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

Magnitude and magnitude homology of filtered set enriched categories

Yasuhiko Asao Department of Applied Mathematics, Fukuoka University [email protected]
Abstract

In this article, we give a framework for studying the Euler characteristic and its categorification of objects across several areas of geometry, topology and combinatorics. That is, the magnitude theory of filtered set enriched categories. It is a unification of the Euler characteristic of finite categories and the magnitude of metric spaces, both of which are introduced by Leinster ([15], [17]). Our definitions cover a class of metric spaces which is broader than the original ones in [15] and [18], so that magnitude (co)weighting of infinite metric spaces can be considered. We give examples of the magnitude from various research areas containing the Poincaré polynomial of ranked posets and the growth series of finitely generated groups. In particular, the magnitude homology gives categorifications of them. We also discuss homotopy invariance of the magnitude homology and its variants. Such a homotopy includes digraph homotopy and rr-closeness of Lipschitz maps. As a benefit of our categorical view point, we generalize the notion of Grothendieck fibrations of small categories to our enriched categories, whose restriction to metric spaces is a notion called metric fibration that is initially introduced in [15]. It is remarkable that the magnitude of such a fibration is a product of those of the fiber and the base. We especially study fibrations of graphs, and give examples of graphs with the same magnitude but are not isomorphic.

1 Introduction

In [17], T. Leinster introduced the Euler characteristic of finite categories, which generalizes that of finite simplicial complexes. Later, he introduced the magnitude of metric spaces ([15]), as an analogue of the above. These two notions have a formalization of great generality, namely the magnitude and the magnitude homology of enriched categories due to Leinster-Shulman ([18]). In this formalization, the magnitude can be defined by choosing a monoidal category VV and a “size function” ObVk\mathrm{Ob}V\longrightarrow k, where kk is an arbitrary semi-ring. In particular, the magnitude is not defined as a single object applicable to all enriched categories simultaneously. For example, the above two magnitude are considered as different things so far.

The aim of this article is to give a foundation of the magnitude and the magnitude homology of filtered set enriched categories, which unifies both of the Euler characteristic for finite categories and the magnitude of metric spaces. We propose to deal with small categories and metric spaces in a single category. Then topological and geometric study for small categories or metric spaces can be generalized to this larger category, which can be a new approach to geometry, topology and combinatorics. In fact, our discussions in this article suggests that there should be a homotopy theory including the magnitude theory for such a larger category, where some topological and geometric studies can be considered in a unified manner.

What we achieve in this article are summarized as follows.

Generalizing magnitude theory

  1. (1)

    We construct a framework for magnitude theory of filtered set enriched categories (Sections 3, 4). Such a framework unifies the Euler characteristic of finite categories and the magnitude of metric spaces with their categorifications, both of which are introduced by Leinster.

  2. (2)

    In our framework, the class of metric spaces for which we can define the magnitude (co)weighting and their categorification is broader than the original one. For example, we can compute the magnitude (co)weighting of locally finite graphs that possibly have infinitely many vertices (Section 3.5). More precisely, we fix a “size function” 𝖢𝖥𝗌𝖾𝗍[[q]]{\sf CFset}\longrightarrow\mathbb{Q}[[q^{\mathbb{R}}]], where 𝖢𝖥𝗌𝖾𝗍{\sf CFset} is the category of collectable filtered sets (Definition 3.2), and [[q]]\mathbb{Q}[[q^{\mathbb{R}}]] is the Novikov series ring (Definition 3.7). Then we can define the magnitude of finite 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-categories and the magnitude (co)weighting of 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-categories of finite type (Definition 3.13), both of which contain finite categories and finite metric spaces. Further, for a special class of 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-categories, namely tame categories (Definition 3.17), we have an explicit formula for the magnitude and the magnitude (co)weighting, which is a generalization of Leinster’s power series expression of the magnitude. It turns out that locally finite graphs are tame categories (Section 3.5.1).

  3. (3)

    To that end, we discuss the convergence of infinite summations of Novikov series. Then we derive conditions for metric spaces and small categories under which the magnitude can be calculated (Section 3.5).

Examples of magnitude

As an example of the magnitude in our framework, we show that the following objects are the magnitude or the magnitude (co)weighting (Section 3.5).

  1. (1)

    magnitude of finite metric spaces in the original sense

  2. (2)

    magnitude (co)weighting of locally finite graphs with infinitely many vertices

  3. (3)

    Euler characteristic of finite categories

  4. (4)

    Euler characteristic of finite simplicial complexes

  5. (5)

    Poincaré polynomial of ranked posets

  6. (6)

    the growth series of finitely generated groups

In particular, the magnitude homology in our sense categorifies them. We have no idea whether the magnitude homology of the Poincaré polynomial or the growth series can have torsions, and what torsions means if any. It is known that the magnitude homology can have torsions in general ([13], [20]).

Magnitude homology as Hochschild homology

Leinster-Shulman pointed out in Remark 5.11 of [18] that the magnitude homology has a form of Hochschild homology in a generalized sense. Here, we give a more ring theoretic description (Section 4.3, Corollary 4.32). Namely we have the following.

Theorem 1.1 (Corollary 4.32).

For a filtered set enriched category CC, we have an isomorphism

𝖬𝖧CGr𝖧𝖧(GrPC(),MC())\displaystyle\operatorname{\sf MH}^{\ell}_{\bullet}C\cong Gr_{\ell}\operatorname{\sf HH}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))

for any 0\ell\in\mathbb{R}_{\geq 0}. Hence we have 0𝖬𝖧C𝖧𝖧(GrPC(),MC())\bigoplus_{\ell\geq 0}\operatorname{\sf MH}^{\ell}_{\bullet}C\cong\operatorname{\sf HH}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z})).

That is, we express the magnitude homology as the Hochschild homology of a generalization of the category algebra. We use techniques of homology theory for small categories to give such an expression. This can also be considered as a kind of generalization of Gerstenhaber-Schack’s theorem ([6]) asserting that the cohomology of a simplicial complex is isomorphic to the Hochschild homology of the incidence algebra (Remark 4.31).

Homotopies for filtered set enriched categories

For a \mathbb{Z}-filtered set enriched categories, we construct a spectral sequence EE that satisfying the following.

Theorem 1.2 (Proposition 4.34).

Let CC be a \mathbb{Z}-filtered set enriched category. Then

  1. (1)

    Ep,q1=𝖬𝖧p+qpCE^{1}_{p,q}=\operatorname{\sf MH}^{p}_{p+q}C.

  2. (2)

    If CC is a digraph, we have Ep,02=H~pCE^{2}_{p,0}=\widetilde{H}_{p}C, where H~\widetilde{H}_{\bullet} denotes the reduced path homology introduced by Grigor’yan–Muranov–Lin–S. -T. Yau et al. [8].

  3. (3)

    It converges to the homology of the underlying small category C¯\underline{C} if it converges, in particular when max{degffMorC}\max\{\deg f\mid f\in\mathrm{Mor}C\} is finite.

Then we discuss homotopy invariance of each page of this spectral sequence. We show that the (r+1)(r+1)-th page of this spectral sequence is invariant under “rr-homotopy” described as follows.

  1. (1)

    0-homotopy contains equivalences of small categories.

  2. (2)

    11-homotopy for digraphs is exactly the digraph homotopy.

  3. (3)

    rr-homotopy for metric spaces is exactly the rr-closeness of Lipschitz maps.

Theorem 1.3 (Theorem 4.37).

The (r+1)(r+1)-page of the above spectral sequence Er+1E^{r+1}_{\bullet\bullet} is invariant under rr-homotopy.

For an example, we have the following.

Theorem 1.4 (Proposition 4.20).

For any generalized metric space XX, we denote its Kolmogorov quotient by 𝖪𝖰X{\sf KQ}X. Then we have 𝖬𝖧X𝖬𝖧𝖪𝖰X\operatorname{\sf MH}^{\ell}_{\bullet}X\cong\operatorname{\sf MH}^{\ell}_{\bullet}{\sf KQ}X. Namely, the magnitude homology is invariant under Kolmogorov quotient.

The above results suggest that there should be “rr-homotopy theories” including the magnitude theory in the category of filtered set enriched categories. In fact, Cirici et al. ([5]) show that there is a cofibrantly generated model structure on the category of 0\mathbb{Z}_{\geq 0}-filtered chain complexes, where weak equivalences are exactly the morphisms inducing quasi isomorphisms on ErE^{r}. Theorem 4.37 suggests that there should be homotopy theories that coincide with the above Ciricis’ structure by taking chain complexes. We also remark that the work [4] by Carranza et al. seems to support this hypothesis. In that paper, they constructed a cofibration category structure on the category of digraphs, where weak equivalences are exactly the morpshisms inducing isomorphisms on the path homology, namely a part of E2E^{2}.

Fibration of filtered set enriched categories

We formulate the fibration of filtered set enriched categories (Section 5). It contains the Grothendieck (op)fibration of small categories and metric fibrations of metric spaces introduced by Leinster ([15]). A remarkable property of this notion is that the magnitude of a fibration splits as the product of those of the “fiber” and the “base”. Namely we have the following.

Proposition 1.5 (Proposition 5.15).

Let EXE\longrightarrow X be a “fibration” with the fibers FF and the base XX admitting the magnitude. Then we have 𝖬𝖺𝗀E=𝖬𝖺𝗀X𝖬𝖺𝗀F{\sf Mag}E={\sf Mag}X\cdot{\sf Mag}F.

Further, we discuss in detail the restriction to metric spaces, namely metric fibrations, and we especially study it for graphs. In particular, we obtain a number of examples of graphs such as in Figure 1 that have the same magnitude but are not isomorphic.

Acknowledgements

The author is grateful to Luigi Caputi for fruitful and helpful comments and feedbacks on the first draft of the paper. He also would like to thank Ayato Mitsuishi and Masahiko Yoshinaga for valuable discussions and comments.

2 Prerequisites

2.1 Enriched categories

For a monoidal category (V,,1)(V,\otimes,1), a category enriched over VV, or VV-category CC is the following data :

  • a set ObC\mathrm{Ob}C,

  • “morphisms” C(a,b)ObVC(a,b)\in\mathrm{Ob}V for any a,bObCa,b\in\mathrm{Ob}C,

  • “a composition rule” C(a,b)C(b,c)C(a,c)C(a,b)\otimes C(b,c)\longrightarrow C(a,c) for any a,b,cObCa,b,c\in\mathrm{Ob}C,

  • “an identity morphism” 1C(a,a)1\longrightarrow C(a,a) for any aObCa\in\mathrm{Ob}C,

which satisfy suitable associativity and unit axioms. We don’t explain details on this subject here. Necessary terminologies are also summarized in Section 1.2 of [15]. For details, we refer to [14] or [19]. We denote the category of small categories, namely 𝖲𝖾𝗍{\sf Set}-categories, by 𝖢𝖺𝗍{\sf Cat}. A small category is finite if it has finitely many objects and morphisms.

2.2 Graphs

  1. (1)

    A graph is a 11-dimensional simplicial complex, that is a pair of sets (V,E)(V,E) such that E{e2V|e|=2}E\subset\{e\in 2^{V}\mid|e|=2\}. A path metric on a graph GG is a function d:V(G)×V(G)0{}d:V(G)\times V(G)\longrightarrow\mathbb{Z}_{\geq 0}\cup\{\infty\} defined by d(a,b)=min{ithere exist a0,,aiV(G) such that a0=a,ai=b,{aj,aj+1}E(G)}d(a,b)=\min\{i\mid\text{there exist }a_{0},\dots,a_{i}\in V(G)\text{ such that }a_{0}=a,a_{i}=b,\{a_{j},a_{j+1}\}\in E(G)\}, or \infty if there are no such sequences. A graph homomorphism f:GHf:G\longrightarrow H is a map f:V(G)V(H)f:V(G)\longrightarrow V(H) such that d(a,b)d(fa,fb)d(a,b)\geq d(fa,fb) for any a,bV(G)a,b\in V(G). Equivalently, that is a map f:V(G)V(H)f:V(G)\longrightarrow V(H) such that either fa=fbfa=fb or {fa,fb}E(H)\{fa,fb\}\in E(H) for any {a,b}E(G)\{a,b\}\in E(G). We denote the category of graphs and graph homomorphisms by Grph.

  2. (2)

    A directed graph or a digraph is a pair of sets (V,E)(V,E) such that EV×VΔVE\subset V\times V\setminus\Delta V, where ΔV\Delta V denotes the diagonal. A directed path metric on a digraph GG is a function d:V(G)×V(G)0{}d:V(G)\times V(G)\longrightarrow\mathbb{Z}_{\geq 0}\cup\{\infty\} defined by d(a,b)=min{ithere exist a0,,aiV(G) such that a0=a,ai=b,(aj,aj+1)E(G)}d(a,b)=\min\{i\mid\text{there exist }a_{0},\dots,a_{i}\in V(G)\text{ such that }a_{0}=a,a_{i}=b,(a_{j},a_{j+1})\in E(G)\}, or \infty if there are no such sequences. A digraph homomorphism f:GHf:G\longrightarrow H is a map f:V(G)V(H)f:V(G)\longrightarrow V(H) such that d(a,b)d(fa,fb)d(a,b)\geq d(fa,fb) for any a,bV(G)a,b\in V(G). Equivalently, that is a map f:V(G)V(H)f:V(G)\longrightarrow V(H) such that either fa=fbfa=fb or (fa,fb)E(H)(fa,fb)\in E(H) for any (a,b)E(G)(a,b)\in E(G). We denote the category of digraphs and digraph homomorphisms by DGrph.

  3. (3)

    We have a functor ι:𝖦𝗋𝗉𝗁𝖣𝖦𝗋𝗉𝗁\iota:{\sf Grph}\longrightarrow{\sf DGrph} defined by V(ιG)=V(G)V(\iota G)=V(G) and E(ιG)={(a,b){a,b}E(G)}E(\iota G)=\{(a,b)\mid\{a,b\}\in E(G)\}, namely by orienting edges in bi-directions. This is a full and faithful functor, and we identify any graphs with their image under ι\iota.

2.3 Metric spaces

  1. (1)

    A generalized metric space (X,d)(X,d) is a set XX with a map d:X×X0{}d:X\times X\longrightarrow\mathbb{R}_{\geq 0}\cup\{\infty\} satisfying

    • d(x,x)=0d(x,x)=0 for any xXx\in X,

    • d(x,y)+d(y,z)d(x,z)d(x,y)+d(y,z)\geq d(x,z) for any x,y,zXx,y,z\in X.

    We sometimes denote a generalized metric space (X,d)(X,d) simply by XX. For generalized metric spaces (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}), a 11-Lipschitz map f:XYf:X\longrightarrow Y is a map satisfying that dX(x,x)dY(fx,fx)d_{X}(x,x^{\prime})\geq d_{Y}(fx,fx^{\prime}) for any x,xXx,x^{\prime}\in X. We denote the category of generalized metric spaces and 1-Lipschitz maps by 𝖦𝖬𝖾𝗍{\sf GMet}.

  2. (2)

    A generalized metric space (X,d)(X,d) is symmetric if it satisfies that d(x,y)=d(y,x)d(x,y)=d(y,x) for any x,yXx,y\in X.

  3. (3)

    A generalized metric space (X,d)(X,d) is non-degenerate if d(x,y)=d(y,x)=0d(x,y)=d(y,x)=0 implies that x=yx=y for any x,yXx,y\in X.

  4. (4)

    A generalized metric space (X,d)(X,d) is a metric space if it is symmetric, non-degenerate and satisfies that Imd\infty\not\in{\rm Im}d. We denote the full subcategory of 𝖦𝖬𝖾𝗍{\sf GMet} that consists of metric spaces and 1-Lipschitz maps by 𝖬𝖾𝗍{\sf Met}. A metric space XX is finite if the set XX is finite.

  5. (5)

    For a generalized metric space (X,d)(X,d) and a point xXx\in X, we define that B(x,):={yXd(x,y)}B(x,\ell):=\{y\in X\mid d(x,y)\leq\ell\} and B(,x):={yXd(y,x)}B(\ell,x):=\{y\in X\mid d(y,x)\leq\ell\}.

3 Magnitude of filtered set enriched categories

In this section, we give a definition of the magnitude of categories enriched over filtered sets, and show basic properties after explaining what precisely ‘filtered sets’ we mean. We also give examples of the magnitude from various areas of topology and geometry containing the Poincaré polynomial of ranked posets and the growth series of finitely generated groups. Our definition is basically the same as the original one in [15], but slightly broader so that we can deal with a broader class of spaces such as infinite graphs, for example.

3.1 Filtered sets

Definition 3.1.
  1. (1)

    A 0\mathbb{R}_{\geq 0}-filtered set, or simply a filtered set is a set XX with subsets XXX_{\ell}\subset X for any 0\ell\in\mathbb{R}_{\geq 0} satisfying that XXX_{\ell}\subset X_{\ell^{\prime}} for any \ell\leq\ell^{\prime} and 0X=X\bigcup_{\ell\in\mathbb{R}_{\geq 0}}X_{\ell}=X. We formally define that X=X_{\ell}=\emptyset for <0\ell<0.

  2. (2)

    Let XX be a filtered set, and let xXx\in X. We define that degx=\deg x=\ell if xX<Xx\in X_{\ell}\setminus\bigcup_{\ell^{\prime}<\ell}X_{\ell^{\prime}}.

  3. (3)

    For filtered sets XX and YY, a filtered map f:XYf:X\longrightarrow Y is a map satisfying that fXYfX_{\ell}\subset Y_{\ell} for any 0\ell\in\mathbb{R}_{\geq 0}.

  4. (4)

    For filtered sets XX and YY, their product X×YX\times Y is defined by (X×Y)=+′′=X×X′′(X\times Y)_{\ell}=\bigcup_{\ell^{\prime}+\ell^{\prime\prime}=\ell}X_{\ell^{\prime}}\times X_{\ell^{\prime\prime}}. Then we have deg(x,y)=degx+degy\deg(x,y)=\deg x+\deg y for any (x,y)X×Y(x,y)\in X\times Y.

  5. (5)

    For a family of filtered sets {Xλ}λΛ\{X_{\lambda}\}_{\lambda\in\Lambda} indexed by a set Λ\Lambda, we define its union λΛXλ\bigcup_{\lambda\in\Lambda}X_{\lambda} by (λΛXλ)=λΛ(Xλ)\left(\bigcup_{\lambda\in\Lambda}X_{\lambda}\right)_{\ell}=\bigcup_{\lambda\in\Lambda}(X_{\lambda})_{\ell}. We sometimes use a notation fYXf\bigcup_{f\in Y}X_{f} for filtered sets YY and XfX_{f}’s, where we consider YY as a set by forgetting the filtration.

We use the following terminologies.

Definition 3.2.
  1. (1)

    A filtered set XX is collectable if XX_{\ell} is a finite set for each 0\ell\in\mathbb{R}_{\geq 0}.

  2. (2)

    We denote a one element filtered set {}\{\ast\} with deg=\deg\ast=\ell by ()\ast(\ell). We formally define that ()=\ast(\infty)=\emptyset.

We denote the category of filtered sets and maps by 𝖥𝗌𝖾𝗍{\sf Fset}. The following is straightforward.

Proposition 3.3.

(𝖥𝗌𝖾𝗍,×)({\sf Fset},\times) is a symmetric monoidal category with the unit object (0)\ast(0).

Let [0,][0,\infty] be the small category with objects 0{}\mathbb{R}_{\geq 0}\cup\{\infty\} and morphisms [0,](,)={{} if ,if <.[0,\infty](\ell,\ell^{\prime})=\begin{cases}\{\ast\}&\text{ if }\ell\geq\ell^{\prime},\\ \emptyset&\text{if }\ell<\ell^{\prime}.\end{cases} We equip [0,][0,\infty] with a symmetric monoidal structure by ++ with the unit object 0. We denote the category of sets with obvious symmetric monoidal structure by 𝖲𝖾𝗍{\sf Set}. The following is also straightforward.

Proposition 3.4.

There is a symmetric monoidal embedding 𝖲𝖾𝗍𝖥𝗌𝖾𝗍{\sf Set}\longrightarrow{\sf Fset} sending a set XX to a filtered set XX with degx=0\deg x=0 for any xXx\in X. There is also a symmetric monoidal embedding [0,]𝖥𝗌𝖾𝗍[0,\infty]\longrightarrow{\sf Fset} sending \ell to ()\ast(\ell).

We denote the category of 𝖥𝗌𝖾𝗍{\sf Fset}-categories by 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat}. Note that it follows from 2.1 that degida=0\deg{\rm id}_{a}=0 for any object aa of a 𝖥𝗌𝖾𝗍{\sf Fset}-category. We also note that the category of [0,][0,\infty]-categories is exactly 𝖦𝖬𝖾𝗍{\sf GMet}. By Proposition 3.4, we have the following.

Proposition 3.5.

There are embeddings 𝖢𝖺𝗍𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Cat}\longrightarrow{\sf Fsetcat} and 𝖦𝖬𝖾𝗍𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf GMet}\longrightarrow{\sf Fsetcat}.

In the following, we denote the full subcategory of 𝖥𝗌𝖾𝗍{\sf Fset} that consists of collectable filtered sets by 𝖢𝖥𝗌𝖾𝗍{\sf CFset}. Note that 𝖢𝖥𝗌𝖾𝗍{\sf CFset} is again a symmetric monoidal category. We also denote the full subcategory of 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat} that consists of 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-categories by 𝖢𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf CFsetcat}.

Remark 3.6.

While we define a filtered set by an “ascending filtration”, we can define it in the descending manner as follows.

  1. (1)

    A 0\mathbb{R}_{\geq 0}-filtered set is a set XX with subsets XXX_{\ell}\subset X for any 0\ell\in\mathbb{R}_{\geq 0} satisfying that X0=XX_{0}=X, XXX_{\ell^{\prime}}\subset X_{\ell} for any \ell\leq\ell^{\prime} and 0X=\bigcap_{\ell\in\mathbb{R}_{\geq 0}}X_{\ell}=\emptyset.

