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

The atomic Leibniz rule

Ben Elias Department of Mathematics, Fenton Hall, University of Oregon, Eugene, OR, 97403-1222, USA [email protected] Hankyung Ko Department of Mathematics, Uppsala University, Box. 480, SE-75106, Uppsala, Sweden [email protected] Nicolas Libedinsky Departamento de Matemáticas, Facultad de Ciencias, Universidad de Chile [email protected]  and  Leonardo Patimo Dipartimento di Matematica, Università di Pisa, Italy [email protected]
Abstract.

The Demazure operator associated to a simple reflection satisfies the twisted Leibniz rule. In this paper we introduce a generalization of the twisted Leibniz rule for the Demazure operator associated to any atomic double coset. We prove that this atomic Leibniz rule is equivalent to a polynomial forcing property for singular Soergel bimodules.

1. Introduction

We introduce a generalization of the twisted Leibniz rule which applies to Demazure operators associated with certain double cosets. We call it the atomic Leibniz rule, and it should play an important role in an eventual description of the singular Hecke category by generators and relations. This result is the algebraic heart which we extract from singular Soergel bimodules and transplant to their diagrammatic calculus.

1.1.

In this paper, we work with general Coxeter systems and quite general actions of these groups on polynomial rings. For ease of exposition, in this section, we assume that R=[x1,,xn]R=\mathbb{Q}[x_{1},\ldots,x_{n}], with the standard action of the symmetric group W=𝐒nW=\mathbf{S}_{n}. Let sis_{i} be the simple reflection swapping ii and i+1i+1, and let S={s1,,sn1}S=\{s_{1},\ldots,s_{n-1}\} be the set of simple reflections. There is a Demazure operator si:RR\partial_{s_{i}}\colon R\to R defined by the formula

(1) si(f)=fsifxixi+1.\partial_{s_{i}}(f)=\frac{f-s_{i}f}{x_{i}-x_{i+1}}.

This operator famously satisfies a twisted Leibniz rule

(2) si(fg)=si(f)si(g)+si(f)g.\partial_{s_{i}}(fg)=s_{i}(f)\partial_{s_{i}}(g)+\partial_{s_{i}}(f)g.

Because the operators si\partial_{s_{i}} satisfy the braid relations, one can unambiguously define an operator w\partial_{w} associated with any w𝐒nw\in\mathbf{S}_{n} by composing the operators si\partial_{s_{i}} along a reduced expression. When computing w(fg)\partial_{w}(fg), one could apply the twisted Leibniz rule repeatedly to obtain a complicated generalization of (2) for w\partial_{w}. We discuss this in §2.2.

What we provide in this paper is the natural generalization of the twisted Leibniz rule to the setting of double cosets. In double coset combinatorics, the analogues of simple reflections are called atomic cosets. We prove a version of (2) for Demazure operators attached to atomic cosets.

1.2.

We recall some definitions so as to precisely state our first theorem. Let (W,S)(W,S) be any Coxeter system. For any subset ISI\subset S, let WIW_{I} be the parabolic subgroup of WW generated by II. We assume that II is finitary, i.e. that WIW_{I} is a finite group. Let RIRR^{I}\subset R be the subring of polynomials invariant under WIW_{I}. Let wIw_{I} denote the longest element of WIW_{I}. For two subsets I,JSI,J\subset S, a double coset pWI\W/WJp\in W_{I}\backslash W/W_{J} will be called an (I,J)(I,J)-coset. Any (I,J)(I,J)-coset pp has a unique minimal element p¯W\underline{p}\in W and a unique maximal element p¯W\overline{p}\in W with respect to the Bruhat order.

In [EKLP23a, §3.4], we introduced a Demazure operator p:RJRI\partial_{p}:R^{J}\to R^{I} for any (I,J)(I,J)-coset pp. In fact, p\partial_{p} is equal to w\partial_{w} for w=p¯wJWw=\overline{p}w_{J}\in W. Normally one views w\partial_{w} as a map RRR\to R, but when the source of this map is restricted from RR to RJR^{J}, then the image is contained in RIR^{I}. When I=J=I=J=\emptyset, so that p={w}p=\{w\} for some wWw\in W, then p=w\partial_{p}=\partial_{w}.

In [EKLP24b] we introduced the notion of an atomic (I,J)(I,J)-coset. Briefly stated, 𝚊{\mathtt{a}} is atomic if

  • there exists s,tSs,t\in S (possibly equal) with sIs\notin I and tJt\notin J such that I{s}=J{t}=:MI\cup\{s\}=J\cup\{t\}=:M.

  • WMW_{M} is a finite group, 𝚊¯=wM\overline{{\mathtt{a}}}=w_{M} and wMs=twMw_{M}s=tw_{M}.

In particular, the subsets II and JJ are conjugate under both 𝚊¯\overline{{\mathtt{a}}} and 𝚊¯\underline{{\mathtt{a}}}. If fRJf\in R^{J} then 𝚊¯(f)RI\underline{{\mathtt{a}}}(f)\in R^{I}. We also have 𝚊¯=𝚊¯wJ\underline{{\mathtt{a}}}=\overline{{\mathtt{a}}}w_{J}, so that 𝚊\partial_{\mathtt{a}} is the restriction of 𝚊¯\partial_{\underline{{\mathtt{a}}}}.

When I=J=I=J=\emptyset, the atomic (I,J)(I,J)-cosets have the form {s}\{s\} for sSs\in S. Atomic cosets in 𝐒n\mathbf{S}_{n} can be described using cabled crossings. For example, the permutation with one-line notation (345671289)(345671289) crosses the first two numbers past the next five while fixing the remaining numbers. It is a prototypical example of 𝚊¯\underline{{\mathtt{a}}} for an atomic coset 𝚊{\mathtt{a}}; in this example WI𝐒2×𝐒5×W_{I}\cong\mathbf{S}_{2}\times\mathbf{S}_{5}\times\cdots and WJ𝐒5×𝐒2×W_{J}\cong\mathbf{S}_{5}\times\mathbf{S}_{2}\times\cdots and WM𝐒7×W_{M}\cong\mathbf{S}_{7}\times\cdots.

1.3.

Our story began with the defining representation of 𝐒n\mathbf{S}_{n} over \mathbb{Q}. Then we considered the polynomial ring RR, its invariant subrings, and the structure derived therefrom. Alternatively, one can start with a realization of a Coxeter system (W,S)(W,S), which is effectively a “reflection representation” of WW over a commutative base ring 𝕜\Bbbk, equipped with a choice of roots and coroots. From this representation we construct its polynomial ring RR, Demazure operators, etcetera. See §3.1 for details.

For each realization111Soergel’s construction depends only on a reflection representation of WW, and not a choice of roots and coroots. Elias and Williamson gave a presentation of Soergel’s category in [EW16] which depends on a choice of realization. The Demazure operators also depend on the choice of roots and coroots., Soergel defined a monoidal category of RR-bimodules called Soergel bimodules, and proved that for certain realizations222Specifically, the base ring 𝕜\Bbbk should be an infinite field of characteristic not equal to 22, and the representation should be (faithful and) reflection-faithful. (that we call SW-realizations) one obtains a categorification of the Hecke algebra. Soergel bimodules and related categorifications of the Hecke algebra are objects of critical importance in geometric representation theory. The following is the first main result of this paper (Theorem 5.10).

Theorem A (The atomic Leibniz rule).

Let (W,S)(W,S) be any Coxeter system equipped with an SW-realization, or the symmetric group 𝐒n\mathbf{S}_{n} with R=[x1,,xn].R=\mathbb{Z}[x_{1},\ldots,x_{n}]. Let 𝚊{\mathtt{a}} be an atomic (I,J)(I,J)-coset and \leq the Bruhat order for double cosets333The Bruhat order on (I,J)(I,J)-cosets can be defined by q𝚊q\leq{\mathtt{a}} if and only if q¯𝚊¯\underline{q}\leq\underline{{\mathtt{a}}}. See [EKLP23b, Thm. 2.16] for equivalent definitions. Note that qq need not be atomic, so q\partial_{q} need not equal (the restriction of) q¯\partial_{\underline{q}}.. For every q<𝚊q<{\mathtt{a}} there is a unique RIJR^{I\cup J}-linear operator on polynomials denoted Tq𝚊T^{\mathtt{a}}_{q}, satisfying the equation

(3) 𝚊(fg)=𝚊¯(f)𝚊(g)+q<𝚊q(Tq𝚊(f)g)\partial_{\mathtt{a}}(f\cdot g)=\underline{{\mathtt{a}}}(f)\partial_{\mathtt{a}}(g)+\sum_{q<{\mathtt{a}}}\partial_{q}(T^{\mathtt{a}}_{q}(f)\cdot g)

for all f,gRJ.f,g\in R^{J}. Polynomials in the image of Tq𝚊T^{\mathtt{a}}_{q} are appropriately invariant, see Definition 3.8 for details. When 𝚊{\mathtt{a}} is fixed, we will often write Tq:=Tq𝚊T_{q}:=T_{q}^{\mathtt{a}}.

The motivation for this result will take some setup, so please bear with us.

Example 1.1.

We let ee denote the identity element. Let ss be a simple reflection and let 𝚊{\mathtt{a}} be the (,)(\emptyset,\emptyset)-coset {s}\{s\}. There is only one coset less than 𝚊{\mathtt{a}}, namely q={e}q=\{e\}. We set Tq𝚊(f)=s(f)T^{\mathtt{a}}_{q}(f)=\partial_{s}(f). Note that 𝚊=s\partial_{\mathtt{a}}=\partial_{s} and q=id\partial_{q}=\operatorname{id}. Then (3) becomes

(4) s(fg)=s(f)s(g)+s(f)g,\partial_{s}(fg)=s(f)\partial_{s}(g)+\partial_{s}(f)g,

which recovers the twisted Leibniz rule.

For more examples see §2. There we also give an example of a non-atomic coset pp for which no equality of the form (3) can hold (with pp replacing 𝚊{\mathtt{a}}).

Remark 1.2.

Though we can prove the existence and uniqueness of such a formula, we do not have an explicit description of the operators TqT_{q}. We consider this an interesting open problem. A more accessible problem is to compute TqT_{q} on a (carefully chosen) set of generators of RJR^{J}, which we accomplish in type AA in Theorem 6.3. This is useful computationally because it gives enough information to apply the atomic Leibniz rule for any pair of elements in RJR^{J} (see the discussion of multiplicativity in §1.6).

Remark 1.3.

In §3.3 we prove that the atomic Leibniz rule for one realization implies the atomic Leibniz rule for related realizations (e.g. obtained via base change, enlargement, or quotient). The atomic Leibniz rule also depends only on the restriction of the realization to the finite parabolic subgroup WMW_{M} associated to 𝚊{\mathtt{a}}. Thus Theorem A implies the atomic Leibniz rule in broad generality for realizations of Coxeter systems in both finite and affine type AA, and in finite characteristic as well, see Example 3.19.

1.4.

Our second main result is a connection between the atomic Leibniz rule and the theory of singular Soergel bimodules [Wil11]. Like Soergel bimodules, singular Soergel bimodules are ubiquitous in geometric representation theory, appearing e.g. in the geometric Satake equivalence and other situations where partial flag varieties play a role. Specifically, we connect the atomic Leibniz rule to a property called polynomial forcing, whose motivation we postpone a little longer.

Let us first be more precise. Singular Soergel bimodules are graded bimodules over graded rings, but we ignore all gradings in this paper. To an atomic coset 𝚊{\mathtt{a}} as above, one can associate the (RI,RJ)(R^{I},R^{J})-bimodule B𝚊:=RIRMRJB_{\mathtt{a}}:=R^{I}\otimes_{R^{M}}R^{J}. This is an indecomposable bimodule; there are also indecomposable (RI,RJ)(R^{I},R^{J})-bimodules BqB_{q} associated to any (I,J)(I,J)-coset qq, which we will not try to describe here. Then B𝚊B_{\mathtt{a}} has a submodule called the submodule of lower terms, spanned by the images of all bimodule endomorphisms of B𝚊B_{\mathtt{a}} which factor through BqB_{q} for q<𝚊q<{\mathtt{a}}.

Atomic polynomial forcing is the statement that 1f1\otimes f and 𝚊¯(f)1\underline{{\mathtt{a}}}(f)\otimes 1 are equal modulo lower terms in B𝚊B_{\mathtt{a}}, for any fRJf\in R^{J}. Williamson has proven for SW-realizations [Wil11, Theorem 6.4] that singular Soergel bimodules have “standard filtrations,” which implies atomic polynomial forcing for SW-realizations. Polynomial forcing (Definition 5.7) is a generalization of atomic polynomial forcing for a bimodule BpB_{p} associated with an arbitrary double coset pp. Using [EKLP23b, EKLP24b] we prove Theorem 5.13, which says that polynomial forcing for any double coset is a consequence of atomic polynomial forcing.

The restrictions imposed on SW-realizations are significant as they rule out some examples of great importance to modular representation theory, e.g. affine Weyl groups in finite characteristic. An almost SW-realization (Definition 4.21) is a realization over a domain 𝕜\Bbbk such that one obtains an SW-realization after base change to the fraction field of 𝕜\Bbbk. An example is [x1,,xn]\mathbb{Z}[x_{1},\ldots,x_{n}] with its action of 𝐒n\mathbf{S}_{n}. The second main result of this paper is the following equivalence (5.5):

Theorem B.

Let (W,S)(W,S) be a Coxeter system equipped with an almost SW-realization. Then the atomic Leibniz rule is equivalent to atomic polynomial forcing.

We deduce the SW-realization case of Theorem A from Theorem B and Williamson’s theory of standard filtrations.

1.5.

Let us now explain the motivation for our results.

The diagrammatic presentation for the category of Soergel bimodules, developed by Elias, Khovanov, Libedinsky, and Williamson [Lib10],[EK10], [Eli16], [EW16] has proven to be an important tool for both abstract and computational reasons. This diagrammatically constructed category is typically called the Hecke category, and is equivalent to the category of Soergel bimodules for SW-realizations. For non-SW-realizations the category of Soergel bimodules need not behave well, but the diagrammatic category does behave well, and continues to provide a categorification of the Hecke algebra. The ability to compute within the Hecke category has led to advances such as [EW14], [EW21],[Wil17],[RW18],[LW18],[Ach+19].

The authors are in the midst of a concerted effort to define the singular Hecke category, a diagrammatic presentation of singular Soergel bimodules. The framework for such a presentation was developed in [ESW17], but this framework lacks some of the relations needed. In [EKLP24a] we described what should be the basis for morphisms in the diagrammatic category, called the double leaves basis, and proved that it descends to a basis for morphisms between singular Soergel bimodules for SW-realizations.

Once the remaining relations are understood, proving that the diagrammatic category is correctly presented reduces to proving that double leaves span. An important part of that proof will be diagrammatic polynomial forcing, which is like polynomial forcing except that the submodule of lower terms is replaced by the span of double leaves associated to smaller elements in the Bruhat order. We abbreviate diagrammatic polynomial forcing to DL\operatorname{DL}-forcing. It is DL\operatorname{DL}-forcing which is our true goal in this paper.

For an atomic coset 𝚊{\mathtt{a}} one can identify End(B𝚊)\operatorname{End}(B_{{\mathtt{a}}}) with B𝚊B_{{\mathtt{a}}} as bimodules. In Section 4.2.3, we explicitly compute the submodule DL<𝚊B𝚊\operatorname{DL}_{<{\mathtt{a}}}\subset B_{\mathtt{a}} corresponding to the 𝕜\Bbbk-linear subspace of End(B𝚊)\mathrm{End}(B_{\mathtt{a}}) spanned by double leaves factoring through q<𝚊q<{\mathtt{a}}. The following theorem (Corollary 4.33 and Theorem 5.5) is proven by direct computation and is the reason we discovered the atomic Leibniz rule in the first place.

Theorem C.

Let (W,S)(W,S) be a Coxeter system equipped with an almost SW-realization. Then the atomic Leibniz rule is equivalent to atomic DL\operatorname{DL}-forcing.

In other words, for a given fRJf\in R^{J}, we prove that 1f𝚊¯(f)1DL<𝚊1\otimes f-\underline{{\mathtt{a}}}(f)\otimes 1\in\operatorname{DL}_{<{\mathtt{a}}} if and only if the atomic Leibniz rule holds for ff and any gRJg\in R^{J}.

For an SW-realization, the results of [EKLP24a] can be used to prove that DL<𝚊\operatorname{DL}_{<{\mathtt{a}}} agrees precisely with the submodule of lower terms (see Corollary 4.19). In §4.4 we use some novel localization tricks to extend this statement to almost-SW-realizations. Thus DL\operatorname{DL}-forcing agrees with polynomial forcing for almost-SW-realizations, thus Theorem C implies Theorem B.

1.6.

One is still motivated to prove DL\operatorname{DL}-forcing (or equivalently, the atomic Leibniz rule) for almost-SW-realizations and more general realizations. We now discuss how this process is simplified by Theorem B.

Let us say the atomic Leibniz rule holds for fRJf\in R^{J} if (3) holds for that specific ff and all gRJg\in R^{J}. It is obvious that if the rule holds for f1f_{1} and f2f_{2} then it holds for f1+f2f_{1}+f_{2}. It is not obvious that if it holds for f1f_{1} and f2f_{2} then it holds for f1f2f_{1}\cdot f_{2}, thus the formula (3) is not obviously multiplicative in ff. However, atomic polynomial forcing is obviously multiplicative in ff. Our equivalence in Theorem B is proven on an element-by-element basis, from which we conclude that the atomic Leibniz rule is actually multiplicative.

As a consequence, the atomic Leibniz rule can be proven without relying on Williamson’s results, by checking it on each generator ff of RJR^{J}. We perform this computation in type AA in §6, for carefully chosen generators ff, via a direct and elementary proof. The operators TqT_{q} are simplified dramatically when applied to these generators.

On the one hand, this proves Theorem A for the symmetric group over [x1,,xn]\mathbb{Z}[x_{1},\ldots,x_{n}], and consequently for other realizations (see Remark 1.3). Via Theorem C, this implies DL\operatorname{DL}-forcing in the same generality. On the other hand, it also gives a computationally effective way to apply the atomic Leibniz rule (or polynomial forcing), even if one does not know the operators TqT_{q} in general: decompose ff as a linear combination of products of generators, and apply the atomic Leibniz rule for generators one term at a time.

Acknowledgments. BE was partially supported by NSF grants DMS-2201387 and DMS-2039316. HK was partially supported by the Swedish Research Council. NL was partially supported by FONDECYT-ANID grant 1230247.

2. Examples and remarks

In this chapter we give some examples of atomic Leibniz rules, in the relatively small Coxeter group W=𝐒4W=\mathbf{S}_{4}. The reader should not be disheartened if these examples are still quite difficult and technical to verify by hand! Indeed, one purpose of this chapter is to showcase the complexity involved in even the simplest examples of the atomic Leibniz rule.

The reader will not miss out on anything important if they skip directly to Section 3.

2.1. Examples

Our examples take place in W=𝐒4W=\mathbf{S}_{4}. We let s=(12)s=(12), t=(23)t=(23), and u=(34)u=(34) denote the simple reflections. We give the examples without justification first, and then discuss the verification afterwards. We require that Tq(f)RKT_{q}(f)\in R^{K} for some KSK\subset S depending on qq (precisely: K=q¯1Iq¯JK=\underline{q}^{-1}I\underline{q}\cap J).

Example 2.1.

There are three (su,su)(su,su)-cosets in 𝐒4\mathbf{S}_{4}: the maximal one pp which is atomic, the submaximal coset qq with q¯=t\underline{q}=t, and the minimal coset rr containing the identity. We claim that for f,gRsuf,g\in R^{su} we have

(5) tsut(fg)=tsut(f)tsut(g)+sut(Tq(f)g)+Tr(f)g,\partial_{tsut}(fg)=tsut(f)\cdot\partial_{tsut}(g)+\partial_{sut}(T_{q}(f)\cdot g)+T_{r}(f)\cdot g,

or in other words

(6) p(fg)=p¯(f)p(g)+q(Tq(f)g)+r(Tr(f)g),\partial_{p}(fg)=\underline{p}(f)\cdot\partial_{p}(g)+\partial_{q}(T_{q}(f)\cdot g)+\partial_{r}(T_{r}(f)\cdot g),

where

(7) Tq(f)=sut(f),Tr(f)=tsut(f)sut(Tq(f)).T_{q}(f)=su\partial_{t}(f),\qquad T_{r}(f)=\partial_{tsut}(f)-\partial_{sut}(T_{q}(f)).

Note that (obviously) Tq(f)R=RT_{q}(f)\in R=R^{\emptyset} while (less obviously) Tr(f)RsuT_{r}(f)\in R^{su}. These are the invariance requirements.

Remark 2.2.

Iterating the ordinary twisted Leibniz rule, it is not hard to deduce the existence of a formula of the form

tsut(fg)=tsut(f)tsut(g)+x<tsutx(Tx(f)g)\partial_{tsut}(fg)=tsut(f)\cdot\partial_{tsut}(g)+\sum_{x<tsut}\partial_{x}(T_{x}(f)g)

for some operators Tx:RRT_{x}\colon R\to R. What is not obvious is that, when f,gRsuf,g\in R^{su}, this formula will simplify so that only the terms where x=sutx=sut and x=1x=1 survive.

Example 2.3.

There are two (tu,st)(tu,st)-cosets in 𝐒4\mathbf{S}_{4}: the maximal coset pp which is atomic, and the minimal coset qq containing the identity element. We claim that for f,gRstf,g\in R^{st} we have

(8) stu(fg)=stu(f)stu(g)+tu(Tq(f)g),\partial_{stu}(fg)=stu(f)\partial_{stu}(g)+\partial_{tu}(T_{q}(f)\cdot g),

or in other words

(9) p(fg)=p¯(f)p(g)+q(Tq(f)g),\partial_{p}(fg)=\underline{p}(f)\partial_{p}(g)+\partial_{q}(T_{q}(f)\cdot g),

where

(10) Tq(f)=stu(f).T_{q}(f)=st\partial_{u}(f).

Note also that Tq(f)RtT_{q}(f)\in R^{t}.

It helps to look at an example of a non-atomic coset, to see that Leibniz rules are not guaranteed.

Example 2.4.

We return to the notation of Example 2.1. Note that rr is the only coset less than qq, and r\partial_{r} is the identity map. Let us argue that there is no naive analogue of (3) for q\partial_{q}. If there were, it would have to have the form

(11) sut(fg)=sut(f)sut(g)+T(f)g\partial_{sut}(f\cdot g)=sut(f)\partial_{sut}(g)+T(f)\cdot g

for some operator TT. By evaluating at g=1g=1 we have T(f)=sut(f)T(f)=\partial_{sut}(f).

Note that f,gRsuf,g\in R^{su}. Iterating the twisted Leibniz rule we have

(12) sut(fg)=\displaystyle\quad\partial_{sut}(f\cdot g)=
sut(f)sut(g)+s(ut(f))ut(g)+u(st(f))st(g)+su(t(f))t(g)+sut(f)g.\displaystyle sut(f)\partial_{sut}(g)+\partial_{s}(ut(f))\partial_{ut}(g)+\partial_{u}(st(f))\partial_{st}(g)+\partial_{su}(t(f))\partial_{t}(g)+\partial_{sut}(f)g.

Only two of these five terms are in (11). If gg is linear and t(g)=1\partial_{t}(g)=1 then the sum of the three missing terms is nonzero for some ff. Thus (11) is false.

