M -LEVEL ROOK PLACEMENTS By Kenneth Barrese A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Mathematics—Doctor of Philosophy 2015 ABSTRACT M -LEVEL ROOK PLACEMENTS By Kenneth Barrese Rook theory focuses on placements of non-attacking rooks on boards of various shapes. An important role is played by the rook numbers which count the number of non-attacking placements of a given number of rooks on a board. Ferrers boards, which are boards indexed by integer partitions, are of particular interest. Briggs and Remmel introduced a generalization of rook placements, called m-level rook placements, where a rook is able to attack a subset of the rows. This manuscript presents generalizations of many of the central results regarding rook placements to the case of m-level rook placements. Goldman, Joichi, and White defined the rook polynomial of a board to be the generating function for the rook numbers of that board in the falling factorial basis. By doing so, they were able to give an elegant factorization of the rook polynomial of a Ferrers board in terms of the various column heights. Briggs and Remmel were able to generalize this factorization to the m-level rook polynomial of a subset of Ferrers boards called singleton boards. We give two factorization theorems for the m-level rook polynomial of a Ferrers board. The first is a generalization of the factorization theorem of Briggs and Remmel, working from similar principles. The second relies on a generalization of transposition which we present, called the l-operator. We are also able to use the factorization to describe a unique representative in any m-level equivalence class of Ferrers boards and count the number of singleton boards in the class.. When generalizing the factorization from singleton boards to all Ferrers boards, we pre- serve the definition of the m-level rook polynomial and alter the factorization to apply to all Ferrers boards. We also consider the dual of this problem, applying the factorization of Briggs and Remmel to all Ferrers boards, then trying to determine what is counted by the coefficients of the polynomial in the m-falling factorial basis. It turns out that the coefficients count weighted file placements on a Ferrers board. We also describe a unique representative in each weighted file placement equivalence class of Ferrers boards, as well as count of the number of Ferrers boards in a given weighted file placement equivalence class. Foata and Sch¨ utzenberger presented explicit bijections between rook placements on any two rook equivalent Ferrers boards as part of their construction of a unique representative in each equivalence class of Ferrers boards. A key tool in their construction was local transposition. We present analogous bijections between m-level rook placements on any two m-level rook equivalent Ferrers boards using the local l-operator. The Garsia-Milne Involution Principle was first used in Garsia and Milne’s bijective proof of the Rogers-Ramanujan identities. We use it to construct two types of explicit bijections. The first is an explicit bijection between m-level rook placements on any two m-level rook equivalent singleton boards. The second bijection is between the sets counted by the mlevel analogue of hit numbers of any two m-level rook equivalent Ferrers boards, providing a bijective proof that m-level equivalent Ferrers boards have the same hit numbers. To Jenn, whose patience with this project often outstripped my own. iv ACKNOWLEDGMENTS I must thank my parents for raising me with a deep seated belief in the value of academic pursuits. I would not be getting a Ph.D. without it. Similarly, that I am here stands testament to the enormous talent of the mathematics educators from whom I have been fortunate enough to study. In public schools these educators began with Mrs. Hanshumaker and ended with my friend Mr. Stallons. Thank you both, and I’m still not calling you Frank. As an undergraduate, I benefited from the courses and advice I received from my eventual thesis advisor, Dr. Flahive. Finally, as a doctoral student I owe great gratitude to my committee members, Dr. Bell, Dr. Hall, Dr. Magyar, and Dr. Meierfrankenfeld, all of whom have led courses that I greatly enjoyed. I would like to thank my collaborators, Dr. Loehr and Dr. Remmel, for sharing their interest in the field of rook placements with me. Finally, I have to thank Dr. Sagan, who has been an inspiring educator, a valuable collaborator, and an essential mentor for me. v TABLE OF CONTENTS LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 1 Introduction and definitions 1.1 Introduction . . . . . . . . . . . . . . . 1.1.1 Background . . . . . . . . . . . 1.1.2 m-level rook placements . . . . 1.2 Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1 1 6 8 rook polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 10 11 15 22 25 29 33 Chapter 3 Bijections on m-level rook placements . . . . . 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Rook equivalence and bijections . . . . . . . . . . . . . . . 3.2.1 Reduction to singleton boards . . . . . . . . . . . . 3.2.2 The l-operator . . . . . . . . . . . . . . . . . . . . 3.2.3 The local l-operator . . . . . . . . . . . . . . . . . . 3.2.4 The Local l-operation on an m-level rook placement 3.2.5 Bijections with m-increasing boards . . . . . . . . . 3.3 A second bijection on m-level rook placements . . . . . . . 3.3.1 A Garsia-Milne bijection for rook placements . . . . 3.4 A bijection for hit numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 35 36 36 39 41 45 47 50 51 58 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 Chapter 2 Factorization for m-level 2.1 Introduction . . . . . . . . . . . . 2.2 The m-Factorization Theorem . . 2.3 Rook equivalence . . . . . . . . . 2.4 Enumeration of singleton boards . 2.5 File placements . . . . . . . . . . 2.6 Weight equivalence classes . . . . 2.7 Weight equivalence class sizes . . BIBLIOGRAPHY vi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii LIST OF FIGURES Figure 1.1 On the left, a board of size 7 and a placement of three rooks on the board. On the right, a Ferrers board, B, also of size 7. . . . . . . . . 2 Figure 1.2 Levels and rook placements 7 Figure 2.1 The zones of (1, 1, 2, 3, 5, 7) when m = 3 . . . . . . . . . . . . . . . 11 Figure 2.2 The board Bx . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 Figure 2.3 A board B and l(B) when m = 2 . . . . . . . . . . . . . . . . . . . 17 Figure 2.4 A file placement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 Figure 3.1 On the left, a placement of two 2-level rooks on B. In the middle, the corresponding placement from Lemma 3.2.3 of two 2-level rooks on singleton board BS . On the right, the placement on l(BS ) from Lemma 3.2.5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 The dashed line goes through the cells counted by the 2-arm length of the fourth column and first level, and the shaded cells are counted by the corresponding 2-leg length. . . . . . . . . . . . . . . . . . . . 42 On the left, B3,2 is shaded within B = (1, 4, 4, 5). Notice that l3,2 is not permissible for B since armlm (4, 2) m < leglm (3, 2), which means l3,2 (B) will not be a singleton board. On the right the shaded cells in l(B3,2 ) illustrate this; in this case l3,2 (B) is not even a Ferrers board. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 On the left, B4,2 is shaded. Here l4,2 is permissible for B and is shown on the right. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 (a) A 2-level rook placement on a Ferrers board. (b) The placement on the singleton board obtained after applying Lemma 3.2.3. (c) The placement obtained after applying Lemma 3.2.8 using l3,1 . (d) The placement obtained on a 2-increasing board after applying Lemma 3.2.8 again using l3,3 . . . . . . . . . . . . . . . . . . . . . . . 50 Figure 3.2 Figure 3.3 Figure 3.4 Figure 3.5 vii . . . . . . . . . . . . . . . . . . . . . . On the top left, an element in S with sign −1. On the top right, the image under I which has sign +1. Beneath each board is its image under f . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 Figure 3.7 The placement corresponding to partition {1, 4}, {2, 3, 5}, {6}. . . . 56 Figure 3.8 The placement on Sq3,2 corresponding to (α1 , α2 , α1 ; (1, 3, 2)). . . . 59 Figure 3.9 On the left, an element in S with sign +1. On the right, the image under I which has sign −1. . . . . . . . . . . . . . . . . . . . . . . . 60 On the left, a 2-level placement of white rooks, black rooks, and circled black rooks on (1, 1, 1, 3, 6) fit inside Sq5,2 . On the right, the corresponding placement under the construction in Theorem 3.4.1. 61 Figure 3.6 Figure 3.10 viii Chapter 1 Introduction and definitions 1.1 1.1.1 Introduction Background The study of rook placements began with a 1946 paper of Kaplansky and Riordan [KR46]. They introduced the concept as a way of unifying various results counting the number of permutations which satisfied certain conditions, or violated a specified number of them. To do so, they considered placing non-attacking rooks on “chessboards” of various shapes. A board B is any finite subset of Z+ × Z+ . Geometrically one can think of a board as a finite number of square cells in a square grid, this is illustrated on the left of Figure 1.1. The squares with darkened edges form a board and the lighter lines show the underlying grid. On the right side of the figure there is another board, this time with the underlying grid omitted. The size of a board B is the number of squares in the board, this is often denoted |B|. For example, both boards in Figure 1.1 have size 7. A placement of non-attacking rooks on B, henceforth called a rook placement, is a subset P ⊆ B such that no two elements of P are in the same row or column. Graphically, one can think of placing rooks in the squares of the board so that no rook can capture another rook in a single move, according to the rules of chess. The board on the left in Figure 1.1 contains a rook placement of 3 rooks. We label each cell by the Cartesian coordinates of its northeast 1 .. . R R ... B = (1, 3, 3) = R Figure 1.1: On the left, a board of size 7 and a placement of three rooks on the board. On the right, a Ferrers board, B, also of size 7. corner, so the rooks on the left of Figure 1.1 are in cells (1, 3), (2, 4), and (3, 1) from left to right. Note that this is neither the English nor the French style of writing Ferrers diagrams, but it is the standard convention in modern rook theory literature. It is useful because we usually consider placing rooks on the board from left to right and enumerating the number of such placements is facilitated by our convention. Kaplansky and Riordan also introduced the generating polynomial of rook placements on a board. For any board B we let rk (B) = the number of rook placements P ⊆ B with k rooks. We call rk (B) the kth rook number of B. Note that for any board r0 (B) = 1, as there is a unique way to place no rooks on the board, and r1 (B) = |B|, because a single rook can be placed in any square of the board without violating the non-attacking condition. For either board in Figure 1.1 r0 (B) = 1, r1 (B) = 7, r2 (B) = 10, r3 (B) = 2, and rk (B) = 0 for k ≥ 4. Kaplansky and Riordan defined the rook polynomial of B to be the the generating function 2 of these rook numbers in the standard polynomial basis, yielding rk (B)xk . (1.1.1) k≥0 By this definition of the rook polynomial, the rook polynomial of either board in Figure 1.1 is 1 + 7x + 10x2 + 2x3 . In 1970, Foata and Sch¨ utzenberger [FS70] introduced the term rook equivalent to describe two boards which have the same rook polynomial. For example, the two boards pictured in Figure 1.1 are rook equivalent because they have all the same rook numbers. Foata and Sch¨ utzenberger restricted their consideration to a subset of boards associated with (integer) partitions. A partition is a weakly increasing sequence (b1 , . . . , bn ) of nonnegative integers. We will use the same notation for the corresponding Ferrers board B = (b1 , . . . , bn ) which consists of the the bi lowest squares in column i for 1 ≤ i ≤ n. The board B = (1, 3, 3) is shown on the right in Figure 1.1, while the board on the left side of Figure 1.1 is not a Ferrers board. A Ferrers board B = (b1 , . . . , bn ) is strictly increasing if bi < bi+1 for all i. Note that the board on the right of Figure 1.1 is not strictly increasing because b2 = b3 = 3. Restricting attention to Ferrers boards, Foata and Sch¨ utzenberger were able to characterize the equivalence classes under the rook equivalence relation by showing that there is a unique strictly increasing Ferrers board in each equivalence class. To do so, they constructed explicit bijections between rook placements on any Ferrers board in the class and rook placements on the unique representative. Theorem 1.1.1 (Theorem 11 [FS70]). Each Ferrers board is rook equivalent to exactly one strictly increasing Ferrers board. 