ebook img

Spin $q$-Whittaker polynomials PDF

0.5 MB·
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Spin $q$-Whittaker polynomials

SPIN q–WHITTAKER POLYNOMIALS ALEXEI BORODIN AND MICHAEL WHEELER Abstract. We introduce and study a one-parameter generalization of the q–Whittaker symmetric functions. This is a family of multivariate symmetric polynomials, whose con- 7 1 structionmay be viewedas anapplicationof the procedureof fusion fromintegrablelattice 0 modelstoavertexmodelinterpretationofaone-parametergeneralizationofHall–Littlewood 2 polynomials from [Bor17, BP16a, BP16b]. n We prove branching and Pieri rules, standard and dual (skew) Cauchy summation iden- a tities, and an integral representation for the new polynomials. J 3 2 ] Contents O C 1. Introduction 1 h. 2. Higher spin vertex model 6 t 3. Spin Hall–Littlewood functions 9 a m 4. Fusion 15 [ 5. Spin q–Whittaker polynomials 20 6. Combinatorial formulae 30 1 v 7. Cauchy identities and Pieri rules 34 2 8. Integral representation of spin q–Whittaker polynomials 38 9 2 References 40 6 0 . 1 0 1. Introduction 7 1 1.1. Background. The q-deformed class one gl Whittaker functions, or q–Whittaker func- : n v tions for short, were defined in [GLO10] as a special class of joint eigenfuctions of the i X q-deformed Toda chain Hamiltonians [Eti99, Rui90] with the support in the positive Weyl r chamber. In the limit q → 1 they reduce to classical gl Whittaker functions [GLO12], while a n for general q they are themselves the limiting case of a more general family of symmetric functions, the Macdonald polynomials [Mac95]. In the last decade they have come to play a prominent role in integrable probability via the q–Whittaker processes [BC14, BP14]. These are a rather general class of stochastic processes, which not only degenerate at q → 1 to the Whittaker processes of O’Connell [O’C12] used to describe random directed polymers but, in a different direction, to the continuous time q-deformed totally asymmetric simple exclu- sion process (q–TASEP). For a survey of these properties, and their connection with KPZ (Kardar–Parisi–Zhang) universality, we refer the reader to [BC14, Chapters 3–5], [BP14]. A somewhat perpendicular approach to KPZ-type observables has recently been proposed in [BP16a, BP16b], in the setting of a higher spin version of the six-vertex model from statistical mechanics [Bax07]. The chief object of study in these works has been the height function of the vertex model, a random (non-negative, integer-valued) variable which lives 1 fusion + (t7→q) Spin Spin Macdonald Hall–Littlewood q–Whittaker (q,t) (s,t) (s,q) s=0 q =0 t=0 s=0 Hall–Littlewood q–Whittaker (t) (q) Figure 1. A table of symmetric functions, indicating their parameter-dependence. on the vertices of the lattice. Once again, the theory of symmetric functions turns out to be indispensable to the calculations: a family of symmetric, rational functions and their associated orthogonality relations play a pivotal role in writing exact integral expressions for the q-moments of the height function. These rational functions were originally introduced in [Bor17] as a one-parameter deformation of the Hall–Littlewood functions [Mac95] (which are yet another family encompassed by the more general Macdonald polynomials). In this work we refer to the former rational functions as spin Hall–Littlewood functions, in reference to the spin parameter s which they carry. Even more recently, some direct equivalences between expectations of observables in Mac- donald processes (and their