Magnitude and magnitude homology of filtered set enriched categories
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 -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 and a “size function” , where 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)
-
(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” , where is the category of collectable filtered sets (Definition 3.2), and is the Novikov series ring (Definition 3.7). Then we can define the magnitude of finite -categories and the magnitude (co)weighting of -categories of finite type (Definition 3.13), both of which contain finite categories and finite metric spaces. Further, for a special class of -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)
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)
magnitude of finite metric spaces in the original sense
-
(2)
magnitude (co)weighting of locally finite graphs with infinitely many vertices
-
(3)
Euler characteristic of finite categories
-
(4)
Euler characteristic of finite simplicial complexes
-
(5)
Poincaré polynomial of ranked posets
-
(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 , we have an isomorphism
for any . Hence we have .
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 -filtered set enriched categories, we construct a spectral sequence that satisfying the following.
Theorem 1.2 (Proposition 4.34).
Let be a -filtered set enriched category. Then
-
(1)
.
-
(2)
If is a digraph, we have , where denotes the reduced path homology introduced by Grigor’yan–Muranov–Lin–S. -T. Yau et al. [8].
-
(3)
It converges to the homology of the underlying small category if it converges, in particular when is finite.
Then we discuss homotopy invariance of each page of this spectral sequence. We show that the -th page of this spectral sequence is invariant under “-homotopy” described as follows.
-
(1)
-homotopy contains equivalences of small categories.
-
(2)
-homotopy for digraphs is exactly the digraph homotopy.
-
(3)
-homotopy for metric spaces is exactly the -closeness of Lipschitz maps.
Theorem 1.3 (Theorem 4.37).
The -page of the above spectral sequence is invariant under -homotopy.
For an example, we have the following.
Theorem 1.4 (Proposition 4.20).
For any generalized metric space , we denote its Kolmogorov quotient by . Then we have . Namely, the magnitude homology is invariant under Kolmogorov quotient.
The above results suggest that there should be “-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 -filtered chain complexes, where weak equivalences are exactly the morphisms inducing quasi isomorphisms on . 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 .
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 be a “fibration” with the fibers and the base admitting the magnitude. Then we have .
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 , a category enriched over , or -category is the following data :
-
•
a set ,
-
•
“morphisms” for any ,
-
•
“a composition rule” for any ,
-
•
“an identity morphism” for any ,
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 -categories, by . A small category is finite if it has finitely many objects and morphisms.
2.2 Graphs
-
(1)
A graph is a -dimensional simplicial complex, that is a pair of sets such that . A path metric on a graph is a function defined by , or if there are no such sequences. A graph homomorphism is a map such that for any . Equivalently, that is a map such that either or for any . We denote the category of graphs and graph homomorphisms by Grph.
-
(2)
A directed graph or a digraph is a pair of sets such that , where denotes the diagonal. A directed path metric on a digraph is a function defined by , or if there are no such sequences. A digraph homomorphism is a map such that for any . Equivalently, that is a map such that either or for any . We denote the category of digraphs and digraph homomorphisms by DGrph.
-
(3)
We have a functor defined by and , namely by orienting edges in bi-directions. This is a full and faithful functor, and we identify any graphs with their image under .
2.3 Metric spaces
-
(1)
A generalized metric space is a set with a map satisfying
-
•
for any ,
-
•
for any .
We sometimes denote a generalized metric space simply by . For generalized metric spaces and , a -Lipschitz map is a map satisfying that for any . We denote the category of generalized metric spaces and 1-Lipschitz maps by .
-
•
-
(2)
A generalized metric space is symmetric if it satisfies that for any .
-
(3)
A generalized metric space is non-degenerate if implies that for any .
-
(4)
A generalized metric space is a metric space if it is symmetric, non-degenerate and satisfies that . We denote the full subcategory of that consists of metric spaces and 1-Lipschitz maps by . A metric space is finite if the set is finite.
-
(5)
For a generalized metric space and a point , we define that and .
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)
A -filtered set, or simply a filtered set is a set with subsets for any satisfying that for any and . We formally define that for .
-
(2)
Let be a filtered set, and let . We define that if .
-
(3)
For filtered sets and , a filtered map is a map satisfying that for any .
-
(4)
For filtered sets and , their product is defined by . Then we have for any .
-
(5)
For a family of filtered sets indexed by a set , we define its union by . We sometimes use a notation for filtered sets and ’s, where we consider as a set by forgetting the filtration.
We use the following terminologies.
Definition 3.2.
-
(1)
A filtered set is collectable if is a finite set for each .