3 Goldman, Joichi, and White [GJW75] gave a different definition of the rook polynomial of a board in 1975. Instead of writing it in the standard basis for polynomials, as seen in 1.1.1, they used the basis of falling factorials x ↓n = x(x − 1)(x − 2) . . . (x − n + 1) for any integer n ≥ 0. In this basis, the rook polynomial of a board B with n columns is n rk (B)x ↓n−k . k≥0 This is the definition of rook polynomial we will be using throughout the rest of this document. Using this definition, the rook polynomial of either board in Figure 1.1, both of which have n = 3, is 1x ↓3 +7x ↓2 +10x ↓1 +2x ↓0 = x(x − 1)(x − 2) + 7x(x − 1) + 10x + 2 = x3 + 4x2 + 5x + 2. Goldman, Joichi, and White were able to show that the falling factorial rook polynomial of any Ferrers board factored over the integers. Theorem 1.1.2 (Theorem 2 [GJW75]). Let B = (b1 , . . . , bn ) be a Ferrers board, then n n rk (B)x ↓n−k = k=0 (x + bi − i + 1). i=0 Continuing the above example, we determined that the rook polynomial of B = (1, 3, 3) 4 is x3 + 4x2 + 5x + 2 = (x + 1)(x + 2)(x + 1) = (x + 1 − 1 + 1)(x + 3 − 2 + 1)(x + 3 − 3 + 1) as stated in Theorem 1.1.2. In 2009 Loehr and Remmel [LR09] introduced another explicit bijection between rook placements on rook equivalent Ferrers board. Their bijection made use of the Garsia-Milne Involution Principle [GM81], which was developed as part of Garsia and Milne’s bijective proof of the Rogers-Ramanujan identities. Another important set of numbers associated with a board are its hit numbers. Given a permutation ω = (ω(1), . . . , ω(N )) in SN , the symmetric group on N elements, one can associate a rook placement of N rooks on an N × N square board with ω in the following way. If ω(i) = j, then place a rook in square (i, N + 1 − j). We will denote this placement R(ω). This is equivalent to the permutation matrix of ω, considered as a left action, where the zeroes in the matrix are empty squares and the ones in the matrix are squares containing rooks. For any board B and any integer N such that B can be considered as a subset of the squares of an N × N board, we can define the kth hit set of B by, Hk,N (B) = {R(ω) | ω ∈ SN and |R(ω) ∩ B| = k}. Then the kth hit number of B is hk,N (B) = |Hk,N (B)|. 5 In their seminal 1946 paper, Kaplansky and Riordan noted that this definition implied that N N hk,N k=0 (B)xk rk (B)(N − k)!(x − 1)k , = k=0 which implies that two rook equivalent boards will have the same hit numbers for any N large enough for both to be considered as subboards of the N ×N square board. In the paper cited above, Loehr and Remmel also used the Garsia-Milne method to give explicit bijections between the hit sets of any two equivalent Ferrers boards. This provided a bijective proof that any two rook equivalent Ferrers boards have the same hit numbers. 1.1.2 m-level rook placements In 2008 Briggs and Remmel [BR06] introduced a generalization of rook placements, called m-level rook placements. The new rook placements correspond with elements of the wreath product of the cyclic group of order m with the symmetric group on N elements, Cm SN , in the same way that Kaplansky and Riordan’s placements correspond to elements of SN as detailed at the end of the last subsection. Using m-level rook placements and the concept of flag descents developed by Adin, Brenti, and Roichman [ABR01], Briggs and Remmel were able to generalize a formula of Frobenius to Cm Sn . They also provided a factorization, similar to that of Goldman, Joichi, and White, for the m-rook polynomial of a specific type of Ferrers board, called a singleton board. In order to make the concepts of the previous paragraph more precise, we first define what a level is. Fix a positive integer m. Partition the rows of Z+ × Z+ into levels where the pth level consists of rows (p − 1)m + 1, (p − 1)m + 2, . . . , pm. The situation for m = 2 is shown on the left in Figure 1.2 where the boundaries between the levels have been thickened. 6 .. . { { R level 2 ... R level 1 R R Figure 1.2: Levels and rook placements Given a board, B, an m-level rook placement (called an m-rook placement by Briggs and Remmel) is P ⊆ B where no two elements of P are in the same level or the same column. Note that when m = 1 we recover the ordinary notion of a rook placement. By way of example, in Figure 1.2, the placement on the middle board is a 2-level rook placement while the one on the right is not since it has two rooks in the first level. We let rk,m (B) = the number of m-level rook placements P ⊆ B with k rooks. In general, we will add a subscript m to quantities when considering their m-level equivalents. For B = (1, 3, 3) we have r0,2 (B) = 1, r1,2 (B) = 7, r2,2 (B) = 6, and rk,2 (B) = 0 for k ≥ 3. To state the Briggs-Remmel generalization of Theorem 1.1.2, we need a few more concepts. One is of an m-falling factoring, a generalization of the falling factorial, defined by x ↓n,m = x(x − m)(x − 2m) . . . (x − (n − 1)m). Now we define an important subset of Ferrers boards when considering m-level rook placements. Having already fixed a positive integer m, a singleton board (called an m-Ferrers board by Briggs and Remmel), is a Ferrers board B = (b1 , . . . , bn ) such that there is 7 at most one bi in each of the intervals (0, m), (m, 2m), . . . where (km, (k + 1)m) = {km + 1, km + 2, . . . , km + m − 1}. Another way of defining singleton boards is to consider the highest level a column has squares in. We say that the ith column of B terminates in level p if p is the largest integer such that the ith column has non-empty intersection with Qp . This gives another characterization of a singleton board, as any Ferrers board such that, for each positive integer p, the set of all columns bi terminating in level p contains at most one i such that bi ≡ 0 mod m. Briggs and Remmel gave a factorization theorem, similar to that of Goldman, Joichi, and White, but only for singleton boards. Theorem 1.1.3 (Theorem 2 [BR06]). If B = (b1 , . . . , bn ) is a singleton board then n n (x + bi − (i − 1)m). rk,m (B)x ↓n−k,m = k=0 1.2 i=1 Organization We proceed to generalize the above results for m-level rook placements. In Chapter 2 we remove the singleton requirement from Theorem 1.1.3. This facilitates generalizing Theorem 1.1.1 to m-level rook placements on Ferrers boards and enumerating the singleton boards in a given equivalence class, an analogue of a result of Goldman, Joichi, and White. Finally we consider what is counted if we take the product on the right side of Theorem 1.1.3 for any Ferrers board and expand it in the m-falling factorial basis. In Chapter 3 we begin by extending the bijections detailed by Foata and Sch¨ utzenberger to m-level rook placements on Ferrers boards. We create a second explicit bijection between m-level rook placements on singleton boards by generalizing that of Loehr and Remmel, 8 again using the Garsia-Milne Involution Theorem. Finally we use similar methods to create bijections between the sets counted by appropriately defined hit numbers of m-level rook equivalent Ferrers boards. 9 Chapter 2 Factorization for m-level rook polynomials 2.1 Introduction The material in this chapter is a result of a collaboration with Nicholas Loehr, Jeffrey Remmel, and Bruce Sagan, and has appeared in the paper [BLRS14]. The first goal of this chapter is to remove the singleton board restriction from Theorem 1.1.3 and prove a generalization of this theorem for any Ferrers board. This will be done in the next section. Call boards B, B m-level rook equivalent if rk,m (B) = rk,m (B ) for all k. In Section 2.3 we extend to all m Theorem 1.1.1 of Foata and Sch¨ utzenberger [FS70] giving a distinguished member of each 1-level rook equivalence class. Goldman, Joichi and White used the Factorization Theorem to enumerate the number of Ferrers boards 1-level rook equivalent to a given board. In Section 2.4 we generalize this formula to count m-level rook equivalent singleton boards for arbitrary m. The rest of the chapter is devoted to the following dual problem. Rather than changing the product side of Theorem 1.1.3, keep the same product for all Ferrers boards and expand it in the m-falling factorial basis. What do the coefficients count? We show in Section 2.5 that they are generating functions for certain weighted file placements, where such placements allow more than one rook in a given row. The next sections investigate properties of the corresponding equivalence classes. In 10 Figure 2.1: The zones of (1, 1, 2, 3, 5, 7) when m = 3 particular, in Section 2.7 we count the number of boards in a given class. 2.2 The m-Factorization Theorem In order to generalize Theorem 1.1.3 to all Ferrers boards, it will be convenient to break a board up into zones depending on the lengths of the columns. Given integers s, t, the interval from s to t will be denoted [s, t] = {s, s + 1, . . . , t}. An m-zone, z, of a board B = (b1 , . . . , bn ) is a maximal interval [s, t] such that bs m = bs+1 m = · · · = bt m . To illustrate this concept, consider m = 3 and the board B = (1, 1, 2, 3, 5, 7) shown in Figure 2.1. In this case the zones are z1 = [1, 3] since b1 3 = b2 3 = b3 3 = 0, z2 = [4, 5] since b4 3 = b5 3 = 3, and z3 = [6, 6] since b6 3 = 6. The zones in Figure 2.1 are separated by thick lines (as are the levels). Note that a Ferrers board is a singleton board if and only if each zone contains at most one column whose length is not a multiple of m. This is the reason for our choice of terminology. 