Notes on Weyl's Formula

Total Page:16

File Type:pdf, Size:1020Kb

Notes on Weyl's Formula THE SIZE OF THE SYMMETRIC GROUP NATHAN WILLIAMS 1. Introduction In our study of the descent algebra and the symmetric group, we haven’t seen a lot of geometry. I’d like to fix that by proving (geometrically!) that the symmetric group has n! elements (interpreting those descent classes geometrically). Let Sn be the symmetric group on [n] := f1; 2; : : : ; ng, defined to be the group of bijections [n] ! [n], where the group operation is composition. You probably know that jSnj = n!—for any element w 2 Sn there are n choices for w(1), leaving n − 1 choices for w(2), and so on. We will take a more geometric point of view—one that generalizes to other (crystallographic) reflection groups. We will interpret Sn+1 as an n-dimensional convex polytope P, so that each of the (n + 1)! permutations corresponds to a certain n-dimensional simplex that is similar to ( n ) X A = (x1; x2; : : : ; xn): xi ≥ 0; xi ≤ 1 : i=1 The union of all of these simplices will be the entire polytope P. 1.1. The simplex A. Let’s look at A a little more closely. 1 Proposition 1. Vol(A) = n! . cn Proof. We show by induction on dimension that Vol(cA) = n! . This is true for n−1 n c R c (c−x) (c−x) the line [0; c]. Then (the slices method) Vol(A) = (n−1)! = − n(n−1)! = x=0 x=0 cn n! : Proof. The number of integer points in cA is the number nonnegative integer so- c+n (c+n)···(c+1) lutions to the equation x0 + x1 + x2 + ··· + xn = c, which is n = n! . 1 By Ehrhart theory, Vol(A) is the leading coefficient of this polynomial, or n! . Proof. For each w 2 Sn, define the region Aw = f(x1; x2; : : : ; xn) : 1 > xw(1) > xw(2) > ··· > xw(n) > 0g: In fancier language, the order polytope of an n-element antichain is a parallelepiped of volume one. It has a triangulation given by the linear extensions of the antichain. There are n! of these. You might be unhappy—why was I allowed to change A to Aw for this last proof? First, I don’t care about the boundaries since those have smaller dimension and therefore don’t change the overall volume. Definition 2. For R a “region” with finite volume and T an n×n invertible matrix, Vol(T R) = jdet(T )jVol(R). If jdet(T )j = 1, then the linear transformation T is volume-preserving. 1 2 N. WILLIAMS Ae1 1 Ee6 Ee7 Aen Ben Ee8 Cen Fe4 Den Ge2 Figure 1. The finite and affine Dynkin diagrams (the affine node is marked in gray). Pi Now A is spanned by ei, while Ae is spanned by j=1 ej (we only need consider Ae because it is similar to all the other Aw). So to map A to Ae, we can just send Pn+1−i ei 7! j=1 ej. The matrix of this transformation is the matrix with ones on and above the diagonal, and zeros below the diagonal. Therefore, its determinant is one and the map that sends A to Ae is volume-preserving. Exercise 3. What is the inverse of this matrix? 1.2. Weyl’s formula. Our proof will cut P by the descent classes of Sn+1, and rearrange them so that they form an (n + 1)-fold cover of the parallelepiped Bn = f(x1; x2; : : : ; xn) : 0 ≤ xi ≤ 1g : This implies that (n + 1)Vol(Bn) = Vol(Pn) = jSn+1j · Vol(A) 1 (n + 1) = jS j · ; n+1 n! so that jSn+1j = (n + 1)!. Example 4. Show Mathematica stuff. Theorem 5. Let W be a Weyl group. Then jW j = f · n! · (a1a2 ··· an); where • f is the index of connection; • n! is the volume of an n-dimensional parallelepiped; where • a1; a2; : : : ; an are the coefficients of the highest root αe when expressed as a sum of simple roots (scaling factors for the parallelepiped). The Weyl groups are classified in Figure 1. The proof I want to show you is due to Hermann Weyl, but we’ll do everything explicitly for the symmetric group. 