1 Introduction

Whitehead’s crossed modules of groups [22] received a lot of attention because of their appearance in many different contexts; see the review articles [17,18,19] and the references in them. In many of the applications they did not appear in their original form; but in the disguise of the equivalent notion of internal groupoid in the category of groups—also called a strict 2-group—(see [6, 7, 13] for proofs of the equivalence).

Groups can be regarded as (distinguished) monoids in the cartesian monoidal category of sets. In our antecedent papers [3] and [4] we worked out the notion of crossed module of monoids in more general, not necessarily cartesian monoidal categories, relative to a chosen suitable class of spans. The main examples described by the theory are crossed modules of monoids in [16], groupoids in [5] and crossed modules of Hopf monoids in [1, 10,11,12, 15].

The aim of this article is to extend to this level of generality a third equivalent description of crossed modules of groups: as simplicial groups whose Moore complex has length 1. Recall that for any simplicial group, the Moore complex is the associated chain complex, which is given at each degree by the joint kernel of the face maps of positive label, and the restriction of the zero’th face map as the differential (this is part of the Dold-Kan correspondence). A chain complex—so in particular the Moore complex—is said to have length \(\ell \) if it is trivial for all degrees greater than \(\ell \).

We consider simplicial monoids; that is, functors from the opposite of the simplicial category \(\Delta \) to the category of monoids in some monoidal category \({\mathsf {C}}\). Assuming the existence of certain relative pullbacks (cf. [3]), in Sect. 2 we associate to a simplicial monoid a sequence of morphisms which yields a chain complex whenever the monoidal unit of \({\mathsf {C}}\) is a terminal (hence zero) object in the category of monoids in \({\mathsf {C}}\). It can be seen as a generalization of the Moore complex of a simplicial group. We also give a meaning to its length, and study the consequences of its having some finite length n.

Section 3 contains some technical material about the invertibility of a certain canonical morphism playing an essential role in the theory.

The main result can be found in Sect. 4 where we prove an equivalence between the category of relative categories in the category of monoids in \({\mathsf {C}}\) (cf. [3]) and the category of relative simplicial monoids in \({\mathsf {C}}\) whose Moore length is 1. The functors establishing the equivalence carry transparent meanings. For obtaining a relative category from a simplicial monoid we truncate it at the first two degrees. In the opposite direction, a simplicial monoid is obtained from a relative category of monoids as the (relative) simplicial nerve.

In the particular monoidal category of spans over a given set, we obtain an equivalence between the categories of certain double categories, and of certain simplicial categories. These equivalent categories contain, as equivalent full subcategories, the category of 2-groupoids on one hand, and the category of crossed modules of groupoids on the other hand.

In the particular monoidal category of comonoids in some symmetric monoidal category \({\mathsf {M}}\), we obtain an equivalence between the categories of certain category objects in the category of bimonoids in \({\mathsf {M}}\), and of certain simplicial bimonoids. These equivalent categories contain, as equivalent full subcategories, the category of internal categories in the category of cocommutative Hopf monoids in \({\mathsf {M}}\) on one hand, and the category of simplicial cocommutative Hopf monoids of Moore length 1 on the other hand. This includes, in particular, the equivalence in [8] between the category objects in the category of cocommutative Hopf algebras over a field, and the category of simplicial cocommutative Hopf algebras whose Moore complex has length 1.

Throughout, we freely use definitions, notation and results from [3, 4]. The composition of some morphisms and in an arbitrary category will be denoted by . Identity morphisms will be denoted by 1 (without any reference to the (co)domain object if it causes no confusion). In any monoidal category \({\mathsf {C}}\) the monoidal product will be denoted by juxtaposition and the monoidal unit will be I. For the monoidal product of n copies of the same object A also the power notation \(A^n\) will be used. For any monoid A in \({\mathsf {C}}\), the multiplication and the unit morphisms will be denoted by and , respectively. If \({\mathsf {C}}\) is also braided, then for the braiding the symbol c will be used.

2 The Moore Length

Recall that a simplicial object in an arbitrary category \({\mathsf {C}}\) is a functor S from the opposite of the simplicial indexing category to \({\mathsf {C}}\). Explicitly, this means objects and morphisms of \({\mathsf {C}}\) in

(1)

subject to the following simplicial relations.

$$\begin{aligned} \begin{array}{ll} \partial _i \cdot \partial _j = \partial _{j-1}\cdot \partial _i &{} \text { if } i<j\\ {}&{} {}\\ \sigma _j\cdot \sigma _i =\sigma _i \cdot \sigma _{j-1} &{} \text { if } i<j \end{array} \qquad \partial _i \cdot \sigma _j = \left\{ \begin{array}{cl} \sigma _{j-1} \cdot \partial _i &{} \text { if } i<j\\ 1 &{} \text { if } i\in \{j,j+1\} \\ \sigma _{j} \cdot \partial _{i-1} &{} \text { if } i>j+1 \end{array} \right. \end{aligned}$$

In the category of groups (and in more general semi-abelian categories [14]), a chain complex—the so called Moore complex—can be associated to any simplicial object. The full subcategory of those simplicial objects whose Moore complex has length 1, turns out to be equivalent to the category of crossed modules and to the category of internal categories in the semi-abelian category in question, see e.g. [19] (in the case of the category of groups) and [9, 21] (more generally).

