This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Quantum Bruhat graphs and tilted Richardson varieties

Jiyang Gao Department of Mathematics, Harvard University, Cambridge, MA 02138 [email protected] Shiliang Gao Department of Mathematics, University of Illinois Urbana-Champaign, Urbana, IL 61801 [email protected]  and  Yibo Gao Beijing International Center for Mathematical Research, Peking University, Beijing 100084 [email protected]
Abstract.

Quantum Bruhat graph is a weighted directed graph on a finite Weyl group first defined by Brenti-Fomin-Postnikov. It encodes quantum Monk’s rule and can be utilized to study the 33-point Gromov-Witten invariants of the flag variety. In this paper, we provide an explicit formula for the minimal weights between any pair of permutations on the quantum Bruhat graph, and consequently obtain an Ehresmann-like characterization for the tilted Bruhat order. Moreover, for any ordered pair of permutations uu and vv, we define the tilted Richardson variety 𝒯u,v\mathcal{T}_{u,v}, with a stratification that gives a geometric meaning to intervals in the tilted Bruhat order. We provide a few equivalent definitions to this new family of varieties that include Richardson varieties, and establish some fundamental geometric properties including their dimensions and closure relations.

1. Introduction

Hilbert’s fifteenth problem, Schubert calculus, concerns the full flag variety

Fln={0V1Vn1n|dimVi=i for i=1,,n1}\mathrm{Fl}_{n}=\{0\subset V_{1}\subset\cdots\subset V_{n-1}\subset\mathbb{C}^{n}\>|\>\dim V_{i}=i\text{ for }i=1,\ldots,n-1\}

and its Schubert decomposition Fln=wSnXw\mathrm{Fl}_{n}=\bigsqcup_{w\in S_{n}}X_{w}^{\circ}. The cohomology ring H(Fln)H^{*}(\mathrm{Fl}_{n}) has a linear basis given by the Schubert varieties {[Xw]}wSn\{[X_{w}]\}_{w\in S_{n}}. The corresponding structure constants cu,wvc_{u,w}^{v}’s, also referred to as the generalized Littlewood-Richardson numbers, are known to be nonnegative from transversal intersection. It has been a long standing open problem to find a combinatorial interpretation of them. The study of flag varieties, Schubert varieties and these structure constants is central in algebraic geometry and algebraic combinatorics.

The small quantum cohomology ring QH(Fln)QH^{*}(\mathrm{Fl}_{n}) is a deformation of the cohomology ring. It shows up naturally in theoretical physics. The structure constants of QH(Fln)QH^{*}(\mathrm{Fl}_{n}) with respect to the Schubert basis are known as the 3-point genus-0 Gromov-Witten invariants. They extend the generalized Littlewood-Richardson numbers in “quantum” direction.

The problem of multiplying Schubert classes in QH(Fln)QH^{*}(\mathrm{Fl}_{n}) can be naturally encoded via the quantum Bruhat graph, first defined by Brenti-Fomin-Postnikov [2] and utilized by Postnikov [12]. The quantum Bruhat graph can be seen as a graphical representation of the quantum Monk’s rule and enjoys very rich algebraic and combinatorial properties. In particular, the minimal degree qdq^{d} that appears in the quantum product [Xu][Xv][X_{u}]*[X^{v}] is the weight of any shortest directed path from uu to vv [12]. The quantum Bruhat graph directly gives rise to the tilted Bruhat order [2], and these are our main combinatorial objects of interest for this paper.

Main result 1: weights in the quantum Bruhat graph:

  1. (1)

    An explicit combinatorial formula for the minimal weight between any pair of permutations uu to vv (Theorem 3.3).

  2. (2)

    An Ehresmann-like characterizaiton for the tilted Bruhat order (Theorem 3.9).

We remark that Theorem 3.3 was also obtained via a combination of Postnikov’s toric Schur polynomials [11] on the quantum cohomology ring of the Grassmannian, and a geometric result by Buch-Chung-Li-Mihalcea [5]. See also [7]. Our proof is independent and combinatorial.

While weights on the quantum Bruhat graph encode important information in the quantum cohomology of the flag variety, we present a novel geometric interpretation of intervals in the tilted Bruhat order with a more classical flavor. For any u,vSnu,v\in S_{n}, we define the tilted Richardson variety 𝒯u,v\mathcal{T}_{u,v} and the open tilted Richardson variety 𝒯u,v\mathcal{T}_{u,v}^{\circ} which reduce to the well-known (open) Richardson variety if uvu\leq v in the Bruhat order.

Main result 2: definitions of the (open) tilted Richardson variety using

  1. (1)

    rank conditions (4.1);

  2. (2)

    cyclically rotated Richardson varieties in the Grassmannian (Theorem 4.6);

  3. (3)

    multi-Plücker coordinates (Theorem 4.12).

The (open) tilted Richardson varieties are our central geometric objects of study. Their geometric properties resemble those of Richardson varieties. However, since an analogue of Schubert varieties do not exist in our setting, most of the analysis requires new insights and techniques.

Main result 3: geometric properties of the (open) tilted Richardson variety:

  1. (1)

    a stratification 𝒯u,v=[x,y][u,v]𝒯x,y\mathcal{T}_{u,v}=\sqcup_{[x,y]\subset[u,v]}\mathcal{T}_{x,y}^{\circ} that relates tilted Bruhat order (Theorem 5.2);

  2. (2)

    dim𝒯u,v=dim𝒯u,v=(u,v)\dim\mathcal{T}_{u,v}=\dim\mathcal{T}_{u,v}^{\circ}=\ell(u,v) in the quantum Bruhat graph (Theorem 5.14);

  3. (3)

    the closure relation 𝒯u,v¯=𝒯u,v\overline{\mathcal{T}_{u,v}^{\circ}}=\mathcal{T}_{u,v} (Theorem 5.22).

In a sequel, we connect tilted Richardson varieties with curve neighborhoods Γd(Xu,Xv)\Gamma_{d}(X_{u},X^{v}). These are subvarieties of Fln\mathrm{Fl}_{n} introduced by Buch-Chaput-Mihalcea-Perrin in [3] (see also [4], [6], [10] and references therein). They are closely related to computations in QH(Fln)QH^{*}(\mathrm{Fl}_{n}) and QK(Fln)QK^{*}(\mathrm{Fl}_{n}).

Our paper is organized as follows. In Section 2, we introduce background information on quantum Bruhat graphs, tilted Bruhat orders, root systems, Grassmannians, flag varieties, Richardson varieties and Plücker coordinates. In Section 3, 4 and 5, we establish each aforementioned main results respectively on the combinatorics of quantum Bruhat graphs, definitions of the (open) tilted Richardson varieties, and their geometry.

2. Preliminaries

Let SnS_{n} be the symmetric group on nn elements, generated by the simple transpositions S={si:=(ii+1) for i=1,,n1}S=\{s_{i}:=(i\ i{+}1)\text{ for }i=1,\ldots,n-1\}. We typically write a permutation ww using its one-line notation w(1)w(2)w(n)w(1)w(2)\cdots w(n) or simplified as w1w2wnw_{1}w_{2}\cdots w_{n}. For wSnw\in S_{n}, let (w)\ell(w) be the Coxeter length of ww, which is the smallest \ell such that w=si1siw=s_{i_{1}}\cdots s_{i_{\ell}} is a product of \ell simple transpositions. Such an expression is called a reduced word of ww. Let R(w)R(w) be the set of reduced words of ww. Let I(w)={(i,j)|i<j,wi>wj}I(w)=\{(i,j)\>|\>i<j,w_{i}>w_{j}\} be the inversion set of ww. It is well-known that (w)\ell(w) equals the number of inversions of ww. Let T={tij:=(ij)| 1i<jn}T=\{t_{ij}:=(i\ j)\>|\>1\leq i<j\leq n\} be the set of transpositions, or equivalently, conjugates of the simple transpositions SS.

2.1. The quantum Bruhat graph and the tilted Bruhat orders

Definition 2.1.

The quantum Bruhat graph Γn\Gamma_{n} is a weighted directed graph on SnS_{n} with the following two types of edges:

{wwtij of weight 1 if (wtij)=(w)+1,wwtij of weight qij:=qiqi+1qj1 if (wtij)=(w)+12(ji),\begin{cases}w\rightarrow wt_{ij}\text{ of weight }1&\text{ if }\ell(wt_{ij})=\ell(w)+1,\\ w\rightarrow wt_{ij}\text{ of weight }q_{ij}:=q_{i}q_{i+1}\cdots q_{j-1}&\text{ if }\ell(wt_{ij})=\ell(w)+1-2(j-i),\end{cases}

where 1i<jn1\leq i<j\leq n. Write wt(wwtij)[q1,,qn1]\operatorname{wt}(w\rightarrow wt_{ij})\in\mathbb{Z}[q_{1},\ldots,q_{n-1}] for the weight.

In other words, wwtijw\rightarrow wt_{ij} if wi<wjw_{i}<w_{j} and for every i<k<ji<k<j, wk>wjw_{k}>w_{j} or w(k)<w(i)w(k)<w(i), which gives an edge in the Hasse diagram of the strong Bruhat order, or wwtijw\rightarrow wt_{ij} if wi>wjw_{i}>w_{j} and for every i<k<ji<k<j, wi>wk>wjw_{i}>w_{k}>w_{j}. We can unify these two cases together using cyclic intervals.

Definition 2.2.

For ab[n]a\neq b\in[n], the (open) cyclic interval from aa to bb is (a,b)c:={a+1,,b1}(a,b)_{c}:=\{a+1,\ldots,b-1\} if a<ba<b, and (a,b)c:={b+1,,n}{1,,a1}(a,b)_{c}:=\{b+1,\ldots,n\}\cup\{1,\ldots,a-1\} if a>ba>b. We can similarly define cyclic intervals [,][,], (,](,] and [,)[,).

Now wwtijw\rightarrow wt_{ij} in the quantum Bruhat graph if and only if wk(wi,wj)cw_{k}\in(w_{i},w_{j})_{c} for all i<k<ji<k<j. It is also clear that k(a,b)ck\in(a,b)_{c} if and only if k+1(a+1,b+1)ck+1\in(a+1,b+1)_{c}. Therefore, we have the following immediate observation on the cyclic symmetry of the quantum Bruhat graph.

Lemma 2.3.

Let τSn\tau\in S_{n} be the long cycle (123n)(123\cdots n). Then wτww\mapsto\tau w is an automorphism of the unweighted quantum Bruhat graph.

The quantum Bruhat graph for n=3n=3 is shown in Figure 1.

\bullet123123\bullet132132\bullet312312\bullet321321\bullet231231\bullet213213q2q_{2}q1q_{1}q2q_{2}q1q_{1}q2q_{2}q1q_{1}q1q2q_{1}q_{2}
Figure 1. Quantum Bruhat graph Γ3\Gamma_{3} (unlabeled edges have weight 11)

For a directed path P:w(0)w(1)w(k)P:w^{(0)}\rightarrow w^{(1)}\rightarrow\cdots\rightarrow w^{(k)} in Γn\Gamma_{n}, we say that PP has length kk, with weight i=1kwt(w(i1)w(i))\prod_{i=1}^{k}\operatorname{wt}(w^{(i-1)}\rightarrow w^{(i)}). For u,vSnu,v\in S_{n}, let (u,v)\ell(u,v) be the length of a shortest path from uu to vv. Note that it is clear that for any u,vSnu,v\in S_{n}, there exists a directed path from uu to vv, as one can always go through the identity. Postnikov [12] established some nice properties regarding weights of shortest paths. Here, we write qαq^{\alpha} for q1α1qn1αn1q_{1}^{\alpha_{1}}\cdots q_{n-1}^{\alpha_{n-1}}.

Lemma 2.4 ([12]).

For any u,vSnu,v\in S_{n}, all shortest paths from uu to vv have the same weight qd(u,v)q^{d(u,v)}. Moreover, the weight of any path from uu to vv is divisible by qd(u,v)q^{d(u,v)}. In addition, if a path from uu to vv has weight qd(u,v)q^{d(u,v)}, it must be a shortest path.

Lemma 2.4 is essentially saying that shortest length paths are precisely minimal weight paths. We remark that the last sentence from Lemma 2.4 is not explicitly written done in [12], but follows directly from the proof of Lemma 1 of [12], with much of the main content established in Lemma 6.7 of [2]. From now on, write qd(u,v)=q1d1(u,v)q2d2(u,v)qn1dn1(u,v)q^{d(u,v)}=q_{1}^{d_{1}(u,v)}q_{2}^{d_{2}(u,v)}\cdots q_{n-1}^{d_{n-1}(u,v)} for the minimal weight from uu to vv, where d(u,v)d(u,v) is a (n1)(n-1)-tuple.

Example 2.5.

Let u=321u=321 and v=213v=213. There are two shortest paths from uu to vv of length 22: 321231213321\rightarrow 231\rightarrow 213 and 321123213321\rightarrow 123\rightarrow 213. We see that both paths have the same weight q1q2q_{1}q_{2} so d(321,213)=(1,1)d(321,213)=(1,1). It is straightfroward to check that any other paths from uu to vv have weight divisible by (and not equal to) q1q2q_{1}q_{2}.

Definition 2.6 (tilted Bruhat order [2]).

For uSnu\in S_{n}, define the tilted Bruhat order DuD_{u} to be the graded partial order on SnS_{n} such that wuvw\leq_{u}v if

(1) (u,w)+(w,v)=(u,v).\ell(u,w)+\ell(w,v)=\ell(u,v).

Equivalently, wuvw\leq_{u}v if there is a shortest path in the quantum Bruhat graph from uu to vv that passes through ww.

Remark 2.7.

A special case of the tilted Bruhat order is the strong Bruhat order. This is obtained by setting u=idu=id.

The tilted Bruhat order D132D_{132} is shown in Figure 2.

\bullet132132\bullet123123\bullet231231\bullet312312\bullet213213\bullet321321
Figure 2. The tilted Bruhat order D132D_{132}
Definition 2.8.

For wuvw\leq_{u}v, define the tilted Bruhat interval to be

[w,v]u={xSn:wuxuv}.[w,v]_{u}=\{x\in S_{n}:w\leq_{u}x\leq_{u}v\}.
Remark 2.9.

It follows from (1) that [w,v]u=[w,v]u[w,v]_{u}=[w,v]_{u^{\prime}} as long as wuvw\leq_{u}v and wuvw\leq_{u^{\prime}}v. Since we always have wwvw\leq_{w}v, we will omit the subscript and write [w,v][w,v] instead of [w,v]u[w,v]_{u} in the remaining part of this paper.

2.2. Root system and reflection ordering

The root system of type An1A_{n-1} consists of Φ={eiej| 1i,jn}\Phi=\{e_{i}-e_{j}\>|\>1\leq i,\neq j\leq n\} with positive roots Φ+={eiej| 1i<jn}\Phi^{+}=\{e_{i}-e_{j}\>|\>1\leq i<j\leq n\} and simple roots Δ={αi:=eiei+1| 1in1}\Delta=\{\alpha_{i}:=e_{i}-e_{i+1}\>|\>1\leq i\leq n-1\}. It’s corresponding Weyl group is identified with the symmetric group SnS_{n}, while the reflection across the hyperplane normal to eieje_{i}-e_{j} is identified with tij=(ij)t_{ij}=(i\ j).

Definition 2.10.

An ordering of Φ+\Phi^{+} (or equivalently, of the reflections TT) as γ1,,γ\gamma_{1},\ldots,\gamma_{\ell} is called a reflection ordering if eieke_{i}-e_{k} appears (not necessarily consecutively) in the middle of eieje_{i}-e_{j} and ejeke_{j}-e_{k} for all i<j<ki<j<k.

The following lemma is very classical. See for example Proposition 3 of [1].

Lemma 2.11.

Reflection orderings are in bijection with reduced words of the longest permutation w0=nn121w_{0}=n\ n{-}1\cdots 21. Given a=a1aR(w0)a=a_{1}\cdots a_{\ell}\in R(w_{0}), its corresponding reflection ordering γ1,,γ\gamma_{1},\ldots,\gamma_{\ell} is constructed via γk=sa1saj1αaj\gamma_{k}=s_{a_{1}}\cdots s_{a_{j-1}}\alpha_{a_{j}} for k=1,,=(n2)k=1,\ldots,\ell={n\choose 2}.

Example 2.12.

Consider the reduced word 4321=s3s1s2s1s3s24321=s_{3}s_{1}s_{2}s_{1}s_{3}s_{2} which corresponds to the reflection ordering written on top of the arrows:

1234{1234}1243{1243}2143{2143}2413{2413}4213{4213}4231{4231}4321.{4321.}e3e4\scriptstyle{e_{3}-e_{4}}e1e2\scriptstyle{e_{1}-e_{2}}e1e4\scriptstyle{e_{1}-e_{4}}e2e4\scriptstyle{e_{2}-e_{4}}e1e3\scriptstyle{e_{1}-e_{3}}e2e3\scriptstyle{e_{2}-e_{3}}

2.3. Grassmaniann, Flag variety, and Plücker coordinates

Define the Grassmaniann Gr(k,n)\mathrm{Gr}(k,n) to be the space of kk-dimensional subspaces VnV\subseteq\mathbb{C}^{n}. The Plücker embedding is a closed embedding ι:Gr(k,n)(Λk(n)))\iota:\mathrm{Gr}(k,n)\hookrightarrow\mathbb{P}(\Lambda^{k}(\mathbb{C}^{n}))) that sends the subspace VV with basis v1,v2,,vk\vec{v}_{1},\vec{v}_{2},\dots,\vec{v}_{k} to [v1v2vk][\vec{v}_{1}\wedge\vec{v}_{2}\wedge\dots\wedge\vec{v}_{k}], the projective equivalence class of v1v2vk\vec{v}_{1}\wedge\vec{v}_{2}\wedge\dots\wedge\vec{v}_{k} in Λk(n)\Lambda^{k}(\mathbb{C}^{n}). For any i1,,ik[n]i_{1},\dots,i_{k}\in[n], let Pi1,,ik(V)P_{i_{1},\dots,i_{k}}(V) be the Plücker coordinate of VV. For any permutation σSk\sigma\in S_{k},

Pi1,,ik(V)=(1)(σ)Pσ(i1),,σ(ik)(V).P_{i_{1},\dots,i_{k}}(V)=(-1)^{\ell(\sigma)}P_{\sigma(i_{1}),\dots,\sigma(i_{k})}(V).

