A proof of the shuffle conjecture

By Erik Carlsson and Anton Mellit

Abstract

We present a proof of the compositional shuffle conjecture by Haglund, Morse and Zabrocki [Canad. J. Math., 64 (2012), 822–844], which generalizes the famous shuffle conjecture for the character of the diagonal coinvariant algebra by Haglund, Haiman, Loehr, Remmel, and Ulyanov [Duke Math. J., 126 (2005), 195–232]. We first formulate the combinatorial side of the conjecture in terms of certain operators on a graded vector space whose degree zero part is the ring of symmetric functions over . We then extend these operators to an action of an algebra acting on this space, and interpret the right generalization of the using an involution of the algebra which is antilinear with respect to the conjugation .

1. Introduction

The shuffle conjecture of Haglund, Haiman, Loehr, Remmel, and Ulyanov Reference HHL05b predicts a combinatorial formula for the Frobenius character of the diagonal coinvariant algebra in pairs of variables, which is a symmetric function in infinitely many variables with coefficients in . By a result of Haiman Reference Hai02, the Frobenius character is given explicitly by

where, up to a sign convention, is the operator which is diagonal in the modified Macdonald basis defined in Reference BGHT99. The original shuffle conjecture states

Here is a Dyck path of length , and is some extra data called a “word parking function” depending on . The functions are statistics associated to a Dyck path and a parking function, and is a monomial in the variables . It was shown in Reference HHL05b that this sum, denoted , is symmetric in the variables, and so it does at least define a symmetric function. It was also shown there that the shuffle conjecture included many previous conjectures and results about the -Catalan numbers, and other special cases Reference GH96Reference GH02Reference Hag03Reference EHKK03Reference Hag04. Remarkably, had not even been proven to be symmetric in the variables until now, even though the symmetry of is obvious. The name “shuffle conjecture” has to do with the fact that the coefficient of in equation Equation 1.1 can be expressed in terms of parking functions that are -shuffles”. See Conjecture 6.1 of Haglund’s book Reference Hag08 for a detailed explanation, and Chapter 6 in general for a thorough introduction to this topic.

In Reference HMZ12 Haglund, Morse, and Zabrocki conjectured a refinement of the original conjecture which partitions by specifying the points where the Dyck path touches the diagonal called the “compositional shuffle conjecture”. The refined conjecture states

Here is a composition, i.e., a finite list of positive integers specifying the gaps between the touch points of . The function is defined below as a composition of creation operators for Hall–Littlewood polynomials in the variable . They proved that

implying that Equation 1.2 does indeed generalize Equation 1.1. The right-hand side of Equation 1.2 will be denoted by . A desirable approach to proving Equation 1.2 would be to determine a recursive formula for and to interpret the result in terms of some commutation relations for . Indeed, this approach has been applied in some important special cases; see Reference GH02Reference GXZ12Reference Hic12. In Reference GXZ12, for instance, the authors devise a recursive formula (Proposition 3.12) to prove the Catalan case of the compositional conjecture, extending the results of Reference GH02. Unfortunately, no such recursion is known in the general case, and so an even more refined function is needed.

In this paper, we will construct the desired refinement as an element of a larger vector space of symmetric functions over with additional variables adjoined, where is the length of the composition ,

In our first result (Theorem 4.11) we will explain how to recover from , which is defined by an explicit recursion. In fact, while they live in different vector spaces, the recursions for are similar to the recursions for the Catalan case in Reference GXZ12. We make this connection precise in Proposition 4.14, which explains how the latter formulas follow as a special case.

We then define a pair of algebras and which are isomorphic by an antilinear isomorphism with respect to the conjugation , as well as an explicit action of each on the direct sum . We will then prove that there is an antilinear involution on which intertwines the two actions (Theorem 7.4) and represents an involutive automorphism on a larger algebra . This turns out to be the essential fact that relates the to .

The compositional shuffle conjecture (Theorem 7.5) then follows as a simple corollary from the following properties:

(i)

There is a surjection coming from

which maps a monomial in the variables to an element which is similar to and maps to , up to a sign.

(ii)

The involution commutes with and maps to .

(iii)

The restriction of to agrees with composed with a conjugation map which essentially exchanges the and .

It then becomes clear that these properties imply Equation 1.2.

While the compositional shuffle conjecture is clearly our main application, the shuffle conjecture has been further generalized in several remarkable directions such as the rational compositional shuffle conjecture, and relationships to knot invariants, double affine Hecke algebras, and the cohomology of the affine Springer fibers; see Reference BGLX14Reference GORS14Reference GN15Reference Neg13Reference Hik14Reference SV11Reference SV13. We hope that future applications to these fascinating topics will be forthcoming.

2. The compositional shuffle conjecture

2.1. Plethystic operators

A -ring is a ring with a family of ring endomorphisms satisfying

Unless stated otherwise, the endomorphisms are defined by for each variable such as . The ring of symmetric functions over the -ring is a free -ring with generator , and it will be denoted . We will employ the standard notation used for plethystic substitution defined as follows: given an element and in some -ring , the plethystic substitution is the image of the homomorphism from defined by replacing by . For instance, we would have

See Reference Hai01 for a reference.

The modified Macdonald polynomials Reference GHT99 will be denoted

where is the integral form of the Macdonald polynomial Reference Mac95, and

The operator is defined by

Note that our definition differs from the usual one from Reference BGHT99 by the sign . We also have the sequences of operators given by the formulas

where is the plethystic exponential and denotes the operation of taking the coefficient of of a Laurent power series. Our definition again differs from the one in Reference HMZ12 by a factor . For any composition , let denote the composition , and similarly for .

Finally, we denote by the involutive automorphism of obtained by sending to . We denote by the -ring automorphism of obtained by sending to and by its composition with , i.e.,

2.2. Parking functions

We now recall the combinatorial background to state the shuffle conjecture, for which we refer to Haglund’s book Reference Hag08. We consider the infinite grid on the top right quadrant of the plane. Its intersection points are denoted as with . For each cell of the grid, its coordinates are the coordinates of the top right corner. Thus indexes the columns and indexes the rows. Let be the set of Dyck paths of all lengths. A Dyck path of length is a grid path, from to consisting of North and East steps, that stays above the main diagonal . For denote by its length . For , let

This is the set of cells between the path and the diagonal. Let denote the number of cells in the row . The area sequence is the sequence and we have .

Let be the cells immediately to the right of the North steps. The sequence is called the coarea sequence, and we have for all .

We have the statistic and the set defined by

For , we say that attacks .

For any , the set of word parking functions associated to is defined by

In other words, the elements of are -tuples of positive integers which, when written from bottom to top to the right of each North step, are strictly decreasing on cells such that one is on top of the other. For any , let

We note that both of these conditions differ from the usual notation in which parking functions are expected to increase rather than decrease, and in which the inequalities are reversed in the definition of . This corresponds to choosing the opposite total ordering on everywhere, which does not affect the final answer and is more convenient for the purposes of this paper.

Let us call the touch composition of if are the lengths of the gaps between the points where touches the main diagonal starting at the lower left. Equivalently, and the numbers , , , are the positions of in the area sequence .

Example 2.1.

Let be the following Dyck path of length described in Figure 1. Then we have

whence , . The labels shown above correspond to the vector , which we can see is an element of because we have , , . We then have

giving .

2.3. The shuffle conjectures

For any infinite set of variables , let . In this notation, the original shuffle conjecture Reference HHL05b states

Conjecture (Reference HHL05b).

We have

In particular, the right-hand side is symmetric in the , and in .

The stronger compositional shuffle conjecture Reference HMZ12 states

Conjecture (Reference HMZ12).

For any composition , we have

2.4. From to

In this paper, we will prove an equivalent version of this conjecture, as obtained in Reference LN14, Theorem 14, by applying the to bijection from Reference HL05 and Reference Hag08. We include our construction of this bijection because it seems to be different from the original one, and to demonstrate that it comes naturally from analysis of the attack relation. An important property of our construction is that it comes with a natural lift from Dyck paths to parking functions.

From any pair , we will obtain a pair , by a procedure described below. After the end of this section we will only work with , so we will drop the apostrophe.

The Dyck path is obtained as follows: sort the cells in the reading order, i.e., in increasing order by the corresponding labels , using the row index to break ties. Equivalently, we read the cells by diagonals from bottom to top, and from left to right in each diagonal. For instance, for the path from Example 2.1, the list would be

Let be the position of the cell in this list. This defines a permutation . In the example case, we would get

Now we observe that for each the cell attacks all the subsequent cells in the reading order whose position is before the position where we would place if it were an element of the list.

