Random sorting networks: edge limit
Abstract
A sorting network is a shortest path from to in the Cayley graph of the symmetric group spanned by adjacent transpositions. The paper computes the edge local limit of the uniformly random sorting networks as . We find the asymptotic distribution of the first occurrence of a given swap and identify it with the law of the smallest positive eigenvalue of a aGUE (an aGUE matrix has purely imaginary Gaussian entries that are independently distributed subject to skew-symmetry). Next, we give two different formal definitions of a spacing — the time distance between the occurrence of a given swap in a uniformly random sorting network. Two definitions lead to two different expressions for the asymptotic laws expressed in terms of derivatives of Fredholm determinants.
Abstract
Un réseau de tri est un chemin le plus court de à dans le graphe de Cayley du groupe symétrique , engendré par des transpositions des éléments adjacents. Dans cet article nous calculons la limite locale au bord des réseaux de tri choisi uniformément quand . Nous trouvons la distribution asymptotique de la première occurrence d’une transposition donnée et l’identifions avec la loi de la plus petite valeur propre positive d’un aGUE (une matrice aGUE a des entrées gaussiennes purement imaginaires qui sont distribuées indépendamment sous condition d’antisymétrie). Ensuite, nous considerons des espacements entre deux occurrences consecutives d’un échange donné (k, k + 1) pour un réseau de tri aléatoire choisi uniformément. Nous prenons deux formalisations pour un choix aléatoire d’un tel espacement. En passant à limite, ces deux définitions conduisent à deux expressions différentes pour des lois asymptotiques exprimées en termes de dérivées des déterminants de Fredholm.
keywords:
[class=MSC]keywords:
1 Introduction
1.1 Motivation
The main object in this article is the uniformly random sorting network which we abbreviate as RSN. Let us start by giving basic definitions. Consider the permutation group of elements . We use the one-row notation for the elements of representing them as sequences . We let be the adjacent swap , , so that
A sorting network of size is a shortest sequence of swaps, whose product equals the reverse permutation . By counting inversions in permutations, one shows that the length of such shortest sequence is . Let denote the set of all sorting networks of size . Thus, elements of are sequences such that
Sorting networks can be drawn as wiring diagrams, see Figure 1. Whenever , we say that swap occurs at time ; in the wiring diagram this corresponds to two wires intersecting at the point .

The study of sorting networks is a fruitful topic, with the first mathematical results going back to [S] where Stanley computed the number of elements in :
A recent point of interest is the study of uniformly random sorting networks initiated by Angel, Holroyd, Romik, and Virag in [AHRV]. The probabilistic results can be split into two groups: global and local. On the global side, [AHRV] proved that the density of swaps approximates the semi-circle law111There is no known direct connection to the Wigner semi-circle law of the random matrix theory and the match seems coincidental. as ; [AHRV] predicted and [DVi, D] proved (among other results) that individual trajectories in the wiring diagram become (random) sine curves; see also [K, RVV] for related results.
Our paper belongs to the second group of results, [AGH, R, ADHV, GR], which studies local limits. The most basic local characteristic of the sorting network is spacing in row , which is a distance between two occurrences of the same swap in the sorting network, cf. Figure 1. [R] and [GR] discovered a link between the asymptotic laws of spacings as and random matrix theory. Namely, [R] matched the asymptotic law of the spacing in row with the distance between two eigenvalues of matrix of Gaussian Orthogonal Ensemble (GOE) of real symmetric random matrices. [GR] dealt with spacings in row with and showed that after proper rescaling they converge to the Gaudin–Mehta law, which is the universal asymptotic law for spacings between eigenvalues of real symmetric matrices of very large sizes.
Comparing the results of [R] and [GR], we see that the former links the spacings in the extreme row (i.e. at the end-point of the edge) to real symmetric matrices, while the latter links the spacings in the bulk rows to the real symmetric matrices of infinite sizes. The observed gap in the asymptotics between and matrices motivated the central question of our paper: we would like to understand, whether distributions related to random matrices can be also found in the asymptotic laws for random sorting networks.
The answer presented in this text is both Yes and No. On one side, we are so far unable to find any connections of sorting networks to eigenvalues of real symmetric matrices (which would be the most direct interpolation between [R] and [GR]). On the other hand, by slightly adjusting our point view, we find that the law of the asymptotic spacing in row can be expressed in terms of the eigenvalues of random anti-symmetric Gaussian matrices — they are known in the literature as aGUE, since their analysis reveals connections to the tools used for the Gaussian Unitary Ensemble222However, up to multiplication by , matrix elements of aGUE are real, rather than complex., see [M, Chapter 13] and [FN].
1.2 Main Results
Here is the precise definition of the random matrix object appearing in our asymptotic results.
Definition 1.1.
Let be an matrix whose entries are i.i.d. standard Gaussian random variables . Then
is called anti-symmetric GUE or aGUE for being short.