In particular, this means Pi1,,ik(V)=0P_{i_{1},\dots,i_{k}}(V)=0 if ia=ibi_{a}=i_{b} for some ab[k]a\neq b\in[k]. For I([n]k)I\in{[n]\choose k}, set PI:=Pi1,,ikP_{I}:=P_{i_{1},\dots,i_{k}} where I={i1<<ik}I=\{i_{1}<\dots<i_{k}\}. For any permutation wSnw\in S_{n}, define

Pw=:k=1n1Pw[k]P_{w}=:\prod_{k=1}^{n-1}P_{w[k]}

where w[k]:={w1,w2,,wk}([n]k)w[k]:=\{w_{1},w_{2},\dots,w_{k}\}\in{[n]\choose k} for k[n1]k\in[n-1]. In particular, Pw[k](w){1,1}P_{w[k]}(w)\in\{-1,1\}. Set PI+i:=Pi1,,ik,iP_{I+i}:=P_{i_{1},\dots,i_{k},i} and PIij:=(1)kjPi1,,ij^,,ikP_{I-i_{j}}:=(-1)^{k-j}P_{i_{1},\dots,\widehat{i_{j}},\dots,i_{k}} for all j[k]j\in[k]. More generally, for J={ij1<<ijr}IJ=\{i_{j_{1}}<\dots<i_{j_{r}}\}\subset I, set PIJ:=P((Iijr))ij1P_{I-J}:=P_{((I-i_{j_{r}})-\dots)-i_{j_{1}}}. The variety Gr(k,n)\mathrm{Gr}(k,n) is the zero locus of the ideal generated by the following Plücker relations.

Definition 2.13.

For any I([n]k+1),J([n]k1)I\in{[n]\choose k+1},J\in{[n]\choose k-1}, the Plücker relation associated to I,JI,J is:

tIPItPJ+t=0.\sum_{t\in I}P_{I-t}P_{J+t}=0.

Let G=GLn()G={GL}_{n}(\mathbb{C}) and B,BGB,B_{-}\subset G be the Borel and opposite Borel subgroup of GG consisting of invertible upper and lower triangular matrices respectively. Let T=BBT=B\cap B_{-} be the maximal torus of diagonal matrices in GG.

The complete flag variety is defined to be Fln=G/B\mathrm{Fl}_{n}=G/B. Fixing a basis of n\mathbb{C}^{n}, we can identify a point gBG/BgB\in G/B with a flag F=0=F0F1F2Fn1Fn=nF_{\bullet}=0=F_{0}\subsetneq F_{1}\subsetneq F_{2}\subsetneq\cdots\subsetneq F_{n-1}\subsetneq F_{n}=\mathbb{C}^{n} where FkGr(k,n)F_{k}\in\mathrm{Gr}(k,n) is the span of the first kk column vectors of any n×nn\times n matrix representative MFgBM_{F}\in gB. We define the multi-Plücker embedding to be the composition

Flnk=0nGr(k,n)k=0n(Λk(n))\mathrm{Fl}_{n}\hookrightarrow\prod_{k=0}^{n}\mathrm{Gr}(k,n)\hookrightarrow\prod_{k=0}^{n}\mathbb{P}(\Lambda^{k}(\mathbb{C}^{n}))

that sends FF_{\bullet} to (ι(F0),ι(F1),,ι(Fn))(\iota(F_{0}),\iota(F_{1}),\dots,\iota(F_{n})). For any kk and I([n]k)I\in{[n]\choose k}, define the Plücker coordinate PI(F)P_{I}(F_{\bullet}) to be PI(Fk)P_{I}(F_{k}). It is also the minor of MFM_{F} in the rows indexed by II in the first kk columns. Similar to the Grassmaniann, the complete flag variety is the zero locus of the following incidence Plücker relations:

Definition 2.14 ([8]).

For any I([n]r),J([n]s)I\in{[n]\choose r},J\in{[n]\choose s} with 1sr<n1\leq s\leq r<n fix AJA\subseteq J, the incidence Plücker relation is:

(2) PIPJ=BI,|B|=|A|P(IB)+AP(JA)+B.P_{I}P_{J}=\sum_{B\subseteq I,|B|=|A|}P_{(I-B)+A}P_{(J-A)+B}.

There are two special cases of Equation (2) that will be of particular importance to us. We now rewrite the incidence Plücker relations in these cases for convenience.

Case I.(A=JA=J and |I|=|J|+1|I|=|J|+1): For I([n]k),J([n]k1)I\in{[n]\choose k},J\in{[n]\choose k-1} where 1<k<n1<k<n,

(3) PIPJ=iIPIiPJ+iP_{I}P_{J}=\sum_{i\in I}P_{I-i}P_{J+i}

Case II.(|A|=1|A|=1 and rs2r-s\geq 2): For I([n]r),J([n]s)I\in{[n]\choose r},J\in{[n]\choose s} with 1sr<n1\leq s\leq r<n,

(4) iIPIiPJ+i=0.\sum_{i\in I}P_{I-i}P_{J+i}=0.

Equivalently,

(5) PIPJ+j=iIPIi+jPJ+i.P_{I}P_{J+j}=-\sum_{i\in I}P_{I-i+j}P_{J+i}.

2.4. Schubert and Richardson varieties in Fln\mathrm{Fl}_{n}

Given a permutation wSnw\in S_{n}, we also view ww as the permutation matrix with 11 at (w(i),i)(w(i),i) and 0 everywhere else111We note that in some literature, this is the permutation matrix that corresponds to w1w^{-1} instead of ww.. The set of TT-fixed points on Fln\mathrm{Fl}_{n} is {ew=wB/B:wSn}\{e_{w}=wB/B:w\in S_{n}\}. Define the Schubert cell Xw=BewX_{w}^{\circ}=Be_{w} to be the Borel orbit of ewe_{w} and the Schubert variety Xw=Xw¯X_{w}=\overline{X_{w}^{\circ}} to be the closure of the Schubert cell. Similarly, define the opposite Schubert cell Ωw=Bew\Omega_{w}^{\circ}=B_{-}e_{w} and the opposite Schubert variety Ωw=Ωw¯\Omega_{w}=\overline{\Omega_{w}^{\circ}}. Here Xw(w)X_{w}^{\circ}\cong\mathbb{C}^{\ell(w)} and Ωw(n2)(w)\Omega_{w}^{\circ}\cong\mathbb{C}^{{n\choose 2}-\ell(w)}.

We will make use of the following equivalent definition of (opposite) Schubert cells and Schubert varieties. For SS any subset of [n][n], let ProjS:n|S|\mathrm{Proj}_{S}:\mathbb{C}^{n}\twoheadrightarrow\mathbb{C}^{|S|} be the projection onto the coordinates with indices in SS. The Schubert cell is

Xw={FFln:dim(Proj[nj+1,n](Fi))=#{w([i])[nj+1,n]} for all i,j[n]},X_{w}^{\circ}=\{F_{\bullet}\in\mathrm{Fl}_{n}:\dim(\mathrm{Proj}_{[n-j+1,n]}(F_{i}))=\#\{w([i])\cap[n-j+1,n]\}\text{ for all }i,j\in[n]\},

and the Schubert variety is

Xw={FFln:dim(Proj[nj+1,n](Fi))#{w([i])[nj+1,n]} for all i,j[n]}.X_{w}=\{F_{\bullet}\in\mathrm{Fl}_{n}:\dim(\mathrm{Proj}_{[n-j+1,n]}(F_{i}))\leq\#\{w([i])\cap[n-j+1,n]\}\text{ for all }i,j\in[n]\}.

Similarly, the opposite Schubert cell is

Ωw={FFln:dim(Proj[j](Fi))=#{w([i])[j]} for all i,j[n]},\Omega_{w}^{\circ}=\{F_{\bullet}\in\mathrm{Fl}_{n}:\dim(\mathrm{Proj}_{[j]}(F_{i}))=\#\{w([i])\cap[j]\}\text{ for all }i,j\in[n]\},

and the opposite Schubert variety is

Ωw={FFln:dim(Proj[j](Fi))#{w([i])[j]} for all i,j[n]}.\Omega_{w}=\{F_{\bullet}\in\mathrm{Fl}_{n}:\dim(\mathrm{Proj}_{[j]}(F_{i}))\leq\#\{w([i])\cap[j]\}\text{ for all }i,j\in[n]\}.

The (opposite) Schubert varieties possess a stratification by (opposite) Schubert cells:

Xw=uwXuand Ωw=wvΩv,X_{w}=\bigsqcup_{u\leq w}X_{u}^{\circ}\ \text{and }\Omega_{w}=\bigsqcup_{w\leq v}\Omega_{v}^{\circ},

where “\leq” denotes the strong Bruhat order on SnS_{n}.

It is perhaps easier to visualize Schubert and opposite Schubert varieties in the following way. For any i,j[n]i,j\in[n] and any MMatn×nM\in\mathrm{Mat}_{n\times n}, define ri,jSW(M)r_{i,j}^{SW}(M) to be the rank of the submatrix of MM obtained by taking the bottom ni+1n-i+1 rows and left jj columns. Define similarly ri,jNW(M)r_{i,j}^{NW}(M) to be the rank of the submatrix obtained by taking the top ii rows and left jj columns. Let MFGM_{F}\in G be an matrix representative of FF_{\bullet} such that FiF_{i} is the span of the first ii column vectors of MFM_{F}. Then

(6) FXwri,jSW(MF)ri,jSW(w)FΩwri,jNW(MF)ri,jNW(w),\displaystyle\begin{split}F_{\bullet}\in X_{w}&\iff r^{SW}_{i,j}(M_{F})\leq r^{SW}_{i,j}(w)\\ F_{\bullet}\in\Omega_{w}&\iff r^{NW}_{i,j}(M_{F})\leq r^{NW}_{i,j}(w)\end{split},

and FF_{\bullet} lies in the (opposite) Schubert cell if (6) holds after replacing “\leq” with “==”.

Define the open Richardson variety to be u,v=XvΩu\mathcal{R}^{\circ}_{u,v}=X^{\circ}_{v}\cap\Omega^{\circ}_{u} and the Richardson variety to be u,v=XvΩu\mathcal{R}_{u,v}=X_{v}\cap\Omega_{u}. In particular, u,v\mathcal{R}^{\circ}_{u,v} and u,v\mathcal{R}_{u,v} are non-empty if and only if uvu\leq v, in which case we have

dim(u,v)=dim(u,v)=(v)(u)=(u,v).\dim(\mathcal{R}_{u,v})=\dim(\mathcal{R}_{u,v}^{\circ})=\ell(v)-\ell(u)=\ell(u,v).

Similar to Schubert varieties, we also have

(7) u,v=u,v¯and u,v=[x,y][u,v]x,y,\mathcal{R}_{u,v}=\overline{\mathcal{R}_{u,v}^{\circ}}\ \text{and }\mathcal{R}_{u,v}=\bigsqcup_{[x,y]\subseteq[u,v]}\mathcal{R}_{x,y}^{\circ},

where the disjoint union is taken over all Bruhat intervals [x,y][x,y] contained in [u,v][u,v].

Moreover, as subvarieties of Fln\mathrm{Fl_{n}}, each Xu,ΩuX_{u},\Omega_{u} and u,v\mathcal{R}_{u,v} can be defined by vanishing of Plücker coordinates:

Xu\displaystyle X_{u} ={FFln:Pw(F)=0 for wu}\displaystyle=\{F_{\bullet}\in\mathrm{Fl}_{n}:P_{w}(F_{\bullet})=0\text{ for }w\nleq u\} Xu\displaystyle X_{u}^{\circ} =Xu{Pu0}\displaystyle=X_{u}\cap\{P_{u}\neq 0\}
(8) Ωu\displaystyle\Omega_{u} ={FFln:Pw(F)=0 for wu}\displaystyle=\{F_{\bullet}\in\mathrm{Fl}_{n}:P_{w}(F_{\bullet})=0\text{ for }w\ngeq u\} Ωu\displaystyle\Omega_{u}^{\circ} =Ωu{Pu0}\displaystyle=\Omega_{u}\cap\{P_{u}\neq 0\}
u,v\displaystyle\mathcal{R}_{u,v} ={FFln:Pw(F)=0 for w[u,v]}\displaystyle=\{F_{\bullet}\in\mathrm{Fl}_{n}:P_{w}(F_{\bullet})=0\text{ for }w\notin[u,v]\} u,v\displaystyle\mathcal{R}^{\circ}_{u,v} =u,v{PuPv0}.\displaystyle=\mathcal{R}_{u,v}\cap\{P_{u}P_{v}\neq 0\}.

2.5. Schubert and Richardson varieties in Grassmannian

For each I([n]k)I\in{[n]\choose k}, define the Grassmannian Schubert variety by

XI={VGr(k,n):dim(Proj[nj+1,n])(V)#(I[nj+1,n]) for all j[n]}X_{I}=\{V\in\mathrm{Gr}(k,n):\dim(\mathrm{Proj}_{[n-j+1,n]})(V)\leq\#(I\cap[n-j+1,n])\text{ for all }j\in[n]\}

and the Grassmannian Schubert cell by

XI={VGr(k,n):dim(Proj[nj+1,n])(V)=#(I[nj+1,n]) for all j[n]}.X_{I}^{\circ}=\{V\in\mathrm{Gr}(k,n):\dim(\mathrm{Proj}_{[n-j+1,n]})(V)=\#(I\cap[n-j+1,n])\text{ for all }j\in[n]\}.

Define the Grassmannian opposite Schubert variety and opposite Schubert cell by

ΩI={VGr(k,n):dim(Proj[j])(V)#(I[j]) for all j[n]}\Omega_{I}=\{V\in\mathrm{Gr}(k,n):\dim(\mathrm{Proj}_{[j]})(V)\leq\#(I\cap[j])\text{ for all }j\in[n]\}

and

ΩI={VGr(k,n):dim(Proj[j])(V)=#(I[j]) for all j[n]}.\Omega_{I}^{\circ}=\{V\in\mathrm{Gr}(k,n):\dim(\mathrm{Proj}_{[j]})(V)=\#(I\cap[j])\text{ for all }j\in[n]\}.

Similar to Schubert varieties in Fln\mathrm{Fl}_{n}, we have XI=XI¯X_{I}=\overline{X_{I}^{\circ}} and ΩI=ΩI¯\Omega_{I}=\overline{\Omega_{I}^{\circ}}. For any two kk-element subsets I,J[n]I,J\subset[n] where I={i1<i2<<ik}I=\{i_{1}<i_{2}<\dots<i_{k}\} and J={j1<j2<<jk}J=\{j_{1}<j_{2}<\dots<j_{k}\}, we say IJI\leq J in the Gale order if irjri_{r}\leq j_{r} for all r[k]r\in[k]. Then the (opposite) Grassmaniann Schubert varieties are disjoint union of (opposite) Grassmaniann Schubert cells:

XI=JIXJand ΩJ=IJΩJ.X_{I}=\bigsqcup_{J\leq I}X_{J}^{\circ}\ \text{and }\Omega_{J}=\bigsqcup_{I\leq J}\Omega_{J}^{\circ}.

When IJI\leq J, define the open Grassmaniann Richardson variety I,J=XJΩI\mathcal{R}_{I,J}^{\circ}=X_{J}^{\circ}\cap\Omega_{I}^{\circ} and its closure the Grassmaniann Richardson variety I,J=XJΩI\mathcal{R}_{I,J}=X_{J}\cap\Omega_{I}. We similarly have

(9) I,J=I,J¯and I,J=[I,J][I,J]I,J.\mathcal{R}_{I,J}=\overline{\mathcal{R}_{I,J}^{\circ}}\ \text{and }\mathcal{R}_{I,J}=\bigsqcup_{[I^{\prime},J^{\prime}]\subseteq[I,J]}\mathcal{R}_{I^{\prime},J^{\prime}}^{\circ}.

One can also define XI,ΩIX_{I},\Omega_{I} and I,J\mathcal{R}_{I,J} by vanishing of Plücker coordinates:

XI\displaystyle X_{I} ={VGr(k,n):PK(V)=0 for KI}\displaystyle=\{V\in\mathrm{Gr}(k,n):P_{K}(V)=0\text{ for }K\nleq I\} XI\displaystyle X_{I}^{\circ} =XI{PI0}\displaystyle=X_{I}\cap\{P_{I}\neq 0\}
(10) ΩI\displaystyle\Omega_{I} ={VGr(k,n):PK(V)=0 for KI}\displaystyle=\{V\in\mathrm{Gr}(k,n):P_{K}(V)=0\text{ for }K\ngeq I\} ΩI\displaystyle\Omega_{I}^{\circ} =ΩI{PI0}\displaystyle=\Omega_{I}\cap\{P_{I}\neq 0\}
I,J\displaystyle\mathcal{R}_{I,J} ={VGr(k,n):PK(V)=0 for K[I,J]}\displaystyle=\{V\in\mathrm{Gr}(k,n):P_{K}(V)=0\text{ for }K\notin[I,J]\} I,J\displaystyle\mathcal{R}^{\circ}_{I,J} =I,J{PIPJ0}.\displaystyle=\mathcal{R}_{I,J}\cap\{P_{I}P_{J}\neq 0\}.

3. Combinatorics of the quantum Bruhat graph

3.1. Minimal weights in the quantum Bruhat graph

Our first theorem provides an explicit formula for the minimal weight qd(u,v)q^{d(u,v)} from any pair of permutation uu to vv in SnS_{n}.

Definition 3.1.

For A,B[n]A,B\subset[n] with |A|=|B||A|=|B|, we construct a lattice path P(A,B)P(A,B) starting at (0,0)(0,0) and ending at (n,0)(n,0) with nn steps as follows. For each i=1,,ni=1,\ldots,n, the ithi^{th} step is

  • an upstep (1,1)(1,1) if iAi\in A and iBi\notin B,

  • a downstep (1,1)(1,-1) if iAi\notin A and iBi\in B,

  • a horizontal step (1,0)(1,0) if iABi\in A\cap B or iABi\notin A\cup B.

Define its depth to be the largest number y0y\geq 0 such that this lattice path passes through (x,y)(x,-y) for some xx, denoted depth(A,B)\operatorname{depth}(A,B).

Example 3.2.

Let n=7n=7 with A={3,4,6,7}A=\{3,4,6,7\} and B={1,2,3,5}B=\{1,2,3,5\}. The lattice path P(A,B)P(A,B) is shown in Figure 3 with depth(A,B)=2\operatorname{depth}(A,B)=2.

\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bulletdepth=2\mathrm{depth}=211223344556677
Figure 3. The lattice path P(A,B)P(A,B)
Theorem 3.3.

Let u,vSnu,v\in S_{n}. All shortest paths from uu to vv have weight qd(u,v)=q1d1(u,v)qn1dn1(u,v)q^{d(u,v)}=q_{1}^{d_{1}(u,v)}\cdots q_{n-1}^{d_{n-1}(u,v)} where dk(u,v)=depth(u[k],v[k])d_{k}(u,v)=\operatorname{depth}(u[k],v[k]). Here, w[k]:={w1,,wk}w[k]:=\{w_{1},\ldots,w_{k}\}.

