Asymptotics of Plancherel measures for symmetric groups

By Alexei Borodin, Andrei Okounkov, and Grigori Olshanski

Abstract

We consider the asymptotics of the Plancherel measures on partitions of as goes to infinity. We prove that the local structure of a Plancherel typical partition in the middle of the limit shape converges to a determinantal point process with the discrete sine kernel.

On the edges of the limit shape, we prove that the joint distribution of suitably scaled 1st, 2nd, and so on rows of a Plancherel typical diagram converges to the corresponding distribution for eigenvalues of random Hermitian matrices (given by the Airy kernel). This proves a conjecture due to Baik, Deift, and Johansson by methods different from the Riemann-Hilbert techniques used in their original papers and from the combinatorial proof given by the second author.

Our approach is based on an exact determinantal formula for the correlation functions of the poissonized Plancherel measures in terms of a new kernel involving Bessel functions. Our asymptotic analysis relies on the classical asymptotic formulas for the Bessel functions and depoissonization techniques.

1. Introduction

1.1. Plancherel measures

Given a finite group , by the corresponding Plancherel measure we mean the probability measure on the set of irreducible representations of which assigns to a representation the weight . For the symmetric group , the set is the set of partitions of the number , which we shall identify with Young diagrams with squares throughout this paper. The Plancherel measure on partitions arises naturally in representation–theoretic, combinatorial, and probabilistic problems. For example, the Plancherel distribution of the first part of a partition coincides with the distribution of the longest increasing subsequence of a uniformly distributed random permutation Reference 31.

We denote the Plancherel measure on partitions of by ,

where is the dimension of the corresponding representation of . The asymptotic properties of these measures as have been studied very intensively; see the References and below.

In the seventies, Logan and Shepp Reference 23 and, independently, Vershik and Kerov Reference 40Reference 42 discovered the following measure concentration phenomenon for as . Let be a partition of and let and be the usual coordinates on the diagrams, namely, the row number and the column number. Introduce new coordinates and by

that is, we flip the diagram, rotate it as in Figure 1, and scale it by the factor of in both directions.

After this scaling, the Plancherel measures converge as (see Reference 23Reference 40Reference 42 for precise statements) to the delta measure supported on the following shape:

where the function is defined by

The function is plotted in Figure 1. As explained in detail in Reference 22, this limit shape is closely connected to Wigner’s semicircle law for distribution of eigenvalues of random matrices; see also Reference 19Reference 20Reference 21.

From a different point of view, the connection with random matrices was observed in Reference 3Reference 4, and also in earlier papers Reference 16Reference 28Reference 29. In Reference 3, Baik, Deift, and Johansson made the following conjecture. They conjectured that in the limit and after proper scaling the joint distribution of , , becomes identical to the joint distribution of properly scaled largest eigenvalues of a Gaussian random Hermitian matrix (which form the so-called Airy ensemble; see Section 1.4). They proved this for the individual distribution of and in Reference 3 and Reference 4, respectively. A combinatorial proof of the full conjecture was given by one of us in Reference 25. It was based on an interplay between maps on surfaces and ramified coverings of the sphere.

In this paper we study the local structure of a typical Plancherel diagram both in the bulk of the limit shape and on its edge, where by the study of the edge we mean the study of the behavior of , , and so on.

We employ an analytic approach based on an exact formula in terms of Bessel functions for the correlation functions of the so-called poissonization of the Plancherel measures (see Theorem 1 in the following subsection), and the so-called depoissonization techniques (see Section 1.4).

The exact formula in Theorem 1 is a limit case of a formula from Reference 8; see also the recent papers Reference 26Reference 27 for more general results. The use of poissonization and depoissonization is very much in the spirit of Reference 3Reference 16Reference 39 and represents a well–known statistical mechanics principle of the equivalence of canonical and grand canonical ensembles.

Our main results are the following two. In the bulk of the limit shape , we prove that the local structure of a Plancherel typical partition converges to a determinantal point process with the discrete sine kernel; see Theorem 3. This result is parallel to the corresponding result for random matrices. On the edge of the limit shape, we give an analytic proof of the Baik-Deift-Johansson conjecture; see Theorem 4. These results will be stated in Sections 1.3 and 1.4 of the present Introduction, respectively.

Simultaneously and independently, results equivalent to our Theorems 2 and 4 were obtained by K. Johansson Reference 17.

1.2. Poissonization and correlation functions

For , consider the poissonization of the measures :

This is a probability measure on the set of all partitions. Our first result is the computation of the correlation functions of the measures .

By correlation functions we mean the following. By definition, set

Also, following Reference 41, define the modified Frobenius coordinates of a partition by

where stands for the symmetric difference of two sets, is the number of squares on the diagonal of , and ’s and ’s are the usual Frobenius coordinates of . Recall that is the number of squares in the th row to the right of the diagonal, and is number of squares in the th column below the diagonal. The equality Equation 1.2 is a well–known combinatorial fact discovered by Frobenius; see Ex. I.1.15(a) in Reference 24. Note that, in contrast to , the set is infinite and, moreover, it contains all but finitely many negative integers.

The sets and have the following nice geometric interpretation. Let the diagram be flipped and rotated as in Figure 1, but not scaled. Denote by a piecewise linear function with whose graph is given by the upper boundary of completed by the lines

Then

In other words, if we consider as a history of a walk on , then are those moments when a step is made in the negative direction. It is therefore natural to call the descent set of . As we shall see, the correspondence is a very convenient way to encode the local structure of the boundary of .

The halves in the definition of have the following interpretation: one splits the diagonal squares in half and gives half to the rows and half to the columns.

Definition 1.1.

The correlation functions of are the probabilities that the sets or, similarly, contain a fixed subset . More precisely, we set

Theorem 1.

For any we have

where the kernel is given by the following formula:

The functions are defined by

where is the Bessel function of order and argument . The diagonal values are determined by the l’Hospital rule.

This theorem is established in Section 2.1; see also Remark 1.2 below. By the complementation principle (see Sections A.3 and 2.2), Theorem 1 is equivalent to the following

Theorem 2.

For any we have

Here the kernel is given by the following formula:

where . The diagonal values are determined by the l’Hospital rule.

Remark 1.2.

Theorem 1 is a limit case of Theorem 3.3 of Reference 8. For the reader’s convenience a direct proof of it is given in Section 2. Various limit cases of the results of Reference 8 are discussed in Reference 9. By different methods, the formula Equation 1.8 was obtained by K. Johansson Reference 17.

A representation–theoretic proof of a more general formula than Theorem 3.3 of Reference 8 has been subsequently given in Reference 27Reference 26; see also Reference 7.

Remark 1.3.

Observe that all Bessel functions involved in the above formulas are of integer order. Also note that the ratios like are entire functions of and because is an entire function of . In particular, the values are well defined. Various denominator–free formulas for the kernel are given in Section 2.1.

1.3. Asymptotics in the bulk of the spectrum

Given a sequence of subsets

where is some fixed integer, we call this sequence regular if the limits

exist, finite or infinite. Here . Observe that if is finite, then for .

In the case when can be represented as and the distance between and goes to as we shall say that the sequence splits; otherwise, we call it nonsplit. Obviously, is nonsplit if and only if all stay at a finite distance from each other.

Define the correlation functions of the measures by the same rule as in Equation 1.4:

We are interested in the limit of as . This limit will be computed in Theorem 3 below. As we shall see, if splits, then the limit correlations factor accordingly.

