ArticlePDF Available

A short proof on the transition matrix from the Specht basis to the Kazhdan-Lusztig basis

Authors:

Abstract

We provide a short proof on the change-of-basis coefficients from the Specht basis to the Kazhdan-Lusztig basis, using Kazhdan-Lusztig theory for parabolic Hecke algebra.
ROCKY MOUNTAIN
JOURNAL OF MATHEMATICS
Volume 51 (2021), No. 5, 1671–1680
DOI: 10.1216/rmj.2021.51.1671 c
Rocky Mountain Mathematics Consortium
A SHORT PROOF ON THE TRANSITION MATRIX
FROM THE SPECHT BASIS TO THE KAZHDAN–LUSZTIG BASIS
MEE SE ONG IM
We provide a short proof on the change-of-basis coefficients from the Specht basis to the Kazhdan–
Lusztig basis, using Kazhdan–Lusztig theory for the parabolic Hecke algebra.
1. Introduction
It is well-known that the irreducible representations for the symmetric group
6d
are the Specht modules
parametrized by the set of standard Young tableaux consisting of
d
boxes. The Specht module has a
purely combinatorial basis (called the Specht basis) described in terms of certain alternating sums of
so-called tabloids.
For the Young tableaux consisting of two rows of equal size, Russell and Tymoczko in [8] compare the
Specht basis with another combinatorial basis (called the web basis) which arises from Temperley–Lieb
algebra and knot theory. In their context, the web basis is a reincarnation of the Kazhdan–Lusztig basis,
but this is not true in general. Their main result is a combinatorial model that gives a different proof
to a special case of a classic theorem by Naruse [7, Theorem 4.1]that the change-of-basis matrix is
unitriangular, altogether with some vanishing conditions. The unitriangularity result is also found in [4],
but without vanishing conditions.
In this paper, we give a short proof of [7, Theorem 4.1]using Kazhdan–Lusztig theory for the parabolic
Hecke algebra. We also notice that the argument applies to all classical types, with a nonstandard notion
of tableaux. For example, in type B we use certain centro-symmetric tableaux which are not the bi-
tableaux that parametrize the irreducibles. That is to say, the analog of the Specht modules here are not
the irreducibles. However, we hope that they correspond to the top nonvanishing cohomology of the
Springer fiber corresponding to the parabolic subgroup. Furthermore, we postpone the details for type D
to future work. In this manuscript, we introduce a Specht basis for this module using certain alternating
sums of centro-symmetric tabloids. We then prove a unitriangular theorem regarding the change-of-basis
matrix between the Specht basis and the Kazhdan–Lusztig basis (see Theorem 3.1).
I would like to thank Chun-Ju Lai, Arik Wilbert, and Jieru Zhu for insightful discussions. I would also like to thank the referee
for constructive feedback on this manuscript. I thank the general manager Jim Hutchinson’s front desk staff at Homewood
Suites by Hilton in New Windsor, NY for their unwavering support to the author and access to quiet work areas during the
covid-19 lockdown period in the United States. I was partially supported by the Summer Collaborators Program at the School
of Mathematics in the Institute for Advanced Study in Princeton, NJ.
2010 AMS Mathematics subject classification: primary 05E10; secondary 20C08, 20C30.
Keywords and phrases: Kazhdan–Lusztig basis, Specht module, parabolic Hecke algebra.
Received by the editors on December 9, 2019, and in revised form on February 9, 2021.
1671
1672 MEE SEONG IM
Figure 1. Dynkin diagrams of type Adand Bd.
1 2
...
d0 1
...
d1
===
Type Ad(d1)Type Bd(d2)
2. Combinatorics
2.1. Weyl groups.
Throughout this manuscript, let
8
be the type A or B. Let
W8
d
be the Weyl group
of type 8dassociated with the Dynkin diagrams in Figure 1.
We define
(2-1) [d] := [1,d] ∩ Z,d] := [−d,d] ∩ Z.
Let
(2-2) IA
d:= [d]and IB
d:= [0,d1] ∩ Z.
We denote by
S8
d= {s8
i:iI8
d}
, the corresponding generators of
W8
d
as Coxeter groups. Define
`8:W8
dZ0
to be the length function, where
`8(u)
of an element
u
in the Weyl group
W8
d
is the
minimal number lsuch that ucan be written as a product of lgenerators.