11 The m-floor function of an integer n is defined as follows, let n m = the largest multiple of m less than or equal to n. For example, 17 3 = 15 since 15 ≤ 17 < 18. In addition to taking m-floors, we will have to consider remainders modulo m. Given an integer n, we denote its remainder on division by m by ρm (n) = n − n m . Continuing the previous example, ρm (17) = 17 − 17 3 = 2. If z is a zone of a Ferrers board B = (b1 , . . . , bn ) then its m-remainder is ρm (bi ). ρm (z) = i∈z In Figure 2.1, the boxes corresponding to the 3-remainders of the zones are shaded. In particular ρ3 (z1 ) = 1 + 1 + 2 = 4, ρ3 (z2 ) = 0 + 2 = 2, and ρ3 (z3 ) = 1. We are now in a position to state and prove our generalization of Theorem 1.1.3. Theorem 2.2.1 (m-Factorization Theorem). If B = (b1 , . . . , bn ) is any Ferrers board then n rk,m (B)x ↓n−k,m = k=0    n  x + bi m − (i − 1)m + ρm (z) if i is the last index in its zone z,   i=1  x + bi m − (i − 1)m otherwise. Proof. Since this is a polynomial identity, it suffices to prove it for an infinite number of values for x. We will do so when x is a nonnegative multiple of m. Consider the board Bx derived from B by adding an x × n rectangle below B. Figure 2.2 shows a schematic representation of Bx . Note that since x is a multiple of m, the zones and remainders of B and Bx are the same. We will show that both the sum and the product count the number of m-rook placements on Bx consisting of n rooks. 12 B Bx = x n Figure 2.2: The board Bx For the sum side, note that any placement of n rooks on Bx must have k rooks in B and n − k rooks in the rectangle for some 0 ≤ k ≤ n. By definition, rk,m (B) counts the number of placements on B. Once these rooks are placed, one must place the remaining rooks in the x × (n − k) subrectangle consisting of those columns of the original rectangle not used for the rooks on B. Placing these rooks from left to right, there will be x choices for the position of the first rook, then x − m choices for the next, and so on, for a total of x ↓n−k,m choices. Thus the sum side is rn,m (Bx ) as desired. On the product side, it will be convenient to consider placing rooks on Bx zone by zone from left to right. So suppose z = [s, t] is a zone and all rooks in zones to its left have been placed. Because z is a zone we have bs m = · · · = bt m = cm for some constant c. Also, among all the rooks placed in the columns of z, there is at most one which is in the set of squares R corresponding to ρm (z). If there are no rooks in R then they all go in a rectangle 13 of height x + cm. Thus, using the same ideas as in the previous paragraph, the number of placements is (x + cm − (s − 1)m)(x + cm − sm) . . . (x + cm − (t − 1)m). (2.2.1) When there is one rook in R, say it is in the column with index i. So there are ρm (bi ) choices for the placement of this rook and the rest of the rooks go in a rectangle of height x + cm. This gives a count of ρm (bi )(x + cm − (s − 1)m)(x + cm − sm) . . . (x + cm − (t − 2)m). (2.2.2) Adding together the contributions from (2.2.1) and (2.2.2) and factoring, we see that the total number of placements is (x + cm − (s − 1)m) . . . (x + cm − (t − 2)m)(x + cm − (t − 1)m + ρm (z)). Remembering that bs m = · · · = bt m = cm, we see that this is exactly the contribution needed for the product. We should show why our result implies the theorems of Goldman-Joichi-White and Briggs-Remmel. In both cases, it is clear that the sum sides correspond, so we will concentrate on the products. For Theorem 1.1.2 we take m = 1. Since n 1 = n for any n, ρ1 (z) = 0 for any zone z and the two cases in Theorem 2.2.1 are the same. So the contribution of the ith column to 14 the product is x + bi 1 − (i − 1) · 1 = x + bi − i + 1 in agreement with the Factorization Theorem. As for Theorem 1.1.3, suppose that B is a singleton board and consider any zone z = [s, t]. If s ≤ i < t then bi is a multiple of m and x + bi m − (i − 1)m = x + bi − (i − 1)m. And if i = t then ρm (z) = ρm (bt ) so that x + bt m − (t − 1)m + ρm (z) = x + bt − (t − 1)m. So in either case one gets the same factor as in the Briggs-Remmel result. 2.3 Rook equivalence Two Ferrers boards B, B are m-level rook equivalent if rk,m (B) = rk,m (B ) for all k ≥ 0. In this case we will write B ≡m B . We will drop the “m” and just say “rook equivalent” if m = 1. Recall that Foata and Sch¨ utzenberger [FS70] proved a beautiful theorem giving a distinguished board in each equivalence class. Theorem 2.3.1 ([FS70]). Every Ferrers board is rook equivalent to a unique strictly increasing board. The purpose of this section is to extend Theorem 2.3.1 to arbitrary m. The FoataSch¨ utzenberger result was reproved by Goldman-Joichi-White using their Factorization The15 orem. To see the connection, suppose that B = (b1 , . . . , bn ) and B = (b1 , . . . , bn ) are two Ferrers boards. Although we are writing the boards with the same number of columns, this is no restriction since we can always pad the shorter board with columns of height 0 on the left. So B and B are rook equivalent if and only if they have the same generating function in the falling factorial basis. By Theorem 1.1.2, this happens if and only if the two vectors ζ(B) = (−b1 , 1−b2 , 2−b3 , . . . , (n−1)−bn ) and ζ(B ) = (−b1 , 1−b2 , 2−b3 , . . . , (n−1)−bn ) are rearrangements of each other since these are the zeros of the corresponding products. We call ζ(B) the root vector of B. For example, if B = (1, 1, 3) and B = (2, 3) then, rewriting B = (0, 2, 3), we have ζ(B) = (−1, 0, −1) and ζ(B ) = (0, −1, −1) and so B ≡ B . We should note that padding a board B with zeros will change the entries of ζ(B). Also, if ζ(B) is a rearrangement of ζ(B ) then the same will be true when padding both B and B with any given number of zeros such that both resulting boards have the same number of columns. We now return to considering general m. Define the m-level root vector of B = (b1 , . . . , bn ) to be ζm (B) = (a1 , . . . , an ) where ai =     (i − 1)m − bi m − ρm (z) if i is the last index in its zone z,    (i − 1)m − bi m otherwise. The next result is immediate from Theorem 2.2.1. Proposition 2.3.2. Ferrers boards B and B satisfy B ≡m B if and only if ζm (B) is a rearrangement of ζm (B ). 16 l(B) = B= Figure 2.3: A board B and l(B) when m = 2 Our first order of business will be to restrict the class representative problem to considering singleton boards B = (b1 , . . . , bn ) since then the ai are simpler to compute. Indeed, the argument in the last paragraph of Section 2.2 shows that in this case ai = (i − 1)m − bi for all i. To describe a singleton board in each equivalence class, let Qp denote the set of squares in the pth level of the tiling Q of the first quadrant and, for any board B = (b1 , . . . , bn ), let lp = |B ∩ Qp |. For example, if m = 2 and B = (1, 3, 3) as shown on the left in Figure 2.3, then l1 = 5, l2 = 2, and li = 0 for i ≥ 3. For any Ferrers board B, if t is the largest index with lt = 0 then we let l(B) = (lt , lt−1 , . . . , l1 ). We call this function the l-operator on boards. The board on the right in Figure 2.3 shows l(1, 3, 3) = (2, 5). Lemma 2.3.3. For any m and any Ferrers board B = (b1 , . . . , bn ) the sequence l(B) = (lt , . . . , l1 ) is a partition, the Ferrers board l(B) is singleton, and B ≡m l(B). Proof. To see that l(B) is a partition first note that, for any p > 1, the set of columns of B ∩ Qp is a subset of the columns of B ∩ Qp−1 . Furthermore, for each of the former columns 17 we have m squares of that column in B ∩ Qp−1 . It follows that lp (B) ≤ lp−1 (B) as desired. To show that l(B) is singleton, it suffices to show that if lp is not a multiple of m then lp m < lp−1 m . Let c be the number of columns of B ∩ Qp which contain m squares and d be the number of columns containing fewer than m squares. By assumption d > 0. Since these c + d columns of B ∩ Qp−1 must all contain m cells we have lp < (c + d)m ≤ lp−1 . Taking floors finishes this part of the proof. To prove rook equivalence, pick any L = {p1 > · · · > pk } ⊆ {1, . . . , t}. It suffices to show that the number of ways to place rooks on the levels of B indexed by L is the same as the number of ways to place rooks in the columns of l(B) indexed by L. For the former, if one places the rooks level by level from top to bottom then, since each rook in a higher level rules out m squares in each level below it, we obtain a count of lp1 (lp2 − m)(lp3 − 2m) . . . (lpk − (k − 1)m). Now consider placing the rooks in l(B) column by column from left to right. Since l(B) is singleton, each rook placed eliminates m squares in each column to its right from consideration. Thus we obtain the same count as before. It is worth noting here that this gives us an alternative form of the factorization from Theorem 2.2.1. Theorem 2.3.4. If B = (b1 , . . . , bn ) is any Ferrers board where t is the largest index with lt = 0 and N is any integer greater than or equal to both n and t, then N N rk,m (B)x ↓N −k,m = k=0 (x + lt+1−i − (i − 1)m) i=1 18 Proof. We know that l(B) is a singleton board by Lemma 2.3.3. To facilitate comparing their m-level rook polynomials, pad both B and l(B) on the left with columns of height 0 until both have N columns. This will not alter the singleton property of l(B), so by Theorem 1.1.3, we know that the product side is the factorization of the m-level rook polynomial of l(B). Furthermore, since we know B ≡m l(B) from Lemma 2.3.3, we know N N rk,m (B)x ↓N −k,m = k=0 rk,m (l(B))x ↓N −k,m , k=0 which completes the proof. The possible ζm -vectors of singleton boards are easy to characterize. Proposition 2.3.5. Consider a vector ζ = (a1 , . . . , an ) where a1 = 0. Let B = (b1 , . . . , bn ) where we define bi = (i − 1)m − ai for all i. We have that ζ = ζm (B) where B is a singleton board if and only if the following conditions are satisfied: (i) ai+1 ≤ ai + m for i ≥ 1, and (ii) if neither of ai+1 , ai are multiples of m then ai+1 m ≤ ai m . Proof. We claim that (i) and the fact that a1 = 0 are equivalent to B being a weakly increasing sequence of nonnegative integers. Since a1 = 0 we have b1 = −a1 = 0 which is nonnegative. And for i ≥ 1 bi+1 − bi = (im − ai+1 ) − ((i − 1)m − ai ) = m + ai − ai+1 . (2.3.1) So (i) is equivalent to B being weakly increasing. A similar argument shows that (ii) is equivalent to the singleton condition. 