Multipoint distribution of periodic TASEP

By Jinho Baik and Zhipeng Liu

Abstract

The height fluctuations of the models in the KPZ class are expected to converge to a universal process. The spatial process at equal time is known to converge to the Airy process or its variations. However, the temporal process, or more generally the two-dimensional space-time fluctuation field, is less well understood. We consider this question for the periodic TASEP (totally asymmetric simple exclusion process). For a particular initial condition, we evaluate the multitime and multilocation distribution explicitly in terms of a multiple integral involving a Fredholm determinant. We then evaluate the large-time limit in the so-called relaxation time scale.

1. Introduction

The models in the KPZ universality class are expected to have the 1:2:3 scaling for the height fluctuations, spatial correlations, and time correlations as time . This means that the scaled two-dimensional fluctuation field

of the height function , where is the spatial variable and is time, is believed to converge to a universal field which depends only on the initial condition.⁠Footnote1 Here are model-dependent constants. Determining the limiting two-dimensional fluctuation field

1

For some initial conditions, such as the stationary initial condition, one may need to translate the space in the characteristic direction.

is an outstanding question.

By now there are several results for the one-point distribution. The one-point distribution of for fixed is given by random matrix distributions (Tracy-Widom distributions) or their generalizations. The convergence is proved for a quite long list of models including PNG, TASEP, ASEP, -TASEP, random tilings, last passage percolations, directed polymers, the KPZ equation, and so on. See, for example, Reference 1Reference 2Reference 6Reference 26Reference 41, and the review article Reference 11. These models were studied using various integrable methods under standard initial conditions. See also the recent papers Reference 13Reference 37 for general initial conditions.

The spatial one-dimensional process, for fixed , is also well understood. This process is given by the Airy process and its variations. However, the convergence is proved rigorously only for a smaller number of models. It was proved for the determinantal models like PNG, TASEP, last passage percolation,⁠Footnote2 but not yet for other integrable models such as ASEP, -TASEP, finite-temperature directed polymers, and the KPZ equation.

2

See, for example, Reference 3Reference 8Reference 9Reference 10Reference 25Reference 27Reference 34 for special initial conditions. See the recent paper Reference 31 for general initial conditions for TASEP.

The two-dimensional fluctuation field, , on the other hand, is less well understood. The joint distribution is known only for the two-point distribution. In 2015, Johansson Reference 28 considered the zero temperature Brownian semidiscrete directed polymer and computed the limit of the two-point (in time and location) distribution.⁠Footnote3 The limit is obtained in terms of rather complicated series involving the determinants of matrices whose entries contain the Airy kernel. The formula is simplified more recently in terms of a contour integral of a Fredholm determinant in Reference 29 in which the author also extended his work to the directed last passage percolation model with geometric weights. Two other papers studied qualitative behaviors of the temporal correlations. Using a variational problem involving two independent Airy processes, Ferrari and Spohn Reference 20 proved in 2016 the power law of the covariance in the time direction in the large- and small-time limits, and . Here, and denote the scaled time parameters. De Nardis and Le Doussal Reference 14 extended this work further and also, augmented by other physics arguments, computed the similar limits of the two-time distribution when one of the arguments is large. It is yet to be seen if one can deduce these results from the formula of Johansson.

3

There are non-rigorous physics papers for the two-time distribution of directed polymers Reference 16Reference 17Reference 18. However, another physics paper Reference 14 indicates that the formulas in these papers are not correct.

The objective of this paper is to study the two-dimensional fluctuation field of spatially periodic KPZ models. Specifically, we evaluate the multipoint distribution of the periodic TASEP (totally asymmetric simple exclusion process) and compute a large-time limit in a certain critical regime.

We denote by the period and by the number of particles per period. Set , the average density of particles. The periodic TASEP (of period and density ) is defined by the occupation function satisfying the spatial periodicity:

Apart from this condition, the particles follow the usual TASEP rules.