However, as in [3] and [4], here we work in categories (of monoids in some monoidal category \({\mathsf {C}}\)) where the existence of zero objects is not assumed. So the notion of chain complex is not available. We shall see below, however, that when certain relative pullbacks exist, we can still associate to a simplicial object S a sequence of composable morphisms (reproducing the Moore complex whenever the monoidal unit of \({\mathsf {C}}\) is terminal in the category of monoids in \({\mathsf {C}}\)). Although the chain condition can not be formulated at this level of generality, there is a natural way to define the length of this sequence that we call the Moore length of the simplicial monoid S.

The construction of this sequence of composable morphisms is based on Assumption 2.1 below.

Recall that a class \({\mathcal {S}}\) of spans in an arbitrary category is said to be admissible if it satisfies the following two properties in ([3], Definition 2.1).

  1. (POST)

    If then too, for any morphisms and .

  2. (PRE)

    If then , for any morphism .

The relative pullback of a cospan with respect to such a class \({\mathcal {S}}\) was introduced in ([3], Definition 3.1) as a span in \({\mathcal {S}}\) satisfying the following properties.

  • Commutativity of the diagram

  • Universality. For any such that \(f\cdot f'=g\cdot g'\), there is a unique morphism which satisfies \(p_A \cdot h =f'\) and \(p_C \cdot h =g'\).

  • Reflection. If both

    belong to \({\mathcal {S}}\) then also belongs to \({\mathcal {S}}\); and symmetrically, if

    belong to \({\mathcal {S}}\) then also belongs to \({\mathcal {S}}\).

A class \({{\mathcal {S}}}\) of spans in a monoidal category is said to be monoidal if it satisfies the following two conditions in ([3], Definition 2.5).

  1. (UNITAL)

    For any morphisms f and g whose domain is the monoidal unit I, .

  2. (MULTIPLICATIVE)

    If both and then also .

Assumption 2.1

Let \({\mathcal {S}}\) be a monoidal admissible class of spans—in the sense of ([3], Definitions 2.1 and 2.5)—in some monoidal category \({\mathsf {C}}\). For any simplicial object S in the category of monoids in \({\mathsf {C}}\) let us use the notation of (1) and consider the following successive assumptions.

  1. (1)

    Assume that for any positive integer n there exists the \({\mathcal {S}}\)-relative pullback—in the sense of ([3], Definition 3.1)—\(S_n^{(1)}\) in

    Since \(\partial _k\) is compatible with the units of the monoids \(S_n\) and \(S_{n-1}\); and by the simplicial relation \(\partial _n\,\cdot \,\partial _k=\partial _k\,\cdot \,\partial _{n+1}\) for any \(0\le k \le n\), we may apply ([3], Proposition 3.5 (1)) to conclude on the existence of the unique morphism rendering commutative the above diagram.

  2. (2)

    In addition to the assumption in item (1) above, assume that for all \(n\ge 2\) there exists the \({\mathcal {S}}\)-relative pullback \(S_{n}^{(2)}\) in

    Since \(\partial _k^{(1)}\) is compatible with the units of the monoids \(S_n^{(1)}\) and \(S_{n-1}^{(1)}\), and by the simplicial relation \(\partial _{n-1}\,\cdot \,\partial _k=\partial _k\,\cdot \,\partial _n\) for any \(0\le k \le n-1\), we may apply ([3], Proposition 3.5 (1)) to conclude on the existence of the unique morphism rendering commutative the above diagram.

    \(\vdots \)

  3. (l)

    In addition to the assumptions in all items (1)\(\cdots \) (l-1) above, assume that for all \(n\ge l\) there exists the \({\mathcal {S}}\)-relative pullback \(S_n^{(l)}\) in

    Since \(\partial _k^{(l-1)}\) is compatible with the units of the monoids \(S_n^{(l-1)}\) and \(S_{n-1}^{(l-1)}\); and by the simplicial relation \(\partial _{n-(l-1)}\,\cdot \,\partial _k=\partial _k\,\cdot \,\partial _{n+1-(l-1)}\) for any \(0\le k \le n-(l-1)\), we may apply ([3], Proposition 3.5 (1)) to conclude on the existence of the unique morphism rendering commutative the above diagram.

    \(\vdots \)

Example 2.2

For this example note that in any monoidal category \({\mathsf {M}}\) the monoidal unit I carries a trivial monoid structure which is initial in the category of monoids in \({\mathsf {M}}\). Symmetrically, I carries a trivial comonoid structure which is terminal in the category of comonoids in \({\mathsf {M}}\). Whenever \({\mathsf {M}}\) is braided, the trivial monoid and comonoid structures of I combine to a bimonoid which is thus the zero object in the category of bimonoids in \({\mathsf {M}}\).

