Department of Statistics Nine month report Exact solvability in directed random polymer models PhD student: Elia Bisi Supervisor: Dr. Nikolaos Zygouras June 25, 2015 Introduction Random polymer models are the object of intense research within probability theory and statistical physics. In particular, this report deals with the theory of directed random polymers immersed in a random potential, which is introduced in section 1. In the last few decades, a few (1+1)-dimensional models have been shown to be exactly solvable, meaning that the probability distributions of quantities such as the polymer partition sum are exactly computable in terms of algebraic objects: links with algebraic combinatorics and representation theory have therefore acquired increasing importance. The exact solvability of such models has turned out to be very useful to prove scaling limit theorems, thus revealing surprising connections to random matrix theory. We analyze two combinatorial algebraic structures that arise in this setting. In section 2, we consider the RSK correspondence, which is related to the zero-temperature limit of polymer models. In section 3, we examine the recently studied geometric RSK correspondence (also referred to as tropical RSK), which is related to finite-temperature polymer models; in this field, we set out possible future developments and open problems. 1 Directed polymers in a random potential The goal of this section is to introduce the statistical mechanics model of directed random polymers immersed in a random potential: polymers are supposed to live in the (1 + d)-dimensional lattice N × Zd and to interact with an environment, being either attracted to or repelled from sites depending on a random potential given by an i.i.d. field on the lattice. We first describe general polymer models, providing the main definitions; then we recall some basic facts about the simple random walk model, which represents the “zero potential” case. Finally, we analyze the phase transition in the model with arbitrary potential: the main results are stated and some proofs are sketched. The main references are [4] and [7]. 1.1 Random polymers From a physical point of view, a polymer is a large molecule consisting of many smaller molecules called monomers and tied together by chemical bonds. Very long concatenated structures are easy to find in nature because of the multivalency of atoms. A polymer is called linear if its monomers only have one reactive group, leading to a linear structure without multiple cross connections: DNA, RNA and proteins are examples of linear polymers. Figure 1. A DNA helix. 1 In Mathematics, linear polymers are modelled as random paths in a lattice (typically Zd ), where vertices represent the monomers and edges represent the chemical bonds connecting the monomers. In this setting, the typical configuration of a polymer varies according to the interactions with itself and the environment it is immersed in. Mathematical research asks how such microscopic interactions determine different macroscopic behaviors, typically in terms of phase transitions, in the limit as the polymer gets long: for instance, spreading out of the polymer’s endpoint as well as its localization are widely studied problems. In each polymer model we consider: • Λ, the lattice where the polymers live; • Xn , a finite set of allowed n-step paths in Λ; • Hn , a (possibly random) Hamiltonian function that associates an energy to each path in Xn ; • Pn , the Gibbs measure on Xn associated to the Hamiltonian Hn and defined by Pn (x) = 1 −Hn (x) e Zn ∀x ∈ Xn , where Zn is its normalizing partition sum. We call Pn polymer measure and we denote by En its associated expectation. The Hamiltonian Hn determines the interaction of the polymer with itself and the environment, and often depends on some parameters such as temperature (or its inverse, denoted by β). According to the intuition, the higher the energy of a path, the lower its probability. In case Hn is chosen to be random, it will depend on a random environment, given by a set of random variables defined on a probability space (Ω, F , P, E). We have then two polymer measures: • the quenched polymer measure, which depends on a fixed configuration ω ∈ Ω of the environment: ω 1 x ∈ Xn ; Pnω (x) = ω e−Hn (x) , Zn • the annealed polymer measure, which is averaged w.r.t. the environment: R ω e−Hn (x) P(dω) E[e−Hn (x) ] ΩR Pn (x) = = , x ∈ Xn , E[Zn ] Znω P(dω) Ω where E[Zn ] is called annealed partition sum. A number of random polymer models have been studied in literature (see [4] for an overview). Here, we will concentrate on directed polymers in a random potential: in this setting, the very first introductory model is the simple random walk on Zd . 2 1.2 The simple random walk model on Zd Let us consider the set Zd with its usual lattice structure: two vertices x and y are connected if the l1 -norm of their difference is 1 (we will write x ∼ y). We set n o Xn = x = (xi )ni=0 ∈ (Zd )n+1 : x0 = 0, xi−1 ∼ xi ∀1 ≤ i ≤ n . We consider Hn = 0, so that Pn is the uniform distribution on Xn , i.e. Pn (x) = (2d)−n for all x. These polymers clearly correspond to the paths of a simple random walk (SRW) in Zd . Using the fact that the endpoint Xn of the polymer is a sum of n i.i.d. increments under Pn , one can easily compute its mean and variance: En [Xn ] = 0, En [|Xn |2 ] = n ∀n ≥ 1 . The behavior of these polymers is diffusive: the continuous-time process interpolating the polymer linearly and defined by Xt := Xbtc + (t − btc)(Xbtc+1 − Xbtc ) ∀t ∈ [0, ∞) , if suitably rescaled, converges in distribution to a standard Brownian Motion: ! n→∞ 1 −−−−−* (Bt )0≤t≤1 √ Xbntc n 0≤t≤1 on the Banach space C[0, 1] of continuous functions on [0, 1]. This follows from the well-known Donsker’s Invariance Principle (see [17, §5.3]). For the SRW model we also have a Local Limit Theorem, which ensures that the position Xn of the polymer’s endpoint spreads out as n → ∞: 1 max Pn (Xn = x) = O d/2 . (1.1) n x∈Zd This can be deduced from the following facts: • If d = 1, then = 0 q Pn (Xn = x) ∼ 2 πn if x . n mod 2 as n → ∞ if x ≡ n mod 2 , where the asymptotics follow from a simple combinatorial argument and the Stirling approximation. (1) (d) • For all d ≥ 1, by independence of the components Xn , . . . , Xn max Pn (Xn = x) = x∈Zd d Y i=1 (i) max Pn Xn = x(i) , x(i) ∈Z and each component is distributed as a SRW on Z. 3 1.3 Directed polymers in a random potential In this section we extend the