License: CC BY 4.0
arXiv:2403.07663v1 [math.CO] 12 Mar 2024

Tilings of Benzels via Generalized Compression

Colin Defant Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA [email protected] Leigh Foster Department of Mathematics, University of Oregon, Eugene, OR 97403, USA [email protected] Rupert Li Massachusetts Institute of Technology, Cambridge, MA 02139, USA [email protected] James Propp Department of Mathematical Sciences, UMass Lowell, Lowell, MA 01854, USA [email protected]  and  Benjamin Young Department of Mathematics, University of Oregon, Eugene, OR 97403, USA [email protected]
Abstract.

Defant, Li, Propp, and Young recently resolved two enumerative conjectures of Propp concerning the tilings of regions in the hexagonal grid called benzels using two types of prototiles called stones and bones (with varying constraints on allowed orientations of the tiles). Their primary tool, a bijection called compression that converts certain k𝑘kitalic_k-ribbon tilings to (k1)𝑘1(k-1)( italic_k - 1 )-ribbon tilings, allowed them to reduce their problems to the enumeration of dimers (i.e., perfect matchings) of certain graphs. We present a generalized version of compression that no longer relies on the perspective of partitions and skew shapes. Using this strengthened tool, we resolve three more of Propp’s conjectures and recast several others as problems about perfect matchings.

1. Introduction

Enumeration of tilings of a region using translates of prescribed prototiles is a well-established topic in combinatorics; one notable example is the problem of counting domino tilings of the Aztec diamond in the square grid [6], which is the setting in which the arctic circle phenomenon was first discovered [3]. We refer readers to [12] for a survey of enumerative tiling methods and results. Conway and Lagarias [4] studied some tiling problems in the hexagonal grid, where both the regions to be tiled and the tiles themselves are composed of regular hexagons and are sometimes referred to as polyhexes. In their setting, the tiles are trihexes, which consist of three hexagons – specifically, these are the trihexes referred to below as stones and bones. In a novel application of combinatorial group theory, Conway and Lagarias devised a new necessary condition for a region to admit a tiling by stones and bones. Thurston [16] expanded upon these results with alternative perspectives, and Lagarias and Romano [9] proved an exact formula for the number of tilings in a particular one-parameter family of trihex tiling problems. Meanwhile, physicists working in an essentially equivalent (dual) setting studied trimer covers [17] of the regular 6-valent planar graph. However, the physicists’ interests were in asymptotics rather than exact enumerations, and their proofs relied on the Bethe ansatz, which has not been rigorously established in all the contexts where it has been applied (even though it tends to give correct answers).

Motivated by finding a trimer analogue of the Aztec diamond (and perhaps new kinds of arctic circle phenomena), Propp [13] proposed using the tiles studied by Conway and Lagarias to tile different sorts of regions in the hexagonal grid not considered in earlier literature, which he dubbed benzels. He made numerous conjectures regarding the exact number of tilings of those regions, and Defant, Li, Propp, and Young [5] resolved some of his conjectures. The key mechanism they developed is compression, which converts the problem from one involving 3-cell tiles to one involving 2-cell tiles. Tilings using 2-cell tiles are equivalent to perfect matchings of graphs, which have been well-studied and are amenable to a host of enumerative techniques.

We follow the conventions of [5], which expand upon the conventions of [13]. We view the hexagonal grid as a tiling of the complex plane with regular hexagons of side length 1, which we refer to as cells of the grid. Specifically, one of the hexagons has vertices 1±ωjplus-or-minus1superscript𝜔𝑗1\pm\omega^{j}1 ± italic_ω start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT for j{0,1,2}𝑗012j\in\{0,1,2\}italic_j ∈ { 0 , 1 , 2 }, where ω=e2πi/3𝜔superscript𝑒2𝜋𝑖3\omega=e^{2\pi i/3}italic_ω = italic_e start_POSTSUPERSCRIPT 2 italic_π italic_i / 3 end_POSTSUPERSCRIPT is a primitive third root of unity. Suppose a𝑎aitalic_a and b𝑏bitalic_b are positive integers satisfying 2a2b2𝑎2𝑏2\leq a\leq 2b2 ≤ italic_a ≤ 2 italic_b and 2b2a2𝑏2𝑎2\leq b\leq 2a2 ≤ italic_b ≤ 2 italic_a; we define the (a,b)𝑎𝑏(a,b)( italic_a , italic_b )-benzel to be the union of the cells lying entirely within the hexagon with vertices ωj(aω+b)superscript𝜔𝑗𝑎𝜔𝑏\omega^{j}(a\omega+b)italic_ω start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ( italic_a italic_ω + italic_b ) and ωj(a+bω)superscript𝜔𝑗𝑎𝑏𝜔-\omega^{j}(a+b\omega)- italic_ω start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ( italic_a + italic_b italic_ω ) for j{0,1,2}𝑗012j\in\{0,1,2\}italic_j ∈ { 0 , 1 , 2 }. This hexagon is invariant under rotation by 120{}^{\circ}start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPT and has sides whose lengths alternate between 2ab2𝑎𝑏2a-b2 italic_a - italic_b and 2ba2𝑏𝑎2b-a2 italic_b - italic_a (hence the inequalities a2b𝑎2𝑏a\leq 2bitalic_a ≤ 2 italic_b and b2a𝑏2𝑎b\leq 2aitalic_b ≤ 2 italic_a). In particular, the equality case corresponds to a triangle, a degenerate hexagon. As discussed in [5, Remark 5.1], the (n,2n2)𝑛2𝑛2(n,2n-2)( italic_n , 2 italic_n - 2 )-benzel, the (n,2n1)𝑛2𝑛1(n,2n-1)( italic_n , 2 italic_n - 1 )-benzel, and the (n,2n)𝑛2𝑛(n,2n)( italic_n , 2 italic_n )-benzel coincide because the extension of the boundary hexagon from the case (n,2n2)𝑛2𝑛2(n,2n-2)( italic_n , 2 italic_n - 2 ) to the case (n,2n)𝑛2𝑛(n,2n)( italic_n , 2 italic_n ) does not add any additional cells in their entirety. The same holds for the (2n2,n)2𝑛2𝑛(2n-2,n)( 2 italic_n - 2 , italic_n )-benzel, the (2n1,n)2𝑛1𝑛(2n-1,n)( 2 italic_n - 1 , italic_n )-benzel, and the (2n,n)2𝑛𝑛(2n,n)( 2 italic_n , italic_n )-benzel. Aside from these cases, the map sending the pair (a,b)𝑎𝑏(a,b)( italic_a , italic_b ) to the (a,b)𝑎𝑏(a,b)( italic_a , italic_b )-benzel is injective, so henceforth we will assume the stronger inequalities 2a2b22𝑎2𝑏22\leq a\leq 2b-22 ≤ italic_a ≤ 2 italic_b - 2 and 2b2a22𝑏2𝑎22\leq b\leq 2a-22 ≤ italic_b ≤ 2 italic_a - 2. Figure 1 shows the (9,11)911(9,11)( 9 , 11 )-benzel.

Refer to caption
Figure 1. The (9,11)911(9,11)( 9 , 11 )-benzel (shaded).

The prototiles in [4, 13] are special types of trihexes. These prototiles come in two forms: a stone consists of three pairwise adjacent cells arranged in a triangle, while a bone consists of three consecutive cells whose centers are colinear. Here, we follow the naming conventions of Propp [13]; Conway and Lagarias [4] and Thurston [16] used different terminology. It will be useful later to distinguish between the different rotations of these stones and bones. As a result, we consider the five prototiles in Figure 2 to be distinct, calling them the left stone, right stone, rising bone, falling bone, and vertical bone, respectively. Tilings of regions in the hexagonal grid using only these five prototiles will be referred to as stones-and-bones tilings.


Refer to caption
Figure 2. The five prototiles: the left stone, right stone, rising bone, falling bone, and vertical bone, respectively.

Tiling problems frequently restrict the set of possible prototiles. Using our five prototiles, there are 251=31superscript251312^{5}-1=312 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT - 1 = 31 nonempty subsets and, consequently, 31 enumeration questions that we can ask for each (a,b)𝑎𝑏(a,b)( italic_a , italic_b )-benzel. However, as benzels have 120superscript120120^{\circ}120 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT rotational symmetry that preserves the two stones, the exact set of allowed bones does not affect the number of stones-and-bones tilings; all that matters is the number of allowed types of bones. This equivalence reduces the number of problems to 2241=152241152\cdot 2\cdot 4-1=152 ⋅ 2 ⋅ 4 - 1 = 15. Propp [13] computed numerical results for these various problems and provided a collection of conjectures and open questions.

Note that the (a,b)𝑎𝑏(a,b)( italic_a , italic_b )-benzel and the (b,a)𝑏𝑎(b,a)( italic_b , italic_a )-benzel are reflections of each other across the real axis; as reflection preserves the two stones and permutes the three bones, each tiling problem is the same for the (a,b)𝑎𝑏(a,b)( italic_a , italic_b )-benzel as for the (b,a)𝑏𝑎(b,a)( italic_b , italic_a )-benzel. We will frequently use this symmetry to assume without loss of generality that ab𝑎𝑏a\leq bitalic_a ≤ italic_b.

As discussed in Section 2, our framework is best suited to enumeration problems forbidding at least one type of bone, so without loss of generality, the vertical bone is forbidden. Defant, Li, Propp, and Young [5, Theorem 1.2] enumerated the stones-and-bones tilings of benzels that use only left stones, rising bones, and falling bones. Such stones-and-bones tilings exist if and only if (a,b){(3n,3n),(3n+1,3n+2),(3n+2,3n+1)}𝑎𝑏3𝑛3𝑛3𝑛13𝑛23𝑛23𝑛1(a,b)\in\{(3n,3n),(3n+1,3n+2),(3n+2,3n+1)\}( italic_a , italic_b ) ∈ { ( 3 italic_n , 3 italic_n ) , ( 3 italic_n + 1 , 3 italic_n + 2 ) , ( 3 italic_n + 2 , 3 italic_n + 1 ) } for some positive integer n𝑛nitalic_n; in this case, the number of such tilings is (2n)!!double-factorial2𝑛(2n)!!( 2 italic_n ) !!.

Defant, Li, Propp, and Young [5, Theorem 1.3] also enumerated the stones-and-bones tilings of benzels that use only right stones, rising bones, and falling bones when a+b2(mod3)not-equivalent-to𝑎𝑏annotated2pmod3a+b\not\equiv 2\pmod{3}italic_a + italic_b ≢ 2 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER; in this case, at most 1 such tiling exists. They were unable to solve the problem when a+b2(mod3)𝑎𝑏annotated2pmod3{a+b\equiv 2\pmod{3}}italic_a + italic_b ≡ 2 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER; in this case, Propp [13, Problem 5] gave a conjectural formula. This conjectural formula is significantly more complicated than the solutions to the aforementioned enumeration problems, suggesting the difficulty of the problem. We prove this formula, stated in the following theorem.

