The sign character of the triagonal fermionic coinvariant ring
Abstract.
We determine the trigraded multiplicity of the sign character of the triagonal fermionic coinvariant ring . As a corollary, this proves a conjecture of Bergeron (2020) that the dimension of the sign character of is . We also give an explicit formula for double hook characters in the diagonal fermionic coinvariant ring , and discuss methods towards calculating the sign character of . Finally, we give a multigraded refinement of a conjecture of Bergeron (2020) that the dimension of the sign character of the -bosonic-fermionic coinvariant ring is , where is a Fibonacci number.
1. Introduction
The diagonal coinvariant ring was introduced by Haiman in 1994 [9], and since then has been studied extensively. Its defining ideal is generated by all polynomials in , with no constant term, which are invariant under the diagonal action of :
(1) |
Haiman found the dimension, bigraded Hilbert series, and bigraded Frobenius series of [10].
There has been much recent interest (see [3, 2, 1, 17]) in studying a more general class of coinvariant rings with sets of commuting (bosonic) variables , , , etc., and sets of anticommuting (fermionic) variables , , , etc. We define the -bosonic-fermionic coinvariant ring by
(2) |
where its defining ideal is generated by all polynomials in , without constant term, which are invariant under the diagonal action of the symmetric group , given by permuting the indices of the variables. Commuting variables commute with all variables. Anticommuting variables anticommute with all anticommuting variables. That is, for all , and mixed products between different sets of fermionic variables likewise anticommute. Note that this implies that .
We recall the definitions of the multigraded Hilbert and Frobenius series of (see for example [2]). For fixed integers , decomposes as a direct sum of multihomogenous components, which are -modules:
(3) |
We denote the multigraded Hilbert series by
(4) |
and the multigraded Frobenius series by
(5) |
where denotes the Frobenius characteristic map and denotes the character. For simplicity, if , we will use for and if , we will use for . Recall that . Furthermore, is a -module (see [2, Section 2]), so its multigraded Frobenius character is a sum of products of three Schur functions, which are irreducible characters of polynomial representations of and , along with a Frobenius character. The main focus of this paper is on the case . We now describe the setting in more detail, from the perspective of both coinvariants and, isomorphically, harmonics.
Now specialize to , which we call the triagonal fermionic coinvariant ring.111Note that is an exterior algebra over the variables . For more on this perspective, see [11]. Its defining ideal is the ideal generated by polynomials in , without constant term, invariant under the diagonal action of . A generating set for the ideal is given by all the nonzero monomial symmetric functions in any of the three sets of variables:
(6) |
Consider the ring in three sets of anticommuting variables, which we arrange into the matrix
(7) |
For each invertible matrix in , multiply on the left by to obtain the product . Then, in any polynomial , replace each variable by the corresponding entry of the matrix . This defines a left -action on .
The ring is naturally trigraded. Denote by the homogeneous component spanned by all monomials that contain exactly variables in , variables in , and variables in . Since acts by linear combinations of the three rows, the -action preserves the total degree . Hence each direct sum is -invariant.
Each permutation corresponds to a permutation matrix , defined so that if and otherwise. Then the right multiplication has the effect on of letting act on the indices of all variables: , , and . Then, in any polynomial , replace each variable by the corresponding entry of the matrix . This defines a right -action on .
Because left multiplication by and right multiplication by commute, these and -actions commute, so is a -module.
Recall the definition of derivative of anticommuting variables (where here each could be from any of the three sets of anticommuting variables) is (see for example [15, Section 1.5]):
(8) |
Define the space of triagonal fermionic harmonics by
(9) |
We also note that since any , we need not check any second derivatives, hence
(10) |
We wish to study as a -module, so we look at the infinitesimal action of the Lie algebra on . Similarly to [9, Section 3.1], we define the operators
(11) |
and
(12) |
Along with two operators given by taking the commutators of the and operators, these generate the Lie algebra .
