Arun Ram
Department of Mathematics and Statistics
University of Melbourne
Parkville, VIC 3010 Australia
aram@unimelb.edu.au
Last update: 10 May 2013
§1. Introduction
Let be a hyperplane system.
In this chapter we consider the question under which circumstances one can prove that the space
is a If this is the case we call the system
a
(1.1) Definition.
A hyperplane system is called simplicial, if the cone on a chamber is a simplicial cone, i.e. of the form
for some independent set
The cone on a set is defined to be
Cf. figure 32 (simplicial) and figure 24 (not simpl.).
The result of Deligne [Del1972] can be stated as (see (3.2) for central):
(1.2) Theorem A simple, central and simplicial hyperplane system is
(1.3) Conjecture. This theorem is still true if one drops the condition that the system has to be central.
After some preliminaries in §2 we construct in §3 an open cover of the universal covering
space of To ensure that the abstract simplicial complex
has simplices with a finite number of vertices we suppose from now on that the configuration satisfies:
(1.4) Assumption. For each facet the
number of facets contained in is finite.
Assuming that the system is
(3.7), by which we mean that the star subsystems (2.9) are
we prove that this cover is simple i.e. that intersections are contractible. Then the geometric realisation
of
has the same homotopy type as
(Weil [Wei1952]). So the
is reduced to a contractibility problem
for a simplicial complex, if one knows that the system is locally
But this problem is
often solved at the same time one solves the contractibility problem for the complex in a positive way, because one can then apply induction on the dimension of
We investigate the simplicial complex. In §4 we do this in the general situation and give a description in terms of galleries. In §5 we describe the
complex in terms of (extended) Artin groups in the cases that the system is associated to a linear Coxeter group or an extended Coxeter group (generalised root system).
The construction is then of the type of the "Coxeter complex" of a Coxeter group.
As example (3.5) shows, not every hyperplane system is a
In the central case the simpliciality takes care for a nice triangulation of a sphere around the origin. The simplicial complex
(defined in (3.6)), which is the complex
used by Deligne, can be described in terms of this triangulation and the galleries. But in this context the use of the simpliciality condition is not essential.
Because we wanted to know where the use of the condition was crucial, we have tried to postpone the use as far as possible. This is a main reason why we construct in
§6 a space
homotopic to
Besides this the space is more related to the structure of the facets and the galleries.
Also in §6 we sketch the way how certain generalisations of the theorem of Deligne (1.2), for instance conjecture (1.3), perhaps can be proved.
One of the conjectures we need (namely "simpliciality implies the direct pieces property (5.23)) can be seen as a generalisation of the "key lemma" of Deligne ([Del1972]
prop. 1.19). Strengthening somewhat this conjecture gives an interesting and simple solution conjecture about the word problem for Artin groups. In general for the
proposed kind of proof of the conjecture it seems necessary to
have knowledge about the minimal representatives of galleries. Hence the connection with word problems.
§2. Stars and their hyperplane systems
Let again be a hyperplane system. We
need some more notions and properties; first of the configuration
(2.1) Definition. Let then the
star of is defined by
(2.2) Proposition. We have
Proof.
(ii) Since for all
is contained in a chamber with respect to
(i) By (ii) The right hand side of (i) equals
Now let then by criterion (I.3.5) for each
we have
so
is contained in the right hand side. And conversely.
(2.3) Corollary. is open in
(2.4) Lemma. Let
Then the following properties are equivalent:
(i)
(ii)
(iii)
For each we have
Proof.
(i) (ii) Let and
So
Then
so
(ii) (iii) is trivial.
(iii) (i) Let
By criterion (I.3.5) we have to show that
Suppose not. Then
where is the other halfspace. So
is an open nonempty set so there exists a chamber which intersects it and since it is also a union of facets,
is even contained in it. By (iii)
so contradiction
with
(2.5) Lemma. Let
then:
(i)
so
(ii)
(iii)
Proof.
(i) If then so
(ii) If then
for all
And with (I.3.5):
It follows by definition of facet that
(iii) follows from (i) and (ii).
(2.6) Proposition. Let
and
Then there is a facet such that
Proof.
is a union of facets (since is so (I.3.4)). Take
with maximal dimension. Suppose
To prove: Choose
If then finished. If
then consider the line through and
Case 1: contains only
Since is open (2.3),
is open in
so there is a
Let be the facet containing the point Then
but
for So by
(2.5)(iii)
But since
is convex, we have
so
Hence
a contradiction.
Case 2: Since is open in
is an open interval on
containing If
then
and we are ready. If
and for some
then there is a
such that and
So and are at different sides of
but
is an intersection of hyperplanes in and closed halfspaces.
is one of them,
so then
contradiction.
(2.7) Definition. Let
If
then we denote the unique facet such that
as exists by (2.6) by
If and are such that there exists a
such that
then by locally finiteness there are only finitely may such facets Then by (2.6) thre is a facet
such that
We denote this facet by
Note that for
(respectively
and this "meet" (respectively "join") is maximal (respectively minimal) with this property, if it exists.
(2.8) Theorem. If
then
is nonempty, if and only if is defined and in this case
Proof.
The equality ($) holds because
(2.9) Definition. (cf. def. (II.4.8)) Let again
and define
Note that
(or simply
if is considered as a map Furthermore note that
(2.10) Proposition.
(cf. prop. (II.4.9)). Let
Then there exists a map
with the properties:
(i)
The map is a homotopy inverse of the inclusion
(ii)
For another
such that exists we have
Proof.
Choose and let be a closed ball around inside
in some metric
is open!). For each
there is a unique
such that
Now is a continuous map
The rest of the proof is analogous to the proof of proposition (II.4.9) from the point ($). Including the proof of (ii) for which we have to remark that
(2.11) Corollary.
The spaces and are homotopy equivalent.
(2.12) Corollary. Let be the inclusion
and
then the homomorphism
is injective.
(2.13) Lemma. Let
such that exists and let
Then
where these fundamental groups are seen as subgroups of
which is done by (2.12).
Proof.
This follows from (2.10)(ii) in the same way as (II.4.12) follows from (II.4.9)(ii).
The following corollary is the key to an important result of this chapter (theorem (3.11)). It makes it possible to prove this result for so called
non central (def. (3.2)) systems. This in contrast to Deligne [Del1972]. Compare remark (3.14).
(2.13) Corollary. For let
and
a curve in
Suppose is homotopic to in
Then exists and the curves
are homotopic in to a curve
in
Proof.
It is easy to see that we may restrict ourselves to the case that and
are loops, say from a point
From it follows that
and so it is an element contained in the left hand side of (2.14). So it is also contained in the right hand side, hence of the form
where is a curve in
(2.16) Definition. Let
Define:
(2.17) Lemma. We have
Proof.
This is a simple inspection of the proof of (I.1.5).
(2.18) Corollary
of (2.15). For let
If
and
then
exists and
for some gallery
Proof.
Let be a curve in such
that
as exists by (2.17). From it follows that
hence So by (2.15) there
is a curve in
such that Then there is by (2.17) a
such that
Then
hence
(2.19) Remark. It is possible and, since we have developed all of the necessary tools, perhaps more natural to prove (2.18) and a "gallery version"
of (2.12) independent from (2.10) and its corollaries with the help of galleries alone. Here is a sketch:
Proof of (2.18) (sketch).
Let for
With the help of (I.3.9) and (I.3.10) we have a map
This induces a map
which respects gallery equivalence. (This map corresponds in a sense with the map in prop. (2.10)). Now let
Then
In fact this is the way (2.18) and hence its consequence (3.11) was discovered. We have given the other approach, because the geometrical background seems more clear.
§3. An open cover of the universal covering space
After these preliminaries we are going to construct an open cover of the universal covering space of
where
is a hyperplane system as before.
(3.1) Definition. We call a facet
minimal if it is so for the natural order relation i.e. if there is no facet contained in the closure
(3.2) Definition. We call a hyperplane system central if and
We call a system non central if there doesn't exist a unique minimal facet.