SRW model by allowing the polymers to be immersed in an environment characterized by the presence of a random potential depending on both time and space. For instance, we may think of hydrophilic polymers wafting in water: if water contains randomly placed hydrophobic molecules that repel the monomers, then the polymers can be viewed as immersed in a random potential. We now define the model: polymers now live in the lattice Λ = N × Zd , i.e. in 1 + d dimensions, where the first coordinate represents time and the others represent space. We define the set of allowed paths by n o Xn = x = (i, xi )ni=0 ∈ (N × Zd )n+1 : x0 = 0, xi−1 ∼ xi ∀1 ≤ i ≤ n . The potential is given by a random field of real valued non-degenerate i.i.d. variables {V (i, x) : i ∈ N, x ∈ Zd } defined on a probability space (Ω, F , P) (expectation denoted by E), with moment generating function h i M(β) = E eβV (1,0) < ∞ ∀β ∈ [0, ∞) . The Hamiltonian is given by β,ω Hn (x) = −β n X V (i, xi ) , x ∈ Xn , i=1 for some choice of β ∈ [0, ∞), which should be thought of as inverse temperature. The β,ω associated quenched polymer measure Pn is defined in terms of the law Pn of the n-step SRW on Zd β,ω 1 β,ω Pn (x) = β,ω e−Hn (x) Pn (x) , x ∈ Xn , Zn β,ω so that the partition sum Zn β,ω Zn β,ω can be expressed as the expectation of e−Hn w.r.t. Pn : = β,ω X e−Hn (x) h βi Pn (x) = En e−Hn . (1.2) x∈Xn As the picture of these paths in N × Zd suggests (see Figure 2), these polymers are called directed. β β β From now on, for the sake of conciseness we will write Hn (x), Pn (x) and Zn instead β,ω β,ω β,ω of Hn (x), Pn (x) and Zn ; however, we stress that these quantities depend on the configuration of the environment, i.e. they are random variables (Ω, F , P) → R. Remark 1.1. The case β = 0 corresponds to the SRW model we have already gone through. Let us see the behavior of the polymer partition sum in the zero temperature limit, i.e. as β → ∞: β Zn −n = (2d) X x∈Xn exp β n X ! −n V (i, xi ) (2d) i=1 4 exp β max x∈Xn n X i=1 ! V (i, xi ) . Figure 2. A (1 + 1)-dim. directed polymer in a random environment. Different vertex colors denote different values of the potential. This establishes a deep connection between the polymer model we are describing and another important model in statistical physics: indeed, Tn∗ := max x∈Xn n X V (i, xi ) = lim 1 β→∞ β i=1 β log Zn (1.3) turns out to be the maximal passage time in the context of directed last passage percolation. We will resume this concept in subsection 2.3. ♦ The main tool in the analysis of these polymers is the so-called normalized partition sum, i.e. the ratio of the quenched and the annealed partition sum: β β Zn Zn β = En [en ] Wn := h β i = M(β)n E Zn where en (x) = n Y eβV (i,xi ) i=1 M(β) ∀n ≥ 0 , (1.4) ∀x ∈ Xn . β The process (Wn )n≥0 turns out to be a positive martingale w.r.t. the natural filtration of the environment (Fn )n≥0 , defined by Fn := σ V (i, x) : 1 ≤ i ≤ n, x ∈ Zd . This follows from the fact that the environment is given in terms of an i.i.d. random field: indeed, en is an (Fn )n≥0 -martingale as a product of mean-one i.i.d. random variables β on (Ω, F , P), so Wn , which is the average of en w.r.t. En (as (1.4) shows), also is. By the β martingale convergence theorem, Wn converges P-a.s., so we can define β β W∞ := lim Wn n→∞ 5 P-a.s. . β Since the event {W∞ = 0} belongs to the tail sigma-algebra of (Fn )n≥0 , the Kolmogorov zero-one law ensures that its probability is either 0 or 1, i.e. one of the two following possibilities holds: β W∞ > 0 P-a.s. , (1.5a) β W∞ P-a.s. . (1.5b) =0 h β 1/2 i Lemma 1.2 (Comets, Yoshida [8]). The fractional moment E W∞ is non-increasing on [0, ∞) as a function of β. h h i i 0 )1/2 = 1, so the latter lemma implies that E W β 1/2 Since Wn0 = 1 for all n, E (W∞ ∞ is strictly positive up to a critical value βc and zero from that value on; of course, the cases βc = 0 and βc = ∞ are also possible. This allows to highlight a phase transition in β from (1.5a) to (1.5b): Theorem 1.3. For any fixed d ≥ 1 and any fixed potential distribution, there exists βc ∈ [0, ∞] such that β W∞ > 0 P-a.s. ∀β ∈ [0, βc ) , β W∞ P-a.s. ∀β ∈ (βc , ∞) . =0 The two phases (1.5a) and (1.5b) are called weak disorder and strong disorder respectively, because the role of the disorder is intuitively significant only when the annealed β partition sum grows faster than the quenched partition sum, i.e. when W∞ = 0. A priori, one cannot deduce from Lemma 1.2 which phase the critical value βc corresponds to. In the next section we will study the behavior of the polymer in the two phases. 1.4 Analysis of the two phases This section is devoted to explain the main results about the two phases of the polymer model introduced in section 1.3. Roughly speaking, for small β, i.e. high temperature, the behavior is diffusive just as in the SRW model (β = 0); for large β, i.e. low temperature, the behavior is (expected to be) superdiffusive. In order to quantify how much β should be “small” or “large” (which of course will depend on the dimension d and the potential distribution), we introduce the following functions for β ∈ [0, ∞): γ1 (β) = log M(2β) − 2 log M(β) , γ2 (β) = β[log M(β)]0 − log M(β) . Since the logarithmic moment generating function log M is strictly convex, both γ1 and γ2 are strictly increasing on [0, ∞). We also define the return probability for the SRW in Zd : πd := P (Xn = 0 for some n ≥ 1) , 6 where P is the law of the SRW; we recall that πd = 1 if d = 1, 2 (recurrent SRW) and πd < 1 if d ≥ 3 (transient SRW). We now consider the following three conditions: d≥3 and γ1 (β) < log(1/πd ) , (1.6a) d≥1 and γ2 (β) > log(2d) , (1.6b) d = 1, 2 and β > 0. (1.6c) Since γ1 and γ2 are strictly increasing, γ1 (0) = γ2 (0) = 0, πd < 1 for d ≥ 3 and 2d > 1, condition (1.6a) is satisfied for large values of β, and condition (1.6b) for small values of β. It has been proven that under condition (1.6a) we have weak disorder phase, and under condition either (1.6b) or (1.6c) we have strong disorder phase (note that (1.6b) implies (1.6c) for d = 1, 2). Theorem 1.4 (Imbrie, Spencer [13], Bolthausen [3], Song, Zhou [21], Carmona, Hu [5], Comets, Shiga, Yoshida [6]). Under condition (1.6a), the following facts hold. (i) Weak disorder phase: β W∞ > 0 P-a.s. (ii) Diffusive behavior: 1 β En [|Xn |2 ] = 1 n→∞ n lim P-a.s. q (iii) Central Limit Theorem: P-a.s., the distribution of the d-dimensional standard normal distribution. d n Xn β under Pn converges to (iv) Delocalization: β lim max Pn (Xn = x) = 0 n→∞ x∈Zd P-a.s. β The proof of (i) follows from the L2 -boundedness of the martingale Wn : this implies β the convergence of Wn also in L1 , so that h βi h βi E W∞ = lim E Wn = 1 , n→∞ β excluding (1.5b). For proving that Wn is L2 -bounded, recalling (1.4) the trick is to write β 2 h (X) (Y ) i Wn = E ⊗ E en en , where E ⊗ E is the expectation w.r.t. the law P ⊗ P of a couple of independent SRWs (Xi , Yi )∞ i=0 . A series of computations shows then that for all n ≥ 1 ∞ k ∞ X k X h β 2 i γ1 (β) E Wn ≤ 1+ e − 1 P ⊗ P (Xi = Yi ) , (1.7) i=1 k=1 and the series over i is actually the expected number of returns to 0 of a SRW in Zd , which is finite for d ≥ 3 and can be easily computed in terms of πd using the Markov property: ∞ X P ⊗ P (Xi = Yi ) = i=1 7 πd . 