Kerodon

$\Newextarrow{\xRightarrow}{5,5}{0x21D2}$ $\newcommand\empty{}$
$\Newextarrow{\xhookrightarrow}{10,10}{0x21AA}$

3.5.5 Higher Groupoids

Recall that a groupoid is a category $\mathcal{G}$ in which every morphism is an isomorphism (Definition 1.3.5.1). By virtue of Propositions 1.3.3.1 and 1.3.5.2, the construction $\mathcal{G} \mapsto \operatorname{N}_{\bullet }( \mathcal{G} )$ determines a fully faithful functor from the category of groupoids to the category of Kan complexes. Consequently, little information is lost by identifying $\mathcal{G}$ with $\operatorname{N}_{\bullet }( \mathcal{G} )$, and thereby viewing groupoids as special kinds of Kan complexes. In this section, we exploit this perspective to introduce a notion of $n$-groupoid for $n \geq 0$.

Definition 3.5.5.1. Let $n$ be a nonnegative integer. An $n$-groupoid is a Kan complex $X$ which satisfies the following condition: for every pair of integers $0 \leq i \leq m$ with $m > n$, the restriction map $\operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Delta ^ m, X) \rightarrow \operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Lambda ^{m}_{i}, X )$ is a bijection.

Remark 3.5.5.2. In the situation of Definition 3.5.5.1, the assumption that $X$ is a Kan complex already guarantees that the restriction map $\operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Delta ^ m, X) \rightarrow \operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Lambda ^{m}_{i}, X )$ is surjective. Consequently, $X$ is an $n$-groupoid if and only if, whenever $\sigma $ and $\tau $ are simplices of $X$ having dimension $m > n$ which satisfy $\sigma |_{ \Lambda ^{m}_{i} } = \tau |_{ \Lambda ^{m}_{i} }$ for some $0 \leq i \leq m$, we have $\sigma = \tau $.

Warning 3.5.5.4. We have now given two a priori different definitions for the notion of $2$-groupoid:

  • According to Definition 2.2.8.24, a $2$-groupoid is a $2$-category $\operatorname{\mathcal{C}}$ such that every $1$-morphism of $\operatorname{\mathcal{C}}$ is an isomorphism and every $2$-morphism of $\operatorname{\mathcal{C}}$ is an isomorphism.

  • According to Definition 3.5.5.1, a $2$-groupoid is a simplicial set $X$ satisfying certain extension conditions.

We will show later that definitions are compatible: a simplicial set $X$ is a $2$-groupoid in the sense of Definition 3.5.5.1 if and only if it is isomorphic to the Duskin nerve $\operatorname{N}_{\bullet }^{\operatorname{D}}(\operatorname{\mathcal{C}})$, where $\operatorname{\mathcal{C}}$ is a $2$-groupoid in the sense of Definition 2.2.8.24 (in this case, the $2$-category $\operatorname{\mathcal{C}}$ is uniquely determined up to non-strict isomorphism; see Theorem 2.3.4.1). See Proposition .

Remark 3.5.5.5. Let $X$ be a simplicial set and let $n \geq 0$ be an integer. Then $X$ is an $n$-groupoid if and only if every connected component of $X$ is an $n$-groupoid.

Remark 3.5.5.6. Let $\{ X_{j} \} _{j \in \operatorname{\mathcal{J}}}$ be a diagram of simplicial sets having limit $X = \varprojlim _{j \in \operatorname{\mathcal{J}}} X_ j$ and let $n$ be a nonnegative integer. If each $X_{j}$ is an $n$-groupoid and $X$ is a Kan complex, then $X$ is also an $n$-groupoid. In particular, any product of $n$-groupoids is an $n$-groupoid (see Example 1.2.5.3).

Proposition 3.5.5.7. A simplicial set $X$ is a $0$-groupoid if and only if it is discrete: that is, if and only if it is isomorphic to a constant simplicial set $\underline{S}$ for some set $S$.

