Detailed balance and invariant measures
for discrete KdV- and Toda-type systems
Abstract.
In order to study the invariant measures of discrete KdV- and Toda-type systems, this article focusses on models, discretely indexed in space and time, whose dynamics are deterministic and defined locally via lattice equations. A detailed balance criterion is presented that, amongst the measures that describe spatially independent and identically/alternately distributed configurations, characterizes those that are temporally invariant in distribution. A condition for establishing ergodicity of the dynamics is also given. These results are applied to various examples of discrete integrable systems, namely the ultra-discrete and discrete KdV equations, for which it is shown that the relevant invariant measures are of exponential/geometric and generalized inverse Gaussian form, respectively, as well as the ultra-discrete and discrete Toda lattice equations, for which the relevant invariant measures are found to be of exponential/geometric and gamma form. Ergodicity is demonstrated in the case of the KdV-type models. Links between the invariant measures of the different systems are presented, as are connections with stochastic integrable models and iterated random functions. Furthermore, a number of conjectures concerning the characterization of standard distributions are posed.
Key words and phrases:
Burke’s property, detailed balance, discrete integrable system, ergodicity, integrable lattice equation, invariant measure, iterated random function, KdV equation, Toda lattice2010 Mathematics Subject Classification:
37K60 (primary), 37K10, 37L40, 60E05, 60J10 (secondary)1. Introduction
So as to capture the local dynamics of discrete KdV- and Toda-type systems, we consider a system of lattice equations with the following two-dimensional structure:
We will think of as the spatial coordinate, and as the temporal one. Moreover, the variables will represent the configuration at time , and a collection of auxiliary variables through which the dynamics from time to are defined. As for the state spaces of the variables and maps , we specialize to two cases:
- Type I (homogeneous) model:
-
The variables take values in a common Polish space . Similarly, the variables take values in a common Polish space . Moreover, for some involution .
- Type II (alternating/bipartite) model:
-
The variables take values in a Polish space if (mod 2), and in Polish space otherwise. Similarly, the variables take values in a Polish space if (mod 2), and in Polish space otherwise. Moreover, for some bijection if (mod 2), and otherwise.
This setting is rich enough to include a number of widely-studied discrete integrable systems, including the discrete and ultra-discrete KdV equations (which are examples of type I models), and the discrete and ultra-discrete Toda equations (which are examples of type II models). We highlight that these models are all important, fundamental examples of integrable systems that arise naturally within the Kadomtsev–Petviashvili hierarchy, which also includes the Korteweg-de Vries equation. See [13, 36] for mathematical and physical background. As we will expand upon shortly, our interest will be in the evolution of such discrete integrable systems started from some random initial configuration. In particular, we give criteria for identifying spatially independent and identically/alternately-distributed (in the case of a type I/type II model, respectively) initial configurations that are distributionally invariant or ergodic in time under the dynamics of the system. These general results will be applied to each of the four aforementioned examples. Furthermore, in the latter part of the article, we discuss the relevance of our results to certain examples of stochastic integrable models, and to iterated random functions.
To give a more detailed description of our main results, let us proceed to define the dynamics associated with a type I/II model precisely. In particular, we start by letting be the set of in for a type I model, or for a type II model, for which there is a unique solution to the initial value problem:
(1.1) |
We then define a function on by supposing , where is given by the unique solution of the initial value problem (1.1) with . For future convenience, we observe that clearly solves
(1.2) |
where we use a superscript to represent the th coordinate of a map. Finally, we define an operator yielding the one time-step dynamics on by supposing is given by
(1.3) |
where is given by the unique solution of the initial value problem (1.1) with . (The shift in the index is included in type II models to ensure that the elements of and that are in the spaces and are the same.) Note that we define the one time-step dynamics similarly on the set of configurations for which there is a unique solution to (1.2). (NB. It is neither the case that nor in general, though on the two definitions of agree.)
Given that the global dynamics of the system arise from locally-defined maps, it is natural to ask whether it is possible to determine which measures supported on are invariant under based on local considerations. In our first result, we show that this is indeed the case for homogeneous/alternating product measures. Before stating the result, we introduce a notion of detailed balance in our setting.
- Detailed balance condition for a type I model:
-
A pair of probability measures on and is said to satisfy the detailed balance condition if
where we define .
- Detailed balance condition for a type II model:
-
A quadruplet of probability measures (,, ,) on , , and is said to satisfy the detailed balance condition if
We then have the following characterization of independent and identically/alternately-distrib- uted configurations, which will be proved in Section 2.
Theorem 1.1 (Detailed balance criteria for invariance).
-
(a)
Type I model. Suppose is a probability measure on and . It is then the case that if and only if there exists a probability measure on such that the pair satisfies the detailed balance condition. Moreover, when this holds, is the distribution of for each , where is distributed according to .
-
(b)
Type II model. Suppose , are probability measures on , and . It is then the case that if and only if there exists probability measures , on , , respectively, such that the quadruplet of probability measures satisfies the detailed balance condition. Moreover, when this holds, then , are the distributions of , , respectively, for each , where is distributed according to .
We remark that the above theorem does not in itself provide a truly local criteria for invariance of homogeneous/alternating product measures under . Indeed, the condition that or depends on knowledge of the global dynamics, and in particular a suitably accessible description of . We do not present a universal approach to this problem here. However, for the KdV- and Toda-type systems already mentioned, the existence and uniqueness of solutions to the initial value problem (1.1) was studied in detail in [13], where it was shown that the associated dynamics could be interpreted in terms of certain ‘Pitman-type transformations’ of related path encodings of the configurations. In this article, we will incorporate as a key ingredient the results of [13] when applying Theorem 1.1 to these examples. (NB. A brief introduction to the results of [13] is presented in [11].)
To prove Theorem 1.1, we proceed in two steps. Firstly, we establish a weaker version (see Theorem 2.1 below), in which the invariance of or under is shown to be equivalent to the detailed balance condition holding with , given by the relevant marginals of . Since it is not trivial to deduce the distribution of from or in general, the latter version of the result is far from straightforward to apply in examples. Towards dealing with this issue, we show that invariant measures on of homogeneous/alternating product form induce stationary/alternating measures of satisfying Burke’s property (see Subsection 2.2 below), and moreover they are the only such measures satisfying this property. Namely, Burke’s property is equivalent to the detailed balance condition or . Combining this observation with Theorem 2.1 yields our main result, i.e. Theorem 1.1. See Section 2, where a sufficient condition for establishing ergodicity of such invariant measures for type I models is also given, for details.
The abstract results discussed above are applied to our concrete KdV- and Toda-type examples of discrete integrable systems in Sections 3 and 4, respectively. In particular, we show that spatially independent and identically/alternately distributed configurations that are also temporally invariant are of exponential/geometric form for the ultra-discrete KdV equation, of generalized inverse Gaussian form for the discrete KdV equation, of exponential/geometric form for the ultra-discrete Toda lattice, and of gamma form for the discrete Toda lattice. Our proofs for checking detailed balance for the various models depends on some well-known characterizations of certain standard distributions, including the exponential, geometric, gamma and generalized inverse Gaussian distributions [8, 18, 17, 30, 28]. Let us also highlight that the lattice structure of the Toda examples is not immediately covered by the framework of this article, with each being based on a map with three inputs and three outputs. Nonetheless, in both the discrete and ultra-discrete cases, it is possible to describe a type II model for which the involution defined by
(1.4) |
gives the appropriate dynamics. For a general involution of this form, we show that invariance under , i.e.
(1.5) |
is equivalent to the detailed balance condition for , i.e. for some , and indeed that both these conditions are equivalent to
(1.6) |
The detailed balance solutions that we derive in our examples yield corresponding invariant measures of the form described above. Our results yield that these satisfy Burke’s property, and we also explore ergodicity for the KdV (type I) models. Moreover, in Section 5, we discuss natural relationships between the detailed balance solutions/invariant measures of the systems in question, which are based on an ultra-discretization procedure, and a certain KdV-Toda correspondence. See Figure 1 below for a summary of these results.
Although in this article we restrict to the case when the maps are deterministic, it is also possible to consider stochastic models, in which the maps themselves are random. In Section 6, we provide some comments on generalizations of our results to this setting, and present links with certain stochastic integrable (solvable) lattice models, specifically last passage percolation, random polymers and higher spin vertex models. We note in particular that the relation at (1.6) is closely related to Burke’s property for two-dimensional stochastic solvable models in integrable probability.
Another strand of literature to which the present article connects is that regarding iterated random functions. Indeed, one can understand (1.2) as a map based on the random function . Such systems arise in many settings, and there are a number of important problems that arise for them, such as the -measurability of . Moreover, if is an independent sequence, then is a Markov chain (homogeneous for type I models, and with alternating transition probabilities for type II models), and one can ask questions about corresponding invariant measures and ergodicity for this process (or suitable variations for type II models). We will discuss how our results can be understood in this context in Section 7.
Finally, in Section 8, we summarize some of the open problems that are left open by this study, and present some conjectures on the characterization of some standard distributions that arise naturally from this study. We also include an appendix containing definitions of some of the probability distributions that appear in earlier sections.
2. Setting and abstract results
In this section, we prove the abstract results outlined in the introduction. We continue to apply the definitions of a type I/II model, the set of configurations for which there exists a unique solution to the initial value problem (1.1), the function , and the operator , as given there. In Subsection 2.1, we prove the weaker version of Theorem 1.1 discussed in the introduction. Moreover, in the type II setting, we establish the characterization of solutions to the detailed balance condition in terms of the conditions at (1.5) and (1.6). In Subsection 2.2, we present our conclusions concerning Burke’s theorem in the present context. These allow us to strengthen the relevant result in Subsection 2.1, and thereby obtain Theorem 1.1. As noted above, this provides our means for checking invariance of homogeneous/alternating product measures under in examples. Finally, in Subsection 2.3, we develop an argument for checking the ergodicity of such invariant measures under for type I models.