If moreover \({\mathsf {M}}\) has equalizers which are preserved by taking the monoidal product with any object, then the category of bimonoids in \({\mathsf {M}}\) has equalizers—see ([3], Example 3.3)—and thus kernels. The kernel of any bimonoid morphism is computed as the equalizer in \({\mathsf {M}}\) of

(2)

and , see [2].

So let \({\mathsf {M}}\) be a symmetric monoidal category in which equalizers exist and are preserved by taking the monoidal product with any object. Let \({\mathsf {C}}\) be the monoidal category of comonoids in \({\mathsf {M}}\) and consider the monoidal admissible class \({\mathcal {S}}\) of spans in \({\mathsf {C}}\) from ([3], Example 2.3). Since \(\varepsilon \) is the counit, the diagram

commutes for any bimonoid morphism g proving that

(3)

From Example 3.3 and Proposition 3.7 in [3] we know that all \({\mathcal {S}}\)-relative pullbacks exist in the category of bimonoids in \({\mathsf {M}}\); hence any simplicial bimonoid in \({\mathsf {M}}\)—that is, any functor from \(\Delta ^{\mathsf {op}}\) to the category of monoids in \({\mathsf {C}}\)—satisfies the successive assumptions of Assumption 2.1 for any positive integer. Still—say, for an easier comparison with [8]—below we present a more explicit description of the objects \(S_n^{(k)}\).

For any positive integer n and any \(0< k \le n\) the desired objects \(S_n^{(k)}\) are constructed as the joint kernels of the morphisms \(\{\partial _n,\partial _{n-1},\cdots , \partial _{n-k+1}\}\) in the category of bimonoids in \({\mathsf {M}}\); that is, as the joint equalizers

in \({\mathsf {M}}\) (where the “hat notation” of (2) is used). By construction they are bimonoids. Using the universality of the equalizer (in \({\mathsf {M}}\)) in the bottom rows, for \(k>1\) we construct bimonoid morphisms in

(note the serial commutativity of the right-hand diagram thanks to the simplicial identities) and show that they give rise to the \({\mathcal {S}}\)-relative pullback

(4)

From (3) we infer . The square of (4) commutes since \(j_{n-1}^{(k-1)}\) in the right vertical side of the commutative diagram

is a monomorphism. In order to check the universality of (4), take a bimonoid morphism such that the exterior of the left-hand diagram of

(5)

commutes (we know that by (3)). Then a filler \(\widetilde{g}\) of the left-hand diagram of (5) is constructed using the universality of the equalizer in the bottom row of the right-hand diagram of (5). The occurring morphism \(j_n^{(k-1)}\,\cdot \,g\) renders commutative the diagrams

for \(n-k+1<i\le n\). Thus since it is a comonoid morphism, it equalizes the parallel morphisms of the right-hand diagram of (5). The so constructed morphism \(\widetilde{g}\) renders commutative the left-hand diagram of (5) since the right column and the bottom row of the left-hand commutative diagram in

(6)

are equal monomorphisms. Finally \(\tilde{g}\) is the unique filler of the left-hand diagram of (5); as if also h makes the left-hand diagram of (5) commute then also the right-hand diagram of (6) commutes. Since \(j_n^{(k)}\) is a monomorphism, this proves \(h=\widetilde{g}\). In order to see that the span (4) satisfies the reflection property of ([3], Definition 3.1) on the right, take bimonoid morphisms such that ; equivalently, the large square on the left of the diagram

commutes. Since is a monomorphism, this is equivalent to the commutativity of the exterior diagram. Since also is a monomorphism, this is further equivalent to \(hg\,\cdot \,c\,\cdot \,\delta =hg\,\cdot \,\delta \); that is, to . Reflectivity on the left follows symmetrically.

Proposition 2.3

Let \({\mathcal {S}}\) be a monoidal admissible class of spans in some monoidal category \({\mathsf {C}}\) and let S be a simplicial monoid in \({\mathsf {C}}\) which satisfies the successive conditions in Assumption 2.1 for any positive integer. For any positive integer n, the morphisms (where \(S_n^{(0)}:=S_n\) and \(\partial _0^{(0)}:=\partial _0\)) render commutative the diagram

In particular, whenever the monoidal unit I is a zero object in the category of monoids in \({\mathsf {C}}\), there is a chain complex

(7)

Proof

The morphisms \(D_n\) are clearly well-defined and they render commutative

\(\square \)

Example 2.4

The definition of cartesian monoidal category includes the fact that the monoidal unit is a terminal object, which is therefore a zero object in the category of monoids. Thus Proposition 2.3 implies the well-known fact that any simplicial monoid in a cartesian monoidal category admits a Moore complex (7).

Example 2.5

If \({\mathsf {M}}\) is a symmetric monoidal category in which equalizers exist and are preserved by taking the monoidal product with any object, then we know from Example 2.2 that Assumption 2.1 holds for any simplicial bimonoid in \({\mathsf {M}}\) (that is, for any functor from \(\Delta ^{\mathsf {op}}\) to the category of monoids in the category of comonoids in \({\mathsf {M}}\)). Since in the category of bimonoids in \({\mathsf {M}}\) the monoidal unit is the zero object, we conclude by Proposition 2.3 that any simplicial bimonoid in \({\mathsf {M}}\) admits a Moore complex (7).

