QUOTIENT POSETS AND THE CHARACTERISTIC POLYNOMIAL By Joshua William Hallam A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Mathematics - Doctor of Philosophy 2015 ABSTRACT QUOTIENT POSETS AND THE CHARACTERISTIC POLYNOMIAL By Joshua William Hallam In this dissertation we consider the generating function for the M¨obius function of finite partially ordered sets. This generating function is called the characteristic polynomial of the partially ordered set. We are primarily interested in explaining why certain families of partially ordered sets have characteristic polynomials where all the roots are nonnegative integers. To this end we introduce the concept of a homogeneous quotient. These quotients allow us to give new proofs of some well-known results in the literature as well as give generalizations of them. We finish by showing how to use homogeneous quotients to give unified proofs of some classic results about the M¨obius function. ACKNOWLEDGMENTS First and foremost I would like to thank my advisor, Dr. Bruce Sagan, for all the support and encouragement he has given me. It has truly been an honor to learn from him both about mathematics as well as how to be a mathematician. I would also like to thank all the graduate students at Michigan State, past and present. In particular, I want to thank Dave Bramer for all the encouragement over the years. Additionally, I would like to thank my parents, William and Kelly Hallam, who always encouraged learning and supported me in all my endeavors. Finally, I would like to thank my wife, Mobashera Hallam. Without her constant support, I would have never been able to complete this work. iii TABLE OF CONTENTS LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 1 Introduction . . . . . . 1.1 Poset Basics . . . . . . . . . . 1.2 The M¨obius Function . . . . . 1.3 The Characteristic polynomial . . . . . . . . v . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1 7 10 Chapter 2 Quotient Posets . . . . . . . 2.1 The Basics . . . . . . . . . . . . . . 2.2 The Standard Equivalence Relation 2.3 Rooted Trees . . . . . . . . . . . . 2.4 Partitions Induced by a Multichain 2.5 Transversal Functions . . . . . . . . 2.6 Complete Transversal Functions . . 2.7 Crosscut-simplicial Lattices . . . . 2.8 LL lattices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 16 20 28 32 41 47 53 57 Chapter 3 Classic Results About the M¨ obius Function . . . . . . . . . . . . 62 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 iv LIST OF FIGURES Figure 1.1: The Boolean algebra, B3 . . . . . . . . . . . . . . . . . . . . . . . . 2 Figure 1.2: The poset of binary words, P3 . . . . . . . . . . . . . . . . . . . . . 3 Figure 1.3: The claw with n atoms . . . . . . . . . . . . . . . . . . . . . . . . . 13 Figure 1.4: Hasse diagrams for the partition lattice example . . . . . . . . . . . 14 Figure 2.1: Hasse diagrams for partition lattice example with new labelings . . . 21 Figure 2.2: Hasse diagrams for rooted trees . . . . . . . . . . . . . . . . . . . . 29 Figure 2.3: A lattice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 Figure 2.4: The Tamari lattice T3 . . . . . . . . . . . . . . . . . . . . . . . . . . 41 Figure 2.5: The weighted partition poset Πw 3 . . . . . . . . . . . . . . . . . . . . 51 Figure 2.6: The Tamari lattice T3 with an SB-labeling . . . . . . . . . . . . . . 56 Figure 2.7: A crosscut-simplicial lattice which does not have an SB-labeling . . 57 v Chapter 1 Introduction 1.1 Poset Basics We will use the following standard notation for certain sets of numbers. • The nonnegative integers will be denoted by N. • The integers will be denoted by Z. • The set {1, 2, . . . , n} will be denoted by [n]. Let us now review some background material on partially ordered sets. See [14, Chapter 3] for a more complete overview. Definition 1.1.1. A partially ordered set or poset is a set P and binary relation ≤ on P such that the following three properties hold for all x, y, z ∈ P : 1. x ≤ x (reflexivity), 2. if x ≤ y and y ≤ x then x = y (antisymmetry), 3. if x ≤ y and y ≤ z then x ≤ z (transitivity). It is common to represent the poset P with binary relation ≤ as either (P, ≤) or just P if the binary relation is clear from context. Additionally, it will sometimes be useful to write ≤ as ≤P , especially in the case when several posets are being considered. 1 {1, 2, 3} {1, 2} {1, 3} {2, 3} {1} {2} {3} ∅ Figure 1.1: The Boolean algebra, B3 We now consider an example of a poset. Let Bn be the set of subsets of {1, 2, . . . , n}. For S, T ∈ Bn , define a binary relation ≤ on Bn as S ≤ T if and only if S ⊆ T . It is easy to verify that all three properties of a poset are satisfied and so (Bn , ⊆) is a poset. It is called the Boolean algebra. It is often useful to consider a graphical representation of a poset. We do this by using Hasse diagrams. Before we can define such a diagram, we will need to develop some more terms associated with posets. If P is a poset and x, y ∈ P , we say y covers x, and write x y, if x < y and there is no z with x < z < y. We call a poset finite if the underlying set is finite. Given a finite poset, P , the Hasse diagram of P is a directed graph such that the vertices of the graph are the elements of P and there is a directed edge from x to y whenever x y in P . We typically do not write arrows on the directed edges, instead we draw the graph so that the edges are directed upwards. Figure 1.1 depicts the Hasse diagram of B3 . Let P and Q be posets. We say a map ϕ : P → Q is order preserving if x ≤P y implies that ϕ(x) ≤Q ϕ(y) for all x, y ∈ P . We say P and Q are isomorphic if there is a bijection ϕ : P → Q such that ϕ and ϕ−1 are order preserving. In other words, we must have x ≤P y 2 111 110 101 011 100 010 001 000 Figure 1.2: The poset of binary words, P3 if and only if ϕ(x) ≤Q ϕ(y) for all x, y ∈ P . If P and Q are isomorphic, then we write P ∼ = Q. If P and Q are finite posets, then checking if P and Q are isomorphic is equivalent to checking that their Hasse diagrams are isomorphic as directed graphs. We now give an example of a poset isomorphism. Let Pn be the set of all binary sequences of length n ordered by saying a1 a2 . . . an ≤ b1 b2 . . . bn if and only if ai ≤ bi for all i. We claim that Pn and Bn are isomorphic. To see why, define ϕ : Pn → Bn so that ϕ(a1 a2 . . . an ) is the subset of {1, 2 . . . , n} obtained by including i if and only if ai = 1. So for example, ϕ(10100) = {1, 3}. First, it is obvious that ϕ is a bijection. Now suppose that a1 a2 . . . an ≤ b1 b2 . . . bn , then bi = 0 implies that ai = 0. Therefore, ϕ(a1 a2 . . . an ) only contains elements that are also in ϕ(b1 b2 . . . bn ) and so ϕ is order preserving. Finally, suppose that S ⊆ T and let ϕ−1 (S) = a1 a2 . . . an and ϕ−1 (T ) = b1 b2 . . . bn . Since S ⊆ T it must be that bi = 0 implies that ai = 0 for all i. This implies that a1 a2 . . . an ≤ b1 b2 . . . bn . It follows that Pn ∼ = Bn . Figure 1.2 is the Hasse diagram of P3 . Comparing this with Figure 1.1 gives a graphical verification that the two posets are indeed isomorphic when n = 3. 3 It will be useful to have names for some special elements of a poset. If a poset, P , has a minimum element then we denote it by ˆ0P or just ˆ0 if P is clear from context. Similarly if P has a maximum element we denote it by ˆ1P or ˆ1. In the example of Bn , ˆ0 is ∅ since every set contains the empty set. Also, the ˆ1 of Bn is {1, 2, . . . , n} because every set is contained in this set. For the rest of the paper, unless otherwise noted, we will assume that all of our posets are finite and have a ˆ0. We, however, do not assume that all of our posets have a ˆ1. Another set of important elements of a poset are the atoms. Let P be a poset, we say a ∈ P is atom if a ˆ0. The atoms of Bn are exactly the singleton subsets of {1, 2, . . . , n}. The set of atoms of a poset P will be denoted by A(P ). Let P be a poset and let S be a subset of P . The lower order ideal generated by S is L(S) = {x ∈ P | x ≤ s for some s ∈ S}. Similarly, we have the upper order ideal generated by S which is defined by U (S) = {x ∈ P | x ≥ s for some s ∈ S}. We say a set, S, is totally ordered if x, y ∈ S implies that either x ≤ y or y ≤ x. A totally ordered subset of a poset is called a chain of the poset. If C : x0 < x1 < · · · < xn is a chain, we say C is saturated if xi chain in Bn is ∅ {1} {1, 2} xi+1 for all i = 0, 1, . . . , n − 1. An example of a saturated ... {1, 2, . . . , n − 1} {1, 2, . . . , n − 1, n}. In terms of a Hasse diagram, a saturated chain of a poset is a directed path. The length of a chain, C, is |C| − 1 where | · | denotes cardinality. The previous example of the saturated chain in Bn has length n. We will also need the notion of a multichain. A multichain of a poset is like 4 a chain where we allow repetition of the elements. We say a multichain is saturated if the underlying set forms a saturated chain. If P has a ˆ0 then we say P is ranked if, for each x ∈ P , every saturated ˆ0–x chain has the same length. Given a ranked poset, we get a rank function ρ : P → N defined by setting ρ(x) to be the length of a ˆ0–x saturated chain. We then define ρ(P ) = max ρ(x). x∈P In the case when P has a ˆ1, we have that ρ(P ) = ρ(ˆ1). Returning to our example of Bn , one can see that ρ(S) = |S| for all S ∈ Bn and that ρ(Bn ) = n. If P