Abstract

The Schwarz reflection principle states that a minimal surface ๐‘† in โ„3 is invariant under reflections in the plane of its principal geodesics and also invariant under 180ยฐ-rotations about its straight lines. We find new examples of embedded triply periodic minimal surfaces for which such symmetries are all of horizontal type.

1. Introduction

During the Clay Mathematics Institute 2001 Summer School on the Global Theory of Minimal Surfaces, M. Weber introduced the following terminology in his first lecture entitled Embedded minimal surfaces of finite topology:

โ€œA horizontal symmetry is a reflection at a vertical plane or a rotation about a horizontal line. A vertical symmetry is a reflection at a horizontal plane or a rotation about a vertical line.โ€

With this terminology, he proved that such symmetries induce symmetries in the cone metrics determined by ๐‘‘โ„Ž, ๐‘”๐‘‘โ„Ž, and ๐‘‘โ„Ž/๐‘” from a WeierstraรŸ pair (๐‘”,๐‘‘โ„Ž) of a minimal surface (see [1, 2] for details).

By classifying the symmetries this way, we sort out the space groups that might admit one, both, or none of them. Since minimal surfaces may model some natural structures, like crystals and copolymers, an example within a given symmetry group might fit an already existing compound, or even hint at nonexisting ones. However, several symmetry groups are not yet represented by any minimal surface (see [3, 4] for details and comments).

Restricted to symmetries given by reflections in the plane of principal geodesics and by 180ยฐ-rotations about straight lines contained in the surface, outside the triply periodic class it is easy to find complete embedded minimal surfaces in โ„3 of which these symmetries are either only horizontal or only vertical. For instance, the Costa surface (see [5โ€“7]) has only horizontal symmetries. The doubly periodic examples found by Meeks and Rosenberg in [8] have only vertical symmetries (see also [9] for nice pictures).

In the class of triply periodic minimal surfaces almost all known examples have either both or none of such symmetries, after suitable motion in โ„3. In fact, this must be true because most of the surfaces in this class have a cubic symmetry group. Examples with only horizontal symmetries do not seem to be well known. Besides the surfaces shown herein, perhaps there are only the โ€œTT-surfacesโ€ as Karcher named them in [10, pp. 297, 328-9] (see also [11]), and a surface from Fischer-Koch [12], which is, however, presented without rigorous proof.

The โ€œTT-surfacesโ€ are generated by an annulus, of which the boundary consists of two twisted equilateral triangles. For edge length 2โˆš3 and height 1, they coincide with the Schwarz P-surface, and hence have further symmetries besides the horizontal ones. Moreover, when a TT-surface has only horizontal symmetries, its translation group cannot be given by an orthogonal lattice.

In the present work, we give existence proofs for examples that are probably the first triply periodic minimal surfaces with only horizontal symmetries, of which the translation group is given by an orthogonal lattice. They are constructed by Karcherโ€™s method [6, 7, 10], although the purpose alone of few symmetries could be accomplished by modern methods introduced, for instance, by Traizet [13] and Fujimori and Weber [14]. However, Traizetโ€™s method is not explicit (in the sense explained in [15]) whereas Fujimori-Weberโ€™s method may turn it hard to analyse the so-called period problems. These are equations involving elliptic integrals with interdependent parameters.

Regarding examples with only vertical symmetries, we believe they have not been found yet.

The examples presented herein are inspired in the surfaces ๐ถ2 and ๐ฟ2,4 from [16, 17]. Any of those is generated by a fundamental piece, which is a surface with boundary in โ„3 with two catenoidal ends. The fundamental piece resembles the Costa surface with its planar end replaced by either symmetry curves or line segments. By suppressing the catenoidal ends, if we pile up several copies of the fundamental piece, we get the pictures in Figures 1 and 2(b). They are also named ๐ถ2 and ๐ฟ2,4.

The reader will notice that the surfaces ๐ถ4, also described in [16, 17], were not mentioned beforehand. This is because, for them, the โ€œpiling upโ€ procedure naturally forces extra symmetries to exist, and one goes back to another famous surface from H. Schwarz (see Figure 2(a)). Notice, for instance, the vertical straight line that comes out in the surface.

We are going to prove the following results:

Theorem 1.1. There exists a one-parameter family of triply periodic minimal surfaces in โ„3, of which the members are called ๐ถ2, and for any of them the following holds. (a)The quotient by its translation group has genus 9.(b)The whole surface is generated by a fundamental piece, which is a surface with boundary in โ„3. The boundary consists of four curves, each contained in a vertical plane. The fundamental piece has a symmetry group generated by reflections in two vertical planes and 180ยฐ-rotations about two line segments. (c)By successive reflections with respect to planes bounding the fundamental domain, and successive vertical translations, one obtains the triply periodic surface.

Theorem 1.2. For ๐‘˜=2,4, there exists a one-parameter family of triply periodic minimal surfaces in โ„3, of which the members are called ๐ฟ๐‘˜, and for any of them the following holds. (a)The quotient by its translation group has genus 2๐‘˜+1. (b)The whole surface is generated by a fundamental piece, which is a surface with boundary in โ„3. The boundary consists of four line segments. The fundamental piece has a symmetry group generated by reflections in two vertical planes and 180ยฐ -rotations about two line segments. Each of these segments makes an angle of ๐œ‹/๐‘˜ with the boundary. (c)By successive rotations about the boundary of the fundamental piece, and successive vertical translations, one obtains the triply periodic surface.

Sections 3 to 7 are devoted to the proof of Theorem 1.1. The proof of Theorem 1.2 follows very similar arguments, and we briefly discuss it in Section 8.

2. Preliminaries

In this section we state some basic definitions and theorems. Throughout this work, surfaces are considered connected and regular. Details can be found in [6, 7, 18โ€“20].

Theorem 2.1. Let ๐‘‹โˆถ๐‘…โ†’๐”ผ be a complete isometric immersion of a Riemannian surface ๐‘… into a three-dimensional complete flat space ๐”ผ. If ๐‘‹ is minimal and the total Gaussian curvature โˆซ๐‘…๐พ๐‘‘๐ด is finite, then ๐‘… is biholomorphic to a compact Riemann surface ๐‘… punched at finitely many points ๐‘1,โ€ฆ,๐‘๐‘Ÿ.