Introduce the following discrete sine kernel which is a translation invariant kernel on the lattice ,

depending on a real parameter :

Note that and for we have

where is the Tchebyshev polynomials of the second kind. We agree that

and also that

The following result describes the local structure of a Plancherel typical partition.

Theorem 3.

Let be a regular sequence and let the numbers , be defined by Equation 1.10, Equation 1.11. If splits, that is, if and the distance between and goes to as , then

If is nonsplit, then

where is the discrete sine kernel and  .

We prove this theorem in Section 3.

Remark 1.4.

Notice that, in particular, Theorem 3 implies that, as , the shape of a typical partition near any point of the limit curve is described by a stationary random process. For distinct points on the curve these random processes are independent.

Remark 1.5.

By complementation (see Sections A.3 and 3.2), one obtains from Theorem 3 an equivalent statement about the asymptotics of the following correlation functions:

Remark 1.6.

The discrete sine kernel was studied before (see Reference 44Reference 45), mainly as a model case for the continuous sine kernel. In particular, the asymptotics of Toeplitz determinants built from the discrete sine kernel was obtained by H. Widom Reference 45 who was answering a question of F. Dyson. We thank S. Kerov for pointing out this reference.

Remark 1.7.

Note that, in particular, Theorem 3 implies that the limit density (the 1-point correlation function) is given by

This is in agreement with the Logan-Shepp-Vershik-Kerov result about the limit shape . More concretely, the function is related to the density Equation 1.14 by

which can be interpreted as follows. Approximately, we have

Set . Then the above relation reads and it should be satisfied on the boundary of the limit shape. Since , we conclude that

as was to be shown.

Remark 1.8.

The discrete sine kernel becomes especially nice near the diagonal, that is, where . Indeed,

1.4. Behavior near the edge of the spectrum and the Airy ensemble

The discrete sine kernel vanishes if . Therefore, it follows from Theorem 3 that the limit correlations vanish if for some . However, as will be shown below in Proposition 4.1, after a suitable scaling near the edge , the correlation functions converge to the correlation functions given by the Airy kernel Reference 12Reference 36

Here is the Airy function:

In fact, the following more precise statement is true about the behavior of the Plancherel measure near the edge . By symmetry, everything we say about the edge applies to the opposite edge .

Consider the random point process on whose correlation functions are given by the determinants

and let

be its random configuration. We call the random variables ’s the Airy ensemble. It is known Reference 12Reference 36 that the Airy ensemble describes the behavior of the (properly scaled) 1st, 2nd, and so on largest eigenvalues of a Gaussian random Hermitian matrix. The distribution of individual eigenvalues was obtained by Tracy and Widom in Reference 36 in terms of certain Painlevé transcendents.

It has been conjectured by Baik, Deift, and Johansson that the random variables

converge, in distribution and together with all moments, to the Airy ensemble. They verified this conjecture for individual distribution of and in Reference 3 and Reference 4, respectively. In particular, in the case of , this generalizes the result of Reference 40Reference 42 that in probability as . The computation of was known as the Ulam problem; different solutions to this problem were given in Reference 1Reference 16Reference 32; see also the survey Reference 2.

Convergence of all expectations of the form

to the corresponding quantities for the Airy ensemble was established in Reference 25. The proof in Reference 25 was based on a combinatorial interpretation of Equation 1.16 as the asymptotics in a certain enumeration problem for random surfaces.

In the present paper we use different ideas to prove the following

Theorem 4.

As , the random variables converge, in joint distribution, to the Airy ensemble.

This is done in Section 4 using methods described in the next subsection. The result stated in Theorem 4 was independently obtained by K. Johansson in Reference 17. See, for example, Reference 13 for an application of Theorem 4.

1.5. Poissonization and depoissonization

We obtain Theorems 3 and 4 from Theorem 1 using the so-called depoissonization techniques. We recall that the fundamental idea of depoissonization is the following.

Given a sequence its poissonization is, by definition, the function

Provided the ’s grow not too rapidly this is an entire function of . In combinatorics, it is usually called the exponential generating function of the sequence . Various methods of extracting asymptotics of sequences from their generating functions are classically known and widely used (see for example Reference 39 where such methods are used to obtain the limit shape of a typical partition under various measures on the set of partitions).

A probabilistic way to look at the generating function Equation 1.17 is the following. If , then is the expectation of where is a Poisson random variable with parameter . Because has mean and standard deviation , one expects that

provided the variations of for are small. One possible regularity condition on which implies Equation 1.18 is monotonicity. In a very general and very convenient form, a depoissonization lemma for nonincreasing nonnegative was established by K. Johansson in Reference 16. We use this lemma in Section 4 to prove Theorem 4.

Another approach to depoissonization is to use a contour integral

where is any contour around . Suppose, for a moment, that is constant, . The function has a unique critical point . If we choose as the contour , then only neighborhoods of size contribute to the asymptotics of Equation 1.19. Therefore, for general , we still expect that provided the overall growth of is under control and the variations of for are small, the asymptotically significant contribution to Equation 1.19 will come from . That is, we still expect Equation 1.18 to be valid. See, for example, Reference 15 for a comprehensive discussion and survey of this approach.

We use this approach to prove Theorem 3 in Section 3. The growth conditions on which are suitable in our situation are spelled out in Lemma 3.1.

In our case, the functions are combinations of the Bessel functions. Their asymptotic behavior as can be obtained directly from the classical results on asymptotics of Bessel functions which are discussed, for example, in the fundamental Watson’s treatise Reference 43. These asymptotic formulas for Bessel functions are derived using the integral representations of Bessel functions and the steepest descent method. The different behavior of the asymptotics in the bulk of the spectrum, near the edges of the spectrum, and outside of is produced by the different location of the saddle point in these three cases.

1.6. Organization of the paper

Section 2 contains the proof of Theorems 1 and 2 and also various formulas for the kernels and . We also discuss a difference operator which commutes with and its possible applications.

Section 3 deals with the behavior of the Plancherel measure in the bulk of the spectrum; there we prove Theorem 3. Theorem 4 and a similar result (Theorem 5) for the poissonized measure are established in Section 4.

At the end of the paper there is an Appendix, where we collected some necessary results about Fredholm determinants, point processes, and convergence of trace class operators.

2. Correlation functions of the measures

2.1. Proof of Theorem 1

As noted above, Theorem 1 is a limit case of Theorem 3.3 of Reference 8. That theorem concerns a family of probability measures on partitions of , where are certain parameters. When the parameters go to infinity, tends to the Plancherel measure . Theorem 3.3 in Reference 8 gives a determinantal formula for the correlation functions of the measure

in terms of a certain hypergeometric kernel. Here and is an additional parameter. As and , the negative binomial distribution in Equation 2.1 tends to the Poisson distribution with parameter . In the same limit, the hypergeometric kernel becomes the kernel of Theorem 1. The Bessel functions appear as a suitable degeneration of hypergeometric functions.

Recently, these results of Reference 8 were considerably generalized in Reference 26, where it was shown how this type of correlation functions can be computed using simple commutation relations in the infinite wedge space.