and Q are posets, we define the product of P and Q written as P × Q as the set P × Q with the binary relation ≤P ×Q such that (p1 , q1 ) ≤ (p2 , q2 ) if and only if p1 ≤p p2 and q1 ≤Q q2 . It is not hard to prove the following lemma. Lemma 1.1.2. Let P and Q be posets, then P × Q is a poset. Moreover, if both P and Q are ranked, then P × Q is ranked as well. In this case, the rank function is given by ρP ×Q (p, q) = ρP (p) + ρQ (q). Our running example, Bn , is in fact a product of smaller posets. To see why, define Ck to be the poset whose elements are {0, 1, . . . , k} and whose order is the normal ordering of the integers. We call Ck the chain of length k. For the Boolean algebra we have that Bn ∼ = C1 × C1 × · · · × C1 . The isomorphism is essentially the isomorphism we gave before n times between Bn and the binary sequences of length n. 5 Given two elements x, y of a poset P , the meet or greatest lower bound of x and y (if it exists) is the element z such that z ≤ x, y with the property that if w ≤ x, y, then w ≤ z. It is denoted by x ∧ y. The join of x and y or least upper bound (if it exists) is the element z such that z ≥ x, y with the property that if w ≥ x, y , then w ≥ z. It is denoted by x ∨ y. A poset is called a lattice if every pair of elements has a meet and a join. Note that in the case of finite posets this is equivalent to saying that every subset of elements has a meet and a join. Moreover, note that every lattice has a ˆ0 and a ˆ1. The poset Bn is a lattice with S ∧ T = S ∩ T and S ∨ T = S ∪ T . Let L be a lattice. We say an element x ∈ L is atomic if there exists atoms a1 , a2 , . . . , an where x = a1 ∨ a2 ∨ · · · ∨ an . As a convention, we will let ˆ0 be the join of the empty set. We say a lattice, L, is atomic if every element of L is atomic. The lattice Bn is atomic since the empty set is the empty union and since every nonempty set is the union of singleton sets. On the other hand, we claim that for k > 1, the chain Ck is not atomic. Recall, that there is at most one atom in any Ck . It follows that there are at most 2 elements of Ck which are atomic. Therefore if k > 1, then Ck is not atomic. Let L be a ranked lattice. We say L is (upper) semimodular if for all x, y ∈ L we have ρ(x) + ρ(y) ≥ ρ(x ∧ y) + ρ(x ∨ y). Note that we will suppress the adjective “upper” and just call such lattices semimodular as is commonly done in the literature. We claim that Bn is semimodular. To see why, let S and T be elements of Bn . Then the Principle of Inclusion-Exclusion implies that |S| + |T | = |S ∪ T | + |S ∩ T |. 6 Recall that if R ∈ Bn , then ρ(R) = |R|. Also recall that S ∪ T = S ∨ T and S ∩ T = S ∧ T . The previous equation now implies that for all S, T ∈ Bn we have ρ(S) + ρ(T ) ≥ ρ(S ∧ T ) + ρ(S ∨ T ). We conclude that Bn is semimodular. Let L be a lattice and let x ∈ L. We say x is left-modular if for all y, z ∈ L with y ≤ z we have y ∨ (x ∧ z) = (y ∨ x) ∧ z. If C is a multichain of L, then we say C is left-modular if all the elements of C are leftmodular. It is not hard to see that every element of Bn is a left-modular element and so any multichain of Bn is left-modular. If L is semimodular and contains a saturated ˆ0–ˆ1 left-modular multichain, then L is called supersolvable. Note that it is typical to define supersolvable using saturated left-modular chains as opposed to saturated left-modular multichains. Since saturated multichains always contain an underlying saturated chain, this will not lead to any problems. Recalling that Bn is semimodular and that every multichain is left-modular, we have that Bn is supersolvable. 1.2 The M¨ obius Function One of the most important functions on a finite poset is its M¨obius function, µ. It is a far-reaching generalization of the M¨obius function from number theory. While there is a two-variable version of the M¨obius function, we will only need the one-variable version. Definition 1.2.1. The (one-variable) M¨obius function of P is a map µ : P → Z defined 7 recursively so that µ(x) = δˆ0,y . (1.1) x≤y It is sometimes convenient to use the following equivalent formulation of µ. µ(y) =      1     − if y = ˆ0, (1.2) µ(x) otherwise. x 0. By relabeling, if necessary, we may assume that I = {1, 2, . . . , k}. It follows from part (a) that the number of atomic transversals in L(Txa ) with support size i is ei (N1 , N2 , . . . , Nk ) where ei is the ith elementary symmetric function. For each atomic transversal t ∈ L(Txa ) we have that µ(t) = (−1)| supp t| . Therefore, k µ(t) = t∈L(Txa ) k (−1)i e i (N1 , N2 , . . . , Nk ) i=0 = (1 − Ni ). i=1 24 Thus the summation condition (2.1) holds for each nonzero element in the quotient if and only if for each nonzero x ∈ L there is an index i such that |Ai ∩ Ax | = 1. Combining the previous result with Corollary 2.1.6 gives conditions under which the product of claws and its quotient have the same characteristic polynomial. We also need to show that there is an isomorphism between L and this quotient. This will give us the desired factorization. Theorem 2.2.3 ([6]). Let L be a ranked lattice and let (A1 , A2 , . . . , An ) be an ordered partition of the atoms of L. Let ∼ be the standard equivalence relation on n i=1 CLAi and let χ(L, t) be the characteristic polynomial. Suppose the following hold. (1) For all x ∈ L, Txa = ∅. (2) If t ∈ Txa , then | supp t| = ρ(x). (3) For each nonzero x ∈ L there is some i with |Ai ∩ Ax | = 1. Then we can conclude the following. (a) For all x ∈ L, µ(x) = (−1)ρ(x) |Txa |. n (t − |Ai |). (b) χ(L, t) = i=1 Proof. Let P = n i=1 CLAi . First, we show that L ∼ = P/ ∼. Define a map ϕ : (P/ ∼) → L by ϕ(Txa ) = x. It is easy to see that ϕ is well-defined. Define ψ : L → (P/ ∼) by ψ(x) = Txa . By assumption Txa = ∅ and so ψ is well-defined. Moreover, it is clear that ϕ and ψ are inverses of each other. To show that ϕ is order preserving, suppose that Txa ≤ Tya . Then just as in the proof of Lemma 2.2.2 part (b), we have that x ≤ y and so ϕ is order preserving. 25 To show that ψ is order preserving, suppose that x ≤ y. Then applying the same technique in the proof of Lemma 2.2.2 part (b) we get that there is a t ∈ Txa and s ∈ Tya with t ≤ s. By the definition of ≤ in P/ ∼ we get that Txa ≤ Tya and so ψ is order preserving. To obtain (a), note that the M¨obius value of an element in the product of claws is µ(t) = (−1)| supp t| . Therefore, using Lemma 2.1.4, we get µ(Txa ) = (−1)| supp t| . µ(t) = t∈Txa t∈Txa Using the isomorphism between L and the quotient as well as the fact that, by assumption (2), all the atomic transversals for x have size ρ(x), we have µ(x) = µ(Txa ) = (−1)ρ(x) |Txa | as desired. Finally, to verify (b) apply Corollary 2.1.6 and Lemma 2.2.2 to get that n (t − |Ai |) = χ(P, t) = χ(P/ ∼, t). i=1 Now part (b) follows immediately since L ∼ = P/ ∼. Let us return to the partition lattice and see how we can apply Theorem 2.2.3 Label the atoms (i, j) as before. Partition the atoms as (A1 , A2 , . . . , An−1 ) where Aj = {(i, j + 1) | i < j + 1} With each atomic transversal t we will associate a graph, Gt on n vertices such that there 26 is an edge between vertex i and vertex j if and only if (i, j) is in t. We will use the graph to verify the assumptions of Theorem 2.2.3 are satisfied for Πn . First, let us show assumption (1) holds. In the case when there is a single block B = {b1 < b2 < · · · < bm }, the elements (b1 , b2 ), (b2 , b3 ), . . . , (bm−1 , bm ) form an atomic transversal and their join is B. Now to get the elements which have more than one nontrivial block, follow the same procedure for each block and take the union of the atomic transversals. It follows every element has an atomic transversal. Next, we prove that assumption (2) holds. We claim that if t ∈ Tπa then Gt is a forest. To see why, suppose that there was a cycle and let c be the largest vertex in the cycle. Then c must be adjacent to two smaller vertices a and b which implies that both (a, c) and (b, c) must be in t. This is impossible since both come from Ac−1 . Since Gt is forest, if Gt has k components then the number of edges in Gt is n − k. It is not hard to see that i and j are in the same block in t if and only if i and j are in the same component of Gt . Moreover, it is well known that if π ∈ Πn and π has k blocks then ρ(π) = n − k. It follows that if t ∈ Tπa and π has k blocks then | supp t| = |E(Gt )| = n − k = ρ(π). We conclude that assumption (2) holds. Finally, to verify assumption (3), let π ∈ Πn with π = ˆ0. Then π contains a nontrivial block. Let i be the second smallest number in this block. We claim that there is only one atom in Ai−1 below π. First note that there is some atom below π in Ai−1 namely (a, i) where a is the smallest element of the block. Suppose there was more than one atom below π in Ai−1 and let (a, i), (b, i) ∈ Ai−1 with (a, i), (b, i) ≤ π. Then (a, i) ∨ (b, i) ≤ π and so a, b and i are all in the same block in π which is impossible since a, b < i but i was chosen to be the second smallest in its block. 27 Now applying the theorem we get that χ(Πn , t) = (t − 1)(t − 2) · · · (t − n + 1) since |Ai | = i for 1 ≤ i ≤ n − 1. More generally, Theorem 2.2.3 can be used to prove Terao’s result [16] about the characteristic polynomial of a hyperplane arrangement with a nice partition. In fact the notion of nice partition is the combination of assumptions (2) and (3) of Theorem 2.2.3 in the special case of a central hyperplane arrangement. 