A convergence framework for Airyβ line ensemble
via pole evolution
Jiaoyang Huang
Department of Statistics and Data Science, University of Pennsylvania, Pennsylvania, PA, USA
[email protected] and Lingfu Zhang
Division of Physics, Mathematics and Astronomy, California Institute of Technology, Pasadena, CA, USA. and Department of Statistics, University of California, Berkeley, CA, USA.
[email protected]
Abstract.
The Airyβ line ensemble is an infinite sequence of random curves. It is a natural extension of the Tracy-Widomβ distributions, and is expected to be the universal edge scaling limit of a range of models in random matrix theory and statistical mechanics. In this work, we provide a framework of proving convergence to the Airyβ line ensemble, via a characterization through the pole evolution of meromorphic functions satisfying certain stochastic differential equations. Our framework is then applied to prove the universality of the Airyβ line ensemble as the edge limit of various continuous time processes, including Dyson Brownian motions with general and potentials, Laguerre processes and Jacobi processes.
During the 18th century, De Moivre established the Gaussian distribution for sums of independent binomial variables, a concept later generalized by Laplace. Gauss popularized the central limit theorem, used for error evaluation in systems characterized by independence.
In recent decades, there has been increasing interest in understanding fluctuations in highly correlated systems, leading to the emergence of a different family of distributions known as Tracy-Widomβ, which are indexed by a positive parameter .
Historically, such Tracy-Widomβ distributions for were first observed in Random Matrix Theory, as the scaling limit of the extreme eigenvalues of the classical matrix ensembles [121, 122, 58, 82, 84, 86].
Later, such extreme eigenvalue statistics are proven to be universal, in the sense that Tracy-Widomβ limits hold for a wide range of random matrix models, including adjacency matrices of random graphs, which are usually sparse. See e.g., [118, 120, 98, 54, 90, 78, 71, 72, 79, 16].
Beyond matrix models, Tracy-Widomβ distributions also appear in lots of different random systems, such as random tilings, 1+1 dimensional random growth models, exclusion processes, planar random geometry such as first/last passage percolation models, and square ice models (six-vertex models).
See e.g., [11, 82, 8, 111, 84, 83, 104, 5, 12, 115, 21, 30, 127, 70, 80, 13, 14, 10, 28, 4, 126].
Many of these models are related to each other through their underlying integrable structures, and are in the so called Kardar-Parisi-Zhang (KPZ) universality class, a topic in probability theory that has been intensively studied in recent years.
A motivation of the current paper is to better understand the mathematical structures behind Tracy-Widomβ distributions, and to develop new methods of proving convergence to them.
The main object studied here is the Airyβ line ensemble (ALEβ) , which can be defined as a random process on or a family of continuous random processes on , with being a parameter, and is ordered, i.e., for any .
They are natural generalizations of the Tracy-Widomβ distributions, akin to Brownian motions being generaliztions of Gaussian distributions, and have Tracy-Widomβ as the one-point distribution of the top line .
These ALEβ are believed to be universal objects, in the sense of being the scaling limit of many random matrix models and interacting particle systems. However, basic properties of these processes as well as these convergences remain quite mysterious so far, except for the special setting of , where a determinantal structure is present and has been largely exploited using algebraic methods.
In this paper, we take a new perspective, and our main result is a characterization of ALEβ in terms of its Stieltjes transform and a system of stochastic differential equations (SDE).
Our result provides a new framework to prove convergence. As some examples, we prove convergence to ALEβ from various random processes, including the classical -Hermite/Laguerre/Jacobi processes and their generalizations. We note that some of these were previously unknown even in the cases, where such convergences can be interpreted as the joint convergence of extreme eigenvalues of correlated real/quaternion random matrices.
Beyond these, our framework should also be applicable to prove convergence to ALEβ for many other models; and for some of them, even the Tracy-Widomβ limits were previously unknown.
Moreover, our characterization also reveals some useful information for ALEβ, such as Hölder properties and collision of adjacent lines.
1.1. Background
We next provide some setup, starting with some more classical processes.
1.1.1. Edge limit of general -ensembles
Tracy-Widomβ distributions for general was introduced and studied in [53, 119] by Edelman and Sutton, and [113] by Ramirez, Rider and Virág, as the scaling limit of the extreme eigenvalue of Gaussian -ensembles.
More generally, -ensemble is a probability distribution on particle system , with probability density:
(1.1)
where is a renormalization constant, and is the weight function.
There are three special cases referred to as the classical ensembles, which are defined by
(1.5)
These classical ensembles for originated from the study of eigenvalue distributions of random matrices. They represent the joint distributions of the eigenvalues of size Gaussian, Wishart, and Jacobi matrices. These matrices and their extreme eigenvalues, with corresponding to the real case, have been widely used in high-dimensional statistical inference (see the survey by Johnstone [87]). Beyond random matrix theory, -ensembles also describe the one-dimensional Coulomb gas in physics [58], and are connected to orthogonal polynomial systems [91].
As mentioned earlier, as , the distribution of the largest eigenvalues converges to the Tracy-Widomβ distribution for , respectively. More generally, one can consider the edge limit, i.e., the joint scaling limit of the top eigenvalues for any arbitrary , as a point process. It has been shown [41, 42, 43] that this edge limit does not depend on the potential function , but it varies for each of . The edge limit is also known as the Airy point process.
For other than , obtaining such edge limit was a challenging problem, partly due to the relative lack of exact-solvable structures.
