One-point asymptotics for half-flat ASEP
Abstract.
We consider the asymmetric simple exclusion process (ASEP) with half-flat initial condition. We show that the one-point marginals of the ASEP height function are described by those of the process, introduced by Borodin-Ferrari-Sasamoto in (Commun. Pure Appl. Math., 61, 1603-1629, 2008). This result was conjectured by Ortmann-Quastel-Remenik (Ann. Appl. Probab., 26, 507-548), based on an informal asymptotic analysis of exact formulas for generating functions of the half-flat ASEP height function at one spatial point. Our present work provides a fully rigorous derivation and asymptotic analysis of the same generating functions, under certain parameter restrictions of the model.
1. Introduction and main results
1.1. Preface
The asymmetric simple exclusion process (ASEP) is a continuous time Markov process, introduced to the mathematical community by Spitzer [Spi70] in 1970 (and also appearing two years earlier in the biology work of MacDonald, Gibbs and Pipkin [MGP68]).
The state space of the process is and we interpret an element as a particle configuration, where indicates the presence of a particle at location , and indicates the presence of a hole at location . The dynamics of the model depend on a parameter , and can be described as follows. Each particle carries an independent exponential clock, which rings at rate . When the clock rings the particle attempts to jump one site to the right with probability , and with probability it attempts to jump one site to the left. The jump is successful if the site to which the particle attempts to jump is a hole; otherwise, the jump is suppressed and the particle stays put. As mentioned in [Spi70], there is some delicacy in formally constructing a Markov process corresponding to the latter dynamics when the number of particles is infinite; however, a number of papers have been published confirming the existence of such a Markov process – see the works of Harris [Har72], Holley [Hol70], and Liggett [Lig72]. We denote this process by . If (or ) the process is called the totally asymmetric simple exclusion process (TASEP), if it is called the symmetric simple exclusion process (SSEP), and if and it is called the (partially) asymmetric simple exclusion process (ASEP). Throughout the paper we will assume that , so that the particles have a drift to the left.
Given , we define through so that , and then we define the height function of ASEP to be
(1.1) |
where is the net number of particles that crossed from site to up to time , i.e. particles that jump from to are counted as and those that jumped from to are counted as .
ASEP is an important member of the one-dimensional Kardar-Parisi-Zhang (KPZ) universality class. For more on the KPZ universality class we refer to the surveys and books [Cor12, HHT15, QS15] and the references therein. In particular, it is expected that the scaled ASEP height function , appropriately shifted, converges as to the fixed time distribution of a certain Markov process known as the KPZ fixed point, started from an initial state that depends on the initial condition for ASEP. We mention that the KPZ fixed point was constructed by Matetski-Quastel-Remenik [MQR21], and the convergence of the ASEP height function to the KPZ fixed point was proved for a class of initial conditions by Quastel-Sarkar [QS22].
Despite the remarkable progress in understanding the asymptotic behavior of ASEP, the results in [QS22] only apply when the initial condition asymptotically is either continuous with moderate growth, or consists of multiple narrow wedges. There are still many natural and important initial conditions that are not handled by [QS22], and for which precise statements are not available. One such example is the case of half-flat initial condition, which corresponds to starting ASEP from , where . The half-flat initial condition asymptotically becomes the function that is for positive and for negative values, so that it is a natural mixture of the two classes of initial conditions handled in [QS22], without belonging to either.
The goal of the present paper is to investigate the large time limit of the ASEP height function , started from , and show that its one-point marginals are asymptotically described by those of the Airy2→1 process, introduced by Borodin-Ferrari-Sasamoto [BFS08]. This convergence result was conjectured by Ortmann-Quastel-Remenik [OQR16], based on an informal analysis of exact formulas for generating functions of the half-flat ASEP height function at one spatial point. Our present work provides a fully rigorous derivation and asymptotic analysis of the same generating functions, under certain parameter restrictions of the model.
The rest of the introduction is structured as follows. In Section 1.2 we introduce some relevant notation and state a certain moment formula derived in [OQR16] – this is Proposition 1.2. Within the same section we state the exact formulas for the generating functions of the half-flat ASEP height function as Theorem 1.5. We mention that Theorem 1.5 already appeared in [OQR16], but as explained in Remarks 1.8 and 2.4, there are aspects of the derivation of that theorem that are not justified in [OQR16]. In fact, we can only prove the well-posedness of our formulas in Theorem 1.5 when is assumed to be sufficiently close to zero. In Section 1.3 we introduce the process, and in Section 1.4 we present our main asymptotic result.
1.2. Exact formulas for half-flat ASEP
We begin by summarizing some of the notation we require for ASEP in the following definition.
Definition 1.1.
We fix and let , , , so that . Let denote the ASEP started from half-flat initial condition for the parameters as in Section 1.1. We denote the distribution of from this initial condition by and write for the expectation with respect to this measure. For and we define
(1.2) |
to be the total number of particles at or to the left of location at time , and note that from (1.1) we have the following relationship between and the height function
(1.3) |
In the remainder of this section we present the exact formulas for certain observables involving that were derived in [OQR16]. We mention that in view of (1.3) all of the observables below can alternatively be expressed in terms of ; however, we will formulate them with , following [OQR16]. Both the observables and the formulas for them depend on various special functions, which we present next.
For and we let be the -Pochhammer symbol
(1.4) |
We also define the -factorial for
(1.5) |
and the -exponential function
(1.6) |
where the first identity is the definition of and is valid for for and the second holds when – see [AAR99, Corollary 10.2.2a].
For the next several functions we fix as in Definition 1.1, , and . We define
(1.7) |
Throughout the paper we denote by the vector and by the complex integration form . We mention that the dimension and the contours over which the integration is performed will be clear from the context.
For complex vectors and we write
(1.8) |
In equation (1.8) and throughout the paper we will always take the principal branch of the logarithm.
With the above notation in place, we can state the exact formulas we use.
Proposition 1.2.
Remark 1.3.
[OQR16, Theorem 1.3] is formulated for general contours , and as explained in that theorem needs to be positively oriented, enclose , , and exclude all other poles of . Our choice for to be the positively oriented circle of radius , centered at the origin, clearly satisfies all of these conditions.
Remark 1.4.