1.3. Eulerian Numbers and Stanley’s Bijection. A related map was used by Stanley to answer a question of Foata, and it is so nice that I’ll give it right away. We will come back to it at the end. Recall that a descent of a permutation w is a position i such that w(i) > w(i+1). If you subdivide the unit n-dimensional cube Bn = f(x1; x2; : : : ; xn) : 0 ≤ xi ≤ 1g THE SIZE OF THE SYMMETRIC GROUP 3 into level sets n X Rn;k := f(x1; x2; : : : ; xn) 2 Bn : k ≤ xi ≤ k + 1g; i=1 the volume of Rn;k is equal to the number of permutations in Sn with k descents. These numbers are called the Eulerian numbers, denoted Vol(Rn;k) = A(n; k). If Pn−1 k we let An(t) := k=0 A(n; k)t be the Eulerian polynomials, then we have 1 X xn t − 1 (1) A (t) = : n n! t − e(t−1)x n=0 Exercise 6. Use Equation (1) to show that 1 X An(t) mntm = : (1 − t)n+1 m=0 The volumes Vol(Rn;k) had been known since Laplace, but it was Foata who gave this combinatorial interpretation of the Eulerian numbers as permutations. Foata asked for a bijective proof: the third proof of Proposition 1 gave a decomposition of the unit cube into permutations—is there a way to rearrange this decomposition (without changing volume) so that the permutations with k descents are sent to the kth level set Rk? Stanley provided a positive solution. Example 7. Draw cubes and explain bijection. Exercise 8. Check that sending a “generic” point (x1; x2; : : : ; xn) in Aw (a point such that xi 6= xi+1 and xn 6= 1) to the point (y1; y2; : : : ; yn) such that ( xi−1 − xi if xi−1 > xi yi := 1 + xi−1 − xi if xi−1 < xi gives the desired map (where x0 = 1). Pn Answer to Exercise 8. We telescope i=1 yi = jdes(w)j + 1 − x1, so the image does indeed lie in the correct level set Rk. The map sends a point x = (x1; x2; : : : ; xn) 2 Aw to the point 0 1 −1 0 ··· 0 0 1 0 1 B 0 1 −1 ··· 0 0 C B C X B . C y = @ eiA + B . ··· . C x: B C i:i2des(w) @ 0 0 0 ··· 1 −1 A 0 0 0 ··· 0 1 Since the determinant of the linear part of this is absolute value one, it is volume- preserving by Definition 2. Note that this matrix is just the inverse of the matrix that we used to prove A and Ae had the same volume! We now invert the map. Starting with y1 = 1 − x1, we obtain y1. In general, we set ( xi−1 − yi if xi−1 > yi xi = 1 + xi−1 − yi if xi−1 < yi; Pi so that 0 < xi < 1. This is the same as xi = 1 − j=1 yj mod 1 4 N. WILLIAMS 1 1 For example, if we have x = 5 (3; 1; 5; 2; 4), then y = 5 (2; 2; 1; 3; 3). And we compute 5x1 = 3 = 5 − 5y5 = 5 − 2 = 3; 5x2 = 5 − (2 + 2) = 1; 5x3 = 5 − (2 + 2 + 1 mod 5) = 5; 5x4 = 5 − (2 + 2 + 1 + 3 mod 5) = 2; 5x5 = 5 − (2 + 2 + 1 + 3 + 3 mod 5) = 4: 2. The Symmetric Group and the Regular Simplex Proposition 9. Sn is isomorphic to the group of symmetries of a regular simplex in Rn−1. Proof. We embed the regular (n − 1)-dimensional simplex into Rn—with basis n feigi=1—as ( n n ) X X ∆n−1 := ciei : ci ≥ 0; ci = 1 : i=1 i=1 n Then the vertices of ∆n−1 are the feigi=1, any permutation of these vertices gives a symmetry of ∆n−1, and these are the only symmetries of ∆n−1. In other words, w 2 Sn acts on the set of vertices of ∆n−1 by w(ei) = ew(i) and therefore on all of Rn by n ! n X X w xiei = xiew(i): i=1 i=1 Example 10. Draw ∆2. n Exercise 11. Check that this representation of Sn on R is not irreducible by Pn showing that any permutation w fixes the space h i=1 eii. Can you write down Pn n generators for the complement of h i=1 eii in R ? Exercise 12. Show that S4 acting on vector space spanned by the six edges heij : 1 ≤ i < j ≤ 4i of ∆3 has the following three invariant subspaces: he12 + e13 + e14 + e23 + e24 + e34i; he12 − e14 − e23 + e34; e13 − e14 − e23 + e24i; and he12 − e34; e13 − e24; e14 − e23i: Can you decompose Sn acting on the (k − 1)-dimensional faces of ∆n−1 into irre- ducible representations? (Hint: expand the product of two homogenous symmetric functions hk · hn−k in the Schur function basis.) Exercise 13. S4 acts by permuting the four long diagonals of a cube. Decompose the corresponding permutation representations on the faces, vertices, and edges of the cube. The point I’m trying to make is that there are lots of “natural” things to act on, and it isn’t clear why the standard permutation representation of Sn ought to be favored! For example, why shouldn’t we just let Sn act on itself? This is called the regular representation.