Furthermore is a -module. Write a trigraded component of as and as . Then we have that as trigraded -modules. Working with harmonics was advocated by Garsia: a benefit of working with harmonics over coinvariants is that one works with polynomials instead of equivalence classes.
Observe that the Schur function is a -analogue of the binomial coefficient ; denote it by
(13) |
Recall the notation . It follows that as a polynomial in with coefficients in and we have
(14) |
and there are two similar expressions for as a polynomial in or in .
A polynomial is called antisymmetric if for all , where acts diagonally by permuting the variables. Let denote the antisymmetric subspace222This is sometimes also referred to as the alternating subspace. of , which consists of all elements which are antisymmetric. This corresponds to the -graded multiplicity of the sign character in , that is, . Our main result is the following.
Theorem 1.1.
The theorem immediately implies the following corollary, which was conjectured by Bergeron [2, Table 3].
Corollary 1.2.
Proof.
Recall that . By evaluating Theorem 1.1 at , the dimension is . ∎
The organization of the paper is as follows. In Section 2, we recall some results of Haglund–Sergel [8] and Kim–Rhoades [11], along with proving some preliminary results. In Section 3, building on the work of Haglund–Sergel and Kim–Rhoades, we give an upper bound on the multiplicity of the sign character of (Corollary 3.2). In Section 4, we construct two elements in the ring of triagonal fermionic harmonics (Proposition 4.4), and study the -representations that they generate (Proposition 4.6), which constructs enough elements in to show that the upper bound on the multiplicity of the sign character is achieved with equality, proving the main theorem.
2. Background and preliminary results
Haglund and Sergel gave a formula for the graded Frobenius series of the fermionic coinvariants .
Lemma 2.1 ([8, Lemma 4.10]).
For ,
Kim and Rhoades gave a formula for the bigraded Frobenius series of the diagonal fermionic coinvariants .
Theorem 2.2 ([11, Theorem 6.1]).
For ,
where denotes the Kronecker product and is interpreted as .
In particular, they found bigraded multiplicities for the trivial, sign, and hook characters.
Proposition 2.3 ([11, Proposition 6.2]).
In , the multiplicity of the trivial character is
the bigraded multiplicity of the sign character is
and for , the bigraded multiplicity of a hook character is
As a consequence of their result on the sign character, we conclude the following.
Corollary 2.4.
-
(1)
If a nonzero harmonic polynomial in is antisymmetric, then it must be of degree exactly .
-
(2)
If a nonzero harmonic polynomial in is antisymmetric, then it must be of degree exactly .
Proof.
The first claim is implied by the second. To show the second, notice that Proposition 2.3 shows that the dimension of each component of the sign character is . ∎
We also will use the following result, which establishes that an antisymmetric polynomial must use enough distinct indices of variables to be nonzero.
Lemma 2.5.
If a polynomial in is antisymmetric, and for the variables that appear, at most unique indices are used, then the polynomial is 0.
Proof.
Without loss of generality, say that the indices used are . Introduce the notation where each letter, when indexed by , is the (ordered) product of any nonempty subset of . We use unique letters, so it is known which subset of is being referred to after applying permutations.
Consider the symmetric group as a set, and partition it into two equal-sized subsets such that for each , we have that , for some , where denotes the simple transposition which interchanges and , and will act on polynomials by first applying and then .
Any antisymmetric polynomial can be obtained by applying an antisymmetrizing operator to some polynomial . Consider each monomial in , say . When and are related by , we have that , since does not contain any variables indexed by or . Hence, by antisymmetrizing the monomial , we get that
where the last line follows since is a simple transposition, so and have opposite signs. ∎
3. An upper bound on degree
The following result is inspired by a similar result of Haglund–Sergel (on a different ring, ) [8, Theorem 4.11].
Proposition 3.1.
For ,
Proof.