Note that the eigenvalues of are real and come in pairs: is an eigenvalue if and only if so is . When is odd, is necessary degenerate, i.e. it has a zero eigenvalue; for even , zero is not an eigenvalue almost surely.
We also would like to deal with eigenvalues of all together.
Definition 1.2.
Let be an infinite aGUE, i.e, an matrix whose corner is a aGUE. The aGUE–corners process is a point process on : we put a particle at whenever is one of the positive eigenvalues of the top-left corner of .
Definition 1.4.
is the smallest positive eigenvalue of a aGUE.
The determinantal structure for the distribution of eigenvalues of (see [M, Chapter 13] and [FN]) leads to an expression for the distribution of as a series, which can be identified with a Fredholm determinant:
where is a correlation kernel, expressed through the Hermite polynomials as
(1.1) |
Here we use the “physicist’s normalization”, so that , , is a polynomial of degree with leading coefficient and such that
(1.2) |
We can now state our first theorem.
Theorem 1.5 (First swapping law).
Fix , and let denote the first occurrence time of the swap in a uniformly random sorting network , :
Then we have the following convergence in law:
(1.3) |
In order to connect the first swaps to the spacings, we are going to use translation invariance of the random sorting networks (cf. [AHRV, Theorem 1, (i)]), which is based on the following combinatorial fact:
Proposition 1.6.
Let . Then if and only if .
Applying Proposition 1.6 one readily proves that for all ,
which can be used to identify the distribution of a spacing in row . Before doing so, we need to specify the definition of a spacing. In fact, there are two natural definitions, which lead to distinct random variables. For the definition it is convenient to extend a sorting network to a –periodic –indexed sequence:
Definition 1.7.
Given , we extend it to a sequence by requiring for all . We call the periodic extension of .
For instance is extended to the infinite –indexed sequence with s at odd positions and s at even positions. The reason for this particular definition is contained in the following straightforward corollary of Proposition 1.6:
Corollary 1.8.
If is a uniformly random element of , then its periodic extension is translation-invariant: for each we have a distributional identity
Definition 1.9 (First definition of the spacing).
Fix , and . Let be the periodic extension of a uniformly random sorting network in . Set
We define the spacing on the row of a sorting network of size as .
Remark 1.10.
While the definition depends on the choice of , Corollary 1.8 implies that the distribution of is the same for all . Hence, we omit from the notation.
The limit of turns out to be connected to of Definition 1.4.
Theorem 1.11.
Fix . We have the following convergence in distribution:
where is a positive random variable of density , , given by
(1.4) |
Here is an alternative definition of the spacing.
Definition 1.12 (Second definition of the spacing).
Fix , and . Let be the periodic extension of a uniformly random sorting network in . Set
We define the spacing on the row of a sorting network of size as a random variable whose law is the distribution of conditional on the event .
Remark 1.13.
As in Definition 1.9, the choice of does not affect the distribution of .
Both definitions of the spacing have their own merits. Definition 1.12 is the one preferred in theoretical physics and random matrix literature, and it has been used implicitly in [R], which studies the scaling limit of the spacing of RSN on the row. However, a disadvantage of Definition 1.12 is that it is hard to sample or observe , as it fails to be a random variable on the state space of all sorting networks; on the other hand, is a function of a sorting network, and this is the definition used in [GR] which studies the limiting behavior of RSN in the bulk. The limit of is still connected to of Definition 1.4, but in a slightly different way.
Theorem 1.14.
Fix . We have the following convergence in distribution:
where is a positive random variable of density , , given by
(1.5) |
Remark 1.15.
In case becomes the absolute value of a Gaussian random variable, . Hence, has density , , which matches the result of [R].
In addition, we present an alternative expression for by identifying it with a marginal of a certain two-dimensional determinantal point process; we refer to [B] and references therein for general discussion of the determinantal point processes.
Definition 1.16.
Let be a determinantal point process on , such that with respect to the product of the counting measure on the first coordinate and the Lebesgue measure on the second coordinate, the correlation kernel is
Here and ; when or is equal to , the kernel is defined as the limit as or tends to 0. are counterclockwise–oriented contours which enclose the integers in and , respectively, and are arranged so that and are disjoint, as in Figure 5.
Remark 1.17.
For , there are particles in the and levels of (i.e. with first coordinates and , respectively), except that on the level there are only particles.
If we denote the particle(from top to bottom) on the level by and set for particles in the level, then the particles of interlace, i.e,
While we do not prove this, we expect that compared to , can be thought as the corner process of aGUE conditioned on the event that the smallest nonnegative eigenvalue of the corner is 0, see Remark 3.2.
Theorem 1.18.
Let be as in Theorem 1.14 and let denote the coordinate of the closest to the origin particle on the level in the point process .
-
•
If , then the law of coincides with that of .
-
•
If , then the law of coincides with that of .
