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

The enumeration of extreme rigid honeycombs

Hari Bercovici and Wing Suet Li Mathematics Department
Indiana University
Bloomington, IN 47405
bercovic@indiana.edu School of Mathematics
Georgia Institute of Technology
Atlanta, GA 30332
li@math.gatech.edu
Abstract.

Rigid tree honeycombs were introduced by Knutson, Tao, and Woodward [9] and they were shown in [2] to be sums of extreme rigid honeycombs, with uniquely determined summands up to permutations. Two extreme rigid honeycombs are essentially the same if they have proportional exit multiplicities and, up to this identification, there are countably many equivalence classes of such honeycombs. We describe two ways to approach the enumeration of these equivalence classes. The first method produces a (finite) list of all rigid tree honeycombs of fixed weight by looking at the locking patterns that can be obtained from a certain quadratic Diophantine equation. The second method constructs arbitrary rigid tree honeycombs from rigid overlays of two rigid tree honeycombs with strictly smaller weights. This allows, in principle, for an inductive construction of all rigid tree honeycombs starting with those of unit weight. We also show that some rigid overlays of two rigid tree honeycombs give rise to an infinite sequence of rigid tree honeycombs of increasing complexity but with a fixed number of nonzero exit multiplicities. This last result involves a new inflation/deflation construction that also produces other infinite sequences of rigid tree honeycombs.

WSL was supported in part by Simons Foundation collaborative grant 416045.

1. Introduction

The notion of a rigid honeycomb was introduced in [9] in relation to the Horn problem. It also played a central role in proving that the Horn inequalities hold in an arbitrary factor of type II1 in [2], where it was shown that every rigid honeycomb arises as a special kind of overlay (more general than the clockwise overlays of [9]) of extreme rigid honeycombs. Extreme rigid honeycombs are scalar multiples of the rigid tree measures (which we now call rigid tree honeycombs) studied in [3].

In retrospect, the results of [2] depended on the study of these extreme rigid honeycombs, and one of the initial difficulties was the lack of examples of sufficiently complicated such honeycombs, which in turn hampered effective experimentation. (Producing just six such honeycombs of weight four required an extensive search; an intersection problem generated by one of these six examples is fully analyzed at the end of [2, Section 6].) Subsequently, several infinite families of extreme rigid honeycombs were constructed in [3].

Our purpose is to address the following two questions:

  1. (a)

    Can one effectively enumerate all the rigid tree honeycombs of a given weight ω?\omega?

  2. (b)

    How can one generate rigid tree honeycombs of high weight from rigid tree honeycombs of smaller weight?

In answer to (a), we offer a method based on the study of certain combinatorial objects called locking patterns. This makes the enumeration of (types of) rigid tree honeycombs possible, though still tedious for large weights. For (b), we show that every rigid tree honeycomb of weight at least two can be constructed from an overlay of honeycombs with strictly lower weights. In principle, this allows one to use the enumeration of rigid tree honeycombs of weight less than a fixed kk\in\mathbb{N} to construct all possible such honeycombs of weight kk. In practice, this is difficult to achieve because there are usually many overlays of two honeycombs of given type and these overlays are hard to construct directly. However, once an overlay of two honeycombs, or even a single honeycomb, is known, we show how to construct honeycombs of much greater complexity and weight. These constructions recover, for instance, some of the infinite families constructed in [3] and produce many additional infinite families of rigid tree honeycombs. The honeycombs constructed this way should be instrumental in the experimental study of related conjectures, such as those of [5].

The paper is organized as follows. Basic definitions about honeycombs, puzzles, and duality are covered in Sections 2 and 3. Locking patterns and their connection to the enumeration of rigid tree honeycombs are also covered in Section 3. (The term locking pattern was suggested by Ken Dykema before we really understood what these overlays are, and this is why the notion is not mentioned in [2].) In Section 4, we describe two constructions that help calculate the exit multiplicities of honeycombs supported in the edges of the puzzle of a given rigid honeycomb. One of these results yields the basic inductive construction of rigid tree honeycombs from other rigid tree honeycombs of smaller weights. Section 5 is dedicated to the study of degenerations of a rigid honeycomb, particularly simple degenerations. These simple degenerations, applied to a rigid tree honeycomb, produce either an extreme rigid honeycomb or an overlay of two extreme rigid honeycombs. Necessary conditions are derived on such overlays and Section 6 is dedicated to showing that these necessary conditions are sufficient as well. The arguments are based on a new construction that starts with a puzzle and a compatible honeycomb and produces another puzzle and compatible honeycomb using an inflation and a partial deflation. The constructions produce infinite sequences of rigid tree honeycombs whose weights increase rapidly (like the powers of a number >1>1) but with a fixed number of nonzero exit multiplicities. Some (but not all) of the infinite families produced in [3] are shown to arise from this construction and some new infinite families are produced.

2. Honeycombs and rigidity

We review the definition of a honeycomb as presented in [8]. Fix vectors u1,u2,u3u_{1},u_{2},u_{3} of length 3/3\sqrt{3}/3 in the plane, arranged in counterclockwise order such that u1+u2+u3=0u_{1}+u_{2}+u_{3}=0 (see Figure 2.1), and define unit vectors

w1=u3u2,w2=u1u3,w3=u2u1.w_{1}=u_{3}-u_{2},\quad w_{2}=u_{1}-u_{3},\quad w_{3}=u_{2}-u_{1}.

Every point X2X\in\mathbb{R}^{2} can be written uniquely as X=x1u1+x2u2+x3u3X=x_{1}u_{1}+x_{2}u_{2}+x_{3}u_{3} such that x1,x2,x3x_{1},x_{2},x_{3}\in\mathbb{R} and x1+x2+x3=0x_{1}+x_{2}+x_{3}=0. We view the numbers xjx_{j} as the coordinates of xx. Given a fixed cc\in\mathbb{R}, the equation xj=cx_{j}=c represents a line parallel to wjw_{j}.

Refer to captionw1w_{1}w2w_{2}w3w_{3}u3u_{3}u2u_{2}u1u_{1}
Figure 2.1. The vectors uju_{j} and wjw_{j}

A honeycomb μ\mu is described by two elements, namely,

  1. (1)

    the support of μ\mu. This consists of a finite union I1Ik2I_{1}\cup\cdots\cup I_{k}\subset\mathbb{R}^{2} where

    1. (a)

      each IjI_{j} is a closed segment, a closed ray, or a line,

    2. (b)

      each IjI_{j} is parallel to one of the vectors w1,w2,w3w_{1},w_{2},w_{3}, and

    3. (c)

      if iji\neq j then IiI_{i} and IjI_{j} have at most one point in common that is an endpoint of both IiI_{i} and IjI_{j}.

  2. (2)

    the multiplicities of the honeycomb. These are (not necessarily integer) numbers μ(I1),,μ(Ik)(0,+)\mu(I_{1}),\dots,\mu(I_{k})\in(0,+\infty).

It is convenient to consider that a honeycomb μ\mu assigns zero multiplicity to segments parallel to some wiw_{i} that intersect the support of μ\mu in finitely many points, and that μ\mu assigns multiplicity μ(Ij)\mu(I_{j}) to all subintervals of IjI_{j}. A branch point of a honeycomb is a point where at least three of the sets IjI_{j} meet. The segments IjI_{j} are called edges of μ\mu. The multiplicities of a honeycomb are subject to the following balance condition: Suppose that that J1,J2,J3,J4,J5,J6J_{1},J_{2},J_{3},J_{4},J_{5},J_{6} are six segments, each parallel to some wiw_{i}, containing no branch points in their interior, having a common endpoint XX and arranged clockwise around XX. Then

(2.1) μ(J1)μ(J4)=μ(J5)μ(J2)=μ(J3)μ(J6).\mu(J_{1})-\mu(J_{4})=\mu(J_{5})-\mu(J_{2})=\mu(J_{3})-\mu(J_{6}).

It may be easier to think of a honeycomb as a Borel measure in the plane that assigns μ(Ij)×length(A)\mu(I_{j})\times\text{length}(A) to each Borel subset AIjA\subset I_{j}. This allows us to add honeycombs, thus giving the collection of all honeycombs the structure of a convex cone.

In this work, we do not distinguish between two honeycombs that are simply translates of each other, with the translation preserving multiplicities.

Suppose that μ\mu is a honeycomb with support I1IkI_{1}\cup\cdots\cup I_{k}. It is easy to see that any endpoint of some IjI_{j} must also be an endpoint of another II_{\ell}. The support of a honeycomb μ\mu must contain lines or rays. Each ray points in the direction of one of the vectors ±wj\pm w_{j}, j=1,2,3j=1,2,3. Figure 2.2 shows a few examples of honeycombs. All edge multiplicities, with one exception indicated by a thicker line, are equal to 11.

Refer to caption
Figure 2.2. Some honeycombs

We denote by \mathcal{M} (respectively, \mathcal{M}_{*}) the collection of all honeycombs with the property that the rays in the support of μ\mu point in the direction of w1,w2,w_{1},w_{2}, or w3w_{3} (respectively w1,w2-w_{1},-w_{2}, or w3-w_{3}). The second and third examples in Figure 2.2 represent a honeycomb in \mathcal{M} and its mirror image (relative to a vertical line) in \mathcal{M}_{*}. Suppose that μ\mu is a honeycomb in \mathcal{M}, cc\in\mathbb{R}, and j{1,2,3}.j\in\{1,2,3\}. If the support of μ\mu contains a ray I{xj=c}I\subset\{x_{j}=c\}, we define

(2.2) μ(j)(c)=μ(I).\mu^{(j)}(c)=\mu(I).

Otherwise, we set μ(j)(c)=0.\mu^{(j)}(c)=0. The numbers μ(j)(c)\mu^{(j)}(c) are called the exit multiplicities of μ\mu. (The ordered collection of the nonzero exit multiplicities will later be called the exit pattern of μ.)\mu.) Of course, μ\mu only has finitely many nonzero exit multiplicities and the balance condition (2.1) is easily seen (see, for instance, [8]) to imply the identity

(2.3) cμ(1)(c)=cμ(2)(c)=cμ(3)(c).\sum_{c\in\mathbb{R}}\mu^{(1)}(c)=\sum_{c\in\mathbb{R}}\mu^{(2)}(c)=\sum_{c\in\mathbb{R}}\mu^{(3)}(c).

This common value is called the weight of μ\mu and is denoted

(2.4) ω(μ)=cμ(1)(c).\omega(\mu)=\sum_{c\in\mathbb{R}}\mu^{(1)}(c).

The exit multiplicities also satisfy the equality

(2.5) cc(μ(1)(c)+μ(2)(c)+μ(3)(c))=0,\sum_{c\in\mathbb{R}}c(\mu^{(1)}(c)+\mu^{(2)}(c)+\mu^{(3)}(c))=0,

sometimes known as the trace identity on account of its connection with linear algebra.

A honeycomb μ\mu\in\mathcal{M} (or μ\mu\in\mathcal{M}_{*}) is said to be rigid if there is no other honeycomb that has the same exit multiplicities as μ\mu. It is easy to see that an arbitrary honeycomb is not rigid if it has a branch point adjacent to six edges in its support; see Figure 2.3.

Refer to caption
Figure 2.3. Three honeycombs with identical exit multiplicities

In fact, one can determine whether a honeycomb μ\mu is rigid just by looking at its support. Let μ\mu be a honeycomb and let A,B,CA,B,C be branch points of μ\mu such that the segments ABAB and BCBC are edges of μ\mu. We say that ABCABC is an evil turn [2, p. 1592] if one of the following situations occurs:

  1. (1)

    C=AC=A, and the support of μ\mu contains edges BX,BY,BZBX,BY,BZ that are 120,180120^{\circ},180^{\circ}, and 240240^{\circ} clockwise from ABAB.

  2. (2)

    BCBC is 120120^{\circ} clockwise from ABAB.

  3. (3)

    CAC\neq A and A,B,CA,B,C are collinear.

  4. (4)

    BCBC is 120120^{\circ} counterclockwise from ABAB and the support of μ\mu contains an edge BXBX that is 120120^{\circ} clockwise from ABAB.

  5. (5)

    BCBC is 6060^{\circ} counterclockwise from ABAB and the support of μ\mu contains edges BX,BYBX,BY which are 120120^{\circ} and 180180^{\circ} clockwise from ABAB.

Refer to captionA=CA=CBBYYXXZZAABBCCAABBCCAABBXXCCAABBCCYYXX
Figure 2.4. Evil turns

The possible evil turns are illustrated in Figure 2.4 where the edges that must be in the support of μ\mu are labeled, but the support may contain all six edges incident to BB. A sequence A1,A2,,An,An+1=A1A_{1},A_{2},\dots,A_{n},A_{n+1}=A_{1} of branch points of μ\mu is called an evil loop if Aj1AjAj+1A_{j-1}A_{j}A_{j+1} is an evil turn for every j=1,,nj=1,\dots,n. Each of the first three configurations in Figure 2.5 contains exactly one evil loop (up to cyclic permutations) but the fourth one contains none.

Refer to caption
Figure 2.5. Some loops

The following result is proved as in [2, Proposition 2.2]. (The proof uses a characterization of rigidity via puzzles; see [9, Theorems 4 and 7].)

Proposition 2.1.

A honeycomb μ\mu\in\mathcal{M} is rigid if and only if its support contains none of the following configurations:

  1. (1)

    Six edges meeting in one branch point.

  2. (2)

    An evil loop.

Suppose that e=ABe=AB and f=BCf=BC are edges of a rigid honeycomb μ\mu. As in [2], we write eμfe\to_{\mu}f (or efe\to f when μ\mu is understood; see Figure 2.6) if one of the following two situations arises:

  1. (1)

    ABC=120\varangle ABC=120^{\circ} and the segment BXBX opposite ABAB is not in the support of μ\mu.

  2. (2)

    ee and ff are opposite and there exists a segment BXBX not in the support of μ\mu such that CBX=60\varangle CBX=60^{\circ}.

Refer to captionBBAACCXXBBCCAAXX
Figure 2.6. The relation ‘\to

The balance condition shows that μ(e)μ(f)\mu(e)\leq\mu(f) whenever efe\to f. In fact, μ(f)\mu(f) equals μ(e)\mu(e) if efe\to f and fef\to e, and it equals μ(e)+μ(e)\mu(e)+\mu(e^{\prime}) when efe\to f, efe^{\prime}\to f, and eee\neq e^{\prime}. The transitive closure \Rightarrow of the relation \to was called descendant in [2]. Thus, efe\Rightarrow f if there exists a descendance path, that is, a sequence e1ene_{1}\dots e_{n} of edges such that e1=e,en=f,e_{1}=e,e_{n}=f, and ejej+1e_{j}\to e_{j+1}, j=1,,n1j=1,\dots,n-1. Given such a path, ee is called an ancestor of ff and ff is called a descendant of ee. We write efe\Leftrightarrow f, and we say that ee and ff are equivalent, if either e=fe=f or both efe\Rightarrow f and fef\Rightarrow e hold. An edge ee of μ\mu is said to be a root edge if the relation fef\Rightarrow e implies efe\Leftrightarrow f. The above observations show that the multiplicity of an arbitrary edge of μ\mu can be written as a linear combination with positive integer coefficients of the multiplicities of root edges. More precisely, suppose that μ\mu is a rigid honeycomb and that e1,,ene_{1},\dots,e_{n} is a maximal sequence of pairwise inequivalent root edges. Then an arbitrary edge ff of μ\mu satisfies

μ(f)=j=1ndjμ(ej),\mu(f)=\sum_{j=1}^{n}d_{j}\mu(e_{j}),

where dj0d_{j}\geq 0 is the number of distinct descendance paths from eje_{j} to ff. Another way of formulating this is to say that

μ=j=1nμ(ej)μj,\mu=\sum_{j=1}^{n}\mu(e_{j})\mu_{j},

where each μj\mu_{j} is itself a rigid honeycomb (see [2, Section 3]). The summands μj\mu_{j} belong to extreme rays in the cone \mathcal{M}. Extreme rigid honeycombs are characterized by the fact that they have a unique equivalence class of root edges. An extreme rigid honeycomb is a tree honeycomb precisely when the multiplicity assigned to its root edges is equal to 11. Alternatively, rigid tree honeycombs are obtained as immersions of a special kind of tree that we now define.

The trees relevant to the construction of rigid honeycombs are finite trees whose edges are labeled 1,2,1,2, or 33, and the following conditions are satisfied:

  1. (1)

    each vertex has order 11 or 33, and

  2. (2)

    the three edges adjacent to a vertex of order 33 have distinct labels.

An immersion of a tree is a map ff that associates to each vertex vv of order 3 of a tree a point f(v)f(v) in the plane such that the following conditions are satisfied:

  1. (i)

    if aa and bb are joined by an edge labeled jj, then f(a)f(b)f(a)\neq f(b) and the segment f(a)f(b)f(a)f(b) is parallel to wjw_{j}, and

  2. (ii)

    if xa,xbxa,xb, and xcxc are three edges adjacent to a vertex xx then the angle between any two of the segments f(x)f(a),f(x)f(b)f(x)f(a),f(x)f(b), and f(x)f(c)f(x)f(c) equals 120120^{\circ}.

Given a tree TT that has at least two vertices of order 33, and an immersion ff of TT, we define a honeycomb μf\mu_{f} as follows. Let abab be an edge of TT. If aa and bb have order 33, we add the segment f(a)f(b)f(a)f(b) to the support of μf\mu_{f} and we add 11 to the multiplicity of this segment. If aa is of order 11, abab is labeled jj, and c,dc,d are the other vertices connected to bb, we add to the support of μf\mu_{f} a ray hh with endpoint f(b)f(b) and parallel to wjw_{j} and we add 11 to the multiplicity of this ray. The ray is chosen such that the angles between any two of h,f(b)f(c)h,f(b)f(c), and f(b)f(d)f(b)f(d) equal 120120^{\circ}. The segments f(a)f(b)f(a)f(b) and the rays constructed above can be viewed as images under the immersion ff of the edges of TT. The honeycomb μf\mu_{f} is simply determined by adding all of the multiplicities just defined. If TT has exactly one vertex of order 33, there are two possible honeycombs associated to an immersion ff, one in \mathcal{M} and the other in \mathcal{M}_{*}. The honeycombs obtained from immersions of trees are called tree honeycombs.

Refer to caption222222222222221111111111111133333333333333
Figure 2.7. A tree and two of its immersions

Not every tree honeycomb is rigid. Figure 2.7 represents a tree TT and two immersions of TT, the second of which yields a rigid honeycomb.

Suppose that μ\mu\in\mathcal{M} is a rigid honeycomb with a unique equivalence class of root edges, all of which have multiplicity 11. We construct a tree T0T_{0} as follows. Fix a root edge e0=XYe_{0}=XY of μ\mu. The vertices of T0T_{0} are x,yx,y, and the descendance paths e0ene_{0}\dots e_{n}. Given such a descendance path, the vertices v=e0en1v=e_{0}\dots e_{n-1} and v=e0env^{\prime}=e_{0}\dots e_{n} are joined by an edge of T0T_{0} labeled jj if ene_{n} is parallel to wjw_{j}. In addition, xx is joined to yy by an edge labeled jj if XYXY is parallel to wjw_{j}, xx is joined to e0e1e_{0}e_{1} if XX is a an endpoint of e1e_{1}, and yy is joined to e0e1e_{0}e_{1} if YY an endpoint of e1e_{1}. The same labeling rule applies to the last two kinds of edges. The tree T0T_{0} obtained this way may have some vertices of order 22. To obtain a tree TT satisfying condition (1) we simply remove the vertices of order 22 as follows: suppose that v0v1vnv_{0}v_{1}\dots v_{n} is a path in T0T_{0} such that v0v_{0} and vnv_{n} have order 11 or 33 and v1,,vn1v_{1},\dots,v_{n-1} have order 22. Then v1,,vn1v_{1},\dots,v_{n-1} are removed from T0T_{0} and an edge joining v0v_{0} and vnv_{n}, carrying the label of v0v1v_{0}v_{1}, is added instead. Clearly, we have μ=μf\mu=\mu_{f} for some immersion ff of TT. The construction of T0T_{0} and TT is illustrated in Figure 2.8 for a particular rigid tree honeycomb.

Refer to captionXXYYxxyyxxyy22222222222222222222111111111111111111111133333333333333333333
Figure 2.8. The construction of T0T_{0} and TT

Suppose that TT is a tree, that ff is an immersion of TT, and that μ=μf\mu=\mu_{f} belongs to \mathcal{M}. Suppose also that there exists an edge xyxy of TT such that μ\mu assigns unit multiplicity to part of f(x)f(y)f(x)f(y). Orient the edges of TT other than xyxy away from xyxy. We say that the immersion ff is coherent if the following condition is satisfied: if a segment II is contained in the images under ff of two different edges of TT, then the two orientations induced on II are the same. The following result is [4, Theorem 4.1].

Theorem 2.2.

Suppose that TT is a tree, that ff is an immersion of TT, and that μ=μf\mu=\mu_{f}\in\mathcal{M}. Then μ\mu is rigid if and only if the following three conditions are satisfied.

  1. (1)

    There exists an edge xyxy of TT such that μ\mu assigns multiplicity 11 to part of f(x)f(y)f(x)f(y).

  2. (2)

    The immersion ff is coherent.

  3. (3)

    There is no branch point XX of μ\mu such that four segment XA1,XA2,XB1,XA_{1},XA_{2},XB_{1}, and XB2XB_{2} have nonzero multiplicity, the points A1XA2A_{1}XA_{2} are collinear, and the points B1XB2B_{1}XB_{2} are collinear; see Figure 2.9.

