<<

arXiv:2108.07604v2 [math.DS] 30 Aug 2021 n fteemp edntb convex. be not need maps these of one lasdfie nconvex on defined always n ed u ovxt sipratfru ee h maps The here. us for important is convexity but field, any ie an Given h successive the . ojcueadResult Let and Conjecture 1.1 Introduction 1 noecncscinadwoeegslei ie agn oaohrcon another to tangent lines in lie edges whose and section section. conic one in nte lis.Mr eeal,a generally, More . another t mg ne utbepoetv rnfraini ovxpoly convex a is of transformation patch affine projective standard suitable the a under image its ∗ e ( Let upre yNSF rn DMS-2102802 Grant N.S.F. by Supported RP T etokCs fPnarmRigidity of Case Textbook A hscs novsatxbo nlsso eclo ellipti of pencil symm a rotational of invol analysis 4-fold one textbook with wi a the 8-gon I involves convex conjecture, case a This the . on of Poncelet map case 3- simple and a maps establish pentagram also diagonal deep the k sgnrclydfie n netbe h aecntuto ok in works construction same The invertible. and defined generically is ,k n, 2 nti ae ilepanargdt ojcueta inter that conjecture rigidity a explain will I paper this In Poncelet n eoetera rjciepae oyo in polygon A . projective real the denote -gon eapi fitgr,ntbt vn with even, both not integers, of pair a be ) k daoasof - P 1 elet we , fi sisrbdi n lis n icmcie about circumscribed and ellipse one in inscribed is it if ihr vnSchwartz Evan Richard n P etme ,2021 1, September gn,tog h mg faconvex a of image the though -gons, 2 RP P = 1 iue1sosti o ( for this shows 1 Figure . T 2 k Abstract See . ( ocltpolygon Poncelet P 1 ethe be ) 1 § . o entos ecl convex a call We definitions. for 2.1 n gnotie yintersecting by obtained -gon soewoevrie lie vertices whose one is ∗ n ≥ ,k n, and 7 RP curves. c T (8 = ) k igthe ving 2 k and n twines is gnunder -gon ∈ etry. convex , (2 T ) The 3). ll k o in gon n/ , − 1 are 2). ic if Figure 1: P1 and P2 = T3(P1).

Figure 1 gives an example of where P1 is convex but P2 is not. Starting j with P0 we define the (n, k)-pentagram orbit Pj where Pj = T (P0). { } k

Conjecture 1.1 Suppose that Pj is an (n, k)-pentagram orbit and that Pj { } is convex for all j Z. Then Pj is convex Poncelet for all j Z. ∈ { } ∈

I proved in [15] that if P0 is a Poncelet polygon, T (P0) and P0 are pro- jectively equivalent. Thus, to prove the conjecture it is enough to prove that the hypotheses force P0 to be convex Poncelet. In this paper I will prove a simple but nontrivial case of the conjecture.

Theorem 1.2 Suppose that P0 is an 8-gon with 4-fold rotational symmetry and Pj is the (8, 3)-pentagram orbit. Then P0 is regular if and only if Pj { } is convex for all j. More precisely

1. P0 has 8-fold dihedral symmetry, with symmetry lines containing the vertices, if and only if Pj is convex for all j 0. ≥

2. P0 has 8-fold dihedral symmetry, with symmetry lines bisecting the edges, if and only if Pj is convex for all j 0. ≤ A version of Theorem 1.2 appears to be true when (8, 3) is replaced by a general pair (n, k) when n is even and k is odd and the n-gons have (n/2)-fold rotational symmetry. The proof should be similar. A much more interesting generalization is to the case of centrally symmetric octagons, because these include all Poncelet octagons. In the last chapter of this paper, I will discuss the progress I have made (mostly experimental so far) towards proving the conjecture in the case of centrally symmetric octagons.

2 The conjecture is not true for n and k both even. In this case, Tk is 2- periodic (modulo scaling) when restricted to the space of n-gons with (n/2)- fold rotational symmetry, and Tk has some non-regular convex fixed points modulo scale.

1.2 Context The classic case of the pentagram map is (n, 2) for n 5. The case n = 5 has been studied e.g. by Clebsch in the 19th century and≥ Motzkin [9] in middle of the 20th century. In 1992 I wrote a paper [13] defining the pentagram map for general n-gons and proving in the convex case that the forward orbit shrinks to a point. Very recently, M. Glick [2] found a kind of formula for this collapse point. It is nice to consider the map T2 as defined on the space n of n-gons mod- P ulo projective transformation. With the correct labeling, T2 is the identity on 5 and 2-periodic on 6. I observed experimentally that the orbits of T2 in generalP seem to lie onP tori. Motivated (for some reason) by the scattering transform for the KdV equation, I found [14] about n algebraically inde- pendent invariants for T2. These invariants are now called the invariants or the pentagram integrals. In [10], V. Ovsienko, S. Tabachnikov, and I showed that T2 has an invari- ant Poisson structure of corank 2 in the odd case and corank 4 in the even case, and that the monodromy invariants Poisson-commute with respect to this structure. This established the complete Arnold-Liouville integrability of the pentagam map on the larger space n of so-called twisted n-gons. Es- T sentially what this means is that the space n, a space of 2n, has a singular by of dimensionT about n such that the restriction of P2 to each is a translation in suitable coordinates. When these manifolds are compact they are necessarily finite unions of tori. Subsequently, we proved in [11] that T2 is Arnold-Liouville integrable on n, which is naturally a codimension 8 subvariety of n. For the subset P T of convex n-gons, the manifolds in the singular foliation are compact and hence finite unions of tori. At the same time, F. Soloviev [18] proved that the pentagram map is algebro-geometrically integrable on n. This implies in particular that the foliation discussed above is naturallyP an abelian fibration, with the individual tori having natural desciptions as Jacobian va- rieties for certain Riemann surfaces. Very recently, M. Weinreich [19] proved that T2 is algebro-geometrically integrable in any field of characteristic not