Let us briefly explain how Proposition 1.2 is proved in [OQR16]. If one defines
it was shown in [BCS14] that is the unique solution to a certain system of differential equations. In [OQR16] the authors discovered a certain -fold contour integral that solved this system, and thus provided formulas for the observables . These formulas can be found in [OQR16, Theorem 1.2], and the derivation is done in Section 2 of that paper. From [BCS14, Lemma 4.18] there is a way to express as a linear combination over with and , which allows one to express as a certain -fold contour integral over different contours. This is done in [OQR16, Proposition 3.2]. Deforming all of these contours to the same one, using a nested contour integral ansatz that is similar to [BC14, Proposition 3.8] and whose proof is inspired by [HO97], one arrives at (1.9).
Using the formulas for the moments of , one can obtain a formula for the generating series of the -Laplace transform of . This formula is given in the following theorem, and forms the basis of our asymptotic analysis.
Theorem 1.5.
Remark 1.6.
Remark 1.7.
The way that the formula in (1.12) is derived is by first showing that for
(1.14) |
which is a direct consequence of (1.6). Afterwards, we use the formulas for the moments from (1.9), take the limit and rearrange the resulting sum. In this way one obtains
(1.15) |
One can express the -fold sums over as -fold contour integrals using a Mellin-Barnes type integral representation from [BC14, Lemma 3.20]. This proves (1.12) for , and then one shows that the equality holds for all by analytic continuation. We mention that in taking the limit, as well as in applying the analytic continuation argument, we require absolute summability of the series in (1.15), which we can ensure only when is sufficiently small. This is the origin of the assumption (1.11). For general the integrals in (1.13) are still well-defined; however, we do not know how to show that the series in (1.12) is convergent. The proof of Theorem 1.5 is the content of Section 2.
Remark 1.8.
Theorem 1.5 appears as Theorem 1.4 in [OQR16], although there is a missing term in the integrand in (1.13) and the result is formulated without the assumption (1.11). The authors derive their result following the outline we gave in Remark 1.7 and it is when the authors apply this Mellin-Barnes integral representation that the term is dropped in [OQR16], see [OQR16, Equation (4.13)]. This is a small typo in the paper. A more serious problem, which we discuss in Remark 2.4, is that the authors do not provide sufficient justification to demonstrate that the series in (1.15) is absolutely convergent for , and that one can analytically extend the equality in (1.12) to . In fact, it is not even clear from the arguments in [OQR16] that the right side of (1.12) is convergent and hence well-defined.
1.3. The process
In this section we define the process, which is the object that describes the large time limit of the height function from (1.3) under the measure from Definition 1.1. The process was introduced by Borodin-Ferrari-Sasamoto [BFS08], where it arises from the limit of the TASEP with half-flat initial condition.
Let us first introduce certain complex contours that arise in our discussion.
Definition 1.9.
For and we define the contour to be the union of and oriented to have increasing imaginary part.
The following is Definition 2.1 from [BFS08].
Definition 1.10.
(The process) Let us set
(1.16) |
and define the kernel
(1.17) |
where are as in Definition 1.9. The Airy2→1 process, denoted by , is the process on with -point joint distribution at given by the Fredholm determinant
(1.18) |
In (1.18) we have that the measure on is the product measure of the counting measure on and the Lebesgue measure on . The function is defined on through . We mention that in [BFS08, Appendix B] it was shown that there is a trace class kernel on that is conjugate to , so that the Fredholm determinant in (1.18) is well-defined.
Remark 1.11.
In [BFS08] the authors chose a different set of contours in the definition of in (1.17), denoted by in that paper. Starting from [BFS08, (2.7)], we can deform to , respectively, without crossing any poles of the integrand and thus without affecting the value of by Cauchy’s theorem. The decay necessary to deform the contours near infinity comes from the cubic terms in the exponential functions. We also mention that there is an extra minus sign in the integrand in (1.17), compared to [BFS08, (2.7)], which comes from the fact that in that paper is oriented in the direction if decreasing imaginary part, whereas is oriented in the direction of increasing imaginary part.
Remark 1.12.
As explained in [BFS08], we have that the finite dimensional distributions of converge to those of as , where is the Airy1 process, introduced by Sasamoto [Sas05]. In addition, converges to as , where is the Airy2 process, introduced by Prähofer and Spohn [PS02]. We refer the interested reader to [QR13, QR14] for more background on the Airy2→1 and other Airy processes.
The way that the formula (1.18) was derived in [BFS08] was by constructing a certain process , related to the TASEP with half-flat initial condition, and showing that
(1.19) |
The fact that the right side of (1.19) is the pointwise limit of distribution functions on does not a priori imply that it is itself a distribution function. As we could not find a place in the literature where it is shown that the right side of (1.19) is a distribution function, we show it in the following lemma, whose proof is given in Section 6. In the same lemma we also establish some properties of the finite-dimensional distribution functions of , which will be required in our arguments.
Lemma 1.13.
The following all hold.
1.4. Main result
Here we present the main asymptotic result of the paper, which describes the large time limit of the height function as in (1.3), under the measure from Definition 1.1.
Theorem 1.14.
Remark 1.15.
We mention that Theorem 1.14 agrees with the analogous result for TASEP (corresponding to and ) from [BFS08] as we explain here. If we let denote the location of the -th leftmost particle of TASEP at time , we have from [BFS08, Theorem 2.2] that for each , and
(1.22) |
We mention that we have set in [BFS08, Theorem 2.2] and also as in that paper, since we have started from the initial condition and , while in [BFS08] the convention and is used. Let us set
and observe that we have the equality of events
where we recall that is as in (1.2). In particular, from (1.3) and (1.22) we get
(1.23) |
Equation (1.23) agrees with (1.21) as . We mention that (1.23) does not follow from Theorem 1.14, as is not allowed, and the point of this computation is to provide a sanity check for the scaling we perform in Theorem 1.14.
Outline
The rest of the paper is organized as follows. In Section 2 we prove Theorem 1.5, which is the starting point of our asymptotic analysis. In Section 3 we present the proof of Theorem 1.14, which relies on taking the limit of the identity (1.12), when is scaled appropriately with and . In order to obtain the limit of (1.12), we need to show that each summand converges, which is the statement of Proposition 3.2, and that we can exchange the order of the series and the limit, which is ensured by Proposition 3.3. Proposition 3.2 is proved in Section 4 and relies on the method of steepest descent, as well as various estimates of the functions in (1.7). Proposition 3.3 is proved in Section 5 and is based on estimating the functions in (1.8) along two classes of contours, depending on whether is large or small relative to . In Section 6 we prove Lemma 1.13.