degenerations) and those in higher spin vertex models have been remarked [Bor16], see also [OP16]. While these equivalences (once guessed) can be read- ily proved by comparing integral formulae, their origin remains fairly mysterious. A more conceptual version of the equivalence [Bor16], albeit in the limit to Hall–Littlewood pro- cesses, has now been exposed [BBW16]. A natural angle towards better understanding these equivalences is to search for a family of symmetric functions which unifies all of the classes listed above, and to study in detail the properties of such a family. This remains beyond the scope of the present work, although we mention that promising steps in this direction have been made in [GdGW17], where a mutual generalization of Macdonald polynomials and spin Hall–Littlewood functions was explicitly constructed. In fact, rather than unifying the known families, in this paper we will add one more family to the list: these we term spin q–Whittaker polynomials. The name is chosen to reflect the fact that they are a one-parameter, s-deformation of q–Whittaker polynomials; they degenerate to the latter at s = 0. This means that the spin q–Whittaker polynomials have a similar footing in the general theory as do the spin Hall–Littlewood functions, but now on the t = 0 side of the Macdonald “coin” (whereas the spin Hall–Littlewood functions lie on the q = 0 side). We refer the reader to Figure 1 for a survey of the symmetric function landscape. This analogy between the spin Hall–Littlewood functions and spin q–Whittaker polyno- mials is not accidental: one can construct the latter directly from the former, using the 2 technique of fusion in integrable lattice models [KRS81, KR87]. To see how this can tran- spire, consider the right hand side of the Cauchy identity for Hall–Littlewood polynomials (here we use a parameter q, rather than the conventional t), m n 1−qu v i j Π(u ,u ,...,u ;v ,v ,...,v ) := , 1 2 m 1 2 n 1−u v i j i=1 j=1 YY and specialize one set of these variables to a geometric progression with ratio q, letting (u ,u ,...,u ) = (u,qu,...,qm−1u). We obtain 1 2 m n 1−qmuv Π(u,qu,...,qm−1u;v ,v ,...,v ) = j . 1 2 n 1−uv j j=1 Y After analytic continuation in qm effected by sending qmu 7→ −x, and setting u = 0, the product turns into n (1 + xv ). This we recognize as the right hand side of the dual j=1 j Cauchy identity between a Hall–Littlewood polynomial and a q–Whittaker polynomial (in Q one variable). On the other hand, such geometric specializations are known to be a key ingredient in fusion. It is therefore reasonable to suggest that q–Whittaker polynomials might be recovered by applying the fusion technique to the lattice model construction of Hall–Littlewood polynomials. This indeed turns out to be the case; surprisingly, we were unable to find a precise statement of this fact in the literature. In light of this observation, one can play the same game taking the spin Hall–Littlewood functions (and their lattice model construction, developed in [Bor17, BP16a, BP16b]) as the starting point: doing so ultimately leads us to our definition of the spin q–Whittaker polynomials. As is common in the theory of symmetric functions [Mac95], we define skew spin q– Whittaker polynomials F (x ,...,x ) parametrized by pairs of partitions λ,ν, as well as λ/ν 1 m non-skew ones F (x ,...,x ) parametrized by single partitions λ; the latter correspond to λ 1 m taking ν = ∅ in the former. Let us list a few of their properties that we prove below. • F (x ,...,x ) is a symmetric, inhomogeneous