Before leaving the world of linear algebra for multilinear algebra (i.e. tensors), let us reformulate the spectral theorem entirely “upstairs” in (as in Lecture 1,
is complex vector space of dimension
equipped with a scalar product).
Theorem 1: If is selfadjoint, then there is a positive integer
real numbers
and selfadjoint operators
in
satisfying
and
such that
The operators above are called orthogonal idempotents. The “idempotent” part is
, and this term is applied more generally in ring theory to refer to anything equal to its own square. The “orthogonal” part is
, the zero operator in
, when
This is a useful representation of
, because orthogonal idempotence immediately implies that
Given your Math 202A background, you have realized immediately that Theorem 1 is equivalent to the eigenvalue/eigenvector form of the Spectral Theorem discussed in Lecture 1, because the set of selfadjoint idempotents acting in
is in canonical bijection with the lattice
of subspaces of
, since these are exactly the operators which act by orthogonal projection onto a subspace: for any
, the corresponding
maps
to the nearest point of the subspace
. The correspondence between selfadjoint idempotents — which are often simply called orthogonal projections — and subspaces even holds in the infinite-dimensional setting provided
is complete with respect to the norm
, i.e.
is a Hilbert space, provided we restrict to closed subspaces. Thus the Grassmannian
is for all practical purposes equivalent to the set
of selfadjoint idempotents/orthogonal projections of rank
. Theorem 1 above is obtained from the eigenvalue/eigenvector form of the Spectral Theorem by replacing the eigenspaces of
with the orthogonal projections onto those spaces, and conversely Theorem 1 gives the eigenvalue/eigenvector form of the Spectral Theorem with
being the distinct eigenvalues of
whose corresponding eigenspaces are the images of the orthogonal idempotents
.
Let us close out this brief recap of eigenvectors and eigenvalues (i.e. Math 202A material) by recalling the algebraic criterion for diagonalizability. An operator is said to be normal if it commutes with its adjoint,
. Obviously, selfadjoint operators
are normal,
, and so are unitary operators
since
.
Theorem 2: Given , there exists an orthonormal basis
of
such that
for each
, where
are scalars, if and only if
is normal. Equivalently,
is a
-linear combination of pairwise orthogonal selfadjoint idempotents if and only if
.
After revisiting this essential piece of 202A technology, my plan was to jump into multilinear algebra — tensor products, exterior products, symmetric products, and their applications to eigenvalue inequalities — and work up to the construction of Schur functors, which interpolate between these using the representation theory of the symmetric groups. This would be an application of the material you learned in 202B wherein the goal is to enhance the material you learned in Math 202A. However, I have since changed my mind. After reviewing Professor Sam’s 202B lecture notes (which I am posting to Piazza for your reference), and speaking briefly with Steven, I have realized that you already have some exposure to this material. Even though there is much left to be done in this direction, namely starting from Section 6 in Steven’s notes and working up to a proof of Theorem 6.4.1 via Schur-Weyl duality, I have ultimately decided to leave this for the second half of the course, and begin instead with some very natural and appealing applications of symmetric group representations which we can sink our teeth into immediately. Also, this will give you exposure to material which cannot be looked up in textbooks, or at least cannot be found presented in the way we’ll do it.
Definition 1: For each , the symmetric group
of degree
consists of the set of bijective functions
i.e. automorphisms of , with the group law being composition of functions. Elements of
are called permutations.