-
(2)
We denote a one element filtered set with by . We formally define that .
We denote the category of filtered sets and maps by . The following is straightforward.
Proposition 3.3.
is a symmetric monoidal category with the unit object .
Let be the small category with objects and morphisms We equip with a symmetric monoidal structure by with the unit object . We denote the category of sets with obvious symmetric monoidal structure by . The following is also straightforward.
Proposition 3.4.
There is a symmetric monoidal embedding sending a set to a filtered set with for any . There is also a symmetric monoidal embedding sending to .
We denote the category of -categories by . Note that it follows from 2.1 that for any object of a -category. We also note that the category of -categories is exactly . By Proposition 3.4, we have the following.
Proposition 3.5.
There are embeddings and .
In the following, we denote the full subcategory of that consists of collectable filtered sets by . Note that is again a symmetric monoidal category. We also denote the full subcategory of that consists of -categories by .
Remark 3.6.
While we define a filtered set by an “ascending filtration”, we can define it in the descending manner as follows.
-
(1)
A -filtered set is a set with subsets for any satisfying that , for any and .
-
(2)
Let be a filtered set, and let . We define that if .
-
(3)
For filtered sets and , a filtered map is a map satisfying that for any .
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 .
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 -adic integers.
Definition 3.7.
We say a function is left finite if the support of restricted to is a finite set for any . We give a “power series like” expression for a left finite function . The universal Novikov series field is defined by pointwise addition and the product . Let be its subring that consists of left finite functions , called a Novikov series ring.
Note that is a non-Archimedean complete valuation field and is its valuation ring. In particular, we have , where . In the following, we denote and by and respectively for readability. To simplify discussions, we introduce the notion of normed groups and rings. See [2], for example.
Definition 3.8.
-
(1)
A pair of an abelian group and a function is a normed group if it satisfies that
-
•
if and only if .
-
•
for any .
We say that is complete if it is complete with respect to the distance function .
-
•
-
(2)
A pair of a (possibly non-commutative) ring and a function is a normed ring if it satisfies that
-
•
is a normed group with respect to the additive structure.
-
•
for any .
We say that is complete or a Banach ring if it is complete as a normed group.
-
•
-
(3)
Let be a normed group and be a normed ring. We say that is a normed (left, right, bi-) -module if is a (left, right, bi-) -module and satisfies that for any and .
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)
For , we define , which makes a complete normed ring.
-
(2)
Let be a set. We equip sets of maps and with norms defined by and respectively. Then and are complete normed groups.
Definition 3.10.
Let be a set.
-
(1)
For every , we denote the set of functions such that
by . We equip with a ring structure by pointwise addition and the product defined by . Note that we have a projection for any .
-
(2)
We define a ring by . Namely, we have . Here we denote the natural projection by .
We consider as a subspace of . Note that, for any , the set is a fundamental system of neighborhoods for . Note also that the set is a fundamental system of neighborhoods for any . Here we denote by .
Lemma 3.11.
The subspace of is closed. Hence is a complete normed ring, where the norm is induced from .
Proof.
Let . Then there exist an and an such that for infinitely many or for infinitely many . If we take , then any element of has the same property. Hence we have , which implies that is closed. This completes the proof. ∎
The following is straightforward.
Lemma 3.12.
The normed abelian groups and are -bimodule. Here, the left and the right action of on is defined by and for any and .
When we take as the set of objects for some category , we abbreviate as for the notations 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 be a symmetric monoidal category, and be a category enriched over .
-
(1)
A map is called a size function if it satisfies the following :
-
(a)
It holds that for any .
-
(b)
It holds that and for any .
-
(a)
-
(2)
We say that is of -finite type (or simply of finite type) if the function defined by is in .
-
(3)
We say that is finite if it has finitely many objects.
-
(4)
Suppose that is of finite type. A function is a weighting if it satisfies that , where for any . Similarly, a function is a coweighting if it satisfies that .
-
(5)
Suppose that is finite. If has both a weighting and a coweighting, we define the magnitude of by . 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 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 be a -category of of finite type. If is invertible in , then it has a unique weighting and a coweighting, and it holds that . Further, if is finite, it holds that .
3.4 Magnitude of -categories
3.4.1 A size function on
Now we define a size function on in a sense of Definition 3.13. Then we can consider the magnitude of finite -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 , we define a size function by .
In the following, we just say a -category 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 -category of finite type. For a morphism , we denote its domain and codomain by and respectively.
Lemma 3.16.