polynomial in (x ,...,x ). λ/ν 1 m 1 m • For any pair of partitions λ,ν the following stability relation holds: F (x ,...,x ,−s) = F (x ,...,x ). λ/ν 1 m−1 λ/ν 1 m−1 At s = 0 this recovers the usual stability property of q–Whittaker polynomials. • The spin q–Whittaker polynomials satisfy a standard branching rule F (x ,...,x ) = F (x ,...,x )F (x ,...,x ). λ/µ 1 m+n ν/µ 1 m λ/ν m+1 m+n ν X • The one-variable skew spin q–Whittaker polynomial is given by (−s/x;q) (−sx;q) (q;q) x|λ|−|ν| λi−νi νi−λi+1 λi−λi+1, λ ≻ ν, (q;q) (q;q) (s2;q) F (x) =  i>1 λi−νi νi−λi+1 λi−λi+1 λ/ν Y    0, otherwise. Together withthe branching rule, this yields a simple “interlacing” construction of the spin  q–Whittaker polynomials (or equivalently, one in terms of Gelfand–Tsetlin patterns). 3 • Define the dual spin q–Whittaker polynomials by (s2;q) (q;q) F∗ (x ,...,x ) = λi−λi+1 νi−νi+1 ·F (x ,...,x ). λ/ν 1 m (q;q) (s2;q) λ/ν 1 m i≥1 λi−λi+1 νi−νi+1 Y Then the following skew Cauchy-type summation identities hold: For arbitrary partitions µ and ν, one has F (x ,...,x )F∗ (y ,...,y ) λ/µ 1 m λ/ν 1 n λ X m n (−sx ;q) (−sy ;q) = i ∞ j ∞ F (x ,...,x )F∗ (y ,...,y ). (s2;q) (x y ;q) ν/κ 1 m µ/κ 1 n ∞ i j ∞ i=1 j=1 κ YY X For µ = ν = ∅, this turns into a Cauchy-type identity (s2;q) m n (−sx ;q) (−sy ;q) F (x ,...,x )F (y ,...,y ) λi−λi+1 = i ∞ j ∞, λ 1 m λ 1 n (q;q) (s2;q) (x y ;q) λ i>1 λi−λi+1 i=1 j=1 ∞ i j ∞ X Y YY which is a multivariate generalization of the q–Gauss summation theorem. Therearealsodual(skew) Cauchy identitiesthatcombinespinq–Whittakerpolynomials and stable spin Hall–Littlewood functions, see Section 7.3 below. • The following integral representation holds: F (x ,...,x ) λ 1 m du1 duL ui −uj L 1−sui λ′i mj=1(1+uixj) = ··· , 2πiu 2πiu u −qu u −s (1−su )m+1 IC 1 IC L 16i<j6L(cid:18) i j(cid:19)i=1(cid:18) i (cid:19) Q i ! Y Y where L = λ denotes the largest part of λ and the contour C is as in Figure 6 below. 1 The vertex model interpretation of the spin Hall–Littlewood functions makes it natural to add countably many inhomogeneity parameters to their definition; most of their properties remain intact [BP16a]. One can do the same for the spin q–Whittaker polynomials; the branching rule remains the same, and in the formula for the one-variable specialization above one needs to add index i to the spin parameter s. This preserves the (skew) dual Cauchy identities (that need to involve stable spin Hall–Littlewood functions with the same inhomogeneities), but appears to destroy the ordinary (skew) Cauchy identities. There is also an integral representation that generalizes the one above. We decided to leave the inhomogeneous case out of this work in order not to cloud the arguments with more involved notation. 1.2. Layout of paper. In Section 2 we recall the basic features of the integrable higher spin vertex model studied in [Bor17, BP16a, BP16b], and then in Section 3 we use it to define the spin Hall–Littlewood rational functions as lattice model partition functions. In Section 4 we gather a number of results on the fusion procedure as applied to the vertex model of Section 2, leading to the construction of integrable Boltzmann weights in equation (35) that form the foundation of the remainder of the paper. We remark that, at s = 0, these weights degenerate into those used by Korff in Sections 3 and 6 of [Kor13]. This is expected, since 4 [Kor13] contains an integrable lattice construction of the q–Whittaker polynomials which is exactly the s = 0 specialization of our results here. In Section 5 we apply the fusion procedure