ABSTRACT Title of Dissertation: A COMBINATORIAL STUDY OF AFFINE DELIGNE-LUSZTIG VARIETIES Arghya Sadhukhan Doctor of Philosophy, 2023 Dissertation Directed by: Professor Jeffrey Adams Department of Mathematics We consider affine Deligne-Lusztig varieties Xw(b) and certain unions X(µ, b) in the affine flag variety of a connected reductive group. They were first introduced by Rapoport to facilitate the study of mod-p reduction of Shimura varieties and moduli spaces of shtukas. We improve upon certain existing results in the study of affine Deligne-Lusztig varieties by weakening the hypothesis to prove them. Such results include a description of generic Newton points in Iwahori double cosets in the loop group of a split reductive group, covering relations in the associated Iwahori-Weyl group, and a dimension formula for X(µ, b) in the case of a quasi-split group. As an application of the work on generic Newton point formula, we obtain a description of the dimension for X(µ, b) associated with the maximal element b in its natural range, under a mild hypothesis on µ but no further restrictions on the group. A COMBINATORIAL STUDY OF AFFINE DELIGNE-LUSZTIG VARIETIES by Arghya Sadhukhan Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2023 Advisory Committee: Professor Jeffrey Adams, Chair/Advisor Professor Thomas J. Haines, Co-chair Professor Xuhua He Professor Yihang Zhu Professor Cole Miller, Dean’s representative © Copyright by Arghya Sadhukhan 2023 Dedication To the teachers who inspire, empower, and propel us to new mathematical horizons. ii Acknowledgments First and foremost, I want to express my heartfelt gratitude to my dissertation advisor Xuhua He, who despite being on the other side of the world, has made a lasting impact on my research and provided invaluable guidance throughout this journey. Navigating a Ph.D. through Zoom meetings presented its challenges, but his dedication, expertise, and unwavering support have been instrumental in my academic growth. I am also immensely grateful to my local supervisor Jeffrey Adams, who has been there for me both mathematically and logistically whenever I needed assistance. His willingness to offer support, answer my questions, and provide guidance has greatly enriched my grad school journey. I would like to extend special thanks to Thomas Haines and Michael Rapoport. Engaging in conversations with them about Shimura varieties and the Langlands program has broadened my understanding and enhanced the depth of this dissertation. Additionally, I want to express my appreciation to the professors in the math depart- ment for offering relevant courses and for their willingness to discuss mathematics outside of classroom, especially Patrick Brosnan, Harry Tamvakis and Yihang Zhu. I am grateful to the staff at the Department of Mathematics and the STEM library at the University of Maryland, especially Haydee, Cristina, Liliana, and Ruth, for their continued support in navigating various bureaucratic systems seamlessly. Going back further, I want to acknowledge the teachers and mentors from the days of my undergraduate and masters studies who have played a pivotal role in shaping my aca- iii demic journey, especially Arijit Ganguly, Parthasarathi Mukhopadhyay, and Amritanshu Prasad. Gratitude is also due to the organizers of the MTTS-2014 summer camp in India, where I was first exposed to mathematical research, sparking my passion for the subject. I am deeply thankful for the public education infrastructure in both India and the United States. Without the opportunities and resources provided by these systems, pur- suing my academic dreams would not have been possible. Additionally, I want to express my gratitude for the support received through the Graduate School Summer Research Fellowship and the Sung Dissertation Fellowship, which have allowed me to focus on my research. I am grateful for the friends I have picked up during my Ph.D. journey. To Chengze, Deric, Dimitri, Gautam, Naren, Prakhar, Rachel, Saurav, Shin, Sohitri, Stavros, Subhayan, Sze-Hong, Tamoghna and Yiannis, thank you for your warm camaraderie, unwavering support, and shared experiences. Your friendship has made this journey more meaningful and enjoyable. I would also like to thank my friends from my undergraduate days, especially Sovanlal, Srijan and Sudipta for their constant companionship and for being memorable travel buddies. During a challenging time post an accident in December 2022, I would like to express my gratitude to the local bike store, Proteus Bicycles, for their assistance and support. I would also like to extend my thanks to the American Film Institute theatre in Silver Spring for providing a much-needed respite and inspiration during the course of this Ph.D. journey. Additionally, I want to express my appreciation to the authorities in charge of maintain- ing and preserving the trails, wilderness, and nature near me, especially the Shenandoah National Park. The serenity and beauty of nature have provided solace and rejuvenation iv during the demanding research process. I would also like to express my gratitude to all the cats I had the privilege to play with, including Jobi-Chan, Django, Elsa, and Cocos. My deepest appreciation goes to my parents, Amitava and Soma, and my sister, Somdutta for their mental support and for allowing me to pursue mathematics away from home for more than a decade. I want to extend a special thank you to my partner, Mita, for being an anchor throughout the better part of this journey. I look forward to the time when we finally solve the two-body problem at hand together! Finally, I would like to express my deepest admiration and gratitude to Robert Lang- lands for his visionary creation of a grand web of compelling conjectures, and to Robert Kottwitz for his invaluable contributions in advancing the field through the infusion of group-theoretic methods into this majestic edifice. To echo the sentiment of Langlands himself, “certainly the best times were when I was alone with mathematics, free of ambition and of pretense, and indifferent to the world.” v Table of Contents Preface ii Foreword ii Dedication ii Acknowledgements iii Table of Contents 1 1 Introduction 2 1.1 Affine Deligne-Lusztig varieties in the affine flag variety . . . . . . . . . . . 3 1.2 Certain union of affine Deligne-Lusztig varieties . . . . . . . . . . . . . . . 6 1.3 Other related work on affine Weyl groups . . . . . . . . . . . . . . . . . . . 11 2 Preliminaries 12 2.1 Notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 2.2 The σ-conjugacy classes of Ğ . . . . . . . . . . . . . . . . . . . . . . . . . 15 2.3 Affine Deligne-Lusztig varieties . . . . . . . . . . . . . . . . . . . . . . . . 18 2.4 Demazure product and its variations . . . . . . . . . . . . . . . . . . . . . 20 2.5 Quantum Bruhat graph . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 3 Generic Newton point and affine Bruhat order 24 3.1 Some combinatorial properties . . . . . . . . . . . . . . . . . . . . . . . . . 25 3.2 Formula for the generic Newton point . . . . . . . . . . . . . . . . . . . . . 27 3.3 Weight of the longest element . . . . . . . . . . . . . . . . . . . . . . . . . 43 3.4 Covering relation in Iwahori-Weyl group . . . . . . . . . . . . . . . . . . . 47 4 A dimension formula for X(µ, b) 64 4.1 Dimension in the quasi-split case . . . . . . . . . . . . . . . . . . . . . . . 65 4.2 Expressing bmax via generic Newton point . . . . . . . . . . . . . . . . . . . 70 4.3 Minimal length elements in certain σ0-conjugacy class . . . . . . . . . . . . 79 4.4 Explicit description of the Newton point of bmax . . . . . . . . . . . . . . . 83 5 Some remarks about the weight function in type An 90 Bibliography 99 1 Chapter 1: Introduction In their seminal paper [DL76], Deligne and Lusztig gave a geometric recipe to con- struct all the irreducible representations for finite groups of Lie type in terms of l-adic cohomology of certain algebraic varieties associated with elements of the corresponding finite Weyl group. On the other hand, the affine Deligne-Lusztig varieties - their coun- terpart for the affine root system - were first introduced by Rapoport [Rap05] and have found substantial application in the geometry of Shimura varieties and moduli spaces of shtukas, and therefore have been a geometric object of recurring interest in the Langlands program. In [Lan77], Langlands outlines a three-part approach to prove that the Hasse- Weil ζ-functions of Shimura varieties are related to L-functions of automorphic forms; this expectation lies at the heart of Langlands program. A key input in this approach is the description of the geometric points of special fibers of suitable integral models of Shimura varieties. A conjectural description of such mod-p points was put forth by Langlands and Rapoport in the cases of good reduction, and subsequently modified by Rapoport and Kottwitz to include cases of parahoric-level bad reduction, and the geometry of associated affine Deligne-Lusztig varieties is a key player in proving such result. For instance, infor- mation about connected components of affine Deligne-Lusztig varieties - a problem that has generated a lot of attention in the past decade and has only been proved in complete 2 generality in a very recent preprint [GL22] - has played an important role in the version of the Langlands-Rapoport conjecture proved in [Kis17] and its subsequent strengthening in [KSZ21]. Due to the combinatorial complexity of the affine Weyl groups, several important results in the study of affine Deligne-Lusztig varieties (in the affine flag variety) in recent years have been established only under the so-called superregularity hypothesis. In this thesis, we weaken the superregularity hypothesis on them and sometimes eliminate it, thus strengthening these existing results. Additionally, we compute the dimension of a certain naturally occurring union X(µ, b) of affine Deligne-Lusztig varieties beyond the setting of quasi-split groups. We now proceed to explain the main results of this thesis in more detail. 1.1 Affine Deligne-Lusztig varieties in the affine flag variety We refer to Chapter 2 for an explanation of notations and definitions. Let G be a reductive group over a non-archimedean local field F , with residue field κF and completed maximal unramified extension F̆ . Denote by σ the Frobenius morphism of F̆ /F and choose a σ-stable Iwahori subgroup I. Then letting Ğ := G(F̆ ), Ĭ := I(F̆ ), we have two natural decompositions of the loop group Ğ: namely, Ğ = ∐ [b]∈B(G) [b] = ∐ w∈W̃ ĬwĬ. Here [b] := {g−1bσ(g) : g ∈ Ğ} is the σ-conjugacy class of b, the so-called Kottwitz set B(G) is the collection of such classes and W̃ is the Iwahori-Weyl group of G. For [b] ∈ B(G) and w ∈ W̃ , the associated affine Deligne-Lusztig variety Xw(b) is a locally closed, reduced 3 subscheme locally of finite type inside the affine flag variety F lG, with geometric points given by Xw(b)(κ̄F ) := {gĬ : g−1bσ(g) ∈ ĬwĬ} ⊂ Ğ/Ĭ. We list some of the major problems in this field below: 1. For which w, b is Xw(b) nonempty? 