A -category is of finite type if and only if it satisfies one of the following equivalent conditions :
-
(1)
For any object and , there are finitely many morphisms with satisfying that or .
-
(2)
For any objects , both of the filtered sets and are collectable.
3.4.2 Tame categories and an explicit formula for magnitude
Next we define a large class of -category of finite type, namely tame categories, whose magnitude is computable. For , we write if .
Definition 3.17.
Let be a -category, and let .
-
(1)
A non-degenerate -path on from to is a sequence of morphisms such that and for any . We define the length of such a sequence by .
-
(2)
A non-degenerate path on is a non-degenerate -path from to for some and .
-
(3)
We say that is quasi-tame if it is of finite type, and there are finitely many non-degenerate paths from to with length for any and .
-
(4)
We say that is tame if it is of finite type, and there exists an for any such that any non-degenerate -path have length .
Lemma 3.18.
Let be a -category of finite type, and let .
-
(1)
The number of non-degenerate -paths from to with length is the coefficient of in .
-
(2)
is quasi-tame if and only if converges in for any .
-
(3)
is tame if and only if converges in .
Proof.
(1) is obvious. In general, converges if and only if converges to for any or since these two metric spaces are complete. Hence (2) and (3) follows. ∎
Lemma 3.19.
A -category of finite type is quasi-tame if it is tame. Further, the converse is true if 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)
Any finite metric space is a finite tame category.
-
(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)
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 be a -category of finite type. If there exists an such that any non-identity morphism satisfies that , then is tame.
Proof.
The existence of such an guarantees that . Hence it follows from Lemma 3.18 (3). This completes the proof. ∎
We call a -category satisfying the condition of Lemma 3.21 uniform. The following gives us an explicit formula of for tame categories.
Proposition 3.22.
Let be a tame category. Then we have
for any .
Proof.
Note that the RHS is equal to , which converges by Lemma 3.18. Since is a topological ring, we have
This completes the proof. ∎
Corollary 3.23.
Let be a tame category . Then we have a weighting and a coweighting
and
for any . Further, if is finite, it holds that
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 . Now we have
Note that the function is in . If not, there should be an object and such that there are infinitely objects and non-degenerate paths from to ’s or from ’s to with length . It implies that there are infinitely many morphisms with or with , which contradicts the assumption that 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 -categories. Since hom-objects of those consist of one element filtered sets , they are also -categories. We have the following characterizations.
Lemma 3.25.
-
(1)
A generalized metric space is of finite type as a -category if and only if the closed balls and are finite sets for any and .
-
(2)
A generalized metric space of finite type is quasi-tame if and only if it is non-degenerate.
-
(3)
A generalized metric space of finite type is tame if there exists an such that for any .
Proof.
-
(1)
It is obvious from Lemma 3.16.
-
(2)
If is not non-degenerate, namely there exists points with and , then is not quasi-tame by the definition. If is not quasi-tame, conversely, there is an infinite family of non-degenerate paths with the length for some and . Since is of finite type and the set is a subset of , it is finite. Hence there should be an infinite sequence with and . It implies that there is a pair of points with and , hence is not non-degenerate.
-
(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 -category if and only if it is a finite metric space. Hence we obtain the following.
Corollary 3.26.
-
(1)
A generalized metric space is a finite -category if and only if it is a finite generalized metric space.
-
(2)
A generalized metric space is a finite tame category if and only if it is a finite non-degenerate generalized metric space.
-
(3)
A metric space is a finite tame category if and only if it is a finite metric space.
3.5.2 Euler characteristic of finite categories
As stated in Proposition 3.5, any small categories are -categories. We note that a small category is a -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 -category. Hence we have the following.
Lemma 3.27.
A small category is a finite -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 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 -category in another way. For a small category and , we consider as an element of a filtered set by setting if is not an identity, and . It is easy to see that is a -category in this setting. When is a preordered set, it is same as considering as a digraph whose directed edges correspond to the relation . 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 -category by setting the degree of all relations by except for identities which have degree . In this setting, any finite poset is a finite non-degenerate generalized metric space, hence is a finite tame category by Lemma 3.26. Therefore Corollary 3.23 implies that
where we denote the order complex of by . In particular, the magnitude of a finite poset in this setting is a polynomial. It implies that is equal to the Euler characteristic of . When is a face poset of a finite simplicial complex , we obtain that . 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 be a finitely generated group with generators . We consider as a -category by and . Here we define the degree of 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 , which is exactly the inverse of the growth series of . On the other hand, let be the Cayley graph of , and we consider it as a metric space by the path metric. Then is a tame category by Lemma 3.25, and we have a weighting . 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 with the minimum element is a ranked poset if it is equipped with a rank function satisfying
Here, we say that covers if .
