Sei sulla pagina 1di 28

REPRESENTATION THEORY OF THE SYMMETRIC

GROUP
MARK WILDON
Recommended reading: [7] G. D. James, Representation theory of
the symmetric groups, Springer Lecture Notes in Mathematics 692,
Springer (1980).
1. Background definitions
Algebraic denitions. Let G be a nite group and let F be a eld.
The group algebra FG is dened to be the F-vector space with basis
g : g G and multiplication dened by linear extension of the
multiplication in G. So given

hH

h
h,

kG

k
k FG we set
(

hH

h
h)(

kG

k
k) =

hH
kK

k
hk =

gG
_

hH

h
1
g
_
g.
An F-vector space V is a FG-module if there is a map V G V ,
which we shall write as (v, g) vg, such that
(i) for each g G, the map v
(g)
vg is F-linear, and
(ii) if g, h G then (g)(h) = (gh) and (1
G
) = 1
V
.
Exercise:
(a) Show that if V is an FG-module as above, with G-action given
by vg = v(g) for g G, then is a group homomorphism from G into
the group GL(V ) of invertible linear transformations of V .
(b) Show conversely that if : G GL(V ) is a group homomor-
phism (so V is a representation of G) then setting vg = v(g) gives V
the structure of an FG-module.
Remark 1.1. I should probably have mentioned that the usual deni-
tion of a module for a ring would require a map V FG V . I didnt
do it this way, because the only extra is the action of the eld F,
and it seems easiest to give this by saying that V is an F-vector space.
Later, I will dene ZG-modules, to which a similar remark will apply.
Remark 1.2. The denition of an FG-module is more technical than
the denition of a representation of G, but, as the exercise shows,
the two notions are equivalent. Module can be more convenient to
work with, because there is less notation, and we can use results from
ring theory without any translation. The language of representations is
preferable if we want to have an explicit map : G GL(V ). A similar
Date: First term 2011/12.
2
situation arises when we have a group acting on a set. Here one can
choose between the equivalent languages of G-actions and permutation
representations. Again, both have their advantages.
If V is a FG-module then a subspace W V is a submodule of V if
wg W for all w W.
Let S
n
denote the symmetric group of degree n. We shall compose
permutations from left to right, so for instance (using cycle notation
for permutations),
(123)(12) = (23).
In the lecture I mentioned the following example as something to
think about, but didnt write anything down.
Example 1.3. Let F be a eld and let V = v
1
, . . . , v
n

F
be an n-
dimensional vector space over F. We make V into an FS
n
-module by
dening
(1) e
i
g = e
ig
for i 1, . . . , n and g S
n
.
Note that it follows from (i) in the denition of a module that
_
n

i=1

i
e
i
_
g =
n

i=1

i
e
ig
for all

n
i=1

i
e
i
V , so (1) really does suce to specify the FS
n
-
module structure on V . The corresponding representation : S
n

GL(V ) represents each g S
n
as a permutation matrix. For example
if n = 3, g = (12) and h = (123) then
(g) =
_
_
0 1 0
1 0 0
0 0 1
_
_
and (h) =
_
_
0 1 0
0 0 1
1 0 0
_
_
.
Combinatorial denitions. A partition of n N
0
is a non-increasing
sequence of natural numbers whose sum is n. For example, the 7
partitions of 5 are (5), (4, 1), (3, 2), (3, 1, 1), (2, 2, 1), (2, 1, 1, 1) and
(1, 1, 1, 1, 1). The numbers making up a partition are its parts. We shall
use powers to indicate multiplicities of parts: for example (2, 1, 1, 1) =
(2, 1
3
).
It is often useful to represent a partition by a rectangular array of
boxes, called its Young diagram
1
. For example, the Young diagram of
1
Young diagrams are named after the Reverend Alfred Young: see www-history.
mcs.st-andrews.ac.uk/Biographies/Young_Alfred.html. In a letter sent to
Frobenius, Young wrote:
I am delighted to nd someone else really interested in the matter.
The worst of modern mathematics is that it is now so extensive
that one nds there is only about one person in the universe really
interested in what you are . . .
3
(4, 2, 1) is
.
Remark 1.4. It is well-known that the partitions of n label the con-
jugacy class of S
n
via cycle types. For example, the conjugacy class of
S
5
labelled by (2, 2, 1) contains (12)(34). We shall see later that the
irreducible CS
n
-modules are also canonically labelled by the partitions
of n.
Let = (
1
, . . . ,
k
) be a partition of n N. A -tableau is an as-
signment of the numbers 1, . . . , n to the boxes of the Young diagram
of , so that each number is used exactly once. For example,
5 6 3 2
7 1
4
is a (4, 2, 1)-tableau. More formally, if we x a numbering of the boxes
of , then we may regard a -tableau t as a function
t : 1, . . . , n 1, . . . , n
such that it = j if and only if box i contains j.
If s and t are -tableaux then s and t are said to be row-equivalent
if for each i 1, . . . , k, the sets of numbers in the ith row of s and t
are equal. For example,
5 6 3 2
7 1
4
and
2 3 5 6
1 7
4
are row-equivalent.
A -tabloid is a row-equivalence class of -tableaux. We denote
the -tabloid corresponding to the -tableau t by t. Tabloids are
drawn by omitting the vertical lines from a representative tableau. For
example if t is the rst tableau above then
t =
5 6 3 2
7 1
4
=
2 3 5 6
1 7
4
= . . . .
A -tableau is said to be (a) row-standard, if its rows are increasing
when read from left to right, (b) column-standard, if its columns are
increasing when read from top to bottom, and (c) standard, if it is both
row-standard and column-standard.
4
Example 1.5. Let = (4, 2). There are 6! = 720 (4, 2)-tableaux, 15
(4, 2)-tabloids and 9 standard (4, 2)-tableaux.
A (4, 2)-tabloid is determined by its second row, so the (4, 2)-tabloids
are in bijection with 2-subsets of 1, 2, 3, 4, 5, 6. For example, the 2-
subset 2, 4 corresponds to the tabloid
1 3 5 6
2 4
The 9 standard (4, 2)-tableaux,
1 3 5 6
2 4
,
1 3 4 6
2 5
,
1 3 4 5
2 6
,
1 2 5 6
3 4
, . . .
will turn out to index a basis of an irreducible 9-dimensional CS
6
-
module.
2. Young permutation modules and Specht modules
Throughout this section let n N and let F be a eld.
The symmetric group S
n
acts on -tableaux. For example, if g =
(235) S
5
then
1 2 5 6
3 4
g =
1 3 2 6
5 4
since 1g = 1, 2g = 3, 5g = 2, 6g = 6, and so on. With the denition
of tableau as function, if t : 1, . . . , n 1, . . . , n is a -tableau and
i 1, . . . , n is the number of a box, then i(tg) = (it)g. Either way, it
is clear that the action of S
n
on -tableaux is regular, i.e. the stabiliser
of a tableau is the trivial group.
Lemma 2.1. Let be a partition of n. There is a well-dened action
of S
n
on the set of -tabloids dened by tg = tg where t is a
-tableau and g S
n
.
Denition 2.2. Let be a partition of n. The Young permutation
module M

F
, dened over F, is the FS
n
-permutation with permutation
basis t : t a -tabloid.
Usually we will write M

rather than M

F
since the ground eld will
be clear from the context.
Example 2.3. (1) If = (n) then there is a unique -tabloid, namely
1 2 n
and so M
(n)