However, now suppose that ft(Rsu)f\in t(R^{su}) instead of RsuR^{su}. Then many terms in (12) vanish, yielding

(13) sut(fg)=sut(f)sut(g)+sut(f)g.\partial_{sut}(f\cdot g)=sut(f)\partial_{sut}(g)+\partial_{sut}(f)g.

This is compatible with (11), with T=sutT=\partial_{sut}.

In view of this example, there is hope of finding some generalization of the atomic Leibniz rule to some non-atomic (I,J)(I,J) cosets qq, letting fq¯1(RI)f\in\underline{q}^{-1}(R^{I}). Be warned that the coset qq considered in this example has various special properties, see Remark 3.12.

The examples above can easily be verified by computer (all the equations are RstuR^{stu}-linear so one need only check the result for f,gf,g in a basis). They can also all be verified using (1) and (2), but this is a very tricky exercise. Doing this exercise may be very instructive for the reader, and emphasizes the difficult and subtlety in these formulae, so we encourage it, and provide some helping hands.

We begin with a few helpful general properties of Demazure operators, which hold in general when mst=3m_{st}=3 and msu=2m_{su}=2.

Lemma 2.5.

We have

(14) s(s(f))=s(f),ss(f)=s(f),s(s(f))=0,\partial_{s}(s(f))=-\partial_{s}(f),\quad s\partial_{s}(f)=\partial_{s}(f),\qquad\partial_{s}(\partial_{s}(f))=0,
(15) sts(f)=t(stf),st(sf)=ts(tf),su(f)=u(sf),st\partial_{s}(f)=\partial_{t}(stf),\quad s\partial_{t}(sf)=t\partial_{s}(tf),\quad s\partial_{u}(f)=\partial_{u}(sf),
(16) αss(f)=fsf,\alpha_{s}\partial_{s}(f)=f-sf,
(17) st(sf)+ts(f)=tst(f).\partial_{st}(sf)+\partial_{ts}(f)=t\partial_{st}(f).

Note that αsi=xixi+1\alpha_{s_{i}}=x_{i}-x_{i+1} above.

Proof.

The reader can verify these relations directly from (1). ∎

Most of these formulae are well-known. Meanwhile, we have not seen (17) before; we only use it in Example 2.9 below. With these relations in hand, one need not refer to the original definition (1) again, and need only use the twisted Leibniz rule.

Here are some example computations using (14) and (15).

(18) s(ts(f))=s(sts(f))=st(stf)=st(tstf).\partial_{s}(t\partial_{s}(f))=-\partial_{s}(st\partial_{s}(f))=-\partial_{s}\partial_{t}(stf)=\partial_{s}\partial_{t}(tstf).
(19) s(ts(f))=s(tss(f))=tsts(f).\partial_{s}(t\partial_{s}(f))=\partial_{s}(ts\partial_{s}(f))=ts\partial_{t}\partial_{s}(f).

Let us consider example 2.3. It helps to observe that

(20) st(u(f))Rst,stu(f)Rstu,\partial_{st}(u(f))\in R^{st},\qquad\partial_{stu}(f)\in R^{stu},

under the assumption that fRstf\in R^{st}. One way to see the equality (8) is to expand both sides using the twisted Leibniz rule. Since gRstg\in R^{st}, any terms containing st(g)\partial_{st}(g) or s(g)\partial_{s}(g) or t(g)\partial_{t}(g) or su(g)\partial_{su}(g) are zero. Now compare the coefficients of gg, u(g)\partial_{u}(g) and tu(g)\partial_{tu}(g) respectively. On the left the coefficient of gg is stu(f)\partial_{stu}(f), while on the right the coefficient of gg is

tustu(f)=tsutuu(f)=tstutu(f)=stsutu(f)=stustu(f)=stu(f).\partial_{tu}st\partial_{u}(f)=\partial_{t}s\partial_{u}tu\partial_{u}(f)=\partial_{t}stu\partial_{tu}(f)=st\partial_{s}u\partial_{tu}(f)=stu\partial_{stu}(f)=\partial_{stu}(f).

We have applied (15) repeatedly and used that stu(f)Rstu\partial_{stu}(f)\in R^{stu}. Thus the coefficients of gg are the same on both sides of (8). We leave the other coefficients to the reader.

Example 2.1 is the hardest. One should first prove the following statements under the critical assumption that f,gRsuf,g\in R^{su}.

(21a) tsut(f)Rstu,ts(ut(f))Rst,\partial_{tsut}(f)\in R^{stu},\qquad\partial_{ts}(ut(f))\in R^{st},
(21b) su(tsu(t(f)))=tsut(sut(f)),\partial_{su}(tsu(\partial_{t}(f)))=tsut(\partial_{sut}(f)),
(21c) t(sut(f))=tsu(sut(f)),t(\partial_{sut}(f))=tsu(\partial_{sut}(f)),
(21d) tsu[sut(f)tsut(f)]=tsu(αttust(f))=(αs+αt+αu)tsut(f),tsu\left[\partial_{sut}(f)-t\partial_{sut}(f)\right]=tsu(\alpha_{t}\partial_{tust}(f))=(\alpha_{s}+\alpha_{t}+\alpha_{u})\partial_{tsut}(f),
(21e) tsu(tf)=tsu(αtt(f))=(αs+αt+αu)tsut(f).\partial_{tsu}(tf)=\partial_{tsu}(-\alpha_{t}\partial_{t}(f))=-(\alpha_{s}+\alpha_{t}+\alpha_{u})\partial_{tsut}(f).

We leave these verifications, and the deduction of (5) therefrom, to the ambitious reader.

2.2. Leibniz rules for permutations

A natural question is raised: for which double cosets pp does one expect an equality of the form (3) to hold? Given the discussion in Example 2.4, one might ask instead: for which double cosets qq does (3) hold under the alternate assumption that fq¯1(RI)f\in\underline{q}^{-1}(R^{I})? Note that fq¯1(RI)f\in\underline{q}^{-1}(R^{I}) is equivalent to fRJf\in R^{J} for atomic cosets, and also for more general cosets called core cosets. We believe these are interesting questions, even though these more general Leibniz rules currently lack a clear connection to the theory of singular Soergel bimodules.

A special case would be when I=J=I=J=\emptyset, so that double cosets are in bijection with elements wWw\in W; all such cosets are core. We start with the following well-known lemma.

Lemma 2.6.

Let wWw\in W and f,gRf,g\in R. Let w=s1snw=s_{1}\ldots s_{n} be a reduced expression and for 𝐞={0,1}n\mathbf{e}=\{0,1\}^{n} we define the element w𝐞=s1e1snenw^{\mathbf{e}}=s_{1}^{e_{1}}\ldots s_{n}^{e_{n}}. For 𝐞={0,1}n\mathbf{e}=\{0,1\}^{n} let

θi𝐞={siif ei=1iif ei=0 and Θ𝐞(f)=θ1𝐞θ2𝐞θn𝐞(f).\theta_{i}^{\mathbf{e}}=\begin{cases}s_{i}&\text{if }e_{i}=1\\ \partial_{i}&\text{if }e_{i}=0\end{cases}\text{ and }\Theta^{\mathbf{e}}(f)=\theta_{1}^{\mathbf{e}}\circ\theta_{2}^{\mathbf{e}}\circ\ldots\circ\theta_{n}^{\mathbf{e}}(f).

Then we have

w(fg)=xwTx(f)x(g)\partial_{w}(fg)=\sum_{x\leq w}T^{\prime}_{x}(f)\partial_{x}(g)

where

Tx(f)=𝐞w𝐞=xΘ𝐞(f).T^{\prime}_{x}(f)=\sum_{\mathbf{e}\mid w^{\mathbf{e}}=x}\Theta^{\mathbf{e}}(f).

As a special case, Tw(f)=w(f)T_{w}^{\prime}(f)=w(f).

Proof.

This is just an iteration of the twisted Leibniz rule. ∎

To summarize, we obtain a generalized Leibniz rule of the form

(22) w(fg)=w(f)w(g)+x<wTx(f)x(g),\partial_{w}(fg)=w(f)\partial_{w}(g)+\sum_{x<w}T^{\prime}_{x}(f)\partial_{x}(g),

for operators Tx:RRT^{\prime}_{x}\colon R\to R (depending on ww). The formula for TxT^{\prime}_{x} one derives in this way is seemingly dependent on the choice of reduced expression for ww, though the operator only depends444Abstractly, the nilHecke algebra is the subalgebra of End(R)\operatorname{End}(R) generated by RR (i.e. multiplication by polynomials) and by Demazure operators. It is well known that the operators {w}\{\partial_{w}\} form a basis of the nilHecke algebra as a free left RR-module. Letting mfm_{f} denote multiplication by ff, (22) can be viewed as an equality wmf=mw(f)w+mTx(f)x\partial_{w}\circ m_{f}=m_{w(f)}\circ\partial_{w}+\sum m_{T^{\prime}_{x}(f)}\circ\partial_{x} in the nilHecke algebra, which rewrites wmf\partial_{w}\circ m_{f} in this basis. Consequently, the coefficients Tx(f)T^{\prime}_{x}(f) in this linear combination depend only on ww and ff. on ww. In practice, confirming the independence of reduced expression can be quite subtle. We are unaware of a formula for TxT^{\prime}_{x} which is obviously independent of the choice of reduced expression.

Meanwhile, one can also deduce an equality of the form

(23) w(fg)=w(f)w(g)+x<wx(Tx(f)g),\partial_{w}(fg)=w(f)\partial_{w}(g)+\sum_{x<w}\partial_{x}(T_{x}(f)\cdot g),

for some operators Tx:RRT_{x}\colon R\to R. We are unaware of any previous study of the operators TxT_{x} and the formula (23).

Remark 2.7.

Later in the paper, we also discuss an atomic Leibniz rule similar to (22) rather than (23).

Here are some examples.

Example 2.8.

Let s=s1s=s_{1} and t=s2t=s_{2}. When w=tsw=ts, applying (2) twice gives

(24) ts(fg)=ts(f)ts(g)+t(sf)s(g)+ts(f)t(g)+ts(f)g.\partial_{ts}(fg)=ts(f)\partial_{ts}(g)+\partial_{t}(sf)\partial_{s}(g)+t\partial_{s}(f)\partial_{t}(g)+\partial_{ts}(f)g.

An equivalent formula is

(25) ts(fg)=ts(f)ts(g)+t(s(f)g)+s(st(sf)g)+st(sf)g.\partial_{ts}(fg)=ts(f)\partial_{ts}(g)+\partial_{t}(\partial_{s}(f)\cdot g)+\partial_{s}(s\partial_{t}(sf)\cdot g)+\partial_{st}(sf)\cdot g.

By applying (2) to the second and third terms on the RHS of equation (25), and with a little help from (14), one obtains (24).

Example 2.9.

With notation as above, when w=stsw=sts, applying (2) thrice gives

sts(fg)\displaystyle\partial_{sts}(fg) =\displaystyle= sts(f)sts(g)+s(tsf)ts(g)+t(stf)st(g)\displaystyle sts(f)\partial_{sts}(g)+\partial_{s}(tsf)\partial_{ts}(g)+\partial_{t}(stf)\partial_{st}(g)
+s(ts(f))t(g)+t(st(f))s(g)+sts(f)g.\displaystyle+\partial_{s}(t\partial_{s}(f))\partial_{t}(g)+\partial_{t}(s\partial_{t}(f))\partial_{s}(g)+\partial_{sts}(f)g.

More honestly, applying (2) thrice gives the above except that the coefficient of s(g)\partial_{s}(g) is st(sf)+sts(f)\partial_{st}(sf)+s\partial_{ts}(f). By applying ss to (17), one obtains

(27) st(sf)+sts(f)=stst(f)=t(stt(f))=t(st(f)).\partial_{st}(sf)+s\partial_{ts}(f)=st\partial_{st}(f)=\partial_{t}(st\partial_{t}(f))=\partial_{t}(s\partial_{t}(f)).

This is how one deduces (2.9).

An equivalent formula is

sts(fg)\displaystyle\partial_{sts}(fg) =\displaystyle= sts(f)sts(g)+ts(t(f)g)+st(s(f)g)\displaystyle sts(f)\partial_{sts}(g)+\partial_{ts}(\partial_{t}(f)\cdot g)+\partial_{st}(\partial_{s}(f)\cdot g)
t(st(f)g)s(ts(f)g)+sts(f)g.\displaystyle-\partial_{t}(\partial_{st}(f)\cdot g)-\partial_{s}(\partial_{ts}(f)\cdot g)+\partial_{sts}(f)\cdot g.

To verify that (2.9) and (2.9) agree, apply (25) to the term ts(t(f)g)\partial_{ts}(\partial_{t}(f)\cdot g), and apply (25) with ss and tt swapped to st(s(f)g)\partial_{st}(\partial_{s}(f)\cdot g). After some additional massaging using (14) and (15), one will recover (2.9).

3. Atomic Leibniz rules

3.1. Realizations

We fix a Coxeter system (W,S)(W,S). We let ee denote the identity element of WW. Recall the definition of a realization from [EW16, §3.1].

Definition 3.1.

A realization of (W,S)(W,S) over 𝕜\Bbbk is the data (𝕜,V,Δ,Δ)(\Bbbk,V,\Delta,\Delta^{\vee}) of a commutative ring 𝕜\Bbbk, a free finite-rank 𝕜\Bbbk-module VV, a set Δ={αs}sS\Delta=\{\alpha_{s}\}_{s\in S} of simple roots inside VV, and a set Δ={αs}sS\Delta^{\vee}=\{\alpha_{s}^{\vee}\}_{s\in S} of simple coroots inside Hom𝕜(V,𝕜)\operatorname{Hom}_{\Bbbk}(V,\Bbbk), satisfying the following properties. One has αs(αs)=2\alpha_{s}^{\vee}(\alpha_{s})=2 for all sSs\in S. The formula

s(v):=vαs(v)αss(v):=v-\alpha_{s}^{\vee}(v)\cdot\alpha_{s}

defines an action of WW on VV. Also, the technical condition [EW16, (3.3)] holds, which is redundant for most base rings 𝕜\Bbbk.

For short, we often refer to the data of a realization simply by reference to the WW-representation VV.

Example 3.2.

The permutation realization of 𝐒n\mathbf{S}_{n} over \mathbb{Z} has V=nV=\mathbb{Z}^{n} with basis {xi}i=1n\{x_{i}\}_{i=1}^{n} and with dual bases {xi}\{x_{i}^{*}\}. For 1in11\leq i\leq n-1 one sets αi=xixi+1\alpha_{i}=x_{i}-x_{i+1} and αi=xixi+1\alpha_{i}^{\vee}=x_{i}^{*}-x_{i+1}^{*}.

Example 3.3.

For a Weyl group WW, the root realization of (W,S)(W,S) over \mathbb{Z} is the free \mathbb{Z}-module with basis Δ\Delta. One defines Δ\Delta^{\vee} so that the pairings αs(αt)\alpha_{s}^{\vee}(\alpha_{t}) agree with the usual Cartan matrix of WW.

Example 3.4.

Let (𝕜,V,Δ,Δ)(\Bbbk,V,\Delta,\Delta^{\vee}) be a realization of (W,S)(W,S), and ISI\subset S. Then (𝕜,V,ΔI,ΔI)(\Bbbk,V,\Delta_{I},\Delta_{I}^{\vee}) is also a realization of (WI,I)(W_{I},I), the restriction of the realization to a parabolic subgroup. Here ΔI={αs}sI\Delta_{I}=\{\alpha_{s}\}_{s\in I}, and similarly for ΔI\Delta_{I}^{\vee}.

Given a realization, let RR be the polynomial ring whose linear terms are VV. We can associate Demazure operators s:RR\partial_{s}\colon R\to R for sSs\in S, which agree with αs\alpha_{s}^{\vee} on VRV\subset R, and are extended by the twisted Leibniz rule. For details, see [EW16, §3.1]. For each finitary subset ISI\subset S, we also consider the subring RIR^{I} of WIW_{I}-invariants in RR. The ring RIR^{I} is graded, and all RIR^{I}-modules will be graded, but we will not keep track of grading shifts in this paper as they will play no significant role. The background on this material in [EKLP23a] should be sufficient.

Definition 3.5.

A (balanced) Frobenius realization is a realization satisfying the following properties, see [EKLP23a, §3.1] for definitions.

  • It is balanced.

  • It satisfies generalized Demazure surjectivity.

  • It is faithful when restricted to each finite parabolic subgroup WIW_{I}.

We assume tacitly throughout this paper that we work with a Frobenius realization. The main implication of these assumptions is that, when ISI\subset S is finitary, the Demazure operator wI:RRI\partial_{w_{I}}\colon R\to R^{I} is well-defined and equips the ring extension RIRR^{I}\subset R with the structure of a Frobenius extension.

The left and right redundancy sets of an (I,J)(I,J)-coset pp are defined and denoted as

(29) LR(p)=Ip¯Jp¯1,RR(p)=p¯1Ip¯J.\operatorname{LR}(p)=I\cap\underline{p}J\underline{p}^{-1},\qquad\operatorname{RR}(p)=\underline{p}^{-1}I\underline{p}\cap J.

An (I,J)(I,J)-coset pp is a core coset if I=LR(p)I=\operatorname{LR}(p) and J=RR(p)J=\operatorname{RR}(p).

For any (I,J)(I,J)-coset qq, in [EKLP23a] we define a Demazure operator

q:RJRI.\partial_{q}:R^{J}\to R^{I}.

By definition, q\partial_{q} is the restriction of the ordinary Demazure operator

q¯wJ1:RR\partial_{\overline{q}w_{J}^{-1}}\colon R\to R

to the subring RJR^{J}. After restriction, the image is contained in RIR^{I}, see [EKLP23a, Lemma 3.9]. Note that q¯wJ1=q¯\overline{q}w_{J}^{-1}=\underline{q} if and only if qq is a core coset.

Remark 3.6.

Some results from [EKLP23a] require further that the realization is faithful, rather than just faithful upon restriction to each finite parabolic subgroup. In particular, the set {p}\{\partial_{p}\} as pp ranges over (I,J)(I,J)-cosets need not be linearly independent when the realization is not faithful.

A multistep (I,J)(I,J)-expression is a sequence of finitary subsets

I=[[I=I0K1I1KmIm=J]].I_{\bullet}=[[I=I_{0}\subset K_{1}\supset I_{1}\subset\ldots K_{m}\supset I_{m}=J]].

The definition of a reduced multistep expression, and of the (I,J)(I,J)-coset that it expresses, can be found in [EK21, Definition 1.4]. When II_{\bullet} is a (reduced) expression which expresses pp, we write pIp\leftrightharpoons I_{\bullet}.

As in [EK21], for x,yWx,y\in W we write x.yx.y for a reduced composition, where (xy)=(x)+(y)\ell(xy)=\ell(x)+\ell(y). We also use this notation for the reduced composition of reduced expressions, or the reduced composition of double cosets, see [EK21] for more details. Demazure operators compose well over reduced compositions: one has p.q=pq\partial_{p.q}=\partial_{p}\circ\partial_{q}, as proven in [EKLP23a, Corollary 3.19].

3.2. Precise statement of atomic Leibniz rules

Remark 3.7.

Note that wJw_{J} is an involution, so wJ=wJ1w_{J}=w_{J}^{-1}. We write q¯wJ1\overline{q}w_{J}^{-1} above to emphasize that q¯=(q¯wJ1).wJ\overline{q}=(\overline{q}w_{J}^{-1}).w_{J}.

Definition 3.8.

Suppose MM is finitary, sMs\in M, and t=wMswMt=w_{M}sw_{M}. Let I=s^:=MsI=\widehat{s}:=M\setminus s, and J=t^:=MtJ=\widehat{t}:=M\setminus t. Let 𝚊{\mathtt{a}} be the (atomic) (I,J)(I,J)-coset containing wMw_{M}. We say a (rightward) atomic Leibniz rule holds for 𝚊{\mathtt{a}} if there exist RMR^{M}-linear operators Tq𝚊T^{\mathtt{a}}_{q} from RJR^{J} to RRR(q)R^{\operatorname{RR}(q)} for each (I,J)(I,J)-coset q<𝚊q<{\mathtt{a}}, such that for any f,gRJf,g\in R^{J} we have

(30) 𝚊(fg)=𝚊¯(f)𝚊(g)+q<𝚊q¯wJ1(Tq𝚊(f)g).\partial_{\mathtt{a}}(f\cdot g)=\underline{{\mathtt{a}}}(f)\partial_{\mathtt{a}}(g)+\sum_{q<{\mathtt{a}}}\partial_{\overline{q}w_{J}^{-1}}(T^{\mathtt{a}}_{q}(f)\cdot g).

We encourage the reader to confirm that Tq𝚊(f)RRR(q)T^{\mathtt{a}}_{q}(f)\in R^{\operatorname{RR}(q)} in the examples of §2. We continue to write TqT_{q} instead of Tq𝚊T^{\mathtt{a}}_{q} when 𝚊{\mathtt{a}} is understood.

Remark 3.9.

We say “an atomic Leibniz rule” rather than “the atomic Leibniz rule” because we are defining a prototype for a kind of formula. If one specifies operators TqT_{q} such that the formula holds, then one has produced “the” atomic Leibniz rule for that coset 𝚊{\mathtt{a}} (indeed, we prove in 5.5 that such operators are unique for certain realizations).

The difference between (30) and (3) is subtle: we have written q¯wJ1\partial_{\overline{q}w_{J}^{-1}} instead of q\partial_{q}. The difference between q¯wJ1\partial_{\overline{q}w_{J}^{-1}} and q\partial_{q} is only a matter of the domain and codomain of the functions: the former is a function RRR\to R, while the latter is its restriction to a function RJRIR^{J}\to R^{I}. Meanwhile, Tq(f)T_{q}(f) lives in RRR(q)R^{\operatorname{RR}(q)}. The inclusion RJRRR(q)R^{J}\subset R^{\operatorname{RR}(q)} is proper unless qq is core. It is therefore inappropriate to apply q\partial_{q} to Tq(f)gT_{q}(f)\cdot g. Having altered notation so that the domain of the operator is appropriate, we still need to worry about the codomain, which we address in the following lemma.

Lemma 3.10.

With notation as in Definition 3.8, we have q¯wJ1(Tq(f)g)RI\partial_{\overline{q}w_{J}^{-1}}(T_{q}(f)\cdot g)\in R^{I}.

Proof.

Recall from [EK21, Proposition 4.28] that any (I,J)(I,J)-coset qq has a reduced expression of the form

(31) q[[ILR(q)]].qcore.[[RR(q)J]].q\leftrightharpoons[[I\supset\operatorname{LR}(q)]].q^{\operatorname{core}}.[[\operatorname{RR}(q)\subset J]].

Let zz be the (I,RR(q))(I,\operatorname{RR}(q))-coset with reduced expression

(32) z[[ILR(q)]].qcore.z\leftrightharpoons[[I\supset\operatorname{LR}(q)]].q^{\operatorname{core}}.

Since (31) is reduced, by [EK21, Proposition 4.3], we have q¯=z¯.(wRR(q)1wJ),\overline{q}=\overline{z}.(w_{\operatorname{RR}(q)}^{-1}w_{J}), so that

(33) q¯wJ1=z¯wRR(q)1.\overline{q}w_{J}^{-1}=\overline{z}w_{\operatorname{RR}(q)}^{-1}.