Consider the limit as with fixed . Since the spatial fluctuations of the usual infinite TASEP is , all of the particles in the periodic TASEP are correlated when . We say that the periodic TASEP is in the relaxation time scale if

If , we expect that the system size has negligible effect and, therefore, the system follows the KPZ dynamics. See, for example, Reference 5. On the other hand, if , then the system is basically in a finite system, and hence we expect the stationary dynamics. See, for example, Reference 15. Therefore, in the relaxation time scale, we predict that the KPZ dynamics and the stationary dynamics are both present.

Even though the periodic TASEP is as natural as the infinite TASEP, the one-point distribution was obtained only recently. Over the last two years, in a physics paper Reference 36 and, independently, in mathematics papers Reference 4Reference 30, the authors evaluated the one-point function of the height function in finite time and computed the large-time limit in the relaxation time scale. The one-point function follows the KPZ scaling but the limiting distribution is different from that of the infinite TASEP.⁠Footnote4 This result was obtained for the three initial conditions of periodic step, flat, and stationary. Some earlier related studies can be found in physics papers Reference 15Reference 21Reference 22Reference 23Reference 24Reference 32Reference 33Reference 35, including results on the large deviation and spectral properties of the system.

4

The formulas obtained in Reference 4Reference 30 and Reference 36 are similar, but different. It is yet to be checked that these formulas are the same.

In this paper, we extend the analysis of the papers Reference 4Reference 30 and compute the multipoint (in time and location) distribution of the periodic TASEP with a special initial condition called the periodic step initial condition. Here we allow any number of points unlike the previous work of Johansson on the infinite TASEP. It appears that the periodicity of the model simplifies the algebraic computation compared with the infinite TASEP. In a separate paper we will consider flat and stationary initial conditions. The main results are the following:

(1)

For arbitrary initial conditions, we evaluate finite-time joint distribution functions of the periodic TASEP at multiple points in the space-time coordinates in terms of a multiple integral involving a determinant of size . See Theorem 3.1 and Corollary 3.3.

(2)

For the periodic step initial condition, we simplify the determinant to a Fredholm determinant. See Theorem 4.6 and Corollary 4.7.

(3)

We compute the large-time limit of the multipoint (in the space-time coordinates) distribution in the relaxation time scale for the periodic TASEP with the periodic step initial condition. See Theorem 2.1.

One way of studying the usual infinite TASEP is the following. First, one computes the transition probability using the coordinate Bethe ansatz method. This means that we solve the Kolmogorov forward equation explicitly after replacing it (which contains complicated interactions between the particles) by the free evolution equation with certain boundary conditions. In Reference 40, Schütz obtained the transition probability of the infinite TASEP. Second, one evaluates the marginal or joint distribution by taking a sum of the transition probabilities. It is important that the resulting expression should be suitable for the asymptotic analysis. This is achieved typically by obtaining a Fredholm determinant formula. In Reference 38, Rákos and Schütz rederived the famous finite-time Fredholm determinant formula of Johansson Reference 26 for the one-point distribution in the case of the step initial condition using this procedure. Subsequently, Sasamoto Reference 39 and Borodin, Ferrari, Prähofer, and Sasamoto Reference 9 obtained a Fredholm determinant formula for the joint distribution of multiple points with equal time. This was further extended by Borodin and Ferrari Reference 7 to the points in spatial directions of each other. However, it was not extended to the case when the points are temporal directions of each other. The third step is to analyze the finite-time formula asymptotically using the method of steepest-descent. See Reference 7Reference 9Reference 26Reference 39 and also a more recent paper Reference 31. In the KPZ 1:2:3 scaling limit, the above algebraic formulas give only the spatial process .

We applied the above procedure to the one-point distribution of the periodic TASEP in Reference 4. We obtained a formula for the transition probability, which is a periodic analogue of the formula of Schütz. Using that, we computed the finite-time one-point distribution for an arbitrary initial condition. The distribution was given by an integral of a determinant of size . We then simplified the determinant to a Fredholm determinant for the cases of the step and flat initial conditions. The resulting expression was suitable for the asymptotic analysis. A similar computation for the stationary initial condition was carried out in Reference 30.