Definition 2.6

Let \({\mathcal {S}}\) be a monoidal admissible class of spans in some monoidal category \({\mathsf {C}}\). We say that a simplicial monoid S in \({\mathsf {C}}\) has Moore length l if the successive conditions in Assumption 2.1 hold for any positive integer and and are mutually inverse isomorphisms for all \(n>l\).

Lemma 2.7

Let \({\mathcal {S}}\) be a monoidal admissible class of spans in some monoidal category \({\mathsf {C}}\). For a simplicial monoid S in \({\mathsf {C}}\) of Moore length l, there are mutually inverse isomorphisms and for any non-negative integer i and any \(n>i+l\).

Proof

We proceed by induction on i.

For \(i=0\) and \(n>l\) the claim holds by definition.

Assume that it holds for some fixed value of i and all \(n>l+i\).

If \(n>l+i+1\) then in the \({\mathcal {S}}\)-relative pullback in the inner square of

(8)

the right column is an isomorphism by the induction hypothesis. We claim that so is then the left vertical of the inner square. By the monoidality of \({\mathcal {S}}\), , hence by ([3], Lemma 3.4 (2)) also . The exterior of (8) commutes by the commutativity of

where the bottom triangle commutes by the induction hypothesis. Then by the universality of the \({\mathcal {S}}\)-relative pullback in (8) there is a unique morphism \(p_{S_n^{(n-i-1)}}^{-1}\) in (8). It is the right inverse of \(p_{S_n^{(n-i-1)}}\) by construction and also the left inverse since are jointly monic and the following diagrams commute

Using that also is an isomorphism by the induction hypothesis, so is the composite morphism , and this completes the proof. \(\square \)

Lemma 2.8

Let \({\mathcal {S}}\) be a monoidal admissible class of spans in some monoidal category \({\mathsf {C}}\) and let S be a simplicial monoid in \({\mathsf {C}}\) for which the successive conditions of Assumption 2.1 hold for any positive integer. If there is some non-negative integer k for which and are mutually inverse isomorphisms for all \(n \ge k\), then also and are mutually inverse isomorphisms for all \(n > k\).

Proof

We need to show that under the standing assumptions the inner square of

is an \({\mathcal {S}}\)-relative pullback for all \(n > k\). Commutativity of this square is immediate by the unitality of the morphism in the bottom row and by the unitality of \({\mathcal {S}}\). For the universality observe that since the verticals are isomorphisms by assumption, the exterior of our diagram commutes if and only if g is equal to ; if and only if f is equal to . Hence the (obviously unique) filler is g. The top row of the inner square is the identity morphism. Therefore the reflection property in ([3], Definition 3.1) of relative pullback trivially holds. \(\square \)

Corollary 2.9

Let \({\mathcal {S}}\) be a monoidal admissible class of spans in some monoidal category \({\mathsf {C}}\). For a simplicial monoid S in \({\mathsf {C}}\) which satisfies the successive conditions of Assumption 2.1 for any positive integer, the following assertions are equivalent.

  1. (i)

    S has Moore length l.

  2. (ii)

    and are mutually inverse isomorphisms for all \(n >l\).

Proof

Statement (i) implies (ii) by Lemma 2.7, and (ii) implies (i) by Lemma 2.8. \(\square \)

3 Invertibility of Some Canonical Morphisms

As in [([4], Theorems 1.1, 2.1 and 3.10), also in the forthcoming equivalence between relative categories and certain simplicial monoids a crucial role is played by the invertibility of some canonical morphisms discussed in this section.

Consider a monoidal admissible class of spans in an arbitrary monoidal category \({\mathsf {C}}\). Take a simplicial monoid S as in (1) in \({\mathsf {C}}\) for which the successive conditions in Assumption 2.1 hold for any positive integer. For any non-negative integer n, for \(0\le i \le n\) and for \(0\le k \le n-i\), we define a morphism iteratively as follows.

  • \(\sigma _i^{(0)}:=\sigma _i\).

  • For \(0<k \le n-i\) we define \(\sigma _i^{(k)}\) as the unique filler in

    It is well-defined by the simplicial identity \(\partial _{(n+1)-(k-1)}\,\cdot \,\sigma _i=\sigma _i\,\cdot \,\partial _{n-(k-1)}\) and the unitality of \(\sigma _i^{(k-1)}\), see ([3], Proposition 3.5).

With these morphisms \(\sigma _i^{(k)}\) at hand, we introduce for any positive integer n and \(0\le k < n\) the morphisms

(9)

Note that they are natural in the following sense. For a simplicial monoid morphism , let us define inductively the morphisms \(f_n^{(0)}:=f_n\) and for \(0<k \le n\) the unique morphism \(f_n^{(k)}\) which renders commutative

(This definition makes sense—see ([3], Proposition 3.5)—since \(\{f_n\}_{n\ge 0}\) is a morphism of simplicial monoids by assumption, hence so is \(\{f_n^{(k-1)}\}_{n\ge 0}\) and therefore

commutes.) These morphisms and \(y_{(n,k)}\) fit into the commutative diagram

(10)

Lemma 3.1

Consider a monoidal admissible class of spans in an arbitrary monoidal category \({\mathsf {C}}\). Take a simplicial monoid S as in (1) in \({\mathsf {C}}\) for which the successive conditions in Assumption 2.1 hold for any positive integer. If for some positive integer n and some \(0\le k < n-1\) the morphism \(y_{(n,k)}\) of (9) is invertible then also \(y_{(n-1,k)}\) is invertible.

