This section and the next one primarily present known material from topology. In several cases we need to adapt results from the literature to our needs, which is sometimes best done by re-proving them. Readers may want to skim these two sections quickly and return to them later when needed.
Here we review two basic ways of building topological spaces from simple pieces: cell complexes and simplicial sets. Cell complexes, also known as CW complexes, are fairly standard in topology, and we will use them for a simple description of the various spaces in our proofs. Simplicial sets are perhaps less well known, and for us, they will mainly be a convenient device for converting cell complexes into simplicial complexes. Moreover, they are of crucial importance in the algorithmic results mentioned in the introduction. For a thorough discussion of simplicial complexes, simplicial sets, cell complexes, and the connections between the three, we refer to [7].
Cell Complexes
In the case of cell complexes, the building blocks are topological disks of various dimensions, called cells, which can be thought of as being completely “flexible” and which can be glued together in an almost arbitrary continuous fashion. Essentially the only condition is that each n-dimensional cell must be attached along its boundary to the (n−1)-skeleton of the space, i.e., to the part that has already been built, inductively, from lower dimensional cells. The formal definition is as follows.
We recall that if X and Y are topological spaces and if f:A→Y is a map defined on a subspace A⊆X, then the space X∪
f
Y obtained by attaching
X
to
Y
via
f is defined as the quotient of the disjoint union X⊔Y under the equivalence relation generated by the identifications a∼f(a), a∈A.
A closed or open
n-cell is a space homeomorphic to the closed n-dimensional unit disk D
n in n-dimensional Euclidean space or its interior \(\mathring{D}^{n}\), respectively; a point is regarded as both a closed and an open 0-cell.
An m-dimensional cell complex
Footnote 11
X is the last term of an inductively constructed sequence of spaces X
(0)⊆X
(1)⊆X
(2)⊆⋯⊆X
(m)=X, called the skeletons of X:
-
1.
X
(0) is a discrete set of points (possibly infinite) that are regarded as 0-cells.
-
2.
Inductively, the n-skeleton X
(n) is formed by attaching closed n-cells \(D^{n}_{i}\) (where i ranges over some arbitrary index set) to X
(n−1) via attaching maps
\(\varphi_{i}\colon S^{n-1}_{i}=\partial D^{n} \to X^{(n-1)}\). Formally, we can consider all attaching maps together as defining a map φ=⊔
i
φ
i
from the disjoint union \(\bigsqcup_{i} S_{i}^{n-1}\) to X
(n−1) and form \(X^{(n)}= (\bigsqcup_{i}D_{i}^{n} ) \cup_{\varphi} X^{(n-1)}\).
For every closed cell \(D_{i}^{n}\), one has a characteristic map
Footnote 12
\(\varPhi_{i}\colon D_{i}^{n} \to X^{(n)}\subseteq X\), which restricts to an embedding on the interior \(\mathring{D}_{i}^{n}\). The image \(\varPhi _{i}(\mathring{D}_{i}^{n})\) is commonly denoted by \(e_{i}^{n}\), and it follows from the construction that every point of X is contained in a unique open cell (note that these are generally not open subsets of X, however).
As a basic example, the n-sphere is a cell complex with one n-cell and one 0-cell, obtained by attaching D
n to a point e
0 via the constant map that maps all of S
n−1 to e
0.
Subcomplexes
A subcomplex
A⊆X is a subspace that is closed and a union of open cells of X. In particular, for each cell in A, the image of its attachment map is contained in A, so A is itself a cell complex (and its cell complex topology agrees with the subspace topology inherited from X).
