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

The Flapping Birds in the Pentagram Zoo

Richard Evan Schwartz Supported by N.S.F. Grant DMS-21082802
Abstract

We study the (k+1,k)(k+1,k) diagonal map for k=2,3,4,k=2,3,4,.... We call this map Δk\Delta_{k}. The map Δ1\Delta_{1} is the pentagram map and Δk\Delta_{k} is a generalization. Δk\Delta_{k} does not preserve convexity, but we prove that Δk\Delta_{k} preserves a subset BkB_{k} of certain star-shaped polygons which we call kk-birds. The action of Δk\Delta_{k} on BkB_{k} seems similar to the action of Δ1\Delta_{1} on the space of convex polygons. We show that some classic geometric results about Δ1\Delta_{1} generalize to this setting.

1 Introduction

1.1 Context

When you visit the pentagram zoo you should certainly make the pentagram map itself your first stop. This old and venerated animal has been around since the place opened up and it is very friendly towards children. When defined on convex pentagons, this map has a very long history. See e.g. [15]. In modern times [19], the pentagram is defined and studied much more generally. The easiest case to explain is the action on convex nn-gons. One starts with a convex nn-gon PP, for n5n\geq 5, and then forms a new convex nn-gon PP^{\prime} by intersecting the consecutive diagonals, as shown Figure 1.1 below.

The magic starts when you iterate the map. One of the first things I proved in [19] about the pentagram map is the successive iterates shrink to a point. Many years later, M. Glick [3] proved that this limit point is an algebraic function of the vertices, and indeed found a formula for it. See also [9] and [1].

[Uncaptioned image]

Figure 1.1: The pentagram map iterated on a convex 77-gon PP.

Forgetting about convexity, the pentagram map is generically defined on polygons in the projective plane over any field except for 𝒁/2\mbox{\boldmath{$Z$}}/2. In all cases, the pentagram map commutes with projective transformations and thereby defines a birational map on the space of nn-gons modulo projective transformations. The action on this moduli space has a beautiful structure. As shown in [17] [18], and independently in [23], the pentagram map is a discrete completely integrable system when the ground field is the reals. ([23] also treats the complex case.) Recently, M. Weinreich [24] generalized the integrability result, to a large extent, to fields of positive characteristic.

The pentagram map has many generalizations. See for example [2], [14], [16], [10], [11], [6]. The paper [2] has the first general complete integrability result. The authors prove the complete integrability of the (k,1)(k,1) diagonal maps, i.e. the maps obtained by intersecting successive kk-diagonals. Figure 1.3 below shows the (3,1)(3,1) diagonal map. (Technically, [2] concentrates on what happens when these maps act on so-called corregated polygons in higher dimensional Euclidean spaces.) The paper [6] proves an integrability result for a very wide class of generalizations, including the ones we study below. (Technically, for the maps we consider here, the result in [6] does not establish the algebraic independence of invariants needed for complete integrability.) The pentagram map and its many generalizations are related to a number of topics: alternating sign matrices [20], dimers [5], cluster algebras [4], the KdV hierarchy [12], [13], spin networks [2], Poisson Lie groups [8], Lax pairs [23], [10], [11], [6], [8], and so forth. The zoo has many cages and sometimes you have to get up on a tall ladder to see inside them.

[Uncaptioned image]

Figure 1.2: The (3,1)(3,1)-diagonal map acting on 88-gons.

The algebraic side of the pentagram zoo is extremely well developed, but the geometric side is hardly developed at all. In spite of all the algebraic results, we don’t really know, geometrically speaking, much about what the pentagram map and its relatives really do to polygons.

Geometrically speaking, there seems to be a dichotomy between convexity and non-convexity. The generic pentagram orbit of a projective equivalance class of a convex polygon lies on a smooth torus, and you can make very nice animations. What you will see, if you tune the power of the map and pick suitable representatives of the projective classes, is a convex polygon sloshing around as if it were moving through water waves. If you try the pentagram map on a non-convex polygon, you see a crazy erratic picture no matter how you try to normalize the images. The situation is even worse for the other maps in the pentagram zoo, because these generally do not preserve convexity. Figure 1.2 shows how the (3,1)(3,1)-diagonal map does not necessarily preserve convexity, for instance. See [21], [22] for more details.

If you want to look at pentagram map generalizations, you have to abandon convexity. However, in this paper, I will show that sometimes there are geometrically appealing replacements. The context for these replacements is the (k+1,k)(k+1,k)-diagonal map, which I call Δk\Delta_{k}, acting on what I call kk-birds. Δk\Delta_{k} starts with the polygon PP and intersects the (k+1)(k+1)-diagonals which differ by kk clicks. (We will give a more formal definition in the next section.) Δk\Delta_{k} is well (but not perfectly) understood algebraically [6]. Geometrically it is not well understood at all.

1.2 The Maps and the Birds

Definition of a Polygon: For us, a polygon is a choice of both vertices and the edges connecting them. Each polygon PP we consider will all be planar, in the sense that there is some projective transformation that maps PP, both vertices and edges, to the affine patch. Our classical example is a regular nn-gon, with the obvious short edges chosen.

The Maps: Given a polygon PP, let PaP_{a} denote the (a)(a)th vertex of PP. Let PabP_{ab} be the line through PaP_{a} and PbP_{b}. The vertices of Δk(P)\Delta_{k}(P) are

Pj,j+k+1Pj+1,jk.P_{j,j+k+1}\cap P_{j+1,j-k}. (1)

Here the indices are taken mod nn. Figure 1.3 shows this for (k,n)=(2,7)(k,n)=(2,7). The polygons in Figure 1.3 are examples of a concept we shall define shortly, that of a kk-bird.

[Uncaptioned image]

Figure 1.3: Δ2\Delta_{2} acting on 22-birds.

We should say a word about how the edges are defined. In the case for the regular nn-gon we make the obvious choice, discussed above. In general, we define the class of polygons we consider in terms of a homotopy from the regular nn-gon. So, in general, we make the edge choices so that the edges vary continuously.

The Birds: Given an nn-gon PP, we let Pa,bP_{a,b} denote the line containing the vertices PaP_{a} and PbP_{b}. We call PP kk-nice if n>3kn>3k, and PP is planar, and the 44 lines

Pi,ik1,Pi,ik,Pi,i+k,Pi,i+k+1P_{i,i-k-1},\hskip 10.0ptP_{i,i-k},\hskip 10.0ptP_{i,i+k},\hskip 10.0ptP_{i,i+k+1} (2)

are distinct for all ii. It is not true that the generic nn-gon is kk-nice, because there are open sets of non-planar polygons. (Consider a neighborhood of PP, where PP the regular 100100-gon with the opposite choice of edges.) However, the generic perturbation of a planar nn-gon is also kk-nice.

We call PP a kk-bird if PP is the endpoint of a path of kk-nice nn-gons that starts with the regular nn-gon. We let Bk,nB_{k,n} be the subspace of nn-gons which are kk-birds. Note that Bk,nB_{k,n} contains the set of convex nn-gons, and the containment is strict when k>1k>1. As Figure 1.3 illustrates, a kk-bird need not be convex for k2k\geq 2. We will show that kk-birds are always star-shaped, and in particular embedded. As we mentioned above, we use the homotopic definition of a kk-bird, to define the edges of Δk(P)\Delta_{k}(P) when PP is a kk-bird.

Example: The homotopy part of our definition looks a bit strange, but it is necessary. To illustrate this, we consider the picture further for the case k=1k=1. In this case, a 11-bird must be convex, though the 11-niceness condition just means planar and locally convex. Figure 1.4 shows how we might attempt a homoropy from the regular octagon to a locally convex octagon which essentially wraps twice around a quadrilateral. The little grey arrows give hints about how the points are moved. At some times, the homotopy must break the 11-niceness condition. The two grey polygons indicate failures and the highlighted vertices indicate the sites of the failures. There might be other failures as well; we are taking some jumps in our depiction.

[Uncaptioned image]

Figure 1.4: A homotopy that cannot stay 11-nice.

One could make similar pictures when k1k\geq 1, but the pictures might be harder to understand.

1.3 The Main Result

Given an embedded planar polygon PP, let PIP^{I} denote the interior of region bounded by PP. We say that PP is strictly star shaped with respect to xPIx\in P^{I} if each ray emanating from xx intersects PP exactly once. More simply, we say that PP is strictly star shaped if it is strictly star shaped with respect to some point xPIx\in P^{I}. Here is the main result.

Theorem 1.1

Let k2k\geq 2 and n>3kn>3k and PBk,nP\in B_{k,n}. Then

  1. 1.

    PP is strictly star-shaped, and in particular embedded.

  2. 2.

    Δk(P)PI\Delta_{k}(P)\subset P^{I}.

  3. 3.

    Δk(Bk,n)=Bk,n\Delta_{k}(B_{k,n})=B_{k,n}.

Remark: The statement that n>3kn>3k is present just for emphasis. Bn,kB_{n,k} is by definition empty when n3kn\leq 3k. The restriction n>3kn>3k is necessary. Figure 1.5 illustrates what would be a counter-example to Theorem 1.1 for the pair (k,n)=(3,9)(k,n)=(3,9). The issue is that a certain triple of 44-diagonals has a common intersection point. This does not happen for n>3kn>3k. See Lemma 3.6.

[Uncaptioned image]

Figure 1.5: Δ3\Delta_{3} acting on a certain convex 99-gon.

1.4 The Energy

We will deduce Statements 1 and 2 of Theorem 1.1 in a geometric way. The key to proving Statement 3 is a natural quantity associated to a kk-bird. We let σa,b\sigma_{a,b} be the slope of the line Pa,bP_{a,b} and we define the cross ratio

χ(a,b,c,d)=(ab)(cd)(ac)(bd).\chi(a,b,c,d)=\frac{(a-b)(c-d)}{(a-c)(b-d)}. (3)

We define

χk(P)=i=1nχ(i,k,P),χ(i,k,P)=χ(σi,ik,σi,ik1,σi,i+k+1,σi,i+k)\chi_{k}(P)=\prod_{i=1}^{n}\chi(i,k,P),\hskip 20.0pt\chi(i,k,P)=\chi(\sigma_{i,i-k},\sigma_{i,i-k-1},\sigma_{i,i+k+1},\sigma_{i,i+k}) (4)

Here we are taking the cross ratio the slopes the lines involved in our definition of kk-nice. When k=1k=1 this is the familiar invariant χ1=OE\chi_{1}=OE for the pentagram map Δ1\Delta_{1}. See [19], [20], [17], [18]. When n=3k+1n=3k+1, a suitable star-relabeling of our polygons converts Δk\Delta_{k} to Δ1\Delta_{1} and χk\chi_{k} to 1/χ11/\chi_{1}. So, in this case χkΔk=χk\chi_{k}\circ\Delta_{k}=\chi_{k}. Figure 1.5 illustrates this for (k,n)=(3,10)(k,n)=(3,10). Note that the polygons suggested by the dots in Figure 1.5 are not convex. Were we to add in the edges we would get a highly non-convex pattern.

[Uncaptioned image]

Figure 1.6: A star-relabeling converts Δ1\Delta_{1} to Δ3\Delta_{3} and 1/χ11/\chi_{1} to χ3\chi_{3}.

In general, χk\chi_{k} is not as clearly related to χ1\chi_{1}. Nonetheless, we will prove

Theorem 1.2

χkΔk=χk\chi_{k}\circ\Delta_{k}=\chi_{k}.

Theorem 1.2 is meant to hold for all nn-gons, as long as all quantities are defined. There is no need to restrict to birds.

1.5 The Collapse Point

When it is understood that PBk,nP\in B_{k,n} it is convenient to write

P=Δk(P)P^{\ell}=\Delta_{k}^{\ell}(P) (5)

We also let P^\widehat{P} denote the closed planar region bounded by PP. Figure 1.7 below shows P^=P^0,P^1,P^2,P^3,P^4\widehat{P}=\widehat{P}^{0},\widehat{P}^{1},\widehat{P}^{2},\widehat{P}^{3},\widehat{P}^{4} for some PB4,13P\in B_{4,13}.

[Uncaptioned image]

Figure 1.7: Δ4\Delta_{4} and its iterates acting on a member of B4,13B_{4,13}.

Define

P^=𝒁P^,P^=𝒁P^.\widehat{P}_{\infty}=\bigcap_{\ell\in\mbox{\boldmath{$Z$}}}\widehat{P}^{\ell},\hskip 30.0pt\widehat{P}_{-\infty}=\bigcup_{\ell\in\mbox{\boldmath{$Z$}}}\widehat{P}^{\ell}. (6)
Theorem 1.3

If PBk,nP\in B_{k,n} then P^\widehat{P}_{\infty} is a point and P^\widehat{P}_{-\infty} is an affine plane.

Our argument will show that PBk,nP\in B_{k,n} is strictly star-shaped with respect to all points in P^n\widehat{P}^{n}. In particular, all polygons in the orbit are strictly star-shaped with respect to the collapse point P^\widehat{P}_{\infty}. See Corollary 7.3.

One might wonder if some version of Glick’s formula works for the P^\widehat{P}_{\infty} in general. I discovered experimentally that this is indeed the case for n=3k+1n=3k+1 and n=3k+2n=3k+2. See §9.2 for a discussion of this and related matters.

Here is a corollary of our results that is just about convex polygons.

Corollary 1.4

Suppose that n>3kn>3k and PP is a convex nn-gon. Then the sequence {Δk(P)}\{\Delta_{k}^{\ell}(P)\} shrinks to a point as \ell\to\infty, and each member of this sequence if strictly star-shaped with respect to the collapse point.

1.6 The Triangulations

In §7.1 we associate to each kk-bird PP a triangulation τP𝑷\tau_{P}\subset\mbox{\boldmath{$P$}}, the projective plane. Here τP\tau_{P} is an embedded degree 66 triangulation of PPP_{-\infty}-P_{\infty}. The edges are made from the segments in the δ\delta-diagonals of PP and its iterates for δ=1,k,k+1\delta=1,k,k+1.

[Uncaptioned image]

Figure 1.8: The triangulation associated to a member of B5,16B_{5,16}.

Figure 1.8 shows this tiling associated to a member of B5,16B_{5,16}. In this figure, the interface between the big black triangles and the big white triangles is some Δ5(P)\Delta_{5}^{\ell}(P) for some smallish value of \ell. (I zoomed into the picture a bit to remove the boundary of the initial PP.) The picture is normalized so that the line PP_{-\infty} is the line at infinity. When I make these kinds of pictures (and animations), I normalize so that the ellipse of inertia of PP is the unit disk.

1.7 Paper Organization

This paper is organized as follows.

  • In §2 we prove Theorem 1.2.

  • In §3 we prove Statement 1 of Theorem 1.1.

  • In §4 we prove Statement 2 of Theorem 1.1.

  • In §5 we prove a technical result called the Degeneration Lemma, which will help with Statement 3 of Theorem 1.1.

  • In §6 we prove Statement 3 of Theorem 1.1.

  • In §7 we introduce the triangulations discussed above. Our Theorem 7.2 will help with the proof of Theorem 1.3.

  • In §8 we prove Theorem 1.3.

  • In §9, an appendix, we sketch an alternate proof of Theorem 1.2 which Anton Izosimov kindly explained. We also discuss Glick’s collapse formula and star relabelings of polygons.

1.8 Visit the Flapping Bird Exhibit

Our results inject some more geometry into the pentagram zoo. Our results even have geometric implications for the pentagram map itself. See §9.3. There are different ways to visit the flapping bird exhibit in the zoo. You could read the proofs here, or you might just want to to look at some images:
http://www.math.brown.edu/\simreschwar/BirdGallery
You can also download and play with the software I wrote:
http://www.math.brown.edu/\simreschwar/Java/Bird.TAR
The software has detailed instructions. You can view this paper as a justification for why the nice images actually exist.

1.9 Acknowledgements

I would like to thank Misha Gekhtman, Max Glick, Anton Izosimov, Boris Khesin, Valentin Ovsienko, and Serge Tabachnikov for many discussions about the pentagram zoo. I would like to thank Anton, in particular, for extensive discussions about the material in §9.

2 The Energy

The purpose of this chapter is to prove Theorem 1.2. The proof, which is similar to what I do in [19], is more of a verification than a conceptual explanation. My computer program allows the reader to understand the technical details of the proof better. The reader might want to just skim this chapter on the first reading. In §9 I will sketch an alternate proof, which I learned from Anton Izosimov. Izosimov’s proof also uses the first two sections of this chapter.

2.1 Projective Geometry

Let 𝑷P denote the real projective plane. This is the space of 11-dimensional subspaces of 𝑹3\mbox{\boldmath{$R$}}^{3}. The projective plane 𝑷P contains 𝑹2\mbox{\boldmath{$R$}}^{2} as the affine patch. Here 𝑹2\mbox{\boldmath{$R$}}^{2} corresponds to vectors of the form (x,y,1)(x,y,1), which in turn define elements of 𝑷P.