More precisely, there is a unique Dyck path for which

The map is the desired bijection. To see the bijectivity one can either use Reference Hag08 or see Remark 2.3.

If is the Dyck path from our example, then would be given by the path in Figure 2.

The above statistics can be translated into new statistics under this bijection. First, it is clear from the construction that . We next explain how to calculate from . For any path, we obtain a new Dyck path called the “bounce path” as follows: Start at the origin , and begin moving North until contact is made with the first East step of . Then start moving East until contacting the diagonal. Then move North until contacting the path again, and so on. Note that contacting the path means running into the left endpoint of an East step, but passing by the rightmost endpoint does not count, as illustrated below. The bounce path splits the main diagonal into the bounce blocks. We number the bounce blocks starting from and define the bounce sequence in such a way that for any the cell belongs to the th block. We then define

Another way to describe this construction is to say that , and if are the smallest indices for which and for some , then is the smallest index with such that . This description and implies , hence ; see Reference Hag08 for an alternative treatment.

For the path above, the bounce path is shown in Figure 3 with the original path in gray. The bounce sequence is given by the numbers written under the diagonal. We have

Next, we show how to reconstruct from . For any path of length , let be the number of North steps from until the first East step, which is the same as the length of the first bounce block. Let be the part of the path such that , the result of beginning with North steps starting at the origin, followed by an East step, followed by the contents of . Define numbers by

Note that the path has length for each , and we have and the go down to . Define

Proposition 2.2.

For every Dyck path

For instance, in the example above we would have ,

Proof.

Consider the th touch point of (we count the touch points starting from , i.e., is the th touch point.) It splits into two parts: followed by . Construct a new path of length by taking a step North, then following a translated copy of , then taking a step East, then following a translated copy of . The new path has length and its area is bigger than the area of by . The (area,dinv) to (bounce,area) map applied to gives precisely the path . Thus we have

So the sizes of the gaps between the touch points of are exactly the differences .

Remark 2.3.

The construction we have used in the proof above can also be used to prove the bijectivity of the (area,dinv) to (bounce,area) map. Here is an idea of a proof. First, every Dyck path arises as above for unique and . On the other hand, every Dyck path can be uniquely written as . Thus, iterating the construction, we obtain every Dyck path on each side of the (area,dinv) to (bounce,area) map in a unique way.

Having analyzed the statistics associated to a Dyck path, we turn to the analysis of what happens to word parking functions. The statistic is straightforward. For any , let

so that

For the value of from Example 2.1, we would have

In particular, .

Finally, we reconstruct the word parking function condition. A cell is called a corner of if it is above the path, but both its Southern and Eastern neighbors are below the path. Denote the set of corners by . One can check that the corners of correspond to pairs of cells with one on top of the other in . For instance, from our example we have . More precisely, we have

We therefore define

so that the condition is equivalent to .

Putting this together, we have

Proposition 2.4.

For any composition we have

where is the right-hand side of Equation 2.2.

3. Characteristic functions of Dyck paths

3.1. Simple characteristic function

We are going to study the summand in as a function of . It is convenient to first introduce a simpler object where we drop the assumption and instead sum over all labelings. Given a Dyck path of length , define as follows:

Definition 3.1.

If and is under , i.e., we say that and attack each other. It is not obvious from the definition that defines a symmetric function. In fact, as we point out in Remark 3.6, is actually an example of an LLT polynomial, but we present a proof in our setup here:

Proposition 3.2.

The expression for above is symmetric in the variables , so that Definition 3.1 correctly defines an element of .

Proof.

We take the main idea from the proof of Lemma 10.2 from Reference HHL05a. First note that for each the correspondence is a bijection between the set of Dyck paths of length and the set of subsets satisfying the property

In the proof we will work with instead of and write , instead of , . For each subset , let . Then again satisfies (*).

It is enough to show that is unaffected by interchange of and for any two neighboring indices . For each subset and a function let be the sub-sum of corresponding to sequences where the set of positions of and in is , and the values of outside of are given by . It is enough to show that is symmetric in and . We have

where depends only on and , but does not depend on the positions of or in . Thus we have

where

So it is enough to show that is symmetric in and for any , satisfying (*). We proceed by induction on the size of , the base case being trivial. Fix and , and pick maximal in the sense that for all and for all . Then satisfies (*). Consider the difference

The coefficient is nonzero only if and . If this happens, then . Consider contributions of pairs of the form , , , to . Since pairs do not contribute anything, and pairs contribute precisely when and . Since pairs do not contribute, and pairs contribute precisely when and . The net contribution is the number of such that , which is . We see that , where , denotes the sequence with the entries and removed. Thus we obtain

By the induction hypothesis and are symmetric in . Hence is also symmetric.

Another way to formulate this property is as follows: For a composition , consider the multiset . Consider the sum

Proposition 3.2 simply says that this sum does not depend on the order of the numbers , or equivalently on the linear order on the set of labels. If is the partition with components , then this sum computes the coefficient of the monomial symmetric function in , so we have (set )

We list here a few properties of so that the reader has a feeling of what kind of object it is.

For a Dyck path denote by the reversed Dyck path, i.e., the path obtained by replacing each North step by East step and each East step by North step and reversing the order of steps. Reversing also the order of the components of in Equation 3.1, we see

Proposition 3.3.

Proofs of the following two statements are essentially taken from Reference HHL05a.

Proposition 3.4.
Proposition 3.5.

where “no attack” means that the summation is only over vectors such that for .

Proofs.

We follow Chapter 4 of Reference HHL05a. For an integer and a subset Gessel’s quasi-symmetric function in is given by

For each sequence , its standardization is the unique permutation such that

In other words, sorts pairs in lexicographic order. We notice the following properties:

where is the descent set of . Thus the sum splits as follows:

Since is symmetric by Proposition 3.2, we can apply Proposition 4.2 in Reference HHL05a. Let be the “super” alphabet

consisting of positive letters and negative letters . Let

Then we have the following expression for , :

where

and the summation is over the sequences of elements of . The statement holds for an arbitrary choice of total ordering on . We work with the following ordering:

We extend the definitions of , , and to sequences of elements of ,

so that the properties Equation 3.2 are satisfied. Therefore, we have

Setting we obtain

where is the number of nonstrict inversions of under the path,

Reversing the order of labels, we have

which implies Proposition 3.4.

To prove Proposition 3.5, we set , in Equation 3.3. Applying the involution from the proof of Lemma 5.1 in Reference HHL05a (flipping the sign of the last label that attacks a label with the same absolute value), we see that the terms for such that for some cancel out. In the remaining terms we have whenever . Therefore the comparison between and depends only on and . So we can first sum over sequences in and then over the choices of signs. The latter summation produces an overall factor of , and we obtain Proposition 3.5.

3.2. Weighted characteristic function

To study the summand of in (Equation 2.5) as a function of , we introduce a more general characteristic function. Given a function on the set of corners of some Dyck path of size , let

so in particular becomes

For a constant function , we recover the simpler characteristic function

It turns out that we can express the weighted characteristic function in terms the unweighted one evaluated at different paths. In particular this implies that is symmetric too.

Remark 3.6.

If is the image of under the bijection from section 2.4, then we have that , where are the path symmetric functions from Haglund’s book Reference Hag08, page 95. As Haglund explains, these functions are examples of LLT polynomials of vertical strips, using the description of Bylund and Haiman. In fact, is also an example of an LLT polynomial, but for a disjoint union of single boxes:

where is the area sequence.

Proposition 3.7.

We have that is symmetric in the variables, and so it defines an element of .

Proof.

Let be a Dyck path, and let be one of its corners. We denote by the weight on which is obtained from by setting the weight of to . Let be the Dyck path obtained from by turning the corner inside out, in other words the Dyck path of smallest area which is both above and above . Let be the weight on which coincides with on all corners of which are also corners of and is on other corners. We claim that

To see this, notice that if we group the terms on the right-hand side, then both sides may be written as a sum over vectors . Split both sums according to terms in which , resulting in an additional factor of , or , resulting in an additional weight factor. It is easy to check that both sums agree on both the left and right sides.

The result now follows because we may recursively express any in terms of , which we have already remarked is symmetric.

Example 3.8.

In particular, we can use this to extract from for all . If is any subset of the set of corners, let denote the path obtained by flipping the corners that are in . Then equation Equation 3.6 implies that

For instance, let be the Dyck path in Figure 4.

Then setting for reduces formula Equation 3.4 to a finite sum over 27 terms, from which we can deduce that

Similarly, if , we have

By formula Equation 3.7 we obtain

Example 3.9.

We can check that the Dyck path from Example 3.8 is the unique one satisfying and that . Therefore, using the calculation that followed, we have that

which can be seen to agree with .