Recommended publications
  • Bijective Proofs for Schur Function Identities Which Imply Dodgson's Condensation Formula and Plücker Relations
    Bijective proofs for Schur function identities which imply Dodgson’s condensation formula and Pl¨ucker relations Markus Fulmek Institut f¨urMathematik der Universit¨atWien Strudlhofgasse 4, A-1090 Wien, Austria [email protected] Michael Kleber Massachusetts Institute of Technology 77 Massachusetts Avenue, Cambridge, MA 02139, USA [email protected] Submitted: July 3, 2000; Accepted: March 7, 2001. MR Subject Classifications: 05E05 05E15 Abstract We present a “method” for bijective proofs for determinant identities, which is based on translating determinants to Schur functions by the Jacobi–Trudi identity. We illustrate this “method” by generalizing a bijective construction (which was first used by Goulden) to a class of Schur function identities, from which we shall obtain bijective proofs for Dodgson’s condensation formula, Pl¨ucker relations and a recent identity of the second author. 1 Introduction Usually, bijective proofs of determinant identities involve the following steps (cf., e.g, [19, Chapter 4] or [23, 24]): Expansion of the determinant as sum over the symmetric group, • Interpretation of this sum as the generating function of some set of combinatorial • objects which are equipped with some signed weight, Construction of an explicit weight– and sign–preserving bijection between the re- • spective combinatorial objects, maybe supported by the construction of a sign– reversing involution for certain objects. the electronic journal of combinatorics 8 (2001), #R16 1 Here, we will present another “method” of bijective proofs for determinant identitities, which involves the following steps: First, we replace the entries ai;j of the determinants by hλ i+j (where hm denotes i− • the m–th complete homogeneous function), Second, by the Jacobi–Trudi identity we transform the original determinant identity • into an equivalent identity for Schur functions, Third, we obtain a bijective proof for this equivalent identity by using the interpre- • tation of Schur functions in terms of nonintersecting lattice paths.
    [Show full text]
  • Affine Partitions and Affine Grassmannians
    AFFINE PARTITIONS AND AFFINE GRASSMANNIANS SARA C. BILLEY AND STEPHEN A. MITCHELL Abstract. We give a bijection between certain colored partitions and the elements in the quotient of an affine Weyl group modulo its Weyl group. By Bott’s formula these colored partitions give rise to some partition identities. In certain types, these identities have previously appeared in the work of Bousquet-Melou-Eriksson, Eriksson-Eriksson and Reiner. In other types the identities appear to be new. For type An, the affine colored partitions form another family of combinatorial objects in bijection with n + 1-core parti- tions and n-bounded partitions. Our main application is to characterize the rationally smooth Schubert varieties in the affine Grassmannians in terms of affine partitions and a generalization of Young’s lattice which refines weak order and is a subposet of Bruhat order. Several of the proofs are computer assisted. 1. Introduction Let W be a finite irreducible Weyl group associated to a simple connected compact Lie Group G, and let Wf be its associated affine Weyl group. In analogy with the Grassmannian manifolds in classical type A, the quotient Wf /W is the indexing set for the Schubert varieties in the affine Grassmannians LG. Let WfS be the minimal length coset representatives for Wf /W . Much of the geometry and topology for the affine Grassmannians can be studied from the combinatorics of WfS and vice versa. For example, Bott [6] showed that the Poincar´eseries for the cohomology ring for the affine Grassmannian is equal to the length generating function for WfS, and this series can be written in terms of the exponents e1, e2, .