2. If non-empty, is Xw(b) equidimensional and can we give a closed formula for its dimension? Let us briefly mention why these questions are of interest from the perspective of arithmetic geometry. Roughly speaking (at least for the Shimura varieties of Hodge type that admits a moduli interpretation), on the special fiber of a suitable integral model of the Shimura variety associated to the Shimura datum (G, µ) with Iwahori level structure, there are two important stratifications: the Newton stratification, coming from grouping together abelian varieties according to the isogeny class of their p-divisible groups, and the Kottwitz- Rapoport stratification, induced from the Iwahori orbits on the associated local model; for an axiomatic approach to these stratifications pertaining to general Shimura varieties, see [HR17]. The strata are indexed by a certain subset of B(G) in the former case, while in the latter case the index set is a certain subset of W̃ governed by the Shimura datum. The affine Deligne-Lusztig varieties capture the delicate interation between these two stratifications; for instance, Xw(b) ̸= ∅ if and only if the Newton stratum indexed by the element b meets the Kottwitz-Rapoport stratum indexed by the element w, cf. [Hai05, §12.3]. However, even this nonemptiness problem has not fully been resolved in complete generality; we mention 4 the work done in [Gör+10], [GHN15], and finally [He21b] for state-of-the-art result, as well as an interesting conjecture made in [Lim23] in this direction. An important feature of B(G) is the poset structure on it, defined via closure relation in Ğ. Recent results of [MV20] highlight certain special elements of W̃ : they show that if the maximal element [bw] of B(G)w := {[b] ∈ B(G) : Xw(b) ̸= ∅} - which coincides with the generic σ-conjugacy class in ĬwĬ - satisfies an explicit group-theoretic condition called cordiality, then the poset B(G)w is saturated; furthermore, all the Newton strata meeting the fixed Kottwitz-Rapoport stratum indexed by w, as well as the affine Deligne-Lusztig varieties associated with w exhibits especially well-behaved geometry. In essence, their theorem gives a condition that can be checked from knowledge of this maximal element bw of B(G)w, but it provides important information about the shape of the entire poset. In light of such results, it becomes important to describe bw explicitly in a way that can be used to check whether w is cordial. By virtue of the concrete parametrization [Kot85] σ-conjugacy classes via their Kottwitz and Newton points (κ, ν) : B(G) ↪→ π1(G)Γ × (X∗(T )Γ0 ⊗Q)+, this amounts to asking the following: Question 1.1.1. Give an explicit closed formula for the generic Newton point ν([bw]). It turns out that such a description comes from the technical framework of the quan- tum Bruhat graph: for elements of W̃ that are sufficiently far away from the walls of any chamber, such a formula was first established in [Mil21]. This combinatorial tool was intro- duced by Brenti, Fomin and Postnikov in [BFP99] to describe the multiplicative structure of the quantum cohomology ring of the complex flag manifold. It is obtained by augmenting the usual Hasse diagram for the Bruhat order on the finite Weyl group W by some quantum 5 edges - certain downward edges that are labeled by the coroot associated to the reflection used to get from one vertex to the other. With this setup, one can define a certain weight function wt : W ×W → Z≥0Φ ∨, cf. Section 2.5. Then we have the following result, see Theorem 3.2.1. Theorem 1.1.2. Suppose that G is a quasi-simple split group of semi-simple rank n. Let w = xtλy be an element of W̃ with x, y ∈ W , and depth(λ) > Ξn, where Ξn is certain explicit linear function of n. Then ν([bw]) = λ− wt(y−1, x). Here for a dominant cocharacter λ its depth(λ) := min{⟨α, λ⟩ : α simple root} quan- tifies its distance from the walls of the fundamental Weyl chamber. Hence this result weakens the hypothesis on the lower bound of the depth from O(n2) in Milićević’s work to O(n). The proof of Theorem 1.1.2 crucially employs the Demazure product and its variations on the finite Weyl group W : this comes from the product in the associated 0-Hecke algebra and induces a monoid structure on W . Using these tools, we show that the generic Newton point ν(bw) is given by the Newton point of the maximal translation element below w, and we identify this element via simple Bruhat order considerations. 1.2 Certain union of affine Deligne-Lusztig varieties Rapoport in [Rap05] predicted “whereas the individual affine Deligne-Lusztig varieties are very difficult to understand, the situation seems to change radically when we form a 6 suitable finite union of them.” Here the union refers to X(µ, b) := ⋃ w∈Adm(µ) Xw(b), where for a dominant cocharacter µ, we define its admissible set as Adm(µ) := {w ∈ W̃ : w ≤ tx(µ) for some x ∈ W}. The interest in studying such a union comes from the fact that (in mixed character- sitics) they arise as the underlying reduced scheme of a formal moduli space of p-divisible groups, known as Rapoport-Zink space, cf. [RV14]. Something analogous holds in the function field case for formal moduli spaces of Shtukas, cf. [Vie18]; in this latter case, µ can be arbitrary. Rapoport-Zink spaces are local analogs of Shimura varieties, and are also related to Shimura varieties themselves by the theory of p-adic uniformization. The problem of describing the l-adic cohomology of Rapoport-Zink spaces may be traced back to Lubin-Tate theory of formal groups and has had important consequences such as the proof of the local Langlands conjecture and the local-global compatibility of the Langlands correspondence in [HT01]. As a first validation of Rapoport’s expectation, let us note that the non-emptiness pattern of X(µ, b) is completely understood (as opposed to that of Xw(b)): settling the Kottwitz-Rapoport conjecture in [He16a], He proves that X(µ, b) is non-empty precisely when [b] lies in the set B(G, µ) of neutrally acceptable σ-conjugacy classes ; this subset of B(G) is defined by a group theoretic reformulation of Mazur’s inequality between the Hodge polygon of an F -crystal and the Newton polygon of its underlying F -isocrystal. As a crucial ingredient toward obtaining a dimension formula for X(µ, b), a useful 7 description of sufficiently large elements in the admissible set Adm(µ) in terms of quantum Bruhat graph theoretic data is given in [HY21], whenever the depth of µ is O(n2) with respect to the semisimple rank n. Based on the work on affine Bruhat order (see next section), we can relax this hypothesis on depth(µ) to O(1); additionally, we exploit the additivity of the admissible set to upgrade it to a necessary numerical criterion that is unconditional on µ. Proposition 1.2.1. Suppose that W is an irreducible Weyl group associated with an affine Weyl group W̃ . Let µ, λ be dominant cocharacters. 1. Assume that depth(µ) is bigger than a certain constant ΘW that is at most 6 in all Cartan types, and further that ⟨ρ, µ−λ⟩ < ⌈depth(µ)−ΘW 2 ⌉. Then for any two elements x, y ∈ W , we have xtλy ∈ Adm(µ) if and only if wt(x, y−1) ≤ µ− λ. 2. Assume that µ is only dominant regular, i.e. depth(µ) ≥ 1. Then ⟨ρ,wt(x, y−1)⟩ ≤ ⟨ρ, µ− λ⟩ if xtλy ∈ Adm(µ). This follows from Proposition 3.4.13 and Proposition 4.1.1. Based on Proposition 1.2.1(2), we show that certain key steps in the proof in loc. sit. of the dimension formula can be carried out differently to bypass the O(n2) superregularity constraint. Thus we establish in Section 4.1 the same dimension formula under just a regularity assumption on µ: Theorem 1.2.2. Suppose that G is a quasi-split group. Let µ be dominant regular and assume that [b] ∈ B(G, µ). Denote by O the σ-conjugacy class of the longest element w0 in W , and let ℓR(O) be the minimal reflection length of elements in O. 8 1. Let us further require that the Galois average µ⋄ ≥ ν([b]) + 2ρ∨. Then dimX(µ, b) = ⟨ρ, µ− ν(b)⟩ − 1 2 defG(b) + 1 2 (ℓ(w0)− ℓR(O)). 2. Furthermore, if G is split, we obtain the same result under the hypothesis that µ ≥ ν([b]) + wt(w0, 1) and depth(µ) > 2. We refer the reader to Section 3.3, Section 2.2.1 and Section 2.3.1 for relevant defi- nitions of the notions reflection length ℓR, Galois average µ⋄ and defect defG, respectively. Note that the quasi-split assumption on G is crucial in the above theorem; unlike in the case of a single affine Deligne-Lusztig variety, here we cannot leverage Proposition 4.2.6 to deduce a dimension formula of X(µ, b) for general G. There is no direct relation be- tween the admissible set Adm(µ) - and hence X(µ, b) - for an arbitrary reductive group and its quasi-split inner form. This adds essential difficulties in the study of X(µ, b) for non-quasi-split groups. Nevertheless, there has been some partial success in the context of general groups. Most notably, the investigation into the structure of X(µ, b) for the element b = bmin := minB(G, µ) has led to the striking observation that in certain cases, the basic locus admits a nice description as a union of classical Deligne-Lusztig varieties, the index set and the closure relations between the strata being encoded in a Bruhat-Tits building attached to the group theoretic data coming from the Shimura variety; such descriptions have been used with great success towards applications in the realm of the Langlands program, for instance in the work toward Zhang’s arithmetic fundamental lemma [RTZ13], as well as the theory of non-archimedean uniformization of Shimura varieties, first pioneered by Cerednik 9 and Drinfeld, and further developed by Rapoport–Zink [RZ96] and Howard–Pappas [HP17]. However, the dimension problem and more qualitative structural information have remained elusive for general elements of B(G, µ). Standing at this juncture, it is therefore natural to look at the other extreme and ask: Question 1.2.3. What is the dimension of X(µ, b) for b = bmax := maxB(G, µ)? Note that if G is quasi-split, this maximal element is just the Galois average µ⋄; hence, the dimension is trivially zero. However, for non-quasi-split groups the description of this maximal element is more subtle and indirect, see [HN18]. However, this is related to Question 1.1.1, as bmax = max x∈W btxµ . Utilizing a recent result in [He21a] expressing the generic Newton point in terms of the Demazure power, we can explicitly describe this maximal element for non-quasi-split groups. Furthermore, combining quantum Bruhat graph theoretic considerations with explicit computation of certain minimal length elements in Frobenius-twisted conjugacy classes of W , we prove the following in Section 4.2.5: Theorem 1.2.4. Assume that the image µad of µ in X∗(Tad)Γ0 has depth at least 2 in every F̄ -simple factor of Gad. Let b = bmax be the maximal element of B(G, µ). Then dimX(µ, b) = rkFssG ∗ − rkFssG, where G∗ is the unique (up to isomorphism) quasi-split inner form of G, and rkFss stands for semisimple F -rank. Note that if G ≃ G∗ is quasi-split, this indeed recovers the trivial case mentioned above, and as such, this theorem also serves as a geometric way to measure how far G is 10 from being quasi-split. A surprising feature of the result is that this dimension does not depend on µ. 1.3 Other related work on affine Weyl groups As is already evidenced from our discussion so far, understanding fundamental prop- erties of the affine Weyl groups is crucial to solving problems related to affine Deligne- Lusztig varieties. Indeed, the key ingredient in Milićević’s original approach to solving Question 1.1.1 was to establish that under a lower bound of order O(n2) on the depth, paths in quantum Bruhat graph encode saturated chains in the Bruhat order on the affine Weyl group. Even though we can only strengthen her result toward Question 1.1.1 to O(n) depth hypothesis, we improve this result on the affine Weyl group to one with an O(1) depth hypothesis. The next result follows from Theorem 3.4.1. Theorem 1.3.1. Suppose that w = xtλy is an element of W̃ such that depth(λ) is bigger than a certain constant (at most 6 in all Cartan types). Then we have a precise description of the cocovers of w, i.e. those elements w′ such that w′ < w and ℓ(w′) = ℓ(w) − 1: such covering relations in affine Bruhat order are in two-to-one correspondence with edges in the quantum Bruhat graph: {cocovers of w = xtλy} ←→ {edges coming into x} ∐ {edges going out of y−1}. The content appearing in Chapter 3, Section 4.1 and Chapter 5 corresponds to the paper [Sad23], while the arxiv preprint [Sad22] makes up the rest of Chapter 4. 