  2. (2)

    Let XX be a filtered set, and let xXx\in X. We define that degx=\deg x=\ell if xX<Xx\in X_{\ell}\setminus\bigcup_{\ell<\ell^{\prime}}X_{\ell^{\prime}}.

  3. (3)

    For filtered sets XX and YY, a filtered map f:XYf:X\longrightarrow Y is a map satisfying that f1YXf^{-1}Y_{\ell}\subset X_{\ell} for any 0\ell\in\mathbb{R}_{\geq 0}.

The categories of filtered sets by the above two definitions are equivalent, however, it seems better to work with descending one since it has good compatibility with the definitions of normed or filtered rings appearing in 3.2 and 4.3. Nevertheless, we adopt the ascending one to give a priority to intuitive understanding of the embeddings 𝖢𝖺𝗍𝖥𝗌𝖾𝗍𝖼𝖺𝗍𝖦𝖬𝖾𝗍{\sf Cat}\longrightarrow{\sf Fsetcat}\longleftarrow{\sf GMet}.

3.2 Novikov series ring

To define the magnitude, we first recall the definition of Novikov series ring following [18]. As explained in [18], the magnitude can take any semi-ring as values. However, to extend the definition in a natural way, we fix a standard one below. We remark that we can choose other suitable valuation ring such as pp-adic integers.

Definition 3.7.

We say a function f:f:\mathbb{R}\longrightarrow\mathbb{Q} is left finite if the support of ff restricted to L\mathbb{R}_{\leq L} is a finite set for any LL\in\mathbb{R}. We give a “power series like” expression for a left finite function f=f()qf=\sum_{\ell}f(\ell)q^{\ell}. The universal Novikov series field ((q))={f:f is left finite}\mathbb{Q}((q^{\mathbb{R}}))=\{f:\mathbb{R}\longrightarrow\mathbb{Q}\mid f\text{ is left finite}\} is defined by pointwise addition and the product fg()=+′′=f()g(′′)f\cdot g(\ell)=\sum_{\ell^{\prime}+\ell^{\prime\prime}=\ell}f(\ell^{\prime})g(\ell^{\prime\prime}). Let [[q]]\mathbb{Q}[[q^{\mathbb{R}}]] be its subring that consists of left finite functions 0\mathbb{R}_{\geq 0}\longrightarrow\mathbb{Q}, called a Novikov series ring.

Note that ((x))\mathbb{Q}((x^{\mathbb{R}})) is a non-Archimedean complete valuation field and [[q]]\mathbb{Q}[[q^{\mathbb{R}}]] is its valuation ring. In particular, we have [[q]]lim[[q]]/(q)\mathbb{Q}[[q^{\mathbb{R}}]]\cong\varprojlim\mathbb{Q}[[q^{\mathbb{R}}]]/(q^{\ell}), where (q)={f:0f is left finite and its support is 0}(q^{\ell})=\{f:\mathbb{R}_{\geq 0}\longrightarrow\mathbb{Q}\mid f\text{ is left finite and its support is }\geq\ell\geq 0\}. In the following, we denote [[q]]\mathbb{Q}[[q^{\mathbb{R}}]] and (q)(q^{\ell}) by AA and mm_{\ell} respectively for readability. To simplify discussions, we introduce the notion of normed groups and rings. See [2], for example.

Definition 3.8.
  1. (1)

    A pair (M,||)(M,|-|) of an abelian group MM and a function ||:M0|-|:M\longrightarrow\mathbb{R}_{\geq 0} is a normed group if it satisfies that

    • |x|=0|x|=0 if and only if x=0x=0.

    • |xy||x|+|y||x-y|\leq|x|+|y| for any x,yMx,y\in M.

    We say that (M,||)(M,|-|) is complete if it is complete with respect to the distance function d(x,y)=|xy|d(x,y)=|x-y|.

  2. (2)

    A pair (R,||)(R,|-|) of a (possibly non-commutative) ring RR and a function ||:R0|-|:R\longrightarrow\mathbb{R}_{\geq 0} is a normed ring if it satisfies that

    • (R,||)(R,|-|) is a normed group with respect to the additive structure.

    • |xy||x||y||xy|\leq|x||y| for any x,yRx,y\in R.

    We say that (R,||)(R,|-|) is complete or a Banach ring if it is complete as a normed group.

  3. (3)

    Let (M,||)(M,|-|) be a normed group and (R,||)(R,|-|) be a normed ring. We say that (M,||)(M,|-|) is a normed (left, right, bi-) (R,||)(R,|-|)-module if MM is a (left, right, bi-) RR-module and satisfies that |rm|,|mr||r||m||rm|,|mr|\leq|r||m| for any rRr\in R and mMm\in M.

Note that the addition and the multiplication of normed groups and normed rings are continuous with respect to the metric induced from the norm. We also note that the action of a normed module is continuous.

Example 3.9.
  1. (1)

    For fAf\in A, we define |f|:=sup{efm}|f|:=\sup\{e^{-\ell}\mid f\in m_{\ell}\}, which makes (A,||)(A,|-|) a complete normed ring.

  2. (2)

    Let SS be a set. We equip sets of maps AS={f:SA}A^{S}=\{f:S\longrightarrow A\} and AS×S={f:S×SA}A^{S\times S}=\{f:S\times S\longrightarrow A\} with norms defined by |f|:=supsS|f(s)||f|:=\sup_{s\in S}|f(s)| and |f|:=sups,tS|f(s,t)||f|:=\sup_{s,t\in S}|f(s,t)| respectively. Then (AS,||)(A^{S},|-|) and (AS×S,||)(A^{S\times S},|-|) are complete normed groups.

Definition 3.10.

Let SS be a set.

  1. (1)

    For every 0\ell\in\mathbb{R}_{\geq 0}, we denote the set of functions f:S×SA/mf:S\times S\longrightarrow A/m_{\ell} such that

    {for any sS,f(s,t)0 for finitely many tS,for any tS,f(s,t)0 for finitely many sS\begin{cases}\text{for any }s\in S,f(s,t)\neq 0\text{ for finitely many }t\in S,\\ \text{for any }t\in S,f(s,t)\neq 0\text{ for finitely many }s\in S\end{cases}

    by MS(A/m)M_{S}(A/m_{\ell}). We equip MS(A/m)M_{S}(A/m_{\ell}) with a ring structure by pointwise addition and the product defined by fg(s,t)=uf(s,u)g(u,t)f\cdot g(s,t)=\sum_{u}f(s,u)g(u,t). Note that we have a projection MS(A/m)MS(A/m)M_{S}(A/m_{\ell^{\prime}})\longrightarrow M_{S}(A/m_{\ell}) for any \ell^{\prime}\geq\ell.

  2. (2)

    We define a ring MS(A)M_{S}(A) by MS(A)=limMS(A/m)M_{S}(A)=\varprojlim M_{S}(A/m_{\ell}). Namely, we have MS(A)={f:S×SAπfMS(A/m) for any 0}M_{S}(A)=\{f:S\times S\longrightarrow A\mid\pi_{\ell}f\in M_{S}(A/m_{\ell})\text{ for any }\ell\in\mathbb{R}_{\geq 0}\}. Here we denote the natural projection AA/mA\longrightarrow A/m_{\ell} by π\pi_{\ell}.

We consider MS(A)M_{S}(A) as a subspace of AS×SA^{S\times S}. Note that, for any fAf\in A, the set {f+m0}\{f+m_{\ell}\mid\ell\in\mathbb{R}_{\geq 0}\} is a fundamental system of neighborhoods for ff. Note also that the set {f+m0}\{f+m_{\ell}\mid\ell\in\mathbb{R}_{\geq 0}\} is a fundamental system of neighborhoods for any fAS×Sf\in A^{S\times S}. Here we denote mS×Sm_{\ell}^{S\times S} by mm_{\ell}.

Lemma 3.11.

The subspace MS(A)M_{S}(A) of AS×SA^{S\times S} is closed. Hence (MS(A),||)(M_{S}(A),|-|) is a complete normed ring, where the norm |||-| is induced from (AS×S,||)(A^{S\times S},|-|).

Proof.

Let fAS×SMS(A)f\in A^{S\times S}\setminus M_{S}(A). Then there exist an 0\ell\in\mathbb{R}_{\geq 0} and an sSs\in S such that πf(s,t)0\pi_{\ell}f(s,t)\neq 0 for infinitely many tSt\in S or πf(t,s)0\pi_{\ell}f(t,s)\neq 0 for infinitely many tSt\in S. If we take L>L>\ell, then any element of f+mLf+m_{L} has the same property. Hence we have f+mLAS×SMS(A)f+m_{L}\subset A^{S\times S}\setminus M_{S}(A), which implies that MS(A)M_{S}(A) is closed. This completes the proof. ∎

The following is straightforward.

Lemma 3.12.

The normed abelian groups ASA^{S} and AS×SA^{S\times S} are (MS(A),||)(M_{S}(A),|-|)-bimodule. Here, the left and the right action of MS(A)M_{S}(A) on ASA^{S} is defined by fv(s)=tSf(s,t)v(t)fv(s)=\sum_{t\in S}f(s,t)v(t) and vf(s)=tSv(t)f(t,s)vf(s)=\sum_{t\in S}v(t)f(t,s) for any fMS(A),vASf\in M_{S}(A),v\in A^{S} and s,tSs,t\in S.

When we take SS as the set of objects ObC\mathrm{Ob}C for some category CC, we abbreviate ObC\mathrm{Ob}C as CC for the notations MS(A),AS,AS×SM_{S}(A),A^{S},A^{S\times S} and so on.

3.3 Magnitude of enriched categories

In the following, we give a general framework for the magnitude of enriched categories, which is slightly broader than the original one ([15], [18]), namely we relax the requirement of finiteness of number of objects. This subsection is basically a slight generalization of arguments by Leinster, for example in [15].

Definition 3.13.

Let (V,,1)(V,\otimes,1) be a symmetric monoidal category, and CC be a category enriched over VV.

  1. (1)

    A map #:ObVA\#:\mathrm{Ob}V\longrightarrow A is called a size function if it satisfies the following :

    1. (a)

      It holds that #u=#v\#u=\#v for any uvObVu\cong v\in\mathrm{Ob}V.

    2. (b)

      It holds that #(1)=1\#(1)=1 and #(uv)=#(u)#(v)\#(u\otimes v)=\#(u)\#(v) for any u,vObVu,v\in\mathrm{Ob}V.

  2. (2)

    We say that CC is of #\#-finite type (or simply of finite type) if the function ζCAC×C\zeta_{C}\in A^{C\times C} defined by ζC(a,b)=#C(a,b)\zeta_{C}(a,b)=\#C(a,b) is in MC(A)M_{C}(A).

  3. (3)

    We say that CC is finite if it has finitely many objects.

  4. (4)

    Suppose that CC is of finite type. A function kACk^{\bullet}\in A^{C} is a weighting if it satisfies that ζCk=1\zeta_{C}k^{\bullet}=1, where 1AC;a11\in A^{C};a\mapsto 1 for any aObCa\in\mathrm{Ob}C. Similarly, a function kACk_{\bullet}\in A^{C} is a coweighting if it satisfies that kζC=1k_{\bullet}\zeta_{C}=1.

  5. (5)

    Suppose that CC is finite. If CC has both a weighting and a coweighting, we define the magnitude of CC by 𝖬𝖺𝗀C=bkb=aka{\sf Mag}C=\sum_{b}k^{b}=\sum_{a}k_{a}. It is straightforward to check that this is well-defined and does not depend on the choice of a weighting and a coweighting.

Note that CC is of finite type if it is finite. The proof of the following is same as that of Lemma 2.3 of [17] together with Lemma 3.12.

Proposition 3.14.

Let CC be a VV-category of of finite type. If ζC\zeta_{C} is invertible in MC(A)M_{C}(A), then it has a unique weighting and a coweighting, and it holds that kb=aζC1(b,a),ka=bζC1(b,a)k^{b}=\sum_{a}\zeta^{-1}_{C}(b,a),k_{a}=\sum_{b}\zeta^{-1}_{C}(b,a). Further, if CC is finite, it holds that 𝖬𝖺𝗀C=a,bζC1(a,b){\sf Mag}C=\sum_{a,b}\zeta^{-1}_{C}(a,b).

3.4 Magnitude of 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-categories

3.4.1 A size function on 𝖢𝖥𝗌𝖾𝗍{\sf CFset}

Now we define a size function on 𝖢𝖥𝗌𝖾𝗍{\sf CFset} in a sense of Definition 3.13. Then we can consider the magnitude of finite 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-categories and the magnitude (co)weighting for ones of finite type. It is straightforward to verify that the following indeed defines a size function.

Definition 3.15.

For X𝖢𝖥𝗌𝖾𝗍X\in{\sf CFset}, we define a size function #\# by #X=xXqdegx[[q]]\#X=\sum_{x\in X}q^{\deg x}\in\mathbb{Q}[[q^{\mathbb{R}}]].

In the following, we just say a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category CC is finite or of finite type if it is so with respect to the above size function. The following is immediate from the definition, but is a useful characterization of 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category of finite type. For a morphism fMorCf\in\mathrm{Mor}C, we denote its domain and codomain by sfsf and tftf respectively.

Lemma 3.16.

A 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category CC is of finite type if and only if it satisfies one of the following equivalent conditions :

  1. (1)

    For any object aObCa\in\mathrm{Ob}C and 0\ell\in\mathbb{R}_{\geq 0}, there are finitely many morphisms fMorCf\in\mathrm{Mor}C with degf\deg f\leq\ell satisfying that sf=asf=a or tf=atf=a.

  2. (2)

    For any objects a,bObCa,b\in\mathrm{Ob}C, both of the filtered sets aObCC(a,b)\bigcup_{a\in\mathrm{Ob}C}C(a,b) and bObCC(a,b)\bigcup_{b\in\mathrm{Ob}C}C(a,b) are collectable.

3.4.2 Tame categories and an explicit formula for magnitude

Next we define a large class of 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category CC of finite type, namely tame categories, whose magnitude is computable. For f,gMorCf,g\in\mathrm{Mor}C, we write fgf\sim g if tf=sgtf=sg.

Definition 3.17.

Let CC be a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category, and let a,bObCa,b\in\mathrm{Ob}C.

  1. (1)

    A non-degenerate nn-path on CC from aa to bb is a sequence of morphisms f1fnf_{1}\sim\dots\sim f_{n} such that sf1=a,tfn=bsf_{1}=a,tf_{n}=b and fiidf_{i}\neq{\rm id}_{\bullet} for any ii. We define the length of such a sequence f1fnf_{1}\sim\dots\sim f_{n} by i=1ndegfi\sum_{i=1}^{n}\deg f_{i}.

  2. (2)

    A non-degenerate path on CC is a non-degenerate nn-path from aa to bb for some n1n\geq 1 and a,bObCa,b\in\mathrm{Ob}C.

  3. (3)

    We say that CC is quasi-tame if it is of finite type, and there are finitely many non-degenerate paths from aa to bb with length \leq\ell for any a,bObCa,b\in\mathrm{Ob}C and 0\ell\in\mathbb{R}_{\geq 0}.

  4. (4)

    We say that CC is tame if it is of finite type, and there exists an N>0N_{\ell}>0 for any 0\ell\in\mathbb{R}_{\geq 0} such that any non-degenerate NN_{\ell}-path have length >>\ell.

Lemma 3.18.

Let CC be a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category of finite type, and let a,bObCa,b\in\mathrm{Ob}C.

  1. (1)

    The number of non-degenerate nn-paths from aa to bb with length \ell is the coefficient of qq^{\ell} in (ζC1)n(a,b)(\zeta_{C}-1)^{n}(a,b).

  2. (2)

    CC is quasi-tame if and only if n=0(1ζC)n(a,b)\sum_{n=0}^{\infty}(1-\zeta_{C})^{n}(a,b) converges in AA for any a,bObCa,b\in\mathrm{Ob}C.

  3. (3)

    CC is tame if and only if n=0(1ζC)n\sum_{n=0}^{\infty}(1-\zeta_{C})^{n} converges in MC(A)M_{C}(A).

Proof.

(1) is obvious. In general, n=0fn\sum_{n=0}^{\infty}f_{n} converges if and only if fnf_{n} converges to 0 for any (fn)nA(f_{n})_{n}\subset A or (fn)nMS(A)(f_{n})_{n}\subset M_{S}(A) since these two metric spaces are complete. Hence (2) and (3) follows. ∎

Lemma 3.19.

A 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category CC of finite type is quasi-tame if it is tame. Further, the converse is true if CC is finite.

Proof.

It follows from Lemma 3.18. ∎

The following is a list of tame categories. They are discussed in the next subsection in detail, including the proof of tameness.

Example 3.20.
  1. (1)

    Any finite metric space is a finite tame category.

  2. (2)

    Any (possibly infinite) digraph is a tame category if it is locally finite, namely, if any vertex is an endpoint of finitely many edges.

  3. (3)

    Any finite category is a tame category if and only if it is skeletal and has no non-trivial endomorphisms. In particular, any finite poset is a finite tame category.

We have the following criterion for tameness.

Lemma 3.21.

Let CC be a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category of finite type. If there exists an ε>0\varepsilon>0 such that any non-identity morphism fMorCf\in\mathrm{Mor}C satisfies that degfε\deg f\geq\varepsilon, then CC is tame.

Proof.

The existence of such an ε>0\varepsilon>0 guarantees that (ζC1)mε(\zeta_{C}-1)\in m_{\varepsilon}. Hence it follows from Lemma 3.18 (3). This completes the proof. ∎

We call a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category satisfying the condition of Lemma 3.21 uniform. The following gives us an explicit formula of ζC1\zeta^{-1}_{C} for tame categories.

Proposition 3.22.

Let CC be a tame category. Then we have

ζC1(a,b)=δ(a,b)+k=1(1)kf1fk,fiid,sf1=a,tfk=bqi=1kdegfi,\zeta_{C}^{-1}(a,b)=\delta(a,b)+\sum_{k=1}^{\infty}(-1)^{k}\sum_{\begin{subarray}{c}f_{1}\sim\dots\sim f_{k},\\ f_{i}\neq{\rm id}_{\bullet},\\ sf_{1}=a,tf_{k}=b\end{subarray}}q^{\sum_{i=1}^{k}\deg f_{i}},

for any a,bObCa,b\in\mathrm{Ob}C.

Proof.

Note that the RHS is equal to (n=0(1ζC)n)(a,b)\left(\sum_{n=0}^{\infty}(1-\zeta_{C})^{n}\right)(a,b), which converges by Lemma 3.18. Since MC(A)M_{C}(A) is a topological ring, we have

ζCn=0(1ζC)n\displaystyle\zeta_{C}\sum_{n=0}^{\infty}(1-\zeta_{C})^{n} =(1(1ζC))n=0(1ζC)n\displaystyle=(1-(1-\zeta_{C}))\sum_{n=0}^{\infty}(1-\zeta_{C})^{n}
=n=0(1ζC)n(n=0(1ζC)n1)\displaystyle=\sum_{n=0}^{\infty}(1-\zeta_{C})^{n}-\left(\sum_{n=0}^{\infty}(1-\zeta_{C})^{n}-1\right)
=1.\displaystyle=1.

This completes the proof. ∎

The following is immediate from Proposition 3.14 and Proposition 3.22.

Corollary 3.23.

Let CC be a tame category CC. Then we have a weighting and a coweighting

ka=1+k=1(1)kf1fk,fiid,sf1=aqi=1kdegfi,k^{a}=1+\sum_{k=1}^{\infty}(-1)^{k}\sum_{\begin{subarray}{c}f_{1}\sim\dots\sim f_{k},\\ f_{i}\neq{\rm id}_{\bullet},sf_{1}=a\end{subarray}}q^{\sum_{i=1}^{k}\deg f_{i}},

and

kb=1+k=1(1)kf1fk,fiid,tfk=bqi=1kdegfi,k_{b}=1+\sum_{k=1}^{\infty}(-1)^{k}\sum_{\begin{subarray}{c}f_{1}\sim\dots\sim f_{k},\\ f_{i}\neq{\rm id}_{\bullet},tf_{k}=b\end{subarray}}q^{\sum_{i=1}^{k}\deg f_{i}},

for any a,bObCa,b\in\mathrm{Ob}C. Further, if CC is finite, it holds that

𝖬𝖺𝗀C=#ObC+k=1(1)kf1fk,fiidqi=1kdegfi.{\sf Mag}C=\#\mathrm{Ob}C+\sum_{k=1}^{\infty}(-1)^{k}\sum_{\begin{subarray}{c}f_{1}\sim\dots\sim f_{k},\\ f_{i}\neq{\rm id}_{\bullet}\end{subarray}}q^{\sum_{i=1}^{k}\deg f_{i}}.
Remark 3.24.

