Kronecker positivity and 2-modular representation theory

By C. Bessenrodt, C. Bowman, and L. Sutton

Abstract

This paper consists of two prongs. Firstly, we prove that any Specht module labelled by a 2-separated partition is semisimple and we completely determine its decomposition as a direct sum of graded simple modules. Secondly, we apply these results and other modular representation theoretic techniques on the study of Kronecker coefficients and hence verify Saxl’s conjecture for several large new families of partitions. In particular, we verify Saxl’s conjecture for all irreducible characters of which are of 2-height zero.

Introduction

This paper brings together, for the first time, the two oldest open problems in the representation theory of the symmetric groups and their quiver Hecke algebras. The first problem is to understand the structure of Specht modules and the second is to describe the decomposition of a tensor product of two Specht modules — the Kronecker problem.

Kronecker positivity

The Kronecker problem is not only one of the central open problems in the classical representation theory of the symmetric groups, but it is also one of the definitive open problems in algebraic combinatorics as identified by Richard Stanley in Reference Sta00. The problem of deciding the positivity of Kronecker coefficients arose in recent times also in quantum information theory Reference Kly04Reference CM06Reference CHM07Reference CDW12 and Kronecker coefficients have subsequently been used to study entanglement entropy Reference CSW18.

A new benchmark for the Kronecker positivity problem is a conjecture of Heide, Saxl, Tiep and Zalesskii Reference HSTZ13 that was inspired by their investigation of the square of the Steinberg character for simple groups of Lie type. It says that for any there is always a complex irreducible character of whose square contains all irreducible characters of as constituents. For a triangular number, an explicit candidate was suggested by Saxl in 2012: Let denote the th staircase partition. Phrased in terms of modules, Saxl’s conjecture states that all simple modules appear in the tensor square of the simple -module . In other words, we have that

with for all partitions of . This conjecture has been studied by algebraists, probabilists, and complexity theorists Reference Bes18Reference Ike15Reference LS17Reference PPV16 yet remains to be proved in general. Positivity of the Kronecker coefficient has been verified for hooks and two-row partitions when is sufficiently large in Reference PPV16, and then for arbitrary and a hook in Reference Ike15Reference Bes18 or a double-hook partition (i.e., when the Durfee size is 2) in Reference Bes18, and for any comparable to in dominance order in Reference Ike15.

This paper begins with the observation that the -module is projective over a field of characteristic , or equivalently, that the character to the Specht module is the character associated to a projective indecomposable -module (via its integral lift to characteristic 0). Therefore, the tensor square of is again a projective module, and the square of is the character to a projective module. This allows us to bring to bear the tools of modular and graded representation theory on the study of the Kronecker coefficients. In particular, we deduce that if is a simple Specht module, then all constituents of the projective cover of must also appear in Saxl’s tensor-square. For example, using this property for the trivial simple module of at characteristic 2 gives all characters of odd degree as constituents in the Saxl square; more generally, we will detect all irreducible characters of 2-height 0 as constituents. Our aim is to understand the columns of the 2-modular graded decomposition matrix which are labelled by simple Specht modules and to utilise these results towards Saxl’s conjecture.

Modular representation theory

The classification of simple Specht modules for symmetric groups and their Hecke algebras has been a massive undertaking involving over 30 years of work Reference Jam78Reference JM96Reference JM97Reference JM99Reference Fay04Reference Fay05Reference JLM06Reference Lyl07Reference FL09Reference Fay10Reference FL13, with some conjectural cases for and still to be verified. The pursuit of a description of semisimple and decomposable Specht modules is similarly old Reference Jam78 and yet has proven a much more difficult nut to crack. The decomposable Specht modules labelled by hook partitions were characterised by Murphy and Speyer Reference Mur80Reference Spe14; the graded decomposition numbers of these Specht modules were calculated by Chuang, Miyachi, and Tan Reference CMT04; the first examples of decomposable Specht modules labelled by non-hook partitions were given by Dodge and Fayers Reference DF12; Donkin and Geranios very recently unified and extended these results to certain “framed staircase” partitions Reference DG18 which we will discuss (within the wider context of “2-separated” partitions) below. It is worth emphasising that for , all Specht modules are indecomposable and therefore questions of decomposability and (non-simple) semisimplicity are inherently 2-modular problems.

For , we show that any Specht module labelled by a 2-separated partition is semisimple and we completely determine its decomposition as a sum of graded simple modules. Our proof makes heavy use of recent results in the graded representation theory of Hecke and rational Cherednik algebras. We shall denote the quantisations of the Specht and simple modules by and respectively over . We completely determine the rows of the graded decomposition matrix of labelled by 2-separated partitions; this serves as a first approximation to our goal and subsumes and generalises the results on decomposability and decomposition numbers of Specht modules for hook partitions (belonging to blocks of small -core) Reference Spe14Reference CMT04, and results on decomposition numbers of Specht modules in blocks of enormous 2-cores Reference JM96.

Graded decomposition numbers of semisimple Specht modules

The partitions of interest to us (for both Saxl’s conjecture and our decomposability classification) are the 2-separated partitions. Such partitions are obtained by taking a staircase partition, , and adding 2 copies of a partition to the right of and 2 copies of a partition to the bottom of in such a way that and do not touch (except perhaps diagonally). Such partitions, denoted , can be pictured as in Figure 1.

Notice that if the weight of a block is small compared to the size of the core, then all partitions in that block are 2-separated. We emphasise that the size of the staircase in the following statement is immaterial (provided that , where denotes the length of the partition ), and so we simply write . For those interested in the extra graded structure, we refer the reader to the full statement in Corollary 4.2.

Theorem A.

Let denote a 2-separated partition of .

The -module is semisimple and decomposes as a direct sum of simples as follows

where is the Littlewood–Richardson coefficient labelled by this triple of partitions.

In particular, there exist many blocks of (those with large cores) for which all Specht modules in the block are semisimple. In Reference DF12 Dodge and Fayers remark that “every known example of a decomposable Specht module is labelled by a 2-separated partition” and “it is interesting to speculate whether the 2-separated condition is necessary for a Specht module to be decomposable”. In fact in Section 6 we show that their speculation is not true by exhibiting two infinite families of decomposable Specht modules obtained by “inflating” the smallest decomposable Specht module (indexed by ).

Theorem A implies that all known examples of decomposable Specht modules for are obtained by reduction modulo from decomposable semisimple Specht modules for .

Applications to Kronecker coefficients

We now discuss the results and insights which 2-modular representation theory affords us in the study of Kronecker coefficients. We verify the positivity of the Kronecker coefficients in Saxl’s conjecture for a large new class of partitions, and propose conjectural strengthened and generalised versions of Saxl’s original conjecture. Our first main theorem on Kronecker coefficients is as follows:

Theorem B.

Let such that is of 2-height 0. Then . In particular, all of odd degree are constituents of the Saxl square.

We now shift focus to the Kronecker coefficients labelled by 2-separated partitions. In what follows, we shall write for the Kronecker coefficient labelled by a staircase of size for some and some 2-separated partition of ; in other words, we do not encumber the notation by explicitly recording the size of the staircases involved.

Theorem C.

For a -Carter–Saxl pair (as in Theorem 5.10) we have that . In particular, all framed staircase partitions appear in the Saxl square.

We do not recall the definition of a -Carter–Saxl pair here, but rather discuss some examples and consequences of Theorem B. In particular, Theorem B implies that every 2-block contains a wealth of constituents of the Saxl square which can be deduced using our techniques. Carter–Saxl pairs cut across hook partitions, partitions of arbitrarily large Durfee size, symmetric and non-symmetric partitions, partitions from arbitrary blocks, and across the full range of the dominance order. (In fact, the only common trait of these partitions is that they label semisimple Specht modules for .) We shall illustrate below that the property of being a Carter–Saxl pair is actually very easy to work with diagrammatically. For example, the above theorem includes the infinite family of “framed staircases” as some of the simplest examples: these are partitions which interpolate hooks and staircases. More explicitly, these are the partitions of the form . These can be pictured as in Figure 3 below.

We wish to provide bounds on the Kronecker coefficients: the maximal possible values obtained by Kronecker products are studied in Reference PPV16, and the Kronecker products whose coefficients are all as small as possible (namely all 0 or 1) are classified in Reference BB17. For constituents to partitions of depth at most 4, explicit formulae for their multiplicity in squares were provided by Saxl in 1987, and later work by Zisser and Vallejo, respectively. For the Kronecker coefficients studied here, the easiest (and well known) non-trivial case is , so the Kronecker coefficients are even unbounded; this also holds for the other families corresponding to partitions of small depth. Lower bounds coming from character values on a specific class were obtained by Pak and Panova in Reference PP17, where also the asymptotic behaviour of the multiplicity of special constituents is studied. Theorem B allows us to provide explicit lower bounds on the Kronecker coefficients for new infinite families of Saxl constituents, where again the multiplicities are unbounded.

We now provide some examples of more complicated Carter–Saxl pairs. For , if we first focus on the (unique) block of weight we find 7 constituents in this block labelled by framed staircases as well as the Carter–Saxl pairs given (up to conjugation) in Figure 4 below.

Finally, we propose two extensions of Saxl’s conjecture based on its modular representation theoretic interpretation. The first conjecture reduces the problem to the case of -regular partitions, but at the expense of working in the more difficult modular setting. We remark that towards Saxl’s conjecture over , it has already been verified that for any 2-regular partition of the Kronecker coefficient is positive Reference Ike15, and so it is natural to hope that this can be extended to positive characteristic.

Strengthened Saxl Conjecture.

Let be a field of characteristic 2. We have that

for any -regular partition of . Equivalently: Saxl’s 2-modular tensor square contains all indecomposable projective modules as direct summands with positive multiplicity.

What could be a suitable candidate for arbitrary , not just triangular numbers?

Generalised Saxl Conjecture.

For there exists a symmetric -core for some such that contains all simple -modules with positive multiplicity.

While this sounds reasonable, in fact, for larger it hardly restricts the search for a good candidate as almost any partition of is then a -core for some . So as a guide towards finding a simple module whose tensor square contains all simples, one would try to find a suitable symmetric -core for a small prime .

1. The Hecke algebra

Let be a commutative integral domain. We let denote the symmetric group on letters, with presentation

We are interested in the representation theory (over ) of symmetric groups and their deformations. Given , we define the Hecke algebra to be the unital associative -algebra with generators , , …, and relations

for . We let be the smallest integer such that or set if no such integer exists. If is a field of characteristic and , then is isomorphic to .

We define a composition, , of to be a finite sequence of non-negative integers whose sum, , equals . If the sequence is weakly decreasing, we say that is a partition; we denote the set of all partitions of by . The number of non-zero parts of a partition, , is called its length, ; the size of the largest part is called the width, . Given , its Young diagram is defined to be the configuration of nodes,

The conjugate partition, , is the partition obtained by interchanging the rows and columns of ; when , the partition is said to be symmetric. Given a node we define the content to be and the (-)residue to be the value of modulo . We now recall the dominance ordering on partitions. Let be partitions. We write if

If and we write . For partitions such that , we define the skew diagram, denoted , to be the set difference between the Young diagrams of and .