= F, the trivial FS
n
-module.
(2) If = (n1, 1) then there are n distinct (n1, 1)-tabloids, with
representative tableaux
t
1
=
2 3 n
1
, t
2
=
1 3 n
2
, . . . , t
n
=
2 3 n
1
.
The top row of a two-row tabloid can be deduced from the second row
so we may write
t
1
=
1
, t
2
=
2
, . . . , t
n
=
n
.
5
It is clear that t
i
g = t
ig
for each g S
n
. Thus M
(n1,1)
aords the
natural permutation representation of S
n
.
(3) More generally, let 1 k < n and let = (n k, k). Ignoring
the rst row of a (nk, k)-tabloid gives a bijection between -tabloids
and k-subsets of 1, . . . , n. Hence M
(nk,k)
is isomorphic to the FS
n
-
permutation module of S
n
acting on k-subsets of 1, . . . , n.
(4) Let 1 k < n and let = (n k, 1
k
). Ignoring the rst row of
a (n k, 1
k
)-tabloid gives a bijection between -tabloids and k-tuples
of distinct elements of S
n
. There is a corresponding interpretation
of M
(nk,1
k
)
. In particular, taking k = (n 1) we see that M
(1
n
)
is
isomorphic to the group algebra FS
n
as a right FS
n
-module.
Denition 2.4. Let be a partition of n and let t be a -tableau. The
column group C(t) consists of all permutations which preserve setwise
the columns of t. The signed column sum b
t
is the element of the group
algebra FS
n
dened by
b
t
=

hC(t)
hsgn(h).
The polytabloid corresponding to t is dened by
e(t) = tb
t
.
Note that while the polytabloid e(t) is a sum of tabloids, it depends
on the tableau t, and not just on the tabloid t. For example if
= (2, 1) then
e
_
1 2
3
_
=
1 2
3

3 2
1
=
1 2
3

2 3
1
e
_
2 1
3
_
=
2 1
3

3 1
2
=
1 2
3

1 3
2
We observe that S
n
permutes the set of all -polytabloids. If t is a
-tableau and g S
n
, then
e(t)g = tb
t
g = tgg
1
b
t
g = tgb
tg
= e(tg).
where we have used that C(tg) = C(t)
g
, and so b
tg
= g
1
b
t
g.
Denition 2.5. Let be a partition of n. The Specht module S

F
, de-
ned over F, is the submodule of M

spanned by all the -polytabloids.


By the observation just made, S

is a well-dened FS
n
-module. It
also shows that S

is cyclic, generated by any single polytabloid.


Example 2.6. (1) Take = (n). Then while there are n! (n)-tableaux,
each gives the same polytabloid, namely e(
1 2 . . . n
) =
1 2 . . . n
.
So S
(n)
= M
(n)
= F.
6
(2) Take = (n 1, 1) and let t
1
, . . . , t
n
be the (n 1, 1)-tableaux
in Example 2.3 above. With the same notation (i.e. omitting the re-
dundant top row in a tabloid), we have
e(t
1
) =
1

2
, e(t
2
) =
2

1
, . . . , e(t
n
) =
n

1
.
Therefore
S
(n1,1)
=

i

1
: 2 i n
_
.
It is easy to see that these vectors are linearly independent, so S
(n1,1)
is (n 1)-dimensional. Problem 5 on the rst problem sheets shows
that S
(n1,1)
is irreducible if the characteristic of F is zero or coprime
to p.
(3) Take = (1
n
) and let t be the (1
n
)-tableau whose single column
has entries 1, 2, . . . , n from top to bottom. If k S
n
then
e(t)k = t

hS
n
hsgn(h)k = t

hS
n
hsgn(h) sgn(k) = e(t) sgn(k).
Hence if s is a (1
n
)-tableau then e(s) = e(t) and S
(1
n
)
aords the
1-dimensional sign representation of S
n
.
Example 2.7. Example 5.2 in James notes on S
(3,2)
is well worth
studying. He shows that S
(3,2)
is the subspace of M
(3,2)
given by impos-
ing all reasonable conditions on a general sum of tabloids. Omitting
the redundant top row in a (3, 2)-tabloid, the reasonable conditions
on

1i<j5

ij i j
M
(3,2)
,
where
ij
F, are
(i)

1i<j5

ij
= 0
(ii)

1j5

ij
= 0 for each i 1, . . . , 5.
Remark 2.8. By the end of the course, at least three equivalent def-
initions of Specht modules will have been seen (four, counting the de-
terminantal version of polytabloids below):
(1) Denition 2.5 above, which denes S

as the subspace of M

spanned by all -tabloids: later we will prove the Standard Basis The-
orem which states that e(t) : t is a standard -tableau is a basis
for S

.
(2) Description by generators and Garnir relations, which denes S

as a quotient of FS
n
(or better, as a quotient of a module

M

, to be
dened later). See 7 and 8 of James notes.
(3) As the subspace of M

given by imposing all reasonable condi-


tions on -tabloids: see Theorem 17.18 in James notes. (Depending
on how time goes, this may only be sketched.)
Remark 2.9. SRB asked, quite reasonably, for more motivation for
the denition of Specht modules. Some historical comments may be
7
helpful. First of all, its worth mentioning that the irreducible charac-
ters of the symmetric groups were constructed by Frobenius in 1904,
long before Spechts construction in 1935 of Specht modules. (See
W. Specht, Die irreduziblen Darstellungen der Symmetrischen Gruppe,
Math. Z. 39 (1935), no. 1, 696711. Specht credits a 1907 paper by
his supervisor I. Schur for many of the ideas.) Frobenius character
formula is naturally expressed in terms of symmetric polynomials, and
Specht constructed his modules inside the polynomial ring in n com-
muting variables. This would also have been a familiar object from
invariant theory.
To give a small example, let R = F[x
1
, x
2
, x
3
, x
4
]. By Spechts de-
nition S
(3,1)
is the submodule of R spanned linearly by the Spechtian
polytabloids
c
_
a b c
d
_
=

x
0
a
x
0
d
x
1
a
x
1
d

x
0
b

x
0
c

= x
d
x
a
.
(In general c(t) is a product of Vandermonde determinants correspond-
ing to the columns of the tableau t.) The map sending x
k
to
k
gives an
isomorphism with the version of S
(3,1)
dened above. To give another
example,
c
_
a b
c d
_
=

x
0
a
x
0
c
x
1
a
x
1
c

x
0
b
x
0
d
x
1
b
x
1
d

= (x
c
x
a
)(x
d
x
b
),
is a FS
4
-generator of S
(2,2)
. Later we will see the Garnir relations
which can express a general polytabloid as a sum of polytabloids e(t)
where t is standard. In Spechts setup, these express (specializations of)
determinantal identities that were probably well known to algebraists
of his time. For example, in S
(2,2)
we have, in James notation,
e
_
2 1
3 4
_
= e
_
1 2
3 4
_
e
_
1 3
2 4
_
.
In Spechts language, this expresses the determinantal identity

1 1
x
2
x
3

1 1
x
1
x
4

1 1
x
1
x
3

1 1
x
2
x
4

1 1
x
1
x
2

1 1
x
3
x
4

,
which is a specialism of the CauchyBinet identity

A C
B D

A
B

C
D

A
B

C
D

.
One feature of Spechts construction is that it makes it very easy
to prove that End
FS
n
S