Example 3.10.

Though we will not need it, this weighted characteristic function can be used to describe an interesting reformulation of the formula for the modified Macdonald polynomial given in Reference HHL05a. Let be a partition of size . Let us list the cells of in the reading order,

Denote the th cell in this list by .

We say that a cell attacks all cells which are after and before . Thus attacks precisely following cells if and all following cells if . Next construct a Dyck path of length in such a way that with is under the path if and only if attacks . More specifically, the path begins with North steps, then it has pairs, of steps East-North, then North steps followed by East-North pairs and so on until we reach the point . We complete the path by performing East steps.

Note that the corners of precisely correspond to the pairs of cells . We set the weight of such a corner to and denote the weight function thus obtained by . Note that in our convention for we should count noninversions in the corners, while in Reference HHL05a they count “descents”, which translates to counting inversions in the corners. Taking this into account, we obtain a translation of their Theorem 2.2:

4. Raising and lowering operators

Now let be the set of Dyck paths from to , which we will call partial Dyck paths, and let be their union over all . For , let denote the number of North steps. Unlike , the union of the sets over all is closed under the operation of adding a North or East step to the beginning of the path, and any Dyck path may be created in such a way starting with the empty path in . This is the set of paths for which we will develop a recursion. More precisely, we will define an extension of the function to a map from to a new vector space , and prove that certain operators on these vector spaces commute with adding North and East steps.

Given a polynomial depending on variables define

We can easily check that . We can recognize these operators as a simple modification of Demazure–Lusztig operators. The following can be checked by direct computation:

Proposition 4.1.

We have the following relations:

Definition 4.2.

Let , and let

Define operators

by

and

for .

Remark 4.3.

Note that the operator is related to the operators,

for which do not depend on .

We now claim the following theorem:

Theorem 4.4.

For any Dyck path of size , let denote the corresponding sequence of plus and minus symbols where a plus denotes an East step, and a minus denotes a North step reading from bottom left to top right. Then

as an element of .

Example 4.5.

Let be the Dyck path from Example 3.8. We have that

which agrees with the value calculated for .

Combining this result with equation Equation 3.7 implies the following:

Corollary 4.6.

The following procedure computes : start with , follow the path from right to left applying for each corner of , and for each North (resp. East) step that is not a side of a corner.

4.1. Rank experiment

The proof of Theorem 4.4 will be divided into several parts. However, before we proceed to the proof of Theorem 4.4, we would like to explain why we expected such a result to hold, and how we obtained it. In fact, the definition of from equation Equation 4.5 in the proof below actually came first, and was discovered using computer experimentation, as we now explain.

First note that the number of Dyck paths of length is given by the Catalan number which grows exponentially with . The dimension of the degree part of is the number of partitions of size , which grows subexponentially. For instance for , we have five Dyck paths but only three partitions. Thus there must be linear dependences between different .

Now fix a partial Dyck path . For each partial Dyck path we can reflect and concatenate it with to obtain a full Dyck path of length . We may then consider its character . We keep , fixed and vary , , thus obtaining a map . The map is a map from to the vector space of maps from to , which is very high dimensional, because both the set is infinite and is infinite dimensional. A priori, it could be the case that the images of the elements of in are linearly independent. However, computer experiments convinced us that it is not the case, and that there should be a vector space whose dimension is generally smaller than the size of . In fact, by restricting to be bounded by some arbitrary but large enough cut-off value, we were able to predict that the dimension of this space stabilizes to a very simple formula, which is the dimension of , the degree component of as it is defined above.

We therefore predicted the existence of a commutative diagram

for some map , whose image spans all of . This ultimately led to the guess of the formula for in Equation 4.5 as the correct extension of . It is not at all trivial to deduce this formula from the dimension of , and indeed, some substantial guesswork was required. However, the validity of any particular guess can be determined experimentally, by testing if its kernel in the -span of agrees with the kernel of . Clearly the existence of a testable criterion such as this makes the problem of determining experimentally much more reasonable.

Once the definition of was conjectured, finding formulas for that satisfy Equation 4.6 turned out to be relatively straightforward.

4.2. Characteristic functions of partial Dyck paths

The following definition is motivated by Proposition 3.5. Let . Let be a tuple of distinct numbers. The elements of will be called special. Let

The second condition on is the “no attack” condition as before. The first condition says that we put the special labels in the positions as prescribed by . Let

Here we use variables .

Suppose is a permutation, i.e., for all . Set for and for . We denote

Let us group the summands in Equation 4.3 according to the positions of special labels. More precisely, let such that and such that for and for , . Set

Let be all the positions not in . Let be the unique Dyck path of length such that if and only if . We have

where

By Proposition 3.5 we have

In particular is a symmetric function in and it makes sense to define

Remark 4.7.

The identity (Equation 4.4) also implies that the coefficients of are polynomials in , and it gives a way to express in terms of the characteristic functions for all .

For , we recover :

Thus, it suffices to prove that

4.3. Raising operator

We begin with the first case. Let so that , and we need to express in terms of . Let be the following sequence:

Then we have a bijection obtained by sending

to

This is possible because does not attack in . We clearly have , which implies

where both sides are written in terms of the variables . When we pass to the variables , on the left, we have

but on the right we have

Thus we need to perform the substitution ,

Performing the transformation Equation 4.5, we obtain

To finish the computation, we need to relate and . We first note that can be obtained from by successively swapping neighboring labels. Let and

so that . It is clear that can be obtained from by interchanging the labels and .

We show below (Proposition 4.8) that this kind of interchange is controlled by the operator :

This implies

When we insert this equation into Equation 4.7, we arrive at

4.4. Swapping operators

Proposition 4.8.

For any , as above and special, suppose that is not special or . Then we have

where is the transposition , for .

Proof.

We decompose both sides as follows. For any let be the set of indices where . For write if and for all . This defines an equivalence relation on . The sum Equation 4.3 is then decomposed as follows:

where

which does not depend on the choice of a representative in the equivalence class , and

Let be the bijection defined by . This bijection respects the equivalence relation , and we have . Moreover, we have . We now make the stronger claim that for any

which would imply the statement by summing over all equivalence classes.

For each the set is decomposed into a disjoint union of runs, i.e., subsets

such that in each run attacks for all and elements of different runs do not attack each other. Because of the nonattacking condition, the labels must alternate between and does not attack . Thus to fix in each equivalence class, it is enough to fix for each run. Suppose the runs of have lengths and the first values of in each run are respectively.

With this information can be computed as follows:

where

For instance, let , and let be the Dyck path in Figure 5, and let

Let . Then we have , which decomposes into two runs and . So we have , , and we obtain

Note that by the assumption on we have , while can take arbitrary values for . This implies

On the other hand we have

Now notice that for all the sum is symmetric in . The operator commutes with multiplication by symmetric functions and satisfies

This establishes Equation 4.10 and the proof is complete.

Remark 4.9.

The arguments used in the proof can be used to show that in the case when are both not special, the function is symmetric in . In particular, we can obtain a direct proof of the fact that is symmetric in the variables for , without use of Proposition 3.5.

4.5. Lowering operator

We now turn to the remaining identity . Assume , so that . We observe that

where

and to get to the second equality, we have summed over all possible values of that do not result in an attack. It is convenient to set . Using Proposition 4.8, we can characterize by

Now notice that there is a unique expansion

The advantage over the more obvious expansion in powers of is that each coefficient is symmetric in the variables As a result, we have that

where

The extra symmetry in the variable is used to pass by multiplication by .

Now we need an explicit formula for :

Proposition 4.10.

Denote , . We have

Proof.

Denote the right-hand side by . The proof goes by induction on . For both sides are equal to . Thus it is enough to show that

Use to write as

Now does not contain the variables , , so we have

Using the formula

which is straightforward to check, we can evaluate

which matches by Equation 4.13.

Now, if we sum over all , we obtain

Thus

This implies

On the other hand were defined in such a way that

Substituting for gives

Comparing Equation 4.15 and Equation 4.16, we obtain

This can be seen to coincide with , establishing the second case of Equation 4.6. Thus the proof of Theorem 4.4 is complete.

4.6. Main recursion

We now show how to express all of using our operators:

Theorem 4.11.

If is a composition of length , we have

where is defined by the recursion relations

Proof.

For any let denote the subset of partial Dyck paths that begin with an East step. For let . Define functions by

Given a composition of length , let

By the definition of , every element of is of the form for a unique element so that by Corollary 4.6 we have

Let

so that . It suffices to show that satisfies the relations Equation 4.17, and so agrees with .

