Random Motzkin paths near boundary
Limits of Random Motzkin paths with KPZ related asymptotics
Abstract.
We study Motzkin paths of length with general weights on the edges and end points. We investigate the limit behavior of the initial and final segments of the random Motzkin path viewed as a pair of processes starting from each of the two end points as becomes large. We then study macroscopic limits of the resulting processes, where in two different regimes we obtain Markov processes that appeared in the description of the stationary measure for the KPZ equation on the half line and of conjectural stationary measure of the hypothetical KPZ fixed point on the half line. Our results rely on the behavior of the Al-Salam–Chihara polynomials in the neighbourhood of the upper end of their orthogonality interval and on the limiting properties of the -Pochhammer and -Gamma functions as .
1. Introduction and main results
In this paper we study Markov processes that arise as limits of random Motzkin paths with random endpoints. A Motzkin path of length is a sequence of steps on the integer lattice that starts at point with the initial altitude and ends at point at the final altitude for some non-negative integers . Steps can be upward, downward, or horizontal, along the vectors , and , respectively, and the path cannot fall below the horizontal axis. The path is uniquely determined by the sequence of altitudes with , .
A random Motzkin path can be generated by assigning a discrete probability measure on the set of all Motzkin paths and choosing at random according to this probability measure. We will write and will use the notation if we need to explicitly indicate its dependence on the parameter . In our main results we are interested in the assignment of probability which depends on two boundary parameters and two parameters and that determine the edge weights. As the probability of selecting a Motzkin path we take the following expression:
(1.1) |
where denotes the usual -number, and is the number of horizontal steps, and is the normalizing constant. This weighting of the Motzkin paths was inspired by a formula in (Barraquand and Le Doussal,, 2023, Section 2.3), where . A more general setup is discussed in Section 2.
The above setup differs from the most commonly studied random Motzkin paths chosen uniformly from all Motzkin paths which start and end at altitude 0. Such random Motzkin paths correspond to random walks conditioned on staying non-negative and returning to at time ; it is well known that their asymptotic fluctuations are described by the Brownian excursion Kaigh, (1976) and their behavior near boundaries is described by an explicit Markov chain, see Keener, (1992). It turns out that new phenomena and new asymptotic fluctuations arise in the presence of boundary parameters , see Bryc and Wang, (2023), Bryc and Wang, (2024). Our goal is to extend Bryc and Wang, (2023) to allow more general weights that depend on parameter , and to study the boundary limit Markov chain in two different asymptotic regimes. As limits in the two asymptotic regimes we recover Markov processes that appeared in the description given in Bryc and Kuznetsov, (2022) of the non-Gaussian term of the stationary measure of the KPZ equation on the half-line, see Barraquand and Le Doussal, (2022), Barraquand and Corwin, (2023), and the non-Gaussian term in the conjectural stationary measure of the hypothetical KPZ fixed point on the half line in Barraquand and Le Doussal, (2022).
We use the following standard notation for the -Pochhammer symbols:
Our first main result is the following limit theorem for the boundaries of the random Motzkin path. Let be a sequence of the initial altitudes of a random Motzkin path of length , appended with for , and let be a sequence of the final altitudes, , , appended with for .
Theorem 1.1.
Suppose that and . Then
as , where on the left hand side we have random Motzkin paths with respect to measure (1.1) and on the right-hand side we have two independent Markov chains , with the same transition probabilities on given by
for and with the initial laws
(1.2) |
where and
with , ; the normalizing constants are
We note that if then , recovering transition probabilities in (Bryc and Wang,, 2023, Theorem 1.1), who use the parameter that is twice our . It is natural to expect that as in (Bryc and Wang,, 2023, Theorem 1.6), there is a version of this result that holds also for . However, this is beyond the scope of this paper.
Theorem 1.1 will be deduced from a more general Theorem 2.4. The proof appears in Section 4 and relies on properties of the Al-Salam–Chihara polynomials, which we discuss in Section 3.
The next two theorems give macroscopic continuous-time limits of the family of Markov processes . In the statements of Theorems 1.2 and 1.3 below, denotes convergence of finite dimensional distributions.
In our first result, we take at an appropriate rate but keep fixed. Then the normalized Markov process converges to the Bessel process, which for appeared in the description of the non-Gaussian term in the conjectural stationary measure for the hypothetical KPZ fixed point on the half-line in (Bryc and Kuznetsov,, 2022, Theorem 2.6 ).
Theorem 1.2.
Fix , and . Let be a Markov process from Theorem 1.1 with the initial law that depends on through . Then
(1.3) |
where is the 3-dimensional Bessel process with transition probabilities
(1.4) |
where , , is the transition kernel of the Brownian motion killed at hitting zero, i.e.