19 We are finally in a position to give distinguished representatives for the m-level equivalence classes. A Ferrers board B = (b1 , . . . , bn ) is called m-increasing if b1 > 0 and bi+1 ≥ bi + m for i ≥ 1. Note that a 1-increasing board is strictly increasing in the sense of Theorem 2.3.1. Also note that, although most properties of Ferrers boards are not affected by padding with columns of length zero, the m-increasing condition will be destroyed. Theorem 2.3.6. Every Ferrers board is m-level rook equivalent to a unique m-increasing board. Proof. Clearly any m-increasing board is a singleton board. So, by Lemma 2.3.3, it suffices to show that any singleton board B is m-rook equivalent to a unique m-increasing board. An example of the construction of this board is given after this proof to illustrate the technique. Let N = |B| + 1 and pad B with columns of zeros so as to write B = (b1 , . . . , bN ). By the choice of N , any board B which is m-equivalent to B can be written as B = (b1 , . . . , bN ) and b1 = b1 = 0. Letting ζm (B) = (a1 , . . . , aN ) and ζm (B ) = (a1 , . . . , aN ), the choice of N also ensures that a1 = a1 = 0 and ai , ai ≥ 0 for all i. We claim that a singleton board B = (b1 , . . . , bn ) will be m-increasing, after discarding any columns of height 0, if and only ζm (B ) = (a1 , . . . , an ) is weakly decreasing. Indeed, this follows directly from (2.3.1). We first show existence. By the previous paragraph, we wish to rearrange ζm (B) in such a way that the portion of the rearrangement corresponding to nonzero entries of the board is weakly decreasing. And the portion of the rearrangement corresponding to zero entries of the board must be of the form 0, m, 2m, . . . , cm for some c. So choose cm to be the largest multiple of m in ζ = ζm (B). We claim that the elements 0, m, . . . , (c − 1)m also occur in ζ. By definition, b1 = 0 so this multiple of m occurs as does cm by assumption. Suppose, 20 towards a contradiction, that jm does not occur where 0 < j < c. Then the definition of jm and condition (i) of Proposition 2.3.5 implies there must be an index i such that we have the inequalities (j − 1)m < ai < jm < ai+1 < (j + 1)m. But now ai m < ai+1 m which contradicts condition (ii) of the same lemma. We now define ζ = (a1 , . . . , aN ) where (a1 , a2 , a3 , . . . , ac+1 ) = (0, m, 2m, . . . , cm) and (ac+2 , ac+3 , . . . , aN ) is the rest of ζ arranged in weakly decreasing order. Since a1 = 0, we can show that ζ corresponds to an m-increasing board by checking conditions (i) and (ii) of Proposition 2.3.5. Condition (i) is clearly true for i ≤ c. For i = c, one can show, by using a proof as in the previous paragraph and the choice of cm as the largest multiple of m in ζ, that ac+2 < (c + 1)m. So (i) also holds in this case. And for i > c the fact that the sequence is weakly decreasing makes (i) a triviality. The same reasoning as for (i) shows that (ii) must also hold. Thus, defining B = (b1 , . . . , bN ) where bi = (i − 1)m − ai for all i results in a singleton board. Furthermore, by construction, b1 = · · · = bc+1 = 0 and (bc+2 , . . . , bN ) is m-increasing. Hence removing the zeros from B leaves the desired m-increasing board. To show uniqueness, suppose ζ = (a1 , . . . , aN ) is a rearrangement of ζ corresponding to a padded m-increasing board. Then ζ must start with 0, m, . . . , cm for some c and be weakly decreasing thereafter by equation (2.3.1). Without loss of generality, we can assume ac+2 = (c + 1)m, since if the two are equal we can just add ac+2 to the initial run of multiples of m. So ζ will be the rearrangement of ζ constructed in the existence proof as long as cm is the largest multiple of m in ζ. But if cm is not the largest multiple of m in ζ then (ac+2 , . . . , aN ) must contain an element greater than or equal to (c + 1)m. And since 21 this portion of ζ is weakly decreasing, this forces ac+2 > (c + 1)m since equality was ruled out earlier. But then ac+1 = cm and ac+2 do not satisfy condition (i) of Proposition 2.3.5, contradicting the fact that ζ corresponds to a singleton board. This finishes the proof of uniqueness and of the theorem. To illustrate the construction in the previous proof, consider m = 2 and the singleton board B = (1, 2, 2, 3). Now N = 1 + 2 + 2 + 3 = 8 and we pad B with zeros to length 8 + 1 = 9, obtaining B = (0, 0, 0, 0, 0, 1, 2, 2, 3). Thus ζ = ζ2 (B) = (0, 2, 4, 6, 8, 9, 10, 12, 13). The largest multiple of 2 in ζ is 12, so we rearrange ζ to begin with the multiples of 2 up through 12 and then decrease. The result is ζ = (0, 2, 4, 6, 8, 10, 12, 13, 9) with associated board B = (0, 0, 0, 0, 0, 0, 0, 1, 7). Removing the initial zeros, we get the 2-increasing board (1, 7) which is easily seen to be 2-level rook equivalent to B. 2.4 Enumeration of singleton boards Goldman, Joichi, and White [GJW75] used their factorization theorem to give a simple formula for the size of a given rook equivalence class. The basic idea is to count, for any board B, the number of rearrangements of ζ1 (B) which correspond to a Ferrers board. To state their result, given any finite vector ν of nonnegative integers, we let n(ν) = (n0 , n1 , . . . ) be defined by ni = the number of copies of i in ν. So ni (ν) = 0 if i < 0 or i is sufficiently large. Theorem 2.4.1 ([GJW75]). Let B = (b1 , . . . , bN ) be a Ferrers board where N = |B| + 1, and suppose n(ζ1 (B)) = (n0 , n1 , . . . ). The number of Ferrers boards in the equivalence class 22 of B is i≥1 ni + ni−1 − 1 . ni Because the entries of ζm (B) are more complicated for m ≥ 2, we will not be able to count all boards in an m-level equivalence class. But we can at least enumerate the singleton boards. The formula will be in terms of multinomial coefficients. Theorem 2.4.2. Let B = (b1 , . . . , bN ) be a singleton board where N = |B| + 1, and suppose n(ζm (B)) = (n0 , n1 , . . . ). Then the number of singleton boards which are m-level rook equivalent to B is i≥0 nim + nim+1 + · · · + nim+m − 1 . nim − 1, nim+1 , nim+2 , . . . , nim+m Proof. By the choice of N we have that ζ = ζm (B) begins with a zero and has all entries nonnegative. So it suffices to compute the number of rearrangements of ζ beginning with 0 and satisfying conditions (i) and (ii) of Proposition 2.3.5. Let d be the maximum entry of ζ. If d = 0 then the result is easy to verify, so assume d > 0. Let cm be the largest nonnegative multiple of m with cm < d. Note that cm exists since d > 0 and also that, by the argument given in the proof of Theorem 2.3.6, ncm > 0. Consider the vector ζ = (a1 , . . . , an ) obtained from ζ by removing all entries which are larger than cm. We claim that ζ = ζm (B ) for some singleton board B . As usual, we use Propostion 2.3.5. Certainly a1 = a1 = 0 since none of the zeros were removed from ζ. Suppose, towards a contradiction, that condition (i) is false in that aj+1 > aj + m for some j. Let ai be the element of ζ corresponding to aj . But since we removed the largest elements of ζ we have ai+1 ≥ aj+1 > aj + m = ai + m 23 which is impossible. A similar contradiction demonstrates (ii), and our claim is proved. Now, by induction, it suffices to show that the number of rearrangements of ζ which come from a given ζ by inserting elements larger than cm is ncm + ncm+1 + · · · + ncm+m − 1 . ncm − 1, ncm+1 , ncm+2 , . . . , ncm+m Consider elements ai , ai+1 in ζ. First note that if ai comes from ζ and ai+1 > cm then we must have ai = cm. If this were not the case then, since ai < cm < ai+1 , to make condition (i) true neither ai nor ai+1 would be multiplies of m. But by the same pair of inequalities we would have ai+1 m ≥ cm > ai m which contradicts condition (ii). Thus we can insert elements greater than cm only after copies of cm itself. This argument is analogous to that given in the proof of Theorem 2.3.6. In addition, any ai+1 with cm < ai+1 ≤ cm + m can come after ai = cm as it is easily verified that we always have conditions (i) and, vacuously, (ii) for such ai+1 . We also claim that the elements larger than cm can be arranged in any order with respect to each other. To see this, suppose cm < ai , ai+1 ≤ cm + m. Condition (i) is immediate because of the given bounds. And if neither is a multiple of m then we have cm < ai , ai+1 < cm + m which implies condition (ii). Finally, if ai > cm and ai+1 comes from ζ then ai+1 < ai and conditions (i) and (ii) are trivial. So an element greater than cm can be followed by any element of ζ . We now enumerate the number of ζ coming from ζ using the structural properties from the previous three paragraphs. There are ncm copies of cm and ncm+1 +ncm+1 +· · ·+ncm+m elements to be inserted after these copies where the space after a copy can be used multiple times. And any element of ζ can follow the inserted elements. So the total number of choices 24 R R R R R Figure 2.4: A file placement for this step is the binomial coefficient ncm + ncm+1 + · · · + ncm+m − 1 . ncm+1 + ncm+2 + · · · + ncm+m Now we need to arrange the elements greater than cm among themselves. This can be done arbitrarily, so the number of choices is ncm+1 + ncm+2 + · · · + ncm+m . ncm+1 , ncm+2 , . . . , ncm+m Multiplying the two displayed expressions and canceling (ncm+1 + ncm+2 + · · · + ncm+m )! gives the desired result. Note that the result just proved does indeed generalize Theorem 2.4.1. This is because when m = 1, every board is a singleton board. And the products clearly conicide in this case. 2.5 File placements Thus far our focus has been to keep the sum side of Theorem 1.1.3 the same and modify the product side to get equality for all Ferrers boards. Another possibility would be to keep the 25 product side the same, expand it in the m-falling factorial basis, and see if the coefficients of the linear combination count anything. This will be our approach in the current section. An equivalent formula in the case where m = 2, albeit in a different rook model, can be found in a paper of Haglund and Remmel [HR01, Theorem 8]. It turns out that these coefficients count weighted file placements. A file placement on a board B is F ⊆ B such that no two rooks (elements) of F are in the same column. However, we permit rooks to be in the same row. Figure 2.4 displays such a placement on the Ferrers board (2, 2, 3, 3, 3, 3). We let fk (B) = the number of file placements F ⊆ B with k rooks. It is easy to count such placements. If B has bi squares in column i for 1 ≤ i ≤ n (B need not be a Ferrers board) then fk (B) = ek (b1 , . . . , bn ) where ek is the kth elementary symmetric function. So in order to get more interesting results, we will weight file placements. Fix, as usual, a positive integer m. Given a board B and a file placement F ⊆ B, let t be the largest index of a row in B and consider y1 , . . . , yt where yj is the number of rooks of F in row j. Define the m-weight of F to be wtm F = 1 ↓y1 ,m 1 ↓y2 ,m . . . 1 ↓yt ,m . In the example of Figure 2.4 with m = 3 we have wtm F = 1 ↓3,3 ·1 ↓0,3 ·1 ↓2,3 = (1)(−2)(−5) · (1) · (1)(−2) = −20. Note that if F is actually a rook placement then wtm F = 1 because 1 ↓0,m = 1 ↓1,m = 1 for 26 any m. Given a board, B, we define the associated m-weighted file placement numbers to be fk,m (B) = wtm F F where the sum is over all file placements F ⊆ B with k rooks. These are the coefficients which we seek. Theorem 2.5.1 (m-weight Factorization Theorem). For any Ferrers board B = (b1 , . . . , bn ) n n (x + bi − (i − 1)m). fk,m (B)x ↓n−k,m = (2.5.1) i=1 k=0 Proof. In the manner to which we have become accustomed, we consider the board Bx obtained by attaching an x × n rectangle R to B where x is a nonnegative multiple of m. Consider mixed placements F ⊆ Bx which are file placements when restricted to B, but satisfy the m-level condition when restricted to R. We will compute S = F wtm F where the sum is over the mixed placements F on Bx with n rooks. We first recover the sum side of equation (2.5.1). The mixed placements with k rooks on B will contribute fk,m (B) to S from these rooks. And x ↓n−k,m will be the contribution from the n − k rooks on R since any rook placement has m-weight equal to one as noted previously. So wtm F k fk,m (B)x ↓n−k,m = (2.5.2) Fk where the sum is over all mixed placements F k ⊆ Bx having k rooks in B. Now summing over k gives us the desired equality. To obtain the product, let B and Bx be B and Bx with their nth columns removed, 27 respectively. By induction on n, it suffices to prove that n n−1 fk,m (B)x ↓n−k,m = (x + bn − (n − 1)m) k=0 fk,m (B )x ↓n−k−1,m . (2.5.3) k=0 Comparing equations (2.5.2) and (2.5.3), we see it suffices to show that, for any mixed placement F ⊆ Bx , (x + bn − (n − 1)m) wtm F = wtm F (2.5.4) F where the sum is over all mixed placements F ⊆ Bx whose restriction to Bx is F . To this end, let y0 be the number of rooks in F which are in R. Also let yj , 1 ≤ j ≤ bn , be the number of rooks in the jth row of F ∩ B . Since every column of F has a rook, we have y0 + y1 + · · · + ybn = n − 1. (2.5.5) We now consider two cases depending on whether the rook in column n of F lies in B or in R. If it lies in R then, by the m-level condition, there are x − y0 m places for the rook. Since these placements are in rows not occupied by rooks of F , each of them contributes a factor of 1 to the weight for a total change in weight of x − y0 m from this case. Now suppose that the rook lies in B, say in the jth row. Then in passing from F to F , the weight is changed from 1 ↓yj ,m to 1 ↓yj +1,m . This means that the weight is multiplied by 1 − yj m when placing a rook in row j of B. Adding together all the contributions and using equation (2.5.5) gives bn x − y0 m + (1 − yj m) = x + bn − (n − 1)m. j=1 28 This completes the proof of equation (2.5.4) and of the theorem. We note that Theorem 2.5.1 is another generalization of the Factorization Theorem. Indeed, when m = 1, then any file placement having a row with y ≥ 2 rooks will have a factor of 1 ↓y,1 = 0. And any rook placement will have a weight of 1. Thus fk,1 (B) = rk,1 (B). 