Based on a tri-diagonal random matrix model discovered by Dumitriu and Edelman [49], this was resolved in [113], where the edge limit of Gaussian and Laguerre -ensemble is shown to be the eigenvalues of the stochastic Airy operator (SAOβ), which is also called the Airyβ point process. In particular, for each , the law of the largest eigenvalue of SAOβ, i.e., the top particle in the Airyβ point process, is then defined as the Tracy-Widomβ distribution.
Later, such Airyβ point process limit is also extended to more general
[95, 25, 18, 17]. Analogous edge limits for discrete -ensembles have been derived in [67].
1.1.2. Airy line ensemble
In another direction, the Tracy-Widom2 distribution and the () Airy point process are extended to the Airy line ensemble (ALE), an ordered family of random processes , where each is continuous, and they are jointly stationary in the direction (see Figure1).
ALE was introduced by Prahöfer and Spohn [109], as the scaling limit for the multi-layer polynuclear growth (PNG) model from the KPZ universality class.
The top line is known as the stationary Airy2 process, whose one-point marginal is the Tracy-Widom2 distribution; and for any , is the Airy point process.
ALE plays a central role in KPZ, in particular through the construction of the directed landscape [37]. (See also [38] for computing passage times in the directed landscape from ALE.)
ALE is particularly useful in KPZ, partly due to its Brownian Gibbs property, which was recognized by Corwin and Hammond [35]. Specifically, for ALE minus a parabola, it inside any domain, conditional on the boundary, is given by non-intersecting Brownian bridges. This fact is later widely used as a powerful tool to study ALE and many KPZ class models. Aggarwal and the first-named author provided a strong characterization of ALE, demonstrating that ALE (minus a parabola) is the only random process on with the Brownian Gibbs property as well as approximating a parabola [6].
Such a strong characterization would be a powerful tool to prove convergence to ALE and establishing KPZ universality for various models; see e.g. [4].
1.2. Airyβ line ensemble
From the success of ALE, the next question is to construct a time dependent evolution for Tracy-Widomβ (and more generally Airyβ point process) for any .
Following [64] where this is formally introduced, we call it the Airyβ line ensemble (ALEβ).
There are several problems in this program:
•
Construction What should it be? How to construct it?
•
Description What are its properties? Ideally, can some precise information be given?
•
Universality Why is ALEβ natural and interesting? Can it be shown to be the universal scaling limit of many natural random processes, as the case?
Towards these questions, there have been many results focusing on different aspects of ALEβ (some tracing back to the studies of Airyβ point process, or for special ): infinite-dimensional SDE [97, 106, 107, 88], correlation function [104, 11, 85, 109], and Laplace transform [118, 102, 103, 117, 81, 64].
In this paper, we provide a different perspective using Stieltjes transform, tailored to the universality problem.
Moreover, we view results presented here and in the concurrent paper [64] (by Gorin, Xu, and the second-named author, to be discussed shortly) complement to each other.
We now give a more detailed account on the state of the art, and further motivate our results.
1.2.1. The edge limit construction and convergence
Beyond ALE where , ALEβ was also constructed and studied for several other special values: in [117] for (and the arguments there should also go through for ), and in [60] for .
For general , even its construction is relatively recent.
One potential way, as inspired by the fact (from [35]) that ALE is the edge scaling limit of the Dyson Brownian motion (DBM), is to consider general DBM:
(1.6)
where and are independent two-sided Brownian motions.
There is a solution to this SDE, such that for any is a Gaussian -ensemble; and this is known as the (stationary) DBM of size with parameter . (See e.g., [9] for some more backgrounds on DBM.)
One can then define ALEβ as its edge limit, i.e., the limit of as .
Note that for any fixed , should be the Airyβ point process, and should follow the Tracy-Widomβ distribution.
Beyond DBM, another potential way of constructing ALEβ is via the edge limit of Gaussian corners processes, which are random Gelfand-Tsetlin patterns, and can be viewed as multi-level generalizations of Gaussian ensembles. For , they can be defined as eigenvalues of the top-left corners of different sizes, for Wigner matrices with Gaussian real, or complex, or quaternionic entries (see e.g. [59, 15, 101]).
To justify the above definition of ALEβ, the above stated convergence should be proved.
In [97], this is achieved for stationary DBM and any via a coupling argument.
In [64] the convergences for both stationary DBM and Gaussian corners process are established, for any .
The proofs are via explicit formulas, also showing that both limits are the same.
Very recently, in [52] a tridiagonal model for DBM is proposed, which may also be used to demonstrate convergence to ALEβ. Additionally, the arguments in this paper provide an alternative proof of DBM convergence.
1.2.2. Description via explicit expressions
Historically, the theory of Tracy-Widomβ distributions used to largely rely on explicit formulas, based on determinantal/Pfaffian structures or matrix models (see e.g., [121, 123, 116, 118, 113, 63]).
As for ALEβ, formulas used to be available only for , i.e., ALE, in the construction by Prahöfer and Spohn [109]; and for in [60].
There are various challenges in obtaining precise formulas for general , primarily due to the lack of structures in this generality.
As above mentioned, the first construction of ALEβ for in [97] was via the edge limit of DBM, using a more abstract convergence argument.
Then in [64], precise formulas for the Laplace transform of ALEβ are obtained.
Such Laplace transforms also determine the law of ALEβ, thereby [64] gives a direct and explicit definition of ALEβ for any .
1.2.3. Uniqueness and universality
As indicated by the convergence of both DBM and Gaussian corners process to the same limit, i.e., ALEβ, in [64], it is natural to expect that ALEβ is also the scaling limit of many other well-known processes.