Acknowledgments
We are grateful to Alexei Borodin, Jeremy Quastel and Daniel Remenik for their useful comments on earlier drafts of the paper. ED is partially supported by NSF grant DMS:2054703.
2. Prelimit formula
In this section we present the proof of Theorem 1.5, which is the starting point of our asymptotic analysis in the next sections. The proof of Theorem 1.5 is given in Section 2.2 and in Section 2.1 below we establish several statements, which will be required. We continue with the same notation as in Section 1.
2.1. Definitions and notation
In this section we present two key lemmas, which will be required in the proof of Theorem 1.5 in the next section, and whose proofs are postponed until Section 2.3. After stating the lemmas we summarize various estimates, which will also be of future use.
The first key result we require is as follows.
Lemma 2.1.
The second key lemma we need is the following. It is a special case of a Mellin-Barnes type integral representation from [BC14, Lemma 3.20].
Lemma 2.2.
In the remainder of this section we summarize several basic estimates for the various functions that appear in (1.8). We observe that given we can find such that if and , then
(2.2) |
Recall the Cauchy determinant formula, see e.g. [Pra94, 1.3],
(2.3) |
The following statement summarizes two different bounds on the above Cauchy determinant.
Lemma 2.3.
[Dim20, Lemma 3.12] Let .
-
(1)
Hadamard’s inequality: If is an matrix and denote the column vectors of , then where for .
-
(2)
Fix with . Let be such that and for . Then,
(2.4)
Let be as in (1.7). Then, we can find , depending on , such that for all (recall this was the positively oriented circle of radius , centered at the origin) and such that we have
(2.5) |
Also if and with we have
(2.6) |
where we used that for the function is maximized when and minimized when .
2.2. Proof of Theorem 1.5
We continue with the same notation as in the statement of the theorem. For clarity we split the proof into two steps.
Step 1. Note that by Lemma 2.1 each summand on the right side of (1.12) is well-defined and finite, and moreover by our assumption on in (1.11) we have that the series on the right side is absolutely convergent and defines an analytic function in on .
We claim that (1.12) holds provided that is such that . We will prove this statement in the next step. Here we assume its validity and proceed to prove that (1.12) holds for all .
From (1.6) we have
(2.10) |
If is compact we observe that so is the set , and the latter is separated from the zeros of the function . In particular, we conclude that there exists a constant , depending on and , such that for all and we have
The latter implies that the right side of (2.10) is the uniform over limit of
as . Since each of the above functions are analytic in on , we conclude the same is true for the expressions in (2.10), cf. [SS03, Chapter 2, Theorem 5.2].
Our work in the above paragraph shows that the left side of (1.12) is analytic in on and from Lemma 2.1 we know the same is true for the right side. As these two functions agree when is such that by assumption, we conclude that they agree for all , cf. [SS03, Chapter 2, Theorem 4.8]. This concludes the proof of the theorem.
Step 2. In this step we fix such that and proceed to prove (1.12). From (1.6)
(2.11) |
where we used Fubini’s theorem to exchange the order of the sum and the expectation. Note that the application of Fubini’s theorem is justified since
where we used that almost surely, and by assumption.
We further have from Proposition 1.2 for each that
(2.12) |
From (2.8) and the fact that has length we have
(2.13) |
In addition, from (2.13) and the dominated convergence theorem we have
The last equation and Lemma 2.2 together imply
(2.14) |
Combining (2.11) and (2.12) we conclude
(2.15) |
where in the last equality we used (2.14). We mention that in exchanging the order of the sum and the limit above we used the dominated convergence theorem as from (2.13) we have
and the latter is summable over . From (2.15) we get (1.12) when , as desired.
Remark 2.4.
Now that we have presented our proof of Theorem 1.5, let us explain how our approach compares with that in [OQR16]. The overall strategy of both proofs is quite similar, in that (1.12) is first established when is such that , and then extended to by analytic continuation. In the case of the main difficulty is in taking the limit of (2.12) and justifying exchanging the order of the series on the right side of that equation and the limit. To analytically extend (1.12) to , the main difficulty is in showing that the series on the right side of (1.12) is absolutely convergent and hence analytic in (as the absolutely convergent series of analytic in functions).
In order to overcome the two challenges above, one needs to find good bounds on the summands in the two series in (1.12) and (2.12). One runs into trouble, since over the best pointwise estimate for (which appears in both and ) is for some . This bound behaves very poorly in , and cannot be compensated by the in the denominator of in (1.13). Faced with this difficulty, the authors [OQR16] suggested to deform to certain -dependent contours , and for general also deform to certain -dependent contours . Over the contours and one can obtain a favorable estimate for . Unfortunately, what was not realized in [OQR16] is that over these deformed contours other parts of the integrands in the definitions of and behave badly and become difficult to control.
The way we overcome the growth of as a function of , is to recognize that one has some decay built into the Cauchy determinant in the definitions of and . This decay is sufficient to control the series in (1.12) and (2.12), only when is sufficiently small – see (1.11). In fact, it still remains a difficulty for us to prove that the right side of (1.12) is convergent and hence well-defined for general .
2.3. Proof of Lemmas 2.1 and 2.2
In this section we give the proofs of the two key lemmas from Section 2.1.
Proof of Lemma 2.1.
Let us fix a compact set and . It follows from (2.2) that there are constants , depending on alone, such that for all
(2.16) |
Combining (2.16) with (2.8) we conclude that the integral in (1.13) is well-defined and finite. Moreover, we see that is the uniform over limit of
The integrand in is analytic in on and is jointly continuous in and , while the contours we are integrating over are compact. The latter implies that is analytic in for each , see e.g. [SS03, Theorem 5.4]. We thus conclude that is analytic as the uniform over compact sets limit of analytic functions, cf. [SS03, Chapter 2, Theorem 5.2].
Using that for we have , that the length of is as well as (2.8) and (2.16) we conclude that
The right side above is summable over , since in view of (1.11). The latter implies that the series in (1.12) is absolutely convergent. Moreover, since each summand is analytic in on , we conclude the same is true for the series by [SS03, Chapter 2, Theorem 5.2]. ∎
Proof of Lemma 2.2.
From (2.8) and the fact that the length of is we conclude that the -th summand in (2.1) is bounded in absolute value by
and the latter is summable over (since by assumption), proving the absolute convergence of the series in (2.1).