For any set
I
, let
Perm(I)
be the group of permutations on
I
, and let
(i,j)Perm(I)
be the trans-
position for
i,jI
. It is a standard fact that
W8
d
can be identified as a group of certain permutations
(see [2]). Here we treat them as subgroups of Perm(d])as follows:
(2-3) WB
d:= {gPerm(d]):g(i)= −g(i)for all i},
WA
d1:= {gWB
d:neg(g)=0},
where the function neg counts the total number of negative entries of {1,...,d}, i.e.,
(2-4) neg :WB
dZ0,g7→ #{k∈ [d] : g(k) < 0}.
The generators can be identified with the following permutations in d]:
(2-5) si=sA
i=sB
i=(i,i+1)(i,i1)for 1 id1,
s0=sB
0=(1,1).
It is sometimes convenient to use the one-line notation
(2-6) w≡ |w(1), w(2), . . . , w(d)|for wW8
d.
We remark that the element
w
is uniquely determined by these
d
values due to the centro-symmetry
condition g(i)= −g(i)for all i.
A partial Bruhat order
on
W8
d
is defined as follows: for
u, w W8
d
,
uw
if and only if
u=sra1· · · srab
, a reduced expression, for some 1
a1<· · · <abm
, where
w=sr1· · · srm
is a
THE TRANSITION MATRIX FROM THE SPECHT BASIS TO THE KAZHDAN–LUSZTIG BASIS 1673
reduced decomposition for
w
. That is,
uw
if any reduced expression for
w
contains a subexpression
which is a reduced expressions for u.
Example 2.1. Let u=(1,2)and w=(1,2)(3,4). Then it is clear that uw.
Example 2.2. Let u=s3s1and w=s1s3s2s1. Then uw.
2.2. Parabolic Hecke algebras.
Let
H8=H(W8
d)
be the Hecke algebra of
W8
d
over
C[q,q1]
, where
8=
A or B. As a free module,
H8
has a basis
{H8
w:wW8
d}
. The multiplication in
H8
is determined
by
(2-7) H8
wH8
x=H8
wxif `8(wx)=`8(w) +`8(x),
(H8
s)2=(q1q)H8
s+H8
eif sS8
d,
where
eW8
d
is the identity, and hence
H8
e
is the identity element in
H8
. By a slight abuse of notation,
we use the same symbol
H8
i
(1
id
1) to denote the element
H8
s8
i
H8
. We write
H8
0:= H8
s8
0
when 8=B. It is a standard fact that H8is generated as a C[q,q1]-algebra by {H8
i:iI8
d}.
For each subset
JI8
d
, denote the corresponding parabolic subgroup and parabolic Hecke algebra
of W8
dby
(2-8) W8
J= hs8
j:jJiand H8
J:= H(W8
J)= hH8
j:jJi,
respectively. We also denote the set of shortest right coset representatives for W8
J\W8
dby
(2-9) D8
J= {wW8
d:`8(wg)=`8(w) +`8(g)for all gW8
J}.
If xyW8
J, then we give it a total order xDyon D8
J.
Denote the induced trivial module corresponding to JI8
dby
(2-10) M8
J=C[q,q1] ⊗H8
J
H8,
where C[q,q1]is regarded as a right H8-module by setting
(2-11) f·H8
i=q1ffor fC[q,q1].
The module M8
Jadmits a standard basis
(2-12) {Mw=1H8
w:wD8
J}.
We note that
M8
J
can also be identified as the right
H8
-module
yλH8
under the assignment
Mw7→ yJH8
w
,
where
yJ=PwW8
Jq`(w) H8
w
, on which
H8
acts by a right multiplication. The action of
H8
on
M8
J
below follows from (2-7) and a straight-forward calculation:
Mw·H8
i=
Mwsiif wsiD8
J, `8(wsi)>`8(w),
Mwsi+(q1q)Mwif wsiD8
J, `8(wsi)<`8(w),
q1Mwif wsi6∈ D8
J.
Following [9, Theorem 3.1],M8
Jadmits a Kazhdan–Lusztig basis {Mw:wD8
J}such that
(2-13) Mw=X
xD8
J
m8
x,w(q1)Mx,
1674 MEE SEONG IM
where m8
x,w Z[q1]. We also define polynomials p8
w,x(q1)Z[q1]such that
(2-14) Mw=X
xD8
J
p8
w,x(q1)Mx.
Below we recall some properties of the (inverse) parabolic Kazhdan–Lusztig polynomials.