to the spin Hall–Littlewood functions, and use the resulting partition functions to define spin q–Whittaker polynomials. Along the way, we also define a stable version of the spin Hall–Littlewood functions, which are later paired with the spin q–Whittaker polynomials in dual Cauchy summation identities. We formulate the spinq–Whittakerpolynomialsalgebraicallyusingcertainmonodromymatrixoperatorsinthe model (35), and derive key commutation relations between these operators. In Section 6 we use the lattice construction of the spin q–Whittaker polynomials to derive their branching rules, and show that they reduce to (ordinary) q–Whittaker polynomials by specializing s = 0. Section 7 contains a list of various Cauchy, dual Cauchy and Pieri identities which thespinq–WhittakerpolynomialsandthestablespinHall–Littlewoodfunctionssatisfy; these are all proven by means of the commutation relations derived in Section 5. We conclude, in Section 8, with a multiple integral expression for the the spin q–Whittaker polynomials. 1.3. Notation. Throughout the paper we use a number of partition-related terminologies, which we summarize below. These are all standard in the combinatorics literature, with one exception: we include parts of size zero in our partitions, rather than the usual practice of truncating a partition after its last positive part. Apartitionλisafinitenon-increasing sequence ofnon-negativeintegersλ > ··· > λ > 0, 1 ℓ where λ is called a part of λ for all 1 6 i 6 ℓ. The empty partition ∅ is the trivial sequence i consisting of no parts. The length of a partition λ is the number of parts which comprise it, and denoted by ℓ(λ). We let Part denote the set {λ > ··· > λ > 0} of all partitions ℓ 1 ℓ of length ℓ. Similarly, Part+ will denote the set {λ > ··· > λ > 1} of all partitions ℓ 1 m 06m6ℓ with purely positive parts, whose length is bounded by ℓ. It is sometimes convenient to write a partition in terms of its part-multiplicities m (λ) = #{j : λ = i}, i.e. by writing i j λ = 0m01m12m2... where m ≡ m (λ). The conjugate of a partition λ, denoted λ′, is the i i partition with parts λ′ = #{j : λ > i}. i j For two positive partitions λ,µ we write λ ⊃ µ if ℓ(λ) > ℓ(µ) and the inequality λ > µ i i holds for all 1 6 i 6 ℓ(µ). Similarly, we write λ ≻ µ and say that λ interlaces µ if 0 6 ℓ(λ) − ℓ(µ) 6 1 and λ > µ > λ for all 1 6 i 6 ℓ(µ) (where the final inequality i i i+1 µ > λ is omitted in the case ℓ(λ) = ℓ(µ)). ℓ(µ) ℓ(µ)+1 We make frequent use of q–Pochhammer symbols, which are defined as follows: (1−aqi−1), m > 0, 16i6m  Q (a;q) = 1, m = 0, m      (1−aq−i)−1, m < 0. 16i6−m   We will also tacitly assume that|qQ| < 1 so that, in particular, (a;q)∞ makes sense. Most of our equations depend on another parameter s; we will also assume that |s| < 1, since this prevents s from taking values which would lead to divergences because of vanishing denominators. 1.4. Acknowledgments. A. B. is supported by the National Science Foundation grant DMS-1607901 and by Fellowships of the Radcliffe Institute for Advanced Study and the Si- monsFoundation. M.W.issupportedbytheAustralianResearchCouncilgrantDE160100958. 