Some examples include DBM with general potentials, non-intersecting random walks [61, 93, 74], various other –corners processes [62, 22, 65] and measures on Gelfand-Tsetlin patterns [29, 108], Macdonald processes [20], and -distributions on lozenge tilings [23, 46].
A robust and general approach to establishing convergence involves a suitable characterization of ALEβ. Specifically, this means identifying easily verifiable properties of ALEβ and demonstrating that these properties uniquely determine ALEβ. To prove convergence, it would then suffice to establish tightness and confirm that any subsequential limit satisfies these properties.
For ALEβ with , an elegant characterization is the Brownian Gibbs property [6] as mentioned above.
However, this does not hold for any .
The next natural candidate of characterization would be an ‘infinite dimensional DBM’, by taking in (1.6).
For example, [97] shows that ALEβ (for ), i.e., the edge limit of finite dimension DBM, is a solution to such an infinite dimensional DBM in a weak sense.
However, there are several technical challenges in developing a characterization and convergence framework along this direction.
First, particles (i.e., those in (1.6)) may collide or adjacent particles may get too close, leading to singularities in the drift term.
In [97], for such singularities are ruled out using existing level replusion estimates (see [97, Theorem 2.2] and [25, Theorem 3.2]), which are known only for stationary DBM whose laws are given by -ensembles. Deriving such estimates for other models could be difficult. Moreover, for collisions are inevitable.
Second, the long-range interactions introduce additional complications when analyzing infinitely many particles. In fact, even the well-posedness of the infinite-dimensional DBM starting from a fixed intial condition appears to be absent from the literature, except for the cases [106, 107, 88] where the underlying algebraic structure is used111Note that for an analogous infinite-dimensional SDE corresponding to the bulk limit of DBM, such well-posedness has been achieved for (see [89, 105, 106, 124]). A key property used in the bulk setting is the cancellation of particle interactions from left and right, and that is absent at the edge..
As a result, for the infinite-dimensional DBM, both proving the uniqueness of solution and verifying it for any subsequential limit face various barriers.
To overcome these difficulties, in this paper we take an alternative approach, and characterize ALEβ as the pole dynamics of meromorphic functions, satisfying a funciton-valued SDE.
In other words, we characterize ALEβ via the dynamics of its Stieltjes transform.
This method completely circumvent the issue of long-range interactions and collisions, and is applicable for any .
1.3. Main characterization result
In the rest of this paper, we fix .
We study (infinite) line ensembles, by which we mean ordered families of continuous random processes, denoted by , for or any interval, satisfying .
As already alluded to, we shall characterize ALEβ using its Stieltjes transform.
If were ALEβ, it would be imperative that (for each ) the Stieltjes transform of (with complex variable ) asymptotically behaves like 222Here and throughout this paper, , or any rational power of , is taken to be the branch on as in the complex plane.
This is because we expect , as a one time slice of ALEβ, to be the Airyβ point process; hence, the particles should be close to the zeros of the Airy function . Therefore, the Stieltjes transform of should exhibit similar asymptotic behaviors as , which is known to behave like as .
We next give a more precise formulation of such asymptotic behaviors.
Below we use to denote the open upper half complex plane.
For Stieltjes transforms, we shall use the notion of Nevanlinna functions, i.e., functions from to that are holomorphic.
Definition 1.1.
A measure on is particle-generated, if it is locally finite, and is in the form of , where (the particles) is an at most countable multiset of real numbers.
A Nevanlinna function has a Nevanlinna representation of the form
(1.7)
where , .
We say that is particle-generated, if the measure in its Nevanlinna representation is particle-generated, namely .
We note that such can be extended to a meromorphic function on with , with being all the poles, and each residue equals 333More precisely, for each , the residue at equals the multiplicity of in ..
Definition 1.2.
For , a Nevanlinna function is -Airy-like, or simply Airy-like, if
(1)
it is particle-generated, and all the poles are ;
(2)
for all with ,
(1.8)
Remark 1.3.
The condition (2) implies the existence of infinitely many poles .
As we will see shortly (2.3), bounds similar to (1) and (2) (but may with a different domain for ) would imply that the density of these poles would be close to in .
Such density closeness can be phrased as quantitative bounds on the distances between the poles and zeros of the Airy function , as stated in 2.3, and will be frequently used in our proofs.
Therefore, as a slight misuse of notions, we will refer to such density closeness as Airy-zero approximation.
As another comment on (2): again note that the domain of is different from that in 2.3, since the domain here is taken to be easily verifiable for the sub-sequential limit of various models, as will be evident in Section7.
In fact, bounds on for closer to can be readily deduced for Airy-like Nevanlinna functions (see 2.5).
Take a family of random Nevanlinna functions . We next state two assumptions.
Assumption 1.4.
For any , is particle-generated. Moreover,
there exists a (deterministic) , a sequence ,
and a tight family of random numbers , so that each is -Airy-like.
Assumption 1.5.
Such is continuous in , and satisfies the following SDE:
(1.9)
where are complex valued Martingales, with quadratic variation given by
(1.10)
and
(1.11)
for .
We now explain where this SDE comes from.
Take the dimensional stationary DBM with parameter , i.e., the stationary solution to (1.6). Let be the Stieltjes transform of , i.e.,
Then one can use Ito’s formula to write out an SDE satisfied by ; by taking an appropriate scaling limit from there, one gets the SDE (1.9). More details on this derivation can be found in Section7.
Our main result states that these two assumptions are sufficient to determine ALEβ uniquely.
Theorem 1.6.
For any satisfying 1.4 and 1.5, its poles give a line ensemble, which has the same law as ALEβ.
Several remarks are in line.
