This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Detailed balance and invariant measures
for discrete KdV- and Toda-type systems

David A. Croydon Research Institute for Mathematical Sciences, Kyoto University, Kyoto 606-8502, Japan croydon@kurims.kyoto-u.ac.jp  and  Makiko Sasada Graduate School of Mathematical Sciences, University of Tokyo, 3-8-1, Komaba, Meguro-ku, Tokyo, 153–8914, Japan sasada@ms.u-tokyo.ac.jp
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 lattice
2010 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:

xnt+2\textstyle{{\begin{array}[]{c}\vdots\\ x_{n}^{t+2}\end{array}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}xn+1t+2\textstyle{{\begin{array}[]{c}\vdots\\ x_{n+1}^{t+2}\end{array}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}un1t+1\textstyle{\cdots u_{n-1}^{t+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fnt+1F_{n}^{t+1}Fn+1t+1F_{n+1}^{t+1}unt+1\textstyle{u_{n}^{t+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fn+1t+1F_{n+1}^{t+1}un+1t+1\textstyle{u_{n+1}^{t+1}\cdots}xnt+1\textstyle{x_{n}^{t+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}xn+1t+1\textstyle{x_{n+1}^{t+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}un1t\textstyle{\cdots u_{n-1}^{t}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}FntF_{n}^{t}Fn+1t+1F_{n+1}^{t+1}unt\textstyle{u_{n}^{t}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fn+1tF_{n+1}^{t}Fn+1t+1F_{n+1}^{t+1}un+1t\textstyle{u_{n+1}^{t}\cdots}xnt\textstyle{\begin{array}[]{c}x_{n}^{t}\\ \vdots\end{array}}xn+1t\textstyle{\begin{array}[]{c}x_{n+1}^{t}\\ \vdots\end{array}}

We will think of nn as the spatial coordinate, and tt as the temporal one. Moreover, the variables (xnt)n(x^{t}_{n})_{n\in\mathbb{Z}} will represent the configuration at time tt, and (unt)n(u^{t}_{n})_{n\in\mathbb{Z}} a collection of auxiliary variables through which the dynamics from time tt to t+1t+1 are defined. As for the state spaces of the variables (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in\mathbb{Z}} and maps (Fnt)n,t(F^{t}_{n})_{n,t\in\mathbb{Z}}, we specialize to two cases:

Type I (homogeneous) model:

The variables xntx_{n}^{t} take values in a common Polish space 𝒳0\mathcal{X}_{0}. Similarly, the variables untu_{n}^{t} take values in a common Polish space 𝒰0\mathcal{U}_{0}. Moreover, FntFF_{n}^{t}\equiv F for some involution F:𝒳0×𝒰0𝒳0×𝒰0F:\mathcal{X}_{0}\times\mathcal{U}_{0}\rightarrow\mathcal{X}_{0}\times\mathcal{U}_{0}.

Type II (alternating/bipartite) model:

The variables xntx_{n}^{t} take values in a Polish space 𝒳0\mathcal{X}_{0} if n+t=0n+t=0 (mod 2), and in Polish space 𝒳~0\tilde{\mathcal{X}}_{0} otherwise. Similarly, the variables untu_{n}^{t} take values in a Polish space 𝒰~0\tilde{\mathcal{U}}_{0} if n+t=0n+t=0 (mod 2), and in Polish space 𝒰0{\mathcal{U}}_{0} otherwise. Moreover, FntFF_{n}^{t}\equiv F_{*} for some bijection F:𝒳0×𝒰0𝒳~0×𝒰~0F_{*}:\mathcal{X}_{0}\times\mathcal{U}_{0}\rightarrow\tilde{\mathcal{X}}_{0}\times\tilde{\mathcal{U}}_{0} if n+t=0n+t=0 (mod 2), and FntF1F_{n}^{t}\equiv F_{*}^{-1} 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 𝒳\mathcal{X}^{*} be the set of (xn)n(x_{n})_{n\in\mathbb{Z}} in 𝒳0\mathcal{X}_{0}^{\mathbb{Z}} for a type I model, or (𝒳0×𝒳~0)(\mathcal{X}_{0}\times\tilde{\mathcal{X}}_{0})^{\mathbb{Z}} for a type II model, for which there is a unique solution to the initial value problem:

(1.1) {Fnt(xnt,un1t)=(xnt+1,unt),n,t,xn0=xn,n.\begin{cases}F_{n}^{t}(x^{t}_{n},u^{t}_{n-1})=(x^{t+1}_{n},u^{t}_{n}),&\forall n,t\in{\mathbb{Z}},\\ x^{0}_{n}=x_{n},&\forall n\in{\mathbb{Z}}.\end{cases}

We then define a function UU on 𝒳\mathcal{X}^{*} by supposing x=(xn)n(Un(x))n:=(un0)nx=(x_{n})_{n\in{\mathbb{Z}}}\mapsto(U_{n}(x))_{n\in{\mathbb{Z}}}:=(u_{n}^{0})_{n\in{\mathbb{Z}}}, where (un0)n(u_{n}^{0})_{n\in{\mathbb{Z}}} is given by the unique solution of the initial value problem (1.1) with xn0=xnx^{0}_{n}=x_{n}. For future convenience, we observe that (Un(x))n(U_{n}(x))_{n\in{\mathbb{Z}}} clearly solves

(1.2) (Fn0)(2)(xn,Un1(x))=Un(x),n,\left(F_{n}^{0}\right)^{(2)}\left(x_{n},U_{n-1}(x)\right)=U_{n}(x),\qquad\forall n\in{\mathbb{Z}},

where we use a superscript (i)(i) to represent the iith coordinate of a map. Finally, we define an operator 𝒯\mathcal{T} yielding the one time-step dynamics on 𝒳\mathcal{X}^{*} by supposing 𝒯(x)=(𝒯(x)n)n\mathcal{T}(x)=(\mathcal{T}(x)_{n})_{n\in{\mathbb{Z}}} is given by

(1.3) 𝒯(x)n={(Fn0)(1)(xn,Un1(x))=xn1,for a type I model,(Fn+10)(1)(xn+1,Un(x))=xn+11,for a type II model,\mathcal{T}(x)_{n}=\begin{cases}\left(F_{n}^{0}\right)^{(1)}\left(x_{n},U_{n-1}(x)\right)=x^{1}_{n},&\text{for a type I model,}\\ \left(F_{n+1}^{0}\right)^{(1)}\left(x_{n+1},U_{n}(x)\right)=x^{1}_{n+1},&\text{for a type II model,}\\ \end{cases}

where (xn1)n(x_{n}^{1})_{n\in{\mathbb{Z}}} is given by the unique solution of the initial value problem (1.1) with xn0=xnx^{0}_{n}=x_{n}. (The shift in the index nn is included in type II models to ensure that the elements of x1x^{1} and x0x^{0} that are in the spaces 𝒳0\mathcal{X}_{0} and 𝒳~0\tilde{\mathcal{X}}_{0} are the same.) Note that we define the one time-step dynamics similarly on the set 𝒳!\mathcal{X}^{\exists!} of configurations (xn)n(x_{n})_{n\in\mathbb{Z}} for which there is a unique solution (Un(x))n(U_{n}(x))_{n\in{\mathbb{Z}}} to (1.2). (NB. It is neither the case that 𝒳!𝒳\mathcal{X}^{\exists!}\subseteq\mathcal{X}^{*} nor 𝒳𝒳!\mathcal{X}^{*}\subseteq\mathcal{X}^{\exists!} in general, though on 𝒳𝒳!\mathcal{X}^{*}\cap\mathcal{X}^{\exists!} the two definitions of 𝒯\mathcal{T} 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 𝒳\mathcal{X}^{*} are invariant under 𝒯\mathcal{T} 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 (μ,ν)(\mu,\nu) on 𝒳0\mathcal{X}_{0} and 𝒰0\mathcal{U}_{0} is said to satisfy the detailed balance condition if

F(μ×ν)=μ×ν,F(\mu\times\nu)=\mu\times\nu,

where we define F(μ×ν):=(μ×ν)F1F(\mu\times\nu):=(\mu\times\nu)\circ F^{-1}.

Detailed balance condition for a type II model:

A quadruplet of probability measures (μ\mu,ν\nu, μ~\tilde{\mu},ν~\tilde{\nu}) on 𝒳0\mathcal{X}_{0}, 𝒰0\mathcal{U}_{0}, 𝒳~0\tilde{\mathcal{X}}_{0} and 𝒰~0\tilde{\mathcal{U}}_{0} is said to satisfy the detailed balance condition if

F(μ×ν)=μ~×ν~.F_{*}(\mu\times\nu)=\tilde{\mu}\times\tilde{\nu}.

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).
  1. (a)

    Type I model. Suppose μ\mu is a probability measure on 𝒳0\mathcal{X}_{0} and μ(𝒳)=1\mu^{{\mathbb{Z}}}(\mathcal{X}^{*})=1. It is then the case that 𝒯μ=μ\mathcal{T}\mu^{{\mathbb{Z}}}=\mu^{{\mathbb{Z}}} if and only if there exists a probability measure ν\nu on 𝒰0\mathcal{U}_{0} such that the pair (μ,ν)(\mu,\nu) satisfies the detailed balance condition. Moreover, when this holds, ν\nu is the distribution of Un(x)U_{n}(x) for each nn, where xx is distributed according to μ\mu^{{\mathbb{Z}}}.

  2. (b)

    Type II model. Suppose μ\mu, μ~\tilde{\mu} are probability measures on 𝒳0\mathcal{X}_{0}, 𝒳~0\tilde{\mathcal{X}}_{0} and (μ×μ~)(𝒳)=1(\mu\times\tilde{\mu})^{{\mathbb{Z}}}(\mathcal{X}^{*})=1. It is then the case that 𝒯(μ×μ~)=(μ×μ~)\mathcal{T}(\mu\times\tilde{\mu})^{{\mathbb{Z}}}=(\mu\times\tilde{\mu})^{{\mathbb{Z}}} if and only if there exists probability measures ν\nu, ν~\tilde{\nu} on 𝒰0\mathcal{U}_{0}, 𝒰~0\tilde{\mathcal{U}}_{0}, respectively, such that the quadruplet of probability measures (μ,ν,μ~,ν~)(\mu,\nu,\tilde{\mu},\tilde{\nu}) satisfies the detailed balance condition. Moreover, when this holds, then ν\nu, ν~\tilde{\nu} are the distributions of U2n1(x)U_{2n-1}(x), U2n(x)U_{2n}(x), respectively, for each nn, where xx is distributed according to (μ×μ~)(\mu\times\tilde{\mu})^{{\mathbb{Z}}}.

We remark that the above theorem does not in itself provide a truly local criteria for invariance of homogeneous/alternating product measures under 𝒯\mathcal{T}. Indeed, the condition that μ(𝒳)=1\mu^{{\mathbb{Z}}}(\mathcal{X}^{*})=1 or (μ×μ~)(𝒳)=1(\mu\times\tilde{\mu})^{{\mathbb{Z}}}(\mathcal{X}^{*})=1 depends on knowledge of the global dynamics, and in particular a suitably accessible description of 𝒳\mathcal{X}^{*}. 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 μ\mu^{\mathbb{Z}} or (μ×μ~)(\mu\times\tilde{\mu})^{\mathbb{Z}} under 𝒯\mathcal{T} is shown to be equivalent to the detailed balance condition holding with ν\nu, ν~\tilde{\nu} given by the relevant marginals of (Un(x))n(U_{n}(x))_{n\in\mathbb{Z}}. Since it is not trivial to deduce the distribution of Un(x)U_{n}(x) from μ\mu or μ×μ~\mu\times\tilde{\mu} 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 𝒳\mathcal{X}^{*} of homogeneous/alternating product form induce stationary/alternating measures of (xnt,unt)n,t(x_{n}^{t},u_{n}^{t})_{n,t\in{\mathbb{Z}}} 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 F(μ×ν)=μ×νF(\mu\times\nu)=\mu\times\nu or F(μ×ν)=μ~×ν~F_{*}(\mu\times\nu)=\tilde{\mu}\times\tilde{\nu}. 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 F:𝒳~0×𝒳0×𝒰0𝒳~0×𝒳0×𝒰0F:\tilde{\mathcal{X}}_{0}\times\mathcal{X}_{0}\times\mathcal{U}_{0}\to\tilde{\mathcal{X}}_{0}\times\mathcal{X}_{0}\times\mathcal{U}_{0} defined by

(1.4) F(a,b,c):=(F(1)(b,c),F1(a,F(2)(b,c)))F(a,b,c):=\left(F_{*}^{(1)}(b,c),F_{*}^{-1}\left(a,F_{*}^{(2)}\left(b,c\right)\right)\right)

gives the appropriate dynamics. For a general involution of this form, we show that invariance under FF, i.e.

(1.5) F(μ~×μ×ν)=μ~×μ×ν,F\left(\tilde{\mu}\times\mu\times\nu\right)=\tilde{\mu}\times\mu\times\nu,

is equivalent to the detailed balance condition for FF_{*}, i.e. F(μ×ν)=μ~×ν~F_{*}(\mu\times\nu)=\tilde{\mu}\times\tilde{\nu} for some ν~\tilde{\nu}, and indeed that both these conditions are equivalent to

(1.6) F(2,3)(μ~×μ×ν)=μ×ν.F^{(2,3)}\left(\tilde{\mu}\times\mu\times\nu\right)=\mu\times\nu.

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 FntF_{n}^{t} 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 Un1UnU_{n-1}\mapsto U_{n} based on the random function fn,xn:=(Fn0)(2)(xn,)f_{n,x_{n}}:=(F_{n}^{0})^{(2)}(x_{n},\cdot). Such systems arise in many settings, and there are a number of important problems that arise for them, such as the (xm)mn(x_{m})_{m\leq n}-measurability of UnU_{n}. Moreover, if (xn)n(x_{n})_{n\in\mathbb{Z}} is an independent sequence, then UnU_{n} 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 𝒳\mathcal{X}^{*} of configurations for which there exists a unique solution to the initial value problem (1.1), the function UU, and the operator 𝒯\mathcal{T}, 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 𝒯\mathcal{T} in examples. Finally, in Subsection 2.3, we develop an argument for checking the ergodicity of such invariant measures under 𝒯\mathcal{T} 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.
  1. (a)

    Type I model. Suppose μ\mu is a probability measure on 𝒳0\mathcal{X}_{0} and μ(𝒳)=1\mu^{{\mathbb{Z}}}(\mathcal{X}^{*})=1. Let ν\nu be the distribution of U1(x)U_{-1}(x), where xx is distributed according to μ\mu^{{\mathbb{Z}}}. It is then the case that 𝒯μ=μ\mathcal{T}\mu^{{\mathbb{Z}}}=\mu^{{\mathbb{Z}}} if and only if the pair (μ,ν)(\mu,\nu) satisfies the detailed balance condition.

  2. (b)

    Type II model. Suppose μ\mu, μ~\tilde{\mu} are probability measures on 𝒳0\mathcal{X}_{0}, 𝒳~0\tilde{\mathcal{X}}_{0} and (μ×μ~)(𝒳)=1(\mu\times\tilde{\mu})^{{\mathbb{Z}}}(\mathcal{X}^{*})=1. Let ν\nu, ν~\tilde{\nu} be the distributions of U1(x)U_{-1}(x), U0(x)U_{0}(x), respectively, where xx is distributed according to (μ×μ~)(\mu\times\tilde{\mu})^{{\mathbb{Z}}}. It is then the case that 𝒯(μ×μ~)=(μ×μ~)\mathcal{T}(\mu\times\tilde{\mu})^{{\mathbb{Z}}}=(\mu\times\tilde{\mu})^{{\mathbb{Z}}} if and only if the quadruplet of probability measures (μ,ν,μ~,ν~)(\mu,\nu,\tilde{\mu},\tilde{\nu}) satisfies the detailed balance condition.

Remark 2.2.

Let 𝒳U\mathcal{X}^{U} be a set of configurations (xn)n(x_{n})_{n\in\mathbb{Z}} for which there is a solution (Un(x))n(U_{n}(x))_{n\in{\mathbb{Z}}} to (1.2) for which UnU_{n} is a function of (xm)mn(x_{m})_{m\leq n} for all nn, and Un(x)=θnU0(x)=U0(θnx)U_{n}(x)=\theta^{n}U_{0}(x)=U_{0}(\theta^{n}x) for a type I model, and U2n=θ2nU0=U0θ2nU_{2n}=\theta^{2n}U_{0}=U_{0}\theta^{2n}, U2n+1=θ2nU1=U1θ2nU_{2n+1}=\theta^{2n}U_{1}=U_{1}\theta^{2n} for a type II model, where θ\theta is the usual shift operator. Moreover, assume that 𝒯𝒳U𝒳U\mathcal{T}\mathcal{X}^{U}\subseteq\mathcal{X}^{U}, and 𝒳U=𝒳U\mathcal{R}\mathcal{X}^{U}=\mathcal{X}^{U}, where 𝒯=𝒯U\mathcal{T}=\mathcal{T}^{U} depends on UU through (1.3), and xn:=x1n\mathcal{R}x_{n}:=x_{1-n} for a type I model and xn:=xn\mathcal{R}x_{n}:=x_{-n} for a type II model. If 𝒯𝒯\mathcal{T}\mathcal{R}\mathcal{T}\mathcal{R} is the identity map on 𝒳U\mathcal{X}^{U}, then Theorem 2.1 holds when we replace 𝒳\mathcal{X}^{*} by 𝒳U\mathcal{X}^{U}. It might be easier to find a space 𝒳U\mathcal{X}^{U} than 𝒳\mathcal{X}^{*} in some cases.

Towards proving Theorem 2.1, we start by setting out a lemma on the measurability of xntx^{t}_{n} and untu^{t}_{n} in terms of the initial configuration xnx_{n}. This is stated in terms of functions XntX^{t}_{n} and UntU^{t}_{n} on 𝒳\mathcal{X}^{*} that are defined via the relation

(Xnt(x),Unt(x))=(xnt,unt),x𝒳,n,t,\left(X^{t}_{n}(x),U^{t}_{n}(x)\right)=(x^{t}_{n},u^{t}_{n}),\qquad\forall x\in\mathcal{X}^{*},\>n,t\in{\mathbb{Z}},

where (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}} is the unique solution of (1.1) with initial condition xx.

Lemma 2.3.

Let mm\in{\mathbb{Z}}.

  1. (a)

    For any nmn\leq m and t0t\geq 0, XntX^{t}_{n} and UntU^{t}_{n} are measurable with respect to (xn)nm(x_{n})_{n\leq m}.

  2. (b)

    For any nm+1n\geq m+1 and t0t\leq 0, XntX^{t}_{n} and Un1t1U^{t-1}_{n-1} are measurable with respect to (xn)nm+1(x_{n})_{n\geq m+1}.

Proof.

(a) Suppose there exist x=(xn)n{x}=(x_{n})_{n\in{\mathbb{Z}}} and y=(yn)ny=(y_{n})_{n\in{\mathbb{Z}}} in 𝒳\mathcal{X}^{*} such that xn=ynx_{n}=y_{n} for all nmn\leq m, but Xnt(x)Xnt(y)X^{t}_{n}({x})\neq X^{t}_{n}({y}) or Unt(x)Unt(y)U^{t}_{n}({x})\neq U^{t}_{n}({y}) for some nmn\leq m, t0t\geq 0. We then define:

{x¯nt:=Xnt(y),u¯nt:=Unt(y),nm,t0;x¯nt:=xn,n>m,t=0;x¯nt:=Xnt(x),u¯nt:=Unt(x),n,t<0.\left\{\begin{array}[]{ll}\bar{x}^{t}_{n}:=X^{t}_{n}({y}),\>\bar{u}^{t}_{n}:=U^{t}_{n}({y}),&{n\leq m,\>t\geq 0;}\\ \bar{x}^{t}_{n}:=x_{n},&{n>m,\>t=0;}\\ \bar{x}^{t}_{n}:=X^{t}_{n}({x}),\>\bar{u}^{t}_{n}:=U^{t}_{n}({x}),&{n\in{\mathbb{Z}},\>t<0.}\end{array}\right.

Moreover, for n>mn>m, t>0t>0, it is clear from the lattice structure that there is a unique solution to (x¯nt,u¯nt1):=Fnt1(x¯nt1,u¯n1t1)(\bar{x}^{t}_{n},\bar{u}^{t-1}_{n}):=F_{n}^{t-1}(\bar{x}^{t-1}_{n},\bar{u}^{t-1}_{n-1}) that is consistent with the previous definitions. Recursively, we have that (x¯nt,u¯nt)n,t(\bar{x}^{t}_{n},\bar{u}^{t}_{n})_{n,t\in{\mathbb{Z}}} solves (1.1) with initial condition xx. Since x¯ntxnt\bar{x}^{t}_{n}\neq x^{t}_{n} or u¯ntunt\bar{u}^{t}_{n}\neq u^{t}_{n} for some nmn\leq m, t0t\geq 0 by assumption, this contradicts the uniqueness of the solution of (1.1) for x𝒳{x}\in\mathcal{X}^{*}. Hence we conclude that XntX^{t}_{n} and UntU^{t}_{n} are measurable with respect to (xn)nm(x_{n})_{n\leq m}.
(b) Appealing to the symmetry of the map (xnt,unt)(x1n1t,unt)(x^{t}_{n},u^{t}_{n})\to(x^{1-t}_{1-n},u^{-t}_{-n}), 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 PP supported on 𝒳\mathcal{X}^{*}, we denote by 𝐏P\mathbf{P}_{P} the probability distribution of (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}}, as defined by the initial value problem (1.1), for which the marginal of (xn0)n(x^{0}_{n})_{n\in{\mathbb{Z}}} is given by PP. We define a spatial shift θ\theta on lattice variables by setting

θ((xnt,unt)n,t):=(xn+1t,un+1t)n,t.\theta\left((x^{t}_{n},u^{t}_{n})_{n,t\in\mathbb{Z}}\right):=\left(x^{t}_{n+1},u^{t}_{n+1}\right)_{n,t\in{\mathbb{Z}}}.

Slightly abusing notation, for elements x𝒳x\in\mathcal{X}^{*}, we similarly suppose θ(x)n=xn+1\theta(x)_{n}=x_{n+1}. The corresponding temporal shift TT is given by

T((xnt,unt)n,t):=(xnt+1,unt+1)n,t.T\left((x^{t}_{n},u^{t}_{n})_{n,t\in\mathbb{Z}}\right):=\left(x^{t+1}_{n},u^{t+1}_{n}\right)_{n,t\in{\mathbb{Z}}}.

Note that if we consider TT as the map on 𝒳\mathcal{X}^{*} given by T(x)n=xn1T(x)_{n}=x^{1}_{n}, then the definition of the dynamics at (1.3) means that, for x𝒳x\in\mathcal{X}^{*},

{𝒯(x)=T(x),for a type I model;𝒯(x)=θT(x),for a type II model.\left\{\begin{array}[]{ll}\mathcal{T}(x)=T(x),&\hbox{for a type I model;}\\ \mathcal{T}(x)=\theta\circ T(x),&\hbox{for a type II model.}\end{array}\right.

NB. From this description, it is easy to see that 𝒯\mathcal{T} is a bijection, with inverse operation 𝒯1=𝒯\mathcal{T}^{-1}=\mathcal{R}\mathcal{T}\mathcal{R}, where \mathcal{R} is defined as in Remark 2.2.

Lemma 2.4.

Let PP be a probability measure supported on 𝒳\mathcal{X}^{*}.

  1. (a)

    For a type I model, 𝒯P=P\mathcal{T}P=P if and only if T𝐏P=𝐏PT\mathbf{P}_{P}=\mathbf{P}_{P}. Also, θP=P\theta P=P if and only if θ𝐏P=𝐏P\theta\mathbf{P}_{P}=\mathbf{P}_{P}.

  2. (b)

    For a type II model, 𝒯P=P\mathcal{T}P=P if and only if θT𝐏P=𝐏P\theta\circ T\mathbf{P}_{P}=\mathbf{P}_{P}. Also, θ2P=P\theta^{2}P=P if and only if θ2𝐏P=𝐏P\theta^{2}\mathbf{P}_{P}=\mathbf{P}_{P}.

Proof.

(a) If 𝒯P=P\mathcal{T}P=P or T𝐏P=𝐏PT\mathbf{P}_{P}=\mathbf{P}_{P} holds, then 𝒯P(𝒳)=1\mathcal{T}P(\mathcal{X}^{*})=1, and so 𝐏𝒯P\mathbf{P}_{\mathcal{T}P} is well-defined. The claim then follows from the fact that T𝐏P=𝐏𝒯PT\mathbf{P}_{P}=\mathbf{P}_{\mathcal{T}P}. The same argument works for θ\theta.
(b) Again, the same argument works. ∎

Combining the previous two lemmas, we have the following.

Corollary 2.5.

Let PP be a probability measure supported on 𝒳\mathcal{X}^{*}, and suppose 𝒯P=P\mathcal{T}P=P. It is then the case that there is a subset of two-dimensional configurations (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}} such that, with probability one on this subset, for any m,sm,s\in{\mathbb{Z}}:

  1. (a)

    for any nmn\leq m and tst\geq s, XntX^{t}_{n} and UntU^{t}_{n} are measurable with respect to (xns)nm(x_{n}^{s})_{n\leq m};

  2. (b)

    for any nm+1n\geq m+1 and tst\leq s, XntX^{t}_{n} and Un1t1U^{t-1}_{n-1} are measurable with respect to (xns)nm+1(x_{n}^{s})_{n\geq m+1}.

Proof.

