You are on page 1of 12

Downloaded from rsfs.royalsocietypublishing.

org on January 10, 2013

Reflections concerning triply-periodic minimal surfaces


Alan H. Schoen Interface Focus 2012 2, doi: 10.1098/rsfs.2012.0023 first published online 30 May 2012

References

This article cites 28 articles, 1 of which can be accessed free

http://rsfs.royalsocietypublishing.org/content/2/5/658.full.html#ref-list-1
Article cited in: http://rsfs.royalsocietypublishing.org/content/2/5/658.full.html#related-urls

Subject collections

Articles on similar topics can be found in the following collections biocomplexity (30 articles) biomathematics (26 articles)

Email alerting service

Receive free email alerts when new articles cite this article - sign up in the box at the top right-hand corner of the article or click here

To subscribe to Interface Focus go to: http://rsfs.royalsocietypublishing.org/subscriptions

Downloaded from rsfs.royalsocietypublishing.org on January 10, 2013

Interface Focus (2012) 2, 658668 doi:10.1098/rsfs.2012.0023 Published online 30 May 2012

DISCUSSION

Reections concerning triply-periodic minimal surfaces


Alan H. Schoen*
Carbondale, IL, USA In recent decades, there has been an explosion in the number and variety of embedded triplyperiodic minimal surfaces (TPMS) identied by mathematicians and materials scientists. Only the rare examples of low genus, however, are commonly invoked as shape templates in scientic applications. Exact analytic solutions are now known for many of the low genus examples. The more complex surfaces are readily dened with numerical tools such as SURFACE EVOLVER software or the Landau Ginzburg model. Even though table-top versions of several TPMS have been placed within easy reach by rapid prototyping methods, the inherent complexity of many of these surfaces makes it challenging to grasp their structure. The problem of distinguishing TPMS, which is now acute because of the proliferation of examples, has been addressed by Lord & Mackay (Lord & Mackay 2003 Curr. Sci. 85, 346 362). Keywords: triply-periodic; minimal; surfaces

In recent decades, there has been an explosion in the number and variety of embedded triply-periodic minimal surfaces (TPMS) identied by mathematicians and materials scientists. Only the rare examples of low genus, however, are commonly invoked as shape templates in scientic applications. Exact analytic solutions are now known for many of the low genus examples. The more complex surfaces are readily dened with numerical tools such as Ken Brakkes SURFACE z dz s Landau EVOLVER software [1] or Holyst and Go Ginzburg model [2]. Even though table-top versions of several TPMS have been placed within easy reach by rapid prototyping methods, the inherent complexity of many of these surfaces makes it challenging to grasp their structure. The problem of distinguishing TPMS, which is now acute because of the proliferation of examples, has been addressed by Eric Lord and Alan Mackay [3]. I describe here some highlights of a journey that culminated unexpectedly in 1966 in a search for new examples of TPMS, exactly 100 years after Schwarz wrote his monumental treatise on TPMS [4] (a more detailed account of my search can be found in [5,6] and at http://www.schoengeometry.com/e_tpms. html). The helicoid and the catenoid were the only minimal surfaces I had encountered before 1966. I knew only the bare rudiments of differential geometry and little more of complex analysis. In those days, soft
*schoenah@gmail.com One contribution of 18 to a Theme Issue Geometry of interfaces: topological complexity in biology and materials. Received 26 April 2012 Accepted 26 April 2012

matter, mesoscopic, block copolymer, MCM-48, photonic crystals, etc., were not yet household words. My principal research interests were random and correlated random walks on lattices, atomic diffusion in crystals, the Mo ssbauer effect, and the combinatorial and geometrical connections between triply-periodic graphs and polyhedra. Some of these topics became fused in my mind, leading me along a tortuous path to TPMS. The rst step on this journey was my discovery in 1956 of a method for distinguishing between the vacancy and interstitial mechanisms for atomic diffusion in crystalline solids (http://www.schoen geometry. com/e_tpms.html) [79]. In 1951, in a mathematical analysis of self-diffusion by the vacancy mechanism, in which the elementary step of a diffusing atom is its exchange of position with a vacancy (vacant lattice site), Conyers Hering found that the correlation between the directions of consecutive steps of the atom reduces the self-diffusion coefcient by the fractional amount 1 2 f, where f 1 kcos ulAv =1 kcos ulAv ; kcos ulAv is the average value of the cosine of the angle u between consecutive jumps of the diffusing atom (cf. table 1); f is known as the BardeenHering correlation factor [9]. I was startled to discover that for self-diffusion by the vacancy mechanism, correlation reduces both the diffusion coefcient and the isotope effect by exactly the same fractional amount. This meant that isotope effect measurements could provide the rst experimental evidence to support the prevailing notion that selfdiffusion in close-packed crystals of cubic symmetry occurs by the vacancy mechanism. By contrast, one

658

This journal is q 2012 The Royal Society

Downloaded from rsfs.royalsocietypublishing.org on January 10, 2013

Discussion. Triply-periodic minimal surfaces

A. H. Schoen

659

Figure 1. Large interstitial sites (red) in the diamond crystal structure (green) (stereo image).