Definition 2.2. Let ๐‘ƒ={๐‘1,โ€ฆ,๐‘๐‘Ÿ} as in Theorem 2.1. An end is the image under ๐‘‹ of a punctured neighbourhood ๐‘‰๐‘ of a point ๐‘โˆˆ๐‘ƒ such that ๐‘ƒโˆฉ๐‘‰๐‘=โˆ…. We say that the surface has no ends when ๐‘ƒ=โˆ….

Theorem 2.3 (WeierstraรŸ Representation). Let ๐‘… be a Riemann surface, ๐‘” and ๐‘‘โ„Ž meromorphic function and 1-differential form on ๐‘…, such that the zeros of ๐‘‘โ„Ž coincide with the poles and zeros of ๐‘”. Suppose that ๐‘‹โˆถ๐‘…โ†’๐”ผ, given by ๎€œ๐‘‹(๐‘)โˆถ=Re๐‘๎€ท๐œ™1,๐œ™2,๐œ™3๎€ธ๎€ท๐œ™,where1,๐œ™2,๐œ™3๎€ธ1โˆถ=2๎‚ต1๐‘”๐‘–โˆ’๐‘”,๐‘”๎‚ถ+๐‘–๐‘”,2๐‘‘โ„Ž,(2.1) is well defined. Then ๐‘‹ is a conformal minimal immersion. Conversely, every conformal minimal immersion ๐‘‹โˆถ๐‘…โ†’๐”ผ can be expressed as (2.1) for some meromorphic function ๐‘” and 1-form ๐‘‘โ„Ž.

Definition 2.4. The pair (๐‘”,๐‘‘โ„Ž) is the WeierstraรŸ data and ๐œ™1, ๐œ™2, ๐œ™3 are the WeierstraรŸ forms on ๐‘… of the minimal immersion ๐‘‹โˆถ๐‘…โ†’๐‘‹(๐‘…)โŠ‚๐”ผ.

Theorem 2.5. Under the assumptions of Theorems 2.1 and 2.3, the WeierstraรŸ data (๐‘”,๐‘‘โ„Ž) extend meromorphically on ๐‘….

The function ๐‘” is the stereographic projection of the GauรŸ map ๐‘โˆถ๐‘…โ†’๐‘†2 of the minimal immersion ๐‘‹. It is a covering map of ๎โ„‚ and โˆซ๐‘†๐พ๐‘‘๐ด=โˆ’4๐œ‹deg(๐‘”). These facts will be largely used throughout this work.

3. The Surfaces ๐‘€ and the Functions ๐‘ง

Consider the surface indicated in Figure 1(a). A reflection in any of its vertical planar curves of the boundary leads to a fundamental piece which represents the quotient of a triply periodic surface ๐‘€ by its translation group. We are going to denote this quotient by ๐‘€. It is not difficult to conclude that it has genus 9. The fundamental domain of ๐‘€ is the shaded region indicated in Figure 3(a).

The surface ๐‘€ is invariant under 180โˆ˜-rotations around the directions ๎‹๐‘ฅ3 and ๎‹๐‘ฅ2. These rotations we call ๐‘Ÿ๐‘ฃ and ๐‘Ÿโ„Ž, respectively, (see Figure 3(a)). Based on this picture, one sees that the fixed points of ๐‘Ÿ๐‘ฃ are ๐‘†, ๐‘†๎…ž, ๐ฟ, ๐ฟ๎…ž, ๐น, ๐น๎…ž and the images of ๐‘† and ๐‘†๎…ž under the symmetries of ๐‘€. They sum up 8 in total. The quotients by ๐‘Ÿ๐‘ฃ and ๐‘Ÿโ„Ž we call ๐œŒ๐‘ฃ and ๐œŒโ„Ž, respectively. The surface ๐œŒ๐‘ฃ(๐‘€) is still invariant under the rotation ๐‘Ÿโ„Ž. In this case, the fixed points of ๐‘Ÿโ„Ž will be ๐œŒ๐‘ฃ(๐ด), ๐œŒ๐‘ฃ(๐ด๎…ž) and their images under the symmetries of ๐œŒ๐‘ฃ(๐‘€). They sum to 8 in total. Because of that, ๐œ’๎‚€๐œŒโ„Ž๎‚€๐œŒ๐‘ฃ๎‚€๐‘€=๎‚๎‚๎‚1โˆ’9+8/22+82=2.(3.1)

Let us define ๐‘งโˆถ=๐œŒโ„Žโˆ˜๐œŒ๐‘ฃโˆถ๐‘€โ†’๐‘†2โ‰ˆ๎โ„‚, such that ๐‘ง(๐‘†)=0, ๐‘ง(๐ฟ)=1 and ๐‘ง(๐ต)=๐‘–. The involutions of ๐‘€ are induced by ๐œŒ๐‘ฃ and ๐œŒโ„Ž on ๎โ„‚, and since all the involutions of ๎โ„‚ are given by Mรถbius transformations, we can conclude the following: ๐‘ง(๐‘†๎…ž)=0, ๐‘ง(๐น๎…ž)=โˆ’๐‘ง(๐น)=โˆ’๐‘ง(๐ฟ๎…ž)=1 and ๐‘ง(๐ต๎…ž)=๐‘–. By applying the symmetries of ๐‘€, one easily reads off the other values of ๐‘ง at the images (under these symmetries) of ๐‘†, ๐‘†๎…ž, ๐ฟ, ๐ฟ๎…ž, ๐น, ๐น๎…ž, ๐ต, and ๐ต๎…ž. Regarding the points ๐ด and ๐ด๎…ž, we have ๎โ„‚๐‘ง(๐ด)=๐‘ฅโˆˆ such that |๐‘ฅ|<1 and Arg(๐‘ฅ)โˆˆ(0,๐œ‹/2). Consequently, ๐‘ง(๐ด๎…ž)=โˆ’๐‘ฅ and one easily gets the other values of ๐‘ง at the images of ๐ด and ๐ด๎…ž under the symmetries of ๐‘€.