Theorem 1.1.

Let a𝑎aitalic_a and b𝑏bitalic_b be integers with 2ab2a2𝑎𝑏2𝑎2\leq a\leq b\leq 2a2 ≤ italic_a ≤ italic_b ≤ 2 italic_a and a+b2(mod3)𝑎𝑏annotated2𝑝𝑚𝑜𝑑3a+b\equiv 2\pmod{3}italic_a + italic_b ≡ 2 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER. Let n=b+1a𝑛𝑏1𝑎n=b+1-aitalic_n = italic_b + 1 - italic_a and k=(2ab1)/3𝑘2𝑎𝑏13k=(2a-b-1)/3italic_k = ( 2 italic_a - italic_b - 1 ) / 3 so that (a,b)=(n+3k,2n+3k1)𝑎𝑏𝑛3𝑘2𝑛3𝑘1(a,b)=(n+3k,2n+3k-1)( italic_a , italic_b ) = ( italic_n + 3 italic_k , 2 italic_n + 3 italic_k - 1 ). Then the (a,b)𝑎𝑏(a,b)( italic_a , italic_b )-benzel has

j=1k(2j)!(2j+2n2)!(j+n1)!(j+n+k1)!superscriptsubscriptproduct𝑗1𝑘2𝑗2𝑗2𝑛2𝑗𝑛1𝑗𝑛𝑘1\displaystyle\prod_{j=1}^{k}\frac{(2j)!(2j+2n-2)!}{(j+n-1)!(j+n+k-1)!}∏ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT divide start_ARG ( 2 italic_j ) ! ( 2 italic_j + 2 italic_n - 2 ) ! end_ARG start_ARG ( italic_j + italic_n - 1 ) ! ( italic_j + italic_n + italic_k - 1 ) ! end_ARG

tilings by right stones, rising bones, and falling bones.

This result, combined with the results of Defant, Li, Propp, and Young [5], completely enumerates the stones-and-bones tilings of benzels that use only right stones, rising bones, and falling bones.

When a+b1(mod3)𝑎𝑏annotated1pmod3a+b\equiv 1\pmod{3}italic_a + italic_b ≡ 1 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER, [5, Theorem 1.1] implies that exactly one stones-and-bones tiling of the (a,b)𝑎𝑏(a,b)( italic_a , italic_b )-benzel exists, consisting entirely of right stones. However, aside from this case, nothing is known about the number of stones-and-bones tilings of benzels that use left stones, right stones, rising bones, and falling bones (i.e., all stones and bones except the vertical bone). Propp [13] stated a number of open questions concerning such enumerations (Problems 8 through 13), and we solve Problems 12 and 13 concerning the (n+2,2n+1)𝑛22𝑛1(n+2,2n+1)( italic_n + 2 , 2 italic_n + 1 )-benzel and the (n+2,2n)𝑛22𝑛(n+2,2n)( italic_n + 2 , 2 italic_n )-benzel in Propositions 3.3 and 3.4, respectively.

In Section 2, we follow the procedure of Defant, Li, Propp, and Young [5] to convert benzel tilings to ribbon-tilings (more specifically 3-ribbon tilings), which have been studied by others (see for instance [10]). In Section 3, we establish the technique of compression in greater generality than in [5]. In Section 3.1, as a warmup, we apply our more generalized compression technique, which converts 3-ribbon tilings into 2-ribbon tilings, to the (n+2,2n+1)𝑛22𝑛1(n+2,2n+1)( italic_n + 2 , 2 italic_n + 1 )-benzel and the (n+2,2n)𝑛22𝑛(n+2,2n)( italic_n + 2 , 2 italic_n )-benzel, solving Problems 12 and 13 of [13]. We use our compression technique to prove Theorem 1.1 in Section 4. Finally, in Section 5, we restate Problems 8 through 11 of [13] in their compressed forms and leave them as open questions, sharpening Problems 9 and 11 in the process.

2. Preliminaries

Following the conventions of [5], we draw the square grid in the complex plane, where each square grid cell has center of the form z=x+iy[i]𝑧𝑥𝑖𝑦delimited-[]𝑖z=x+iy\in\mathbb{Z}[i]italic_z = italic_x + italic_i italic_y ∈ blackboard_Z [ italic_i ] with x+y𝑥𝑦x+yitalic_x + italic_y odd, and the cell centered at z𝑧zitalic_z has vertices of the form z+ij𝑧superscript𝑖𝑗z+i^{j}italic_z + italic_i start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT for j{0,1,2,3}𝑗0123j\in\{0,1,2,3\}italic_j ∈ { 0 , 1 , 2 , 3 }. To avoid confusion, hexagonal cells will continue to be called cells, while square cells will be called boxes. We refer to a specific box by identifying it with its center. To leverage the existing literature for tilings of regions in the square grid, we follow the “squarification” procedure of [5] to convert tilings in the hexagonal grid to tilings in this square grid. Note that the unit-width vertical strips {z:n22Rez<n2}conditional-set𝑧𝑛22Re𝑧𝑛2\{z\in\mathbb{C}:\frac{n-2}{2}\leq\operatorname{Re}z<\frac{n}{2}\}{ italic_z ∈ blackboard_C : divide start_ARG italic_n - 2 end_ARG start_ARG 2 end_ARG ≤ roman_Re italic_z < divide start_ARG italic_n end_ARG start_ARG 2 end_ARG } for n3𝑛3n\in 3\mathbb{Z}italic_n ∈ 3 blackboard_Z are traversed only by horizontal edges of the hexagonal grid. See Figure 3 for an example of the squarification process applied to the (9,11)911(9,11)( 9 , 11 )-benzel; the horizontal edges, contained in the aforementioned unit-width vertical strips, are highlighted in green in the leftmost pane.

Refer to caption
Figure 3. The squarification process applied to the (9,11)-benzel.

Removing these strips and compressing the complex plane appropriately yields a bijection map** hexagons to rhombuses in the resulting rhombic grid; see the third pane from the left in Figure 3, where the second pane illustrates an intermediate state in this step. Rotating this grid by 90{}^{\circ}start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPT counterclockwise and rescaling the axes suitably transforms this rhombic grid to our aforementioned square grid. Boxes with the same real part form a column.

The four prototiles other than the vertical bone (which we have excluded) appear in Figure 2; these prototiles become the four prototiles in Figure 4 under squarification and rotation. Because of the rotation process, the names of the prototiles are no longer well-suited; to resolve this issue, the authors of [5] introduced new names for the prototiles in the rotated setting: we rename the right stone, left stone, rising bone, and falling bone the mountain stone, valley stone, negative bone, and positive bone, respectively. Recall that we excluded the vertical bone; this ensures that all squarified tiles are ribbon tiles, although our grid is rotated by 45superscript4545^{\circ}45 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT relative to the usual pictures that appear in the literature on ribbon tilings. Within our rotated square grid framework, a ribbon is a connected union of boxes in the square grid occupying consecutive columns, and a k𝑘kitalic_k-ribbon is a ribbon consisting of exactly k𝑘kitalic_k boxes. Our four remaining prototiles are precisely the four 3-ribbons. Tilings of regions in the square grid using ribbons are referred to as ribbon tilings, and ribbon tilings only using k𝑘kitalic_k-ribbons are k𝑘kitalic_k-ribbon tilings.

Refer to caption
Figure 4. The four squarified tiles after 90superscript9090^{\circ}90 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT rotation: the valley stone, mountain stone, negative bone, and positive bone.

An additional tool is the following factorization lemma, which we will use after applying the compression technique. Recall that compression converts 3-ribbon tilings into 2-ribbon tilings, which are in bijection with perfect matchings of a graph.

Lemma 2.1.

Let G𝐺Gitalic_G be a bipartite graph with vertex set V𝑉Vitalic_V, with vertices colored black and white in such a way that every edge of G𝐺Gitalic_G joins vertices of opposite colors. Suppose also that V=AB𝑉square-union𝐴𝐵V=A\sqcup Bitalic_V = italic_A ⊔ italic_B, where A𝐴Aitalic_A contains as many white vertices as black vertices and all edges joining A𝐴Aitalic_A and B𝐵Bitalic_B connect a white vertex in A𝐴Aitalic_A to a black vertex in B𝐵Bitalic_B. Then no perfect matching of G𝐺Gitalic_G contains an edge joining A𝐴Aitalic_A and B𝐵Bitalic_B, so the number of perfect matchings of G𝐺Gitalic_G equals the product of the number of perfect matchings of the induced subgraph on A𝐴Aitalic_A and the number of perfect matchings of the induced subgraph on B𝐵Bitalic_B.

Proof.

Every perfect matching of G𝐺Gitalic_G contains (up to) four kinds of edges, according to whether the white vertex lies in A𝐴Aitalic_A or B𝐵Bitalic_B and according to whether the black vertex lies in A𝐴Aitalic_A or B𝐵Bitalic_B. Since A𝐴Aitalic_A has as many white vertices as black vertices, the number of edges joining a white vertex in A𝐴Aitalic_A to a black vertex in B𝐵Bitalic_B must equal the number of edges joining a black vertex in A𝐴Aitalic_A to a white vertex in B𝐵Bitalic_B. But by assumption, the latter number is zero, so the former number must be zero as well. ∎

Lemma 2.1 is a special case of a more general result that applies with k𝑘kitalic_k color classes and k𝑘kitalic_k-ribbon tilings; for instance, the case k=3𝑘3k=3italic_k = 3 appears in [5], from which the generalization to other k𝑘kitalic_k is apparent.

3. Compression

Compression is a streamlined version of the combinatorial transformation that in [5] was described in terms of the abacus bijection of Gordon James. Under that earlier approach, we take the region we wish to tile by k𝑘kitalic_k-ribbons, view it as a Young diagram, and disassemble the diagram into a core and a quotient (a k𝑘kitalic_k-tuple of partitions, one of which turns out to be empty); we then remove the empty partition from the quotient and obtain a (k1)𝑘1(k-1)( italic_k - 1 )-quotient that can be used to assemble a new Young diagram whose (k1)𝑘1(k-1)( italic_k - 1 )-ribbon tilings are in bijection with k𝑘kitalic_k-ribbon tilings of the original Young diagram. Care is required when the core of the original Young diagram is nonempty.