Proof

It follows by the commutativity of both diagrams

(11)

for any \(0\le i < n-k-1\) together with the simplicial identity \(\partial _i\,\cdot \,\sigma _i=1\) that \(y_{(n-1,k)}\) has the inverse

\(\square \)

Note that in addition to the diagrams of (11), \(y_{(n,k)}\) renders commutative also

(12)

Proposition 3.2

Let \({\mathcal {S}}\) be a monoidal admissible class of spans in some monoidal category \({\mathsf {C}}\). For a simplicial monoid S in \({\mathsf {C}}\) of Moore length l, the following assertions hold.

(1):

The morphism \(y_{(n,k)}\) of (9) is invertible for all \(l<k<n\).

(2):

For any \(n>l\) the following are equivalent.

(i):

\(y_{(n,l)}\) of (9) is invertible.

(ii):

is invertible.

(iii):

The morphisms are equal isomorphisms. Their inverse is equal to \(\partial _{0}^{(l)}=\partial _{1}^{(l)}=\cdots =\partial _{n-l}^{(l)}\).

Proof

  1. (1)

    In the unitality diagram

    the vertical arrows are isomorphisms by Lemma 2.7. Then the bottom row is an isomorphism too.

  2. (2)

    The equivalence of (i) and (ii) follows by the commutativity of the diagram

    whose left vertical arrow is an isomorphism by Lemma 2.7.

Assertion (iii) trivially implies (ii). Conversely, if \(\sigma _{n-1-l}^{(l)}\) is invertible then by the simplicial relations its inverse is \(\partial _{n-1-l}^{(l)}=\partial _{n-l}^{(l)}\) which is then invertible too. Again by the simplicial relations the inverse of \(\partial _{n-1-l}^{(l)}=\partial _{n-l}^{(l)}\) is \(\sigma _{n-1-l}^{(l)}=\sigma _{n-2-l}^{(l)}\). Iterating this reasoning we conclude that (iii) holds. \(\square \)

Example 3.3

For a simplicial category S (that is, a functor S from \(\Delta ^{\mathsf {op}}\) to the category of monoids in the category of spans over a fixed set in [4], Example 1.2]), \(S_n^{(k+1)}\) is the subcategory of those morphisms in \(S_n^{(k)}\) which are taken by the functor \(\partial _{n-k}^{(k)}\) to an identity morphism. Hence for a simplicial groupoid S, the category \(S_n^{(k)}\) is a groupoid for all \(0\le k \le n\). In this case all morphisms \(\{y_{(n,k)}\}_{0\le k<n}\) of (9) are invertible by the same argument applied to the morphism (1.3) in ([4], Example 1.2).

Example 3.4

Let \({\mathsf {M}}\) be a symmetric monoidal category in which equalizers exist and are preserved by taking the monoidal product with any object. Let \({\mathsf {C}}\) be the monoidal category of comonoids in \({\mathsf {M}}\) and let \({\mathcal {S}}\) be the monoidal admissible class of spans in \({\mathsf {C}}\) from ([3], Example 2.3). Take a simplicial monoid S in \({\mathsf {C}}\) (that is, a simplicial bimonoid in \({\mathsf {M}}\)) such that for all non-negative integers n, \(S_n\) is a cocommutative Hopf monoid in \({\mathsf {M}}\). Then \(y_{(n,k)}\) of (9) is invertible for all \(0\le k <n\). This can be seen as follows.

Recall from Example 2.2 that \(S_n^{(k)}\) is now a joint kernel in the category of bimonoids in \({\mathsf {M}}\), hence it is a sub bimonoid of the cocommutative bimonoid \(S_n\). Thus \(S_n^{(k)}\) is a cocommutative bimonoid. Moreover, by the cocommutativity of \(S_n\), its antipode is a comonoid morphism for all \(n\ge 0\). So we can use the universality of the equalizer in \({\mathsf {C}}\) in the bottom row of

(13)

to define the antipode \(z_n^{(k)}\) for all \(n>0\) and \(0\le k\le n\) as the restriction of \(z_n\). The diagram of (13) is serially commutative since any bimonoid morphism between Hopf monoids is compatible with the antipodes, thus so are in particular the parallel morphisms of the rows. This makes \(S_n^{(k)}\) a cocommutative Hopf monoid for all \(0\le k \le n\). By construction the morphisms and are morphisms of bimonoids and therefore of Hopf monoids.

Summarizing, is a split epimorphism of bimonoids in \({\mathsf {M}}\) for which conditions (1.a) and (1.b) of ([4], Proposition 1.5) hold; hence by ([4], Proposition 1.5) (see also [20]) the corresponding morphism \(y_{(n,k)}\) in ([4], Theorem 1.1 (1.b)) is invertible.

