<<

f-VECTORS OF BARYCENTRIC SUBDIVISIONS

FRANCESCO BRENTI AND VOLKMAR WELKER

Abstract. For a or more generally Boolean cell complex ∆ we study the behavior of the f- and h-vector under barycentric subdivision. We show that if ∆ has a non-negative h-vector then the h-polynomial of its barycentric subdivision has only simple and real zeros. As a consequence this implies a strong version of the Charney-Davis conjecture for spheres that are the subdivision of a Boolean cell complex or the subdivision of the boundary complex of a simple . For a general (d − 1)-dimensional simplicial complex ∆ the h- polynomial of its n-th iterated subdivision shows convergent be- havior. More precisely, we show that among the zeros of this h- polynomial there is one converging to infinity and the other d − 1 converge to a set of d − 1 real numbers which only depends on d.

1. Introduction This work is concerned with the effect of barycentric subdivision on the enumerative structure of a simplicial complex. More precisely, we study the behavior of the f- and h-vector of a simplicial complex, or more generally, Boolean cell complex, under barycentric subdivisions. In our main result we show that if the h-vector of a simplicial complex or Boolean cell complex is non-negative then the h-polynomial (i.e., the generating polynomial of the h-vector) of its barycentric subdivi- sion has only real simple zeros. Moreover, if one applies barycentric subdivision iteratively then there is a limiting behavior of the zeros of the h-polynomial, with one zero going to infinity and the other zeros converging. The limit values of the zeros only depend on the of the complex. In this introductory section we first review the basic concepts around f- and h-vectors and give the basic definitions, then we outline the structure of this work.

Key words and phrases. Barycentric subdivision, f-vector, real-rootedness, unimodality. Both authors partially supported by EU Research Training Network “Algebraic Combinatorics in Europe”, grant HPRN-CT-2001-00272 and the program on “Al- gebraic Combinatorics” at the Mittag-Leffler Institut in Spring 2005. 1 2 FRANCESCO BRENTI AND VOLKMAR WELKER

Our results hold in the generality of Boolean cell complexes. Recall that if ∆ is a CW-complex then for two open cells A and A0 in ∆ 0 0 we define A ≤∆ A if A is contained in the closure of A in ∆. If ∆ is a regular CW-complex then this partial order on its cells already encodes the of ∆ up to homeomorphism. In this text we consider the empty cell as a cell of any CW-complex ∆ which then serves as the least element of the partial order on ∆. In the sequel, we identify ∆ with the partial order on its cells. In particular, we write A ∈ ∆ if we want to express that A is a cell of ∆. We call a regular CW-complex ∆ a Boolean cell complex if for each A ∈ ∆ the lower interval [∅,A] := {B ∈ ∆ | ∅ ≤∆ B ≤∆ A} is a Boolean lattice – i.e. the lattice of subsets of a set. The prime example of a Boolean cell complex is a simplicial complex. Most of the time we consider a simplicial complex as an abstract simplicial complex and therefore identify its open cells with subsets of the ground set of ∆. Adopting notions from simplicial complexes, we call a cell A ∈ ∆ a face of ∆. By dim(A) we denote the dimension of the cell A. The dimension dim(∆) of ∆ is the maximal dimension of one of its faces. def ∆ For −1 ≤ i ≤ d − 1 (= dim(∆)) we write fi for the number of i- ∆ ∆ ∆ dimensional faces of ∆. The vector f = (f−1, . . . , fd−1) is called the f-vector of ∆. Often it will be convenient to encode the f-vector of ∆ Pd ∆ d−i a Boolean cell complex as a polynomial. By f (t) = i=0 fi−1t we denote the f-polynomial of the (d − 1)-dimensional Boolean cell complex ∆. Information equivalent to the one carried by the f-vector ∆ ∆ ∆ is encoded in the h-vector. It is defined as h = (h0 , . . . , hd ) where ∆ Pd ∆ d−i ∆ h (t) = i=0 hi t is the expansion of f (t − 1) in terms of t-powers. The barycentric subdivision of a Boolean cell complex ∆ is the sim- plicial complex sd(∆), that as an abstract simplicial complex is defined on the ground set ∆ \ {∅} – the set of non-empty faces of ∆ – with i-faces the strictly increasing flags A0 <∆ ··· <∆ Ai of faces in ∆ \ {∅}. It is well known that ∆ and sd(∆) are homeomorphic. Thus ∆ and sd(∆) define cellulations and triangulations of the same space. Throughout the paper we will use [n] to denote the set {1, . . . , n} for a natural number n ≥ 0.

2. The h-Vector of a Barycentric Subdivision For the sake of completeness we sketch the proof of the following well known result. f-VECTORS OF BARYCENTRIC SUBDIVISIONS 3

Lemma 2.1. Let ∆ be a (d−1)-dimensional Boolean cell complex then

d sd(∆) X ∆ fj = fi−1 (j + 1)! S(i, j + 1), i=0 where S(j, i) is the Stirling number of the second kind.

Proof. By definition a j-face of sd(∆) is a flag A0 <∆ ··· <∆ Aj of faces in ∆ \ {∅}. Let us fix j and Aj. Since ∆ is a Boolean cell complex, the interval [∅,Aj] is the Boolean lattice of subsets of a (dim(Aj) + 1)- element set. Thus we may identify each of the cells A0,...,Aj with a subset of this set. Using this identification, the map sending A0 <∆ ··· <∆ Aj to (A0,A1 \ A0, ··· ,Aj \ Aj−1) is a bijection between j-faces of sd(∆) with Aj as its largest element with respect to ≤∆ and ordered partitions of Aj into j+1 non-empty blocks. The latter are enumerated by (j + 1)!S(i, j + 1), where i = dim(Aj) + 1. Now summing over all Aj ∆ of fixed dimension i − 1 we get fi−1(j + 1)!S(i, j + 1). Summing over all i then yields the desired formula. 

Less obvious is the representation of the h-vector of a barycentric subdivision in terms of the h-vector of the original complex. In the formulation of the proposition we write

D(σ) = i ∈ [d − 1] | σ(i) > σ(i + 1) for the descent set of the permutation σ, des(σ) := #D(σ) for its number of descents and Sd for the symmetric group on [d]. For 1 ≤ d, 1 ≤ j ≤ d and 0 ≤ i ≤ d − 1 , we denote by A(d, i, j) the number of permutations σ ∈ Sd such that σ(1) = j and des(σ) = i. We define A(d, i, j) for all d ≥ 1 and all integers i and j. In particular, A(d, i, j) = 0 if i ≤ −1 or i ≥ d. A formula similar to the one in the following theorem appears in a slightly different context in [13, Chapter 3, Ex. 71] and [12, Theorem 8.3]. We were not able to deduce the following result from these formulas directly.