11 Chapter 2: Preliminaries 2.1 Notations Let G be a connected reductive group over a non-archimedean local field F . Let F̆ be the completion of the maximal unramified extension of F and σ be the Frobenius morphism of F̆ /F . The residue field of F is a finite field Fq and the residue field of F̆ is the algebraically closed field F̄q. We write Ğ for G(F̆ ). We use the same symbol σ for the induced Frobenius morphism on Ğ. Let S be a maximal F̆ -split torus of G defined over F , which contains a maximal F -split torus. Let A be the apartment of GF̆ corresponding to SF̆ . We fix a σ-stable alcove a in A, and let Ĭ ⊂ Ğ be the Iwahori subgroup corresponding to a. Then Ĭ is σ-stable. Let T be the centralizer of S in G. Then T is a maximal torus. We denote by N the normalizer of T in G. The Iwahori–Weyl group (associated to S) is defined as W̃ = N(F̆ )/T (F̆ ) ∩ Ĭ. For any w ∈ W̃ , we choose a representative ẇ in N(F̆ ); however if there is no possibility of confusion we will call the lift w too. The action σ on Ğ induces a natural action of σ on W̃ , which we still denote by σ. We will sometimes identify the element w ∈ W̃ with wa, 12 the (extended) alcove that one obtains as image of the base alcove a under w. Similarly, we may say w lies in some given chamber to express that wa is an alcove belonging to that chamber. We denote by ℓ the length function on W̃ determined by the base alcove a and denote by S̃ the set of simple reflections in W̃ . Let Wa be the subgroup of W̃ generated by S̃. Then Waff is an affine Weyl group. Let Ω ⊂ W̃ be the subgroup of length-zero elements (or equivalently, the stabilizer of a) in W̃ . Then W̃ = Wa ⋊ Ω. Since the length function is compatible with the σ-action, the semi-direct product decom- position W̃ = Wa ⋊ Ω is also stable under the action of σ. Note that Wa is the Coxeter group associated to S̃, and hence it comes equipped with an associated Bruhat order, which we denote by ≤. This is extended to W̃ as follows. For two elements w̃1, w̃2 ∈ W̃ , use the above decomposition to write w̃i = wiζi with wi ∈ Wa, ζi ∈ Ω for i = 1, 2. We then declare w̃1 ≤ w̃2 if w1 ≤ w2 and ζ1 = ζ2. The following properties of ℓ and ≤ are well-known, e.g. see [Mil21, Lemma 4.1], [Kna02, exercise 23, Chapter 2] and [BB05, exercise 21, Chapter 2 ] respectively. • Let w = utλv ∈ W̃ , where λ ∈ X∗(T ) is regular dominant and u, v ∈ W . Then ℓ(w) = ℓ(u) + ℓ(tλ)− ℓ(v) = ℓ(u) + ⟨2ρ, λ⟩ − ℓ(v). (2.1.1) • For any element x of W , let Inv(x) = {α ∈ Φ+ : xα ∈ −Φ+} be its inversion set, 13 and denote its complement by Inv(x)c, i.e. Inv(x)c = Φ+ \ Inv(x). Then for any two elements x, y ∈ W , we have ℓ(xy) = ℓ(x)+ℓ(y)−2|Inv(x)∩Inv(y−1)| = ℓ(x)−ℓ(y)+2|Inv(x)c∩Inv(y−1)|. (2.1.2) • Let w1, w2, v ∈ W̃ be three elements such that ℓ(wiv) = ℓ(wi) + ℓ(v) for i = 1, 2. Then w1 ≥ w2 is equivalent to w1v ≥ w2v. (2.1.3) Let W = N(F̆ )/T (F̆ ) be the relative Weyl group. We denote by S the subset of S̃ consisting of simple reflections generating W . We let Φ (resp. ∆) denote the set of roots (resp. simple roots) for W . We write Γ for Gal(F̄ /F ), and write Γ0 for the inertia subgroup of Γ. Then fixing a special vertex of the base alcove a, we have the splitting W̃ = X∗(T )Γ0 ⋊W = {tλw;λ ∈ X∗(T )Γ0 , w ∈ W}. When considering an element λ ∈ X∗(T )Γ0 as an element of W̃ , we write tλ. Note that if G is not quasi-split over F , then there does not exist a σ-stable special vertex in a and thus the splitting W̃ = X∗(T )Γ0 ⋊W is not σ-stable. For an irreducible Weyl group W of rank n, we follow the labeling of roots as in [Bou02] and we usually write si instead of sαi , where ∆ = {αi : 1 ≤ i ≤ n}. Let w0 be the longest element in W , and for i ∈ [1, n] we let wi,0 be the longest element of the parabolic subgroup of W corresponding to ∆ \ {αi}. Let ρ be the dominant weight with 14 ⟨α∨, ρ⟩ = 1 for any α ∈ ∆. Let {ϖ∨ i : 1 ≤ i ≤ n} be the set of fundamental coweights. If ϖ∨ i is minuscule, we denote the image of tϖ ∨ i under the projection W̃ → Ω by τi; then conjugation by τi is a length preserving automorphism of W̃ , which we denote by Ad(τi). 2.2 The σ-conjugacy classes of Ğ We say that two elements b, b′ ∈ Ğ are σ-conjugate if there is some g ∈ Ğ such that b′ = gbσ(g)−1. Let B(G) be the set of σ-conjugacy classes on Ğ. By the work of Kottwitz in [Kot85] and [Kot97], any σ-conjugacy class [b] is determined by two invariants: • The Kottwitz point κ([b]) ∈ π1(G)Γ, where π1(G) := X∗(T )/ZΦ∨ is the Borovoi fundamental group and π1(G)Γ is the set of Γ-coinvariants in π1(G); • The Newton point ν([b]) ∈ ((X∗(T )Γ0,Q) +)⟨σ⟩, where X∗(T )Γ0,Q := X∗(T )Γ0 ⊗ Q = X∗(T ) Γ0 ⊗ Q, and ((X∗(T )Γ0,Q) +)⟨σ⟩ is defined to be the set of ⟨σ⟩-invariants of the intersection of X∗(T )Γ0,Q with the set X∗(T ) + Q of dominant elements in X∗(T )Q. We denote by ⩽ the dominance order on X∗(T ) + Q, i.e., for ν, ν ′ ∈ X∗(T ) + Q, we have ν ⩽ ν ′ if and only if ν ′−ν is a non-negative (rational) linear combination of positive coroots over F̆ . The dominance order on X∗(T ) + Q extends to a partial order on B(G). Namely, for [b], [b′] ∈ B(G), we say that [b] ⩽ [b′] if and only if κ([b]) = κ([b′]) and ν([b]) ⩽ ν([b′]). Denote by Jb the σ-centralizer group of b; this is a reductive group over F with F -rational points given by Jb(F ) = {g ∈ Ğ | g−1bσ(g) = b}. 15 For any reductive group H over F , we denote by rkssFH the semisimple F -rank of H. The following result is implicit in [Kot06, §1.9]. Proposition 2.2.1. Let G be quasi-simple over F and assume τ ∈ Ω. Then rkssF Jτ̇ =| Ad τ ◦ σ orbits on S̃ | −1. 2.2.1 The straight σ-conjugacy classes Note that the action of σ on Ğ gives rise to an action on W̃ , still denoted by σ. The set of σ-conjugacy classes in W̃ is denoted by B(W̃ , σ). Let w ∈ W̃ , n ∈ N. The n-th σ-twisted power of w is defined by wσ,n = wσ(w)σ2(w) · · ·σn−1(w). Note that this is nothing but the image of the n-th power of wσ ∈ W̃ ⋊ ⟨σ⟩ under the quotient map W̃ ⋊ ⟨σ⟩ → W̃ , since (wσ)n = wσ,nσn. Then by definition, w is called a σ-straight element if ℓ(wσ,n) = nℓ(w), for all n ∈ N. A σ-conjugacy class of W̃ is called straight if it contains a σ-straight element, and we denote the collection of such straight σ-conjugacy classes by B(W̃ , σ)str. We have a map Ψ : B(W̃ , σ) → B(G), coming from the assignment w → ẇ. We can also define Kottwitz and Newton maps, denoted by the same symbols κ, ν resp. from B(W̃ , σ) with the same targets as before, cf. [He14, §1.7], [Gör+10, §7.2]. The importance of the straight σ-conjugacy classes is illustrated in the following result. 16 Theorem 2.2.2. [He14, Theorem 3.7] The restriction of Ψ induces a bijective map Ψ : B(W̃ )str → B(G). Moreover, We have the following commutative diagram B(W̃ )str Ψ // (κ,ν) (( B(G) (κ,ν)uu π1(G)Γ × ((X∗(T )Γ0,Q) +)⟨σ⟩ Given w ∈ W̃ , define B(G)w := {[b] ∈ B(G) : [b] ∩ ĬwĬ ̸= ∅}. It is easy to see that B(G)w has an unique maximal element [bw], which coincides with the generic σ-conjugacy class in ĬwĬ. We will write νw to denote ν([bw]). Let {µ} be a conjugacy class of cocharacters over F̄ . Choose µ be a dominant repre- sentative of {µ} and denote by µ its image in X∗(T )Γ0 . We also have the set of neutrally acceptable elements, cf. [KR03] B(G, µ) = {[b] ∈ B(G) | κ([b]) = µ♮, ν([b]) ≤ µ⋄}. Here µ♮ denotes the common image of µ ∈ {µ} in π1(G)Γ, and µ⋄ denotes the average of the σ0-orbit of µ. Here σ0 denotes the L-action of σ, see [GHN20, definition 2.1] for details. The set B(G, µ) inherits a partial order from B(G). Since the Kottwitz map κ is constant on B(G, µ), we may view it as a subset of X∗(T )Γ0,Q via the Newton map. The following description of B(G, µ) is obtained in [HN18, Theorem 1.1 & Lemma 2.5]. To state the 17 result, for each σ0 orbit O on S, we set ϖO = ∑ i∈O ϖi. Theorem 2.2.3. [HN18] (1) Let v ∈ X∗(T )Γ0,Q. Then v ∈ B(G, µ) if and only if σ0(v) = v is dominant, and for any σ0-orbit O on S with ⟨v, αi⟩ ≠ 0 for each (or equivalently, some) i ∈ O, we have that ⟨µ+ σ(0)− v,ϖO⟩ ∈ Z and ⟨µ− v,ϖO⟩ ≥ 0. (2) The set B(G, µ) contains a unique maximal element. 2.3 Affine Deligne-Lusztig varieties For b ∈ Ğ and w ∈ W̃ , the associated affine Deligne-Lusztig variety Xw(b) is a locally closed, reduced subscheme locally of finite type inside the affine flag variety F lG, with geometric points given by Xw(b)(κ̄F ) := {gĬ : g−1bσ(g) ∈ ĬwĬ} ⊂ F lG(κ̄F ) = Ğ/Ĭ. We will also discuss certain finite unions of affine Deligne-Lusztig varieties. As before, we denote by µ the image of a dominant cocharacter µ in X∗(T )Γ0 Following [Rap05], we then define the associated admissible set as Adm(µ) = {w ∈ W̃ : w ≤ tx(µ) for some x ∈ W}. We will need the following additivity property of admissible sets. Theorem 2.3.1. [He16a, Theorem 5.1][HH17, Theorem 1.4] Let µ, µ′ ∈ X∗(T ) be domi- nant. Then we have Adm(µ) · Adm(µ′) = Adm(µ+ µ′). 18 For any b ∈ Ğ, we set X(µ, b) = ⋃ w∈Adm(µ) Xw(b). We remark that Xw(b) and X(µ, b) are subschemes of the affine flag variety in the usual sense in equal characteristic, and in the sense of Zhu [Zhu17], Bhatt and Scholze [BS17] in mixed characteristic. Settling the Kottwitz-Rapoport conjecture made in [KR03] and [Rap05] about the non-emptiness pattern for X(µ, b), He proves the following result in [He16a]. Theorem 2.3.2. [He16a, Theorem A] X(µ, b) ̸= ∅ if and only if [b] ∈ B(G, µ). 2.3.1 Virtual dimension of affine Deligne-Lusztig variety Suppose that G is quasi-split; then σ preserves the set of finite simple roots and hence acts on W . Note that w ∈ W̃ can be written in a unique way as w = utλv with λ dominant, u, v ∈ W such that tλv is a minimum length element in the right W -coset in W̃ determined by w. In this case, we set ησ(w) = σ−1(v)u. Let b ∈ Ğ. The defect of b is defined by defG(b) = rankF G− rankF Jb. Following [He14, §10.1], we then define the virtual dimension for Xw(b) to be dw(b) = 1 2 ( ℓ(w) + ℓ(η(w))− defG(b) ) − ⟨ρ, ν([b])⟩. The justification of defining such an expression lies in a result proved by He in [He14, Corollary 10.4] that says dimXw(b) ≤ dw(b) whenever κ(w) = κ([b]). Recent work by Milićević and Viehmann in [MV20] singles out those w ∈ W̃ for which dimXw(bw) = dw(bw) 19 holds; these are called cordial elements. It is shown in loc. sit. that B(G)w exhibits remarkable properties for such w, see [MV20, Theorem 1.1] for a discussion of this. 2.4 Demazure product and its variations We now discuss three operations ∗,�,� : W̃ × W̃ → W̃ . Here ∗ is the Demazure product and �,� are the left and right downward Demazure products, respectively. We describe these operations in form of the following lemma. Lemma 2.4.1. [HL15, Section 2.1] Let x, y ∈ W̃ . 1. The subset {uv : u ≤ x, v ≤ y} contains a unique maximal element, which we denote by x ∗ y. Moreover, x ∗ y = u′y = xv′ for some u′ ≤ x and v′ ≤ y and ℓ(x ∗ y) = ℓ(u′) + ℓ(y) = ℓ(x) + ℓ(v′). 2. The subset {uy : u ≤ x} contains a unique minimal element, which we denote by x� y. Moreover, x� y = u′′y for some u′′ ≤ x with ℓ(x� y) = ℓ(y)− ℓ(u′′). 3. The subset {xv : v ≤ y} contains a unique minimal element, which we denote by x� y. Moreover, x� y = xv′′ for some v′′ ≤ y with ℓ(x� y) = ℓ(x)− ℓ(v′′). The Demazure product is related to product of closure of two Iwahori double cosets in Ğ. More precisely, for any w ∈ W̃ , ĬwĬ = ⋃ w′≤w Ĭw′Ĭ is a closed admissible subset of G in the sense of [He16a]. Then for any x, y ∈ W̃ , we have ĬxĬ · ĬyĬ = ⋃ u≤x ĬuĬ · ⋃ v≤y ĬvĬ = ⋃ w≤x∗y ĬwĬ = Ĭ(x ∗ y)Ĭ . 20 The above equation also implies that ∗ is an associative binary operation on W̃ . We will need the following results about these operations. Lemma 2.4.2. [He09, Lemma 2 and Lemma 3] 1. If w ≥ w′, v ≤ v′ are elements of W̃ , then w � v ≤ w′ � v′. 2. For any three elements x, y, z ∈ W̃ , we have x� (y� z) = (x∗y)� z. In other words, the action (W̃ , ∗)× W̃ → W̃ , (x, y)→ x� y is a left action of the monoid (W̃ , ∗). 2.5 Quantum Bruhat graph We recall the quantum Bruhat graph introduced by Brenti, Fomin and Postnikov in [BFP99]. By definition, a quantum Bruhat graph ΓΦ is a directed graph with • vertices given by the elements of W ; • upward edges w ⇀ wsα for some α ∈ Φ+ with ℓ(wsα) = ℓ(w) + 1; • downward edges w ⇁ wsα for some α ∈ Φ+ with ℓ(wsα) = ℓ(w)− ⟨2ρ, α∨⟩+ 1. Note that by [BFP99, Lemma 4.3], ℓ(sα) ≤ ⟨2ρ, α∨⟩ − 1 for any α ∈ Φ+, so we have ℓ(wsα) ≥ ℓ(w)− ℓ(sα) ≥ ℓ(w)− ⟨2ρ, α∨⟩+ 1. Therefore the condition for downward edges can be rephrased to saying that ℓ(wsα) = ℓ(w)− ℓ(sα) with ℓ(sα) = ⟨2ρ, α∨⟩ − 1. Following [Len+15], we call α ∈ Φ+ to be a quantum root if ℓ(sα) = ⟨2ρ, α∨⟩− 1. We have the following description of quantum roots. 