The above assumption about the cocommutativity of each \(S_n\) may look quite strong. Note however, that for the application of ([4], Proposition 1.5) we need the assumption that belongs to the class \({\mathcal {S}}\). By ([3], Lemma 2.4)Footnote 1 this is equivalent to the cocommutativity of \(S_{n-1}^{(k)}\). This should hold for all \(n>0\) and \(0\le k <n\); so in particular for \(k=0\).

4 Equivalence of Relative Categories and Simplicial Monoids of Moore Length 1

Theorem 4.1

Consider a monoidal admissible class \({\mathcal {S}}\) of spans in a monoidal category \({\mathsf {C}}\) such that there exist the \({\mathcal {S}}\)-relative pullbacks of those cospans in \({\mathsf {C}}\) whose legs are in \({\mathcal {S}}\) (cf. ([3], Assumption 4.1])). The equivalent categories \(\mathsf {CatMon}_{{\mathcal {S}}}({\mathsf {C}})\) and \(\mathsf {Xmod}_{{\mathcal {S}}}({\mathsf {C}})\) of ([4], Theorem 3.10) are equivalent also to the category \(\mathsf {Simp}^{1}\mathsf {Mon}_{{\mathcal {S}}}({\mathsf {C}})\) whose objects are simplicial monoids in \({\mathsf {C}}\) (that, is functors S from \(\Delta ^{\mathsf {op}}\) to the category of monoids in \({\mathsf {C}}\)) such that the following conditions hold.

(a):

S has Moore length 1.

(b):

Using the notation from (1), and belong to \({\mathcal {S}}\).

(c):

The morphisms \(y_{(n,k)}\) of (9) and the morphisms

are invertible for all \(n> 0\) and \(0\le k < n\).

morphisms are simplicial monoid morphisms (that is, natural transformations between the functors from \(\Delta ^{\mathsf {op}}\) to the category of monoids in \({\mathsf {C}}\)).

Proof

The proof consists of the construction of mutually inverse equivalence functors between \(\mathsf {Simp}^1\mathsf {Mon}_{{\mathcal {S}}}({\mathsf {C}})\) and \(\mathsf {CatMon}_{{\mathcal {S}}}({\mathsf {C}})\).

The functor \(\mathsf {Simp}^1\mathsf {Mon}_{{\mathcal {S}}}({\mathsf {C}}) \rightarrow \mathsf {CatMon}_{{\mathcal {S}}}({\mathsf {C}})\) sends an object S in (1) to the \({\mathcal {S}}\)-relative category whose underlying reflexive graph is . By construction this is an object of the category \(\mathsf {ReflGraphMon}_{{\mathcal {S}}}({\mathsf {C}})\) in ([4], Theorem 2.1) for which the morphisms \(q_n\) of ([4], (3.3)) are invertible. By ([4], Proposition 3.8) it extends uniquely to an object of \(\mathsf {CatMon}_{{\mathcal {S}}}({\mathsf {C}})\) since the following diagram commutes.

Recall from Proposition 3.2 that and are mutually inverse isomorphisms. Hence the regions of the above diagram sharing the dashed arrow commute because both \(q_2\,\cdot \,\partial _0^{(1)}1\) and \(y_{(2,0)}\) are multiplicative by [3], Lemma 1.5], with respect to the multiplications induced by the respective distributive laws

whose equality follows by the commutativity of the diagrams

A morphism in \(\mathsf {Simp}^1\mathsf {Mon}_{{\mathcal {S}}}({\mathsf {C}})\) is sent to the morphism of reflexive graphs ; it is an \({\mathcal {S}}\)-relative functor by ([4], Proposition 3.9).

In the opposite direction \(\mathsf {CatMon}_{{\mathcal {S}}}({\mathsf {C}}) \rightarrow \mathsf {Simp}^1\mathsf {Mon}_{{\mathcal {S}}}({\mathsf {C}})\), we send an \({\mathcal {S}}\)-relative category to its ‘\({\mathcal {S}}\)-relative nerve’

where for any positive integer n we put

By the functoriality of —cf. ([3], Proposition 3.5)—they constitute a simplicial monoid which obeys property (b) by construction and for which the morphisms \(q_n\) of part (c) are invertible. In order to see that it has Moore length 1, note first that exists; see ([4], Theorem 1.1). We claim that also for any \(n>0\) there is an \({\mathcal {S}}\)-relative pullback

determining . By construction hence by (POST) also . If some morphisms f and g render commutative the left-hand diagram of

then they make commute the right-hand diagram as well by the commutativity of

Whenever also by (POST). Hence by the universality of the \({\mathcal {S}}\)-relative pullback in the right-hand diagram, it has a unique filler h. But the same morphism h is a filler also for the left-hand diagram by the commutativity of both diagrams

The unique filler of the right-hand diagram is the unique filler for the left-hand diagram since any filler h of the left-hand diagram is clearly a filler for the right-hand diagram as well. For the reflection property let us use again that is a monomorphism split by \(p_1\). Then if then by (POST) also . So, whenever also , it follows from the reflection property of the relative pullback span (in the sense of ([3], Definition 3.1)) that . The reflection property on the left is proven symmetrically.