Proof. By virtue of Remark 3.5.5.5, we may assume without loss of generality that $X$ is connected. Assume that $X$ is a $0$-groupoid; we wish to show that the projection map $X \rightarrow \Delta ^0$ is an isomorphism (the converse follows immediately from the definition). To prove this, it will suffice to show that for every pair of $m$-simplices $\sigma , \tau : \Delta ^ m \rightarrow X$, we have $\sigma = \tau $. Our proof proceeds by induction on $m$. If $m > 0$, then our inductive hypothesis guarantees that $\sigma |_{ \Lambda ^{m}_{0} } = \tau |_{ \Lambda ^{m}_{0} }$, so that $\sigma = \tau $ by virtue of our assumption that $X$ is a $0$-groupoid. It will therefore suffice to treat the case $m = 0$, so that $\sigma $ and $\tau $ can be identified with vertices $x,y \in X$. Since $X$ is a connected Kan complex, there exists an edge $e: x \rightarrow y$ with source $x$ and target $y$ (Proposition 1.2.5.10). Our assumption that $X$ is a $0$-groupoid then guarantees that $e = \operatorname{id}_{x}$, so that $x = y$ as desired. $\square$

Proposition 3.5.5.8. Let $X$ be a simplicial set. The following conditions are equivalent:

$(1)$

There exists a groupoid $\mathcal{G}$ and an isomorphism of simplicial sets $X \xrightarrow {\sim } \operatorname{N}_{\bullet }( \mathcal{G} )$.

$(2)$

The simplicial set $X$ is a $1$-groupoid (in the sense of Definition 3.5.5.1).

$(3)$

The simplicial set $X$ is a Kan complex, and the tautological map $X \rightarrow \operatorname{N}_{\bullet }( \pi _{\leq 1}(X) )$ is an isomorphism.

Proof. We first show that $(1)$ implies $(2)$. For every groupoid $\mathcal{G}$, Proposition 1.3.5.2 guarantees that the simplicial set $\operatorname{N}_{\bullet }( \mathcal{G} )$ is a Kan complex. To show that $\operatorname{N}_{\bullet }( \mathcal{G} )$ is a $1$-groupoid, we must prove that if $\sigma , \tau : \Delta ^{m} \rightarrow \operatorname{N}_{\bullet }( \mathcal{G} )$ are $m$-simplices for $m \geq 2$ which have the same restriction to some horn $\Lambda ^{m}_{i} \subset \Delta ^ m$, then $\sigma = \tau $. For $m > 2$, this is immediate (since $\Lambda ^{m}_{i}$ contains the $1$-skeleton of $\Delta ^{m}$). In the case $m=2$, we can identify $m$-simplices of $\operatorname{N}_{\bullet }(\mathcal{G} )$ with commutative diagrams

\[ \xymatrix@R =50pt@C=50pt{ & Y \ar [dr]^{g} & \\ X \ar [ur]^{f} \ar [rr]^{h} & & Z } \]

in the groupoid $\operatorname{\mathcal{G}}$. The desired result then follows from the observation that any two of the morphisms $f$, $g$, and $h$ determine the third.

The implication $(3) \Rightarrow (1)$ is immediate. We will complete the proof by showing that $(2)$ implies $(3)$. Assume that $X$ is a $1$-groupoid and let $\mathcal{G} = \pi _{\leq 1}(X)$ be its fundamental groupoid. We wish to show that the tautological map $u: X \rightarrow \operatorname{N}_{\bullet }( \mathcal{G} )$ is an isomorphism: that is, it is bijective on $m$-simplices for $m \geq 0$. The proof proceeds by induction on $m$. The case $m = 0$ is immediate from the definitions. For $m \geq 2$, we have a commutative diagram

\[ \xymatrix@R =50pt@C=50pt{ \operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Delta ^{m}, X ) \ar [r]^-{u \circ } \ar [d] & \operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Delta ^ m, \operatorname{N}_{\bullet }(\mathcal{G}) ) \ar [d] \\ \operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Lambda ^{m}_{0}, X) \ar [r]^-{u \circ } & \operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Lambda ^{m}_{0}, \operatorname{N}_{\bullet }( \mathcal{G} ) ) } \]