Let 𝑷\mbox{\boldmath{$P$}}^{*} denote the dual projective plane, namely the space of lines in 𝑷P. The elements in 𝑷\mbox{\boldmath{$P$}}^{*} are naturally equivalent to 22-dimensional subspaces of 𝑹3\mbox{\boldmath{$R$}}^{3}. The line in 𝑷P such a subspace Π\Pi defines is equal to the union of all 11-dimensional subspaces of Π\Pi.

Any invertible linear transformation of 𝑹3\mbox{\boldmath{$R$}}^{3} induces a projective transformation of 𝑷P, and also of 𝑷\mbox{\boldmath{$P$}}^{*}. These form the projective group PSL3(𝑹)PSL_{3}(\mbox{\boldmath{$R$}}). Such maps preserve collinear points and coincident lines.

A duality from 𝑷P to 𝑷\mbox{\boldmath{$P$}}^{*} is an analytic diffeomorphism 𝑷𝑷\mbox{\boldmath{$P$}}\to\mbox{\boldmath{$P$}}^{*} which maps collinear points to coincidence lines. The classic example is the map which sends each linear subspace of 𝑹3\mbox{\boldmath{$R$}}^{3} to its orthogonal complement.

A PolyPoint is a cyclically ordered list of points of 𝑷P. When there are nn such points, we call this an nn-Point. A PolyLine is a cyclically ordered list of lines in 𝑷P, which is the same as a cyclically ordered list of points in 𝑷\mbox{\boldmath{$P$}}^{*}. A projective duality maps PolyLines to PolyPoints, and vice versa.

Each nn-Point determines 2n2^{n} polygons in 𝑷P because, for each pair of consecutive points, we may choose one of two line segments to join them. As we mentioned in the introduction, we have a canonical choice for kk-birds. Theorem 1.2 only involves PolyPoints, and our proof uses PolyPoints and PolyLines.

Given a nn-Point PP, we let PjP_{j} be its jjth point. We make a similar definition for nn-Lines. We always take indices mod nn.

2.2 Factoring the Map

Like the pentagram map, the map Δk\Delta_{k} is the product of 22 involutions. This factorization will be useful here and in later chapters.

Given a PolyPoint PP, consisting of points P1,,PnP_{1},...,P_{n}, we define Q=Dm(P)Q=D_{m}(P) to be the PolyLine whose successive lines are P0,mP_{0,m}, P1,m+1P_{1,m+1}, etc. Here P0,mP_{0,m} denotes the line through P0P_{0} and PmP_{m}, etc. We labed the vertices so that

Qmi=Pi,i+m.Q_{-m-i}=P_{i,i+m}. (7)

This is a convenient choice. We define the action of DmD_{m} on PolyLines in the same way, switching the roles of points and lines. For PolyLines, P0,mP_{0,m} is the intersection of the line P0P_{0} with the line PmP_{m}. The map DmD_{m} is an involution which swaps PolyPoints with PolyLines. We have the compositions

Δk=DkDk+1,Δk1=Dk+1Dk.\Delta_{k}=D_{k}\circ D_{k+1},\hskip 30.0pt\Delta_{k}^{-1}=D_{k+1}\circ D_{k}. (8)

The energy χk\chi_{k} makes sense for nn-Lines as well as for nn-Points. The quantities χkDk(P)\chi_{k}\circ D_{k}(P) and χkDk+1(P)\chi_{k}\circ D_{k+1}(P) can be computed directly from the PolyPoint PP. Figure 2.1 shows schematically the 44-tuples associated to χ(0,k,Q)\chi(0,k,Q) for Q=PQ=P and Dk(P)D_{k}(P) and Dk+1(P)D_{k+1}(P). In each case, χk(Q)\chi_{k}(Q) is a product of nn cross ratios like these. If we want to compute the factor of χk(Dk(P))\chi_{k}(D_{k}(P)) associated to index ii we subtract (rather than add) ii from the indices shown in the middle figure. A similar rule goes for Dk+1(P)D_{k+1}(P).

[Uncaptioned image]

Figure 2.1: Computing the kk-energy.

Theorem 1.2 follows from the next two results.

Lemma 2.1

χkDk=χk\chi_{k}\circ D_{k}=\chi_{k}.

Lemma 2.2

χkDk+1=χk\chi_{k}\circ D_{k+1}=\chi_{k}.

These results have almost identical proofs. We consider Lemma 2.1 in detail and then explain the small changes needed for Lemma 2.2.

2.3 Proof of the First Result

We study the ratio

R(P)=χkDk(P)χk(P).R(P)=\frac{\chi_{k}\circ D_{k}(P)}{\chi_{k}(P)}. (9)

We want to show that R(P)R(P) equals 11 wherever it is defined. We certainly have R(P)=1R(P)=1 when PP is the regular nn-Point.

Given a PolyPoint PP we choose a pair of vertices a,ba,b with |ab|=k|a-b|=k. We define P(t)P(t) to be the PolyPoint obtained by replacing PaP_{a} with

(1t)Pa+tPb.(1-t)P_{a}+tP_{b}. (10)

Figure 2.2 shows what we are talking about, in case k=3k=3. We have rotated the picture so that PaP_{a} and PbP_{b} both lie on the XX-axis.

[Uncaptioned image]

Figure 2.2: Connecting one PolyPoint to another by sliding a point.

The two functions

f(t)=χk(P(t)),g(t)=χkDk(P(t))f(t)=\chi_{k}(P(t)),\hskip 30.0ptg(t)=\chi_{k}\circ D_{k}(P(t)) (11)

are each rational functions of tt. Our notation does not reflect that ff and gg depend on P,a,bP,a,b.

A linear fractional transformation is a map of the form

tαt+βγt+δ,α,β,γ,δ𝑹,αδβγ0.t\to\frac{\alpha t+\beta}{\gamma t+\delta},\hskip 30.0pt\alpha,\beta,\gamma,\delta\in\mbox{\boldmath{$R$}},\hskip 30.0pt\alpha\delta-\beta\gamma\not=0.
Lemma 2.3 (Factor I)

If n4k+2n\geq 4k+2 and PP is a generically chosen nn-Point, then f(t)f(t) and g(t)g(t) are each products of 44 linear fractional transformations. The zeros of ff and gg occur at the same points and the poles of ff and gg occur at the same points. Hence f/gf/g is constant.

The only reason we choose n4k+2n\geq 4k+2 in the Factor Lemma is so that the various diagonals involved in the proof do not have common endpoints. The Factor Lemma I works the same way for all kk and for all choices of (large) nn. We write PQP\leftrightarrow Q if we can choose indices a,ba,b and some t𝑹t\in\mbox{\boldmath{$R$}} such that Q=P(t)Q=P(t). The Factor Lemma implies that when P,QP,Q are generic and PQP\leftrightarrow Q we have R(P)=R(Q)R(P)=R(Q). The result for non-generic choices of PP follows from continuity. Any nn-Point QQ can be included in a finite chain

P0P1P2n=Q,P_{0}\leftrightarrow P_{1}\leftrightarrow\cdots\leftrightarrow P_{2n}=Q,

where P0P_{0} is the regular nn-Point. Hence R(Q)=R(P0)=1R(Q)=R(P_{0})=1. This shows that Lemma 2.1 holds for (k,n)(k,n) where k2k\geq 2 and n4k+2n\geq 4k+2. (The case k=1k=1 is a main result of [19], and by now has many proofs.)

Lemma 2.4

If Lemma 2.1 is true for all large values of nn, then it is true for all values of nn.

Proof: If we are interested in the result for small values of nn, we can replace a given PolyPoint PP with its mm-fold cyclic cover mPmP. We have χk(mP)=χk(P)m\chi_{k}(mP)=\chi_{k}(P)^{m} and χk(Dk(mP))=χk(Dk(p))m\chi_{k}(D_{k}(mP))=\chi_{k}(D_{k}(p))^{m}. Thus, the result for large nn implies the result for small nn. \spadesuit

In view of Equation 4 we have

f(t)=f1(t)fn(t),fj(t)=χ(j,k,P(t)).f(t)=f_{1}(t)...f_{n}(t),\hskip 30.0ptf_{j}(t)=\chi(j,k,P(t)). (12)

Thus f(t)f(t) is the product of nn “local” cross ratios. We call an index jj asleep if none of the lines involved in the cross ratio fj(t)f_{j}(t) depend on tt. In other words, the lines do not vary at all with tt. Otherwise we call jj awake.

As we vary tt, only the diagonals P0,hP_{0,h} change for h=k,k1,k+1,kh=-k,-k-1,k+1,k. From this fact, it is not surprising that there are precisely 44 awake indices. These indices are

j0=0,j1=k+1,j2=k1,j3=k.j_{0}=0,\hskip 10.0ptj_{1}=k+1,\hskip 10.0ptj_{2}=-k-1,\hskip 10.0ptj_{3}=-k. (13)

The index kk is not awake because the diagonal P0,k(t)P_{0,k}(t) does not move with tt.

We define a chord of P(t)P(t) to be a line defined by a pair of vertices of P(t)P(t). The point P0(t)P_{0}(t) moves at linear speed, and the 44 lines involved in the calculation of fcj(t)f_{c_{j}}(t) are distinct unless P0(t)P_{0}(t) lies in one of the chords of P(t)P(t). Thus fcj(t)f_{c_{j}}(t) only has zeros and poles at the corresponding values of tt. It turns out that only the following chords are involved.

kk1kk+1k1k2k1k11k12k1k+11k+12k+1\begin{matrix}-k\cr-k-1\end{matrix}\hskip 15.0pt\begin{matrix}-k\cr k+1\end{matrix}\hskip 15.0pt\begin{matrix}-k\cr 1\end{matrix}\hskip 15.0pt\begin{matrix}-k\cr-2k-1\end{matrix}\hskip 15.0pt\begin{matrix}-k-1\cr-1\end{matrix}\hskip 15.0pt\begin{matrix}-k-1\cr-2k-1\end{matrix}\hskip 15.0pt\begin{matrix}k+1\cr 1\end{matrix}\hskip 15.0pt\begin{matrix}k+1\cr 2k+1\end{matrix} (14)

We call these c0,,c7c_{0},...,c_{7}. For instance, c0c_{0} is the line through PkP_{-k} and Pk1P_{-k-1}. Let tjt_{j} denote the value of tt such that P(tj)cjP(t_{j})\in c_{j}.

The PolyPoint Q(t)=Dk(P(t))Q(t)=D_{k}(P(t)) has the same structure as P(t)P(t). Up to projective transformations Q(t)Q(t) is also obtained from the regular PolyPoint by moving a single vertex along one of the kk-diagonals. The pattern of zeros and poles is not precisely the same because the chords of Q(t)Q(t) do not correspond to the chords of P(t)P(t) in a completely straightforward way. The kk-diagonals of Q(t)Q(t) correspond to the vertices of P(t)P(t) and vice versa. The (k+1)(k+1) diagonals of Q(t)Q(t) correspond to the vertices of Δk1(P(t))\Delta_{k}^{-1}(P(t)). This is what gives us our quadruples of points in the middle picture in Figure 2.1.

We now list the pattern of zeros and poles. We explain our notation by way of example. The quadruple (f,2,4,5)(f,2,4,5) indicates that fc2f_{c_{2}} has a simple zero at f4f_{4} and a simple pole at t5t_{5}.

(f,0,0,1),(f,1,6,7),(f,2,4,5),(f,3,2,3).(f,0,0,1),\hskip 20.0pt(f,1,6,7),\hskip 20.0pt(f,2,4,5),\hskip 20.0pt(f,3,2,3). (15)
(g,0,6,5),(g,1,0,3),(g,2,2,1),(g,3,4,7).(g,0,6,5),\hskip 20.0pt(g,1,0,3),\hskip 20.0pt(g,2,2,1),\hskip 20.0pt(g,3,4,7). (16)

Since these functions have holomorphic extensions to 𝑪C with no other zeros and poles, these functions are linear fractional transformations. This pattern establishes the Factor Lemma I.

Checking that the pattern is correct is just a matter of inspection. We give two example checks.

  • To see why fc2f_{c_{2}} has a simple zero at t4t_{4} we consider the quintuple

    (k1,2k1,2k2,0,1).(-k-1,-2k-1,-2k-2,0,-1).

    At t4t_{4} the two diagonals Pk1,0P_{-k-1,0} and Pk1,1P_{-k-1,-1} coincide. In terms of the cross ratios of the slopes we are computing χ(a,b,c,d)\chi(a,b,c,d) with a=ba=b. The point P0(t)P_{0}(t) is moving with linear speed and so the zero is simple.

  • To see why gc2g_{c_{2}} has a simple pole at t1t_{1} we consider the 44 points

    P2k+2,k+2P1,k+1,Pk+1,P1,P1,k+1Pk,0.P_{2k+2,k+2}\cap P_{1,k+1},\hskip 10.0ptP_{k+1},\hskip 10.0ptP_{1},\hskip 10.0ptP_{1,k+1}\cap P_{-k,0}. (17)

    These are all contained in the kk-diagonal P1,k+1P_{1,k+1}, which corresponds to the vertex (k1)(-k-1) of Dk(P)D_{k}(P). At t=t1t=t_{1} the three points P0(t)P_{0}(t) and PkP_{-k} and Pk+1P_{k+1} are collinear. This makes the 22nd and 44th listed point coincided. In terms of our cross ratio χ(a,b,c,d)\chi(a,b,c,d) we have b=db=d. This gives us a pole. The pole is simple because the points come together at linear speed.

The other explanations are similar. The reader can see graphical illustrations of these zeros and poles using our program.

2.4 Proof of the Second Result

The proof of Lemma 2.2 is essentially identical to the proof of Lemma 2.1. Here are the changes. The Factor Lemma II has precisely the same statement as the Factor Lemma I, except that

  • When defining P(t)P(t) we use points PaP_{a} and PbP_{b} with |ab|=k+1|a-b|=k+1.

  • We are comparing P(t)P(t) with Dk+1(P(t))D_{k+1}(P(t)).

This changes the definition of the functions ff and gg. With these changes made, the Factor Lemma I is replaced by the Factor Lemma II, which has an identical statement. This time the chords involved are as follows.

k1kk1kk11k12k1k1k2k1k1k2k+1\begin{matrix}-k-1\cr-k\end{matrix}\hskip 15.0pt\begin{matrix}-k-1\cr k\end{matrix}\hskip 15.0pt\begin{matrix}-k-1\cr-1\end{matrix}\hskip 15.0pt\begin{matrix}-k-1\cr-2k-1\end{matrix}\hskip 15.0pt\begin{matrix}-k\cr 1\end{matrix}\hskip 15.0pt\begin{matrix}-k\cr-2k-1\end{matrix}\hskip 15.0pt\begin{matrix}k\cr-1\end{matrix}\hskip 15.0pt\begin{matrix}k\cr 2k+1\end{matrix} (18)

This time the 44 awake indices are:

j0=0,j1=k,j2=k1,j3=k.j_{0}=0,\hskip 10.0ptj_{1}=k,\hskip 10.0ptj_{2}=-k-1,\hskip 10.0ptj_{3}=-k. (19)

Here is the pattern of zeros and poles.

(f,0,1,0),(f,1,7,6),(f,2,3,2),(f,3,5,4).(f,0,1,0),\hskip 20.0pt(f,1,7,6),\hskip 20.0pt(f,2,3,2),\hskip 20.0pt(f,3,5,4). (20)
(g,0,5,6),(g,1,3,0),(g,2,7,4),(g,3,1,2).(g,0,5,6),\hskip 20.0pt(g,1,3,0),\hskip 20.0pt(g,2,7,4),\hskip 20.0pt(g,3,1,2). (21)

The pictures in these cases look almost identical to the previous case. The reader can see these pictures by operating my computer program. Again, the zeros of ff and gg are located at the same places, and likewise the poles of ff and gg are located at the same places. Hence f/gf/g is constant. This completes the proof the Factor Lemma II, which implies Lemma 2.2.

3 The Soul of the Bird

3.1 Goal of the Chapter

Given a polygon P𝑹2P\subset\mbox{\boldmath{$R$}}^{2}, let P^\widehat{P} be the closure of the bounded components of 𝑹2P\mbox{\boldmath{$R$}}^{2}-P and let PIP^{I} be the interior of P^\widehat{P}. (Eventually we will see that birds are embedded, so P^\widehat{P} will be a closed topological disk and PIP^{I} will be an open topological disk.)

Suppose now that P(t)P(t) for t[0,1]t\in[0,1] is a path in Bn,kB_{n,k} starting at the regular nn-gon P(0)P(0). We can adjust by a continuous family of projective transformations so that P(t)P(t) is a bounded polygon in 𝑹2\mbox{\boldmath{$R$}}^{2} for all t[0,1]t\in[0,1]. We orient P(0)P(0) counter-clockwise around PI(0)P^{I}(0). We extend this orientation choice continuously to P(t)P(t). We let Pab(t)P_{ab}(t) denote the diagonal through vertices Pa(t)P_{a}(t) and Pb(t)P_{b}(t). We orient Pa,b(t)P_{a,b}(t) so that it points from Pa(t)P_{a}(t) to Pb(t)P_{b}(t). We take indices mod nn.