3 equal to 2. In [3], M. Glick related the pentagram map to a cluster algebra. By now there are many generalizations of the pentagram map, and also a number of ways to generate invariant functions and the invariant Poisson structure. In [1], M. Gekhtman, M. Shapiro, S. Tabachnikov, A. Vainshtein generalized the pentagram map to similar maps using longer diagonals, and defined on spaces of so-called corrugated polygons in higher . The work in [1] also generalizes Glick’s cluster algebra and establishes the com- plete integrability of these maps in some form. In [7], G. Mari-Beffa defines higher dimensional generalizations of the pentagram map and relates their continuous limits to various families of integrable PDEs. See also [8]. In [5], B. Khesin and F. Soloviev obtain definitive results about higher dimensional analogues of the pentagram map, their integrability, and their connection to KdV-type equations. The little survey above is not meant to be complete. Now let me explain how these various results are related to the Pentagram Rigidity Conjecture above. First of all, the map Tk is the one used in [1]. It would be nice if one could conclude from [1] that Tk is completely integrable on n, but this has not been directly worked out. The spaces of corrugated polygP ons are somewhat different than spaces of ordinary polygons, though in some sense closely related. Ordinary polygons are limits of corrugated polygons under a kind of flattening operation. Let me just leave it by saying that Tk is certainly believed to be completely integrable on n in some sense. The Pentagram Rigidity Conjecture is reallyP about the global geometry of the torus foliation of n (presumably) associated to Tk. The smaller space P n of convex n-gons modulo projective transformations is a subset of n. For C P k [3, n/2) the tori in this foliation probably are not contained in n. So, if ∈ C Tk is not the identity on one of these tori, and moreover the intersection of the torus with n is not too large, then the orbit of Tk on this torus cannot stay C in n. (See Lemma 2.1 below.) This observation would prove the conjecture C for the n-gons corresponding to this torus. The Poncelet polygons in n also lie on these tori, but T is the identity there. C Motivated by the Pentagram Rigidity Conjecture, A. Izosimov [4] has recently proved that if n is odd and P n is a fixed point of T2 then in fact ∈ C P is a Poncelet polygon. So, we now can say that for n odd points in n are C fixed by T2 if and only if they are Poncelet. The convexity is important here. 2 Izosimov gave some easy examples of n-gons in CP that are fixed by T2 but not Poncelet. The parity of n is also important. As I mentioned above, the result is not true when n is even. Presumably, Izosimov’s result would also

4 work for general pairs (n, k) where both numbers are not even. In the case I consider, that of T3 acting on 8-gons with 4-fold rotational symmetry, there is just a single invariant for the map, and its level curves are nonsingular elliptic curves except when they contain points corresponding to octagons having 8-fold dihedral symmetery. These are the curves I analyze in §2.2. These elliptic indeed stretch outside 8 in the appropriate sense, and this is enough to prove Theorem 1.2. See FigureC 2 in §3.4 The various invariant-generating machines for the pentagram map prob- ably would turn up the invariant I found, but these machines are better developed for T2 than they are for T3. I just guessed the invariant for T3 by looking at the picture, and then checked algebraically that it works. I will explain in §3.2 what led me to the invariant. There are two other connections I want to make between the Penta- gram Rigidity Conjecture and other of . When I originally thought of this conjecture, about 30 years ago, I had imagined it as a projec- tive geometry analogue of the packing rigidity theorem [12] of B. Rodin and D. Sullivan. Much more recently, it occured to me that the conjecture is something like a discrete analogue of the Birkhoff-Poritsky Conjecture about billiards in strictly convex ovals. This conjecture says roughly that if a neigh- borhood of the boundary of the (cylindrical) billiard phase space is foliated by invariant curves (corresponding to caustics) then the oval is an ellipse.

1.3 Organization In §2 I will give some background information about and also analyze the family of elliptic curves that arises in the proof of Theorem 1.2. I will also present a few well-known results about complex tori. In §3 I give the proof of Theorem 1.2. In §4 I discuss the generalization of Theorem 1.2 to the set of centrally symmetric octagons. The interested reader can download the computer program I wrote, which does experiments with the 3-diagonal map on centrally symmetric octagons. The location of the program is http://www.math.brown.edu/ res/Java/OCTAGON.tar: ∼ 1.4 Acknowledgements I would like to thank Misha Bialy, Misha Gehktman, Anton Izosimov, Joe Silverman, Sergei Tabachnikov, and Max Weinreich for helpful conversations.