We can prove Proposition 3.22 for any quasi-tame categories by a direct calculation as follows. We put the right hand side m(a,b)Am(a,b)\in A. Now we have

bζC(a,b)m(b,c)\displaystyle\sum_{b}\zeta_{C}(a,b)m(b,c)
=ζC(a,a)m(a,c)+baζC(a,b)δ(b,c)+bak=1(1)kζC(a,b)f1fk,fiid,sf1=b,tfk=cqi=1kdegfi\displaystyle=\zeta_{C}(a,a)m(a,c)+\sum_{b\neq a}\zeta_{C}(a,b)\delta(b,c)+\sum_{b\neq a}\sum_{k=1}^{\infty}(-1)^{k}\zeta_{C}(a,b)\sum_{\begin{subarray}{c}f_{1}\sim\dots\sim f_{k},\\ f_{i}\neq{\rm id}_{\bullet},\\ sf_{1}=b,tf_{k}=c\end{subarray}}q^{\sum_{i=1}^{k}\deg f_{i}}
=ζC(a,a)m(a,c)+ζC(a,c)(1δ(a,c))+(m(a,c)+δ(a,c)(ζC(a,c)δ(a,c)))\displaystyle=\zeta_{C}(a,a)m(a,c)+\zeta_{C}(a,c)(1-\delta(a,c))+(-m(a,c)+\delta(a,c)-(\zeta_{C}(a,c)-\delta(a,c)))
(ζC(a,a)1)k=1(1)kf1fk,fiid,sf1=a,tfk=cqi=1kdegfi\displaystyle-(\zeta_{C}(a,a)-1)\sum_{k=1}^{\infty}(-1)^{k}\sum_{\begin{subarray}{c}f_{1}\sim\dots\sim f_{k},\\ f_{i}\neq{\rm id}_{\bullet},\\ sf_{1}=a,tf_{k}=c\end{subarray}}q^{\sum_{i=1}^{k}\deg f_{i}}
=ζC(a,a)m(a,c)+ζC(a,c)(1δ(a,c))+(m(a,c)+δ(a,c)(ζC(a,c)δ(a,c)))\displaystyle=\zeta_{C}(a,a)m(a,c)+\zeta_{C}(a,c)(1-\delta(a,c))+(-m(a,c)+\delta(a,c)-(\zeta_{C}(a,c)-\delta(a,c)))
(ζC(a,a)1)(m(a,c)δ(a,c))\displaystyle-(\zeta_{C}(a,a)-1)(m(a,c)-\delta(a,c))
=ζC(a,c)δ(a,c)+δ(a,c)+ζC(a,a)δ(a,c)\displaystyle=-\zeta_{C}(a,c)\delta(a,c)+\delta(a,c)+\zeta_{C}(a,a)\delta(a,c)
=δ(a,c).\displaystyle=\delta(a,c).

Note that the function m:ObC×ObCAm:\mathrm{Ob}C\times\mathrm{Ob}C\longrightarrow A is in MC(A)M_{C}(A). If not, there should be an object aObCa\in\mathrm{Ob}C and 0\ell\in\mathbb{R}_{\geq 0} such that there are infinitely objects bλObCb_{\lambda}\in\mathrm{Ob}C and non-degenerate paths from aa to bλb_{\lambda}’s or from bλb_{\lambda}’s to aa with length \leq\ell. It implies that there are infinitely many morphisms fMorCf\in\mathrm{Mor}C with sf=asf=a or tf=atf=a with degf\deg f\leq\ell, which contradicts the assumption that CC is of finite type.

3.5 Examples

In the following, we show examples of quasi-tame or tame categories and their magnitude or (co)weighting. Some of them are examples of the magnitude or Euler characteristic in the sense of the original definitions in [17] or [15]. However, we need to extend the definition to include 3.5.4.

3.5.1 Magnitude of generalized metric spaces

As stated in Proposition 3.5, any generalized metric spaces, in particular any metric spaces are 𝖥𝗌𝖾𝗍{\sf Fset}-categories. Since hom-objects of those consist of one element filtered sets ()\ast(\ell), they are also 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-categories. We have the following characterizations.

Lemma 3.25.
  1. (1)

    A generalized metric space XX is of finite type as a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category if and only if the closed balls B(x,)B(x,\ell) and B(,x)B(\ell,x) are finite sets for any xXx\in X and 0\ell\in\mathbb{R}_{\geq 0}.

  2. (2)

    A generalized metric space XX of finite type is quasi-tame if and only if it is non-degenerate.

  3. (3)

    A generalized metric space XX of finite type is tame if there exists an ε>0\varepsilon>0 such that d(x,y)>εd(x,y)>\varepsilon for any xyXx\neq y\in X.

Proof.
  1. (1)

    It is obvious from Lemma 3.16.

  2. (2)

    If XX is not non-degenerate, namely there exists points x,yXx,y\in X with d(x,y)=d(y,x)=0d(x,y)=d(y,x)=0 and xyx\neq y, then XX is not quasi-tame by the definition. If XX is not quasi-tame, conversely, there is an infinite family {(x,x1λ,,xnλλ,y)}λ\{(x,x^{\lambda}_{1},\dots,x^{\lambda}_{n_{\lambda}},y)\}_{\lambda} of non-degenerate paths with the length \leq\ell for some x,yXx,y\in X and 0\ell\in\mathbb{R}_{\geq 0}. Since XX is of finite type and the set {xiλ}λ,i\{x^{\lambda}_{i}\}_{\lambda,i} is a subset of B(x,)B(x,\ell), it is finite. Hence there should be an infinite sequence {aj}j\{a_{j}\}_{j} with aj{xiλ}λ,i,ajaj+1a_{j}\in\{x^{\lambda}_{i}\}_{\lambda,i},a_{j}\neq a_{j+1} and d(aj,aj+1)=0d(a_{j},a_{j+1})=0. It implies that there is a pair of points (xiλ0,xjλ1)(x^{\lambda_{0}}_{i},x^{\lambda_{1}}_{j}) with d(xiλ0,xjλ1)=d(xiλ0,xjλ1)=0d(x^{\lambda_{0}}_{i},x^{\lambda_{1}}_{j})=d(x^{\lambda_{0}}_{i},x^{\lambda_{1}}_{j})=0 and xiλ0xjλ1x^{\lambda_{0}}_{i}\neq x^{\lambda_{1}}_{j}, hence XX is not non-degenerate.

  3. (3)

    It follows from Lemma 3.21.

This completes the proof. ∎

By the above lemma, any finite metric spaces are of finite type and tame. We note that a metric space is a finite 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category if and only if it is a finite metric space. Hence we obtain the following.

Corollary 3.26.
  1. (1)

    A generalized metric space is a finite 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category if and only if it is a finite generalized metric space.

  2. (2)

    A generalized metric space is a finite tame category if and only if it is a finite non-degenerate generalized metric space.

  3. (3)

    A metric space is a finite tame category if and only if it is a finite metric space.

It is easy to see that the definition of the magnitude in Definition 3.13 is restricted to the original one for finite metric spaces introduced by Leinster ([16]).

3.5.2 Euler characteristic of finite categories

As stated in Proposition 3.5, any small categories are 𝖥𝗌𝖾𝗍{\sf Fset}-categories. We note that a small category is a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category if and only if each hom-set is a finite set. Furthermore, we also note that a finite category is obviously of finite type as a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category. Hence we have the following.

Lemma 3.27.

A small category is a finite 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category if and only if it is a finite category.

The following shows that many finite categories including finite posets are tame. See also Proposition 2.11 of [17]

Lemma 3.28.

A finite category CC is a tame category if and only if it is skeletal and has no non-trivial endomorphisms.

Proof.

It can be proved similarly to the proof of Lemma 3.25 (2). ∎

It is easy to see that the definition of the magnitude in Definition 3.13 is restricted to the definition of Euler characteristic for finite categories introduced by Leinster ([17]).

We remark that any small category can be considered as a 𝖥𝗌𝖾𝗍{\sf Fset}-category in another way. For a small category CC and fMorCf\in\mathrm{Mor}C, we consider ff as an element of a filtered set by setting degf=1\deg f=1 if ff is not an identity, and degid=0\deg{\rm id_{\bullet}}=0. It is easy to see that CC is a 𝖥𝗌𝖾𝗍{\sf Fset}-category in this setting. When CC is a preordered set, it is same as considering CC as a digraph whose directed edges correspond to the relation \leq. Then it is also a generalized metric space.

3.5.3 Euler characteristic of finite simplicial complexes

As explained in the last paragraph of 3.5.2, any poset can be considered as a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category by setting the degree of all relations by 11 except for identities which have degree 0. In this setting, any finite poset PP is a finite non-degenerate generalized metric space, hence is a finite tame category by Lemma 3.26. Therefore Corollary 3.23 implies that

𝖬𝖺𝗀P\displaystyle{\sf Mag}P =#P+k=1(1)ka0akqd(a0,a1)++d(ak1,ak)\displaystyle=\#P+\sum_{k=1}^{\infty}(-1)^{k}\sum_{a_{0}\neq\dots\neq a_{k}}q^{d(a_{0},a_{1})+\dots+d(a_{k-1},a_{k})}
=#P+k=1(1)ka0<<akqk\displaystyle=\#P+\sum_{k=1}^{\infty}(-1)^{k}\sum_{a_{0}<\dots<a_{k}}q^{k}
=k=0dimΔ(P)(1)k#{a0<<ak}qk,\displaystyle=\sum_{k=0}^{\dim\Delta(P)}(-1)^{k}\#\{a_{0}<\dots<a_{k}\}q^{k},

where we denote the order complex of PP by Δ(P)\Delta(P). In particular, the magnitude of a finite poset in this setting is a polynomial. It implies that (𝖬𝖺𝗀P)|q=1({\sf Mag}P)|_{q=1} is equal to the Euler characteristic of Δ(P)\Delta(P). When PP is a face poset of a finite simplicial complex SS, we obtain that (𝖬𝖺𝗀P)|q=1=χ(S)({\sf Mag}P)|_{q=1}=\chi(S). Hence the magnitude covers the ordinary Euler characteristic of finite simplicial complexes. We again refer to Proposition 2.11 of [17].

3.5.4 Growth seriess of finitely generated groups

Let Γ\Gamma be a finitely generated group with generators SS. We consider (Γ,S)(\Gamma,S) as a 𝖥𝗌𝖾𝗍{\sf Fset}-category by Ob(Γ,S)={}\mathrm{Ob}(\Gamma,S)=\{\ast\} and Mor(Γ,S)=Γ\mathrm{Mor}(\Gamma,S)=\Gamma. Here we define the degree of gMor(Γ,S)g\in\mathrm{Mor}(\Gamma,S) by its word length denoted by wl in the following. Then it is a finite tame category. Hence we can consider its magnitude, and we have 𝖬𝖺𝗀(Γ,S)=(gΓqwlg)1[[q]]{\sf Mag}(\Gamma,S)=(\sum_{g\in\Gamma}q^{{\rm wl}g})^{-1}\in\mathbb{Q}[[q^{\mathbb{R}}]], which is exactly the inverse of the growth series of (Γ,S)(\Gamma,S). On the other hand, let Cay(Γ,S){\rm Cay}(\Gamma,S) be the Cayley graph of (Γ,S)(\Gamma,S), and we consider it as a metric space by the path metric. Then Cay(Γ,S){\rm Cay}(\Gamma,S) is a tame category by Lemma 3.25, and we have a weighting w=(gΓqwlg)1w_{\bullet}=(\sum_{g\in\Gamma}q^{{\rm wl}g})^{-1} . We note that this coincidence is not accidental, which will be explained in Example 5.20.

3.5.5 Poincaré polynomials of ranked posets

In the following, we see that the well-known invariant the Poincaré polynomial is a magnitude weighting.

Definition 3.29.

A poset PP with the minimum element 0 is a ranked poset if it is equipped with a rank function r:P0r:P\longrightarrow\mathbb{Z}_{\geq 0} satisfying

{r(0)=0,r(b)>r(a)if a<b,r(b)=r(a)+1if b covers a.\begin{cases}r(0)=0,\\ r(b)>r(a)&\text{if $a<b$},\\ r(b)=r(a)+1&\text{if $b$ covers $a$}.\end{cases}

Here, we say that bb covers aa if {cacb}={a,b}\{c\mid a\leq c\leq b\}=\{a,b\}.

Example 3.30.

Let VV be a vector space over an arbitrary field. A finite collection of affine subspaces of VV is called subspace arrangement. For a subspace arrangement SS, let I(S)={tTtTS}I(S)=\{\cap_{t\in T}t\mid T\subset S\}, where we set tt=V\cap_{t\in\emptyset}t=V. We equip I(S)I(S) with a poset structure by defining xyx\leq y if xyx\supset y. Then I(S)I(S) has the minimum element VV, and we can define a rank function by r(x)=codimxr(x)={\rm codim\ }x with codim=dimV+1{\rm codim\ }\emptyset=\dim V+1, which makes it a ranked poset.

We metrize a finite ranked poset (P,r)(P,r) by d(x,y)={r(b)r(a)if ab,otherwise.d(x,y)=\begin{cases}r(b)-r(a)&\text{if }a\leq b,\\ \infty&\text{otherwise}.\end{cases} Note that this is equivalent to considering the directed Hasse diagram of PP as a generalized metric space, where a directed edge is spanned from aa to bb if bb covers aa. By Lemma 3.26, it is a finite tame category. Then we have the following by Proposition 3.22 and Corollary 3.23.

Proposition 3.31.

For a finite ranked poset (P,φ)(P,\varphi), we have

ζP1(a,b)=δ(a,b)+k=1(1)ka0<<ak,a0=a,ak=bqr(b)r(a),\zeta_{P}^{-1}(a,b)=\delta(a,b)+\sum_{k=1}^{\infty}(-1)^{k}\sum_{\begin{subarray}{c}a_{0}<\dots<a_{k},\\ a_{0}=a,a_{k}=b\end{subarray}}q^{r(b)-r(a)},

and

ka=1+bk=1(1)ka0<<ak,a0=a,ak=bqr(b)r(a).k^{a}=1+\sum_{b}\sum_{k=1}^{\infty}(-1)^{k}\sum_{\begin{subarray}{c}a_{0}<\dots<a_{k},\\ a_{0}=a,a_{k}=b\end{subarray}}q^{r(b)-r(a)}.

Now we recall the Möbius function of a poset.

Definition 3.32.

For a finite poset PP, we define a square matrix ξPM|P|()\xi_{P}\in M_{|P|}(\mathbb{Z}) by ξP(x,y)={1if xy,0otherwise.\xi_{P}(x,y)=\begin{cases}1&\text{if }x\leq y,\\ 0&\text{otherwise}.\end{cases} Then μP:=ξP1M|P|()\mu_{P}:=\xi_{P}^{-1}\in M_{|P|}(\mathbb{Z}) exists, which we call the Möbius function of PP.

Note that, in the above definition, we consider a finite poset as a finite tame category, as in 3.5.3. Hence it has a magnitude, and the existence of ξP1\xi_{P}^{-1} follows by substituting q=1q=1 to the polynomial magnitude of it. Furthermore, we immediately obtain the following by Proposition 3.22.

Proposition 3.33.

We have

μP(a,b)=δ(a,b)+k=1(1)ka0<<ak,a0=a,ak=b1.\mu_{P}(a,b)=\delta(a,b)+\sum_{k=1}^{\infty}(-1)^{k}\sum_{\begin{subarray}{c}a_{0}<\dots<a_{k},\\ a_{0}=a,a_{k}=b\end{subarray}}1.

Note that we have ζP1|q=1=μP\zeta_{P}^{-1}|_{q=1}=\mu_{P}. The following definition is fundamental and pivotal in the study of subspace arrangements.

Definition 3.34.

The Poincaré polynomial πP\pi_{P} of a ranked poset PP with the minimum element 0 is defined by

πP(q):=aPμP(0,a)(q)r(a).\pi_{P}(q):=\sum_{a\in P}\mu_{P}(0,a)(-q)^{r(a)}.
Example 3.35.

Let SS be a subspace arrangement. Then the Poincaré polynomial of SS is defined as that of (I(S),codim)(I(S),{\rm codim}).

The following shows that the Poincaré polynomial of a ranked poset is essentially the weighting of PP at 0.

Proposition 3.36.
πP(q)=aPζP1(0,a)=k0.\pi_{P}(-q)=\sum_{a\in P}\zeta_{P}^{-1}(0,a)=k^{0}.
Proof.

By Propositions 3.31 and 3.33, we have

aPζP1(0,a)\displaystyle\sum_{a\in P}\zeta_{P}^{-1}(0,a) =1+0<ak=1(1)ka0<<ak,a0=0,ak=aqr(a)\displaystyle=1+\sum_{0<a}\sum_{k=1}^{\infty}(-1)^{k}\sum_{\begin{subarray}{c}a_{0}<\dots<a_{k},\\ a_{0}=0,a_{k}=a\end{subarray}}q^{r(a)}
=aqr(a)(δ(0,a)+k=1(1)ka0<<ak,a0=0,ak=a1)\displaystyle=\sum_{a}q^{r(a)}\left(\delta(0,a)+\sum_{k=1}^{\infty}(-1)^{k}\sum_{\begin{subarray}{c}a_{0}<\dots<a_{k},\\ a_{0}=0,a_{k}=a\end{subarray}}1\right)
=aqr(a)μP(0,a)\displaystyle=\sum_{a}q^{r(a)}\mu_{P}(0,a)
=πP(q).\displaystyle=\pi_{P}(-q).

This completes the proof. ∎

Remark 3.37.

The above coincidence of the Poincaré polynomial and the weighting can be generalized as follows. Let PP be a finite poset with the minimum element 0, equipped with an order preserving map φ:P0\varphi:P\longrightarrow\mathbb{Z}_{\geq 0} satisfying that φ(0)=0\varphi(0)=0. For such a PP, we define dφ:P×P[0,]d_{\varphi}:P\times P\longrightarrow[0,\infty] by dφ(a,b)={φ(b)φ(a)if ab,otherwise,d_{\varphi}(a,b)=\begin{cases}\varphi(b)-\varphi(a)&\text{if }a\leq b,\\ \infty&\text{otherwise},\end{cases} which makes (P,dφ)(P,d_{\varphi}) a non-degenerate finite generalized metric space, hence a finite tame category. The Poincaré polynomial π(P,φ)\pi_{(P,\varphi)} of such a PP is defined as π(P,φ)(q)=aPμP(0,a)(q)φ(a)\pi_{(P,\varphi)}(q)=\sum_{a\in P}\mu_{P}(0,a)(-q)^{\varphi(a)}. We can prove that π(P,φ)\pi_{(P,\varphi)} coincides with the weighting k0k^{0} of (P,dφ)(P,d_{\varphi}) by the same argument of Proposition 3.36.

Example 3.38.
  1. (1)

    For a ranked poset (P,r)(P,r), we can choose φ=r\varphi=r to adapt to the above situation. Then the definitions above coincide with the original ones.

  2. (2)

    Let SS be a subspace arrangement. The power set P(S)P(S) is naturally equipped with a poset structure by inclusion, that is, xyx\leq y if xyx\subset y. It has the minimum element \emptyset, and we can define an order preserving map φ:P(S)0\varphi:P(S)\longrightarrow\mathbb{Z}_{\geq 0} by φ(x)=codimtxt\varphi(x)={\rm codim}\cap_{t\in x}t. Then the Poincaré polynomial of (P(S),codim)(P(S),{\rm codim}) is the following :

    π(P(S),codim)(q)=TSμ(P(S),codim)(0,T)(q)codimtTt.\pi_{(P(S),{\rm codim})}(q)=\sum_{T\subset S}\mu_{(P(S),{\rm codim})}(0,T)(-q)^{{\rm codim\ }\cap_{t\in T}t}.

    It is natural to ask how (I(S),codim)(I(S),{\rm codim}) and (P(S),codim)(P(S),{\rm codim}) differ. We will see that the Poincaré polynomials of those coincide in Example 4.24.

4 Magnitude homology of filtered set enriched categories

In this section, we define the magnitude homology of 𝖥𝗌𝖾𝗍{\sf Fset}-categories as a functor to the category of bi-graded abelian groups, by following [18], and show its properties. From Examples 3.5.4 and 3.5.5, it turns out to be a categorification of the Poincaré polynomial and the growth series. After defining it, we describe the magnitude homology as a Hochschild homology of the “incidence algebra” of 𝖥𝗌𝖾𝗍{\sf Fset}-categories. We also show invariance of the magnitude homology under the adjointness of functors in the setting of filtered sets enrichment. Finally we consider a spectral sequence whose first page is isomorphic to the magnitude homology. We introduce a relation with Grigor’yan–Muranov–Lin–Yau’s path homology, studied by the author in [1]. We also discuss homotopy invariance of each page of this spectral sequence.

4.1 Definition and the categorification property

4.1.1 Un-normalized magnitude chain complex

First we prepare basic terminologies.

Definition 4.1.
  1. (1)

    An 0\mathbb{R}_{\geq 0}-filtered abelian group, or simply a filtered abelian group is a filtered set AA equipped with an abelian group structure such that deg(x+y)max{degx,degy}\deg(x+y)\leq\max\{\deg x,\deg y\} for any x,yAx,y\in A. A filtered homomorphism between filtered abelian groups is a filtered map that is also a group homomorphism. We denote the category of filtered abelian groups and homomorphisms by 𝖥𝖠𝖻{\sf FAb}. We define a functor :𝖥𝗌𝖾𝗍𝖥𝖠𝖻\mathbb{Z}:{\sf Fset}\longrightarrow{\sf FAb} by freely generating filtered abelian groups. We also define functors ,<:𝖥𝗌𝖾𝗍𝖥𝖠𝖻\mathbb{Z}_{\ell},\mathbb{Z}_{<\ell}:{\sf Fset}\longrightarrow{\sf FAb} by X=X\mathbb{Z}_{\ell}X=\mathbb{Z}X_{\ell} and <X=<X\mathbb{Z}_{<\ell}X=\mathbb{Z}\bigcup_{\ell^{\prime}<\ell}X_{\ell^{\prime}}. Note here that XX_{\ell} and <X\bigcup_{\ell^{\prime}<\ell}X_{\ell^{\prime}} are also filtered sets for any filtered set XX and 0\ell\in\mathbb{R}_{\geq 0}.

  2. (2)

    A filtered chain complex is a collection {n:AnAn1}n\{\partial_{n}:A_{n}\longrightarrow A_{n-1}\}_{n\in\mathbb{Z}} of filtered abelian groups AnA_{n}’s and filtered homomorphisms n\partial_{n}’s such that nn+1=0\partial_{n}\circ\partial_{n+1}=0 for any nn\in\mathbb{Z}. We suppose that chain complexes are non-negative, that is An=0A_{n}=0 for n<0n<0. A filtered chain map between filtered chain complexes is a family of filtered homomorphisms that is a usual chain map when forgetting the filtration. We denote the category of non-negative filtered chain complexes and filtered chain maps by 𝖥𝖢𝗁0{\sf FCh}_{\geq 0}.