Example 3.4.

Consider u=7364152u=7364152 and v=2513746v=2513746 in S7S_{7}. We need to figure out dk(u,v)d_{k}(u,v) for k=1,,n1k=1,\ldots,n-1. For k=1k=1, u[1]={7}u[1]=\{7\} and v[1]={2}v[1]=\{2\} so we need to construct a lattice path with an upstep at position 77 and a downstep at position 22. This lattice path has depth 11, meaning that d1(u,v)=1d_{1}(u,v)=1. We continue this procedure for all kk’s. For example, at k=4k=4, the lattice path P(u[4],v[4])P(u[4],v[4]) is discussed in Example 3.2 with d4(u,v)=2d_{4}(u,v)=2. In the end, we arrive at qd(u,v)=q1q2q32q42q5q6q^{d(u,v)}=q_{1}q_{2}q_{3}^{2}q_{4}^{2}q_{5}q_{6}.

We make use of the following technical tool by Brenti-Fomin-Postnikov. In the quantum Bruhat graph Γn\Gamma_{n}, label each directed edge wwtijw\rightarrow wt_{ij} by eieje_{i}-e_{j}. The “label” here does not have indications towards the weight.

Theorem 3.5 ([2]).

Fix a reflection ordering τ=τ1,,τ(n2)\tau=\tau_{1},\ldots,\tau_{n\choose 2} of Φ+\Phi^{+} and fix u,vSnu,v\in S_{n}. Then there is a unique directed path from uu to vv in Γn\Gamma_{n} such that its sequence of labels is strictly increasing with respect to τ\tau. Moreover, this path has length (u,v)\ell(u,v).

Therefore, to prove Theorem 3.3, we pick a specific reflection ordering, and find its corresponding directed path and compute its weight, recalling that shortest paths are precisely minimal weight paths (Lemma 2.4). We need one more definition.

Definition 3.6.

For r[n]r\in[n], let r\leq_{r} be the shifted linear order on [n][n] given by

r<rr+1<r<rn<r1<r<rr1.r<_{r}r+1<_{r}\cdots<_{r}n<_{r}1<_{r}\cdots<_{r}r-1.

For A={a1<r<rak},B={b1<r<rbk}[n]A=\{a_{1}<_{r}\cdots<_{r}a_{k}\},B=\{b_{1}<_{r}\cdots<_{r}b_{k}\}\subset[n], define the shifted Gale order r\leq_{r} as

ArBairbi for all i[k].A\leq_{r}B\iff a_{i}\leq_{r}b_{i}\text{ for all }i\in[k].
Proof of Theorem 3.3.

Choose the reflection ordering τ=e1e2,e1e3,,e1en,e2e3,e2en,,en1en\tau=e_{1}-e_{2},e_{1}-e_{3},\ldots,e_{1}-e_{n},e_{2}-e_{3},\ldots e_{2}-e_{n},\ldots,e_{n-1}-e_{n} that starts with all positive roots involving e1e_{1} and then e2e_{2} and so on. This is a valid reflection order because for any i<j<ki<j<k, eieje_{i}-e_{j}, eieke_{i}-e_{k} and ejeke_{j}-e_{k} appear in this order. One can also check that this ordering corresponds to the reduced word w0=(s1s2sn1)(s1s2sn2)(s1)w_{0}=(s_{1}s_{2}\cdots s_{n-1})(s_{1}s_{2}\cdots s_{n-2})\cdots(s_{1}).

Fix uu and vv. Theorem 3.5 says that there is a unique directed path from uu to vv in Γn\Gamma_{n} by uuti1j1uti1j1ti2j2uti1j1tij=vu\rightarrow ut_{i_{1}j_{1}}\rightarrow ut_{i_{1}j_{1}}t_{i_{2}j_{2}}\rightarrow\cdots\rightarrow ut_{i_{1}j_{1}}\cdots t_{i_{\ell}j_{\ell}}=v such that ei1ej1,,eieje_{i_{1}}-e_{j_{1}},\ldots,e_{i_{\ell}}-e_{j_{\ell}} appear in the order of τ\tau. We describe this unique directed path explicitly. Consider the first n1n-1 roots e1e2,,e1ene_{1}-e_{2},\ldots,e_{1}-e_{n}. After choosing e1ep1,e1ep2,,e1epme_{1}-e_{p_{1}},e_{1}-e_{p_{2}},\ldots,e_{1}-e_{p_{m}} such that we have a directed path uut1p1ut1p1t1pm=uu\rightarrow ut_{1p_{1}}\rightarrow\cdots\rightarrow ut_{1p_{1}}\cdots t_{1p_{m}}=u^{\prime}, we must need u(1)=v(1)=v1u^{\prime}(1)=v(1)=v_{1} since the rest of positive roots do not involve the index 11. At the same time, as long as u(1)=v(1)u^{\prime}(1)=v(1), we can continue finding the path via induction on nn.

We can choose pip_{i} to be the smallest index greater than pi1p_{i-1} (with the convention that p0=1p_{0}=1) such that u(pi)>v1+1u(pi1)u(p_{i})>_{v_{1}+1}u(p_{i-1}) in the shifted linear order on [n][n] where v1v_{1} is declared the largest number. Such choices work because by construction, there is an edge from ut1p1t1pi1ut_{1p_{1}}\cdots t_{1p_{i-1}} to ut1p1t1piut_{1p_{1}}\cdots t_{1p_{i}} in Γn\Gamma_{n} and we inevitably end up at pm=u1(v1)p_{m}=u^{-1}(v_{1}) since v1v_{1} is the largest number in this linear order. Here is an example with v1=4v_{1}=4:

permutationrootweightu6¯57¯913428e1e317¯569¯13428e1e419¯5671¯3428e1e5q1q2q3q41¯56793¯428e1e613¯567914¯28e1e71456791328v4 .\begin{tabular}[]{c|c|c|c}&permutation&root&weight\\ \hline\cr$u$&$\underline{\textbf{6}}5\underline{\textbf{7}}913428$&$e_{1}-e_{3}$&1\\ \hline\cr&$\underline{\textbf{7}}56\underline{\textbf{9}}13428$&$e_{1}-e_{4}$&1\\ \hline\cr&$\underline{\textbf{9}}567\underline{\textbf{1}}3428$&$e_{1}-e_{5}$&$q_{1}q_{2}q_{3}q_{4}$\\ \hline\cr&$\underline{\textbf{1}}5679\underline{\textbf{3}}428$&$e_{1}-e_{6}$&1\\ \hline\cr&$\underline{\textbf{3}}56791\underline{\textbf{4}}28$&$e_{1}-e_{7}$&1\\ \hline\cr&$\textbf{4}56791328$&&\\ &$\cdots$&&\\ \hline\cr$v$&$\textbf{4}\rule{41.54121pt}{0.42677pt}$&&\\ \hline\cr\end{tabular}.

We now examine how the weight accumulated. The weight of this path uuu\rightarrow\cdots\rightarrow u^{\prime} equals q1q2qpk1q_{1}q_{2}\cdots q_{p_{k}-1} if u(pk)<u(pk1)u(p_{k})<u(p_{k-1}) and u(pk)>v1+1u(pk1)u(p_{k})>_{v_{1}+1}u(p_{k-1}), and equals 11 if no such kk exists, i.e. u(1)<v(1)u(1)<v(1). It’s clear that at most one such kk exists which is intuitively at the place where the shifted linear order “wraps around”. Denote this weight by wt1(u,v)\operatorname{wt}_{1}(u,v). Recall that the weight of a minimal weight path from uu to vv is denoted as qd(u,v)q^{d(u,v)}. Also let depth(u,v)\operatorname{depth}(u,v) be a vector such that depthj(u,v)=depth(u[j],v[j])\operatorname{depth}_{j}(u,v)=\operatorname{depth}(u[j],v[j]) as in Definition 3.1.

Our goal is to show that d(u,v)=depth(u,v)d(u,v)=\operatorname{depth}(u,v) and we use induction on nn. By Theorem 3.5, wt1(u,v)qd(u,v)=qd(u,v)\operatorname{wt}_{1}(u,v)q^{d(u^{\prime},v)}=q^{d(u,v)}, and by induction hypothesis, qd(u,v)=qdepth(u,v)q^{d(u^{\prime},v)}=q^{\operatorname{depth}(u^{\prime},v)} since u(1)=v(1)u^{\prime}(1)=v(1). It remains to prove that

(11) wt1(u,v)qdepth(u,v)=qdepth(u,v).\operatorname{wt}_{1}(u,v)q^{\operatorname{depth}(u^{\prime},v)}=q^{\operatorname{depth}(u,v)}.

We study Equation (11) coordinate by coordinate.

Case I: (u1>v1u_{1}>v_{1}). There exists a unique kk such that u(pk)<u(pk1)u(p_{k})<u(p_{k-1}) and u(pk)>v1+1u(pk1)u(p_{k})>_{v_{1}+1}u(p_{k-1}), with wt1(u,v)=q1q2qpk1\operatorname{wt}_{1}(u,v)=q_{1}q_{2}\cdots q_{p_{k}-1}. In fact, all of u(1),,u(pk1)u(1),\ldots,u(p_{k}-1) are strictly greater than v1v_{1} by construction. Now compare the exponent of qjq^{j} on both sides of Equation (11). In other words, we will compare the two lattice paths P(u[j],v[j])P(u[j],v[j]) and P(u[j],v[j])P(u^{\prime}[j],v[j]).

For j<pkj<p_{k}, u[j]u^{\prime}[j] is obtained from u[j]u[j] by deleting its largest element and replacing it by v(1)v(1) which is smaller than all values in u[j]u[j]. Thus, the path P(u[j],v[j])P(u[j],v[j]) is strictly below the xx-axis when the xx-coordinate is v(1)v(1), and is weakly above the xx-axis when the xx-coordinate is maxu[j]\max u[j]. The lattice path P(u[j],v[j])P(u^{\prime}[j],v[j]) is obtained from P(u[j],v[j])P(u[j],v[j]) by moving the subpath from (v1,)(v_{1},-) to (maxu[j],)(\max u[j],-) one step up. As a result, depth(u[j],v[j])=depth(u[j],v[j])+1\operatorname{depth}(u[j],v[j])=\operatorname{depth}(u^{\prime}[j],v[j])+1 as desired. A cartoon for visualization for this scenario is seen in Figure 4.

\bullet\bullet\bullet\bullet(v1,)(v_{1},-)\bullet(maxu[j],)(\max u[j],-)\bullet\bullet\bullet\bullet\bullet\bullet\bullet(v1,)(v_{1},-)\bullet(maxu[j],)(\max u[j],-)\bullet\bullet\bullet
Figure 4. A comparison for P(u[j],v[j])P(u[j],v[j]) (left) and P(u[j],v[j])P(u^{\prime}[j],v[j]) (right) for j<pkj<p_{k} in case 1 of the proof of Theorem 3.3 where the red curve stands for any lattice subpath

For jpmj\geq p_{m}, u[j]=u[j]u^{\prime}[j]=u[j] so depth(u[j],v[j])=depth(u[j],v[j])\operatorname{depth}(u^{\prime}[j],v[j])=\operatorname{depth}(u[j],v[j]) as desired.

For pkj<pmp_{k}\leq j<p_{m}, v1u[j]v_{1}\notin u[j] and u[j]u^{\prime}[j] is obtained from u[j]u[j] by deleting the largest number smaller than v1v_{1}, called bb, and replacing it by v1v_{1}. Thus, P(u[j],v[j])P(u^{\prime}[j],v[j]) and P(u[j],v[j])P(u[j],v[j]) only differ from step bb to step v1v_{1}. During this period,

P(u[j],v[j]):\displaystyle P(u[j],v[j]):\ ( or )+(a sequence of  and ’s)+(),\displaystyle(\nearrow\text{ or }\rightarrow)\ +\ (\text{a sequence of }\rightarrow\text{ and }\searrow\text{'s})\ +\ (\searrow),
P(u[j],v[j]):\displaystyle P(u^{\prime}[j],v[j]):\ ( or )+(a sequence of  and ’s)+().\displaystyle(\rightarrow\text{ or }\searrow)\ +\ (\text{a sequence of }\rightarrow\text{ and }\searrow\text{'s})\ +\ (\rightarrow).

This local change does not modify the depth. So depth(u[j],v[j])=depth(u[j],v[j])\operatorname{depth}(u[j],v[j])=\operatorname{depth}(u^{\prime}[j],v[j]).

Case II: (u1v1u_{1}\leq v_{1}). Here wt1(u,v)=1\operatorname{wt}_{1}(u,v)=1 and we need depth(u[j],v[j])=depth(u[j],v[j])\operatorname{depth}(u^{\prime}[j],v[j])=\operatorname{depth}(u[j],v[j]) for all jj. The sub-case for jpmj\geq p_{m} is exactly the same as the sub-case for jpmj\geq p_{m} in Case 1, and the sub-case for j<pmj<p_{m} is exactly the same as the sub-case for pkj<pmp_{k}\leq j<p_{m} in Case 1.

The induction step goes through and we conclude that d(u,v)=depth(u,v)d(u,v)=\operatorname{depth}(u,v). ∎

3.2. Characterization of the tilted Bruhat order

A crucial consequence of Theorem 3.3 is a succinct description (Theorem 3.9) of the tilted Bruhat order defined by Brenti-Fomin-Postnikov. We now build up some intuition regarding the relationship between the shifted Gale order r\leq_{r} and the lattice path construction (Definition 3.1).

Lemma 3.7.

For all A,B[n]A,B\subset[n] with |A|=|B||A|=|B|, there exists r[n]r\in[n] such that ArBA\leq_{r}B. In fact, ArBA\leq_{r}B if and only if the lattice path P(A,B)P(A,B) passes through (r1,depth(A,B))(r-1,-\operatorname{depth}(A,B)).

Proof.

Note that ABA\leq B in the Gale order if and only if the lattice path P(A,B)P(A,B) does not go below the xx-axis, i.e. P(A,B)P(A,B) is a Dyck path. Likewise, ArBA\leq_{r}B if and only if the lattice path P(A+r1,B+r1)P(A+r-1,B+r-1) does not go below the xx-axis, where A+r1={a+r1|aA}A+r-1=\{a+r-1\>|\>a\in A\} where the values are taken modulo nn. Therefore, for general AA and BB with |A|=|B||A|=|B|, let the steps of P(A,B)P(A,B) be g1,g2,,gng_{1},g_{2},\ldots,g_{n} where each gi{,,}g_{i}\in\{\rightarrow,\nearrow,\searrow\} based on Definition 3.1. Then the lattice path P(A+r1,B+r1)P(A+r-1,B+r-1) is then constructed via the steps gr,gr+1,,gn,g1,,gr1g_{r},g_{r+1},\ldots,g_{n},g_{1},\ldots,g_{r-1} in this order. We now see that P(A+r1,B+r1)P(A+r-1,B+r-1) does not go below the xx-axis if and only if the lattice path P(A,B)P(A,B) goes through its lowest point at xx-coordinate equals r1r-1. ∎

If ArBA\leq_{r}B, let [A,B]r:={I:ArIrB}[A,B]_{r}:=\{I:A\leq_{r}I\leq_{r}B\} denote the interval in shifted Gale order r\leq_{r}. The following lemma explains the relation between different shifted Gale orders.

Lemma 3.8.

For all A,B[n]A,B\subset[n] and |A|=|B||A|=|B|, if there exists rr[n]r\neq r^{\prime}\in[n] such that ArBA\leq_{r}B and ArBA\leq_{r^{\prime}}B, then the intervals [A,B]r=[A,B]r[A,B]_{r}=[A,B]_{r^{\prime}}. Moreover, for any I[A,B]rI\in[A,B]_{r},

#(A[r,r)c)=#(I[r,r)c)=#(B[r,r)c).\#(A\cap[r,r^{\prime})_{c})=\#(I\cap[r,r^{\prime})_{c})=\#(B\cap[r,r^{\prime})_{c}).
Proof.

Without loss of generality assume r<rr<r^{\prime}. Since ArBA\leq_{r}B and ArBA\leq_{r^{\prime}}B, by Lemma 3.7, the lattice path P(A,B)P(A,B) passes through (r,d)(r,-d) and (r,d)(r^{\prime},-d), where d=depth(A,B)d=\operatorname{depth}(A,B). Since they have the same yy-coordinate, there are the same number of \nearrow and \searrow between the two points and thus

#(A[r,r)c)=#(B[r,r)c).\#(A\cap[r,r^{\prime})_{c})=\#(B\cap[r,r^{\prime})_{c}).

For any I[A,B]rI\in[A,B]_{r}, we have

IrA#(I[r,r)c)#(A[r,r)c),I\geq_{r}A\implies\#(I\cap[r,r^{\prime})_{c})\leq\#(A\cap[r,r^{\prime})_{c}),
IrB#(I[r,r)c)#(B[r,r)c).I\leq_{r}B\implies\#(I\cap[r,r^{\prime})_{c})\geq\#(B\cap[r,r^{\prime})_{c}).

Therefore,

#(I[r,r)c)=#(A[r,r)c)=#(B[r,r)c).\#(I\cap[r,r^{\prime})_{c})=\#(A\cap[r,r^{\prime})_{c})=\#(B\cap[r,r^{\prime})_{c}).

As a result, the lattice path P(A,I)P(A,I) have the same yy-coordinate at x=rx=r and rr^{\prime}. Since ArIA\leq_{r}I, P(A,I)P(A,I) reaches its lowest point at both xx-coordinate rr and rr^{\prime}. Similarly, P(I,B)P(I,B) reaches its lowest point at both xx-coordinate rr and rr^{\prime}. By Lemma 3.7, ArIrBA\leq_{r^{\prime}}I\leq_{r^{\prime}}B. Therefore [A,B]r[A,B]r[A,B]_{r}\subseteq[A,B]_{r^{\prime}}. The opposite direction follows by same reasoning. ∎

Theorem 3.9.

For u,v,wSnu,v,w\in S_{n}, the following are equivalent:

  1. (1)

    wuvw\leq_{u}v;

  2. (2)

    w[u,v]w\in[u,v];

  3. (3)

    for all sequence 𝐚=a1an1\mathbf{a}=a_{1}\ldots a_{n-1} such that {u1,,uk}ak{v1,,vk}\{u_{1},\ldots,u_{k}\}\leq_{a_{k}}\{v_{1},\ldots,v_{k}\} and for all k[n1]k\in[n-1], we have {u1,,uk}ak{w1,wk}ak{v1,,vk}\{u_{1},\ldots,u_{k}\}\leq_{a_{k}}\{w_{1},\ldots w_{k}\}\leq_{a_{k}}\{v_{1},\ldots,v_{k}\};

  4. (4)

    there exists a sequence 𝐚=a1an1\mathbf{a}=a_{1}\ldots a_{n-1} such that for all k[n1],{u1,,uk}ak{w1,wk}ak{v1,,vk}k\in[n-1],\{u_{1},\ldots,u_{k}\}\leq_{a_{k}}\{w_{1},\ldots w_{k}\}\leq_{a_{k}}\{v_{1},\ldots,v_{k}\}.