2.1. The detailed balance condition and invariance
Recalling the definition of the detailed balance condition for type I/II models from the introduction, the first goal of this subsection is to prove the following variation on Theorem 1.1, which provides a link between detailed balance solutions and invariant measures.
Theorem 2.1.
-
(a)
Type I model. Suppose is a probability measure on and . Let be the distribution of , where is distributed according to . It is then the case that if and only if the pair satisfies the detailed balance condition.
-
(b)
Type II model. Suppose , are probability measures on , and . Let , be the distributions of , , respectively, where is distributed according to . It is then the case that if and only if the quadruplet of probability measures satisfies the detailed balance condition.
Remark 2.2.
Let be a set of configurations for which there is a solution to (1.2) for which is a function of for all , and for a type I model, and , for a type II model, where is the usual shift operator. Moreover, assume that , and , where depends on through (1.3), and for a type I model and for a type II model. If is the identity map on , then Theorem 2.1 holds when we replace by . It might be easier to find a space than in some cases.
Towards proving Theorem 2.1, we start by setting out a lemma on the measurability of and in terms of the initial configuration . This is stated in terms of functions and on that are defined via the relation
where is the unique solution of (1.1) with initial condition .
Lemma 2.3.
Let .
-
(a)
For any and , and are measurable with respect to .
-
(b)
For any and , and are measurable with respect to .
Proof.
(a) Suppose there exist and in such that for all , but or for some , . We then define:
Moreover, for , , it is clear from the lattice structure that there is a unique solution to that is consistent with the previous definitions. Recursively, we have that solves (1.1) with initial condition . Since or for some , by assumption, this contradicts the uniqueness of the solution of (1.1) for . Hence we conclude that and are measurable with respect to .
(b) Appealing to the symmetry of the map , we can apply the same proof as for part (a).
∎
In the next lemma, we rephrase spatial/temporal invariance of the law of an initial configuration as invariance under appropriate shifts of the induced law on variables on the entire lattice. Specifically, for a probability measure supported on , we denote by the probability distribution of , as defined by the initial value problem (1.1), for which the marginal of is given by . We define a spatial shift on lattice variables by setting
Slightly abusing notation, for elements , we similarly suppose . The corresponding temporal shift is given by
Note that if we consider as the map on given by , then the definition of the dynamics at (1.3) means that, for ,
NB. From this description, it is easy to see that is a bijection, with inverse operation , where is defined as in Remark 2.2.
Lemma 2.4.
Let be a probability measure supported on .
-
(a)
For a type I model, if and only if . Also, if and only if .
-
(b)
For a type II model, if and only if . Also, if and only if .
Proof.
(a) If or holds, then , and so is well-defined. The claim then follows from the fact that . The same argument works for .
(b) Again, the same argument works.
∎
Combining the previous two lemmas, we have the following.
Corollary 2.5.
Let be a probability measure supported on , and suppose . It is then the case that there is a subset of two-dimensional configurations such that, with probability one on this subset, for any :
-
(a)
for any and , and are measurable with respect to ;
-
(b)
for any and , and are measurable with respect to .
Proof.
Before proceeding, we note the following consequence of the above measurability results, which is somewhat related to Burke’s property, as will be introduced in the next subsection. The particular statement will not be used later, but we believe it is of independent interest to observe that we do not require spatial stationarity of the initial configuration to establish temporal independence of the random variables .
Corollary 2.6.
Let be a probability measure supported on , and suppose is an independent sequence under .
-
(a)
For a type I model, if it holds that , then is an independent and identically distributed (i.i.d.) sequence under .
-
(b)
For a type II model, if it holds that , then is an independent and alternately-distributed sequence under .
Proof.
(a) Since , it readily follows that the sequence is stationary. As for the independence claim, we note that, by Corollary 2.5, is a measurable function of , and is a measurable function of . Since and are independent, the result follows.
(b) The proof is similar.
∎
We are nearly read to prove Theorem 2.1. As the final ingredient, we give an elementary lemma regarding independence of sigma-algebras.
Lemma 2.7.
Let be sigma-algebras on a probability space. If and are independent, and and are independent, then and are independent.
Proof.
Denoting by the probability measure on the relevant space, we have that, for any , , . The result follows. ∎
Proof of Theorem 2.1.
(a) Suppose . By definition, we have that and . Moreover, by invariance under , we have that . And, since , Lemma 2.4 yields that , and so the distribution of is also . Now, by Corollary 2.5, we have that is a measurable function of , and is a measurable function of . In particular, it follows that is independent of , and is independent of , i.e. it holds that and . Since , we thus obtain that satisfies the detailed balance condition in this case.
Next, suppose that . By Lemma 2.3, is measurable with respect to , so and are independent for all . By assumption . Moreover, by assumption and the invariance given by Lemma 2.4, . Hence the distribution of is , and also and are independent, for all . Since, by Lemma 2.3, and are both measurable with respect to , it follows from Lemma 2.7 that and are independent. Therefore, since is measurable with respect to , it must be the case that and are independent. Finally, since by Lemma 2.4, we obtain that is an i.i.d. sequence with marginal distribution , and so .
(b) Essentially the same argument as for part (a) applies.
∎
We complete the subsection by proving the alternative characterizations of the detailed balance condition for type II models that were presented in the introduction.
Proposition 2.8.
Let be a bijection, and define the involution as at (1.4). For a triplet of probability measures on , and , the following three conditions are then equivalent.
-
(a)
.
-
(b)
.
-
(c)
There exists a probability measure on such that the quadruplet of probability measures satisfies the detailed balance condition with respect to .
Proof.
(b) (a): This is obvious.
(c) (b): Let , , be independent random variables, and define . By (c), . Moreover, by Lemma 2.7, and are independent. Now, by definition, , and, by the detailed balance condition, , so (b) holds.
(a) (c): Let , and , , be independent random variables. Since and the distribution of is , (a) implies .
∎
2.2. Burke’s property
Burke’s theorem is a classical result in queueing theory, which states that, for an queue, the departure process at stationarity has the same law as the arrivals process, and that the departure process prior to a given time is independent of the current queue length [4]. This result has been generalized to many settings, see Section 6 for discussion in the context of stochastic integrable systems in particular. In this subsection we present a definition of Burke’s property for our model, and relate it to the study of the detailed balance condition and invariant homogeneous/alternating product measures. This allows us to complete the proof of Theorem 1.1.
- Burke’s property for a type I model:
-
We say that a distribution supported on configurations satisfying satisfies Burke’s property if:
-
•:
the sequences and are each i.i.d., and independent of each other;
-
•:
the distribution of is translation invariant, that is, for any ,
-
•:
- Burke’s property for a type II model:
-
We say that a distribution supported on configurations satisfying satisfies Burke’s property if:
-
•:
the sequences , , and are each i.i.d., and independent of each other;
-
•:
the distribution of is translation invariant, that is, for any such that (mod 2),
-
•:
We make the obvious remark that, in the case of a type I model, if the distribution of satisfies Burke’s property, then is i.i.d. for each , and is i.i.d. for each . A similar property holds for type II models.
In the main result of this subsection, we show that the existence of a solution to the detailed balance condition implies the existence of a distribution satisfying Burke’s property. Moreover, in the case that the relevant marginal of this measure is supported on configurations for which (1.1) has a unique solution, we are able to describe both the distributions of and in terms of the detailed balance solution.
Proposition 2.9 (Burke’s property).
-
(a)
Type I: If a pair of probability measures satisfies the detailed balance condition, then there exists a distribution supported on configurations satisfying for which Burke’s property holds. Moreover, if it holds that , then and satisfies Burke’s property.
-
(b)
Type II: If a quadruplet of probability measures satisfies the detailed balance condition, then there exists a distribution supported on configurations satisfying for which Burke’s property holds. Moreover, if it holds that , then , , and satisfies Burke’s property.
Proof.
(a) Let be independent random variables satisfying and . For , define
recursively. By induction and the detailed balance condition, one readily obtains that , and and are independent for all . Moreover, for any , and are measurable with respect to , and is measurable with respect to . So, applying Lemma 2.7, we find that and are independent. Hence is an i.i.d. sequence with the marginal . Now, since is measurable with respect to , it further holds that and are independent. Letting and , we thus have that are independent random variables satisfying , and
for all . In particular, , which implies
By the same argument, one can show that
and so
for any . Finally, by constructing the distributions of for each by translation, we can construct the distribution of by applying the Daniell-Kolmogorov extension theorem, see [24, Theorem 5.14], for example. (This is the one place in our arguments where we require the state spaces to be Polish.)
Moreover, if , then there is a unique distribution of that is supported on configurations satisfying and with marginal . Hence it must be the one satisfying Burke’s property, as constructed above.
In particular, for all .
(b) The same argument as for part (a) works.
∎
We conclude the subsection with a corollary that establishes, when the marginal of is supported on , Burke’s property is actually equivalent to the detailed balance condition. As with Theorem 1.1, it readily follows from Theorem 2.1 and Proposition 2.9.
Corollary 2.10.
-
(a)
Type I: Suppose that is a probability measure on such that . Let be the distribution of , where . It is then the case that there exists a distribution of satisfying and Burke’s property if and only if satisfies the detailed balance condition.
-
(b)
Type II: Suppose that is a probability measure on such that . Let , be the distributions of , , respectively, where . It is then the case that if there exists a distribution of satisfying and Burke’s property if and only if satisfies the detailed balance condition.
Proof.