(1.5) |
and with the initial distribution
(1.6) |
In our second result, we take both and at appropriate rates. Then, under appropriate centering and normalization, the Markov process converges to the Markov process on , which for appeared in the description of the non-Gaussian term in the representation of the stationary measure for the KPZ equation on the half line given in (Bryc and Kuznetsov,, 2022, Theorem 2.3). (This is only a one-parameter subset of the two-parameter family of such measures conjectured in Barraquand and Le Doussal, (2022) and proved rigorously in Barraquand and Corwin, (2023).) In the statement, is a modified Bessel K function (Macdonald function) of imaginary index and positive argument , see e.g. (Olver et al.,, 2023, 10.32.E9).
Theorem 1.3.
The paper is organized as follows. In Section 2 we introduce random Motzkin paths with general weights, we prove matrix representation, integral representation, and a general boundary limit theorem. In Section 3 we recall some properties of the Al–Salam–Chihara polynomials, and we prove new asymptotic results for their behavior near the boundary of the interval of orthogonality. In Section 4 we use these results to prove Theorem 1.1 and two local limit theorems, which we use to complete proofs of Theorems 1.2 and 1.3. In Section 5 we discuss properties of the -Pochhammer and -Gamma functions and derive the limits and bounds needed in our proofs.
2. General limit theorem for random Motzkin paths at the boundary
In this section we introduce a more general setting and prove a version of Theorem 1.1 for more general random Motzkin paths.
Recall that a Motzkin path of length is a sequence of lattice points such that . An edge is called a (-th) step, and only three types of steps are allowed: upward steps, downward steps, and horizontal steps. The edge is an upward step if , a downward step if , and a horizontal step if , see, e.g., Flajolet and Sedgewick, (2009, Definition V.4, p. 319) or Viennot, (1985). Each such path can be identified with a sequence of non-negative integers that specify the starting point and consecutive values along the vertical axis at step . We shall write with for such a sequence and refer to as a Motzkin path. By we denote the family of all Motzkin paths of length with the initial altitude and the final altitude . We also refer to and as the boundary/end points of the path.
We are interested in probabilistic properties of random Motzkin paths. To introduce a probability measure on the set of the Motzkin paths, we assign weights to the edges and to the end-points of a Motzkin path. The weights for the edges arise multiplicatively from three sequences of positive real numbers. For a path we define its (edge) weight
(2.1) |
where
That is, the edge weight is multiplicative in the edges, we take , and as the weights of the upward steps, horizontal steps and downward steps, and the weight of a step depends on the altitude of the left-end of an edge. Note that the value of does not contribute to . Since is a finite set, the normalization constants
are well defined for all .
In addition to the weights of the edges, we also assign weights to the two boundary points. To this end we choose two additional non-negative sequences and such that
(2.2) |
With finite normalizing constant (2.2), the countable set becomes a probability space with the discrete probability measure
(2.3) |
For an illustration, see Figure 1. We now consider a random vector sampled from , and extend it to an infinite process denoted by
(2.4) |
Similarly, we also introduce the infinite reversed process by
(2.5) |
We are interested in the limit processes for both and as . In these limits, we do not scale the sequences nor the locations.
Recall that weak convergence of discrete-time processes means convergence of finite-dimensional distributions (Billingsley,, 1999). For integer-valued random variables, the latter follows from convergence of probability generating functions. We will therefore fix , and and the goal is to determine two discrete time processes and such that
(2.6) |
Indeed, the above expressions uniquely determine the corresponding probability generating functions for small enough arguments. For example,
with and .
2.1. Matrix representation and integral representation
We will need a convenient representation for the left hand side of (2.6). We introduce a tri-diagonal matrix
and infinite column vectors
We use the following matrix representation for the left hand side of (2.6).
Lemma 2.1 (matrix ansatz).
For every such that , we have
(2.7) |
where
(2.8) |
is the normalization constant.
Proof.
The key step in our proof of boundary limit theorem is an integral representation for the left hand side of (2.6) based on the orthogonality measure of the orthogonal polynomials determined by the edge weights of Motzkin paths.
Following Viennot, (1985) and Flajolet and Sedgewick, (2009) with the sequences of edge weights for the Motzkin paths that appeared in (2.1) we now associate real polynomials , defined by the three step recurrence
(2.11) |
(With the usual conventions that , and , the recursion determines polynomials uniquely, with .)
By Favard’s theorem (Ismail,, 2009), polynomials are orthogonal with respect to a probability measure on the real line, which we assume to be compactly supported, and thus unique. It is well known that the norm is given by the formula
(2.12) |
We need some conditions on the weights of the end points in (2.3).
Assumption 2.1.
We assume that
-
()
The series
(2.13) converge absolutely on the support of for all .
-
()
The function is integrable with respect to the measure , and
(2.14)
Consider the following two infinite column vectors
where
(2.15) |
Note that with , polynomials satisfy the three step recurrence
In particular,
(2.16) |
The left hand side of the expression (2.7) has the following integral representation.
Lemma 2.2.
Proof.
We use expression (2.9) with of the form . Proceeding somewhat formally, we write the identity matrix as the integral, and rewrite the product of matrices in (2.7) as
Therefore, the right hand side of (2.7) becomes
and the result follows from the symmetry of the dot product, . (Assumption (2.14) is responsible for Fubini’s theorem, which is used to switch the integral with the infinite sums in the last expression. This avoids formal integration of the product of infinite matrices.) ∎
We will need the following elementary lemma. (See also (Bryc and Wang,, 2023, Lemma A.1).)