Proof.

(1)(2)(1)\iff(2) is clear by definition. We will show that (1)(3)(4)(1)(1)\implies(3)\implies(4)\implies(1). (3)(4)(3)\implies(4) is immediate given that Lemma 3.7 establishes the existence of aka_{k}’s satisfying {u1,,uk}ak{v1,,vk}\{u_{1},\ldots,u_{k}\}\leq_{a_{k}}\{v_{1},\ldots,v_{k}\}.

(1)(3):(1)\implies(3): Recall that the long cycle τ=(123n)Sn\tau=(123\cdots n)\in S_{n} is an automorphism of the unweighted quantum Bruhat graph (Lemma 2.3). Fix a sequence 𝐚=a1,an1\mathbf{a}=a_{1}\ldots,a_{n-1} such that {u1,,uk}ak{v1,,vk}\{u_{1},\ldots,u_{k}\}\leq_{a_{k}}\{v_{1},\ldots,v_{k}\} for all kk. As wuvw\leq_{u}v, there is a shortest path from uu to vv that passes through ww. As shortest length paths equal minimal weight paths (Lemma 2.4), we also have d(u,v)=d(u,w)+d(w,v)d(u,v)=d(u,w)+d(w,v), where d(u,v)d(u,v) is the exponent vector of the weight of any shortest path from uu to vv. Since τ\tau is an automorphism, the same logic also applies to τ1akwτ1akuτ1akv\tau^{1-a_{k}}w\leq_{\tau^{1-a_{k}}u}\tau^{1-a_{k}}v so

d(τ1aku,τ1akv)=d(τ1aku,τ1akw)+d(τ1akw,τ1akv).d(\tau^{1-a_{k}}u,\tau^{1-a_{k}}v)=d(\tau^{1-a_{k}}u,\tau^{1-a_{k}}w)+d(\tau^{1-a_{k}}w,\tau^{1-a_{k}}v).

By Theorem 3.3, the kthk^{th} coordinate of d(τ1aku,τ1akv)d(\tau^{1-a_{k}}u,\tau^{1-a_{k}}v) is given by the depth of the lattice path P(τ1aku[k],τ1akv[k])P(\tau^{1-a_{k}}u[k],\tau^{1-a_{k}}v[k]) which is 0 because {u1,,uk}ak{v1,,vk}\{u_{1},\ldots,u_{k}\}\leq_{a_{k}}\{v_{1},\ldots,v_{k}\}. As a result, the kthk^{th} coordinate of both d(τ1aku,τ1akw)d(\tau^{1-a_{k}}u,\tau^{1-a_{k}}w) and d(τ1akw,τ1akv)d(\tau^{1-a_{k}}w,\tau^{1-a_{k}}v) must be 0 as well, which translate to {u1,,uk}ak{w1,,wk}\{u_{1},\ldots,u_{k}\}\leq_{a_{k}}\{w_{1},\ldots,w_{k}\} and {w1,,wk}ak{v1,,vk}\{w_{1},\ldots,w_{k}\}\leq_{a_{k}}\{v_{1},\ldots,v_{k}\}.

(4)(1):(4)\implies(1): Fix such a sequence 𝐚\mathbf{a} and look at a coordinate kk. Since u[k]akw[k]akv[k]u[k]\leq_{a_{k}}w[k]\leq_{a_{k}}v[k], the three lattice paths P(u[k],v[k])P(u[k],v[k]), P(u[k],w[k])P(u[k],w[k]) and P(w[k],v[k])P(w[k],v[k]) all reach their lowest points at xx-coordinate ak1a_{k}-1 by Lemma 3.7. Then by construction (3.1),

depth(u[k],v[k])=|{bv[k]:b<ak}||{bu[k]:b<ak}|0.\operatorname{depth}(u[k],v[k])=|\{b\in v[k]:b<a_{k}\}|-|\{b\in u[k]:b<a_{k}\}|\geq 0.

This description allows us to conclude depth(u[k],v[k])=depth(u[k],w[k])+depth(w[k],v[k])\operatorname{depth}(u[k],v[k])=\operatorname{depth}(u[k],w[k])+\operatorname{depth}(w[k],v[k]). Then d(u,v)=d(u,w)+d(w,v)d(u,v)=d(u,w)+d(w,v) by Theorem 3.3 which implies wuvw\leq_{u}v by Lemma 2.4 and Definition 2.6. ∎

The following is immediate from Theorem 3.9:

Corollary 3.10.

Let [x,y][u,v][x,y]\subset[u,v] be any subinterval. Let 𝐚\mathbf{a} be any sequence such that u[k]akv[k]u[k]\leq_{a_{k}}v[k] for all k[n1]k\in[n-1], then for all k[n1]k\in[n-1],

x[k]aky[k].x[k]\leq_{a_{k}}y[k].

4. Tilted Richardson varieties

From this point on, we turn our attention to geometry. The goal of this section is to give three equivalent definitions to (open) tilted Richardson varieties 𝒯u,v\mathcal{T}_{u,v}^{\circ} and 𝒯u,v\mathcal{T}_{u,v} for any pair of permutations u,vSnu,v\in S_{n}. These varieties of interest will later be proved to provide a geometric meaning to the tilted Bruhat order (Theorem 5.2). In particular, if uvu\leq v in the Bruhat order, the (open) tilted Richardson varieties coincide with the usual (open) Richardson varieties u,v\mathcal{R}_{u,v}^{\circ} and u,v\mathcal{R}_{u,v}. These three equivalent definitions use:

  1. (1)

    rank conditions (4.1);

  2. (2)

    cyclically rotated Richardson varieties in the Grassmannian (Theorem 4.6);

  3. (3)

    multi-Plücker coordinates (Theorem 4.12).

As an intermediate step, we first work with 𝒯u,v,𝐚\mathcal{T}^{\circ}_{u,v,\mathbf{a}} and 𝒯u,v,𝐚\mathcal{T}_{u,v,\mathbf{a}} and then show that they do not depend on the choice of sequence 𝐚\mathbf{a} (4.10). Here, given u,vSnu,v\in S_{n} and a sequence 𝐚=(a1,,an1)\mathbf{a}=(a_{1},\ldots,a_{n-1}), we write u𝐚vu\leq_{\mathbf{a}}v if u[k]akv[k]u[k]\leq_{a_{k}}v[k] for all k[n1]k\in[n-1].

When ab[n]a\neq b\in[n], recall the definition of cyclic intervals [a,b]c,[a,b)c,(a,b]c,(a,b)c[a,b]_{c},[a,b)_{c},(a,b]_{c},(a,b)_{c} in Definition 2.2. For convenience, we extend the definition by setting [j,0]c:=[j,n]={j,j+1,,n},[j,j]c:={j}[j,0]_{c}:=[j,n]=\{j,j+1,\dots,n\},[j,j]_{c}:=\{j\} and [j,j)c=(j,j]c:=[j,j)_{c}=(j,j]_{c}:=\emptyset.

4.1. The first definition: rank conditions

Our first definition is motivated by our characterization of tilted Bruhat order (Theorem 3.9).

Definition 4.1.

Define the open tilted Richardson variety with respect to 𝐚\mathbf{a} to be:

(14) 𝒯u,v,𝐚={FFln:dim(Proj[ai,j]c(Fi))=#{u[i][ai,j]c},dim(Proj[j,ai1]c(Fi))=#{v[i][j,ai1]c}i,j[n]}.\displaystyle\mathcal{T}_{u,v,\mathbf{a}}^{\circ}=\left\{F_{\bullet}\in\mathrm{Fl}_{n}:\begin{array}[]{c}\dim(\mathrm{Proj}_{[a_{i},j]_{c}}(F_{i}))=\#\{u[i]\cap[a_{i},j]_{c}\},\\ \dim(\mathrm{Proj}_{[j,a_{i}-1]_{c}}(F_{i}))=\#\{v[i]\cap[j,a_{i}-1]_{c}\}\end{array}\forall i,j\in[n]\right\}.

Define the tilted Richardson variety with respect to 𝐚\mathbf{a} to be:

(17) 𝒯u,v,𝐚={FFln:dim(Proj[ai,j]c(Fi))#{u[i][ai,j]c},dim(Proj[j,ai1]c(Fi))#{v[i][j,ai1]c}i,j[n]}.\displaystyle\mathcal{T}_{u,v,\mathbf{a}}=\left\{F_{\bullet}\in\mathrm{Fl}_{n}:\begin{array}[]{c}\dim(\mathrm{Proj}_{[a_{i},j]_{c}}(F_{i}))\leq\#\{u[i]\cap[a_{i},j]_{c}\},\\ \dim(\mathrm{Proj}_{[j,a_{i}-1]_{c}}(F_{i}))\leq\#\{v[i]\cap[j,a_{i}-1]_{c}\}\end{array}\forall i,j\in[n]\right\}.

Indeed, for any u𝐚vu\leq_{\mathbf{a}}v, dim(Proj[ai,j]c(wB/B))=#{w[i][ao,j]c}\dim(\mathrm{Proj}_{[a_{i},j]_{c}}(wB/B))=\#\{w[i]\cap[a_{o},j]_{c}\}. Therefore

ew𝒯u,v,𝐚w[u,v].e_{w}\in\mathcal{T}_{u,v,\mathbf{a}}\iff w\in[u,v].

In particular, ew𝒯u,v,𝐚e_{w}\in\mathcal{T}_{u,v,\mathbf{a}} is independent of the choice of 𝐚\mathbf{a}. We will see later (4.10) that in fact 𝒯u,v,𝐚\mathcal{T}_{u,v,\mathbf{a}} and 𝒯u,v,𝐚\mathcal{T}_{u,v,\mathbf{a}}^{\circ} are both independent of 𝐚\mathbf{a}.

It is easier to visualize tilted Richardson varieties as follows. For any S[n]S\subseteq[n], k[n]k\in[n] and MMatn×nM\in\mathrm{Mat}_{n\times n}, define rS,k(M)r_{S,k}(M) to be the rank of the submatrix of MM obtained by taking the rows in SS and the left kk columns. Let MFM_{F} be a matrix representative of FF_{\bullet}. Then

(18) F𝒯u,v,𝐚{r[ai,j]c,i(MF)r[ai,j]c,i(u)r[j,ai1]c,i(MF)r[j,ai1]c,i(v), for all i,j[n]\displaystyle F_{\bullet}\in\mathcal{T}_{u,v,\mathbf{a}}\iff\begin{cases}\begin{split}r_{[a_{i},j]_{c},i}(M_{F})&\leq r_{[a_{i},j]_{c},i}(u)\\ r_{[j,a_{i}-1]_{c},i}(M_{F})&\leq r_{[j,a_{i}-1]_{c},i}(v)\end{split}\end{cases},\text{ for all }i,j\in[n]

and F𝒯u,v,𝐚F_{\bullet}\in\mathcal{T}_{u,v,\mathbf{a}}^{\circ} if we replace “\leq” with “==” in (18).

Example 4.2.

Let u=4321u=4321 and v=3142v=3142. Then u𝐚vu\leq_{\mathbf{a}}v for all choices of a1{4},a2{2,3,4}a_{1}\in\{4\},a_{2}\in\{2,3,4\} and a3{2}a_{3}\in\{2\}. See Figure 5 for an illustration where \star and \bullet represent uu and vv respectively. The red horizontal line segment in column kk represent the cutoff of [n][n] under ak\leq_{a_{k}} for different choices of aka_{k}, where k{1,2,3}k\in\{1,2,3\}.

Let 𝐚=(4,4,2)\mathbf{a}=(4,4,2). For F𝒯u,v,𝐚F_{\bullet}\in\mathcal{T}_{u,v,\mathbf{a}}, there are 88 rank conditions imposed on F2F_{2} as in (17). The condition dim(Proj{1,2,3}(F2))2=#v[2]{1,2,3}\dim(\mathrm{Proj}_{\{1,2,3\}}(F_{2}))\leq 2=\#v[2]\cap\{1,2,3\} is interpreted as the rank of the shaded submatrix in Figure 5 being at most 22, the number of \bullet in said region.

4433221111223344\star\star\star\star\bullet\bullet\bullet\bullet
Figure 5. u=4321,v=3142u=4321,v=3142
Remark 4.3.

If uvu\leq v in strong Bruhat order, namely u𝐚vu\leq_{\mathbf{a}}v where 𝐚=(11)\mathbf{a}=(1\ldots 1), then 𝒯u,v,𝐚\mathcal{T}_{u,v,\mathbf{a}}^{\circ} and 𝒯u,v,𝐚\mathcal{T}_{u,v,\mathbf{a}} are the (open) Richardson variety u,v\mathcal{R}_{u,v}^{\circ} and u,v\mathcal{R}_{u,v} respectively.

4.2. The second definition: pullback of cyclically rotated Richardson varieties

For VGr(k,n)V\in\mathrm{Gr}(k,n), let V~\widetilde{V} be any matrix such that VV is the column span of V~\widetilde{V}. Let χ:Gr(k,n)Gr(k,n)\chi:\mathrm{Gr}(k,n)\rightarrow\mathrm{Gr}(k,n) be the cyclic rotation such that for

(19) V~=[v1v2vn], set χ(V~):=[vnv1vn1].\widetilde{V}=\begin{bmatrix}\text{---}&\vec{v}_{1}&\text{---}\\ \text{---}&\vec{v}_{2}&\text{---}\\ \ &\vdots&\ \\ \text{---}&\vec{v}_{n}&\text{---}\\ \end{bmatrix}\text{, set }\chi(\widetilde{V}):=\begin{bmatrix}\text{---}&\vec{v}_{n}&\text{---}\\ \text{---}&\vec{v}_{1}&\text{---}\\ \ &\vdots&\ \\ \text{---}&\vec{v}_{n-1}&\text{---}\\ \end{bmatrix}.

For I={i1,,ik}[n]I=\{i_{1},\dots,i_{k}\}\subset[n], denote χ(I):={i1+1,,ik+1}\chi(I):=\{i_{1}+1,\dots,i_{k}+1\}, identifying n+1n+1 with 11.

Definition 4.4.

For any I,J[n]I,J\subset[n] with |I|=|J||I|=|J| and r[n]r\in[n] such that IrJI\leq_{r}J, define the cyclically rotated (open) Grassmaniann Richardson variety

I,J,r\displaystyle\mathcal{R}^{\circ}_{I,J,r} :=χr1(χ1r(I),χ1r(J)),\displaystyle:=\chi^{r-1}(\mathcal{R}^{\circ}_{\chi^{1-r}(I),\chi^{1-r}(J)}),
I,J,r\displaystyle\mathcal{R}_{I,J,r} :=χr1(χ1r(I),χ1r(J)).\displaystyle:=\chi^{r-1}(\mathcal{R}_{\chi^{1-r}(I),\chi^{1-r}(J)}).

The cyclically rotated Richardson varieties are instances of Positroid varieties (see [9, Section 6]). Similar to Grassmaniann Richardson varieties, cyclically rotated Richardson varieties can also be defined via vanishing of certain Plücker coordinates.

Proposition 4.5 ([9]).

For I,J[n]I,J\subset[n], r[n]r\in[n] with |I|=|J||I|=|J| and IrJI\leq_{r}J,

I,J,r={VGr(k,n):PK(V)=0 for K[I,J]r}.\mathcal{R}_{I,J,r}=\{V\in\mathrm{Gr}(k,n):P_{K}(V)=0\text{ for }K\notin[I,J]_{r}\}.

The corresponding open cell I,J,r=I,J,r{PIPJ0}\mathcal{R}^{\circ}_{I,J,r}=\mathcal{R}_{I,J,r}\cap\{P_{I}P_{J}\neq 0\}.

Proof.

Following from (2.5) and the fact that χ(PK)=Pχ(K)\chi_{\ast}(P_{K})=P_{\chi(K)},

I,J,r={VGr(k,n):Pχr1(K)(V)=0 for K[χ1r(I),χ1r(J)]}.\mathcal{R}_{I,J,r}=\{V\in\mathrm{Gr}(k,n):P_{\chi^{r-1}(K)}(V)=0\text{ for }K\notin[\chi^{1-r}(I),\chi^{1-r}(J)]\}.

The statement then follows from the fact that ArBχ1r(A)χ1r(B)A\leq_{r}B\iff\chi^{1-r}(A)\leq\chi^{1-r}(B). ∎

For each k[n1]k\in[n-1], define πk:FlnGr(k,n)\pi_{k}:\mathrm{Fl}_{n}\rightarrow\mathrm{Gr}(k,n) to be the projection onto the kk-th flag.

Theorem 4.6.

For u𝐚vu\leq_{\mathbf{a}}v, we have

𝒯u,v,𝐚=k=1n1πk1(u[k],v[k],ak) and 𝒯u,v,𝐚=k=1n1πk1(u[k],v[k],ak).\mathcal{T}_{u,v,\mathbf{a}}=\bigcap_{k=1}^{n-1}\pi_{k}^{-1}(\mathcal{R}_{u[k],v[k],a_{k}})\quad\text{ and }\quad\mathcal{T}^{\circ}_{u,v,\mathbf{a}}=\bigcap_{k=1}^{n-1}\pi_{k}^{-1}(\mathcal{R}^{\circ}_{u[k],v[k],a_{k}}).
Proof.

This is done by unpacking the rank conditions and comparing with (14) and (17). ∎

Since each πk1(u[k],v[k],ak)\pi_{k}^{-1}(\mathcal{R}_{u[k],v[k],a_{k}}) is a closed subvariety of Fln\mathrm{Fl}_{n} and u[k],v[k],aku[k],v[k],ak\mathcal{R}^{\circ}_{u[k],v[k],a_{k}}\subseteq\mathcal{R}_{u[k],v[k],a_{k}} is open, we have the following corollary.

Corollary 4.7.

𝒯u,v,𝐚\mathcal{T}_{u,v,\mathbf{a}} is a closed subvariety of Fln\mathrm{Fl}_{n} and 𝒯u,v,𝐚𝒯u,v,𝐚\mathcal{T}_{u,v,\mathbf{a}}^{\circ}\subseteq\mathcal{T}_{u,v,\mathbf{a}} is open.