Theorem 2.2. Let ∆ be a (d − 1)-dimensional Boolean cell complex. Then d sd(∆) X ∆ hj = A(d + 1, j, r + 1) hr r=0 for 0 ≤ j ≤ d. 4 FRANCESCO BRENTI AND VOLKMAR WELKER

Proof. For all 0 ≤ j ≤ d we have that j sd(∆) X d−i sd(∆) h = (−1)j−if j j−i i−1 i=0 j d X d−i X = (−1)j−i f ∆ S(k, i) i! j−i k−1 i=0 k=0 j d k X X d−i X d−r  = (−1)j−i S(k, i) i! h∆ j−i d−k r i=0 k=0 r=0 d d j ! X X X d−i d−r  = (−1)j−i S(k, i) i! h∆. j−i d−k r r=0 k=0 i=0

Now notice that if S = {s1, . . . , sk} ⊆ [d] and s1 < . . . < sk then

n D(σ) ⊆ S, o  sk   d−r  # σ ∈ Sd+1 = . σ(d + 1) = d + 1 − r sk−sk−1,...,s2−s1,s1 d−sk So, X n D(σ) ⊆ S, o # σ ∈ S d+1 σ(d + 1) = d + 1 − r {S⊆[d], #S=k}

X  sk   d−r  = sk−sk−1,...,s2−s1,s1 d−sk 1≤s1<...

j X d−i X n D(σ) ⊆ S, o = (−1)j−i # σ ∈ S j−i d+1 σ(d + 1) = d + 1 − r i=0 {S⊆[d], #S=i}

X d−#S  n D(σ) ⊆ S, o = (−1)j−#S # σ ∈ S j−#S d+1 σ(d + 1) = d + 1 − r {S⊆[d], #S≤j}

X d−#S  X n D(σ) = T, o = (−1)j−#S # σ ∈ S j−#S d+1 σ(d + 1) = d + 1 − r {S⊆[d], #S≤j} T ⊆S f-VECTORS OF BARYCENTRIC SUBDIVISIONS 5

X n D(σ) = T, o X d−#S  = # σ ∈ S (−1)j−#S d+1 σ(d + 1) = d + 1 − r j−#S {T ⊆[d], #T ≤j} {S⊇T, #S≤j}

j X n D(σ) = T, o X d−i d−#T  = # σ ∈ S (−1)j−i . d+1 σ(d + 1) = d + 1 − r j−i i−#T {T ⊆[d], #T ≤j} i=#T But j j X d−i d−#T  d−#T  X j−#T  (−1)j−i = (−1)j−i j−i i−#T j−#T i−#T i=#T i=#T

= δj,#T Hence d j X X d−i d−r  (−1)j−i S(k, i)i! j−i d−k k=0 i=0 X n D(σ) = T, o = # σ ∈ S d+1 σ(d + 1) = d + 1 − r {T ⊆[d], #T =j} n des(σ) = j, o = # σ ∈ S , d+1 σ(d + 1) = d + 1 − r and the result follows.  Corollary 2.3. Let ∆ be a (d − 1)-dimensional Boolean cell complex ∆ such that hi ≥ 0 for 1 ≤ i ≤ d. Then for 0 ≤ i ≤ d sd(∆) ∆ hi ≥ hi . sd(∆) Furthermore, hj ≥ 1 for 0 ≤ j ≤ d − 1.

sd(∆) Proof. We have from Theorem 2.2 and our hypotheses that hj ≥ ∆ A(d + 1, j, j + 1) hj so the result follows since A(d + 1, j, j + 1) ≥ 1 for sd(∆) all 0 ≤ j ≤ d. Furthermore, if hj = 0 for some j < d then, since A(d + 1, j, r + 1) > 0 for all r = 0, . . . , d, we conclude from Theorem ∆ 2.2 that h0 = 0, which is a contradiction.  Note that for any (d − 1)-dimensional Boolean cell complex ∆ we ∆ ∆ d have h0 = 1 and hd = (−1) ·χe(∆) for the reduced Euler-characteristic sd(∆) χe(∆) of ∆. In particular, h0 = 1 and by the topological invariance sd(∆) ∆ of the Euler-characteristic also hd = hd . Of course both identities are also implicit in Theorem 2.2. 6 FRANCESCO BRENTI AND VOLKMAR WELKER

sd(∆) ∆ Example 2.4. It is not true that hi ≥ hi for any Boolean cell complex ∆ and any i. For example let ∆ be the disjoint union of a 2- and a with f-vector f∆ = (1, 6, 6, 1) and h-vector h∆ = (1, 3, −3, 0). Then fsd(∆) = (1, 13, 18, 6) and hsd(∆) = (1, 10, −5, 0).

Before we proceed with the next corollary we summarize some basic facts about the numbers A(d, i, j).

Lemma 2.5.

(i)

j−1 d X X A(d, i, j) = A(d − 1, i − 1, l) + A(d − 1, i, l − 1), l=1 l=j+1

for 2 ≤ d, 1 ≤ j ≤ d, and 0 ≤ i ≤ d − 1. (ii)

A(d, i, j) = A(d, d − 1 − i, d + 1 − j)

for 1 ≤ d, 1 ≤ j ≤ d and 0 ≤ i ≤ d − 1.

Proof. For (i) the recursion formula just reflects the enumeration of the permutations in Sd with σ(1) = j and des(σ) = i by l = σ(2). For (ii) the assertion reflects the fact that the map σ(1) ··· σ(d) ↔ d + 1 − σ(1) ··· d + 1 − σ(d) is a bijection between the sets enumerated by the two numbers. 

We call the h-vector h∆ (resp. h-polynomial h∆(t)) of a Boolean ∆ ∆ ∆ cell complex ∆ reciprocal if hi = hd−i for 0 ≤ i ≤ d (resp. h (t) = d ∆ 1 ∆ ∆ t h ( t )). Clearly, h is reciprocal if and only if h (t) is. Either con- dition is equivalent to the h-vector satisfying the Dehn-Sommerville relations (see [1, Theorem 5.4.2]).