Given , we define a tableau of shape to be a filling of the nodes of the Young diagram of with the numbers . We define a standard tableau to be a tableau in which the entries increase along both the rows and columns of each component. We let denote the set of all standard tableaux of shape . We extend this to (standard) skew tableaux of shape in the obvious fashion. Given , we set . Given , we let be the subtableau of whose entries belong to the set . We write if for all and refer to this as the dominance order on .

We let and denote the most and least dominant tableaux respectively. We let be the permutation such that . For example, and

Definition 1.1.

Given a partition of , we set and we set

and we define the Specht module, , to be the left -module

Remark 1.2.

Letting and specialising we have that is isomorphic to . In this case, we drop the subscript on the Specht modules and we have that

provide a complete set of non-isomorphic simple -modules. We let denote the character of the complex irreducible module .

1.1. Modular representation theory

Let be a field and . The group algebra of the symmetric group is a semisimple algebra if and only if is a field of characteristic . By a result of Dipper and James, the Hecke algebra of the symmetric group is a non-semisimple algebra if is a primitive th root of unity for some or and is a field of characteristic . We shall now recall the basics of the non-semisimple representation theory of these algebras.

Modular representation theory seeks to deconstruct the non-semisimple representations of an algebra in terms of their simple constituents. To this end, we define the radical of a finite-dimensional -module , denoted , to be the smallest submodule of such that the corresponding quotient is semisimple. We then let and inductively define the radical series, , of by . We have a finite chain

In the non-semisimple case, the Specht modules are no longer simple but they continue to play an important role in the representation theory of as we shall now see. We say that a partition is -regular if there is no such that . We let denote the set of all -regular partitions of . Occasionally, we will also use the notation in place of . For an arbitrary field, we have that

provides a full set of non-isomorphic simple -modules. Of course, the radical of a Specht module is not easy to compute! The passage between the Specht and simple modules is recorded in the decomposition matrix,

where denotes the multiplicity of as a composition factor of . This matrix is uni-triangular with respect to the dominance ordering on . We have already seen in equation Equation 1.1 that every column of the decomposition matrix contains an entry equal to 1; namely if then . We now recall James’ regularisation theorem, which states that every row of the decomposition matrix contains an entry equal to 1 (and identifies this entry).

Example 1.3.

We picture a partition and its 2-regularisation in Figure 5. We have highlighted which nodes are moved and to where they have been moved.

We define the (-)ladder number of a node to be . The th ladder of is defined to be the set

The -regularisation of is the partition obtained by moving all of the nodes of as high along their ladders as possible. When , each ladder of is a complete north-east to south-westerly diagonal in . In particular, when the partition is obtained from by sliding nodes as high along their south-west to north-easterly diagonals as possible.

Theorem 1.4 (James’ regularisation theorem).

Let be a partition of and be an arbitrary field. We have that is equal to 1 if and is zero unless .

1.2. Brauer–Humphrey’s reciprocity

Given an -regular partition and the corresponding simple -module, we let denote its projective cover. Brauer–Humphrey’s reciprocity states that has a Specht filtration,

such that for each , we have for some dependent on and such that the multiplicity, , in this filtration is given by

In other words, the th column of the decomposition matrix determines the multiplicities in a Specht filtration of . This will be a key observation for our applications to Kronecker coefficients in Section 5.

1.3. 2-blocks

We first recall the block-structure of Hecke algebras in (quantum) characteristic (which will be the main case of interest in this paper). Throughout this section and can be taken to be an arbitrary field (although we are mainly interested in the cases when or is of characteristic ). The algebra decomposes as a direct sum of primitive 2-sided ideals, called blocks. All questions concerning modular representation theory break-down block-by-block according to this decomposition: in particular each simple/Specht module belongs to a unique block.

The rim of the Young diagram of is the collection of nodes . Given , we define the associated rim-hook to be the set of nodes . If , then we refer to as a removable -hook; if we refer to as a removable domino. Removing from gives the Young diagram of a partition of . It is easy to see that a partition has no removable dominoes if and only if it is of the form for some , in which case we say that it is a 2-core. We let denote the 2-core partition obtained by successively removing all removable dominoes from (this defines a unique partition). The number of dominoes removed from is referred to as the weight of the partition and is denoted . Given , we define to be the corresponding combinatorial 2-block. The set decomposes as the disjoint union of the non-empty . We note that it makes sense to speak of the weight of a 2-block since any two partitions in the same 2-block necessarily have the same weight. Two simple -modules (or irreducible characters of ) belong to the same 2-block if and only if their labelling partitions belong to the same combinatorial 2-block (the same is true of Specht modules).

Example 1.5.

The partition has 4 removable dominoes: two -dominoes and and two -dominoes and . One can continue to successively remove such dominoes until one is left with the -core as depicted on the left-hand-side of Figure 6.

Definition 1.6.

Let be arbitrary and and such that . We let denote the partition

We say that any partition, , of this form is 2-separated.

Remark 1.7.

We note that 2-separated partitions appear across all 2-blocks of the Hecke algebra. If the weight of a block is small compared to the size of the core, then all partitions in that block are 2-separated.

Remark 1.8.

While the name “2-separated” may seem odd to some readers, it is motivated by the form this partition takes on a 2-abacus. In Reference JM96 these partitions are referred to as “2-quotient separated”.

1.4. Characters of 2-height zero

We now wish to discuss the defect groups of 2-blocks of symmetric groups and their characters of 2-height zero (see Reference JK for background and more details). Write where ; we set . For , let be the largest 2-power dividing . Then is the size of a Sylow 2-subgroup of . Let be a -block of of weight ; then a defect group of is isomorphic to a Sylow 2-subgroup of , and thus is of cardinality ; the number is called the defect of .

Example 1.9.

The five 2-blocks of are indexed by the 2-cores and . These blocks are of weight 18, 15, 13, 4, and 0 respectively. Since , the 2-block of weight 18 has defect .

We now recall the important notion of 2-height 0 characters and simple modules. First we recall the fact that the dimension of any simple module or belonging to a -block of the symmetric group is divisible by . Such a module is said to be of 2-height 0 (or just height 0 if the prime is fixed in the context) if this 2-power is the largest 2-power dividing its dimension. We also say that the character associated to the Specht module is a 2-height 0 character (or just height 0 character if the prime is fixed).

For the 2-block , we set

Generalising an earlier result of Macdonald on characters of odd degree, a combinatorial description for the partition labels of height 0 characters was given in terms of the so-called 2-core tower by Olsson (see Reference O76Reference O93). A new characterisation was recently given in Reference GMT18, Section 3.2, again generalising an earlier version for the principal 2-block. This says that a partition in a 2-block of weight , where , labels a height 0 character if and only if there is a sequence

of partitions such that is a rim-hook for , …, .

A formula for the number of height 0 characters in a 2-block of weight was already given by Olsson Reference O76, and was also deduced from the description above in Reference GMT18. With and given as above, we have

Since we get this number for each 2-block, the set of height 0 characters for constitutes quite a large class of irreducible characters.

Example 1.10.

The irreducible characters of height 0 belonging to the principal 2-block of are precisely the characters of odd degree.

Example 1.11.

The partitions and the partitions all label 2-height zero characters. These partitions are depicted in Figure 2 in such a manner as to illustrate their combinatorial construction via adding rim hooks (detailed above).

Example 1.12.

The five 2-blocks of , their weights , the 2-adic expansions of , and the number of height 0 characters in the 2-block are recorded in the table below.

In particular, there are in total 1417 height 0 characters in 2-blocks of , amongst which there are 128 of odd degree.

The following theorem will be one of the key results we use later on. It says that while there are many complex characters of height 0, there is only one simple module of height 0 in each 2-block.

Theorem 1.13 (Reference KOW12, Theorem 1.4).

Let be a field of characteristic 2. For any 2-block of weight of the symmetric group , the module is the unique simple -module in of height 0.

2. KLR algebras and coloured tableaux

In this section, we assume is a primitive th root of unity. Given and an indeterminate we define the quantum integers and quantum factorials

and given a partition of length , we set

We now define the quantum binomial coefficients to be

for all . The motivating observation for studying Hecke algebras is the following. Let be a field of characteristic and let be an element of order : then is isomorphic to . This gives us a way of factorising representation theoretic questions into two steps: firstly specialise the quantum parameter to be a th root of unity (in and compatibly in ) and study the non-semisimple algebra ; then reduce modulo by studying . This allows us to factorise the problem of understanding decomposition matrices as follows,

On the right-hand side of the equality we have two matrices: the first is the decomposition matrix for and the second is known as “James’ adjustment matrix”. Therefore understanding the decomposition matrix of serves as a first step toward understanding the decomposition matrix of .

We now recall the manner in which the grading can be incorporated into the picture and its immense power in understanding the decomposition matrix for (and hence, by equation Equation 2.1 gives us a method for attacking the problem of calculating decomposition numbers for symmetric groups). Let be an indeterminate over . The following theorem provides us with a -graded presentation (which we record with respect to the indeterminate ) of the Hecke algebra.

Theorem 2.1 (Reference BK09aReference KL09Reference Rou08a).

The Hecke algebra admits a graded presentation with generators

subject to a list of relations given in Reference BK09a, Main Theorem. The -grading on is given by

The importance of Theorem 2.1 is that it allows to consider an extra, richer graded structure on the Specht modules. We now recall the definition of this grading on the tableau basis of the Specht module. Let and . We let denote the node in containing the integer . Given , we let , (respectively ) denote the set of all addable -nodes (respectively all removable -nodes) of the partition which are above , i.e. those in an earlier row. Let and . We define the degree of as follows:

Given we define the residue sequence of as follows:

and we write . Let be an indeterminate over . If is a free graded -module, then its graded dimension is the Laurent polynomial

If is a graded -module and , define to be the same module with for all . We call this a degree shift by . The graded dimensions of Specht modules admit a combinatorial description as follows:

Theorem 2.2 (Reference BKW11).

The Specht module is a free -graded -module with basis and where .

Of course, this theorem gives us an added level of graded structure to consider: the graded decomposition numbers of symmetric groups and their Hecke algebras. By Theorem 2.2, we obtain a grading on the module . We define the graded decomposition number to be the polynomial

which records the composition multiplicity of each simple module and its relevant degree shift. In particular upon specialisation the polynomials of equation Equation 2.3 specialise to be the usual decomposition numbers. While one might expect this grading to increase the level of difficulty of our question, we find that by keeping track of this extra grading information we are rewarded with an incredibly powerful algorithm for understanding the decomposition numbers of .

Equation Equation 2.1 hints that we could first study the decomposition numbers of as an intermediary first step toward understanding the decomposition numbers of symmetric groups in positive characteristic. In fact, this approach has been incredibly successful: Lascoux, Leclerc and Thibon provided an iterative algorithm for understanding the graded decomposition numbers of in Reference LLT96. We now provide an elementary tableau-theoretic re-interpretation of this algorithm (using the work of Kleshchev and Nash Reference KN10).