Refer to captionA1A_{1}A2A_{2}XXB1B_{1}B2B_{2}
Figure 2.9. A crossing

Having obtained a description of rigid tree honeycombs, it is important to determine which linear combinations c1μ1++cnμnc_{1}\mu_{1}+\cdots+c_{n}\mu_{n}, c1,,cn>0c_{1},\dots,c_{n}>0, of rigid tree honeycombs are rigid. Such combinations are called rigid overlays and their description involves a bilinear map [3] defined on honeycombs in \mathcal{M} as follows: given μ,ν\mu,\nu\in\mathcal{M}, we set

(2.6) Σν(μ)=c<d[μ(1)(c)ν(1)(d)+μ(2)(c)ν(2)(d)+μ(3)(c)ν(3)(d)]ω(μ)ω(ν),\Sigma_{\nu}(\mu)=\sum_{c<d}[\mu^{(1)}(c)\nu^{(1)}(d)+\mu^{(2)}(c)\nu^{(2)}(d)+\mu^{(3)}(c)\nu^{(3)}(d)]-\omega(\mu)\omega(\nu),

where μ(j)(c)\mu^{(j)}(c) is as defined above (2.2). It was seen in [3, end of Section 2] that

(2.7) Σμ(μ)=12[ω(μ)2c[μ(1)(c)2+μ(2)(c)2+μ(3)(c)2]].\Sigma_{\mu}(\mu)=\frac{1}{2}\left[\omega(\mu)^{2}-\sum_{c\in\mathbb{R}}[\mu^{(1)}(c)^{2}+\mu^{(2)}(c)^{2}+\mu^{(3)}(c)^{2}]\right].

The following result is [3, Theorem 4.2].

Theorem 2.3.

Let μ\mu and ν\nu be two distinct rigid tree honeycombs in \mathcal{M}. Then Σν(μ)0\Sigma_{\nu}(\mu)\geq 0 and Σν(ν)=1\Sigma_{\nu}(\nu)=-1.

It is possible that Σν(μ)>0\Sigma_{\nu}(\mu)>0 and Σμ(ν)>0\Sigma_{\mu}(\nu)>0 for two rigid tree honeycombs μ\mu and ν\nu, see Figure 2.10.

Refer to caption
Figure 2.10. Σμ(ν)>0\Sigma_{\mu}(\nu)>0 and Σν(μ)>0\Sigma_{\nu}(\mu)>0

It follows from (2.7) that the second assertion in Theorem 2.3 can be written as

(2.8) c[μ(1)(c)2+μ(2)(c)2+μ(3)(c)2]=ω(μ)2+2.\sum_{c\in\mathbb{R}}[\mu^{(1)}(c)^{2}+\mu^{(2)}(c)^{2}+\mu^{(3)}(c)^{2}]=\omega(\mu)^{2}+2.
Remark 2.4.

Suppose that μ\mu and ν\nu are distinct rigid honeycombs in \mathcal{M} obtained from immersions ff of TT and gg of SS, respectively. Then [3, Theorem 5.1] provides a method for the calculation of Σν(μ)\Sigma_{\nu}(\mu). Fix an edge ee in SS that is mapped by gg onto a root edge of ν\nu that is not contained in the support of μ\mu. Orient all the other edges of SS away from ee. Then Σν(μ)\Sigma_{\nu}(\mu) is the total number of occurrences of events of the four types indicated in the Figure 2.11.

Refer to captionμ\muν\nuμ\muν\nuμ\muν\nuμ\muν\nu
Figure 2.11. Calculation of Σν(μ)\Sigma_{\nu}(\mu)

In these diagrams, the dashed lines represent the images under gg of edges of SS, the dotted lines represent the images under ff of edges of TT, and the solid lines represent common edges of μ\mu and ν\nu. (Note that, in counting these events we consider edges of the trees rather than edges of the honeycombs themselves. Thus, when looking at edges of the honeycombs, one must take into account their multiplicities.)

Refer to captionμ\muν\nu
Figure 2.12. μ\mu is clockwise from ν\nu

The following characterization of rigid overlays is [3, Corollary 4.5] (see also [2, Theorem 3.8]).

Theorem 2.5.

Let μ1,,μn\mu_{1},\dots,\mu_{n} be distinct rigid tree honeycombs and let c1,,cnc_{1},\dots,c_{n} be positive numbers. Then the following are equivalent:

  1. (1)

    The honeycomb j=1ncjμj\sum_{j=1}^{n}c_{j}\mu_{j} is rigid.

  2. (2)

    There exists a permutation ν1,,νn\nu_{1},\dots,\nu_{n} of μ1,,μn\mu_{1},\dots,\mu_{n} such that Σνi(νj)=0\Sigma_{\nu_{i}}(\nu_{j})=0 for i>ji>j.

3. Puzzles and duality

Suppose that μ\mu\in\mathcal{M}. The puzzle of μ\mu is obtained by a process of inflation that we describe next. Cut the plane along the edges of μ\mu to obtain a collection of white puzzle pieces and translate these pieces away from each other in such a way that, given an edge ABAB of μ\mu, the parallelogram formed by the two translates of ABAB has two sides of length μ(AB)\mu(AB) that are 6060^{\circ} clockwise from ABAB. We call this parallelogram the inflation of ABAB and we color it dark gray. The balance condition for μ\mu implies that these white and dark gray pieces fit together and leave a space corresponding to each branch point of μ\mu. These extra spaces are colored light gray. The light gray piece associated to a branch point XX of μ\mu is a convex polygon with as many sides as there are edges incident to XX. One can recover the honeycomb μ\mu, up to a translation, from its puzzle by deflation. This process consists simply of moving the white pieces together and replacing each dark gray parallelogram (or half-infinite strip) by a segment parallel to its white sides and with multiplicity equal to the length of its light gray side. There are similar processes of *-inflation and *-deflation which are the same as inflation and deflation except that ‘clockwise’ is replaced by ‘counterclockwise’ in the construction of the *-inflation.

Starting with a nonzero honeycomb μ\mu\in\mathcal{M}, one can construct the puzzle of μ\mu and apply *-deflation to this puzzle. The only difficulty is that the segments arising from dark gray infinite strips would be assigned infinite multiplicity. To overcome this difficulty, choose a closed equilateral triangle Δ\Delta containing all the branch points of μ\mu whose sides are 6060^{\circ} clockwise from those rays in the support of μ\mu that they intersect. Apply now the process of inflation only to the part of the support of μ\mu that is contained in Δ\Delta and apply *-deflation to the resulting puzzle. The result of these operations is the part of a honeycomb μΔ\mu_{\Delta}^{*}\in\mathcal{M}_{*} contained in an equilateral triangle Δ\Delta_{*} with sides equal to ω(μ)\omega(\mu). (A simple example is illustrated in Figure 3.1.)

Refer to caption
Figure 3.1. A honeycomb μ\mu in Δ\Delta, its puzzle, the *-deflation, and the honeycombs μΔ\mu_{\Delta}^{*} and μ\mu^{*}

Choosing a different triangle Δ\Delta changes the multiplicities of μΔ\mu_{\Delta}^{*} only on the boundary of Δ\Delta_{*}. Denote by Δ0\Delta_{0} the smallest triangle satisfying the requirements above. We define μ\mu^{*} to be equal to μΔ0\mu_{\Delta_{0}}^{*}. The honeycomb μ\mu^{*} has the property that each side of Δ\Delta_{*} contains a segment with zero multiplicity.

In general, we observe that the sides of Δ\Delta_{*} are divided by the rays in the support of μ\mu^{*} into segments with lengths equal to the exit multiplicities of μ\mu. (The triangle Δ\Delta_{*} is partitioned by the support of μ\mu^{*} into light gray pieces that form the degeneracy graph of μ\mu introduced in [8].) If μ\mu has only one branch point, we define μ=0\mu^{*}=0.

Remark 3.1.

As just observed, the nonzero exit multiplicities of a honeycomb μ\mu are precisely the lengths of the segments on Δ\partial\Delta_{*} between consecutive exit rays of μ\mu^{*}. For instance, suppose that the support of μ\mu^{*} is contained in the support of ν.\nu^{*.} If exit(μ)=exit(ν),{\rm exit(\mu)={\rm exit}(\nu)},then μ\mu^{*} and ν\nu^{*} share their exit rays, and therefore μ\mu and ν\nu have precisely the same nonzero exit multiplicities. On the other hand, if exit(μ)=exit(ν)1{\rm exit}(\mu)={\rm exit}(\nu)-1, then μ\mu^{*} has one fewer exit ray, and thus μ\mu has the same nonzero exit multiplicities, except that the two nonzero exit multiplicities of ν\nu corresponding to segments separated by the extra exit ray are replaced by their sum. (If exit(μ){\rm exit}(\mu) is even smaller, more of the multiplicities of ν\nu are aggregated in this manner into multiplicities of μ\mu.)

It is easily seen that a honeycomb μ\mu\in\mathcal{M} is rigid if and only if μ\mu^{*}\in\mathcal{M}_{*} is rigid (see [2, p. 1591]). Given a rigid honeycomb μ\mu\in\mathcal{M}, we denote by exit(μ){\rm exit}(\mu) the number of nonzero exit multiplicities of μ\mu, and we denote by root(μ){\rm root}(\mu) the number of equivalence classes of root edges in the support of μ\mu. As noted earlier, μ\mu is a linear combination with positive coefficients of root(μ){\rm root}(\mu) distinct rigid tree honeycombs. The following result is equivalent to [1, Theorem 1.1]. (To see the equivalence, observe that the number of attachment points of μ\mu relative to the minimal triangle Δ0\Delta_{0} considered above is equal to exit(μ)3{\rm exit}(\mu)-3.)

Theorem 3.2.

For every rigid honeycomb μ\mu\in\mathcal{M} we have

root(μ)+root(μ)=exit(μ)2.{\rm root}(\mu)+{\rm root}(\mu^{*})={\rm exit}(\mu)-2.

In order to study the exit multiplicities of a rigid tree honeycomb, we require an auxiliary result of convex analysis.

Lemma 3.3.

Let K1K_{1} and K2K_{2} be two convex sets such that K2K_{2} is finite-dimensional, and let f:K1K2f:K_{1}\to K_{2} be a surjective affine map. Suppose that there exists an interior point b0K2b_{0}\in K_{2} such that f1(b0)f^{-1}(b_{0}) is a singleton. Then ff is injective.

Proof.

Let a0K1a_{0}\in K_{1} satisfy f(a0)=b0f(a_{0})=b_{0} and suppose that a1,a1′′K1a^{\prime}_{1},a_{1}^{\prime\prime}\in K_{1} are such that f(a1)=f(a1′′)=b1f(a_{1}^{\prime})=f(a_{1}^{\prime\prime})=b_{1}. Since b0b_{0} is an interior point of K2K_{2}, there exist b2K2b_{2}\in K_{2} and t(0,1)t\in(0,1) such that b0=tb1+(1t)b2b_{0}=tb_{1}+(1-t)b_{2}. Choose a2f1(b2)a_{2}\in f^{-1}(b_{2}). Since ff is affine, we deduce that

f(ta1+(1t)a2)\displaystyle f(ta_{1}^{\prime}+(1-t)a_{2}) =tb1+(1t)b2=b0,\displaystyle=tb_{1}+(1-t)b_{2}=b_{0},
f(ta1′′+(1t)a2)\displaystyle f(ta_{1}^{\prime\prime}+(1-t)a_{2}) =tb1+(1t)b2=b0,\displaystyle=tb_{1}+(1-t)b_{2}=b_{0},

and thus

ta1+(1t)a2=ta1′′+(1t)a2=a0.ta_{1}^{\prime}+(1-t)a_{2}=ta_{1}^{\prime\prime}+(1-t)a_{2}=a_{0}.

It follows immediately that a1=a1′′a^{\prime}_{1}=a_{1}^{\prime\prime}. ∎

Fix a positive number rr and consider an equilateral triangle Δ=ABC\Delta=ABC with side length rr and sides parallel to w1,w2,w_{1},w_{2}, and w3w_{3}. For simplicity, suppose that A,B,A,B, and CC are arranged counterclockwise around Δ\Delta and ABAB is parallel to w1w_{1}.

Refer to captionA1A_{1}A2A_{2}B1B_{1}B2B_{2}B3B_{3}C1C_{1}C2C_{2}
Figure 3.2. The pattern 2,1,2|1,1,1,2|3,1,12,1,2|1,1,1,2|3,1,1

A pattern consists of a sequence A0=A,A1,,Ak11,Ak1=B=B0,B1,,Bk21,Bk2=C=C0,C1,,Ck31,Ck3=AA_{0}=A,A_{1},\dots,A_{k_{1}-1},A_{k_{1}}=B=B_{0},B_{1},\dots,B_{k_{2}-1},B_{k_{2}}=C=C_{0},C_{1},\dots,C_{k_{3}-1},C_{k_{3}}=A of points arranged counterclockwise on the boundary of Δ\Delta; see Figure 3.2. Alternatively, a pattern is determined by the distances α1(1),,αk1(1)|α1(2),,αk2(2)|α1(3),,αk3(3)\alpha_{1}^{(1)},\dots,\alpha_{k_{1}}^{(1)}|\alpha_{1}^{(2)},\dots,\alpha_{k_{2}}^{(2)}|\alpha_{1}^{(3)},\dots,\alpha_{k_{3}}^{(3)} defined by

αj(1)=length(Aj1,Aj),αj(2)=length(Bj1,Bj),αj(3)=length(Cj1,Cj).\alpha_{j}^{(1)}=\text{length}(A_{j-1},A_{j}),\ \alpha_{j}^{(2)}=\text{length}(B_{j-1},B_{j}),\ \alpha_{j}^{(3)}=\text{length}(C_{j-1},C_{j}).

These positive numbers are subject to

j=1k1αj(1)=j=1k2αj(2)=j=1k3αj(3)=r.\sum_{j=1}^{k_{1}}\alpha_{j}^{(1)}=\sum_{j=1}^{k_{2}}\alpha_{j}^{(2)}=\sum_{j=1}^{k_{3}}\alpha_{j}^{(3)}=r.
Definition 3.4.

We say that a pattern is a locking pattern if the following condition is satisfied: Suppose that μ\mu\in\mathcal{M} is such that the branch points of μ\mu are contained in Δ\Delta and that the boundary of Δ\Delta only intersects the rays in the support of μ\mu in a subset of {Aj}j=0k1{Bj}j=0k2{Cj}j=0k3.\{A_{j}\}_{j=0}^{k_{1}}\cup\{B_{j}\}_{j=0}^{k_{2}}\cup\{C_{j}\}_{j=0}^{k_{3}}. Then μ\mu is rigid.

It is easy to see that \mathcal{M} can be replaced by \mathcal{M}_{*} in the above definition. Moreover, the above condition need only be satisfied for honeycombs μ\mu\in\mathcal{M} with rays intersecting Δ\partial\Delta in a subset of {Aj}j=0k11{Bj}j=0k21{Cj}j=0k31\{A_{j}\}_{j=0}^{k_{1}-1}\cup\{B_{j}\}_{j=0}^{k_{2}-1}\cup\{C_{j}\}_{j=0}^{k_{3}-1}; such honeycombs are allowed to have a ray meeting Δ\partial\Delta in A0(=Ck3)A_{0}(=C_{k_{3}}) that is parallel to the other rays passing through some AjA_{j} with j0j\neq 0 but not such a ray that is parallel to the ones passing through some CjC_{j} for jk3.j\neq k_{3}. A similar observation (with A0,B0,C0A_{0},B_{0},C_{0} in place of Ak1,Bk2,Ck3A_{k_{1}},B_{k_{2}},C_{k_{3}}) applies to honeycombs in \mathcal{M}_{*}. It is also obvious that a pattern that has a locking refinement must itself be locking.

Suppose now that μ\mu\in\mathcal{M} and the only nonzero exit multiplicities of μ\mu are αj(1)=μ(1)(cj(1))\alpha_{j}^{(1)}=\mu^{(1)}(c_{j}^{(1)}), αj(2)=μ(2)(cj(2))\alpha_{j}^{(2)}=\mu^{(2)}(c_{j}^{(2)}), and αj(3)=μ(3)(cj3)\alpha_{j}^{(3)}=\mu^{(3)}(c_{j}^{3}), where c1(1)<<ck1(1)c_{1}^{(1)}<\cdots<c_{k_{1}}^{(1)}, c1(2)<<ck2(2)c_{1}^{(2)}<\cdots<c_{k_{2}}^{(2)}, and c1(3)<<ck3(3)c_{1}^{(3)}<\cdots<c_{k_{3}}^{(3)}. These numbers define a pattern with r=ω(μ)r=\omega(\mu) that we call the exit pattern of μ\mu.

Proposition 3.5.

Suppose that μ\mu\in\mathcal{M} is a rigid tree honeycomb. Then the exit pattern of μ\mu is a locking pattern.

Proof.

Choose rigid tree honeycombs ρ1,,ρn\rho_{1},\dots,\rho_{n}\in\mathcal{M}_{*} and c1,,cn>0c_{1},\dots,c_{n}>0 such that μ=j=1ncjρj\mu^{*}=\sum_{j=1}^{n}c_{j}\rho_{j}. Theorem 3.2 and the hypothesis imply that n=exit(μ)3n=\text{exit}(\mu)-3. Recall that all the branch points of μ\mu^{*} are contained in a triangle Δ\Delta_{*} with side lengths ω(μ)\omega(\mu) and that the rays of μ\mu^{*} determine precisely the exit pattern of μ\mu on the boundary of Δ\Delta_{*}. Denote by K1K_{1} the collection of all honeycombs in \mathcal{M}_{*} whose branch points are contained in Δ\Delta_{*} and whose rays intersect the boundary of Δ\Delta_{*} in a subset of {Aj}j=1k1{Bj}j=1k2{Cj}j=1k3.\{A_{j}\}_{j=1}^{k_{1}}\cup\{B_{j}\}_{j=1}^{k_{2}}\cup\{C_{j}\}_{j=1}^{k_{3}}. Observe that k1+k2+k3=exit(μ)k_{1}+k_{2}+k_{3}={\rm exit}(\mu). Given νK1\nu\in K_{1}, denote by f(ν)exit(μ)f(\nu)\in\mathbb{R}^{{\rm exit}(\mu)} the vector (a1,,ak1,b1,,bk2,c1,,ck3)(a_{1},\dots,a_{k_{1}},b_{1},\dots,b_{k_{2}},c_{1},\dots,c_{k_{3}}), where aja_{j} is the exit multiplicity that ν\nu assigns to the ray passing through AjA_{j}, and bj,cjb_{j},c_{j} are defined similarly. Denote by K2exit(μ)K_{2}\subset\mathbb{R}^{{\rm exit}(\mu)} the range of ff. In order to apply Lemma 3.3, we verify that f(μ)f(\mu^{*}) is an interior point of K2K_{2}. First, we note that the dimension of K2K_{2} is at most exit(μ)3{\rm exit}(\mu)-3 because every honeycomb must satisfy (2.3) and (2.5). On the other hand, ff is injective on the set S={j=1ntjρj:t1,,tn>0}S=\{\sum_{j=1}^{n}t_{j}\rho_{j}:t_{1},\dots,t_{n}>0\} and thus f(K1)f(K_{1}) has dimension at least exit(μ)3{\rm exit}(\mu)-3. It follows that K2K_{2} has dimension exactly exit(μ)3{\rm exit}(\mu)-3 and that each point in f(S)f(S) is an interior point of K2K_{2}. The proposition follows from Lemma 3.3. ∎

The multiplicity pattern of a rigid tree honeycomb has an additional property that helps determine whether a given pattern arises from such a honeycomb. Suppose that an arbitrary pattern A0,,Ak1,B0,,Bk2,C0,Ck3A_{0},\dots,A_{k_{1}},B_{0},\dots,B_{k_{2}},C_{0},\dots C_{k_{3}} is given on a triangle Δ\Delta with sides parallel to w1,w2,w_{1},w_{2}, and w3w_{3}. A convex region ΩΔ\Omega\subset\Delta is said to be flat relative to this pattern if every honeycomb μ\mu\in\mathcal{M}, whose rays intersect Δ\partial\Delta in a subset of {Aj,Bj,Cj}\{A_{j},B_{j},C_{j}\}, assigns zero multiplicity to all segments contained in the interior of Ω\Omega.

Proposition 3.6.

Let μ\mu\in\mathcal{M} be a rigid tree honeycomb. Then the light gray pieces determined in Δ\Delta_{*} by the support of μ\mu^{*} are flat relative to the exit pattern of μ\mu.

Proof.

We denote by 𝒩\mathcal{N}\subset\mathcal{M}_{*} the collection of those honeycombs that have branch points contained in Δ\Delta_{*} and exit points among A0,,Ak1,B0,,Bk2,C0,Ck3A_{0},\dots,A_{k_{1}},B_{0},\dots,B_{k_{2}},C_{0},\dots C_{k_{3}}. Suppose that IΔI\subset\Delta_{*} is a segment parallel to wjw_{j} for some j=1,2,3j=1,2,3. We say that II is flat if the following condition is satisfied for every ν𝒩\nu\in\mathcal{N}:

  1. (\dagger)

    There is no edge JJ of ν\nu such that JIJ\cap I is a single point in the interior of II.

If ABAB and ACAC are two flat edges, each parallel to w1,w2,w_{1},w_{2}, or w3,w_{3}, and BAC=60\varangle BAC=60^{\circ}, then the smallest convex polygon with sides parallel to w1,w2,w_{1},w_{2}, or w3w_{3} containing A,B,A,B, and CC is flat relative to the exit pattern of μ\mu. This polygon is either an equilateral triangle or a trapezoid. It suffices to show that all the edges of μ\mu^{*} in Δ\Delta_{*} are flat. These edges are of the form ee^{*}, with ee an edge of μ\mu, where ee^{*} is a segment of length μ(e)\mu(e) that is 6060^{\circ} clockwise from ee. Denote by SS the set of all edges of μ\mu and by S1SS_{1}\subset S the collection of those edges ee with the property that ee^{*} is flat. It is clear that the rays of μ\mu belong to S1S_{1}. We show that the set S1S_{1} has the following properties (see Figure 3.3):

Refer to captione1e_{1}e2e_{2}e3e_{3}e1e_{1}e2e_{2}e3e_{3}e4e_{4}
Figure 3.3. The edges eje_{j}
  1. (1)

    If e1,e2,e_{1},e_{2}, and e3e_{3} are the only three edges meeting at a branch point, and if e1,e2S1e_{1},e_{2}\in S_{1}, then e3S1e_{3}\in S_{1}.

  2. (2)

    If e1,e2,e3e_{1},e_{2},e_{3}, and e4e_{4} are four edges meeting at a branch point of μ\mu such that μ(e1)=μ(e2)+μ(e4)\mu(e_{1})=\mu(e_{2})+\mu(e_{4}), and if e1,e2S1e_{1},e_{2}\in S_{1}, then e3,e4S1e_{3},e_{4}\in S_{1}.