In this paper, we extend the analysis of Reference 4Reference 30 to multipoint distributions. For general initial conditions, we evaluate the joint distribution by taking a multiple sum of the transition probabilities obtained in Reference 4. The computation can be reduced to an evaluation of a sum involving only two arbitrary points in the space-time coordinates (with different time coordinates). The main technical result of this paper, presented in Proposition 3.4, is the evaluation of this sum in a compact form. The key point, compared with the infinite TASEP Reference 7Reference 9Reference 31, is that the points do not need to be restricted to the spatial directions.⁠Footnote5 The final formula is suitable for the large-time asymptotic analysis in relaxation time scale.

5

In the large-time limit, we add a certain restriction when the rescaled times are equal. See Theorem 2.1. The outcome of the above computation is that we find the joint distribution in terms of a multiple integral involving a determinant of size . For the periodic step initial condition, we simplify the determinant further to a Fredholm determinant.

If we take the period to infinity while keeping other parameters fixed, the periodic TASEP becomes the infinite TASEP. Moreover, it is easy to check (see Section 8 below) that the joint distributions of the periodic TASEP and the infinite TASEP are equal even for fixed if is large enough compared with the times. Hence, the finite-time joint distribution formula obtained in this paper (Theorem 3.1 and Corollary 3.3) in fact gives a formula of the joint distribution of the infinite TASEP; see the equations Equation 8.7 and Equation 8.8. This formula contains an auxiliary parameter which has no meaning in the infinite TASEP. From this observation, we find that if we take the large-time limit of our formula in the subrelaxation time scale, , then the limit, if it exists, is the joint distribution of the two-dimensional process in Equation 1.2. However, it is not clear at this moment if our formula is suitable for the asymptotic analysis in the subrelaxation time scale; the kernel of the operator in the Fredholm determinant does not seem to converge in the subrelaxation time scale while it converges in the relaxation time scale. The question of computing the limit in the subrelaxation time scale, and hence the multipoint distribution of the infinite TASEP, will be left as a later project.

This paper is organized as follows. We state the limit theorem in Section 2. The finite-time formula for general initial conditions is in Section 3. Its simplification for the periodic step initial condition is obtained in Section 4. In Section 5, we prove Proposition 3.4, the key algebraic computation. The asymptotic analysis of the formula obtained in Section 4 is carried out in Section 6, proving the result in Section 2. We discuss some properties of the limit of the joint distribution in Section 7. In Section 8 we show that the finite-time formulas obtained in Sections 3 and 4 are also valid for infinite TASEP for all large enough .

2. Limit theorem for multipoint distribution

2.1. Limit theorem

Consider the periodic TASEP of period with particles per period. We set , the average particle density. We assume that the particles move to the right. Let be the occupation function of periodic TASEP: if the site is occupied at time , otherwise , and it satisfies the periodicity . We consider the periodic step initial condition defined by

and .

We state the results in terms of the height function

Here and are the unit coordinate vectors in the spatial and time directions, respectively. The height function is defined by

where counts the number of particles jumping through the bond from to during the time interval . The periodicity implies that

for integers .

See Figure 1 for the evolution of the density profile and Figure 2 for the limiting height function. Note that the step initial condition Equation 2.1 generates shocks.⁠Footnote6. By solving the Burgers’ equation in a periodic domain, one could derive the explicit formulas of the density profile, the limiting height function and the shock location. These computations were done in Reference 5.

6

These shocks are generated when faster particles from the lower density region enter the higher density region and are forced to slow down. See, for example, Reference 12Reference 19 for the study of similar behaviors in infinite TASEP.

We represent the space-time position in new coordinates. Let

be a vector parallel to the characteristic directions. If we represent in terms of and , then

Consider the region

See Figure 3. Due to the periodicity, the height function in determines the height function in the whole space-time plane.

The following theorem is the main asymptotic result. We take the limit as follows. We take in such a way that the average density is fixed, or more generally stays in a compact subset of the interval . We consider distinct points in the space-time plane such that their temporal coordinate and they satisfy the relaxation time scale . The relative distances of the coordinates are scaled as in the 1:2:3 KPZ prediction: , , and the height at each point is scaled by .