2.1. Coloured tableaux

We now recast ideas from Reference KN10 in terms of orbits of tableaux which we encode as “coloured tableaux”. This “colouring” comes from the observation that each idempotent truncation of a Specht module, , has a homogeneous basis indexed by the such that , by Theorem 2.2. We now develop these ideas further.

Let a partition of and a composition of . We define a Young tableau of shape and weight to be a filling of the nodes of with the entries

We say that a tableau is row standard if the entries are weakly increasing along the rows of ; we denote the set of such tableaux by . We say that the Young tableau is semistandard if the entries are weakly increasing along the rows and are strictly increasing along the columns of ; we denote the set of such tableaux by .

Definition 2.3.

Given , we let denote the composition such that

where we have that by definition. We define a semistandard coloured tableaux, , to be a semistandard tableau of weight such that the entry of any node is congruent to its residue. We denote the set of all such tableaux of shape by . We let denote the unique element of . We set .

Example 2.4.

For , and , we have that is equal to , and respectively. All semistandard coloured tableaux (up to conjugation) for the principal 2-block of are listed in the table below.

The importance of coloured semistandard tableaux of weight is that they encode an -orbit of standard Young tableaux; we shall now make this idea more precise. Given a composition and , we set . Let be an -regular partition and let be a standard Young tableau of shape such that the residue sequence of is given by

for ; we refer to such an as a ladder tableau of ladder weight . Then define to be the coloured tableau obtained from by replacing each entry for in by the entry for .

We identify a coloured semistandard Young tableau, , of weight with the set of standard Young tableaux, . Given we let denote the unique most dominant tableau in .

Example 2.5.

Continuing with Example 2.4, we let depicted above. We have that

as an orbit of standard tableaux (which we have coloured in order to facilitate comparison).

For , we let be the unique element of .

Proposition 2.6.

We have that

Proof.

We have that for any (by definition of the ladder tableau) and so the first and second equalities hold by Theorem 1.4 and the definition of the ladder tableau as an orbit. The final equality is not difficult, but is explicitly proven in Reference KN10, Lemma 3.4.

By the above, the orbit sum is invariant under the bar map interchanging and so we keep track of this by formally setting . We now provide a general definition of the degree of a coloured tableau which allows us to calculate the graded characters of weight spaces of Specht modules in terms of coloured tableaux. Let be a node of residue and , . We let denote the set of all addable -nodes of the partition

which are above . We let denote the set of all removable -nodes of the partition

which are above . We then define the degree of the node to be . We define to be the sum over the degrees of all nodes .

We have seen that the tableaux of are simply the orbits of tableaux from with a given residue sequence. Therefore, by comparing the degree function for coloured tableaux with that of standard tableaux we obtain

And so coloured standard tableaux provide a combinatorial description of the ladder-weight multiplicity as defined in Reference KN10, Section 3.3.

Example 2.7.

Continuing with in Example 2.5, we have that

and therefore

Therefore . The four distinct standard tableaux are depicted in Example 2.5; these four tableaux are obtained from each other by permuting the pairs in the third ladder and the pairs in the fourth ladder. We have that

With our new tableaux theoretic combinatorics in place, we can recast the (LLT) algorithm from Reference KN10, Section 4 in this combinatorial setting.

Example 2.8.

We record the graded degrees of the coloured tableaux appearing in Example 2.4 and their conjugates (which are not pictured). Notice that conjugation does not preserve the degrees of tableaux.

We are almost ready to restate the LLT algorithm in terms of our combinatorics, we simply require two observations about the graded structure of the Hecke algebra. The first is almost trivial, but the proof of the latter depends on incredibly deep geometric or categorical insights.

Theorem 2.9 (Reference BK09b, Theorem 4.18).

For and , the polynomial is bar-invariant (i.e., fixed under interchanging and ).

Theorem 2.10 (Reference VV99).

Let . For and with , .

Rearranging Reference KN10, Theorem 3.8 in terms of our coloured tableaux, we obtain the following relationships between coloured tableaux, simple characters, and graded decomposition numbers:

Proposition 2.11.

For and we have that

Moreover, the following hold:

if , then and ;

we have ;

we have that, for ,

Now we set . The right-hand side of the equation in Proposition 2.11 is calculated by induction along the dominance ordering. Any polynomial in can be written uniquely as the sum of a bar-invariant polynomial from and a polynomial from . Putting together Theorems 2.9 and 2.10 we deduce that the left-hand-side is uniquely determined by the right-hand-side and induction on the dominance order.

Example 2.12.

We continue with Example 2.4. Using the equation in Proposition 2.11, we obtain the first 5 rows of the graded decomposition matrix of the principal block of and as follows:

Notice that if we multiply these two matrices together we obtain the matrix from Example 2.8. The remaining entries of the table can be deduced by applying the sign automorphism to the Specht modules (although this automorphism is not of degree zero and so the entries will differ by a degree shift). Comparing with the table in Example 2.8, we observe that the entry in the row labelled by and column labelled by is bar-invariant in Example 2.8 and so does not contribute to the decomposition matrix, but instead contributes a vector in the simple module . Then in the row labelled by we see another discrepancy between the two tables: this is because is a composition factor of and so it contributes to the sum in Proposition 2.11.

We are now ready to provide new upper bounds for (graded) decomposition numbers in terms of our coloured tableaux.

Theorem 2.13.

For and and an arbitrary field, we have that

for and in particular, .

Proof.

It is immediate from equation Equation 2.4 that

and indeed this is just rephrasing a classical observation due to Gordon James. The new observation is that by Proposition 2.11, we know that divides both

and the result follows by induction on the dominance ordering and the equation in Proposition 2.11. In more detail, our base case for induction is when mentioned above. Now, by Proposition 2.11 and our inductive assumption, the result holds for all such that . Putting this together with Proposition 2.11, we deduce that divides as required.

Example 2.14.

If then the graded decomposition matrix of is given by the table in Example 2.8. In other words, the bounds of Theorem 2.13 are sharp.

Remark 2.15.

The inductive approach to calculating decomposition numbers of highlighted in the equation in Proposition 2.11 above is used in the arXiv appendix to this paper to prove decomposability of an infinite family of Specht modules. In Section 4 the above algorithm will not work (as the set of 2-separated partitions is not saturated in the dominance order). However, we provide an analogous algorithm for calculating 2-separated decomposition numbers using “2-dilated” coloured tableaux.

3. The Cherednik algebra and a simple criterion for semisimplicity of a Specht module

The group acts on the algebra, , of polynomials in non-commuting variables. The rational Cherednik algebra is a quotient of the semidirect product algebra by commutation relations in the ’s and ’s that are similar to those of the Weyl algebra but involve an error term in (see Reference EG02, Section 1 for the full list of relations). In particular, these relations tell us that the ’s commute with each other and so do the ’s. The algebra has three distinguished subalgebras: , , and the group algebra . The PBW theorem Reference EG02, Theorem 1.3 asserts that multiplication gives a vector space isomorphism

called the triangular decomposition of , by analogy with the triangular decomposition of the universal enveloping algebra of a semisimple Lie algebra. We define the category to be the full subcategory consisting of all finitely generated -modules on which , …, act locally nilpotently. The category is a highest weight category with respect to the poset . The standard modules are constructed as follows. Extend the action of on to an action of by letting , …, act by . The algebra is a subalgebra of and we define the Weyl modules,

where the last equality is only as -modules and follows from the triangular decomposition. We let denote the unique irreducible quotient of . In Reference RSVV16, Theorem 7.4. (see also Reference Los16Reference Web13) it is shown that is standard Koszul. We do not recall the definition of a standard Koszul algebra here, but merely the following useful proposition. The following proposition is proven in Reference BGS96, Proposition 2.4.1 in the generality of all Koszul algebras.

Proposition 3.1.

For we have that

Proof.

By Reference BGS96, Corollary 2.3.3, any Koszul algebra is quadratic. Therefore, since is simple and concentrated in degree zero, the radical filtration of coincides with the grading filtration of by Reference BGS96, Proposition 2.4.1.

Now, there exists an exact functor (the Knizhnik–Zamolodchikov functor) relating the module categories of Cherednik and Hecke algebras,

A construction of this functor is given in Reference GGOR03, here we will only need the fact (from Reference GGOR03, Section 6) that

This allows us to prove the following criterion for decomposability of Specht modules, denoted , for the Hecke algebra .

Theorem 3.2.

Fix . Suppose that for all -regular partitions , we have that

for some fixed (independent of ) and some scalars . It follows that the Specht module is semisimple.

Proof.

Throughout the proof, we let be an arbitrary partition. For an -regular partition , we have that

by equation Equation 3.1. Putting together Proposition 3.1 and our assumption in equation Equation 3.2, we have that

for any -regular partition and any . Therefore

Therefore

and the result follows.

Remark 3.3.

We have seen the grading and radical structure of standard -modules are intimately related. It is unknown as to whether or not the Schur functor preserves this property. Thus Theorem 3.2 represents all that is currently known about the relationship between the grading and radical structure on Specht modules for .

Remark 3.4.

For the reader unfamiliar with Cherednik algebras, one can also deduce the results of this section using the language of quiver Schur algebras and appealing to Reference SW11.

4. The Hecke algebra and -separated partitions

Throughout this section, we shall consider the representation theory of the Hecke algebra as a first approximation to the -modular representation theory of symmetric groups. We focus on the Specht modules labelled by -separated partitions. We shall prove that these modules are semisimple and decompose them as a direct sum of graded simple modules.

Theorem 4.1.

We set . We have that

Before embarking on the proof, we note the following immediate corollaries.

Corollary 4.2.

We have that

Therefore, as an -module, any Specht module labelled by a 2-separated partition is semisimple and decomposes as follows

In particular, the Specht -module is simple if and only if or is equal to .

Proof.

We first note that the composition factors of are all of the form and so we need only consider 2-quotient separated partitions for the remainder of the proof. To see this, assume that and that is 2-quotient separated, but is not. The partition must have at least complete ladders (as it is not 2-separated) and so . On the other hand, is 2-quotient separated and so it has at most complete ladders, . Now, implies that . The result follows from Theorem 1.4.

As in Section 2, it suffices to restrict our attention to the dimensions of “weight spaces” given by the ladder tableaux for 2-regular partitions (in other words, we consider the dimensions of ). We further note that, if , then and therefore the character of , given in equation Equation 4.1, is bar-invariant for all and ; this implies that by Theorem 2.10. In particular, we note that

as an obvious special case of equation Equation 4.1, using Proposition 2.11(ii). For an arbitrary 2-separated simple module, this implies that

Finally, we have that

by the definition of the Littlewood–Richardson coefficients, hence equation Equation 4.2 holds. Equation Equation 4.3 follows immediately by induction on the dominance ordering (as in the LLT algorithm of Subsection 2.1).

Proof of Theorem 4.1.