(3.3) Remark. As the terminology suggests, we can suppose that a system is either central or non central. For, suppose there exists a unique minimal
facet Then since each facet contains at least one minimal facet, we have for all
that
Hence
Hence
The only hyperplanes, which meet are the elements of
so
By (2.11) we have
Since we are only interested in the homotopy type of we see that indeed we can suppose that the system is central in
the sense of def. (3.2), if there exists a unique minimal facet. Note that we have:
(3.5) Definition. We agree that in the non central case
(this fits in with (3.4)!). We define the set of proper facets by:
(hence
in the non central case). A proper facet is called properly minimal if its closure doesn't contain a different proper facet. The set of
properly minimal facets is denoted by
(3.6) Definition. Let be the universal
covering. For define:
the collection of components of
and:
If the system is central, then is finite, hence
is locally finite on For
we define: the collection of components of
and
and we agree that if the system is non central. Furthermore we define:
(3.7) Definition. We call a hyperplane system
if
is a for each
This means that
and hence (by (2.11)) are
(3.8) Example. A simple, simplicial system is
(central or non central). This follows from the theorem of Deligne (1.2) and the observation that the local system
is simplicial if the original system is so, and this local system is clearly central.
(3.9) Example. If it is true that a simple, simplicial system is
(conjecture (1.3)) at least in the case of an affine linear Coxeter group then the hyperplane systems associated to the extended Coxeter groups (III.1.6) are
(3.10) Example. (The "triangle"). Let be a plane and let
consist of three different lines with
but
(so a non central system!). See Figure 24. Consider the simple hyperplane system of this configuration. The local systems are easily seen to be simplicial, hence
So the system is
But it is not
This is well known and can be seen by the method of
§4 (see (4.17)).
(3.11) Theorem. Let
be a hyperplane system which is
Then as defined in (3.6) is a simple covering of the universal covering space of
In particular this universal covering space is homotopy equivalent to the geometric realisation
of the abstract simplicial complex the nerf
of the cover
Proof.
First we prove the following remark: If
for and
then exists and
i.e.
is closed under intersections. Now
because so it equals
(theorem (2.8)). Let
Then we have
The right hand side is a union of intersections of elements in
and and each intersection is in turn a union of components.
So, for instance is a union of elements in
and we have to show that this union consists of just one element, that is we have
to see that is connected.
For this purpose suppose and let
be a curve in from
to (each is connected). We have
in because is simply connected.
So in
where
Now is in so
by corollary (2.15) we know that where is a
curve in Let
be the unique lift of to with
then
in
in particular
Furthermore is inside so we have connected and
by a curve in hence is connected. This proves the first remark i.e.
that is closed under intersections.
It can be verified that for
is a universal covering. For instance the simply connectedness of follows from the fact that
induces an injection of fundamental groups (2.12).
The assumption ensures now that
is a
if Hence
is contractible then. Together with the first remark it now follows that all intersections of elements of are
contractible. In the non central case the elements of cover and so this is then indeed
a simple cover.
In the central case we still have to cover the points of with
This can be done with the elements of Since these elements are mutually disjoint, we still have only to
prove that is contractible (if nonempty) for
and
where
and
We have:
Argueing as after (3.12) we can conclude from (3.13) that is a union of components of the right hand side of
(3.13). We will see that this union consist of just one element i.e. That is connected. The only thing we have
to do is to prove the following analog of corollary (2.15): Let (respectively
be a curve in
(respectively in
where we consider and suppose
in Then
there is a curve in
such that in
To prove this we define a curve by
Let be a homotopy from to then
gives a homotopy from to So
as we wanted.
So is indeed connected.
is again a universal covering. The base space is convex, hence contractible (and
is even a homeomorphism). So
is indeed a simple open cover.
As said before, the second statement of the theorem follows from Weil [Wei1952].
(3.14) Remark. Note that we did not suppose simpliciality nor centrality. This in contrast with Deligne. His ingredient which is comparable with
corollary (2.15) is [Del1972] 1.31. And there he uses the existence of a gallery
which depends on the fact that the situation is central.
(3.15) Remark. We have the same result for the open cover
and the abstract simplicial complex
as easily can be seen. The space looks
like a "thickening" of the space
§4. The Simplicial Complex in terms of Galleries
In this paragraph we study the simplicial complex defined in the previous paragraph with the help of the galleries of the hyperplane system. We fix a chamber
a point
and a point
above where
is the universal covering.
(4.1) Definition. For define:
(Note that
and recall (2.16):
(4.2) Definition. Let
for We write
if
(i)
(ii)
There is a such that
(gallery equivalence (I.4.7))
Note that
by (i), so
so (ii) make sense. Note also that this relation is transitive.
(4.3) Definition. If
then we call and
notation:
if
(then automatically
i.e. if there is a such that
The
classes are denoted by
If then
Note that for all
where is gallery equivalence. Furthermore
(note that this is a disjoint union).
(4.4) Definition. Let
for Then
means that
where is in the class
It is easily checked that this indeed does not depend on the choice of the representatives.
(4.5) Proposition
(i)
The relation is transitive.
(ii)
Suppose
and
then there is a unique
such that
(iii)
If and
there exists a such that
then there exists a unique
such that
which is the smallest, i.e. if then
(iv)
If and
there exists a such that
then there exists a unique
such that
which is the biggest, i.e. if then
(v)
If
then if and only if
as sets.
(vi)
If
then if and only if there is
a such that
Proof.
(i) is trivial.
(ii) Let G2∈g2. Then
G2∈GalCF2⊂GalCF1,
so G2 also determines a g1≤g2,g1∈galCF1.
If also g1′≤g2 and
G1∈g1′,
then G2∼G1G for some
G∈GalCF1,
by the definition of g1′≤g2.
But this says at the same time that G1∼F1G2,
so g1=g1′.
(iii) Suppose gi∈galCFi,g∈galCF and gi≤g.
Then Fi⊂F‾, so
F1∨F2⊂F‾
is defined. By (ii) there is a unique g1∨g2∈galC,F1∨F2
such that g1∨g2≤g.Fi⊂F1∨F2‾,
so by (ii) there is a unique gi′∈galFi,
such that gi′≤g1∨g2.
By (i) gi′≤g and again by (ii) we have
gi′=gi. We still
have to prove that this construction of g1∨g2 does not depend on the choice of
g, so suppose h∈galC,F1∨F2
has also the property that gi≤h. Choose representatives
G1,G2,G and
H of g1,g2,g and
h respectively and let Gi′ (respectively
Gi′′) be an element
of GalFi, such that
G∼Gi·Gi′
(respectively H∼Gi·Gi′′),
as exists, since gi≤g1∨g2 (respectively
g1≤h). Then
G·(Gi′)-1∼H·(Gi′′)-1,
so we have g-1·G∼(Gi′′)-1·Gi′∈GalFi.
Then by corollary (2.18) we have (Gi′′)-1·Gi′∼G′
for some g′∈GalF1∨F2.
Then H-1·G∼G′, hence
G∼F1∨F2H,
so g1∨g2=h.
(iv) If g∈gal there are only finitely many h∈gal
such that h≤g, as follows from (ii) and (1.4). Therefore (iv) follows from (iii).
(v) Suppose gi∈galCFi and
g1≤g2. So
F1⊂F‾2.
Let G2∈g2, then
G2∈GalCF2⊂GalCF1.
If G1∈g1, then since
g1≤g2, there is a
G∈GalF1 such that
G2∼G1G, so
G1∼F1G2.
Hence G2∈g1. So
g2⊂g1. Now suppose
g2⊂g1. First we prove
F1⊂F‾2.
For each chamber A⊂St(F2) there is a
G2∈g2 such that
A=end(G2). Since
g2⊂g1, we also have
A⊂St(F2).
By lemma (2.4) it follows that F1⊂F‾2.
Now let Gi∈gi, then
G2∈g1, so
G2∼F1G1.