Lemma 2.3.
Let be a probability measure with with . Let be a real function which is bounded on and left-continuous at .
Then
(2.18) |
Proof.
Let be a random variable with the law . First we will derive some estimates on the moments of and . Fix such that . Since , we have . And since we have
The inequalities imply that the term converges to zero as , therefore for large enough . Next, we use the inequality , which is valid for all and , and obtain
which gives us an upper bound , valid for large enough.
Now we are ready to establish (2.18). By taking instead of in (2.18), without loss of generality, we can assume that (clearly, remains left-continuous at ). Now we estimate, for large enough
where we used the estimates:
for the first term and
for the second term. Since converges to zero when , we obtain
Taking the limit , gives us the desired result (2.18) in the case . ∎
2.2. Limit theorem
The following result is a version of (Bryc and Wang,, 2023, Theorem 1.1) for non-constant weights of edges. It describes the joint limiting behavior of processes and introduced in (2.4) and (2.5).
Theorem 2.4.
Suppose that , the support of the orthogonality measure of the polynomials (2.11), satisfies with and that and from Assumption 2.1 hold. Then , and
as , where and are independent Markov chains with the same transition probabilities
(2.19) |
with , and with the initial laws given by
(2.20) |
where , are the normalizing constants and with given in (2.15).
Proof.
We first verify that transition probabilities (2.19) and (2.24) are well defined, i.e. , . Clearly, . For , the polynomial has leading term , thus as . If were negative, this would imply that has a zero outside of the support of the measure of orthogonality, which is impossible, see e.g. Ismail, (2009). The case is also impossible, since the interlacing property of the zeroes of orthogonal polynomials would imply that has a zero in the interval .
To determine the limit as of the left hand side of (2.6), we re-write the right hand side of (2.17) as
Using Lemma 2.3, each of the factors above converges as and we get
(2.21) |
We will show that the latter expressions matches the product on the right hand side of (2.6).
To this end it suffices to prove
(2.22) |
where , as well as
(2.23) |
where , and
(2.24) |
Note that the processes and have the same transition probabilities, since
We prove only (2.22) since the proof of (2.23) follows along the same lines. We use induction with respect to .
The case of is immediate since the left hand side of (2.22) is . For we first note that the left hand side of (2.22) can be written as
where , . Consequently, by induction assumption for we get
Formula (2.21) shows that the limit processes and are independent. This ends the proof. ∎
For proofs of local limit theorems we need to determine the limit of as for suitably chosen sequences and . The following formula is useful in computing such limits.
Proposition 2.5.
For a Markov process , we have
(2.25) |
(This resembles (Karlin and McGregor,, 1959, p. 67).)
3. Properties of the Al-Salam–Chihara polynomials
The general theory developed in Section 2 can be advanced further when it is specialized to the setting considered in Section 1. To proceed, we need to recall the definition and properties of the Al-Salam–Chihara polynomials.
We will be working with Al-Salam-Chihara polynomials in real variable , defined by the three-step recursion
(3.1) |
where parameters are either real or complex conjugate, , and . As usual, we initialize the recursion with and .
It is known, see Ismail, (2009); Koekoek et al., (2010) or Koekoek and Swarttouw, (1994), that if , then the polynomials are orthogonal with respect to the unique probability density supported on and given by
(3.2) |
It is also known that the generating function of is well defined for complex , and is given by the infinite product:
(3.3) |
Invoking the -binomial theorem, see e.g. (Gasper and Rahman,, 2004, (1.3.2)), the latter implies the following formula, which can also be recalculated from (Berg and Ismail,, 1996, p. 50) or (Ismail,, 2009, (15.1.12)):
(3.4) |
In particular, at the boundary of the interval of orthogonality, we have
(3.5) |
3.1. Maximum of the Al-Salam–Chihara polynomials
Chebyshev polynomials and on the interval are bounded in absolute value by their value at . Similar result holds for the -Hermite polynomials, see (Ismail,, 2009, Theorem 13.1.2). For our proofs we need to extend this property to the Al-Salam–Chihara polynomials.
Proposition 3.1.
Let , and be real or a complex conjugate pair such that and . Then for all we have
(3.6) |
Remark 3.1.
Corollary 3.1.
Let , , . Denote and . Then for all bound (3.6) holds.
Proof.
Indeed, and ∎
3.2. Pointwise asymptotics near the end of the interval of orthogonality
We will need pointwise asymptotics for the Al-Salam–Chihara polynomials at the upper endpoint of the orthogonality interval. Such pointwise limits have been studied for orthogonal polynomials both on finite intervals and on unbounded domains Aptekarev, (1993); Baik et al., (2003); Deift et al., (2001, 1999); Ismail and Li, (2013); Ismail, (1986, 2005); Ismail and Wilson, (1982); Ismail et al., (2022); Kuijlaars et al., (2004); Lubinsky, (2020); Nevai, (1984). In particular, (Aptekarev,, 1993, Theorem 1) gives general conditions on the orthogonality measure and on the three-step recursion of the orthonormal polynomials which imply
(3.7) |
uniformly on compact subsets of complex plane, where is the Bessel function.