Example 3.30.
Let be a vector space over an arbitrary field. A finite collection of affine subspaces of is called subspace arrangement. For a subspace arrangement , let , where we set . We equip with a poset structure by defining if . Then has the minimum element , and we can define a rank function by with , which makes it a ranked poset.
We metrize a finite ranked poset by Note that this is equivalent to considering the directed Hasse diagram of as a generalized metric space, where a directed edge is spanned from to if covers . 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 , we have
and
Now we recall the Möbius function of a poset.
Definition 3.32.
For a finite poset , we define a square matrix by Then exists, which we call the Möbius function of .
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 follows by substituting to the polynomial magnitude of it. Furthermore, we immediately obtain the following by Proposition 3.22.
Proposition 3.33.
We have
Note that we have . The following definition is fundamental and pivotal in the study of subspace arrangements.
Definition 3.34.
The Poincaré polynomial of a ranked poset with the minimum element is defined by
Example 3.35.
Let be a subspace arrangement. Then the Poincaré polynomial of is defined as that of .
The following shows that the Poincaré polynomial of a ranked poset is essentially the weighting of at .
Proposition 3.36.
Remark 3.37.
The above coincidence of the Poincaré polynomial and the weighting can be generalized as follows. Let be a finite poset with the minimum element , equipped with an order preserving map satisfying that . For such a , we define by which makes a non-degenerate finite generalized metric space, hence a finite tame category. The Poincaré polynomial of such a is defined as . We can prove that coincides with the weighting of by the same argument of Proposition 3.36.
Example 3.38.
-
(1)
For a ranked poset , we can choose to adapt to the above situation. Then the definitions above coincide with the original ones.
-
(2)
Let be a subspace arrangement. The power set is naturally equipped with a poset structure by inclusion, that is, if . It has the minimum element , and we can define an order preserving map by . Then the Poincaré polynomial of is the following :
It is natural to ask how and 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 -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 -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)
An -filtered abelian group, or simply a filtered abelian group is a filtered set equipped with an abelian group structure such that for any . 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 . We define a functor by freely generating filtered abelian groups. We also define functors by and . Note here that and are also filtered sets for any filtered set and .
-
(2)
A filtered chain complex is a collection of filtered abelian groups ’s and filtered homomorphisms ’s such that for any . We suppose that chain complexes are non-negative, that is for . 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 .
-
(3)
A filtered simplicial set is an object of the functor category . A filtered simplicial abelian group is an object of the functor category . We have functors induced from respectively.
-
(4)
We define a functor by and .
The following is a refinement of the nerve functors for small categories in the setting of filtered sets enrichment.
Definition 4.2.
Let . Let be its underlying small category structure.
-
(1)
We define a filtered nerve functor by
The face and the degeneracy maps are defined similarly to the usual nerve . We denote simplicial abelian groups and by and respectively.
-
(2)
We denote the composition by . Explicitly, we have the following for :
-
•
,
-
•
,
-
•
as filtered sets.
We also denote chain complexes and by and respectively. Note that and are subchain complexes of .
-
•
Now we define the un-normalized magnitude chain complex as follows.
Definition 4.3.
For and , we define the un-normalized magnitude chain complex of as the quotient chain complex . Explicitly, we define a chain complex by
-
•
,
-
•
-
•
.
4.1.2 Magnitude chain complex and the categorification property
We define the magnitude chain complex as the normalization of . To that end, we recall the following fundamental fact. See Section III-2 of [7], for example.
Lemma 4.4.
Let be a simplicial abelian group. Let be the associated chain complex and be the subchain complex of degenerated simplices. Then the quotient is a homotopy equivalence.
Since is a subsimplicial abelian group of , its quotient
is again a simplicial abelian group. The chain complex of degenerated simplices of this quotient is a subchain complex of described as
Then the quotient is a homotopy equivalence by Lemma 4.4. We define described as
-
•
,
-
•
-
•
.
We define the magnitude homology as the homology of the above homotopy equivalent chain complexes.
Definition 4.5.
We call the above chain complex the magnitude chain complex of . We denote its homology by .
Note that and define functors from .
Definition 4.6.
Let be a -category and . We say that is a -object, or simply , if there is no morphism such that and . We also say that is if every object is .