Hence there is a G∈GalF1 such that
G2∼G1G. So
g1≤g2.
(vi) If there is a g∈gal such that
gi≤g, then g⊂gi
by (v), so g1∩g2≠∅.
If G∈g1∩g2, then
end(G)⊂St(F1)∩St(F2).
So F1∨F2 exists and G determines a
g∈galC,F1∨F2
such that gi≤g.
□
(4.6) Remark.g1 and g2 is not necessarily an element of
galC,end(g1)∧end(g2).
(4.7) Definition. We define a mapping
u:{(G,F)|F∈F*(ℳ),G∈GalCF}→𝒰
as follows: Let γ be a curve in Y such that
φ(G)={γ}.
Then γ(1)∈Y(F).
Let γ∼ be the unique lift with
γ∼(0)=c∼.
Then γ∼(1) is determined by G
and is contained in a unique u(G,F)∈𝒰(F).
In the central case we define also a mapping
v:GalCC×K(pℳ)→𝒱
as follows: Let (G,A)∈GalCC×K(pℳ)
and suppose γ is a loop in Y such that
φ(G)={γ}.
Then the unique lift of γ[-1c,-1c+a],
where a∈A, that starts in c∼,
has his endpoint in a uniquely determined v(G,A)∈𝒱(A).
(4.8) Lemma. Let Ui≔u(Gi,Fi)
for i=1,2. Then
U1⊂U2 if and only if
(G2,F2)⊆(G1,F1).
Proof.
Let U1⊂U2, then clearly
St(F1)⊂St(F2),
so F2⊂F‾1.
Let γ∼i be as in definition (4.7). Let
γ∼ be a curve in U2 from
γ∼1(1) to
γ∼2(1) and
γ≔p∘γ∼.
Then γ is a curve in Y(F2),
so by lemma (2.17) we have φ(G)={γ}
for some G∈GalF. Now
γ2=γ1γ so by lemma (2.17)
we have φ(G)={γ} for some
G∈GalF. Now
γ2=γ1γ so by (I.4.10)
G2∼G1·G.
So indeed (G2,F2)⊆(G1,F1).
The reverse implication can also easily be seen.
□
(4.9) Theorem
(i)
The mapping u induces a bijection between gal and
𝒰, the vertices of Σ.
(ii)
The mapping v induces a bijection between galCC×K(pℳ)
and 𝒱, the vertices of Σ^,
which are not vertices of Σ.
Proof.
(i) Let U∈𝒰(F),A∈K(ℳ),A⊂St(F),a∈A and a∼ a point in U above
-1a,γ∼ a curve from
c∼ to a∼ and
γ=p∘γ∼.
According to (I.4.10) there is a G∈GalCA⊂GalCF,
such that φ(G)={γ}.
Then u(G,F)=U, so
u is surjective. From lemma (4.8) it follows that
u(G1,F1)=u(G2,F2)
if and only if F1=F2≕F and
G1∼FG2.
Hence u can be defined on gal and is then injective.
(ii) Let V∈𝒱(A),A∈K(pℳ).
Let γ∼1 be a curve from c∼
to a point in V. Now γ1≔p∘γ∼1
is a curve in Y from -1c to a point
γ1(1)∈Re-1(A).
We may assume that γ1(1)=a+-1c
for some a∈A. Then
γ1[a+-1c,-1c]
is a loop in Y on -1c, so by (I.4.10) there is a
G1∈GalCC such that
φ(G1)={γ1[a+-1c,-1c]},
hence the surjectivity. It is also straightforward to see that
v(G1,A1)=v(G2,A2)
if and only if A1=A2 and
G1∼G2. Hence we have a bijection
v:galCC×K(pℳ)→𝒱.
□
So from now on we can identify the vertices of Σ with the set gal.
We proceed by describing the simplices of Σ:
(4.10) Theorem. A set {g1,…,gk},gi∈gal constitutes a simplex in Σ if
and only if there is a g∈gal such that
gi≤g for i=1,…,k.
Proof.
Let Gi∈GalCFi
for i=1,…,k and
Ui≔u(Gi,Fi).
Then by definition {U1,…,Uk}
constitutes a simplex if and only if U≔U1∩…∩Uk≠∅.
Then U∈𝒰(F) for some
F∈F*(ℳ), hence
U=u(G,F) for some
G∈GalCF.
Since U⊂Ui, it follows from lemma (4.8) that
(Gi,Fi)⊆(G,F).
□
(4.11) Corollary. The simplicial complex Σ is isomorphic to
nerf(gal).
Proof.
This follows from (4.10) with the help of proposition (4.5)(vi).
□
(4.12) Remark. If we define galmin≔⋃F∈Fmin*galCF,
then it can be seen also, that the simplicial complex Σmin is isomorphic to
nerf(galmin).
(4.13) Definition. An element g∈gal (or a representative
G∈g) as in theorem (4.10) is said to point to that simplex
{g1,…,gk}.
Let g∈galCF and
{F1,…,Fk} a set of facets in
F‾. Then by proposition (4.5)(ii) there is a uniquely determined
gi∈galFi with
gi≤g and they constitute a simplex. In particular, if
{F1,…,Fk}
is the set of all the facets in F‾, then this simplex is denoted by
σ(g). The set of all simplices to which g
points is exactly the set of all subsimplices of σ(g).
The maximal simplices of Σ are the simplices of the form
σ(g) for
g∈galCA for some
A∈K(ℳ).
(4.14) Definition. Let the system be central. For a vertex V of
Σ^, which is not a vertex of
Σ (i.e. an element of 𝒱) we denote by
Σ(V) the subcomplex of
Σ consisting of all simplices that form with V a simplex of
Σ^. The corresponding subcomplex of
Σ^ is denoted by
Σ^(V).
For (g,A)∈galCC×K(pℳ)
we write Σ(g,A) (respectively
Σ^(g,A))
instead of Σ(v(g,A))
(resp. Σ^(v(g,A))).
We define Σmin(V) and
Σ^min(V) analogously.
In remark (6.21) we will see that Σ(g,A)
(resp. Σ(g,A)) has the
homotopy type of a sphere (respectively cell). Now we still have to finish our description of Σ^ in the central case:
(4.15) Theorem.
Suppose that the hyperplane system is central and let (g,A)∈galCC×K(pℳ)
and G∈g. Then the galleries that point to simplices of
Σ(g,A) are equivalent to
galleries of the form G·Gd where Gd is a direct
replacing gallery of CED with
D∈K(ℳ) and
E≔pr(p(C,D)⊂pℳ)(A).
Proof.
Let G1≔G and let the curve γ1 and the point
a∈A be as in the proof of (4.9)(ii). Let G2∈GalCF
point to a simplex {U1,…,Uk} of
Σ (definition (4.13)); suppose that this simplex forms a simplex in Σ^
with V=v(G1,A).
This means that if U≔u(G2,F)
then U∩V≠∅. For we may suppose that, if
Ui∈𝒰(Fi) then
F=F1∨…∨Fk,
hence U=⋂i=1kUi.
Now φ(G2)={γ2},
where γ2 is a curve in Y from
-1c to -1d, with
d∈D∈K(ℳ),D⊂St(F). Let
γ∼ be a curve in Y from c∼ to a point
in U∩V above a+-1d and put
γ≔p∘γ∼. Now define
the "straight" curves:
Then ?? ≃γλ2-1
and γ2≃γλ3.
Hence φ(G1)={γλ2-1λ1-1},φ(G2)={γλ3}.
Now φ(G1-1G2)={λ1λ2λ3}
and by the definition of the functor φ (I.1.3) the latter equals φ(CED),
where E≔pr(p(C,D)⊂pℳ)(A).
So we have G2∼G1·CED
as has to be proven.
Figure 26.
□
If moreover the hyperplane system is simple, then we have still another description of the extension
Σ⊂Σ^:
(4.16) Theorem. Suppose we have a simple central hyperplane system.
