,d^{\prime}]. Then there exist Mn𝒢nIM_{n}\in\mathscr{G}^{I}_{n} with x(Mn)=x^x(M_{n})=\hat{x}. By Claim 4, a subsequence Mn(i)M_{n(i)} will converge to a limit set M𝒞Γ()M\in\mathscr{C}_{\Gamma(\cdot)} with x(M)=x^x(M)=\hat{x}. The existence of a subsequence Mn(i)𝒢n(i)IM_{n(i)}\in\mathscr{G}^{I}_{n(i)} converging to MM means, by the definition of lim sup\limsup of sets (Definition C.1) that

Mlim supn𝒢nI.M\in\limsup_{n}\mathscr{G}^{I}_{n}.

But since the 𝒢nI\mathscr{G}^{I}_{n} converge, the limsup is the same as the limit, so M𝒢M\in\mathscr{G}.

We have shown that for each x^[d,d]\hat{x}\in[d,d^{\prime}], there exist M𝒢M\in\mathscr{G} with x(M)=x^x(M)=\hat{x}, so

[d,d]{x(M):M𝒢I}.[d,d^{\prime}]\subset\{x(M):M\in\mathscr{G}^{I}\}.

But the reverse inclusion (50) also holds, so the two sets are equal. ∎

Remark 20.3.

In the discussion at the beginning of the proof of Theorem 20.2, we asserted that 𝒫~\tilde{\mathscr{P}} was a Banach manifold. This requires a little explanation, since, in general, the space of CkC^{k} (or Ck,αC^{k,\alpha}, or Ck,α+)C^{k,\alpha+}) compact submanifolds of 𝐑N\mathbf{R}^{N} is not a Banach manifold.

If Γ𝒫~\Gamma\in\tilde{\mathscr{P}}, let fΓ:SS1(0,)2f_{\Gamma}:\SS^{1}\to(0,\infty)^{2} be the map such that for each pSS1p\in\SS^{1}, fΓ(p)f_{\Gamma}(p) is the unique (rin,rout)(0,)2(r_{\textnormal{in}},r_{\textnormal{out}})\in(0,\infty)^{2} such that rinpΓinr_{\textnormal{in}}p\in\Gamma_{\textnormal{in}} and routpΓoutr_{\textnormal{out}}p\in\Gamma_{\textnormal{out}}. Then

ΓfΓ\Gamma\mapsto f_{\Gamma}

maps 𝒫~\tilde{\mathscr{P}} homeomorphically onto an open subset of Csymk,α+(SS1,𝐑2)C^{k,\alpha+}_{\textnormal{sym}}(\SS^{1},\mathbf{R}^{2}), the space of fCk,α+(SS1,𝐑2)f\in C^{k,\alpha+}(\SS^{1},\mathbf{R}^{2}) that have the symmetries f(x,y)f(x,y)f(x,y)f(x,y)\equiv f(-x,y)\equiv f(x,-y). Thus we can regard 𝒫~\tilde{\mathscr{P}} as a smooth, separable Banach manifold.

For various technical reasons, [white87] works with spaces of parametrized boundaries. But since each Γ𝒫~\Gamma\in\tilde{\mathscr{P}} has a canonical parametrization fΓf_{\Gamma}, here we need not distinguish between parametrized and unparametrized boundaries.

21. A connected family of Annuloids in 𝒜(b)\mathscr{A}(b)

Lemma 21.1.

Let bπ/2b\geq\pi/2 and a>0a>0.

  1. (1)

    If x^(0,a)\hat{x}\in(0,a), there exists an MM\in\mathscr{R} with a(M)aa(M)\geq a, b(M)=bb(M)=b, and x(M)=x^x(M)=\hat{x}.

  2. (2)

    If I=[d,d]I=[d,d^{\prime}] is a compact interval in (0,a)(0,a), then there exists a compact, connected subset 𝒢I\mathscr{G}^{I} of 𝐊(𝐑3)\mathscr{R}\subset\mathbf{K}(\mathbf{R}^{3}) such for each M𝒢IM\in\mathscr{G}^{I},

    a(M)\displaystyle a(M) a,\displaystyle\geq a,
    b(M)\displaystyle b(M) =b,\displaystyle=b,

    and such that

    {x(M):M𝒢I}=I.\{x(M):M\in\mathscr{G}^{I}\}=I.
Proof.

Define a one-parameter family t[0,1]Γ(t)=Γin(t)Γout(t)t\in[0,1]\mapsto\Gamma(t)=\Gamma_{\textnormal{in}}(t)\cup\Gamma_{\textnormal{out}}(t) as follows:

  1. (i)

    Γin(t)\Gamma_{\textnormal{in}}(t) is the boundary of [a,a]×[b,b][-a,a]\times[-b,b].

  2. (ii)

    Γout(t)\Gamma_{\textnormal{out}}(t) is the boundary of [A(t),A(t)]×[B(t),B(t)][-A(t),A(t)]\times[-B(t),B(t)], where A()A(\cdot) and B()B(\cdot) are continuous, strictly increasing functions such that A(0)=aA(0)=a, A(1)=A(1)=\infty, B(0)=bB(0)=b, and B(1)=b+πB(1)=b+\pi.

Assertions (1) and (2) now follow immediately from Theorem 20.2. ∎

Theorem 21.2.

Suppose bπ/2b\geq\pi/2. The space 𝒜(b)\mathscr{A}(b) contains a closed, connected subset =(b)\mathscr{F}=\mathscr{F}(b) such that

{x(M):M}=(0,).\{x(M):M\in\mathscr{F}\}=(0,\infty).

We do not know whether there is a unique such subset \mathscr{F}.

Proof.

Let I=[d,d]I=[d,d^{\prime}] be a compact interval in (0,)(0,\infty). Let ana_{n} be a sequence of numbers such that an>da_{n}>d^{\prime} and such that ana_{n}\to\infty. By Lemma 21.1, there exists a compact, connected subset 𝒢nI\mathscr{G}^{I}_{n} of 𝐊(𝐑3)\mathscr{R}\subset\mathbf{K}(\mathbf{R}^{3}) such that for each M𝒢nIM\in\mathscr{G}^{I}_{n},

a(M)an,\displaystyle a(M)\geq a_{n},
b(M)=b,\displaystyle b(M)=b,

and such that

{x(M):M𝒢nI}=I.\{x(M):M\in\mathscr{G}_{n}^{I}\}=I.

Let

nI={M(0,0,z(M)):M𝒢nI}.\mathscr{F}^{I}_{n}=\{M-(0,0,z(M)):M\in\mathscr{G}^{I}_{n}\}.

Then nI\mathscr{F}^{I}_{n} is a compact, connected subset of 𝐊(𝐑3)\mathbf{K}(\mathbf{R}^{3}), and

(51) {x(M):MnI}=I.\{x(M):M\in\mathscr{F}_{n}^{I}\}=I.

By passing to a subsequence, we can assume (by Theorem C.2) that the nI\mathscr{F}^{I}_{n} converge to a compact, connected subset I\mathscr{F}^{I} of 𝐊(𝐑3)\mathbf{K}(\mathbf{R}^{3}).

By definition of 𝒜(b)\mathscr{A}(b), I𝒜(b)\mathscr{F}^{I}\subset\mathscr{A}(b). Letting nn\to\infty in (51) gives

(52) {x(M):MI}=I.\{x(M):M\in\mathscr{F}^{I}\}=I.

We have shown

(53) For every compact interval I(0,)I\subset(0,\infty), there is a compact,
connected family I\mathscr{F}^{I} of 𝒜(b)\mathscr{A}(b) such that (52) holds.

Note that 𝒜(b)\mathscr{A}(b) is a metric space (it is a subspace of 𝐊(𝐑3)\mathbf{K}(\mathbf{R}^{3})), and that

(54) Mx(M)M\mapsto x(M) is a continuous, proper map from 𝒜(b)\mathscr{A}(b) to (0,)(0,\infty).

By a general theorem about metric spaces (Theorem C.6), statements (53) and (54) imply that the space 𝒜(b)\mathscr{A}(b) contains a closed, connected subset (b)\mathscr{F}(b) such that

{x(M):M(b)}=(0,).\{x(M):M\in\mathscr{F}(b)\}=(0,\infty).

22. Capped and Uncapped Annuloids

Suppose M𝒜M\in\mathscr{A} is an annuloid with B(M)>π/2B(M)>\pi/2. Recall (see (3) in Theorem 14.1) that the limit L:=limx(ϕupper)(x)\displaystyle L:=\lim_{x\to\infty}(\phi^{\textnormal{upper}})^{\prime}(x) exists and is either s(B)s(B) or s(B)-s(B). If L=s(B)L=-s(B), we say that the annuloid is capped. If L=s(B)L=s(B), we say that it is uncapped.

(We do not defined capped and uncapped for annuloids MM with B(M)=π/2B(M)=\pi/2.)

Theorem 22.1.

Suppose that M𝒜M\in\mathscr{A} and that B=B(M)>π/2B=B(M)>\pi/2. If MM is uncapped, then z()|Mz(\cdot)|M is not bounded above. If MM is capped, then z()|Mz(\cdot)|M attains its maximum.

Proof.

The first assertion follows immediately from the definition. For the second, recall that the function

ψM(t):=maxM{x=t}z()\psi_{M}(t):=\max_{M\cap\{x=t\}}z(\cdot)

is a Lipschitz function. Also, ψM(x)=ϕupper(x)\psi_{M}(x)=\phi^{\textnormal{upper}}(x) for xx(M)+πx\geq x(M)+\pi (by Theorem 10.3), so limxψM(x)=\lim_{x\to\infty}\psi_{M}(x)=-\infty. The theorem follows immediately. ∎

Lemma 22.2.

Let h(x,y)=log(cosy)h(x,y)=\log(\cos y) and H(x,y,z)=zh(x,y)H(x,y,z)=z-h(x,y) for |y|<π/2|y|<\pi/2. If M𝒞M\in\mathscr{C} or M𝒜M\in\mathscr{A}, then

𝖭(H|M{x>0})1.\mathsf{N}(H|M\cap\{x>0\})\leq 1.
Proof.

By lower semicontinuity, it suffices to prove it for M𝒞M\in\mathscr{C}. We use

𝖭(H|M)|S|χ(MW)\mathsf{N}(H|M)\leq|S|-\chi(M\cap W)

where SS is the set of local minima of H|MH|\partial M and WW is the slab {|y|<π/2}\{|y|<\pi/2\}. Note that SS consists of four points, namely (±a(M),0,0)(\pm a(M),0,0) and (±A(M),0,0)(\pm A(M),0,0). Also, by symmetry,

𝖭(H|M)=2𝖭(H|M{x>0})+𝖭(H|M{x=0}).\mathsf{N}(H|M)=2\mathsf{N}(H|M\cap\{x>0\})+\mathsf{N}(H|M\cap\{x=0\}).

Thus

2𝖭(H|M{x>0})+𝖭(H|M{x=0})+χ(MW)4,2\mathsf{N}(H|M\cap\{x>0\})+\mathsf{N}(H|M\cap\{x=0\})+\chi(M\cap W)\leq 4,

Hence to prove the lemma, it suffices to show that

(55) 𝖭(H|M{x=0})+χ(MW)1.\mathsf{N}(H|M\cap\{x=0\})+\chi(M\cap W)\geq 1.

There are two cases: y(M)<π/2y(M)<\pi/2 and y(M)π/2y(M)\geq\pi/2.