Our new approach does not involve cores or quotients. Note that the squarification process of Figure 3 results in a row-convex and column-convex polyomino that has been rotated by 45superscript4545^{\circ}45 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT. Let S𝑆Sitalic_S denote such a rotated shape, which is to be tiled by k𝑘kitalic_k-ribbon tiles so that each k𝑘kitalic_k-ribbon consists of boxes from k𝑘kitalic_k consecutive columns. Let r𝑟ritalic_r be the number of columns of S𝑆Sitalic_S containing one or more boxes, and index these columns as 1,,r1𝑟1,\ldots,r1 , … , italic_r from left to right. Let misubscript𝑚𝑖m_{i}italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT denote the number of boxes in the i𝑖iitalic_i-th column of S𝑆Sitalic_S, and let M(x)=imixi𝑀𝑥subscript𝑖subscript𝑚𝑖superscript𝑥𝑖M(x)=\sum_{i}m_{i}x^{i}italic_M ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT be the corresponding generating function. For each k𝑘kitalic_k-ribbon tiling T𝑇Titalic_T of S𝑆Sitalic_S and for each i{1,,r}𝑖1𝑟i\in\{1,\ldots,r\}italic_i ∈ { 1 , … , italic_r }, let pi(T)subscript𝑝𝑖𝑇p_{i}(T)italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_T ) be the number of k𝑘kitalic_k-ribbon tiles in T𝑇Titalic_T whose leftmost box is in column i𝑖iitalic_i, and let P(x)=ipi(T)xi𝑃𝑥subscript𝑖subscript𝑝𝑖𝑇superscript𝑥𝑖P(x)=\sum_{i}p_{i}(T)x^{i}italic_P ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_T ) italic_x start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT be the corresponding generating function.

Every box in the i𝑖iitalic_i-th column belongs to a tile of T𝑇Titalic_T whose leftmost box belongs to some column between the (ik+1)𝑖𝑘1(i-k+1)( italic_i - italic_k + 1 )-th and the i𝑖iitalic_i-th, inclusive, so M(x)=(1+x++xk1)P(x)𝑀𝑥1𝑥superscript𝑥𝑘1𝑃𝑥M(x)=(1+x+\dots+x^{k-1})P(x)italic_M ( italic_x ) = ( 1 + italic_x + ⋯ + italic_x start_POSTSUPERSCRIPT italic_k - 1 end_POSTSUPERSCRIPT ) italic_P ( italic_x ). Since the polynomial M(x)𝑀𝑥M(x)italic_M ( italic_x ) does not depend on T𝑇Titalic_T, neither does P(x)𝑃𝑥P(x)italic_P ( italic_x ), so the coefficients pi(T)subscript𝑝𝑖𝑇p_{i}(T)italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_T ) do not depend on T𝑇Titalic_T either. Thus, we will henceforth write pisubscript𝑝𝑖p_{i}italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT instead of pi(T)subscript𝑝𝑖𝑇p_{i}(T)italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_T ).

Arguments similar to those employed in the preceding paragraph were originally proposed by Pak, though without the use of generating functions; see the paragraph that immediately follows the proof of Theorem 8.2 in [10]. The requirement that M(x)/(1+x++xk1)𝑀𝑥1𝑥superscript𝑥𝑘1M(x)/(1+x+\dots+x^{k-1})italic_M ( italic_x ) / ( 1 + italic_x + ⋯ + italic_x start_POSTSUPERSCRIPT italic_k - 1 end_POSTSUPERSCRIPT ) be a polynomial with nonnegative coefficients provides a criterion for tileability that is stronger than the usual coloring argument; for example, the region shown in Figure 5

Refer to caption
Figure 5. A color-balanced region that cannot be tiled by 2-ribbons.

has equal numbers of shaded and unshaded boxes, but since (2x1+1x2+1x3+2x4)/(1+x)2superscript𝑥11superscript𝑥21superscript𝑥32superscript𝑥41𝑥(2x^{1}+1x^{2}+1x^{3}+2x^{4})/(1+x)( 2 italic_x start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT + 1 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 1 italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + 2 italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) / ( 1 + italic_x ) equals 2x1x2+2x32superscript𝑥1superscript𝑥22superscript𝑥32x^{1}-x^{2}+2x^{3}2 italic_x start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (a polynomial with a negative coefficient), the region does not admit a 2-ribbon (i.e., domino) tiling.

Our shape S𝑆Sitalic_S is vertically convex so that all the boxes in the i𝑖iitalic_i-th column are corner-to-corner contiguous. If piksubscript𝑝𝑖𝑘p_{i-k}italic_p start_POSTSUBSCRIPT italic_i - italic_k end_POSTSUBSCRIPT and pisubscript𝑝𝑖p_{i}italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT both vanish, then there does not exist a k𝑘kitalic_k-ribbon tile whose rightmost box is in the (i1)𝑖1(i-1)( italic_i - 1 )-th column or whose leftmost box is in the i𝑖iitalic_i-th column. This implies that mi1=misubscript𝑚𝑖1subscript𝑚𝑖m_{i-1}=m_{i}italic_m start_POSTSUBSCRIPT italic_i - 1 end_POSTSUBSCRIPT = italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, but it also implies more: if T𝑇Titalic_T is a tiling of S𝑆Sitalic_S and 1jmi1𝑗subscript𝑚𝑖1\leq j\leq m_{i}1 ≤ italic_j ≤ italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, then the tile of T𝑇Titalic_T that contains the j𝑗jitalic_j-th box in column i1𝑖1i-1italic_i - 1 must also contain the j𝑗jitalic_j-th box in column i𝑖iitalic_i (where boxes in a column are indexed from top to bottom).

Proposition 3.1.

Let S𝑆Sitalic_S be a horizontally and vertically convex shape, with the quantities misubscript𝑚𝑖m_{i}italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and pisubscript𝑝𝑖p_{i}italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT as defined above. Suppose there is an integer i0subscript𝑖0i_{0}italic_i start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT such that pi=0subscript𝑝𝑖0p_{i}=0italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 for all ii0(modk)𝑖annotatedsubscript𝑖0𝑝𝑚𝑜𝑑𝑘i\equiv i_{0}\pmod{k}italic_i ≡ italic_i start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_MODIFIER ( roman_mod start_ARG italic_k end_ARG ) end_MODIFIER. Define a smaller (“compressed”) shape Ssuperscript𝑆normal-′S^{\prime}italic_S start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT by identifying the misubscript𝑚𝑖m_{i}italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT boxes in column i𝑖iitalic_i with the respective misubscript𝑚𝑖m_{i}italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT boxes in column i1𝑖1i-1italic_i - 1 from top to bottom, for every ii0(modk)𝑖annotatedsubscript𝑖0𝑝𝑚𝑜𝑑𝑘i\equiv i_{0}\pmod{k}italic_i ≡ italic_i start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_MODIFIER ( roman_mod start_ARG italic_k end_ARG ) end_MODIFIER. Then the k𝑘kitalic_k-ribbon tilings of S𝑆Sitalic_S are in bijection with the (k1)𝑘1(k-1)( italic_k - 1 )-ribbon tilings of Ssuperscript𝑆normal-′S^{\prime}italic_S start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT.

Proof.

Every k𝑘kitalic_k-ribbon tile must contain paired boxes from two paired columns, and by identifying the two boxes, we turn the k𝑘kitalic_k-ribbon tile into a (k1)𝑘1(k-1)( italic_k - 1 )-ribbon tile. ∎

Before returning to benzels, we illustrate the procedure with an example taken from an article by Chen and Kargin [1]. The left part of Figure 6 shows (a rotated version of) Figure 3 from their article. We have

M(x)=3x1+3x2+4x3+3x4+3x5+4x6+3x7+3x8+4x9+3x10+3x11,𝑀𝑥3superscript𝑥13superscript𝑥24superscript𝑥33superscript𝑥43superscript𝑥54superscript𝑥63superscript𝑥73superscript𝑥84superscript𝑥93superscript𝑥103superscript𝑥11M(x)=3x^{1}+3x^{2}+4x^{3}+3x^{4}+3x^{5}+4x^{6}+3x^{7}+3x^{8}+4x^{9}+3x^{10}+3x% ^{11},italic_M ( italic_x ) = 3 italic_x start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT + 3 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + 3 italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT + 3 italic_x start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT + 4 italic_x start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT + 3 italic_x start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT + 3 italic_x start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT + 4 italic_x start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT + 3 italic_x start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT + 3 italic_x start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT ,

and the polynomial

P(x)=3x1+1x3+2x4+2x6+1x7+3x9𝑃𝑥3superscript𝑥11superscript𝑥32superscript𝑥42superscript𝑥61superscript𝑥73superscript𝑥9P(x)=3x^{1}+1x^{3}+2x^{4}+2x^{6}+1x^{7}+3x^{9}italic_P ( italic_x ) = 3 italic_x start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT + 1 italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + 2 italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT + 2 italic_x start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT + 1 italic_x start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT + 3 italic_x start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT

does not have any terms with exponents congruent to 2 modulo 3333. Hence we may identify the boxes in columns 1 and 2, columns 4 and 5, columns 7 and 8, and columns 10 and 11 using the “sutures” that are shown, obtaining the region shown in the right part of Figure 6.

Refer to caption
Figure 6. Compressing a region into an Aztec diamond.

Therefore the 3-ribbon tilings of the region (S𝑆Sitalic_S) on the left are equinumerous with the 2-ribbon tilings of the region (Ssuperscript𝑆S^{\prime}italic_S start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT) on the right, which is the Aztec diamond of order 3. Indeed, [1, Theorem 2.5] can be proved using compression. (The bijection used in Chen and Kargin’s proof appears to be the same as ours, but we have not verified this.)