1.3 Methods
Here is a sketch of our proof of Theorem 1.5:
- •
-
•
We replace standard Young tableau by a continuous version, which we call Poissonized Young tableau. In Section 3.2 we show that the error of this replacement is negligible in our limit regime.
-
•
We use the formula of [GR] for the correlation kernel of the determinantal point process, describing the entries of a Poissonized Young tableau.
- •
-
•
Expanding the integrals in residues, performing resummations, and using the Gibbs (conditional uniformity) properties of the point processes under consideration, we reveal in Section 3.4 a match to the distribution of eigenvalues of aGUE matrices.
We remark that (in contrast to the results in the bulk, i.e., for of order , , as in [GR, ADHV]) we do not claim any results for the joint asymptotic distribution of several swaps; Theorem 1.5 only deals with one-dimensional marginal distribution. A technical reason is that the Edelman-Green bijection does not have any straightforward continuous version acting on the eigenvalues of aGUE of finite sizes: it seems that one needs to deal simultaneously with aGUE of all sizes, which is not a particularly transparent operation. In future, it would be interesting to find a way to overcome this difficulty.
Theorems 1.11 and 1.14 are proven in Section 4 as corollaries of Theorem 1.5. The central idea is to develop discrete versions of (1.4), (1.5) valid for random sorting networks of finite sizes. In fact, these discrete statements are valid for any periodic point processes and, thus, can be of independent interest, see Proposition 4.2.
Finally, Theorem 1.18 is proven in Section 4 by repeating the arguments of Theorem 1.5 for the sorting networks conditional on the event . This requires asymptotic analysis of entries of standard Young tableau of a slightly different shape, which is still possible by our methods. Thus, we arrive at the identities of Theorem 1.18 in an indirect way, by showing that the two sides of the identity are limits of the same discrete distributions.
Acknowledgements
We thank the anonymous referee for helpful comments. The work of V.G. was partially supported by NSF Grants DMS-1664619, DMS-1949820, DMS-2246449, and by the Office of the Vice Chancellor for Research and Graduate Education at the University of Wisconsin–Madison with funding from the Wisconsin Alumni Research Foundation.
2 Preliminaries
One of the key tools in studying sorting networks is the Edelman-Green bijection (see Section 4.2 for the details), which maps them to Standard Young Tableaux (SYT). In this and the next section we develop asymptotic results for SYT, which will be ultimately used in Section 4 to prove the theorems about the random sorting networks announced in the introduction.
In addition, in the last subsection of this section we recall the properties of the eigenvalues of aGUE, which will be useful later on.
2.1 Young diagrams and Young tableaux
A partition is a sequence of non-negative integers such that . The length of , denoted by , is the number of strictly positive .
We identify a partition with a collection of boxes, such that there are boxes in row . We use coordinate to denote the j-th box at row i. In other words, the coordinates of boxes are . Such a combinatorial object is named a Young diagram, still denoted as . In particular, we say the Young diagram is of staircase shape and write for some , if for . We also use the diagram for , which has for and .
A standard Young tableau (SYT) of shape is an insertion of numbers 1,2,…, into boxes of without repetitions, such that the numbers in each row and column strictly increase from left to right and from top to bottom. The numbers within a SYT are its entries. Fixing the Young diagram , we can get a uniformly random SYT of shape , denoted as by considering a uniform measure on the space of all SYT of shape . See Figure 2 for an example with .
A Poissonized Young tableau (PYT) of shape is an insertion of real numbers in into boxes of , such that the numbers strictly increase along each row and column. We use to denote the space of all PYT of shape . Note that a given PYT of shape can be transformed canonically to a SYT of the same shape, by replacing the k-th smallest entry with k. In the opposite direction, we can get a uniformly random PYT of shape , denoted as by the following steps: first generate a uniformly random SYT of shape , then independently sample i.i.d. random variables uniform in [0,1], and replace the entry with the k-th smallest random variable.

2.2 Rotation of Young Tableaux
Let denote a standard Young tableau of shape . We make a change of coordinates for the entries in by letting
We let denote the entry of on the row and the column, where and are reconstructed from and by inverting the above formulas. In more detail, we have
The allowed values for are: , , . The allowed values for are and . The transformation is essentially a clockwise rotation by 45 degrees, as can be seen in Figure 2.
Similarly, for standard Young tableau we make the same change of coordinates as above and deal with . Formally, the entry at does not exist, because it corresponds to the removed box at , but we can also think about this entry as being ; in this way becomes with the additional restriction .
For Poissonized Young tableau (PYT) of shapes or , we define the change of coordinate in exactly the same way and use coordinate system.
2.3 Point processes associated with PYT
Let be a of shape . As in Section 2.2, we will focus on the case or , for which we induce a point process on from its entries.
Definition 2.1.
The projection of a with shape or is a point configuration on , such that there’s a particle at , if for some and ,
By definition, is the lowest particle on level (except for level in , where is missing), and the particles on neighboring levels interlace by the setting of PYT, see Figure 3.