    [Show full text]
  • Crystal Symmetry Groups
    X-Ray and Neutron Crystallography rational numbers is a group under Crystal Symmetry Groups multiplication, and both it and the integer group already discussed are examples of infinite groups because they each contain an infinite number of elements. ymmetry plays an important role between the integers obey the rules of In the case of a symmetry group, in crystallography. The ways in group theory: an element is the operation needed to which atoms and molecules are ● There must be defined a procedure for produce one object from another. For arrangeds within a unit cell and unit cells example, a mirror operation takes an combining two elements of the group repeat within a crystal are governed by to form a third. For the integers one object in one location and produces symmetry rules. In ordinary life our can choose the addition operation so another of the opposite hand located first perception of symmetry is what that a + b = c is the operation to be such that the mirror doing the operation is known as mirror symmetry. Our performed and u, b, and c are always is equidistant between them (Fig. 1). bodies have, to a good approximation, elements of the group. These manipulations are usually called mirror symmetry in which our right side ● There exists an element of the group, symmetry operations. They are com- is matched by our left as if a mirror called the identity element and de- bined by applying them to an object se- passed along the central axis of our noted f, that combines with any other bodies.
    [Show full text]
  • Molecular Symmetry
    Molecular Symmetry 1 I. WHAT IS SYMMETRY AND WHY IT IS IMPORTANT? Some object are ”more symmetrical” than others. A sphere is more symmetrical than a cube because it looks the same after rotation through any angle about the diameter. A cube looks the same only if it is rotated through certain angels about specific axes, such as 90o, 180o, or 270o about an axis passing through the centers of any of its opposite faces, or by 120o or 240o about an axis passing through any of the opposite corners. Here are also examples of different molecules which remain the same after certain symme- try operations: NH3, H2O, C6H6, CBrClF . In general, an action which leaves the object looking the same after a transformation is called a symmetry operation. Typical symme- try operations include rotations, reflections, and inversions. There is a corresponding symmetry element for each symmetry operation, which is the point, line, or plane with respect to which the symmetry operation is performed. For instance, a rotation is carried out around an axis, a reflection is carried out in a plane, while an inversion is carried our in a point. We shall see that we can classify molecules that possess the same set of symmetry ele- ments, and grouping together molecules that possess the same set of symmetry elements. This classification is very important, because it allows to make some general conclusions about molecular properties without calculation. Particularly, we will be able to decide if a molecule has a dipole moment, or not and to know in advance the degeneracy of molecular states.
    [Show full text]
  • Math 958–Topics in Discrete Mathematics Spring Semester 2018 TR 09:30–10:45 in Avery Hall (AVH) 351 1 Instructor
    Math 958{Topics in Discrete Mathematics Spring Semester 2018 TR 09:30{10:45 in Avery Hall (AVH) 351 1 Instructor Dr. Tri Lai Assistant Professor Department of Mathematics University of Nebraska - Lincoln Lincoln, NE 68588-0130, USA Email: [email protected] Website: http://www.math.unl.edu/$\sim$tlai3/ Office: Avery Hall 339 Office Hours: By Appointment 2 Prerequisites Math 450. 3 Contacting me The best way to contact with me is by email, [email protected]. Please put \[MATH 958]" at the beginning of the title and make sure to include your whole name in your email. Using your official UNL email to contact me is strongly recommended. My office is Avery Hall 339. My office hours are by appoinment. 4 Course Description Bijective Combinatorics is a branch of combinatorics focusing on bijections between mathematical objects. The fact is that there many totally different objects in mathematics that are actually equinumerous, in this case, we always want to seek for a simple bijection between them. For example, the number of ways to divide that a convex (n + 2)-gon into triangles by non-intersecting diagonals is equal to the number of full binary trees with n + 1 leaves. Both objects are counted 1 2n by the well known Catalan number Cn = n+1 n . Do you know that there are more than two hundred (!!) mathematical objects are counted by the Catalan number? In this course, we will go over a number of such \Catalan objects" and investigate the beautiful bijections between them. One more example is the well-known theorem by Euler about integer partitions stating that the number of ways to write a positive integer n as the sum of distinct positive integers is equal to the number of ways to write n as the sum of odd positive integers.