Lemma 2.3. We have the following:
(a) If x 6≤ wwith respect to the Bruhat order on W 8
d,then m8
x,w =0=p8
x,w.
(b) m8
x,x=1=p8
x,x.
Proof. It follows from [9, Theorem 3.1]that the matrix
(m8
x,w(q1))x,w
is upper unitriangular, so its
inverse matrix (p8
x,w(q1))x,w is also upper unitriangular.
2.3. Young tableaux.
Now we want to associate to each parabolic subgroup
W8
JW8
d
a “Young
tableau” beyond type A. The idea here is to use the corresponding composition (cf. [1;3]), where in type
A, a composition of dis an integer vector of positive integers summing to d. To be precise, we set
(2-15) 3A(d):= G
n0
3A(n,d)and 3B(d):= G
n0
3B(n,d),
where the legit n-part (signed) compositions are defined as
3A(n,d)=nλ=i)i∈[n]Zn
>0:
n
P
i=1
λi=do,(2-16)
3B(n,d)=
i)i∈[±r]Zn
>0:λ01+2Z,Pr
i=−rλi=2d+1,
λi=λiiif n=2r+1,
{λ3B(2r+1,d):λ0=1}if n=2r.
(2-17)
In other words, the corresponding parabolic subgroups W8
Jare generated by
(2-18) SA
d− {sλ1,sλ1+λ2,...,sλ1+···+λn1}for λ3A(n,d),
and generated by
(2-19) SB
d{sλ1,sλ1+λ2,...,sλ1+···+λr1}if n=2r,
sλ\
0
,sλ\
0+λ1,...,sλ\
0+λ1+···+λr1if n=2r+1,for λ3B(n,d)
(letting
λ\
0:= bλ0/2c
, where
bxc
is the greatest integer less than or equal to
x
). We also write
W8
λ
and
H8
λ
to denote the corresponding parabolic subgroups and the parabolic Hecke algebras, respectively. Note
that rfor 8=B corresponds to nfor 8=A.
For type A, any composition λ=1, . . . , λn)3A(n,d)defines a Young subgroup
WA
λ'6λ1×6λ2×···×6λn.
That is, any two compositions giving the same Young subgroup (and hence parabolic Hecke algebra)
correspond to the same partition. We denote the set of type A partitions by
5A(d)=Fn15A(n,d),
where
(2-20) 5A(n,d)= {λ3A(n,d):λ1λ2≥ · · · λn}.
THE TRANSITION MATRIX FROM THE SPECHT BASIS TO THE KAZHDAN–LUSZTIG BASIS 1675
However, for type B, a composition λ=0, . . . , λr)3B(n,d)defines a Young subgroup
WB
λ'WB
λ\
0
×6λ1×6λ2×···×6λr.
In particular,
WB
λ
is always a product of symmetric groups when
n
is even. Hence, compositions
λ, µ
describe the same Young subgroup (and hence parabolic Hecke algebra) if
λ0=µ0
and also
1, . . . , λr)
corresponds to the same partition as
1, . . . , µr)
. We denote the set of type B “partitions” by
5B(d):=
Fn15B(n,d), where
(2-21) 5B(n,d)= {λ3B(n,d):λ1λ2≥ · · · λr=λd(n1)/2e}
and dxeis the least integer greater than or equal to x.
Example 2.4.
Let
d=
3. There are 4 and 8 subsets of
IB
d
for
n
even and odd, respectively, and hence
we have
r n λ3B(3)WB
JJDB
J
1 2 (3,1,3)WB
3{0,1,2} {e}
2 4 (2,1,1,1,2)hsB
0,s2i ' 62×62{0,2} hs1i
2 4 (1,2,1,2,1)hsB
0,s1i ' WB
2{0,1} hs2i
3 6 (1,1,1,1,1,1,1)hsB
0i ' 62{0} hs1,s2i
0 1 (7)WB
3{0,1,2} {e}
1 3 (3,1,3)hs1,s2i ' 63{1,2} hsB
0i
1 3 (2,3,2)hsB
0,s2i ' 62×62{0,2} hs1i
1 3 (1,5,1)hsB
0,s1i ' WB
2{0,1} hs2i
2 5 (2,1,1,1,2)hs2i ' 62{2} hsB
0,s1i
2 5 (1,2,1,2,1)hs1i ' 62{1} hsB
0,s2i
2 5 (1,1,3,1,1)hsB
0i ' 62{0} hs1,s2i
3 7 (1,1,1,1,1,1,1){e}WB
3
Let
λ38(n,d)
. For now we assume that
8=
A. Let its corresponding Young diagram be
λ=
{(i,j):j≤ −
1
,
1
iλj}
, which is a corner justified set of boxes. A Young tableau of shape
λ
(or
λ-tableau for short) is a bijection
T:λ→ [d].