5 2. Higher spin vertex model 2.1. Vertex weights and Yang–Baxter equation. Following Section 2 of [BP16b], we define a vertex model consisting of SW → NE oriented paths on a square grid. Horizontal edges of the grid can be occupied by at most one lattice path, but no restriction is placed on the number of paths which traverse a vertical edge. Every intersection of horizontal and vertical gridlines constitutes a vertex, and each vertex is assigned a Boltzmann weight that depends on the local configuration of lattice paths about that intersection. Assuming conservation of lattice paths through a vertex, four types of vertex are possible. We indicate these vertices and their explicit weights below: g g g +1 g 0 0 0 1 1 0 1 1 g g +1 g g (1) w (g,0;g,0) w (g +1,0;g,1) w (g,1;g+1,0) w (g,1;g,1) u u u u 1−sqgu (1−s2qg)u 1−qg+1 u−sqg 1−su 1−su 1−su 1−su where g is any non-negative integer (representing the number of paths which sit at a vertical edge), s and q are fixed global parameters of the model, and u is a local variable called the spectral parameter. We denote the Boltzmann weight of a vertex in two equivalent ways, and interchange between the two according to convenience: k (2) wu j ℓ ≡ wu(i,j;k,ℓ), i,k ∈ Z>0, 0 6 j,ℓ 6 1.   i   It is conventional to relax the constraint of lattice path conservation through each vertex, whichcanbedonebyextending theBoltzmannweights(2)toallvaluesofi,j,k,ℓandassum- ing that w (i,j;k,ℓ) = 0 for all i+j 6= k+ℓ. The common factor 1−su in the denominator u of each vertex (1) is to ensure that the empty vertex has weight 1, i.e. w (0,0;0,0) = 1. u Define an n-vertex by concatenating n vertices vertically, with summation assumed over all internal vertical edges: k   j ℓ n n . . (3) w  .. ..  ≡ w i,{j ,...,j };k,{ℓ ,...,ℓ } {u1,...,un}  {u1,...,un} 1 n 1 n  j1 ℓ1  (cid:16) (cid:17)      i      It is easily seen that the weight of an n-vertex (3) is always factorized into weights of individual vertices (1): knowing three of the edge states surrounding a vertex determines the fourth by lattice path conservation, and this constrains each of the internal edges in (3) to assume a unique value (or it causes the whole weight to vanish if i + j + ··· + j 6= 1 n k +ℓ +···+ℓ , when conservation is impossible). 1 n 6 Definition 2.1. Let W = Span{|ji} ∼= C2 be a two-dimensional vector space, and for 06j61 all 1 6 i 6 n let W denote a copy of W. The n-vertex operator W (i;k) acts linearly i {u1,...,un} on W ⊗···⊗W as follows: 1 n W (i;k) : |ℓ i ⊗···⊗|ℓ i {u1,...,un} 1 1 n n 7→ w i,{j ,...,j };k,{ℓ ,...,ℓ } |j i ⊗···⊗|j i . {u1,...,un} 1 n 1 n 1 1 n n 06j1X,...,jn61 (cid:16) (cid:17) This in fact defines an infinite family of operators, since i and k can be any non-negative integers. Proposition 2.2. Let W (i;k) ∈ End(W ⊗ W ) be a 2-vertex operator as defined {u1,u2} 1 2 above, with i,k ∈ Z . The Yang–Baxter equation holds: >0 (4) P ◦R(u /u )◦W (i;k) = W (i;k)◦P ◦R(u /u ), 2 1 {u1,u2} {u2,u1} 2 1 where the R-matrix is given by 1−qu 0 0 0 0 q(1−u) 1−q 0 (5) R(u) =   ∈ End(W ⊗W ), 0 (1−q)u 1−u 0 1 2  0 0 0 1−qu     and P ∈ End(W ⊗W ) is the permutation operator, with action P : |ai ⊗|bi 7→ |bi ⊗|ai 1 2 1 2 1 2 for all vectors |ai,|bi ∈ W. Proof. By direct computation, using the vertex weights (1) and the explicit realization of the permutation operator, 1 0 0 0 0 0 1 0 P =   ∈ End(W ⊗W ). 0 1 0 0 1 2 0 0 0 1     See also Section 2 of [Bor17] for the derivation of (4) from the U (sl ) Yang–Baxter equation q 2 in the tensor product W ⊗W ⊗V , where V denotes a highest weight representation with 1 2 I I weight I. c (cid:3) 2.2. Dual vertex weights. We will at times adopt an alternative convention for the vertex weights (1). The change in convention is brought about inverting the spectral parameter u in (1), then multiplying all vertices by (u−s)/(1−su). Since the Yang–Baxter equation (4) is preserved under this transformation (up to inversion of u and u ), this does not damage 1 2 the integrability of