Abstract

We develop the theory of poset pinball, a combinatorial game introduced by Harada-Tymoczko to study the equivariant cohomology ring of a GKM-compatible subspace 𝑋 of a GKM space; Harada and Tymoczko also prove that, in certain circumstances, a successful outcome of Betti poset pinball yields a module basis for the equivariant cohomology ring of 𝑋. First we define the dimension pair algorithm, which yields a successful outcome of Betti poset pinball for any type 𝐴 regular nilpotent Hessenberg and any type 𝐴 nilpotent Springer variety, considered as GKM-compatible subspaces of the flag variety. The algorithm is motivated by a correspondence between Hessenberg affine cells and certain Schubert polynomials which we learned from Insko. Second, in a special case of regular nilpotent Hessenberg varieties, we prove that our pinball outcome is poset-upper-triangular, and hence the corresponding classes form a 𝐻𝑆1(pt)-module basis for the S1-equivariant cohomology ring of the Hessenberg variety.

1. Introduction

The purpose of this paper is to further develop the theory of poset pinball, a combinatorial game introduced in [1] for the purpose of computing in equivariant cohomology rings (all cohomology rings in this note are with coefficients), in certain cases of type 𝐴 nilpotent Hessenberg varieties. One of the main uses of poset pinball in [1] is to construct module bases for the equivariant cohomology rings of GKM-compatible subspaces of GKM spaces [1, Definition  4.5]. In the context of this paper, the ambient GKM space is the flag variety 𝑎𝑔𝑠(𝑛) equipped with the action of the diagonal subgroup 𝑇 of 𝑈(𝑛,), and the GKM-compatible subspaces are the nilpotent Hessenberg varieties. It is well recorded in the literature (e.g., [2] and references therein) that GKM spaces often have geometrically and/or combinatorially natural module bases for their equivariant cohomology rings; the basis of equivariant Schubert classes {𝜎𝑤}𝑤𝑆𝑛 for 𝐻𝑇(𝑎𝑔𝑠(𝑛)) is a famous example. The results of this paper represent first steps towards the larger goal of using poset pinball to construct a similarly computationally effective and convenient module bases for a GKM-compatible subspace by exploiting the structure of the ambient GKM space.

We briefly recall the setting of our results. Let 𝑁𝑛𝑛 be a nilpotent operator. Let {1,2,,𝑛}{1,2,,𝑛} be a function satisfying (𝑖)𝑖 for all 1𝑖𝑛 and (𝑖+1)(𝑖) for all 1𝑖<𝑛. The associated Hessenberg variety Hess(𝑁,) is then defined as the following subvariety of 𝑎𝑔𝑠(𝑛):Hess𝑉(𝑁,)==0𝑉1𝑉2𝑉𝑛1𝑉𝑛=𝑛𝑁𝑉𝑖𝑉(𝑖).𝑖=1,,𝑛(1.1) Since we deal exclusively with type 𝐴 in this paper, henceforth we omit this phrase from our terminology. Two special cases of Hessenberg varieties are of particular interest in this paper: when 𝑁 is the principal nilpotent operator (in this case Hess(𝑁,) is called a regular nilpotent Hessenberg variety) and when is the identity function (𝑖)=𝑖 for all 1𝑖𝑛 (in this case Hess(𝑁,) is called a nilpotent Springer variety and is sometimes denoted 𝒮𝑁). Hessenberg varieties arise in many areas of mathematics, including geometric representation theory [35], numerical analysis [6], mathematical physics [7, 8], combinatorics [9], and algebraic geometry [10, 11], so it is of interest to explicitly analyze their topology, for example, the structure of their (equivariant) cohomology rings. We do so through poset pinball and Schubert calculus techniques, as initiated and developed in [1, 12, 13] and briefly recalled below.

The following relationship between two group actions on the nilpotent Hessenberg variety and the flag variety, respectively, allows us to use the theory of GKM-compatible subspaces and poset pinball. There is a natural 𝑆1 subgroup of the unitary diagonal matrices 𝑇 which acts on Hess(𝑁,) (defined precisely in Section 2). The group 𝑇, the maximal torus of 𝑈(𝑛,), acts on 𝑎𝑔𝑠(𝑛) in the standard fashion. It turns out that the 𝑆1-fixed points Hess(𝑁,)𝑆1 are a subset of the 𝑇-fixed points 𝑎𝑔𝑠(𝑛)𝑇𝑆𝑛. Moreover, the inclusion of Hess(𝑁,) into 𝑎𝑔𝑠(𝑛) and the inclusion of groups 𝑆1 into 𝑇 then induce a natural ring homomorphism𝐻𝑇(𝑎𝑔𝑠(𝑛))𝐻𝑆1(Hess(𝑁,)).(1.2) As mentioned above, it is well known in Schubert calculus that the equivariant Schubert classes {𝜎𝑤}𝑤𝑆𝑛 are a computationally convenient 𝐻𝑇(pt)-module basis for 𝐻𝑇(𝑎𝑔𝑠(𝑛)). We refer to the images in 𝐻𝑆1(Hess(𝑁,)) of the equivariant Schubert classes {𝜎𝑤}𝑤𝑆𝑛 via the projection (1.2) as Hessenberg Schubert classes. Given this setup and following [1], the game of poset pinball uses the data of the fixed points 𝑎𝑔𝑠(𝑛)𝑇𝑆𝑛 (considered as a partially ordered set with respect to Bruhat order) and the subsetHess(𝑁,)𝑆1𝑎𝑔𝑠(𝑛)𝑇𝑆𝑛(1.3) to determine a set of rolldowns in 𝑆𝑛. It is shown in [1] that, under certain circumstances (one of which is discussed in more detail below), such a set of rolldowns in turn specifies a subset of the Hessenberg Schubert classes which form a 𝐻𝑆1(pt)-module basis of 𝐻𝑆1(Hess(𝑁,)). Thus poset pinball is an important tool for building computationally effective module bases for the equivariant cohomology of Hessenberg varieties. Indeed, the results of [13] accomplish precisely this goal—that is, of constructing a module basis via poset pinball techniques—in the special case of the Peterson variety, which is the regular nilpotent Hessenberg variety with Hessenberg function defined by (𝑖)=𝑖+1 for 1𝑖𝑛1 and (𝑛)=𝑛. Exploiting this explicit module basis, in [13, Theorem  6.12] the second author and Tymoczko give a manifestly positive Monk formula for the product of a degree-2 Peterson Schubert class with an arbitrary Peterson Schubert class, expressed as a 𝐻𝑆1(pt)-linear combination of Peterson Schubert classes. This is an example of equivariant Schubert calculus in the realm of Hessenberg varieties, and it is an open problem to generalize the results of [13] to a wider class of Hessenberg varieties.

We now describe our main results. First, we explain in detail an algorithm which we dub the dimension pair algorithm and which associates to each 𝑆1-fixed point 𝑤Hess(𝑁,)𝑆1 a permutation in 𝑆𝑛, which we call the rolldown of 𝑤 following terminology in [1] and denoted 𝑟𝑜(𝑤)𝑆𝑛. In the special cases of regular nilpotent Hessenberg varieties and nilpotent Springer varieties, we show that the set {𝑟𝑜(𝑤)}𝑤Hess(𝑁,)𝑆1 can be interpreted as the result of a successful game of Betti pinball (in the sense of [1]). The main motivation for our construction is that a successful outcome of Betti pinball can, under some circumstances, produce a module basis for the associated equivariant cohomology ring (cf. [1, Section  4.3]). In this sense, our algorithm represents a significant step towards the construction of module bases for the equivariant cohomology rings of general nilpotent Hessenberg varieties, thus extending the theory developed in [1, 13]. Although we formulate our algorithm in terms of dimension pairs and permissible fillings following terminology of Mbirika [14], the essential idea comes from a correspondence between Hessenberg affine cells and certain Schubert polynomials which we learned from Erik Insko.

Second, for a specific case of a regular nilpotent Hessenberg variety which we call a 334-type Hessenberg variety, we prove that the set of rolldowns {𝑟𝑜(𝑤)}𝑤Hess(𝑁,)𝑆1 obtained from the dimension pair algorithm is in fact poset-upper-triangular in the sense of [1]. As shown in [1], this is one of the possible circumstances under which we can conclude that the corresponding set of Hessenberg Schubert classes forms a module basis for the 𝑆1-equivariant cohomology ring of the variety. Thus our result gives rise to a new family of examples of Hessenberg varieties (and GKM-compatible subspaces) for which poset pinball successfully produces explicit module bases. We mention that the dimension pair algorithm also produces module bases in a special case of Springer varieties [15]. Although we do not know whether the dimension pair algorithm always succeeds in producing module bases for the 𝑆1-equivariant cohomology rings for a general nilpotent Hessenberg variety, the evidence thus far is suggestive. We leave further investigation to future work.

We give a brief summary of the contents of this manuscript. In Section 2 we recall some definitions and constructions necessary for later statements. In Section 3.1 we describe the dimension pair algorithm and prove that the result of the algorithm satisfies the conditions to be the outcome of a successful game of Betti poset pinball in the special cases of regular nilpotent Hessenberg varieties and nilpotent Springer varieties. We briefly review in Section 3.2 the theory developed in [1] which show that if the rolldown set obtained from a successful game of Betti poset pinball also satisfies poset-upper-triangularity conditions, then it yields a module basis in equivariant cohomology. In Sections 4 and 5 we prove that the dimension pair algorithm produces a poset-upper-triangular module basis in a special class of regular nilpotent Hessenberg varieties which we call 334-type Hessenberg varieties. We close with some open questions in Section 6.

2. Background

We begin with necessary definitions and terminology for what follows. In Section 2.1 we recall the geometric objects and the group actions under consideration. In Section 2.2 we recall some combinatorial definitions associated to Young diagrams. We recall a bijection between Hessenberg fixed points and certain fillings of Young diagrams in Section 2.3. The discussion closely follows previous work (e.g., [1, 13] and also [16]) so we keep exposition brief.

2.1. Hessenberg Varieties, Highest Forms, and Fixed Points

By the flag variety we mean the homogeneous space 𝐺𝐿(𝑛,)/𝐵 which is also identified with𝑎𝑔𝑠(𝑛𝑉)=={0}𝑉1𝑉2𝑉𝑛1𝑉𝑛=𝑛dim𝑉𝑖.=𝑖(2.1) A Hessenberg function is a function {1,2,,𝑛}{1,2,,𝑛} satisfying (𝑖)𝑖 for all 1𝑖𝑛 and (𝑖+1)(𝑖) for all 1𝑖<𝑛. We frequently denote a Hessenberg function by listing its values in sequence, =((1),(2),,(𝑛)=𝑛). Let 𝑁𝑛𝑛 be a linear operator. The Hessenberg variety Hess(𝑁,) is defined as the following subvariety of 𝑎𝑔𝑠(𝑛):Hess𝑉(𝑁,)=𝑎𝑔𝑠(𝑛)𝑁𝑉𝑖𝑉(𝑖)𝑖=1,,𝑛𝑎𝑔𝑠(𝑛).(2.2) If 𝑁 is nilpotent, we say Hess(𝑁,) is a nilpotent Hessenberg variety, and if 𝑁 is the principal nilpotent operator (i.e., has one Jordan block with eigenvalue 0), then Hess(𝑁,) is called a regular nilpotent Hessenberg variety. If 𝑁 is nilpotent and is the identity function (𝑖)=𝑖 for all 1𝑖𝑛, then Hess(𝑁,) is called a nilpotent Springer variety and often denoted 𝒮𝑁. In this manuscript we study in some detail the regular nilpotent case, and as such sometimes notate Hess(𝑁,) as Hess() when 𝑁 is understood to be the standard principal nilpotent operator.

Suppose 𝑁 is a nilpotent matrix in standard Jordan canonical form. It turns out that for many of our statements below we must use a choice of conjugate of 𝑁 which is in highest form [16, Definition  4.2]. We recall the following.

Definition 2.1 (see [16, Definitions 4.1 and 4.2]). (i) Let 𝑋 be any 𝑚×𝑛 matrix. We call the entry 𝑋𝑖𝑘 a pivot of 𝑋 if 𝑋𝑖𝑘 is nonzero and if all entries below and to its left vanish, that is, 𝑋𝑖𝑗=0 if 𝑗<𝑘 and 𝑋𝑗𝑘=0 if 𝑗>𝑖. Moreover, given 𝑖, define 𝑟𝑖 to be the row of 𝑋𝑟𝑖,𝑖 if the entry is a pivot, and 0 otherwise.
(ii) Let 𝑁 be an upper-triangular nilpotent 𝑛×𝑛 matrix. Then we say 𝑁 is in highest form if its pivots form a nondecreasing sequence, namely, 𝑟1𝑟2𝑟𝑛.

We do not require the details of the theory of highest forms of linear operators; for the purposes of the present manuscript it suffices to remark firstly that when 𝑁 is the principal nilpotent matrix, then 𝑁 is already in highest form, and secondly that any nilpotent matrix can be conjugated by an appropriate 𝑛×𝑛 permutation matrix 𝜎 so that 𝑁𝑓=𝜎𝑁𝜎1 is in highest form. However, the following observation will be relevant in Section 2.3.

Remark 2.2. In this manuscript we always assume that our highest form 𝑁𝑓=𝜎𝑁𝜎1 has been chosen in accordance to the recipe described by Tymoczko in [16, Section 4]. Since the precise method of this construction is not relevant for the rest of the present manuscript, we omit further explanation here. In the case when 𝑁 is principal nilpotent, we take 𝑁𝑓=𝑁 since 𝑁 is already in highest form and this is the form chosen by Tymoczko in [16]. A more detailed discussion of highest forms as it pertains to poset pinball theory is in [15].