= F, provided char F ,= 2. (This result also


holds in many cases when char F = 2, but is harder to prove). For
example, let : S
(2,2)
S
(2,2)
be an endomorphism and let
x = c
_
1 2
3 4
_
.
8
We must have (x)(13) = x and x(24) = x, so x is divisible,
in R, by x
1
x
3
and x
2
x
4
. Hence x = x for some x F. Over the
complex numbers, a module is irreducible if and only if it has trivial
endomorphism ring, and so Spechts construction leads to a quick proof
of the irreducibility of Specht modules.
Finally, it is worth noting that Specht worked only over the complex
numbers. It is not clear to me whether he realised that one of the nicest
features of his construction is that it gives modules that are dened in
a uniform way over all elds.
3. James Submodule Theorem
Denition 3.1. Let F be a eld and let be a partition of n N.
We dene a symmetric bilinear form , on M

F
by linear extension
of
s, t =
_
1 if s = t
0 otherwise.
Therefore given x =

s
s and y =

t
t M

, where the
sums all over distinct tabloids, we have
x, y =

s,t

t
s, t =

t
.
When F = C it is also possible to dene the form so that it is Hermit-
ian, i.e. with the notation as above, x, y =

t
. See the remark
about Theorem 3.5 below.
If U M

we shall write U

for the orthogonal space to U, i.e.


U

= v M

: x, v = 0 for all x U.
It is useful to note that the form , is S
n
-invariant, i.e.
xg, yg = x, y
for all x, y M

and g S
n
. (This follows easily from the case when
x and y are tabloids.) Suppose that U M

is an FS
n
-submodule.
Then if v U

and g S
n
we have
vg, x =

v, xg
1
_
= 0
for all x U. Hence U

is also an FS
n
-submodule.
Example 3.2. Let F be a eld. By denition S
(2,2)
F
is spanned linearly
by the polytabloids e(t) where t is a (2, 2)-tableau. But since e(t)h =
sgn(h)e(t) for any h C(t), we need only take those polytabloids that
are column standard. Hence if
x = e
_
1 3
2 4
_
, y = e
_
1 2
3 4
_
, z =
_
1 4
2 3
_
9
then S
(2,2)
= x, y, z. By writing out x, y and z as linear combinations
of tabloids, one can check that z = x+y, and that x and y are linearly
independent. (See also Remark 2.9 above.) Therefore
S
(2,2)
= x, y
F
.
Consider the restriction of , to S
(2,2)
. The Gram matrix of this form
is, with respect to the basis x, y,
G =
_
x, x x, y
y, x y, y
_
=
_
4 2
2 4
_
.
Let p = char F. If p = 2 then the restriction of , is zero. If p = 3
then
G =
_
1 1
1 1
_
and so we see that
S
(2,2)
(S
(2,2)
)

= x + y .
Calculation shows that
x + y =
1 2
3 4
+
1 3
2 4
+
1 4
2 3
+
2 3
1 4
+
2 4
1 3
+
3 4
1 2
is the sum of all distinct (2, 2)-tabloids. Therefore x + y aords the
trivial representation of S
4
.
Exercise:
(1) Show that S
(2,2)
/ x + y aords the sign representation of S
4
.
(2) Let V
4
= (12)(34), (13)(24) S
4
. Show that, over any eld,
the kernel of S
(2,2)
is V
4
, and that, thought of as a representation of
S
4
/V
4

= S
3
, there is an isomorphism S
(2,2)

= S
(2,1)
.
We now give the critical combinatorial lemma needed to prove James
Submodule Theorem. James states this in more general form (see his
Lemma 4.6), but the version below suces for this section. (The asser-
tion about h was part of the proof given in the lecture: it may clarify
the lemma to put it in the statement.)
Lemma 3.3. Let be a partition of n. If u and t are -tableaux then
either ub
t
= 0, or there exists h C(t) such that uh = t and
ub
t
= e(t).
Proof. Suppose that ub
t
,= 0. Then any two numbers i, j that appear in
the same row of u must appear in dierent columns of t. It follows that
there exists h C(t) such that u = th. (First choose h
1
C(t)
so that u and th
1
have the same set of entries in their rst rows, then
choose h
2
C(t) so that h
2
xes the rst rows of u and th
1
, and u and
th
1
h
2
agree in their second rows, and so on.) Hence
ub
t
= thb
t
= tb
t
= e(t).
10
Theorem 3.4 (James Submodule Theorem). Let F be a eld and let
be a partition of n N. If U is an FS
n
-submodule of M

then either
U S

or U (S

.
Proof. Suppose that U is not contained in (S

. Since S

is spanned
by the -polytabloids, there must exist v S

and a -tableau t such


that v, tb
t
,= 0. Now using the S
n
-invariance of , we get that
vb
t
,= 0. But, by Lemma 3.3, M

b
t
= ub
t
= e(t) so e(t) S

.
As remarked after Denition 2.5, S

is cyclic, generated by any single


polytabloid, so we have S

U, as required.
Corollary 3.5. Let F be a eld of characteristic zero and let be a
partition of n N. Then S

F
is irreducible.
Proof. Let U be a non-zero submodule of S

F
. By James Submodule
Theorem, either U = S

, or U S

(S

F
)

. So it will suce to show


that S

(S

F
)

= 0.
If F = Q or F = R, this is clear, because then the bilinear form ,
is an inner product, and so if V is any subspace of M

then V V

= 0.
If F = C the same argument holds, provided we dene the form to be
a Hermitian inner product.
To deal with the general case, James argues as follows: since the
polytabloids span S

, we may choose, as in Example 2.2, a basis of S

consisting of polytabloids. Let G be the Gram matrix of , restricted


to S

, with respect to this basis. The entries of G lie in Z, so the rank


of G does not depend on F. (Generally, if A is a d d-matrix over a
eld F, and K is an extension eld of F, then the rank of F does not
depend on whether the rows are thought of as a vectors in F
d
, or as
vectors in K
d
.) When F = Q the matrix has full rank. Therefore ,
is non-degenerate when restricted to S

, and so S

(S

= 0.
4. Homomorphisms (Part I)
Denition 4.1. Let n N and let and be partitions of n, with
k() and k() parts, respectively. We say that dominates , and
write , if
j

i=1

i

j

i=1

i
for all j min(k(), k()).
Remarks:
(1) The relation denes a partial order on the set of partitions of
n. It is not a total order; for example, (4, 1, 1) and (3, 3) are incompara-
ble. In the lecture I dened the dominance order between partitions of
dierent numbers, but this is problematic since some arbitrary choices
are then needed to make it a partial order. The denition as above will
suce for this course.
11
(2) Let n N and let F be a eld. For each partition of n, let
J() Mat
n
(F) be a nilpotent matrix of Jordan type . For instance,
we may take J() = diag(B

1
, . . . , B

k
) where
B
d
=
_
_
_
_
_
_
0 1 0 0
0 0 1 0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0 0 0 1
0 0 0 0
_
_
_
_
_
_
dd
.
Dene N() = J()
GL
n
(F)
. So J() is the set of all n n nilpotent
matrices over F whose Jordan type is . Then
N() N() .
Here N() means the closure of N() in either the Euclidean topology
if F = R or F = C, or otherwise in the Zariski topology. For example,
the closure of the nilpotent cone
N(2) =
__
a b
c a
_
: a
2
+ bc = 0
_

__
0 0
0 0
__
.
contains the zero-matrix, which is the only element of N(1, 1).
(3) The dominance order appears in many other settings. In lectures
I mentioned the GaleRyser Theorem as a combinatorial example. Re-
call that if A is an m n-matrix then the sequence of row-sums of
A is (

n
j=1
A
1j
, . . . ,

n
j=1
A
mj
) and the sequence of column-sums of A
is (

m
i=1
A
i1
, . . . ,

m
i=1
A
in
). A matrix is said to be zero-one if all its
entries are either 0s or 1s. We also need the conjugate partition to a
partition , which is dened by
(2)

j
= [i :
i
j[.
The Young diagram of

is obtained by reecting the Young diagram


of . For example, (4, 2)

= (2, 2, 1, 1), as seen below.