    [Show full text]
  • Finding Direct Partition Bijections by Two-Directional Rewriting Techniques
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Elsevier - Publisher Connector Discrete Mathematics 285 (2004) 151–166 www.elsevier.com/locate/disc Finding direct partition bijections by two-directional rewriting techniques Max Kanovich Department of Computer and Information Science, University of Pennsylvania, 3330 Walnut Street, Philadelphia, PA 19104, USA Received 7 May 2003; received in revised form 27 October 2003; accepted 21 January 2004 Abstract One basic activity in combinatorics is to establish combinatorial identities by so-called ‘bijective proofs,’ which consists in constructing explicit bijections between two types of the combinatorial objects under consideration. We show how such bijective proofs can be established in a systematic way from the ‘lattice properties’ of partition ideals, and how the desired bijections are computed by means of multiset rewriting, for a variety of combinatorial problems involving partitions. In particular, we fully characterizes all equinumerous partition ideals with ‘disjointly supported’ complements. This geometrical characterization is proved to automatically provide the desired bijection between partition ideals but in terms of the minimal elements of the order ÿlters, their complements. As a corollary, a new transparent proof, the ‘bijective’ one, is given for all equinumerous classes of the partition ideals of order 1 from the classical book “The Theory of Partitions” by G.Andrews. Establishing the required bijections involves two-directional reductions technique novel in the sense that forward and backward application of rewrite rules heads, respectively, for two di?erent normal forms (representing the two combinatorial types). It is well-known that non-overlapping multiset rules are con@uent.
    [Show full text]
  • Diagrams of Affine Permutations and Their Labellings Taedong
    Diagrams of Affine Permutations and Their Labellings by Taedong Yun Submitted to the Department of Mathematics in partial fulfillment of the requirements for the degree of Doctor of Philosophy at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY June 2013 c Taedong Yun, MMXIII. All rights reserved. The author hereby grants to MIT permission to reproduce and to distribute publicly paper and electronic copies of this thesis document in whole or in part in any medium now known or hereafter created. Author.............................................................. Department of Mathematics April 29, 2013 Certified by. Richard P. Stanley Professor of Applied Mathematics Thesis Supervisor Accepted by . Paul Seidel Chairman, Department Committee on Graduate Theses 2 Diagrams of Affine Permutations and Their Labellings by Taedong Yun Submitted to the Department of Mathematics on April 29, 2013, in partial fulfillment of the requirements for the degree of Doctor of Philosophy Abstract We study affine permutation diagrams and their labellings with positive integers. Bal- anced labellings of a Rothe diagram of a finite permutation were defined by Fomin- Greene-Reiner-Shimozono, and we extend this notion to affine permutations. The balanced labellings give a natural encoding of the reduced decompositions of affine permutations. We show that the sum of weight monomials of the column-strict bal- anced labellings is the affine Stanley symmetric function which plays an important role in the geometry of the affine Grassmannian. Furthermore, we define set-valued balanced labellings in which the labels are sets of positive integers, and we investi- gate the relations between set-valued balanced labellings and nilHecke words in the nilHecke algebra. A signed generating function of column-strict set-valued balanced labellings is shown to coincide with the affine stable Grothendieck polynomial which is related to the K-theory of the affine Grassmannian.