  3. (3)

    If e1,e2,e3e_{1},e_{2},e_{3}, and e4e_{4} are four edges meeting at a branch point of μ\mu such that μ(e1)=μ(e2)+μ(e4)\mu(e_{1})=\mu(e_{2})+\mu(e_{4}) and e1e_{1} is collinear with e4e_{4}, and if e1S1e_{1}\in S_{1}, then e4S1e_{4}\in S_{1}.

Supposing for the moment that properties (1)–(3) have been proved, we argue by contradiction that S1=SS_{1}=S. If S1SS_{1}\neq S, choose an edge eS\S1e\in S\backslash S_{1} with the property that the length of a descendance path from some root edge to ee is the largest possible. All the descendants of ee have longer descendance paths from the same root edge and hence they belong to S1S_{1}. Since ee is not a ray, it must have descendants. There are three possibilities:

  1. (a)

    ee has two descendants e2e_{2} and e3e_{3} and there is no other edge adjacent to their common endpoint. This possibility is excluded by property (1) above because e2,e3S1e_{2},e_{3}\in S_{1}.

  2. (b)

    ee has two descendants e1,e3e_{1},e_{3}, and there is a fourth edge e4e_{4} adjacent to their common endpoint. This possibility is excluded by (2) for similar reasons.

  3. (c)

    ee has only one descendant e1e_{1}, in which case there must exist e2e_{2} and e3e_{3} adjacent to their common point such that μ(e1)=μ(e)+μ(e2)\mu(e_{1})=\mu(e)+\mu(e_{2}). This is excluded by (3).

All three possibilities being excluded, we conclude that S1=SS_{1}=S.

Finally, we verify properties (1)–(3). In case (1), the edges e1,e2,e_{1}^{*},e_{2}^{*}, and e3e_{3}^{*} form an equilateral triangle. Suppose that ν𝒩\nu\in\mathcal{N} has an edge that intersects e3e_{3}^{*} in a single interior point. It is easy to see that in this case there must also be some edge of ν\nu that intersects either e1e_{1}^{*} or e2e_{2}^{*} in its interior, thus contradicting the flatness of these segments.

In cases (2) and (3), the edges e1e_{1}^{*} and e4e_{4}^{*} are the two parallel edges of an isosceles trapezoid and e2,e3e_{2}^{*},e_{3}^{*} are the other two edges. Suppose that ν𝒩\nu\in\mathcal{N} has an edge that intersects e3e_{3}^{*} in a single interior point. Then ν\nu must also have an edge that intersects either e1e_{1}^{*} or e2e_{2}^{*} in a single interior point, so e3S1e_{3}\in S_{1} if e1,e2S1e_{1},e_{2}\in S_{1}. Finally, suppose that ν𝒩\nu\in\mathcal{N} has an edge that intersects e4e_{4}^{*} in a single interior point. Then ν\nu must also have an edge that intersects e1e_{1}^{*} in a single interior point, so e4S1e_{4}\in S_{1} if e1S1e_{1}\in S_{1}. This concludes the proof. ∎

The last two results indicate an algorithm for checking whether a given pattern arises from the exit multiplicities of a rigid tree honeycomb. Thus, starting with a pattern, we construct flat segments and flat regions in stages, as follows. At stage 0, we have the collection of flat segments determined by the pattern on the boundary of the triangle. Suppose that the flat segments and flat regions of stage kk have been constructed, and let ABAB and CDCD be two nonparallel flat edges that appear in this configuration. Suppose that ABAB and CDCD lie on different sides of a 6060^{\circ} angle, and that the interiors of two equilateral triangles Δ1\Delta_{1} and Δ2\Delta_{2}, based on ABAB and CDCD respectively, intersect. Then we construct segments ABΔ1A^{\prime}B^{\prime}\subset\Delta_{1} parallel to ABAB with A,BΔ1A^{\prime},B^{\prime}\in\partial\Delta_{1} and CDΔ2C^{\prime}D^{\prime}\subset\Delta_{2} parallel to CDCD with C,DΔ2C^{\prime},D^{\prime}\subset\partial\Delta_{2} (see Figure 3.4)

Refer to captionAABBAABBAA^{\prime}BB^{\prime}AA^{\prime}BB^{\prime}DDDDCCCCCC^{\prime}CC^{\prime}
Figure 3.4. Flat regions, D=BD^{\prime}=B^{\prime}

that have one endpoint in common. Denote by Ω\Omega the smallest trapezoid (or equilateral triangle) with edges parallel to w1,w2,w_{1},w_{2,} or w3w_{3} that contains ABA^{\prime}B^{\prime} and CDC^{\prime}D^{\prime}. Then, in stage k+1k+1 we declare that Ω\Omega and all of its sides are flat. If the pattern is in fact the exit pattern of a rigid tree honeycomb, then the preceding result shows that this process ends in a finite number of stages and that the flat regions it creates (which are only equilateral triangles and trapezoids) tile the original triangle. Moreover, the shortest flat segments, corresponding to root edges of the rigid tree honeycomb, must have unit length.

In order to verify that a given rigid pattern, for which the preceding process yields a collection of flat triangles and trapezoids that tile the entire triangle Δ\Delta, really arises as the exit pattern of some rigid tree honeycomb μ\mu, we must construct μ\mu or some honeycomb homologous to it. One way to do this is to find a honeycomb ν\nu in \mathcal{M}_{*} whose support consists of the flat segments created in the process, construct μ\mu such that μ=ν\mu^{*}=\nu, and verify that it is in fact a rigid tree honeycomb. This may be a laborious process. Fortunately, there exists a simpler process that yields such as honeycomb μ\mu: each flat region being a triangle or a trapezoid, it has a circumcenter. The segment joining the circumcenters of two adjacent flat regions is part of the perpendicular bisector of the common edge. Add to these segments rays contained in the perpendicular bisector of each flat segment in Δ\partial\Delta terminating at the circumcenter of the adjacent flat region. Assign to each bisecting segment a multiplicity equal to the length of the original flat segment. This way, we obtain a honeycomb, rotated counterclockwise by 3030^{\circ}. It is fairly easy to check whether this is a rigid tree honeycomb since we know what its root edges are.

Refer to caption222211111111
Figure 3.5. Determining flat regions

Figure 3.5 illustrates the two above processes for the pattern 3,1|2,1,1|2,1,13,1|2,1,1|2,1,1. The flat regions are determined in two stages, they tile Δ\Delta, and the shortest flat edges have unit length. However, the (red) honeycomb constructed using circumcenters is rigid but not extreme because not all the root edges are equivalent. For the pattern 1,2,1|1,2,1|1,2,11,2,1|1,2,1|1,2,1 shown in Figure 3.6, all the flat regions are determined in stage 11 and they do not cover Δ\Delta, so this is not the exit pattern of a rigid tree honeycomb.

Refer to caption
Figure 3.6. Flat regions do not cover Δ\Delta

On the other hand, for the pattern 4,2|2,2,2|2,2,24,2|2,2,2|2,2,2 the flat regions do cover Δ\Delta but the shortest flat edge has length 22. In this case, the circumcenter construction yields twice a rigid tree honeycomb. There are locking patterns for which the flat regions created by the method described above overlap and thereby create greater flat regions that may not by triangles or trapezoids. Such an example is 3,2|1,3,2|2,2,13,2|1,3,2|2,2,1, illustrated in Figure 3.7.

Refer to caption
Figure 3.7. A rigid pattern

Suppose that μ\mu\in\mathcal{M} is a rigid tree honeycomb and its dual is written as in the above proof μ=j=1ncjρj\mu^{*}=\sum_{j=1}^{n}c_{j}\rho_{j}. If t1,,tn>0t_{1},\dots,t_{n}>0 and ν\nu\in\mathcal{M} is such that ν=j=1ntjρj\nu^{*}=\sum_{j=1}^{n}t_{j}\rho_{j} then μ\mu and ν\nu are homologous in the sense defined in [3]. In other words, up to a translation, ν\nu can be obtained by stretching or shrinking the edges of μ\mu while preserving their multiplicities (see Figure 3.8).

Refer to caption
Figure 3.8. Two homologous honeycombs with ω=6\omega=6

Conversely, if ν\nu is an arbitrary rigid tree honeycomb of weight ω(μ)\omega(\mu) that determines the same pattern on Δ\Delta_{*}, then ν\nu must satisfy ν=j=1ntjρj\nu^{*}=\sum_{j=1}^{n}t_{j}\rho_{j} for some t1,,tn>0t_{1},\dots,t_{n}>0. Thus, in order to classify the rigid tree honeycombs of a given weight ω\omega, it suffices to find the locking patterns on a triangle Δ\Delta with sides equal to ω\omega that also correspond to rigid tree honeycombs. There is only a finite number of such patterns because the lengths of the segments in the boundary of Δ\Delta determined by the relevant patterns are exit multiplicities of a tree honeycomb and are therefore integers.

Let μ\mu\in\mathcal{M} be a rigid tree honeycomb of weight ω\omega. According to (2.8), the integers μ(j)(c)\mu^{(j)}(c), j=1,2,3,cj=1,2,3,c\in\mathbb{R}, are subject to the following requirements:

cμ(j)(c)\displaystyle\sum_{c\in\mathbb{R}}\mu^{(j)}(c) =ω,j=1,2,3,\displaystyle=\omega,\quad j=1,2,3,
c[μ(1)(c)2+μ(2)(c)2+μ(3)(c)2]\displaystyle\sum_{c\in\mathbb{R}}[\mu^{(1)}(c)^{2}+\mu^{(2)}(c)^{2}+\mu^{(3)}(c)^{2}] =ω2+2.\displaystyle=\omega^{2}+2.

If we arrange the nonzero exit multiplicities of μ\mu in a nonincreasing list α1,,αN\alpha_{1},\dots,\alpha_{N}, we have

(3.1) j=1Nαj=3ω,j=1Nαj2=ω2+2.\sum_{j=1}^{N}\alpha_{j}=3\omega,\quad\sum_{j=1}^{N}\alpha_{j}^{2}=\omega^{2}+2.

This system of Diophantine equations has finitely many solutions that are fairly easy to list. We observe that subtracting the first equation above from the second we obtain

(3.2) j=1Nαj(αj1)2=(ω1)(ω2)2.\sum_{j=1}^{N}\frac{\alpha_{j}(\alpha_{j}-1)}{2}=\frac{(\omega-1)(\omega-2)}{2}.

In other words, we need to write the triangular number (ω1)(ω2)/2(\omega-1)(\omega-2)/2 as a sum of triangular numbers, and this sum will determine precisely the values of those αj\alpha_{j} that are greater than 11. Once numbers αj\alpha_{j} have been determined, there may be several ways that they can be arranged to make a pattern on a triangle with sides ω\omega. For each of these arrangements, we can use the algorithm outlined above to determine whether it is the exit pattern of a rigid tree honeycomb. We illustrate this for small values of ω\omega. For ω2\omega\leq 2 we have (ω1)(ω2)/2=0(\omega-1)(\omega-2)/2=0 and thus αj=1\alpha_{j}=1 for all jj. The corresponding types of rigid tree honeycombs are pictured in Figure 3.9(A) and (B).

Refer to caption
(a) ω=1\omega=1
Refer to caption
(b) ω=2\omega=2
Refer to caption
(c) ω=3\omega=3
Figure 3.9. Rigid tree honeycombs with small weight

For ω=3\omega=3 we have (ω1)(ω2)/2=1(\omega-1)(\omega-2)/2=1 so the only numbers αj\alpha_{j} satisfying (3.1) are 2,1,1,1,1,1,1,12,1,1,1,1,1,1,1. These lengths can be arranged around a triangle Δ\Delta in six different ways but all these arrangements can be obtained from one of them by rotations or reflections. It is seen immediately that these arrangements are in fact the exit patterns of rigid tree honeycombs. Thus, there are exactly six types of rigid tree honeycombs of weight 3, all of which can be obtained by rotations and reflections from the one pictured in Figure 3.9(C) (the thicker edge has multiplicity 2).

For ω=4\omega=4, we have (ω1)(ω2)/2=3(\omega-1)(\omega-2)/2=3, and this can be written as 3=33=3 and 3=1+1+13=1+1+1. The corresponding solutions αj\alpha_{j} are

3,1,1,1,1,1,1,1,1,1,\displaystyle 3,1,1,1,1,1,1,1,1,1, and
2,2,2,1,1,1,1,1,1.\displaystyle 2,2,2,1,1,1,1,1,1.

We find that there are 4444 distinct types of rigid tree honeycombs, all of which can be obtained from one of the eight depicted in

Refer to caption
Figure 3.10. Rigid tree honeycombs with ω=4\omega=4

Figure 3.10 by rotations and reflections. Each of the types pictured, except for the first one, yields six different types via rotations and reflections. The first one is invariant under rotations by 120120^{\circ}. For larger values of ω\omega, the number of types of rigid tree honeycombs grows fairly rapidly. For instance, we get 272 such types for ω=5\omega=5 and more than 1000 for ω=6\omega=6. Rigid tree honeycombs with a rotational symmetry exist for every weight ω\omega that is not a multiple of 33. Some examples of such honeycombs are provided in [3] but for large ω\omega there are other examples. To find them we denote by β1βk\beta_{1}\geq\dots\geq\beta_{k} the possible exit multiplicities of such a honeycomb in the direction of w1w_{1}, and note that these numbers satisfy

j=1kβj=ω,j=1kβj2=ω2+23.\sum_{j=1}^{k}\beta_{j}=\omega,\quad\sum_{j=1}^{k}\beta_{j}^{2}=\frac{\omega^{2}+2}{3}.

These equations combine to yield

j=1kβj(βj1)2=(ω1)(ω2)6.\sum_{j=1}^{k}\frac{\beta_{j}(\beta_{j}-1)}{2}=\frac{(\omega-1)(\omega-2)}{6}.

For instance, if ω=8\omega=8, we must write (ω1)(ω2)/6=7(\omega-1)(\omega-2)/6=7 as a sum of triangular numbers. The acceptable possibilities are

6+1,3+3+1,6+1,3+3+1,

and they correspond to the following sequences of βj\beta_{j}:

4,2,1,1\displaystyle 4,2,1,1
3,3,2.\displaystyle 3,3,2.

(The representation 7=1+1+1+1+1+1+17=1+1+1+1+1+1+1 requires that β1==β7=2\beta_{1}=\cdots=\beta_{7}=2 but 7×2>87\times 2>8. Similarly, 7=3+1+1+1+17=3+1+1+1+1 requires β1=3\beta_{1}=3 and β2==β5=2\beta_{2}=\cdots=\beta_{5}=2.) We obtain, up to symmetry, the 4 distinct rigid tree honeycomb types depicted in Figure 3.11.

Refer to caption
Figure 3.11. Four honeycombs with ω=8\omega=8

4. Honeycombs compatible with a puzzle

Suppose that we have the puzzle of a honeycomb ν\nu\in\mathcal{M} and, in addition, a honeycomb μ\mu\in\mathcal{M} superposed on it. We say that μ\mu is compatible with the puzzle of ν\nu if the dark gray parallelogram contain no branch points of μ\mu in their interior and the edges of μ\mu only cross such a parallelogram along segments parallel to an edge of the parallelogram. This concept appeared in [9] where it was used to explain saturated Horn inequalities. It was also employed in [3, Section 7] to give a different argument for some factorizations of Littlewood-Richardson coefficients first found in [7, 6]. Here we are mostly interested in the existence and construction of such compatible honeycombs μ\mu, particularly rigid tree honeycombs. It is useful to observe that a tree honeycomb that is compatible with the puzzle of ν\nu crosses the gray parallelograms along segments that are either all parallel to the white edges or all parallel to the light gray edges (see

Refer to caption
Figure 4.1. ν\nu, its puzzle, a compatible μ\mu, and μν\mu_{\nu}

Figure 4.1). Deflating the puzzle of ν\nu yields, in addition to the original honeycomb ν\nu, a honeycomb μν\mu_{\nu} constructed as follows:

  1. (1)

    the parts of the support of μ\mu inside the white puzzle pieces are translated along with those white pieces,

  2. (2)

    the parts of the support of μ\mu that cross dark gray parallelograms between white puzzle pieces are discarded, and

  3. (3)

    if ABAB is an edge of ν\nu and the support of μ\mu crosses the inflation of ABAB on segments II parallel to ABAB, then ABAB is also contained in the support of μν\mu_{\nu} and its multiplicity is the sum of the multiplicities of all such segments II.

A similar construction, using the *-deflation of the puzzle of ν\nu, yields a honeycomb μν\mu_{\nu^{*}} with ω(μν)=ω(μν)=ω(μ)\omega(\mu_{\nu^{*}})=\omega(\mu_{\nu})=\omega(\mu). Moreover, it was shown in [3] that

(4.1) μ=(μν)+(μν).\mu^{*}=(\mu_{\nu})^{*}+(\mu_{\nu^{*}})^{*}.

Our inductive procedure for constructing honeycombs from rigid overlays depends on the construction of honeycombs that are compatible with the puzzle of a rigid honeycomb. We begin with compatible honeycombs whose support does not intersect the interior of any puzzle piece. Such a support must be contained in the edges of the dark gray parallelograms. Thus, suppose that ν\nu\in\mathcal{M} is a rigid honeycomb and regard the union of the edges of the dark gray parallelograms in its puzzle as a prospective support for a honeycomb. For simplicity, we simply call puzzle edges the edges of the dark gray parallelogram. These edges are either white or light gray, according to the color of the other neighboring puzzle piece.

Lemma 4.1.

Let ν\nu be a honeycomb in \mathcal{M}. Every evil loop contained in the puzzle edges of ν\nu is a gentle loop, that is, consecutive edges in the loop form 120120^{\circ} angles.

Proof.

Since all the branch points in this support are ‘rakes’ (as in the second part of Figure 3.3), it follows that every evil turn of an evil loop is either a gentle turn (of 120120^{\circ}) or a 180180^{\circ} turn along the ‘handle’ of a rake. Suppose that an evil loop e1,,en=e1e_{1},\dots,e_{n}=e_{1} contains at least one 180180^{\circ} turn. We may assume that e1=ABe_{1}=AB and e2=BAe_{2}=BA, in which case it follows that the edges e2,,eke_{2},\dots,e_{k} point to the acute angles of the adjacent parallelograms until we meet again a 180180^{\circ} turn ekek+1e_{k}e_{k+1} (see Figure 4.2).

Refer to captionAAAAAAAABBBBBBBB
Figure 4.2. Evil turns in a rigid puzzle

Continuing this way around the loop, we conclude that e1e_{1} also points to the acute angle of the adjacent parallelogram, an obvious contradiction. ∎

According to [9, Theorems 4 and 7], the puzzle of a rigid honeycomb ν\nu contains no gentle (and, by Lemma 4.1, no evil) loops. It follows that any honeycomb supported by the puzzle edges of ν\nu is necessarily rigid. Suppose for a moment that there exists a honeycomb μ\mu whose support is the union of all the edges of the puzzle of ν\nu. The root edges of such a honeycomb are easily identified. For each ray ee in the support of ν\nu there is one equivalence class of root edges containing the ray in the inflation

Refer to caption
Figure 4.3.

of ee incident to the acute angle of that inflation(see Figure 4.3 for an example with these root edges colored red). We now orient all edges towards the acute angles of the parallelogram to which they belong (or away from the obtuse angles in the case of some of the rays). (Note that the opposite orientation was used in [9, Section 4] for different purposes.) Thus, if ff is a ray in the support of ν\nu, the boundary of its inflation has two edges that are rays: an incoming ray pointing to the acute angle of the inflation, and an outgoing ray. This orientation is useful because, except for the root edges, it is precisely the orientation of the descendant relation, that is, each edge (with the exception of the root edges) is oriented from an ancestor to a descendant. The resulting oriented graph contains no loops and, in addition, for each edge ee there is at least one path from one of the root edges to ee. It is easy now to see that there actually exists a rigid tree honeycomb rooted at each of the root edges of the hypothetical honeycomb μ\mu, and that the union of the supports of these honeycombs contains all the edges of the puzzle. We summarize these observations in Theorem 4.2. The last assertion in the statement follows by observing that none of the events described in Remark 2.4 is possible for the honeycombs μ1\mu_{1} and μ2\mu_{2}. Figure 4.3 shows (in red) the support of one of the honeycombs constructed in the manner just described.

Theorem 4.2.

Let ν\nu\in\mathcal{M} be a rigid honeycomb. Then there are exactly exit(ν){\rm exit}(\nu) rigid tree honeycombs whose support is contained in the union of the puzzles edges of ν\nu. The root of the honeycomb corresponding to a ray ee of ν\nu is the incoming ray in the inflation of ee. Every edge of the puzzle of ν\nu is contained in the support of one of these tree honeycombs. If μ1\mu_{1} and μ2\mu_{2} are two of these tree honeycombs, then Σμ1(μ2)=0\Sigma_{\mu_{1}}(\mu_{2})=0.

The following observation about arbitrary compatible honeycombs is useful in Section 6.

Lemma 4.3.

Let ν\nu\in\mathcal{M} be a rigid honeycomb and let μ\mu\in\mathcal{M} be a honeycomb that is compatible with the puzzle of ν\nu. Suppose that the support of μ\mu contains an edge ee of the puzzle. Then the support of μ\mu contains every descendant of ee relative to the orientation of the puzzle edges of ν\nu.

Proof.

It suffices to show that, given two edges e1e_{1} and e2e_{2} in the puzzle of ν\nu such that e1e2e_{1}e_{2} is a gentle path and e1e_{1} is contained in the support of μ\mu, it follows that e2e_{2} is also contained in the support of μ\mu. The three possible configurations, up to rotations and reflections, are shown in Figure

Refer to captione1e_{1}e1e_{1}e1e_{1}e2e_{2}e2e_{2}e2e_{2}
Figure 4.4.