To conclude the proof of the lemma, it suffices to show
(2.17) |
Let and set , , , . Denote by the contour, which goes from vertically up to , by the contour, which goes from horizontally to , by the contour, which goes from vertically down to and by the contour, which goes from horizontally to . Also let traversed in order, see Figure 1.

We observe by the Residue Theorem that for each
(2.18) |
where we recall that was defined in (1.8). In deriving the last expression we used that in each variable for the function is analytic in the region enclosed by except at the points where the function has a simple pole coming from , the fact that
and also that is negatively oriented.
We next note by the dominated convergence theorem that
(2.19) |
In deriving the last statement we used (2.8) and (2.16), which justify the dominated convergence theorem with dominating function (here are as in (2.16) and we have written for ).
In addition, we have from (2.2) and (2.8), the fact that the length of is , the length of is for and that for
(2.20) |
Using also that the length of is , and the same statements as above we get
(2.21) |
We mention that in (2.20) and (2.21) the constant is as in (2.2) for , and we used that are at least distance from by construction.
Since by assumption, we have that the right sides of (2.20) and (2.21) both converge to zero as . Combining (2.18), (2.19), (2.20) and (2.21) we conclude (2.17), which concludes the proof of the lemma.
∎
3. Weak convergence
The goal of this section is to prove Theorem 1.14, for which we need to study equation (1.12) in Theorem 1.5 as . In Seciton 3.1 we explain how we need to scale the parameters in (1.12), and formulate two key asymptotic statements about the summands – see Propositions 3.2 and 3.3. In Section 3.2 we use these two propositions to complete the proof of Theorem 1.14. In Section 3.3 we present a useful way to rewrite , which will help us establish Propositions 3.2 and 3.3, whose proofs are given in Sections 4 and 5, respectively. Throughout this section we continue with the same notation as in Sections 1 and 2.
3.1. Two key propositions
In this section we formulate the key results we require in the proof of Theorem 1.14 in Section 3.2. We begin by stating our assumptions on parameters and their scaling.
Definition 3.1.
The first key proposition we require is as follows.
Proposition 3.2.
The second key proposition we require is as follows.
Proposition 3.3.
Assume the same notation as in Definition 3.1. There exists sufficiently small, so that the following holds. If , we can find constants , depending on , such that for and
(3.3) |
We end this section with the following elementary probability lemma from [BC14], which will also be required in our arguments. We mention that analogues of the below lemma have been known for a while in the physics literature, see e.g. [CLDR10, Equation (14)].
Lemma 3.4.
[BC14, Lemma 4.39]. Suppose that is a sequence of functions , such that for each , is strictly decreasing in with a limit of at and at . Assume that for each one has on , uniformly. Let be a sequence of random variables such that for each
and assume that is a continuous probability distribution function. Then converges in distribution to a random variable , such that .
3.2. Proof of Theorem 1.14
In this section we present the proof of Theorem 1.14. For clarity, we split the proof into two steps.
Step 1. Let be as in Proposition 3.3, and let be sufficiently small so that and for we have that (1.11) holds. This specifies in the statement of the theorem. In the remainder we fix and a sequence , such that as .
For and we define
We further introduce the random variables
(3.4) |
where is as in (1.2). We claim that for each we have
(3.5) |
where is as in Definition 1.10. We prove (3.5) in Step 2. Here, we assume its validity and conclude the proof of the theorem.
From [FV15, Lemma 5.1] we have that
(3.6) |
is strictly decreasing for all . Moreover, for each one has uniformly on as . The latter implies that satisfy the conditions of Lemma 3.4. We further note that the right side of (3.5) is continuous in from the second part of Lemma 1.13. In particular, we see that the conditions of Lemma 3.4 are all satisfied, and so
(3.7) |
Step 2. In this step we prove (3.5). From Theorem 1.5 (here we use that ) and Definition 3.1 we know that
(3.8) |
We observe that each summand in (3.8) converges by Proposition 3.2, and also from Proposition 3.3 we may exchange the order of the series and the limit by the dominated convergence theorem with dominating series . Consequently, we conclude from Propositions 3.2 and 3.3 that
We may now apply the change of variables , for to get
(3.9) |
In the last integral we may deform the contours to without crossing any poles of the integrands, and thus without affecting the value of the integrals. The decay necessary to deform the contours near infinity comes from the cubic terms in the exponential functions. At this point, we see that the right side of (3.9) agrees with the right side of the first line of (1.20) with and . From (3.9) and Lemma 1.13 we conclude (3.5).
3.3. Change of variables
In this section we rewrite the function from (1.13) in a way that is suitable for our asymptotic analysis in the next two sections, see Lemma 3.8. In order to state our new formula we require a bit of notation, and our exposition here follows [Dim20, Section 3].
Definition 3.5.
Fix . Let be such that , for any and . For such a set of parameters we define the function
(3.10) |
where everywhere we take the principal branch of the logarthm, i.e. if with and we set . Observe that
which implies that each of the summands in (3.10) is well-defined and finite.
Remark 3.6.
In equation (3.10) we chose the principal branch of the logarithm for expressing and . However, we could have chosen different branches for and , and note that then would shift by for some . Since the sum in the definition of is over we see that such a shift does not change the value of . So even though the logarithm is a multi-valued function for fixed the function as a function of is single valued and well-defined as long as for some .
We next summarize the main properties we require for the function from Definition 3.5 in the following lemma.
Lemma 3.7.
[Dim20, Lemma 3.9] Fix and such that . Denote by the annulus of inner radius and outer radius that has been centered at the origin. Then, the function from Definition 3.5 is well-defined for and is jointly continuous in those variables (for fixed ) over . If we fix and then as a function of , is analytic on ; analogously, if we fix and then as a function of , is analytic on . Finally, if we fix with then is analytic in as a function of .
With the above notation in place we can state the main result of this section.
Lemma 3.8.
Proof.
We first note that by Cauchy’s theorem we may deform to as in Proposition 1.2, and to without affecting the value of the integral. In the latter statement we used that is analytic separately in and , in view of Lemma 3.7, and that in the process of deformation we do not cross any poles of the integrand. We proceed to denote by and by in the remainder of the proof.