5 2 Preliminaries

2.1 Projective Geometry The real RP 2 is the space of lines through the origin in R3. Equivalently it is the space of scale-equivalence-classes of nonzero vectors in R3. Points in RP 2 will be denote by [x : y : z]. This point represents the through the origin and (x, y, z). The quotient map R3 RP 2 is called projectivization. → There is a natural inclusion R2 RP 2 given by → (x, y) [x : y : 1]. (1) → The image of this inclusion is known as the standard affine patch. I often identify R2 with its image under this inclusion, and when speaking about points in the affine patch I will often write (x, y) for [x : y : 1]. The inclusion in Equation 1 has an inverse, given by [x : y : z] (x/z,y/z). (2) → The line at infinity is the subset of RP 2 outside the standard affine patch. The line at infinity consists of points of the form [x : y : 0]. More generally, a line in RP 2 is the set of members represented by lines in a 2-dimensional subspace of R3. We can also represent lines by triples [a : b : c]. This point represents the linear subspace given by the equation ax + by + cz = 0. Conveniently, the line through 2 points is represented by the cross product of the corresponding vectors. Likewise, the intersection of 2 lines is given by the cross product of the corresponding vectors. These facts make computations with the pentagram map very easy. 2 2 The dual RP ∗ is the space of lines in RP . As our 2 2 notation suggests, there is an isomorphism between RP and RP ∗. The isomorphism sends the point represented by [u : v : w] to the line represented by [u : v : w]. This isomorphism sends collinear points to coincident lines. Note that all the same words apply with the field R replacing the field C. Thus CP 2 is the complex projective plane. A projective variety in CP 2 is the projectivization of the set V (x, y, z) = 0 where V (x, y, z) is a homogeneous polynomial in 3 variables. This variety is called nonsingular if the (formal) gradient V is everywhere nonzero on the set V (x, y, z) = 0. When V ∇ is cubic and nonsingular the corresponding projective variety is a smooth Riemann surface of genus 1, also known as a complex torus. See [17].

6 2.2 A Family of Cubics In this section we study the solutions to the equation

(x y)(x2 + y2 1) − − = λ, (x y)(x2 y2 1) λxy =0. (3) xy − − − − The second equation is a rearrangement of the first one. To bring this equa- tion into the form we mentioned at the end of the last section, we expand it out and then homogenize it by padding the z-variable. This gives us the equation

V (x, y, z)= x3 y3 x2y + xy2 xz2 + yz2 λxyz =0. (4) − − − − Let Eλ denote the complex projective variety corresponding to V = 0. Let ρ be reflection in the line y = x . Call this line L. Call a subset of RP 2 bounded if it lies in R2 and{ otherwise− } unbounded. Below, I will prove two results:

1. For all λ = 0, 2, 4i√2 the variety Eλ is nonsingular, and hence a complex torus.6 ± ±

2 2. When λ R 2, 0, 2 , the set Eλ RP consists of 2 smooth loops, both ρ-invariant,∈ −{− one bounded} and one∩ unbounded. The bounded loop intersects L twice and the unbounded loop intersects L once.

Figure 2 in the next chapter shows a rough but topologically accurate picture 2 of Eλ R for λ (0, 2). ∩ ∈ First Statement: We want to see that the gradient never vanishes on the V = 0. We compute

2 2 2 Vx λyz +3x 2xy + y z − 2 − 2 − 2 V = Vy =  λxz x +2xy 3y + z  (5) ∇ − − − Vz 2z(y x) λxy    − −  To analyze this, let us first consider the points in the line at infinity that belong to V = 0. Indpendent of λ, these are the 3 points

[1 : 1 : 0], [i : 1 : 0], [ i : 1 : 0]. −

7 Since both x, y = 0 and z = 0 for these points, we see from the equation that the third coordinate6 of V is nonzero. This takes care of these points. ∇ To consider the remaining points of V we can set z = 1. If V = 0 we ∇ have Vx + Vy = 0. This gives one of two equations: 2x λ y = x, y = − . (6) − 2 When y = x we have − λx +6x2 1 − V =  λx 6x2 +1 (7) − − ∇ x(λx 4)  −  This can only vanish when x =4/λ. But then 4 4 256 8 V , , 1 = + , λ −λ  λ3 λ and this vanishes only if λ = 4i√2. ± When y = (2x λ)/2 we have − λ(4 λ2) V (x, y, 1) = − . 8 This can only vanish when λ =0, 2. ± ♠ 2 Second Statement: Let λ R 2, 0, 2 . The set Eλ RP is a finite ∈ − {− } ∩ disjoint union of smooth loops, permuted by ρ. If C is a bounded component and ρ(C) = C then C ρ(C) would in- tersect some line 4 times, a contradiction. Hence6 ρ preserves∪ each bounded component, and each bounded component intersects L twice at right angles. 2 2 Since Eλ intersects CP C three times, and exactly one of these inter- − 2 2 section points, namely [1 : 1 : 0], lies in RP , we see that Eλ RP has one unbounded component. ∩ Note that Eλ L always consists of 3 points, namely (x, x) for x = 0 ∩ − and λ √λ2 + 32 x == − ± =0. 8 6 2 We conclude that Eλ RP must have exactly two components, both ρ- invariant, one bounded∩ and one unbounded, and the intersections are as claimed. ♠

8 2.3 Uniformization Let E CP 2 be nonsingular cubic variety. As we mentioned above, E is a complex⊂ torus. Let f : E E be some birational map which is also invertible. The birational nature→ of f implie that all the singularities of f on E are removable. This means that f is a biholomorphic map of E and orientation preserving. As is well known, there is also a biholomorphic map φ : E C/Λ where Λ C is a lattice. The map φ conjugates f to an of→C/Λ. Thus, we⊂ may simply equip E with the coordinates coming from φ and treat E as a flat torus and f : E E as an orientation preserving isometry. We call this the flat structure on→E.