Edges of a puzzle in the support of a compatible honeycomb

4.4. We observe first that at least part of e2e_{2} must be contained in the support of μ\mu because otherwise the support of μ\mu would contain one edge crossing a dark gray parallelogram in a direction that is not allowed by compatibility. Similarly, if Ie2I\subset e_{2} is the largest segment in the support of μ\mu that contains e1e2e_{1}\cap e_{2}, then II must equal e2e_{2} because otherwise the endpoint of II (other than e1e2)e_{1}\cap e_{2}) would need to be a branch point for μ\mu with one edge of μ\mu crossing the neighboring dark gray parallelogram in the direction forbidden by compatibility. ∎

It is of interest to calculate the exit pattern of the tree honeycombs described in Theorem 4.2, and we do eventually provide an answer amounting to an inductive procedure depending on the number of extreme summands of the honeycomb ν\nu (see Remark 4.15). We start with the special case of an extreme rigid honeycomb. The direction of the rays corresponding to the multiplicities αj\alpha_{j} are not specified in the following statement, but the multiplicity α1αj\alpha_{1}\alpha_{j} corresponds to the outgoing ray in the inflation of the ray with multiplicity αj\alpha_{j}, while the multiplicities α121\alpha_{1}^{2}-1 and 11 correspond to outgoing and incoming rays, respectively, in the inflation of the ray with multiplicity α1\alpha_{1}.

Proposition 4.4.

Suppose that ν\nu is a rigid tree honeycomb and that the nonzero exit multiplicities of ν\nu are, in counterclockwise order, α1,,αn\alpha_{1},\dots,\alpha_{n}. Let μ\mu be the rigid tree honeycomb, supported in the union of the puzzle edges of ν\nu, and rooted in the incoming ray in the inflation of the ray with multiplicity α1\alpha_{1}. Then the nonzero exit multiplicities of μ\mu are, in counterclockwise order,

α121,1,α1α2,,α1αn,\alpha_{1}^{2}-1,1,\alpha_{1}\alpha_{2},\dots,\alpha_{1}\alpha_{n},

where the first term is omitted if α1=1\alpha_{1}=1. In particular, ω(μ)=α1ω(ν)\omega(\mu)=\alpha_{1}\omega(\nu),

(4.2) exit(μ)=exit(ν)+1{\rm exit}(\mu)={\rm exit}(\nu)+1

if α1>1\alpha_{1}>1, and μ\mu is homologous to ν\nu if α1=1\alpha_{1}=1.

Proof.

Suppose that the exit multiplicities of μ\mu are β1,1,β2,,βn\beta_{1},1,\beta_{2},\dots,\beta_{n}, where βj\beta_{j} is the multiplicity of the outgoing ray corresponding to αj\alpha_{j}. The presence of 11 is explained by the choice of root edge which happens to be an incoming ray. Deflate now the puzzle of ν\nu and obtain thereby a honeycomb μν\mu_{\nu} with exit multiplicities β1+1,β2,,βn.\beta_{1}+1,\beta_{2},\dots,\beta_{n}. The honeycomb μν\mu_{\nu} has support contained in that of ν\nu and is therefore of the form kνk\nu for some positive integer kk. We deduce that the exit multiplicities of μ\mu are kα11,1,kα2,,kαnk\alpha_{1}-1,1,k\alpha_{2},\dots,k\alpha_{n} and ω(μ)=kω(ν).\omega(\mu)=k\omega(\nu). Thus,

j=1nαj2\displaystyle\sum_{j=1}^{n}\alpha_{j}^{2} =ω(ν)2+2,\displaystyle=\omega(\nu)^{2}+2,
(kα11)2+1+j=2n(kαj)2\displaystyle(k\alpha_{1}-1)^{2}+1+\sum_{j=2}^{n}(k\alpha_{j})^{2} =(kω(ν))2+2,\displaystyle=(k\omega(\nu))^{2}+2,

and subtracting k2k^{2} times the first equation from the second we obtain

2kα1+2=2k2+2.-2k\alpha_{1}+2=-2k^{2}+2.

The only positive solution of this equation is k=α1k=\alpha_{1}. ∎

Corollary 4.5.

Suppose that ν\nu is a rigid tree honeycomb and that the nonzero exit multiplicities of ν\nu are, in counterclockwise order, α1,,αn\alpha_{1},\dots,\alpha_{n}. There exists a rigid tree honeycomb μ\mu whose nonzero exit multiplicities are, in counterclockwise order,

1,α121,α1α2,,α1αn,1,\alpha_{1}^{2}-1,\alpha_{1}\alpha_{2},\dots,\alpha_{1}\alpha_{n},

where the second term is omitted if α1=1\alpha_{1}=1. In particular, ω(μ)=α1ω(ν)\omega(\mu)=\alpha_{1}\omega(\nu).

Proof.

Let ν\nu^{\prime} be a honeycomb obtained by reflecting ν\nu in a line parallel to w1w_{1}. Thus the nonzero exit multiplicities of ν\nu^{\prime} can be listed, in counterclockwise order, as α1,αn,αn1,,α2\alpha_{1},\alpha_{n},\alpha_{n-1},\dots,\alpha_{2}. Proposition 4.4, applied to ν\nu^{\prime}, shows that there exists a rigid tree honeycomb μ\mu^{\prime} whose nonzero exit multiplicities are, in counterclockwise order,

α121,1,α1αn,α1αn1,,α1α2.\alpha_{1}^{2}-1,1,\alpha_{1}\alpha_{n},\alpha_{1}\alpha_{n-1},\dots,\alpha_{1}\alpha_{2}.

The honeycomb μ\mu is now obtained by reflecting μ\mu^{\prime} in a line parallel to w1w_{1}. ∎

Remark 4.6.

An alternate proof of the preceding corollary uses the *-inflation of μ\mu.

We next construct compatible honeycombs starting with a rigid overlay of two rigid tree honeycombs. The following result was proved in [3, Theorem 5.2].

Theorem 4.7.

Suppose that ν1\nu_{1} and ν2\nu_{2} are rigid tree honeycombs and Σν2(ν1)=0\Sigma_{\nu_{2}}(\nu_{1})=0. Then there exists a honeycomb ν~2\widetilde{\nu}_{2}, homologous to ν2\nu_{2} and compatible with the puzzle of ν1\nu_{1}, such that the parts of the support of ν~2\widetilde{\nu}_{2} contained in the white puzzle pieces are simply translates of the corresponding parts of the support of ν2\nu_{2}.

We recall briefly the construction of ν~2\widetilde{\nu}_{2}. Choose a root edge ee of ν2\nu_{2} that is not contained in the support of ν1\nu_{1} and orient all the edges of ν2\nu_{2} away from ee (that is, in the direction of the descendance paths from ee. If a portion ff of an edge of ν2\nu_{2} is contained in an edge of ν1\nu_{1}, attach ff to the white puzzle piece on its right relative to this orientation. This way we obtain the part of the support of ν~2\widetilde{\nu}_{2} that is contained in the boundary of a white puzzle piece. The part of the support of ν~2\widetilde{\nu}_{2} in the interior of the white pieces is obtained simply by translation, as in the statement above. Finally, one reconnects the edges of ν~2\widetilde{\nu}_{2} that are transversal to the support of ν1\nu_{1} and this is done by adding a number of segments that cross dark gray parallelograms and are parallel to their light gray sides. The requirement that Σν2(ν1)=0\Sigma_{\nu_{2}}(\nu_{1})=0 makes this construction possible. This process is illustrated in Figure 4.5 where the supports of ν2\nu_{2} and ν~2\widetilde{\nu}_{2} are drawn with dashed lines. (An interesting feature of the overlay in Figure 4.5 is that ν1\nu_{1} and ν2\nu_{2} assign nonzero multiplicity to precisely the same rays.)

Refer to caption
Figure 4.5. The honeycombs ν1\nu_{1}, ν2\nu_{2}, and ν~2\widetilde{\nu}_{2}
Proposition 4.8.

Suppose that ν1\nu_{1} and ν2\nu_{2} are rigid tree honeycombs such that Σν2(ν1)=0\Sigma_{\nu_{2}}(\nu_{1})=0, and let ff be a ray of ν1\nu_{1} to which ν1\nu_{1} assigns exit multiplicity α1\alpha_{1}. Let α20\alpha_{2}\geq 0 be the exit multiplicity that ν2\nu_{2} assigns to the same ray. Denote by ν~1\widetilde{\nu}_{1} the rigid tree honeycomb whose support is contained in the union of the puzzle edges of ν1\nu_{1} and with one root edge equal to the incoming ray of the inflation of ff. Finally, let ν~2\widetilde{\nu}_{2} be the rigid tree honeycomb, homologous to ν2\nu_{2}, constructed in Theorem 4.7. Then Σν~2(ν~1)=0\Sigma_{\widetilde{\nu}_{2}}(\widetilde{\nu}_{1})=0 and Σν~1(ν~2)=α2+α1Σν1(ν2)\Sigma_{\widetilde{\nu}_{1}}(\widetilde{\nu}_{2})=\alpha_{2}+\alpha_{1}\Sigma_{\nu_{1}}(\nu_{2}).

Proof.

Suppose that ν2(j)(c)ν1(j)(d)0\nu_{2}^{(j)}(c)\nu_{1}^{(j)}(d)\neq 0 for some j=1,2,3j=1,2,3 and some c<dc<d. Unless the ray ff is on the line xj=dx_{j}=d, there exist numbers c<dc^{\prime}<d^{\prime} such that ν~2(j)(c)=ν2(j)(c)\widetilde{\nu}_{2}^{(j)}(c^{\prime})=\nu_{2}^{(j)}(c) and ν~1(j)(d)=α1ν1(j)(d)\widetilde{\nu}_{1}^{(j)}(d^{\prime})=\alpha_{1}\nu_{1}^{(j)}(d). If the ray ff is on the line xj=dx_{j}=d, we find c<d<d′′c^{\prime}<d^{\prime}<d^{\prime\prime} such that ν~2(j)(c)=ν2(j)(c)\widetilde{\nu}_{2}^{(j)}(c^{\prime})=\nu_{2}^{(j)}(c), ν~1(j)(d)=α121\widetilde{\nu}_{1}^{(j)}(d^{\prime})=\alpha_{1}^{2}-1, and ν~1(j)(d′′)=1\widetilde{\nu}_{1}^{(j)}(d^{\prime\prime})=1. This describes all the pairs c<dc^{\prime}<d^{\prime} (or c<d′′c^{\prime}<d^{\prime\prime}) such that ν~2(j)(c)ν~1(j)(d)0\widetilde{\nu}_{2}^{(j)}(c^{\prime})\widetilde{\nu}_{1}^{(j)}(d^{\prime})\neq 0 with one exception arising from the ray ff that corresponds to ν1(j)(c)=α1\nu_{1}^{(j)}(c)=\alpha_{1} and ν2(j)(c)=α2\nu_{2}^{(j)}(c)=\alpha_{2}. In this case, we obtain d<d′′d^{\prime}<d^{\prime\prime} (corresponding to the outgoing and incoming rays in the inflation of ff) such that ν~2(j)(d)=α2\widetilde{\nu}_{2}^{(j)}(d^{\prime})=\alpha_{2} and ν~1(j)(d′′)=1\widetilde{\nu}_{1}^{(j)}(d^{\prime\prime})=1. We conclude that

c<dν~2(j)(c)ν~1(j)(d)\displaystyle\sum_{c^{\prime}<d^{\prime}}\widetilde{\nu}_{2}^{(j)}(c^{\prime})\widetilde{\nu}_{1}^{(j)}(d^{\prime}) =α2+α1c<dν2(j)(c)ν1(j)(d)\displaystyle=\alpha_{2}+\alpha_{1}\sum_{c<d}\nu_{2}^{(j)}(c)\nu_{1}^{(j)}(d)
=α2+α1Σν2(ν1)+α1ω(ν1)ω(ν2),\displaystyle=\alpha_{2}+\alpha_{1}\Sigma_{\nu_{2}}(\nu_{1})+\alpha_{1}\omega(\nu_{1})\omega(\nu_{2}),

and thus

Σν~1(ν~2)\displaystyle\Sigma_{\widetilde{\nu}_{1}}(\widetilde{\nu}_{2}) =α2+α1Σν1(ν2)+α1ω(ν1)ω(ν2)ω(ν~1)ω(ν~2)\displaystyle=\alpha_{2}+\alpha_{1}\Sigma_{\nu_{1}}(\nu_{2})+\alpha_{1}\omega(\nu_{1})\omega(\nu_{2})-\omega(\widetilde{\nu}_{1})\omega(\widetilde{\nu}_{2})
=α2+α1Σν1(ν2)\displaystyle=\alpha_{2}+\alpha_{1}\Sigma_{\nu_{1}}(\nu_{2})

because ω(ν~1)=α1ω(ν1)\omega(\widetilde{\nu}_{1})=\alpha_{1}\omega(\nu_{1}) and ω(ν~2)=ω(ν2)\omega(\widetilde{\nu}_{2})=\omega(\nu_{2}). The calculation of Σν~2(ν~1)\Sigma_{\widetilde{\nu}_{2}}(\widetilde{\nu}_{1}) is somewhat simpler because there is no extra pair c>dc^{\prime}>d^{\prime} such that ν~1(j)(c)ν~2(j)(d)0\widetilde{\nu}_{1}^{(j)}(c^{\prime})\widetilde{\nu}_{2}^{(j)}(d^{\prime})\neq 0. ∎

Remark 4.9.

The preceding proposition can also be proved using the calculation of Σ\Sigma described in Remark 2.4 though the extra α2\alpha_{2} crossings needed for Σν~1(ν~2)\Sigma_{\widetilde{\nu}_{1}}(\widetilde{\nu}_{2}) may be difficult to locate.

Remark 4.10.

Suppose that ν1,ν2,\nu_{1},\nu_{2}, and ν3\nu_{3} are three rigid tree honeycombs such that Σν2(ν1)=Σν3(ν1)=0\Sigma_{\nu_{2}}(\nu_{1})=\Sigma_{\nu_{3}}(\nu_{1})=0. Let ν~1\widetilde{\nu}_{1} and ν~2\widetilde{\nu}_{2} be as in Proposition 4.8, and construct a third honeycomb ν~3\widetilde{\nu}_{3} that is homologous to ν3\nu_{3} and compatible with the puzzle of ν1\nu_{1} (the same way that ν~2\widetilde{\nu}_{2} was constructed). Then the entire overlay ν~2+ν~3\widetilde{\nu}_{2}+\widetilde{\nu}_{3} is homologous to ν2+ν3,\nu_{2}+\nu_{3}, and therefore

(4.3) Σν~2(ν~3)=Σν2(ν3) and Σν~3(ν~2)=Σν3(ν2).\Sigma_{\widetilde{\nu}_{2}}(\widetilde{\nu}_{3})=\Sigma_{\nu_{2}}(\nu_{3})\text{ and }\Sigma_{\widetilde{\nu}_{3}}(\widetilde{\nu}_{2})=\Sigma_{\nu_{3}}(\nu_{2}).

There is a second way to construct a rigid tree honeycomb from a rigid overlay of two tree honeycombs. We show in Theorem 5.1 that every rigid tree honeycomb with weight at least 22 can be obtained from this construction applied to an overlay of honeycombs with smaller weights.

Theorem 4.11.

Suppose that ν1\nu_{1} and ν2\nu_{2} are rigid tree honeycombs, Σν1(ν2)=σ>0\Sigma_{\nu_{1}}(\nu_{2})=\sigma>0 and Σν2(ν1)=0\Sigma_{\nu_{2}}(\nu_{1})=0. Then there exists a rigid tree honeycomb ν^1\widehat{\nu}_{1} compatible with the puzzle of ν2\nu_{2}, such that the parts of the support of ν^1\widehat{\nu}_{1} contained in the white puzzle pieces are simply translates of the corresponding parts of the support of ν1\nu_{1}. Moreover, the exit pattern of ν^1\widehat{\nu}_{1} is the same as the exit pattern of ν1+σν2\nu_{1}+\sigma\nu_{2}.

Proof.

Fix a root edge ee of ν1\nu_{1} that is disjoint from the support of ν2\nu_{2} and orient all the other edges of μ1\mu_{1} away from ee, that is, in the direction given by descendance paths. We construct a collection 𝒞\mathcal{C} of segments and rays containing:

  1. (1)

    the edges of the puzzle of ν2\nu_{2},

  2. (2)

    the segments obtained as translates of segments in the support of ν1\nu_{1} that are contained in the interior of a white puzzle piece, and

  3. (3)

    the inflation of every branch point of ν1+ν2\nu_{1}+\nu_{2} that is in the interior of a side of a white puzzle piece. That is, if ZZ is a branch point in the interior of an edge ABAB of ν2\nu_{2}, and if AB,A′′B′′A^{\prime}B^{\prime},A^{\prime\prime}B^{\prime\prime} are the two white edges of the inflation of ABAB, then we include in our collection the segment ZZ′′Z^{\prime}Z^{\prime\prime} joining the points on ABA^{\prime}B^{\prime} and A′′B′′A^{\prime\prime}B^{\prime\prime} that are the translates of ZZ.

We orient all the segments in 𝒞\mathcal{C}, with the exception of the translate of the root edge ee, as follows:

  1. (1)

    the edges of every dark gray parallelogram point to the acute angles of the parallelogram,

  2. (2)

    the translate of a (part of) an edge ff of ν1\nu_{1} other than ee is pointing in the same direction as ff,

  3. (3)

    the segments ZZ′′Z^{\prime}Z^{\prime\prime} described above are oriented as in Figure 4.6 where the six possible configurations (up to rotations) are shown.

    Refer to captionZZZZZZZZZZZZZZ^{\prime}ZZ^{\prime}ZZ^{\prime}ZZ^{\prime}ZZ^{\prime}ZZ^{\prime}Z′′Z^{\prime\prime}Z′′Z^{\prime\prime}Z′′Z^{\prime\prime}Z′′Z^{\prime\prime}Z′′Z^{\prime\prime}Z′′Z^{\prime\prime}
    Figure 4.6. Orientation of ZZ′′Z^{\prime}Z^{\prime\prime}

    (Here, the horizontal edges are in the support of ν2\nu_{2} and are represented by dashed lines. The support of ν1\nu_{1} is represented by dotted lines and the solid lines represent common parts of the two supports. No other configurations are possible because Σν2(ν1)=0\Sigma_{\nu_{2}}(\nu_{1})=0; see Remark 2.4.)

We treat 𝒞\mathcal{C} as the prospective support of a honeycomb and, accordingly, we say that a point where three or more of these segments meet is a branch point. Denote by e=xye^{\prime}=xy the translate of ee and construct (as in Section 3) a tree T0T_{0} by following the descendance paths in 𝒞\mathcal{C}. Thus, the vertices of T0T_{0} are x,yx,y, and the gentle paths ee1ene^{\prime}e_{1}\dots e_{n}, where ej𝒞,e_{j}\in\mathcal{C}, e1e_{1} points away from xx or from yy, and eje_{j} points away from ej1e_{j-1} for j=2,,nj=2,\dots,n. If ee1ene^{\prime}e_{1}\dots e_{n} is such a path, then it is joined to ee1en1e^{\prime}e_{1}\dots e_{n-1} by an edge labeled jj if ene_{n} is parallel to wjw_{j}. In addition, xx is joined to yy by an edge labeled jj if ee^{\prime} is parallel to wjw_{j}. Finally, either xx or yy is joined to e1e_{1} by an edge labeled using the same rule.

We first argue that the tree T0T_{0} is finite. In the contrary case, there would exist an infinite gentle oriented path. Because 𝒞\mathcal{C} is finite, some edge of 𝒞\mathcal{C} must appear twice in this path, and we conclude that there exists an oriented gentle loop e1ene_{1}\dots e_{n} formed by edges in 𝒞\mathcal{C}. Some of the edges eje_{j}, but no two consecutive ones, may be of the form ZZ′′Z^{\prime}Z^{\prime\prime} as in the case (3) above. Each other eje_{j} is the translate of some edge fjf_{j} of ν1+ν2\nu_{1}+\nu_{2}. These edges fjf_{j} form a loop in the support of ν1+ν2.\nu_{1}+\nu_{2}. We claim that this loop is evil, thus contradicting the rigidity of ν1+ν2\nu_{1}+\nu_{2}. This is verified by considering any two, three, or four consecutive edges that form an oriented gentle path in 𝒞\mathcal{C}. Such a sequence of edges arises from a branch point of ν1+ν2\nu_{1}+\nu_{2} and the relevant evil turns are observed by examining the situations depicted in Figures 4.6 and 4.7.

Refer to caption
Figure 4.7.

All other possibilities for the branch points of ν1+ν2\nu_{1}+\nu_{2} are precluded by the condition Σν2(ν1)=0\Sigma_{\nu_{2}}(\nu_{1})=0.

We also note that ends of the tree T0T_{0} are precisely those paths ee1ene^{\prime}e_{1}\dots e_{n} such that ene_{n} is either an outgoing ray in the puzzle of ν2\nu_{2} or the translate of a ray of ν1\nu_{1}. We define now multiplicities ν^1(e)\widehat{\nu}_{1}(e) for each e𝒞e\in\mathcal{C} by setting ν^1(e)=1\widehat{\nu}_{1}(e^{\prime})=1 and, for other edges, ν^1(e)\widehat{\nu}_{1}(e) is the number of gentle descendance paths ee1ene^{\prime}e_{1}\dots e_{n} satisfying en=ee_{n}=e.