If are -objects, then we have subsimplicial abelian groups and of , where
and
We define chain complexes , and . Further, we define the normalizations of these and described as
and
The quotient induces quotients and , which are all homotopy equivalences by Lemma 4.4. We denote the homology of these equivalent chain complexes by and respectively. When is , we have decompositions of chain complexes
where . Hence they induce decompositions
We note that is spanned by non-degenerate paths on with length . Hence we have the following.
Lemma 4.7.
Let be a -category of finite type.
-
(1)
If is tame, then the chain complex is bounded for any .
-
(2)
If is quasi-tame and , then the chain complex is bounded for any and .
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 be a tame category. When is , then it holds that
and
Proposition 4.9.
Let be a finite tame category. Then it holds that
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 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 , which are straightforward generalization of those in .
Definition 4.12.
Let be -categories. We define a tensor product by and .
Let be the poset considered as a -category via the inclusions . Namely, has two objects and one non-trivial morphism with degree .
Definition 4.13.
Let . A natural transformation is a functor such that and .
Note that the underlying functor is a natural transformation in . For natural transformations and , we denote their “horizontal” and “vertical” compositions by and respectively whenever they are defined. We also denote the identity natural transformation of a functor by .
Definition 4.14.
Let . We say that is left adjoint to , or equivalently is right adjoint to , denoted by , if there are natural transformations and , called the unit and the counit, satisfying the following commutative diagrams :
We remark that if and only if there is a natural isomorphism in .
Example 4.15.
Let . Then if and only if for any and .
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 . If there is a natural transformation , then and are chain homotopic. Further, if and are , then and are chain homotopic.
Proof.
We construct a filtration preserving chain homotopy . 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 , we define by
where is the arrow . Then we can check that . Since , it does not decrease the degree. Hence the above homotopy induces a homotopy between and , which are chain maps from to . Further, if and are , then it is easily checked that this homotopy is restricted to . Hence the latter assertion follows. This completes the proof. ∎
The following is immediate from Proposition 4.16.
Corollary 4.17.
Let . If , then they induce isomorphisms between and . Further, if and are , and , then the above isomorphisms are restricted to isomorphisms between and .
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 , we define an equivalence relation on by if and only if . Then the quotient set is again a generalized metric space by defining , which is obviously non-degenerate. We call this generalized metric space the Kolmogorov quotient of , and denote it by .
Note that if is non-degenerate.
Lemma 4.19.
For any generalized metric space , the quotient map is a left adjoint functor in .
Proof.
It is straightforward to check that the quotient map is a -Lipschitz map, hence a functor in . Let be a section of the map . Then is obviously a -Lipschitz map, and we have
for any and . Hence we have by Example 4.15. This completes the proof. ∎
Proposition 4.20.
For any generalized metric space , we have . Namely, the magnitude homology is invariant under Kolmogorov quotient.
4.2.2 Galois connection of posets
Definition 4.21.
Let and be posets, and let and be order preserving maps. The pair is a Galois connection if in and thus in , namely it satisfies that and for all and .
Proposition 4.22.
Let and be finite posets with minimum elements and . Let and be order preserving maps satisfying that and . If a Galois connection , where and , satisfies that and , then is an adjoint pair between generalized metric spaces and defined in Remark 3.37.
Proof.
Let and . Then we have
and
Since is a Galois connection, we obtain that , hence is an adjoint pair by Example 4.15. This completes the proof. ∎
Corollary 4.23.
Under the same assumption as in Proposition 4.22, we have .
Example 4.24.
Let be a subspace arrangement. Let be posets with order preserving maps and as in Examples 3.30 and 3.38. We define poset maps and by and . Then we have and for any and . Hence the pair is a Galois connection. Further, by the definitions of and , it is straightforward to check the assumptions in Proposition 4.22. Thus we obtain that 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 -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 with a unital associative ring structure such that .
Definition 4.26.
Let be a -category. We define a filtered ring by
-
•
with as a filtered abelian group.
-
•
For any , we define an associative product by
We call the category algebra of . We also define an action of on the abelian group from the right and the left by and
Now we recall the Hochschild homology. Let be a unital associative ring, and be an -bimodule, which are not filtered. The Hochschild chain complex of and is defined by and
where is taken over . Its homology is called Hochschild homology of . If is filtered, then the filtration induces again a filtration on , hence also on , by
and
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 and with respect to the filtrations in the following.
Lemma 4.28.
Let be a -category. The homomorphisms
for , and define an injective filtered chain map .
Proof.
It is straightforward. ∎
We consider the chain complex as a subcomplex of in the following.
Lemma 4.29.
Let be a -category. The chain complexes
and
are contractible for any .
Proof.