The Homotopy Extension Property
An important fact is that cell complexes have the homotopy extension property: Suppose that X is a cell complex and that A⊆X is a subcomplex. If we are given a map f
0:A→Y into some space Y, an extension \(\bar{f}_{0}\colon X\to Y\) of f
0 and a homotopy H:A×[0,1] between f
0 and some other map f
1:A→Y, then H can be extended to a homotopy \(\bar{H}\colon X\times[0,1]\to Y\) between \(\bar{f}_{0}\) and some extension \(\bar{f}_{1}\colon X\to Y\) of f
1. The following is an immediate consequence.
Corollary 3.1
For a cell complex
X, subcomplex
A⊆X, and a space
Y, the extendability of a map
f:A→Y
to
X
depends only on the homotopy class of
f
in [A,Y]. Moreover, the map
f:A→Y
has an extension
\(\bar{f}\colon X\to Y\)
iff there exists a map
g:X→Y
such that the diagram
commutes up to homotopy, i.e., gi∼f.
Cellular Maps and Cellular Approximation
A map f:X→Y between cell complexes is called cellular if it maps skeletons to skeletons, i.e., f(X
(n))⊆Y
(n) for every n.
The cellular approximation theorem (see [9, Theorem 4.8]) states that every continuous map f:X→Y between cell complexes is homotopic to a cellular one; moreover, if the given map f is already cellular on some subcomplex A⊆X, then the homotopy can be taken to be stationary on A (i.e., the image of every point in A remains fixed throughout).
Simplicial Sets
For certain constructions it is advantageous to use a special type of cell complex with an additional structure that allows for a purely combinatorial description; the latter also facilitates representing and manipulating the objects in question, simplicial sets, on a computer. We refer to [8] for a very friendly and thorough introduction to simplicial sets.
Intuitively, a simplicial set can be thought of as a kind of hybrid or compromise between a simplicial complex (more special) on the one hand and a cell complex (more general) on the other hand. As in the case of simplicial complexes, the building blocks (cells) of which a simplicial set is constructed are simplices (vertices, edges, triangles, tetrahedra, …), and the boundary of each n-simplex Δn is attached to the lower dimensional skeleton by identifications that are linear on each proper face (subsimplex) of Δn; thus, these identifications can be described combinatorially by maps between the vertex sets of the simplices.Footnote 13 However, the attachments are more general than the one permitted for simplicial complexes; for example, one may have several 1-dimensional simplices connecting the same pair of vertices, a 1-simplex forming a loop, two edges of a 2-simplex identified to create a cone, or the boundary of a 2-simplex all contracted to a single vertex, forming an S
2.
Moreover, one keeps track of certain additional information that might seem superfluous but turns out to be very useful for various constructions. For example, even if the identifications force some n-simplex to be collapsed to something lower dimensional (so that it could be discarded for the purposes of describing the space as a cell complex), it will still be formally kept on record as a degenerate
n-simplex; for instance, the edges of the triangle with a boundary contracted to a point (the last example above) do not disappear—formally, each of them keeps a phantom-like existence of a degenerate 1-simplex.
Formally, a simplicial set X is given by a sequence (X
0,X
1,X
2,…) of mutually disjoint sets, where the elements of X
n
are called the n-simplices of
X (we note that, unlike the case of simplicial complexes, a simplex in a simplicial set need not be determined by the set of its vertices; indeed, there can be many simplices with the same vertex set). The 0-simplices are also called vertices.
For every n≥1, there are n+1 mappings ∂
0,…,∂
n
:X
n
→X
n−1 called face operators; the intuitive meaning is that for a simplex σ∈X
n
, ∂
i
σ is the face of σ opposite to the ith vertex. Moreover, there are n+1 mappings s
0,…,s
n
:X
n
→X
n+1 called the degeneracy operators; the approximate meaning of s
i
σ is the degenerate simplex which is geometrically identical to σ, but with the ith vertex duplicated. A simplex is called degenerate if it lies in the image of some s
i
; otherwise, it is nondegenerate. We write X
ndg for the set of all nondegenerate simplices of X. A simplicial set is called finite if it has only finitely many nondegenerate simplices (if X is nonempty, there are always infinitely many degenerate simplices, at least one for every positive dimension).
There are natural axioms that the ∂
i
and the s
i
have to satisfy, but we will not list them here, since we won’t really use them. Moreover, the usual definition of simplicial sets uses the language of category theory and is very elegant and concise; see, e.g., [7, Sect. 4.2].