Define now a honeycomb ν^1\widehat{\nu}_{1} by setting ν^1(e)=1\widehat{\nu}_{1}(e^{\prime})=1 and, given an arbitrary edge ff in the collection 𝒞\mathcal{C}, ν^1(f)\widehat{\nu}_{1}(f) is the number of distinct oriented descendance paths from ee^{\prime} to ff. The balance condition at each branch point is checked by looking at the various cases from Figures 4.6 and 4.7. It is clear that ν^1\widehat{\nu}_{1} is a tree honeycomb, and Theorem 2.2 implies that it is rigid. Since the edges in the support of ν^1\widehat{\nu}_{1} are either translates of the original edges in the support of ν1+ν2\nu_{1}+\nu_{2} or segments of the form ZZ′′Z^{\prime}Z^{\prime\prime}, it follows that the support of the honeycomb (ν^1)ν2(\widehat{\nu}_{1})_{\nu_{2}} (obtained by deflating the puzzle of ν2\nu_{2}) is contained in the support of ν1+ν2\nu_{1}+\nu_{2}. Thus (ν^1)ν2(\widehat{\nu}_{1})_{\nu_{2}} is a rigid honeycomb of the form c1ν1+c2ν2c_{1}\nu_{1}+c_{2}\nu_{2} for some c1,c20c_{1},c_{2}\geq 0. Clearly, (ν^1)ν2(\widehat{\nu}_{1})_{\nu_{2}} assigns unit mass to ee^{\prime}, so c1=1c_{1}=1. To determine c2c_{2}, we note that, since (ν^1)ν2(\widehat{\nu}_{1})_{\nu_{2}} and ν1+c2ν2\nu_{1}+c_{2}\nu_{2} have same exit pattern, relation (2.7) shows that

1\displaystyle-1 =Σ(ν^1)ν2((ν^1)ν2)=Σν1+c2ν2(ν1+c2ν2)\displaystyle=\Sigma_{(\widehat{\nu}_{1})_{\nu_{2}}}((\widehat{\nu}_{1})_{\nu_{2}})=\Sigma_{\nu_{1}+c_{2}\nu_{2}}(\nu_{1}+c_{2}\nu_{2})
=Σν1(ν1)+c2Σν1(ν2)+c2Σν2(ν1)+c22Σν2(ν2)\displaystyle=\Sigma_{\nu_{1}}(\nu_{1})+c_{2}\Sigma_{\nu_{1}}(\nu_{2})+c_{2}\Sigma_{\nu_{2}}(\nu_{1})+c_{2}^{2}\Sigma_{\nu_{2}}(\nu_{2})
=1+0+c2σc22.\displaystyle=-1+0+c_{2}\sigma-c_{2}^{2}.

This equality is only satisfied by c2=0c_{2}=0 and c2=σc_{2}=\sigma. If σ>0\sigma>0, the alternative c2=0c_{2}=0 can be dismissed because in this case one of the first two configurations depicted in Figure 4.6 or the last configuration from Figure 4.7 occurs, thus ensuring that the honeycomb ν^1\widehat{\nu}_{1} assigns positive multiplicity to some white edges of the puzzle of ν2\nu_{2}, and so c2>0c_{2}>0. ∎

Corollary 4.12.

Suppose that ν1\nu_{1} and ν2\nu_{2} are rigid tree honeycombs such that Σν2(ν1)=0\Sigma_{\nu_{2}}(\nu_{1})=0, and set σ=Σν1(ν2)\sigma=\Sigma_{\nu_{1}}(\nu_{2}). Then there exist a rigid tree honeycomb μ1\mu_{1} (respectively, μ2\mu_{2}) that has the same exit pattern as ν1+σν2\nu_{1}+\sigma\nu_{2} (respectively, σν1+ν2\sigma\nu_{1}+\nu_{2}).

Proof.

The honeycomb μ1=ν^1\mu_{1}=\widehat{\nu}_{1} satisfies the requirement. To prove the existence of μ2\mu_{2}, we argue as in Corollary 4.5. Thus, we consider the reflections ν1\nu^{\prime}_{1} and ν2\nu^{\prime}_{2} of ν1\nu_{1} and ν2\nu_{2} in a line parallel to w1w_{1}. These rigid tree honeycombs satisfy Σν1(ν2)=0\Sigma_{\nu_{1}^{\prime}}(\nu_{2}^{\prime})=0 and Σν2(ν1)=σ\Sigma_{\nu_{2}^{\prime}}(\nu_{1}^{\prime})=\sigma, so there exists a rigid tree honeycomb μ2\mu_{2}^{\prime} with the same exit pattern as σν1+ν2\sigma\nu^{\prime}_{1}+\nu_{2}^{\prime}. We obtain μ2\mu_{2} as the reflection of μ2\mu_{2}^{\prime} in a line parallel to w1w_{1}. ∎

Once we know that a pattern comes from a rigid tree honeycomb μ\mu, one can find a honeycomb homologous to μ\mu using the flat region construction from Section 3. This allows for a fairly efficient construction of measures μ1\mu_{1} and μ2\mu_{2} with the exit patterns prescribed by Corollary 4.12.

Remark 4.13.

With the notation of Theorem 4.11, relation (4.1) applies to the honeycomb ν^1\widehat{\nu}_{1} to yield

(ν^1)\displaystyle(\widehat{\nu}_{1})^{*} =((ν^1)ν2)+((ν^1)ν2)\displaystyle=((\widehat{\nu}_{1})_{\nu_{2}})^{*}+((\widehat{\nu}_{1})_{\nu_{2}^{*}})^{*}
=(ν1+σν2)+((ν^1)ν2).\displaystyle=(\nu_{1}+\sigma\nu_{2})^{*}+((\widehat{\nu}_{1})_{\nu_{2}^{*}})^{*}.

The support of (ν^1)ν2(\widehat{\nu}_{1})_{\nu_{2}^{*}} inside the triangle Δ\Delta_{*} is always contained in the support of ν2\nu_{2}^{*}. In many cases, the support of (ν^1)ν2(\widehat{\nu}_{1})_{\nu_{2}^{*}} contains the entire support of ν2\nu_{2}^{*} (inside Δ\Delta_{*}). When this occurs, the honeycomb ((ν^1)ν2)((\widehat{\nu}_{1})_{\nu_{2}^{*}})^{*} is in fact a tree honeycomb with the same exit pattern as ν2\nu_{2}, so ((ν^1)ν2)((\widehat{\nu}_{1})_{\nu_{2}^{*}})^{*} is homologous (after a 6060^{\circ} rotation) to ν2\nu_{2}. Here is one particular way to form overlays with this special property. Given an arbitrary rigid tree honeycomb ν2\nu_{2}, find another rigid tree honeycomb ν1\nu_{1} such that ω(ν1)=3\omega(\nu_{1})=3, ν2\nu_{2} is clockwise from ν1\nu_{1}, and every exit ray of ν2\nu_{2} crosses an edge of ν1\nu_{1}. (See Figure 4.8 for a particular case. The honeycomb ν2\nu_{2} is pictured in black and ν2\nu_{2} in red. The second part of the figure shows the puzzle of ν2\nu_{2} and the support of ν^1\widehat{\nu}_{1}.) Since the support of ν^1\widehat{\nu}_{1} contains a part of each incoming ray of the puzzle of ν2\nu_{2}, this support must also contain the descendants of these incoming rays, and these include all the parallelogram edges other than the incoming rays. It is clear that the support of (ν^1)ν2(\widehat{\nu}_{1})_{\nu_{2}^{*}} contains all the edges of the light gray puzzle pieces. We record this fact as follows.

Refer to caption
Figure 4.8. Clockwise overlay with Σν1(ν2)=ω(ν1)ω(ν2)\Sigma_{\nu_{1}}(\nu_{2})=\omega(\nu_{1})\omega(\nu_{2})
Proposition 4.14.

Let μ\mu\in\mathcal{M} be an arbitrary rigid tree honeycomb. There exist rigid tree honeycombs ν\nu\in\mathcal{M} and ρ\rho\in\mathcal{M}_{*} such that ρν\rho\leq\nu^{*} and ρ\rho is homologous to a 6060^{\circ} rotation of μ\mu.

Remark 4.15.

At this point, we have enough information to calculate the exit pattern of any of the rigid tree honeycombs whose support is contained in the union of the boundaries of the dark gray parallelograms in a rigid puzzle. Thus, suppose that ν\nu is a rigid honeycomb, and write it as ν=j=1ncjνj\nu=\sum_{j=1}^{n}c_{j}\nu_{j}, where each νj\nu_{j} is a rigid tree honeycomb and Σνi(νj)=0\Sigma_{\nu_{i}}(\nu_{j})=0 for i>ji>j. Replacing cjc_{j} by 11 does not alter the structure of the puzzle of ν\nu, so we assume that cj=1c_{j}=1 for j=1,,nj=1,\dots,n. Let ff be a ray in the support of ν\nu, and let αj0\alpha_{j}\geq 0 be the corresponding exit multiplicity of νj\nu_{j}. The main observation is that the rigid honeycomb μ\mu, supported by the edges of parallelograms in the puzzle of ν\nu and rooted in the incoming ray in the inflation of ff, can be obtained as a result of a sequence of operations that we now describe. Suppose for the moment that α1>0\alpha_{1}>0.

  1. (1)

    Construct rigid tree honeycombs ν1(1),,νn(1)\nu_{1}(1),\dots,\nu_{n}(1), compatible with the puzzle of ν1\nu_{1} as follows: ν1(1)\nu_{1}(1) is supported by the puzzle edges and is rooted in the inflation of ff while, for j2j\geq 2, νj(1)\nu_{j}(1) is obtained by translating the support of νj\nu_{j} along with the white pieces as in Theorem 4.7. As seen above (Proposition 4.8 and Remark 4.10), these new tree honeycombs form again a rigid overlay.

  2. (2)

    Construct rigid tree honeycombs ν2(2),,νn(2)\nu_{2}(2),\dots,\nu_{n}(2), compatible with the puzzle of ν2(1)\nu_{2}(1), as follows: ν2(2)\nu_{2}(2) is obtained by translating the support of ν1(1)\nu_{1}(1) as in Theorem 4.11 and, for j3j\geq 3, νj(2)\nu_{j}(2) is obtained using the procedure of Theorem 4.7. These tree honeycombs also form a rigid overlay.

  3. (3)

    For k=3,,nk=3,\dots,n, construct rigid tree honeycombs νk(k),νk+1(k),,νn(k)\nu_{k}(k),\nu_{k+1}(k),\dots,\nu_{n}(k). The construction is done by induction and the inductive step is the procedure described in (2). Thus, supposing that these honeycombs are constructed for some k<nk<n, we construct the puzzle of νk+1(k)\nu_{k+1}(k) and we construct νk+1(k+1)\nu_{k+1}(k+1) by an application of Theorem 4.11 to νk(k)\nu_{k}(k) and we construct νk+1(j)\nu_{k+1}(j) by an application of Theorem 4.7. to νk(j)\nu_{k}(j), j=k+2,,nj=k+2,\dots,n.

  4. (4)

    The rigid tree honeycomb νn(n)\nu_{n}(n) is precisely the desired honeycomb μ\mu.

One can calculate the exit patterns of all the honeycombs constructed above, in particular, the exit pattern of μ\mu.

  1. (1)

    After the first operation described above, the honeycomb νj(1)\nu_{j}(1) is homologous to νj\nu_{j} for j2j\geq 2 and the exit pattern of ν1(1)\nu_{1}(1) is obtained by multiplying the exit pattern of ν1\nu_{1} by α1\alpha_{1}, except for the ray ff that gives rise to two exit multiplicities equal to α121\alpha_{1}^{2}-1 and 11. Moreover, we have Σν1(i)(ν1(j))=Σνi(νj)\Sigma_{\nu_{1}(i)}(\nu_{1}(j))=\Sigma_{\nu_{i}}(\nu_{j}) for i,j2,i,j\geq 2, and Σν1(1)(ν1(j))=αj+α1Σν1(νj)\Sigma_{\nu_{1}(1)}(\nu_{1}(j))=\alpha_{j}+\alpha_{1}\Sigma_{\nu_{1}}(\nu_{j}) for j2j\geq 2 (see Proposition 4.8).

  2. (2)

    After the second operation, the honeycomb ν2(j)\nu_{2}(j) is homologous to νj\nu_{j} for j3j\geq 3 and the exit pattern of ν2(2)\nu_{2}(2) is the same as the one for ν1(1)+σν1(2)\nu_{1}(1)+\sigma\nu_{1}(2), where σ=Σν1(1)(ν1(2))\sigma=\Sigma_{\nu_{1}(1)}(\nu_{1}(2)) (see Theorem 4.11). Moreover, we have Σν2(i)(ν2(j))=Σνi(νj)\Sigma_{\nu_{2}(i)}(\nu_{2}(j))=\Sigma_{\nu_{i}}(\nu_{j}) for i,j3,i,j\geq 3, and an easy calculation shows that

    Σν2(2)(ν2(j))=Σν1(1)(ν1(j))+σΣν1(2)(ν1(j)),j3.\Sigma_{\nu_{2}(2)}(\nu_{2}(j))=\Sigma_{\nu_{1}(1)}(\nu_{1}(j))+\sigma\Sigma_{\nu_{1}(2)}(\nu_{1}(j)),\quad j\geq 3.

    Further operations follow this model and they eventually yield the exit pattern of μ\mu.

In case α1=0\alpha_{1}=0, we look for the first nonzero αj\alpha_{j} and we can simply discard the honeycombs ν1,,νj1\nu_{1},\dots,\nu_{j-1} because they do not contribute to the exit pattern of μ\mu.

Example 4.16.

The special case of a rigid overlay of two tree honeycombs is used in Section 6. Thus, let ν1\nu_{1} and ν2\nu_{2} be rigid tree honeycombs such that Σν2(ν1)=0\Sigma_{\nu_{2}}(\nu_{1})=0 and Σν1(ν2)=σ>0\Sigma_{\nu_{1}}(\nu_{2})=\sigma>0. Let ff be a ray to which νj\nu_{j} assigns multiplicity αj\alpha_{j}, j=1,2j=1,2. The first operation in the preceding remark provides ν1(1)\nu_{1}(1) and ν1(2)\nu_{1}(2) such that Σν1(2)(ν1(1))=0\Sigma_{\nu_{1}(2)}(\nu_{1}(1))=0 and Σν1(1)(ν1(2))=α2+σα1\Sigma_{\nu_{1}(1)}(\nu_{1}(2))=\alpha_{2}+\sigma\alpha_{1}. The exit multiplicities assigned by ν1(1)\nu_{1}(1) to the outgoing and incoming rays in the inflation of ff are α121\alpha_{1}^{2}-1 and 11, respectively, while ν1(2)\nu_{1}(2) assigns multiplicity α2\alpha_{2} to the outgoing ray. If ee is another ray such that νj(e)=βj\nu_{j}(e)=\beta_{j}, j=1,2j=1,2, then the corresponding (outgoing) ray is assigned multiplicities α1β1\alpha_{1}\beta_{1} and β2\beta_{2}, respectively, by ν1(1)\nu_{1}(1) and ν1(2)\nu_{1}(2). After the second operation, we obtain a honeycomb μ=ν2(2)\mu=\nu_{2}(2) with typical exit multiplicities α1β1+(α2+σα1)β2\alpha_{1}\beta_{1}+(\alpha_{2}+\sigma\alpha_{1})\beta_{2}, and with exit multiplicities α121+(α2+σα1)α2\alpha_{1}^{2}-1+(\alpha_{2}+\sigma\alpha_{1})\alpha_{2} and 11 assigned to the outgoing and incoming rays in the inflation of ff (in the puzzle of ν1+ν2\nu_{1}+\nu_{2}). This process is illustrated in Figure 4.9, where ν1\nu_{1} is drawn in black, ν2\nu_{2} in red, and the numbers represent exit multiplicities. For this example, σ=2\sigma=2, α1=2\alpha_{1}=2, and α2=1\alpha_{2}=1. For comparison, we show in Figure 4.10 the same final honeycomb constructed directly on the puzzle of ν1+ν2\nu_{1}+\nu_{2}.

Refer to caption2+12+\color[rgb]{1,0,0}122221\color[rgb]{1,0,0}11111111111111\color[rgb]{1,0,0}11\color[rgb]{1,0,0}11\color[rgb]{1,0,0}13+13+\color[rgb]{1,0,0}11144442222222222225\color[rgb]{1,0,0}5445\color[rgb]{1,0,0}52222222222223+53+\color[rgb]{1,0,0}51144
Figure 4.9. The measures ν1,ν2,ν~1,ν~2\nu_{1},\nu_{2},\widetilde{\nu}_{1},\widetilde{\nu}_{2}, and the final honeycomb for Example 4.16
Refer to caption
Figure 4.10. The same honeycomb constructed using the puzzle of ν1+ν2\nu_{1}+\nu_{2}

5. Degeneration of a rigid tree honeycomb

Consider an arbitrary rigid honeycomb ν\nu\in\mathcal{M} and write its dual as

ν=c1ρ1++cnρn,\nu^{*}=c_{1}\rho_{1}+\cdots+c_{n}\rho_{n},

where n=root(ν)=exit(ν)root(ν)2,n={\rm root}(\nu^{*})={\rm exit}(\nu)-{\rm root(\nu)-2}, c1,,cn>0c_{1},\dots,c_{n}>0, and each ρj\rho_{j} is a rigid tree honeycomb in \mathcal{M}_{*} (see Theorem 3.2). As we pointed out before, the honeycombs μ\mu\in\mathcal{M} that are homologous to ν\nu are precisely the ones that satisfy

(5.1) μ=t1ρ1++tnρn\mu^{*}=t_{1}\rho_{1}+\cdots+t_{n}\rho_{n}

for some t1,,tn>0t_{1},\dots,t_{n}>0. Decreasing one of the coefficients tjt_{j} amounts to decreasing the lengths of some of the edges of μ\mu. If we replace some of the coefficients tjt_{j} by 0, the honeycomb μ\mu satisfying (5.1) is no longer homologous to ν\nu. We call it a degeneration of ν\nu. A simple degeneration of μ\mu is obtained by replacing exactly one of the tjt_{j} by 0. All the degenerations of ν\nu satisfy ω(μ)=ω(ν)\omega(\mu)=\omega(\nu). Moreover, the exit multiplicities of μ\mu are sums of one or several (consecutive) exit multiplicities of ν\nu (see Remark 3.1). For instance, if all the coefficients tjt_{j} are replaced by 0, we obtain the degeneration μ=ω(ν)ν0\mu=\omega(\nu)\nu_{0}, where ν0\nu_{0} is a rigid tree honeycomb of unit weight. Suppose that μ\mu is an arbitrary degeneration of ν\nu. An application of Theorem 3.2 to μ\mu and to ν\nu yields

root(μ)\displaystyle{\rm root(\mu)} =exit(μ)root(μ)2,\displaystyle={\rm exit}(\mu)-{\rm root}(\mu^{*})-2,
root(ν)\displaystyle{\rm root}(\nu) =exit(ν)root(ν)2,\displaystyle={\rm exit}(\nu)-{\rm root}(\nu^{*})-2,

and subtracting these equalities we obtain the equation

(5.2) root(μ)root(ν)=[root(ν)root(μ)][exit(ν)exit(μ)],{\rm root}(\mu)-{\rm root}(\nu)=[{\rm root}(\nu^{*})-{\rm root}(\mu^{*})]-[{\rm exit}(\nu)-{\rm exit}(\mu)],

where root(ν)root(μ)>0{\rm root}(\nu^{*})-{\rm root}(\mu^{*})>0 and exit(ν)exit(μ)0{\rm exit}(\nu)-{\rm exit}(\mu)\geq 0.

Suppose now that ν\nu is a rigid tree honeycomb and that μ\mu is a simple degeneration of ν\nu. Then we have root(ν)root(μ)=1{\rm root}(\nu^{*})-{\rm root}(\mu^{*})=1 and (5.2) allows for two possibilities:

  1. (1)

    root(μ)=2{\rm root}(\mu)=2 and exit(μ)=exit(ν){\rm exit}(\mu)={\rm exit}(\nu), or

  2. (2)

    root(μ)=1{\rm root}(\mu)=1 and exit(μ)=exit(ν)1{\rm exit}(\mu)={\rm exit}(\nu)-1.

We show that the first situation arises for at least one of the simple degenerations of ν\nu.

Theorem 5.1.

Suppose that ν\nu\in\mathcal{M} is a rigid tree honeycomb such that ω(ν)3\omega(\nu)\geq 3. Then there exist rigid tree honeycombs ν1,ν2\nu_{1},\nu_{2}\in\mathcal{M} satisfying Σν1(ν2)=0\Sigma_{\nu_{1}}(\nu_{2})=0 and such that the ν\nu has the same exit pattern as either ν1+σν2\nu_{1}+\sigma\nu_{2} or σν1+ν2\sigma\nu_{1}+\nu_{2}, where σ=Σν2(ν1)0\sigma=\Sigma_{\nu_{2}}(\nu_{1})\neq 0.

Proof.

The support of ν=c1ρ1++cnρn\nu^{*}=c_{1}\rho_{1}+\cdots+c_{n}\rho_{n} inside the triangle Δ\Delta_{*} of size ω(ν)\omega(\nu) has at least one flat segment ee of unit length, dual to a root edge of ν\nu. We can select j,k,{1,,n}j,k,\ell\in\{1,\dots,n\} such that ee is also an edge of ρj+ρk+ρ\rho_{j}+\rho_{k}+\rho_{\ell} (we may actually need three such honeycombs, one whose support contains ee and two more whose supports contain just the endpoints of ee.) Observe that n5n\geq 5 because ω(ν)3\omega(\nu)\geq 3, and thus we can choose i{1,,n}\{j,k,}.i\in\{1,\dots,n\}\backslash\{j,k,\ell\}. Consider the simple degeneration μ\mu of ν\nu defined by μ=νciρi\mu^{*}=\nu^{*}-c_{i}\rho_{i}. The choice of ii ensures that ee is an edge of μ\mu^{*}, so μ\mu has some edges with multiplicity 11.