2.6 Weight equivalence classes Given m, define two boards B, B to be m-weight file equivalent, written B ≈m B , if fk,m (B) = fk,m (B ) for all k ≥ 0. Our goal in this section is to find distinguished representatives for the m-weight file equivalence classes. Interestingly, our result will be dual to the one for m-level rook equivalence in the sense that the inequalities will be reversed. In order to define the representatives, we will have to assume that all our boards start with at least one zero. So for this section we will write our Ferrers boards in the form B = (b0 , b1 , . . . , bn ) where b0 = 0. We can use Theorem 2.5.1 to test m-weight file equivalence. The m-weight root vector of a Ferrers board B = (b0 , b1 , . . . , bn ) is ωm (B) = (−b0 , m − b1 , 2m − b2 , . . . , nm − bn ). From the m-weight Factorization Theorem we immediately get the following. Proposition 2.6.1. Ferrers boards B and B satisfy B ≈m B if and only if ωm (B) is a rearrangement of ωm (B ). We will also need a characterization of the vectors which can be m-weight root vectors. The proof of the next result is similar to that of Proposition 2.3.5 and so is omitted. 29 Proposition 2.6.2. Consider a vector ω = (a0 , a1 , . . . , an ). Let B = (b0 , b1 , . . . , bn ) where we define bi = im − ai for all i. We have that ω = ωm (B) where B is a Ferrers board if and only if the following conditions are satisfied: (i) a0 = 0, and (ii) ai+1 ≤ ai + m for i ≥ 0. Now define a Ferrers board B = (b0 , b1 , . . . , bn ) to be m-restricted if bi+1 ≤ bi + m for all i ≥ 0. We now show that the m-weight file equivalence class of any Ferrers board contains a unique m-restricted board. An example of the construction of this board follows the proof. Theorem 2.6.3. Every Ferrers board B = (b0 , . . . , bn ) is m-weight file equivalent to a unique m-restricted board. Proof. Similarly to the proof of Theorem 2.3.6, we rewrite B = (b0 , . . . , bN ) where N = |B|. This assures us that any equivalent board B = (b0 , . . . , bN ) has ωm (B ) which is nonnegative and starts with zero. Also, using equation (2.3.1), we see that B is m-restricted if and only if ω(B ) is weakly increasing. So consider ω = (a0 , . . . , aN ) which is the unique weakly increasing rearrangement of ω = ωm (B) = (a0 , . . . , aN ). It suffices to show that ω = ωm (B ) for some board B . So we just need to check the two conditions of Proposition 2.6.2. Condition (i) follows from the choice of N and the fact that ω is nonnegative. For condition (ii), assume, towards a contradiction, that there is an index i such that ai+1 > ai + m. Let aj be the element of ω which was rearranged to become ai+1 . Then aj = ai+1 > 0 and so there must be an element ak with k < j and ak < aj . Let k be the largest such index. By the choice of k and the fact that ω satisfies (ii), we must have ak ≥ aj − m. Thus ai < ai+1 − m = aj − m ≤ ak < aj = ai+1 . 30 But then when rearranging ω in weakly increasing order, ak should have been placed between ai and ai+1 , a contradiction. By way of illustration, suppose that we take m = 2 and B = (1, 5). This gives N = |B| = 6 and ω = ω2 (B) = (0, 2, 4, 6, 8, 10, 12) − (0, 0, 0, 0, 0, 1, 5) = (0, 2, 4, 6, 8, 9, 7). The weakly increasing rearrangement of ω is ω = (0, 2, 4, 6, 7, 8, 9) and so B = (0, 2, 4, 6, 8, 10, 12) − (0, 2, 4, 6, 7, 8, 9) = (0, 0, 0, 0, 1, 2, 3). There is a close relationship between the m-increasing boards introduced in Section 2.3 and m-restricted boards. This is easy to see if m = 1. In this case, board B is 1-increasing if and only if its transpose B t (obtained by interchanging rows and columns) is 1-restricted. Indeed, a Ferrers board is 1-increasing if and only if the northwestern boundary of B contains no horizontal line segment of length at least 2. And a board is 1-restricted if and only if this boundary contains no vertical line segment of length at least 2. Note also that when m = 1, the l-operator of Section 2.3 satisfies l(B) = B t . In generalizing these ideas to all m, it is the l-operator which is key as the next result shows. Proposition 2.6.4. The l-operator has the following properties. (i) If B is m-restricted then l(B) is m-increasing. (ii) If B is m-increasing then l(B) is m-restricted. (iii) If B is a singleton board then l2 (B) = B, so l is an involution on singleton boards. Proof. (i) Let B = (b1 , . . . , bn ) be m-restricted and l(B) = (lt , . . . , l1 ). Keeping in mind that the subscripts in l(B) are decreasing, we wish to show that lp ≥ lp+1 + m. Let Bi be the set of cells in column i and let ci = |Bi ∩ Qp | and di = |Bi ∩ Qp+1 | for all i. Now lp − lp+1 = i (ci − di ) and ci − di ≥ 0 for all i. Since B is m-restricted, there is an index 31 k such that Bk has its highest cell in Qp . Let k be the largest such index. Using the fact that B is m-restricted again forces Bk+1 to have its highest cell in Qp+1 . Thus, using the m-restricted condition a third time, lp −lp+1 ≥ (ck −dk )+(ck+1 −dk+1 ) = (bk −(p−1)m)−0+m−(bk+1 −pm) = bk −bk+1 +2m ≥ m. which is what we wished to prove (ii) This is similar to (i), finding an upper bound for lp − lp+1 using the fact that, for an m-increasing board, there is at most one Bk with its highest square in Qp for each p. Details are left to the reader. (iii) Induct on |B|. Given B, let B be the board obtained by removing its first level. Since B is singleton so is B and thus, by induction, l2 (B ) = B . If l(B) = (lt , . . . , l1 ) then the definition of the l-operator shows that l(B ) = (lt , . . . , l2 ) and l1 = |B ∩ Q1 |. Applying l to l(B) we see that the column for l1 in l(B) adds m to every column of l2 (B ). Hence every column of l2 (B) which contains cells in Qp for p ≥ 2 agrees with the corresponding column in B. Also, again by definition of l, those columns of l2 (B) which lie wholly in Q1 are obtained from the column for l1 in l(B) by breaking it into columns of length m and a column of length ρm (l1 ). Since this is also the unique way to complete l2 (B) so that it is a singleton board, it must be that l2 (B) = B as desired. We now return to considering m-level rook equivalence classes as in Section 2.3. Using the proposition just proved, we obtain a second distinguished representative in each m-level rook equivalence class. Corollary 2.6.5. Every Ferrers board is m-level rook equivalent to a unique m-restricted 32 singleton board. Proof. By Theorem 2.3.6, we know that each class has a representative B which is mincreasing and so also a singleton board. Applying the previous proposition and Lemma 2.3.3, we see that l(B) is an m-restricted singleton board in the class. If there is a second such board B = l(B) then, by Proposition 2.6.4 again, l(B ) and l2 (B) = B will be distinct m-increasing singleton boards in the class, contradicting the uniqueness part of Theorem 2.3.6. 2.7 Weight equivalence class sizes In this section we will generalize Theorem 2.4.1 to m-weight equivalence classes. Theorem 2.7.1. Let B = (b0 , . . . , bN ) be a Ferrers board where N = |B|, and suppose n(ωm (B)) = (n0 , n1 , . . . ). The number of Ferrers boards in the m-weight equivalence class of B is i≥1 ni + ni−1 + · · · + ni−m − 1 . ni Proof. We use Proposition 2.6.1 and count the number of rearrangements of ω = ω(B) which correspond to a Ferrers board. Let d be the maximum value of an entry of ω. Our assumptions imply that d is nonnegative and all entries of ω are between between 0 and d inclusive. Consider ω which is obtained from ωm (B) by removing all values equal to d. Using Proposition 2.6.2, it is easy to see that ω = ωm (B ) for some Ferrers board B . By induction, it suffices to show that the number of rearrangements of ω which come from a given ω is nd + nd−1 + · · · + nd−m − 1 . nd 33 (2.7.1) By condition (ii) in the proposition just cited, we can insert d after any element of ω which is at least d − m. So the number of places for insertion is nd−m + · · · + nd−1 . Since we need to insert nd copies of d and more than one copy can go in a given place, the number of choices is given by (2.7.1). 34 Chapter 3 Bijections on m-level rook placements 3.1 Introduction The material in this chapter is a result of a collaboration with Nicholas Loehr, Jeffrey Remmel, and Bruce Sagan, and will appear in a forthcoming paper. The purpose of this chapter is to generalize the bijection of Foata and Sch¨ utzenberger and those of Loehr and Remmel to m-level rook placements. The remainder of this section gives the background terminology necessary to begin this task. In Section 3.2 we generalize the bijection used by Foata and Sch¨ utzenberger. Although this bijection is the composition of many intermediary bijections, and is therefore not direct, it does provide an explicit bijection between m-level rook placements on arbitrary m-level rook equivalent Ferrers boards. We will need this bijection again in Section 3.4. In Section 3.3 we generalize the construction of Loehr and Remmel. In this case the bijection can only be specified for singleton boards, a subset of all Ferrers boards. However, the construction leads to an explicit calculation of the m-level rook numbers for such boards using elementary symmetric functions and Stirling numbers of the second kind. In Section 3.4, we generalize a second bijection of Loehr and Remmel, and in doing so prove that any two m-level rook equivalent Ferrers boards have the same hit numbers. The last two bijections involve the Garsia-Milne Involution Principle [GM81]. Recall that we say that the ith column of B terminates in level p if p is the largest integer such that the ith column has non-empty intersection with Qp , which provides a character35 7 6 R 5 2 4 R 6 B= 7 5 3 1 4 2 3 1 R 4 2 3 1 6 4 2 5 3 1 R BS = R 7 4 4 3 3 2 2 1 1 R l(BS ) = Figure 3.1: On the left, a placement of two 2-level rooks on B. In the middle, the corresponding placement from Lemma 3.2.3 of two 2-level rooks on singleton board BS . On the right, the placement on l(BS ) from Lemma 3.2.5. ization of singleton boards, specifically, for any given level, a singleton board contains at most one column that terminates in that level and has a height that is not a multiple of m. The Ferrers board on the left in Figure 3.1 is not a singleton board, as two different columns terminate in the second level without having 2 cells in that level, while the Ferrers boards in the middle and on the right are singleton boards. 