If A and X are simplicial sets such that A
n
⊆X
n
for every n and the face and degeneracy operators of A are the restrictions of the corresponding operators of X, then we call A a simplicial subset of X.
Examples
Here we sketch some basic examples of simplicial sets; again, we won’t provide all details, referring to [8]. Let Δp denote the standard p-dimensional simplex regarded as a simplicial set. For p=0, (Δ0)
n
consists of a single simplex, denoted by 0n, for every n=0,1,…; 00 is the only nondegenerate simplex. The face and degeneracy operators are defined in the only possible way.
For p=1, Δ1 has two 0-simplices (vertices), say 0 and 1, and in general there are n+2 simplices in (Δ1)
n
; we can think of the ith one as containing i copies of the vertex 0 and n+1−i copies of the vertex 1, i=0,1,…,n+1. For p arbitrary, the n-simplices of Δp can be thought of as all nondecreasing (n+1)-term sequences with entries in {0,1,…,p}; the ones with all terms distinct are nondegenerate.
In a similar fashion, every simplicial complex K can be converted into a simplicial set X in a canonical way; first, however, we need to fix a linear ordering of the vertices. The nondegenerate n-simplices of X are in one-to-one correspondence with the n-simplices of K, but many degenerate simplices show up as well.
Geometric Realization
Like a simplicial complex, every simplicial set X defines a topological space |X|, the geometric realization of
X, which is unique up to homeomorphism. More specifically, |X| is a cell complex with one n-cell for every nondegenerate
n-simplex of X, and these cells are glued together according to the identifications implied by the face and degeneracy operators (we omit the precise definition of the attachments, since we will not really use it and refer to the literature, e.g., to [8] or [7, Sect. 4.3]).
Simplicial Maps
Simplicial sets serve as a combinatorial way of describing a topological space; in a similar way, simplicial maps provide a combinatorial description of continuous maps.
A simplicial map
f:X→Y of simplicial sets X,Y consists of maps f
n
:X
n
→Y
n
, n=0,1,…, that commute with the face and degeneracy operators.
A simplicial map f:X→Y induces a continuous, in fact, a cellular map |f|:|X|→|Y| of the geometric realizations in a natural way (we again omit the precise definition). Often we will take the usual liberty of omitting |⋅| and not distinguishing between simplicial sets and maps and their geometric realizations.
Of course, not all continuous maps are induced by simplicial maps. However, simplicial maps can be used to approximate arbitrary continuous maps up to homotopy. The simplicial approximation theorem (which may be most familiar in the context of simplicial complexes) says that for an arbitrary continuous map φ:|X|→|Y| between the geometric realizations of simplicial sets, with X finite, there exist a sufficiently fine subdivision
X′ of X and a simplicial map f:X′→Y whose geometric realization is homotopic to φ; see Sect. 3.4 for more details.
Encoding Finite Simplicial Sets
A finite simplicial complex can be encoded in a straightforward way by listing the vertices of each simplex.
For simplicial sets, the situation is a bit more complicated, since the simplices are no longer uniquely determined by their vertices, but if X is finite, then we can encode X by the set X
ndg of its nondegenerate simplices (which we assume to be numbered from 1 to N, where N is the total number of nondegenerate simplices), plus a little bit of additional information.
The simple but crucial fact (see, e.g. [7, Theorem 4.2.3]) we need is that every simplex σ can be written uniquely as σ=sτ, where τ is nondegenerate and s is a degeneracy, i.e., a composition \(s=s_{i_{k}}\ldots s_{i_{1}}\) of degeneracy operators where \(k=\mathop{\mathrm{dim}}\nolimits \sigma-\mathop{\mathrm{dim}}\nolimits \tau\) (in particular, σ is nondegenerate itself if σ=τ and s is the identity). Thus, as mentioned above, degenerate simplices σ do not need to be encoded explicitly but can be represented by sτ when needed, where the degeneracy s can be encoded by the sequence (i
k
,…,i
1) of indices of its components.Footnote 14 The extra information we need to encode X, in addition to the list of its nondegenerate simplices, is how these fit together. Specifically, for \(\sigma\in X_{n}^{\mathrm{ndg}}\) and 0≤i≤n, the ith face can be written uniquely as ∂
i
σ∈X
n−1=sτ with τ nondegenerate, and for each σ, we record the (n+1)-tuple of pairs (τ,s).
