1. Introduction
Representation theory of symmetric groups is a classical and rich subject. The study of their irreducible representations has been initiated by Schur (cf. [
1]). Today, we have a well-developed theory of representations of symmetric groups in characteristic zero. Symmetric groups form a tower of groups
via the inclusions
,
, for example. Hence, once the representation category of symmetric groups is understood, one can study the induction and restriction functors on these categories.
The functors
and
are biadjoint and are related through Mackey’s theorem. In 2010, using this relation, Khovanov described a category (known as the Heisenberg category) which governs the natural transformations between induction and restriction functors between representations of the symmetric groups (cf. [
2]).
Today, we have more general Heisenberg categories, which are quantized, have a central charge
c, and depend on a choice of a graded Frobenius superalgebra
F, where
play the role of the symmetric groups in Khovanov’s construction, see e.g., [
3,
4,
5,
6,
7,
8].
In this paper, we investigate induction and restriction functors arising from a different tower, namely the tower of iterated wreath products of symmetric groups of order two. Intuitively, we work with the tower , where , and for and study the categories with objects bimodules and with morphisms bimodule homomorphisms. Induction and restriction functors can be identified with tensoring with certain bimodules; therefore, the morphisms in the category correspond to natural transformations between these functors. We will denote by the -linear additive category whose objects are generated by compositions of induction and restriction functors between groups , which start from . Morphisms of are natural transformations between the induction and restriction functors.
Since every finite-dimensional group algebra is a Frobenius algebra, the tower consists of Frobenius algebras. However, this tower is fundamentally different than the towers in Savage’s papers in that we do not take wreath products with symmetric groups increasing in size. Instead, we are iterating the wreath product.
We mention some other work related to the iterated wreath products for the interested reader. In [
9], the authors consider the iterated wreath products of
and describe a correspondence between their irreducible representations and certain rooted trees. Such a correspondence allows them to determine the Bratelli diagram for their tower of groups. In [
10], the first author and Angela Wu generalize these results to iterated wreath products of cyclic groups of various orders. As a follow up, in [
11], they generalize their results to iterated wreath products of symmetric groups of various sizes. Our focus is not the representations themselves but the induction and restriction between them as well as natural transformations between induction and restriction. In [
12], the authors introduce a notion of foams and relations on them to interpret functors and natural transformations in categories of representations of certain iterated wreath products, leading toward a discussion of patched surfaces with certain defects and their connections to separable field extensions and Galois theory.
In order to describe the category
or
, one needs to establish certain structural properties of the groups
. In
Section 2, we provide background on rooted trees and its connection to the construction of iterated wreath products. In
Section 3, we prove results about the right cosets, centers, centralizer algebras, and double cosets of
’s. In
Section 4, we consider
, where
,
and
. We provide an explicit vector space basis for
in Proposition 9, and we construct a generating set of
in Theorem 4. In
Section 5, we conclude by discussing many interesting open problems, in particular, in connection to the Heisenberg categories.
2. Rooted Trees and Iterated Wreath Products
Iterated wreath products of cyclic groups or permutation groups can be seen as automorphism groups of certain rooted trees (cf. [
9,
10,
11,
12]). A
rooted tree is a connected simple graph with no cycles and with a distinguished vertex, which is called a
root. A vertex is in the
j-th level of a rooted tree if it is at distance
j from the root. The
branching factor of a vertex is its number of children, and a
leaf is a vertex with branching factor zero.
A complete binary tree of height n is a rooted tree with all of its leaves at level n, and whose vertices that are not leaves all have branching factor 2.
A labeled complete binary tree is a complete binary tree whose children are labeled . Without loss of generality, we will also denote a labeled tree by .
The wreath product of the group with is with the multiplication given by , where corresponds to swapping the components of an element if is not the identity of , and if is the identity of . The symmetric group of degree has as its Sylow 2-subgroup the n-step iterated wreath product of . Writing to be the identity element in the wreath product and e to be the identity element in , the group and . This identification shows that the wreath product is precisely the semi-direct product of by .
Another way to think of the wreath product is through automorphism groups of rooted trees. An automorphism of a tree with vertex set V is a bijection such that are adjacent if and only if and are adjacent. These automorphisms of form a group under composition. So, the wreath product of the group with may be identified with the automorphism group of complete binary tree of height n that permutes the two children of each node.