If y(M)π/2y(M)\geq\pi/2, then (MW){x=0}=(M\cap W)\cap\{x=0\}=\emptyset, so MWM\cap W has two components, both simply connected, and thus χ(MW)=2\chi(M\cap W)=2, so (55) holds.

If y(M)<π/2y(M)<\pi/2, note that the function HH is \infty on the endpoints of the curve M{x=0}{y>0}M\cap\{x=0\}\cap\{y>0\} and thus it has at least one critical point on that curve. By symmetry, that point is also a critical point of H|MH|M. Thus 𝖭(H|M{x=0})>0\mathsf{N}(H|M\cap\{x=0\})>0, so (55) holds. (Recall that MWM\cap W has nonnegative Euler characteristic by Lemma 6.5.) ∎

Corollary 22.3.

Suppose that M𝒜M\in\mathscr{A} and that B(M)>π/2B(M)>\pi/2. Then either

  1. (1)

    ϕupper\phi^{\textnormal{upper}} is strictly increasing, with no critical points, or

  2. (2)

    ϕupper\phi^{\textnormal{upper}} has exactly one critical point, an absolute maximum.

In the first case, MM is uncapped, and in the second case, MM is capped.

Proof.

Since any critical point of ϕupper\phi^{\textnormal{upper}} is also a critical point of H|M{x>0}H|M\cap\{x>0\}, we see (from the lemma) that ϕupper\phi^{\textnormal{upper}} either has no critical point or exactly one critical point, a nondegenerate one. The corollary now follows from the fact that (ϕupper)(x(M))=(\phi^{\textnormal{upper}})^{\prime}(x(M))=\infty. ∎

Theorem 22.4.

Let 𝒮\mathscr{S} be the set of M𝒜M\in\mathscr{A} such that B(M)>π/2B(M)>\pi/2 and such that MM is capped. Then 𝒮\mathscr{S} is an open subset of 𝒜\mathscr{A}.

Proof.

Suppose that Mn𝒜M_{n}\in\mathscr{A} converge to M𝒮M\in\mathscr{S}. By Corollary 22.3, ϕupper\phi^{\textnormal{upper}} has a nondegenerate local maximum (its absolute maximum). Thus ϕnupper\phi_{n}^{\textnormal{upper}} has a local maximum for all sufficiently large nn. For such nn, MnM_{n} is capped by Corollary 22.3. ∎

Theorem 22.5.

Suppose bπ/2b\geq\pi/2.

  1. (1)

    There is a λ=λ(b)\lambda=\lambda(b) such that if M𝒜(b)M\in\mathscr{A}(b) and x(M)>λx(M)>\lambda, then B(M)>bB(M)>b, and therefore MM is uncapped.

  2. (2)

    If b>π/2b>\pi/2, there is an ϵ=ϵ(b)\epsilon=\epsilon(b) with the following property. If M𝒜(b)M\in\mathscr{A}(b) and if x(M)<ϵx(M)<\epsilon, then MM is capped, and therefore B(M)=bB(M)=b.

Proof.

Let Mn𝒜(b)M_{n}\in\mathscr{A}(b) with x(Mn)x(M_{n})\to\infty. Then B(Mn)b+πB(M_{n})\to b+\pi (by Theorem 15.3(9)), so B(Mn)>bB(M_{n})>b for all sufficiently large nn. If B(M)>b(M)B(M)>b(M), then MM is uncapped by Theorem 14.1(4). This proves Assertion (1).

Now suppose that b>π/2b>\pi/2, that Mn𝒜(b)M_{n}\in\mathscr{A}(b), and that x(Mn)0x(M_{n})\to 0. Then for each x>0x>0, (ϕnupper)(x)<0(\phi_{n}^{\textnormal{upper}}{})^{\prime}(x)<0 for all sufficiently large nn by Theorem 16.1[5]. If ϕnupper(x)<0\phi_{n}^{\textnormal{upper}}{}^{\prime}(x)<0 for some xx, then MnM_{n} is capped by Corollary 22.3 and therefore B(Mn)=bB(M_{n})=b by Assertion (1). This proves Assertion (2). ∎

Theorem 22.6.

Suppose b>π/2b>\pi/2. There exist capped annuloids Mn𝒜(b)M_{n}\in\mathscr{A}(b) that converge to an uncapped annuloid M𝒜(b)M\in\mathscr{A}(b). Let pnp_{n} be the highest point on graph(ϕnupper)\operatorname{graph}(\phi_{n}^{\textnormal{upper}}). Then MnpnM_{n}-p_{n} converges smoothly as nn\to\infty to the Δ\Delta-wing in the slab {|y|<b}\{|y|<b\} whose highest point is the origin.

See Figure 8 for a sketch of Mi{y=0}{x>0}M_{i}\cap\{y=0\}\cap\{x>0\} for large ii.

Refer to caption
Figure 8. The curve Mi{y=0}M_{i}\cap\{y=0\} in Theorem 22.6.
Proof.

Recall (Theorem 21.2) that 𝒜(b)\mathscr{A}(b) has a connected subset (b)\mathscr{F}(b) containing annuloids MM of every necksize. Thus by Theorem 22.5, (b)\mathscr{F}(b) contains both capped and uncapped annuloids. Since the set of capped M(b)M\in\mathscr{F}(b) is relatively open, it cannot be closed. Thus there are capped Mn(b)M_{n}\in\mathscr{F}(b) that converge to an uncapped MM in \mathscr{F}. Since the MnM_{n} are capped,

B(Mn)=b.B(M_{n})=b.

Let pn=(xn,0,ϕnupper(xn))p_{n}=(x_{n},0,\phi_{n}^{\textnormal{upper}}(x_{n})). Recall (see Theorem 12.6) that Mn{xx(Mn)+π}M_{n}\cap\{x\geq x(M_{n})+\pi\} is the graph of a function

un:[x(Mn)+π,)×(b,b)𝐑.u_{n}:[x(M_{n})+\pi,\infty)\times(-b,b)\to\mathbf{R}.

The gradient bound (4) in Theorem 12.6) implies that un(xn+x,y)un(xn,0)u_{n}(x_{n}+x,y)-u_{n}(x_{n},0) converges smoothly (perhaps after passing to a subsequence) to a smooth translator

u:𝐑×(b,b)𝐑.u:\mathbf{R}\times(-b,b)\to\mathbf{R}.

Since xun(xn,0)=0\frac{\partial}{\partial x}u_{n}(x_{n},0)=0, we see that xu(0,0)=0\frac{\partial}{\partial x}u(0,0)=0 and thus that uu is a Δ\Delta-wing. (We are assuming that b>π/2b>\pi/2, so uu cannot be an untilted grim reaper surface.) ∎

Corollary 22.7.

If b>π/2b>\pi/2, then there exist an uncapped annuloid M𝒜(b)M\in\mathscr{A}(b) for which B(M)=bB(M)=b.

Proof.

Let M𝒜(b)M\in\mathscr{A}(b) be as in Theorem 22.6. Since B(Mn)=bB(M_{n})=b, we see that B(M)=bB(M)=b (by Theorem 17.2). ∎

23. A graphical property of uncapped annuloids and prongs

Here we show that if MM is an uncapped annuloid in 𝒜\mathscr{A}, then M{x>0}M\cap\{x>0\} is a sideways graph x=x(y,z)x=x(y,z). It follows (Theorem 23.6) that if MM is a prong, then MM is also a sideways graph.

For c𝐑c\in\mathbf{R}, let

Ic\displaystyle I_{c} =(cπ/2,c+π/2),\displaystyle=(c-\pi/2,c+\pi/2),
Wc\displaystyle W_{c} =𝐑×Ic×𝐑.\displaystyle=\mathbf{R}\times I_{c}\times\mathbf{R}.

Let hc:𝐑×Ic𝐑h_{c}:\mathbf{R}\times I_{c}\to\mathbf{R} be the untilted grim reaper surface with hc(x,c)0h_{c}(x,c)\equiv 0, and let

Hc:Wc𝐑,\displaystyle H_{c}:W_{c}\to\mathbf{R},
Hc(x,y,z)=zhc(x,y).\displaystyle H_{c}(x,y,z)=z-h_{c}(x,y).

For MM\in\mathscr{R}, we will use the formula (see Theorem 4.3)

(56) 𝖭(Hc|M)|S||T|χ(MWc).\mathsf{N}(H_{c}|M)\leq|S|-|T|-\chi(M\cap W_{c}).

where SS is the set of local minima of Hc|MH_{c}|\partial M that are also local minima of Hc|MH_{c}|M, and where TT is the set of local maxima of Hc|MH_{c}|\partial M that are not local maxima of Hc|MH_{c}|M.

By symmetry,

𝖭(Hc|M)=𝖭(Hc|M{x=0})+2𝖭(Hc|M{x>0}).\mathsf{N}(H_{c}|M)=\mathsf{N}(H_{c}|M\cap\{x=0\})+2\mathsf{N}(H_{c}|M\cap\{x>0\}).

Thus we can rewrite (56) as

(57) χ(MWc)+𝖭(Hc|M{x=0})+2𝖭(Hc|M{x>0})|S||T|.\chi(M\cap W_{c})+\mathsf{N}(H_{c}|M\cap\{x=0\})+2\mathsf{N}(H_{c}|M\cap\{x>0\})\leq|S|-|T|.

Recall that

y(M):=inf{|y|:(0,y,z)M}.y(M):=\inf\{|y|:(0,y,z)\in M\}.
Lemma 23.1.

Let MM\in\mathscr{R} and suppose c0c\geq 0. If y(M)Ic-y(M)\in I_{c}, then Hc|MH_{c}|M has a critical point in M{x=0}{y<0}M\cap\{x=0\}\cap\{y<0\}.

Proof.

Let

Γ=M{x=0}{(cπ/2)<y<0}.\Gamma=M\cap\{x=0\}\cap\{(c-\pi/2)<y<0\}.

Note that Γ\Gamma is a compact curve with both endpoints on the line {(0,cπ/2)}×𝐑\{(0,c-\pi/2)\}\times\mathbf{R}. Now H|ΓH|\Gamma is a continuous map to (,](-\infty,\infty], and H=H=\infty on each of the endpoints of Γ\Gamma. Thus H|ΓH|\Gamma has an interior minimum. By symmetry, that point is a critical point of H|MH|M. ∎

Corollary 23.2.

Suppose MM\in\mathscr{R} and MWcM\cap W_{c} is nonempty. Then

(58) χ(MWc)+𝖭(Hc|M{x=0})1,\chi(M\cap W_{c})+\mathsf{N}(H_{c}|M\cap\{x=0\})\geq 1,

and thus (by (57))

(59) 1+2𝖭(Hc|M{x>0})|S||T|.1+2\mathsf{N}(H_{c}|M\cap\{x>0\})\leq|S|-|T|.
Proof.

By symmetry, we can assume that c0c\geq 0. If y(M)Ic-y(M)\notin I_{c}, then MWcM\cap W_{c} does not contain a curve that winds around ZZ, so MWcM\cap W_{c} is a union of disks. Thus (in this case) χ(MWc)1\chi(M\cap W_{c})\geq 1, so the inequality (58) holds. If y(M)Ic-y(M)\in I_{c}, then (by Lemma 23.1) 𝖭(Hc|M{x=0})1\mathsf{N}(H_{c}|M\cap\{x=0\})\geq 1, so the inequality (58) holds. ∎

In the rest of this section, MM will be in 𝒜\mathscr{A} and MnM_{n}\in\mathscr{R} will be a sequence such that suitable vertical translates of the MnM_{n} converge to MM. We write ana_{n}, bnb_{n}, AnA_{n}, and BnB_{n} for a(Mn)a(M_{n}), b(Mn)b(M_{n}), A(Mn)A(M_{n}), and B(Mn)B(M_{n}). We let LnL_{n} be the edge

[An,An]×{Bn}×{0}[-A_{n},A_{n}]\times\{B_{n}\}\times\{0\}

and n\ell_{n} be the edge

[an,an]×{bn}×{0}.[-a_{n},a_{n}]\times\{b_{n}\}\times\{0\}.
Proposition 23.3.