For type I models, it is possible to deduce from Lemma 2.4(a) that xs=(xns)n𝒳x^{s}=(x^{s}_{n})_{n\in{\mathbb{Z}}}\in\mathcal{X}^{*} for all ss\in{\mathbb{Z}}, 𝐏P\mathbf{P}_{P}-a.s. Since Xnt(x)=Xnts(xs)X^{t}_{n}(x)=X^{t-s}_{n}(x^{s}) and Unt(x)=Unts(xs)U^{t}_{n}(x)=U^{t-s}_{n}(x^{s}) when xs𝒳x^{s}\in\mathcal{X}_{*}, Lemma 2.3 completes the proof. The same argument works for type II model. ∎

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 (u0t)t(u^{t}_{0})_{t\in{\mathbb{Z}}}.

Corollary 2.6.

Let PP be a probability measure PP supported on 𝒳\mathcal{X}^{*}, and suppose (xn)n(x_{n})_{n\in{\mathbb{Z}}} is an independent sequence under PP.

  1. (a)

    For a type I model, if it holds that 𝒯P=P\mathcal{T}P=P, then (u0t)t(u^{t}_{0})_{t\in{\mathbb{Z}}} is an independent and identically distributed (i.i.d.) sequence under 𝐏P\mathbf{P}_{P}.

  2. (b)

    For a type II model, if it holds that 𝒯P=P\mathcal{T}P=P, then (u0t)t(u^{t}_{0})_{t\in{\mathbb{Z}}} is an independent and alternately-distributed sequence under 𝐏P\mathbf{P}_{P}.

Proof.

(a) Since u0t=U0(xt)u^{t}_{0}=U_{0}(x^{t}), it readily follows that the sequence (u0t)t(u^{t}_{0})_{t\in{\mathbb{Z}}} is stationary. As for the independence claim, we note that, by Corollary 2.5, u0tu^{t}_{0} is a measurable function of (xnt)n0(x^{t}_{n})_{n\leq 0}, and (u0s)s<t(u^{s}_{0})_{s<t} is a measurable function of (xnt)n>0(x^{t}_{n})_{n>0}. Since (xnt)n0(x^{t}_{n})_{n\leq 0} and (xnt)n>0(x^{t}_{n})_{n>0} 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 𝒢1,𝒢2,𝒢3\mathcal{G}_{1},\mathcal{G}_{2},\mathcal{G}_{3} be sigma-algebras on a probability space. If 𝒢1\mathcal{G}_{1} and 𝒢2\mathcal{G}_{2} are independent, and σ(𝒢1𝒢2)\sigma(\mathcal{G}_{1}\cup\mathcal{G}_{2}) and 𝒢3\mathcal{G}_{3} are independent, then 𝒢1\mathcal{G}_{1} and σ(𝒢2𝒢3)\sigma(\mathcal{G}_{2}\cup\mathcal{G}_{3}) are independent.

Proof.

Denoting by PP the probability measure on the relevant space, we have that, for any Ei𝒢iE_{i}\in\mathcal{G}_{i}, i=1,2,3i=1,2,3, P(E1E2E3)=P(E1E2)P(E3)=P(E1)P(E2)P(E3)P(E_{1}\cap E_{2}\cap E_{3})=P(E_{1}\cap E_{2})P(E_{3})=P(E_{1})P(E_{2})P(E_{3}). The result follows. ∎

Proof of Theorem 2.1.

(a) Suppose 𝒯μ=μ\mathcal{T}\mu^{{\mathbb{Z}}}=\mu^{{\mathbb{Z}}}. By definition, we have that x00μx_{0}^{0}\sim\mu and u10νu_{-1}^{0}\sim\nu. Moreover, by invariance under 𝒯\mathcal{T}, we have that x01μx^{1}_{0}\sim\mu. And, since θμ=μ\theta\mu^{{\mathbb{Z}}}=\mu^{{\mathbb{Z}}}, Lemma 2.4 yields that θ𝐏μ=𝐏μ\theta\mathbf{P}_{\mu^{{\mathbb{Z}}}}=\mathbf{P}_{\mu^{{\mathbb{Z}}}}, and so the distribution of u00u^{0}_{0} is also ν\nu. Now, by Corollary 2.5, we have that u10u_{-1}^{0} is a measurable function of (xn0)n1(x_{n}^{0})_{n\leq-1}, and u00u_{0}^{0} is a measurable function of (xn1)n1(x_{n}^{1})_{n\geq 1}. In particular, it follows that u10u_{-1}^{0} is independent of x00x^{0}_{0}, and u00u_{0}^{0} is independent of x01x_{0}^{1}, i.e. it holds that (x00,u10)μ×ν(x^{0}_{0},u_{-1}^{0})\sim\mu\times\nu and (x01,u00)μ×ν(x^{1}_{0},u_{0}^{0})\sim\mu\times\nu. Since F(x00,u10)=(x01,u00)F(x_{0}^{0},u_{-1}^{0})=(x_{0}^{1},u_{0}^{0}), we thus obtain that μ×ν\mu\times\nu satisfies the detailed balance condition in this case.

Next, suppose that F(μ×ν)=μ×νF(\mu\times\nu)=\mu\times\nu. By Lemma 2.3, un10u_{n-1}^{0} is measurable with respect to (xm0)mn1(x_{m}^{0})_{m\leq n-1}, so xn0x_{n}^{0} and un10u_{n-1}^{0} are independent for all nn\in{\mathbb{Z}}. By assumption xn0μx_{n}^{0}\sim\mu. Moreover, by assumption and the invariance θ𝐏P=𝐏P\theta\mathbf{P}_{P}=\mathbf{P}_{P} given by Lemma 2.4, un10νu_{n-1}^{0}\sim\nu. Hence the distribution of xn1=F(1)(xn0,un11)x^{1}_{n}=F^{(1)}(x_{n}^{0},u^{1}_{n-1}) is μ\mu, and also xn1x^{1}_{n} and un0u^{0}_{n} are independent, for all nn\in{\mathbb{Z}}. Since, by Lemma 2.3, u00u^{0}_{0} and x01x^{1}_{0} are both measurable with respect to (xn)n0(x_{n})_{n\leq 0}, it follows from Lemma 2.7 that x01x^{1}_{0} and σ(u00,x10,x20,x30,)\sigma(u^{0}_{0},x^{0}_{1},x^{0}_{2},x^{0}_{3},\dots) are independent. Therefore, since (xn1)n1(x^{1}_{n})_{n\geq 1} is measurable with respect to σ(u01,x10,x20,x30,)\sigma(u^{1}_{0},x^{0}_{1},x^{0}_{2},x^{0}_{3},\dots), it must be the case that x01x^{1}_{0} and (xn1)n1(x^{1}_{n})_{n\geq 1} are independent. Finally, since θ𝐏μ=𝐏μ\theta\mathbf{P}_{\mu^{{\mathbb{Z}}}}=\mathbf{P}_{\mu^{{\mathbb{Z}}}} by Lemma 2.4, we obtain that (xn1)n(x^{1}_{n})_{n\in{\mathbb{Z}}} is an i.i.d. sequence with marginal distribution μ\mu, and so 𝒯μ=μ\mathcal{T}\mu^{{\mathbb{Z}}}=\mu^{{\mathbb{Z}}}.
(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 F:𝒳0×𝒰0𝒳~0×𝒰~0F_{*}:\mathcal{X}_{0}\times\mathcal{U}_{0}\to\tilde{\mathcal{X}}_{0}\times\tilde{\mathcal{U}}_{0} be a bijection, and define the involution F:𝒳~0×𝒳0×𝒰0𝒳~0×𝒳0×𝒰0F:\tilde{\mathcal{X}}_{0}\times\mathcal{X}_{0}\times\mathcal{U}_{0}\to\tilde{\mathcal{X}}_{0}\times\mathcal{X}_{0}\times\mathcal{U}_{0} as at (1.4). For a triplet of probability measures (μ,ν,μ~)(\mu,\nu,\tilde{\mu}) on 𝒳0\mathcal{X}_{0}, 𝒰0\mathcal{U}_{0} and 𝒳~0\tilde{\mathcal{X}}_{0}, the following three conditions are then equivalent.

  1. (a)

    F(2,3)(μ~×μ×ν)=μ×νF^{(2,3)}(\tilde{\mu}\times\mu\times\nu)=\mu\times\nu.

  2. (b)

    F(μ~×μ×ν)=μ~×μ×νF(\tilde{\mu}\times\mu\times\nu)=\tilde{\mu}\times\mu\times\nu.

  3. (c)

    There exists a probability measure ν~\tilde{\nu} on 𝒰~0\tilde{\mathcal{U}}_{0} such that the quadruplet of probability measures (μ,ν,μ~,ν~)(\mu,\nu,\tilde{\mu},\tilde{\nu}) satisfies the detailed balance condition with respect to FF_{*}.

Proof.

(b) \Rightarrow (a): This is obvious.
(c) \Rightarrow (b): Let X0μX_{0}\sim\mu, U0νU_{0}\sim\nu, X~0μ~\tilde{X}_{0}\sim\tilde{\mu} be independent random variables, and define (X~0,U~0):=F(X0,U0)(\tilde{X}^{\prime}_{0},\tilde{U}_{0}):=F_{*}(X_{0},U_{0}). By (c), (X~0,U~0)μ~×ν~(\tilde{X}^{\prime}_{0},\tilde{U}_{0})\sim\tilde{\mu}\times\tilde{\nu}. Moreover, by Lemma 2.7, X~0,X~0\tilde{X}_{0},\tilde{X}^{\prime}_{0} and U~0\tilde{U}_{0} are independent. Now, by definition, F(X~0,X0,U0)=(X~0,F1(X~0,U~0))F(\tilde{X}_{0},X_{0},U_{0})=(\tilde{X}^{\prime}_{0},F_{*}^{-1}(\tilde{X}_{0},\tilde{U}_{0})), and, by the detailed balance condition, F1(μ~×ν~)=μ×νF_{*}^{-1}(\tilde{\mu}\times\tilde{\nu})=\mu\times\nu, so (b) holds.
(a) \Rightarrow (c): Let ν~:=F(2)(μ×ν)\tilde{\nu}:=F_{*}^{(2)}(\mu\times\nu), and X0μX_{0}\sim\mu, U0νU_{0}\sim\nu, X~0μ~\tilde{X}_{0}\sim\tilde{\mu} be independent random variables. Since F(2,3)(X~0,X0,U0)=F1(X0~,F(2)(X0,U0))F^{(2,3)}(\tilde{X}_{0},X_{0},U_{0})=F_{*}^{-1}(\tilde{X_{0}},F_{*}^{(2)}(X_{0},U_{0})) and the distribution of (X0~,F(2)(X0,U0))(\tilde{X_{0}},F_{*}^{(2)}(X_{0},U_{0})) is μ~×ν~\tilde{\mu}\times\tilde{\nu}, (a) implies F1(μ~×ν~)=μ×νF_{*}^{-1}(\tilde{\mu}\times\tilde{\nu})=\mu\times\nu. ∎

2.2. Burke’s property

Burke’s theorem is a classical result in queueing theory, which states that, for an M/M/1M/M/1 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 (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}} satisfying Fnt(xnt,un1t)=(xnt+1,unt)F_{n}^{t}(x_{n}^{t},u_{n-1}^{t})=(x_{n}^{t+1},u_{n}^{t}) satisfies Burke’s property if:

  • •:

    the sequences (xn0)n1(x_{n}^{0})_{n\geq 1} and (u0t)t0(u_{0}^{t})_{t\geq 0} are each i.i.d., and independent of each other;

  • •:

    the distribution of (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}} is translation invariant, that is, for any m,sm,s\in{\mathbb{Z}},

    Tsθm((xnt,unt)n,t)=d(xnt,unt)n,t.T^{s}\theta^{m}\left(\left(x^{t}_{n},u^{t}_{n}\right)_{n,t\in{\mathbb{Z}}}\right)\buildrel d\over{=}\left(x^{t}_{n},u^{t}_{n}\right)_{n,t\in{\mathbb{Z}}}.
Burke’s property for a type II model:

We say that a distribution supported on configurations (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}} satisfying Fnt(xnt,un1t)=(xnt+1,unt)F_{n}^{t}(x_{n}^{t},u_{n-1}^{t})=(x_{n}^{t+1},u_{n}^{t}) satisfies Burke’s property if:

  • •:

    the sequences (x2n0)n1(x_{2n}^{0})_{n\geq 1}, (x2n10)n1(x_{2n-1}^{0})_{n\geq 1}, (u02t)t0(u_{0}^{2t})_{t\geq 0} and (u02t1)t1(u_{0}^{2t-1})_{t\geq 1} are each i.i.d., and independent of each other;

  • •:

    the distribution of (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}} is translation invariant, that is, for any m,sm,s\in{\mathbb{Z}} such that m+s=0m+s=0 (mod 2),

    Tsθm((xnt,unt)n,t)=d(xnt,unt)n,t.T^{s}\theta^{m}\left(\left(x^{t}_{n},u^{t}_{n}\right)_{n,t\in{\mathbb{Z}}}\right)\buildrel d\over{=}\left(x^{t}_{n},u^{t}_{n}\right)_{n,t\in{\mathbb{Z}}}.

We make the obvious remark that, in the case of a type I model, if the distribution of (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}} satisfies Burke’s property, then (xnt)n(x_{n}^{t})_{n\in{\mathbb{Z}}} is i.i.d. for each tt\in{\mathbb{Z}}, and (unt)t(u^{t}_{n})_{t\in{\mathbb{Z}}} is i.i.d. for each nn\in\mathbb{N}. 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 xntx_{n}^{t} and untu_{n}^{t} in terms of the detailed balance solution.

Proposition 2.9 (Burke’s property).
  1. (a)

    Type I: If a pair of probability measures (μ,ν)(\mu,\nu) satisfies the detailed balance condition, then there exists a distribution supported on configurations (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}} satisfying Fnt(xnt,un1t)=(xnt+1,unt)F_{n}^{t}(x_{n}^{t},u_{n-1}^{t})=(x_{n}^{t+1},u_{n}^{t}) for which Burke’s property holds. Moreover, if it holds that μ(𝒳)=1\mu^{{\mathbb{Z}}}(\mathcal{X}^{*})=1, then u10νu^{0}_{-1}\sim\nu and 𝐏μ\mathbf{P}_{\mu^{\mathbb{Z}}} satisfies Burke’s property.

  2. (b)

    Type II: If a quadruplet of probability measures (μ,ν,μ~,ν~)(\mu,\nu,\tilde{\mu},\tilde{\nu}) satisfies the detailed balance condition, then there exists a distribution supported on configurations (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}} satisfying Fnt(xnt,un1t)=(xnt+1,unt)F_{n}^{t}(x_{n}^{t},u_{n-1}^{t})=(x_{n}^{t+1},u_{n}^{t}) for which Burke’s property holds. Moreover, if it holds that (μ×μ~)(𝒳)=1(\mu\times\tilde{\mu})^{{\mathbb{Z}}}(\mathcal{X}^{*})=1, then u10νu^{0}_{-1}\sim\nu, u00ν~u^{0}_{0}\sim\tilde{\nu}, and 𝐏(μ×μ~)\mathbf{P}_{(\mu\times\tilde{\mu})^{{\mathbb{Z}}}} satisfies Burke’s property.

Proof.

(a) Let (xn0,u0t)n1,t0(x_{n}^{0},u_{0}^{t})_{n\geq 1,t\geq 0} be independent random variables satisfying xn0μx_{n}^{0}\sim\mu and u0tνu_{0}^{t}\sim\nu. For n,tn,t\in{\mathbb{N}}, define

(xnt,unt1):=F(xnt1,un1t1)\left(x_{n}^{t},u_{n}^{t-1}\right):=F\left(x_{n}^{t-1},u_{n-1}^{t-1}\right)

recursively. By induction and the detailed balance condition, one readily obtains that xn1μx^{1}_{n}\sim\mu, un0νu^{0}_{n}\sim\nu and xn1x^{1}_{n} and un0u^{0}_{n} are independent for all nn\in{\mathbb{N}}. Moreover, for any nn\in{\mathbb{N}}, xn1x^{1}_{n} and un0u^{0}_{n} are measurable with respect to σ(u00,x10,x20,,xn0)\sigma(u^{0}_{0},x_{1}^{0},x_{2}^{0},\dots,x_{n}^{0}), and (xm1)mn+1(x_{m}^{1})_{m\geq n+1} is measurable with respect to σ(un0,xn+10,xn+20,)\sigma(u^{0}_{n},x_{n+1}^{0},x_{n+2}^{0},\dots). So, applying Lemma 2.7, we find that xn1x^{1}_{n} and (xm1)mn+1(x_{m}^{1})_{m\geq n+1} are independent. Hence (xn1)n(x^{1}_{n})_{n\in{\mathbb{N}}} is an i.i.d. sequence with the marginal μ\mu. Now, since (xn1)n(x^{1}_{n})_{n\in{\mathbb{N}}} is measurable with respect to σ(u00,(xn0)n)\sigma(u^{0}_{0},(x_{n}^{0})_{n\in{\mathbb{N}}}), it further holds that (xn1)n(x^{1}_{n})_{n\in{\mathbb{N}}} and (u0t)t1(u^{t}_{0})_{t\geq 1} are independent. Letting ynt:=xnt+1y^{t}_{n}:=x^{t+1}_{n} and vnt:=unt+1v^{t}_{n}:=u^{t+1}_{n}, we thus have that (yn0,v0t)n1,t0(y_{n}^{0},v_{0}^{t})_{n\geq 1,t\geq 0} are independent random variables satisfying yn0μy_{n}^{0}\sim\mu, v0tνv_{0}^{t}\sim\nu and

(ynt,vnt1)=F(ynt1,vn1t1)\left(y_{n}^{t},v_{n}^{t-1}\right)=F\left(y_{n}^{t-1},v_{n-1}^{t-1}\right)

for all n,tn,t\in{\mathbb{N}}. In particular, (xnt,unt)n1,t0=d(ynt,vnt)n1,t0(x_{n}^{t},u_{n}^{t})_{n\geq 1,t\geq 0}\buildrel d\over{=}(y_{n}^{t},v_{n}^{t})_{n\geq 1,t\geq 0}, which implies

(xnt+1,unt+1)n1,t0=d(xnt,unt)n1,t0.\left(x_{n}^{t+1},u_{n}^{t+1}\right)_{n\geq 1,t\geq 0}\buildrel d\over{=}\left(x_{n}^{t},u_{n}^{t}\right)_{n\geq 1,t\geq 0}.

By the same argument, one can show that

(xn+1t,un+1t)n1,t0=d(xnt,unt)n1,t0,\left(x_{n+1}^{t},u_{n+1}^{t}\right)_{n\geq 1,t\geq 0}\buildrel d\over{=}\left(x_{n}^{t},u_{n}^{t}\right)_{n\geq 1,t\geq 0},

and so

(xn+mt+s,un+mt+s)n1,t0=d(xnt,unt)n1,t0,\left(x_{n+m}^{t+s},u_{n+m}^{t+s}\right)_{n\geq 1,t\geq 0}\buildrel d\over{=}\left(x_{n}^{t},u_{n}^{t}\right)_{n\geq 1,t\geq 0},

for any m,sm,s\in{\mathbb{N}}. Finally, by constructing the distributions of (xnt,unt)nk+1,tk(x^{t}_{n},u^{t}_{n})_{n\geq k+1,t\geq k} for each kk\in{\mathbb{Z}} by translation, we can construct the distribution of (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}} 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 μ(𝒳)=1\mu^{{\mathbb{Z}}}(\mathcal{X}^{*})=1, then there is a unique distribution of (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}} that is supported on configurations satisfying Fnt(xnt,un1t)=(xnt+1,unt)F_{n}^{t}(x_{n}^{t},u_{n-1}^{t})=(x_{n}^{t+1},u_{n}^{t}) and with marginal (xn0)nμ(x_{n}^{0})_{n\in{\mathbb{Z}}}\sim\mu^{{\mathbb{Z}}}. Hence it must be the one satisfying Burke’s property, as constructed above. In particular, untνu^{t}_{n}\sim\nu for all n,tn,t\in{\mathbb{Z}}.
(b) The same argument as for part (a) works. ∎

Proof of Theorem 1.1.

Combine Theorem 2.1 and Proposition 2.9. ∎

We conclude the subsection with a corollary that establishes, when the marginal of (xn0)n(x_{n}^{0})_{n\in{\mathbb{Z}}} is supported on 𝒳\mathcal{X}^{*}, 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.
  1. (a)

    Type I: Suppose that μ\mu is a probability measure on 𝒳0\mathcal{X}_{0} such that μ(𝒳)=1\mu^{{\mathbb{Z}}}(\mathcal{X}^{*})=1. Let ν\nu be the distribution of U1(x)U_{-1}(x), where xμx\sim\mu^{{\mathbb{Z}}}. It is then the case that there exists a distribution of (xnt,unt)n,t(x_{n}^{t},u_{n}^{t})_{n,t\in{\mathbb{Z}}} satisfying (xn0)nμ(x_{n}^{0})_{n\in{\mathbb{Z}}}\sim\mu^{{\mathbb{Z}}} and Burke’s property if and only if (μ,ν)(\mu,\nu) satisfies the detailed balance condition.

  2. (b)

    Type II: Suppose that μ×μ~\mu\times\tilde{\mu} is a probability measure on 𝒳0×𝒳~0\mathcal{X}_{0}\times\tilde{\mathcal{X}}_{0} such that (μ×μ~)(𝒳)=1(\mu\times\tilde{\mu})^{{\mathbb{Z}}}(\mathcal{X}^{*})=1. Let ν\nu, ν~\tilde{\nu} be the distributions of U1(x)U_{-1}(x), U0(x)U_{0}(x), respectively, where x(μ×μ~)x\sim(\mu\times\tilde{\mu})^{{\mathbb{Z}}}. It is then the case that if there exists a distribution of (xnt,unt)n,t(x_{n}^{t},u_{n}^{t})_{n,t\in{\mathbb{Z}}} satisfying (xn0)n(μ×μ~)(x_{n}^{0})_{n\in{\mathbb{Z}}}\sim(\mu\times\tilde{\mu})^{{\mathbb{Z}}} and Burke’s property if and only if (μ,ν,μ~,ν~)(\mu,\nu,\tilde{\mu},\tilde{\nu}) 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 (xnt,unt)n,t(x_{n}^{t},u_{n}^{t})_{n,t\in{\mathbb{Z}}} satisfying (xn0)nμ(x_{n}^{0})_{n\in{\mathbb{Z}}}\sim\mu^{{\mathbb{Z}}} and Burke’s property. Since μ(𝒳)=1\mu^{{\mathbb{Z}}}(\mathcal{X}^{*})=1, the measure must be 𝐏μ\mathbf{P}_{\mu^{{\mathbb{Z}}}}. By the second condition of Burke’s property, T𝐏μ=𝐏μT\mathbf{P}_{\mu^{{\mathbb{Z}}}}=\mathbf{P}_{\mu^{{\mathbb{Z}}}} holds. Hence, by Lemma 2.4, we must have that 𝒯μ=μ\mathcal{T}\mu^{{\mathbb{Z}}}=\mu^{{\mathbb{Z}}} 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 𝒯\mathcal{T} for i.i.d. invariant measures. To state the result, we introduce an involution Fˇ:𝒰0×𝒳0𝒰0×𝒳0\check{F}:\mathcal{U}_{0}\times\mathcal{X}_{0}\to\mathcal{U}_{0}\times\mathcal{X}_{0} by setting

Fˇ=πFπ,\check{F}=\pi\circ F\circ\pi,

where π(u,x):=(x,u)\pi(u,x):=(x,u). We consider Fˇ\check{F} the dual of FF.

Theorem 2.11.

Suppose we have a type I model, and that μ\mu is a probability measure on 𝒳0\mathcal{X}_{0} such that μ(𝒳)=1\mu^{{\mathbb{Z}}}(\mathcal{X}^{*})=1 and 𝒯μ=μ\mathcal{T}\mu^{{\mathbb{Z}}}=\mu^{{\mathbb{Z}}}. If it holds that, for 𝐏μ\mathbf{P}_{\mu^{{\mathbb{Z}}}}-a.e. u0=(u0t)tu_{0}=(u_{0}^{t})_{t\in\mathbb{Z}}, there exists at most one x=(xt)t𝒳0x=(x^{t})_{t}\in\mathcal{X}_{0}^{\mathbb{Z}} such that

Fˇ(2)(u0t,xt)=xt+1,t,\check{F}^{(2)}(u_{0}^{t},x^{t})=x^{t+1},\qquad\forall t\in{\mathbb{Z}},

then μ\mu^{{\mathbb{Z}}} is ergodic under 𝒯\mathcal{T}.

Remark 2.12.

We note that, by Theorem 2.1 and Proposition 2.9, under 𝐏μ\mathbf{P}_{\mu^{{\mathbb{Z}}}}, u0=(u0t)tu_{0}=(u_{0}^{t})_{t\in\mathbb{Z}} has law ν\nu^{{\mathbb{Z}}}, where ν\nu is the distribution of u10u_{-1}^{0} under 𝐏μ\mathbf{P}_{\mu^{{\mathbb{Z}}}}. In particular, one could replace ‘𝐏μ\mathbf{P}_{\mu^{{\mathbb{Z}}}}-a.e.’ with ‘ν\nu^{{\mathbb{Z}}}-a.e.’ in the above statement.

Remark 2.13.

Under the assumptions of Theorem 2.11, in addition to ergodicity, the same proof gives the measure-preserving transformation 𝒯\mathcal{T} is metrically isomorphic to a two-sided Bernoulli shift, cf. [25].

The proof of the above theorem will depend on the following lemma. For the statement of this, we define a function Λ:𝒳𝒰0\Lambda:\mathcal{X}^{*}\to\mathcal{U}_{0}^{\mathbb{Z}} by setting

Λ(x):=(u0t)t,\Lambda(x):=(u^{t}_{0})_{t\in{\mathbb{Z}}},

where (xnt,unt)n,t(x_{n}^{t},u_{n}^{t})_{n,t\in{\mathbb{Z}}} is the unique solution of (1.1) with initial condition xx. Note that, as is consistent with the idea that TT is a temporal shift, we set T((u0t)t):=(u0t+1)tT((u^{t}_{0})_{t\in{\mathbb{Z}}}):=(u^{t+1}_{0})_{t\in{\mathbb{Z}}}.