Table 1. Calculated values of the Bardeen Hering correlation factor f [10] for four two-dimensional and four three-dimensional structures. Z is the coordination number of the graph. structure linear chain honeycomb layer square layer triangle layer diamond simple cubic body-centred cubic face-centred cubic Z 2 3 4 6 4 6 8 12 2 kcos ulAv 1 1/2 p 1 2 2/p 5/6 2 3/p 1/3 0.209841 3 4 1 1 2 G4(1 4)8p 2 8p/G (4) 0.123. . . f 0 1/3 0.466942 0.566057 1/2 0.653120 0.727194 0.781. . .

would expect that small foreign atoms in dilute solution in silicon or germanium would occupy the comparatively large interstitial spaces (cf. gure 1) and diffuse by strictly random walk, as they hop from one interstitial site to another. Because the directions of consecutive steps of an atom in such a case are uncorrelated, f 1. These expectations were conrmed: in self-diffusion isotope effect experiments in single crystals of palladium, Peterson obtained results consistent with those of the vacancy mechanism [10], while measurements of the isotope effect for lithium diffusion in silicon by Pell [11] implied diffusion by random walk, which is consistent only with interstitial diffusion. The computational techniques used before 1960 to calculate correlation factors yielded only approximate values. Even though these were accurate enough for comparison with experiment, I decided to derive exact values, using a direct combinatorial approach. In 1960, with the help of a key clue suggested to me by the theoretical physicist J. Kanamori (1960, personal communication), R. W. Lowen Jr and I obtained solutions for f expressed as elliptic integrals [12], yielding exact values for six two- and three-dimensional structures, plus a six-gure numerical value for the simple cubic (sc) lattice. Our results are summarized in table 1. I do not recall whether we succeeded in obtaining analytic expressions for the face-centred cubic ( fcc) lattice. The approximate
Interface Focus (2012)

fcc value listed there, which Bob Lowen and I were able to conrm, was computed by Hering [9]. The smaller the value of f, the easier it is to obtain accurate experimental values of the isotope effect. The entries in table 1 suggested to me that for self-diffusion by the vacancy mechanism in a monatomic cubic crystal with Z 3, f would probably be less than 0.5. (The data in the table are consistent with the approximation kcos ulAv 1=Z 1; for both two- and three-dimensional structures.) Although I was fascinated in 1958 by Wells stereoscopic images [13] of an intertwined pair of enantiomorphic Laves graphs [1417], for each of which Z 3, I was disappointed to learn that Wells considered it highly unlikely that there exists a monatomic crystal in which the atomic positions correspond to the vertices of just one Laves graph [18]. In 1960, I constructed a model of an intertwined pair of Laves graphs, and six years later it became my guide when I conjectured the existence of a TPMSthe gyroidthat separates the two graphs (gure 2). While pondering the classical concepts of dual maps and reciprocal polyhedra described by Coxeter [19,20], I attempted to develop a systematic general procedure for constructing a kind of dual relation for pairs of triplyperiodic graphs whose edges coincide with hypothetical diffusion pathwaysone graph for self-diffusion and the other for interstitial diffusion. I wondered what restrictions would be required on the properties of a graph in

Downloaded from rsfs.royalsocietypublishing.org on January 10, 2013

660

Discussion. Triply-periodic minimal surfaces

A. H. Schoen

Figure 2. Two dual enanatiomorphic Laves graphs.

order for such a dual recipe to be effective. For example, would the graph have to be symmetrici.e. both edgetransitive and vertex-transitive? I searched for crystals of cubic symmetry in which there is only one kind of interstitial site of atomic proportions, because I assumed that interstitial diffusion would approximate random walk most closely in such crystals. In the spring of 1966, at the suggestion of Konrad Wachsmann [21], architecture chairman at the University of Southern California, I paid a call on Peter Pearce [22,23], a Los Angeles architect/designer who was studying polyhedral packings and triply-periodic networks (graphs). Peters studio was lled with ball-and-stick models of crystal structures, two of which I found especially intriguing. One of them modelled the diamond crystal structure and the other the body-centred cubic lattice. I will call these two the diamond graph and the bcc graph. A single interstitial cavity in each of them was occupied by what Peter called a saddle polyhedronan object whose faces are skew polygons congruent to the smallest edge circuit in the graph (cf. gures 3b and 4b). Peter had been inspired by a museum exhibit, designed by his former mentor Charles Eames [24] and the mathematician Ray Redheffer [25], in which a regular skew quadrangular boundary frame was periodically immersed in a beaker of soapy water, spanning a minimal surface each time it emerged from the beaker. When I saw Peters two saddle polyhedra, I recognized immediately that at least for the graphs sc, bcc, Laves and diamondand probably also for other graphsit was true that: a point at the centre of the saddle polyhedron P1 that is interstitial with respect to the graph G1 is a vertex of a second graph G2, and a point at the centre of the saddle polyhedron P2 that is interstitial with respect to the graph G2 is a vertex of the original graph G1. The edges of G1 protrude through the faces of P2, and the edges of G2 protrude through the faces of P1. I concluded that saddle polyhedra might serve as the basis for a three-dimensional dual relation analogous to the conventional duality of planar graphs. I called G1 a
Interface Focus (2012)

(a)

(b)

Figure 3. (a; stereo image) The dual graphs bcc (blue vertices and edges) and WP (orange vertices and edges); (b; stereo image) Tetragonal tetrahedron, interstitial polyhedron of the bcc graph.

nodal graph and G2 an interstitial graph, but a saddle polyhedron that is interstitial in one of the graphs is nodal in its dual and vice versa. Figures 3a,b and 4a,b illustrate this duality for the bcc and I-WP graphs. A few graphs, such as sc, diamond and Laves, are selfdual, but the graphs of most dual pairs, such as the bcc graph (cf. gure 3a) and the WP graph (cf. gure 4a), are dissimilar. Two dual Laves graphs are enantiomorphic because the Laves graph is chiral. The nodal polyhedron for the Laves graph is shown in gure 5. During the weeks following my rst meeting with Peter Pearce, I tested my dual recipe on a variety of other graphs, not only symmetric ones, obtaining unambiguous results in every case. In some cases, the saddle polyhedra degenerated into convex polyhedra. In the 6-valent sc graph and the 12-valent fcc graph, for example, both the interstitial and nodal polyhedra produced by the recipe are the Voronoi polyhedra of the vertices of their respective graphs. For the fcc graph, there are two shapes of interstitial polyhedrathe

Downloaded from rsfs.royalsocietypublishing.org on January 10, 2013

Discussion. Triply-periodic minimal surfaces


(a)

A. H. Schoen

661

(b)

Figure 5. nodal polyhedron of the Laves graph. Its mirror image is the interstitial polyhedron of the graph (stereo image).