For the reader’s convenience, we present here a direct and elementary proof of Theorem 1 which uses the same ideas as in Reference 8 plus an additional technical trick, namely, differentiation with respect to which kills denominators. This trick yields a denominator–free integral formula for the kernel ; see Proposition 2.7. Our proof here is a verification, not a derivation. For more conceptual approaches the reader is referred to Reference 26Reference 27Reference 7.

Let . Introduce the following kernel :

We shall consider the kernels and as operators in the space on .

We recall that simple multiplicative formulas (for example, the hook formula) are known for the number in Equation 1.1. For our purposes, it is convenient to rewrite the hook formula in the following determinantal form. Let be the Frobenius coordinates of ; see Section 1.2. We have

The following proposition is a straightforward computation using Equation 2.2.

Proposition 2.1.

Let be a partition. Then

where are the modified Frobenius coordinates of defined in Equation 1.2.

Let be the push-forward of under the map . Note that the image of consists of sets having equally many positive and negative elements. For other , the right-hand side of Equation 2.3 can be easily seen to vanish. Therefore is a determinantal point process (see the Appendix) corresponding to , that is, its configuration probabilities are determinants of the form Equation 2.3.

Corollary 2.2.

.

This follows from the fact that is a probability measure. This is explained in Propositions A.1 and A.4 in the Appendix. Note that, in general, one needs to check that is a trace class operator.⁠Footnote1 However, because of the special form of , it suffices to check a weaker claim – that is a Hilbert–Schmidt operator, which is immediate.

1

Actually, is of trace class because the sum of the absolute values of its matrix elements is finite. We are grateful to P. Deift for this remark.

Theorem 1 now follows from general properties of determinantal point processes (see Proposition A.6 in the Appendix) and the following

Proposition 2.3.

.

We shall need three identities for Bessel functions which are degenerations of the identities (3.13–15) in Reference 8 for the hypergeometric function. The first identity is due to Lommel (see Reference 43, Section 3.2, or Reference 14, 7.2.(60)):

The other two identities are the following.

Lemma 2.4.

For any and any we have

Proof.

Another identity due to Lommel (see Reference 43, Section 5.23, or Reference 14, 7.15.(10)) reads

Substituting we get Equation 2.5. Substituting yields

Let be the difference of the left-hand side and the right-hand side in Equation 2.6. Using Equation 2.7 and the recurrence relation

we find that . Hence for any it is a periodic function of and it suffices to show that . Clearly, the left-hand side in Equation 2.6 goes to 0 as . From the defining series for it is clear that

which implies that the right-hand side of Equation 2.6 also goes to as . This concludes the proof.

Proof of Proposition 2.3.

It is convenient to set . Since the operator is invertible we have to check that

This is clearly true for ; therefore, it suffices to check that

where and . Using the formulas

one computes

where . Similarly,

Now the verification of Equation 2.10 becomes a straightforward application of the formulas Equation 2.5 and Equation 2.6, except for the occurrence of the singularity in those formulas. This singularity is resolved using Equation 2.4. This concludes the proof of Proposition 2.3 and Theorem 1.

2.2. Proof of Theorem 2

Recall that by construction

Let us check that this and Proposition A.8 imply Theorem 2. In Proposition A.8 we substitute

By definition, set

We have the following

Lemma 2.5.

.

It is clear that since the -factors cancel out of all determinantal formulas, this lemma and Proposition A.8 establish the equivalence of Theorems 1 and 2.

Proof.

Using the relation

and the definition of one computes

Clearly, the relation Equation 2.12 remains valid for . It remains to consider the case . In this case we have to show that

Rewrite it as

By Equation 2.14 this is equivalent to

Examine the right-hand side. The terms with vanish because then . The term with is equal to 1, which corresponds to 1 in the left-hand side. Next, the terms with vanish because for these values of , the expression vanishes. Finally, for , set . Then the th term in the second sum is equal to minus the th term in the first sum. Indeed, this follows from the trivial relation

This concludes the proof.

2.3. Various formulas for the kernel

Recall that since is an entire function of , the function is entire in and . We shall now obtain several denominator–free formulas for the kernel .

Proposition 2.6.
Proof.

Straightforward computation using a formula due to Nielsen (see Section 5.41 of Reference 43 or Reference 14, formula 7.2.(48)).

Proposition 2.7.

Suppose . Then

Proof.

Follows from a computation done in the proof of Proposition 2.3,

and the following corollary of Equation 2.14:

Remark 2.8.

Observe that by Proposition 2.7 the operator is a sum of two operators of rank 1.

Proposition 2.9.
Proof.

Our argument is similar to an argument due to Tracy and Widom; see the proof of the formula (4.6) in Reference 36. The recurrence relation Equation 2.8 implies that

Consequently, the difference between the left-hand side and the right-hand side of Equation 2.15 is a function which depends only on . Let and go to infinity in such a way that remains fixed. Because of the asymptotics Equation 2.9 both sides in Equation 2.15 tend to zero and, hence, the difference actually is 0.

In the same way as in Reference 36 this results in the following

Corollary 2.10.

For any , the restriction of the kernel to the subset defines a nonnegative trace class operator in the space on that subset.

Proof.

By Proposition 2.9, the restriction of on is the square of the kernel . Since the latter kernel is real and symmetric, the kernel is nonnegative. Hence, it remains to prove that its trace is finite. Again, by Proposition 2.9, this trace is equal to

This sum is clearly finite by Equation 2.9.

Remark 2.11.

The kernel resembles a Christoffel–Darboux kernel and, in fact, the operator in defined by the kernel is an Hermitian projection operator. Recall that , where is of the form

One can prove that this together with Lemma 2.5 imply that is an Hermitian projection kernel. However, in contrast to a Christoffel–Darboux kernel, it projects to an infinite–dimensional subspace.

Note that in Reference 17 the restriction of the kernel to was obtained as a limit of Christoffel–Darboux kernels for Charlier polynomials.

2.4. Commuting difference operator

Consider the difference operators and on the lattice ,

Note that as operators on . Consider the following second order difference Sturm–Liouville operator:

where and are operators of multiplication by certain functions , . The operator Equation 2.17 is self–adjoint in . A straightforward computation shows that

It follows that if for a certain , then the space of functions vanishing for is invariant under .

Proposition 2.12.

Let denote the operator in obtained by restricting the kernel to . Then the difference Sturm–Liouville operator Equation 2.17 commutes with provided

Proof.

Since is the square of the operator with the kernel , it suffices to check that the latter operator commutes with , with the above choice of and . But this is readily checked using Equation 2.18.

This proposition is a counterpart of a known fact about the Airy kernel; see Reference 36. Moreover, in the scaling limit when and

the difference operator becomes, for a suitable choice of the constant, the differential operator

which commutes with the Airy operator restricted to . The above differential operator is exactly that of Tracy and Widom Reference 36.

Remark 2.13.

Presumably, this commuting difference operator can be used to obtain, as was done in Reference 36 for the Airy kernel, asymptotic formulas for the eigenvalues of , where and . Such asymptotic formulas may be very useful if one wishes to refine Theorem 4 and to establish convergence of moments in addition to convergence of distribution functions. For individual distributions of and the convergence of moments was obtained, by other methods, in Reference 3Reference 4.

3. Correlation functions in the bulk of the spectrum

3.1. Proof of Theorem 3

We refer the reader to Section 1.3 of the Introduction for the definition of a regular sequence and the statement of Theorem 3. Also, in this section, we shall be working in the bulk of the spectrum, that is, we shall assume that all numbers defined in Equation 1.10 lie inside . The edges of the spectrum and its exterior will be treated in the next section.