Then there is a bijection between galCC×K(pℳ)
and ⋃A∈K(ℳ)galCA,
the set of equivalence classes of galleries starting in C. So the subcomplexes
Σ(g,A) can also be labelled by the
latter. If g∈galCA and
G∈g, then the galleries that point to simplices of
Σ(g) are equivalent with galleries of the form
G·P where P is a direct positive gallery.
Proof.
Let us have the situation of theorem (4.15) and suppose moreover that the system is simple. Then we have
K(pℳ)=K(ℳ),
so A⊂K(ℳ). Let
C,C1,…,Cn=A
and A=D0,D1,…,Dk=D
be direct Tits galleries and E1≔pr(ℳ(C,A)⊂ℳ)(A)
and E2≔pr(ℳ(A,D)⊂ℳ)(A) Then
∅≠A⊂E∩E1∩E2, so
Gd∼RCED∼RCE1AE2D∼RC-C1-…Cn·D0+D1+…Dk.
(where ∼R denotes pregallery equivalence, see (I.1.4)). By proposition (I.4.8) the first and the last gallery are
also gallery equivalent, so G·Gd∼G′·P,
where P is the positive gallery
D0+D1…+Dk
and G′≔G·C-C1-…-Cn.
Now each gallery is equivalent to a gallery of this form and this induces the desired bijection between
galCC×K(pℳ) and
⋃A∈K(ℳ)galCA.
□
(4.17) Example. As an illustration let us determine the complex Σ^min
of the hyperplane system "the triangle", example (3.10) (equals Σmin in this non
central case). It is easily seen that the group galCC (which is in general the
fundamental group of Y !) is isomorphic to ℤ3, by the isomorphism
(n1,n2,n3)→g1n1g2n2g3n3,
where gi is the equivalence class of the gallery
Gi≔ii (in the notation used in this kind of examples, see
§(I.5)). For instance the commutation relation G1·G2∼G2·G1
is illustrated by flip relations around the plinth P3:
11221212122121212211Figure 27.
The minimal facets are the points: Fmin*(ℳ)={P1,P2,P3}.
For each g∈galCPi there is a unique
representative Gin∈g. So the vertices of
Σmin:galmin=⋃i=13galCPi
can be labelled by a disjoint union of three copies of ℤ:ℤ×{1,2,3}.
The two-simplices {(n1,1),(n2,2),(n3,3)}
are the maximal simplices of Σmin. The system is
locally-K(π,1), so by theorem (3.11) or in fact
remark (3.15) we have |Σmin|≃Y∼.
Now we can see why the system is not K(π,1): Take a subcomplex
Σ′⊂Σmin
consisting of the simplices {(n1,1),(n2,2),(n3,3)}
with ni∈{a,b}, where
a and b are two different integers. Then |Σ′|
is a 2-sphere (the surface of the platonic solid in figure 28). Since the simplicial complex
|Σmin| is two dimensional
|Σ′| generates a nontrivial element in
π2(Y∼).
Figure 28.
§5. The Artin Complex and the Extended Artin Complex
In the case that the hyperplane system is associated to a linear Coxeter group or an extended Coxeter group (i.e. a generalised root system) the simplicial complex
constructed in §3 can be described in terms of the groups involved. The construction turns out to be of the following general type:
(5.1) Definition. Let G be a group and ℋ a finite collection of subgroups. We define an abstract
simplicial complex, called the cosetnerf of the group G and the collection ℋ as follows:
(5.2) Remark. Since ∅≠{1}⊂⋂H∈ℋH,
the collection ℋ itself is a (maximal) simplex of cnerf(G,ℋ).
This simplex serves as a fundamental domain for the natural action of G on cnerf(G,ℋ).
(5.3) Lemma. the map G∋x→{xH|H∈ℋ}
induces a bijection between G/(⋂H∈ℋH)
and the set of maximal simplices of cnerf(G,ℋ).
Proof.
Straightforward.
□
(5.4) Lemma.
If G⊂G′, then the space
|cnerf(G′,ℋ)|
is homeomorphic to |cnerf(G,ℋ)|×G′/G,
where we give G′/G the discrete topology. Hence
|cnerf(G′,ℋ)|
is a disjoint union of spaces homeomorphic to |cnerf(G,ℋ)|.
Proof.
Let H∈ℋ. Then each left coset of G in
G′ is a disjoint union of left cosets of H.
So for each element of G′/G we have a subcomplex of
|cnerf(G′,ℋ)|.
The space |cnerf(G′,ℋ)|
is a disjoint union of this kind of subcomplexes and a point in one such a subcomplex cannot be connected to a point in another. Furthermore they are all isomorphic to
|cnerf(G,ℋ)|,
as can be seen from the action of G′ on
|cnerf(G′,ℋ)|.
□
(5.5) Lemma. The space
|cnerf(G,ℋ)| is
connected if and only if G is generated by ⋃H∈ℋH.
Proof.
If |cnerf(G,ℋ)|
is connected, then by (5.3) the index of the subgroup generated by the elements of ℋ is one. Conversely, let G be
generated by ⋃H∈ℋH. Then an arbitrary
vertex of nerf(G,ℋ) is of the form
xH,x∈H and H∈ℋ.
Write x=a1a2…ak,ai∈Hi∈ℋ for
i=1,…,k and let
xi≔ai…ai.
Then, since aiHi=xi…xiHi=x1…xi-1Hi=ai-1Hi
the sequence
|{xH,xkHk}|,|{xk-1Hk,xk-1Hk-1}|,…,|{x1H2,x1H1}|
is a sequence of 1-simplices connecting the vertex xH to a vertex of the fundamental simplex.
□
(5.6) Remark. It is convenient to restrict ourselves to collections ℋ, which have the following two properties:
(i)
⋂H∈ℋH={1}.
(ii)
G is generated by
⋃H∈ℋH.
By lemma (5.4) we do not lose much generality by supposing (ii). If (ii) is supposed we will suppress G in the notation:
cnerf(ℋ).
Property (i) ensures with the help of lemma (5.3) that we have a bijection
(5.7)σ:G→{maximal simplices of cnerf(ℋ)}.
In all the cases we will consider, the properties (i) and (ii) will be satisfied.
(5.8) Example. Let (W,{σ1,…,σℓ})
be a finite, affine or compact hyperbolic system i.e. the subgroups
Wi≔W{1,…,ℓ}-{i}
for i=1,…,ℓ are finite. Then the collection
ℋ≔{Wi|i∈{1,…,ℓ}}
satisfies (5.6) (i) and (ii). Now cnerf(ℋ) is the well known Coxeter complex.
Note that if W is finite, then |cnerf(ℋ)|
is homeomorphic to the ℓ-1-sphere.
(5.9) Theorem. Suppose a simple hyperplane system of a linear Coxeter group W is given. Let
(AW,{s1,…,sℓ})
be the associated Artin system. Then:
(i)
The simplicial complex Σ of the hyperplane system as defined in (3.6) is isomorphic to the
cosetnerf of the collection of subgroups of AW of the form
AWJ, where
J⫋{1,…,ℓ} and
WJ is finite. This collection satisfies (5.6) (i) and (ii).
(ii)
The system is central if and only if W is finite and in this case there is a bijection m ("midpoint") from
AW to the vertices of Σ^,
which are not vertices of Σ. If
x∈AW, then the maximal simplices of the subcomplex
Σ(m(x)) of
Σ are the simplices of the form:
σ(xw),w∈W,
where W has to be interpreted as a subset of AW by remark (II.3.19).
Proof.
The fact that the collection of subgroups of AW as in (i) satisfies (5.61)(i) follows from theorem (II.4.13)(ii).
