ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P2.03 https://doi.org/10.26493/1855-3974.2894.b07 (Also available at http://amc-journal.eu) A non-associative incidence near-ring with a generalized Möbius function* John Johnson †, Max Wakefield ‡ US Naval Academy, 572-C Holloway Rd, Annapolis MD, 21402 USA This paper is dedicated to the memory of John Johnson. Received 1 June 2022, accepted 27 February 2023, published online 20 September 2023 Abstract There is a convolution product on 3-variable partial flag functions of a locally finite poset that produces a generalized Möbius function. Under the product this generalized Möbius function is a one sided inverse of the zeta function and satisfies many generaliza- tions of classical results. In particular we prove analogues of Phillip Hall’s Theorem on the Möbius function as an alternating sum of chain counts, Weisner’s Theorem, and Rota’s Crosscut Theorem. A key ingredient to these results is that this function is an overlapping product of classical Möbius functions. Using this generalized Möbius function we define analogues of the characteristic polynomial and Möbius polynomials for ranked lattices. We compute these polynomials for certain families of matroids and prove that this generalized Möbius polynomial has -1 as root if the matroid is modular. Using results from Ardila and Sanchez we prove that this generalized characteristic polynomial is a matroid valuation. Keywords: Incidence algebra, matroid, Möbius function, valuation. Math. Subj. Class. (2020): 37K15, 42A99, 60E05, 05A17 *The authors are very thankful for detailed comments by the reviewer. The reviewers suggestions have signif- icantly improved the article. The authors are thankful for discussions with Carolyn Chun, Joel Lewis, and Will Traves. The authors are also thankful to George Andrews for help on Lemma 6.14. Frederico Ardila and Mario Sanchez significantly helped with the material on valuations for which the authors are very thankful. Also, Jose Bastidas made multiple excellent comments for which the authors are very thankful. The authors would like to thank the US Naval Academy trident program for support during this project. †Supported by the US Naval Academy as a Trident Scholar. ‡Corresponding author. E-mail addresses: m213162@usna.edu (John Johnson), wakefiel@usna.edu (Max Wakefield) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 2 Ars Math. Contemp. 24 (2024) #P2.03 1 Introduction Combinatorial invariants in incidence algebras play a central role in many areas of com- binatorics as well as in number theory, algebraic topology, algebraic geometry, and repre- sentation theory. In particular, the Möbius function appears in the inverse of the Riemann zeta function as well as the coefficients of the chromatic polynomial for graphs. In this note we study a generalization of the classical incidence algebra by looking at three vari- able incidence functions. A large portion of this study is focussed on studying a 3-variable generalized Möbius function inside this generalized incidence structure. Incidence algebras and Möbius functions were popularized by Rota in [26]. Rota char- acterized the classical Möbius function from number theory (see [20] and [14]) as the inverse of the constant function 1 on the intervals of the poset which is called the zeta function. In [26] Rota gives many results on the Möbius function, including his Crosscut Theorem. Since then, many advances can be attributed to Möbius functions. Of particular importance are the counting theorems of Zaslavsky in [33] and Terao’s factorization the- orem (see [29]) using the Möbius function in the form of the characteristic polynomial of a hyperplane arrangement. The main motivation for this work is to build invariants which are finer than the classical Möbius function and characteristic polynomial to obtain more information about the underlying combinatorial structure. More recently, there has been considerable developments in understanding of some classical invariants on matroids. One generalization came from Krajewski, Moffatt, and Tanasa who built Tutte polynomials from a Hopf algebra in [18]. Taking this a little fur- ther, in [11] Dupont, Fink and Moci construct a categorical framework to view various combinatorial invariants and they prove some convolution formulas. The work of Aguiar and Ardila in [1] framed many combinatorial structures like matroids in terms of general- ized permutahedra, where there is a natural Hopf monoid governing classical operations. One possible starting place for this study could be the work of Joni and Rota in [16]. Then, in [6], Ardila and Sanchez use this Hopf monoid structure to build a concrete method for investigating valuations on many combinatorial structures. Another aim of this study is to add another invariant to the list of valuations. Concretely, we use the methods of Ardila and Sanchez to show that one of our invariants is a valuation on matroids. One view that one can take for many combinatorial structures is that of posets (e.g. matroids are geometric lattices) and this is the view that we take here. The starting point for our study is the collection of 3-variable functions on ordered triples of elements in a poset. The set of these 3-variable functions also appears in the book [2] by Aguiar and Mahajan in Appendix C4 where they study 2-cochains