For details on the following facts we refer the reader to, for example, [1, 13, 16] and references therein. Let 𝑁 be an 𝑛×𝑛 nilpotent matrix in Jordan canonical form and let 𝜎 denote a permutation matrix such that 𝑁𝑓=𝜎𝑁𝜎1 is in highest form. It is known and straightforward to show that the following 𝑆1 subgroup of 𝑈(𝑛,) preserves Hess(𝑁,) for 𝑁 as above and any Hessenberg function :𝑆1=𝑡𝑛000𝑡𝑛1000000𝑡𝑡,𝑡=1𝑇𝑛𝑈(𝑛,).(2.3) Here 𝑇𝑛 is the standard maximal torus of 𝑈(𝑛,) consisting of diagonal unitary matrices.

This implies that the conjugate circle subgroup 𝜎𝑆1𝜎1 preserves Hess(𝑁𝑓,). By abuse of notation we will denote both circle subgroups by 𝑆1, since it is clear by context which is meant. The 𝑆1-fixed points of Hess(𝑁,) and Hess(𝑁𝑓,) are isolated and are a subset of the 𝑇𝑛-fixed points of 𝑎𝑔𝑠(𝑛). Since the set of 𝑇𝑛-fixed points 𝑎𝑔𝑠(𝑛)𝑇𝑛 may be identified with the Weyl group 𝑊=𝑆𝑛, and since Hess(𝑁,)𝑆1 (resp., Hess(𝑁𝑓,)𝑆1) is a subset of 𝑎𝑔𝑠(𝑛)𝑇𝑛, any Hessenberg fixed point may be thought of as a permutation 𝑤𝑆𝑛.

2.2. Permissible Fillings, Dimension Pairs, Lists of Top Parts, and Associated Permutations

Recall that there is a bijective correspondence between the set of conjugacy classes of nilpotent 𝑛×𝑛 complex matrices 𝑁 and Young diagrams (we use English notation for Young diagrams) with 𝑛 boxes, given by associating to 𝑁 the Young diagram 𝜆 with row lengths the sizes of the Jordan blocks of 𝑁 listed in weakly decreasing order. We will use this bijection to often treat such 𝑁 and 𝜆 as the same data; we sometimes denote by 𝜆𝑁 the Young diagram given as above corresponding to a nilpotent 𝑁.

For more details on the following see [14].

Definition 2.3. Let 𝜆 be a Young diagram with 𝑛 boxes. Let {1,2,,𝑛}{1,2,,𝑛} be a Hessenberg function. A filling of 𝜆 by the alphabet {1,2,,𝑛} is an injective placing of the integers {1,2,,𝑛} into the boxes of 𝜆. A filling of 𝜆 is called a (𝐡,𝜆)-permissible filling if for every horizontal adjacency

254235.fig.001
in the filling we have 𝑘(𝑗).

Remark 2.4. In this manuscript the 𝜆 and will frequently be understood by context. When there is no danger of confusion we simply refer to permissible fillings.

Example 2.5. Let 𝑛=5. Suppose 𝜆=(5) and =(3,3,4,5,5). Then

254235.fig.002
is a permissible filling, whereas
254235.fig.003
is not, since 4(1).
We denote a permissible filling of 𝜆 by 𝑇, in analogy with standard notation for Young tableaux. Next we focus attention on certain pairs of entries in a permissible filling 𝑇.

Definition 2.6. Let {1,2,,𝑛}{1,2,,𝑛} be a Hessenberg function and 𝜆 a Young diagram with 𝑛 boxes. A pair (𝑎,𝑏) is a dimension pair of an (,𝜆)-permissible filling 𝑇 of 𝜆 if the following conditions hold:(1)𝑏>𝑎, (2)𝑏 is either(i)below 𝑎 in the same column of 𝑎, or(ii)anywhere in a column strictly to the left of the column of 𝑎, and(3)if there exists a box with filling 𝑐 directly adjacent to the right of 𝑎, then 𝑏(𝑐).

For a dimension pair (𝑎,𝑏) of 𝑇, we will refer to 𝑏 as the top part of the dimension pair.

Example 2.7. Let 𝜆, be as in Example 2.5. The dimension pairs in the permissible filling

254235.fig.004
are (1,2), (1,3), and (1,4). Note that (3,4) is not a dimension pair because 1 is directly to the right of the 3 and 4(1).

Given a permissible filling 𝑇 of 𝜆, we follow [14] and denote by 𝐷𝑃𝑇 the set of dimension pairs of 𝑇. For each integer with 2𝑛, let 𝑥||=(𝑎,)(𝑎,)𝐷𝑃𝑇||(2.4) so 𝑥 is the number of times occurs as a top part in the set of dimension pairs of 𝑇. From the definitions it follows that 0𝑥1forall2𝑛. We call the integral vector 𝐱=(𝑥2,𝑥3,,𝑥𝑛) the list of top parts of 𝑇.

To each such 𝐱 we associate a permutation in 𝑆𝑛 as follows. As a preliminary step, for each with 2𝑛 define𝑢(𝑠𝐱)=1𝑠2𝑠𝑥,if𝑥>0,1,if𝑥=0,(2.5) where 𝑠𝑖 denotes the simple transposition (𝑖,𝑖+1) in 𝑆𝑛 and 1 denotes the identity permutation. Now define the association𝐱𝜔(𝐱)=𝑢2(𝐱)𝑢3(𝐱)𝑢𝑛(𝐱)𝑆𝑛.(2.6) It is not difficult to see that (2.6) is a bijection between the set of integral vectors 𝐱𝑛1 satisfying 0𝑥1 for all 2𝑛1 and the group 𝑆𝑛. In fact the word given by (2.6) is a reduced word decomposition of 𝜔(𝐱) and the 𝑥 count the number of inversions in 𝜔(𝐱) with as the higher integer. The following simple fact will be used later.

Fact 1. Suppose 𝐱=(𝑥2,,𝑥𝑛), 𝐲=(𝑦2,,𝑦𝑛)𝑛10 are both lists of top parts. Suppose further that for all 2𝑛, we have 𝑥𝑦. Then 𝜔(𝐱)𝜔(𝐲) in Bruhat order. This follows immediately from the definition  (2.6).

Example 2.8. Continuing with Examples 2.5 and 2.7, for the permissible filling

254235.fig.005
, the set 𝐷𝑃𝑇 of top parts of dimension pairs is {2,3,4}, yielding the integer vector 𝐱=(1,1,1,0). The associated permutation 𝜔(𝐱) is then 𝑠1𝑠2𝑠3.

Example 2.9. Let 𝜆, be as in Example 2.5. The filling

254235.fig.006
is also permissible, with dimension pairs (1,2), (1,3), (1,4), (2,3). Hence 𝐱=(1,2,1,0) and the associated permutation 𝜔(𝐱) is 𝑠1(𝑠2𝑠1)𝑠3.

2.3. Bijection between Fixed Points and Permissible Fillings

For nilpotent Hessenberg varieties, the 𝑆1-fixed points Hess(𝑁,)𝑆1 are in bijective correspondence with the set of permissible fillings of the Young diagram 𝜆=𝜆𝑁, as we now describe. We will use this correspondence in the formulation of our dimension pair algorithm.

Suppose 𝜆 is a Young diagram with 𝑛 boxes. We begin by defining a bijective correspondence between the set 𝑖(𝜆) of all fillings (not necessarily permissible) of 𝜆 with permutations in 𝑆𝑛. Given a filling, read the entries of the filling by reading along each column from the bottom to the top, starting with the leftmost column and proceeding to the rightmost column. The association 𝑖(𝜆)𝑆𝑛 is then given by interpreting the resulting word as the one-line notation of a permutation. For example, the filling254235.fig.007(2.7) has associated permutation 641523. It is easily seen that this is a bijective correspondence. Given a filling 𝑇 of 𝜆, we denote its associated permutation by 𝜙𝜆(𝑇).

Remark 2.10. In the case when 𝑁 is the principal nilpotent 𝑛×𝑛 matrix, the corresponding Young diagram 𝜆=𝜆𝑁=(𝑛) has only one row, so the above correspondence simply reads off the (one row of the) filling from left to right. In this case we abuse notation and denote 𝜙𝜆1(𝑤) by just 𝑤. For instance, the permissible filling of 𝜆=(5) in Example 2.9 has associated permutation 43215.
Now let 𝒫𝐹𝑖(𝜆,)(2.8) denote the set of (,𝜆)-permissible fillings of 𝜆. Recall that elements in Hess(𝑁,)𝑆1 are viewed as permutations in 𝑆𝑛 via the identification 𝑎𝑔𝑠(𝑛)𝑇𝑛𝑆𝑛. The next proposition follows from the definitions and some linear algebra. It is proven and discussed in more detail in [15], where the notation used is slightly different.

Proposition 2.11. Fix 𝑛 a positive integer. Let {1,2,,𝑛}{1,2,,𝑛} be a Hessenberg function and 𝜆 a Young diagram with 𝑛 boxes. Suppose 𝑁𝑓 is a nilpotent operator in highest form as chosen in [16] (cf. Remark 2.2) with 𝜆𝑁𝑓=𝜆. Let Hess(𝑁𝑓,) denote the associated nilpotent Hessenberg variety. Then the map from the 𝑆1-fixed points Hess(𝑁𝑓,)𝑆1 to the set of permissible fillings 𝒫𝐹𝑖(𝜆,)𝑤Hess()𝑆1𝑆𝑛𝜙𝜆1𝑤1𝒫𝐹𝑖(𝜆,)(2.9) is well defined and is a bijection.

Remark 2.12. In the case when 𝑁 is the principal nilpotent 𝑛×𝑛 matrix, 𝜆 is the Young diagram with only one row. Thus the map (2.9) above simplifies to 𝑤𝑤1 where we abuse notation (cf. Remark 2.10) and denote 𝜙𝜆1(𝑤1) by 𝑤1.

3. The Dimension Pair Algorithm for Betti Poset Pinball for Nilpotent Hessenberg Varieties

In this section we first explain the dimension pair algorithm which associates to any Hessenberg fixed point a permutation in 𝑆𝑛. The name is due to the fact that the construction proceeds by computing dimension pairs in appropriate permissible fillings. We then interpret this algorithm as a method for choosing rolldowns associated to the Hessenberg fixed points in a game of Betti poset pinball in the sense of [1]. The algorithm makes sense for any nilpotent Hessenberg variety, so it is defined in that generality in Section 3.1. However, our proof that the algorithm produces a successful outcome of Betti poset pinball in the sense of [1] is only for the special cases of regular nilpotent Hessenberg varieties and nilpotent Springer varieties. In Section 3.2 we briefly recall the setup and necessary results of poset pinball which allow us to conclude that our poset pinball result yields an explicit module basis for equivariant cohomology.

3.1. The Dimension Pair Algorithm for Nilpotent Hessenberg Varieties

Let 𝑁𝑓 be a nilpotent 𝑛×𝑛 matrix in highest form chosen as in Remark 2.2 and let 𝜆=𝜆𝑁𝑓. Let {1,2,,𝑛}{1,2,,𝑛} be a Hessenberg function and Hess(𝑁𝑓,) the corresponding nilpotent Hessenberg variety.

The definition of the dimension pair algorithm is pure combinatorics. It produces for each Hessenberg fixed point 𝑤Hess(𝑁𝑓,)𝑆1 an element in 𝑆𝑛. Following terminology of poset pinball, we denote this function by𝑟𝑜Hess𝑁𝑓,𝑆1𝑆𝑛.(3.1)

Definition 3.1 (the dimension pair algorithm). We define 𝑟𝑜Hess(𝑁𝑓,)𝑆1𝑆𝑛 as follows.(1)Let 𝑤Hess(𝑁𝑓,)𝑆1 and let 𝜙𝜆1(𝑤1) be its corresponding permissible filling as defined in (2.9).(2)Let 𝐷𝑃𝜙𝜆1(𝑤1) be the set of dimension pairs in the permissible filling 𝜙𝜆1(𝑤1).(3)For each with 2𝑛, set 𝑥|||=(𝑎,)(𝑎,)𝐷𝑃𝜙𝜆1(𝑤1)|||(3.2) as in (2.4) and define 𝐱=(𝑥2,,𝑥𝑛).(4)Define 𝑟𝑜(𝑤)=(𝜔(𝐱))1 where 𝜔(𝐱) is the permutation associated to the integer vector 𝐱 defined in (2.6).

Example 3.2. Let 𝜆, be as in Example 2.5. The permutation 𝑤=43215𝑆𝑛 is in Hess(𝑁𝑓,)𝑆1, as can be checked. The associated permissible filling is

254235.fig.008
. In Example 2.9 we saw that the associated permutation is 𝑠1(𝑠2𝑠1)𝑠3, so we conclude 𝑟𝑜(𝑤)=𝑠3(𝑠1𝑠2)𝑠1.

We next show that the rolldown function 𝑟𝑜Hess()𝑆1𝑆𝑛 defined by the dimension pair algorithm above satisfies the conditions to be a successful outcome of Betti poset pinball as in [1] in certain cases of nilpotent Hessenberg varieties. The statement of one of the conditions requires advance knowledge of the Betti numbers of nilpotent Hessenberg varieties, for which we recall the following result (reformulated in our language) from [16].