Corollary 2.6. Let ∆ be a (d − 1)-dimensional Boolean cell complex with reciprocal h-vector then sd(∆) also has reciprocal h-vector. f-VECTORS OF BARYCENTRIC SUBDIVISIONS 7

Proof. By Lemma 2.5 and Theorem 2.2 we get d sd(∆) X ∆ hj = A(d + 1, j, i + 1)hi . i=0 d X ∆ = A(d + 1, d − j, d − i + 1)hd−i i=0 d X ∆ = A(d + 1, d − j, i + 1)hi i=0 sd(∆) = hd−j

 3. Main Result For the formulation of the following result we recall that a sequence 2 (a0, . . . , ad) of real numbers is called log-concave if ai ≥ ai−1ai+1 for 1 ≤ i ≤ d − 1. We say that (a0, . . . , ad) has internal zeros if there are 0 ≤ i+1 < j ≤ d such that ai, aj 6= 0 and ai+1 = ai+2 = ··· = aj−1 = 0. It is a well known and easy to prove fact that if (a0, . . . , ad) is a log- concave sequence of non-negative real numbers and has no internal zeros then it is unimodal; that is, there is a 0 ≤ j ≤ d such that a0 ≤ · · · ≤ aj ≥ · · · ≥ ad. By another well known fact (see, e.g., [3, d §2.5]) it follows that if the polynomial a0 + a1t + ··· + adt has only real zeros then (a0, . . . , ad) is log-concave. Theorem 3.1. Let ∆ be a (d − 1)-dimensional Boolean cell complex ∆ such that hi ≥ 0 for 0 ≤ i ≤ d. Then d sd(∆) X sd(∆) d−i h (t) = hi t i=0 sd(∆) sd(∆) sd(∆) has only simple and real zeros. In particular, h = (h0 , . . . , hd ) is a log-concave and unimodal sequence. Before we can prove Theorem 3.1 we need a few definitions and a lemma about permutation statistics of Coxeter groups of types Ad and Bd. The Coxeter group of type Ad is the symmetric group Sd+1 and we have defined the respective statistics already in Section 2. The Coxeter group of type Bd is the group of all bijections σ of {±1,..., ±d} in itself such that σ(−i) = −σ(i) for 1 ≤ i ≤ d. An element of Bd is called a signed permutation. For σ ∈ Bd one conveniently sets σ(0) := 0. 8 FRANCESCO BRENTI AND VOLKMAR WELKER

Then i ∈ [0, d − 1] is called a descent of σ ∈ Bd if σ(i) > σ(i + 1). In particular, 0 is a descent if and only if σ(1) < 0. By desB(σ) we denote the number of descents of the signed permutation σ. By N(σ) we denote the number of i ∈ [d] such that σ(i) < 0. Let A(d + 1, i, j) be the set of pairs (S, σ) of (j − 1)-subsets S of [d] and σ ∈ Sd+1 such that σ has i descents and σ(1) = j. Lemma 3.2. Let d ≥ 1, j ∈ [d + 1] and 0 ≤ i ≤ d. There is a bijection between A(d + 1, i, j) and the set of signed permutations τ ∈ Bd with i descents and N(τ) = j − 1.

Proof. For (S, σ) ∈ A(d + 1, i, j) we define τ(S,σ) ∈ Bd as follows. Let S = {s1 > . . . > sj−1} and [d] \ S = {sj < . . . < sd}. Then define τ(S,σ) ∈ Bd by  −sσ(r+1) if σ(r + 1) < j τ(S,σ)(r) = , sσ(r+1)−1 if σ(r + 1) > j

(and τ(S,σ)(0) = 0). We show that for 0 ≤ r ≤ d:

(3.1) τ(S,σ)(r) > τ(S,σ)(r + 1) ⇔ σ(r + 1) > σ(r + 2). This is clear for r = 0, so assume 1 ≤ r ≤ d. Let us distinguish the four possible cases:

. σ(r + 1), σ(r + 2) < j. In this case τ(S,σ)(r) > τ(S,σ)(r + 1) ⇔ sσ(r+1) < sσ(r+2), which is equivalent to σ(r + 1) > σ(r + 2). . σ(r + 1), σ(r + 2) > j. In this case τ(S,σ)(r) > τ(S,σ)(r + 1) ⇔ sσ(r+1)−1 > sσ(r+2)−1, which is equivalent to σ(r +1) > σ(r +2). . σ(r+1) < j, σ(r+2) > j. In this case clearly σ(r+1) < σ(r+2) and τ(S,σ)(r) < 0 < τ(S,σ)(r + 1). . σ(r+1) > j, σ(r+2) < j. In this case clearly σ(r+1) > σ(r+2) and τ(S,σ)(r) > 0 > τ(S,σ)(r + 1).

Thus desB(τ(S,σ)) = des(σ). Clearly N(τ(S,σ)) = j − 1. One easily checks that the map (S, σ) 7→ τ(S,σ) is injective and surjective from A(d + 1, i, j) to the set of signed permutations in Bd with i descents and N(τ) = j − 1.  Proof of Theorem 3.1. By Theorem 2.2 the following identity holds.

d d X sd(∆) i X i hi t d A(d + 1, i, j + 1)t i=0 X i=0 (3.2) = h∆ · (1 − t)d+1 j (1 − t)d+1 j=0 f-VECTORS OF BARYCENTRIC SUBDIVISIONS 9

If B(d, i, j) is the number of signed permutations in Bd such that desB(σ) = i and N(σ) = j then it follows from [4, Corollary 3.9, Eq.(25)] that:

d X B(d, i, j)ti   i=0 X d (3.3) = rj(r + 1)d−jtr (1 − t)d+1 j r≥0

Using Lemma 3.2 it follows from (3.3) that

d X A(d + 1, i, j + 1)ti i=0 X (3.4) = rj(r + 1)d−jtr. (1 − t)d+1 r≥0

Therefore, by (3.2)

d X sd(∆) i hi t d i=0 X X (3.5) = ( h∆rj(r + 1)d−j) tr. (1 − t)d+1 j r≥0 j=0

By [5, Theorem 4.2] it follows that if we write

d d X X r (3.6) h∆rj(r + 1)d−j = k∆ j j j j=0 j=0

d X ∆ j then the polynomial kj r has non-negative coefficients and only j=0 simple and real zeros. Using the Binomial Theorem we obtain, by (3.5) and (3.6) 10 FRANCESCO BRENTI AND VOLKMAR WELKER