the model, although it will be essential to ensure that certain infinite partition functions which we study are well-defined. The alternative weights are as shown 7 below: g g g +1 g 0 0 0 1 1 0 1 1 (6) g g +1 g g u−sqg 1−s2qg (1−qg+1)u 1−sqgu 1−su 1−su 1−su 1−su Graphically, we distinguish such vertices from their counterparts (1) by using a coloured background. If we complement the states that live on horizontal edges (i.e. draw a path if the edge is unoccupied, delete a path if the edge is occupied), we find that the vertices (6) are converted to g g g +1 g 1 1 1 0 0 1 0 0 g g +1 g g (7) w∗(g,1;g,1) w∗(g +1,1;g,0) w∗(g,0;g+1,1) w∗(g,0;g,0) u u u u u−sqg 1−s2qg (1−qg+1)u 1−sqgu 1−su 1−su 1−su 1−su where we have also reversed the orientation of all paths, so that they propagate NW → SE. In this way we define a set of dual vertex weights w∗(i,j;k,ℓ), whose non-zero values u are indicated in (7). There is a strong similarity between the vertices (7) and those of the starting vertex model (1). Indeed, by reflecting the vertices (7) about their central horizontal axis, we almost recover those of (1) – the only difference is that the weights of the two middle vertices have changed slightly. It turns out that this discrepancy can be cured by a simple gauge transformation of the weights. More precisely, find that (q;q) (s2;q) (8) w (i,j;k,ℓ) = k w∗(k,j;i,ℓ) i, i,k ∈ Z , 0 6 j,ℓ 6 1. u (s2;q) u (q;q) >0 k i 2.3. Partition states. Let V = Span{|mi} be an infinite-dimensional vector space, m>0 and for all i > 0 let V denote a copy of V. Further, let V = ⊗ V , the global vector space i i>0 i obtained by tensoring the local spaces V for all i > 0. It can be viewed as the span of pure i tensors ⊗ |m i with m ∈ Z , all but finitely many of which are 0. i>0 i i i >0 Consider a partition λ = 0m01m12m2... expressed in terms of its of part multiplicities m (λ). The part multiplicities can be obtained as the difference between adjacent parts in i the conjugate partition λ′, as follows: m (λ) = λ′ −λ′ , i i i+1 since λ′ = m (λ) for all i > 0. To every partition λ we associate a unique state in V: i j>i j (9) P |λi = |m i = |λ′ −λ′ i ∈ V. i i i i+1 i i>0 i>0 O O 8 Similarly, one can define dual partition states hλ| = ⊗ hm | ∈ V∗, with the orthonormal i>0 i i action hλ|µi = δ for all partitions λ,µ. λ,µ Inthecomingsections, wewillstudythevertexmodel(1)onasquarelatticewithinfinitely many columns, labelled from left to right by non-negative integers. In that situation, we shall identify the vector space V with the ith column of the lattice, with |mi encoding a i i vertical edge in that column which is occupied by m paths. In Section 5 we will usually require infinitely many paths on the vertical edges of the 0th column. In this setting, the natural objects are partitions λ with strictly positive parts (i.e. the traditional notion of a partition), since the number of zero parts in λ becomes irrelevant. This leads us to define, for all (positive) partitions λ = 1m12m2 ..., the state (10) |λ⟫ = |m i ∈ V, where V = V , i i i i>1 i>1 O O e e and dual state ⟪λ| = ⊗ hm | ∈ V∗. i>1 i i 3. Spin Heall–Littlewood functions 3.1. Setup of the lattice for F (u ,...,u ). Following [BP16b], we study the vertex model λ 1 ℓ (1) on the quadrant Z ×Z . Horizontal lines are oriented from left to right and numbered >0 >1 from bottom to top, while vertical lines are oriented from bottom to top and numbered from left to right. The ith horizontal line is assigned spectral parameter u , where i ∈ Z . The i >1 boundary conditions are fixed as follows: 