Figure 4. (a; stereo image) WP graph; (b; stereo image) Expanded octahedron, interstitial polyhedron of the WP graph.

regular octahedron and the regular tetrahedron. The dual graph in this case is not k-regular, because it has both degree 4 and degree 8 vertices. Nevertheless, the nodal polyhedron produced by the recipe for fcc is the rhombic dodecahedron, as one might expect. In spite of all these favourable indications, I recognized that the recipe, which is described in [6], is essentially an ad hoc heuristic. Expecting it to fail eventually, I remained on the lookout for the failure case. In recent years, several authors [26,27] have applied DelaneyDress tiling theory to formalize the concept of saddle polyhedra, treating a greatly expanded set of graphs, with results equivalent to those sketched here. The treatment of the subject has become simplied, in part because these authors have devised improved conventions for naming both graphs and saddle polyhedra. Peter Pearce fabricated each face of his saddle polyhedra by pushing a metal tool in the shape of the face boundary into a heated sheet of transparent vinyl, stretching it into an approximation of a minimal surface. I constructed my own plastic models of a variety of saddle polyhedra, using my childrens toy vacuumforming machine and tools made in my tiny garage shop. I made each mould for vacuum-forming face polygons as a solid cast of polyester resin, rst stretching a rubber membrane across the boundary of the face tool and then pouring resin onto the membrane and allowing it to harden. My rst encounter with TPMS occurred soon after I met Peter Pearce. While I was thinking about how to name saddle polyhedra, I was startled to discover two spectacular surfaces (cf. gure 6a,c) I had never seen before. Let me call the regular skew hexagonal face of the expanded octahedron shown in gure 4b Fp/2 because it has 908 face angles. Adjacent faces of this saddle polyhedron are related by mirror reection in the plane that contain a shared pair of consecutive edges. But if adjacent faces are related instead by a half-turn about a common edge, their union denes a
Interface Focus (2012)

smooth surface spanned by a 10-sided skew polygon. No matter how many additional replicas of Fp/2 are attached in this fashion, the emerging triply-periodic labyrinthine surface remains free of self-intersections. Applying the same procedure to the regular skew hexagon Fp/3, which has 608 face angles, yields a second triply-periodic surface. I named the Fp/2 surface D and the Fp/3 surface P. The lattice for D is fcc and the lattice for P is sc. Eight Fp/2 hexagons dene a lattice fundamental domain for D and four Fp/3 hexagons dene one for P. In June 1966, I telephoned the minimal surface expert Hans Nitsche for information about the surfaces I was calling D and P. I described them as optimally smoothed versions of the three innite regular skew polyhedra Coxeter and Petrie discovered as schoolboys in the 1920s [28]. Hans informed me that D and P are the two adjoint minimal surfaces investigated in 1866 by Schwarz [4], who proved that they are described by conjugate harmonic functions in Weierstrass integrals. He explained that the smoothness at the junction between their hexagonal or quadrangular faces is a consequence of Schwarzs reection principle [4,29]. A few weeks later, Norman Johnson [30] visited me at my home, where we discussed the Coxeter maps f6,4j4g, f4,6j4g and f6,6j3g [24] and their relevance to these surfaces. The symbol f6,4j4g, for example, describes a tiling by regular six-gons, with four incident on each vertex, and holes that are regular four-gons. Now I began to study Schwarz [4], Eisenhart [31], Hilbert & Cohn-Vossen [32] and a few other authors. While I was learning more about the mathematics of minimal surfaces, I conducted a variety of wire-frame experiments with soap lms, starting with the catenoid and the helicoid. I was mildly curious about the shape of the curve around the waist of the square catenoid (cf. gure 7). It was clear from soap lm experiments with two square rings at different separation distances that this curve is not a circle (cf. the circular waist of the true catenoid), but to prove that it is not, it is necessary to invoke Bjo rlings theorem [33]: if two minimal surfaces contain a curve C at all corresponding points of which the surface tangent planes are the same, then the surfaces are the same. Early in 1968, my colleague Jim Wixson and I computed the shape of this square catenoid waist curve, using Schwarzs equations for the P surface [4]. As I anticipated, it bulges slightly outward in the neighbourhood of the four points closest to the corners of the squares.

Downloaded from rsfs.royalsocietypublishing.org on January 10, 2013

662

Discussion. Triply-periodic minimal surfaces


(a) (b)

A. H. Schoen
(c)

Figure 6. The Coxeter-Petrie f6,4j4gmap [28] on the surfaces (a) D, (b) G and (c) P.

Figure 7. The linear asymptotics and plane geodesics in the square catenoid of Schwarzs P surface (stereo image).

(In September 1968, I made use of Schwarzs analytic expression for the length of this curve to derive the Bonnet angle [4] of the gyroid.) In the summer of 1966, I began calling the intertwined pair of labyrinth graphs in a TPMS skeletal graphs. I believed that there must exist other examples of TPMS besides Schwarzs D, P, H and CLP surfaces [4], but I did not undertake a very systematic search. In July, I began to suspect that there exists a TPMS I named L (for Laves), with two intertwined labyrinths that are essentially swollen versions of enantiomorphic Laves graphs [14 17]. I later changed its name to gyroid, or G, which is what I will call it here. I believed in the existence of G partly because its skeletal graphslike the sc and diamond skeletal graphs of P and Dwould be symmetric, i.e. both vertex-transitive and edge-transitive. Even now I know of no other examples of dual pairs of symmetric graphs on a cubic lattice. The genus of G would be three, as it is for D and P, and the space lattice of G would be bcc, implying that there is a TPMS of genus three for each of the three cubic lattices (the space lattices of D and P are fcc and sc, respectively). Schwarz demonstrated how to derive the Weierstrass integrals that dene the coordinates of sufciently simple symmetrical minimal surfaces bounded by either straight lines or plane geodesics (or both) [4]. Because it was impossible for G to contain either straight lines or plane geodesics, I had no idea how to construct it. It didnt occur to me that the key to the gyroid problem was Ossian Bonnets associate surface transformation. Bonnet proved in 1853 [34] that every simply-connected minimal surface S can be bent in such a way that (a) the orientation of the tangent plane at every point is unchanged, (b) the Gaussian curvature at every point is unchanged, and (c) the mean curvature at every point remains zero.
Interface Focus (2012)