For example, when
λ=(m,m)
or
(m,m,m)
, a
λ
-tableau represents a 2
×m
or 3
×m
grid, respectively,
whose boxes are filled in by every number, exactly once, from 1 to
d=
2
m
or
d=
3
m
, respectively. We
also write
Sh(T)=λ
to mean the composition corresponding to the function
T
, and
Sh(λ)
to mean the
set of all Young tableaux having composition
λ
. The set of all Young tableaux admits a (right) action of
the symmetric group WA
d1=6dby permuting the letters that are filled in the boxes.
A Young tableau is called row standard if the entries in each row are increasing; it is called standard
if the entries in each row and each column are increasing. Denote by
rStdA(λ)
and
StdA(λ)
as the sets of
row standard and standard Young tableaux of shape
λ
, respectively. We remark that the size of
StdA(λ)
is counted by the hook formula, and hence
(2-22) #StdA((m,m)) =1
m+12m
mand #StdA((m,m,m)) =2·(3m)!
m! · (m+1)! · (m+2)!.
1676 MEE SEONG IM
We say
SrStdA(λ)
is in standard form if integers from 1 to
d
are written in each box, increasing
along each row from left to right, and then moving down the consecutive rows from top to bottom.
Write
rStdA
J(λ)
to be those Young tableaux in
rStdA(λ)
containing standard form Young tableau
S
,
together with those of the form S·w, where wDA
J.
It is well-known that the following assignment is a bijection:
(2-23) rStdA
J(λ) DA
J,T7→ wT,
where DA
J= hsλ1,sλ1+λ2,...,sλ1+···+λn1i.
Example 2.5.
Let
λ=(
3
,
2
)3A(
2
,
5
)
and
J= {
1
,
2
,
4
}
. Then
WA
J= hs1,s2,s4i
and
DA
J= hs3i
. Let
T=123
4 5 rStdA
J(λ)
, which is in standard form. Then
wT=e
. Furthermore,
T·s3=124
3 5
, and
wT·s3=(3,4).
We now generalize this bijection to type
8=
B by defining the Young tableaux of classical type. For
λ3B(n,d), denote the corresponding Young diagram by
(2-24) λ=(i,0): −λ\
0iλ\
0t {±(i,j):j≤ −1,0iλj1}if n=2r+1,
(i,j):j≤ −1,1iλj}if n=2r.
Note that when
n
is even, the corresponding Young diagram is just a type A Young diagram, with a copy
obtained by rotating it 180 degrees.
Example 2.6.
For
d=
2,
5B(
2
,
2
)= {(
2
,
1
,
2
)}
and
5B(
4
,
2
)= {(
1
,
1
,
1
,
1
,
1
)}
, and hence the Young
diagrams are
and ,
respectively. For d=3, we have
λ5B(3) (7) (1,5,1) (2,3,2) (3,1,3) (1,1,3,1,1) (1,2,1,2,1) (1,1,1,1,1,1,1)
λ
A
λ
-tableau (Young tableau) of type
8=
B is again a bijection
T:λd]
for
n
odd, and
T:λd]\{
0
}
for
n
even. We also write
Sh(T)=λ
by a slight abuse of notation, and we write
Sh(λ)
to be the set of
all Young tableaux with composition
λ
. Such a
λ
-tableau is called row standard if the entries in each
row of
λ
-tableau are increasing, while it is called standard if the entries in each row and each column of
λ
-tableau are increasing. Denote by
rStdB(λ)
and
StdB(λ)
the sets of type B row standard and standard
λ-tableaux, respectively.
For
n
even, we say
SrStdB(λ)
is in standard form if integers
{−r,...,r}\{
0
}
are written in each
box, increasing along each row from left to right, and moving down the rows from top to bottom. For
n
odd, we say
SrStdB(λ)
is in standard form if integers from
r
to
r
are written in each box, increasing
along each row from left to right, and moving down the rows from top to bottom.
Write
rStdB
J(λ) rStdB(λ)
to be those Young tableaux containing standard form Young tableau
S
,
together with those of the form S·w, where wDB
J.