4. The ๐‘”-Function on ๐‘€ in Terms of ๐‘ง

First of all, observe that Jorge-Meeksโ€™ formula gives deg(๐‘”)=9โˆ’1=8. Let us then consider Figure 3(b). We will have ๐‘”โˆ’๐‘”โˆ’1=โˆž if and only if ๐‘งโˆ’๐‘งโˆ’1โˆˆ{0,โˆž}. Moreover, ๐‘”โˆ’๐‘”โˆ’1=0 if and only if ๐‘งโˆˆ{โˆ’๐‘ฅ,๐‘ฅ,๐‘ฅโˆ’1,โˆ’๐‘ฅโˆ’1,๐‘–๐‘Ž,๐‘–๐‘Žโˆ’1}, where ๐‘Žโˆˆ(0,1). From this point on we introduce the following notation: ๐‘โˆถ=๐‘งโˆ’๐‘งโˆ’1,๐‘‹โˆถ=๐‘ฅโˆ’1โˆ’๐‘ฅ,๐’œโˆถ=๐‘Ž+๐‘Žโˆ’1.(4.1)

By following Karcherโ€™s method in [6, 7], Figure 3 represents the surfaces whose existence we want to prove. From this picture we read off the necessary conditions for Theorem 1.1 to be valid. Afterwards, these will prove Theorem 1.1. The first condition is an algebraic relation between ๐‘” and ๐‘ง. Hence, based on Figure 3 and Karcherโ€™s method, it is not difficult to conclude that ๎‚ต1๐‘”โˆ’๐‘”๎‚ถ2=โˆ’๐‘–๐‘๐‘3โ‹…(๐‘โˆ’๐‘–๐’œ)2๎‚€(๐‘โˆ’๐‘‹)๐‘+๐‘‹๎‚,(4.2) where ๐‘ is a positive constant. Now we define ๐‘€ as a member of the family of compact Riemann surfaces given by the algebraic equation (4.2). Later on, we are going to verify that ๐‘€ has genus 9 indeed. But first we derive some conditions on the variables ๐‘Ž,๐‘ฅ, and ๐‘ in order to guarantee that ๐‘”2=โˆ’1 at ๐‘ง=โˆ’๐‘–๐‘Žยฑ1. This will be the case if ๐’œ๐‘=๐’œ2||๐‘‹||+2๐’œIm{๐‘‹}+2.(4.3)

Since |๐‘‹|2=Im2{๐‘‹}+Re2{๐‘‹}, one easily sees that ๐‘ is positive.

Now we analyse what happens to (4.2) under the map ๐‘งโ†’๐‘ง. In this case we will get ๐‘”โ†’๐‘–๐‘” or ๐‘”โ†’โˆ’๐‘–๐‘”. Therefore ๎‚ต1๐‘”+๐‘”๎‚ถ2=โˆ’๐‘–๐‘๐‘3โ‹…(๐‘+๐‘–๐’œ)2๎‚€๐‘โˆ’๐‘‹๎‚(๐‘+๐‘‹).(4.4)

At this point we are ready to prove that ๐‘€ has genus 9. The function ๐‘ง is a four-sheeted branched covering of the sphere. The values 0, โˆž, ยฑ1, ยฑ๐‘ฅยฑ1, ยฑ๐‘ฅยฑ1 correspond to the only branch points of ๐‘ง, all of them of order 2, and each of these values is taken twice on ๐‘€. Therefore, from the Riemann-Hurwitzโ€™s formula we have ๎‚€genus๐‘€๎‚=12โ‹…(2โˆ’1)โ‹…22โˆ’4+1=9.(4.5)

Now we are ready to find some relations that the parameters ๐‘Ž, ๐‘, and ๐‘ฅ will have to satisfy. These relations will make (4.2) and (4.4) consistent with the values of ๐‘” on the symmetry curves and lines of ๐‘€.

5. Conditions on the Parameters ๐‘Ž, ๐‘, and ๐‘ฅ

Consider the curves ๐‘†๎…ž๐ฟ and ๐น๎…ž๐‘† represented in Figure 3. The same picture shows how we have positioned our coordinate system. On the curve ๐‘†๎…ž๐ฟ, we expect that ๐‘”โˆˆ๐‘’๐‘–๐œ‹/4โ„ and on ๐น๎…ž๐‘† one should have ๐‘”โˆˆ๐‘’โˆ’๐‘–๐œ‹/4โ„. Let us now verify under which conditions this will really happen.

On ๐‘†๎…ž๐ฟ, we ought to have Re{(๐‘”โˆ’๐‘”โˆ’1)2}=โˆ’2. By taking ๐‘ง(๐‘ก)=๐‘ก, 0<๐‘ก<1, defining ๐‘‡โˆถ=๐‘กโˆ’๐‘กโˆ’1, and applying it to (4.2) we get the following equality: ๎€ท๐‘”โˆ’๐‘”โˆ’1๎€ธ2|||๐‘ง(๐‘ก)=โˆ’๐‘–๐‘๐‘‡3โ‹…(๐‘‡โˆ’๐‘–๐’œ)2๎‚€๐‘‡2||๐‘‹||โˆ’2๐‘–Im{๐‘‹}โ‹…๐‘‡โˆ’2๎‚.(5.1)

Therefore, ๎‚†๎€ทRe๐‘”โˆ’๐‘”โˆ’1๎€ธ2๎‚‡=โˆ’2๐‘๐‘‡2โ‹…๎‚€Im{๐‘‹}โ‹…๐‘‡2+๐’œ๐‘‡2โˆ’Im{๐‘‹}โ‹…๐’œ2||๐‘‹||โˆ’๐’œ2๎‚,(5.2) on the curve ๐‘ง(๐‘ก). Since we want Re{(๐‘”โˆ’๐‘”โˆ’1)2}=โˆ’2 on this curve, (5.2) will then give rise to the following conditions 1๐‘=,||๐‘‹||๐’œ+Im{๐‘‹}(5.3)๐’œ=โˆ’2.Im{๐‘‹}(5.4)

Equation (5.3) can be deduced from (4.3) and (5.4) by a simple calculation. Equation (5.4) will restrict the definition domain of our parameters. Since ๐‘Žโˆˆ(0,1), then ๐’œ>2, and by taking ๐‘ฅ=|๐‘ฅ|๐‘’๐‘–๐œƒ one clearly sees that Im{๐‘‹}<0 for ๐œƒโˆˆ(0,๐œ‹/2). From (5.4) we finally get the following restriction for the ๐‘ฅ-variable Re2{๐‘‹}>โˆ’2Im{๐‘‹}โˆ’Im2{๐‘‹}.(5.5)

Figure 4 illustrates the ๐‘‹-domain established by (5.5), and we recall that |๐‘ฅ|<1 and ๐œƒโˆˆ(0,๐œ‹/2).