For a composition of length and , we have a map as follows: . Clearly . From the definition of , we see the following relation:

Next we compute . For , we have

This implies

so in particular depends only on and .

Since every nonempty Dyck path can be obtained as in a unique way, we obtain for every composition of length ,

It is not hard to see that this identity precisely translates to the relations Equation 4.17 for .

Example 4.12.

Using Theorem 4.11, we find that

where We may then check that

Example 4.13.

Let be a composition of norm and length and consider the polynomials in given by

which encode the Catalan case of the compositional shuffle conjecture. This case of the conjecture was proved in Reference GXZ12, using recursions (Proposition 3.12) that are similar to those defining Equation 4.17, and it is natural to ask if they follow as a special case. Indeed, this can be achieved with help of the following proposition:

Proposition 4.14.

Let be Dyck path of size ending with East steps, and let denote the corresponding sequence of plus and minus symbols, as in Theorem 4.4. Then for , we have

Proof.

We prove this by induction on , the number of symbols.

First, suppose the final step is a plus. Let denote the homomorphism on the right-hand side of Equation 4.18. Using the definition Equation 4.1 of , and the observation that , we have that

The case now follows from the above equation, the induction step, and the fact that the path remains the same.

The second case follows easily from the formula

for homogeneous of degree in the symmetric function part. The base case is obvious.

The recursions of Reference GXZ12 now follow fairly easily from Equation 4.17 as a special case, if Theorem 4.11 is taken as the definition of . The relations for were not discovered this way, and the authors only noticed this connection after following up on a useful suggestion from one of the referees, for which the authors are grateful.

5. Operator relations

We have operators

where is the projection onto , and the others are defined as above. It is natural to ask for a complete set of relations between them. They are formalized in the following algebra:

Definition 5.1.

The Dyck path algebra (over ) is the path algebra of the quiver with vertex set , arrows from to , arrows from to for , and loops from to subject to the following relations:

where in each identity denotes the index of the vertex where the respective paths begin. We have used the same letters to label the th loop at every node to match with the previous notation. To distinguish between different nodes, we will use where is the idempotent associated with node .

We will prove:

Theorem 5.2.

The operators Equation 5.1 define a representation of on . Furthermore, we have an isomorphism of representations

which sends to , and maps isomorphically onto .

The proof will occupy the rest of this section. We begin by establishing that we have a defined a representation of the algebra.

Lemma 5.3.

The operators and satisfy the relations of Definition 5.1.

Proof.

The first line is just Proposition 4.1, and most follow from definition. The first one that does not is the commutation relation of with . We have

using the braiding relations.

For the next, we have

The last can be removed because its argument is symmetric in and , and we obtain .

The next identity is more technical. The operator image of consists of elements of the form , where is symmetric in and . Thus we need to check that vanishes on such elements. Let us evaluate on

where does not contain the variables and , and . We obtain

This expression is antisymmetric in , by Corollary 3.4 of Reference HMZ12, which implies our identity.

Next using the previous relations and Lemma 5.4 below write

Similarly,

To establish the isomorphism, we first show that we can produce the operators of multiplication by from .

Lemma 5.4.

For , we have

Proof.

First, we endow with the following twisted action of :

for , and It can be checked that the operators , intertwine this action:

For the second one, for instance, it suffices to assume that . Then by the definition of given in Equation 4.2, we have

We will not need this, but in fact, if , , and is their concatenation, then we must also have that

Since the operators on both sides commute with the twisted action of introduced above, we may assume without loss of generality that is a polynomial of .

Write the left-hand side of the first identity as

The operator in the first summand involves only the variable . Thus we can write the left-hand side as

Hence it is enough to prove

It is clear that none of the operations involve the variables . Thus we can assume for without loss of generality. Direct computation gives

Thus the left-hand side equals

The second relation is easy.

The operators of multiplication by are characterized by these relations and therefore come from elements of . We next establish the relations these operators satisfy within :

Lemma 5.5.

For define elements by solving for in the identities Equation 5.4. Then the following identities hold in :

Proof.

Note that can be written as

Our task becomes easier if we notice that it is enough to check the first identity for and , the second one for , the third one for and the last one for , . The other cases can be deduced from these by applying the -operators.

For we have

Similarly, we verify that commutes with hence with for .

Reversing the arguments in Equation 5.2 and Equation 5.3, we verify the second and the third identities.

Thus it is left to check that . We assume . Write the left-hand side as

using the already established commutation relations and that to swap and . Performing the cancellation, we obtain .

The following lemma completes the proof of the theorem:

Lemma 5.6.

The elements of the form

with form a basis of . Furthermore, the representation maps these elements to a basis of .

Proof.

We first show that elements of the form Equation 5.6, with no condition on the span . It suffices to check that the span of these elements is invariant under , , and . This can be done by applying the following reduction rules that follow from the definition of and Lemma 5.5:

The next step is to reduce the spanning set. We can use the following identity, which follows from :

Note that commutes with . Suppose . Then we can rewrite the above identity as

Using , for , and , we can rewrite the identity as vanishing of a linear combination of terms of the form Equation 5.6, and the lexicographically smallest term is precisely

Thus we can always reduce terms of the form Equation 5.6 which violate the condition to a linear combination of lexicographically greater terms, showing that the subspace in the lemma at least spans .

We now show that they map to a basis of , which also establishes that they are independent, completing the proof. Consider the image of the elements of our spanning set

Notice that is a partition, so

is a multiple of the Hall–Littlewood polynomial . These polynomials form a basis of the space of symmetric functions, thus the elements Equation 5.7 form a basis of .

6. Conjugate structure

It is natural to ask if there is a way to extend to the spaces recovering the original operator at . What we have found is that it is simpler to extend the composition

where the conjugation simply makes the substitution , is the Weyl involution up to a sign, and denotes the composition. This is a very interesting operator, which in fact is an antilinear involution on corresponding to dualizing vector bundles and tensoring with in the Haiman–Bridgeland–King–Reid picture, which identifies with the equivariant -theory of the Hilbert scheme of points in the complex plane Reference BKR01. The key to our proof is to extend this operator to an antilinear involution on every , suggesting that should have a geometric interpretation as well. Since this paper was completed, E. Gorsky and the authors have discovered this connection in terms of a smooth subscheme of the flag Hilbert scheme. This will be explained in an upcoming paper, which also explains a family of ways to extend the operator itself in addition to the involution , also in terms of line bundles.

We will define the operator, which was discovered experimentally to have nice properties, by explicitly constructing the action of conjugated by the conjectural involution . Let , and label the corresponding generators by . Denote by the image of under the isomorphism from to which sends generators to generators, and which is antilinear with respect to .

Theorem 6.1.

There is an action of on given by the assignment

where and is the operator which sends to for and to . Furthermore, it satisfies the additional relations

for any .

The statement is equivalent to validity of a certain set of relations satisfied by the operators. These will be verified in the following propositions.

First we list the obvious relations:

Proposition 6.2.
Proof.

Easy from the definition.

To verify the rest of the relations, we use an approach similar to the one used in the proof of Lemma 5.4. Now we need not just one, but a family of twisted multiplications: For , , put

It is not hard to show that they satisfy

where , and , the first identity holds for , and the second one for .

Let us first verify

Proposition 6.3.
Proof.

Rewrite it as

Multiplying both sides by produces an equivalent relation, which can be reduced to

Note that the image of consists of elements which are symmetric in , . Let

It is enough to show that vanishes on elements of that are symmetric in , . We have (recall that )

Thus it is enough to verify vanishing of on symmetric polynomials of . We evaluate on :

where is the operator . For any monomial and integer , we have operator identities

thus we have

Performing the cancellations, we arrive at

This expression is antisymmetric in , by Reference HMZ12, Corollary 3.4. Thus Equation 6.5 is true.

Next we have to check:

Proposition 6.4.
Proof.

Multiplying both sides by and using the easier relations, rewrite it as

Again, because of the commutation relations with the twisted multiplication by symmetric functions and , it is enough to evaluate the left-hand side on for all . We obtain

We use the identity to write the left-hand side as a linear combination of terms with . The coefficient in front of each term with is

By a direct computation,

and we are done.

At this point, we have established the fact that the operators given by Equation 6.2 define an action of on . Also we have established the second relation in Equation 6.3. The last relation is obvious. The first and the third are verified below:

Proposition 6.5.
Proof.

Using Equation 6.4 we see that the operator satisfies the two properties,

for , , . By definition (on )

thus (again, on )

From this expression we see, using Equation 6.4 once again, that satisfies the same two properties as the operator ,

for , , . Thus it is enough to verify the first identity on . The right-hand side is . The left-hand side is

so the first identity holds.