  3. (3)

    A filtered simplicial set is an object of the functor category 𝖥𝗌𝖾𝗍Δop{\sf Fset}^{\Delta^{\rm op}}. A filtered simplicial abelian group is an object of the functor category 𝖥𝖠𝖻Δop{\sf FAb}^{\Delta^{\rm op}}. We have functors ,,<:𝖥𝗌𝖾𝗍Δop𝖥𝖠𝖻Δop\mathbb{Z},\mathbb{Z}_{\ell},\mathbb{Z}_{<\ell}:{\sf Fset}^{\Delta^{\rm op}}\longrightarrow{\sf FAb}^{\Delta^{\rm op}} induced from ,,<:𝖥𝗌𝖾𝗍𝖥𝖠𝖻\mathbb{Z},\mathbb{Z}_{\ell},\mathbb{Z}_{<\ell}:{\sf Fset}\longrightarrow{\sf FAb} respectively.

  4. (4)

    We define a functor 𝖢:𝖥𝖠𝖻Δop𝖥𝖢𝗁0{\sf C}_{\bullet}:{\sf FAb}^{\Delta^{\rm op}}\longrightarrow{\sf FCh}_{\geq 0} by 𝖢nA=An{\sf C}_{n}A_{\bullet}=A_{n} and n=i(1)idi\partial_{n}=\sum_{i}(-1)^{i}d_{i}.

The following is a refinement of the nerve functors for small categories in the setting of filtered sets enrichment.

Definition 4.2.

Let C𝖥𝗌𝖾𝗍𝖼𝖺𝗍C\in{\sf Fsetcat}. Let C¯\underline{C} be its underlying small category structure.

  1. (1)

    We define a filtered nerve functor 𝖥𝖭:𝖥𝗌𝖾𝗍𝖼𝖺𝗍𝖥𝗌𝖾𝗍Δop{\sf FN}_{\bullet}:{\sf Fsetcat}\longrightarrow{\sf Fset}^{\Delta^{\rm op}} by

    𝖥𝖭nC=aiObCC(a0,a1)×C(a1,a2)××C(an1,an).{\sf FN}_{n}C=\bigcup_{a_{i}\in\mathrm{Ob}C}C(a_{0},a_{1})\times C(a_{1},a_{2})\times\dots\times C(a_{n-1},a_{n}).

    The face and the degeneracy maps are defined similarly to the usual nerve 𝖭C¯{\sf N}_{\bullet}\underline{C}. We denote simplicial abelian groups 𝖥𝖭C\mathbb{Z}_{\ell}\circ{\sf FN}_{\bullet}C and <𝖥𝖭C\mathbb{Z}_{<\ell}\circ{\sf FN}_{\bullet}C by (𝖥𝖭C)(\mathbb{Z}{\sf FN}_{\bullet}C)_{\ell} and (𝖥𝖭C)<(\mathbb{Z}{\sf FN}_{\bullet}C)_{<\ell} respectively.

  2. (2)

    We denote the composition 𝖢𝖥𝖭:𝖥𝗌𝖾𝗍𝖼𝖺𝗍𝖥𝖢𝗁0{\sf C}_{\bullet}\circ\mathbb{Z}\circ{\sf FN}_{\bullet}:{\sf Fsetcat}\longrightarrow{\sf FCh}_{\geq 0} by 𝖥𝖢{\sf FC}_{\bullet}. Explicitly, we have the following for C𝖥𝗌𝖾𝗍𝖼𝖺𝗍C\in{\sf Fsetcat} :

    • 𝖥𝖢0C=ObC{\sf FC}_{0}C=\mathbb{Z}\mathrm{Ob}C,

    • 𝖥𝖢nC={(f1,,fn)(MorC)ntfi=sfi+1}{\sf FC}_{n}C=\mathbb{Z}\{(f_{1},\dots,f_{n})\in(\mathrm{Mor}C)^{n}\mid tf_{i}=sf_{i+1}\},

    • {(𝖥𝖢0C)0=𝖥𝖢0C,(𝖥𝖢nC)={(f1,,fn)𝖥𝖢nCidegfi},\begin{cases}({\sf FC}_{0}C)_{0}={\sf FC}_{0}C,\\ ({\sf FC}_{n}C)_{\ell}=\mathbb{Z}\{(f_{1},\dots,f_{n})\in{\sf FC}_{n}C\mid\sum_{i}\deg f_{i}\leq\ell\},\end{cases} as filtered sets.

    We also denote chain complexes 𝖢(𝖥𝖭C){\sf C}_{\bullet}(\mathbb{Z}{\sf FN}_{\bullet}C)_{\ell} and 𝖢(𝖥𝖭C)<{\sf C}_{\bullet}(\mathbb{Z}{\sf FN}_{\bullet}C)_{<\ell} by (𝖥𝖢C)({\sf FC}_{\bullet}C)_{\ell} and (𝖥𝖢C)<({\sf FC}_{\bullet}C)_{<\ell} respectively. Note that (𝖥𝖢C)({\sf FC}_{\bullet}C)_{\ell} and (𝖥𝖢C)<({\sf FC}_{\bullet}C)_{<\ell} are subchain complexes of 𝖥𝖢C{\sf FC}_{\bullet}C.

Now we define the un-normalized magnitude chain complex as follows.

Definition 4.3.

For C𝖥𝗌𝖾𝗍𝖼𝖺𝗍C\in{\sf Fsetcat} and 0\ell\in\mathbb{R}_{\geq 0}, we define the un-normalized magnitude chain complex 𝖬𝖢~C\widetilde{\operatorname{\sf MC}}^{\ell}_{\bullet}C of CC as the quotient chain complex (𝖥𝖢C)/(𝖥𝖢C)<({\sf FC}_{\bullet}C)_{\ell}/({\sf FC}_{\bullet}C)_{<\ell}. Explicitly, we define a chain complex (𝖬𝖢~C,)(\widetilde{\operatorname{\sf MC}}^{\ell}_{\bullet}C,\partial^{\ell}_{\bullet}) by

  • 𝖬𝖢~nC={(f1,,fn)(MorC)ntfi=sfi+1,idegfi=}\widetilde{\operatorname{\sf MC}}^{\ell}_{n}C=\mathbb{Z}\{(f_{1},\dots,f_{n})\in(\mathrm{Mor}C)^{n}\mid tf_{i}=sf_{i+1},\sum_{i}\deg f_{i}=\ell\},

  • n,i(f1,,fn)={(f1,,fi+1fi,,fn) if degfi+1fi=degfi+1+degfi,0otherwise,\partial^{\ell}_{n,i}(f_{1},\dots,f_{n})=\begin{cases}(f_{1},\dots,f_{i+1}\circ f_{i},\dots,f_{n})&\text{ if }\deg f_{i+1}\circ f_{i}=\deg f_{i+1}+\deg f_{i},\\ 0&\text{otherwise},\end{cases}

  • n=i(1)in,i\partial^{\ell}_{n}=\sum_{i}(-1)^{i}\partial^{\ell}_{n,i}.

4.1.2 Magnitude chain complex and the categorification property

We define the magnitude chain complex as the normalization of 𝖬𝖢~C\widetilde{\operatorname{\sf MC}}^{\ell}_{\bullet}C. To that end, we recall the following fundamental fact. See Section III-2 of [7], for example.

Lemma 4.4.

Let AA_{\bullet} be a simplicial abelian group. Let 𝖢A{\sf C}_{\bullet}A be the associated chain complex and 𝖢D{\sf C}_{\bullet}D be the subchain complex of degenerated simplices. Then the quotient 𝖢A𝖢A/𝖢D{\sf C}_{\bullet}A\longrightarrow{\sf C}_{\bullet}A/{\sf C}_{\bullet}D is a homotopy equivalence.

Since (𝖥𝖭C)<(\mathbb{Z}{\sf FN}_{\bullet}C)_{<\ell} is a subsimplicial abelian group of (𝖥𝖭C)(\mathbb{Z}{\sf FN}_{\bullet}C)_{\ell}, its quotient

(𝖥𝖭C)/(𝖥𝖭C)<(\mathbb{Z}{\sf FN}_{\bullet}C)_{\ell}/(\mathbb{Z}{\sf FN}_{\bullet}C)_{<\ell}

is again a simplicial abelian group. The chain complex of degenerated simplices 𝖢D{\sf C}_{\bullet}D of this quotient is a subchain complex of 𝖬𝖢~C\widetilde{\operatorname{\sf MC}}^{\ell}_{\bullet}C described as

𝖢nD={(f1,,fn)𝖬𝖢~nCfi=id for some i}.{\sf C}_{n}D=\mathbb{Z}\{(f_{1},\dots,f_{n})\in\widetilde{\operatorname{\sf MC}}^{\ell}_{n}C\mid f_{i}={\rm id}_{\bullet}\text{ for some }i\}.

Then the quotient 𝖬𝖢~C𝖬𝖢~C/𝖢D\widetilde{\operatorname{\sf MC}}^{\ell}_{\bullet}C\longrightarrow\widetilde{\operatorname{\sf MC}}^{\ell}_{\bullet}C/{\sf C}_{\bullet}D is a homotopy equivalence by Lemma 4.4. We define 𝖬𝖢C:=𝖬𝖢~C/𝖢D\operatorname{\sf MC}^{\ell}_{\bullet}C:=\widetilde{\operatorname{\sf MC}}^{\ell}_{\bullet}C/{\sf C}_{\bullet}D described as

  • 𝖬𝖢nC={(f1,,fn)(MorC)ntfi=sfi+1,idegfi=,fiid for any i}\operatorname{\sf MC}^{\ell}_{n}C=\mathbb{Z}\{(f_{1},\dots,f_{n})\in(\mathrm{Mor}C)^{n}\mid tf_{i}=sf_{i+1},\sum_{i}\deg f_{i}=\ell,f_{i}\neq{\rm id}_{\bullet}\text{ for any }i\},

  • n,i(f1,,fn)={(f1,,fi+1fi,,fn) if degfi+1fi=degfi+1+degfi and fi+1fiid,0otherwise,\partial^{\ell}_{n,i}(f_{1},\dots,f_{n})=\begin{cases}(f_{1},\dots,f_{i+1}\circ f_{i},\dots,f_{n})&\begin{subarray}{c}\text{ if }\deg f_{i+1}\circ f_{i}=\deg f_{i+1}+\deg f_{i}\\ \text{ and }f_{i+1}\circ f_{i}\neq{\rm id}_{\bullet}\end{subarray},\\ 0&\text{otherwise},\end{cases}

  • n=i(1)in,i\partial^{\ell}_{n}=\sum_{i}(-1)^{i}\partial^{\ell}_{n,i}.

We define the magnitude homology as the homology of the above homotopy equivalent chain complexes.

Definition 4.5.

We call the above chain complex 𝖬𝖢C\operatorname{\sf MC}^{\ell}_{\bullet}C the magnitude chain complex of CC. We denote its homology by 𝖬𝖧C\operatorname{\sf MH}^{\ell}_{\bullet}C.

Note that 𝖬𝖢\operatorname{\sf MC}^{\ell}_{\bullet} and 𝖬𝖧\operatorname{\sf MH}^{\ell}_{\bullet} define functors from 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat}.

Definition 4.6.

Let CC be a 𝖥𝗌𝖾𝗍{\sf Fset}-category and aObCa\in\mathrm{Ob}C. We say that aa is a T1T_{1}-object, or simply T1T_{1}, if there is no morphism fMorCf\in\mathrm{Mor}C such that sf=a or tf=a,degf=0sf=a\text{ or }tf=a,\deg f=0 and fidf\neq{\rm id}_{\bullet}. We also say that CC is T1T_{1} if every object is T1T_{1}.

If a,bObCa,b\in\mathrm{Ob}C are T1T_{1}-objects, then we have subsimplicial abelian groups 𝖥𝖭,aC\mathbb{Z}{\sf FN}^{\ell,a}_{\bullet}C and 𝖥𝖭,a,bC\mathbb{Z}{\sf FN}^{\ell,a,b}_{\bullet}C of 𝖥𝖭C:=(𝖥𝖭C)/(𝖥𝖭C)<\mathbb{Z}{\sf FN}^{\ell}_{\bullet}C:=(\mathbb{Z}{\sf FN}_{\bullet}C)_{\ell}/(\mathbb{Z}{\sf FN}_{\bullet}C)_{<\ell}, where

𝖥𝖭n,aC={(f1,,fn)𝖥𝖭nCsf1=a},\mathbb{Z}{\sf FN}^{\ell,a}_{n}C=\mathbb{Z}\{(f_{1},\dots,f_{n})\in\mathbb{Z}{\sf FN}^{\ell}_{n}C\mid sf_{1}=a\},

and

𝖥𝖭n,a,bC={(f1,,fn)𝖥𝖭nCsf1=a,tfn=b}.\mathbb{Z}{\sf FN}^{\ell,a,b}_{n}C=\mathbb{Z}\{(f_{1},\dots,f_{n})\in\mathbb{Z}{\sf FN}^{\ell}_{n}C\mid sf_{1}=a,tf_{n}=b\}.

We define chain complexes 𝖬𝖢~C=𝖢𝖥𝖭C\widetilde{\operatorname{\sf MC}}^{\ell}_{\bullet}C={\sf C}_{\bullet}\mathbb{Z}{\sf FN}^{\ell}_{\bullet}C, 𝖬𝖢~,aC=𝖢𝖥𝖭,aC\widetilde{\operatorname{\sf MC}}^{\ell,a}_{\bullet}C={\sf C}_{\bullet}\mathbb{Z}{\sf FN}^{\ell,a}_{\bullet}C and 𝖬𝖢~,a,bC=𝖢𝖥𝖭,a,bC\widetilde{\operatorname{\sf MC}}^{\ell,a,b}_{\bullet}C={\sf C}_{\bullet}\mathbb{Z}{\sf FN}^{\ell,a,b}_{\bullet}C. Further, we define the normalizations of these 𝖬𝖢,aC\operatorname{\sf MC}^{\ell,a}_{\bullet}C and 𝖬𝖢,a,bC\operatorname{\sf MC}^{\ell,a,b}_{\bullet}C described as

𝖬𝖢n,aC={(f1,,fn)𝖬𝖢nCsf1=a},\operatorname{\sf MC}^{\ell,a}_{n}C=\mathbb{Z}\{(f_{1},\dots,f_{n})\in\operatorname{\sf MC}^{\ell}_{n}C\mid sf_{1}=a\},

and

𝖬𝖢n,a,bC={(f1,,fn)𝖬𝖢nCsf1=a,tfn=b}.\operatorname{\sf MC}^{\ell,a,b}_{n}C=\mathbb{Z}\{(f_{1},\dots,f_{n})\in\operatorname{\sf MC}^{\ell}_{n}C\mid sf_{1}=a,tf_{n}=b\}.

The quotient 𝖬𝖢~C𝖬𝖢C\widetilde{\operatorname{\sf MC}}^{\ell}_{\bullet}C\longrightarrow\operatorname{\sf MC}^{\ell}_{\bullet}C induces quotients 𝖬𝖢~,aC𝖬𝖢,aC\widetilde{\operatorname{\sf MC}}^{\ell,a}_{\bullet}C\longrightarrow\operatorname{\sf MC}^{\ell,a}_{\bullet}C and 𝖬𝖢~,a,bC𝖬𝖢,a,bC\widetilde{\operatorname{\sf MC}}^{\ell,a,b}_{\bullet}C\longrightarrow\operatorname{\sf MC}^{\ell,a,b}_{\bullet}C, which are all homotopy equivalences by Lemma 4.4. We denote the homology of these equivalent chain complexes by 𝖬𝖧,aC\operatorname{\sf MH}^{\ell,a}_{\bullet}C and 𝖬𝖧,a,bC\operatorname{\sf MH}^{\ell,a,b}_{\bullet}C respectively. When CC is T1T_{1}, we have decompositions of chain complexes

𝖬𝖢CaObC𝖬𝖢,aCa,bObC𝖬𝖢,a,bC,\operatorname{\sf MC}^{\ell}_{\bullet}C\cong\bigoplus_{a\in\mathrm{Ob}C}\operatorname{\sf MC}^{\ell,a}_{\bullet}C\cong\bigoplus_{a,b\in\mathrm{Ob}C}\operatorname{\sf MC}^{\ell,a,b}_{\bullet}C,

where 𝖬𝖢n,a,bC={(f1,,fn)𝖬𝖢nCsf1=a,tfn=b}\operatorname{\sf MC}^{\ell,a,b}_{n}C=\mathbb{Z}\{(f_{1},\dots,f_{n})\in\operatorname{\sf MC}^{\ell}_{n}C\mid sf_{1}=a,tf_{n}=b\}. Hence they induce decompositions

𝖬𝖧CaObC𝖬𝖧,aCa,bObC𝖬𝖧,a,bC.\operatorname{\sf MH}^{\ell}_{\bullet}C\cong\bigoplus_{a\in\mathrm{Ob}C}\operatorname{\sf MH}^{\ell,a}_{\bullet}C\cong\bigoplus_{a,b\in\mathrm{Ob}C}\operatorname{\sf MH}^{\ell,a,b}_{\bullet}C.

We note that 𝖬𝖢C\operatorname{\sf MC}^{\ell}_{\bullet}C is spanned by non-degenerate paths on CC with length \ell. Hence we have the following.

Lemma 4.7.

Let CC be a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category of finite type.

  1. (1)

    If CC is tame, then the chain complex 𝖬𝖢C\operatorname{\sf MC}^{\ell}_{\bullet}C is bounded for any 0\ell\in\mathbb{R}_{\geq 0}.

  2. (2)

    If CC is quasi-tame and T1T_{1}, then the chain complex 𝖬𝖢,a,bC\operatorname{\sf MC}^{\ell,a,b}_{\bullet}C is bounded for any a,bObCa,b\in\mathrm{Ob}C and 0\ell\in\mathbb{R}_{\geq 0}.

The next propositions follow from Propositions 3.22, 3.23, and Lemma 4.7. They show that the magnitude homology categorifies the magnitude and the magnitude (co)weighting. It is a slight generalization of Corollary 7.15 in [18]

Proposition 4.8.

Let CC be a tame category. When CC is T1T_{1}, then it holds that

ζC1(a,b)=n,(1)n(rk𝖬𝖧n,a,bC)q\zeta_{C}^{-1}(a,b)=\sum_{n,\ell}(-1)^{n}(\operatorname{rk}\operatorname{\sf MH}^{\ell,a,b}_{n}C)q^{\ell}

and

ka=n,(1)n(rk𝖬𝖧n,aC)q.k_{a}=\sum_{n,\ell}(-1)^{n}(\operatorname{rk}\operatorname{\sf MH}^{\ell,a}_{n}C)q^{\ell}.
Proposition 4.9.

Let CC be a finite tame category. Then it holds that

𝖬𝖺𝗀C=n,(1)n(rk𝖬𝖧nC)q.{\sf Mag}C=\sum_{n,\ell}(-1)^{n}(\operatorname{rk}\operatorname{\sf MH}^{\ell}_{n}C)q^{\ell}.
Remark 4.10.

We have ζC1(a,b)=n,(1)n(rk𝖬𝖧n,a,bC)q\zeta_{C}^{-1}(a,b)=\sum_{n,\ell}(-1)^{n}(\operatorname{rk}\operatorname{\sf MH}^{\ell,a,b}_{n}C)q^{\ell} for a quasi-tame category CC by Remark 3.24 and Lemma 4.7.

Remark 4.11.

As shown in Examples 3.5.4 and 3.5.5, the magnitude homology gives categorifications of the growth series of finitely generated groups and Poincaré polynomials of ranked posets. We have no idea whether they can have torsions, and what torsions means if any. It is known that the magnitude homology can have torsions ([13], [20]).

4.2 Homotopy invariance of MH

In the following, we prove that 𝖬𝖧\operatorname{\sf MH} is invariant under category equivalence in the setting of filtered sets enrichment. As applications, we show that the magnitude homology of generalized metric spaces is invariant under the Kolmogorov quotient, and we also consider the Galois connection of posets. To that end, we first write down the definitions of some categorical notions in 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat}, which are straightforward generalization of those in 𝖢𝖺𝗍{\sf Cat}.

Definition 4.12.

Let C,DC,D be 𝖥𝗌𝖾𝗍{\sf Fset}-categories. We define a tensor product C×DC\times D by Ob(C×D)=ObC×ObD\mathrm{Ob}(C\times D)=\mathrm{Ob}C\times\mathrm{Ob}D and (C×D)((a,b),(a,b))=C(a,b)×D(a,b)(C\times D)((a,b),(a^{\prime},b^{\prime}))=C(a,b)\times D(a^{\prime},b^{\prime}).

Let I0I_{0} be the poset {0<1}\{0<1\} considered as a 𝖥𝗌𝖾𝗍{\sf Fset}-category via the inclusions 𝖯𝗈𝗌𝖾𝗍𝖢𝖺𝗍𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Poset}\longrightarrow{\sf Cat}\longrightarrow{\sf Fsetcat}. Namely, I0I_{0} has two objects and one non-trivial morphism 010\to 1 with degree 0.

Definition 4.13.

Let F,G:CD𝖥𝗌𝖾𝗍𝖼𝖺𝗍F,G:C\longrightarrow D\in{\sf Fsetcat}. A natural transformation τ:FG\tau:F\Rightarrow G is a functor τ:C×I0D\tau:C\times I_{0}\longrightarrow D such that τ(,0)=F()\tau(-,0)=F(-) and τ(,1)=G()\tau(-,1)=G(-).

Note that the underlying functor τ¯:C×I0¯D¯\underline{\tau}:\underline{C\times I_{0}}\longrightarrow\underline{D} is a natural transformation in 𝖢𝖺𝗍{\sf Cat}. For natural transformations τ\tau and τ\tau^{\prime}, we denote their “horizontal” and “vertical” compositions by ττ\tau\circ\tau^{\prime} and ττ\tau\bullet\tau^{\prime} respectively whenever they are defined. We also denote the identity natural transformation of a functor FF by idF{\rm id}_{F}.

Definition 4.14.