Lubinsky (Lubinsky,, 2020, Theorem 1.1) proves a version of (3.7) which can be re-written as
(3.8) |
uniformly on compact subsets of complex plane. For our density function (3.2) the assumptions in (Lubinsky,, 2020, Theorem 1.1) are satisfied with . Indeed, one can re-write (3.2) as
(3.9) |
Note that for , the right hand side of (3.8) simplifies to and that the limit under normalization (3.7) is discussed in (Lubinsky,, 2020, Theorem 1.3).
Pointwise asymptotics for the Al-Salam–Chihara polynomials at the endpoints of the interval of orthogonality follows from the above results. However, our proof is quite direct, and the approach extends to the case of varying orthogonality measures, so we include it for completeness. Our asymptotic result for the Al-Salam-Chihara polynomials with varying orthogonality measure as seem to be new and does not follow from Levin and Lubinsky, (2020).
The following asymptotics holds for fixed .
Theorem 3.2.
Fix real or complex conjugate with , and . Let be the Al-Salam–Chihara polynomials (3.1). If , then
(3.10) |
Remark 3.2.
We will also need the case , with
(3.11) |
We also note the following extension of bound (3.6): there is a constant such that for all and we have
(3.12) |
Proof of Theorem 3.2.
The use of in (3.10) is solely to avoid notation clash when we use it in the proof of Theorem 4.2, so for the proof, we replace by .
The series (3.3) converges for all , thus we can write
(3.13) |
where and the integration is done in the counter-clockwise direction. We fix now , so that and we note that the integrand
has two simple poles at inside the disk . By the Cauchy Residue Theorem we can write
(3.14) |
where is in the interval .
Since under our assumptions with is such that , we see that
Consequently, returning to (3.14) we get
∎
Proof of Remark 3.2.
The following asymptotics holds for . The statement is somewhat cumbersome as we will need to apply this result to that vary with .
Theorem 3.3.
If are real in , or a complex conjugate pair with , then with , and , we have
(3.15) |
Proof.
The proof relies on two technical estimates that we will prove in Section 5. In the proof, for simplicity of notation we suppress the dependence of on . Fix and let , , where .
In view of (3.13) with , we can write
(3.16) |
where we are integrating counterclockwise along a circle of radius . Next we change the variable of integration , so that , and obtain
(3.17) |
From (3.17), in view of (5.2), we get
(3.18) |
where
(3.19) |
We now write . In order to use the dominated convergence theorem, we verify that for every (recall that depends on ) there are constants such that for all and all real we have
(3.20) |
To verify (3.20) we write , where
From (5.24) we see that there exist constants such that for all real . From (5.26) we see that . With
we see that
Thus (3.20) holds.
We can now apply the dominated convergence theorem and pass to the limit inside the integral (3.18). We fix and compute by computing the limit for the factors in (3.19).
Clearly, as and by (5.3) we get
Since , from (5.1) we get
Finally, since
we get
Putting all the factors together,
Thus by the dominated convergence theorem,
This completes the proof by the Mellin-Barnes type formula for the Bessel K function:
where , . (This is (Olver et al.,, 2023, 10.32.13) [10.32.13] https://dlmf.nist.gov/10.32.E13 with , , and shifted variable of integration .) ∎
4. Motzkin paths with -number weights
We now return to the setting from Section 1, providing additional details and further results for the case of edge weights , , .
We note that with the above choice of the edge weights, probability measure (1.1) is the same as (2.3) with the boundary weights and . Indeed, we note that for an upward step we have , for a horizontal step we have and for a downward step we have . So formula (2.1) gives and thus (2.3) gives (1.1).
Lemma 4.1.
Suppose . Polynomials are orthogonal with respect to the density supported on the interval with
(4.3) |
Moreover,
(4.4) |
where
(4.5) |
We note that are the solutions of the system of equations
(4.6) |
Proof.
To determine the orthogonality measure as well as the sequence , we establish a connection between polynomials and monic Al-Salam-Chihara polynomials , defined by the three step recursion (3.1) with complex conjugate parameters (4.5).
It is straightforward to verify that if satisfy recursion (4.1) then polynomials
satisfy recursion (3.1) with parameters (4.5). Conversely, with , we have
(4.7) |
Therefore, with defined by (3.2) (recall, that is supported on ), polynomials are orthogonal with respect to the weight function (probability density function)
(4.8) |
which is supported on the interval given by (4.3). Combining (3.5) and (4.7) we see that
(4.9) |
With the above preparations, we can now prove Theorem 1.1.
Proof of Theorem 1.1.