We have a surjective map f:Gal→AW as the
composite Gal→Gal/W→φ/Wπ/W→≅AW,
where the first is canonical projection and the two others are discussed in the proof of (II.3.8)(see (II.3.161). Now we claim that
gal∋g→f(g)⊂AW
induces a bijection between gal and the cosets of the subgroups of AW mentioned
in (i). For, if G1∈g,g∈galCF,
then the elements of g are equivalent to galleries of the form G1G2
with G2∈GalF. After dividing
out the action of W, we can suppose F⊂C‾,
then f(GalF)=AWJ,
where J⫋{1,…,ℓ} is of the type
F, i.e. J={j|ϕj(F)={0}}.WJ is finite because F⊂I∘ and
J is a proper subset, because F∈F*(ℳ).
Now f(g) is the coset
f(G1)AWJ.
Conversely, if WJ is finite, but
J≠{1,…,ℓ} and x∈AW,
it is easy to see that f-1(xAWJ)∈galCF,
for some proper facet F determined by π(x)∈W
and WJ. (i) now follows from corollary (4.11). (ii) follows from (4.16).
□
(5.10) Definition. Let AW∼ be an extended Artin group of rank
ℓ as in definition (III.2.13). For J⊂{1,…,ℓ}
we define the following subgroups:
AW∼J≔the group generated bysj,tj,j∈J.π-1(W)J≔the group generated bysj(n),j∈Jandn∈ℤ.
(5.11) Remarks. (i) We have π-1(W)J⊂π-1(WJ),
but the other inclusion need not hold (take for instance J=∅). What we do have is:
π-1(W)J=(π|AW∼J)-1(WJ)=AW∼Jπ-1(WJ).
(ii) Analogous to theorem (II.4.13) it can be proved that the flip relations form a complete set of relations for
π-1(W)J and also:
(5.12)π-1(W)J1∩π-1(W)J2=π-1(W)J1∩J2.
(iii) From (ii) it can be seen that (as the notation already suggests)
(AW∼J,{sj,tj}j∈J)
forms an extended Artin system and we have
AW∼J1∩AW∼J2=AW∼J1∩J2.
(5.13) Theorem.
Suppose a hyperplane system of an extended Coxeter group of rank ℓ is given. Let AW∼
be the associated extended Artin group of ℓ. Then:
(i)
The simplicial complex Σ of the hyperplane system is isomorphic to the cosetnerf of the
collection subgroups of π-1(W)
of the form π-1(W)J
where J⫋{1,…,ℓ} and
WJ is finite. This collection satisfies (5.6)(i) and (ii).
(ii)
The system is central if and only if W is finite and in this case there is a bijection
m ("midpoints") from AW∼ to the vertices of
Σ^ which are not vertices of
Σ. If
x·eq∈AW∼ is
the unique way to write an element of AW∼ as product of
x∈π-1(W) and
q∈Q, then the maximal simplices of the subcomplex
Σ(m(xeq))
are the simplices of the form
σ(xλq(w)),w∈W
where W has to be interpreted as a subset of
π-1(W), which can be
done by remark (II.3.19) and lemma (III.4.1) and λ is the pseudo conjugation the action of
Q on the set π-1(W)
as defined in (III.4.14).
Proof.
(i) is proved in a similar way as the corresponding statements of (5.9). Note that (5.6)(i) is satisfied by (5.12). It is in principle possible to give a bijection
between AW∼ and
galCC×K(pℳ)
and prove (ii) with the help of (4.15). But let us give a more direct and instructive proof: Let
V∈𝒱(A) be a vertex of
Σ^, which is not a vertex of
Σ. Here
A=w(q+U0)∈K(pℳ)
for unique w∈W,q∈Q and U0 as in the proof of (III.3.5)(iii).
Figure 29.
A maximal simplex of
Σ is determined by a U∈𝒰(n),D=w′(C)∈K(ℳ)
for a unique w′∈W. This simplex forms with
V a simplex in Σ^ (i.e. is a simplex of
Σ(V)) iff
U∩V≠∅. We may assume, that there is a point in this inter¬section,
which projects (by Y∼→Y) on
a+-1d∈Re-1(A)∩Im-1(D)⊂Y,
where a≔w(q+c)∈w(q+U0)=A
and d≔w′(c)∈D
(recall that c∈U0⊂C). Let
γ∼ be a curve in Y∼ from
c∼ to this point in U∩V. Then
γ≔p∘γ∼ is a curve in Y
from -1c to a+-1d.
Define the "straight" curves λ1,…,λ4
as follows
Note that λ2⊂Re-1(A)
and λ3,λ4⊂Im-1(D).
Now the curve γλ2-1λ1-1
projects on a loop in X representing an element in AW∼=π(X,*)
of the form x·eq,x∈π-1(W),q∈Q as above. Conversely each element in AW∼
determines a curve in Y from -1c to a point of the form
w(q+-1c).
Composing this curve with λ1 and lifting to Y∼ the endpoint lies
in a V∈𝒱. This way we get the desired bijection
m:AW∼→𝒱.
The curve γλ3λ4 projects on a loop representing
an element y of π-1(W),
such that π(y)=w′.
This describes the bijection between π-1(W)
and the maximal simplices of Σ given by (i).
Since
γλ3λ4≃γλ2-1λ1-1λ1λ2λ3λ4
we have
($)y=x·eq·z,
for some z∈AW∼, represented by the projection of
λ1λ2λ3λ4.
To analyse this element z, we note that λ1λ2λ3
projects on the same loop in X as λ1′λ2′λ3′,
where
the transforms of λi under
e-q·w-1∈W∼.
Inspection of these formulas gives that λ1′λ2′λ3′
projects on a loop representing the element w-1w′∈W⊂π-1(W).
The curve λ4 projects on the same loop as its transform under
(w′)-1·e-w(q)∈W∼:[-1c,(w′)-1(-w(q))+-1c],
so it projects on a loop representing
(w′)-1(w(-q))∈Q⊂AW∼.
we conclude that
z=w-1w′·e(w′)-1w(-q).
The reasoning has shown that if σ(y) is a simplex of
Σ(m(x·eq)),
then y=x·eq·z=x·eq·w′′·e(w′′)-1(-q)≕x·λq(w′′),
where w′′≔w-1w′.
Conversely, it can be checked that σ(x·λq(w))
is a simplex of Σ(m(x·eq))
for all x∈π-1(W),q∈Q and w∈W.
□
(5.14) Remark.
In the situation of a linear Coxeter group as well the situation of an extended Coxeter group the simplicial complex
Σ^min can be described by adding to the
conditions on J in (5.9)(i) (respectively (5.13)(i), namely that
J⫋{1,…,ℓ} and
WJ finite, that J is maximal with respect to these properties.
(5.15) Remark. In the case that W is finite the sets
J⊂{1,…,ℓ} with the properties mentioned
in (5.14) are precisely the sets J with |J|=ℓ-1,
hence of the form {1,…,ℓ}-{j}
for a j∈{1,…,ℓ}.
Hence for a given x∈AW (respectively
x·eq∈AW∼)
we have a bijection between the maximal simplices of the Coxeter complex of W (example (5.8)) and the maximal simplices of
Σmin(m(x))
(respectively Σmin(m(x·eq))).
This bijection induces a homeomorphism between the ℓ-1-sphere and
|Σmin(m(x))|
(respectively |Σmin(m(x·eq))|).
§6. The Contractibility Problem
If the space |Σ^| is
contractible and the system is locally-K(π,1),
then the system is K(π,1), as follows from theorem (3.11). To
investigate the question of contractibility of |Σ^|
we construct a space Z^=Z^(𝕍,H,ℳ,p)
in terms of facets and galleries, which turns out to be homotopy equivalent to |Σ^|, as follows:
(6.1) Definition.
Let S≔H-Fmin.
Note that in the non central case S=H and in the central case
S=𝕍-⋂M∈ℳM,
which has the homotopy type of a sphere, hence the use of the symbol S.
(6.2) Definition. For F∈F*(ℳ)
we define F‾^≔F‾-Fmin
(hence, by (3.5) ≔F‾ in the non central case).
We give Z‾g′ the topology of
end(g)‾^⊂𝕍
and we give Z′ the topology of the disjoint union of the spaces
Z‾g′ (Note that the set
Z′ is a disjoint union of the setsZ‾g′,g∈gal). Furthermore for
g∈gal:Zg′≔end(g)×{g}.