d X sd(∆) h ti i d   i=0 X X r = ( k∆ ) tr (1 − t)d+1 j j r≥0 j=0 d X X r = k∆ tr j j j=0 r≥0 d X tj = k∆ j (1 − t)j+1 j=0 d X ∆ j d−j kj t (1 − t) j=0 = . (1 − t)d+1 This finishes the proof of the main assertion. Since polynomials with only real zeros have a log-concave coefficient sequence and since by Corollary 2.3 hsd(∆) is non-negative and has no sd(∆) internal zeros the unimodality of h follows.  d+1 Remark 3.3. Let (h0, . . . , hd) ∈ R such that hi ≥ 0 for 0 ≤ i ≤ d. d 0 X If we set hj = A(d + 1, j, i + 1)hi for 0 ≤ j ≤ d then the polynomial i=0 0 0 0 d hd + hd−1t + ··· h0t has only real and simple zeros. ∆ Proof. The proof of Theorem 3.1 never uses the fact that hi is the ∆ h-vector of a Boolean cell complex. The proof only relies on hi ≥ 0 ∆ sd(∆) and the formula transforming hi into hi . Therefore, the remark follows from the proof of Theorem 3.1.  Note that since sd(∆) is a flag simplicial complex, that is, all mini- mal non-faces are of size 2, it is known that hsd(∆)(t) has at least one real zero. This fact for flag simplicial complexes appears in [11] or [7]. As mentioned in [11] this is actually a well known fact in commuta- tive algebra, where it appears as a theorem saying that the numerator polynomial of the Hilbert-series of a standard graded Koszul k-algebra has at least one real zero. The following example shows that the as- sumptions of the theorem are actually needed. Example 3.4. Let ∆ be the disjoint union of a 2-simplex and a 1- simplex. Then h∆ = (1, 2, −3, 1) and hsd(∆) = (1, 7, −3, 1). In particu- lar, hsd(∆)(t) has only one real zero. f-VECTORS OF BARYCENTRIC SUBDIVISIONS 11

Corollary 3.5. Let ∆ be a (d−1)-dimensional simplicial complex that is Cohen-Macaulay over some field k. Then d sd(∆) X sd(∆) d−i h (t) = hi t i=0

sd(∆) sd(∆) sd(∆) has only simple and real zeros. In particular, h = (h0 , . . . , hd ) is a log-concave and unimodal sequence. In particular, all conclusions hold for Gorenstein, Gorenstein∗ simplicial complexes, for simplicial spheres and boundary complexes of simplicial . Proof. All mentioned simplicial complex are Cohen-Macaulay over some field k. Therefore, it suffices to recall that Cohen-Macaulay simplicial complexes have a non-negative h-vector without internal zeros. This well known fact follows since if ∆ is Cohen-Macaulay the sequence ∆ ∆ ∆ (h0 , . . . , hd ) is the Hilbert-function hi = dimk Ai of a 0-dimensional standard graded k-algebra A = A0 ⊕ · · · ⊕ Ad (see [1, Chapter 5] for details). Being the dimension of a k-vector space clearly implies that ∆ hi ≥ 0. Now the assertion immediately follows from Theorem 3.1.  The preceding corollary assures that there is an abundance of in- teresting simplicial complexes satisfying the assumption of Theorem 3.1. As another corollary we can treat barycentric subdivisions of simple polytopes. Here we leave the setting of Boolean cell complexes. Note, that for a (general) polytope P the barycentric subdivision sd(∆) of its boundary complex ∆ is the simplicial complex of all linearly ordered subsets of the proper part of its face lattice L(P ). We refer to Ziegler’s book [15] for the definition of a simple polytopes, its f-vector, polytope duality, and the face lattice L(P ) of a polytope P . Corollary 3.6. Let ∆ be the boundary complex of a simple d-polytope. Then d sd(∆) X sd(∆) d−i h (t) = hi t i=0 sd(∆) sd(∆) sd(∆) has only simple and real zeros. In particular, h = (h0 , . . . , hd ) is a log-concave and unimodal sequence. Proof. Let ∆ be the boundary complex of the simple polytope P . Since P is a simple polytope it is the dual of a simplicial polytope Q. The face lattices L(P ) of P and L(Q) of Q are order dual; that is, they are isomorphic with the order relation reversed. In particular, there is a bijection between the (i + 1) element chains in the proper parts 12 FRANCESCO BRENTI AND VOLKMAR WELKER of L(P ) and L(Q). In particular, if ∆Q is the boundary complex of Q, then sd(∆) and sd(∆Q) are isomorphic simplicial complexes. Thus f sd(∆)(t) = f sd(∆Q)(t). Now the assertion follows from Corollary 3.5.  At this point we would like to add examples that show that the additional generality of Boolean cell complexes enriches the class of complexes for which the theorem holds by interesting examples.

Example 3.7. An injective word over the alphabet [n] is a word i1 ··· ij for which ir 6= is for all 1 ≤ r < s ≤ j. We call {i1, . . . , ij} the con- tent of the injective word i1 ··· ij. We order injective words by saying 0 0 i1 ··· il  ii . . . ij if there is a sequence of indices 1 ≤ s1 < ··· < sl ≤ j 0 0 such that i1 ··· il = is1 ··· isl . For a simplicial complex ∆ over ground set [n] we denote by Γ(∆) the partially ordered set of all injective words with content in ∆. It is easily seen that Γ(∆) is the face poset of a Boolean cell complex. In [9], where the complexes Γ(∆) are defined, it is shown that all Γ(∆) have non-negative h-vector, thus providing a big class of Boolean cell complexes satisfying the assumptions of The- orem 3.1. It should be mentioned that if ∆ is the full simplex then the Boolean cell complex Γ(∆) is a well studied object (see the references in [9]). At this point we are in position to relate our results to the Charney- Davis conjecture [6]. The original conjecture states that if ∆ is a flag (d − 1)-dimensional Gorenstein∗ simplicial complex and d is even then b d c ∆ ∆ d ∆ (−1) 2 (h0 −h1 +···+(−1) hd ) ≥ 0. Alternatively, the assertion can b d c ∆ ∗ be phrased as (−1) 2 h (−1) ≥ 0. Since d odd implies for Gorenstein ∆ ∆ ∆ simplicial complexes by hi = hd−i that h (−1) = 0 we will drop the assumption d even and say that a simplicial complex ∆ with reciprocal b d c ∆ h-vector satisfies the Charney-Davis Conjecture if (−1) 2 h (−1) ≥ 0. The latter number is also called Charney-Davis quantity of ∆. Barycen- tric subdivisions already appear in one of the first instances where the Charney-Davis conjecture was verified. Stanley [14] shows that if ∆ is the barycentric subdivision of the boundary complex of a (not neces- sarily simplicial) polytope then the Charney-Davis conjecture holds for ∆. More recently, the work of Karu [10] implies the conjecture for order complexes of Gorenstein∗ posets. As a corollary of Theorem 3.5 we ob- tain the following result which strengthens Karu’s results in the case of subdivisions of Boolean cell complexes whose face poset is Gorenstein∗ and subdivisions of boundary complexes of simple polytopes. f-VECTORS OF BARYCENTRIC SUBDIVISIONS 13