Lemma 2.14.

Let PP be a distribution on 𝒳\mathcal{X}^{*}. Suppose there exists a set 𝒰𝒰0\mathcal{U}^{*}\subseteq\mathcal{U}_{0}^{\mathbb{Z}} and a function

Λ~:𝒰𝒳0\tilde{\Lambda}:\mathcal{U}^{*}\to\mathcal{X}_{0}^{\mathbb{Z}}

such that ΛP(𝒰)=1\Lambda P(\mathcal{U}^{*})=1 and Λ~Λ\tilde{\Lambda}\circ\Lambda is the identity map on the set {x𝒳:Λ(x)𝒰}\{x\in\mathcal{X}^{*}:\>\Lambda(x)\in\mathcal{U}^{*}\}. The following statements then hold.

  1. (a)

    PP is invariant under 𝒯\mathcal{T} if and only if ΛP\Lambda P is invariant under TT.

  2. (b)

    PP is invariant and ergodic under 𝒯\mathcal{T} if and only if ΛP\Lambda P is invariant and ergodic under TT.

Proof.

(a) Define 𝒳:={x𝒳:Λ(x)𝒰}\mathcal{X}^{**}:=\{x\in\mathcal{X}^{*}:\>\Lambda(x)\in\mathcal{U}^{*}\} and 𝒰:=Λ(𝒳)𝒰\mathcal{U}^{**}:=\Lambda(\mathcal{X}^{*})\cap\mathcal{U}^{*}. We first check that Λ:𝒳𝒰\Lambda:\mathcal{X}^{**}\to\mathcal{U}^{**} is a bijection with inverse function Λ~\tilde{\Lambda}. Clearly Λ(𝒳)𝒰\Lambda(\mathcal{X}^{**})\subseteq\mathcal{U}^{**}. Moreover, by assumption, Λ~Λ(x)=x\tilde{\Lambda}\circ\Lambda(x)=x for all x𝒳x\in\mathcal{X}^{**}. Hence it remains to show that

ΛΛ~(u)=u,u𝒰.\Lambda\circ\tilde{\Lambda}(u)=u,\qquad\forall u\in\mathcal{U}^{**}.

For any u𝒰Λ(𝒳)u\in\mathcal{U}^{**}\subseteq\Lambda(\mathcal{X}^{*}), there exists xu𝒳x_{u}\in\mathcal{X}^{**} such that Λ(xu)=u\Lambda(x_{u})=u. It follows that

ΛΛ~(u)=ΛΛ~Λ(xu)=Λ(xu)=u,\Lambda\circ\tilde{\Lambda}(u)=\Lambda\circ\tilde{\Lambda}\circ\Lambda(x_{u})=\Lambda(x_{u})=u,

as required. Next, since P(𝒳)=ΛP(𝒰)=1P(\mathcal{X}^{*})=\Lambda P(\mathcal{U}^{*})=1, we have that P(𝒳)=1P(\mathcal{X}^{**})=1, and thus also ΛP(𝒰)=1\Lambda P(\mathcal{U}^{**})=1. Consequently, if 𝒯P=P\mathcal{T}P=P, then it PP-a.s. holds that x:=(xn)nx:=(x_{n})_{n\in{\mathbb{Z}}} and 𝒯(x)\mathcal{T}(x) take values in 𝒳\mathcal{X}^{**}, and so

Λ(𝒯(x))=(u0t+1)t=T((u0t)t)=TΛ(x).\Lambda(\mathcal{T}(x))=(u^{t+1}_{0})_{t\in{\mathbb{Z}}}=T\left((u^{t}_{0})_{t\in{\mathbb{Z}}}\right)=T\Lambda(x).

It follows that TΛP=Λ𝒯P=ΛPT\Lambda P=\Lambda\mathcal{T}P=\Lambda P. On the other hand, if TΛP=Λ𝒯P=ΛPT\Lambda P=\Lambda\mathcal{T}P=\Lambda P, then it ΛP\Lambda P-a.s. holds that u:=(u0t)tu:=(u^{t}_{0})_{t\in{\mathbb{Z}}} and T(u)T(u) takes values in 𝒰\mathcal{U}^{**}, and so

Λ~(T(u))=Λ~((u0t+1)t)=𝒯(x)=𝒯Λ~(u).\tilde{\Lambda}(T(u))=\tilde{\Lambda}\left((u^{t+1}_{0})_{t\in{\mathbb{Z}}}\right)=\mathcal{T}(x)=\mathcal{T}\tilde{\Lambda}(u).

Hence 𝒯P=𝒯Λ~ΛP=Λ~TΛP=Λ~ΛP=P\mathcal{T}P=\mathcal{T}\tilde{\Lambda}\Lambda P=\tilde{\Lambda}T\Lambda P=\tilde{\Lambda}\Lambda P=P.
(b) By the proof of (a), for any subset E𝒳E\subseteq\mathcal{X}^{**}, Λ(𝒯(E))=T(Λ(E))\Lambda(\mathcal{T}(E))=T(\Lambda(E)), and so 𝒯E=E\mathcal{T}E=E is equivalent to TΛE=ΛET\Lambda E=\Lambda E. 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 Λ(μ)=ν\Lambda(\mu^{{\mathbb{Z}}})=\nu^{{\mathbb{Z}}}. Moreover, ν\nu^{{\mathbb{Z}}} is clearly invariant and ergodic under TT. Hence, by Lemma 2.14, we only need to show the existence of a set 𝒰𝒰0\mathcal{U}^{*}\subseteq\mathcal{U}_{0}^{\mathbb{Z}} and a function Λ~:𝒰𝒳0\tilde{\Lambda}:\mathcal{U}^{*}\to\mathcal{X}_{0}^{\mathbb{Z}} such that ν(𝒰)=1\nu^{{\mathbb{Z}}}(\mathcal{U}^{*})=1 and Λ~Λ\tilde{\Lambda}\circ\Lambda is the identity map on the set {x𝒳:Λ(x)𝒰}\{x\in\mathcal{X}^{*}:\>\Lambda(x)\in\mathcal{U}^{*}\}. To this end, let 𝒰,0𝒰0\mathcal{U}^{*,0}\subseteq\mathcal{U}_{0}^{\mathbb{Z}} be the set of u=(ut)tu=(u^{t})_{t\in{\mathbb{Z}}} such that there is at most one x=(xt)t𝒳0x=(x^{t})_{t\in{\mathbb{Z}}}\in\mathcal{X}_{0}^{\mathbb{Z}} satisfying

Fˇ(2)(ut,xt)=xt+1,t.\check{F}^{(2)}(u^{t},x^{t})=x^{t+1},\qquad\forall t\in{\mathbb{Z}}.

By assumption, ν(𝒰,0)=1\nu^{{\mathbb{Z}}}(\mathcal{U}^{*,0})=1. Since ν=ν\mathcal{R}\nu^{{\mathbb{Z}}}=\nu^{{\mathbb{Z}}}, where ut:=ut\mathcal{R}u^{t}:=u^{-t}, and un:=(unt)tνu_{n}:=(u_{n}^{t})_{t\in{\mathbb{Z}}}\sim\nu^{{\mathbb{Z}}} under 𝐏μ\mathbf{P}_{\mu^{{\mathbb{Z}}}} for all nn, it follows that

(2.1) 𝐏μ(un𝒰,0𝒰,0,n)=1.\mathbf{P}_{\mu^{{\mathbb{Z}}}}\left(u_{n}\in\mathcal{U}^{*,0}\cap\mathcal{R}\mathcal{U}^{*,0},\forall n\in{\mathbb{Z}}\right)=1.

Now, define 𝒳\mathcal{X}^{**} to be the set of x𝒳x\in\mathcal{X}^{*} such that un(x)𝒰,0𝒰,0u_{n}(x)\in\mathcal{U}^{*,0}\cap\mathcal{R}\mathcal{U}^{*,0} for all nn, where un=(unt)tu_{n}=(u_{n}^{t})_{t\in{\mathbb{Z}}} is given by the solution of the initial value problem (1.1) with initial condition xx. Moreover, set 𝒰:=Λ(𝒳)\mathcal{U}^{*}:=\Lambda(\mathcal{X}^{**}), and note that, by (2.1), we have that ν(𝒰)=1\nu^{{\mathbb{Z}}}(\mathcal{U}^{*})=1. We next claim that for any u𝒰u\in\mathcal{U}^{*}, there is a unique x𝒳x\in\mathcal{X}^{*} such that Λ(x)=u\Lambda(x)=u, and moreover that x𝒳x\in\mathcal{X}^{**}. Indeed, if x𝒳x\in\mathcal{X}^{**}, x𝒳x^{\prime}\in\mathcal{X}^{*} and Λ(x)=u=Λ(x)\Lambda(x)=u=\Lambda(x^{\prime}), then

F(x1t,u0t)=(x1t+1,u1t),F(x1t,u0t)=(x1t+1,u1t),F\left(x_{1}^{t},u^{t}_{0}\right)=\left(x_{1}^{t+1},{u}_{1}^{t}\right),\qquad F\left({x^{\prime}}_{1}^{t},{u^{\prime}}^{t}_{0}\right)=\left({x^{\prime}}_{1}^{t+1},{u^{\prime}}_{1}^{t}\right),

where (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}} and (xnt,unt)n,t({x^{\prime}}^{t}_{n},{u^{\prime}}^{t}_{n})_{n,t\in{\mathbb{Z}}} are the solutions of the initial value problem (1.1) with initial conditions xx and xx^{\prime}, respectively. Hence,

Fˇ(2)(u0t,x1t)=x1t+1,Fˇ(2)(u0t,x1t)=x1t+1.\check{F}^{(2)}\left(u^{t}_{0},x_{1}^{t}\right)=x_{1}^{t+1},\qquad\check{F}^{(2)}\left({u^{\prime}}^{t}_{0},{x^{\prime}}_{1}^{t}\right)={x^{\prime}}_{1}^{t+1}.

Since (u0t)t=Λ(x)=Λ(x)=(u0t)t(u^{t}_{0})_{t\in{\mathbb{Z}}}=\Lambda(x)=\Lambda(x^{\prime})=({u^{\prime}}^{t}_{0})_{t\in{\mathbb{Z}}} is an element of 𝒰,0\mathcal{U}^{*,0}, it must therefore be the case that x1t=x1tx_{1}^{t}={x^{\prime}}_{1}^{t} for all tt\in{\mathbb{Z}}. It moreover follows that u1t=u1tu_{1}^{t}={u^{\prime}}_{1}^{t} for all tt\in{\mathbb{Z}}. Since x𝒳x\in\mathcal{X}^{**} implies unt𝒰,0u_{n}^{t}\in\mathcal{U}^{*,0} for all nn, iterating this argument yields that xnt=xntx_{n}^{t}={x^{\prime}}_{n}^{t} for all tt\in{\mathbb{Z}} and n0n\geq 0. To deal with negative nn, note that

F(x0t,u1t)=(x0t+1,u0t),F(x0t,u1t)=(x0t+1,u0t),F\left(x_{0}^{t},u^{t}_{-1}\right)=\left(x_{0}^{t+1},{u}_{0}^{t}\right),\qquad F\left({x^{\prime}}_{0}^{t},{u^{\prime}}^{t}_{-1}\right)=\left({x^{\prime}}_{0}^{t+1},{u^{\prime}}_{0}^{t}\right),

is equivalent to

(x0t,u1t)=F(x0t+1,u0t),(x0t,u1t)=F(x0t+1,u0t),\left(x_{0}^{t},u^{t}_{-1}\right)=F\left(x_{0}^{t+1},{u}_{0}^{t}\right),\qquad\left({x^{\prime}}_{0}^{t},{u^{\prime}}^{t}_{-1}\right)=F\left({x^{\prime}}_{0}^{t+1},{u^{\prime}}_{0}^{t}\right),

and so

Fˇ(2)(u0t,x0t+1)=x0t,Fˇ(2)(u0t,x0t+1)=x0t.\check{F}^{(2)}\left(u^{t}_{0},x_{0}^{t+1}\right)=x_{0}^{t},\qquad\check{F}^{(2)}\left({u^{\prime}}^{t}_{0},{x^{\prime}}_{0}^{t+1}\right)={x^{\prime}}_{0}^{t}.

Applying the reflection \mathcal{R} thus yields

Fˇ(2)(u0t,x0t+1)=x0t,Fˇ(2)(u0t,x0t+1)=x0t.\check{F}^{(2)}\left(u^{-t}_{0},x_{0}^{-t+1}\right)=x_{0}^{-t},\qquad\check{F}^{(2)}\left({u^{\prime}}^{-t}_{0},{x^{\prime}}_{0}^{-t+1}\right)={x^{\prime}}_{0}^{-t}.

Since u0𝒰,0\mathcal{R}u_{0}\in\mathcal{U}^{*,0}, this implies x0t=x0tx_{0}^{t}={x^{\prime}}_{0}^{t} for all tt\in{\mathbb{Z}}. Again, we can iterate this argument to conclude that xnt=xntx^{t}_{n}={x^{\prime}}^{t}_{n} for all t,nt,n\in{\mathbb{Z}}, as desired. Hence the function Λ~:𝒰𝒳\tilde{\Lambda}:\mathcal{U}^{*}\to\mathcal{X}^{**} given by Λ(x)x\Lambda(x)\mapsto x is well-defined, and Λ~Λ(x)=x\tilde{\Lambda}\circ\Lambda(x)=x for all x𝒳x\in\mathcal{X}^{**}. Moreover, we have from the above argument that 𝒳={x𝒳:Λ(x)𝒰}\mathcal{X}^{**}=\{x\in\mathcal{X}^{*}:\>\Lambda(x)\in\mathcal{U}^{*}\}, 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, J,K{}J,K\in{\mathbb{R}}\cup\{\infty\}, and is based on the following lattice map:

(udKdV) FudK(J,K)(x,u)\displaystyle F^{(J,K)}_{udK}(x,u)
:=(umax{x+uJ,0}+max{x+uK,0},xmax{x+uK,0}+max{x+uJ,0}),\displaystyle:=\left(u-\max\{x+u-J,0\}+\max\{x+u-K,0\},x-\max\{x+u-K,0\}+\max\{x+u-J,0\}\right),

where the variables xx and uu are {\mathbb{R}} valued. When the variables are positive, one can think of xx as the amount of mass currently at a lattice site, which has capacity JJ. Moreover, uu represents the amount of mass that a ‘carrier’, which has capacity KK, is bringing to this site. Simultaneously, the carrier deposits what it can, i.e. min{u,Jx}\min\{u,J-x\}, and collects what it can, i.e. min{x,Ku}\min\{x,K-u\}. This leaves a mass of

x+min{u,Jx}min{x,Ku}=(FudK(J,K))(1)(x,u)x+\min\{u,J-x\}-\min\{x,K-u\}=\left(F^{(J,K)}_{udK}\right)^{(1)}(x,u)

at the site, and the carrier moves forward (rightwards) to the next lattice site carrying a mass of

umin{u,Jx}+min{x,Ku}=(FudK(J,K))(2)(x,u);u-\min\{u,J-x\}+\min\{x,K-u\}=\left(F^{(J,K)}_{udK}\right)^{(2)}(x,u);

one discrete time step of the lattice dynamics is given by a complete pass of the carrier from n=n=-\infty to n=+n=+\infty. We note that the original udKdV equation corresponds to setting K=K=\infty. We also highlight that if J,KJ,K\in{\mathbb{N}} and we restrict the possible values of the variables so that x{0,1,,J}x\in\{0,1,\dots,J\} and u{0,1,,K}u\in\{0,1,\dots,K\}, then the dynamics associated with FudK(J,K)F^{(J,K)}_{udK} correspond to the box-ball system with box capacity JJ and carrier capacity KK, which we denote by BBS(JJ,KK).

Remark 3.1.

Similarly to the discussion for BBS(JJ,KK) in [10], the map (udKdV) admits various symmetries, including the following.

Involution:

For any (x,u)2(x,u)\in\mathbb{R}^{2}, it holds that

(3.1) FudK(J,K)FudK(J,K)(x,u)=(x,u).F_{udK}^{(J,K)}\circ F_{udK}^{(J,K)}(x,u)=(x,u).
Configuration-carrier duality:

If π(x,u):=(u,x)\pi(x,u):=(u,x), then

(3.2) FudK(J,K)=πFudK(K,J)π.F^{(J,K)}_{udK}=\pi\circ F_{udK}^{(K,J)}\circ\pi.
Empty space-particle duality:

Suppose J,K<J,K<\infty. If σJ,K(x,u):=(Jx,Ku)\sigma_{J,K}(x,u):=(J-x,K-u), then

(3.3) FudK(J,K)=σJ,KFudK(J,K)σJ,K.F^{(J,K)}_{udK}=\sigma_{J,K}\circ F^{(J,K)}_{udK}\circ\sigma_{J,K}.
Shift invariance:

If rr\in\mathbb{R}, then for any (x,u)2(x,u)\in\mathbb{R}^{2} it holds that

(3.4) FudK(J2r,K2r)(xr,ur)=((FudK(J,K))(1)(x,u)r,(FudK(J,K))(2)(x,u)r).F_{udK}^{(J-2r,K-2r)}(x-r,u-r)=\left(\left(F_{udK}^{(J,K)}\right)^{(1)}(x,u)-r,\left(F_{udK}^{(J,K)}\right)^{(2)}(x,u)-r\right).
Scale invariance:

If λ\lambda\in\mathbb{R}, then for any (x,u)2(x,u)\in\mathbb{R}^{2} it holds that

(3.5) FudK(λJ,λK)(λx,λu)=λFudK(J,K)(x,u).F_{udK}^{(\lambda J,\lambda K)}(\lambda x,\lambda u)=\lambda F_{udK}^{(J,K)}(x,u).

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 JKJ\leq K. 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 JKJ\leq K. 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(JJ,KK) 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 μ×ν\mu\times\nu satisfy FudK(J,K)(μ×ν)=μ×νF^{(J,K)}_{udK}(\mu\times\nu)=\mu\times\nu.

  1. (a)

    Suppose J,K{}J,K\in{\mathbb{R}}\cup\{\infty\}.

    1. (i)

      For λ\lambda\in\mathbb{R} if max{J,K}<\max\{J,K\}<\infty, or λ>0\lambda>0 if max{J,K}=\max\{J,K\}=\infty, and finite c<min{J2,K2}c<\min\{\frac{J}{2},\frac{K}{2}\},

      μ×ν=stExp(λ,c,Jc)×stExp(λ,c,Kc).\mu\times\nu=\mathrm{stExp}(\lambda,c,J-c)\times\mathrm{stExp}(\lambda,c,K-c).
    2. (ii)

      For finite c<min{J2,K2}c<\min\{\frac{J}{2},\frac{K}{2}\} and m>0m>0 such that c,J,Km{}c,J,K\in m\mathbb{Z}\cup\{\infty\},

      μ×ν=sstbGeo(1θ,cm,Jcm,κ,m)×sstbGeo(1θ,cm,Kcm,κ,m),\mu\times\nu=\mathrm{sstbGeo}\left(1-\theta,\frac{c}{m},\frac{J-c}{m},\kappa,m\right)\times\mathrm{sstbGeo}\left(1-\theta,\frac{c}{m},\frac{K-c}{m},\kappa,m\right),

      where it is further supposed that: either J2c,K2cm{}J-2c,K-2c\in m\mathbb{Z}\cup\{\infty\}, θ(0,1)\theta\in(0,1), κ=1\kappa=1; or J2c,K2c2m{}J-2c,K-2c\in 2m\mathbb{Z}\cup\{\infty\}, θ(0,1)\theta\in(0,1), κ(0,)\{1}\kappa\in(0,\infty)\backslash\{1\}; or J2c,K2cmJ-2c,K-2c\in m\mathbb{Z}, θ1\theta\geq 1, κ=1\kappa=1; or J2c,K2c2m{}J-2c,K-2c\in 2m\mathbb{Z}\cup\{\infty\}, θ1\theta\geq 1, κ(0,)\{1}\kappa\in(0,\infty)\backslash\{1\}.

  2. (b)

    Suppose J=KJ=K. For any measure mm on \mathbb{R},

    μ×ν=m×m.\mu\times\nu=m\times m.
  3. (c)

    Suppose J<KJ<K.

    1. (i)

      For any measure mm supported on (,J2](-\infty,\frac{J}{2}],

      μ×ν=m×m.\mu\times\nu=m\times m.
    2. (ii)

      For any measure mm supported on [J2,KJ2][\frac{J}{2},K-\frac{J}{2}],

      μ×ν=δJ2×m,\mu\times\nu=\delta_{\frac{J}{2}}\times m,

      where for xx\in\mathbb{R}, δx\delta_{x} is the probability measure placing all of its mass at xx.

    3. (iii)

      Suppose further that K<K<\infty. For any measure mm supported on [J2,)[\frac{J}{2},\infty),

      μ×ν=m×(m+L),\mu\times\nu=m\times(m+L),

      where L:=KJL:=K-J and (m+L)(A)=m({xL:xA})(m+L)(A)=m(\{x-L:\>x\in A\}).

Proof.

Since FudK(J,K)F^{(J,K)}_{udK} preserves mass, i.e.

(FudK(J,K))(1)(x,u)+(FudK(J,K))(2)(x,u)=x+u,\left(F_{udK}^{(J,K)}\right)^{(1)}(x,u)+\left(F_{udK}^{(J,K)}\right)^{(2)}(x,u)=x+u,

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 FudK(J,K)F^{(J,K)}_{udK}, and so their proofs are omitted. ∎

Proposition 3.3.
  1. (a)

    Suppose J=KJ=K. It is then the case that the product measures given in Proposition 3.2(b) are the only solutions to FudK(J,K)(μ×ν)=μ×νF^{(J,K)}_{udK}(\mu\times\nu)=\mu\times\nu.

  2. (b)

    Suppose J<KJ<K and a product measure μ×ν\mu\times\nu satisfies FudK(J,K)(μ×ν)=μ×νF^{(J,K)}_{udK}(\mu\times\nu)=\mu\times\nu. It is then the case that one of the following statements hold.

    1. (i)

      The product measure μ×ν\mu\times\nu is given in Proposition 3.2(c).

    2. (ii)

      There exists c[,J2)c\in[-\infty,\frac{J}{2}) such that

      infsupp(μ)=infsupp(ν)=c,\inf\mathrm{supp}(\mu)=\inf\mathrm{supp}(\nu)=c,
      supsupp(μ)=Jc,supsupp(ν)=Kc,\sup\mathrm{supp}(\mu)=J-c,\qquad\sup\mathrm{supp}(\nu)=K-c,

      where supp(μ)\mathrm{supp}(\mu) and supp(ν)\mathrm{supp}(\nu) are the support of μ\mu and ν\nu, respectively.

    Moreover, if (ii) holds and μ\mu and ν\nu have smooth (twice differentiable), strictly positive densities on the intervals [c,Jc][c,J-c] and [c,Kc][c,K-c] respectively, then they given by Proposition 3.2(a)(i). And, if (ii) holds and neither supp(μ)\mathrm{supp}(\mu) nor supp(ν)\mathrm{supp}(\nu) contains an accumulation point, then they are given by Proposition 3.2(a)(ii).

Proof.

(a) Since FudK(J,J)(x,u)=(u,x)F^{(J,J)}_{udK}(x,u)=(u,x), this part of the result is obvious.
(b) Let a1:=infsupp(μ)a_{1}:=\inf\mathrm{supp}(\mu), a2:=supsupp(μ)a_{2}:=\sup\mathrm{supp}(\mu), b1:=infsupp(ν)b_{1}:=\inf\mathrm{supp}(\nu), b2:=supsupp(ν)b_{2}:=\sup\mathrm{supp}(\nu). Since

0max{x+uJ,0}max{x+uK,0}L,(x,u)2,0\leq\max\{x+u-J,0\}-\max\{x+u-K,0\}\leq L,\qquad\forall(x,u)\in\mathbb{R}^{2},

where L:=KJL:=K-J, FudK(J,K)(x,u)=(y,v)F^{(J,K)}_{udK}(x,u)=(y,v) implies uLyuu-L\leq y\leq u and xvx+Lx\leq v\leq x+L. Thus it holds that

a1b1a1+L,a2b2a2+L.a_{1}\leq b_{1}\leq a_{1}+L,\qquad a_{2}\leq b_{2}\leq a_{2}+L.

Also, by definition, it holds that:

FudK(J,K)(x,u)={(u,x),if x+uJ,(Jx,u+2xJ),if Jx+uK,(uL,x+L),if x+uK,F^{(J,K)}_{udK}(x,u)=\left\{\begin{array}[]{ll}(u,x),&\mbox{if }x+u\leq J,\\ (J-x,u+2x-J),&\mbox{if }J\leq x+u\leq K,\\ (u-L,x+L),&\mbox{if }x+u\geq K,\end{array}\right.

and, in particular, FudK(J,K)(x,u)F^{(J,K)}_{udK}(x,u) is continuous with respect to (x,u)(x,u). We now consider three cases separately: (I) a1+b1<Ja_{1}+b_{1}<J, (II) Ja1+b1<KJ\leq a_{1}+b_{1}<K, (III) a1+b1Ka_{1}+b_{1}\geq K.

  1. (I)

    If a1+b1<Ja_{1}+b_{1}<J, then FudK(J,K)(a1,b1)=(b1,a1)F^{(J,K)}_{udK}(a_{1},b_{1})=(b_{1},a_{1}). This implies a1b1a_{1}\leq b_{1}, b1a1b_{1}\leq a_{1}, and so a1=b1<J2a_{1}=b_{1}<\frac{J}{2}.