We apply Theorem 2.4. With and , conditions and from Assumption 2.1 hold by (4.7), (4.2) and (3.12). The assumptions on the support of the orthogonality measure follow from Lemma 4.1. The probabilities are given by (4.4), so the transition matrix is just a recalculation of (2.19). Recalling that , see (4.2), the initial laws (1.2) arise from (2.20). Finally, the formulas for the normalization constants are calculated from the relation (4.7) by applying (3.3) with and . ∎
4.1. Local limit theorems
We will deduce Theorem 1.2 and Theorem 1.3 from two local limit theorems. These theorems use (2.25) to study convergence of the transition probabilities for the Markov process from Theorem 1.1 under the 1:2 scaling of space and time.
We have the following local limit theorem for fixed .
Theorem 4.2.
Let be the Markov process introduced in Theorem 1.1. If , and then for we have
(4.10) |
where , , is given in (1.5).
Furthermore, if varies with for some fixed constant , then denoting by a random variable with the initial law (1.2), for we have
(4.11) |
Proof.
We use (2.25) with
(4.12) |
(To avoid notation clash, we shall use and instead of for the variables of integration, and in formulas such as (4.7).)
In view of (3.11) and (4.7), it is clear that for we have
(4.13) |
Indeed, the constant on the right hand side of (3.11) is non-zero, as our have modulus , see (4.5). So, see (4.8) and (4.3), we need to find the limit of the expression
(Here, we used (4.7) and identity/substitution .)
Next we observe that the integral over the interval is negligible. Indeed, on this interval , which together with (3.12) shows that
as , since are given by (4.12) and .
It remains to analyze the case of the integral over . Changing the variable of integration by setting , i.e. , we get
(4.14) |
Clearly, . The next observation is that for and , we have
(4.15) |
We now determine the asymptotics for the remaining factors of the integrand on the right hand side of (4.14). We apply Theorem 3.2 with and with so that . Recalling (4.12), we see that and thus
(4.17) |
as . The same argument with from (4.12) gives
(4.18) |
To verify uniform integrability, we apply elementary bounds
within (3.9). We get
Together with an elementary bound for and the fact that this show that the integrand on the right hand side of (4.14) is bounded by an integrable function , so we can use dominated convergence theorem to pass to the limit under the integral. Combining (4.13), (4.14), (4.15), (4.16), (4.17) and (4.18) we conclude that the left hand side of (4.10) becomes
as required.
Next we prove a local limit theorem with . Note that if parameter varies with , then the law of the Markov process from Theorem 1.1 also varies with , so we denote this process by .
Theorem 4.3.
In addition, if for some fixed constant , then
(4.22) |
Proof.
We write the parameters (4.5) in trigonometric form , with , .
First, we prove (4.22). As in (4.19) we have
Denote . Then with and we see that as defined in Theorem 3.3 assumes the form . Consequently, we can rewrite this expression as follows
where
(4.23) |
with the limit deduced from (3.15). Clearly,
(4.24) |
Finally, relying on (5.1), we get
(4.26) |
as . Combining (4.23), (4.24), (4.25) and (4.26) we get (4.22).
Second, we prove (4.21). Our starting point is formula (2.25) which we use with
(4.27) |
Even though , note that are non-negative integers for large enough . In particular, we note that and as . It is also clear that . (To avoid notation clash, we use and instead of for the variables of integration, or in formulas like (4.7).)
By Theorem 3.3 used with , , and or we get
(4.29) |
This gives the pair of Bessel functions on the right-hand side of (4.21).
Next, we show that the contribution from the integral in (4.28) over is negligible. Noting that for from this interval , similarly as in the proof of Theorem 4.2, the contribution of this integral is bounded by
Here is a constant from boundedness of the convergent sequence in (4.29), and the last equality follows from (4.2) and (4.7). To compute the limit we used the observation that and .
Substituting into (4.28) and discarding the non-contributing part of the integral, we have
(4.30) |
where
Referring first to (3.2) and then using and (5.2) we get
Thus can be written as
(4.31) |
Using again, we see that the first factor is asymptotically constant, . Furthermore,
By Lemma 5.1, . From Theorem 3.3, we see that for real , recall (4.20),
(4.32) |
In view of Lemma 5.1
Note that , so this factor does not contribute to the limit of .
To summarize, we see that
(4.33) |
We now justify that one can pass to the limit under the integral (4.30). For we have
for some . Recall that are fixed.
By Proposition 3.1 and Theorem 3.3 we know that for real , recall (4.32), we have
therefore the expression on the left-hand side above is uniformly bounded in . By Lemma 5.9, there exist , and such that
for all and all . By Lemma 5.7, for . Moreover, recalling that , the convergent sequence is bounded. Combining all these bounds, we see that there are constants such that for all function is bounded on by the integrable function
so we can invoke the dominated convergence theorem. Taking the limit (4.33) inside the integral in (4.30), we see that
which ends the proof by (1.9). ∎
4.2. Proofs of Theorems 1.2 and 1.3
Both proofs are very similar.
Proof of Theorem 1.2.
Proof of Theorem 1.3.
Denote by the Markov process from the conclusion of Theorem 1.1 with and .