If we take to be uniformly random, then the projection of becomes a point process on . Note that this random point configuration is simple almost surely. We also would like to rescale the –coordinate.
Definition 2.2.
The point process on is obtained by taking a uniformly random , projecting its entries on , and then rescaling the second coordinate of each particle by the map . Similarly, is obtained from uniformly random PYT by the same procedure.
We study the asymptotic behavior of and in Section 3.3. Our analysis uses the technique of determinantal point processes with correlation kernels given by double contour integrals, which we present next.
2.4 Determinantal Point Process
Here’s a brief review of some standard definitions, for a thorough introduction see [B, DVe]. Let be a locally compact Polish space, say or . We consider the space of discrete subsets in , which consists of countable subsets of without accumulation points, and identify each such point configuration with the measure . The topology of this space is given as the weak topology of Radon measures on . We further use Borel algebra corresponding to this topology.
A point process is a random discrete subset of , and we assume that it is simple a.s, i.e, all the points have distinct positions. A determinantal point process on is a point process with correlation kernel , and a reference measure on , such that for any with compact support,
(2.1) |
Expectations of the form given by the l.h.s. of (2.1) determine the law of under mild conditions on , see [Le]. This will always be the case in this text as the correlation kernels we consider will be continuous.
The determinantal point processes appearing in this paper live in space or , with reference measure being the product of counting measure and Lebesgue measure, denoted by or resp. We use the following lemma, whose proof we omit, see [DVe], [Le].
Lemma 2.3.
Let be a determinantal point process on with reference measure and correlation kernel . Then the sequence converges weakly as to a determinantal point process with correlation kernel on (with reference measure ) and with almost surely no particles at , , if
uniformly on compact subsets of .
2.5 Uniformly Random Projection
When and are uniformly random, their projections are point processes on , whose correlation functions were computed in [GR]. We restate the result there for our case in the following theorem, which is a stepping stone of the proofs of our main results.
Theorem 2.4.
The processes and of Definition 2.2 are determinantal point process on with correlation kernel given for and , by
(2.2) | ||||
where
with for and for . The contours and are as shown in Figure 4. Both are counter-clockwise, encloses only the integers in the respective half open intervals and , and are arranged so that and are disjoint, as in Figure 4.

Remark 2.5.
The –functions in double-contour integrals are used for the convenience of notations, but the integral is, in fact, a rational function of and multiplied by . Indeed, we have
(2.3) |
Proof of Theorem 2.4.
Lemma 2.6.
Let be a determinantal point process on with reference measure and correlation kernel . For , , we define the rescaled and shifted point process
Then, the correlation kernel of with reference measure is
Remark 2.7.
We can extend and to point processes on , in such a way that almost surely there is no particle at for any . Moreover, the kernel can be extended to the same space based on its regular behavior at or .
In order to see that the double contour integral in the definition of the kernel is well-behaved as or , we expand the integral as a sum of residues at the poles inside the integration contours. We start from the case with contours of the top panel of Figure 4. In this case the denominator is never singular inside the integration contours, the –residues are at simple poles at finitely many non-negative integers and –residues are also at simple poles at finitely many non-negative integers. The residue at , for has the form , where does not depend on . Hence, the residue is continuous as and so is the double integral.
Proceeding to the case with contours of the bottom panel of Figure 4, the additional feature is the potential singularity of the denominator . Let us first compute the –integral as a sum of the residues, getting a sum of the terms of the form (one-dimensional -integral) with non-negative values for . Up to the factors not depending on or , the corresponding –integral has the form
The simple poles of the last integral lead to –dependence of the form , , which is again continuous at . It would be more complicated, if the integral had a double pole: the residue at such a pole would have involved the –derivative of , which leads to the appearance of factor that is singular at . However, we claim that there is no double pole. Indeed, using (2.3), . Therefore, for the pole of , we have . Hence, this pole is outside the set of poles of .
2.6 Anti-symmetric GUE
Let us recall a statement from [FN, Theorems 2.9 and 3.3]:
Proposition 2.8.
For in Definition 1.2 and we have:
-
(a)
The corner of has real eigenvalues.
-
(b)
The eigenvalues come in pairs, i.e, is an eigenvalue of the corner if and only if is an eigenvalue of the corner. When is odd, is necessarily an eigenvalue of the corner.
-
(c)
The eigenvalues of the corner interlace with the eigenvalues of the corner a.s, which means that
(2.4) where and denotes the largest eigenvalue of the corner.
-
(d)
Conditionally on the positive eigenvalues of the corner, the positive eigenvalues of the corners, , are jointly distributed uniformly on the polytope determined by the interlacing inequalities (2.4).
-
(e)
The joint density of positive eigenvalues of the corner is
where is an explicitly known normalization constant.