Theorem 2.1 (Limit of multipoint joint distribution for periodic TASEP).

Fix two constants and satisfying . Let be a sequence of integers such that for all sufficiently large . Consider the periodic TASEP of period and average particle density . Assume the periodic step initial condition Equation 2.1. Let be a positive integer. Fix points , , in the region

Assume that

Let be points⁠Footnote7 in the region shown in Figure 3, where and , with

7

Since should have an integer value for its spatial coordinate, to be precise, we need to take the integer part of for the spatial coordinate. This small distinction does not change the result since the limits are uniform in the parameters . Therefore, we suppress the integer value notation throughout this paper.

Then, for arbitrary fixed ,

where the function is defined in Equation 2.15. The convergence is locally uniform in , and . If for some , then Equation 2.11 still holds if we assume that .

Remark 2.2.

Suppose that we have arbitrary distinct points in . Then we may rearrange them so that . If are all different, we can apply the above theorem since the result holds for arbitrarily ordered . If some of the are equal, then we may rearrange the points further so that are ordered with those , and use the theorem if are distinct. The only case which is not covered by the above theorem is when some of the are equal and the corresponding are also equal.

The case when was essentially obtained in our previous paper Reference 4 (and also Reference 36). In that paper, we considered the location of a tagged particle instead of the height function, but it is straightforward to translate the result to the height function.

Remark 2.3.

We will check that is periodic with respect to each of the space coordinates in Subsection 2.2. By this spatial periodicity, we can remove the restrictions in the above theorem.

Remark 2.4.

Since we expect the KPZ dynamics in the subrelaxation scale , we expect that the limit of the above result should give rise to a result for the usual infinite TASEP. Concretely, we expect that the limit

exists and it is the limit of the multitime, multilocation joint distribution of the height function of the usual TASEP with step initial condition,

In particular, we expect that when , Equation 2.12 is , the Tracy-Widom GUE distribution; we expect that when , Equation 2.12 is equal to the corresponding joint distribution of the Airy process Reference 34 ; and when , Equation 2.12 is expected to match the two-time distribution obtained by Johansson Reference 28Reference 29. See also Section 8.

2.2. Formula of the limit of the joint distribution

2.2.1. Definition of

Definition 2.5.

Fix a positive integer . Let for each , where and

Define, for ,

where and the contours are nested circles in the complex plane satisfying . Set , , and . The function is defined by Equation 2.21 and it depends on and but not on . The function depends on all , , and , and it is given by the Fredholm determinant defined in Equation 2.38.

The functions in the above definition satisfy the following properties. The proofs of (P1), (P3), and (P4) are scattered in this section while (P2) is proved later in Lemma 7.1.

(P1)

For each , is a meromorphic function of in the disk . It has simple poles at for .

(P2)

For each , is analytic in the punctured disk .

(P3)

For each , does not change if we replace by . Therefore, is periodic, with period , in the parameter for each .

(P4)

If , the function is still well-defined for .

Remark 2.6.

It is not easy to check directly from the formula that defines a joint distribution function. Nonetheless, we may check them indirectly. From the fact that is a limit of a sequence of joint distribution functions, and is a non-decreasing function of for each . It also follows from the fact that the joint distribution can be majorized by a marginal distribution that converges to if any coordinate since it was shown in (4.10) of Reference 4 that the case is indeed a distribution function; see Lemma 7.5 below. The most difficult property to prove is the consistency which should satisfy as a coordinate . We prove this property in Section 7 by finding a probabilistic interpretation of the formula of when the -contours are not nested; see Theorem 7.3 and Proposition 7.4.

2.2.2. Definition of

Let be the principal branch of the logarithm function with cut . Let be the polylogarithm function defined by

It has an analytic continuation using the formula

Set

For , set

where the integral contours are the vertical lines and with constants and satisfying . The equality of the double integral and the series is easy to check (see (9.27)–(9.30) in Reference 4 for a similar calculation). Note that . When , we can also check that

Definition 2.7.

Define

where we set .

Since are analytic inside the unit circle, it is clear from the definition that satisfies property (P1) in Subsubsection 2.2.1.