D and P are examples of adjoint minimal surfaces. Plane geodesics in one surface of an adjoint pair S1 and S2 correspond to straight lines, orthogonal to that plane, in the other surface. If a point O common to S1 and S2 is xed, and r1 and r2 are corresponding points of S1 and S2, then r*(u), the image of r1 and r2 under bending, is given by r*(u) r1 cos u r2 sin u. Hence the points r1, r2, and r*(u) lie on an ellipse centered at O. For S1 and S2, u 0 and p/2, respectively. Every surface produced by Bonnet bending is called associate to S1 and S2 and is parameterized by the Bonnet angle u. If S1 D, S2 P, and the Bonnet angle uG cot21 (K0 [1/2]/ K[1/2]) 38.0147748, the resulting intersection-free associate surface has all of the properties I had anticipated for G. Figure 8 shows ellipses through sets of corresponding points on D, P, and G. These ellipses have the same eccentricity. In May, 1968, I made an incomplete survey of the regular and uniform tilings on D, G, and P by straight-edged skew polygons. (In unpublished work, Norman Johnson later lled in the gaps in my inventory.) Because I was beginning to receive not-so-subtle pressure from NASA headquarters to do something useful, I decided to apply my analysis of these tilings to the design of expandable spaceframes, including one based on the Laves graph. I spent the next two months developing these designs, writing NASA patent applications, andwith the assistance of Charles Strauss and Bob Davismaking computeranimated movies of what I called the collapse transformation for several examples of triply-periodic graphs, including the Laves graph [6]. Let us consider the kinematics of the collapse transformation applied to the innite Laves graph. At rst we regard the graph as embedded in Schwarzs D surface, with adjacent vertices of the graph at the centres of adjacent skew hexagonal faces of D, as in gure 9a. The initial directions of the vertex displacements are along perpendiculars to D, adjacent vertices moving oppositely with respect to the two sides of D. But it is convenient instead to describe all of the vertex trajectories relative to the position of a single xed vertex at the origin O. Now the initial directions of the vertex displacements are as shown in gure 9b. The trajectory of every vertex V is an ellipse centred at the origin. The three vertices nearest the origin a,b, and c in gure 10rotate on circular arcs in orthogonal coordinate planes. Figure 11 shows the circular

Downloaded from rsfs.royalsocietypublishing.org on January 10, 2013

Discussion. Triply-periodic minimal surfaces


G1 P1 G2 D2 D3 G6 P2 G3 O P2 D6 D5 D4 P3 G4 P4 G5 D4 G4 P3 P5 G3 P6 D1 G2 P1 D 2 D3 G6 O P5 G1 P6 D1

A. H. Schoen

663

D6

D5 P4

G5

Figure 8. (stereo image) Ellipses, all of the same eccentricity, through corresponding points on D, G and P. There is a xed point O at the centre of the hexagonal patches.

(a) d O b c b a d O c a

(b) d a d a O c b c

O b

Figure 9. (a; stereo image) Directions of initial displacements of vertices of the Laves graph embedded in Schwarzs D surface. (b; stereo image) Directions of initial displacements of vertices of the Laves graph embedded in Schwarzs D surface, with one vertex ( yellow) xed at the origin.

trajectories of vertices a, b, and c of gure 10 as well as the elliptical trajectories of three vertices shown in gure 9 that are farther from the origin than a, b and c are. Every vertex in the graph belongs to one of four classes1, 2, 3, or 4according to whether it is related by a translational symmetry of the graph to the vertex a, b, c, or O. In gure 10, the points A, B, C, and O, with coordinates rA, rB, rC, and (0,0,0), respectively, lie at the corners of the regular tetrahedron ABCO. Every vertex of the graph is mapped onto one of
Interface Focus (2012)

the four vertices of ABCO. If a vertex is in class 1, with initial position r1, its collapse trajectory is r r1 cos u rA sin u. The collapse trajectories of vertices in class 2 and 3, with initial positions r2 and r3, are r r2 cos u rB sin u and r r3 cos u rC sin u, respectively. The trajectory of a vertex in class 4, with initial position r4, is along the line through O, r r4 cos u. For a vertex V with initial position (x,y,z), the major radius of its trajectory ellipse is equal to j(x,y,z)j. If V is in class 1, 2, or 3, the minor radius of the ellipse is equal

Downloaded from rsfs.royalsocietypublishing.org on January 10, 2013

664

Discussion. Triply-periodic minimal surfaces


d a

A. H. Schoen
d a

b C, D

O b A C, D c

O A c

Figure 10. Collapse trajectories of vertices a ! A, b ! B, c ! C, d ! D in the Laves graph for rotation angle u [ [0, p p/2]. The vertices a, b and c move on circular paths. The vertex d, for example, moves on an ellipse of eccentricity 2=3 (stereo image).