(a) The ‘if’ part is shown in Proposition 2.9. We prove the ‘only if’ part. Suppose that there exists a distribution of satisfying and Burke’s property. Since , the measure must be . By the second condition of Burke’s property, holds. Hence, by Lemma 2.4, we must have that holds. Consequently, by Theorem 2.1, the detailed balance condition holds.
(b) The same argument as for part (a) works.
∎
2.3. Ergodicity
We now turn our attention to the issue of ergodicity. In this part of the article, we consider only type I models. Our main result gives a sufficient condition for the ergodicity of for i.i.d. invariant measures. To state the result, we introduce an involution by setting
where . We consider the dual of .
Theorem 2.11.
Suppose we have a type I model, and that is a probability measure on such that and . If it holds that, for -a.e. , there exists at most one such that
then is ergodic under .
Remark 2.12.
Remark 2.13.
The proof of the above theorem will depend on the following lemma. For the statement of this, we define a function by setting
where is the unique solution of (1.1) with initial condition . Note that, as is consistent with the idea that is a temporal shift, we set .
Lemma 2.14.
Let be a distribution on . Suppose there exists a set and a function
such that and is the identity map on the set . The following statements then hold.
-
(a)
is invariant under if and only if is invariant under .
-
(b)
is invariant and ergodic under if and only if is invariant and ergodic under .
Proof.
(a) Define and . We first check that is a bijection with inverse function . Clearly . Moreover, by assumption, for all . Hence it remains to show that
For any , there exists such that . It follows that
as required. Next, since , we have that , and thus also . Consequently, if , then it -a.s. holds that and take values in , and so
It follows that . On the other hand, if , then it -a.s. holds that and takes values in , and so
Hence .
(b) By the proof of (a), for any subset , , and so is equivalent to . The claim follows.
∎
Remark 2.15.
The same result was shown in [10] in the setting of the box-ball system of finite box and/or carrier capacity.
Proof of Theorem 2.11.
As per Remark 2.12, we know that . Moreover, is clearly invariant and ergodic under . Hence, by Lemma 2.14, we only need to show the existence of a set and a function such that and is the identity map on the set . To this end, let be the set of such that there is at most one satisfying
By assumption, . Since , where , and under for all , it follows that
(2.1) |
Now, define to be the set of such that for all , where is given by the solution of the initial value problem (1.1) with initial condition . Moreover, set , and note that, by (2.1), we have that . We next claim that for any , there is a unique such that , and moreover that . Indeed, if , and , then
where and are the solutions of the initial value problem (1.1) with initial conditions and , respectively. Hence,
Since is an element of , it must therefore be the case that for all . It moreover follows that for all . Since implies for all , iterating this argument yields that for all and . To deal with negative , note that
is equivalent to
and so
Applying the reflection thus yields
Since , this implies for all . Again, we can iterate this argument to conclude that for all , as desired. Hence the function given by is well-defined, and for all . Moreover, we have from the above argument that , and so the proof is complete. ∎
3. Type I examples: KdV-type discrete integrable systems
Two important examples of discrete integrable systems are the discrete and ultra-discrete KdV equations, which are obtained from the original KdV equation by natural discretization and ultra-discretization procedures. See [13, 36] and the references therein for background. Both are examples of type I systems, and the aim of this section is to explain how our general results for such can be applied to identify examples of invariant and ergodic measures for them.
3.1. Ultra-discrete KdV equation
3.1.1. The model
The (modified) ultra-discrete KdV equation incorporates two parameters, , and is based on the following lattice map:
(udKdV) | ||||
where the variables and are valued. When the variables are positive, one can think of as the amount of mass currently at a lattice site, which has capacity . Moreover, represents the amount of mass that a ‘carrier’, which has capacity , is bringing to this site. Simultaneously, the carrier deposits what it can, i.e. , and collects what it can, i.e. . This leaves a mass of
at the site, and the carrier moves forward (rightwards) to the next lattice site carrying a mass of
one discrete time step of the lattice dynamics is given by a complete pass of the carrier from to . We note that the original udKdV equation corresponds to setting . We also highlight that if and we restrict the possible values of the variables so that and , then the dynamics associated with correspond to the box-ball system with box capacity and carrier capacity , which we denote by BBS(,).
Remark 3.1.
Similarly to the discussion for BBS(,) in [10], the map (udKdV) admits various symmetries, including the following.
- Involution:
-
For any , it holds that
(3.1) - Configuration-carrier duality:
-
If , then
(3.2) - Empty space-particle duality:
-
Suppose . If , then
(3.3) - Shift invariance:
-
If , then for any it holds that
(3.4) - Scale invariance:
-
If , then for any it holds that
(3.5)
Note that, whilst we will not dwell on it here, the property (3.1) implies that the time-reversal of the (udKdV) system can be studied in exactly the same way as the original system. As for (3.2), this means that it will suffice to solve the detailed balance equation for . Properties (3.3), (3.4) and (3.5) yield corresponding relationships between solutions of the detailed balance equation for (udKdV) of various parameters.
3.1.2. Detailed balance solutions
We now address the detailed balance equation for (udKdV); as per Remark 3.1, it will be enough to do this for . We give two results. The first, Proposition 3.2 lists a number of solutions of the detailed balance equation. We highlight that the detailed balance equation was completely solved for the BBS(,) in [10], and the discrete part of the following result (i.e. (a)(ii)) is essentially a restatement of the result from that paper. We refer the reader to the appendix for definitions of the probability distributions that appear. Our second result, Proposition 3.3 shows, up to a technical condition, that these are all the solutions of the detailed balance equation in this setting.
Proposition 3.2.
The following product measures satisfy .
-
(a)
Suppose .
-
(i)
For if , or if , and finite ,
-
(ii)
For finite and such that ,
where it is further supposed that: either , , ; or , , ; or , , ; or , , .
-
(i)
-
(b)
Suppose . For any measure on ,
-
(c)
Suppose .
-
(i)
For any measure supported on ,
-
(ii)
For any measure supported on ,
where for , is the probability measure placing all of its mass at .
-
(iii)
Suppose further that . For any measure supported on ,
where and .
-
(i)
Proof.
Since preserves mass, i.e.
and the absolute value of the associated Jacobian determinant is equal to one (Lebesgue almost-everywhere), part (a)(i) is straightforward to check. As already noted, part (a)(ii) was proved in [10]. Parts (b) and (c) readily follow from the definition of , and so their proofs are omitted. ∎
Proposition 3.3.
-
(a)
Suppose . It is then the case that the product measures given in Proposition 3.2(b) are the only solutions to .
-
(b)
Suppose and a product measure satisfies . It is then the case that one of the following statements hold.
-
(i)
The product measure is given in Proposition 3.2(c).
-
(ii)
There exists such that
where and are the support of and , respectively.
Moreover, if (ii) holds and and have smooth (twice differentiable), strictly positive densities on the intervals and respectively, then they given by Proposition 3.2(a)(i). And, if (ii) holds and neither nor contains an accumulation point, then they are given by Proposition 3.2(a)(ii).
-
(i)
Proof.
(a) Since , this part of the result is obvious.
(b) Let , , , . Since
where , implies and . Thus it holds that
Also, by definition, it holds that:
and, in particular, is continuous with respect to . We now consider three cases separately: (I) , (II) , (III) .
-
(I)
If , then . This implies , , and so .
-
(II)
If , then . Hence , , which implies in turn that and . If , namely is the measure , then must be concentrated on . If , then there exist , such that , . In particular, we can take small enough so that . If , then, , but this contradicts with the fact that . On the other hand, if , then . However, , which again contradicts with . Thus we have shown that it is not possible that . Consequently, in this case, if holds, then and .
-
(III)
If , then for all , so holds if and only if .
We next consider the corresponding three cases for the suprema of the support: (I’) , (II’) , (III’) . By a similar argument to above, we have the following.
-
(I’)
If , then holds if and only if .
-
(II’)
If and holds, then , .
-
(III’)
If , then and .
Putting together the above discussion, there are only four possible cases: (I”-1) and ; (I”-2) , ; (I”-3) and ; (II”) , and , . The cases (I”-1), (I”-2), (I”-3) correspond to Proposition 3.2(c)(i), (ii), (iii), respectively. It remains to check that the case (II”) corresponds to part (b)(ii) of the current proposition. In this case, there exist such that and . Suppose . Then, . If , then , and so . The latter inequality is equivalent to , which contradicts . If , then , which implies . However, this contradicts . Hence . A similar argument allows one to deduce the reverse inequality, and thus we obtain . In conclusion, letting , we obtain the desired result.
To complete the proof, we study the special cases where and have densities, or they are discrete. Let , be densities of and . For and , we then have that
Letting and and taking derivatives of the relation
with respect to first and then with respect to , for satisfying , we have . For any , by letting and
we have , , and , so for all . Therefore, there exists such that for all . Also, by taking the derivative of with respect to , for satisfying , we have . Since for any , by taking , we have and , hence for all . Therefore, since and are probability measures, must be finite. Moreover, if , then must be positive.
Finally, we consider the case where and are discrete. We first prove that . If , then for any , there exists such that , and vice versa. Since for such we have , we conclude that . Moreover, by noting for , it is an elementary exercise to check that . Next, note that if , then for any with , it holds that and
However, since , this implies that for any , for all and for some , which can not happen since is a probability measure. Similarly, if , then for any satisfying and , we have , and
In particular, applying this relation with satisfying with , we have . Iterating this argument yields for all , and
for all . Moreover, since for ,
we have for all , for some . Again, this can not happen since is a probability measure. We can therefore conclude that .
