In this previous post (about J.-C. Yoccoz 2011-2012 course at College de France), we reviewed the distinct points of view on origamis, and we started the discussion of the action of the automorphism group on the first absolute homology group of the origami. We quickly recall the main points of the previous post for the reader’s convenience.
Given an origami , we saw how to canonically describe it (through its monodromy group) by a finite group
generated by two elements
and
, and a choice of subgroup
such that
, so that one has
. Also,
and
act on the set of squares of
by
and
. Furthermore, denoting by
the normalizer of
in
, one has
acting as
, for
.
Next, given a subfield , we decomposed
and we reduced the problem of studying the action of
on
to understanding the action of
on
, and, to do so, we considered the set
, and we proved that
as -module. In this way, since
and
, we were led to the study of the action of
on
. At this stage, we noticed that
is the set of orbits of
acting to the right on
or equivalently the set of orbits of
acting on
by
.
Nevertheless, we introduced the following notation: is the point of
associated to
,
is the stabilizer of
for the action of
on
, and
is the order of the commutator
.
Finally, if we denote by the smallest integer such that
, we showed (in the very end of the previous post) the following lemma:
Lemma. .
After this quick revision, we will enter (below the fold) the content of Yoccoz’s 2nd lecture (on January 18, 2012).
Analogously to the definition of , we denote by
the smallest integer such that
. Observe that, by construction,
divides
, and
divides
. For later use, in the next proposition we collect a series of elementary facts:
Proposition. It holds:
is a cyclic group of order
;
is a cyclic group of order
;
- the orbit
of
in
has cardinality
;
;
.
Proof. The first two items follow from the definitions of and
. The third item follows from the second one (as the size of an orbit is the order of the group divided by the size of the stabilizer and this last quantity is given by item 2.). The fourth item is a matter of counting correctly the involved objects:
, so that the natural “reflex” is to count
through
and
. However, we should be slightly careful here because the representation
is not unique. To solve this (easy) problem we observe that two representations
verify
, so that we have an unique representation
once the condition
, and hence the fourth item follows. Finally, the fifth item is a direct consequence of the third and fourth items.
Following the “general plan” mentioned in the previous post, we will dedicate today’s discussion to the study of the action of on
in the case
, i.e.,
Standing Assumption. In the sequel, .
Recall that , so that it suffices to understand the action of
on each piece.
For this reason, it is natural to introduce the character of the representation of
(or
) on
, and
the character of the representation of
on
.
By the fifth item of the proposition above,
On the other hand, since is the induced representation of the trivial representation of
, one has
for . Here,
is the characteristic function of
.
Notation. We denote by the class determined by an element
.
By the 2nd item of the proposition above,
Now, we consider a character of an irreducible representation (over
) of
, and we denote by
the multiplicity of
in
.
At this point, we observe that the action of on
is “determined” by the knowledge of
: indeed, morally speaking, the representation theory of finite groups (such as
) is “well-known” in the sense that irreducible characters
are “determined”, at least when
is a “classical” finite group (such as symmetric groups); therefore, the computation of
permits (in principle) to calculate
from character tables of finite groups.
In any event, recall that the multiplicity is given by
By using the formula of above, we can expand the right-hand side to get
where is the usual Kronecker’s delta.
Since is the character of a
-representation, one has
, and, hence,
It follows that
On the other hand, . By the second item of the proposition above,
where
, so that
Now, denoting by the degree of
, we note that if
are the eigenvalues of
, then
are the eigenvalues of
. On the other hand,
, so that
and, hence,
In particular
where is the subspace of fixed vectors under the action of
(a subspace of dimension
, by definition).
By putting these identities together, we obtain the following statement:
Proposition. The multiplicity of
in
is
Next, we observe that is the sum of
copies of the regular representation of
. Since the regular representation contains (all) irreducible representations with a multiplicity equal to its degree, we have that the multiplicity of
in
is
.
In particular, the multiplicity of
in
is
. By the previous proposition, we can write
Observe that the right-hand side of this identity depends only on the class : indeed,
,
, implies
and, a fortiori,
, i.e., the elements
and
are conjugated by an element
. Therefore, we can rewrite
Now we proceed to rewrite this formula by using induced representations. More precisely, we think of the irreducible representation of
as a representation of
of
on a vector space
, and we denote by
the induced representation of
on
, where
are representants of the classes
and
. Denote by
the permutation of
determined by the action of
, i.e.,
. In this notation, we have, by definition,
.