Let 𝐚=(a1an1)\mathbf{a}=(a_{1}\dots a_{n-1}) and 𝐚=(a1an1)\mathbf{a}^{\prime}=(a_{1}^{\prime}\ldots a_{n-1}^{\prime}) be any two sequences such that u𝐚vu\leq_{\mathbf{a}}v and u𝐚vu\leq_{\mathbf{a}^{\prime}}v. Our next goal is to assert that the (open) tilted Richardson variety is independent of the choice of 𝐚\mathbf{a}. We first need the following lemma. We note that this can be seen as an alternate interpretation of Lemma 3.8.

Lemma 4.8.

For I,J[n]I,J\subset[n] with |I|=|J|=k|I|=|J|=k, suppose IrJI\leq_{r}J and IrJI\leq_{r^{\prime}}J for some rr[n]r\neq r^{\prime}\in[n]. Then I,J,r=I,J,r\mathcal{R}_{I,J,r}=\mathcal{R}_{I,J,r^{\prime}} and I,J,r=I,J,r\mathcal{R}^{\circ}_{I,J,r}=\mathcal{R}^{\circ}_{I,J,r^{\prime}}. Furthermore, take any VI,J,aV\in\mathcal{R}_{I,J,a} and its n×kn\times k matrix representative MVM_{V} whose column span equals VV. Then

span{vi:i[r,r)c}span{vi:i[r,r)c}={0},\mathrm{span}\{\vec{v}_{i}:i\in[r,r^{\prime})_{c}\}\cap\mathrm{span}\{\vec{v}_{i}:i\in[r^{\prime},r)_{c}\}=\{0\},

where vi\vec{v_{i}} is the ii-th row of MVM_{V}.

Proof.

Since I,J,r=I,J,r{PIPJ0}\mathcal{R}^{\circ}_{I,J,r}=\mathcal{R}_{I,J,r}\cap\{P_{I}P_{J}\neq 0\}, we only need to prove the closed case. From Proposition 4.5, we have

I,J,r\displaystyle\mathcal{R}_{I,J,r} ={VGr(k,n):PK(V)=0 for K[I,J]r},\displaystyle=\{V\in\mathrm{Gr}(k,n):P_{K}(V)=0\text{ for }K\notin[I,J]_{r}\},
I,J,r\displaystyle\mathcal{R}_{I,J,r^{\prime}} ={VGr(k,n):PK(V)=0 for K[I,J]r}.\displaystyle=\{V\in\mathrm{Gr}(k,n):P_{K}(V)=0\text{ for }K\notin[I,J]_{r^{\prime}}\}.

Since Lemma 3.8 implies [I,J]r=[I,J]r[I,J]_{r}=[I,J]_{r^{\prime}}, we have I,J,r=I,J,r\mathcal{R}_{I,J,r}=\mathcal{R}_{I,J,r^{\prime}}.

For the second part, Lemma 3.8 implies that there exists some integer 0dk0\leq d\leq k satisfying

#(I[r,r)c)\displaystyle\#(I\cap[r,r^{\prime})_{c}) =#(J[r,r)c)=d,\displaystyle=\#(J\cap[r,r^{\prime})_{c})=d,
#(I[r,r)c)\displaystyle\#(I\cap[r^{\prime},r)_{c}) =#(J[r,r)c)=kd.\displaystyle=\#(J\cap[r^{\prime},r)_{c})=k-d.

By definition, for any VI,J,rV\in\mathcal{R}^{\circ}_{I,J,r}, we have

#(I[r,r)c)=d\displaystyle\#(I\cap[r,r^{\prime})_{c})=d dim(span{vi:i[r,r)c})d,\displaystyle\implies\dim(\mathrm{span}\{\vec{v}_{i}:i\in[r,r^{\prime})_{c}\})\leq d,
#(J[r,r)c)=kd\displaystyle\#(J\cap[r^{\prime},r)_{c})=k-d dim(span{vi:i[r,r)c})kd.\displaystyle\implies\dim(\mathrm{span}\{\vec{v}_{i}:i\in[r^{\prime},r)_{c}\})\leq k-d.

Notice that

dim(span{vi:i[r,r)c})+dim(span{vi:i[r,r)c})dim(V)=k,\dim(\mathrm{span}\{\vec{v}_{i}:i\in[r,r^{\prime})_{c}\})+\dim(\mathrm{span}\{\vec{v}_{i}:i\in[r^{\prime},r)_{c}\})\geq\dim(V)=k,

and equality holds if and only if the two subspaces span{vi:i[r,r)c}\mathrm{span}\{\vec{v}_{i}:i\in[r,r^{\prime})_{c}\} and span{vi:i[r,r)c}\mathrm{span}\{\vec{v}_{i}:i\in[r^{\prime},r)_{c}\} are linearly independent. This concludes the proof. ∎

Example 4.9.

Continuing Example 4.2, Lemma 4.8 implies that both the green shaded area and the purple hatched area in Figure 6 have rank 11. Moreover, the row spans of the two areas are independent.

4433221111223344\star\star\star\star\bullet\bullet\bullet\bullet
Figure 6. The two horizontal red line segments represents two choices of a2a_{2}. Row spans of the green area and purple area are independent.
Corollary 4.10.

𝒯u,v,𝐚=𝒯u,v,𝐚\mathcal{T}_{u,v,\mathbf{a}}^{\circ}=\mathcal{T}_{u,v,\mathbf{a}^{\prime}}^{\circ} and 𝒯u,v,𝐚=𝒯u,v,𝐚\mathcal{T}_{u,v,\mathbf{a}}=\mathcal{T}_{u,v,\mathbf{a}^{\prime}} for u𝐚vu\leq_{\mathbf{a}}v and u𝐚vu\leq_{\mathbf{a}^{\prime}}v.

Proof.

This follows from Proposition 4.6 and Lemma 4.8. ∎

As a consequence of 4.10, we will denote the (open) tilted Richardson variety as 𝒯u,v\mathcal{T}_{u,v} (𝒯u,v\mathcal{T}_{u,v}^{\circ}). In particular, if uvu\leq v in strong Bruhat order, 𝒯u,v=u,v\mathcal{T}_{u,v}^{\circ}=\mathcal{R}_{u,v}^{\circ} and 𝒯u,v=u,v\mathcal{T}_{u,v}=\mathcal{R}_{u,v}.

4.3. The third definition: vanishing of Plücker coordinates

Lemma 4.11.

For any flag FFlnF_{\bullet}\in\mathrm{Fl}_{n} and subset I([n]k)I\in\binom{[n]}{k}, if PI(F)0P_{I}(F_{\bullet})\neq 0, then there exists a permutation wSnw\in S_{n} such that w[k]=Iw[k]=I, and Pw(F)0P_{w}(F_{\bullet})\neq 0.

Proof.

Let MFGM_{F}\in G be an matrix representative of FF_{\bullet} such that FiF_{i} is the span of the first ii column vectors of MFM_{F}. The k×kk\times k submatrix MM^{\prime} of MFM_{F} in the rows indexed by II in the first kk columns has full rank. Therefore, there exists a (k1)×(k1)(k-1)\times(k-1) submatrix of MM^{\prime} in the first (k1)(k-1) columns that also has full rank. Denote the row index of this submatrix as II^{\prime}. Then |I|=k1|I^{\prime}|=k-1 and PI(F)0P_{I^{\prime}}(F_{\bullet})\neq 0.

Apply this process inductively to find a chain of subsets =I0I1Ik1=IIk=I\emptyset=I_{0}\subsetneq I_{1}\subsetneq\cdots\subsetneq I_{k-1}=I^{\prime}\subsetneq I_{k}=I such that PIj(F)0P_{I_{j}}(F_{\bullet})\neq 0 for any 0jk0\leq j\leq k. Using similar idea, we can also find a chain I=IkIk+1In=[n]I=I_{k}\subsetneq I_{k+1}\subsetneq\cdots\subsetneq I_{n}=[n] such that PIj(F)0P_{I_{j}}(F_{\bullet})\neq 0 for kjnk\leq j\leq n. Let wSnw\in S_{n} be the permutation such that wk=IkIk1w_{k}=I_{k}\setminus I_{k-1}. Then Pw(F)0P_{w}(F_{\bullet})\neq 0 and w[k]=Iw[k]=I. ∎

Theorem 4.12.

For any permutations u,vSnu,v\in S_{n}, we have

𝒯u,v={FFln:Pw(F)=0 for w[u,v]},\mathcal{T}_{u,v}=\{F_{\bullet}\in\mathrm{Fl}_{n}:P_{w}(F_{\bullet})=0\text{ for }w\notin[u,v]\},

and the corresponding open cell 𝒯u,v=𝒯u,v{PuPv0}\mathcal{T}^{\circ}_{u,v}=\mathcal{T}_{u,v}\cap\{P_{u}P_{v}\neq 0\}.

Proof.

Fix a sequence 𝐚\mathbf{a} such that u𝐚vu\leq_{\mathbf{a}}v. For any F𝒯u,vF_{\bullet}\in\mathcal{T}_{u,v} and w[u,v]w\notin[u,v], we first show Pw(F)=0P_{w}(F_{\bullet})=0. By Theorem 3.9, there exists k[n1]k\in[n-1] such that w[k][u[k],v[k]]akw[k]\notin[u[k],v[k]]_{a_{k}}. Since Fku[k],v[k],akF_{k}\in\mathcal{R}_{u[k],v[k],a_{k}}, by Proposition 4.5, Pw[k](Fk)=0P_{w[k]}(F_{k})=0 and Pw(F)=0P_{w}(F_{\bullet})=0. We have proven \subseteq.

We now prove \supseteq. For any F𝒯u,v,𝐚F_{\bullet}\notin\mathcal{T}_{u,v,\mathbf{a}}, there exists k[n1]k\in[n-1] such that Fku[k],v[k],akF_{k}\notin\mathcal{R}_{u[k],v[k],a_{k}}. By Proposition 4.5, there exists subset I([n]k)I\in{[n]\choose k} such that I[u[k],v[k]]akI\notin[u[k],v[k]]_{a_{k}} and PI(Fk)0P_{I}(F_{k})\neq 0. By Lemma 4.11, there exists a permutation wSnw\in S_{n} such that Pw(F)0P_{w}(F_{\bullet})\neq 0 and w[k]=Iw[k]=I. However, I[u[k],v[k]]akI\notin[u[k],v[k]]_{a_{k}} implies w[u,v]w\notin[u,v].

For the open cell, notice that

u[k],v[k],ak=u[k],v[k],ak{Pu[k]Pv[k]0}.\mathcal{R}^{\circ}_{u[k],v[k],a_{k}}=\mathcal{R}_{u[k],v[k],a_{k}}\cap\{P_{u[k]}P_{v[k]}\neq 0\}.

Therefore, by Theorem 4.6, 𝒯u,v=𝒯u,v{PuPv=k=1n1Pu[k]Pv[k]0}\mathcal{T}^{\circ}_{u,v}=\mathcal{T}_{u,v}\cap\{P_{u}P_{v}=\prod_{k=1}^{n-1}P_{u[k]}P_{v[k]}\neq 0\} as desired. ∎

5. Geometric properties of tilted Richardson varieties

In this section, we explore some further geometric properties of the tilted Richardson varieties, analogous to those of the classical Richardson varieties. We do note, however, that most of our analysis are significantly different and more technical than the classical case, due to the lack of an analogue notion of Schubert varieties and the constant usage of the sequence 𝐚=(a1,,an1)\mathbf{a}=(a_{1},\ldots,a_{n-1}) that keeps varying.

5.1. Stratification of 𝒯u,v\mathcal{T}_{u,v} by open tilted Richardson varieties

We start with a useful lemma that controls the choices of aka_{k} to some extent.

Lemma 5.1.

For any permutation u,vu,v and any k[n2]k\in[n-2], there exists aka_{k} such that u[k]akv[k]u[k]\leq_{a_{k}}v[k] and u[k+1]akv[k+1]u[k+1]\leq_{a_{k}}v[k+1].

Proof.

Consider the lattice path Pk:=P(u[k],v[k])P_{k}:=P(u[k],v[k]). Since u[k+1]=u[k]{u(k+1)}u[k+1]=u[k]\cup\{u(k+1)\} and v[k+1]=v[k]{v(k+1)}v[k+1]=v[k]\cup\{v(k+1)\}, we obtain Pk+1:=P(u[k+1],v[k+1])P_{k+1}:=P(u[k+1],v[k+1]) from PkP_{k} by,

  • -

    at the u(k+1)u(k+1)-th step, changing a \searrow to a \rightarrow or changing a \rightarrow to a \nearrow; and

  • -

    at the v(k+1)v(k+1)-th step, changing a \nearrow to a \rightarrow or changing a \rightarrow to a \searrow.

In other words, we obtain Pk+1P_{k+1} by moving the lattice points

(20) (a,b)Pk{(a,b+1) if u(k+1)a<v(k+1)(a,b1) if v(k+1)a<u(k+1)(a,b) otherwise.\displaystyle(a,b)\in P_{k}\longrightarrow\begin{cases}(a,b+1)&\text{ if }u(k+1)\leq a<v(k+1)\\ (a,b-1)&\text{ if }v(k+1)\leq a<u(k+1)\\ (a,b)&\text{ otherwise}\end{cases}.

Consider now Ak:={ak1:u[k]akv[k]}A_{k}:=\{a_{k}-1:u[k]\leq_{a_{k}}v[k]\}. By Lemma 3.7, u[k]akv[k]u[k]\leq_{a_{k}}v[k] if and only if PkP_{k} hits the its lowest point with xx-coordinate in AkA_{k}. Since the yy-coordinate changes by at most 11 from PkP_{k} to Pk+1P_{k+1}, if Ak+1Ak=A_{k+1}\cap A_{k}=\emptyset, then all lattice points in PkP_{k} with xx-coordinate ak1a_{k}-1 are moved up by 11 and all lattice points with xx-coordinate ak+11a_{k+1}-1 are moved down by 11. This is impossible by (20). Therefore AkAk+1A_{k}\cap A_{k+1}\neq\emptyset and there exists aka_{k} such that u[k]akv[k]u[k]\leq_{a_{k}}v[k] and u[k+1]akv[k+1]u[k+1]\leq_{a_{k}}v[k+1]. ∎

Theorem 5.2.

𝒯u,v=[x,y][u,v]𝒯x,y\mathcal{T}_{u,v}=\bigsqcup_{[x,y]\subseteq[u,v]}\mathcal{T}_{x,y}^{\circ}.

Proof.

It is straightforward from definition that the right hand side is a disjoint union and that 𝒯x,y𝒯u,v\mathcal{T}_{x,y}^{\circ}\subset\mathcal{T}_{u,v} for all [x,y][u,v][x,y]\subseteq[u,v]. To prove the \subsetneq direction, pick F𝒯u,vF_{\bullet}\in\mathcal{T}_{u,v}. We will show that F𝒯x,yF_{\bullet}\in\mathcal{T}_{x,y}^{\circ} for some [x,y][u,v][x,y]\subseteq[u,v]. Pick 𝐚\mathbf{a} such that u𝐚vu\leq_{\mathbf{a}}v, for k[n1]k\in[n-1], set

Ik=\displaystyle I_{k}= {i[n]:dim(Proj[ak,j)c(Fk))<dim(Proj[ak,j]c(Fk))},\displaystyle\{i\in[n]:\dim(\mathrm{Proj}_{[a_{k},j)_{c}}(F_{k}))<\dim(\mathrm{Proj}_{[a_{k},j]_{c}}(F_{k}))\},
Jk=\displaystyle J_{k}= {j[n]:dim(Proj(i,ak1]c(Fk))<dim(Proj[i,ak1]c(Fk))}.\displaystyle\{j\in[n]:\dim(\mathrm{Proj}_{(i,a_{k}-1]_{c}}(F_{k}))<\dim(\mathrm{Proj}_{[i,a_{k}-1]_{c}}(F_{k}))\}.

Then FkIk,Jk,akF_{k}\in\mathcal{R}^{\circ}_{I_{k},J_{k},a_{k}}. In particular, IkakJkI_{k}\leq_{a_{k}}J_{k} for all k[n]k\in[n]. Furthermore, Fku[k],v[k],akF_{k}\in\mathcal{R}_{u[k],v[k],a_{k}} implies that [I,J]ak[u[k],v[k]]ak[I,J]_{a_{k}}\subseteq[u[k],v[k]]_{a_{k}} is a subinterval in the shifted Gale order ak\leq_{a_{k}}.

Fix k[n1]k\in[n-1], by Lemma 3.8 and 4.10, the sets IkI_{k} and JkJ_{k} are independent of the choice of aka_{k} as long as u[k]akv[k]u[k]\leq_{a_{k}}v[k]. Combining with Lemma 5.1, we can choose 𝐚\mathbf{a} such that ak=ak+1a_{k}=a_{k+1}. Let MGLnM\in{GL}_{n} be any matrix representative of FF_{\bullet} and let MkM_{k} and Mk+1M_{k+1} denote the submatrix of MM consisting of the first kk and k+1k+1 columns respectively. If

dim(Proj[ak+1,j)c(Fk+1))=dim(Proj[ak+1,j]c(Fk+1)),\dim(\mathrm{Proj}_{[a_{k+1},j)_{c}}(F_{k+1}))=\dim(\mathrm{Proj}_{[a_{k+1},j]_{c}}(F_{k+1})),

then row jj of Mk+1M_{k+1} lies in the row span of Mk+1M_{k+1} with row index in [ak+1,j)c[a_{k+1},j)_{c}. Thus row jj of MkM_{k} also lies in the row span of MkM_{k} with row index in [ak,j)c[a_{k},j)_{c}. This implies that IkIk+1I_{k}\subset I_{k+1}. The claim JkJk+1J_{k}\subset J_{k+1} follows from similar reasoning.

Let x,ySnx,y\in S_{n} be permutations such that xk=IkIk1x_{k}=I_{k}\setminus I_{k-1} and yk=JkJk1y_{k}=J_{k}\setminus J_{k-1} for all k[n]k\in[n]. It then follows from definition of IkI_{k} and JkJ_{k} that F𝒯x,yF_{\bullet}\in\mathcal{T}_{x,y}^{\circ} where u𝐚xu\leq_{\mathbf{a}}x and y𝐚vy\leq_{\mathbf{a}}v. Since x[k]=IkakJk=y[k]x[k]=I_{k}\leq_{a_{k}}J_{k}=y[k] for all k[n1]k\in[n-1], namely x𝐚yx\leq_{\mathbf{a}}y, we conclude that any F𝒯u,vF_{\bullet}\in\mathcal{T}_{u,v} lies in 𝒯x,y\mathcal{T}_{x,y}^{\circ} for some [x,y][u,v][x,y]\subset[u,v]. ∎

5.2. Dimension of 𝒯u,v\mathcal{T}_{u,v}^{\circ} and 𝒯u,v\mathcal{T}_{u,v}

In this subsection, we show that 𝒯u,v\mathcal{T}_{u,v}^{\circ} and 𝒯u,v\mathcal{T}_{u,v} have dimension (u,v)\ell(u,v) (Theorem 5.14), the distance between u,vu,v in the quantum Bruhat graph (Section 2). We do this by analyzing 𝒯u,v\mathcal{T}_{u,v}^{\circ} from different charts.