First suppose . We then have for some with , and for some . Additionally, for some with , and for some . By a similar argument to the previous paragraph, it is possible to check that, for an appropriate choice of , one may take , and moreover , for each . Hence, by making the change of variables , we can apply [10, Lemma 4.5] to complete the proof. To establish the result when , one can proceed in the same way to check that for some , and then use the identity for to derive the full support of , from which point one can again apply [10, Lemma 4.5] to obtain the desired result. ∎
3.1.3. Invariant measures
Much of the hard work for identifying invariant product measures for (udKdV) has now been done. Indeed, up to the technical restriction of Proposition 3.3, Theorem 1.1 tells us that the marginals of invariant product measures must be described within the statement of Proposition 3.2 (as in the case , and in the case ).
We start by restricting our attention to . The reason for this is that it allows us to apply the approach of [9, 10, 13], which provides a description of the dynamics in terms of certain Pitman-type transformations of path encodings of configurations, to give an explicit set upon which the initial value problem (1.1) has a unique solution. In particular, we will now consider the initial value problem (1.1) with for all , where . For , we set . For , we take
where for , the sum should be interpreted as , and, for ,
Lemma 3.4.
Suppose . If , then there exists a unique solution of (1.1) with for all .
Proof.
To handle the case , we consider the set
(3.6) |
and for , the set
(3.7) |
The subsequent result gives that if we start from a configuration within these sets, then it is not possible to give multiple definitions for the one time-step dynamics.
Lemma 3.5.
Suppose . If , then there exists at most one sequence such that
(3.8) |
Proof.
We first prove that if , then . Since and
it must hold that . Hence , and so
Similarly, if and , then , where . Indeed, since , in this case we have that , and the result follows. As a consequence, if , then there exists a sequence such that is determined by . For , the relation (3.8) means that is uniquely defined by such that and , and so the proof is complete. ∎
Putting together Theorem 1.1, Proposition 3.2, Lemma 3.4 and Lemma 3.5, we complete this section by describing a number of invariant product measures for (udKdV). We write for the dynamics given by , as defined at (1.3).
Theorem 3.6.
Proof.
(a) The case is obvious.
(b) In the case , for one of the measures from Proposition 3.2 to satisfy , it will suffice to check that . For this, the law of large numbers tells us that it is sufficient for . The measures given in the statement of the theorem are readily checked to satisfy this requirement. Finally, for , it will again be enough to determine measures from Proposition 3.2 that satisfy . The latter constraint simply rules out the trivial measure , and so the result readily follows.
(c) Let us continue for the moment to suppose that . We will appeal to the configuration-carrier duality of (3.2) to prove the result, and as a first step we take to be one of the measures identified in part (b). If is given by solving the initial value problem (1.1) with initial condition , it then readily follows from Proposition 2.9 that, for each , is i.i.d., with marginal given by the corresponding from Proposition 3.2. Now, as long as , then it is clear that . This means that, -a.s., uniquely determines , with , where represents the dynamics given by (cf. the proof of Theorem 2.11). In particular, we have demonstrated that . Reversing the role of and gives the result.
∎
3.1.4. Ergodicity
Finally, we study the ergodicity of the operator . The next result is an immediate application of Theorem 2.11, together with the observations we made in the proof of Theorem 3.6, and so we simply state the conclusion.
Theorem 3.7.
Suppose . Let be a product measure satisfying , as given by Proposition 3.2, with and . It is then the case that is ergodic under , and is ergodic under .
3.2. Discrete KdV equation
3.2.1. The model
Our next model, the (modified) discrete KdV equation also incorporates two parameters, in this case given by , and is based on the following lattice map:
(dKdV) |
where we now assume the variables and are valued. We note that satisfies the Yang-Baxter relation, and may be derived from the 3d-consistency condition of the discrete potential KdV equation or the discrete BKP equation, see [32, 23]. Moreover, if , then gives the discrete KdV equation.
Remark 3.8.
Similarly to Remark 3.1, the lattice map (dKdV) admits a number of symmetries.
- Involution:
-
For any , it holds that
- Configuration-carrier duality:
-
If , then
- Empty space-particle duality:
-
Suppose . If , then
- Scale invariance:
-
If , then for any it holds that
We note that scale invariance in this setting corresponds to the shift invariance of (udKdV).
3.2.2. Detailed balance solutions
For (dKdV), we are unable to characterize the solutions of the detailed balance equation, even up to a technical condition as we did for (udKdV). Nonetheless, we are able to describe a family of solutions based on the GIG distribution. As we explain in Section 5, this family naturally corresponds to the stExp solutions of the (udKdV) detailed balance equation, as presented in Proposition 3.2.
Proposition 3.9.
The following product measures satisfy .
-
(a)
For any if , or if , and ,
-
(b)
Suppose . For any measure on ,
In the case , there are no other non-trivial (i.e. non-Dirac measure) solutions to the detailed balance equation.
Proof.
(a) To verify the claim, given that absolute value of the associated Jacobian determinant of is equal to one, it suffices to check that the following relation between joint densities:
where and . This is a simple consequence of the identities and , which can be checked directly.
(b) Since , the result is obvious.
For the final part of the result, suppose . In this case, the map of interest becomes
Now, in [28, Theorem 4.1], it is shown that if and are strictly positive independent random variables such that at least one of and has a non-trivial distribution, and and are also independent, then must have a generalized inverse Gaussian distribution and must have a gamma distribution with related parameters. (NB. This result builds on [27].) Considering the form of the map as given above, and applying [28, Theorem 4.1] with , yields the result. ∎
3.2.3. Invariant measures
We now show how the measures of Proposition 3.9 yield invariant product measures for , that is, the operator describing the (dKdV) dynamics. Apart from the trivial case , we restrict our attention to the case when . (We list the case amongst the open problems in Section 8.) The reason for this is that it will allow the application of the path encoding results from [13] concerning the initial value problem (1.1). In particular, consider the latter problem with for all , where and . Letting
we have the following result (see [13, Theorem 2.2]).
Lemma 3.10.
Suppose . If , then there exists a unique solution of (1.1) with for all .
In the case , , we consider the set
where . The parallel to Lemma 3.5 that we apply in the discrete setting is the following.
Lemma 3.11.
Suppose , . If , then there exists precisely one sequence such that
(3.9) |
which is explicitly given by the infinite continued fraction
Proof.
The relation (3.9) can be written as
which is equivalent to
Hence, the sequence defined by
satisfies (3.9). Indeed, the condition ensures that the infinite continued fraction converges in (see [29, Chapter 8], for example). Suppose that we have another solution to (3.9). It is then the case that
Iterating this, we find that for any ,
Taking the limit as , the defining properties of imply that , as desired. ∎
Arguing as for Theorem 3.6, we have that Theorem 1.1, Proposition 3.9 and Lemmas 3.10 and 3.11 yield the subsequent result. For the proof of part (c) of the result, the one additional useful observation is that if and , then (this ensures that the given condition is enough to ensure that both marginals of the solution to the relevant detailed balance equation satisfy the required logarithmic moment bound).
Theorem 3.12.
The product measure satisfies for the following measures .
-
(a)
Suppose . Any measure on .
-
(b)
Suppose , . The measure for any parameters such that .
-
(c)
Suppose , . The measure for any parameters such that .
3.2.4. Ergodicity
Regarding the ergodicity of , combining the results of the previous section with Theorem 2.11 gives the next result.
Theorem 3.13.
Suppose , . Let be a product measure satisfying , as given by Proposition 3.9 (i.e. . If it holds that , it is then the case that is ergodic under , and is ergodic under .
4. Type II examples: Toda-type discrete integrable systems
The type II examples that we study arise from two other important discrete integrable systems, namely the discrete and ultra-discrete Toda equations. Again, see [7, 15] and the references therein for background. As in the previous section, our aim is to identify solutions of the corresponding detailed balance equations and invariant measures. For type II systems, we do not have a strategy for checking ergodicity.
4.1. Ultra-discrete Toda equation
4.1.1. The model
The ultra-discrete Toda equation is described as follows:
(udToda) |
where take values in . We summarise this evolution as , highlighting that is an involution on , and represent the lattice structure diagrammatically as
(4.1) |
Whilst this system might not immediately appear to link with (udKdV) or the BBS, we note that if we restrict to non-negative integer-valued variables, and view as the length of the th interval containing balls, as the length of the th empty interval (at time ), and as the carrier load at the relevant lattice location, then the dynamics of these variables coincides with that given by the BBS. (In the case of infinite balls, there is an issue of how to enumerate the intervals.) Moreover, although the lattice structure at (4.1) does not immediately fit into our general framework, it is possible to decompose the single map with three inputs and three outputs into two maps and , each with two inputs and two outputs:
where we generically take . Including the additional lattice variables, we can thus view the system as type II locally-defined dynamics, as defined in the introduction, with the maps alternating between the bijection and its inverse, which are given explicitly by
Note that the decomposition of into and is not unique. The form of chosen here is slightly simpler than the corresponding map in [13] (see also [11]), since we do not need to satisfy the additional constraint that yields a ‘Pitman-type transformation map’.
4.1.2. Detailed balance solutions
For , we are able to completely solve the detailed balance equation, see Proposition 4.1. In the subsequent result, Proposition 4.2, we show how this yields a complete solution to the corresponding problem for .
Proposition 4.1.
The following measures satisfy .
-
(a)
For any and ,
-
(b)
For any , and ,
-
(c)
For any and measure supported on ,
-
(i)
, , , ,
-
(ii)
, , , ,
-
(iii)
, , , .
NB. Case (c)(i) is contained in cases (c)(ii) and (c)(iii).
-
(i)
It is further the case that there are no other quadruples of probability measures that satisfy .
Proof.