(a)

(b)

Figure 11. (a,b) Circular and elliptical collapse trajectories of Laves graph vertices for rotation angles u [ [0, p/2].

to the edge length of the graph, and if V is in class 4, the minor radius is equal to zero. Consequently, the ellipse eccentricities are not all equal, unlike the eccentricities of the elliptical trajectories in the associate surface trnsformation. The one-to-one mapping of points in the associate surface transformation is unrelated to the mapping of vertices in the collapse transformation, which is many-to-one. In June 1966, in the course of my campaign to test the robustness of the dual graph recipe, I dened a symmetric graph on a given set of vertices as defective if not all pairs of nearest-neighbor vertices are joined by an edge [6]. Aside from an uninteresting 3-valent graph on the vertices of the sc lattice, the only example of a defective symmetric graph I investigated on that occasion was FCC6, a 6-valent graph on the vertices of the fcc lattice. The duality recipe yielded the
Interface Focus (2012)

interstitial and nodal polyhedra shown in gures 12 and 13 [6] (http://www.schoengeometry.com/e_tpms. html). If an innite set of the interstitial polyhedra for this graph is arranged so that every adjacent pair of polyhedra share quadrangular faces, the hexagonal faces dene the P surface, and if every adjacent pair of polyhedra share hexagonal faces, the quadrangular faces dene the D surface. In July 1967, at the invitation of the physicist Lester Van Atta, I joined the staff of the NASA Electronics Research Center (ERC) in Cambridge, MA, where he was Associate Director. My responsibilities were not very clearly dened. Van Atta told me just to follow my nose. But to give my position some bureaucratic heft, he created for me the Ofce of Geometrical Applications. I understood, of course, that I was expected eventually to produce something NASA might regard as useful to its mission. Exploiting the vagueness of my job description, however, I resumed what I had barely begun a year earlierexploring the connections among polyhedra, triply-periodic graphs, and minimal surfaces. Van Atta provided generous support for my research and never tried to inuence my choices of what to work on. On 14 February 1968, it occurred to me that when I constructed the FCC6 graph on the vertices of the fcc lattice a year earlier, I forgot to devise a defective symmetric graph on the vertices of the bcc lattice. I quickly discovered BCC6, a symmetric graph of degree six, which turned out to provide the long anticipated counterexample to the duality recipe [6]. The breakdown in the recipe occurred at the rst step. When each quadrangular interstice of BCC6 is spanned by a minimal surface, the resulting interstitial polyhedron is an innite saddle polyhedron that I call M4 (cf. gure 14). Its skeletal graphs are enantiomorphic Laves graphs. Weeks later I designed an improvised nodal polyhedron [5,6] (http://www.schoengeometry.com/e_tpms.html) for BCC6 (cf. gure 15).

Downloaded from rsfs.royalsocietypublishing.org on January 10, 2013

Discussion. Triply-periodic minimal surfaces

A. H. Schoen

665

Figure 14. M4, an innite warped polyhedron (stereo image). Figure 12. Doubly expanded tetrahedron, interstitial polyhedron of the FCC6 graph (stereo image).

Figure 13. FCC6 pinwheel polyhedron, nodal polyhedron of the FCC6 graph (stereo image).

Figure 15. BCC6 pinwheel polyhedron, improvised nodal polyhedron of BCC6 graph (stereo image).

It seemed plausible to me that M4 could somehow be transformed into the gyroid TPMS I had imagined in 1966. If the dihedral angle w at an edge shared by two congruent skew polygons is equal to p, as in the case of the skew hexagons in Schwarzs D surface (cf. gure 6a), the angle decit d p 2 w is zero, but if w , p, d is positive. I judged that d would be smaller in M6the dual of M4than in M4, and a calculation conrmed that it is: for M4, d p/3, and for M6, d p 2 cos21(5/7) 44.4158. On 1 April 1968, after constructing a physical model of M6 (cf. gure 16), I decided to replace the innite regular skew helical polygons that wind around the outside of the tunnels parallel to the cube axes of M6 by helices, believing that M6 might thereby be transformed into a TPMS. The straight edges of each hexagonal face of M6 would then become helical arcs, each of one-quarter pitch, in a sequence of alternating handedness. When I constructed a plastic model of this modied surface (G in gure 6b), I found that it did indeed appear to resemble a TPMS, but I had no idea how to prove that it is one. I telephoned Bob Osserman and explained my predicament: I had a hexagonal surface patch, bounded by helical arcs, that looked like a plausible candidate for the unit patch of an embedded TPMS, but I had no idea how to solve the equations for the patch. After I sent him a model of G, he asked his PhD student Blaine Lawson to investigate. I introduced myself to Blaine by telephone, told him everything I knew about the surface and decided to wait for him to prove that G is an embedded TPMS. At the end of August, 1968, I returned to Cambridge from an AMS summer meeting in Madison, where I had given a talk about the gyroid. When I submitted my abstract [5] for this talk a few months earlier, I naively assumed that Blaine would surely complete his proof by August. When I telephoned him early in September to
Interface Focus (2012)

ask whether he had nished the proof, he replied that he was too busy nishing his dissertation and would be unable to devote any more time to my problem. I begged him not to give up, saying that I felt mysteriously condent that the gyroid is a minimal surface and that the proof must be right around the corneror words to that effect (even though I hadnt the slightest idea how to carry it out!). Then I abruptly changed the subject and described what had been my principal concern for the previous several monthsinvestigating the values of the skewness of the faces, vertex gures, and holes of the Coxeter maps {6,4|4}, {4,6|4}, and {6,6|3} on G and comparing them to their values on D and P. I explained how this analysis [5] had led me to design expandable spaceframes based on the geometry of collapsing graphs [35]. I had not previously even mentioned this subject to Blaine. I tried, clumsily, to explain what I called an amusing coincidence: both the vertices in the collapsing graph transformation and also points on Schwarzs D and P surfaces subjected to the associate surface transformation move on elliptical trajectories, although the ellipses in the two transformations are not related. Suddenlyin what was to become a Eureka momentBlaine asked me if I was saying that the gyroid is associate to Schwarzs D and P surfaces. I had not said that, because this quite obvious idea had not crossed my mind. But I immediately became greatly excited, because I recognized that Blaine had pointed to the solution of the problem! For two months I had been blindly applying truncations and other algebraically complicated operations to the three Coxeter maps on D, P, and G, without stopping to think about why these maps match the combinatorial structure of all three surfaces, why the tangent planes at corresponding

Downloaded from rsfs.royalsocietypublishing.org on January 10, 2013

666

Discussion. Triply-periodic minimal surfaces

A. H. Schoen

Figure 16. M6, dual of M4 viewed from near [100] axis (stereo image).

