llll ‘ llllll'llllllllllllllllllllll 3 1293 01688 5 This is to certify that the dissertation entitled iiD€\€*1OT\-CDV\JTVQC\V‘\OH Teckniyxes :0? fine. (\fiVOMQHC Symmetric Pundits“ 0? a 6?ka presented by "Dana D. seem-A has been accepted towards fulfillment of the requirements for ?\'\o D: degree in Ma‘i’kemaltlcS l Major proféor g; Date é/s/‘I 8 MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 ' {LIBRARY Michigan State Unlverslty PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MTE DUE DATE DUE DATE DUE 1/98 chlRCIDabDuapGS—p.“ Deletion-Contraction Techniques for the Chromatic Symmetric Function of a Graph By David D. G ebhard A Dissertation Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mathematics 1998 ABSTRACT Deletion-Contraction Techniques for the Chromatic Symmetric Function of a Graph By David D. G ebhard Recently, R. P. Stanley defined and studied a symmetric function, XG, which generalizes the chromatic polynomial of a graph, G. This generalization has both ad— vantages and disadvantages. The main advantage is that it gives us more information about the colorings of G than the chromatic polynomial. However, one disadvantage is that this new symmetric function does not satisfy a deletion-contraction recurrence similar to the one for the chromatic polynomial. In this thesis, we define a similar graph invariant called YG. .This invariant is defined using noncommutative variables, and from it we can recover XG by allowing the variables to commute. This new invariant is also a symmetric function. More importantly, by using noncommutative variables we will be able to obtain a deletion- contraction recurrence for YG. We may then obtain some of Stanley’s results for X G in a uniform manner by using induction. In addition, this will allow us to make some progress on the 3+1 Conjecture of Stanley and Stembridge. ACKNOWLEDGMENTS I would like to thank my advisor, Bruce Sagan, for just about everything you can imagine. The patience he showed as he went over the first draft of this thesis was nothing short of remarkable. His help and support on my way to finishing this work were wonderful, and for that I give him my thanks. The way he went over this manuscript with a fine-toothed comb and made it a readable document was impressive, and for that you should give him thanks! I would also like to thank Mark McCormick for all his help with my [MEX files, for his friendship and for helping to keep me in shape. Similarly, I would like to thank Kevin and Melissa Dennis for their friendship, their food, and fun and games. Finally, I need to thank my family, especially my parents and my brother for their love, support, and encouragement through this (to them) seemingly endless journey towards my Ph.D. iii TABLE OF CONTENTS INTRODUCTION 1 Preliminaries 1.1 Symmetric Functions ........................... 1.2 The Chromatic Symmetric Function, XG ................ The N oncommutative Case 2.1 Symmetric Functions in Noncommuting Variables ........... 2.2 Development and Results for YG .................... Orientations and Sinks 3.1 Acyclic Orientations ........................... 3.2 The Modified Blass-Sagan Algorithm .................. Results on e-positivity 4.1 Inducing e,r ................................ 4.2 Some e-positivity Results ......................... 4.3 The (3+1)-free Conjecture ........................ Open Problems and Conjectures 5.1 Partitioning Acyclic Orientations .................... 5.2 X0 and Trees ............................... BIBLIOGRAPHY iv 1 7 7 10 15 15 20 29 29 34 41 41 48 56 67 67 70 74 INTRODUCTION As early as 1912, Whitney [19] began to study graph colorings from a mathematical point of View. Today the theory of graph coloring has many applications to both scheduling problems and efficient network design. [11]. Here we will use symmetric functions to enumerate graph colorings. While this section contains much of the background leading up to our study, we will try to introduce notation and definitions as they are needed throughout the text, rather than all at once. We will generally follow Stanley [15, 14] for combinatorial notation or anything specifically related to the symmetric function of a graph, X0, and MacDonald [10] for symmetric functions in general. To begin, let G be a finite graph with vertex set V(G) and edge set E (G), where the edges consist of unordered pairs of the vertices. We mention here that if the edge set consisted of ordered pairs of vertices we would have had a graph with directed edges, referred to as a digraph. A vl—vn walk in a graph is a sequence of vertices, 221,112,... ,2)” such that 22,--122, is an edge for all 2 g i S n. A graph, G, is connected if there is a u,v walk for every pair of vertices, u and v in V(G). The connected components of G are just the maximal connected subgraphs of G. Finally, H is a spanning subgraph of G if V(H) = V(G) and E(H) Q E(G). In our study we will actually consider multigraphs, in which multiple edges and loops are allowed. The other definitions above extend in the natural way to multigraphs. Since our main interest here is in coloring graphs, we define a coloring of G to be 1 2 2 c o 0 v1 v2 123 Figure 1. A coloring (not proper) of P3. 1 2 3 1 2 1 k t o o 4 0 ’U 1 02 ’03 ’01 v2 03 Figure 2. Two prOper colorings of P3 a map a : V(G) -—> C, where C is the color set. In particular, a proper coloring of G is a coloring such that no two adjacent vertices are the same color, i.e., 0(1),) 75 01(2),) if vivj is an edge of the graph. For an example, we show a coloring for the path on three vertices, P3, which is not a proper coloring in Figure 1 and two proper colorings for P3 in Figure 2. Whitney’s object of study was the chromatic polynomial of a graph, XG(n), which is defined to be the number of ways to properly color G using the color set C = {1,2, . . . ,n déf [n]. For P3, since there are n ways to color 111 from a set of n colors, and n—1 ways to color each of the remaining vertices, we see that Xp3(n) = n(n—1)2. It is somewhat surprising that XG(n) is always going to be a polynomial in n. One easy way to see this is to use induction along with the Deletion-Contraction Lemma, which we will now discuss. Given a graph G and an edge e E E (G), we can define the graph G -— e to be the graph G with the edge e deleted from its edge set. The contraction of G by e, G / e, is obtained from G by contracting e (in the topological sense) to a single vertex. Given these definitions, the Deletion-Contraction Lemma states that XG(n) = XG_e(n) — XG/e(n). This gives us a recursive way to compute the chromatic polynomial of a graph, as well as to establish various properties of XG(n) by induction. Two of Whitney’s results that can be proven using this method are stated here. Theorem 1 [1.9] For a finite graph, G, 260(71): 2: (—1)'3'nc<5>, sc_:E(c) where C(S) is the number of connected components of the spanning subgraph of G with edge set S, which by abuse of notation we just denote by S. I As an illustration, we will use this theorem to again calculate Xp3(n). If we let the edge set of P3 be {e1,e2}, where el = 111112 and e2 = v2v3, then we can make the following table. S C E(G) (—1)lSl 7245) 1 n3 e1 —1 72.2 e2 —1 n2 e1,e2 1 nl This shows us that according to the Theorem, Xp3(n) = n3 — 2n2 + n :2 n(n - 1)2, which agrees with our previous calculation. The other theorem of Whitney’s in which we will be interested is the one known as the Broken Circuit Theorem. A cycle or circuit is a closed walk with distinct vertices and edges, v1,v2,. .. ,vm,u1, for m 2 1. If we fix a total order on E(G), a broken circuit is a circuit with its largest edge (with respect to the total order) removed. Let the broken circuit complex 80 of G denote the set of all S g E (G) which do not contain a broken circuit in our fixed ordering on the edges. The Broken Circuit Theorem then asserts: Theorem 2 [19] For any finite graph, G, on d vertices we have xG(n) = Z (—1)|Slnd-IS|. 