If root(μ)=1{\rm root}(\mu)=1, it follows that μ\mu, having an edge of multiplicity 11, must itself be a rigid tree honeycomb and we are in the second situation described above, that is, exit(μ)=exit(ν)1{\rm exit}(\mu)={\rm exit}(\nu)-1. margin: review this proof It follows that the nonzero exit multiplicities of ν\nu can be arranged in counterclockwise order α1,,αn+3\alpha_{1},\dots,\alpha_{n+3} (see Theorem 3.2) so the nonzero exit multiplicities of μ\mu are α1+α2,α3,,αn+3\alpha_{1}+\alpha_{2},\alpha_{3},\dots,\alpha_{n+3}. We see then that

ω(μ)2\displaystyle\omega(\mu)^{2} =2+(α1+α2)2+j=3n+3αj2\displaystyle=-2+(\alpha_{1}+\alpha_{2})^{2}+\sum_{j=3}^{n+3}\alpha_{j}^{2}
=2α1α22+j=1n+3αj2\displaystyle=2\alpha_{1}\alpha_{2}-2+\sum_{j=1}^{n+3}\alpha_{j}^{2}
=2α1α2+ω(ν)2,\displaystyle=2\alpha_{1}\alpha_{2}+\omega(\nu)^{2},

and this is impossible because ω(μ)=ω(ν)\omega(\mu)=\omega(\nu) and α1α20\alpha_{1}\alpha_{2}\neq 0. We conclude that we must be in the first situation described above, that is, exit(μ)=exit(ν){\rm exit}(\mu)={\rm exit}(\nu) and μ\mu is a sum of two extreme honeycombs. One of these summands is a tree honeycomb because an edge of μ\mu has unit multiplicity, so μ=ν1+σν2\mu=\nu_{1}+\sigma\nu_{2} for some rigid tree honeycombs ν1\nu_{1} and ν2\nu_{2}, and σ>0\sigma>0. Since μ\mu is rigid, one of the numbers Σν1(ν2)\Sigma_{\nu_{1}}(\nu_{2}) and Σν2(ν1)\Sigma_{\nu_{2}}(\nu_{1}) is equal to zero. Finally, the equality exit(μ)=exit(ν){\rm exit}(\mu)={\rm exit(\nu)} implies that μ\mu and ν\nu have the same exit pattern (see Remark 3.1), in particular Σμ(μ)=Σν(ν)=1\Sigma_{\mu}(\mu)=\Sigma_{\nu}(\nu)=-1. Thus,

1+σΣν2(ν1)+σΣν1(ν2)σ2=1,-1+\sigma\Sigma_{\nu_{2}}(\nu_{1})+\sigma\Sigma_{\nu_{1}}(\nu_{2})-\sigma^{2}=-1,

so σ\sigma is equal to either Σν2(ν1)\Sigma_{\nu_{2}}(\nu_{1}) or Σν1(ν2)\Sigma_{\nu_{1}}(\nu_{2}). The theorem follows. ∎

It may happen that all the simple degenerations of ν\nu yield honeycombs ν1\nu_{1} and ν2\nu_{2} as in the preceding result. For instance, this is the case for all rigid tree honeycombs ν\nu of weight at most 55. This however is not true for all rigid tree honeycombs. The smallest examples correspond to the patterns 2,3,2|2,2,3|2,3,22,3,2|2,2,3|2,3,2 and 2,3,1|2,2,2|2,2,22,3,1|2,2,2|2,2,2. They are pictured in Figure 5.1

Refer to caption
Figure 5.1. (A) and (B) examples

along with their duals and one particular extreme rigid summand in the dual (drawn in red) that yields a simple degeneration unlike the ones envisaged in the proof of Theorem 5.1. One of the following situations arises in these examples:

  1. (A)

    A simple degeneration μ\mu of ν\nu has exit(μ)=exit(ν){\rm exit}(\mu)={\rm exit}(\nu) and the decomposition μ=m1ν1+m2ν2,\mu=m_{1}\nu_{1}+m_{2}\nu_{2}, where ν1\nu_{1} and ν2\nu_{2} are rigid tree honeycombs, has integer coefficients m1,m22m_{1},m_{2}\geq 2.

  2. (B)

    A simple degeneration μ\mu of ν\nu has exit(μ)=exit(ν)1{\rm exit}(\mu)={\rm exit}(\nu)-1. In this case, μ=cν0\mu=c\nu_{0} for some rigid tree honeycomb ν0\nu_{0} and some integer c2c\geq 2.

In both cases, all the edges of the dual honeycomb μ\mu^{*} have length greater than 11. Later, we describe all the rigid tree honeycombs ν\nu that have a degeneration of one of these two types (see Theorems 6.2 and 6.8). Here we find necessary conditions for such honeycombs. We start with case (A). In this situation, μ\mu has the same exit pattern as ν\nu and so Σμ(μ)=Σν(ν)=1\Sigma_{\mu}(\mu)=\Sigma_{\nu}(\nu)=-1. Suppose without loss of generality that Σν2(ν1)=0\Sigma_{\nu_{2}}(\nu_{1})=0 and calculate

Σμ(μ)=m12m22+σm1m2,\Sigma_{\mu}(\mu)=-m_{1}^{2}-m_{2}^{2}+\sigma m_{1}m_{2},

where σ=Σν1(ν2)\sigma=\Sigma_{\nu_{1}}(\nu_{2}). Thus, the pair (m1,m2)(m_{1},m_{2}) is a solution of the quadratic Diophantine equation

(5.3) m12+m22σm1m2=1.m_{1}^{2}+m_{2}^{2}-\sigma m_{1}m_{2}=1.

In case (B), we arrange the nonzero exit multiplicities of ν\nu in a counterclockwise sequence α1,,αm\alpha_{1},\dots,\alpha_{m} such that the nonzero exit multiplicities of μ\mu are α1+α2,α3,,αm\alpha_{1}+\alpha_{2},\alpha_{3},\dots,\alpha_{m}. Note that the exit multiplicities of ν0\nu_{0} are β2=(α1+α2)/c,β3=α3/c,,βm=αm/c\beta_{2}=(\alpha_{1}+\alpha_{2})/c,\beta_{3}=\alpha_{3}/c,\dots,\beta_{m}=\alpha_{m}/c. We have Σμ(μ)=Σcν0(cν0)=c2\Sigma_{\mu}(\mu)=\Sigma_{c\nu_{0}}(c\nu_{0})=-c^{2} and an application of (2.7) to ν\nu and to μ\mu yields

j=1mαj2=ω(ν)2+2,(α1+α2)2+j=3mαj2=ω(μ)2+2c2.\sum_{j=1}^{m}\alpha_{j}^{2}=\omega(\nu)^{2}+2,\quad(\alpha_{1}+\alpha_{2})^{2}+\sum_{j=3}^{m}\alpha_{j}^{2}=\omega(\mu)^{2}+2c^{2}.

Recalling that ω(μ)=ω(ν)\omega(\mu)=\omega(\nu), subtract the two equalities to obtain

α1α2=c21.\alpha_{1}\alpha_{2}=c^{2}-1.

Now substitute β2cα1\beta_{2}c-\alpha_{1} for α2\alpha_{2} to obtain

(5.4) c2+α12β2α1c=1,c^{2}+\alpha_{1}^{2}-\beta_{2}\alpha_{1}c=1,

and similarly,

(5.5) c2+α22β2α2c=1.c^{2}+\alpha_{2}^{2}-\beta_{2}\alpha_{2}c=1.

These equations are of the same kind as (5.3)\ref{eq:Diophantine c1 and c2}) with β2\beta_{2} in place of σ\sigma.

We discuss briefly the structure of the nonnegative integer solutions (p,q)(p,q) of the equation

(5.6) p2+q2σpq=1p^{2}+q^{2}-\sigma pq=1

for a given integer σ1\sigma\geq 1. If σ=1,\sigma=1, the equation can be rewritten as

(pq)2+pq=1,(p-q)^{2}+pq=1,

and one immediately deduces that the only nonnegative integer solutions are (0,1)(0,1), (1,0)(1,0), and (1,1)(1,1).

If σ=2\sigma=2, the equation becomes (pq)2=1(p-q)^{2}=1, so the nonnegative integer solutions are (n,n+1)(n,n+1) and (n+1,n)(n+1,n) for n=0,1,n=0,1,\dots.

If σ3\sigma\geq 3, a simple substitution transforms (5.6) into a standard Pell equation (see [10, Chapter 8]). It is easier however to treat the equation directly. Denote by ξ\xi and ξ\xi^{\prime} the two solutions of the quadratic equation ξ2σξ+1=0\xi^{2}-\sigma\xi+1=0, so ξ+ξ=σ\xi+\xi^{\prime}=\sigma and ξξ=1\xi\xi^{\prime}=1. We set

[ξ]={pξq:p,q},\mathbb{Z}[\xi]=\{p-\xi q:p,q\in\mathbb{Z}\},

and we define the multiplicative function (sometimes called norm) N:[ξ]N:\mathbb{Z}[\xi]\to\mathbb{Z} by

N(pξq)=(pξq)(pξq)=p2+q2σpq,pξq[ξ].N(p-\xi q)=(p-\xi q)(p-\xi^{\prime}q)=p^{2}+q^{2}-\sigma pq,\quad p-\xi q\in\mathbb{Z}[\xi].

We need to study the multiplicative group G={z[ξ]:N(z)=1}G=\{z\in\mathbb{Z}[\xi]:N(z)=1\}. This group is generated, for instance, by 1-1 and ξ\xi^{\prime}. Define a sequence {pn:n=0,1,}\{p_{n}:n=0,1,\dots\} of nonnegative integers by setting

(5.7) p0=0,p1=1,pn+1=σpnpn1,n.p_{0}=0,p_{1}=1,p_{n+1}=\sigma p_{n}-p_{n-1},\quad n\in\mathbb{N}.

An induction argument shows that

ξn=pn+1pnξ and ξn=(pn1pnξ),n.\xi^{\prime n}=p_{n+1}-p_{n}\xi\text{ and $\xi^{\prime-n}=-(p_{n-1}-p_{n}\xi),\quad n\in\mathbb{N}.$}

We deduce that the nonnegative integer solutions of (5.6) are precisely the pairs (pn,pn+1)(p_{n},p_{n+1}) and (pn+1,pn)(p_{n+1},p_{n}) for n=0,1,n=0,1,\dots.

We note that the numbers pn=np_{n}=n also satisfy the identity pn+1=2pnpn1p_{n+1}=2p_{n}-p_{n-1}. In other words, the same description of the solutions of (5.6) applies to the case σ=2\sigma=2.

We note for further use some identities that the sequence {pn}n=0\{p_{n}\}_{n=0}^{\infty} satisfies. First, consider the column vectors vn=[pn+1pn]v_{n}=\left[\begin{array}[]{c}p_{n+1}\\ p_{n}\end{array}\right] that satisfy vn+1=σvnvn1v_{n+1}=\sigma v_{n}-v_{n-1} for n>0.n>0. We deduce that

det[vn+1,vn+2]=σdet[vn+1,vn+1]det[vn+1,vn]=det[vn,vn+1],\det[v_{n+1},v_{n+2}]=\sigma\det[v_{n+1},v_{n+1}]-\det[v_{n+1},v_{n}]=\det[v_{n},v_{n+1}],

and thus det[vn,vn+1]\det[v_{n},v_{n+1}] does not depend on nn. Evaluating this determinant for n=0n=0 we see that

pn+12pnpn+2=det[vn,vn+1]=1,n0.p_{n+1}^{2}-p_{n}p_{n+2}=\det[v_{n},v_{n+1}]=1,\quad n\geq 0.

Equivalently,

(5.8) pn+121=pnpn+2,n0.p_{n+1}^{2}-1=p_{n}p_{n+2},\quad n\geq 0.

The inductive argument above is easily seen to yield the more general equality

det[vn,vn+k]=pk,k,n0,\det[v_{n},v_{n+k}]=p_{k},\quad k,n\geq 0,

or, equivalently,

pn+1pn+kpnpn+k+1=pk,k,n0.p_{n+1}p_{n+k}-p_{n}p_{n+k+1}=p_{k},\quad k,n\geq 0.

In Section 6 we need the special case k=2k=2. Since p2=σp_{2}=\sigma, this can be rewritten as

(5.9) pnpn+1σ=pn1pn+2,n1.p_{n}p_{n+1}-\sigma=p_{n-1}p_{n+2},\quad n\geq 1.

These identities show, for instance, that pnp_{n} and pn+1p_{n+1} are relatively prime and that the greatest common divisor of pnp_{n} and pn+2p_{n+2} is σ\sigma if nn is even.

The first few terms of the sequence pnp_{n} are

0,1,σ,σ21,σ32σ,σ43σ2+1,σ54σ3+3σ,0,1,\sigma,\sigma^{2}-1,\sigma^{3}-2\sigma,\sigma^{4}-3\sigma^{2}+1,\sigma^{5}-4\sigma^{3}+3\sigma,

and a closed formula for these numbers is

pn=ξnξnξξ,n0.p_{n}=\frac{\xi^{n}-\xi^{\prime n}}{\xi-\xi^{\prime}},\quad n\geq 0.
Remark 5.2.

If σ=2\sigma=2 and pnp_{n} is defined inductively by (5.7), then pn=np_{n}=n. Thus, even in this case, the nonnegative solutions of (5.6) are the pairs (pn,pn+1)(p_{n},p_{n+1}) and (pn+1,pn)(p_{n+1},p_{n}). If σ=1\sigma=1, then the sequence pnp_{n} defined inductively by (5.7) is periodic and the nonnegative solutions of (5.6) are the pairs (pn,pn+1)(p_{n},p_{n+1}) with n=0,1,2n=0,1,2.

Returning now to the discussion of the degenerations of a rigid tree honeycomb, we have the following result.

Proposition 5.3.

Let ν\nu be a rigid tree honeycomb and let μ\mu be a simple degeneration of ν\nu. Then one of the following cases occurs:

  1. (1)

    There exist rigid tree honeycombs ν1\nu_{1} and ν2\nu_{2} such that Σν1(ν2)=1\Sigma_{\nu_{1}}(\nu_{2})=1, Σν2(ν1)=0\Sigma_{\nu_{2}}(\nu_{1})=0, and μ=ν1+ν2.\mu=\nu_{1}+\nu_{2}.

  2. (2)

    There exist rigid tree honeycombs ν1\nu_{1} and ν2\nu_{2} such that Σν2(ν1)=0\Sigma_{\nu_{2}}(\nu_{1})=0, Σν1(ν2)=σ>1\Sigma_{\nu_{1}}(\nu_{2})=\sigma>1, and μ=m1ν1+m2ν2,\mu=m_{1}\nu_{1}+m_{2}\nu_{2}, where m1m_{1} and m2m_{2} are consecutive terms in the sequence {pn}n=1\{p_{n}\}_{n=1}^{\infty} defined by (5.7).

  3. (3)

    There exists a rigid tree honeycomb ν0\nu_{0} that has an exit multiplicity σ>1\sigma>1 and there exist three consecutive terms pn,pn+1,pn+2p_{n},p_{n+1},p_{n+2} in the sequence {pn}n=1\{p_{n}\}_{n=1}^{\infty} defined by (5.7), such that the exit multiplicities of μ\mu are obtained by replacing each exit multiplicity β\beta of ν0\nu_{0} by pn+1βp_{n+1}\beta except for the multiplicity σ\sigma which is replaced by two consecutive exit multiplicities equal to pnp_{n} and pn+2p_{n+2}, respectively.

Proof.

Parts (1) and (2) follow immediately from the above discussion of the equation (5.6). For (3), set σ=β2\sigma=\beta_{2} in equations (5.4) and (5.5) to see that c,α1c,\alpha_{1} and c,α2c,\alpha_{2} must be consecutive pairs of elements in the sequence pnp_{n}. Moreover, α2=σcα1\alpha_{2}=\sigma c-\alpha_{1}, showing that α1,c,α2\alpha_{1},c,\alpha_{2} are three consecutive terms of this sequence (in either increasing or decreasing order). ∎

The proof of Theorem 5.1 shows that all but at most three of the simple degenerations of a rigid tree honeycomb fall under either case (1) or case (2) above with min{m1,m2}=1\min\{m_{1},m_{2}\}=1. Case (A) above corresponds to (2) with min{m1,m2}>1\min\{m_{1},m_{2}\}>1 and (B) corresponds to (3). The following result shows that at most two degenerations are as in (B).

Theorem 5.4.

Suppose that ν\nu is a rigid tree honeycomb. Then:

  1. (1)

    If ν\nu has two simple degenerations that are extreme honeycombs, then ν\nu has no simple degeneration of the form m1ν1+m2ν2m_{1}\nu_{1}+m_{2}\nu_{2} with ν1,ν2\nu_{1},\nu_{2} rigid tree honeycombs and m1,m22m_{1},m_{2}\geq 2.

  2. (2)

    At most two of the simple degenerations of ν\nu are extreme honeycombs.

  3. (3)

    If two of the simple degenerations of ν\nu are extreme honeycombs, then there exists a rigid tree honeycomb ν\nu^{\prime}, with exit multiplicities δ1,,δk\delta_{1},\dots,\delta_{k}, listed in counterclockwise order, such that:

    1. (a)

      δ1>1\delta_{1}>1.

    2. (b)

      If the sequence {pn}n=0\{p_{n}\}_{n=0}^{\infty} is defined by (5.7) with σ=δ1\sigma=\delta_{1}, then there exists n>1n>1 such that ω(μ)=pnpn+1ω(ν)\omega(\mu)=p_{n}p_{n+1}\omega(\nu^{\prime}), and the exit multiplicities of μ\mu, arranged in counterclockwise order, are either

      pn21,1,pn+121,pnpn+1δ2,,pnpn+1δk,p_{n}^{2}-1,1,p_{n+1}^{2}-1,p_{n}p_{n+1}\delta_{2},\dots,p_{n}p_{n+1}\delta_{k},

      or

      pn+121,1,pn21,pnpn+1δ2,,pnpn+1δk,.p_{n+1}^{2}-1,1,p_{n}^{2}-1,p_{n}p_{n+1}\delta_{2},\dots,p_{n}p_{n+1}\delta_{k},.

      where the first three exit multiplicities correspond to parallel rays.

Proof.

Part (1) follows immediately from part (3) because the nonzero exit multiplicities of m1ν1+m2ν2m_{1}\nu_{1}+m_{2}\nu_{2}, and hence those of ν\nu, are at least 22. As observed already, the exit multiplicities of a simple degeneration that is an extreme honeycomb are obtained simply by replacing two neighboring exit multiplicities of ν\nu by their sum and leaving the others multiplicities unchanged.

If we write ν=j=1ncjρj\nu^{*}=\sum_{j=1}^{n}c_{j}\rho_{j} as above, this situation corresponds to the fact that there exists a ray II such that ρj0(I)>0\rho_{j_{0}}(I)>0 but ρj(I)=0\rho_{j}(I)=0 for every jj0.j\neq j_{0}. Suppose that ν\nu has at least two simple degenerations of this type. Reordering the honeycombs ρj\rho_{j}, we may assume that these degenerations μ1\mu_{1} and μ2\mu_{2} satisfy μ1=νc1ρ1\mu_{1}^{*}=\nu^{*}-c_{1}\rho_{1} and μ2=νc2ρ2\mu_{2}^{*}=\nu^{*}-c_{2}\rho_{2}. Consider also the degeneration μ0\mu_{0} such that μ0=νc1ρ1c2ρ2\mu_{0}^{*}=\nu^{*}-c_{1}\rho_{1}-c_{2}\rho_{2}. Since there are two distinct rays I1,I2I_{1},I_{2} to which ν\nu^{*} assigns positive multiplicity such μi\mu_{i}^{*} assigns zero multiplicity to μi\mu_{i}, i=1,2,i=1,2, we see that exit(μ0)exit(μ)2{\rm exit}(\mu_{0}^{*})\leq{\rm exit}(\mu)-2. By Theorem 3.2, we have exit(μ)=n+3{\rm exit}(\mu)=n+3 and

root(μ0)+root(μ0)=exit(μ0)2exit(μ)4=n1.{\rm root}(\mu_{0}^{*})+{\rm root}(\mu_{0})={\rm exit(\mu_{0})-2\leq{\rm{\rm exit}(\mu)}-4}=n-1.

Since root(μ0)=n2{\rm root}(\mu_{0}^{*})=n-2, we see that root(μ0)1{\rm root}(\mu_{0})\leq 1, and thus μ0\mu_{0} is an extreme rigid honeycomb. Thus, there are rigid tree honeycombs ν0,ν1,ν2\nu_{0},\nu_{1},\nu_{2} and integers k0,k1,k2>1k_{0},k_{1},k_{2}>1 such that μj=kjνj\mu_{j}=k_{j}\nu_{j} for j=0,1,2j=0,1,2. Moreover, since μ0\mu_{0} is also a simple degeneration of μ1\mu_{1}, k0>k1k_{0}>k_{1}; similarly, k0>k2k_{0}>k_{2} and, in fact, k0k_{0} is a multiple of both k1k_{1} and k2k_{2}. The exit pattern of μ1\mu_{1} is the same as that of ν\nu, except that two consecutive exit multiplicities α1,α2\alpha_{1},\alpha_{2} of ν\nu are replaced by α1+α2\alpha_{1}+\alpha_{2}. We next exclude the possibility that the exit pattern of μ2\mu_{2} is the same as that of ν\nu, except that two consecutive exit multiplicities α3,α4\alpha_{3},\alpha_{4}, other than α1,α2\alpha_{1},\alpha_{2}, are replaced by α3+α4\alpha_{3}+\alpha_{4}. Suppose that this possibility does arise. In this case, the exit multiplicities of the honeycombs involved can be listed as follows.