As noted in [8, equation (4.9)], following from [4], for any -modules and tensored under the diagonal action of , we have that
where are the Kronecker coefficients. In our case, we are interested in
(15) |
since . By Lemma 2.1, which states that only hook Schur functions appear in , we reduce to
(16) |
where we applied Proposition 2.3. After a bit of algebra, we obtain the claimed formula. ∎
Now the following result gives us an upper bound on the occurrences of the sign character in . For multivariate polynomials and , the notation means that is a sum of monomials with only nonnegative coefficients.
Corollary 3.2.
For ,
Proof.
First note that any occurrence of the sign character in is predicated on it occurring in , since is a quotient of under the diagonal action of (see [8]).
By permuting which sets of variables out of are assigned to and to in Proposition 3.1, we conclude that a sign character in must have trigraded multiplicity bounded above by
(17) |
(18) |
and
(19) |
Note that if a polynomial satisfies , , and , then . Hence by taking the bounds given by equations (17-19) together, we conclude that the sign character in must have trigraded multiplicity bounded above by
(20) |
∎
4. Construction of basis elements
Now we will work towards the proof of the main theorem, by constructing two explicit elements in , which are highest weight vectors for certain -representations. We ultimately show that the upper bound given in Corollary 3.2 is obtained with equality.
Definition 4.1.
For , define the primary theta-seed by
Definition 4.2.
For , define the secondary theta-seed by
When or in the definitions of the primary and secondary theta-seed, respectively, the empty product of ’s is interpreted as .
We now prove a technical lemma.
Lemma 4.3.
The antisymmetrization operator commutes with the differential operator .
Proof.
Let . Observe that for any ,
(21) |
since addition is commutative. Thus,
(22) |
so the operators commute as claimed. ∎
Now we show that and are harmonic.
Proposition 4.4.
-
(1)
The primary theta-seed is in .
-
(2)
The secondary theta-seed is in .
Proof.
We show (1). Consider the primary theta-seed
As discussed in the introduction, we only need to show that for Since only variables appear in , it only remains to check . Consider , which is what we will antisymmetrize to get the primary theta-seed. By Lemma 4.3, the differential operator commutes with the antisymmetrization operator. We write
(23) |
where each monomial on the right hand side is in only variables. Upon antisymmetrization, by Lemma 2.5, this becomes .
We show (2). For , consider the secondary theta-seed
We must check that is harmonic. We only need to show that for the following seven choices of :
Case 1: . Let , which is what we will antisymmetrize to get the secondary theta-seed. For concision, write for the ordered product . We write that
(24) |
Then upon antisymmetrization, both of the terms and will be due to Lemma 2.5, since they only contain indices for variables.
For the term , there are indices for variables, so we cannot apply Lemma 2.5 directly. However, we can apply a similar argument as in the proof of Lemma 2.5. Partition into two equal-sized subsets such that for each , we have that , where denotes the simple transposition which interchanges 1 and 2. Upon symmetrization, we write
(25) |
Case 2: . Before antisymmetrization, we write that
Observe that becomes 0 upon antisymmetrization, by Lemma 2.5. Additionally, becomes 0 upon antisymmetrization by the same argument used in Case 1.
Case 3: . The same argument as in Case 2 applies.
Case 4: . Before antisymmetrization, we write that
which becomes 0 upon antisymmetrization by Lemma 2.5.
Case 5: . The same argument as in Case 4 applies.
Case 6: . The same argument as in Case 4 applies.
Case 7: . Before antisymmetrization, we write that
which remains 0 upon antisymmetrization.
This completes the proof that the secondary theta-seed is harmonic. ∎
A priori, a harmonic polynomial in could equal zero if certain term cancellations occur. However, the following result demonstrates that this does not happen for and .
Proposition 4.5.
-
(1)
The primary theta-seed is nonzero in .
-
(2)
The secondary theta-seed is nonzero in .
Proof.
We show (1). Consider in ,
When the antisymmetrization operator is applied to , consider which will output the monomial . Such a must fix , but can be permuted. For any in the symmetric group on letters , we have that (see for example [5, Chapter III, Section 7.3, Proposition 5]). Hence antisymmetrization produces copies of , so is nonzero in , as desired.