Note that this fits in with the notation, because Z‾g′
is indeed the closure of Zg′ in Z′.
Now we come to the definition of the space Z:
(6.4) Definition. We define an equivalence relation on Z′ as follows: If
(xi,gi)∈Z′
for i=1,2, then
(x1,g1)∼(x2,g2)
if x1=x2 and if
F⊂end(gi)‾
is the facet containing xi and gi′ the
unique element of galCF (see (4.5)(ii)) such that
gi′≤gi, then
g1′=g2′.
Let Z denotes the quotient space Z′/∼ and
b:Z′→Z the projection.
(6.5) Remark. From the definition it follows that each equivalence class contains a unique element
(x,g) with the property that
x∈end(g). So
Z=⋃g∈galZg,
where Zg≔b(Zg′).
This is a disjoint union. The sets Zg are the facets of the space Z.
(6.6) Remark. To clarify the picture of the space Z (see also fig. 30) we note the following: Since each point in
S is contained in A‾ for some chamber A we could
have restricted ourselves to elements g such that end(g)∈K(ℳ)
(and g is then nothing but an equivalence class of galleries). I.e. Z can also be considered as a quotient space
⋃g∈galCA,A∈K(ℳ)Z‾g′.
H=𝕍=ℝandℳ={0,1}Figure 30.
(6.7) Definition. Let g∈gal. We define the
star of the facet Zg (and denote it by Str to avoid confusion with the notations St and Star) by:
Str(g)≔⋃g≤hZh
if the l.h.s. is nonempty, because the r.h.s. of (6.9) is open in Z‾k′.
Let x∈end(k)‾^
and b(x,k)∈Str(g).
By the definition of Str(g) (6.7) we have
b(x,k)∈Zh for some
h∈gal with g≤h. Then
(x,k)∼(x,h) and
x∈end(h), which entails
h≤k. So first of all we see that, if the l.h.s. of (6.9) is nonempty, then
g≤k. Furthermore from g≤h it follows that
end(g)⊂end(h)‾,
so x∈St(end(g)),
which proves that the l.h.s. is contained in the r.h.s.
Now we assume that the l.h.s. is nonempty (hence as we saw above we know that g≤k) and prove
that the r.h.s. is contained in the l.h.s.: Suppose (x,k)∈Z‾k′ and
x∈St(end(g)).
Then x∈F for some facet F with
end(g)⊂F‾.
Furthermore F⊂end(k)‾,
so there is a unique h∈galCF,
such that h≤k.end(g)⊂F‾,
so there is a unique element g′∈galC,end(g)
such that g′≤h. So
g′≤k, but we also know
that g≤k, so g=g′
and hence g≤h. From
b(x,k)=b(x,h)∈Zh
and g≤h, we conclude that
b(x,k)∈Str(g).
Hence (x,k) is element of the r.h.s. of (6.9).
□
(6.10) Proposition. Let g∈gal. Then:
(i)
b(Z‾g′)=⋃h≤gZh
(ii)
b(Z‾g′)is closed in Z, henceZ‾g=b(Z‾g′).
Proof.
(i) is clear from def. (6.4) and remark (6.5).
(ii) Suppose k∈gal is cuh that
Zk∩b(Z‾g′)=∅,
then by (i): k≰g. We show that
Str(k)∩b(Z‾g′)=∅,
which proves (ii): Suppose Zk⊂Str(k)∩b(Z‾g′),
then k≤h and h≤g,
hence k≤g, contradiction.
□
(6.11) Remark. We have Z=⋃g∈galZ‾g
and a map defined on Z is continuous, if and only if the restrictions to
Z‾g,g∈gal
are continuous. If we wish we even can restrict ourselves to g∈galCA with
A∈K(ℳ).
(6.12) Proposition. Let g1,g2∈gal.
Then Str(g1)∩Str(g2)
is nonempty if and only if g1∨g2 exists and then it equals
Str(g1∨g2).
Proof.
If nonempty, then there is a Zg contained in the intersection and then
gi≤gi=1,2.
So then g1∨g2 exists. And conversely.
Str(g1)∩Str(g2)=⋃gi≤hZh=⋃g1∨g2≤hZh≕Str(g1∨g2).
□
(6.13) Proposition.
Let g∈gal. Then
Str(g) is contractible.
Proof.
Choose a point x0∈end(g).
Then b(x0,g)∈Zg⊂Str(g)
and we are going to contract Str(g) on
b(x0,g). Let
k∈gal and suppose, that the left hand side of (6.9) is nonempty (i.e.
g≤k). By the equality (6.9) we see that we can define a continuous map
ϕk:A×[0,1]→A
by ϕk((x,k),t)≔(tx0+(1-t)x,k),
where A is the set in (6.9). It is not difficult to check, that
(x1,k1)∼(x2,k2)
then
ϕk1((x1,k1),t)∼(ϕk((x2,k2),t)).
So we have a continuous map ϕ:Str(g)×[0,1]→Str(g),
which is the desired contraction on b(x0,g).
□
(6.14) Theorem. The space Z is homotopy equivalent to the space
|Σ|.
Proof.
It follows from (6.8) and (6.12) that {Str(g)}g∈gal
is an open cover of Z, whose nerf is isomorphic to Σ.
This cover is simple by (6.13) and (6.12), so the theorem follows.
□
(6.15) Discussion.
We have a map q:Z→S, sending the class of
(x,g) to x. Suppose we have a central system.
For A∈K(pℳ) and
g∈galCC we define a section
s=s(g,A):S→Z
of q as follows:
If x∈S, then
x∈D‾^ for some
D∈K(ℳ). Now let
k∈galCD be the class of
G·Gd, where G∈g
and Gd is a direct replacing gallery of the direct pregallery
CED with
E≔pr(p(C,D)⊂pℳ)(A).
Then s(x)≔b(x,k),
We have to check that this does not depend on the choices involved. We sketch this check for the choice of another
D′∈K(ℳ),
such that x∈D‾^′.
Let F⊂D‾∩D‾′
be the facet such that x∈F. Then, as we saw in (I.3.9) and (I.3.10) the map
prF restricted to {B∈K(ℳ)|F⊂B‾}
is a bijection on K(ℳF). So we can take
B≔prF-1(prF(C)).
Then a direct Tits gallery from C to B, followed by a direct Tits gallery from
B to D, is a direct Tits gallery from C to D.
So Gd can be chosen of the form
G1G2,G2∈GalF
and Gd′=G1G2′,G2′∈GalF.
Then Gd and Gd′ are
F-equivalent.
Figure 31.
The map s is continuous, because it is so when one restricts it to a set D‾^
and S is a finite union of such in S closed sets. Furthermore it is clear that
q∘s=idS and
s∘q|S(g,A)=idS(g,A),
where S(g,A)≔s(g,A)(S).
So we conclude from this discussion, that we have a collection of spheres in Z in the central case, labelled by
galCC×K(pℳ).
Now the hope is that we can prove in some cases that Z is just a bouquet of these spheres. I.e. if one attach cells to these spheres one
gets a contractible space. Indeed, constructing the space Z^, we are going to attach
for each (g,A)∈galCC×K(pℳ)
a cell B(g,A). Now, if it could happen that
S(g1,A1)=S(g2,A2),
with (g1,A2)≠(g2,A2),
then the contractibility of Z^ would be doubtful. But we have:
(6.16) Lemma. If
S(g1,A1)=S(g2,A2),
then (g1,A1)=(g2,A2).
Proof (sketch).
Let x∈C, then
s(g1,A1)(x)=s(g2,A2)(x),
so (x,g1)∼(x,g2),
which implies g1=g2. From
s(g1,A1)(-x)=s(g2,A2)(-x)
it follows that G·CA1(-C)∼RG·CA2(-C),
where G∈g1=g2. So
CA1(-C)∼RCA2(-C),
from which it can be proved that A1=A2, by a simple application
of an invariant technique, which we do not want to develop here.