21 Lemma 2.5.1. [Len+15, Lemma 4.2] We have that α ∈ Φ+ is a quantum root if and only if 1. α is a long root, or 2. α is a short root, and if α = ∑ αi∈∆ ciαi then we have ci = 0 for any long simple root αi. Here for simply laced root systems we consider all roots to be long. Thus in a simply laced type, all roots are quantum. Examples of quantum root, in general, include the simple roots as well as the highest root. We now recall some graph theoretic notions related to ΓΦ; we refer to [Mil21, section 3.1] for more details. The weight of an upward edge is defined to be 0 and the weight of a downward edge w ⇁ wsα is defined to be α∨. For two elements w,w′, a directed path between them is defined to be concatenation of directed edges joining a sequence of vertices in ΓΦ. The weight of a path in ΓΦ is defined to be the sum of weights of the edges in the path. The length of path is defined to be the number of edges present in it. For any x, y ∈ W , we denote by dΓ(x, y) the minimal length among all paths in ΓΦ from x to y. Any path between x and y affording dΓ(x, y) as its length is called a shortest path between them. Lemma 2.5.2. [Pos05, Lemma 1], [BFP99, Lemma 6.7] Let x, y ∈ W . Then 1. There exists a directed path (consisting of possibly both upward and downward edges) in ΓΦ from x to y. 22 2. Any two shortest paths in ΓΦ from x to y have the same weight, which we denote by wt(x, y). 3. Any path in ΓΦ from x to y has weight ⩾ wt(x, y). 23 Chapter 3: Generic Newton point and affine Bruhat order For a split connected reductive group, an explicit description of νw was first derived by Milićević in [Mil21, Theorem 3.2] for elements w ∈ W̃ with superregular dominant translation part. More precisely, a uniform bound is given in [Mil21, Corollary 3.3] in the quasi-simple case, which says that the description of νw in loc. sit. is valid whenever the following depth hypothesis is satisfied by λ: depth(λ) ≥  8ℓ(w0), if G is of classical type; 16ℓ(w0), if G is of exceptional type. To establish this result, Milićević first derives in [Mil21, Proposition 4.2] a characteriza- tion of the covering relations in W̃ , again under certain superregularity hypothesis of the following nature: depth(λ) ≥  2ℓ(w0) + 2, if G is not of type G2; 3ℓ(w0) + 3, if G is of type G2. This characterization is of independent interest, and more recently it has been utilized by He and Yu in [HY21] to deduce a partial description of the admissible sets in W̃ . In this 24 chapter, we improve upon these results by weakening the superregularity hypothesis to prove them. 3.1 Some combinatorial properties We start by proving a monotonicity property for the weight function. Let us define the function wt : W → X∗(T ) by wt(x) := wt(x, 1). Lemma 3.1.1. Let x1, x2 ∈ W ; if x1 ≤ x2 in Bruhat order, then wt(x1) ≤ wt(x2) in dominance order. Proof. It suffices to show this when x2 is a cover of x1, i.e. x2 = x1sα for some α ∈ Φ+ such that ℓ(x2) = ℓ(x1)+1. We construct a path from x1 to 1 by concatenating the upward edge x1 ⇀ x2 with a path of shortest length from x2 to 1. Since the first upward edge has weight 0, the weight of this particular path is wt(x2). Appealing to Lemma 2.5.2(3), we then get wt(x2) ≥ wt(x1). We will now reduce our calculation of the maximal Newton point of an arbitrary element of W̃ to that of a suitable element lying in the dominant chamber. Proposition 3.1.2. If λ is dominant regular, then [butλv] = [btλ(v�σ(u)]. Proof. Note that by Equation (2.1.1), ℓ(utλv) = ℓ(u)+ ℓ(tλv), hence utλv = u∗ (tλv). Thus ĬutλvĬ = ĬuĬ · ĬtλvĬ. Now, recall that Ĭ is σ-stable. Therefore, if w = w1w2 is an element with w1 ∈ ĬuĬ, w2 ∈ ĬtλvĬ, then w−1 1 ·σw = w2σ(w1). This shows that Ğ ·σ (ĬuĬ · ĬtλvĬ) ⊂ Ğ ·σ (ĬtλvĬ · 25 Ĭσ(u)Ĭ). One obtains the other inclusion similarly. This shows that Ğ ·σ (ĬuĬ · ĬtλvĬ) = Ğ ·σ (ĬtλvĬ · ĬuĬ). Hence, we conclude that Ğ ·σ ĬutλvĬ = Ğ ·σ Ĭ(tλv) ∗ σ(u)Ĭ . Since G ·σ ĬwĬ is the set of σ-conjugacy classes intersecting ĬwĬ, by appealing to [Vie14, Corollary 5.6] we get that [butλv] = [b(tλv)∗σ(u)]. (3.1.1) Thus it suffices to show that (a) if λ is dominant regular, then (tλv) ∗ u′ = tλ(v � u′) for elements u′, v ∈ W . By Lemma 2.4.2(2), we only need to argue the case when u′ is a simple reflection s ∈ S. Then this statement is equivalent to the assertion that ℓ(vs) = ℓ(v)− 1 if and only if ℓ(tλvs) = ℓ(tλv) + 1. This follows directly from a computation using Equation (2.1.1). Hence we are done. Corollary 3.1.3. Let x, y ∈ W . Then wt(x, y) = wt(x−1 � y). Proof. Suppose that λ ∈ X∗(T ) + is superregular. Applying the formula for maximal New- ton point established in [Mil21, Theorem 3.2] to w := xtλy and w′ := tλ(y � x), we get that νw = λ− wt(y−1, x) and νw′ = λ− wt((y � x)−1) = λ− wt(y � x). Since νw = νw′ by Proposition 3.1.2, we get wt(y−1, x) = wt(y�x). Now replacing the pair (x, y) by (y, x−1), we get the conclusion. We now close this section by giving another interpretation of the weight of a shortest path in the quantum Bruhat graph. 26 Definition 3.1.4. For an element x ∈ W , we say that x = sβ1 · · · sβk is a reduced quantum reflection decomposition if 1. {βi : 1 ≤ i ≤ k} is a collection of (not necessarily simple) quantum roots, i.e. l(sβi ) = ⟨2ρ, β∨ i ⟩ − 1. 2. ℓ(x) = ∑k i=1 ℓ(sβi ). 3. Subject to the first two conditions, k is minimal. In this case, we say that k is the minimal reduced length of x in terms of reflections associated to quantum roots, and write ℓ↓(x) = k. The choice of notation is suggested from the fact that {Reduced quantum reflection decompositions for x}xy {Shortest paths in ΓΦ from x to 1 that uses only downward edges} By [MV20, Proposition 4.11], the weight of any path in the latter set is equal to wt(x) and dΓ(x, 1) = ℓ↓(x). Therefore, the problem of computing wt(x) reduces to one of finding a suit- able decomposition for x: if x = sβ1 · · · sβk is such a decomposition then wt(x) = ∑k i=1 β ∨ i . In view of Corollary 3.1.3, we can thus reformulate wt(x, y) in terms of a decomposition of x−1 � y. 3.2 Formula for the generic Newton point The goal of this section is to prove the following result. 27 Theorem 3.2.1. Assume that G is a quasi-simple split group of rank n. Let w = tλx be an element of its Iwahori-Weyl group such that λ is dominant and depth(λ) > Ξ, where Ξ is an integer depending on n. Then the maximal Newton point associated to w is given by νw = λ− wt(x). For each Cartan type, the lower bound Ξ for the depth mentioned above is given in the following table. Type An Bn/Cn Dn E6 E7 E8 F4 G2 Ξ 3n+ 1 6n− 2 6n− 6 23 33 57 23 9 Remark 3.2.2. The statement of Theorem 3.2.1 seems weaker than the statement of Theorem 1.1.2 in the sense that the former gives a formula only for certain elements of W̃ (associated to a quasi-simple split group G) in the dominant chamber for whereas the latter promises to handle elements in arbitrary chambers (for arbitrary split groups). Hence let us first explain how Theorem 3.2.1 implies Theorem 1.1.2 - for quasi-simple groups; in the next remark, we further remove this condition on the group. By Proposition 3.1.2, we know that νutλv = νtλ(v�σ(u). Under the hypothesis on λ, this element tλ(v � σ(u)) lies in the dominant chamber. Then by Theorem 3.2.1 it follows that νtλ(v�σ(u) = λ− wt(v � σ(u)). By Corollary 3.1.3 we finally deduce that νutλv = λ − wt(v−1, σ(u)). This concludes our 28 discussion. Remark 3.2.3. One can handle the case of a connected reductive split group G via a routine reduction procedure, which we discuss now. In that case, we have W = W1 × · · · ×Wl, where Wi are irreducible Weyl groups associated to F -simple factors of Gad. This gives rise to a partition of m, where m is the semisimple rank of G. Any element x of W is of the form (x1, · · · , xl). Similarly, we write λ as (λ1, · · · , λl). We have w0 = (w0,1, · · · , w0,l), where w0,i is the longest element of Wi. In the same way, we write the sum of positive roots as (2ρ∨1 , · · · , 2ρ∨l ) and the highest root as (θ1, · · · , θl). We observe that while dealing with the quasi-simple case, we introduce a restriction on depth in two stages - as found in Section 3.2.1 and Section 3.2.4. By the last paragraph we therefore need to require for each i depth(λi) ≥ Ξi, where Ξi is the lower bound given for the Weyl group Wi, to be read off from the table above. Note that Ξi is then some linear expression of m, depending on the type of i-th factor and the partition of m. We remark that under such hypotheses, Theorem 3.2.1 applies to each quasi-simple factor of Gad and thus we need to justify the formula for such groups. Remark 3.2.4. Let us note that Theorem 3.2.1 has since been strengthened due to the 29 work of multiple authors, by weakening the bound on depth of the relevant coweight. In [HN21], He and Nie prove the formula for generic Newton point under the assumption that the relevant cocharacter is of depth at least 2. Finally, Schremmer completely solves the problem for arbitrary cocharacter in [Sch22b] by removing the hypothesis on depth, albeit the description gets more complicated in this situation. We now give a technical layout of the proof. As mentioned before, under the regularity assumption on its associated dominant coweight as in Proposition 3.1.2, we only need to handle elements w ∈ W̃ that lie in the dominant chamber. In Proposition 3.2.6, we show that w is bigger than a certain translation element, thus giving us a lower bound for νw. This holds under a certain depth hypothesis proposed in Section 3.2.1, which relies on explicit calculation in Section 3.3 of wt(w0) for the longest element w0 in each Cartan type. Using this lower bound, we also show in Lemma 3.2.7 that νw is given by Newton point of some translation tγ smaller than w. Note that if this translation can be shown to be lying in the dominant chamber (i.e. γ is dominant), we are practically done - as one can then multiply the inequality tλy ≥ tγ by a large enough dominant translation element, thereby making the translation parts in the resulting elements superregular and hence suitable for applying the existing formula for νw in [Mil21, Theorem 3.2]. However, we cannot show this and rather bypass this problem by converting the Bruhat inequality tλy ≥ tγ to one involving elements whose translation parts are dominant. This is done in Lemma 3.2.9 by applying techniques from the theory of the Demazure product on W̃ . Then we can carry out the aforementioned trick of artificially adding a large translation part, applying [Mil21, Theorem 3.2] and then finally subtracting it - as shown in Section 4.2.5. 30 3.2.1 Proposed linear bound on depth For the rest of Section 3.2, we assume that W is an irreducible Weyl group for the root system of Cartan type Xn. We first establish an upper bound for M =MXn := max{⟨αi,wt(x)⟩ : αi ∈ ∆, x ∈ W} This is based on an explicit computation of wt(w0) for the longest element w0 in each type, which we defer to the next section for the sake of continuity. By Lemma 3.1.1, we have {wt(x) : x ∈ W} ⊂ {ν ∈ ZΦ∨ : ν ≤ wt(w0)}. Therefore, M≤ max{⟨αi, ν⟩ : ν ∈ ZΦ∨, ν ≤ wt(w0), αi ∈ ∆}. It is easy to see that maximum of the latter set is realized at ν = κlα ∨ l , where κlα ∨ l is a summand of wt(w0) such that κl = max{κj : κjα ∨ j is a summand of wt(w0)}. Since the maximum value of ⟨αi, α ∨ l ⟩ is 2 in every Cartan type, this shows that MXn is bounded above by twice the largest coefficient in the expression of wt(w0) in each type Xn. This is tabulated in the list below, where we denote by M̃ = M̃Xn the upper bound obtained in this manner. Type An Bn/Cn Dn E6 E7 E8 F4 G2 M̃ n+ 1 2n 2n 12 16 28 12 4 31 3.2.2 Proof of the easier inequality We need the following result about Bruhat order on W̃ . Lemma 3.2.5. [Rap05, Remark 3.9] Let λ ∈ X∗(T ) be dominant. Let β ∈ Φ+ such that λ− β∨ is dominant. Then tλ−β∨ ≤ tλsβ ≤ tλ in W̃ . We start by establishing the easier inequality. Proposition 3.2.6. Let w = tλx ∈ W̃ such that depth(λ) ≥ M̃. Then w ≥ tλ−wt(x) and therefore νw ≥ λ− wt(x). Proof. Suppose that x = sβ1 · · · sβl is a reduced quantum reflection decomposition with ℓ↓(x) = l > 1, and therefore wt(x) = l∑ i=1 β∨ i . Let us first note that: (a) sβ1 · · · sβk is a reduced quantum reflection decomposition for xk := xsβl · · · sβk+1 for all k ≤ l. Suppose otherwise; then we can find a reduced quantum reflection decomposition of the form xk = sγ1 · · · sγj with j < k. But then x = xksβk+1 · · · sβl = sγ1 · · · sγjsβk+1 · · · sβl satisfies length additivity: ℓ(x) = ℓ(xk) + l∑ i=k+1 β∨ i = j∑ i=1 γ∨ i + l∑ i=k+1 β∨ i . Note that all associated roots in this new decomposition of x are quantum, and it has j + l − k < l factors, hence it contradicts the fact that ℓ↓(x) = l. So statement (a) is proved. 