Expanding the determinant on the right side of (3.11) and the Cauchy determinant in the definition of , see (1.8) and (1.13), we see that to prove (3.11) it suffices to show that for each (the permutation group of elements) and we have
Writing
and recalling the definition of from (1.7), and that , we see that to conclude (3.11) it suffices to show that for each and we have
(3.13) |
We next note that
(3.14) |
We mention that in deriving the last equality we used that for all and the fact that , depend on only through , while depends on only through and , see (1.7).
Applying the change of variables for in (3.14) we conclude that
(3.15) |
where we mention that the extra sign in (3.15) comes from the fact that is positively oriented while covers with negative orientation as varies from upward to . From Fubini’s theorem and (2.2) we can put the sums in (3.15) inside the integral at which point we obtain (3.13), as desired. ∎
4. Asymptotic analysis: Part I
The goal of this section is to prove Proposition 3.2. In Section 4.1 we use Lemma 3.8 to express from Proposition 3.2 as a -fold contour integral, see (4.5), which is suitable for asymptotic analysis. The formula for in (4.5) involves several functions and in Section 4.2 we establish various estimates for these functions. The proof of Proposition 3.2 is presented in Section 4.3. It is based on the method of steepest descent and relies on the formula for from (4.5) and the results from Section 4.2. The lemmas in Section 4.1 and 4.2 are proved in Section 4.4.
4.1. Formula for
The goal of this section is to get a formula for from Proposition 3.2 that is suitable for asymptotic analysis. We begin with a lemma, which is proved in Section 4.4.
Lemma 4.1.
Let , be such that and for . Then we have
(4.1) |
The next definition introduces various contours, which will be required in our arguments.
Definition 4.2.
For each and , we let and be two contours, where
The contours are oriented to have increasing imaginary part.
For we define the contours
All of the above contours are oriented in the direction of increasing imaginary part.
We also define and to be the positively oriented contours obtained from and , respectively, under the map . Observe that and are piecewise smooth positively oriented contours, is contained in the closed unit disc in , which in turn is contained in the region enclosed by . Some of the contours in the definition are depicted in Figure 2.

The following lemma introduces a certain analytic function , which will appear in our analysis, and establishes a few of its properties. Its proof is given in Section 4.4.
Lemma 4.3.
For such that define the function
(4.2) |
There exist universal constants , , such that
(4.3) |
where are as in Definition 4.2. Furthermore, for any we have
(4.4) |
Definition 4.4.
With the above notation, we are ready to state our formula for . Let be as in Definition 4.4, and for let , be as in Definition 4.2. We then have the following formula for all and
(4.5) |
where
(4.6) |
(4.7) |
Let us briefly explain the origin of (4.5). Our starting point is the formula for from (3.11) in Lemma 3.8, where we take and . At this point we can deform the contours in (3.11) to as in Definition 4.2 without crossing any poles and thus without affecting the value of the integral by Cauchy’s theorem. Here we used that is analytic in for fixed , as follows from Lemma 3.7. We mention that the integrand in (3.11) has singularities at for all , coming from , at , coming from , at for all , coming from and , and all of these singularities lie outside of the annulus of inner radius and outer radius , where the deformation takes place. There is also a singularity at , coming from , which is also not crossed in the deformation, see the left part of Figure 2. After we deform all the contours to we proceed to deform the contours to as in Definition 4.2. Again, in the process of deformation we do not cross any poles, and by Cauchy’s theorem we do not change the value of the integral. We mention that the only poles in the annulus of inner radius and outer radius , where the deformation takes place, come from (from ), (from ), and at (from and ). As is strictly contained in the unit circle, and is outside of it we do not cross the and the poles. On the other hand, the fact that is to the left of , as in Definition 4.2, ensures that we also do not cross .
4.2. Preliminary estimates
In this section we state various estimates for the functions for and for from (4.6) and (4.7) in a sequence of lemmas, whose proofs are given in Section 4.4. These estimates will be required in the proof of Proposition 3.2 in Section 4.3. In the lemmas below are as in Definition 4.2.
Lemma 4.5.
Let and be as in Lemma 4.3, and . We can find , depending on alone, such that the following holds. If ,
(4.8) |
Lemma 4.6.
Let be as in Lemma 4.3, and be sufficiently small so that . We can find , depending on alone, such that the following holds. If , , ,
(4.9) |
Lemma 4.7.
Let be as in Lemma 4.3, and . We can find , depending on alone, such that the following holds. If ,
(4.10) |
Lemma 4.8.
Let be as in Lemma 4.3, and be sufficiently small so that . We can find , depending on alone, such that the following holds. If , ,
(4.11) |
Lemma 4.9.
4.3. Proof of Proposition 3.2
In this section we present the proof of Proposition 3.2, and for clarity we split the proof into three steps.
Step 1. Let be as in Definition 4.4, and let be sufficiently large, so that and , where is as in Definition 4.4, and are as in Lemmas 4.5-4.10. For and , we define
(4.14) |
where are as in Definition 4.2, are as in (4.6), and are as in (4.7). We claim that
(4.15) |
We prove (4.15) in the steps below. Here, we assume its validity and conclude the proof of (3.2).
It follows from (4.5) that
(4.16) |
It follows from Lemmas 4.5, 4.6, 4.7, 4.8, 4.9, and 4.10 (all lemmas are applied for as in Definition 4.4 and ) that for , , and , we have
(4.17) |
Observe that if , we have
Using the latter and (4.17), we conclude that there exist constants , such that for
(4.18) |
Combining (4.16), (4.18) with the fact that the lengths of for are at most , we conclude that for all large we have
Step 2. From (4.14) we have that
(4.19) |
We claim that for all and we have
(4.20) |
We prove (4.20) in the next step. Here, we assume its validity and conclude the proof of (4.15).
In view of (4.20) and the dominated convergence theorem, with dominating function given by the right side of (4.17), we may take the limit in (4.19) to get (4.15).
Step 3. In this step we prove (4.20), and in the sequel we assume that , and are fixed. We begin to investigate the limits of and in the definition of .
From the first line of (4.3) we have for all large and fixed
In particular, we conclude that
(4.21) |
We also observe that for each
where we can exchange the order of the sum and the limit by (2.2), and only the summand contributes to the limit. We conclude that
(4.22) |
By a direct Taylor expansion we have for each and
from which we conclude that
(4.23) |
For every we have
from which we conclude that
(4.24) |
Finally, we have the straightforward limits
(4.25) |
4.4. Proof of the lemmas from Sections 4.1 and 4.2
Proof of Lemma 4.1.