For any permutation xx, the permuted opposite Schubert cell xΩidx\Omega_{id}^{\circ} is the set of flags satisfying Px0P_{x}\neq 0. These could also be realized as matrices with 11’s in (xi,i)(x_{i},i) for all i[n]i\in[n] and 0’s to the right of 11’s.

Example 5.3.

Let x=3142x=3142, then xΩidx\Omega_{id}^{\circ} can be identified as the following affine space with each * being a copy of \mathbb{C}:

443322111122334411111111000000******
Proposition 5.4.

For any permutations u,vSnu,v\in S_{n},

  1. (1)

    if x[u,v]x\notin[u,v], then 𝒯u,vxΩid=𝒯u,vxΩid=\mathcal{T}_{u,v}\cap x\Omega_{id}^{\circ}=\mathcal{T}_{u,v}^{\circ}\cap x\Omega_{id}^{\circ}=\emptyset,

  2. (2)

    𝒯u,v=𝒯u,vuΩidvΩid\mathcal{T}_{u,v}^{\circ}=\mathcal{T}_{u,v}\cap u\Omega_{id}^{\circ}\cap v\Omega_{id}^{\circ}.

Proof.

This follows directly from Theorem 4.12. ∎

Definition 5.5.

For u𝐚vu\leq_{\mathbf{a}}v, define D𝐚(u),D𝐚(v)[n]2D_{\mathbf{a}}^{\downarrow}(u),D_{\mathbf{a}}^{\uparrow}(v)\subset[n]^{2} to be the tilted Rothe diagrams

D𝐚(u)=\displaystyle D_{\mathbf{a}}^{\downarrow}(u)= {(i,k):i<akuk,u1(i)>k},\displaystyle\{(i,k):i<_{a_{k}}u_{k},u^{-1}(i)>k\},
D𝐚(v)=\displaystyle D_{\mathbf{a}}^{\uparrow}(v)= {(i,k):i>akvk,v1(i)>k}.\displaystyle\{(i,k):i>_{a_{k}}v_{k},v^{-1}(i)>k\}.
Example 5.6.

We demonstrate 5.5 with u=4321,v=3142u=4321,v=3142 and 𝐚=(4,4,2)\mathbf{a}=(4,4,2) using diagrams in Figure 7(A) and Figure 7(B). Put \bullet in position (uk,k)(u_{k},k) and (vk,k)(v_{k},k). Draw a red horizontal line in column kk immediately above aka_{k}. These red lines represent the “floor” in each column. For D𝐚(u)D_{\mathbf{a}}^{\downarrow}(u), draw death rays to the right and down of each \bullet until they hit the red horizontal line or right boundary. Similarly, draw death rays to the right and up for D𝐚(v)D_{\mathbf{a}}^{\uparrow}(v). The boxes remain are D𝐚(u)D_{\mathbf{a}}^{\downarrow}(u) and D𝐚(v)D_{\mathbf{a}}^{\uparrow}(v).

4433221111223344\bullet\bullet\bullet\bullet
(A) D𝐚(u)={(1,2),(2,2)}D_{\mathbf{a}}^{\downarrow}(u)=\{(1,2),(2,2)\}
4433221111223344\bullet\bullet\bullet\bullet
(B) D𝐚(v)={(2,2)}D_{\mathbf{a}}^{\uparrow}(v)=\{(2,2)\}

One may notice similarities between Figure 7(A) and Rothe diagram of a permutation. Indeed, if 𝐚=(1,,1)\mathbf{a}=(1,\dots,1), then D𝐚(u)D_{\mathbf{a}}^{\downarrow}(u) is precisely the Rothe diagram D(u1)D(u^{-1}). Moreover, on uΩidu\Omega_{id}^{\circ}, the vanishing of Plücker coordinates as in 5.7(1) gives us the parametrization of Ωu\Omega_{u}^{\circ} (see e.g. [13, Section 2.2]).

Definition 5.7.

For every pair of permutations (u,v)(u,v) and every choice of 𝐚=(a1,,an1)\mathbf{a}=(a_{1},\dots,a_{n-1}) such that u𝐚vu\leq_{\mathbf{a}}v, define u,v,𝐚\mathcal{I}_{u,v,\mathbf{a}} as the following set of equations:

  1. (1)

    for every box (i,k)D𝐚(u)(i,k)\in D_{\mathbf{a}}^{\downarrow}(u), set Pu[k1]+i=0P_{u[k-1]+i}=0,

  2. (2)

    for every box (i,k)D𝐚(v)(i,k)\in D_{\mathbf{a}}^{\uparrow}(v), set Pv[k1]+i=0P_{v[k-1]+i}=0.

Remark 5.8.

An alternative way to phrase u,v,𝐚\mathcal{I}_{u,v,\mathbf{a}} is, for column k[n1]k\in[n-1]:

  • under the chart uΩidu\Omega_{id}^{\circ} as in Example 5.3, set entries in the diagram Da(u)D^{\downarrow}_{\mathrm{a}}(u) to be 0.

  • under the chart vΩidv\Omega_{id}^{\circ}, set entries in the diagram Da(v)D^{\uparrow}_{\mathrm{a}}(v) to be 0.

Up until now, we have been working with any sequence 𝐚\mathbf{a} such that u𝐚vu\leq_{\mathbf{a}}v. There are many such choices for 𝐚\mathbf{a} and any one of them works. However, we are now going to distinguish some choices from the rest, that are particularly nice.

Definition 5.9.

For a pair of permutations u,vu,v, a sequence 𝐚=(a1,,an1)\mathbf{a}=(a_{1},\dots,a_{n-1}) is flat if

  • u𝐚vu\leq_{\mathbf{a}}v and

  • u[k1]akv[k1]u[k-1]\leq_{a_{k}}v[k-1] for all k[2,n1]k\in[2,n-1].

We note that this is equivalent to choosing ak(uk,vk]ca_{k}\in(u_{k},v_{k}]_{c} whenever possible. These are also the choices of 𝐚\mathbf{a}’s such that the number of equations in u,v,𝐚\mathcal{I}_{u,v,\mathbf{a}} is maximized.

Lemma 5.10.

For any pair of permutations u,vu,v, there exists 𝐚\mathbf{a} that is flat. Moreover, for any such 𝐚\mathbf{a}, it is also flat with respect to all pairs of (x,y)(x,y) such that [x,y][u,v][x,y]\subseteq[u,v].

Proof.

This follows from Lemma 5.1 and Corollary 3.10. ∎

Lemma 5.11.

For any u,vu,v and any flat 𝐚\mathbf{a}, there are (n2)(u,v){n\choose 2}-\ell(u,v) equations in u,v,𝐚\mathcal{I}_{u,v,\mathbf{a}}.

Proof.

We proceed by induction on (u,v)\ell(u,v). If (u,v)=0\ell(u,v)=0, we have u=vu=v and any sequence 𝐚\mathbf{a} is flat. For any k[n1]k\in[n-1] and any aka_{k}, the number of entries in column kk of D𝐚(u)D_{\mathbf{a}}^{\downarrow}(u) and D𝐚(v)D_{\mathbf{a}}^{\uparrow}(v) sum up to nkn-k. Therefore u,v,𝐚\mathcal{I}_{u,v,\mathbf{a}} consists of (n2){n\choose 2} equations.

Suppose now that the statement holds when (u,v)=\ell(u,v)=\ell. For any pair u,vu,v such that (u,v)=+1\ell(u,v)=\ell+1, let x=vtp,q[u,v]x=vt_{p,q}\in[u,v] be any permutation such that (u,x)=\ell(u,x)=\ell. Fix any flat 𝐚\mathbf{a} with respect to uu and vv, by Lemma 5.10, 𝐚\mathbf{a} is also flat for u,xu,x. We are then left to show that there is one more equation from Definition 5.7 (2)(2) in u,x,𝐚\mathcal{I}_{u,x,\mathbf{a}} than in u,v,𝐚\mathcal{I}_{u,v,\mathbf{a}}. We compare the diagrams D𝐚(v)D_{\mathbf{a}}^{\uparrow}(v) and D𝐚(x)D_{\mathbf{a}}^{\uparrow}(x) in each column kk.

For k<pk<p or k>qk>q, since vk=xkv_{k}=x_{k} and v[k]=x[k]v[k]=x[k],

(21) (i,k)D𝐚(v)(i,k)D𝐚(x).(i,k)\in D_{\mathbf{a}}^{\uparrow}(v)\iff(i,k)\in D_{\mathbf{a}}^{\uparrow}(x).

Consider now k[p+1,q1]k\in[p+1,q-1]. In this case we still have xk=vkx_{k}=v_{k}. Since xv=xtp,qx\rightarrow v=xt_{p,q} is an edge in quantum Bruhat graph, we have xk[xp,xq]cx_{k}\notin[x_{p},x_{q}]_{c}. Since x<𝐚vx<_{\mathbf{a}}v, for k[p,q1]k\in[p,q-1], x[k]<akv[k]=x[k]{xp}{xq}x[k]<_{a_{k}}v[k]=x[k]\setminus\{x_{p}\}\cup\{x_{q}\}. Therefore xp<akxqx_{p}<_{a_{k}}x_{q} and thus ak(xp,xq]ca_{k}\notin(x_{p},x_{q}]_{c}. Therefore either both xp,xq[xk,ak1]cx_{p},x_{q}\in[x_{k},a_{k}-1]_{c}, in which case

(22) {i:(i,k)D𝐚(v)}={i:(i,k)D𝐚(x)}{xp}{xq},\{i:(i,k)\in D_{\mathbf{a}}^{\uparrow}(v)\}=\{i:(i,k)\in D_{\mathbf{a}}^{\uparrow}(x)\}\setminus\{x_{p}\}\cup\{x_{q}\},

or xp,xq[xk,ak1]cx_{p},x_{q}\notin[x_{k},a_{k}-1]_{c}, in which case

(23) {i:(i,k)D𝐚(v)}={i:(i,k)D𝐚(x)}.\{i:(i,k)\in D_{\mathbf{a}}^{\uparrow}(v)\}=\{i:(i,k)\in D_{\mathbf{a}}^{\uparrow}(x)\}.

In both cases, D𝐚(v)D_{\mathbf{a}}^{\uparrow}(v) and D𝐚(x)D_{\mathbf{a}}^{\uparrow}(x) have the same number of boxes in column kk.

We are left to consider column pp and qq. Since xp<apxqx_{p}<_{a_{p}}x_{q} and x[p1]=v[p1]x[p-1]=v[p-1],

(24) {i:(i,p)D𝐚(x)}={i:(i,p)D𝐚(v)}{i[xp,xq]c:x1(i)>p}.\{i:(i,p)\in D_{\mathbf{a}}^{\uparrow}(x)\}=\{i:(i,p)\in D_{\mathbf{a}}^{\uparrow}(v)\}\sqcup\{i\in[x_{p},x_{q}]_{c}:x^{-1}(i)>p\}.

Since 𝐚\mathbf{a} is flat, x[q1]<aqv[q1]=x[q1]{xp}{xq}x[q-1]<_{a_{q}}v[q-1]=x[q-1]\setminus\{x_{p}\}\cup\{x_{q}\} and thus xp<aqxqx_{p}<_{a_{q}}x_{q}. Therefore

(25) {i:(i,q)D𝐚(v)}={i:(i,q)D𝐚(x)}{i[xp,xq]c:x1(i)>q}.\{i:(i,q)\in D_{\mathbf{a}}^{\uparrow}(v)\}=\{i:(i,q)\in D_{\mathbf{a}}^{\uparrow}(x)\}\sqcup\{i\in[x_{p},x_{q}]_{c}:x^{-1}(i)>q\}.

Since xk[xp,xq]cx_{k}\notin[x_{p},x_{q}]_{c} for all k[p+1,q1]k\in[p+1,q-1],

{i[xp,xq]c:x1(i)>p}={i[xp,xq]c:x1(i)>q}{xq}.\{i\in[x_{p},x_{q}]_{c}:x^{-1}(i)>p\}=\{i\in[x_{p},x_{q}]_{c}:x^{-1}(i)>q\}\cup\{x_{q}\}.

Therefore D𝐚(x)}D_{\mathbf{a}}^{\uparrow}(x)\} has one more box than D𝐚(v)}D_{\mathbf{a}}^{\uparrow}(v)\} in column pp and qq combined.

Combining Equations (21)-(25), we conclude that there is one more equation from Definition 5.7 (2)(2) in u,x,𝐚\mathcal{I}_{u,x,\mathbf{a}} than in u,v,𝐚\mathcal{I}_{u,v,\mathbf{a}}. We are then done by induction. ∎

Example 5.12.

We use the following example to help understand the proof of Lemma 5.11. Let v=465123,u=263145,x=265143v=465123,u=263145,x=265143 and 𝐚=(2,2,2,6,6)\mathbf{a}=(2,2,2,6,6). One can verify that x[u,v]x\in[u,v] and that u𝐚vu\leq_{\mathbf{a}}v. We draw D𝐚(v)D_{\mathbf{a}}^{\uparrow}(v) and D𝐚(x)D_{\mathbf{a}}^{\uparrow}(x) as in 5.6. Notice that both diagrams have the same number of boxes in columns 2,3,4,62,3,4,6, and the number of boxes in column 1,51,5 combined differ by 11.

665544332211112233445566\bullet\bullet\bullet\bullet\bullet\bullet
(A) D𝐚(v)D_{\mathbf{a}}^{\uparrow}(v)
665544332211112233445566\bullet\bullet\bullet\bullet\bullet\bullet
(B) D𝐚(x)D_{\mathbf{a}}^{\uparrow}(x)
Proposition 5.13.

For any u,vu,v and any flat 𝐚\mathbf{a}, 𝒯u,v=uΩidvΩidV(u,v,𝐚)\mathcal{T}^{\circ}_{u,v}=u\Omega_{id}^{\circ}\cap v\Omega_{id}^{\circ}\cap V(\mathcal{I}_{u,v,\mathbf{a}}), where V(u,v,𝐚)V(\mathcal{I}_{u,v,\mathbf{a}}) is the vanishing locus of equations in u,v,𝐚\mathcal{I}_{u,v,\mathbf{a}}.

Proof.

We first show \subseteq. This is equivalent to proving that for every flag F𝒯u,vF_{\bullet}\in\mathcal{T}^{\circ}_{u,v}, FF_{\bullet} satisfies the equations in u,v,𝐚\mathcal{I}_{u,v,\mathbf{a}}. There are two type of equations in u,v,𝐚\mathcal{I}_{u,v,\mathbf{a}}:

Type (1). For (i,k)D𝐚(u)(i,k)\in D^{\downarrow}_{\mathbf{a}}(u), since i<akuku[k1]{i}<aku[k]i<_{a_{k}}u_{k}\implies u[k-1]\cup\{i\}<_{a_{k}}u[k], by Proposition 4.5

Fπk1(u[k],v[k],ak)Pu[k1]+i(F)=0;F_{\bullet}\in\pi_{k}^{-1}(\mathcal{R}_{u[k],v[k],a_{k}})\implies P_{u[k-1]+i}(F_{\bullet})=0;

Type (2). For (i,k)D𝐚(v)(i,k)\in D^{\uparrow}_{\mathbf{a}}(v), since i>akvkv[k1]{i}>akv[k]i>_{a_{k}}v_{k}\implies v[k-1]\cup\{i\}>_{a_{k}}v[k], by Proposition 4.5

Fπk1(u[k],v[k],ak)Pv[k1]+i(F)=0.F_{\bullet}\in\pi_{k}^{-1}(\mathcal{R}_{u[k],v[k],a_{k}})\implies P_{v[k-1]+i}(F_{\bullet})=0.

This implies 𝒯u,vV(u,v,𝐚)\mathcal{T}^{\circ}_{u,v}\subseteq V(\mathcal{I}_{u,v,\mathbf{a}}). Combining with Proposition 5.4, we are done.

We now show \supseteq. It is enough to show that πk1(u[k],v[k],ak)V(u,v,𝐚)\pi_{k}^{-1}(\mathcal{R}_{u[k],v[k],a_{k}})\supseteq V(\mathcal{I}_{u,v,\mathbf{a}}) on the chart uΩidvΩidu\Omega_{id}^{\circ}\cap v\Omega_{id}^{\circ} for all k[n1]k\in[n-1]. We proceed by induction on kk. Define u,v,𝐚ku,v,𝐚\mathcal{I}^{k}_{u,v,\mathbf{a}}\subseteq\mathcal{I}_{u,v,\mathbf{a}} to be the subset of equations that correspond to boxes in the kk-th column of D𝐚(u)D^{\downarrow}_{\mathbf{a}}(u) and D𝐚(v)D^{\uparrow}_{\mathbf{a}}(v). For k=1k=1, Fπ11(u[k],v[k],ak)F_{\bullet}\in\pi_{1}^{-1}(\mathcal{R}_{u[k],v[k],a_{k}}) if and only if Pi(F)=0P_{i}(F_{\bullet})=0 for all i[u1,v1]ci\notin[u_{1},v_{1}]_{c}. Notice that these are precisely the equations in u,v,𝐚1\mathcal{I}^{1}_{u,v,\mathbf{a}}. Therefore π11(u1,v1,a1)V(u,v,𝐚)\pi_{1}^{-1}(\mathcal{R}_{u_{1},v_{1},a_{1}})\supseteq V(\mathcal{I}_{u,v,\mathbf{a}}).

The induction step is to show that for any FuΩidvΩidF_{\bullet}\in u\Omega_{id}^{\circ}\cap v\Omega_{id}^{\circ}:

Fπk11(u[k1],v[k1],ak1)V(u,v,𝐚k)Fπk1(u[k],v[k],ak).F_{\bullet}\in\pi_{k-1}^{-1}(\mathcal{R}_{u[k-1],v[k-1],a_{k-1}})\cap V(\mathcal{I}^{k}_{u,v,\mathbf{a}})\implies F_{\bullet}\in\pi_{k}^{-1}(\mathcal{R}_{u[k],v[k],a_{k}}).

Since 𝐚\mathbf{a} is flat, by Lemma 4.8,

u[k1],v[k1],ak1=u[k1],v[k1],ak.\mathcal{R}_{u[k-1],v[k-1],a_{k-1}}=\mathcal{R}_{u[k-1],v[k-1],a_{k}}.