Corollary 3.8. The Charney-Davis conjecture holds for: (i) Subdivisions of (d − 1)-dimensional Boolean cell complexes ∆ ∆ ∆ ∆ for which hi ≥ 0 and hi = hd−i for 0 ≤ i ≤ d. (ii) Subdivisions of boundary complexes of simple d-polytopes. If d is even then the assumptions imply that the Charney-Davis quantity of sd(∆) is strictly positive. Proof. The case (ii) follows from (i) by the arguments used in the proof of Corollary 3.6. Therefore we can restrict our attention to (i). First let us verify the Charney-Davis conjecture for sd(∆) for (i). By Corollary 2.6 the assumptions imply that sd(∆) has a reciprocal h-polynomial hsd(∆)(t). By Theorem 3.1 we also know that hsd(∆) has only real zeros. Since the coefficients of hsd(∆)(t) are non-negative and sd(∆) sd(∆) h0 = 1 it follows that the zeros of h (t) are all strictly negative. Being a reciprocal polynomial with non-zero constant coefficient also 1 implies that if α is a zero then α is a zero. Thus the zeros are either 1 −1 or come in pairs α < −1 < α < 0. If −1 is a zero then the assertion follows immediately. If −1 is not a zero then d must be even and the d −1 sd(∆) 2 factors (−1 − α)(−1 − α ) of h (−1) have all negative sign. Thus sd(∆) d h (−1) has sign (−1) 2 , which implies Charney-Davis. Let again d be even. Since there is an even number of zeros of hsd(∆)(t) which by Theorem 3.1 are simple and real it follows that −1 1 can only be a simple zero. But since all zeros come in pairs α, α this is impossible. Thus hsd(∆)(−1) 6= 0 and it follows from the first part of the proof that the Charney-Davis quantity is strictly positive.  Note that the arguments showing that a monic real reciprocal h- polynomial with only real zeros and non-negative coefficients satisfies the Charney-Davis conjecture can also be found in [11]. Remark 3.9. It had been asked in [6] whether for flag Gorenstein∗ simplicial complexes ∆ the h-polynomial h∆(t) has only real zeros. This conjecture has been shown to fail by Gal [7] for dim ∆ ≥ 5 even for the class of flag boundary complexes of simplicial polytopes. For d ≤ 4 an even more general assertion is true. Namely, in [11, Corollary 4.14] it is shown that if A is a finitely generated Koszul Gorenstein k-algebra A hA(t) such that the numerator h (t) of its Hilbert-series Hilb(A, t) = (1−t)d has degree ≤ 3 then hA(t) has only real zeros. Question 3.10. Is the conclusion of Theorem 3.1 true for barycen- tric subdivisions sd(∆) of boundary complexes ∆ of (not necessarily simplicial or simple) polytopes ? More precisely, if ∆ is the boundary 14 FRANCESCO BRENTI AND VOLKMAR WELKER complex of a (not necessarily simplicial or simple) polytope is it true that fsd(∆)(t) has only real and simple roots ? In the remaining section we study the effect of barycentric subdivi- sion on the Stanley-Reisner ring of a simplicial complex. Recall, that for a simplicial complex ∆ on ground set Ω the Stanley-Reisner ideal Q I∆ is the ideal generated within S = k[xω | ω ∈ Ω] by xA = ω∈A xω for A 6∈ ∆. In particular, I∆ is minimally generated by xA for A a minimal non-face of ∆. By k[∆] we denote the quotient S/I∆. Since I∆ is a homogeneous ideal, the quotient k[∆] is a standard graded k- ∼ L ∆ P ∆ algebra k[∆] = i≥0 Ai . By Hilb(k[∆], t) = i≥0 dimk Ai we denote the Hilbert-series of k[∆]. In the process of passing from ∆ to sd(∆) the ring theoretic properties change as follows: . The Krull dim(k[∆]) and dim(k[sd(∆)]) coincide. . The Hilbert-series of k[∆] and k[sd(∆)] are given as: h∆ + ··· + h tt Hilb(k[∆], t) = 0 d (1 − t)d and hsd(∆) + ··· + hsd(∆)td Hilb(k[sd(∆)], t) = 0 d . (1 − t)d . k[sd(∆)] is Koszul. Question 3.11. Is there a version of the main result for graded rings ? What is the “barycentric subdivision” of a standard graded k-algebra ? For that we have in mind an operator sd(·) that transforms a standard graded algebra A into a standard graded algebra sd(A) such that: • sd(k[∆]) = k[sd(∆)] for simplicial complexes ∆. • dim(A) = dim(sd(A)). • The numerator polynomials of the Hilbert-series of A and sd(A) transform ”similarly” as the ones of k[∆] and k[sd(∆)]. • sd(A) is Koszul.

4. Limiting Behavior For a number n ≥ 1 and a Boolean cell complex ∆ we denote by sdn(∆) the result of an n-fold application of the subdivision operator sd to ∆. If we fix a geometric realization |∆| of ∆ in some real vector space then one may regard the geometric realizations | sdn(∆)| of the complexes sdn(∆) as triangulations of the same space with simplices of volume converging to 0. f-VECTORS OF BARYCENTRIC SUBDIVISIONS 15