We now recall a definition from the introduction: When PP is embedded, we say that PP is strictly star shaped with respect to xPIx\in P^{I} if each ray emanating from xx intersects PP exactly once.

[Uncaptioned image]

Figure 3.1: The soul of a 33-bird

Each such (k+1)(k+1)-diagonal defines an oriented line that contains it, and also the (closed) distinguished half plane which lies to the left of the oriented line. These nn half-planes vary continuously with tt. The soul of P(t)P(t), which we denote S(t)S(t), is the intersection of the distinguished half-planes. Figure 3.1 shows the an example.

The goal of this chapter is to prove the following result.

Theorem 3.1

Let PP be a bird and let SS be its soul. Then:

  1. 1.

    SS is has non-empty interior.

  2. 2.

    SPIS\subset P^{I}.

  3. 3.

    PP is strictly star-shaped with respect to any point in SS.

Theorem 3.1 immediately implies Statement 1 of Theorem 1.1.

We are going to give a homotopical proof of Theorem 3.1. We say that a value t[0,1]t\in[0,1] is a good parameter if Theorem 3.1 holds for P(t)P(t). All three conclusions of Theorem 3.1 are open conditions. Finally, 0 is a good parameter. For all these reasons, it suffices to prove that the set of good parameters is closed. By truncating our path at the first supposed failure, we reduce to the case when Theorem 3.1 holds for all t[0,1)t\in[0,1).

3.2 The Proof

For ease of notation we set X=X(1)X=X(1) for any object XX associated to P(1)P(1).

Lemma 3.2

If PP is any kk-bird, then P0P_{0} and P2k+1P_{2k+1} lie to the left of Pk,k+1P_{k,k+1}. The same goes if all indices are cyclically shifted by the same amount.

Proof: Consider the triangle with vertices P0(t)P_{0}(t) and Pk(t)P_{k}(t) and Pk+1(t)P_{k+1}(t). The kk-niceness condition implies that this triangle is non-degenerate for all t[0,1]t\in[0,1]. Since P0(t)P_{0}(t) lies to to the left of Pk,k+1(t)P_{k,k+1}(t), the non-degeneracy implies the same result for t=1t=1. The same argument works for the triple (2k+1,k,k+1)(2k+1,k,k+1). \spadesuit

Lemma 3.3

SS is non-empty and contained in PIP^{I}.

Proof: By continuity, SS is nonempty and contained in PPIP\cup P^{I}. By the kk-niceness property and continuity, P1P_{1} lies strictly to the right of P0,k+1P_{0,k+1}. Hence the entire half-open edge [P0,P1)[P_{0},P_{1}) lies strictly to the right of P0,k+1P_{0,k+1}. Hence [P0,P1)[P_{0},P_{1}) is disjoint from SS. By cyclic relabeling, the same goes for all the other half-open edges. Hence SP=S\cap P=\emptyset. Hence SPIS\subset P^{I}. \spadesuit

Lemma 3.4

PP is strictly star-shaped with respect to any point of SS.

Proof: Since P(t)P(t) is strictly star-shaped with respect to all points of S(t)S(t) for t<1t<1, this lemma can only fail if there is an edge of PP whose extending line contains a point xSx\in S. We can cyclically relabel so that the edge of P0P1¯\overline{P_{0}P_{1}}.

[Uncaptioned image]

Figure 3.2: The diagonal P0,k+1P_{0,k+1} does not separate 11 from xx.

Since xPx\not\in P, either P1P_{1} lies between P0P_{0} and xx or P0P_{0} lies in between xx and P1P_{1}. In the first case, both P1P_{1} and xx lie on the same side of the diagonal P0,k+1P_{0,k+1}. This is a contradiction: P1P_{1} is supposed to lie on the right and xx is supposed to lie on the left. In the second case we get the same kind of contradiction with respect to the diagonal Pk,1P_{-k,1}. \spadesuit

We say that PP has opposing (k+1)(k+1)-diagonals if it has a pair of (k+1)(k+1)-diagonals which lie in the same line and point in opposite directions. In this case, the two left half-spaces are on opposite sides of the common line.

Lemma 3.5

PP does not have opposing (k+1)(k+1)-diagonals.

Proof: We suppose that PP has opposing diagonals and we derive a contradiction. In this case SS, which is the intersection of all the associated left half-planes, must be a subset of the line LL containing these diagonals. But then PP intersects LL in at least 44 points, none of which lie in SS. But then PP cannot be strictly star-shaped with respect to any point of SS. This is a contradiction. \spadesuit

We call three (k+1)(k+1)-diagonals of P(t)P(t) interlaced if the intersection of their left half-spaces is a triangle. See Figure 3.3.

[Uncaptioned image]

Figure 3.3: Interlaced diagonals on P(t)P(t).

Given interlaced (k+1)(k+1)-diagonals, and a point xx in the intersection, the circle of rays emanating from xx encounters the endpoints of the diagonals in an alternating pattern: a1,b3,a2,b1,a3,b2a_{1},b_{3},a_{2},b_{1},a_{3},b_{2}, where a1,a2,a3a_{1},a_{2},a_{3} are the tail points and b1,b2,b3b_{1},b_{2},b_{3} are the head points. Here a1a_{1} names the vertex Pa1(t)P_{a_{1}}(t), etc.

Lemma 3.6

P(t)P(t) cannot have interlaced diagonals for t<1t<1.

Proof: Choose xS(t)x\in S(t). The nn-gon P(t)P(t) is strictly star-shaped with respect to xx. Hence, the vertices of PP are encountered in order (mod nn) by a family of rays that emanate from xx and rotates around full-circle. Given the order these vertices are encountered, we have aj+1=aj+ηja_{j+1}=a_{j}+\eta_{j}, where ηjk\eta_{j}\leq k. Here we are taking the subscripts mod 33 and the vertex values mod nn. This tells us that n=η1+η2+η33kn=\eta_{1}+\eta_{2}+\eta_{3}\leq 3k. This contradicts the fact that n>3kn>3k. \spadesuit

It only remains to show that SS has non-empty interior. A special case of Helly’s Theorem says the following: If we have a finite number of convex subsets of 𝑹2\mbox{\boldmath{$R$}}^{2} then they all intersect provided that every 33 of them intersect. Applying Helly’s Theorem to the set of interiors of our distinguished half-planes, we conclude that we can find 33 of these open half-planes whose triple intersection is empty. On the other hand, the triple intersection of the closed half-planes contains xx. Since PP has no opposing diagonals, this is only possible if the 33 associated diagonals are interlaced for tt sufficiently close to 11. This contradicts Lemma 3.6. Hence SS has non-empty interior.

4 The Feathers of the Bird

4.1 Goal of the Chapter

Recall that PIP^{I} is the interior of the region bounded by PP. We call the union of black triangles in Figure 4.1 the feathers of the bird. the black region in the center is the soul.

[Uncaptioned image]

Figure 4.1 The feathers of a 33-bird.

Each feather FF of a kk-bird PP is the convex hull of its base, an edge ee of PP, and its tip, a vertex of Δk(P)\Delta_{k}(P).

The goal of this chapter is to prove the following result, which says that the simple topological picture shown in Figure 4.1 always holds.

Theorem 4.1

The following is true.

  1. 1.

    Let FF be an feather with base ee. Then F{e}PIF-\{e\}\subset P^{I}.

  2. 2.

    Distinct feathers can only intersect at a vertex of PP.

  3. 3.

    The line segment connecting two consecutive feather tips lies in PIP^{I}.

When we apply Δk\Delta_{k} to PP we are just specifying the points of Δk(P)\Delta_{k}(P). We define the polygon Δk(P)\Delta_{k}(P) so that the edges are the bounded segments connecting the consecutive tips of the feathers of PP. With this definion, Statement 2 of Theorem 1.1 follows immediately from Theorem 4.1.

4.2 The Proof

There is one crucial idea in the proof of Theorem 4.1: The soul of PP lies in the sector FF^{*} opposite any of its feathers FF. See Figure 4.2.

[Uncaptioned image]

Figure 4.2 The soul lies in the sectors opposite the feathers.

We will give a homotopical proof of Theorem 4.1. By truncating our path of birds, we can assume that Theorem 4.1 holds for all t[0,1)t\in[0,1). We set P=P(1)P=P(1), etc.

Statement 1: Figure 4.3 shows the 22 ways that Statement 1 could fail:

  1. 1.

    The tip vv of the feather FF could coincide with some pPp\in P.

  2. 2.

    Some pPp\in P could lie in the interior point of Fe\partial F-e.

[Uncaptioned image]

Figure 4.3: Case 1 (left) and Case 2 (right).

For all xFx\in F^{*}, the ray xp\overrightarrow{xp} intersects PP both at pp and at a point pep^{\prime}\in e. This contradicts the fact that for any xSFx\in S\subset F^{*}, the polygon PP is strictly star-shaped with respect to xx. This establishes Statement 1 of Theorem 4.1.

Statement 2: Let F1F_{1} and F2F_{2} be two feathers of PP, having bases e1e_{1} and e2e_{2}. For Statement 2, it suffices to prove that F1e1F_{1}-e_{1} and F2e2F_{2}-e_{2} are disjoint.

The same homotopical argument as for Statement 1 reduces us to the case when F1F_{1} and F2F_{2} have disjoint interiors but F1e1\partial F_{1}-e_{1} and F2e2\partial F_{2}-e_{2} share a common point xx. If F1\partial F_{1} and F2\partial F_{2} share an entire line segment then, thanks to the fact that all the feathers are oriented the same way, we would have two (k+1)(k+1) diagonals of PP lying in the same line and having opposite orientation. Lemma 3.5 rules this out.

In particular xx must be the tip of at least one feather. Figure 4.4 shows the case when x=v1x=v_{1}, the tip of F1F_{1}, but xv2x\not=v_{2}. The case when x=v1=v2x=v_{1}=v_{2} has a similar treatment.

[Uncaptioned image]

Figure 4.4: Opposiing sectors are disjoint

In this case, the two sectors F1F_{1}^{*} and F2F_{2}^{*} are either disjoint or intersect in a single point. This contradicts the fact that SF1F2S\subset F_{1}^{*}\subset F_{2}^{*} has non-empty interior. This contradiction establishes Statement 2 of Theorem 4.1.

Statement 3: Recall that P^=PPI\widehat{P}=P\cup P^{I}. Let F1F_{1} and F2F_{2} be consecutive feathers with bases e1e_{1} and e2e_{2} respeectively. Let ff be the edge connecting the tips of F1F_{1} and F2F_{2}. Our homotopy idea reduces us to the case when fP^f\subset\widehat{P} and fPf\cap P\not=\emptyset. Figure 4.5 shows the situation.

[Uncaptioned image]

Figure 4.5: The problem a common boundary point

Note that fPf\cap P must be strictly contained in the interior of ff because (by Statement 1 of Theorem 4.1) the endpoints of ff lie in PIP^{I}. But then, fPf\cap P is disjoint from F1F2F_{1}^{*}\cap F_{2}^{*}, which is in turn contained in the shaded region GG. For any xGx\in G and each vertex pp of ff, the ray the ray xp\overrightarrow{xp} also intersects PP at a point pe1e2p^{\prime}\in e_{1}\cup e_{2}. This gives the same contradiction as above when we take xSF1F2Gx\in S\subset F_{1}^{*}\cap F_{2}^{*}\subset G. This completes the proof of Statement 3 of Theorem 4.1.

5 The Degeneration of Birds

5.1 Statement of Result

Let Bk,nB_{k,n} denote the space of nn-gons which are kk-birds. Let χk\chi_{k} denote the kk-energy. With the value of kk fixed in the background, we say that a degenerating path is a path Q(t)Q(t) of nn-gons such that

  1. 1.

    Q(t)Q(t) is planar for all t[0,1]t\in[0,1].

  2. 2.

    All vertices of Q(t)Q(t) are distinct for all t[0,1]t\in[0,1].

  3. 3.

    Q(t)Bk,nQ(t)\in B_{k,n} for all t[0,1)t\in[0,1) but Q(1)Bk,nQ(1)\not\in B_{k,n}.

  4. 4.

    χk(Q(t))>ϵ0>0\chi_{k}(Q(t))>\epsilon_{0}>0 for all t[0,1]t\in[0,1].

In this chapter we will prove the following result, which will help us prove that Δk(Bk,n)Bk,n\Delta_{k}(B_{k,n})\subset B_{k,n} in the next chapter. The reader should probably just use the statement as a black box on the first reading.

Lemma 5.1 (Degeneration)

If Q()Q(\cdot) is a degenerating path, then all but at most one vertex of Q(1)Q(1) lies in a line segment.

Remark: Our proof only uses the fact that QQ has nontrivial edges, nontrivial kk-diagonals, and nontrivial (k+1)(k+1)-diagonals. Some of the other vertices could coincide and it would not matter. Also, the same proof works if, instead of a continuous path, we have a convergent sequence {Q(tn)}\{Q(t_{n})\} with tn1t_{n}\to 1 and a limiting polygon Q(1)=limQ(tn)Q(1)=\lim Q(t_{n}).

Example: Let us give an example for the case k=1k=1 and n=5n=5. Figure 5.0 shows a picture of a pentagon Q(t)Q(t) for t=1st=1-s.

[Uncaptioned image]

Figure 5.0: A degenerating path in the case k=1k=1 and n=5n=5.

Here ss ranges from 11 to 0 as tt ranges from 0 to 11. We have labeled some of the slopes to facility the calculation (which we leave to the reader) that χ1(Q(t))\chi_{1}(Q(t)) remains uniformly bounded away from 0.

5.2 Distinguished Diagonals

We orient Q(t)Q(t) so that it goes counter-clockwise around the region it bounds. We orient the diagonal QabQ_{ab} so that it points from QaQ_{a} to QbQ_{b}. For t<1t<1 the vertices Q1(t)Q_{1}(t) and Qk(t)Q_{k}(t) lie to the right of the diagonal Q0,k+1(t)Q_{0,k+1}(t), in the sense that a person walking along this diagonal according to its orientation would see that points in the right. This has the same proof as Lemma 3.2. The same rule holds for all cyclic relabelings of these points. The rule holds when t<1t<1. Taking a limit, we get a weak version of the rule: Each of Q1(1)Q_{1}(1) and Qk(1)Q_{k}(1) either lies to the right of the diagonal Q0,k+1(1)Q_{0,k+1}(1) or on it. The same goes for cyclic relabeings. We call this the Right Hand Rule.

Say that a distinguished diagonal of Q(t)Q(t) is either a kk-diagonal or a (k+1)(k+1)-diagonal. There are 2n2n of these, and they come in a natural cyclic order:

Q0,k(t)Q0,k+1(t),Q1,k+1(t),Q1,k+2(t),Q_{0,k}(t)\hskip 10.0ptQ_{0,k+1}(t),\hskip 10.0ptQ_{1,k+1}(t),\hskip 10.0ptQ_{1,k+2}(t),... (22)

The pattern alternates between kk and (k+1)(k+1)-diagonals. We say that a diagonal chain is a consecutive list of these.

We say that one oriented segment L2L_{2} lies ahead of another one L1L_{1} if we can rotate L1L_{1} by θ(0,π)\theta\in(0,\pi) radians counter-clockwise to arrive at a segment parallel to L2L_{2}, In this case we write L1L2L_{1}\prec L_{2}. We have

Q0,k+1(t)Q1,k+1(t)Q1,k+2(t)Q2,k+2(t).Q_{0,k+1}(t)\prec Q_{1,k+1}(t)\prec Q_{1,k+2}(t)\prec Q_{2,k+2}(t). (23)
[Uncaptioned image]

Figure 5.1: The turning rule

This certainly holds when t=0t=0. By continuity and the Right Hand Rule, this holds for all t<1t<1. Taking a limit, we see that the kk-diagonals of Q(1)Q(1) weakly turn counter-clockwise in the sense that either L1L2L_{1}\prec L_{2} for consecutive diagonals or else L1L_{1} and L2L_{2} lie in the same line and point in the same direction. Moreover, the total turning is 2π2\pi. If we start with one distinguished diagonal and move through the cycle then the turning angle increases by jumps in [0,π][0,\pi] until it reaches 2π2\pi. We call these observations the Turning Rule.

5.3 Collapsed Diagonals

Figure 5.2 shows the distinguished diagonals incident to Q0Q_{0}. We always take indices mod nn. Thus k1=nk1-k-1=n-k-1 mod nn.

[Uncaptioned image]

Figure 5.2: The 44 distinguished diagonals incident to Q0(t)Q_{0}(t).