With this we proved that exists for any \(n>0\) and it is isomorphic to . The morphism \(\partial _{n-1}^{(1)}\) is the identity morphism, being defined as the unique morphism fitting the left-hand commutative diagram of

Since by the unitality of \({\mathcal {S}}\), the right-hand diagram is obviously an \({\mathcal {S}}\)-relative pullback. This proves that for any \(n>1\)

exists and it is isomorphic to I. Then by Corollary 2.9S has Moore length 1. Above we proved that is invertible; then so is its inverse \(\sigma _{n-2}^{(1)}\). Therefore by Proposition 3.2 the morphism \(y_{(n,k)}\) of (9) is invertible for all \(0<k<n\). For any \(n>0\) the morphism \(y_{(n,0)}\) takes now the form

in which we recognize the invertible morphism \(q_n\). This proves that the nerve of an object of \(\mathsf {CatMon}_{{\mathcal {S}}}({\mathsf {C}})\) is indeed an object of \(\mathsf {Simp}^1\mathsf {Mon}_{{\mathcal {S}}}({\mathsf {C}})\). A morphism

is sent to the family which is clearly a morphism of simplicial monoids.

It remains to see that the above constructed functors are mutually inverse equivalences. Sending an \({\mathcal {S}}\)-relative category (and \({\mathcal {S}}\)-relative functor, respectively) to its nerve and then truncating it as above, we clearly re-obtain the \({\mathcal {S}}\)-relative category (and \({\mathcal {S}}\)-relative functor, respectively) that we started with. Composing the functors in the opposite order, an object S of \(\mathsf {Simp}^1\mathsf {Mon}_{{\mathcal {S}}}({\mathsf {C}})\) is sent to

(14)

where (see ([4], Proposition 3.8)) and for any positive integer n,

Note that together with the family of morphisms \(\{q_n\}_{n> 0}\) they render commutative the following diagrams. For all \(0\le i <n\),

(15)

For \(1< i <n\),

(16)

and for \(0\le i \le 1<n\) the analogous one

(17)

We claim that a natural isomorphism from S to its image in (14) can be constructed iteratively for all \(n\ge 0\) as

(18)

Note that it gives \(w_1=1\), but non-trivial higher components. Let us prove by induction on n the equality

(19)

For \(n=1\) both sides yield the identity morphism thus the equality holds. If it holds for some \(n>0\) then the following diagram commutes

proving (19) for all positive n. (The region marked by (IH) commutes by the induction hypothesis.)

The morphisms \(w_n\) of (18) are composites of isomorphisms (see Proposition 3.2) hence they are invertible. They are compatible with the units of the domain and codomain monoids by the unitality of the constituent monoid morphisms. Multiplicativity is checked by induction on n again. Trivially, \(w_0=1\) is multiplicative. If \(w_{n-1}\) is multiplicative for some \(n>0\) then by ([3], Corollary 1.7) so are both isomorphisms

(20)

with respect to the multiplications induced by the respective distributive laws

Their equality follows by the commutativity of the diagrams of Fig. 1, whose right vertical arrows are jointly monomorphic.

Fig. 1
figure 1

Multiplicativity of \(w_n\)

Since \(w_n\) is the composite of the right-hand morphism of (20) with the inverse of the left-hand one, we conclude that it is multiplicative too.

Next we check by induction on n that \(\{w_n \}_{n\ge 0}\) is a simplicial morphism; that is,

(21)

commute for all \(n\ge 0\) and \(0\le i \le n\) and \(0\le j \le n+1\). Note that the induction must be started with \(n=1\) because the left-hand diagram of (24) below only makes sense for \(n>0\). For \(n=0\) the diagrams of (21) commute because \(w_0\) and \(w_1\) are the identity morphisms and the equalities and hold for \(i=0,1\) by construction. For \(n=1\) commutativity of the diagrams of (21) is checked as follows.

Now assume that the left-hand diagram of (21) commutes for some \(n>0\) and all \(0\le i \le n\). By the commutativity of the diagrams

(where the region marked by (IH) commutes by the induction hypothesis) we conclude that the left-hand diagram of (21) commutes for all \(n>0\) and all \(0\le i \le n\).

Assume next that the right-hand diagram of (21) commutes for some \(n>0\) and all \(0\le j \le n+1\). Then the following diagrams commute for all \(0\le j \le n\).

(22)
(23)

The missing case \(j=n+1\) follows by the commutativity of the following diagrams whose vertical arrows are jointly monomorphic.

(24)

This proves that the right-hand diagram of (21) commutes for all stated values of n and j and thus w is a simplicial morphism.