32 To prove the proposition, we argue by induction on ℓ↓(x). We note our depth hy- pothesis ensures that λ−β∨ is dominant for all β ∈ Φ+: for αi ∈ ∆, we have ⟨αi, λ−β∨⟩ = ⟨αi, λ⟩−⟨αi, β ∨⟩ > 0, since maximum value of ⟨αi, β ∨⟩ equals 2 in every Cartan type, except for G2 - in which case it equals 3, cf. [Bou02, Chapter VI, Section 1, no. 3]. Therefore Lemma 3.2.5 covers the base case, i.e. when x = sβ for some β ∈ Φ+. Let us now apply the induction hypothesis to xsβl = sβ1 · · · sβl−1 . Since wt(xsβl ) = l−1∑ i=1 β∨ i by virtue of statement (a), this gives tλsβ1 · · · sβl−1 ≥ t λ− l−1∑ i=1 β∨ i . (3.2.1) To perform the induction step, we apply a property of Bruhat order as described in Equa- tion (2.1.3). We set w1 = tλx,w2 = t λ− l−1∑ i=1 β∨ i sβl , v = sβl and check the required length additivity. Note that λ− l−1∑ i=1 β∨ i is dominant under our depth hypothesis: for αi ∈ ∆, ⟨αi, λ− l−1∑ i=1 β∨ i ⟩ = ⟨αi, λ⟩ − ⟨αi,wt(xsβl )⟩ > ⟨αi, λ⟩ − M̃ ≥ 0, therefore the length formula in Equation (2.1.1) applies. We see that 1. ℓ(w1v) = ℓ(tλsβ1 · · · sβl−1 ) = ℓ(tλ)− l−1∑ i=1 ℓ(sβi ) = ℓ(tλ)− l∑ i=1 ℓ(sβi )+ℓ(sβl ) = ℓ(w1)+ℓ(v). 2. ℓ(w2v) = ℓ(t λ− l−1∑ i=1 β∨ i ) = ℓ(t λ− l−1∑ i=1 β∨ i )− ℓ(sβl ) + ℓ(sβl ) = ℓ(w2) + ℓ(v). In presence of Equation (3.2.1), we therefore get w1 ≥ w2, i.e. tλx ≥ t λ− l−1∑ i=1 β∨ i sβl . Finally, 33 we note that λ−wt(x) = λ− l∑ i=1 β∨ i is dominant (by similar argument as before) and thus Lemma 3.2.5 applies to give t λ− l−1∑ i=1 β∨ i sβl ≥ t λ− l∑ i=1 β∨ i . Combining these inequalities, we get tλx ≥ tλ−wt(x). By [Vie14, Corollary 5.6], we have νw = max{ν(u) : u ≤ w, u ∈ W̃}. (3.2.2) Appealing to Equation (3.2.2), we thus get νw ≥ λ− wt(x). This finishes the proof. Lemma 3.2.7. Let w = tλx ∈ W̃ such that depth(λ) > M̃. Then νw = max{γ+ : tγ ≤ w}. Proof. By Equation (3.2.2), we can assume that νw = ν(tµz) for some tµz ∈ W̃ = X∗(T )⋊ W , such that w ≥ tµz. Suppose that z ̸= 1, therefore the order of z equals m ≥ 2. Then ν(tµz) = ( 1 m m∑ i=1 zi(µ))+. Note that z · m∑ i=1 zi(µ) = m∑ i=1 zi(µ), so the element m∑ i=1 zi(µ) lies on the wall of some Weyl chamber, cf. [Bou02, Chapter V, Section 3.3, Remark 3]. Therefore ( 1 m m∑ i=1 zi(µ))+ lies on the wall of C+, and hence νw is singular. By Proposition 3.2.6, we can assume that νw = λ − ϵ, with ϵ ≤ wt(x) ≤ wt(w0). Hence, there is some αi ∈ ∆ such that ⟨αi, λ − ϵ⟩ = 0, and thus ⟨ai, λ⟩ = ⟨αi, ϵ⟩ ≤ M̃ , which is a contradiction to hypothesis on depth(λ). Therefore, νw cannot be singular and hence z = 1. 3.2.3 Toward computing a downward Demazure product We need to understand the largest translation dominated by w to determine its asso- ciated generic Newton point νw. In view of Proposition 3.2.6, let us denote this translation element by tyγ for some y ∈ W and γ ≥ λ− wt(x). Of course, this would be equality if λ 34 is superregular. We want to leverage this result to establish such equality with some linear bound on depth here. The following lemma lays the path towards proving that. Lemma 3.2.8. If λ−2ρ∨ is dominant, we have tλx ≥ tyγ if and only if tλ−2ρ∨ ≥ t−2ρ∨�tyγ. Proof. Note that if λ− 2ρ∨ is dominant, ℓ(tλ−2ρ∨) = ⟨2ρ, λ− 2ρ∨⟩ = ℓ(tλ)− ℓ(t−2ρ∨); also, since λ ≥ λ − 2ρ∨ are dominant coweights, tλ ≥ tλ−2ρ∨ , cf. [Rap05, proof of Proposition 3.5]. Hence, t−2ρ∨ � tλ = tλ−2ρ∨ . Now, it suffices to show the following: (a) If a, b are two elements in a Coxeter group and s is a simple reflection, then a ≥ b if and only if s� a ≥ s� b. This follows from the well known lifting property of Bruhat order. Note that we only need to consider the case when s� a = sa, i.e. a ≥ sa. Then there are two further cases. 1. s � b = b: in this case, sb ≥ b. Hence a ≥ b is equivalent to sa ≥ b, and hence s� a ≥ s� b. 2. s � b = sb: in this case, b ≥ sb. Hence a ≥ b is equivalent to sa ≥ sb, and hence s� a ≥ s� b. In Section 3.2.4 we will focus on computing the downward Demazure product that appears in Lemma 3.2.8. Before proceeding any further to do that, let us first sketch what the answer should look like. We first apply Lemma 2.4.2 with w = t−2ρ∨ , w′ = w0 and v = v′ = tyµ; if we require that λ is of large enough depth (in a sense made precise below) so that γ is dominant regular, we get w0 � ytγy−1 = tγy−1 ≥ t−2ρ∨ � tyγ . We 35 next apply Lemma 2.4.2 with v′ = tyγ, v = tγy−1 and w = w′ = t−2ρ∨ . If we now require λ to be of sufficiently large depth so that γ − 2ρ∨ is also dominant, we get t−2ρ∨ � tyγ ≥ t−2ρ∨ � tγy−1 = tγ−2ρ∨y−1. Combining these, we see that whenever λ has “large” depth (to be specified later) we have tγy−1 ≥ t−2ρ∨ � tyγ ≥ tγ−2ρ∨y−1. Therefore, it is natural to expect that this downward Demazure product depends only on y whenever λ has a “large” depth. In the next section, we quantify this depth condition and show that we indeed get the desired conclusion. 3.2.4 A technical lemma For each irreducible Cartan type Xn, we let S = SXn be twice the sum of all coef- ficients appearing in the expression of the highest root in terms of simple roots; in other words, S = ⟨θ, 2ρ∨⟩, where θ is the highest root. This quantity will become relevant for our next lemma. We record this integer for each type in the table below. Type An Bn/Cn Dn E6 E7 E8 F4 G2 S 2n 4n− 2 4n− 6 11 17 29 11 5 Note that Ξ = M̃ + S is the final lower bound that appears in Theorem 3.2.1. We now prove a key lemma that would help us achieve the harder inequality. Lemma 3.2.9. Let depth(λ) > Ξ. We continue to assume that w = tλx ≥ tyγ such that 36 γ = νw ≥ λ − wt(x). Then there exists a coweight µy depending only on y such that t−2ρ∨ � tyγ = tγ−µyy−1. Proof. We begin by noting that we can compute this downward Demazure product from a specific length additive decomposition of tyγ. More precisely, we have the following desiderata • (D1) a decomposition tyγ = a1t µ1b1 · a2tµ2b2, where µi is dominant and tµibi ∈ SW̃ for i = 1, 2; • (D2) ℓ(tyγ) = ℓ(a1t µ1b1) + ℓ(a2t µ2b2); • (D3) a1t µ1b1 is the largest element dominated by t2ρ ∨ , subject to the first two condi- tions. Lemma 2.4.1 asserts the existence of such decomposition, and then we have t−2ρ∨ � tγy = a2t µ2b2. We now proceed to identify these elements ait µibi in the following three steps. 3.2.4.1 Relating relevant coweights We first determine how the coweights associated to translation part of these elements are related. By (D1), tyγ = a1b1a2t a−1 2 b−1 1 µ1+µ2b2, so we have γ = (a−1 2 b−1 1 µ1 + µ2) +, i.e. there exists some v ∈ W such that γ = v(a−1 2 b−1 1 µ1 + µ2). We will now show that v = 1. Assume the contrary and let β ∈ Inv(v). Now, µ2 = v−1γ − a−1 2 b−1 1 µ1 is dominant, therefore ⟨β, v−1γ − a−1 2 b−1 1 µ1⟩ ≥ 0. 37 By (D3), we have t2ρ ∨ ≥ a1t µ1b1, thus 2ρ∨ ≥ µ1; by a geometric characterization of dominance order, e.g. see [AB83, Lemma 12.14], this gives µ1 = ∑ ζ∈W aζζ · 2ρ∨ for some aζ ∈ R≥0 such that ∑ ζ∈W aζ = 1. Substituting this expression of µ1 in the pairing above and using its W -invariance, we get ⟨vβ, γ⟩ − ∑ ζ∈W aζ⟨ζ−1b1a2β, 2ρ ∨⟩ ≥ 0. Let β′ = −vβ ∈ Φ+. Note that ζ−1b1a2β ≥ −θ for any element ζ ∈ W , and therefore ⟨ζ−1b1a2β, 2ρ ∨⟩ ≥ −⟨θ, 2ρ∨⟩. Putting all these together, we obtain ⟨β′, γ⟩ = − ∑ ζ∈W aζ⟨ζ−1b1a2β, 2ρ ∨⟩ ≤ ∑ ζ∈W aζ⟨θ, 2ρ∨⟩ = S. Now recall that γ = νw = λ−ϵ is dominant with ϵ ≤ wt(x) ≤ wt(w0). Choose a simple root α ≤ β′. Then we get ⟨α, γ⟩ ≤ ⟨β′, γ⟩ ≤ S, and therefore ⟨α, λ⟩ ≤ S + ⟨α, ϵ⟩ ≤ S + M̃. Since this yields a contradiction to the depth hypothesis of λ, we are done. Therefore v = 1 and γ = a−1 2 b−1 1 µ1 + µ2. 3.2.4.2 Showing that certain coweights are regular Note that γ is regular, since for any simple root αi we have ⟨αi, γ⟩ = 0 =⇒ ⟨αi, λ⟩ = ⟨αi, ϵ⟩ ≤ M̃, thereby contradicting the depth hypothesis imposed on λ. Hence, knowing that the translation parts in tyγ = a1b1a2t a−1 2 b−1 1 µ1+µ2b2 are equal allows us to relate the finite Weyl group components from both sides. Namely, we have y = a1b1a2 = b−1 2 . We next 38 simplify these relations further in the following way. Let us first observe that µ2 is regular, since otherwise we can find a simple root α such that ⟨α, γ⟩ = ∑ ζ∈W aζ⟨α, ζ ·a−1 2 b−1 1 2ρ∨⟩ =∑ ζ∈W ⟨b1a2ζα, 2ρ∨⟩ ≤ ∑ ζ∈W aζ⟨θ, 2ρ∨⟩ = S; but this would yield contradiction to the imposed depth condition as before. 3.2.4.3 Showing that the product describes a dominant chamber alcove We can now show that a2 = 1. Recall that t−2ρ∨ � tyγ = a2t µ2b2. Let us rewrite t−2ρ∨ = w0t 2ρ∨w0 and apply w0� to the former equation. This gives w0 � {(w0t 2ρ∨w0)� tyγ} = w0 � (a2t µ2b2). We now apply Lemma 2.4.2(2) on the left hand side, and use the fact that µ2 is regular in the right hand side. We get {w0 ∗ (w0t 2ρ∨w0)}� tyγ = (w0 � a2)t µ2b2. Note that w0t 2ρ∨w0 = w0 ∗ t2ρ ∨ w0, and hence we can use associativity of operation ∗ to rewrite the element in the parenthesis on the left hand side as w0 ∗ (w0 ∗ t2ρ ∨ w0) = (w0 ∗ w0) ∗ t2ρ ∨ w0 = w0 ∗ t2ρ ∨ w0 = w0t 2ρ∨w0 = t−2ρ∨ . Therefore, we finally deduce t−2ρ∨ � tyγ = tµ2b2. This proves our claim. 39 Therefore, the relations arising from (D1) are γ = b−1 1 µ1 + µ2, y = a1b1 = b−1 2 . (3.2.3) Let us now utilize (D2). It says that we have length additivity in the decomposition tyγ = a1t µ1b1 · tµ2b2, i.e. ⟨γ, 2ρ⟩ = ℓ(a1) + ⟨µ1, 2ρ⟩ − ℓ(b1) + ⟨µ2, 2ρ⟩ − ℓ(b2). By use of Equation (3.2.3), the last equation reduces to ℓ(a1b1) = ℓ(a1)− ℓ(b1) + ⟨µ1 − b−1 1 µ1, 2ρ⟩. (3.2.4) Therefore we are led to consider the following subset of W̃ : Ωy := {atµb ∈ W̃ : y = ab, atµb ≤ t2ρ ∨ , ℓ(ab) = ℓ(a)− ℓ(b) + ⟨µ− b−1µ, 2ρ⟩}. Note that the definition of Ωy encapsulates all information amongst a1, b1, µ1 coming out of the first two items in our desiderata. By (D3), the element a1t µ1b1 is therefore uniquely specified as the maximal element of Ωy. Hence, the element b−1 1 µ1 depends only on y. Denoting it by µy, we see that Equation (3.2.3) implies µ2 = γ − µy. This finishes the proof. 40 3.2.5 Finishing the proof Proof of Theorem 3.2.1. By combining Lemma 3.2.8 and Lemma 3.2.9, we observe that our standing assumption tλx ≥ tyγ implies tλ−2ρ∨x ≥ tγ−µyy−1. Let us now choose λ′ to be large enough so that λ + λ′ is superregular, e.g. take λ′ = nρ∨ for large n. We note that 1. ℓ(tλ ′ · tλ−2ρ∨x) = ℓ(tλ+λ′−2ρ∨x) = ⟨2ρ, λ+ λ′− 2ρ∨⟩ − ℓ(x) = ⟨2ρ, λ′⟩+ ⟨2ρ, λ− 2ρ∨⟩ − ℓ(x) = ℓ(tλ ′ ) + ℓ(tλ−2ρ∨x). 2. ℓ(tλ ′ · tγ−µyy−1) = ℓ(tλ ′+γ−µyy−1) = ⟨2ρ, λ′ + γ − µy⟩ − ℓ(y−1) = ⟨2ρ, λ′⟩ + ⟨2ρ, γ − µy⟩ − ℓ(y−1) = ℓ(tλ ′ ) + ℓ(tγ−µyy−1). Hence, we can apply Equation (2.1.3) in this situation, and get tλ+λ′−2ρ∨x ≥ tγ+λ′−µyy−1. Now, we apply Lemma 3.2.8 again to deduce from the previous equation tλ+λ′ x ≥ ty(γ+λ′). Therefore, by Equation (3.2.2) we have νtλ+λ′x ≥ νty(γ+λ′) . Since the formula for the maximal Newton point in [Mil21, Theorem 3.2] applies to elements of W̃ having λ+λ′ as its 41 dominant translation part, we get that λ+λ′−wt(x) ≥ γ+λ′, whence λ−wt(x) ≥ γ = νw. Combining this with Proposition 3.2.6, we get νw = λ−wt(x). This finishes the proof. 3.2.6 A remark about Bruhat order on W̃ Proposition 3.2.10. Suppose that G is a quasi-simple split group. Assume depth(λ) > Ξ. Then utλv ≥ tyγ for some dominant γ implies that tλ(v � u) ≥ tγ. Proof. Note that utλv ≥ tyγ implies νutλv ≥ νtyγ by Equation (3.2.2). By Theorem 3.2.1 and Corollary 3.1.3, we then have λ − wt(v � u) ≥ γ. By Proposition 3.2.6, we have that tλ(v � u) ≥ tλ−wt(v�u). Since λ − wt(v � u), γ are both dominant, we also have that tλ−wt(v�u) ≥ tγ, cf. [Rap05, proof of Proposition 3.5]. Putting together the last two inequalities, we finally get tλ(v � u) ≥ tγ. Remark 3.2.11. Note that if u = 1, this says that tλv ≥ tyγ implies tλv ≥ tγ. In other words, if some W -conjugate of a dominant translation element lies below an element in the dominant chamber, so does the dominant translation element itself. We point out that in the presence of Proposition 3.2.6 and the fact that maximal Newton point formula is avail- able for elements with superregular translation part, this is equivalent to Theorem 3.2.1. Hence, an independent proof of Proposition 3.2.10 would help us get rid of the bound im- posed in Section 3.2.4; similarly, an improved estimate forM would weaken the final depth hypothesis required in Theorem 3.2.1. 