We now consider the region obtained from the (a,b)𝑎𝑏(a,b)( italic_a , italic_b )-benzel via contraction of horizontal edges, squarification, and rotation, as described in Section 2. When S𝑆Sitalic_S is this region obtained from the (a,b)𝑎𝑏(a,b)( italic_a , italic_b )-benzel, write the polynomials M(x)𝑀𝑥M(x)italic_M ( italic_x ) and P(x)𝑃𝑥P(x)italic_P ( italic_x ) as Ma,b(x)subscript𝑀𝑎𝑏𝑥M_{a,b}(x)italic_M start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT ( italic_x ) and Pa,b(x)subscript𝑃𝑎𝑏𝑥P_{a,b}(x)italic_P start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT ( italic_x ), respectively. Then Ma,b(x)=(1+x+x2)Pa,b(x)subscript𝑀𝑎𝑏𝑥1𝑥superscript𝑥2subscript𝑃𝑎𝑏𝑥M_{a,b}(x)=(1+x+x^{2})P_{a,b}(x)italic_M start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT ( italic_x ) = ( 1 + italic_x + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_P start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT ( italic_x ), where the polynomial Pa,b(x)subscript𝑃𝑎𝑏𝑥P_{a,b}(x)italic_P start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT ( italic_x ) has a structure that depends in a simple way on the parameters a𝑎aitalic_a and b𝑏bitalic_b, with different behavior according to a+b𝑎𝑏a+bitalic_a + italic_b (mod 3). For 1ik1𝑖𝑘1\leq i\leq k1 ≤ italic_i ≤ italic_k, let

Qi,k(x)=m=0i1(m+1)xm+m=ik1ixm=1x0+2x1+3x2++ixi1+ixi++ixk1subscript𝑄𝑖𝑘𝑥superscriptsubscript𝑚0𝑖1𝑚1superscript𝑥𝑚superscriptsubscript𝑚𝑖𝑘1𝑖superscript𝑥𝑚1superscript𝑥02superscript𝑥13superscript𝑥2𝑖superscript𝑥𝑖1𝑖superscript𝑥𝑖𝑖superscript𝑥𝑘1Q_{i,k}(x)=\sum_{m=0}^{i-1}(m+1)x^{m}+\sum_{m=i}^{k-1}ix^{m}=1x^{0}+2x^{1}+3x^% {2}+\dots+ix^{i-1}+ix^{i}+\dots+ix^{k-1}italic_Q start_POSTSUBSCRIPT italic_i , italic_k end_POSTSUBSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_m = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i - 1 end_POSTSUPERSCRIPT ( italic_m + 1 ) italic_x start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_m = italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k - 1 end_POSTSUPERSCRIPT italic_i italic_x start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT = 1 italic_x start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT + 2 italic_x start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT + 3 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ⋯ + italic_i italic_x start_POSTSUPERSCRIPT italic_i - 1 end_POSTSUPERSCRIPT + italic_i italic_x start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT + ⋯ + italic_i italic_x start_POSTSUPERSCRIPT italic_k - 1 end_POSTSUPERSCRIPT

and let Ri,k(x)=xk1Qi,k(1/x)subscript𝑅𝑖𝑘𝑥superscript𝑥𝑘1subscript𝑄𝑖𝑘1𝑥R_{i,k}(x)=x^{k-1}Q_{i,k}(1/x)italic_R start_POSTSUBSCRIPT italic_i , italic_k end_POSTSUBSCRIPT ( italic_x ) = italic_x start_POSTSUPERSCRIPT italic_k - 1 end_POSTSUPERSCRIPT italic_Q start_POSTSUBSCRIPT italic_i , italic_k end_POSTSUBSCRIPT ( 1 / italic_x ) be the polynomial of degree k1𝑘1k-1italic_k - 1 obtained from Qi,k(x)subscript𝑄𝑖𝑘𝑥Q_{i,k}(x)italic_Q start_POSTSUBSCRIPT italic_i , italic_k end_POSTSUBSCRIPT ( italic_x ) by reversing the order of the coefficients. For example, we have

Q3,5(x)=1+2x+3x2+3x3+3x4andR3,5(x)=3+3x+3x2+2x3+x4.formulae-sequencesubscript𝑄35𝑥12𝑥3superscript𝑥23superscript𝑥33superscript𝑥4andsubscript𝑅35𝑥33𝑥3superscript𝑥22superscript𝑥3superscript𝑥4Q_{3,5}(x)=1+2x+3x^{2}+3x^{3}+3x^{4}\quad\text{and}\quad R_{3,5}(x)=3+3x+3x^{2% }+2x^{3}+x^{4}.italic_Q start_POSTSUBSCRIPT 3 , 5 end_POSTSUBSCRIPT ( italic_x ) = 1 + 2 italic_x + 3 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 3 italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + 3 italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT and italic_R start_POSTSUBSCRIPT 3 , 5 end_POSTSUBSCRIPT ( italic_x ) = 3 + 3 italic_x + 3 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT .

Also, for 1i,jkformulae-sequence1𝑖𝑗𝑘1\leq i,j\leq k1 ≤ italic_i , italic_j ≤ italic_k, let

Ti,j,k(x)subscript𝑇𝑖𝑗𝑘𝑥\displaystyle T_{i,j,k}(x)italic_T start_POSTSUBSCRIPT italic_i , italic_j , italic_k end_POSTSUBSCRIPT ( italic_x ) =m=0ki(i+m)xm+m=ki+12kij(2kim)xmabsentsuperscriptsubscript𝑚0𝑘𝑖𝑖𝑚superscript𝑥𝑚superscriptsubscript𝑚𝑘𝑖12𝑘𝑖𝑗2𝑘𝑖𝑚superscript𝑥𝑚\displaystyle=\sum_{m=0}^{k-i}(i+m)x^{m}+\sum_{m=k-i+1}^{2k-i-j}(2k-i-m)x^{m}= ∑ start_POSTSUBSCRIPT italic_m = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k - italic_i end_POSTSUPERSCRIPT ( italic_i + italic_m ) italic_x start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_m = italic_k - italic_i + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_k - italic_i - italic_j end_POSTSUPERSCRIPT ( 2 italic_k - italic_i - italic_m ) italic_x start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT
=ix0+(i+1)x1++(k1)xki1+kxki+(k1)xki+1++jx2kij.absent𝑖superscript𝑥0𝑖1superscript𝑥1𝑘1superscript𝑥𝑘𝑖1𝑘superscript𝑥𝑘𝑖𝑘1superscript𝑥𝑘𝑖1𝑗superscript𝑥2𝑘𝑖𝑗\displaystyle=ix^{0}+(i+1)x^{1}+\dots+(k-1)x^{k-i-1}+kx^{k-i}+(k-1)x^{k-i+1}+% \dots+jx^{2k-i-j}.= italic_i italic_x start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT + ( italic_i + 1 ) italic_x start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT + ⋯ + ( italic_k - 1 ) italic_x start_POSTSUPERSCRIPT italic_k - italic_i - 1 end_POSTSUPERSCRIPT + italic_k italic_x start_POSTSUPERSCRIPT italic_k - italic_i end_POSTSUPERSCRIPT + ( italic_k - 1 ) italic_x start_POSTSUPERSCRIPT italic_k - italic_i + 1 end_POSTSUPERSCRIPT + ⋯ + italic_j italic_x start_POSTSUPERSCRIPT 2 italic_k - italic_i - italic_j end_POSTSUPERSCRIPT .

For example,

T1,2,4(x)=1+2x+3x2+4x3+3x4+2x5.subscript𝑇124𝑥12𝑥3superscript𝑥24superscript𝑥33superscript𝑥42superscript𝑥5T_{1,2,4}(x)=1+2x+3x^{2}+4x^{3}+3x^{4}+2x^{5}.italic_T start_POSTSUBSCRIPT 1 , 2 , 4 end_POSTSUBSCRIPT ( italic_x ) = 1 + 2 italic_x + 3 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + 3 italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT + 2 italic_x start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT .
Lemma 3.2.

Fix integers a,b1𝑎𝑏1a,b\geq 1italic_a , italic_b ≥ 1 with a2b𝑎2𝑏a\leq 2bitalic_a ≤ 2 italic_b and b2a𝑏2𝑎b\leq 2aitalic_b ≤ 2 italic_a. Let γ𝛾\gammaitalic_γ be the unique element of {1,2,3}123\{1,2,3\}{ 1 , 2 , 3 } such that a+bγ(mod3)𝑎𝑏annotated𝛾𝑝𝑚𝑜𝑑3a+b\equiv\gamma\pmod{3}italic_a + italic_b ≡ italic_γ start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER, and let s=(2ab3+γ)/3𝑠2𝑎𝑏3𝛾3s=(2a-b-3+\gamma)/3italic_s = ( 2 italic_a - italic_b - 3 + italic_γ ) / 3 and t=(2ba3+γ)/3𝑡2𝑏𝑎3𝛾3t=(2b-a-3+\gamma)/3italic_t = ( 2 italic_b - italic_a - 3 + italic_γ ) / 3. Then

Pa,b(x)={x3Ts+1,t+1,(a+b1)/3(x3)if γ=1,xRs,(a+b2)/3(x3)+x2Qt,(a+b2)/3(x3)if γ=2,xRs,(a+b3)/3(x3)+x3Qt,(a+b3)/3(x3)if γ=3.subscript𝑃𝑎𝑏𝑥casessuperscript𝑥3subscript𝑇𝑠1𝑡1𝑎𝑏13superscript𝑥3if 𝛾1𝑥subscript𝑅𝑠𝑎𝑏23superscript𝑥3superscript𝑥2subscript𝑄𝑡𝑎𝑏23superscript𝑥3if 𝛾2𝑥subscript𝑅𝑠𝑎𝑏33superscript𝑥3superscript𝑥3subscript𝑄𝑡𝑎𝑏33superscript𝑥3if 𝛾3P_{a,b}(x)=\begin{cases}x^{3}T_{s+1,t+1,(a+b-1)/3}(x^{3})&\text{if }\gamma=1,% \\ xR_{s,(a+b-2)/3}(x^{3})+x^{2}Q_{t,(a+b-2)/3}(x^{3})&\text{if }\gamma=2,\\ xR_{s,(a+b-3)/3}(x^{3})+x^{3}Q_{t,(a+b-3)/3}(x^{3})&\text{if }\gamma=3.\end{cases}italic_P start_POSTSUBSCRIPT italic_a , italic_b end_POSTSUBSCRIPT ( italic_x ) = { start_ROW start_CELL italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT italic_s + 1 , italic_t + 1 , ( italic_a + italic_b - 1 ) / 3 end_POSTSUBSCRIPT ( italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) end_CELL start_CELL if italic_γ = 1 , end_CELL end_ROW start_ROW start_CELL italic_x italic_R start_POSTSUBSCRIPT italic_s , ( italic_a + italic_b - 2 ) / 3 end_POSTSUBSCRIPT ( italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_Q start_POSTSUBSCRIPT italic_t , ( italic_a + italic_b - 2 ) / 3 end_POSTSUBSCRIPT ( italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) end_CELL start_CELL if italic_γ = 2 , end_CELL end_ROW start_ROW start_CELL italic_x italic_R start_POSTSUBSCRIPT italic_s , ( italic_a + italic_b - 3 ) / 3 end_POSTSUBSCRIPT ( italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) + italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_Q start_POSTSUBSCRIPT italic_t , ( italic_a + italic_b - 3 ) / 3 end_POSTSUBSCRIPT ( italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) end_CELL start_CELL if italic_γ = 3 . end_CELL end_ROW

(The lemma contains more information than is required for this article; in order to apply compression, all we need to know is the existence of suitable polynomials Q𝑄Qitalic_Q, R𝑅Ritalic_R, and T𝑇Titalic_T, not their precise forms. However, we include the formulas here as they might be useful to future researchers.)

Proof.