and 2-cocycles. We differ from the work in Appendix C4 [2] by equipping this set of functions with a spe- cial convolution product. The motivation for this product comes from trying to symmetrize a more natural convolution product that was studied by the second author in [30] as well as making new invariants with special properties. This product provides a 3-variable Möbius function which is a sort of left inverse of the 3-variable analogue of the zeta function. We call this function the J-function and study many of its properties. It turns out that it is essentially a staggered product of the classical Möbius functions and hence satisfies gen- eralizations of many of the classical theorems on the classical Möbius function. To prove these results we develop and use certain operations and formulas these 3-variable func- tions satisfy that give maps between various different types of incidence algebras. In [8] Jose Bastidas studies Type B Hopf monoids and defines an antipode via some convolution formulas which seem to have some similar properties to the work presented here. J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 3 As an application we build two different polynomials from the J-function: a general- ized characteristic polynomial and a generalized Möbius polynomial (see [17] and [21] for Möbius polynomials). It turns out that these polynomials have some interesting properties that are not apparent from the surface. In the case of matroids, the generalized character- istic polynomial has positive coefficients. We then compute these polynomials for certain families of matroids and find special roots. Of particular interest is that the generalized Möbius function has −1 as a root for modular matroids, which mimics Theorem 1 in [21]. However, we show that the converse is not true and so one is led to question what do these polynomial count? Could there be some chromatic generalization for the generalized poly- nomials, or some lattice point or finite field counting formula for these polynomials (like [9] or [7])? Also, in [8], Bastidas defines some polynomial invariants via characters of a Hopf monoid. Can the polynomials we define here be put in the framework of [8]? We finish by employing the methods of Ardila and Sanchez in [6] to show that our generalized characteristic polynomial is a matroid valuation. This follows from the fact that the J-function splits as a product of Möbius functions. In the case of the Möbius polynomial, we are not sure whether or not it is a valuation, yet we show that it does have a decomposition in terms of the classical characteristic polynomials. We find it interest- ing that this decomposition looks very similar to the recursive definition of the matroid Kazhdan-Lusztig polynomial originally defined in [12]. We begin this study with reviewing classical results on incidence algebras and Möbius functions in Section 2. Then we define our 3-variable incidence structure in Section 3. There we show that this structure has some interesting properties but that it is neither as- sociative nor distributive. However, in Section 4 we develop multiple operations which give nice formulas between these different kinds of incidence functions. Using these for- mulas we define a generalized Möbius function, the J-function, and study its properties in Section 5. Finally in Section 6 we define our generalized characteristic and Möbius polynomials. 2 Incidence Algebras Let R be a commutative ring and P be a locally finite poset. We follow [28] and [4] for combinatorics on posets. For the remainder of this note we refer to the order in P by ≤. Also, for n ∈ N let [n] = {1, 2, 3, . . . , n}. In this section we review basic material of incidence algebras where we follow [27]. First we define the poset of partial flags. Definition 2.1. The poset of partial flags of length k on P is F lk(P) = { (x1, x2, . . . , xk) ∈ Pk| x1 ≤ x2 ≤ · · · ≤ xk } with order given by (x1, . . . , xk) ⪯ (y1, . . . , yk) if and only if for all i ∈ [k] we have xi ≤ yi. Now we define the classical incidence algebras. Definition 2.2. The incidence algebra on P is the set I(P, R) = Hom(F l2(P), R) where R is a commutative ring. Addition in I(P, R) is given by (f + g)(x, y) = f(x, y) + g(x, y), 4 Ars Math. Contemp. 24 (2024) #P2.03 the multiplication is given by convolution (f ∗ g)(x, y) = ∑ x≤a≤b f(x, a)g(a, y), and the scalar product is given by (rf)(x, y) = rf(x, y) for all r ∈ R. In this note, we will examine multiple different operations on functions on posets. For this reason we will reserve juxtaposition only for products of elements in the ring R. Oth- erwise we will denote products of functions with specific operation names like ∗. It turns out that I(P, R) is a non-commutative R-algebra with identity element given by the Kronecker delta function δ(x, y) = { 1 if x = y, 0 else. There are two other very important elements in I(P, R). Definition 2.3. The zeta function ζ ∈ I(P, R) is defined as the constant function on F l2(P) ζ(x, y) = 1 for all (x, y) ∈ F l2(P). The Möbius function µ ∈ I(P, R) is defined by∑ x≤a≤y µ(x, a) = ∑ x≤a≤y µ(a, y) = δ(x, y) for all (x, y) ∈ F l2(P). The Möbius function was originally defined by Möbius (see [20]) on the poset of the natural numbers ordered by division for the purpose of inverting the Riemann zeta function. Since then the Möbius function has been