It is not difficult to prove that (5.5) is equivalent to the following inequality: 1|๐‘ฅ|<2๎ƒฏโˆšsin๐œƒ+1+3cos2๎‚™๐œƒโˆ’๎‚€โˆš2sin๐œƒ1+3cos2๎‚๎ƒฐ๐œƒโˆ’sin๐œƒ.(5.6)

Of course, the right-hand side of (5.6) is one of the two roots of a second-degree equation. One easily proves that the other root is bigger than 1. Its inverse is exactly the right-hand side of (5.6), and this shows that it is positive and smaller than 1.

Regarding our remaining restriction, namely, Re{(๐‘”โˆ’๐‘”โˆ’1)2}=2 on ๐น๎…ž๐‘†, it is not difficult to verify that it leads to the same conditions (5.3) and (5.4). Therefore, we are now ready to write down Table 1, which summarizes some special involutions of ๐‘€.

Notice that the points (๐‘”,๐‘ง)=(ยฑ๐‘–,โˆ’๐‘–๐‘Žยฑ1) do not come out as fixed points of ๐œŒโ„Ž in Table 1. This is because the germs of the function ๐‘” at these points are not the same (see [21] for details). This has to do with the fact that the power of (๐‘+๐‘–๐’œ) is a multiple of the power of (๐‘”+๐‘”โˆ’1) in (4.4).

6. The Height Differential ๐‘‘โ„Ž on ๐‘€

Since the surface ๐‘€ has no ends, ๐‘‘โ„Ž must be a holomorphic differential form on it. The zeros of ๐‘‘โ„Ž are exactly at the points where ๐‘”=0 or ๐‘”=โˆž, and ord(๐‘‘โ„Ž)=|ord(๐‘”)| at these points. They should sum up 16 in total, which is consistent with deg(๐‘‘โ„Ž)=โˆ’๐œ’(๐‘€). Let us now analyse the differential ๐‘‘๐‘ง. Based on Figure 3, one sees that ๐‘‘๐‘ง has a simple zero at the points ๐‘งโˆ’1({0,ยฑ1,ยฑ๐‘ฅยฑ1,ยฑ๐‘ฅยฑ1}) and a pole of order 3 at the points ๐‘งโˆ’1({โˆž}). Let the symbol ~ indicate that two meromorphic functions on ๐‘€ differ by a nonzero proportional constant. It is not difficult to conclude that ๎‚€๐‘‘โ„Ž๎‚๐‘‘๐‘ง2โˆผ๎€ท๐‘ง2๎€ธโˆ’12๎€ท๐‘ง2โˆ’๐‘ฅ2๐‘ง๎€ธ๎€ท2โˆ’๐‘ฅโˆ’2๎€ธ๎‚€๐‘ง2โˆ’๐‘ฅ2๐‘ง๎‚๎‚€2โˆ’๐‘ฅโˆ’2๎‚.(6.1)

If we had a well-defined square root of the function at the right-hand side of (6.1), then we could get an explicit formula for ๐‘‘โ„Ž in terms of ๐‘ง and ๐‘‘๐‘ง. This square root exists indeed. By multiplying (4.2) and (4.4) it follows that ๎‚€(๐‘โˆ’๐‘‹)๐‘+๐‘‹๎‚๎‚€๐‘โˆ’๐‘‹๎‚(๐‘+๐‘‹)=โˆ’๐‘6๐‘2๎€ท๐‘2+๐’œ2๎€ธ2๎‚ต๐‘”2โˆ’1๐‘”2๎‚ถ2,(6.2) which allows us to define ๎‚™๎€ท๐‘2โˆ’๐‘‹2๎€ธ๎‚€๐‘2โˆ’๐‘‹2๎‚โˆถ=๐‘–๐‘3๐‘๎€ท๐‘2+๐’œ2๎€ธ๎‚ต๐‘”2โˆ’1๐‘”2๎‚ถ.(6.3)

Now we apply (6.3) to (6.1) and obtain ๐‘๐‘‘โ„Ž=๎‚™๎€ท๐‘2โˆ’๐‘‹2๎€ธ๎‚€๐‘2โˆ’๐‘‹2๎‚โ‹…๐‘‘๐‘ง๐‘ง.(6.4)

At (6.4) the equality sign holds because we want Re{๐‘‘โ„Ž}=0 on the straight line segment ๐‘†๐ต (see Figure 3(a)). On this segment ๐‘ง is purely imaginary and then we can fix both sides of (6.4) to be equal. Let us now verify if the symmetry curves and lines of ๐‘€ really exist. From Table 1 and (6.4) we write down Table 2.

From Table 2 it follows that ๐‘‘๐‘”/๐‘”โ‹…๐‘‘โ„Ž is purely imaginary on ๐‘†๐ต and ๐‘†๎…ž๐ต๎…ž. It is real on the other paths, confirming that ๐‘€ will have the expected symmetry curves and lines.

7. Solution of the Period Problems

The analysis of the period problems can be reduced to the analysis of the fundamental domain of our minimal immersion. If this fundamental domain is contained in a rectangular prism of โ„3, and if the boundary of the former is contained in the border of the latter, we will have that the fundamental piece of our minimal surface will be free of periods.

In order to obtain such a prism, a little reflection will show us that the following two conditions will be enough.(1)The symmetry ๐œŒโ„Ž really exists in โ„3.(2)After an orthonormal projection of the fundamental domain in the direction ๐‘ฅ3, we will have ๐‘†=๐‘†๎…ž and ๐ต=๐ต๎…ž (see Figure 5).