= .
Theorem (GaleRyser). Let and be partitions of n. There is a
zero-one matrix with row sums and column sums

if and only .
The necessity of the condition is fairly easily seen. Suppose that A is
a zero-one matrix with row sums and column sums

. There are
1
1s in the rst row of A so at least
1
of the column sums are non-zero.
Hence
1

1
. There are
1
+
2
1s in the rst two rows of A and
there are
1

2
columns that can contain at most a single 1, and
2
that can contain two 1s. Hence

1
+
2
(
1

2
) + 2
2
=
1
+
2
.
12
Repeating this argument shows that . The proof that the condition
is also it sucient is harder. (It will follow from a later result on the
structure of Young permutation modules.)
Lemma 4.2. Let n N and let and be partitions of n. If there is
a -tableau u and a -tableau t such that ub
t
,= 0 then .
Proof. By the argument in Lemma 3.3, any two numbers in the same
row of u must appear in dierent columns of t. Let k be the number
of parts of . We shall argue that for each i such that 1 i k, there
exists a permutation h
i
C(t) such that all the numbers in the rst i
rows of u lie in the rst i rows of th
i
. Clearly this implies the result,
because there are
1
+ +
i
numbers in the rst i rows of u, and

1
+ +
i
numbers in the rst i rows of t (or th
i
).
There are
1
numbers in the rst row of u, and they must all lie in
dierent columns of t. Take h
1
C(t) to be the permutation moving
these numbers into the rst row of t.
Suppose inductively we have h
i1
with the required property. Con-
sider the numbers in row i of u. These all lie in dierent columns of
th
i1
. Take k
i
to be the permutation which moves these numbers as
high as possible in th
i1
, while xing the entries of the rst i 1 rows
of u. There is at most one entry in each column of t that has to be
moved up by h
i
, so the entries of row i of u lie in rows 1, . . . , i in
th
i1
k
i
. So we may take h
i
= h
i1
k
i
.
Remark: if = then the permutation h
k
constructed in the proof
is the h required in Lemma 3.3. The argument with the h
i
was sug-
gested in the lecture as an improvement on the traditional hand-waving
approach.
Theorem 4.3. Let and be partitions of n. If there is a non-
zero homomorphism : S

that extends to a homomorphism

: M

then .
Proof. Since S

is spanned linearly by the polytabloids e(t) where t is


a -tableau, there exists a -tableau t such that e(t) ,= 0. Suppose
that
t

{u}

{u}
u
where the sum is over all -tabloids u. Then, since e(t) = tb
t
and
t

b
t
,= 0, there exists a -tabloid u such that ub
t
,= 0. Now apply
Lemma 4.2.
In the lecture the following corollary was stated for modules over the
complex numbers, but the same proof works for any eld of character-
istic zero.
13
Corollary 4.4. Let F be a eld of characteristic zero. If S

= S

F
then = .
Proof. Since F has characteristic zero, the FS
n
-modules are completely
reducible (by Maschkes Theorem). Hence M

F
= S

F
C for some
complementary FS
n
-module C. If S

= S

F
then there is an injective
map : S

F
M

F
. We extend to a map

: M

F
M

F
by setting
C

= 0. By Theorem 4.3 we have . By symmetry , so


= .
Remark: We will prove later that the extension condition in Theo-
rem 4.3 always holds if the ground eld does not have characteristic
2. Therefore Corollary 4.4 also holds over such elds. The corollary
is denitely false over F
2
; for example, S
(n)

= S
(1
n
)
for any n N.
(More generally, if S

is an irreducible Specht module over F


2
then
S


= S

.) The following example gives a related phenomenon.


Example 4.5. Over F
2
there is an isomorphism
S
(5,1,1)

= S
(5,2)

S
(7)
.
Hence there is a non-zero homomorphism from S
(5,1,1)
to S
(5,2)
. See
Question 1 on Sheet 2 for a proof of this decomposition.
If G is any nite group then the number of complex irreducible rep-
resentations of G is equal to the number of conjugacy classes of G.
There are [Par(n)[ conjugacy classes in S
n
, so the Specht modules S

C
form a complete set of non-isomorphic irreducible CS
n
-modules, as
varies over the partitions of n. (In fact this holds over any eld of
characteristic zero.)
Over C any module is completely reducible. So Theorem 4.3 implies
that M

C
is a direct sum of Specht modules labelled by partitions
such that . For example, it follows from Questions 2 and 3 on the
second problem sheet that if n N then
M
(nr,r)
C
= S
(nr,r)
C
+ S
(nr+1,r1)
C
+ + S
(n1,1)
C
+ S
(n)
C
for any r such that 2r n. To give another example, since M
(1
n
)
C
is
isomorphic to the regular representation of CS
n
we have
M
(1
n
)
C

n
(dimS

)S

C
.
5. Construction of simple modules in characteristic p
Overview: Let F be a eld. By James Submodule Theorem, if U S

and U ,= S

then U S

(S

. Hence if S

/(S

) is non-zero
then it is a simple FS
n
-module. When F has characteristic 0 we saw
in Corollary 3.5 that S

(S

= 0 for all partitions . To describe


the situation in prime characteristic p we need the following denition.
14
Denition 5.1. Let be a partition of n with exactly a
i
parts of
length i for each i 1, 2, . . . , n. Given p N, we say that is
p-regular if a
i
< p for all i.
In this lecture and the next we shall prove that the following theorem.
(James proof is in 10 and 11 of [7].)
Theorem. Let F be a eld of prime characteristic p. If is a partition
of n then S