2.3 Rooted Trees One of the drawbacks of Theorem 2.2.3 is that assumption (1) requires that every element of the lattice has an atomic transversal. This forces L to be atomic. However, by generalizing the notion of a claw to that of a rooted tree, we will be able to remove this assumption and derive Theorem 2.3.2 below which applies to a wider class of lattices. Definition 2.3.1. Let P be a poset and S be a subset of P containing ˆ0. Let C be the collection of saturated chains of P which start at ˆ0 and use only elements of S. The rooted tree with respect to S is the poset obtained by ordering C by inclusion and will be denoted by RTS . It is easy to see that given any subset S of a poset containing ˆ0, the Hasse diagram of RTS always contains a ˆ0 and has no cycles. This explains the choice of rooted tree for the name of the poset. Strictly speaking the elements of RTS are chains of L. However, it will be useful to think 28 123 123 123 12/3 13/2 1/23 1/2/3 1/2/3 RTS1 RTS2 Figure 2.2: Hasse diagrams for rooted trees of the elements of RTS as elements of L where we associate a chain C with its top element. One can still recover the full chain by considering the unique path from ˆ0 to C in RTS . Let us consider an example in Π3 . As before, partition the atom set as A1 = {12/3} and A2 = {13/2, 1/23}. Let S1 , S2 be the upper order ideals generated by A1 , A2 , respectively, together with ˆ0. Then we we get RTS1 and RTS2 as in Figure 2.2. Note that we label the chains ˆ0 < 12/3 < 123, ˆ0 < 13/2 < 123 and ˆ0 < 1/23 < 123 in S1 and S2 all by 123 in RTS1 and RTS2 since each of these chains terminates at 123. In the previous sections, we used a partition of the atom set to form claws. In this section, we will use the partition of the atom set to form rooted trees. Given an ordered partition of the atoms of a lattice (A1 , A2 , . . . , An ), for each i we form the rooted tree RTUˆ (A ) where i Uˆ (Ai ) is the upper order ideal generated by Ai together with ˆ0. We will call rooted trees of the form RTUˆ (A ) complete rooted trees. We will study them in more detail in Section 2.6. i Note that since (A1 , A2 , . . . , An ) is a partition of the atoms, every element of the lattice appears in an RTUˆ (A ) for some i. i Given (A1 , A2 , . . . , An ), we call t ∈ n ˆ (A ) i=1 RTU i a transversal. We will use the notation, n Tx = t∈ RTUˆ (A ) : i i=1 29 t=x and call such elements transversals of x. If t consists of only atoms of L or ˆ0 then t is called an atomic transversal. This agrees with the terminology we used for claws. The set of atomic transversals for x will be denoted Txa as before. There is very little change in the approach using rooted trees as opposed to claws. As before, given a partition (A1 , A2 , . . . , An ) of the atom set of L, we will put the standard equivalence relation on n ˆ (A ) . i=1 RTU i Note that one can take the join using all the elements of a chain or just the top element as the results will be equal. Since we are using rooted trees, n ˆ (A ) i=1 RTU i the natural map from / ∼ to L is automatically surjective. In other words, we can remove the condition that every element of L has an atomic transversal. Additionally, since each Hasse diagram is a tree, when we take the product of the trees, the M¨obius value of any transversal which is not atomic is zero and so does not affect χ. Therefore, we get the following improvement on Theorem 2.2.3. Theorem 2.3.2 ([6]). Let L be a ranked lattice and let (A1 , A2 , . . . , An ) be an ordered partition of the atoms of L. Let ∼ be the standard equivalence relation on n ˆ (A ) i=1 RTU i and let χ be the characteristic polynomial. Suppose the following hold. (1) If t ∈ Txa , then | supp t| = ρ(x). (2) For each nonzero x ∈ L there is some i with |Ai ∩ Ax | = 1. Then we can conclude the following. (a) For all x ∈ L, µ(x) = (−1)ρ(x) |Txa |. n (b) χ(L, t) = tρ(L)−n (t − |Ai |). i=1 Proof. Let P = n ˆ (A ) . i=1 RTU i We need to show that P/ ∼ is homogeneous. The first condition of the definition is obvious. For the second, suppose that Tx ≤ Ty and t ∈ Tx . It 30 follows that x ≤ y. Let i be an index such that Ai ∩ Ay = ∅ so that y ∈ Uˆ (Ai ). If t ∈ Tx , then tj ≤ x ≤ y for all j. Therefore, t(y i ) ∈ Ty and t ≤ t(y i ). It follows that P/ ∼ is homogeneous. In the proof of Theorem 2.2.3, we showed that the lattice and the quotient of the product of claws were isomorphic. The proof that L and P/ ∼ are isomorphic is essentially the same. If we define ϕ and ψ analogously, then the only difference is showing ψ is order preserving in which case one can use the same ideas as in the previous paragraph to complete the demonstration. Now we verify that the summation condition (2.1) holds for all nonzero elements of P/ ∼. We only need to modify the proof that we gave in Lemma 2.2.2 part (c) slightly. Analogously to the proof of part (a) of that lemma, one sees that L(Tx ) = {t : ti ≤ x for all i}. Using this and the fact that only atomic transversals have nonzero M¨obius values, the proof of Lemma 2.2.2 part (c) goes through as before with Txa replaced by Tx . Now applying Lemma 2.1.4 and the fact that µ(t) = 0 if t is not atomic, we get µ(Tx ) = µ(t) = µ(t). (2.4) t∈Txa t∈Tx Then applying the same proof as in Theorem 2.2.3 gives us (a). To finish the proof we define a modification of the characteristic polynomial for any ranked poset P , µ(x)t−ρ(x) . χ(P, ¯ t) = x∈P We claim that χ(P, ¯ t) = χ(P/ ¯ ∼, t). Applying assumption (1) and the isomorphism L ∼ = 31 P/ ∼, we get that for every t ∈ Txa we have ρ(t) = | supp t| = ρ(x) = ρ(Tx ). This combined with equation (2.4), proves the claim. Now if RT is a rooted tree with k atoms then χ(RT ¯ ) = t−1 (t − k). It follows that n χ(P, ¯ t) = t−n (t − |Ai |). i=1 Since χ¯ is preserved by isomorphism, n χ(L, ¯ t) = χ(P/ ¯ ∼, t) = χ(P, ¯ t) = t−n (t − |Ai |). i=1 Multiplying by tρ(L) gives us part (b). 2.4 Partitions Induced by a Multichain It turns out that under certain circumstances we can show that assumption (2) of Theorem 2.3.2 and factorization of the characteristic polynomial are equivalent. To be able to prove this equivalence, we will not be able to take an arbitrary partition of the atoms, but rather we will need the partition to be induced by a multichain in the lattice. If L is a lattice and C : ˆ0 = x0 ≤ x1 ≤ · · · ≤ xn = ˆ1 is a ˆ0–ˆ1 multichain of L we get an ordered partition (A1 , A2 , . . . , An ) of the set A(L) of atoms of L by defining the set Ai as Ai = {a ∈ A(L) | a ≤ xi and a 32 xi−1 }. In this case we say (A1 , A2 , . . . , An ) is induced by the multichain C. Note that we do not insist that our multichain be a chain nor does it need to be saturated as is usually done in the literature. This also implies that some of the Ai ’s can be empty. Partitions induced by multichains have several nice properties. The first property will apply to any lattice (Lemma 2.4.2), but for the second we will need the lattice to be semimodular (Lemma 2.4.5). Before we get to these properties, we need a modification of Lemma 2.1.4. Lemma 2.4.1. Suppose that P/ ∼ is a homogeneous quotient and that for all non-maximal, nonzero X ∈ P/ ∼ we have that µ(y) = 0 y∈L(X) Then for all X ∈ P/ ∼ µ(X) =      µ(x)     x∈X               if X is not maximal, µ(x) − x∈X µ(y) if X is maximal. y∈L(X) Proof. If X is not maximal, then the proof of Lemma 2.1.4 goes through as before. Now suppose that X is maximal. If X = ˆ0 then the result holds because the quotient is homogeneous. So suppose X = ˆ0. In the proof of Lemma 2.1.4, we derived equation (2.3) without using the summation condition (2.1) and so it still holds. Moreover, it is easy to see that this equation is equivalent to the one for maximal X in the statement of the current result. Given a lattice and a partition of the atoms, it will be useful to know when elements 33 of a lattice do not satisfy condition (2) of Theorem 2.3.2. This is possible to do when the partition of the atoms is induced by a multichain. Lemma 2.4.2. Let L be a lattice and let (A1 , A2 , . . . , An ) be induced by a multichain C : ˆ0 = x0 ≤ x1 ≤ · · · ≤ xn = ˆ1. Let Ni be the number of atoms below an element x ∈ L in Ai . If Ni = 1 for all i and x = ˆ0 is minimal with respect to this property, then for all but one i, Ni = 0. Proof. Suppose that x is minimal, but that Ni > 1 for at least two i. Let k be the smallest index with Nk = 0, and B ⊆ Ak be the atoms below x in Ak so |B| ≥ 2. Let y = B. So, by the choice of B, y ≤ xk which implies that the atoms below y are in Ai for i ≤ k. So the choice of Ak forces the set of atoms below y to be B which is a proper subset of the set of atoms below x, and thus y < x. Since |B| ≥ 2, this contradicts the choice of x. The next definition gives one of the conditions equivalent to factorization when the atom partition is induced by a multichain. Definition 2.4.3. Let L be a lattice and let C : ˆ0 = x0 ≤ x1 ≤ · · · ≤ xn = ˆ1 be a ˆ0–ˆ1 multichain. For atomic x ∈ L, x neither ˆ0 nor an atom, let i be the index such that x ≤ xi but x ≤ xi−1 . We say that C satisfies the meet condition if, for each such x, we have x ∧ xi−1 = ˆ0. We are now in a position to give a list of equivalent conditions to factorization. Theorem 2.4.4. Let L be a ranked lattice and let (A1 , A2 , . . . , An ) be induced by a ˆ0–ˆ1 multichain, C. Suppose that if t ∈ Txa , then | supp t| = ρ(x). 