1 − πd It follows that the series over k in (1.7) converges if and only if πd < 1, eγ1 (β) − 1 1 − πd which is equivalent to condition (1.6a). The proof of (ii) and (iii) involves the analysis of some classes of martingales that β generalize (Wn )n≥0 . Similarly to the SRW model, for which the Local Limit Theorem (1.1) holds, under condition (1.6a) the position Xn of the polymer’s endpoint spreads out in the limit as n → ∞, leading to the delocalization stated in (iv). Let us define for n ≥ 1 the following random variables on (Ω, F , P): X β β β β Jn := max Pn (Xn = x) , In := Pn (Xn = x)2 . (1.8) x∈Zd x∈Zd Since clearly β 2 β β 0 ≤ Jn ≤ In ≤ Jn , β β Jn converges to 0 if and only if In does. Thus, the proof of (iv) is carried out by analyzing β β β β In instead of Jn . (In )n≥1 and W∞ are connected by the following fundamental fact: for all β>0 ∞ β X β P W∞ > 0 = P In < ∞ , n=1 β whose proof is based on Doob’s decomposition for the process − log Wn . Under conβ β dition (1.6a), (i) ensures that the series of In is finite P-a.s., so that In converges to 0 P-a.s.. Theorem 1.5 (Carmona, Hu [5], Comets, Shiga, Yoshida [6]). Under either condition (1.6b) or condition (1.6c), the following facts hold. (i) Strong disorder phase: β W∞ = 0 P-a.s. (ii) Localization to the favorite sites: β lim sup max Pn (Xn = x) ≥ c n→∞ x∈Zd P-a.s. for some c = c(d, β) > 0. Note that (ii) highlights a qualitative behavior of the polymer in contrast with (iv) of Theorem 1.4: in the limit as n → ∞, the polymer’s endpoint concentrates on some favorite sites instead of spreading out. Furthermore, √ the location of these favorite sites is predicted to be at a distance of order larger than n, highlighting superdiffusive behavior: only partial results have been obtained, for example in [20]. Let us sketch the proof of Theorem 1.5 under condition (1.6b). The key is to provide h β θ i an estimate on the fractional moment E Wn for some θ ∈ (0, 1). Using the Markov 8 property of the SRW and the inequality (u + v)θ ≤ u θ + v θ , it is not difficult to prove that for all θ ∈ (0, 1) h β θ i h β θ i E Wn ≤ r(θ) E Wn−1 , 1−θ r(θ) := (2d) " βV (1,1) !θ # e E . M(β) By induction, it follows that h β θ i E Wn ≤ r(θ)n . (1.9) Since θ → log r(θ) is convex on (0, 1) and log(2d) = log r(0) > log r(1) = 0, there exists θ ∈ (0, 1) such that log r(θ) < 0 (i.e. r(θ) < 1) if and only if (log r)0 (1) > 0: the latter h β θ i condition is indeed equivalent to (1.6b). Choosing such a θ, (1.9) implies E Wn →0 as n → ∞. By Fatou’s Lemma: h β θ i h β θ i h β θ i E W∞ = E lim inf Wn ≤ lim inf E Wn = 0. n→∞ n→∞ β β β Since W∞ ≥ 0 P-a.s., this proves (i). A little more work involving estimates on Jn , In (as β defined in (1.8)) and − log Wn is needed to prove (ii). Finally, from Theorems 1.4 and 1.5 we can deduce the following about the critical value βc introduced in Theorem 1.3: • βc = 0 for d = 1, 2; • βc ∈ [βc1 , βc2 ] for d ≥ 3, where βc1 is the maximum β such that γ1 (β) < log(1/πd ) and βc2 is the minimum β such that γ2 (β) > log(2d). 2 RSK correspondence and related probabilistic models This section deals with the classical RSK correspondence, so called because it dates back to Robinson’s, Schensted’s and Knuth’s mathematical works. Subsection 2.1 is devoted to provide the general combinatorial framework. In the next subsections, we will explain the fundamental role that RSK plays in understanding some probabilistic models closely related to random polymers: we will talk about longest increasing subsequences in random permutations and maximal directed paths on Poisson points (subsection 2.2), directed last passage percolation (subsection 2.3) and polynuclear growth model (subsection 2.4). 2.1 Robinson-Schensted-Knuth correspondence The Robinson-Schensted-Knuth correspondence (RSK) is a combinatorial algorithm providing a bijection between matrices of non-negative integers and semistandard Young tableaux of the same shape. In this section we explain it in detail and show its main properties. A partition of n of length l is a weakly decreasing sequence λ = (λ1 , . . . , λl ) of l positive integers that sum up to n: we write λ ` n and l(λ) = l. The (only) partition of length 0 is λ = . A graphical representation of a partition λ is a Young diagram, i.e. a collection of l(λ) left-justified rows of boxes such that i-th row contains λi boxes; row lengths turn out 9 to be weakly decreasing. The following is an example of partition of 8 of length 3, with its corresponding Young diagram: λ = (4, 3, 1) ←→ A Young tableau P is obtained by filling the boxes of a Young diagram with positive integer numbers; the shape of P , denoted by sh(P ), is the partition which corresponds to the Young diagram; the size of P is its total number of boxes; the type of P is the vector (P1 , P2 , . . . ) such that Pj is the number of j’s in P (since the size is finite, Pj = 0 for j large enough). A Young tableau is called semistandard if its rows are weakly increasing and its columns are strictly increasing. Here, we show an example of semistandard Young tableaux of shape λ = (4, 3, 1), therefore of size 8, and type (2, 2, 1, 3): 1 1 2 3 2 4 4 4 We now define the insertion of a positive integer k into a semistandard Young tableau P = (pi,j ). In row i, starting from i = 1, we search for the smaller j such that pi,j > k. If such a j does not exist, we simply add a box filled with k at the end of row i; if such a j does exist, we fill box (i, j) with k and bump the old entry pi,j to the next row i + 1, where we will try to insert it in the same way. Clearly, the procedure must stop in a finite number of steps, producing a new semistandard Young tableau whose size is increased by one. For example, inserting 2 in the tableau above, we obtain: 1 1 2 2 2 3 4 4 4 For any m × n matrix W = (wi,j ) with non-negative integer entries, we consider two words of the same length composed of weakly increasing words in the alphabet of nonnegative integer numbers: (i) the word w := w1 . . . wm such that wi := 1wi,1 . . . nwi,n ; (ii) the word w0 := w10 . . . wn0 such that wj0 := 1w1,j . . . mwm,j . The RSK correspondence is defined as the map that associates a matrix W to the pair of semistandard Young tableaux (P , Q) such that P is obtained by inserting all numbers appearing in w successively (starting from the empty tableau) and Q is obtained in the same way using w0 instead. 