Clearly, the centering sequence at the left-hand side of (1.7) in Theorem 1.3 can be replaced by , .
5. Auxiliary results on special functions
5.1. Some useful limits for
The following result appears in Ch. 16 of Ramanujan’s notebook, see (Adiga et al.,, 1985, Entry 1), (Berndt,, 1991, p. 13, Entry 1(i)) or (Gasper and Rahman,, 2004, (I34)) for the statement. The proof in (Koornwinder,, 1990, Proposition A.2) is for real , which is not enough for our purposes. The proof in (Berndt,, 1991, p. 13) uses analytic continuation from . We note that we always use principal branch of the logarithm and of the power function.
Lemma 5.1 (Ramanujan).
For all and in cut complex plane we have
(5.1) |
as . The convergence is uniform in on compact subsets of .
Proof.
Denote the left-hand side of (5.1) by . Then
Thus as we have
This convergence clearly holds pointwise, for all . To prove uniform convergence on compact subsets, we use Montel’s theorem, which requires us to show that the functions are uniformly bounded on compact subsets of as . For small , we define to be the closed domain
The domain is shown on Figure 2.
Now we take an arbitrary compact set . There exists small enough such that . Since , for any (in fact, for any ) and any we have .
Next, noting that , there is such that and for all . This implies that for any we have . Indeed, if then either or . In the first case we have . In the second case, .
We have thus shown that there exists such that if then, with , both and are in for all and all . But the distance from to the set is strictly positive. This means that there is such that for all , and every we have
which implies an upper bound
for some constant .
Thus we have proved that the functions converge to as , uniformly on compact subsets of . Then, integrating over the segment from to
as , also uniformly on compact subsets of . ∎
Recall (Gasper and Rahman,, 2004, Section 1.10) that for and , the -Gamma function is defined by
(5.2) |
The following result appears Ch. 16 of Ramanujan notebook. For the statement, see (Adiga et al.,, 1985, Entry 1), or (Berndt,, 1991, p. 13, Entry 1(ii)). For the proof, see (Koornwinder,, 1990, Theorem B.2).
Lemma 5.2.
If then
(5.3) |
We will also need the following.
Lemma 5.3.
Let . For any the function
is periodic with period , it is increasing for and decreasing for .
Proof.
Referring to (5.2) we note that
As changes over the interval , the point goes around a circle , thus goes around a circle centered at and having radius . Thus, is periodic with period , has minimum at and maximum at and it is clearly monotone between these points, which means that for every , the function
is increasing on and decreasing on . Taking the product of positive increasing/decreasing functions gives us an increasing/decreasing function. ∎
5.2. Bounds on
Following numerous references, such as Corwin and Knizel, (2021); Zhang, (2008) we will use the Jacobi theta functions to derive the bounds we need.
We assume that , and set . The Jacobi theta functions are defined as
(5.4) | ||||
(5.5) |
We will need the following modular transformations of theta functions and .
(5.6) | ||||
(5.7) |
The Jacobi’s triple product identity for is
(5.8) |
That is,
(5.9) |
All of these results can be found in (Rademacher,, 1973, Chapter 10).
The next results also require the following elementary bounds: for all
(5.10) |
We leave their proofs to the reader.
Lemma 5.4.
Let . There exists and a universal constant such that
(5.11) |
for all and all .
Proof.
In view of the product formula (5.8) combined with (5.2) we have the following representation
(5.12) |
We use (5.7) to rewrite (5.12) as follows
where
(5.13) |
with
In view of formula (35) in Zhang, (2008) with , and (or (Corwin and Knizel,, 2021, Proposition 2.3)) we have
Consequently, for and large enough we have
(5.14) |
For , applying bounds (5.10) we obtain:
Consequently,
for an absolute constant . Therefore, for and for large enough we have
(5.15) |
for all . In other words, we proved that
(5.16) |
uniformly in .
Lemma 5.5.
Let and assume that . There exist and (which may depend on ) such that
(5.18) |
for all and .
Proof.
Expression (5.12) with modular transformation (5.7) applied twice gives
(5.19) |
We consider two cases. If , then and the first arguments in both functions arguments satisfy the assumptions and . Using estimate (5.16) we obtain
uniformly in . We also write
The second and fourth factors in the right hand side are bounded by one in absolute value, and using the last bound in (5.10) we see that the product of the first and third factors is bounded by . Using these results and the bound on we see that in the case we have
for some absolute constant , and this implies (5.18).
We use the same method as in the proof of Lemma 5.4. In view of (5.5) we write
(5.21) |
and estimate
as follows: for any small , and we have
for an absolute constant ,
From the above estimate and (5.21) it follows that for arbitrary , as ,
(5.22) |
Since and , it is easy to see that
Lemma 5.6.
Fix and let , . Then there exist constants that depend only on such that for all we have
(5.24) |
for all .
Proof.
We consider two cases. If , then (5.24) follows from Lemma 5.4. On the other hand, if , then with we have . Recalling that , we get .