Similarly, if f:X→Y is a simplicial map between finite simplicial sets, then given the encodings of X and Y, we can encode f by expressing, for each \(\sigma\in X_{n}^{\mathrm{ndg}}\), the image f(σ)=sτ, with \(\tau\in Y_{m}^{\mathrm{ndg}}\) and recording the list of triples (σ,τ,s).
For a finite simplicial set X, we define \(\operatorname {\mathsf {size}}(X)\) as the number of nondegenerate simplices. If the dimension of X is bounded by some number d, then the number of bits in the encoding of X described above is bounded by \(O(\operatorname {\mathsf {size}}(X)\log \operatorname {\mathsf {size}}(X))\), with the constant of proportionality depending only on d.
The notion of size will be a convenient tool that allows us to ensure that our reductions can be carried out in polynomial time, without analyzing the running time in complete detail, which we feel would be cumbersome and not very enlightening.
More specifically, our reductions will be composed of a sequence of various basic constructions of simplicial sets, which will be described in the next subsection.
For each of these basic constructions, it is straightforward to checkFootnote 15 that when we apply them to finite simplicial sets of bounded dimension, both the running time of the construction (the number of steps needed to compute the encoding of the output from the encoding of the input) as well as the size of the output simplicial set are polynomial in the size of the input. Thus, to ensure that the overall reduction is polynomial, it will be enough to take care that we combine only a polynomial number of such basic constructions, that the size of every intermediate simplicial set constructed during the reduction remains polynomial in the initial input, and that the dimension remains bounded.
Basic Constructions
In this subsection, we review several basic constructions for cell complexes and simplicial sets. (One advantage of simplicial sets over simplicial complexes is that various operations on topological spaces, in particular Cartesian products and quotients, have natural counterparts for simplicial sets. This is where the degeneracy operators and degenerate simplices turn out to be necessary.) For more details, we refer to [7, 9].
Pointed and k-Reduced Simplicial Sets and Cell Complexes
Several of the constructions are defined for pointed spaces. We recall that a pointed space (X,x
0) is a topological space X with a choice of a distinguished point x
0∈X (the basepoint). If X is a cell complex or a simplicial set, then we will always assume the basepoint to be a vertex (i.e., a 0-cell or 0-simplex, respectively). A pointed map (X,x
0)→(Y,y
0) of pointed spaces (cell complexes, simplicial sets) is a continuous (cellular, simplicial) map sending x
0 to y
0. Homotopies of pointed maps are also meant to be pointed; i.e., they must keep the image of the basepoint fixed. The reader may recall that, for example, the homotopy groups π
k
(Y) are defined as homotopy classes of pointed maps. The set of pointed homotopy classes of pointed maps X→Y will be denoted by [X,Y]∗.
A simplicial set X is called k-reduced, k≥0, if it has a single vertex and no nondegenerate simplices in dimensions 1 through k. Similarly, a cell complex X is k-reduced if it has a single vertex and no cells of dimensions 1 up to k. It is then necessarily k-connected.
If (Y,y
0) is a 0-reduced cell complex (or simplicial set), then any cellular (or simplicial) map from a pointed complex (X,x
0) into Y is automatically pointed. Moreover, if Y is 1-reduced, then every homotopy is pointed, too, and thus [X,Y] is canonically isomorphic to [X,Y]∗.
Products
If X and Y are cell complexes, then their Cartesian product X×Y has a natural cell complex structure whose n-cells are products e
p×e
q, where p+q=n and e
p and e
q range over the p-cells of X and the q-cells of Y, respectively.