2.2.3. Definition of

The function is given by a Fredholm determinant. Before we describe the operator and the space, we first introduce a few functions.

For , define the function

and

The integration contour lies in the half-plane , and is given by the union of the interval on the real axis and the line segment from to . Since on the integration contour, is well-defined. Thus, we find that the integrals are well-defined using as .

Observe the symmetry

We also have

This identity can be obtained by the power series expansion and using the fact that for with ; see (4.8) of Reference 4. From Equation 2.24 and Equation 2.25, we find that

for any fixed satisfying .

Let , , and be the parameters in Definition 2.5. We set

for , where we set .

Now we describe the space and the operators. For a non-zero complex number , consider the roots of the equation . The roots are on the contour . It is easy to check that if , the contour consists of two disjoint components, one in and the other in . See Figure 4. The asymptotes of the contours are the straight lines of slope . For , we define the discrete sets

For distinct complex numbers satisfying , define the sets

and

See Figure 5. Now we define two operators

by kernels as follows. If

for some , then we set

Similarly, if

for some , then we set

Here the delta function if or otherwise. We also set so that

and the functions and are defined by

Definition 2.8.

Define

for , where and are distinct.

In this definition, we temporarily assumed that are distinct in order to ensure that the term in the denominators in Equation 2.33 and Equation 2.35 does not vanish. However, as we stated in (P2) in Subsubsection 2.2.1, is still well-defined when the are equal. See Lemma 7.1.

The definition of and implies that as along and as along . Hence, due to the cubic term in Equation 2.27, super exponentially as on the set if . Hence, using the property Equation 2.26 of , we see that the kernels decay super-exponentially fast as on the spaces. Therefore, the Fredholm determinant is well-defined if .

We now check property (P4). If , the exponent of has no cubic term . The quadratic term contributes to since for , and hence . On the other hand, the linear term in the exponent of has a negative real part if . Hence, if and , then exponentially as along and hence the kernel decays exponentially fast as on the spaces. Therefore, the Fredholm determinant is still well-defined if and . This proves (P4).

2.3. Matrix kernel formula of and

Due to the delta functions, only when

for some integer , and similarly only when

for some integer . Thus, if we represent the kernels as matrix kernels, then they have block structures.

For example, consider the case when . Let us use and to represent variables in and , respectively:

The matrix kernels are given by

and

where the empty entries are zeros and the function is given below. When is odd, the structure is similar. On the other hand, when is even, consists only of blocks and contains an additional non-zero block at the bottom right corner.

We now define . For , writing

we define

This means that

and so on. The term is defined by the entry of Equation 2.45 with , where we set . The term is defined by the entry of Equation 2.45 with , where we set .

2.4. Series formulas for

We present two series formulas for the function . The first one Equation 2.52 is the series expansion of Fredholm determinant using the block structure of the matrix kernel. The second formula Equation 2.53 is obtained after evaluating the finite determinants in Equation 2.52 explicitly.

To simplify formulas, we introduce the following notation.

Definition 2.9 (Notational conventions).

For complex vectors and , we set

For a function of single variable, we write

We also use the notation

for finite sets and .

The next lemma follows from a general result whose proof is given in Subsection 4.3 below.

Lemma 2.10 (Series formulas for ).

We have

with for , where can be expressed in the following two ways.

(i)

We have, for ,

where , with , , and where

and

Here, we set .

(ii)

We also have

with

where

Recall Equation 2.22Equation 2.23, and Equation 2.27 for the definition of and .

Property (P3) in Subsubsection 2.2.1 follows easily from Equation 2.56. Note that only appears in the factor for or . If we replace by , then and are changed by and if , or by and if . But has the same number of components as for each . Therefore does not change. We can also check (P3) from the original Fredholm determinant formula.

The analyticity property (P2) is proved in Lemma 7.1 later using the series formula.

3. Joint distribution function for the general initial condition

We obtain the limit theorem of the previous section from a finite-time formula of the joint distribution. In this section, we describe a formula of the finite-time joint distribution for an arbitrary initial condition. We simplify the formula further in the next section for the case of the periodic step initial condition.