It is enough to verify the second identity for because the general case can be deduced from this one by applying the -operators. For the identity , expressing , in terms of , and the -operators, we arrive at the equivalent identity

If we denote by either of the two sides, we have

for , , . Thus it is enough to verify the identity on (). Applying to both sides, the identity to be verified is

The left-hand side is evaluated to

The right-hand side is evaluated to

with

and the identity follows.

This completes our proof of Theorem 6.1.

We also have the following proposition, which we will use to connect the conjugate action with .

Proposition 6.6.

For a composition of length let

Then the following recursions hold:

Proof.

The first identity easily follows from the explicit formula for . For , we have

Therefore it is enough to verify the following identity for any :

We group the terms on the right-hand side by and the sum becomes

We have used the identity

which can be obtained by applying to Proposition 5.2 of Reference HMZ12:

Thus the right-hand side of (Equation 6.6) is evaluated to the expression

Thus we need to prove

Then the left-hand side as a polynomial in indeed has the right coefficient of . The coefficient of for is

So it is enough to show

Using (Equation 6.7) again, we see that the left-hand side equals

because by Reference HMZ12, Proposition 3.5 and .

7. The involution

Definition 7.1.

Consider and as algebras over , and let be the quotient of the free product of and by the relations

Remark 7.2.

For any the affine Hecke algebra is the algebra generated over by modulo relations

The positive part is defined as the subalgebra of generated by and , or equivalently as the algebra generated over by modulo the same relations. We have a natural homomorphism which can be shown to be injective using Lemma 5.6. It is tempting to guess that in a similar way the subalgebra of generated by , , and is isomorphic to the positive part of the double affine Hecke algebra . To fix a version of which is close to our notations, we start with relations Reference SV11, (2.1)–(2.7) and perform substitutions , , , , followed by reversal of the order of generators in each monomial. So is defined over by generators and relations of two copies of (the second copy is transformed by , )

and two extra relations. The first one is

which can be deduced in from Equation 5.4 and . The second relation is

The following identity can be deduced from the rest of the relations:

Thus we expect to have

However this does not hold in . Instead we have

So we see that we do not obtain a natural homomorphism . One way to repair the situation is to introduce the “partially symmetrized” by starting with the DAHA in infinitely many generators , , (), and then symmetrizing in generators with . For instance for we expect to coincide with the positive part of the elliptic Hall algebra, which is the stable limit of spherical DAHAs as shown in Reference SV11. Details of this construction will be provided in a future publication.

We now prove

Theorem 7.3.

The operations , , , , define an action of on . Furthermore, the kernel of the natural map that sends to is given by where is the ideal generated by

In particular, we have an isomorphism .

Proof.

Theorem 6.1 shows that we have a map of modules that restricts to the isomorphism of Theorem 5.2 on the subspace , so in particular is surjective. Furthermore, the last relation of Equation 6.3 shows that it descends to a map , which must still be surjective. We have the following commutative diagram.

Thus we have an inclusion and it remains to show that the image of in is the entire space. We do so by induction: notice that both and have a grading by the total degree in , , and , as all the relations are homogeneous. For instance, and have degree , and has degree for all . Denote the space of elements of degree in , by , respectively. We need to prove . The base cases , are clear.

For the induction step, suppose , for and let . It is enough to show that . By Lemma 5.6, we can assume that is in the canonical form Equation 5.6. We therefore must check three cases: for , for , and for . In the first case we have . In the second case we have . In the third case we have

Now we use expansion of in terms of the generators , , and . Because of the commutation relations between and it is enough to consider two cases, and for . In the first case we have if and if . In the second case we have . In all cases the claim is reduced to the induction hypothesis.

Now by looking at the defining relations of , we make the remarkable observation that there exists an involution of that permutes and and is antilinear with respect to the conjugation on the ground field ! Furthermore, this involution preserves the ideal , and therefore induces an involution on via the isomorphism of Theorem 7.3.

Theorem 7.4.

There exists a unique antilinear degree-preserving automorphism satisfying

Moreover, we have

(i)

is an involution, i.e., .

(ii)

For any composition we have

(iii)

On , we have , where is the involution sending , , to , , , respectively (see Equation 6.1).

Proof.

The automorphism is induced from the involution of , from which part (i) follows immediately. Part (ii) follows from applying to the relations of Proposition 6.6.

Finally, let be the operators

and let be the operator of multiplication by . It is easy to verify that

Thus it follows that

Let on . Then

It was shown in Reference GHT99 that satisfies the same commutation relations, and that one can obtain all symmetric functions starting from and successively applying and . Thus , proving part (iii).

The compositional shuffle conjecture now follows easily:

Theorem 7.5.

For a composition of length , we have

Proof.

Using Theorems 4.11 and 7.4, we have

Acknowledgments

The authors would like to thank François Bergeron, Adriano Garsia, Mark Haiman, Jim Haglund, Fernando Rodriguez-Villegas, and Guoce Xin for many valuable discussions on this and related topics. The authors also acknowledge the International Center for Theoretical Physics, Trieste, Italy, where most of the research for this paper was performed. Erik Carlsson was also supported by the Center for Mathematical Sciences and Applications at Harvard University during some of this period, which he gratefully acknowledges.

Figures

Figure 1.

Example of a Dyck path of size 8.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \draw[help lines] (0,0) grid (8,8); \draw[dashed,color=gray] (0,0)--(8,8); \draw[->,very thick] (0,0)--(0,1); \draw[->,very thick] (0,1)--(1,1); \draw[->,very thick] (1,1)--(1,2); \draw[->,very thick] (1,2)--(1,3); \draw[->,very thick] (1,3)--(1,4); \draw[->,very thick] (1,4)--(2,4); \draw[->,very thick] (2,4)--(2,5); \draw[->,very thick] (2,5)--(2,6); \draw[->,very thick] (2,6)--(3,6); \draw[->,very thick] (3,6)--(4,6); \draw[->,very thick] (4,6)--(5,6); \draw[->,very thick] (5,6)--(6,6); \draw[->,very thick] (6,6)--(6,7); \draw[->,very thick] (6,7)--(6,8); \draw[->,very thick] (6,8)--(7,8); \draw[->,very thick] (7,8)--(8,8); \node at (.5,.5) {9}; \node at (1.5,1.5) {5}; \node at (1.5,2.5) {2}; \node at (1.5,3.5) {1}; \node at (2.5,4.5) {5}; \node at (2.5,5.5) {2}; \node at (6.5,6.5) {3}; \node at (6.5,7.5) {2}; \end{tikzpicture}
Figure 2.

Image of the path from Figure 1 under the to bijection.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \draw[help lines] (0,0) grid (8,8); \draw[dashed,color=gray] (0,0)--(8,8); \draw[->,very thick] (0,0)--(0,1); \draw[->,very thick] (0,1)--(0,2); \draw[->,very thick] (0,2)--(0,3); \draw[->,very thick] (0,3)--(1,3); \draw[->,very thick] (1,3)--(2,3); \draw[->,very thick] (2,3)--(2,4); \draw[->,very thick] (2,4)--(3,4); \draw[->,very thick] (3,4)--(3,5); \draw[->,very thick] (3,5)--(4,5); \draw[->,very thick] (4,5)--(4,6); \draw[->,very thick] (4,6)--(4,7); \draw[->,very thick] (4,7)--(5,7); \draw[->,very thick] (5,7)--(6,7); \draw[->,very thick] (6,7)--(7,7); \draw[->,very thick] (7,7)--(7,8); \draw[->,very thick] (7,8)--(8,8); \end{tikzpicture}
Figure 3.