In our proof, we shall follow the strategy explained in Section 1.5. Namely, in order to compute the limit of we shall use the contour integral

compute the asymptotics of for , and estimate away from . Both tasks will be accomplished using classical results about the Bessel functions.

We start our proof with the following lemma which formalizes the above informal depoissonization argument. The hypothesis of this lemma is very far from optimal, but it is sufficient for our purposes. For the rest of this section, we fix a number which shall play an auxiliary role.

Lemma 3.1.

Let be a sequence of entire functions

and suppose that there exist constants and such that

as . Then

Proof.

By replacing by , we may assume that . By Cauchy and Stirling formulas, we have

Choose some large and split the circle into two parts as follows:

The inequality Equation 3.1 and the equality

imply that the integral decays exponentially provided is large enough. On , the inequality Equation 3.2 applies for sufficiently large and gives

Therefore, the integral is of the following integral:

Hence, and the lemma follows.

Definition 3.2.

Denote by the algebra (with respect to term-wise addition and multiplication) of sequences which satisfy the properties Equation 3.1 and Equation 3.2 for some, depending on the sequence, constants and . Introduce the map

which is clearly a homomorphism.

Remark 3.3.

Note that we do not require to be entire. Indeed, the kernel may have a square root branching; see the formula Equation 2.14.

By Theorem 2, the correlation functions belong to the algebra generated by sequences of the form

where the sequence is regular which, we recall, means that the limits

exist, finite or infinite. Therefore, we first consider such sequences.

Proposition 3.4.

If is regular, then

In the proof of this proposition it will be convenient to allow . For complex sequences we shall require ; the number may be arbitrary.

Lemma 3.5.

Suppose that a sequence is as above and, additionally, suppose that , are bounded and . Then the sequence satisfies Equation 3.2 with and certain .

Proof of Lemma 3.5.

We shall use Debye’s asymptotic formulas for Bessel functions of complex order and large complex argument; see, for example, Section 8.6 in Reference 43. Introduce the function

The formula Equation 1.9 can be rewritten as

The asymptotic formulas for Bessel functions imply that

where

provided that in such a way that stays in some neighborhood of ; the precise form of this neighborhood can be seen in Figure 22 in Section 8.61 of Reference 43. Because we assume that

and because , the ratios , stay close to . For future reference, we also point out that the constant in in Equation 3.4 is uniform in provided is bounded away from the endpoints .

First we estimate . The function clearly takes real values on the real line. From the obvious estimate

and the boundedness of , , and we obtain an estimate of the form

If , then because of the denominator in Equation 3.3 the estimate Equation 3.5 implies that

Since , it follows that in this case the lemma is established.

Assume, therefore, that is finite. Observe that for any bounded increment we have

and, in particular, the last term is . Using the trigonometric identity

and observing that

we compute

Since, by hypothesis,

and , the lemma follows.

Remark 3.6.

Below we shall need this lemma for a variable sequence . Therefore, let us spell out explicitly under what conditions on the estimates in Lemma 3.5 remain uniform. We need the sequences and to converge uniformly; then, in particular, the ratios and are uniformly bounded away from . Also, we need and to be uniformly bounded. Finally, we need to be uniformly bounded from below.

Proof of Proposition 3.4.

First, we check the condition Equation 3.2. In the case this was done in the previous lemma. Suppose, therefore, that is a regular sequence in and consider the asymptotics of .

Because the function is an entire function of and we have

where is arbitrary; we shall take to be some small but fixed number. From the previous lemma we know that

From the above remark it follows that this estimate is uniform in . This implies the property Equation 3.2 for .

To prove the estimate Equation 3.1 we use Schläfli’s integral representation (see Section 6.21 in Reference 43)

which is valid for and even for provided or .

If , then the second summand in Equation 3.8 vanishes and the first summand is uniformly in . This implies the estimate Equation 3.1 provided .

It remains, therefore, to check Equation 3.1 for where is a regular sequence. Again, we use Equation 3.7. Observe that since , the second summand in Equation 3.8 is uniformly small provided is bounded from above and is bounded from below. Therefore, Equation 3.7 produces the Equation 3.1 estimate for . For we use the relation Equation 2.13 and the reccurence Equation 2.16 to obtain the estimate.

Proof of Theorem 3.

Let be a regular sequence and let the numbers and be defined by Equation 1.10, Equation 1.11. We shall assume that for all . The validity of the theorem in the case when for some will be obvious from the results of the next section.

We have

where the first line is the definition of and the second is Theorem 2. From Equation 3.9 it is obvious that is entire. Therefore, we can apply Lemma 3.1 to it. It is clear that Lemma 3.1, together with Proposition 3.4, implies Theorem 3. The factorization Equation 1.12 follows from the vanishing .

3.2. Asymptotics of

Recall that the correlation functions were defined by

The asymptotics of these correlation functions can be easily obtained from Theorem 3 by complementation (see Sections A.3 and 2.2), and the result is the following.

Let be a regular sequence. If it splits, then factors as in Equation 1.12. Suppose therefore, that is nonsplit. Here one has to distinguish two cases. If or , then we shall say that this sequence is off-diagonal. Geometrically, it means that corresponds to modified Frobenius coordinates of only one kind: either the row ones or the column ones. For off-diagonal sequences we obtain from Theorem 3 by complementation that

where is the discrete sine kernel and .

If is nonsplit and diagonal, that is, if it is nonsplit and includes both positive and negative numbers, then one has to assume additionally that the number of positive and negative elements of stabilizes for sufficiently large . In this case the limit correlations are given by the kernel

Remark that this kernel is not translation invariant. Note, however, that

provided and have the same sign and similarly for . Therefore, if the subsets and , , have the same number of positive and negative elements, then

4. Edge of the spectrum: Convergence to the Airy ensemble

4.1. Results and strategy of proof

In this section we prove Theorem 4 which was stated in Section 1.4 of the Introduction. We refer the reader to Section 1.4 for a discussion of the relation between Theorem 4 and the results obtained in Reference 3Reference 4Reference 25.

Recall that the Airy kernel was defined as

where is the Airy function Equation 1.15. The Airy ensemble is, by definition, a random point process on , whose correlation functions are given by

This ensemble was studied in Reference 36. We denote by a random configuration of the Airy ensemble. Theorem 4 says that after a proper scaling and normalization, the rows of a Plancherel random partition converge in joint distribution to the Airy ensemble. Namely, the random variables ,

converge, in joint distribution, to the Airy ensemble as .

In the proof of Theorem 4, we shall follow the strategy explained in Section 1.5 of the Introduction. First, we shall prove that under the poissonized measure on the set of partitions , the random variables converge, in joint distribution, to the Airy ensemble as . This result is stated below as Theorem 5. From this, using certain monotonicity and Lemma 4.7 which is due to K. Johansson, we shall conclude that the same is true for the measures as .

The proof of Theorem 5 will be based on the analysis of the behavior of the correlation functions of , , near the point . From the expression for correlation functions of given in Theorem 1 it is clear that this amounts to the study of the asymptotics of when . This asymptotics is classically known and from it we shall derive the following

Proposition 4.1.

Set . We have

uniformly in and on compact sets of .

The prefactor corresponds to the fact that we change the local scale near to get nonvanishing limit correlations.