Note that the above five properties uniquely characterize the joint distribution of . An alternative characterization is through the correlation functions of the corresponding point process :
3 Local Limit of Standard Staircase Shaped Tableaux
The aim of this section is to investigate the scaling limit near the bottom-left corner for the uniformly random standard Young tableaux of shapes and as .
3.1 Statement of the result
We use the coordinate system for SYT, as in Section 2.2. We recall that is the point process corresponding to the corners of aGUE as in Definition 1.2. Let denote the particle of in the level , sorted in the increasing order; in other words, is the positive eigenvalue of corner of the infinite aGUE matrix (with corresponding to the smallest positive and to the largest positive eigenvalue).
Further, we set to be the point process of Definition 1.16 with an extra particle inserted at . We let denote the particle of in the level , sorted in the increasing order.
Theorem 3.1.
For each , let and be uniformly random standard Young tableaux of shapes and , respectively, in the coordinate system of Section 2.2. Recall . Then for each fixed , as we have
(3.1) |
(3.2) |
in the sense of convergence of finite-dimensional distributions.
Remark 3.2.
The SYT can be thought as conditioned on its largest entry being at , and in (3.2), is set to be , corresponding to on the right. Hence, Theorem 3.1 suggests a conjecture: should have the same law as conditioned on having a particle at , i.e. on . Note that additional efforts are needed to prove this, because the topology of the convergence in Theorem 3.1 is not strong enough to make conclusions about the conditional distributions.
The rest of this section is devoted to the proof of Theorem 3.1. We first couple the two types of SYT with the uniformly random PYT of the same shape, and induce two point processes on respectively from these two PYT, as in Definition 2.2. Then we prove that and as , where and are determinantal point processes on with explicit correlation kernels given by double contour integrals. Next, we identify with , the corners process of aGUE of Definition 1.2.
3.2 Coupling with PYT
The first step of the proof is to introduce a coupling between SYT and PYT of the same shape and to show that under this coupling the difference between random SYT and PYT is negligible in the asymptotic regime of our interest.
Let and be uniformly random SYT and PYT, respectively, of shape . We couple these two tableaux in the following way: for , let be a uniformly random point on the simplex . Given a uniformly random SYT , we replace entry in by for ; it is straightforward to check that the result is a uniformly random PYT . In the opposite direction, is reconstructed by by replacing the th largest entry in by for each .
Lemma 3.3.
Fix any pair of integers with and . Using the coupling described above, we have for each
in probability, as .
Proof.
Let , where is a random variable. Then, by the construction of the coupling, . By Chebyshev’s inequality,
(3.3) |
For a fixed , the random variable is distributed according to Beta distribution with parameters and , which has mean and variance (see, e.g., [JKB, Chapter 25]).
(3.4) |
Thus,
(3.5) |
Since , (3.3) and (3.5) imply that as
In a similar way, we couple uniformly random SYT and PYT of shape : we let be a uniformly random point on the simplex . Given a uniformly random SYT , we replace entry in by for ; it is straightforward to check that the result is a uniformly random PYT . Repeating the proof of Lemma 3.3 we arrive at:
Lemma 3.4.
Fix any pair of integers with and . Using the coupling described above, we have for each
in probability, as .
3.3 Limiting Processes
In this section we analyze asymptotic behavior of the point processes associated to random PYT and by the procedure of Section 2.3.
Theorem 3.5.
The point process (corresponding to random PYT via Definition 2.2) converges weakly as to a point process on . is a determinantal point process (with respect to reference measure ) with correlation kernel
Here and ; when or is equal to , the kernel is defined as the limit as or tends to 0. are counterclockwise–oriented contours which enclose the integers in and , respectively, and are arranged so that and are disjoint, as in Figure 5.


Theorem 3.6.
The point process (corresponding to random PYT via Definition 2.2) converges weakly to a point process on . is a determinantal point process (with respect to reference measure ) with correlation kernel given by
Here and ; the value of when or equals 0 is to be understood as the limit as or tends to 0. The contours and are the same as in Theorem 3.5.
In the proof we use a known asymptotic property of the Gamma-function:
(3.6) |
where term is uniform as long as is uniformly bounded.
Proof of Theorem 3.5.
We show the convergence in distribution by verifying conditions of Lemma 2.3, namely, we show that the correlation kernel of , where is given by (2.2), converges to uniformly on compact subsets of and . Note that the just introduced multiplication by does not change the value of the correlations functions . Let us fix arbitrary , strictly positive , , and analyze the asymptotic behavior of the double contour integral in (2.2).
The first step is to deform the contours of integration from the ones of Figure 4 to the ones of Figure 5. Using (2.3), the –dependent factors of the integrand are
(3.7) |
The form of (3.7) implies that there are no –poles of the integrand to the right from the contour of Figure 4; hence, the value of the integral does not change in the deformation. For each fixed and large enough , on the new contours of Figure 5, the expression (3.7) rapidly decays as (uniformly in ). This observation allows us to control the part of the integral corresponding to large , and it remains to study the asymptotics of the integrand for finite and .