(i) Essentiality of 1.4 (in characterizing ALEβ). Both (1) and (2) in Definition1.2 are necessary:
without (1), the line ensemble may be ALEβ plus some additional lines (see [2, 45] for an example in the setting); while (2) rules out the possibility that the line ensemble is ALEβ shifted by a (deterministic or random) constant.
As already mentioned in Remark1.3, we’ve aimed to make 1.4 as minimal as possible to ensure broad applicability of our convergence framework. As will be seen in Section7, Assumption 1.4 is straightforward to verify in these examples.
(ii) DBM convergence. Our proof of 1.6 does not a priori assume the convergence at the edge of stationary DBM.
Instead, in Section7, we show the tightness at the edge of stationary DBM, and that any subsequential limit satisfies 1.4 and 1.5. Therefore, we essentially provide an alternative construction of ALEβ.
(iii) Stationarity. We also note that in 1.6, we do not assume that is stationary. Rather, it is a consequence of the theorem that the poles of converge to ALEβ, which is stationary, and hence is stationary as well. Our proof of 1.6 in fact establishes a natural relaxation for the SDE (1.9). Specifically, for a family of random particle-generated Nevanlinna functions , if there is a (deterministic) and a random number such that is -Airy-like, and satisfies 1.5, then for , the poles of converge to ALEβ, under the uniform in compact topology.
(iv) Stieltjes transform and poles dynamics.
Stieltjes transforms and Nevanlinna functions have been widely used to investigate and characterize eigenvalue distributions of random matrix ensembles. See [57] for studies on eigenvalue rigidity, [56, 7] for bulk limits, and [98, 90, 31] for edge limits.
The concept of characterizing the evolution of interacting particle systems through the pole dynamics of meromorphic functions has been explored previously. In integrable systems, it has been demonstrated that the movement of poles in certain solutions of various nonlinear PDEs can be formally linked to the dynamics of particle systems interacting through simple two-body potentials. This discovery was initially made in [34] for equations such as the Korteweg-de Vries and Burgers-Hopf equations, and in [99] for specific integrable Hamiltonian systems. Subsequently, these observations were extended to include elliptic solutions of equations such as the Kadomtsev-Petviashvili equation [94], the Korteweg–de Vries equation [40], the Kadomtsev-Petviashvili hierarchy [125], and the Toda lattice hierarchy [110].
Our results can be interpreted as a stochastic counterpart to these findings, wherein ALEβ is characterized as the pole evolution of the SDE (1.9).
1.4. Convergence framework
Given the characterization presented in 1.6, to prove convergence to ALEβ, it suffices to
(1)
establish the tightness of the Stieltjes transforms of the empirical particle density at the microscopic scale, and
(2)
verify that the scaling limit is Airy-like, and satisfies the SDE (1.9).
As a demonstration of this approach, we prove the convergence to ALEβ for several continuous interacting particle systems.
We next give the formal statement of our result.
We use a strong topology of uniform in compact convergence for line ensembles. More precisely, for a sequence of ordered families of functions , they converge to under the uniform in compact topology, if for each , uniformly in any compact interval.
Theorem 1.7.
ALEβ is the edge scaling limit of stationary DBM with certain general potentials (satisfying 7.1 below), stationary Laguerre process, and stationary Jacobi process, all with parameter , under the uniform in compact topology. We refer to 7.2 for a more detailed statement.
The definitions and background of these processes, as well as the precise statement and proof, will be given in Section7.
We emphasize that these convergence results are new even for the classical cases of (except for the DBM with ), which can be viewed as the joint convergence of eigenvalues of time-evolved classical ensembles with real or quaternion entries.
We remark that the developed framework can also be applied to prove convergence to ALEβ for the other mentioned models. The main remaining task is to establish the desired tightness given in (1).
While such tightness are not available from [26], a plausible way is to utilize the dynamical loop equation, as in [75] where local laws down to any mesoscopic scale have been proven for random tilings.
We leave this for future works.
1.5. Other properties
In addition to proving convergence to ALEβ, our new characterization can be leveraged to further investigate its properties.
First, we can study the regularity of ALEβ. The Brownian regularity for the ALE has been intensively studied in [35, 69, 68, 36]. For ALEβ with , it has been established in [97] that the lines of ALEβ are locally Brownian.
In Section4 we show that the lines of ALEβ are Hölder continuous with an exponent for any .
The second property we study is the collision of lines. For , it has been established in [97] that the lines of ALEβ do not collide. Conversely, for , collisions among lines are anticipated. We prove in Section5 that the occurrence of collisions is almost surely of measure zero.
1.6. Proof ideas
We give an outline of our proofs, highlighting the main difficulties and ideas.
To prove the uniqueness in law as stated in 1.6,
the overall strategy is to establish a certain sense of ‘mixing in time’ of the dynamics (given by 1.5).
More precisely, we take two families of random particle-generated Nevanlinna functions, both satisfying the two assumptions.
Using the SDE (1.9) we reconstruct the dynamics of the poles, which are ‘infinite dimensional DBM’ in a certain sense.
We couple the two ‘infinite dimensional DBM’ obtained from both functions, by coupling the driven Brownian motions. Then we show that the poles get closer in time under this coupling.
Thus since both dynamics start from time , necessarily they are the same.
For both the reconstruction of DBM and the coupling, an essential input is that the poles have Airy-zero approximation, uniformly in time.
This is implied by the uniform in time Airy-like property, as explained in Remark1.3.
Then under 1.4, it remains to show that such an approximation propagates in time, for which we again resort to the SDE (1.9).
In summary, three tasks are inline: propagation of Airy-zero approximation, reconstruction of DBM, and coupling.