    [Show full text]
  • Special Unitary Group - Wikipedia
    Special unitary group - Wikipedia https://en.wikipedia.org/wiki/Special_unitary_group Special unitary group In mathematics, the special unitary group of degree n, denoted SU( n), is the Lie group of n×n unitary matrices with determinant 1. (More general unitary matrices may have complex determinants with absolute value 1, rather than real 1 in the special case.) The group operation is matrix multiplication. The special unitary group is a subgroup of the unitary group U( n), consisting of all n×n unitary matrices. As a compact classical group, U( n) is the group that preserves the standard inner product on Cn.[nb 1] It is itself a subgroup of the general linear group, SU( n) ⊂ U( n) ⊂ GL( n, C). The SU( n) groups find wide application in the Standard Model of particle physics, especially SU(2) in the electroweak interaction and SU(3) in quantum chromodynamics.[1] The simplest case, SU(1) , is the trivial group, having only a single element. The group SU(2) is isomorphic to the group of quaternions of norm 1, and is thus diffeomorphic to the 3-sphere. Since unit quaternions can be used to represent rotations in 3-dimensional space (up to sign), there is a surjective homomorphism from SU(2) to the rotation group SO(3) whose kernel is {+ I, − I}. [nb 2] SU(2) is also identical to one of the symmetry groups of spinors, Spin(3), that enables a spinor presentation of rotations. Contents Properties Lie algebra Fundamental representation Adjoint representation The group SU(2) Diffeomorphism with S 3 Isomorphism with unit quaternions Lie Algebra The group SU(3) Topology Representation theory Lie algebra Lie algebra structure Generalized special unitary group Example Important subgroups See also 1 of 10 2/22/2018, 8:54 PM Special unitary group - Wikipedia https://en.wikipedia.org/wiki/Special_unitary_group Remarks Notes References Properties The special unitary group SU( n) is a real Lie group (though not a complex Lie group).
    [Show full text]
  • Chapter 1 Group and Symmetry
    Chapter 1 Group and Symmetry 1.1 Introduction 1. A group (G) is a collection of elements that can ‘multiply’ and ‘di- vide’. The ‘multiplication’ ∗ is a binary operation that is associative but not necessarily commutative. Formally, the defining properties are: (a) if g1, g2 ∈ G, then g1 ∗ g2 ∈ G; (b) there is an identity e ∈ G so that g ∗ e = e ∗ g = g for every g ∈ G. The identity e is sometimes written as 1 or 1; (c) there is a unique inverse g−1 for every g ∈ G so that g ∗ g−1 = g−1 ∗ g = e. The multiplication rules of a group can be listed in a multiplication table, in which every group element occurs once and only once in every row and every column (prove this !) . For example, the following is the multiplication table of a group with four elements named Z4. e g1 g2 g3 e e g1 g2 g3 g1 g1 g2 g3 e g2 g2 g3 e g1 g3 g3 e g1 g2 Table 2.1 Multiplication table of Z4 3 4 CHAPTER 1. GROUP AND SYMMETRY 2. Two groups with identical multiplication tables are usually considered to be the same. We also say that these two groups are isomorphic. Group elements could be familiar mathematical objects, such as numbers, matrices, and differential operators. In that case group multiplication is usually the ordinary multiplication, but it could also be ordinary addition. In the latter case the inverse of g is simply −g, the identity e is simply 0, and the group multiplication is commutative.
    [Show full text]
  • Arxiv:1305.5974V1 [Math-Ph]
    INTRODUCTION TO SPORADIC GROUPS for physicists Luis J. Boya∗ Departamento de F´ısica Te´orica Universidad de Zaragoza E-50009 Zaragoza, SPAIN MSC: 20D08, 20D05, 11F22 PACS numbers: 02.20.a, 02.20.Bb, 11.24.Yb Key words: Finite simple groups, sporadic groups, the Monster group. Juan SANCHO GUIMERA´ In Memoriam Abstract We describe the collection of finite simple groups, with a view on physical applications. We recall first the prime cyclic groups Zp, and the alternating groups Altn>4. After a quick revision of finite fields Fq, q = pf , with p prime, we consider the 16 families of finite simple groups of Lie type. There are also 26 extra “sporadic” groups, which gather in three interconnected “generations” (with 5+7+8 groups) plus the Pariah groups (6). We point out a couple of physical applications, in- cluding constructing the biggest sporadic group, the “Monster” group, with close to 1054 elements from arguments of physics, and also the relation of some Mathieu groups with compactification in string and M-theory. ∗[email protected] arXiv:1305.5974v1 [math-ph] 25 May 2013 1 Contents 1 Introduction 3 1.1 Generaldescriptionofthework . 3 1.2 Initialmathematics............................ 7 2 Generalities about groups 14 2.1 Elementarynotions............................ 14 2.2 Theframeworkorbox .......................... 16 2.3 Subgroups................................. 18 2.4 Morphisms ................................ 22 2.5 Extensions................................. 23 2.6 Familiesoffinitegroups ......................... 24 2.7 Abeliangroups .............................. 27 2.8 Symmetricgroup ............................. 28 3 More advanced group theory 30 3.1 Groupsoperationginspaces. 30 3.2 Representations.............................. 32 3.3 Characters.Fourierseries . 35 3.4 Homological algebra and extension theory . 37 3.5 Groupsuptoorder16..........................