  2. (II)

    If Ja1+b1<KJ\leq a_{1}+b_{1}<K, then FudK(J,K)(a1,b1)=(Ja1,b1+2a1J)F^{(J,K)}_{udK}(a_{1},b_{1})=(J-a_{1},b_{1}+2a_{1}-J). Hence a1Ja1a_{1}\leq J-a_{1}, b1b1+2a1Jb_{1}\leq b_{1}+2a_{1}-J, which implies in turn that a1=J2a_{1}=\frac{J}{2} and J2b1<KJ2\frac{J}{2}\leq b_{1}<K-\frac{J}{2}. If a1=a2a_{1}=a_{2}, namely μ\mu is the measure δJ/2\delta_{J/2}, then ν\nu must be concentrated on [b1,KJ2][b_{1},K-\frac{J}{2}]. If a1<a2a_{1}<a_{2}, then there exist ε>0\varepsilon>0, ε0\varepsilon^{\prime}\geq 0 such that a1+εsupp(μ)a_{1}+\varepsilon\in\mathrm{supp}(\mu), b1+εsupp(ν)b_{1}+\varepsilon^{\prime}\in\mathrm{supp}(\nu). In particular, we can take ε\varepsilon^{\prime} small enough so that a1+b1+ε<Ka_{1}+b_{1}+\varepsilon^{\prime}<K. If a1+b1+ε+εKa_{1}+b_{1}+\varepsilon+\varepsilon^{\prime}\leq K, then, J(a1+ε)=J2εsupp(μ)J-(a_{1}+\varepsilon)=\frac{J}{2}-\varepsilon\in\mathrm{supp}(\mu), but this contradicts with the fact that a1=J2a_{1}=\frac{J}{2}. On the other hand, if a1+b1+ε+ε>Ka_{1}+b_{1}+\varepsilon+\varepsilon^{\prime}>K, then b1+εLsupp(μ)b_{1}+\varepsilon^{\prime}-L\in\mathrm{supp}(\mu). However, b1+εL<Ka1L=J2b_{1}+\varepsilon^{\prime}-L<K-a_{1}-L=\frac{J}{2}, which again contradicts with a1=J2a_{1}=\frac{J}{2}. Thus we have shown that it is not possible that a1<a2a_{1}<a_{2}. Consequently, in this case, if FudK(J,K)(μ×ν)=μ×νF^{(J,K)}_{udK}(\mu\times\nu)=\mu\times\nu holds, then μ=δJ2\mu=\delta_{\frac{J}{2}} and supp(ν)[J2,KJ2]\mathrm{supp}(\nu)\subseteq[\frac{J}{2},K-\frac{J}{2}].

  3. (III)

    If a1+b1Ka_{1}+b_{1}\geq K, then FudK(J,K)(x,u)=(uL,x+L)F^{(J,K)}_{udK}(x,u)=(u-L,x+L) for all x,u[a1,a2]×[b1,b2]x,u\in[a_{1},a_{2}]\times[b_{1},b_{2}], so FudK(J,K)(μ×ν)=μ×νF^{(J,K)}_{udK}(\mu\times\nu)=\mu\times\nu holds if and only if ν=μ+L\nu=\mu+L.

We next consider the corresponding three cases for the suprema of the support: (I’) a2+b2Ja_{2}+b_{2}\leq J, (II’) J<a2+b2KJ<a_{2}+b_{2}\leq K, (III’) a2+b2>Ka_{2}+b_{2}>K. By a similar argument to above, we have the following.

  1. (I’)

    If a2+b2Ja_{2}+b_{2}\leq J, then FudK(J,K)(μ×ν)=μ×νF^{(J,K)}_{udK}(\mu\times\nu)=\mu\times\nu holds if and only if ν=μ\nu=\mu.

  2. (II’)

    If J<a2+b2KJ<a_{2}+b_{2}\leq K and FudK(J,K)(μ×ν)=μ×νF^{(J,K)}_{udK}(\mu\times\nu)=\mu\times\nu holds, then μ=δJ2\mu=\delta_{\frac{J}{2}}, supp(ν)[J2,KJ2]\mathrm{supp}(\nu)\subseteq[\frac{J}{2},K-\frac{J}{2}].

  3. (III’)

    If a2+b2>Ka_{2}+b_{2}>K, then b2=a2+Lb_{2}=a_{2}+L and a2>J2a_{2}>\frac{J}{2}.

Putting together the above discussion, there are only four possible cases: (I”-1) μ=ν\mu=\nu and a2J2a_{2}\leq\frac{J}{2}; (I”-2) μ=δJ2\mu=\delta_{\frac{J}{2}}, supp(ν)[J2,KJ2]\mathrm{supp}(\nu)\subseteq[\frac{J}{2},K-\frac{J}{2}]; (I”-3) μ+L=ν\mu+L=\nu and a1J2a_{1}\geq\frac{J}{2}; (II”) a1=b1a_{1}=b_{1}, a2=b2La_{2}=b_{2}-L and a1<J2a_{1}<\frac{J}{2}, a2>J2a_{2}>\frac{J}{2}. 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 c1,c2>0c_{1},c_{2}>0 such that a1=b1=J2c1a_{1}=b_{1}=\frac{J}{2}-c_{1} and a2=b2L=J2+c2a_{2}=b_{2}-L=\frac{J}{2}+c_{2}. Suppose c1>c2c_{1}>c_{2}. Then, a1+b2=Jc1+c2+L=Kc1+c2<Ka_{1}+b_{2}=J-c_{1}+c_{2}+L=K-c_{1}+c_{2}<K. If Ja1+b2<KJ\leq a_{1}+b_{2}<K, then FudK(J,K)(a1,b2)=(Ja1,b2+2a1J)F^{(J,K)}_{udK}(a_{1},b_{2})=(J-a_{1},b_{2}+2a_{1}-J), and so Ja1a2J-a_{1}\leq a_{2}. The latter inequality is equivalent to c1c2c_{1}\leq c_{2}, which contradicts c1>c2c_{1}>c_{2}. If a1+b2<Ja_{1}+b_{2}<J, then FudK(J,K)(a1,b2)=(b2,a1)F^{(J,K)}_{udK}(a_{1},b_{2})=(b_{2},a_{1}), which implies b2a2b_{2}\leq a_{2}. However, this contradicts a2=b2La_{2}=b_{2}-L. Hence c1c2c_{1}\leq c_{2}. A similar argument allows one to deduce the reverse inequality, and thus we obtain c1=c2c_{1}=c_{2}. In conclusion, letting c=J2c1c=\frac{J}{2}-c_{1}, we obtain the desired result.

To complete the proof, we study the special cases where μ\mu and ν\nu have densities, or they are discrete. Let fμf_{\mu}, fνf_{\nu} be densities of μ\mu and ν\nu. For x[c,Jc]x\in[c,J-c] and u[c,Kc]u\in[c,K-c], we then have that

fμ(x)fν(u)={fμ(u)fν(x),if x+uJ,fμ(Jx)fν(u+2xJ),if Jx+uK,fμ(uL)fν(x+L),if x+uK.f_{\mu}(x)f_{\nu}(u)=\left\{\begin{array}[]{ll}f_{\mu}(u)f_{\nu}(x),&\mbox{if }x+u\leq J,\\ f_{\mu}(J-x)f_{\nu}(u+2x-J),&\mbox{if }J\leq x+u\leq K,\\ f_{\mu}(u-L)f_{\nu}(x+L),&\mbox{if }x+u\geq K.\end{array}\right.

Letting hμ(x):=logfμ(x)h_{\mu}(x):=\log f_{\mu}(x) and hν(u):=logfν(u)h_{\nu}(u):=\log f_{\nu}(u) and taking derivatives of the relation

hμ(x)+hν(u)=hμ(Jx)+hν(u+2xJ)h_{\mu}(x)+h_{\nu}(u)=h_{\mu}(J-x)+h_{\nu}(u+2x-J)

with respect to xx first and then with respect to uu, for (x,u)(x,u) satisfying Jx+uKJ\leq x+u\leq K, we have hν′′(u+2xJ)=0h^{\prime\prime}_{\nu}(u+2x-J)=0. For any v[c,Kc]v\in[c,K-c], by letting ε:=vcK2c[0,1]\varepsilon:=\frac{v-c}{K-2c}\in[0,1] and

x=c+ε(J2c),u=c+(1ε)(J2c)+ε(KJ),x=c+\varepsilon(J-2c),\qquad u=c+(1-\varepsilon)(J-2c)+\varepsilon(K-J),

we have x[c,Jc]x\in[c,J-c], u[c,Kc]u\in[c,K-c], Jx+uKJ\leq x+u\leq K and v=u+2xJv=u+2x-J, so hν′′(v)=0h^{\prime\prime}_{\nu}(v)=0 for all v[c,Kc]v\in[c,K-c]. Therefore, there exists λ\lambda\in{\mathbb{R}} such that hν(u)=λh^{\prime}_{\nu}(u)=\lambda for all u[c,Kc]u\in[c,K-c]. Also, by taking the derivative of hμ(x)+hν(u)=hμ(u)+hν(x)h_{\mu}(x)+h_{\nu}(u)=h_{\mu}(u)+h_{\nu}(x) with respect to xx, for (x,u)(x,u) satisfying x+uJx+u\leq J, we have hμ(x)=hν(x)h^{\prime}_{\mu}(x)=h^{\prime}_{\nu}(x). Since for any x[c,Jc]x\in[c,J-c], by taking u=Jxu=J-x, we have u[c,Kc]u\in[c,K-c] and x+uJx+u\leq J, hence hμ(x)=λh^{\prime}_{\mu}(x)=\lambda for all x[c,Jc]x\in[c,J-c]. Therefore, since μ\mu and ν\nu are probability measures, cc must be finite. Moreover, if K=K=\infty, then λ\lambda must be positive.

Finally, we consider the case where μ\mu and ν\nu are discrete. We first prove that c>c>-\infty. If c=c=-\infty, then for any xsupp(μ)x\in\mathrm{supp}(\mu), there exists usupp(ν)u\in\mathrm{supp}(\nu) such that x+uJx+u\leq J, and vice versa. Since for such x,ux,u we have FudK(J,K)(x,u)=(u,x)F_{udK}^{(J,K)}(x,u)=(u,x), we conclude that supp(μ)=supp(ν)\mathrm{supp}(\mu)=\mathrm{supp}(\nu). Moreover, by noting μ({x})ν({u})=μ({u})ν({x})\mu(\{x\})\nu(\{u\})=\mu(\{u\})\nu(\{x\}) for x+uJx+u\leq J, it is an elementary exercise to check that μ=ν\mu=\nu. Next, note that if K<K<\infty, then for any x+uKx+u\geq K with x,usupp(μ)x,u\in\mathrm{supp}(\mu), it holds that x+L,uLsupp(μ)x+L,u-L\in\mathrm{supp}(\mu) and

μ({x})μ({x+L})=μ({uL})μ({u}).\frac{\mu(\{x\})}{\mu(\{x+L\})}=\frac{\mu(\{u-L\})}{\mu(\{u\})}.

However, since supsupp(μ)=\sup\mathrm{supp}(\mu)=\infty, this implies that for any xsupp(μ)x\in\mathrm{supp}(\mu), x+nLsupp(μ)x+nL\in\mathrm{supp}(\mu) for all nn\in{\mathbb{Z}} and μ({x+nL})=μ({x})λn\mu(\{x+nL\})=\mu(\{x\})\lambda^{n} for some λ0\lambda\neq 0, which can not happen since μ\mu is a probability measure. Similarly, if K=K=\infty, then for any x,n,mx,n,m satisfying J2+nx,J2+mxsupp(μ)\frac{J}{2}+nx,\frac{J}{2}+mx\in\mathrm{supp}(\mu) and n+m0n+m\geq 0, we have J2nx\frac{J}{2}-nx, J2+(2n+m)xsupp(μ)\frac{J}{2}+(2n+m)x\in\mathrm{supp}(\mu) and

μ({J2+nx})μ({J2+(2n+m)x})=μ({J2nx})μ({J2+mx}).\frac{\mu(\{\frac{J}{2}+nx\})}{\mu(\{\frac{J}{2}+(2n+m)x\})}=\frac{\mu(\{\frac{J}{2}-nx\})}{\mu(\{\frac{J}{2}+mx\})}.

In particular, applying this relation with x0x\geq 0 satisfying J2+xsupp(μ)\frac{J}{2}+x\in\mathrm{supp}(\mu) with n=m=1n=m=1, we have J2x,J2+3xsupp(μ)\frac{J}{2}-x,\frac{J}{2}+3x\in\mathrm{supp}(\mu). Iterating this argument yields J2+(2n+1)xsupp(μ)\frac{J}{2}+(2n+1)x\in\mathrm{supp}(\mu) for all nn\in{\mathbb{Z}}, and

μ({J2+x})μ({J2x})=μ({J2+(2n+1)x})μ({J2+(2n1)x})\frac{\mu(\{\frac{J}{2}+x\})}{\mu(\{\frac{J}{2}-x\})}=\frac{\mu(\{\frac{J}{2}+(2n+1)x\})}{\mu(\{\frac{J}{2}+(2n-1)x\})}

for all n0n\geq 0. Moreover, since for n1n\leq-1,

μ({J2+(2n1)x})μ({J2+(2n+1)x})=μ({J2(2n1)x})μ({J2(2n3)x}),\frac{\mu(\{\frac{J}{2}+(2n-1)x\})}{\mu(\{\frac{J}{2}+(2n+1)x\})}=\frac{\mu(\{\frac{J}{2}-(2n-1)x\})}{\mu(\{\frac{J}{2}-(2n-3)x\})},

we have μ({J2+(2n+1)x})=μ({J2+x}))λn\mu(\{\frac{J}{2}+(2n+1)x\})=\mu(\{\frac{J}{2}+x\}))\lambda^{n} for all nn\in\mathbb{Z}, for some λ0\lambda\neq 0. Again, this can not happen since μ\mu is a probability measure. We can therefore conclude that c>c>-\infty.

First suppose K<K<\infty. We then have supp(μ)={x0,x1,,xn}\mathrm{supp}(\mu)=\{x_{0},x_{1},\dots,x_{n}\} for some nn with c=x0<x1<<xn=Jcc=x_{0}<x_{1}<\dots<x_{n}=J-c, and supp(μ)c+b\mathrm{supp}(\mu)\subset c+b\mathbb{Z} for some b>0b>0. Additionally, supp(ν)={u0,u1,,um}\mathrm{supp}(\nu)=\{u_{0},u_{1},\dots,u_{m}\} for some mm with c=u0<u1<<um=Kcc=u_{0}<u_{1}<\dots<u_{m}=K-c, and supp(ν)c+b\mathrm{supp}(\nu)\subset c+b^{\prime}\mathbb{Z} for some b>0b^{\prime}>0. By a similar argument to the previous paragraph, it is possible to check that, for an appropriate choice of bb, one may take b=bb^{\prime}=b, and moreover xi=c+ibx_{i}=c+ib, ui=c+ibu_{i}=c+ib for each ii. Hence, by making the change of variables xxc,uuc,JJ2c,KK2cx\to x-c,u\to u-c,J\to J-2c,K\to K-2c, we can apply [10, Lemma 4.5] to complete the proof. To establish the result when K=K=\infty, one can proceed in the same way to check that supp(ν)[c,Jc]=supp(μ)=(c+b)[c,Jc]\mathrm{supp}(\nu)\cap[c,J-c]=\mathrm{supp}(\mu)=(c+b\mathbb{Z})\cap[c,J-c] for some b>0b>0, and then use the identity μ({c})ν({J+x})=μ({Jc})ν({x+2c})\mu(\{c\})\nu(\{J+x\})=\mu(\{J-c\})\nu(\{x+2c\}) for xcx\geq-c to derive the full support of ν\nu, 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 μ\mu in the case JKJ\leq K, and ν\nu in the case JKJ\geq K).

We start by restricting our attention to JKJ\leq K. 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 Fn,t=FudK(J,K)F_{n,t}=F^{(J,K)}_{udK} for all n,tn,t, where JKJ\leq K. For J=KJ=K, we set 𝒳J,K:=\mathcal{X}^{*}_{J,K}:=\mathbb{R}^{\mathbb{Z}}. For J<K=J<K=\infty, we take

𝒳J,K:={(xn)n:lim|n|k=1n(J2xk)n>0},\mathcal{X}^{*}_{J,K}:=\left\{(x_{n})_{n\in\mathbb{Z}}:\>\lim_{|n|\to\infty}\frac{\sum_{k=1}^{n}\left(J-2x_{k}\right)}{n}>0\right\},

where for n<0n<0, the sum k=1n\sum_{k=1}^{n} should be interpreted as n+10-\sum^{0}_{n+1}, and, for J<K<J<K<\infty,

𝒳J,K:={(xn)n:lim supn±|k=1n(J2xk)|=}.\mathcal{X}^{*}_{J,K}:=\left\{(x_{n})_{n\in\mathbb{Z}}:\>\limsup_{n\to\pm\infty}\left|\sum_{k=1}^{n}\left(J-2x_{k}\right)\right|=\infty\right\}.

From results of [10, 13], we then have the following.

Lemma 3.4.

Suppose JKJ\leq K. If (xn)n𝒳J,K(x_{n})_{n\in\mathbb{Z}}\in\mathcal{X}^{*}_{J,K}, then there exists a unique solution of (1.1) with Fn,t=FudK(J,K)F_{n,t}=F^{(J,K)}_{udK} for all n,tn,t.

Proof.

In the case J=KJ=K, we have FudK(J,K)(x,u)=(u,x)F^{(J,K)}_{udK}(x,u)=(u,x), and so the result is clear. In the case J<K=J<K=\infty, the result is given by [13, Theorem 2.1]. For the case J<K<J<K<\infty, the result is given for BBS(JJ,KK) in [10], i.e. for J,KJ,K\in{\mathbb{N}} and x{0,1,2,,J}x\in\{0,1,2,\dots,J\}, u{0,1,2,,K}u\in\{0,1,2,\dots,K\}. The same proof applies in the more general case. ∎

To handle the case =J>K\infty=J>K, we consider the set

(3.6) 𝒳J,K!:={(xn)n:lim supn𝟏{xn+xn+1K}=1},\mathcal{X}^{!}_{J,K}:=\left\{(x_{n})_{n\in\mathbb{Z}}:\>\limsup_{n\to-\infty}\mathbf{1}_{\{x_{n}+x_{n+1}\leq K\}}=1\right\},

and for >J>K\infty>J>K, the set

(3.7) 𝒳J,K!:={(xn)n:lim supn𝟏{xn+xn+1K}{xn+xn+12JK}=1}.\mathcal{X}^{!}_{J,K}:=\left\{(x_{n})_{n\in\mathbb{Z}}:\>\limsup_{n\to-\infty}\mathbf{1}_{\{x_{n}+x_{n+1}\leq K\}\cup\{x_{n}+x_{n+1}\geq 2J-K\}}=1\right\}.

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 J>KJ>K. If (xn)n𝒳J,K!(x_{n})_{n\in\mathbb{Z}}\in\mathcal{X}^{!}_{J,K}, then there exists at most one sequence (un)n(u_{n})_{n\in\mathbb{Z}} such that

(3.8) (Fudk(J,K))(2)(xn,un1)=un,n.\left(F^{(J,K)}_{udk}\right)^{(2)}(x_{n},u_{n-1})=u_{n},\qquad\forall n\in\mathbb{Z}.
Proof.

We first prove that if xn+xn+1Kx_{n}+x_{n+1}\leq K, then un+1=xn+1u_{n+1}=x_{n+1}. Since J>KJ>K and

un=xnmax{xn+un1K,0}+max{xn+un1J,0},u_{n}=x_{n}-\max\{x_{n}+u_{n-1}-K,0\}+\max\{x_{n}+u_{n-1}-J,0\},

it must hold that unxnu_{n}\leq x_{n}. Hence xn+1+unxn+1+xnKx_{n+1}+u_{n}\leq x_{n+1}+x_{n}\leq K, and so

un+1=xn+1max{xn+1+unK,0}+max{xn+1+unJ,0}=xn+1.u_{n+1}=x_{n+1}-\max\{x_{n+1}+u_{n}-K,0\}+\max\{x_{n+1}+u_{n}-J,0\}=x_{n+1}.

Similarly, if >J>K\infty>J>K and xn+xn+12JKx_{n}+x_{n+1}\geq 2J-K, then un+1=xn+1Lu_{n+1}=x_{n+1}-L, where L=JKL=J-K. Indeed, since unxnLu_{n}\geq x_{n}-L, in this case we have that xn+1+un2JKL=Jx_{n+1}+u_{n}\geq 2J-K-L=J, and the result follows. As a consequence, if (xn)n𝒳J,K!(x_{n})_{n\in\mathbb{Z}}\in\mathcal{X}^{!}_{J,K}, then there exists a sequence nkn_{k}\downarrow-\infty such that unku_{n_{k}} is determined by xnkx_{n_{k}}. For n{nk:k1}n\notin\{n_{k}:k\geq 1\}, the relation (3.8) means that unu_{n} is uniquely defined by unku_{n_{k}} such that nk<nn_{k}<n and (xm)nk+1mn(x_{m})_{n_{k}+1\leq m\leq n}, 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 𝒯udK(J,K)\mathcal{T}_{udK}^{(J,K)} for the dynamics given by FudK(J,K)F_{udK}^{(J,K)}, as defined at (1.3).

Theorem 3.6.

The product measure μ\mu^{\mathbb{Z}} satisfies 𝒯udK(J,K)μ=μ\mathcal{T}_{udK}^{(J,K)}\mu^{{\mathbb{Z}}}=\mu^{{\mathbb{Z}}} for the following measures μ\mu.

  1. (a)

    Suppose J=KJ=K. Any measure μ\mu on \mathbb{R}.

  2. (b)

    Suppose J<KJ<K. Excluding μ=δJ/2\mu=\delta_{J/2}, any measure μ\mu given by Proposition 3.2(a) or Proposition 3.2(c).

  3. (c)

    Suppose J>KJ>K. Excluding μ=δK/2\mu=\delta_{K/2} and μ=δJK/2\mu=\delta_{J-K/2}, any measure μ\mu given by Proposition 3.2(a) or supported on (,K2](-\infty,\frac{K}{2}] or [JK2,)[J-\frac{K}{2},\infty).

Proof.

(a) The case J=KJ=K is obvious.
(b) In the case J<K=J<K=\infty, for one of the measures μ\mu^{\mathbb{Z}} from Proposition 3.2 to satisfy 𝒯μ=μ\mathcal{T}\mu^{{\mathbb{Z}}}=\mu^{{\mathbb{Z}}}, it will suffice to check that μ(𝒳J,K)=1\mu^{{\mathbb{Z}}}(\mathcal{X}^{*}_{J,K})=1. For this, the law of large numbers tells us that it is sufficient for xμ(dx)<J/2\int x\mu(dx)<J/2. The measures given in the statement of the theorem are readily checked to satisfy this requirement. Finally, for J<K<J<K<\infty, it will again be enough to determine measures μ\mu^{\mathbb{Z}} from Proposition 3.2 that satisfy μ(𝒳J,K)=1\mu^{{\mathbb{Z}}}(\mathcal{X}^{*}_{J,K})=1. The latter constraint simply rules out the trivial measure μ=δJ/2\mu=\delta_{J/2}, and so the result readily follows.
(c) Let us continue for the moment to suppose that J<KJ<K. We will appeal to the configuration-carrier duality of (3.2) to prove the result, and as a first step we take μ\mu to be one of the measures identified in part (b). If (xnt,unt)n,t(x_{n}^{t},u_{n}^{t})_{n,t\in\mathbb{Z}} is given by solving the initial value problem (1.1) with initial condition (xn)nμ(x_{n})_{n\in\mathbb{Z}}\sim\mu^{\mathbb{Z}}, it then readily follows from Proposition 2.9 that, for each nn\in\mathbb{Z}, (unt)t(u_{n}^{t})_{t\in\mathbb{Z}} is i.i.d., with marginal given by the corresponding ν\nu from Proposition 3.2. Now, as long as ν((,J2][KJ2,))>0\nu((-\infty,\frac{J}{2}]\cup[K-\frac{J}{2},\infty))>0, then it is clear that ν(𝒳K,J!)=1\nu^{\mathbb{Z}}(\mathcal{X}_{K,J}^{!})=1. This means that, μ\mu^{\mathbb{Z}}-a.s., (unt)t(u_{n}^{t})_{t\in\mathbb{Z}} uniquely determines (xn+1t,un+1t)t(x_{n+1}^{t},u_{n+1}^{t})_{t\in\mathbb{Z}}, with un+1=𝒯udK(K,J)unu_{n+1}=\mathcal{T}_{udK}^{(K,J)}u_{n}, where 𝒯udK(K,J)\mathcal{T}_{udK}^{(K,J)} represents the dynamics given by FudK(K,J)F_{udK}^{(K,J)} (cf. the proof of Theorem 2.11). In particular, we have demonstrated that 𝒯udK(K,J)ν=ν\mathcal{T}_{udK}^{(K,J)}\nu^{\mathbb{Z}}=\nu^{\mathbb{Z}}. Reversing the role of JJ and KK gives the result. ∎

3.1.4. Ergodicity

Finally, we study the ergodicity of the operator 𝒯udK(J,K)\mathcal{T}_{udK}^{(J,K)}. 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 JKJ\leq K. Let μ×ν\mu\times\nu be a product measure satisfying FudK(J,K)(μ×ν)=μ×νF_{udK}^{(J,K)}(\mu\times\nu)=\mu\times\nu, as given by Proposition 3.2, with μδJ/2\mu\neq\delta_{J/2} and ν((,J2][KJ2,))>0\nu((-\infty,\frac{J}{2}]\cup[K-\frac{J}{2},\infty))>0. It is then the case that μ\mu^{{\mathbb{Z}}} is ergodic under 𝒯udK(J,K)\mathcal{T}_{udK}^{(J,K)}, and ν\nu^{{\mathbb{Z}}} is ergodic under 𝒯udK(K,J)\mathcal{T}_{udK}^{(K,J)}.