10 Example 2.1. 1 2 1 1 W = 0 1 1 0 3 0 0 1 0 w = 12234 23 1114 | {z } |{z} |{z} w1 w2 w = 1333 112 12 13 |{z} |{z} |{z} |{z} w10 w3 w20 ? w30 w40 ? 1 1 1 1 3 4 P= 2 2 2 3 4 1 1 1 1 1 3 Q= 2 2 3 3 3 ♦ It is no accident that P and Q have the same shape in the latter example. Indeed, Q can also be constructed at the same time as P this way: when inserting a number of wi into P , the diagram increases by one box; a box must be added to Q in the same position and filled with i. It is clear from this alternative construction that P and Q have finally (actually, at every insertion step) the same shape. Moreover, the denominations insertion tableau and recording tableau for P and Q respectively are now natural. We also observe that the j-th column of W sum up to the number of j’s in P , i.e. Pj . Similarly, the i-th row of W sum up to the number of i’s in Q, i.e. Qi . We then have the following fundamental theorem, whose proof can be found for example in [22, §7.11]. Theorem 2.2. The RSK correspondence is a bijection between matrices W = (wi,j ) with non-negative integer entries and pairs (P , Q) of semistandard Young tableaux with the same shape. Under this bijection, the type of P and Q is determined by: X X Pj = wi,j ∀j , Qi = wi,j ∀i . (2.1) i j Transposing matrix W , the roles of P and Q are interchanged: RSK RSK Theorem 2.3. If W −−−−→ (P , Q), then W T −−−−→ (Q, P ). Let us now consider the special case of permutation matrices of order n, i.e. n × n matrices such that every row and every column has one entry equal to 1 and all the others equal to 0. In this case, the corresponding words w and w0 are permutations of {1, . . . , n}, inverse of each other. The resulting tableaux under RSK are standard Young tableaux of size n: namely, they are composed of n boxes occupied by the numbers 1, . . . , n, and both rows and columns are strictly increasing. 11 Example 2.4. 0 0 W = 0 1 1 0 0 0 0 0 1 0 0 1 0 0 - w0 = 4132 w = 2431 ? ? 1 3 P= 2 4 1 2 Q= 3 4 ♦ The restriction of RSK to permutations is the so-called called Robinson-Schensted (RS) correspondence, and was actually studied first: Theorem 2.5. The RS correspondence is a bijection between permutations of n objects and pairs of standard Young tableaux of size n with the same shape. We finally introduce another combinatorial object which is in a bijective correspondence with semistandard Young tableaux. Let P be a semistandard Young tableau such that maxi,j pi,j = n and let us define zi,j as the total number of 1’s, 2’s, ..., i’s in row j of P . By the column strict rule, below row i there cannot be numbers ≤ i, so that zi,j = 0 for all j > i; for the same reason, zi,j = 0 for all i > n. We can thus arrange all (possibly) nonzero zi,j , j ≤ i ≤ n, in a triangular array: z1,1 z2,1 z3,1 Z= . .. zn,1 z2,2 z3,2 ··· z3,2 .. ··· zn,2 ··· . ··· zn,n The i-th row (zi,1 , . . . , zi,i ) of Z is the shape of the semistandard Young tableau obtained by removing all numbers > i from the original tableau P ; it is hence clear that the last row (zn,1 , . . . , zn,n ) is the shape of P , and it is called shape of Z by analogy. On the other hand, by definition the type of P can be recovered from Z through the following equations: Pj = j X i=1 zj,i − j−1 X zj−1,i ∀j = 1, . . . , n , (2.2) i=1 where the empty sum is set to 0. The latter numbers define the type of Z by analogy. Another important property of Z is the interlacing condition: zi+1,j ≥ zi,j ≥ zi+1,j+1 12 ∀1 ≤ j ≤ i < n , (2.3) which follows from the column strict rule for P as well. Triangular arrays satisfying condition (2.3) are called Gelfand-Tsetlin patterns. It is thus easy to see that there is a bijective correspondence between semistandard Young tableaux and Gelfand-Tsetlin patterns. Consequently, Theorem 2.2 can be reformulated this way: Theorem 2.6. The RSK correspondence induces a bijection between matrices W = (wi,j ) with non-negative integer entries and pairs (Z, Z 0 ) of Gelfand-Tsetlin patterns with the same shape. Under this bijection, the type of Z and Z 0 is determined by: j X i=1 zj,i − j−1 X zj−1,i = X i=1 wi,j i X ∀j , i 0 zi,j − j=1 i−1 X 0 zi−1,j = X j=1 wi,j ∀i . j Example 2.7. The Gelfand-Tsetlin patterns associated to tableaux P and Q of Example 2.1 are respectively: 4 4 Z= 5 6 3 3 5 3 0 2 5 Z = 1 6 0 2 3 2 ♦ 2.2 Longest increasing subsequence of random permutations and directed polymers on Poisson points Let Sn be the symmetric group of order n. For each permutation σ ∈ Sn , we say that a subsequence (n1 , . . . , nk ) of the sequence (σ (1), . . . , σ (n)) is an increasing subsequence of length k if n1 < · · · < nk . Let Ln (σ ) be the length of the longest increasing subsequence for the permutation σ . If we think of Sn as a probability space with the uniform distribution, we can investigate the asymptotic law of the random variable Ln for large n: this is a classical problem, proposed by Ulam [23] in 1961. A graphical formulation of the problem is the following. Let us define a directed path in R2 from (x0 , y0 ) to (x, y) as any piecewise linear path between these endpoints that is increasing in both coordinates; it is determined by a finite sequence of points (x1 , y1 ), . . . , (xl , yl ) (endpoints of the linear pieces) such that x0 < x1 < · · · < xl < x and y0 < y1 < · · · < yl < y; l is called length of the path. Let us now fix a square, say [0, 1]2 , and n points (x1 , y1 ), . . . , (xn , yn ) in the square with distinct abscissas and ordinates. Assuming that x1 < x2 < · · · < xn , let σ ∈ Sn be the permutation such that yσ (1) < yσ (2) < · · · < yσ (n) . Then increasing subsequences of σ are clearly in bijective correspondence with directed paths from (0, 0) to (1, 1) on the points (x1 , y1 ), . . . , (xn , yn ) (i.e., consisting of a subset of these points). Therefore, if we consider the n points