Consequently, the same operator q¯wJ1:RR\partial_{\overline{q}w_{J}^{-1}}\colon R\to R restricts to both q:RJRI\partial_{q}\colon R^{J}\to R^{I} and z:RRR(q)RI\partial_{z}\colon R^{\operatorname{RR}(q)}\to R^{I}. In particular, this operator sends RRR(q)R^{\operatorname{RR}(q)} to RIR^{I}. ∎

Further elaboration will be helpful in subsequent chapters. By [EKLP23a, Lemma 3.17], the reduced expression (31) implies that the map q\partial_{q} is a composition of three Demazure operators. Recall that [[ILR(q)]]\partial_{[[I\supset\operatorname{LR}(q)]]} is the Frobenius trace map RLR(q)RIR^{\operatorname{LR}(q)}\to R^{I} often denoted as ILR(q)\partial^{\operatorname{LR}(q)}_{I}. Recall also that [[RR(q)J]]\partial_{[[\operatorname{RR}(q)\subset J]]} is the inclusion map RJRRR(q)R^{J}\subset R^{\operatorname{RR}(q)}. We denote this inclusion map ιJRR(q)\iota^{\operatorname{RR}(q)}_{J} below. So we have

(34) q=ILR(q)qcoreιJRR(q).\partial_{q}=\partial^{\operatorname{LR}(q)}_{I}\circ\partial_{q^{\operatorname{core}}}\circ\iota^{\operatorname{RR}(q)}_{J}.

By (32) we have

(35) z=ILR(q)qcore,\partial_{z}=\partial^{\operatorname{LR}(q)}_{I}\circ\partial_{q^{\operatorname{core}}},

which agrees with the restriction of q¯wJ1\partial_{\overline{q}w_{J}^{-1}} to RRR(q)R^{\operatorname{RR}(q)}. Thus one has the following reformulation of (30):

(36) 𝚊(fg)=𝚊¯(f)𝚊(g)+q<𝚊ILR(q)qcore(Tq(f)ιJRR(q)(g)).\partial_{\mathtt{a}}(f\cdot g)=\underline{{\mathtt{a}}}(f)\partial_{\mathtt{a}}(g)+\sum_{q<{\mathtt{a}}}\partial^{\operatorname{LR}(q)}_{I}\partial_{q^{\operatorname{core}}}(T_{q}(f)\cdot\iota^{\operatorname{RR}(q)}_{J}(g)).

Now the polynomial Tq(f)T_{q}(f) appears more appropriately in the “middle” of this factorization of q\partial_{q}. This discussion of the “placement” of the polynomial Tq(f)T_{q}(f) will play a role in our diagrammatic proof of polynomial forcing.

We are now prepared to discuss another version of the atomic Leibniz rule, using the factorization (34). It should not be obvious that these two atomic Leibniz rules are related, though the equivalence with polynomial forcing will shed light on this issue.

Definition 3.11.

Use the notation from Definition 3.8 and from (34). We say that a leftward atomic Leibniz rule holds for 𝚊{\mathtt{a}} if there exist RMR^{M}-linear operators TqT^{\prime}_{q} from RJR^{J} to RLR(q)R^{\operatorname{LR}(q)} for each (I,J)(I,J)-coset q<𝚊q<{\mathtt{a}}, such that for any f,gRJf,g\in R^{J} we have

(37) 𝚊(fg)=𝚊¯(f)𝚊(g)+q<𝚊ILR(q)(Tq(f)qcore(ιJRR(q)(g))).\partial_{\mathtt{a}}(f\cdot g)=\underline{{\mathtt{a}}}(f)\partial_{\mathtt{a}}(g)+\sum_{q<{\mathtt{a}}}\partial^{\operatorname{LR}(q)}_{I}(T^{\prime}_{q}(f)\cdot\partial_{q^{\operatorname{core}}}(\iota_{J}^{\operatorname{RR}(q)}(g))).
Remark 3.12.

The fact that Tq(f)T_{q}(f) lives in RRR(q)R^{\operatorname{RR}(q)} and not in RJR^{J} is easy to overlook, but overlooking it is dangerous. We have attempted to prove atomic Leibniz-style rules for more general families of cosets (core cosets, cosets whose core is atomic, etcetera). Each time what prevents one from bootstrapping from the atomic case to more general cases is the fact that Tq(f)T_{q}(f) does not live in RJR^{J}. The generalization in Example 2.4 has the special feature that the lower cosets are all core, so that their right redundancy equals JJ. (It also has the special feature that qcoreq^{\operatorname{core}} is atomic.)

3.3. Changing the realization

We argue that the atomic Leibniz rule for some realizations implies the atomic Leibniz rule for others. Given a realization, one can obtain another realization by applying base change ()𝕜𝕜(-)\otimes_{\Bbbk}\Bbbk^{\prime} to VV, and choosing new roots and coroots in the natural way. We call this a specialization. Here are two other common ways to alter the realization.

Definition 3.13.

Let (V,Δ,Δ)(V,\Delta,\Delta^{\vee}) be a realization of (W,S)(W,S) over 𝕜\Bbbk. Let NN be a free 𝕜\Bbbk-module acted on trivially by WW. Then (VN,Δ0,Δ0)(V\oplus N,\Delta\oplus 0,\Delta^{\vee}\oplus 0) is a realization, called a WW-invariant enlargement of the original. More precisely, the new roots are the image of the old roots under the inclusion map, and the new coroots kill the summand NN.

Definition 3.14.

Let (V,Δ,Δ)(V,\Delta,\Delta^{\vee}) be a realization of (W,S)(W,S) over 𝕜\Bbbk. Suppose one has a decomposition V=XYV=X\oplus Y of free 𝕜\Bbbk-modules, such that WW acts trivially on XX, though WW need not preserve YY. Note that the coroots necessarily annihilate XX. Then (Y,Δ¯,ΔY)(Y,\overline{\Delta},\Delta^{\vee}_{Y}) is a realization, called a WW-invariant quotient of the original. Here, we identify YY as the quotient V/XV/X, and Δ¯\overline{\Delta} represents the image of Δ\Delta under the quotient map. The functionals Δ\Delta^{\vee} kill XX, so they descend to functionals ΔY\Delta^{\vee}_{Y} on YY. We also make the technical assumption555This assumption is required for the WW-invariant quotient to satisfy Demazure surjectivity. that αs\alpha_{s} induces a surjective map Y𝕜Y^{*}\to\Bbbk.

Example 3.15.

Let (W,S)(W,S) have type A~n1\widetilde{A}_{n-1}, with simple reflections sis_{i} for 1in1\leq i\leq n. Let VV be the free \mathbb{Z}-module spanned by {xi}i=1n\{x_{i}\}_{i=1}^{n} and δ\delta. Let {xi}{δ}\{x_{i}^{*}\}\cup\{\delta^{*}\} denote the dual basis in Hom𝕜(V,𝕜)\operatorname{Hom}_{\Bbbk}(V,\Bbbk). With indices considered modulo nn, let αi=xixi+1+δ\alpha_{i}=x_{i}-x_{i+1}+\delta, and αi=xixi+1\alpha_{i}^{\vee}=x_{i}^{*}-x_{i+1}^{*}. This is a realization of (W,S)(W,S) called the affine permutation realization. Note that i=0n1αi=nδ\sum_{i=0}^{n-1}\alpha_{i}=n\delta, which is WW-invariant. Let XX be the span of δ\delta, and YY be the span of {xi}i=1n\{x_{i}\}_{i=1}^{n}. Note that WW does not preserve YY, since the roots are not contained in YY. There is a valid WW-invariant quotient (Y,Δ¯,ΔY)(Y,\overline{\Delta},\Delta^{\vee}_{Y}) which agrees, upon restriction to the parabolic subgroup 𝐒n\mathbf{S}_{n} generated by {si}i=1n1\{s_{i}\}_{i=1}^{n-1}, with the permutation representation.

Example 3.16.

Continuing the previous example, let yi=xiiδy_{i}=x_{i}-i\delta. Then we can also view VV as having basis {yi}i=1n{δ}\{y_{i}\}_{i=1}^{n}\cup\{\delta\}, and αi=yiyi+1\alpha_{i}=y_{i}-y_{i+1} for ini\neq n. Upon restriction to the parabolic subgroup 𝐒n\mathbf{S}_{n}, we see that VV is isomorphic to the WW-invariant enlargement of the permutation representation of 𝐒n\mathbf{S}_{n} (with basis {yi}\{y_{i}\}) by the WW-invariant span of δ\delta.

Indeed, WW has nn distinct maximal parabolic subgroups isomorphic to 𝐒n\mathbf{S}_{n} as groups. A similar construction will show that the restriction of VV to any maximal parabolic subgroup (a copy of 𝐒n\mathbf{S}_{n}) will be isomorphic to an invariant enlargement of its permutation representation.

Lemma 3.17.

If a rightward (resp. leftward) atomic Leibniz rule holds for a Frobenius realization, then it also holds for specializations, WW-invariant enlargements, and WW-invariant quotients.

Proof.

Let RR be the ring associated to the original realization, and RnewR_{\operatorname{new}} be the realization associated to the specialization, enlargement, or quotient. All three cases are united by the fact that RnewR_{\operatorname{new}} is a tensor product of the form RABR\otimes_{A}B, where ARA\subset R is a subring on which WW acts trivially, and BB is a ring on which WW acts trivially. For specializations we have Rnew=R𝕜𝕜R_{\operatorname{new}}=R\otimes_{\Bbbk}\Bbbk^{\prime}; for enlargements we have Rnew=R𝕜RNR_{\operatorname{new}}=R\otimes_{\Bbbk}R_{N}, where RNR_{N} is the polynomial ring of NN; for quotients we have Rnew=RRX𝕜R_{\operatorname{new}}=R\otimes_{R_{X}}\Bbbk, where RXR_{X} is the polynomial ring of XX, and 𝕜\Bbbk is its quotient by the ideal of positive degree elements. For wWw\in W, its action on RnewR_{\operatorname{new}} is given by widw\otimes\operatorname{id}. The roots in RnewR_{\operatorname{new}} are given by αs1\alpha_{s}\otimes 1, and the Demazure operators snew\partial^{\operatorname{new}}_{s} on RnewR_{\operatorname{new}} have the form sid\partial_{s}\otimes\operatorname{id}.

The important point in all three cases is that for each ISI\subset S finitary we have RnewI=RIABR_{\operatorname{new}}^{I}=R^{I}\otimes_{A}B. We now prove this somewhat subtle point. There is an obvious inclusion RIABRnewIR^{I}\otimes_{A}B\subset R_{\operatorname{new}}^{I}, so we need only show the other inclusion.

It is straightforward to verify that the new realization satisfies generalized Demazure surjectivity. A consequence is that the typical properties of Demazure operators are satisfied. For example, the kernel and the image of snew\partial^{\operatorname{new}}_{s} are both equal to RnewsR_{\operatorname{new}}^{s}, and snew\partial^{\operatorname{new}}_{s} is RnewsR^{s}_{\operatorname{new}}-linear. It follows that Inew\partial^{\operatorname{new}}_{I} is also RnewIR_{\operatorname{new}}^{I}-linear.

Suppose that gRnewIg\in R_{\operatorname{new}}^{I}, and write g=fibig=\sum f_{i}\otimes b_{i}. Choose some PRP\in R with I(P)=1\partial_{I}(P)=1, which exists by generalized Demazure surjectivity. Then Inew(P1)=11\partial^{\operatorname{new}}_{I}(P\otimes 1)=1\otimes 1 in RnewR_{\operatorname{new}}. Thus

(38) g=gInew(P1)=Inew(g(P1))=I(fiP)bi.g=g\partial^{\operatorname{new}}_{I}(P\otimes 1)=\partial^{\operatorname{new}}_{I}(g\cdot(P\otimes 1))=\sum\partial_{I}(f_{i}\cdot P)\otimes b_{i}.

Thus gRIABg\in R^{I}\otimes_{A}B.

The rest of the proof is straightforward. Fix an atomic coset 𝚊{\mathtt{a}}. For each q<𝚊q<{\mathtt{a}}, given operators TqT_{q} for the original realization satisfying (30), we define Tq,new:=TqidT_{q,\operatorname{new}}:=T_{q}\otimes\operatorname{id}. By linearity, we need only check (30) for RnewR_{\operatorname{new}} on elements in RnewJR_{\operatorname{new}}^{J} of the form fb1f\otimes b_{1} and gb2g\otimes b_{2} for f,gRJf,g\in R^{J}. It is easy to verify (30) for RnewR_{\operatorname{new}} on such elements, since all operators (like 𝚊¯\underline{{\mathtt{a}}} or q¯wJ1\partial_{\overline{q}w_{J}^{-1}}) are applied only to the first tensor factor, where we can use the atomic Leibniz rule from RR. We conclude by noting that Tq,newT_{q,\operatorname{new}} has the appropriate codomain as well. ∎

We do not claim that any statements about the unicity of the operators TqT_{q} will extend from a realization to its specializations, enlargements, or quotients.

Lemma 3.18.

Let (𝕜,V,Δ,Δ)(\Bbbk,V,\Delta,\Delta^{\vee}) be a realization of (W,S)(W,S). If one can prove an atomic Leibniz rule for the restriction of VV to WMW_{M}, for all (maximal) finitary subsets MSM\subset S, then an atomic Leibniz rule holds for WW.

Proof.

Every atomic coset in WW lives within WMW_{M} for some finitary MM (which lives within a maximal finitary subset), and the same atomic Leibniz rule which works for WMW_{M} will work for WW. ∎

Example 3.19.

Suppose one can prove an atomic Leibniz rule for the permutation realization of 𝐒n\mathbf{S}_{n} over \mathbb{Z}. Then by enlargement, one obtains an atomic Leibniz rule for the affine permutation realization restricted to any finite parabolic subgroup, see Example 3.16. By the previous lemma, an atomic Leibniz rule holds for the affine permutation realization of the affine Weyl group of type A~n1\widetilde{A}_{n-1}.

4. Lower terms

In this section we give an explicit description of the ideal of lower terms for an atomic coset using the technology of singular light leaves.

4.1. Definition of lower terms

Definition 4.1.

Let

I=[[I=I0K1I1KmIm=J]]I_{\bullet}=[[I=I_{0}\subset K_{1}\supset I_{1}\subset\ldots K_{m}\supset I_{m}=J]]

be a multistep (I,J)(I,J) expression. To this expression we associate a (singular) Bott-Samelson bimodule

BS(I):=RI0RK1RI1RK2RKmRIm.\operatorname{BS}(I_{\bullet}):=R^{I_{0}}\otimes_{R^{K_{1}}}R^{I_{1}}\otimes_{R^{K_{2}}}\cdots\otimes_{R^{K_{m}}}R^{I_{m}}.

This is an (RI,RJ)(R^{I},R^{J})-bimodule. The collection of all Bott-Samelson bimodules is closed under tensor product, and forms (the set of objects in) a full sub-2-category of the 2-category of bimodules. This sub-2-category is denoted 𝐒𝐁𝐒𝐁𝐢𝐦\mathbf{SBSBim}.

For two (RI,RJ)(R^{I},R^{J})-bimodules BB and BB^{\prime}, Hom(B,B)\operatorname{Hom}(B,B^{\prime}) denotes the space of bimodule maps. Moreover, Hom(B,B)\operatorname{Hom}(B,B^{\prime}) is itself an (RI,RJ)(R^{I},R^{J})-bimodule in the usual way.

Inside any linear category, given a collection of objects, their identity maps generate a two-sided ideal. This ideal consists of all morphisms which factor through one of those objects, and linear combinations thereof. In the context of (RI,RJ)(R^{I},R^{J})-bimodules, the actions of RIR^{I} and RJR^{J} commute with any morphism, and thus preserve the factorization of morphisms. Hence the morphisms within any such ideal form a sub-bimodule of the original Hom space.

Definition 4.2.

Let pp be an (I,J)(I,J)-coset. Consider the set of reduced expressions MM_{\bullet} for any (I,J)(I,J)-coset qq with q<pq<p. Let Hom<p\operatorname{Hom}_{<p} denote the ideal in the category of (RI,RJ)(R^{I},R^{J})-bimodules generated by the identity maps of BS(M)\operatorname{BS}(M_{\bullet}) for such expressions. Then Hom<p\operatorname{Hom}_{<p} is a two-sided ideal, the ideal of lower terms relative to pp. The ideal Homp\operatorname{Hom}_{\leq p} is defined similarly.

So Hom<p(B,B)\operatorname{Hom}_{<p}(B,B^{\prime}) is a subset of Hom(B,B)\operatorname{Hom}(B,B^{\prime}), and is a sub-bimodule for (RI,RJ)(R^{I},R^{J}). We write End<p(B)\operatorname{End}_{<p}(B) instead of Hom<p(B,B)End(B)\operatorname{Hom}_{<p}(B,B)\subset\operatorname{End}(B).

We now focus on the case of atomic cosets. We use the letter 𝚊{\mathtt{a}} to denote an atomic coset and let [[IMJ]][[I\subset M\supset J]] denote the unique reduced expression of 𝚊{\mathtt{a}}. We let

(39) B𝚊:=BS([[IMJ]])=RIRMRJ.B_{\mathtt{a}}:=\operatorname{BS}([[I\subset M\supset J]])=R^{I}\otimes_{R^{M}}R^{J}.

Because B𝚊B_{\mathtt{a}} is generated by 111\otimes 1 as a bimodule, any endomorphism is determined by where it sends this element. Thus

(40) End(B𝚊)RIRMRJ\operatorname{End}(B_{\mathtt{a}})\cong R^{I}\otimes_{R^{M}}R^{J}

as (RI,RJ)(R^{I},R^{J})-bimodules, via the operations of left and right multiplication. Hence End(B𝚊)B𝚊\operatorname{End}(B_{\mathtt{a}})\cong B_{\mathtt{a}} as (RI,RJ)(R^{I},R^{J})-bimodules666We have ignored gradings in this paper. Using traditional grading conventions for Bott-Samelson bimodules, End(B𝚊)\operatorname{End}(B_{\mathtt{a}}) and B𝚊B_{\mathtt{a}} are only isomorphic up to shift. The identity map of End(B𝚊)\operatorname{End}(B_{\mathtt{a}}) is in degree zero, while 11B𝚊1\otimes 1\in B_{\mathtt{a}} is not.. It is easy to deduce that B𝚊B_{\mathtt{a}} is indecomposable (when 𝕜\Bbbk is a domain) since there are no non-trivial idempotents in End(B𝚊)B𝚊\operatorname{End}(B_{\mathtt{a}})\cong B_{\mathtt{a}}.

4.2. Atomic double leaves

The goal of the section is to describe a large family of morphisms in End(B𝚊)\operatorname{End}(B_{{\mathtt{a}}}) called double leaves, most of which are in End<𝚊(B𝚊)\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}}}) by construction. We use the diagrammatic technology originally found in [ESW17] and developed further in [EKLP24a].

We assume a Frobenius realization, see Definition 3.5. In particular, the ring inclusions RIRJR^{I}\subset R^{J} are Frobenius extensions. Under these assumptions, a diagammatic 2-category 𝐅𝐫𝐨𝐛\mathbf{Frob} is constructed in [ESW17], and it comes equipped with a 2-functor to 𝐒𝐁𝐒𝐁𝐢𝐦\mathbf{SBSBim}. This 2-functor is essentially surjective, but is not expected to be an equivalence; the category 𝐅𝐫𝐨𝐛\mathbf{Frob} is missing a number of relations.

Double leaves are to be constructed either as morphisms in 𝐅𝐫𝐨𝐛\mathbf{Frob}, or as their images in 𝐒𝐁𝐒𝐁𝐢𝐦\mathbf{SBSBim}, depending on the context.

The objects in 𝐅𝐫𝐨𝐛\mathbf{Frob} are indexed not by multistep expressions but by singlestep expressions. An (I,J)(I,J) singlestep expression is a sequence

I=[I=I0,I1,,Id=J]I_{\bullet}=[I=I_{0},I_{1},\ldots,I_{d}=J]

where each IiI_{i} is a finitary subset of SS, and each IiI_{i} and Ii+1I_{i+1} differ by the addition or removal of a single simple reflection. We use single brackets for singlestep expressions, and double brackets for multistep expressions.

Throughout this section, we fix IM=IsSI\subset M=Is\subset S finitary, and let t=wMswMt=w_{M}sw_{M} and J=MtJ=M\setminus t, so that

𝚊[I,M,J]{\mathtt{a}}\leftrightharpoons[I,M,J]

is an atomic coset. We also fix the (I,M)(I,M)-coset

n=WIeWM.n=W_{I}eW_{M}.

4.2.1. Elementary light leaves for atomic Grassmannian pairs

By definition of atomic, 𝚊¯=wM\overline{{\mathtt{a}}}=w_{M}. Then for an (I,J)(I,J)-coset qq, then condition q𝚊q\leq{\mathtt{a}} is equivalent (see [EKLP23b, Thm 2.16]) to q¯wM\overline{q}\leq w_{M}, which in turn is equivalent to qn=WM.q\subseteq n=W_{M}.

For an (I,J)(I,J)-coset qq contained in WMW_{M}, the pair qnq\subset n is Grassmannian in the sense of [EKLP24a, Definition 2.7]. Associated to such a pair, [EKLP24a, Section 7.3] constructs a distinguished map called an elementary light leaf. The map (and codomain of the map) depends on a choice we make now: we fix a reduced expression XqX_{q} of the form

(41) Xq=[[ILR(q)]]Xqcore[[RR(q)J]]X_{q}=[[I\supset\operatorname{LR}(q)]]\circ X_{q}^{\operatorname{core}}\circ[[\operatorname{RR}(q)\subset J]]

where XqcoreX_{q}^{\operatorname{core}} is a reduced expression of qcoreq^{\operatorname{core}}.

Definition 4.3.

Let qq be an (I,J)(I,J)-coset contained in nn. The elementary light leaf associated to [n,q][n,q] (and XqX_{q}) is the (RI,RJ)(R^{I},R^{J})-bimodule morphism

𝔼𝕃𝕃([n,q]):BS([I,M,J])BS(Xq)\mathbb{ELL}([n,q]):\operatorname{BS}([I,M,J])\to\operatorname{BS}(X_{q})

sending the generator 11RIRMRJ=BS([I,M,J])1\otimes 1\in R^{I}\otimes_{R^{M}}R^{J}=\operatorname{BS}([I,M,J]) to the element 1:=11BS(Xq)1^{\otimes}:=1\otimes\cdots\otimes 1\in\operatorname{BS}(X_{q}). Equivalently, 𝔼𝕃𝕃([n,q])\mathbb{ELL}([n,q]) is defined by the diagrammatic construction in [EKLP24a, Section 7.3] (see also [EKLP24a, Lemma 8.10]).

We refer to [EKLP24a] and [ESW17] for a diagrammatic exhibition of morphisms between Bott-Samelson bimodules. The morphism 𝔼𝕃𝕃([n,q])\mathbb{ELL}([n,q]) is determined by the condition that its diagram consists only of counterclockwise cups and right-facing crossings, as in the following diagram.

\labellist\hair1pt\pinlabelLR(q)[]at6050\pinlabelRR(q)[]at24550\pinlabelI[]at3514\pinlabelM[]at15014\pinlabelJ[]at28014\endlabellist[Uncaptioned image]{\labellist\small\hair 1pt\pinlabel{\operatorname{LR}(q)}[]at6050\pinlabel{\operatorname{RR}(q)}[]at24550\pinlabel{I}[]at3514\pinlabel{M}[]at15014\pinlabel{J}[]at28014\endlabellist\centering\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 129.00195pt\hbox{{\hbox{\kern-129.00195pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-27.4626pt\hbox{$\textstyle{\includegraphics[scale={1.1}]{./EPS/ELLat}}$}}}}}}}}}}\@add@centering}

In our examples, we color the simple reflection s{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}s} in strawberry, and t{\color[rgb]{0,.5,.5}\definecolor[named]{pgfstrokecolor}{rgb}{0,.5,.5}t} in teal. Sometimes s=ts=t, which will force us to change our convention.