We use (3.6) to compute the asymptotic behavior of in (3.7):
Similarly,
Thus, for , we have
Hence, the integrand in the double contour integral part of , converges as to
Combining with the straightforward asymptotics of the indicator term, we conclude that for all strictly positive and ,
Clearly, the convergence is uniform over in compact subsets of . It remains to show that there is no explosion for or near , which is done in exactly the same way as we did in Remark 2.7, i.e. by interpreting the integral as a sum of residues at simple and poles at non-negative integers. ∎
3.4 Connection to Anti-symmetric GUE
In this section we prove that the distribution of the limiting process of Theorem 3.5 coincides with the corners process of aGUE.
Proposition 3.7.
of Theorem 3.5 for , , is equal to
(3.8) |
Remark 3.8.
When , we simplify . There’s only one particle on level which corresponds to the limit of the entry of the SYT at lower corner, with coordinate .
When , we simplify . There are two particles on level which correspond to the two entries of the SYT with coordinates and .
Proof of Proposition 3.7.
of Theorem 3.5 is given by
The –integral is evaluated as the sum of residues at simple poles at even integers and we have
which matches the –dependent factors in (3.8). The remaining –integral for the –th term becomes
The last integral is evaluated as the sum of the residues at . Using the fact that the residue of the Gamma function at a simple pole at , is , we get
which matches the –dependent factors in (3.8). ∎
Proposition 3.9.
Proof.
Our task is to show that (3.8) and (3.9) match. The proof is induction in . For , this is a content of Remark 3.8. Subtracting (3.8) at and at , we get
(3.10) |
On the other hand, subtracting (3.9) at and at , we get
(3.11) |
The explicit expression for the Hermite polynomials [Sz, Chapter 5.5] is
In order to match with –dependent part of (3.10) we write
Hence, (3.10) gets transformed into
Comparing with (3.11), it remains to show that:
Or, equivalently,
(3.12) |
The identity (3.12) is a corollary of the Rodrigues formula for Hermite polynomials [Sz, Chapter 5.5]
Indeed, we have
(3.13) |
Theorem 3.10.
Proof.
Step 1. For each , each of the processes and has particles on level (i.e. with the first coordinate ). Let us show the marginal distributions describing the joint law of these particles are the same. By Proposition 3.9, for these particles form a determinantal point process with kernel (3.9). By Theorem 2.9, for these particles also form a determinantal point process with kernel given by plugging into (2.5). The kernels match and, hence, so do the point processes.
Step 2. Let us now fix and compare the joint distribution of particles in and in on levels . By part d in Proposition 2.8, the conditional law of the particles on levels given particles on level is uniform (subject to interlacing conditions) for . On the other hand, possesses the same conditional uniformity, because it was obtained by a scaling limit from uniformly random and conditional uniformity is preserved in limit transitions. Combining with Step 1, we conclude that the joint distributions of particles on levels are the same for and . Since is arbitrary, we are done. ∎
3.5 Proof of Theorem 3.1
Theorems 3.5 and 3.6 show that the point process , induced from , the rescaled entries of PYT of shapes and , converge weakly as to and , respectively. In particular, treating the entries of the PYT of these two shapes as infinite random vectors, they converges in the sense of finite dimensional distribution to the positions of the corresponding particles in and , respectively.
On the other hand, by Lemma 3.3, there’s a coupling of PYT and SYT that, for any ,
converges to in probability. Combining the above two results, we obtain the convergence for finite dimensional distributions of
to the limits given by particles of and . It remains to use Theorem 3.10 to match with . ∎
4 Asymptotics for spacings
We start this section by studying general rotationally-invariant point processes on a circle, giving two different definitions of a spacing between particles in such processes, and proving that their distribution functions satisfy relations, which are discrete versions of (1.4) and (1.5). When specialized to the random sorting networks setting, these spacings are precisely the ones discussed in the introduction. After that we recall the Edelman-Greene bijection between standard Young tableaux and sorting networks. In the last subsection we combine all the ingredients to finish the proofs of Theorems 1.5, 1.11, 1.14, and 1.18.
4.1 Generalities on spacings
Consider a random point process on a discrete circle of length . We index the possible positions of the particles by , , …, in the clockwise order and refer to the midpoint between and as the origin, see Figure 6. Throughout this section we assume that the distribution of the point process is invariant under rotations of the circle; we also silently assume that almost surely has at least one particle.


We use to define several random variables with positive integer values. We define the waiting time to be the smallest value of such that there is a particle at position . We let the spacing to be the distance between two adjacent particles on the circle: one to the left from the origin and one to the right from the origin. By distance we mean here one plus the number of the holes between the particles (counted along the arc including the origin), so that if there are particles at positions and , then the distance is .