used in many different contexts and broadened by the work of Rota in [26]. For our discussion, it is important to note that µ is the multiplica- tive inverse of the zeta function µ ∗ ζ = ζ ∗ µ = δ. Now we review how the incidence algebra functor factors over products. Recall that for posets P and Q the product poset is P ×Q with order given by (x1, x2) ≤ (y1, y2) if and only if x1 ≤ y1 and x2 ≤ y2. Proposition 2.4 (Proposition 2.1.12 [27]). If P and Q are locally finite posets then I(P, R)⊗R I(Q, R) ∼= I(P ×Q, R). Because of Proposition 2.4 we define the following operation on functions. In order to the make the exposition clear in the case when we are dealing with functions over different posets, we will put the poset in the subscript. For fP ∈ I(P, R) and gQ ∈ I(Q, R) define fP × gQ ∈ I(P ×Q, R) by (fP × gQ)((x1, x2), (y1, y2)) = fP(x1, y1)gQ(x2, y2). We will use this notation and the following consequence of Proposition 2.4 in our study in Section 5. J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 5 Corollary 2.5. If P and Q are locally finite posets then µP × µQ = µP×Q. Next we recall how the Möbius function counts chains (or is an Euler characteristic for the order complex). For (x, y) ∈ F l2(P) let ci(x, y) = ∣∣{(a0, . . . , ai) ∈ F li+1 : ∀k, ak < ak+1 and a0 = x and ai = y}∣∣ be the number of chains of length i between x and y. Theorem 2.6 (Phillip Hall’s Theorem [13]; Proposition 3.8.5 [28]). If P is a locally finite poset and (x, y) ∈ F l2(P) then µ(x, y) = ∑ i (−1)ici(x, y). Now we review Rota’s Crosscut Theorem. Let L be a finite lattice with 0̂ the minimum element and 1̂ the maximum element. Usually, Rota’s Crosscut Theorem is stated globally in the lattice giving a formula for µ(0̂, 1̂). However, for our generalization we will need a local version. Definition 2.7. Let (x, y) ∈ F l2(L). A lower crosscut of the interval [x, y] = {a ∈ L|x ≤ a ≤ y} is a set Sx,y ⊆ [x, y]\{x} such that if b ∈ [x, y]\(Sx,y ∪ {x}) then there is some a ∈ Sx,y with a < b. A upper crosscut of the interval [x, y] is a set Tx,y ⊆ [x, y]\{y} such that if a ∈ [x, y]\(Tx,y ∪ {y}) then there is some b ∈ Tx,y with a < b. This definition gives Rota’s famous Crosscut Theorem which we state in the style of Lemma 2.35 in [22] for use in arrangement theory. Theorem 2.8 ([26, Theorem 3]). If L is a lattice, (x, y) ∈ F l2(L), and Sx,y is a lower crosscut of [x, y] then µ(x, y) = ∑ A⊆Sx,y∨ A=y (−1)|A|. Dually, if Tx,y is an upper crosscut of [x, y] then µ(x, y) = ∑ B⊆Tx,y∧ B=x (−1)|B|. Next we consider Weisner’s Theorem (see [31]). Theorem 2.9 ([28, Weisner’s Theorem, Corollary 3.9.3]). If L is a finite lattice with at least two elements and 1̂ ̸= a ∈ L then∑ x∈L x∧a=0̂ µ(x, 1̂) = 0. Now we recall one more result that follows from this classical result for matroids: the Mobius function of the lattice of flats of a matroid alternates in sign. Lemma 2.10. If L is a finite semimodular lattice then sgn(µ(x, y)) = (−1)rk(x)+rk(y). 6 Ars Math. Contemp. 24 (2024) #P2.03 3 A 3-variable incidence non-associative near-ring In this section, we define the algebraic structures where our invariants live. It turns out that these algebraic structures support various operations that can yield nice formulas. Later these formulas will be used to show certain formulas and relations on our new invariants. Definition 3.1. Let R be a commutative ring and P be a locally finite poset. Define the 3-variable incidence left near-ring as J(P, R) = Hom(F l3(P),R) with binary operations as follows: • For f, g ∈ J(P, R) we define addition by (f + g)(x, y, z) = f(x, y, z) + g(x, y, z). • For f, g ∈ J (P, R) we define a multiplication by (f  g)(x, y, z) = ∑ (a,b)⊴(x,y,z) f(x, a, a)g(a, y, b)f(b, b, z) where the juxtaposition in each term is multiplication in the ring R and (a, b) ⊴ (x, y, z) means x ≤ a ≤ y ≤ b ≤ z in P . First we show that J(P, R) is indeed left distributive. Proposition 3.2. If P is any poset then the multiplication  in J(P, R) is left distributive. Proof. Let f, g, h ∈ J(P, R) and (x, y, z) ∈ F l3(P). Then (f  (g + h))(x, y, z) = ∑ (a,b)⊴(x,y,z) f(x, a, a)(g + h)(a, y, b)f(b, b, z) = ∑ (a,b)⊴(x,y,z) f(x, a, a)(g(a, y, b) + h(a, y, b))f(b, b, z) = ∑ (a,b)⊴(x,y,z) f(x, a, a)g(a, y, b)f(b, b, z) + ∑ (a,b)⊴(x,y,z) f(x, a, a)h(a, y, b)f(b, b, z) = (f  g)(x, y, z) + (f  h)(x, y, z). Remark 3.3. With this + the set J(P, R) is an abelian group. It would be convenient if J(P, R) were naturally an R-algebra. However, this is far from the case as we will see. Even the natural action of R on J(P, R) is flawed. Let r ∈ R and f, g ∈ J(P, R) then r · (f  g) = f  (r · g) but (r · f)  g = r2 · (f  g). Fortunately, though, there are a few special functions in J(P, R) that provide substantial information. We will use these to study the structure of J(P, R) and define other special elements later. J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 7 Definition 3.4. Assume that 1 is the multiplicative identity and 0 is the additive identity in R. • Define δ3 ∈ J(P, R) by δ3(x, y, z) = { 1 if x = y = z 0 otherwise • Define ζ3 ∈ J(P, R) by setting ζ3(x, y, z) = 1 for all (x, y, z) ∈ F l3(P). With these functions we can investigate basic properties of J(P, R). Proposition 3.5. The element δ3 ∈ J(P, R) is a left multiplicative identity. Proof. Let f ∈ J(P, R) and (x, y, z) ∈ F l3(P). Then (δ3  f)(x, y, z) = ∑ (a,b)⊴(x,y,z) δ3(x, a, a)f(a, y, b)δ3(b, b, z) = δ3(x, x, x)f(x, y, z)δ3(z, z, z) = f(x, y, z). In the next three propositions we note that in general J(P, R) is not commutative, as- sociative, or right distributive. We could do this with a single example, however these propositions show that J(P, R) is basically never commutative, associative, or right dis- tributive. Proposition 3.6. If P is a non-trivial poset (it has at least two comparable elements) or the base ring is not Boolean (not idempotent), then the multiplication  in J(P, R) is non- commutative and δ3 is not a right multiplicative identity. Proof. Let (x, y, z) ∈ F l3(P) and suppose that either x < y or that y < z in P or that R is not Boolean. Under these assumptions we can construct a function f ∈ J(P, R) that has f(x, y, z) ̸= f(x, y, y)f(y, y, z). Then from Proposition 3.5 we have (δ3  f)(x, y, z) = f(x, y, z) but (f  δ3)(x, y, z) = f(x, y, y)f(y, y, z). The proof for the next fact is very similar. Proposition 3.7. If P is a poset with three elements x, y, z satisfying x < y < z or the base ring is not Boolean (not idempotent), then the multiplication  in J(P, R) is non- associative. Proof. Let (x, y, z) ∈ F l3(P) be three elements satisfying x < y < z in P or that R is not Boolean. Under these assumptions we can construct a function f ∈ J(P, R) that has f(x, y, y)f(y, y, y)2f(y, y, z) ̸= f(x, y, y)f(y, y, z). Compute ((f  δ3)  δ3))(x, y, z) = ∑ (a,b)⊴(x,y,z) (f  δ3)(x, a, a)δ3(a, y, b)(f  δ3)(b, b, z) = [(f  δ3)(x, y, y)][(f  δ3)(y, y, z)] = [f(x, y, y)f(y, y, y)][f(y, y, y)f(y, y, z)]. Then from Proposition 3.5 we have (f  (δ3  δ3))(x, y, z) = (f  δ3)(x, y, z) = f(x, y, y)f(y, y, z) which is different from ((f  δ3)  δ3))(x, y, z) by our assumption on f . 8 Ars Math. Contemp. 24 (2024) #P2.03 Proposition 3.8. If P is a non-trivial poset (it has at least two comparable elements) and R is any non-trivial commutative ring, then the multiplication  in J(P, R) is not right distributive. Proof. Let (x, y, z) ∈ F l3(P) and f ∈ J(P, R) be any function such that f(x, y, y) + f(y, y, z) ̸= 0. Then ((f + ζ3)  δ3)(x, y, z) = ∑ (a,b)⊴(x,y,z) (f + ζ3)(x, a, a)δ3(a, y, b)(f + ζ3)(b, b, z) = [(f + ζ3)(x, y, y)][(f + ζ3)(y, y, z)] = f(x, y, y)f(y, y, z) + f(x, y, y) + f(y, y, z) + 1. On the other hand we have ((f  δ3) + (ζ3  δ3))(x, y, z) = f(x, y, y)f(y, y, z) + ζ3(x, y, y)ζ3(y, y, z) = f(x, y, y)f(y, y, z) + 1 which by the hypothesis on f we have the right distributive property not holding. With Propositions 3.5, 3.6, 3.7, 3.2, and 3.8 we conclude that J(P, R) is a left only unital, non-commutative, non-associative, near-ring (see [25] for this terminology). Also, note that there is the zero function Z ∈ J(P, R) which satisfies Z  f = f  Z = Z for all f ∈ J(P, R). Further note that addition in J(P, R) is abelian. Hence J(P, R) is an abelian, zero-symmetric, left only unital, non-commutative, non-associative, near-ring. It is worth noting that in general J(P, R) is not even close to being associative on both sides and is not an alternative algebra or any similar generalization. Now we look at a few special cases that do not satisfy the hypothesis of some of these propositions. Example 3.9. Let P = B0 = {0} be the poset with just one element and R any commu- tative ring. Then as a set J(B0, R) = R, but multiplication is given by a  b = aba = a2b. If R is Boolean then J(B0, R) ∼= R. Otherwise, this near-ring is not associative, not com- mutative, and is only left unital. Example 3.10. Let P = B1 = {0, 1} be the Boolean poset of rank 1 and R be any Boolean ring (one example would be F2). Then the hypothesis of Proposition 3.7 is not satisfied and the non-equality f(x, y, y)f(y, y, y)2f(y, y, z) ̸= f(x, y, y)f(y, y, z) used in the proof is always equal. It turns out that in this case J(B1, R) is associative and we prove this now. In order to shorten the calculation we will denote (0, 0, 0) by 0⃗ and (1, 1, 1) by 1⃗. First we see that ((f  g)  h)(⃗0) = f (⃗0)g(⃗0)h(⃗0) = (f  (g  h))(⃗0). Then for the non-trivial tuple (0, 0, 1) we compute ((f  g)  h)(0, 0, 1) =(f  g)(⃗0)h(⃗0)(f  g)(0, 0, 1) + (f  g)(⃗0)h(0, 0, 1)(f  g)(⃗1) =f (⃗0)g(⃗0)h(⃗0)[f (⃗0)g(⃗0)f(0, 0, 1) + f (⃗0)g(0, 0, 1)f (⃗1)] + f (⃗0)g(⃗0)h(0, 0, 1)f (⃗1)g(⃗1) =f (⃗0)g(⃗0)h(⃗0)f(0, 0, 1) + f (⃗0)g(⃗0)h(⃗0)g(0, 0, 1)f (⃗1) + f (⃗0)g(⃗0)h(0, 0, 1)f (⃗1)g(⃗1). J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 9 Then the other side of the associative identity is (f  (g  h))(0, 0, 1) =f (⃗0)(g  h)(⃗0)f(0, 0, 1) + f (⃗0) [ (g  h)(0, 0, 1) ] f (⃗1) =f (⃗0)g(⃗0)h(⃗0)f(0, 0, 1) + f (⃗0) [ g(⃗0)h(⃗0)g(0, 0, 1) + g(⃗0)h(0, 0, 1)g(⃗1) ] f (⃗1) =((f  g)  h)(0, 0, 1). Hence J(B1, R) is associative. This example does satisfy the hypothesis of Proposition 3.8. Hence J(B1, R) is a (associative) left abelian (addition is commutative) near-ring. That’s about as good as it gets though. For example, if R = F2 then J(B1,F2) is not a near-field because any function with f (⃗0) = 0 and f(0, 0, 1) = 1 does not have an inverse. For exactly the same reason δ3 ∈ J(B1,F2) is still not a right identity element. 