Using this and verifying certain tail estimates we obtain the following

Theorem 5.

For any fixed and any we have

where is the Airy ensemble.

Observe that the limit behavior of is, obviously, identical with the limit behavior of similarly scaled st, nd, and so on maximal Frobenius coordinates.

Proofs of Proposition 4.1 and Theorem 5 are given Section 4.2. In Section 4.3, using a depoissonization argument based on Lemma 4.7 we deduce Theorem 4.

Remark 4.2.

We consider the behavior of any number of first rows of , where is a Plancherel random partition. By symmetry, same results describe the behavior of any number of first columns of .

4.2. Proof of Theorem 5

Suppose we have a point process on with determinantal correlation functions

for some kernel . Let be a possibly infinite interval . By we denote the operator in obtained by restricting the kernel on . Assume is a trace class operator. Then the intersection of the random configuration with is finite almost surely and

In particular, the probability that is empty is equal to

More generally, if is a finite family of pairwise nonintersecting intervals such that the operators are trace class, then

Here operators are considered to be acting in the same Hilbert space, for example, in .

In the case of intersecting intervals , the probabilities

are finite linear combinations of expressions of the form Equation 4.2. Therefore, in order to show the convergence in distribution of point processes with determinantal correlation functions, it suffices to show the convergence of expressions of the form Equation 4.2.

The formula Equation 4.2 is discussed, for example, in Reference 37. See also Theorem 2 in Reference 35. It remains valid for processes on a lattice such as in which case the kernel should be an operator in .

As verified, for example, in Proposition A.11 in the Appendix, the right-hand side of Equation 4.2 is continuous in each with respect to the trace norm. We shall show that after a suitable embedding of into the kernel converges to the Airy kernel as .

Namely, we shall consider a family of embeddings , indexed by a positive number , which are defined by

where is the characteristic function of the point and, similarly, the function on the right is the characteristic function of a segment of length . Observe that this embedding is isometric. Let denote the kernel on that is obtained from the kernel on using the embedding Equation 4.3. We shall establish the following

Proposition 4.3.

We have

in the trace norm for all uniformly on compact sets in .

This proposition immediately implies Theorem 5 as follows.

Proof of Theorem 5.

Consider the left-hand side of Equation 4.1 and choose for each a pair of functions such that

and as . Then, on the one hand, the probability in the left-hand side of Equation 4.1 lies between the corresponding probabilities for and . On the other hand, the probabilities for and can be expressed in the form Equation 4.2 for the kernel and by Proposition 4.3 and continuity of the Airy kernel they converge to the corresponding probabilities given by the Airy kernel as .

Now we get to the proofs of Propositions 4.1 and 4.3 which will require some computations. Recall that the Airy function can be expressed in terms of Bessel functions as follows:

(see Section 6.4 in Reference 43). Also recall that

(see, for example, the formula 7.23 (1) in Reference 43).

Lemma 4.4.

For any we have

moreover, the constant in is uniform in on compact subsets of .

Proof.

Assume first that . We denote

It will be convenient to use the following notation:

The formula 8.43(4) in Reference 43 reads

where . We have the following estimates as :

Substituting this into Equation 4.7, we obtain the claim Equation 4.6 for .

Assume now that . Denote

Introduce the notation

The formula 8.43 (5) in Reference 43 reads

where . Again we have the estimates as

These estimates after substituting into Equation 4.8 produce Equation 4.6 for .

Lemma 4.5.

There exist such that for any and we have

for all .

Proof.

First suppose that . Set . We shall use Equation 4.7 with . Provided is chosen small enough and is sufficiently large, will be close to and we will be able to use Taylor expansions. For we have

and, similarly,

Since the function is concave, we have

The constant here is strictly positive.

Since (see, for example, the formula 7.23 (1) in Reference 43) we obtain

where we used that . Finally, we note that

and this completes the proof of Equation 4.9.

The estimate Equation 4.10 follows directly from the formulas 8.5 (9), (4), (5) in Reference 43.

Lemma 4.6.

For any there exists such that for all and large enough

Proof.

From Equation 2.15 we have

Let us split the sum in Equation 4.11 into two parts,

that is, one sum for and the other for , and apply Lemma 4.5 to these two sums. Note that here corresponds to in Lemma 4.5; this produces factors of and does not affect the estimate.

Let the ’s stand for some positive constants not depending on . From Equation 4.9 we obtain the following estimate for the first sum:

where

Therefore,

We can choose so that .

For the second sum we use Equation 4.10 and obtain

Clearly, this is less than for .

Proof of Proposition 4.1.

As shown in Reference 10Reference 36, the Airy kernel has the following integral representation:

The formula Equation 4.11 implies that for any integer

Let us fix and pick according to Lemma 4.6. Since, by assumption, and lie in a compact set of , we can fix such that . Set

Then the inequalities

are satisfied for all in our compact set and Lemma 4.4 applies to the sum in Equation 4.13. We obtain

because the number of summands is and is bounded on subsets of which are bounded from below. Note that

is a Riemann integral sum for the integral

and it converges to this integral as . Since the absolute value of the second term in the right-hand side of Equation 4.13 does not exceed by the choice of , we get

as , and this estimate is uniform on compact sets. Now let and . By Equation 4.5 the integral Equation 4.12 converges uniformly in and on compact sets and we obtain the claim of the proposition.

Proof of Proposition 4.3.

It is clear that Proposition 4.1 implies the convergence of to in the weak operator topology. Therefore, by Proposition A.9, it remains to prove that as . We have

where the correction comes from the fact that may not be a number of the form , . By Equation 4.11 we have

Similarly,

Since we already established the uniform convergence of kernels on compact sets, it is enough to show that both Equation 4.14 and Equation 4.15 go to zero as and . For the Airy kernel this is clear from Equation 4.5. For the kernel it is equivalent to the following statement: for any there exists such that for all and large enough we have

We shall employ Lemma 4.5 for . Again, we split the sum in Equation 4.16 into two parts:

For the first sum Lemma 4.5 gives

where

and the ’s are some positive constants that do not depend on . Since we obtain

This can be made arbitrarily small by taking sufficiently large.

For the other part of the sum we have the estimate

which, evidently, goes to zero as .

4.3. Depoissonization and proof of Theorem 4

Fix some and denote by the distribution function of under the Plancherel measure ,

Also, set

This is the distribution function corresponding to the measure .

The measures can be obtained as distribution at time of a certain random growth process of a Young diagram; see e.g. Reference 42. This implies that

Also, by construction, is monotone in and similarly

We shall use these monotonicity properties together with the following lemma.

Lemma 4.7 (Johansson, Reference 16).

There exist constants and such that for any nonincreasing sequence ,

and its exponential generating function

we have for all the following inequalities:

This lemma implies that for all

Set

Theorem 5 asserts that

where is the corresponding distribution function for the Airy ensemble. Note that is continuous.

Denote . Then for

Hence, for any and all sufficiently large we have

for . By Equation 4.17 this implies that

From this and Equation 4.18 we obtain

From this and Equation 4.19 we conclude that

as . Since is arbitrary and is continuous we obtain

which is the statement of Theorem 4.

Appendix A. General properties of determinantal point processes

In this Appendix, we collect some necessary facts about determinantal point processes, their correlation functions, Fredholm determinants, and convergence of trace class operators.