That is, let and for . The iterated wreath product may be seen as the automorphism group of , since means take two copies of and permute them. One can see the two branches coming out of the root of as two copies of the tree , and an automorphism of can leave those two branches fixed or swap them (which then becomes the action of on two copies of the wreath product ).
We will denote the root of a rooted tree with an empty circle. See, e.g.,
.
With the tree picture in mind, there are two natural choices for embedding the wreath product into . One can either add new branches to the top of a tree, or one can take two copies of a tree of height n and connect them through a new root.
The first one preserves the root of the tree; hence, we call it
tree embedding of the wreath product and denote it as
. For example,
The second one, although it does not preserve the root of
, is more convenient to work with when one wants to identify elements of
with permutations of the leaves. If we label the leaves of a tree from 1 to
, then each tree automorphism can be seen as a permutation in
. In the second embedding, the expression of a permutation and its image are the same in cycle notation, since the labeling of the leaves is preserved. Hence, this embedding will be called
permutation embedding and will be denoted as
. For example,
.
Automorphisms of are determined by whether two children of each vertex are swapped or not. We will represent these swaps by dotted arrows on the diagram of a rooted tree. When reading a tree automorphism from the tree diagram, we apply the swaps from the bottom-most layer to the top-most layer, where the root of the tree is considered to be at the bottom.
From this point forward, we will work with the permutation embedding, and write .
Example 1. The wreath product can be thought of as the set of automorphisms of the tree
where the vertices of the top layer are labeled 1 and 2 when reading from left to right. That is, there is an isomorphism of groups, where Example 2. Consider . Given the complete binary tree
,
the map is a group homomorphism, where the isomorphism is indicated via sending
.
We thus have an embedding of , where maps to the automorphisms of that fixes vertices 3 and 4, i.e., the right branch in the blue circle is fixed. See Figure 1. This gives us a map , which is given by .
Note that there is also an embedding of the wreath product into the symmetric group on 4 letters, where one just keeps track of the labels on the top of the labeled tree . For example and .
Example 3. The group has a canonical identification with the automorphism group of the tree
.
Furthermore, the permutation embedding gives rise to the identification between and the automorphisms of that fixes the children of level 2 vertices fixed on the right branch: Therefore, is isomorphic to the automorphism group of a complete rooted binary tree of height n, and is isomorphic to the subgroup of the automorphism group of , which fixes the children of level vertices.
We will denote by
the image of
inside
induced by the map
which sends the label
k to the label
, where
, and by
, the permutation
.
Note that in terms of the tree , is the automorphism of the children of the leftmost branch growing from level 0, is the automorphism of the children of the rightmost branch growing from level 0, and is the automorphism of the tree which swaps and .
Example 4. When , is the automorphism of the subtree of indicated by the dotted edges:
,
and is the tree automorphism given below:
.
We will make a systematic use of the following observations in the rest of the paper:
Lemma 1. The groups and commute.
Proof. only acts on the label set and only acts on the label set . Hence, acting on disjoint labels, these two subgroups commute with each other, i.e., . □
Remark 1. The automorphism is the only link between the two branches and .
Lemma 2. Conjugating by gives .
Proof. Conjugating by is same as applying the permutation to the labels of g; hence, the resulting element is an element of obtained by replacing i with in the cycle notation of g. Therefore, one has . □
Example 5. Consider the transposition . Thenby Lemma 2. Note that is a set of generators for . In terms of the tree diagrams, these elements correspond to swaps at leftmost nodes at every level of the tree. The automorphism corresponds to the top level, furthest away from the root.
We will also be referring to groups
and
as the subgroups of
(
Figure 2).
For example, in , we have
where
,
and
.
3. Structural Properties of the Iterated Wreath Products of
This section provides results about the centers, centralizers, cosets and their representatives for the groups , which will be needed in the rest of the paper.
Recall that is the permutation on the leftmost vertex at the level i of the binary tree . We start by citing a generators and relations description of iterated wreath products of symmetric groups.
Theorem 1 (Theorem 1.2, [
9]).
The group is given by generators with relations- 1.
for ,
- 2.
for , , and
- 3.
for and .
Proposition 1. The group decomposes as the disjoint unionwhere Proof. We will prove by induction on l.