Next we show (2). Consider in ,
When the operator is applied to , consider which will include the monomial in its output. Since the and have the same index, it must come from an antisymmetrization of (and not of nor ). Thus such a must fix and , but can be permuted. For any in the symmetric group on letters , it follows that . This implies that . Hence antisymmetrization produces copies of , so is nonzero in , as desired. ∎
Starting with and , the operators defined in equation (11) and create -representations.
Proposition 4.6.
-
(1)
The primary theta-seed is the highest weight vector for the -representation with character .
-
(2)
The secondary theta-seed is the highest weight vector for the -representation with character .
Proof.
We show (1). Start with the primary theta-seed
Since we constructed it using an antisymmetrizing operator, it is antisymmetric. By Proposition 4.4, it is in , and by Proposition 4.5, it is nonzero. Observe that and both kill . Thus is a highest weight vector for the irreducible -representation with highest weight , where the weight of any is given by . Thus the -character of this representation is .
We show (2). Consider the secondary theta-seed
Note that this is antisymmetric, as it is constructed using an antisymmetrization operator. By Corollary 2.4, there are no antisymmetric elements of degree solely in any one or two sets of variables. This implies that and both kill . By Proposition 4.4, it is in , and by Proposition 4.5, it is nonzero. Thus is a highest weight vector for the irreducible -representation with highest weight . Thus the -character of this representation is . ∎
Now we can prove the main theorem.
Theorem 1.1.
Proof.
Remark 4.7.
While it is not necessarily true in general that the Lie algebra operators and will correspond to crystal operators, in the present case, it is true since every weight space for the representations studied in Proposition 4.6 is one-dimensional. The representation with highest weight has crystal structure isomorphic to a crystal of tableaux and the representation with highest weight has crystal structure isomorphic to a crystal of tableaux (see for example [6, Chapter 3]).
5. Double hook characters of the diagonal fermionic coinvariant ring
With Proposition 2.3, Kim and Rhoades gave explicit formulas for the trivial, sign, and hook characters of . In this section, we extend the analysis to give an explicit formula for double hook characters of . All characters of indexed by shapes not contained in a double hook are .
Rosas gave a combinatorial formula for the Kronecker coefficients , for any shape . These Kronecker coefficients are only nonzero if is contained in a double hook shape. Here, we recall her formulas for double hook shapes, rows, and hook shapes.
Theorem 5.1 ([14, Theorem 3]).
Let and be hook shapes.
-
(1)
If is not contained in a double hook, then .
-
(2)
Let where be a double hook. Then
where is if the proposition is true and if it is false.
-
(3)
If , then .
-
(4)
If is a hook shape , then
Note that this implies that these coefficients are in [14, Corollary 4]. We are ready to prove the following result, establishing a formula for double hook characters in .
Theorem 5.2.
Let be a partition contained in a double hook shape, which is not itself contained within a hook shape. That is, where . Let . Then if , we have that
(26) |
If , we have that
(27) |
Proof.
Recall from Theorem 2.2 that
Let where . Hence,
(28) |
By Theorem 5.1,
(29) |
Define . We analyze a component of equation (5):
(30) |
which is equivalent to
(31) |
Since the bigraded component whenever , we exclude these cases from our analysis. Thus is satisfied if or (if ), if (if ), if (if ), etc. On the other hand, is satisfied if (if ), if (if ), etc. Putting these together, we get that equation (31) is 1 when and 0 otherwise.
We analyze when condition is satisfied. At , the condition is true exactly for integers which satisfy
(32) |
Summing up over such gives
(33) |
At , the condition is true exactly for integers which satisfy
(34) |
Summing up over such gives
(35) |
Next we analyze another component of equation (5):
(36) |
which is equivalent to
(37) |
We have that is satisfied if (if ), if (if ), if (if ), etc. On the other hand, is satisfied if (if ), if (if ), etc. Putting these together, when , we get that equation (37) is 1 when and 0 otherwise. When , we get that equation (37) is 1 when and 0 otherwise.