We say that QQ has collapsed diagonals at a vertex if QQ if the 44 distinguished diagonals incident to QkQ_{k} do not all lie on distinct lines. We set Q=Q(1)Q=Q(1). We set X=X(1)X=X(1) for each object XX associated to Q(1)Q(1).

Since QQ is planar but not kk-nice, QQ must have collapsed diagonals at some vertex. We relabel so that the collapsed diagonals are at Q0Q_{0}.

Lemma 5.2

If QQ has collapsed diagonals at Q0Q_{0} then Qk1,0Q_{-k-1,0} and Q0,k+1Q_{0,k+1} point in opposite directions or Qk,0Q_{-k,0} and Q0,kQ_{0,k} point in the same direction.

Proof: Associated to each diagonal incident to Q0Q_{0} is the ray which starts at Q0Q_{0} and goes in the direction of the other endpoint of the diagonal. (Warning: The ray may have the opposite orientation than the diagonal it corresponds to.) If the angle between any of the rays tends to π\pi as t1t\to 1 then the angle between the outer two rays tends to π\pi. In this case Qk,0Q_{-k,0} and Q0,kQ_{0,k} point in the same directions. If the angle between non-adjacent rays tends to 0 then Qk1,0Q_{-k-1,0} and Q0,k+1Q_{0,k+1} are squeezed together and point in opposite directions.

Suppose that the angle between adjacent rays tends to 0. If the two adjacent rays are the middle ones, we have the case just considered. Otherwise, either the angle between the two left rays tends to 0 or the angle between the two right rays tends to 0. In either case, the uniform lower bound on the cross ratio forces a third diagonal to point either in the same or the opposite direction as these adjacent diagonals when t=1t=1. Any situation like this leads to a case we have already considered. \spadesuit

5.4 The Case of Aligned Diagonals

We say that QQ has aligned diagonals at the vertex Q0Q_{0} if Qk,0Q_{-k,0} and Q0,kQ_{0,k} are parallel. This is the second option in Lemma 5.2. We make the same kind of definition at other vertices, with the indices shifted in the obvious way,.

Lemma 5.3

Suppose QQ does not lie in a single line. Suppose also that QQ has aligned diagonals at Q0Q_{0}. Then the diagonals Qk,0,Qk,1,,Q1,k,Q0,kQ_{-k,0},Q_{-k,1},...,Q_{-1,k},Q_{0,k} all are parallel and (hence) the 2k+12k+1 points Qk,,Q0,,QkQ_{-k},...,Q_{0},...,Q_{k} are contained in the line defined by these diagonals.

Proof: These two diagonals define a short chain of diagonals, which starts with the first listed diagonal and ends with the second one. They also define a long chain, which starts with the second and ends with the first. The total turning of the diagonals is 2π2\pi, so one of the two chains defined by our diagonals turns 2π2\pi and the other turns 0. Suppose first that the long chain has 0 turning. This chain involves all points of QQ, and forces all points of QQ to be on the same line. So, the short chain must consist of parallel diagonals. \spadesuit

All we use in the rest of the proof is that Qk,,QkQ_{-k},...,Q_{k} are all contained in a line LL. By shifting our indices, we can assume that Qk+1LQ_{k+1}\not\in L. This relabeling trick comes with a cost. Now we cannot say whether the points Qk.QkQ_{-k}....Q_{k} come in order on LL. We now regain this control.

Lemma 5.4

The length 2k2k-diagonal chain Qk,0Q0,kQ_{-k,0}\to...\to Q_{0,k} consists entirely of parallel diagonals. There is no turning at all.

Proof: The diagonals Qk,0Q_{-k,0} and Q0,kQ_{0,k}. are either parallel or anti-parallel. If they are anti-parallel, then the angle between the corresponding rays incident Q0(t)Q_{0}(t) tends to 0 as t1t\to 1. But these are the outer two rays. This forces the angle between all 44 rays incident to Q0(t)Q_{0}(t) to tend to 0. The whole picture just folds up like a fan. But one or these rays corresponds to Q0,k+1(t)Q_{0,k+1}(t). This picture forces Qk+1LQ_{k+1}\in L. But this is not the case.

Now we know that Qk,0Q_{-k,0} and Q0,kQ_{0,k} are parallel. All the diagonals in our chain are either parallel or anti-parallel to the first and last ones in the chain. If we ever get an anti-parallel pair, then the diagonals in the chain must turn 2π2\pi around. But then none of the other distinguished diagonals outside our chain turns at all. As in Lemma 5.3, this gives QLQ\subset L, which is false. \spadesuit

We rotate the picture so that LL coincides with the XX-axis and so that Q0,kQ_{0,k} points in the positive direction. Since we are already using the words left and right for another purpose, we say that pLp\in L is forward of of qLq\in L if pp has larger XX-coordinate. Likewise we say that qq is backwards of pp in this situation. We say that Q0,kQ_{0,k} points forwards. We have established that Qk,0,,Q0,kQ_{-k,0},...,Q_{0,k} all point forwards.

Lemma 5.5

Qk+2LQ_{k+2}\in L and both Q1,k+2Q_{1,k+2} and Q2,k+2Q_{2,k+2} point backwards.

Proof: We have arranged that Qk+1LQ_{k+1}\not\in L. Let us first justify the fact that Qk+1Q_{k+1} lies above LL. This follows from Right Hand Rule applied to Q0,k+1Q_{0,k+1} and QkQ_{k} and the fact that Q0,kQ_{0,k} points forwards. Since Qk,Qk+1,Q1Q_{-k},Q_{-k+1},Q_{1} are collinear, QQ has collapsed diagonals at Q1Q_{1}. But QQ cannot have aligned diagonals because Q1,k+1Q_{1,k+1} is not parallel to Qk,1Q_{-k,1}. Hence QQ has folded diagonals at 11. This means that the diagonals Qk,1Q_{-k,1} and Q1,k+2Q_{1,k+2} point in opposite directions. This forces Qk+2LQ_{k+2}\in L and morever we can say that Q1,k+2Q_{1,k+2} points backwards.

We have Q2LQ_{2}\in L because 2k2\leq k. We want to see that Q2,k+2Q_{2,k+2} points forwards and they Suppose not. We consider the 33 distinguished diagonals

Q0,k,Q1,k+2,Q2,k+2.Q_{0,k},\hskip 10.0ptQ_{1,k+2},\hskip 10.0ptQ_{2,k+2}.

These diagonals respectively point forwards, backwards, forwards and they all point one direction or the other along LL. But then, in going from Q0,kQ_{0,k} to Q2,k+2Q_{2,k+2}, the diagonals have already turned 2π2\pi. Since the total turn is 2π2\pi, the diagonals Q2,k+2,Q3,k+3,,Qn,n+kQ_{2,k+2},\ Q_{3,k+3},...,Q_{n,n+k} are all parallel. But then Q2,,QnLQ_{2},...,Q_{n}\in L. This contradicts the fact that Qk+1LQ_{k+1}\not\in L. \spadesuit

Lemma 5.6

For at least one of the two indices j{2k+2,2k+3}j\in\{2k+2,2k+3\} we have QjLQ_{j}\in L and Qk+2,jQ_{k+2,j} points forwards.

Proof: Since Q1,Q2,Qk+2Q_{1},Q_{2},Q_{k+2} are collinear, QQ has collapsed diagonals at Qk+2Q_{k+2}. So, by Lemma 5.2, we either have folded diagonals at Qk+2Q_{k+2} or aligned diagonals at Qk+2Q_{k+2}. The aligned case gives Q2k+2LQ_{2k+2}\in L and the folded case gives Q2k+3LQ_{2k+3}\in L. We need to work out the direction of pointing in each case.

Consider the aligned case. Suppose Qk+2,2k+2Q_{k+2,2k+2} points backwards, as shown in Figure 5.3.

[Uncaptioned image]

Figure 5.3: Violation of the Right Hand Rule

This violates the Right Hand Rule for Qk+2Q_{k+2} and Qk+1,2k+2Q_{k+1,2k+2} because Qk+1Q_{k+1} lies above LL.

Consider the folded case. Since Qk+2,2k+3Q_{k+2,2k+3} and Q1,k+2Q_{1,k+2} point in opposite directions, and Q1,k+2Q_{1,k+2} points backwards (by the previous lemma), Qk+2,2k+3Q_{k+2,2k+3} points forwards. \spadesuit

Let j{2k+2,2k+3}j\in\{2k+2,2k+3\} be the index from Lemma 5.6. Consider the 33 diagonals

Q0,k,Q1,k+1,Qk+2,j.Q_{0,k},\hskip 10.0ptQ_{1,k+1},\hskip 10.0ptQ_{k+2,j}.

These diagonals are all parallel to LL and respectively point in the forwards, backwards, forwards direction. This means that the diagonals in the chain Q0,kQk+2,jQ_{0,k}\to...\to Q_{k+2,j} have already turned 2π2\pi radians. But this means that the diagonals

Qk+2,2k+3,Qk+3,2k+3,Qk+3,2k+4,Q0,k=Qn,n+kQ_{k+2,2k+3},\hskip 10.0ptQ_{k+3,2k+3},\hskip 10.0ptQ_{k+3,2k+4},\hskip 10.0pt...\hskip 10.0ptQ_{0,k}=Q_{n,n+k}

are all parallel and point forwards along LL. Hence Qk+2,Qk+3,,QnLQ_{k+2},Q_{k+3},...,Q_{n}\in L. Hence all points but Qk+1Q_{k+1} lie in LL.

5.5 The Case of Double Folded Diagonals

We fix a value of kk. Say that two indices a,b𝒁/na,b\in\mbox{\boldmath{$Z$}}/n are far if their distance is at least kk in 𝒁/n\mbox{\boldmath{$Z$}}/n. We say that QQ has far folded diagonals if QQ has folded diagonals at QaQ_{a} and QQ has folded diagonals at bb and a,ba,b are far.

In this case we have two parallel diagonals Qa,a+k+1Q_{a,a+k+1} and Qb,b+k+1Q_{b,b+k+1}. As in the proof of Lemma 5.3, one of the two diagonal chains defined by these diagonals consists of parallel diagonals. The far condition guarantees that at least 2k+12k+1 consecutive points are involved in each chain. But then we get 2k+12k+1 collinear points. So, if QQ has far folded diagonals, then the same proof as in the previous section shows that the conclusion of the Degeneration Lemma holds for QQ.

5.6 Good Folded Diagonals

We say that the folded diagonals Qk1,0Q_{-k-1,0} and Q0,k+1Q_{0,k+1} are good if all the points Qk+1,Qk+2,,Qnk1Q_{k+1},Q_{k+2},...,Q_{n-k-1} are collinear. This notion is empty when k=2k=2 and n=7n=7 but otherwise it has content. In this section we treat the case when QQ has a pair of good folded diagonals. We start by discussing an auxiliary notion.

We say that QQ has backtracked edges at QaQ_{a} if the angle between the edges Qa,a+1Q_{a,a+1} and Qa,a1Q_{a,a-1} is either 0 or 2π2\pi.

Lemma 5.7

If QQ has backtracked edges at QaQ_{a} then QQ has folded diagonals at QaQ_{a}.

Proof: For t[0,1)t\in[0,1), the edges of QQ emanating from aa divide the plane into 44 sectors, and one of these sectors, C(t)C(t) contains all the distinguished diagonals emanating from Qa(t)Q_{a}(t). The sector C(t)C(t) is the one which locally intersects Q(t)Q(t) near Qa(t)Q_{a}(t). The angle of C(t)C(t) tends to 0 as t1t\to 1, forcing all the distinguished diagonals emanating from Qa(t)Q_{a}(t) to squeeze down as t1t\to 1. This gives us the folded diagonals. \spadesuit

We will use Lemma 5.7 in our analysis of good folded edges. Now we get to it. We rotate so that our two diagonals are in the line LL, which is the XX-axis. We normalize so that Q0Q_{0} is the origin, and Qk+1Q_{k+1} and Qk1Q_{-k-1} are forward of Q0Q_{0}.

Lemma 5.8

If n>3k+1n>3k+1 and Qk1,0,Q0,k+1Q_{-k-1,0},Q_{0,k+1} are good folded diagonals, then the Degeneration Lemma is true for QQ.

Proof: Suppose first that Q1LQ_{1}\in L. Then QQ has folded diagonals at Qk+1Q_{k+1}. When n>3k+1n>3k+1 the indices (k+1)(k+1) and (k1)(-k-1) are kk-far. This gives QQ far folded diagonals, a case we have already treated.

To finish our proof, we show that Q1LQ_{1}\in L. We explore some of the other points. We know that Qk+1,,Qnk1LQ_{k+1},...,Q_{n-k-1}\in L. We can relabel dihedrally so that Qnk1Q_{n-k-1} is forwards of Qk+1Q_{k+1}. We claim that Qk+2Q_{k+2} is forwards of Qk+1Q_{k+1}. Suppose not. Then there is some index a{k+2,,k2}a\in\{k+2,...,-k-2\} such that QaQ_{a} is backwards of Qa±1Q_{a\pm 1}. What is going on is that our points would start by moving backwards on LL and eventually they have to turn around. The index aa is the turn-around index. This gives us backtracked edges at QaQ_{a}. By Lemma 5.7, we have folded diagonals at QaQ_{a}. But aa and 0 are kk-far indices. This gives QQ far-folded diagonals.

The only way out of the contradiction is that Qk+2Q_{k+2} is forwards of Qk+1Q_{k+1}.

[Uncaptioned image]

Figure 5.4: A contradiction involving Q1Q_{1}.

Suppose Q1LQ_{1}\not\in L. by the Right Hand Rule applied to the diagonal Q0,k+1Q_{0,k+1}, the point Q1Q_{1} lies beneath LL, as shown in Figure 5.4. But then Qk+1Q_{k+1} lies to the left of the diagonal Q1,k+2Q_{1,k+2}. This violates the Right Hand Rule. Now we know that Q1LQ_{1}\in L. \spadesuit

Lemma 5.9

Suppose n=3k+1n=3k+1 and k>2k>2. If Qk1,0,Q0,k+1Q_{-k-1,0},Q_{0,k+1} are good folded diagonals, then the Degeneration Lemma is true for QQ.

Proof: The same argument as in Lemma 5.8 establishes the key containment Q1LQ_{1}\in L. (We need k>2k>2 for this.) From here, as in Lemma 5.8, we deduce that Qk1,0Q_{-k-1,0} and Qk+1,2k+2Q_{k+1,2k+2} are parallel. This time the conclusion we get from this is not as good. We get a run of kk points in LL, and then a point not necessarily in LL, and then a run of kk additional points in LL.

The points are Qk+1,,Q2k+1,,Q0Q_{k+1},...,Q_{2k+1},...,Q_{0} with the point QkQ_{-k} omitted. But then QQ has folded diagonals at each of these points except the outer two, Qk+1Q_{k+1} and Q0Q_{0}. But then For each such index hh, we see that both Qh±(k+1)Q_{h\pm(k+1)} belong to LL. This gives us all but one point in LL.

It is instructive to consider an example, say k=4k=4 and n=13n=13. In this case, our initial run of points in LL is Q5,Q6,Q7,Q8,Q10,Q11,Q12,Q13.Q_{5},Q_{6},Q_{7},Q_{8},Q_{10},Q_{11},Q_{12},Q_{13}. The folded diagonals at Q6,Q7,Q8Q_{6},Q_{7},Q_{8} respectively give Q1,Q2,Q3LQ_{1},Q_{2},Q_{3}\in L. The folded diagonals at Q10,Q11,Q12Q_{10},Q_{11},Q_{12} respectively give Q2,Q3,Q4LQ_{2},Q_{3},Q_{4}\in L. \spadesuit

Finally we consider the case k=2k=2 and n=7n=7. In this case all we know is that Q0,Q3,Q4LQ_{0},Q_{3},Q_{4}\in L with Q3,Q4Q_{3},Q_{4} forwards of Q0Q_{0}. We can dihedrally relabel to that Q4Q_{4} is forwards of Q3Q_{3}. Here Q3=Qk+1Q_{3}=Q_{k+1} and Q4=Qk+2Q_{4}=Q_{k+2}. So, now we can run the same argument as in Lemma 5.9 to conclude that Q1LQ_{1}\in L. Now we proceed as in the proof of Lemma 5.9.

5.7 Properties of the Soul

We define S=S(1)S=S(1) to be the set of all accumulation points of sequences {p(tn)}\{p(t_{n})\} where p(tn)S(tn)p(t_{n})\in S(t_{n}) and tn1t_{n}\to 1. Here S(tn)S(t_{n}) is the soul of P(tn)P(t_{n}). We have one more case to analyze, namely ungood folded diagonals. To make our argument go smoothly, we first prove some properties about SS.