Theorem 3.3 (see [16, Theorem  1.1]). Let 𝑁𝑓𝑛𝑛 be a nilpotent matrix in highest form chosen as in Remark 2.2 and let 𝜆=𝜆𝑁𝑓. Let {1,2,,𝑛}{1,2,,𝑛} be a Hessenberg function and let Hess(𝑁𝑓,) denote the corresponding nilpotent Hessenberg variety. There is a paving by (complex) affine cells of Hess(𝑁𝑓,) such that(i)the affine cells are in one-to-one correspondence with Hess(𝑁𝑓,)𝑆1, and(ii)the (complex) dimension of the affine cell 𝐶𝑤 corresponding to a fixed point 𝑤Hess(𝑁,)𝑆1 isdim𝐶𝑤=|||𝐷𝑃𝜙𝜆1(𝑤1)|||.(3.3)

In particular, Theorem 3.3 implies that the odd Betti numbers of Hess(𝑁𝑓,) are 0, and the 2𝑘th even Betti number is precisely the number of fixed points 𝑤 in Hess(𝑁𝑓,)𝑆1 such that |𝐷𝑃𝜙𝜆1(𝑤1)|=𝑘. Given the regular nilpotent Hessenberg variety Hess(𝑁𝑓,), denote by 𝑏𝑘 its 2𝑘th Betti number, that is, 𝑏𝑘=dim𝐻2𝑘Hess𝑁𝑓.,(3.4) We may now formulate the conditions that guarantee that 𝑟𝑜Hess(𝑁𝑓,)𝑆1𝑆𝑛 is a successful outcome of Betti pinball. For more details we refer the reader to [1, Section 3]. It suffices to check the following(1)𝑟𝑜Hess(𝑁𝑓,)𝑆1𝑆𝑛 is injective,(2)for every 𝑤Hess(𝑁𝑓,)𝑆1, we have 𝑟𝑜(𝑤)𝑤 in Bruhat order, and(3)for every 𝑘0, 𝑘, we have 𝑏𝑘=|||𝑟𝑜(𝑤)𝑤Hess𝑁𝑓,𝑆1with|||,(𝑟𝑜(𝑤))=𝑘(3.5) where (𝑟𝑜(𝑤)) denotes the Bruhat length of ro(𝑤)𝑆𝑛.

We prove each claim in turn. For the first assertion we restrict to two special cases of Hessenberg varieties.

Lemma 3.4. Suppose that Hess(𝑁𝑓,) is either a regular nilpotent Hessenberg variety or a nilpotent Springer variety. Then the function roHess(𝑁𝑓,)𝑆1𝑆𝑛 is injective.

Proof. Since the association 𝐱=(𝑥2,𝑥3,,𝑥𝑛)𝜔(𝐱) given in (2.6) is a bijection, it suffices to show that the map which sends a Hessenberg fixed point 𝑤Hess()𝑆1 to the list of top parts 𝐱 of its associated permissible filling is injective. Mbirika shows that, in the cases of regular nilpotent Hessenberg varieties and nilpotent Springer varieties, there exists an inverse to this map (Mbirika works with monomials in 𝑛1 variables constructed from the list of top parts, but this is equivalent data) [14, Section 3.2]. The result follows.

Lemma 3.5. For every 𝑤Hess()𝑆1, one has 𝑟𝑜(𝑤)𝑤 in Bruhat order.

Proof. Since Bruhat order is preserved under taking inverses, it suffices to prove that 𝜔(𝐱) is Bruhat-less than 𝑤1. For any permutation 𝑢𝑆𝑛, set 𝑦={(𝑎,)(𝑎,)isaninversionin𝑢}(3.6) and let 𝐲=(𝑦2,𝑦3,,𝑦𝑛). Then the association (2.6) applied to the vector 𝐲 recovers the permutation 𝑢. By definition of 𝜙𝜆 and the definition of dimension pairs, the set 𝐷𝑃𝜙𝜆1(𝑤1) is always a subset of the set of inversions of the permutation 𝑤1. From Fact 1 it follows that the permutation 𝜔(𝐱) is Bruhat-less than 𝑤1 as desired.

Lemma 3.6. Let 𝑁𝑓𝑛𝑛 be a nilpotent matrix in highest form chosen as in Remark 2.2 and let 𝜆=𝜆𝑁𝑓. Let {1,2,,𝑛}{1,2,,𝑛} be a Hessenberg function and Hess(𝑁𝑓,) the associated nilpotent Hessenberg variety. For every 𝑘0, 𝑘, one has 𝑏𝑘=|||ro(𝑤)𝑤Hess()𝑆1with(ro|||,(𝑤))=𝑘(3.7) where (𝑟𝑜(𝑤)) denotes the Bruhat length of 𝑟𝑜(𝑤)𝑆𝑛.

Proof. By construction, 𝑟𝑜(𝑤) has a reduced word decomposition consisting of precisely |𝐷𝑃𝜙𝜆1(𝑤1)| simple transpositions. Hence its Bruhat length is |𝐷𝑃𝜙𝜆1(𝑤1)|. By Theorem 3.3, 𝑏𝑘 is precisely the number of fixed points 𝑤 with |𝐷𝑃𝜙𝜆1(𝑤1)|=𝑘 so the result follows.

The following is immediate from the above lemmas and the definition of Betti pinball given in [1, Section 3].

Proposition 3.7. Suppose that Hess(𝑁𝑓,) is either a regular nilpotent Hessenberg variety or a nilpotent Springer variety. Then the association 𝑤ro(𝑤) given by the dimension pair algorithm is a possible outcome of a successful game of Betti poset pinball played with ambient partially ordered set 𝑆𝑛 equipped with Bruhat order, rank function 𝜌=𝑆𝑛 given by Bruhat length, initial subset Hess()𝑆1𝑆𝑛, and target Betti numbers 𝑏𝑘=dim𝐻2𝑘(Hess();).

Remark 3.8. Lemmas 3.5 and 3.6 hold for general nilpotent 𝑁𝑓 and Hessenberg functions . Hence to prove that Proposition 3.7 holds for more general cases of nilpotent Hessenberg varieties, it suffices to check that the injectivity assertion (1) above holds. We do not know counterexamples where the injectivity fails. It would be of interest to clarify the situation for more general 𝑁𝑓 and .

3.2. Betti Pinball, Poset-Upper-Triangularity, and Module Bases

In the context of a GKM-compatible subspace of a GKM space [1, Definition  4.5], it is explained in [1, Section 4] that the outcome of a game of poset pinball may be interpreted as specifying a set of equivariant cohomology classes which, under additional conditions, yields a module basis for the equivariant cohomology of the GKM-compatible subspace. In this paper, the GKM space is the flag variety 𝑎𝑔𝑠(𝑛) with the standard 𝑇𝑛-action and the GKM-compatible subspace is Hess(𝑁𝑓,) with the 𝑆1-action specified above. Consider the 𝐻𝑇𝑛(pt)-module basis for 𝐻𝑇𝑛(𝑎𝑔𝑠(𝑛)) given by the equivariant Schubert classes {𝜎𝑤}𝑤𝑆𝑛. The dimension pair algorithm then specifies the set𝑝𝑟𝑜(𝑤)𝑤Hess𝑁𝑓,𝑆1𝐻𝑆1Hess𝑁𝑓,,(3.8) where for any 𝑢𝑆𝑛 the class 𝑝𝑢=𝜋(𝜎𝑢) is defined to be the image of 𝜎𝑢 under the natural projection map𝜋𝐻𝑇𝑛(𝑎𝑔𝑠(𝑛))𝐻𝑆1Hess𝑁𝑓,(3.9) induced by the inclusion of groups 𝑆1𝑇𝑛 and the 𝑆1-equivariant inclusion of spaces Hess(𝑁𝑓,)𝑎𝑔𝑠(𝑛). We refer to the images 𝑝𝑢 as Hessenberg Schubert classes.

Following the methods of [1] we view 𝐻𝑇𝑛(𝑎𝑔𝑠(𝑛)) and 𝐻𝑆1(Hess(𝑁𝑓,)) as subrings of𝐻𝑇𝑛(𝑎𝑔𝑠(𝑛))𝑇𝑛𝑤𝑆𝑛𝐻𝑇𝑛(pt)respectively𝐻𝑆1Hess𝑁𝑓,𝑆1𝑤Hess𝑁𝑓,𝑆1𝐻𝑆1(pt).(3.10) We denote by 𝜎𝑤(𝑤), 𝑝𝑟𝑜(𝑤)(𝑤) the value of the 𝑤th coordinate in the direct sums above, for 𝑤, 𝑤𝑆𝑛 or 𝑤, 𝑤Hess(𝑁𝑓,)𝑆1, respectively. If𝑝𝑟𝑜(𝑤)(𝑤)0,𝑝𝑟𝑜(𝑤)𝑤=0if𝑤𝑤(3.11) for all 𝑤, 𝑤Hess(𝑁𝑓,)𝑆1, then the set {𝑝ro(𝑤)𝑤Hess(𝑁𝑓,)𝑆1} in 𝐻𝑆1(Hess(𝑁𝑓,)) is called poset-upper-triangular (with respect to the partial order on Hess(𝑁𝑓,)𝑆1𝑆𝑛 induced from Bruhat order) [1, Definition 2.3]. Finally, recall that the cohomology degree of an equivariant Schubert class 𝜎𝑤 (and hence also the corresponding Hessenberg Schubert class 𝑝𝑤) is 2(𝑤).

The following is immediate from [1, Proposition  4.14] and the above discussion.

Proposition 3.9. Let Hess(𝑁𝑓,) be either a regular nilpotent Hessenberg variety or a nilpotent Springer variety. Let roHess(𝑁𝑓,)𝑆1𝑆𝑛 be the dimension pair algorithm defined above. Suppose (3.11) holds for all 𝑤Hess(𝑁𝑓,)𝑆1. Then the Hessenberg Schubert classes {𝑝ro(𝑤)𝑤Hess(𝑁𝑓,)𝑆1} form a 𝐻𝑆1(pt)-module basis for the 𝑆1-equivariant cohomology ring 𝐻𝑆1(Hess(𝑁𝑓,)).

Therefore, in order to prove that the Hessenberg Schubert classes above form a module basis as desired, it suffices to show that they satisfy the upper-triangularity conditions (3.11) for all 𝑤, 𝑤Hess(𝑁𝑓,)𝑆1. The proof of this assertion, for a special class of regular nilpotent Hessenberg varieties closely related to Peterson varieties, is the content of Sections 4 and 5.

We close the section with a brief discussion of matchings. Following [1, Section 4.3], definedegHess(𝑁𝑓,)(𝑤)=dim𝐶𝑤(3.12) to be the (complex) dimension of the affine cell 𝐶𝑤 containing the fixed point 𝑤 in Tymoczko’s paving by affines of Hess(𝑁𝑓,) in Theorem 3.3. Then from the discussion above we knowdegHess(𝑁𝑓,)|||(𝑤)=𝐷𝑃𝜙𝜆1(𝑤1)|||=(𝑟𝑜(𝑤)),(3.13) and since the cohomology degree of 𝑝ro(𝑤) is 2(𝑟𝑜(𝑤)), we see that the association 𝑤𝑟𝑜(𝑤) from Hess(𝑁𝑓,)𝑆1𝑆𝑛 is also a matching in the sense of [1] with respect to degHess(𝑁𝑓,) and rank function 𝜌 on 𝑆𝑛 given by Bruhat length. Thus the fact that the {𝑝𝑟𝑜(𝑤)𝑤Hess(𝑁𝑓,)𝑆1} form a module basis can also be deduced from [1, Theorem  4.18].

4. Poset-Upper-Triangularity of Rolldown Classes for 334-Type Hessenberg Varieties

In this section and in Section 5 we analyze in detail the dimension pair algorithm in the case of a Hessenberg variety which is closely related to the Peterson variety and in particular prove that the algorithm produces a poset-upper-triangular module basis for its 𝑆1-equivariant cohomology ring. Here and below the nilpotent operator 𝑁 under consideration is always the principal nilpotent, so we omit the 𝑁 from the notation and write Hess(). Similarly the corresponding Young diagram is always 𝜆=(𝑛) so we omit the 𝜆 from notation and write 𝒫𝐹𝑖() instead of 𝒫𝐹𝑖(𝜆,).

We fix for this discussion the Hessenberg function given by(1)=(2)=3,(𝑖)=𝑖+1for3𝑖𝑛1,(𝑛)=𝑛.(4.1) The only difference between this function and the Hessenberg function for the Peterson variety studied in [13] is that the value of (1) is 3 instead of 2. In this sense this is “close’’ to the Peterson case. Thus it is natural that much of our analysis follows that for Peterson varieties in [13], although it is still necessary to introduce new ideas and terminology to handle the Hessenberg fixed points in Hess()𝑆1 which do not arise in the Peterson case.

The Hessenberg function in (4.1) is trivial if 𝑛=3 since in that case (1)=(2)=(3)=3 which implies that the corresponding Hessenberv variety Hess() is equal to the full flag variety 𝑎𝑔𝑠(3). Hence we assume 𝑛4 throughout. Under this assumption and following the notation introduced in Section 2, the Hessenberg function is of the form =(3,3,4,). As such, for the purposes of this manuscript, we refer to this family of regular nilpotent Hessenberg varieties as 334-type Hessenberg varieties.

Our main result is the following theorem.

Theorem 4.1. Let 𝑛4 and let Hess() be the 334-type Hessenberg variety in 𝑎𝑔𝑠(𝑛). Let 𝑟𝑜Hess()𝑆1𝑆𝑛 be the dimension pair algorithm defined in Section 3. Then 𝑝𝑟𝑜(𝑤)(𝑤)0,𝑝𝑟𝑜(𝑤)𝑤=0if𝑤𝑤(4.2) for all 𝑤, 𝑤Hess()𝑆1. In particular the Hessenberg Schubert classes {𝑝𝑟𝑜(𝑤)𝑤Hess()𝑆1} form a 𝐻𝑆1(pt)-module basis for the 𝑆1-equivariant cohomology ring 𝐻𝑆1(Hess()).

For ease of exposition, and because the arguments required are of a somewhat different nature, we prove Theorem 4.1 by proving the two assertions in (4.2) separately, as follows.

Proposition 4.2. Let 𝑛, , Hess() and 𝑟𝑜 be as above. Then, 𝑝ro(𝑤)(𝑤)0(4.3) for all 𝑤Hess()𝑆1.