We let and we assume for notational purposes that is even; the odd case is identical except that the residues 0 and 1 must be transposed. We let . Given , we have that if and only if where

and we let denote the set of all tableaux of the required residue sequence. We have that . All that remains is to show that

Given an integer we have that there exists a unique corresponding integer such that

these integers will record the weight of the semistandard tableaux in the statement of equation Equation 4.4. Namely, we record a skew-tableau by placing both the usual entry but we also add a subscript . An example is depicted on the left-hand side of Figure 7. Recall that we can think of the partition as being obtained by adding -dominoes to the right of and -dominoes to the bottom of in an intuitive fashion demonstrated in Figure 6. Take the partition and add a total of nodes of residue 0; the resulting partition has precisely addable 1-nodes , …, : namely, those which belong to the - and -dominoes containing the nodes , …, . Repeating this observation as necessary, we deduce that any two nodes in the same domino of a tableau have the same subscript. Furthermore, we note that the fact that the residue sequence is of the form

implies that no two -dominoes of the same subscript can be added in the same row and no two -dominoes of the same subscript can be added in the same column. Therefore we obtain a well-defined map

given by scaling the sizes of all the dominoes by , conjugating , and recording only the subscripts (i.e., deleting the integers ). An example is depicted in Figure 7.

All that remains to show is that

for any . The set consists of an orbit

of standard tableaux. In other words, are such if and only if they differ by permuting nodes whose subscripts and residues are both matching. Therefore, we have that

and so the ungraded version of equation Equation 4.5 follows.

It remains to consider the grading. We first cut the diagram of any 2-separated partition into four regions by drawing a vertical line immediately after the th column of and a horizontal line immediately below the th row. An example is depicted in Figure 8. We label the three of the four quarters of the diagram , , and as suggested in Figure 8. The intersection of with the region is equal to the staircase partition of width ; we set .

We shall calculate the left-hand-side of equation Equation 4.5 by peeling off a row of at a time, in a manner which we now make precise. Fix a partition and set and . We define

for . We set and . We consider the associated sequence of partitions

an example is given in Figure 7. Setting , we will show that

and hence deduce the result. Clearly we can calculate the degree contribution of the node to the tableau

by considering the addable/removable nodes above from the regions

separately. We let , and denote these respective contributions. We first consider the contribution from . By definition, the sum total

is independent of the tableau . For a staircase partition of , we have that

Therefore is equal to the number of -bricks in region . In other words .

Finally, it remains to prove that

We set . It is easy to see that

where the first/second multiplicand on the right-hand-side counts the contribution of all the residue 0-boxes/1-boxes respectively. We set for . We have that

which is equal to . Here the first equality follows by considering both the degree of the node (which is equal to ) and the resulting shift to the degrees of each of the nodes below . The second equality holds by induction. Therefore the result follows.

5. Kronecker coefficients and Saxl’s conjecture

Let be partitions of . For the remainder of the paper, we will let denote the unquantised simple -module. We define the Kronecker coefficients to be the coefficients in the expansion

We now recall Saxl’s conjecture concerning the positivity of these coefficients. We let denote the complex irreducible -character to the partition of , i.e., the character of the Specht module .

Saxl’s conjecture.

Let and . For all , the multiplicity of in the Kronecker product is strictly positive.

In Reference HSTZ13, Heide, Saxl, Tiep and Zalesski verified that for almost all finite simple groups of Lie type the square of the Steinberg character contains all irreducible characters as constituents. They also conjectured that for all alternating groups there is some irreducible character with this property. For symmetric groups to triangular numbers , Saxl then suggested the candidate as stated above. Saxl’s conjecture has been attacked by algebraists and complexity theorists using a variety of combinatorial and probabilistic methods Reference Bes18Reference Ike15Reference LS17Reference PPV16. From our perspective, a particularly useful result is the following.

Theorem 5.1 (Reference Ike15, Theorem 2.1).

Let and . If is a partition of such that or , then .

We let be a field of characteristic 2. We keep the notation and , and we note that is a simple projective -module. Therefore its tensor square is also projective and decomposes as a direct sum of indecomposable projective modules labelled by 2-regular partitions; we let denote the corresponding coefficients as follows

Equivalently, on the level of complex characters we have for the irreducible character the decomposition of its square into characters to projective indecomposable modules (i.e., to integral lifts of the projective modules at characteristic 2):

We wish to pass information back and forth between the 2-modular coefficients and the Kronecker coefficients in order to make headway on Saxl’s conjecture. The following observation is immediate

We first want to explain how to apply this to obtain positivity for new classes of Kronecker coefficients. Recall from Subsection 1.3 that the 2-blocks of are parameterized by the common 2-core of the partitions labelling the simple modules in characteristic 2 and the Specht modules in the block, together with the weight. Further recall from Subsection 1.1 that the decomposition matrix for each block is unitriangular with respect to the dominance ordering. In particular, the most dominant partition, , in a given 2-block of weight labels a Specht module with simple reduction mod 2. Ikenmeyer’s result implies that appears with positive multiplicity; therefore appears as a subquotient of some projective -module; by maximality we know that the only projective containing as a subquotient is, in fact, itself. Thus appears as a direct summand in equation Equation 5.1 and so .

Hence, any non-zero entry of the first column (labelled by ) of the 2-decomposition matrix of any 2-block corresponds to a non-zero Kronecker coefficient. This allows us to verify Kronecker positivity in Saxl’s conjecture for two new infinite families of partitions:

Theorem 5.2.

Let , and such that is of height 0. Then . In particular, all of odd degree are constituents of the Saxl square.

Proof.

Let be the 2-block of to which belongs. Because is a character of height 0, the modulo 2 reduction must have a composition factor of height zero and this composition factor must appear with odd multiplicity (simply by comparing the dimensions, see Subsection 1.4). By Theorem 1.13, the 2-block contains a unique simple module of height 0, with the most dominant partition belonging to the block. The discussion preceding the theorem now implies

Since we get the large number of height 0 irreducible characters for each 2-block , this constitutes quite a large class of constituents in the Saxl square.

Example 5.3.

We have already seen (in Example 1.12) that the block contains 1024 height 0 characters, which all appear in Saxl’s tensor square.

Example 5.4.

We consider the partition . This is the first of the four partitions pictured in Figure 2. In this case, the desired positivity of cannot be deduced using the available non-vanishing criteria in the literature Reference Bes18Reference PPV16, and is incomparable to in the dominance order, so Reference Ike15 does not apply. The character belongs to the 2-block of weight and 2-core , and it is of height 0. Instead of computing the degree explicitly, this can also be seen by applying one of the combinatorial descriptions for labels of height 0 characters, e.g., the one due to Reference GMT18 recalled in Subsection 1.4. Finally, referring forward in this paper: we remark that is not 2-separated. Thus Theorem 5.2 provides us with constituents of the Saxl square that cannot be deduced using any other results in the literature.

For the second new family, we will also need to apply our new results on Specht modules for the Hecke algebra.

Theorem 5.5.

For any framed staircase partition of , we have that .

Proof.

Let be the weight of the 2-block to which belongs. We have that is the most dominant partition in and so the corresponding Specht module is simple. Now

and so the result follows by the discussion above.

For reasons that will soon become apparent, we now recall Carter’s criterion explicitly.

Theorem 5.6 (Reference JM99).

We let be a field of characteristic 2. Let be a partition. Then the Specht module is simple if and only if one of the following conditions holds:

modulo for all ;

the transpose partition, , satisfies ;

,

where here is the least non-negative integer such that . We say that any partition as in satisfies Carter’s criterion.

Example 5.7.

The most dominant partition, , in a 2-block of weight satisfies Carter’s criterion.

Example 5.8.

In a 2-block of weight , we find the partition that satisfies Carter’s criterion.

If is a 2-regular partition, then all the rows of are of distinct length. It immediately follows that and therefore . If furthermore the partition satisfies Carter’s criterion, then by equation Equation 1.2 we have that is the unique projective module in which appears as a composition factor of a Specht filtration. Putting these two statements together (in light of equation Equation 5.2) we obtain the following.

Proposition 5.9.

Let and . If satisfies Carter’s criterion, then we have that .

The following result is immediate by equation Equation 5.2. It is the key to all of our results on Kronecker positivity (as it relates this problem to that of determining the positivity of modular decomposition numbers) and vastly generalises Theorem 5.5.

Theorem 5.10.

Let and . If there exists some satisfying Carter’s criterion such that , then . We refer to such a pair as an -Carter–Saxl pair.

We are now ready to use the results of Sections 3 and 4 toward the Kronecker problem.

Theorem 5.11.

Let , and . Then for any pair such that we have that

Proof.

Clearly, is a partition of . We restrict our attention to the block of weight in with 2-core . For , the partition belongs to this block, and it satisfies Carter’s criterion. Note that implies that , and also belongs to . Finally, we have that

and the result follows from Theorem 5.10.

We remark that none of the partitions above are covered by existing results in the literature. It is clear that they are not hooks or double-hooks and providing that neither or is the empty partition, then these partitions are not comparable with in the dominance order (as is both wider and longer that ).

An explicit new infinite family with unbounded Kronecker coefficients is given in the following corollary.

Corollary 5.12.

For , and , we have that

Proof.

We have

from which the result follows from Theorem 5.11.

Example 5.13.

In particular, for , for any pair such that . Examples of such partitions are pictured in Figure 4.

Example 5.14.

Let and consider the 2-block of weight 20 with 2-core . There exist 35 Carter–Saxl pairs belonging to pairs such that is equal to either or . There are many more Carter–Saxl pairs in this block.

Finally, we conclude this section by remarking that we have only used positivity of decomposition numbers for the Hecke algebra over . These are the easiest decomposition numbers to calculate, but only provide lower bounds for the decomposition numbers of symmetric groups.

6. More semisimple decomposable Specht modules

Semisimplicity and decomposability of Specht modules has long been a subject of major interest: the highlight being the recent progress on the classification of simple Specht modules for symmetric groups and their Hecke algebras Reference FL13Reference Jam78Reference JLM06Reference JM99Reference Fay05Reference JM96Reference Lyl07Reference JM97Reference Fay04Reference FL09Reference Fay10. Progress towards understanding the wider family of semi-simple and decomposable modules has been snail-like in comparison Reference Mur80Reference Spe14Reference CMT04Reference DF12Reference Rou08bReference FS16 and reserved solely to near-hook partitions. All examples of decomposable Specht modules discovered to date have been labelled by 2-separated partitions. Our Theorem A proves that for any 2-separated partition, the corresponding Specht module for the algebra is decomposable. It is natural to ask whether the converse is true: are all decomposable Specht modules for and more generally indexed by 2-separated partitions?. In Reference DF12, Section 8.2, Dodge and Fayers asked exactly this question for the symmetric group with . In this section, we provide counterexamples to this question for the Hecke algebra.

