This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A Categorical Approach to Möbius Inversion via Derived Functors

Alex Elchesen Department of Mathematics, Colorado State University Amit Patel Department of Mathematics, Colorado State University
Abstract

We develop a cohomological approach to Möbius inversion using derived functors in the enriched categorical setting. For a poset PP and a closed symmetric monoidal abelian category 𝒞\mathcal{C}, we define Möbius cohomology as the derived functors of an enriched hom functor on the category of PP-modules. We prove that the Euler characteristic of our cohomology theory recovers the classical Möbius inversion, providing a natural categorification. As a key application, we prove a categorical version of Rota’s Galois Connection. Our approach unifies classical ideas from combinatorics with homological algebra.

1 Introduction

Möbius inversion is a fundamental principle in combinatorics, generalizing the classical inclusion-exclusion principle to partially ordered sets. Originally emerging from number theory in the 19th century, the concept was later vastly generalized by Gian-Carlo Rota [14], who recognized its deeper structural role in combinatorics. At its core, Möbius inversion provides a systematic way to recover local information from global summaries. Given a function f:Pf:P\to\mathcal{R} from a finite poset PP to a commutative ring \mathcal{R}, its upper Möbius inversion is the unique function +f:P\partial_{+}f:P\to\mathcal{R} such that for all aPa\in P,

f(a)=b:ab+f(b).f(a)=\sum_{b:a\leq b}\partial_{+}f(b).

Dually, its lower Möbius inversion is the unique function f:P\partial_{-}f:P\to\mathcal{R} such that for all bPb\in P,

f(b)=a:abf(a).f(b)=\sum_{a:a\leq b}\partial_{-}f(a).

Our entry into Möbius inversion stems from its fundamental role in persistent homology, a key tool in topological data analysis [12, 8]. In pursuit of generalizations, Patel and Skraba developed Möbius homology for poset modules M:P𝒞M:P\to\mathcal{C} (functors valued in an abelian category 𝒞\mathcal{C}[13]. Their theory associates a homology object HdM(a)H_{d}^{\downarrow}M(a) to every aPa\in P and index dd such that its Euler characteristic recovers the lower Möbius inversion of the dimension function m:P(𝒞)m:P\to\mathcal{R}(\mathcal{C}). They construct these homology objects via a simplicial cosheaf over the order complex of PP.

In this paper, we develop a cohomological approach to Möbius inversion through derived functors. Rather than dualizing the construction of Patel and Skraba to simplicial sheaves, we provide a more general and functorial perspective through enriched hom functors. Specifically, given a point aPa\in P, we define the indicator module 𝟏a{\mathbf{1}}_{a}. When 𝒞\mathcal{C} is closed monoidal, we show that the set of natural transformations Nat(𝟏a,M)\mathrm{Nat}({\mathbf{1}}_{a},M) forms an object Hom¯(𝟏a,M)\underline{\mathrm{Hom}}({\mathbf{1}}_{a},M) of 𝒞\mathcal{C} (Definition 3.5). The functor Hom¯(𝟏a,):𝒞P𝒞\underline{\mathrm{Hom}}({\mathbf{1}}_{a},-):\mathcal{C}^{P}\to\mathcal{C} is left-exact, and we define its right-derived functors Ext¯d(𝟏a,M)\underline{\mathrm{Ext}}^{d}({\mathbf{1}}_{a},M) as our Möbius cohomology objects.

This derived functor approach offers several key advantages. First, it generalizes naturally to handle cohomology for arbitrary spreads and modules, extending beyond the traditional scope of Möbius inversion. Second, it reveals connections with classical homological algebra, allowing us to leverage established theoretical machinery. Our framework culminates in a categorical version of Rota’s Galois Connection Theorem, expressed as an enriched adjunction.

Previous Work

The study of the total cohomology of PP-modules using derived functors has a rich history, with foundational contributions by Deheuvels [5]. Baclawski made significant advances in understanding the homological aspects of posets through two key papers. His 1975 work [1] developed a homological interpretation of Whitney numbers of geometric lattices, while his 1977 paper [2] established deep connections between Galois connections and spectral sequences. While this body of work laid important groundwork, the explicit connection to Möbius inversions remained unexplored until the recent work of Patel and Skraba [13].

A parallel development in the categorical treatment of Möbius inversion emerged through the work of Leroux [9], who introduced the concept of Möbius categories. This framework was significantly expanded by Content, Lemay, and Leroux [3], who developed a comprehensive categorical setting for Möbius inversion. While their work established important functorial properties of incidence algebras and Möbius functions, it did not pursue the homological aspects that we develop here.

Our investigation originates from questions in persistent homology, a fundamental tool in topological data analysis. While we do not explicitly address persistent homology in this paper, the discussion of persistent homology in [13] dualizes to our setting. In this context, our Möbius cohomology theory shares important connections with recent work by Oudot and Scoccola [11] on bigraded Betti numbers.

Recent work by Gülen and McCleary [7] demonstrated how Rota’s Galois connection theorem reveals the functorial character of Möbius inversion. Our Theorem 6.6 substantially generalizes their observation, providing a comprehensive categorical framework that unifies the combinatorial and functorial perspectives on Möbius inversion.

Our contributions

The main contributions of this paper are:

  1. 1.

    We develop a cohomological theory for Möbius inversion using derived functors in the enriched categorical setting (Section 5). We prove that the Euler characteristic of our cohomology theory coincides with the classical Möbius inversion (Theorem 5.7), providing a natural categorification while simultaneously offering a more general framework for studying poset modules.

  2. 2.

    We establish an explicit formula for computing Möbius cohomology using the standard cofree resolution (Section 4). Specifically, we show that applying Hom¯(𝟏a,)\underline{\mathrm{Hom}}({\mathbf{1}}_{a},-) to the standard cofree resolution (Equation 2) yields a cochain complex (Equation 5) whose cohomology groups recover Möbius cohomology. This formula reveals how our derived functor approach recovers the construction using sheaves on the order complex.

  3. 3.

    We prove a categorical version of Rota’s Galois Connection Theorem (Theorem 6.6) that implies the classical result (Theorem 6.7). This generalization takes the form of an enriched adjunction between the pushforward and pullback functors induced by a Galois connection (Section 6).

  4. 4.

    We demonstrate that our derived functor approach unifies various perspectives on Möbius inversion through several key isomorphisms (Proposition 3.8, Proposition 4.8, and Theorem 6.10), connecting classical combinatorial methods, homological algebra, and categorical techniques in a single coherent framework.

Throughout, we have made an effort to keep the exposition self-contained, providing background material and motivation for key concepts in Section 2. This makes the paper accessible to readers with a basic understanding of category theory and homological algebra.

Outline

The paper is organized as follows. Section 2 establishes the categorical preliminaries, including key concepts from homological algebra and enriched category theory. Section 3 introduces poset modules and develops the fundamental constructions needed for our cohomology theory. Section 4 defines derived hom functors and explores their properties in the context of poset modules. Section 5 introduces Möbius cohomology and establishes its relationship to classical Möbius inversion. Finally, Section 6 examines Galois connections between posets and proves our categorical version of Rota’s theorem.

2 Categorical Preliminaries

In this section, we introduce the key categorical concepts, constructions, and notation used throughout the paper. These preliminaries provide the necessary framework for later sections, focusing on abelian and monoidal categories, ends, and basic homological algebra. For a deeper study, see [10, 16].

2.1 Adjunctions

For a category 𝒞\mathcal{C}, we denote by Hom𝒞(A,B)\mathrm{Hom}_{\mathcal{C}}(A,B) the set of morphisms from AA to BB. Recall that an adjunction between two categories 𝒞\mathcal{C} and 𝒟\mathcal{D} consists of a pair of functors F:𝒞𝒟F:\mathcal{C}\to\mathcal{D} and G:𝒟𝒞G:\mathcal{D}\to\mathcal{C}, together with natural isomorphisms of sets Φ:Hom𝒟(F(A),B)Hom𝒞(A,G(B))\Phi:\mathrm{Hom}_{\mathcal{D}}(F(A),B)\cong\mathrm{Hom}_{\mathcal{C}}(A,G(B)). In this situation, we say FF is left adjoint to GG and GG is right adjoint to FF, denoted FGF\dashv G. An important property of left adjoints is that they are cocontinuous, while right adjoints are continuous, meaning left adjoints preserve colimits and right adjoints preserve limits.

2.2 Abelian Categories

A category 𝒞\mathcal{C} is preadditive if for every set Hom𝒞(A,B)\mathrm{Hom}_{\mathcal{C}}(A,B) there is an abelian group structure, and composition is bilinear, i.e.,

f(g+h)=(fg)+(fh),(f+g)h=(fh)+(gh).f\circ(g+h)=(f\circ g)+(f\circ h),\quad(f+g)\circ h=(f\circ h)+(g\circ h).

A functor F:𝒞𝒟F:\mathcal{C}\to\mathcal{D} between preadditive categories is additive if for all objects A,BA,B, the induced map Hom𝒞(A,B)Hom𝒟(F(A),F(B))\mathrm{Hom}_{\mathcal{C}}(A,B)\to\mathrm{Hom}_{\mathcal{D}}(F(A),F(B)) is a group homomorphism.

A biproduct of objects A1,,AnA_{1},\ldots,A_{n} in a preadditive category 𝒞\mathcal{C} is an object A1AnA_{1}\oplus\cdots\oplus A_{n} with morphisms πk:A1AnAk\pi_{k}:A_{1}\oplus\cdots\oplus A_{n}\to A_{k} and ιk:AkA1An\iota_{k}:A_{k}\to A_{1}\oplus\cdots\oplus A_{n} satisfying:

πkιk=idAk,πlιk=0 for kl.\pi_{k}\circ\iota_{k}={\mathrm{id}}_{A_{k}},\quad\pi_{l}\circ\iota_{k}=0\text{ for }k\neq l.

In this case, (A1An,πk)\big{(}A_{1}\oplus\cdots\oplus A_{n},\pi_{k}\big{)} is a product, and (A1An,ιk)\big{(}A_{1}\oplus\cdots\oplus A_{n},\iota_{k}\big{)} is a coproduct. A preadditive category is additive if it admits all finite biproducts.

A preadditive category 𝒞\mathcal{C} is abelian if it has a zero object, all binary biproducts, kernels and cokernels, and every monomorphism and epimorphism is normal. Functors between abelian categories are typically additive, and if FGF\dashv G, both FF and GG are automatically additive.

Exactness is central in abelian categories. An additive functor F:𝒞𝒟F:\mathcal{C}\to\mathcal{D} between abelian categories is:

  • left-exact if it preserves exactness of sequences of the form

    0ABC,0\to A\to B\to C,
  • right-exact if it preserves exactness of sequences of the form

    ABC0.A\to B\to C\to 0.