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 α,β0\alpha,\beta\geq 0, and is based on the following lattice map:

(dKdV) FdK(α,β)(x,u)=(u(1+βxu)1+αxu,x(1+αxu)1+βxu),\displaystyle F^{(\alpha,\beta)}_{dK}(x,u)=\left(\frac{u(1+\beta xu)}{1+\alpha xu},\frac{x(1+\alpha xu)}{1+\beta xu}\right),

where we now assume the variables xx and uu are (0,)(0,\infty) valued. We note that FdK(α,β)F^{(\alpha,\beta)}_{dK} 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 β=0\beta=0, then FdK(α,β)F^{(\alpha,\beta)}_{dK} 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 (x,u)(0,)2(x,u)\in(0,\infty)^{2}, it holds that

FdK(α,β)FdK(α,β)(x,u)=(x,u).F^{(\alpha,\beta)}_{dK}\circ F^{(\alpha,\beta)}_{dK}(x,u)=(x,u).
Configuration-carrier duality:

If π(x,u):=(u,x)\pi(x,u):=(u,x), then

FdK(α,β)=πFdK(β,α)π.F^{(\alpha,\beta)}_{dK}=\pi\circ F^{(\beta,\alpha)}_{dK}\circ\pi.
Empty space-particle duality:

Suppose α,β>0\alpha,\beta>0. If σα,β(x,u):=(1αx,1βu)\sigma_{\alpha,\beta}(x,u):=(\frac{1}{\alpha x},\frac{1}{\beta u}), then

FdK(α,β)=σα,βFdK(α,β)σα,β.F^{(\alpha,\beta)}_{dK}=\sigma_{\alpha,\beta}\circ F^{(\alpha,\beta)}_{dK}\circ\sigma_{\alpha,\beta}.
Scale invariance:

If λ>0\lambda>0, then for any (x,u)(0,)2(x,u)\in(0,\infty)^{2} it holds that

FdK(λ2α,λ2β)(λx,λu)=λFdK(α,β)(x,u).F^{(\lambda^{-2}\alpha,\lambda^{-2}\beta)}_{dK}(\lambda x,\lambda u)=\lambda F^{(\alpha,\beta)}_{dK}(x,u).

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 μ×ν\mu\times\nu satisfy FdK(α,β)(μ×ν)=μ×νF^{(\alpha,\beta)}_{dK}(\mu\times\nu)=\mu\times\nu.

  1. (a)

    For any λ\lambda\in\mathbb{R} if αβ>0\alpha\beta>0, or λ>0\lambda>0 if αβ=0\alpha\beta=0, and c>0c>0,

    μ×ν=GIG(λ,cα,c)×GIG(λ,cβ,c).\mu\times\nu=\mathrm{GIG}(\lambda,c\alpha,c)\times\mathrm{GIG}(\lambda,c\beta,c).
  2. (b)

    Suppose α=β\alpha=\beta. For any measure mm on (0,)(0,\infty),

    μ×ν=m×m.\mu\times\nu=m\times m.

In the case αβ=0\alpha\beta=0, 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 FdK(α,β)F_{dK}^{(\alpha,\beta)} is equal to one, it suffices to check that the following relation between joint densities:

xλ1ecαxcx1uλ1ecβucu1=yλ1ecαycy1vλ1ecβvcv1,x^{-\lambda-1}e^{-c\alpha x-cx^{-1}}u^{-\lambda-1}e^{-c\beta u-cu^{-1}}=y^{-\lambda-1}e^{-c\alpha y-cy^{-1}}v^{-\lambda-1}e^{-c\beta v-cv^{-1}},

where y=u(1+βxu)1+αxuy=\frac{u(1+\beta xu)}{1+\alpha xu} and v=x(1+αxu)1+βxuv=\frac{x(1+\alpha xu)}{1+\beta xu}. This is a simple consequence of the identities xu=yvxu=yv and αx+x1+βu+u1=αy+y1+βv+v1\alpha x+x^{-1}+\beta u+u^{-1}=\alpha y+y^{-1}+\beta v+v^{-1}, which can be checked directly.
(b) Since FdK(α,α)(x,u)=(u,x)F^{(\alpha,\alpha)}_{dK}(x,u)=(u,x), the result is obvious.
For the final part of the result, suppose α>0=β\alpha>0=\beta. In this case, the map of interest becomes

FdK(α,0)(x,u)=(u1+αxu,x(1+αxu))=(1αx+u1,α1(1αx1αx+u1)1).F^{(\alpha,0)}_{dK}(x,u)=\left(\frac{u}{1+\alpha xu},x(1+\alpha xu)\right)=\left(\frac{1}{\alpha x+u^{-1}},\alpha^{-1}\left(\frac{1}{\alpha x}-\frac{1}{\alpha x+u^{-1}}\right)^{-1}\right).

Now, in [28, Theorem 4.1], it is shown that if XX and YY are strictly positive independent random variables such that at least one of XX and YY has a non-trivial distribution, and (X+Y)1(X+Y)^{-1} and X1(X+Y)1X^{-1}-(X+Y)^{-1} are also independent, then XX must have a generalized inverse Gaussian distribution and YY must have a gamma distribution with related parameters. (NB. This result builds on [27].) Considering the form of the map FdK(α,0)F^{(\alpha,0)}_{dK} as given above, and applying [28, Theorem 4.1] with X=αxX=\alpha x, Y=u1Y=u^{-1} yields the result. ∎

3.2.3. Invariant measures

We now show how the measures of Proposition 3.9 yield invariant product measures for 𝒯dKα,β\mathcal{T}_{dK}^{\alpha,\beta}, that is, the operator describing the (dKdV) dynamics. Apart from the trivial case α=β\alpha=\beta, we restrict our attention to the case when αβ=0\alpha\beta=0. (We list the case αβ>0\alpha\beta>0 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 Fn,t=FdK(α,β)F_{n,t}=F^{(\alpha,\beta)}_{dK} for all n,tn,t, where α>0\alpha>0 and β=0\beta=0. Letting

𝒳α:={(xn)n(0,):lim|n|k=1n(logα2logxk)n>0},\mathcal{X}^{*}_{\alpha}:=\left\{(x_{n})_{n\in\mathbb{Z}}\in(0,\infty)^{{\mathbb{Z}}}:\>\lim_{|n|\to\infty}\frac{\sum_{k=1}^{n}(-\log\alpha-2\log x_{k})}{n}>0\right\},

we have the following result (see [13, Theorem 2.2]).

Lemma 3.10.

Suppose α>0\alpha>0. If (xn)n𝒳α(x_{n})_{n\in\mathbb{Z}}\in\mathcal{X}^{*}_{\alpha}, then there exists a unique solution of (1.1) with Fn,t=FdK(α,0)F_{n,t}=F^{(\alpha,0)}_{dK} for all n,tn,t.

In the case α=0\alpha=0, β>0\beta>0, we consider the set

𝒳β!:={(xn)n(0,):n=0xn1=limnSn=,limn(logxn)Sn1=0},\mathcal{X}^{\exists!}_{\beta}:=\left\{(x_{n})_{n\in\mathbb{Z}}\in(0,\infty)^{{\mathbb{Z}}}:\>\sum_{n=-\infty}^{0}x_{n}^{-1}=\lim_{n\rightarrow\infty}-S_{-n}=\infty,\>\lim_{n\rightarrow-\infty}(\log x_{n})S_{n}^{-1}=0\right\},

where Sn:=m=n+10(logβ2logxm)S_{-n}:=\sum_{m=-n+1}^{0}(-\log\beta-2\log x_{m}). The parallel to Lemma 3.5 that we apply in the discrete setting is the following.

Lemma 3.11.

Suppose α=0\alpha=0, β>0\beta>0. If (xn)n𝒳β!(x_{n})_{n\in\mathbb{Z}}\in\mathcal{X}^{\exists!}_{\beta}, then there exists precisely one sequence (un)n(0,)(u_{n})_{n\in\mathbb{Z}}\in(0,\infty)^{{\mathbb{Z}}} such that

(3.9) (FdK(0,β))(2)(xn,un1)=un,n,\left(F^{(0,\beta)}_{dK}\right)^{(2)}(x_{n},u_{n-1})=u_{n},\qquad\forall n\in\mathbb{Z},

which is explicitly given by the infinite continued fraction

un=1β1(βxn)1+1(βxn1)1+.u_{n}=\frac{1}{\sqrt{\beta}}\frac{1}{(\sqrt{\beta}x_{n})^{-1}+\frac{1}{(\sqrt{\beta}x_{n-1})^{-1}+\dots}}.
Proof.

The relation (3.9) can be written as

un=xn1+βxnun1,u_{n}=\frac{x_{n}}{1+\beta x_{n}u_{n-1}},

which is equivalent to

βun=1(βxn)1+βun1.\sqrt{\beta}u_{n}=\frac{1}{(\sqrt{\beta}x_{n})^{-1}+\sqrt{\beta}u_{n-1}}.

Hence, the sequence defined by

βun=1(βxn)1+1(βxn1)1+\sqrt{\beta}u_{n}=\frac{1}{(\sqrt{\beta}x_{n})^{-1}+\frac{1}{(\sqrt{\beta}x_{n-1})^{-1}+\dots}}

satisfies (3.9). Indeed, the condition n=0xn1=\sum_{n=-\infty}^{0}x_{n}^{-1}=\infty ensures that the infinite continued fraction converges in (0,)(0,\infty) (see [29, Chapter 8], for example). Suppose that we have another solution (u~n)n(0,)(\tilde{u}_{n})_{n\in\mathbb{Z}}\in(0,\infty)^{{\mathbb{Z}}} to (3.9). It is then the case that

|unu~n|=unu~n|un1u~n1|=βunu~n|un1u~n1|βxn2|un1u~n1|.\left|u_{n}-\tilde{u}_{n}\right|=u_{n}\tilde{u}_{n}\left|u_{n}^{-1}-\tilde{u}_{n}^{-1}\right|=\beta u_{n}\tilde{u}_{n}\left|u_{n-1}-\tilde{u}_{n-1}\right|\leq\beta x_{n}^{2}\left|u_{n-1}-\tilde{u}_{n-1}\right|.

Iterating this, we find that for any mnm\leq n,

|unu~n|k=mn(βxk2)×xm1=exp(k=mn(logβ+2logxk)+logxm1).\left|u_{n}-\tilde{u}_{n}\right|\leq\prod_{k=m}^{n}(\beta x_{k}^{2})\times x_{m-1}=\exp\left(\sum_{k=m}^{n}(\log\beta+2\log x_{k})+\log x_{m-1}\right).

Taking the limit as mm\rightarrow-\infty, the defining properties of 𝒳β!\mathcal{X}^{\exists!}_{\beta} imply that un=u~nu_{n}=\tilde{u}_{n}, 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 μ=GIG(λ,cα,c)\mu=GIG(\lambda,c\alpha,c) and ν=IG(λ,c)\nu=IG(\lambda,c), then 2log(x)μ(dx)2log(x)ν(dx)2\int\log(x)\mu(dx)\leq 2\int\log(x)\nu(dx) (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 μ\mu^{\mathbb{Z}} satisfies 𝒯dK(α,β)μ=μ\mathcal{T}_{dK}^{(\alpha,\beta)}\mu^{{\mathbb{Z}}}=\mu^{{\mathbb{Z}}} for the following measures μ\mu.

  1. (a)

    Suppose α=β\alpha=\beta. Any measure μ\mu on \mathbb{R}.

  2. (b)

    Suppose α>0\alpha>0, β=0\beta=0. The measure μ=GIG(λ,cα,c)\mu=GIG(\lambda,c\alpha,c) for any parameters λ,c>0\lambda,c>0 such that 2log(x)μ(dx)<logα2\int\log(x)\mu(dx)<-\log\alpha.

  3. (c)

    Suppose α=0\alpha=0, β>0\beta>0. The measure μ=IG(λ,c)\mu=IG(\lambda,c) for any parameters λ,c>0\lambda,c>0 such that 2log(x)μ(dx)<logβ2\int\log(x)\mu(dx)<-\log\beta.

3.2.4. Ergodicity

Regarding the ergodicity of 𝒯dK(α,β)\mathcal{T}_{dK}^{(\alpha,\beta)}, combining the results of the previous section with Theorem 2.11 gives the next result.

Theorem 3.13.

Suppose α>0\alpha>0, β=0\beta=0. Let μ×ν\mu\times\nu be a product measure satisfying FdK(α,0)(μ×ν)=μ×νF_{dK}^{(\alpha,0)}(\mu\times\nu)=\mu\times\nu, as given by Proposition 3.9 (i.e. μ×ν=GIG(λ,cα,c)×IG(λ,c)\mu\times\nu=GIG(\lambda,c\alpha,c)\times IG(\lambda,c). If it holds that 2log(x)ν(dx)<logα2\int\log(x)\nu(dx)<-\log\alpha, it is then the case that μ\mu^{{\mathbb{Z}}} is ergodic under 𝒯dK(α,0)\mathcal{T}_{dK}^{(\alpha,0)}, and ν\nu^{{\mathbb{Z}}} is ergodic under 𝒯dK(0,α)\mathcal{T}_{dK}^{(0,\alpha)}.

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) {Qnt+1=min{Unt,Ent},Ent+1=Qn+1t+EntQnt+1,Un+1t=Unt+Qn+1tQnt+1,\begin{cases}Q_{n}^{t+1}=\min\{U_{n}^{t},E_{n}^{t}\},\\ E_{n}^{t+1}=Q_{n+1}^{t}+E_{n}^{t}-Q_{n}^{t+1},\\ U_{n+1}^{t}=U_{n}^{t}+Q_{n+1}^{t}-Q_{n}^{t+1},\end{cases}

where (Qnt,Ent,Unt)n,t(Q_{n}^{t},E_{n}^{t},U_{n}^{t})_{n,t\in\mathbb{Z}} take values in \mathbb{R}. We summarise this evolution as (Qnt+1,Ent+1,Un+1t)=FudT(Qn+1t,Ent,Unt)(Q_{n}^{t+1},E_{n}^{t+1},U_{n+1}^{t})=F_{udT}(Q_{n+1}^{t},E_{n}^{t},U_{n}^{t}), highlighting that FudTF_{udT} is an involution on 3\mathbb{R}^{3}, and represent the lattice structure diagrammatically as

(4.1) Qnt+1\textstyle{Q_{n}^{t+1}}Ent+1\textstyle{E_{n}^{t+1}}Unt\textstyle{U_{n}^{t}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Un+1t.\textstyle{U_{n+1}^{t}.}Ent\textstyle{E_{n}^{t}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Qn+1t\textstyle{Q_{n+1}^{t}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

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 QntQ_{n}^{t} as the length of the nnth interval containing balls, EntE_{n}^{t} as the length of the nnth empty interval (at time tt), and UntU_{n}^{t} 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 FudTF_{udT} with three inputs and three outputs into two maps FudTF_{udT^{*}} and FudT1F_{udT^{*}}^{-1}, each with two inputs and two outputs:

            FudT\textstyle{\boxed{F_{udT^{*}}\vphantom{F_{udT^{*}}^{-1}}}}min{b,c}\textstyle{\min\{b,c\}}FudT1\textstyle{\boxed{F_{udT^{*}}^{-1}}}a+max{bc,0}\textstyle{a+\max\{b-c,0\}}c\textstyle{c\ignorespaces\ignorespaces\ignorespaces\ignorespaces}bc\textstyle{b-c\ignorespaces\ignorespaces\ignorespaces\ignorespaces}amin{bc,0},\textstyle{a-\min\{b-c,0\},}b\textstyle{b\ignorespaces\ignorespaces\ignorespaces\ignorespaces}a\textstyle{a\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

where we generically take (a,b,c)=(Qn+1t,Ent,Unt)(a,b,c)=(Q_{n+1}^{t},E_{n}^{t},U_{n}^{t}). 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 FudT:22F_{udT^{*}}:{\mathbb{R}}^{2}\to{\mathbb{R}}^{2} and its inverse, which are given explicitly by

FudT(x,u)=(min{x,u},xu),FudT1(x,u)=(x+max{u,0},xmin{u,0}).F_{udT^{*}}(x,u)=\left(\min\{x,u\},x-u\right),\quad F_{udT^{*}}^{-1}(x,u)=\left(x+\max\{u,0\},x-\min\{u,0\}\right).

Note that the decomposition of FudTF_{udT} into FudTF_{udT^{*}} and FudT1F_{udT^{*}}^{-1} is not unique. The form of FudTF_{udT^{*}} 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 FudTF_{udT^{*}}, 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 FudTF_{udT}.

Proposition 4.1.

The following measures μ,ν,μ~,ν~\mu,\nu,\tilde{\mu},\tilde{\nu} satisfy FudT(μ×ν)=μ~×ν~F_{udT^{*}}(\mu\times\nu)=\tilde{\mu}\times\tilde{\nu}.

  1. (a)

    For any λ1,λ2>0\lambda_{1},\lambda_{2}>0 and cc\in{\mathbb{R}},

    μ=sExp(λ1,c),ν=sExp(λ2,c),μ~=sExp(λ1+λ2,c),ν~=AL(λ1,λ2).\mu=\mathrm{sExp}(\lambda_{1},c),\quad\nu=\mathrm{sExp}(\lambda_{2},c),\quad\tilde{\mu}=\mathrm{sExp}(\lambda_{1}+\lambda_{2},c),\quad\tilde{\nu}=\mathrm{AL}(\lambda_{1},\lambda_{2}).
  2. (b)

    For any θ1,θ2(0,1)\theta_{1},\theta_{2}\in(0,1), MM\in{\mathbb{Z}} and m(0,)m\in(0,\infty),

    μ=ssGeo(1θ1,M,m),ν=ssGeo(1θ2,M,m),\mu=\mathrm{ssGeo}(1-\theta_{1},M,m),\quad\nu=\mathrm{ssGeo}(1-\theta_{2},M,m),
    μ~=ssGeo(1θ1θ2,M,m),ν~=sdAL(1θ1,1θ2,m).\tilde{\mu}=\mathrm{ssGeo}(1-\theta_{1}\theta_{2},M,m),\quad\tilde{\nu}=\mathrm{sdAL}(1-\theta_{1},1-\theta_{2},m).
  3. (c)

    For any c1,c2c_{1},c_{2}\in\mathbb{R} and measure mm supported on [0,)[0,\infty),

    1. (i)

      μ=δc1\mu=\delta_{c_{1}}, ν=δc2\nu=\delta_{c_{2}}, μ~=δmin{c1,c2}\tilde{\mu}=\delta_{\min\{c_{1},c_{2}\}}, ν~=δc1c2\tilde{\nu}=\delta_{c_{1}-c_{2}},

    2. (ii)

      μ=δc1\mu=\delta_{c_{1}}, ν=m(c1)\nu=m(\cdot-c_{1}), μ~=δc1\tilde{\mu}=\delta_{c_{1}}, ν~=m()\tilde{\nu}=m(-\cdot),

    3. (iii)

      μ=m(c1)\mu=m(\cdot-c_{1}), ν=δc1\nu=\delta_{c_{1}}, μ~=δc1\tilde{\mu}=\delta_{c_{1}}, ν~=m\tilde{\nu}=m.

    NB. Case (c)(i) is contained in cases (c)(ii) and (c)(iii).

It is further the case that there are no other quadruples of probability measures (μ,ν,μ~,ν~)(\mu,\nu,\tilde{\mu},\tilde{\nu}) that satisfy FudT(μ×ν)=μ~×ν~F_{udT^{*}}(\mu\times\nu)=\tilde{\mu}\times\tilde{\nu}.

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 XX and YY are two non-constant, independent random variables. It is then the case that min{X,Y}\min\{X,Y\} and XYX-Y are independent if and only if XX and YY 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

FudT(a,b,c)=(FudT(1)(b,c),FudT1(a,FudT(2)(b,c))).F_{udT}(a,b,c)=\left(F_{udT^{*}}^{(1)}(b,c),F_{udT^{*}}^{-1}\left(a,F_{udT^{*}}^{(2)}(b,c)\right)\right).

This enables us to deduce from Propositions 2.8 and 4.1 the subsequent result.

Proposition 4.2.

The following product measures μ~×μ×ν\tilde{\mu}\times\mu\times\nu satisfy FudT(μ~×μ×ν)=μ~×μ×νF_{udT}(\tilde{\mu}\times\mu\times\nu)=\tilde{\mu}\times\mu\times\nu.

  1. (a)

    For any λ1,λ2>0\lambda_{1},\lambda_{2}>0 and cc\in{\mathbb{R}},

    μ~×μ×ν=sExp(λ1+λ2,c)×sExp(λ1,c)×sExp(λ2,c).\tilde{\mu}\times\mu\times\nu=\mathrm{sExp}(\lambda_{1}+\lambda_{2},c)\times\mathrm{sExp}(\lambda_{1},c)\times\mathrm{sExp}(\lambda_{2},c).
  2. (b)

    For any θ1,θ2(0,1)\theta_{1},\theta_{2}\in(0,1), MM\in{\mathbb{Z}} and m(0,)m\in(0,\infty),

    μ~×μ×ν=ssGeo(1θ1θ2,M,m)×ssGeo(1θ1,M,m)×ssGeo(1θ2,M,m).\tilde{\mu}\times\mu\times\nu=\mathrm{ssGeo}(1-\theta_{1}\theta_{2},M,m)\times\mathrm{ssGeo}(1-\theta_{1},M,m)\times\mathrm{ssGeo}(1-\theta_{2},M,m).
  3. (c)

    For any c1,c2c_{1},c_{2}\in\mathbb{R} and measure mm supported on [0,)[0,\infty),

    1. (i)

      μ~×μ×ν=δmin{c1,c2}×δc1×δc2\tilde{\mu}\times\mu\times\nu=\delta_{\min\{c_{1},c_{2}\}}\times\delta_{c_{1}}\times\delta_{c_{2}},

    2. (ii)

      μ~×μ×ν=δc1×δc1×m(c1)\tilde{\mu}\times\mu\times\nu=\delta_{c_{1}}\times\delta_{c_{1}}\times m(\cdot-c_{1}),

    3. (iii)

      μ~×μ×ν=δc1×m(c1)×δc1\tilde{\mu}\times\mu\times\nu=\delta_{c_{1}}\times m(\cdot-c_{1})\times\delta_{c_{1}}.

    NB. Again, case (c)(i) is contained in cases (c)(ii) and (c)(iii).

Moreover, if a product measure is invariant under FudTF_{udT}, 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 X,YX,Y and ZZ be independent random variables satisfying

FudT(X,Y,Z)=d(X,Y,Z).F_{udT}(X,Y,Z)\buildrel d\over{=}(X,Y,Z).

Let W:=FudT(2)(Y,Z)W:=F_{udT^{*}}^{(2)}(Y,Z), then, by assumption,

FudT1(X,W)=FudT1(X,FudT(2)(Y,Z))=d(Y,Z).F_{udT^{*}}^{-1}(X,W)=F_{udT^{*}}^{-1}\left(X,F_{udT^{*}}^{(2)}(Y,Z)\right)\buildrel d\over{=}(Y,Z).

Hence FudT(Y,Z)=d(X,W)F_{udT^{*}}(Y,Z)\buildrel d\over{=}(X,W). Since X,Y,ZX,Y,Z are independent, XX and WW are independent. Therefore the marginals of (Y,Z,X,W)(Y,Z,X,W) must be given by one of the collections (μ,ν,μ~,ν~)(\mu,\nu,\tilde{\mu},\tilde{\nu}) 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 (Q0,E0)(2)(Q^{0},E^{0})\in(\mathbb{R}^{2})^{\mathbb{Z}}, find (Qnt,Ent,Unt)n,t(Q^{t}_{n},E^{t}_{n},U^{t}_{n})_{n,t\in\mathbb{Z}} such that (udToda) holds for all n,tn,t. This was solved in [13] for initial conditions in the set

𝒳udT:={(Q,E)(2):limnm=1n(QmEm)n=limnm=1n(QmEm)+Qn+1n<0,limnm=1n(QmEm)n=limnm=1n(QmEm)+Enn<0}.\mathcal{X}_{udT}:=\left\{(Q,E)\in(\mathbb{R}^{2})^{\mathbb{Z}}\>:\>\begin{array}[]{l}\lim_{n\to\infty}\frac{\sum_{m=1}^{n}(Q_{m}-E_{m})}{n}=\lim_{n\to\infty}\frac{\sum_{m=1}^{n}(Q_{m}-E_{m})+Q_{n+1}}{n}<0,\\ \lim_{n\to-\infty}\frac{\sum_{m=1}^{n}(Q_{m}-E_{m})}{n}=\lim_{n\to-\infty}\frac{\sum_{m=1}^{n}(Q_{m}-E_{m})+E_{n}}{n}<0\end{array}\right\}.

In particular, the subsequent result was established.

Lemma 4.3 ([13, Theorem 2.3]).

If (Q0,E0)𝒳udT(Q^{0},E^{0})\in\mathcal{X}_{udT}, then there exists a unique collection (Qnt,Ent,Unt)n,t(Q^{t}_{n},E^{t}_{n},U^{t}_{n})_{n,t\in\mathbb{Z}} such that (udToda) holds for all n,tn,t.

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

𝒯udT(Q0,E0):=(Q1,E1).\mathcal{T}_{udT}(Q^{0},E^{0}):=(Q^{1},E^{1}).

In what is the main result of this section, we characterize invariant product measures for the resulting evolution.

Theorem 4.4.