    [Show full text]
  • Affine Permutations and Rational Slope Parking Functions
    AFFINE PERMUTATIONS AND RATIONAL SLOPE PARKING FUNCTIONS EUGENE GORSKY, MIKHAIL MAZIN, AND MONICA VAZIRANI ABSTRACT. We introduce a new approach to the enumeration of rational slope parking functions with respect to the area and a generalized dinv statistics, and relate the combinatorics of parking functions to that of affine permutations. We relate our construction to two previously known combinatorial constructions: Haglund’s bijection ζ exchanging the pairs of statistics (area, dinv) and (bounce, area) on Dyck paths, and the Pak-Stanley labeling of the regions of k-Shi hyperplane arrangements by k-parking functions. Essentially, our approach can be viewed as a generalization and a unification of these two constructions. We also relate our combinatorial constructions to representation theory. We derive new formulas for the Poincar´epolynomials of certain affine Springer fibers and describe a connection to the theory of finite dimensional representations of DAHA and nonsymmetric Macdonald polynomials. 1. INTRODUCTION Parking functions are ubiquitous in the modern combinatorics. There is a natural action of the symmetric group on parking functions, and the orbits are labeled by the non-decreasing parking functions which correspond natu- rally to the Dyck paths. This provides a link between parking functions and various combinatorial objects counted by Catalan numbers. In a series of papers Garsia, Haglund, Haiman, et al. [18, 19], related the combinatorics of Catalan numbers and parking functions to the space of diagonal harmonics. There are also deep connections to the geometry of the Hilbert scheme. Since the works of Pak and Stanley [29], Athanasiadis and Linusson [5] , it became clear that parking functions are tightly related to the combinatorics of the affine symmetric group.
    [Show full text]
  • Rational Catalan Combinatorics 1
    Rational Catalan Combinatorics Drew Armstrong et al. University of Miami www.math.miami.edu/∼armstrong March 18, 2012 This talk will advertise a definition. Here is it. Definition Let x be a positive rational number written as x = a=(b − a) for 0 < a < b coprime. Then we define the Catalan number 1 a + b (a + b − 1)! Cat(x) := = : a + b a; b a!b! !!! Please note the a; b-symmetry. This talk will advertise a definition. Here is it. Definition Let x be a positive rational number written as x = a=(b − a) for 0 < a < b coprime. Then we define the Catalan number 1 a + b (a + b − 1)! Cat(x) := = : a + b a; b a!b! !!! Please note the a; b-symmetry. This talk will advertise a definition. Here is it. Definition Let x be a positive rational number written as x = a=(b − a) for 0 < a < b coprime. Then we define the Catalan number 1 a + b (a + b − 1)! Cat(x) := = : a + b a; b a!b! !!! Please note the a; b-symmetry. Special cases. When b = 1 mod a ::: I Eug`eneCharles Catalan (1814-1894) (a < b) = (n < n + 1) gives the good old Catalan number n 1 2n + 1 Cat(n) = Cat = : 1 2n + 1 n I Nicolaus Fuss (1755-1826) (a < b) = (n < kn + 1) gives the Fuss-Catalan number n 1 (k + 1)n + 1 Cat = : (kn + 1) − n (k + 1)n + 1 n Special cases. When b = 1 mod a ::: I Eug`eneCharles Catalan (1814-1894) (a < b) = (n < n + 1) gives the good old Catalan number n 1 2n + 1 Cat(n) = Cat = : 1 2n + 1 n I Nicolaus Fuss (1755-1826) (a < b) = (n < kn + 1) gives the Fuss-Catalan number n 1 (k + 1)n + 1 Cat = : (kn + 1) − n (k + 1)n + 1 n Special cases.
    [Show full text]