Next, we define the conditional spacing to be a random variable whose distribution is that of conditional on having a particle at position . Note that and are random variables (functions) on the probability space of , but should be defined on a different probability space.
Example 4.1.
Figure 6 gives two possible configurations of a point process on the discrete circle of length . On the first configuration, and . On the second configuration, . Since there is a particle at , we can also think of being equal to for the second configuration.
Let , , and be the probability mass functions of , , and , respectively:
We also define to the the probability that there is a particle at position (in other words, this is the first correlation function or density of the point process ).
Proposition 4.2.
For any rotationally invariant point process , we have
(4.1) |
(4.2) |
where is the forward difference operator: .
Proof.
We claim that
(4.3) |
Indeed, using rotational invariance of the law of , we have
Remark 4.3.
A continuous version of Proposition 4.2 for translationally invariant point processes on the real line says that the following is true (under technical regularity assumptions on the point process, which we do not spell out, and which are needed to guarantee the existence of all the densities below): Suppose that , and are spacing (distance between closest particles to the left and to the right from the origin), conditional spacing, and waiting time for an arrival of a particle, and let , , , be probability densities of these random variables. Then we have
where is equal to the first correlation function for the process.
4.2 Edelman-Greene bijection
The relation of the random Young tableaux (which we were studying in Sections 2 and 3) with the random sorting networks relies on the Edelman–Green bijection [EG], which we now present.
The bijection takes a standard Young tableau of shape as an input and outputs a sorting network. This correspondence maps the uniform measure on standard Young tableaux of shape to the uniform measure on sorting networks of size .
The bijection proceeds through the following algorithm, in which we use the standard coordinate system for Young tableaux, as in Section 2.1:
-
1.
Given a standard Young tableau , find the box which contains the largest entry of . Necessarily, this entry is and .
-
2.
Set the first swap of the corresponding sorting network to be .
-
3.
Define the sliding path in the following way: the first box of the path is . Compare the entries at and at , and take the box with the larger entry as the second box (if only one of the boxes is inside the Young diagram, then take that one). Repeat this procedure (each time decreasing by either the first of the second coordinate of the box and moving in the direction of the larger entry), until you arrive at the box , which is the last box of the sliding path.
-
4.
Slide the entries on the sliding path in such a way that the maximal entry is removed and all the remaining entries on the path are moved to the neighboring boxes on the path. No entry remains at after the sliding. Then we increase all entries by , and fill the box with new entry .
-
5.
After transformations of steps 1-4, becomes a new standard Young tableau of shape . Repeat steps 1-4 additional times to get swaps ,,…, of the sorting network.
Using the Edelman-Green bijection, the random variable of Theorem 1.5 — the first time swap occurs — gets recast in terms of the uniformly random Young tableau.
Lemma 4.4.
Let be a uniformly random standard Young tableau of shape in the rotated coordinate system of Section 2.2. The following distributional identity holds:
(4.5) |
Proof.
In each iteration of the Edelman–Greene algorithm the entry (in the standard coordinate system for Young tableaux, as in Section 2.1) grows by until it becomes , at which point the swap is added to the sorting network for the first time. The entry at in the coordinate system is the same as the entry in the rotated coordinate system, leading to (4.5). ∎
We can also compute the distribution of the conditional spacing of Definition 1.12.
Lemma 4.5.
Let be a uniformly random standard Young tableau of shape in the rotated coordinate system of Section 2.2. The following distributional identity holds:
(4.6) |
Proof.
In order to obtain the law of , we need to condition on the first swap being ; through the Edelman-Green bijection this is the same as conditioning on the largest entry in the tableau to be in the box in the coordinate system. Note that if we condition a uniformly random standard Young tableau of shape on the position of largest entry , then the rest is a uniformly random standard Young tableau of shape .
Once we do conditioning, the value of becomes the number of the iterations of the Edelman-Greene algorithm when the largest entry of the tableau is at for the second time. After the first step of the algorithm, the entry at is
and at each further step the entry grows by , until it reaches . Hence, we need additional iterations, matching the first case in (4.6). The second and the third cases correspond to the situations when has only one neighboring box in the tableau. ∎
4.3 Proofs of the main theorems
Proof of Theorem 1.5.
By Lemma 4.4, we need to find the asymptotics of
Recalling and using (3.1) in Theorem 3.1, we get (1.3) with the right-hand side being times the law of the smallest eigenvalue in aGUE. The law of the latter is given by the Fredholm determinant through Theorem 2.9 for the case and general properties of the determinantal point processes (cf.[B]). ∎
Remark 4.6.
Proof of Theorem 1.18.
Remark 4.7.
Proof of Theorem 1.11.
Let and . We have , and . Note that .
By translation invariance of Corollary 1.8,
if and . Hence, for any by Theorem 1.5
where is the distribution function . Smoothness of and the above computation show that converges in distribution as to a random vector , whose probability density we denote , . We also let ; we already know that , . Denoting and the partial derivatives of in the first and second coordinates, respectively, we have . Recall that is the scaled limit of that we are interested in. We have:
Differentiating the last identity in we get the desired (1.4). ∎
Proof of Theorem 1.14.