Finally, naturality of w is proven by induction on n. For any simplicial monoid morphism and \(n=1\) the left-hand diagram of

evidently commutes. If the left-hand diagram commutes for some non-negative integer n then so does the right-hand one too. \(\square \)

The functor \(\mathsf {Simp}^1\mathsf {Mon}_{{\mathcal {S}}}({\mathsf {C}}) \rightarrow \mathsf {CatMon}_{{\mathcal {S}}}({\mathsf {C}})\) in the above proof can be composed with the functor \(\mathsf {CatMon}_{{\mathcal {S}}}({\mathsf {C}}) \rightarrow \mathsf {Xmod}_{{\mathcal {S}}}({\mathsf {C}})\) in the proof of ([4], Theorem 3.10). The resulting equivalence functor sends an object S to

where \(D_0\) is the same morphism in Proposition 2.3.

Example 4.2

As in [4], Example 1.2, take the (evidently admissible and monoidal) class of all spans in the category \({\mathsf {C}}\) of spans over a given set X. The equivalent categories of [4], Example 3.11 are equivalent also to the following category.

\(\mathsf {Simp}^{1}\mathsf {Mon}({\mathsf {C}})\) whose

objects are simplicial categories S such that the object set of \(S_n\) for each \(n\ge 0\) is the given set X and the following conditions hold.

  1. (a)

    The Moore complex of S has length 1.

  2. (c)

    The morphisms \(q_n\) and \(y_{(n,k)}\) of Theorem 4.1 are invertible for all \(0\le k<n\) (equivalently, \(q_1\), \(y_{(n,0)}\) and \(y_{(n,1)}\) are invertible for all \(0<n\)).

(There is no condition (b) because we are working with the class of all spans.) morphisms are the simplicial functors.

This category contains as a full subcategory the category of simplicial groupoids of Moore length 1; which is therefore equivalent to the category of internal categories in the category of groupoids; and also to the category of crossed modules of groupoids, see [4], Example 3.11].

Example 4.3

Let \({\mathsf {M}}\) be a symmetric monoidal category in which equalizers exist and are preserved by taking the monoidal product with any object. Take \({\mathsf {C}}\) to be the category of comonoids in \({\mathsf {M}}\) with the monoidal admissible class \({\mathcal {S}}\) in ([3], Example 2.3) of spans in \({\mathsf {C}}\). The equivalent categories of ([4], Example 3.12) are also equivalent to the category

\(\mathsf {Simp}^{1}\mathsf {Mon}_{{\mathcal {S}}}({\mathsf {C}})\) whose

objects are simplicial bimonoids S in \({\mathsf {M}}\) such that

  1. (a)

    For all \(n>0\) and for for \(0<i\le n\),

    (25)

    is a joint equalizer in \({\mathsf {M}}\) (that is; the Moore complex of S has length 1, see Examples 2.2 and 2.5).

  2. (b)

    \(\partial _0 1\,\cdot \,\delta =\partial _0 1\,\cdot \,c\,\cdot \,\delta \) and \(\partial _1 1\,\cdot \,\delta =\partial _1 1\,\cdot \,c\,\cdot \,\delta \).

  3. (c)

    The morphisms \(q_n\) and \(y_{(n,k)}\) of Theorem 4.1 are invertible for \(0\le k<n\) (equivalently, \(q_1\), \(y_{(n,0)}\) and \(y_{(n,1)}\) are invertible for all \(0<n\)).

morphisms are the simplicial bimonoid morphisms.

As a simple consequence we obtain the following result in [8].

Proposition 4.4

Let \({\mathsf {M}}\) be a symmetric monoidal category in which equalizers exist and are preserved by taking the monoidal product with any object. Take \({\mathsf {C}}\) to be the category of comonoids in \({\mathsf {M}}\) with the monoidal admissible class \({\mathcal {S}}\) in ([3], Example 2.3) of spans in \({\mathsf {C}}\). The equivalent categories in ([4], Example 3.12) and in Example 4.3 have equivalent full subcategories as follows.

  • The full subcategory of \(\mathsf {CatMon}_{{\mathcal {S}}}({\mathsf {C}})\) for whose objects both A and B are cocommutative Hopf monoids in \({\mathsf {M}}\).

  • The full subcategory of \(\mathsf {Xmod}_{{\mathcal {S}}}({\mathsf {C}})\) for whose objects both Y and B are cocommutative Hopf monoids in \({\mathsf {M}}\).

  • The full subcategory of \(\mathsf {ReflGraphMon}_{{\mathcal {S}}}({\mathsf {C}})\) for whose objects the following conditions hold.

    • A and B are cocommutative Hopf monoids (with antipodes z)

    • for the morphisms

      the following diagram commutes.

      (26)
  • The full subcategory of \(\mathsf {Simp}^1\mathsf {Mon}_{{\mathcal {S}}}({\mathsf {C}})\) for whose objects S the following conditions hold.

    • \(S_n\) is a cocommutative Hopf monoid in \({\mathsf {M}}\) for all \(n\ge 0\).

    • The Moore complex of S has length 1; that is, (25) is a joint equalizer in \({\mathsf {M}}\) for all \(n>1\).

Proof

We need to show that the category listed last is a subcategory of the category in Example 4.3. Condition (b) of Example 4.3 becomes trivial thanks to the cocommutativity assumption. Concerning condition (c), the morphisms \(q_n\) are invertible by ([4], Proposition 3.13) and the morphisms \(y_{(n,k)}\) are invertible by Example 3.4. \(\square \)