By Lemma 5.3, function is periodic with period and is increasing on the interval . With we have so we get
where we used Lemma 5.4 and then the bound . Thus in this case (5.24) follows, too.
∎
We also need the following result.
Lemma 5.7.
There exist constants such that with , for all and all we have
(5.25) |
5.3. Additional technical lemmas
The following technical Lemma is needed in proof of Theorem 3.3.
Lemma 5.8.
Let be either real or complex conjugates. If , , then for and all we have
(5.26) |
Proof.
We first prove the bound for the case of complex-conjugate parameters. Write , with and . (The case is trivial.) Consider one factor of the infinite product:
As varies through the real line, the point moves on the circle . One readily checks that the product in the numerator is largest when is minimal, i.e., when . Thus
To end the proof, we note that
The proof for the case of real requires some minor modifications and is omitted. (This case is not used in the proof of Theorem 3.3.) ∎
The proof of Theorem 4.3 uses the following technical estimate:
Lemma 5.9.
Let . If , and then there exist , and such that
(5.27) |
for all and all .
Acknowledgements
The first author was supported in part by Simons Foundation/SFARI Award Number: 703475, US. The second author was supported in part by Natural Sciences and Engineering Research Council of Canada. The authors acknowledge support from the Taft Research Center at the University of Cincinnati.
References
- Adiga et al., (1985) Adiga, C., Berndt, B. C., Bhargava, S., and Watson, G. N. (1985). Ch 16 of Ramanujan’s Second Notebook: Theta-Functions and -Series: Theta Functions and Q-series, volume 315. American Mathematical Soc.
- Aptekarev, (1993) Aptekarev, A. I. (1993). Asymptotics of orthogonal polynomials in a neighborhood of the endpoints of the interval of orthogonality. Sbornik: Mathematics, 76(1):35.
- Baik et al., (2003) Baik, J., Kriecherbauer, T., McLaughlin, K.-R., and Miller, P. D. (2003). Uniform asymptotics for polynomials orthogonal with respect to a general class of discrete weights and universality results for associated ensembles: announcement of results. International Mathematics Research Notices, 2003(15):821–858.
- Barraquand and Corwin, (2023) Barraquand, G. and Corwin, I. (2023). Stationary measures for the log-gamma polymer and KPZ equation in half-space. Ann. Probab., 51(5):1830–1869.
- Barraquand and Le Doussal, (2022) Barraquand, G. and Le Doussal, P. (2022). Steady state of the KPZ equation on an interval and Liouville quantum mechanics. Europhysics Letters, 137(6):61003. ArXiv preprint with Supplementary material: https://arxiv.org/abs/2105.15178.
- Barraquand and Le Doussal, (2023) Barraquand, G. and Le Doussal, P. (2023). Stationary measures of the KPZ equation on an interval from Enaud-Derrida’s matrix product ansatz representation. J. Phys. A, 56(14):Paper No. 144003, 14.
- Berg and Ismail, (1996) Berg, C. and Ismail, M. E. H. (1996). -Hermite polynomials and classical orthogonal polynomials. Canad. J. Math., 48(1):43–63.
- Berndt, (1991) Berndt, B. C. (1991). Ramanujan’s Notebooks. Part III. Springer-Verlag, New York.
- Billingsley, (1999) Billingsley, P. (1999). Convergence of Probability Measures. Wiley Series in Probability and Statistics: Probability and Statistics. John Wiley & Sons Inc., New York, second edition. A Wiley-Interscience Publication.
- Bryc and Kuznetsov, (2022) Bryc, W. and Kuznetsov, A. (2022). Markov limits of steady states of the KPZ equation on an interval. ALEA, Lat. Am. J. Probab. Math. Stat., 19:1329–1351. (http://arxiv.org/abs/2109.04462).
- Bryc and Wang, (2023) Bryc, W. and Wang, Y. (2023). Limit theorems for random Motzkin paths near boundary. Bernoulli (to appear). http://arxiv.org/abs/2304.00924.
- Bryc and Wang, (2024) Bryc, W. and Wang, Y. (2024). Fluctuations of random Motzkin paths II. ALEA, Lat. Am. J. Probab. Math. Stat., 21:73–94. http://arxiv.org/abs/2304.12975.
- Corwin and Knizel, (2021) Corwin, I. and Knizel, A. (2021). Stationary measure for the open KPZ equation. Communications on Pure and Applied Mathematics, to appear. arXiv preprint arXiv:2103.12253, DOI: 10.1002/cpa.22174.
- Deift et al., (2001) Deift, P., Kriecherbauer, T., McLaughlin, K.-R., Venakides, S., and Zhou, X. (2001). A Riemann–Hilbert approach to asymptotic questions for orthogonal polynomials. Journal of Computational and Applied Mathematics, 133(1-2):47–63.
- Deift et al., (1999) Deift, P., Kriecherbauer, T., McLaughlin, K. T.-R., Venakides, S., and Zhou, X. (1999). Strong asymptotics of orthogonal polynomials with respect to exponential weights. Communications on Pure and Applied Mathematics, 52(12):1491–1552.