Figure 17. One-fourth of a lattice fundamental domain of the minimal surface C(D) (genus 19).

points of these surfaces are parallel, why the surfaces have the same genus, etc. Now I understood for the rst time that the spiral curves in each lattice fundamental domain of G are simply intermediate curves in the morphing of the round waists of P into the straight lines of D. It was at last clear that these spiral curves could not be the perfect helices I had naively imagined them to be, since their pre-imagesthe round waists of Pare not quite perfect circles. I soon used computeranimated stereoscopic images to support these conclusions and also to demonstrate that G is the only embedded surface associate to D and P. Eighteen years later, Karsten Grosse-Brauckmann and Meinhard Wohlgemuth published a rigorous proof that the gyroid is embedded [36]. I immediately proposed to Blaine that we publish a joint paper announcing the gyroid, but he declined, explaining that he had done no more than misunderstand what I said about the ellipses of the collapse transformation and the ellipses of the associate surface transformation. When I insisted that it was impossible to know how long it would have taken me to discover the facts by myself if he had not asked that crucial question, Blaine reluctantly agreed to co-publish. Two days later, when Dr Van Atta returned to ERC from an out-of-town trip, I told him the exciting news. To my surprise, he became angry and insisted that I telephone Blaine, explain that I had made a serious mistake and publish alone. Reluctantly, I acceded to his demand. In the spring of 1967, I was puzzled by a reference on p. 271 of Hilbert and Cohn-Vossens Geometry and the Imagination [32] to a TPMS with the symmetry of the diamond structure, discovered by E. R. Neovius, Schwarzs doctoral student. In 1969, when at last I examined Neoviuss dissertation, I was startled to see a drawing of a lattice fundamental domain of an embedded TPMS of genus nine [37] in which the set of embedded straight lines is exactly the same as the set in Schwarzs P surface! When I inspected my model of the D surface, I conrmed that it too has an embedded complement of genus nineteenthat I named C(D) (http://www. schoengeometry.com/e_tpms.html) [38]. It is shown in gure 17. In 1970, I conjectured from soap lm experiments inside transparent polyhedral cells that both D and P have unbounded arithmetic sequences of complements of increasing genus and that complements exist for other TPMS as well [6]. Ken Brakkes experiments with his SURFACE EVOLVER [1,39], beginning in 1999, provide qualied support for these conjectures.
Interface Focus (2012)

While studying a 1934 paper by Stessmann [40], I conjectured the existence of several intersection-free TPMS, bounded by plane geodesics, that contain no embedded straight lines. The elementary patch of each of these surfaces, which include I-WP and F-RD, is a simplyconnected surface in a stationary state inside a Coxeter cell [6], derived from its adjoint surface. In 1989, Karcher constructed a rigorous proof of the existence of these surfaces and derived Weierstrass parametrizations for many of them [41]. Fogden and Hyde independently derived the Weierstrass parametrizations for these and other examples of TPMS [4244]. In 1969, I invented a technique for grafting handles onto TPMS [6] (http://www.schoengeometry.com/ e_tpms.html). It is based on the construction of an embedded surface g that is a hybrid of two known examples, a and b, of embedded surfaces. First, one constructs the surface g, which is a linear combination of the straight-edged polygons a and b, the adjoints of a and b. The surface g is then derived as the adjoint of the adjoint: g ( g). Translational periodicity for g and the absence of self-intersections in g are achieved by properly adjusting relative weights for a and b, a technique that came to be called killing periods. The rst example of a hybrid surface for which I constructed a physical model was the surface O,C-TO [6], an unattractive genus 10 hybrid of P and I-WP. While preparing this article, I found in my les a long-forgotten 1969 sketch of a genus 14 hybrid of P and Neoviuss C(P) [6,37], the rst example of a conjectured hybrid surface. Not long after I circulated that sketch among mathematicians in 1969, grafting handles onto minimal surfacesand not just those of the periodic variety became rather fashionable. Many attractive surfaces were devised by handle grafting. Richard Schoen (to whom I am not closely related!) proved a famous NoGo theorem that asserts the impossibility of grafting a horizontal handle inside a catenoid that is coaxial with a vertical line (cf. Karchers interactive applet [45], which beautifully illustrates Schoens theorem). The P C(P) hybrid I sketched in 1969 nally sprang to life recently when I sent a copy of the original hybrid proposal to Ken Brakke. Ken used his SURFACE EVOLVER program [1] to derive a lattice fundamental domain (cf. gure 18), and he named the surface N14. In 1968, with the collaboration of my colleague, the sculptor/model-maker Harald Robinson, I developed a technique that uses a laser to measure the orientation of the surface normal at points near the boundary of a

Downloaded from rsfs.royalsocietypublishing.org on January 10, 2013

Discussion. Triply-periodic minimal surfaces

A. H. Schoen

667

U.S. Senator. In September 1969, thanks to the kindness of Robert Osserman and Lipman Bers, I participated in an international conference in Tbilisi, Georgia, USSR on Optimal Control Theory, Partial Differential Equations and Minimal Surfaces (http:// www.schoengeometry.com/e_tpms.html). In 1972 and 1974, I made two black-and-white videos on the subject of TPMS. I edited these videos in 1999, and I expect to make them available soon.
Figure 18. Views of Ken Brakkes SURFACE EVOLVER solution for two lattice fundamental domains of N14, a genus 14 hybrid of P and C(P) (images courtesy of Ken Brakke). This article is dedicated to the memory of Alexander F. Jumbo Wells, without whose inspiration I would almost certainly not have embarked on my journey. I cherish my memories of a day spent with Wells during a 1968 visit to his home in Storrs, Connecticut. Jim Tanaka has written an oral history of Wells life and work, which can be found at http:// schoengeometry.com/a_f_wells_oral_history.pdf. I am grateful to an anonymous referee for suggesting improvements of this article. Note: All of the stereoscopic image pairs shown above are arranged for cross-eyed viewing. If they are viewed with a stereoscope, the left and right images should be exchanged.