We next explain each of them in more details.
1.6.1. Propagation of Airy-zero approximation
Our 1.4 concerns specific times, implying that the -th pole remains constant away from the -th zero of the Airy function. Utilizing the SDE (1.9), we manage to get refined estimates: the -th pole approximates the -th zero with a polynomially small error over arbitrarily long time intervals with high probability, as demanded for later steps. To achieve this, in Section3, we analyze (1.9) along certain characteristics which offset the singularity of the nonlinear term. This idea has previously been used (see e.g., [76, 1, 24]) to study DBM down to any mesoscopic scale, where the distance from the spectral parameter to the particles is much bigger than particle fluctuations.
However, in our analysis of (1.9), we operate at a microscopic scale, where the distance from to the poles is of the same order as their fluctuation size. While a straightforward union bound over characteristic flows from polynomially many points suffices at the mesoscopic scale, our case demands careful selection of characteristic flows and precise estimation of error terms in the SDE, tailored to their initial positions.
1.6.2. DBM reconstruction
As already mentioned, there are significant challenges in analyzing DBM due to the singular repulsive interaction and possible particle collisions, in particular when . Even for finite dimensional DBM, establishing the existence and uniqueness of a strong solution requires the theory of multivalued SDE, see [33, 32]. Our approach through pole evolutions circumvents these issues entirely. Notably, there are no singularities even when poles collide.
On the other hand, a key challenge of our method lies in reconstructing the dynamics of each pole, which requires ruling out the possibility of poles adhering to each other for prolonged periods. To address these, in Section4 we first establish that the trajectory of each pole is Hölder continuous solely utilizing (1.9). Together with the Airy-zero approximation, for any short time interval, we can identify a large index such that the -th and -th poles remain bounded away from each other. This enables us to localize the system and study the evolution of the first poles, treating the remaining poles’ influence as an additional potential.
For this poles system, in Section5,
by employing classical Itô calculus on certain elementary symmetric polynomials, we show that the time of collisions almost surely has measure zero. We note that similar ideas have been employed to show the instant
diffraction of the particles for DBM [66].
Subsequently, the evolution of each pole can be reconstructed by performing a contour integral of (1.9).
As a result, the poles system can be interpreted as a -dimensional DBM with a time-dependent random drift, which exhibits a monotonicity property.
1.6.3. Uniqueness via coupling
In Section6, we take two solutions to (1.9), and design a coupling where the poles get closer in time.
Consider the -dimensional DBMs with random drifts reconstructed in the previous step, for these two solutions respectively.
Our coupling is by using the same set of driven Brownian motions for both.
Note that such -dimensional DBMs with random drifts are constructed with random , and only for a short time interval; but we need a coupling for a long time (to let the poles get closer).
A trick here is to concatenate these short intervals, and allow for different in each of them, as long as is always large enough.
There is a monotonicity property: if the -th pole of the initial data for the first solution dominates that of the second solution for each , then at any time after the -th pole of the first solution dominates that of the second solution for each .
Then we can sandwich one solution between affine shifts of the other, while keeping the error arising from the affine shifts arbitrarily small. Such sandwiching forces the poles of the two solutions to get closer in time. By taking long enough time intervals, they must coincide, establishing the uniqueness as desired. Such coupling and sandwiching strategies have been used to establish local statistics universality in random lozenge tilings [3, 5, 77].
Notations
In the rest of this paper, for any , we let .
For any , we use to denote some , satisfying for some universal constant .
We also write for .
Acknowledgement
The research of J.H. is supported by NSF grant DMS-2331096 and DMS-2337795, and the Sloan research award.
The research of L.Z. is supported by NSF grant DMS-2246664 and partially by the Miller Institute for Basic Research in Science.
Part of this project was done when L.Z. was visiting University of Pennsylvania in the spring of 2023, and he thanks them for their hospitality.
The authors would like to thank Paul Bourgade, Vadim Gorin, Benjamin Landon, and Bálint Virág for helpful discussions.
2. Preliminaries and decomposition
In this section, we set up some preliminaries of our arguments.
We start with an explicit expression for any Airy-like Nevanlinna from Definition1.2.
The expression involves the Airy function, which is usually denoted by , and is a special function that appears in various areas of mathematics and physics.
It can be defined as an entire function, and the solution to the Airy equation: with as along .
All the zeros of are on the real line, and are all negative, and we denote them as .
We now give the expression.
Proposition 2.1.
For any particle-generated Nevanlinna function with infinitely many poles , if i) ; and ii) there exists a sequence of complex numbers along any direction in , such that , then
The Nevanlinna representation (1.7), for any particle-generated Nevanlinna function with poles (a multiset), can be written as
(2.2)
where , . We remark that it is possible that contains only finitely many numbers, and the summation in (2.2) is finite.
Then to prove 2.1, it remains to determine and in (2.2) for , and establish that
the sum converges.
To start with, we first collect some basic estimates on the Airy function and Nevanlinna functions, which will also be used repeatedly in the rest of this paper.
2.1. Airy function
The Airy function has the following asymptotic formula.
For ,
where .
In particular, there is
for any , where is a continuous function.
See e.g., [47, Subsection 9.7.iv] and [100, Appendix B].
It follows that
(2.3)
for any with large enough and .
The Weierstrass representation gives
(2.4)
It is also known that is around . More precisely, for any we have
(2.5)
2.2. Estimates on Nevanlinna functions
We now present some estimates on (particle-generated) Nevanlinna functions, which will be used in the Airy-like function part of 2.1. We note that some of them are also used repeatedly in subsequent sections.