ν\nu α1,α2\alpha_{1},\alpha_{2} α3,α4\alpha_{3},\alpha_{4} α5,\alpha_{5},\dots
μ1\mu_{1} α1+α2=k1β1\alpha_{1}+\alpha_{2}=k_{1}\beta_{1} α3=k1β3,α4=k1β4\alpha_{3}=k_{1}\beta_{3},\alpha_{4}=k_{1}\beta_{4} α5=k1β5,\alpha_{5}=k_{1}\beta_{5},\dots
μ2\mu_{2} α1=k2γ1,α2=k2γ2\alpha_{1}=k_{2}\gamma_{1},\alpha_{2}=k_{2}\gamma_{2} α3+α4=k2γ3\alpha_{3}+\alpha_{4}=k_{2}\gamma_{3} α5=k2γ5,\alpha_{5}=k_{2}\gamma_{5},\dots
μ0\mu_{0} α1+α2=k0δ1\alpha_{1}+\alpha_{2}=k_{0}\delta_{1} α3+α4=k0δ3\alpha_{3}+\alpha_{4}=k_{0}\delta_{3} α5=k0δ5,\alpha_{5}=k_{0}\delta_{5},\dots

Here, βj,γj,δj\beta_{j},\gamma_{j},\delta_{j} represent the nonzero exit multiplicities of ν1,ν2,ν0\nu_{1},\nu_{2},\nu_{0}, respectively. Note that β2,γ4,δ2\beta_{2},\gamma_{4},\delta_{2}, and δ4\delta_{4} are not listed because μ1,μ2,\mu_{1},\mu_{2}, and μ0\mu_{0} have fewer nonzero exit multiplicities than ν\nu. We now apply (2.8) to the four rigid tree honeycombs involved to obtain

α12+α22+α32+α42+j5αj2\displaystyle\alpha_{1}^{2}+\alpha_{2}^{2}+\alpha_{3}^{2}+\alpha_{4}^{2}+\sum_{j\geq 5}\alpha_{j}^{2} =ω(ν)2+2,\displaystyle=\omega(\nu)^{2}+2,
(α1+α2)2+α32+α+42j5αj2\displaystyle(\alpha_{1}+\alpha_{2})^{2}+\alpha_{3}^{2}+\alpha{}_{4}^{2}+\sum_{j\geq 5}\alpha_{j}^{2} =k12(ω(ν1)2+2)=ω(ν)2+2k12,\displaystyle=k_{1}^{2}(\omega(\nu_{1})^{2}+2)=\omega(\nu)^{2}+2k_{1}^{2},
α12+α22+(α3+α4)2+j5αj2\displaystyle\alpha_{1}^{2}+\alpha_{2}^{2}+(\alpha_{3}+\alpha_{4})^{2}+\sum_{j\geq 5}\alpha_{j}^{2} =ω(ν)2+2k22,\displaystyle=\omega(\nu)^{2}+2k_{2}^{2},
(α1+α2)2+(α3+α4)2+j5αj2\displaystyle(\alpha_{1}+\alpha_{2})^{2}+(\alpha_{3}+\alpha_{4})^{2}+\sum_{j\geq 5}\alpha_{j}^{2} =ω(ν)2+2k02.\displaystyle=\omega(\nu)^{2}+2k_{0}^{2}.

Subtract now the first equality from the other three to see that

α1α2=k121,α3α4=k221,α1α2+α3α4=k021,\alpha_{1}\alpha_{2}=k_{1}^{2}-1,\quad\alpha_{3}\alpha_{4}=k_{2}^{2}-1,\quad\alpha_{1}\alpha_{2}+\alpha_{3}\alpha_{4}=k_{0}^{2}-1,

and thus

k02+1=k12+k22.k_{0}^{2}+1=k_{1}^{2}+k_{2}^{2}.

Since k0k_{0} is a common multiple of k1k_{1} and k2k_{2}, it follows that k1k_{1} and k2k_{2} are relatively prime, so k0=mk1k2k_{0}=mk_{1}k_{2} for some mm\in\mathbb{N}. Therefore

k12+k22=m2k12k22+1k12k22+1,k_{1}^{2}+k_{2}^{2}=m^{2}k_{1}^{2}k_{2}^{2}+1\geq k_{1}^{2}k_{2}^{2}+1,

hence (k121)(k221)0(k_{1}^{2}-1)(k_{2}^{2}-1)\leq 0, and this is impossible because kj>1k_{j}>1.

We see right away that it is not possible to have three degenerations of ν\nu of this type because two of them would have to involve disjoint sets of exit multiplicities. This proves (2).

Let now μ1,μ2,μ0\mu_{1},\mu_{2},\mu_{0}, ν1,ν2,ν0\nu_{1},\nu_{2},\nu_{0}, and k0,k1,k2k_{0},k_{1},k_{2} be as above. The preceding argument shows that the exit multiplicities of these honeycombs can be arranged as follows.

ν\nu α1,α2,α3\alpha_{1},\alpha_{2},\alpha_{3} α4,\alpha_{4},\dots
μ1\mu_{1} α1+α2=k1β1,α3=k1β3\alpha_{1}+\alpha_{2}=k_{1}\beta_{1},\alpha_{3}=k_{1}\beta_{3} α4=k1β4,\alpha_{4}=k_{1}\beta_{4},\dots
μ2\mu_{2} α1=k2γ1,α2+α3=k2γ2\alpha_{1}=k_{2}\gamma_{1},\alpha_{2}+\alpha_{3}=k_{2}\gamma_{2} α4=k2γ4,\alpha_{4}=k_{2}\gamma_{4},\dots
μ0\mu_{0} α1+α2+α3=k0δ1\alpha_{1}+\alpha_{2}+\alpha_{3}=k_{0}\delta_{1} α4=k0δ4,\alpha_{4}=k_{0}\delta_{4},\dots

Of course, α1,α2,\alpha_{1},\alpha_{2}, and α3\alpha_{3} must be consecutive exit multiplicities corresponding to parallel rays. As in the situation discussed above, the numbers k1k_{1} and k2k_{2} must be relatively prime. Indeed, a common factor dd of these integers must divide ω(ν)\omega(\nu), α1,\alpha_{1},α3,\alpha_{3}, and αj\alpha_{j} for j4j\geq 4, and therefore dd also divides α2=3ω(ν)j2αj\alpha_{2}=3\omega(\nu)-\sum_{j\neq 2}\alpha_{j}. Now, the relation j1αj2=ω(ν)2+2\sum_{j\geq 1}\alpha_{j}^{2}=\omega(\nu)^{2}+2 shows that d2d^{2} divides 22, so d=1d=1. Thus k0=mk1k2k_{0}=mk_{1}k_{2} for some mm\in\mathbb{N}. We apply again (2.8) to obtain

α12+α22+α32+j4αj2\displaystyle\alpha_{1}^{2}+\alpha_{2}^{2}+\alpha_{3}^{2}+\sum_{j\geq 4}\alpha_{j}^{2} =ω(ν)2+2,\displaystyle=\omega(\nu)^{2}+2,
(α1+α2)2+α32+j4αj2\displaystyle(\alpha_{1}+\alpha_{2})^{2}+\alpha_{3}^{2}+\sum_{j\geq 4}\alpha_{j}^{2} =ω(ν)2+2k12,\displaystyle=\omega(\nu)^{2}+2k_{1}^{2},
α12+(α2+α3)2+j4αj2\displaystyle\alpha_{1}^{2}+(\alpha_{2}+\alpha_{3})^{2}+\sum_{j\geq 4}\alpha_{j}^{2} =ω(ν)2+2k22,\displaystyle=\omega(\nu)^{2}+2k_{2}^{2},
(α1+α2+α3)2+j4αj2\displaystyle(\alpha_{1}+\alpha_{2}+\alpha_{3})^{2}+\sum_{j\geq 4}\alpha_{j}^{2} =ω(ν)2+2k02,\displaystyle=\omega(\nu)^{2}+2k_{0}^{2},

and we subtract the first equality from the others to see that

α1α2=k121,α2α3=k221,α1α2+α2α3+α1α3=k021.\alpha_{1}\alpha_{2}=k_{1}^{2}-1,\quad\alpha_{2}\alpha_{3}=k_{2}^{2}-1,\quad\alpha_{1}\alpha_{2}+\alpha_{2}\alpha_{3}+\alpha_{1}\alpha_{3}=k_{0}^{2}-1.

Thus

α1α3=m2k12k22k12k22+1(k121)(k221),\alpha_{1}\alpha_{3}=m^{2}k_{1}^{2}k_{2}^{2}-k_{1}^{2}-k_{2}^{2}+1\geq(k_{1}^{2}-1)(k_{2}^{2}-1),

and since α1=(k121)/α2,\alpha_{1}=(k_{1}^{2}-1)/\alpha_{2}, α3=(k221)/α2\alpha_{3}=(k_{2}^{2}-1)/\alpha_{2}, these relations are only possible when m=α2=1,m=\alpha_{2}=1, α1=k121\alpha_{1}=k_{1}^{2}-1, and α3=k221\alpha_{3}=k_{2}^{2}-1. Now set σ=δ1\sigma=\delta_{1} and let {pn}n=0\{p_{n}\}_{n=0}^{\infty} be defined by (5.7). Since μ0\mu_{0} is a simple degeneration of μ1\mu_{1}, it follows from Proposition 5.3(3) that β1,k2=k0/k1,\beta_{1},k_{2}=k_{0}/k_{1}, and β3\beta_{3} are consecutive terms of this sequence. Finally, β1=(α1+α2)/k1=k1\beta_{1}=(\alpha_{1}+\alpha_{2})/k_{1}=k_{1}, and therefore k1k_{1} and k2k_{2} are consecutive terms in the sequence {pn}n=0\{p_{n}\}_{n=0}^{\infty}. This concludes the proof of (3). ∎

6. Regeneration

In Section 5, we found necessary conditions for a honeycomb to be one of the simple degenerations of a rigid tree honeycomb. Our purpose in this section is to show that these necessary conditions are sufficient as well. The argument requires a basic construction of rigid tree honeycombs that involves a combination of an inflation with a partial deflation.

Construction

We start with the following data:

  1. (1)

    a rigid honeycomb ν\nu.

  2. (2)

    a rigid honeycomb μ\mu that is compatible with the puzzle of ν\nu. We set ν=μν\nu^{\prime}=\mu_{\nu}, where μν\mu_{\nu} is defined in Section 4. It follows from [6, Theorem 7.14] and [7, Theorem 1.4] (see also the discussion following [3, Theorem 7.2]) that ν\nu^{\prime} is also rigid.

  3. (3)

    a ray ff in the support of μ\mu.

We proceed in three steps.

  1. (i)

    Construct the puzzle of μ\mu and preserve the coloring of the ‘white’ pieces. Color the remainder of the pieces as follows:

    1. (a)

      The pieces corresponding to branch points of μ\mu are colored light red.

    2. (b)

      The inflation of (part of) an edge is colored dark red if it is contained in the closure of a white piece of the puzzle of ν\nu or if it crosses a dark gray parallelogram and is parallel to its white sides. Two of the sides of a dark red parallelograms are white and the other two are light red.

    3. (c)

      The inflation of (part of) an edge is colored black if it is contained in the closure of a light gray piece of the puzzle of ν\nu or if it crosses a dark gray parallelogram and is parallel to its light gray sides.

  2. (ii)

    Construct the measure μ~\widetilde{\mu} with support in the edges of the puzzle of μ\mu and rooted in the incoming ray corresponding to ff (Theorem 4.7).

  3. (iii)

    Remove all the black, dark gray, and light gray areas of the puzzle constructed in (i), translate the remaining pieces to tile the plane, and preserve the multiplicities assigned by μ~\widetilde{\mu} for those (parts of) edges that are contained in the closure of a white piece or in a dark gray parallelogram and are parallel to its white sides. Add multiplicities in case several such edges are translated to the same segment. Call μ\mu^{\prime} the collection of multiplicities obtained this way.

Refer to caption
Figure 6.1. The basic construction inside a white puzzle piece
Refer to caption
Figure 6.2. The basic construction inside a dark gray parallelogram
Refer to caption
Figure 6.3. The basic construction inside a light gray puzzle piece
Proposition 6.1.

The construction above yields:

  1. (1)

    The puzzle of ν\nu^{\prime}, with the color gray replaced by red, and

  2. (2)

    A rigid honeycomb μ\mu^{\prime} that is compatible with the puzzle of ν\nu^{\prime}.

Proof.

The white areas in the puzzle of ν\nu are divided by the support of μ\mu into smaller regions, and it is these regions that are the white puzzle pieces that remain after the operation described in (iii) above. Suppose that two such areas G1G_{1} and G2G_{2} are separated by an edge ee of μ\mu. Then their counterparts after operation (iii) are separated by a dark red parallelogram with two edges parallel to ee and two edges of length μ(e)\mu(e) that are 6060^{\circ} clockwise from ee. Of course, ee is the translate of some edge ee^{\prime} of ν\nu^{\prime} with multiplicity ν(e)=μ(e)\nu(e^{\prime})=\mu(e), and thus the dark red parallelogram is the one that normally appears in the construction of the puzzle of ν\nu^{\prime}. Another possibility is that G1G_{1} and G2G_{2} are separated by a dark gray parallelogram in the puzzle of ν\nu. Suppose that ABAB and ABA^{\prime}B^{\prime} are the two white sides of that parallelogram, and that I1,,ImI_{1},\dots,I_{m} are the edges of μ\mu that cross the parallelogram and are parallel to its white sides. Then, in the picture produced by (iii), G1G_{1} and G2G_{2} are separated by a dark red parallelogram that is decomposed into m+2m+2 parallelograms with light red sides of lengths μ(AB),μ(I1),,μ(Im),μ(AB)\mu(AB),\mu(I_{1}),\dots,\mu(I_{m}),\mu(A^{\prime}B^{\prime}). The sum of these lengths is precisely ν(e)\nu^{\prime}(e^{\prime}), where ee^{\prime} is the edge of ν\nu with inflation ABABABA^{\prime}B^{\prime}. Similarly, the light red pieces assemble in step (iii) to complete the puzzle of ν\nu^{\prime}. This process is illustrated in Figure 6.1 for white pieces in the puzzle of ν\nu, in Figure 6.3 for light gray puzzle pieces, and in Figure 6.2 for dark gray puzzle pieces. These figures are sufficiently general (they deal with branch points of order five contained in the interior, on the boundary, or in a corner of a puzzle piece) to demonstrate that the white, dark red, and light red pieces do actually fit together to form the puzzle of ν\nu^{\prime}. The multiplicities chosen for these figures are the smallest positive integers that satisfy the the balance condition on the (red) edges of μ\mu. Choosing different multiplicities (possibly zero) produce essentially the same picture (possibly without some of the dark red parallelograms).

In order to verify that μ\mu^{\prime} also satisfies the balance condition we observe that μ\mu^{\prime} is obtained from μ~\widetilde{\mu} by decreasing the lengths of its edges without changing their multiplicities, except in those cases in which two edges are translated to the same segment and their multiplicities are added. In other words, μ\mu^{\prime} is a degeneration of μ~\widetilde{\mu} and is therefore a rigid honeycomb. ∎

Theorem 6.2.

Suppose that ν1\nu_{1} and ν2\nu_{2} are rigid tree honeycombs such that Σν1(ν2)=σ1\Sigma_{\nu_{1}}(\nu_{2})=\sigma\geq 1 and Σν2(ν1)=0\Sigma_{\nu_{2}}(\nu_{1})=0. Let {pn}n=0\{p_{n}\}_{n=0}^{\infty} be the sequence defined by (5.7). Then, for every n1n\geq 1, there exists a rigid tree honeycomb μn\mu_{n} such that:

  1. (1)

    μn\mu_{n} and pnν1+pn+1ν2p_{n}\nu_{1}+p_{n+1}\nu_{2} have the same exit pattern, and

  2. (2)

    μn\mu_{n} is compatible with the puzzle of pn1ν1+pnν2p_{n-1}\nu_{1}+p_{n}\nu_{2}.

Proof.

We proceed by induction. The existence of μ1=ν^1\mu_{1}=\widehat{\nu}_{1} is a consequence of Theorem 4.11. Suppose that nn\in\mathbb{N} and that μn\mu_{n} has been constructed. Set ν=pn1ν1+pnν2\nu=p_{n-1}\nu_{1}+p_{n}\nu_{2} and ν=(μn)ν=pnν1+pn+1ν2\nu^{\prime}=(\mu_{n})_{\nu}=p_{n}\nu_{1}+p_{n+1}\nu_{2}. The nonzero exit multiplicity of μn\mu_{n} corresponding to a typical ray ee is of the form pnα+pn+1βp_{n}\alpha+p_{n+1}\beta, where α\alpha and β\beta are exit multiplicities of ν1\nu_{1} and ν2\nu_{2}, respectively, and α+β>0\alpha+\beta>0. Choose a particular ray e0e_{0} of μn\mu_{n} with exit multiplicity pnα0+pn+1β0p_{n}\alpha_{0}+p_{n+1}\beta_{0} such that β0>0\beta_{0}>0. We apply the construction above to the measures ν\nu, μ=μn,\mu=\mu_{n}, and the incoming ray ff in the inflation of e0e_{0}. We obtain a honeycomb μ\mu^{\prime} compatible with the puzzle of ν\nu^{\prime}. The honeycomb μ~\widetilde{\mu} constructed in step (ii) has exit multiplicities (pnα0+pn+1β0)21(p_{n}\alpha_{0}+p_{n+1}\beta_{0})^{2}-1 and 11 corresponding to the outgoing and incoming rays of the puzzle of μ\mu corresponding to e0e_{0}. The other nonzero exit multiplicities correspond to the remaining outgoing rays and they are of the form (pnα0+pn+1β0)(pnα+pn+1β)(p_{n}\alpha_{0}+p_{n+1}\beta_{0})(p_{n}\alpha+p_{n+1}\beta). It follows, in particular, that

μν=(pnα0+pn+1β0)(pnν1+pn+1ν2).\mu^{\prime}_{\nu^{\prime}}=(p_{n}\alpha_{0}+p_{n+1}\beta_{0})(p_{n}\nu_{1}+p_{n+1}\nu_{2}).

Next, observe that μ\mu^{\prime} assigns unit multiplicity to the incoming ray in the inflation of e0e_{0} in the puzzle of ν\nu^{\prime}. Denote by ρ\rho the rigid tree honeycomb supported by the puzzle edges of ν\nu^{\prime} and rooted in the incoming ray corresponding to e0e_{0}. Lemma 4.3 implies that there exists a rigid honeycomb μ′′\mu^{\prime\prime} such that μ=ρ+μ′′\mu^{\prime}=\rho+\mu^{\prime\prime}. Suppose that ff0f\neq f_{0} is an arbitrary ray of ν\nu^{\prime} such that ν1(f)=α\nu_{1}(f)=\alpha and ν2(f)=β.\nu_{2}(f)=\beta. Then, according to Example 4.16, the multiplicity that ρ\rho assigns to the exit ray corresponding to ee0e\neq e_{0} in the puzzle of ν\nu^{\prime} is α0α+(β0+σα0)β\alpha_{0}\alpha+(\beta_{0}+\sigma\alpha_{0})\beta. Therefore ρμ\rho\neq\mu^{\prime}, so μ′′0\mu^{\prime\prime}\neq 0 and thus root(μ)2{\rm root}(\mu^{\prime})\geq 2. Now, ν\nu^{\prime} is a degeneration of μ\mu^{\prime} and exit(μ)=exit(ν){\rm exit}(\mu^{\prime})={\rm exit}(\nu^{\prime}), so (5.2) implies

2root(μ)=root(ν)root(μ)=root(μ)root(ν)0.2-{\rm root}(\mu^{\prime})={\rm root}(\nu^{\prime})-{\rm root}(\mu^{\prime})={\rm root}(\mu^{\prime*})-{\rm root}(\nu^{\prime*})\geq 0.

We conclude that root(μ)=2{\rm root}(\mu^{\prime})=2, and thus μ′′=γμ′′′\mu^{\prime\prime}=\gamma\mu^{\prime\prime\prime} for some rigid tree measure μ′′′\mu^{\prime\prime\prime} and some γ>0\gamma>0. We conclude the proof by showing that μ′′′\mu^{\prime\prime\prime} has the same exit pattern as pn+1ν1+pn+2ν2p_{n+1}\nu_{1}+p_{n+2}\nu_{2} and thus μn+1=μ′′′\mu_{n+1}=\mu^{\prime\prime\prime} satisfies the conclusion of the theorem with nn replaced by n+1n+1. We start with a typical ray ee0e\neq e_{0} in the support of ν\nu^{\prime} and calculate the density that μ′′\mu^{\prime\prime} assigns to the outgoing ray in the inflation of ν\nu^{\prime}:

(pnα0+pn+1β0)(pnα+pn+1β)α0α(β0+σα0)β=cα+dβ.(p_{n}\alpha_{0}+p_{n+1}\beta_{0})(p_{n}\alpha+p_{n+1}\beta)-\alpha_{0}\alpha-(\beta_{0}+\sigma\alpha_{0})\beta=c\alpha+d\beta.

Here

c\displaystyle c =(pnα0+pn+1β0)pnα0\displaystyle=(p_{n}\alpha_{0}+p_{n+1}\beta_{0})p_{n}-\alpha_{0}
=(pn21)α0+pnpn+1β0\displaystyle=(p_{n}^{2}-1)\alpha_{0}+p_{n}p_{n+1}\beta_{0}
=pn1pn+1α0+pnpn+1β0=(pn1α0+pnβ0)pn+1,\displaystyle=p_{n-1}p_{n+1}\alpha_{0}+p_{n}p_{n+1}\beta_{0}=(p_{n-1}\alpha_{0}+p_{n}\beta_{0})p_{n+1},

where we used (5.8). Similarly, using (5.8) and (5.9) we obtain

d\displaystyle d =(pnα0+pn+1β0)pn+1β0σα0\displaystyle=(p_{n}\alpha_{0}+p_{n+1}\beta_{0})p_{n+1}-\beta_{0}-\sigma\alpha_{0}
=(pnpn+1σ)α0+(pn+121)β0\displaystyle=(p_{n}p_{n+1}-\sigma)\alpha_{0}+(p_{n+1}^{2}-1)\beta_{0}
=pn1pn+2α0+pnpn+2β0=(pn1α0+pnβ0)pn+2.\displaystyle=p_{n-1}p_{n+2}\alpha_{0}+p_{n}p_{n+2}\beta_{0}=(p_{n-1}\alpha_{0}+p_{n}\beta_{0})p_{n+2}.

Similar calculations apply to the outgoing ray in the inflation of e0e_{0}, while the outgoing ray in that inflation is assigned multiplicity 11=01-1=0 by μ′′\mu^{\prime\prime}. Thus, the exit pattern of μ′′\mu^{\prime\prime} is the same as that of (pn1α0+pnβ0)(pn+1ν1+pn+2ν2)(p_{n-1}\alpha_{0}+p_{n}\beta_{0})(p_{n+1}\nu_{1}+p_{n+2}\nu_{2}), and the desired conclusion follows along with the equality γ=pn1α0+pnβ0\gamma=p_{n-1}\alpha_{0}+p_{n}\beta_{0}. ∎

Remark 6.3.

The proof above would be slightly simpler if one could choose α0=0\alpha_{0}=0 and β0>0\beta_{0}>0. This however is not always possible because there are rigid overlays ν1+ν2\nu_{1}+\nu_{2} such that ν1\nu_{1} and ν2\nu_{2} assign positive multiplicity to precisely the same rays, as seen in the two examples in Figure 6.4 (the edges of ν2\nu_{2} outside the support of ν1\nu_{1} are drawn with dotted lines).