5630 I If we again calculate Xp3(n) using this theorem, we will come out with exactly what we had before, only with n3 and n1 reversing positions in the table, since P3 contains no circuits and hence no broken circuits. As a less trivial example, we will use this theorem to verify that the chromatic polynomial for K 3, the complete graph on 3 vertices is indeed given by n(n — 1)(n — 2), which can be obtained by noticing that there are n ways to color the first vertex, n —1 colors left available for the second vertex, and n — 2 colors allowed for the last vertex. We label E(K3) = {e1,e2, e3}, where the fixed order on the edges is the obvious one induced by the subscripts. Since the only circuit in K 3 is {e1, e2, e3}, the only broken circuit will be {e1, e2}. This gives us the following table, where we notice that here d = 3. S 6 Ba (—1)|5| rid—'5' d) 1 n3 e1 —1 n2 e2 —1 n2 e3 —1 77.2 e1, e3 1 it1 e2, e3 1 n1 This gives us XK3 = n3 — 3n2 + 2n, which again agrees with the previous calculation. Following these early results, some of the more interesting applications are those of Zaslavsky in [20, 21, 22]. In that series of papers he introduces the notion of colorings for certain generalizations of graphs called signed graphs. These colorings have very nice connections to characteristic polynomials of certain types of hyperplane arrangements. A related result by Zaslavsky and Greene [7] concerns the sinks of acyclic orientations for G. An orientation of G is a digraph D obtained by assigning a unique direction to each edge of G. An orientation is acyclic if it has no directed cycles. We also define a sink of D to be a vertex v E V(D) such that fit’ ¢ E (D) for all a: E V(D). Also, for notational convenience we adopt the convention that XG(n) = a0 + aln + agn2 + - - - + akn". Theorem 3 ([7] Theorem 7.3) Let v0 be any vertex of G. The number of acyclic orientations of G with a unique sink at v0 is [a1]. I This theorem is related to one of Stanley, which states: Theorem 4 [13] The number of acyclic orientations ofG is Z, |a,~[. I All of these theorems are actually specializations of results which can be obtained from Stanley’s symmetric function generalization of the chromatic polynomial. The first three theorems listed previously can all easily be derived from the recurrence relation for the chromatic polynomial. However, this symmetric function does not satisfy any similar deletion-contraction recursion, which eliminates induction as a tool for these proofs. In what follows we will extend the Stanley’s definition by using symmetric functions in noncommutative variables. This setting will allow us to establish a recurrence and again allow induction as a valid approach to our proofs. CHAPTER 1 Preliminaries 1 . 1 Symmetric Functions Here we will review the basic facts about symmetric functions in commuting variables. Our development will closely mirror that found in Sagan’s book [12]. The interested reader should consult either MacDonald [10] or Sagan [12] for a more comprehensive discussion. We will begin with the monomial symmetric functions. Let x = {$1, $2,223, . . .} be a countably infinite set of commutative variables, and let A = (A1,A2, . .. ,Ak) be an integer partition of n, denoted A l- n, where the A, form a weakly decreasing sequence of positive integers such that 2le A,- = n. If we allow r,- to be the number of parts of A equal to i, then we may also express A = (1", 2’2, . . . ,n'") as an alternate notation. The monomial symmetric function corresponding to A is given by _§:A1A2,H )‘k mA— (11111132.? (III-k, where the sum is over all distinct monomials having exponents A1, . .. ,Ak. As an example we can see that 2 2 2 2 2 m(2,1) —$1$2+$1$3+”'+$2$1+$2$3+°"+$3$1 +"' . We then define the ring of symmetric functions as the vector space over C spanned by the monomial symmetric functions. It is an elementary fact that the monomial symmetric functions are actually linearly independent over C and so form a basis for the vector space of symmetric functions. It is important to note here that while we will usually consider the symmetric functions as a vector Space in this thesis, the fact that they form a closed set under multiplication also makes the symmetric functions a ring. While it is clear from our development that the monomial symmetric functions form a standard basis for the symmetric functions, there are other nice bases for this vector space which are routinely used. These include the elementary, power sum, and complete homogeneous symmetric functions as well as the Schur functions. We will define the power sum symmetric functions and the elementary symmetric functions here, as they will be relevant to the rest of this thesis. For a description of the complete homogeneous symmetric functions and the Schur functions, please see either [10], [12], or [4] The rth power sum symmetric function is Pr =m IP, and IP’ is the set of positive integers. Note that XG is homogeneous of degree d = [V(G)], where [4:] denotes cardinality. We also notice that if G has loops this sum is empty, giving X0 = 0. To illustrate this definition, we will compute the chromatic symmetric function for our standard example of the path on three vertices, P3. We can see that any proper coloring of this graph will have one of two possible types: the coloring could have v1 and v3 one color with v2 3 different color, or it could have all three vertices different colors. Since there are 6 different ways to color the three vertices with the same set of three different COIOTS, we Obtain _ 2 + 2 Q . I 6 + 6 + o . . X123 — 313:2 11321131 + + 1131172233 231332234 It should be clear from the definition that X0 is a symmetric function, since any permutation of the subscripts simply permutes the colors and doesn’t affect the set of colorings. We can also see it more explicitly in this case, since the previous expression 11 clearly shows that X133 2 m(2,1) + 67720,”). For A l- n having r,- parts of size i, we can also use the notation A = (1’1,2’2,... ,n’"), and define [A] = r1! - - -r,,!. We say that a partition of the ver- tex set of G is stable if no block of the partition contains adjacent vertices. Then it is not hard to see [15] that X0 = 2A a,\|A|m,\, where a), is the number of stable partitions whose block sizes correspond exactly to the parts of A. We are also easily able to see that for the disjoint union of two graphs, G = H as], we have X G = X EX 1. We can verify that this symmetric function is a generalization of the chromatic polynomial, XG(n), since setting :51 = x2 = = a2" = 1 and 1:,- = 0 for all i > n in X0, denote by XG(1"), yields XG(n). To see this, note that this substitution will produce a term equal to 1 for each monomial in X0 which comes from a proper coloring of the graph using the first it colors, and a term equal to zero for each monomial arising from a proper coloring which uses a color not in [77.]