where the vertical maps are bijective (since $X$ and $\operatorname{N}_{\bullet }(\mathcal{G} )$ are $1$-groupoids) and the bottom horizontal map is bijective (by virtue of our inductive hypothesis); it follows that the upper horizontal map is bijective as well. It will therefore suffice to treat the case $m = 1$. Let $e, e': x \rightarrow y$ be edges of the simplicial set $X$ having the same source and target; we wish to show that if the homotopy classes $[e]$ and $[e']$ coincide (as morphisms in the category $\mathcal{G} = \pi _{\leq 1}(X)$), then $e = e'$. Let $\sigma $ be a $2$-simplex of $X$ which is a homotopy from $e$ to $e'$: that is, a $2$-simplex whose boundary is depicted in the diagram

\[ \xymatrix@R =50pt@C=50pt{ & y \ar [dr]^{\operatorname{id}_ y} & \\ x \ar [ur]^{e} \ar [rr]^{e'} & & y } \]

(see Definition 1.4.3.1). Let $\tau $ be the right-degenerate $2$-simplex $s^{1}_{1}(e)$. Then $\sigma $ and $\tau $ have the same restriction to the horn $\Lambda ^{2}_{1} \subset \Delta ^2$. Invoking our assumption that $X$ is a $1$-groupoid, we conclude that $\sigma = \tau $. In particular, we have $e' = d^2_{1}(\sigma ) = d^{2}_{1}( \tau ) = e$. $\square$

Proposition 3.5.5.9. Let $n$ be a nonnegative integer, let $A_{\ast }$ be a chain complex of abelian groups, and let $X = \mathrm{K}( A_{\ast } )$ denote the associated Eilenberg-MacLane space (Construction 2.5.6.3). Then $X$ is an $n$-groupoid if and only if the abelian groups $A_{m}$ vanish for $m > n$.

Proof. Fix a pair of integers $0 \leq i \leq m$ with $m > n$. Let $\sigma : \Delta ^{m} \rightarrow \Delta ^{m}$ be the identity map, which we identify with its image in the normalized chain complex $\mathrm{N}_{\ast }( \Delta ^{m}; \operatorname{\mathbf{Z}})$. Then the chain complex $\mathrm{N}_{\ast }( \Delta ^{m}; \operatorname{\mathbf{Z}})$ splits as a direct sum of $\mathrm{N}_{\ast }( \Lambda ^{m}_{i}; \operatorname{\mathbf{Z}})$ with the subcomplex $C_{\ast }$ spanned by the $\sigma $ and $\partial (\sigma )$. We therefore obtain a canonical bijection

\begin{eqnarray*} \operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Delta ^{m}, X ) & \simeq & \operatorname{Hom}_{ \operatorname{Ch}(\operatorname{\mathbf{Z}}) }( \mathrm{N}_{\ast }( \Delta ^ m; \operatorname{\mathbf{Z}}), A_{\ast } ) \\ & \simeq & \operatorname{Hom}_{ \operatorname{Ch}(\operatorname{\mathbf{Z}}) }( \mathrm{N}_{\ast }( \Lambda ^ m_{i}; \operatorname{\mathbf{Z}}), A_{\ast } ) \times \operatorname{Hom}_{ \operatorname{Ch}(\operatorname{\mathbf{Z}}) }( C_{\ast }, A_{\ast } ) \\ & \simeq & \operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Lambda ^{m}_{i}, X) \times A_{m}. \end{eqnarray*}

It follows that the restriction map $\operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Delta ^{m}, X ) \rightarrow \operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Lambda ^{m}_{i}, X)$ is a bijection if and only if the abelian group $A_ m$ vanishes. The desired result follows by allowing the integers $0 \leq i \leq m$ to vary. $\square$

Proposition 3.5.5.10. Let $n$ be a nonnegative integer and let $X$ be a Kan complex. Then:

$(1)$

If $X$ is weakly $(n-1)$-coskeletal, then it is an $n$-groupoid.

$(2)$

If $X$ is an $n$-groupoid, then it is weakly $n$-coskeletal.