2.4 Minor Subsets of Tori In this section I will prove a general lemma about flat tori. The only case required for the proof of Theorem 1.2 is that of the circle R/Z, but the general case might be useful for a more general version of the conjecture. The general case rather quickly reduces to the circle case anyway. Say that a subset S Rn/Zn is minor if there is some translation φ of Rn/Zn such that φ(S) ⊂ (0, 1/2)n. In general, say that a subset S of a flat ⊂ n n torus is minor if an affine isomorphism from the flat torus to R /Z carries S to a minor subset. Lemma 2.1 Suppose p S Y where S is a minor subset of the flat torus Y . Suppose that f : Y ∈ Y⊂is a nontrivial translation. Then the forward → orbit f k(p) k > 0 is not contained in S. { | } Proof: By affine symmetry it suffices to prove this when Y = Rn/Zn. The translation f has a nontrivial action in at least one coordinate. Let f : Rn/Zn R/Z be the projection onto this coordinate. By construction f(S) is minor→ in R/Z and f covers a nontrivial translation of R/Z. This reduction shows that it suffices to prove our result for Y = R/Z. This is what we do. If f(p) S then we are done. Otherwise p f(p) < 1/2. 6∈ | − | But then the forward orbit of p is at least 1/4-dense. This means that every point of R/Z is within 1/4 of some point in the forward orbit. In particular, the point ζ diametrically opposed from the midpoint of S has this property. But then the orbit point that is within 1/4 of ζ is disjoint from S. ♠

9 3 Proof of the Result

3.1 Formulas Let denote the space of labeled 8-gons with 4-fold rotational symmetry moduloX similarities in the plane. We normalize so that the 4-fold symmetry in question is the map

ρ([x : y : z]) = [ y : x : z]. (8) − This map fixes the origin (0, 0) in the affine patch and preserves the whole affine patch. It is just rotation by 90 degrees counterclockwise. One possibility is that ρ cycles the labels by 2 and the other pos- sibility is that ρ cycles the vertex labels by 2. We only consider the first possibility; the second possibility is essentially− treated by symmetry. For the purpose of getting formulas, we ignore for now the members of which have X points on the line at infinity. We call the remaining members finite. In other words, the finite members lie entirely in the standard affine patch.

The Map: Every finite member of has a canonical representative P (x, y) with vertices X

(1, 0), (x, y), (0, 1), ( y, x), ( 1, 0), ( x, y), (0, 1), (y, x). (9) − − − − − − Here (x, y) is really [x : y : 1], etc. Expressed in these coordinates, and with a suitable labeling scheme, the ′ ′ map T3 is given by T3(x, y)=(x ,y ), where

x′ = Ax(B 2xy), y′ =+Ay(B +2xy), (10) − −

α10 + α20 i j i j A = , αij = x y + y x . (1 + α10)(α20 +2α30 + α40 α11 2α12 + α22) − −

i j i j B = β10 +2β20 + β30 + β12, βij = x y y x . (11) − ′ ′ In other words, T3 sends the polygon P (x, y) to the polygon P (x ,y ). I computed this map (and everything else in the paper) using Mathematica [6].

10 The Invariant: Define the (x y)(x2 y2 1) Ψ(x, y)= − − − . (12) xy

A direct calculation shows that Ψ T3 = Ψ. This is the invariant mentioned in the introduction. In the next section◦ I will explain where it comes from.

Projective Duality: Each 8-gon P , defined by its vertices, gives rise to an 8-gon P ∗ in the dual space defined by the successive lines. The suc- cessive “vertices” of P ∗ are the successive lines extending the edges of P . Using our isomorphism, we get a second polygon (P ∗)# in . The operation P (P ∗)# is an involution given algebraically by the mapX → y (x2 x+y2 y) y(x + y 1) D(x, y)= − − , − (13)  − x (x2 2x + y2 + 1) x (x2 2x + y2 + 1) − − Direct calculations show −1 −1 Ψ D =Ψ, DT3D = T . (14) ◦ 3 One can also deduce these equations from abstract properties of projective duality. I will leave this to the interested reader.

Symmetries and Factorization: Define

σ1(x, y)=(y, x), σ2(x, y)=( x, y). (15) − − A direct calculation shows that −1 −1 −1 σ1T3σ1 = T3, σ2T3σ2 = T3 . (16)

Geometrically, the map σ2 swaps the regular and star-regular 8-gons. Beau- tifully, a calculation shows that 2 T3 =(D σ2) . (17) ◦ In other words T3 is the square of a simpler map. Readers familiar with the pentagram map will not be surprised by this kind of factorization. The map D σ2 satisfies the rule ◦ Ψ (D σ2)= Ψ. (18) ◦ ◦ − This equation would probably be the quickest way for the reader to show, without symbolic manipulation, that Ψ is an invariant for T3.