The first part follows by direct computation. The uniqueness claim relies on a well-known fact [18, 17, 8] about exponential and geometric distributions. Namely, suppose that and are two non-constant, independent random variables. It is then the case that and are independent if and only if and are sExp-distributed random variables with the same location parameter or ssGeo-distributed random variables with the same location and scale parameters. The trivial solutions of part (c) are covered by [18, Theorem 1 (and the following comment)]. ∎
By construction, we have that
This enables us to deduce from Propositions 2.8 and 4.1 the subsequent result.
Proposition 4.2.
The following product measures satisfy .
-
(a)
For any and ,
-
(b)
For any , and ,
-
(c)
For any and measure supported on ,
-
(i)
,
-
(ii)
,
-
(iii)
.
NB. Again, case (c)(i) is contained in cases (c)(ii) and (c)(iii).
-
(i)
Moreover, if a product measure is invariant under , then it must be one of the above.
Proof.
The first part follows directly from Propositions 2.8 and 4.1. To show uniqueness, let and be independent random variables satisfying
Let , then, by assumption,
Hence . Since are independent, and are independent. Therefore the marginals of must be given by one of the collections described in Proposition 4.1. ∎
4.1.3. Invariant measures
The initial value problem for the ultra-discrete Toda equation that we consider is: for , find such that (udToda) holds for all . This was solved in [13] for initial conditions in the set
In particular, the subsequent result was established.
Lemma 4.3 ([13, Theorem 2.3]).
If , then there exists a unique collection such that (udToda) holds for all .
In the case when a unique solution to (udToda) exists, it makes sense to define the dynamics of the system similarly to (1.3), i.e. set
In what is the main result of this section, we characterize invariant product measures for the resulting evolution.
Theorem 4.4.
Suppose that is an i.i.d. sequence with marginal given by , where one of the following holds:
-
(a)
for some and ,
-
(b)
for some , and ,
-
(c)
for some and measure supported on with ,
It is then the case that . Moreover, there are no other non-trivial measures such that is an i.i.d. sequence, with independent of , and .
Proof.
If is an i.i.d. sequence with marginal of one of the given forms, then it is a simple application of the law of large numbers to check that, -a.s., . It readily follows from Lemma 4.3 that, -a.s., the corresponding type II lattice equations have a unique solution with initial condition , where and . Thus we can apply Theorem 1.1 and Proposition 4.1 to deduce the result. ∎
4.2. Discrete Toda equation
4.2.1. The model
The discrete Toda equation is given by:
(dToda) |
Here, the variables take values in , and we can summarise the above dynamics by , where is an involution on . Similarly to (4.1), in this case we have a lattice structure
which can be decomposed into two maps, and , as follows:
where we generically take . So, again including the additional lattice variables, we can view the system as type II locally-defined dynamics, as defined in the introduction, with the maps alternating between the bijection and its inverse, which are given explicitly by:
As in the ultra-discrete case, we note that the decomposition of into and is not unique, with the form of chosen here being slightly simpler than the corresponding map in [13] (see also [11]).
4.2.2. Detailed balance solutions
As in the ultra-discrete case, we are also able to completely solve the detailed balance equation for , see Proposition 4.5. In the subsequent result, Proposition 4.6, we apply this to deduce a complete solution to the corresponding problem for .
Proposition 4.5.
The following measures satisfy .
-
(a)
For any and ,
-
(b)
For any , , , , .
It is further the case that there are no other quadruples of probability measures that satisfy .
Proof.
The first part follows by direct computation. The uniqueness relies on a well-known fact [30] about gamma distributions. Namely, suppose that and are two non-constant, independent, positive random variables. Then and are independent if and only if and are gamma-distributed random variables with the same scale parameter. Applying the fact that and are independent if and only if is a constant random variable, the trivial solutions of part (b) are readily checked to be the only other option. ∎
In this case, by construction, we have that
This enables us to deduce from Propositions 2.8 and 4.5 the following result.
Proposition 4.6.
The following product measures satisfy .
-
(a)
For any and ,
-
(b)
For any ,
Moreover, if a product measure is invariant under , then it must be one of the above.
Proof.
The proof is same as that of Proposition 4.2. ∎
4.2.3. Invariant measures
The initial value problem for the discrete Toda equation that we consider is: for , find such that (dToda) holds for all . This was solved in [13] for initial conditions in the set
In particular, the following result was established.
Lemma 4.7 ([13, Theorem 2.5]).
If it holds that , then there exists a unique collection such that (dToda) holds for all .
As in the ultra-discrete case, in the case when a unique solution to (dToda) exists, it makes sense to define the dynamics of the system similarly to (1.3), i.e. set
In what is the main result of this section, we characterize invariant product measures for the resulting evolution.
Theorem 4.8.
Suppose that is an i.i.d. sequence with marginal given by , where the following holds: for some and ,
It is then the case that . Moreover, there are no other non-trivial measures such that is an i.i.d. sequence, with independent of , and .
Proof.
The proof is the same as that of Theorem 4.4. ∎
5. Links between discrete integrable systems
In this section, we explain how the well-known links between the systems (udKdV), (dKdV), (udToda) and (dToda) extend to invariant measures. Our results are summarised in Figure 1.
5.1. Ultra-discretization
The systems (udKdV) and (udToda) arise as ultra-discrete limits of (dKdV) and (dToda), respectively. In particular, it is straightforward to check that if
then
(5.1) |
Similarly, if
then
(5.2) |
As a consequence of the following proposition, we have that making corresponding changes of parameters for certain invariant measures for and yields invariant measures for and (see Remark 5.2).
Proposition 5.1.
-
(a)
Suppose that , where and , for some , if , if , and . It then holds that
in distribution, where .
-
(b)
Suppose that , where , for some and . It then holds that
in distribution, where .
Proof.
(a) Write . By making a standard change of variables, we see that this has density proportional to
Observe that, for , we have that . Hence, by the dominated convergence theorem, for any compact interval , we have
Moreover, if , the monotone convergence theorem yields that
Similarly,
Combining the previous three limits, the result readily follows.
(b) Writing , we find that has density proportional to
Given this, arguing similarly to the first part of the proof gives the desired conclusion. ∎
5.2. KdV-Toda correspondence
In [13], a correspondence was established between one time-step solutions of the ultra-discrete Toda equation with a particular symmetry and solutions of the ultra-discrete KdV equation (with ), and similarly for the discrete models. Here, we use these relations to connect invariant measures for the various systems.
5.2.1. Ultra-discrete case
To describe the story in the ultra-discrete case, first observe that the preserves the space . In particular, we have that
Combining the first two coordinates, we introduce an involution by setting , or equivalently,
Moreover, we note that this is simply a change of coordinates from . Indeed, if , then we have that
The above sequence of operations incorporates the ‘self-convolution’ procedure of [13, Section 6, and Proposition 6.5 in particular], with the reverse procedure from to involving the ‘splitting’ operation of [13, Section 6]. NB. The presentation of this article differs by a unimportant factor of 2 from that of [13], where such a factor was needed to define a ‘Pitman-type transformation map’. Now, it is an elementary exercise to check that the invariant measure (with and ) for of Proposition 3.2 corresponds to the following invariant measure for :
(5.3) |
Returning to the coordinates of the (udToda) system, this gives that if has the above distribution, then is invariant for . We note that this solution relates to the product invariant measure of Proposition 4.2. Indeed, it is readily checked that if with and , then
(5.4) |
in distribution as , where . Since it holds that , the left-hand side of (5.4) has a distribution that is invariant under , and thus the continuous mapping theorem implies that so does the right-hand side. Reparameterising gives that has distribution as at (5.3), which establishes that, in the case , Proposition 3.2(a)(i) can alternatively be obtained as a consequence of Proposition 4.2.
5.2.2. Discrete case
The discrete case is similar to the ultra-discrete one. Indeed, preserves the space , with
In this case, we introduce an involution by setting
and note that if , then
Again, these operations essentially describe the self-convolution procedure of [13, Section 6], with the reverse procedure from to involving the splitting procedure of [13, Section 6]. The invariant measure (with ) for of Proposition 3.9 corresponds to the following invariant measure for :
(5.5) |
Hence, if has the above distribution, then the law of is invariant under . Moreover, it is possible to check that the solution relates to the product invariant measure of Proposition 4.6. For, if , then one may verify that
(5.6) |
in distribution as , where . Since it holds that , the left-hand side of (5.6) has a distribution that is invariant under , and thus the continuous mapping theorem implies that so does the right-hand side. Reparameterising gives that has distribution as at (5.5), which establishes that, in the case , Proposition 3.9(a) can alternatively be obtained as a consequence of Proposition 4.6.
6. Connection to stochastic integrable models
In this section, we discuss links between our framework and results, and studies on stochastic integrable models. To expand slightly, stochastic two-dimensional lattice integrable (explicitly solvable) models have been intensively studied in recent years in the context of KPZ universality. These include last passage percolation with exponential/geometric weights, the log-gamma, strict-weak, beta, and inverse-beta directed random polymer models, and also higher spin vertex models. An important common property of these systems is that they admit stationary measures that satisfy an appropriate version of Burke’s property. We will describe how the arguments of Subsection 2.2 can be extended to cover the stochastic setting, and explain how this applies in a number of examples. We highlight that we are able to make explicit connections between the Toda-type systems of Section 4 and certain polymer models. This part of the study is continued in [12], wherein the techniques of this article are used to explore the stationary solutions of random polymer models and their zero-temperature limits.
A typical setting for the stochastic models of interest here is the following: for a given boundary condition , where the are random variables taking values in a space and the are random variables taking values in a space , the random variables are defined recursively via the equations:
(6.1) |
where is a deterministic function, and are i.i.d. random variables, independent of . In particular, for a given realization of the variables , we have a two-dimensional system of equations of the form described in the first sentence of the article with . For these models, we define the following notion of Burke’s property.
