In this post, we discuss a few ways in which the symmetric and alternating groups can be realized as finite collections of self-maps on the Riemann sphere. Throughout, our group operation will be composition of functions: as such, the maps we choose will necessarily be homeomorphisms of
. Within this broad framework, two classes are of particular interest:
1. The group of biholomorphic maps (those that respect the structure of
as a Riemann surface). It is well-known that such maps are given by Möbius transformations, i.e. rational functions of the form
satisfying . The group of Möbius transformations (also known as the Möbius Group and herein denoted
) is naturally isomorphic to
, the projective (special) linear group, via:
2. The group of conformal maps , denoted
for brevity. To be clear, here we refer to those maps
which preserve unsigned angle measure. (In contrast, some authors require conformal maps to preserve orientation as well.) We recall the fundamental result that such maps contain the Möbius group as a subgroup of index two. To be specific, any conformal self-map on
is either biholomorphic (returning to case (1)), or bijective and anti-holomorphic: a biholomorphic function of the complex conjugate
.
In this post, we enact a two-part program to calculate the maximal such that the symmetric group
injects into
(resp.
). Along the way, we study injections of the alternating group into
, and highlight some exceptional cases in which our injections can be attached to group actions on a finite invariant set.
— PART I (Injections ) —
Our first goal will be to verify the existence of injective maps . While it suffices here to provide a single example, it is infinitely more enlightening to show how such an example may be found. As a happy consequence, we’ll end up classifying all possible images
up to inner automorphism. To be specific,
Theorem 1.1: There exists an injection . Moreover, this injection is unique (up to inner automorphism on
).
Proof: Recall the group presentation
Now, suppose that is given, and let
denote the images of
, respectively. Because
is elliptic, we may assume
(up to inner automorphism on
), in which
is primitive. Since
has order two if and only if
,
takes the form
Secondly, the relation (with plenty of algebra!) forces
as a polynomial in . Moreover, since
has order exactly four, we find
(else
has order dividing two). Next, because
, we can rewrite this as
, hence
. In particular,
is non-zero, by a determinant calculation. We may thus projectivize
, such that
(1)
in which we have used that . Because the centralizer
contains the subgroup of diagonal (projective) matrices, we may freely conjugate
by diagonal elements. In particular, we note that
hence is conjugate (via
) to a Möbius transformation mapping
. In other words, we may assume
in line (1) – again, up to inner automorphism on
. In this sense,
is uniquely determined by
, whereby there exist at most two injections
modulo inner automorphism: one for each choice of
among the roots
of
.
As it happens, each choice of induces a valid injection
(this is finite computation). As a final claim, we leave to the reader to show that conjugation by the Möbius transformation
fixes
while interchanging the choices for
associated to
and
. That is, any two injections
differ by an inner automorphism of
, precisely as claimed.
Next, we claim that the constant taken in Theorem 1.1 is maximal. Here, two avenues of proof seem promising:
- A direct proof (along the lines of Theorem 1.1), built from a manageable group presentation for
.
- A proof built upon Theorem 1.1, noting that any injection
restricts to a map on
of the type studied in Theorem 1.1.
Both work well. The first is given below, while the second is sketched in the Exercises:
Theorem 1.2: There exists no injection .
Proof: We begin with the group presentation
known at the time of Burnside’s article Note on the Symmetric Group (1897). If exists, let
denote the images of
, respectively. For definiteness, we may assume that
(up to inner automorphism). For
, we recall that
if and only if
. Thus
takes the form
and some computer-assisted algebra quickly gives us the following relations:
The first of these implies , whereas the second yields
Since , we obtain
, which contradicts that
.
— PART II (Injections ) —
The questions answered in the previous section admit natural analogues with in place of
. Answering them, however, will be a bit harder than before, owing to the greater complexity of the group structure on
. For this reason, we try when possible to reduce questions regarding
to questions of
alone.
For example, the identification induces a semi-direct product structure on
; namely,
,
in which the -action is given by complex conjugation. It follows that any injective map
induces a restricted map
, i.e. an injection of the alternating group
into the Möbius group. To see how this might benefit us, consider the following:
Proposition 2.1: There exists no injection . It follows that
does not inject into
.
Proof: See the Exercises.
As it turns out, there does exists an injection of into
, which is maximal in the sense of the preceding Proposition. Unfortunately, this fact does not a priori imply that
injects into
(so Proposition 2.1 won’t serve a major role in our classification of symmetric subgroups). In fact, the opposite is true:
Theorem 2.2: There exists no injection .
Proof: As in Theorem 1.2, we begin with the group presentation
.
Suppose that exists. Since
has cycle type
, it follows that
, hence
lies in
. Regarded as an element of
,
is torsion (hence elliptic), so we may assume that
, in which
(up to inner automorphism). As
by Theorem 1.2, we may write
(in general form). The relation holds if and only if
(2) ,
, and
.