42 3.3 Weight of the longest element In this section, we give a reduced quantum reflection decomposition for the longest element w0 in Weyl group W of each irreducible Cartan type. For an element x ∈ W , the reflection length of x is the smallest number l such that x can be written as a product of l reflections in W . We denote by ℓR(x) the reflection length of x. Note that in general ℓR(x) ≤ ℓ↓(x), and strict inequality can occur. We enumerate reflection length for w0 for all the Cartan types below. Type An Bn/Cn Dn E6 E7 E8 F4 G2 ℓR(w0) ⌈n 2 ⌉ n 2⌊n 2 ⌋ 4 7 8 4 2 We compute wt(w0) by exhibiting suitable decomposition of w0 in each type, and on the way we see that ℓR(w0) = ℓ↓(w0). We first write down expression of the longest element affording its reflection length; for the classical types, we extract this from [Bla09], and for the exceptional ones we can check this directly by hand or using a computer algebra system such as [TheYY]. Once we find these decompositions, we see that they only involve reflections corresponding to quantum roots; furthermore, these expressions do satisfy the length additivity condition as well. Taken together, these observations establish that we have found reduced quantum reflection decomposition for w0 in these types. Finally, we add up the coroots associated to the reflections appearing in each such decompositions and deduce wt(w0) from that. See Remark 3.3.1 for an alternative recipe for computing wt(w0). 43 (a) Type An: here w0 =  sα1+···+α2k sα2+···+α2k−1 · · · sαk+αk+1 , if n = 2k; sα1+···+α2k+1 sα2+···+α2k · · · sαk+1 , if n = 2k + 1. Hence, wt(w0) =  α∨ 1 + 2α∨ 2 + · · ·+ (k − 1)α∨ k−1 + kα∨ k + kα∨ k+1 + (k − 1)α∨ k+2 + · · ·+ α∨ 2k, if n = 2k; α∨ 1 + 2α∨ 2 + · · ·+ kα∨ k + (k + 1)α∨ k+1 + kα∨ k+2 + · · ·+ α∨ 2k+1, if n = 2k + 1. (b) Type Bn: here w0 =  sα1+2α2+···+2α2k sα1sα3+2α4+···+2α2k sα3 · · · sα2k−1+2α2k sα2k−1 , if n = 2k; sα1+2α2+···+2α2k+1 sα1sα3+2α4+···+2α2k+1 sα3 · · · sα2k−1+2α2k+2α2k+1 sα2k+1 , if n = 2k + 1. Hence, wt(w0) =  2α∨ 1 + 2α∨ 2 + 4α∨ 3 + 4α∨ 4 + · · ·+ 2(k − 1)α∨ 2k−3 + 2(k − 1)α∨ 2k−2 +2kα∨ 2k−1 + kα∨ 2k, if n = 2k; 2α∨ 1 + 2α∨ 2 + 4α∨ 3 + 4α∨ 4 + · · ·+ 2kα∨ 2k−1 + 2kα∨ 2k + (k + 1)α∨ 2k+1, if n = 2k + 1. 44 (c) Type Cn: here w0 = s2α1+···+2αn−1+αns2α2+···+2αn−1+αn · · · s2αn−1+αnsαn , and thus wt(w0) = α∨ 1 + 2α∨ 2 + · · ·+ nα∨ n . (d) Type Dn: here w0 =  sα1+2α2+···+2α2k−2+α2k−1+α2k sα1sα3+2α4+···+2α2k−2+α2k−1+α2k sα3 · · · sα2k−3+2α2k−2+α2k−1+α2k sα2k−3 sα2k , if n = 2k; sα1+2α2+···+2α2k−2+α2k−1+α2k sα1sα3+2α4+···+2α2k−2+α2k−1+α2k sα3 · · · sα2k−3+2α2k−2+2α2k−1+α2k+α2k+1 sα2k−1+α2k+α2k+1 sα2k−1 , if n = 2k + 1. Hence, wt(w0) =  2α∨ 1 + 2α∨ 2 + 4α∨ 3 + 4α∨ 4 + · · ·+ (2k − 2)α∨ 2k−3 +(2k − 2)α∨ 2k−2 + kα∨ 2k−1 + kα∨ 2k, if n = 2k; 2α∨ 1 + 2α∨ 2 + 4α∨ 3 + 4α∨ 4 + · · ·+ (2k − 2)α∨ 2k−3 + (2k − 2)α∨ 2k−2 +2kα∨ 2k−1 + kα∨ 2k + kα∨ 2k+1, if n = 2k + 1. (e) Type E6: here w0 = sα1+2α2+2α3+3α4+2α5+α6sα1+α3+α4+α5+α6sα3+α4+α5sα4 , and thus wt(w0) = 2α∨ 1 + 2α∨ 2 + 4α∨ 3 + 6α∨ 4 + 4α∨ 5 + 2α∨ 6 . (f) Type E7: here w0 = s2α1+2α2+3α3+4α4+3α5+2α6+α7sα2+α3+2α4+2α5+2α6+α7sα2+α3+2α4+α5sα2 45 sα3sα5sα7 , and thus wt(w0) = 2α∨ 1 + 5α∨ 2 + 6α∨ 3 + 8α∨ 4 + 7α∨ 5 + 4α∨ 6 + 3α∨ 7 . (g) Type E8: here w0 = s2α1+3α2+4α3+6α4+5α5+4α6+3α7+2α8s2α1+2α2+3α3+4α4+3α5+2α6+α7 sα2+α3+2α4+2α5+2α6+α7sα2+α3+2α4+α5sα2sα3sα5sα7 , and thus wt(w0) = 4α∨ 1 + 8α∨ 2 + 10α∨ 3 + 14α∨ 4 + 12α∨ 5 + 8α∨ 6 + 6α∨ 7 + 2α∨ 8 . (h) Type F4: here w0 = s2α1+3α2+4α3+2α4sα2+2α3+2α4sα2+2α3sα2 , and thus wt(w0) = 2α∨ 1 + 6α∨ 2 + 4α∨ 3 + 2α∨ 4 . (i) Type G2: here w0 = s3α1+2α2sα1 , and thus wt(w0) = 2α∨ 1 + 2α∨ 2 . Remark 3.3.1. We note that the weight of longest element found above appears in a rather different context, cf. [Lus18]. In that paper, Lusztig points out that the coroots appearing as a summand of wt(w0) as above form a so-called cascade (terminology due to Kostant). He lists out these coroots and their sums in loc. sit. section 1.2 and section 1.8. In fact, a map x 7→ rx is defined from the set of involutions IW in an irreducible Weyl group W in loc. sit. and this is crucially used in constructing certain lifts of involutions in 46 associated reductive groups. We compare this map defined on the set of involutions with our weight function in Chapter 5. 3.4 Covering relation in Iwahori-Weyl group The main result of this section characterizes the covering relation of Bruhat order for most of the elements in W̃ . Theorem 3.4.1. Suppose that G is a split reductive group and w = utλv be an element of W̃ such that depth(λ) is bigger than a certain constant. Let rβ = tmuα∨ suα be an affine reflection for some positive root α and integer m, and let w′ = rβw. Then w ⋗ w′ is a covering relation, i.e. w ≥ w′ and ℓ(w) = ℓ(w′) + 1, if and only if one of the following conditions holds: 1. m = 0 and ℓ(usα) = ℓ(u)− 1; in this case, w′ = usαt λv. 2. m = 1 and ℓ(usα) = ℓ(u) + ⟨2ρ, α∨⟩ − 1; in this case, w′ = usαt λ−α∨ v. 3. m = ⟨α, λ⟩ and ℓ(sαv) = ℓ(v) + 1; in this case, w′ = utλsαv. 4. m = ⟨α, λ⟩ − 1 and ℓ(sαv) = ℓ(v)− ⟨2ρ, α∨⟩+ 1; in this case, w′ = utλ−α∨ sαv. In fact, it suffices to show the following result instead; see the remarks following the theorem below. Theorem 3.4.2. Suppose that W̃ is an Iwahowi Weyl group of a quasi-simple split reductive group. Let tmα∨ sα be an affine reflection, with (α,m) ∈ Φ+ × Z and assume w = tλy is an 47 element of W̃ such that depth(λ) ≥  3, if W is of simply laced type; 4, if W is of non-simply laced type but not of type G2; 6, if W is of type G2. (3.4.1) Then w′ := tmα∨ sαw is a cocover of w1 if and only if one of the following holds. 1. m = 1 and ℓ(sα) = ⟨2ρ, α∨⟩ − 1; in this case, w′ = sαt λ−α∨ y. 2. m = ⟨α, λ⟩ and ℓ(sαy) = ℓ(y) + 1; in this case, w′ = tλsαy. 3. m = ⟨α, λ⟩ − 1 and ℓ(sαy) = ℓ(y)− ⟨2ρ, α∨⟩+ 1; in this case, w′ = tλ−α∨ sαy. Remark 3.4.3. We explain how Theorem 3.4.2 implies Theorem 3.4.1. Let us first note that we can again reduce to the case of irreducible factors as in Remark 3.2.3. Note that if λ is regular, ℓ(utλv) = ℓ(u) + ℓ(tλv) by Equation (2.1.1). Choose reduced expressions u = si1 · · · sil and tλv = sj1 · · · sjkζ, where sip , sjq ∈ S for 1 ≤ p ≤ l, 1 ≤ q ≤ k and ζ ∈ Ω. Then utλv = si1 · · · silsj1 · · · sjkζ is a reduced expression, and a cocover of utλv must be of the form of either si1 · · · ŝim · · · silsj1 · · · sjkζ for some p ∈ [1, l] or si1 · · · silsj1 · · · ŝjn · · · sjkζ for some q ∈ [1, k]. In other words, any cocover of utλv is obtained by • (C1) either multiplying a cocover of u with tλv, or • (C2) multiplying u with a cocover of tλv. 1for two elements w,w′ of W̃ , we say that w′ is a cocover of w if w is a cover of w′, i.e. w > w′ and ℓ(w′) = ℓ(w)− 1. We also shorthand these last two conditions by writitng w ⋗ w′. 48 The procedure listed in (C1) is easy to describe. A cocover of u is of the form usα for some α ∈ Φ+ such that ℓ(usα) = ℓ(u)− 1. This corresponds to case (1) in theorem B. We claim that the three other cases there, i.e. (2)-(4), come from the procedure described in (C2), and hence they correspond to those listed in Theorem 3.4.2. Clearly, the third and fourth case in theorem B corresponds respectively to the second and third case in Theorem 3.4.2. Finally, suppose that w′ := sat λ−α∨ v is a cocover of w := tλv such that uw′ is a cocover of uw. By Equation (2.1.1), ℓ(uw′) = ℓ(usa) + ℓ(tλ−α∨ )− ℓ(v) = ℓ(usa) + ⟨2ρ, λ− α∨⟩ − ℓ(v). We use dominance of λ − α∨ in the last equality; this is true as ⟨αi, λ − a∨⟩ = ⟨αi, λ⟩ − ⟨αi, α ∨⟩ ≥ 0 by our earlier discussion about the maximum value of ⟨αi, α ∨⟩. Note that ℓ(usα) ≤ ℓ(u) + ℓ(sa) ≤ ℓ(u) + ⟨2ρ, α∨⟩ − 1. (3.4.2) Since ℓ(utλv) = ℓ(u) + ⟨2ρ, λ⟩ − ℓ(v), the cocover condition uw ⋗ uw′ gives ℓ(u) + ⟨2ρ, λ⟩ − ℓ(v)− 1 = ℓ(usa) + ⟨2ρ, λ− α∨⟩ − ℓ(v) ≤ ℓ(u) + ⟨2ρ, α∨⟩ − 1 + ⟨2ρ, λ− α∨⟩ − ℓ(v). Note that ℓ(u)+⟨2ρ, λ⟩−ℓ(v)−1 = ℓ(usa)+⟨2ρ, λ−α∨⟩−ℓ(v) ≤ ℓ(u)+⟨2ρ, α∨⟩−1+⟨2ρ, λ−α∨⟩−ℓ(v). 49 Hence we deduce that all of the inequalities in Equation (3.4.2) must be equality, and thus we get ℓ(usα) = ℓ(u) + ⟨2ρ, α∨⟩ − 1. This is exactly the situation listed in the second item in theorem B. Remark 3.4.4. We note that in a recent preprint [Sch22a][proposition 4.5], Schremmer gives several equivalent criteria that describes the covering relation, without any assumption on the depth of the relevant coweight. We will only focus on proving the necessity of the above conditions. For the sufficiency part, the argument in [Mil21, Section 4.1] toward the end of proposition 4.2 can be applied. We divide our discussion of the proof in the following subsections. 3.4.1 Some useful inequalities In this subsection, we lay out estimate of some relevant quantities that we shall use in the next subsection. We shall resume the assumption on the depth of λ stated in Theorem 3.4.2 throughout our discussion after the first lemma below. Note that we can assume throughout that w ∈ Wa, i.e. the length zero component of w is 1 ∈ Ω. Lemma 3.4.5. Let w = tλy be an element of W̃ with λ dominant and w ∈ SW̃ . Suppose w⋗w′. Then we must have w′ = tmα∨ sαw for some affine reflection tmα∨ sα with 1 ≤ m ≤ ⟨α, λ⟩. Proof. Let w = si1 · · · sil be a reduced expression, where sij ∈ S̃. This describes a reduced gallery from a to wa via the chain of adjacent alcoves a→ si1a→ · · · → si1 · · · sila. Since the gallery is reduced, it cannot cross any hyperplane twice - hence all the alcoves lie in 50 the dominant chamber. Let rj be the affine reflection with respect to the common wall between aj−1 := si1 · · · sij−1 a and aj := si1 · · · sija, that is rj = si1 · · · sij−1 sijsij−1 · · · si1 . Now suppose that w′ = si1 · · · ŝik · · · sil is a cocover; as before, let {a′ j : 1 ≤ j ≤ l− 1} be the collection of alcoves describing the reduced gallery corresponding to this expression of w′ and let r′j be the associated affine reflections. Then we see that rj = r′j and aj = a′ j for 1 ≤ j ≤ k − 1, and the rest of the gallery for w′ is the reflection of portion of the gallery for w from ak+1 onward with respect to the hyperplane corresponding to rk. This is because for j ≥ k + 1 we have a′ j−1 = si1 · · · sik−1 sik+1 · · · sija = si1 · · · sik−1 siksik−1 · · · si1(si1 · · · sija) = rk(aj). In particular, w′ = rkw. Therefore, to create a cocover of w, we must reflect wa with respect to some hy- perplane Hα−m lying between a and itself. Since the number of hyperplanes between the alcove wa and the wall Hα is ⟨α, λ⟩, this restricts the possibility of m to asserted values above. Now, w′ = tmα∨ sαw = sαt λ−mα∨ y. By Lemma 3.4.5, the coweight associated to the translation part of w′ belongs to following list of coweights: λ− α∨, λ− 2α∨, · · · , λ− (⟨α, λ⟩ − 1)α∨ = sα(λ− α∨), λ− ⟨α, λ⟩α∨ = sα(λ). (3.4.3) 51 Let us now define k = kλ,α := max{m : λ−mα∨ ∈ C+, 1 ≤ m ≤ ⟨α, λ⟩,m ∈ Z}. For a chamber Cx, we denote its closure by Cx. We now give a lower bound for k. Lemma 3.4.6. We have that k ≥  1, if W is of simply laced type; 2, if W is of non-simply laced type. Proof. By the definition of k, there is a real number m̃ with k ≤ m̃ < k + 1 such that λ− m̃α∨ lies on at least one of the walls of C+. It follows that for some simple root αi, we have ⟨αi, λ− m̃α∨⟩ = 0. Therefore, m̃ = ⟨αi,λ⟩ ⟨αi,α∨⟩ - hence giving k = ⌊ ⟨αi,λ⟩ ⟨αi,α∨⟩⌋. Recalling the estimate about maximum value of ⟨αi, α ∨⟩ in each Cartan type, we get the desired bound on k from the depth condition on λ. Our next lemma estimates the length of the translation elements corresponding to the coweights in the list (6.1). This proof closely follows a part of the proof of [Mil21, Proposition 4.2]. We include it here for completeness’ sake. Lemma 3.4.7. Suppose that for some integer m ∈ [1, ⟨α, λ⟩], λ−mα∨ lies in the closure of a chamber different from C+ or Csα. Then ℓ(tλ−mα∨ ) ≤ ℓ(tλ−kα∨ ). Proof. Following [LS10], define the function f : R→ R≥0 by linearly extending the function f̃ : Z→ Z≥0 defined by f̃(m) = ℓ(tλ−mα∨ ). 