Proof. We first prove $(1)$. Assume that $X$ is weakly $(n-1)$-coskeletal. Suppose we are given an integer $m > n$ and a pair of $m$-simplices $\sigma , \tau : \Delta ^{m} \rightarrow X$ which satisfy $\sigma |_{ \Lambda ^{m}_{i} } = \tau |_{ \Lambda ^{m}_{i} }$ for some $0 \leq i \leq m$; we wish to show that $\sigma = \tau $. Note that $d^{m}_{i}(\sigma )$ and $d^{m}_{i}(\tau )$ are $(m-1)$-simplices of $X$ which coincide on the boundary $\operatorname{\partial \Delta }^{m-1}$. Since $X$ is weakly $(n-1)$-coskeletal, it follows that $d^{m}_{i}(\sigma ) = d^{m}_{i}(\tau )$, so that $\sigma $ and $\tau $ coincide on the boundary $\operatorname{\partial \Delta }^{m}$. Invoking our assumption that $X$ is weakly $(n-1)$-coskeletal again, we conclude that $\sigma = \tau $.

We now prove $(2)$. Suppose that $X$ is an $n$-groupoid; we wish to show that it is weakly $n$-coskeletal. By virtue of Exercise 3.5.4.11, it will suffice to show that for every integer $m > n$, the restriction map

\[ \theta : \operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Delta ^{m}, X ) \rightarrow \operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \operatorname{\partial \Delta }^{m}, X) \]

is injective. This is clear: our assumption that $X$ is an $n$-groupoid guarantees that the composition of $\theta $ with the restriction map $\operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \operatorname{\partial \Delta }^{m}, X) \rightarrow \operatorname{Hom}_{\operatorname{Set_{\Delta }}}( \Lambda ^{m}_{0}, X)$ is a bijection. $\square$

Corollary 3.5.5.11. Let $n$ be a nonnegative integer and let $X$ be a Kan complex. If $X$ is $(n-1)$-coskeletal, then it is an $n$-groupoid. if $X$ is an $n$-groupoid, then it is $(n+1)$-coskeletal.

Proposition 3.5.5.12. Let $n \geq 0$ be an integer and let $X$ be a Kan complex which is weakly $n$-coskeletal. Then $X$ is an $n$-groupoid if and only if it satisfies the following condition:

$(\ast )$

Let $\sigma _0, \sigma _1: \Delta ^ n \rightarrow X$ be $n$-simplices which are homotopic relative to $\operatorname{\partial \Delta }^{n}$ (Definition 3.2.1.3). Then $\sigma _0 = \sigma _1$.

Proof. Assume first that $X$ is an $n$-groupoid and let $\sigma _0, \sigma _1: \Delta ^ n \rightarrow X$ be $n$-simplices of $X$ which are homotopic relative to $\operatorname{\partial \Delta }^ n$. Since $X$ is a Kan complex, there exists a homotopy $h: \Delta ^1 \times \Delta ^ n \rightarrow X$ from $\sigma _0$ to $\sigma _1$ which is constant along $\operatorname{\partial \Delta }^ n$ (Proposition 3.2.1.4). For $0 \leq i \leq n$, let $\alpha _{i}: [n+1] \rightarrow [1] \times [n]$ denote the nondecreasing function given by the formula

\[ \alpha _{i}(j) = \begin{cases} (0, j) & \text{ if } j \leq i \\ (1, j-1) & \text{ if } j > i, \end{cases} \]

and let $\tau _{i}$ denote the $(n+1)$-simplex of $X$ given by the composition

\[ \Delta ^{n+1} \xrightarrow { \alpha _{i} } \Delta ^1 \times \Delta ^ n \xrightarrow {h} X. \]

Let $\rho _{i}, \rho '_{i}: \Delta ^{n} \rightarrow X$ be the $n$-simplices of $X$ given by $\rho _{i} = d^{n+1}_{i}( \tau _ i )$ and $\rho '_{i} = d^{n+1}_{i+1}( \tau _ i )$; by construction, we have

\[ \sigma _0 = \rho '_{n} \quad \quad \rho _{n} = \rho '_{n-1} \quad \quad \cdots \quad \quad \rho _{1} = \rho '_{0} \quad \quad \rho _0 = \sigma _1. \]

We will complete the proof by showing that $\rho _{i} = \rho '_{i}$ for $0 \leq i \leq n$. Using our assumption that the homotopy $h$ is constant along the boundary $\operatorname{\partial \Delta }^{n}$, we see that the degenerate $(n+1)$-simplex $s^{n}_{i}( \rho _{i} )$ coincides with $\tau _ i$ on the horn $\Lambda ^{n+1}_{i} \subset \Delta ^{n+1}$. Invoking our assumption that $X$ is an $n$-groupoid, we conclude that $\tau _ i = s^{n}_{i}( \rho _{i} )$. Applying the face operator $d^{n+1}_{i+1}$, we obtain $\rho _{i} = \rho '_{i}$.