Lemma 5.10

Suppose that QQ has folded diagonals at Q0Q_{0}. If the Degeneration Lemma is false for QQ, then SS is contained in the line segment joining Q0Q_{0} to Qk+1Q_{k+1}

Proof: Here is a general statement about SS. Since S(t)S(t) is non-empty and closed for all t<1t<1, we see by compactness that SS is also a non-empty closed subset of the closed region bounded by QQ. By continuity SS lies to the left of all the closed half-planes defined by the oriented (k+1)(k+1) diagonals (or in their boundaries). Since SS lies to the left of (or on) each (k+1)(k+1) diagonal, SS is a subset of the line LL common to the folded diagonals and indeed SS lies to one side of the fold point Q0Q_{0}. From the way we have normalized, SS lies in the XX-axis forward of Q0Q_{0}. (The point Q0Q_{0} might be an endpoint of SS.)

If SS contains points of LL that lie forward of Qk+1Q_{k+1} then either the diagonal Qk+1,2k+2Q_{k+1,2k+2} points along the positive XX-axis or into the lower half-plane. In the former cases, the diagonals Q0,k+1,Qk+1,2k+2Q_{0,k+1},Q_{k+1,2k+2} are parallel and we get at least 2k+12k+1 collinear points and so the Degeneration Lemma holds for QQ.

If Qk+1,2k+2Q_{k+1,2k+2} points into the negative half-plane, then the diagonal Q0,k+1Q_{0,k+1} turns more than π\pi degrees before reaching Qk+1,2k+2Q_{k+1,2k+2}. But then the diagonals in the chain Qk1,0Q0,k+1Qk+1,2k+2Q_{-k-1,0}\to...\to Q_{0,k+1}...\to Q_{k+1,2k+2} turn more than 2π2\pi degrees. This is a contradiction. \spadesuit

Remark: The same argument works with Qk1Q_{-k-1} in place of Qk+1Q_{k+1}.

Lemma 5.11

If the Degeneration Lemma is false for QQ then SS cannot intersect QQ in the interior of an edge of QQ.

Proof: Suppose this happens. We relabel so that the edge is Q0,1Q_{0,1}. By the Right Hand Rule, the point Q1Q_{1} is not on the left of the diagonal Q0,k+1Q_{0,k+1}. At the same time, SS is not on the right of the diagonal. The only possibility is that Q1,Q0,Qk+1Q_{1},Q_{0},Q_{k+1} are collinear. Likewise Qk,Q0,Q1Q_{-k},Q_{0},Q_{1} are collinear. Furtheremore, the (k+1)(k+1)-diagonals Qk,1Q_{-k,1} and Q0,k+1Q_{0,k+1} are parallel. Figure 5.5 shows the situation for Q(t)Q(t) and S(t)S(t) when tt is very near 11.

[Uncaptioned image]

Figure 5.5: The relevant points and lines.

But now we have two (k+1)(k+1)-diagonals that are parallel and which start at indices that are kk apart in 𝒁/n\mbox{\boldmath{$Z$}}/n. This gives us 2k+12k+1 consecutive collinear points on the line containing our edge. We know how to finish the Degeneration Lemma in this case. The only way out is that SS cannot intersect QQ in the interior of an edge of QQ. \spadesuit

Lemma 5.12

If the Degeneration Lemma is false for QQ, then SS cannot contain a vertex of QQ.

Proof: We relabel so that Q0SQ_{0}\in S. The same analysis as in the previous lemma shows that Q1,Q0,QkQ_{1},Q_{0},Q_{-k} are collinear. Figure 5.6. shows the situation for tt near 11. At the same time, the points Q1,Q0,QkQ_{-1},Q_{0},Q_{k} are collinear.

[Uncaptioned image]

Figure 5.6: The relevant points and lines

To avoid a case of the Degeneration Lemma we have already done, QQ must have folded diagonals at QkQ_{-k}. Likewise QQ must have folded diagonals at QkQ_{k}. But then QQ has far folded diagonals, and the Degeneration Lemma holds for QQ. \spadesuit

Now let us bring back our assumptions: QQ has folded diagonals at Q0Q_{0} and the points Q0,Qk+1,Qk1Q_{0},Q_{k+1},Q_{-k-1} all lie in the XX-axis in the forward order listed.

Corollary 5.13

If the Degeneration Lemma is false for QQ then SS lies in the open interval bounded by Q0Q_{0} and Qk+1Q_{k+1} and no point of SS lies in QQ. In particular, SS contains a point xx, forwards of Q0Q_{0} and backwards of both Qk+1Q_{k+1} and Qk1Q_{-k-1}, that is disjoint from QQ.

5.8 Ungood Folded Diagonals

The only case left is when QQ does not have 2k+12k+1 consecutive collinear points, and when all folded diagonals of QQ are ungood. Without loss of generality, we will consider the case when QQ has ungood folded diagonals at Q0Q_{0}. We normalize as in the previous section, so that Q0,Qk+1,Qk1Q_{0},Q_{k+1},Q_{-k-1} lie in forward order on LL, which is the XX-axis. Let xx be a point from Corollary 5.13.

We call an edge of QQ escaping if eLe\cap L is a single point. We call two different edges of QQ, in the labeled sense, twinned if they are both escaping and if they intersect in an open interval. Even if two distinctly labeled edges of QQ coincide, we consider them different as labeled edges.

Lemma 5.14

QQ cannot have twinned escaping edges.

Proof: Consider Q(t)Q(t) for tt near 11. This polygon is strictly star shaped with respect to a point x(t)x(t) near xx.

[Uncaptioned image]

Figure 5.7: Rays intersecting the twinned segments

There is a disk DD about xx such that every pDp\in D contains a ray which intersects the twinned edges in the middle third portion of their intersection. Figure 5.7 shows what we mean. Once tt is sufficiently near 11, the soul S(t)S(t) will intersect DD, and for all points pDp\in D there will be a ray which intersects Q(t)Q(t) twice. This contradicts the fact that Q(t)Q(t) is strictly star-shaped with respect to all points of S(t)S(t). \spadesuit

We say that an escape edge rises above LL if it intersects the upper half plane in a segment.

Lemma 5.15

QQ cannot have two escape edges which rise above LL and intersect QQ on the same side of the point xx.

Proof: This situation is similar to the previous proof. In this case, there is a small disk DD about xx such that every point pDp\in D has a ray which intersects both rising escape edges transversely, and in the middle third of each of the two subsegments of these escape edges that lie above LL. Figure 5.8 shows this situation.

[Uncaptioned image]

Figuren 5.8: Rays intersecting the rising segments.

In this case, some part of Q(t)Q(t) closely shadows our two escape edges for tt near 11. But then, once tt is sufficiently near 11, each ray we have been talking about intersects Q(t)Q(t) at least twice, once by each escaping edge. This gives the same contradiction as in the previous lemma. \spadesuit

We define falling escape segments the same way. The same statement as in Lemma 5.15 works for falling escape segments. Since xQx\not\in Q we conclude that QQ can have at most 44 escaping segments total.

But Q=Q+QQ=Q_{+}\cup Q_{-}, where Q±Q_{\pm} is an arc of QQ that starts at Qk+1Q_{k+1} and ends at Qk1Q_{-k-1}. Since both these arcs start and end on LL, and since both do not remain entirely on LL, we see that each arc has at least 22 escape edges, and none of these are twinned. This means that both Q+Q_{+} and QQ_{-} have exactly two escape edges.

Now for the moment of truth: Consider Q+Q_{+}. Since Q+Q_{+} just has 22 escape edges, they both have to be either rising or falling. Also, since Q+Q_{+} starts and ends on the same side of xx, and cannot intersect xx, both the escape edges for Q+Q_{+} are on the same side of xx. This is a contradiction. The same argument would work for QQ_{-} but we don’t need to make it.

6 The Persistence of Birds

In this chapter we prove Statement 3 of Theorem 1.1, namely the fact that Δk(Bn,k)=Bn,k\Delta_{k}(B_{n,k})=B_{n,k}. First we use the Degeneration Lemma to prove that Δk(Bn,k)Bn,k\Delta_{k}(B_{n,k})\subset B_{n,k}. Then we deduce the opposite containment from projective duality and from the factoring of Δk\Delta_{k} given in §2.2.

6.1 Containment

Suppose for the sake of contradiction that there is some PBk,nP\in B_{k,n} such that Δ(P)Bk,n\Delta(P)\not\in B_{k,n}. Recall that there is a continuous path P(t)P(t) for t[0,1]t\in[0,1] such that P(0)P(0) is the regular nn-gon.

Define Q(t)=Δk(P(t))Q(t)=\Delta_{k}(P(t)). There is some t0[0,1]t_{0}\in[0,1] such that Q(t0)Bk,nQ(t_{0})\not\in B_{k,n}. We can truncate our path so that t0=1t_{0}=1. In other words, Q(t)Bn,kQ(t)\in B_{n,k} for t[0,1)t\in[0,1) but Q(1)Bk,nQ(1)\not\in B_{k,n}.

Lemma 6.1

Q()Q(\cdot) is a degenerating path.

Proof: Note that Q()Q(\cdot) is planar and hence satisfies Property 1 for degenerating paths. Let P=P(1)P=P(1) and Q=Q(1)Q=Q(1). If QQ doe not have all distinct vertices then two different feathers of PP intersect at a point which (by Statement 2 of Theorem 1.1) lies in PIP^{I}. This contradicts Statement 2 of Theorem 4.1. Hence Q()Q(\cdot) satisfies Property 2 for degenerating paths. By construction, Q(t)Bn,kQ(t)\in B_{n,k} for all t[0,1)t\in[0,1). Hence Q()Q(\cdot) satisfies Property 3. The energy χk\chi_{k} is well-defined and continuous on Bk,nB_{k,n}. Hence, by compactness, χk(P(t))>ϵ0\chi_{k}(P(t))>\epsilon_{0} for some ϵ0>0\epsilon_{0}>0 and all t[0,1]t\in[0,1]. Now for the crucial step: We have already proved that χkΔk=χk\chi_{k}\circ\Delta_{k}=\chi_{k}. Hence χk(Q(t))>ϵ0\chi_{k}(Q(t))>\epsilon_{0} for all t[0,1]t\in[0,1]. That is, Q()Q(\cdot) satisfies Property 4 for degenerating paths. \spadesuit

Now we apply the Degeneration Lemma to Q()Q(\cdot). We conclude that all but at most 11 vertex of Q(1)Q(1) lies in a line LL. Stating this in terms of P(1)P(1), we can say that all but at most one of the feathers of P(1)P(1) have their tips in a single line LL. Call an edge of P(1)P(1) ordinary if the feather associated to it has its tip in LL. We call the remaining edge, if there is one, special. Thus, all but at most one edge of PP is ordinary.

Let S(t)S(t) be the soul of P(t)P(t). We know that S(1)S(1) has non-empty interior by Theorem 3.1. For ease of notation we set P=P(1)P=P(1) and S=S(1)S=S(1).

Lemma 6.2

PP cannot have ordinary edges e1e_{1} and e2e_{2} that lie on opposite sides of LL and are disjoint from LL.

Proof: Suppose this happens. Figure 6.1 shows the situation.

[Uncaptioned image]

Figure 6.1: Two feathers on opposite sides of LL.

Let F1F_{1} and F2F_{2} be the two associated feathers. Then the opposite sector F1F_{1}^{*} lies above LL, and the opposite sector F2F_{2}^{*} lies below LL and the two tips are distinct. But then S(1)S(1), which must lie in the intersection of these sectors, is empty. \spadesuit

Lemma 6.3

PP cannot have more than 22 ordinary edges which intersect LL.

Proof: Note that an ordinary edge cannot lie in LL because then the tip would not. So, an ordinary edge that intersects LL does so either at a single vertex or at an interior point. As we trace along LL in one direction or the other we encounter the first intersecting edge and then the last one and then some other intersecting edge. Let F1.F2.F3F_{1}.F_{2}.F_{3} be the two feathers, as shown in Figure 6.3. Let eje_{j} be the edge of FjF_{j} that belongs to PP. Let vjv_{j} be the tip of FjF_{j}. (Figure 6.3 shows the case when ejLe_{j}\cap L is an interior point of eje_{j} for each j=1,2,3j=1,2,3, but the same argument would work if some of these intersection points were vertices.)

[Uncaptioned image]

Figure 6.2: Three or more crossing edges

One of the two arcs α\alpha of QQ joining v1v_{1} to v2v_{2} stays in LL, namely the one avoiding the one point of QQ not on LL. However, α\alpha passes right through F3F_{3} and in particular crosses e3e_{3} transversely. However, one side of F3F_{3} is outside PP. Hence α\alpha is not contained in PIP^{I}, the interior of the region bounded by PP. This contradicts Statement 2 of Theorem 1.1, which says that QPIQ\subset P^{I}. \spadesuit

The line LL divides the plane into two open half-planes, which we call the sides of LL. Lemma 6.2 says that PP cannot have ordinary edges contained in opposite sides of LL. Lemma 6.3 says that at most 22 ordinary edges can intersect LL. Hence, all but at most 22 of the ordinary edges of PP lie on one side of LL. Call this the abundant side of LL. Call the other side the barren side. The barren side contains no ordinary edges at all, and perhaps the special edge. In particular, at most two vertices of PP lie in the barren side.

[Uncaptioned image]

Figure 6.3: Following the diagonals bounding a feather

At the same time, each ordinary edge on the abundant side contributes two vertices to the barren side: We just follow the diagonals comprising the corresponding feather. These diagonals cross LL from the abundant side into the barren side. Two different ordinary edges contribute at least 33 distinct vertices to the barren side. This is a contradiction.

We have ruled out all possible behavior for P=P(1)P=P(1) assuming that Q=Q(1)Q=Q(1) is degenerate. Hence, Q(1)Q(1) is not degenerate. This means that Q(1)Q(1) is a bird. This completes the proof that

Δk(Bk,n)Bk,n.\Delta_{k}(B_{k,n})\subset B_{k,n}. (24)

6.2 Equality

We use the notation from §2.2. Equation 8 implies that

Δk1=Dk+1ΔkDk+1.\Delta_{k}^{-1}=D_{k+1}\circ\Delta_{k}\circ D_{k+1}. (25)

So far, Equation 25 makes sense in terms of PolyPoints and PolyLines.

Below we will explain how to interpret Dk+1D_{k+1} as a map from polygons in 𝑷P to polygons in 𝑷\mbox{\boldmath{$P$}}^{*} and also as a map from polygons in 𝑷\mbox{\boldmath{$P$}}^{*} to polygons in 𝑷P. Since the dual projective plane 𝑷\mbox{\boldmath{$P$}}^{*} is an isomorphic copy of 𝑷P, it makes sense to define Bk.nB_{k.n}^{*}. This space is just the image of Bk,nB_{k,n} under any projective duality. Below we will prove

Theorem 6.4

Dk+1(Bk,n)Bk,nD_{k+1}(B_{k,n})\subset B^{*}_{k,n}.

It then follows from projective duality that Dk+1(Bk,n)Bk,nD_{k+1}(B^{*}_{k,n})\subset B_{k,n}. Combining these equations with Equation 25 we see that Δk1(Bn,k)Bn,k\Delta_{k}^{-1}(B_{n,k})\subset B_{n,k}. This combines with Equation 24 to finish the proof of Theorem 1.1.

Now we prove Theorem 6.4.

Lemma 6.5

If PBk,nP\in B_{k,n}, then we can enhance Dk+1(P)D_{k+1}(P) in such a way that Dk+1(P)D_{k+1}(P) is a planar polygon in 𝐏\mbox{\boldmath{$P$}}^{*}. The enhancement varies continuously.

Proof: A polygon is a PolyPoint together with additional data specifying an edge in 𝑷P joining each consecutive pair of points. Dually, we get a polygon in 𝑷\mbox{\boldmath{$P$}}^{*} from a PolyLine by specifying, for each pair of consecutive lines Lj,Lj+1L_{j},L_{j+1}, an arc of the pencil of lines through the intersection point which connects LjL_{j} to Lj+1L_{j+1}.

Specifying an enhancement of Dk+1(P)D_{k+1}(P) is the same as specifing, for each consecutive pair L1,L2L_{1},L_{2} of (k+1)(k+1) diagonals of PP, an arc of the pencil through their intersection that connects L1,L2L_{1},L_{2}. There are two possible arcs. One of them avoids the interior of the soul of PP and the other one sweeps through the soul of PP. We choose the arc that avoids the soul interior. Figure 6.4 shows that we mean for a concrete example.

[Uncaptioned image]

Figure 6.4: Enhancing a PolyLine to a polygon: Avoid the soul.

Since the soul of PP has non-empty interior, there exists a point xPx\in P which is disjoint from all these pencil-arcs. Applying duality, this exactly says that there is some line in 𝑷\mbox{\boldmath{$P$}}^{*} which is disjoint from all the edges of our enhanced Dk+1(P)D_{k+1}(P). Hence, this enhancement makes Dk+1(P)D_{k+1}(P) planar. Our choice also varies continuously on Bn,kB_{n,k}. \spadesuit

Lemma 6.6

Dk+1D_{k+1} maps a member of Bk,nB_{k,n} to an nn-gon which is kk-nice.