We first prove the latter contractibility. Let
For each , let be the minimum number such that , where we put . We define a homomorphism
by
for and
Then it is straightforward to check that . Hence this defines a chain homotopy between the identity and the zero homomorphism on
It is easily seen that the homotopy extends to a homotopy
which shows the contractibility of . This completes the proof. ∎
Corollary 4.30.
Let be a -category. We have homotopy equivalences
and
for any .
Proof.
By Lemma 4.29, the inclusions
and
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 , 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 is a face poset of a finite simplicial complex as in Example 3.5.3, the algebra is known as the incidence algebra of , and Corollary 4.30 can be considered as a homological version of well-known Gerstenhaber-Schack’s theorem ([6]) that asserts . 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 by
-
•
with as a filtered abelian group.
-
•
For any we define an associative product by
We also define an action of on similarly to that of . Note that
and
are direct summands. Hence we have
and
Finally we obtain the following description of the magnitude homology as the Hochschild homology.
Corollary 4.32.
Let be a -category. The chain complexes and
are homotopy equivalent for any . In particular, we have an isomorphism
for any . Hence we have .
Proof.
By Lemma 4.29, the inclusions
and
are quasi isomorphisms. Hence the five lemma shows that we have a quasi isomorphism
Since they are level-wise free, this is a homotopy equivalence. The latter follows from the identification
and the direct sum decomposition of . This completes the proof. ∎
4.4 MH, a spectral sequence and homotopy relations
In the following, we construct a spectral sequence for a -category whose morphisms have integral degrees. It converges to the homology of underlying small category , 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 is a -filtered set if for every . We denote the full subcategory of that consists of -filtered sets by .
Note that the filtered chain complex is -filtered for a -category . We denote the spectral sequence associated with this filtered chain complex by . 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 -category , we have the following.
-
(1)
.
-
(2)
If is a digraph, we have , where denotes the reduced path homology introduced by Grigor’yan–Muranov–Lin–S. -T. Yau et al. [8].
-
(3)
It converges to the homology of the underlying small category if it converges, in particular when is finite.
Proof.
(1), (3) is straightforward. (2) is shown in Theorem 1.2 of [1]. ∎
For , let be a -category with two objects and one non-trivial morphism with degree . The following is a generalization of the definition of natural transformation in , which induces a series of homotopy relations in as it will be stated in Theorem 4.37.
Definition 4.35.
Let . An -natural transformation is a functor such that and .
Example 4.36.
-
(1)
A -natural transformation is the one defined in Definition 4.13.
-
(2)
A -natural transformation restricted to the category DGrph is exactly the -step homotopy of digraph homomorphisms defined in [8]. It is shown in the same paper that the path homology is invariant under -step homotopies.
-
(3)
Let be 1-Lipschitz maps with being metric spaces. Then the existence of an -natural transformation is equivalent to the condition that for any , which is called -close in the metric geometry.
-
(4)
An -natural transformation induces an -natural transformation via the natural morphism if .
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 -page is invariant under -natural transformations. That is, induces identical homomorphisms if there exists an -natural transformation .
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 -filtered chain complexes, where weak equivalences are exactly the morphisms inducing quasi isomorphisms on . Theorem 4.37 suggests that there should be some homotopy theory on that coincides with the above Ciricis’ structure when we apply a functor . 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 .
5 Fibrations in
In this section, we study “fibrations” in , whose restriction to is originally introduced as metric fibrations by Leinster in [15]. It contains the Grothendieck (op)fibration in . 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 . Further, we restrict our notion of fibrations to , 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 ” in the following. While there are variants of definitions of “fibrations” in , we adopt pre-opfibration here. We remark that the other notions of fibrations can be considered similarly in , and they all coincide when we restrict them to .
Definition 5.1.
Let be a small category. The following data is called a normal oplax functor.
-
(1)
is a small category for any .
-
(2)
is a functor for any .
-
(3)
For any , .
-
(4)
For any , there is a natural transformation satisfying the following. For any , we have , and the following commutative diagram :
Definition 5.2.
Let be a normal oplax functor. We construct a small category as follows.
-
(1)
,
-
(2)
,
-
(3)
For and , , where is a component of the natural transformation .
Note that we have a natural projection functor defined by and . In the following, we use the convention “” for arbitrary category if and .
Definition 5.3.
Let and . We say that is weakly -cartesian if it satisfies the following : For any with , there uniquely exists such that and .
Definition 5.4.
Let . We say that is a pre-opfibration if it satisfies the following : For any , there exists a weakly -cartesian morphism such that .