We state the results in terms of particle locations instead of the height function used in the previous section. It is easy to convert one to another; see Equation 6.5. The particle locations are denoted by , where

Due to the periodicity of the system, we have for all integers .

The periodic TASEP can be described if we keep track of consecutive particles, say . If we focus only on these particles, they follow the usual TASEP rules plus the extra condition that for all . Define the configuration space

We call the process of the particles TASEP in . We use the same notation , , to denote the particle locations in the TASEP in . We state the result for the TASEP in first and then for the periodic TASEP as a corollary.

For , consider the polynomial of degree given by

Denote the set of the roots by

The roots are on the level set . It is straightforward to check the following properties of the level set. Set

where, as before, . The level set becomes larger as increases; see Figure 6. If , the level set consists of two closed contours, one in enclosing the point and the other in enclosing the point . When , the level set has a self-intersection at . If , then the level set is a connected closed contour. Now consider the set of roots . Note that if , then . It is also easy to check that if a non-zero satisfies , then the roots of are all simple. On the other hand, if , then there is a double root at and the remaining roots are simple. For the results in this section, we take to be any non-zero complex number. But in the next section, we restrict .

Theorem 3.1 (Joint distribution of TASEP in for the general initial condition).

Consider the TASEP in . Let and assume that . Fix a positive integer . Let be distinct points in . Assume that . Let for . Then

where the contours are nested circles in the complex plane satisfying . Here , , , and . The functions in the integrand are

and

with

where we set .

Remark 3.2.

The limiting joint distribution in the previous section was not defined for all parameters: when , we need to put the restriction . See property (P4) in Subsubsection 2.2.1. The finite-time joint distribution does not require such restrictions. The sums in the entries of the determinant are over finite sets, and hence there is no issue with the convergence. Therefore, the right-hand side of Equation 3.6 is well-defined for all real numbers and integers and .

Corollary 3.3 (Joint distribution of periodic TASEP for the general initial condition).

Consider the periodic TASEP with a general initial condition determined by and its periodic translations; for all and . Then Equation 3.6 holds for all without the restriction that .

Proof.

The particles in the periodic TASEP satisfy for every integer . Hence if is not between and , we may translate it. This amounts to changing to and to for some integer . Hence it is enough to show that the right-hand side of Equation 3.6 is invariant under these changes. Under these changes, the term is multiplied by the factor if and by if . On the other hand, produces the multiplicative factor which is by Equation 3.4. Taking this factor outside the determinant Equation 3.8, we cancels out the factor from . Similarly produces a factor which cancel out if .

Before we prove the theorem, let us comment on the analytic property of the integrand in the formula Equation 3.6. The function is clearly analytic in each . Consider the function . Note that

Hence, if is an analytic function of in and

then

for any and such that all roots of lie in the region . Note that is an entire function of for each . Since we may take arbitrarily large and arbitrarily small and positive, the right-hand side of Equation 3.11 defines an analytic function of . Now the entries of the determinant in Equation 3.8 are of the form

for a function which is analytic in each variable in as long as for all . The last condition is due to the factor in the denominator. Note that if , then . Hence by using Equation 3.11 times, each entry of Equation 3.8, and hence , is an analytic function of each in the region where all are distinct.

When , the product in Equation 3.7 is set to be and the formula Equation 3.6 in this case was obtained in Proposition 6.1 in Reference 4. For , as we mentioned in the Introduction, we prove Equation 3.6 by taking a multiple sum of the transition probability. The main new technical result is a summation formula and we summarize it in Proposition 3.4 below.

The transition probability was obtained in Proposition 5.1 of Reference 4. Denote by the transition probability from to in time ,

where the integral is over any simple closed contour in which contains inside. The integrand is an analytic function of for by using Equation 3.11.

Proof of Theorem 3.1.

It is enough to consider . It is also sufficient to consider the case when the times are distinct, , because both sides of Equation 3.6 are continuous functions of . Note that Equation 3.8 involves only finite sums.