We wish to show that if Iaku[k]I\not\geq_{a_{k}}u[k] or Iakv[k]I\not\leq_{a_{k}}v[k] then PIP_{I} vanishes on FF_{\bullet}. We first fix any Iaku[k]I\not\geq_{a_{k}}u[k] and consider the incidence Plücker relation in (3) associated to u[k1]u[k-1] and II:

(26) Pu[k1]PI=iIPIiPu[k1]+i.P_{u[k-1]}P_{I}=\sum_{i\in I}P_{I-i}P_{u[k-1]+i}.

We claim that the right hand side of (26) vanishes. There are two cases:

  1. (1)

    If iakuki\geq_{a_{k}}u_{k}. Since Iaku[k]I\not\geq_{a_{k}}u[k], we have I{i}aku[k1]I\setminus\{i\}\not\geq_{a_{k}}u[k-1]. This implies that PIiP_{I-i} vanishes since Fπk11(u[k1],v[k1],ak1)F_{\bullet}\in\pi_{k-1}^{-1}(\mathcal{R}_{u[k-1],v[k-1],a_{k-1}});

  2. (2)

    If i<akuki<_{a_{k}}u_{k}. Then (i,k)(i,k) is a box in D𝐚(u)D^{\downarrow}_{\mathbf{a}}(u) and Pu[k1]+iP_{u[k-1]+i} vanishes.

Therefore the both sides of (26) vanish. Since FuΩidF_{\bullet}\in u\Omega^{\circ}_{id}, Pu[k1](F)0P_{u[k-1]}(F_{\bullet})\neq 0 and thus PIP_{I} vanishes on FF_{\bullet}. The case where Iakv[k]I\nleq_{a_{k}}v[k] follows from the same reasoning. ∎

Theorem 5.14.

𝒯u,v\mathcal{T}^{\circ}_{u,v} is equidimensional of dimension (u,v)\ell(u,v).

Proof.

Since 𝒯u,v\mathcal{T}^{\circ}_{u,v} is carved out by (n2)(u,v){n\choose 2}-\ell(u,v) equations in Proposition 5.13, by Krull’s principal ideal theorem, every irreducible component of 𝒯u,v\mathcal{T}^{\circ}_{u,v} has dimension at least (u,v)\ell(u,v).

Now we need to prove dim(𝒯u,v)(u,v)\dim(\mathcal{T}^{\circ}_{u,v})\leq\ell(u,v). If not, let Z𝒯u,vZ\subseteq\mathcal{T}^{\circ}_{u,v} be an irreducible component with dim(Z)(u,v)+1\dim(Z)\geq\ell(u,v)+1. Denote Z¯\overline{Z} as the closure of ZZ in FlnFl_{n} and Z:=Z¯Z\partial Z:=\overline{Z}\setminus Z. Since ZZ is closed in the chart uΩidvΩidu\Omega_{id}^{\circ}\cap v\Omega_{id}^{\circ} (by Proposition 5.13), we have Z=Z¯(Fln(uΩidvΩid))=Z¯V(PuPv=0)\partial Z=\overline{Z}\cap(Fl_{n}\setminus(u\Omega_{id}^{\circ}\cap v\Omega_{id}^{\circ}))=\overline{Z}\cap V(P_{u}P_{v}=0). There are two cases.

  1. (1)

    Z\partial Z\neq\emptyset. Since Z\partial Z is the intersection of Z¯\overline{Z} and a union of hyperplanes V(PuPv=0)V(P_{u}P_{v}=0), the dimensions of Z\partial Z and Z¯\overline{Z} differ by at most one: dim(Z)dim(Z¯)1(u,v)\dim(\partial Z)\geq\dim(\overline{Z})-1\geq\ell(u,v). However, since Z𝒯u,v(:=𝒯u,v𝒯u,v)\partial Z\subseteq\partial\mathcal{T}_{u,v}(:=\mathcal{T}_{u,v}\setminus\mathcal{T}^{\circ}_{u,v}) and 𝒯u,v=[x,y][u,v]𝒯x,y\partial\mathcal{T}_{u,v}=\sqcup_{[x,y]\subsetneq[u,v]}\mathcal{T}^{\circ}_{x,y}by Theorem 5.2, dim(Z)(u,v)1\dim(\partial Z)\leq\ell(u,v)-1 by induction on \ell, which is a contradiction.

  2. (2)

    Z=\partial Z=\emptyset. This implies that Z=Z¯Z=\overline{Z} is a projective variety. On the other hand, ZuΩidvΩidZ\subseteq u\Omega_{id}^{\circ}\cap v\Omega_{id}^{\circ} is quasi-affine. This cannot happen unless Z={pt}Z=\{\text{pt}\}, contradicting dim(Z)(u,v)+11\dim(Z)\geq\ell(u,v)+1\geq 1.

Since we reach a contradiction in both cases, we conclude that dim(𝒯u,v)=(u,v)\dim(\mathcal{T}_{u,v}^{\circ})=\ell(u,v). ∎

5.3. Closure of 𝒯u,v\mathcal{T}_{u,v}^{\circ}

Certainly, we expect 𝒯u,v¯=𝒯u,v\overline{\mathcal{T}_{u,v}^{\circ}}=\mathcal{T}_{u,v} (Theorem 5.22). One direction 𝒯u,v¯𝒯u,v\overline{\mathcal{T}_{u,v}^{\circ}}\subset\mathcal{T}_{u,v} is clear. For the other direction, by Theorem 5.2 and induction hypothesis, it suffices to show that 𝒯u,x𝒯u,v¯\mathcal{T}_{u,x}^{\circ}\subset\overline{\mathcal{T}_{u,v}^{\circ}} for x[u,v]x\in[u,v] such that (u,x)=(u,v)1\ell(u,x)=\ell(u,v)-1, and dually, 𝒯y,v𝒯u,v¯\mathcal{T}_{y,v}^{\circ}\subset\overline{\mathcal{T}_{u,v}^{\circ}} for y[u,v]y\in[u,v] such that (y,v)=(u,v)1\ell(y,v)=\ell(u,v)-1. We focus on such an xx.

Throughout this section, let x[u,v]x\in[u,v] such that (u,x)=(u,v)1\ell(u,x)=\ell(u,v)-1.

Lemma 5.15.

𝒯u,vuΩidxΩid=𝒯u,x(𝒯u,vxΩid)\mathcal{T}_{u,v}\cap u\Omega_{id}^{\circ}\cap x\Omega_{id}^{\circ}=\mathcal{T}_{u,x}^{\circ}\sqcup(\mathcal{T}_{u,v}^{\circ}\cap x\Omega_{id}^{\circ}).

Proof.

By Theorem 5.4, the direction \subseteq follows from the fact that there are only two strata in 𝒯u,v=[x,y][u,v]𝒯x,y\mathcal{T}_{u,v}=\bigsqcup_{[x,y]\subseteq[u,v]}\mathcal{T}_{x,y}^{\circ} that intersect both uΩidu\Omega_{id}^{\circ} and xΩidx\Omega_{id}^{\circ}: 𝒯u,x\mathcal{T}_{u,x} and 𝒯u,v\mathcal{T}_{u,v}. The other direction \supseteq follows from 𝒯u,xuΩidxΩid\mathcal{T}_{u,x}^{\circ}\subseteq u\Omega_{id}^{\circ}\cap x\Omega_{id}^{\circ} and 𝒯u,vuΩid\mathcal{T}_{u,v}^{\circ}\subseteq u\Omega_{id}^{\circ} in Proposition 5.4. ∎

The rest of this section is mainly devoted to controling the dimensions of this quasi-affine variety in Lemma 5.15 by finding explicit equations in said charts.

Definition 5.16.

For every pair of permutations (u,v)(u,v) and every choice of flat 𝐚=(a1,,an1)\mathbf{a}=(a_{1},\dots,a_{n-1}) such that u𝐚vu\leq_{\mathbf{a}}v, let x[u,v]x\in[u,v] such that (u,x)=(u,v)1\ell(u,x)=\ell(u,v)-1. Assume x=vtpqx=vt_{pq} for p<qp<q. Define u,v,𝐚,x\mathcal{I}_{u,v,\mathbf{a},x} as the following set of equations:

  1. (1)

    for every box (i,k)D𝐚(u)(i,k)\in D_{\mathbf{a}}^{\downarrow}(u), set Pu[k1]+i=0P_{u[k-1]+i}=0,

  2. (2)

    for every box (i,k)D𝐚(v)(i,k)\in D_{\mathbf{a}}^{\uparrow}(v) that is also in D𝐚(x)D_{\mathbf{a}}^{\uparrow}(x), set Px[k1]+i=0P_{x[k-1]+i}=0,

  3. (3)

    for every box (i,k)D𝐚(v)(i,k)\in D_{\mathbf{a}}^{\uparrow}(v) that is not in D𝐚(x)D_{\mathbf{a}}^{\uparrow}(x) there are two cases:

    1. (a)

      i=xpi=x_{p} and k(p,q)k\in(p,q), set Px[k1]+xq=0P_{x[k-1]+x_{q}}=0,

    2. (b)

      k=qk=q and i(xp,xq)ci\in(x_{p},x_{q})_{c}, set Px[q1]+iPx[p1]+xqPx[p1]+iPx[q1]+xq=0P_{x[q-1]+i}P_{x[p-1]+x_{q}}-P_{x[p-1]+i}P_{x[q-1]+x_{q}}=0.

Example 5.17.

We demonstrate equations of type (2)(2) and (3)(3) in 5.16 using u,v,x,𝐚u,v,x,\mathbf{a} as in 5.12. Here, p=1p=1 and q=5q=5. On xΩidx\Omega_{id}^{\circ}, equations from (2)(2) correspond to having 0’s in the green shaded boxes in Figure 9(C). Equations from (3a)(3a) and (3b)(3b) correspond to having 0 in position (4,4)(4,4) and vanishing of the 2×22\times 2 minor in blue brackets in Figure 9(C).

665544332211112233445566\bullet\bullet\bullet\bullet\bullet\bullet
(A) D𝐚(v)D_{\mathbf{a}}^{\uparrow}(v)
665544332211112233445566\bullet\bullet\bullet\bullet\bullet\bullet
(B) D𝐚(x)D_{\mathbf{a}}^{\uparrow}(x)
66554433221111223344556611111111111100000000000000000aaabab0***0**000bb
(C) xΩidx\Omega_{id}^{\circ}

We need the following technical lemma and its consequence to prove Proposition 5.20.

Lemma 5.18.

For any I([n]a),J([n]b)I\in{[n]\choose a},J\in{[n]\choose b} with 1b<a<n1\leq b<a<n, if there exists k(b,a)k\in(b,a) and K1,K2([n]k)K_{1},K_{2}\in{[n]\choose k} such that FK1,K2,r=K1,K2,rF_{\bullet}\in\mathcal{R}_{K_{1},K_{2},r}=\mathcal{R}_{K_{1},K_{2},r^{\prime}} for some r,r[n]r,r^{\prime}\in[n], then

(*) iI[r,r)cPIiPJ+i(F)=0.\sum_{i\in I\cap[r,r^{\prime})_{c}}P_{I-i}P_{J+i}(F_{\bullet})=0.
Proof.

Fix any K([n]k)K\in{[n]\choose k} such that PK(F)0P_{K}(F_{\bullet})\neq 0. Then Equation (*5.18) is equivalent to

iI[r,r)cPIiPJ+iPK(F)=0.\sum_{i\in I\cap[r,r^{\prime})_{c}}P_{I-i}P_{J+i}P_{K}(F_{\bullet})=0.

By expanding PJ+iPKP_{J+i}P_{K} using (5), this is equivalent to

(27) iI,i[r,r)c,jKPIiPJ+jPKj+i(F)=0.\sum_{i\in I,i\in[r,r^{\prime})_{c},j\in K}P_{I-i}P_{J+j}P_{K-j+i}(F_{\bullet})=0.

Since K1rK2K_{1}\leq_{r}K_{2} and K1rK2K_{1}\leq_{r^{\prime}}K_{2}, PK(F)0P_{K^{\prime}}(F_{\bullet})\neq 0 only if #(K[r,r)c)=#(K1[r,r)c)\#(K^{\prime}\cap[r,r^{\prime})_{c})=\#(K_{1}\cap[r,r^{\prime})_{c}) for any K([n]k)K^{\prime}\in{[n]\choose k}. Thus PKj+i(F)0P_{K-j+i}(F_{\bullet})\neq 0 only if j[r,r)cj\in[r,r^{\prime})_{c}. Equation (27) is then equivalent to

iI[r,r)c,jK[r,r)cPIiPJ+jPKj+i(F)=0.\sum_{i\in I\cap[r,r^{\prime})_{c},j\in K\cap[r,r^{\prime})_{c}}P_{I-i}P_{J+j}P_{K-j+i}(F_{\bullet})=0.

By the same reasoning, this is equivalent to

(28) iI,jK[r,r)cPIiPJ+jPKj+i(F)=0.\sum_{i\in I,j\in K\cap[r,r^{\prime})_{c}}P_{I-i}P_{J+j}P_{K-j+i}(F_{\bullet})=0.

Since iIPIiPKj+i=0\sum_{i\in I}P_{I-i}P_{K-j+i}=0 for jK[r,r)cj\in K\cap[r,r^{\prime})_{c} by (4), Equation 28 and thus (*5.18) holds. ∎

Corollary 5.19.

Given u𝐚vu\leq_{\mathbf{a}}v, for any I([n]q),J([n]p1)I\in{[n]\choose q},J\in{[n]\choose p-1} with 1p<q<n1\leq p<q<n, if flag FF_{\bullet} satisfies Fπk1(u[k],v[k],ak)F_{\bullet}\in\pi_{k}^{-1}(\mathcal{R}_{u[k],v[k],a_{k}}) for every pk<qp\leq k<q, then

(*) iI[ap,aq)cPIiPJ+i(F)=0.\sum_{\begin{subarray}{c}i\in I\cap[a_{p},a_{q})_{c}\end{subarray}}P_{I-i}P_{J+i}(F_{\bullet})=0.
Proof.

By Lemma 5.10, there is a flat 𝐚=(a1,,an1)\mathbf{a^{\prime}}=(a_{1}^{\prime},\dots,a_{n-1}^{\prime}) such that u𝐚vu\leq_{\mathbf{a^{\prime}}}v. By Lemma 5.18,

iI[ap,ap)cPIiPJ+i(F)=0 and iI[aq,aq)cPIiPJ+i(F)=0.\sum_{\begin{subarray}{c}i\in I\cap[a_{p},a_{p}^{\prime})_{c}\end{subarray}}P_{I-i}P_{J+i}(F_{\bullet})=0\text{ and }\sum_{\begin{subarray}{c}i\in I\cap[a_{q}^{\prime},a_{q})_{c}\end{subarray}}P_{I-i}P_{J+i}(F_{\bullet})=0.

Moreover, for every pk<qp\leq k<q, by flatness of 𝐚\mathbf{a}^{\prime}:

iI[ak,ak+1)cPIiPJ+i(F)=0;\sum_{\begin{subarray}{c}i\in I\cap[a_{k}^{\prime},a_{k+1}^{\prime})_{c}\end{subarray}}P_{I-i}P_{J+i}(F_{\bullet})=0;

By summing the above equations and subtracting iIPIiPJ+i(F)=0\sum_{i\in I}P_{I-i}P_{J+i}(F_{\bullet})=0 appropriate times, we obtain (*5.19). ∎

Proposition 5.20.

For any flat 𝐚\mathbf{a}, 𝒯u,vuΩidxΩid=uΩidxΩidV(u,v,𝐚,x)\mathcal{T}_{u,v}\cap u\Omega_{id}^{\circ}\cap x\Omega_{id}^{\circ}=u\Omega_{id}^{\circ}\cap x\Omega_{id}^{\circ}\cap V(\mathcal{I}_{u,v,\mathbf{a},x}).

Proof.

We first show \subseteq. This is equivalent to proving that FF_{\bullet} satisfies the equations in u,v,𝐚,x\mathcal{I}_{u,v,\mathbf{a},x} if F𝒯u,vuΩidxΩidF_{\bullet}\in\mathcal{T}_{u,v}\cap u\Omega_{id}^{\circ}\cap x\Omega_{id}^{\circ}. There are four types of equations in u,v,𝐚,x\mathcal{I}_{u,v,\mathbf{a},x}:

Type (1). Here (i,k)D𝐚(u)(i,k)\in D^{\downarrow}_{\mathbf{a}}(u). Since i<akuki<_{a_{k}}u_{k}, we have u[k1]{i}<aku[k]u[k-1]\cup\{i\}<_{a_{k}}u[k], by Proposition 4.5

Fπk1(u[k],v[k],ak)Pu[k1]+i(F)=0;F_{\bullet}\in\pi_{k}^{-1}(\mathcal{R}_{u[k],v[k],a_{k}})\implies P_{u[k-1]+i}(F_{\bullet})=0;

Type (2). Here (i,k)D𝐚(v)D𝐚(x)(i,k)\in D^{\uparrow}_{\mathbf{a}}(v)\cap D^{\uparrow}_{\mathbf{a}}(x). Since i>akvki>_{a_{k}}v_{k} and xp>akvpx_{p}>_{a_{k}}v_{p} if k(p,q]k\in(p,q], we have x[k1]{i}>akv[k]x[k-1]\cup\{i\}>_{a_{k}}v[k], and

Fπk1(u[k],v[k],ak)Px[k1]+i(F)=0.F_{\bullet}\in\pi_{k}^{-1}(\mathcal{R}_{u[k],v[k],a_{k}})\implies P_{x[k-1]+i}(F_{\bullet})=0.

Type (3a). Here (xp,k)D𝐚(v)D𝐚(x)(x_{p},k)\in D^{\uparrow}_{\mathbf{a}}(v)\setminus D^{\uparrow}_{\mathbf{a}}(x) where i=xpi=x_{p} and k(p,q)k\in(p,q). Since xp>akvkx_{p}>_{a_{k}}v_{k}, we have x[k1]{xq}>akv[k]x[k-1]\cup\{x_{q}\}>_{a_{k}}v[k], and

Fπk1(u[k],v[k],ak)Px[k1]+xq(F)=0.F_{\bullet}\in\pi_{k}^{-1}(\mathcal{R}_{u[k],v[k],a_{k}})\implies P_{x[k-1]+x_{q}}(F_{\bullet})=0.

Type (3b). Here (i,q)D𝐚(v)D𝐚(x)(i,q)\in D^{\uparrow}_{\mathbf{a}}(v)\setminus D^{\uparrow}_{\mathbf{a}}(x) where k=qk=q and i(xp,xq)ci\in(x_{p},x_{q})_{c}. By (5), we have

(29) Px[p1]+iPv[q]=k[p,q]Px[p1]+vkPv[q]vk+i.P_{x[p-1]+i}P_{v[q]}=-\sum_{k\in[p,q]}P_{x[p-1]+v_{k}}P_{v[q]-v_{k}+i}.