We now prove the converse. Assume that $X$ satisfies condition $(\ast )$; we wish to show that the Kan complex $X$ is an $n$-groupoid. Fix a pair of integers $0 \leq i \leq m$ with $m > n$ and a pair of $m$-simplices $\tau _0, \tau _1: \Delta ^{m} \rightarrow X$ which coincide on the horn $\Lambda ^{m}_{i} \subset \Delta ^{m}$; we wish to show that $\tau _0 = \tau _1$. Since $X$ is weakly $n$-coskeletal, it will suffice to prove that $\tau _0$ and $\tau _1$ coincide on the boundary $\operatorname{\partial \Delta }^{m}$: that is, to show that the $(m-1)$-simplices $\sigma _0 = d^{m}_{i}(\tau _0)$ and $\sigma _1 = d^{m}_{i}(\tau _1)$ coincide. Note that $\sigma _0$ and $\sigma _1$ have the same restriction to the boundary $\operatorname{\partial \Delta }^{m-1}$. Consequently, if $m \geq n+2$, the desired result follows from our assumption that $X$ is weakly $n$-coskeletal. We may therefore assume that $m = n+1$. By virtue of $(\ast )$, it will suffice to show that the $(m-1)$-simplices $\sigma _0$ and $\sigma _1$ are homotopic relative to $\operatorname{\partial \Delta }^{m-1}$. In fact, we will prove a stronger claim: the $m$-simplices $\tau _0$ and $\tau _1$ are homotopic relative to the horn $\Lambda ^{m}_{i} \subset \Delta ^{m}$. This follows from the observation that the restriction map $\operatorname{Fun}( \Delta ^{m}, X) \rightarrow \operatorname{Fun}( \Lambda ^{m}_{i}, X)$ is a trivial Kan fibration; see Corollary 3.1.3.6. $\square$

Proof. It follows from Corollary 3.1.3.4 that $\operatorname{Fun}(K,X)$ is a Kan complex. Since $X$ is weakly $n$-coskeletal (Proposition 3.5.5.10), it follows that $\operatorname{Fun}(K,X)$ is also weakly $n$-coskeletal (Corollary 3.5.4.13). We will complete the proof by showing that $\operatorname{Fun}(K,X)$ satisfies condition $(\ast )$ of Proposition 3.5.5.12. Suppose we are given a pair of $n$-simplices $\sigma _0, \sigma _1: \Delta ^ n \rightarrow \operatorname{Fun}(K,X)$ which are homotopic relative to $\operatorname{\partial \Delta }^{n}$; we wish to show that $\sigma _0 = \sigma _1$. Let us identify $\sigma _0$ and $\sigma _1$ with morphisms $f_0, f_1: \Delta ^ n \times K \rightarrow X$. Since $X$ is weakly $n$-coskeletal, it will suffice to show that $f_0$ and $f_1$ coincide on $m$-simplices $\tau = (\tau ', \tau '')$ of $\Delta ^ n \times K$ for $m \leq n$. If $\tau '$ factors through the boundary $\operatorname{\partial \Delta }^ n$, this follows immediately from the equality $\sigma _0|_{\operatorname{\partial \Delta }^ n} = \sigma _1|_{ \operatorname{\partial \Delta }^ n}$. We may therefore assume without loss of generality that $m = n$ and that $\tau ': \Delta ^ m \rightarrow \Delta ^ n$ is the identity map. In this case, our assumption that $\sigma _0$ and $\sigma _1$ are homotopic relative to $\operatorname{\partial \Delta }^ n$ guarantees that $f_0(\tau )$ and $f_1(\tau )$ are homotopic relative to $\operatorname{\partial \Delta }^ n$, so that $f_0( \tau ) = f_1(\tau )$ by virtue of Proposition 3.5.5.12. $\square$

Corollary 3.5.5.14. Let $n$ be a nonnegative integer and let $f: X \rightarrow Y$ be a morphism of Kan complexes. Assume that $Y$ is an $n$-groupoid and that $f$ is bijective on $m$-simplices for $m < n$. The following conditions are equivalent:

$(1)$

The morphism $f$ is a Kan fibration.