Let us assume γ=3𝛾3\gamma=3italic_γ = 3; a similar argument handles the other two cases. Recall (see [8]) that the boundary of the (a,b)𝑎𝑏(a,b)( italic_a , italic_b )-benzel is given by the word

(𝐛,𝐚,𝐛,𝐜)s(𝐜,𝐚,𝐛,𝐚)t(𝐜,𝐛,𝐜,𝐚)s(𝐚,𝐛,𝐜,𝐛)t(𝐚,𝐜,𝐚,𝐛)s(𝐛,𝐜,𝐚,𝐜)tsuperscript𝐛superscript𝐚𝐛superscript𝐜𝑠superscript𝐜superscript𝐚𝐛superscript𝐚𝑡superscript𝐜superscript𝐛𝐜superscript𝐚𝑠superscript𝐚superscript𝐛𝐜superscript𝐛𝑡superscript𝐚superscript𝐜𝐚superscript𝐛𝑠superscript𝐛superscript𝐜𝐚superscript𝐜𝑡({\rm\bf{b}},{\rm\bf{a}}^{\prime},{\rm\bf{b}},{\rm\bf{c}}^{\prime})^{s}({\rm% \bf{c}},{\rm\bf{a}}^{\prime},{\rm\bf{b}},{\rm\bf{a}}^{\prime})^{t}({\rm\bf{c}}% ,{\rm\bf{b}}^{\prime},{\rm\bf{c}},{\rm\bf{a}}^{\prime})^{s}({\rm\bf{a}},{\rm% \bf{b}}^{\prime},{\rm\bf{c}},{\rm\bf{b}}^{\prime})^{t}({\rm\bf{a}},{\rm\bf{c}}% ^{\prime},{\rm\bf{a}},{\rm\bf{b}}^{\prime})^{s}({\rm\bf{b}},{\rm\bf{c}}^{% \prime},{\rm\bf{a}},{\rm\bf{c}}^{\prime})^{t}( bold_b , bold_a start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , bold_b , bold_c start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT ( bold_c , bold_a start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , bold_b , bold_a start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT ( bold_c , bold_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , bold_c , bold_a start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT ( bold_a , bold_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , bold_c , bold_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT ( bold_a , bold_c start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , bold_a , bold_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT ( bold_b , bold_c start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , bold_a , bold_c start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT

where 𝐚,𝐛,𝐜,𝐚𝐛𝐜𝐚𝐛𝐜superscript𝐚superscript𝐛superscript𝐜{\rm\bf{a}},{\rm\bf{b}},{\rm\bf{c}},{\rm\bf{a}}^{\prime}{\rm\bf{b}}^{\prime}{% \rm\bf{c}}^{\prime}bold_a , bold_b , bold_c , bold_a start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT bold_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT bold_c start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT are the unit vectors pointing from 00 to 1,ω,ω2,1,ω,ω21𝜔superscript𝜔21𝜔superscript𝜔21,\omega,\omega^{2},-1,-\omega,-\omega^{2}1 , italic_ω , italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , - 1 , - italic_ω , - italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, respectively (where ω𝜔\omegaitalic_ω is a primitive 3rd root of unity as earlier), with s=(2ab)/3𝑠2𝑎𝑏3s=(2a-b)/3italic_s = ( 2 italic_a - italic_b ) / 3 and t=(2ba)/3𝑡2𝑏𝑎3t=(2b-a)/3italic_t = ( 2 italic_b - italic_a ) / 3. When we apply the compression, squarification, and rotation operations described in Section 2, two of these vectors shrink away and the remaining four change, but the combinatorial description remains the same, so the method used to prove [8, Theorem 1] applies here as well. ∎

We are mostly interested in the case in which a+b2(mod3)𝑎𝑏annotated2pmod3a+b\equiv 2\pmod{3}italic_a + italic_b ≡ 2 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER, but remarks on the other two cases are in order.

When a+b1(mod3)𝑎𝑏annotated1pmod3a+b\equiv 1\pmod{3}italic_a + italic_b ≡ 1 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER, the coefficient of xisuperscript𝑥𝑖x^{i}italic_x start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT in P(x)𝑃𝑥P(x)italic_P ( italic_x ) vanishes for every i1,2(mod3)𝑖1annotated2pmod3i\equiv 1,2\pmod{3}italic_i ≡ 1 , 2 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER. In this case, the bijection in Proposition 3.1 can be used twice to let us put 3-ribbon tilings of the region in question in bijection with 1-ribbon tilings of a reduced region, but every region has exactly one 1-ribbon tiling (and in fact the uncompressed tiling is composed exclusively of right stones).

The case in which a+b0(mod3)𝑎𝑏annotated0pmod3a+b\equiv 0\pmod{3}italic_a + italic_b ≡ 0 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER is subtler. Here the coefficient of xisuperscript𝑥𝑖x^{i}italic_x start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT in P(x)𝑃𝑥P(x)italic_P ( italic_x ) vanishes whenever i2𝑖2i\equiv 2italic_i ≡ 2 (mod 3), so we can apply compression (once) to put 3-ribbon tilings of the region in question in bijection with 2-ribbon tilings of a reduced region. These 2-ribbon tilings correspond to perfect matchings of graphs like the one shown in Figure 7. Here we set s=(2ab)/3𝑠2𝑎𝑏3s=(2a-b)/3italic_s = ( 2 italic_a - italic_b ) / 3 and t=(2ba)/3𝑡2𝑏𝑎3t=(2b-a)/3italic_t = ( 2 italic_b - italic_a ) / 3. (We have flipped the graph across a horizontal axis in Figure 7 so as to be more consistent with Figure 8, the analog for the case a+b2(mod3)𝑎𝑏annotated2pmod3a+b\equiv 2\pmod{3}italic_a + italic_b ≡ 2 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER; the flip of course has no effect on the number of perfect matchings.) Edges colored teal correspond to right/mountain stones, edges colored yellow correspond to left/valley stones, and black edges correspond to bones. Thus, forbidding right (respectively, left) stones in the stones-and-bones tiling of an (a,b)𝑎𝑏(a,b)( italic_a , italic_b )-benzel corresponds to forbidding teal (respectively, yellow) edges in the associated perfect matching.

Refer to caption
Figure 7. The graph dual to the compressed (8,10)810(8,10)( 8 , 10 )-benzel. In this example, we have s=2𝑠2s=2italic_s = 2 and t=4𝑡4t=4italic_t = 4.

Finally, we have the case in which a+b2(mod3)𝑎𝑏annotated2pmod3a+b\equiv 2\pmod{3}italic_a + italic_b ≡ 2 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER, which is relevant to Theorem 1.1. Here, the coefficient of xisuperscript𝑥𝑖x^{i}italic_x start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT in P(x)𝑃𝑥P(x)italic_P ( italic_x ) vanishes whenever i0𝑖0i\equiv 0italic_i ≡ 0 (mod 3), so we can apply compression (once) to put 3-ribbon tilings of the region in question in bijection with 2-ribbon tilings of a reduced region. These 2-ribbon tilings correspond to perfect matchings of graphs like the one shown in Figure 8, where the “sutures” of this compression process, along with an example 3-ribbon tiling that compresses to a 2-ribbon tiling, are shown in Figure 9. Here, we set s=(2ab1)/3𝑠2𝑎𝑏13s=(2a-b-1)/3italic_s = ( 2 italic_a - italic_b - 1 ) / 3 and t=(2ba1)/3𝑡2𝑏𝑎13t=(2b-a-1)/3italic_t = ( 2 italic_b - italic_a - 1 ) / 3. As in the previous case, edges colored teal correspond to right stones while edges colored yellow correspond to left stones.

Refer to caption
Figure 8. The graph dual to the compressed (8,9)89(8,9)( 8 , 9 )-benzel. In this example, we have s=2𝑠2s=2italic_s = 2 and t=3𝑡3t=3italic_t = 3.
Refer to caption
Refer to caption
Figure 9. Compressing the (8,9)89(8,9)( 8 , 9 )-benzel with an example 3-ribbon tiling to obtain a 2-ribbon tiling, or equivalently a perfect matching, below.

3.1. Two examples

We now provide two simple examples of compression, which yield proofs of two conjectures of Propp [13, Problems 12 and 13]. For both examples, we consider tilings of benzels using both stones and two types of bones; we assume without loss of generality that the vertical bone is forbidden. Since both left and right stones are permitted, these tilings after compression correspond to perfect matchings of Figure 7 that use edges yellow and teal edges as well as black edges.

The following result, which resolves Problem 12 of [13], states that the number of tilings of the (n+2,2n+1)𝑛22𝑛1(n+2,2n+1)( italic_n + 2 , 2 italic_n + 1 )-benzel by left stones, right stones, rising bones, and falling bones is the n𝑛nitalic_n-th large Schröder number (see [14, Sequence A006318]).

Proposition 3.3.

Let Tnsubscript𝑇𝑛T_{n}italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT denote the number of tilings of the (n+2,2n+1)𝑛22𝑛1(n+2,2n+1)( italic_n + 2 , 2 italic_n + 1 )-benzel by left stones, right stones, rising bones, and falling bones (with the convention T0=1subscript𝑇01T_{0}=1italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1). Then

(1) n=0Tnxn=1x16x+x22x.superscriptsubscript𝑛0subscript𝑇𝑛superscript𝑥𝑛1𝑥16𝑥superscript𝑥22𝑥\sum_{n=0}^{\infty}T_{n}x^{n}=\frac{1-x-\sqrt{1-6x+x^{2}}}{2x}.∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = divide start_ARG 1 - italic_x - square-root start_ARG 1 - 6 italic_x + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG start_ARG 2 italic_x end_ARG .
Proof.

Using the characterization of benzels with a+b0(mod3)𝑎𝑏annotated0pmod3a+b\equiv 0\pmod{3}italic_a + italic_b ≡ 0 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER from Section 3, we have s=1𝑠1s=1italic_s = 1 and t=n𝑡𝑛t=nitalic_t = italic_n. Using these parameters in the schematic of Figure 7, we see by applying compression that our stones-and-bones tilings (with the vertical bone forbidden) of the (n+2,2n+1)𝑛22𝑛1(n+2,2n+1)( italic_n + 2 , 2 italic_n + 1 )-benzel correspond to perfect matchings of the Aztec triangle of order n𝑛nitalic_n; see Figure 10 for the case n=4𝑛4n=4italic_n = 4, identical to [2, Figure 13]. Ciucu [2, Theorem 4.1] proved that the number of such perfect matchings is the n𝑛nitalic_n-th large Schröder number, which completes the proof. ∎

Refer to caption
Figure 10. The Aztec triangle of order 4.

Next we build on Proposition 3.3 to solve Problem 13 of [13]. The relevant enumeration sequence is sequence A006319 in [14].

Proposition 3.4.

Let Tnsuperscriptsubscript𝑇𝑛normal-′T_{n}^{\prime}italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT denote the number of tilings of the (n+2,2n)𝑛22𝑛(n+2,2n)( italic_n + 2 , 2 italic_n )-benzel by left stones, right stones, rising bones, and falling bones (with the convention T0=1superscriptsubscript𝑇0normal-′1T_{0}^{\prime}=1italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 1). Then

(2) n=0Tnxn=(1x)1x16x+x22x.superscriptsubscript𝑛0superscriptsubscript𝑇𝑛superscript𝑥𝑛1𝑥1𝑥16𝑥superscript𝑥22𝑥\sum_{n=0}^{\infty}T_{n}^{\prime}x^{n}=(1-x)\frac{1-x-\sqrt{1-6x+x^{2}}}{2x}.∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = ( 1 - italic_x ) divide start_ARG 1 - italic_x - square-root start_ARG 1 - 6 italic_x + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG start_ARG 2 italic_x end_ARG .
Proof.