Suppose M𝒜M\in\mathscr{A}. If BIc¯B\in\overline{I_{c}}, then

𝖭(M{x>0})=0.\mathsf{N}(M\cap\{x>0\})=0.

Note that this proposition is true for capped and uncapped annuloids.

Proof.

The condition BIc¯B\in\overline{I_{c}} is equivalent to |cB|π/2|c-B|\leq\pi/2. By semicontinuity, it suffices to prove it for a dense set of cc with |cB|<π/2|c-B|<\pi/2. Thus we can assume that BIcB\in I_{c} and that cc is not equal to bb or to BB. By (59),

(60) 1+2𝖭(Hc|Mn{x>0})|Sn||Tn|.1+2\mathsf{N}(H_{c}|M_{n}\cap\{x>0\})\leq|S_{n}|-|T_{n}|.

Claim: |Sn||Tn|2|S_{n}|-|T_{n}|\leq 2 for all sufficiently large nn.

Assuming the claim, we see that

2𝖭(Hc|Mn{x>0})12\mathsf{N}(H_{c}|M_{n}\cap\{x>0\})\leq 1

for large nn. Thus for such nn,

𝖭(Hc|Mn{x>0})=0.\mathsf{N}(H_{c}|M_{n}\cap\{x>0\})=0.

Now letting nn\to\infty gives 𝖭(M{x>0}=0\mathsf{N}(M\cap\{x>0\}=0. Thus it remain only to prove the claim.

Proof of Claim.

We divide the proof of the claim into the three cases c>Bc>B, b<c<Bb<c<B, and c<bc<b.

Case 1: c>Bc>B. Then c>Bnc>B_{n} for large nn, and thus SnS_{n} consists of LnL_{n} and (if bnIcb_{n}\in I_{c}) also of n\ell_{n}. Thus |Sn||Sn||Tn|2|S_{n}|\leq|S_{n}|-|T_{n}|\leq 2.

Case 2: b<c<Bb<c<B. Then bn<c<Bnb_{n}<c<B_{n} for large nn. In this case, the local minima of Hc|MnH_{c}|\partial M_{n} are the two points (±An,c,0)(\pm A_{n},c,0) and also (if bnIcb_{n}\in I_{c}) the “point” n\ell_{n}. So |Sn|3|S_{n}|\leq 3.

Now LnL_{n} is a local maximum of Hc|MnH_{c}|\partial M_{n}. But for large nn, LnL_{n} is not a local maximum of Hc|MnH_{c}|M_{n}.

(To see that, note that Tan(Mn,(An,0,0))\operatorname{Tan}(M_{n},(A_{n},0,0)) is nearly vertical. It follows (for large nn) that HcH_{c} restricted to the curve Mn{x=0}M_{n}\cap\{x=0\} has a strict local minimum at the point (An,0,0)(A_{n},0,0).

Thus |Tn|1|T_{n}|\geq 1, so

|Sn||Tn|312.|S_{n}|-|T_{n}|\leq 3-1\leq 2.

Case 3: 0c<b0\leq c<b. Then 0<c<bn0<c<b_{n} for large nn.

In this case, Hc|MnH_{c}|\partial M_{n} has exactly four local minima, namely (±an,c,0)(\pm a_{n},c,0) and (±An,c,0)(\pm A_{n},c,0). Thus |Sn|4|S_{n}|\leq 4.

Also, Hc|MnH_{c}|\partial M_{n} has two local maxima, n\ell_{n} and LnL_{n}. As in the case 2, when nn is large, those points are not local maxima of Hc|MnH_{c}|M_{n}. Thus |Tn|2|T_{n}|\geq 2. Thus (in case 3)

|Sn||Tn|42=2.|S_{n}|-|T_{n}|\leq 4-2=2.

This completes the proof of the claim, and therefore the proof of Proposition 23.3. ∎

Theorem 23.4.

Suppose M𝒜M\in\mathscr{A}. Suppose also that

  1. (1)

    B=π/2B=\pi/2, or

  2. (2)

    B>π/2B>\pi/2 and MM is uncapped.

Then 𝖭(Hc|M{x>0})=0\mathsf{N}(H_{c}|M\cap\{x>0\})=0.

Proof.

By Proposition 23.3, it suffices to prove it when Ic¯\overline{I_{c}} is contained in (B,B)(-B,B). In particular, this means we are in Case (2) of the theorem: B>π/2B>\pi/2 and MM is uncapped.

Let α>x(M)+π\alpha>x(M)+\pi. Then α>x(Mn)+π\alpha>x(M_{n})+\pi for all large nn. We may suppose that α>x(Mn)+π\alpha>x(M_{n})+\pi for all nn.

Then Mupper{xα}M^{\textnormal{upper}}\cap\{x\geq\alpha\} is the graph of a function

u:[α,)×(B,B)𝐑.u:[\alpha,\infty)\times(-B,B)\to\mathbf{R}.

Likewise, Mnupper{xα}M^{\textnormal{upper}}_{n}\cap\{x\geq\alpha\} is the graph of

un:[α,An]×(Bn,Bn)𝐑.u_{n}:[\alpha,A_{n}]\times(-B_{n},B_{n})\to\mathbf{R}.

Recall that

yxu(x,y)y\mapsto\frac{\partial}{\partial x}u(x,y)

converges to s(B)s(B) uniformly on compact subsets of (B,B)(-B,B).

Thus there is a Λ\Lambda such that

xu>s(B)/2>0\frac{\partial}{\partial x}u>s(B)/2>0

on

Ic¯×[Λ,).\overline{I_{c}}\times[\Lambda,\infty).

Now let λ>Λ\lambda>\Lambda. Then for all sufficiently large nn,

(61) xun>0\frac{\partial}{\partial x}u_{n}>0 on {λ}×Ic¯\{\lambda\}\times\overline{I_{c}}.

Let

En+:=Mnupper{x>λ}E_{n}^{+}:=M^{\textnormal{upper}}_{n}\cap\{x>\lambda\}

and let EnE_{n}^{-} be the image of EnE_{n} under reflection in the plane {x=0}\{x=0\}. Let

M~n:=Mn(En+En),\tilde{M}_{n}:=M_{n}\setminus(E_{n}^{+}\cap E_{n}^{-}),

Now we wish to apply the formula (see Theorem 4.3)

χ(M~nWc)+𝖭(Hc|M~n)\displaystyle\chi(\tilde{M}_{n}\cap W_{c})+\mathsf{N}(H_{c}|\tilde{M}_{n}) |Sn|\displaystyle\leq|S_{n}|

(where SnS_{n} is the set of local minima of Hc|MnH_{c}|\partial M_{n} that are also local minima of Hc|MnH_{c}|M_{n}), which we can rewrite as

χ(M~nWc)+𝖭(M~n{x=0})+2𝖭(M~n{x>0})|Sn|.\chi(\tilde{M}_{n}\cap W_{c})+\mathsf{N}(\tilde{M}_{n}\cap\{x=0\})+2\mathsf{N}(\tilde{M}_{n}\cap\{x>0\})\leq|S_{n}|.

By Corollary 23.2, the sum of the first two terms on the left is at least 11. Thus

(62) 1+2𝖭(M~n{x>0})|Sn|.1+2\mathsf{N}(\tilde{M}_{n}\cap\{x>0\})\leq|S_{n}|.

The boundary of M~n\tilde{M}_{n} contains two curved portions, namely

Γn+:=Mnupper{x=λ}\Gamma_{n}^{+}:=M^{\textnormal{upper}}_{n}\cap\{x=\lambda\}

and its mirror image Γn\Gamma_{n}^{-} under reflection it the plane {x=0}\{x=0\}. These portions Γn±\Gamma_{n}^{\pm} are the portions of M~n\partial\tilde{M}_{n} in {z>0}\{z>0\}. Because

xun>0\frac{\partial}{\partial x}u_{n}>0

on In¯×{λ}\overline{I_{n}}\times\{\lambda\}, we see that Hc|M~nH_{c}|\tilde{M}_{n} has no local minima on Γn+\Gamma_{n}^{+}. By symmetry, there are none on Γn\Gamma_{n}^{-}. Thus

(63) |Sn{z>0}|=0.|S_{n}\cap\{z>0\}|=0.

If c<bnc<b_{n}, then Sn{z=0}S_{n}\cap\{z=0\} has just the 22 points (±an,c,0)(\pm a_{n},c,0). If bncb_{n}\leq c, then Sn{z=0}S_{n}\cap\{z=0\} has either no points (if bnIcb_{n}\notin I_{c}) or one point (namely n\ell_{n}) if bnIcb_{n}\in I_{c}. Thus

(64) |Sn{z=0}|2.|S_{n}\cap\{z=0\}|\leq 2.

Thus in either case (c<bnc<b_{n} or cbnc\geq b_{n}), we have (64), so by (63), we see that |Sn|2|S_{n}|\leq 2.

Thus by (62),

1+2𝖭(Hc|M~n{x>0})2.1+2\mathsf{N}(H_{c}|\tilde{M}_{n}\cap\{x>0\})\leq 2.

which implies that

𝖭(Hc|M~n{x>0})=0\mathsf{N}(H_{c}|\tilde{M}_{n}\cap\{x>0\})=0

for all sufficiently large nn. Thus

𝖭(Hc|M{0<x<λ})\displaystyle\mathsf{N}(H_{c}|M\cap\{0<x<\lambda\}) lim inf𝖭(Hc|Mn{0<x<λ}\displaystyle\leq\liminf\mathsf{N}(H_{c}|M_{n}\cap\{0<x<\lambda\}
=lim inf𝖭(Hc|M~n{0<x})\displaystyle=\liminf\mathsf{N}(H_{c}|\tilde{M}_{n}\cap\{0<x\})
=0.\displaystyle=0.

This holds for each λΛ\lambda\geq\Lambda. Hence 𝖭(Hc|M)=0\mathsf{N}(H_{c}|M)=0. ∎

Theorem 23.5.

Suppose M𝒜M\in\mathscr{A}.

  1. (1)

    if b=π/2b=\pi/2, or

  2. (2)

    if b>π/2b>\pi/2 and MM is uncapped,

then M{x>0}M\cap\{x>0\} is the graph of a function x=x(y,z)x=x(y,z) over a domain in the yzyz-plane.

Proof.

We claim that 𝐞1ν\mathbf{e}_{1}\cdot\nu never vanishes in M{x>0}M\cap\{x>0\}. For suppose to the contrary that 𝐞1ν=0\mathbf{e}_{1}\cdot\nu=0 at point pp with x(p)>0x(p)>0. If Tan(M,p)\operatorname{Tan}(M,p) were vertical, then 𝐯:=ν(M,p)\mathbf{v}:=\nu(M,p) would be ±𝐞2\pm\mathbf{e}_{2}, and then (x(p),y(p),z(p))(-x(p),y(p),z(p)) would be a second point with ν=𝐯\nu=\mathbf{v}, contrary to Corollary 6.7. Thus Tan(M,p)\operatorname{Tan}(M,p) is not vertical, so there is a strip 𝐑×(cπ/2,c+π/2)\mathbf{R}\times(c-\pi/2,c+\pi/2) and a grim reaper surface over that strip that is tangent to MM at pp. But then 𝖭(Hc|M{x>0})1\mathsf{N}(H_{c}|M\cap\{x>0\})\geq 1, contrary to Theorem 23.4.