□
(6.17) Definition. We define a space
Z^=Z^(𝕍,H,ℳ,p)
as follows: If the system is non central then Z^≔Z.
If the system is central, then we attach a cell B(g,A) to each sphere
S(g,A) of Z.
Precisely: Z^ is a quotient of
Z∪(𝕍×(galCC×K(pℳ))),
where points of S(g,A) are identified with points of
𝕍×{(g,A)} by
S(g,A)→q≅S⊂𝕍≅𝕍×{(g,A)}
Note that we can extend q:Z^→𝕍 in the central case, so we have a projection
q:Z^→H in both cases.
(6.18) Theorem. The space Z^ is homotopy equivalent to
the space |Σ^|.
Proof.
For the non central case this is theorem (6.14). So suppose central. Then: The in Z open sets
Str(g) for g∈gal
induce open sets in Z^ This follows from the fact that
q(Str(g))=St(end(g))
is open in 𝕍. These open sets still do not cover Z^.
But for each (g,A)∈galCC×K(pℳ)𝕍×{(g,A)} induces an open set in
Z^ and together with the first sets this gives an open cover of
Z^. It can be verified, that this cover is simple and that its nerf is
isomorphic to Σ^.
□
(6.19) Corollary. If the system is locally-K(π,1),
then the universal covering space Y∼ is homotopy equivalent to Z^.
(6.20) Remark. It can be proved that the homotopy equivalence of (6.19) can be realised by the following map
ψ:Y∼→Z^,
which makes the diagram shown commutative. Let y∈Y∼ and set
x≔Im(p(y))∈H.
If x∈Fmin, then
y∈V∈𝒱 and ψ(y)
is the point in Z^ determined by x and
v-1(V)∈galCC×K(pℳ).
If x∈S, then x∈F for an
F∈F*(ℳ).
Then y∈U∈𝒰(F) and
ψ(y) is the point in Z^
determined by b(x,u-1(U))∈Z.
(Recall the maps u and v from (4.9)).
(6.21) Remark. The spaces |Σ(g,A)|
are homotopy equivalent to a sphere, since they can be seen to correspond to the spheres S(g,A) by the homotopy equivalence of theorem (6.14).
Y∼⟶ψZ^p↓Y↙qIm↓H
(6.22) Remark. If the system is simplicial and if ⋂M∈ℳM={0},
then the hyperplanes cut out a triangulation of a sphere around the origin. Deligne uses this triangulation to define his complex (which coincides with
|Σ^min|).
But the use of the simpliciality condition for this purpose is not the essential use as we can see from the constructions of this chapter. The geometric simplicial
complex on the sphere is exactly |Σmin(g,A)|
(cf. (5.14) and (5.15)).
Now we are going to define some conditions on the galleries, which seem to be satisfied in many cases and which are sufficient to prove that Z
is contractible in the non central case (for a simple system it follows then for a central system too).
(6.23) Definition. Let G∈GalCF.
Recall (I.4.1) that |G| is the length of G.
The gallery G is called F-minimal, if
|G|≤|G′|
for all G′ in the F-equivalence class of
G. A gallery G is called minimal, if it is
end(G)-minimal, i.e. if its length is minimal for the gallery equivalence
class to which it belongs. Note that F-minimality implies minimality, but in general not conversely.
If g∈galCF, we define the
length of g, notation also |g|
to be the length of an F-minimal representative of g.
(6.24) Definition. Let G,G′∈Gal.G is said to be left (resp. right) contained in G′,
if G′≈G·G′′
(resp. G′≈G′′·G)
for some G′′∈Gal.
Recall that ≈ denotes flip equivalence (I.4.5).
(6.25) Definition. Let g be an equivalence class of galleries. Then a gallery G
is called a direct starting (resp. ending) piece of g, if G
is direct and left (resp. right) contained in some minimal representative of g.
(6.26) Example. The galleries 1 2 and 3_ 2 are direct starting pieces of the
gallery class that is depicted in figure 12 (example (I.5.5)). The galleries 2 1_ and 2 3
are the direct ending pieces.
(6.27) Example.
Let us consider the configuration of figure 32. The galleries G1,…,G6,
starting in C, where
(Example is due to the computer) are all the minimal representatives of a gallery class g.
Figure 32.
We see that the (maximal) direct starting nieces of g are: 1_
2 5, 1_3_
5 (≈3_1_5) and
3_ 4 5. The (maximal) direct ending pieces: 5 2_ 1,
5 3 1 (≈513) and 5
4_ 3, ending in chamber C. This example is interesting
because this equivalence class splits up in three flip equivalence classes:
{G1},{G2,G3,G4,G5}
and {G6}.
(6.28) Definition. A gallery class g (with A≔beg(g))
is said to satisfy the direct starting pieces property (DSPP), if there exists a direct gallery Gd,
such that:
(i)
Each direct starting piece of g is left contained in Gd.
(ii)
For each M∈ℳA∩ℳ(A,end(Gd))
there is a minimal representative of g, which starts through M, i.e. which
Deligne gallery is of the form A,spM∩A‾(A),… (definition of spF: (I.3.11)).
The direct ending pieces property (DEPP) is defined analogously. It is easily seen that all the gallery classes of a hyperplane system satisfy DSPP
if and only if all the classes satisfy DEPP and in this case the hyperplane system is said to satisfy the direct pieces property (DPP).
(6.29) Example. For the examples (6.26) and (6.27) the classes mentioned have both properties:
Gd=123_≈3_21
for (6.26) and for example (6.27): Gd=1_253_4
for the direct starting pieces (note that Gd is direct, because the underlying Tits gallery is so and chamber
A prescribes the signs). And Gd=4_352_1
for the direct ending pieces (now chamber B prescribes the signs of Gd).
(6.30) Definition. A hyperplane system is said to satisfy the facet minimality property, if for each
F∈F(ℳ) and
G1∈GalCF,G1 is F-minimal and
G2∈GalF,G2 is minimal, then G1G2 is minimal
(if defined).
(6.31) Theorem. Let (𝕍,H,ℳ,p)
be a non central hyperplane system, which satisfies the direct pieces property and the facet minimality property. Then
Z(𝕍,H,ℳ,p) is contractible.
For the proof we need a simple lemma, which tells us how we can project the closure of a chamber on its boundary in the direction of another chamber:
Then E⊂∂A is a strong deformation retract of
A‾.
Proof.
Choose b∈B. for
x∈A‾,A‾∩[b,x]
is of the form [d(x),x] for some unique
d(x)∈A‾.
We prove that d(A‾)=E.
First remark that,
Let x∈A‾. If there is a
M∈ℳA∩ℳ(A,B) with
x∈M, then
[b,x]∩DM(A)‾={x},
for b∉DM(A,B)‾.
So then x∈d(A‾).
Conversely, if there is no such M, then for all M∈ℳA we
have [b,x]∩DM(A)‾≠{x}.
So then [b,x]∩A‾=⋂M∈ℳA([b,x]∩DM(A)‾)≠{x}.
This shows that d(A‾)=E.
The continuity of d is now also clear. By the convexity of A‾,
we see that A‾×[0,1]∋(x,t)→(1-t)x+t·d(x)∈A‾
gives a deformation retraction of A‾ on E.
□
Proof of theorem (6.31).
We construct a filtration of Z: For n∈ℤ+
we define zn≔ union of all
Z‾g, where
g∈galCA,A∈K(ℳ) and
|g|≤n. Then
C‾≅Z0⊂Z1⊂Z2⊂…⊂Zn⊂Zn+1⊂…⊂Z
and Z=⋃n∈ℤ+Zn,
so it suffices to show that Zn is a (strong) deformation retract of
Zn+1.
Therefore we define a continuous map κt:Zn+1→Zn+1,t∈[0,1],
which is a (strong) deformation retraction of Zn+1 on
Zn. Let us define this first on Z‾g
for a g∈galCA,A∈K(ℳ) and
|g|≤n+1. If
|g|≤n, then we define
κt to be the identity on Z‾g.