34 Under these conditions the following are equivalent. 1. For every nonzero x ∈ L, there is an index i such that |Ai ∩ Ax | = 1. 2. For every element which is the join of two elements from the same Aj , there is an index i such that |Ai ∩ Ax | = 1. 3. The multichain C satisfies the meet condition. 4. We have that n χ(L, t) = tρ(L)−n (t − |Ai |) i=1 where χ is the characteristic polynomial. We note that it is possible to have n > ρ(L) in part 4 of the previous theorem. In this case the exponent on the outside of the factorization will be negative. This is possible since we are using multichains and so repeating elements in the chain will give rise to as many empty blocks in the partition of the atom set as we wish. However, for each such block, we get a corresponding factor (t − 0). Thus χ(L, t) is still a polynomial since the negative power of t on the outside of the product will be canceled by the positive powers of t on the inside of the product. Proof. (1) ⇒ (4) This is Theorem 2.3.2. (4) ⇒ (2) We actually show that (4) ⇒ (1) (the fact that (1) ⇒ (2) is trivial). We do so by proving the contrapositive. By assumption, there must be a nonzero x ∈ L such that for each i the number of atoms below x in Ai is different from one. Let k be the smallest value of ρ(x) for which elements of L have this property. We show that the coefficients of tρ(L)−k in χ(L, t) and in χ(P, t) = tρ(L)−n n i=1 (t − |Ai |) 35 are different, where P = n ˆ (A ) . i=1 RTU i Using the same proof as we did in Theorem 2.3.2, we can show that L ∼ = P/ ∼. So it suffices to show that the coefficient of tρ(L)−k in χ(P/ ∼, t) is different from the coefficient in χ(P, t). Let Q be the poset obtained by removing all the elements of P/ ∼ which have rank more than k. Let x1 , x2 , . . . , xl be the elements of L at rank k such that the number of atoms below xi in each block of the partition is different from one. Then by Lemma 2.4.2, each xi has atoms above exactly one block. Now let S = {Tx1 , Tx2 , . . . , Txl } be the set of the corresponding transversals. In Q, the elements of S are maximal and all the other non-maximal elements in Q satisfy the hypothesis of Lemma 2.4.1 which can be verified as in the proof of Theorem 2.3.2. Therefore we can calculate the M¨obius values of the elements of rank k in Q using Lemma 2.4.1. Once we know these values we can find the coefficient of tρ(L)−k in χ(P/ ∼, t). Each xi is above at least two atoms and is above only atoms in one block. Therefore the only atomic transversals which are in L(Txi ) are transversals with single atoms and the transversal with only zeros. Since only atomic transversals have nonzero M¨obius values we get that for all elements of S, def µ(t) = 1 − |Axi | < 0. ci = t∈L(Txi ) We know that ci < 0 since the number of atoms below each xi is at least two. Let Qk be the set of elements of Q at rank k. Using Lemma 2.4.1, we see that the sum of the M¨obius 36 values of Qk is l µ(Tx ) = Tx ∈Qk µ(Txi ) + i=1 l = i=1 µ(Tx ) Tx ∈Qk \S     µ(t) − ci  +   t∈Txi  Tx ∈Qk \S  µ(t) . t∈Tx As recently noted, only elements of L which have atomic transversals have nonzero M¨obius values. Using this and the assumption that | supp t| = ρ(x) = ρ(Tx ), we get that the coefficient of of tρ(L)−k in χ(P/ ∼, t) is l µ(t) − ci i=1 | supp t|=k where the first sum is over atomic t. As we saw before, each ci is negative and all are nonzero and so the coefficient of tρ(L)−k is different from µ(t) | supp t|=k which is the coefficient of tρ(L)−k in χ(P, t). This completes the proof that (4) ⇒ (2). (2) ⇒ (3) We show the contrapositive holds. Suppose that C does not satisfy the meet condition. Then there is some atomic x which is neither an atom nor ˆ0 such that x ≤ xi , x ≤ xi−1 , and x ∧ xi−1 = ˆ0. It follows that x is only above atoms in Ai . Since x is atomic, but not an atom, there are at least two atoms, a, b below x in Ai . Let y = a ∨ b. Since y ≤ x, y can only be above atoms in Ai . Therefore, for all indices j, |Aj ∩ Ay | = 1 and y is the join of two atoms. 37 ˆ1 c d a b ˆ0 Figure 2.3: A lattice (3) ⇒ (1) First let us note that if x is an atom then the result is obvious. For x ∈ L let i be the index such that x ≤ xi and x ≤ xi−1 . We now induct on i. If i = 1 then it suffices to show that |A1 | = 1 since then every nonzero x ≤ x1 is only above the unique element of A1 . However if a, b are distinct atoms in A1 then x = a ∨ b is atomic but not an atom or zero. Further x ≤ x1 but x ∧ xi−1 = x ∧ ˆ0 = ˆ0 which contradicts the meet condition. This finishes the i = 1 case. Now suppose that i > 1 and x is not an atom. Let z = Ax . Then z is atomic and Az = Ax . Let y = z ∧ xi−1 . Since C satisfies the meet condition, y = ˆ0. By construction y < xi−1 and so by induction, there is some index j ≤ i − 1 with Aj ∩ Ay = {a}. Suppose that there was some other atom b ∈ Aj ∩ Az . Then y ∨ b is less than or equal to both z and xi−1 and so y ∨ b ≤ z ∧ xi−1 = y. However, this is impossible since then Aj ∩ Ay ⊇ {a, b}. It follows that 1 = |Aj ∩ Az | = |Aj ∩ Ax | and so (1) holds. It would be nice if all atomic transversals had the correct support size when using a partition induced by a multichain since then we could remove this assumption from the previous theorem. Unfortunately this does not always occur. To see why, consider the lattice in Figure 2.3. The left-most saturated ˆ0–ˆ1 chain induces the ordered partition ({a}, {b}). 38 It is easy to see that the support size of the transversal with both elements is not the rank of their join. Note, however, that if we had the relation a < d, then the support size would be the rank of the join. Moreover, note that this would also make the lattice semimodular. We see in the next lemma that semimodularity always implies transversals induced by a multichain have the correct support size. Lemma 2.4.5. Let L be a semimodular lattice and let (A1 , A2 , . . . , An ) be induced by the multichain C : ˆ0 = x0 ≤ x1 ≤ x2 ≤ · · · ≤ xn = ˆ1. If ∼ is the standard equivalence, then for all x ∈ L we have that t ∈ Txa implies | supp t| = ρ(x). Proof. Given an atomic t ∈ Txa we induct on | supp t|. If | supp t| = 0 the result is obvious. Now suppose that | supp t| = k > 0. Let i be the largest index in supp t. Let s = t(ˆ0i ), then | supp s| = k − 1. Suppose that s ∈ Tya , then ρ(y) = k − 1 by induction. Let j be the largest index such that j ∈ supp s. Then y = i > j. Thus x = s ≤ xj by definition of j and ti ≤ xj since t = ( s) ∨ ti > y. Therefore ρ(x) > ρ(y) = k − 1 and so ρ(x) ≥ k. Since | supp t| = k, ρ(x) ≤ k as L is semimodular. We conclude that ρ(x) = k = | supp t| and so our result holds by induction. Let us now consider supersolvable semimodular lattices. Recall that every supersolvable semimodular lattice contains a saturated ˆ0–ˆ1 left-modular chain. It turns out that saturated ˆ0–ˆ1 left-modular chains satisfy the meet condition as we see in the next lemma. Lemma 2.4.6. Let L be a lattice. If C : ˆ0 = x0 ≤ x1 ≤ x2 < · · · ≤ xn = ˆ1 is a left-modular saturated ˆ0–ˆ1 multichain then C satisfies the meet condition. 39 Proof. We claim that it is without loss of generality to assume that C is a chain as opposed to a multichain. To see why, suppose D is a multichain and that xj = xj+1 in D. Moreover, suppose that there was some x which was atomic and not an atom with x ≤ xj+2 , but x ≤ xj+1 . Then x ∧ xj = x ∧ xj+1 and so x ∧ xj = ˆ0 if and only if x ∧ xj+1 = ˆ0. It follows that we can remove the repeated element without changing whether or not D satisfies the meet condition. Let x ∈ L be atomic and neither an atom nor ˆ0. Let i be such that x ≤ xi and x ≤ xi−1 . Then we claim that there is some atom a with a < x and a ≤ xi−1 . To verify the claim, suppose that no such a existed. Since x is not an atom, it must be that all the atoms below x are also below xi−1 . However, x being atomic implies that x = Ax and so x ≤ xi−1 which is impossible. By the claim, xi−1 < a ∨ xi−1 ≤ xi . Since xi−1 xi we have that a ∨ xi−1 = xi . Now (xi−1 , x) is a modular pair and a < x so, by the definition of a modular pair, a ∨ (xi−1 ∧ x) = (a ∨ xi−1 ) ∧ x = xi ∧ x = x. But a < x so xi−1 ∧ x = ˆ0 and thus C satisfies the meet condition. We now get Stanley’s Supersolvability Theorem as a corollary of Theorem 2.4.4, Lemma 2.4.5, and Lemma 2.4.6. Theorem 2.4.7 (Stanley’s Supersolvability Theorem [15]). Let L be a semimodular lattice with partition of the atoms (A1 , A2 , . . . , An ) induced by a saturated ˆ0–ˆ1 left-modular multichain. Then n χ(L, t) = tρ(L)−n (t − |Ai |). i=1 40 (x1 (x2 (x3 x4 ))) (x1 ((x2 x3 )(x4 )) ((x1 x2 )(x3 x4 )) ((x1 (x2 x3 ))(x4 ) (((x1 x2 )x3 )x4 ) Figure 2.4: The Tamari lattice T3 2.5 Transversal Functions In the previous sections we had to make the assumption that our poset was a lattice and that it was ranked. In this section we develop the tools to deal with the general case when P is not necessarily a lattice nor ranked. To start, let us do an example and calculate the generalized characteristic polynomial of an unranked poset. We will consider the Tamari lattices [3, 8], which will be denoted by Tn . One way to define Tn is as the set of parenthesizations of the word x1 x2 · · · xn+1 with ordering given by saying π is covered by σ if there exists subwords A, B, and C such that π = . . . ((AB)C) . . . and