Let F:CD,G:DC𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:C\longrightarrow D,G:D\longrightarrow C\in{\sf Fsetcat}. We say that FF is left adjoint to GG, or equivalently GG is right adjoint to FF, denoted by FGF\dashv G, if there are natural transformations ε:idCGF\varepsilon:{\rm id}_{C}\Rightarrow GF and η:FGidD\eta:FG\Rightarrow{\rm id}_{D}, called the unit and the counit, satisfying the following commutative diagrams :

F\textstyle{F\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idFε\scriptstyle{{\rm id}_{F}\bullet\varepsilon}FGF\textstyle{FGF\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ηidF\scriptstyle{\eta\bullet{\rm id}_{F}}G\textstyle{G\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}εidG\scriptstyle{\varepsilon\bullet{\rm id}_{G}}GFG\textstyle{GFG\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idGη\scriptstyle{{\rm id}_{G}\bullet\eta}F,\textstyle{F,}G.\textstyle{G.}

We remark that FGF\vdash G if and only if there is a natural isomorphism C(,G())D(F(),)C(-,G(-))\cong D(F(-),-) in 𝖥𝗌𝖾𝗍{\sf Fset}.

Example 4.15.

Let F:XY,G:YX𝖦𝖬𝖾𝗍F:X\longrightarrow Y,G:Y\longrightarrow X\in{\sf GMet}. Then FGF\dashv G if and only if d(Fx,y)=d(x,Gy)d(Fx,y)=d(x,Gy) for any xXx\in X and yYy\in Y.

The following is essentially shown in Theorem 5.12 of [18], but we write down the proof again in our setting.

Proposition 4.16.

Let F,G:CD𝖥𝗌𝖾𝗍𝖼𝖺𝗍F,G:C\longrightarrow D\in{\sf Fsetcat}. If there is a natural transformation τ:FG\tau:F\Rightarrow G, then 𝖬𝖢F\operatorname{\sf MC}^{\ell}_{\bullet}F and 𝖬𝖢G\operatorname{\sf MC}^{\ell}_{\bullet}G are chain homotopic. Further, if Fa=Ga(=:b)Fa=Ga(=:b) and a,ba,b are T1T_{1}, then 𝖬𝖢,aF\operatorname{\sf MC}^{\ell,a}_{\bullet}F and 𝖬𝖢,bG\operatorname{\sf MC}^{\ell,b}_{\bullet}G are chain homotopic.

Proof.

We construct a filtration preserving chain homotopy H:𝖥𝖢C𝖥𝖢+1DH_{\bullet}:{\sf FC}_{\bullet}C\longrightarrow{\sf FC}_{\bullet+1}D. Note that this construction is well-known for small categories, namely when we forget the filtration. Hence it suffices to check the chain homotopy preserves the filtration. For each k0k\in\mathbb{Z}_{\geq 0}, we define HkH_{k} by

Hk(f1,,fk)=i=1k(1)i(Ff1,,Ffi,τ(tfi),Gfi,,Gfk),H_{k}(f_{1},\dots,f_{k})=\sum_{i=1}^{k}(-1)^{i}(Ff_{1},\dots,Ff_{i},\tau(tf_{i}),Gf_{i},\dots,Gf_{k}),

where τ(tfi)\tau(tf_{i}) is the arrow τ((tfi,0),(tfi,1)):F(tfi)G(tfi)\tau((tf_{i},0),(tf_{i},1)):F(tf_{i})\to G(tf_{i}). Then we can check that k+1Hk+Hk1k=𝖥𝖢kF𝖥𝖢kG\partial_{k+1}H_{k}+H_{k-1}\partial_{k}={\sf FC}_{k}F-{\sf FC}_{k}G. Since degτ(tfi)=0\deg\tau(tf_{i})=0, it does not decrease the degree. Hence the above homotopy induces a homotopy between 𝖬𝖢~kF\widetilde{\operatorname{\sf MC}}^{\ell}_{k}F and 𝖬𝖢~kG\widetilde{\operatorname{\sf MC}}^{\ell}_{k}G, which are chain maps from (𝖥𝖢C)/(𝖥𝖢C)<({\sf FC}_{\bullet}C)_{\ell}/({\sf FC}_{\bullet}C)_{<\ell} to (𝖥𝖢D)/(𝖥𝖢D)<({\sf FC}_{\bullet}D)_{\ell}/({\sf FC}_{\bullet}D)_{<\ell}. Further, if Fa=Ga(=:b)Fa=Ga(=:b) and a,ba,b are T1T_{1}, then it is easily checked that this homotopy is restricted to 𝖬𝖢~,aC𝖬𝖢~+1,bD\widetilde{\operatorname{\sf MC}}^{\ell,a}_{\bullet}C\longrightarrow\widetilde{\operatorname{\sf MC}}^{\ell,b}_{\bullet+1}D. Hence the latter assertion follows. This completes the proof. ∎

The following is immediate from Proposition 4.16.

Corollary 4.17.

Let F:CD,G:DC𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:C\longrightarrow D,G:D\longrightarrow C\in{\sf Fsetcat}. If FGF\dashv G, then they induce isomorphisms between 𝖬𝖧C\operatorname{\sf MH}^{\ell}_{\bullet}C and 𝖬𝖧D\operatorname{\sf MH}^{\ell}_{\bullet}D. Further, if aObCa\in\mathrm{Ob}C and bObDb\in\mathrm{Ob}D are T1T_{1}, and Fa=b,Gb=aFa=b,Gb=a, then the above isomorphisms are restricted to isomorphisms between 𝖬𝖧,aC\operatorname{\sf MH}^{\ell,a}_{\bullet}C and 𝖬𝖧,bD\operatorname{\sf MH}^{\ell,b}_{\bullet}D.

As an application of Corollary 4.17, we consider the Kolmogorov quotient of generalized metric spaces, and the Galois connection of posets.

4.2.1 Kolmogorov quotient of generalized metric spaces

Definition 4.18.

For a generalized metric space XX, we define an equivalence relation \sim on XX by xyx\sim y if and only if d(x,y)=d(y,x)=0d(x,y)=d(y,x)=0. Then the quotient set X/X/\sim is again a generalized metric space by defining d([x],[y])=d(x,y)d([x],[y])=d(x,y), which is obviously non-degenerate. We call this generalized metric space the Kolmogorov quotient of XX, and denote it by 𝖪𝖰X{\sf KQ}X.

Note that X=𝖪𝖰XX={\sf KQ}X if XX is non-degenerate.

Lemma 4.19.

For any generalized metric space XX, the quotient map p:X𝖪𝖰Xp:X\longrightarrow{\sf KQ}X is a left adjoint functor in 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat}.

Proof.

It is straightforward to check that the quotient map pp is a 11-Lipschitz map, hence a functor in 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat}. Let q:𝖪𝖰XXq:{\sf KQ}X\longrightarrow X be a section of the map pp. Then qq is obviously a 11-Lipschitz map, and we have

dX(x,q[y])=dX(x,y)=d𝖪𝖰X(px,[y])d_{X}(x,q[y])=d_{X}(x,y)=d_{{\sf KQ}X}(px,[y])

for any xXx\in X and [y]𝖪𝖰X[y]\in{\sf KQ}X. Hence we have pqp\dashv q by Example 4.15. This completes the proof. ∎

Now we obtain the following by Corollary 4.17 and Lemma 4.19.

Proposition 4.20.

For any generalized metric space XX, we have 𝖬𝖧X𝖬𝖧𝖪𝖰X\operatorname{\sf MH}^{\ell}_{\bullet}X\cong\operatorname{\sf MH}^{\ell}_{\bullet}{\sf KQ}X. Namely, the magnitude homology is invariant under Kolmogorov quotient.

Note that Proposition 4.20 together with Proposition 4.9 implies that any finite generalized metric space have the magnitude by factoring its Kolmogorov quotient, even if it is not tame, namely not non-degenerate.

4.2.2 Galois connection of posets

Definition 4.21.

Let PP and QQ be posets, and let F:PQF:P\longrightarrow Q and G:QPG:Q\longrightarrow P be order preserving maps. The pair (F,G)(F,G) is a Galois connection if FGF\dashv G in 𝖢𝖺𝗍{\sf Cat} and thus in 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat}, namely it satisfies that xGFxx\leq GFx and FGyyFGy\leq y for all xPx\in P and yQy\in Q.

Proposition 4.22.

Let PP and QQ be finite posets with minimum elements 0P0_{P} and 0Q0_{Q}. Let φ:P0\varphi:P\longrightarrow\mathbb{Z}_{\geq 0} and ψ:Q0\psi:Q\longrightarrow\mathbb{Z}_{\geq 0} be order preserving maps satisfying that φ1{0}={0P}\varphi^{-1}\{0\}=\{0_{P}\} and ψ1{0}={0Q}\psi^{-1}\{0\}=\{0_{Q}\}. If a Galois connection (F,G)(F,G), where F:PQF:P\longrightarrow Q and G:QPG:Q\longrightarrow P, satisfies that ψF=φ\psi F=\varphi and φG=ψ\varphi G=\psi, then (F,G)(F,G) is an adjoint pair between generalized metric spaces (P,dφ)(P,d_{\varphi}) and (Q,dψ)(Q,d_{\psi}) defined in Remark 3.37.

Proof.

Let xPx\in P and yQy\in Q. Then we have

dψ(Fx,y)={ψ(y)ψ(Fx)Fxy,Fxy,={ψ(y)φ(x)Fxy,Fxy,d_{\psi}(Fx,y)=\begin{cases}\psi(y)-\psi(Fx)&Fx\leq y,\\ \infty&Fx\not\leq y,\end{cases}=\begin{cases}\psi(y)-\varphi(x)&Fx\leq y,\\ \infty&Fx\not\leq y,\end{cases}

and

dφ(x,Gy)={φ(Gy)φ(x)xGy,xGy,={ψ(y)φ(x)xGy,xGy.d_{\varphi}(x,Gy)=\begin{cases}\varphi(Gy)-\varphi(x)&x\leq Gy,\\ \infty&x\not\leq Gy,\end{cases}=\begin{cases}\psi(y)-\varphi(x)&x\leq Gy,\\ \infty&x\not\leq Gy.\end{cases}

Since (F,G)(F,G) is a Galois connection, we obtain that dψ(Fx,y)=dφ(x,Gy)d_{\psi}(Fx,y)=d_{\varphi}(x,Gy), hence (F,G)(F,G) is an adjoint pair by Example 4.15. This completes the proof. ∎

The following is immediate from Corollary 3.23, Proposition 3.36 and Corollary 4.17.

Corollary 4.23.

Under the same assumption as in Proposition 4.22, we have π(P,φ)(q)=π(Q,ψ)(q)\pi_{(P,\varphi)}(q)=\pi_{(Q,\psi)}(q).

Example 4.24.

Let SS be a subspace arrangement. Let I(S),P(S)I(S),P(S) be posets with order preserving maps ψ:I(S)0\psi:I(S)\longrightarrow\mathbb{Z}_{\geq 0} and φ:P(S)0\varphi:P(S)\longrightarrow\mathbb{Z}_{\geq 0} as in Examples 3.30 and 3.38. We define poset maps F:P(S)I(S)F:P(S)\longrightarrow I(S) and G:I(S)P(S)G:I(S)\longrightarrow P(S) by Fx=txtFx=\cap_{t\in x}t and Gy={sSys}Gy=\{s\in S\mid y\subset s\}. Then we have xGFxx\leq GFx and FGy=yFGy=y for any xP(S)x\in P(S) and yI(S)y\in I(S). Hence the pair (F,G)(F,G) is a Galois connection. Further, by the definitions of ψ\psi and φ\varphi, it is straightforward to check the assumptions in Proposition 4.22. Thus we obtain that π(I(S),ψ)(q)=π(P(S),φ)\pi_{(I(S),\psi)}(q)=\pi_{(P(S),\varphi)} by Corollary 4.23.

4.3 MH as a Hochschild homology

In the following, we describe the magnitude homology as the Hochschild homology. A relevant claim appears in Remark 5.11 of [18], however, the description we give here is more ring theoretic. That is, we express the magnitude homology as the Hochschild homology of an algebra associated with the 𝖥𝗌𝖾𝗍{\sf Fset}-category, which is a generalization of the category algebra. We start from basic conventions.

Definition 4.25.

A filtered ring is a filtered abelian group RR with a unital associative ring structure such that degabdega+degb\deg ab\leq\deg a+\deg b.

Definition 4.26.

Let CC be a 𝖥𝗌𝖾𝗍{\sf Fset}-category. We define a filtered ring PC()P_{C}(\mathbb{Z}) by

  • PC()=MorCP_{C}(\mathbb{Z})=\mathbb{Z}\mathrm{Mor}C with (PC())={fMorCdegf}(P_{C}(\mathbb{Z}))_{\ell}=\mathbb{Z}\{f\in\mathrm{Mor}C\mid\deg f\leq\ell\} as a filtered abelian group.

  • For any f,gMorCf,g\in\mathrm{Mor}C, we define an associative product \cdot by fg={gf if tf=sg,0 otherwise.f\cdot g=\begin{cases}g\circ f&\text{ if }tf=sg,\\ 0&\text{ otherwise}.\end{cases}

We call PC()P_{C}(\mathbb{Z}) the category algebra of CC. We also define an action of PC()P_{C}(\mathbb{Z}) on the abelian group MC():=(ObX×ObX)M_{C}(\mathbb{Z}):=\mathbb{Z}(\mathrm{Ob}X\times\mathrm{Ob}X) from the right and the left by f(a,b)={(sf,b) if tf=a,0 otherwise,f\cdot(a,b)=\begin{cases}(sf,b)&\text{ if }tf=a,\\ 0&\text{ otherwise},\end{cases} and (a,b)f={(a,tf) if sf=b,0 otherwise.(a,b)\cdot f=\begin{cases}(a,tf)&\text{ if }sf=b,\\ 0&\text{ otherwise}.\end{cases}

Now we recall the Hochschild homology. Let RR be a unital associative ring, and MM be an RR-bimodule, which are not filtered. The Hochschild chain complex 𝖧𝖢(R,M)\operatorname{\sf HC}_{\bullet}(R,M) of RR and MM is defined by 𝖧𝖢n(R,M):=RnM\operatorname{\sf HC}_{n}(R,M):=R^{\otimes n}\otimes M and

n(r1rnm)\displaystyle\partial_{n}(r_{1}\otimes\dots\otimes r_{n}\otimes m) =r2rnmr1\displaystyle=r_{2}\otimes\dots\otimes r_{n}\otimes mr_{1}
+i=1n1(1)ir1riri+1rnm\displaystyle+\sum_{i=1}^{n-1}(-1)^{i}r_{1}\otimes\dots\otimes r_{i}r_{i+1}\otimes\dots\otimes r_{n}\otimes m
+(1)nr1rnm,\displaystyle+(-1)^{n}r_{1}\otimes\dots\otimes r_{n}m,

where \otimes is taken over \mathbb{Z}. Its homology 𝖧𝖧(R,M)\operatorname{\sf HH}_{\bullet}(R,M) is called Hochschild homology of (R,M)(R,M). If RR is filtered, then the filtration induces again a filtration FF_{\bullet} on 𝖧𝖢(R,M)\operatorname{\sf HC}_{\bullet}(R,M), hence also on 𝖧𝖧(R,M)\operatorname{\sf HH}_{\bullet}(R,M), by

F𝖧𝖢n(R,M):={r1rnmi=1ndegri},F_{\ell}\operatorname{\sf HC}_{n}(R,M):=\mathbb{Z}\{r_{1}\otimes\dots\otimes r_{n}\otimes m\mid\sum_{i=1}^{n}{\rm deg\ }r_{i}\leq\ell\},

and

F𝖧𝖧n(R,M):=Im(𝖧nF𝖧𝖢(R,M)𝖧𝖧n(R,M)).F_{\ell}\operatorname{\sf HH}_{n}(R,M):={\rm Im}({\sf H}_{n}F_{\ell}\operatorname{\sf HC}_{\bullet}(R,M)\longrightarrow\operatorname{\sf HH}_{n}(R,M)).
Remark 4.27.

Such a filtration on the Hochschild chain complex is also considered by Brylinski ([3]) in the study of Poisson manifolds.

Now we compare the chain complexes 𝖥𝖢C{\sf FC}_{\bullet}C and 𝖧𝖢(PC(),MC())\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z})) with respect to the filtrations in the following.

Lemma 4.28.

Let CC be a 𝖥𝗌𝖾𝗍{\sf Fset}-category. The homomorphisms

𝖥𝖢nC𝖧𝖢n(PC(),MC());(f1,,fn)f1fn(tfn,sf1){\sf FC}_{n}C\longrightarrow\operatorname{\sf HC}_{n}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}));(f_{1},\dots,f_{n})\mapsto f_{1}\otimes\dots\otimes f_{n}\otimes(tf_{n},sf_{1})\,

for n1n\geq 1, and 𝖥𝖢0C𝖧𝖢0(PC(),MC());a(a,a){\sf FC}_{0}C\longrightarrow\operatorname{\sf HC}_{0}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}));a\mapsto(a,a) define an injective filtered chain map 𝖥𝖢C𝖧𝖢(PC(),MC()){\sf FC}_{\bullet}C\longrightarrow\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z})).

Proof.

It is straightforward. ∎

We consider the chain complex 𝖥𝖢C{\sf FC}_{\bullet}C as a subcomplex of 𝖧𝖢(PC(),MC())\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z})) in the following.

Lemma 4.29.

Let CC be a 𝖥𝗌𝖾𝗍{\sf Fset}-category. The chain complexes

𝖧𝖢(PC(),MC())/𝖥𝖢C,\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/{\sf FC}_{\bullet}C,

and

F𝖧𝖢(PC(),MC())/(𝖥𝖢C)F_{\ell}\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/({\sf FC}_{\bullet}C)_{\ell}

are contractible for any 0\ell\in\mathbb{R}_{\geq 0}.

Proof.

We first prove the latter contractibility. Let

α=λΛcλf1λfnλ(aλ,bλ)F𝖧𝖢n(PC(),MC())/(𝖥𝖢nC).\alpha=\sum_{\lambda\in\Lambda}c_{\lambda}f^{\lambda}_{1}\otimes\dots\otimes f^{\lambda}_{n}\otimes(a^{\lambda},b^{\lambda})\in F_{\ell}\operatorname{\sf HC}_{n}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/({\sf FC}_{n}C)_{\ell}.

For each λΛ\lambda\in\Lambda, let k(λ)0k(\lambda)\geq 0 be the minimum number such that tfk(λ)λsfk(λ)+1λtf^{\lambda}_{k(\lambda)}\neq sf^{\lambda}_{k(\lambda)+1}, where we put tf0λ=bλtf^{\lambda}_{0}=b^{\lambda}. We define a homomorphism

Bn:F𝖧𝖢n(PC(),MC())/(𝖥𝖢nC)F𝖧𝖢n+1(PC(),MC())/(𝖥𝖢n+1C)B^{\ell}_{n}:F_{\ell}\operatorname{\sf HC}_{n}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/({\sf FC}_{n}C)_{\ell}\longrightarrow F_{\ell}\operatorname{\sf HC}_{n+1}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/({\sf FC}_{n+1}C)_{\ell}

by

Bnα:=λΛ(1)k(λ)+1cλf1λfk(λ)λidsfk(λ)+1λfk(λ)+1λfnλ(aλ,bλ)B^{\ell}_{n}\alpha:=\sum_{\lambda\in\Lambda}(-1)^{k(\lambda)+1}c_{\lambda}f^{\lambda}_{1}\otimes\dots\otimes f^{\lambda}_{k(\lambda)}\otimes\mathrm{id}_{sf^{\lambda}_{k(\lambda)+1}}\otimes f^{\lambda}_{k(\lambda)+1}\otimes\dots\otimes f^{\lambda}_{n}\otimes(a^{\lambda},b^{\lambda})

for n1n\geq 1 and

B0(a,b):=ida(a,b).B^{\ell}_{0}(a,b):=-\mathrm{id}_{a}\otimes(a,b).

Then it is straightforward to check that n+1Bnα+Bn1nα=α\partial_{n+1}B^{\ell}_{n}\alpha+B^{\ell}_{n-1}\partial_{n}\alpha=\alpha. Hence this defines a chain homotopy between the identity and the zero homomorphism on

F𝖧𝖢(PC(),MC())/(𝖥𝖢C).F_{\ell}\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/({\sf FC}_{\bullet}C)_{\ell}.

It is easily seen that the homotopy BB^{\ell}_{\bullet} extends to a homotopy

B:𝖧𝖢(PC(),MC())/𝖥𝖢C𝖧𝖢+1(PC(),MC())/𝖥𝖢+1C,B_{\bullet}:\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/{\sf FC}_{\bullet}C\longrightarrow\operatorname{\sf HC}_{\bullet+1}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/{\sf FC}_{\bullet+1}C,

which shows the contractibility of 𝖧𝖢(PC(),MC())/𝖥𝖢C\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/{\sf FC}_{\bullet}C. This completes the proof. ∎

Corollary 4.30.

Let CC be a 𝖥𝗌𝖾𝗍{\sf Fset}-category. We have homotopy equivalences

𝖥𝖢C𝖧𝖢(PC(),MC()){\sf FC}_{\bullet}C\simeq\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))

and

(𝖥𝖢C)F𝖧𝖢(PC(),MC())({\sf FC}_{\bullet}C)_{\ell}\simeq F_{\ell}\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))

for any 0\ell\in\mathbb{R}_{\geq 0}.

Proof.

By Lemma 4.29, the inclusions

𝖥𝖢C𝖧𝖢(PC(),MC()){\sf FC}_{\bullet}C\longrightarrow\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))

and

(𝖥𝖢C)F𝖧𝖢(PC(),MC())({\sf FC}_{\bullet}C)_{\ell}\longrightarrow F_{\ell}\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))

defined in Lemma 4.28 are quasi isomorphisms. Since these chain complexes are level-wise free, which implies that they are fibrant-cofibrant objects in the projective model structure of 𝖢𝗁0{\sf Ch}_{\geq 0}, it is a homotopy equivalence by the Whitehead theorem (we refer to section 1.4 and 1.5 of [21] for more concrete discussion). This completes the proof. ∎

Remark 4.31.