long-lasting polyoxyethylene lm spanned by a straightedged skew polygon. I used this data to derive the approximate shape of the plane geodesic boundary curves of a surface patch of the adjoint TPMS, which has no embedded straight lines. In many of these examples, it was necessary to kill periods in order to determine the relative lengths of the edges of the patch. Fortunately, the SURFACE EVOLVER program [1] has made this tedious experimental procedure obsolete. In 1999, Ken Brakke and I began a collaboration to validate examples of these TPMS, whose existence I had conjectured between 1969 and 1972 (http://www. schoengeometry.com/e_tpms.html). In my last eighteen months at NASA/ERC, I became aware of several possible scientic applications of TPMS. A literature search in 1966 revealed that the structure of the prolamellar body for some etiolated green plants [46,47] invites comparison with Schwarzs P surface. In 1969, I conferred twice with the Harvard biologists Lawrence Bogorad and Christopher Woodcock, who are experts in this eld. When I telephoned the biophysicist Donald Caspar [48] in 1969 to inquire whether he knew of any chemical compounds with space group Ia3d (the space group of the gyroid), he instantly referred me to Polymorphism of lipids, a 1966 paper by Luzzati and Spegt [49] on the structure of a high-temperature phase of divalent cation soaps. That was the rst hint I had of the existence of crystalline matter that incorporates the geometry of an enantiomorphic pair of Laves graphs. The rst signicant report of the gyroid in liquid crystals was an article by Stephen Hyde et al. [50]. Additional recommended readings are in [51 53]. In lectures between 1969 and 1975, I described several potential scientic applications of TPMS, but I never published anything about them. In 1969, Arthur Drexler, Director of Design at the N.Y. Museum of Modern Art, commissioned me to build a large sculpture of the gyroid for a 1970 exhibition at MOMA. Dr. Van Atta obtained funds for the project from NASA Headquarters, but President Nixon closed NASA/ERC just a few months later, and the project was terminated very soon after Jim Wixson and I began work on the CAD/CAM phase of the project. In 1971, I was asked by the NSF College Science Curriculum Improvement Program to design a packaged course about TPMS, graphs, and polyhedra, but after NSF approved my preliminary proposal, funding for the entire CSCIP was cancelled by an inuential
Interface Focus (2012)

REFERENCES
1 Brakke, K. 1996 The surface evolver and the stability of liquid surfaces. Phil. Trans. R. Soc. Lond. A 354, 21432157. (doi:10.1098/rsta.1996.0095) z dz , W. & Holyst, R. 1996 High genus gyroid surfaces of 2 Go non-positive Gaussian curvature Phys. Rev. Lett. 76, 2726. (doi:10.1103/PhysRevLett.76.2726) 3 Lord, E. A. & Mackay, A. L. 2003 Periodic minimal surfaces of cubic symmetry. Curr. Sci. 85, 346362. (http://materials.iisc.ernet.in/~lord/webles/tpmbs.pdf.) 4 Schwarz, H. A. 1970 Gesammelte mathematische Abhandlungen, vol. I, AMS. Berlin, Germany: Springer. 5 Schoen, A. H. 1968 Innite Regular Warped Polyhedra (IRWP) and Innite Periodic Minimal Surfaces (IPMS). Abstract 658 30, Amer. Math. Soc. 26 30. 6 Schoen, A. H. 1970 Innite periodic surfaces without selfintersections. NASA TN D-5541. Springeld, VA: Federal Scientic and Technical Information. 7 Schoen, A. H. 1959 Correlation and the isotope effect for diffusion in crystalline solids. Phys. Rev. Lett. 1, 138140. (doi:10.1103/PhysRevLett.1.138) 8 Tharmalingam, K. & Lidiard, A. B. 1959 Isotope effect in vacancy diffusion. Phil. Mag. 8, 899906. (doi:10.1080/ 14786435908238264) 9 Bardeen, J. & Hering, C. 1951 Atom movements, p. 87. Cleveland, OH: The American Society for Metals. 10 Peterson, N. L. 1964 Isotope effect in self-diffusion in palladium. Phys. Rev. 136, A568A574. (doi:10.1103/ PhysRev.136.A568) 11 Pell, E. M. 1960 Diffusion of Li in Si at high T and the isotope effect. Phys. Rev. 119, 10141021. (doi:10.1103/ PhysRev.119.1014) 12 Schoen, A. H. & Lowen, R. W. 1960 Evaluation of Bardeen-Hering correlation factors. Bull. Am. Phys. Soc. 4, 280. 13 Wells, A. F. 1956 Third dimension in chemistry. Oxford, UK: Oxford University Press. 14 Laves, F. 1932 Zur Klassikation der Silikate. Z. Kristallogr. 82, 114. ber du 15 Heesch, H., Laves, F. & Kristallogr, Z. 1933 U nne kugelpackungen. Z. Kristallogr. 85, 443 453. 16 Coxeter, H. S. M. 1955 On Laves graph of girth ten. Can. J. Math. Soc. 7, 1823. (doi:10.4153/CJM-1955003-7)