, (S

if and only if is p-regular, and


_
S

(S

: p-regular
_
is a complete set of non-isomorphic simple FS
n
-modules.
Remark: SRB asked for a theoretical reason for this theorem. I men-
tioned that quasi-hereditary algebras as an example of a situation where
something even nicer happens: if A is a quasi-hereditary algebra then
there is a canonical set of standard modules for A such that any projec-
tive A-module is ltered by standard modules, and the top composition
factors of the standard modules form a complete set of non-isomorphic
simple A-modules: see [3] for an introduction to this area.
In particular, this setup holds for polynomial modules for GL
n
(F)
where F is an innite eld; in this case the algebra A is the Schur
algebra (see [6]). For any n N there is a Schur algebra A and
an idempotent e A such that FS
n
= eAe. This means that the
symmetric group algebra FS
n
has some of the desirable properties of
quasi-hereditary algebras.
I should also have mentioned cellular algebras, which generalize quasi-
hereditary algebras in a way modelled on symmetric group algebras.
For an introduction, see [9]. An analogous version of the theorem above
holds for the simple modules of any cellular algebra.
Example 5.2. (1) In Example 3.2 we saw that the restriction of ,
to S
(2,2)
F
is zero if and only if F has characteristic 2. Correspondingly,
(2, 2) is p-regular if and only if p > 2.
(2) Let = (2, 1, 1). Let
t
1
=
2 1
3
4
, t
2
=
1 2
3
4
, t
3
=
1 3
2
4
, t
4
=
1 4
2
3
be the four column-standard (2, 1, 1)-tableaux. Reasoning as in Exam-
ple 3.2 we see that if F is any eld then
S
(2,1,1)
F
= e(t
1
), e(t
2
), e(t
3
), e(t
4
) .
These polytabloids are not linearly independent: one can check that
e(t
1
) = e(t
2
) e(t
3
) + e(t
4
).
15
However e(t
2
), e(t
3
), e(t
4
) are linearly independent, because each in-
volves a tabloid not appearing in the other two. For instance, since 1
and 2 appear in the rst column of both t
3
and t
4
, e(t
2
) is the unique
polytabloid having
1 2
3
4
as a summand. Each polytabloid e(t
i
) has 6 summands, so e(t
i
), e(t
i
) =
6 for each i. To calculate e(t
2
), e(t
3
) we argue that any common
tabloid must have 2 and 3 in its rst row. Calculation shows that
e(t
2
) =
3 2
1
4
+
3 2
4
1
+ , e(t
3
) =
2 3
1
4
+
2 3
4
1
+
so e(t
2
), e(t
3
) = 2. Hence the matrix for , on S
(2,1,1)
is
_
_
6 2 2
2 6 2
2 2 6
_
_
.
It is clear this matrix is zero if the ground eld has characteristic 2,
and otherwise has full rank. Correspondingly (2, 1, 1) is p-regular if
and only if p > 2.
It is an instructive exercise to generalize Example 5.2(2) to the par-
titions (2, 1
n2
). (Question 6 on Sheet 1 is relevant.)
Remark: BK suggested dividing every entry by 2 and then taking the
radical with respect to the new form, and pointed out that this can be
justied even when F has characteristic 2, because the new form is still
symmetric and S
4
-invariant. The matrix of the new form (over F
2
) is
_
_
1 1 1
1 1 1
1 1 1
_
_
so its radical is V = e(t
1
) e(t
2
), e(t
1
) e(t
3
). One can check that
V is simple. (In fact V

= S
(3,1)
/S
(3,1)
(S
(3,1)
)

. There is also an
isomorphism V

= S
(2,2)
.)
This idea is usually applied to integral Specht modules. Given an
integral Specht module S

Z
and a prime p we can dene a ltration:
S

Z
= U
0
U
1
U
2
. . .
where
U
i
= u S

Z
: p
i
[ u, v for all v S

Z
.
If we then reduce mod p we get a ltration of the Specht module S

F
;
the JantzenSchaper formula gives some information about the simple
modules that appear. See [1] for one account. In [4] Fayers used the
16
JantzenSchaper formula to determine all the Schaper layers for Specht
modules labelled by hook partitions (nr, 1
r
) and two-row partitions.
The following two lemmas will imply that S

, (S

if and only
if is p-regular. The lemmas make sense for Specht modules dened
over any eld but are strongest for Specht modules dened over Z.
Lemma 5.3. Let be a partition with exactly a
j
parts of size j for
each j 1, . . . , n. If t and t

are -tableaux then e(t), e(t

) is a
multiple of

n
j=1
a
j
!.
Proof. We shall say that two -tabloids u and v are Foulkes equiv-
alent if it is possible to obtain v from u by reordering the rows
of u. (This term is not standard.) Let
T =
_
u : u is a summand in both e(t) and e(t

)
_
.
If u T and v is Foulkes equivalent to u then there exist per-
mutations h C(t) and h

C(t

) which permutes the rows of u as


blocks for their action, such that
(i) uh = uh

= v,
(ii) sgn h = sgn h

.
(Here (ii) holds because we can swap any two rows of length r in u
by a product of r disjoint transpositions in either C(t) or C(t

).) Hence
T is a union of Foulkes equivalence classes. Moreover, if u T then
the contribution from the Foulkes class of u to e(t), e(t

) is equal
to the size of the class, namely

n
j=1
a
j
!.
Example 5.4. Let
t =
1 2 3
4 5 6
7 8 9
and t

=
1 4 7
2 5 8
3 6 9
One Foulkes class of tabloids common to both e(t) and e(t

) has repre-
sentative u where
u =
1 5 9
4 8 3
7 2 6
since
t(258)(396) =
1 5 9
4 8 3
7 2 6
= t

(23)(45)(79).
The Foulkes equivalence class of u consists of the six tabloids
_ 1 5 9
3 4 8
2 6 7
,
1 5 9
2 6 7
3 4 8
,
3 4 8
1 5 9
2 6 7
,
3 4 8
2 6 7
1 5 9
,
2 6 7
1 5 9
3 4 8
,
2 6 7
3 4 8
1 5 9
_
.
17
For example, to swap the rst two rows in u we may either apply
(14)(58)(39) C(t) or (13)(45)(89) C(t

). As claimed in the proof,


these permutations both have sign 1.
There is one other Foulkes equivalence class of tabloids common to
both e(t) and e(t

), so e(t), e(t

) = 12.
Lemma 5.5. Let be a partition with exactly a
j
parts of size j for
each j 1, . . . , n. Let t be a -tableau and let t

be the -tableau
obtained from t by reversing each row. Then
e(t), e(t

) =
n

j=1
a
j
!
j
.
Before proving the lemma we shall give an example which shows the
key ideas in the proof. (This example is a trivial variation on the one
given by James in his proof: see [7, Lemma 10.4].)
Example 5.6. Let = (4, 3, 3, 2) and let t and t

be as shown below.
t =
1 2 3 4
5 6 7
8 9 10
11 12
Suppose that u is a summand in both e(t) and e(t

). So u =
th = t

for some h C(t) and h

C(t

). Working down the


rst column of t we see that
(1) 1h

is in row 1 of t

. Hence 1h

= 1h = 1;
(2) 5h

and 8h

are in rows 1, 2 or 3 of t

, and in rows 2, 3 or 4
of th. Hence 5, 8h = 5, 8h

= 5, 8;
(3) 11h is in row 4 of th. Hence 11h = 11h

= 11.
The diagram below shows the positions of the elements in the sets
1, 5, 8, 11 that are known to be permuted by h and h

in the
tableaux t and t

.
1 2 3 4
5 6 7
8 9 10
11 12
4 3 2 1
7 6 5
10 9 8
12 11
Observe that if we remove the rst column of t and the nal entries in
the rows of t

we can repeat this argument, showing that h and h

also
permute the elements in the sets 2, 6, 9, 10, and 3, 7, 10,
12 and nally 4. It is now clear that h = h