σ = . . . (A(BC)) . . . Figure 2.4 displays the Hasse diagrams for T3 . As one can see from the Hasse diagram, T3 is not ranked. In order to calculate the generalized characteristic polynomial for T3 we need a function, ρ. We will use generalized rank which was introduced in [1]. To define generalized rank, let us set up some notation. As done previously, let Ax be the set of atoms below x in P . If (A1 , A2 , . . . , An ) is an ordered 41 partition of the atoms of P , the generalized rank of an element x is given by ρ(x) = |{i : Ai ∩ Ax = ∅}|. (2.5) In other words, ρ(x) counts the number of blocks in the partition that x is above. Returning to the T3 example, let us partition the atoms into A1 = {((x1 x2 )(x3 x4 ))}, A2 = {((x1 (x2 x3 ))(x4 )}). Given this partition, we see that the generalized rank of the bottom element is 0, the three middle elements all have generalized rank 1 and the top element has generalized rank 2. We take m = 3 which is the the length of the longest chain in T3 . Using the definition of the generalized characteristic polynomial (equation (1.3)), we get χ(T3 , t) = t3 − 2t2 + t = t(t − 1)2 . We see that χ(T3 , t) factors with roots 0 and 1. One might ask if we can decompose T3 into the product of two smaller posets. Using this reasoning, we might guess that T3 is the product of two chains since chains have characteristic polynomials with roots 0 and 1. Of course, this cannot be the case since chains are ranked and so their products are too, but T3 is not ranked. However, it is possible to take the product of the chains, collapse elements in the Hasse diagram without changing the characteristic polynomial and also get a poset isomorphic to T3 . However, unlike the method we explained when our posets were ranked, we will need to collapse elements of different ranks. In order to accomplish, we introduce the notion of a transversal function. 42 As we did in Section 2.3 we will be using rooted trees. When the poset we are interested in is a lattice, the standard equivalence relation (Definition 2.2.1) is the canonical choice for taking quotients. Since we are interested in posets which are not necessarily lattices we need to generalize this idea. To do this, we quotient out by the kernel of a special type of map from the product of rooted trees to the poset. Definition 2.5.1. Let P be a poset and let (S1 , S2 , . . . , Sn ) be an ordered collection of subsets of P each containing ˆ0. We say f : n i=1 RTSi → P is a transversal function if it has the following properties: 1. The function f is order preserving. 2. The function f is surjective. 3. If f (t) = ˆ0, then ti = ˆ0 for all i. If f is a transversal function, the kernel of f , denoted ker f , is the equivalence relation ∼ given by s ∼ t if and only if f (s) = f (t). Since we will often be referring to equivalence classes and the elements of these classes we need names for these objects. Some of the names we use were defined in earlier sections, but we use the same names since they are generalizations. Definition 2.5.2. Let P be a poset and let (S1 , S2 , . . . , Sn ) be an ordered collection of subsets of P . Let f : n i=1 RTSi → P be a transversal function. If t ∈ n i=1 RTSi then we say t is a transversal for x if f (t) = x. We say t is atomic or an atomic transversal if all the elements of t are atoms of RTSi or ˆ0. The set of all transversals for x will be denoted by Tx and the set of all atomic transversals will be denoted by Txa . We also define the support of a transversal, t, as supp t = {i : ti = ˆ0}. 43 From the definitions it is evident that the set of equivalence classes of n i=1 RTSi / ker f is {Tx : x ∈ P }. Moreover, it is clear that the size of the support of an atomic transversal for x is also its rank in the product of the rooted trees. We are now in a position to give another factorization theorem. The factorization result we provide in Section 2.6 will be a special case of this one. Recall that we are using the notation A(P ) to denote the atom set of a poset P . In particular, A(RTS ) denotes the atoms in the rooted tree RTS . Theorem 2.5.3. Let P be a poset with ρ : P → N and let m ∈ N such that ρ(P ) ≤ m. Moreover, let (S1 , S2 , . . . , Sn ) be an ordered collection of subsets of P which contain ˆ0 and let f be a transversal function. Suppose the following hold. (1) If x ≤ y and s ∈ Tx , there exists t ∈ Ty with s ≤ t. (2) If t ∈ Txa , then | supp t| = ρ(x). (3) The summation condition (2.1) holds for all Tx . We can conclude the following. (a) We have an isomorphism n P ∼ = RTSi / ker f. i=1 (b) For each x ∈ P , µ(x) = (−1)ρ(x) |Txa |. (c) The generalized characteristic polynomial of P with respect to ρ and m (equation (1.3)) is given by n χ(P, t) = tm−n (t − |A(RTSi )|). i=1 44 Proof. First, we need to show the quotient is a homogeneous quotient. Conditions (2) and (3) in the definition of a transversal function (Definition 2.5.1) imply condition (1) of a homogeneous quotient (Definition 2.1.2). To show condition (2) holds, suppose that Tx ≤ Ty . Then there is a q ∈ Tx and a r ∈ Ty with q ≤ r. Since f is order preserving, x = f (q) ≤ f (r) = y. By assumption (1) of the theorem, given a s ∈ Tx there is a t ∈ Ty with s ≤ t and so condition (2) of a homogeneous quotient is satisfied. Now we show (a). Let f¯ : n i=1 RTSi / ker f → P be the induced quotient map sending Tx to x. Since f is surjective, it follows easily that f¯ is a bijection and so has an inverse say g. Next we show that f¯ is order preserving. Recall that the elements of the quotient, ( n i=1 RTSi )/ ker f , are of the form Tx for some x ∈ P . Suppose that Tx ≤ Ty . Then again, since f is order preserving, x ≤ y and so f¯ is order preserving. To finish the proof of (a), we show g is order preserving. Suppose that x ≤ y. Since f is surjective, Tx = ∅. Therefore, by assumption (1), there are s ∈ Tx and t ∈ Ty with s ≤ t. Using the definition of a quotient poset, we get that that Tx ≤ Ty and so g(x) ≤ g(y). Now we verify (b). By Lemma 2.1.4, assumption (3), and the fact that isomorphisms preserve M¨obius values, we have that µ(x) = µ(t). t∈Tx Since only atomic transversals have nonzero M¨obius value we have µ(x) = µ(t). t∈Txa 45 By assumption (2), all the atomic transversals have the same support size which is the rank of x. It follows that each atomic transversal for x has M¨obius value (−1)ρ(x) . Therefore we have that µ(x) = (−1)ρ(x) |Txa |. Finally we show (c). By definition, µ(x)tm−ρ(x) . χ(P, t) = x∈P Using part (b), we get (−1)ρ(x) |Txa |tm−ρ(x) . χ(P, t) = x∈P We can break this sum into parts, depending on the rank of x. Note that by assumption (2) and part (b), every element with rank larger than n has M¨obius value zero. Thus we have,  n χ(P, t) =  (−1)k |Txa |tm−k  .  k=0 ρ(x)=k Neither (−1)k nor tm−k depend on x so we can pull them out to get,  n χ(P, t) =  (−1)k tm−k  k=0 |Txa | . ρ(x)=k Using assumption (2) and denoting the k th elementary symmetric function as ek , we have the inner sum is exactly ek (|A(RTS1 )|, |A(RTS2 )|, . . . , |A(RTSn )|). It follows that, n (−1)k ek (|A(RTS1 )|, |A(RTS2 )|, . . . , |A(RTSn )|)tm−k . χ(P, t) = k=0 46 Pulling out a factor of tm−n permits us to rewrite the sum as a product n χ(P, t) = tm−n (t − |A(RTSi )|) i=1 completing the proof. 2.6 Complete Transversal Functions By definition, a transversal function must be surjective. However, if we impose more structure on the choice of subsets used to build the rooted trees, we can remove this assumption. In order to show this, we begin with a definition. Definition 2.6.1. Let P be a poset and let A be a set of atoms. The complete tree (with respect to A) is the rooted tree RTUˆ (A) where Uˆ (A) is the upper order ideal generated by the set A together with ˆ0. Along with this new definition, we have a new type of function. Definition 2.6.2. Let P be a poset and let (A1 , A2 , . . . , An ) be an ordered partition of A(P ). We say f : n ˆ (A ) i=1 RTU i → P is a complete transversal function if it is order preserving and has the property that if in t we have ti = ˆ0 or ti = x for all i, then f (t) = x. Note that it may appear that complete transversal functions are not transversal functions because we dropped the condition that they are surjective. However, we will see in the next lemma that, among other nice properties, the surjectivity of the function is a consequence of the definition. We also note that if we have a lattice, then f (t) = ∨t is a complete transversal function where ∨t = t1 ∨ . . . · · · ∨ tn . 47 Lemma 2.6.3. Let P be a poset and let (A1 , A2 , . . . , An ) be an ordered partition of A(P ). Let f be a complete transversal function. Then we can conclude the following. (a) The function f is surjective and f is a transversal function. (b) For all j, tj ≤ f (t1 , t2 , . . . , tn ). (c) If x ≤ y and s ∈ Tx , there exists t ∈ Ty with s ≤ t. (d) For x ∈ P , let Ni be the number of atoms below x in Ai , then n (1 − Ni ). µ(s) = (2.6) i=1 s∈L(Tx ) (e) The summation condition (2.1) holds for all Tx if and only if for all nonzero x ∈ P , there is an index i such that |Ai ∩ Ax | = 1. Proof. First we show (a). Let ˆ 0 be the transversal having all components equal to ˆ0. Since we are using complete trees and a partition of the atom set, for every x ∈ P there exists some i such that ˆ 0(xi ) is a transversal. It follows from the definition of a complete transversal function that f (ˆ 0(xi )) = x and so f is surjective. To show that the third condition for a transversal function holds, suppose that f (t) = ˆ0. By definition of a complete transversal function, f (ˆ 0(tii )) = ti . Since f is order preserving and ˆ 0(tii ) ≤ t we get that ti = f (ˆ 0(tii )) ≤ f (t) = ˆ0. Therefore, if f (t) = ˆ0, then t = ˆ 0. This completes the proof that f is a transversal function. For (b), we noted in the previous paragraph that j f (ˆ 0(tj )) = tj . 