In type AA the expression XqcoreX_{q}^{\operatorname{core}} takes a simple form, and thus 𝔼𝕃𝕃([n,q])\mathbb{ELL}([n,q]) could be described more explicitly.

Example 4.4.

Let WM=𝐒a+bW_{M}=\mathbf{S}_{a+b} be a symmetric group, for some aba\neq b, and let I=s^I=\widehat{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}s}} be such that WI=𝐒b×𝐒a𝐒a+bW_{I}=\mathbf{S}_{b}\times\mathbf{S}_{a}\subset\mathbf{S}_{a+b}. Then J=t^J=\widehat{{\color[rgb]{0,.5,.5}\definecolor[named]{pgfstrokecolor}{rgb}{0,.5,.5}t}} is such that WJ=𝐒a×𝐒b𝐒a+bW_{J}=\mathbf{S}_{a}\times\mathbf{S}_{b}\subset\mathbf{S}_{a+b}. As will be explained in Section 6 (Equation (88)), each (I,J)(I,J)-coset qq in WMW_{M} has a unique reduced expression XqX_{q} of the form (41).

  1. (1)

    If q=WIeWJq=W_{I}eW_{J} then we have

    qXq=[[s^s^t^]][s^t^][[s^t^t^]]=[s^t+s].q\leftrightharpoons X_{q}=[[\widehat{s}\supset\widehat{s}\widehat{t}]]\circ[\widehat{s}\widehat{t}]\circ[[\widehat{s}\widehat{t}\subset\widehat{t}]]=[\widehat{s}-{\color[rgb]{0,.5,.5}\definecolor[named]{pgfstrokecolor}{rgb}{0,.5,.5}t}+{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}s}].

    Here we have

    𝔼𝕃𝕃([n,q])=[Uncaptioned image].\mathbb{ELL}([n,q])=\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 14.09265pt\hbox{{\hbox{\kern-14.09265pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-14.09265pt\hbox{$\textstyle{\includegraphics[scale={.6}]{./EPS/rightcross}}$}}}}}}}}}}.
  2. (2)

    If q=WIwMWJq=W_{I}w_{M}W_{J} then we have

    q=𝚊Xq=[I,M,J].q={\mathtt{a}}\leftrightharpoons X_{q}=[I,M,J].

    Here we have 𝔼𝕃𝕃([n,q])=idBS([I,M,J])\mathbb{ELL}([n,q])=\operatorname{id}_{\operatorname{BS}([I,M,J])}.

  3. (3)

    Otherwise, we have

    qXq=[[s^s^k^^]][s^k^^k^^t^k^^][[t^k^^t^]]q\leftrightharpoons X_{q}=[[\widehat{s}\supset\widehat{s}{\widehat{k}\widehat{\ell}}]]\circ[\widehat{s}{\widehat{k}\widehat{\ell}}\subset\widehat{k}\widehat{\ell}\supset\widehat{t}\widehat{k}\widehat{\ell}]\circ[[\widehat{t}\widehat{k}\widehat{\ell}\subset\widehat{t}]]

    for distinct s,k,,tMs,k,\ell,t\in M. Here we have

    𝔼𝕃𝕃([n,q])=[Uncaptioned image].\mathbb{ELL}([n,q])=\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 19.5129pt\hbox{{\hbox{\kern-19.5129pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-14.45401pt\hbox{$\textstyle{\includegraphics[scale={.6}]{./EPS/qELL1}}$}}}}}}}}}}.

Here is a non-type A example.

Example 4.5.

Let (W,S)(W,S) be of type E6E_{6} where SS is indexed as in the Dynkin diagram 665544332211. Let M=SM=S and s=3s=3. Then wM3wM=5w_{M}3w_{M}=5 and thus

𝚊[3^+35]=[3^,M,5^]{\mathtt{a}}\leftrightharpoons[\widehat{3}+3-5]=[\widehat{3},M,\widehat{5}]

is an atom. For the (3^,5^)(\widehat{3},\widehat{5})-coset q<𝚊q<{\mathtt{a}} with reduced expression

qXq=[[3^{4,6}]][+34+5+2+1623+45][[{1,4}5^]],q\leftrightharpoons X_{q}=[[\widehat{3}\supset\{4,6\}]]\circ[+3-4+5+2+1-6-2-3+4-5]\circ[[\{1,4\}\subset\widehat{5}]],

the elementary light leaf 𝔼𝕃𝕃([n,q])\mathbb{ELL}([n,q]) is

\labellist\hair2pt\pinlabel1[]at683\pinlabel2[]at1883\pinlabel5[]at3083\pinlabel3[]at4683\pinlabel4[]at5883\pinlabel5[]at7083\pinlabel2[]at8283\pinlabel1[]at9483\pinlabel6[]at10683\pinlabel2[]at11883\pinlabel3[]at13083\pinlabel4[]at14283\pinlabel5[]at15483\pinlabel3[]at17083\pinlabel2[]at18283\pinlabel6[]at19483\endlabellist [Uncaptioned image] {\labellist\small\hair 2pt\pinlabel{1}[]at683\pinlabel{2}[]at1883\pinlabel{5}[]at3083\pinlabel{3}[]at4683\pinlabel{4}[]at5883\pinlabel{5}[]at7083\pinlabel{2}[]at8283\pinlabel{1}[]at9483\pinlabel{6}[]at10683\pinlabel{2}[]at11883\pinlabel{3}[]at13083\pinlabel{4}[]at14283\pinlabel{5}[]at15483\pinlabel{3}[]at17083\pinlabel{2}[]at18283\pinlabel{6}[]at19483\endlabellist\centering\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 74.07675pt\hbox{{\hbox{\kern-74.07675pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-31.07611pt\hbox{$\textstyle{\includegraphics[scale={1}]{./EPS/qELL2}}$}}}}}}}}}}\@add@centering}

4.2.2. Atomic double leaves

There is a contravariant (but monoidally-covariant) “duality” functor 𝒟\mathcal{D} from 𝐅𝐫𝐨𝐛\mathbf{Frob} to itself defined as follows:

  • it preserves objects and 11-morphisms,

  • on 22-morphisms, it flips each diagram upside-down and reverses all the orientations.

This functor is an involution.

Definition 4.6.

Given q𝚊q\leq{\mathtt{a}} and bRRR(q)b\in R^{\operatorname{RR}(q)}, the associated (right-sprinkled) double leaf 𝔻𝕃𝕃r(q,b)\mathbb{DLL}_{r}(q,b) is the composition

(42) B𝚊𝔼𝕃𝕃([n,q])BS(Xq)𝑏BS(Xq)𝒟(𝔼𝕃𝕃([n,q]))B𝚊.B_{\mathtt{a}}\xrightarrow{\mathbb{ELL}([n,q])}\operatorname{BS}(X_{q})\xrightarrow{b}\operatorname{BS}(X_{q})\xrightarrow{\mathcal{D}(\mathbb{ELL}([n,q]))}B_{\mathtt{a}}.

The middle map in (42) uses that XqX_{q} has the form (41): the map is multiplication by the element

1b1BS([[ILR(q)]]Xqcore)RRR(q)RRR(q)RJRJ.1^{\otimes}\otimes b\otimes 1\in\operatorname{BS}([[I\supset\operatorname{LR}(q)]]\circ X_{q}^{\operatorname{core}})\otimes_{R^{\operatorname{RR}(q)}}R^{\operatorname{RR}(q)}\otimes_{R^{J}}R^{J}.

In diagrams, we have

𝔻𝕃𝕃r(q,b)=\labellist\hair1pt\pinlabelb[]at24564\endlabellist[Uncaptioned image]\mathbb{DLL}_{r}(q,b)={\labellist\small\hair 1pt\pinlabel{b}[]at24564\endlabellist\centering\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 105.51419pt\hbox{{\hbox{\kern-105.51419pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-40.83255pt\hbox{$\textstyle{\includegraphics[scale={.9}]{./EPS/DLLat}}$}}}}}}}}}}\@add@centering}
Definition 4.7.

Given q𝚊q\leq{\mathtt{a}} and bRLR(q)b\in R^{\operatorname{LR}(q)}, the associated left-sprinkled double leaf 𝔻𝕃𝕃l(q,b)\mathbb{DLL}_{l}(q,b) is the composition

(43) B𝚊𝔼𝕃𝕃([n,q])BS(Xq)𝑏BS(Xq)𝒟(𝔼𝕃𝕃([n,q]))B𝚊,B_{\mathtt{a}}\xrightarrow{\mathbb{ELL}([n,q])}\operatorname{BS}(X_{q})\xrightarrow{b}\operatorname{BS}(X_{q})\xrightarrow{\mathcal{D}(\mathbb{ELL}([n,q]))}B_{\mathtt{a}},

whose diagram is

\labellist\hair1pt\pinlabelb[]at6064\endlabellist[Uncaptioned image].{\labellist\small\hair 1pt\pinlabel{b}[]at6064\endlabellist\centering\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 105.51419pt\hbox{{\hbox{\kern-105.51419pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-40.83255pt\hbox{$\textstyle{\includegraphics[scale={.9}]{./EPS/DLLat}}$}}}}}}}}}}\@add@centering}.

The maps provided here (when bb ranges over a basis for RRR(q)R^{\operatorname{RR}(q)} or RLR(q)R^{\operatorname{LR}(q)}) are the same as the “double leaves basis” from [EKLP24a]. This is verified in the following remark, intended for a reader familiar with [EKLP24a].

Remark 4.8.

Let us verify that

(44) 𝔻𝕃𝕃l(q,b)=𝔻𝕃𝕃(q,([p,n,q],1),([p,n,q],1),b),\mathbb{DLL}_{l}(q,b)=\mathbb{DLL}(q,([p,n,q],1),([p,n,q],1),b),

where pp is the identity (I,I)(I,I)-coset, by following the stages in [EKLP24a, Chapter 7]. First, we construct 𝕃𝕃(q,([p,n,q],1))\mathbb{LL}(q,([p,n,q],1)). The single step light leaf for the first step [p,n][p,n] is the identity map since [p,n][p,n] is reduced. Since the left redundancy for pp and nn are the same, In the second step, we have the coset pair [n,q][n,q] which is already a Grassmannian pair, thus the single step light leaf is the elementary light leaf 𝔼𝕃𝕃([n,q])\mathbb{ELL}([n,q]).

Then we choose Xq=YqX_{q}=Y_{q} and take the dual map for the upside-down light leaf from BS(Xq)\operatorname{BS}(X_{q}) to B𝚊B_{\mathtt{a}}. Altogether we obtain the double leaf of the form (43). No rex moves were used at any stage of the process.

Moreover, for each q𝚊q\leq{\mathtt{a}}, there is one subordinate path with terminus qq, namely [p,n,q][p,n,q]. It follow that the morphisms 𝔻𝕃𝕃(q,([p,n,q],1),([p,n,q],1),b)\mathbb{DLL}(q,([p,n,q],1),([p,n,q],1),b) form a double leaves basis in the sense of [EKLP24a].

Remark 4.9.

We know that End(B𝚊)\operatorname{End}(B_{{\mathtt{a}}}) is an (RI,RJ)(R^{I},R^{J})-bimodule, so it is natural to ask how the actions of RIR^{I} and RJR^{J} interact with the bases presented in Propositions 4.18 and 4.20. For gRJg\in R^{J} we claim that

(45) 𝔻𝕃𝕃r(q,b)g=𝔻𝕃𝕃r(q,bg).\mathbb{DLL}_{r}(q,b)\cdot g=\mathbb{DLL}_{r}(q,b\cdot g).

Consider the diagram in Definition 4.6, and right-multiply by gRJg\in R^{J}. Since gg is also in RRR(q)R^{\operatorname{RR}(q)}, it can be slid from the right side to the region where bb lives.

However, the left action of fRIf\in R^{I} is more mysterious. For fixed qq, it does not preserve the span of {𝔻𝕃𝕃r(q,b)bRRR(q)}\{\mathbb{DLL}_{r}(q,b)\mid b\in R^{\operatorname{RR}(q)}\}. Indeed, the comparison between the left action and the right action is controlled by polynomial forcing for BS(Xq)\operatorname{BS}(X_{q}), which involves lower terms.

Similarly, for fRIf\in R^{I}, the left action on left-sprinkled double leaves is straightforward,

f𝔻𝕃𝕃l(q,b)=𝔻𝕃𝕃l(q,fb),f\cdot\mathbb{DLL}_{l}(q,b)=\mathbb{DLL}_{l}(q,f\cdot b),

whereas the right action of RJR^{J} is mysterious.

4.2.3. Evaluation of double leaves

The following crucial computation links double leaves with the description End(B𝚊)RIRMRJ\operatorname{End}(B_{\mathtt{a}})\cong R^{I}\otimes_{R^{M}}R^{J}. Let ΔM,(1)J\Delta^{J}_{M,(1)} and ΔM,(2)J\Delta^{J}_{M,(2)} be dual bases of RJR^{J} over RMR^{M}, where we use Sweedler notation.

Lemma 4.10.

The double leaf 𝔻𝕃𝕃r(q,b)\mathbb{DLL}_{r}(q,b) coincides with multiplication by the element

(46) q¯wJ1(bΔM,(1)J)ΔM,(2)J.\partial_{\overline{q}w_{J}^{-1}}\left(b\cdot\Delta^{J}_{M,(1)}\right)\otimes\Delta^{J}_{M,(2)}.

This can also be written as

(47) ILR(q)qcore(bιJRR(q)ΔM,(1)J)ΔM,(2)J.\partial_{I}^{\operatorname{LR}(q)}\partial_{q^{\operatorname{core}}}(b\cdot\iota_{J}^{\operatorname{RR}(q)}\Delta^{J}_{M,(1)})\otimes\Delta^{J}_{M,(2)}.
Proof.

The proof is immediate from [EKLP24a, Algorithm 8.12]. It is proven exactly as [EKLP24a, Lemma 6.10]. ∎

A similar computation involving left-sprinkled double leaves gives the following.

Lemma 4.11.

The double leaf 𝔻𝕃𝕃l(q,b)\mathbb{DLL}_{l}(q,b) coincides with multiplication by the element

(48) ILR(q)(bqcore(ιJRR(q)ΔM,(1)J))ΔM,(2)J.\partial_{I}^{\operatorname{LR}(q)}(b\cdot\partial_{q^{\operatorname{core}}}(\iota_{J}^{\operatorname{RR}(q)}\Delta^{J}_{M,(1)}))\otimes\Delta^{J}_{M,(2)}.
Example 4.12.

If q𝚊q\leq{\mathtt{a}} is the minimal (I,J)(I,J)-coset, namely q=WIeWJq=W_{I}eW_{J}, then we have two cases.

  1. (1)

    When t:=wMswMs{\color[rgb]{0,.5,.5}\definecolor[named]{pgfstrokecolor}{rgb}{0,.5,.5}t}:=w_{M}{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}s}w_{M}\neq{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}s}, we have

    q[I,K,J]=[It+s]q\leftrightharpoons[I,K,J]=[I-{\color[rgb]{0,.5,.5}\definecolor[named]{pgfstrokecolor}{rgb}{0,.5,.5}t}+{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}s}]

    for K=IJK=I\cap J. In this case, both left and right redundancies are KK, and for bRKb\in R^{K} the double leaf 𝔻𝕃𝕃r(q,b)=𝔻𝕃𝕃l(q,b)\mathbb{DLL}_{r}(q,b)=\mathbb{DLL}_{l}(q,b) is the left diagram in the equality

    (49) \labellist\hair2pt\pinlabelb[]at3132\endlabellist[Uncaptioned image]=IK(bΔM,(1)J) [Uncaptioned image] [Uncaptioned image] ΔM,(2)J.{\labellist\small\hair 2pt\pinlabel{b}[]at3132\endlabellist\centering\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 21.31964pt\hbox{{\hbox{\kern-21.31964pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-20.9583pt\hbox{$\textstyle{\includegraphics[scale={.9}]{./EPS/R2hard1_colorswap}}$}}}}}}}}}}\@add@centering}=\partial^{K}_{I}(b\cdot\Delta^{J}_{M,(1)})\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 7.94969pt\hbox{{\hbox{\kern-7.94969pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-21.681pt\hbox{$\textstyle{\includegraphics[scale={.9}]{./EPS/ups}}$}}}}}}}}}}\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 7.94969pt\hbox{{\hbox{\kern-7.94969pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-21.681pt\hbox{$\textstyle{\includegraphics[scale={.9}]{./EPS/downt}}$}}}}}}}}}}\Delta^{J}_{M,(2)}.

    Thus we have

    𝔻𝕃𝕃r(q,b)=𝔻𝕃𝕃l(q,b)=IK(bΔM,(1)J)ΔM,(2)J\mathbb{DLL}_{r}(q,b)=\mathbb{DLL}_{l}(q,b)=\partial^{K}_{I}(b\cdot\Delta^{J}_{M,(1)})\otimes\Delta^{J}_{M,(2)}

    (see [EKLP24a, Equations (89) and (91)]).

  2. (2)

    When wMswM=sw_{M}{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}s}w_{M}={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}s}, we have q[I]q\leftrightharpoons[I]. In this case, we have LR(q)=I=RR(q)\operatorname{LR}(q)=I=\operatorname{RR}(q), and for bRIb\in R^{I} the double leaf has a capcup diagram

    (50) \labellist\hair2pt\pinlabelb[]at1532\endlabellist[Uncaptioned image].{\labellist\small\hair 2pt\pinlabel{b}[]at1532\endlabellist\centering\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 20.9583pt\hbox{{\hbox{\kern-20.9583pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-21.31966pt\hbox{$\textstyle{\includegraphics[scale={.9}]{./EPS/cupcap}}$}}}}}}}}}}\@add@centering}.

    Thus we have 𝔻𝕃𝕃r(q,b)=𝔻𝕃𝕃l(q,b)=bΔMI=ΔMIb\mathbb{DLL}_{r}(q,b)=\mathbb{DLL}_{l}(q,b)=b\cdot\Delta_{M}^{I}=\Delta_{M}^{I}\cdot b.

Definition 4.13.

For an atomic coset 𝚊[I,M,J]{\mathtt{a}}\leftrightharpoons[I,M,J], let DL<𝚊\operatorname{DL}_{<{\mathtt{a}}} denote the 𝕜\Bbbk-linear subspace of End(B𝚊)\operatorname{End}(B_{{\mathtt{a}}}) spanned by right-sprinkled double leaves factoring through q<𝚊q<{\mathtt{a}}. More explicitly we have

DL<𝚊:=Spanq<𝚊{𝔻𝕃𝕃r(q,b)bRRR(q)}=Spanq<𝚊{q¯wJ1(RRR(q)ΔM,(1)J)ΔM,(2)J}.\operatorname{DL}_{<{\mathtt{a}}}:=\operatorname{Span}_{q<{\mathtt{a}}}\{\mathbb{DLL}_{r}(q,b)\mid b\in R^{\operatorname{RR}(q)}\}=\operatorname{Span}_{q<{\mathtt{a}}}\{\partial_{\overline{q}w_{J}^{-1}}(R^{\operatorname{RR}(q)}\cdot\Delta_{M,(1)}^{J})\otimes\Delta_{M,(2)}^{J}\}.
Lemma 4.14.

We have DL<𝚊End<𝚊(B𝚊)\operatorname{DL}_{<{\mathtt{a}}}\subset\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}}}).

Proof.

By construction, every double leaf associated to q<𝚊q<{\mathtt{a}} factors through a reduced expression for qq, and thus lives in End<𝚊(B𝚊)\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}}}). ∎

We also note a consequence of Remark 4.9.

Corollary 4.15.

For each gRJg\in R^{J} and bRRR(q)b\in R^{\operatorname{RR}(q)} we have

(51) q¯wJ1(bΔM,(1)J)ΔM,(2)Jg=q¯wJ1(bgΔM,(1)J)ΔM,(2)J.\partial_{\overline{q}w_{J}^{-1}}(b\cdot\Delta_{M,(1)}^{J})\otimes\Delta_{M,(2)}^{J}\cdot g=\partial_{\overline{q}w_{J}^{-1}}(b\cdot g\cdot\Delta_{M,(1)}^{J})\otimes\Delta_{M,(2)}^{J}.

In particular, DL<𝚊\operatorname{DL}_{<{\mathtt{a}}} is a right RJR^{J}-module.

Proof.

This follows from (45) and Lemma 4.10. ∎

4.3. Double leaves and lower terms: part I

The main result of [EKLP24a] is that double leaves form a basis for morphisms between Bott-Samelson bimodules. However, [EKLP24a] relies on Williamson’s theory of standard filtrations, which relies on several assumptions originally made by Soergel.

Definition 4.16.

We call a realization a Soergel-Williamson realization or an SW-realization for short, if it is a Frobenius realization (see Definition 3.5), and it also satisfies the following assumptions.

  • The realization is reflection faithful, i.e. it is faithful, and the reflections in WW are exactly those elements that fix a codimension-one subspace.

  • The ring 𝕜\Bbbk is an infinite field of characteristic not equal to 22.

Remark 4.17.

Abe [Abe24] has recently developed a theory of singular Soergel bimodules that works for Frobenius realizations, without the extra restrictions of an SW-realization. One expects that the results of [EKLP24a] can be straightforwardly generalized to Abe’s setting. For simplicity and because of the current state of the literature, we will work with Williamson’s category of bimodules.

Proposition 4.18.

Assume an SW-realization. Let 𝔹q\mathbb{B}_{q} be a 𝕜\Bbbk-basis of RRR(q)R^{\operatorname{RR}(q)}, for each q𝚊q\leq{\mathtt{a}}. Then

(52) {𝔻𝕃𝕃r(q,b)}b𝔹q\{\mathbb{DLL}_{r}(q,b)\}_{b\in\mathbb{B}_{q}}

gives a basis of Endq(B𝚊)/End<q(B𝚊)\operatorname{End}_{\leq q}(B_{{\mathtt{a}}})/\operatorname{End}_{<q}(B_{{\mathtt{a}}}) over 𝕜\Bbbk. In particular,

(53) {𝔻𝕃𝕃r(q,b)|q𝚊,b𝔹q}\{\mathbb{DLL}_{r}(q,b)\ |\ q\leq{\mathtt{a}},b\in\mathbb{B}_{q}\}

is a 𝕜\Bbbk-basis of End(B𝚊)\operatorname{End}(B_{{\mathtt{a}}}), and the subset indexed by q<𝚊q<{\mathtt{a}} is a basis for End<𝚊(B𝚊)\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}}}).

Proof.

This is a special case of [EKLP24a, Lemma 8.35 and Theorem 7.49], as confirmed in Remark 4.8. ∎

Corollary 4.19.

Assume an SW-realization. We have

(54) DL<𝚊=End<𝚊(B𝚊)=q<𝚊q¯wJ1(RRR(q)ΔM,(1)J)ΔM,(2)J.\operatorname{DL}_{<{\mathtt{a}}}=\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}}})=\bigoplus_{q<{\mathtt{a}}}\partial_{\overline{q}w_{J}^{-1}}(R^{\operatorname{RR}(q)}\cdot\Delta_{M,(1)}^{J})\otimes\Delta_{M,(2)}^{J}.

In particular, DL<𝚊\operatorname{DL}_{<{\mathtt{a}}} is an (RI,RJ)(R^{I},R^{J})-bimodule.

Proof.

As DL<𝚊\operatorname{DL}_{<{\mathtt{a}}} is the span of double leaves factoring through q<𝚊q<{\mathtt{a}}, the first equality follows from Proposition 4.18. The second equality follows from the linear independence of double leaves. Since End<𝚊(B𝚊)\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}}}) is an (RI,RJ)(R^{I},R^{J})-bimodule, so is DL<𝚊\operatorname{DL}_{<{\mathtt{a}}}. ∎

Similarly, double leaves provides a left-sprinkled basis.