A functor is exact if it is both left-exact and right-exact. If FGF\dashv G, then FF is automatically right-exact and GG is left-exact.

2.3 Closed Monoidal Categories

A monoidal category consists of a category 𝒞\mathcal{C}, a bifunctor :𝒞×𝒞𝒞\otimes:\mathcal{C}\times\mathcal{C}\to\mathcal{C} (called the tensor product), an object 𝟏{\mathbf{1}} (the unit object), and natural isomorphisms:

αA,B,C:(AB)CA(BC),λA:𝟏AA,ρA:A𝟏A.\alpha_{A,B,C}:(A\otimes B)\otimes C\cong A\otimes(B\otimes C),\quad\lambda_{A}:{\mathbf{1}}\otimes A\cong A,\quad\rho_{A}:A\otimes{\mathbf{1}}\cong A.

These isomorphisms satisfy certain coherence conditions (see [10, VII]). A monoidal category is symmetric if there is a natural isomorphism sA,B:ABBAs_{A,B}:A\otimes B\cong B\otimes A that satisfies its own coherence conditions. A (symmetric) monoidal category (𝒞,,𝟏)(\mathcal{C},\otimes,{\mathbf{1}}) is closed if for each A𝒞A\in\mathcal{C}, the functor A:𝒞𝒞-\otimes A:\mathcal{C}\to\mathcal{C} has a right adjoint. This right adjoint is denoted om𝒞(A,):𝒞𝒞{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(A,-):\mathcal{C}\to\mathcal{C} and is called the internal hom.

Proposition 2.1:

For any closed symmetric monoidal category (𝒞,,𝟏)(\mathcal{C},\otimes,{\mathbf{1}}) and for all objects BB in 𝒞\mathcal{C}, we have om𝒞(𝟏,B)B.{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}({\mathbf{1}},B)\cong B.

Proof.

For any A𝒞A\in\mathcal{C}, we have Hom𝒞(A,om𝒞(𝟏,B))Hom𝒞(A𝟏,B)Hom𝒞(A,B)\mathrm{Hom}_{\mathcal{C}}(A,{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}({\mathbf{1}},B))\cong\mathrm{Hom}_{\mathcal{C}}(A\otimes{\mathbf{1}},B)\cong\mathrm{Hom}_{\mathcal{C}}(A,B). The result follows now from the Yoneda lemma. ∎

2.4 Abelian Symmetric Monoidal Categories

We will focus on categories that are both abelian and closed symmetric monoidal. A key compatibility condition between the abelian and monoidal structures is that the tensor product must be additive in both variables. When the monoidal structure is closed, this compatibility condition is automatically satisfied because biproducts serve as both limits and colimits, and the tensor product is left adjoint to the internal hom functor.

Henceforth, 𝒞\mathcal{C} is a small, complete, abelian, and closed symmetric monoidal category.

2.5 Ends

Fix a functor S:𝒞op×𝒞𝒟S:\mathcal{C}^{\mathrm{op}}\times\mathcal{C}\to\mathcal{D}. A wedge of SS, denoted e:WSe:W\to S, is an object W𝒟W\in\mathcal{D} together with maps eA:WS(A,A)e_{A}:W\to S(A,A) such that for every f:ABf:A\to B, the diagram

W\displaystyle{W}S(B,B)\displaystyle{S(B,B)}S(A,A)\displaystyle{S(A,A)}S(B,A)\displaystyle{S(B,A)}eB\scriptstyle{e_{B}}eA\scriptstyle{e_{A}}S(B,f)\scriptstyle{S(B,f)}S(f,B)\scriptstyle{S(f,B)}

commutes. The end of SS, denoted AS(A,A)\int_{A}S(A,A), is the universal wedge, meaning any wedge e:WSe:W\to S factors uniquely through it. That is, if e:WSe:W\to S is any wedge of SS, then there is a unique map ϕ:WAS(A,A)\phi:W\to\int_{A}S(A,A) satisfying πAϕ=eA\pi_{A}\circ\phi=e_{A}, for all objects A𝒞A\in\mathcal{C}.

The end of an ordinary functor T:𝒞𝒟T:\mathcal{C}\to\mathcal{D}, denoted AT(A)\int_{A}T(A), is defined by taking the end of the composition S=𝒞op×𝒞p𝒞T𝒟S=\mathcal{C}^{\mathrm{op}}\times\mathcal{C}\stackrel{{\scriptstyle p}}{{\to}}\mathcal{C}\stackrel{{\scriptstyle T}}{{\to}}\mathcal{D}, where pp denotes the projection functor onto the second factor. In this case, the end reduces to the limit of TT.

Given functors F,G:𝒞𝒟F,G:\mathcal{C}\to\mathcal{D}, the set of natural transformations Nat(F,G)\mathrm{Nat}(F,G) from FF to GG is given (up to isomorphism) by the formula

Nat(F,G)AHom𝒟(F(A),G(A)).\mathrm{Nat}(F,G)\cong\int_{A}\mathrm{Hom}_{\mathcal{D}}(F(A),G(A)). (1)

This formula and generalizations of it are central to our constructions. If 𝒟\mathcal{D} is a closed category, then we can replace Hom𝒟(F(A),G(A))\mathrm{Hom}_{\mathcal{D}}(F(A),G(A)) in the formula above by om𝒟(F(A),G(A)){\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{D}}(F(A),G(A)) to make Nat(F,G)\mathrm{Nat}(F,G) an object of 𝒟\mathcal{D} (see Definition 3.5 below).

2.6 Injective Resolutions

An object II in 𝒞\mathcal{C} is called injective if, for every monomorphism f:ABf:A\hookrightarrow B and any morphism g:AIg:A\to I, there exists a morphism h:BIh:B\to I making the following diagram commute:

A\displaystyle{A}B\displaystyle{B}I\displaystyle{I}f\scriptstyle{f}g\scriptstyle{g}h\scriptstyle{h}

Arbitrary products of injective objects are injective. Moreover, in an abelian category, finite direct sums (i.e., biproducts) of injective objects remain injective since finite sums coincide with products. However, arbitrary direct sums do not necessarily preserve injectivity.

An abelian category 𝒞\mathcal{C} is said to have enough injectives if every object AA admits a monomorphism AIA{\hookrightarrow}I into an injective object II.

An injective resolution of an object XX in 𝒞\mathcal{C} is an exact sequence

0XI0I1I20\to X\hookrightarrow I^{0}\to I^{1}\to I^{2}\to\cdots

where each IdI^{d} is an injective object in 𝒞\mathcal{C}.

If 𝒞\mathcal{C} has enough injectives, every object XX admits an injective resolution via the standard construction. Begin by embedding XX into an injective object I0I_{0}, and successively embed each cokernel into a new injective object Ii+1I_{i+1} as follows:

X\displaystyle{X}I0\displaystyle{I_{0}}cokerϵ0\displaystyle{{\mathrm{coker}}\,\epsilon_{0}}I1\displaystyle{I_{1}}cokerϵ1\displaystyle{{\mathrm{coker}}\,\epsilon_{1}}I2\displaystyle{I_{2}}cokerϵ2\displaystyle{{\mathrm{coker}}\,\epsilon_{2}}\displaystyle{\cdots}ϵ0\scriptstyle{\epsilon_{0}}ϵ1\scriptstyle{\epsilon_{1}}ϵ2\scriptstyle{\epsilon_{2}}ϵ3\scriptstyle{\epsilon_{3}}

2.7 Grothendieck Ring

The Grothendieck ring of a category 𝒞\mathcal{C}, denoted (𝒞)\mathcal{R}(\mathcal{C}), is the free abelian group generated by the isomorphism classes [A][A] of objects AA in 𝒞\mathcal{C}, subject to the relation [B]=[A]+[C][B]=[A]+[C] for every short exact sequence 0ABC00\to A\to B\to C\to 0 in 𝒞\mathcal{C}. The ring structure on (𝒞)\mathcal{R}(\mathcal{C}) is induced by the monoidal structure of 𝒞\mathcal{C}, where the product is given by [A][B]=[AB][A]\cdot[B]=[A\otimes B], with \otimes denoting the tensor product in 𝒞\mathcal{C}. The resulting ring (𝒞)\mathcal{R}(\mathcal{C}) is commutative and unital, with the class of the unit object [𝟏][\mathbf{1}] acting as the multiplicative identity: [A][𝟏]=[A][A]\cdot[\mathbf{1}]=[A]. The multiplication is both commutative and associative, reflecting the symmetric monoidal structure of 𝒞\mathcal{C}.

3 Poset Modules

In this section, we introduce poset modules, which are functors from a poset (viewed as a category) to the target category 𝒞\mathcal{C}. We establish key definitions and notation that will be used throughout the paper, focusing on constructions such as cofree modules, enriched hom-objects, tensor products, and injective resolutions.

3.1 First Constructions

We begin by defining posets and their interpretation as categories, along with the notion of PP-modules. Principal cofree modules and cofree modules are introduced as building blocks for working with PP-modules.

Definition 3.1:

A poset PP consists of a set PP equipped with a binary relation \leq that is reflexive, antisymmetric, and transitive. A poset can be regarded as a category, where the objects are the elements of PP and where there is a morphism aba\to b whenever aba\leq b.

Henceforth, we assume all posets are finite.

Definition 3.2:

A PP-module valued in a category 𝒞\mathcal{C} is a functor M:P𝒞M:P\to\mathcal{C}. Let 𝒞P\mathcal{C}^{P} denote the category of PP-modules, with morphisms given by natural transformations. Denote by Nat(M,N)\mathrm{Nat}(M,N) the set of morphisms from MM to NN.

Definition 3.3:

For aPa\in P and X𝒞X\in\mathcal{C}, define the PP-module XaX^{\downarrow a} by

Xa(b):={Xif ba,0otherwise.X^{\downarrow a}(b):=\begin{cases}X&\text{if }b\leq a,\\ 0&\text{otherwise}.\end{cases}

The map Xa(bc)X^{\downarrow a}(b\leq c) is the identity on XX if bcb\leq c, and 0 otherwise. Such PP-modules are called principal cofree modules.

Definition 3.4:

A PP-module MM is cofree if it is isomorphic to a product of principal cofree modules, i.e., if MaPXaaM\cong\prod_{a\in P}X_{a}^{\downarrow a} for objects Xa𝒞X_{a}\in\mathcal{C}.

Having introduced principal cofree modules, we now shift our focus to enriched hom-objects and tensor product constructions for PP-modules.