48 Using the fact that f is order preserving, we get that j tj = f (ˆ 0(tj )) ≤ f (t1 , t2 , . . . , tn ). Next we prove (c). This is trivial if x = ˆ0 so assume x is nonzero. Let s ∈ Tx . Then by by part (b), si ≤ x for all i. Let t be given by ti = y for all i with i ∈ supp s and ti = ˆ0 for all other i. Such a t is a valid transversal since si ≤ x ≤ y and we are using complete trees. Note also that since x = ˆ0 it must be that t has at least one nonzero coordinate. It follows that t ∈ Ty and s ≤ t. Next, let us show (d). We start by showing that L(Tx ) = {t a transversal : ti ≤ x for all i}. (2.7) To see that L(Tx ) is contained in the other set, let t ∈ L(Tx ). Then for each i we have ˆ 0(tii ) ∈ L(Tx ). By definition of a complete transversal function, f (ˆ 0(tii )) = ti . Since f is order preserving, ti = f (ˆ 0(tii )) ≤ f (t) = x. For the reverse inclusion, suppose that t is a transversal with ti ≤ x for all i. Let s be the transversal obtained from t by replacing all the nonzero ti with x. We know that s is a valid transversal because we are using complete trees. Since f is a complete transversal function, f (s) = x and so s ∈ L(Tx ). By construction, t ≤ s and therefore t ∈ L(Tx ). Let I be the set of indices, i, such that there is an atom below x in Ai . By relabeling, if necessary, we may assume that I = {1, 2, . . . , j}. Since Ni = 0 implies that 1 − Ni = 1, j n (1 − Ni ) = i=1 (1 − Ni ). i=1 49 From equation (2.7) we can conclude that the number of atomic transversals in L(Tx ) with support size i is ei (N1 , N2 , . . . , Nj ) where ei is the ith elementary symmetric function. Now for each atomic transversal s ∈ L(Tx ) we have that µ(s) = (−1)| supp(s)| and all other transversals have M¨obius value zero. Therefore, j j (−1)i ei (N1 , N2 , . . . , Nj ) µ(s) = s∈L(Tx ) = i=0 (1 − Ni ) i=1 which completes part (d). Finally, (e) follows immediately from (d) and the definition of the summation condition. Given this lemma, we can use Theorem 2.5.3 to immediately obtain the following. Theorem 2.6.4. Let P be a poset with ρ : P → N and let m ∈ N such that ρ(P ) ≤ m. Moreover, let (A1 , A2 , . . . , An ) be an ordered partition of A(P ) and let f be a complete transversal function. Suppose the following hold. (1) If t ∈ Txa , then | supp t| = ρ(x). (2) For all nonzero x ∈ P , there is an index i such that |Ai ∩ Ax | = 1. We can conclude the following. (a) We have an isomorphism n P ∼ = i=1 RTUˆ (A ) i / ker f. (b) For each x ∈ P , µ(x) = (−1)ρ(x) |Txa |. 50 1230 120 /30 1231 130 /20 10 /230 1232 121 /30 131 /30 10 /231 10 /20 /30 Figure 2.5: The weighted partition poset Πw 3 (c) The generalized characteristic polynomial of P with respect to ρ and m is given by n χ(P, t) = tm−n (t − |Ai |). i=1 The reader may be wondering why we did not just assume from the start that we were using complete transversal functions. By doing so, we reduce the number of things we need to check and we still get the same conclusions as in Theorem 2.5.3. However, there are situations where the first theorem applies but the second does not. Let us give an example were the summation condition (2.1) for Tx needed in Theorem 2.5.3 holds, but the second condition of Theorem 2.6.4 does not. We will consider the weighted w partition poset, Πw n introduced in [2]. The elements of Πn are set partitions of [n] where each block Bi has one of the following weights {0, 1, . . . , |Bi | − 1}. The weighted partitions w w will be denoted by B1 1 /B2 2 / . . . /Bnwn where wi is the weight of block Bi . The ordering is given by v v v w w A11 /A22 / . . . /Akk ≤ B1 1 /B2 2 / . . . /Bnwn if and only if 51 1. We have A1 /A2 / . . . /Ak ≤ B1 /B2 / . . . /Bn in the (unweighted) partition lattice Πn . 2. If Bl = Ai1 ∪ Ai2 ∪ · · · ∪ Aim , then vl − (wi1 + wi2 + · · · + wim ) ∈ {1, 2, . . . , m − 1}. The weighted partition poset Πw 3 is shown in Figure 2.5. It is easy to check that the characteristic polynomial of this poset factors as 2 χ(Πw 3 , t) = (t − 3) . Consider the sets A1 = {120 /30 , 130 /20 , 121 /30 , ˆ0} and A2 = {10 /230 , 131 /30 , 10 /231 , ˆ0}. Additionally, consider the transversal function f : 2 i=1 RTAi → Πw 3 which sends any pair which contains ˆ0 to the other element in the pair and sends any pair with two non-zero elements to 123i where i is the sum of their exponents. It is easy to check that f is a transversal function and that the summation condition (2.1) is satisfied. However, the element 1231 is above every atom so it is impossible that it is above only one atom of either A1 or A2 . One can also check that all the conditions of Theorem 2.5.3 are satisfied and so we have verified that the characteristic polynomial does factor using our method. 52 We should also point out that, as was shown in [4], the characteristic polynomial of the w n−1 . This was shown using different weighted partition poset Πw n factors as χ(Πn , t) = (t−n) methods than presented here. As of now, we do not have a transversal function which gives us the factorization. 2.7 Crosscut-simplicial Lattices As defined in [9], a crosscut-simplicial lattice is a lattice, L, such that if [x, y] is an interval of L then any proper subset of the atoms of [x, y] has a join different from y. Examples of crosscut-simplicial lattices include the Tamari lattices, the weak Bruhat order on Coxeter groups and more generally the Cambrian lattices [10]. In this section we show that a lattice is crosscut-simplicial if and only if the generalized characteristic polynomial, with respect to a function we define below, of every interval only has roots 0 and 1. In addition to using Theorem 2.6.4 to prove this result, we will also make use of a special case of Rota’s Crosscut Theorem [11]. We will give a proof of the full theorem using quotient posets in Chapter 3. Theorem 2.7.1 (Rota’s Crosscut Theorem [11]). Let L be a lattice. For x ∈ L, let ai (x) be the number of subsets of A(L) of size i whose join is x. Then (−1)i ai (x). µ(x) = i≥0 Theorem 2.7.2. Let L be lattice and let I be any interval in L. Let ρI (x) be the number of atoms below x in I and let m(I) be the length of the longest chain in I. Suppose that χ(I, t) is the generalized characteristic polynomial with respect to ρI and m(I). The lattice 53 L is crosscut-simplicial if and only if χ(I, t) = tm(I)−|A(I)| (t − 1)|A(I)| for every interval I of L. Note that if we partition the atoms of an interval I into singleton blocks, then ρI is just the generalized rank of I. However, since “generalized” is ambiguous without knowing the partition of the atom set, we have decided to define it differently. Proof. (⇒) Given an interval I, partition the atoms of I as (A1 , A2 , . . . , Ak ) where each Ak has exactly one atom. Let f be the complete transversal function defined as f (t) = ∨t. With this partition we trivially get condition (2) of Theorem 2.6.4. By the definition of ρI , we must show that the join of any j atoms is above exactly j atoms in order to show condition (1). Suppose there was some subset of j atoms of I whose join is above at least j + 1 atoms. Let x be the meet of these j + 1 atoms and y be the join of the j + 1 atoms. Then L cannot be crosscut-simplicial since there is a proper subset of atoms in [x, y] whose join is y. Thus we have a contradiction and so condition (1) of Theorem 2.6.4 is satisfied. Finally, we must show that m(I) ≥ ρI (I) where m(I) is the length of the longest chain in I since this was required in the definition of the generalized characteristic polynomial. Let x0 be the ˆ0 element of I and for each 1 ≤ i ≤ k define i xi = al l=1 where ai is the unique element of Ai . Since the join of j atoms is above exactly those j 54 atoms, we know that all the xi ’s are distinct. It follows that I contains a chain of length k, namely the chain x0 < x1 < · · · < xk . Using the definition of ρI , we see that if m(I) is the length of the longest chain in I then ρI (I) = k ≤ m(I). Applying Theorem 2.6.4 now yields this direction. (⇐) We prove the contrapositive. Suppose that L is not crosscut-simplicial. Then there must be some interval I of L where there is a set of atoms whose join is above more than just those atoms. Let j be the minimum number of atoms needed to form a set whose join is above more than just those j atoms. We claim that the coefficient of tm(I)−j in χ(I, t) cannot be the same as the coefficient of tm(I)−j in tm(I)−|A(I)| (t − 1)|A(I)| . Suppose that {a1 , a2 , . . . , al } ⊆ A(I) with ρI (a1 ∨a2 ∨· · ·∨al ) = j. By the definition of ρI , we have that l ≤ j since ai ∨a2 ∨· · ·∨al is above at least a1 , a2 , . . . , al . If l < j, then there would be a subset of A(I) with l < j elements whose join was above more than l elements. This contradicts the definition of j. It follows that if ρI (a1 ∨ a2 ∨ · · · ∨ al ) = j, then l = j. In other words, any subset of A(I) whose join has rank j has exactly j atoms in it. Applying Theorem 2.7.1, we get that the coefficient of tm(I)−j in χ(I, t) is (−1)j multiplied by the number of j-element subsets of A(I) which are above exactly those j atoms. Since there is at least one j-element subset whose join is above more than j atoms, it must be that the coefficient of tm(I)−j in χ(I, t) is not (−1)j |A(I)| . However, (−1)j |A(I)| is the j j coefficient of tm(I)−j in tm(I)−|A(I)| (t − 1)|A(I)| and so χ(I, t) = tm(I)−|A(I)| (t − 1)|A(I)| . While the Cambrian lattices provide a large family of crosscut-simplicial lattices, we now discuss another way to find such lattices. To do so, we will consider the notion of an edge labeling of a poset. Let P be a poset, let H be its Hasse diagram and let E(H) be the set 55 (x1 (x2 (x3 x4 ))) 3 2 (x1 ((x2 x3 )(x4 )) ((x1 x2 )(x3 x4 )) 3 3 ((x1 (x2 x3 ))(x4 ) 2 (((x1 x2 )x3 )x4 ) Figure 2.6: The Tamari lattice T3 with an SB-labeling of edges of H. An edge labeling of P is a map λ : E(H) → S, where S is some set. In other words, we label the edges of the Hasse diagram of the poset. Introduced in [7, Definition 3.4], an SB-labeling of a lattice L is an edge labeling of L such that for all u, v, w ∈ L with u v, w then we have the following. 