Example 4.1. If dim(∆) = 0 then sd(∆) = ∆ which clearly implies that the h-vector remains invariant under barycentric subdivision. n n sd (∆) ∆ sd (∆) ∆ If dim(∆) > 0 then still h0 = h0 = 1 and hd = hd . But experimentally one observes that all the other coefficients of hsdn(∆)(t) grow rapidly in absolute value. The following result shows that this growth is reflected in the behavior of a single root of hsdn(∆)(t). More- over, all other roots converge to real numbers, which only depend on the dimension of the simplicial complex. Theorem 4.2. For a number d ≥ 2 there are real negative numbers α1, . . . , αd−2 such that for every Boolean cell complex ∆ of dimension (n) d − 1 there are sequences (βi )n≥1, 1 ≤ i ≤ d of complex numbers such that (n) (i) βi , 1 ≤ i ≤ d, are real for n sufficiently large. (n) (ii) lim βi = αi for 1 ≤ i ≤ d − 2. n→∞ (iii) lim β(n) = 0. n→∞ d−1 (iv) lim β(n) = −∞. n→∞ d (v) For n ≥ 1 d Y (n) sdn(∆) (t − βi ) = h (t). i=1 Before we can come to a proof of Theorem 4.2 we need to analyze the transformation from h∆ to hsd(∆) more closely. As a byproduct we will prove the following proposition. Proposition 4.3. Let ∆ be a (d−1)-dimensional Boolean cell complex for some d ≥ 1. Then there is a number N(∆) such that for n ≥ N(∆) we have sdn(∆) hi > 0 for 0 ≤ i ≤ d − 1. In particular, any (d − 1)-dimensional triangulable topological space X has a triangulation ∆ such that ∆ hi > 0 for 0 ≤ i ≤ d − 1. First we set up the matrices of the transformations that send f- and h-vectors of simplicial complexes to those of their barycentric (d−1) subdivision. Let Fd−1 = (fij )0≤i,j≤d be the matrix with entries (d−1) (d−1) fij := (i + 1)!S(j, i + 1). Let Hd−1 = (hij )0≤i,j≤d be the ma- (d−1) trix with entries hij := A(d + 1, i, j + 1). Now Lemma 2.1 and Theorem 2.2 can be rephrased as follows: 16 FRANCESCO BRENTI AND VOLKMAR WELKER

sd(∆) ∆ f = Fd−1f ,

sd(∆) ∆ h = Hd−1h . In the sequel we summarize some simple results on the matrices Hd−1 and Fd−1 that will be used in the proof of the main theorem of this section. Lemma 4.4. Let d ≥ 1. Then:

(i) The matrices Fd−1 and Hd−1 are similar. (ii) The matrices Fd−1 and Hd−1 are diagonizable with eigenvalue 1 of multiplicity 2 and eigenvalues 2!, 3!, . . . , d! of multiplicities 1. In particular, there is a basis of rational eigenvectors. Proof. Since the transformation from f∆ to h∆ is an invertible linear transformation the matrices Fd−1 and Hd−1 are similar by Lemma 2.1 and Theorem 2.2. For the second assertion consider Fd−1. Clearly, Fd−1 is an upper triangular matrix with diagonal 1, 1, 2!, 3!, . . . , d!. Thus it remains to be shown that the eigenspace for eigenvalue 1 is of dimension 2. But this follows since the first and the second unit vectors are eigenvectors for the eigenvalue 1.  The following lemmas give the crucial information on the eigenvec- tors of Hd−1. (1) (2) Lemma 4.5. Let d ≥ 2 and v1 (d), v1 (d), v2(d), . . . , vd(d) be a basis of d (1) (2) R consisting of eigenvectors of the matrix Fd−1, where v1 (d), v1 (d) are eigenvectors for the eigenvalue 1 and vi(d) is an eigenvector for the (1) (1) eigenvalue i!, 2 ≤ i ≤ d. Then the vectors v1 (d + 1) := (v1 (d), 0), (2) (2) v1 (d + 1) := (v1 (d), 0) and vi(d + 1) := (vi(d), 0) are eigenvectors of Fd for the eigenvalues 1, 1, 2!, . . . , d!.

Proof. The assertion follows from the fact that Fd−1 and Fd are upper triangular and that if one deletes the (d + 1)-st column and row from Fd then one obtains Fd−1.  (1) (2) Lemma 4.6. Let d ≥ 2 and let w1 , w1 , w2, w3, . . . , wd be a basis of (1) (2) eigenvectors of the matrix Hd−1, where w1 , w1 are eigenvectors for the eigenvalue 1 and wi is an eigenvector for the eigenvalue i!, 2 ≤ i ≤ d. (i) Let ∆ be a (d − 1)-dimensional Boolean cell complex. If we ∆ (1) (1) (2) (2) Pd expand h = a1 w1 + a1 w1 + i=2 aiwi in terms of the eigenvectors, then ad 6= 0. (ii) The first and the last coordinate entry in w2, . . . , wd is zero. f-VECTORS OF BARYCENTRIC SUBDIVISIONS 17

(1) (2) (iii) The vectors w1 and w1 can be chosen such that

(1) (2) w1 = (1, i1, . . . , id−1, 0) and w1 = (0, j1, . . . , jd−1, 1).

(iv) The vector wd can be chosen such that wd = (0, b1, . . . , bd−1, 0) for strictly positive rational numbers bi, 1 ≤ i ≤ d − 1. Proof. (i) From Lemma 4.5 we deduce that if one expands the f- vector of ∆ in terms of a basis of eigenvectors of Fd−1 then by ∆ fd−1 6= 0 the coefficient of the eigenvector for the eigenvalue d! is non-zero. Since Fd−1 and Hd−1 are similar the assertion follows. (ii) From the definition of Hd−1 it is clear that the first row is the first unit vector and the last row the (d + 1)-st unit vector. 0 Thus if we denote by Hd−1 the matrix which results from Hd−1 by deleting the first and (d + 1)-st rows and columns, then the 2 characteristic polynomial of Hd−1 splits into (1 − t) times the 0 0 characteristic polynomial of Hd−1. In particular, Hd−1 has eigen- 0 0 values 2!, . . . , d! and therefore is diagonizable. Let w2, . . . , wd 0 be eigenvectors of Hd−1 for the eigenvalues 2!, . . . , d!. Again by the fact that the first and last row of Hd−1 are the first and 0 0 the (d + 1)-st unit vector it follows that (0, w2, 0),..., (0, wd, 0) are eigenvectors of Hd−1 for the eigenvalues 2!, . . . , d!. Since all eigenspaces are of dimension 1 the assertion follows. (iii) Follows immediately from (ii). 0 0 (iv) Let Hd−1 be as in the proof of (ii). The entries of Hd−1 are A(d + 1, j, i) for 1 ≤ j ≤ d − 1 and 2 ≤ i ≤ d. It is easily seen that these numbers are strictly positive. Therefore, by the Perron-Frobenius Theorem [8, Theorem 8.2.11] it follows that 0 there is an eigenvector wd for the eigenvalue d! with strictly positive entries. Now the assertion follows by (ii).  Now we can prove Proposition 4.3.