1. There is an incoming lattice path at the external left edge of every horizontal line. 2. The external bottom edge of every vertical line is unoccupied. Definition 3.1. Let λ = (λ > ··· > λ > 0) ∈ Part be a partition. The spin Hall– 1 ℓ ℓ Littlewood function F (u ,...,u ) is defined as the partition function of Z ×{1,...,ℓ} in λ 1 ℓ >0 the model (1), whose left and bottom edge boundary conditions are given by 1 and 2 as above, and whose ith external top edge is occupied by exactly m (λ) paths for all i > 0. See i Figure 2, left panel. Remark 3.2. The lattice used in the definition of F (u ,...,u ) has infinitely many vertical λ 1 ℓ lines, but it remains a meaningful definition. Indeed, for a given λ, only the vertical lines numbered 0 to λ will contribute non-trivially to the partition function. It is obvious that 1 all vertical lines beyond this will be unoccupied, giving rise to infinitely many copies of 0 the vertex 0 0 , which has weight 1. A similar remark applies to all infinite partition 0 functions studied in this section. By allowing more general boundary conditions at the base of the lattice, we can extend the definition to skew Young diagrams: Definition 3.3. Let λ = (λ > ··· > λ > 0) ∈ Part and µ = (µ > ··· > µ > 0) ∈ 1 ℓ+n ℓ+n 1 n Part be two partitions. The skew spin Hall–Littlewood function F (u ,...,u ) is defined n λ/µ 1 ℓ as the partition function of Z × {1,...,ℓ} in the model (1), whose left edge boundary >0 satisfies 1 as above, whose ith external top edge is occupied by exactly m (λ) paths, and i whose ith external bottom edge is occupied by exactly m (µ) paths for all i > 0. The i 9 u4 u4 u3 u3 u2 u2 u1 u1 Figure 2. Left panel: boundary conditions used to calculate F (u ,...,u ) λ 1 4 for λ = (4,3,3,1). Right panel: boundary conditions used to calculate F (u ,...,u ) for λ = (5,4,4,2,2,1,0), µ = (4,1,1). λ/µ 1 4 v5 v5 v4 v4 v3 v3 v2 v2 v1 v1 Figure 3. Left panel: boundary conditions used to calculate G (v ,...,v ) λ 1 5 for λ = (5,4,1,0). Right panel: boundary conditions used to calculate G (v ,...,v ) for λ = (4,3,3,1), µ = (3,2,0,0). λ/µ 1 5 (ordinary) spin Hall–Littlewood function F is recovered in the special case µ = ∅. See λ Figure 2, right panel. 3.2. Setup of the lattice for G (v ,...,v ). Here we use an alternative set of boundary λ 1 n conditions: 1. The external left edge of every horizontal line is unoccupied. 2. The external bottom edge of each vertical line is unoccupied, with the exception of the 0th, which is occupied by ℓ incoming paths, for some ℓ > 0. Definition 3.4. Let λ = (λ > ··· > λ > 0) ∈ Part be a partition. The dual spin Hall– 1 ℓ ℓ Littlewood function G (v ,...,v ) is defined as the partition function of Z × {1,...,n} λ 1 n >0 in the model (1), whose left and bottom edge boundary conditions are given by 1 and 2 as above, and whose ith external top edge is occupied by exactly m (λ) paths for all i > 0. i Notice that we impose no relation between the values of ℓ and n. See Figure 3, left panel. Once again, this definition can be extended to skew Young diagrams by allowing for more general boundary conditions at the base of the lattice: Definition 3.5. Let λ = (λ > ··· > λ > 0) ∈ Part and µ = (µ > ··· > µ > 1 ℓ ℓ 1 ℓ 0) ∈ Part be two partitions of equal length. The dual skew spin Hall–Littlewood function ℓ G (v ,...,v ) is defined as the partition function of Z × {1,...,n} in the model (1), λ/µ 1 n >0 whose left edge boundary conditions are given by 1 as above, whose ith external top edge is 10

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.