The proof we present here is an adaptation of the one in [GT20, Section A]. We proceed to prove (4.1) by induction on with base case being obvious. Assuming the result for we proceed to prove it for . Let us denote for convenience
By dividing each row of by the entry in the first column, and subtracting the first row from all other rows, we obtain
We next note by a direct computation that
Combining the last two equalities and the multi-linearity of the determinant we get
Applying the induction hypothesis to the last determinant we arrive at (4.1) for , which completes the induction step. ∎
Proof of Lemma 4.3.
We note that for any
By direct computation, we conclude that
The last equality for and implies (4.4). In the remainder we prove (4.3).
We note that is a meromorphic function on with simple poles at , and so it is analytic in the zero-centered disc of radius . If is the Taylor expansion of at the origin, we directly compute and . The latter suggests that is an analytic function in the zero-centered disc of radius , so that there is a constant , such that for all we have
This specifies our choice of and proves the first inequality in (4.3) for all .
Let be sufficiently small so that . We also let . This specifies our choice of and and we proceed to prove the second and third line in (4.3). As the proofs of the second and third inequality in (4.3) are quite similar, we only establish the second.
We fix , such that . We also fix , such that , and note that , where , while . In particular, we see that , so that from the first inequality in (4.3) we have
(4.26) |
We also have from that
(4.27) |
Proof of Lemma 4.5.
Let be as in Lemma 4.3. We will prove that if
(4.29) |
Notice that (4.29) implies the statement of the lemma with , , and . As the proofs of the two lines of (4.29) are quite similar, we only establish the second.
Notice that if and , we have from the third line of (4.3) with
(4.30) |
Suppose that is such that , and let , where the sign is chosen to agree with the imaginary part of . Then, we observe that
(4.31) |
where in the first inequality we used the second inequality in (4.4) with , in the second inequality we used (4.30) with , and in the last inequality we used that , . Equations (4.30) and (4.31) imply the second line of (4.29), as for all . ∎
Proof of Lemma 4.6.
We set and note that for , , and we have
The latter implies that the conditions of [Dim18, Lemma 4.5] are satisfied with , , (here we used that as in Lemma 4.3, and ). From the proof of [Dim18, Lemma 4.5], see the displayed equation after [Dim18, (6.15)], we conclude that
The latter, the triangle inequality, and the fact that for , imply that the left side of (4.9) is bounded by
which proves (4.9) with and . ∎
Proof of Lemma 4.7.
Let . We claim that we can find , such that for we have
(4.32) |
If (4.32) holds, then (4.10) would follow with the same choice of and , since and for we have , .
In the remainder we prove (4.32). By Taylor expansion, we can find , such that for the function is well-defined, and
As always, the logarithm is with respect to the principal branch. This shows that for
(4.33) |
Since is continuous and does not vanish on (note that the zeros are at ), we conclude that there is a constant such that for
In particular, we see that if and we have
(4.34) |
Proof of Lemma 4.8.
We set and proceed to estimate the various terms that appear in (4.11) for . Throughout we will use frequently that , which follows from our assumptions.
We note that if and , then
(4.35) |
Since is entire, we can find a constant , such that if we have
Since for we have for all , we conclude that
(4.36) |
where in going to the second line we also used (4.35), in the second inequality we used that , and in the last one we used that .
Since for , we see that
(4.37) |
where again we used .
Proof of Lemma 4.9.
Let and be small enough so that for we have
We note that if we have and for and . The latter implies that for , and we have
(4.40) |
Next, we let be the compact set . We observe that does not vanish on , and so we can find , such that for all
Notice that for , and we have , upto a shift by for some , belongs to . The latter implies that for , and we have
(4.41) |
Combining (4.40), (4.41) and Hadamard’s inequality from Lemma 2.3, we get for
which proves (4.12) with . ∎
5. Asymptotic analysis: Part II
In this section we present the proof of Proposition 3.3. In Section 5.1 we present several lemmas, which will be required for our arguments, and whose proofs are given in Section 5.4. The proof of Proposition 3.3 is established by considering the cases and separately in Sections 5.2 and 5.3, respectively.
5.1. Preliminary results
In this section we summarize various estimates that will be required in the proof of Proposition 3.3. As we explained earlier, the estimate of in Proposition 3.3 is established by considering the cases and separately. In the case , we will use the formula for from (4.5) and to upper bound it we will use the estimates in Lemmas 4.5-4.9. In addition, we will use Lemma 5.1 below in place of Lemma 4.10 to upper bound the term from (4.7). For the case , we use a different formula for , established in (5.3). This formula involves a different set of contours and functions and and we estimate these new functions in Lemmas 5.3-5.7. The proofs of all lemmas from this section can be found in Section 5.4.
Lemma 5.1.
There is a universal constant such that the following holds for . If and with for , then
(5.1) |
where
(5.2) |
Definition 5.2.
Let , and let be the positively oriented circles, centered at the origin, of radii and . We observe that the choice of satisfies , since .
Using the definition of from Definition 3.1 and Lemma 3.8 we conclude that
(5.3) |
where are as in Definition 5.2, and
(5.4) |
(5.5) |
In equations (5.5), and (5.4) we take the principal branch of the logarithm everywhere.
In the remainder of this section, we estimate the functions that appear in , over the contours in a sequence of lemmas.
Lemma 5.3.
Let be as in Definition 5.2. For , and we have
(5.6) |
Lemma 5.4.
Let be as in Definition 5.2. There exists a constant , depending on alone, such that for all , , , and we have
(5.7) |
Lemma 5.5.
Let be as in Definition 5.2. There exists a constant , depending on alone, such that for all , , , and we have
(5.8) |
Lemma 5.6.
Let be as in Definition 5.2. There exists a constant , depending on alone, such that for all , , , and we have
(5.9) |
Lemma 5.7.
Let be as in Definition 5.2, and fix . Suppose that , , for . Then, we have
(5.10) |
Lemma 5.8.
Let be as in Definition 5.2. We can find a universal constant , such that for , , , for
(5.11) |
5.2. Proof of Proposition 3.3: Part I
In this section we specify the choice of , and as in the statement of Proposition 3.3 and prove (3.3) when .
Let be as in Lemma 4.3 and as in Lemma 5.1. We assume that is sufficiently small so that the following all hold for
(5.12) |
where as in Lemma 4.5, is as in (5.2), and is as in Lemma 5.8. This specifies in the statement of the proposition. Below we fix .