3.2 3.2.1 Rook equivalence and bijections Reduction to singleton boards In order to produce bijections between m-level rook placements on Ferrers boards, it is convenient to restrict our attention to singleton boards. In order to do this we prove the following two lemmas. First we show that for every Ferrers board there is a unique singleton board which has the same number of cells at each level. Then we prove that there is a bijection between the rook placements on a Ferrers board and those on the singleton board guaranteed in the first lemma. These lemmas together imply that every Ferrers board is 36 m-level rook equivalent to a singleton board and that there is an explicit bijection between the corresponding rook placements. Lemma 3.2.1. Given a Ferrers board B, there exists a unique singleton board BS which has the same number of cells at each level as B. Proof. Let B have lp cells in the pth level. In order for BS to be a singleton board with lp cells in the pth level, the cells of the pth level must be arranged uniquely as follows. If lp = cm + r with 0 ≤ r < m, then level p of BS must have one column with r cells followed on the right by c columns with a full m cells in the level. This is because a singleton board may have at most one column which intersects a given level non-trivially in fewer than m cells. Thus BS must be unique if it exists. In order to show that BS exists, we shall construct it. Arrange each level as specified above and line up the furthest right column in each level to create the furthest right column of BS . This yields a Ferrers board because every column which has any cells in the pth level of B must have a full m cells in the (p − 1)st level of B. Thus the total number of columns in the pth level of BS will be less than or equal to the number of columns in the (p − 1)st level of BS containing m cells at that level. Hence a singleton board BS exists and is unique. Ignoring the rook placement, Figure 3.1 shows a board B and its corresponding board BS . Since we know that an arbitrary Ferrers board B has the same number of cells at each level as a unique singleton board BS , we wish to provide an explicit bijection between rook placements on the two boards. In order to do so we require the following numbering on a Ferrers board. Definition 3.2.2. A level numbering of board B assigns a number to each cell of B in the following way. Proceeding level by level in B, number the cells in the level by numbering each 37 column from bottom to top, starting with the rightmost column and working left. In each level begin the numbering with 1. Figure 3.1 presents two examples of this numbering, on the left and middle boards, and also illustrates the bijection of the next lemma. Lemma 3.2.3. Given a Ferrers board B, there is an explicit bijection between m-level rook placements of k rooks on B and m-level rook placements of k rooks on BS , where BS is as constructed in Lemma 3.2.1. Proof. Give both B and BS a level numbering as shown in Figure 3.1. Since both boards have the same number of cells in each level, corresponding levels will each be numbered with the same set of numbers. Given any m-level rook placement on B, place rooks on BS initially so that each rook occupies the same numbered cell in the same level as it does in B. This may not provide an m-level rook placement on BS since two rooks could end up in the same column, so we will modify it as follows. Notice that if a rook in column i and level p of B is not in the same cell in BS , then column i must be to the left of a column of B that intersects Qp in less than m cells. Furthermore, if the rook ends up in column i in BS , then all columns in the interval [i, i ] have a full m rooks in levels below p in B. Thus, if any of the rooks that move create a column with two or more rooks, there will be exactly two rooks in the column and the upper rook will have moved while the lower rook remained stationary. To rectify the situation, whenever a rook is moved from column i in b to column i in BS , move all other rooks in columns (i, i ] one column to the left, preserving their row. This is possible since these columns must contain m cells in the lower level of both B and BS in order for the upper rook to have been in that column in B. Rearranging the rooks at each level in this fashion provides a function from 38 m-level rook placements on B to m-level rook placements on BS . Figure 3.1 illustrates this map on a rook placement, including moving a lower rook one column to the left. To see that this is a bijection, use the level numbering to produce a set of rooks on B from those on BS . All the rooks will return to their initial positions once the appropriate right shift is applied. Similarly one can show that applying the map first to BS and then to B is the identity. Thus we have a bijection between m-level rook placements on B and m-level rook placements on BS . Lemma 3.2.1 and Lemma 3.2.3 guarantee that every Ferrers board is m-level rook equivalent to a singleton board. Additionally, there is an explicit bijection between m-level rook placements on the two boards. This permits us to restrict our attention to singleton boards henceforth. 3.2.2 The l-operator Transposition of boards plays a central role in the Foata-Sch¨ utzenberger construction of bijections between rook-equivalent Ferrers boards when m = 1. We will use the l-operator, defined in Section 2.3 of the previous chapter, as a generalization of this operation for arbitrary m. Definition 3.2.4. Given a Ferrers board B the l-operator applied to B is defined as follows. If t is the largest index of a non-empty level of B and the number of cells in the pth level of B is lp , then l(B) = (lt , lt−1 , . . . , l1 ). Figure 3.1 contains an example board BS as well as l(BS ). The fact that l(B) is a Ferrers board was shown in Lemma 2.3.3 and is analogous to an argument in the proof of 39 Lemma 3.2.1. In particular, if B is a Ferrers board then its pth level must fit above its (p − 1)st level which implies lp m ≤ lp−1 m , with strict inequality if lp ≡ 0 mod m. It follows that l(B) is a weakly increasing sequence and so l(B) is a Ferrers board and, because of the strict inequality for non-multiples of m, a singleton board. To see that the l-operator is a generalization of transposition, note that if m = 1 then the levels of B are individual rows and these become the columns of l(B). Furthermore, when restricted to the set of singleton boards the l-operator is an involution. This is shown in Proposition 2.6.4 of the previous chapter. Thus, the l-operator is a surjection from the set of Ferrers boards onto the set of singleton boards with B = l(l(B)) when B is singleton. We now provide a bijection between m-level rook placements on B and m-level rook placements on l(B) to generalize the well-known bijection for transposition. Lemma 3.2.5. Given a singleton board B and a non-negative integer k, there is an explicit bijection between m-level rook placements of k rooks on B and m-level rook placements of k rooks on l(B). Proof. Give B a level numbering, then number the columns of l(B) from bottom to top beginning with the number 1 in each column. Note that in this case the numbering of a level of B will consist of the same set of numbers as the numbering of the corresponding column of l(B). Assume that B has t non-empty levels. For a given m-level rook placement of k rooks on B, place rooks on l(B) in the following way. If a rook was in the cell numbered n of level p in B, then place a rook in the cell numbered n in column t − p + 1 in l(B). See Figure 3.1 for an example of this map for a 2-level placement. 40 We must show that this gives a valid m-level rook placement on l(B). If two rooks end up in the same column of l(B) they must have originated in the same level of B, contradicting having an m-level rook placement on B. Similarly, if two rooks end up in the same level of l(B), then they must have originated in the same column of B, since B is a singleton board. The inverse of this map acts as follows. If a rook is in the cell numbered a of column t − p + 1 in l(B) then it is placed in the cell numbered a in level p of B. The proof that this gives a rook placement is similar to the one in the previous paragraph and so is omitted. Note that Lemma 3.2.3 and Lemma 3.2.5 combine to provide an explicit bijection between m-level rook placements on any Ferrers board B and on its m-transpose, l(B) = l(BS ). 3.2.3 The local l-operator For any set S, let #S be the cardinality of S. Given a column i and a level p define the m-arm length of column i, level p by armlm (i, p) = #{ci,j ∈ B | ci,j is strictly above level p}. In Figure 3.2 the cells counted by the 2-arm length of column 4, level 1 have a dashed line through them. (Reflecting our boards to put them in English notation will result in the arm being the usual set of squares when m = 1.) We let armlm (i, p) = ∞ if the number of columns in B is less than i for reasons detailed in Lemma 3.2.7. Similarly, define the m-leg length of column i, level p to be leglm (i, p) = #{ci ,j ∈ B | ci ,j is in level p and i < i}. 41 Figure 3.2: The dashed line goes through the cells counted by the 2-arm length of the fourth column and first level, and the shaded cells are counted by the corresponding 2-leg length. The cells counted by the 2-leg length of column 4, level 1 are shaded in Figure 3.2. As before, this is equivalent to the usual notion of leg length in the m = 1 case. We also let leglm (i, 0) = ∞ by convention. Since the l operation generalizes the transposition of a Ferrers board, one would expect that some sort of local l operation would be the appropriate generalization of the local transposition introduced by Foata and Sch¨ utzenberger. This is indeed the case, and we define the local l operation as follows. Given a Ferrers board B with non-empty intersection of the ith column and pth level, let Bi,p denote the subboard of B consisting of all cells in or above the pth level and in or to the left of the ith column: see Figures 3.3 and 3.4 for examples. Note that if B is a singleton board, then Bi,p is also, because the set of rows in level p of Bi,p will be the same as the set of rows in level p + p − 1 of B. If B is a Ferrers board then the local l operation at (i, p) is the result of applying the l operator to the subboard Bi,p and leaving the rest of B fixed. We will denote the resulting board by li,p (B). As defined above li,p (B) may not be a Ferrers board, let alone a singleton board. We now develop a pair of conditions to determine if li,p (B) will be a singleton board. Definition 3.2.6. The operation li,p is permissible for a singleton board B if armlm (i, p) ≤ leglm (i, p − 1) m and 42 leglm (i, p) ≤ armlm (i + 1, p) m . l3,2 (B) = B= Figure 3.3: On the left, B3,2 is shaded within B = (1, 4, 4, 5). Notice that l3,2 is not permissible for B since armlm (4, 2) m < leglm (3, 2), which means l3,2 (B) will not be a singleton board. On the right the shaded cells in l(B3,2 ) illustrate this; in this case l3,2 (B) is not even a Ferrers board. See Figure 3.3 for an example of a local l-operation not permissible for the given board, and Figure 3.4 for a local l-operation which is permissible. Lemma 3.2.7. Let a singleton Ferrers board B have a non-empty intersection of the ith column and pth level. Then li,p is permissible for B if and only if li,p (B) is a singleton Ferrers board. Proof. If column i, level p in B contains fewer than m cells, then li,p (B) = B since B is singleton, and there is nothing to prove. Henceforth, assume that column i, level p in B contains m cells. We know that B, Bi,p , and l(Bi,p ) are all singleton Ferrers boards. It follows that li,p (B) will be a singleton Ferrers board if and only if these three conditions hold for the board li,p (B). (a) The lowest row of level p is weakly shorter than the highest row of level p − 1; (b) column i is weakly lower than column i + 1; and (c) if columns i and i + 1 terminate at the same level, then the height of column i + 1 is a multiple of m. Condition (c) is needed to ensure li,p (B) will be singleton. 