Proof of Proposition 4.3. For d − 1 = 0 there is nothing to prove. As- (1) (2) sume d ≥ 2 and let w1 , w1 , w2, w3, . . . , wd be a basis of eigenvectors (1) (2) of the matrix Hd−1, where w1 , w1 are eigenvectors for the eigenvalue 1 and wi is an eigenvector for the eigenvalue i!, 2 ≤ i ≤ d. By Lemma 4.6 (iv) we may assume that wd = (0, b1, . . . .bd−1, 0) for strictly positive bi, 1 ≤ i ≤ d − 1. Let

∆ (1) (1) (2) (2) h = a1 w1 + a1 w1 + a2w2 + ··· + adwd 18 FRANCESCO BRENTI AND VOLKMAR WELKER

∆ be the expansion of h as a linear combination of eigenvectors of Hd−1. By Lemma 4.6 we have ad 6= 0. Clearly, for n ≥ 1, sdn(∆) (1) (1) (2) (2) n n h = a1 w1 + a1 w1 + 2 a2w2 + ··· + (d!) adwd. Hence hsdn(∆) lim = adwd. n→∞ d!n sdn(∆) In particular, for sufficiently large n the signs of adbi and hi , 1 ≤ i ≤ d − 1, all coincide. Since for any (d − 1)-dimensional Boolean cell complex the h1-entry in its h-vector equals the number of vertices minus d, it follows that sdn(∆) sdn(∆) h1 = f0 − d. sdn(∆) sdn(∆) Thus from limn→∞ f0 = +∞ we can deduce that h1 is strictly positive for sufficiently large n. From this we deduce that adb1 must be sdn(∆) strictly positive and therefore also ad. But then hi ,1 ≤ i ≤ d − 1, is strictly positive for sufficiently large n. Now let X be a (d − 1)-dimensional triangulable topological space and ∆ triangulation of X. It is well known that ∆ and sd(∆) are home- omorphic. It follows from the first part that for sufficiently large n the n sdn(∆) simplicial complex sd (∆) is a triangulation of X such that hi > 0, 1 ≤ i ≤ d − 1.  We mention a fact derived in the proof of Proposition 4.3 that will become important later.

Remark 4.7. The vector wd and its coefficient ad from Lemma 4.6 (i) can be chosen to be simultaneously non-negative.

Lemma 4.8. Let wd = (0, b1, . . . , bd−1, 0) be an eigenvector of Hd−1 for the eigenvalue d! such that bi > 0 for all 1 ≤ i ≤ d − 1. Then the d−1 polynomial bd−1t + ··· b1t has only real and simple zeroes.

Proof. The vector wd satisfies the assumptions of Remark 3.3. There- d fore, for (h0, . . . , hd) = Hd−1wd it follows that hd + hd−1t + ··· + h0t has only real and simple zeros. Since wd is an eigenvector of Hd−1 for the eigenvalue d! it follows that h0 = hd = 0 and hi = d!bi. Thus, d−1 d d−1 d!(bd−1t + ··· + b1t ) = hd + hd−1t + ··· + h0t and bd−1t + ··· b1t has only real and simple zeros.  We now formulate a technical lemma on the behavior of sequences of polynomials and its zeros which will serve as the key ingredient in the proof of Theorem 4.2. f-VECTORS OF BARYCENTRIC SUBDIVISIONS 19

Lemma 4.9. Let (gn(t))n≥0 be a sequence of real polynomials of degree d − 2, f(t) another real polynomial of degree d − 2 and ρ > 1, hd real numbers such that: n d−1 . limn→∞ gn(t)/ρ = 0, where the limit is taken in R . . All the roots of the polynomial f(t) are strictly negative and all coefficients of f(t) are strictly positive. (n) Then there are real numbers αi, 1 ≤ i ≤ d − 2 and sequences (βi )n≥0, 1 ≤ i ≤ d of complex numbers such that: (n) (i) βi , 1 ≤ i ≤ d, are real for n sufficiently large. (n) (ii) limn→∞ βi = αi, 1 ≤ i ≤ d − 2. (n) (iii) limn→∞ βd−1 = 0. (n) (iv) limn→∞ βd = −∞. Qd−1 (n) n d (v) i=0 (t − βi ) = hd + tgn(t) + ρ tf(t) + t .

Proof. Let −∞ < γ1 < ··· < γd−2 < 0 be the roots of f(t). Consider a zero γi of the polynomial f(t). Since γi is a simple zero of f(t) there is an  > 0 such that either (A) f(t) > 0 if t ∈ (γi − , γi) and f(t) < 0 if t ∈ (γi, γi + ) or (B) f(t) < 0 if t ∈ (γi − , γi) and f(t) > 0 if t ∈ (γi, γi + ). Without loss of generality we may assume (A). Set − n d Rn = min (hd + tgn(t) + ρ tf(t) + t ) t∈[γi−,γi]

+ n d Rn = max (hd + tgn(t) + ρ tf(t) + t ) t∈[γi,γi+] n By assumption limn→∞ gn(t)/ρ = 0 and hence n d lim 1/ρ · (hd + tgn(t) + t ) = 0(4.1) n→∞ n d Set Mn := max 1/ρ · |hd + tgn(t) + t |. From (4.1) it follows t∈[γi−,γi+] that limn→∞ Mn = 0. Thus for any δ > 0 there is an N(δ) such that − 1 − |Mn| < δ for n > N(δ). Set δ := − 2 min tf(t) and t ∈ [γi − , γi] t∈[γi−,γi] − 1 − − − − such that δ = 2 t f(t ). By assumption δ > 0. Then for n > N(δ )

− − − − d n − − Rn ≤ hd + t g(t ) + (t ) + ρ t f(t ) n n − . ≤ ρ Mn − 2ρ δ < ρnδ− − 2ρnδ− = −ρnδ− ≤ −δ− < 0 + 1 + If we set δ := 2 max tf(t) then analogous arguments show Rn > t∈[γi,γi+] δ+ > 0 for n > N(δ+). In particular, for all n > N(δ−),N(δ+) there is n d is a zero of hd + tgn(t) + ρ tf(t) + t in the interval [γi − , γi + ]. If 20 FRANCESCO BRENTI AND VOLKMAR WELKER we choose  < 1/4 min |γi+1 − γi| then the intervals [γi − , γi + ] are 1≤i≤d−3 (n) disjoint and hence the zeros distinct. In this situation we denote by βi n d the zero of hd + tgn(t) + ρ tf(t) + t in the interval [γi − , γi + ]. Our (n) n d arguments show that at βi the polynomial hd + tgn(t) + ρ tf(t) + t changes sign in the same way as tf(t) does at γi. From now on we assume n ≥ N(δ−),N(δ+). We now have to distin- guish between odd and even d. First, we consider the case d is even. From the assumption it follows (n) that tf(t) < 0 for t < γ1 and for γd−2 < t < 0. Thus at β1 the n d polynomial hd + tgn(t) + ρ tf(t) + t changes sign from − to + and at (n) n d βd−2 from + to −. By limt→±∞(hd + tgn(t) + ρ tf(t) + t ) = +∞ it (n) (n) (n) (n) follows that there is a zero βd−1 > βd−2 and a zero βd < β1 . If d is (n) (n) odd analogous arguments prove the existence of zeros βd−1 > βd−2 and (n) (n) βd < β1 . So far we have shown that for n large enough all zeros of hd +tgn(t)+ ρntf(t) + td are simple and real. (n) For fixed 1 ≤ i ≤ d−2 and large enough n each sequence (βi )n≥1,1 ≤ i ≤ d−2, lies in a compact interval on the real line. Since we can choose the length of the interval to be arbitrarily small it follows that the se- (n) quences (βi )n≥1, 1 ≤ i ≤ d−2, converge to fixed real numbers αi. The i n d coefficients of t , 1 ≤ i ≤ d−1, in the polynomial hd+tgn(t)+ρ tf(t)+t n converge to +∞ by limn→∞ gn(t)/ρ = 0 and the fact that all coeffi- cients of f(t) are strictly positive. It follows that there must be a zero of large absolute value. On the other hand from the fact that the con- stant term hd is independent of n it follows that there must also be a (n) zero close to 0. Since βd is bounded from above by γ1 it follows that (n) (n) limn→∞ βd = −∞. But this in turn implies limn→∞ βd−1 = 0. 