Observe that is precisely the length of the cycle
of
containing
, and the restriction of
to
is conjugated to the action of
on
. Thus, we get the following statement:
Theorem. The multiplicity is given by the formula
where is the dimension of the subspace of
fixed by
.
Even though this statement is a simple consequence of our discussion so far, we decided to call it a “theorem” because it has some interesting consequences. In fact, the rest of today’s post will be dedicated to the proof of the following “corollary” to this theorem:
Corollary 1. For any origami , and for any
, one has
.
We begin the discussion of the proof of the previous statement by showing the following result:
Corollary 2. The multiplicity is always greater than or equal to the multiplicity
of the trivial representation.
Proof. By the theorem, . Since
and
for every
, we conclude
so that the argument is complete.
Let us consider now the case of regular origamis (in the sense of D. Zmiaikou), i.e., ,
. In this context, the formulas for the multiplicities simplify to
and
.
Proposition. In the case of regular origamis:
- If
then
;
- If
then
.
Proof. If , we have that
is a homomorphism and hence
as
is a commutator. Therefore,
in this case.
If , we have that
, where
denotes the representation with character
. We claim that
. Otherwise,
, and hence the action of
factorizes into the action of a commutative group (as
and
generate
). However, this would force
(since commutative groups only have
-dimensional irreducible representations), a contradiction proving our claim. Now, we observe that the claim and the fact that
imply that
has at least two eigenvalues
, so that
.
Partly motivated by this discussion, we introduce the following definition:
Definition. An origami is quasi-regular if the multiplicity
of the trivial representation is zero.
Remark. By the previous proposition, any regular origami is quasi-regular (so that the nomenclature is coherent).
The following proposition allows to characterize quasi-regular origamis in terms of the commutator :
Proposition 1. is quasi-regular
the normal subgroup generated by
is contained in
.
Proof. We know that . Thus,
for all
cycle
acts trivially on
.
Proposition 2. If is not quasi-regular, then
for all
.
Proof. By Corollary 2 above, it suffices to check that . Since
and
is not quasi-regular (i.e.,
), we have that the permutation
(associated to the action of
) is not the identity. On the other hand, since
is a commutator, the permutation
is even. In particular,
is not a transposition, and, consequently,
.
Before proceeding further, let us give an example of a quasi-regular but not regular origami.
Example. Let
be a finite Heisenberg group. Observe that
is generated by the two elements
We choose
. Note that
because
Hence, this data defines an origami
. Since
,
is not a regular origami. On the other hand, the normalizer
of
is
is a normal subgroup (coinciding with the centralizer of
). Note also that
By Proposition 1,
is a quasi-regular origami. However, this is not the most interesting example of quasi-regular origami (from the representation theory point of view) because
is an Abelian group. In any case, this quasi-regular example shows two interesting features:
is normal and
. As we are going to see below, this is a “general phenomenon” for quasi-regular origamis.
Proposition. is quasi-regular if and only if
is normal and
is Abelian generated by two elements
(i.e.,
is either cyclic group or a product of two cyclic groups).
Proof. If is normal and
is Abelian, then the commutator
is mapped into the identity under the natural map
, and, by Proposition 1,
is quasi-regular.
Conversely, if is quasi-regular, then
. Thus,
is generated by two elements
satisfying
inside
, i.e.,
is an Abelian group (generated by two elements). Since
is a subgroup of the Abelian group
, it follows that
is normal and
is Abelian.
At this point, we are ready to complete today’s discussion by proving Corollary 1. We start by noticing that, by Proposition 2, it suffices to consider the case of quasi-regular origamis. The idea of the proof in this case is similar to the one in the case of regular origamis (see the proposition before Proposition 1), despite the fact that we need the following little trick.
Let be a quasi-regular origami and denote by
and
the orders of
and
in
. Define
and
.
Note that is a complete system of representatives of
. Define
, for
,
, so that
.
Observe that
and
so that is the product of
,
,
.
In this language, the Theorem implies that
Proposition. If is a quasi-regular origami, then
.
On the other hand, since is a commutator of elements of
, we obtain
We have the following possibilities:
for all
: here either
for all
and, a fortiori,
, or
for some
, so that
(by the condition on the determinant);
for at least two pairs of indices
: in this situation,
for
and, a fortiori,
.
In any event, we find that . This completes the proof of Corollary 1.
Leave a Reply