11 3.2 Special Cases

I first noticed that T3 behaved nicely on the sets described in this section. I then systematically tested Laurent monomials in the defining functions for these sets and this led me to Ψ.

The Coordinate Axes: First of all

T3(x, 0)=( x, 0), T3(0,y)=(0, y), (19) − −

So, T3 preserves the coordinate axes and is an involution there. The corre- sponding octagons look (to me) like the blades of a circular saw.

The Diagonal Line: Let ∆ denote the diagonal line x = y. The 8-gon P0 = P (x, x) has 8-fold dihedral symmetry, with the lines of symmetry going through the vertices. To study the (8, 3)-pentagram orbit Pj we compute { } 1+ x T (x, x)=(x′, x′), x′ = . (20) 3 1+2x

The map T3 is given by a projective transformation of ∆. The fixed points are p± = (1/√2, 1/√2). (21) ± The fixed point p+, which corresponds to the regular 8-gon, is attracting. The fixed point p−, which corresponds to the star-regular 8-gon, is repelling. Thus, every orbit on ∆ aside from p−, is attracted to p+ So, if P0 is not −1 regular then Pj is convex for all j 0. The inverse map T has p− as an ≥ 3 attracting fixed point and p+ as a repelling fixed point. Hence Pj is non- convex for all j sufficiently negative.

The Unit Circle: Let S1 denote the unit circle. Here S1 corresponds to 8-gons with 8-fold dihedral symmetry in which the lines of symmetry bisect the sides. These 8-gons are dual to the ones on ∆ and indeed the map D 1 −1 defined above has the property that D(∆) = S . Thus, the action of T3 1 on S is conjugate to the action of T3 on ∆. In particular, if we start with C0 convex then Cj is convex for all j 0 but Cj is non-convex for all j ≤ sufficiently positive. This is a case that the reader can easily experiment with. Just draw a “stop sign” and see what the 3-diagonal map does.

12 Other Special Orbits: The material here is not needed for the proof of Theorem 1.2 but it is nice. Let L±2,1 denote the diagonal line y = x 1 ∓ and let L±2,2 denote the circle of radius 1/√2 centered at ( 1/2, 1/2). The ∓ ± union L±2 = L±2,1 L±2 is the level set Ψ = 2. Each of these two level sets ∪ ± 2 is the disjoint union of a diagonal line and a circle. The map T3 preserves each component and acts there with order 3. For example

′′ ′′ ′′ 1 x T 2(x, x +1)=(x , x + 1), x = − − . (22) 3 x 3.3 Nontriviality 2 Let λ R 2, 0, 2 . We know that Eλ R contains one bounded loop ∈ − {− } ∩ 2 2 and one unbounded loop. Since T3 preserves Eλ and R and RP , we see 2 2 that f = T preserves both components of Eλ RP . Let Υ be one of these. 3 ∩ Lemma 3.1 f cannot be the identity on Υ.

Proof: Suppose f is the identity on Υ. Since f is a orientation preserving isometry Eλ in the flat coordinates, we see that f must be the identity on Eλ. We saw in §2.2 that Eλ intersects the line y = x in 3 distinct { − } points. Hence f(x, x)=(x, x) for two distinct nonzero points (x1, x1) − − − and (x2, x2) in Eλ. Setting the sum of the coordinates of f(x, x) equal to − − 0, we get (4x2( 1 2x2 + x4 6x6 + 32x10)) − − − =0. (23) (( 1+ x2 +4x4)(1 2x2 + x4 + 24x6 + 16x8)) − − The only nonzero real roots are 1 1 x = (1 + √17). ±2r2

But the corresponding points satisfy Ψ(x1, x1)= Ψ(x2, x2) = 0 so these − − − 6 points cannot both lie in Eλ. This is a contradiction. ♠

Remark: The points (x1, x1) and (x2, x2) constructed in the previous proof lie on level sets where−f has order 2.− In particular, f 2 is the identity on these two level sets.

We call an orientation preserving isometry of a metric circle a translation.

13 Lemma 3.2 f is a nontrivial translation of Υ in the flat coordinates.

Proof: Since f is a nontrivial isometry of Υ, we see that f is either a trans- lation or an orientation reversing isometry on Υ. In the latter case, f must have order 2 on Υ. If f is orientation reversing on Υ then f is orientation re- versing on nearby level sets. This means that f 2 is the identity on all nearby level sets. But then f 2 is the identity on an open subset of R2. Since f is a birational map, this forces f 2 = T 4 to be the identity. This is false. 3 ♠