.
Conversely we have
C(t) C(t

) = S
{5,8}
S
{6,9}
S
{7,10}
and so, given any k C(t) C(t

), we have tk = tk

. Hence the
tabloids that appear as summands in both e(t) and e(t

) are precisely
18
the tabloids tk for k C(t) C(t

) and so e(t), e(t

) = 2!
3
, as
stated by Lemma 5.5.
We now prove Lemma 5.5. Suppose that th = t

is a summand of
both
e(t) =

hC(t)
thsgn(h) and
e(t

) =

C(t

)
t

sgn(h

).
Claim: h = h

and if i 1, . . . , n and ih = ih

j then i and j lie in


rows of equal length of .
Proof of claim. We shall prove the claim by induction on the number
of columns of the Young diagram of .
Let
1
, . . . ,

be the sequence of distinct part sizes in in decreasing


order of size. Suppose that there are m
i
parts of length
i
. For each
i 1, . . . , let X
i
be the set of entries in the rst column of t lying
in rows of length
i
. Working down the rst column of t we see that
X
i
h

= X
i
and X
i
h = X
i
for each i. It is now clear that the restrictions
of h and h

to X
1
X

agree, and permute entries within rows of


equal length as required by the claim.
Let s be the tableau obtained by removing the rst column from t
and let s

be the tableau obtained by removing the nal entry in each


row of t

. By induction the restrictions of h and h

to s and s

agree,
and permute entries within rows of equal length. This completes the
proof of the claim.
It is easy to see that k C(t) C(t

) if and only if for each i


1, . . . , n, the rows of t (or equivalently, of t

) containing i and ik
have equal length. It therefore follows from the claim that the set of
tabloids u such that u is a summand in both e(t) and e(t

) is
exactly
_
tk : k C(t) C(t

)
_
.
Hence
e(t), e(t

) =

kC(t)C(t

)
tk sgn(k), t

k sgn(k)
= [C(t) C(t

)[.
Finally we argue that the subgroup of C(t) C(t

) permuting entries
within the rows of length j has order a
j
!
j
, and so
[C(t) C(t

)[ =
n

j=1
a
j
!
j
.
This completes the proof of Lemma 5.5.
19
Theorem 5.7. If is a partition of n N and F is a eld of prime
characteristic p then S

F
, (S

F
)

if and only if is p-regular.


Proof. Suppose that has exactly a
j
parts of length j, for each j
1, 2, . . . , n.
By Lemma 5.3, if t and t

are any -tableaux, then we have


e(t), e(t

) =
n

j=1
a
j
!
for some F. By denition, S

F
is the F-linear span of all -
polytabloids, so if is not p-regular then S

(S

.
Conversely, suppose that is p-regular. Let t be any -tableau and
let t

be the tableau obtained by reversing the rows of t. By Lemma 5.5


we have
e(t), e(t

) =
n

j=1
a
j
!
j
,= 0
Hence the restriction of , to S

is non-zero, and so S

, (S

.
To show that the irreducible modules coming from the previous the-
orem are non-isomorphic we need the following result. In its proof, we
shall on two occasions use the corollary of Lemma 3.3, that if t is any
-tableau, then
M

b
t
= e(t) .
(See Question 2 on Sheet 3 for one reason why this result may seem to
be applied in a slightly unexpected way.)
Theorem 5.8. Suppose that and are partitions of n and that
is p-regular. Let V be a submodule of M

. If there exists a non-zero


homomorphism : S

/V then . Moreover, if = and t


is a -tableaux, then e(t) e(t) + V .
Proof. Suppose that has exactly a
j
parts of size j for each j
1, 2, . . . , n. Let t be a -tableau and let t

denote the tableau ob-


tained from t by reversing each of its rows. Since
e(t

)b
t
M

b
t
= e(t) ,
we have e(t

)b
t
= e(t) for some F. Using e(t), t = 1 we can
determine as follows:
= e(t

)b
t
, t = e(t

), tb
t
= e(t

), e(t) =
n

j=1
a
j
!
j
where in the nal step we have used Lemma 5.5. In particular, ,= 0.
Since e(t) generates S

and is non-zero we have e(t) ,= 0. Therefore


e(t

)b
t
= e(t) ,= 0
and so, setting x = e(t

) M

we have xb
t
,= 0. It follows that there
exists a -tabloid u such that a multiple of u + V is a summand
20
of x and ub
t
,= 0. Hence, by Lemma 3.3 . Moreover, if =
then since M

b
t
= e(t) we have x e(t) + V .
Remark: it is worth comparing this theorem with Theorem 4.3. In
Theorem 5.8 we assume that is p-regular, whereas in Theorem 4.3
we make no assumption on , but instead assume that the map :
S

extends to a non-zero map



: M

. In either case,
the conclusion is the same, that .
Corollary 5.9. Let F be a eld of prime characteristic p and let
and be p-regular partitions of n N. If
S

(S

=
S

(S

then = . Moreover,
End
FS
n
_
S

(S

)
_

= F.
Proof. Let be an isomorphism as in the statement of the corollary.
We may lift to a non-zero map : S

/V where V = S

(S

by taking the composition of the maps


S

(S

(S

/V.
Therefore, by Theorem 5.8, we have . Moreover if = then this
theorem we have
e(t) = (e(t) + V )
for some F. Since e(t) generates S

/S

(S

, this implies that


= 1.
Finally we must show that this construction gives every simple rep-
resentation of a symmetric group in prime characteristic. For this we
shall use the following theorem of Brauer. Recall that if G is a nite
group and p is a prime, then g G is said to be p-regular if the order
of g is not divisible by p.
Theorem (Brauer). Let G be a nite group and let F be a splitting
eld for G. The number of isomorphism classes of simple FG-modules
is equal to the number of conjugacy classes of p-regular elements of G.
In fact any eld is a splitting eld for S
n
. I skipped the proof in the
lecture as it is rather technical. See James lecture notes [7, page 40] for
a proof using two results from Curtis & Reiner [2]. James also mentions
an alternative approach using an extension of Brauers result, namely
that over any eld F, the number of isomorphism classes of absolutely
irreducible simple FG-modules is at most the number of conjugacy
classes of p-regular elements (see [2, 82.6]). By the moreover part
of Corollary 5.9 and the following lemma the simple modules we have
constructed are absolutely irreducible.
21
Lemma 5.10. Let G be a nite group, let F be a eld and let U be an
FG-module such that End
F
(U)

= F. Then U is absolutely irreducible.
2
Proof. Let K : F be an extension eld of F. We must show that
U
F
K is an irreducible KG-module. Since End
FG
(U)

= F it follows
from Jacobsons Density Theorem that the image of the action map
FG End
F
(U) is all of End
F
(U). Hence the image of the map KG
End
K
(U
F
K) is also all of End
K
(U). Therefore KG acts transitively
on the non-zero vectors in U
F
K and so U
F
K is an irreducible
FG-module.
It is clear that a conjugacy class in S
n
is p-regular if and only if all the
parts of its labelling partition have size not divisible by p. Therefore,
adopting either approach, the proof that we have a complete set of
simple modules is completed by the following proposition.
Proposition 5.11. Let p 2. The number of p-regular partitions of
n is equal to the number of partitions of n with no part divisible by p.
Proof. The generating function for p-regular partitions is
F(x) =

i1
(1 + x
i
+ x
2i
+ + x
(p1)i
);
a choice of ia where a N from the ith term corresponds to a partition
with a parts of size i. Hence
F(x) =

i1
1 x
ip
1 x
i
=

i1
p| i
1
1 x
i
which is the generating function for partitions with no part divisible
by p.
6. Standard Basis Theorem
Denition 6.1. If t is a standard tableau then we say that the corre-
sponding polytabloid e(t) is standard.
The object of this section is to prove the following theorem giving a
basis for S