6.1. Two new infinite families of decomposable Specht modules

Given and , we define and , respectively, to be the partitions

For example, the Young diagrams of the partitions and , along with their residues, are drawn as follows:

In the arXiv appendix to this paper, we prove that

are decomposable for all and odd . The graded composition factors of these modules for can be computed with the Hecke package in GAP (and shown to be concentrated in one degree). Thus decomposability and semisimplicity can be deduced from Theorem 3.2. These provide the first examples of decomposable Specht modules indexed by partitions which are not 2-separated. Thus, while our Theorem A provides the largest family of decomposable (and semisimple) Specht modules discovered to date, it is worth noting that our list is not exhaustive.

In the arXiv Appendix, we show that both Specht modules have a direct summand equal to a (different) simple Specht module. Namely,

The proof of decomposability is not difficult, but it does involve twenty pages of extensive calculations. The basic idea is to (1) show that

using results on semistandard homomorphisms and prove that

by counting corresponding coloured tableaux. One hence deduces that this simple composition factor occurs exactly once as a composition factor but in both the head and the socle of , and thus is a direct summand. We refer the reader to the arXiv appendix for more details.

Conjecture 6.1.

For , we set . Then we expect that

By -induction, we conjecture the direct sum decomposition of .

Conjecture 6.2.

For , we set . Then we expect that

6.2. Other decomposable Specht modules

We are indebted to Matt Fayers for sharing the following examples (which he discovered by computer) after we posited that the two families in Subsection 6.1 might be the only counterexamples to the quantised version of his question Reference DF12, Section 8.2.

We hope that the examples in Figure 9 serve as inspiration for further work towards a classification of semisimple Specht modules. Several more examples can be obtained from those in Figure 9 by -induction for (analogously to Subsection 6.1) namely: , and . Finally we have one partition which breaks the mould: which is the only partition in this section not equal to its own conjugate.

6.3. Patterns

The examples of Subsections 6.1 and 6.2 do share several striking similarities. Firstly, all the examples in Subsections 6.1 and 6.2 have a direct summand which is isomorphic to a simple Specht module. Secondly, all those of Subsection 6.2 decompose as a direct sum of simples concentrated in one degree and so are semisimple by Theorem 3.2. We conjecture this is also true of the infinite families in Theorem 3.2. It is interesting to speculate whether the converse of Theorem 3.2 is also true: is semisimplicity of a Specht module equivalent to its composition factors being focused in one degree?

Acknowledgments

We would like to thank Matt Fayers for helpful discussions on simple Specht modules and for sharing his extensive computer calculations which were fundamental in writing this paper. We also wish to thank Liron Speyer for playing “match-maker” in this collaboration. We are also very grateful to the referees for their helpful comments and suggestions.

Table of Contents

  1. Abstract
  2. Introduction
    1. Kronecker positivity
    2. Modular representation theory
    3. Graded decomposition numbers of semisimple Specht modules
    4. Theorem A.
    5. Applications to Kronecker coefficients
    6. Theorem B.
    7. Theorem C.
    8. Strengthened Saxl Conjecture.
    9. Generalised Saxl Conjecture.
  3. 1. The Hecke algebra
    1. Definition 1.1.
    2. 1.1. Modular representation theory
    3. Example 1.3.
    4. Theorem 1.4 (James’ regularisation theorem).
    5. 1.2. Brauer–Humphrey’s reciprocity
    6. 1.3. 2-blocks
    7. Example 1.5.
    8. Definition 1.6.
    9. 1.4. Characters of 2-height zero
    10. Example 1.9.
    11. Example 1.10.
    12. Example 1.11.
    13. Example 1.12.
    14. Theorem 1.13 (KOW12, Theorem 1.4).
  4. 2. KLR algebras and coloured tableaux
    1. Theorem 2.1 (BK09aKL09Rou08a).
    2. Theorem 2.2 (BKW11).
    3. 2.1. Coloured tableaux
    4. Definition 2.3.
    5. Example 2.4.
    6. Example 2.5.
    7. Proposition 2.6.
    8. Example 2.7.
    9. Example 2.8.
    10. Theorem 2.9 (BK09b, Theorem 4.18).
    11. Theorem 2.10 (VV99).
    12. Proposition 2.11.
    13. Example 2.12.
    14. Theorem 2.13.
    15. Example 2.14.
  5. 3. The Cherednik algebra and a simple criterion for semisimplicity of a Specht module
    1. Proposition 3.1.
    2. Theorem 3.2.
  6. 4. The Hecke algebra and -separated partitions
    1. Theorem 4.1.
    2. Corollary 4.2.
  7. 5. Kronecker coefficients and Saxl’s conjecture
    1. Saxl’s conjecture.
    2. Theorem 5.1 (Ike15, Theorem 2.1).
    3. Theorem 5.2.
    4. Example 5.3.
    5. Example 5.4.
    6. Theorem 5.5.
    7. Theorem 5.6 (JM99).
    8. Example 5.7.
    9. Example 5.8.
    10. Proposition 5.9.
    11. Theorem 5.10.
    12. Theorem 5.11.
    13. Corollary 5.12.
    14. Example 5.13.
    15. Example 5.14.
  8. 6. More semisimple decomposable Specht modules
    1. 6.1. Two new infinite families of decomposable Specht modules
    2. Conjecture 6.1.
    3. Conjecture 6.2.
    4. 6.2. Other decomposable Specht modules
    5. 6.3. Patterns
  9. Acknowledgments

Figures

Figure 1.

A 2-separated partition (see Definition 1.6)

Graphic without alt text
Figure 2.

Examples of partitions which label 2-height 0 characters for and (and therefore label constituents of Saxl’s tensor square by Theorem B). There are 672 and 1417 such characters for these groups, respectively. A combinatorial construction of all such partitions (for arbitrary ) is given in Subsection 1.4.

Graphic without alt text
Figure 3.

Some examples of framed staircases: , , , and are all partitions of . Up to conjugation, there are 35 framed staircase partitions of . The classification of decomposable Specht modules labelled by framed staircases is the main result of Reference DG18. In Theorem 5.5 we prove Saxl’s conjecture for all framed staircase partitions. The key ingredient in our proof is that is a 1-Carter–Saxl pair for .

Graphic without alt text
Figure 4.

More examples of coefficients . These belong to the block of weight 6 for the symmetric group of rank . We have that each of belongs to a Carter–Saxl pair of the form .

Graphic without alt text
Figure 5.

The partition , its 2-regularisation

Figure 6.

The partitions and for

Graphic without alt text
Figure 7.

A -decorated standard tableau of shape and the corresponding element of . The associated sequence of partitions (as in equation Equation 4.6) is .

Graphic without alt text
Figure 8.

Dividing the partition into regions , and . In this case is the copy of the partition in region .

Figure 9.

More partitions labelling semisimple Specht modules. The graded composition factors of these modules can be computed with the Hecke package in GAP (and shown to be concentrated in one degree). Thus decomposability and semisimplicity are deduced from Theorem 3.2.

Mathematical Fragments

Theorem A.

Let denote a 2-separated partition of .

The -module is semisimple and decomposes as a direct sum of simples as follows

where is the Littlewood–Richardson coefficient labelled by this triple of partitions.

Theorem B.

Let such that is of 2-height 0. Then . In particular, all of odd degree are constituents of the Saxl square.

Equation (1.1)
Theorem 1.4 (James’ regularisation theorem).

Let be a partition of and be an arbitrary field. We have that is equal to 1 if and is zero unless .

Equation (1.2)
Example 1.12.

The five 2-blocks of , their weights , the 2-adic expansions of , and the number of height 0 characters in the 2-block are recorded in the table below.

In particular, there are in total 1417 height 0 characters in 2-blocks of , amongst which there are 128 of odd degree.

Theorem 1.13 (Reference KOW12, Theorem 1.4).

Let be a field of characteristic 2. For any 2-block of weight of the symmetric group , the module is the unique simple -module in of height 0.

Equation (2.1)
Theorem 2.1 (Reference BK09aReference KL09Reference Rou08a).

The Hecke algebra admits a graded presentation with generators

subject to a list of relations given in Reference BK09a, Main Theorem. The -grading on is given by

Theorem 2.2 (Reference BKW11).

The Specht module is a free -graded -module with basis and where .

Equation (2.3)
Example 2.4.

For , and , we have that is equal to , and respectively. All semistandard coloured tableaux (up to conjugation) for the principal 2-block of are listed in the table below.

Example 2.5.

Continuing with Example 2.4, we let depicted above. We have that

as an orbit of standard tableaux (which we have coloured in order to facilitate comparison).

Equation (2.4)
Example 2.8.

We record the graded degrees of the coloured tableaux appearing in Example 2.4 and their conjugates (which are not pictured). Notice that conjugation does not preserve the degrees of tableaux.

Theorem 2.9 (Reference BK09b, Theorem 4.18).

For and , the polynomial is bar-invariant (i.e., fixed under interchanging and ).

Theorem 2.10 (Reference VV99).

Let . For and with , .

Proposition 2.11.

For and we have that

Moreover, the following hold:

if , then and ;

we have ;

we have that, for ,

Theorem 2.13.

For and and an arbitrary field, we have that

for and in particular, .

Proposition 3.1.

For we have that

Equation (3.1)
Theorem 3.2.

Fix . Suppose that for all -regular partitions , we have that

for some fixed (independent of ) and some scalars . It follows that the Specht module is semisimple.

Theorem 4.1.

We set . We have that

Corollary 4.2.

We have that

Therefore, as an -module, any Specht module labelled by a 2-separated partition is semisimple and decomposes as follows

In particular, the Specht -module is simple if and only if or is equal to .

Equation (4.4)
Equation (4.5)
Equation (4.6)
Equation (5.1)
Equation (5.2)
Theorem 5.2.

Let , and such that is of height 0. Then . In particular, all of odd degree are constituents of the Saxl square.

Theorem 5.5.

For any framed staircase partition of , we have that .

Theorem 5.10.

Let and . If there exists some satisfying Carter’s criterion such that , then . We refer to such a pair as an -Carter–Saxl pair.

Theorem 5.11.