Let be a countable set, let be the set of subsets of and denote by the set of finite subsets of . We call elements of configurations. Let be a kernel on , that is, a function on also viewed as a matrix of an operator in .

By a determinantal point process on we mean a probability measure on such that

Here the determinant in the numerator is the usual determinant of linear algebra, whereas the determinant in the denominator is, in general, a Fredholm determinant. Some sufficient conditions under which makes sense are described in the following subsection.

A.1. Fredholm determinants and determinantal processes

Let be a complex Hilbert space, the algebra of bounded operators in , and , the ideals of trace class and Hilbert–Schmidt operators, respectively.

Assume we are given a splitting . According to this splitting, write operators in block form, , where

The algebra is equipped with a natural -grading. Specifically, given , its even part and odd part are defined as follows:

Denote by the set of operators such that is in the trace class while is in the Hilbert–Schmidt class . It is readily seen that is an algebra. We endow it with the topology induced by the trace norm on the even part and the Hilbert–Schmidt norm on the odd part.

It is well known that the determinant makes sense if . It can be characterized as the only function which is continuous in with respect to the trace norm and which coincides with the usual determinant when is a finite–dimensional operator. See, e.g., Reference 33.

Proposition A.1.

The function admits a unique extension to , which is continuous in the topology of that algebra.

Proof.

For , set

As is well known (e.g., Reference 33),

is a continuous map from to . Next, evidently is a continuous function on . Consequently, Equation A.1 is well defined and is a continuous function on . When , Equation A.1 agrees with the conventional definition, because then

This concludes the proof.

Corollary A.2.

If is an ascending sequence of even projection operators in such that strongly, then

Proof.

Indeed, approximates in the topology of .

Corollary A.3.

If , then

Proof.

Indeed, this is true for finite–dimensional , and then we use the continuity argument.

In our particular case, the splitting of will come from a splitting of into two complementary subsets as follows:

An operator in will be viewed as an infinite matrix whose rows and columns are indexed by elements of . Given , we denote by the corresponding finite submatrix in .

Proposition A.4.

If , then

where summation is taken over all finite subsets including the empty set with the understanding that .

The exact meaning of the sum in the left-hand side is explained in the proof.

Proof.

Given a finite subset , we assign to it, in the natural way, a projection operator . Note that is even. By elementary linear algebra, we have

Assume becomes larger and larger, so that in the limit it covers the whole . Then the left-hand side tends to the left-hand side of Equation A.2. More precisely, this is evident if all the minors are nonnegative. In general, instead of proving that the sum in the left-hand side of Equation A.2 is absolutely convergent we simply define it as

On the other hand, the right-hand side tends to by Corollary A.2.

Remark A.5.

Suppose that , where is of Hilbert–Schmidt class. Then . It is readily seen that for all , and it is worth noting that unless . By Proposition A.4, we can define a probability measure on finite subsets of by

A.2. Correlation functions of determinantal processes

Given , let be the probability that a random configuration contains , that is,

We call the correlation functions. The fundamental fact about determinantal point processes is that their correlation functions again have a determinantal form.

Proposition A.6.

Let be as above and set . Then .

Proof.

We follow the argument in Reference 11, Exercise 5.4.7. Let be an arbitrary function on such that for all but a finite number of ’s. Form the probability generating functional:

Then, viewing as a diagonal matrix, we get

where the last equality is justified by Proposition A.4 applied to the operator .

Now, set , so that for all but finitely many ’s. Then we can rewrite this relation as

where the last equality follows by Corollary A.3.

Next, as is in (it is even finite–dimensional), this can be rewritten as

On the other hand, by the very definition of ,

This implies , as desired.

Remark A.7.

If , then

In the recent survey Reference 35, the determinantal formula for the correlation functions is taken as a definition. The paper Reference 35 contains a more general and detailed discussion of the basics of the theory of determinantal processes which in Reference 35 are called determinantal random point fields.

A.3. Complementation principle

In this section we discuss a simple but useful observation which was communicated to us by S. Kerov. Consider an arbitrary probability measure on such that its correlation functions

have a determinantal form

for some kernel .

Let be an arbitrary subset of . Consider the symmetric difference mapping

which is an involution in . Let be the image of our probability measure under and let be the correlation functions of the measure . Define a new kernel as follows. Let be the complement of and write the matrix in block form with respect to the decomposition :

By definition, set

We have the following

Proposition A.8.

 .

Proof.

Set , . By the inclusion-exclusion principle we have

This alternating sum is easily seen to be identical by linearity to the expansion of using

A.4. Convergence of trace class operators

Let and be Hermitian nonnegative operators in . The following proposition is a special case of Theorem 2.20 in the book Reference 34 (we are grateful to P. Deift for this reference). For the reader’s convenience we give a proof here.

Proposition A.9.

The following conditions are equivalent:

(i) ;

(ii) and in the weak operator topology.

First, we prove a lemma:

Lemma A.10.

Let be a nonnegative operator matrix. Then .

Proof of Lemma A.10.

Without loss of generality one can assume that the block is a nonnegative diagonal matrix, . Write the blocks and as matrices, too, and let and be their diagonal entries. Since , we have and therefore

Proof of Proposition A.9.

Clearly, (i) implies (ii). To check the converse claim, write in block form,

where is of finite size and is small. Write all the ’s in block form with respect to the same decomposition of the Hilbert space,

Since weakly, we have convergence of finite blocks, , which implies . Since , we get , so that all the traces are small together with provided that is large enough.

Write

and similarly for . Then

In the right-hand side, the first and the third summands are small because of the lemma, while the second summand is small because it is equal to .

Proposition A.11.

The map defines a continuous map from to the algebra of entire functions in variables with the topology of uniform convergence on compact sets.

Proof.

The fact that is holomorphic in for any trace class operators is proved in Reference 33. The continuity of the map follows from the inequality

which holds for any ; see Reference 30Reference 34.

Acknowledgments

In many different ways, our work was inspired by the work of J. Baik, P. Deift, and K. Johansson, on the one hand, and by the work of A. Vershik and S. Kerov, on the other. It is our great pleasure to thank them for this inspiration and for many fruitful discussions.

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. 1.1. Plancherel measures
    2. 1.2. Poissonization and correlation functions
    3. Definition 1.1.
    4. Theorem 1.
    5. Theorem 2.
    6. 1.3. Asymptotics in the bulk of the spectrum
    7. Theorem 3.
    8. 1.4. Behavior near the edge of the spectrum and the Airy ensemble
    9. Theorem 4.
    10. 1.5. Poissonization and depoissonization
    11. 1.6. Organization of the paper
  3. 2. Correlation functions of the measures
    1. 2.1. Proof of Theorem 1
    2. Proposition 2.1.
    3. Corollary 2.2.
    4. Proposition 2.3.
    5. Lemma 2.4.
    6. 2.2. Proof of Theorem 2
    7. Lemma 2.5.
    8. 2.3. Various formulas for the kernel
    9. Proposition 2.6.
    10. Proposition 2.7.
    11. Proposition 2.9.
    12. Corollary 2.10.
    13. 2.4. Commuting difference operator
    14. Proposition 2.12.
  4. 3. Correlation functions in the bulk of the spectrum
    1. 3.1. Proof of Theorem 3
    2. Lemma 3.1.
    3. Definition 3.2.
    4. Proposition 3.4.
    5. Lemma 3.5.
    6. 3.2. Asymptotics of
  5. 4. Edge of the spectrum: Convergence to the Airy ensemble
    1. 4.1. Results and strategy of proof
    2. Proposition 4.1.
    3. Theorem 5.
    4. 4.2. Proof of Theorem 5
    5. Proposition 4.3.
    6. Lemma 4.4.
    7. Lemma 4.5.
    8. Lemma 4.6.
    9. 4.3. Depoissonization and proof of Theorem 4
    10. Lemma 4.7 (Johansson, 16).
  6. Appendix A. General properties of determinantal point processes
    1. A.1. Fredholm determinants and determinantal processes
    2. Proposition A.1.
    3. Corollary A.2.
    4. Corollary A.3.
    5. Proposition A.4.
    6. A.2. Correlation functions of determinantal processes
    7. Proposition A.6.
    8. A.3. Complementation principle
    9. Proposition A.8.
    10. A.4. Convergence of trace class operators
    11. Proposition A.9.
    12. Lemma A.10.
    13. Proposition A.11.
  7. Acknowledgments

Figures

Figure 1.