THE TRANSITION MATRIX FROM THE SPECHT BASIS TO THE KAZHDAN–LUSZTIG BASIS 1677
The following assignment is an obvious bijection:
(2-25) rStdB
J(λ) DB
J,T7→ wT,
where
DB
J=hsλ1,sλ1+λ2,...,sλ1+···+λr1iif nis even,
hsλ\
0
,sλ\
0+λ1,...,sλ\
0+λ1+···+λr1iif nis odd.
Example 2.7. Given tableau T=3210123in standard form, we have wT=e.
Example 2.8.
Given tableau
T=32
1 0 1
2 3
in standard form, we have
wT=e
. As for
T·s1=31
2 0 2
1 3
,
wT·s1=s1=(1,2)(1,2).
We define RrTfor R,TrStd8(λ) if there exists transpositions sr1,...,srmsuch that
R=S·sra1· · · srab,T=S·sr1· · · srm,
where 1
a1<· · · <abm
,
S
is in standard form, and
sra1· · · srab
and
sr1· · · srm
are reduced expressions.
2.4. Specht module.
We refer to [10;5;6] for an extensive background on Specht modules. For now
we assume that
8=
A. Let
Sh(λ)
be the set of all Young tableaux of composition
λ
. Let
{T}
be the
equivalence class (called a tabloid) of the
λ
-tableau
T
under the (right) action of
WA
λWA
d
. That is,
two tableaux are row equivalent if one may be obtained from the other from a permutation within rows,
and
{T}
is the row equivalence class of the tableau
T
. It is clear that each class contains a unique row
standard λ-tableau.
Example 2.9. All tabloids of composition (3,2)of type A are of the form
n123
4 5 o,n124
3 5 o,n125
3 4 o,n134
2 5 o,n135
2 4 o,
n145
2 3 o,n234
1 5 o,n235
1 4 o,n245
1 3 o,n345
1 2 o.
We define the Specht vector corresponding to TSh(λ) by
(2-26) vA
T=X
wcol(T)
(1)`A(w){T·w},
where
col(T)
is the subset of the symmetric group
6d
consisting of the permutations which reorder the
columns of T. For TStd8(λ),RrStd8(λ), and λ38(d), define cA
R,TZby
(2-27) vA
T=X
RrStdA(λ)
cA
R,T{R}.
Note that vA
Tform a standard basis for the Specht module indexed by standard λ-tableaux.
We now generalize this to type B by defining
(2-28) vB
T=X
RrStdB(λ)
cB
R,T{R}for TSh(λ), λ 3B(n,d), and Tin standard form.
1678 MEE SEONG IM
Example 2.10.
Let
λ=(
3
,
2
)
and
J= {
1
,
2
,
4
}
. Then for
8=
A, the standard Specht vectors are of the
form
vA
123
4 5
=n123
4 5 on423
1 5 on153
4 2 o+n453
1 2 o,
vA
124
3 5
=n124
3 5 on324
1 5 on154
3 2 o+n354
1 2 o.
In type B, for
T=54
321
123
4 5
and T·s3=53
421
124
3 5
the standard Specht vectors are of the form
vB
T=
54
321
1 2 3
4 5
51
324
4 2 3
1 5
24
351
1 5 3
4 2
+
21
354
4 5 3
1 2
,
vB
T·s3=
53
421
1 2 4
3 5
51
423
3 2 4
1 5
23
451
1 5 4
3 2
+
21
453
3 5 4
1 2
.
Lemma 2.11. If c8
R,T6= 0, then wRwTwith respect to the Bruhat order.
Proof. Let
c8
R,T6=
0. Recall the bijection
rStd8
J(λ) D8
J
in (2
-
23) and (2
-
25), where
T7→ wT
. Then
there exists
wcol(T)
such that
R=T·w
. If
w=Id
, then we are done. So assume
w6= Id
. Since
TStd8(λ)
, permute the entries within each row of
R
such that it is in
rStd8(λ)
. Without loss of
generality, rename this representative by
R
. Note that
R
may not be in
Std8(λ)
since
T
is in
Std8(λ)
.