Furthermore, if X and Y are simplicial sets, then there is a formally very simple way to define their product X×Y: one sets (X×Y)
n
:=X
n
×Y
n
for every n, and the face and degeneracy operators work componentwise; e.g., ∂
i
(σ,τ):=(∂
i
σ,∂
i
τ). As one would expect from a good definition, the product of simplicial sets corresponds to the Cartesian product of their geometric realizations, i.e., |X×Y|≅|X|×|Y|.Footnote 16 The apparent simplicity of the definition hides some intricacies, though, as one can guess after observing that, for example, the product of two 1-simplices is not a simplex—so the above definition must imply some canonical way of triangulating the product.
Remark 3.2
A pair (sσ,tτ) of degenerate simplices in the factors may yield a nondegenerate simplex in the product, if the degeneracies s and t are composed of different degeneracy operators s
i
. However, \(\mathop{\mathrm{dim}}\nolimits (X\times Y)=\mathop{\mathrm{dim}}\nolimits X +\mathop{\mathrm{dim}}\nolimits Y\), so the product contains no nondegenerate simplices of dimension larger than \(\mathop{\mathrm{dim}}\nolimits X+\mathop{\mathrm{dim}}\nolimits Y\), and hence \(\operatorname {\mathsf {size}}(X\times Y)\) is at most \(\operatorname {\mathsf {size}}(X)\times \operatorname {\mathsf {size}}(Y)\) times some factor that depends only on the dimensionFootnote 17
\(\mathop{\mathrm{dim}}\nolimits (X\times Y)\).
Moreover, if the dimensions are bounded, the product can be constructed in polynomial time.
Quotients and Attachments
If X, Y, and A are cell complexes with A⊆X and if f:A→Y is a cellular map, then the space X∪
f
Y obtained by attaching X to Y along f is also a cell complex in a natural way (see, e.g., [7, Sect. 2.3]). In particular, X/A is a cell complex, with cells corresponding to the cells of X not contained in A, plus one additional 0-cell (corresponding to the image of A under the quotient map).
Similarly, if X is a simplicial set and if ∼ is an equivalence relation on each X
n
that is compatible with the face and degeneracy operators, then the quotient X/∼ is also a simplicial set. In particular, this includes simplicial attachments
X∪
f
Y of simplicial sets along a simplicial map f:A→Y defined on a simplicial subset A⊆X, and quotients X/A by simplicial subsets. These constructions are compatible with geometric realizations, e.g., |X∪
f
Y|≅|X|∪|f||Y|.
Moreover, the size of X∪
f
Y is at most the size of X plus the size of Y, and in bounded dimension, the attachment can be constructed in polynomial time.
Wedge Sum (or Wedge Product)
If X
1,…,X
m
are pointed spaces, then their wedge sum
X
1∨⋯∨X
m
is simply the disjoint union of the X
i
with the basepoints identified (this is a very special type of attachment). If the X
i
are cell complexes or simplicial sets, then so is their wedge sum.
Later we will need the following bijection:
$$ [X_1\vee X_2\vee\cdots\vee X_{m},Y]_*\xrightarrow{\cong}[X_1,Y]_*\times[X_2,Y]_*\times\dots\times [X_{m},Y]_* $$
(5)
where the components of this map are given by the restrictions to the respective X
i
.
Mapping Cylinder and Mapping Cone
For a map f:X→Y, the mapping cylinder of f is the space \(\operatorname {Cyl}(f)\) defined as the quotient of (X×[0,1])⊔Y under the identifications (x,0)∼f(x) for each x∈X. The mapping cone
\(\operatorname {Cone}(f)\) is defined as the quotient \(\operatorname {Cyl}(f)/(X\times\{1\})\) of \(\operatorname {Cyl}(f)\) with the subspace X×{1} collapsed into a point.
By the discussion concerning attachments, if X and Y are cell complexes and f is cellular, then \(\operatorname {Cyl}(f)\) and \(\operatorname {Cone}(f)\) are cell complexes as well. Moreover, if f is a simplicial map between simplicial sets, then by taking the analogous simplicial attachments and quotients, we obtain simplicial sets, denoted by \(\operatorname {Cyl}(f)\) and \(\operatorname {Cone}(f)\) as well, and called the simplicial mapping cylinder and simplicial mapping cone, respectively. The simplicial constructions are compatible with geometric realizations, for example, \(|\operatorname {Cyl}(f)|\cong \operatorname {Cyl}(|f|)\).