43 To determine when these conditions hold, first note that applying li,p to B exchanges armlm (i, p) and leglm (i, p). Because B is singleton, the top row of level p − 1 in B (and in li,p (B)) extends left of column i by leglm (i, p − 1) m cells. On the other hand, the new bottom row of level p in li,p (B) extends left of column i by armlm (i, p) m cells. Thus, condition (a) holds iff armlm (i, p) m ≤ leglm (i, p − 1) m . Since both sides are multiples of m, this inequality is equivalent to armlm (i, p) ≤ leglm (i, p− 1) m , which is the first condition in the definition of permissibility. Now consider the heights of columns i and i + 1 in li,p (B). Both column i and column i + 1 have a full m cells in level p. So, in both B and li,p (B), column i + 1 has height pm + armlm (i + 1, p). On the other hand, the new column i in li,p (B) has height pm + leglm (i, p). So condition (b) will hold iff leglm (i, p) ≤ armlm (i + 1, p). To deal with condition (c), consider two cases. First suppose that armlm (i + 1, p) is a multiple of m. Then condition (c) must hold, and here condition (b) will hold iff leglm (i, p) ≤ armlm (i + 1, p) m . Now suppose that armlm (i + 1, p) is not a multiple of m. Given that condition (b) holds, the new board li,p (B) will be singleton iff the strengthened inequality leglm (i, p) ≤ armlm (i + 1, p) m is true. Thus, this last inequality is equivalent to the truth of (b) and (c) in all cases. 44 3.2.4 The Local l-operation on an m-level rook placement Since there is a bijection between rook placements on B and l(B) when B is singleton, it stands to reason that it would generalize to a bijection between rook placements on B and li,p (B). The following lemma makes this precise. Lemma 3.2.8. For a singleton board B, suppose li,p is permissible for B. Then there is an explicit bijection between m-level rook placements of k rooks on B and m-level rook placements of k rooks on li,p (B). Proof. Use the bijection induced by the l operation in Lemma 3.2.5 on the subboard transposed by li,p , not moving the rooks on the part of board B which is fixed. However, this may cause a rook in the transposed subboard to occupy the same column or level of li,p (B) as one of the rooks which was fixed. We deal with this possibility next. In order for two rooks to end up in the same column, there must be rooks placed on B beneath Bi,p , so we can assume p > 1 without losing generality. Consider the set of columns of B which do not contain rooks in Bi,p , and the set of columns of li,p (B) which do not contain rooks in l(Bi,p ). By our assumption on p, these two sets have the same cardinality and so we can put a canonical bijection on them by pairing the leftmost columns in each set and moving to the right. If there is a rook lower than level p in one of these columns of B, use this bijection on the columns to move it to the cell in the same row of the corresponding column of li,p (B). After doing so, there must be at most one rook in each column of li,p (B). For example, in Figure 3.4 the rook in c3,2 is in the second column from the left of B which does not contain a rook in B4,2 . Thus it moves to column 2, which is the second column from the left of l4,2 (B) that does not contain a rook in l(B4,2 ). If two rooks end up in the same level we treat them similarly where we can assume, 45 without loss of generality, that the i-th column is not the rightmost column of B. There is a canonical bijection between the levels of B which do not contain rooks in Bi,p and those of li,p (B) that do not contain rooks in l(Bi,p ). Adjust the levels of all rooks to the right of column i using this bijection, fixing the column of the rook that moves. Furthermore, fix the height of the rook that moves within the level, that is, if the rook was in cell cx,y , move the rook to cell cx,y in the appropriate level with y ≡ y (mod m). Note that since B is a singleton board, columns to the right of column i will contain a full m cells at any level which contained a rook in the subboard Bi,p . To see that this is a bijection, we construct its inverse. Recall that the l operator is an involution on singleton boards. Thus, since Bi,p is a singleton subboard, li,p (li,p (B)) = B. Similarly, applying the bijection from Lemma 3.2.5 and then its inverse returns the original placement of rooks on Bi,p . All that remains to check is that any rooks moved outside of Bi,p return to their original cells. Since the rooks return to their original placement on Bi,p , the set of columns that gain a rook in Bi,p after the first application of l will be the same set as those that lose a rook in Bi,p after the second application of l. Thus the bijection on the columns induced by the first application of l will be the inverse of the bijection induced by the second application, and any rook required to move in li,p (B) will move back in li,p (li,p (B)). A similar argument holds for levels, noting that since li,p (B) is singleton ensures that any level that gains a rook in Bi,p after applying li,p contains a full m cells in every column to the right of column i. Thus this yields a bijection between rook placements on B and li,p (B). Figure 3.4 illustrates this bijection. 46 6 5 R 4 1 3 2 R B= 6 4 2 5 3 1 R 1 l4,2 (B) = R R 1 R Figure 3.4: On the left, B4,2 is shaded. Here l4,2 is permissible for B and is shown on the right. 3.2.5 Bijections with m-increasing boards Foata and Sch¨ utzenberger proved there is a unique Ferrers board in every rook equivalence class whose column lengths are strictly increasing and used this board as a target for their bijections. To accomplish the same thing, we need the the analogous concepts for m-level rook placements, developed in the previous chapter. Definition 3.2.9. A Ferrers board B = (b1 , b2 , . . . , bn ) is called m-increasing if bi+1 ≥ bi +m for all 1 ≤ i ≤ n − 1. Recall that when m = 1 strictly increasing and m-increasing are equivalent. Theorem 3.2.10 (2.3.6). Every Ferrers board is m-level rook equivalent to a unique mincreasing board. We are now almost ready to prove the main result of this section, Theorem 3.2.12 below. However, to do so we must put an order on Ferrers boards. Once we have established this order, we will be able to give an explicit bijection between m-level rook placements on an arbitrary Ferrers board B and on an m-level rook equivalent Ferrers board which 47 is greater than B in this order, if such a board exists. Additionally, the set of all Ferrers boards equivalent to B will have a unique maximum element under this order, namely the m-increasing board guaranteed by [BLRS14]. To define this order, if B = (b1 , . . . , bn ) then consider the reversal of B, B r = (bn , . . . , b1 ). Now let B < B if B r is lexicographically smaller than (B )r . It is important to note that when applying Lemma 3.2.3 we will always have BS ≥ B (3.2.1) since in BS all the cells in each level are as far to the right as possible. Lemma 3.2.11. Given a singleton board B containing a column i and a level p with the property that armlm (i, p) < leglm (i, p), (3.2.2) there is a singleton board B = li ,p (B) with i ≥ i and B > B. Furthermore, if B is not m-increasing then a column i and level p satisfying equation 3.2.2 must exist. Proof. To prove the first statement, let i ≥ i be the maximum index such that armlm (i , p) < leglm (i , p). Note that by our convention on armlm , we must have that i is at most the number of columns of B. We claim that it suffices to show that li ,p is permissible for B. This is because if li ,p is permissible for B, then the resulting board B must satisfy B > B. Indeed, li ,p (B) increases the length of column i by leglm (i , p) − armlm (i , p), which must be greater than 0, and column i is the rightmost column affected by li ,p (B). Thus B > B. If li ,p is not permissible for B, then we claim that we have armlm (i +1, p) < leglm (i +1, p) 48 which will contradict the maximality of i and complete this part of the proof. Note that armlm (i , p) < leglm (i , p) ≤ leglm (i , p − 1) m . So li ,p not being permissible for B implies that armlm (i +1, p) m < leglm (i , p) = leglm (i + 1, p) − m since B is singleton. This implies the desired contradiction that armlm (i + 1, p) < leglm (i + 1, p). To prove the second statement of the theorem, note that if B is not m-increasing there are two possible cases, either there are two adjacent columns i − 1, i of B which terminate at the same level, or column i − 1 terminates in level p and B has exactly r1 cells in the pth level of column i − 1 and exactly r2 cells in the (p + 1)st level of column i where r1 > r2 > 0. Case 1: Let columns i − 1 and i both terminate at level p. Then armlm (i, p) = 0, by the assumption that column i terminates at level p, but leglm (i, p) ≥ 1 since column i − 1 also terminates at the pth level. Thus armlm (i, p) < leglm (i, p) as desired. Case 2: By assumption armlm (i, p) = r2 < r1 ≤ leglm (i, p) which completes the proof. We are now in a position to prove our main theorem of this section. Theorem 3.2.12. Given any two m-level rook equivalent Ferrers boards, there is an explicit bijection between m-level rook placements of k rooks on them. Proof. Given any Ferrers board B, let Bm be the unique m-increasing board in the m-level rook equivalence class of B guaranteed by Theorem 3.2.10. It suffices to show that there is an explicit bijection between the m-level rook placements of k rooks on B and those on Bm . This is trivial if B = Bm so assume B = Bm . By Lemma 3.2.3, we have an explicit bijection 49 R R R R R (a) R R (b) (c) R R R (d) R R Figure 3.5: (a) A 2-level rook placement on a Ferrers board. (b) The placement on the singleton board obtained after applying Lemma 3.2.3. (c) The placement obtained after applying Lemma 3.2.8 using l3,1 . (d) The placement obtained on a 2-increasing board after applying Lemma 3.2.8 again using l3,3 . between the placements on B and those on BS where BS ≥ B by equation 3.2.1. If BS = Bm then we are done. Otherwise, apply the local l operator defined in Lemma 3.2.11 which will give B = li,p (BS ) with B > BS and, by Lemma 3.2.8, another explicit bijection between rook placements. We now repeat this process if necessary. Since there are only finitely many boards in an m-level rook equivalence class and the lexicographic order increases with every application of Lemma 3.2.8, we must eventually terminate. And, by Lemma 3.2.11 again, termination must occur at Bm . Composing all the bijections finishes the proof. See Figure 3.5 for a short example of this process. 