So for
T7→ wT
,
R=T·w7→ wT·w=wR
,
`8(wTw) `8(w)
since
w
is not the identity permutation
and wTD8
Jis a reduced word while wW8
d\D8
J. This shows RrTimplies wRwT.
Example 2.12. Let r=2, n=5. So d=3,
T=
3
2
1 0 1
2
3
,T·s1=
3
1
2 0 2
1
3
,T·s1s2=
2
1
3 0 3
1
2
,
and
vB
T= {T} − {T·s2}, v B
T·s1= {T·s1} − {T·s1(s2s1s2)}, v B
T·s1s2= {T·s1s2} − {T·s1s2s1}.
We denote the (classical) Specht module over Ccorresponding to λ58(d)by
(2-29) S8
λ=M
TStd8(λ)
Cv8
T,
where the (right) action of
W8
d
is given by
v8
T·si=v8
T·si
. We call
{v8
T:TStd8(λ)}
the tableaux basis
for S8
λ.
THE TRANSITION MATRIX FROM THE SPECHT BASIS TO THE KAZHDAN–LUSZTIG BASIS 1679
Note that for
TStd8(λ)
, it is not guaranteed that
T·siStd8(λ)
, and hence
v8
T·si
is, in general, not
a single Specht vector corresponding to a standard
λ
-tableau, but a linear combination of Specht vectors.
3. The change-of-basis matrix
Now we embed the Specht module
S8
J
into the induced trivial module
M8
J
by identifying the tabloid
{R}
for RrStd8(d), with the standard basis element M8
wRin M8
J, where wRD8
J.
Under the embedding, we denote by
a8
x,T
(for
TrStd8(λ)
and
xD8
J
) the change-of-basis coeffi-
cients from the Specht basis to the Kazhdan–Lusztig basis, i.e.,
(3-1) v8
T=X
xD8
J
a8
x,TMx.
Recall that is the Bruhat order on D8
JW8
d.
Theorem 3.1. For J I8
d,fix total orders Don D8
Jand ron rStd8
J(λ) satisfying
xyD8
JxDy and wRwTD8
JRrT for R,TrStd8
J(λ),
respectively. The matrix (a8
x,T)x,Tis upper unitriangular in the sense that
a8
wT,T=1 for TrStd8
J(λ), a8
x,T=0 if x6DwT.
Moreover,a8
x,T=0if x 6≤ wT.
Note that if a partial ordering does not exist, then the appropriate change-of-basis coefficient is zero.
Proof. Combining Lemmas 2.3 and 2.11, we obtain the following:
(3-2) v8
T=X
RrStd8(λ)
wRwT
c8
R,TX
xwR
m8
x,wR(1)Mx=X
xwTX
RrStd8(λ)
wRwT
c8
R,Tm8
x,wR(1)Mx.
It follows from the unitriangularity of (c8
R,T)and (m8
wT,wR(1)) that
a8
wT,T=X
RrStd8(λ)
wRwT
c8
R,Tm8
wT,wR(1)=1.
Also, if x6≤DwTthen x6wT, and so a8
x,T=0.
Remark 3.2.
For type A, Theorem 3.1 recovers [7, Theorem 4.1]for arbitrary
J
. If
J= [
2
n] − {n}
, then
it also recovers [8, Theorems 5.5 and 5.7].
References
[1]
H. Bao, J. Kujawa, Y. Li, and W. Wang, “Geometric Schur duality of classical type”,Transform. Groups
23
:2 (2018),
329–389.
[2] A. Björner and F. Brenti, Combinatorics of Coxeter groups, Grad. Texts in Math. 231, Springer, 2005.
[3]
R. Dipper and G. James, “Representations of Hecke algebras of general linear groups”,Proc. London Math. Soc.
(
3
)52
:1
(1986), 20–52.
1680 MEE SEONG IM
[4]
A. M. Garsia and T. J. McLarnan, “Relations between Young’s natural and the Kazhdan–Lusztig representations of
Sn
,
Adv. in Math. 69:1 (1988), 32–92.
[5] G. D. James, The representation theory of the symmetric groups, Lecture Notes in Math. 682, Springer, 1978.
[6]
G. James and A. Kerber, The representation theory of the symmetric group, Encyclopedia of Mathematics and its Applica-
tions 16, Addison-Wesley, Reading, MA, 1981.
[7] H. Naruse, “On an isomorphism between Specht module and left cell of Sn,Tokyo J. Math. 12:2 (1989), 247–267.
[8]
H. M. Russell and J. S. Tymoczko, “The transition matrix between the Specht and web bases is unipotent with additional
vanishing entries”,Int. Math. Res. Not. 2019:5 (2019), 1479–1502.