$(2)$

The morphism $f$ is surjective on $n$-simplices.

$(3)$

The morphism $f$ is $n$-connective.

Proof. We first show that $(1)$ implies $(2)$. Here we may assume that $n > 0$ (otherwise, the result is a special case of Proposition 3.5.1.22). Let $\overline{\tau }$ be an $n$-simplex of $Y$, and set $\overline{\tau }_0 = \overline{\tau }|_{ \Lambda ^{n}_{0} }$. Since $f$ is bijective on $m$-simplices for $m < n$, we can lift $\overline{\tau }_0$ to a morphism $\tau _0: \Lambda ^{n}_{0} \rightarrow X$. If $f$ is a Kan fibration, then the lifting problem

\[ \xymatrix@R =50pt@C=50pt{ \Lambda ^{n}_{0} \ar [r]^-{\tau _0} \ar [d] & X \ar [d]^{f} \\ \Delta ^{n} \ar@ {-->}[ur]^{\tau } \ar [r]^-{\overline{\tau }} & Y } \]

admits a solution, given by an $n$-simplex $\tau $ of $X$ satisfying $f( \tau ) = \overline{\tau }$.

The implication $(2) \Rightarrow (3)$ is a special case of Corollary 3.5.2.2. We will complete the proof by showing that $(3)$ implies $(1)$. Assume that $f$ is $n$-connective and fix a pair of integers $0 \leq i \leq m$ with $m > 0$; we wish to show that every lifting problem

\[ \xymatrix@R =50pt@C=50pt{ \Lambda ^{m}_{i} \ar [r]^-{ \sigma _0 } \ar [d] & X \ar [d]^{f} \\ \Delta ^{m} \ar@ {-->}[ur]^{\sigma } \ar [r]^-{ \overline{\sigma } } & Y } \]

admits a solution. We consider three cases:

  • Suppose that $m < n$. In this case, our assumption that $f$ is bijective on $m$-simplices guarantees that there is a unique $m$-simplex $\sigma $ of $X$ satisfying $f(\sigma ) = \overline{\sigma }$. By construction, we have $(f \circ \sigma )|_{ \Lambda ^{m}_{i} } = \overline{\sigma }|_{ \Lambda ^{m}_{i} } = f \circ \sigma _0$. Since $f$ is bijective on simplices of dimension $< m$, it follows that $\sigma _0 = \sigma |_{ \Lambda ^{m}_{i} }$.

  • Suppose that $m > n$. In this case, our assumption that $X$ is a Kan complex guarantees that we can extend $\sigma _0$ to an $m$-simplex $\sigma $ of $X$. By construction, we have

    \[ (f \circ \sigma )|_{ \Lambda ^{m}_{i} } = f \circ \sigma _0 = \overline{\sigma }|_{ \Lambda ^{m}_{i} }. \]

    Since $Y$ is an $n$-groupoid, it follows that $f \circ \sigma = \overline{\sigma }$.

  • Suppose that $m = n$. Since $f$ is bijective on $(n-1)$-simplices, the morphism $\sigma _0$ admits a unique extension $\sigma _1: \operatorname{\partial \Delta }^{n} \rightarrow X$ satisfying $f \circ \sigma _1 = \overline{\sigma }|_{ \operatorname{\partial \Delta }^{n} }$. The morphism $f$ factors as a composition

    \[ X \xrightarrow {i} X \times _{ \operatorname{Fun}( \{ 0\} , Y) } \operatorname{Fun}( \Delta ^1, Y ) \xrightarrow {q} Y, \]

    where $i$ is a homotopy equivalence and $q$ is a Kan fibration (see Example 3.1.7.10). Since $f$ is $n$-connective, the Kan fibration $q$ is also $n$-connective (Proposition 3.5.1.26). Applying Proposition 3.5.2.1, we conclude that there is an $n$-simplex $\sigma $ of $X$ satisfying $\sigma _1 = \sigma |_{ \operatorname{\partial \Delta }^{n} }$ and a homotopy from $f(\sigma )$ to $\overline{\sigma }$ which is constant when restricted to $\operatorname{\partial \Delta }^{n}$. Since $Y$ is an $n$-groupoid, Proposition 3.5.5.12 guarantees that $f(\sigma ) = \overline{\sigma }$.