3.2 Enriched Hom

This subsection defines enriched hom-objects and describes how the symmetric monoidal structure extends from the base category to the category of PP-modules. We introduce tensor products of PP-modules and establish an adjunction between the enriched hom functor and the tensor product functor.

If 𝒞\mathcal{C} is an abelian category, then the category of PP-modules, 𝒞P\mathcal{C}^{P}, is also abelian. Moreover, if (𝒞,,𝟏)(\mathcal{C},\otimes,{\mathbf{1}}) is a symmetric monoidal category, then (𝒞P,,𝟏)(\mathcal{C}^{P},\otimes,{\mathbf{1}}) inherits a symmetric monoidal structure as follows: for two PP-modules MM and NN, their tensor product MNM\otimes N is given by the composition

P\displaystyle{P}𝒞×𝒞\displaystyle{\mathcal{C}\times\mathcal{C}}𝒞.\displaystyle{\mathcal{C}.}M×N\scriptstyle{M\times N}\scriptstyle{\otimes}

The unit PP-module, denoted 𝟏:P𝒞{\mathbf{1}}:P\to\mathcal{C}, assigns the object 𝟏{\mathbf{1}} to each element of PP, with identity morphisms acting between them. The coherence conditions for this structure follow directly from those in 𝒞\mathcal{C}.

While we do not require 𝒞P\mathcal{C}^{P} to be closed, we do require it to be enriched over 𝒞\mathcal{C}. Specifically, for each PP-module MM, there must be a functor

Hom¯(M,):𝒞P𝒞\underline{\mathrm{Hom}}(M,-):\mathcal{C}^{P}\to\mathcal{C}

that is right adjoint to a functor of the form A:𝒞𝒞P-\otimes A:\mathcal{C}\to\mathcal{C}^{P}, for each object A𝒞A\in\mathcal{C}.

Definition 3.5:

Define the enriched hom-object between two PP-modules MM and NN as

Hom¯(M,N):=aPom𝒞(M(a),N(a)).\underline{\mathrm{Hom}}(M,N):=\int_{a\in P}{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(a),N(a)).

This assignment extends, by the universal property of the end, to a functor

Hom¯(,):(𝒞P)op×𝒞P𝒞.\underline{\mathrm{Hom}}(-,-):(\mathcal{C}^{P})^{\mathrm{op}}\times\mathcal{C}^{P}\to\mathcal{C}.

Before proving the enriched hom adjunction, we introduce constant PP-modules. For an object A𝒞A\in\mathcal{C}, denote by AP:P𝒞A^{P}:P\to\mathcal{C} as the AA-constant PP-module. That is AP(a)=AA^{P}(a)=A, for all aPa\in P, all the morphisms are set to the identity morphism on AA.

Definition 3.6:

Define the tensor product M:𝒞𝒞P-\otimes M:\mathcal{C}\to\mathcal{C}^{P} as the functor induced by sending every object A𝒞A\in\mathcal{C} to the tensor product of PP-modules APMA^{P}\otimes M.

Proposition 3.7:

The tensor product M:𝒞𝒞P-\otimes M:\mathcal{C}\to\mathcal{C}^{P} is left-adjoint to the enriched-hom Hom¯(M,):𝒞P𝒞\underline{\mathrm{Hom}}(M,-):\mathcal{C}^{P}\to\mathcal{C}. That is, for every object A𝒞A\in\mathcal{C} and PP-modules MM and NN, there is a natural isomorphism of sets

Nat(APM,N)Hom𝒞(A,Hom¯(M,N)).\mathrm{Nat}\big{(}A^{P}\otimes M,N\big{)}\cong\mathrm{Hom}_{\mathcal{C}}\big{(}A,\underline{\mathrm{Hom}}(M,N)\big{)}.
Proof.

For each M,N𝒞PM,N\in\mathcal{C}^{P} and A𝒞A\in\mathcal{C}, we have

Nat(APM,N)\displaystyle\mathrm{Nat}\big{(}A^{P}\otimes M,N\big{)} =aPHom𝒞(AM(a),N(a))\displaystyle=\int_{a\in P}\mathrm{Hom}_{\mathcal{C}}\big{(}A\otimes M(a),N(a)\big{)} by definition of APMA^{P}\otimes M and by (1)
aPHom𝒞(A,om𝒞(M(a),N(a))\displaystyle\cong\int_{a\in P}\mathrm{Hom}_{\mathcal{C}}\big{(}A,{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(a),N(a)\big{)} by the hom-tensor adjunction in 𝒞\mathcal{C}
Hom𝒞(A,aPom𝒞(M(a),N(a)))\displaystyle\cong\mathrm{Hom}_{\mathcal{C}}\left(A,\int_{a\in P}{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}\big{(}M(a),N(a)\big{)}\right) since Hom\mathrm{Hom} commutes with limits
Hom𝒞(A,Hom¯(M,N)).\displaystyle\cong\mathrm{Hom}_{\mathcal{C}}\big{(}A,\underline{\mathrm{Hom}}(M,N)\big{)}.

These isomorphisms are natural in AA, MM, and NN which completes the proof. ∎

Proposition 3.8:

If NaPXaaN\cong\prod_{a\in P}X_{a}^{\downarrow a} is a cofree PP-module and MM is any PP-module, then

Hom¯(M,N)aPom𝒞(M(a),Xa).\underline{\mathrm{Hom}}(M,N)\cong\prod_{a\in P}{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}\big{(}M(a),X_{a}\big{)}.
Proof.

Since the functor Hom¯(M,)\underline{\mathrm{Hom}}(M,-) is a right adjoint, it preserves limits, and in particular, it commutes with products. Therefore, it suffices to demonstrate that for every object X𝒞X\in\mathcal{C} and element aPa\in P, there is an isomorphism

Hom¯(M,Xa)om𝒞(M(a),X).\underline{\mathrm{Hom}}(M,X^{\downarrow a})\cong{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(a),X).

Consider the case where N=XaN=X^{\downarrow a}. For each pair bcb\leq c in PP, examine the following commutative diagram:

om𝒞(M(a),X)\displaystyle{{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(a),X)}om𝒞(M(c),Xa(c))\displaystyle{{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(c),X^{\downarrow a}(c))}om𝒞(M(b),Xa(b))\displaystyle{{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(b),X^{\downarrow a}(b))}om𝒞(M(b),Xa(c))\displaystyle{{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(b),X^{\downarrow a}(c))}ϕc\scriptstyle{\phi_{c}}ϕb\scriptstyle{\phi_{b}}

Here, for each pPp\in P, the morphism ϕp:om𝒞(M(a),X)om𝒞(M(p),Xa(p))\phi_{p}:{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(a),X)\to{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(p),X^{\downarrow a}(p)) is defined by

ϕp={om𝒞(M(ap),X)if pa,0otherwise.\phi_{p}=\begin{cases}{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(a\leq p),X)&\text{if }p\leq a,\\ 0&\text{otherwise}.\end{cases}

It is straightforward to verify that this diagram commutes for every bcb\leq c in PP. Consequently, the collection (om𝒞(M(a),X),ϕp)pP\left({\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(a),X),\phi_{p}\right)_{p\in P} forms a wedge for the functor om𝒞(M(),Xa()){\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(-),X^{\downarrow a}(-)).

Now, let LL be any other such wedge with morphisms ep:Lom𝒞(M(p),Xa(p))e_{p}:L\to{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(p),X^{\downarrow a}(p)) for each pPp\in P. In particular, there is a morphism ea:Lom𝒞(M(a),X)e_{a}:L\to{\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(a),X) corresponding to p=ap=a. It can be checked that for every pPp\in P, the following equality holds: ϕpea=ep.\phi_{p}\circ e_{a}=e_{p}. This demonstrates that the morphisms epe_{p} uniquely factor through om𝒞(M(a),X){\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(a),X). Therefore, om𝒞(M(a),X){\mathcal{H}\kern-0.5pt\mathrm{om}}_{\mathcal{C}}(M(a),X) is the universal wedge. ∎

3.3 Injective Resolutions

We define principal injective modules and elementary injective modules, which are necessary for constructing injective resolutions for PP-modules. This section parallels the theory of injective resolutions in the underlying abelian category.

Definition 3.9:

For an injective object Q𝒞Q\in\mathcal{C} and cPc\in P, define the PP-module QcQ^{\downarrow c} by:

Qc(b):={Qif bc,0otherwise.Q^{\downarrow c}(b):=\begin{cases}Q&\text{if }b\leq c,\\ 0&\text{otherwise}.\end{cases}

The map Qc(ab)Q^{\downarrow c}(a\leq b) is the identity on QQ for bcb\leq c, and 0 otherwise. Such PP-modules are called principal injective PP-modules.

Definition 3.10:

A PP-module MM is called an elementary injective PP-module if it is isomorphic to a product of principal injective PP-modules. That is, MaPQaaM\cong\prod_{a\in P}Q_{a}^{\downarrow a}, where each QaQ_{a} is an injective object in 𝒞\mathcal{C}.

Clearly, principal injective modules are principal cofree modules, and elementary injective modules are cofree. With the concept of elementary injective modules in hand, we now establish that every elementary injective module is injective in the category of PP-modules.

Proposition 3.11:

Every elementary injective PP-module is injective in 𝒞P\mathcal{C}^{P}.

Proof.

Since products of injective objects are injective, it suffices to prove the proposition for a principal injective module QbQ^{\downarrow b}. Let MM and NN be PP-modules, ϕ:MN\phi:M\hookrightarrow N a monomorphism, and ψ:MQb\psi:M\to Q^{\downarrow b} any morphism. We construct a map of PP-modules μ:NQb\mu:N\to Q^{\downarrow b} such that ϕμ=ψ\phi\circ\mu=\psi. Since Qb(b)=QQ^{\downarrow b}(b)=Q is injective, there exists a map μ~b:N(b)Q\tilde{\mu}_{b}:N(b)\to Q in 𝒞\mathcal{C} such that the diagram

M(b)\displaystyle{M(b)}N(b)\displaystyle{N(b)}Q\displaystyle{Q}ϕb\scriptstyle{\phi_{b}}ψb\scriptstyle{\psi_{b}}μ~b\scriptstyle{\tilde{\mu}_{b}}

commutes. Define μ:NQb\mu:N\to Q^{\downarrow b} by setting μb=μ~b\mu_{b}=\tilde{\mu}_{b} and then extending to all of PP as follows:

μa:={μbN(ab)if ab,0otherwise.\mu_{a}:=\begin{cases}\mu_{b}\circ N(a\leq b)&\textup{if }a\leq b,\\ 0&\textup{otherwise}.\end{cases}

It is straightforward to verify that μ\mu is a natural transformation and that ϕμ=ψ\phi\circ\mu=\psi. ∎

3.4 Enough Injectives

We show that if the base category has enough injectives, then the category of PP-modules inherits this property, enabling us to construct injective resolutions for all PP-modules.