4 Operations on incidence functions In this section we look at a relationship between the classical incidence algebra I(P, R) and J(P, R). For f, g ∈ I(P, R) we define f♢g ∈ J(P, R) by setting (f♢g)(x, y, z) = f(x, y)g(y, z). We can use the ♢ operation to construct interesting elements in J(P, R). There are rela- tionships between the operations ∗ in I(P, R),  in J(P, R), and ♢. Proposition 4.1. If f, g, r, s ∈ I(P, R) and f(b, b)g(a, a) = 1 for all a, b ∈ P then (f♢g)  (r♢s) = (f ∗ r)♢(s ∗ g). Proof. Let (x, y, z) ∈ F l3(P) and f, g, r, s ∈ I(P, R). Then ((f♢g)  (r♢s))(x, y, z) = ∑ (a,b)⊴(x,y,z) (f♢g)(x, a, a)(r♢s)(a, y, b)(f♢g)(b, b, z) = ∑ (a,b)⊴(x,y,z) f(x, a)g(a, a)r(a, y)s(y, b)f(b, b)g(b, z) =  ∑ x≤a≤y f(x, a)r(a, y)  ∑ y≤b≤z s(y, b)g(b, z)  = [(f ∗ r)(x, y)] [(s ∗ g)(y, z)] =((f ∗ r)♢(s ∗ g))(x, y, z) where the third equality only holds due the the assumption. One can see from the proof that without the hypothesis on f and g that the equality will not hold. Hence there is no hope for this to give any kind of near-ring homomorphism from a twisted product version of I(P, R)× I(P, R). Also, the natural addition homomorphism assumption does not hold. Instead we have the following proposition which does not have special hypothesis on the functions. For this proposition there are two different additions, for I(P, R) and J(P, R), which for brevity we use the same addition symbol. 10 Ars Math. Contemp. 24 (2024) #P2.03 Proposition 4.2. If f, g, r, s ∈ I(P, R) then (f + g)♢(r + s) = (f♢r) + (f♢s) + (g♢r) + (g♢s). Proof. For all (x, y, z) ∈ F l3(P) ((f + g)♢(r + s))(x, y, z) =(f(x, y) + g(x, y))(r(y, z) + s(y, z)) =f(x, y)r(y, z) + f(x, y)s(y, z) + g(x, y)r(y, z)+ g(x, y)s(y, z) =((f♢r) + (f♢s) + (g♢r) + (g♢s))(x, y, z) which is the identity we are looking for. We can also define products of functions on products of posets over 3-flags. We prefer to limit our study of J(P, R) to this product definition since the technicalities of tensor products over non-associative near-rings would present significant and unnecessary com- plications. Definition 4.3. Let P and Q be locally finite posets, fP ∈ J(P, R), and gQ ∈ J(Q, R). Define fP × gQ ∈ J(P ×Q, R) by (fP × gQ)((x1, x2), (y1, y2), (z1, z2)) = fP(x1, y1, z1)gQ(x2, y2, z2). Now we show how the ♢ operation is compatible with products of posets. Proposition 4.4. If P and Q are locally finite posets, fP , gP ∈ I(P, R), and rQ, sQ ∈ I(Q, R) then (fP♢gP)× (rQ♢sQ) = (fP × rQ)♢(gP × sQ). Proof. Let ((x1, x2), (y1, y2), (z1, z2)) ∈ F l3(P ×Q). Then ((f♢g)× (r♢s))((x1, x2), (y1, y2), (z1, z2)) = [(f♢g)(x1, y1, z1)] [(r♢s)(x2, y2, z2)] = [f(x1, y1)g(y1, z1)] [r(x2, y2)s(y2, z2)] = [f(x1, y1)r(x2, y2)] [g(y1, z1)s(y2, z2)] = [(f × r)((x1, y1), (x2, y2))] [(g × s)((y1, z1), (y2, z2))] = ((f × r)♢(g × s))((x1, x2), (y1, y2), (z1, z2)) which completes the proof. As in Proposition 4.4 we will now show how the operations × and  factor over prod- ucts of posets. We use subscripts on these operations to keep track of which poset the operation is applied. Proposition 4.5. If P and Q be locally finite posets, fP , gP ∈ J(P, R), and rQ, sQ ∈ J(Q, R) then (fP P gP)× (rQ Q sQ) = (fP × rQ) P×Q (gP × sQ). J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 11 Proof. Let x = (x1, x2), y = (y1, y2), z = (z1, z2) ∈ P ×Q so that (x, y, z) ∈ F l3(P × Q) and a = (a1, a2), b = (b1, b2) ∈ P ×Q so that (a, b) ∈ F l2(P ×Q). Then ((fP×rQ) P×Q (gP × sQ))(x, y, z) = ∑ (a,b)⊴(x,y,z) (fP × rQ)(x, a, a)(gP × sQ)(a, y, b)(fP × rQ)(b, b, z) = ∑ (a1,b1) ∑ (a2,b2) fP(x1, a1, a1)gP(a1, y1, b1)fP(b1, b1, z1) rQ(x2, a2, a2)sQ(a2, y2, b2)rQ(b2, b2, z2) = [(fP  gP)(x1, y1, z1)] [(rQ  sQ)(x2, y2, z2)] = ((fP P gP)× (rQ Q sQ))(x, y, z) which is the required identity. 5 The J-function Let P be a locally finite poset. In this section we define the central invariant of this note which we call the J function. This function is a generalization of the classical Möbius function µ. We show that it satisfies generalizations of the classical theorems on µ. A key ingredient for these results is the operation ♢. Definition 5.1. Define J : F l3(P) → Z for all fixed (x, y, z) ∈ F l3(P ) by∑ (a,b)⊴(x,y,z) J(a, y, b) = δ3(x, y, z). This function is well defined because either x = y = z with J(x, y, z) = 1 or otherwise all of the following summations are finite J(x, y, z) =− ∑ x 2. The lattice Pn consists of 0̂, 1̂, and n atoms α1, . . . , αn. Now JPn(0̂, 0̂, 1̂) = n − 1 and JPn(0̂, 1̂, 1̂) = n − 1 are the only JPn values that do not have αn as an entry and incorporate αn in it’s recursive definition. So, JPn(0̂, 0̂, 1̂) = JPn−1(0̂, 0̂, 1̂) + 1 and similarly for (0̂, 1̂, 1̂). Incorporating this difference into the calculation we get that M(Pn, t) =M(Pn−1, t) + t4 + t2 + J(0̂, 0̂, αn)t5 + J(0̂, αn, αn)t4 + J(0̂, αn, 1̂)t3 + J(αn, αn, αn)t 3 + J(αn, αn, 1̂)t 2 + J(αn, 1̂, 1̂)t =(t2 − (n− 1)t+ 1)(t+ 1)2(t− 1)2 − (t5 − 2t3 + t) =(t2 − nt+ 1)(t+ 1)2(t− 1)2 which is the desired formula. Now we consider a decomposition of M(L, t) for a finite lattice L. If L is a finite lattice then Lop is the same underlying set as L but with the order reversed (i.e. x ≤op y in Lop if and only if x ≥ y in L). Also for y ∈ L let Ly = {x ∈ L|x ≤ y} and Ly = {x ∈ L|x ≥ y}. Now we can state the result. Proposition 6.10. If L is a finite ranked lattice then M(L, t) = trk(L) ∑ y∈L tcrk(y)χ(Ly, t)χ((Lop)y, t−1). J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 17 Proof. First we note that for x ≤ y ∈ L the Möbius function on Lop has µop(y, x) = µ(x, y) and that rank is corank in Lop. Then again using Theorem 5.2 we compute M(L, t) = ∑ (x,y,z)∈Fl3(P) J(x, y, z)tρ(x,y,z) = ∑ y∈L ∑ x≤y ∑ z≥y µ(x, y)µ(y, z)tcrk(x)+crk(y)+crk(z) = ∑ y∈L tcrk(y) ∑ x≤y µ(x, y)tcrk(x) ∑ z≥y µ(y, z)tcrk(z) = ∑ y∈L tcrk(y)χ(Ly, t) ∑ x≤y µ(x, y)trk(L)−rk(x) = ∑ y∈L tcrk(y)χ(Ly, t)trk(L) ∑ x≥opy µop(y, x)t−rk(x) =trk(L) ∑ y∈L tcrk(y)χ(Ly, t)χ((Lop)y, t−1). We can use Proposition 6.10 to compute M(P, t) for cases where χ(P, t) is well known. Let Lnq be the modular lattice of all subspaces in Fnq , a vector space of dimen- sion n over a field with q elements. The Möbius function and the characteristic polynomial of Lnq are well known. Proposition 6.11 ([34, Proposition 7.5.3]). In Lnq we have µ(0̂, 1̂) = (−1)nq( n 2) and χ(Lnq , t) = n−1∏ i=0 (t− qi). Using this we can get a nice formulation for M(Lnq , t). First we need to recall some terminology from q-series. Let[ n k ] q = (qn − 1) · · · (q − 1) (qk − 1) · · · (q − 1) · (qn−k − 1) · · · (q − 1) be the q-binomial coefficient (aka Gaussian coefficient). Also, we denote by[ n k1, k2, . . . , km ] q = [ n k1 ] q [ n− k1 k2 ] q · · · [ n− (k1 + · · · km−1) km ] q the q-multinomial coefficient. We also use the q-Pochhammer symbol (a; q)n = n−1∏ i=0 (1− aqi). 18 Ars Math. Contemp. 24 (2024) #P2.03 We use [3] for a general reference for q-series. Using Proposition 6.11 we get the following. Proposition 6.12. If Lnq is the modular lattice of subspaces of Fnq then M(Lnq , t) = ∑ 0≤i≤j≤k≤n (−1)k−i [ n i, j − i, k − j, n− k ] q q( j−i 2 )+( k−j 2 )t3n−i−j−k. Proof. Use that [ n k ] q counts the number of subspaces of dimension k in Fnq and apply Theorem 5.2 to J in M(Lnq , t) together with Proposition 6.11. Now we can reformulate Proposition 6.12 using Proposition 6.10 together with Propo- sition 6.11 to get a nice identity in q-series. Proposition 6.13. If Lnq is the modular lattice of subspaces of Fnq , then M(Lnq , t) = tn ∑ 0≤k≤n tn−k [ n k ] q n−k−1∏ i=0 (t− qi) k−1∏ j=0 (t− qj). It turns out that −1 is a root of M(Lnq , t). We need a few results in order to prove this. First we present a formula or q-identity which seems to be a kind of q-generalized binomial theorem (the authors could not find it in the literature). It’s interesting that in the odd case the sum trivially collapses but not for the even case. Lemma 6.14. If n > 0 then n∑ k=0 (−1)k [ n k ] q (−1 : q)n−k(−1; q)k = 0. Proof. Let S(n) = n−1∑ k=0 (−1)k [ n k ] q (−1; q)n−k(−1; q)k (−1; q)n which is the left hand side up to the n − 1 term divided by the nth term. Using tech- niques from [24] and Mathematica [15] we build a recursion for S(n). We compute J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 19 (1 + qn−1)S(n) = n−1∑ k=0 (−1)k ( qk [ n− 1 k ] q + [ n− 1 k − 1 ] q ) (−1; q)n−k(−1; q)k (−1; q)n−1 = n−1∑ k=0 (−1)k [ n− 1 k ] q (−1; q)n−k−1(−1; q)k (−1; q)n−1 (qk + qn−1) + n−1∑ k=1 (−1)k [ n− 1 k − 1 ] q (−1; q)n−k(−1; q)k (−1; q)n−1 =(−1)n−12qn−1 + n−2∑ k=0 (−1)k [ n− 1 k ] q (−1; q)n−k−1(−1; q)k (−1; q)n−1 qk + n−2∑ k=0 (−1)k [ n− 1 k ] q (−1; q)n−k−1(−1; q)k (−1; q)n−1 qn−1 + n−2∑ k=0 (−1)k+1 [ n− 1 k ] q (−1; q)n−(k+1)(−1; q)k+1 (−1; q)n−1 =(−1)n−12qn−1 + n−2∑ k=0 (−1)k [ n− 1 k ] q (−1; q)n−k−1(−1; q)k (−1; q)n−1 qk + qn−1S(n− 1)− n−2∑ k=0 (−1)k [ n− 1 k ] q (−1; q)n−k−1(−1; q)k (−1; q)n−1 (1 + qk) =(−1)n−12qn−1 + qn−1S(n− 1)− S(n− 1). Now we prove with induction that S(n) = (−1)n−1. First we see that S(1) = 1. Then using the recursion above we have (1 + qn−1)S(n) = (−1)n−12qn−1 − qn−1(−1)n−1 + (−1)n−1 = (−1)n−1(qn−1 + 1) which finishes the proof. Proposition 6.15. If Lnq is the modular lattice of subspaces of Fnq then M(Lnq ,−1) = 0. Proof. Evaluate the expression in Proposition 6.13 and apply Lemma 6.14. Now we can prove the main result of this section. Theorem 6.16. If L is a modular geometric lattice (modular matroid) then M(L,−1) = 0. Proof. Use the classical result that a modular geometric lattice is product of Boolean and projective spaces (see 12.1 Theorem 4 in [32] or Proposition 6.9.1 in [23]). Then the result follows from Propositions 6.15, 6.8, and 6.6. Remark 6.17. The proof of Theorem 6.16 is done in cases. It would be interesting if there was a case free proof just using the modular property. 20 Ars Math. Contemp. 24 (2024) #P2.03 Remark 6.18. At first when looking at examples of M on the lattice of flats L(M) of a matroid M it seems that the converse of Theorem 6.16 might be true. As for even the simplest non-modular matroid U3,4 has M polynomial M(L(U3,4), t) = (t− 1)(t8 − 3t7 − t6 + 12t5 − 2t4 − 12t3 + 3t2 + 5t− 1) which does not have a factor of (t + 1). However, the converse is false, but the example seems rather special. Using the SageMath computer algebra system [10] we compute M(L(M∗(K3,3)), t) = (t10 − 9t9 + 22t8 + 12t7 − 81t6 + 21t5 + 69t4 − 18t3 − 34t2 + 15t− 1)(t+ 1)(t− 1) where M∗(K3,3) is the dual matroid of the graphic matroid corresponding to the complete bipartite graph K3,3. Since M∗(K3,3) is a connected non-modular matroid (it does not have a modular direct summand) this example gives a connected non-modular matroid that has −1 as a root of M. This example and Theorem 6.16 motivate a few questions. Question 6.19. Is there a rank 3 non-modular connected matroid M such that M(M,−1) = 0? Question 6.20. Is there a classification of all matroids whose M polynomial has -1 as a root? Question 6.21. Is there a nice enumerative combinatorial interpretation for M(M,−1) where M is a matroid (i.e. what does it count)? 