The first condition is easy to prove. Take a path ๐‘ƒโ†’๐ดโ†’๐‘ƒ๎…ž on ๐‘€ as indicated in Figure 6. Consider that ๐ดโ†’๐‘ƒ๎…ž with reversed orientation is the image of ๐‘ƒโ†’๐ด under the involution (๐‘”,๐‘ง)โ†’(โˆ’1/๐‘”,๐‘ง). Now we compute in โ„3 what happens to the coordinates of our minimal surface: ๎€ท๐‘ฅ1,๐‘ฅ2,๐‘ฅ3๎€ธ||(๐‘”,๐‘ง)โ†’(โˆ’1/๐‘”,๐‘ง)๎€œ=Re๐ด=(โˆ’๐‘–,๐‘ฅ)๐‘ƒ=(๐‘”0,๐‘ง0)๎€ท๐œ™1,๐œ™2,๐œ™3๎€ธ๎€œ=Re๐‘ƒ๐ด=(โˆ’๐‘–,๐‘ฅ)โ€ฒ=(โˆ’1/๐‘”0,๐‘ง0)๎€ท๐œ™1,โˆ’๐œ™2,๐œ™3๎€ธ๎€œ=Re๐ด=(โˆ’๐‘–,๐‘ฅ)๐‘ƒ=(๐‘”0,๐‘ง0)๎€ทโˆ’๐œ™1,๐œ™2,โˆ’๐œ™3๎€ธ=๎€ทโˆ’๐‘ฅ1,๐‘ฅ2,โˆ’๐‘ฅ3๎€ธ.(7.1)

Therefore, our minimal surface is really invariant under 180ยฐ-rotations around the ๐‘ฅ2-axis. This proves the existence of the symmetry ๐œŒโ„Ž of our initial assumptions.

Now we are ready to deal with the second condition. Consider Figure 5 with the segments ๐‘†๐ต and ๐ต๐ฟ on it. The period will be zero if and only if these segments have the same length, or equivalently ๎€œRe๐‘†๐ต๐œ™2๎€œ=Re๐ต๐ฟ๐œ™1.(7.2)

On ๐‘†๐ต we can take ๐‘(๐‘ก)=๐‘–๐‘ก, 2<๐‘ก<โˆž. This implies that โˆš๐‘‘๐‘ง/๐‘ง=โˆ’๐‘‘๐‘ก/๐‘ก2โˆ’4. From (4.4) and (6.4) we have ๐œ™2||๐‘(๐‘ก)=๐‘–๐‘ก=๐‘1/2(๐‘ก+๐’œ)๐‘ก1/2๎‚€๐‘ก2||๐‘‹||โˆ’2Im{๐‘‹}โ‹…๐‘ก+2๎‚1/2โ‹…๐‘‘๐‘กโˆš๐‘ก2โˆ’4.(7.3)

On ๐ต๐ฟ we can take ๐‘ง(๐‘ก)=๐‘’๐‘–๐‘ก, 0<๐‘ก<๐œ‹/2. From (4.2) and (6.4) it follows that ๐œ™1||๐‘ง(๐‘ก)=๐‘’๐‘–๐‘ก=1โˆš2โ‹…๐‘1/2(๐’œโˆ’2sin๐‘ก)๎‚€4sin2||๐‘‹||๐‘ก+4Im{๐‘‹}โ‹…sin๐‘ก+2๎‚1/2โ‹…๐‘‘๐‘กโˆšsin๐‘ก.(7.4)

Now define ๐ผ1โˆšโˆถ=(1/โˆซ2๐‘)๐ต๐ฟ๐œ™1 and ๐ผ2โˆšโˆถ=(1/โˆซ2๐‘)๐‘†๐ต๐œ™2. For ๐ผ1 apply the change of variables ๐‘ข2=sin๐‘ก and for ๐ผ2, ๐‘ก=2๐‘ขโˆ’2. A simple reckoning will lead to the following equalities: ๐ผ1=๎€œ10๐’œโˆ’2๐‘ข2๎‚€4๐‘ข4+4Im{๐‘‹}โ‹…๐‘ข2+||๐‘‹||2๎‚1/2โ‹…๐‘‘๐‘ขโˆš1โˆ’๐‘ข4,๐ผ(7.5)2=๎€œ102+๐’œ๐‘ข2๎‚€4โˆ’4Im{๐‘‹}โ‹…๐‘ข2+||๐‘‹||2๐‘ข4๎‚1/2โ‹…๐‘‘๐‘ขโˆš1โˆ’๐‘ข4.(7.6)

The next proposition will solve the period problem given by (7.2).

Proposition 7.1. For any fixed positive value of Re{๐‘‹} one has that the following limit exists and is positive limIm{๐‘‹}โ†’0๎€ท๐ผ(โˆ’Im{๐‘‹})โ‹…1โˆ’๐ผ2๎€ธ.(7.7) For Im{๐‘‹}=โˆ’1 one has that lim๐’œโ†’2(๐ผ1โˆ’๐ผ2) exists and is negative.