When CC is a face poset of a finite simplicial complex SS as in Example 3.5.3, the algebra PC()P_{C}(\mathbb{Z}) is known as the incidence algebra of SS, and Corollary 4.30 can be considered as a homological version of well-known Gerstenhaber-Schack’s theorem ([6]) that asserts H(S)HH(PC())H^{\ast}(S)\cong HH^{\ast}(P_{C}(\mathbb{Z})). We also remark that Grigor’yan–Muranov–S.-T. Yau ([9]) give a proof of Gerstenhaber-Schack’s theorem via path cohomology. Our proofs of Lemma 4.29 and Corollary 4.30 are inspired by Lemma 5.1 of [9].

Now we define a filtered ring GrPC(){\rm Gr}P_{C}(\mathbb{Z}) by

  • PC()=MorCP_{C}(\mathbb{Z})=\mathbb{Z}\mathrm{Mor}C with (PC())={fMorCdegf}(P_{C}(\mathbb{Z}))_{\ell}=\mathbb{Z}\{f\in\mathrm{Mor}C\mid\deg f\leq\ell\} as a filtered abelian group.

  • For any f,gMorCf,g\in\mathrm{Mor}C we define an associative product \cdot by

    fg={gf if tf=sg and degf+degg=deggf,0 otherwise.f\cdot g=\begin{cases}g\circ f&\text{ if }tf=sg\text{ and }\deg f+\deg g=\deg g\circ f,\\ 0&\text{ otherwise}.\end{cases}

We also define an action of GrPC(){\rm Gr}P_{C}(\mathbb{Z}) on MC()M_{C}(\mathbb{Z}) similarly to that of PC()P_{C}(\mathbb{Z}). Note that

<F𝖧𝖢(GrPC(),MC())F𝖧𝖢(GrPC(),MC())\bigcup_{\ell^{\prime}<\ell}F_{\ell^{\prime}}\operatorname{\sf HC}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))\subset F_{\ell}\operatorname{\sf HC}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))

and

F𝖧𝖢(GrPC(),MC())𝖧𝖢(GrPC(),MC())F_{\ell}\operatorname{\sf HC}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))\subset\operatorname{\sf HC}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))

are direct summands. Hence we have

𝖧(F𝖧𝖢(GrPC(),MC())/<F𝖧𝖢(GrPC(),MC()))\displaystyle{\sf H}_{\bullet}\left(F_{\ell}\operatorname{\sf HC}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/\bigcup_{\ell^{\prime}<\ell}F_{\ell^{\prime}}\operatorname{\sf HC}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))\right)
F𝖧𝖧(GrPC(),MC())/<F𝖧𝖧(GrPC(),MC())\displaystyle\cong F_{\ell}\operatorname{\sf HH}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/\bigcup_{\ell^{\prime}<\ell}F_{\ell^{\prime}}\operatorname{\sf HH}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))
=:Gr𝖧𝖧(GrPC(),MC()),\displaystyle=:{\rm Gr}_{\ell}\operatorname{\sf HH}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z})),

and

𝖧𝖧(GrPC(),MC())0Gr𝖧𝖧(GrPC(),MC()).\operatorname{\sf HH}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))\cong\bigoplus_{\ell\geq 0}{\rm Gr}_{\ell}\operatorname{\sf HH}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z})).

Finally we obtain the following description of the magnitude homology as the Hochschild homology.

Corollary 4.32.

Let CC be a 𝖥𝗌𝖾𝗍{\sf Fset}-category. The chain complexes 𝖬𝖢C\operatorname{\sf MC}^{\ell}_{\bullet}C and

F𝖧𝖢(PC(),MC())/<F𝖧𝖢(PC(),MC())F_{\ell}\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/\bigcup_{\ell^{\prime}<\ell}F_{\ell^{\prime}}\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))

are homotopy equivalent for any 0\ell\in\mathbb{R}_{\geq 0}. In particular, we have an isomorphism

𝖬𝖧CGr𝖧𝖧(GrPC(),MC())\displaystyle\operatorname{\sf MH}^{\ell}_{\bullet}C\cong Gr_{\ell}\operatorname{\sf HH}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))

for any 0\ell\in\mathbb{R}_{\geq 0}. Hence we have 0𝖬𝖧C𝖧𝖧(GrPC(),MC())\bigoplus_{\ell\geq 0}\operatorname{\sf MH}^{\ell}_{\bullet}C\cong\operatorname{\sf HH}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z})).

Proof.

By Lemma 4.29, the inclusions

(𝖥𝖢C)F𝖧𝖢(PC(),MC())({\sf FC}_{\bullet}C)_{\ell}\longrightarrow F_{\ell}\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))

and

<(𝖥𝖢C)<F𝖧𝖢(PC(),MC())\bigcup_{\ell^{\prime}<\ell}({\sf FC}_{\bullet}C)_{\ell^{\prime}}\longrightarrow\bigcup_{\ell^{\prime}<\ell}F_{\ell^{\prime}}\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))

are quasi isomorphisms. Hence the five lemma shows that we have a quasi isomorphism

𝖬𝖢~CF𝖧𝖢(PC(),MC())/<F𝖧𝖢(PC(),MC()).\widetilde{\operatorname{\sf MC}}^{\ell}_{\bullet}C\longrightarrow F_{\ell}\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/\bigcup_{\ell^{\prime}<\ell}F_{\ell^{\prime}}\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z})).

Since they are level-wise free, this is a homotopy equivalence. The latter follows from the identification

F𝖧𝖢(PC(),MC())/<F𝖧𝖢(PC(),MC())\displaystyle F_{\ell}\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/\bigcup_{\ell^{\prime}<\ell}F_{\ell^{\prime}}\operatorname{\sf HC}_{\bullet}(P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))
=F𝖧𝖢(GrPC(),MC())/<F𝖧𝖢(GrPC(),MC()),\displaystyle=F_{\ell}\operatorname{\sf HC}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z}))/\bigcup_{\ell^{\prime}<\ell}F_{\ell^{\prime}}\operatorname{\sf HC}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z})),

and the direct sum decomposition of 𝖧𝖢(GrPC(),MC())\operatorname{\sf HC}_{\bullet}({\rm Gr}P_{C}(\mathbb{Z}),M_{C}(\mathbb{Z})). This completes the proof. ∎

4.4 MH, a spectral sequence and homotopy relations

In the following, we construct a spectral sequence for a 𝖥𝗌𝖾𝗍{\sf Fset}-category CC whose morphisms have integral degrees. It converges to the homology of underlying small category C¯\underline{C}, and its first page is isomorphic to the magnitude homology. Further, a part of the second page is a well-known invariant path homology of digraphs. We also discuss the homotopy invariance of each page.

Definition 4.33.

A filtered set XX is a 0\mathbb{Z}_{\geq 0}-filtered set if degx0\deg x\in\mathbb{Z}_{\geq 0} for every xXx\in X. We denote the full subcategory of 𝖥𝗌𝖾𝗍{\sf Fset} that consists of 0\mathbb{Z}_{\geq 0}-filtered sets by 𝖥𝗌𝖾𝗍(){\sf Fset}(\mathbb{Z}).

Note that the filtered chain complex 𝖥𝖢C{\sf FC}_{\bullet}C is 0\mathbb{Z}_{\geq 0}-filtered for a 𝖥𝗌𝖾𝗍(){\sf Fset}(\mathbb{Z})-category CC. We denote the spectral sequence associated with this filtered chain complex by EE^{\bullet}_{\bullet\bullet}. The following proposition, which is proved by the author in [1], shows a remarkable connection between the magnitude homology and the homotopy theory of digraphs.

Proposition 4.34.

For a 𝖥𝗌𝖾𝗍(){\sf Fset}(\mathbb{Z})-category CC, we have the following.

  1. (1)

    Ep,q0=𝖬𝖢~p+qpC,Ep,q1=𝖬𝖧p+qpCE^{0}_{p,q}=\widetilde{\operatorname{\sf MC}}^{p}_{p+q}C,E^{1}_{p,q}=\operatorname{\sf MH}^{p}_{p+q}C.

  2. (2)

    If CC is a digraph, we have Ep,02=H~pCE^{2}_{p,0}=\widetilde{H}_{p}C, where H~\widetilde{H}_{\bullet} denotes the reduced path homology introduced by Grigor’yan–Muranov–Lin–S. -T. Yau et al. [8].

  3. (3)

    It converges to the homology of the underlying small category C¯\underline{C} if it converges, in particular when max{degffMorC}\max\{\deg f\mid f\in\mathrm{Mor}C\} is finite.

Proof.

(1), (3) is straightforward. (2) is shown in Theorem 1.2 of [1]. ∎

For r0r\in\mathbb{R}_{\geq 0}, let IrI_{r} be a 𝖥𝗌𝖾𝗍{\sf Fset}-category with two objects 0,10,1 and one non-trivial morphism 010\to 1 with degree rr. The following is a generalization of the definition of natural transformation in 𝖢𝖺𝗍{\sf Cat}, which induces a series of homotopy relations in 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat} as it will be stated in Theorem 4.37.

Definition 4.35.

Let F,G:CD𝖥𝗌𝖾𝗍𝖼𝖺𝗍F,G:C\longrightarrow D\in{\sf Fsetcat}. An rr-natural transformation τ:FG\tau:F\Rightarrow G is a functor τ:C×IrD\tau:C\times I_{r}\longrightarrow D such that τ(,0)=F()\tau(-,0)=F(-) and τ(,1)=G()\tau(-,1)=G(-).

Example 4.36.
  1. (1)

    A 0-natural transformation is the one defined in Definition 4.13.

  2. (2)

    A 11-natural transformation restricted to the category DGrph is exactly the 11-step homotopy of digraph homomorphisms defined in [8]. It is shown in the same paper that the path homology is invariant under 11-step homotopies.

  3. (3)

    Let F,G:CDF,G:C\longrightarrow D be 1-Lipschitz maps with C,DC,D being metric spaces. Then the existence of an rr-natural transformation τ:FG\tau:F\Rightarrow G is equivalent to the condition that d(Fa,Ga)rd(Fa,Ga)\leq r for any aCa\in C, which is called rr-close in the metric geometry.

  4. (4)

    An rr-natural transformation induces an rr^{\prime}-natural transformation via the natural morphism IrIrI_{r^{\prime}}\longrightarrow I_{r} if rrr^{\prime}\geq r.

The following is a generalization of Proposition 5.7 in [1] and Theorem 3.3 in [8]. It also contains Theorem 4.16.

Theorem 4.37.

The (r+1)(r+1)-page Er+1E^{r+1}_{\bullet\bullet} is invariant under rr-natural transformations. That is, F,G:CDF,G:C\longrightarrow D induces identical homomorphisms Er+1CEr+1DE^{r+1}_{\bullet\bullet}C\longrightarrow E^{r+1}_{\bullet\bullet}D if there exists an rr-natural transformation τ:FG\tau:F\Rightarrow G.

Proof.

It is an immediate consequence of Proposition 3.9 in [5]. ∎

Remark 4.38.

In [5], Cirici et al. show that there is a cofibrantly generated model structure on the category of 0\mathbb{Z}_{\geq 0}-filtered chain complexes, where weak equivalences are exactly the morphisms inducing quasi isomorphisms on ErE^{r}. Theorem 4.37 suggests that there should be some homotopy theory on 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat} that coincides with the above Ciricis’ structure when we apply a functor 𝖥𝖢:𝖥𝗌𝖾𝗍𝖼𝖺𝗍𝖥𝖢𝗁0{\sf FC}_{\bullet}:{\sf Fsetcat}\longrightarrow{\sf FCh}_{\geq 0}. The work [4] by Carranza et al. seems to support this hypothesis. In that paper, they constructed a cofibration category structure on the category of digraphs, where weak equivalences are exactly the morpshisms inducing isomorphisms on the path homology, namely a part of E2E^{2}.

5 Fibrations in 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat}

In this section, we study “fibrations” in 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat}, whose restriction to 𝖬𝖾𝗍{\sf Met} is originally introduced as metric fibrations by Leinster in [15]. It contains the Grothendieck (op)fibration in 𝖢𝖺𝗍{\sf Cat}. A remarkable property of this notion is that the magnitude of a fibration split as the product of those of the “fiber” and the “base”. It is well-known that there is a one to one correspondence between Grothendieck (op)fibrations and the (op)lax functors (2-category equivalence in precise). We generalize it to 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat}. Further, we restrict our notion of fibrations to 𝖬𝖾𝗍{\sf Met}, and give some examples. Due to the above one to one correspondence, we can find new examples of metric fibrations that are not considered in [15]. We remark that the Euler characteristic of categorical fibrations is also considered in [17], and we also generalize it.

5.1 pre-opfibrations

We first recall “fibrations in 𝖢𝖺𝗍{\sf Cat}” in the following. While there are variants of definitions of “fibrations” in 𝖢𝖺𝗍{\sf Cat}, we adopt pre-opfibration here. We remark that the other notions of fibrations can be considered similarly in 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat}, and they all coincide when we restrict them to 𝖬𝖾𝗍{\sf Met}.

Definition 5.1.

Let XX be a small category. The following data F:X𝖢𝖺𝗍F:X\longrightarrow{\sf Cat} is called a normal oplax functor.

  1. (1)

    FxFx is a small category for any xObXx\in\mathrm{Ob}X.

  2. (2)

    Ff:FxFxFf:Fx\to Fx^{\prime} is a functor for any f:xxMorXf:x\to x^{\prime}\in\mathrm{Mor}X.

  3. (3)

    For any xObXx\in\mathrm{Ob}X, Fidx=idFxF{\rm id}_{x}={\rm id}_{Fx}.

  4. (4)

    For any f:xx,g:xx′′MorXf:x\to x^{\prime},g:x^{\prime}\to x^{\prime\prime}\in\mathrm{Mor}X, there is a natural transformation τf,g:F(gf)FgFf\tau_{f,g}:F(g\circ f)\Rightarrow Fg\circ Ff satisfying the following. For any f:xx,g:xx′′,h:x′′x′′′MorXf:x\to x^{\prime},g:x^{\prime}\to x^{\prime\prime},h:x^{\prime\prime}\to x^{\prime\prime\prime}\in\mathrm{Mor}X, we have τf,idx=idFf,τidx,f=idFf\tau_{f,{\rm id}_{x^{\prime}}}={\rm id}_{Ff},\tau_{{\rm id}_{x},f}={\rm id}_{Ff}, and the following commutative diagram :

    F(h(gf))\textstyle{F(h\circ(g\circ f))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τgf,h\scriptstyle{\tau_{g\circ f,h}}FhF(gf)\textstyle{Fh\circ F(g\circ f)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idFhτf,g\scriptstyle{{\rm id}_{Fh}\bullet\tau_{f,g}}Fh(FgFf)\textstyle{Fh\circ(Fg\circ Ff)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F((hg)f)\textstyle{F((h\circ g)\circ f)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τf,hg\scriptstyle{\tau_{f,h\circ g}}F(hg)Ff\textstyle{F(h\circ g)\circ Ff\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τg,hidFf\scriptstyle{\tau_{g,h}\bullet{\rm id}_{Ff}}(FhFg)Ff.\textstyle{(Fh\circ Fg)\circ Ff.}
Definition 5.2.

Let F:X𝖢𝖺𝗍F:X\longrightarrow{\sf Cat} be a normal oplax functor. We construct a small category E(F)E(F) as follows.

  1. (1)

    ObE(F)={(x,a)aObFx,xObX}\mathrm{Ob}E(F)=\{(x,a)\ \mid a\in\mathrm{Ob}Fx,x\in\mathrm{Ob}X\},

  2. (2)

    E(F)((x,a),(x,b))={(f,ξ)f:xxMorX,ξ:FfabMorFx}E(F)((x,a),(x^{\prime},b))=\{(f,\xi)\mid f:x\to x^{\prime}\in\mathrm{Mor}X,\xi:Ffa\to b\in\mathrm{Mor}Fx^{\prime}\},

  3. (3)

    For (f,ξ):(x,a)(x,b)(f,\xi):(x,a)\to(x^{\prime},b) and (g,θ):(x,b)(x′′,c)(g,\theta):(x^{\prime},b)\to(x^{\prime\prime},c), (g,θ)(f,ξ)=(gf,θFg(ξ)τf,ga)(g,\theta)\circ(f,\xi)=(g\circ f,\theta\circ Fg(\xi)\circ\tau_{f,g}a), where τf,ga:F(gf)aFgFfa\tau_{f,g}a:F(g\circ f)a\to Fg\circ Ffa is a component of the natural transformation τf,g\tau_{f,g}.

Note that we have a natural projection functor πF:E(F)X\pi_{F}:E(F)\longrightarrow X defined by πF(x,a)=x\pi_{F}(x,a)=x and πF(f,ξ)=f\pi_{F}(f,\xi)=f. In the following, we use the convention “φ:uuU\varphi:u\longrightarrow u^{\prime}\in U” for arbitrary category UU if u,uObUu,u^{\prime}\in\mathrm{Ob}U and φMorU\varphi\in\mathrm{Mor}U.

Definition 5.3.

Let π:XY𝖢𝖺𝗍\pi:X\longrightarrow Y\in{\sf Cat} and f:xxXf:x\to x^{\prime}\in X. We say that ff is weakly π\pi-cartesian if it satisfies the following : For any g:xx′′Xg:x\to x^{\prime\prime}\in X with πg=πf\pi g=\pi f, there uniquely exists h:xx′′Xh:x^{\prime}\to x^{\prime\prime}\in X such that g=hfg=h\circ f and πh=idπx\pi h={\rm id}_{\pi x^{\prime}}.

Definition 5.4.

Let π:XY𝖢𝖺𝗍\pi:X\longrightarrow Y\in{\sf Cat}. We say that π\pi is a pre-opfibration if it satisfies the following : For any f:πxyYf:\pi x\to y\in Y, there exists a weakly π\pi-cartesian morphism f~:xzMorX\tilde{f}:x\to z\in\mathrm{Mor}X such that πf~=f\pi\tilde{f}=f.

The following well-known propositions show that there is a one to one correspondence between pre-opfibtaions and normal oplax functors. For the detail, see part B of [12], for example.

Proposition 5.5.

For any normal oplax functor F:X𝖢𝖺𝗍F:X\longrightarrow{\sf Cat}, the natural projection πF:E(F)X\pi_{F}:E(F)\to X is a pre-opfibration.

Proof.

It will be shown in Proposition 5.11. ∎

Proposition 5.6.

For any pre-opfibration π:XY\pi:X\longrightarrow Y, we can construct a normal oplax functor Fπ:Y𝖢𝖺𝗍F_{\pi}:Y\longrightarrow{\sf Cat} such that there is an isomorphism XE(Fπ)𝖢𝖺𝗍X\cong E(F_{\pi})\in{\sf Cat} that commutes with π\pi and πFπ\pi_{F_{\pi}}. Furthermore, we have F=FπFF=F_{\pi_{F}} for any normal oplax functor F:X𝖢𝖺𝗍F:X\longrightarrow{\sf Cat}.

Sketch of proof.

For yObYy\in\mathrm{Ob}Y, we define the small category π1y\pi^{-1}y by Obπ1y={xObXπx=y}\mathrm{Ob}\pi^{-1}y=\{x\in\mathrm{Ob}X\mid\pi x=y\} and Morπ1y={fMorXπf=idy}\mathrm{Mor}\pi^{-1}y=\{f\in\mathrm{Mor}X\mid\pi f={\rm id}_{y}\}. For g:yyYg:y\to y^{\prime}\in Y and xObXx\in\mathrm{Ob}X with πx=y\pi x=y, we fix a weakly π\pi-cartesian lift g~x\tilde{g}_{x} of gg with sg~x=xs\tilde{g}_{x}=x by the axiom of choice. We choose them so that (idy)~x=idx\widetilde{({\rm id}_{y})}_{x}={\rm id}_{x}. Note that any morphism f:xxπ1yf:x\to x^{\prime}\in\pi^{-1}y induce a unique morphism tg~xtg~xt\tilde{g}_{x}\to t\tilde{g}_{x^{\prime}} that produces a functor g~:π1yπ1y\tilde{g}:\pi^{-1}y\to\pi^{-1}y^{\prime}. We define the normal oplax functor Fπ:Y𝖢𝖺𝗍F_{\pi}:Y\longrightarrow{\sf Cat} by Fπy=π1yF_{\pi}y=\pi^{-1}y and Fπg=g~F_{\pi}g=\tilde{g} for any yObYy\in\mathrm{Ob}Y and g:yyYg:y\to y^{\prime}\in Y. Then it is straightforward to check that FπF_{\pi} is indeed a normal oplax functor. Next we show that XE(Fπ)𝖢𝖺𝗍X\cong E(F_{\pi})\in{\sf Cat} with suitable compatibility. Note that we have ObE(Fπ)={(y,x)ObY×ObXπx=y}\mathrm{Ob}E(F_{\pi})=\{(y,x)\in\mathrm{Ob}Y\times\mathrm{Ob}X\mid\pi x=y\} and E(Fπ)((y,x),(y,x))={(g,f)g:yy,f:tg~xx}E(F_{\pi})((y,x),(y^{\prime},x^{\prime}))=\{(g,f)\mid g:y\to y^{\prime},f:t\tilde{g}_{x}\to x^{\prime}\}. We define the functor φ:E(Fπ)X\varphi:E(F_{\pi})\longrightarrow X by φ(y,x)=x\varphi(y,x)=x and φ(g,f)=fg~x\varphi(g,f)=f\circ\tilde{g}_{x} for any (y,x),(y,x)ObE(Fπ)(y,x),(y^{\prime},x^{\prime})\in\mathrm{Ob}E(F_{\pi}) and (g,f)E(Fπ)((y,x),(y,x))(g,f)\in E(F_{\pi})((y,x),(y^{\prime},x^{\prime})). It is straightforward to check that φ\varphi is an isomorphism with the desired compatibility. Finally, it is easily checked that we can choose the lifts of gg’s to construct FπF_{\pi} so that F=FπFF=F_{\pi_{F}}. This completes the proof. ∎

Now we generalize the above definitions to 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat} in the following.