$\square$

Corollary 3.5.5.15. Let $n$ be a nonnegative integer and let $f: X \rightarrow Y$ be a homotopy equivalence of $n$-groupoids. If $f$ is bijective on $m$-simplices for $m < n$, then $f$ is an isomorphism.

Proof. It follows from Corollary 3.5.5.14 that $f$ is a Kan fibration. Applying Proposition 3.2.7.2, we deduce that $f$ is a trivial Kan fibration. In particular, $f$ admits a section $g: Y \hookrightarrow X$. To complete the proof, it will suffice to show that $g$ is an epimorphism of simplicial sets. This follows from Corollary 3.5.5.14, since $g$ is also bijective on $m$-simplices for $m < n$. $\square$

Proposition 3.5.5.16. Let $X$ be a Kan complex and let $n \geq 0$ be an integer. The following conditions are equivalent:

$(1)$

The Kan complex $X$ is isomorphic to an Eilenberg-MacLane space $\mathrm{K}(G,n)$. Here $G$ is a set if $n=0$, a group if $n =1$, and an abelian group if $n \geq 2$ (see Construction 2.5.6.9).

$(2)$

The Kan complex $X$ is an $n$-groupoid having a single $m$-simplex for each $m < n$.

Proof. If $n = 0$, the desired result follows from Proposition 3.5.5.7. If $n = 1$, then $X$ is an $n$-groupoid if and only if it is isomorphic to the nerve $\operatorname{N}_{\bullet }(\operatorname{\mathcal{C}})$, where $\operatorname{\mathcal{C}}$ is a groupoid (Proposition 3.5.5.8). In this case, the assumption that $X$ has a single vertex is equivalent to the requirement that the category $\operatorname{\mathcal{C}}$ contains a single object $C$, in which case we can identify $\operatorname{N}_{\bullet }(\operatorname{\mathcal{C}})$ with the classifying simplicial set $\mathrm{K}(G,1) =B_{\bullet }G$ for $G = \operatorname{Aut}_{\operatorname{\mathcal{C}}}(C)$. We may therefore assume that $n \geq 2$. The implication $(1) \Rightarrow (2)$ follows from Proposition 3.5.5.9. For the converse, assume that $X$ is an $n$-groupoid having a single $m$-simplex for each $m < n$. Let $x$ be the unique vertex of $X$ and set $G = \pi _{n}(X,x)$. For every $n$-simplex $\sigma $ of $X$, the restriction $\sigma |_{ \operatorname{\partial \Delta }^{n} }$ is the constant map taking the value $x$, so the homotopy class $[\sigma ]$ can be regarded as an element of the group $G$. Our assumption that $X$ is an $n$-groupoid guarantees that the assignment $\sigma \mapsto [\sigma ]$ determines a bijection from the collection of $n$-simplices of $X$ to the group $G$, which determines an isomorphism $f_0$ from the $n$-skeleton of $X$ to the $n$-skeleton of $\mathrm{K}(G,n)$. Invoking Theorem 3.2.2.10, we see that a morphism $\tau _0: \operatorname{\partial \Delta }^{n+1} \rightarrow X$ can be extended to an $(n+1)$-simplex of $X$ if and only if the composite map

\[ \operatorname{\partial \Delta }^{n+1} \xrightarrow {\tau _0} \operatorname{sk}_{n}(X) \xrightarrow {f_0} \mathrm{K}(G,n) \]

can be extended to an $(n+1)$-simplex of $\mathrm{K}(G,n)$. Since $\mathrm{K}(G,n)$ is weakly $n$-coskeletal, it follows that $f_0$ extends uniquely to a morphism of simplicial sets $f: X \rightarrow \mathrm{K}(G,n)$ (Proposition 3.5.4.12) which is surjective on $(n+1)$-simplices. In particular, $f$ exhibits $\mathrm{K}(G,n)$ as a weak $n$-coskeleton of $X$ (Definition 3.5.4.14). Since $X$ is weakly $n$-coskeletal (Proposition 3.5.5.10), we conclude that $f$ is an isomorphism. $\square$