We analyze when condition is satisfied. At , the condition is true exactly for integers which satisfy
(38) |
Summing up over such gives
(39) |
When , at , the condition is true exactly for integers which satisfy
(40) |
Summing up over such gives
(41) |
For concision, we use . When , i.e., , summing equations (33), (35), and (39) proves equation (26). When , i.e., , summing equations (33), (35), (39), and (41) proves equation (27).
∎
6. Four sets of fermions
A natural further question is to determine for , of which the simplest next case is . If the hook characters in are all known, then a similar process as in Proposition 3.1 using can be attempted. Using Theorem 5.1, we can first bound .
Example 6.1.
Let . Suppose we want to calculate . Consider that any in must appear in . We compute that
(42) |
Using Rosas’ formula for Kronecker coefficients of two hooks, we determine that the only pairs of and which do not have multiplicity of are ; ; ; ; and , which all have multiplicity of . Using Proposition 2.3 and Lemma 2.1, we obtain
(43) |
By permuting which sets of variables are assigned to and , we can similarly obtain
(44) |
(45) |
Summing the monomials which appear in all three equations, we obtain
(46) |
Now a computer calculation finds that
(47) |
which shows that the bound is not tight in this case due to the monomial . It will require additional work to determine how to cut out the extra monomials in general.
Example 6.2.
Let . Suppose we know all of the hook characters in ; in this case . Consider that any sign character in must appear in . We compute that
(48) |
using Lemma 2.1. This simplifies to
(49) |
which is unchanged under changing which sets of variables are assigned to and . In this example, equation (49) is equal to determined by computer calculation.
Example 6.3.
The previous example also gives tight bounds for , with graded multiplicity , and for , with graded multiplicity .
Another approach is to use to bound directly. As Kim and Rhoades note, only Schur functions of shapes contained within double hooks occur in , so the shapes which index sign characters are limited.
Example 6.4.
Let . Consider that any sign character in must appear in . We compute that
(50) |
In this case, there are only three partitions of : . Using Proposition 2.3, we get
(51) |
(52) |
(53) |
Putting these together, we obtain
(54) |
By changing which sets of variables are assigned to each , we can similarly obtain
(55) |
(56) |
In this case, all three are equal to each other, and expand to give
(57) |
Now a computer calculation finds that
(58) |
which shows that the bound is not tight in this case due to the monomial . It will require additional work to determine how to cut out the extra monomials in general.
7. One set of bosons and three sets of fermions
In this section, we study the sign character in , the -bosonic-fermionic coinvariant ring. We first recall the “Theta conjecture” of D’Adderio, Iraci, and Vanden Wyngaerd, which expresses the multigraded Frobenius series of in terms of certain Theta operators and the nabla operator.
Conjecture 7.1 ([7, Conjecture 8.2]).
For all ,
The Theta conjecture specialized at is the following.
Conjecture 7.2 (D’Adderio, Iraci, and Vanden Wyngaerd).
For all ,
We recall the following result on the conjectural hook characters in .
Theorem 7.3 ([12, Theorem 8.4]).
If the Theta conjecture specialized at (Conjecture 7.2) is true, then
Now we compute an example.
Example 7.4.
Let . Consider that any sign character in must appear in . We compute that
(59) |
In this example, this is equal to determined by computer calculation.
Proposition 7.5.
If the Theta conjecture specialized at (Conjecture 7.2) is true, then we have the following upper-bound:
Proof.
Based on data for , we propose the following conjecture.
Conjecture 7.6.
Notice that upon specializing all to , the conjecture becomes
(61) |
Let be the th Fibonacci number, defined by the initial conditions , , and recurrence for all . Inspired by a conjecture of Zabrocki on [16], Bergeron made a general conjecture on how the Frobenius series of bosonic-fermionic coinvariant rings could be derived from the Frobenius series of purely bosonic coinvariant rings [2, Conjecture 1]. While collecting computational evidence, Bergeron conjectured the following.