Proof. By recalling (5.4), a simple reckoning will show that limIm{๐‘‹}โ†’0(โˆ’Im{๐‘‹})โ‹…๐ผ1=๎€œ10Re2{๐‘‹}๎€ท4๐‘ข4+Re2๎€ธ{๐‘‹}1/2โ‹…๐‘‘๐‘ขโˆš1โˆ’๐‘ข4,limIm{๐‘‹}โ†’0(โˆ’Im{๐‘‹})โ‹…๐ผ2=๎€œ10Re2{๐‘‹}โ‹…๐‘ข2๎€ท4+Re2{๐‘‹}โ‹…๐‘ข4๎€ธ1/2โ‹…๐‘‘๐‘ขโˆš1โˆ’๐‘ข4.(7.8)
Since ๐‘ข2๎€ท4+Re2{๐‘‹}โ‹…๐‘ข4๎€ธ1/2<1๎€ท4๐‘ข4+Re2๎€ธ{๐‘‹}1/2(7.9) for every Re{๐‘‹}>0 and ๐‘ขโˆˆ(0,1), from (7.8) the first assertion of Proposition 7.1 follows.
By fixing Im{๐‘‹}=โˆ’1 and recalling (5.5), the convergence ๐’œโ†’2 is equivalent to Re{๐‘‹}โ†’1. This means that ๐‘‹ approaches the point 1โˆ’๐‘– indicated in Figure 4. An easy calculation will give us lim๐’œโ†’2๐ผ1=โˆš2๎€œ10๎€ท2๐‘ข4โˆ’2๐‘ข2๎€ธ+1โˆ’1/2โ‹…๎‚ต1โˆ’๐‘ข21+๐‘ข2๎‚ถ1/2๐‘‘๐‘ข,(7.10)lim๐’œโ†’2๐ผ2=โˆš2๎€œ10๎€ท2+2๐‘ข2+๐‘ข4๎€ธโˆ’1/2โ‹…๎‚ต1+๐‘ข21โˆ’๐‘ข2๎‚ถ1/2๐‘‘๐‘ข.(7.11)
The integrand of (7.10) can be rewritten as ๎‚€๐‘ข4+๎€ท1โˆ’๐‘ข2๎€ธ2๎‚โˆ’1/2โ‹…๎‚ต1โˆ’๐‘ข21+๐‘ข2๎‚ถ1/2=๎ƒฌ๐‘ข4๎€ท1โˆ’๐‘ข2๎€ธ2๎ƒญ+1โˆ’1/2โ‹…1โˆš1โˆ’๐‘ข4,(7.12) while one rewrites the integrand of (7.11) as ๎‚€๐‘ข4+๎€ท1+๐‘ข2๎€ธ2๎‚โˆ’1/2โ‹…๎‚ต1+๐‘ข21โˆ’๐‘ข2๎‚ถ1/2=๎ƒฌ๐‘ข4๎€ท1+๐‘ข2๎€ธ2๎ƒญ+1โˆ’1/2โ‹…1โˆš1โˆ’๐‘ข4.(7.13)
Since ๐‘ข4๎€ท1โˆ’๐‘ข2๎€ธ2>๐‘ข4๎€ท1+๐‘ข2๎€ธ2(7.14) for every ๐‘ขโˆˆ(0,1), the last assertion of Proposition 7.1 follows.

Proposition 7.1 provides a family of triply periodic surfaces of which a member is depicted in Figure 1(a). By looking at Figure 4, this family can be represented by the values of ๐‘‹ which belong to a curve ๐’ž contained in the shaded region. All members of this family will have only three periods, as suggested by Figure 1(a). Nevertheless, a priori there might be some nonembedded members, but it will not be the case. This is the subject of our next section.

8. Embeddedness of the Triply Periodic Surfaces

From now on we will denote our triply periodic surfaces by ๐‘€๐‘‹, where ๐‘‹โˆˆ๐’ž. Figure 6 shows that the projection of the unitary normal on a fundamental domain of ๐‘€๐‘‹ is contained in the lower hemisphere of ๎โ„‚. This means that (๐‘ฅ1,๐‘ฅ3) is an immersion of ๐’Ÿโˆถ={๐‘งโˆˆโ„‚โˆถ|๐‘ง|<1and0<Arg(๐‘ง)<๐œ‹/2} in โ„2. Figure 7 shows a possible image of this map in โ„2:

It is not difficult to prove that the contour of the shaded region in Figure 7 is a monotone curve. The ๐‘ฅ1-coordinate of the curve ๐ต๐ฟ is given by the integral of โˆ’๐œ™1 as in (7.4). The integrand is clearly positive, hence this stretch is monotone. Regarding ๐ฟ๐‘†๎…ž, where we can take ๐‘(๐‘ก)=๐‘ก, 0>๐‘ก>โˆ’โˆž, a simple reckoning gives us ||๐‘‘โ„Ž๐‘(๐‘ก)=๐‘ก=๐‘ก๎”๐‘ก4๎€ฝ๐‘‹โˆ’Re2๎€พ๐‘ก2+||๐‘‹||4โ‹…๐‘‘๐‘กโˆš๐‘ก2+4.(8.1)

Hence, the stretch ๐ฟ๐‘†๎…ž is also monotone. By using the symmetry ๐œŒโ„Ž, it follows that the whole contour indicated in Figure 7 is a monotone curve. Since the third coordinate of ๐ต๐ฟ is increasing, the projections ๐ต๐ฟ and ๐ฟ๐‘†๎…ž will intersect only at the point ๐ฟ. Nevertheless, it can happen that the projection ๐ฟ๐‘†๎…ž crosses ๐ต๎…ž๐น๎…ž. If we prove that this is not the case, the contour will have no self-intersections. The shaded region will then be simply connected, and we will conclude that the fundamental domain is a graph, hence embedded.

But even so, it can happen that the expanded triply periodic surface will not be embedded. We do not know whether the curve ๐ฟ๐‘†๎…ž crosses the ๐‘ฅ3-axis or not. A little reflection will show that, if ๐‘” does not take the value โˆ’๐‘’๐‘–๐œ‹/4 along ๐ฟ๐‘†๎…ž, then this curve does not intersect the vertical axis. Consequently, its projection will not intersect ๐ต๎…ž๐น๎…ž. In this case, since the triply periodic surface is expanded horizontally by reflections only, and vertically by rotations only, the whole surface will then be embedded.

By using the maximum principle, if we find an embedded member of our family in the curve ๐’ž, the whole family will then consist of embedded surfaces. The following proposition gives us such a member and will conclude this section.

Proposition 8.1. There is an โˆš๐‘‹โˆˆโˆ’๐‘–+(1,22) such that ๐‘‹โˆˆ๐’ž and ๐‘€๐‘‹ is embedded.