Comparing (2) with (1), we see that it suffices to prove that Tn=TnTn1superscriptsubscript𝑇𝑛subscript𝑇𝑛subscript𝑇𝑛1T_{n}^{\prime}=T_{n}-T_{n-1}italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_T start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT. Using the characterization of benzels with a+b2(mod3)𝑎𝑏annotated2pmod3a+b\equiv 2\pmod{3}italic_a + italic_b ≡ 2 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER from Section 3, we have s=1𝑠1s=1italic_s = 1 and t=n1𝑡𝑛1t=n-1italic_t = italic_n - 1. Using these parameters in the schematic of Figure 8, we see by applying compression that our stones-and-bones tilings (with the vertical bone forbidden) of the (n+2,2n)𝑛22𝑛(n+2,2n)( italic_n + 2 , 2 italic_n )-benzel correspond to perfect matchings of a shape obtained by removing one “square” from the bottom corner of the Aztec triangle of order n𝑛nitalic_n (that is, removing the two lower-left-most vertices and the three edges they participate in). See the left pane of Figure 11 for the case n=4𝑛4n=4italic_n = 4.

Refer to caption
Figure 11. The graph dual to the compressed (6,8)68(6,8)( 6 , 8 )-benzel and its relation to the Aztec triangle of order 4.

Adding this square back in amounts to adding two new vertices u𝑢uitalic_u and v𝑣vitalic_v, so we see that perfect matchings of this shape correspond to perfect matchings of the Aztec triangle of order n𝑛nitalic_n that have these two vertices matched (i.e., perfect matchings that use the blue dotted edge (u,v)𝑢𝑣(u,v)( italic_u , italic_v ) in Figure 11). Since the augmented graph (the Aztec triangle of order n𝑛nitalic_n) has Tnsubscript𝑇𝑛T_{n}italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT perfect matchings, it suffices to show there are Tn1subscript𝑇𝑛1T_{n-1}italic_T start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT perfect matchings of the Aztec triangle of order n𝑛nitalic_n that do not use the edge (u,v)𝑢𝑣(u,v)( italic_u , italic_v ). After removing pairs of vertices that must be matched to each other (shaded in teal in Figure 11), the resulting shape (drawn in green in Figure 11) is the Aztec triangle of order n1𝑛1n-1italic_n - 1. Thus, there are indeed Tn1subscript𝑇𝑛1T_{n-1}italic_T start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT such perfect matchings. ∎

4. Decomposition

We now prove Theorem 1.1. Let us fix integers k0𝑘0k\geq 0italic_k ≥ 0 and n1𝑛1n\geq 1italic_n ≥ 1 with (k,n)(0,1)𝑘𝑛01(k,n)\neq(0,1)( italic_k , italic_n ) ≠ ( 0 , 1 ), and let (a,b)=(n+3k,2n+3k1)𝑎𝑏𝑛3𝑘2𝑛3𝑘1(a,b)=(n+3k,2n+3k-1)( italic_a , italic_b ) = ( italic_n + 3 italic_k , 2 italic_n + 3 italic_k - 1 ). As a+b2(mod3)𝑎𝑏annotated2pmod3a+b\equiv 2\pmod{3}italic_a + italic_b ≡ 2 start_MODIFIER ( roman_mod start_ARG 3 end_ARG ) end_MODIFIER, the parameters s𝑠sitalic_s and t𝑡titalic_t from Section 3 are s=(2ab1)/3=k𝑠2𝑎𝑏13𝑘s=(2a-b-1)/3=kitalic_s = ( 2 italic_a - italic_b - 1 ) / 3 = italic_k and t=(2ba1)/3=n+k1𝑡2𝑏𝑎13𝑛𝑘1t=(2b-a-1)/3=n+k-1italic_t = ( 2 italic_b - italic_a - 1 ) / 3 = italic_n + italic_k - 1. As we are only allowed to use right stones, rising bones, and falling bones, compression converts our tiling problem into a perfect matching problem of the form of Figure 8, where the edges corresponding to left stones in Figure 8 (drawn in yellow) are forbidden; see Figure 12 for a visual example of the resulting graph after removing these edges. There are a number of forced matches, as marked in green in Figure 12. After removing these vertices, we obtain a graph as in Figure 13.

Refer to caption
Figure 12. The graph dual to the compressed (14,18)1418(14,18)( 14 , 18 )-benzel, with edges corresponding to left stones removed. Forced matches are marked in green.
Refer to caption
Figure 13. The graph dual to the compressed (14,18)1418(14,18)( 14 , 18 )-benzel, with forced matches removed. A bipartition of the vertices is indicated via a black-and-white coloring. A pink dotted line separates the graph into two pieces as in Lemma 2.1.

Consider the pink dotted line in Figure 13. By deleting the edges that cross this dotted line, we break the graph into two connected components. Let A𝐴Aitalic_A be the connected component on the southwest side of the dotted line and B𝐵Bitalic_B be the connected component on the northeast side; see Figure 14, where A𝐴Aitalic_A appears on the left and B𝐵Bitalic_B appears on the right.

Refer to caption
Refer to caption
Figure 14. The two pieces of the compressed (14,18)1418(14,18)( 14 , 18 )-benzel, with black-and-white bipartitions of the vertices. In the piece on the left, pink dotted lines separate the graph into three Λsans-serif-Λ\mathsf{\Lambda}sansserif_Λ-shaped pieces to which we can apply Lemma 2.1.

We have colored the vertices black and white in a checkerboard fashion, so that each edge is incident to one black and one white vertex. Note that A𝐴Aitalic_A contains as many white vertices as black vertices, and likewise for B𝐵Bitalic_B. All edges joining A𝐴Aitalic_A and B𝐵Bitalic_B (i.e., edges that cross the dotted line) connect a white vertex in A𝐴Aitalic_A to a black vertex in B𝐵Bitalic_B. Thus, by Lemma 2.1, the number of perfect matchings of the full graph is the product of the number of perfect matchings of A𝐴Aitalic_A and the number of perfect matchings of B𝐵Bitalic_B.

We first enumerate the perfect matchings of A𝐴Aitalic_A. One can observe that vertically reflecting A𝐴Aitalic_A yields the same graph as that obtained by applying compression to the shape λssubscript𝜆𝑠\lambda_{s}italic_λ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT from [5], where right stones, i.e., mountain stones, are forbidden; this vertical reflection corresponds to our reflection of Figure 7 to align with Figure 8. See [5, Figure 11] for a visual aid for this compression process (the compression amounts to removing the green columns in that figure). Working in the uncompressed environment, Defant, Li, Propp, and Young [5, Proposition 6.2] proved that the number of perfect matchings of A𝐴Aitalic_A is (2s)!!=(2s)(2s2)(4)(2)=2ss!double-factorial2𝑠2𝑠2𝑠242superscript2𝑠𝑠{(2s)!!=(2s)(2s-2)\cdots(4)(2)=2^{s}s!}( 2 italic_s ) !! = ( 2 italic_s ) ( 2 italic_s - 2 ) ⋯ ( 4 ) ( 2 ) = 2 start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT italic_s !. Let us briefly sketch the proof of this enumerative result, but in our compressed environment. We can apply Lemma 2.1 to split A𝐴Aitalic_A along the dotted lines in Figure 14, just as we split the original graph along the dotted line in Figure 13. (As noted after Lemma 2.1, this is where, working in the uncompressed environment, the article [5] used a 3-color variant of Lemma 2.1.) It then suffices to show that the i𝑖iitalic_i-th smallest of these “Λsans-serif-Λ\mathsf{\Lambda}sansserif_Λ-shapes” (the article [5] used the term “𝖵𝖵\mathsf{V}sansserif_V-shape” in the reflected setting) has 2i2𝑖2i2 italic_i perfect matchings. It is straightforward to see that exactly one of the 2i2𝑖2i2 italic_i edges that are “perpendicular to the Λsans-serif-Λ\mathsf{\Lambda}sansserif_Λ” (i.e., the 2i22𝑖22i-22 italic_i - 2 edges in the interior and the 2 edges at the two tips of the Λsans-serif-Λ\mathsf{\Lambda}sansserif_Λ-shape) must be used in a perfect matching, and this choice determines the rest of the perfect matching.

Finally, recalling that s=k𝑠𝑘s=kitalic_s = italic_k and dividing (2k)!!double-factorial2𝑘(2k)!!( 2 italic_k ) !! from the expression in Theorem 1.1, we find that it remains to show that the number of perfect matchings of B𝐵Bitalic_B is the quantity

(3) Jk,n=j=1k(2j1)!(2j+2n2)!(j+n1)!(j+n+k1)!.subscript𝐽𝑘𝑛superscriptsubscriptproduct𝑗1𝑘2𝑗12𝑗2𝑛2𝑗𝑛1𝑗𝑛𝑘1J_{k,n}=\prod_{j=1}^{k}\frac{(2j-1)!(2j+2n-2)!}{(j+n-1)!(j+n+k-1)!}.italic_J start_POSTSUBSCRIPT italic_k , italic_n end_POSTSUBSCRIPT = ∏ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT divide start_ARG ( 2 italic_j - 1 ) ! ( 2 italic_j + 2 italic_n - 2 ) ! end_ARG start_ARG ( italic_j + italic_n - 1 ) ! ( italic_j + italic_n + italic_k - 1 ) ! end_ARG .

We can deform the “brickwork” pattern of B𝐵Bitalic_B into a hexagonal grid. There is a well-known bijection between perfect matchings of this hexagonal grid and plane partitions; see [18] for a discussion of this correspondence from brickwork to hexagonal grid to plane partitions. Specifically, perfect matchings of B𝐵Bitalic_B correspond to plane partitions of the staircase shape

(ts,ts1,,1)=(n1,n2,,1)𝑡𝑠𝑡𝑠11𝑛1𝑛21(t-s,t-s-1,\dots,1)=(n-1,n-2,\dots,1)( italic_t - italic_s , italic_t - italic_s - 1 , … , 1 ) = ( italic_n - 1 , italic_n - 2 , … , 1 )

with parts no larger than s=k𝑠𝑘s=kitalic_s = italic_k.

Refer to caption
Figure 15. The B𝐵Bitalic_B component of the (14,18)1418(14,18)( 14 , 18 )-benzel, deformed into a hexagonal grid with its maximal matching, and the corresponding plane partition.