Proposition 4.3. Let 𝑛, , Hess() and 𝑟𝑜 be as above. Then, 𝑝𝑟𝑜(𝑤)𝑤=0if𝑤𝑤(4.4) for all 𝑤, 𝑤Hess()𝑆1.

The proof of Proposition 4.2 is the content of Section 5. The main result of the present section is the upper-triangularity property asserted in Proposition 4.3. Its proof requires a number of preliminary results. We first begin by reformulating the problem in terms of Bruhat relations among the fixed points.

Lemma 4.4. Let 𝑛, , Hess() and 𝑟𝑜 be as above. If for all 𝑤, 𝑤Hess()𝑆1 one has 𝑟𝑜(𝑤)𝑤𝑤𝑤(4.5) in Bruhat order, then the Hessenberg Schubert classes {𝑝𝑟𝑜(𝑤)𝑤Hess()𝑆1} satisfy (4.4).

Proof. Recall that the equivariant Schubert classes are poset-upper-triangular with respect to Bruhat order on 𝑆𝑛. In particular, for all 𝑤, 𝑤𝑆𝑛 we have 𝜎𝑤(𝑤)=0 if ̸𝑤𝑤. Since the Hessenberg Schubert classes are images of the Schubert classes and the diagram 𝐻𝑇𝑛(𝑎𝑔𝑠(𝑛))𝐻𝑇𝑛(𝑎𝑔𝑠(𝑛))𝑇𝑛𝑤𝑊𝐻𝑇𝑛(pt)𝐻𝑆1(Hess())𝐻𝑆1(Hess())𝑆1𝑤Hess()𝑆1𝐻𝑆1(pt)(4.6) commutes, it follows that if for all 𝑤, 𝑤Hess()𝑆1, we have 𝑟𝑜(𝑤)𝑤𝑤𝑤(4.7) in Bruhat order, then (4.4) follows.

The rest of this section is devoted to the proof of (4.5), which by Lemma 4.4 then proves Proposition 4.3.

4.1. Fixed Points and Associated Subsets for the 334-Type Hessenberg Variety

In this section we enumerate the fixed points in the 334-type Hessenberg variety and also associate to each fixed point in Hess()𝑆1 a subset of {1,2,,𝑛1}. As we show below, the set of fixed points in the Peterson variety is a subset of the fixed points of the 334-type Hessenberg variety, so the main task is to describe the new fixed points which arise in the 334-type case. We begin with a general observation.

Lemma 4.5. Let 𝑛 and let ,{1,2,,𝑛}{1,2,,𝑛} be two Hessenberg functions. If (𝑖)(𝑖) for all 𝑖, 1𝑖𝑛, then HessHess().(4.8) The inclusion Hess()Hess() is 𝑆1-equivariant and in particular Hess()𝑆1Hess()𝑆1 and 𝒫𝐹𝑖()𝒫𝐹𝑖().

Proof. Let 𝑉=(𝑉𝑖) denote an element in 𝑎𝑔𝑠(𝑛). By definition the regular nilpotent Hessenberg variety Hess() associated to is Hess𝑉=𝑎𝑔𝑠(𝑛)𝑁𝑉𝑖𝑉(𝑖),,1𝑖𝑛(4.9) where 𝑁 is the principal nilpotent operator. Since 𝑉𝑖𝑉𝑖+1 for all 1𝑖𝑛1 by definition of flags and 𝑉𝑛=𝑛 for all flags, if (𝑖)(𝑖) for all 𝑖, then 𝑁𝑉𝑖𝑉(𝑖) automatically implies 𝑁𝑉𝑖𝑉(𝑖). We conclude Hess()Hess(). The 𝑆1-equivariance of the inclusion Hess()Hess() follows from the definition of the 𝑆1-action of (2.3).

Applying Lemma 4.5 to the Hessenberg function(𝑖)=𝑖+1for1𝑖𝑛1,(𝑛)=𝑛(4.10) corresponding to the Peterson variety Hess() and the 334-type Hessenberg function (4.1), we conclude that all fixed points in Hess()𝑆1 also arise as fixed points in Hess()𝑆1. We refer to the elements of Hess()𝑆1 (viewed as elements of Hess()𝑆1) as Peterson-type fixed points. It therefore remains to describe Hess()𝑆1Hess()𝑆1. It turns out to be convenient to do this by first describing 𝒫𝐹𝑖()𝒫𝐹𝑖().

We first introduce some terminology. Given a permutation 𝑤=(𝑤(1)𝑤(2)𝑤(𝑛)) in one-line notation and some 𝑖, , we say that the entries {𝑤(𝑖),𝑤(𝑖+1),,𝑤(𝑖+)} form a decreasing staircase, or simply a staircase, if 𝑤(𝑗+1)=𝑤(𝑗)1 for all 𝑖𝑗<𝑖+. For example, for 𝑤=4327516, the segment 432 is a staircase, but 751, though the entries decrease, is not. We will say that a consecutive series of staircases is an increasing sequence of staircases (or simply increasing staircases) if each entry in a given staircase is smaller than any entry in any following staircase (reading from left to right). For instance, 𝑤=654987321 is a sequence of staircases 654, 987, and 321, but is not an increasing sequence of staircases since the entries 4,5,6 are not smaller than the entries in the later staircase 321. However, 𝑤=321654987 is an increasing sequence of (three) staircases 321, 654, and 987.

It is shown in [13] that the 𝑆1-fixed points of the Peterson variety Hess() consist precisely of those permutations 𝑤𝑆𝑛 such that the one-line notation of 𝑤 is an increasing sequence of staircases. Since such 𝑤 are equal to their own inverses, the permissible fillings 𝒫𝐹𝑖() corresponding to Hess() are precisely those which are increasing sequences of staircases (cf. Remark 2.12). We now describe the permissible fillings 𝒫𝐹𝑖() which are not Peterson-type fillings. We use the language of -tableau trees introduced by Mbirika; see [14, Section 3.1] for definitions. Recall from Remark 2.10 that we identify permissible fillings with permutations in 𝑆𝑛 via one-line notation.

Lemma 4.6. Let 𝑛4 and let Hess() be the 334-type Hessenberg variety in 𝑎𝑔𝑠(𝑛). Let 𝑤𝒫𝐹𝑖() be a permissible filling for Hess() which is not of Peterson type, that is, 𝑤𝒫𝐹𝑖()𝒫𝐹𝑖(). Then precisely one of the following hold.(i)The one-line notation of 𝑤 is of the form 𝑤𝑤312𝑤,(4.11) where 𝑤 is a (possibly empty) staircase such that 𝑤3 is also a staircase, and 𝑤 is an increasing sequence of staircases. We refer to these as 312-type permissible fillings. (ii)The one-line notation of 𝑤 is of the form 2𝑤31𝑤,(4.12) where 𝑤 is a (possibly empty) staircase such that 𝑤3 is also a staircase, and 𝑤 is an increasing sequence of staircases. We refer to these as 231-type permissible fillings.
Moreover, any filling satisfying either of the above conditions appears in 𝒫𝐹𝑖()𝒫𝐹𝑖().

Proof of Lemma 4.6. For any Hessenberg function {1,2,,𝑛}{1,2,,𝑛}, Mbirika shows in [14, Section 3.2] that the Level 𝑛 fillings in an -tableau tree are precisely the permissible fillings with respect to . For the Peterson Hessenberg function in (4.10) Mbirika’s corresponding -tableau tree has the property that for every 𝑘 with 1𝑘𝑛1 and every vertex at Level 𝑘, there are precisely 2 edges going down from that vertex to a Level 𝑘+1 vertex (this is because the corresponding degree tuple 𝛽 [14, Definition  3.1.1] has 𝛽𝑖=2 for all 1𝑖𝑛1). In the case of the 334-type Hessenberg function, by definition the -tableau tree also has precisely 2 edges going down from every vertex at Level 𝑘 for all 𝑘2, 1𝑘𝑛1. However, at Level 2, each vertex has not 2 but 3 edges pointing down to a vertex at Level 3.
From [14, Section 3] (cf. in particular [14, Definition 3.1.9]) it can be seen that for the case of the Peterson Hessenberg function, the corresponding -tableau tree at Level 2 has vertices 21 and 12, whereas for the 334-type Hessenberg function, the Level 2 vertices have the form 21 and 12. Here the bullets indicate the locations of the -permissible positions available for the placement of the next index 3, in the sense of [14, Section 3] (cf. in particular [14, Lemma  3.1.8]). In particular, since we saw above that the edges going down from Level 3 onwards are identical in both the Peterson and 334-type Hessenberg case, it follows that the branches of the tree emanating downwards from the two Level 3 vertices 321, 213 (coming from 21) and the two vertices 132, 123 (coming from 12) are identical to the corresponding branches in the -tableau tree for the Peterson Hessenberg function. Hence all permissible fillings at the final Level 𝑛 of these branches are of Peterson type. In contrast, the branches emanating from 231 and 312 do not appear in the Peterson -tableau tree, and none of the fillings appearing at Level 𝑛 in these branches can be Peterson permissible fillings since a 3 appears directly before a 1. Hence it is precisely these branches which account for the permissible fillings which are not of Peterson type. As noted above, the rest of the branch only has 2 edges going down from each vertex with -permissible positions determined exactly as in the Peterson case. In particular, except for the exceptional 3 appearing directly to the left of a 1, the fillings must consist of decreasing staircases and all possible arrangements of decreasing staircases do appear. The result follows.

Example 4.7. Suppose 𝑛=8. Then 𝑤=54312876 is an example of a 312-type permissible filling where 𝑤=54 and 𝑤=876. An example of a 231-type permissible filling is 𝑤=25431876 where 𝑤=54 and 𝑤=876. Neither of these are permissible with respect to the Peterson Hessenberg function since a 3 appears directly to the left of a 1. Nevertheless, both of these fillings are closely related to the Peterson-type permissible filling 𝑤=54321876; this relationship is closely analyzed and used below.

We now give explicit descriptions of the corresponding non-Peterson-type elements in Hess()𝑆1, obtained by taking inverses of the permissible fillings described in Lemma 4.6.

Definition 4.8. Let 𝑤Hess()𝑆1. We say 𝑤 is a 312-type (resp., 231-type) fixed point if its inverse 𝑤1 is a permissible filling of 312-type (resp., 231-type).
As observed above, since Peterson-type permissible fillings are equal to their own inverses, in that case there is no distinction between the fillings and their associated fixed points. For the 312 and 231 types, however, this is not the case. We record the following. The proof is a straightforward computation and is left to the reader.

Lemma 4.9. Let 𝑤 be a 312-type (resp., 231-type) permissible filling. Let 𝑎2 be the integer such that 𝑎2+1 is the first entry (resp., second entry) in the one-line notation of 𝑤. Let 𝑤1 be the corresponding 312-type (resp., 231 type) fixed point. Then,(i)the one-line notation of 𝑤1 is the same as that of 𝑤 for all th entries with >𝑎2+1,(ii)if 𝑤 is 312-type, then the first 𝑎2+1 entries of the one-line notation of 𝑤1 are 𝑎2𝑎2+1𝑎21𝑎2221,(4.13)(iii)if 𝑤 is 231-type, then the first 𝑎2+1 entries of the one-line notation of 𝑤1 are 𝑎2+11𝑎2𝑎2132.(4.14)

In the case of the Peterson variety, there is a convenient bijective correspondence between the set of 𝑆1-fixed points of the Peterson variety and subsets 𝒜 of {1,2,,𝑛1} given as follows [13, Section 2.3]. Let 𝑤 be a Peterson-type fixed point. Then the corresponding subset is𝒜={𝑖1𝑖𝑛1,𝑤(𝑖)=𝑤(𝑖+1)+1}{1,2,,𝑛1}.(4.15) In the case of the 334-type Hessenber variety, it is also useful to assign a subset of {1,2,,𝑛1} to each fixed point as follows.

Definition 4.10. Let 𝑤Hess()𝑆1. The associated subset of {1,2,,𝑛}corresponding to 𝑤, notated 𝒜(𝑤), is defined as follows.(i)Suppose 𝑤 is of Peterson type. Then 𝒜(𝑤) is defined to be the set 𝒜 in (4.15).(ii)Suppose 𝑤 is 312-type. Consider the permutation 𝑤=𝑤𝑠1 (i.e., swap the 𝑎2 and the 𝑎2+1 in the one-line notation (4.13)). This is a fixed point of Peterson type. Define 𝒜(𝑤)=𝒜(𝑤).(iii)Suppose 𝑤 is 231-type. Consider the permutation 𝑤=𝑤𝑠2𝑠3𝑠𝑎2(4.16) (i.e., move the 1 to the right of the 2 in the one-line notation (4.14)). This is a fixed point of Peterson type.  Define 𝒜(𝑤)=𝒜(𝑤).

Example 4.11. Suppose 𝑛=8.(i)Suppose 𝑤 is the Peterson-type fixed point 𝑤=54321876. Then 𝒜(𝑤)={1,2,3,4}{6,7}. This agrees with the association 𝑤𝒜(𝑤) used in [13].(ii)Suppose 𝑤 is the 312-type fixed point 𝑤=34217658 (corresponding to the 312-type permissible filling 43127658). Then 𝑤=𝑤𝑠1=43217658 and 𝒜(𝑤)=𝒜(𝑤)={1,2,3}{5,6}.(iii)Suppose 𝑤 is the 231-type fixed point 𝑤=51432768 (corresponding to the 231-type permissible filling 25431768). Then 𝑤=54321768 and 𝒜(𝑤)=𝒜(𝑤)={1,2,3,4}{6}.

Remark 4.12. The three fixed points 𝑤=54321876, 𝑤=45321876, and 𝑤=51432876, which are, respectively, of Peterson type, 312 type, and 231-type, all have the same associated subset 𝒜(𝑤)={1,2,3,4}{6,7}.

It is useful to observe that the 312-type and 231-type fixed points have associated subsets that always contain 1 and 2.