Proposition 4.20.

Assume an SW-realization. Let 𝔹q\mathbb{B}_{q} be a 𝕜\Bbbk-basis of RLR(q)R^{\operatorname{LR}(q)}, for each q𝚊q\leq{\mathtt{a}}. Then

(55) {𝔻𝕃𝕃l(q,b)}b𝔹q\{\mathbb{DLL}_{l}(q,b)\}_{b\in\mathbb{B}_{q}}

gives a basis of Endq(B𝚊)/End<q(B𝚊)\operatorname{End}_{\leq q}(B_{{\mathtt{a}}})/\operatorname{End}_{<q}(B_{{\mathtt{a}}}) over 𝕜\Bbbk. In particular,

(56) {𝔻𝕃𝕃l(q,b)|q𝚊,b𝔹q}\{\mathbb{DLL}_{l}(q,b)\ |\ q\leq{\mathtt{a}},b\in\mathbb{B}_{q}\}

is a 𝕜\Bbbk-basis of End(B𝚊)\operatorname{End}(B_{{\mathtt{a}}}), and the subset indexed by q<𝚊q<{\mathtt{a}} is a basis for End<𝚊(B𝚊)\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}}}).

4.4. Double leaves and lower terms: part II

Definition 4.21.

An almost-SW realization is a Frobenius realization, together with the following assumptions.

  • The ring 𝕜\Bbbk is a domain with fraction field 𝔽{\mathbb{F}}.

  • After base change to 𝔽{\mathbb{F}}, the result is an SW-realization.

  • Finitely-generated projective modules over 𝕜\Bbbk are free.

Example 4.22.

The defining representation of 𝐒n\mathbf{S}_{n} over \mathbb{Z} is an almost-SW realization [Dem73, Lemma 5].

Example 4.23.

The root realization of a Weyl group is almost-SW when defined over 𝕜=[1/N]\Bbbk=\mathbb{Z}[1/N] for small NN (N=30N=30 will suffice for all Weyl groups by [Dem73, Proposition 8]).

Lemma 4.24.

Let ISI\subset S be finitary. Then RIR^{I} is a free 𝕜\Bbbk-module.

Proof.

As a polynomial ring over a free 𝕜\Bbbk-module VV, RR is a free 𝕜\Bbbk-module. By the assumption of generalized Demazure surjectivity, RR is free as an RIR^{I}-module when ISI\subset S is finitary. Thus RIR^{I} is a direct summand of RR, and is therefore projective as a 𝕜\Bbbk-module. Since both RR and RIR^{I} are finitely-generated as 𝕜\Bbbk-modules in each graded degree, we deduce that RIR^{I} is also a free module over 𝕜\Bbbk. ∎

Our goal in this section is to generalize the results of the previous section to almost-SW realizations. In all the lemmas in this section, we assume an almost-SW realization. First we note the compatibility of base change with most of the constructions above.

We let RR be the polynomial ring of the realization over 𝕜\Bbbk, and let R𝔽:=R𝕜𝔽R_{\mathbb{F}}:=R\otimes_{\Bbbk}{\mathbb{F}} be the polynomial ring of the realization after base change. Let R𝔽IR𝔽R^{I}_{\mathbb{F}}\subset R_{{\mathbb{F}}} be the invariant subring.

Lemma 4.25.

We have R𝔽IRI𝕜𝔽R^{I}_{{\mathbb{F}}}\cong R^{I}\otimes_{\Bbbk}{\mathbb{F}}.

Proof.

There is a natural map RI𝕜𝔽R𝔽R^{I}\otimes_{\Bbbk}{\mathbb{F}}\to R_{{\mathbb{F}}}, and since scalars are WW-invariant, the image lies within R𝔽IR_{{\mathbb{F}}}^{I}. The map is injective since 𝔽{\mathbb{F}} is flat over 𝕜\Bbbk. We now argue that the map RI𝕜𝔽R𝔽IR^{I}\otimes_{\Bbbk}{\mathbb{F}}\to R_{{\mathbb{F}}}^{I} is surjective. If fR𝔽If\in R_{{\mathbb{F}}}^{I}, then there is some c𝕜c\in\Bbbk such that cfRcf\in R (e.g. letting cc be the product of the denominators of each monomial in ff). Clearly cfRIcf\in R^{I}, whence ff is the image of cf1ccf\otimes\frac{1}{c}.

Let B𝚊,𝔽=R𝔽IR𝔽MR𝔽JB_{{\mathtt{a}},{\mathbb{F}}}=R^{I}_{\mathbb{F}}\otimes_{R^{M}_{\mathbb{F}}}R^{J}_{\mathbb{F}}. If I=[[I=I0K1I1KmIm=J]]I_{\bullet}=[[I=I_{0}\subset K_{1}\supset I_{1}\subset\ldots K_{m}\supset I_{m}=J]] is a multistep (I,J)(I,J)-expression, let

BS𝔽(I):=R𝔽I0R𝔽K1R𝔽I1R𝔽K2R𝔽KmR𝔽Im.\operatorname{BS}_{\mathbb{F}}(I_{\bullet}):=R^{I_{0}}_{\mathbb{F}}\otimes_{R_{\mathbb{F}}^{K_{1}}}R_{\mathbb{F}}^{I_{1}}\otimes_{R_{\mathbb{F}}^{K_{2}}}\cdots\otimes_{R_{\mathbb{F}}^{K_{m}}}R_{\mathbb{F}}^{I_{m}}.
Lemma 4.26.

The natural inclusion map BS(I)BS(I)𝕜𝔽\operatorname{BS}(I_{\bullet})\to\operatorname{BS}(I_{\bullet})\otimes_{\Bbbk}{\mathbb{F}} is injective. We have BS𝔽(I)BS(I)𝕜𝔽\operatorname{BS}_{\mathbb{F}}(I_{\bullet})\cong\operatorname{BS}(I_{\bullet})\otimes_{\Bbbk}{\mathbb{F}}. As a consequence we have an injective map

(57) Hom(BS(I),BS(I))Hom(BS𝔽(I),BS𝔽(I)).\operatorname{Hom}(\operatorname{BS}(I_{\bullet}),\operatorname{BS}(I^{\prime}_{\bullet}))\to\operatorname{Hom}(\operatorname{BS}_{{\mathbb{F}}}(I_{\bullet}),\operatorname{BS}_{{\mathbb{F}}}(I^{\prime}_{\bullet})).
Proof.

By our assumptions from §3.1, RIR^{I} is free over RKR^{K} whenever IKI\subset K. We fix a basis {biI,K}\{b_{i}^{I,K}\} for this extension. Hence the Bott-Samelson bimodule BS(I)\operatorname{BS}(I_{\bullet}) is free as a right RImR^{I_{m}}-module with basis

(58) {bi1I0,K1bimIm1,Km1}.\left\{b_{i_{1}}^{I_{0},K_{1}}\otimes\ldots\otimes b_{i_{m}}^{I_{m-1},K_{m}}\otimes 1\right\}.

It is also free as a right 𝕜\Bbbk-module by 4.24. So base change is injective on bimodules. Notice that {biI,K1}\{b_{i}^{I,K}\otimes 1\} is a basis of R𝔽IR^{I}_{\mathbb{F}} over R𝔽KR^{K}_{\mathbb{F}}. Hence (58) gives also a basis of BS𝔽(I)\operatorname{BS}_{{\mathbb{F}}}(I_{\bullet}) over R𝔽ImR^{I_{m}}_{\mathbb{F}}. Since it sends a basis to a basis, we deduce that the natural map BS(I)𝕜𝔽BS𝔽(I)\operatorname{BS}(I_{\bullet})\otimes_{\Bbbk}{\mathbb{F}}\to\operatorname{BS}_{\mathbb{F}}(I_{\bullet}) is an isomorphism.

The localization functor gives the map in (57). For a morphism ϕ\phi between bimodules over 𝕜\Bbbk, let ϕ1\phi\otimes 1 denote its image, a morphism between bimodules over 𝔽{\mathbb{F}}. The restriction of ϕ1\phi\otimes 1 to the subset BS(I)BS𝔽(I)\operatorname{BS}(I_{\bullet})\subset\operatorname{BS}_{{\mathbb{F}}}(I_{\bullet}) is the original morphism ϕ\phi. Hence if ϕ1\phi\otimes 1 is the zero morphism, so is ϕ\phi. ∎

Lemma 4.27.

For q𝚊q\leq{\mathtt{a}} and bRRR(q)b\in R^{\operatorname{RR}(q)}, let us temporarily write 𝔻𝕃𝕃r,𝕜(q,b)\mathbb{DLL}_{r,\Bbbk}(q,b) for the double leaf as a morphism between Bott-Samelson bimodules over 𝕜\Bbbk, and 𝔻𝕃𝕃r,𝔽(q,b)\mathbb{DLL}_{r,{\mathbb{F}}}(q,b) for the double leaf as a morphism between Bott-Samelson bimodules over 𝔽{\mathbb{F}}, where for the latter we identify bb with its image in R𝔽RR(q)R_{{\mathbb{F}}}^{\operatorname{RR}(q)}. Under the map (57), we have

𝔻𝕃𝕃r,𝕜(q,b)𝔻𝕃𝕃r,𝔽(q,b).\mathbb{DLL}_{r,\Bbbk}(q,b)\mapsto\mathbb{DLL}_{r,{\mathbb{F}}}(q,b).
Proof.

The calculus from [ESW17] for interpreting diagrams is invariant under base change. Alternatively, B𝚊B_{{\mathtt{a}}} is generated as a bimodule by 111\otimes 1, and elementary light leaves 𝔼𝕃𝕃([n,q])\mathbb{ELL}([n,q]) are determined uniquely in their Hom space by the fact that they send 111\otimes 1 to 111\otimes\ldots\otimes 1. This property is preserved by base change. ∎

Lemma 4.28.

Let 𝔹q\mathbb{B}_{q} be a basis of RRR(q)R^{\operatorname{RR}(q)} over 𝕜\Bbbk. Then the set {𝔻𝕃𝕃r,𝕜(q,b)|q𝚊,b𝔹q}\{\mathbb{DLL}_{r,\Bbbk}(q,b)\ |\ q\leq{\mathtt{a}},b\in\mathbb{B}_{q}\} is linearly independent. Let DL<𝚊,𝕜\operatorname{DL}_{<{\mathtt{a}},\Bbbk} and DL<𝚊,𝔽\operatorname{DL}_{<{\mathtt{a}},{\mathbb{F}}} be defined as before for their respective realizations. The map (57) induces an isomorphism

(59) DL<𝚊,𝕜𝕜𝔽DL<𝚊,𝔽.\operatorname{DL}_{<{\mathtt{a}},\Bbbk}\otimes_{\Bbbk}\,{\mathbb{F}}\stackrel{{\scriptstyle\sim}}{{\to}}\operatorname{DL}_{<{\mathtt{a}},{\mathbb{F}}}.
Proof.

By definition DL<𝚊\operatorname{DL}_{<{\mathtt{a}}} is the span (over 𝕜\Bbbk or 𝔽{\mathbb{F}}) of the double leaf morphisms. The map (57) restricts to a map DL<𝚊,𝕜Hom(BS𝔽(I),BS𝔽(I))\operatorname{DL}_{<{\mathtt{a}},\Bbbk}\to\operatorname{Hom}(\operatorname{BS}_{{\mathbb{F}}}(I_{\bullet}),\operatorname{BS}_{{\mathbb{F}}}(I^{\prime}_{\bullet})). By the previous lemma, the image of this map is contained in DL<𝚊,𝔽\operatorname{DL}_{<{\mathtt{a}},{\mathbb{F}}}. Thus one has an induced map DL<𝚊,𝕜𝕜𝔽DL<𝚊,𝔽\operatorname{DL}_{<{\mathtt{a}},\Bbbk}\otimes_{\Bbbk}{\mathbb{F}}\to\operatorname{DL}_{<{\mathtt{a}},{\mathbb{F}}}.

Note that 𝔹q\mathbb{B}_{q} is sent by base change to a basis of R𝔽RR(q)R^{\operatorname{RR}(q)}_{{\mathbb{F}}} over 𝔽{\mathbb{F}}. The elements {𝔻𝕃𝕃r,𝕜(q,b)1}\{\mathbb{DLL}_{r,\Bbbk}(q,b)\otimes 1\} (ranging over the appropriate index set) form an 𝔽{\mathbb{F}}-spanning set for the left-hand side of (59), and are sent to {𝔻𝕃𝕃r,𝔽(q,b)}\{\mathbb{DLL}_{r,{\mathbb{F}}}(q,b)\}, which form a 𝔽{\mathbb{F}}-basis for DL<𝚊,𝔽\operatorname{DL}_{<{\mathtt{a}},{\mathbb{F}}} by Proposition 4.18. Thus the map (59) is an isomorphism, and the elements {𝔻𝕃𝕃r,𝕜(q,b)1}\{\mathbb{DLL}_{r,\Bbbk}(q,b)\otimes 1\} are linearly independent over 𝔽{\mathbb{F}}. Consequently, {𝔻𝕃𝕃r,𝕜(q,b)}\{\mathbb{DLL}_{r,\Bbbk}(q,b)\} are linearly independent over 𝕜\Bbbk. ∎

Lemma 4.29.

We have

DL<𝚊,𝕜=q<𝚊q¯wJ1(RRR(q)ΔM,(1)J)ΔM,(2)J.\operatorname{DL}_{<{\mathtt{a}},\Bbbk}=\bigoplus_{q<{\mathtt{a}}}\partial_{\overline{q}w_{J}^{-1}}(R^{\operatorname{RR}(q)}\cdot\Delta_{M,(1)}^{J})\otimes\Delta_{M,(2)}^{J}.
Proof.

The span in Definition 4.13 in indeed a direct sum of subspaces, by the linear independence shown in the previous lemma. ∎

Henceforth we return to the 𝕜\Bbbk-linear setting by default (i.e. in the absence of a subscript). Now the question remains: is the inclusion DL<𝚊End<𝚊(B𝚊)\operatorname{DL}_{<{\mathtt{a}}}\subset\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}}}) an equality over 𝕜\Bbbk, knowing that the result holds over 𝔽{\mathbb{F}}? In Corollary 4.33 below, we prove that the answer is yes.

Lemma 4.30.

For any ϕEnd<𝚊(B𝚊)\phi\in\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}}}) there is some n𝕜n\in\Bbbk such that nϕDL<𝚊n\phi\in\operatorname{DL}_{<{\mathtt{a}}}.

Proof.

In view of 4.26 and (57), we can regard both DL<𝚊\operatorname{DL}_{<{\mathtt{a}}} and End(B𝚊)\operatorname{End}(B_{{\mathtt{a}}}) as 𝕜\Bbbk-submodules of End(B𝚊,𝔽)\operatorname{End}(B_{{\mathtt{a}},{\mathbb{F}}}), which we identify with B𝚊,𝔽B_{{\mathtt{a}},{\mathbb{F}}}. We have End<𝚊(B𝚊)End<𝚊(B𝚊,𝔽)\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}}})\subset\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}},{\mathbb{F}}}) by definition. Meanwhile, 4.19 holds over 𝔽{\mathbb{F}} and thus

(60) End<𝚊(B𝚊,𝔽)q<𝚊q¯wJ1(R𝔽RR(q)ΔM,(1)J)ΔM,(2)J.\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}},{\mathbb{F}}})\cong\bigoplus_{q<{\mathtt{a}}}\partial_{\overline{q}w_{J}^{-1}}(R_{\mathbb{F}}^{\operatorname{RR}(q)}\cdot\Delta_{M,(1)}^{J})\otimes\Delta_{M,(2)}^{J}.

So any ϕEnd<𝚊(B𝚊)\phi\in\operatorname{End}_{<{\mathtt{a}}}(B_{\mathtt{a}}) is a 𝔽{\mathbb{F}}-multiple of an element of DL<𝚊\operatorname{DL}_{<{\mathtt{a}}}. Multiplying by the denominator, there exists n𝕜n\in\Bbbk such that nϕDL<𝚊n\phi\in\operatorname{DL}_{<{\mathtt{a}}}. ∎

We continue with a divisibility lemma on Demazure operators, which ensures that Demazure operators do not “produce” additional divisibility by elements of 𝕜\Bbbk.

Lemma 4.31.

Let bRb\in R and let n𝕜n\in\Bbbk. If nwMwJ1(bg)n\mid\partial_{w_{M}w_{J}^{-1}}(bg) for all gRJg\in R^{J}, then nbn\mid b.

Proof.

Assume that nbn\nmid b. We need to find gRJg\in R^{J} such that nwMwJ1(bg)n\nmid\partial_{w_{M}w_{J}^{-1}}(bg).

Recall from Lemma 2.6 that for wWw\in W and f,gRf,g\in R we have

w(fg)=xwTx(f)x(g)\partial_{w}(fg)=\sum_{x\leq w}T^{\prime}_{x}(f)\partial_{x}(g)

for certain operators TxT^{\prime}_{x} defined over 𝕜\Bbbk, where Tw(f)=w(f)T^{\prime}_{w}(f)=w(f). Let WJW^{J} be the subset of elements in WW which are minimal (for the Bruhat order) in their right WJW_{J}-coset. We have x(g)=0\partial_{x}(g)=0 when xWJx\not\in W^{J} and gRJg\in R^{J}, so we have

w(fg)=xw,xWJTx(f)x(g).\partial_{w}(fg)=\sum_{x\leq w,x\in W^{J}}T^{\prime}_{x}(f)\partial_{x}(g).

We apply this formula when w=wMwJ1w=w_{M}w_{J}^{-1}. Let WMJ=WJWMW^{J}_{M}=W^{J}\cap W_{M}. Since nbn\nmid b, then also nwMwJ1(b)=TwMwJ1(b)n\nmid w_{M}w_{J}^{-1}(b)=T_{w_{M}w_{J}^{-1}}^{\prime}(b). So there exists some yWMJy\in W_{M}^{J} (not necessarily unique) which is minimal with respect to the property that nTy(b)n\nmid T^{\prime}_{y}(b). Now we have

(61) wMwJ1(bg)=xWMJTx(b)x(g)xy,xWMJTx(b)x(g)(modn).\partial_{w_{M}w_{J}^{-1}}(bg)=\sum_{x\in W_{M}^{J}}T^{\prime}_{x}(b)\partial_{x}(g)\equiv\sum_{x\not<y,x\in W_{M}^{J}}T^{\prime}_{x}(b)\partial_{x}(g)\pmod{n}.

Recall that (xwM)=(wM)(x)\ell(xw_{M})=\ell(w_{M})-\ell(x) for all xWMx\in W_{M}. Let z=y1wMz=y^{-1}w_{M} so that y.z=wMy.z=w_{M}. We have

(wJy1wM)=(wM)(wJy1)=(wM)(y1)(wJ)=(z)(wJ),\ell(w_{J}y^{-1}w_{M})=\ell(w_{M})-\ell(w_{J}y^{-1})=\ell(w_{M})-\ell(y^{-1})-\ell(w_{J})=\ell(z)-\ell(w_{J}),

thus the left descent set of zz contains JJ. By [EKLP23a, Lemma 3.2] we have Im(z)RJ\operatorname{Im}(\partial_{z})\subset R^{J}. By our assumption of generalized Demazure surjectivity, we can choose PMRP_{M}\in R such that wM(PM)=1\partial_{w_{M}}(P_{M})=1. Set g=z(PM)RJg=\partial_{z}(P_{M})\in R^{J} and note that y(g)=1\partial_{y}(g)=1.

Let xWMx\in W_{M}. If x.zx.z is not reduced, then x(g)=xz(PM)=0\partial_{x}(g)=\partial_{x}\partial_{z}(P_{M})=0 by [EKLP23a, (27))]. If x.zx.z is reduced, then

(wM)(yx1)=(x.y1wM)=(x)+(wM)(y).\ell(w_{M})-\ell(yx^{-1})=\ell(x.y^{-1}w_{M})=\ell(x)+\ell(w_{M})-\ell(y).

It follows that (yx1)+(x)=(y)\ell(yx^{-1})+\ell(x)=\ell(y), so yx1.x=yyx^{-1}.x=y and, in particular, xyx\leq y. This means that x(g)0\partial_{x}(g)\neq 0 only if xyx\leq y.

Finally, we plug g=z(PM)g=\partial_{z}(P_{M}) into (61) and we observe that wMwJ1(bg)Ty(b)0(modn)\partial_{w_{M}w_{J}^{-1}}(bg)\equiv T^{\prime}_{y}(b)\not\equiv 0\pmod{n}. ∎

Now we can prove that DL<𝚊\operatorname{DL}_{<{\mathtt{a}}}, as a submodule of End(B𝚊)\operatorname{End}(B_{\mathtt{a}}), is closed under division by elements in 𝕜\Bbbk (when that makes sense).

Proposition 4.32.

Let ϕDL<𝚊\phi\in\operatorname{DL}_{<{\mathtt{a}}} and assume there exists n𝕜n\in\Bbbk such that 1nϕEnd(B𝚊)\frac{1}{n}\phi\in\operatorname{End}(B_{\mathtt{a}}). Then 1nϕDL<𝚊\frac{1}{n}\phi\in\operatorname{DL}_{<{\mathtt{a}}}.

Proof.

If nn is a unit in 𝕜\Bbbk the result is trivial, so assume otherwise.

Let ϕDL<𝚊\phi\in\operatorname{DL}_{<{\mathtt{a}}} and assume that 1nϕEnd(B𝚊)RIRMRJ\frac{1}{n}\phi\in\operatorname{End}(B_{\mathtt{a}})\cong R^{I}\otimes_{R^{M}}R^{J}. We can write

ϕ=q<𝚊q¯wJ1(bqΔM,(1)J)ΔM,(2)J\phi=\sum_{q<{\mathtt{a}}}\partial_{\overline{q}w_{J}^{-1}}(b_{q}\cdot\Delta_{M,(1)}^{J})\otimes\Delta_{M,(2)}^{J}

for some unique bqRRR(q)b_{q}\in R^{\operatorname{RR}(q)}. For clarity we choose to unravel Sweedler’s notation. Choose dual bases {ci}\{c_{i}\} and {di}\{d_{i}\} for RJR^{J} over RMR^{M} relative to the Frobenius trace map MJ\partial^{J}_{M}. We have

ϕ=iq<𝚊q¯wJ1(bqci)di.\phi=\sum_{i}\sum_{q<{\mathtt{a}}}\partial_{\overline{q}w_{J}^{-1}}(b_{q}\cdot c_{i})\otimes d_{i}.

Since did_{i} is a basis of RJR^{J} over RMR^{M}, any element of RIRMRJR^{I}\otimes_{R^{M}}R^{J} is uniquely expressible as ifidi\sum_{i}f_{i}\otimes d_{i} for fiRIf_{i}\in R^{I}. In particular, if nn divides ifidi\sum_{i}f_{i}\otimes d_{i} we have

1nifidi=ifidi\frac{1}{n}\sum_{i}f_{i}\otimes d_{i}=\sum_{i}f^{\prime}_{i}\otimes d_{i}

for some unique fif^{\prime}_{i}. Then fidi=infidi\sum f_{i}\otimes d_{i}=\sum_{i}nf^{\prime}_{i}\otimes d_{i}, and by the unicity mentioned before, this implies that nfi=finf_{i}^{\prime}=f_{i} for all i.i.