Proof: Let Q=Dk+1(P)Q=D_{k+1}(P). A (k+1)(k+1)-diagonal of QQ is just a vertex of PP. A kk diagonal of QQ is a vertex of Δk(p)\Delta_{k}(p). Thus, to check the kk-nice property for QQ we need to take nn-collections of 44-tuples of points and check that they are distinct. In each case, the points are collinear because the lines of QQ are coincident.

[Uncaptioned image]

Figure 6.5 One of the nn different 44-tuples we need to check.

Once we make this specification, there is really combinatorially only possibility for which collections we need to check. Figure 6.5 shows one such 44-tuple, a,b,c,da,b,c,d. The shaded triangles are the two feathers of PP whose tips are b,cb,c. But a,b,c,da,b,c,d are distinct vertices of PΔk(P)P\cup\Delta_{k}(P) and so they are distinct. That is all there is to it. \spadesuit

To show that Q=Dk+1(P)Q=D_{k+1}(P) is a kk-bird, we consider a continuous path P(t)P(t) from the regular nn-gon P(0)P(0) to P=P(1)P=P(1). We set Q(t)=P(t)Q(t)=P(t). By construction, Q(0)Q(0) is a copy of the regular nn-gon in 𝑷\mbox{\boldmath{$P$}}^{*}, and Q(t)Q(t) is kk-nice for all t[0,1]t\in[0,1], and Q(t)Q(t) is a planar polygon for all t[0,1]t\in[0,1]. By definition Q=Q(1)Q=Q(1) is a kk-bird. This completes the proof of Theorem 6.4.

7 The Triangulation

7.1 Basic Definition

In this section we gather together the results we have proved so far and explain how we construct the triangulation τP\tau_{P} associated to a bird PBk,nP\in B_{k,n}.

Since Δk(Bk,n)Bk,n\Delta_{k}(B_{k,n})\subset B_{k,n}, we know that Δk(P)\Delta_{k}(P) is also a kk-bird. Combining this with Theorem 3.1 and Theorem 4.1 we can say that Δk(P)\Delta_{k}(P) is one embedded nn-gon contained in PIP^{I}, the interior of the region bounded by the embedded PP. The region between PP and Δk(P)\Delta_{k}(P) is a topological annulus. Moreover, Δk(P)\Delta_{k}(P) is obtained from PP by connecting the tips of the feathers of PP. The left side Figure 7.1 shows how this region is triangulated. The black triangles are the feathers of PP and each of the white triangles is made from an edge of Δk(P)\Delta_{k}(P) and two edges of adjacent feathers.

[Uncaptioned image]

Figure 7.1: The triangulation of the annulus

Lemma 7.1

For every member PBk,nP\in B_{k,n}, the associated 2n2n triangles have pairwise disjoint interiors, and thus triangulate the annular region between PP and Δk(P)\Delta_{k}(P).

Proof: As usual, we make a homotopical argument. If this result is false for some PP, then we can look at path which starts at the regular nn-gon (for which it is true) and stop at the first place where it fails. Theorem 4.1 tells us that nothing goes wrong with the feathers of PP. The only thing that can go wrong is Δk(P)\Delta_{k}(P) fails to be an embedded polygon. Since this does not happen, we see that in fact there is no counter-example at all. \spadesuit

We can now iterate, and produce 2n2n triangles between Δk(P)\Delta_{k}(P) and Δk2(P)\Delta_{k}^{2}(P), etc. The right side of Figure 7.1 shows the result of doing this many times. The fact that Δk(Bk,n)=Bk,n\Delta_{k}(B_{k,n})=B_{k,n} allows us to extend outward as well. When we iterate forever in both directions, we get an infinite triangulation of a (topological) cylinder that has degree 66 everywhere. This is what Figure 1.6 is showing. We call this bi-infinite triangulation τP\tau_{P}.

7.2 Some Structural Results

The following result will help with the proof of Theorem 1.3.

Theorem 7.2

Let PBn,kP\in B_{n,k}. Let SS be the soul of BB. Then for n\ell\geq n we have Δk(P)S\Delta^{\ell}_{k}(P)\subset S.

Proof: We first note the existence of certain infinite polygonal arcs in τP\tau_{P}. We start at a vertex of PP and then move inward to a vertex of Δk(P)\Delta_{k}(P) along one of the edges. We then continue through this vertex so that 33 triangles are on our left and 33 on our right. Figure 7.2 below shows the two paths like this that emanate from the same vertex of PP.

[Uncaptioned image]

Figure 7.2: The spiral paths.

The usual homotopical argument establishes the fact that the spiral paths are locally convex. One can understand their combinatrics, and how they relate to the polygons in the orbit, just by looking at the case of the regular nn-gon. We call the two spiral paths in Figure 7.2 partners. In the regular nn-gon the partners intersect infinitely often. So this is true in general. Each spiral path has an initial segment joining the initial endpoint on PP to the first intersection point with the partner. We define a petal to be the region bounded by the initial paths of the two partners.

It is convenient to write P=Δk(P)P^{\ell}=\Delta_{k}^{\ell}(P). In the regular case, PP^{\ell} is contained in the petal for >n1\ell>n-1.. Hence, the same goes in the general case. Because the initial segments are locally convex, the petal lies to the left of the lines extending the edges e1e_{1} and e2e_{2} when these edges are oriented according to the (k+1)(k+1)-diagonals of PP. But this argument works for every pair of partner spiral paths which start at a vertex of PP. We conclude that for n\ell\geq n, the polygon PP^{\ell} lies to the left of all the (k+1)(k+1)-diagonals of PP. But the soul of PP is exactly the intersection of all these left half planes. \spadesuit

Theorem 7.2 in turn gives us information about the nesting properties of birds within an orbit. Let SS_{\ell} denote the soul of PP^{\ell}. Let

S=𝒁S,S=𝒁S.S_{\infty}=\bigcap_{\ell\in\mbox{\boldmath{$Z$}}}S_{\ell},\hskip 30.0ptS_{-\infty}=\bigcup_{\ell\in\mbox{\boldmath{$Z$}}}S_{\ell}. (26)

It follows from Theorem 7.2 that P^=S\widehat{P}_{\infty}=S_{\infty} and P^=S\widehat{P}_{-\infty}=S_{-\infty}, because

S+nP+nSP.S_{\ell+n}\subset P^{\ell+n}\subset S_{\ell}\subset P^{\ell}. (27)

Hence these sets are all convex subsets of an affine plane.

Corollary 7.3

Any PBk,nP\in B_{k,n} is strictly star-shaped with respect to all points in the convex hull of Δkn(P)\Delta_{k}^{n}(P).

Proof: Since P+nSP^{\ell+n}\subset S_{\ell}, and PP^{\ell} is strictly star shaped with respect to all points of SS^{\ell}, we see that PP^{\ell} is strictly star shaped with respect to all points of P+nP^{\ell+n\/}. Since SS_{\ell} is convex, we can say more strongly that PP^{\ell} is strictly star-shaped with respect to all points of the convex hull of P+nP^{\ell+n}. Now we just set =0\ell=0 and recall the meaning of our notation, we get the exact statement of the result. \spadesuit

An immediate corollary is that PP is strictly star-shaped with respect to P^\widehat{P}_{\infty}. (Theorem 1.3 says that this is a single point.)

8 Nesting Properties of Birds

8.1 Duality

In this chapter we prove Theorem 1.3. In this first section we show how Statement 1 of Theorem 1.3 implies Statement 2. We want to prove that the “backwards union” P^\widehat{P}_{-\infty} is an affine plane. Here PBn,kP\in B_{n,k} is a kk-bird.

We take 0\ell\geq 0 and consider P=Δk(P)P^{-\ell}=\Delta_{k}^{-\ell}(P). Since PP^{-\ell} is planar, there is a closed set Λ\Lambda_{\ell} of lines in 𝑷P which miss PP^{-\ell}. These sets of lines are nested: Λ1Λ2Λ3\Lambda_{1}\supset\Lambda_{2}\supset\Lambda_{3}.... The intersection is non-empty and contains some line LL. We can normalize so that LL is the line at infinity. Thus all PP^{-\ell} lie in 𝑹2\mbox{\boldmath{$R$}}^{2}. We want to see that P^=𝑹2\widehat{P}_{-\infty}=\mbox{\boldmath{$R$}}^{2}.

Let Dk+1D_{k+1} be the map from §2.2 and §6.2. From Equation 8 we see that Dk+1D_{k+1} conjugates Δk\Delta_{k} to Δk1\Delta_{k}^{-1}. With Theorem 6.4 in mind, define the following “dual” kk-birds:

Π=Δk(Dk+1(P))=Dk+1(P).\Pi^{\ell}=\Delta_{k}^{\ell}(D_{k+1}(P))=D_{k+1}(P^{-\ell}). (28)

From Statement 1 of Theorem 1.3, the sequence of kk-birds {Π}\{\Pi^{\ell}\} shrinks to a point in the dual plane 𝑷\mbox{\boldmath{$P$}}^{*}. The vertices of Π\Pi^{\ell} are the (k+1)(k+1)-diagonals of PP^{-\ell}. Because the vertices of Π\Pi^{\ell} shrink to a single point, all the (k+1)(k+1)-diagonals of PP^{-\ell} converge to a single line LL^{\prime}.

Lemma 8.1

LL^{\prime} is the line at infinity.

Proof: Suppose not. When \ell is large, all the (k+1)(k+1)-diagonals point nearly in the same direction as LL^{\prime}. In particular, this is true of the subset of these diagonals which define the soul SS^{-\ell}. But these special diagonals turn monotonically and by less than π\pi radians as we move from one to the next. Hence, some of these diagonals nearly point in one direction along LL^{\prime} and some point nearly in the opposite direction. But then SS^{-\ell} converges to a subset of LL^{\prime}. This is a contradiction, \spadesuit

The soul SS^{-\ell} is a convex set, containing the origin, and is bounded by some of the (k+1)(k+1) diagonals. If SS^{-\ell} does not converge to the whole plane, then some (k+1)(k+1)-diagonal intersects a uniformly bounded region in 𝑹2\mbox{\boldmath{$R$}}^{2} for each \ell. But this produces a sequence of (k+1)(k+1)-diagonals that does not converge to the line at infinity. This is a contradiction. Hence SS^{-\ell} converges to all of 𝑹2\mbox{\boldmath{$R$}}^{2}. But then so does PP^{-\ell}.

8.2 The Pre-Compact Case

The rest of the chapter is devoted to proving Statrement 1 of Theorem 1.3. Let PBn,kP\in B_{n,k} and let P=Δ(P)P^{\ell}=\Delta^{\ell}(P). We take =0,1,2,3\ell=0,1,2,3....

Conjecture 8.2

The sequence {P}\{P^{\ell}\} is pre-compact modulo affine transformations. That is, this sequence has a convergent subsequence which converges to another element of Bn,kB_{n,k}.

In this section I will prove the P^\widehat{P}_{\infty} is a single point under the assumption that {P}\{P^{\ell}\} is pre-compact.

We would like to see that the diameter of PP^{\ell} steadily shrinks, but the notion of diameter is not affinely natural. We first develop a notion of affinely natural diameter. For each direction vv in the plane, we let Sv\|S\|_{v} denote the maximum length of LSL\cap S where LL is a straight line parallel to vv. We then define

δ(S1,S2)=supvS1vS2v[0,1].\delta(S_{1},S_{2})=\sup_{v}\frac{\|S_{1}\|_{v}}{\|S_{2}\|_{v}}\in[0,1]. (29)

The quantity δ(S1,S2)\delta(S_{1},S_{2}) is affine invariant, and (choosing a direction μ\mu which realizes the diamater of S1S_{1}) we have

diam(S1)diam(S2)S1μS2μδ(S1,S2).\frac{{\rm diam\/}(S_{1})}{{\rm diam\/}(S_{2})}\leq\frac{\|S_{1}\|_{\mu}}{\|S_{2}\|_{\mu}}\leq\delta(S_{1},S_{2}). (30)

Let SS^{\ell} be the soul of PP^{\ell}. By Theorem 5.11 we have S+nSS^{\ell+n}\subset S^{\ell}. More precisely, the former set is contained in the interior of the latter set. Under the pre-compactness assumption, there are infinitely many indices j\ell_{j} and some ϵ>0\epsilon>0 such that

δ(Sj+n,Sj)<1ϵ.\delta(S^{\ell_{j}+n},S^{\ell_{j}})<1-\epsilon. (31)

But then

diam(Sj+n)diam(Sj)<1ϵ\frac{{\rm diam\/}(S^{\ell_{j}+n})}{{\rm diam\/}(S^{\ell_{j}})}<1-\epsilon (32)

infinitely often. This forces diam(S)0{\rm diam\/}(S^{\ell})\to 0. But P^\widehat{P}_{\infty} is contained in this nested intersection and hence is a point.

If we knew the truth of Conjecture 8.2 then our proof of Theorem 1.3 would be done. Since we don’t know this, we have to work much harder to prove Statement 1 in general.

8.3 Normalizing by Affine Transformations

Henceforth we assume that the forward orbit {P}\{P^{\ell}\} of PP under Δk\Delta_{k} is not pre-compact modulo affine transformations.

Lemma 8.3

There is a sequence {T}\{T_{\ell}\} of affine transformations such that

  1. 1.

    T(P)T_{\ell}(P^{\ell}) has (the same) 33 vertices which make a fixed equilateral triangle.

  2. 2.

    TT_{\ell} expands distances on PP^{\ell} for all \ell.

  3. 3.

    T(P)T_{\ell}(P^{\ell}) is contained in a uniformly bounded subset of 𝑹2\mbox{\boldmath{$R$}}^{2}.

Proof: To PP^{\ell} we associate the triangle τ\tau_{\ell} made from 33 vertices of PP^{\ell} and having maximal area. The diameter of τ\tau_{\ell} is uniformly small, so we can find a single equilateral triangle TT and an expanding affine map T:τTT_{\ell}:\tau_{\ell}\to T. Let dd be the side length of TT. Every vertex of T(P)T_{\ell}(P^{\ell}) is within dd of all the sides of TT, because otherwise we’d have a triangle of larger area. The sequence {T}\{T_{\ell}\} has the advertised properties. \spadesuit

Let Q=T(P)Q^{\ell}=T_{\ell}(P^{\ell}). By compactness we can pass to a subsequence so that the limit polygon QQ exists, in the sense that the vertices and the edges converge. Let Q0,Q1Q_{0},Q_{1}, etc. be the vertices of QQ. Perhaps some of these coincide. Each distinguished diagonal of QQ^{\ell} defines the unit vector which is parallel to it. Thus QQ^{\ell} defines a certain list of 2n2n unit vectors. We can pass to a subsequence so that all these unit vectors converge. Thus QQ still has well defined distinguished diagonals even when the relevant points coincide.

We now define the “limiting soul”. Let S=S(Q)S^{\ell}=S(Q^{\ell}), the soul of QQ^{\ell}. As in §5.7. let SS be the set of accumulation points of sequences {p}\{p^{\ell}\} with pSp^{\ell}\in S^{\ell}. Since SQS^{\ell}\subset Q^{\ell} for all \ell we have SQS\subset Q. Now we define a related object. We have a left half-plane associated to each diagonal of QQ. We define Σ\Sigma to be the intersection of all these half-planes. We will use the set Σ\Sigma at various places below to get control over the set SS.

Lemma 8.4

SΣS\subset\Sigma.

Proof: Fix ϵ>0\epsilon>0. If this is not the case, then by compactness we can find a convergent sequence {p}\{p^{\ell}\}, with pSp^{\ell\/}\in S^{\ell}, which does not converge to a point of Σ\Sigma. But pp^{\ell} lies in every left half plane associated to QQ^{\ell}. But then, by continuity, the accumulation point pp lies in every left half plane associated to QQ. Hence pΣp\in\Sigma. \spadesuit

8.4 Structure of the Normalized Limits

We work under the assumption that P^\widehat{P}_{\infty} is not a single point. The goal of this section is to establish several structural properties about the sets SS and QQ. Our first property guarantees that there is a chord SS^{*} of SS connecting vertices of QQ. Once we establish this, we show that QQ is a union of two “monotone” arcs joining the endpoints of SS^{*}. These structural properties will be used repeatedly in subsequent sections of this chapter.

Let HQH_{Q} denote the convex hull of QQ. Note that SQHQS\subset Q\subset H_{Q}.

Corollary 8.5

Suppose that P^\widehat{P}_{\infty} is not a single point. Then δ(S,HQ)=1\delta(S,H_{Q})=1.