Proposition 3.12:

If 𝒞\mathcal{C} has enough injectives, then the category 𝒞P\mathcal{C}^{P} has enough injectives.

Proof.

We show that 𝒞P\mathcal{C}^{P} has enough injectives by constructing, for a fixed PP-module MM, an injective PP-module QQ and a monomorphism j:MQj:M{\hookrightarrow}Q.

Since 𝒞\mathcal{C} has enough injectives, for each aPa\in P, there exists an injective object Qa𝒞Q_{a}\in\mathcal{C} and a monomorphism ia:M(a)Qai_{a}:M(a){\hookrightarrow}Q_{a}. Set

Q:=aPQaa.Q:=\prod_{a\in P}Q_{a}^{\downarrow a}.

By Proposition 3.11, QQ is an elementary injective PP-module and hence is injective.

For each bPb\in P, define a map jb:MQbj_{b}:M\to Q^{\downarrow b} by

jb(a)={ibM(ab)if ab,0otherwise.j_{b}(a)=\begin{cases}i_{b}\circ M(a\leq b)&\text{if }a\leq b,\\ 0&\text{otherwise}.\end{cases}

It is straightforward to verify that each jbj_{b} is a natural transformation. Note that jb(b)j_{b}(b) is a monomorphism. Finally, the desired map j:MQj:M{\hookrightarrow}Q is obtained using the universal property of the product. ∎

By Proposition 3.12, every PP-module admits a monomorphism into an elementary injective module, leading to the following corollary.

Corollary 3.13:

If 𝒞\mathcal{C} is an abelian category with enough injectives, then every PP-module M𝒞PM\in\mathcal{C}^{P} has an injective resolution by elementary injective PP-modules.

4 Derived Hom Functors

In this section, we introduce the Ext-functor as a derived functor of the enriched hom. Next, we introduce cofree resolutions as a computational tool. Finally, we explore the Euler characteristic of Ext-functors.

4.1 Ext Functors and Injective Resolutions

This subsection introduces Ext-functors, which arise as derived functors of the enriched hom.

Recall Proposition 3.7: for a fixed PP-module MM, the tensor product functor

M:𝒞𝒞P-\otimes M:\mathcal{C}\to\mathcal{C}^{P}

is left adjoint to the enriched hom functor

Hom¯(M,):𝒞P𝒞.\underline{\mathrm{Hom}}(M,-):\mathcal{C}^{P}\to\mathcal{C}.

This adjunction makes Hom¯(M,)\underline{\mathrm{Hom}}(M,-) a left-exact functor between the two abelian categories.

Now, let NN be a second PP-module. Consider an injective resolution

0NI0I1I2,0\to N\hookrightarrow I^{0}\to I^{1}\to I^{2}\to\cdots,

and apply the functor Hom¯(M,)\underline{\mathrm{Hom}}(M,-) to obtain the following cochain complex in 𝒞\mathcal{C}:

0Hom¯(M,I0)Hom¯(M,I1)Hom¯(M,I2).0\to\underline{\mathrm{Hom}}(M,I^{0})\to\underline{\mathrm{Hom}}(M,I^{1})\to\underline{\mathrm{Hom}}(M,I^{2})\to\cdots.

The cohomology objects of this complex, denoted by RdHom¯(M,N)R^{d}\underline{\mathrm{Hom}}(M,N), are the higher derived functors of the enriched hom and are commonly referred to as Ext-functors. These are denoted by

Ext¯d(M,N).\underline{\mathrm{Ext}}^{d}(M,N).

We are particularly interested in the Ext-functor for a specific class of indicator PP-modules MM, which we now describe.

Definition 4.1:

A subposet ZPZ\subset P is called a spread if, for all aca\leq c in ZZ and for all abca\leq b\leq c in PP, we have bZb\in Z. Define the PP-module 𝟏Z:P𝒞{\mathbf{1}}_{Z}:P\to\mathcal{C} as follows:

𝟏Z(b):={𝟏if bZ,0otherwise.{\mathbf{1}}_{Z}(b):=\begin{cases}{\mathbf{1}}&\text{if }b\in Z,\\ 0&\text{otherwise}.\end{cases}

Additionally, the morphism 𝟏Z(ab){\mathbf{1}}_{Z}(a\leq b) is the identity morphism for aba\leq b in ZZ and zero otherwise.

As a direct consequence of Propositions 2.1 and 3.8, we have the following corollary.

Corollary 4.2:

For a spread ZPZ\subset P and a cofree PP-module NaPXaaN\cong\prod_{a\in P}X_{a}^{\downarrow a}, we have

Hom¯(𝟏Z,N)aZXa.\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},N)\cong\prod_{a\in Z}X_{a}.

Henceforth, all Ext-functors will be of the form Ext¯(𝟏Z,N)\underline{\mathrm{Ext}}^{\ast}({\mathbf{1}}_{Z},N), where ZPZ\subset P is a spread.

4.2 Standard Cofree Resolutions and Ext Calculations

This subsection focuses on calculating Ext-functors using cofree resolutions. While injective resolutions are theoretically useful for defining the Ext-functor, cofree resolutions offer computational advantages, particularly for explicit calculations.

A cofree resolution of a PP-module NN is an exact sequence in 𝒞P\mathcal{C}^{P}:

0NF0F1F20\to N\to F^{0}\to F^{1}\to F^{2}\to\cdots

where each FdF^{d} is a cofree PP-module. A specific cofree resolution, called the standard cofree resolution, is particularly useful for computations.

Definition 4.3:

The order complex of a poset PP, denoted by ΔP\Delta P, is a simplicial complex whose nn-simplices are chains σ=a0<<an\sigma=a_{0}<\cdots<a_{n} consisting of n+1n+1 distinct elements of PP. A simplex σ\sigma is a face of a simplex τ\tau, written στ\sigma\unlhd\tau, if σ\sigma is a subchain of τ\tau.

We use σiτ\sigma\lhd_{i}\tau to indicate that dimτdimσ=i\dim\tau-\dim\sigma=i. For a simplex τ=a0<<ad\tau=a_{0}<\cdots<a_{d}, we denote minτ\min\tau as a0a_{0} and maxτ\max\tau as ada_{d}. If στ\sigma\unlhd\tau, then minτminσ\min\tau\leq\min\sigma and maxσmaxτ\max\sigma\leq\max\tau.

Definition 4.4:

The standard cofree resolution of a PP-module NN is the exact sequence:

0\displaystyle{0}N\displaystyle{N}F0N\displaystyle{F^{0}N}F1N\displaystyle{F^{1}N}\displaystyle{\cdots}ϵ\scriptstyle{\epsilon}δ0\scriptstyle{\delta^{0}}δ1\scriptstyle{\delta^{1}} (2)

Define FdN:P𝒞F^{d}N:P\to\mathcal{C} as

FdN:=σΔP:dimσ=dN(maxσ)minσ.F^{d}N:=\prod_{\sigma\in\Delta P:\dim\sigma=d}N(\max\sigma)^{\downarrow\min\sigma}.

For σ1τ\sigma\lhd_{1}\tau, define the map

δd|σ1τ:N(maxσ)minσN(maxτ)minτ\delta^{d}|_{\sigma\lhd_{1}\tau}:N(\max\sigma)^{\downarrow\min\sigma}\to N(\max\tau)^{\downarrow\min\tau}

as

δd|σ1τ(a)={[τ:σ]N(maxσmaxτ)if aminσ,0otherwise.\delta^{d}|_{\sigma\lhd_{1}\tau}(a)=\begin{cases}[\tau:\sigma]\cdot N(\max\sigma\leq\max\tau)&\textup{if }a\leq\min\sigma,\\ 0&\textup{otherwise}.\end{cases}

The universal property of the product induces the full coboundary morphism δd:FdNFd+1N\delta^{d}:F^{d}N\to F^{d+1}N.

The canonical monomorphism ϵ:NF0N\epsilon:N{\hookrightarrow}F^{0}N is defined as follows. For a 0-simplex σΔP\sigma\in\Delta P (i.e., an element b=maxσ=minσb=\max\sigma=\min\sigma), define the component

ϵ|σ:NN(maxσ)minσ\epsilon|_{\sigma}:N\to N(\max\sigma)^{\downarrow\min\sigma}

as

ϵ|σ(a)={N(amaxσ)if aminσ,0otherwise.\epsilon|_{\sigma}(a)=\begin{cases}N(a\leq\max\sigma)&\textup{if }a\leq\min\sigma,\\ 0&\textup{otherwise}.\end{cases}

The universal property of the product induces the full monomorphism ϵ\epsilon.

Proposition 4.5 ([5]):

The standard cofree resolution of a PP-module is exact.

Now consider an elementary cofree PP-module XaaX_{a}^{\downarrow a}. Since 𝒞\mathcal{C} has enough injectives, XaX_{a} admits an injective resolution:

0XaIa0Ia1Ia2,0\to X_{a}\hookrightarrow I^{0}_{a}\to I^{1}_{a}\to I^{2}_{a}\to\cdots,

which induces an injective resolution of the elementary PP-module:

0Xaa(Ia0)a(Ia1)a(Ia2)a.0\to X_{a}^{\downarrow a}\hookrightarrow(I^{0}_{a})^{\downarrow a}\to(I^{1}_{a})^{\downarrow a}\to(I^{2}_{a})^{\downarrow a}\to\cdots.

Applying the functor Hom¯(𝟏Z,)\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},-), and using Propositions 3.8 and 2.1, we obtain, assuming aZa\in Z, the following acyclic cochain complex:

0Ia0Ia1Ia20\to I^{0}_{a}\to I^{1}_{a}\to I^{2}_{a}\to\cdots (3)

If aZa\notin Z, the resulting chain complex is zero. Thus, Ext¯d(𝟏Z,Xaa)=0\underline{\mathrm{Ext}}^{d}({\mathbf{1}}_{Z},X_{a}^{\downarrow a})=0 for all d>0d>0. In other words, elementary cofree PP-modules, and hence cofree PP-modules in general, are Hom¯(𝟏Z,)\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},-)-acyclic, leading to the following proposition.