F
over any eld F.
Theorem 6.2 (Standard Basis Theorem). Let be a partition of n.
If F is a eld then the standard -polytabloids form a basis for S

.
Remarks: (1) In Example 2.6(2) the basis we found of S
(n1,1)
is in
fact the basis of standard (n 1, 1)-polytabloids. In Example 3.2 we
saw that the Standard Basis Theorem holds for S
(2,2)
.
(2) An important consequence of the Standard Basis Theorem is
that if is a partition and F is a eld then the dimension of S

F
is
2
The converse result, that if U is absolutely irreducible then End
F
(U)

= F also
holds, but we do not need this here.
22
equal to the number of standard -tableaux, and so is independent of
the eld F. The Hook-Formula (see [7, Chapter 20] or the reference
in Problem Sheet 3) is a remarkable combinatorial formula for this
number.
(3) Another corollary of the Standard Basis Theorem is the isomor-
phism
_
r
S
(n1,1)
F
= S
(n1,1
r
)
F
over any eld F: see Question 6 on
Problem Sheet 3.
6.1. Linear independence. To show that the standard polytabloids
are linearly independent we need the following order on tabloids.
Denition 6.3. Let be a partition of n. Given distinct -tabloids
t and s, we dene
t > s
the greatest number appearing in a dierent row
in t than s lies in a lower row of t than s
It is straightforward to show that > is a total order on the set of
all -tabloids for any partition . Since we compare on the largest
number that appears in a dierent position, it is clear that if t is the
greatest -tabloid under > then n must appear in the bottom row of t.
Similarly, n 1 must appear as low down as possible, and so on. For
example, the largest (4, 3, 2)-tabloid under > is
1 2 3 4
5 6 7
8 9
.
Exercise: Show that the 10 distinct (3, 2)-tabloids are, in increasing
order
3 4 5
1 2
<
2 4 5
1 3
<
1 4 5
2 3
<
2 3 5
1 4
<
1 3 5
2 4
<
1 2 5
3 4
<
2 3 4
1 5
<
1 3 4
2 5
<
1 2 4
3 5
<
1 2 3
4 5
.
See Problem Sheet 3, Question 8 for a connection with the colexico-
graphic order on subsets of N.
Lemma 6.4. Let be a partition of n and let s be a column-standard
-tableau. If h C(s) then s > sh.
Proof. Let m = maxj 1, 2, . . . , n : jh ,= j. Suppose that h = m,
and that lies in row r

of s and m lies in row r


m
of s. Then in sh, m
lies in row r

. Since s is standard r

< r
m
. Hence the greatest number
that appears in a dierent row in s and sh appears in a lower row
in s than in sh.
Proposition 6.5. Let be a partition of n and let F be a eld. Work-
ing over F, the set
e(s) : s a standard -tableau
23
is F-linearly independent.
Proof. Suppose, for a contradiction, that

t
e(t)
where the sum is over all standard -tableaux t is a non-trivial linear
dependency. Let t
max
be the greatest t in the order > such that

t
,= 0. Now suppose that s is a standard -tableaux such that
s
,= 0.
Since s and t
max
are row-standard, either s = t
max
or s ,= t
max
and so
t
max
> s. Moreover, if h C(s) then by Lemma 6.4 we have
t
max
s > sh.
Hence
0 =

t
e(t) =
t
t + y
where y M

F
is an F-linear combination of tabloids u such that
t
max
> u. But t and y are linearly independent. This contradic-
tion completes the proof.
6.2. Garnir relations. The spanning part of the proof of Theo-
rem 6.2 is best presented as a result about integral Specht modules.
Given a partition of n, let M

Z
denote the free abelian group with
basis all -tabloids. Dene
S

Z
= e(t) : t a -tableau
Z
to be the free Z-submodule generated by the -polytabloids.
It is clear that M

Z
is a module for the integral group algebra ZS
n
,
which is dened as on page 1, but replacing the eld F with Z. Before
Denition 2.5 we showed that if t is a -tableau and g S
n
then
e(t)g = e(tg), so S

is a Z-free ZS
n
-submodule of M

Z
.
Denition 6.6. Let be a partition of n and let 1 i <
i
. Let t be
a -tableau. Let X be a subset of the entries in column i of t and let
Y be a subset of the entries in column i +1 of t. Let S
X
, S
Y
and S
XY
denote the full symmetric groups on X, Y and X Y , respectively. If
S
XY
=
c
_
j=1
(S
X
S
Y
)g
j
where the union is disjoint, then we say that the element
c

j=1
g
j
sgn(g
j
) ZS
n
is a Garnir element for X and Y .
Note that there is an arbitrary choice of coset representatives in-
volved in this denition. We will see that the choice made is irrelevant
in all our applications of Garnir elements.
24
Example 6.7. Let = (2, 1) and let t =
2 1
3
. Taking X = 2, 3
and Y = 1 we see that a set of coset representatives for S
X
S
Y
in
S
XY
is 1, (12) and (13) so we can take
G
X,Y
= 1 (12) (13).
Observe that
e(t)G
X,Y
=
_
2 1
3

3 1
2
_
_
1 (12) (13)
_
=
2 1
3

3 1
2

1 2
3
+
3 2
1

2 3
1
+
1 3
2
= 0.
This example is a special case of a more general result that leads to
an algorithm for writing an arbitrary polytabloid as an integral linear
combination of standard polytabloids. In the following theorem,

denotes the conjugate of the partition , as dened in (2) on page 11.


Theorem 6.8. Let be a partition of n and let t be a -tableau. Let
1 i <
1
, let X be a subset of the entries in column i of t, and let
Y be a subset of the entries in column i +1 of t. Let G
X,Y
ZS
n
be a
Garnir element for these data. If [X[ +[Y [ >

i
then
e(t)G
X,Y
= 0.
Proof. Let
G
XY
=

gS
XY
g sgn(g).
It is clear that
G
XY
=
_

hS
X
hsgn(h)
__

kS
Y
k sgn(k)
_
G
X,Y
.
Moreover, since e(t)hsgn(h) = e(t) and e(t)k sgn(k) = e(t) for all
h S
X
and k S
Y
we have
e(t)G
XY
= [X[![Y [!e(t)G
X,Y
.
Therefore, since we work over Z and e(t) =

hC(t)
th, it will suce
to show that
thG
XY
= 0
for each h C(t). But if h C(t) then, since [X[ + [Y [ >

i
, there
exist x X and y Y such that x and y lie in the same row of th.
Now
thG
XY
= th(1 (xy))
d

k=1
g
k
sgn(g
k
) = 0
where g
1
, . . . , g
d
is a set of coset representatives for (xy) inside S
XY
.
The theorem follows.
25
6.3. The standard polytabloids span S

Z
. Let be a partition of
n N. If t is a -tableau and h C(t) then e(th) = e(t) sgn(h) =
e(t). So it is clear that
S