[9] W. Soergel, “Kazhdan–Lusztig-Polynome und eine Kombinatorik für Kipp-Moduln”,Represent. Theory 1(1997), 37–68.
[10] W. Specht, “Die irreduziblen Darstellungen der symmetrischen Gruppe”,Math. Z. 39:1 (1935), 696–711.
MEE SE ON G IM:meeseongim@gmail.com
Department of Mathematics, United States Naval Academy, Annapolis, MD, United States
RMJ — prepared by msp for the
Rocky Mountain Mathematics Consortium
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
We compare two important bases of an irreducible representation of the symmetric group: the web basis and the Specht basis. The web basis has its roots in the Temperley-Lieb algebra and knot-theoretic considerations. The Specht basis is a classic algebraic and combinatorial construction of symmetric group representations which arises in this context through the geometry of varieties called Springer fibers. We describe a graph that encapsulates combinatorial relations between each of these bases, prove that there is a unique way (up to scaling) to map the Specht basis into the web representation, and use this to recover a result of Garsia-McLarnan that the transition matrix between the Specht and web bases is upper-triangular with ones along the diagonal. We then strengthen their result to prove vanishing of certain additional entries unless a nesting condition on webs is satisfied. In fact we conjecture that the entries of the transition matrix are nonnegative and are nonzero precisely when certain directed paths exist in the web graph.
Article
Full-text available
This is a generalization of the classic work of Beilinson, Lusztig and MacPherson. We show that the quantum algebras obtained via a BLM-type stabilization procedure in the setting of partial flag varieties of type B/C are two (modified) coideal subalgebras of the quantum general linear Lie algebra. These coideal algebras and a new variant of canonical bases arose in a recent approach to Kazhdan-Lusztig theory of type B developed by two of the authors, where a Schur-type duality between such a coideal algebra and Iwahori-Hecke algebra of type B was also established. Here we provide a geometric realization of this Schur-type duality. The monomial bases and canonical bases of the modified coideal algebras are constructed for the first time. Moreover, precise connections between the two modified coideal algebras and their distinguished bases are developed.
Article
Background from representation theory.- The symmetric group.- Diagrams, tableaux and tabloids.- Specht modules.- Examples.- The character table of .- The garnir relations.- The standard basis of the specht module.- The branching theorem.- p-regular partitions.- The irreducible representations of .- Composition factors.- Semistandard homomorphisms.- Young's rule.- Sequences.- The Littlewood-richardson rule.- A specht series for M?.- Hooks and skew-hooks.- The determinantal form.- The hook formula for dimensions.- The murnaghan-nakayama rule.- Binomial coefficients.- Some irreducible specht modules.- On the decomposition matrices of .- Young's orthogonal form.- Representations of the general linear group.
Article
This article gives a selfcontained treatment of the theory of Kazh- dan-Lusztig polynomials with special emphasis on ane reflection groups. There are only a few new results but several new proofs. We close with a conjectural character formula for tilting modules, which formed the starting point of these investigations.
Article
Our main result here is that, under a suitable order of standard tableaux, the classical representation of Sn introduced by Young (in “The Collected Papers of Alfred Young, 1873–1940,” Univ. of Toronto Press, Toronto) (QSA IV), and usually referred as the Natural representattion, the the more recently discovered (Invent. Math.53 (1979), 165–184) Kazhdan-Lusztig (K-L) representation are related by an upper triangular integral matrix with unit diagonal elements. We have been led to this discovery by a numerical exploration. We noted it in each of the irreducible representations of Sn up to n = 6. The calculations in these cases were carried out by constructing the corresponding Kazhdan-Lusztig graphs from tables (M. Goresky, Tables of Kazhdan-Lusztig polynomials, unpublished) of K-L polynomials. To extend the calculations to n = 7 we have used graphs obtained by means of an algorithm given by Lascoux and Schützenberger (Polynomes de Kazhdan & Lusztig pour les Grassmanniennes, preprint). Remarkably, the same property holds also for these graphs. These findings appear to confirm the assertion made by these authors that their algorithm does indeed yield K-L graphs. For the case of hook shapes we have obtained an explicit construction of the transforming matrices, a result which was also suggested by our numerical data. For general shapes, the transforming matrices are less explicit and our proof is based on certain properties of the Kazhdan-Lusztig representations given in their article (Invent. Math.53 (1979), 165–184) and on a purely combinatorial construction of the natural representation.