(Case 1): If , we may scale
; then
and
. It follows that
, in which
or
. If
, we then have
, which implies that
, a contradiction. Thus
, and we may write
On the other hand, that forces
, an impossibility. This case is therefore untenable, and we turn to:
(Case 2): If , we may projectivize to assume
; then
and
. This last expression is both real and (purely) imaginary, hence
. Thus
, and
takes the form
As for the relation , it follows that
. Since
, this implies
. After substitutiting this into our expression for
, we find that the relation
forces
Coupled with , we find
, but this contradicts that
has order five. Thus there exists no injection
.
We note that Theorem 2.2 recovers Theorem 1.2. Regardless, Theorem 2.2 can’t be used to prove Theorem 1.2, because we have invoked that first theorem in assuming . Avoiding such circularity would require — in essence — reproving Theorem 1.2 as an early claim.
For free, we obtain
Corollary 2.3: There exists an injection if and only if
.
— PART III (Group Actions and Invariant Sets) —
We’ve seen that contains no symmetric subgroups
beyond those that inject into
, despite the fact that
contains
properly as an index two subgroup. Why, then, might we be interested in subgroups
not contained in
?
For starters, let be any injection, and consider the sequence of maps
in which denotes the quotient of
by its normal subgroup
. As mentioned previously, we have
. (In particular, we cannot have
, Klein’s four group, despite the fact that
is normal in
.) Thus
if and only if
. When these equivalent conditions hold, membership in
is detected by the sign character on
. In this one sense, the theory of symmetric subgroups in
is more robust than the analogue theory in
.
There is a second, much more compelling reason to study symmetric subgroups in , which begins with the following construction:
Let be any finite set,
, and suppose that
is a subgroup of conformal mappings such that each
permutes the points of
. (In other words,
is an invariant set for the action of
.) This gives an induced map
, known as the permutation representation (associated to the group action
). Our interest in such maps is simple: if
bijects, then
is exactly the sort of map we’ve been looking for.
If this permutation representation bijects, then our general theory implies. Of these
, the extremal case
naturally presents the greatest interest. Here, there are two cases to consider:
- All points in
lie on a circle in
. Up to inner automorphism, we may assume that
, with
.
- The points in
lie on no common circle, but we may assume after conjugation that
is given by
, with
.
If case (1) holds, note that we may assume , by modding out by the trivial action of
. As it turns out, however, case (1) cannot actually occur:
Example 3.1: Suppose that surjects, and that case (1) holds. We may assume
. If
corresponds to the 3-cycle
, brief computation shows that
Since fixes
, we must have
. But this polynomial has no real roots, which contradicts that
. It follows that case (1) cannot occur.
Note: If case (2) can be realized, this gives strong motive to consider injections of into
, as opposed to injections into the smaller group
. As it happens, case (2) is realized, in an essentially unique way.
To prove this result, it will be advantageous to present first a general Lemma. For the moment, let’s relax our assumptions and suppose that the permutation representation merely surjects. By the previous Example, it follows that the invariant set
has
(after inner automorphism). As the following (general) Lemma shows, the map
injects without further hypotheses:
Lemma 3.2: Suppose that contains four points not lying on a circle in
. If
fixes each point in
, then
. In other words, the permutation representation
injects.
Proof: Suppose that a non-identity element fixes
. Conjugating
by an element of
, we may assume that
contains
, with
. We note that
, as
acts sharply 3-transitively on
. Thus, writing
with
, it follows that
fixes the three points
, and
. Because
acts sharply 3-transitively on
, we have
. Thus
, which fixes
(and nothing else). This contradicts that
fixes
, whence no such
exists.
And now, our final Theorem:
Theorem 3.3: There exists a subgroup and an invariant set
,
, such that the permutation representation
bijects. If
and
are as
and
above, there exists an element
such that
and
. It follows that the associated injection
is unique up to inner automorphism.
Proof: Suppose that and
exist as above. Up to inner automorphism by
, we may assume that
, with
. Moreover, Example 3.1 implies that
is a root of
. In the group presentation
we may take , wherein
is as in Example 3.1. Since
the image of
(
, say) is anti-holomorphic, hence a simple transposition (by order). Moreover,
can be chosen to fix
while transposing
, as
is complete. Thus
takes the form
, for which we note that
surjects.
By Lemma 3.2, we have . As such,
is uniquely determined by
, and
is determined by choice of
(among the roots of
). These two choices of
are related by conjugation by
, which gives uniqueness of
(in the sense of this Theorem). It follows that
is unique up to inner automorphism, as claimed.
— EXERCISES —
Exercise: Herein, we outline a second proof of Theorem 1.2. Assuming that injects, let
denote any subgroup isomorphic to
. By Theorem 1.1, we may assume that
is generated by
and
,
in which is some primitive cube root of unity. Show that
contains an element
that commutes with
such that
and
. Use these relations to find a contradiction.
Exercise: Using the group presentation
,
show that contains a subgroup isomorphic to
(which is unique up to inner automorphism). Then, using the group presentation
,
show that no injection exists.