Z
= e(t) : t a column standard -tableau .
To show that it suces to take just the standard tableaux we need the
following denition which is the analogue of Denition 6.3 (provided
we identify -tabloids with row-standard -tableaux).
Denition 6.9. Let be a partition of n. Given distinct column
standard -tableaux t and s, we dene
t > s
the greatest number appearing in a dierent
column in t than s lies in a column fur-
ther to the right in t than s.
Note that this is the opposite order to that used by James on page
30 of [7]. See Question 7 on Sheet 3 for one reason for preferring the
order as dened above.
The column standard (2, 2)-tableaux are, in increasing order
3 1
4 2
<
2 1
4 3
<
1 2
4 3
<
2 1
3 4
<
1 2
3 4
<
1 3
2 4
For instance, since we compare on the largest number in a dierent
column, and tableau with 4 in the rst column comes before any tableau
with 4 in the second column.
Lemma 6.10. Let be a partition of n and let be a column standard
-tableau. Then e(s) S

Z
is a Z-linear combination of standard -
polytabloids.
Proof. We may suppose that s is not standard. By induction we may
assume that if t > s then e(t) is a Z-linear combination of standard
-polytabloids.
Since s is not standard there is some row, say row q that is not
increasing. Thus there exists i <
1
such that the entries in rows q up
to

i
of column i of s are
x
q
< x
q+1
< < x

i
;
the entries in rows 1 up to q of column i + 1 of s are
y
1
< y
2
< < y
q
;
and x
q
> y
q
. Set X = x
q
, x
q+1
, . . . , x

i
and Y = y
1
, y
2
, . . . , y
q
.
Let G
X,Y
be a Garnir element corresponding to the sets X and Y . We
may suppose that G
X,Y
=

k
j=1
g
j
sgn(g
j
) where g
1
= 1 corresponds
to the identity coset of S
X
S
Y
in S
XY
. By Theorem 6.8 we have
e(s)G
X,Y
= 0, so
e(s) =
k

j=2
sgn(g
j
)e(sg
j
).
26
It is possible that some of the sg
i
are not column-standard. (We can
assume that the entries of sg
i
in the positions occupied by X and Y
are increasing, but not anything stronger than this: see Example 6.11
below.) For each j 2, . . . , k let u
j
be the column standard tableau
whose columns agree setwise with sg
i
. Then
e(s) =
k

j=2
e(u
j
)
for some appropriate choice of signs. It therefore suces to show that
u
j
> s for each j.
Since (S
X
S
Y
)g
j
,= S
X
S
Y
we have XY g
j
,= . Let x
max
be the
greatest element of XY g
j
. Since g
j
xes all elements not in columns
i and i + 1, and if x X and y Y then x > y, x
max
is the largest
element that appears in a dierent column in s and sg
j
. It follows that
u
j
> s, as required.
Example 6.11. Let t =
2 1
4 3
. Following the proof of Lemma 6.10 we
might take X = 4 and Y = 1, 3. A Garnir element for these sets
is
G
X,Y
= 1 (14) (34),
so by Theorem 6.8 we have
e
_
2 1
4 3
_
= e
_
2 4
1 3
_
+ e
_
2 1
3 4
_
= e
_
1 3
2 4
_
+ e
_
2 1
3 4
_
.
Note that by taking the alternative coset representative
(134) (S
X
S
Y
)(14) = (13) (14)
we would get e
_
2 1
4 3
_
(134) = e
_
2 3
1 4
_
which has its second column
in increasing order. But in both cases the rst column must have 1 in
its second row, so we cannot immediately obtain a column standard
tableau.
The second tableau in the sum is still not standard, so we repeat the
argument with X = 2, 3 and Y = 1. This leads to
e
_
2 1
4 3
_
= e
_
1 2
3 4
_
.
This can also be seen directly from the denition of polytabloids. (See
Question 8 on Sheet 3.)
Remarks:
(1) There are other relations that can be used to write an arbitrary -
polytabloid as a linear combination of standard polytabloids. Fultons
27
quadratic relations (see [5, 7.4]) are often easier to use in practice than
the Garnir relations.
(2) It follows from [8, Proposition 4.1] that if is a partition of n
and t is any column-standard -tableau then
e(t) = e(

t) + x
where

t is the row-standard -tableau obtained from t by rearranging
its rows into increasing order, and x is a Z-linear combination of poly-
tabloids e(t) such that t > s. (By Question 4 on Sheet 1, the tableau

t is in fact standard.)
We are now ready to prove the Standard Basis Theorem. Let F be
a eld and let be a partition of n. By Proposition 6.5 the standard
-polytabloids are linearly independent over F. By the argument at
the start of 6.3, it suces to show that
S

F
= e(t) : t a column-standard -tableau
F
.
Let t be a column-standard -tableau. Thinking of e(t) as an element
in S

Z
, Lemma 6.10 implies that
e(t) =

s
c
s
e(s)
where c
s
Z and the sum is over all standard -tableaux s. But the
same equation holds if we think of e(s) as an element of S

F
, and regard
the c
s
as elements of F. So the standard -tableaux also span S

F
.
We remark that the Standard Basis Theorem also holds for inte-
gral Specht modules: linear independence over Z is clear from the
argument in Proposition 6.5, and the harder part, that the standard
-polytabloids span S

Z
, follows from Lemma 6.10.
We end with two corollaries of the Standard Basis Theorem.
Corollary 6.12. Let be a partition of n N.
(i) The dimension of S

depends only on and not on the eld of


denition; it is equal to the number of standard -tableau.
(ii) The denitions of Specht modules over dierent elds are com-
patible, in the sense that if F is any eld then there is an isomorphism
of FS
n
-modules
S

= S

Z
F
In particular,
S

F
p

=
S

Z
pS

Z
.
The Hook-Formula (see the suggestions for further reading on Sheet 3)
is a remarkable formula for the number of standard -tableau.
28
References
[1] Benson, D. J. Some remarks on the decomposition numbers for the symmetric
groups In The Arcata Conference on Representations of Finite Groups, Proc.
Symp. Pure Math. AMS 47, Part 1 (1987), pp. 381394. Available from http:
//www.abdn.ac.uk/
~
mth192/html/archive/benson.html.
[2] Curtis, C. and Reiner, I. Representation theory of nite groups and asso-
ciative algebras. AMS Chelsea 1962.
[3] Dlab, V., and Ringel, C. The module theoretic approach to quasi-hereditary
algebras. In Representations of Algebras and Related Topics (1992), vol. 168 of
London Mathematical Society Lecture Note Series, Cambridge University Press,
pp. 200224.
[4] Fayers, M. On the structure of Specht modules. Journal of the London Math.
Soc. (Series 2) 67, (2003), 85102.
[5] W. Fulton. Young Tableaux, volume 35 of London Mathematical Society Student
Texts. Cambridge University Press, 1997.
[6] Green, J. A. Polynomial representations of GL
n
, vol. 830 of Lecture Notes in
Mathematics. Springer-Verlag, Berlin, 1980.
[7] James, G. D. The representation theory of the symmetric groups, vol. 682 of
Lecture Notes in Mathematics. Springer, Berlin, 1978.
[8] Wildon, M. Vertices of Specht modules and defect groups of the symmetric
group. J. Alg. 323, (2010), 22432256.
[9] Xi, C. Cellular algebras, Advanced School and Conference on Representation
Theory and Related Topics (9 - 27 January 2006). Available from http://www.
math.jussieu.fr/
~
keller/ictp2006/lecturenotes/xi.pdf.

Potrebbero piacerti anche