Since M{x>0}M\cap\{x>0\} is connected and since ν(x(M),0,0)𝐞1=1<0\nu(x(M),0,0)\cdot\mathbf{e}_{1}=-1<0, it follows that ν𝐞1<0\nu\cdot\mathbf{e}_{1}<0 at all points of M{x>0}M\cap\{x>0\}.

Since MM is connected and properly embedded in 𝐑3\mathbf{R}^{3}, MM divides 𝐑3\mathbf{R}^{3} into two components. Let KK be the component such that ν\nu is the unit normal to MM that points out of KK. Let LL be a line parallel to the xx-axis. Noe LL cannot M{x>0}M\cap\{x>0\} in more than one point, since if it did, the sign of ν𝐞1\nu\cdot\mathbf{e}_{1} would alternate from one point to the next.

Thus LL intersects M{x>0}M\cap\{x>0\} in at most one point, and the intersection is transverse. ∎

Theorem 23.6.

Suppose that MM is a prong. Then MM is the graph of a function x=x(y,z)x=x(y,z) over an open subset of the yzyz-plane.

Proof.

Recall that MM is a limit of Mn:=Mn(x(Mn),0,0)M_{n}^{\prime}:=M_{n}-(x(M_{n}),0,0) where Mn𝒜(b)M_{n}\in\mathscr{A}(b) and x(Mn)x(M_{n})\to\infty. Since Bnb+πB_{n}\to b+\pi, we see that the MnM_{n} are uncapped for large nn. For such nn, ν𝐞1<0\nu\cdot\mathbf{e}_{1}<0 everywhere in Mn{x>x(Mn)}M_{n}^{\prime}\cap\{x>-x(M_{n})\}. Thus ν𝐞10\nu\cdot\mathbf{e}_{1}\leq 0 everywhere on MM. Now MM is connected, and ν𝐞1\nu\cdot\mathbf{e}_{1} is a Jacobi field, so if it vanished anywhere, it would vanish everywhere by the strong maximum principle. Since ν𝐞1=1\nu\cdot\mathbf{e}_{1}=-1 at the origin, we see that ν𝐞1<0\nu\cdot\mathbf{e}_{1}<0 everywhere on MM.

Let LL be a line parallel to the xx-axis. If LL intersected MM in more than one point, then for large nn, it would intersect Mn{x>0}M_{n}\cap\{x>0\} in more than one point, contrary to Theorem 23.5. ∎

Appendix A Some Useful Barriers

Theorem A.1.

For every a>0a>0 and b>0b>0, there is a translator

u=ua,b:[0,a]×[b,b][0,]u=u_{a,b}:[0,a]\times[-b,b]\to[0,\infty]

such that

u(0,)=u(,±b)=0,and\displaystyle u(0,\cdot)=u(\cdot,\pm b)=0,\quad\text{and}
u(a,)=.\displaystyle u(a,\cdot)=\infty.

Suppose MM is a translator in [0,a]×[b,b]×𝐑[0,a]\times[-b,b]\times\mathbf{R}.

  1. (i)

    If M\partial M lies in {zu(x,y)}\{z\leq u(x,y)\}, then MM lies in {zu(x,y)}\{z\leq u(x,y)\}.

  2. (ii)

    If M\partial M lies in {zu(x,y)}\{z\geq u(x,y)\}, then MM lies in {zu(x,y)}\{z\geq u(x,y)\}.

Furthermore, if MM is any translator in [0,a]×[b,b]×𝐑[0,a]\times[-b,b]\times\mathbf{R} with M=(graph(u))\partial M=\partial(\operatorname{graph}(u)), then M=graph(u)M=\operatorname{graph}(u).

Proof.

Existence of uu can be proved by solving the translator equation on [0,a]×[b,b][0,a]\times[-b,b] with boundary values nn on the side {a}×(b,b)\{a\}\times(-b,b) and 0 on the other three sides, and then letting nn\to\infty; see [Gama]*Theorem 9.

Refer to caption
Figure 9. The barrier uu.

Suppose that M\partial M is contained in {zu(x,y)}\{z\leq u(x,y)\}. We claim that

(65) If (xi,yi,zi)M(x_{i},y_{i},z_{i})\in M and ziz_{i}\to\infty, then xiax_{i}\to a.

This can be proved using rotationally symmetric translating annuli as barriers. Alternatively, let

x~=limζinfM{zζ}x().\tilde{x}=\lim_{\zeta\to\infty}\inf_{M\cap\{z\geq\zeta\}}x(\cdot).

Let (xi,yi,zi)M(x_{i},y_{i},z_{i})\in M with ziz_{i}\to\infty and xix~x_{i}\to\tilde{x}. Then, after passing to a subsequence, (xi,yi,0)(x_{i},y_{i},0) converges to a point (x^,y^,0)(\hat{x},\hat{y},0) and the sets M(0,0,zi)M-(0,0,z_{i}) converges (as sets) to a limit set MM^{\prime}. (In the language of [white16], M{x<a}M^{\prime}\cap\{x<a\} is a “(2,0)(2,0)-set” with respect to the translator metric.) Note that x()|Mx(\cdot)|M^{\prime} attains its minimum at the point (x^,y^,0)(\hat{x},\hat{y},0). If x^<a\hat{x}<a, then by the strong maximum principle in [white16], MM^{\prime} would contain the plane {x=x^}\{x=\hat{x}\}, which is impossible since MM^{\prime} is contained in [0,a]×[b,b]×𝐑[0,a]\times[-b,b]\times\mathbf{R}. This completes the proof of (65).

Suppose, contrary to (i), that MM contains points with z>u(x,y)z>u(x,y). By (65), there would be a largest t>0t>0 such that M+(t,0,t)M+(t,0,-t) intersects graph(f)\operatorname{graph}(f). At the point of contact, the maximum principle would be violated. Thus MM lies in {zu(x,y)}\{z\leq u(x,y)\}. This completes the proof of (i).

Now suppose, contrary to (ii), that M\partial M is contained in {zu(x,y)}\{z\geq u(x,y)\} but that MM contains points with z<u(x,y)z<u(x,y). Then there would be a largest t>0t>0 such that M+(t,0,t)M+(-t,0,t) intersects graph(u)\operatorname{graph}(u). At the point of contact, the maximum principle would be violated. Thus MM lies in {zu(x,y)}\{z\geq u(x,y)\}.

The last statement of the theorem follows immediately from (ii) and (i). ∎

Corollary A.2.

For each (x,y)[0,a)×[b,b](x,y)\in[0,a)\times[-b,b],

ua,b(x,y)=ua,b(x,y),u_{a,b}(x,y)=u_{a,b}(x,-y),

and u(x,y)u(x,y) is a decreasing function of |y||y|. In particular,

ua,b(x,y)ua,b(x,0)u_{a,b}(x,y)\leq u_{a,b}(x,0)

on [0,a)×[b,b][0,a)\times[-b,b].

Proof.

Let MM be the graph of ua,b(x,y)u_{a,b}(x,-y). By the last statement of Theorem A.1, M=graph(ua,b)M=\operatorname{graph}(u_{a,b}) and thus u(x,y)u(x,y)u(x,-y)\equiv u(x,y).

Let 0y1<y2b0\leq y_{1}<y_{2}\leq b and let y^=(y1+y2)/2\hat{y}=(y_{1}+y_{2})/2. Let MM be the image of graph(ua,b)|[0,a]×[y^,b]\operatorname{graph}(u_{a,b})|[0,a]\times[\hat{y},b] under reflection in the plane {y=y^}\{y=\hat{y}\}. Then M\partial M lies in {zua,b(x,y)}\{z\leq u_{a,b}(x,y)\}, so, by Theorem A.1, MM lies in {zua,b}\{z\leq u_{a,b}\}. Thus

ua,b(x,y2)ua,b(x,y1).u_{a,b}(x,y_{2})\leq u_{a,b}(x,y_{1}).

Hence ua,b(x,y)u_{a,b}(x,y) is a decreasing function of |y||y| for y[0,b]y\in[0,b]. ∎

Corollary A.3.

Let MM be a translator in a slab {|y|b}\{|y|\leq b\}. If infMx()>\inf_{M}x(\cdot)>-\infty, then

infMx()=infMx().\inf_{M}x(\cdot)=\inf_{\partial M}x(\cdot).

Likewise, if supMx()<\sup_{M}x(\cdot)<\infty, then supMx()=supMx()\sup_{M}x(\cdot)=\sup_{\partial M}x(\cdot).

In particular, if MM is a complete, nonempty translator in a slab {|y|b}\{|y|\leq b\}, then M{x=t}M\cap\{x=t\} is nonempty for every tt.

Proof.

Suppose the statement about infima is false. By translating, we can assume that

(66) 0<infMx()<a<infMx()0<\inf_{M}x(\cdot)<a<\inf_{\partial M}x(\cdot)

for some 0<a<0<a<\infty. By Theorem A.1, M{xa}M\cap\{x\leq a\} lies below the graph of ua,bu_{a,b}. The same is true for any vertical translate of MM. Thus M{x<a}M\cap\{x<a\} is empty, contrary to (66).

The statement about suprema follows by reflection. ∎

Theorem A.4.

Let

sb=infa>0xua,b(0,0).s_{b}=\inf_{a>0}\frac{\partial}{\partial x}u_{a,b}(0,0).

Then sb<s_{b}<\infty.

Suppose MM is a translator in {|y|b}\{|y|\leq b\}, and let

ψM(t):=supM{x=t}z(),\psi_{M}(t):=\sup_{M\cap\{x=t\}}z(\cdot),

If M{x=x0}\partial M\subset\{x=x_{0}\}, then

(*) ψM(x0+h)ψ(x0)+sb|h|\psi_{M}(x_{0}+h)\leq\psi(x_{0})+s_{b}|h|

for all hh. In particular, if MM has no boundary, then (*A.4) holds for all x0x_{0} and hh.

Proof.

Note that sa,b:=xua,b(0,0)<s_{a,b}:=\frac{\partial}{\partial x}u_{a,b}(0,0)<\infty by the strong maximum principle. Thus sb<s_{b}<\infty.

By symmetry, it suffices to prove (*A.4) for h>0h>0. First we claim that

(67) ψ(x+h)ψ(x)+ua,b(h,0)if xx0 and a>h0.\psi(x^{\prime}+h)\leq\psi(x^{\prime})+u_{a,b}(h,0)\quad\text{if $x^{\prime}\geq x_{0}$ and $a>h\geq 0$}.

We may assume that ψ(x)<\psi(x^{\prime})<\infty, as otherwise (67) is trivially true. Let ψM(x)<ζ<\psi_{M}(x^{\prime})<\zeta<\infty, and let M=(M(x,0,ζ)){0xa}M^{\prime}=(M-(x^{\prime},0,\zeta))\cap\{0\leq x\leq a\}. Then M\partial M^{\prime} lies in {zua,b(x,y)}\{z\leq u_{a,b}(x,y)\}, so MM^{\prime} lies in {zua,b(x,y)}\{z\leq u_{a,b}(x,y)\} by Theorem A.1. Hence

ψM(h)ua,b(h,0),\psi_{M^{\prime}}(h)\leq u_{a,b}(h,0),

which is equivalent (by translation) to

ψM(x+h)ζ+ua,b(h,0).\psi_{M}(x^{\prime}+h)\leq\zeta+u_{a,b}(h,0).