3.4 The End of the Proof

Let denote the subset of convex 8-gons. Let − and + respectively denoteC ⊂ the X subsets of corresponding to 8-gons withC 8-fold dihedralC symme- C try with the lines of symmetry respectively going through the vertices and bisecting the edges. Note that + − is precisely the point representing the regular 8-gon. C ∩C The analysis of the special sets in §3.2 reduces us to considering polygons in + −. To finish the proof it suffices to show that when we have C − C − C P0 − + there are indices i< 0 < j such that Pi,Pj . Our choice ∈C−C −C 6∈ C of P0 implies that Ψ(p) R 2, 0, 2 . Here p is the point in representing ∈ 2 −{− } C P0. But then p Eλ R , where Eλ is the complex torus discussed in §2.2. ∈ ∩ Figure 2 shows a picture of the relevant sets. The shaded semidisk is . The lightly shaded disk is the unit disk. The sets + and − are the C C C intersection of with the unit circle and with the diagonal line respectively. C 2 The blue curve is a rough but topologically accurate sketch of Eλ RP when λ (0, 2). If λ ( 2, 0) the picture would be reflected in the∩ line ∈ ∈ − y = x . { }

14 2 Figure 2: A topologically accurate sketch of Eλ R for λ (0, 2). ∩ ∈ 2 Let Υ be the component of Eλ RP containing p. We equip Υ with its 2 ∩ flat coordinates. Let f = T3 as in the previous section. We know that f is a nontrivial translation of Υ. Let ρ be the reflection in the line y = x . This is the red diagonal in Figure 2. We showed in §2.2 that ρ{(Υ) =− Υ.} Being holomorphic, ρ is an isometry of Υ in the flat coordinates. The two sets Υ and ρ(Υ ) are disjoint and have equal length. Hence Υ is minor in∩ C the sense of∩ Lemma C 2.1. By Lemma 2.1, we have f j(p) Υ ∩ Cfor − 6∈ ∩ C some j > 0. Applying the same argument to f 1 gives some i< 0 such that f i(p) . This completes the proof of Theorem 1.2. 6∈ C

15 4 Centrally Symmetric Octagons

4.1 The Invariant Revisited We begin this chapter by discussing the invariant Ψ in more detail. At the end of §3.2 we discussed two other special level sets for Ψ, namely the level sets Ψ = 2. Each of these level sets consists of a disjoint union of a line and a circle.± The function (1 + x y)( x + x2 + y + y2) F (x, y)= − − (24) ( 1+ x y)(x + x2 y + y2) − − − equals 0 on one of these level sets and on the other one. One can compute directly that ∞ 2 Ψ F = − . (25) 2+Ψ

Thus, F is also an invariant for the map T3 on the set of 4-fold symmetric octagons. I discovered the two invariants for T3 on the larger space of cen- trally symmetric octagons by taking the natural generalizations of F 2 and Ψ.

4.2 Anatomy of an Octagon We will consider the space Ω of centrally symmetric octagons modulo affine transformations. We can uniquely specify a point in Ω by a quadruple (a,b,c,d). The vertices of the corresponding octagon are:

(1, 0), (a, b), (0, 1), ( d,c), ( 1, 0), ( a, b), (0, 1), (d, c). − − − − − − We choose coordinates this way so that points of the form (a, b, a, b) corre- spond to octagons with 4-fold rotational symmetry. Let denote this subset S of Ω. Let Ω+ and Ω− respectively denote the subset of Ω consisting of octagons circumscribed about, and inscribed in, an . We have

(a,b,c,d) Ω+ a b + c d =0. (26) ∈ ⇐⇒ − −

1 a2 b2 1 c2 d2 (a,b,c,d) Ω− − − + − − =0. (27) ∈ ⇐⇒ ab cd

16 Equation 27 has a nice geometric interpretation. The points (a, b) and ( d,c) respectively lie on of the form − x2 + y2 + Axy =1, and x2 + y2 + Bxy =1, where 1 a2 b2 1 c2 d2 A = − − , B = − − . ab − cd The equation on the right side of Equation 27 says that these two ellipses coincide. I don’t know a nice geometric explanation for Equation 26. The two equations are not symmetric with respect to each other because our coordinates are not natural with respect to projective dualities. Let be the set of convex centrally symmetric octagons. Let ± = Ω±. On theC subset , Equations 26 and 27 respectively pick out theC diagonalC∩ line S a = b and the unit circle. In particular, our definition of ± extends the definition given in the previous chapter. C In the previous chapter we observed that + − is just the regular C ∩ C octagon. In this more general setting + − consists of the convex Poncelet C ∩C octagons. The larger intersection Ω+ Ω− is exactly the set of Poncelet ∩ octagons. It turns out that the Poncelet polygons with an even number of vertices are centrally symmetric, and in particular this is true for Poncelet octagons.

4.3 The Map In these coordinates the 3 diagonal map is given by

′ ′ ′ ′ T3(a,b,c,d)=(a , b ,c ,d ) where

2 − − − 2 − − 2 − ′ c(ac + a + bd + d) ac acd + ac + bcd bc bd 2bd b + c cd + c a = − , (c + d + 1) a2c2 + a2c + 2abcd − abc − abd − ab + ac2 − acd + ac + b2d2 + b2d − bcd + bd2 + bd − cd 