Suppose that (Qn0,En0)n(Q^{0}_{n},E^{0}_{n})_{n\in\mathbb{Z}} is an i.i.d. sequence with marginal given by μ~×μ\tilde{\mu}\times{\mu}, where one of the following holds:

  1. (a)

    for some λ1,λ2>0\lambda_{1},\lambda_{2}>0 and cc\in{\mathbb{R}},

    μ~×μ=sExp(λ1+λ2,c)×sExp(λ1,c);\tilde{\mu}\times\mu=\mathrm{sExp}(\lambda_{1}+\lambda_{2},c)\times\mathrm{sExp}(\lambda_{1},c);
  2. (b)

    for some θ1,θ2(0,1)\theta_{1},\theta_{2}\in(0,1), MM\in{\mathbb{Z}} and m(0,)m\in(0,\infty),

    μ~×μ=ssGeo(1θ1θ2,M,m)×ssGeo(1θ1,M,m);\tilde{\mu}\times\mu=\mathrm{ssGeo}(1-\theta_{1}\theta_{2},M,m)\times\mathrm{ssGeo}(1-\theta_{1},M,m);
  3. (c)

    for some cc\in\mathbb{R} and measure mm supported on [c,)[c,\infty) with mδcm\neq\delta_{c},

    μ~×μ=δc×m(c).\tilde{\mu}\times\mu=\delta_{c}\times m(\cdot-c).

It is then the case that 𝒯udT(Q0,E0)=d(Q0,E0)\mathcal{T}_{udT}(Q^{0},E^{0})\buildrel d\over{=}(Q^{0},E^{0}). Moreover, there are no other non-trivial measures such that (Qn0,En0)n(Q^{0}_{n},E^{0}_{n})_{n\in\mathbb{Z}} is an i.i.d. sequence, with Qn0Q_{n}^{0} independent of En0E_{n}^{0}, and 𝒯udT(Q0,E0)=d(Q0,E0)\mathcal{T}_{udT}(Q^{0},E^{0})\buildrel d\over{=}(Q^{0},E^{0}).

Proof.

If (Qn0,En0)n(Q^{0}_{n},E^{0}_{n})_{n\in\mathbb{Z}} is an i.i.d. sequence with marginal μ~×μ\tilde{\mu}\times{\mu} of one of the given forms, then it is a simple application of the law of large numbers to check that, (μ~×μ)(\tilde{\mu}\times{\mu})^{\mathbb{Z}}-a.s., (Q0,E0)𝒳udT(Q^{0},E^{0})\in\mathcal{X}_{udT}. It readily follows from Lemma 4.3 that, (μ~×μ)(\tilde{\mu}\times{\mu})^{\mathbb{Z}}-a.s., the corresponding type II lattice equations have a unique solution with initial condition (xn)n(x_{n})_{n\in\mathbb{Z}}, where x2n:=En0x_{2n}:=E^{0}_{n} and x2n+1:=Qn+10x_{2n+1}:=Q^{0}_{n+1}. 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) {Int+1=Jnt+Unt,Jnt+1=In+1tJnt(Int+1)1,Un+1t=In+1tUnt(Int+1)1.\begin{cases}I_{n}^{t+1}=J_{n}^{t}+U_{n}^{t},\\ J_{n}^{t+1}={I_{n+1}^{t}J_{n}^{t}}(I_{n}^{t+1})^{-1},\\ U_{n+1}^{t}={I_{n+1}^{t}U_{n}^{t}}(I_{n}^{t+1})^{-1}.\end{cases}

Here, the variables (Int,Jnt,Unt)n,t(I_{n}^{t},J_{n}^{t},U_{n}^{t})_{n,t\in\mathbb{Z}} take values in (0,)(0,\infty), and we can summarise the above dynamics by (Int+1,Jnt+1,Un+1t)=FdT(In+1t,Jnt,Unt)(I_{n}^{t+1},J_{n}^{t+1},U_{n+1}^{t})=F_{dT}(I_{n+1}^{t},J_{n}^{t},U_{n}^{t}), where FdTF_{dT} is an involution on (0,)3(0,\infty)^{3}. Similarly to (4.1), in this case we have a lattice structure

Int+1\textstyle{I_{n}^{t+1}}Jnt+1\textstyle{J_{n}^{t+1}}Unt\textstyle{U_{n}^{t}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Un+1t,\textstyle{U_{n+1}^{t},}Jnt\textstyle{J_{n}^{t}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}In+1t\textstyle{I_{n+1}^{t}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

which can be decomposed into two maps, FdTF_{dT^{*}} and FdT1F_{dT^{*}}^{-1}, as follows:

FdT\textstyle{\boxed{F_{dT^{*}}\vphantom{F_{udT^{*}}^{-1}}}}b+c\textstyle{b+c}FdT1\textstyle{\boxed{F_{dT^{*}}^{-1}}}abb+c\textstyle{\frac{ab}{b+c}}c\textstyle{c\ignorespaces\ignorespaces\ignorespaces\ignorespaces}bb+c\textstyle{\frac{b}{b+c}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}acb+c,\textstyle{\frac{ac}{b+c},}b\textstyle{b\ignorespaces\ignorespaces\ignorespaces\ignorespaces}a\textstyle{a\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

where we generically take (a,b,c)=(In+1t,Jnt,Unt)(a,b,c)=(I_{n+1}^{t},J_{n}^{t},U_{n}^{t}). 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 FdT:(0,)2(0,)2F_{dT^{*}}:(0,\infty)^{2}\rightarrow(0,\infty)^{2} and its inverse, which are given explicitly by:

FdT(x,y)=(x+y,xx+y),FdT1(x,y)=(xy,x(1y)).F_{dT^{*}}(x,y)=\left(x+y,\frac{x}{x+y}\right),\qquad F_{dT^{*}}^{-1}(x,y)=\left(xy,x(1-y)\right).

As in the ultra-discrete case, we note that the decomposition of FdTF_{dT} into FdTF_{dT^{*}} and FdT1F_{dT^{*}}^{-1} is not unique, with the form of FdTF_{dT^{*}} 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 FdTF_{dT^{*}}, see Proposition 4.5. In the subsequent result, Proposition 4.6, we apply this to deduce a complete solution to the corresponding problem for FdTF_{dT}.

Proposition 4.5.

The following measures μ,ν,μ~,ν~\mu,\nu,\tilde{\mu},\tilde{\nu} satisfy FdT(μ×ν)=μ~×ν~F_{dT^{*}}(\mu\times\nu)=\tilde{\mu}\times\tilde{\nu}.

  1. (a)

    For any λ1,λ2>0\lambda_{1},\lambda_{2}>0 and c>0c>0,

    μ=Gam(λ1,c),ν=Gam(λ2,c),μ~=Gam(λ1+λ2,c),ν~=Be(λ1,λ2).\mu=\mathrm{Gam}(\lambda_{1},c),\quad\nu=\mathrm{Gam}(\lambda_{2},c),\quad\tilde{\mu}=\mathrm{Gam}(\lambda_{1}+\lambda_{2},c),\quad\tilde{\nu}=\mathrm{Be}(\lambda_{1},\lambda_{2}).
  2. (b)

    For any c1,c2(0,)c_{1},c_{2}\in(0,\infty), μ=δc1\mu=\delta_{c_{1}}, ν=δc2\nu=\delta_{c_{2}}, μ~=δc1+c2\tilde{\mu}=\delta_{c_{1}+c_{2}}, ν~=δc1/(c1+c2)\tilde{\nu}=\delta_{c_{1}/(c_{1}+c_{2})}.

It is further the case that there are no other quadruples of probability measures (μ,ν,μ~,ν~)(\mu,\nu,\tilde{\mu},\tilde{\nu}) that satisfy FdT(μ×ν)=μ~×ν~F_{dT^{*}}(\mu\times\nu)=\tilde{\mu}\times\tilde{\nu}.

Proof.

The first part follows by direct computation. The uniqueness relies on a well-known fact [30] about gamma distributions. Namely, suppose that XX and YY are two non-constant, independent, positive random variables. Then X+YX+Y and XX+Y\frac{X}{X+Y} are independent if and only if XX and YY are gamma-distributed random variables with the same scale parameter. Applying the fact that XX and 1/X1/X are independent if and only if XX 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

FdT(a,b,c):=(FdT(1)(b,c),FdT1(a,FdT(2)(b,c))).F_{dT}(a,b,c):=\left(F_{dT^{*}}^{(1)}(b,c),F_{dT^{*}}^{-1}\left(a,F_{dT^{*}}^{(2)}(b,c)\right)\right).

This enables us to deduce from Propositions 2.8 and 4.5 the following result.

Proposition 4.6.

The following product measures μ~×μ×ν\tilde{\mu}\times\mu\times\nu satisfy FdT(μ~×μ×ν)=μ~×μ×νF_{dT}(\tilde{\mu}\times\mu\times\nu)=\tilde{\mu}\times\mu\times\nu.

  1. (a)

    For any λ1,λ2>0\lambda_{1},\lambda_{2}>0 and c>0c>0,

    μ~×μ×ν=Gam(λ1+λ2,c)×Gam(λ1,c)×Gam(λ2,c).\tilde{\mu}\times\mu\times\nu=\mathrm{Gam}(\lambda_{1}+\lambda_{2},c)\times\mathrm{Gam}(\lambda_{1},c)\times\mathrm{Gam}(\lambda_{2},c).
  2. (b)

    For any c1,c2(0,)c_{1},c_{2}\in(0,\infty),

    μ~×μ×ν=δc1+c2×δc1×δc2.\tilde{\mu}\times\mu\times\nu=\delta_{c_{1}+c_{2}}\times\delta_{c_{1}}\times\delta_{c_{2}}.

Moreover, if a product measure is invariant under FdTF_{dT}, 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 (I0,J0)((0,)2)(I^{0},J^{0})\in((0,\infty)^{2})^{\mathbb{Z}}, find (Int,Jnt,Unt)n,t(I^{t}_{n},J^{t}_{n},U^{t}_{n})_{n,t\in\mathbb{Z}} such that (dToda) holds for all n,tn,t. This was solved in [13] for initial conditions in the set

𝒳dT:=\displaystyle\mathcal{X}_{dT}:=
{(I,J)((0,)2):limnm=1n(logJmlogIm)n=limnm=1n(logJmlogIm)logIn+1n<0,limnm=1n(logJmlogIm)n=limnm=1n(logJmlogIm)logJnn<0}.\displaystyle\left\{(I,J)\in((0,\infty)^{2})^{\mathbb{Z}}\>:\>\begin{array}[]{l}\lim_{n\to\infty}\frac{\sum_{m=1}^{n}(\log J_{m}-\log I_{m})}{n}=\lim_{n\to\infty}\frac{\sum_{m=1}^{n}(\log J_{m}-\log I_{m})-\log I_{n+1}}{n}<0,\\ \lim_{n\to-\infty}\frac{\sum_{m=1}^{n}(\log J_{m}-\log I_{m})}{n}=\lim_{n\to-\infty}\frac{\sum_{m=1}^{n}(\log J_{m}-\log I_{m})-\log J_{n}}{n}<0\end{array}\right\}.

In particular, the following result was established.

Lemma 4.7 ([13, Theorem 2.5]).

If it holds that (I0,J0)𝒳dT(I^{0},J^{0})\in\mathcal{X}_{dT}, then there exists a unique collection (Int,Jnt,Unt)n,t(I^{t}_{n},J^{t}_{n},U^{t}_{n})_{n,t\in\mathbb{Z}} such that (dToda) holds for all n,tn,t.

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

𝒯dT(I0,J0):=(I1,J1).\mathcal{T}_{dT}(I^{0},J^{0}):=(I^{1},J^{1}).

In what is the main result of this section, we characterize invariant product measures for the resulting evolution.

Theorem 4.8.

Suppose that (In0,Jn0)n(I^{0}_{n},J^{0}_{n})_{n\in\mathbb{Z}} is an i.i.d. sequence with marginal given by μ~×μ\tilde{\mu}\times{\mu}, where the following holds: for some λ1,λ2>0\lambda_{1},\lambda_{2}>0 and cc\in{\mathbb{R}},

μ~×μ=Gam(λ1+λ2,c)×Gam(λ1,c).\tilde{\mu}\times\mu=\mathrm{Gam}(\lambda_{1}+\lambda_{2},c)\times\mathrm{Gam}(\lambda_{1},c).

It is then the case that 𝒯dT(I0,J0)=d(I0,J0)\mathcal{T}_{dT}(I^{0},J^{0})\buildrel d\over{=}(I^{0},J^{0}). Moreover, there are no other non-trivial measures such that (In0,Jn0)n(I^{0}_{n},J^{0}_{n})_{n\in\mathbb{Z}} is an i.i.d. sequence, with In0I_{n}^{0} independent of Jn0J_{n}^{0}, and 𝒯dT(I0,J0)=d(I0,J0)\mathcal{T}_{dT}(I^{0},J^{0})\buildrel d\over{=}(I^{0},J^{0}).

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.

Discrete KdV (α,β)(\alpha,\beta): GIG(λ,cα,c)×\mathrm{GIG}(\lambda,c\alpha,c)\times

GIG(λ,cβ,c)\mathrm{GIG}(\lambda,c\beta,c)

Ultra-discretization: λ(ε)=λε\lambda(\varepsilon)=\lambda\varepsilon c(ε)=ec/εc(\varepsilon)=e^{c/\varepsilon} α(ε)=eJ/ε\alpha(\varepsilon)=e^{-J/\varepsilon} β(ε)=eK/ε\beta(\varepsilon)=e^{-K/\varepsilon}          

Discrete Toda:

Gam(λ1+λ2,c)×\mathrm{Gam}(\lambda_{1}+\lambda_{2},c)\times

Gam(λ1,c)×Gam(λ2,c)\mathrm{Gam}(\lambda_{1},c)\times\mathrm{Gam}(\lambda_{2},c)

Ultra-discretization: λ1(ε)=λ1ε\lambda_{1}(\varepsilon)=\lambda_{1}\varepsilon λ2(ε)=λ2ε\lambda_{2}(\varepsilon)=\lambda_{2}\varepsilon c(ε)=ec/εc(\varepsilon)=e^{c/\varepsilon} Self-convolution: β=0\beta=0, (λ,cα)(λ2,c)(\lambda,c\sqrt{\alpha})\leftrightarrow(\lambda_{2},c)

Ultra-discrete KdV (J,K)(J,K): stExp(λ,c,Jc)×stExp(λ,c,Kc)\mathrm{stExp}(\lambda,c,J-c)\times\mathrm{stExp}(\lambda,c,K-c)

Self-convolution: K=K=\infty, (λ,cJ2)(λ2,c)(\lambda,c-\frac{J}{2})\leftrightarrow(\lambda_{2},c)

Ultra-discrete Toda:

sExp(λ1+λ2,c)×\mathrm{sExp}(\lambda_{1}+\lambda_{2},c)\times

sExp(λ1,c)×sExp(λ2,c)\mathrm{sExp}(\lambda_{1},c)\times\mathrm{sExp}(\lambda_{2},c)

Figure 1. Links between some of the product invariant measures of Propositions 3.2, 3.9, 4.2 and 4.6, as discussed in Section 5. In particular, the two solid arrows are essentially given by the weak convergence statements of Proposition 5.1, see Remark 5.2. The two dashed arrows indicate how particular conditionings of the invariant measures for the Toda-type systems give rise to the invariant measures for the KdV-type systems, see Subsection 5.2 for details.

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

x=limε0εlogx(ε),u=limε0εlogu(ε),x=\lim_{\varepsilon\downarrow 0}\varepsilon\log x(\varepsilon),\qquad u=\lim_{\varepsilon\downarrow 0}\varepsilon\log u(\varepsilon),
J=limε0εlogα(ε),K=limε0εlogβ(ε),J=\lim_{\varepsilon\downarrow 0}-\varepsilon\log\alpha(\varepsilon),\qquad K=\lim_{\varepsilon\downarrow 0}-\varepsilon\log\beta(\varepsilon),

then

(5.1) limε0εlog(FdK(α(ε),β(ε)))(i)(x(ε),u(ε))=(FudK(J,K))(i)(x,u),i=1,2.\lim_{\varepsilon\downarrow 0}\varepsilon\log\left(F_{dK}^{(\alpha(\varepsilon),\beta(\varepsilon))}\right)^{(i)}(x(\varepsilon),u(\varepsilon))=\left(F_{udK}^{(J,K)}\right)^{(i)}(x,u),\qquad i=1,2.

Similarly, if

a=limε0εloga(ε),b=limε0εlogb(ε),c=limε0εlogc(ε),a=\lim_{\varepsilon\downarrow 0}-\varepsilon\log a(\varepsilon),\qquad b=\lim_{\varepsilon\downarrow 0}-\varepsilon\log b(\varepsilon),\qquad c=\lim_{\varepsilon\downarrow 0}-\varepsilon\log c(\varepsilon),

then

(5.2) limε0εlogFdT(i)(a(ε),b(ε),c(ε))=FudT(i)(a,b,c),i=1,2,3.\lim_{\varepsilon\downarrow 0}-\varepsilon\log F_{dT}^{(i)}(a(\varepsilon),b(\varepsilon),c(\varepsilon))=F_{udT}^{(i)}(a,b,c),\qquad i=1,2,3.

As a consequence of the following proposition, we have that making corresponding changes of parameters for certain invariant measures for FdK(α,β)F_{dK}^{(\alpha,\beta)} and FdTF_{dT} yields invariant measures for FudK(J,K)F_{udK}^{(J,K)} and FudTF_{udT} (see Remark 5.2).

Proposition 5.1.
  1. (a)

    Suppose that X(ε)GIG(ελ,c(ε)α(ε),c(ε))X(\varepsilon)\sim\mathrm{GIG}(\varepsilon\lambda,c(\varepsilon)\alpha(\varepsilon),c(\varepsilon)), where c(ε):=ec/εc(\varepsilon):=e^{c/\varepsilon} and α(ε):=eL/ε\alpha(\varepsilon):=e^{-L/\varepsilon}, for some L{}L\in\mathbb{R}\cup\{\infty\}, λ\lambda\in\mathbb{R} if L<L<\infty, λ>0\lambda>0 if L=L=\infty, and c<L/2c<L/2. It then holds that

    limε0εlogX(ε)=X\lim_{\varepsilon\downarrow 0}\varepsilon\log X(\varepsilon)=X

    in distribution, where XstExp(λ,c,Lc)X\sim\mathrm{stExp}(\lambda,c,L-c).

  2. (b)

    Suppose that X(ε)Gam(ελ,c(ε))X(\varepsilon)\sim\mathrm{Gam}(\varepsilon\lambda,c(\varepsilon)), where c(ε):=ec/εc(\varepsilon):=e^{c/\varepsilon}, for some λ>0\lambda>0 and cc\in\mathbb{R}. It then holds that

    limε0εlogX(ε)=X\lim_{\varepsilon\downarrow 0}-\varepsilon\log X(\varepsilon)=X

    in distribution, where XsExp(λ,c)X\sim\mathrm{sExp}(\lambda,c).

Proof.

(a) Write Y(ε):=εlogX(ε)Y(\varepsilon):=\varepsilon\log X(\varepsilon). By making a standard change of variables, we see that this has density proportional to

fε(y):=eλyexp(c(ε)α(ε)ey/εc(ε)ey/ε),y.f_{\varepsilon}(y):=e^{-\lambda y}\exp\left(-c(\varepsilon)\alpha(\varepsilon)e^{y/\varepsilon}-c(\varepsilon)e^{-y/\varepsilon}\right),\qquad y\in\mathbb{R}.

Observe that, for y(c,Lc)y\in(c,L-c), we have that fε(y)eλyf_{\varepsilon}(y)\rightarrow e^{-\lambda y}. Hence, by the dominated convergence theorem, for any compact interval I[c,Lc]I\subseteq[c,L-c], we have

Ifε(y)𝑑yIeλy𝑑y.\int_{I}f_{\varepsilon}(y)dy\rightarrow\int_{I}e^{-\lambda y}dy.

Moreover, if L<L<\infty, the monotone convergence theorem yields that

Lcfε(y)𝑑yLceλyexp(e(yL+c)/ε)𝑑y0.\int_{L-c}^{\infty}f_{\varepsilon}(y)dy\leq\int_{L-c}^{\infty}e^{-\lambda y}\exp\left(-e^{(y-L+c)/\varepsilon}\right)dy\rightarrow 0.

Similarly,

cfε(y)𝑑yceλyexp(e(yc)/ε)𝑑y0.\int_{-\infty}^{c}f_{\varepsilon}(y)dy\leq\int_{-\infty}^{c}e^{-\lambda y}\exp\left(-e^{-(y-c)/\varepsilon}\right)dy\rightarrow 0.

Combining the previous three limits, the result readily follows.
(b) Writing Z(ε):=εlogX(ε)Z(\varepsilon):=-\varepsilon\log X(\varepsilon), we find that Z(ε)Z(\varepsilon) has density proportional to

gε(y):=eλyexp(c(ε)ey/ε),y.g_{\varepsilon}(y):=e^{-\lambda y}\exp\left(-c(\varepsilon)e^{-y/\varepsilon}\right),\qquad y\in\mathbb{R}.

Given this, arguing similarly to the first part of the proof gives the desired conclusion. ∎

Remark 5.2.

Applying Proposition 3.9(a), (5.1) and Proposition 5.1(a) gives another proof of Proposition 3.2(a)(i). Similarly, applying Proposition 4.6, (5.2) and Proposition 5.1(b) gives another proof of Proposition 4.2.

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 K=K=\infty), 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 FudTF_{udT} preserves the space {(a,b,c)3:a+b=0}\{(a,b,c)\in\mathbb{R}^{3}:\>a+b=0\}. In particular, we have that

FudT(a,a,b)=(min{a,b},min{a,b},bamin{a,b}).F_{udT}(-a,a,b)=\left(\min\{a,b\},-\min\{a,b\},b-a-\min\{a,b\}\right).

Combining the first two coordinates, we introduce an involution KudT:22K_{udT}:\mathbb{R}^{2}\rightarrow\mathbb{R}^{2} by setting KudT(a,b):=(FudT(2)(a,a,b),FudT(3)(a,a,b))K_{udT}(a,b):=(F_{udT}^{(2)}(-a,a,b),F_{udT}^{(3)}(-a,a,b)), or equivalently,

KudT(a,b)=(min{a,b},bamin{a,b}).K_{udT}(a,b)=\left(-\min\{a,b\},b-{a}-{\min\{a,b\}}\right).

Moreover, we note that this is simply a change of coordinates from FudK(J,)F_{udK}^{(J,\infty)}. Indeed, if A(J)(x,u):=(J2x,uJ2)A^{(J)}(x,u):=(\frac{J}{2}-x,u-\frac{J}{2}), then we have that

FudK(J,)=(A(J))1KudTA(J).F_{udK}^{(J,\infty)}=(A^{(J)})^{-1}\circ K_{udT}\circ A^{(J)}.

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 FudK(J,)F_{udK}^{(J,\infty)} to FudT(a,a,b)F_{udT}(-a,a,b) 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 stExp(λ,c,Jc)×sExp(λ,c)\mathrm{stExp}(\lambda,c,J-c)\times\mathrm{sExp}(\lambda,c) (with λ>0\lambda>0 and cJ/2c\leq J/2) for FudK(J,)F_{udK}^{(J,\infty)} of Proposition 3.2 corresponds to the following invariant measure for KudTK_{udT}:

(5.3) stExp(λ,cJ2,J2c)×sExp(λ,cJ2),\mathrm{stExp}\left(-{\lambda},c-\frac{J}{2},\frac{J}{2}-c\right)\times\mathrm{sExp}\left(\lambda,c-\frac{J}{2}\right),

Returning to the coordinates of the (udToda) system, this gives that if (A,B)(A,B) has the above distribution, then (A,A,B)(-A,A,B) is invariant for FudTF_{udT}. We note that this solution relates to the product invariant measure of Proposition 4.2. Indeed, it is readily checked that if (A,B,C)sExp(λ1+λ2,c)×sExp(λ1,c)×sExp(λ2,c)(A,B,C)\sim\mathrm{sExp}(\lambda_{1}+\lambda_{2},c)\times\mathrm{sExp}(\lambda_{1},c)\times\mathrm{sExp}(\lambda_{2},c) with λ1,λ2>0\lambda_{1},\lambda_{2}>0 and c<0c<0, then

(5.4) (A,B,C){|A+B|ε}(A~,A~,B~),\left(A,B,C\right)\>\vline\>\left\{|A+B|\leq\varepsilon\right\}\>\rightarrow\left(-\tilde{A},\tilde{A},\tilde{B}\right),

in distribution as ε0\varepsilon\rightarrow 0, where (A~,B~)stExp(λ2,c,c)×sExp(λ2,c)(\tilde{A},\tilde{B})\sim\mathrm{stExp}(-\lambda_{2},c,-c)\times\mathrm{sExp}(\lambda_{2},c). Since it holds that a+b=FudT(1)(a,b,c)+FudT(2)(a,b,c)a+b=F^{(1)}_{udT}(a,b,c)+F^{(2)}_{udT}(a,b,c), the left-hand side of (5.4) has a distribution that is invariant under FudTF_{udT}, and thus the continuous mapping theorem implies that so does the right-hand side. Reparameterising gives that (A~,B~)(\tilde{A},\tilde{B}) has distribution as at (5.3), which establishes that, in the case K=K=\infty, 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, FdTF_{dT} preserves the space {(a,b,c)(0,)3:ab=1}\{(a,b,c)\in(0,\infty)^{3}:\>ab=1\}, with

FdT(a1,a,b)=(a+b,(a+b)1,ba(a+b)).F_{dT}\left(a^{-1},a,b\right)=\left(a+b,(a+b)^{-1},\frac{b}{a(a+b)}\right).

In this case, we introduce an involution KdT:(0,)2(0,)2K_{dT}:(0,\infty)^{2}\rightarrow(0,\infty)^{2} by setting

KdT(a,b):=(FdT(2)(a,a,b),FdT(3)(a,a,b))=(1a+b,ba(a+b)),K_{dT}(a,b):=\left(F_{dT}^{(2)}(-a,a,b),F_{dT}^{(3)}(-a,a,b)\right)=\left(\frac{1}{a+b},\frac{b}{a(a+b)}\right),

and note that if A(α)(x,u):=(xα,1uα)A^{(\alpha)}(x,u):=(x\sqrt{\alpha},\frac{1}{u\sqrt{\alpha}}), then

FdK(α,0)=(A(α))1KdTA(α).F^{(\alpha,0)}_{dK}=(A^{(\alpha)})^{-1}\circ K_{dT}\circ A^{(\alpha)}.