This holds for all ζ>ψM(x)\zeta>\psi_{M}(x^{\prime}), and therefore it holds for ζ=ψM(x)\zeta=\psi_{M}(x^{\prime}). Thus we have proved (67).

Now let a>0a>0 and let nn be a positive integer such that h/n<ah/n<a. Then

ψM(x0+kh/n)ψM(x0+(k1)h/n)+ua,b(h/n,0)\psi_{M}(x_{0}+kh/n)\leq\psi_{M}(x_{0}+(k-1)h/n)+u_{a,b}(h/n,0)

for all integers kk, and, in particular, for k=1,2,,nk=1,2,\dots,n. Thus

ψM(x0+h)\displaystyle\psi_{M}(x_{0}+h) ψM(x0)+nua,b(h/n,0)\displaystyle\leq\psi_{M}(x_{0})+nu_{a,b}(h/n,0)
=ψM(x0)+h(n/h)ua,b(h/n,0).\displaystyle=\psi_{M}(x_{0})+h(n/h)u_{a,b}(h/n,0).

Letting nn\to\infty gives

ψM(x0+h)ψM(x0)+hxua,b(0,0)=ψM(x0)+hsa,b.\psi_{M}(x_{0}+h)\leq\psi_{M}(x_{0})+h\,\frac{\partial}{\partial x}u_{a,b}(0,0)=\psi_{M}(x_{0})+h\,s_{a,b}.

Taking the infimum over b>0b>0 gives (*A.4). ∎

Appendix B Translators in half-slabs

In this section, we investigate the asymptotic behavior translators in a half-slab as |x||x|\rightarrow\infty. Some of the ideas of this section are inspired by [Chini]*Theorem 10.

Definition B.1.

If M𝐑3M\subset\mathbf{R}^{3}, we let Φ(M)\Phi(M) be the union of all subsequential limits of M+(0,0,z)M+(0,0,z) as zz\to\infty. Equivalently, Φ(M)\Phi(M) is the set of (x,y,z)(x,y,z) such that there exists (xi,yi,zi)M(x_{i},y_{i},z_{i})\in M with (xi,yi)(x,y)(x_{i},y_{i})\to(x,y) and ziz_{i}\to-\infty.

Note that if KK is a compact subset of 𝐑2\mathbf{R}^{2} with K×𝐑K\times\mathbf{R} disjoint from Φ(M)\Phi(M), then

infM(K×𝐑)z()>.\inf_{M\cap(K\times\mathbf{R})}z(\cdot)>-\infty.

Let bπ/2b\geq\pi/2, let I=(cb,c+b)I=(c-b,c+b) and let w:𝐑×I𝐑w:\mathbf{R}\times I\to\mathbf{R} be a complete translating graph. Define f:𝐑2𝐑f:\mathbf{R}^{2}\to\mathbf{R} by

f(x,y)={w(x,y)if yI,if yI,f(x,y)=\begin{cases}w(x,y)&\text{if $y\in I$},\\ -\infty&\text{if $y\notin I$},\end{cases}

and define F:𝐑2×(,](,]F:\mathbf{R}^{2}\times(-\infty,\infty]\to(-\infty,\infty] by

F(x,y,z)=zf(x,y).F(x,y,z)=z-f(x,y).

Note that ff and FF are continuous.

Theorem B.2.

Suppose M𝐑3M\subset\mathbf{R}^{3} is a translator such that

infMy()>,\inf_{M}y(\cdot)>-\infty,

and such that

(68) (M){|x|a}(\partial M)\cap\{|x|\geq a\} is contained in {yc}\{y\geq c\} for some a0a\geq 0.

Suppose also that

(69) Φ(M)\Phi(M) is contained in {y>c+b}\{y>c+b\}.

Then

(70) infMF=infMF.\inf_{M}F=\inf_{\partial M}F.
Lemma B.3.

Suppose CC is a closed set in 𝐑3\mathbf{R}^{3} such that Φ(C)\Phi(C) is disjoint from the the closed slab {yI¯}\{y\in\overline{I}\}. Suppose also that C{|y|<b}C\cap\{|y|<b\} is contained in K×𝐑K\times\mathbf{R} for some compact set KK, and that

infCF<λ<.\inf_{C}F<\lambda<\infty.

Then C{Fλ}C^{\prime}\cap\{F\leq\lambda\} is compact, and therefore F|CF|C attains its minimum: there is a point qCq\in C such that F(q)=infCFF(q)=\inf_{C}F.

Proof of Lemma B.3.

Note that {yI}={F<}\{y\in I\}=\{F<\infty\}, so

(71) CC{yI}K×𝐑.C^{\prime}\subset C\cap\{y\in I\}\subset K\times\mathbf{R}.

Note also that

(72) Φ(C)=\Phi(C^{\prime})=\emptyset

because Φ(C)\Phi(C) is contained in {yI¯}\{y\in\overline{I}\} and also in Φ(C)\Phi(C). Thus

ζ:=infCz()>.\zeta:=\inf_{C^{\prime}}z(\cdot)>-\infty.

by (71) and (72). Note that f|Kf|K is bounded above, so η:=supKf<\eta:=\sup_{K}f<\infty. For (x,y,z)C(x,y,z)\in C^{\prime}, we have (by definition of FF)

z\displaystyle z =F(x,y,z)+f(x,y)\displaystyle=F(x,y,z)+f(x,y)
λ+η.\displaystyle\leq\lambda+\eta.

Thus CC^{\prime} is contained in K×[ζ,λ+η]K\times[\zeta,\lambda+\eta], so CC^{\prime} is compact. ∎

Proof of Theorem B.2.

It suffices to prove the theorem for c=0c=0. Let β\beta be a negative number such that βinfMy()\beta\leq\inf_{M}y(\cdot).

By replacing MM by M{yb}M\cap\{y\leq b\}, we can assume that

(73) MM lies in the slab {βyb}\{\beta\leq y\leq b\}.

Note that (73) and (69) imply that M+(0,0,z)M+(0,0,z) converges to the empty set as zz\to\infty. Equivalently,

(74) infM(K×𝐑)z()>for every compact K𝐑2.\inf_{M\cap(K\times\mathbf{R})}z(\cdot)>-\infty\quad\text{for every compact $K\subset\mathbf{R}^{2}$.}

To prove the theorem, it suffices to prove that

(75) infMFinfMF,\inf_{M}F\geq\inf_{\partial M}F,

as the reverse inequality is trivially true.

By horizontal translation, we may assume that if uu is a Δ\Delta-wing, then it is centered at the origin: Du(0,0)=0Du(0,0)=0.

We may choose the aa in (68) so that a>ba>b.

Claim 5.

Let M=M{xa}M^{\prime}=M\cap\{x\geq a\}. Then

infMF=infMF.\inf_{M^{\prime}}F=\inf_{\partial M^{\prime}}F.

If uu is a grim reaper surface with ux0\frac{\partial u}{\partial x}\leq 0 or if uu is a Δ\Delta-wing (in which case ux<0\frac{\partial u}{\partial x}<0 on (0,)×(b,b)(0,\infty)\times(-b,b)), let LL be the vertical line through (0,b,0)(0,b,0). If uu is a grim reaper surface with ux>0\frac{\partial u}{\partial x}>0, let L=ZL=Z. Let R(θ)R(\theta) denote counterclockwise rotation about LL through angle θ\theta. We also let R(θ)R(\theta) denote the corresponding rotation in the xyxy-plane. For each p[a,)×[0,b]p\in[a,\infty)\times[0,b], note that

θ[0,π/2)f(R(θ)p)\theta\in[0,\pi/2)\mapsto f(R(\theta)p)

is a decreasing function. Thus, for each p=(x,y,z)[a,)×[0,b]×𝐑p=(x,y,z)\in[a,\infty)\times[0,b]\times\mathbf{R} (and thus, in particular, for each pMp\in\partial M^{\prime}),

θ[0,π/2)F(R(θ)p)=zf(R(θ)(x,y))\theta\in[0,\pi/2)\mapsto F(R(\theta)p)=z-f(R(\theta)(x,y))

is an increasing function. Consequently,

(76) infR(θ)MFinfMF.\inf_{R(\theta)\partial M^{\prime}}F\geq\inf_{\partial M^{\prime}}F.

For θ(0,π/2)\theta\in(0,\pi/2)), we claim that

(77) infR(θ)MF=infR(θ)MF.\inf_{R(\theta)M^{\prime}}F=\inf_{R(\theta)\partial M^{\prime}}F.

To prove (77), we may assume that the left side is <<\infty, as otherwise (77) is trivially true. Now

R(θ)M{|y|<b}T×𝐑,R(\theta)M^{\prime}\cap\{|y|<b\}\subset T\times\mathbf{R},

where T=TθT=T_{\theta} is the compact triangular region

R(θ)([0,)×[β,B]){yb}.R(\theta)([0,\infty)\times[\beta,B])\cap\{y\leq b\}.

By Lemma B.3, the infimum of FF on R(θ)MR(\theta)M^{\prime} is attained at a point qq. By the strong maximum principle, qq is in the boundary R(θ)M=R(θ)M\partial R(\theta)M^{\prime}=R(\theta)\partial M^{\prime}. Hence (77) holds.

Now let pMp\in M^{\prime}. By (77) and (76),

F(R(θ)p)\displaystyle F(R(\theta)p) infR(θ)MF\displaystyle\geq\inf_{R(\theta)M^{\prime}}F
=infR(θ)MF\displaystyle=\inf_{R(\theta)\partial M^{\prime}}F
infMF.\displaystyle\geq\inf_{\partial M^{\prime}}F.

Thus

F(R(θ)p)infMF.F(R(\theta)p)\geq\inf_{\partial M^{\prime}}F.

Letting θ0\theta\to 0 gives

F(p)infMF.F(p)\geq\inf_{\partial M^{\prime}}F.

Taking the infimum over pMp\in M^{\prime} gives (75). Thus we have proved Claim 5.

Using Claim 5, we now prove that

(78) infMF=infMF.\inf_{M}F=\inf_{\partial M}F.

We may assume that MM contains a point p0p_{0} with F(p0)<F(p_{0})<\infty, as otherwise (78) is trivially true.

Let

α:=infM{|x|a}F=infM(K×𝐑)F,\alpha:=\inf_{M\cap\{|x|\leq a\}}F=\inf_{M\cap(K\times\mathbf{R})}F,

where KK is the rectangle [a,a]×[b,b][-a,a]\times[-b,b]. Since p0M{|x|a}p_{0}\in M\cap\{|x|\leq a\},

αF(p0)<.\alpha\leq F(p_{0})<\infty.

By Lemma B.3, the infimum is attained: M{|x|a}M\cap\{|x|\leq a\} contains a point pp^{\prime} such that

F(p)=infM{|x|a}F.F(p^{\prime})=\inf_{M\cap\{|x|\leq a\}}F.

Suppose first that

(79) infMF=infM{|x|a}F.\inf_{M}F=\inf_{M\cap\{|x|\leq a\}}F.

Then F|MF|M attains its minimum at pp^{\prime}, and therefore pMp^{\prime}\in\partial M by the strong maximum principle. Thus we have proved (78) in case (79) holds.

Now suppose that (79) does not hold. Then

(80) infMF<infM{|x|a}F=α.\inf_{M}F<\inf_{M\cap\{|x|\leq a\}}F=\alpha.