- Burke’s property for a stochastic model:
-
We say that the distribution of the random variables satisfies Burke’s property if:
-
•:
the sequences and are each i.i.d., and independent of each other;
-
•:
the distribution of is translation invariant, that is, for any ,
-
•:
By applying the same argument as that used to prove Proposition 2.9, we can obtain the following.
Proposition 6.1 (Burke’s property for a stochastic model).
Suppose are probability measures on respectively satisfying
(6.2) |
If are independent random variables whose marginals are and respectively, then the random variables defined by the relation (6.1) satisfy Burke’s property for a stochastic model.
Just as in the deterministic case, it is also possible to consider inhomogeneous stochastic models. For the purposes of the subsequent discussion, in this direction we suppose that are -valued random variables, are -valued random variables, are random variables, and there exists a sequence of deterministic functions
Given a (random) boundary condition , we then define by
(6.3) |
For such dynamics, we define Burke’s property as follows.
- Burke’s property for an inhomogeneous stochastic model:
-
We say that the distribution of the random variables satisfies Burke’s property if there exist a sequence of probability measures , with supported on , and , with supported on such that:
-
•:
for all ;
-
•:
for all ;
-
•:
for any , are independent random variables.
-
•:
The above notion of Burke’s property was discussed in [19] in the study of the stochastic higher spin six vertex model introduced in [7] (see Subsection 6.4 below). We can prove the following in the same way as the homogeneous case.
Proposition 6.2 (Burke’s property for an inhomogeneous stochastic model).
Suppose , , are probability measures on , , , respectively, satisfying
If , , are independent random variables whose marginals are , and , respectively, then the random variables defined by the relation (6.3) satisfy Burke’s property for an inhomogeneous stochastic model.
The type I and type II models considered in the earlier deterministic part of the article can be understood as special cases of the stochastic models in the following ways.
-
•
Firstly, the local dynamics of a type I model clearly match those of a homogeneous stochastic model for which the map at (6.1) does not depend on . More generally, one could apply Proposition 6.2 to study an inhomogeneous deterministic model. For example, if we set , then we have that and satisfy
and so there is a distribution on that satisfies the inhomogeneous version of Burke’s property.
-
•
Secondly, to connect to type II models, we observe that the condition at (6.2) matches the condition on of Proposition 2.8(a). Hence, if is given by a map
(6.4) where is a bijection (i.e. similarly to (1.4) with and ), then the detailed balance condition for the type II model given by is equivalent to (6.2). Consequently, for any type II model, we can construct stochastic counterpart by (6.4), and the detailed balance condition for implies the existence of distributions satisfying Burke’s property both for the deterministic and stochastic models. Note that the configuration for the deterministic model is , where , for (mod 2), and , for (mod 2), whereas, for the stochastic model, satisfies and for all .
We next proceed to discuss a number of examples of stochastic integrable systems. In particular, we will observe that
where is the function for directed last passage percolation, and is the function for the directed polymer with site weights (see Subsections 6.1 and 6.2, respectively). We will further see that , the function for the directed polymer with edge weights, can also written in terms of a bijection . For the latter model, the solutions of the detailed balance equation were, up to a regularity condition, characterized in [5].
6.1. Directed last passage percolation in two dimensions
In the study of directed last passage percolation on , a key quantity of interest is the partition function
where the maximum is taken over ‘up-right paths’ on , and are i.i.d. random variables. One readily sees that this partition function satisfies the following recursion:
(6.5) |
By setting and , the recursive equation at (6.5) can be rewritten as
where
In particular, , and we obtain from Proposition 2.8 that
Apart from trivial solutions, we have from Proposition 4.1 that the above identities imply that is a (possibly scaled and shifted) exponential/geometric distribution; note that when has an i.i.d. exponential/geometric distribution, the directed last passage percolation model is known to be exactly solvable. Moreover, the solution of the detailed balance equation for of Proposition 4.1 further yields the existence of the stationary distribution , , and its geometric distribution version, cf. [1].
6.2. Directed random polymer with site weights
For this model, which is a positive temperature version of directed last passage percolation, the partition function is given by
where the sum is taken over ‘up-right paths’ on , and are i.i.d. random variables. In this case, we have a recursive equation for the partition function of the form
(6.6) |
Letting , , the recursive equation (6.6) can be rewritten as
where
We thus see that , and we obtain from Proposition 2.8 that
From Proposition 4.5, we have that the only non-trivial solution to these equations has being a gamma distribution, and, similarly to the comment in the previous example, it is of note to observe that when has an i.i.d. inverse gamma distribution, the model is exactly solvable. Furthermore, it also follows from Proposition 4.5 that we have the existence of a stationary distribution with , , , cf. [33, 34].
6.3. Directed random polymer with edge weights
Similarly to the previous subsection, the model we next consider has partition function
where again the sum is taken over ‘up-right paths’ on , and
where is a positive function on , and are i.i.d. random variables. This partition function satisfies
(6.7) |
and by letting , , the recursive equation (6.7) can be rewritten as
where
Note that, whilst in the previous example we wrote in terms of in order to fit closely with the map , here we write in terms of to better fit the discussion in [5]. In particular, in [5], up to technical conditions, the authors characterize distributions , and such that . To expand on this, under the assumptions of [5], whenever and are in the support of , the function , where , has an inverse function on the support of . It follows that the function
is a bijection (on the support of ), with inverse function given by
and putting these together yields
Hence the condition is equivalent to for some , and also to
where
In [5], the authors show that Burke’s property holds for the directed random polymer with edge weights only if for some such that . NB. In this case, the above map is of the form
Moreover, they characterize all distributions satisfying Burke’s property. Up to linear transformations, these fall into one of the following four classes:
- Inverse gamma:
-
For , , i.e. ,
- Gamma:
-
For , , i.e. ,
- Beta:
-
For , , i.e. ,
- Inverse beta:
-
For , , i.e. ,
To obtain the results in the cases and , the well-known characterization of gamma distributions from [30] was applied, cf. our argument characterising the invariant measures for the discrete Toda lattice. (Note that if , then, up to inversion of the variables, the dynamics of matches that of .) In the cases and , a similar result for the beta distribution is used, see [35].
Remark 6.3.
The equation (6.7) with corresponds to a recursion equation for the distribution function of the random walk in a Beta-distributed random environment, as studied in [3]. Specifically, the environment of the latter model is given by an i.i.d. collection of random variables , and conditional on this, the process is the (discrete-time) Markov chain with transition probabilities given by
It is readily checked that satisfies
Reparameterising by setting , , we obtain (6.7).
6.4. Higher spin vertex models
In this subsection, we explain how Proposition 6.2 applies to higher spin vertex models. The state spaces for such models are given by , and for some . In the case , the dynamics of the model are given by the probabilities
for an appropriate choice of , see [7] for details. For simplicity, we consider the case and . If
and is an i.i.d. collection of uniform random variables on , we then have that
By direct computation, one can check that
for any , where is the uniform distribution on , and qNB is a -negative binomial distribution (see the appendix for details). Note in particular that is a Bernoulli distribution with parameter . In [19], the authors introduce a change of parameters from to with , , , , so that , , . With this, we have
Moreover, in [19], the parameters are allowed to be inhomogeneous, so that , and . To align with this framework, we set . It then follows that, for any fixed ,
where , .
For more general , the model is defined by a fusion operation, see [7]. This gives the stochastic matrix
for , , and so there exists such that
with i.i.d. uniform random variables on . Noting that a random variable can be written as , where (see Proposition 2.3 of [19]), the fusion procedure gives that
In [19], the inhomogeneous version was also studied in the same way as above. Namely, for for , for any fixed ,
where , . Hence Proposition 6.3 applies.
As a final remark, we note that the role of the distribution of and the function are different in the higher spin vertex model and the other models discussed here. Indeed, for models other than the higher spin vertex model, the function reflects the structure of the model, or more precisely the recursion equation of the partition function, independent of the distribution of . On the other hand, for the higher spin vertex model, the function and the distribution of do not have any meaning in themselves, but rather the pair together determines the stochastic matrix from the input to , which determines the model.
7. Iterated random functions
As noted in the introduction, our models can be understood as a special class of iterated random functions. In this section, we discuss how our contributions relate to some known results in the literature regarding such systems. To introduce iterated random functions, we will follow the notation of Diaconis and Freedman’s article [14], which is a comprehensive survey on this subject (up to its year of writing). Let be a topological space equipped with its Borel -algebra, be a measurable space, be a collection of continuous maps , and be a probability measure on . Let be an i.i.d. sequence with marginal . The object of interest is the Markov chain constructed by iterating random functions on the state space , that is
where for some . Diaconis and Freedman showed that when ‘ is contracting on average’ (see [14] for a precise definition), has a unique stationary distribution, which is independent of . We highlight that a key ingredient in the proof of this theorem is the proposition that the backward iteration defined by
converges almost surely, at an exponential rate, to a random variable that does not depend on (see [14, Proposition 5.1]).
We now explain how our setting is embedded into the iterated random function framework, starting with type I models. Recall in this case, we have an involution , and that, for a given , we are interested in the existence and uniqueness of such that
(Cf. (1.2).) Letting , and for , it is clear that if is an i.i.d. sequence with marginal and we are given , then is the Markov chain constructed by the iterating random functions . If we know that the backward iteration converges almost surely to a limit which does not depend on , then for any , the limit
also exists almost surely and does not depend on (cf. [25, Section 4]). In particular, is measurable with respect to , and for all . Setting and , it follows that
and is measurable with respect to . In conclusion, for -a.e. realization of , there exists at least one satisfying and is measurable with respect to . Moreover, the distribution of is translation invariant, being given by the stationary distribution for the Markov chain constructed by the iterated random functions .