Case . This is easy to see when one considers as the automorphism group of the truncated tree . Such an automorphism either contains a swap at the root of the tree, i.e., , or not. Automorphisms that do not contain consist of combinations of swaps at higher levels of the tree, so they are elements of . Note that since elements of and commute, this set is the same as . The remaining automorphisms should contain and may contain any other swap, so they are elements of . Hence, .
Note that one can also express the set as using Lemma 2.
Inductive step. Suppose the proposition holds for
for some
l; that is,
is the disjoint union
Since
, by replacing
with the expression in (
1) one obtains
This concludes the proof. □
Proposition 2. The set of right cosets decomposes as the disjoint unionwhere Proof. It is enough to show that the sets
that partition
are invariant under left multiplication by
. However, this is obvious, since
implies that
□
Proposition 3. A set of representatives for the right cosets is given by the disjoint union .
Proof. By Proposition 2, we have
which completes the proof. □
Remark 2. For Proposition 3 reduces to the statement that a set of representatives for the right cosets is given by .
Proof. Let
and consider the generating set
of
where
. We know that
should satisfy
for all
. Conjugation by a permutation has the effect of relabeling; hence, we obtain
This equation puts a restriction on
for each
.
For , we obtain , so either fixes 1 and 2, or it swaps 1 and 2.
For , we obtain , so if fixes , we must have , which forces to fix 3 and 4 as well. Otherwise, if it swaps 1 and 2, we obtain , which forces to swap 3 and 4 as well.
Since Equation (
7) holds for
, repeating similar arguments, we obtain that either
, or
. □
Define the centralizer of
in
as
We note that the usual notation in group theory for the centralizer is .
Lemma 4. The centralizer of in is equal to .
Proof. Again, it will suffice to work with the generators of . Let , where and . Then, , where the second equality holds, since commutes with (see Lemma 1) and the third equality holds, since a is in the center of .
On the other hand, for , the condition for forces to either fix , or to swap 1 and 2, 3 and 4,⋯, and , and there are no restrictions on the labels . Since is the subgroup of that permutes these labels in any possible way, is of the form , where and . Therefore, . □
Lemma 5. is an -double coset of with elements.
Proof. We want to show that the elements are all distinct for . So, suppose , where . Then, . However, conjugation by has the effect of sending labels to . So, the permutations and are acting on disjoint sets in a partition of all the labels for . So, implies and . This proves that are all distinct for . Hence, is a -double coset of with elements. □
Proposition 4. The elements form a complete set of representatives for -double cosets of .
Proof. We will provide a proof by showing that each of these representatives gives disjoint double cosets, and by using a counting argument, i.e., their total cardinality is equal to the cardinality of .
For , we obtain since elements in and commute. This gives us disjoint double cosets (they are disjoint because each b corresponds to a different swap at the right branch of the tree ), each with elements. So, the double cosets for count for many elements of .
Since the cardinality of is , there are still elements we have to account for.
Note that . It follows from Lemma 4 that each element of gives a distinct double coset of size , and it follows from Lemma 5 that gives a double coset of size . This counts for all the elements of . □
Proposition 5. A basis of the subalgebra is given by the orbit sums of the conjugation action .
This is an application of the classical yet extremely powerful idea of averaging over group elements ubiquitous in representation theory. We will now prove Proposition 5.
Proof. Let
be an orbit and let
. Then for any
, we have
Hence, .
Conversely, let
where
is a basis of
and suppose that
for all
. Then, we have
Since this is true for all
, all the basis vectors
for
in the orbit of
should be present in
v for all
. Moreover, in order for the above equation to hold, one needs the coefficients of the basis vectors indexed by elements from the same orbit to be equal to each other. Therefore,
is a basis of
. □
There is a special element whose orbit exhibits a notable behavior. The orbit of is the same under the conjugation actions of and .
Proposition 6. Let be the orbit of under the conjugation action of on , and let be the orbit of under the conjugation action of on , i.e., the conjugacy class of in . Then, .
Proof. Note that is a product of transpositions of the form where , . Conjugation by has the effect of permuting only the labels from the set .
Conjugation by affects both the labels and . Every g can be written as for and . Here, only can exchange labels between and ; however, note that no matter which element of we choose, we can never obtain a permutation containing since if is in , then should be in .
Therefore, conjugates of are products of transpositions , which can all be obtained by conjugation with . Hence . □
Remark 3. Let denote the sum of the elements in the orbit , i.e., . Note that is a central element of since it is a conjugacy class sum. As a result, we obtain that is also central in .