It follows that infMF\inf_{M}F is equal to one or both of infM{xa}F\inf_{M\cap\{x\geq a\}}F and infM{xa}F\inf_{M\cap\{x\leq-a\}}F. By symmetry, it is enough to consider the case

infMF=infMF.\inf_{M}F=\inf_{M^{\prime}}F.

where M=M{xa}M^{\prime}=M\cap\{x\geq a\}. By Claim 5,

infMF\displaystyle\inf_{M^{\prime}}F =infMF\displaystyle=\inf_{\partial M^{\prime}}F
min{infM{x=a}F,inf(M){x>a}F}\displaystyle\geq\min\{\inf_{M\cap\{x=a\}}F,\inf_{(\partial M)\cap\{x>a\}}F\}
min{α,infMF}.\displaystyle\geq\min\{\alpha,\inf_{\partial M}F\}.

Thus by (80),

α>infMFmin{α,infMF},\alpha>\inf_{M}F\geq\min\{\alpha,\inf_{\partial M}F\},

which immediately implies that

infMFinfMF.\inf_{M}F\geq\inf_{\partial M}F.

The reverse inequality is trivially true. ∎

Corollary B.4.

In Theorem B.2,

  1. (1)

    If MM is complete or, more, generally, if (M){yI}(\partial M)\cap\{y\in I\} is empty, then M{yI}M\cap\{y\in I\} is empty.

  2. (2)

    If supMF<\sup_{M}F<\infty and if M{x=0}\partial M\subset\{x=0\}, then F|MF|M attains its minimum at a point q(M){yI}q\in(\partial M)\cap\{y\in I\}, and thus

    minMF=F(q)>.\min_{M}F=F(q)>-\infty.
  3. (3)

    If M\partial M is contained in {x=0}\{x=0\}, and if JJ is a closed interval in II, then there is γ=γJ\gamma=\gamma_{J} such that

    zs(b)x+γfor (x,y,z)M{yJ}.z\geq s(b)x+\gamma\quad\text{for $(x,y,z)\in M\cap\{y\in J\}$.}
Proof.

If (M){yI}(\partial M)\cap\{y\in I\} is empty, then infMF=\inf_{\partial M}F=\infty, so infMF=\inf_{M}F=\infty (by Theorem B.2), and thus M{yI}M\cap\{y\in I\} is empty.

Now suppose that M\partial M is contained in {x=0}\{x=0\} and that infMF<\inf_{M}F<\infty. Then

\displaystyle\infty >infMF\displaystyle>\inf_{M}F
=infMF.\displaystyle=\inf_{\partial M}F.

Now

(M){yI}K×𝐑,(\partial M)\cap\{y\in I\}\subset K\times\mathbf{R},

where KK is the compact set {0}×I¯\{0\}\times\overline{I}. By Lemma B.3, the infimum is attained at some point qq in M\partial M. Since F(q)<F(q)<\infty, the point qq lies in the slab {yI}\{y\in I\}.

To prove Assertion (3), let w:𝐑×I𝐑w:\mathbf{R}\times I\to\mathbf{R} be the grim reaper surface with wxs(b)\frac{\partial w}{\partial x}\equiv s(b). We may assume that infMF<\inf_{M}F<\infty, as otherwise the assertion is trivially true. Let qq be as in Assertion (2). Then

<F(q)F(x,y,z)=zf(x,y),-\infty<F(q)\leq F(x,y,z)=z-f(x,y),

so

z\displaystyle z F(q)+f(x,y)\displaystyle\geq F(q)+f(x,y)
=F(q)+f(0,y)+s(b)x\displaystyle=F(q)+f(0,y)+s(b)x
F(q)+minyJf(0,y)+s(b)x.\displaystyle\geq F(q)+\min_{y\in J}f(0,y)+s(b)x.

Corollary B.5.

Suppose that MM is a translator such that

infMy()>\inf_{M}y(\cdot)>-\infty

and that Φ(M)\Phi(M) is contained in the halfspace {yd}\{y\geq d\}. If MM is complete, or, more generally, if M\partial M is contained in {yd}\{y\geq d\}, then MM is contained in {yd}\{y\geq d\}.

Proof.

Let II be any open interval of width π\geq\pi whose closure is in (,d)(-\infty,d). Then by Theorem B.2 (applied to any translating graph w:𝐑×I𝐑w:\mathbf{R}\times I\to\mathbf{R}), MM is disjoint from the slab {yI}\{y\in I\}. ∎

Corollary B.6.

Suppose that MM is a translator in {x0}\{x\geq 0\} with M\partial M in {x=0}\{x=0\}. Suppose also that

infMy()>\inf_{M}y(\cdot)>-\infty

and that Φ(M)\Phi(M) is contained in the halfspace {yd}\{y\geq d\}.

  1. (1)

    If (xi,yi,zi)M(x_{i},y_{i},z_{i})\in M with xix_{i}\to\infty and yiy^<dy_{i}\to\hat{y}<d, then zi/xiz_{i}/x_{i}\to\infty.

  2. (2)

    If MM is contained in the region {zmx+c}\{z\leq mx+c\} for some mm and cc, and if d<dd^{\prime}<d, then M{yd}M\cap\{y\leq d^{\prime}\} is compact.

Proof.

To prove Assertion (1), note that for every bπ/2b\geq\pi/2, there is an open interval II of width 2b\geq 2b such that y^I\hat{y}\in I and such that I¯(,d)\overline{I}\subset(-\infty,d). By Assertion (3) of Corollary B.4,

lim supizi/xis(b).\limsup_{i\to\infty}z_{i}/x_{i}\geq s(b).

Now let bb\to\infty.

To prove Assertion (2), let pi=(xi,yi,zi)p_{i}=(x_{i},y_{i},z_{i}) be a sequence of points in M{yd}M\cap\{y\leq d^{\prime}\}. Note that yiy_{i} is a bounded sequence (since infMy()>\inf_{M}y(\cdot)>-\infty). If xix_{i}\to\infty, then zi/xiz_{i}/x_{i}\to\infty by Assertion (1), which is impossible since zimxi+cz_{i}\leq mx_{i}+c. Thus xix_{i} is bounded. Since zimxi+cz_{i}\leq mx_{i}+c, ziz_{i} is bounded above. After passing to a subsequence, (xi,yi)(x_{i},y_{i}) converges to point (x^,y^)(\hat{x},\hat{y}) with y^<d\hat{y}<d. Since y()dy(\cdot)\geq d on Φ(M)\Phi(M), we see that ziz_{i} is bounded below. We have shown that every sequence in M{yd}M\cap\{y\leq d^{\prime}\} is bounded. Thus M{yd}M\cap\{y\leq d^{\prime}\} is compact. ∎

Theorem B.7.

Suppose MM is a translator in {x0}\{x\geq 0\} such that

  1. (1)

    M\partial M is contained in {x=0}\{x=0\}.

  2. (2)

    z()z(\cdot) is bounded above on M\partial M.

  3. (3)

    supM|y()|<\sup_{M}|y(\cdot)|<\infty.

  4. (4)

    Every subsequential limit as zz\to\infty of M+(0,0,z)M+(0,0,z) is contained in the slab {|y|b}\{|y|\leq b\}.

Then

  1. (i)

    If β>b\beta>b, then M{|y|β}M\cap\{|y|\geq\beta\} is compact.

  2. (ii)

    If b<π/2b<\pi/2, then MM is contained in the plane {x=0}\{x=0\}.

Proof.

Let β>b\beta>b. By Theorem A.4, MM is contained in a region of the form {zmx+c}\{z\leq mx+c\}. Thus, M{yβ}M\cap\{y\leq-\beta\} is compact by Corollary B.6. Likewise, M{yβ}M\cap\{y\geq\beta\} is compact. Thus M{|y|β}M\cap\{|y|\geq\beta\} is compact. This completes the proof of Asssertion (i).

Now suppose that b<π/2b<\pi/2 and let b<β<β<π/2b<\beta<\beta^{\prime}<\pi/2. Since M{|y|β}M\cap\{|y|\geq\beta\} is compact, there is an a[0,)a\in[0,\infty) such that

(81) M{xa}{|y|<β}.M\cap\{x\geq a\}\subset\{|y|<\beta\}.

By [scherk], there is an a>aa^{\prime}>a and a translator

u:[a,a]×[β,β]𝐑u:[a,a^{\prime}]\times[-\beta^{\prime},\beta^{\prime}]\to\mathbf{R}

such that u(a,)=u(a,)=u(a,\cdot)=u(a^{\prime},\cdot)=\infty and u(,β)=u(,β)=u(\cdot,-\beta^{\prime})=u(\cdot,\beta^{\prime})=-\infty.

If M{a<x<a}M\cap\{a<x<a^{\prime}\} were nonempty, then zu(x,y)z-u(x,y) would attain a maximum on MM, violating the strong maximum principle.

Thus M{xa}{xa}M\subset\{x\leq a\}\cup\{x\geq a^{\prime}\}.

By Corollary A.3, MM is contained in {x=0}\{x=0\}. ∎

Corollary B.8.

Suppose MM is a complete translator in 𝐑3\mathbf{R}^{3} such that supM|y()|<\sup_{M}|y(\cdot)|<\infty and such that M+(0,0,z)M+(0,0,z) converges as zz\to\infty to planes {y=bi}\{y=b_{i}\}, with b1b2bkb_{1}\leq b_{2}\leq\dots\leq b_{k}. Then

b1\displaystyle b_{1} =infMy(),\displaystyle=\inf_{M}y(\cdot),
bk\displaystyle b_{k} =supMy().\displaystyle=\sup_{M}y(\cdot).
Proof.

It suffices to prove the equation for bkb_{k}. Trivially bksupMy()b_{k}\leq\sup_{M}y(\cdot). The reverse inequality holds by Corollary B.5. Hence bk=supMy()b_{k}=\sup_{M}y(\cdot). ∎

Theorem B.9.

Suppose MM is a finite-type translator in {x0}\{x\geq 0\} such that

  1. (1)

    M{x=0}\partial M\subset\{x=0\}.

  2. (2)

    B:=supMy()<B:=\sup_{M}y(\cdot)<\infty.

  3. (3)

    MM is invariant under reflection in the plane {y=0}\{y=0\}.

  4. (4)

    M+(0,0,z)M+(0,0,z) converges smoothly, as zz\to\infty, to the halfplanes {y=±b}{x0}\{y=\pm b\}\cap\{x\geq 0\}.

  5. (5)

    M{y=0}M\cap\{y=0\} is graph of a function z=ϕ(x)z=\phi(x).

Then the limit L=limxϕ(x)L=\lim_{x\to\infty}\phi^{\prime}(x) exists, and L=±s(b)L=\pm s(b).

Proof.

Theorem 5.9, the limit LL exists, and M(x,0,ϕ(x))M-(x,0,\phi(x)) converges (as xx\to\infty) to a grim reaper surface GG containing the line {(x,0,Lx)}\{(x,0,Lx)\}. By Theorem B.7(i), GG is contained in the slab {|y|b}\{|y|\leq b\}. Thus

(82) |L|s(b).|L|\leq s(b).

If b=π/2b=\pi/2, then s(b)=0s(b)=0, so L=0L=0 by (82). Thus we may assume that b>π/2b>\pi/2.

We prove the theorem by assuming that Ls(b)L\neq-s(b) and then proving that L=s(b)L=s(b). Since Ls(b)L\neq-s(b),

L>s(b)L>-s(b)

by (82). By choosing β(π/2,b)\beta\in(\pi/2,b) close to bb, we can ensure that

(83) L>s(β)>s(b).L>-s(\beta)>-s(b).