6.1 No Deletion-Contraction We now show that J and M are not some evaluation of the Tutte polynomial for matroids. We first recall the following definition. Definition 6.22. We say that a function f from matroids to a ring R is a generalized Tutte- Grothendieck invariant (following [4] Sec 1.8.6) if there exists a, b ∈ R such that for every matroid M and element of the ground set e ∈ M f(M) =  f(M\e)f(L) if e is a loop f(M/e)f(C) if e is a coloop af(M\e) + bf(M/e) otherwise. where L is the matroid consisting of exactly one loop and C is the matroid consisting of exactly one coloop. Let Ur,n be the uniform matroid of rank r on n elements and recall that Ur,r ∼= Br are Boolean or free matroids. Then, a direct computation gives J (B1, t) = t+ 1 and J (U2,n, t) = (n− 1)t2 + nt+ n− 1. Hence J(U2,3, t) = 2t2 + 3t + 2. Then any deletion is U2,3\e ∼= B2 and any contraction is U2,3/e ∼= U1,2. Putting this together with Definition 6.22 and assuming that J is a Tutte-Grothendieck invariant 2t2 + 3t+ 2 = a(t2 + 2t+ 1) + b(t+ 1). J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 21 However, this is a contradiction since t+ 1 is not a factor of the left hand side. The same result for M needs two more steps. Looking at the same matroid and using Proposition 6.9 we get M(U2,3, t) = (t2 − 3t+ 1)(t+ 1)2(t− 1)2 = a(t+ 1)2(t− 1)4 + b(t+ 1)(t− 1)2 which reduces to b = (t+ 1)(t2 − 3t+ 1)− a(t+ 1)(t− 1)2. Then we look at U2,4 and again assume M is a Tutte-Grothendieck invariant M(U2,4, t) = (t2−4t+1)(t+1)2(t−1)2 = a(t2−3t+1)(t+1)2(t−1)2+b(t+1)(t−1)2. Inserting the above value for b and reducing we get t2 − 4t+ 1 = a(t2 − 3t+ 1) + (t2 − 3t+ 1)− a(t− 1)2 which gives a = 1 and makes b = −t(t+ 1). But then M(U3,4, t) = (t− 1)(t8 − 3t7 − t6 + 12t5 − 2t4 − 12t3 + 3t2 + 5t− 1) which does not have a factor of t+ 1. This is a contradiction since the right hand side M(U3,4\e, t)− t(t+ 1)M(U3,4/e, t) = M(U3,3, t)− t(t+ 1)M(U2,3, t) does have a t+ 1 factor. 6.2 Valuations Here we study the invariant M over matroid subdivisions. One could focus on a wider range combinatorial objects like posets but we are motived by applications to matroid the- ory. First we recall the basis matroid polytope (using [6] as our general reference for this material). A matroid M can be defined via its set of bases B(M) which are all the inde- pendent sets of M whose size is the rank of M . Then, the matroid polytope of M is P (M) = Conv{eB |B ∈ B(M)} where eB = ei1 + · · ·+ eir with B = {i1, . . . , ir}. Now we need a few key definitions to state our main result. Definition 6.23. A matroid polyhedral subdivision of a matroid polytope P (M) is a col- lection of polyhedra {Pi} such that ⋃ Pi = P (M), each Pi is a matroid polytope whose vertices are vertices of P (M), and for i ̸= j if Pi ⋂ Pj ̸= ∅, then Pi ⋂ Pj is a proper face of both Pi and Pj . Now we want to know how invariants decompose across subdivisions which gives rise to valuations. We will use what is called a weak valuation in [6] but we follow [5] and just say valuation. This makes sense since by Theorem 4.2 in [6] for matroids weak valuations are actually strong valuations. 22 Ars Math. Contemp. 24 (2024) #P2.03 Definition 6.24. Let P be the collection of matroid polytopes and R a commutative ring. A function f : P → R is a (weak) valuation if for any matroid polytope P (M) and any matroid polyhedral subdivision with maximal pieces {P (M1), . . . , P (Mk)} we have that f(∅) = 0 and f(P (M)) = ∑ {j1,...,ji}⊆[k] (−1)if(P (Mj1) ∩ · · · ∩ P (Mji)). Finally we can state the result for the invariant J in terms of valuations. Proposition 6.25. The polynomial J is a valuation on matroids. Proof. Using the decomposition of the J-function given in Theorem 5.2 we know that J (M, t) = (−1)rk(M) ∑ X∈L(M) µ(∅, X)µ(X, 1̂)trk(X) where 1̂ is the maximal flat of M . Hence as a function from the collection of matroids to Z[t] we can represent the function J as J = (±1) ∑ f1 ⋆ f2 where f1 ⋆ f2 = m ◦ (f1 ⊗ f2) ◦ ∆S,T from the notation in Theorem C in [6] and f1 = χM (0) and f2 = χM (0)trk(M). Since f1 and f2 are both Tutte-Grothendieck invariants for matroids and are evaluations of the Tutte polynomial we can conclude that f1 and f2 are both valuations from Proposition 7.5 in [6]. Finally putting it all together Theorem C in [6] finished the result. We conclude with a natural question. The polynomial M(L, t) is slightly more com- plicated but has promising properties that seems to imply it should be a valuation. Question 6.26. Is the polynomial M a matroid valuation? It seems that Proposition 6.10 with Proposition 7.5 and Theorem C in [6] is essentially the proof. However that would use that the characteristic polynomial on the flipped lattice of flats L(M)op is a matroid valuation. References [1] M. Aguiar and F. Ardila, Hopf monoids and generalized permutahedra, 2017, arXiv:1709.075048 [math.CO]. [2] M. Aguiar and S. Mahajan, Topics in Hyperplane Arrangements, volume 226 of Mathematical Surveys and Monographs, American Mathematical Society, Providence, RI, 2017, doi:10.1090/ surv/226, https://doi.org/10.1090/surv/226. [3] G. E. Andrews, q-series: Their Development and Application in Analysis, Nnumber Theory, Combinatorics, Physics, and Computer Algebra, volume 66 of CBMS Regional Conference Series in Mathematics, Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 1986. [4] F. Ardila, Algebraic and geometric methods in enumerative combinatorics, in: Handbook of enumerative combinatorics, CRC Press, Boca Raton, FL, Discrete Math. Appl., pp. 3–172, 2015. J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 23 [5] F. Ardila, A. Fink and F. Rincón, Valuations for matroid polytope subdivisions, Canad. J. Math. 62 (2010), 1228–1245, doi:10.4153/cjm-2010-064-9, https://doi.org/10. 