We let be sufficiently large so that for we have
(5.13) |
where is as in Lemma 4.3, and are as in Definition 4.2, is as in Lemma 4.5, is as in Lemma 4.6, is as in Lemma 4.7, is as in Lemma 4.8, is as in Lemma 4.9, and is as in the statement of the proposition. In (5.13) the integrals are with respect to arc-length. This specifies in the statement of the proposition.
We now fix sufficiently large so that for we have
(5.14) |
where for are as in Lemmas 4.5 - 4.9 (here and so ). In addition, is as in (5.2), is as in Lemma 5.4, is as in Lemma 5.5, is as in Lemma 5.6, are as in the statement of the proposition. This specifies in the statement of the proposition.
In the remainder of this section we assume , , and proceed to prove (3.3). It follows from (4.5) that
(5.15) |
where for are as in (4.6), for are as in (4.7), , and . We mention that in obtaining (5.15) we implicitly used that , as follows from the second inequality in (5.14), and that as follows from the first inequality in (5.12).
It follows from Lemmas 4.5-4.9 that if and we have
In addition, from Lemma 5.1 we have
We next observe that our assumptions give the inequalities
where the first inequality used the first inequality in (5.12) and the third inequality in (5.14).
Combining the last three inequalities, we get
(5.16) |
We next claim that
(5.17) |
To see why (5.17) holds let us denote by the set indices in such that and , where we recall that were defined in Definition 4.2. We also set . Using that and , we see that
We also have for that or (or both), which implies that for
We remark that in the first inequality we used that and in the second one we used the third inequality in (5.12) and the fact that for .
Finally, we note that for we have and where and . The latter implies that
Combining the last three inequalities, we deduce (5.17).
5.3. Proof of Proposition 3.3: Part II
In this section we assume , , and proceed to prove (3.3). It follows from (5.3) that
(5.19) |
where for are as in (5.5), for are as in (5.4), , and and , denote integration with respect to arc-length. We mention that in obtaining (5.19) we implicitly used that as follows from the first inequality in (5.12).
From Lemmas 5.3-5.8, and the fact that , we have for and that
We mention that in applying Lemmas 5.3-5.8 we used that , and . Combining the latter with (5.19), the fourth inequality in (5.12), and the fact that the length of is and that of is , we conclude that
(5.20) |
In view of (5.20) and the second line in (5.14), we conclude that as by construction. This proves (3.3) when .
5.4. Proof of the lemmas from Section 5.1
In this section we present the proofs of the seven lemmas from Section 5.1.
Proof of Lemma 5.1.
We first note that if and then
where in the last line the order of the sums can be exchanged by Fubini’s theorem as
In particular, if we set
we see that for and we have
(5.21) |
and the latter is an absolutely convergent in the closed disc of radius .
Proof of Lemma 5.3.
Proof of Lemma 5.4.
Proof of Lemma 5.5.
Using that and , we see that
Combining the latter with the fact that for we have , we conclude (5.8) with . ∎
Proof of Lemma 5.6.
Proof of Lemma 5.7.
6. Appendix: Proof of Lemma 1.13
In the proof we use the same notation as in Definition 1.10. For clarity, we split the proof into four steps.
Step 1. For we define
(6.1) |
In this step we show that is a distribution function (DF) on .
From the displayed equation preceding Remark 1.1 in [QR13], we see that as in (6.1) is the same as from [QR13, (1.7)]. In particular, from [QR13, Theorem 1] we have for
(6.2) |
where is the Airy2 process. From the almost sure continuity of the Airy2 process, see [Joh03, Theorem 1.2], we conclude that is the DF of an extended real-valued random variable taking values in , and so it suffices to show that
(6.3) |
To prove (6.3), we note from (6.2) that for each and
(6.4) |
where is the GOE Tracy-Widom distribution [TW96]. We mention that the equality in (6.4) was derived by Johansson [Joh03] (the factor is omitted in that paper, see [CQR13, Theorem 1] for the correct statement). Using that is a DF on , we have . The latter statement and (6.4) imply (6.3), and so is a DF on . We mention that (6.3) also follows from [QR13, Proposition 1.2].
Step 2. In this step we prove the third part of the lemma. By using the Fredholm determinant expansion, we have for all
(6.5) |
and . We mention that in deriving the formula for we used the definition of from (1.17) and the multilinearity of the determinant function.
We next seek to exchange the order of the integrals in . We observe that we can find a constant , depending on , such that
(6.6) |
where , denote integration with respect to arc length. The latter follows from the cubic terms in the exponential functions.
Note that for , we have . The latter and Hadamard’s inequality from Lemma 2.3 imply
(6.7) |
In addition, since , we have
(6.8) |
Combining (6.7) and (6.8) with the fact that , we conclude that
(6.9) |
where , .
One consequence of (6.9) is that
(6.10) |
which is summable over . Another consequence of (6.9) is that we can exchange the order of the integrals in without affecting the value of the integral by Fubini’s theorem. The result is
(6.11) |
where we used the multilinearity of the determinant, the fact that the determinant of a matrix is equal to that of its transpose, and that for we have
Equations (6.5) and (6.11) establish (1.20), and so we conclude the third part of the lemma.
Step 3. In this step we prove the second part of the lemma. We first show that for each the function from (6.1) is continuous. Let for be such that . We seek to show that
(6.12) |
We define for , and
Let be sufficiently large so that for all . We observe that the functions satisfy the conditions of [Dim18, Lemma 2.3] with , and functions
From [Dim18, Lemmas 2.2 and 2.3] we conclude that
(6.13) |
From (1.20), which we proved in Step 2 above, and (6.1) we see that
Let us fix , and define for the function
(6.14) |
In the remainder of this step we prove that, for fixed , is a continuous function on . Let for be such that . We seek to show that
(6.15) |
In principle, it is possible to use the Fredholm determinant expansion formula in the right side of (6.14) and repeat our work from Step 2 and the first part of Step 3 to prove (6.15). As the computations are quite involved, we instead deduce the continuity of from the continuity of (i.e. ), which we showed in (6.12) and the fact that the arise as limits of distribution functions of the process in [BFS08]. We provide the details below.