By translating MM vertically, we can assume that

(84) {0}×[β,β]×𝐑\{0\}\times[-\beta,\beta]\times\mathbf{R} is disjoint from M\partial M.

For a0a\geq 0, let ha:𝐑×(β,β)𝐑h_{a}:\mathbf{R}\times(-\beta,\beta)\to\mathbf{R} be the Δ\Delta-wing such that ha(0,0)=0h_{a}(0,0)=0 and Dha(a,0)=0Dh_{a}(a,0)=0. Let

Ha(x,y,z)={zha(x,y)if |y|<β, andif |y|β.H_{a}(x,y,z)=\begin{cases}z-h_{a}(x,y)&\text{if $|y|<\beta$, and}\\ \infty&\text{if $|y|\geq\beta$}.\end{cases}

Let M+:=M{y0}M^{+}:=M\cap\{y\geq 0\}.

Then the boundary M+\partial M^{+} of M+M^{+} consists of (M){y0}(\partial M)\cap\{y\geq 0\} together with {(x,0,ϕ(x))\{(x,0,\phi(x)), x0x\geq 0. Now

Ha(x,0,ϕ(x))\displaystyle H_{a}(x,0,\phi(x)) =ϕ(x)ha(x,0),\displaystyle=\phi(x)-h_{a}(x,0),

which tends to \infty as xx\to\infty by (83). (Note that ϕ(x)/xL\phi(x)/x\to L and that ha(x,0)/xs(β)h_{a}(x,0)/x\to-s(\beta) as xx\to\infty.)

Consequently, there is a t<t<\infty such that

infM+Ha=inf(M+){|x|t}Ha.\inf_{\partial M^{+}}H_{a}=\inf_{(\partial M^{+})\cap\{|x|\leq t\}}H_{a}.

By Lemma B.3 (applied to the infimum on the right), the infimum is attained: there is a qM+q\in\partial M^{+} such that

infM+Ha=Ha(q).\inf_{\partial M^{+}}H_{a}=H_{a}(q).

Thus by Theorem B.2,

minM+Ha=Ha(q).\min_{M^{+}}H_{a}=H_{a}(q).

By symmetry,

(85) minMHa=Ha(q).\min_{M}H_{a}=H_{a}(q).

By the strong maximum principle, qM=M{x=0}q\in\partial M=M\cap\{x=0\}. Thus

Ha(q)\displaystyle H_{a}(q) =z(q)ha(0,y(q))0,\displaystyle=z(q)-h_{a}(0,y(q))\geq 0,

since z(q)0z(q)\geq 0 by (84) and ha(0,y(q))ha(0,0)=0h_{a}(0,y(q))\leq h_{a}(0,0)=0.

Thus

Ha(p)0H_{a}(p)\geq 0

for all pMp\in M. In particular, for p=(x,0,ϕ(x))p=(x,0,\phi(x)),

ϕ(x)ha(x,0)0\phi(x)-h_{a}(x,0)\geq 0

As aa\to\infty, ha(x,0)s(β)xh_{a}(x,0)\to s(\beta)x. Thus

ϕ(x)s(β)x\phi(x)\geq s(\beta)x

for all xx, so Ls(β)L\geq s(\beta). Letting βb\beta\to b gives Ls(b)L\geq s(b) and thus L=s(b)L=s(b) by (82).

B.1. Some properties of annuloids

Recall that an annuloid is a complete, properly embedded translating annulus MM in 𝐑3\mathbf{R}^{3} with the following properties:

  1. (a)

    MM is contained in a slab {|y|<B}\{|y|<B\}, for some B>0B>0.

  2. (b)

    MM is symmetric under reflection in the planes {x=0}\{x=0\} and {y=0}\{y=0\}.

  3. (c)

    MM is disjoint from the axis ZZ.

  4. (d)

    As zz\to-\infty, MM is smoothly asymptotic to the planes y=±by=\pm b and y=±By=\pm B, where b=b(M)b=b(M) and B=B(M)B=B(M) and 0<bB<0<b\leq B<\infty.

  5. (e)

    M(0,0,z)M-(0,0,z) converges, as zz\to\infty, to the empty set.

The aim of this subsection is to prove the following:

Theorem B.10.

If MM is an annuloid of finite type, then b(M)π/2.b(M)\geq\pi/2.

Proof.

Since MM is symmetric about {y=0}\{y=0\}, M{y=0}M\cap\{y=0\} is a 11-manifold. Since MM is connected, M{y=0}M\cap\{y=0\} is nonempty. Let Γ\Gamma be a component of M{y=0}M\cap\{y=0\}. Since MZ=M\cap Z=\emptyset, we can assume (by symmetry) that Γ\Gamma is in {y=0,x>0}\{y=0,\,x>0\}. By Lemma 3.5, Γ\Gamma is not compact, because that would violate the maximum principle.

Since MM is of finite type, x()|Mx(\cdot)|M has finitely many critical points. Let aa be a number greater than every critical value of x()|Mx(\cdot)|M. Note that every critical point of x()|Γx(\cdot)|\Gamma is also a critical point of x()|Γx(\cdot)|\Gamma, so aa is greater than every critical value of x()|Γx(\cdot)|\Gamma.

By Assertion (5) of Theorem 5.9, x()x(\cdot)\to\infty on each end of Γ\Gamma. Thus for tat\geq a, Γ{x=t}\Gamma\cap\{x=t\} consists of two points (t,0,z1(t))(t,0,z_{1}(t)) and (t,0,z2(t))(t,0,z_{2}(t)), where z1(t)<z2(t)z_{1}(t)<z_{2}(t).

For tat\geq a, M{x=t}M\cap\{x=t\} is a smooth manifold with exactly four ends, corresponding to the points at infinity (t,±b,)(t,\pm b,-\infty) and (t,±B,)(t,\pm B,-\infty). (By a slight abuse of notation, we regard these as four distinct points even if b=Bb=B.) Reasoning as above, M{x=t}M\cap\{x=t\} has no closed curve components. Thus M{x=t}M\cap\{x=t\} has exactly two components. By symmetry and embeddedness, the component C1(t)C_{1}(t) containing (t,0,z1(t))(t,0,z_{1}(t)) has ends tending to (t,±b,)(t,\pm b,-\infty), and the component C2(t)C_{2}(t) containing (t,0,z2(t))(t,0,z_{2}(t)) has ends tending to (t,±B,)(t,\pm B,-\infty).

Let Σ=tx~C1(t)\Sigma=\cup_{t\geq\tilde{x}}C_{1}(t). Then Σ=C1(a)\partial\Sigma=C_{1}(a), a curve (in the plane {z=a}\{z=a\}) on which z()z(\cdot) is bounded above. Also, Σ+(0,0,z)\Sigma+(0,0,z) converges as zz\to\infty to the to the halfplanes {xa}{y=±b}\{x\geq a\}\cap\{y=\pm b\}. Thus bπ/2b\geq\pi/2 by Theorem B.7. ∎

Appendix C Convergence of Sets

Here we describe basic facts about convergence of sets in metric spaces. According to the wikipedia article on Kuratowski convergence, these notions were introduced in lectures by Painlevé in 1902 and popularized in books by Hausdorff and Kuratowski.

Let XX be a compact metric space. Let 𝐊(X)\mathbf{K}(X) be the set of nonempty closed subsets of XX. We define a metric dd on 𝐊(X)\mathbf{K}(X) by

d(K1,K2)\displaystyle d(K_{1},K_{2}) =dist(,K1),dist(,K2)C0\displaystyle=\|\operatorname{dist}(\cdot,K_{1}),\operatorname{dist}(\cdot,K_{2})\|_{C_{0}}
=maxpX|dist(p,K1)dist(p,K2)|.\displaystyle=\max_{p\in X}|\operatorname{dist}(p,K_{1})-\operatorname{dist}(p,K_{2})|.

Note that Ki𝐊(X)K_{i}\in\mathbf{K}(X) converges to K𝐊(X)K\in\mathbf{K}(X) if and only if the following holds: dist(p,Ki)0\operatorname{dist}(p,K_{i})\to 0 for each pKp\in K and lim infd(p,Ki)>0\liminf d(p,K_{i})>0 for each pKp\notin K.

Note that 𝐊(X)\mathbf{K}(X) is also compact, because if Ki𝐊(X)K_{i}\in\mathbf{K}(X), then by Arzela-Ascoli, the functions dist(,Ki)\operatorname{dist}(\cdot,K_{i}) will converge uniformly (after passing to a subsequence) to a limit ff. Let K={p:f(p)=0}K=\{p:f(p)=0\}. Then f=dist(,K)f=\operatorname{dist}(\cdot,K) and so KiKK_{i}\to K.

There are also useful notions of lim sup\limsup and lim inf\liminf in the space 𝐊(X)\mathbf{K}(X):

Definition C.1.

If KnK_{n} is a sequence in 𝐊(X)\mathbf{K}(X), we let

lim supnKn\displaystyle\limsup_{n}K_{n} :={pX:lim infnd(p,Kn)=0},\displaystyle:=\{p\in X:\liminf_{n}d(p,K_{n})=0\},
lim infnKn\displaystyle\liminf_{n}K_{n} :={pX:lim supnd(p,Kn)=0}.\displaystyle:=\{p\in X:\limsup_{n}d(p,K_{n})=0\}.

Thus plim supnKnp\in\limsup_{n}K_{n} if and only there are n(i)n(i)\to\infty and pn(i)Kn(i)p_{n(i)}\in K_{n(i)} such the pn(i)p_{n(i)} converge to pp. Note that lim infnKnlim supnKn\liminf_{n}K_{n}\subset\limsup_{n}K_{n}. If equality holds, then KnK_{n} converges, and, conversely, if KnK_{n} converges then

lim infnKn=lim supnKn=limnKn.\liminf_{n}K_{n}=\limsup_{n}K_{n}=\lim_{n}K_{n}.
Theorem C.2.

Suppose that KnK_{n} is a sequence of nonempty closed subsets of a compact metric space XX. Then, after passing to subsequence, the KnK_{n} converge to a closed set KK. Furthermore, if each KnK_{n} is connected, then KK is also connected.

Proof.

We already proved the first assertion. Assume that each KnK_{n} is connected. Let CC be a nonempty closed subset of KK such that KCK\setminus C is also closed. We must show that KCK\setminus C is empty. Let f:X𝐑f:X\to\mathbf{R} be a continuous function that is 0 on CC and >0>0 at each point of XCX\setminus C. (For example, f(p)=dist(p,C)f(p)=\operatorname{dist}(p,C)). Now f(Kn)f(K_{n}) is a compact, connected subset of 𝐑\mathbf{R}, so f(K)f(K) is also. Thus f(K)=[0,r]f(K)=[0,r] for some r0r\geq 0. Now f(KC)=[0,r]{0}f(K\setminus C)=[0,r]\setminus\{0\} is compact since KCK\setminus C is compact. Thus r=0r=0 and so KCK\setminus C is empty. ∎

We would like to apply this theory when the metric space XX is 𝐑3\mathbf{R}^{3}, but 𝐑3\mathbf{R}^{3} is not compact. We get around this by using the one-point compactification 𝐑3{}\mathbf{R}^{3}\cup\{\infty\}, with the metric that comes from stereographic projection. Let 𝐊(𝐑3)\mathbf{K}(\mathbf{R}^{3}) be the family of all closed subsets of 𝐑3\mathbf{R}^{3}, including the empty set. If K𝐊(𝐑3)K\in\mathbf{K}(\mathbf{R}^{3}), we let

K~:={Kif K is compact and nonempty,{}if K=,K{}if K is non-compact.\tilde{K}:=\begin{cases}K&\text{if $K$ is compact and nonempty},\\ \{\infty\}&\text{if $K=\emptyset$,}\\ K\cup\{\infty\}&\text{if $K$ is non-compact}.\end{cases}

We define a metric δ(,)\delta(\cdot,\cdot) on 𝐊(𝐑3)\mathbf{K}(\mathbf{R}^{3}) by setting letting

δ(K,K)=d(K~,K~),\delta(K,K^{\prime})=d(\tilde{K},\tilde{K}^{\prime}),

where on the right side, d(,)d(\cdot,\cdot) is the metric on 𝐊(𝐑3{})\mathbf{K}(\mathbf{R}^{3}\cup\{\infty\}). Then Ki𝐊(𝐑3)K_{i}\in\mathbf{K}(\mathbf{R}^{3}) converges to K𝐊(𝐑3)K\in\mathbf{K}(\mathbf{R}^{3}) if and only if the following holds: every pKp\in K is a limit of piKip_{i}\in K_{i}, and every qKq\notin K has a neighborhood UU such that UKiU\cap K_{i} is empty for all sufficiently large ii.