Proposition 4.6:

[6, Theorem III.6.16] Let 0NF0NF1NF2N0\to N\to F^{0}N\to F^{1}N\to F^{2}N\to\cdots be the standard cofree resolution of a PP-module NN, and let 0NI0I1I20\to N\to I^{0}\to I^{1}\to I^{2}\to\cdots be any injective resolution of NN. For all spreads ZPZ\subset P, the two cochain complexes

0Hom¯(𝟏Z,I0)Hom¯(𝟏Z,I1)Hom¯(𝟏Z,I2)0\to\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},I^{0})\to\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},I^{1})\to\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},I^{2})\to\cdots

and

0Hom¯(𝟏Z,F0N)Hom¯(𝟏Z,F1N)Hom¯(𝟏Z,F2N)0\to\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},F^{0}N)\to\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},F^{1}N)\to\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},F^{2}N)\to\cdots

are quasi-isomorphic, i.e., their cohomology objects are isomorphic.

4.3 Euler Characteristic of Ext-Functors

Having established how Ext-functors are computed using cofree resolutions, we now define the Euler characteristic as an element of the Grothendieck ring.

Definition 4.7:

The Euler characteristic of the Ext-functor Ext¯(𝟏Z,N)\underline{\mathrm{Ext}}^{\ast}({\mathbf{1}}_{Z},N) is defined as:

χ(𝟏Z,N):=d>0(1)d[Ext¯d(𝟏Z,N)].\chi({\mathbf{1}}_{Z},N):=\sum_{d>0}(-1)^{d}\big{[}\underline{\mathrm{Ext}}^{d}({\mathbf{1}}_{Z},N)\big{]}.

The following proposition gives an explicit formula for computing χ(𝟏Z,N)\chi({\mathbf{1}}_{Z},N).

Proposition 4.8:

For a spread ZPZ\subset P and a PP-module NN, the Euler characteristic is given by:

χ(𝟏Z,N)=d0(1)dσΔPdimσ=dminσZ[N(maxσ)].\chi({\mathbf{1}}_{Z},N)=\sum_{d\geq 0}(-1)^{d}\sum_{\begin{subarray}{c}\sigma\in\Delta P\\ \dim\sigma=d\\ \min\sigma\in Z\end{subarray}}\left[N(\max\sigma)\right].
Proof.

We compute the Euler characteristic by analyzing the cochain complex:

0δ1Hom¯(𝟏Z,F0N)δ0Hom¯(𝟏Z,F1N)δ1Hom¯(𝟏Z,F2N)δ20\stackrel{{\scriptstyle\delta^{-1}}}{{\to}}\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},F^{0}N)\stackrel{{\scriptstyle\delta^{0}}}{{\to}}\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},F^{1}N)\stackrel{{\scriptstyle\delta^{1}}}{{\to}}\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},F^{2}N)\stackrel{{\scriptstyle\delta^{2}}}{{\to}}\cdots (4)

For each degree dd, let

Zd(𝟏Z,N):=ker(δd)andBd(𝟏Z,N):=im(δd1),Z_{d}({\mathbf{1}}_{Z},N):={\mathrm{ker}}\,(\delta^{d})\quad\text{and}\quad B_{d}({\mathbf{1}}_{Z},N):={\mathrm{im}}\,(\delta^{d-1}),

yielding the short exact sequence:

0Zd(𝟏Z,N)Hom¯(𝟏Z,FdN)Bd+1(𝟏Z,N)0.0\to Z_{d}({\mathbf{1}}_{Z},N)\hookrightarrow\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},F^{d}N)\twoheadrightarrow B_{d+1}({\mathbf{1}}_{Z},N)\to 0.

By Corollary 4.2, the middle term can be expressed as:

Hom¯(𝟏Z,FdN)σΔPdimσ=dminσZN(maxσ).\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},F^{d}N)\cong\prod_{\begin{subarray}{c}\sigma\in\Delta P\\ \dim\sigma=d\\ \min\sigma\in Z\end{subarray}}N(\max\sigma).

Thus, in the Grothendieck ring:

σΔPdimσ=dminσZ[N(maxσ)]=[Hom¯(𝟏Z,FdN)]=[Zd(𝟏Z,N)]+[Bd+1(𝟏Z,N)].\sum_{\begin{subarray}{c}\sigma\in\Delta P\\ \dim\sigma=d\\ \min\sigma\in Z\end{subarray}}\left[N(\max\sigma)\right]=\left[\underline{\mathrm{Hom}}({\mathbf{1}}_{Z},F^{d}N)\right]=[Z_{d}({\mathbf{1}}_{Z},N)]+[B_{d+1}({\mathbf{1}}_{Z},N)].

Since the cohomology of Equation (4) computes the Ext-functors Ext¯d(𝟏Z,N)\underline{\mathrm{Ext}}^{d}({\mathbf{1}}_{Z},N), we have the following short exact sequence:

0Bd(𝟏Z,N)Zd(𝟏Z,N)Ext¯d(𝟏Z,N)0,0\to B_{d}({\mathbf{1}}_{Z},N)\hookrightarrow Z_{d}({\mathbf{1}}_{Z},N)\twoheadrightarrow\underline{\mathrm{Ext}}^{d}({\mathbf{1}}_{Z},N)\to 0,

which gives:

[Ext¯d(𝟏Z,N)]=[Zd(𝟏Z,N)][Bd(𝟏Z,N)].[\underline{\mathrm{Ext}}^{d}({\mathbf{1}}_{Z},N)]=[Z_{d}({\mathbf{1}}_{Z},N)]-[B_{d}({\mathbf{1}}_{Z},N)].

Using the fact that [B0(𝟏Z,N)]=0[B_{0}({\mathbf{1}}_{Z},N)]=0, we obtain:

σΔPdimσ=dminσZ(1)d[N(maxσ)]\displaystyle\sum_{\begin{subarray}{c}\sigma\in\Delta P\\ \dim\sigma=d\\ \min\sigma\in Z\end{subarray}}(-1)^{d}\left[N(\max\sigma)\right] =d0(1)d([Zd(𝟏Z,N)]+[Bd+1(𝟏Z,N)])\displaystyle=\sum_{d\geq 0}(-1)^{d}\left([Z_{d}({\mathbf{1}}_{Z},N)]+[B_{d+1}({\mathbf{1}}_{Z},N)]\right)
=d0(1)d[Zd(𝟏Z,N)]d1(1)d[Bd(𝟏Z,N)]\displaystyle=\sum_{d\geq 0}(-1)^{d}[Z_{d}({\mathbf{1}}_{Z},N)]-\sum_{d\geq 1}(-1)^{d}[B_{d}({\mathbf{1}}_{Z},N)]
=d0(1)d([Zd(𝟏Z,N)][Bd(𝟏Z,N)])\displaystyle=\sum_{d\geq 0}(-1)^{d}\left([Z_{d}({\mathbf{1}}_{Z},N)]-[B_{d}({\mathbf{1}}_{Z},N)]\right)
=d0(1)d[Ext¯d(𝟏Z,N)]\displaystyle=\sum_{d\geq 0}(-1)^{d}[\underline{\mathrm{Ext}}^{d}({\mathbf{1}}_{Z},N)]
=χ(𝟏Z,N).\displaystyle=\chi({\mathbf{1}}_{Z},N).\qed

5 Möbius Cohomology

In this section, we introduce Möbius cohomology as a special case of Ext-cohomology that applies to indicator modules associated with individual elements of a poset. The discussion concludes with an overview of Möbius inversions and how Möbius cohomology serves as its categorification.

5.1 Definition

We begin by specializing the notion of indicator modules 𝟏Z{\mathbf{1}}_{Z} to singleton sets. For any element aPa\in P, the set {a}P\{a\}\subset P is a spread, allowing us to define its indicator module in the same way as for more general indicator PP-modules. For simplicity, we denote the PP-module 𝟏{a}{\mathbf{1}}_{\{a\}} by 𝟏a{\mathbf{1}}_{a}.

Definition 5.1:

The Möbius cohomology of a PP-module NN at aPa\in P is given by the Ext-functor:

Ext¯d(𝟏a,N).\underline{\mathrm{Ext}}^{d}({\mathbf{1}}_{a},N).

By Equation 4 and Corollary 4.2, the Möbius cohomology corresponds to the cohomology of the following cochain complex:

0σΔPdimσ=0minσ=aN(maxσ)aδ0σΔPdimσ=1minσ=aN(maxσ)aδ1σΔPdimσ=2minσ=aN(maxσ)aδ20\to\prod_{\begin{subarray}{c}\sigma\in\Delta P\\ \dim\sigma=0\\ \min\sigma=a\end{subarray}}N(\max\sigma)^{\downarrow a}\stackrel{{\scriptstyle\delta^{0}}}{{\to}}\prod_{\begin{subarray}{c}\sigma\in\Delta P\\ \dim\sigma=1\\ \min\sigma=a\end{subarray}}N(\max\sigma)^{\downarrow a}\stackrel{{\scriptstyle\delta^{1}}}{{\to}}\prod_{\begin{subarray}{c}\sigma\in\Delta P\\ \dim\sigma=2\\ \min\sigma=a\end{subarray}}N(\max\sigma)^{\downarrow a}\stackrel{{\scriptstyle\delta^{2}}}{{\to}}\cdots (5)
Remark 5.2:

Möbius homology, introduced by Patel and Skraba [13], provides a dual view. It is defined as the homology of the chain complex:

σΔPdimσ=2maxσ=aN(minσ)aσΔPdimσ=1maxσ=aN(minσ)aσΔPdimσ=0maxσ=aN(minσ)a0.\cdots\longrightarrow\prod_{\begin{subarray}{c}\sigma\in\Delta P\\ \dim\sigma=2\\ \max\sigma=a\end{subarray}}N(\min\sigma)^{\downarrow a}\longrightarrow\prod_{\begin{subarray}{c}\sigma\in\Delta P\\ \dim\sigma=1\\ \max\sigma=a\end{subarray}}N(\min\sigma)^{\downarrow a}\longrightarrow\prod_{\begin{subarray}{c}\sigma\in\Delta P\\ \dim\sigma=0\\ \max\sigma=a\end{subarray}}N(\min\sigma)^{\downarrow a}\longrightarrow 0.

The above chain complex is precisely the dual of Equation 5, demonstrating that Möbius cohomology and homology are dual constructions.

5.2 Background: Möbius Inversion

This subsection introduces the classical combinatorial tool of Möbius inversion, which, in the next subsection, is connected to Möbius cohomology through the Euler characteristic.

For any pair aca\leq c in a poset PP, the interval between them is defined as:

[a,c]:={bPabc}.[a,c]:=\{b\in P\mid a\leq b\leq c\}.

Let IntP\mathrm{Int}\,P denote the set of all intervals in PP.