Proof: Suppose not. Note that HQSnH_{Q^{\ell}}\subset S^{\ell-n} by Theorem 7.2 and convexity. Then for \ell large we have

δ(Qn)=δ(S,Sn)δ(S,HQ)<δ(S,HQ)+ϵ,\delta(Q^{\ell-n})=\delta(S^{\ell},S^{\ell-n})\leq\delta(S^{\ell},H_{Q^{\ell}})<\delta(S,H_{Q})+\epsilon,

and we can make ϵ\epsilon as small as we like. This gives us a uniform δ<1\delta<1 such that δ(Q)<δ\delta(Q^{\ell})<\delta once \ell is large enough. The argument in the compact case now shows that P^\widehat{P}_{\infty} is a single point. \spadesuit

Corollary 8.5 says that SS and QQ have the same diameter. Hence there is a chord SSS^{*}\subset S which has the same diameter as QQ. Since QQ is a polygon, this means that QQ must have vertices at either endpoint of SS^{*}. We normalize so that SS^{*} is the unit segment joining (0,0)(0,0) to (1,0)(1,0).

Lemma 8.6

Let QQQ^{\prime}\subset Q be an arc of QQ that joins (0,0)(0,0) to (1,0)(1,0).

  1. 1.

    The vertices of QQ^{\prime} must have non-decreasing xx-coordinates.

  2. 2.

    If consecutive vertices of QQ^{\prime} have the same xx-coordinate, they coincide.

  3. 3.

    Either QSQ^{\prime}\subset S^{*} or QQ^{\prime} intersects SS^{*} only at (0,0)(0,0) and (1,0)(1,0).

Proof: Suppose the Statement 1 is false. Then we can find a vertical line Λ\Lambda which intersects SS^{*} at a relative interior point and which intersects QQ^{\prime} transversely at 33 points. But then once \ell is sufficiently large, QQ^{\ell} will intersect all vertical lines sufficiently close to Λ\Lambda in at least 33 points and moreover some of these lines will contain points of SS^{\ell}. This contradicts the fact that QQ^{\ell} is strictly star-shaped with respect to all points of QQ^{\ell}.

For Statement 2, we observe that QQ^{\prime} does not contain any point of the form (0,y)(0,y) or (1,y)(1,y) for y0y\not=0. Otherwise QQ has larger diameter than 11. This is to say that once QQ^{\prime} leaves (0,0)(0,0) it immediately moves forward in the XX-direction. Likewise, once QQ^{\prime} (traced out the other way) leaves (1,0)(1,0) it immediately moves backward in the XX-direction. If Statement 2 is false, ten we can find a non-horizontal line Λ\Lambda^{\prime} which intersects SS^{*} in a relative interior point and which intersects QQ^{\prime} transversely at 33 points. The slope is Λ\Lambda^{\prime} depends on which of the two vertices of QQ^{\prime} lies above the other. Once we have Λ\Lambda^{\prime} we play the same game as for the first statement, and get the same kind of contradiction.

Suppose Statement 3 is false. We use the kind of argument we had in §5.8. By Statements 1 and 2 together, QQ^{\prime} must have an escape edge which touches SS^{*} in a relative interior point. Moreover, this one escape edge is paired with another escape edge. Thus we can find a point xSx\in S^{*} which strictly lies on the same side of both of these same-type escape edges. The argument in §5.8 now shows that QQ^{\ell} is not strictly star-shaped with respect to points of SS^{\ell} very near xx. \spadesuit

Corollary 8.7

Suppose 0a<b<n0\leq a<b<n and Qa=QbQ_{a}=Q_{b}. Then either we have Qa=Qa+1==QbQ_{a}=Q_{a+1}=...=Q_{b} or else we have Qb=Qb+1==Qa+nQ_{b}=Q_{b+1}=...=Q_{a+n}.

Proof: In view of Lemma 8.6 it suffices to show that our two monotone arcs comprising QQ are disjoint except at their endpoints.

Let UU denote the open upper halfplane, bounded by the XX-axis. After reflecting in the XX-axis we can guarantee that one of our monotone arcs α\alpha has a point in UU. But then, by Lemma 8.6, all of α\alpha lies in UU except for its endpoints. If the other monotone arc β\beta intersects α\alpha away from the endpoints, then β\beta has a point in UU, but then, by Lemma 8.6, all of β\beta lies in UU except for the endpoints. But then SS lies in UU, except for the points (0,0)(0,0) and (1,0)(1,0). This contradicts the fact that SSS^{*}\subset S. \spadesuit

Our argument shows in particular that QQ is embedded, up to adding repeated vertices. However, we will not directly use this property in our proof below.

8.5 The Triangular Case

We continue with the assumption that P^\widehat{P} is not a single point. Here we pick off a special case:

  • There is a line LL such that Q0LQ_{0}\not\in L.

  • Qk,Qk+1,,Qnk1,QnkLQ_{k},Q_{k+1},...,Q_{n-k-1},Q_{n-k}\in L and

  • QkQnkQ_{k}\not=Q_{n-k}.

Figure 8.1 shows the situation. As always, the notation QkQ_{-k} and QnkQ_{n-k} names the same point. All but 2k12k-1 points are on LL, and except for Q0Q_{0} we don’t know where these other 2k12k-1 points are.

[Uncaptioned image]

Figure 8.1: The triangular limit QQ.

Given the constant energy of our orbit, the cross ratio of the lines

Q0,k,Q0,k+1,Qnk1,0,Qnk,0Q_{0,k},\ Q_{0,k+1},\ Q_{n-k-1,0},\ Q_{n-k,0}

is at least ϵ0\epsilon_{0}. Also, these lines are cyclically ordered about 0 as indicated in Figure 8.1, thanks to the kk-niceness property and continuity. Also, the two lines containing Q0,kQ_{0,k} and Qk,0Q_{-k,0} are not parallel because Q0LQ_{0}\not\in L. Hence SS is contained in the shaded region in Figure 8.1, namely the triangle with vertices Q0Q_{0} and Q±(k+1)Q_{\pm(k+1)}. But this shaded region has diameter strictly smaller than the triangle τ\tau with vertices Q0Q_{0} and Q±kQ_{\pm k}. Hence diam(S)<diam(τ)diam(Q){\rm diam\/}(S)<{\rm diam\/}(\tau)\leq{\rm diam\/}(Q). This contradicts Corollary 8.5 which says, in particular, that SS and QQ have the same diameter.

8.6 The Case of No Folded Diagonals

We work under the assumption that P^\widehat{P}_{\infty} is not a single point. The notions of collapsed diagonals, folded diagonals, and aligned diagonals from §5 make sense for QQ because the concepts just involve the directions of the diagonals. The proof of Lemma 5.3 also works the same way.

Lemma 8.8

QQ must either have a trivial edge, a trivial distinguished diagonal, or collapsed diagonals,

Proof: As remarked in §5, the proof of the Degeneration Lemma works for sequences as well as paths, and only uses the fact that the limiting polygon has nontrivial edges and nontrivial distinguished diagonals. So, if QQ has no trivial edges and no trivial distinguished diagonals, then all but one vertex of QQ lies in a single line. But then QQ has collapsed diagonals. \spadesuit

Remark: Here is a second, more direct proof. If Lemma 8.8 is false then we have a picture as in the left side of Figure 7.1. The feathers defined in §4.1 would be all non-degenerate and the segments joining the tips of consecutive feathers would be nontrivial. This would force SS to lie in the interior of QQ. But then diam(S)<diam(Q){\rm diam\/}(S)<{\rm diam\/}(Q), contradicting Corollary 8.5.

If QQ has a trivial distinguished diagonal, then by Lemma 8.7, we see that QQ also has a trivial edge. If QQ has a trivial edge, say Q1=Q0Q_{-1}=Q_{0}, then the diagonals at QQ are collapsed at QkQ_{k}. So, in all cases, QQ has collapsed diagonals. We assume in this section that QQ has no folded diagonals anywhere. This means that QQ has aligned diagonals, say at QkQ_{k}. Thus Q0,kQ_{0,k} and Qk,2kQ_{k,2k} are parallel. Since QQ does not lie in a line, Lemma 5.3 tells us that the chain of 2k+12k+1 parallel distinguished diagonals:

Q0,k,Q0,k+1,Q1,k+1,Q1,k+2,,Qk1,2k,Qk,2kQ_{0,k},Q_{0,k+1},Q_{1,k+1},Q_{1,k+2},...,Q_{k-1,2k},Q_{k,2k} (33)

Now we have a “runaway situation”. The two diagonals Q2k,kQ_{2k,k} and Q2k,k1Q_{2k,k-1} (which are just the reversals of the last two in Equation 33) are parallel. Thus QQ has collapsed diagonals at Q2kQ_{2k}. Since QQ has no folded diagonals, QQ has aligned diagonals at Q2kQ_{2k}. But then, applying Lemma 5.3 again, we can extend that chain in Equation 33 so that it contines as ,,Q2k1,3k,Q2k,3k,...,Q_{2k-1,3k},Q_{2k,3k}. But now QQ has collapsed diagonals at Q3kQ_{3k}. And so on. Continuing this way, we end up with all points on QQ. This is a contradiction.

The only way out is that QQ must have folded diagonals somewhere

8.7 The Case of Folded Diagonals

We continue to work under the assumption that P^\widehat{P}_{\infty} is not a single point. Now we consider the case when QQ has folded diagonals at, say, Q0Q_{0}. What this means that the diagonals Q0,k+1Q_{0,k+1}, Q0,k1Q_{0,-k-1} are parallel. (Again, these diagonals are well defined even when their endpoints coincide; we are just using a notational convention to name them here.) But then the corresponding half planes intersect along a single line LL, forcing ΣL\Sigma\subset L. By Lemma 8.4, the soul SS is contained in Σ\Sigma. Hence, SLS\subset L. Letting SS^{*} be the chord from §8.4, we also have S=SS=S^{*}. This is because SS and SS^{*} are segments of the same diagonal and in the same line. We will use SS and SS^{*} interchangeably below.

We normalize so that SS is the line segment connecting (0,0)(0,0) to (1,0)(1,0). As in §8.4, both these points are vertices of QQ. The folding condition forces Σ\Sigma (and hence SS) to lie to one side of these points. Hence, we have either Q0=(0,0)Q_{0}=(0,0) or Q0=(1,0)Q_{0}=(1,0). Without loss of generality we consider the case when Q0=(0,0)Q_{0}=(0,0). Note that points of QSQ-S do not belong to LL, because QQ and SS have the same diameter. We break the analysis down into cases.

Case 1: Suppose that Qk+1Q_{k+1} is not an endpoint of SS^{*} and Qnk1(0,0)Q_{n-k-1}\not=(0,0). Consider the arc QQ^{\prime} given by Q0Qk+1Qβ=(1,0)Q_{0}\to...\to Q_{k+1}\to...\to Q_{\beta}=(1,0). Here β\beta is some index we do not know explicitly, but we take β\beta as large as possible, in the sense that Qβ+1(1,0)Q_{\beta+1}\not=(1,0). The arc QQ^{\prime} connects (0,0)(0,0) to (1,0)(1,0) and intersets SS^{*} at Qk+1Q_{k+1}, a point which is neither (0,0)(0,0) or (1,0)(1,0). By Lemma 8.6, we have QSQ^{\prime}\subset S^{*}. We conclude that Q0,,QβSQ_{0},...,Q_{\beta}\subset S^{*}.

If β\beta does not lie in the index interval (k+1,nk1)(k+1,n-k-1) then we have just shown that Qk+1,,Qnk1SQ_{k+1},...,Q_{n-k-1}\in S^{*}. If β=nk1\beta=n-k-1 we have the same result. Here is what we do if β\beta does lie in (k+1,nk1)(k+1,n-k-1). We apply our same argument as in the previous paragraph to the arc QβQnk1Q_{\beta}\to...\to Q_{n-k-1}, and see that Qβ,,Qnk1SQ_{\beta},...,Q_{n-k-1}\in S. So, in all cases, we see that Qk+1,,Qnk1SQ_{k+1},...,Q_{n-k-1}\in S.

In short, QjLQ_{j}\in L unless j{k,,1}j\in\{-k,...,-1\}. All but kk vertices belong to LL. In particular, we have an index h{k,,1}h\in\{-k,...,-1\} such that QhLQ_{h}\not\in L but Qh+k,Qh+k+1,,Qh+nk1,Qh+nkLQ_{h+k},Q_{h+k+1},...,Q_{h+n-k-1},Q_{h+n-k}\in L. Now we are close to the Triangular case from §8.5 except that all the indices are shifted by hh. If it happens that Qh+kQh+nkQ_{h+k}\not=Q_{h+n-k} then we have the Triangular Case and we are done.

The other possibility is that Qh+k=Qh+nkQ_{h+k}=Q_{h+n-k}. In this case, Lemma 8.7 gives us Qh+k=Qh+k+1=Qh+nk1=Qh+nkQ_{h+k}=Q_{h+k+1}=Q_{h+n-k-1}=Q_{h+n-k}. In particular, the diagonals Qh,h+k+1Q_{h,h+k+1} and Qh,h+nk1Q_{h,h+n-k-1} are folded at QhQ_{h}. Since QhLQ_{h}\not\in L this means that there is some other line LL^{\prime} such that SLS\subset L^{\prime}. This is a contradiction.

Case 2: Suppose Qk1=Qk+1=(1,0)Q_{-k-1}=Q_{k+1}=(1,0). Before analyzing this case, we remember a lesson from the end of Case 1: It is not possible for QQ to have folded diagonals at a point not on SS.

Corollary 8.7 says that Qk+1==Qnk1=(1,0).Q_{k+1}=...=Q_{n-k-1}=(1,0). This is a run of k+βk+\beta points, where β=n(3k+1)0\beta=n-(3k+1)\geq 0. There is some index h{±1,±k}h\in\{\pm 1,...\pm k\} such that QhLQ_{h}\not\in L. Without loss of generality we will take h{1,,k}h\in\{1,...,k\}.

Suppose first that n>3k+1n>3k+1. Then there are at least k+1k+1 consecutive vertices sitting at (1,0)(1,0) and so both diagonals Qh,k+hQ_{h,k+h} and Qh,k+h+1Q_{h,k+h+1} point from QhQ_{h} to (1,0)Qh(1,0)\not=Q_{h}. This means that QQ has collapsed diagonals at QhQ_{h}. Remembering our lesson, we know that QQ does not have folded diagonals at QhQ_{h}. Hence QQ has aligned diagonals at QhQ_{h}.

Now we have the same runaway situation we had in §8.6. The diagonals in the chain Qhk,hQh,h+kQ_{h-k,h}...Q_{h,h+k} point are all pointing along the line connecting (1,0)(1,0) to QhQ_{h}, and they are pointing away from (1,0)(1,0). This gives us collapsed diagonals at Qh+kQ_{h+k}. Remembering our lesson, we see that QQ has aligned diagonals at Qh+kQ_{h+k}. And so on. All the points after QhQ_{h} get stuck on LL^{\prime} and we have a contradiction.

If n=3k+1n=3k+1, then the same argument works as long as h±kh\not=\pm k. So, we just have to worry about the case when all points of QQ belong to SS except for QkQ_{k} and QkQ_{-k}, which do not belong to SS. Applying Lemma 8.6 to the arc Q0Q1Qk(1,0)Q_{0}\to Q_{1}\to...\to Q_{k}\to(1,0) we conclude that Q0==Qk1=(0,0)Q_{0}=...=Q_{k-1}=(0,0). Applying Lemma 8.6 to the arc Q0Q1Qk(1,0)Q_{0}\to Q_{-1}\to...\to Q_{-k}\to(1,0) we conclude that Q0==Qk1=(0,0)Q_{0}=...=Q_{k-1}=(0,0). But now we have a run of 2k1k+12k-1\geq k+1 points sitting at (0,0)(0,0) and we can run the same argument as in the case n>3k+1n>3k+1, with (0,0)(0,0) in place of (1,0)(1,0).

Case 3: The only cases left to consider is when one or both of Q±(k+1)Q_{\pm(k+1)} equals (0,0)(0,0). We suppose without loss of generality that Qk1=(0,0)Q_{-k-1}=(0,0). Since we also have Q0=(0,0)Q_{0}=(0,0), Lemma 8.7 gives Qk1==Q0=(0,0)Q_{-k-1}=...=Q_{0}=(0,0). This is a run of k+2k+2 consecutive points sitting at (0,0)(0,0).

There is some smallest h>0h>0 so that QhSQ_{h}\not\in S. Applying Lemma 8.6 to the arc Q0Qk(1,0)Q_{0}\to...\to Q_{k}\to...\to(1,0), we conclude that Qh1==Q1=(0,0)Q_{h-1}=...=Q_{1}=(0,0). (Otherwise Lemma 8.6 would force QhSQ_{h}\in S.)

Now we know that QQ has collapsed diagonals at QhLQ_{h}\not\in L. We now get a contradiction from the same runaway situation as in Case 2.

9 Appendix

9.1 The Energy Invariance Revisited

In this section we sketch Anton Izosimov’s proof that χkΔk=χk\chi_{k}\circ\Delta_{k}=\chi_{k}. This proof requires the machinery from [6]. (The perspective comes from [8], but the needed result for Δk\Delta_{k} is in the follow-up paper [6].)