For any particle-generated Nevanlinna function , from (2.2) we have
(2.6)
Lemma 2.2.
The quantity is monotone in ; the derivatives of satisfy
(2.7)
where is the -th derivative of , for any integer .
which is increasing in . Using (2.2), the derivative of satisfies
(2.9)
And by
the case follows.
∎
As already alluded to, if a particle-generated Nevanlinna function is close to the function , its poles would be close to the Airy function zeros.
More precisely, we have the following estimate.
Lemma 2.3.
Take any parameters and . Suppose that a particle-generated Nevanlinna function satisfies the following conditions:
•
there is no pole of in ;
•
for any with and , we have
Then has infinitely many poles , and for any , where is a universal constant.
In particular, these conditions are satisfied by Airy-like Nevanlinna functions (with large and ).
Corollary 2.4.
For any -Airy-like Nevanlinna function (with poles ), there exists depending only on , such that for each .
The proof of 2.3 relies on the Helffer-Sjöstrand formula, which has become standard in random matrix theory. Therefore, we defer it to AppendixA.
As we shall derive convergence to ALEβ from Stieltjes transforms, we will need several statements on the functional spaces, which we provide here.
Definition 2.6.
For any locally finite measures on , we say that they converge in the vague topology to another locally finite , if , for any that is compactly supported and smooth.
Such vague topology arises naturally from Nevanlinna function convergence.
Lemma 2.7.
Take Nevanlinna functions and such that as , uniformly in any compact subset of .
Suppose the corresponding measures (in their Nevanlinna representation) are and , respectively,
then in the vague topology.
Proof.
Take any that is compactly supported and smooth, and let be a large enough number such that outside .
Take a smooth function , such that on , and on .
By A.1, we have
We note that whenever .
Also, whenever ; and for , from Nevalinna representation we have that .
Therefore, we have that the integrand in the above integral is non-zero only in ; and it is bounded by a constant, which is independent of by the uniform convergence of in .
Thus we can apply dominated convergence theorem to deduce that the above integral converges to
So the conclusion follows.
∎
On the other hand, in the setting of particle-generated measures, vague topology convergence can imply pole convergence.
Lemma 2.8.
For a sequence of particle-generated measures , such that as in the vague topology, the limit must also be particle-generated.
Moreover, if there is some such that for each , then the followings are true.
We denote by the -th largest pole of (with the convention that if there are less than poles).
For each , either as , or exists and is a pole of .
Also, all the poles of are given by such limits.
Proof.
By vague topology convergence, for any , we have that , and .
Then for any with , we must have for large enough, since each must be an integer.
Therefore, for large enough, and .
These imply that in any open interval is either or zero.
Therefore, in any compact interval, is supported on finitely many points.
Now take any where . Take any such that .
Then for large enough, ; so .
These imply that is particle-generated.
Under the additional assumption, there is . So we can write the poles of as (with the convention that if there are less than poles).
Then we can show via induction in , using that for each with , for any large enough.
∎
3. Pole evolution: uniform rigidity in time
In this section, we prove a uniform in time estimate for the poles. More precisely, the following proposition states that for the SDE (1.9),
with high probability, its pole evolution gives a line ensemble (i.e., all the poles are bounded from above, and the trajectories are continuous), and the poles are close to the zeros of the Airy function, uniformly in time.
Proposition 3.1.
For any , there exist small and large , such that the following holds.
Take any particle-generated satisfying 1.5, and large . Conditional on the event that is -Airy-like,
with probability at least , we have
(1)
The poles of give a line ensemble , and for each and ,
Our general strategy is to obtain uniform in time estimates for , and to apply 2.3.
The main tasks are (1) to estimate bulk pole densities, via bounding for in a reasonable domain contained in (in particular, allowing for polynomially close to the real axis, when ); (2) to bound the first pole .
Both these are to be achieved through analyzing the SDE (1.9).
3.1. Characteristic flow
We consider the characteristic flow,
(3.4)
which can be solved as
(3.5)
By plugging this characteristic flow into (1.9), itô’s formula gives the following semi-martingale decomposition for
(3.6)
with the Martingale term , whose quadratic variations are given by (using (1.10)):
(3.7)
In the rest of this section, we fix , and take satisfying 1.5, and (unless otherwise noted) conditional on which is -Airy-like.
All the constants below (including those in and ) can depend on and .
We take to be a large number, and set .
3.2. Estimates for the bulk
We next prove the following bound of for in a domain contained in .
It will be used to bound the bulk pole densities.
Proposition 3.3.
There exist small such that the following holds. Define the spectral domain
(3.8)
With probability , for any , and , it holds
(3.9)
Thanks to the (away from the real axis) Lipschitz property of in 2.2, we only need to prove (3.9) for a set of carefully chosen mesh points. Namely, we consider the following mesh of points in the upper half plane:
(3.10)
Lemma 3.4.
For any as defined in (3.8) and , there exists such that , and
(3.11)
Proof.
Suppose that for some integer , we can take , therefore . Suppose that for some , we can take . Then , and , and . Also , thus .
∎
Now 3.3 follows from the following estimate on one point.
Lemma 3.5.
The following holds true for small enough .
Take any , and let for each . Conditional on with , with probability ,
The quadratic variation of the martingale term is given by (3.7); then (again using 2.2)
By integrating in time, we have
Take . By the Burkholder-Davis-Gundy inequality, for any the following holds with probability444Here and in the rest of this proof, is used to denote a large constant, whose value may change from line to line. :
Then we can take a union bound over times
for , and get that with probability ,
(3.21)
Now by (3.6), and using (3.20) and (3.21), we have that for any ,
where in the last inequality we used , which is much smaller than provided is large enough; and
(3.25)
Therefore, since is continuous in , we conclude that with probability .