Consequently, ϕ\phi is divisible by nn if and only if for all ii we have

nq<𝚊q¯wJ1(bqci).n\mid\sum_{q<{\mathtt{a}}}\partial_{\overline{q}w_{J}^{-1}}(b_{q}\cdot c_{i}).

We want to show that all the bqb_{q} are actually divisible by nn, so that 1nϕDL<𝚊\frac{1}{n}\phi\in\operatorname{DL}_{<{\mathtt{a}}}.

Assume for contradiction that there exists a minimal rr such that nbrn\nmid b_{r}. Let zz be such that z.r¯wJ1=wMwJ1z.\overline{r}w_{J}^{-1}=w_{M}w_{J}^{-1}. Then by similar arguments to the previous lemma we have

0z(q<𝚊q¯wJ1(bqci))wMwJ1(brci)(modn)0\equiv\partial_{z}\left(\sum_{q<{\mathtt{a}}}\partial_{\overline{q}w_{J}^{-1}}(b_{q}\cdot c_{i})\right)\equiv\partial_{w_{M}w_{J}^{-1}}(b_{r}\cdot c_{i})\pmod{n}

for each ii. The subset of RJR^{J} consisting of those cc for which wMwJ1(brc)0(modn)\partial_{w_{M}w_{J}^{-1}}(b_{r}\cdot c)\equiv 0\pmod{n} is evidently an RMR^{M}-submodule. Since this submodule contains a basis {ci}\{c_{i}\} for RJR^{J} over RMR^{M}, it must contain all of RJR^{J}. Thus

nwMwJ1(brg)n\mid\partial_{w_{M}w_{J}^{-1}}(b_{r}\cdot g)

for all gRJg\in R^{J}. By 4.31 we deduce nbrn\mid b_{r}, leading to a contradiction. ∎

Corollary 4.33.

For an almost SW-realization we have DL<𝚊=End<𝚊(B𝚊)\operatorname{DL}_{<{\mathtt{a}}}=\operatorname{End}_{<{\mathtt{a}}}(B_{\mathtt{a}}). In particular, DL<𝚊\operatorname{DL}_{<{\mathtt{a}}} is an (RI,RJ)(R^{I},R^{J})-bimodule.

Proof.

The containment DL<𝚊End𝚊(B𝚊)\operatorname{DL}_{<{\mathtt{a}}}\subset\operatorname{End}_{{\mathtt{a}}}(B_{{\mathtt{a}}}) was already shown in 4.14. Now pick an arbitrary element ψEnd<𝚊(B𝚊)\psi\in\operatorname{End}_{<{\mathtt{a}}}(B_{\mathtt{a}}). By 4.30, there is some n𝕜n\in\Bbbk such that nψDL<𝚊n\psi\in\operatorname{DL}_{<{\mathtt{a}}}. Then ψ=1n(nψ)End<𝚊(B𝚊)\psi=\frac{1}{n}(n\psi)\in\operatorname{End}_{<{\mathtt{a}}}(B_{\mathtt{a}}), so by 4.32 we deduce that ψDL<𝚊\psi\in\operatorname{DL}_{<{\mathtt{a}}}. ∎

5. Polynomial forcing and atomic Leibniz

5.1. Polynomial forcing for atomic cosets

Now we explain the concept of polynomial forcing. We consider first the case of an atomic coset 𝚊[[IMJ]]{\mathtt{a}}\leftrightharpoons[[I\subset M\supset J]].

Recall that 𝚊¯J=I𝚊¯\underline{{\mathtt{a}}}J=I\underline{{\mathtt{a}}} and LR(𝚊)=I\operatorname{LR}({\mathtt{a}})=I since 𝚊{\mathtt{a}} is core. Thus there is an isomorphism

(62) RJRI,f𝚊¯(f).R^{J}\to R^{I},\qquad f\mapsto\underline{{\mathtt{a}}}(f).
Definition 5.1.

Let 𝚊{\mathtt{a}} be an atomic coset with reduced expression [[IMJ]][[I\subset M\supset J]] and let fRJf\in R^{J}. We say that polynomial forcing holds for ff and 𝚊{\mathtt{a}} if we have

(63) 1f𝚊¯(f)1End<𝚊(B𝚊).1\otimes f-\underline{{\mathtt{a}}}(f)\otimes 1\in\operatorname{End}_{<{\mathtt{a}}}(B_{\mathtt{a}}).

We say that polynomial forcing holds for 𝚊{\mathtt{a}} if (63) holds for all fRJf\in R^{J}.

Lemma 5.2.

Suppose that (63) holds for f1f_{1} and for f2f_{2}, with both f1,f2RJf_{1},f_{2}\in R^{J}. Then it holds for f1+f2f_{1}+f_{2} and f1f2f_{1}f_{2}.

Proof.

Additivity is trivial, because Hom<𝚊\operatorname{Hom}_{<{\mathtt{a}}} is closed under addition. Now consider the following:

(64) 1f1f2𝚊¯(f1f2)1=(1f1𝚊¯(f1)1)f2+𝚊¯(f1)(1f2𝚊¯(f2)1).1\otimes f_{1}\cdot f_{2}-\underline{{\mathtt{a}}}(f_{1}\cdot f_{2})\otimes 1=(1\otimes f_{1}-\underline{{\mathtt{a}}}(f_{1})\otimes 1)\cdot f_{2}+\underline{{\mathtt{a}}}(f_{1})\cdot(1\otimes f_{2}-\underline{{\mathtt{a}}}(f_{2})\otimes 1).

Since End<𝚊(B𝚊)\operatorname{End}_{<{\mathtt{a}}}(B_{\mathtt{a}}) is closed under right and left multiplication, both terms on the right-hand side above are in End<𝚊(B𝚊)\operatorname{End}_{<{\mathtt{a}}}(B_{\mathtt{a}}), and the result is proven. ∎

Before continuing, let us contrast polynomial forcing with an a priori different notion.

Definition 5.3.

Consider DL<𝚊\operatorname{DL}_{<{\mathtt{a}}} from Definition 4.13. We say that DL\operatorname{DL}-forcing holds for 𝚊{\mathtt{a}} and ff if 1f𝚊¯(f)1DL<𝚊End(B𝚊)1\otimes f-\underline{{\mathtt{a}}}(f)\otimes 1\in\operatorname{DL}_{<{\mathtt{a}}}\subset\operatorname{End}(B_{{\mathtt{a}}}). We say that DL\operatorname{DL}-forcing holds for 𝚊{\mathtt{a}} if it holds for 𝚊{\mathtt{a}} and ff, for all fRJ.f\in R^{J}.

For an almost-SW realization, DL\operatorname{DL}-forcing is equivalent to polynomial forcing, since DL<𝚊=End<𝚊(B𝚊)\operatorname{DL}_{<{\mathtt{a}}}=\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}}}) by Corollary 4.33. In general, it is not obvious that DL\operatorname{DL}-forcing is multiplicative. The proof of multiplicativity in Lemma 5.2 relied on the fact that End<𝚊(B𝚊)\operatorname{End}_{<{\mathtt{a}}}(B_{{\mathtt{a}}}) is an (RI,RJ)(R^{I},R^{J})-bimodule, whereas DL<𝚊\operatorname{DL}_{<{\mathtt{a}}} is only a priori a right RJR^{J}-module.

5.2. Equivalence

Now we prove the equivalence between atomic Leibniz rules and polynomial forcing. To formulate an intermediate condition in the proof, which is also of importance for the next section, we agree to say the following. Given an atomic (I,J)(I,J)-coset 𝚊{\mathtt{a}} and an element fRJf\in R^{J}, an atomic Leibniz rule for 𝚊{\mathtt{a}} and ff is said to hold if there exist elements Tq(f)T_{q}(f) such that equation (30) is satisfied for all gRJg\in R^{J}. Since this condition is stated for one polynomial ff at a time, there is no requirement that TqT_{q} is an RMR^{M}-linear operator.

Proposition 5.4.

Let 𝚊[I,M,J]{\mathtt{a}}\leftrightharpoons[I,M,J] be an atomic (I,J)(I,J)-coset, and fRJf\in R^{J}. We have a rightward atomic Leibniz rule for 𝚊{\mathtt{a}} and ff if and only if DL\operatorname{DL}-forcing holds for 𝚊{\mathtt{a}} and ff. Moreover, for an almost-SW realization, if DL\operatorname{DL}-forcing holds for 𝚊{\mathtt{a}} and ff, then the atomic Leibniz rule is unique, i.e., the elements Tq(f)RRR(q)T_{q}(f)\in R^{\operatorname{RR}(q)} in (30) are uniquely determined.

Proof.

Note that RJRMR^{J}\subset R^{M} is a Frobenius extension, see [EMTW20, Section 24.3.2]. The trace map is

MJ:=wMwJ1=𝚊.\partial^{J}_{M}:=\partial_{w_{M}w_{J}^{-1}}=\partial_{\mathtt{a}}.

Let ΔMJRJRMRJ\Delta^{J}_{M}\in R^{J}\otimes_{R^{M}}R^{J} denote the coproduct element (the image of 1RJ1\in R^{J} under the coproduct map), which we often denote using Sweedler notation. Then [ESW17, (2.2) with f=1f=1] implies that

(65) 11=𝚊(ΔM,(1)J)ΔM,(2)J.1\otimes 1=\partial_{\mathtt{a}}(\Delta^{J}_{M,(1)})\otimes\Delta^{J}_{M,(2)}.

Multiplying 𝚊¯(f)\underline{{\mathtt{a}}}(f) on the left we get

𝚊¯(f)1=𝚊¯(f)𝚊(ΔM,(1)J)ΔM,(2)J.\underline{{\mathtt{a}}}(f)\otimes 1=\underline{{\mathtt{a}}}(f)\cdot\partial_{\mathtt{a}}(\Delta^{J}_{M,(1)})\otimes\Delta^{J}_{M,(2)}.

Meanwhile, [ESW17, (2.2)] implies that

(66) 1f=𝚊(fΔM,(1)J)ΔM,(2)J.1\otimes f=\partial_{\mathtt{a}}(f\cdot\Delta^{J}_{M,(1)})\otimes\Delta^{J}_{M,(2)}.

Thus we have

(67) 1f𝚊¯(f)1=[𝚊(fΔM,(1)J)𝚊¯(f)𝚊(ΔM,(1)J)]ΔM,(2)J.1\otimes f-\underline{{\mathtt{a}}}(f)\otimes 1=\left[\partial_{\mathtt{a}}(f\cdot\Delta^{J}_{M,(1)})-\underline{{\mathtt{a}}}(f)\cdot\partial_{\mathtt{a}}(\Delta^{J}_{M,(1)})\right]\otimes\Delta^{J}_{M,(2)}.

Letting g=ΔM,(1)Jg=\Delta^{J}_{M,(1)}, a rightward atomic Leibniz rule for 𝚊{\mathtt{a}} and ff gives

(68) 𝚊(fg)𝚊¯(f)𝚊(g)=q<𝚊q¯wJ1(Tq(f)g).\partial_{{\mathtt{a}}}(f\cdot g)-\underline{{\mathtt{a}}}(f)\cdot\partial_{{\mathtt{a}}}(g)=\sum_{q<{\mathtt{a}}}\partial_{\overline{q}w_{J}^{-1}}(T_{q}(f)\cdot g).

Thus we have

(69) 1f𝚊¯(f)1=q<𝚊q¯wJ1(Tq(f)ΔM,(1)J)ΔM,(2)J.1\otimes f-\underline{{\mathtt{a}}}(f)\otimes 1=\sum_{q<{\mathtt{a}}}\partial_{\overline{q}w_{J}^{-1}}\left(T_{q}(f)\cdot\Delta^{J}_{M,(1)}\right)\otimes\Delta^{J}_{M,(2)}.

which lies in DL<𝚊\operatorname{DL}_{<{\mathtt{a}}} by definition.

We prove now the other direction. We have

1f𝚊¯(f)1DL<𝚊.1\otimes f-\underline{{\mathtt{a}}}(f)\otimes 1\in\operatorname{DL}_{<{\mathtt{a}}}.

By (67), we obtain

[𝚊(fΔM,(1)J)𝚊¯(f)𝚊(ΔM,(1)J)]ΔM,(2)JEnd<𝚊(B𝚊).\left[\partial_{\mathtt{a}}(f\cdot\Delta^{J}_{M,(1)})-\underline{{\mathtt{a}}}(f)\cdot\partial_{\mathtt{a}}(\Delta^{J}_{M,(1)})\right]\otimes\Delta^{J}_{M,(2)}\in\operatorname{End}_{<{\mathtt{a}}}(B_{\mathtt{a}}).

By definition of DL<𝚊\operatorname{DL}_{<{\mathtt{a}}} we deduce that

𝚊(fΔM,(1)J)\displaystyle\partial_{\mathtt{a}}(f\cdot\Delta^{J}_{M,(1)}) ΔM,(2)J=\displaystyle\otimes\Delta^{J}_{M,(2)}=
(70) 𝚊¯(f)𝚊(ΔM,(1)J)ΔM,(2)J+q<𝚊q¯wJ1(Tq(f)ΔM,(1)J)ΔM,(2)J\displaystyle\underline{{\mathtt{a}}}(f)\cdot\partial_{\mathtt{a}}(\Delta^{J}_{M,(1)})\otimes\Delta^{J}_{M,(2)}+\sum_{q<{\mathtt{a}}}\partial_{\overline{q}w_{J}^{-1}}(T_{q}(f)\cdot\Delta_{M,(1)}^{J})\otimes\Delta_{M,(2)}^{J}

for some Tq(f)RRR(q)T_{q}(f)\in R^{\operatorname{RR}(q)}. For an almost-SW realization, Lemma 4.29 implies that the Tq(f)T_{q}(f) are unique.

Note that ΔM,(1)J\Delta^{J}_{M,(1)} and ΔM,(2)J\Delta^{J}_{M,(2)} run over dual bases of RJR^{J} over RMR^{M}. The elements 𝔹={1ΔM,(2)J}\mathbb{B}=\{1\otimes\Delta^{J}_{M,(2)}\} form a basis for B𝚊B_{\mathtt{a}}, when viewed as a left RIR^{I}-module. Thus in order for the equation (70) to hold, it must be an equality for each coefficient with respect to the basis 𝔹\mathbb{B}. Hence we conclude

(71) 𝚊(fΔ)=𝚊¯(f)𝚊(Δ)+q<𝚊q(Tq(f)Δ)\partial_{\mathtt{a}}(f\cdot\Delta)=\underline{{\mathtt{a}}}(f)\cdot\partial_{\mathtt{a}}(\Delta)+\sum_{q<{\mathtt{a}}}\partial_{q}(T_{q}(f)\cdot\Delta)

for all Δ\Delta ranging through a basis of RJR^{J} over RMR^{M}.

Using the linearity of (71) over RMR^{M}, we deduce that it continues to hold when Δ\Delta is replaced by any element gRJg\in R^{J}. Thus the atomic Leibniz rule for ff is proven. ∎

Theorem 5.5.

Assume an almost-SW realization (see Definition 4.21). Let 𝚊[I,M,J]{\mathtt{a}}\leftrightharpoons[I,M,J] be an atomic (I,J)(I,J)-coset. Then the following are equivalent.

  1. (1)

    A rightward atomic Leibniz rule holds for 𝚊{\mathtt{a}}.

  2. (2)

    A leftward atomic Leibniz rule holds for 𝚊{\mathtt{a}}.

  3. (3)

    For a set of generators {ci}\{c_{i}\} of the RMR^{M}-algebra RJR^{J}, a rightward atomic Leibniz rule holds for 𝚊{\mathtt{a}} and each cic_{i}.

  4. (4)

    For a set of generators {ci}\{c_{i}\} of the RMR^{M}-algebra RJR^{J}, a leftward atomic Leibniz rule holds for 𝚊{\mathtt{a}} and each cic_{i}.

  5. (5)

    Polynomial forcing holds for 𝚊{\mathtt{a}}.

Moreover, in this case, there are unique operators Tq,TqT_{q},T^{\prime}_{q} that satisfy atomic Leibniz rules.

Proof.

First, we observe that polynomial forcing holds for fRMf\in R^{M}. Clearly 1f=f11\otimes f=f\otimes 1. Moreover, 𝚊WM{\mathtt{a}}\subset W_{M} so 𝚊¯(f)=f\underline{{\mathtt{a}}}(f)=f.

That (1) implies (3) is clear.

Suppose that (3) holds. By Proposition 5.4, DL\operatorname{DL} forcing holds for all cic_{i}. By Corollary 4.33, DL\operatorname{DL}-forcing for cic_{i} is equivalent to polynomial forcing for cic_{i}. By Lemma 5.2, the subset of RJR^{J} consisting of those ff for which polynomial forcing holds is a subring. As explained above, this subring includes RMR^{M}, so if it includes {ci}\{c_{i}\} then it must be all of RJR^{J}. In this way, (3) implies (5).

Suppose that (5) holds. Once again, Proposition 5.4 and Corollary 4.33 imply that, for each fRJf\in R^{J}, a rightward atomic Leibniz rule holds for ff, with the elements Tq(f)RRR(q)T_{q}(f)\in R^{\operatorname{RR}(q)} being unique. To prove that a rightward atomic Leibniz rule holds, it remains to prove that the operators Tq:RJRRR(q)T_{q}\colon R^{J}\to R^{\operatorname{RR}(q)} are RMR^{M}-linear. We do this below, finishing the proof that (5) implies (1).

Let gRMg\in R^{M}. Multiplying both sides of equation (70) on the left by gg, and pulling gg into various RMR^{M}-linear operators (namely 𝚊\partial_{{\mathtt{a}}} and 𝚊¯\underline{{\mathtt{a}}} and q¯wJ1\partial_{\overline{q}w_{J}^{-1}}), we obtain

𝚊(gfΔM,(1)J)\displaystyle\partial_{\mathtt{a}}(gf\cdot\Delta^{J}_{M,(1)}) ΔM,(2)J=\displaystyle\otimes\Delta^{J}_{M,(2)}=
𝚊¯(gf)𝚊(ΔM,(1)J)ΔM,(2)J+q<𝚊q¯wJ1(gTq(f)ΔM,(1)J)ΔM,(2)J\displaystyle\underline{{\mathtt{a}}}(gf)\cdot\partial_{\mathtt{a}}(\Delta^{J}_{M,(1)})\otimes\Delta^{J}_{M,(2)}+\sum_{q<{\mathtt{a}}}\partial_{\overline{q}w_{J}^{-1}}(gT_{q}(f)\cdot\Delta_{M,(1)}^{J})\otimes\Delta_{M,(2)}^{J}

This is exactly (70) with gfgf replacing ff, except that gTq(f)gT_{q}(f) appears instead of Tq(gf)T_{q}(gf). By uniqueness, we deduce that Tq(gf)=gTq(f)T_{q}(gf)=gT_{q}(f).

We have thus shown the equivalence of (1),(3) and (5). A similar argument will imply the equivalence of (2) and (4) and (5), and the uniqueness of TqT^{\prime}_{q}. This similar argument replaces 𝔻𝕃𝕃r(q,b)\mathbb{DLL}_{r}(q,b) with 𝔻𝕃𝕃l(q,b)\mathbb{DLL}_{l}(q,b), using Proposition 4.20 and Lemma 4.11. The left analogue of the remaining arguments (e.g. Proposition 5.4 and Corollary 4.33) is left to the reader. ∎

Remark 5.6.

The intermediate conditions (3) and (4) do not play a significant role in the proof. We have included them to make it easier to prove the atomic Leibniz rule by establishing it on a set of generators.

5.3. Polynomial forcing for general cosets

Now let pp be an arbitrary (I,J)(I,J)-coset, with a reduced expression II_{\bullet}. We wish to avoid the technicalities of changing the reduced expression II_{\bullet} in this paper. Instead we focus on the special case when II_{\bullet} is an atomic-factored reduced expression, i.e. it has the following form:

(72) I=[[ILR(p)]]I[[RR(p)J]]I_{\bullet}=[[I\supset\operatorname{LR}(p)]]\circ I^{\prime}_{\bullet}\circ[[\operatorname{RR}(p)\subset J]]

where II^{\prime}_{\bullet} is an atomic reduced expression (see below) for pcorep^{\operatorname{core}}.

An atomic reduced expression for a core coset pcorep^{\operatorname{core}} is a reduced expression of the form

(73) I=[[LR(p)=N0M1N1MmNm=RR(p)]],I^{\prime}_{\bullet}=[[\operatorname{LR}(p)=N_{0}\subset M_{1}\supset N_{1}\subset\cdots\subset M_{m}\supset N_{m}=\operatorname{RR}(p)]],

where each [[NiMi+1Ni+1]][[N_{i}\subset M_{i+1}\supset N_{i+1}]] is a reduced expression for an atomic coset 𝚊i+1{\mathtt{a}}_{i+1}. In particular, p=𝚊1.𝚊2..𝚊mp={\mathtt{a}}_{1}.{\mathtt{a}}_{2}.\cdots.{\mathtt{a}}_{m}. Any core coset has an atomic reduced expression, see [EKLP24b, Cor. 2.17] and thus any coset has an atomic-factored reduced expression by [EK21, Proposition 4.28].

We have

(74) BS(I)=RLR(p)RM1RN1RM2RMmRRR(p)\operatorname{BS}(I_{\bullet})=R^{\operatorname{LR}(p)}\otimes_{R^{M_{1}}}R^{N_{1}}\otimes_{R^{M_{2}}}\cdots\otimes_{R^{M_{m}}}R^{\operatorname{RR}(p)}

viewed as an (RI,RJ)(R^{I},R^{J})-bimodule. Meanwhile, BS(I)\operatorname{BS}(I^{\prime}_{\bullet}) is the same abelian group, but is viewed as an (RLR(p),RRR(p))(R^{\operatorname{LR}(p)},R^{\operatorname{RR}(p)})-bimodule. There is an action of each RNiR^{N_{i}} on BS(I)\operatorname{BS}(I_{\bullet}) by multiplication in the ii-th tensor factor of (74). Indeed, this induces an injective map

(75) BS(I)End(BS(I))\operatorname{BS}(I_{\bullet})\to\operatorname{End}(\operatorname{BS}(I_{\bullet}))

which is not surjective in general.

An arbitrary reduced expression for pp might never factor through the subset LR(p)\operatorname{LR}(p) or RR(p)\operatorname{RR}(p). The first advantage of an atomic-factored expression is that there is an obvious action of RLR(p)R^{\operatorname{LR}(p)} on BS(I)\operatorname{BS}(I_{\bullet}) by left-multiplication, and an obvious action of RRR(p)R^{\operatorname{RR}(p)} by right-multiplication. The goal is to prove that these two actions agree up to a twist by p¯\underline{p}, modulo lower terms.

We denote by IdI\operatorname{Id}_{I_{\bullet}} the identity morphism of BS(I)\operatorname{BS}(I_{\bullet}).

Definition 5.7.

Let II_{\bullet} be an atomic-factored reduced expression as in (72). We say that polynomial forcing holds for II_{\bullet} if for all fRRR(p)f\in R^{\operatorname{RR}(p)}, within End(BS(I))\operatorname{End}(\operatorname{BS}(I_{\bullet})) as described in (74) and (75), we have

(76) p¯(f)IdIIdIf modulo End<p(BS(I)).\underline{p}(f)\cdot\operatorname{Id}_{I_{\bullet}}\equiv\operatorname{Id}_{I_{\bullet}}\cdot f\text{ modulo }\operatorname{End}_{<p}(\operatorname{BS}(I_{\bullet})).

We say that polynomial forcing holds for a double coset pp if it holds for all atomic-factored reduced expressions II_{\bullet} satisfying IpI_{\bullet}\leftrightharpoons p.