as independent random variables uniformly distributed on the square, the length of the longest directed path on these points has the same law as Ln . The key in analysing Ulam’s problem is the Young tableau representation of permutations, i.e. the RS correspondence introduced in subsection 2.1. In this respect, the following theorem is fundamental, see [12]: 13 RS Theorem 2.8. Let σ ∈ Sn such that σ −−→ (P , Q), and let λ be the common shape of (k) tableaux P and Q. For k ≤ n, let Ln (σ ) be the length of the longest subsequence of σ consisting of k disjoint increasing subsequences. Then (k) Ln (σ ) = λ1 + · · · + λk . (1) In particular, the length Ln (σ ) = Ln (σ ) of the longest increasing subsequence of σ is the length of the first row of the Young tableaux associated to σ via RS. The asymptotic law of Ln can therefore be studied by analyzing the behavior of λ1 , with the law inherited by the uniform distribution on Sn , which is called Plancherel measure (see Theorem 2.5): Pl(λ) = P dλ2 2 µ`n dµ ∀λ ` n , where dµ is the number of semistandard Young tableaux of shape µ, for any µ ` n. Using this representation, Vershik and Kerov [24] carried out the first important step in attacking the problem, proving that E[L ] lim √ n = 2 . n→∞ n Only in 1999, Baik, Deft and Johansson [1] finally solved Ulam’s problem, proving that the fluctuations of Ln are of order n1/6 and its asymptotic law is F2 , the Tracy-Widom distribution for the Gaussian Unitary Ensemble (GUE): √ Ln − 2 n lim P ≤ s = F2 (s), s ∈ R. (2.4) n→∞ n1/6 They proved this result using a closely related problem, called Poissonized version. Instead of fixing the length n, they considered it as a Poisson random variable with mean λ, and studied the asymptotic law of Ln as λ gets large; afterwards, they used the obtained result to solve the original problem. As Ulam’s problem has a graphical interpretation in terms of directed paths, also its Poissonized version can be visualized as a point-to-point problem of directed paths on Poisson points. Let ω be a configuration of points in the square [0, t]2 , corresponding to the realization of a Poisson process of intensity 1 on the plane. Let us define L(t) as the length of the longest direct path from (0, 0) to (t, t) on these Poisson points. Since t 2 is the 2 area of the square, the probability of having k points in the square is just e−t (t 2 )k /k!: this shows the equivalence to the Poissonized version of Ulam’s problem, setting λ := t 2 . In this setting, the fluctuations are of order t 1/3 and the convergence in law stated in (2.4) becomes: L(t) − 2t lim P ≤ s = F2 (s), s ∈ R. t→∞ t 1/3 Finally, we mention a modification of the latter model, called point-to-line problem of directed paths on Poisson points. On point configurations arising from a Poisson process of intensity 1 on the plane, we now consider the length L∗ (t) of the longest directed path starting from (0, 0) and ending at any point (x, y) of the first quadrant on the line 14 (m, n) (m + n − 2, −m + n) (0, 0) (1, 1) Figure 3. Two graphical representations of directed paths on the (1 + 1)-dim. lattice. {x + y = 2t}. Fluctuations are still of order t 1/3 , but the asymptotic law is now F1 , the Tracy-Widom distribution for the Gaussian Orthogonal Ensemble (GOE), see [2]: lim P t→∞ 2.3 L∗ (t) − 2t ≤ s = F1 (s), t 1/3 s ∈ R. Directed last passage percolation Let us consider the following discrete model on the lattice N2 . Let us define Πm,n as the set of up-right paths from (1, 1) to (m, n): namely, every path π ∈ Πm,n is a sequence ((i1 , j1 ), (i2 , j2 ), . . . , (im+n , jm+n )) such that (i1 , j1 ) = (1, 1), (im+n , jm+n ) = (m, n) and (ik+1 , jk+1 ) − (ik , jk ) is either (1, 0) or (0, 1). Assigning to every point (i, j) ∈ N2 a non-negative weight wi,j , we define: X Tm,n := max π∈Πm,n wi,j ∀m, n ≥ 1 . (2.5) (i,j)∈π If the weights wi,j are interpreted as waiting times, Tm,n is the maximal passage time for directed paths from (1, 1) to (m, n), and turns out to be the point-to-point version of the directed last passage percolation defined in (1.3). To see this, let us identify in the obvious way a path π ∈ Πm,n with a path x = (k, xk )m+n−2 ∈ (N × Z)m+n−1 such that x0 = 0, xk−1 ∼ xk k=0 for 1 ≤ i ≤ m + n − 2 and xm+n−2 = −m + n: namely, up steps (resp., right steps) in the first path corresponds to up-right steps (resp., down-right steps) in the latter, see also Figure 3. Also, set V (k, xk ) := wi,j if k = i + j − 2 and xk = −i + j. Then, Tm,n = max x m+n−2 X V (k, xk ) , k=0 where the maximum is over all paths x defined above. The latter equation is clearly a point-to-point version of (1.3). We note that Tm,n can also be defined by the recursive formula: Tm,n = max(Tm−1,n , Tm,n−1 ) + wm,n ∀m, n ≥ 1 , (2.6) setting Tm,n := 0 whenever m or n are zero. We now explain why directed last passage percolation is strongly related to RSK. For this, we need a generalization of Theorem 2.8 from the RS setting to the RSK setting (see for example [11]): 15 Theorem 2.9. Let W = (wi,j ) be an m × n matrix of non-negative integers such that RSK W −−−−→ (P , Q), and let λ be the common shape of tableaux P and Q. For k ≤ n, let us define X (k) Tm,n := max wi,j , π(1) ,...,π(k) (i,j)∈π(1) ∪···∪π(k) where the maximum is over all k-tuples of non-intersecting directed lattice paths π(i) from (1, i) to (m, n − k + i), for i = 1, . . . , k. Then (k) Tm,n = λ1 + · · · + λk . (1) In particular, taking k = 1, Tm,n = Tm,n turns out to be equal to the length λ1 of the first row of the tableaux P and Q obtained from the weight matrix (wi,j ) via RSK. Let us now consider independent weights Wi,j geometrically distributed with parameter pj qi , where (pj )j≥1 and (qi )i≥1 are two sequences of numbers in (0, 1): P(Wi,j = wi,j ) = (1 − pj qi )(pj qi )wi,j ∀wi,j ≥ 0 . We are interested in calculating the law of Tm,n . Let us fix an m × n matrix (wi,j ) and the corresponding tableaux (P , Q) via RSK. Denoting by (P1 , . . . , Pn ) and (Q1 , . . . , Qm ) the type of P and Q respectively, by independence of the weights and Theorem 2.2: h i RSK P (Wi,j )i≤m,j≤n −−−−→ (P , Q) = P(Wi,j = wi,j ∀i ≤ m, j ≤ n) = n Y m Y (1 − pj qi )(pj qi )wi,j j=1 i=1 = "Y n Y m # "Y # "Y # n m Pj Qi (1 − pj qi ) · pj · qi . j=1 i=1 | j=1 {z i=1 } =:cm,n If sh is the (random) shape of the tableaux obtained from (Wi,j )i≤m,j≤n via RSK, then for a fixed partition λ of length ≤ min(m, n) P(sh = λ) = X h i RSK P (Wi,j )i≤m,j≤n −−−−→ (P , Q) (P ,Q): sh(P )=sh(Q)=λ n m X Y X Y Pj Q = cm,n · pj · qi i sh(P )=λ j=1 sh(Q)=λ i=1 = cm,n · sλ (p1 , . . . , pn ) · sλ (q1 , . . . , qm ) . Here sλ is a Schur polynomial, which is a well-known symmetric polynomial in n variables p1 , . . . , pn , indexed by a partition λ of length ≤ n and defined by n X Y sλ (p1 , . . . , pn ) = sh(P )=λ j=1 16 Pj pj , where the sum is over all semistandard Young tableaux of shape λ in the alphabet {1, . . . , n}. Since Tm,n = λ1 by Theorem 2.9, we can now express the distribution function of Tm,n in terms of Schur polynomials: Theorem 2.10. If the weights Wi,j are independent and geometrically distributed with parameter pj qi , then for all m, n ≥ 1 and t ∈ R X P(Tn,m ≤ t) = cm,n sλ (p1 , . . . , pn ) · sλ (q1 , . . . , qm ) , λ: λ1 ≤t where the sum is over all partitions λ of length ≤ min(m, n), and n Y m Y cn,m := (1 − pj qi ) . j=1 i=1 This model with geometric weights is therefore exactly solvable: the law of Tn,m can be explicitly computed in terms of Schur polynomials, which also have an explicit determinantal expression (see [22, § 7.15]). In the special case of i.i.d. geometric weights, Johansson [14] proved that Tn,n , appropriately scaled, converges in law to the Tracy-Widom distribution for the GUE. The case of independent exponentially distributed weights can be also proven to be exactly solvable; furthermore, it is of particular interest because, if all the weights Wi,j have exponential distribution with mean 1, Tm,n also describes passage times of particles in the well-known totally asymmetric exclusion process (TASEP). This is a continuous time Markov process (ηt )t≥0 on the state space {0, 1}Z , interpreted as an interacting particle system: any η ∈ {0, 1}Z is a configuration of particles and holes (1’s and 0’s respectively) on the one dimensional integer lattice. A particle at site k jumps with exponential rate one to k + 1, provided that this site is vacant, otherwise nothing happens. Considering the initial configuration η such that η(k) = 1 if and only if k ≤ 0 (i.e. all non-negative sites are occupied and all positive sites are empty), let us call Wi,j the time that the particle starting at site −i has to wait to perform its j-th jump once that site −j + i is empty. Then Tm,n is the time that the particle at site −m needs to wait to go beyond site −m + n; in particular, Tn,n is the time that the particle at site −n needs to wait to go beyond site 0. Indeed, it is easily seen that Tm,n defined in such a way satisfies the recursive formula (2.6). 2.4 Polynuclear growth model In statistical mechanics, surface growth models simulate the behavior of atoms or molecules that, once ejected onto a surface, attach to each other and form growing islands. Every growth model is given by a random function h(x, t), which represents the height of the surface and depends on a d-dimensional spatial coordinate x and a time coordinate t. KPZ universality class is the most studied class of random growth models characterized by both local evolution and local randomness: for such models, the height function is solution of a (d + 1)-dimensional nonlinear stochastic equation introduced by Kardar, Parisi and Zhang [15]. The polynuclear growth (PNG) model is a surface growth model in one spatial dimension belonging to the KPZ universality class, and can be thought of as a further graphical representation of RSK correspondence, see [10] and [9]. We now wish to give 17 a discrete time version of the PNG model. Starting from an m × n matrix (wi,j ) of independent non-negative random weights (to simplify notation we assume m = n), we will construct a growing surface of height h(x, t); in fact, for a complete characterization of matrix (wi,j ), we will describe an ensemble of height functions {hl }l=0,−1,−2,... such that h0 ≡ h and hl (x, t) ≥ hl−1 (x, t) for all level l and for all (x, t). For all t and l, hl (·, t) will be constant on any spatial interval [x − 1/2, x + 1/2), x ∈ Z; therefore, we can see each hl (·, t) as a profile of columns of unit width centered on the integers. This way, it suffices to define the evolution on all integers x. We start from the flat initial condition: at time 0, hl (x, 0) = 0 for all x ∈ Z, l ≤ 0. For t ≥ 1, this is the outline of the evolution: • in the time interval from t − 1 to t, at any level l every column grows one unit to the left and one unit to the right, if it is higher than its neighbors; • if two columns meet, they merge and the highest column “wins”, creating an overlap; • each overlap created at level l falls down to level l − 1, growing the corresponding column; • at level 0, weights wi,j “fall down” to grow columns further on: precisely, column at site x also grows by wi,j if i − j = x and i + j − 1 = t; • the evolution stops after instant 2n − 1. We now describe this evolution formally, constructing from matrix (wi,j ) a new n × n matrix (ti,j ) = T ((wi,j )) which will contain all the information about the height functions. This map T : (R≥0 )n×n → (R≥0 )n×n will be expressed as a composition of simple maps li,j called local moves, as explained in [19, §8]. For all i, j = 1, . . . , n, let us define li,j : (R≥0 )n×n → (R≥0 )n×n as the map that takes as input an n × n matrix (wi,j ) and replaces the submatrix " # wi−1,j−1 wi−1,j wi,j−1 wi,j with its image under the map " # " # min(b, c) − a b a b → . c d c max(b, c) + d (2.7) To give a meaning to li,j also for i = 1 or j = 1, we simply set wi,j = 0 whenever i = 0 or j = 0; in other words: • l1,1 is the identity map; • for j = 2, . . . , n, l1,j replaces w1,j with w1,j−1 + w1,j ; • for i = 2, . . . , n, li,1 replaces wi,1 with wi−1,1 + wi,1 . We note that li,j is bijective. For i, j = 1, . . . , n, set: l1,j−i+1 ◦ · · · ◦ li−1,j−1 ◦ li,j %ji := li−j+1,1 ◦ · · · ◦ li−1,j−1 ◦ li,j i≤j i ≥j. (2.8) For convenience, we set %ji to be the identity map if (i, j) < {1, . . . , n}2 . For t = 1, . . . , 2n − 1, we also set 2 Dt := %1t ◦ %2t−1 ◦ · · · ◦ %t−1 ◦ %t1 . (2.9) 18 The desired map T : (R≥0 )n×n → (R≥0 )n×n is then defined by T := D2n−1 ◦ · · · ◦ D1 . (2.10) The height profiles at time 2n − 1 are now determined by ti,j = hmax(i,j) (i − j, 2n − 1) ∀(i, j) ∈ {1, . . . , n}2 , and hl (x, 2n − 1) = 0 for any other x ∈ Z and l ≤ 0. Since the local moves are bijective, T is also bijective, and it turns out to be a version of RSK extended to all matrices of non-negative real numbers (not necessarily integers). More precisely, let us change notation of (ti,j ) this way: t1,1 t1,2 t1,n t2,1 tn−1,n tn,n−1 tn,n tn,1 