− 2 − − − − 2 − 2 ′ d(ac + bd + b + c)  ac + acd 2ac ad a bcd + bd + bd cd + d + d b = − , (c + d + 1) a2c2 + a2c + 2abcd − abc − abd − ab + ac2 − acd + ac + b2d2 + b2d − bcd + bd2 + bd − cd  2 2 − − − − 2 − − ′ a(ac + bd + b + c) a c + a abc + abd ab + ac ad + a b d 2bd d c = − , (a + b + 1) a2c2 + a2c + 2abcd − abc − abd − ab + ac2 − acd + ac + b2d2 + b2d − bcd + bd2 + bd − cd  2 − − − − 2 2 − − ′ b(ac + a + bd + d) a ( c)+ abc abd ab 2ac + b d + b bc + bd + b c d = − (a + b + 1) a2c2 + a2c + 2abcd − abc − abd − ab + ac2 − acd + ac + b2d2 + b2d − bcd + bd2 + bd − cd 

17 This map has some symmetries. We have

I1T3I1 = T3, I1(a,b,c,d)=(c,d,a,b),

I2T3I2 = T3, I1(a,b,c,d)=(d,c,b,a), −1 I3T3I3 = T , I3(a,b,c,d)= (a,b,c,d). 3 − −1 In particular T3 has a formula very much like T3. Typically the orbits of T3 look like they are dense in a union of two 2- dimensional tori. Figure 4 shows an example of an orbit, projected into the (a, b)-plane. This orbit is not quite typical because the first 215 points of the orbit appear to lie in rather intricate torus curves. To see the torus, you need to blow the picture up at the pinch points ( 1, 0) and (0, 1). My computer program allows you to drag the points around± and change± the orbit. I dragged the points around until I found a nice one.

Figure 4: An orbit

18 4.4 The Invariants

The torus picture suggests that the map T3 has 2 algebraically independent invariants on Ω. One could perhaps find the invariant G below using A. Izosi- mov’s construction involving corrugated polygons, described in the appendix of his paper [4]. Perhaps one can find the invariant F below as well in this manner. I found F and G using some mixture of ad hoc geometric reasoning and computer experimentation. Motivated by Equations 27 and 26, I discovered that the quantity 1 1 a2 b2 1 c2 d2 Ψ(a,b,c,d)= (a b + c d) − − + − − (28) 4 − −  ab cd  is an invariant. This quantity agrees with the invariant from the previous chapter on . That is, Ψ(a, b, a, b) = Ψ(a, b). Note that Ψ vanishes exactly S on on Ω+ Ω−. Motivated by Equation 31, we define ∪ 2 Ψ F = − . (29) 2+Ψ F is the first invariant. The second invariant is (1+a b)(1+c d)(b c+ac+bd)( a+ac+d+bd) G(a,b,c,d)= − − − − (30) ( 1+a b)( 1+c d)( b+c+ac+bd)(a+ac d+bd). − − − − − − In view of Equation 25 we have

F 2(a, b, a, b)= G(a, b, a, b). (31)

This equation makes it more natural for us to work with F rather than Ψ. The invariants F and G are algebraically independent. A single calculation, showing the gradients F and G are not parallel, suffices to establish this. ∇ ∇ For instance 32 F (1, 1, 1, 2) = 1, 2, 2, 2 , G(1, 1, 1, 2)= (0, 0, 1, 1). ∇ 121 − −  ∇ −

The pair of maps (F,G) determines a map from Ω into R2. Since we are mostly intersected in the set , it is nicer to use the map C µ =(F 2,G). (32)

19 The map F is positive on so we don’t lose any information by taking the square. Also, µ interacts nicelyC with Equation 31. Thanks to Equation 31, the image µ( ) is the diagonal line y = x. By constructionS µ( ±)= (1, 1) . (33) C { } It seems that µ( ) = ∆1 ∆2, where ∆1 is the with vertices (0, 0), C ∪ (1, 0), (1, 1) and ∆2 is the image of ∆2 under the map (x, y) (1/x, 1/y). These two in RP 2 have (1, 1) as a common vertex. →

Let me close this section with a few more words addressed to people famil- iar with the pentagram map. Two of the classic invariants for the pentagram map on an n-gon are

On = x1, ..., xn, En = y1, ..., yn, (34) where x1,y1, ..., xn,yn are the corner invariant coordinates for the space of n-gons modulo projective transformations. These two quantities are the fun- damental invariants for the 2-diagonal map, the classical pentagram map. In general, On/En is an invariant for the 3-diagonal map, though the individual quantities On and En are not. For me, this is just an experimentally ob- served fact, though presumably one can deduce rigorously it from Izosimov’s construction, as mentioned above. I think that one could also prove it by just writing out the 3-diagonal map in corner invariant coordinates and doing some elementary algebra. On Ω we have O n = G±6 (35) En Which sign we get depends on the labeling conventions for the corner in- variant coordinates. I do not know the expression for F in terms of these invariants.

4.5 Hyperbolic Dynamics In [16], S. Tabachnikov and I observed that

T3(Ω±) = Ω±, (36) We phrased things differently, and also did not require the octagons to be centrally symmetric for this to be true.