Refer to caption
Figure 6.4. Rigid overlays where the nonzero exit multiplicities of ν1\nu_{1} and ν2\nu_{2} correspond to the same rays
Example 6.4.

We illustrate the entire process in the proof of Theorem 6.2 for the overlay pictured in Figure 6.5. In this example, Σν1(ν2)=1\Sigma_{\nu_{1}}(\nu_{2})=1, and the red lines in the second part of the picture represent the support of the honeycomb μ1=ν^1\mu_{1}=\widehat{\nu}_{1}, compatible with the puzzle of ν2\nu_{2}.

Refer to caption
Figure 6.5. A simple overlay with Σν1(ν2)=1\Sigma_{\nu_{1}}(\nu_{2})=1

The inflation of μ1\mu_{1}, colored as in the proof above is shown in Figure 6.6. The support of μ\mu^{\prime} is colored blue and the black dot indicates the incoming ray ff. The second part of the picture represents the support of μ2\mu_{2} (in black), and the parts of the support of ρ\rho not covered by the support of μ2\mu_{2} (in blue). Since σ=1\sigma=1, the honeycomb μ2\mu_{2} has the exit pattern of p2ν1+p3ν2=ν1p_{2}\nu_{1}+p_{3}\nu_{2}=\nu_{1}, so μ2\mu_{2} is homologous to ν1\nu_{1}.

Refer to caption
Figure 6.6. The inflation of μ1\mu_{1}, the support of μ′′=μ~1\mu^{\prime\prime}=\widetilde{\mu}_{1}. Also, the support of μ′′\mu^{\prime\prime} (black) and the edges of ρ\rho not contained in the support of μ=μ2\mu^{\prime}=\mu_{2}
Corollary 6.5.

With the notation of the Theorem 6.2, for every nn\in\mathbb{N} there exists a rigid tree honeycomb μn\mu_{n}^{\prime} with the same exit pattern as pn+1ν1+pnν2p_{n+1}\nu_{1}+p_{n}\nu_{2}.

Proof.

Reflection in a line parallel to w1w_{1} changes the roles of ν1\nu_{1} and ν2\nu_{2}. Apply Theorem 6.2 to these reflected honeycombs to get a honeycomb μn\mu{}_{n}. Finally, reflect μn\mu{}_{n} in a line parallel to w1w_{1} to obtain μn\mu^{\prime}_{n}. ∎

Example 6.6.

The preceding corollary, applied to the first two overlays in Figure 6.7 shows that

n,n,n+1|n+1,n,n|n+1,n,n\displaystyle n,n,n+1|n+1,n,n|n+1,n,n
n+1,n+1,n|n,n+1,n+1|n,n+1,n+1\displaystyle n+1,n+1,n|n,n+1,n+1|n,n+1,n+1
n,n+1,n|n,n,n+1|n,n+1,n\displaystyle n,n+1,n|n,n,n+1|n,n+1,n
n+1,n,n+1|n+1,n+1,n|n+1,n,n+1\displaystyle n+1,n,n+1|n+1,n+1,n|n+1,n,n+1

are the exit patterns of rigid tree honeycombs. These four families were already described in [3, Section 8]. Similarly, the third overlay yields the patterns

n,n+1,n+1,n+1|n+1,n+1,n+1,n|2n+2,2n+1\displaystyle n,n+1,n+1,n+1|n+1,n+1,n+1,n|2n+2,2n+1
n+1,n,n,n|n,n,n,n+1|2n,2n+1.\displaystyle n+1,n,n,n|n,n,n,n+1|2n,2n+1.
Refer to caption
Figure 6.7. Three rigid overlays with Σν1(ν2)=2\Sigma_{\nu_{1}}(\nu_{2})=2

The overlays in Figure 6.8 satisfy Σν1(ν2)=3\Sigma_{\nu_{1}}(\nu_{2})=3.

Refer to caption
Figure 6.8. Two rigid overlays with Σν1(ν2)=3\Sigma_{\nu_{1}}(\nu_{2})=3

We deduce that

pn,pn+1,pn+1,pn+1|pn+1,pn+1,pn+1,pn|2pn+1,pn+1,pn\displaystyle p_{n},p_{n+1},p_{n+1},p_{n+1}|p_{n+1},p_{n+1},p_{n+1},p_{n}|2p_{n+1},p_{n+1},p_{n}
pn+1,pn,pn,pn|pn,pn,pn,pn+1|2pn,pn,pn+1\displaystyle p_{n+1},p_{n},p_{n},p_{n}|p_{n},p_{n},p_{n},p_{n+1}|2p_{n},p_{n},p_{n+1}
pn,pn+1,pn,pn+1|pn+1,pn+1,pn,pn|pn+pn+1,pn+1,pn\displaystyle p_{n},p_{n+1},p_{n},p_{n+1}|p_{n+1},p_{n+1},p_{n},p_{n}|p_{n}+p_{n+1},p_{n+1},p_{n}
pn+1,pn,pn+1,pn|pn,pn,pn+1,pn+1|pn+pn+1,pn,pn+1\displaystyle p_{n+1},p_{n},p_{n+1},p_{n}|p_{n},p_{n},p_{n+1},p_{n+1}|p_{n}+p_{n+1},p_{n},p_{n+1}

are the exit patterns of rigid tree honeycombs, where the sequence {pn}n=0={0,3,8,21,55,144,}\{p_{n}\}_{n=0}^{\infty}=\{0,3,8,21,55,144,\dots\} is defined by (5.7) with σ=3\sigma=3.

Remark 6.7.

The preceding two results (combined with Theorem 4.11 for σ=1\sigma=1) could be paraphrased as follows. Suppose that ν\nu is a rigid honeycomb with root(ν)=2{\rm root}(\nu)=2. If Σν(ν)=1\Sigma_{\nu}(\nu)=-1, then there exists a rigid tree honeycomb μ\mu that has the same exit pattern as ν\nu. One may wonder whether one can remove the restriction on root(ν).{\rm root}(\nu). The answer is negative. If ν1,ν2\nu_{1},\nu_{2}, and ν3\nu_{3} are three rigid honeycombs of weight 11 forming a clockwise overlay (see Figure 6.9),

Refer to captionν1\nu_{1}ν2\nu_{2}ν3\nu_{3}
Figure 6.9. Σνj(νi)=0\Sigma_{\nu_{j}}(\nu_{i})=0 for i<ji<j and Σνj(νi)=1\Sigma_{\nu_{j}}(\nu_{i})=1 for i>ji>j

then ν=c1ν1+c2ν2+c3ν3\nu=c_{1}\nu_{1}+c_{2}\nu_{2}+c_{3}\nu_{3}, c1,c2,c3,c_{1},c_{2},c_{3}\in\mathbb{N}, satisfies Σν(ν)=1\Sigma_{\nu}(\nu)=-1 precisely when.

c12+c22+c32c1c2c2c3c1c3=1.c_{1}^{2}+c_{2}^{2}+c_{3}^{2}-c_{1}c_{2}-c_{2}c_{3}-c_{1}c_{3}=1.

Rewriting this equation as

(c1c2)2+(c2c3)2+(c1c3)2=2,(c_{1}-c_{2})^{2}+(c_{2}-c_{3})^{2}+(c_{1}-c_{3})^{2}=2,

we see that, for each positive solution (c1,c2,c3)(c_{1},c_{2},c_{3}), two of the cjc_{j} must be equal and differ from the third by 11. Among the resulting solutions, it seen that nν1+(n+1)ν2+nν3n\nu_{1}+(n+1)\nu_{2}+n\nu_{3} and (n+1)ν1+nν2+(n+1)ν3(n+1)\nu_{1}+n\nu_{2}+(n+1)\nu_{3} are not the exit patterns of rigid tree honeycombs for any nn\in\mathbb{N}. The other solutions are the exit patterns of rigid tree honeycombs, as seen from Example 6.10 below.

Theorem 6.8.

Suppose that μ\mu is a rigid tree honeycomb and that its nonzero exit multiplicities, listed in counterclockwise order as α1,,αk\alpha_{1},\dots,\alpha_{k} are such that α1=σ>1\alpha_{1}=\sigma>1. Let {pn}n=0\{p_{n}\}_{n=0}^{\infty} be the sequence defined by (5.7). Then, for every n2n\geq 2, there exists a rigid tree honeycomb μn\mu_{n} such that:

  1. (1)

    The exit multiplicities of μn\mu_{n}, listed in counterclockwise order, are

    pn+1,pn1,pnα2,,pnαk.p_{n+1},p_{n-1},p_{n}\alpha_{2},\dots,p_{n}\alpha_{k}.

    In other words, μn\mu_{n} has the same exit pattern as pnμp_{n}\mu, except that pnα1p_{n}\alpha_{1} is replaced by pn+1p_{n+1} and pn1p_{n-1}.

  2. (2)

    μn\mu_{n} is compatible with the puzzle of pn1μp_{n-1}\mu.

Proof.

Let e0e_{0} be the ray in the support of μ\mu corresponding to the multiplicity α0\alpha_{0}. We prove the existence of μn\mu_{n} satisfying the following additional property: the rays in the support of μn\mu_{n} are:

  1. (i)

    the outgoing rays in the puzzle of pn1μp_{n-1}\mu, and

  2. (ii)

    one other ray contained in inflation of e0e_{0}.

The existence of μ2\mu_{2} follows from Theorem 4.2. In this case, the additional ray in (ii) is the incoming ray in the inflation of e0e_{0}. Suppose that μn\mu_{n} has been constructed for some n2n\geq 2. We apply the basic construction with pn1μp_{n-1}\mu in place of ν\nu and μn\mu_{n} in place of μ\mu and with the outgoing ray in the inflation of e0e_{0} in place of ff. The important observation is that (using notation from part (iii) of the construction, is that μ~\widetilde{\mu} has consecutive rays with multiplicities pn+121,p_{n+1}^{2}-1,11, and pn+1pn1p_{n+1}p_{n-1}, and that the rays with densities 11 and pn1pn+1p_{n-1}p_{n+1} are only separated by a dark gray strip (a translated part of the inflation of e0e_{0}). It follows that these three exit multiplicities collapse to two exit multiplicities pn+121p_{n+1}^{2}-1 and pn1pn+1+1p_{n-1}p_{n+1}+1 of μ\mu^{\prime}. Since

pn+121\displaystyle p_{n+1}^{2}-1 =pnpn+2,\displaystyle=p_{n}p_{n+2},
pn1pn+1+1\displaystyle p_{n-1}p_{n+1}+1 =pnpn,\displaystyle=p_{n}p_{n},

it suffices to show that μ\mu^{\prime} is extreme, in which case μn+1=μ/pn\mu_{n+1}=\mu^{\prime}/p_{n} satisfies the conclusion of the theorem with n+1n+1 in place of nn. To see this, we note that ν=pnμ\nu^{\prime}=p_{n}\mu and μν=pnpn+1μ\mu^{\prime}_{\nu^{\prime}}=p_{n}p_{n+1}\mu, so root(μν)=1{\rm root}(\mu^{\prime}_{\nu^{\prime}})=1 and exit(μν)=k{\rm exit}(\mu^{\prime}_{\nu^{\prime}})=k. An application of (5.2) shows that

1root(μ)\displaystyle 1-{\rm root}(\mu^{\prime}) =root(μν)root(μ)\displaystyle={\rm root}(\mu^{\prime}_{\nu^{\prime}})-{\rm root}(\mu^{\prime})
=[root(μ)root(μν)][exit(μ)exit(μν)]\displaystyle=[{\rm root}(\mu^{\prime*})-{\rm root}(\mu_{\nu^{\prime}}^{\prime*})]-[{\rm exit}(\mu^{\prime})-{\rm exit}(\mu^{\prime}_{\nu^{\prime}})]
=[root(μ)root(μν)]1.\displaystyle=[{\rm root}(\mu^{\prime*})-{\rm root}(\mu_{\nu^{\prime}}^{\prime*})]-1.

Since root(μ)1{\rm root}(\mu^{\prime})\geq 1 and root(μ)root(μν)>0{\rm root}(\mu^{\prime*})-{\rm root}(\mu_{\nu^{\prime}}^{\prime*})>0, we conclude that root(μ)=1{\rm root}(\mu^{\prime})=1, as desired. ∎

A reflection argument, like the one used in the proof of Corollary 6.2, immediately yields the following result.

Corollary 6.9.

Under the hypothesis of Theorem 6.8, for every n2n\geq 2 there exists a rigid tree honeycomb μn\mu_{n}^{\prime} such that the exit multiplicities of μn\mu_{n}^{\prime}, listed in counterclockwise order, are

pn1,pn+1,pnα2,,pnαk.p_{n-1},p_{n+1},p_{n}\alpha_{2},\dots,p_{n}\alpha_{k}.
Example 6.10.

An application of Theorem 6.8 and Corollary 6.9 to the rigid tree honeycomb with exit pattern 1,1,1|2,1|1,1,11,1,1|2,1|1,1,1 shows that

n,n,n|n+1,n1,n|n,n,n\displaystyle n,n,n|n+1,n-1,n|n,n,n
n,n,n|n1,n+1,n|n,n,n\displaystyle n,n,n|n-1,n+1,n|n,n,n

are the exit patterns of rigid tree honeycombs. These examples were first described in [3, Section 8]. Another rigid tree honeycomb has exit pattern 2,2,1|1,3,1|1,1,2,12,2,1|1,3,1|1,1,2,1. If we apply the results above with the first 22 in the role of α1\alpha_{1}, we deduce that

n1,n+1,2n,n|n,3n,n|n,n,2n,n\displaystyle n-1,n+1,2n,n|n,3n,n|n,n,2n,n
n+1,n1,2n,n|n,3n,n|n,n,2n,n\displaystyle n+1,n-1,2n,n|n,3n,n|n,n,2n,n

are the exit patterns of rigid tree measures. The tree honeycombs in the first series are represented in Figure 6.10

Refer to caption
Figure 6.10. Rigid tree honeycombs with exit pattern n1,n+1,2n,n|n,3n,n|n,n,2n,nn-1,n+1,2n,n|n,3n,n|n,n,2n,n, n=1,2,3n=1,2,3

for n=1,2,3n=1,2,3. It is easy to see how further honeycombs in this series are constructed. The results can also be applied with 33 in place of α1\alpha_{1} to generate the exit patterns

2pn,2pn,pn|pn,pn+1,pn1,pn|pn,pn,2pn,pn,\displaystyle 2p_{n},2p_{n},p_{n}|p_{n},p_{n+1},p_{n-1},p_{n}|p_{n},p_{n},2p_{n},p_{n},
2pn,2pn,pn|pn,pn1,pn+1,pn|pn,pn,2pn,pn,\displaystyle 2p_{n},2p_{n},p_{n}|p_{n},p_{n-1},p_{n+1},p_{n}|p_{n},p_{n},2p_{n},p_{n},

where the numbers pnp_{n} come from the sequence 0,3,8,21,0,3,8,21,\dots.

The following result is extracted from the proof of Theorem 6.8. The second part follows from the first using reflection.

Corollary 6.11.

Under the hypothesis of Theorem 6.8, for every n2n\geq 2 there exists a rigid tree honeycomb τn\tau_{n} such that the exit multiplicities of τn\tau_{n}, listed in counterclockwise order, are

pn+121,1,pn21,pnpn+1α2,,pnpn+1αk.p_{n+1}^{2}-1,1,p_{n}^{2}-1,p_{n}p_{n+1}\alpha_{2},\dots,p_{n}p_{n+1}\alpha_{k}.

Similarly, there exists a rigid tree measure τn\tau_{n}^{\prime} such that the exit multiplicities of τn\tau_{n}^{\prime}, listed in counterclockwise order, are

pn21,1,pn+121,pnα2,,pnαk.p_{n}^{2}-1,1,p_{n+1}^{2}-1,p_{n}\alpha_{2},\dots,p_{n}\alpha_{k}.

The following result shows that all the possibilities described in Theorem 5.4 actually arise.

Proposition 6.12.

Each of the honeycombs τn\tau_{n} and τn\tau^{\prime}_{n} of Corollary 6.11 has two simple degenerations that are extreme honeycombs.

Proof.

Fix n2n\geq 2 and consider the honeycomb τn\tau_{n}. We know from the above results that there exist rigid tree honeycombs μn\mu_{n} and μn+1\mu_{n+1} with exit multiplicities

pn+1,pn1,pnα2,,pnαkp_{n+1},p_{n-1},p_{n}\alpha_{2},\dots,p_{n}\alpha_{k}

and

pn+2,pn,pn+1α2,,pn+1αk,p_{n+2},p_{n},p_{n+1}\alpha_{2},\dots,p_{n+1}\alpha_{k},

respectively. Thus the extreme rigid honeycombs pn+1μnp_{n+1}\mu_{n} and pnμn+1p_{n}\mu_{n+1} have exit multiplicities

pn+12,pn+1pn1,pn+1pnα2,,pn+1pnαkp_{n+1}^{2},p_{n+1}p_{n-1},p_{n+1}p_{n}\alpha_{2},\dots,p_{n+1}p_{n}\alpha_{k}

and

pnpn+2,pn2,pn+1pnα2,,pn+1pnαk,p_{n}p_{n+2},p_{n}^{2},p_{n+1}p_{n}\alpha_{2},\dots,p_{n+1}p_{n}\alpha_{k},

respectively. Since pn+12=pn+121+1p_{n+1}^{2}=p_{n+1}^{2}-1+1, pn+1pn1=pn21p_{n+1}p_{n-1}=p_{n}^{2}-1, pnpn+2=pn+121,p_{n}p_{n+2}=p_{n+1}^{2}-1, and pn2=pn21+1p_{n}^{2}=p_{n}^{2}-1+1, we see immediately that pn+1μnp_{n+1}\mu_{n} and pnμn+1p_{n}\mu_{n+1} are simple degenerations of τn\tau_{n}. The case of τn\tau^{\prime}_{n} is treated similarly. ∎

Example 6.13.

The smallest illustration of Proposition 6.12 arises from the rigid tree honeycomb μ\mu with exit pattern 1,1,1|2,1|1,1,11,1,1|2,1|1,1,1. For n=2,n=2, we obtain the rigid tree honeycomb with exit pattern 6,6,6|3,1,8,6|6,6,66,6,6|3,1,8,6|6,6,6. Figure 6.11

Refer to caption
Figure 6.11. A rigid tree honeycomb with exit pattern 6,6,6|3,1,8,6|6,6,66,6,6|3,1,8,6|6,6,6 and two extreme summands of its dual that yield extreme simple degenerations

shows the support of τ2\tau_{2}^{*} along with the supports of the two extreme summands of τ2\tau_{2}^{*} whose elimination yields the two extreme simple degenerations of τ2\tau_{2}.

Remark 6.14.

The simple degenerations ν\nu of a rigid tree honeycomb fall into three categories:

  1. (G)

    σν1+ν2\sigma\nu_{1}+\nu_{2}, where ν1\nu_{1} and ν2\nu_{2} are distinct tree honeycombs and σ\sigma is a positive integer. According to Theorem 5.1, these simple degenerations are generic: at most three simple degenerations are not of this kind.

  2. (A)

    c1ν1+c2ν2c_{1}\nu_{1}+c_{2}\nu_{2}, where ν1\nu_{1} and ν2\nu_{2} are distinct tree honeycombs and c1,c22c_{1},c_{2}\geq 2 are integers.

  3. (B)

    cνc\nu, where ν\nu is a rigid tree measure and cc is a positive integer.

Thus, in addition to generic simple degenerations, a rigid tree measure might have

  1. (1)

    no other degenerations,

  2. (2)

    one degeneration of type (B),

  3. (3)

    one degeneration of type (A),

  4. (4)

    two degenerations of type (B),

  5. (5)

    two degenerations of type (A),

  6. (6)

    one degeneration of type (B) and one of type (A),

  7. (7)

    two degenerations of type (B) and one of type (A),

  8. (8)

    two degenerations of type (A) and one of type (B),

  9. (9)

    three degenerations of type (B),

  10. (10)

    three degenerations of type (A).

We have seen examples in the first four categories, and Theorem 5.4 shows that (7) and (9) are impossible. Extensive experimentation has not produced any rigid tree honeycombs in categories (5–10) and we conjecture that no such examples exist.

References

  • [1] C. Angiuli and H. Bercovici, The number of extremal components of a rigid measure, J. Combin. Theory Ser. A 118 (2011), 1925–1938.
  • [2] H. Bercovici, B. Collins, K. Dykema, W. S. Li, and D. Timotin, Intersections of Schubert varieties and eigenvalue inequalities in an arbitrary finite factor, J. Funct. Anal. 258 (2010), 1579–1627.
  • [3] H. Bercovici, W. S. Li, and D. Timotin, A family of reductions for Schubert intersection problems, J. Algebraic Combin. 33 (2011), 609–649.
  • [4] H. Bercovici, W. S. Li, and L. Truong, Extremal measures and clockwise overlays, Discrete Math. 315 (2014), 53–64.
  • [5] V. I. Danilov and G. A. Koshevoy, Discrete convexity and Hermitian matrices, Tr. Mat. Inst. Steklova 241 (2003), 68–89; translation in Proc. Steklov Inst. Math. 241 (2003), 58–78.
  • [6] H. Derksen and J. Weyman, The combinatorics of quiver representations, Ann. Inst. Fourier (Grenoble) 61 (2011), 1061–1131.
  • [7] R. C. King, C. Tollu, and F. Christophe, Factorisation of Littlewood-Richardson coefficients, J. Combin. Theory Ser. A 116 (2009), 314–333.
  • [8] A. Knutson and T. Tao, The honeycomb model of GL()n{}_{n}(\mathbb{C}) tensor products. I. Proof of the saturation conjecture, J. Amer. Math. Soc. 12 (1299), 1055–1090.
  • [9] A. Knutson, T. Tao, and C. Woodward, The honeycomb model of GL()n{}_{n}(\mathbb{C}) tensor products. II. Puzzles determine facets of the Littlewood-Richardson cone, J. Amer. Math. Soc. 17 (2004), 19–48.
  • [10] W. J. LeVeque, Topics in number theory, Vol. 1, Addison-Wesley Publishing Co., Inc., Reading, Mass., 1956.