We will use the mapping cylinder in our construction to replace an arbitrary map f:X→Y by an inclusion \(X\hookrightarrow \operatorname {Cyl}(f)\), which has the same homotopy properties as f. A more precise statement is given in the following lemma (see, e.g., [9, Corollary 0.21]).
Lemma 3.3
Let
f:X→Y
be a continuous map between topological spaces. We consider
X≅X×{1} and
Y
as subspaces of
\(\operatorname {Cyl}(f)\)
and denote the corresponding inclusion maps
Footnote 18
by
\(i_{X}\colon X\hookrightarrow \operatorname {Cyl}(f)\)
and
\(i_{Y}\colon Y\hookrightarrow \operatorname {Cyl}(f)\).
-
(a)
Y
is a strong deformation retract
Footnote 19
of
\(\operatorname {Cyl}(f)\).
-
(b)
X (considered as a subspace via
i
X
) is a strong deformation retract of
\(\operatorname {Cyl}(f)\)
iff
f
is a homotopy equivalence.
-
(c)
i
X
∼i
Y
f
are homotopic as maps
\(X\to \operatorname {Cyl}(f)\).
-
(d)
If
f:X→Y
is a homotopy equivalence and if
g:Y→X
is a homotopy inverse for
f, then
i
X
g∼i
Y
as well.
Reduced Mapping Cone and Mapping Cylinder
If X and Y and f are pointed, with basepoints x
0 and y
0, it will be technically convenient, particularly in Sect. 6, to consider the spaces \(\mathop{\mathrm{\widetilde {\operatorname {Cyl}}}}\nolimits (f)\) and \(\mathop{\mathrm{\widetilde {\operatorname {Cone}}}}\nolimits (f)\), called the reduced mapping cylinder and the reduced mapping cone, respectively, that are obtained from \(\operatorname {Cyl}(f)\) and \(\operatorname {Cone}(f)\) by collapsing the segment x
0×[0,1] (whose lower end is identified with y
0) to a single point. We will apply this construction only to cellular or simplicial mapping cylinders and cones, in which case contracting the subcomplex x
0×[0,1] is a homotopy equivalence.
Moreover, if f is a homotopy equivalence, then we may assume that its homotopy inverse g is pointed as well and that the homotopies fg≃id
Y
and gf≃id
X
keep the basepoints fixed (see [9, Corollary 0.19]). It follows that Lemma 3.3 remains true if we take \(C=\mathop{\mathrm{\widetilde {\operatorname {Cyl}}}}\nolimits (f)\) as the reduced mapping cylinder (the inclusions are given as those into \(\operatorname {Cyl}(f)\), followed by the quotient map \(\operatorname {Cyl}(f)\to \mathop{\mathrm{\widetilde {\operatorname {Cyl}}}}\nolimits (f)\), which does not make any identifications within X or within Y).
By the remarks concerning the size of simplicial products and attachments, the size of the (reduced or unreduced) simplicial mapping cylinder or cone is at most the size of X plus the size of Y, times a factor depending only on \(\mathop{\mathrm{dim}}\nolimits X\).
Subdivisions and Simplicial Approximation
For simplicial complexes, there is the well-known notion of barycentric subdivision (see, e.g., [18, §15]). An analogous notion of subdivision, called normal subdivision, can also be defined for simplicial sets. Informally speaking, the normal subdivision \(\operatorname {Sd}(X)\) of a simplicial set X is defined by barycentrically subdividing each simplex of X and then gluing these subdivided simplices together according to the identifications implied by the face and degeneracy operators of X. We refer to [7, Sect. 4.6] for the precise formal definition and just state the facts that we will need in what follows.