Definition 5.7.

Let XX be a 𝖥𝗌𝖾𝗍{\sf Fset}-category. The following data F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat} is called a normal oplax functor.

  1. (1)

    FF is a normal oplax functor in the sense of Definition 5.1 when we forget the filtrations,

  2. (2)

    Ff:FxFx𝖥𝗌𝖾𝗍𝖼𝖺𝗍Ff:Fx\to Fx^{\prime}\in{\sf Fsetcat} for any f:xxMorXf:x\to x^{\prime}\in\mathrm{Mor}X,

  3. (3)

    For any f:xx,g:xx′′MorXf:x\to x^{\prime},g:x^{\prime}\to x^{\prime\prime}\in\mathrm{Mor}X and for any aFxa\in Fx, degτf,gadegf+deggdeggf\deg\tau_{f,g}a\leq\deg f+\deg g-\deg g\circ f.

Definition 5.8.

Let F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat} be a normal oplax functor. We construct a 𝖥𝗌𝖾𝗍{\sf Fset}-category E(F)E(F) as follows.

  1. (1)

    When we forget the filtration, E(F)E(F) coincides with the one defined in Definition 5.2.

  2. (2)

    deg(f,ξ):=degf+degξ\deg(f,\xi):=\deg f+\deg\xi for any (f,ξ)MorE(F)(f,\xi)\in\mathrm{Mor}E(F).

It is straightforward to check that the above E(F)E(F) is indeed a 𝖥𝗌𝖾𝗍{\sf Fset}-category. Note that we have a natural projection πF:E(F)X𝖥𝗌𝖾𝗍𝖼𝖺𝗍\pi_{F}:E(F)\longrightarrow X\in{\sf Fsetcat} defined by πF(x,a)=x\pi_{F}(x,a)=x and πF(f,ξ)=f\pi_{F}(f,\xi)=f. For a normal oplax functor F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat}, we can consider XX and FxFx’s as the “base” and the “fibers” of E(F)E(F) respectively. The following definitions contain Definitions 5.3 and 5.4 by considering the inclusion 𝖢𝖺𝗍𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Cat}\longrightarrow{\sf Fsetcat} of Proposition 3.5.

Definition 5.9.

Let π:XY𝖥𝗌𝖾𝗍𝖼𝖺𝗍\pi:X\longrightarrow Y\in{\sf Fsetcat} and f:xxXf:x\to x^{\prime}\in X. We say that ff is weakly π\pi-cartesian if it satisfies the following : For any g:xx′′Xg:x\to x^{\prime\prime}\in X with πg=πf\pi g=\pi f, there uniquely exists h:xx′′Xh:x^{\prime}\to x^{\prime\prime}\in X such that g=hfg=h\circ f and πh=idπx\pi h={\rm id}_{\pi x^{\prime}}. Further, it satisfies degg=degf+degh\deg g=\deg f+\deg h.

Definition 5.10.

Let π:XY𝖥𝗌𝖾𝗍𝖼𝖺𝗍\pi:X\longrightarrow Y\in{\sf Fsetcat}. We say that π\pi is a pre-opfibration if it satisfies the following : For any f:πxyYf:\pi x\to y\in Y, there exists a weakly π\pi-cartesian morphism f~:xzMorX\tilde{f}:x\to z\in\mathrm{Mor}X such that πf~=f\pi\tilde{f}=f and degf=degf~\deg f=\deg\tilde{f}.

Now we have the following propositions that are generalizations of Propositions 5.5 and 5.6.

Proposition 5.11.

Let F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat} be a normal oplax functor. Then the natural projection πF:E(F)X\pi_{F}:E(F)\longrightarrow X is a pre-opfibration.

Proof.

Let f:xxXf:x\to x^{\prime}\in X. For any aFxa\in Fx, we define f~=(f,idFfa):(x,a)(x,Ffa)\tilde{f}=(f,{\rm id}_{Ffa}):(x,a)\to(x^{\prime},Ffa). It is immediate to obtain πFf~=f\pi_{F}\tilde{f}=f and degf=degf~\deg f=\deg\tilde{f}. Now we show that f~\tilde{f} is weakly πF\pi_{F}-cartesian. Let bFxb\in Fx^{\prime} and ξ:FfabFx\xi:Ffa\to b\in Fx^{\prime}. If we have a morphism h=(u,θ):(x,Ffa)(x,b)h=(u,\theta):(x^{\prime},Ffa)\to(x^{\prime},b) with hf~=(f,ξ)h\circ\tilde{f}=(f,\xi) and πFh=idx\pi_{F}h={\rm id}_{x^{\prime}}, then it is immediate to see that we should have h=(idx,ξ)h=({\rm id}_{x^{\prime}},\xi). Conversely, we have (idx,ξ)(f,idFfa)=(f,ξτf,idx)=(f,ξ)({\rm id}_{x^{\prime}},\xi)\circ(f,{\rm id}_{Ffa})=(f,\xi\circ\tau_{f,{\rm id}_{x^{\prime}}})=(f,\xi). Further, we have deg(f,ξ)=degf+degξ=deg(f,idFfa)+deg(idx,ξ)\deg(f,\xi)=\deg f+\deg\xi=\deg(f,{\rm id}_{Ffa})+\deg({\rm id}_{x^{\prime}},\xi). This completes the proof. ∎

Proposition 5.12.

For any pre-opfibration π:XY\pi:X\longrightarrow Y, we can construct a normal oplax functor Fπ:Y𝖥𝗌𝖾𝗍𝖼𝖺𝗍F_{\pi}:Y\longrightarrow{\sf Fsetcat} such that there is an isomorphism XE(Fπ)𝖥𝗌𝖾𝗍𝖼𝖺𝗍X\cong E(F_{\pi})\in{\sf Fsetcat} that commutes with π\pi and πFπ\pi_{F_{\pi}}. Furthermore, we have F=FπFF=F_{\pi_{F}} for any normal oplax functor F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat}.

Proof.

The construction of FπF_{\pi} is same as Proposition 5.6, however, we should check that FπF_{\pi} is indeed a normal oplax functor in the sense of Definition 5.7. Namely, we should check that i) FπgF_{\pi}g is a morphism in 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat}, and ii) degτg,gxdegg+deggdeggg\deg\tau_{g,g^{\prime}}x\leq\deg g+\deg g^{\prime}-\deg g^{\prime}\circ g for any g:yy,g:yy′′Yg:y\to y^{\prime},g^{\prime}:y^{\prime}\to y^{\prime\prime}\in Y and xπ1yx\in\pi^{-1}y.

  1. i)

    It is enough to show that FπgF_{\pi}g preserves filtrations. Let f:xxπ1yf:x\to x^{\prime}\in\pi^{-1}y. Since (Fπg)f(F_{\pi}g)f is induced by the universality of weakly π\pi-cartesian morphism g~x\tilde{g}_{x}, we have degg~x+deg(Fπg)f=degg~xfdegg~x+degf\deg\tilde{g}_{x}+\deg(F_{\pi}g)f=\deg\tilde{g}_{x^{\prime}}\circ f\leq\deg\tilde{g}_{x^{\prime}}+\deg f. Since we have degg=degg~x=degg~x\deg g=\deg\tilde{g}_{x}=\deg\tilde{g}_{x^{\prime}}, we obtain (degFπg)fdegf(\deg F_{\pi}g)f\leq\deg f.

  2. ii)

    Since the morphism τg,gx\tau_{g,g^{\prime}}x is induced from the universality of the weakly π\pi-cartesian morphism (gg~)x(\widetilde{g^{\prime}\circ g})_{x}, we have deg(gg~)x+degτg,gx=degg~xg~xdegg~x+degg~x\deg(\widetilde{g^{\prime}\circ g})_{x}+\deg\tau_{g,g^{\prime}}x=\deg\tilde{g^{\prime}}_{x^{\prime}}\circ\tilde{g}_{x}\leq\deg\tilde{g}_{x}+\deg\tilde{g^{\prime}}_{x^{\prime}}, where we put x=tg~xx^{\prime}=t\tilde{g}_{x}. Hence we obtain degτg,gxdegg~x+degg~xdeg(gg~)x=degg+deggdeggg\deg\tau_{g,g^{\prime}}x\leq\deg\tilde{g}_{x}+\deg\tilde{g^{\prime}}_{x^{\prime}}-\deg(\widetilde{g^{\prime}\circ g})_{x}=\deg g+\deg g^{\prime}-\deg g^{\prime}\circ g.

Next we check that the functor φ:E(Fπ)X\varphi:E(F_{\pi})\longrightarrow X defined in the proof of Proposition 5.6 is also an isomorphism in 𝖥𝗌𝖾𝗍𝖼𝖺𝗍{\sf Fsetcat}. We should check that degφ(g,f)=deg(g,f)=degg+degf\deg\varphi(g,f)=\deg(g,f)=\deg g+\deg f for any (g,f)MorE(Fπ)(g,f)\in\mathrm{Mor}E(F_{\pi}). Recall that we define φ(g,f)=fg~x\varphi(g,f)=f\circ\tilde{g}_{x}, where g:yyYg:y\to y^{\prime}\in Y and f:xxπ1yf:x\to x^{\prime}\in\pi^{-1}y. Since g~x\tilde{g}_{x} is weakly π\pi-cartesian, we have degfg~x=degf+degg~x=deg(g,f)\deg f\circ\tilde{g}_{x}=\deg f+\deg\tilde{g}_{x}=\deg(g,f). Finally, we note that we have F=FπFF=F_{\pi_{F}} similarly to Proposition 5.6. This completes the proof. ∎

Corollary 5.13.

There is a one to one correspondence between normal oplax functors FF and pre-opfibrations π\pi given by πF\pi_{F} and FπF_{\pi}.

Proof.

It is immediate from Proposition 5.11 and Proposition 5.12. ∎

5.2 Magnitude of pre-opfibrations

Now we show a remarkable property that the magnitude of a pre-opfibration splits as the product of those of the “fiber” and the “base” if they have. The following two propositions are inspired by Lemma 1.14 of [17] and Theorem 2.3.11 of [15].

Proposition 5.14.

Let F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat} be a normal oplax functor. Suppose that XX, FxFx for any xObXx\in\mathrm{Ob}X and E(F)E(F) are all 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-categories of finite type. If XX and FxFx for any xObXx\in\mathrm{Ob}X have weightings kk^{\bullet}, then E(F)E(F) has a weighting k(x,a)=kxkak^{(x,a)}=k^{x}k^{a}.

Proof.

It follows from the following calculation.

(x,a)ObE(F)#E(F)((x,b),(x,a))kxka\displaystyle\sum_{(x,a)\in\mathrm{Ob}E(F)}\#E(F)((x^{\prime},b),(x,a))k^{x}k^{a}
=xObXaFxfX(x,x)ξFx(Ffb,a)qdegf+degξkxka\displaystyle=\sum_{x\in\mathrm{Ob}X}\sum_{a\in Fx}\sum_{f\in X(x^{\prime},x)}\sum_{\xi\in Fx(Ffb,a)}q^{\deg f+\deg\xi}k^{x}k^{a}
=xObXfX(x,x)qdegfkx\displaystyle=\sum_{x\in\mathrm{Ob}X}\sum_{f\in X(x^{\prime},x)}q^{\deg f}k^{x}
=1.\displaystyle=1.

Proposition 5.15.

Let F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat} be a normal oplax functor with FxFxFx\cong Fx^{\prime} for any x,xObXx,x^{\prime}\in\mathrm{Ob}X. Suppose that XX, FxFx, E(F)E(F) are all finite 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-categories. If XX, FxFx and E(F)E(F) all have the magnitude, then 𝖬𝖺𝗀E(F)=𝖬𝖺𝗀X𝖬𝖺𝗀Fx{\sf Mag}E(F)={\sf Mag}X\cdot{\sf Mag}Fx.

Proof.

It follows from Proposition 5.15. ∎

In the following, we discuss when the assumptions in the above propositions are satisfied. We say that a subset U0U\subset\mathbb{R}_{\geq 0} is left finite if it is a support of some left finite function f:0f:\mathbb{R}_{\geq 0}\longrightarrow\mathbb{Q}.

Lemma 5.16.

Let F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat} be a normal oplax functor. Suppose that XX and any FxFx for xObXx\in\mathrm{Ob}X are 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-categories. Then E(F)E(F) is also a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category.

Proof.

Let X(x,x)():={fX(x,x)degf=}X(x,x^{\prime})(\ell):=\{f\in X(x,x^{\prime})\mid\deg f=\ell\} for any x,xObXx,x^{\prime}\in\mathrm{Ob}X and 0\ell\in\mathbb{R}_{\geq 0}. Let U:={0X(x,x)()}U:=\{\ell\in\mathbb{R}_{\geq 0}\mid X(x,x^{\prime})(\ell)\neq\emptyset\}, which is left finite since X(x,x)X(x,x^{\prime}) is collectable. Then we have

E(F)((x,a),(x,b))\displaystyle E(F)((x,a),(x^{\prime},b)) ={(f,ξ)fX(x,x),ξFx(Ffa,b)}\displaystyle=\{(f,\xi)\mid f\in X(x,x^{\prime}),\xi\in Fx^{\prime}(Ffa,b)\}
=U{(f,ξ)fX(x,x)(),ξFx(Ffa,b)}.\displaystyle=\bigcup_{\ell\in U}\{(f,\xi)\mid f\in X(x,x^{\prime})(\ell),\xi\in Fx^{\prime}(Ffa,b)\}.

Since each X(x,x)()X(x,x^{\prime})(\ell) is a finite set and each Fx(Ffa,b)Fx^{\prime}(Ffa,b) is collectable, the filtered set {(f,ξ)fX(x,x)(),ξFx(Ffa,b)}\{(f,\xi)\mid f\in X(x,x^{\prime})(\ell),\xi\in Fx^{\prime}(Ffa,b)\} is collectable, whose elements have degree \geq\ell. Hence the left finiteness of λ\lambda implies that the filtered set S{(f,ξ)fX(x,x)(),ξFx(Ffa,b)}\bigcup_{\ell\in S}\{(f,\xi)\mid f\in X(x,x^{\prime})(\ell),\xi\in Fx^{\prime}(Ffa,b)\} is collectable. This completes the proof. ∎

Now we have the following two obvious corollaries.

Corollary 5.17.

Let F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat} be a normal oplax functor. Suppose that XX and any FxFx for xObXx\in\mathrm{Ob}X are finite 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-categories. Then E(F)E(F) is also a finite 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category.

Corollary 5.18.

Let F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat} be a normal oplax functor. Suppose that XX and any FxFx for xObXx\in\mathrm{Ob}X are finite tame categories. Then E(F)E(F) is also a finite tame category.

Lemma 5.19.

Let F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat} be a normal oplax functor. Suppose that XX and any FxFx for xObXx\in\mathrm{Ob}X are 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-categories of finite type. If any fiber of Ff:ObFxObFxFf:\mathrm{Ob}Fx\longrightarrow\mathrm{Ob}Fx^{\prime} is a finite set for any f:xxXf:x\to x^{\prime}\in X, then E(F)E(F) is also a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category of finite type.

Proof.

We have

(x,b)ObE(F)E(F)((x,a),(x,b))\displaystyle\bigcup_{(x^{\prime},b)\in\mathrm{Ob}E(F)}E(F)((x,a),(x^{\prime},b)) ={(f,ξ)fX(x,x),ξFx(Ffa,b)}\displaystyle=\{(f,\xi)\mid f\in X(x,x^{\prime}),\xi\in Fx^{\prime}(Ffa,b)\}
=xObXbObFxfX(x,x)(degf)×Fx(Ffa,b)\displaystyle=\bigcup_{x^{\prime}\in\mathrm{Ob}X}\bigcup_{b\in\mathrm{Ob}Fx^{\prime}}\bigcup_{f\in X(x,x^{\prime})}\ast(\deg f)\times Fx^{\prime}(Ffa,b)
=xObXfX(x,x)((degf)×bObFxFx(Ffa,b)).\displaystyle=\bigcup_{x^{\prime}\in\mathrm{Ob}X}\bigcup_{f\in X(x,x^{\prime})}\left(\ast(\deg f)\times\bigcup_{b\in\mathrm{Ob}Fx^{\prime}}Fx^{\prime}(Ffa,b)\right).

Here, since each FxFx^{\prime} is of finite type, fX(x,x)((degf)×bObFxFx(Ffa,b))=:C(x,x)\bigcup_{f\in X(x,x^{\prime})}\left(\ast(\deg f)\times\bigcup_{b\in\mathrm{Ob}Fx^{\prime}}Fx^{\prime}(Ffa,b)\right)=:C(x,x^{\prime}) is a collectable filtered set by the same argument of Proposition 5.16 and Lemma 3.16. Let us denote the lowest degree of elements of X(x,x)X(x,x^{\prime}) by (x,x)\ell(x,x^{\prime}). We define ObX(x,):={xObX(x,x)=}\mathrm{Ob}X(x,\ell):=\{x^{\prime}\in\mathrm{Ob}X\mid\ell(x,x^{\prime})=\ell\}. Then a subset U:={0ObX(x,)}0U:=\{\ell\in\mathbb{R}_{\geq 0}\mid\mathrm{Ob}X(x,\ell)\neq\emptyset\}\subset\mathbb{R}_{\geq 0} is left finite since XX is of finite type. Furthermore, the filtered set xObX(x,)C(x,x)\bigcup_{x^{\prime}\in\mathrm{Ob}X(x,\ell)}C(x,x^{\prime}) is collectable with lowest degree \geq\ell for each 0\ell\in\mathbb{R}_{\geq 0}. Hence xObXC(x,x)=UxObX(x,)C(x,x)\bigcup_{x^{\prime}\in\mathrm{Ob}X}C(x,x^{\prime})=\bigcup_{\ell\in U}\bigcup_{x^{\prime}\in\mathrm{Ob}X(x,\ell)}C(x,x^{\prime}) is collectable. On the other hand, we have

(x,a)ObE(F)E(F)((x,a),(x,b))\displaystyle\bigcup_{(x,a)\in\mathrm{Ob}E(F)}E(F)((x,a),(x^{\prime},b)) ={(f,ξ)fX(x,x),ξFx(Ffa,b)}\displaystyle=\{(f,\xi)\mid f\in X(x,x^{\prime}),\xi\in Fx^{\prime}(Ffa,b)\}
=xObXaObFxfX(x,x)(degf)×Fx(Ffa,b)\displaystyle=\bigcup_{x\in\mathrm{Ob}X}\bigcup_{a\in\mathrm{Ob}Fx}\bigcup_{f\in X(x,x^{\prime})}\ast(\deg f)\times Fx^{\prime}(Ffa,b)
=xObXfX(x,x)((degf)×aObFxFx(Ffa,b)),\displaystyle=\bigcup_{x\in\mathrm{Ob}X}\bigcup_{f\in X(x,x^{\prime})}\left(\ast(\deg f)\times\bigcup_{a\in\mathrm{Ob}Fx^{\prime}}Fx^{\prime}(Ffa,b)\right),

where fX(x,x)((degf)×aObFxFx(Ffa,b))\bigcup_{f\in X(x,x^{\prime})}\left(\ast(\deg f)\times\bigcup_{a\in\mathrm{Ob}Fx^{\prime}}Fx^{\prime}(Ffa,b)\right) turns out to be collectable by the assumption for fibers. Hence (x,a)ObE(F)E(F)((x,a),(x,b))\bigcup_{(x,a)\in\mathrm{Ob}E(F)}E(F)((x,a),(x^{\prime},b)) is collectable by the similar argument to the above. This completes the proof. ∎

Example 5.20.

Let (Γ,S)(\Gamma,S) be a finitely generated group considered as a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category as in 3.5.4. We denote the underlying set of the group Γ\Gamma by |Γ||\Gamma|. We also denote the automorphism of |Γ||\Gamma| by the left multiplication of gΓg\in\Gamma by aga_{g}. We construct a normal oplax functor F:(Γ,S)𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:(\Gamma,S)\longrightarrow{\sf Fsetcat} by F=|Γ|F\ast=|\Gamma| and Fg=agFg=a_{g}. Then FF satisfies the condition of Lemma 5.19, hence E(F)E(F) is a 𝖢𝖥𝗌𝖾𝗍{\sf CFset}-category of finite type. Actually, E(F)E(F) is isomorphic to the Cayley graph Cay(Γ,S){\rm Cay}(\Gamma,S), and the coincidence of the weightings of (Γ,S)(\Gamma,S) and Cay(Γ,S){\rm Cay}(\Gamma,S) observed in 3.5.4 can be explained by Proposition 5.15.

5.3 Restriction to metric spaces

In the following, we consider the restrictions of pre-opfibrations and oplax functors to 𝖬𝖾𝗍{\sf Met}, which will turn out to coincide with Leinster’s metric fibrations ([15]). We give some examples of metric fibrations by utilizing Corollary 5.13.

For a normal oplax functor F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat}, the 𝖥𝗌𝖾𝗍{\sf Fset}-category E(F)E(F) is not necessarily a metric space in general, even if XX and FxFx are metric spaces for any xObXx\in\mathrm{Ob}X. To consider a restriction of normal oplax funtors to 𝖬𝖾𝗍{\sf Met}, we need the following notion.

Definition 5.21.

A normal oplax functor F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat} is a metric action if it satisfies the following:

  1. (1)

    FxFx is a metric space for any xObXx\in\mathrm{Ob}X.

  2. (2)

    Let XmX_{m} be the full subcategory of XX consisting of objects xx with FxFx\neq\emptyset. Then XmX_{m} is a metric space.

  3. (3)

    Let fXm(x,x)f\in X_{m}(x,x^{\prime}) and gXm(x,x)g\in X_{m}(x^{\prime},x) be the unique morphisms between xx and xx^{\prime}. Then we have FgFf=idFxFgFf={\rm id}_{Fx} and FfFg=idFxFfFg={\rm id}_{Fx^{\prime}}.

Proposition 5.22.

Let F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat} be a normal oplax functor. Then E(F)E(F) is a metric space if and only if FF is a metric action.

Proof.