The {\mathbb{Z}}-incidence algebra of PP, denoted Inc(P){\mathrm{Inc}}(P), consists of functions α:IntP\alpha:\mathrm{Int}\,P\to{\mathbb{Z}} with operations of scaling, addition, and multiplication. Multiplication is given by:

(αβ)[a,c]:=b:abcα[a,b]β[b,c].(\alpha\ast\beta)[a,c]:=\sum_{b:a\leq b\leq c}\alpha[a,b]\cdot\beta[b,c].

The multiplicative identity in this algebra is:

𝟏[a,b]={1if a=b,0otherwise.{\mathbf{1}}[a,b]=\begin{cases}1&\text{if }a=b,\\ 0&\text{otherwise}.\end{cases}

The zeta function, denoted by ζ\zeta, assigns the value 11 to every interval: ζ[a,b]=1\zeta[a,b]=1. The inverse of ζ\zeta, denoted μ\mu, is known as the Möbius function.

Lemma 5.3 (Philip Hall’s Theorem, Prop 3.8.5 [15]):

For aba\leq b in PP, let nin_{i} denote the number of chains of length ii from aa to bb. Then:

μ[a,b]=i>0(1)ini.\mu[a,b]=\sum_{i>0}(-1)^{i}\cdot n_{i}.
Definition 5.4:

Let PP be a finite poset, and let \mathcal{R} be a unital commutative ring. The (upper) Möbius inversion of a function f:Pf:P\to\mathcal{R} is defined as:

f(a):=b:abf(b)μ[a,b].\partial f(a):=\sum_{b:a\leq b}f(b)\cdot\mu[a,b]. (6)

The Möbius inversion is the unique \mathcal{R}-valued function on PP satisfying the following identity for all aPa\in P:

b:abf(b)\displaystyle\sum_{b:a\leq b}\partial f(b) =b:abf(b)ζ[a,b]\displaystyle=\sum_{b:a\leq b}\partial f(b)\cdot\zeta[a,b]
=b:ab(c:bcf(c)μ[b,c])ζ[a,b]\displaystyle=\sum_{b:a\leq b}\left(\sum_{c:b\leq c}f(c)\cdot\mu[b,c]\right)\cdot\zeta[a,b]
=c:acf(a)(b:abcζ[a,b]μ[b,c])\displaystyle=\sum_{c:a\leq c}f(a)\left(\sum_{b:a\leq b\leq c}\zeta[a,b]\cdot\mu[b,c]\right)
=c:acf(c)𝟏[a,c]\displaystyle=\sum_{c:a\leq c}f(c)\cdot{\mathbf{1}}[a,c]
=f(a).\displaystyle=f(a).

Note that the set P\mathcal{R}^{P} of all functions PP\to\mathcal{R} forms an \mathcal{R}-module. Möbius inversion can then be viewed as a module homomorphism :PP\partial:\mathcal{R}^{P}\to\mathcal{R}^{P}. We will also write P\partial_{P} when we wish to emphasize the poset over which the Möbius inversion is being taken.

Remark 5.5:

While our focus is on the upper Möbius inversion, which sums over elements greater than or equal to a given element aPa\in P, there is an alternative form known as the lower Möbius inversion. Specifically, for a function f:Pf:P\to\mathcal{R}, the lower inversion is defined as:

f(a)=b:baf(b)μ[b,a].\partial_{-}f(a)=\sum_{b:b\leq a}f(b)\cdot\mu[b,a].

Although this form is not required in our framework, it plays a dual role to the upper inversion.

5.3 Euler Characteristic of Möbius Cohomology

We now examine the relationship between Möbius cohomology and Möbius inversion through the Euler characteristic.

Definition 5.6:

The dimension function of a PP-module NN is the function n:P(𝒞)n:P\to\mathcal{R}(\mathcal{C}) given by:

n(b):=[N(b)].n(b):=[N(b)].
Theorem 5.7:

For any PP-module NN and aPa\in P, the Möbius inversion of the dimension function equals the Euler characteristic of the Möbius cohomology:

n(a)=d0(1)d[Ext¯d(𝟏a,N)].\partial n(a)=\sum_{d\geq 0}(-1)^{d}\big{[}\underline{\mathrm{Ext}}^{d}({\mathbf{1}}_{a},N)\big{]}.
Proof.

For each aPa\in P, we have

d0(1)d[Ext¯(𝟏a,N)]\displaystyle\sum_{d\geq 0}(-1)^{d}\big{[}\underline{\mathrm{Ext}}({\mathbf{1}}_{a},N)\big{]} =σΔPdimσ=dminσ=aN(maxσ)\displaystyle=\prod_{\begin{subarray}{c}\sigma\in\Delta P\\ \dim\sigma=d\\ \min\sigma=a\end{subarray}}N(\max\sigma) by Proposition 4.8
=d0(1)db:abnd(a,b)[N(b)]\displaystyle=\sum_{d\geq 0}(-1)^{d}\sum_{b:a\leq b}n_{d}(a,b)\cdot[N(b)]
=b:abd0(1)dnd(a,b)[N(b)]\displaystyle=\sum_{b:a\leq b}\sum_{d\geq 0}(-1)^{d}n_{d}(a,b)\cdot[N(b)]
=b:abμ(a,b)[N(b)]\displaystyle=\sum_{b:a\leq b}\mu(a,b)\cdot[N(b)] by Lemma 5.3
=m(a).\displaystyle=\partial m(a). by Equation (6)\displaystyle\text{by Equation~{}(\ref{eq:mobius_inversion})}\qed
Remark 5.8:

The lower Möbius inversion arises naturally in the decategorification of Möbius homology. This reflects the duality between homology and cohomology: while Möbius cohomology decategorifies to the upper inversion, Möbius homology decategorifies to the lower inversion.

6 Galois Connections

This section examines Galois connections, which are adjunctions between posets. We begin by defining the fundamental functors between poset modules that arise from monotone functions, followed by a discussion of the adjoint functors induced by Galois connections. The central focus of this section is Theorem 6.6, which serves as a categorification of Rota’s classical theorem on Galois connections and Möbius functions. We then review Rota’s classical theorem before demonstrating how it is categorified by our theorem.

6.1 Adjoint Functors from Monotone Functions

In this subsection, we explore the functors that naturally arise from monotone functions between posets. These functors—pushforward, pushforward with open supports, and pullback—allow us to transfer structure between categories of modules over different posets.

Let f:PQf:P\to Q be a monotone function between two posets. A function ff is monotone if aba\leq b in PP implies f(a)f(b)f(a)\leq f(b) in QQ. A monotone function generates three distinct but related functors:

  • The pushforward functor f:𝒞P𝒞Qf_{\ast}:\mathcal{C}^{P}\to\mathcal{C}^{Q} sends a PP-module MM to the QQ-module fMf_{\ast}M, where

    fM(x):=𝗅𝗂𝗆aP:f(a)xM(a).f_{\ast}M(x):={\mathsf{lim}}\,_{a\in P:f(a)\geq x}M(a).
  • The pushforward with open supports functor f:𝒞P𝒞Qf_{\dagger}:\mathcal{C}^{P}\to\mathcal{C}^{Q} sends a PP-module MM to the QQ-module fMf_{\dagger}M, where

    fM(x):=𝖼𝗈𝗅𝗂𝗆aP:f(a)xM(a).f_{\dagger}M(x):={\mathsf{colim}}\,_{a\in P:f(a)\leq x}M(a).
  • The pullback functor f:𝒞Q𝒞Pf^{\ast}:\mathcal{C}^{Q}\to\mathcal{C}^{P} sends a QQ-module NN to the PP-module fNf^{\ast}N, where

    fN:=Nf.f^{\ast}N:=N\circ f.

The relationships among these functors are summarized in the following diagram:

𝒞P\displaystyle{\mathcal{C}^{P}}𝒞Q\displaystyle{\mathcal{C}^{Q}}f\scriptstyle{f_{\ast}}f\scriptstyle{f_{\dagger}}f\scriptstyle{f^{\ast}}
Proposition 6.1:

[4, Theorems 3.14 and 3.15] For every monotone function f:PQf:P\to Q, the following adjunctions hold:

ffandff.f_{\dagger}\dashv f^{\ast}\quad\text{and}\quad f^{\ast}\dashv f_{\ast}.

Since right adjoints are left-exact, and left adjoints are right-exact, the following properties hold:

  • ff_{\ast} is left-exact.

  • ff^{\ast} is exact (both left-exact and right-exact).

  • ff_{\dagger} is right-exact.

6.2 Galois Connections and Functorial Equality

In this subsection, we explore how Galois connections induce equalities of functors.

Definition 6.2:

A Galois connection between two posets PP and QQ consists of two monotone functions f:PQf:P\to Q and g:QPg:Q\to P such that:

f(a)xag(x),f(a)\leq x\iff a\leq g(x),

for all aPa\in P and xQx\in Q. We write f:PQ:gf:P\rightleftarrows Q:g and refer to ff as the left adjoint and gg as the right adjoint.

Proposition 6.3:

For a Galois connection f:PQ:gf:P\rightleftarrows Q:g, the following functorial equalities hold:

f=gandf=g.f^{\ast}=g_{\ast}\quad\text{and}\quad f_{\dagger}=g^{\ast}.
Proof.

For N𝒞QN\in\mathcal{C}^{Q} and aPa\in P, we have:

gN(a)=𝗅𝗂𝗆xQ:ag(x)N(x)=N(f(a))=fN(a).g_{\ast}N(a)={\mathsf{lim}}\,_{x\in Q:a\leq g(x)}N(x)=N(f(a))=f^{\ast}N(a).

Similarly, for M𝒞PM\in\mathcal{C}^{P} and xQx\in Q, we obtain:

fM(x)=𝖼𝗈𝗅𝗂𝗆aP:f(a)xM(a)=M(g(x))=gM(x).f_{\dagger}M(x)={\mathsf{colim}}\,_{a\in P:f(a)\leq x}M(a)=M(g(x))=g^{\ast}M(x).\qed

6.3 Categorical Galois Connection Theorem

Galois connections naturally give rise to an enriched adjunction leading us to the main theorem of this section.

Given a Galois connection f:PQ:gf:P\rightleftarrows Q:g, Propositions 6.1 and 6.3 establish the following adjunctions:

g=ffandg=ff.g^{*}=f_{\dagger}\dashv f^{*}\quad\text{and}\quad g_{*}=f^{*}\dashv f_{*}.

Thus, we have the following natural isomorphisms:

Nat(gM,N)Nat(M,fN)andNat(gN,M)Nat(N,fM),\mathrm{Nat}(g^{*}M,N)\cong\mathrm{Nat}(M,f^{*}N)\quad\text{and}\quad\mathrm{Nat}(g_{\ast}N,M)\cong\mathrm{Nat}(N,f_{\ast}M),

for all M𝒞PM\in\mathcal{C}^{P} and N𝒞QN\in\mathcal{C}^{Q}.