Proof. We will prove that ๐‘”โ‰ โˆ’๐‘’๐‘–๐œ‹/4 along ๐ฟ๐‘†๎…ž, for any โˆš๐‘‹โˆˆโˆ’๐‘–+(1,22). Moreover, (๐ผ1โˆ’๐ผ2)|โˆš๐‘‹=22โˆ’๐‘– will be positive. These two facts together with Proposition 7.1 will conclude Proposition 8.1.
By recalling (5.1), we would have ๐‘”=โˆ’๐‘’๐‘–๐œ‹/4 for some ๐‘‡โˆˆ(โˆ’โˆž,0) if and only if ๎€ท๐‘‡2โˆ’๐’œ2๎€ธ๎‚€๐‘‡2โˆ’||๐‘‹||2๎‚=4๐’œIm{๐‘‹}โ‹…๐‘‡2.(8.2)
Equation (8.2) will not be fulfilled by any ๐‘‡2โˆˆ(0,โˆž) providing |||๐’œ2+||๐‘‹||2|||||๐‘‹||+4๐’œIm{๐‘‹}<2๐’œ,(8.3) or equivalently โˆ’Re{๐‘‹}โˆšIm{๐‘‹}<22.(8.4)
We have fixed Im{๐‘‹}=โˆ’1, hence ๐‘”โ‰ โˆ’๐‘’๐‘–๐œ‹/4 along ๐ฟ๐‘†๎…ž for any โˆš๐‘‹โˆˆโˆ’๐‘–+(1,22). Let us now verify that (๐ผ1โˆ’๐ผ2)|โˆš๐‘‹=22โˆ’๐‘–>0. From (7.5) we have ๐ผ1||โˆš๐‘‹=22โˆ’๐‘–=๎€œ109โˆ’2๐‘ข2๎€ท4๐‘ข4โˆ’4๐‘ข2๎€ธ+91/2โ‹…๐‘‘๐‘ขโˆš1โˆ’๐‘ข4,(8.5) and from (7.6) it follows that ๐ผ2||โˆš๐‘‹=22โˆ’๐‘–=๎€œ102+9๐‘ข2๎€ท4+4๐‘ข2+9๐‘ข4๎€ธ1/2โ‹…๐‘‘๐‘ขโˆš1โˆ’๐‘ข4.(8.6)
But 9โˆ’2๐‘ข2๎€ท4๐‘ข4โˆ’4๐‘ข2๎€ธ+91/22>3โˆ’3๐‘ข2,โˆ€๐‘ขโˆˆ(0,1),(8.7) and if we define โˆš๐‘Žโˆถ=1โˆ’11/17 it is possible to prove that 2+9๐‘ข2๎€ท4+4๐‘ข2+9๐‘ข4๎€ธ1/2<๐‘Ž๐‘ข2โˆ’2๐‘Ž๐‘ข+1,โˆ€๐‘ขโˆˆ(0,1).(8.8)
But ๎‚๐ผ1๎€œโˆถ=10๎€ท3โˆ’2๐‘ข2๎€ธ/3๐‘‘๐‘ขโˆš1โˆ’๐‘ข4=34๐ต๎‚€14,12๎‚โˆ’16๐ต๎‚€34,12๎‚,๎‚๐ผ2๎€œโˆถ=10๎€ท๐‘Ž๐‘ข2๎€ธโˆ’2๐‘Ž๐‘ข+1๐‘‘๐‘ขโˆš1โˆ’๐‘ข4=๐‘Ž4๐ต๎‚€34,12๎‚โˆ’๐‘Ž2๐ต๎‚€12,12๎‚+14๐ต๎‚€14,12๎‚.(8.9)
Now we use ๐ต(๐‘š,๐‘›)=ฮ“(๐‘š)ฮ“(๐‘›)/ฮ“(๐‘š+๐‘›), ฮ“(1/4)=3,625600โ€ฆ, โˆšฮ“(1/2)=๐œ‹ and ฮ“(3/4)=1,225417โ€ฆ in order to conclude that ๎‚๐ผ1>๎‚๐ผ2.(8.10)
Together with (8.5)โ€“(8.9), (8.10) shows that (๐ผ1โˆ’๐ผ2)|โˆš๐‘‹=22โˆ’๐‘– is positive.

9. The Surfaces ๐ฟ2,4

In order to prove Theorem 1.2, one follows very similar ideas already explained in Sections 3 to 7. For the surfaces ๐ฟ2, consider Figures 8(a) and 8(b). The fundamental piece ๐‘€ has genus 5, and ๐‘‚๐‘ฅ2 passes through point ๐ด. The piece is invariant under ๐‘Ÿ๐‘ฃ and ๐‘Ÿโ„Ž, with quotient functions ๐œŒ๐‘ฃ and ๐œŒโ„Ž, respectively.