∎
3.3. Estimates for the first pole
Under the same setup as the previous subsection (in particular, is taken to be any large number, and ),
we next prove the following proposition, which states that with high probability all the poles are bounded by .
Proposition 3.6.
There exists a small number such that the following holds. With probability , for any , has no pole in .
We introduce the following stopping time, which is the first time the largest pole exceeds ,
(3.26)
We now denote for (in particular ),
and consider a mesh of points:
Then for any , it holds , and , provided is large enough.
Now 3.6 follows from the following estimate on one point.
Lemma 3.7.
For any , the following holds true for small enough .
Take any , with , and let for each . Conditional on with , with probability ,
This leads to a contradiction.
Therefore (with probability ) there is no pole in at any time in . Recall the definition of from (3.26). Using that is continuous in , and 2.7 and 2.8,
we conclude that , and 3.6 follows.
∎
As before, take large enough and . 3.6 and 3.3 verify the two assumptions in 2.3, respectively. Thus with probability , for any , has infinitely many poles , satisfying
(3.36)
This gives the second statement in 3.1, after replacing by .
The statement (3.36) also verifies the first assumption in 2.1. Moreover, in (3.9) we can take a sequence of complex numbers , so that as . This verifies the second assumption in 2.1, from which (3.1) holds.
Finally, for each , the continuity of in follows from the continuity of in , and 2.7, 2.8.
∎
4. Hölder Regularity
In this section, we upgrade the trajectory continuity into Hölder regularity.
Proposition 4.1.
For any , there exist large and small such that the following holds.
Take any satisfying 1.5, and that its poles are given by a line ensemble , and that for any .
Take large enough and any .
Then conditional on the event that
(4.1)
with probability we have
Proof.
We first prove the (with conditional probability ) upper bound
(4.2)
The lower bound can be proven in the same way.
By (4.1) with (2.5), we conclude that there exists a large constant such that
(4.3)
Take small . Then (4.3) implies that there exists some , such that . We take the smallest such , then .
Let , and take a small and . Next we show that, provided ,
(4.4)
For this, from (4.1) and (2.5), we have for each .
It follows that
using that .
By a similar argument, we can also upper bound the summation over , and (4.4) follows.
For the Martingale term,
using (1.10) and 2.2, it follows that (for )
Therefore, by the Burkholder-Davis-Gundy inequality, there exists a small constant ,
(4.6)
By plugging (4.6) into (4.5), it follows that with probability ,
(4.7)
provided that .
Now we take a large , and assume that . Then for any , and using (4.4), we have
which, in particular, implies that . Since , this further implies that , for any .
In summary, we conclude that with probability , , for any .
Since , it follows that (with probability ) for any ,
Finally, we choose the parameter and , satisfying all the above constraints:
Then we can take and .
By taking large enough and small enough (depending on , , and ), the constraint is also satisfied since ; and (4.2) follows.
∎
5. Recover Dyson Brownian Motion
In this section, we localize any line ensemble given by the poles of the SDE (1.9), by deriving another SDE satisfied by the first finitely many poles (in the sense of weak solution).
More precisely, for the line ensemble of poles,
we prove that, if for some large , and are bounded away from each other for certain amount of time, the evolution of is then described by DBM plus a drift term, describing the effect of .
Proposition 5.1.
For any the following is true.
Take any satisfying 1.5, and that its poles are given by a line ensemble .
Fix a large , and denote the stopping time
(5.1)
Conditional on the event ,
there exist independent Brownian motions adapted to the filtration , satisfying
(5.2)
where is a random meromorphic function, defined as
Moreover, almost surely the following holds:
(5.3)
The rest of this section is devoted to proving 5.1.
Using that are poles of , we derive (5.2) from 1.5, using a contour integral.
For this, we need first establish that the poles do not collide at almost every time (i.e., (5.3)).
The idea to establish the collision time estimate is to consider the process for some , showing that its level- local time equals . We mainly follow the standard argument used to study the Bessel process, see [114, Chapter XI, Section 1].
To analyze such processes we again resort to contour integrals, and therefore an induction will be used.
We next give the semi-martingale decomposition of a process, which is the sum of for in an interval.
For any line ensemble satisfying 1.5, and any with , denote
(5.4)
and
Lemma 5.2.
In the above setup, take any , and denote the stopping time
where we use the convention that , when .
Then in satisfies ,
where
and is the Martingale term, with quadratic variation
(5.5)
Proof.
For simplicity of notations, we fix , and write , , an within this proof.
For with and , we take a contour enclosing , but not any for or . Then by 1.5, we have
(5.6)
Note that
and
Note that the poles of are , except for .
Then we have
(5.7)
using that .
Similarly, we have
(5.8)
Here for the last equality, we used that
and that .
Now using (5.6) and (5.8), and Ito’s formula, we can write , with
Here we used (1.11) in the first equality.
For the second equality, it is by evaluating the contour integral in and , via integration by parts.
As for , we have
By (1.11), the quadratic variation therefore equals
By taking the residues at , we get
By further taking the residues at , this equals
and the conclusion follows.
∎
We next establish the collision time estimate, for poles whose indices are in an interval.
We use the local time of to analyze its boundary behavior at zero.
According to 5.2, for is a semi-martingale. We let be the level local time in (with for each ). Then by [114, Chapter VI, Theorem 1.7], almost surely is continuous in and cadlag (right continuous) in , and satisfies
(5.9)
where for the second equality, we used that if , then and .