- Flajolet and Sedgewick, (2009) Flajolet, P. and Sedgewick, R. (2009). Analytic Combinatorics. Cambridge University Press, Cambridge.
- Gasper and Rahman, (2004) Gasper, G. and Rahman, M. (2004). Basic Hypergeometric Series, volume 96. Cambridge university press.
- Ismail, (2005) Ismail, M. E. (2005). Asymptotics of -orthogonal polynomials and a -Airy function. International Mathematics Research Notices, 2005(18):1063–1088.
- Ismail and Li, (2013) Ismail, M. E. and Li, X. (2013). Plancherel–Rotach asymptotics for -orthogonal polynomials. Constructive Approximation, 37:341–356.
- Ismail et al., (2022) Ismail, M. E., Zhang, R., and Zhou, K. (2022). Orthogonal polynomials of Askey-Wilson type. arXiv preprint arXiv:2205.05280.
- Ismail, (1986) Ismail, M. E. H. (1986). Asymptotics of the Askey-Wilson and -Jacobi polynomials. SIAM J. Math. Anal., 17(6):1475–1482.
- Ismail, (2009) Ismail, M. E. H. (2009). Classical and Quantum Orthogonal Polynomials in One Variable, volume 98 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge.
- Ismail and Wilson, (1982) Ismail, M. E. H. and Wilson, J. A. (1982). Asymptotic and generating relations for the -Jacobi and polynomials. J. Approx. Theory, 36(1):43–54.
- Kaigh, (1976) Kaigh, W. D. (1976). An invariance principle for random walk conditioned by a late return to zero. The Annals of Probability, pages 115–121.
- Karlin and McGregor, (1959) Karlin, S. and McGregor, J. (1959). Random walks. Illinois Journal of Mathematics, 3(1):66–81.
- Keener, (1992) Keener, R. W. (1992). Limit theorems for random walks conditioned to stay positive. The Annals of Probability, pages 801–824.
- Koekoek et al., (2010) Koekoek, R., Lesky, P. A., Swarttouw, R. F., Koekoek, R., Lesky, P. A., and Swarttouw, R. F. (2010). Hypergeometric Orthogonal Polynomials. Springer.
- Koekoek and Swarttouw, (1994) Koekoek, R. and Swarttouw, R. F. (1994). The askey-scheme of hypergeometric orthogonal polynomials and its -analogue. https://homepage.tudelft.nl/11r49/askey/.
- Koornwinder, (1990) Koornwinder, T. H. (1990). Jacobi functions as limit cases of -ultraspherical polynomials. Journal of Mathematical Analysis and Applications, 148(1):44–54.
- Kuijlaars et al., (2004) Kuijlaars, A. B. J., McLaughlin, K. T.-R., Van Assche, W., and Vanlessen, M. (2004). The Riemann-Hilbert approach to strong asymptotics for orthogonal polynomials on . Adv. Math., 188(2):337–398.
- Levin and Lubinsky, (2020) Levin, E. and Lubinsky, D. S. (2020). Local limits for orthogonal polynomials for varying measures via universality. Journal of Approximation Theory, 254:105394.
- Lubinsky, (2020) Lubinsky, D. S. (2020). Pointwise asymptotics for orthonormal polynomials at the endpoints of the interval via universality. International Mathematics Research Notices, 2020(4):961–982.
- Nevai, (1984) Nevai, P. (1984). Asymptotics for orthogonal polynomials associated with . SIAM journal on mathematical analysis, 15(6):1177–1187.
- Olver et al., (2023) Olver, F. W. J., Daalhuis, A. B. O., Lozier, D. W., Schneider, B. I., Boisvert, R. F., Clark, C. W., Millerand, B. R., Saunders, B. V., Cohl, H. S., and McClain, M. A. (2023). NIST digital library of mathematical functions. https://dlmf.nist.gov/, 2023-12-15 DLMF Update; Version 1.1.12.
- Rademacher, (1973) Rademacher, H. (1973). Topics in Analytic Number Theory. Die Grundlehren der mathematischen Wissenschaften, Band 169. Springer-Verlag, New York-Heidelberg. Edited by E. Grosswald, J. Lehner and M. Newman.
- Szwarc, (1992) Szwarc, R. (1992). Connection coefficients of orthogonal polynomials. Canadian Mathematical Bulletin, 35(4):548–556.
- Viennot, (1985) Viennot, G. (1985). A combinatorial theory for general orthogonal polynomials with extensions and applications. In Brezinski, C., Draux, A., Magnus, A. P., Maroni, P., and Ronveaux, A., editors, Polynômes Orthogonaux et Applications, pages 139–157, Berlin, Heidelberg. Springer Berlin Heidelberg.
- Yakubovich, (2011) Yakubovich, S. (2011). The heat kernel and Heisenberg inequalities related to the Kontorovich-Lebedev transform. Commun. Pure Appl. Anal., 10(2):745–760.
- Zhang, (2008) Zhang, R. (2008). On asymptotics of -gamma functions. J. Math. Anal. Appl., 339(2):1313–1321.