20 It appears that for each p Ω−, representing an inscribed centrally sym- metric octagon, the following∈ is true:

• T n(p) converges to a star-convex Poncelet octagon as n + . 3 → ∞ • T n(p) converges to a convex Poncelet octagon as n . 3 → −∞

The dynamics is exactly reversed for points in Ω+. It is also worth pointing out that on Ω− the quantity

1 a2 b2 A = − − ab is an invariant. I think that the quantity a b should be an invariant on Ω+, just by duality, but I did not check this. − More generally, Ω+ and Ω− are respectively stable and unstable manifolds for T3. Their intersection is precisely the set of Poncelet octagons. This picture extends the picture discussed in the previous chapter. Away from the set Ω+ Ω− it seems that the space Ω, when complexified, has a singular ∪ foliation by complex tori of complex dimension 2 and in the corresponding 2 flat coordinates the map T3 is a translation on each such torus. This picture almost seems incompatible with the hyperbolic dynamics just described but there is no contradiction. This kind of dynamics is certainly possible, as we say one complex dimension down in the previous chapter.

4.6 The Expected Result Here is the theorem I expect:

Conjecture 4.1 Suppose that P0 is a centrally symmetric 8-gon and Pj { } is the (8, 3)-pentagram orbit. Then P0 is convex Poncelet if and only if Pj is convex for all j. More precisely

1. P0 is convex and superscribed about an ellipse if and only if Pj is convex for all j 0. ≥

2. P0 is convex and inscribed in an ellipse if and only if Pj is convex for all j 0. ≤

21 At least in outline the proof ought to be similar to what we did in the last chapter. In the last chapter we could draw on standard facts about el- liptic curves to identify the level curves of our invariants as complex tori. In the present setting we would have to analyze the common level sets of two invariants – i.e., the fibers of the map (F,G). Let me at least offer the barest sketch of a proof. I hope to fill in the details eventually.

Very Rough Sketch: The set X where F and G are parallel is T3- invariant. This suggests that X is a union∇ of common∇ level sets of F and G. A dimension count then suggests that µ(X) is contained in a subset of (C )2 having complex dimension 1. Assuming that we find a suitable compactification∪∞ for Ω(C), the complexified version of Ω, this situation says that the set of regular fibers is connected. If we can show that one of them is a complex torus, then they are all complex tori by “analytic continuation”. Assuming that all the regular fibers are (unions of) complex tori, the next thing would be to show that all the level sets containing points of are regular, except those containing points of the slice . Experimentally,C this seems to be true. Here is a heuristic argument. Experimentally,S every such fiber contains a point of the form (1, b, 1,d). Let Y denote the set of such points. If X is really a union of fibers then it suffices to show that

X Y . (37) ∩ ⊂ S The projection of F into the tangent space of Y is parallel to (1, 1) and the ∇ projection of G into the tangent space to Y is parallel to (1, 1) only when b = d. This establishes∇ Equation 37. The next step would be so show that for each complex torus τ the in- tersection τ is minor in the sense of Lemma 2.1. For this purpose the ∩ C symmetries I1 and I2 which commute with T3 ought to help. The final step 2 would be to show that T3 is nontrivial each connected component of a fiber that intersects . From here the proof works as in the previous chapter. C

22 References

[1] M. Gekhtman, M. Shapiro, S.Tabachnikov, A. Vainshtein, Higher penta- gram maps, weighted directed networks, and cluster dynamics, Electron. Res. Announc. Math. Sci. 19 21012, 1–17 [2] M. Glick, The Limit Point of the Pentagram Map, International Math- ematics Research Notices 9 (2020) 2818–2831 [3] M. Glick, The pentagram map and Y -patterns, Adv. Math. 227, 2012, 1019–1045. [4] A. Izosimov, The pentagram map, Poncelet polygons, and commuting difference operators, arXiv 1906.10749 (2019) [5] B. Khesin, F. Soloviev Integrability of higher pentagram maps, Mathem. Annalen. (to appear) 2013 [6] Wolfram Research Inc., Mathematica, Wolfram Programming Lab, Champaign, IL (2021) [7] G. Mari Beffa, On Generalizations of the Pentagram Map: Discretiza- tions of AGD Flows, arXiv:1303.5047, 2013 [8] G. Mari Beffa, On integrable generalizations of the pentagram map arXiv:1303.4295, 2013 [9] Th. Motzkin, The in the projective plane, with a comment on Napier’s rule, Bull. Amer. Math. Soc. 52, 1945, 985–989. [10] V. Ovsienko, R. Schwartz, S. Tabachnikov, The pentagram map: A dis- crete , Comm. Math. Phys. 299, 2010, 409–446. [11] V. Ovsienko, R. Schwartz, S. Tabachnikov, Liouville-Arnold integrability of the pentagram map on closed polygons, to appear in Duke Math. J. [12] B. Rodin and S. Sullivan, The convergence of circle packings to the Riemann maping, J. Diff. Geom. 26 (1987) 349–360 [13] R. Schwartz, The pentagram map, Experiment. Math. 1, 1992, 71–81. [14] R. Schwartz, Discrete monodromy, , and the method of con- densation, J. of Fixed Point Theory and Appl. 3, 2008, 379–409.

23 [15] R. Schwartz, The Poncelet Grid, Advances in Geometry 7 (2006)

[16] R. Schwartz, S. Tabachnikov, Elementary surprises in projective geom- etry, Math. Intelligencer (2010)

[17] J. Silverman, The arithmetic of elliptic curves, Springer

[18] F. Soloviev Integrability of the Pentagram Map, Duke Math J.

[19] M. Weinreich, The Algebraic Dynamics of the Pentagram Map, (2021) arXiv: 2104.06211

24