Let , and . Then for any pair such that we have that

References

Reference [BB17]
Christine Bessenrodt and Christopher Bowman, Multiplicity-free Kronecker products of characters of the symmetric groups, Adv. Math. 322 (2017), 473–529, DOI 10.1016/j.aim.2017.10.009. MR3720803,
Show rawAMSref \bib{cbcb}{article}{ label={BB17}, author={Bessenrodt, Christine}, author={Bowman, Christopher}, title={Multiplicity-free Kronecker products of characters of the symmetric groups}, journal={Adv. Math.}, volume={322}, date={2017}, pages={473--529}, issn={0001-8708}, review={\MR {3720803}}, doi={10.1016/j.aim.2017.10.009}, }
Reference [Bes18]
Christine Bessenrodt, Critical classes, Kronecker products of spin characters, and the Saxl conjecture, Algebr. Comb. 1 (2018), no. 3, 353–369, DOI 10.5802/alco. MR3856528,
Show rawAMSref \bib{MR3856528}{article}{ label={Bes18}, author={Bessenrodt, Christine}, title={Critical classes, Kronecker products of spin characters, and the Saxl conjecture}, journal={Algebr. Comb.}, volume={1}, date={2018}, number={3}, pages={353--369}, review={\MR {3856528}}, doi={10.5802/alco}, }
Reference [BGS96]
Alexander Beilinson, Victor Ginzburg, and Wolfgang Soergel, Koszul duality patterns in representation theory, J. Amer. Math. Soc. 9 (1996), no. 2, 473–527, DOI 10.1090/S0894-0347-96-00192-0. MR1322847,
Show rawAMSref \bib{MR1322847}{article}{ label={BGS96}, author={Beilinson, Alexander}, author={Ginzburg, Victor}, author={Soergel, Wolfgang}, title={Koszul duality patterns in representation theory}, journal={J. Amer. Math. Soc.}, volume={9}, date={1996}, number={2}, pages={473--527}, issn={0894-0347}, review={\MR {1322847}}, doi={10.1090/S0894-0347-96-00192-0}, }
Reference [BK09a]
Jonathan Brundan and Alexander Kleshchev, Blocks of cyclotomic Hecke algebras and Khovanov-Lauda algebras, Invent. Math. 178 (2009), no. 3, 451–484, DOI 10.1007/s00222-009-0204-8. MR2551762,
Show rawAMSref \bib{MR2551762}{article}{ label={BK09a}, author={Brundan, Jonathan}, author={Kleshchev, Alexander}, title={Blocks of cyclotomic Hecke algebras and Khovanov-Lauda algebras}, journal={Invent. Math.}, volume={178}, date={2009}, number={3}, pages={451--484}, issn={0020-9910}, review={\MR {2551762}}, doi={10.1007/s00222-009-0204-8}, }
Reference [BK09b]
Jonathan Brundan and Alexander Kleshchev, Graded decomposition numbers for cyclotomic Hecke algebras, Adv. Math. 222 (2009), no. 6, 1883–1942, DOI 10.1016/j.aim.2009.06.018. MR2562768,
Show rawAMSref \bib{bk09}{article}{ label={BK09b}, author={Brundan, Jonathan}, author={Kleshchev, Alexander}, title={Graded decomposition numbers for cyclotomic Hecke algebras}, journal={Adv. Math.}, volume={222}, date={2009}, number={6}, pages={1883--1942}, issn={0001-8708}, review={\MR {2562768}}, doi={10.1016/j.aim.2009.06.018}, }
Reference [BKW11]
Jonathan Brundan, Alexander Kleshchev, and Weiqiang Wang, Graded Specht modules, J. Reine Angew. Math. 655 (2011), 61–87, DOI 10.1515/CRELLE.2011.033. MR2806105,
Show rawAMSref \bib{bkw11}{article}{ label={BKW11}, author={Brundan, Jonathan}, author={Kleshchev, Alexander}, author={Wang, Weiqiang}, title={Graded Specht modules}, journal={J. Reine Angew. Math.}, volume={655}, date={2011}, pages={61--87}, issn={0075-4102}, review={\MR {2806105}}, doi={10.1515/CRELLE.2011.033}, }
Reference [CDW12]
Matthias Christandl, Brent Doran, and Michael Walter, Computing multiplicities of Lie group representations, 2012 IEEE 53rd Annual Symposium on Foundations of Computer Science—FOCS 2012, IEEE Computer Soc., Los Alamitos, CA, 2012, pp. 639–648. MR3186652,
Show rawAMSref \bib{MR3186652}{article}{ label={CDW12}, author={Christandl, Matthias}, author={Doran, Brent}, author={Walter, Michael}, title={Computing multiplicities of Lie group representations}, conference={ title={2012 IEEE 53rd Annual Symposium on Foundations of Computer Science---FOCS 2012}, }, book={ publisher={IEEE Computer Soc., Los Alamitos, CA}, }, date={2012}, pages={639--648}, review={\MR {3186652}}, }
Reference [CSW18]
Matthias Christandl, M. Burak Şahinoğlu, and Michael Walter, Recoupling coefficients and quantum entropies, Ann. Henri Poincaré 19 (2018), no. 2, 385–410, DOI 10.1007/s00023-017-0639-1. MR3748296,
Show rawAMSref \bib{entropy}{article}{ label={CSW18}, author={Christandl, Matthias}, author={\c {S}ahino\u {g}lu, M. Burak}, author={Walter, Michael}, title={Recoupling coefficients and quantum entropies}, journal={Ann. Henri Poincar\'{e}}, volume={19}, date={2018}, number={2}, pages={385--410}, issn={1424-0637}, review={\MR {3748296}}, doi={10.1007/s00023-017-0639-1}, }
Reference [CHM07]
Matthias Christandl, Aram W. Harrow, and Graeme Mitchison, Nonzero Kronecker coefficients and what they tell us about spectra, Comm. Math. Phys. 270 (2007), no. 3, 575–585, DOI 10.1007/s00220-006-0157-3. MR2276458,
Show rawAMSref \bib{MR2276458}{article}{ label={CHM07}, author={Christandl, Matthias}, author={Harrow, Aram W.}, author={Mitchison, Graeme}, title={Nonzero Kronecker coefficients and what they tell us about spectra}, journal={Comm. Math. Phys.}, volume={270}, date={2007}, number={3}, pages={575--585}, issn={0010-3616}, review={\MR {2276458}}, doi={10.1007/s00220-006-0157-3}, }
Reference [CM06]
Matthias Christandl and Graeme Mitchison, The spectra of quantum states and the Kronecker coefficients of the symmetric group, Comm. Math. Phys. 261 (2006), no. 3, 789–797, DOI 10.1007/s00220-005-1435-1. MR2197548,
Show rawAMSref \bib{MR2197548}{article}{ label={CM06}, author={Christandl, Matthias}, author={Mitchison, Graeme}, title={The spectra of quantum states and the Kronecker coefficients of the symmetric group}, journal={Comm. Math. Phys.}, volume={261}, date={2006}, number={3}, pages={789--797}, issn={0010-3616}, review={\MR {2197548}}, doi={10.1007/s00220-005-1435-1}, }
Reference [CMT04]
Joseph Chuang, Hyohe Miyachi, and Kai Meng Tan, A -analogue of Peel’s theorem, J. Algebra 280 (2004), no. 1, 219–231, DOI 10.1016/j.jalgebra.2004.04.017. MR2081930,
Show rawAMSref \bib{MR2081930}{article}{ label={CMT04}, author={Chuang, Joseph}, author={Miyachi, Hyohe}, author={Tan, Kai Meng}, title={A $v$-analogue of Peel's theorem}, journal={J. Algebra}, volume={280}, date={2004}, number={1}, pages={219--231}, issn={0021-8693}, review={\MR {2081930}}, doi={10.1016/j.jalgebra.2004.04.017}, }
Reference [DF12]
Craig J. Dodge and Matthew Fayers, Some new decomposable Specht modules, J. Algebra 357 (2012), 235–262, DOI 10.1016/j.jalgebra.2012.01.035. MR2905253,
Show rawAMSref \bib{MR2905253}{article}{ label={DF12}, author={Dodge, Craig J.}, author={Fayers, Matthew}, title={Some new decomposable Specht modules}, journal={J. Algebra}, volume={357}, date={2012}, pages={235--262}, issn={0021-8693}, review={\MR {2905253}}, doi={10.1016/j.jalgebra.2012.01.035}, }
Reference [DG18]
Stephen Donkin and Haralampos Geranios, Decompositions of some Specht modules I, J. Algebra 550 (2020), 1–22, DOI 10.1016/j.jalgebra.2019.12.017. MR4053571,
Show rawAMSref \bib{DG}{article}{ label={DG18}, author={Donkin, Stephen}, author={Geranios, Haralampos}, title={Decompositions of some Specht modules I}, journal={J. Algebra}, volume={550}, date={2020}, pages={1--22}, issn={0021-8693}, review={\MR {4053571}}, doi={10.1016/j.jalgebra.2019.12.017}, }
Reference [EG02]
Pavel Etingof and Victor Ginzburg, Symplectic reflection algebras, Calogero-Moser space, and deformed Harish-Chandra homomorphism, Invent. Math. 147 (2002), no. 2, 243–348, DOI 10.1007/s002220100171. MR1881922,
Show rawAMSref \bib{EtingofGinzburg}{article}{ label={EG02}, author={Etingof, Pavel}, author={Ginzburg, Victor}, title={Symplectic reflection algebras, Calogero-Moser space, and deformed Harish-Chandra homomorphism}, journal={Invent. Math.}, volume={147}, date={2002}, number={2}, pages={243--348}, issn={0020-9910}, review={\MR {1881922}}, doi={10.1007/s002220100171}, }
Reference [Fay04]
Matthew Fayers, Reducible Specht modules, J. Algebra 280 (2004), no. 2, 500–504, DOI 10.1016/j.jalgebra.2003.09.053. MR2089249,
Show rawAMSref \bib{MR2089249}{article}{ label={Fay04}, author={Fayers, Matthew}, title={Reducible Specht modules}, journal={J. Algebra}, volume={280}, date={2004}, number={2}, pages={500--504}, issn={0021-8693}, review={\MR {2089249}}, doi={10.1016/j.jalgebra.2003.09.053}, }
Reference [Fay05]
Matthew Fayers, Irreducible Specht modules for Hecke algebras of type , Adv. Math. 193 (2005), no. 2, 438–452, DOI 10.1016/j.aim.2004.06.001. MR2137291,
Show rawAMSref \bib{MR2137291}{article}{ label={Fay05}, author={Fayers, Matthew}, title={Irreducible Specht modules for Hecke algebras of type ${\mathbf {A}}$}, journal={Adv. Math.}, volume={193}, date={2005}, number={2}, pages={438--452}, issn={0001-8708}, review={\MR {2137291}}, doi={10.1016/j.aim.2004.06.001}, }
Reference [Fay10]
Matthew Fayers, On the irreducible Specht modules for Iwahori-Hecke algebras of type A with , J. Algebra 323 (2010), no. 6, 1839–1844.
Reference [FL09]
Matthew Fayers and Sinéad Lyle, Some reducible Specht modules for Iwahori-Hecke algebras of type with , J. Algebra 321 (2009), no. 3, 912–933, DOI 10.1016/j.jalgebra.2008.11.006. MR2488560,
Show rawAMSref \bib{MR2488560}{article}{ label={FL09}, author={Fayers, Matthew}, author={Lyle, Sin\'{e}ad}, title={Some reducible Specht modules for Iwahori-Hecke algebras of type $A$ with $q=-1$}, journal={J. Algebra}, volume={321}, date={2009}, number={3}, pages={912--933}, issn={0021-8693}, review={\MR {2488560}}, doi={10.1016/j.jalgebra.2008.11.006}, }
Reference [FL13]
Matthew Fayers and Sinéad Lyle, The reducible Specht modules for the Hecke algebra , J. Algebraic Combin. 37 (2013), no. 2, 201–241, DOI 10.1007/s10801-012-0360-6. MR3011340,
Show rawAMSref \bib{MR3011340}{article}{ label={FL13}, author={Fayers, Matthew}, author={Lyle, Sin\'{e}ad}, title={The reducible Specht modules for the Hecke algebra $\mathcal {H}_{\mathbb {C},-1}(\mathfrak {S}_n)$}, journal={J. Algebraic Combin.}, volume={37}, date={2013}, number={2}, pages={201--241}, issn={0925-9899}, review={\MR {3011340}}, doi={10.1007/s10801-012-0360-6}, }
Reference [FS16]
Matthew Fayers and Liron Speyer, Generalised column removal for graded homomorphisms between Specht modules, J. Algebraic Combin. 44 (2016), no. 2, 393–432, DOI 10.1007/s10801-016-0674-x. MR3533560,
Show rawAMSref \bib{MR3533560}{article}{ label={FS16}, author={Fayers, Matthew}, author={Speyer, Liron}, title={Generalised column removal for graded homomorphisms between Specht modules}, journal={J. Algebraic Combin.}, volume={44}, date={2016}, number={2}, pages={393--432}, issn={0925-9899}, review={\MR {3533560}}, doi={10.1007/s10801-016-0674-x}, }
Reference [GMT18]
Eugenio Giannelli, John Murray, and Joan Tent, Alperin-McKay natural correspondences in solvable and symmetric groups for the prime , Ann. Mat. Pura Appl. (4) 197 (2018), no. 4, 999–1016, DOI 10.1007/s10231-017-0712-x. MR3829558,
Show rawAMSref \bib{MR3829558}{article}{ label={GMT18}, author={Giannelli, Eugenio}, author={Murray, John}, author={Tent, Joan}, title={Alperin-McKay natural correspondences in solvable and symmetric groups for the prime $p=2$}, journal={Ann. Mat. Pura Appl. (4)}, volume={197}, date={2018}, number={4}, pages={999--1016}, issn={0373-3114}, review={\MR {3829558}}, doi={10.1007/s10231-017-0712-x}, }
Reference [GGOR03]
Victor Ginzburg, Nicolas Guay, Eric Opdam, and Raphaël Rouquier, On the category for rational Cherednik algebras, Invent. Math. 154 (2003), no. 3, 617–651, DOI 10.1007/s00222-003-0313-8. MR2018786,
Show rawAMSref \bib{GGOR03}{article}{ label={GGOR03}, author={Ginzburg, Victor}, author={Guay, Nicolas}, author={Opdam, Eric}, author={Rouquier, Rapha\"{e}l}, title={On the category $\mathcal {O}$ for rational Cherednik algebras}, journal={Invent. Math.}, volume={154}, date={2003}, number={3}, pages={617--651}, issn={0020-9910}, review={\MR {2018786}}, doi={10.1007/s00222-003-0313-8}, }
Reference [HSTZ13]
Gerhard Heide, Jan Saxl, Pham Huu Tiep, and Alexandre E. Zalesski, Conjugacy action, induced representations and the Steinberg square for simple groups of Lie type, Proc. Lond. Math. Soc. (3) 106 (2013), no. 4, 908–930, DOI 10.1112/plms/pds062. MR3056296,
Show rawAMSref \bib{MR3056296}{article}{ label={HSTZ13}, author={Heide, Gerhard}, author={Saxl, Jan}, author={Tiep, Pham Huu}, author={Zalesski, Alexandre E.}, title={Conjugacy action, induced representations and the Steinberg square for simple groups of Lie type}, journal={Proc. Lond. Math. Soc. (3)}, volume={106}, date={2013}, number={4}, pages={908--930}, issn={0024-6115}, review={\MR {3056296}}, doi={10.1112/plms/pds062}, }
Reference [Ike15]
Christian Ikenmeyer, The Saxl conjecture and the dominance order, Discrete Math. 338 (2015), no. 11, 1970–1975, DOI 10.1016/j.disc.2015.04.027. MR3357782,
Show rawAMSref \bib{I15}{article}{ label={Ike15}, author={Ikenmeyer, Christian}, title={The Saxl conjecture and the dominance order}, journal={Discrete Math.}, volume={338}, date={2015}, number={11}, pages={1970--1975}, issn={0012-365X}, review={\MR {3357782}}, doi={10.1016/j.disc.2015.04.027}, }
Reference [Jam78]
G. D. James, The representation theory of the symmetric groups, Lecture Notes in Mathematics, vol. 682, Springer, Berlin, 1978. MR513828,