. Hence the sum of all these monomial terms after this substitution will just be the number of proper colorings of G which only use the first n colors. This is precisely XG(n). Once we are assured that this is a generalization of the chromatic polynomial, one might expect that previous results about the chromatic polynomial should also generalize. It is also natural to study the expansion of this chromatic symmetric function in terms of the different symmetric function bases. The calculation of X0 for various specific graphs is also of interest. Stanley pursues all of these lines of inquiry in his paper. Several of Stanley’s results for XG are extensions of Whitney’s [19] theorems for the chromatic polynomial. For example, Stanley’s symmetric function extension of Theorem 1 utilizes the power sum symmetric functions. 12 Theorem 1.2.2 [15, Theorem 2. 5] Let G be a finite graph of order d. We have X0 = Z (‘Ulslpusp 599(0) where A(S) = (A1, A2, . .. ,Ak) is the integer partition of d with A,- being the number of vertices in the it” component of S. I We can see that this result directly implies Whitney’s first theorem by noticing that pr(1") = n for any r, and so pA(1") = n“), where l(A) is the number of parts of A. Hence p,\(3)(1") = n45), completing the reduction. Not surprisingly, Stanley also has a generalization of Theorem 2. Theorem 1.2.3 [15, Theorem 2.9] For any finite graph G, we have XG : Z (—1)lslp,\(5). 3630 I In this thesis, we will be studying an analogue of Stanley’s chromatic symmet- ric function X0, called Ya, which is defined using noncommutative variables. We wish to consider this analogue because we know that many results for the chromatic polynomial can be proven easily using induction and the deletion-contraction recur- rence. Unfortunately, Stanley’s symmetric function has no such deletion-contraction property, which deprives him of induction as a tool for his proofs. To see where the problem lies, note that X0 is homogeneous of degree d, while XG/e is homogeneous of degree d — 1. In order to find a recurrence, we would need to add another variable to each monomial in X G )8. But which variable? In the proof of the deletion-contraction rule for the chromatic polynomial, we have proper colorings of G / e corresponding to colorings of G - e with u and v the same color, where e = uv. However, while XG/e gives us more information about the colorings of G / e than the chromatic polynomial, it does not give us the explicit information we need to fix the 13 problem: namely, what color was assigned to the vertex obtained by contracting uv. We lost that information when we allowed the variables to commute. To correct this difficulty, in Chapter 2 we introduce an analogue of X0 which is a symmetric function in noncommutative variables. That is, for any multigraph G with vertices labeled v1, v2, . . . ,vd in a fixed order we define the analogue of X0 as Y0 = 2%(voxnaz) ' ' '$n(v.) where again the sum is over all proper colorings of G, but the 12:,- are now noncom- muting variables. The reason for using noncommuting variables is so that we can keep track of the color which n assigns to each vertex. Since we still have the homogeneity problem in YG/e, we define an operation on the non-commutative symmetric functions which will allow us to use deletion-contraction techniques for computing Ya. In this chapter we also provide some basic expansions of YG which closely resemble Whitney’s and Stanley’s theorems. In Chapter 3 we will further explore the interrelationships between chromatic polynomials, chromatic symmetric functions, acyclic orientations and sinks. There is an interesting connection between Theorem 2 and the result of Green and Zaslavsky, Theorem 3. From Theorem 2 we can interpret the coefficients of the chromatic poly- nomial as the number of sets S E BG of a certain size. From Theorem 3, we see the coefficient of n in XG(n) is the number of acyclic orientations of G with a unique sink at any fixed vertex of G. It follows that the number of acyclic orientations of G with a unique sink at the fixed vertex is the same as the number of sets S 6 BC with IS] = d — 1. Elementary graph theory tells us that this is also be the number of spanning trees of G which contain no broken circuits. We will provide a bijective proof of this fact by modifying an algorithm due to Blass and Sagan [1]. Finally, 14 we will also extract some of the information that the non-commutative chromatic symmetric function can give us about acyclic orientations and sinks. In Chapter 4 we will consider a conjecture about the coefficients of Xg, when it is expanded in terms of the elementary symmetric function basis. We will make some progress here on the (3 + 1)-free conjecture of Stanley and Stembridge, proving it in some special cases. We finish in Chapter 5 with some other partial results about acyclic orientations and sinks, as they relate to the (3 + l)-free conjecture. We conclude with some open problems, as well as a few ideas on how they might be approached using our techniques. Before we begin, however, we will need to discuss symmetric functions in noncommuting variables and our analogue of X0 in that setting. This is the focus of our next chapter. CHAPTER 2 The Noncommutative Case 2.1 Symmetric Functions in Noncommuting Vari- ables We begin with some background on symmetric functions in noncommuting variables. Much of this follows from the work of Doubilet [4], although he does not explicitly mention these functions in his work. These noncommutative symmetric functions will be indexed by set partitions, which form a lattice under a certain partial order, which we will define here. A lattice is a poset (partially ordered set) .C such that every pair 2:, y E C has a least upper bound (or join) denoted by a: V y and a greatest lower bound (or meet) denoted :1: /\ y. Any finite lattice has a unique minimal element denoted by D and a unique maximal element denoted by 1. We let IId denote the set partitions of {1,2, - - - , d} = [d]. This forms the set partition lattice, where the partial order is defined as follows. Ifo = Al/Ag/ - - -/A;c and r = Bl/Bg/ - n/Bm, then a g r if and only if for all 1 g i _<_ k there exists some 3' with 1 S j S m 30011 that Ai g B)" That this partial order on 114 actually forms a lattice is an elementary result. As an example, We have included the Hasse diagram for I14 in Figure 2.1. Given a poset, 15 Figure 2.1. The partition lattice I14. P, we also recursively define the Mb'bius function, u, of P on intervals [33, y] in P by p(:r,:c) = 1 and u(a:,y) = — Z n(x, z) for all x,y E P. zSz1r [1(0,T) a)T= 552T where it is a proper coloring of G /e, and the vertex obtained from the contraction is labeled vd_1. Thus we have Yg_e 2 Ya + YG/e T. Rearranging the terms gives YG = YG—e — YG/eT- I We note that if e is a repeated edge, then the proper colorings of G — e are exactly the same as those of G. The fact that there are no proper colorings of the second type corresponds to the fact that G /e has a loop, and so it has no proper colorings. Also note that since we allow multiple edges when contracting e we have |E(G - ell = |E(G/6)| = |E(G)l - 1- The Deletion-Contraction Proposition for Ya gives us a tool for using induction. For 7? 