Now we are in position to provide a proof of Theorem 4.2.

(1) (2) Proof of Theorem 4.2. Let w1 , w1 , w2, . . . , wd be eigenvectors of Hd−1 for the eigenvalues 1, 1, 2, . . . , d! satisfying (iii) and (iv) of Lemma 4.6. ∆ ∆ (1) (1) (2) (2) If we expand h as a linear combination h = a1 w1 + a1 w1 + a2w2 + ··· + adwd then by Lemma 4.6 (i) and Remark 4.7 we know ∆ that ad > 0. The h-polynomial h (t) decomposes analogously into (1) (1) (2) (2) (1) (2) a1 H1 (t)+a1 H1 (t)+a2H2(t)+···+adHd(t), where H1 (t),H1 (t), (1) H2(t),...,Hd(t) are the generating polynomials of the vectors w1 , (2) w1 , w2, . . . , wd when read backward; that is, if (v0, . . . , vd) is the vector d then vd + vd−1t + ··· + v0t is its generating polynomial when read f-VECTORS OF BARYCENTRIC SUBDIVISIONS 21

(1) ∆ (2) backward. By Lemma 4.6 (iii) it follows that a1 = hd and a1 = ∆ sdn(∆) ∆ (1) (2) n h0 = 1. Clearly, h (t) = hd H1 (t) + H1 (t) + 2 a2H2(t) + ··· + n (d!) adHd(t). We set

1 g (t) := (h∆H(1)(t) + H(2)(t) + 2na H (t) + ··· n t d 1 1 2 2 n ∆ d ··· + (d − 1)! ad−1Hd−1(t) − hd − t ) and

a H (t) f(t) := d d . d!t

By Lemma 4.6 (ii)-(iv) it follows that f(t) and gn(t) are polynomials of degree d−2. Moreover, by Lemma 4.6 and Remark 4.7 all coefficients of f(t) are strictly positive and

sdn(∆) ∆ n d h (t) = hd + tgn(t) + (d!) tf(t) + t .

By Lemma 4.8 the polynomial Hd(t) and therefore the polynomial f(t) has only simple and real zeros that are all strictly negative. Thus gn(t) and f(t) satisfy the assumption of Lemma 4.9 and therefore the assertion follows by Lemma 4.9. 

It seems to be a challenging problem to determine the numbers αi, 1 ≤ i ≤ d − 2, from Theorem 4.2. As our proofs show this in turn is closely related to determining eigenvectors of the matrices Fd−1 and Hd−1 for the eigenvalue d!. We were not able to give an explicit de- scription of such an eigenvector. Problem 4.10. Give a description of an eigenvector of the matrices Fd−1 and Hd−1 for the eigenvalue d!.

Acknowledgment We thank the referee for her/his numerous suggestions. In partic- ular, we are grateful for asking the question that led us to the proof Proposition 4.3.

References [1] Bruns, W., Herzog, J.: Cohen-Macaulay rings. Cambridge University Press, Cambridge (1993) [2] Bj¨orner,A., Farley, J.D.: Chain polynomials of distributive lattices are 75% unimodal. Electr. J. of Comb. 12, 1-7 (2005) 22 FRANCESCO BRENTI AND VOLKMAR WELKER

[3] Brenti, P.: Unimodal log-concave and P´olya frequency sequences in combina- torics. Memoirs of the Amer. Math. Soc. 81, American Mathematical Society, Providence (1989) [4] Brenti, F.: q-Eulerian polynomials arising from Coxeter groups. European J. Combin. 15, 417–441 (1994) [5] Br¨anden, P.: On linear transformations preserving the P´olya frequency prop- erty. Trans. Amer. Math. Soc. 358, 3697-3716 (2006) [6] Charney, R., Davis, M.: Euler characteristic of a nonpositively curved, piece- wise Euclidean manifold. Pac. J. Math. 171, 117–137 (1995) [7] Gal, S.R.: The real root conjecture fails for five- and higher-dimensional spheres. Disc. and Comp. Geom. 34, 269-284 (2005) [8] Horn, R.A., Johnson, C.R.: Matrix analysis. Cambridge University Press, Cambridge (1985) [9] Jonsson, J., Welker, V.: Complexes of injective words. in preparation. [10] Karu, K.: The cd-index of fans and lattices. Compositio Math. 142, 701-718 (2006) [11] Reiner, V., Welker, V.: On the Charney-Davis and Neggers-Stanley conjec- tures. J. Comb. Theory, Ser A, 109, 247-280 (2005) [12] Stanley, R.P.: Some aspects of groups acting on posets. J. Comb. Theory, Ser A, 32, 132-161 (1002) [13] Stanley, R.P.: Enumerative combinatorics, vol. I. Cambridge University Press, Cambridge (1997) [14] Stanley, R.P.: Flag f-vectors and the cd-index. Math. Z. 216, 483-499 (1994) [15] Ziegler, G.M.: Lectures on polytopes. Springer Verlag, New York (1994)

Dipartimento di Matematica, Universita’ di Roma ”Tor Vergata”, Via della Ricerca Scientifica, 1, 00133, Roma, Italy E-mail address: [email protected]

Fachbereich Mathematik und Informatik, Philipps-Universitat¨ Mar- burg, 35032 Marburg, Germany E-mail address: [email protected]