Show rawAMSref \bib{James}{book}{ label={Jam78}, author={James, G. D.}, title={The representation theory of the symmetric groups}, series={Lecture Notes in Mathematics}, volume={682}, publisher={Springer, Berlin}, date={1978}, pages={v+156}, isbn={3-540-08948-9}, review={\MR {513828}}, }
Reference [JK]
Gordon James and Adalbert Kerber, The representation theory of the symmetric group, Encyclopedia of Mathematics and its Applications, vol. 16, Addison-Wesley Publishing Co., Reading, Mass., 1981. With a foreword by P. M. Cohn; With an introduction by Gilbert de B. Robinson. MR644144,
Show rawAMSref \bib{JK}{book}{ label={JK}, author={James, Gordon}, author={Kerber, Adalbert}, title={The representation theory of the symmetric group}, series={Encyclopedia of Mathematics and its Applications}, volume={16}, note={With a foreword by P. M. Cohn; With an introduction by Gilbert de B. Robinson}, publisher={Addison-Wesley Publishing Co., Reading, Mass.}, date={1981}, pages={xxviii+510}, isbn={0-201-13515-9}, review={\MR {644144}}, }
Reference [JLM06]
Gordon James, Sinéad Lyle, and Andrew Mathas, Rouquier blocks, Math. Z. 252 (2006), no. 3, 511–531, DOI 10.1007/s00209-005-0863-0. MR2207757,
Show rawAMSref \bib{MR2207757}{article}{ label={JLM06}, author={James, Gordon}, author={Lyle, Sin\'{e}ad}, author={Mathas, Andrew}, title={Rouquier blocks}, journal={Math. Z.}, volume={252}, date={2006}, number={3}, pages={511--531}, issn={0025-5874}, review={\MR {2207757}}, doi={10.1007/s00209-005-0863-0}, }
Reference [JM96]
Gordon James and Andrew Mathas, Hecke algebras of type with , J. Algebra 184 (1996), no. 1, 102–158, DOI 10.1006/jabr.1996.0251. MR1402572,
Show rawAMSref \bib{MR1402572}{article}{ label={JM96}, author={James, Gordon}, author={Mathas, Andrew}, title={Hecke algebras of type ${\mathbf {A}}$ with $q=-1$}, journal={J. Algebra}, volume={184}, date={1996}, number={1}, pages={102--158}, issn={0021-8693}, review={\MR {1402572}}, doi={10.1006/jabr.1996.0251}, }
Reference [JM97]
Gordon James and Andrew Mathas, A -analogue of the Jantzen-Schaper theorem, Proc. London Math. Soc. (3) 74 (1997), no. 2, 241–274, DOI 10.1112/S0024611597000099. MR1425323,
Show rawAMSref \bib{MR1425323}{article}{ label={JM97}, author={James, Gordon}, author={Mathas, Andrew}, title={A $q$-analogue of the Jantzen-Schaper theorem}, journal={Proc. London Math. Soc. (3)}, volume={74}, date={1997}, number={2}, pages={241--274}, issn={0024-6115}, review={\MR {1425323}}, doi={10.1112/S0024611597000099}, }
Reference [JM99]
Gordon James and Andrew Mathas, The irreducible Specht modules in characteristic , Bull. London Math. Soc. 31 (1999), no. 4, 457–462, DOI 10.1112/S0024609399005822. MR1687552,
Show rawAMSref \bib{MR1687552}{article}{ label={JM99}, author={James, Gordon}, author={Mathas, Andrew}, title={The irreducible Specht modules in characteristic $2$}, journal={Bull. London Math. Soc.}, volume={31}, date={1999}, number={4}, pages={457--462}, issn={0024-6093}, review={\MR {1687552}}, doi={10.1112/S0024609399005822}, }
Reference [KL09]
Mikhail Khovanov and Aaron D. Lauda, A diagrammatic approach to categorification of quantum groups. I, Represent. Theory 13 (2009), 309–347, DOI 10.1090/S1088-4165-09-00346-X. MR2525917,
Show rawAMSref \bib{MR2525917}{article}{ label={KL09}, author={Khovanov, Mikhail}, author={Lauda, Aaron D.}, title={A diagrammatic approach to categorification of quantum groups. I}, journal={Represent. Theory}, volume={13}, date={2009}, pages={309--347}, review={\MR {2525917}}, doi={10.1090/S1088-4165-09-00346-X}, }
Reference [Kly04]
A. Klyachko, Quantum marginal problem and representations of the symmetric group, arXiv:quant-ph/0409113, (2004).
Reference [KN10]
Alexander Kleshchev and David Nash, An interpretation of the Lascoux-Leclerc-Thibon algorithm and graded representation theory, Comm. Algebra 38 (2010), no. 12, 4489–4500, DOI 10.1080/00927870903386536. MR2764833,
Show rawAMSref \bib{KN10}{article}{ label={KN10}, author={Kleshchev, Alexander}, author={Nash, David}, title={An interpretation of the Lascoux-Leclerc-Thibon algorithm and graded representation theory}, journal={Comm. Algebra}, volume={38}, date={2010}, number={12}, pages={4489--4500}, issn={0092-7872}, review={\MR {2764833}}, doi={10.1080/00927870903386536}, }
Reference [KOW12]
Masao Kiyota, Tetsuro Okuyama, and Tomoyuki Wada, The heights of irreducible Brauer characters in 2-blocks of the symmetric groups, J. Algebra 368 (2012), 329–344, DOI 10.1016/j.jalgebra.2012.07.001. MR2955236,
Show rawAMSref \bib{KOW}{article}{ label={KOW12}, author={Kiyota, Masao}, author={Okuyama, Tetsuro}, author={Wada, Tomoyuki}, title={The heights of irreducible Brauer characters in 2-blocks of the symmetric groups}, journal={J. Algebra}, volume={368}, date={2012}, pages={329--344}, issn={0021-8693}, review={\MR {2955236}}, doi={10.1016/j.jalgebra.2012.07.001}, }
Reference [LLT96]
Alain Lascoux, Bernard Leclerc, and Jean-Yves Thibon, Hecke algebras at roots of unity and crystal bases of quantum affine algebras, Comm. Math. Phys. 181 (1996), no. 1, 205–263. MR1410572,
Show rawAMSref \bib{LLT}{article}{ label={LLT96}, author={Lascoux, Alain}, author={Leclerc, Bernard}, author={Thibon, Jean-Yves}, title={Hecke algebras at roots of unity and crystal bases of quantum affine algebras}, journal={Comm. Math. Phys.}, volume={181}, date={1996}, number={1}, pages={205--263}, issn={0010-3616}, review={\MR {1410572}}, }
Reference [Los16]
Ivan Losev, Proof of Varagnolo-Vasserot conjecture on cyclotomic categories , Selecta Math. (N.S.) 22 (2016), no. 2, 631–668, DOI 10.1007/s00029-015-0209-7. MR3477332,
Show rawAMSref \bib{Losev}{article}{ label={Los16}, author={Losev, Ivan}, title={Proof of Varagnolo-Vasserot conjecture on cyclotomic categories $\mathcal {O}$}, journal={Selecta Math. (N.S.)}, volume={22}, date={2016}, number={2}, pages={631--668}, issn={1022-1824}, review={\MR {3477332}}, doi={10.1007/s00029-015-0209-7}, }
Reference [LS17]
Sammy Luo and Mark Sellke, The Saxl conjecture for fourth powers via the semigroup property, J. Algebraic Combin. 45 (2017), no. 1, 33–80, DOI 10.1007/s10801-016-0700-z. MR3591371,
Show rawAMSref \bib{LS}{article}{ label={LS17}, author={Luo, Sammy}, author={Sellke, Mark}, title={The Saxl conjecture for fourth powers via the semigroup property}, journal={J. Algebraic Combin.}, volume={45}, date={2017}, number={1}, pages={33--80}, issn={0925-9899}, review={\MR {3591371}}, doi={10.1007/s10801-016-0700-z}, }
Reference [Lyl07]
Sinéad Lyle, Some -analogues of the Carter-Payne theorem, J. Reine Angew. Math. 608 (2007), 93–121, DOI 10.1515/CRELLE.2007.054. MR2339470,
Show rawAMSref \bib{MR2339470}{article}{ label={Lyl07}, author={Lyle, Sin\'{e}ad}, title={Some $q$-analogues of the Carter-Payne theorem}, journal={J. Reine Angew. Math.}, volume={608}, date={2007}, pages={93--121}, issn={0075-4102}, review={\MR {2339470}}, doi={10.1515/CRELLE.2007.054}, }
Reference [Mur80]
Gwendolen Murphy, On decomposability of some Specht modules for symmetric groups, J. Algebra 66 (1980), no. 1, 156–168, DOI 10.1016/0021-8693(80)90117-9. MR591250,
Show rawAMSref \bib{MR591250}{article}{ label={Mur80}, author={Murphy, Gwendolen}, title={On decomposability of some Specht modules for symmetric groups}, journal={J. Algebra}, volume={66}, date={1980}, number={1}, pages={156--168}, issn={0021-8693}, review={\MR {591250}}, doi={10.1016/0021-8693(80)90117-9}, }
Reference [O76]
Jørn B. Olsson, McKay numbers and heights of characters, Math. Scand. 38 (1976), no. 1, 25–42, DOI 10.7146/math.scand.a-11614. MR409625,
Show rawAMSref \bib{O76}{article}{ label={O76}, author={Olsson, J\o rn B.}, title={McKay numbers and heights of characters}, journal={Math. Scand.}, volume={38}, date={1976}, number={1}, pages={25--42}, issn={0025-5521}, review={\MR {409625}}, doi={10.7146/math.scand.a-11614}, }
Reference [O93]
Jørn B. Olsson, Combinatorics and representations of finite groups, Vorlesungen aus dem Fachbereich Mathematik der Universität GH Essen [Lecture Notes in Mathematics at the University of Essen], vol. 20, Universität Essen, Fachbereich Mathematik, Essen, 1993. MR1264418,
Show rawAMSref \bib{O-LN}{book}{ label={O93}, author={Olsson, J\o rn B.}, title={Combinatorics and representations of finite groups}, series={Vorlesungen aus dem Fachbereich Mathematik der Universit\"{a}t GH Essen [Lecture Notes in Mathematics at the University of Essen]}, volume={20}, publisher={Universit\"{a}t Essen, Fachbereich Mathematik, Essen}, date={1993}, pages={ii+94}, review={\MR {1264418}}, }
Reference [PPV16]
Igor Pak, Greta Panova, and Ernesto Vallejo, Kronecker products, characters, partitions, and the tensor square conjectures, Adv. Math. 288 (2016), 702–731, DOI 10.1016/j.aim.2015.11.002. MR3436397,
Show rawAMSref \bib{PP16}{article}{ label={PPV16}, author={Pak, Igor}, author={Panova, Greta}, author={Vallejo, Ernesto}, title={Kronecker products, characters, partitions, and the tensor square conjectures}, journal={Adv. Math.}, volume={288}, date={2016}, pages={702--731}, issn={0001-8708}, review={\MR {3436397}}, doi={10.1016/j.aim.2015.11.002}, }
Reference [PP17]
Igor Pak and Greta Panova, Bounds on certain classes of Kronecker and -binomial coefficients, J. Combin. Theory Ser. A 147 (2017), 1–17, DOI 10.1016/j.jcta.2016.10.004. MR3589885,
Show rawAMSref \bib{PP17}{article}{ label={PP17}, author={Pak, Igor}, author={Panova, Greta}, title={Bounds on certain classes of Kronecker and $q$-binomial coefficients}, journal={J. Combin. Theory Ser. A}, volume={147}, date={2017}, pages={1--17}, issn={0097-3165}, review={\MR {3589885}}, doi={10.1016/j.jcta.2016.10.004}, }
Reference [Rou08a]
R. Rouquier, -Kac–Moody algebras, arXiv:0812.5023, 2008.
Reference [Rou08b]
Raphaël Rouquier, -Schur algebras and complex reflection groups (English, with English and Russian summaries), Mosc. Math. J. 8 (2008), no. 1, 119–158, 184, DOI 10.17323/1609-4514-2008-8-1-119-158. MR2422270,
Show rawAMSref \bib{MR2422270}{article}{ label={Rou08b}, author={Rouquier, Rapha\"{e}l}, title={$q$-Schur algebras and complex reflection groups}, language={English, with English and Russian summaries}, journal={Mosc. Math. J.}, volume={8}, date={2008}, number={1}, pages={119--158, 184}, issn={1609-3321}, review={\MR {2422270}}, doi={10.17323/1609-4514-2008-8-1-119-158}, }
Reference [RSVV16]
Raphaël Rouquier, Peng Shan, Michela Varagnolo, and Eric Vasserot, Categorifications and cyclotomic rational double affine Hecke algebras, Invent. Math. 204 (2016), no. 3, 671–786, DOI 10.1007/s00222-015-0623-7. MR3502064,
Show rawAMSref \bib{RSVV}{article}{ label={RSVV16}, author={Rouquier, Rapha\"{e}l}, author={Shan, Peng}, author={Varagnolo, Michela}, author={Vasserot, Eric}, title={Categorifications and cyclotomic rational double affine Hecke algebras}, journal={Invent. Math.}, volume={204}, date={2016}, number={3}, pages={671--786}, issn={0020-9910}, review={\MR {3502064}}, doi={10.1007/s00222-015-0623-7}, }
Reference [Spe14]
Liron Speyer, Decomposable Specht modules for the Iwahori-Hecke algebra , J. Algebra 418 (2014), 227–264, DOI 10.1016/j.jalgebra.2014.07.011. MR3250450,
Show rawAMSref \bib{MR3250450}{article}{ label={Spe14}, author={Speyer, Liron}, title={Decomposable Specht modules for the Iwahori-Hecke algebra $\mathcal {H}_{\mathbb {F},-_1}(\mathfrak {S}_n)$}, journal={J. Algebra}, volume={418}, date={2014}, pages={227--264}, issn={0021-8693}, review={\MR {3250450}}, doi={10.1016/j.jalgebra.2014.07.011}, }
Reference [Sta00]
Richard P. Stanley, Positivity problems and conjectures in algebraic combinatorics, Mathematics: frontiers and perspectives, Amer. Math. Soc., Providence, RI, 2000, pp. 295–319. MR1754784,
Show rawAMSref \bib{Sta00}{article}{ label={Sta00}, author={Stanley, Richard P.}, title={Positivity problems and conjectures in algebraic combinatorics}, conference={ title={Mathematics: frontiers and perspectives}, }, book={ publisher={Amer. Math. Soc., Providence, RI}, }, date={2000}, pages={295--319}, review={\MR {1754784}}, }
Reference [SW11]
C. Stroppel and B. Webster, Quiver Schur algebras and -Fock space, arXiv:1110.1115, 2011.
Reference [VV99]
Michela Varagnolo and Eric Vasserot, On the decomposition matrices of the quantized Schur algebra, Duke Math. J. 100 (1999), no. 2, 267–297, DOI 10.1215/S0012-7094-99-10010-X. MR1722955,
Show rawAMSref \bib{MR1722955}{article}{ label={VV99}, author={Varagnolo, Michela}, author={Vasserot, Eric}, title={On the decomposition matrices of the quantized Schur algebra}, journal={Duke Math. J.}, volume={100}, date={1999}, number={2}, pages={267--297}, issn={0012-7094}, review={\MR {1722955}}, doi={10.1215/S0012-7094-99-10010-X}, }
Reference [Web13]
Ben Webster, Rouquier’s conjecture and diagrammatic algebra, Forum Math. Sigma 5 (2017), Paper No. e27, 71, DOI 10.1017/fms.2017.17. MR3732238,
Show rawAMSref \bib{Webster}{article}{ label={Web13}, author={Webster, Ben}, title={Rouquier's conjecture and diagrammatic algebra}, journal={Forum Math. Sigma}, volume={5}, date={2017}, pages={Paper No. e27, 71}, review={\MR {3732238}}, doi={10.1017/fms.2017.17}, }