Again, these operations essentially describe the self-convolution procedure of [13, Section 6], with the reverse procedure from FdK(α,0)F^{(\alpha,0)}_{dK} to FdT(a1,a,b)F_{dT}(a^{-1},a,b) involving the splitting procedure of [13, Section 6]. The invariant measure GIG(λ,cα,c)×IG(λ,c)\mathrm{GIG}(\lambda,c\alpha,c)\times\mathrm{IG}(\lambda,c) (with λ,c>0\lambda,c>0) for F(α,0)F^{(\alpha,0)} of Proposition 3.9 corresponds to the following invariant measure for KdTK_{dT}:

(5.5) GIG(λ,cα,cα)×Gam(λ,cα).\mathrm{GIG}\left(\lambda,c\sqrt{\alpha},c\sqrt{\alpha}\right)\times\mathrm{Gam}\left(\lambda,{c}{\sqrt{\alpha}}\right).

Hence, if (A,B)(A,B) has the above distribution, then the law of (A1,A,B)(A^{-1},A,B) is invariant under FdTF_{dT}. Moreover, it is possible to check that the solution relates to the product invariant measure of Proposition 4.6. For, if (A,B,C)Gam(λ1+λ2,c)×Gam(λ1,c)×Gam(λ2,c)(A,B,C)\sim\mathrm{Gam}(\lambda_{1}+\lambda_{2},c)\times\mathrm{Gam}(\lambda_{1},c)\times\mathrm{Gam}(\lambda_{2},c), then one may verify that

(5.6) (A,B,C){|AB1|ε}(A~1,A~,B~),\left(A,B,C\right)\>\vline\>\left\{|AB-1|\leq\varepsilon\right\}\>\rightarrow\left(\tilde{A}^{-1},\tilde{A},\tilde{B}\right),

in distribution as ε0\varepsilon\rightarrow 0, where (A~,B~)GIG(λ2,c,c)×Gam(λ2,c)(\tilde{A},\tilde{B})\sim\mathrm{GIG}(\lambda_{2},c,c)\times\mathrm{Gam}(\lambda_{2},c). Since it holds that ab=FdT(1)(a,b,c)FdT(2)(a,b,c)ab=F^{(1)}_{dT}(a,b,c)F^{(2)}_{dT}(a,b,c), the left-hand side of (5.6) has a distribution that is invariant under FdTF_{dT}, and thus the continuous mapping theorem implies that so does the right-hand side. Reparameterising gives that (A~,B~)(\tilde{A},\tilde{B}) has distribution as at (5.5), which establishes that, in the case β=0\beta=0, 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 (Xn0,U0t)n1,t0(X^{0}_{n},U^{t}_{0})_{n\geq 1,t\geq 0}, where the Xn0X^{0}_{n} are random variables taking values in a space 𝒳0\mathcal{X}_{0} and the U0tU^{t}_{0} are random variables taking values in a space 𝒰0\mathcal{U}_{0}, the random variables (Xnt,Unt)n1,t0(X^{t}_{n},U^{t}_{n})_{n\geq 1,t\geq 0} are defined recursively via the equations:

(6.1) (Xnt+1,Unt)=R(X~nt,Xnt,Un1t)\left(X_{n}^{t+1},U_{n}^{t}\right)=R\left(\tilde{X}_{n}^{t},X^{t}_{n},U^{t}_{n-1}\right)

where R:𝒳~0×𝒳0×𝒰0𝒳0×𝒰0R:\tilde{\mathcal{X}}_{0}\times\mathcal{X}_{0}\times\mathcal{U}_{0}\to\mathcal{X}_{0}\times\mathcal{U}_{0} is a deterministic function, and (X~nt)n1,t0(\tilde{X}_{n}^{t})_{n\geq 1,t\geq 0} are i.i.d. random variables, independent of (Xn0,U0t)n1,t0(X^{0}_{n},U^{t}_{0})_{n\geq 1,t\geq 0}. In particular, for a given realization of the variables (X~nt)n1,t0(\tilde{X}_{n}^{t})_{n\geq 1,t\geq 0}, we have a two-dimensional system of equations of the form described in the first sentence of the article with Fnt=R(X~nt,,)F_{n}^{t}=R(\tilde{X}_{n}^{t},\cdot,\cdot). 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 (Xnt,Unt)n1,t0(X^{t}_{n},U^{t}_{n})_{n\geq 1,t\geq 0} satisfies Burke’s property if:

  • •:

    the sequences (Xn0)n1(X_{n}^{0})_{n\geq 1} and (U0t)t0(U_{0}^{t})_{t\geq 0} are each i.i.d., and independent of each other;

  • •:

    the distribution of (Xnt,Unt)n1,t0(X^{t}_{n},U^{t}_{n})_{n\geq 1,t\geq 0} is translation invariant, that is, for any m,s+m,s\in{\mathbb{Z}}_{+},

    (Xm+ns+t,Um+ns+t)n1,t0=d(Xnt,Unt)n1,t0.\left(X^{s+t}_{m+n},U^{s+t}_{m+n}\right)_{n\geq 1,t\geq 0}\buildrel d\over{=}\left(X^{t}_{n},U^{t}_{n}\right)_{n\geq 1,t\geq 0}.

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 μ~,μ,ν\tilde{\mu},\mu,\nu are probability measures on 𝒳~0,𝒳0,𝒰0\tilde{\mathcal{X}}_{0},\mathcal{X}_{0},\mathcal{U}_{0} respectively satisfying

(6.2) R(μ~×μ×ν)=μ×ν.R(\tilde{\mu}\times\mu\times\nu)=\mu\times\nu.

If (Xn0)n1,(U0t)t0,{X~nt}n1,t0(X^{0}_{n})_{n\geq 1},(U^{t}_{0})_{t\geq 0},\{\tilde{X}^{t}_{n}\}_{n\geq 1,t\geq 0} are independent random variables whose marginals are μ,ν\mu,\nu and μ~\tilde{\mu} respectively, then the random variables (Xnt,Unt)n1,t0(X^{t}_{n},U^{t}_{n})_{n\geq 1,t\geq 0} 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 XntX^{t}_{n} are 𝒳n\mathcal{X}_{n}-valued random variables, UntU^{t}_{n} are 𝒰t\mathcal{U}_{t}-valued random variables, X~nt\tilde{X}^{t}_{n} are 𝒳~nt\tilde{\mathcal{X}}^{t}_{n} random variables, and there exists a sequence of deterministic functions

Rn,t:𝒳~nt×𝒳n×𝒰t𝒳n×𝒰t.R_{n,t}:\tilde{\mathcal{X}}^{t}_{n}\times\mathcal{X}_{n}\times\mathcal{U}_{t}\to\mathcal{X}_{n}\times\mathcal{U}_{t}.

Given a (random) boundary condition (Xn0,U0t)n1,t0(X^{0}_{n},U^{t}_{0})_{n\geq 1,t\geq 0}, we then define (Xnt,Unt)n1,t0(X^{t}_{n},U^{t}_{n})_{n\geq 1,t\geq 0} by

(6.3) (Xnt+1,Unt)=Rn,t(X~nt,Xnt,Un1t).(X_{n}^{t+1},U_{n}^{t})=R_{n,t}(\tilde{X}_{n}^{t},X^{t}_{n},U^{t}_{n-1}).

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 (Xnt,Unt)n1,t0(X^{t}_{n},U^{t}_{n})_{n\geq 1,t\geq 0} satisfies Burke’s property if there exist a sequence of probability measures (μn)n1(\mu_{n})_{n\geq 1}, with μn\mu_{n} supported on 𝒳n\mathcal{X}_{n}, and (νt)t0(\nu_{t})_{t\geq 0}, with νt\nu_{t} supported on 𝒰t\mathcal{U}_{t} such that:

  • •:

    XntμnX^{t}_{n}\sim\mu_{n} for all n1,t0n\geq 1,t\geq 0;

  • •:

    UntνtU^{t}_{n}\sim\nu_{t} for all n1,t0n\geq 1,t\geq 0;

  • •:

    for any m,s+m,s\in{\mathbb{Z}}_{+}, (Xm+ns)n1,(Ums+t)t0(X^{s}_{m+n})_{n\geq 1},(U^{s+t}_{m})_{t\geq 0} 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 μ~nt\tilde{\mu}^{t}_{n}, μn\mu_{n}, νt\nu_{t} are probability measures on 𝒳~nt\tilde{\mathcal{X}}_{n}^{t}, 𝒳n\mathcal{X}_{n}, 𝒰t\mathcal{U}_{t}, respectively, satisfying

Rn,t(μ~nt×μn×νt)=μn×νt.R_{n,t}(\tilde{\mu}^{t}_{n}\times\mu_{n}\times\nu_{t})=\mu_{n}\times\nu_{t}.

If (Xn0)n1(X^{0}_{n})_{n\geq 1}, (U0t)t0(U^{t}_{0})_{t\geq 0}, (X~nt)n1,t0(\tilde{X}^{t}_{n})_{n\geq 1,t\geq 0} are independent random variables whose marginals are μn\mu_{n}, νt\nu_{t} and μ~nt\tilde{\mu}^{t}_{n}, respectively, then the random variables (Xnt,Unt)n1,t0(X^{t}_{n},U^{t}_{n})_{n\geq 1,t\geq 0} 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 RR at (6.1) does not depend on X~nt\tilde{X}_{n}^{t}. More generally, one could apply Proposition 6.2 to study an inhomogeneous deterministic model. For example, if we set Fn,t=FudKJn,KtF_{n,t}=F^{J_{n},K_{t}}_{udK}, then we have that μn=stExp(λ,c,Jnc)\mu_{n}=\mathrm{stExp}(\lambda,c,J_{n}-c) and νt=stExp(λ,c,Ktc)\nu_{t}=\mathrm{stExp}(\lambda,c,K_{t}-c) satisfy

    FudKJn,Kt(μn×νt)=μn×νt,F^{J_{n},K_{t}}_{udK}\left(\mu_{n}\times\nu_{t}\right)=\mu_{n}\times\nu_{t},

    and so there is a distribution on (xnt,unt)n1,t0(x^{t}_{n},u^{t}_{n})_{n\geq 1,t\geq 0} that satisfies the inhomogeneous version of Burke’s property.

  • Secondly, to connect to type II models, we observe that the condition R(μ~×μ×ν)=μ×νR(\tilde{\mu}\times\mu\times\nu)=\mu\times\nu at (6.2) matches the condition on FF of Proposition 2.8(a). Hence, if RR is given by a map

    (6.4) R(a,b,c)=R1(a,R(2)(b,c)),R(a,b,c)=R_{*}^{-1}\left(a,R_{*}^{(2)}(b,c)\right),

    where R:𝒳0×𝒰0𝒳~0×𝒰~0R_{*}:\mathcal{X}_{0}\times\mathcal{U}_{0}\to\tilde{\mathcal{X}}_{0}\times\tilde{\mathcal{U}}_{0} is a bijection (i.e. similarly to (1.4) with R=F(2,3)R=F^{(2,3)} and R=FR_{*}=F_{*}), then the detailed balance condition for the type II model given by RR_{*} 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 F=RF_{*}=R_{*} 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 (xnt,unt)n,t(x^{t}_{n},u^{t}_{n})_{n,t\in{\mathbb{Z}}}, where xnt𝒳0x^{t}_{n}\in\mathcal{X}_{0}, unt𝒰~0u^{t}_{n}\in\tilde{\mathcal{U}}_{0} for n+t=0n+t=0 (mod 2), and xnt𝒳~0x^{t}_{n}\in\tilde{\mathcal{X}}_{0}, unt𝒰0u^{t}_{n}\in{\mathcal{U}}_{0} for n+t=1n+t=1 (mod 2), whereas, for the stochastic model, (Xnt,Unt)n1,t0(X^{t}_{n},U^{t}_{n})_{n\geq 1,t\geq 0} satisfies Xnt𝒳0X^{t}_{n}\in\mathcal{X}_{0} and Unt𝒰0U^{t}_{n}\in\mathcal{U}_{0} for all n,tn,t.

We next proceed to discuss a number of examples of stochastic integrable systems. In particular, we will observe that

RDLPP=FudT(2,3),RRPs=FdT(2,3),R_{DLPP}=F_{udT}^{(2,3)},\qquad R_{RPs}=F_{dT}^{(2,3)},

where RDLPPR_{DLPP} is the function RR for directed last passage percolation, and RRPsR_{RPs} is the function RR for the directed polymer with site weights (see Subsections 6.1 and 6.2, respectively). We will further see that RRPeR_{RPe}, the function RR for the directed polymer with edge weights, can also written in terms of a bijection R=RRPeR_{*}=R_{RPe^{*}}. 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 2{\mathbb{N}}^{2}, a key quantity of interest is the partition function

Zn,m=maxπ:(1,1)(n,m){(k,)πXk,},m,n,Z_{n,m}=\max_{\pi:(1,1)\to(n,m)}\left\{\sum_{(k,\ell)\in\pi}X_{k,\ell}\right\},\qquad m,n\in\mathbb{N},

where the maximum is taken over ‘up-right paths’ π:(1,1)(n,m)\pi:(1,1)\to(n,m) on 2{\mathbb{N}}^{2}, and (Xn,m)n,m(X_{n,m})_{n,m\in{\mathbb{N}}} are i.i.d. random variables. One readily sees that this partition function satisfies the following recursion:

(6.5) Zn,m=Xn,m+max{Zn1,m,Zn,m1}.Z_{n,m}=X_{n,m}+\max\left\{Z_{n-1,m},Z_{n,m-1}\right\}.

By setting Un,m:=Zn,mZn1,mU_{n,m}:=Z_{n,m}-Z_{n-1,m} and Vn,m:=Zn,mZn,m1V_{n,m}:=Z_{n,m}-Z_{n,m-1}, the recursive equation at (6.5) can be rewritten as

RDLPP(Xn,m,Un,m1,Vn1,m)=(Un,m,Vn,m),R_{DLPP}(X_{n,m},U_{n,m-1},V_{n-1,m})=(U_{n,m},V_{n,m}),

where

RDLPP(a,b,c)=(a+bmin{b,c},a+cmin{b,c}).R_{DLPP}\left(a,b,c\right)=\left(a+b-\min\{b,c\},a+c-\min\{b,c\}\right).

In particular, RDLPP=FudT(2,3)R_{DLPP}=F_{udT}^{(2,3)}, and we obtain from Proposition 2.8 that

RDLPP(μ~×μ×ν)=μ×νFudT(μ×ν)=μ~×ν~ for some ν.R_{DLPP}(\tilde{\mu}\times\mu\times\nu)=\mu\times\nu\qquad\Leftrightarrow\qquad F_{udT^{*}}(\mu\times\nu)=\tilde{\mu}\times\tilde{\nu}\mbox{ for some }\nu.

Apart from trivial solutions, we have from Proposition 4.1 that the above identities imply that μ~\tilde{\mu} is a (possibly scaled and shifted) exponential/geometric distribution; note that when (Xn,m)n,m(X_{n,m})_{n,m\in\mathbb{N}} 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 FudTF_{udT^{*}} of Proposition 4.1 further yields the existence of the stationary distribution Un,mExp(λ1)U_{n,m}\sim\mathrm{Exp}(\lambda_{1}), Vn,mExp(λ2)V_{n,m}\sim\mathrm{Exp}(\lambda_{2}), Xn,mExp(λ1+λ2)X_{n,m}\sim\mathrm{Exp}(\lambda_{1}+\lambda_{2}) 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

Zn,m=π:(1,1)(n,m){(k,)πXk,},m,n,Z_{n,m}=\sum_{\pi:(1,1)\to(n,m)}\left\{\prod_{(k,\ell)\in\pi}X_{k,\ell}\right\},\qquad m,n\in\mathbb{N},

where the sum is taken over ‘up-right paths’ π:(1,1)(n,m)\pi:(1,1)\to(n,m) on 2{\mathbb{N}}^{2}, and (Xn,m)n,m(X_{n,m})_{n,m\in{\mathbb{N}}} are i.i.d. random variables. In this case, we have a recursive equation for the partition function of the form

(6.6) Zn,m=Xn,m(Zn1,m+Zn,m1).Z_{n,m}=X_{n,m}\left(Z_{n-1,m}+Z_{n,m-1}\right).

Letting Un,m=Zn,m/Zn1,mU_{n,m}=Z_{n,m}/Z_{n-1,m}, Vn,m=Zn,m/Zn,m1V_{n,m}=Z_{n,m}/Z_{n,m-1}, the recursive equation (6.6) can be rewritten as

RRPs(Xn,m1,Un,m11,Vn1,m1)=(Un,m1,Vn,m1),R_{RPs}\left(X_{n,m}^{-1},U_{n,m-1}^{-1},V_{n-1,m}^{-1}\right)=\left(U_{n,m}^{-1},V_{n,m}^{-1}\right),

where

RRPs(a,b,c)=(abb+c,acb+c).R_{RPs}\left(a,b,c\right)=\left(\frac{ab}{b+c},\frac{ac}{b+c}\right).

We thus see that RRPs=FdT(2,3)R_{RPs}=F_{dT}^{(2,3)}, and we obtain from Proposition 2.8 that

RRPs(μ~×μ×ν)=μ×νFdT(μ×ν)=μ~×ν~ for some ν.R_{RPs}(\tilde{\mu}\times\mu\times\nu)=\mu\times\nu\qquad\Leftrightarrow\qquad F_{dT^{*}}(\mu\times\nu)=\tilde{\mu}\times\tilde{\nu}\mbox{ for some }\nu.

From Proposition 4.5, we have that the only non-trivial solution to these equations has μ~\tilde{\mu} being a gamma distribution, and, similarly to the comment in the previous example, it is of note to observe that when (Xn,m)n,m(X_{n,m})_{n,m\in\mathbb{N}} 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 Un,m1Gam(λ1,c)U_{n,m}^{-1}\sim\mathrm{Gam}(\lambda_{1},c), Vn,m1Gam(λ2,c)V_{n,m}^{-1}\sim\mathrm{Gam}(\lambda_{2},c), Xn,m1Gam(λ1+λ2,c)X_{n,m}^{-1}\sim\mathrm{Gam}(\lambda_{1}+\lambda_{2},c), cf. [33, 34].

6.3. Directed random polymer with edge weights

Similarly to the previous subsection, the model we next consider has partition function

Zn,m=π:(0,0)(n,m){eiπYei},m,n,Z_{n,m}=\sum_{\pi:(0,0)\to(n,m)}\left\{\prod_{e_{i}\in\pi}Y_{e_{i}}\right\},\qquad m,n\in\mathbb{N},

where again the sum is taken over ‘up-right paths’ π:(1,1)(n,m)\pi:(1,1)\to(n,m) on 2{\mathbb{N}}^{2}, and

Yei:={Xk,l,if ei=((k1,),(k,)),h(Xk,),if ei=((k,1),(k,)),Y_{e_{i}}:=\left\{\begin{array}[]{ll}X_{k,l},&\hbox{if }e_{i}=((k-1,\ell),(k,\ell)),\\ h(X_{k,\ell}),&\hbox{if }e_{i}=((k,\ell-1),(k,\ell)),\end{array}\right.

where hh is a positive function on +{\mathbb{R}}_{+}, and (Xn,m)n,m(X_{n,m})_{n,m\in{\mathbb{N}}} are i.i.d. random variables. This partition function satisfies

(6.7) Zn,m=Xn,mZn1,m+h(Xn,m)Zn,m1,Z_{n,m}=X_{n,m}Z_{n-1,m}+h(X_{n,m})Z_{n,m-1},

and by letting Un,m:=Zn,m/Zn1,mU_{n,m}:=Z_{n,m}/Z_{n-1,m}, Vn,m:=Zn,m/Zn,m1V_{n,m}:=Z_{n,m}/Z_{n,m-1}, the recursive equation (6.7) can be rewritten as

RRPe(Xn,m,Un,m1,Vn1,m)=(Un,m,Vn,m),R_{RPe}\left(X_{n,m},U_{n,m-1},V_{n-1,m}\right)=\left(U_{n,m},V_{n,m}\right),

where

RRPe(a,b,c)=(a+h(a)bc,h(a)+acb).R_{RPe}\left(a,b,c\right)=\left(a+\frac{h(a)b}{c},h(a)+\frac{ac}{b}\right).

Note that, whilst in the previous example we wrote RRPsR_{RPs} in terms of (Xn,m1,Un,m11,Vn1,m1)(X_{n,m}^{-1},U_{n,m-1}^{-1},V_{n-1,m}^{-1}) in order to fit closely with the map FdTF_{dT^{*}}, here we write RRPeR_{RPe} in terms of (Xn,m,Un,m1,Vn1,m)(X_{n,m},U_{n,m-1},V_{n-1,m}) to better fit the discussion in [5]. In particular, in [5], up to technical conditions, the authors characterize distributions μ~\tilde{\mu}, μ\mu and ν\nu such that RRPe(μ~×μ×ν)=μ×νR_{RPe}(\tilde{\mu}\times\mu\times\nu)=\mu\times\nu. To expand on this, under the assumptions of [5], whenever bb and cc are in the support of μ×ν\mu\times\nu, the function Hs(a):=as+h(a)H_{s}(a):=as+h(a), where s=cbs=\frac{c}{b}, has an inverse function Hs1H_{s}^{-1} on the support of μ~\tilde{\mu}. It follows that the function

RRPe(x,u)=(Hux1(u),ux),R_{RPe^{*}}(x,u)=\left(H^{-1}_{\frac{u}{x}}(u),\frac{u}{x}\right),

is a bijection (on the support of μ×ν\mu\times\nu), with inverse function given by

RRPe1(x,u)=(1uHu(x),Hu(x))=(x+h(x)u,h(x)+xu),R_{RPe^{*}}^{-1}(x,u)=\left(\frac{1}{u}H_{u}(x),H_{u}(x)\right)=\left(x+\frac{h(x)}{u},h(x)+xu\right),

and putting these together yields

RRPe(a,b,c)=RRPe1(a,RRPe(2)(b,c)).R_{RPe}\left(a,b,c\right)=R_{RPe^{*}}^{-1}\left(a,R_{RPe^{*}}^{(2)}(b,c)\right).

Hence the condition RRPe(μ~×μ×ν)=μ×νR_{RPe}(\tilde{\mu}\times\mu\times\nu)=\mu\times\nu is equivalent to R,RPe(μ×ν)=μ~×ν~R_{*,RPe}(\mu\times\nu)=\tilde{\mu}\times\tilde{\nu} for some ν~\tilde{\nu}, and also to

FRPe(μ~×μ×ν)=μ~×μ×ν,F_{RPe}(\tilde{\mu}\times\mu\times\nu)=\tilde{\mu}\times\mu\times\nu,

where

FRPe(a,b,c)=(RRPe(1)(b,c),RRPe1(a,R,RPe(2)(b,c))).F_{RPe}(a,b,c)=\left(R_{RPe^{*}}^{(1)}(b,c),R_{RPe^{*}}^{-1}\left(a,R_{*,RPe}^{(2)}(b,c)\right)\right).

In [5], the authors show that Burke’s property holds for the directed random polymer with edge weights only if h(x)=Ax+Bh(x)=Ax+B for some A,BA,B\in{\mathbb{R}} such that max{A,B}>0\max\{A,B\}>0. NB. In this case, the above map is of the form

(aB)bAb+c\textstyle{\frac{(a-B)b}{Ab+c}}a(Ab+c)c+Bbc\textstyle{\frac{a(Ab+c)}{c}+\frac{Bb}{c}}c\textstyle{c\ignorespaces\ignorespaces\ignorespaces\ignorespaces}a(Ab+c)b+B.\textstyle{\frac{a(Ab+c)}{b}+B.}b\textstyle{b\ignorespaces\ignorespaces\ignorespaces\ignorespaces}a\textstyle{a\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

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 A=1A=1, B=0B=0, i.e. h(x)=xh(x)=x,

Un,mIG(λ1,c),Vn,mIG(λ2,c),Xn,mIG(λ1+λ2,c);U_{n,m}\sim\mathrm{IG}(\lambda_{1},c),\qquad V_{n,m}\sim\mathrm{IG}(\lambda_{2},c),\qquad X_{n,m}\sim\mathrm{IG}(\lambda_{1}+\lambda_{2},c);
Gamma:

For A=0A=0, B=1B=1, i.e. h(x)=1h(x)=1,

Un,mGam(λ1+λ2,c),Vn,mBe1(λ1,λ2),Xn,mGam(λ2,c);U_{n,m}\sim\mathrm{Gam}(\lambda_{1}+\lambda_{2},c),\qquad V_{n,m}\sim\mathrm{Be}^{-1}(\lambda_{1},\lambda_{2}),\qquad X_{n,m}\sim\mathrm{Gam}(\lambda_{2},c);
Beta:

For A=1A=-1, B=1B=1, i.e. h(x)=1xh(x)=1-x,

Un,mBe(λ1+λ2,λ3),Vn,m1Be(λ1,λ2),Xn,mBe(λ2,λ3);U_{n,m}\sim\mathrm{Be}(\lambda_{1}+\lambda_{2},\lambda_{3}),\qquad V_{n,m}^{-1}\sim\mathrm{Be}(\lambda_{1},\lambda_{2}),\qquad X_{n,m}\sim\mathrm{Be}(\lambda_{2},\lambda_{3});
Inverse beta:

For A=1A=1, B=1B=-1, i.e. h(x)=x1h(x)=x-1,

Un,m1Be(λ1,λ3),(Vn,m+1)1Be(λ2,λ1+λ3),Xn,m1Be(λ1+λ2,λ3).U_{n,m}^{-1}\sim\mathrm{Be}(\lambda_{1},\lambda_{3}),\qquad(V_{n,m}+1)^{-1}\sim\mathrm{Be}(\lambda_{2},\lambda_{1}+\lambda_{3}),\qquad X_{n,m}^{-1}\sim\mathrm{Be}(\lambda_{1}+\lambda_{2},\lambda_{3}).