Conjecture 7.7 ([2, Table 3]).
For all ,
Beginning at , the sequence is (see [13, Sequence A001076]). We show the following result connecting the two proposed enumerations.
Proposition 7.8.
For ,
Proof.
Shifting indices, the claim is equivalent to
(62) |
for all . From [13, Sequence A001076] we have that for all , , where is defined by , , and for . Thus the claim is equivalent to
(63) |
for all .
First, we check that the left hand side of equation (63) satisfies the initial conditions and . For , the sum is . For , the sum is .
Remark 7.9.
Acknowledgements
The author would like to thank François Bergeron, Sylvie Corteel, Nicolle González, and Mark Haiman for helpful conversations. The author was partially supported by the National Science Foundation Graduate Research Fellowship DGE-2146752.
References
- [1] François Bergeron, Jim Haglund, Alessandro Iraci, and Marino Romero, Bosonic-fermionic diagonal coinvariants and theta operators, preprint (2023), https://www2.math.upenn.edu/~jhaglund/preprints/BF2.pdf.
- [2] François Bergeron, The bosonic-fermionic diagonal coinvariant modules conjecture, preprint (2020), https://arxiv.org/abs/2005.00924.
- [3] by same author, -modules of multivariate diagonal harmonics, Open Problems in Algebraic Combinatorics (Christine Berkesch, Benjamin Brubaker, Gregg Musiker, Pavlo Pylyavskyy, and Victor Reiner, eds.), Proc. Symp. Pure Math., vol. 110, Providence, RI: American Mathematical Society (AMS), 2024, pp. 1–22.
- [4] Christine Bessenrodt, Tensor products of representations of the symmetric groups and related groups, RIMS Kokyuroku (Proceedings of Research Insititute for Mathematical Sciences) 1149 (2000), 1–15.
- [5] Nicolas Bourbaki, Algebra I, Elements of mathematics, Springer-Verlag, Berlin, 1989 (English).
- [6] Daniel Bump and Anne Schilling, Crystal bases. Representations and combinatorics, Hackensack, NJ: World Scientific, 2017 (English).
- [7] Michele D’Adderio, Alessandro Iraci, and Anna Vanden Wyngaerd, Theta operators, refined delta conjectures, and coinvariants, Adv. Math. 376 (2021), 60, Id/No 107447.
- [8] James Haglund and Emily Sergel, Schedules and the delta conjecture, Annals of Combinatorics 25 (2020), no. 1, 1?31.
- [9] Mark Haiman, Conjectures on the quotient ring by diagonal invariants, J. Algebr. Comb. 3 (1994), no. 1, 17–76.
- [10] by same author, Vanishing theorems and character formulas for the Hilbert scheme of points in the plane, Invent. Math. 149 (2002), no. 2, 371–407.
- [11] Jongwon Kim and Brendon Rhoades, Lefschetz theory for exterior algebras and fermionic diagonal coinvariants, Int. Math. Res. Not. 2022 (2022), no. 4, 2906–2933.
- [12] John Lentfer, A conjectural basis for the -bosonic-fermionic coinvariant ring, preprint (2024), https://arxiv.org/abs/2406.19715.
- [13] OEIS Foundation Inc. (2024), The On-Line Encyclopedia of Integer Sequences, Published electronically at https://oeis.org.
- [14] Mercedes H. Rosas, The Kronecker product of Schur functions indexed by two-row shapes or hook shapes, Journal of Algebraic Combinatorics 14 (2001), no. 2, 153–173.
- [15] Joshua P. Swanson and Nolan R. Wallach, Harmonic differential forms for pseudo-reflection groups II. Bi-degree bounds, Combinatorial Theory 3 (2023), no. 3, Paper no. 17.
- [16] Mike Zabrocki, A module for the Delta conjecture, preprint (2019), https://arxiv.org/abs/1902.08966.
- [17] by same author, Coinvariants and harmonics, 2020, https://realopacblog.wordpress.com/2020/01/26/coinvariants-and-harmonics/.