6 I'Id_1 we let it + (d) 6 IL; denote the partition 7r with d inserted into the 24 block containing (1 — 1. From equations (2.1) and (2.2) it is easy to see that mnT: m1r+(d) and 19an pn+(d)- With this notation we can now provide an example of the deletion-contraction propo- sition for P3, where the vertices are labeled sequentially, and the distinguished edge is e = v2v3. YP3 : YP2BJ{U3} _ YP2T ' It is not difficult to compute YP2w{v3} = m1/2/3 + m1/23 + "113/2, YP2 = ml/Za YP2T = 7711/23- This gives us YPs = m1/2/3 + "ll/23 + m13/2 — m13/2 — m1/2/3 + "113/2, which agrees with our previous calculation. We may use this recurrence to provide noncommutative analogues of the previ- ously cited results of Stanley, where the proofs now follow from induction. Theorem 2.2.5 For any finite labeled graph G, Y0 = Z(—1)Islpn(5), s<_:E where 7r(S) denotes the partition of {1, 2, . . . ,d} associated with the partition of V(G) 25 into the connected components of S. Proof. We induct on the number of non-loops in E (G) If E (G) consists only of n loops, for n 2 0, then for all S Q E(G), we will have 7r(S) = 1/2/ ' - -/d. This shows us that P1/2/.../d if n = 0, Z(—1)|Slp7r(s) =Z(-1)ISIP1/2/.../d :2: (i) (_1) P1/2/.../d = i=0 0 if n > 0. sge ng This agrees with equation (2.6). Now, if G has edges which are not 100ps, we use the Relabeling Pr0position to obtain a labeling for G with e = vd_1vd. From the Deletion-Contraction Proposition we know that YG = YG__e — Yg/eT and that both G —— e and G / e have one fewer edge than G. This allows us to apply induction to YG_e and Ya/e, obtaining Ya: Z) (—1>'S'p.(5)— Z (—1>'§'p.(S-,T. S§E(G—e) S§E(G/e) It should be clear that Z (‘1)IS'PMS): Z (-1)'S'pn(3). S§E(G—e) S§E(G) e¢S Hence it suffices to show that if e E S, — Z (‘Ulglpfigfl = Z (—1)'S|Pn(5)- S§E(G/e) sgsga) 8E To do so we define a map 6 : {S g E(G/e)} -> {S Q E(G) 3 6 E S} by 9(S) = S, where S = S U e. 26 Then 9 is a bijection, since we allow multiple edges to occur when we contract e to vd_1 . We know that [S] +1 2 IS] and n(S) = 7r(S) + (d), giving p,,(5) = pfigfl‘. Thus we have " Z (‘1l'S'Pn(S)T = Z (_lllswl’rw?)T S§E(G/e) S§E(G/e) = Z (-1)'S'P«(S)- S§E(G) e65 This completes the proof. I By letting the x,- commute, we then obtain Theorem 1.2.2 as a corollary. There are also other results which we may obtain by this method, such as Stanley’s gener- alization of Whitney’s Broken Circuit Theorem. Before we prove this, however, we will need the following lemma, which appeared in [1]. For the sake of completeness, we include a proof here. Lemma 2.2.6 For any non-loop e, there is a bijection between BG and 80.3 U Bg/e given by ~ S=S—eeBG/e ifeeS SZSEBG_3 ife¢S, S——> where we take 6 to be the first edge ofG in the total order on the edges . Proof. It is enough to show that this map is well-defined, and that it has a well- defined inverse. To show that it is well-defined, we let S E BC with e E S. Notice that e is neither a loop nor a multiple edge: if e were a loop, then d) is a broken circuit, and so e gt S. If e and e’ have the same endpoints, then by the assumption that e is minimal, e is 27 a broken circuit, and so again e g! S. If e a! S it is clear that S E BG_e, since any broken circuit of G — e is also a broken circuit of G. If e E S, we let S = {e, e2, . .. ,ek} be listed in increasing order, according to the total order, 3, on the edges. Then S 2 {e2, . .. ,ek} does not contain any broken circuits of G / e, for if S’ was a broken circuit of G / e contained in S, then S’ U{em} would be a circuit of G/e for some em larger than any element of S’. But then S’ U{em, e} would contain a circuit of G. So S’ U{e} would contain a broken circuit of G, since e g em. This contradicts S 6 Ba. To construct the inverse, we simply map In - S If S 6 80-8 S ——-> ~ ~ SU{€} if S E BG/e- It is clear that this is the inverse, provided again that the map is well-defined. An argument similar to the one given above shows that this is indeed the case. I We can now obtain a characterization of YG in terms of the broken circuit complex of G for any fixed total ordering on the edges. Theorem 2.2.7 We have ya = Z (‘1)ISIPn(S), 3686 where again n(S) denotes the partition of {1,2, . .. ,d} with blocks corresponding to the connected components of S. I Proof. We again induct on the number of non-loops in E (G). If the edge set consists only of n loops, it should be clear that for n > O we will have every edge 28 being a circuit, and so the empty set is a broken circuit. Thus we have Zse¢(—1)|S|Pn(5) : 0 if n > 0, Zse{¢}(-1)IS|P7r(S) = P1/2/.../d if R = 0. Y0 = This again matches equation (2.6). For n > 0 and e a non-loop, we again consider Y0 = YG_e — YG/CT, where G —- e and G / e both have one less edge than G, and so induction applies. From the preceding lemma and arguments as in Proposition 2.2.5, we have 2 (-1)Sp..(5)= Z (-1)Spn(5) SGBG SEB _ eQS G c and Z (—1)'S'p.(3)=— Z (—1)'§'p.(S-,t S93(0) SEB c 865 G/ which gives the result. I CHAPTER 3 Orientations and Sinks 3.1 Acyclic Orientations As we have seen in the introduction, there are some interesting results which relate the chromatic polynomial of a graph to the number of acyclic orientations of the graph and to the sinks of these acyclic orientations. We begin here by proving Theorem 3. While the result is not new, we do offer a new proof here more in keeping with the spirit of the other proofs in this thesis. We denote the set of acyclic orientations of G by A(G), and the set of acyclic orientations of G with a unique sink at v0 by A(G, v0). We also recall our notation that XG(n) = a0 + aln + ' - - + aknk. Lemma 3.1.1 For any fixed vertex v0, and any edge e = uvo, u 75 v0, the map D—e€A(G—e,v0) ifD—eEA(G——e,v0) D/e E A(G/e,v0) ifD — e ¢ A(G — e, v0), D—-> is a bijection between A(G, v0) and A(G—e, vo) ErJA(G/e, v0), where the vertex of G/e formed by contracting e is labeled v0. Proof. We must first prove that this map is well-defined, by showing that in both cases we actually obtain an acyclic orientation with unique sink at v0. This is clear 29 30 in the first case by definition. In the second, where D — e 5! A(G — 6, v0), it must be true that D — c has sinks both at u and at v0 (since deleting a directed edge of D will not change the acyclic property of the orientation, nor can it cause us to lose the sink at v0). So the orientation D/e will be in A(G / e, v0): since u and v0 were the only sinks in D — uvo the contraction must have a unique sink at v0, and there will be no new cycles formed. Hence this map is well-defined. To see that this is actually a bijection, we need only exhibit the inverse. This is obtained by simply orienting all edges of G as in D — uvo or D/uvo as appropriate, and then adding in the oriented edge 2E3. It should be clear that this map is also well-defined. I Lemma 3.1.2 IfG is connected, then any D E [A(G)| has at least one sink. Proof. While this is a well-known graph theory result, we prove it here for complete- ness, by way of contradiction. Consider the finite set of directed walks in A(G) given by S={’Uil—)’Ui2—>...