Denoting by the configuration of the particles at time , the joint distribution function on the left-hand side of Equation 3.6 is equal to

Applying the Cauchy-Binet formula to Equation 3.13, we have

where, for ,

and

Here the factor in the denominator comes from the Cauchy-Binet formula; it will eventually disappear since we will apply the Cauchy-Binet identity backward again at the end of the proof.

We insert Equation 3.15 into Equation 3.14 and interchange the order of the sums and the integrals. Assuming that the series converges absolutely so that the interchange is possible, the joint distribution is equal to

where and

Here we set

for a pair of complex vectors and . Let us now show that it is possible to exchange the sums and integrals if we take the -contours properly. We first consider the convergence of Equation 3.20 and the sum in Equation 3.19. Note that, shifting the summation variable to ,

The right-hand side of Equation 3.20 is the sum of the above formula over . Hence Equation 3.20 converges absolutely and the convergence is uniform for if is in a compact subset of . Similarly, the sum of in Equation 3.19 converges if . Therefore, Equation 3.19 converges absolutely if the intermediate variables satisfy

We now show that it is possible to choose the contours of so that Equation 3.22 is achieved. Since , satisfies the equation . Hence as . Therefore, if we take the contours where and are large enough (where ), then Equation 3.22 is satisfied. Thus, Equation 3.20 and the sum in Equation 3.19 converge absolutely. It is easy to see that the convergences are uniform. Hence we can exchange the sums and integrals, and therefore, the joint distribution is indeed given by Equation 3.18 if we take the contours of to be large nested circles.

We simplify Equation 3.18. The terms are evaluated in Proposition 3.4 below. Note that since the -contours are the large nested circles, we have Equation 3.22, and hence the assumptions in Proposition 3.4 are satisfied. On the other hand, the sum of in Equation 3.18 was computed in Reference 4. Lemma 6.1 in Reference 4 implies that for ,

Hence, from the geometric series, for ,

if . The last condition is satisfied for . We thus find that Equation 3.19 is equal to an explicit factor times a product of Cauchy determinants times a Vandermonde determinant. By using the Cauchy-Binet identity times, we obtain Equation 3.6 assuming that the -contours are large nested circles.

Finally, using the analyticity of the integrand on the right-hand side of Equation 3.6, which was discussed before the start of this proof, we can deform the contours of to any nested circles, not necessarily large circles. This completes the proof.

The main technical part of this section is the following summation formula. We prove it in Section 5.

Proposition 3.4.

Let and be two non-zero complex numbers satisfying . Let and . Suppose that . Consider defined in Equation 3.20. Then for any and integer ,

4. Periodic step initial condition

We now assume the following periodic step condition:

In the previous section, we obtained a formula for general initial conditions. In this section, we find a simpler formula for the periodic step initial condition which is suitable for the asymptotic analysis. We express as a Fredholm determinant times a simple factor. The result is described in terms of two functions and . We first define them and then state the result.

Throughout this section, we fix a positive integer , and fix parameters , , as in the previous section.

4.1. Definitions

Recall the function for complex in Equation 3.3 and the set of its roots

in Equation 3.4. Set

as in Equation 3.5. We discussed in the previous section that if , then the contour consists of two closed contours, one in enclosing the point and the other in enclosing the point . Now, for , set

It is not difficult to check that

See the left picture in Figure 6 in Section 3. (Note that if , then the roots are with multiplicity and with multiplicity .) From the definitions, we have

In Theorem 3.1, we took the contours of as nested circles of arbitrary sizes. In this section, we assume that the circles satisfy

Hence and are all well-defined.

We define two functions and of , both of which depend on the parameters . The first one is the following. Recall the notational convention introduced in Definition 2.9. For example, .

Definition 4.1.

Define

where

for , and .

It is easy to see that all terms in other than are analytic for within the disk . Hence is analytic in the disk except for the simple poles when , .

We now define . It is given by a Fredholm determinant. Set

Define

Also, set

Define, for ,

Note that . Set

When , we define and hence .

Define two sets

and

See Figure 7. We define two operators

by kernels. If and for some , we set

Similarly, if and for some , we set