Let be the process, defined in [BFS08, Equation (2.5)]. From the work in [BFS08, Section 4] we have that for each , and
(6.16) |
Equation (6.16) shows that for each , and for fixed is non-decreasing in the variables. Let us define the sequences through
As is non-decreasing, we see that for each
(6.17) |
In addition, by subadditivity we have for each
Taking the limit in the last inequality, using (6.16), we get
(6.18) |
Since is continuous from (6.12), and , we may conclude (6.15) by combining (6.17) and (6.18).
Step 4. In this final step we prove the first part of the lemma. From our work in Steps 1 and 3, we know that for each , the function as in (6.14) is a continuous DF on . This and (6.16) show that weakly converge to a random variable with DF as . In particular, we conclude that is a tight sequence of random variables. The latter implies that for all and , we have is a tight sequence of random vectors in . The tightness of , equation (6.16) and the continuity of together imply that is a continuous DF on and converge weakly to a random vector in with DF as .
For , we let denote a copy of and endow it with the Borel -algebra . If is a finite set, and , with , we let and endow with the product -algebra . We also let denote the unique measure on , whose DF is given by
Here, we have placed the subscript to indicate that . Since (6.16) holds for each , and , we conclude that for finite sets , we have
From the Kolmogorov existence theorem, see [Kal97, Theorem 5.16], we conclude that there is a probability space and a real-valued process on that space, whose finite-dimensional distribution functions are given by . This proves the first part of the lemma.
References
- [AAR99] G. Andrews, R. Askey, and R. Roy. Special functions. Cambridge University Press, Cambridge, 1999.
- [BC14] A. Borodin and I. Corwin. Macdonald processes. Probab. Theory Relat. Fields, 158:225–400, 2014.
- [BCS14] A. Borodin, I. Corwin, and T. Sasamoto. From duality to determinants for -TASEP and ASEP. Ann. Probab., 42(6):2314–2382, 2014.
- [BFS08] A. Borodin, P.L. Ferrari, and T. Sasamoto. Transition between Airy1 and Airy2 processes and TASEP fluctuations. Commun. Pure Appl. Math., 61(11):1603–1629, 2008.
- [CLDR10] P. Calabrese, P. Le Doussal, and A. Rosso. Free-energy distribution of the directed polymer at high temperature. EPL (Europhysics Letters), 90:p.20002, 2010.
- [Cor12] I. Corwin. The Kardar-Parisi-Zhang equation and universality class. Random Matrices: Theory Appl., 1, 2012.
- [CQR13] I. Corwin, J. Quastel, and D. Remenik. Continuum statistics of the Airy2 process. Commun. Math. Phys., 317:347–362, 2013.
- [Dim18] E. Dimitrov. KPZ and Airy limits of Hall-Littlewood random plane partitions. Ann. Inst. H. Poincaré Probab. Statist., 54:640–693, 2018.
- [Dim20] E. Dimitrov. Two-point convergence of the stochastic six-vertex model to the Airy process. 2020. Preprint, arXiv:2006.15934.
- [FV15] P. L. Ferrari and B. Vető. Tracy-Widom asymptotics for -TASEP. Ann. Inst. H. Poincaré Probab. Statist., 51:1465–1485, 2015.
- [GT20] J.M.N. Garcia and A. Torrielli. Norms and scalar products for . J. Phys. A Math., 53(14):145401, 2020.
- [Har72] T.E. Harris. Nearest-neighbor Markov interaction processes on multidimensional lattices. Adv. Math., 9:66–89, 1972.
- [HHT15] T. Halpin-Healy and K. Takeuchi. A KPZ cocktail-shaken, not stirred: Toasting 30 years of kinetically roughened surfaces. J. Stat. Phys., 160:794–814, 2015.
- [HO97] G.J. Heckman and E.M. Opdam. Yang’s system of particles and Hecke algebras. Ann. Math., 145(1):139–173, 1997.
- [Hol70] R. Holley. A class of interactions in an infinite particle system. Adv. Math., 5:291–309, 1970.
- [Joh03] K. Johansson. Discrete polynuclear growth and determinantal processes. Comm. Math. Phys., 242:277–329, 2003.
- [Kal97] O. Kallenberg. Foundations of Modern Probability. Springer-Verlag, New York, 1997.
- [Lig72] T. M. Liggett. Existence theorems for infinite particle systems. Trans. Amer. Math. Soc., 165:471–481, 1972.
- [MGP68] C.T. MacDonald, J.H. Gibbs, and A.C. Pipkin. Kinetics of biopolymerization on nucleic acid templates. Biopolymers, 6:1–25, 1968.
- [MQR21] K. Matetski, J. Quastel, and D. Remenik. The KPZ fixed point. Acta Math., 227:115–203, 2021.
- [OQR16] J. Ortmann, J. Quastel, and D. Remenik. Exact formulas for random growth with half-flat initial data. Ann. Appl. Probab., 26(1):507–548, 2016.
- [Pra94] V. Prasolov. Problems and Theorems in Linear Algebra. Amer. Math. Soc., Providence, RI, 1994.
- [PS02] M. Prähofer and H. Spohn. Scale invariance of the PNG droplet and the Airy process. J. Stat. Phys., 108:1071–1106, 2002.
- [QR13] J. Quastel and D. Remenik. Supremum of the Airy2 process minus a parabola on a half line. J. Stat. Phys., 150:442–456, 2013.
- [QR14] J. Quastel and D. Remenik. Airy processes and variational problems. In Topics in percolative and disordered systems, pages 121–171. Springer, New York, 2014.
- [QS15] J. Quastel and H. Spohn. The one-dimensional KPZ equation and its universality class. J. Stat. Phys., 160:965–984, 2015.
- [QS22] J. Quastel and S. Sarkar. Convergence of exclusion processes and the KPZ equation to the KPZ fixed point. J. Amer. Math. Soc., 2022. https://doi.org/10.1090/jams/999.
- [Rob55] H. Robbins. A remark on Stirling’s formula. Amer. Math. Monthly, 62:26–29, 1955.
- [Sas05] T. Sasamoto. Spatial correlations of the D KPZ surface on a flat substrate. J. Phys. A Math. Gen., 38(33):L549, 2005.
- [Spi70] F. Spitzer. Interaction of Markov processes. Adv. Math., 5:246–290, 1970.
- [SS03] E. Stein and R. Shakarchi. Complex analysis. Princeton University Press, Princeton, NJ, 2003.
- [TW96] C.A. Tracy and H. Widom. On orthogonal and symplectic matrix ensembles. Comm. Math. Phys., 177(3):727–754, 1996.