Lemma 4.13. Let 𝑤 be a 334-type Hessenberg fixed point. Suppose further that 𝑤 is not of Peterson type. Then {1,2}𝒜(𝑤).

Proof. From the explicit descriptions of the one-line notation of the 312 type (resp., 231-type) fixed points given above, we know that the initial segment 𝑎2𝑎2+121 (resp, 𝑎2+11𝑎232) in the one-line notation is such that 𝑎22. From Definition 4.10 it follows that the first decreasing staircase of the associated Peterson-type fixed point 𝑤𝑠1 (resp., 𝑤𝑠2𝑠3𝑠𝑎2) is of length at least 3. In particular, the first staircase starts with an integer 𝑘 which is ≥ 3. The result follows.

As noted in Remark 4.12, the association 𝑤𝒜(𝑤) given in Definition 4.10 is not one-to-one and hence in particular not a bijective correspondence. This makes our analysis more complicated than in [13], but the notion is still useful for our arguments below.

4.2. Reduced Word Decompositions for 334-Type Fixed Points and Rolldowns

In this section we fix particular choices of reduced word decompositions for the fixed points in Hess()𝑆1 which we use in our arguments below. We also compute, and fix choices of reduced words for, the rolldowns 𝑟𝑜(𝑤) of the fixed points.

The association 𝑤𝒜(𝑤) of the previous section allows us to describe these reduced word decompositions in relation to that of the Peterson-type fixed points. Let 𝑎 be a positive integer and 𝑘 a nonnegative integer. Recall that a reduced word decomposition of the maximal element (the full inversion) in the subgroup 𝑆{𝑎,𝑎+1,,𝑎+𝑘+1}𝑆𝑛 is given by𝑠𝑎𝑠𝑎+1𝑠𝑎𝑠𝑎+2𝑠𝑎+1𝑠𝑎𝑠𝑎+𝑘𝑠𝑎+𝑘1𝑠𝑎+1𝑠𝑎.(4.17) For the purposes of this manuscript, we call this the standard reduced word (decomposition) for the maximal element. (This is different from the choice of reduced word decomposition used in [13, Section 2.3].) We denote a consecutive set of integers {𝑎,𝑎+1,,𝑎+𝑘} for 𝑎 positive and 𝑘 a nonnegative integer by [𝑎,𝑎+𝑘]. We say that [𝑎,𝑎+𝑘] is a maximal consecutive substring of 𝒜 if [𝑎,𝑎+𝑘]𝒜 and neither 𝑎1 nor 𝑎+𝑘+1 are in 𝒜. It is straightforward that any subset 𝒜 of {1,2,,𝑛1} uniquely decomposes into a disjoint union of maximal consecutive substrings𝑎𝒜=1,𝑎2𝑎3,𝑎4𝑎𝑚1,𝑎𝑚.(4.18) For instance, for 𝒜={1,2,3,5,6,9,10,11}, the decomposition is 𝒜=[1,3][5,6][9,11]. For any [𝑎,𝑏], denote by 𝑤[𝑎,𝑏] the full inversion in the subgroup 𝑆[𝑎,𝑏+1]. Then it follows from Definition 4.10 (see also [13, Section 2.3]) that the Peterson-type fixed point associated to 𝒜, which we denote by 𝑤𝒜, is the product𝑤𝒜=𝑤[𝑎1,𝑎2]𝑤[𝑎3,𝑎4]𝑤[𝑎5,𝑎6]𝑤[𝑎𝑚1,𝑎𝑚].(4.19) We fix a choice of reduced word decomposition of 𝑤𝒜 given by taking the product of the standard reduced words (4.17) for each of the full inversions 𝑤[𝑎𝑗,𝑎𝑗+1] appearing in (4.19). For the purposes of this manuscript we call this the standard reduced word decomposition of a Peterson-type fixed point 𝑤𝒜.

Example 4.14. Let 𝑛=7 and let 𝑤=4321765 be a Peterson-type fixed point. Then the two decreasing staircases are 4321 and 765, the associated subset 𝒜(𝑤) is {1,2,3}{5,6} with maximal consecutive strings [1,3]={1,2,3} and [5,6]={5,6}. The standard reduced word decomposition of 𝑤 is 𝑤{1,2,3}{5,6}=𝑤[1,3]𝑤[5,6]=𝑠1𝑠2𝑠1𝑠3𝑠2𝑠1𝑠5𝑠6𝑠5.(4.20)

We now fix a reduced word decomposition of the non-Peterson-type fixed points.

Lemma 4.15. Let 𝑤Hess()𝑆1 be a fixed point which is not of Peterson type and let 𝒜(𝑤)=[𝑎1,𝑎2][𝑎3,𝑎4][𝑎𝑚1,𝑎𝑚] be the associated subset with its decomposition into maximal consecutive substrings.(i)If 𝑤 is 312-type, then a reduced word decomposition for 𝑤 is given by 𝑠1𝑠2𝑠1𝑠𝑎2𝑠𝑎21𝑠3𝑠2𝑤[𝑎3,𝑎4]𝑤[𝑎𝑚1,𝑎𝑚].(4.21)(ii)If 𝑤 is 231-type, then a reduced word decomposition for 𝑤 is given by 𝑠2𝑠3𝑠2𝑠𝑎21𝑠𝑎22𝑠3𝑠2𝑠𝑎2𝑠𝑎21𝑠2𝑠1𝑤[𝑎3,𝑎4]𝑤[𝑎𝑚1,𝑎𝑚],(4.22) where the 𝑤[𝑎,𝑎+1] in the above expressions are assumed to be given the reduced word decomposition described in (4.17).

Proof. For the first assertion, observe that the explicit description of the one-line notation 312-type fixed points in (4.13) implies that 𝑤 has precisely 1 fewer inversion than 𝑤𝒜(𝑤). An explicit computation shows that the given word (4.21) is equal to 𝑤, so it is a word decomposition of 𝑤 with exactly as many simple transpositions as the Bruhat length of 𝑤. In particular it must be reduced. A similar argument proves the second assertion.

Example 4.16. Suppose 𝑛=7. Suppose 𝑤=3421765 is a 312-type fixed point. Then the reduced word decomposition of 𝑤 given in Lemma 4.15 is 𝑤=𝑠1𝑠2𝑠1𝑠3𝑠2𝑠5𝑠6𝑠5.(4.23) Similarly suppose 𝑤=4132765 is a 231-type fixed point. Then the reduced word decomposition of 𝑤 given in Lemma 4.15 is 𝑤=𝑠2𝑠3𝑠2𝑠1𝑠5𝑠6𝑠5.(4.24)

Henceforth, we always use the reduced words given above.

Next we explicitly describe the rolldowns 𝑟𝑜(𝑤) associated to each 𝑤 in Hess()𝑆1 by the dimension pair algorithm. We begin with the Peterson-type fixed points. It turns out there are two important subcases of Peterson-type fixed points.

Definition 4.17. We say that a Peterson-type fixed point 𝑤contains the string 321 (or simply contains 321) if, in the one-line notation of 𝑤, the string 321 appears (equivalently, if {1,2}𝒜(𝑤)). We say 𝑤does not contain the string 321 (or simply does not contain 321) otherwise.

Remark 4.18. Note that Definition 4.17 is different from the standard notion of pattern-containing or pattern-avoiding permutations since here we require the one-line notation of 𝑤 to contain the string 321 exactly.

Given a subset 𝒜={𝑗1<𝑗2<<𝑗𝑘}{1,2,,𝑛1} and corresponding Peterson-type fixed point 𝑤𝒜, we call the permutation𝑠𝑗𝑘𝑠𝑗𝑘1𝑠𝑗2𝑠𝑗1𝑆𝑛(4.25) the Peterson case rolldown of 𝑤𝒜. Note that the word (4.25) is in fact a reduced word decomposition of this permutation; we always use this choice of reduced word. The terminology is motivated by the fact that (4.25) is the (inverse of the) permutation given in [13, Definition  4.1]. (The fact that it is the inverse of the permutation used in [13] does not affect the theory very much, as is explained in [13, Proposition  5.16]).

Lemma 4.19. Let 𝑛4 and let Hess() be the 334-type Hessenberg variety in 𝑎𝑔𝑠(𝑛). Let 𝑤 be a Peterson-type fixed point and let 𝒜(𝑤)={𝑗1<𝑗2<<𝑗𝑘} be its associated subset.(i)Suppose 𝑤 does not contain 321. Then 𝑟𝑜(𝑤) is the Peterson case rolldown of 𝑤𝒜(𝑤).(ii)Suppose 𝑤 does contain 321, that is, 𝒜(𝑤)={𝑗1<𝑗2<<𝑗𝑘} for 𝑘2 and 𝑗1=1 and 𝑗2=2. Then 𝑟𝑜(𝑤) is𝑟𝑜(𝑤)=𝑠𝑗𝑘𝑠𝑗𝑘1𝑠𝑗3𝑠1𝑠2𝑠1.(4.26) In particular, if a Peterson-type fixed point 𝑤 contains 321, then its rolldown 𝑟𝑜(𝑤) is Bruhat-greater and has Bruhat length 1 greater than the Peterson case rolldown of 𝑤.

Proof. If 𝑤 contains a 321, then by Definition 2.6, the pairs (1,3),(2,3), and (1,2) are all dimension pairs in 𝑤. Hence 3 appears precisely twice as a top part of a dimension pair and 2 appears precisely once. Thus by constructing the dimension pair algorithm the permutation 𝜔(𝐱) begins with the word 𝑠1(𝑠2𝑠1). With respect to all other indices 𝑗𝒜(𝑤), the 334-type Hessenberg function is identical to the Peterson Hessenberg function and hence, for each such 𝑗, the index 𝑗+1 appears precisely once as a top part of a dimension pair of 𝑤 and thus contributes precisely one 𝑠𝑗 to 𝜔(𝐱). Taking the inverse yields (4.26) as desired.
If 𝑤 does not contain 321, then 3 appears at most once as the top part of a dimension pair in 𝑤, and again for all other indices the computations are identical to the Peterson case as above. Hence 𝑟𝑜(𝑤) is identical to the Peterson case rolldown. This completes the proof.

Next, we give an explicit description, along with a choice of reduced word decomposition, of the rolldowns corresponding to the non-Peterson-type fixed points.

Lemma 4.20. Let 𝑤Hess()𝑆1 and suppose that 𝑤 is not of Peterson type. Let 𝒜(𝑤)={𝑗1=1<𝑗2=2<𝑗3<<𝑗𝑘} for some 𝑘2.(1)If 𝑤 is of 312-type, then the dimension pair algorithm associates to 𝑤 the permutation 𝑟𝑜(𝑤)=𝑠𝑗𝑘𝑠𝑗𝑘1𝑠𝑗4𝑠𝑗3𝑠1𝑠2.(4.27)(2)If 𝑤 is 231-type, then the dimension pair algorithm associates to 𝑤 the permutation𝑟𝑜(𝑤)=𝑠𝑗𝑘𝑠𝑗𝑘1𝑠𝑗4𝑠𝑗3𝑠2𝑠1.(4.28)

Proof. Suppose 𝑤 is a 312-type fixed point so 𝜙𝜆1(𝑤1) is a 312-type permissible filling. By definition of dimension pairs, 2 does not appear as the top part of any dimension pair (since it appears to the right of a 1). Also by definition, 3 appears as a top part of the two dimension pairs (1,3) and (2,3). The form of the 312-type permissible fillings described in Lemma 4.6 and the definition of 𝒜(𝑤) imply that the other dimension pairs are precisely the pairs (𝑗,𝑗+1) for 𝑗𝒜(𝑤) (for 𝑗1,2), from which it follows that 𝜔(𝐱)=𝑠2𝑠1𝑠𝑗3𝑠𝑗4𝑠𝑗𝑘1𝑠𝑗𝑘. Taking inverses yields (4.27). The proof of the second assertion is similar.

Example 4.21. (i) Suppose 𝑤=54321876. This is of Peterson type. Then ro(𝑤)=(𝜔(𝐱))1=𝑠7𝑠6𝑠4𝑠3𝑠1𝑠2𝑠1.(ii)Suppose 𝑤=45321876. This is 312-type. Then ro(𝑤)=(𝜔(𝐱))1=𝑠7𝑠6𝑠4𝑠3𝑠1𝑠2.(iii)Suppose 𝑤=51432876. This is 231-type. Then ro(𝑤)=(𝜔(𝐱))1=𝑠7𝑠6𝑠4𝑠3𝑠2𝑠1.

We conclude the section with a computation of the one-line notation of the rolldowns for different types; we leave proofs to the reader.

Lemma 4.22. Let 𝑤 be a 334-type Hessenberg fixed point and let 𝒜(𝑤)=[𝑎1,𝑎2][𝑎𝑚1,𝑎𝑚] be its associated subset with its decomposition into maximal consecutive substrings. Suppose 𝑤 is of Peterson type that contains 321, 312-type, or of 231 type. Then 𝑎1=1, 𝑎22 and the first 𝑎2+1 entires of the one-line notation of 𝑟𝑜(𝑤) are 𝑎2+12134𝑎2(4.29) for 𝑤 of Peterson type that contains 321, 2𝑎2+1134𝑎2(4.30) for 𝑤312-type, and 𝑎2+1123𝑎2(4.31) for 𝑤231-type.

4.3. Bruhat Order Relations

In this section we analyze the properties of the association 𝑤𝒜(𝑤) with respect to comparisons in Bruhat order.

The first two lemmas are straightforward and proofs are left to the reader.

Lemma 4.23. Let 𝒜{1,2,,𝑛1} and let 𝑤𝒜 be the Peterson-type filling associated to 𝒜. Then 𝑤𝒜 is maximal in the subgroup 𝑆𝒜 of 𝑆𝑛 generated by the simple transpositions {𝑠𝑖}𝑖𝒜. In particular, 𝑤𝒜 is Bruhat-bigger than any permutation 𝑤𝑆𝒜.