—)’Uikivi,EV(G) fOI‘ISlSk, andk§|V(G)[+1}. Clearly S 75 d5, since for any vertex v E G, the trivial walk given by v will be an element of S. So we may consider a walk, W, in S with maximum k. If k = |V(G)|+1, then W contains a cycle, contradicting D E A(G). If k _<_ |V(G)|, then we claim that wk will be a sink of D. If it is not a sink, then there is a directed edge e = 17,15, for some w E E (G) Adding this directed edge to W will again give us a walk in S. But this contradicts our choice of W. Hence vi, must be a sink of D, completing the proof. I As an immediate corollary we have the following result. 31 Corollary 3.1.3 For any D E |A(G)|, the number of sinks is greater than or equal to the number of components of G. I Using Lemma 3.1.1 and Corollary 3.1.3, we may now prove Theorem 3 by showing that the boundary conditions and recurrence relations for [all and the number of acyclic orientations with a unique sink at a fixed vertex, v0, are the same. Theorem 3.1.4 [7] For any fixed vertex v0, the number of acyclic orientations of G with a unique sink at v0 is [a1]. Proof. We prove this theorem by showing that the boundary conditions and recur— rence relations for |a1| and |A(G, v0)| both match. Our recurrence will only be valid when there is a non-loop incident with v0, and so our boundary condition will occur in the case where only loops are incident with v0. If d = 1, then Tl If G 2 K1, X601) = 0 if G has loops. So in this case, 1 If G 2 K1, [all = = me, 210)]. 0 if G has loops If d > 1, then having only loops incident with v0 is equivalent to having at least two components in G. In this case we see from Theorem 1 that [a1] = 0 and from Corollary 3.1.3 that [A(G, v0)[ = 0 as well. Thus the boundary conditions match. Now if there is a non-loop e incident with v0, we can see that [all is the sum of the absolute values of the coefficients of n from XG_e and XG/e since the signs on the coefficients of the chromatic polynomial alternate. Hence it follows from Lemma 3.1.1 that the recurrence relations are also the same and so the theorem is proven. I Stanley has a stronger version of this result. 32 Theorem 3.1.5 [15] If X0 = Z, cAeA, then the number of acyclic orientations ofG with j sinks is given by Z c,\. I l(A)=J’ We can prove an analogue of this this theorem in the noncommutative setting by using techniques similar to his, have not been able to do so using induction. We can inductively demonstrate the weaker versions which follow. Theorem 3.1.6 Let YG = 2 one”. Then for any fixed vertex, v0, the number of «611,, acyclic orientations of G with a unique sink at v0 is (d — 1)!c[d]. Proof. We again induct on the number of non-loops in G. In the base case, if all the edges of G are 100ps, then el/g/W/d if G has no edges Y0 = 0 if G has loops. Cid] = = [A(G,’U0)|. 0 ifd>1orGhasloops If G has non-loops, then by the Relabeling Proposition , we may let e = vd_1vd. We know that Y0 = YG\e - YG/e T. Since we will only be interested in the leading coefficient, let Y0 = aem + Z aaea, o<[d] Ya. = bet] + Z baa... a<[d] 33 and YG/e : C€[d—1] + Z Coed o<[d-—1] where S is the partial order on set partitions. By induction and Lemma 3.1.1, it is enough to show that (d — 1)!a = (d — 1)!b + (d — 2)!c. For 7r 6 IId_1 we recall that 7r + (d) 6 IL; is obtained by inserting d into the block of it which contains d — 1. It was noted before that for it E 11.1-1, we have p,,T= p,,+(d). We utilize the change of basis formulae from equations (2.2) and (2.4) to obtain enT= Z ”(0 a) )2 p( 7‘, o +( (3.1) )r<0'+(d) With this formula, we compute the coefficient of em from YG/CT. The only term which contributes comes from ce[d_1]T, which gives us ce[d_1]T = c X [NO/fooffil) )2 p(,(dro+ o€Hd_1)r 220 path for every 3: E V(D). (d) The unoriented part of D contains no broken circuit. From these conditions, it should be clear that Do is indeed the set of acyclic orientations of G with a unique sink at no, since any finite weakly connected acyclic digraph has at least one sink. It is also clear that any element of R, will be an acyclic, connected graph, which implies that the elements of Q, must be trees with exactly d — 1 edges. So provided the algorithm produces a bijection at each step, we will produce the desired bijection between acyclic orientations of G with a unique sink at 110, and edge sets of size d — 1 which contain no broken circuits. We should also note here that conditions (b) and (c) together imply that c(D) must have a unique sink which occurs at the vertex identified with uo. That this is the only possible sink of c(D) is clear from condition (c). We also know that 220 must be a sink of c(D), since if it is not, then there is a vertex u and are a = m in c(D). But from condition (c) there would have to be a u —-> v0 path in D. This contradicts (b), which asserts that D is acyclic. To show that the algorithm does indeed produce a bijection at each step, we use the following three lemmas. We also use the notational convention that a digraph in D1, will be denoted by Dk. Lemma 3.2.2 Ak maps ’Dk_1 into Dk. Proof. We need only prove that each of the properties (a)-(d) listed above is satisfied after the algorithm is applied at the kth step. We proceed to prove each one in turn. (a) Since at the kth step the algorithm will either delete or unorient the kth edge, this is clear. 