The bounce path of the path in Figure 2.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \draw[help lines] (0,0) grid (8,8); \draw[dashed,color=gray] (0,0)--(8,8); \draw[->,very thick,color=gray] (0,0)--(0,1); \draw[->,very thick,color=gray] (0,1)--(0,2); \draw[->,very thick,color=gray] (0,2)--(0,3); \draw[->,very thick,color=gray] (0,3)--(1,3); \draw[->,very thick,color=gray] (1,3)--(2,3); \draw[->,very thick,color=gray] (2,3)--(2,4); \draw[->,very thick,color=gray] (2,4)--(3,4); \draw[->,very thick,color=gray] (3,4)--(3,5); \draw[->,very thick,color=gray] (3,5)--(4,5); \draw[->,very thick,color=gray] (4,5)--(4,6); \draw[->,very thick,color=gray] (4,6)--(4,7); \draw[->,very thick,color=gray] (4,7)--(5,7); \draw[->,very thick,color=gray] (5,7)--(6,7); \draw[->,very thick,color=gray] (6,7)--(7,7); \draw[->,very thick,color=gray] (7,7)--(7,8); \draw[->,very thick,color=gray] (7,8)--(8,8); \draw[->,very thick] (0,0)--(0,1); \draw[->,very thick] (0,1)--(0,2); \draw[->,very thick] (0,2)--(0,3); \draw[->,very thick] (0,3)--(1,3); \draw[->,very thick] (1,3)--(2,3); \draw[->,very thick] (2,3)--(3,3); \draw[->,very thick] (3,3)--(3,4); \draw[->,very thick] (3,4)--(3,5); \draw[->,very thick] (3,5)--(4,5); \draw[->,very thick] (4,5)--(5,5); \draw[->,very thick] (5,5)--(5,6); \draw[->,very thick] (5,6)--(5,7); \draw[->,very thick] (5,7)--(6,7); \draw[->,very thick] (6,7)--(7,7); \draw[->,very thick] (7,7)--(7,8); \draw[->,very thick] (7,8)--(8,8); \node at (.7,.3) {0}; \node at (1.7,1.3) {0}; \node at (2.7,2.3) {0}; \node at (3.7,3.3) {1}; \node at (4.7,4.3) {1}; \node at (5.7,5.3) {2}; \node at (6.7,6.3) {2}; \node at (7.7,7.3) {3}; \end{tikzpicture}
Figure 4.
\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \draw[help lines] (0,0) grid (3,3); \draw[dashed, color=gray] (0,0)--(3,3); \draw[->,very thick] (0,0)--(0,1); \draw[->,very thick] (0,1)--(0,2); \draw[->,very thick] (0,2)--(1,2); \draw[->,very thick] (1,2)--(2,2); \draw[->,very thick] (2,2)--(2,3); \draw[->,very thick] (2,3)--(3,3); \end{tikzpicture}
Figure 5.
\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \draw[help lines] (0,0) grid (8,8); \draw[dashed, color=gray] (0,0)--(8,8); \draw[->,very thick] (0,3)--(0,4); \draw[->,very thick] (0,4)--(1,4); \draw[->,very thick] (1,4)--(2,4); \draw[->,very thick] (2,4)--(2,5); \draw[->,very thick] (2,5)--(3,5); \draw[->,very thick] (3,5)--(4,5); \draw[->,very thick] (4,5)--(4,6); \draw[->,very thick] (4,6)--(5,6); \draw[->,very thick] (5,6)--(5,7); \draw[->,very thick] (5,7)--(5,8); \draw[->,very thick] (5,8)--(6,8); \draw[->,very thick] (6,8)--(7,8); \draw[->,very thick] (7,8)--(8,8); \node at (.7,.3) {1}; \node at (1.7,1.3) {3}; \node at (2.7,2.3) {2}; \node at (3.7,3.3) {7}; \node at (4.7,4.3) {1}; \node at (5.7,5.3) {7}; \node at (6.7,6.3) {1}; \node at (7.7,7.3) {2}; \end{tikzpicture}

Mathematical Fragments

Equation (1.1)
Equation (1.2)
Example 2.1.

Let be the following Dyck path of length described in Figure 1. Then we have

whence , . The labels shown above correspond to the vector , which we can see is an element of because we have , , . We then have

giving .

Conjecture (Reference HMZ12).

For any composition , we have

Remark 2.3.

The construction we have used in the proof above can also be used to prove the bijectivity of the (area,dinv) to (bounce,area) map. Here is an idea of a proof. First, every Dyck path arises as above for unique and . On the other hand, every Dyck path can be uniquely written as . Thus, iterating the construction, we obtain every Dyck path on each side of the (area,dinv) to (bounce,area) map in a unique way.

Proposition 2.4.

For any composition we have

where is the right-hand side of Equation 2.2.

Definition 3.1.
Proposition 3.2.

The expression for above is symmetric in the variables , so that Definition 3.1 correctly defines an element of .

Equation (3.1)
Proposition 3.4.
Proposition 3.5.

where “no attack” means that the summation is only over vectors such that for .

Equation (3.2)
Equation (3.3)
Equation (3.4)
Remark 3.6.

If is the image of under the bijection from section 2.4, then we have that , where are the path symmetric functions from Haglund’s book Reference Hag08, page 95. As Haglund explains, these functions are examples of LLT polynomials of vertical strips, using the description of Bylund and Haiman. In fact, is also an example of an LLT polynomial, but for a disjoint union of single boxes:

where is the area sequence.

Equation (3.6)
Example 3.8.

In particular, we can use this to extract from for all . If is any subset of the set of corners, let denote the path obtained by flipping the corners that are in . Then equation Equation 3.6 implies that

For instance, let be the Dyck path in Figure 4.

Then setting for reduces formula Equation 3.4 to a finite sum over 27 terms, from which we can deduce that

Similarly, if , we have

By formula 3.7 we obtain

Proposition 4.1.

We have the following relations:

Definition 4.2.

Let , and let

Define operators

by

and

for .

Theorem 4.4.

For any Dyck path of size , let denote the corresponding sequence of plus and minus symbols where a plus denotes an East step, and a minus denotes a North step reading from bottom left to top right. Then

as an element of .

Corollary 4.6.

The following procedure computes : start with , follow the path from right to left applying for each corner of , and for each North (resp. East) step that is not a side of a corner.

Equation (4.3)
Equation (4.4)
Equation (4.5)
Equation (4.6)
Equation (4.7)
Proposition 4.8.

For any , as above and special, suppose that is not special or . Then we have

where is the transposition , for .

Equation (4.10)
Equation (4.13)
Equation (4.15)
Equation (4.16)
Theorem 4.11.

If is a composition of length , we have

where is defined by the recursion relations

Example 4.13.

Let be a composition of norm and length and consider the polynomials in given by

which encode the Catalan case of the compositional shuffle conjecture. This case of the conjecture was proved in Reference GXZ12, using recursions (Proposition 3.12) that are similar to those defining Equation 4.17, and it is natural to ask if they follow as a special case. Indeed, this can be achieved with help of the following proposition:

Proposition 4.14.

Let be Dyck path of size ending with East steps, and let denote the corresponding sequence of plus and minus symbols, as in Theorem 4.4. Then for , we have

Proof.

We prove this by induction on , the number of symbols.

First, suppose the final step is a plus. Let denote the homomorphism on the right-hand side of 4.18. Using the definition Equation 4.1 of , and the observation that , we have that

The case now follows from the above equation, the induction step, and the fact that the path remains the same.

The second case follows easily from the formula

for homogeneous of degree in the symmetric function part. The base case is obvious.

The recursions of Reference GXZ12 now follow fairly easily from Equation 4.17 as a special case, if Theorem 4.11 is taken as the definition of . The relations for were not discovered this way, and the authors only noticed this connection after following up on a useful suggestion from one of the referees, for which the authors are grateful.

Equation (5.1)
Definition 5.1.

The Dyck path algebra (over ) is the path algebra of the quiver with vertex set , arrows from to , arrows from to for , and loops from to subject to the following relations:

where in each identity denotes the index of the vertex where the respective paths begin. We have used the same letters to label the th loop at every node to match with the previous notation. To distinguish between different nodes, we will use where is the idempotent associated with node .

Theorem 5.2.

The operators Equation 5.1 define a representation of on . Furthermore, we have an isomorphism of representations

which sends to , and maps isomorphically onto .

Equation (5.2)
Equation (5.3)
Lemma 5.4.

For , we have

Lemma 5.5.

For define elements by solving for in the identities Equation 5.4. Then the following identities hold in :

Lemma 5.6.

The elements of the form

with form a basis of . Furthermore, the representation maps these elements to a basis of .

Equation (5.7)
Equation (6.1)
Theorem 6.1.

There is an action of on given by the assignment

where and is the operator which sends to for and to . Furthermore, it satisfies the additional relations

for any .

Equation (6.4)
Proposition 6.3.
Proposition 6.6.

For a composition of length let

Then the following recursions hold:

Equation (6.6)
Equation (6.7)
Theorem 7.3.

The operations , , , , define an action of on . Furthermore, the kernel of the natural map that sends to is given by where is the ideal generated by

In particular, we have an isomorphism .

Theorem 7.4.

There exists a unique antilinear degree-preserving automorphism satisfying

Moreover, we have

(i)

is an involution, i.e., .

(ii)

For any composition we have

(iii)

On , we have , where is the involution sending , , to , , , respectively (see Equation 6.1).

Theorem 7.5.