Lemma 4.24. Let 𝑤Hess()𝑆1. Suppose 𝑤 is not of Peterson type. Then 𝑤 is Bruhat-less than the Peterson type fixed point 𝑤𝒜(𝑤) corresponding to 𝒜(𝑤).

We also observe that a Bruhat relation 𝑤<𝑤 implies a containment relation of the associated subsets.

Lemma 4.25. Let 𝑤, 𝑤Hess()𝑆1 and let 𝒜(𝑤), 𝒜(𝑤) be the respective associated subsets. Let 𝑠𝑖 be a simple transposition. Then,(1)𝑠𝑖<𝑤 if and only if 𝑖𝒜(𝑤),(2)𝑠𝑖<𝑟𝑜(𝑤) if and only if 𝑖𝒜(𝑤),(3)if 𝑤𝑤 or 𝑟𝑜(𝑤)𝑤, then 𝒜(𝑤)𝒜(𝑤).

Proof. Bruhat order is independent of choice of reduced word decomposition for 𝑤. Therefore, a simple transposition 𝑠𝑖 is less than 𝑤 in Bruhat order if and only if 𝑠𝑖 appears in a (and hence any) reduced word decomposition of 𝑤. In particular, to prove the first claim it suffices to observe that by the definitions of 𝒜(𝑤), the index 𝑖 appears in 𝒜(𝑤) precisely when 𝑠𝑖 appears in the choice of reduced word for 𝑤 given above. A similar argument using the explicit reduced words given for 𝑟𝑜(𝑤) in Lemmas 4.19 and 4.20 proves the second claim. The last claim follows from the first two.

We have just seen that 𝑤𝑤 implies 𝒜(𝑤)𝒜(𝑤). In the case of the Peterson variety Hess() these Bruhat relations are precisely encoded by the partial ordering given by containment of the 𝒜(𝑤); specifically, by Lemma 4.23, 𝑤𝒜𝑤 if and only if 𝒜. In our 334-type Hessenberg case this is no longer true although the sets 𝒜(𝑤) do still encode the Bruhat data. The precise statements occupy the next several lemmas.

We take a moment to recall the tableau criterion for determining Bruhat order in the Weyl group 𝑆𝑛 (see, e.g., [17]) which will be useful in the discussion below. For 𝑤𝑆𝑛, denote by 𝐷𝑅(𝑤) the descent set of 𝑤, namely,𝐷𝑅(𝑤)={𝑖𝑤(𝑖)>𝑤(𝑖+1),1𝑖𝑛1}.(4.32) For example, for 𝑤=368475912 the descent set is 𝐷𝑅(𝑤)={3,5,7}.

Theorem 4.26 (the tableau criterion [17, Theorem  2.6.3]). For 𝑤, 𝑣𝑆𝑛, let 𝑤𝑖,𝑘 be the 𝑖th element in the increasing rearrangement of 𝑤(1),𝑤(2),,𝑤(𝑘), and similarly for 𝑣𝑖,𝑘. Then 𝑤𝑣 in Bruhat order if and only if 𝑤𝑖,𝑘𝑣𝑖,𝑘,𝑘𝐷𝑅(𝑤),1𝑖𝑘.(4.33)

For example, suppose 𝑤=368475912 and 𝑣=694287531. Since 𝐷𝑅(𝑤)={3,5,7}, we examine the three increasing rearrangements of initial segments of 𝑤 and 𝑣 of lengths 3, 5, and 7, respectively, which we may organize into Young tableaux: 254235.fig.0010(4.34)

Comparing corresponding entries, there are two violations of the tableau condition of the proposition (3>2) in the upper-left corner, so we conclude that 𝑤̸<𝑣.

Now we observe that some Bruhat relations never arise.

Lemma 4.27. Let 𝑤, 𝑤Hess()𝑆1. Let 𝒜(𝑤)=[𝑎1,𝑎2][𝑎𝑚1,𝑎𝑚] be the associated subset of 𝑤 with its decomposition into maximal consecutive substrings. Suppose one of the following conditions hold:(1)𝑤 is of Peterson type that does not contain 321 while 𝑤 is not,(2)𝑤 is 231-type while 𝑤 is either of Peterson type that contains 321 or is 312-type. Then 𝑤̸<𝑤 and ̸𝑟𝑜(𝑤)<𝑤.

Proof. If 𝑤 is of Peterson type that does not contain 321, then ̸{1,2}𝒜(𝑤) by definition of the associated subsets. All other types (Peterson type that contains 321, or 312-type, or 231-type) have associated subsets containing {1,2} by Lemma 4.13 and by definition of 𝒜(𝑤). The claim (1) now follows from Lemma 4.25.
Next suppose 𝑤 is 231-type and 𝑤 is of Peterson type that contains 321. Then the first two entries of the one-line notation of 𝑤 must be both strictly greater than 1, and 2𝐷𝑅(𝑤). Similarly if 𝑤 is a 312-type fixed point, then 𝑎22. From (4.13) it follows that the first two entries in the one-line notation of 𝑤 are also strictly greater than 1, and 2𝐷𝑅(𝑤). On the other hand, the one-line notation for a 231-type fixed point, in (4.14) has a 1 in the second entry. By the tableau criterion, if 𝑤<𝑤, then, since 2𝐷𝑅(𝑤) in both cases under consideration, we must have that one of the first two entries of 𝑤 is equal to 1, but we have just seen that is impossible. Hence 𝑤̸<𝑤. The assertion that ̸𝑟𝑜(𝑤)<𝑤 follows by a similar argument using (4.29), (4.30), and (4.37).

For the next lemma and below, we say two fixed points are of the same type if both are Peterson-type, or both are 312-type, or both are 231-type.

Lemma 4.28. Let 𝑤, 𝑤Hess()𝑆1. Suppose one of the following conditions hold:(i)𝑤 and 𝑤 are of the same type, or(ii)𝑤 is of Peterson type and does not contain 321, and 𝑤 is either 312-type or 231-type, or(iii)𝑤 is either 312-type or 231-type, and 𝑤 is of Peterson type.Then, 𝑤<𝑤i𝒜𝑤(𝑤)𝒜.(4.35)

Proof. Since the lemma above shows that 𝑤<𝑤 implies 𝒜(𝑤)𝒜(𝑤), for all cases it suffices to show the reverse implication. First suppose 𝑤 and 𝑤 are of the same type and 𝒜(𝑤)𝒜(𝑤). An examination of the reduced word decompositions of the 334-type fillings given in the above discussion and an argument similar to that in [13] implies 𝑤<𝑤. Now suppose 𝑤 is of Peterson type and does not contain 321 and 𝑤 is either of 312 type or 231-type. Then since ̸{1,2}𝒜(𝑤), either 1𝒜(𝑤) or 2𝒜(𝑤). From the explicit reduced word decompositions of 312 or 231-type fixed points chosen above it can be seen that 𝑤 is Bruhat-greater than both 𝑤𝒜(𝑤){1} and 𝑤𝒜(𝑤){2}. The claim now follows from Lemma 4.23. Finally suppose 𝑤 is either 312-type or 231-type and 𝑤 is of Peterson type. Since 𝒜(𝑤)𝒜(𝑤) we know from Lemma 4.23 that 𝑤𝒜(𝑤)<𝑤𝒜(𝑤)=𝑤. Lemma 4.24 shows that 𝑤<𝑤𝒜(𝑤) so the result follows.

The next step is to show that Bruhat relations between certain Hessenberg fixed points are connected to lengths of initial maximal consecutive substrings in the associated subsets. We need some notation. Let 𝒜{1,2,,𝑛1}. Recall we denote by 𝑤𝒜 the Peterson-type fixed point associated to 𝒜. For the purposes of this discussion we let 𝑢𝒜 (resp., 𝑣𝒜) denote the 312-type (resp., 231-type) fixed point with associated subset 𝒜. Thus for 𝒜=[1,𝑎] for some 𝑎 with 2𝑎𝑛1, we have𝑢[1,𝑎]=(𝑎+1𝑎312𝑎+2𝑎+3𝑛)1=𝑎𝑎+1𝑎1𝑎221𝑎+2𝑎+3𝑛,(4.36)𝑣[1,𝑎]=(2𝑎+1𝑎431𝑎+2𝑎+3𝑛)1=𝑎+11𝑎𝑎132𝑎+2𝑎+3𝑛(4.37) in one-line notation. For general subsets𝑎𝒜=1,𝑎2𝑎3,𝑎4𝑎𝑚1,𝑎𝑚,(4.38) with 𝑎1=1 and 𝑎22, the definitions 312-type and 231-type fixed points imply that𝑢𝒜=𝑢[𝑎1,𝑎2]𝑤[𝑎3,𝑎4]𝑤[𝑎𝑚1,𝑎𝑚],(4.39)𝑣𝒜=𝑣[𝑎1,𝑎2]𝑤[𝑎3,𝑎4]𝑤[𝑎𝑚1,𝑎𝑚].(4.40)

Lemma 4.29. Let 𝒜, be subsets of {1,2,,𝑛1} and let 𝑎𝒜=1,𝑎2𝑎3,𝑎4𝑎𝑚1,𝑎𝑚𝑏,=1,𝑏2𝑏3,𝑏4𝑏𝑚1,𝑏𝑚(4.41) be the respective decompositions into maximal consecutive substrings. Assume both 𝒜 and contain {1,2}. Let 𝑤𝒜 (resp., 𝑣𝒜) be the Peterson-type (resp., 231-type) fixed point corresponding to 𝒜 and let 𝑢 be the 312-type fixed point corresponding to . Then 𝑤𝒜<𝑢resp.,𝑣𝒜<𝑢i𝒜,𝑏2𝑎2+1.(4.42)

Proof. We begin by recalling two basic observations about Bruhat order in 𝑆𝑛. Both follow straightforwardly from its definition in terms of reduced word decompositions. Suppose 𝑤,𝑤𝑆𝑛 and assume that 𝑤 and 𝑤 do not share any simple transpositions in their reduced word decompositions, that is, 𝑠𝑖<𝑤 implies 𝑠𝑖̸<𝑤 and vice versa. Then firstly, 𝑤𝑤<𝑤 for 𝑤𝑆𝑛 if and only if both 𝑤<𝑤 and 𝑤<𝑤. Secondly, 𝑤<𝑤𝑤 if and only if 𝑤<𝑤.
Recall that 𝑤𝒜 can be written as 𝑤𝒜=𝑤[𝑎1,𝑎2]𝑤[𝑎3,𝑎4]𝑤[𝑎𝑚1,𝑎𝑚].(4.43) Moreover, each factor appearing in the decomposition (4.43) (resp. (4.40) and (4.39)) for 𝑤𝒜 (resp., 𝑣𝒜 and 𝑢) has the property that it does not share any simple transpositions with any other factor appearing in the decomposition.
Now suppose 𝑣𝒜 (resp., 𝑤𝒜) is Bruhat-less than 𝑢. Then we know from Lemma 4.25 that 𝒜 so it suffices to prove 𝑏2𝑎2+1. From Lemma 4.9 and the definition of the Peterson type fixed points we know that the one-line notation for 𝑣𝒜 (resp., 𝑤𝒜) has first 𝑎2+1 entries 𝑎2+11𝑎2𝑎2132(4.44) (resp., 𝑎2+1𝑎2𝑎21321) while the one-line notation of 𝑢 has first 𝑏2+1 entries given by 𝑏2𝑏2+1𝑏21321.(4.45) In particular 1𝐷𝑅(𝑣𝒜) and also 1𝐷𝑅(𝑤𝒜). By the tableau criterion, this implies that the first entry of the one-line notation of 𝑣𝒜 and 𝑤𝒜 must be less than or equal to the first entry of that of 𝑢. Hence 𝑎2+1𝑏2 as desired.
Conversely suppose 𝒜 and 𝑏2𝑎2+1. Then an examination of the one-line notation of 𝑣[𝑎1,𝑎2] (resp., 𝑤[𝑎1,𝑎2]) compared to that of 𝑢[𝑏1,𝑏2] and another application of the tableau criterion implies that 𝑣[𝑎1,𝑎2]<𝑢[𝑏1,𝑏2] and 𝑤[𝑎1,𝑎2]<𝑢[𝑏1,𝑏2]. In particular 𝑣[𝑎1,𝑎2] and 𝑤[𝑎1,𝑎2] are also Bruhat-less than 𝑢. Moreover, since 𝒜 it follows that [𝑎3,𝑎4][𝑎𝑚1,𝑎𝑚] so Lemma 4.23 implies 𝑤=𝑤[𝑎3,𝑎4]𝑤[𝑎𝑚1,𝑎𝑚]<𝑤=𝑢𝑠1, where the last equality follows from Lemma 4.9. Since 𝑠1 does not appear in any factor of 𝑤 the general fact above implies 𝑤<𝑢. Finally since neither 𝑤[𝑎1,𝑎2] nor 𝑣[𝑎1,𝑎2] share any simple transpositions with 𝑤, the other general fact above yields 𝑣𝒜<𝑢, 𝑤𝒜<𝑢 as desired.

4.4. Proof of Proposition 4.3

We may now prove the upper-triangular vanishing property of 334-type Hessenberg Schubert classes.