This definition generalizes Definition 5.1 because atomic cosets have only one reduced expression.

Let pp be an arbitrary (I,J)(I,J)-coset. Since LR(p)I\operatorname{LR}(p)\subset I, there is an inclusion of rings RIRLR(p)R^{I}\subset R^{\operatorname{LR}(p)}, and RLR(p)R^{\operatorname{LR}(p)} is naturally an RIR^{I}-module. Similarly, RRR(p)R^{\operatorname{RR}(p)} is an RJR^{J}-module. If fRRR(p)f\in R^{\operatorname{RR}(p)} then p¯(f)RLR(p)\underline{p}(f)\in R^{\operatorname{LR}(p)}. We can identify the rings RLR(p)R^{\operatorname{LR}(p)} and RRR(p)R^{\operatorname{RR}(p)} via p¯\underline{p}. In this way, RLR(p)R^{\operatorname{LR}(p)} becomes an (RI,RJ)(R^{I},R^{J})-bimodule.

Definition 5.8.

Let pp be an (I,J)(I,J)-coset. The standard bimodule associated to pp, denoted RpR_{p}, is RLR(p)R^{\operatorname{LR}(p)} as a left RIR^{I}-module. If fRJf\in R^{J} and mRpm\in R_{p} then

(77) mf:=p¯(f)m.m\cdot f:=\underline{p}(f)m.

We identify RpR_{p} with either RLR(p)R^{\operatorname{LR}(p)} (with right action twisted) or RRR(p)R^{\operatorname{RR}(p)} (with left action twisted), as is more convenient.

Let qu:End(BS(I))End(BS(I))/End<p(BS(I))\operatorname{qu}:\operatorname{End}(\operatorname{BS}(I_{\bullet}))\to\operatorname{End}(\operatorname{BS}(I_{\bullet}))/\operatorname{End}_{<p}(\operatorname{BS}(I_{\bullet})) denote the quotient map.

Lemma 5.9.

Let pp be a core (I,J)(I,J)-coset and let II_{\bullet} be a reduced expression for pp. The bimodule map

(78) RpEnd(BS(I))/End<p(BS(I)),1qu(IdI)R_{p}\to\operatorname{End}(\operatorname{BS}(I_{\bullet}))/\operatorname{End}_{<p}(\operatorname{BS}(I_{\bullet})),\qquad 1\mapsto\operatorname{qu}(\operatorname{Id}_{I_{\bullet}})

is well-defined if and only if polynomial forcing holds for II_{\bullet}.

Proof.

The right action of fRJf\in R^{J} on 1RJ=Rp1\in R^{J}=R_{p} yields fRJf\in R^{J}, and the right action on qu(IdI)\operatorname{qu}(\operatorname{Id}_{I_{\bullet}}) yields qu(IdIf)\operatorname{qu}(\operatorname{Id}_{I_{\bullet}}\cdot f). The left action of p¯(f)RI\underline{p}(f)\in R^{I} on 1RJ=Rp1\in R^{J}=R_{p} yields fRJf\in R^{J}, and the left action on qu(IdI)\operatorname{qu}(\operatorname{Id}_{I_{\bullet}}) yields qu(p¯(f)IdI)\operatorname{qu}(\underline{p}(f)\cdot\operatorname{Id}_{I_{\bullet}}). These agree if and only if the bimodule map is well-defined, and if and only if (76) holds. ∎

In conclusion, we have shown the equivalence of three ideas (for almost SW-realizations) for an atomic coset 𝚊{\mathtt{a}}: the well-definedness of the morphism (78) when p=𝚊Ip={\mathtt{a}}\leftrightharpoons I_{\bullet}, atomic polynomial forcing, and the atomic Leibniz rule.

For SW-realizations, (78) is an isomorphism by the theory of singular Soergel bimodules. We can thus prove one of our main theorems.

Theorem 5.10.

For an SW-realization, atomic polynomial forcing and atomic Leibniz hold. Moreover, the operators Tq,TqT_{q},T_{q}^{\prime} in the Leibniz formulas are unique.

Proof.

Assume 𝚊[I,M,J]{\mathtt{a}}\leftrightharpoons[I,M,J] is an atomic coset. Then B𝚊BS([I,M,J])B_{\mathtt{a}}\cong\operatorname{BS}([I,M,J]).

Recall from [Wil11, §4.5] the definition of the submodule Γ<𝚊B𝚊\Gamma_{<{\mathtt{a}}}B_{\mathtt{a}} of elements supported on lower cosets. By [EKLP23b, Lemma 3.31] we have a short exact sequence

(79) 0End(B𝚊,Γ<𝚊B𝚊)End(B𝚊)Hom(B𝚊,B𝚊/Γ<𝚊B𝚊)00\to\operatorname{End}(B_{\mathtt{a}},\Gamma_{<{\mathtt{a}}}B_{\mathtt{a}})\to\operatorname{End}(B_{\mathtt{a}})\to\operatorname{Hom}(B_{\mathtt{a}},B_{\mathtt{a}}/\Gamma_{<{\mathtt{a}}}B_{\mathtt{a}})\to 0

and, by [EKLP23b, Theorem 3.30], the first term in (79) is isomorphic to End<𝚊(B𝚊)\operatorname{End}_{<{\mathtt{a}}}(B_{\mathtt{a}}). Moreover, since B𝚊B_{\mathtt{a}} is indecomposable, by [Wil11, Theorem 7.10] we have B𝚊/Γ<𝚊B𝚊R𝚊B_{\mathtt{a}}/\Gamma_{<{\mathtt{a}}}B_{{\mathtt{a}}}\cong R_{\mathtt{a}}. The Soergel–Williamson hom formula [Wil11, Theorem 7.9] implies that we have an isomorphism Hom(B𝚊,R𝚊)R𝚊\operatorname{Hom}(B_{\mathtt{a}},R_{\mathtt{a}})\cong R_{\mathtt{a}}777As in the rest of this paper, we are ignoring degrees here. given by ff(11)f\mapsto f(1\otimes 1). Putting all together, we obtain an isomorphism

(80) End(B𝚊)/End<𝚊(B𝚊)R𝚊\operatorname{End}(B_{\mathtt{a}})/\operatorname{End}_{<{\mathtt{a}}}(B_{\mathtt{a}})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}R_{\mathtt{a}}

which sends id𝚊\operatorname{id}_{\mathtt{a}} to 1R𝚊1\in R_{\mathtt{a}}

By Lemma 5.9, the existence of the isomorphism implies polynomial forcing for 𝚊{\mathtt{a}}. By Theorem 5.5, this is in turn equivalent to the atomic Leibniz rule for 𝚊{\mathtt{a}}. Moreover, as proven in 5.5, the operators Tq,TqT_{q},T_{q}^{\prime} in the Leibniz formulas are unique. ∎

Remark 5.11.

For a SW-realization, there is an equivalent module-theoretic (rather than morphism-theoretic) version of polynomial forcing. We first recall from [EKLP24b, Definition 3.23] the filtration on Soergel bimodules

N<p(B)=fHom(BS(I),B),Iq<pIm(f).N_{<p}(B)=\sum_{f\in\operatorname{Hom}(\operatorname{BS}(I_{\bullet}),B),I_{\bullet}\leftrightharpoons q<p}\operatorname{Im}(f).

In [EKLP24b, Proposition 3.25] we have showed that this coincides with the support filtration Γ<p\Gamma_{<p} introduced in [Wil11].

Let 𝚊[I,M,J]{\mathtt{a}}\leftrightharpoons[I,M,J] be an atomic coset. We say that (module-theoretic) polynomial forcing holds for 𝚊{\mathtt{a}} and ff if

(81) 𝚊¯(f)11fN<𝚊(B𝚊).\underline{{\mathtt{a}}}(f)\otimes 1-1\otimes f\in N_{<{\mathtt{a}}}(B_{\mathtt{a}}).

There is an isomorphism B𝚊End(B𝚊)B_{\mathtt{a}}\cong\operatorname{End}(B_{\mathtt{a}}), where bbB𝚊b\otimes b^{\prime}\in B_{\mathtt{a}} is sent to multiplication by bbb\otimes b^{\prime}. Moreover, by [EKLP24b, Theorem 3.30] we have

Hom(B𝚊,N<𝚊B𝚊)End<𝚊(B𝚊).\operatorname{Hom}(B_{\mathtt{a}},N_{<{\mathtt{a}}}B_{\mathtt{a}})\cong\operatorname{End}_{<{\mathtt{a}}}(B_{\mathtt{a}}).

Hence, (81) holds if and only if multiplication by 𝚊¯(f)11f\underline{{\mathtt{a}}}(f)\otimes 1-1\otimes f induces a morphism in End<𝚊(B𝚊)\operatorname{End}_{<{\mathtt{a}}}(B_{\mathtt{a}}), that is, if an only if (morphism-theoretic) polynomial forcing holds for ff.

5.4. Polynomial forcing: atomic and general

In the diagrammatic category, we intend to use the atomic Leibniz rule to prove polynomial forcing, and not vice versa. In that context, polynomial forcing is to be interpreted as the morphism-theoretic statement that (78) is a well-defined morphism, when II_{\bullet} is an atomic-factored reduced expression. The goal of this section is to prove that atomic polynomial forcing implies general polynomial forcing.

In [EKLP23b], a compatibility between the Bruhat order and concatenation of reduced expression is proven, which implies the following result.

Proposition 5.12 ([EKLP23b, Prop. 3.7]).

Let PpP_{\bullet}\leftrightharpoons p and QqQ_{\bullet}\leftrightharpoons q and RrR_{\bullet}\leftrightharpoons r be reduced expressions such that PQRp.q.rP_{\bullet}\circ Q_{\bullet}\circ R_{\bullet}\leftrightharpoons p.q.r is reduced. Then

(82) idPEnd<q(Q)idREnd<p.q.r(PQR).\operatorname{id}_{P_{\bullet}}\otimes\operatorname{End}_{<q}(Q_{\bullet})\otimes\operatorname{id}_{R_{\bullet}}\subset\operatorname{End}_{<p.q.r}(P_{\bullet}\circ Q_{\bullet}\circ R_{\bullet}).
Theorem 5.13.

Assume an SW-realization. Then polynomial forcing holds for all double cosets.

Proof.

We first treat the case where p=pcorep=p^{\operatorname{core}} is a core (I,J)(I,J)-coset. Consider an atomic reduced expression II_{\bullet} for pp, yielding atomic cosets 𝚊i{\mathtt{a}}_{i} such that p=𝚊1.𝚊2..𝚊mp={\mathtt{a}}_{1}.{\mathtt{a}}_{2}.\cdots.{\mathtt{a}}_{m}. Since 𝚊i{\mathtt{a}}_{i} are core cosets, by [EKLP24b, Lem. 2.10] we have p¯=𝚊1¯𝚊2¯𝚊m¯\underline{p}=\underline{{\mathtt{a}}_{1}}\cdot\underline{{\mathtt{a}}_{2}}\cdots\underline{{\mathtt{a}}_{m}}. Now within BS(I)\operatorname{BS}(I_{\bullet}) we have

(83) Id𝚊1Id𝚊mfId𝚊1𝚊m¯(f)Id𝚊m𝚊1¯((𝚊m¯(f)))Id𝚊1Id𝚊m,\operatorname{Id}_{{\mathtt{a}}_{1}}\otimes\cdots\otimes\operatorname{Id}_{{\mathtt{a}}_{m}}f\equiv\operatorname{Id}_{{\mathtt{a}}_{1}}\otimes\cdots\otimes\underline{{\mathtt{a}}_{m}}(f)\operatorname{Id}_{{\mathtt{a}}_{m}}\equiv\ldots\equiv\underline{{\mathtt{a}}_{1}}(\cdots(\underline{{\mathtt{a}}_{m}}(f)))\operatorname{Id}_{{\mathtt{a}}_{1}}\otimes\cdots\otimes\operatorname{Id}_{{\mathtt{a}}_{m}},

where \equiv indicates equality modulo lower terms. At each step we applied polynomial forcing for an atomic coset as proved in 5.10, and used Proposition 5.12 to argue that lower terms for 𝚊i{\mathtt{a}}_{i} embed into lower terms for pp. Thus polynomial forcing holds for pp.

If pp is not a core coset, let II_{\bullet} be a special reduced expression for pp as in (72). Polynomial forcing for pcorep^{\operatorname{core}} implies that

p¯(f)IdIIdIf\underline{p}(f)\cdot\operatorname{Id}_{I_{\bullet}}\equiv\operatorname{Id}_{I_{\bullet}}\cdot f

modulo

id[[ILR(p)]]End<pcore(I)id[[RR(p)J]].\operatorname{id}_{[[I\supset\operatorname{LR}(p)]]}\otimes\operatorname{End}_{<p^{\operatorname{core}}}(I^{\prime}_{\bullet})\otimes\operatorname{id}_{[[\operatorname{RR}(p)\subset J]]}.

By Proposition 5.12, they are also equivalent modulo End<p(I)\operatorname{End}_{<p}(I_{\bullet}), as desired. ∎

The reader familiar with Soergel bimodules might be familiar with the following example, which showcases how atomic polynomial forcing implies the general case. It also relates our new concept of polynomial forcing to the concept previously in the literature.

Example 5.14.

Consider the (,)(\emptyset,\emptyset)-coset p={s}p=\{s\} for a simple reflection ss, and the reduced expression I=[,s,]I_{\bullet}=[\emptyset,s,\emptyset]. Inside Bs:=BS(I)=RRsRB_{s}:=\operatorname{BS}(I_{\bullet})=R\otimes_{R^{s}}R we have

(84) 1fs(f)1=s(f)12(αs1+1αs).1\otimes f-s(f)\otimes 1=\partial_{s}(f)\cdot\frac{1}{2}(\alpha_{s}\otimes 1+1\otimes\alpha_{s}).

The term on the right hand side is in End<s(I)\operatorname{End}_{<s}(I_{\bullet}), a consequence of the so-called polynomial forcing relation in the Hecke category, see e.g. [EW16, (5.2)].

Now consider the (,)(\emptyset,\emptyset) coset p={w}p=\{w\} for some wWw\in W, and a reduced expression I=[,s1,,s2,,,sd,]I_{\bullet}=[\emptyset,s_{1},\emptyset,s_{2},\ldots,\emptyset,s_{d},\emptyset]. By applying the polynomial forcing relation for BsdB_{s_{d}}, we see that 11f1sd(f)11\otimes\cdots\otimes 1\otimes f\equiv 1\otimes\cdots\otimes s_{d}(f)\otimes 1 modulo maps which factor through [,s1,,,,sd1,][\emptyset,s_{1},\emptyset,\ldots,\emptyset,s_{d-1},\emptyset]. Continuing, we can apply polynomial forcing for each BsiB_{s_{i}} to force ff across all the tensors, at the cost of maps which factor through subexpressions of s1sds_{1}\cdots s_{d}. By the subexpression property of the Bruhat order, such maps consist of lower terms.

6. Atomic Leibniz rule in type AA

In this section we explicitly prove condition (3) in Theorem 5.5 in type AA. We establish this result for 𝕜=\Bbbk=\mathbb{Z} in this section, rather than over a field. In addition to extending our results over \mathbb{Z}, we feel the ability to be explicit in a key example is its own reward.

In order to achieve this, we first prove in 6.3 explicitly an atomic Leibniz rule for a specific set of generators when WMW_{M} is the entire symmetric group. In Section 6.3 we extend our results to the case when WMW_{M} is a product of symmetric groups. This handles all atomic cosets in type AA.

6.1. Notation in type AA

We fix notation under the assumption that WMW_{M} is an irreducible Coxeter group of type AA. For s,tMs,t\in M we write s^:=M{s}\widehat{s}:=M\setminus\{s\} and s^t^:=M{s,t}\widehat{s}\widehat{t}:=M\setminus\{s,t\}, etcetera.

Fix a,b1a,b\geq 1 and let n=a+bn=a+b. Let WM=𝐒nW_{M}=\mathbf{S}_{n}. Let t=sat=s_{a} and s=sb=wMtwMs=s_{b}=w_{M}tw_{M}, so that WJ=𝐒a×𝐒bW_{J}=\mathbf{S}_{a}\times\mathbf{S}_{b} and WI=𝐒b×𝐒aW_{I}=\mathbf{S}_{b}\times\mathbf{S}_{a}. Let 𝚊{\mathtt{a}} be the (𝐒b×𝐒a,𝐒a×𝐒b)(\mathbf{S}_{b}\times\mathbf{S}_{a},\mathbf{S}_{a}\times\mathbf{S}_{b})-coset containing wMw_{M}. The coset 𝚊{\mathtt{a}} is depicted as follows, with its minimal element 𝚊¯\underline{{\mathtt{a}}} being the string diagram visible.

(85) 𝚊=[Uncaptioned image].{\mathtt{a}}=\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 16.26074pt\hbox{{\hbox{\kern-16.26074pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-13.73131pt\hbox{$\textstyle{\includegraphics[scale={1}]{./EPS/cosetq0}}$}}}}}}}}}}.

Drawn is the example a=3a=3 and b=5b=5.

For each 0kmin(a,b)0\leq k\leq\min(a,b), there is an (s^,t^)(\widehat{s},\widehat{t})-coset qkq_{k} depicted as follows.

(86) q1=[Uncaptioned image],q2=[Uncaptioned image],q3=[Uncaptioned image].q_{1}=\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 16.26074pt\hbox{{\hbox{\kern-16.26074pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-13.73131pt\hbox{$\textstyle{\includegraphics[scale={1}]{./EPS/cosetq1}}$}}}}}}}}}},\qquad q_{2}=\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 16.26074pt\hbox{{\hbox{\kern-16.26074pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-13.73131pt\hbox{$\textstyle{\includegraphics[scale={1}]{./EPS/cosetq2}}$}}}}}}}}}},\qquad q_{3}=\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 16.26074pt\hbox{{\hbox{\kern-16.26074pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-13.73131pt\hbox{$\textstyle{\includegraphics[scale={1}]{./EPS/cosetq3}}$}}}}}}}}}}.

Then p=q0p=q_{0}, and {qk}0kmin(a,b)\{q_{k}\}_{0\leq k\leq\min(a,b)} is an enumeration of all the (s^,t^)(\widehat{s},\widehat{t})-cosets. The Bruhat order is a total order in this case:

q0>q1>q2>>qmin(a,b).q_{0}>q_{1}>q_{2}>\ldots>q_{\min(a,b)}.

The left redundancy subgroup (see [EKLP24b, Section 1.2] to see how to calculate redundancies and cores) of qkq_{k} is 𝐒k×𝐒bk×𝐒ak×𝐒k\mathbf{S}_{k}\times\mathbf{S}_{b-k}\times\mathbf{S}_{a-k}\times\mathbf{S}_{k}. For brevity, let :=nk\ell:=n-k. Then (with one exception) LR(qk)=b^k^^:=M{b,k,}\operatorname{LR}(q_{k})=\widehat{b}\widehat{k}\widehat{\ell}:=M\setminus\{b,k,\ell\} and RR(qk)=a^k^^\operatorname{RR}(q_{k})=\widehat{a}\widehat{k}\widehat{\ell}. The core of qkq_{k} is the double coset depicted as

(87) q1core=[Uncaptioned image],q2core=[Uncaptioned image],q3core=[Uncaptioned image].q_{1}^{\operatorname{core}}=\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 19.15155pt\hbox{{\hbox{\kern-19.15155pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-13.73131pt\hbox{$\textstyle{\includegraphics[scale={1}]{./EPS/coreq1}}$}}}}}}}}}},\qquad q_{2}^{\operatorname{core}}=\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 19.15155pt\hbox{{\hbox{\kern-19.15155pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-13.73131pt\hbox{$\textstyle{\includegraphics[scale={1}]{./EPS/coreq2}}$}}}}}}}}}},\qquad q_{3}^{\operatorname{core}}=\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 19.15155pt\hbox{{\hbox{\kern-19.15155pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-13.73131pt\hbox{$\textstyle{\includegraphics[scale={1}]{./EPS/coreq3}}$}}}}}}}}}}.

Note that the core of qkq_{k} is itself atomic, except for k=min(a,b)k=\min(a,b) when the core is an identity coset.

The case k=a=b=k=a=b=\ell is relatively special. We denote this special coset as qa=bq_{a=b}. We have LR(qa=b)=RR(qa=b)=I=J\operatorname{LR}(q_{a=b})=\operatorname{RR}(q_{a=b})=I=J. Unlike other qiq_{i}, qa=bq_{a=b} is core.

Remark 6.1.

In type AA the following statement is always true: if 𝚊{\mathtt{a}} is atomic and q<𝚊q<{\mathtt{a}} then qcoreq^{\operatorname{core}} is either atomic or an identity coset. We do not know for which atomic cosets 𝚊{\mathtt{a}} this property holds in other types.

A reduced expression for qkq_{k}, which factors through the core, is

(88) qk[[b^b^k^^k^^a^k^^a^]].q_{k}\leftrightharpoons[[\widehat{b}\supset\widehat{b}{\widehat{k}\widehat{\ell}}\subset\widehat{k}\widehat{\ell}\supset\widehat{a}\widehat{k}\widehat{\ell}\subset\widehat{a}]].

The exception is when a=b=ka=b=k, in which case

(89) qa=b[b^],q_{a=b}\leftrightharpoons[\widehat{b}],

that is, the identity expression of I=b^I=\widehat{b} is a reduced expression for the length zero coset qa=bq_{a=b}.

Finally, let us write yk=qk¯wJ1y_{k}=\overline{q_{k}}w_{J}^{-1}. Then qk=yk\partial_{q_{k}}=\partial_{y_{k}}. Here are examples of yky_{k}.

(90) y0=[Uncaptioned image],y1=[Uncaptioned image],y2=[Uncaptioned image],y3=[Uncaptioned image].y_{0}=\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 13.36995pt\hbox{{\hbox{\kern-13.36995pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-10.8405pt\hbox{$\textstyle{\includegraphics[scale={1}]{./EPS/y0}}$}}}}}}}}}},\qquad y_{1}=\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 13.36995pt\hbox{{\hbox{\kern-13.36995pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-10.8405pt\hbox{$\textstyle{\includegraphics[scale={1}]{./EPS/y1}}$}}}}}}}}}},\qquad y_{2}=\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 13.36995pt\hbox{{\hbox{\kern-13.36995pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-10.8405pt\hbox{$\textstyle{\includegraphics[scale={1}]{./EPS/y2}}$}}}}}}}}}},\qquad y_{3}=\vbox{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 13.36995pt\hbox{{\hbox{\kern-13.36995pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise-10.8405pt\hbox{$\textstyle{\includegraphics[scale={1}]{./EPS/y3}}$}}}}}}}}}}.

Of course y0=𝚊¯y_{0}=\underline{{\mathtt{a}}}. Meanwhile yky_{k} is obtained from y0y_{0} by removing a k×kk\times k square of crossings from the top. In the special case of the coset qa=bq_{a=b}, we have ya=b=ey_{a=b}=e, the identity of WW.

6.2. Complete symmetric polynomials

Fix a,b1a,b\geq 1 and continue to use the notation from the previous section. The standard action of 𝐒n\mathbf{S}_{n} on n\mathbb{Z}^{n} (with the standard choice of roots and coroots) we call the permutation realization. Let R=[x1,,xn]R=\mathbb{Z}[x_{1},\ldots,x_{n}]. All Demazure operators preserve RR. By a result of Demazure [Dem73, Lemme 5], Frobenius surjectivity holds in type AA over \mathbb{Z}, that is, for any IMI\subset M we can find PIRP_{I}\in R such that I(PI)=1\partial_{I}(P_{I})=1.