52 More precisely, f is the function associated to the graph obtained by joining f̃(m) and f̃(m + 1) by the line passing through them for every m ∈ Z. It is easy to see that f is a convex function, cf. [LS10], proof of proposition 4; in fact, it is a piece-wise linear function, and is given by a single expression linear in m as long as λ−mα∨ is in the same chamber. Since λ, λ− α∨ are both dominant due to the imposed depth hypothesis, we see that f(1) = ⟨2ρ, λ− α∨⟩, f(⟨α, λ⟩) = ⟨2ρ, λ⟩. We now show that f is decreasing around 1 and increasing around ⟨α, λ⟩. For example, if m ∈ (0, 3 2 ) and αi ∈ ∆ then by our earlier discussion about maximum value of ⟨αi, α ∨⟩ we have ⟨αi, λ−mα∨⟩ = ⟨αi, λ⟩ −m⟨αi, α ∨⟩ ≥  depth(λ)− 3, if W is not of type G2; depth(λ)− 9 2 , if W is of type G2. Hence, our depth hypothesis ensures that ⟨αi, λ−mα∨⟩ ≥ 0. Therefore, λ−mα∨ is in C+ whenever m ∈ (0, 3 2 ), and thus f(m) = ⟨2ρ, λ−mα∨⟩ = ⟨2ρ, λ⟩ −m⟨2ρ, α∨⟩ (3.4.4) is clearly decreasing in this neighbourhood. Similarly, we note that if m ∈ (⟨α, λ⟩ − 1 2 , ⟨α, λ⟩+ 1 2 ) and αi ∈ ∆ we have ⟨αi, sα(λ−mα∨)⟩ = ⟨αi, λ⟩ − (⟨α, λ⟩ −m)⟨αi, α ∨⟩ ≥  depth(λ)− 1, if W is not of type G2; depth(λ)− 3 2 , if W is of type G2. 53 Again, our depth hypothesis ensures that ⟨αi, sα(λ − mα∨)⟩ > 0. Hence λ − mα∨ ∈ Csα whenever m ∈ (⟨α, λ⟩ − 1 2 , ⟨α, λ⟩+ 1 2 ), and thus f(m) = ⟨2ρ, sa(λ−mα∨)⟩ = ⟨2ρ, λ⟩−(⟨α, λ⟩−m)⟨2ρ, α∨⟩ = ⟨2ρ, sa(λ)⟩+m⟨2ρ, α∨⟩ (3.4.5) is clearly increasing in the neighbourhood. Since the intervals (0, 3 2 ) and (⟨α, λ⟩ − 1 2 , ⟨α, λ⟩ + 1 2 ) are disjoint, by convexity of f we infer the global shape of the graph of f . Namely, f(m) steadily decreases and is defined by Equation (3.4.4) as long as λ−mα∨ ∈ C+; as λ−mα∨ traverses through other chambers with value of m increasing, it (weakly) decreases further until it reaches local minimum, then it starts increasing; and finally, once λ −mα∨ enters Csα , f(m) is defined by Equation (3.4.5) and steadily increases. Therefore, if we define k′ = k′ λ,α := min{m : λ−mα∨ ∈ Csα , 1 ≤ m ≤ ⟨α, λ⟩,m ∈ Z}, we must have f(m) ≤ max{(f(k), f(k′)} for allm ∈ [k, k′]∩Z. Since λ−kα∨ = sα(λ−k′α∨) by symmetry, we have f(k) = f(k′) = ℓ(tλ−kα∨ ). Hence we are done. Recall from Lemma 3.4.5 that m ≤ ⟨α, λ⟩. We now define z ∈ W to be such that λ−mα∨ ∈ Cz. Note that if λ−mα∨ is singular, z is not uniquely specified by this condition. However, the precise choice of z would be immaterial in what follows, and henceforth fix one such z satisfying the above condition. Let us denote the left and right weak Bruhat order by ≺left and ≺right respectively. Lemma 3.4.8. We have z ≺right sα. As a consequence, ℓ(sαz) + ℓ(z) = ℓ(sα). 54 Proof. It suffices to show that Inv(z−1) ⊂ Inv(sα), because this is equivalent to z−1 ≺left sα, which in turn is equivalent to the claim above. We start by noting that for β ∈ Inv(z−1), we have ⟨β, λ −mα∨⟩ ≤ 0. This follows from rewriting ⟨β, λ − mα∨⟩ = ⟨z−1β, z−1(λ − mα∨)⟩ and noting that z−1(λ − mα∨) is a dominant coweight and z−1β ∈ −Φ+. Therefore we get ⟨β, λ⟩ ≤ m⟨β, α∨⟩. Since λ dominant regular, ⟨β, λ⟩ > 0 and hence ⟨β, α∨⟩ > 0 as well. Dividing out both side of the inequality by ⟨β, α∨⟩, we get m ≥ ⟨β,λ⟩ ⟨β,α∨⟩ . Combining this with Lemma 3.4.5, we obtain ⟨α, λ⟩ ≥ ⟨β, λ⟩ ⟨β, α∨⟩ . But then ⟨β, α∨⟩⟨α, λ⟩ ≥ ⟨β, λ⟩, whence ⟨sα(β), λ⟩ = ⟨β − ⟨β, α∨⟩α, λ⟩ ≤ 0. Since λ is dominant regular, this gives sα(β) ∈ −Φ+, hence β ∈ Inv(sα). Thus we have shown Inv(z−1) ⊂ Inv(sα). Finally, let us note that ℓ(sαz)+ℓ(z) = ℓ(sα)−ℓ(z)+2|Inv(sα)c∩Inv(z−1)|+ℓ(z) = ℓ(sα)+2|Inv(sα)c∩Inv(z−1)| = ℓ(sα). This finishes the proof. 3.4.2 Two distinct possibilities In this subsection we further pin down the possible values of z. Recall that z−1(λ− mα∨) is dominant; let us temporarily denote this coweight by µ. Let J ⊂ ∆ be such that Stab(µ) = WJ , the associated standard parabolic subgroup. Abbreviate ζ = z−1y. 55 By standard fact about Coxeter groups, ζ has an unique factorization as ζ = ζJζ J with ζJ ∈ WJ and ζJ ∈ JW . Note that w′ = sαt λ−mα∨ y = sαzt µz−1y = sαzt µζJζ J = sαzζJt µζJ . (3.4.6) Lemma 3.4.9. The alcove tµζJa is in the dominant chamber. Proof. Define ω̄ := 1 n ∑n i=1 ω∨ i ni , where ni is the coefficient of αi in the highest root θ. Since µ+ ζJ ω̄ is centroid of the alcove tµζJa, we can make the following observation: tµζJa is in the dominant chamber if and only if µ+ ζJ ω̄ is dominant. We now show that the latter statement holds true. Note that if αi ∈ J , then ⟨αi, µ+ ζJ ω̄⟩ = ⟨(ζJ)−1αi, ω̄⟩ > 0, since by definition siζ J > ζJ , or equivalently (ζJ)−1si > (ζJ)−1, thereby giving (ζJ)−1αi ∈ Φ+. Now let αi ∈ ∆ \ J , then ⟨αi, µ⟩ ≥ 1; since (ζJ)−1αi ≥ −θ, we have ⟨αi, µ+ ζJ ω̄⟩ ≥ 1 + ⟨−θ, ω̄⟩ = 1− 1 = 0. This proves the claim. Applying the previous lemma, we get from Equation (3.4.6) ℓ(w′) = ℓ(sαzζJ) + ⟨2ρ, µ⟩ − ℓ(ζJ). Therefore the cocover condition gives ⟨2ρ, λ⟩ − ℓ(y)− 1 = ℓ(sαzζJ) + ⟨2ρ, µ⟩ − ℓ(ζJ). 56 In other words, we get ℓ(tλ)− ℓ(tµ) = 1 + ℓ(sαzζJ) + ℓ(y)− ℓ(ζJ). (3.4.7) Write y = z · z−1y = zζJζ J . Note that Equation (2.1.2) then gives ℓ(y) = ℓ(z · ζJζJ) = ℓ(z) + ℓ(ζJ) + ℓ(ζJ)− 2|Inv(z) ∩ Inv((ζJ)−1ζ−1 J )|. Similarly, ℓ(sazζJ) = ℓ(sαz)− ℓ(ζJ) + 2|Inv(sαz)c ∩ Inv(ζ−1 J )| Hence we can rewrite the right hand side of Equation (3.4.7) as 1 + ℓ(sαz) + ℓ(z) + 2{|Inv(sαz)c ∩ Inv(ζ−1 J )| − |Inv(z) ∩ Inv((ζJ)−1ζ−1 J )|}. Since ζ−1 J ≺left (ζ J)−1ζ−1 J , we have that Inv(ζ−1 J ) ⊂ Inv((ζJ)−1ζ−1). Combining this with the fact that |A| − |B| ≤ |A \B| for two sets A,B, we therefore conclude that ℓ(tλ)− ℓ(tµ) ≤ 1 + ℓ(sαz) + ℓ(z) + 2|Inv(sαz)c ∩ Inv(z)c ∩ Inv(ζ−1 J )|. (3.4.8) Lemma 3.4.10. We have Inv(sαz) c ∩ Inv(z)c ∩ Inv(ζ−1 J ) = ∅. Proof. The argument here is similar to the proof of Lemma 3.4.8. Suppose β ∈ Inv(sαz) c∩ Inv(z)c∩ Inv(ζ−1 J ) ⊂ Inv(sαz) c∩ Inv(z)c∩Φ+ J . Thus ⟨β, z−1(λ−mα∨)⟩ = 0, hence ⟨zβ, λ⟩ = 57 m⟨zβ, α∨⟩. Since zβ is a positive root, both sides are positive and m = ⟨zβ, λ⟩ ⟨zβ, α∨⟩ ≤ ⟨α, λ⟩. Therefore, ⟨zβ, λ⟩ − ⟨zβ, α∨⟩⟨α, λ⟩ = ⟨zβ, sα(λ)⟩ ≤ 0. Since β ∈ Inv(sαz) c, we have sα(zβ) ∈ Φ+. Hence we have ⟨zβ, sα(λ)⟩ = ⟨sα(zβ), λ⟩ ≤ 0, but that is a contradiction since λ is dominant regular. This shows that the purported set must be empty and we are done. Therefore, combining Equation (3.4.8) with Lemma 3.4.8 and Lemma 3.4.10 gives ℓ(tλ)− ℓ(tµ) ≤ 1 + ℓ(sα). (3.4.9) Now, suppose that z ̸= 1, sα. By Lemma 3.4.7 this gives ℓ(tλ)− ℓ(tµ) ≥ ℓ(tλ)− ℓ(tλ−kα∨ ) = k⟨2ρ, α∨⟩. Combining this with Equation (3.4.9), we get k⟨2ρ, α∨⟩ ≤ 1 + ℓ(sα) ≤ ⟨2ρ, α∨⟩ This gives k = 1. For the non-simply laced types, this is a contradiction with pre- viously established lower bound in Lemma 3.4.6. Therefore in such cases, we get that z ∈ {1, sα}. Before going forward, we make the following observation about a simply laced root 58 system: if ⟨β, α∨⟩ = 2 for a fixed coroot α, then β = α. This can be checked directly in type An and Dn, where the positive roots are given by {ei − ej : 1 ≤ i < j ≤ n} and {ei ± ej : 1 ≤ i < j ≤ n} ∪ {ei + en : 1 ≤ i < n} respectively. Since root systems of type E6, E7 arise as subsystem of type E8, we just give an argument for root system of type E8 to conclude. Note that for root system of type E8, the positive roots are of two kinds - given by {±ei + ej : 1 ≤ i < j ≤ 8}, and {1 2 (e8 + ∑7 i=1(−1)ν(i)ei):∑7 i=1 ν(i) ∈ 2Z}. We can easily see that pairing between a root γ from the first set with a coroot α∨ corresponding to elements of either sets cannot be 2 unless γ = α. The remaining possibility is ⟨1 2 (e8 + ∑7 i=1(−1)ν1(i)ei), 1 2 (e8 + ∑7 i=1(−1)ν2(i)ei)⟩ = 2, but that gives ⟨e8 + ∑7 i=1(−1)ν1(i)ei, e8 + ∑7 i=1(−1)ν2(i)ei⟩ = 8 - therefore forcing ν1 = ν2. We now resume the discussion about estimating k and restrict ourselves to the sim- ply laced types. Recall that k = 1. This means that λ − α∨ ∈ C+, but λ − 2α∨ /∈ C+; hence there exists β ∈ Φ+ such that ⟨β, λ− 2α∨⟩ ≤ 0. By the depth condition, this yields 3 ≤ ⟨β, λ⟩ ≤ 2⟨β, α∨⟩. Therefore ⟨β, α∨⟩ = 2, and ⟨β, λ⟩ is equal to either 3 or 4. But this gives β = α, and hence ⟨α, λ⟩ is either 3 or 4. If ⟨α, λ⟩ = 3, the relevant coweights are λ − α∨, λ − 2α∨ = sα(λ − α∨), λ − 3α∨ = sα(λ); the first one is in C+ by the depth condition, and the last two are in Csα . If ⟨α, λ⟩ = 4, then the relevant coweights are λ−α∨, λ− 2α∨, λ− 3α∨ = sα(λ−α∨), λ− 4α∨ = sα(λ); the first one is in C+, the last two are in Csα and λ− 2α∨ lies on the wall Hα, so it lies in both C+ and Csα . We summarize the content of this subsection as follows. 59 Proposition 3.4.11. If sαt λ−mα∨ y is a cocover of tλy, then λ − mα∨ is either in C+ or Csα. 3.4.3 Finishing the proof Proof of Theorem 6.1. We deal with the two cases found above. 1. When λ−mα∨ is in C+: in this case, Equation (3.4.9) gives us ⟨2ρ, λ⟩ − ⟨2ρ, λ−mα∨⟩ ≤ 1 + ℓ(sα) ≤ ⟨2ρ, α∨⟩. Therefore we get m ≤ 1; hence m = 1 and 1 + ℓ(sα) = ⟨2ρ, α∨⟩. This corresponds to the first case in Theorem 3.4.2. 2. When λ−mα∨ is in Csα : in this case, Equation (3.4.9) gives us ⟨2ρ, λ⟩ − ⟨2ρ, sα(λ−mα∨)⟩ ≤ 1 + ℓ(sα). (3.4.10) Since ℓ(sa) ≤ ⟨2ρ, α∨⟩ − 1, we thus get (⟨α, λ⟩ −m)⟨2ρ, α∨⟩ ≤ ⟨2ρ, α∨⟩. Therefore we get ⟨α, λ⟩ − m ≤ 1. Since m ⩽ ⟨α, λ⟩, this means that either m = ⟨α, λ⟩ − 1 or m = ⟨α, λ⟩. Suppose first that it is the former case. Then substituting this value of m in Equa- tion (3.4.10) yields ℓ(sα) = ⟨2ρ, α∨⟩ − 1. Note that λ−mα∨ = sα(λ− α∨) is regular 60 (since λ−α∨ is regular dominant due to the depth hypothesis) and thus J = ∅; substi- tuting µ = sα(λ−α∨), z = sα, ζJ = 1, ζJ = sαy in Equation (3.4.7) therefore produces ⟨2ρ, α∨⟩ = 1+ℓ(y)−ℓ(sαy), whence we get ℓ(sαy) = ⟨2ρ, α∨⟩−1−ℓ(y) = ℓ(sα)−ℓ(y). This corresponds to the third case in Theorem 3.4.2. Now, if it is the latter case we can carry out a similar substitution in Equation (3.4.7) to get 0 = 1+ ℓ(y)− ℓ(sαy); Equation (3.4.10) does not yield any information in this case. This gives ℓ(say) = ℓ(y) + 1 and hence it corresponds to the second case in Theorem 3.4.2. This completes the proof. Remark 3.4.12. We know a posteriori that only the first and the last two choices from the string of coweights in Equation (3.4.3) are viable - so we must necessarily put a depth condition on λ to ensure that λ − α∨ is dominant; in that case, the first coweight is in the dominant chamber and the last two are in Csα . In other words, the least stringent hypothesis on λ under which one can expect to prove a result like this would be that λ−α∨ is dominant. The depth condition that we impose above for the simply laced types is almost as weak an assumption as this, in the sense that we require λ− α∨ to be dominant regular. Calculations in small rank seem to provide evidence for our this speculation. In other words, we suspect that the depth hypothesis can be lowered to 2 uniformly in all the cases. 61 3.4.4 Admissible subsets of W̃ A useful description of the admissible set is given in [HY21] in terms of weight of minimal length paths in the quantum Bruhat graph. This relies on the characterization of covering relation in W̃ as established in [Mil21, Proposition 4.2] and hence as such it necessarily brings into picture the superregularity condition, cf. [HY21, Proposition 3.3]. Since we can strengthen the result about covering relation, we automatically get the following improvement in the description of the admissible set. Proposition 3.4.13. Suppose that W is an irreducible Weyl group. Assume that depth(µ) ≥  3, if W is of simply laced type; 4, if W is of non-simply laced type but not of type G2. 