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 , the natural projection is a pre-opfibration.
Proof.
It will be shown in Proposition 5.11. ∎
Proposition 5.6.
For any pre-opfibration , we can construct a normal oplax functor such that there is an isomorphism that commutes with and . Furthermore, we have for any normal oplax functor .
Sketch of proof.
For , we define the small category by and . For and with , we fix a weakly -cartesian lift of with by the axiom of choice. We choose them so that . Note that any morphism induce a unique morphism that produces a functor . We define the normal oplax functor by and for any and . Then it is straightforward to check that is indeed a normal oplax functor. Next we show that with suitable compatibility. Note that we have and . We define the functor by and for any and . It is straightforward to check that is an isomorphism with the desired compatibility. Finally, it is easily checked that we can choose the lifts of ’s to construct so that . This completes the proof. ∎
Now we generalize the above definitions to in the following.
Definition 5.7.
Let be a -category. The following data is called a normal oplax functor.
-
(1)
is a normal oplax functor in the sense of Definition 5.1 when we forget the filtrations,
-
(2)
for any ,
-
(3)
For any and for any , .
Definition 5.8.
Let be a normal oplax functor. We construct a -category as follows.
-
(1)
When we forget the filtration, coincides with the one defined in Definition 5.2.
-
(2)
for any .
It is straightforward to check that the above is indeed a -category. Note that we have a natural projection defined by and . For a normal oplax functor , we can consider and ’s as the “base” and the “fibers” of respectively. The following definitions contain Definitions 5.3 and 5.4 by considering the inclusion of Proposition 3.5.
Definition 5.9.
Let and . We say that is weakly -cartesian if it satisfies the following : For any with , there uniquely exists such that and . Further, it satisfies .
Definition 5.10.
Let . We say that is a pre-opfibration if it satisfies the following : For any , there exists a weakly -cartesian morphism such that and .
Proposition 5.11.
Let be a normal oplax functor. Then the natural projection is a pre-opfibration.
Proof.
Let . For any , we define . It is immediate to obtain and . Now we show that is weakly -cartesian. Let and . If we have a morphism with and , then it is immediate to see that we should have . Conversely, we have . Further, we have . This completes the proof. ∎
Proposition 5.12.
For any pre-opfibration , we can construct a normal oplax functor such that there is an isomorphism that commutes with and . Furthermore, we have for any normal oplax functor .
Proof.
The construction of is same as Proposition 5.6, however, we should check that is indeed a normal oplax functor in the sense of Definition 5.7. Namely, we should check that i) is a morphism in , and ii) for any and .
-
i)
It is enough to show that preserves filtrations. Let . Since is induced by the universality of weakly -cartesian morphism , we have . Since we have , we obtain .
-
ii)
Since the morphism is induced from the universality of the weakly -cartesian morphism , we have , where we put . Hence we obtain .
Next we check that the functor defined in the proof of Proposition 5.6 is also an isomorphism in . We should check that for any . Recall that we define , where and . Since is weakly -cartesian, we have . Finally, we note that we have similarly to Proposition 5.6. This completes the proof. ∎
Corollary 5.13.
There is a one to one correspondence between normal oplax functors and pre-opfibrations given by and .
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 be a normal oplax functor. Suppose that , for any and are all -categories of finite type. If and for any have weightings , then has a weighting .
Proof.
It follows from the following calculation.
∎
Proposition 5.15.
Let be a normal oplax functor with for any . Suppose that , , are all finite -categories. If , and all have the magnitude, then .
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 is left finite if it is a support of some left finite function .
Lemma 5.16.
Let be a normal oplax functor. Suppose that and any for are -categories. Then is also a -category.
Proof.
Let for any and . Let , which is left finite since is collectable. Then we have
Since each is a finite set and each is collectable, the filtered set is collectable, whose elements have degree . Hence the left finiteness of implies that the filtered set is collectable. This completes the proof. ∎
Now we have the following two obvious corollaries.
Corollary 5.17.
Let be a normal oplax functor. Suppose that and any for are finite -categories. Then is also a finite -category.
Corollary 5.18.
Let be a normal oplax functor. Suppose that and any for are finite tame categories. Then is also a finite tame category.
Lemma 5.19.
Let be a normal oplax functor. Suppose that and any for are -categories of finite type. If any fiber of is a finite set for any , then is also a -category of finite type.
Proof.