For type II models, the story is similar. In this case we have a bijection , and taking , , and as
we can repeat the discussion of the preceding paragraph. For the ultra-discrete Toda model in particular, we have that
which can be analysed in the same way as the queue considered in [14, Section 4]. More specifically, in the latter example, the map of interest is given by
Although this is not a strict contraction, it is nonetheless shown in [14] that, under a certain condition on the distribution , which includes the case when , the backward iteration converges almost surely to a limit which does not depend on . To transfer the argument to the ultra-discrete Toda case, we first make the elementary observation that
where . We can then apply the identity given at [14, (4.4)] (that originally appeared in [16]) to obtain that the above expressions are both equal to
It readily follows that if (cf. the requirement on configurations in in Section 4), then this backward iteration converges almost-surely, for any , to the finite random variable . As is shown in [13, Theorem 2.3], this precisely corresponds to the value of given by the unique solution to the initial value problem for (udToda) with initial condition . One can similarly reconstruct , and indeed the dynamics for all time using this iterated random function approach.
Remark 7.1.
The connection between the ultra-discrete Toda lattice and queueing theory is further highlighted by a comparison of the framework and results of the present paper with those of [15]. Indeed, in the latter work, the local dynamics of the model studied precisely correspond to those given by the map , with the variables in our notation being the analogues of in that of [15]. In particular, [15] gives a version of Burke’s theorem for the queuing process in question, with exponential/geometric invariant measures. (Cf. the discussion concerning directed last passage percolation in Subsection 6.1.)
Next, returning to type I models, if the backward iteration converges, one can further consider the question of invariance. Namely, when is it the case that , as defined by , has the same distribution as , where is defined by the backward iteration? On this issue, in [25], it is shown that when:
-
(i)
the Markov chain has reversible transition probabilities,
-
(ii)
for each , the map is injective,
if we set
where is the inverse function of , then is an i.i.d. sequence with marginal . NB. It is straightforward to check that, for a type I model and a measure on such that , we almost-surely have that . It is moreover shown in [25, Theorem 4.1] that is independent of , which yields that is an i.i.d. sequence, where is defined recursively. In addition, if and are one-to-one almost surely, then the dynamics given by are ergodic (actually Bernoulli) with respect to (cf. [25, Theorem 2.2] and Theorem 2.11 above). As an example, the authors of [25] study a discrete-time version of the M/M/1 queue, the dynamics of which are equivalent to BBS started from an i.i.d. configuration. The aim of their paper was to establish the ergodicity of the dynamics, and it was left as an open problem to identify under what conditions is ergodic more generally. Whilst we do not address that question here, we do provide further examples of models satisfying the various conditions, namely the ultra-discrete and discrete KdV equations with appropriate i.i.d. marginals, as described in Section 3.
Remark 7.2.
The conditions (i) and (ii) above imply that there exists an involution (at least, on the support of appropriate measures) that is an extension of the map . More precisely, if we assume that is injective for each , and that the set is symmetric in the two coordinates, then such an is given by . Note that, even if there exists such an extension, however, we can not expect that is contracting on average in general. Indeed, although the relevant backward iteration converges, the example studied in [25] does not satisfy the latter property.
Another approach to demonstrating convergence of the backward iteration for a certain iterated random function system is set out in [37, 38]. In the latter works, a key notion is that of a ‘synchronizing sequence’, which represents a finite string such that the image of contains exactly one point. If such a string occurs infinitely often under the measure , then it is easy to see that the backward iteration converges. Observe that we have applied the same idea in the proof of Lemma 3.5, with the conditions on given in (3.6) and (3.7) being ‘synchronizing’ for the ultra-discrete KdV system with .
Finally, we further note that there has also been a series of works on the stochastic equation:
where is the ‘evolution process’, and is an unknown process, with both taking values in a compact group (see the survey [39] and the references therein). It is clear that this model is in the setting of iterated random functions with and . Moreover, it is obvious that in this case there exists an involution such that , as given by . These studies are motivated by Tsirelson’s equation, and in particular, it is shown that the Markov chains given by this type of iterated random function system can have a quite different behaviour to the models discussed above. Namely, depending on the distribution of , the Markov chain might or might not have a unique stationary distributional solution or a strong solution (i.e. for which is measurable with respect to ), and surprisingly, when the uniqueness of the stationary distributional solution holds, then there does not exist a strong solution, and on the other hand, when there is a strong solution, then there exist multiple strong solutions (for details, see [39]).
8. Open problems and conjectures
8.1. Problems for KdV- and Toda-type discrete integrable systems
Problem 8.1.
Problem 8.2.
Completely characterize the detailed balance solutions for (dKdV), i.e. extend the final claim of Proposition 3.9 to general (see Conjecture 8.6 below for our expectation in this direction). Moreover, describe a reasonable subset of when , so that the invariance and ergodicity results can be extended to these cases. (As commented above, the results of [13] do not apply.)
Problem 8.3.
Give an argument for establishing the ergodicity of invariant measures for type II models, and in particular apply this in the case of the discrete and ultra-discrete Toda lattice equations. (Ergodicity of a polymer model related to the discrete Toda lattice, cf. Subsection 6.2, is studied in [21].)
Problem 8.4.
In Section 6, we presented some basic connections between the ultra-discrete/ discrete Toda lattices and certain stochastic integrable systems that explain why the invariant measures of the corresponding systems match up. In the last few decades, an important aspect of research in stochastic integrable systems has been the development of machinery to study models in the Kardar-Parisi-Zhang (KPZ) universality class (see [6] for background). Remarkably, it has recently been seen that the KPZ fixed point can be linked to the Kadomtsev-Petviashivili (KP) equation, which is a two-dimensional version of the KdV equation [20]. These observations naturally lead one to wonder where else there might be parallels between deterministic integrable systems of KdV/Toda-type, and stochastic integrable systems in the KPZ universality class, and to what extent these might be used to transfer knowledge between the two areas.
8.2. Characterizations of standard distributions
In the course of this work, and in particular when solving the various detailed balance equations, we have applied several classical results of the form: if and are independent, then so are and , where for a given , if and only if the distribution of falls into a certain class. Perhaps the most famous result in this direction is that first proved by Kac in 1939: ‘if and are independent, then so are and if and only if both and have normal distributions with a common variance’ (see [22], as described in [18]). In this subsection, alongside recalling other known results for specific involutions or bijections , we formulate a number of natural conjectures that arise from our study. NB. In what follows, we say that random variables are ‘non-trivial’ if they are non-Dirac).
As a first example, we recall the characterization of the product of GIG and gamma distributions from [28]. Similar results are sometimes described in the literature as being of ‘Matsumoto-Yor type’, after [31], where the ‘if’ part of the result was established (see [26], for example).
Theorem 8.5 ([28]).
Let be the involution given by
Let and be non-trivial -valued independent random variables. It is then the case that are independent if and only if there exist such that
and in this case, and . Hence, if moreover has the same distribution as , then and for some .
As a direct corollary, by making the change of variables , one can check a similar result for the involution given by
(8.1) |
In this case, the random variables and have the same distribution as in Theorem 8.5, but and . Now, the above map is precisely , and indeed it was the conclusion of [28] that we applied in the proof of Proposition 3.9 to characterize the solutions of the detailed balance equation for with . In light of the conclusion of Proposition 3.9, we conjecture that for general , a similar result holds.
Conjecture 8.6.
Let with , and recall the definition of from (dKdV). Let and be non-trivial -valued independent random variables. It is then the case that are independent if and only if there exist such that
and in this case and . Hence, if moreover has the same distribution as , then , for some .
The next statement was applied in the proof of Proposition 4.5 when characterising the solutions of the detailed balance equation for the discrete Toda system. Moreover, this and the subsequent two results were used in [5] to characterize directed random polymer models having stationary measures satisfying Burke’s property. We note that Corollary 8.8 is a direct consequence of Theorem 8.7.
Theorem 8.7 ([30]).
Let be the bijection given by
NB. . Let and be non-trivial -valued independent random variables. It is then the case that are independent if and only if there exist such that
and in this case, and .
Corollary 8.8.
Let be the bijection given by
Let and be non-trivial -valued and -valued, respectively, independent random variables. It is then the case that are independent if and only if there exist such that
and in this case, and .
Theorem 8.9 ([35]).
Let be the involution given by
Let and be non-trivial -valued independent random variables. It is then the case that are independent if and only if there exist such that
and in this case, and . Hence, if moreover has the same distribution as , then , .
Just as we related solutions of the detailed balance equations for the discrete and ultra-discrete KdV- and Toda-type systems in Section 5, it is possible to ultra-discretize the above statements, and this leads to a number of further conjectures. To do this, we transform variables taking values in to via the bijection (the inverse of which is ). The ultra-discretization procedure is then given by applying the limit
where we take for Conjectures 8.10 and 8.11, and in the remaining cases. Precisely, we arrive at Conjecture 8.10 from the map at (8.1), Conjecture 8.11 from Conjecture 8.6, Theorem 8.13/Corollary 8.14 from Theorem 8.7/Corollary 8.8, and Conjecture 8.15 from Theorem 8.9.
Conjecture 8.10.
If , then is an involution. For any , is an involution, and for any , is a bijection. Let and be absolutely continuous -valued independent random variables satisfying . It is then the case that are independent if and only if there exist such that
and in this case, , . Hence, if moreover has the same distribution as , then , for some .
Conjecture 8.11.
If for some , then is an involution. Also, for any , is an involution, and for any , is a bijection. Let and be absolutely continuous -valued independent random variables satisfying . It is then the case that are independent if and only if there exist and such that
and in this case, , . Hence, if moreover has the same distribution as , then , for some .
Remark 8.12.