Moreover, for any JIJ\subset I the ring RIR^{I} is Frobenius over RJR^{J} and we can choose dual bases ΔI,(1)J\Delta^{J}_{I,(1)} and ΔI,(1)J\Delta^{J}_{I,(1)} accordingly.

It is well-known that the subring RJ=R𝐒a×𝐒bR^{J}=R^{\mathbf{S}_{a}\times\mathbf{S}_{b}} is generated over R𝐒nR^{\mathbf{S}_{n}} by the complete symmetric polynomials

hi(x1,,xa)=1k1k2kiaxk1xk2xkih_{i}(x_{1},\ldots,x_{a})=\sum_{1\leq k_{1}\leq k_{2}\leq\ldots\leq k_{i}\leq a}x_{k_{1}}x_{k_{2}}\cdots x_{k_{i}}

in the first aa variables. In this section, we directly prove the atomic Leibniz rule for 𝚊{\mathtt{a}} when f=hi(x1,,xa)f=h_{i}(x_{1},\ldots,x_{a}).

One of the great features of complete symmetric polynomials is their behavior under Demazure operators. For example, we have

(91) 3(hi(x1,x2,x3))=hi1(x1,x2,x3,x4).\partial_{3}(h_{i}(x_{1},x_{2},x_{3}))=h_{i-1}(x_{1},x_{2},x_{3},x_{4}).

As a consequence, 23(hi(x1,x2,x3))=0\partial_{2}\partial_{3}(h_{i}(x_{1},x_{2},x_{3}))=0, a fact which is false if hih_{i} is replaced by some general polynomial inside R𝐒3×𝐒n3R^{\mathbf{S}_{3}\times\mathbf{S}_{n-3}}. Indeed, the only elements ws1s2s3w\leq s_{1}s_{2}s_{3} for which w(hi(x1,x2,x3))0\partial_{w}(h_{i}(x_{1},x_{2},x_{3}))\neq 0 are w=s3w=s_{3} and w=ew=e. This will simplify the computation considerably.

Below we shall use letters like XX, YY, and ZZ to denote subsets of {1,,n}\{1,\ldots,n\}. We write hi(X)h_{i}(X) for the ii-th complete symmetric polynomial in the variables xjx_{j} for jXj\in X.

Lemma 6.2.

We have

(92) j(hi(X))={hi1(X{j+1})if jXand j+1Xhi1(X{j})if j+1Xand jX0otherwise.\partial_{j}(h_{i}(X))=\begin{cases}h_{i-1}(X\cup\{j+1\})&\text{if }j\in X\text{and }j+1\not\in X\\ -h_{i-1}(X\cup\{j\})&\text{if }j+1\in X\text{and }j\not\in X\\ 0&\text{otherwise.}\end{cases}
Proof.

Clearly, hi(X)h_{i}(X) is sjs_{j}-invariant if both jj and j+1j+1 are inside or outside XX, so j(hi(X))=0\partial_{j}(h_{i}(X))=0. Assume now that jXj\in X and j+1Xj+1\not\in X. We have hi(X)=hi(X{j})+hi1(X)xjh_{i}(X)=h_{i}(X\setminus\{j\})+h_{i-1}(X)x_{j}. Hence,

j(hi(X))=j(hi1(X)xj).\partial_{j}(h_{i}(X))=\partial_{j}(h_{i-1}(X)x_{j}).

Applying the twisted Leibniz rule, by induction on |X||X|, we obtain

j(hi(X))\displaystyle\partial_{j}(h_{i}(X)) =hi2(X{j+1})xj+sj(hi1(X))j(xj)\displaystyle=h_{i-2}(X\cup\{j+1\})x_{j}+s_{j}(h_{i-1}(X))\partial_{j}(x_{j})
=hi2(X{j+1})xj+hi1(X{j+1}{j})\displaystyle=h_{i-2}(X\cup\{j+1\})x_{j}+h_{i-1}(X\cup\{j+1\}\setminus\{j\})
=hi1(X{j+1}).\displaystyle=h_{i-1}(X\cup\{j+1\}).

The case where j+1Xj+1\in X and jXj\notin X follows because j(sj(f))=j(f)\partial_{j}(s_{j}(f))=-\partial_{j}(f), so we have

j(hi(X))=j(hi(X{j}{j+1}))=hi1(X{j}).\partial_{j}(h_{i}(X))=-\partial_{j}(h_{i}(X\cup\{j\}\setminus\{j+1\}))=-h_{i-1}(X\cup\{j\}).\qed
Theorem 6.3.

Use the notation of §6.1. Fix a,b1a,b\geq 1, and let n=a+b.n=a+b. Let X={1,,a}X=\{1,\ldots,a\} and Y={n,n1,,n+1a}Y=\{n,n-1,\ldots,n+1-a\}. Recall y0,y1𝐒ny_{0},y_{1}\in\mathbf{S}_{n} from (90).

Then for any i0i\geq 0 and any gR𝐒a×𝐒b=RJg\in R^{\mathbf{S}_{a}\times\mathbf{S}_{b}}=R^{J} we have

(93) y0(hi(X)g)=hi(Y)y0(g)+y1(hi1(Xn)g).\partial_{y_{0}}(h_{i}(X)\cdot g)=h_{i}(Y)\cdot\partial_{y_{0}}(g)+\partial_{y_{1}}(h_{i-1}(X\cup n)\cdot g).

As p=y0,\partial_{p}=\partial_{y_{0}}, p¯=y0,\underline{p}=y_{0}, and y0(hi(X))=hi(Y)y_{0}(h_{i}(X))=h_{i}(Y), this is compatible with (30), where Tq1(hi(X))=hi1(Xn)T_{q_{1}}(h_{i}(X))=h_{i-1}(X\cup n), and Tqk(hi(X))T_{q_{k}}(h_{i}(X)) is zero for all k>1k>1. Most of the terms in (30) are zero for complete symmetric polynomials, making the formula much easier than the general case.

Proof.

We will do a proof by example, for the example a=3a=3 and b=4b=4. The general proof is effectively the same only the notation is more cumbersome.

In this proof, we write 123\partial_{123} for 123\partial_{1}\circ\partial_{2}\circ\partial_{3} (and not the Frobenius trace associated to the longest element w123w_{123}). We use parenthesization for emphasis, so that (12)3\partial_{(12)3} is the same thing as 123\partial_{123}, but emphasizes that (12)3=123\partial_{(12)3}=\partial_{12}\circ\partial_{3}.

Remember that gg is invariant under anything except s3s_{3}, so j(g)=0\partial_{j}(g)=0 if j3j\neq 3. This implies, for example, that 3\partial_{3} kills 23(g)\partial_{23}(g), and that 2\partial_{2} kills 3243(g)\partial_{3243}(g), etcetera.

We claim that

(94) 123(hi(123)g)=hi1(1234)23(g)+hi(234)123(g).\partial_{123}(h_{i}(123)g)=h_{i-1}(1234)\partial_{23}(g)+h_{i}(234)\partial_{123}(g).

One proof is to apply the ordinary twisted Leibniz rule repeatedly using (92). After one step we obtain

123(hi(123)g)=12(hi1(1234)g+hi(124)3(g)).\partial_{123}(h_{i}(123)g)=\partial_{12}\left(h_{i-1}(1234)g+h_{i}(124)\partial_{3}(g)\right).

The first term on the right-hand side is invariant under s2s_{2}, so it is killed by 2\partial_{2}. Thus we have

123(hi(123)g)=1(2(hi(124)3(g)))=1(hi1(1234)3(g)+hi(134)23(g)).\partial_{123}(h_{i}(123)g)=\partial_{1}(\partial_{2}(h_{i}(124)\partial_{3}(g)))=\partial_{1}(h_{i-1}(1234)\partial_{3}(g)+h_{i}(134)\partial_{23}(g)).

Again the first term on the right-hand side is invariant under s1s_{1}, so it is killed by 1\partial_{1}. A final application of the twisted Leibniz rule to 1(hi(134)23(g))\partial_{1}(h_{i}(134)\partial_{23}(g)) gives (94). Essentially, this proof is by iterating the twisted Leibniz rule and arguing that the first term vanishes in every application but the last, because the first term is appropriately invariant. We call this the easy invariance argument.

Note that 123(g)\partial_{123}(g) is invariant under everything but s4s_{4}. More generally, 12a(g)\partial_{12\ldots a}(g) is killed by j\partial_{j} for all ja+1j\neq a+1. This is true when j=1j=1 since 11=0\partial_{1}\partial_{1}=0. This is true when j>a+1j>a+1 since j12a=12aj\partial_{j}\partial_{12\ldots a}=\partial_{12\ldots a}\partial_{j} and j(g)=0\partial_{j}(g)=0. This is true when 2ja2\leq j\leq a because j12a=12aj1\partial_{j}\partial_{12\ldots a}=\partial_{12\ldots a}\partial_{j-1} and j1(g)=0\partial_{j-1}(g)=0.

By the easy invariance argument again, but with indices shifted and gg replaced by 123(g)\partial_{123}(g), we have

(95) 234(hi(234)123(g))=hi1(2345)34123(g)+hi(345)234123(g).\partial_{234}(h_{i}(234)\partial_{123}(g))=h_{i-1}(2345)\partial_{34123}(g)+h_{i}(345)\partial_{234123}(g).

This is how we treat the second term in the right side of (94).

Note that all computations above are unchanged by adding new variables to our complete symmetric polynomials which are untouched by any of the simple reflections used by the formula. For example, adding 77 to every hih_{i} in (94) we get

(96) 123(hi(1237)g)=hi1(12347)23(g)+hi(2347)123(g).\partial_{123}(h_{i}(1237)g)=h_{i-1}(12347)\partial_{23}(g)+h_{i}(2347)\partial_{123}(g).

In this example, we call 77 an irrelevant index.

Now we examine the first term in the right side of (94). Note that 23(g)\partial_{23}(g) is invariant under all simple reflections except s1s_{1} and s4s_{4}. For the next computation, the index 11 is irrelevant. The easy invariance argument again implies that

(97) 234(hi1(1234)23(g))=hi2(12345)3423(g)+hi1(1345)23423(g).\partial_{234}(h_{i-1}(1234)\partial_{23}(g))=h_{i-2}(12345)\partial_{3423}(g)+h_{i-1}(1345)\partial_{23423}(g).

However, as 23423=32434,23423=32434, we have 23423(g)=0\partial_{23423}(g)=0, so one has the simpler formula

(98) 234(hi1(1234)23(g))=hi2(12345)3423(g).\partial_{234}(h_{i-1}(1234)\partial_{23}(g))=h_{i-2}(12345)\partial_{3423}(g).

Overall, we see that

(234)(123)(hi(123)g)\displaystyle\partial_{(234)(123)}(h_{i}(123)g) =\displaystyle= hi(345)(234)(123)(g)\displaystyle h_{i}(345)\partial_{(234)(123)}(g)
+\displaystyle+ hi1(2345)(34)(123)(g)\displaystyle h_{i-1}(2345)\partial_{(34)(123)}(g)
+\displaystyle+ hi2(12345)(34)(23)(g).\displaystyle h_{i-2}(12345)\partial_{(34)(23)}(g).

The pattern is relatively straightforward. Here’s the next one in the pattern:

(345)(234)(123)(hi(123)g)\displaystyle\partial_{(345)(234)(123)}(h_{i}(123)g) =\displaystyle= hi(456)(345)(234)(123)(g)\displaystyle h_{i}(456)\partial_{(345)(234)(123)}(g)
+\displaystyle+ hi1(3456)(45)(234)(123)(g)\displaystyle h_{i-1}(3456)\partial_{(45)(234)(123)}(g)
+\displaystyle+ hi2(23456)(45)(34)(123)(g)\displaystyle h_{i-2}(23456)\partial_{(45)(34)(123)}(g)
+\displaystyle+ hi3(123456)(45)(34)(23)(g).\displaystyle h_{i-3}(123456)\partial_{(45)(34)(23)}(g).

The word whose Demazure is applied to gg is obtained from the concatenation of triples (345)(234)(123)(345)(234)(123) by removing the first index from some of the triples; more specifically, from a prefix of the set of triples. The indices that get removed are instead added to the complete symmetric polynomial. The reason triples appear is because a=3a=3.

The inductive proof of this pattern is the same as above. One takes (6.2) and applies 345\partial_{345}. The first term splits in two, giving the first two terms of (6.2), similar to (94) or (95). Each other term contributes one term in (6.2), similar to (98).

Note that we could have added the irrelevant index 77 to every set in sight within (6.2), without any issues. This will be important later.

Repeating until one applies y0\partial_{y_{0}}, we calculate y0(hi(X)g)\partial_{y_{0}}(h_{i}(X)\cdot g):

(456)(345)(234)(123)(hi(123)g)\displaystyle\partial_{(456)(345)(234)(123)}(h_{i}(123)g) =\displaystyle= hi(567)(456)(345)(234)(123)(g)\displaystyle h_{i}(567)\partial_{(456)(345)(234)(123)}(g)
+\displaystyle+ hi1(4567)(56)(345)(234)(123)(g)\displaystyle h_{i-1}(4567)\partial_{(56)(345)(234)(123)}(g)
+\displaystyle+ hi2(34567)(56)(45)(234)(123)(g)\displaystyle h_{i-2}(34567)\partial_{(56)(45)(234)(123)}(g)
+\displaystyle+ hi3(234567)(56)(45)(34)(123)(g)\displaystyle h_{i-3}(234567)\partial_{(56)(45)(34)(123)}(g)
+\displaystyle+ hi4(1234567)(56)(45)(34)(23)(g).\displaystyle h_{i-4}(1234567)\partial_{(56)(45)(34)(23)}(g).

The fact that there were four triples is because b=4b=4. Note that the first term in the RHS is hi(Y)y0(g).h_{i}(Y)\partial_{y_{0}}(g).

Let us now compute y1(hi1(1237)g)\partial_{y_{1}}(h_{i-1}(1237)\cdot g). Note that y1=(56)(345)(234)(123)y_{1}=(56)(345)(234)(123). To compute (345)(234)(123)(hi1(1237)g)\partial_{(345)(234)(123)}(h_{i-1}(1237)g), we take (6.2), add the irrelevant index 77 to all variable lists, and reduce ii by one. Now we need only apply (56)\partial_{(56)} to the result. The key thing to note here is that each h(567)h_{\bullet}(\cdots 567) is invariant already under s5s_{5} and s6s_{6}. Thus both operators in (56)\partial_{(56)} simply apply to the gg term. From this we can compute y1(hi1(Xn)g)\partial_{y_{1}}(h_{i-1}(X\cup n)\cdot g):

(56)(345)(234)(123)(hi1(1237)g)\displaystyle\partial_{(56)(345)(234)(123)}(h_{i-1}(1237)g) =\displaystyle= hi1(4567)(56)(345)(234)(123)(g)\displaystyle h_{i-1}(4567)\partial_{(56)(345)(234)(123)}(g)
+\displaystyle+ hi2(34567)(56)(45)(234)(123)(g)\displaystyle h_{i-2}(34567)\partial_{(56)(45)(234)(123)}(g)
+\displaystyle+ hi3(234567)(56)(45)(34)(123)(g)\displaystyle h_{i-3}(234567)\partial_{(56)(45)(34)(123)}(g)
+\displaystyle+ hi4(1234567)(56)(45)(34)(23)(g).\displaystyle h_{i-4}(1234567)\partial_{(56)(45)(34)(23)}(g).

This exactly matches all terms from (6.2) except the first term. Thus the theorem is proven! ∎

Theorem 6.4.

The atomic Leibniz rule and atomic polynomial forcing both hold for atomic cosets 𝚊[I,M,J]{\mathtt{a}}\leftrightharpoons[I,M,J] when WM=𝐒nW_{M}=\mathbf{S}_{n} when R=[x1,,xn]R=\mathbb{Z}[x_{1},\ldots,x_{n}].

Proof.

Theorem 6.3 proved property (3) from Theorem 5.5 in this case. Thus conditions (1) and (5) also hold in this case. ∎

6.3. Reduction to the connected case

The previous section proves an atomic Leibniz rule under the assumption WM=𝐒nW_{M}=\mathbf{S}_{n}. Now we do the general case.

Let W=𝐒nW=\mathbf{S}_{n}. An arbitrary atomic coset in WW contains the longest element of the reducible Coxeter group WM=𝐒n1××𝐒nkW_{M}=\mathbf{S}_{n_{1}}\times\cdots\times\mathbf{S}_{n_{k}} where ni=n\sum n_{i}=n. It is a coset for (s^,t^)(\widehat{s},\widehat{t}), where ss and tt are simple reflections in the same irreducible component 𝐒ni\mathbf{S}_{n_{i}} of WMW_{M}. We can prove polynomial forcing for an arbitrary atomic coset in type AA if we can bootstrap the result from 𝐒ni\mathbf{S}_{n_{i}} to WMW_{M}.

In this discussion, there is no difference between type AA and a general Coxeter type. Thus let MM be finitary, with s,tMs,t\in M and wMswM=tw_{M}sw_{M}=t. Let 𝚊{\mathtt{a}} be the atomic (s^,t^)(\widehat{s},\widehat{t})-coset containing wMw_{M}.

Now suppose that M=C1C2CkM=C_{1}\sqcup C_{2}\sqcup\ldots\sqcup C_{k} is a disjoint union of connected components (the simple reflections in CiC_{i} commute with those in CjC_{j} for iji\neq j). Suppose without loss of generality that sC1s\in C_{1}, and let D=C2CkD=C_{2}\sqcup\ldots\sqcup C_{k}. Then tC1t\in C_{1} as well, and t=wC1swC1t=w_{C_{1}}sw_{C_{1}}. Let 𝚊{\mathtt{a}}^{\prime} denote the atomic (C1s,C1t)(C_{1}\setminus s,C_{1}\setminus t)-coset containing wC1w_{C_{1}}. Then 𝚊{\mathtt{a}} and 𝚊{\mathtt{a}}^{\prime} are related by the operation +D+D described in [EK21, §4.10].

Lemma 6.5.

With the notation as above, polynomial forcing holds for 𝚊{\mathtt{a}} if and only if it holds for 𝚊{\mathtt{a}}^{\prime}.

Lemma 6.6.

With the notation as above, an atomic Leibniz rule holds for 𝚊{\mathtt{a}} if and only if it holds for 𝚊{\mathtt{a}}^{\prime}.

Proof.

The proof is straightforward and left to the reader, but we wish to point out the available ingredients. Many basic properties of the operator +D+D are given in [EK21, §4.10]. There is a bijection between cosets q<𝚊q<{\mathtt{a}} and cosets q<𝚊q^{\prime}<{\mathtt{a}}^{\prime}, and also a bijection between their reduced expressions. Note that q¯wt^1=q¯wC1t1\partial_{\overline{q}w_{\widehat{t}^{-1}}}=\partial_{\overline{q^{\prime}}w_{C_{1}\setminus t}^{-1}} as operators RRR\to R. Finally, dual bases for the Frobenius extension RMRMtR^{M}\subset R^{M\setminus t} can also be chosen as dual bases for the Frobenius extension RC1RC1tR^{C_{1}}\subset R^{C_{1}\setminus t}. ∎

Theorem 6.7.

The atomic Leibniz rule and polynomial forcing hold for any atomic coset in type An1A_{n-1} when R=[x1,,xn]R=\mathbb{Z}[x_{1},\ldots,x_{n}].

Proof.

The restriction of the permutation realization to any 𝐒ni𝐒n\mathbf{S}_{n_{i}}\subset\mathbf{S}_{n} is a WW-invariant enlargement of the permutation realization of 𝐒ni\mathbf{S}_{n_{i}}. Thus the result follows from the previous two lemmas, Theorem 6.4, and Lemma 3.17. ∎

Applying 3.17, we also deduce the atomic Leibniz rule for a host of other realizations, including when R=𝕜[x1,,xn]R=\Bbbk[x_{1},\ldots,x_{n}] for any commutative ring 𝕜\Bbbk.

References

  • [Abe24] Noriyuki Abe “Singular Soergel bimodules for realizations”, 2024 arXiv:2401.04994
  • [Ach+19] Pramod N. Achar, Shotaro Makisumi, Simon Riche and Geordie Williamson “Koszul duality for Kac-Moody groups and characters of tilting modules” In J. Amer. Math. Soc. 32.1, 2019, pp. 261–310
  • [Dem73] Michel Demazure “Invariants symétriques entiers des groupes de Weyl et torsion” In Invent. Math. 21, 1973, pp. 287–301
  • [EK10] Ben Elias and Mikhail Khovanov “Diagrammatics for Soergel categories” In Int. J. Math. Math. Sci., 2010, pp. Art. ID 978635\bibrangessep58
  • [EK21] Ben Elias and Hankyung Ko “A singular Coxeter presentation”, 2021 arXiv:2105.08563
  • [EKLP23a] Ben Elias, Hankyung Ko, Nicolas Libedinsky and Leonardo Patimo “Demazure operators for double cosets”, 2023 arXiv:2307.15021
  • [EKLP23b] Ben Elias, Hankyung Ko, Nicolas Libedinsky and Leonardo Patimo “Subexpressions and the Bruhat order for double cosets”, 2023 arXiv:arXiv:2307.15726
  • [EKLP24a] Ben Elias, Hankyung Ko, Nicolas Libedinsky and Leonardo Patimo “Singular Light Leaves”, 2024 arXiv:arXiv:2401.03053
  • [EKLP24b] Ben Elias, Hankyung Ko, Nicolas Libedinsky and Leonardo Patimo “On reduced expressions for core double cosets”, 2024 arXiv:arXiv:2402.08673
  • [Eli16] Ben Elias “The two-color Soergel calculus” In Compos. Math. 152.2, 2016, pp. 327–398
  • [EMTW20] Ben Elias, Shotaro Makisumi, Ulrich Thiel and Geordie Williamson “Introduction to Soergel bimodules” 5, RSME Springer Series Springer, Cham, 2020, pp. xxv+588
  • [ESW17] Ben Elias, Noah Snyder and Geordie Williamson “On cubes of Frobenius extensions” In Representation theory—current trends and perspectives, EMS Ser. Congr. Rep. Eur. Math. Soc., Zürich, 2017, pp. 171–186
  • [EW14] Ben Elias and Geordie Williamson “The Hodge theory of Soergel bimodules” In Ann. of Math. (2) 180.3, 2014, pp. 1089–1136
  • [EW16] Ben Elias and Geordie Williamson “Soergel calculus” In Represent. Theory 20, 2016, pp. 295–374
  • [EW21] Ben Elias and Geordie Williamson “Relative hard Lefschetz for Soergel bimodules” In J. Eur. Math. Soc. (JEMS) 23.8, 2021, pp. 2549–2581
  • [Lib10] Nicolas Libedinsky “Presentation of right-angled Soergel categories by generators and relations” In J. Pure Appl. Algebra 214.12, 2010, pp. 2265–2278
  • [LW18] George Lusztig and Geordie Williamson “Billiards and Tilting Characters for SL3\rm SL_{3} In SIGMA Symmetry Integrability Geom. Methods Appl. 14, 2018, pp. Paper No. 015\bibrangessep22
  • [RW18] Simon Riche and Geordie Williamson “Tilting modules and the pp-canonical basis” In Astérisque, 2018, pp. ix+184
  • [Wil11] Geordie Williamson “Singular Soergel bimodules” In Int. Math. Res. Not. IMRN, 2011, pp. 4555–4632
  • [Wil17] Geordie Williamson “Schubert calculus and torsion explosion” With a joint appendix with Alex Kontorovich and Peter J. McNamara In J. Amer. Math. Soc. 30.4, 2017, pp. 1023–1046