The left hand side of (29) is Px[p1]+iPv[q]=(1)#x[q1]>xqPx[p1]+iPx[q1]+xqP_{x[p-1]+i}P_{v[q]}=(-1)^{\#x[q-1]>x_{q}}P_{x[p-1]+i}P_{x[q-1]+x_{q}}.

For k=pk=p, the corresponding term in right hand side of (29) is

Px[p1]+vpPv[q]vp+i=(1)#x[q1]>xqPx[p1]+xqPx[q1]+i.-P_{x[p-1]+v_{p}}P_{v[q]-v_{p}+i}=(-1)^{\#x[q-1]>x_{q}}P_{x[p-1]+x_{q}}P_{x[q-1]+i}.

For k=qk=q, since i(xp,xq)ci\in(x_{p},x_{q})_{c} and aq(xp,xq)ca_{q}\notin(x_{p},x_{q})_{c}, we have i>aqvqi>_{a_{q}}v_{q} and thus v[q]vq+i>aqv[q]v[q]-v_{q}+i>_{a_{q}}v[q]. Therefore Fπq1(u[q],v[q],aq)Pv[q]vq+i(F)=0F_{\bullet}\in\pi_{q}^{-1}(\mathcal{R}_{u[q],v[q],a_{q}})\implies P_{v[q]-v_{q}+i}(F_{\bullet})=0 for all i(xp,xq)ci\in(x_{p},x_{q})_{c}.

For p<k<qp<k<q, if vk>apvpv_{k}>_{a_{p}}v_{p}, then x[p1]+vk>apv[p]x[p-1]+v_{k}>_{a_{p}}v[p] and Px[p1]+vk(F)=0P_{x[p-1]+v_{k}}(F_{\bullet})=0. Similarly, if i>aqvki>_{a_{q}}v_{k}, then v[q]vk+i>aqv[q]v[q]-v_{k}+i>_{a_{q}}v[q] and Pv[q]vk+i(F)=0P_{v[q]-v_{k}+i}(F_{\bullet})=0. Therefore, for each ii,

(30) k(p,q)Px[p1]+vkPv[q]vk+i=k(p,q),iaqvkapvpPx[p1]+vkPv[q]vk+i.\sum_{k\in(p,q)}P_{x[p-1]+v_{k}}P_{v[q]-v_{k}+i}=\sum_{k\in(p,q),i\leq_{a_{q}}v_{k}\leq_{a_{p}}v_{p}}P_{x[p-1]+v_{k}}P_{v[q]-v_{k}+i}.

Since vk[xp,xq]cv_{k}\notin[x_{p},x_{q}]_{c} for k(p,q)k\in(p,q),

iaqvkvk(xq,aq)c and vkapvpvk[ap,xq)c.i\leq_{a_{q}}v_{k}\iff v_{k}\in(x_{q},a_{q})_{c}\text{ and }v_{k}\leq_{a_{p}}v_{p}\iff v_{k}\in[a_{p},x_{q})_{c}.

The right hand side of (30) equals

k:vk[ap,aq)cPx[p1]+vkPv[q]vk+i.\sum_{k:v_{k}\in[a_{p},a_{q})_{c}}P_{x[p-1]+v_{k}}P_{v[q]-v_{k}+i}.

Since i[ap,aq)ci\notin[a_{p},a_{q})_{c}, this is 0 by 5.19. Therefore

F𝒯u,vuΩidxΩidPx[q1]+iPx[p1]+xqPx[p1]+iPx[q1]+xq(F)=0.F_{\bullet}\in\mathcal{T}_{u,v}\cap u\Omega_{id}^{\circ}\cap x\Omega_{id}^{\circ}\implies P_{x[q-1]+i}P_{x[p-1]+x_{q}}-P_{x[p-1]+i}P_{x[q-1]+x_{q}}(F_{\bullet})=0.

Since we have exhausted all equations in u,v,𝐚,x\mathcal{I}_{u,v,\mathbf{a},x}, we conclude the proof of \subseteq direction.

We now prove the \supseteq direction. Namely for FuΩidxΩidF_{\bullet}\in u\Omega_{id}^{\circ}\cap x\Omega_{id}^{\circ} and k[n1]k\in[n-1], if FV(u,v,𝐚,x)F_{\bullet}\in V(\mathcal{I}_{u,v,\mathbf{a},x}) then Fπk1(u[k],v[k],ak)F_{\bullet}\in\pi_{k}^{-1}(\mathcal{R}_{u[k],v[k],a_{k}}). By the same reasoning in 5.13, if FV(u,v,𝐚,x)uΩidF_{\bullet}\in V(\mathcal{I}_{u,v,\mathbf{a},x})\cap u\Omega_{id}^{\circ} then PK(F)=0P_{K}(F_{\bullet})=0 for all K([n]k)K\in{[n]\choose k} such that Kaku[k]K\ngeq_{a_{k}}u[k]. We can then focus on Equations in u,v,𝐚,x\mathcal{I}_{u,v,\mathbf{a},x} of type (2)(2) and (3)(3) as in 5.16. It is enough to show that if FxΩidV(u,v,𝐚,x)F_{\bullet}\in x\Omega_{id}^{\circ}\cap V(\mathcal{I}_{u,v,\mathbf{a},x}) then PK(F)=0P_{K}(F_{\bullet})=0 for all Kakv[k]K\nleq_{a_{k}}v[k].

We proceed by induction on kk. For k=1k=1, Piu,v,𝐚,xi>a1v1P_{i}\in\mathcal{I}_{u,v,\mathbf{a},x}\iff i>_{a_{1}}v_{1}. Therefore FV(u,v,𝐚,x)Fπ11(u1,v1,a1)F_{\bullet}\in V(\mathcal{I}_{u,v,\mathbf{a},x})\implies F_{\bullet}\in\pi_{1}^{-1}(\mathcal{R}_{u_{1},v_{1},a_{1}}). Suppose the statement holds for k1k-1, we fix K([n]k)K\in{[n]\choose k} such that Kv[k]K\nleq v[k] and consider the following two cases:

Case I (kqk\neq q): By (3), we have

(31) Px[k1]PK=iKPx[k1]+iPKi.P_{x[k-1]}P_{K}=\sum_{i\in K}P_{x[k-1]+i}P_{K-i}.

If i>akvki>_{a_{k}}v_{k}, then Px[k1]+i(F)=0P_{x[k-1]+i}(F_{\bullet})=0 by (2)(2) of 5.16. If iakvki\leq_{a_{k}}v_{k}, then Kiakv[k1]K-i\nleq_{a_{k}}v[k-1]. Since 𝐚\mathbf{a} is flat, Kiak1v[k1]K-i\nleq_{a_{k-1}}v[k-1] and thus PKi(F)=0P_{K-i}(F_{\bullet})=0 by inductive hypothesis. The right hand side of (31) is then 0. Since FxΩidF_{\bullet}\in x\Omega_{id}^{\circ}, Px[k1](F)0P_{x[k-1]}(F_{\bullet})\neq 0. Thus PK(F)=0P_{K}(F_{\bullet})=0.

Case II (k=qk=q): Again by (3),

(32) Px[q1]PK=iKPx[q1]+iPKi.P_{x[q-1]}P_{K}=\sum_{i\in K}P_{x[q-1]+i}P_{K-i}.

If i>aqxqi>_{a_{q}}x_{q}, then either ix[q1]i\in x[q-1] or (i,q)D𝐚(v)D𝐚(x)(i,q)\in D_{\mathbf{a}}^{\uparrow}(v)\cap D_{\mathbf{a}}^{\uparrow}(x). In both cases, we have Px[q1]+i(F)=0P_{x[q-1]+i}(F_{\bullet})=0. If iaqxp=vqi\leq_{a_{q}}x_{p}=v_{q}, then Kiv[q1]K-i\nleq v[q-1] and PKi(F)=0P_{K-i}(F_{\bullet})=0 by inductive hypothesis. The right hand side of (32) then equals

(33) iK(xp,xq]cPx[q1]+iPKi.\sum_{i\in K\cap(x_{p},x_{q}]_{c}}P_{x[q-1]+i}P_{K-i}.

In order to show the above equation vanishes, we wish to multiply by Px[p1]+xqP_{x[p-1]+x_{q}} and apply the equation in (3b)(3b) of 5.16. However, it is not guaranteed that Px[p1]+xq(F)0P_{x[p-1]+x_{q}}(F_{\bullet})\neq 0 for FxΩiduΩidV(u,v,𝐚,x)F_{\bullet}\in x\Omega_{id}^{\circ}\cap u\Omega_{id}^{\circ}\cap V(\mathcal{I}_{u,v,\mathbf{a},x}). So we further divide into two cases.

Case IIA (Px[p1]+xq(F)0P_{x[p-1]+x_{q}}(F_{\bullet})\neq 0): Using (3b)(3b) from 5.16, we have

(34) iKPx[p1]+xqPx[q1]+iPKi=iKPx[q1]+xqPx[p1]+iPKi.\sum_{i\in K}P_{x[p-1]+x_{q}}P_{x[q-1]+i}P_{K-i}=\sum_{i\in K}P_{x[q-1]+x_{q}}P_{x[p-1]+i}P_{K-i}.

By (4), Equation 34 equals 0. Since PKi(F)=0P_{K-i}(F_{\bullet})=0 for i[aq,xp]ci\in[a_{q},x_{p}]_{c} and Px[p1]+i(F)=0P_{x[p-1]+i}(F_{\bullet})=0 for i(xq,ap)ci\in(x_{q},a_{p})_{c}, we have

iK([ap,aq)c(xp,xq]c)Px[q1]+xqPx[p1]+iPKi=0.\sum_{i\in K\cap([a_{p},a_{q})_{c}\cup(x_{p},x_{q}]_{c})}P_{x[q-1]+x_{q}}P_{x[p-1]+i}P_{K-i}=0.

Since [ap,aq)c[a_{p},a_{q})_{c} and (xp,xq]c(x_{p},x_{q}]_{c} are disjoint, by 5.19, we have

iK(xp,xq]cPx[q1]+xqPx[q1]+iPKi=0.\sum_{i\in K\cap(x_{p},x_{q}]_{c}}P_{x[q-1]+x_{q}}P_{x[q-1]+i}P_{K-i}=0.

Since Px[p1]+xq(F),Px[q1](F)0P_{x[p-1]+x_{q}}(F_{\bullet}),P_{x[q-1]}(F_{\bullet})\neq 0, by (32) and (33), we must have PK(F)=0P_{K}(F_{\bullet})=0.

Case IIB (Px[p1]+xq(F)=0P_{x[p-1]+x_{q}}(F_{\bullet})=0): For i(xp,xq)ci\in(x_{p},x_{q})_{c}, since we have Px[q1]+iPx[p1]+xq=Px[p1]+iPx[q1]+xqP_{x[q-1]+i}P_{x[p-1]+x_{q}}=P_{x[p-1]+i}P_{x[q-1]+x_{q}}, we have Px[p1]+i(F)=0P_{x[p-1]+i}(F_{\bullet})=0. Notice by 5.7 and 5.16, these are the equations in u,x,𝐚u,v,𝐚,x\mathcal{I}_{u,x,\mathbf{a}}\setminus\mathcal{I}_{u,v,\mathbf{a},x}. Therefore

FV(u,x,𝐚)xΩiduΩid=𝒯u,x.F_{\bullet}\in V(\mathcal{I}_{u,x,\mathbf{a}})\cap x\Omega_{id}^{\circ}\cap u\Omega_{id}^{\circ}=\mathcal{T}_{u,x}^{\circ}.

By Lemma 5.15, we have V(u,v,𝐚,x)xΩiduΩid𝒯u,vxΩiduΩidV(\mathcal{I}_{u,v,\mathbf{a},x})\cap x\Omega_{id}^{\circ}\cap u\Omega_{id}^{\circ}\subseteq\mathcal{T}_{u,v}\cap x\Omega_{id}^{\circ}\cap u\Omega_{id}^{\circ} as desired. ∎

Proposition 5.21.

𝒯u,vuΩidxΩid\mathcal{T}_{u,v}\cap u\Omega_{id}^{\circ}\cap x\Omega_{id}^{\circ} is equidimensional of dimension (u,v)\ell(u,v).

Proof.

Since 𝒯u,vuΩidxΩid\mathcal{T}_{u,v}\cap u\Omega_{id}^{\circ}\cap x\Omega_{id}^{\circ} is the zero locus of (n2)(u,v){n\choose 2}-\ell(u,v) equations by Proposition 5.20, every irreducible component has dimension (u,v)\geq\ell(u,v). On the other hand, Theorem 5.14 and Lemma 5.15 implies that every irreducible component has dimension at most (u,v)\ell(u,v). ∎

Theorem 5.22.

𝒯u,v=𝒯u,v¯\mathcal{T}_{u,v}=\overline{\mathcal{T}_{u,v}^{\circ}}.

Proof.

We prove by induction on (u,v)\ell(u,v). For (u,v)=0\ell(u,v)=0, u=vu=v and 𝒯u,u=𝒯u,u={eu}\mathcal{T}_{u,u}=\mathcal{T}_{u,u}^{\circ}=\{e_{u}\}. Now assume (u,v)>0\ell(u,v)>0. By Theorem 5.2, it is enough to show that 𝒯u,x𝒯u,v¯\mathcal{T}_{u,x}^{\circ}\subseteq\overline{\mathcal{T}_{u,v}^{\circ}} for all x[u,v]x\in[u,v] such that (u,x)=(u,v)1\ell(u,x)=\ell(u,v)-1. Pick any such xx, by Lemma 5.15, we have

𝒯u,vuΩidxΩid¯=𝒯u,x¯𝒯u,vxΩid¯\overline{\mathcal{T}_{u,v}\cap u\Omega_{id}^{\circ}\cap x\Omega_{id}^{\circ}}=\overline{\mathcal{T}_{u,x}^{\circ}}\cup\overline{\mathcal{T}^{\circ}_{u,v}\cap x\Omega_{id}^{\circ}}

as quasi-affine varieties in uΩidxΩidu\Omega_{id}^{\circ}\cap x\Omega_{id}^{\circ}. Let Z𝒯u,vuΩidxΩid¯Z\subseteq\overline{\mathcal{T}_{u,v}\cap u\Omega_{id}^{\circ}\cap x\Omega_{id}^{\circ}} be any irreducible component. Since dim(Z)=(u,v)>dim(𝒯u,x¯)\dim(Z)=\ell(u,v)>\dim(\overline{\mathcal{T}_{u,x}^{\circ}}) by Proposition 5.21, Z𝒯u,x¯Z\nsubseteq\overline{\mathcal{T}_{u,x}^{\circ}}. Since ZZ is irreducible, Z𝒯u,vxΩid¯Z\subseteq\overline{\mathcal{T}^{\circ}_{u,v}\cap x\Omega_{id}^{\circ}}, and since this holds for all irreducible components ZZ, 𝒯u,vuΩidxΩid¯𝒯u,vxΩid¯\overline{\mathcal{T}_{u,v}\cap u\Omega_{id}^{\circ}\cap x\Omega_{id}^{\circ}}\subseteq\overline{\mathcal{T}^{\circ}_{u,v}\cap x\Omega_{id}^{\circ}}. Thus, 𝒯u,x𝒯u,x¯𝒯u,vxΩid¯𝒯u,v¯\mathcal{T}^{\circ}_{u,x}\subseteq\overline{\mathcal{T}^{\circ}_{u,x}}\subseteq\overline{\mathcal{T}^{\circ}_{u,v}\cap x\Omega_{id}^{\circ}}\subseteq\overline{\mathcal{T}_{u,v}^{\circ}} as desired. ∎

Acknowledgements

We are grateful to Alex Postnikov for introducing us to quantum Bruhat graphs. We thank Anders Buch, Allen Knutson, Thomas Lam, Leonardo Mihalcea, Lauren Williams, Weihong Xu and Alexander Yong for many helpful conversations. SG is partially supported by an NSF Graduate Research Fellowship under grant No. DGE-1746047.

References

  • [1] Anders Björner. Orderings of Coxeter groups. In Combinatorics and algebra (Boulder, Colo., 1983), volume 34 of Contemp. Math., pages 175–195. Amer. Math. Soc., Providence, RI, 1984.
  • [2] Francesco Brenti, Sergey Fomin, and Alexander Postnikov. Mixed Bruhat operators and Yang-Baxter equations for Weyl groups. Internat. Math. Res. Notices, (8):419–441, 1999.
  • [3] Anders S. Buch, Pierre-Emmanuel Chaput, Leonardo C. Mihalcea, and Nicolas Perrin. Finiteness of cominuscule quantum KK-theory. Ann. Sci. Éc. Norm. Supér. (4), 46(3):477–494, 2013.
  • [4] Anders S. Buch, Pierre-Emmanuel Chaput, Leonardo C. Mihalcea, and Nicolas Perrin. Projected Gromov-Witten varieties in cominuscule spaces. Proc. Amer. Math. Soc., 146(9):3647–3660, 2018.
  • [5] Anders S. Buch, Sjuvon Chung, Changzheng Li, and Leonardo C. Mihalcea. Euler characteristics in the quantum KK-theory of flag varieties. Selecta Math. (N.S.), 26(2):Paper No. 29, 11, 2020.
  • [6] Anders S. Buch and Leonardo C. Mihalcea. Curve neighborhoods of Schubert varieties. J. Differential Geom., 99(2):255–283, 2015.
  • [7] W. Fulton and C. Woodward. On the quantum product of Schubert classes. J. Algebraic Geom., 13(4):641–661, 2004.
  • [8] William Fulton. Young tableaux, volume 35 of London Mathematical Society Student Texts. Cambridge University Press, Cambridge, 1997. With applications to representation theory and geometry.
  • [9] Allen Knutson, Thomas Lam, and David E. Speyer. Positroid varieties: juggling and geometry. Compos. Math., 149(10):1710–1752, 2013.
  • [10] Changzheng Li and Leonardo C. Mihalcea. K-theoretic Gromov-Witten invariants of lines in homogeneous spaces. Int. Math. Res. Not. IMRN, (17):4625–4664, 2014.
  • [11] Alexander Postnikov. Affine approach to quantum Schubert calculus. Duke Math. J., 128(3):473–509, 2005.
  • [12] Alexander Postnikov. Quantum Bruhat graph and Schubert polynomials. Proc. Amer. Math. Soc., 133(3):699–709, 2005.
  • [13] Alexander Woo and Alexander Yong. A Gröbner basis for Kazhdan-Lusztig ideals. Amer. J. Math., 134(4):1089–1137, 2012.