0 0 λn zn−1,n−1 z1,1 zn−1,n−1 = 0 zn−1,1 z1,1 zn−1,1 λ1 . The main diagonal of T in reverse order 0 0 ) (tn,n , . . . , t1,1 ) = (λ1 , . . . , λn ) = (zn,1 , . . . , zn,n ) = (zn,1 , . . . , zn,n turns out to be the common shape of two Gelfand-Tsetlin patterns Z and Z 0 , corresponding to the entries of T below and above the diagonal respectively. If wi,j are integers, W → (Z, Z 0 ) coincides with the version of RSK stated in Theorem 2.6. This new representation of RSK correspondence in terms of local moves will be especially useful in subsection 3.1, where we will talk about geometric RSK. Geometric RSK and polymers in 1 + 1 dimensions 3 This section deals with the geometric lifting of the RSK correspondence introduced by Kirillov in [16], and applications to some exactly solvable random polymer models. In subsection 3.1, we describe the geometric RSK, following [19] (the construction is equivalent to the one given in [18]). In subsection 3.2, we show the exact solvability of the random polymer model with inverse gamma weights [19]. In subsection 3.3, we sketch more polymer models with inverse gamma weights characterized by symmetries or other constraints, and we also set out some open problems. 3.1 Geometric RSK Consider the classical RSK correspondence for square matrices as a map (R≥0 )n×n → (R≥0 )n×n , as explained in subsection 2.4: the output matrix (ti,j ) is obtained from the input matrix (wi,j ) by applying a sequence of local moves li,j defined by (2.7). By construction, such local moves only involve sums, subtractions, max and min operations: all of them can be defined in the (max, +) semiring, i.e. R ∪ {−∞} with the algebraic structure defined by the binary operations max (“addition”) and + (“multiplication”)† . † Note that min(b, c) = − max(−b, −c). 19 In brief, the geometric RSK is the mapping T : (R>0 )n×n → (R>0 )n×n consisting in the composition of the analogous sequence of maps li,j , where the operations in the (max, +) semiring (combinatorial setting) are formally replaced with the corresponding ones in the usual (+, ·) ring (geometric setting): a+b → a·b, a−b → a , b max(a, b) → a + b . In other words, the new local move li,j , for all i, j = 1, . . . , n, is defined as the map (R>0 )n×n → (R>0 )n×n that takes as input an n × n matrix (wi,j ) and replaces the submatrix " wi−1,j−1 wi−1,j wi,j−1 wi,j # with its image under the map " # " # a b bc/[a(b + c)] b → . c d c (b + c)d To give a meaning to li,j also for i = 1 or j = 1, the convention is that wi,j = 0 whenever i = 0 or j = 0, but w0,1 + w1,0 = 1. In other words: • l1,1 is the identity map; • for j = 2, . . . , n, l1,j replaces w1,j with w1,j−1 · w1,j ; • for i = 2, . . . , n, li,1 replaces wi,1 with wi−1,1 · wi,1 . With this new definition of local moves, we set T to be the same sequence of local moves defined in (2.8), (2.9) and (2.10). Since many local moves commute with each other (li,j ◦ li 0 ,j 0 = li 0 ,j 0 ◦ li,j whenever |i − i 0 | + |j − j 0 | > 2), there are other equivalent ways to define the order of local moves in the sequence, see [19, § 3]. Since the local moves li,j are clearly birational, i.e. bijective rational mappings, T is also birational. Similarly to the RSK case, we will also adopt the notation: t1,1 t1,2 t1,n t2,1 tn−1,n tn,n−1 tn,n tn,1 0 0 λn zn−1,n−1 z1,1 zn−1,n−1 = 0 zn−1,1 z1,1 zn−1,1 λ1 . This way, we can see the output matrix T as a pair (Z, Z 0 ) of triangular patterns of height n 0 z1,1 z1,1 .. Z= zn,1 . .. ··· Z0 = , . .. 0 zn,1 zn,n . .. ··· where 0 0 (zn,1 , . . . , zn,n ) = (zn,1 , . . . , zn,n ) = (λ1 , . . . , λn ) = (tn,n , . . . , t1,1 ) 20 , . 0 zn,n will be referred to as the (common) shape of Z and Z 0 , as in the case of Gelfand-Tsetlin gRSK gRSK patterns. We will write either W −−−−−→ T or W −−−−−→ (Z, Z 0 ). We define the type (P1 , . . . , Pn ) of a pattern Z of height n by Qj i=1 zj,i Pj := Qj−1 i=1 zj−1,i ∀j = 1, . . . , n , (3.1) which is the analog of (2.2) in the geometric setting (here the empty product is equal to 1 by convention). We now state some properties of geometric RSK, most of which have a straightforward equivalent in the combinatorial setting, see [18] and [19]. gRSK Theorem 3.1. Let W ∈ (R>0 )n×n and W −−−−−→ T = (Z, Z 0 ). Let (P1 , . . . , Pn ) and (Q1 , . . . , Qn ) denote the types of Z and Z 0 respectively, and (λ1 , . . . , λn ) their common shape. Then: Pj = n Y wi,j and ∀j Qi = n Y i=1 wi,j ∀i ; (3.2) j=1 gRSK W T −−−−−→ T T = (Z 0 , Z) ; X Y λ1 · · · λk = (3.3) ∀k ≤ n ; wi,j (3.4) π(1) ,...,π(k) (i,j)∈π(1) ∪···∪π(k) X 1 1 = + wi,j t1,1 i,j X ti−1,j + ti,j−1 (i,j),(1,1) ti,j . (3.5) In (3.4), the sum is over all k-tuples of non-intersecting directed lattice paths π(i) from (1, i) to (n, n − k + i), for i = 1, . . . , k. In (3.5), the convention is that ti−1,j = 0 if i = 1 and ti,j−1 = 0 if j = 1. Formulas (3.2) are the equivalent of (2.1) in the geometric setting. Formulas (3.3) and (3.4) are the analogs of Theorem 2.3 and Theorem 2.9 respectively in the geometric setting. Formula (3.5) is the cornerstone of [19] and can be proved by induction on n. 3.2 Inverse gamma polymers From (3.4) for k = 1, we see that tn,n = λ1 = X Y wi,j ∀n ≥ 1 , (3.6) π∈Πn,n (i,j)∈π where Πn,n is the set of directed paths from (1, 1) to (n, n) in N2 . Similarly to what we noticed in subsection 2.3 for last passage percolation, (3.6) is a point-to-point version of the (1 + 1)-dimensional directed polymer partition sum defined in (1.2). Again, to see 2n−1 such that x = 0, this we identify a path π ∈ Πn,n with a path x = (k, xk )2n−2 0 k=0 ∈ (N × Z) xk−1 ∼ xk for 1 ≤ i ≤ 2n − 2 and x2n−2 = 0. If we set wi,j = exp[βV (k, xk )] 21 for k = i + j − 2 and xk = −i + j, then tn,n = X " 2n−2 # X exp β V (k, xk ) , x k=0 where the maximum is over all paths x defined above. The latter equation is clearly a point-to-point version of (1.2). We now wish to study the distribution of the polymer partition sum tn,n in the special case of inverse gamma distributed weights. We recall that X −1 ∼ Γ (α, β) if on R>0 ! β α −α−1 β dx . P(X ∈ dx) = x exp − Γ (α) x −1 If (wi,j )1≤i,j≤n are independent and wi,j ∼ Γ (αj + α̂i , 1) for all i, j, then the joint law of matrix W is given by Y Y −α −α̂ X 1 ! Y dwi,j 1 j i να,α̂ (dw) = wi,j . exp − Γ (αj + α̂i ) wi,j wi,j i,j i,j i,j i,j We now need the following lemma: Lemma 3.2. The gRSK mapping in logarithmic variables log(wi,j )1≤i,j≤n → log(ti,j )1≤i,j≤n has Jacobian ±1. The proof is based on the local move decomposition of the gRSK mapping: it is easy to see that all maps li,j have Jacobian ±1. This lemma permits to make the change of variables from (wi,j ) to (ti,j ) in the measure να,α̂ : Theorem 3.3. The push-forward of να,α̂ under the gRSK map T is given by να,α̂ ◦ T −1 (dt) = Y i,j −α 1 1 −α −α̂ P Q exp − − Γ (αj + α̂i ) t1,1 −α X (i,j),(1,1) ti−1,j + ti,j−1 Y dti,j . t t i,j i,j i,j Here, P −α = P1 1 · · · Pn n , where (P1 , . . . , Pn ) is the type of Z as usual, and the same for Q−α̂ . This theorem follows from (3.2), (3.5) and Lemma 3.2. We are now able to give an expression for the Laplace transform of the polymer 0 partition sum, using the push forward of να,α̂ and writing (ti,j ) in terms of (zi,j ), (zi,j ) and 22 λ: for θ ∈ R, i Y h Γ (αj + α̂i )E e−θtn,n i,j Z = e −θtn,n −α P Q (R>0 )n×n Z 1 − exp − t 1,1 e−θλ1 = −α̂ (R>0 )n×n n Z zj,1 . . . zj,j j=1 zj−1,1 . . . zj−1,j−1 = (R>0 )n (i,j) i,j e−θλ1 −1/λn ψα (λ)ψα̂ (λ) i,j i,j 0 0 zj,1 . . . zj,j i,j !