Figure 15 shows an example of this correspondence for the (14,18)1418(14,18)( 14 , 18 )-benzel, the same example as in Figure 14; on the left, we show the hexagonal grid obtained by deforming B𝐵Bitalic_B, along with an example matching, in the middle we show how this matching becomes a plane partition, and on the right we show this plane partition by itself. This matching is the maximal plane partition, which contains the plane partitions corresponding to all of the other matchings. In our example of the (14,18)1418(14,18)( 14 , 18 )-benzel, where s=3𝑠3s=3italic_s = 3 and t=7𝑡7t=7italic_t = 7, this maximal plane partition is the staircase shape (4,3,2,1)4321(4,3,2,1)( 4 , 3 , 2 , 1 ) with parts all of size 3. Proctor [11, Corollary 4.1] found that the number of such plane partitions is the quantity

(4) Kk,n=i=1n1(k+iij=2i2k+i+j1i+j1)subscript𝐾𝑘𝑛superscriptsubscriptproduct𝑖1𝑛1𝑘𝑖𝑖superscriptsubscriptproduct𝑗2𝑖2𝑘𝑖𝑗1𝑖𝑗1\displaystyle K_{k,n}=\prod_{i=1}^{n-1}\left(\frac{k+i}{i}\prod_{j=2}^{i}\frac% {2k+i+j-1}{i+j-1}\right)italic_K start_POSTSUBSCRIPT italic_k , italic_n end_POSTSUBSCRIPT = ∏ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT ( divide start_ARG italic_k + italic_i end_ARG start_ARG italic_i end_ARG ∏ start_POSTSUBSCRIPT italic_j = 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT divide start_ARG 2 italic_k + italic_i + italic_j - 1 end_ARG start_ARG italic_i + italic_j - 1 end_ARG ) =1i<jn2k+i+j1i+j1absentsubscriptproduct1𝑖𝑗𝑛2𝑘𝑖𝑗1𝑖𝑗1\displaystyle=\prod_{1\leq i<j\leq n}\frac{2k+i+j-1}{i+j-1}= ∏ start_POSTSUBSCRIPT 1 ≤ italic_i < italic_j ≤ italic_n end_POSTSUBSCRIPT divide start_ARG 2 italic_k + italic_i + italic_j - 1 end_ARG start_ARG italic_i + italic_j - 1 end_ARG

(see also [15, Exercise 7.101a]). To prove that the quantities in (3) and (4) are equal, one can induct on k𝑘kitalic_k. Note that they are trivially equal for k=0𝑘0k=0italic_k = 0. Moreover,

Jk,nJk1,n=(2k1)!(2k+2n2)!(n+2k1)!(n+2k2)!,subscript𝐽𝑘𝑛subscript𝐽𝑘1𝑛2𝑘12𝑘2𝑛2𝑛2𝑘1𝑛2𝑘2\displaystyle\frac{J_{k,n}}{J_{k-1,n}}=\frac{(2k-1)!(2k+2n-2)!}{(n+2k-1)!(n+2k% -2)!},divide start_ARG italic_J start_POSTSUBSCRIPT italic_k , italic_n end_POSTSUBSCRIPT end_ARG start_ARG italic_J start_POSTSUBSCRIPT italic_k - 1 , italic_n end_POSTSUBSCRIPT end_ARG = divide start_ARG ( 2 italic_k - 1 ) ! ( 2 italic_k + 2 italic_n - 2 ) ! end_ARG start_ARG ( italic_n + 2 italic_k - 1 ) ! ( italic_n + 2 italic_k - 2 ) ! end_ARG ,

while

Kk,nKk1,nsubscript𝐾𝑘𝑛subscript𝐾𝑘1𝑛\displaystyle\frac{K_{k,n}}{K_{k-1,n}}divide start_ARG italic_K start_POSTSUBSCRIPT italic_k , italic_n end_POSTSUBSCRIPT end_ARG start_ARG italic_K start_POSTSUBSCRIPT italic_k - 1 , italic_n end_POSTSUBSCRIPT end_ARG =1i<jn2k+i+j12k+i+j3absentsubscriptproduct1𝑖𝑗𝑛2𝑘𝑖𝑗12𝑘𝑖𝑗3\displaystyle=\prod_{1\leq i<j\leq n}\frac{2k+i+j-1}{2k+i+j-3}= ∏ start_POSTSUBSCRIPT 1 ≤ italic_i < italic_j ≤ italic_n end_POSTSUBSCRIPT divide start_ARG 2 italic_k + italic_i + italic_j - 1 end_ARG start_ARG 2 italic_k + italic_i + italic_j - 3 end_ARG
=i=1n1(2k+i+n1)!(2k+2i3)!(2k+2i1)!(2k+i+n3)!absentsuperscriptsubscriptproduct𝑖1𝑛12𝑘𝑖𝑛12𝑘2𝑖32𝑘2𝑖12𝑘𝑖𝑛3\displaystyle=\prod_{i=1}^{n-1}\frac{(2k+i+n-1)!(2k+2i-3)!}{(2k+2i-1)!(2k+i+n-% 3)!}= ∏ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT divide start_ARG ( 2 italic_k + italic_i + italic_n - 1 ) ! ( 2 italic_k + 2 italic_i - 3 ) ! end_ARG start_ARG ( 2 italic_k + 2 italic_i - 1 ) ! ( 2 italic_k + italic_i + italic_n - 3 ) ! end_ARG
=i=1n1(2k+i+n1)(2k+i+n2)(2k+2i1)(2k+2i2)absentsuperscriptsubscriptproduct𝑖1𝑛12𝑘𝑖𝑛12𝑘𝑖𝑛22𝑘2𝑖12𝑘2𝑖2\displaystyle=\prod_{i=1}^{n-1}\frac{(2k+i+n-1)(2k+i+n-2)}{(2k+2i-1)(2k+2i-2)}= ∏ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT divide start_ARG ( 2 italic_k + italic_i + italic_n - 1 ) ( 2 italic_k + italic_i + italic_n - 2 ) end_ARG start_ARG ( 2 italic_k + 2 italic_i - 1 ) ( 2 italic_k + 2 italic_i - 2 ) end_ARG
=(2k+2n2)!(2k+2n3)!(2k1)!(2k+n1)!(2k+n2)!(2k+2n3)!absent2𝑘2𝑛22𝑘2𝑛32𝑘12𝑘𝑛12𝑘𝑛22𝑘2𝑛3\displaystyle=\frac{(2k+2n-2)!(2k+2n-3)!(2k-1)!}{(2k+n-1)!(2k+n-2)!(2k+2n-3)!}= divide start_ARG ( 2 italic_k + 2 italic_n - 2 ) ! ( 2 italic_k + 2 italic_n - 3 ) ! ( 2 italic_k - 1 ) ! end_ARG start_ARG ( 2 italic_k + italic_n - 1 ) ! ( 2 italic_k + italic_n - 2 ) ! ( 2 italic_k + 2 italic_n - 3 ) ! end_ARG
=(2k1)!(2k+2n2)!(n+2k1)!(n+2k2)!.absent2𝑘12𝑘2𝑛2𝑛2𝑘1𝑛2𝑘2\displaystyle=\frac{(2k-1)!(2k+2n-2)!}{(n+2k-1)!(n+2k-2)!}.= divide start_ARG ( 2 italic_k - 1 ) ! ( 2 italic_k + 2 italic_n - 2 ) ! end_ARG start_ARG ( italic_n + 2 italic_k - 1 ) ! ( italic_n + 2 italic_k - 2 ) ! end_ARG .

This completes the proof by induction and concludes the proof of Theorem 1.1.

5. Conclusion and Open Problems

Propp [13, Problems 8 to 13] stated a number of open questions concerning the number of stones-and-bones tilings of benzels that use all stones and bones except the vertical bone. Propositions 3.3 and 3.4 addressed Problems 12 and 13. Using our compression technique, we restate Problems 8 to 11 as perfect matching problems and leave them as open questions, ho** that their simplified form will facilitate further progress with these problems.

Problem 8 concerns the (3n,3n)3𝑛3𝑛(3n,3n)( 3 italic_n , 3 italic_n )-benzel. Referencing Figure 7, we find the graph dual of the compressed (3n,3n)3𝑛3𝑛(3n,3n)( 3 italic_n , 3 italic_n )-benzel is as depicted in Figure 16, which shows the case n=3𝑛3n=3italic_n = 3.

Refer to caption
Figure 16. The graph dual to the (9,9)99(9,9)( 9 , 9 )-benzel.
Conjecture 5.1 ([13, Problem 8]).

Let Tnsubscript𝑇𝑛T_{n}italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT denote the number of stones-and-bones tilings of the (3n,3n)3𝑛3𝑛(3n,3n)( 3 italic_n , 3 italic_n )-benzel in which the vertical bone is forbidden. We have

TnTn+2Tn+12=256(2n+3)2(4n+1)(4n+3)2(4n+5)27(3n+1)(3n+2)2(3n+4)2(3n+5)subscript𝑇𝑛subscript𝑇𝑛2superscriptsubscript𝑇𝑛12256superscript2𝑛324𝑛1superscript4𝑛324𝑛5273𝑛1superscript3𝑛22superscript3𝑛423𝑛5\frac{T_{n}T_{n+2}}{T_{n+1}^{2}}=\frac{256(2n+3)^{2}(4n+1)(4n+3)^{2}(4n+5)}{27% (3n+1)(3n+2)^{2}(3n+4)^{2}(3n+5)}divide start_ARG italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_n + 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = divide start_ARG 256 ( 2 italic_n + 3 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 4 italic_n + 1 ) ( 4 italic_n + 3 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 4 italic_n + 5 ) end_ARG start_ARG 27 ( 3 italic_n + 1 ) ( 3 italic_n + 2 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 3 italic_n + 4 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 3 italic_n + 5 ) end_ARG

for all n1𝑛1n\geq 1italic_n ≥ 1.

Problem 9 concerns the (3n+1,3n+1)3𝑛13𝑛1(3n+1,3n+1)( 3 italic_n + 1 , 3 italic_n + 1 )-benzel. Referencing Figure 8, we find the graph dual of the compressed (3n+1,3n+1)3𝑛13𝑛1(3n+1,3n+1)( 3 italic_n + 1 , 3 italic_n + 1 )-benzel is as depicted in Figure 17, which shows the case n=3𝑛3n=3italic_n = 3.

Refer to caption
Figure 17. The graph dual to the (10,10)1010(10,10)( 10 , 10 )-benzel.

Propp [13] empirically observed that for small n𝑛nitalic_n, the number of such stones-and-bones tilings had no prime factor greater than or equal to 4n4𝑛4n4 italic_n, suggesting the enumeration could have a nice factored form. However, obtaining data for large values of n𝑛nitalic_n was difficult. Knowing now how to recast the problem in terms of dimers, we were able to use existing technology for enumerating perfect matchings (specifically the determinant method of Kasteleyn [7]) to obtain more data. This allows us to supplant Propp’s original Problem 9 with the following more precise conjecture.