4153/cjm-2010-064-9. [6] F. Ardila and M. Sanchez, Valuations and the Hopf monoid of generalized permutahedra, Int. Math. Res. Not. IMRN (2023), 4149–4224, doi:10.1093/imrn/rnab355, https://doi.org/ 10.1093/imrn/rnab355. [7] C. A. Athanasiadis, Characteristic polynomials of subspace arrangements and finite fields, Adv. Math. 122 (1996), 193–233, doi:10.1006/aima.1996.0059, https://doi.org/10.1006/ aima.1996.0059. [8] J. Bastidas, Species and hyperplane arrangements, Ph.D. thesis, Cornell University, 2021. [9] A. Cameron and A. Fink, The Tutte polynomial via lattice point counting, J. Comb. Theory Ser. A 188 (2022), Paper No. 105584, doi:10.1016/j.jcta.2021.105584, https://doi.org/ 10.1016/j.jcta.2021.105584. [10] T. S. Developers, Sage Mathematics Software (Version 8.1), 2020, http://www. sagemath.org. [11] C. Dupont, A. Fink and L. Moci, Universal Tutte characters via combinatorial coalgebras, Al- gebr. Comb. 1 (2018), 603–651, doi:10.5802/alco, https://doi.org/10.5802/alco. [12] B. Elias, N. Proudfoot and M. Wakefield, The Kazhdan-Lusztig polynomial of a matroid, Adv. Math. 299 (2016), 36–70, doi:10.1016/j.aim.2016.05.005, https://doi.org/10.1016/ j.aim.2016.05.005. [13] P. Hall, A Contribution to the Theory of Groups of Prime-Power Order, Proc. London Math. Soc. (2) 36 (1934), 29–95, doi:10.1112/plms/s2-36.1.29, https://doi.org/10.1112/ plms/s2-36.1.29. [14] G. H. Hardy and E. M. Wright, An Introduction to the Theory of Numbers, Oxford University Press, Oxford, sixth edition, 2008. [15] W. R. Inc., Mathematica, Version 12.3.1, Champaign, IL, 2021, https://www.wolfram. com/mathematica. [16] S. A. Joni and G.-C. Rota, Coalgebras and bialgebras in combinatorics, Stud. Appl. Math. 61 (1979), 93–139, doi:10.1002/sapm197961293, https://doi.org/10.1002/ sapm197961293. [17] R. Jurrius, Relations between Möbius and coboundary polynomials, Math. Comput. Sci. 6 (2012), 109–120, doi:10.1007/s11786-012-0117-6, https://doi.org/10.1007/ s11786-012-0117-6. [18] T. Krajewski, I. Moffatt and A. Tanasa, Hopf algebras and Tutte polynomials, Adv. in Appl. Math. 95 (2018), 271–330, doi:10.1016/j.aam.2017.12.001, https://doi.org/10. 1016/j.aam.2017.12.001. [19] J. L. Martin, Lecture notes on algebraic combinatorics, 2012 . [20] A. F. Möbius, Über eine besondere Art von Umkehrung der Reihen, J. Reine Angew. Math. 9 (1832), 105–123, doi:10.1515/crll.1832.9.105, https://doi.org/10.1515/crll. 1832.9.105. [21] W. Murray, Möbius polynomials, Math. Mag. 85 (2012), 376–383, doi:10.4169/math.mag.85. 5.376, https://doi.org/10.4169/math.mag.85.5.376. [22] P. Orlik and H. Terao, Arrangements of Hyperplanes, volume 300 of Grundlehren der Math- ematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], Springer- Verlag, Berlin, 1992, doi:10.1007/978-3-662-02772-1, https://doi.org/10.1007/ 978-3-662-02772-1. 24 Ars Math. Contemp. 24 (2024) #P2.03 [23] J. Oxley, Matroid Theory, volume 21 of Oxford Graduate Texts in Mathematics, Oxford Uni- versity Press, Oxford, 2nd edition, 2011, doi:10.1093/acprof:oso/9780198566946.001.0001, https://doi.org/10.1093/acprof:oso/9780198566946.001.0001. [24] P. Paule and A. Riese, A Mathematica q-analogue of Zeilberger’s algorithm based on an alge- braically motivated approach to q-hypergeometric telescoping, in: Special functions, q-series and related topics (Toronto, ON, 1995), Amer. Math. Soc., Providence, RI, volume 14 of Fields Inst. Commun., pp. 179–210, 1997. [25] G. Pilz, Near-Rings, volume 23 of North-Holland Mathematics Studies, North-Holland Pub- lishing Co., Amsterdam, 2nd edition, 1983, the theory and its applications. [26] G.-C. Rota, On the foundations of combinatorial theory. I. Theory of Möbius functions, Z. Wahrscheinlichkeitstheorie und Verw. Gebiete 2 (1964), 340–368 (1964), doi:10.1007/ bf00531932, https://doi.org/10.1007/bf00531932. [27] E. Spiegel and C. J. O’Donnell, Incidence Algebras, volume 206 of Monographs and Textbooks in Pure and Applied Mathematics, Marcel Dekker, Inc., New York, 1997. [28] R. P. Stanley, Enumerative Combinatorics. Volume 1, volume 49 of Cambridge Studies in Ad- vanced Mathematics, Cambridge University Press, Cambridge, 2nd edition, 2012. [29] H. Terao, Free arrangements of hyperplanes and unitary reflection groups, Proc. Japan Acad. Ser. A Math. Sci. 56 (1980), 389–392, http://projecteuclid.org/euclid.pja/ 1195516722. [30] M. Wakefield, Partial flag incidence algebras, 2022, arXiv:1605.01685 [math.CO]. [31] L. Weisner, Abstract theory of inversion of finite series, Trans. Am. Math. Soc. 38 (1935), 474–484, doi:10.2307/1989808, https://doi.org/10.2307/1989808. [32] D. J. A. Welsh, Matroid Theory, Academic Press [Harcourt Brace Jovanovich, Publishers], London-New York, 1976, l. M. S. Monographs, No. 8. [33] T. Zaslavsky, Facing up to arrangements: face-count formulas for partitions of space by hyper- planes, Mem. Am. Math. Soc. 1 (1975), vii+102. [34] T. Zaslavsky, The Möbius function and the characteristic polynomial, in: Combinatorial Ge- ometries, Cambridge Univ. Press, Cambridge, volume 29 of Encyclopedia Math. Appl., pp. 114–138, 1987.