It is straightforward to check that E(F)E(F) is a metric space when FF is a metric action. We show the converse. Suppose that E(F)E(F) is a metric space. Since FxFx can be embedded to E(F)E(F) for any xObXx\in\mathrm{Ob}X by ObFxa(x,a)\mathrm{Ob}Fx\ni a\mapsto(x,a) and MorFxξ(idx,ξ)\mathrm{Mor}Fx\ni\xi\mapsto({\rm id}_{x},\xi), it should be a metric space. Further, there should be exactly one morphism between any two objects x,xObXx,x^{\prime}\in\mathrm{Ob}X with Fx,FxFx,Fx^{\prime}\neq\emptyset, because E(F)E(F) has only one morphism between two objects. In particular, we have X(x,x)={idx}X(x,x)=\{{\rm id}_{x}\} for any xObXx\in\mathrm{Ob}X. Let x,xObXx,x^{\prime}\in\mathrm{Ob}X with Fx,FxFx,Fx^{\prime}\neq\emptyset, xxx\neq x^{\prime}, aFxa\in Fx and X(x,x)={f},X(x,x)={g}X(x,x^{\prime})=\{f\},X(x^{\prime},x)=\{g\}. Denoting the distance function of E(F)E(F) by dd, we have

d((x,a),(x,Ffa))\displaystyle d((x,a),(x^{\prime},Ffa)) =degf,\displaystyle=\deg f,
d((x,Ffa),(x,a))\displaystyle d((x^{\prime},Ffa),(x,a)) =degg+degτf,ga.\displaystyle=\deg g+\deg\tau_{f,g}a.

Hence we obtain that degfdegg\deg f\geq\deg g. On the other hand, we have

d((x,Ffa),(x,FgFfa))\displaystyle d((x^{\prime},Ffa),(x,FgFfa)) =degg,\displaystyle=\deg g,
d((x,FgFfa),(x,Ffa))\displaystyle d((x,FgFfa),(x^{\prime},Ffa)) =degf+degτg,fFfa.\displaystyle=\deg f+\deg\tau_{g,f}Ffa.

Hence we obtain that deggdegf\deg g\geq\deg f, and thus degf=degg>0,degτf,ga=0\deg f=\deg g>0,\deg\tau_{f,g}a=0. This implies that the full subcategory of XX consisting of objects xx with FxFx\neq\emptyset is a metric space, and FgFf=idFxFgFf={\rm id}_{Fx} for any such x,xObXx,x^{\prime}\in\mathrm{Ob}X and fX(x,x),gX(x,x)f\in X(x,x^{\prime}),g\in X(x^{\prime},x). This completes the proof. ∎

Definition 5.23.

Let X,YX,Y be metric spaces. A 11-Lipschitz map π:XY\pi:X\longrightarrow Y is a metric fibration if it satisfies the following: For any xXx\in X and yYy\in Y, there uniquely exists zπ1yz\in\pi^{-1}y such that

  1. (1)

    d(x,z)=d(πx,y)d(x,z)=d(\pi x,y),

  2. (2)

    d(x,w)=d(x,z)+d(w,z)d(x,w)=d(x,z)+d(w,z) for any wπ1yw\in\pi^{-1}y.

The following shows that the restriction of pre-opfibrations to 𝖬𝖾𝗍{\sf Met} is exactly the metric fibrations.

Proposition 5.24.

A 11-Lipschitz map π:XY\pi:X\longrightarrow Y is a pre-opfibration if and only if it is a metric fibration.

Proof.

Suppose that π\pi is a metric fibration. For any xX,yYx\in X,y\in Y and f:πxyf:\pi x\to y, there exist zπ1yz\in\pi^{-1}y and f~:xz\tilde{f}:x\to z satisfying (1) and (2) of Definition 5.23. By (1)(1), we have degf=degf~\deg f=\deg\tilde{f}. Further, f~\tilde{f} is weakly π\pi-cartesian as follows. Let g:xwg:x\to w with πg=f\pi g=f, namely wπ1yw\in\pi^{-1}y. Then there exists h:zwh:z\to w with degh+degf~=degg\deg h+\deg\tilde{f}=\deg g by (2). Moreover, such a hh is unique since XX is a metric space. Conversely, suppose that π\pi is a pre-opfibration. Let xX,yYx\in X,y\in Y and let f:πxyf:\pi x\to y be the unique morphism in Y(πx,y)Y(\pi x,y). Then there exists a f~:xzX\tilde{f}:x\to z\in X with πf~=f\pi\tilde{f}=f and degf~=degf\deg\tilde{f}=\deg f. Namely, we have zπ1yz\in\pi^{-1}y and d(x,z)=d(πx,y)d(x,z)=d(\pi x,y). If there exists another f~:xz\tilde{f}^{\prime}:x\to z^{\prime}, then the condition in Definition 5.9 implies that z=zz=z^{\prime}. Hence such a zz is unique. Moreover, the same condition implies (2) of Definition 5.23. This completes the proof. ∎

Proposition 5.25.

For any metric fibration π:XY\pi:X\longrightarrow Y, the normal oplax functor FπF_{\pi} is a metric action.

Proof.

We only verify (3) of Definition 5.21. Let y,yYy,y^{\prime}\in Y and fY(y,y),gY(y,y)f\in Y(y,y^{\prime}),g\in Y(y^{\prime},y) be unique morphisms. For any xFπyx\in F_{\pi}y, we have d(y,y)=d(x,Fπfx)=d(Fπfx,x)=d(Fπfx,FπgFπfx)+d(FπgFπfx,x)=d(y,y)+d(FπgFπfx,x)d(y,y^{\prime})=d(x,F_{\pi}fx)=d(F_{\pi}fx,x)=d(F_{\pi}fx,F_{\pi}gF_{\pi}fx)+d(F_{\pi}gF_{\pi}fx,x)=d(y^{\prime},y)+d(F_{\pi}gF_{\pi}fx,x). Hence we obtain that d(FπgFπfx,x)=0d(F_{\pi}gF_{\pi}fx,x)=0, namely FπgFπf=idFπyF_{\pi}gF_{\pi}f={\rm id}_{F_{\pi}y}. Similarly, we have FπfFπg=idFπyF_{\pi}fF_{\pi}g={\rm id}_{F_{\pi}y^{\prime}}. This completes the proof. ∎

Corollary 5.26.

The one to one correspondence of Corollary 5.13 is restricted to a one to one correspondence between metric actions and metric fibrations.

Proof.

It follows from Corollary 5.13, Proposition 5.24 and Proposition 5.25. ∎

Before giving examples of metric fibrations, we show that the genuine functoriality of a metric action implies the triviality of the fibration.

Definition 5.27.

Let π:XY\pi:X\longrightarrow Y and π:XY\pi^{\prime}:X^{\prime}\longrightarrow Y be metric fibrations. We say that π\pi and π\pi^{\prime} are isomorphic if there is an isometry φ:XX\varphi:X\cong X^{\prime} with πφ=π\pi^{\prime}\circ\varphi=\pi.

Proposition 5.28.

Let XX and YY be metric spaces. Let F:X𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:X\longrightarrow{\sf Fsetcat} be a metric action with Fx=YFx=Y for a fixed xXx\in X. If FF is a functor, namely τf,g\tau_{f,g}’s are all identities for any f,gMorXf,g\in\mathrm{Mor}X, then the metric fibration πF\pi_{F} is isomorphic to the projection X×YXX\times Y\longrightarrow X.

Proof.

Fix a point xXx\in X. Then any points of FxFx^{\prime} for any xXx^{\prime}\in X can be expressed as F(x,x)yF(x,x^{\prime})y for some yFxy\in Fx, where (x,x)(x,x^{\prime}) is the unique morphism from xx to xx^{\prime}. We define a 11-Lipschitz map φ:E(F)X×Y\varphi:E(F)\longrightarrow X\times Y by φ(x,F(x,x)y)=(x,y)\varphi(x^{\prime},F(x,x^{\prime})y)=(x^{\prime},y). Then it preserves distances since we have

d((x,F(x,x)y),(x′′,F(x,x′′)y))\displaystyle d((x^{\prime},F(x,x^{\prime})y),(x^{\prime\prime},F(x,x^{\prime\prime})y^{\prime})) =d(x,x′′)+d(F(x,x′′)F(x,x)y,F(x,x′′)y)\displaystyle=d(x^{\prime},x^{\prime\prime})+d(F(x^{\prime},x^{\prime\prime})F(x,x^{\prime})y,F(x,x^{\prime\prime})y^{\prime})
=d(x,x′′)+d(F(x,x′′)y,F(x,x′′)y)\displaystyle=d(x^{\prime},x^{\prime\prime})+d(F(x,x^{\prime\prime})y,F(x,x^{\prime\prime})y^{\prime})
=d(x,x′′)+d(y,y)\displaystyle=d(x^{\prime},x^{\prime\prime})+d(y,y^{\prime})
=d((x,y),(x′′,y)).\displaystyle=d((x^{\prime},y),(x^{\prime\prime},y^{\prime})).

Further, φ\varphi is apparently a bijection. Hence φ\varphi gives the desired isomorphism. This completes the proof. ∎

Example 5.29.

Let KnK_{n} be the complete graph with nn vertices. Let T:K2K2{\mathrm{T}}:K_{2}\longrightarrow K_{2} be the non-trivial involution. We construct a metric action F:K3𝖥𝗌𝖾𝗍F:K_{3}\longrightarrow{\sf Fset} by Fv=K2Fv=K_{2} for each vK3v\in K_{3}, Fe=idK2Fe={\rm id}_{K_{2}} for two edges of K3K_{3} and Fe=TFe^{\prime}={\mathrm{T}} for the rest edge ee^{\prime}. Then E(F)E(F) is a graph shown in Figure 1 right. The left graph in Figure 1 is K3×K2K_{3}\times K_{2}, hence both have the same magnitude by Proposition 5.15 and Corollary 5.18, although they are not isomorphic (they have different girth). Further, they are diagonal since the left is a product of diagonal graphs and the right is a Pawful graph ( Theorem 4.4 of [10]). Thus they have the same magnitude homology. Furthermore, they both have trivial path homologies by Theorem 1.3 in [1].

Figure 1: The left is K3×K2K_{3}\times K_{2}, and the right is E(F)E(F) for FF defined in Example 5.29. The right one is also isomorphic to the complete bipartite graph K3,3K_{3,3}. It is also shown in Example 3.7 of [16] that K3×K2K_{3}\times K_{2} and K3,3K_{3,3} has the same magnitude.

In the case that the base of a metric fibration is a cyclic graph, it is isomorphic to one obtained by “twisting a fiber along only one edge”, as follows.

Proposition 5.30.

Let CnC_{n} be the cyclic graph with n(3)n(\geq 3) vertices, and let π:XCn\pi:X\longrightarrow C_{n} be a metric fibration. We label the vertices of CnC_{n} by V(Cn)={1,,n}V(C_{n})=\{1,\dots,n\}, and we denote π11\pi^{-1}1 by YY. Then there exists an isometry θ:YY\theta:Y\longrightarrow Y, and π\pi is isomorphic to a metric fibration πθ\pi_{\theta} constructed from θ\theta as follows : we construct a metric action Fθ:Cn𝖥𝗌𝖾𝗍𝖼𝖺𝗍F_{\theta}:C_{n}\longrightarrow{\sf Fsetcat} by Fθi=YF_{\theta}i=Y, F(i,i+1)=idYF(i,i+1)={\rm id}_{Y} for any 1in1\leq i\leq n except for F(n,1)=θF(n,1)=\theta. For the other pair of vertices (j,k)(j,k), we define F(j,k)F(j,k) by the composition of F(i,i+1)F(i,i+1)’s along the shortest edge path connecting the vertices jj and kk. Further, π\pi is isomorphic to the projection Y×CnCnY\times C_{n}\longrightarrow C_{n} if nn is even.

Proof.

Note that the metric space E(Fθ)E(F_{\theta}) consists of points V(Cn)×YV(C_{n})\times Y with the distance function dEd_{E} defined by

dE((i,y),(j,y))\displaystyle d_{E}((i,y),(j,y^{\prime}))
={dCn(i,j)+dY(y,y) if the shortest path dose not contain the edge {1,n},dCn(i,j)+dY(y,θy) if the shortest path contains the edge {1,n}.\displaystyle=\begin{cases}d_{C_{n}}(i,j)+d_{Y}(y,y^{\prime})&\text{ if the shortest path dose not contain the edge }\{1,n\},\\ d_{C_{n}}(i,j)+d_{Y}(y,\theta y^{\prime})&\text{ if the shortest path contains the edge }\{1,n\}.\end{cases}

Let θ=Fπ(n,1)Fπ(n1,n)Fπ(1,2)\theta=F_{\pi}(n,1)\circ F_{\pi}(n-1,n)\circ\dots\circ F_{\pi}(1,2). We define a map φ:X=E(Fπ)E(Fθ)\varphi:X=E(F_{\pi})\longrightarrow E(F_{\theta}) by φx=(i,Fπ1(1,2)Fπ1(i1,i)x)\varphi x=(i,F^{-1}_{\pi}(1,2)\circ\dots\circ F^{-1}_{\pi}(i-1,i)x) for xπ1ix\in\pi^{-1}i and 1in1\leq i\leq n. Now we verify that φ\varphi is an isometry. Let xπ1ix\in\pi^{-1}i and xπ1jx^{\prime}\in\pi^{-1}j with 1ijn1\leq i\leq j\leq n. When the shortest path connecting ii and jj does not contain the edge {1,n}\{1,n\}, we have

dE(φx,φx)\displaystyle d_{E}(\varphi x,\varphi x^{\prime})
=dCn(i,j)+dY(Fπ1(1,2)Fπ1(i1,i)x,Fπ1(1,2)Fπ1(j1,j)x)\displaystyle=d_{C_{n}}(i,j)+d_{Y}(F^{-1}_{\pi}(1,2)\circ\dots\circ F^{-1}_{\pi}(i-1,i)x,F^{-1}_{\pi}(1,2)\circ\dots\circ F^{-1}_{\pi}(j-1,j)x^{\prime})
=dCn(i,j)+dY(Fπ(j1,j)Fπ(i,i+1)x,x)\displaystyle=d_{C_{n}}(i,j)+d_{Y}(F_{\pi}(j-1,j)\circ\dots\circ F_{\pi}(i,i+1)x,x^{\prime})
=dX(x,x).\displaystyle=d_{X}(x,x^{\prime}).

Note here that we have Fπ(j1,j)Fπ(i,i+1)=Fπ(i,j)F_{\pi}(j-1,j)\circ\dots\circ F_{\pi}(i,i+1)=F_{\pi}(i,j) because the sequence i,i+1,,ji,i+1,\dots,j is the shortest path and FπF_{\pi} should satisfy (3) of Definition 5.7. When the shortest path connecting ii and jj contains the edge {1,n}\{1,n\}, we have

dE(φx,φx)\displaystyle d_{E}(\varphi x,\varphi x^{\prime})
=dCn(i,j)+dY(Fπ1(1,2)Fπ1(i1,i)x,θFπ1(1,2)Fπ1(j1,j)x)\displaystyle=d_{C_{n}}(i,j)+d_{Y}(F^{-1}_{\pi}(1,2)\circ\dots\circ F^{-1}_{\pi}(i-1,i)x,\theta F^{-1}_{\pi}(1,2)\circ\dots\circ F^{-1}_{\pi}(j-1,j)x^{\prime})
=dCn(i,j)+dY(Fπ1(1,2)Fπ1(i1,i)x,Fπ(n,1)Fπ(j,j+1)x)\displaystyle=d_{C_{n}}(i,j)+d_{Y}(F^{-1}_{\pi}(1,2)\circ\dots\circ F^{-1}_{\pi}(i-1,i)x,F_{\pi}(n,1)\circ\dots\circ F_{\pi}(j,j+1)x^{\prime})
=dCn(i,j)+dY(x,Fπ(i1,i)Fπ(1,2)Fπ(n,1)Fπ(j,j+1)x)\displaystyle=d_{C_{n}}(i,j)+d_{Y}(x,F_{\pi}(i-1,i)\circ\dots\circ F_{\pi}(1,2)\circ F_{\pi}(n,1)\circ\dots\circ F_{\pi}(j,j+1)x^{\prime})
=dX(x,x).\displaystyle=d_{X}(x,x^{\prime}).

Note that we have Fπ(i1,i)Fπ(1,2)Fπ(n,1)Fπ(j,j+1)=Fπ(j,i)F_{\pi}(i-1,i)\circ\dots\circ F_{\pi}(1,2)\circ F_{\pi}(n,1)\circ\dots\circ F_{\pi}(j,j+1)=F_{\pi}(j,i) similarly to the above. Hence φ\varphi gives the desired isomorphism. Next, we suppose that n=2Nn=2N. Since we have

Fπ(1,N+1)\displaystyle F_{\pi}(1,N+1) =Fπ(N,N+1)Fπ(1,2)\displaystyle=F_{\pi}(N,N+1)\circ\dots\circ F_{\pi}(1,2)
=Fπ(N+2,N+1)Fπ(1,2N)\displaystyle=F_{\pi}(N+2,N+1)\circ\dots\circ F_{\pi}(1,2N)

from (3) of Definition 5.7, we obtain that θ=idY\theta={\rm id}_{Y}. Hence π\pi is isomorphic to the projection Y×CnCnY\times C_{n}\longrightarrow C_{n} by Proposition 5.28. This completes the proof. ∎

Remark 5.31.

We don’t establish the determination of the magnitude homology of metric fibrations even for the case of graphs. As far as we calculate for some examples by using Hepworth-Willerton’s computer program ([11]), we have not found any difference between rational magnitude homology of metric fibrations and direct products.

Example 5.32.

Let CRC_{R} be a circle of radius RR in 2\mathbb{R}^{2} with the metric induced from 2\mathbb{R}^{2}. We construct a metric action F:CR𝖥𝗌𝖾𝗍𝖼𝖺𝗍F:C_{R}\longrightarrow{\sf Fsetcat} with Fc=C2R/πFc=C_{2R/\pi} for cCRc\in C_{R}. For c,cCRc,c^{\prime}\in C_{R}, we denote the unique morphism ccc\to c^{\prime} by (c,c)(c,c^{\prime}). We define isometries F(c,c):C2R/πC2R/πF(c,c^{\prime}):C_{2R/\pi}\longrightarrow C_{2R/\pi} by the (anti-)clockwise πd(c,c)/2R\pi d(c,c^{\prime})/2R-rotation if the shortest geodesic from cc to cc^{\prime} is (anti-)clockwise. It is easy to check that FF is a metric action. We don’t know whether they are isormorphic or not.

References

  • [1] Y. Asao, Magnitude homology and Path homology, (2022), Bull. London Math. Soc., 55(1)(2023), 375–398.
  • [2] S. Bosch, U. Guntzer, R. Remmert, Non-archimedean analysis : a systematic approach to rigid analytic geometry, Berlin, New York : Springer-Verlag, 1984.
  • [3] J-L. Brylinski, A differential complex for Poisson manifolds, Differential Geom. 28(1)(1988), 93–114.
  • [4] D. Carranza, B. Doherty, M. Opie, M. Sarazola, and L.-Z. Wong, Cofibration category of digraphs for path homology, preprint, arXiv:2212.12568, (2022).
  • [5] J. Cirici, D. -E. Santander, M. Livernet, S. Whitehouse, Model category structures and spectral sequences, Proc. Royal Soc. Edinburgh Sect. A: Mathematics, 150 (6)(2020) , 2815 – 2848.
  • [6] M. Gerstenhaber and S.D. Schack, Simplicial cohomology is Hochschild cohomology, J. Pure Appl. Algebra 30 (1983), 143–156.
  • [7] P. Goerss and J. F. Jardine, Simplicial homotopy theory, Progress in Mathematics, Birkhäuser (1999) Modern Birkhäuser Classics (2009). arXiv:1902.07044, 2019.
  • [8] A. Grigor’yan, Y. Lin, Y. Muranov and S.-T. Yau, Homotopy theory for digraphs, Pure Appl. Math. Q. 10 (2014), no. 4, 619–674.
  • [9] A. Grigor’yan, Y. Muranov and S.-T. Yau,On a cohomology of digraphs and Hochschild cohomology, J. Homotopy Relat. Struct. 11 (2016), 209–230.
  • [10] Y. Gu, Graph magnitude homology via algebraic Morse theory, preprint, arXiv:1809.07240, (2018).
  • [11] R. Hepworth and S. Willerton, Categorifying the magnitude of a graph, Homology Homotopy Appl. 19 (2017), 31–60.
  • [12] P. T. Johnstone, Sketches of an Elephant : A Topos Theory Compendium, Oxford: Oxford University Press (2002).
  • [13] R. Kaneta and M. Yoshinaga, Magnitude homology of metric spaces and order complexes, Bulletin of the London Mathematical Society 53(3) (2021), 893–905.
  • [14] G. M. Kelly, Basic Concepts of Enriched Category Theory, London Mathematical Society Lecture Note Series 64, Cambridge University Press, Cambridge, 1982.
  • [15] T. Leinster, The magnitude of metric spaces, Documenta Mathematica 18 (2013), 857–905.
  • [16] T. Leinster, The magnitude of a graph , Mathematical Proceedings of the Cambridge Philosophical Society 166 (2019), 247–264.
  • [17] T. Leinster, The Euler characteristic of a category, Documenta Mathematica 13 (2008), 21–49.
  • [18] T. Leinster and M. Shulman, Magnitude homology of enriched categories and metric spaces, Alg. Geom. Topol. 21 (2021), 2175–2221.
  • [19] S. Mac Lane, Categories for the Working Mathematician, Graduate Texts in Mathematics 5, Springer, Berlin, 1971.
  • [20] R. Sazdanovic and V. Summers, Torsion in the magnitude homology of graphs, Journal of Homotopy and Related Structures 16(2) (2021), 275–296.
  • [21] C.A. Weibel, An introduction to homological algebra, Cambridge Studies in Advanced Mathematics, 38. Cambridge University Press, Cambridge, 1994.