Proof of Proposition 4.3. Let 𝑤,𝑤Hess()𝑆1 and let 𝒜𝑎(𝑤)=1,𝑎2𝑎3,𝑎4𝑎𝑚1,𝑎𝑚,𝒜𝑤=𝑎1,𝑎2𝑎3,𝑎4𝑎𝑟1,𝑎𝑟(4.46) be the respective associated subsets decomposed into maximal consecutive substrings. By Lemmas 3.5 and 4.4 it suffices to prove that if 𝑟𝑜(𝑤)𝑤, then 𝑤𝑤. So suppose 𝑟𝑜(𝑤)𝑤. By Lemma 4.25 this implies 𝒜(𝑤)𝒜(𝑤). By Lemma 4.28 we can conclude 𝑤𝑤 if one of the following holds:(i)𝑤 and 𝑤 are of the same type, or(ii)𝑤 is of Peterson type that does not contain 321, and 𝑤 is either 312-type or 231-type, or(iii)𝑤 is either 312-type or 231-type, and 𝑤 is of Peterson type.Now suppose one of the following holds:(i)𝑤 is of Peterson type that does not contain 321 and 𝑤 is not,(ii)𝑤 is 231-type and 𝑤 is of Peterson type that contains 321, or(iii)𝑤 is 231-type and 𝑤 is 312-type.In these cases, Lemma 4.27 implies that ̸𝑟𝑜(𝑤)<𝑤 so there is nothing to prove.
It remains to discuss the cases when(i)𝑤 is of Peterson type that contains 321 and 𝑤 is 312-type, or(ii)𝑤 is 231-type and 𝑤 is 312-type.By Lemma 4.29 it suffices to show that 𝑎2𝑎2+1. Suppose 𝑤 is of Peterson type that contains 321. In particular 𝑎22. From (4.29) we know that 1𝐷𝑅(𝑟𝑜(𝑤)) and the first entry in the one-line notation of 𝑟𝑜(𝑤) is 𝑎2+1. On the other hand (4.36) implies the one-line notation of 𝑤 begins with 𝑎2. So if 𝑟𝑜(𝑤)<𝑤, then the tableau criterion implies 𝑎2𝑎2+1 as desired. Now suppose 𝑤 is of 231-type. Again 𝑎22 and from (4.31) we know 𝑟𝑜(𝑤) has 1𝐷𝑅(𝑟𝑜(𝑤)) and 𝑎2+1 as its first entry. By the same argument, 𝑎2𝑎2+1 as desired. The result follows.

5. Combinatorial Formulae for Restrictions to Fixed Points of 334-Type Hessenberg Schubert Classes

Our goal in this section is to give a combinatorial formula for 𝑝𝑟𝑜(𝑤)(𝑤) from which it follows as a corollary that it is nonzero. This proves Proposition 4.2 and hence Theorem 4.1. Although not strictly necessary for the proof of Proposition 4.2, we choose to prove the explicit formula (Proposition 5.8 below) since such a formula is a first step towards a derivation of a Monk formula for 334-type Hessenberg varieties and because it conveys a flavor of the combinatorics embedded in the GKM theory of Hessenberg varieties which are larger than the Peterson varieties in [13]. Many of our computations are analogues of those in [13, Section 5]. Our main tool is Billey’s formula. We briefly recall some definitions and results (see also discussion in [13, Section 4]).

Definition 5.1 (see [13, Definition  4.7]). Given a permutation 𝑤𝑆𝑛, an index 𝑗{1,2,,(𝑤)}, and a choice of reduced word decomposition 𝐛=(𝑏1,𝑏2,,𝑏(𝑤)) (corresponding to the word 𝑤=𝑠𝑏1𝑠𝑏2𝑠𝑏(𝑤)) for 𝑤, define 𝑟(𝑗,𝐛)=𝑠𝑏1𝑠𝑏2𝑠𝑏𝑗1𝑡𝑏𝑗𝑡𝑏𝑗+1.(5.1)
From the definition it follows that 𝑟(𝑗,𝐛) is an element of 𝐻𝑇(pt)Sym(𝑡)[𝑡1,𝑡2,,𝑡𝑛] of the form 𝑡𝑡𝑘 for some , 𝑘. These elements 𝑟(𝑗,𝐛) are the building blocks of Billey’s formula [18, Theorem  4] which computes the restrictions 𝜎𝑣(𝑤) of equivariant Schubert classes 𝜎𝑣 at arbitrary permutations 𝑤 in 𝑆𝑛.

Theorem 5.2 (Billey’s formula, [18, Theorem  4]). Let 𝑤𝑆𝑛. Fix a reduced word decomposition 𝑤=𝑠𝑏1𝑠𝑏2𝑠𝑏(𝑤) and let 𝐛=(𝑏1,𝑏2,,𝑏(𝑤)) be the sequence of its indices. Let 𝑣𝑆𝑛. Then the restriction 𝜎𝑣(𝑤) of the Schubert class 𝜎𝑣 at the 𝑇-fixed point 𝑤 is given by 𝜎𝑣(𝑟𝑗𝑤)=1𝑟𝑗,𝐛2𝑗,𝐛𝑟(𝑣),,𝐛(5.2) where the sum is taken over subwords 𝑠𝑏𝑗1𝑠𝑏𝑗2𝑠𝑏𝑗(𝑣) of 𝐛 that are reduced words for 𝑣.

We record the following fact, used in the proof below, which follows straightforwardly from the Billey formula.

Fact 2. Suppose 𝑣, 𝑤𝑆𝑛 with 𝑣𝑤 in Bruhat order. Suppose there exists a decomposition 𝑤=𝑤𝑤 for 𝑤, 𝑤𝑆𝑛 where 𝑣𝑤 and, for all simple transpositions 𝑠𝑖 such that 𝑠𝑖<𝑣, we have 𝑠𝑖𝑤. Then 𝜎𝑣(𝑤)=𝜎𝑣(𝑤).

Following terminology in [13], we refer to an individual summand of the expression in the right hand side of (5.2), corresponding to a single reduced subword 𝑣=𝑠𝑏𝑗1𝑠𝑏𝑗2𝑠𝑏𝑗(𝑣) of 𝑤, as a summand in Billey’s formula. In order to derive formulas for 𝑝𝑣(𝑤) where 𝑝𝑣 is a Hessenberg Schubert class, we use the linear projection 𝜋𝑆1𝑡Lie(𝑆1) dual to the inclusion of our circle subgroup 𝑆1 into 𝑇 given by (2.3). More specifically, since the diagram (4.6) commutes, we have𝑝𝑣(𝜋𝑤)=𝑆1𝑟𝑗1𝜋,𝐛𝑆1𝑟𝑗2,𝐛𝜋𝑆1𝑟𝑗(𝑣).,𝐛(5.3) We refer to the right hand side of the above equality as Billey’s formula for 𝑝𝑣(𝑤). Recall 𝜋𝑆1(𝑡𝑡𝑘+1)=(𝑘+1)𝑡 for a positive root 𝑡𝑡𝑘+1 [13, Section 5].

We also use the following.

Definition 5.3 (see [13, Definition 5.4]). Fix 𝒜{1,2,,𝑛1}. Define 𝒜𝒜𝒜 by 𝒜(𝑗)=themaximalelementinthemaximalconsecutivesubstringof𝒜containing𝑗.(5.4)

Definition 5.4 (see [13, Definition 5.5]). Fix 𝒜{1,2,,𝑛1}. Define 𝒯𝒜𝒜𝒜 by 𝒯𝒜(𝑗)=theminimalelementinthemaximalconsecutivesubstringof𝒜containing𝑗.(5.5)

We proceed to some preliminary computations. Let 𝐛=(𝑏1,,𝑏(𝑤)) be a reduced word decomposition 𝑤=𝑠𝑏1𝑠𝑏2𝑠𝑏(𝑤) of 𝑤 and let 𝑖 be an index appearing in 𝐛, that is, 𝑏=𝑖 for some 1(𝑤). Our first computation, Lemma 5.5, gives an expression for 𝜋𝑆1(𝑟(,𝐛)) which shows in particular that the value of 𝜋𝑆1(𝑟(,𝐛)) depends only on the value of the index 𝑏=𝑖 and not on its location in the word 𝐛. Note that if 𝑣=𝑠𝑖, then the summands in Billey’s formula for 𝑝𝑣(𝑤)=𝑝𝑠𝑖(𝑤) are precisely equal to 𝑟(,𝑏) for each such that 𝑏=𝑖. Thus an equivalent formulation of the claim is that the summands in Billey’s formula for 𝑝𝑠𝑖(𝑤) are all equal. This is analogous to a result in the Peterson case [13, Lemma  5.2] except that, in our situation, the form of the formulas depend on the index 𝑖 as well as on the type of the fixed point 𝑤 in question.

Lemma 5.5. Let 𝑤Hess()𝑆1 and let 𝐛=(𝑏1,,𝑏(𝑤)) be the reduced word decomposition of 𝑤 chosen in Section 4. Let 𝒜(𝑤)=[𝑎1,𝑎2][𝑎3,𝑎4][𝑎𝑚1,𝑎𝑚] be the associated subset of 𝑤 decomposed into maximal consecutive substrings. Let 𝑖{1,2,,𝑛1}.(1)If 𝑖𝒜(𝑤), then each summand in Billey’s formula for 𝑝𝑠𝑖(𝑤) is 0. In particular, 𝑝𝑠𝑖(𝑤)=0.(2)Suppose 𝑖𝒜(𝑤) and suppose one of the following conditions hold:(a)𝑤 is of Peterson type, or(b)𝑤 is 312-type, or(c)𝑤 is 231-type and 𝑖[𝑎1,𝑎2]. Then each summand in Billey’s formula for 𝑝𝑠𝑖(𝑤) is equal to 𝑖𝒯𝒜(𝑤)(𝑖)+1𝑡.(5.6)(3)Suppose 𝑤 is 231-type and 𝑖=1. Then each summand in Billey’s formula for 𝑝𝑠𝑖(𝑤) is equal to 𝑎2𝑡=𝒜(𝑤)(1)𝑡.(5.7)(4)Suppose 𝑤 if 231-type and 𝑖[2,𝑎2]. Then each summand in Billey’s formula for 𝑝𝑠𝑖(𝑤) is equal to 𝑖𝒯𝒜(𝑤)(𝑖)𝑡=(𝑖1)𝑡.(5.8)

Proof. If 𝑖 does not occur in 𝒜(𝑤), then each summand is 0 by Billey’s formula for 𝜎𝑠𝑖(𝑤), since 𝑠𝑖̸<𝑤 and thus never appears in the reduced word decomposition of 𝑤. For the next claim, the fact that each summand is equal to (𝑖𝒯𝒜(𝑤)(𝑖)+1)𝑡 for the listed cases follows from examination of the chosen reduced word decompositions of 𝑤 and an argument identical to that in [13]. Thus it remains to check the cases in which the summand differs from the case of Peterson varieties. First suppose 𝑤 is 231-type and that 𝑖=1. From the choice of explicit reduced word decomposition for such 𝑤 given in (4.22) and Billey’s formula, it follows that each summand in Billey’s formula for 𝜎𝑠1(𝑤) is equal to 𝑟2𝑟3𝑟2𝑟𝑎21𝑟𝑎22𝑟3𝑟2𝑟𝑎2𝑟𝑎21𝑟2𝑡1𝑡2=𝑟2𝑟3𝑟2𝑟𝑎21𝑟𝑎22𝑟3𝑟2𝑡1𝑡𝑎2+1=𝑡1𝑡𝑎2+1(5.9) since the reflection 𝑟𝑗 switches 𝑡𝑗 and 𝑡𝑗+1. Hence we have 𝑝𝑠1(𝑤)=𝜋𝑆1(𝑡1𝑡𝑎2+1)=(𝑎2+11)𝑡=𝑎2𝑡=𝒜(𝑤)(1)𝑡. Now suppose 𝑤 is 231-type and 𝑖[2,𝑎2]. The factor in the reduced word decomposition (4.22) corresponding to [1=𝑎1,𝑎2] is equal to 𝑤[2,𝑎2]𝑠1. By Fact 2, for 𝑖>1 the presence of the extra 𝑠1 does not affect the Billey computation, so each summand is equal to that for the Peterson type fixed point 𝑤[2,𝑎2] and hence is equal to 𝑖𝒯𝒜(𝑤)=(𝑖)(𝑖1)𝑡,(5.10) as desired.

Our next lemma concerns the summands in Billey’s formula for 𝑝𝑟𝑜(𝑤)(𝑤) for 𝑤Hess()𝑆1.