−α̂j ... 0 0 zj−1,1 . . . zj−1,j−1 j=1 0 0 0 n zi−1,j + zi+1,j+1 Y dzi,j Y dzi,j Y dλi + 0 0 zi,j λi zi,j zi,j 1≤j≤i<n n Y dλ i=1 (i,j),(1,1) ti−1,j + ti,j−1 Y dti,j t t !−αj Y n n Y 1 X zi−1,j + zi+1,j+1 − . . . exp − λ z X λi i 1≤j≤i<n i=1 . The convention is that zi,j = 0 whenever (i, j) does not satisfy 1 ≤ j ≤ i ≤ n. In the latter formula, ψα and ψα̂ (where α = (α1 , . . . , αn ) and α̂ = (α̂1 , . . . , α̂n )) are GL(n, R)-Whittaker functions, which play an analogous role of Schur polynomials in the geometric setting: !−αj Z Y n zj,1 . . . zj,j X zi−1,j + zi+1,j+1 Y dzi,j ψα (λ) = exp − , zj−1,1 . . . zj−1,j−1 zi,j zi,j j=1 (i,j) 1≤j≤i<n where the integral is over all triangular patterns (zi,j ) of height n and fixed shape λ = (zn,1 , . . . , zn,n ). 3.3 Variations and future developments In [19], O’Connell, Seppäläinen and Zygouras also analyze the case of symmetric input matrix with inverse gamma distributed weights: in this case, by (3.3), the output matrix is also symmetric. They prove that an analog of Lemma 3.2 holds, i.e. the geometric RSK map in logarithmic variables has Jacobian ±1: again, this leads to a successful change of variable in the polymer measure, and to an expression of the Laplace transform for the polymer partition sum in terms of Whittaker functions. In the same spirit, one might try to figure out what happens with antisymmetric matrices, for example finding a sequence of local moves that somehow connects the antisymmetric case to the symmetric one, and then computing the Jacobian of such a transformation. More variations on the same theme are possible, imposing other constraints on the geometry of the input matrices, and thus of the polymers. In [19], the case of triangular input arrays is also analyzed: the corresponding polymers are forced to stay “below a hard wall”. A further interesting open problem consists in analyzing the behavior of the geometric RSK and the corresponding polymers for paths not allowed to go outside a diagonal strip. References [1] J. Baik, P.A. Deift, K. Johansson, On the distribution of the length of the longest increasing subsequence of random permutations, J. Amer. Math. Soc. 12 (1999), 1119-1178. 23 [2] J. Baik, E.M. Rains, Symmetrized random permutations, Random Matrix Models and Their Applications, vol. 40, Cambridge University Press (2001), 1-19. [3] E. Bolthausen, A note on diffusion of directed polymers in a random environment. Commun. Math. Phys. 123 (1989), 529-534. [4] F. Caravenna, F. den Hollander, N. Pétrélis, Lectures on Random Polymers. Clay Mathematics Proceedings 15 (2012). [5] P. Carmona, Y. Hu, On the partition function of a directed polymer in a Gaussian random environment. Probab. Theory Relat. Fields 124 (2002), 431-457. [6] F. Comets, T. Shiga, N. Yoshida, Directed polymers in random environment: Path localization and strong disorder. Bernoulli 9 (2003), 705-723. [7] F. Comets, T. Shiga, N. Yoshida, Probabilistic analysis of directed polymers in a random environment: a review. Stochastic analysis on large scale interacting systems, Adv. Stud. Pure Math. 39, Math. Soc. Japan, Tokyo (2004), 115-142. [8] F. Comets, N. Yoshida, Directed polymers in random environment are diffusive at weak disorder. Ann. Probab. 34 (2006), 1746-1770. [9] P. Ferrari, Polynuclear growth on a flat substrate and edge scaling of GOE eigenvalues. Comm. Math. Phys. 252 (2004), 77-109. [10] P. Ferrari, Shape fluctuations of crystal facets and surface growth in one dimension. PhD Thesis, Technische Universität München (2004), http: //wt.iam.uni-bonn.de/fileadmin/WT/Inhalt/people/Patrik_Ferrari/ download/PhD_ThesisSmall.pdf. [11] W. Fulton, Young Tableaux. London Mathematical Society Student Texts, Vol. 35, Cambridge University Press (1997). [12] C. Greene, An extension of Schensted’s theorem. Adv. Math. 14 (1974), 254-265. [13] J. Z. Imbrie, T. Spencer, Diffusion of directed polymers in a random environment. J. Stat. Phys. 52 (1988), 609-626. [14] K. Johansson, Shape fluctuations and random matrices. Comm. Math. Phys. 209 (2000), 437-476. [15] K. Kardar, G. Parisi, Y.Z. Zhang, Dynamic scaling of growing interfaces. Phys. Rev. Lett. 56 (1986), 889-892. [16] A.N. Kirillov, Introduction to tropical combinatorics. Physics and Combinatorics 2000, Proceedings of the Nagoya 2000 International workshop, Eds. A.N. Kirillov and N. Liskova, World Scientific (2001), 82-150. [17] P. Mörters, Y. Peres, Brownian Motion. Cambridge Series in Statistical and Probabilistic Mathematics (2010). [18] M. Noumi, Y. Yamada, Tropical Robinson-Schensted-Knuth correspondence and birational Weyl group actions. Adv. Stud. Pure Math. 40 (2004), 371-442. 24 [19] N. O’Connell, T. Seppäläinen, N. Zygouras, Geometric RSK correspondence, Whittaker functions and symmetrized random polymers. Inventiones mathematicae , Vol. 197, Number 2 (2014), 361-416. [20] M. Petermann, Superdiffusivity of directed polymers in random environment. Ph.D. thesis, University of Zürich (2000). [21] R. Song, X. Y. Zhou, A remark on diffusion of directed polymers in random environments. J. Stat. Phys. 85, Nos. 1/2 (1996), 277-289. [22] R.P. Stanley, Enumerative Combinatorics Volume 2, Cambridge Studies in Advanced Mathematics 62 (2001). [23] S.M. Ulam, Monte Carlo simulations in problems of mathematical physics. Modern Mathematics for the Engineer (E.F. Beckenbach ed.), vol. II, McGraw-Hill (1961), 261-277. [24] A.M. Vershik, S.V. Kerov, Asymptotics of Plancherel measure of symmetric group and the limiting form of Young tables. Sov. Math. Dolk. 18 (1977), 527-531. 25