In what follows, we extend these adjunctions to the enriched hom setting, formalized in Theorem 6.6. To do so, we first establish the necessary and sufficient conditions under which an adjunction between functors extends to the enriched hom.

Lemma 6.4:

Let L:𝒞P𝒞Q:RL:\mathcal{C}^{P}\rightleftarrows\mathcal{C}^{Q}:R be a pair of adjoint functors with LRL\dashv R. Then

Hom¯(L(M),N)Hom¯(M,R(N))\underline{\mathrm{Hom}}(L(M),N)\cong\underline{\mathrm{Hom}}(M,R(N))

for all M𝒞PM\in\mathcal{C}^{P} and N𝒞QN\in\mathcal{C}^{Q} if and only if

L(APM)AQL(M)L(A^{P}\otimes M)\cong A^{Q}\otimes L(M)

for all M𝒞PM\in\mathcal{C}^{P}. Moreover, the first isomorphism is natural in MM and NN if and only if the second is natural in MM.

Proof.

Assume that Hom¯(L(M),N)Hom¯(M,R(N))\underline{\mathrm{Hom}}(L(M),N)\cong\underline{\mathrm{Hom}}(M,R(N)) for all M𝒞PM\in\mathcal{C}^{P} and N𝒞QN\in\mathcal{C}^{Q}. Then for any M𝒞PM\in\mathcal{C}^{P} and N𝒞QN\in\mathcal{C}^{Q}, we have:

Nat(L(APM),N)\displaystyle\mathrm{Nat}(L(A^{P}\otimes M),N) Nat(APM,R(N))\displaystyle\cong\mathrm{Nat}(A^{P}\otimes M,R(N)) since LR\displaystyle\text{since }L\dashv R
Hom𝒞(A,Hom¯(M,R(N)))\displaystyle\cong\mathrm{Hom}_{\mathcal{C}}(A,\underline{\mathrm{Hom}}(M,R(N))) by Proposition 3.7
Hom𝒞(A,Hom¯(L(M),N))\displaystyle\cong\mathrm{Hom}_{\mathcal{C}}(A,\underline{\mathrm{Hom}}(L(M),N)) by assumption
Nat(AQL(M),N)\displaystyle\cong\mathrm{Nat}(A^{Q}\otimes L(M),N) by Proposition 3.7.\displaystyle\text{by Proposition~{}\ref{prop:enriched_hom}}.

Since N𝒞QN\in\mathcal{C}^{Q} was arbitrary, by the Yoneda lemma, we have L(APM)AQL(M)L(A^{P}\otimes M)\cong A^{Q}\otimes L(M). These isomorphisms are natural in MM and NN, assuming the isomorphism Hom¯(L(M),N)Hom¯(M,R(N))\underline{\mathrm{Hom}}(L(M),N)\cong\underline{\mathrm{Hom}}(M,R(N)) is natural. Hence, the isomorphism L(APM)AQL(M)L(A^{P}\otimes M)\cong A^{Q}\otimes L(M) is natural in MM.

Conversely, suppose L(APM)AQL(M)L(A^{P}\otimes M)\cong A^{Q}\otimes L(M) for all M𝒞PM\in\mathcal{C}^{P}. For every object A𝒞A\in\mathcal{C}, we have:

Hom𝒞(A,Hom¯(L(M),N))\displaystyle\mathrm{Hom}_{\mathcal{C}}(A,\underline{\mathrm{Hom}}(L(M),N)) Nat(AQL(M),N)\displaystyle\cong\mathrm{Nat}(A^{Q}\otimes L(M),N) by Proposition 3.7
Nat(L(APM),N)\displaystyle\cong\mathrm{Nat}(L(A^{P}\otimes M),N) by assumption
Nat(APM,R(N))\displaystyle\cong\mathrm{Nat}(A^{P}\otimes M,R(N)) since LR\displaystyle\text{since }L\dashv R
Hom𝒞(A,Hom¯(M,R(N)))\displaystyle\cong\mathrm{Hom}_{\mathcal{C}}(A,\underline{\mathrm{Hom}}(M,R(N))) by Proposition 3.7.\displaystyle\text{by Proposition~{}\ref{prop:enriched_hom}}.

By the Yoneda lemma, we conclude that Hom¯(L(M),N)Hom¯(M,R(N))\underline{\mathrm{Hom}}(L(M),N)\cong\underline{\mathrm{Hom}}(M,R(N)). All of these isomorphisms are natural in MM and NN, provided that L(APM)AQL(M)L(A^{P}\otimes M)\cong A^{Q}\otimes L(M) is natural in MM. This completes the proof. ∎

Next, we verify that pullbacks and pushforwards with open supports satisfy the conditions of the preceding lemma.

Lemma 6.5:

Let h:PQh:P\to Q be a monotone function between posets. Then h(APM)AQhMh_{\dagger}(A^{P}\otimes M)\cong A^{Q}\otimes h_{\dagger}M and h(AQN)APhNh^{*}(A^{Q}\otimes N)\cong A^{P}\otimes h^{*}N for all M𝒞PM\in\mathcal{C}^{P} and N𝒞QN\in\mathcal{C}^{Q}. Moreover, these isomorphisms are natural in MM and NN.

Proof.

By Proposition 3.7, the tensor product M:𝒞𝒞P-\otimes M:\mathcal{C}\to\mathcal{C}^{P} is a left-adjoint and therefore commute with colimits. Let us verify the isomorphism h(APM)AQhMh_{\dagger}(A^{P}\otimes M)\cong A^{Q}\otimes h_{\dagger}M. For each xQx\in Q, we have

h(APM)(x)\displaystyle h_{\dagger}(A^{P}\otimes M)(x) =𝖼𝗈𝗅𝗂𝗆aP:h(a)x(APM)(a)\displaystyle={\mathsf{colim}}\,_{a\in P:h(a)\leq x}(A^{P}\otimes M)(a) by definition of hh_{\dagger}
A𝖼𝗈𝗅𝗂𝗆aP:h(a)xM(a)\displaystyle\cong A\otimes{\mathsf{colim}}\,_{a\in P:h(a)\leq x}M(a) since \otimes commutes with colimts
=(AQhM)(x)\displaystyle=(A^{Q}\otimes h_{\dagger}M)(x) by definition of hh_{\dagger}.

Each of these isomorphisms is natural in xx and MM, so we obtain an isomorphism h(APM)AQhMh_{\dagger}(A^{P}\otimes M)\cong A^{Q}\otimes h_{\dagger}M of QQ-modules, natural in MM.

Next, we verify the isomorphism h(AQN)APhNh^{*}(A^{Q}\otimes N)\cong A^{P}\otimes h^{*}N. For each aPa\in P, we have:

h(AQN)(a)=(AQN)(h(a))=AN(h(a))=AhN(a)=(APhN)(a).h^{*}(A^{Q}\otimes N)(a)=(A^{Q}\otimes N)(h(a))=A\otimes N(h(a))=A\otimes h^{*}N(a)=(A^{P}\otimes h^{*}N)(a).

Thus, h(AQN)h^{*}(A^{Q}\otimes N) and APhNA^{P}\otimes h^{*}N are pointwise equal. Moreover, the internal maps of both PP-modules are the same, so we in fact have the equality h(AQN)=APhNh^{*}(A^{Q}\otimes N)=A^{P}\otimes h^{*}N, which is trivially natural in NN. ∎

Theorem 6.6 (Categorical Rota’s Galois Connection Theorem):

Let f:PQ:gf:P\rightleftarrows Q:g be a Galois connection. Then:

Hom¯(gM,N)Hom¯(M,fN)andHom¯(gN,M)Hom¯(N,fM),\underline{\mathrm{Hom}}(g^{*}M,N)\cong\underline{\mathrm{Hom}}(M,f^{*}N)\quad\text{and}\quad\underline{\mathrm{Hom}}(g_{*}N,M)\cong\underline{\mathrm{Hom}}(N,f_{*}M),

for all M𝒞PM\in\mathcal{C}^{P} and N𝒞QN\in\mathcal{C}^{Q}.

Proof.

By Lemma 6.4 and Lemma 6.5, we have:

Hom¯(hM,N)Hom¯(M,hN)andHom¯(hN,M)Hom¯(N,hM).\underline{\mathrm{Hom}}(h_{\dagger}M,N)\cong\underline{\mathrm{Hom}}(M,h^{*}N)\quad\text{and}\quad\underline{\mathrm{Hom}}(h^{*}N,M)\cong\underline{\mathrm{Hom}}(N,h_{*}M).

Using Proposition 6.3, we substitute f=gf_{\dagger}=g^{\ast} and f=gf^{\ast}=g_{\ast}:

Hom¯(gM,N)Hom¯(M,fN),andHom¯(gN,M)Hom¯(N,fM).\underline{\mathrm{Hom}}(g^{\ast}M,N)\cong\underline{\mathrm{Hom}}(M,f^{\ast}N),\quad\text{and}\quad\underline{\mathrm{Hom}}(g_{\ast}N,M)\cong\underline{\mathrm{Hom}}(N,f_{\ast}M).\qed

6.4 Background: Rota’s Theorem

In this subsection, we review Rota’s original theorem, which establishes a fundamental relationship between Möbius functions on two posets connected by a Galois connection.

Theorem 6.7 (Rota’s Galois Connection Theorem [14]):

Let f:PQ:gf:P\rightleftarrows Q:g be a Galois connection. For all aPa\in P and yQy\in Q, the following equality holds:

xQ:g(x)=aμQ(x,y)=bP:f(b)=yμP(a,b).\sum_{x\in Q:g(x)=a}\mu_{Q}(x,y)=\sum_{b\in P:f(b)=y}\mu_{P}(a,b).

This theorem connects the Möbius functions μP\mu_{P} and μQ\mu_{Q} associated with the respective posets PP and QQ. We now present an equivalent result using Möbius inversions.

Given a monotone function f:PQf:P\to Q, we define two key operations:

  • The pushforward f#m:Qf_{\#}m:Q\to\mathcal{R} of a function m:Pm:P\to\mathcal{R} is given by:

    f#m(z):=af1(z)m(a).f_{\#}m(z):=\sum_{a\in f^{-1}(z)}m(a).
  • The pullback f#n:Pf^{\#}n:P\to\mathcal{R} of a function n:Qn:Q\to\mathcal{R} is defined by:

    f#n(a):=n(f(a)).f^{\#}n(a):=n(f(a)).