It is also possible to write down discrete versions of the previous two conjectures, replacing the distribution with the one, cf. Proposition 3.2. The appearance of the bipartite version in the discrete case of these results is an interesting consequence of the particular structure of the ultra-discrete KdV system. Similarly, one might also make a discrete version of Conjecture 8.15 below involving the distribution.
Theorem 8.13 ([8]).
Let be the bijection given by
NB. . Let and be non-trivial -valued independent random variables. It is then the case that are independent if and only if there exist such that
or such that
and in this case , , or , , respectively.
Corollary 8.14.
Let be the bijection given by
Let and be non-trivial -valued independent random variables. It is then the case that are independent if and only if there exist such that
or such that
and in this case , , or , , respectively.
Conjecture 8.15.
Let be the involution given by
Let and be absolutely continuous -valued independent random variables. It is then the case that are independent if and only if there exist such that
and in this case, , . Hence, if moreover has the same distribution as , then , .
Acknowledgements
This research was supported by JSPS Grant-in-Aid for Scientific Research (B), 19H01792. The research of DC was also supported by JSPS Grant-in-Aid for Scientific Research (C), 19K03540, and the Research Institute for Mathematical Sciences, an International Joint Usage/Research Center located in Kyoto University. This work was completed while MS was kindly being hosted by the Courant Institute, New York University.
Appendix A Probability distributions
In the following list, we give definitions of the various probability distributions that appear within this article.
- Shifted truncated exponential distribution:
-
For with , the shifted truncated exponential distribution with parameters , which we denote , has density
where is a normalizing constant.
- Shifted exponential distribution:
-
For , , the shifted exponential distribution with parameters , which we denote , has density
where is a normalizing constant. We use the convention that when .
- Shifted scaled (truncated bipartite) geometric distribution:
-
For , , such that , and , we say a random variable has shifted scaled truncated bipartite geometric distribution with parameters , , , and if
where and is a normalising constant; in this case we write . Note that, if , then we require that for the distribution to be defined. We observe that is simply the distribution of the usual parameter geometric distribution conditioned to take a value in . In the special case when , , , we say that has shifted scaled geometric distribution with parameters , and , and write .
- Asymmetric Laplace distribution:
-
For , the asymmetric Laplace distribution with parameters , which we denote , has density
where is a normalizing constant.
- Scaled discrete asymmetric Laplace distribution:
-
For and , we say a random variable has scaled discrete asymmetric Laplace distribution with parameters if
where is a normalizing constant; in this case we write .
- Gamma distribution:
-
For , the gamma distribution with parameters , which we denote , has density
where is a normalizing constant.
- Inverse gamma distribution:
-
For , the inverse gamma distribution with parameters , which we denote , has density
where is a normalizing constant.
- Generalized inverse Gaussian distribution:
-
For , , the generalized inverse Gaussian distribution with parameters , which we denote , has density
where is a normalizing constant. We use the convention that .
- Beta distribution:
-
For , the beta distribution with parameters , which we denote , has density
where is a normalizing constant.
- -negative binomial distribution:
-
Fix . For or , for some , we say a random variable has -negative binomial distribution with parameters if
where for , , and is a normalising constant, which can be given explicitly as ; in this case we write . Note that, if , then the support of is , and if and for some , then the support of is .
References
- [1] M. Balázs, E. Cator, and T. Seppäläinen, Cube root fluctuations for the corner growth model associated to the exclusion process, Electron. J. Probab. 11 (2006), no. 42, 1094–1132.
- [2] K. B. Bao and C. Noack, Characterizations of the generalized inverse Gaussian, assymetric Laplace, and shifted (truncated) exponential laws via independence properties, preprint appears at arXiv:2107.01394, 2021.
- [3] G. Barraquand and I. Corwin, Random-walk in beta-distributed random environment, Probab. Theory Related Fields 167 (2017), no. 3-4, 1057–1116.
- [4] P. J. Burke, The output of a queuing system, Operations Res. 4 (1956), 699–704 (1957).
- [5] H. Chaumont and C. Noack, Characterizing stationary dimensional lattice polymer models, Electron. J. Probab. 23 (2018), Paper No. 38, 19.
- [6] I. Corwin, The Kardar-Parisi-Zhang equation and universality class, Random Matrices Theory Appl. 1 (2012), no. 1, 1130001, 76.
- [7] I. Corwin and L. Petrov, Stochastic higher spin vertex models on the line, Comm. Math. Phys. 343 (2016), no. 2, 651–700.
- [8] G. B. Crawford, Characterization of geometric and exponential distributions, Ann. Math. Statist. 37 (1966), 1790–1795.
- [9] D. A. Croydon, T. Kato, M. Sasada, and S. Tsujimoto, Dynamics of the box-ball system with random initial conditions via Pitman’s transformation, to appear in Mem. Amer. Math. Soc., preprint appears at arXiv:1806.02147, 2018.
- [10] D. A. Croydon and M. Sasada, Duality between box-ball systems of finite box and/or carrier capacity, RIMS Kôkyûroku Bessatsu B79 (2020), 63–107.
- [11] by same author, Discrete integrable systems and Pitman’s transformation, Adv. Stud. Pure Math. 87 (2021), 381–402.
- [12] by same author, On the stationary solutions of random polymer models and their zero-temperature limits, preprint appears at arXiv:2104.03458, 2021.
- [13] D. A. Croydon, M. Sasada, and S. Tsujimoto, General solutions for KdV- and Toda-type discrete integrable systems based on path encodings, preprint appears at arXiv:2011.00690, 2020.
- [14] P. Diaconis and D. Freedman, Iterated random functions, SIAM Rev. 41 (1999), no. 1, 45–76.
- [15] M. Draief, J. Mairesse, and N. O’Connell, Queues, stores, and tableaux, J. Appl. Probab. 42 (2005), no. 4, 1145–1167.
- [16] W. Feller, An introduction to probability theory and its applications. Vol. II, Second edition, John Wiley & Sons, Inc., New York-London-Sydney, 1971. MR 0270403
- [17] T. S. Ferguson, A characterization of the exponential distribution, Ann. Math. Statist. 35 (1964), 1199–1207.
- [18] by same author, A characterization of the geometric distribution, Amer. Math. Monthly 72 (1965), 256–260.
- [19] T. Imamura, M. Mucciconi, and T. Sasamoto, Stationary stochastic Higher Spin Six Vertex Model and -Whittaker measure, Probab. Theory Related Fields 177 (2020), no. 3-4, 923–1042.
- [20] D. Remenik J. Quastel, KP governs random growth off a one dimensional substrate, preprint appears at arXiv:1908.10353, 2019.
- [21] C. Janjigian and F. Rassoul-Agha, Uniqueness and Ergodicity of Stationary Directed Polymers on , J. Stat. Phys. 179 (2020), no. 3, 672–689.
- [22] M. Kac, On a characterization of the normal distribution, Amer. J. Math. 61 (1939), 726–728.
- [23] S. Kakei, J. J. C. Nimmo, and R. Willox, Yang-Baxter maps from the discrete BKP equation, SIGMA Symmetry Integrability Geom. Methods Appl. 6 (2010), Paper 028, 11.
- [24] O. Kallenberg, Foundations of modern probability, second ed., Probability and its Applications (New York), Springer-Verlag, New York, 2002.
- [25] M. Keane and N. O’Connell, The queue is Bernoulli, Colloq. Math. 110 (2008), no. 1, 205–210.
- [26] A. E. Koudou and P. Vallois, Independence properties of the Matsumoto-Yor type, Bernoulli 18 (2012), no. 1, 119–136.
- [27] G. Letac and V. Seshadri, A characterization of the generalized inverse Gaussian distribution by continued fractions, Z. Wahrsch. Verw. Gebiete 62 (1983), no. 4, 485–489.
- [28] G. Letac and J. Wesołowski, An independence property for the product of GIG and gamma laws, Ann. Probab. 28 (2000), no. 3, 1371–1383.
- [29] P. Loya, Amazing and aesthetic aspects of analysis, Springer, New York, 2017.
- [30] E. Lukacs, A characterization of the gamma distribution, Ann. Math. Statist. 26 (1955), 319–324.
- [31] H. Matsumoto and M. Yor, An analogue of Pitman’s theorem for exponential Wiener functionals. II. The role of the generalized inverse Gaussian laws, Nagoya Math. J. 162 (2001), 65–86.
- [32] V. G. Papageorgiou, A. G. Tongas, and A. P. Veselov, Yang-Baxter maps and symmetries of integrable equations on quad-graphs, J. Math. Phys. 47 (2006), no. 8, 083502, 16.
- [33] T. Seppäläinen, Scaling for a one-dimensional directed polymer with boundary conditions, Ann. Probab. 40 (2012), no. 1, 19–73.
- [34] by same author, Erratum to “Scaling for a one-dimensional directed polymer with boundary conditions”, Ann. Probab. 45 (2017), no. 3, 2056–2058.
- [35] V. Seshadri and J. Wesołowski, Constancy of regressions for beta distributions, Sankhyā 65 (2003), no. 2, 284–291.
- [36] T. Tokihiro, Ultradiscrete systems (cellular automata), Discrete integrable systems, Lecture Notes in Phys., vol. 644, Springer, Berlin, 2004, pp. 383–424.
- [37] K. Yano, Random walk in a finite directed graph subject to a road coloring, J. Theoret. Probab. 26 (2013), no. 1, 259–283.
- [38] K. Yano and K. Yasutomi, Realization of an ergodic Markov chain as a random walk subject to a synchronizing road coloring, J. Appl. Probab. 48 (2011), no. 3, 766–777.
- [39] K. Yano and M. Yor, Around Tsirelson’s equation, or: The evolution process may not explain everything, Probab. Surv. 12 (2015), 1–12.