The following theorem is an immediate consequence of Theorem C.2:

Theorem C.3.

Let KnK_{n} be a sequence of closed subsets of 𝐑3\mathbf{R}^{3}. After passing to a subsequence, the KnK_{n} converge to a closed KK. Suppose each KnK_{n} is connected. If KK is compact, it is connected, and if KK is noncompact, then K{}K\cup\{\infty\} is connected.

Remark C.4.

Theorem C.3 remains true, with the same proof, in any locally compact space whose one-point compactification is metrizable.

Theorem C.5.

Let pqp\neq q be two points in a compact, connected metric space MM. Then there is a compact, connected set XX containing pp and qq such that X{p,q}X\setminus\{p,q\} is also connected.

Proof.

Let \mathscr{F} be the family of closed sets SS in MM with the following property: {p,q}S\{p,q\}\subset S, and if TT is a clopen subset of SS containing one of the two points pp and qq, then it also contains the other. Then \mathscr{F} is nonempty since MM\in\mathscr{F}.

Claim 6.

If XX\in\mathscr{F} and if X{p,q}X\setminus\{p,q\} is not connected, then there is an XX^{\prime}\in\mathscr{F} with XXX^{\prime}\subsetneq X.

To prove the claim, note that, by hypothesis, X{p,q}=YZX\setminus\{p,q\}=Y\cup Z, where YY and ZZ are nonempty, disjoint, sets that are relatively closed in X{p,q}X\setminus\{p,q\}. It suffices to show that at least one of the sets Y¯\overline{Y} and Z¯\overline{Z} belongs to \mathscr{F}, since then we can let XX^{\prime} be that set.

If every nonempty clopen subset of Y¯\overline{Y} contains both pp and qq, then Y¯\overline{Y} is a connected set containing pp and qq, so Y¯\overline{Y} is in \mathscr{F}, and we are done proving the claim. Now suppose that there is a nonempty clopen subset SS of Y¯\overline{Y} that does not contain both pp and qq. We may assume that SS does not contain qq. Let TT be a clopen subset of Z¯\overline{Z} that contains pp. Then Y¯T\overline{Y}\cup T is a clopen subset of XX containing pp, so it also contains qq. Since qq is not in Y¯\overline{Y}, qTq\in T. Hence Z¯\overline{Z} is in \mathscr{F}. This completes the proof of the claim.

By Zorn’s Lemma, there is a set XX in \mathscr{F} such that no proper subset of XX is in \mathscr{F}. By the claim, X{p,q}X\setminus\{p,q\} is connected.

(Here is the Zorn’s Lemma argument. Let \mathscr{L} be nonempty subcolletion of \mathscr{F} that is linearly ordered by inclusion. Let K=XXK=\cup_{X\in\mathscr{L}}X. We must show that KK\in\mathscr{F}. It not, we could write KK as the disjoint union of two closed sets SS and TT with pSp\in S and qTq\in T. Let f:M[0,1]f:M\to[0,1] be a continuous function such that S=f1S=f^{-1} and T=f1(1)T=f^{-1}(1). Since the sets X{f=1/2}X\cap\{f=1/2\} (with XX\in\mathscr{L}) are nested compact sets whose intersection K{f=1/2}K\cap\{f=1/2\} is empty, we see that X{f=1/2}X\cap\{f=1/2\} is empty for some XX\in\mathscr{L}. But then the set Q:=X{f1/2}=X{f<1/2}Q:=X\cap\{f\leq 1/2\}=X\cap\{f<1/2\} is a clopen subset of XX containing pp but not qq, a contradiction.) ∎

Theorem C.6.

Let XX be a metric space and f:XIf:X\to I be a proper, continuous map to an open interval I𝐑I\subset\mathbf{R}. Suppose XX has the following property: if JIJ\subset I is a compact interval, then there is a compact, connected set YXY\subset X such that f(Y)=Jf(Y)=J. Then there is a closed, connected subset QQ of XX such that f(Q)=If(Q)=I.

Proof.

We can assume that the interval is (0,1)(0,1). Add two points pp and qq to XX to get X~=X{p,q}\tilde{X}=X\cup\{p,q\}.

Add two points pp and qq to XX to get X~=X{p,q}\tilde{X}=X\cup\{p,q\}. Extend ff to X~\tilde{X} by setting f(p)=0f(p)=0 and f(q)=1f(q)=1. Let δ\delta be a metric on X~\tilde{X} such that X~\tilde{X} is compact, f:X~[0,1]f:\tilde{X}\to[0,1] is continuous, and δ\delta and dd induce the same topology on XX.

(Such a metric δ\delta can be defined as follows. First we extend d(,)d(\cdot,\cdot) to X{p,q}X\cup\{p,q\} by setting d(y,z)=|f(y)f(z)|d(y,z)=|f(y)-f(z)| if either yy or zz. This d(,)d(\cdot,\cdot) will not satisfy the triangle inequality. We define δ(y,y)\delta(y,y^{\prime}) to be the infimum of i=1nd(yi1,yi)\sum_{i=1}^{n}d(y_{i-1},y_{i}) among finite sequences y0,y1,,yny_{0},y_{1},\dots,y_{n} such that y0=yy_{0}=y and yn=yy_{n}=y^{\prime}. It is straightforward to show that δ\delta is a metric with the asserted properties.)

Let J1,J2,J_{1},J_{2},\dots be a sequence of compact subintervals of (0,1)(0,1) whose union is (0,1)(0,1). By hypothesis, there are compact, connected sets KnXK_{n}\subset X such that f(Kn)=Jnf(K_{n})=J_{n}. By Theorem C.2, there is a subsequence Kn(i)K_{n(i)} that converges to a compact, connected subset KK of X~\tilde{X}. Note that f(K)=[0,1]f(K)=[0,1]. By Theorem C.5, there is a closed connected set KK^{\prime} containing pp and qq such that X:=K{p,q}X:=K^{\prime}\setminus\{p,q\} is connected. Since f(K)=[0,1]f(K^{\prime})=[0,1] (by the intermediate value theorem), f(X)=(0,1)f(X)=(0,1). ∎

Remark C.7.

Theorem C.6 holds for any interval II, not just open intervals. It holds trivially if II is a closed interval. The case when II is half-open reduces to the open case by doubling.

Appendix D Notation

For the reader’s convenience, we provide a list of notation used in this paper. Items are listed alphabetically according to how they are spelled in English. Thus Γ\Gamma (“Gamma”) appears after F𝐯F_{\mathbf{v}} and before “Grim reaper surface”. Some notations that are used in only a small portion of the paper are not listed here; of course, those notations are explained where they occur.

  1. a(M)a(M), A(M)A(M). Definition 6.1.

  2. |A(M,p)||A(M,p)|: norm of the (Euclidean) second fundamental form of MM at pp.

  3. 𝒜\mathscr{A}, 𝒜(b)\mathscr{A}(b), 𝒜(b,x^)\mathscr{A}(b,\hat{x}), 𝒜(b,B,x^)\mathscr{A}(b,B,\hat{x}). Definition 7.4.

  4. Bowl solition. §2

  5. 𝒞\mathscr{C}: a space of translating annuli. Definition 6.1.

  6. 𝒞Γ(t)\mathscr{C}_{\Gamma(t)}. Defined just before Lemma 20.1.

  7. Cin(t)C_{\rm in}(t), Cout(t)C_{\rm out}(t). Defined just before Lemma 20.1.

  8. b(M)b(M): inner width of MM. §1, and also Definition 6.1.

  9. B(M)B(M): (outer) width of MM. §1, and also Definition 6.1.

  10. Δ\Delta-wing. §2.

  11. fb(x,y)f_{b}(x,y): Δ\Delta-wing over 𝐑×(b,b)\mathbf{R}\times(-b,b) with fb(0,0)=0f_{b}(0,0)=0 and Dfb(0,0)=0Df_{b}(0,0)=0. §2.

  12. fMf_{M}. Definition 12.1.

  13. F𝐯F_{\mathbf{v}}: The function F𝐯(p)=𝐯pF_{\mathbf{v}}(p)=\mathbf{v}\cdot p. See (4) in §4.

  14. Γ(t)\Gamma(t), Γin(t)\Gamma_{\rm in}(t), Γout(t)\Gamma_{\rm out}(t). Defined just before Lemma 20.1.

  15. Grim reaper surface. §2

  16. Minimal foliation function. Defined just before Theorem 4.3.

  17. MlowerM^{\textnormal{lower}}, MupperM^{\textnormal{upper}}. Theorems 8.5 and 10.5.

  18. MinnerYM^{Y}_{\textnormal{inner}}, MouterYM^{Y}_{\textnormal{outer}}. Definition 12.1.

  19. 𝖭(|M)\mathsf{N}(\mathscr{F}|M): number of critical points. Definition 4.1.

  20. 𝖭(F|M)\mathsf{N}(F|M). Defined just before Theorem 4.3.

  21. Ωin\Omega^{\textnormal{in}}, Ωout\Omega^{\textnormal{out}}. Theorem 10.3.

  22. 𝒫\mathscr{P}. Defined just before Lemma 19.5.

  23. innerM\partial_{\textnormal{inner}}M, outerM\partial_{\textnormal{outer}}M: inner and outer boundaries of MM. Definition 6.1.

  24. ϕupper(x)\phi^{\textnormal{upper}}(x), ϕlower(x)\phi^{\textnormal{lower}}(x). Theorems 8.1 and 8.2.

  25. \mathscr{R}: a space of translating annuli with rectangular boundaries. Definition 6.2.

  26. s(b)s(b): slope ub/x\partial u_{b}/\partial x of the grim reaper function ubu_{b}. See Item (2) in §2.

  27. Translator metric. Equation (3).

  28. ub(x,y)u_{b}(x,y): a grim reaper surface over 𝐑×(b,b)\mathbf{R}\times(-b,b). §2.

  29. uMu_{M}. Definition 11.3.

  30. waist(M)\operatorname{waist}(M). Definition 11.1.

  31. wb(x,y)w_{b}(x,y). Definition 13.1.

  32. x(M)x(M): necksize of MM. Definition 1.2.

  33. y(M)y(M). Equation (21).

  34. yinner(x,z)y^{\textnormal{inner}}(x,z), youter(x,z)y^{\textnormal{outer}}(x,z). Theorem 10.3.

  35. z(M)z(M). Definition 7.2.

References

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