We have
Here, since each is of finite type, 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 by . We define . Then a subset is left finite since is of finite type. Furthermore, the filtered set is collectable with lowest degree for each . Hence is collectable. On the other hand, we have
where turns out to be collectable by the assumption for fibers. Hence is collectable by the similar argument to the above. This completes the proof. ∎
Example 5.20.
Let be a finitely generated group considered as a -category as in 3.5.4. We denote the underlying set of the group by . We also denote the automorphism of by the left multiplication of by . We construct a normal oplax functor by and . Then satisfies the condition of Lemma 5.19, hence is a -category of finite type. Actually, is isomorphic to the Cayley graph , and the coincidence of the weightings of and 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 , 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 , the -category is not necessarily a metric space in general, even if and are metric spaces for any . To consider a restriction of normal oplax funtors to , we need the following notion.
Definition 5.21.
A normal oplax functor is a metric action if it satisfies the following:
-
(1)
is a metric space for any .
-
(2)
Let be the full subcategory of consisting of objects with . Then is a metric space.
-
(3)
Let and be the unique morphisms between and . Then we have and .
Proposition 5.22.
Let be a normal oplax functor. Then is a metric space if and only if is a metric action.
Proof.
It is straightforward to check that is a metric space when is a metric action. We show the converse. Suppose that is a metric space. Since can be embedded to for any by and , it should be a metric space. Further, there should be exactly one morphism between any two objects with , because has only one morphism between two objects. In particular, we have for any . Let with , , and . Denoting the distance function of by , we have
Hence we obtain that . On the other hand, we have
Hence we obtain that , and thus . This implies that the full subcategory of consisting of objects with is a metric space, and for any such and . This completes the proof. ∎
Definition 5.23.
Let be metric spaces. A -Lipschitz map is a metric fibration if it satisfies the following: For any and , there uniquely exists such that
-
(1)
,
-
(2)
for any .
The following shows that the restriction of pre-opfibrations to is exactly the metric fibrations.
Proposition 5.24.
A -Lipschitz map is a pre-opfibration if and only if it is a metric fibration.
Proof.
Suppose that is a metric fibration. For any and , there exist and satisfying (1) and (2) of Definition 5.23. By , we have . Further, is weakly -cartesian as follows. Let with , namely . Then there exists with by (2). Moreover, such a is unique since is a metric space. Conversely, suppose that is a pre-opfibration. Let and let be the unique morphism in . Then there exists a with and . Namely, we have and . If there exists another , then the condition in Definition 5.9 implies that . Hence such a is unique. Moreover, the same condition implies (2) of Definition 5.23. This completes the proof. ∎
Proposition 5.25.
For any metric fibration , the normal oplax functor is a metric action.
Proof.
We only verify (3) of Definition 5.21. Let and be unique morphisms. For any , we have . Hence we obtain that , namely . Similarly, we have . 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.
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 and be metric fibrations. We say that and are isomorphic if there is an isometry with .
Proposition 5.28.
Let and be metric spaces. Let be a metric action with for a fixed . If is a functor, namely ’s are all identities for any , then the metric fibration is isomorphic to the projection .
Proof.
Fix a point . Then any points of for any can be expressed as for some , where is the unique morphism from to . We define a -Lipschitz map by . Then it preserves distances since we have
Further, is apparently a bijection. Hence gives the desired isomorphism. This completes the proof. ∎
Example 5.29.
Let be the complete graph with vertices. Let be the non-trivial involution. We construct a metric action by for each , for two edges of and for the rest edge . Then is a graph shown in Figure 1 right. The left graph in Figure 1 is , 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].
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 be the cyclic graph with vertices, and let be a metric fibration. We label the vertices of by , and we denote by . Then there exists an isometry , and is isomorphic to a metric fibration constructed from as follows : we construct a metric action by , for any except for . For the other pair of vertices , we define by the composition of ’s along the shortest edge path connecting the vertices and . Further, is isomorphic to the projection if is even.
Proof.
Note that the metric space consists of points with the distance function defined by
Let . We define a map by for and . Now we verify that is an isometry. Let and with . When the shortest path connecting and does not contain the edge , we have
Note here that we have because the sequence is the shortest path and should satisfy (3) of Definition 5.7. When the shortest path connecting and contains the edge , we have
Note that we have similarly to the above. Hence gives the desired isomorphism. Next, we suppose that . Since we have
from (3) of Definition 5.7, we obtain that . Hence is isomorphic to the projection 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 be a circle of radius in with the metric induced from . We construct a metric action with for . For , we denote the unique morphism by . We define isometries by the (anti-)clockwise -rotation if the shortest geodesic from to is (anti-)clockwise. It is easy to check that 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.