Article Information

MSC 2020
Primary: 05E10 (Combinatorial aspects of representation theory), 20C30 (Representations of finite symmetric groups)
Author Information
C. Bessenrodt
Institut für Algebra, Zahlentheorie und Diskrete Mathematik, Leibniz Universität Hannover, D-30167 Hannover, Germany
bessen@math.uni-hannover.de
MathSciNet
C. Bowman
Department of Mathematics, University of York, Heslington, York, YO10 5DD, United Kingdom
Chris.Bowman-Scargill@york.ac.uk
MathSciNet
L. Sutton
Okinawa Institute of Science and Technology, Okinawa, Japan 904-0495
louise.sutton@oist.jp
ORCID
MathSciNet
Additional Notes

The second author would like to thank both the Alexander von Humboldt Foundation and the Leibniz Universität Hannover for financial support and an enjoyable summer. The third author was supported by Singapore MOE Tier 2 AcRF MOE2015-T2-2-003.

Journal Information
Transactions of the American Mathematical Society, Series B, Volume 8, Issue 33, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2021 by the authors under Creative Commons Attribution 3.0 License (CC BY 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/btran/70
  • MathSciNet Review: 4350547
  • Show rawAMSref \bib{4350547}{article}{ author={Bessenrodt, C.}, author={Bowman, C.}, author={Sutton, L.}, title={Kronecker positivity and 2-modular representation theory}, journal={Trans. Amer. Math. Soc. Ser. B}, volume={8}, number={33}, date={2021}, pages={1024-1055}, issn={2330-0000}, review={4350547}, doi={10.1090/btran/70}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.