Conjecture 5.2 (c.f. [13, Problem 9]).

Let Tnsubscript𝑇𝑛T_{n}italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT denote the number of stones-and-bones tilings of the (3n+1,3n+1)3𝑛13𝑛1(3n+1,3n+1)( 3 italic_n + 1 , 3 italic_n + 1 )-benzel in which the vertical bone is forbidden. We have

TnTn1=22n(4n1)!(4n2)!n!(3n)!(3n1)!(3n2)!subscript𝑇𝑛subscript𝑇𝑛1superscript22𝑛4𝑛14𝑛2𝑛3𝑛3𝑛13𝑛2\frac{T_{n}}{T_{n-1}}=\frac{2^{2n}(4n-1)!(4n-2)!n!}{(3n)!(3n-1)!(3n-2)!}divide start_ARG italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT end_ARG = divide start_ARG 2 start_POSTSUPERSCRIPT 2 italic_n end_POSTSUPERSCRIPT ( 4 italic_n - 1 ) ! ( 4 italic_n - 2 ) ! italic_n ! end_ARG start_ARG ( 3 italic_n ) ! ( 3 italic_n - 1 ) ! ( 3 italic_n - 2 ) ! end_ARG

for all n2𝑛2n\geq 2italic_n ≥ 2.

Problem 10 yields the pattern depicted in Figure 18, which shows the case n=3𝑛3n=3italic_n = 3. Similarly to Problem 8, it also has a conjectured formula for the second quotient.

Refer to caption
Figure 18. The graph dual to the (10,11)1011(10,11)( 10 , 11 )-benzel.
Conjecture 5.3 ([13, Problem 10]).

Let Tnsubscript𝑇𝑛T_{n}italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT denote the number of stones-and-bones tilings of the (3n+1,3n+2)3𝑛13𝑛2(3n+1,3n+2)( 3 italic_n + 1 , 3 italic_n + 2 )-benzel where the vertical bone is forbidden. We have

TnTn+2Tn+12=65536(2n+3)(2n+5)2(2n+7)(4n+3)(4n+5)2(4n+7)2(4n+9)2(4n+11)729(3n+2)(3n+4)3(3n+5)2(3n+7)2(3n+8)3(3n+10)subscript𝑇𝑛subscript𝑇𝑛2superscriptsubscript𝑇𝑛12655362𝑛3superscript2𝑛522𝑛74𝑛3superscript4𝑛52superscript4𝑛72superscript4𝑛924𝑛117293𝑛2superscript3𝑛43superscript3𝑛52superscript3𝑛72superscript3𝑛833𝑛10\frac{T_{n}T_{n+2}}{T_{n+1}^{2}}=\frac{65536(2n+3)(2n+5)^{2}(2n+7)(4n+3)(4n+5)% ^{2}(4n+7)^{2}(4n+9)^{2}(4n+11)}{729(3n+2)(3n+4)^{3}(3n+5)^{2}(3n+7)^{2}(3n+8)% ^{3}(3n+10)}divide start_ARG italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_n + 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = divide start_ARG 65536 ( 2 italic_n + 3 ) ( 2 italic_n + 5 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 2 italic_n + 7 ) ( 4 italic_n + 3 ) ( 4 italic_n + 5 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 4 italic_n + 7 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 4 italic_n + 9 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 4 italic_n + 11 ) end_ARG start_ARG 729 ( 3 italic_n + 2 ) ( 3 italic_n + 4 ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( 3 italic_n + 5 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 3 italic_n + 7 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 3 italic_n + 8 ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( 3 italic_n + 10 ) end_ARG

for all n1𝑛1n\geq 1italic_n ≥ 1.

Problem 11 yields the pattern depicted in Figure 19, which shows the case n=3𝑛3n=3italic_n = 3. As for Problem 9, Propp [13] empirically observed that for small n𝑛nitalic_n, the number of such stones-and-bones tilings had no prime factor greater than or equal to 4n4𝑛4n4 italic_n, suggesting the enumeration could have a nice factored form. As was the case for Problem 9, recasting Problem 11 as a question about perfect matchings allowed us to obtain a conjectural formula.

Refer to caption
Figure 19. The graph dual to the (8,9)89(8,9)( 8 , 9 )-benzel.
Question 5.4 ([13, Problem 11]).

Let Tnsubscript𝑇𝑛T_{n}italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT denote the number of stones-and-bones tilings of the (3n1,3n)3𝑛13𝑛(3n-1,3n)( 3 italic_n - 1 , 3 italic_n )-benzel in which the vertical bone is forbidden. We have

TnTn1=22n3(4n2)!(4n4)!(n1)!!(n3)!!(3n1)!!(3n2)!(3n3)!(3n5)!!subscript𝑇𝑛subscript𝑇𝑛1superscript22𝑛34𝑛24𝑛4double-factorial𝑛1double-factorial𝑛3double-factorial3𝑛13𝑛23𝑛3double-factorial3𝑛5\frac{T_{n}}{T_{n-1}}=\frac{2^{2n-3}(4n-2)!(4n-4)!(n-1)!!(n-3)!!}{(3n-1)!!(3n-% 2)!(3n-3)!(3n-5)!!}divide start_ARG italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT end_ARG = divide start_ARG 2 start_POSTSUPERSCRIPT 2 italic_n - 3 end_POSTSUPERSCRIPT ( 4 italic_n - 2 ) ! ( 4 italic_n - 4 ) ! ( italic_n - 1 ) !! ( italic_n - 3 ) !! end_ARG start_ARG ( 3 italic_n - 1 ) !! ( 3 italic_n - 2 ) ! ( 3 italic_n - 3 ) ! ( 3 italic_n - 5 ) !! end_ARG

for all n2𝑛2n\geq 2italic_n ≥ 2.

There are probably more applications of the central idea behind compression. Consider (as a trivial example) the problem of using straight n𝑛nitalic_n-ominoes (that is, 1-by-n𝑛nitalic_n rectangles) to tile an (n+1)𝑛1(n+1)( italic_n + 1 )-by-(n+1)𝑛1(n+1)( italic_n + 1 ) square from which the central (n1)𝑛1(n-1)( italic_n - 1 )-by-(n1)𝑛1(n-1)( italic_n - 1 ) square has been removed. One way to prove that there are only two tilings is to note that the n1𝑛1n-1italic_n - 1 middle boxes on each side of the big square must belong to the same n𝑛nitalic_n-omino and so can be compressed into a single box, reducing the problem to that of using dominoes to tile a 3-by-3 square from which the central box has been removed. In this case, the compression can be achieved geometrically, but all that is required for purposes of enumeration is that it can be done topologically.

More broadly, given a tiling problem in which certain unions of cells are permitted as tiles, we can look at all nonempty regions that can arise as the intersection of two tiles that actually occur in tilings of the entire region we are trying to tile. In some cases, these will just be the individual cells; in other cases, however, these “pseudocells” will be unions of two or more cells, and in this case it may be helpful to imagine identifying all the cells that belong to a common pseudocell. As can be seen in [5] and the current article, determining when two cells must always be occupied by the same tile in any tiling of a large region can involve non-local arguments that are sensitive to the shape of the boundary of the large region.

Acknowledgements

Colin Defant was supported by the National Science Foundation under Award No. 2201907 and by a Benjamin Peirce Fellowship at Harvard University. James Propp was supported by a Travel Support for Mathematicians gift from the Simons Foundation.

References

  • [1] Y. Chen and V. Kargin, On enumeration and entropy of ribbon tilings, Electron. J. Combinatorics 30 (2023), P2.15.
  • [2] M. Ciucu, Perfect matchings of cellular graphs, J. Algebraic Combin. 5 (1996), no. 2, 87–103. MR1382040 doi:10.1023/A:1022408900061
  • [3] H. Cohn, N. Elkies, and J. Propp, Local statistics for random domino tilings of the Aztec diamond, Duke Math. J. 85 (1996), no. 1, 117–166. MR1412441 doi:10.1215/S0012-7094-96-08506-3
  • [4] J. H. Conway and J. C. Lagarias, Tiling with polyominoes and combinatorial group theory, J. Combin. Theory Ser. A 53 (1990), no. 2, 183–208. MR1041445
  • [5] C. Defant, R. Li, J. Propp, and B. Young, Tilings of benzels via the abacus bijection, Comb. Theory 3 (2023), no. 2, Paper No. 16, 24. MR4646097 doi:10.5070/c63261995
  • [6] N. Elkies, G. Kuperberg, M. Larsen, and J. Propp, Alternating-sign matrices and domino tilings. I, J. Algebraic Combin. 1 (1992), no. 2, 111–132. MR1226347 doi:10.1023/A:1022420103267
  • [7] P. W. Kasteleyn, The statistics of dimers on a lattice : I. the number of dimer arrangements on a quadratic lattice, Physica 27 (1961), 1209–1225.
  • [8] J. Kim and J. Propp, A pentagonal number theorem for tribone tilings, Electron. J. Combinatorics 30 (2023), P3.26.
  • [9] J. C. Lagarias and D. S. Romano, A polyomino tiling problem of Thurston and its configurational entropy, J. Combin. Theory Ser. A 63 (1993), no. 2, 338–358. MR1223689
  • [10] I. Pak, Ribbon tile invariants, Trans. Amer. Math. Soc. 352 (2000), no. 12, 5525–5561. MR1781275
  • [11] R. A. Proctor, Odd symplectic groups, Invent. Math. 92 (1988), no. 2, 307–332. MR936084 doi:10.1007/BF01404455
  • [12] J. Propp, Enumeration of tilings, Handbook of Enumerative Combinatorics (M. Bóna, ed.), CRC Press, 2015.
  • [13] J. Propp, Trimer covers in the triangular grid: twenty mostly open problems, arXiv:2206.06472 [math.CO] (2022).
  • [14] N. J. A. Sloane et al., The On-line Encyclopedia of Integer Sequences, Published electronically at oeis.org (2021).
  • [15] R. P. Stanley, Enumerative combinatorics. Vol. 2, second ed., Cambridge Studies in Advanced Mathematics, vol. 208, Cambridge University Press, Cambridge, [2024] ©2024, With an appendix by Sergey Fomin. MR4621625
  • [16] W. P. Thurston, Conway’s tiling groups, Amer. Math. Monthly 97 (1990), no. 8, 757--773. MR1072815
  • [17] A. Verberkmoes and B. Nienhuis, Bethe ansatz solution of triangular trimers on the triangular lattice, Phys. Rev. E 63 (2001), no. 6, 066122.
  • [18] B. Young, Computing a pyramid partition generating function with dimer shuffling, J. Combin. Theory Ser. A 116 (2009), no. 2, 334--350. MR2475021 doi:10.1016/j.jcta.2008.06.006