38 (b) Since any edge which would form a cycle if unoriented will be deleted by the algorithm, this also is clear. (c) Since unorienting an arc can never destroy an a: ——> '00 path, we need only consider the case where the algorithm deletes an arc. In fact, if the arc a 2 17b in Dk_1 was deleted, we need only show that there is still a w —+ uo path. Now, if the arc a = 171% in Dk-1 was deleted for the first reason, then we must have had another (different) w —+ u path in Dk_1. Since there was a u ——> v0 path in Dk_1, (in fact, one which didn’t use the arc a) we can then extend our other w —> u path into a walk containing a w —> 110 path in Dk. If the arc a = 21713 in Dk_1 was deleted for the second reason, again we need only consider the possibility that for the vertex w, there is no w —> uo path in DC. But then there is no oriented arc 1172’ with u 75 u’, since otherwise all u’ —+ on paths must also use the arc a, as there are no w —> v0 paths in Dk. Thus Dk_1 would have a cycle containing w. Contracting all unoriented arcs from w and repeating this argument as necessary, we see that w would then be a sink of c(Dk_1) — a, which contradicts our reason for deleting a. (d) Suppose for the sake of contradiction that the unoriented part of D, contains a broken circuit, C — x, where :1: is the greatest element of the cycle G. Since the un- oriented part of Dk_1 didn’t contain any broken circuits, and since the only difference between Dk_1 and Dk is at the kth edge a, we see that a must be unoriented in Dk and that a E C — 1:. But then a: is greater than a, and so :2: is present in DC in one of its orientations. But all the other edges in C’ are also present and unoriented. Hence, C forms a cycle in Dk, contradicting the previously verified fact that Dk is acyclic. I Lemma 3.2.3 Ak is one-to-one. Proof. Suppose D1 and D2 are two distinct elements of Dbl which are both mapped to D by the algorithm. Since the algorithm only affects the kth edge, we 39 note that D1 and D2 (and consequently c(Dl) and c(D2)) must only differ in the are a. Without loss of generality, we may assume that a has an abnormal orientation in D2 and a normal orientation in D1. We note that D was not obtained from D1 and D2 by the deletion of the arc a: we know that a could not have been deleted from either D1 or D2 for forming a cycle, as the other would then have contained that cycle. Also, if a was deleted by the algorithm from both D1 and D2 for the second reason, then a was abnormally oriented in both D1 and D2, which is not possible. If a was not deleted by the algorithm, then c(Dg) — a has an additional sink. So if a = u'J—b is the arc in C(Dg), then w 79 220 since 220 is a sink, and so w must be the additional sink in c(D2) — a . But this means that w was already a sink in c(Dl), contradicting D1 6 ’Dk_1. I Lemma 3.2.4 A, maps Dk_1 onto Dk. Proof. Given Dk 6 Pk we must construct Dk_1 E ’Dk_1 which maps onto it. Hence for any digraph, D1,, 6 Dk, we must construct a digraph Dk_1 and verify that the algorithm does indeed map Dk_1 onto D1,, and that Dk_1 satisfies properties (a)- (d). For all of the following cases, it will be immediate that the Dk_1 we construct will satisfy properties (a), (b), and ((1), so we will only show the verification of property (c). Let a be the kth edge of G. There are two cases. The first case is when a $4 Dk. If there exists a unique orientation of a in which Dk would remain acyclic, we give a that orientation in Dk_1. If both orientations of a would preserve the acyclicity of D1,, then we choose the abnormal orientation for a in Dk-1. We note that at least one of the orientations of a must preserve acyclicity, since otherwise a completes two different cycles in Dk_1. These two cycles together would contain a cycle in DC, which is a contradiction. 40 That the algorithm maps the digraph Dk_1 obtained in the previous paragraph to D, is obvious when only one orientation of a produces an acyclic orientation of Dk_1. However, if both produce acyclic orientations, we need to check that c(Dk_1) —a has a unique sink at 220. This is true, since it is easy to see that c(Dk_1) —a = c(Dk_1— a) = c(Dk). To verify that c(Dk_1) constructed above still satisfies property (c), we note that adding an arc cannot destroy any existing paths. For the other case, we suppose that a is present in Dk as an unoriented edge e, and so neither orientation can produce a cycle in Dk_1 . We note that there must be at least one orientation of e = wu such that there remains an a: —> 220 path for every x E Dk_1. If all a: —> 110 paths p use the are a = M for some 1', and if all y —+ 210 paths q use a’ = M, then the a: —+ w portion of P together with the w —> v0 portion of Q contains an a: —> v0 path avoiding a, which contradicts our assumption about 11:. If there is a unique orientation of e = wu so that there remains an a: ——> uo path for every :1: E Dk_1 we choose that one to maintain property (c) for Dk_1, say a = 13%. Using the same argument we used to prove the second case of (c) in Lemma 3.2.2, it is easy to verify that the algorithm will take the Dk_1 so constructed and map it to D], by unorienting a since c(Dk_1) — a has an additional sink at w . In the subcase where e = uw is present in D], as an unoriented edge and we would still retain prOperty (c) with either orientation of e, we will consider the digraph Dk_1 obtained from D by giving 6 the normal orientation, say a = 27715. It is clear that the algorithm maps Dk_1 to D1,, since Dk_1 + a’ = D1,, is acyclic and a has the normal orientation. I CHAPTER 4 Results on e-positivity 4. 1 Inducing e7r We now turn our attention to the expansion of Y0 in terms of the elementary sym- metric function basis. We recall that for any fixed 1r 6 IL, we use 7r + (d+ 1) to denote the partition of [d + 1] formed by inserting the element (d + 1) into the block of 7r which contains d. We will denote the block of 7r which contains d by B,,. We also let 7r/d + 1 be the partition of [d + 1] formed by adding the block {(1 + 1} to 7r. In order obtain information about the coefficients for the expansion of Y0 in non- commutative elementary symmetric functions using our deletion-contraction results, it is necessary for us to understand the coefficients arising in e,r T. We have seen that the expression for ewT is rather complicated (see equation (3.1)). However, if the terms in the expression of enT are grouped properly, the coefficients in many of the groups will sum to zero. To see that such a grouping should exist, we use the following lemma. Lemma 4.1.1 For 7r 6 11¢, let esz Z: CTeT. Then 76 r1«1+1 41 42 1/|B,,| if /\('r) = /\(7r/d+1), Z 0.: —1/|B,.