Now, we are in a position to give an explicit vector space basis of .
Proposition 7. The number of orbits of the conjugation action is where is the number of conjugacy classes of . Moreover, if we denote by the conjugacy class of , then the orbits are and .
Proof. We will use the decomposition . Since and commute, the action of on results in orbits of the form for where is the conjugacy class of a in .
As for the action on
, note that we have the equalities
Therefore, if we choose to work with
, for
and
,
Hence, we obtain orbits
. Since
is equal to the conjugacy class
in
by Proposition 6 and any
can be expressed as
, we obtain
Therefore, the orbits are of the form
for
. □
This determines a basis for and its dimension is . To obtain an explicit formula for this dimension, we need an explicit formula for the number of conjugacy classes of , which is same as the number of irreducible representations of .
We restrict Theorem 2.2 in [
9] to
to obtain the number of conjugacy classes:
Theorem 2 (Orellana–Orrison–Rockmore).
The number of irreducible representations of is given by and the recursion The following proposition describes the orbits of the action . An explicit vector space basis for is given by their orbit sums.
Proposition 8. The number of orbits of the conjugation action is where is the number of conjugacy classes of , is the sequence defined recursively as , for .
Moreover, if we denote by the conjugacy class sum of g in , then the orbits are
- 1.
for and
- 2.
,
where are from the set .
Proof. We will give a proof by induction on k. The base case is proved in Proposition 7. Now, assume the proposition holds for some k. We will make use of the decomposition . This decomposition allows us to describe the orbits by studying the action of on and on separately.
For the action of on , since and commute, for , , , we obtain . Therefore, the orbits of the action of on are given by orbits of on times an element of .
Now, for the action
, since we have
using this relation inductively, one ends up with
and obtains
For
,
, we want to study
. Note that it is possible to express
as
, where
. Hence, we obtain
ending up with orbits of the form
, where
∈
and
.
□
4. Natural Transformation between Induction and Restriction
Let the symbol denote the as an -bimodule.
Objects of the category - are generated by -bimodules for a fixed n. Since and correspond to and , respectively, one can identify the bimodules and with and , respectively. Therefore, it is possible to describe the category in terms of compositions of and and natural transformations between them as well. Let us denote by the category whose objects are generated by compositions of induction and restriction functors between groups , where the compositions start from induction or restriction from .
Given a sequence of inductions and restrictions, using the formula (
11) below, it is possible to move inductions to the left and push restrictions to the right. Hence, objects of
are given by direct sums of
starting from
-bimodules.
Morphisms of consist of sets . Since we start by applying the restrictions to an -bimodule, it is clear that these hom spaces are trivial when or . In addition, note that when , we are talking about bimodule maps between two bimodules over different rings. Therefore, in this case, the hom space is trivial as well.
From now on, we will assume that , and . We would like to describe the spaces , or for short, as a vector space and as an algebra.
In this section, we describe all these spaces implicitly and provide explicit descriptions for the cases (hence ) and .
4.1. Mackey Theorem
Let us start by recalling the general form of Mackey theorem for induction and restriction.
Theorem 3 (Mackey Theorem).
Let G be a finite group and be two subgroups of G. Thenwhere I is a set of -double coset representatives of G and . In our case, by letting
and
, we obtain
where
.
Using Proposition 4 which informs us about double cosets, we can express the above formula more explicitly once we say something about for double coset representatives g. We will work with the double coset representatives .
For , , since and commute. Hence, the intersection of and is again .
For
, we have
; hence,
. Therefore, the Mackey formula for the groups
gives
The fact that the number of summands of
on the right-hand side of (
11) depends on
n suggests that this line of investigation should be different than the Heisenberg categories in the literature. More precisely, there should not exist an abstract 1-category
encapsulating the induction and restriction functors on the tower
and natural transformation between them which will act on the categories
-
for all
n as bimodule homomorphisms. Therefore, we will focus on describing the explicit categories
-
.
Note that given the induction functor
-⟶-,
we want to describe the natural transformations such that the diagram
commutes, where is an -bimodule homomorphism. Suppose . Then, for any and , , where is multiplication by a from the right. On the other hand, . So, for any implies if and only if the bimodule homomorphism f is multiplication by an element from the centralizer .