For a composition of length , we have

References

[BGHT99]
F. Bergeron, A. M. Garsia, M. Haiman, and G. Tesler, Identities and positivity conjectures for some remarkable operators in the theory of symmetric functions, Methods Appl. Anal. 6 (1999), no. 3, 363–420. Dedicated to Richard A. Askey on the occasion of his 65th birthday, Part III. MR1803316, Show rawAMSref\bib{Bergeron99identitiesand}{article}{ label={BGHT99}, author={Bergeron, F.}, author={Garsia, A. M.}, author={Haiman, M.}, author={Tesler, G.}, title={Identities and positivity conjectures for some remarkable operators in the theory of symmetric functions}, note={Dedicated to Richard A.\ Askey on the occasion of his 65th birthday, Part III}, journal={Methods Appl. Anal.}, volume={6}, date={1999}, number={3}, pages={363--420}, issn={1073-2772}, review={\MR {1803316}}, } Close amsref.
[BGLX14]
F. Bergeron, A. Garsia, E. Sergel Leven, and G. Xin, Compositional -shuffle conjectures, Int. Math. Res. Not. IMRN 14 (2016), 4229–4270. MR3556418, Show rawAMSref\bib{bergeron2014rational}{article}{ label={BGLX14}, author={Bergeron, Francois}, author={Garsia, Adriano}, author={Sergel Leven, Emily}, author={Xin, Guoce}, title={Compositional $(km,kn)$-shuffle conjectures}, journal={Int. Math. Res. Not. IMRN}, date={2016}, number={14}, pages={4229--4270}, issn={1073-7928}, review={\MR {3556418}}, } Close amsref.
[BKR01]
T. Bridgeland, A. King, and M. Reid, The McKay correspondence as an equivalence of derived categories, J. Amer. Math. Soc. 14 (2001), no. 3, 535–554. MR1824990, Show rawAMSref\bib{BKR}{article}{ label={BKR01}, author={Bridgeland, Tom}, author={King, Alastair}, author={Reid, Miles}, title={The McKay correspondence as an equivalence of derived categories}, journal={J. Amer. Math. Soc.}, volume={14}, date={2001}, number={3}, pages={535--554}, issn={0894-0347}, review={\MR {1824990}}, } Close amsref.
[EHKK03]
E. S. Egge, J. Haglund, K. Killpatrick, and D. Kremer, A Schröder generalization of Haglund’s statistic on Catalan paths, Electron. J. Combin. 10 (2003), Research Paper 16, 21. MR1975766, Show rawAMSref\bib{egge2003sch}{article}{ label={EHKK03}, author={Egge, E. S.}, author={Haglund, J.}, author={Killpatrick, K.}, author={Kremer, D.}, title={A Schr\"oder generalization of Haglund's statistic on Catalan paths}, journal={Electron. J. Combin.}, volume={10}, date={2003}, pages={Research Paper 16, 21}, issn={1077-8926}, review={\MR {1975766}}, } Close amsref.
[GH96]
A. M. Garsia and M. Haiman, A remarkable -Catalan sequence and -Lagrange inversion, J. Algebraic Combin. 5 (1996), no. 3, 191–244. MR1394305, Show rawAMSref\bib{garsia1996remarkable}{article}{ label={GH96}, author={Garsia, A. M.}, author={Haiman, M.}, title={A remarkable $q,t$-Catalan sequence and $q$-Lagrange inversion}, journal={J. Algebraic Combin.}, volume={5}, date={1996}, number={3}, pages={191--244}, issn={0925-9899}, review={\MR {1394305}}, } Close amsref.
[GH02]
A. M. Garsia and J. Haglund, A proof of the -Catalan positivity conjecture, Discrete Math. 256 (2002), no. 3, 677–717. LaCIM 2000 Conference on Combinatorics, Computer Science and Applications (Montreal, QC). MR1935784, Show rawAMSref\bib{garsia2002catalan}{article}{ label={GH02}, author={Garsia, A. M.}, author={Haglund, J.}, title={A proof of the $q,t$-Catalan positivity conjecture}, note={LaCIM 2000 Conference on Combinatorics, Computer Science and Applications (Montreal, QC)}, journal={Discrete Math.}, volume={256}, date={2002}, number={3}, pages={677--717}, issn={0012-365X}, review={\MR {1935784}}, } Close amsref.
[GHT99]
A. M. Garsia, M. Haiman, and G. Tesler, Explicit plethystic formulas for Macdonald -Kostka coefficients, Sém. Lothar. Combin. 42 (1999), Art. B42m, 45. The Andrews Festschrift (Maratea, 1998). MR1701592, Show rawAMSref\bib{garsia1998explicit}{article}{ label={GHT99}, author={Garsia, A. M.}, author={Haiman, M.}, author={Tesler, G.}, title={Explicit plethystic formulas for Macdonald $q,t$-Kostka coefficients}, note={The Andrews Festschrift (Maratea, 1998)}, journal={S\'em. Lothar. Combin.}, volume={42}, date={1999}, pages={Art. B42m, 45}, issn={1286-4889}, review={\MR {1701592}}, } Close amsref.
[GN15]
E. Gorsky and A. Neguţ, Refined knot invariants and Hilbert schemes (English, with English and French summaries), J. Math. Pures Appl. (9) 104 (2015), no. 3, 403–435. MR3383172, Show rawAMSref\bib{gorsky2015refined}{article}{ label={GN15}, author={Gorsky, Eugene}, author={Negu\c t, Andrei}, title={Refined knot invariants and Hilbert schemes}, language={English, with English and French summaries}, journal={J. Math. Pures Appl. (9)}, volume={104}, date={2015}, number={3}, pages={403--435}, issn={0021-7824}, review={\MR {3383172}}, } Close amsref.
[GORS14]
E. Gorsky, A. Oblomkov, J. Rasmussen, and V. Shende, Torus knots and the rational DAHA, Duke Math. J. 163 (2014), no. 14, 2709–2794. MR3273582, Show rawAMSref\bib{gorsky2014torus}{article}{ label={GORS14}, author={Gorsky, Eugene}, author={Oblomkov, Alexei}, author={Rasmussen, Jacob}, author={Shende, Vivek}, title={Torus knots and the rational DAHA}, journal={Duke Math. J.}, volume={163}, date={2014}, number={14}, pages={2709--2794}, issn={0012-7094}, review={\MR {3273582}}, } Close amsref.
[GXZ12]
A. M. Garsia, G. Xin, and M. Zabrocki, Hall-Littlewood operators in the theory of parking functions and diagonal harmonics, Int. Math. Res. Not. IMRN 6 (2012), 1264–1299. MR2899952, Show rawAMSref\bib{GXZ2012}{article}{ label={GXZ12}, author={Garsia, A. M.}, author={Xin, G.}, author={Zabrocki, M.}, title={Hall-Littlewood operators in the theory of parking functions and diagonal harmonics}, journal={Int. Math. Res. Not. IMRN}, date={2012}, number={6}, pages={1264--1299}, issn={1073-7928}, review={\MR {2899952}}, } Close amsref.
[Hag03]
J. Haglund, Conjectured statistics for the -Catalan numbers, Adv. Math. 175 (2003), no. 2, 319–334. MR1972636, Show rawAMSref\bib{haglund2003conjectured}{article}{ label={Hag03}, author={Haglund, J.}, title={Conjectured statistics for the $q,t$-Catalan numbers}, journal={Adv. Math.}, volume={175}, date={2003}, number={2}, pages={319--334}, issn={0001-8708}, review={\MR {1972636}}, } Close amsref.
[Hag04]
J. Haglund, A proof of the -Schröder conjecture, Internat. Math. Res. Notices 11 (2004), 525–560.
[Hag08]
J. Haglund, The ,-Catalan numbers and the space of diagonal harmonics, University Lecture Series, vol. 41, American Mathematical Society, Providence, RI, 2008. With an appendix on the combinatorics of Macdonald polynomials. MR2371044, Show rawAMSref\bib{haglund2008catalan}{book}{ label={Hag08}, author={Haglund, James}, title={The $q$,$t$-Catalan numbers and the space of diagonal harmonics}, series={University Lecture Series}, volume={41}, note={With an appendix on the combinatorics of Macdonald polynomials}, publisher={American Mathematical Society, Providence, RI}, date={2008}, pages={viii+167}, isbn={978-0-8218-4411-3}, isbn={0-8218-4411-3}, review={\MR {2371044}}, } Close amsref.
[Hai01]
M. Haiman, Hilbert schemes, polygraphs and the Macdonald positivity conjecture, J. Amer. Math. Soc. 14 (2001), no. 4, 941–1006. MR1839919, Show rawAMSref\bib{haiman2001polygraphs}{article}{ label={Hai01}, author={Haiman, Mark}, title={Hilbert schemes, polygraphs and the Macdonald positivity conjecture}, journal={J. Amer. Math. Soc.}, volume={14}, date={2001}, number={4}, pages={941--1006}, issn={0894-0347}, review={\MR {1839919}}, } Close amsref.