Let PP be an nn-gon. We let V1,,VnV_{1},...,V_{n} be points in 𝑹3\mbox{\boldmath{$R$}}^{3} representing the consecutive vertices of PP. Thus the vertex PjP_{j} is the equivalence class of VjV_{j}. We can choose periodic sequences {ai}\{a_{i}\}, {bi}\{b_{i}\}, {ci}\{c_{i}\}, {di}\{d_{i}\} such that

aiVi+biVi+k+ciVi+k+1+diVi+2k+1=0,i.a_{i}V_{i}+b_{i}V_{i+k}+c_{i}V_{i+k+1}+d_{i}V_{i+2k+1}=0,\hskip 30.0pt\forall i. (34)

Recall from §2.2 that Δk=DkDk+1\Delta_{k}=D_{k}\circ D_{k+1}.

Lemma 9.1

One of the cross ratio factors of χkDk+1\chi_{k}\circ D_{k+1} is (a0dk)/(c0bk)(a_{0}d_{-k})/(c_{0}b_{-k}).

Proof: One of the factors is the cross ratio of P0,y,x,Pk+1P_{0},y,x,P_{k+1}, where

x=P0,k+1Pk,2k+1,y=Pk,1P0,k+1.x=P_{0,k+1}\cap P_{k,2k+1},\hskip 30.0pty=P_{-k,1}\cap P_{0,k+1}.

(Compare the right side of Figure 2.1, shifting all the indices there by k+1k+1.)

The points xx and yy respectively are represented by vectors

X=a0V0+c0Vk+1=b0Vkd0V2k+1,X=a_{0}V_{0}+c_{0}V_{k+1}=-b_{0}V_{k}-d_{0}V_{2k+1},
Y=akVkckV1=bkV0+dkVk+1.Y=-a_{-k}V_{-k}-c_{-k}V_{1}=b_{-k}V_{0}+d_{-k}V_{k+1}.

The point here is that the vector XX lies in the span of {V0,Vk+1}\{V_{0},V_{k+1}\} and in the span of {Vk,V2k+1}\{V_{k},V_{2k+1}\} and projectively this is exactly what is required. A similar remark applies to YY.

Setting Ω=V0×Vk+1\Omega=V_{0}\times V_{k+1}, we compute the relevant cross ratio as

V0×YV0×XX×Vk+1Y×Vk+1=dkΩc0Ω×a0ΩbkΩ=dka0bkc0,\frac{V_{0}\times Y}{V_{0}\times X}\cdot\frac{X\times V_{k+1}}{Y\times V_{k+1}}=\frac{d_{-k}\Omega}{c_{0}\Omega}\times\frac{a_{0}\Omega}{b_{-k}\Omega}=\frac{d_{-k}a_{0}}{b_{-k}c_{0}}, (35)

which is just a rearrangement of the claimed term. \spadesuit

The other cross ratio factors are obtained by shifting the indices in an obvious way. As an immediate corollary, we see that

χk(Dk+1(P))=i=1naidibici.\chi_{k}(D_{k+1}(P))=\prod_{i=1}^{n}\frac{a_{i}d_{i}}{b_{i}c_{i}}. (36)

Let us call this quantity μk(P)\mu_{k}(P).

Lemma 9.2

If μkΔk=μk\mu_{k}\circ\Delta_{k}=\mu_{k} then χkΔk=χk\chi_{k}\circ\Delta_{k}=\chi_{k}.

Proof: If μkΔk=μk\mu_{k}\circ\Delta_{k}=\mu_{k} then μkΔk1=μk\mu_{k}\circ\Delta_{k}^{-1}=\mu_{k}. Equation 36 says that

χkDk+1=μk,μkDk+1=χk.\chi_{k}\circ D_{k+1}=\mu_{k},\hskip 30.0pt\mu_{k}\circ D_{k+1}=\chi_{k}. (37)

The first equation implies the second because Dk+1D_{k+1} is an involution. Since Dk+1D_{k+1} conjugates Δk\Delta_{k} to Δk1\Delta_{k}^{-1} we have

χkΔk=χkDk+1Δk1Dk+1=μkΔk1Dk+1=μkDk+1=χk.\chi_{k}\circ\Delta_{k}=\chi_{k}\circ D_{k+1}\circ\Delta_{k}^{-1}\circ D_{k+1}=\mu_{k}\circ\Delta_{k}^{-1}\circ D_{k+1}=\mu_{k}\circ D_{k+1}=\chi_{k}.

This completes the proof. \spadesuit

Let P~=Δk(P)\widetilde{P}=\Delta_{k}(P). Let {a~i}\{\widetilde{a}_{i}\}, etc., be the sequences associated to P~\widetilde{P}. We want to show that

i=1naidibici=i=1na~id~ib~ic~i.\prod_{i=1}^{n}\frac{a_{i}d_{i}}{b_{i}c_{i}}=\prod_{i=1}^{n}\frac{\widetilde{a}_{i}\widetilde{d}_{i}}{\widetilde{b}_{i}\widetilde{c}_{i}}. (38)

This is just a restatement of the equation μkΔk=μk\mu_{k}\circ\Delta_{k}=\mu_{k}.

Now we use the formalism from [6] to establish Equation 38. We associate to our polygon PP operator DD on the space 𝒱\cal V of bi-infinite sequences {Vi}\{V_{i}\} of vectors in 𝑹3\mbox{\boldmath{$R$}}^{3}. The definition of DD is given coordinate-wise as

D(Vi)=aiVi+biTk(Vi)+ciTk+1(Vi)+diT2k+1(Vi).D(V_{i})=a_{i}V_{i}+b_{i}T^{k}(V_{i})+c_{i}T^{k+1}(V_{i})+d_{i}T^{2k+1}(V_{i}). (39)

Here TT is the shift operator, whose action is T(Vi)=Vi+1T(V_{i})=V_{i+1}. If we take {Vi}\{V_{i}\} to be a periodic bi-infinite sequence of vectors corresponding to our polygon PP, then DD maps {Vi}\{V_{i}\} to the 0-sequence.

Next, we write D=D++DD=D_{+}+D_{-} where coordinate-wise

D+(Vi)=aiVi+ciTk+1(Vi),D(Vi)=biTk(Vi)+diT2k+1(Vi).D_{+}(V_{i})=a_{i}V_{i}+c_{i}T^{k+1}(V_{i}),\hskip 30.0ptD_{-}(V_{i})=b_{i}T^{k}(V_{i})+d_{i}T^{2k+1}(V_{i}). (40)

The pair (D+,D)(D_{+},D_{-}) is associated to the polygon PP.

Let D~\widetilde{D} and (D~+,D~)(\widetilde{D}_{+},\widetilde{D}_{-}) be the corresponding operators associated to P~\widetilde{P}. One of the main results of [6] is that the various choices can be made so that

D~+D=D~D+.\widetilde{D}_{+}D_{-}=\widetilde{D}_{-}D_{+}. (41)

This is called refactorization. Equating the lowest (respectively highest) terms of the relation in Equation 41 gives us the identity a~ibi=b~iai+k\widetilde{a}_{i}b_{i}=\widetilde{b}_{i}a_{i+k} (respectively c~idi+k+1=d~ici+2k+1\widetilde{c}_{i}d_{i+k+1}=\widetilde{d}_{i}c_{i+2k+1}.) These relations hold for all ii and together imply Equation 38.

9.2 Extensions of Glick’s Formula

Theorem 1.1 in [3] says that the coordinates for the collapse point of the pentagram map Δ1\Delta_{1} are algebraic functions of the coordinates of the initial polygon. In Equation 1.1 of [3], Glick goes further and gives a formula for the collapse point. I will explain his formula. Let (x,y)(x^{*},y^{*}) denote the accumulation point of the forward iterates of PP under Δ1\Delta_{1}. Let P^=(x,y,1)\widehat{P}_{\infty}=(x^{*},y^{*},1) be the collapse point. In somewhat different notation, Glick introduces the operator

TP=nI3GP,GP(v)=i=1n|Pi1,v,Pi+1||Pi1,Pi,Pi+1|Pi.T_{P}=nI_{3}-G_{P},\hskip 30.0ptG_{P}(v)=\sum_{i=1}^{n}\frac{|P_{i-1},v,P_{i+1}|}{|P_{i-1},P_{i},P_{i+1}|}P_{i}. (42)

Here |a,b,c||a,b,c| denotes the determinant of the matrix with rows a,b,ca,b,c and I3I_{3} is the 3×33\times 3 identity matrix. It turns out TPT_{P} is a Δ1\Delta_{1}-invariant operator, in the sense that TΔ0(P)=TPT_{\Delta_{0}(P)}=T_{P}. Moreover PP_{\infty} is an eigenvector of TPT_{P}. This is Glick’s formula for P^\widehat{P}_{\infty}. Actually, one can say more simply that GPG_{P} is a Δ0\Delta_{0}-invariant operator and that P^\widehat{P}_{\infty} is a fixed point of the projective action of GpG_{p}. This means that the vectors representing these points in 𝑹3\mbox{\boldmath{$R$}}^{3} are eigenvectors for the operator. The reason Glick uses the more complicated expression nI3GPnI_{3}-G_{P} is that geometrically it is easier to work with.

Define GP,a,bG_{P,a,b} by the formula

GP,a,b(v)=i=1n|Pia,v,Pi+b||Pia,Pi,Pi+b|Pi.G_{P,a,b}(v)=\sum_{i=1}^{n}\frac{|P_{i-a},v,P_{i+b}|}{|P_{i-a},P_{i},P_{i+b}|}P_{i}. (43)

Let P^,k\widehat{P}_{\infty,k} be the limit point of the forward iterates of PP under Δk\Delta_{k}.

A lot of experimental evidence suggests the following conjecture.

Conjecture 9.3

Let k2k\geq 2. If n=3k+1n=3k+1 the point P^\widehat{P}_{\infty} is a fixed point for the projective action of GP,k,kG_{P,k,k}. If n=3k+2n=3k+2 the point P^\widehat{P}_{\infty} is a fixed point for the projective action of GP,k+1,k+1G_{P,k+1,k+1}. In particular, in these cases the coordinates of P^\widehat{P}_{\infty} are algebraic functions of the vertices of PP.

Anton Izosimov kindly explained the following lemma, which seems like a big step in proving the conjecture. (I am still missing the geometric side of Glick’s argument in this new setting.)

Lemma 9.4

When n=3k+1n=3k+1 the operator GP,k,kG_{P,k,k} is invariant under Δk\Delta_{k}. When n=3k+2n=3k+2 the operator GP,k+1,k+1G_{P,k+1,k+1} is invariant under Δk\Delta_{k}.

Proof: These operators are Glick’s operator in disguise. When n=3k+1n=3k+1 we can relabel our nn-gons in a way that converts Δk\Delta_{k} to the pentagram map. The corresponding space of birds Bn,kB_{n,k} corresponds to some strange set of “relabeled kk-birds”. This relabeling converts GP,k,kG_{P,k,k} respectively to Glick’s original operator. This proves the invariance of GP,k,kG_{P,k,k} under Δk\Delta_{k} when n=3k+1n=3k+1. A similar thing works for n=3k+2n=3k+2, but this time the relabeling converts Δk\Delta_{k} to the inverse of the pentagram map. \spadesuit

I was not able to find any similar formulas when n>3k+2n>3k+2.

Question 9.5

When n>3k+2n>3k+2 and PP is a kk-bird, are the coordinates of the collapse point P^\widehat{P}_{\infty} algebraic functions of the vertices of PP?

Here is one more thing I have wondered about. Suppose that nn is very large and PP is a convex nn-gon. Then PP can be considered as a kk-bird for all k=1,2,,βk=1,2,...,\beta, where β\beta is the largest integer such that n3β+1n\geq 3\beta+1. When we apply the map Δk\Delta_{k} for these various values of kk we get potentially β\beta different collapse points. All I can say, based on experiments, is that these points are not generally collinear.

Question 9.6

Does the collection of β\beta collapse points in this situation have any special meaning?

9.3 Star Relabelings

Let us further take up the theme in the proof of Lemma 9.4. Given an nn-gon PP and and some integer rr relatively prime to nn, we define a new nn-gon PrP^{*r} by the formula

Pjr=Prj.P^{*r}_{j}=P_{rj}. (44)

Figure 1.5 shows the P(3)P^{*(-3)} when PP is the regular 1010-gon.

As we have already mentioned, the action of Δ1\Delta_{1} on the P(k)P^{*(-k)} is the same as the action of Δk\Delta_{k} on PP when n=3k+1n=3k+1. So, when n=3k+1n=3k+1, the pentagram map has another nice invariant set (apart from the set of convex nn-gons), namely

Bk,n(k)={P(k)|PBk,n}.B_{k,n}^{*(-k)}=\{P^{*(-k)}|\ P\in B_{k,n}\}.

The action of the pentagram map on this set is geometrically nice. If we suitably star-relabel, we get star-shaped (and hence embedded) polygons. A similar thing works when n=3k+2n=3k+2.

References

  • [1] Q. Aboud and A. Izosimov, The Limit Point of the Pentagram Map and Infinitesimal Monodromy, I.M.R.N., Vol 7, (2022)
  • [2] M. Gekhtman, M. Shapiro, S. Tabachnikov, A. Vainshtein, Integrable cluster dynamics of diected networks and pentagram maps, Adv. Math. 300 (2016), pp 390-450
  • [3] M. Glick, The Limit Point of the Pentagram Map, International Mathematics Research Notices 9 (2020) pp. 2818–2831
  • [4] M. Glick, The pentagram map and YY-patterns, Adv. Math. 227, 2012, pp. 1019–1045.
  • [5] A. B. Goncharov and R. Kenyon, Dimers and Cluster Integrable Systems, Ann. Sc. Ec. Norm. Super. (4) 46 (2013) no. 5 pp 747–813
  • [6] A. Izosimov and B. Khesin, Long Diagonal Pentagram Maps, Bulletin of the L.M.S., vol. 55, no. 3, (2023) pp. 1-15
  • [7] A. Izosimov, The pentagram map, Poncelet polygons, and commuting difference operators, Compos. Math. 158 (2022) pp 1084-1124
  • [8] A. Izosimov, Pentagram maps and refactorization in Poisson-Lie groups, Advances in Mathematics, vol. 404 (2022)
  • [9] A. Izosimov, Intersecting the Sides of the Polygon, Proc. A.M.S. 150 (2022) 639-649.
  • [10] B. Khesin, F. Soloviev Integrability of higher pentagram maps, Mathem. Annalen. Vol. 357 no. 3 (2013) pp. 1005–1047
  • [11] B. Khesin, F. Soloviev The geometry of dented pentagram maps, J. European Math. Soc. Vol 18 (2016) pp. 147 – 179
  • [12] G. Mari Beffa, On Generalizations of the Pentagram Map: Discretizations of AGD Flows, Journal of Nonlinear Science, Vol 23, Issue 2 (2013) pp. 304–334
  • [13] G. Mari Beffa, On integrable generalizations of the pentagram map
    Int. Math. Res. Notices (2015) (12) pp. 3669-3693
  • [14] R Felipe and G. Mari-Beffa, The pentagram map on Grassmannians, Ann. Inst. Fourier (Grenoble) 69 (2019) no. 1 pp 421–456
  • [15] Th. Motzkin, The pentagon in the projective plane, with a comment on Napier’s rule, Bull. Amer. Math. Soc. 52, 1945, pp. 985–989.
  • [16] N. Ovenhouse, The Non-Commutative Integrability of the Grassman Pentagram Map, arXiv 1810.11742 (2019)
  • [17] V. Ovsienko, R. E. Schwartz, S. Tabachnikov, The pentagram map: A discrete integrable system, Comm. Math. Phys. 299, 2010, pp. 409–446.
  • [18] V. Ovsienko, R. E. Schwartz, S. Tabachnikov, Liouville-Arnold integrability of the pentagram map on closed polygons, Duke Math. J. Vol 162 No. 12 (2012) pp. 2149–2196
  • [19] R. E. Schwartz, The pentagram map, Exper. Math. 1, 1992, pp. 71–81.
  • [20] R. E. Schwartz, Discrete monodromy, pentagrams, and the method of condensation, J. of Fixed Point Theory and Appl. 3, 2008, pp. 379–409.
  • [21] R. E. Schwartz, A Textbook Case of Pentagram Rigidity, arXiv 2108-07604 preprint (2021)[
  • [22] R. E. Schwartz, Pentagram Rigidity for Centrally Symmetric Octagons, I.M.R.N. (2024) pp 9535–9561
  • [23] F. Soloviev Integrability of the Pentagram Map, Duke Math J. Vol 162. No. 15, (2012) pp. 2815 – 2853
  • [24] M. Weinreich, The Algebraic Dynamics of the Pentagram Map, Ergodic Theory and Dynamical Systems 43 (2023) no. 10, pp. 3460 – 3505