4.2. Morphism Spaces as Vector Spaces
In this subsection, we give a description of the morphism spaces between induction and restriction functors on the representations of the tower
as vector spaces.
Unlike the case for symmetric groups (cf. Theorem 2.1 in [
13]), the centralizer for our tower of group algebras forms a non-commutative algebra. The size of our groups grows exponentially, and as a consequence,
is not multiplicity free for this tower.
consists of natural transformation satisfying the following diagram:
In terms of bimodules, this amounts to
where
. Therefore, if
, for
, we require that
This calculation shows that
For
, we obtain that
which is the centralizer of the image of
inside
. Let us denote it with
. A vector space basis of
, hence of
, is given explicitly by orbit sums described in Proposition 8.
Using the biadjointness between and , one immediately obtains a similar result for .
Lemma 6. is isomorphic to as a vector space, and it is isomorphic to its opposite algebra as an algebra.
Proof. The induction and restriction functors form a cyclic biadjoint pair for a finite group
G and one of its subgroups
H, i.e., see Section 3.2 in [
2]. This relationship allows one to deduce the structure of
as a vector space and as an algebra.
Alternatively, in terms of bimodule homomorphisms, the natural transformations between correspond to bimodule maps of multiplication on the right, and the ones for correspond to bimodule maps of multiplication on the left. Hence, one obtains the same vector space, and the algebra multiplication reverses the order. □
Although we can provide a nice description for the vector space structure of and , the picture changes slightly for when since in this case, a vector space basis is given by the orbit sums of the conjugation action of on the tensor product , which is not a group algebra. So, even though one can give an explicit description of orbit representatives, it is difficult to express the orbit sums explicitly.
Proposition 9. A vector space basis of is given bywhere . Proof. We start by noting that an element of
can be written as
where
,
and
is a representative from the quotient
after pushing all possible terms to the left of the tensor. Therefore,
is a basis of
. We will choose
from the set of representatives in the disjoint union
given in Proposition 3.
Let us look at the case first. In this case, the representatives are . Given a basis element , b is either in or in .
For
,
and
, we obtain
and
More generally, for any
l, if
, then
and
where
. □
We do not have an explicit description for , where but .
In the next section, we will describe the algebra structure of the endomorphism spaces.
4.3. Algebra Structure of Morphism Spaces
In [
14], the author develops a general approach for obtaining a nice set of multiplicative generators for the Gelfand–Tsetlin subalgebra
. His approach does not apply directly to our tower of groups, since we are in a non-multiplicity free setting.
The set is closed under precomposing with elements of the center of the group algebra. Therefore, the algebra can be decomposed as , where conjugacy class sums of form a basis of and is a spanning set for the algebra .
Since the group is generated by , a generating set for the algebra is given by . The relations among the generators in are described in Theorem 1. is a central element of by Remark 3; therefore, it commutes with the generators in , i.e., . However, this is not enough to give a generators and relations description of the algebra . In particular, the only relation missing in order to give a complete generators and relations description is the expression of in terms of the generators. The only nice pattern we have observed is for and k odd. In fact, closed formulas are very difficult to find for all the other cases, i.e., and k even, and . However, working in a concrete setting has the advantage of defining as the subalgebra of generated by .
Thus, to determine the algebra structure of , it remains to describe the algebra structure of . It is difficult to express the structure coefficients if one chooses to work with the conjugacy class sums as a basis. Khonvanov’s Heisenberg category provides a way to express centers of symmetric group algebras with generators and relations using a different basis defined via diagrammatics as certain bubbles. We lack a diagrammatic description for our category. We leave the algebra structure of as an open problem.
Next, we will write down the generating set for the algebra structure of the endomorphism algebra
. Similar as before,
decomposes as
, where
span the algebra
. Since
is generated by the swaps
and
is generated by
, where
are the leftmost swaps of
and
are the leftmost swaps in each row in
, respectively, a generating set for the algebra
is given by
More generally, consider
acting by conjugation on the group algebra
. Then,
decomposes as
, where the disjoint union
spans the algebra
. The group
is generated by
where
are the leftmost swaps in each row in
,
. This leads us to the following theorem:
Theorem 4. The endomorphism algebra decomposes aswhere a generating set for the algebra is the disjoint unionand , (Figure 3). As mentioned above, since the relations among the algebra generators are very complicated, we will refrain from explicitly writing them but leave this as an open problem.