Since ๐œ’๎‚€๐œŒโ„Ž๎‚€๐œŒ๐‘ฃ๎‚€๐‘€=๎‚๎‚๎‚1โˆ’5+8/22+42=2,(9.1) we may define ๐‘งโˆถ=๐œŒโ„Žโˆ˜๐œŒ๐‘ฃโˆถ๐‘€โ†’๐‘†2โ‰ˆ๎โ„‚, such that ๐‘ง(๐‘†)=0,๐‘ง(๐ต)=1 and ๐‘ง(๐ฟ)=โˆž. The symmetries imply ๐‘ง(๐‘†๎…ž)=0,๐‘ง(๐ต๎…ž)=1, and ๐‘ง(๐ฟ๎…ž)=โˆž whereas ๐‘ง(๐ด) is a certain complex ๐‘ฅ in the first quadrant. Moreover, there is a point in the segment ๐ต๐‘† at which ๐‘”=1. After analysing the divisors of ๐‘ง and ๐‘” on ๐‘€, together with the behaviour of the unitary normal on symmetry curves and lines, we get ๎‚ต1๐‘”+๐‘”๎‚ถ2=1/๐‘Žโˆ’๐‘Ž|๐‘ฅโˆ’๐‘Ž|2โ‹…๎€ท(๐‘งโˆ’๐‘ฅ)๐‘งโˆ’๐‘ฅ๎€ธ(๐‘ง+๐‘Ž)2๐‘ง๎€ท1โˆ’๐‘ง2๎€ธ.(9.2) Since there is a point in the segment ๐น๐‘† at which ๐‘”=โˆ’๐‘–, we should also have ๎‚ต1๐‘”โˆ’๐‘”๎‚ถ2=1/๐‘Žโˆ’๐‘Ž|๐‘ฅโˆ’๐‘Ž|2โ‹…๎€ท(๐‘ง+๐‘ฅ)๐‘ง+๐‘ฅ๎€ธ(๐‘งโˆ’๐‘Ž)2๐‘ง๎€ท1โˆ’๐‘ง2๎€ธ.(9.3) In order to have equivalence between (9.2) and (9.3), a necessary and sufficient condition is ๐’œ=๐‘Ž+๐‘Žโˆ’1=(|๐‘ฅ|2+1)/Re{๐‘ฅ}. Now, it is easy to get ๐‘‘โ„Ž=๐‘–๐‘‘๐‘ง๎‚™๎€ท๐‘ง2โˆ’๐‘ฅ2๎€ธ๎‚€๐‘ง2โˆ’๐‘ฅ2๎‚,(9.4) with a well-defined square root in the denominator. One checks the assumed symmetries the same way we did in Tables 1 and 2. The unique period problem is again (7.2), which can be visualised again by Figure 5. Therefore, (7.2) is equivalent to ๐ฝ1=๐ฝ2, where ๐ฝ1=๎€œ10(๐‘ก+๐‘Ž)๐‘‘๐‘ก๎”|๐‘ก+๐‘ฅ|๐‘ก๎€ท1โˆ’๐‘ก2๎€ธ,๐ฝ2=๎€œโˆž1(๐‘กโˆ’๐‘Ž)๐‘‘๐‘ก๎”|๐‘กโˆ’๐‘ฅ|๐‘ก๎€ท๐‘ก2๎€ธ.โˆ’1(9.5) The change ๐‘กโ†ฆ1/๐‘ก for ๐ฝ2 makes clear that ๐ฝ1<๐ฝ2 (๐ฝ1>๐ฝ2) providing ๐‘…1<๐‘…2 (๐‘…1>๐‘…2), where ๐‘…1=(๐‘ก+๐‘Ž)/(1โˆ’๐‘Ž๐‘ก) and ๐‘…2=|(๐‘ก+๐‘ฅ)/(1โˆ’๐‘ฅ๐‘ก)|, 0<๐‘ก<1. On the one hand, for a fixed ๐‘Ÿ=Re{๐‘ฅ}>1, if Im{๐‘ฅ}โ†’0 then ๐‘Žโ†’1/๐‘Ÿ, and consequently ๐‘…1<๐‘…2. On the other hand, by fixing Im{๐‘ฅ} and letting Re{๐‘ฅ}โ†’0, then ๐‘Žโ†’0 and so ๐‘…1>๐‘…2. In this case, notice that the singularity at ๐‘ก=1 of both integrands in (9.5) is easily removable with a change of variables. This means, no matter if we have ๐‘…1|๐‘ก=1=๐‘…2|๐‘ก=1, it still holds ๐ฝ1>๐ฝ2.

For the surfaces ๐ฟ4, consider Figures 9(a) and 9(b). The fundamental piece ๐‘€ has genus 9, and ๐‘‚๐‘ฅ2 passes through point ๐ด. The piece is invariant under ๐‘Ÿ๐‘ฃ and ๐‘Ÿโ„Ž, with quotient functions ๐œŒ๐‘ฃ and ๐œŒโ„Ž, respectively. We will have ๐‘”โˆ’๐‘”โˆ’1=โˆž if and only if ๐‘ง+๐‘งโˆ’1โˆˆ{ยฑ๐‘–,0,โˆž}. Moreover, ๐‘”โˆ’๐‘”โˆ’1=0 if and only if ๐‘งโˆˆ{โˆ’๐‘ฅ,๐‘ฅ,โˆ’๐‘ฅโˆ’1,๐‘ฅโˆ’1,๐‘–๐‘Ž,โˆ’๐‘–๐‘Žโˆ’1}, where ๐‘Žโˆˆ(0,1).

From this point on we redefine the following: ๐‘โˆถ=๐‘งโˆ’1+๐‘ง,๐‘‹โˆถ=๐‘ฅโˆ’1+๐‘ฅ,๐’œโˆถ=๐‘Žโˆ’1โˆ’๐‘Ž.(9.6)

Based on Figure 9 it is not difficult to conclude that ๎‚ต1๐‘”โˆ’๐‘”๎‚ถ2=๐‘–๐‘๐‘3โ‹…(๐‘+๐‘–๐’œ)2๎‚€(๐‘+๐‘‹)๐‘โˆ’๐‘‹๎‚,(9.7) where ๐‘ is given by (4.3) again. Moreover, (9.7) is equivalent to ๎‚ต1๐‘”+๐‘”๎‚ถ2=๐‘–๐‘๐‘3โ‹…(๐‘โˆ’๐‘–๐’œ)2๎‚€๐‘+๐‘‹๎‚(๐‘โˆ’๐‘‹).(9.8) Similar arguments as in Section 5 will give again (5.3) and (5.4), but unlike (5.5) there is no restriction now. Regarding ๐‘‘โ„Ž, it still holds (6.4), but unlike Figure 5 the period problem is now illustrated by Figure 10.

Integrals ๐ผ1 and ๐ผ2 are again given by (7.5) and (7.6), but now the period is solved when 2๐ผ1=๐ผ2. This will come with the following.

Proposition 9.1. For any fixed positive value of Re{๐‘‹}one has that the following limit exists and is positive: limIm{๐‘‹}โ†’0๎€ท(โˆ’Im{๐‘‹})โ‹…2๐ผ1โˆ’๐ผ2๎€ธ.(9.9) For Im{๐‘‹}=โˆ’1 one has that lim๐’œโ†’0(I1โˆ’2I2) exists and is negative.

The proof of Proposition 9.1 is quite similar to the proof of Proposition 7.1, and so we will omit it here. The arguments for the embeddedness of ๐ฟ2,4 are even easier than the ones used in Section 8 for ๐ถ2, because now the contours are given by four straight line segments and two curves, pairwise congruent.

Acknowledgments

For this present paper, V. R. Batista was supported by the Grants โ€œBolsa de Produtividade Cientรญficaโ€ from CNPqโ€”Conselho Nacional de Desenvolvimento Cientรญfico e Tecnolรณgico, and โ€œBolsa de Pรณs-Doutoradoโ€ FAPESP 2000/07090-5.