| if/\(r)=)\(7r+(d+1)), Tend-H A(T)fixed 0 8188. Proof. Let 7r : Bl/Bg/ . . . /B,,/ . . . /B,c E Iln. We also let b,- denote the size of the block 8,. We may now consider the graphs on d + 1 vertices given by GwZKB,EHKBQEHH'HZWKB,+6)EH°"L+JKBk, where K B, is the complete graph on b, vertices labeled with the elements of Bi, and where K 3,, + e is the complete graph on b,r dr‘if b vertices labeled with the elements of B, which also contains an additional vertex labeled (1 + 1, and an additional edge from d to d + 1. By the recurrence relation for Ya, we can see that Y0, = YG”_8 - Yaw/8T. Equiv- alently, this gives us that YG,/eT= Ya”-.. — YG,- It is easy to see that YGt/e = e,” and so this gives us that eWTz Y0”-.. — YG,- If we let G be the operator which allows the variables to commute, we get C(ewT) = XG,_e -— X0". We now proceed to calculate XGFC and X0". Since it is easy to see that G,r — e 2 K3, EH KB, ha - - ° t6 K13,r H5 K{d+1} Lu - - - t6 K3,, we may use the product rule for XG to show that XG,__e = 7r!e(,\(,r),1). To calculate X0, we again will use the product rule, together with the fact that XKB, = Z: (Ml/Wm)” Al—d-l—l where a,\ is the number of stable partitions of the vertex set of K B. of type A. Letting u = A(Bl/Bg/ - - - /B3,/ - - - /Bk), we can easily count the number of stable partitions of K B. to obtain the equation 43 X0, = uleu [(b —1)(b — 1)!m(1b—1,2) + (b +1)!m(1b+1)] . We may then use the fact that ”bub—1’2) = e_ 1. Finally let fl denote the partition obtained from r by merging the blocks of T which contain d and d + 1, allowing [3 = r if d and d + 1 are in the same block of r. Replacing a + (d + 1) by a E IId+1 in equation (4.2), we see that —1 CT: 2: lBa|_1#(T,0'). fiSaSn+(d+l) Now for any B Q [d + 1] we will consider the sets L(B)={o€lld+1:{d,d+1}§B€o, wherefigogr+(d+1)}. The nonempty L(B) partition the interval [5, it + (d + 1)] according to the content of the block containing {d, d + 1} and so we may express —1 01:; IBl—l 2 ”(7,0) o€L(B) To compute the inner sum, we need to consider the following 2 cases. Case 1) For some k > q + 2, 8,. is strictly contained in a block of 7r + (d + 1). In this case, we see that each non-empty L(B) forms a non-trivial cross-section of a 46 product of partition lattices, and so for this case 2 p.(,)7'0’ )20. a€L( B() Thus partitions in this case will not contribute to 2 c7. TEP(a) Case 2) For all k > q+2, Bk is a block of7r+ (d+1). In this case, we have q 2 0, since otherwise we must have T = 7r+ (d+ 1) which we have already considered. Then we can show __ q+1 , _ QLRM If B : Bir+(d+1) a€L(B) 0 else. (4.3) It is easy to see that if B 2 BMW“) then L(B) = {7r+(d+1)} and so that part is clear. Also, if B = B,,+(d+1)\B,- for some 2 S i S q+2, then we have L(B) ”é H1 and so ZUELU?) n(T, o) = (—1)"q!. Otherwise, L(B) again forms a non-trivial cross-section of a product of partition lattices, and again gives us no net contribution to the sum. We notice that since {d,d + 1} C B, the second case in (4.3) will only occur if d E Bj for j 79 i . So adding up all these contributions gives q+2 1 q+1 T=_1qv ________ c ( )q §m_ai m i¢j In order to compute the sum over all T E P(a), it will be convenient to consider all possible choices for the block of T containing d. So for T 3 7r + (d + 1) and 47 1_<_j£q+2,let P(C¥,j) = {(B1,Bg,... ,B() I Bl/Bg/H./B( E P(C¥),d€ Bj}. The sequence (Bl, B2, . . . aj—l Ifj=1 6, : Otj else, so m —1 P a, . = . I ( j)| (61,...,6j—1,...,6q+2) Thus we can see that 2 m—l ) (1+ 1 2: CT:( (—1)qq! Z T€P(a,j) 61,co.,6j_1,...,6q+2 :;§m_ai ,Bg) is called an ordered set partition. Also define _‘I_+_1 m To obtain the sum over all T E P(a) we need to sum over all P(a, j) for 1 S j S q+2. However, for 1 S r g m + 1, if we let k, be the number of blocks 3,, 1 S i S q + 2 which have size r, then in the sum over all P(a, j), each T E P(a) appears 113:1] kr! times. Combining all this information, we see that Z _ (-1)"q! (if m—l “if 1 _q+1 CT_Hm+1kl. 61... 6—1... 6 2 . m—(S m T€P(a) r=l 7‘ 3:1 ’ ’ J ’ 1 11+ 2.:2‘ z 1963 Hence it suffices to show that §( m _1 ) (If: 1 — q H 0 j=1 61,...,5j—1,...,(5q+2 i—2m_6' m ' 48 Using the multinomial recurrence, q+2 2( H m > , 5,,...,5,—1,...,5,+2 5,,...,5,,...,5,,2’ 1:1 we need only show that if m—l 3‘: 1 _q+1 m , 61,...,(5j—1,...,6q+2 , m—6,_ m 61,...,6j,...,6q+2 . 1:1 i=2 #1 However, we may express q+2 q+2 q+2 m , (1+2 2( m — 1 ) Z 1 _ (61,...,6j,...,6q+2)61 Z 1 j=1 61,...,(Sj—1,...,6q+2 {am—6,. j=l m izzm—bi #2 iii ( )q+2 q+2 _ 515q+2) j=1 i: 2m #1 ( m ) q+2 q+2 _ 61 6; 6q+2 1 (S — m m — 6 J i=2 ‘ j=1 #i ( m (1+2 61 .,6 ,....5 +2) 2 1 = J Q (m 6') m , m — 61 4.2 Some e-positivity Results We wish to use this result to prove some positivity theorems about YG’s expansion in the elementary symmetric function basis. If the coefficients of the elementary symmetric functions in this expansion are all non-negative, then we say that YG is e-positive. Unfortunately, even for some of the simplest graphs, Ya is usually not 6— positive. The only graphs which are obviously e—positive are the complete 49 graphs on n vertices, for which we have YKn = em], and their complements, which yield Y?" = e1/2/.../,,. Even paths, with the vertices labeled sequentially, are not e—positive, for we can compute that Yp3 = %€12/3 — %el3/2 + 'é'81/23 + %8123. However, in this example we can notice that while Y]:3 is not e-positive, if we group the terms according to their type and the size of the block which contains 3, the sum of these coefficients will be non-negative. This observation along with the proof of the previous lemma inspires us to define equivalence classes reflecting the sets P(a). That is, if the block of 0 containing i is B0,, and the block of T containing i is B”, we define a E,- T iff A(o) = /\(T) and IBM] 2 Wm]- In a similar manner, we can define eazieTiffoE,T and let (T) and em denote the equivalence classes of T and 6, respectively. We can take formal sums of these equivalence classes and write expressions such as 2 eye” E,- Z C(r)e(r) where em = 2 ca. 06114 (Tlgld ”E(T) We will refer to this equivalence relation as congruence modulo i. Using this notation, we have Yp3 E3 %e(12/3) + §e<123), since 813/2 53 61/23. We will say that a labeled graph G (and similarly YG) is (e)— positive if cm is non-negative for all (T) g IId under some labeling of G and suitably chosen congruence. We notice that the expression of YG for a labeled graph G may have all non-negative amalgamated coefficients for congruence modulo i, but not for congruence modulo j, making the definition for (e)-positivity seem to depend on the labeling of the graph. However, 50 the (e)-positivity of a graph is in fact not dependent on the labeling of the graph. If a different labeling for an (e)-positive graph is chosen, then we need only change the congruence class appropriately, to again see the (e)-positivity. This should be clear from the Relabeling Proposition. That is, if 7(G) = H for some '7 such that 7(i) = j, and Ya E,- 2mg,“ cmem, then YH E,- 2mg” c(7(7))e(,(,)). So if G is (e)-positive modulo i, then H is seen to be (e)-positive modulo j. Hence relabeling a graph does not change its (e)-positivity. We now turn our attention to showing that paths, cycles, and complete graphs with one edge deleted are all (e)-positive. We begin with a few more preliminary results about this congruence relation and how it affects our induction of e,,. We note that in the proof of Lemma 4.1.2, the roles played by the elements at and ~ d+ 1 are essentially interchangeable. That is, if we let P(a) be the set of all partitions of [d+ 1] which are less than or equal to 7r + (d+1), have blocks of size 011, . . . ,a; and for which d is in a block of size a1, and let it E Hd be the partition 71 with d replaced by d + 1, then the same proof will show that l/lerl if Pia) = {if/d}, ~ 20; -1/|B«| ifP(a)={7“r+(d)}. TEP(O) 0 else. Note that here fr + (d) is the partition obtained from fr by inserting the element d into the block of ii containing at + 1. This allows us to state a corollary in terms of the congruence relationship