Now let |g|=n+1.
Let Gd be the direct gallery, such that all the direct ending pieces of g are right contained in
Gd (so in particular end(Gd)=A).
Let B≔beg(Gd).
Then A≠B, for |g|>0.
Since the system is non central, we have A‾=A‾^,
so Z‾g′=A‾×{g}≅A‾.
So we can define κt on Z‾g
with the help of the deformation retraction of lemma (6.32). Let us check that κt is now well defined on
Zn+1 by proving that the definitions coincide on
Z‾g∩Z‾h.
There is only something to prove in the case that at least one, say g, of the gallery classes has length
n+1. So |g|=n+1
and we use the same notations for g and its attributes as above. Suppose
h∈galCD,D∈K(ℳ).Z‾g∩Z‾h≠∅
implies that A‾∩D‾≠∅,
so this intersection equals F‾, where
F≔A∧D (2.7). Now there is a unique
k∈galCF such that
k≤g and k≤h. Choose an
F-minimal representative G1 of k and let
G2∈GalF be a minimal gallery such that
G1G2 is a representative of g.
Let also G3∈GalF be minimal, such that
G1G3 is a representative of h.
It follows from the facet minimality property, that G1G2 and
G1G3 are minimal, in particular
|G1G2|=n+1.
Now we may suppose that |G2|≥1,
for otherwise |G1|=n+1,
hence |G3|=0,
so then A=D=F and g=h and there is nothing to prove,
go we can write G2=G2′G2′′
with |G2′′|=1
and let G2′′ pass through the wall
M∈ℳA. Since G2′′
is a direct ending piece of g, we have M∈ℳA∩ℳ(A,B)
by (DEPP). Now G2∈GalF, so
G2′′ also and hence
M∈ℳF. So
F⊂E, where E is as in lemma (6.32). It follows that
κt restricted to the intersection Z‾g∩Z‾h
is the identity for each t∈[0,1], in particular
coincidence of the definition on Z‾g and on
Z‾h So κt is veil defined
and it is continuous by remark (6.11).
We still have to prove κ1(Zn+1)⊂Zn.
Let g and its attributes be again as above and suppose x∈A‾.
We have to show, that κ1(b(x,g)),
which equals b(d(x),g)
by definition, is contained in Zn, where
d:A‾→A‾
is the map constructed in the proof of lemma (6.32). By this lemma d(x)∈M
for some M∈ℳA∩ℳ(A,B),
so by (DEPP)(ii), we know that g has a minimal representative of the form
G1G2, where
|G2|=1 and
G2 passes through M. So
d(x) is element of the closure F‾
of a common face F of A and end(G1).
The pair (G1,F) determines an element
h∈galCF and G1
and G1G2 are F-equivalent, so
(d(x),h)∼(d(x),g)
and since |h|=n, it follows that
b(d(x),h), hence
also b(d(x),g) is contained in
Zn.
□
(6.33) Corollary. Let we have a system as in theorem (6.31). Suppose the system is also
locally-K(π,1).
Then the system is K(π,1).
In the simple case there is no need to restrict ourselves to non central systems:
(6.34) Theorem. Let a simple hyperplane system be given, which satisfies the direct pieces property and the facet
minimality property. If the system is locally-K(π,1),
then it is K(π,1).
Proof.
In the noncentral case this is corollary (6.33). So suppose the system is central. Choose a hyperplane N∈ℳ and a form
ϕ∈𝕍*, such that
N=ϕ-1(0).
Let H≔ϕ-1((0,∞)).
Consider ϕ as an element of (𝕍ℂ)*.
Let Y′≔ϕ-1(1)∩Y.
Then it is easily seen that Y′ is a strong deformation retract of
Y′′≔Y(ℂ,H,ℳ-{N}).
It is also not difficult to see, that ℂ*×Y′∋(t,z)→tz∈Y
defines a homeomorphism. So the system (𝕍,𝕍,ℳ) is
K(π,1) if and only if
(𝕍,H,ℳ-{N}) is
K(π,1). And the latter is
K(π,1) by (6.33).
□
§7. The Direct Piece Property; Conjectures
(7.1) Conjecture. A simplicial system has the direct pieces property and the facet minimality property.
If this conjecture is true, at least for simple systems one can prove conjecture (1.3). This proof is independent of the result of Deligne (1.2), because to prove
the locally-K(π,1)-property
one can apply induction on the codimension of ⋂M∈ℳM
in 𝕍.
In this paragraph we want to give a sketch of a possible proof of the conjectured implication: simpliciality ⇒ direct pieces property.
First we remark that it is only necessary to prove (6.28)(i), because in the simplicial case it can be proved that (ii) is satisfied if one takes the length of
Gd of (i) minimal. Now we need a definition:
(7.2) Definition. We say that a gallery Greduces to a gallery
G′, notation
G′∠G,
if we can obtain G′ from G by length reducing cancel operations.
If we can take these cancel operations as follows:
(Gdm·G1·AUBUA·G2,Gdm·G1·G2),
where Gdm is the maximal direct starting piece of the second gallery, we say that Greduces toG′respecting the direct starting pieces, notation
G′∠!G.
If gfl is a flip class of galleries, then
G′∠gfl
(resp. G′∠!gfl)
means that there is a G∈gfl, such that
G′∠G (resp.
G′∠!G).
Now by very general reasoning it can be seen that if {G1,…,Gk}
is a finite collection of representatives of a gallery class g, then there exists a flip class
gfl⊂g such that
Gi∠gfl for
i=1,…,k. In
particular there exists such a class for all the minimal representatives of g. Now (6.28) would be
proved if we have:
(7.3) Conjecture. By the minimality of the representatives it follows that we can choose gfl
such that we even have Gi∠!gfl for all the minimal
representatives Gi of g.
and:
(7.4) Proposition. (6.28) is true for flip equivalence.
Proof.
This can be proved analogously as the key lemma of Deligne [Del1972] 1.19.
□
Here is the essential use of the simpliciality.
It is interesting to ask if a somewhat stronger form of the direct pieces property is true, namely if we replace the word "minimal" by "irreducible", where:
(7.5) Definition. We call a gallery irreducible, if it is not flip equivalent to a gallery, that contains a cancellation, i.e. that is of the form
G1·AUBUA·G2.
Note that minimal implies irreducible, but not conversely:
(7.6) Example. Consider G≔1221_1_2_
in figure 12, page 40. Then no flip operations nor length reducing cancel operations are possible, so G is irreducible. But
G∼3_2231_2_≈3_222_1_3∼3_21_3,
so G is not minimal.
If we not only replace "minimal" by "irreducible", but require an additional condition (which is satisfied in all the examples we have considered so far)
we get the following:
(7.7) Definition. A hyperplane system as said to satisfy the strong direct pieces property, if for each gallery class
g, there exists a direct gallery Gd,
such that direct starting pieces of irreducible representatives of g are left contained in Gd
and moreover if gi⊂g for
i=1,2 is a flip class of irreducible galleries and Gi
is the maximal direct starting piece of gi, then
|G1|+|G2|>|Gd|.
(7.8) Theorem. If the hyperplane system satisfies the strong direct pieces property, then we have the following
solution to the word problem of the galleries: Let G∈Gal and
beg(A)≕A. If
G∼∅A, then G is reducible.
Proof.
If G is irreducible, then G=G1·G2-1
for two irreducible and equivalent galleries G1 and G2.
Using the second condition in def.(7.7), we get a contradiction. The details are left to the reader.
□
If simplicial systems would satisfy the strong direct pieces property, then theorem (7.8) would give a very simple solution to the word problem for Artin groups.
Note that for braid groups it would mean the following: "If a braid is trivial, then this can be shown without increasing the number of crossings".
Notes and References
This is an excerpt of a thesis entitled The Homotopy Type of Complex Hyperplane Complements, written by Harm van der Lek in 1983.