For the standard simplex Δp, the nondegenerate k-simplices of \(\operatorname {Sd}(\Delta^{p})\) correspond to chains of proper inclusions of nondegenerate simplices (faces) of Δp. It follows that \(\operatorname {Sd}(\Delta^{p})\) has (p+1)! nondegenerate p-simplices and, in general, at most 2p+1(p+1)! nondegenerate simplices of any dimension. Consequently, for any simplicial set X, the size of \(\operatorname {Sd}(X)\) is at most the size of X times a factor that depends only on \(\mathop{\mathrm{dim}}\nolimits X\) and which can be bounded from above by \(2^{\mathop{\mathrm{dim}}\nolimits X+1}(\mathop{\mathrm{dim}}\nolimits X+1)!\). Moreover, if the dimension is bounded, \(\operatorname {Sd}(X)\) can be constructed in time polynomial in \(\operatorname {\mathsf {size}}(X)\).
If f:X→Y is a simplicial map, then subdivision also induces a map \(\operatorname {Sd}(f)\colon \operatorname {Sd}(X)\to \operatorname {Sd}(Y)\), and this is compatible with compositions, i.e., \(\operatorname {Sd}(fg)= \operatorname {Sd}(f) \operatorname {Sd}(g)\).
For each simplicial set X, there is a simplicial map \(\mathop{\mathrm{lastv}}\nolimits _{X}\colon \operatorname {Sd}(X)\to X\), called the last vertex map,Footnote 20 which is a homotopy equivalence that is compatible with simplicial maps f:X→Y, i.e., \(f \mathop{\mathrm{lastv}}\nolimits _{X}=\mathop{\mathrm{lastv}}\nolimits _{Y} \operatorname {Sd}(f)\).Footnote 21
,
Footnote 22
There is also a simplicial approximation theorem for simplicial sets, which uses iterated normal subdivisions. Specifically, the t-fold iterated normal subdivision of a simplicial set is defined inductively as \(\operatorname {Sd}^{t}(X):=\operatorname {Sd}(\operatorname {Sd}^{t-1}(X))\), where \(\operatorname {Sd}^{0}(X):=X\).
Theorem 3.4
[7, Theorem 4.6.25]
Let
X
and
Y
be simplicial sets such that
X
has only finitely many nondegenerate simplices, and let
f:|X|→|Y| be a continuous map. Then there exist a finite integer
t (which depends on
f) and a simplicial map
\(g\colon \operatorname {Sd}^{t}(X)\to Y\)
such that |g| is homotopic to the composition
\(f|\mathop{\mathrm{lastv}}\nolimits _{X}^{t}|\)
of
f
with the iterated last vertex map
\(\mathop{\mathrm{lastv}}\nolimits _{X}^{t}\colon \operatorname {Sd}^{t}(X)\to Y\).
To convert arbitrary simplicial sets into homotopy equivalent (in fact, homeomorphic) simplicial complexes, another subdivision-like operation is needed (see, e.g., [11]). Given a simplicial set Z, one can define a simplicial complex B
∗(Z) inductively, by introducing a new vertex v
σ
for every nondegenerate simplex σ, and then replacing σ by the cone with apex v
σ
over B
∗(∂σ). If the simplicial set Z has a certain regularity property—which is satisfied, for instance, if \(Z=\operatorname {Sd}(X)\)—then B
∗(Z) and Z are homotopy equivalent (in fact, homeomorphic).Footnote 23 We summarize the properties that we need in the following proposition (for completeness, we provide a proof in Appendix A).
Proposition 3.5
If
X
is a simplicial set, then the twofold subdivision \(B_{*}(\operatorname {Sd}(X))\)
is a simplicial complex. Moreover, there is a simplicial map
\(\gamma_{X}\colon B_{*}(\operatorname {Sd}(X))\to X\), which is a homotopy equivalence. For a simplicial subset
A⊆X, \(B_{*}(\operatorname {Sd}(A))\)
is a subcomplex of
\(B_{*}(\operatorname {Sd}(X))\)
and
γ
X
|
A
=γ
A
.Footnote 24
If
X
is finite and of bounded dimension, there are algorithms that construct the simplicial complex
\(B_{*}(\operatorname {Sd}(X))\)
and evaluate the map
γ
X
, both in polynomial time.