Note that the pushforward and pullback gives rise to \mathcal{R}-module homomorphisms f:PQf_{\sharp}:\mathcal{R}^{P}\to\mathcal{R}^{Q} and f:QPf^{\sharp}:\mathcal{R}^{Q}\to\mathcal{R}^{P}, respectively.

We now introduce an equivalent result to Rota’s theorem, which highlights how these pushforward and pullback operations interact with Möbius inversions. This alternative formulation was first observed by Aziz Gülen and Alex McCleary for the dual setting of lower Möbius inversions [7].

Theorem 6.8:

For a Galois connection f:PQ:gf:P\rightleftarrows Q:g, the following identity holds:

Pf#=g#Q.\partial_{P}\circ f^{\#}=g_{\#}\circ\partial_{Q}.
Proof.

Let n:Qn:Q\to\mathcal{R} be an arbitrary function. For each aPa\in P, we have

b:ab(g#Q)(n)(b)\displaystyle\sum_{b:a\leq b}(g_{\#}\circ\partial_{Q})(n)(b) =b:abqg1(b)Qn(q)\displaystyle=\sum_{b:a\leq b}\sum_{q\in g^{-1}(b)}\partial_{Q}n(q)
=qQ:ag(q)Qn(q)\displaystyle=\sum_{q\in Q:a\leq g(q)}\partial_{Q}n(q)
=qQ:f(a)qQn(q)\displaystyle=\sum_{q\in Q:f(a)\leq q}\partial_{Q}n(q) since fgf\dashv g
=n(f(a))\displaystyle=n(f(a))
=fn(a).\displaystyle=f^{\sharp}n(a).

By uniqueness of the Möbius inversion, it follows that (gQ)n=(Pf)n(g_{\sharp}\circ\partial_{Q})n=(\partial_{P}\circ f^{\sharp})n. Since this is true for every n:Qn:Q\to\mathcal{R}, the result follows. ∎

The equivalence between Theorem 6.8 and Theorem 6.7 can be demonstrated by applying the Möbius inversion formula to a specific case. Choose aPa\in P and yQy\in Q, and let 1y:Q1_{y}:Q\to\mathcal{R} be the function defined by:

1y(z)={1if z=y,0otherwise.1_{y}(z)=\begin{cases}1&\text{if }z=y,\\ 0&\text{otherwise}.\end{cases}

Evaluating both sides of the identity in Theorem 6.8 using this function, we obtain, for all aPa\in P and yQy\in Q:

(g#Q)(1y)(a)\displaystyle(g_{\#}\circ\partial_{Q})(1_{y})(a) =(Pf#)(1y)(a)\displaystyle=(\partial_{P}\circ f^{\#})(1_{y})(a)
\displaystyle\iff\qquad xg1(a)Q1y(x)\displaystyle\sum_{x\in g^{-1}(a)}\partial_{Q}1_{y}(x) =bP:ab(1yf)(b)μP(a,b)\displaystyle=\sum_{b\in P:a\leq b}(1_{y}\circ f)(b)\cdot\mu_{P}(a,b)
\displaystyle\iff\qquad xg1(a)zQ:xz1y(z)μQ(x,z)\displaystyle\sum_{x\in g^{-1}(a)}\sum_{z\in Q:x\leq z}1_{y}(z)\cdot\mu_{Q}(x,z) =bP:f(b)=yμP(a,b)\displaystyle=\sum_{b\in P:f(b)=y}\mu_{P}(a,b)
\displaystyle\iff\qquad xg1(a)μQ(x,y)\displaystyle\sum_{x\in g^{-1}(a)}\mu_{Q}(x,y) =bP:f(b)=yμP(a,b)\displaystyle=\sum_{b\in P:f(b)=y}\mu_{P}(a,b)
\displaystyle\iff\qquad xQ:g(x)=aμQ(x,y)\displaystyle\sum_{x\in Q:g(x)=a}\mu_{Q}(x,y) =bP:f(b)=yμP(a,b).\displaystyle=\sum_{b\in P:f(b)=y}\mu_{P}(a,b).

Thus, Rota’s Galois connection theorem is equivalent to the statement (g#Q)(1y)=(Pf#)(1y)(g_{\#}\circ\partial_{Q})(1_{y})=(\partial_{P}\circ f^{\#})(1_{y}) for all yQy\in Q. Since gQg_{\sharp}\circ\partial_{Q} and Pf\partial_{P}\circ f^{\sharp} are \mathcal{R}-module homomorphism, and since Q\mathcal{R}^{Q} is generated by functions of the form 1y1_{y}, the preceding statement is true if and only if gQ=Pfg_{\sharp}\circ\partial_{Q}=\partial_{P}\circ f^{\sharp}.

6.5 Euler Characteristic and Rota’s Theorem

We now prove in Theorem 6.10 how Theorem 6.6 decategorifies to Theorem 6.8 via the Euler characteristic of the Ext-functor.

The following corollary is an immediate consequence of Theorem 6.6.

Corollary 6.9:

Let f:PQ:gf:P\rightleftarrows Q:g be a Galois connection, let N𝒞QN\in\mathcal{C}^{Q}, and let aPa\in P. Then,

Ext¯d(𝟏a,fN)Ext¯d(g𝟏a,N).\underline{\mathrm{Ext}}^{d}({\mathbf{1}}_{a},f^{*}N)\cong\underline{\mathrm{Ext}}^{d}(g^{*}{\mathbf{1}}_{a},N).
Theorem 6.10:

Let f:PQ:gf:P\rightleftarrows Q:g be a Galois connection and let N𝒞QN\in\mathcal{C}^{Q}. Then, for aPa\in P,

(Pf#)(n)(a)=χ(𝟏a,fN)=χ(g𝟏a,N)=(g#Q)(n)(a),(\partial_{P}\circ f^{\#})(n)(a)=\chi({\mathbf{1}}_{a},f^{\ast}N)=\chi(g^{\ast}{\mathbf{1}}_{a},N)=(g_{\#}\circ\partial_{Q})(n)(a), (7)

where n:Q(𝒞)n:Q\to\mathcal{R}(\mathcal{C}) is the dimension function of NN.

Proof.

We will prove the equalities in equation (7) one by one.

The middle equality follows directly from Corollary 6.9.

The function f#n:P(𝒞)f^{\#}n:P\to\mathcal{R}(\mathcal{C}) is the dimension function for fN:Q𝒞f^{\ast}N:Q\to\mathcal{C}. Thus, by Theorem 5.7, we have:

(Pf#n)(a)=χ(𝟏a,fN).(\partial_{P}\circ f^{\#}n)(a)=\chi({\mathbf{1}}_{a},f^{\ast}N).

To establish the right equality, consider the spread

Z:={xQ:g(x)=a}.Z:=\{x\in Q:g(x)=a\}.

By the definition of gg^{\ast}, it follows that g𝟏a𝟏Zg^{\ast}{\mathbf{1}}_{a}\cong{\mathbf{1}}_{Z}. Therefore, we can express the Euler characteristic as follows:

χ(g𝟏a,N)\displaystyle\chi(g^{*}{\mathbf{1}}_{a},N) =d0(1)dτΔQdimτ=dminτZ[N(maxτ)]\displaystyle=\sum_{d\geq 0}(-1)^{d}\sum_{\begin{subarray}{c}\tau\in\Delta Q\\ \dim\tau=d\\ \min\tau\in Z\end{subarray}}\big{[}N(\max\tau)\big{]} by Proposition 4.8
=xZd0(1)dτΔQdimτ=dminτ=x[N(maxτ)]\displaystyle=\sum_{x\in Z}\sum_{d\geq 0}(-1)^{d}\sum_{\begin{subarray}{c}\tau\in\Delta Q\\ \dim\tau=d\\ \min\tau=x\end{subarray}}\big{[}N(\max\tau)\big{]}
=xZQn(x)\displaystyle=\sum_{x\in Z}\partial_{Q}n(x) by Proposition 4.8
=(g#Q)(n)(a).\displaystyle=(g_{\#}\circ\partial_{Q})(n)(a). by definition of g#\displaystyle\text{ by definition of $g_{\#}$}\qed

References

  • Bac [75] Kenneth Baclawski. Whitney numbers of geometric lattices. Advances in Mathematics, 16:125–138, 1975.
  • Bac [77] Kenneth Baclawski. Galois connections and the Leray spectral sequence. Advances in Mathematics, 25(2):191–215, 1977.
  • CLL [80] Mireille Content, François Lemay, and Pierre Leroux. Catégories de Möbius et fonctorialité: un cadre général pour l’inversion de Möbius. Journal of Combinatorial Theory, Series A, 28(2):169–190, 1980.
  • Cur [18] Justin Michael Curry. Dualities between cellular sheaves and cosheaves. Journal of Pure and Applied Algebra, 222(4):966–993, 2018.
  • Deh [62] R. Deheuvels. Homologie des ensembles ordonnés et des espaces topologiques. Bulletin de la Société Mathématique de France, 90:261–321, 1962.
  • GM [03] I. Gelfand and M. Manin. Methods of Homological Algebra. Springer Monographs in Mathematics. Springer-Verlag, Berlin, 2nd edition, 2003.
  • GM [22] Aziz Burak Gülen and Alexander McCleary. Galois connections in persistent homology. arXiv, July 2022.
  • KM [23] Woojin Kim and Facundo Mémoli. Persistence over posets. Notices of the American Mathematical Society, 70(8), September 2023.
  • Ler [75] Pierre Leroux. Les catégories de Möbius. Cahiers de Topologie et Géométrie Différentielle Catégoriques, 16(3):280–282, 1975.
  • ML [98] Saunders Mac Lane. Categories for the Working Mathematician. Springer, 2nd edition, 1998.
  • OS [24] Steve Oudot and Luis Scoccola. On the stability of multigraded Betti numbers and Hilbert functions. SIAM Journal on Applied Algebra and Geometry, 8(1):54–88, 2024.
  • Pat [18] Amit Patel. Generalized persistence diagrams. Journal of Applied and Computational Topology, 1(3-4):397–419, 2018.
  • PS [23] Amit Patel and Primoz Skraba. Möbius homology. arXiv preprint arXiv:2307.01040, 2023.
  • Rot [64] Gian-Carlo Rota. On the foundations of combinatorial theory I. Theory of Möbius functions. Zeitschrift für Wahrscheinlichkeitstheorie und Verwandte Gebiete, 2(4):340–368, 1964.
  • Sta [11] Richard P. Stanley. Enumerative Combinatorics, volume 1 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2nd edition, 2011.
  • Wei [94] Charles A. Weibel. An Introduction to Homological Algebra. Cambridge University Press, 1994.