just defined. Corollary 4.2.1 For 7T 6 Ed, _ 1 1 €nT=d+1 38(1r/d+1) - 36(1r+(d+1)) 51 and CnTZd Sew/d) - 36(7'r+(d))’ where b: |B,,|. I The next lemma simply verifies that the induction operation respects the congru- ence relation and should be clear without proof. Lemma 4.2.2 Ify Ed T, then 67T2d+1 6,. I From this we can extend induction to congruence classes in a well-defined manner: 600‘]: Z cmem if and only if en: 2 cTeT. (Tlgnd+l TEHd+1 In order to use induction to prove the (e)-positivity of a graph G, we will usually try to delete a set of edges which will isolate either a single vertex or a complete graph from G in hope of obtaining a simpler (e)-positive graph. In order to see how this procedure will affect Ya, we use the following lemma. Lemma 4.2.3 Given a graph, G on d vertices define H = GErJKm. If Y0 = Z coed, 0611,, then YH = E Caea/d+1,d+2,...,d+m- 0611,, Proof. From previous statements it follows that Y}; = Yaqm] = Z: caeaqm] 06114 = E Coed/d+1,d+2,...,d+m- dead The result above naturally suggests that we use the notation G / vd+1 for the graph GU{ud+1}. We are now in a position to prove the (e)-positivity of paths. 52 Proposition 4.2.4 For all d 2 1, Ypd is (e)—positiue. Proof. We proceed by induction, having labeled Pd so that E(Pd) = {v1v2,v2v3,... ,ud_1vd}. If d = 1, then we have Yp, = el and the proposition is clearly true. So we assume by induction that YP. Ed 2 Gwen), (7)6111: where cm 2 0 for all (T) 6 11d. From our deletion-contraction recurrence, Corollary 4.2.1 and Lemma 4.2.3, we see that using e 2 mod“, YPd+1 : YPd/vd+l — YPdT Ed.“ 2 C(T)e(T/d+1) _ Z C(T)e(T)T (agnd (7)914 1 C(T) Ed+l Z (1 — m) C(r)e(T/d+1) + Z WBUHHID' (ecu. T (”End T Since we know that cm 2 O, and |BT| 2 1 for all T, this completes the induction step and the proof. I In the commutative context we will say that the symmetric function X0 is 6- positive if all the coefficients in the expansion of the elementary symmetric functions are non-negative. It is easy to see that we can use the (e)-positivity result for Ypd and specialize it to show the e-positivity of X pd. Corollary 4.2.5 X]:d is e—positiue. I It is then natural to ask if cycles will also be (e)—positive, for when we delete an edge of a cycle we obtain a path. While it is true that cycles will be (e)—positive, 53 a stronger relationship exists between paths and cycles. In fact, it turns out that the coefficients in the (e)-expansion of paths will appear again as the coefficients in the (e)-expansion of cycles with only a slight modification. For labeling purposes, however, we will need the following lemma which follows easily from the Relabeling Proposition. Lemma 4.2.6 Ify 6 8d fixes d, then Y,(G) Ed Ya. Proposition 4.2.7 For all d 2 2, if YR, Ed ZOMBm, then Yo“, Ed+1 ZC(T)6(T+(d+l))7 where we have labeled the graphs so E(Pd) = {v1u2, v2u3, . .. ,vd_1vd} and E(Cd+1) = {111112, 712713, - -- Wat—1714, vdvd+1a vd+lvll° Proof. We proceed by induction on d. If d = 2, then Yp2 2 cm and Y0, = em, and so the proposition holds for d = 2. For the induction step, we assume that Ya.-. Eat—1 2 Como and also that You Ed 2 C914 i=0 ”+1 where b = [3,], and Tr=7r+i/d+i+1,... ,d+m, f=rr+i+(d+m)/d+i+1,... ,d+m—1. Proof. We induct on m. If m = 1, then YG+K2 = YGwK, — YGTZ+1 . This shows that c (b — 1) c ,, YG+K2 Ed+1 2: (Lb—_ehr/d+1) + %e(r+(d+1))) a (7') which verifies the base case. To begin the induction step, we repeatedly utilize the deletion-contraction recur- rence to delete the edges ud+,vd+m+1 for 0 S i S m, and obtain YG+Km+2 Ed+m+1 YG+Km+1wvd+m+1 — mYG+Km+1Tgim+1 —YG+K,.,+1 3m“ - (4-8) Note that we are able to combine all the terms from YG+Km+1TZ::n+1 together by the previous observation, Lemma 4.3.3 since for all 1 g i g m there is a permutation satisfying the correct conditions. We now expand each of the terms in equation (4.8). For the first, using Lemma 4.2.3, m-l C 71’ (m _ 1>i . . YG+Km+lwvd+m+1 Ed+m+1 Z Z ( )(b)'+1 [(b _ m + z)e(1r1) + (2 +1)e(1r2)] a (w) i=0 ' 63 where T1=T+i/d+i+1,... ,d+m/d+m+1, T227r+i+(d+m)/d+i+1,... ,d+m—1/d+m+1, and For the second term, using Corollary 4.2.1, we have +1.. mYG+Km+1Td+m+ :d+m+l C(W)( (771),.“ b— m+t _ i+1 _ (2) m:- (b) )1+1 [ (60“) E(H)) + b + ’i + 1 (6(172) 6(7T4l) m—Z (7r) i=0 where 7r3=7r+i/d+i+1,... ,d+m+1, and 7T4=7r+i+(d+m)+(d+m+1)/d+i+1,... ,d+m—1. And finally, using Lemma 4.3.2, Cn)< YG+Km+1Td+m+ l=d+m+1 22:76—- —(_—‘) e(_7r3) 60%)) (1r) i=0 b)i+l where 7r5=7r+i+(d+m+1)/d+i+1,...,d+m. Grouping the terms appropriately and shifting indices where needed we have mas- sive cancellation, which yields m c,r (m),- (b—(m+1)+i)e,r +(z‘+1)e,r YG+Km+2 Ed+m+1 z: ( ) l (b)- ( 3) ( 5)]. (1r) i=0 1+1 This completes the induction step and the proof. I Examining this lemma, we can see that in YG+Km+1 we have the same sign on all the coefficients as we had in Y6, with the possible exception of the terms where 64 b< m — i. But it is easy to see that in this case we have e(7r+i/d+i+1,...,d+m) Ed+m e(1r+m—i—lc—1+(d+m)/d+m—i—k,...,d+m—1)- This means that in the expression for YG+ K... H as a sum over congruence classes mod- ulo d + m, we can combine the coefficients on these terms. And so upon simplification, the coefficient on e(,,+,-/d+,-+1P_.,d+m) will be: (k — m + Z)i (m -" Z — k)(m _1>m—i—k—1 ( (k),+1 + (klm—i—k ) Cm, where cm is the coefficient on em in Y0. Adding these fractions by finding a common denominator, we see that this coef- ficient is actually zero. corresponding This gives us the next result. Theorem 4.3.5 If Ya is (e)-positive, then Your", is also (e)-positive. I Notice that Proposition 4.2.4 follows easily from Theorem 4.3.5 and induction, since for paths Pm+1 = Pm + K2. As a more general result we have the following Corollary. Corollary 4.3.6 If G is a KA-chain, then YG is (e)-positive. Hence, X0 is also e-positive. I We can also describe another class of (e)-positive graphs. We define a diamond to be the indifference graph on the collection of intervals {[1, 3], [2,4]}. So a diamond consists of two K 3’s sharing a common edge. Then the following holds. Theorem 4.3.7 Let D be a diamond. IfG is (e)-positive, then so is G + D. Proof. The proof of this result is analogous to the proof for the case of G + Km. We note that if G is a graph on d vertices, then we can think of G + D as 65 G with the additional vertices ud+1,ud+2,vd+3, and new edges constructed from the intervals {[d, d + 2], [d + 1, d + 3]}. If we now delete and contract the edges ud+1vd+3 and vd+2vd+3, we see (using the apprOpriate symmetry) that YG+D E(1+3 YG+K3s{d+3} — 2YG+K3T313 - (49) By Lemma 4.3.4, if YG Ed Somem then (T) _ b—2 b—l 1 2 Yo+K3 =d+2 2%) Teen) + Ween) + gem) + mew.) . (1r) where 7T1=7T/d+1,d+2, W2=7T+(d+1)/d+2, mzuw+w+ayd+L 7r, =7r+(d+1)+(d+2), b = |B,,|. So by Corollary 4.2.1, b—2 b—l YG+K3T$igzd+3 E036”) [(W) (eh-n) — 80(1)) + ‘71));— (e(7'r2) _ 6&9) 1 2 + (72—2) (eons) — 6W) + @ (em) _ 6650)] where fr,- = n,/d+ 3, and 71;: T.- + (d+ 3). Using Lemma 4.2.3 and plugging these equations into (4.9), we have 2e,-r b—2e,,l 2b—1e,,g 2e,“ 4e“: [(