Downloaded from rsfs.royalsocietypublishing.org on January 10, 2013

668

Discussion. Triply-periodic minimal surfaces

A. H. Schoen
39 Brakke, K. Triply periodic minimal surfaces. See http:// www.susqu.edu/brakke/evolver/examples/periodic/perio dic.html. 40 Stessmann, B. 1934 Periodische minimala chen. Mathematische Zeitschrift. 38, 417442. (doi:10.1007/ BF01170644) 41 Karcher, H. 1989 The triply periodic minimal surfaces of Alan Schoen and their constant mean curvature companions. Manuscripta Math. 64, 291357. (doi:10.1007/ BF01165824) 42 Fogden, A. & Hyde, S. T. 1992 Parametrization of triply periodic minimal surfaces. I. Mathematical basis of the construction algorithm for the regular class. Acta Cryst. A48, 442451. (doi:10.1107/S0108767 391015167) 43 Fogden, A. & Hyde, S. T. 1992 Parametrization of triply periodic minimal surfaces. II. Regular class solutions. Acta Cryst. A48, 575591. (doi:10.1107/S01087673 92002885) 44 Fogden, A. & Hyde, S. T. 1993 Parametrization of triply periodic minimal surfaces. III. General algorithm and specic examples for the irregular class. Acta Cryst. A49, 409421. (doi:10.1107/S0108767392010456) 45 Karcher, H. Schoen No-Go theorem. See http://virtual mathmuseum.org/Surface/schoen_no-go_thm/schoen_nogo_thm.html. 46 Gunning, B. E. S. & Jagoe, M. P. 1965 The prolamellar body, biochemistry of the chloroplasts, vol. 2, pp. 655676. London, UK: Academic Press. 47 Gunning, B. E. S. 2001 Membrane geometry of open prolamellar bodies. Protoplasma 215, 4 15. (doi:10.1007/ BF01280299) 48 Caspar, D. See http://en.wikipedia.org/wiki/Donald_ Caspar. 49 Luzzati, V. & Spegt, P. A. 1967 Polymorphism of lipids. Nature 215, 701704. (doi:10.1038/215701a0) 50 Hyde, S. T., Andersson, S., Ericsson, B., Larsson, K. & Kristallogr, Z. 1984 A cubic structure consisting of a lipid bilayer forming an innite periodic minimum surface of the gyroid type in the glycerolmonooleate water system. Z. Kristallogr. 168, 213219. (doi:10.1524/zkri. 1984.168.1-4.213) 51 Hyde, S. T., OKeeffe, M. & Proserpio, D. M. 2008 A short history of an elusive yet ubiquitous structure in chemistry, material, and mathematics. Angew. Chem. Int. Ed. 47, 79968000. (doi:10.1002/anie.200801519) 52 Hyde, S. T., Andersson, S., Ericsson, B. & Larsson, K. 1984 A cubic structure consisting of a lipid bilayer forming an innite periodic minimal surface of the gyroid type in the glycerolmonooleate water system. Zeit. Kristallogr. 168, 213219. (doi:10.1524/zkri.1984.168.1-4.213) 53 Dierkes, U., Hildebrandt, S. & Sauvigny, F. Second edition 2010, Minimal Surfaces. Springer, 708 p., 140 illus. See http://www.springer.com/mathematics/book/978-3-64211697-1.

17 Sunada, T. Crystals that nature might miss creating. See http://www.ams.org/notices/200802/tx080200208p.pdf. 18 Wells, A. F. 1954 The geometrical basis of crystal chemistry. Acta Crystallogr. 7, 849 853. (doi:10.1107/ S0365110X54002587) 19 Coxeter, H. S. M. 1973 Regular polytopes, 3rd edn. New York, NY: Dover Publications. 20 Rouse Ball, W. W. & Coxeter, H. S. M. 1987 Mathematical recreations and essays. New York, NY: Dover Publications Inc. 21 Wachsmann, K. Short life history: Konrad Wachsmann. See http://www.einstein-website.de/biographies/wachs mann_content.html. 22 Pearce, P. J. 1990 Structure in nature is a strategy for design. Cambridge, MA: The MIT Press. 23 Pearce, P. J. See http://www.pjpearcedesign.com/. 24 Charles, E. & Ray, E. See http://en.wikipedia.org/wiki/ Charles_and_Ray_Eames. 25 Gamelin, T. W. 2005 Raymond Redheffer, Professor of Mathematics, Emeritus, Los Angeles 1921 2005. See http://www.universityofcalifornia.edu/senate/ inmemoriam/raymon dredheffer.htm. 26 Delgado-Friedrichs, O., OKeeffe, M. & Yaghi, O. M. 2003 Three-periodic nets and tillings: regular and quasiregular nets. Acta Cryst. A59, 22 27. 27 Bonneau, C., Delgado-Friedrichs, O., OKeeffe, M. & Yaghi, O. M. 2004 Three-periodic nets and tillings: minimal nets. Acta Cryst. A60, 517 520. 28 Coxeter, H. S. M. 1937 Regular skew polyhedra in three and four dimensions and their topological analogues. Proc. Lond. Math. Soc. 2, 43. 29 Rowland, T. Schwarz reection principle, Wolfram MathWorld. See http://mathworld.wolfram.com/Schwarz ReectionPrinciple.html. 30 Johnson, N. See http://en.wikipedia.org/wiki/Norman_ Johnson_(mathematician). 31 Eisenhart, L. P. 1960 A treatise on the differential geometry of curves and surfaces. New York, NY: Dover Publications. 32 Hilbert, D. & Cohn-Vossen, S. 1952 Geometry and the imagination. New York, NY: Chelsea Publishing Co. 33 Bjo r Mathematik und Physik, rling, E. G. 1844 Archiv fu IV. Greifswald, Germany: C.A. Koch. 34 Bonnet, O. 1853 Sur la theorie generale des surfaces. Comptes Rendus. 37, 529 532. 35 Schoen, A. H. et al. 1973 Expandable space-frames. US Patent 3,757,476, 11 September. 36 Grosse-Brauckmann, K. & Wohlgemuth, M. 1996 The gyroid is embedded and has constant mean curvature companions. Cal. Var. Partial Differ. Equ 4, 499 523. (doi:10. 1007/BF01261761) 37 Neovius, E. R. 1883 Bestimmung zweier speziellen periodischen Minimalachen. Helsingfors, Finland: Akad. Abhandlung. 38 Schoen, A. H. 1969 A fth intersection-free innite periodic minimal surface [IPMS] of cubic symmetry. Abstract 664 664. Amer. Math. Soc.

Interface Focus (2012)

You might also like