Thanks to the occupation time formula [114, Chapter VI, Corollary 1.6], we have
where we used (5.5) which gives
, and .
Since is right continuous in , it follows that , and hence (5.9) implies that almost surely .
Thus the conclusion follows.
∎
Step 1: Non-collision.
We will first show (5.3), i.e., almost everywhere poles do not collide. More precisely, we will prove inductively on
(5.10)
The claim of 5.1 follows from the case of in (5.10).
For the base case where , it follows from 5.3 with and , and .
(Note that in this case, we always have )
We next give the induction step: if (5.10) holds for some , then it holds for .
Under the induction hypothesis, almost everywhere, there exist such that
(5.11)
Fix the indices , the set of time such that (5.11) holds is a random open set, and we denote it by .
Next we show that almost surely, for almost every , take at least distinct values. This implies that (5.10) holds for .
For the convenience of notations, we denote and .
Since , there exists some such that .
We then apply 5.3 with and , and taking any rational numbers.
We note that the union of all such would cover , therefore
Thus for almost every , would take at least two distinct values, so we finish the induction step.
Then by induction principle, we finish the proof of (5.3).
Step 2: Dyson Brownian motion.
We next prove (5.2). For that we need to construct the Brownian motions in (5.2).
By (5.3) and that each is continuous, for any outside a closed measure zero set (i.e., in a countable union of open intervals, whose closure is ), we have for each .
Then we can take a small contour enclosing but not any other poles.
From (5.7) in the proof of 5.2, we have
(5.12)
where
We can then further extend to all of as a continuous process. The quadratic variations are given by
which equals .
Then it follows that are independent Brownian motions.
Another approach to study DBM developed in [66] is based on applying Ito’s formula to the elementary symmetric functions . There a large family of Dyson type interacting particle systems are considered.
For , they show that if the initial data has some particles at the same location, they will separate instantly. The method there could potentially be adapted and derive non-collision in the above proof as well.
6. Coupling and Uniqueness
In this section we prove the uniqueness part of 1.6.
Take any satisfying 1.4 and 1.5. Let be the line ensemble given by its poles (from 3.2).
We also take another line ensemble through the same way.
Proposition 6.1.
The two line ensembles and have the same law.
Our general strategy is to construct a coupling of the dynamics in , where and would get closer as increases.
Then by sending the starting time of the dynamics to , one concludes that these two line ensembles must equal in law.
The coupling.
There are four parameters in the definition of this coupling.
Here are small and large real numbers;
is among the sequence in 1.4, and is large enough depending on ; and we let .
We shall mainly consider the dynamics of the first order many paths, for .
For each , let be the event where
for each .
For each , denote .
Under , by (2.5), we let and be the smallest numbers in ]], such that
where is a large enough universal constant. Note that when is large enough depending on , such and exist.
We introduce a stopping time (with respect to the filtration ), as follows.
If there exists any such that does not hold, or if there is any and , such that
we let be the smallest such .
Otherwise, we let .
Lemma 6.2.
For any , there exist , such that .
Proof.
By 3.1, for small enough , large enough and , we have .
Then by the Hölder continuity estimate 4.1,
where is small enough depending on , and the second inequality is by taking large.
∎
By 5.1, we can find a family of independent Brownian motions , such that for each and , ,
(6.1)
where we used 3.2 for the expression of .
We can similarly find a family of independent Brownian motions , such that for each and , ,
(6.2)
We now couple and so that they equal almost surely.
Thereby, we get a coupling between and .
The following proposition states that under this coupling, these two line ensembles are close to each other with high probability.
Proposition 6.3.
Fix any and . Then there exist , such that under the above coupling with probability ,
In the following, we will prove that with probability ,
(6.3)
The lower bound that can be proven in the same way.
Our strategy is to consider a shifted version of , which at is much larger than ;
then we show that it is larger than for (under the coupling), while the amount of shift is in .
We now define the shifted version of . For any and , we let
where and are taken as follows.
By 1.4 and 2.4, we take taken large enough (depending only on ) such that with probability ,
(6.4)
We then take such that
Then for large enough (depending on ), the above choice of parameters imply
(6.5)
We can now rewrite (6.2) in terms of . For each and , , we have
(6.6)
Lemma 6.4.
There exist , such under the above coupling the following hodls. Take any . Assuming that
then
(6.7)
Assuming this lemma, we can now finish proving the uniqueness in law of line ensembles.
As already alluded to, it suffices to prove (6.3).
From our choice of (see (6.4)), we have that for all . Then by repeatedly applying 6.4 for , and 6.2, we conclude that with probability , we have and for all and . In particular for , this gives
The conclusion follows from taking to zero and to infinity in 6.3.
∎
The rest of this section is devoted to proving 6.4.
The idea is straightforward: from the coupling we take the difference between (6.1) and (6.6), to cancel out the Brownian motions; and the rest are deterministic arguments.
For simplicity of notation, in this proof we fix , and write and . Recall that ]].
We can take the difference between (6.1) and (6.6), so that for any ,
(6.8)
Denote the stopping time to be the first time after , such that there exists at least one index with (if there were multiple such indices, take to be the smallest one). We will prove that then (6.7) holds.
We prove by contradiction, and assume that . By the definition of the stopping time , for each , , and , we have
(6.9)
We let (resp. ) be the smallest (resp. largest) index with (resp. );
and we let be the corresponding indices for .
By (6.9), and that for each ,
necessarily .
Now for (LABEL:e:yxdiff), by summing over , and integrating from to for a sufficiently small , we have