1. If λ(u, v) and λ(u, w) are the labels of the edges from u to v and u and w respectively, then λ(u, v) = λ(u, w). 2. Every inclusion-maximal chain in the interval [u, v ∨ w] uses only the labels λ(u, v) and λ(u, w) with each label occurring at least once. An example of an SB-labeling of the Tamari Lattice T3 is given in Figure 2.6 where the labels are in boldface. This labeling is the labeling explained in [7, Theorem 5.5]. The authors in [7, Theorem 3.7] showed that if L has an SB-labeling, then it is crosscut-simplicial. As pointed out in [9], the converse of this theorem is not true. One way to see this is to let P be the poset formed by starting with B2 and and replacing one of its atoms with copy of B3 . Figure 2.7 contains the Hasse diagram of P . It is not hard to check that P is crosscut-simplicial. Since P only has two atoms whose join is ˆ1, only 2 labels are allowed to be used to label the entire poset. However, the B3 that we inserted needs at least 3 different 56 . . . . . . . . . . . Figure 2.7: A crosscut-simplicial lattice which does not have an SB-labeling labels and so there is no SB-labeling of P . The author of [9] suggested a generalization of an SB-labeling which avoids this type of problem. He then goes on to ask the question if a lattice being crosscut-simplicial is equivalent to having one of these generalizations of SB-labeling. It would interesting if one could use the equivalent formulation of crosscutsimplicial we described earlier using generalized characteristic polynomials to investigate this question. 2.8 LL lattices Having shown what Theorem 2.6.4 can say about crosscut-simplicial lattices, we now turn our attention to seeing how it implies a theorem of [1]. Earlier we showed how to use the fact that partitions induced by saturated ˆ0–ˆ1 left-modular multichains imply assumption (2) of Theorem 2.6.4 to prove Stanley’s Supersolvability Theorem [15]. We will use this fact to prove Blass and Sagan’s result about LL lattices [1] which is a generalization of the supersolvability result. In order to explain this result, we need to define the level condition. Let (A1 , A2 , . . . , An ) 57 be induced by C : ˆ0 = x0 ≤ x1 ≤ x2 ≤ · · · ≤ xn = ˆ1 . This multichain also induces a partial ordering on the atoms denoted by . It is defined by saying a b if a ∈ Ai and b ∈ Aj with i < j. We say that a lattice L with multichain C satisfies the level condition if a b1 ··· b2 bk implies that k a≤ bi . i=1 The lattice L is called an LL lattice if it contains a left-modular multichain C and L together with C satisfy the level condition. We are now in a position to state Blass and Sagan’s result. Theorem 2.8.1 ([1]). Let L be a lattice and let (A1 , A2 , . . . , An ) be induced by a left-modular saturated multichain such that L is an LL lattice. Let ρ be generalized rank and let m be the length of the longest ˆ0–ˆ1 chain. Then n χ(L, t) = tm−n (t − |Ai |). i=1 Proof. We wish to use Theorem 2.6.4. First, note that since we are using generalized rank we have that ρ(P ) is at most the number of nonempty blocks in the partition. Since our partition is induced by a multichain and since m is the length of the largest chain in the lattice, we have that ρ(P ) ≤ m. Define the complete transversal function to be f (t) = ∨t. Although it is not worded in the same way, the authors in [1, Theorem 6.3 and Lemma 6.4] proved assumption (1) of Theorem 2.6.4 holds. Finally, as noted before, it was shown in Lemma 2.4.6 that satu58 rated left-modular multichains satisfy the meet condition and so satisfy assumption (2) of Theorem 2.6.4. The theorems presented so far have provided conditions which imply factorization. We would like to finish this section with a theorem where we provide a condition which is equivalent to factorization. Theorem 2.8.2. Let P be a poset and let ρ : P → N with m ∈ N such that ρ(P ) ≤ m. Let χ(P, t) be the generalized characteristic polynomial with respect to ρ and m. Let (A1 , A2 , . . . , An ) be an ordered partition of A(P ) and let f : n ˆ (A ) i=1 RTU i complete transversal function. Finally, define T = {x ∈ P \ ˆ0 : |Ai ∩ Ax | = 1 for all i}. Suppose that the following hold. 1. If t ∈ Txa then | supp(t)| = ρ(x). 2. If x, y ∈ P and x < y, then ρ(x) < ρ(y). 3. For all minimal elements x, y ∈ T , the cardinality of the sets {i : |Ai ∩ Ax | = 0} and {i : |Ai ∩ Ay | = 0} have the same parity. Under these conditions, n χ(P, t) = tm−n (t − |Ai |) i=1 59 → P be a if and only if for every nonzero x ∈ P there is an index i such that |Ai ∩ Ax | = 1. This theorem is a generalization of Theorem 2.4.4. The two proofs are quite similar so we only provide a sketch below. Sketch of proof. First, note that the backwards direction is Theorem 2.6.4. For the forward direction, we will prove the contrapositive. Note that the assumption in this direction implies that T = ∅. Let k be the smallest value of ρ applied to the elements of T . We show that the coefficient of tm−k in χ(P, t) and in tm−n Define R = n ˆ (A ) i=1 RTU i n i=1 (t − |Ai |) are different. / ker f . We claim that R is a homogeneous quotient and that P ∼ = R. Since f is a complete transversal function, Lemma 2.6.3 part (c) implies that assumption (1) of Theorem 2.5.3 is satisfied. Note that the proof of part (a) of Theorem 2.5.3 only requires assumption (1). Therefore, R is homogeneous and P ∼ = R. Since P ∼ = R, it is enough to show that the coefficient of tm−k in χ(R, t) and in tm−n n i=1 (t − |Ai |) are not the same. Let x1 , x2 , . . . , xl be the set of elements of T with ρ(xi ) = k for all i and let S = {Tx1 , Tx2 , . . . , Txl } be the corresponding equivalence classes. Moreover, define Q to be the poset obtained from R by removing all elements of R with ρ value larger than k. Using assumption (2), we can see that the M¨obius value of elements with ρ at most k in R and Q are the same. In Q all elements with ρ value k are maximal. By assumption (2) and the assumption on k any element of Q which is not maximal cannot be in the set T . Then Lemma 2.6.3 part (e) implies that every non-maximal element satisfies the summation condition (2.1). Thus, we can apply Lemma 2.4.1 to conclude that µ(t) − µ(Txi ) = t∈Txi µ(s). s∈L(Txi ) 60 Let ci = µ(s). s∈L(Txi ) We claim that all the ci ’s are either 0 or have the same sign. By equation (2.6), if ci = 0, then the sign of ci is (−1)ki where ki is the number of blocks with atoms below xi . By assumption (3), for the ci ’s which are not equal to 0, the corresponding ki ’s have the same parity. Therefore, the signs of the nonzero ci ’s are the same. Since there is at least one element of T = ∅ in Q, there is at least one ci = 0. Using the same argument as in the proof of Theorem 2.4.4, we get the coefficient of tm−k in χ(R, t) is l µ(t) − ci i=1 | supp t|=k in which the first sum ranges over atomic transversals. Since there is at least one ci which is nonzero and all the nonzero ci ’s have the same sign, we see that this coefficient is not the same as µ(t). | supp t|=k However, the previous expression is the coefficient of tm−k in tm−n that n χ(P, t) = tm−n (t − |Ai |) i=1 which is what we wished to show. 61 n i=1 (t − |Ai |). It follows Chapter 3 Classic Results About the M¨ obius Function In this chapter, we will give a new method to prove an array of classic results about the M¨obius function. To explain the idea, we need a definition. Let P be a poset with a ˆ1. A coatom of L is an element c ∈ P such that c ˆ1. The idea of this new method is to use induction on the size of the poset. In order to do this, we will collapse a coatom and the ˆ1 of the poset. We begin with a lemma that explains the simple nature of the values of µ for the original poset and the poset obtained by collapsing a coatom and ˆ1. In the lemma and throughout the rest of the section, we will use [x] to denote the equivalence class which contains x. Lemma 3.0.3 ([6]). Let P be a poset with a ˆ0 and ˆ1 and at least 3 elements. Suppose c is a coatom and let ∼ be the equivalence relation identifying c and ˆ1. Then P/ ∼ is homogeneous and µ([ˆ1]) = µ(c) + µ(ˆ1). Moreover, if P is a lattice, then P/ ∼ is a lattice with [x] ∨ [y] = [x ∨ y] for all x, y ∈ P and [x] ∧ [y] = [x ∧ y] provided [x], [y] = [ˆ1]. Proof. First, let us show that P/ ∼ is homogeneous. Since there are at least 3 elements and we are collapsing a coatom and ˆ1, we have that ˆ0 is in its own equivalence class. Now 62 suppose that [x] < [y]. It follows that [x] = {c, ˆ1} since [x] < [y] and [c] = [ˆ1] is the ˆ1 of the quotient. Therefore, [x] = {x} and so it is obvious that P/ ∼ is a homogeneous quotient. To show that µ([ˆ1]) = µ(c) + µ(ˆ1) note that since every element of P is below ˆ1 and every other equivalence class has only one element, we get µ(y) = µ(y) = 0 y≤x y∈L([x]) for all nonzero x = c. By Lemma 2.1.4 this implies that µ([ˆ1]) = µ(c) + µ(ˆ1) which is what we wished to prove. Now suppose that P is a lattice. It is not hard to see that (P/ ∼) ∼ = (P \{c}). Therefore, if x ∨ y = c, we immediately get that [x] ∨ [y] exists and [x] ∨ [y] = [x ∨ y]. If x ∨ y = c, then ˆ1 is the only element in P \ {c} which is an upper bound for both x and y. It follows that [x] ∨ [y] = [ˆ1] = [c] = [x ∨ y]. Since P \ {c} clearly has a ˆ0, we conclude P/ ∼ is a lattice. Finally, if [x], [y] = [ˆ1] then [x] = {x} and [y] = {y} and so [x] ∧ [y] = [x ∧ y]. Let us now use Lemma 3.0.3 to prove some classic results. Corollary 3.0.4 (Hall’s Theorem [5]). Let P be a finite poset, then (−1)i ci µ(x, y) = i≥0 where ci is the number of chains of length i which start at x and terminate at y. Proof. Without loss of generality we may assume that x = ˆ0 and y = ˆ1 since all chains which 63 start at x and terminate at y are in the interval [x, y]. We prove the theorem by inducting on |P |. If |P | = 1 or |P | = 2 then the result is obvious. Now suppose that |P | > 2. Let P/ ∼ be obtained by identifying a coatom c and ˆ1. Consider the sum (−1)i ci i≥0 where ci is the number of ˆ0–ˆ1 chains of length i in P . Let ai be the number chains of length i which do not contain c and let bi be the number chains of length i containing c. Then (−1)i ci = i≥0 (−1)i ai + i≥0 (−1)i bi . i≥0 There exists a bijection between ˆ0–ˆ1 chains in P not containing c and [ˆ0]–[ˆ1] chains in P/ ∼ which preserves length. Moreover, there is a bijection between ˆ0–ˆ1 chains in P containing c and ˆ0–c chains in [0, c]. Note that in this bijection, the chains decrease by one in length. Since |P/ ∼ | < |P |, using induction we get that (−1)i ai . µ([ˆ1]) = i≥0 Similarly since |[0, c]| < |P | we get that (−1)i bi µ(c) = − i≥0 where we have multiplied the sum by −1 since the chains have decreased by one in length. 64 By Lemma 3.0.3, we have that µ([ˆ1]) = µ(c) + µ(ˆ1) or equivalently µ(ˆ1) = µ([ˆ1]) − µ(c). Therefore, (−1)i ai + µ(ˆ1) = i≥0 (−1)i bi = i≥0 (−1)i ci i≥0 which is what we wished to prove. Next, we prove a theorem of Weisner. Corollary 3.0.5 (Weisner’s Theorem [18]). Let L be a lattice and let ˆ0 = a ∈ L. If |L| ≥ 2, then µ(ˆ1) = − µ(x). x=ˆ1,x∨a=ˆ1 Proof. Let us note that if a = ˆ1, then the result is just restating the definition of µ, so we assume that a = ˆ1 for the rest of the proof. We prove the result by induction. We have already covered the case |L| = 2, since then a must be ˆ1. Now suppose that |L| > 2. Let c be a coatom such that a ≤ c. Consider, the lattice L/ ∼ obtained by identifying c and ˆ1. Since |L/ ∼ | < |L|, we get that µ([ˆ1]) = − µ([x]). [x]=[ˆ1],[x]∨[a]=[ˆ1] 65 Using the facts that [ˆ1] = {c, 1}, [x] ∨ [a] = [x ∨ a], and µ([x]) = µ(x) for [x] = [ˆ1], we obtain, µ([ˆ1]) = − µ(x). x=c,ˆ1,x∨a=c,ˆ1 Since joins are unique, we can break the sum into two parts as, µ([ˆ1]) = − µ(x) − x=c,ˆ1,x∨a=c µ(x). x=c,ˆ1,x∨a=ˆ1 If x ∨ a = c, then it is clear that x ∈ [0, c]. Moreover, since a ≤ c, it is clear that c ∨ a = ˆ1. Thus, we can remove the x = ˆ1 condition in the first sum and remove the x = c condition in the second. This gives, µ([ˆ1]) = − µ(x) − x=c,x∨a=c µ(x). x=ˆ1,x∨a=ˆ1 Now the first sum is only over [0, c] and |[0, c]| < |L| so by induction, µ([ˆ1]) = µ(c) − µ(x). x=ˆ1,x∨a=ˆ1 Using the fact that µ([ˆ1]) = µ(ˆ1) + µ(c), we immediately obtain the result. Our next corollary will make use of crosscuts. We remind the reader of a few definitions here. Let P be a poset, a subset C ∈ P is called an antichain if whenever x, y ∈ C, then x ≤ y and x ≥ y. Definition 3.0.6. Let L be a lattice. A crosscut of L is a set C with the following properties: 1. ˆ0, ˆ1 ∈ / C. 66 2. C is an antichain. 3. Every maximal ˆ0–ˆ1 chain intersects C. Although we will not need it to state the next theorem, we will need the notion of the dual of a poset in the proof. Let P be a poset. The dual of P , written as P ∗ , is the poset with the same underlying set and binary relation given by saying x ≤P ∗ y if and only if y ≤P x. Theorem 3.0.7 (Rota’s Crosscut Theorem [11]). Let L be a lattice and let C be a crosscut. Then (−1)|B| µ(ˆ1) = ∨B=ˆ1,∧B=ˆ0 where the sum ranges over all B ⊆ C such that ∨B = ˆ1 and ∧B = ˆ0. Proof. We first consider the special case when every coatom is also an atom. In this case, the crosscut must be the atom set. Moreover, a subset of the crosscut has meet ˆ0 and join ˆ1 if and only if it has at least two elements. Therefore, if L has n atoms we obtain the following n (−1)|B| = ∨B=ˆ 1,∧B=ˆ 0 (−1)|B| |B|≥2 (−1)k = k=2 n k = n − 1. This agrees with the value of µ(ˆ1) when L has n atoms and every coatom is an atom. Thus, the result holds in this special case. Recall that if L∗ is the dual lattice of L, then µL (ˆ1) = µL∗ (ˆ1). Moreover, in L∗ , joins and meets reverse roles. Therefore, if we have a crosscut consisting of only coatoms, then we can consider the dual lattice. As a result, we may now assume that there is always at least one coatom in the lattice which is not in the crosscut. With this in mind we proceed 67 by induction on |L|. If |L| ≤ 3, then it must be that |L| = 3 since smaller lattices do not have crosscuts. We have already done the case when |L| = 3. Suppose that |L| > 3 and let c be a coatom that is not in the crosscut. Consider the lattice L/ ∼ where we collapse c and ˆ1. Since c was not in the crosscut we still have the same crosscut. By induction, we know that (−1)|B| . µ([ˆ1]) = ∨B=[ˆ1],∧B=[ˆ0] Lemma 3.0.3 implies that ∨B = [ˆ1] in L/ ∼ if and only if ∨B = c or ∨B = ˆ1 in L. Additionally, since C does not contain c nor ˆ1, Lemma 3.0.3 also implies that ∧B = [ˆ0] in L/ ∼ if and only if ∧B = ˆ0 in L. Therefore, we can break the previous sum as follows (−1)|B| + µ([ˆ1]) = ∨B=c,∧B=ˆ0 (−1)|B| . ∨B=ˆ1,∧B=ˆ0 Note that if ∨B = c, then B must only have elements in [ˆ0, c]. Thus the first sum in the previous equation is over B contained in [ˆ0, c] ∩ C such that ∨B = c and ∧B = ˆ0. Since |[ˆ0, c]| < |L|, induction implies that (−1)|B| . µ([ˆ1]) = µ(c) + ∨B=ˆ1,∧B=ˆ0 Subtracting µ(c) from both sides and applying Lemma 3.0.3 we see that (−1)|B| µ(ˆ1) = ∨B=ˆ1,∧B=ˆ0 68 which completes the proof. 69 BIBLIOGRAPHY 70 BIBLIOGRAPHY [1] Andreas Blass and Bruce E. Sagan. M¨obius functions of lattices. Adv. Math., 127(1):94– 123, 1997. [2] V. V. Dotsenko and A. S. Khoroshkin. Character formulas for the operad of a pair of compatible brackets and for the bi-Hamiltonian operad. Funktsional. Anal. i Prilozhen., 41(1):1–22, 96, 2007. [3] Haya Friedman and Dov Tamari. Probl`emes d’associativit´e: Une structure de treillis finis induite par une loi demi-associative. J. Combinatorial Theory, 2:215–242, 1967. [4] R. S. Gonz´alez D’Le´on and M.L. Wachs. On the (co)homology of the poset of weighted partitions. preprint. [5] P. Hall. The eulerian functions of a group. Quart J. Math, 7(1):134–151, 1936. [6] Joshua Hallam and Bruce Sagan. Factorization of the Characteristic Polynomial of a Lattice using Quotient Posets. 26th International Conference on Formal Power Series and Algebraic Combinatorics. DMTCS Proceedings 2014, 125-136. [7] Patricia Hersh and Karola M´esz´aros. SB-labelings and posets with each interval homotopy equivalent to a sphere or a ball. preprint arXiv:1407.5311 [8] Samuel Huang and Dov Tamari. Problems of associativity: A simple proof for the lattice property of systems ordered by a semi-associative law. J. Combinatorial Theory Ser. A, 13:7–13, 1972. [9] Thomas McConville. Crosscut-simplicial Lattices preprint arXiv:1409.6269. [10] Nathan Reading. Cambrian Lattices Advances in Mathematics, 205(2): 313–353, 2006. [11] Gian-Carlo Rota. On the foundations of combinatorial theory. I. Theory of M¨obius functions. Z. Wahrscheinlichkeitstheorie und Verw. Gebiete, 2:340–368 (1964), 1964. [12] Bruce E. Sagan. Why the characteristic polynomial factors. Bull. Amer. Math. Soc. (N.S.), 36(2):113–133, 1999. 71 [13] Kyoji Saito. Theory of logarithmic differential forms and logarithmic vector fields. J. Fac. Sci. Univ. Tokyo Sect. IA Math., 27(2):265–291, 1980. [14] R. P. Stanley. Enumerative combinatorics. Volume 1 Cambridge Studies in Advanced Mathematics, 2012. [15] R. P. Stanley. Supersolvable lattices. Algebra Universalis, 2:197–217, 1972. [16] Hiroaki Terao. Factorizations of the Orlik-Solomon algebras Adv. Math. 91:45–53, 1992. [17] Hiroaki Terao. Generalized exponents of a free arrangement of hyperplanes and Shepherd-Todd-Brieskorn formula. Invent. Math., 63(1):159–179, 1981. [18] Louis Weisner. Abstract theory of inversion of finite series. Trans. Amer. Math. Soc., 38(3):474–484, 1935. [19] Thomas Zaslavsky. Signed graph coloring. Discrete Math., 39(2):215–228, 1982. 72