3.3 A second bijection on m-level rook placements Our next two main results will require the Garsia-Milne Involution Principle. The first will use the Involution Principle to construct another explicit bijection between two arbitrary m-level rook placements of k rooks on m-level rook equivalent singleton boards. Theorem 3.3.1 (Garsia-Milne Involution Principle [GM81]). Consider a triple (S, T, I) where S is a signed set, I is a sign-reversing involution on S, and the set T of fixed points of 50 I is required to be a subset of the positive part S + of S. Let (S , T , I ) be defined similarly. Then, given an explicit sign-preserving bijection f from S to S , one can construct an explicit bijection between T and T . The way that Garsia and Milne define the explicit bijection is as follows. Start with an element t ∈ T ⊆ S + . If f (t) ∈ T , then apply (f ◦ I ◦ f −1 ◦ I ) to f (t). This takes f (t) ∈ S + to S − , then to S − , then S + , and finally back to S + . Iterating this procedure must ultimately yield an element of T which is considered the image of t under the desired bijection. 3.3.1 A Garsia-Milne bijection for rook placements We will use the Involution Principle to construct a bijection between m-level rook placements on two m-level rook equivalent singleton boards. We must first construct a signed set and a sign-reversing involution so that the m-level rook placements are the fixed points under the involution. We do this as follows. Given two Ferrers boards, B and B , we shall say B fits inside B if juxtaposing the two boards with their lower right cells in the same position makes the cells of B a subset of the cells of B . Figure 3.6 shows that the thick bordered B = (2, 3) fits inside B = (0, 2, 4, 6). The shading and rook placement may be ignored for now. Let ∆n,m denote the triangular Ferrers board (0, m, 2m, . . . , (n − 1)m). Given a singleton board B, fix N large enough that B fits inside ∆N,m . If B has fewer than N columns, expand B on the left with columns of height zero so B = (b0 , b1 , . . . , bN −1 ) has the same number of columns as ∆N,m . Fix a non-negative integer k with k < N and let the integer i vary over 0 ≤ i ≤ k. Then S will consist of all configurations C constructed as follows. Take ∆N,m with B fitting inside and 51 W C= R W I(C) = ∈S R W W f (C) = R W f (I(C)) = R R W R Figure 3.6: On the top left, an element in S with sign −1. On the top right, the image under I which has sign +1. Beneath each board is its image under f . place white rooks W in i cells of ∆N,m that are outside of B so that no two white rooks are in the same column. Next, place k − i black rooks R forming an m-level rook placement on the subboard ∆N −i,m which is located in the columns of ∆N,m which do not contain a white rook. We will call this the inset ∆N −i,m board. Note that the columns of the inset ∆N −i,m may not be contiguous. See the top left board of Figure 3.6 for an example of such an object C where m = 2. The singleton board B = (0, 0, 2, 3) fits inside ∆4,2 . Here k = 3 < 4 and there is i = 1 white rook on the board ∆4,2 \ B and k − i = 2 black rooks on the board ∆3,2 which is represented by the grey shaded cells inside ∆4,2 . The rooks on ∆3,2 form a 2-level rook placement, but there is both a black rook and a white rook in the second level of ∆4,2 . Note that each column of ∆N,m may contain at most one white rook or black rook. On the other hand, a level of ∆N,m will contain at most one black rook, but may contain any number of white rooks. Further, define the sign of such a placement to be (−1)i . The sign 52 of the placement on the top left in Figure 3.6 is −1. To define I on an element C ∈ S, if all rooks of C are in B, and therefore black, then C is a fixed point. Otherwise, examine the columns of C from left to right until coming to a column with a rook outside of B. If that column contains a black rook, change the rook to a white rook, increase i by one, and move every black rook above and to the right of the cell containing the new white rook down m cells. If that column contains a white rook, change it to a black rook, decrease i by one, and move every black rook to the right and at the same level or higher as the new rook up m cells. The placement on the top right in Figure 3.6 illustrates what happens to the board on the left under I. Similarly, I takes the placement on the right to the placement on the left. We must show that I(C) will be an element of S. Clearly each column has at most one rook. When a black rook becomes white, it is clear that there will be at most one black rook per level on the resulting board, as there was at most one black rook per level on the original board. If a black rook is added, each level will have at most one black rook since when a black rook is added all black rooks at its level or above to the right of the new rook move up one level. Furthermore, there can be no black rooks at the same level or higher to the left of the new black rook if we change a white rook to a black rook. This is because the new black rook was a white rook which, by definition, was above board B. Since B is a singleton board, no columns of B to the left of the white rook in question will terminate in the level of the white rook. Thus if there were a black rook at the same level or higher to the left, it too would be outside of board B, which contradicts the white rook being the leftmost rook outside of board B. When a black rook is added, the black rooks must be placed on a board ∆N −i+1,m where the column in which the new black rook is placed is added to the columns in the initial inset 53 ∆N −i,m . Since there are no white rooks to the left of the new black rook, there will be no omitted columns to the left of the column containing the new black rook, thus all cells of that column will be in the inset ∆N −i+1,m and the new rook must be inside ∆N −i+1,m . This means that all the columns to the right of the new black rook that do not contain a white rook will contain m more squares as the inset ∆N −i+1,m than they did as the inset ∆N −i,m . Thus moving black rooks to the right of the new black rook up m cells will keep them within the new ∆N −i+1,m . Similarly, changing a black rook to a white rook will decrease the number of cells in the columns of ∆N −i−1,m to the right of the new white rook by m, but all black rooks to the right of the new white rook and at a higher level than it are moved down m cells, so they will be in ∆N −i−1,m because they were in ∆N −i,m originally. Finally, if there are any black rooks below the level of the new white rook but to its right, they will remain in ∆N −i−1,m because the first column in ∆N −i−1,m to the right of the new white rook must go up to at least the level of the new white rook since previously it was a black rook contained in ∆N −i,m . By construction, I is an involution. The fixed set of I will be denoted T . It is the set of all configurations which only have rooks on the subboard B and, by definition, these rooks must be black. As such, T is equal to the set of m-level rook placements of k rooks on B. Furthermore, if a board is not in T , then I either increases or decreases the number of white rooks on the board by one. Either way I will change the sign of the board. And if a board is in T , then it has positive sign. Given a singleton board B , define N , S , T , and I similarly for B contained in ∆N ,m . Without a loss of generality, assume N = N . Let B = (b0 , b1 , . . . , bN −1 ). If B and B are m-level rook equivalent singleton boards we can use I and I to construct an explicit bijection between m-level rook placements of k rooks on B and m-level rook placements of k rooks on 54 B . We do this by constructing a sign-preserving bijection between S and S . We will need the following characterization of when two singleton boards are m-level rook equivalent. The root vector of B is ζm = (−b0 , m − b1 , . . . , (N − 1)m − bN −1 ). The following result of Briggs and Remmel determines when two singleton boards are m-level rook equivalent simply by considering their root vectors. Theorem 3.3.2 (Briggs-Remmel Theorem 2 [BR06]). Two singleton boards are m-level rook equivalent if and only if they have the same root vector up to rearrangement for a sufficiently large N . We are now ready to apply the Garsia-Milne Involution Principle. Theorem 3.3.3. Let B and B be m-level rook equivalent singleton boards. Then there exists an explicit Garsia-Milne bijection between m-level rook placements of k rooks on B and m-level rook placements of k rooks on B . Proof. By Theorem 3.3.1 and what we have already established, it suffices to find a signpreserving bijection f : S → S . We construct f as follows. For clarity of notation, let B be placed in ∆N,m and B be placed in a copy ∆N,m of ∆N,m . Notice that the kth element of the root vector of B, km − bk , is the number of cells in the kth column of ∆N,m which lie outside of board B. Since B and B are m-level rook equivalent, the root vector for B is a rearrangement of the root vector for B. Therefore there is a length-preserving bijection between the columns of the set difference ∆N,m \B and the columns of ∆N,m \ B which takes the leftmost column of a given length in ∆N,m \ B 55 R R R 1 2 3 4 5 6 5 4 3 2 1 Figure 3.7: The placement corresponding to partition {1, 4}, {2, 3, 5}, {6}. to the leftmost column with that length in ∆N,m \ B and so forth. This bijection induces a bijection on the placement of the white rooks. If a white rook appears in the jth cell above B, place a white rook in the jth cell above B in the associated column. Once all the white rooks are placed, create a copy of ∆N −i,m inside of ∆N,m using the columns which do not contain a white rook. Place the black rooks on the board with relation to the ∆N −i,m subboard exactly as they are placed on the original board with relation to the original ∆N −i,m subboard. Each placement on the bottom of Figure 3.6 is the image under f of the corresponding placement on the top where B = (0, 0, 2, 3) and B = (0, 0, 1, 4). Notice that in the top left board, the white rook is at the top of the second column from the left which has two cells above B. In the board on the bottom left the white rook is still at the top of the second column from the left which has two cells above B . Under this map the white rooks must be placed inside ∆N,m but outside B , and the black rooks are placed inside ∆N −i,m , so f maps S to S . Further this map preserves the number of white rooks placed on the board, so it is sign preserving. Therefore we may conclude from the Involution Principle that there is an explicit bijection between m-level rook placements of k rooks on B and m-level rook placements of k rooks on B . To obtain a consequence of this construction, we will need some background on symmetric functions and Stirling numbers. For d ≤ n both non-negative integers, let ed (x1 , x2 , . . . , xn ) 56 denote the elementary symmetric function of degree d in n variables, that is, ed (x1 , x2 , . . . , xn ) = 1≤i1