[Hai02]
M. Haiman, Vanishing theorems and character formulas for the Hilbert scheme of points in the plane, Invent. Math. 149 (2002), no. 2, 371–407. MR1918676, Show rawAMSref\bib{Hai02}{article}{ label={Hai02}, author={Haiman, Mark}, title={Vanishing theorems and character formulas for the Hilbert scheme of points in the plane}, journal={Invent. Math.}, volume={149}, date={2002}, number={2}, pages={371--407}, issn={0020-9910}, review={\MR {1918676}}, } Close amsref.
[HHL05a]
J. Haglund, M. Haiman, and N. Loehr, A combinatorial formula for Macdonald polynomials, J. Amer. Math. Soc. 18 (2005), no. 3, 735–761. MR2138143, Show rawAMSref\bib{haglund2005combinatorial}{article}{ label={HHL05a}, author={Haglund, J.}, author={Haiman, M.}, author={Loehr, N.}, title={A combinatorial formula for Macdonald polynomials}, journal={J. Amer. Math. Soc.}, volume={18}, date={2005}, number={3}, pages={735--761}, issn={0894-0347}, review={\MR {2138143}}, } Close amsref.
[HHL05b]
J. Haglund, M. Haiman, N. Loehr, J. B. Remmel, and A. Ulyanov, A combinatorial formula for the character of the diagonal coinvariants, Duke Math. J. 126 (2005), no. 2, 195–232. MR2115257, Show rawAMSref\bib{haglund2005diagcoinv}{article}{ label={HHL{${+}$}05b}, author={Haglund, J.}, author={Haiman, M.}, author={Loehr, N.}, author={Remmel, J. B.}, author={Ulyanov, A.}, title={A combinatorial formula for the character of the diagonal coinvariants}, journal={Duke Math. J.}, volume={126}, date={2005}, number={2}, pages={195--232}, issn={0012-7094}, review={\MR {2115257}}, } Close amsref.
[Hic12]
A. S. Hicks, Two parking function bijections: a sharpening of the , -Catalan and Shröder theorems, Int. Math. Res. Not. IMRN 13 (2012), 3064–3088. MR2946232, Show rawAMSref\bib{hicks2012sharpening}{article}{ label={Hic12}, author={Hicks, Angela S.}, title={Two parking function bijections: a sharpening of the $q$, $t$-Catalan and Shr\"oder theorems}, journal={Int. Math. Res. Not. IMRN}, date={2012}, number={13}, pages={3064--3088}, issn={1073-7928}, review={\MR {2946232}}, } Close amsref.
[Hik14]
T. Hikita, Affine Springer fibers of type and combinatorics of diagonal coinvariants, Adv. Math. 263 (2014), 88–122. MR3239135, Show rawAMSref\bib{hikita2014affine}{article}{ label={Hik14}, author={Hikita, Tatsuyuki}, title={Affine Springer fibers of type $A$ and combinatorics of diagonal coinvariants}, journal={Adv. Math.}, volume={263}, date={2014}, pages={88--122}, issn={0001-8708}, review={\MR {3239135}}, } Close amsref.
[HL05]
J. Haglund and N. Loehr, A conjectured combinatorial formula for the Hilbert series for diagonal harmonics, Discrete Math. 298 (2005), no. 1-3, 189–204. MR2163448, Show rawAMSref\bib{haglund2005conjectured}{article}{ label={HL05}, author={Haglund, J.}, author={Loehr, N.}, title={A conjectured combinatorial formula for the Hilbert series for diagonal harmonics}, journal={Discrete Math.}, volume={298}, date={2005}, number={1-3}, pages={189--204}, issn={0012-365X}, review={\MR {2163448}}, } Close amsref.
[HMZ12]
J. Haglund, J. Morse, and M. Zabrocki, A compositional shuffle conjecture specifying touch points of the Dyck path, Canad. J. Math. 64 (2012), no. 4, 822–844. MR2957232, Show rawAMSref\bib{haglund2012compositional}{article}{ label={HMZ12}, author={Haglund, J.}, author={Morse, J.}, author={Zabrocki, M.}, title={A compositional shuffle conjecture specifying touch points of the Dyck path}, journal={Canad. J. Math.}, volume={64}, date={2012}, number={4}, pages={822--844}, issn={0008-414X}, review={\MR {2957232}}, } Close amsref.
[LN14]
N. A. Loehr and E. Niese, New combinatorial formulations of the shuffle conjecture, Adv. in Appl. Math. 55 (2014), 22–47. MR3176715, Show rawAMSref\bib{loehr2014new}{article}{ label={LN14}, author={Loehr, Nicholas A.}, author={Niese, Elizabeth}, title={New combinatorial formulations of the shuffle conjecture}, journal={Adv. in Appl. Math.}, volume={55}, date={2014}, pages={22--47}, issn={0196-8858}, review={\MR {3176715}}, } Close amsref.
[Mac95]
I. G. Macdonald, Symmetric functions and Hall polynomials, 2nd ed., Oxford Mathematical Monographs, The Clarendon Press, Oxford University Press, New York, 1995. With contributions by A. Zelevinsky; Oxford Science Publications. MR1354144, Show rawAMSref\bib{Mac}{book}{ label={Mac95}, author={Macdonald, I. G.}, title={Symmetric functions and Hall polynomials}, series={Oxford Mathematical Monographs}, edition={2}, note={With contributions by A. Zelevinsky; Oxford Science Publications}, publisher={The Clarendon Press, Oxford University Press, New York}, date={1995}, pages={x+475}, isbn={0-19-853489-2}, review={\MR {1354144}}, } Close amsref.
[Neg13]
A. Negut, The shuffle algebra revisited, Int. Math. Res. Not. IMRN 22 (2014), 6242–6275. MR3283004, Show rawAMSref\bib{negut2013shuffle}{article}{ label={Neg13}, author={Negut, Andrei}, title={The shuffle algebra revisited}, journal={Int. Math. Res. Not. IMRN}, date={2014}, number={22}, pages={6242--6275}, issn={1073-7928}, review={\MR {3283004}}, } Close amsref.
[SV11]
O. Schiffmann and E. Vasserot, The elliptic Hall algebra, Cherednik Hecke algebras and Macdonald polynomials, Compos. Math. 147 (2011), no. 1, 188–234. MR2771130, Show rawAMSref\bib{shiff2011elliptic}{article}{ label={SV11}, author={Schiffmann, O.}, author={Vasserot, E.}, title={The elliptic Hall algebra, Cherednik Hecke algebras and Macdonald polynomials}, journal={Compos. Math.}, volume={147}, date={2011}, number={1}, pages={188--234}, issn={0010-437X}, review={\MR {2771130}}, } Close amsref.
[SV13]
O. Schiffmann and E. Vasserot, The elliptic Hall algebra and the -theory of the Hilbert scheme of , Duke Math. J. 162 (2013), no. 2, 279–366. MR3018956, Show rawAMSref\bib{shiff2013elliptic}{article}{ label={SV13}, author={Schiffmann, Olivier}, author={Vasserot, Eric}, title={The elliptic Hall algebra and the $K$-theory of the Hilbert scheme of $\mathbb {A}^2$}, journal={Duke Math. J.}, volume={162}, date={2013}, number={2}, pages={279--366}, issn={0012-7094}, review={\MR {3018956}}, } Close amsref.

Article Information

MSC 2010
Primary: 05E10 (Combinatorial aspects of representation theory)
Secondary: 05E05 (Symmetric functions and generalizations), 05A30 (-calculus and related topics), 33D52 (Basic orthogonal polynomials and functions associated with root systems)
Author Information
Erik Carlsson
International Centre for Theoretical Physics, Str. Costiera, 11, 34151 Trieste, Italy
Address at time of publication: Department of Mathematics, University of California, Davis, 1 Shields Ave., Davis, California 95616
ecarlsson@math.ucdavis.edu
MathSciNet
Anton Mellit
International Centre for Theoretical Physics, Str. Costiera, 11, 34151 Trieste, Italy
Address at time of publication: Faculty of Mathematics, University of Vienna, Oskar-Morgenstern-Platz 1, 1090 Vienna, Austria
anton.mellit@univie.ac.at
MathSciNet
Journal Information
Journal of the American Mathematical Society, Volume 31, Issue 3, ISSN 1088-6834, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , , and published on .
Copyright Information
Copyright 2017 American Mathematical Society
Article References

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.