The limit shape of a typical diagram.

Graphic without alt text

Mathematical Fragments

Equation (1.1)
Equation (1.2)
Definition 1.1.

The correlation functions of are the probabilities that the sets or, similarly, contain a fixed subset . More precisely, we set

Theorem 1.

For any we have

where the kernel is given by the following formula:

The functions are defined by

where is the Bessel function of order and argument . The diagonal values are determined by the l’Hospital rule.

Theorem 2.

For any we have

Here the kernel is given by the following formula:

where . The diagonal values are determined by the l’Hospital rule.

Remark 1.2.

Theorem 1 is a limit case of Theorem 3.3 of Reference 8. For the reader’s convenience a direct proof of it is given in Section 2. Various limit cases of the results of Reference 8 are discussed in Reference 9. By different methods, the formula Equation 1.8 was obtained by K. Johansson Reference 17.

A representation–theoretic proof of a more general formula than Theorem 3.3 of Reference 8 has been subsequently given in Reference 27Reference 26; see also Reference 7.

Equations (1.10), (1.11)
Theorem 3.

Let be a regular sequence and let the numbers , be defined by Equation 1.10, Equation 1.11. If splits, that is, if and the distance between and goes to as , then

If is nonsplit, then

where is the discrete sine kernel and  .

Remark 1.7.

Note that, in particular, Theorem 3 implies that the limit density (the 1-point correlation function) is given by

This is in agreement with the Logan-Shepp-Vershik-Kerov result about the limit shape . More concretely, the function is related to the density 1.14 by

which can be interpreted as follows. Approximately, we have

Set . Then the above relation reads and it should be satisfied on the boundary of the limit shape. Since , we conclude that

as was to be shown.

Equation (1.15)
Equation (1.16)
Theorem 4.

As , the random variables converge, in joint distribution, to the Airy ensemble.

Equation (1.17)
Equation (1.18)
Equation (1.19)
Equation (2.1)
Equation (2.2)
Proposition 2.1.

Let be a partition. Then

where are the modified Frobenius coordinates of defined in Equation 1.2.

Proposition 2.3.

.

Equation (2.4)
Lemma 2.4.

For any and any we have

Equation (2.7)
Equation (2.8)
Equation (2.9)
Equation (2.10)
Lemma 2.5.

.

Equation (2.12)
Equation (2.13)
Proposition 2.6.
Proposition 2.7.

Suppose . Then

Proposition 2.9.
Equation (2.16)
Equation (2.17)
Equation (2.18)
Lemma 3.1.

Let be a sequence of entire functions

and suppose that there exist constants and such that

as . Then

Proposition 3.4.

If is regular, then

Lemma 3.5.

Suppose that a sequence is as above and, additionally, suppose that , are bounded and . Then the sequence satisfies Equation 3.2 with and certain .

Equation (3.3)
Equation (3.4)
Equation (3.5)
Equation (3.7)
Equation (3.8)
Equations (3.9), (3.10)
Proposition 4.1.

Set . We have

uniformly in and on compact sets of .

Theorem 5.

For any fixed and any we have

where is the Airy ensemble.

Equation (4.2)
Equation (4.3)
Proposition 4.3.

We have

in the trace norm for all uniformly on compact sets in .

Equation (4.5)
Lemma 4.4.

For any we have

moreover, the constant in is uniform in on compact subsets of .

Equation (4.7)
Equation (4.8)
Lemma 4.5.

There exist such that for any and we have

for all .

Lemma 4.6.

For any there exists such that for all and large enough

Equation (4.11)
Equation (4.12)
Equation (4.13)
Equation (4.14)
Equation (4.15)
Equation (4.16)
Equation (4.17)
Lemma 4.7 (Johansson, Reference 16).

There exist constants and such that for any nonincreasing sequence ,

and its exponential generating function

we have for all the following inequalities:

Equation (4.18)
Equation (4.19)
Proposition A.1.

The function admits a unique extension to , which is continuous in the topology of that algebra.

Equation (A.1)
Corollary A.2.

If is an ascending sequence of even projection operators in such that strongly, then

Corollary A.3.

If , then

Proposition A.4.

If , then

where summation is taken over all finite subsets including the empty set with the understanding that .

Proposition A.6.

Let be as above and set . Then .

Proposition A.8.

 .

Proposition A.9.

The following conditions are equivalent:

(i) ;

(ii) and in the weak operator topology.

Lemma A.10.

Let be a nonnegative operator matrix. Then .

Proposition A.11.

The map defines a continuous map from to the algebra of entire functions in variables with the topology of uniform convergence on compact sets.