We fix and deal with periodic extension of the sorting network of Definition 1.7. Given a random sorting network , we define a point process on discrete circle of length by declaring that a spot is occupied by a particle if and only if ; in other words, this is the point process of times of swaps . By Corollary 1.8 the process is rotationally invariant and, therefore, the results of Section 4.1 apply. Since we would like to find the distribution of the conditional spacing, we rely on (4.2) — the desired quantity is in that formula, which is being connected to and . The asymptotics of is computed in Theorem 1.5. In order to find , we notice that by the Edelman–Greene bijection (cf. [AHRV, Proposition 9]), it is computed as
Both numerator and denominator are explicitly known (they can be computed by the hook length formula [FRT]). Applying the Stirling’s formula, we get as :
(4.7) |
Now take and sum (4.2) over . We get
(4.8) |
We send using Theorem 1.5, (4.7) and the following lemma, whose proof we leave as an exercise for the reader (the monotonicity condition is implied by (4.2)).
Lemma 4.8.
Let be –valued random variables, such that for each the probabilities depend on in a monotone way. Suppose that we have a distributional limit , where is an absolutely continuous random variable with density . Then for each
(4.9) |
Remark 4.9.
References
- [ADHV] O. Angel, D. Dauvergne, A. E. Holroyd, B. Virag, The Local Limit of Random Sorting Networks, Annales Institut Henri Poincare: Probability and Statististics 55, no. 1 (2019), 412–440. arXiv:1702.08368.
- [AGH] O. Angel, V. Gorin, A. E. Holroyd, A pattern theorem for random sorting networks. Electronic Journal of Probability, 17, paper 99 (2012), 1–16. arXiv:1110.0160.
- [AHRV] O. Angel, A. Holroyd, D. Romik, B. Virág, Random sorting networks, Advances in Mathematics 215, no. 2 (2007), 839–864, arXiv:0609538.
- [B] A. Borodin, Determinantal point processes, Oxford Handbook of Random Matrix Theory, Oxford University Press, 2011, arXiv:0911.1153.
- [DVe] D. Daley, D. Vere-Jones, An introduction to the theory of point processes: Vol. I. Elementary theory and methods, Springer–Verlag, New York, 2003.
- [D] D. Dauvergne, The Archimedean limit of random sorting networks, to appear in Journal of the American Mathematical Society, arXiv:1802.08934.
- [DVi] D. Dauvergne, B. Virag, Circular support in random sorting networks, Transactions of the American Mathematical Society 373 (2020), 1529-1553 arXiv:1802.08933.
- [EG] P. Edelman, C. Greene, Balanced tableaux, Advances in Mathematics 63, no. 1 (1987), 42–99.
- [GR] V. Gorin, M. Rahman, Random sorting networks: local statistics via random matrix laws, Probability Theory and Related Fields 175, no. 1 (2019), 45–96.
- [F] P. J. Forrester, Log-Gases and Random Matrices, Princeton University Press, 2010.
- [FRT] J. S. Frame, G. B. Robinson, and R. M. Thrall. The hook graphs of the symmetric groups. Canadian Journal of Mathematics, 6 (1954), 316–324.
- [FN] P. J. Forrester, E. Nordenstam, The anti-symmetric GUE minor process, Moscow Mathematical Journal 9, no. 4 (2009), 749–774, arXiv:0804.3293.
- [JKB] N. Johnson, S. Kotz, N. Balakrishnan, Continuous Univariate Distributions, vol. 2 (2nd edition). Wiley, 1995.
- [K] M. Kotowski, Limits of random permuton processes and large deviations for the interchange process, PhD Thesis, University of Toronto, 2016.
- [Le] A. Lenard, Correlation functions and the uniqueness of the state in classical statistical mechanics, Communications in Mathematical Physics 30 (1973), 35–44.
- [M] M. L. Mehta, Random matrices, third edition, Pure and Applied Mathematics (Amsterdam), vol. 142, Elsevier/Academic Press, Amsterdam, 2004.
- [P] L. Petrov, Asymptotics of random lozenge tilings via Gelfand-Tsetlin schemes, Probability Theory and Related Fields 160, np. 3 (2014), 429–487.
- [RVV] M. Rahman, B. Virag, M. Vizer, Geometry of Permutation Limits, Combinatorica 39 (2019), 933–960. arXiv:1609.03891.
- [R] A. Rozinov, Statistics of Random Sorting Networks, PhD Thesis, Courant Institute, NYU, 2016.
- [S] R. P. Stanley, On the number of reduced decompositions of elements of Coxeter groups, European Journal of Combinatorics 5, no. 4 (1984), 359–372.
- [Sz] G. Szego, Orthogonal polynomials. American Mathematical Society, 1939.