Lemma 5.6. Let 𝑤Hess()𝑆1.(i)Suppose 𝑤 is 312-type or 𝑤 is of Peterson type that does not contain 321. Then each summand for Billey’s formula for 𝑝𝑟𝑜(𝑤)(𝑤) is equal to 𝑖𝒜(𝑤)𝑖𝒯𝒜(𝑤)(𝑖)+1𝑡|𝒜(𝑤)|.(5.11)(ii)Suppose 𝑤 is 231-type. Then each summand for Billey’s formula for 𝑝𝑟𝑜(𝑤)(𝑤) is equal to 𝒜(𝑤)(1)𝒜(𝑤)(1)𝑖=2(𝑖1)𝑖𝒜(𝑤)[𝒯𝒜(𝑤)(1),𝒜(𝑤)(1)]𝑖𝒯𝒜(𝑤)𝑡(𝑖)+1|𝒜(𝑤)|.(5.12)(iii)Suppose 𝑤 is of Peterson type that contains 321. Then each summand for Billey’s formula for 𝑝𝑟𝑜(𝑤)(𝑤) is equal to 𝑖𝒜(𝑤)𝑖𝒯𝒜(𝑤)(𝑖)+1𝑡|𝒜(𝑤)|+1.(5.13)

Proof. Before considering the separate cases we make a general observation. By Lemma 5.5 and the discussion before Lemma 5.5 we know that the summands in Billey’s formula for 𝑝𝑠𝑖(𝑤) for 𝑖𝒜(𝑤) are exactly the terms 𝜋𝑆1(𝑟(,𝐛)) for such that 𝑏=𝑖. Suppose in addition that 𝑤Hess()𝑆1 is such that 𝑟𝑜(𝑤) contains at most one simple transposition 𝑠𝑖 for each 𝑖{1,2,,𝑛1}, that is, 𝑟𝑜(𝑤)=𝑠𝑖1𝑠𝑖2𝑠𝑖(𝑟𝑜(𝑤)) is a reduced word for 𝑟𝑜(𝑤) where all 𝑖𝑘 are distinct for 1𝑘(𝑟𝑜(𝑤)). This implies that any subword of a reduced word decomposition 𝐛 of 𝑤 which is a reduced word of 𝑟𝑜(𝑤) also must contain precisely one 𝑠𝑖𝑘 for each 1𝑘(𝑟𝑜(𝑤)). From Billey’s formula (5.3) for 𝑝𝑟𝑜(𝑤)(𝑤) we know that a summand is of the form 𝜋𝑆1𝑟𝑗1𝜋,𝐛𝑆1𝑟𝑗2,𝐛𝜋𝑆1𝑟𝑗(𝑣),,𝐛(5.14) where 𝑠𝑏𝑗1𝑠𝑏𝑗2𝑠𝑏𝑗(𝑟𝑜(𝑤)) is a reduced word of ro(𝑤). Since{𝑏𝑗1,,𝑏𝑗(𝑟𝑜(𝑤))}={𝑖1,𝑖2,,𝑖(𝑟𝑜(𝑤))} for each such summand, the quantity (5.14) is equal to (𝑟𝑜(𝑤))𝑘=1𝑝𝑠𝑖𝑘(𝑤).(5.15)
We now take cases. Suppose 𝑤 is not a Peterson-type that contains 321. Then from the explicit descriptions of 𝑟𝑜(𝑤) given in Section 4 it follows that 𝑟𝑜(𝑤) contains in its reduced word a single 𝑠𝑖 for each 𝑖𝒜(𝑤). Thus we are in the situation described in the above paragraph and the claims follow from the computations given in Lemma 5.5.
Suppose 𝑤 is Peterson type and contains 321. Let 𝐛 be the standard reduced word decomposition (cf. (4.17) and (4.19)) of 𝑤. We claim that the only reduced word decompositions of 𝑟𝑜(𝑤) that occur as a subword of 𝐛 are those which contain two 𝑠1’s, only one 𝑠2, and precisely one 𝑠𝑗 for all other 𝑗. Indeed let 𝒜(𝑤)=[𝑎1,𝑎2][𝑎3,𝑎4][𝑎𝑚1,𝑎𝑚] be the decomposition of 𝒜(𝑤) into maximal consecutive substrings. Recall 𝑎22 and 𝑎1=1 in this case. The rolldown 𝑟𝑜(𝑤) is 𝑠𝑎𝑚𝑠𝑎𝑚1𝑠𝑎𝑚1𝑠𝑎4𝑠𝑎41𝑠𝑎3𝑠𝑎2𝑠𝑎21𝑠1𝑠2𝑠1(5.16) and 𝑤 is 𝑤=𝑤[𝑎1,𝑎2]𝑤[𝑎3,𝑎4]𝑤[𝑎𝑚1,𝑎𝑚].(5.17) Let >1. There is only one reduced word decomposition of the factor 𝑠𝑎+1𝑠𝑎+11𝑠𝑎 in 𝑟𝑜(𝑤), so it remains to analyze the subwords of 𝑤[𝑎1,𝑎2] which are reduced words of 𝑠𝑎2𝑠𝑎21𝑠1𝑠2𝑠1. Let 𝐛 denote the standard reduced word of 𝑤[𝑎1,𝑎2]. Note that another valid reduced word of 𝑠𝑎2𝑠𝑎21𝑠1𝑠2𝑠1 is 𝑠𝑎2𝑠𝑎21𝑠2𝑠1𝑠2. Since 𝑠2 does not commute with 𝑠1, the rightmost 𝑠2 in the word 𝑠𝑎2𝑠𝑎21𝑠2𝑠1𝑠2 must appear to the right of the 𝑠𝑎2; in particular, there are two 𝑠2’s to the right of the 𝑠𝑎2 in this word. Since there is only one 𝑠2 appearing to the right of the 𝑠𝑎2 in 𝐛, we conclude that the reduced word of 𝑠𝑎2𝑠𝑎21𝑠1𝑠2𝑠1 containing two copies of 𝑠2 never appears as a subword of 𝐛. Hence the only subwords of 𝐛 contributing to summands in Billey’s formula for 𝑝𝑟𝑜(𝑤)(𝑤) contain two 𝑠1’s and one 𝑠2, as claimed. Now since 𝑗1=1 and 𝑗2=2 and 𝑝𝑠1(𝑤)=(1𝒯𝒜(𝑤)(1)+1)𝑡=(11+1)𝑡=𝑡 by Lemma 5.5, the claim follows.

We have just seen that all summands in Billey’s formula for 𝑝𝑟𝑜(𝑤)(𝑤) are equal for all fixed points 𝑤Hess()𝑆1. In order to finish the computation we must now compute the number of summands which occur.

Lemma 5.7. Let 𝑤Hess()𝑆1.(i)Suppose 𝑤 is of Peterson type that contains 321. Then the number of summands in Billey’s formula for 𝑝𝑟𝑜(𝑤)(𝑤) is 𝒜(𝑤)(1)1.(ii)Suppose 𝑤 is of Peterson type that does not contain 321. Then the number of summands in Billey’s formula for 𝑝𝑟𝑜(𝑤)(𝑤) is 1.(iii)Suppose 𝑤 is 312-type. Then the number of summands in Billey’s formula for 𝑝𝑟𝑜(𝑤)(𝑤) is 𝒜(𝑤)(1)1.(iv)Suppose 𝑤 is 231-type. Then the number of summands in Billey’s formula for 𝑝𝑟𝑜(𝑤)(𝑤) is 1.

Proof. We consider each case in turn. Suppose 𝑤 is of Peterson type that contains 321. Let 𝒜(𝑤)=[𝑎1,𝑎2][𝑎3,𝑎4][𝑎𝑚1,𝑎𝑚] be the decomposition of 𝒜(𝑤) into maximal consecutive substrings. Recall 𝑎22 and 𝑎1=1 in this case. The rolldown 𝑟𝑜(𝑤) is 𝑠𝑎𝑚𝑠𝑎𝑚1𝑠𝑎𝑚1𝑠𝑎4𝑠𝑎41𝑠𝑎3𝑠𝑎2𝑠𝑎21𝑠1𝑠2𝑠1(5.18) and 𝑤 is 𝑤=𝑤[𝑎1,𝑎2]𝑤[𝑎3,𝑎4]𝑤[𝑎𝑚1,𝑎𝑚].(5.19) Let >1. As observed in the proof of Lemma 5.6, there is only one reduced word decomposition of the factor 𝑠𝑎+1𝑠𝑎+11𝑠𝑎 in 𝑟𝑜(𝑤). Moreover, by examination it is evident that it appears only once in the standard reduced word decomposition of the corresponding 𝑤[𝑎,𝑎+1] factor in 𝑤. Hence in order to count the number of ways 𝑟𝑜(𝑤) appears in 𝑤 it suffices to count the number of subwords of the standard reduced word decomposition 𝐛 of 𝑤[𝑎1,𝑎2] which are reduced subwords of 𝑠𝑎2𝑠𝑎21𝑠1𝑠2𝑠1. We already saw in the proof of Lemma 5.6 that the reduced word 𝑠𝑎2𝑠𝑎21𝑠2𝑠1𝑠2 never appears in 𝐛. On the other hand, since 𝑠1 commutes with any 𝑠𝑘 with 𝑘3, another reduced word decomposition of 𝑠𝑎2𝑠𝑎21𝑠1𝑠2𝑠1 is 𝑠1𝑠𝑎2𝑠𝑎21𝑠2𝑠1. From examination of b it can be seen that the word 𝑠1𝑠𝑎2𝑠𝑎21𝑠2𝑠1 appears as a subword in the standard reduced word of 𝑤[𝑎1,𝑎2] precisely 𝑎21=𝒜(𝑤)(1)1 times and that these are the only subwords of 𝐛 which equal 𝑠𝑎2𝑠𝑎21𝑠1𝑠2𝑠1. The claim follows.
Suppose 𝑤 is of Peterson type that does not contain 321. Then the rolldown 𝑟𝑜(𝑤) is the Peterson case rolldown so the claim follows from explicit examination of the standard reduced word of 𝑤 (alternatively from [13, Fact  4.5]).
Suppose 𝑤 is 312-type. Then the rolldown 𝑟𝑜(𝑤) is of the form 𝑠𝑟𝑜(𝑤)=𝑎𝑚𝑠𝑎𝑚1𝑠𝑎𝑚1𝑠𝑎4𝑠𝑎41𝑠𝑎3𝑠𝑎2𝑠𝑎21𝑠1𝑠2(5.20) from Lemma 4.20. By an argument similar to the case of Peterson type that contains 321 it suffices to analyze only the factors in both 𝑟𝑜(𝑤) and 𝑤 corresponding to the initial maximal consecutive substring [𝑎1,𝑎2]. As above we have 𝑠𝑎2𝑠𝑎21𝑠1𝑠2=𝑠1𝑠𝑎2𝑠𝑎21𝑠2(5.21) and again it follows from examination of the standard reduced word of 𝑢[𝑎1,𝑎2] that 𝑠1𝑠𝑎2𝑠𝑎21𝑠2 appears precisely 𝑎21=𝒜(𝑤)(1)1 times.
Finally suppose 𝑤 is 231-type. Then the rolldown 𝑟𝑜(𝑤) coincides with the Peterson case rolldown of 𝑤𝒜(𝑤) and the claim follows from examination of the reduced word decomposition (4.22).

The following is immediate from Lemmas 5.6 and 5.7.

Proposition 5.8. Let 𝑤Hess()𝑆1.(i)Suppose 𝑤 is of Peterson type that contains 321. Then 𝑝𝑟𝑜(𝑤)(𝑤)=𝒜(𝑤)(1)1𝑖𝒜(𝑤)𝑖𝒯𝒜(𝑤)(𝑖)+1𝑡|𝒜(𝑤)|+1.(5.22)(ii)Suppose 𝑤 is of Peterson type that does not contain 321. Then 𝑝𝑟𝑜(𝑤)(𝑤)=𝑖𝒜(𝑤)𝑖𝒯𝒜(𝑤)𝑡(𝑖)+1|𝒜(𝑤)|.(5.23)(iii)Suppose 𝑤 is of type 312. Then, 𝑝𝑟𝑜(𝑤)(𝑤)=𝒜(𝑤)(1)1𝑖𝒜(𝑤)𝑖𝒯𝒜(𝑤)(𝑖)+1𝑡|𝒜(𝑤)|.(5.24)(iv)Suppose 𝑤 is of type 231. Then,𝑝𝑟𝑜(𝑤)(𝑤)=𝒜(𝑤)(1)𝒜(𝑤)(1)𝑖=2(𝑖1)𝑖𝒜(𝑤)[𝒯𝒜(𝑤)(1),𝒜(𝑤)(1)]×𝑖𝒯𝒜(𝑤)𝑡(𝑖)+1|𝒜(𝑤)|.(5.25)

The proofs of the main results are now immediate.

Proof of Proposition 4.2. Let 𝑤Hess()𝑆1. From the explicit formulas given in Proposition 5.8 it follows that 𝑝𝑟𝑜(𝑤)(𝑤)0 for all possible types of fixed points 𝑤.

Proof of Theorem 4.1. Since both (4.3) and (4.4) are satisfied for all 𝑤, 𝑤Hess()𝑆1 by Propositions 4.3 and 4.2, respectively, the result follows.

6. Open Questions

This manuscript raises more questions than it answers. We close by mentioning some of them.

Question 1. For 𝑛4, Theorem 4.1 shows that for the case when 𝑁𝑛𝑛 is the principal nilpotent operator and is the 334-type Hessenberg function, the dimension pair algorithm produces a set of Hessenberg Schubert classes {𝑝𝑟𝑜(𝑤)}𝑤Hess(𝑁,)𝑆1 which are poset-upper-triangular and hence form a 𝐻𝑆1(pt)-module basis for 𝐻𝑆1(Hess(𝑁,)).(1)What are other examples of 𝑁 and such that the conclusion of Lemma 3.4 holds (cf. Remark 3.8)?(2)What are other examples of 𝑁 and for which the dimension pair algorithm produces a successful outcome of Betti poset pinball which is also poset-upper-triangular? Are there necessary and sufficient conditions on 𝑁 and that guarantee poset-upper-triangularity?(3)What are other examples of 𝑁 and for which the dimension pair algorithm produces a successful outcome of Betti poset pinball which corresponds to a linearly independent set of classes and hence a module basis? Are there necessary and sufficient conditions on 𝑁 and that guarantee this?

Question 2. In [13] the explicit module basis consisting of Peterson Schubert classes is used to derive a manifestly positive Monk formula in the 𝑆1-equivariant cohomology of Peterson varieties. Preliminary investigation suggests that an analogous Monk formula for the 334-type Hessenberg varieties, using the module basis of Hessenberg Schubert classes derived in this manuscript, would be computationally much more complex. Thus, we may ask the following.(1)Does there exist a combinatorially elegant or computationally effective Monk formula for the 334-type Hessenberg varieties?(2)Can such a Monk formula be further generalized to a larger family of regular nilpotent Hessenberg varieties? For instance, can our techniques be generalized to give new insights to the equivariant Schubert calculus of the full flag variety 𝑎𝑔𝑠(𝑛) (which is an example of a regular nilpotent Hessenberg variety)?(3)In [12] the Monk formula for Peterson varieties is used to derive a Giambelli formula. Does there also exist a combinatorially elegant and/or computationally effective Giambelli formula for other cases of regular nilpotent Hessenberg varieties?

Acknowledgment

M. Harada is partially supported by an NSERC Discovery Grant, an NSERC University Faculty Award, and an Ontario Ministry of Research and Innovation Early Researcher Award.