As you know from Math 202B (and likely much before), any permutation can be factored into a product of disjoint cyclic permutations, and this factorization is unique up to the order of the factors. Our starting point will be a rather different unique factorization theorem for permutations. To explain, let us consider
from the perspective of geometric group theory. Let
denote the set of transpositions, i.e
is the conjugacy class of
-cycles,
As you are also no doubt aware, generates
, which means that any permutation
can be written as a product of transpositions,
This representation is very far from being unique, and indeed the question of counting the number of factorizations of a given permutation into a given number of transpositions is a classical problem in algebraic combinatorics known as the Hurwitz enumeration problem; despite its elementary formulation, this question has deep ties to enumerative algebraic geometry and mathematical physics, some of which we will hopefully be able to discuss. Nevertheless, for each there is a minimal value of
for which such a factorization exists, and this is called the word norm of
and denoted
, where the subscript reminds that this quantity depends on the choice of generating set
We will omit this subscript in order to lighten the notation, and if we switch to a different generating set
inducing a different word norm on
, we will say so loud and clear. The following is elementary combinatorics, but you should take the time to prove it.
Proposition 1: The length of any factorization
of a permutation
into transpositions satisfies
for some
. That is, all factorizations of
into transpositions have the same parity.
Somewhat less elementary but no less important is the fact that there is a tight connection between the word norm of a permutation and the number of factors
in its unique decomposition into disjoint cycles.
Proposition 2 (join-cut relation): We have
The key idea in the proof of Proposition 2 is expressed in its name: multiplying a permutation by a transposition will either join two of its cycles into one, or cut one of its cycles into two. Thus,
The word norm makes the symmetric group into a metric space space.
Proposition 3: The function
defined by | is a metric.
We may thus for example speak of the open ball of radius centered at a given permutation
,
and its boundary, the sphere
The metric is nothing but the graph theory distance (or geodesic distance) on
the Cayley graph of the symmetric group as generated by the conjugacy class
of transpositions. The vertex set of this graph is
itself, and two permutations
are joined by an edge if and only if there is
such that
Note that this is an undirected graph because
is the identity permutation for every
so that
is equivalent to
Note also that
is a regular graph of degree
Another feature of
is that it is a graded graph: it decomposes into a disjoint union of independent sets or “levels,”
with being the sphere of radius
centered at the identity permutation
. By the join-cut relation,
may also be described as the set of permutations whose disjoint cycle decomposition contains
factors.
The Hurwitz enumeration problem has an equivalent formulation in terms of the Cayley graph: it asks for the number of
-step walks from
to
on
This description also points to a refinement of the Hurwitz problem. Let us introduce the edge-labeling on
in which each edge corresponding to the transposition
is marked by
, the larger of the two numbers it interchanges. We will call this the Biane-Stanley edge labeling of the symmetric group, after Philippe Biane and Richard Stanley, in whose work it appears implicitly.

Given an -step walks
on
, we define its signature to be the corresponding sequence
of edge labels. We say that a walk is strictly monotone if its signature is a strictly increasing sequence. Thus, an
-step strictly monotone walk from
to
is the same thing as a strictly monotone factorization of
into
transpositions, i.e. a factorization
such that
Theorem (unique monotone factorization): Every permutation admits a unique strictly monotone factorization, and the length of this factorization is equal to
That is,
Proof: The proof is by induction on .
In the base case we have with
and
The complete answer to the Hurwitz enumeration problem for
is thus given by
and
This corresponds to the fact that if and
have the same parity, then we have the unique length
factorization
but no factorization exists if and
have opposite parity. Imposing the strictly monotone constraint whittles this down to
and
For the inductive step, there are two cases to consider. Let and let
be any permutation. If
is a fixed point of
, i.e.
, then
has a unique strictly monotone factorization by the induction hypothesis. If on the other hand
is one of the
permutations which contains in its support (meaning
), then
is of the form
for a unique pair and
By the induction hypothesis,
admits a unique strictly monotone factorization of length equal to
, and the result follows.
—- Q.E.D.
Graphically, the unique monotone factorization theorem gives a presentation of the symmetric group as a starlike tree, which we call the Jucys tree after the Lithuanian physicist Algimantas Adolfas Jucys (not to be confuse with his father, the Lithuanian physicist Adolfas Jucys), in whose work it appears implicitly. We will embed this construction in the group algebra
in Lecture 3, and start using it in conjunction with linear representations of the group algebra.