References

Reference [1]
D. Aldous and P. Diaconis, Hammersley’s interacting particle process and longest increasing subsequences, Prob. Theory and Rel. Fields, 103, 1995, 199–213. MR 96k:60017
Reference [2]
—, Longest increasing subsequences: From patience sorting to the Baik-Deift-Johansson theorem, Bull. Amer. Math. Soc., 36, 1999, 413–432. CMP 99:17
Reference [3]
J. Baik, P. Deift, K. Johansson, On the distribution of the length of the longest increasing subsequence of random permutations, Journal of AMS, 12, 1999, 1119–1178. CMP 99:15
Reference [4]
—, On the distribution of the length of the second row of a Young diagram under Plancherel measure, math.CO/9901118.
[5]
P. Biane, Permutation model for semi-circular systems and quantum random walks, Pacific J. Math., 171, no. 2, 1995, 373–387. MR 96m:46117
[6]
—, Representations of symmetric groups and free probability, Adv. Math. 138, 1998, no. 1, 126–181. CMP 99:01
Reference [7]
A. Borodin, Riemann-Hilbert problem and the discrete Bessel kernel, math.CO/9912093, to appear in Intern. Math. Res. Notices.
Reference [8]
A. Borodin and G. Olshanski, Distribution on partitions, point processes, and the hypergeometric kernel, math.RT/9904010, to appear in Comm. Math. Phys.
Reference [9]
—, Z-measures on partitions, Robinson-Schensted-Knuth correspondence, and random matrix ensembles, math.CO/9905189, to appear in Proceedings of the 1999 MSRI Workshop on Random Matrices and their Applications.
Reference [10]
P. A. Clarkson and J. B. McLeod, A connection formula for the second Painlevé transcendent, Arch. Rat. Mech. Anal., 103, 1988, 97–138. MR 89e:34010
Reference [11]
D. J. Daley and D. Vere-Jones, An introduction to the theory of point processes, Springer series in statistics, Springer Verlag, 1988. MR 90e:60060
Reference [12]
P. J. Forrester, The spectrum edge of random matrix ensembles, Nuclear Phys. B, 402, 1993, no. 3, 709–728. MR 94h:82031
Reference [13]
—, Random walks and random permutations, math.CO/9907037.
Reference [14]
Higher Transcendental functions, Bateman Manuscript Project, McGraw-Hill, New York, 1953. MR 15:419i
Reference [15]
P. Jacquet and W. Szpankowski, Analytical depoissonization and its applications, Theor. Computer Sc., 201, 1998, 1–62. MR 99f:68099
Reference [16]
K. Johansson, The longest increasing subsequence in a random permutation and a unitary random matrix model, Math. Res. Letters, 5, 1998, 63–82. MR 99e:60033
Reference [17]
—, Discrete orthogonal polynomials and the Plancherel measure, math.CO/9906120.
[18]
S. Kerov, Gaussian limit for the Plancherel measure of the symmetric group, C. R. Acad. Sci. Paris, 316, Série I, 1993, 303–308. MR 93k:20106
Reference [19]
—, Transition probabilities of continual Young diagrams and the Markov moment problem, Func. Anal. Appl., 27, 1993, 104–117. MR 95g:82045
Reference [20]
—, The asymptotics of interlacing roots of orthogonal polynomials, St. Petersburg Math. J., 5, 1994, 925–941. MR 96a:33010
Reference [21]
—, A differential model of growth of Young diagrams, Proceedings of the St. Petersburg Math. Soc., 4, 1996, 167–194. CMP 2000:07
Reference [22]
—, Interlacing measures, Amer. Math. Soc. Transl., 181, Series 2, 1998, 35–83. MR 99h:30034
Reference [23]
B. F. Logan and L. A. Shepp, A variational problem for random Young tableaux, Adv. Math., 26, 1977, 206–222. MR 98e:05108
Reference [24]
I. G. Macdonald, Symmetric functions and Hall polynomials, Oxford University Press, 1995. MR 96h:05207
Reference [25]
A. Okounkov, Random matrices and random permutations, math.CO/9903176
Reference [26]
—, Infinite wedge and measures on partitions, math.RT/9907127
Reference [27]
—, and -measures, math.RT/0002135, to appear in Proceedings of the 1999 MSRI Workshop on Random Matrices and their Applications.
Reference [28]
E. M. Rains, Increasing subsequences and the classical groups, Electr. J. of Combinatorics, 5(1), 1998. MR 98k:05146
Reference [29]
A. Regev, Asymptotic values for degrees associated with strips of Young diagrams, Adv. Math., 41, 1981, 115-136. MR 82h:20015
Reference [30]
E. Seiler and B. Simon, On finite mass renormalization in the two-dimensional Yukawa model, J. of Math. Phys., 16, 1975, 2289–2293. MR 53:7295
Reference [31]
C. Schensted, Longest increasing and decreasing subsequences, Canad. J. Math., 13, 1961, 179-191. MR 22:12047
Reference [32]
T. Seppäläinen, A microscopic model for Burgers equation and longest increasing subsequences, Electron. J. Prob., 1, no. 5, 1996. MR 97d:60162
Reference [33]
B. Simon, Notes on Infinite Determinants of Hilbert Space Operators, Adv. Math., 24, 1977, 244–273. MR 58:2401
Reference [34]
—, Trace ideals and their applications, London Math. Soc. Lecture Note Ser., 35, Cambridge University Press, 1979, 134 pp. MR 80k:47048
Reference [35]
A. Soshnikov, Determinantal random point fields, math.PR/0002099.
Reference [36]
C. A. Tracy and H. Widom, Level-spacing distributions and the Airy kernel, Commun. Math. Phys., 159, 1994, 151–174. MR 95e:82003
Reference [37]
—, Introduction to random matrices, Geometric and quantum aspects of integrable systems, Lecture Notes in Phys., 424, 1993, Springer Verlag, pp. 103–130. MR 95a:82050
[38]
—, On the distribution of the lengths of the longest monotone subsequences in random words, math.CO/9904042.
Reference [39]
A. Vershik, Statistical mechanics of combinatorial partitions and their limit configurations, Func. Anal. Appl., 30, no. 2, 1996, 90–105. MR 99d:82008
Reference [40]
A. Vershik and S. Kerov, Asymptotics of the Plancherel measure of the symmetric group and the limit form of Young tableaux, Soviet Math. Dokl., 18, 1977, 527–531.
Reference [41]
—, Asymptotic theory of the characters of a symmetric group, Func. Anal. Appl., 15, 1981, no. 4, 246-255. MR 84a:22016
Reference [42]
—, Asymptotics of the maximal and typical dimension of irreducible representations of symmetric group, Func. Anal. Appl., 19, 1985, no.1. MR 86k:11051
Reference [43]
G. N. Watson, A treatise on the theory of Bessel functions, Cambridge University Press, 1944. MR 6:64a
Reference [44]
H. Widom, Random Hermitian matrices and (nonrandom) Toeplitz matrices, Toeplitz Operators and Related Topics, the Harold Widom anniversary volume, edited by E. L. Basor and I. Gohberg, Operator Theory: Advances and Application, Vol. 71, Birkhäuser Verlag, 1994, pp. 9–15. MR 95h:15036
Reference [45]
—, The strong Szegö limit theorem for circular arcs, Indiana U. Math. J., 21, 1971, 271–283. MR 44:5693

Article Information

MSC 1991
Primary: 05E10 (Tableaux, representations of the symmetric group), 60C05 (Combinatorial probability)
Author Information
Alexei Borodin
Department of Mathematics, University of Pennsylvania, Philadelphia, Pennsylvania 19104–6395 and Dobrushin Mathematics Laboratory, Institute for Problems of Information Transmission, Bolshoy Karetny 19, 101447, Moscow, Russia
borodine@math.upenn.edu
Andrei Okounkov
University of Chicago, Department of Mathematics, 5734 University Ave., Chicago, Illinois 60637
Address at time of publication: Department of Mathematics, University of California at Berkeley, Evans Hall, Berkeley, California 94720-3840
okounkov@math.berkeley.edu
ORCID
MathSciNet
Grigori Olshanski
Dobrushin Mathematics Laboratory, Institute for Problems of Information Transmission, Bolshoy Karetny 19, 101447, Moscow, Russia
olsh@glasnet.ru
MathSciNet
Additional Notes

The second author is supported by NSF grant DMS-9801466, and the third author is supported by the Russian Foundation for Basic Research under grant 98-01-00303.

Journal Information
Journal of the American Mathematical Society, Volume 13, Issue 3, ISSN 1088-6834, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2000 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/S0894-0347-00-00337-4
  • MathSciNet Review: 1758751
  • Show rawAMSref \bib{1758751}{article}{ author={Borodin, Alexei}, author={Okounkov, Andrei}, author={Olshanski, Grigori}, title={Asymptotics of Plancherel measures for symmetric groups}, journal={J. Amer. Math. Soc.}, volume={13}, number={3}, date={2000-07}, pages={481-515}, issn={0894-0347}, review={1758751}, doi={10.1090/S0894-0347-00-00337-4}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.