6, if W is of type G2. Let λ be dominant and assume that ⟨ρ, µ− λ⟩ <  ⌈depth(µ)−3 2 ⌉, if W is of simply laced type; ⌈depth(µ)−4 2 ⌉, if W is of non-simply laced type but not of type G2; ⌈depth(µ)−6 2 ⌉, if W is of type G2. Then xtλy ∈ Adm(µ) if and only if wt(x, y−1) ≤ µ− λ. The proof of this proposition is identical to the argument made in [HY21], cf. section 3.3 and proof of proposition 3.3 in there and hence we do not repeat it here. The additional 62 ingredient is Theorem 3.4.2, and the conditions on µ and λ are a direct reflection of that. 63 Chapter 4: A dimension formula for X(µ, b) The primary goal of this chapter is to investigate the dimension of the union X(µ, b) of affine Deligne-Lusztig varieties. Such a geometric object is the group-theoretic model for the Newton stratum associated with the element [b] lying B(G, µ). We remark that in the situations where the He-Rapoport axioms in [HR17] hold, this sought-after dimension is equal to the dimension of the associated Newton stratum minus the dimension of the corresponding central leaf, see [He16a, §2.12]. The said union of affine Deligne-Lusztig varieties can be defined purely in terms of group-theoretic data without any reference to Shimura varieties, and as before we resort to this latter setup. In Section 4.1, we obtain a dimension formula in the case of a quasi-split group, improving an earlier work of [HY21]. Let us now place ourselves in the context of a general group G, and let bmin be the unique minimal (equivalently, basic) element of B(G, µ). Then the work of Görtz, He and Rapoport in [GHR22] gives characterizations of when X(µ, bmin) can be of minimal dimension zero, or of maximal dimension ⟨2ρ, µ⟩; the model cases of such phenomena are the Lubin-Tate case and the Drinfeld case, respectively. Let us also remark that in [GHN20], Görtz, He and Nie provide a sharp lower bound for the dimension of X(µ, bmin), and they are also able to classify completely when this becomes an equality. Besides these results, there is no dimension formula available for X(µ, b), nor is there any conjectural description. 64 Standing at this juncture, it is therefore natural to investigate the dimension problem for the other extremal element of B(G, µ), namely bmax. In the rest of this chapter, we compute this dimension with a mild depth hypothesis on µ. 4.1 Dimension in the quasi-split case In this section, we show that the dimension formula provided in [HY21, Theorem 6.1] holds in the absence of the superregularity condition imposed there. Our argument closely follows the proof scheme established in [HY21]. Therefore, we replace most proofs with references to the corresponding arguments made in loc. sit. and only point out how to bypass certain steps that will enable us to drop the superregularity condition. We first prove an enhanced version of [HY21, Proposition 4.4]. First, we set dAdm(µ)(b) = max w∈Adm(µ) dw(b). Proposition 4.1.1. Assume that G is quasi-split. Let µ be regular. Then dAdm(µ)(b) = ⟨ρ, µ− ν([b])⟩ − 1 2 defG(b) + 1 2 ℓ(w0)− 1 2 min{dΓ(x, σ(x)w0) : x ∈ W}. (4.1.1) Proof. We first show that (i) if xtλy ∈ Adm(µ), then xtλ+µ′ y ∈ Adm(µ+ µ′) for any dominant µ′. We write xtλ+µ′ y = xtλyy−1tµ ′ y, and note that y−1tµ ′ y ∈ Adm(µ′) by definition. Hence the claim follows from an application of Theorem 2.3.1. Now we argue that statement (a) in the course of the proof in loc. sit. holds true in the absence of the superregularity condition. In other words, we need to show that (ii) if xtλy ∈ Adm(µ) where µ is regular, then ⟨ρ,wt(x, y−1)⟩ ≤ ⟨ρ, µ− λ⟩. 65 Let xtλy ∈ Adm(µ). Choose µ′ to be superregular, e.g. µ′ = nρ∨ for large enough n. Apply the statement in (i) to obtain xtλ+µ′ y ∈ Adm(µ + µ′). Now, µ + µ′ is superregular and thus statement (a) in loc. sit. applies to give us ⟨ρ,wt(x, y−1)⟩ ≤ ⟨ρ, (µ+ µ′)− (λ+ µ′)⟩ = ⟨ρ, µ− λ⟩. This finishes the proof of (ii). Therefore, one can use this enhanced version of state- ment (a) to get the established upper bound (i.e. the expression in the right hand side of Equation (4.1.1)) for dAdm(µ) by resorting to the same proof method in loc. sit. In the remaining part of the proof, the authors construct an explicit w ∈ Adm(µ)∩Cx for some x ∈ W and argue that dw(b) = ⟨ρ, µ− ν([b])⟩ − 1 2 defG(b) + 1 2 ℓ(w0)− 1 2 dΓ(x, σ(x)w0). (4.1.2) To prove this, they employ the description of Adm(µ) established in proposition 3.3 in loc. sit. and hence it relies on the superregularity hypothesis. However, one can bypass this simply by choosing a different candidate. Namely, let w = xtµ(σ(x)w0) −1 for some x ∈ W such that σ(x)w0 ≥ x. Let us first explain how this can be obtained using [HY21, Theorem 5.1]. Let O be the σ-conjugacy class of w0 and define ℓR(O) = min{ℓR(w) : w ∈ O}. Then the aforementioned theorem asserts that for any finite Coxeter group W and a length 66 preserving graph automorphism on W , we have that ℓ(w0)− ℓR(O) = 2max{ℓ(x) : x ≤ σ(x)w0}. Therefore, the existence of our required element is equivalent to the assertion that ℓ(w0) > ℓR(O). The latter statement is true in all types arising from quasi-split groups, cf. section 5.1 in loc. sit. Then we have w ≤ σ(x)w0t µ(σ(x)w0) −1 by a combination of Equation (2.1.1) and Equation (2.1.3). This is where we use that µ is regular. Hence, w ∈ Adm(µ). As has been shown in loc. sit., it now follows from definition that Equation (4.1.2) holds true for this element. This completes the proof that Equation (4.1.1) holds for dominant regular µ. Let us record an immediate consequence of the above result. Corollary 4.1.2. Suppose that µ is regular. Then dimX(µ, b) ≤ ⟨ρ, µ− ν([b])⟩ − 1 2 defG(b) + 1 2 ℓ(w0)− 1 2 min{dΓ(x, σ(x)w0) : x ∈ W}. Proof of Theorem 1.2.2[(1). ] We note that once the formula for dAdm(µ)(b) in Equation (4.1.1) is established, the remaining part of [HY21] focuses on showing min{dΓ(x, σ(x)w0) : x ∈ W} = ℓR(O). However, this is a calculation done purely on the finite Weyl group W , and the superregu- larity condition does not appear in establishing this. 67 Finally, the characterization of admissible sets is used once more in the proof of the theorem in section 6 in loc. sit. to show that xtµw0σ(x) −1 ∈ Adm(µ) for certain element x ∈ W satisfying σ(x)w0 ≥ x. However, this follows directly as we have shown above. Hence, the superregularity hypothesis on µ can be avoided completely. This finishes the proof. Remark 4.1.3. We remark that [HY21, Remark 6.2] gives an example where µ is singular and W is of type C4, and they point out that the dimension formula fails in this case. Therefore, the depth hypothesis on µ is optimal, subject to the other condition. However, we do not know if the formula for dAdm(µ) given in Proposition 4.1.1 is valid in the absence of regularity condition on µ. When G is split, we can prove the following statement. Theorem 4.1.4. The same assertion as above holds under the hypothesis that µ has depth at least 3 and µ ≥ ν([b]) + wt(w0, 1). Note that wt(w0, 1) is substantially smaller than 2ρ∨, cf. Section 3.3; for instance, in type A2n, we have wt(w0, 1) = ϖ∨ n + ϖ∨ n+1. We remark that unlike our above proof of the dimension formula in the quasi-split case, the proof of Theorem 4.1.4 is completely independent of the arguments in [HY21]. Basically, the restriction µ⋄ ≥ ν(b) + 2ρ∨ enters in the proof of [HY21][Theorem 6.1] because the dimension for the affine Deligne-Lusztig variety associated to w := xtµw0σ(x) −1 - the element chosen in the proof of Theorem 6.1. in loc. sit. - can be asserted to be equal to dw(b) by [He21a] only under that stated restriction. Note that it is then shown for this element, we have dw(b) = dAdm(µ)(b) - by which one concludes that dimX(µ, b) = dAdm(µ)(b). 68 In other words, for this particular choice of w, the associated affine Deligne-Lusztig variety has a known dimension by design, and also makes it to the top dimensional component of X(µ, b). Proof. Instead of working with the element w as above, let us consider the element w′ := w0t µ−wt(w0,1). We claim that (a) w′ ∈ Adm(µ). Indeed, proceeding as in the proof of Proposition 3.2.6 and utilizing the decomposition of w0 in Section 3.3, we have tµ−wt(w0,1) ≤ tµw0, and hence w′ ≤ w0t µw0. This last inequality needs depth at least 3 to ensure that µ − wt(w0, 1) is dominant regular, as well the coweights appearing in the intermediate steps (arising from application of the proof technique in Proposition 3.2.6) are dominant. Hence, w′ ∈ Adm(µ). Now, we claim that (b) dw′(b) = dAdm(µ)(b). To that end, let us compute dw′(b) = 1 2 {ℓ(w′) + ℓ(w0)− def(b)− ⟨2ρ, ν([b])⟩} = 1 2 {ℓ(w0) + ⟨2ρ, µ− wt(w0, 1)⟩+ ℓ(w0)− def(b)− ⟨2ρ, ν([b])⟩} = 1 2 {⟨2ρ, µ− ν(b)⟩ − def(b)}+ {ℓ(w0)− ⟨ρ,wt(w0, 1)⟩}. The claim then would follow if we can show that the part in the second curly braces 69 is equal to ℓ(w0) − ℓR(w0). Note that Equation (5.0.2) gives ⟨2ρ,wt(w0, 1)⟩ = ℓ(w0) + d(w0, 1), but the explicit decomposition of w0 in Section 3.3 shows that d(w0, 1) = ℓR(w0). Combining, we have ℓ(w0)−⟨ρ,wt(w0, 1)⟩ = ℓ(w0)− 1 2 (ℓ(w0)+ℓR(w0)) = 1 2 {ℓ(w0)−ℓR(w0)}. Finally, we have that (c) dimXw′(b) = dw′(b), whenever Xw′(b) ̸= ∅. By [MV20, Theorem 1.2] every element in the antidominant chamber is cordial. Therefore w′ is cordial, and then above claim follows from [MV20, Corollary 3.17]. Finally, note further that the basic element in B(G) satisfying κ(b) = κ(w′) is indeed the minimal element ofB(G)w′ by [GHN16][Theorem B]. Hence, B(G)w′ = {[b] : [b] ≤ [bw′ ]}. By [He21b, Theorem 4.2], we have ν([bw′ ]) = µ − wt(w0, 1). By Equation (3.2.2), this in turn enforces the condition µ ≥ ν([b]) + wt(w0, 1) in order to ensure that Xw′(b) ̸= ∅. We are done. 4.2 Expressing bmax via generic Newton point In the rest of this chapter, we focus on the dimension problem for X(µ, b) associated to the maximal element b = bmax of B(G, µ). We first identify bmax as the maximum of generic σ-conjugacy classes associated with the maximal translation elements in the µ-admissible set. Then we proceed to express such generic σ-conjugacy classes in two distinct ways, first in terms of σ-twisted Demazure power and then via a reduction to the quasi-split case. Such considerations allow us to express the dimension of individual affine Deligne-Lusztig varieties associated to such generic σ-conjugacy classes in terms of certain statistics on the quantum Bruhat graph. Further analysis in Section 4.2.5 shows 70 that dimX(µ, bmax) lies between the length and the reflection length of certain minimal length elements of some Frobenius-twisted conjugacy classes in the finite Weyl group. We then show that such minimal length elements are, in fact, partial Coxeter elements via a case-by-case calculation in Section 4.3, and hence their length (which is equal to their reflection length) matches the asserted dimension in Theorem 1.2.4, thereby finishing the proof. Throughout the rest of this chapter, we make use of the following strengthened formula for the generic Newton point (for quasi-split groups) as established in [HN21]. Theorem 4.2.1. [HN21, Proposition 3.1] Suppose that G is quasi-split and adjoint over F . Let x, y ∈ W and µ ∈ X∗(T ) + Γ0 be such that depth(µ) ≥ 2. Then νxtµy is the average of the σ-orbit of µ− wt(y−1, σ(x)). 4.2.1 Dimension of a generic affine Deligne-Lusztig variety Following [He21a], we define the n-th σ-twisted Demazure power of w by setting w∗σ,n = w ∗ σ(w) ∗ σ2(w) ∗ · · · ∗ σn−1(w). We need the following result, which on the one hand describes the dimension of a single affine Deligne-Lusztig variety associated with generic σ-conjugacy class, and on the other hand quantifies such generic σ-conjugacy class in terms of σ-twisted Demazure power. For the first equality below, see [He16a, Theorem 2.23] and [MV20, Lemma 3.2], whereas for the second equality, see [He21a, Theorem 0.1]. Theorem 4.2.2. Let w ∈ W̃ . Then dimXw(bw) = ℓ(w)− ⟨2ρ, νw⟩ = ℓ(w)− lim n→∞ ℓ(w∗σ,n) n . 71 Note that there can be elements w ∈ W̃ that are not pure translations but still satisfy bw = bmax; this can be seen e.g. using the Deligne-Lusztig reduction method, see [He14, Proposition 4.2]. However, if w contributes to top dimensional components - i.e., dimX(µ, bmax) = dimXw(bmax) - then w is necessarily a translation element of the form txµ for some x ∈ W , by the first equality in Theorem 4.2.2. Now, again by an application of the first equality in Theorem 4.2.2 we see that dimXtxµ(btxµ) is a decreasing function of btxµ ; therefore, finding top dimensional Xw(bmax) inside X(µ, bmax) boils down to understanding elements x ∈ W such that dimXtxµ(btxµ) is minimized, and this minimum value of the dimension is indeed the dimension of X(µ, bmax). We approach the problem of computing dimXtxµ(btxµ) below in two different ways. In the rest of this chapter, we omit the underline for simplicity and write µ instead of µ throughout. 4.2.2 A standard reduction Let G be a connected reductive group over F , and let Gad be its adjoint group. Let Tad be the image of T in Gad, and denote by µad the image of µ in X∗(Tad)Γ0 . Similarly for any b ∈ Ğ, we denote by bad its image in Ğad. By [Kot97, Proposition 4.10], we then have an isomorphism of posets B(G, µ)→ B(Gad, µad), via [b] 7→ [bad]. Next, we have a decomposi