To obtain the results in the cases h(x)=xh(x)=x and h(x)=1h(x)=1, 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 h(x)=xh(x)=x, then, up to inversion of the variables, the dynamics of RRPeR_{RPe} matches that of RRPsR_{RPs}.) In the cases h(x)=1xh(x)=1-x and h(x)=x1h(x)=x-1, a similar result for the beta distribution is used, see [35].

Remark 6.3.

The equation (6.7) with h(x)=1xh(x)=1-x 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 Be(α,β)\mathrm{Be}(\alpha,\beta) random variables (Bn,t)n,t0(B_{n,t})_{n\in\mathbb{Z},t\geq 0}, and conditional on this, the process (Yt)t0(Y_{t})_{t\geq 0} is the (discrete-time) Markov chain with transition probabilities given by

𝐏B(Yt+1=n+1Yt=n)=Bn,t=1𝐏B(Yt+1=n1Yt=n).\mathbf{P}^{B}\left(Y_{t+1}=n+1\>\vline\>Y_{t}=n\right)=B_{n,t}=1-\mathbf{P}^{B}\left(Y_{t+1}=n-1\>\vline\>Y_{t}=n\right).

It is readily checked that Z~(t,n):=PB(Ytt2n+2)\tilde{Z}(t,n):=P^{B}(Y_{t}\geq t-2n+2) satisfies

Z~(t,n)=Bt,nZ~(t1,n)+(1Bt,n)Z~(t1,n1).\tilde{Z}(t,n)=B_{t,n}\tilde{Z}(t-1,n)+(1-B_{t,n})\tilde{Z}(t-1,n-1).

Reparameterising by setting Zn,m:=Z~(n+m,m)Z_{n,m}:=\tilde{Z}(n+m,m), Xn,m:=Bn+m,mX_{n,m}:=B_{n+m,m}, 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 𝒳0:={0,1,2,,}\mathcal{X}_{0}:=\{0,1,2,\dots,\}, and 𝒰0:={0,1,,J}\mathcal{U}_{0}:=\{0,1,\dots,J\} for some JJ\in{\mathbb{N}}. In the case J=1J=1, the dynamics of the model are given by the probabilities

𝐏((Xnt+1,Unt)=(i,0)(Xnt,Un1t=(i,0))\displaystyle\mathbf{P}\left((X^{t+1}_{n},U^{t}_{n})=(i,0)\>\vline\>(X^{t}_{n},U^{t}_{n-1}=(i,0)\right) =1+αqi1+α=:ci,0,\displaystyle=\frac{1+\alpha q^{i}}{1+\alpha}=:c_{i,0},
𝐏((Xnt+1,Unt)=(i1,1)(Xnt,Un1t=(i,0))\displaystyle\mathbf{P}\left((X^{t+1}_{n},U^{t}_{n})=(i-1,1)\>\vline\>(X^{t}_{n},U^{t}_{n-1}=(i,0)\right) =α(1qi)1+α,\displaystyle=\frac{\alpha(1-q^{i})}{1+\alpha},
𝐏((Xnt+1,Unt)=(i+1,0)(Xnt,Un1t=(i,1))\displaystyle\mathbf{P}\left((X^{t+1}_{n},U^{t}_{n})=(i+1,0)\>\vline\>(X^{t}_{n},U^{t}_{n-1}=(i,1)\right) =1νqi1+α=:ci,1,\displaystyle=\frac{1-\nu q^{i}}{1+\alpha}=:c_{i,1},
𝐏((Xnt+1,Unt)=(i,1)(Xnt,Un1t=(i,1))\displaystyle\mathbf{P}\left((X^{t+1}_{n},U^{t}_{n})=(i,1)\>\vline\>(X^{t}_{n},U^{t}_{n-1}=(i,1)\right) =α+νqi1+α,\displaystyle=\frac{\alpha+\nu q^{i}}{1+\alpha},

for an appropriate choice of α,ν,q\alpha,\nu,q, see [7] for details. For simplicity, we consider the case α0\alpha\geq 0 and ν,q[0,1)\nu,q\in[0,1). If

RHSVα,ν,q(u,i,j):=(i+j𝟏{uci,j},𝟏{uci,j}),R_{HSV}^{\alpha,\nu,q}\left(u,i,j\right):=\left(i+j-\mathbf{1}_{\{u\geq c_{i,j}\}},\mathbf{1}_{\{u\geq c_{i,j}\}}\right),

and (X~nt)n1,t0(\tilde{X}_{n}^{t})_{n\geq 1,t\geq 0} is an i.i.d. collection of uniform random variables on (0,1)(0,1), we then have that

(Xnt+1,Unt)=RHSVα,ν,q(X~nt,Xnt,Un1t).\left(X^{t+1}_{n},U^{t}_{n}\right)=R_{HSV}^{\alpha,\nu,q}\left(\tilde{X}_{n}^{t},X^{t}_{n},U^{t}_{n-1}\right).

By direct computation, one can check that

RHSVα,ν,q(Uni(0,1)×qNB(ν,pα)×qNB(q1,qp))=qNB(ν,pα)×qNB(q1,qp)R_{HSV}^{\alpha,\nu,q}\left(\mathrm{Uni}(0,1)\times\mathrm{qNB}\left(\nu,\frac{p}{\alpha}\right)\times\mathrm{qNB}\left(q^{-1},-qp\right)\right)=\mathrm{qNB}\left(\nu,\frac{p}{\alpha}\right)\times\mathrm{qNB}\left(q^{-1},-qp\right)

for any 0pα0\leq p\leq\alpha, where Uni(0,1)\mathrm{Uni}(0,1) is the uniform distribution on (0,1)(0,1), and qNB is a qq-negative binomial distribution (see the appendix for details). Note in particular that qNB(q1,qp)\mathrm{qNB}(q^{-1},-qp) is a Bernoulli distribution with parameter p1+p\frac{p}{1+p}. In [19], the authors introduce a change of parameters from (ν,α,p)(\nu,\alpha,p) to (s,ξ,u,v)(s,\xi,u,v) with 0s<10\leq s<1, ξ>0\xi>0, u<0u<0, 0v<sξ0\leq v<s\xi, so that α=sξu\alpha=-s\xi u, ν=s2\nu=s^{2}, p=uvp=-uv. With this, we have

RHSVsξu,s2,q(Uni(0,1)×qNB(s2,vsξ)×qNB(q1,quv))=qNB(s2,vsξ)×qNB(q1,quv).R_{HSV}^{-s\xi u,s^{2},q}\left(\mathrm{Uni}(0,1)\times\mathrm{qNB}\left(s^{2},\frac{v}{s\xi}\right)\times\mathrm{qNB}\left(q^{-1},quv\right)\right)=\mathrm{qNB}\left(s^{2},\frac{v}{s\xi}\right)\times q\mathrm{NB}\left(q^{-1},quv\right).

Moreover, in [19], the parameters (s,ξ,u)(s,\xi,u) are allowed to be inhomogeneous, so that s=sns=s_{n}, ξ=ξn\xi=\xi_{n} and u=utu=u_{t}. To align with this framework, we set Rn,t=RHSVsnξnut,sn2,qR_{n,t}=R_{HSV}^{-s_{n}\xi_{n}u_{t},s_{n}^{2},q}. It then follows that, for any fixed 0v<infnsnξn0\leq v<\inf_{n}{s_{n}\xi_{n}},

Rn,t(Uni(0,1)×μn×νt)=μn×νt,R_{n,t}\left(\mathrm{Uni}(0,1)\times\mu_{n}\times\nu_{t}\right)=\mu_{n}\times\nu_{t},

where μn=qNB(sn2,vsnξn)\mu_{n}=\mathrm{qNB}(s_{n}^{2},\frac{v}{s_{n}\xi_{n}}), νt=qNB(q1,qutv)\nu_{t}=\mathrm{qNB}(q^{-1},qu_{t}v).

For more general JJ\in{\mathbb{N}}, the model is defined by a fusion operation, see [7]. This gives the stochastic matrix

𝐏((Xnt+1,Unt)=(i,j)(Xnt,Un1t=(i,j))=𝟏{i+j=i+j}pi,j\mathbf{P}\left((X^{t+1}_{n},U^{t}_{n})=(i^{\prime},j^{\prime})\>\vline\>(X^{t}_{n},U^{t}_{n-1}=(i,j)\right)=\mathbf{1}_{\{i^{\prime}+j^{\prime}=i+j\}}p_{i^{\prime},j^{\prime}}

for i,i{0,1,2,,}i,i^{\prime}\in\{0,1,2,\dots,\}, j,j{0,1,,J}j,j^{\prime}\in\{0,1,\dots,J\}, and so there exists RHSVJ,α,ν,qR_{HSV}^{J,\alpha,\nu,q} such that

(Xnt+1,Unt)=RHSVJ,α,ν,q(X~nt,Xnt,Un1t)(X^{t+1}_{n},U^{t}_{n})=R_{HSV}^{J,\alpha,\nu,q}\left(\tilde{X}_{n}^{t},X^{t}_{n},U^{t}_{n-1}\right)

with (X~nt)n1,t0(\tilde{X}_{n}^{t})_{n\geq 1,t\geq 0} i.i.d. uniform random variables on (0,1)(0,1). Noting that a random variable XqNB(qJ,qJp)X\sim q\mathrm{NB}(q^{-J},-q^{J}p) can be written as X=Y1+Y2+YJX=Y_{1}+Y_{2}+\dots Y_{J}, where YiqNB(q1,qip)=Ber(qi1p1+p)Y_{i}\sim q\mathrm{NB}(q^{-1},-q^{i}p)=\mathrm{Ber}(\frac{q^{i-1}p}{1+p}) (see Proposition 2.3 of [19]), the fusion procedure gives that

RHSVJ,α,ν,q(Uni(0,1)×qNB(ν,pα)×qNB(qJ,qJp))=qNB(ν,pα)×qNB(qJ,qJp).R_{HSV}^{J,\alpha,\nu,q}\left(\mathrm{Uni}(0,1)\times\mathrm{qNB}\left(\nu,\frac{p}{\alpha}\right)\times\mathrm{qNB}\left(q^{-J},-q^{J}p\right)\right)=\mathrm{qNB}\left(\nu,\frac{p}{\alpha}\right)\times\mathrm{qNB}\left(q^{-J},-q^{J}p\right).

In [19], the inhomogeneous version was also studied in the same way as above. Namely, for Rn,t=RHSVJ,snξnut,sn2,qR_{n,t}=R_{HSV}^{J,-s_{n}\xi_{n}u_{t},s_{n}^{2},q} for n,t2n,t\in{\mathbb{N}}^{2}, for any fixed 0v<infnsnξn0\leq v<\inf_{n}{s_{n}\xi_{n}},

Rn,t(Uni(0,1)×μn×νt)=μn×νt,R_{n,t}\left(\mathrm{Uni}(0,1)\times\mu_{n}\times\nu_{t}\right)=\mu_{n}\times\nu_{t},

where μn=qNB(sn2,vsnξn)\mu_{n}=\mathrm{qNB}(s_{n}^{2},\frac{v}{s_{n}\xi_{n}}), νt=qNB(qJ,qJutv)\nu_{t}=\mathrm{qNB}(q^{-J},q^{J}u_{t}v). Hence Proposition 6.3 applies.

As a final remark, we note that the role of the distribution of X~\tilde{X} and the function RR 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 RR reflects the structure of the model, or more precisely the recursion equation of the partition function, independent of the distribution of X~\tilde{X}. On the other hand, for the higher spin vertex model, the function RR and the distribution of X~\tilde{X} do not have any meaning in themselves, but rather the pair together determines the stochastic matrix from the input (Xnt,Un1t)(X_{n}^{t},U^{t}_{n-1}) to (Xnt+1,Unt)(X_{n}^{t+1},U^{t}_{n}), 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 SS be a topological space equipped with its Borel σ\sigma-algebra, (Θ,)(\Theta,\mathcal{F}) be a measurable space, {fθ:θΘ}\{f_{\theta}:\>\theta\in\Theta\} be a collection of continuous maps fθ:SSf_{\theta}:S\to S, and μ\mu be a probability measure on Θ\Theta. Let (θn)n(\theta_{n})_{n\in{\mathbb{Z}}} be an i.i.d. sequence with marginal μ\mu. The object of interest is the Markov chain XnX_{n} constructed by iterating random functions on the state space SS, that is

Xn:=fθn(Xn1)=(fθnfθn1fθ2fθ1)(X0),X_{n}:=f_{\theta_{n}}\left(X_{n-1}\right)=\left(f_{\theta_{n}}\circ f_{\theta_{n-1}}\circ f_{\theta_{2}}\dots f_{\theta_{1}}\right)(X_{0}),

where X0=sX_{0}=s for some sSs\in S. Diaconis and Freedman showed that when ‘(fθ)θΘ(f_{\theta})_{\theta\in\Theta} is contracting on average’ (see [14] for a precise definition), XnX_{n} has a unique stationary distribution, which is independent of ss. We highlight that a key ingredient in the proof of this theorem is the proposition that the backward iteration defined by

Yn:=(fθ1fθ2fθ3fθn)(s)Y_{n}:=\left(f_{\theta_{1}}\circ f_{\theta_{2}}\circ f_{\theta_{3}}\dots f_{\theta_{n}}\right)(s)

converges almost surely, at an exponential rate, to a random variable YY_{\infty} that does not depend on ss (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 F:𝒳0×𝒰0𝒳0×𝒰0F:\mathcal{X}_{0}\times\mathcal{U}_{0}\to\mathcal{X}_{0}\times\mathcal{U}_{0}, and that, for a given (xn)n𝒳0(x_{n})_{n\in\mathbb{Z}}\in\mathcal{X}_{0}^{\mathbb{Z}}, we are interested in the existence and uniqueness of (un)n𝒰0(u_{n})_{n\in{\mathbb{Z}}}\in\mathcal{U}_{0}^{\mathbb{Z}} such that

un=F(2)(xn,un1).u_{n}=F^{(2)}(x_{n},u_{n-1}).

(Cf. (1.2).) Letting S:=𝒰0S:=\mathcal{U}_{0}, Θ:=𝒳0\Theta:=\mathcal{X}_{0} and fθ:=F(2)(θ,)f_{\theta}:=F^{(2)}(\theta,\cdot) for θΘ\theta\in\Theta, it is clear that if (xn)n(x_{n})_{n\in\mathbb{Z}} is an i.i.d. sequence with marginal μ\mu and we are given uNu_{N}, then (un)nN(u_{n})_{n\geq N} is the Markov chain constructed by the iterating random functions fxnf_{x_{n}}. If we know that the backward iteration YnY_{n} converges almost surely to a limit which does not depend on ss, then for any nn\in{\mathbb{Z}}, the limit

Zn:=limm(fθn+1fθn+2fθm)(s)Z_{n}:=\lim_{m\to\infty}(f_{\theta_{n+1}}\circ f_{\theta_{n+2}}\dots f_{\theta_{m}})(s)

also exists almost surely and does not depend on ss (cf. [25, Section 4]). In particular, ZnZ_{n} is measurable with respect to (θm)m1+n(\theta_{m})_{m\geq 1+n}, and Zn=fθn+1(Zn+1)Z_{n}=f_{\theta_{n+1}}(Z_{n+1}) for all nn. Setting xn:=θ1nx_{n}:=\theta_{1-n} and un:=Znu_{n}:=Z_{-n}, it follows that

un=fxn(un1),u_{n}=f_{x_{n}}(u_{n-1}),

and unu_{n} is measurable with respect to (xm)mn(x_{m})_{m\leq n}. In conclusion, for μ\mu^{{\mathbb{Z}}}-a.e. realization of (xn)n(x_{n})_{n\in\mathbb{Z}}, there exists at least one (un)n(u_{n})_{n\in{\mathbb{Z}}} satisfying un=F(2)(xn,un1)u_{n}=F^{(2)}(x_{n},u_{n-1}) and unu_{n} is measurable with respect to (xm)mn(x_{m})_{m\leq n}. Moreover, the distribution of (un)n(u_{n})_{n\in{\mathbb{Z}}} is translation invariant, being given by the stationary distribution for the Markov chain constructed by the iterated random functions fxnf_{x_{n}}.

For type II models, the story is similar. In this case we have a bijection F:𝒳0×𝒰0𝒳~0×𝒰~0F_{*}:\mathcal{X}_{0}\times\mathcal{U}_{0}\to\tilde{\mathcal{X}}_{0}\times\tilde{\mathcal{U}}_{0}, and taking S=𝒰0S=\mathcal{U}_{0}, Θ=𝒳0×𝒳~0\Theta=\mathcal{X}_{0}\times\tilde{\mathcal{X}}_{0}, and fθf_{\theta} as

fx,x~(s)=(F1)(2)(x~,F(2)(x,s)),f_{x,\tilde{x}}(s)=(F^{-1}_{*})^{(2)}\left(\tilde{x},F_{*}^{(2)}(x,s)\right),

we can repeat the discussion of the preceding paragraph. For the ultra-discrete Toda model in particular, we have that

fb,audT(c)=a+max{cb,0},f^{udT}_{b,a}(c)=a+\max\{c-b,0\},

which can be analysed in the same way as the G/G/1G/G/1 queue considered in [14, Section 4]. More specifically, in the latter example, the map of interest is given by

fθG/G/1(s)=max{x+θ,0}.f^{G/G/1}_{\theta}(s)=\max\{x+\theta,0\}.

Although this is not a strict contraction, it is nonetheless shown in [14] that, under a certain condition on the distribution μ\mu, which includes the case when θ𝑑μ<0\int\theta d\mu<0, the backward iteration converges almost surely to a limit which does not depend on ss. To transfer the argument to the ultra-discrete Toda case, we first make the elementary observation that

fE1,Q0udTfE2,Q1udTfEn,Qn+1udT(s)=Q0+fθ1G/G/1fθ2G/G/1fθn1G/G/1(sEn),f^{udT}_{E_{-1},Q_{0}}\circ f^{udT}_{E_{-2},Q_{-1}}\circ\cdots\circ f^{udT}_{E_{-n},Q_{-n+1}}(s)=Q_{0}+f^{G/G/1}_{\theta_{1}}{\circ}f^{G/G/1}_{\theta_{2}}\circ\cdots\circ f^{G/G/1}_{\theta_{n-1}}(s-E_{-n}),

where θi:=QiEi\theta_{i}:=Q_{-i}-E_{-i}. 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

Q0+max{0,θ1,θ1+θ2,,θ1+θ2++θn1,θ1+θ2++θn1+sEn}.Q_{0}+\max\left\{0,\theta_{1},\theta_{1}+\theta_{2},\dots,\theta_{1}+\theta_{2}+\dots+\theta_{n-1},\theta_{1}+\theta_{2}+\dots+\theta_{n-1}+s-E_{-n}\right\}.

It readily follows that if 𝐄(QnEn)<0\mathbf{E}(Q_{n}-E_{n})<0 (cf. the requirement on configurations in 𝒳udT\mathcal{X}_{udT} in Section 4), then this backward iteration converges almost-surely, for any ss, to the finite random variable Q0+max{0,θ1,θ1+θ2,}Q_{0}+\max\{0,\theta_{1},\theta_{1}+\theta_{2},\dots\}. As is shown in [13, Theorem 2.3], this precisely corresponds to the value of U00U^{0}_{0} given by the unique solution to the initial value problem for (udToda) with initial condition (Qn,En)n(Q_{n},E_{n})_{n\in\mathbb{Z}}. One can similarly reconstruct (Un0)n(U^{0}_{n})_{n\in\mathbb{Z}}, 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 FudTF_{udT}, with the variables (Qn,En,Un,𝒯udTQn,𝒯udTEn)(Q_{n},E_{n},U_{n},\mathcal{T}_{udT}Q_{n},\mathcal{T}_{udT}E_{n}) in our notation being the analogues of (sn,an,wn+sn,rn,dn)(s_{n},a_{n},w_{n}+s_{n},r_{n},d_{n}) 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 (𝒯(x)n)n(\mathcal{T}(x)_{n})_{n\in\mathbb{Z}}, as defined by 𝒯(x)n:=F(1)(xn,un1)\mathcal{T}(x)_{n}:=F^{(1)}(x_{n},u_{n-1}), has the same distribution as (xn)n(x_{n})_{n\in\mathbb{Z}}, where unu_{n} is defined by the backward iteration? On this issue, in [25], it is shown that when:

  1. (i)

    the Markov chain (un)n(u_{n})_{n\in\mathbb{Z}} has reversible transition probabilities,

  2. (ii)

    for each sSs\in S, the map θ(s,fθ(s))\theta\mapsto(s,f_{\theta}(s)) is injective,

if we set

𝒯~(x)n=ϕ(un,un1),\tilde{\mathcal{T}}(x)_{n}=\phi(u_{n},u_{n-1}),

where ϕ\phi is the inverse function of θ(s,fθ(s))\theta\mapsto(s,f_{\theta}(s)), then (𝒯~(x)n)n(\tilde{\mathcal{T}}(x)_{n})_{n\in\mathbb{Z}} is an i.i.d. sequence with marginal μ\mu. NB. It is straightforward to check that, for a type I model and a measure μ\mu on 𝒳0\mathcal{X}_{0} such that μ(𝒳)=1\mu^{\mathbb{Z}}(\mathcal{X}^{*})=1, we almost-surely have that 𝒯=𝒯~\mathcal{T}=\tilde{\mathcal{T}}. It is moreover shown in [25, Theorem 4.1] that u0u_{0} is independent of 𝒯(x)0,𝒯(x)1,{\mathcal{T}}(x)_{0},{\mathcal{T}}(x)_{-1},\dots, which yields that (u0t)t(u^{t}_{0})_{t\in{\mathbb{Z}}} is an i.i.d. sequence, where (unt)n,t(u^{t}_{n})_{n\in\mathbb{Z},t\in\mathbb{Z}} is defined recursively. In addition, if (xn)n(x_{n})_{n\in{\mathbb{Z}}} and (u0t)t(u^{t}_{0})_{t\in{\mathbb{Z}}} are one-to-one almost surely, then the dynamics given by 𝒯{\mathcal{T}} are ergodic (actually Bernoulli) with respect to μ\mu^{{\mathbb{Z}}} (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(1,)(1,\infty) 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 𝒯\mathcal{T} 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 F:Θ×SΘ×SF:\Theta\times S\to\Theta\times S (at least, on the support of appropriate measures) that is an extension of the map (θ,s)fθ(s)(\theta,s)\mapsto f_{\theta}(s). More precisely, if we assume that θ(s,fθ(s))\theta\to(s,f_{\theta}(s)) is injective for each sSs\in S, and that the set {(s,fθ(s))S2:θΘ}S2\{(s,f_{\theta}(s))\in S^{2}:\>\theta\in\Theta\}\subseteq S^{2} is symmetric in the two coordinates, then such an F:Θ×SΘ×SF:\Theta\times S\to\Theta\times S is given by F(θ,s):=(ϕ(fθ(s),s),fθ(s))F(\theta,s):=(\phi(f_{\theta}(s),s),f_{\theta}(s)). Note that, even if there exists such an extension, however, we can not expect that (fθ)θΘ(f_{\theta})_{\theta\in\Theta} 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 θ1,θ2,,θn\theta_{1},\theta_{2},\dots,\theta_{n} such that the image of fθ1fθ2fθnf_{\theta_{1}}\circ f_{\theta_{2}}\circ\cdots\circ f_{\theta_{n}} contains exactly one point. If such a string occurs infinitely often under the measure μ\mu^{\mathbb{Z}}, 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 xn+xn+1x_{n}+x_{n+1} given in (3.6) and (3.7) being ‘synchronizing’ for the ultra-discrete KdV system with J>KJ>K.

Finally, we further note that there has also been a series of works on the stochastic equation:

ηk=ξkηk1,k,\eta_{k}=\xi_{k}\eta_{k-1},\qquad k\in\mathbb{Z},

where (ξk)k(\xi_{k})_{k\in\mathbb{Z}} is the ‘evolution process’, and (ηk)k(\eta_{k})_{k\in\mathbb{Z}} is an unknown process, with both taking values in a compact group GG (see the survey [39] and the references therein). It is clear that this model is in the setting of iterated random functions with Θ=S=G\Theta=S=G and fθ(s)=θsf_{\theta}(s)=\theta s. Moreover, it is obvious that in this case there exists an involution F:Θ×SΘ×SF:\Theta\times S\to\Theta\times S such that F(2)(θ,s)=fθ(s)F^{(2)}(\theta,s)=f_{\theta}(s), as given by F(θ,s)=(θ1,θs)F(\theta,s)=(\theta^{-1},\theta s). 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 θn\theta_{n}, the Markov chain might or might not have a unique stationary distributional solution or a strong solution (i.e. for which ηk\eta_{k} is measurable with respect to (ξm)mk(\xi_{m})_{m\leq k}), 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.

Completely characterize the detailed balance solutions for (udKdV), i.e. remove the technical conditions from Proposition 3.3.

Problem 8.2.

Completely characterize the detailed balance solutions for (dKdV), i.e. extend the final claim of Proposition 3.9 to general α,β0\alpha,\beta\geq 0 (see Conjecture 8.6 below for our expectation in this direction). Moreover, describe a reasonable subset of 𝒳\mathcal{X}^{*} when αβ>0\alpha\beta>0, 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 XX and YY are independent, then so are UU and VV, where (U,V)=F(X,Y)(U,V)=F(X,Y) for a given FF, if and only if the distribution of (X,Y)(X,Y) falls into a certain class. Perhaps the most famous result in this direction is that first proved by Kac in 1939: ‘if XX and YY are independent, then so are X+YX+Y and XYX-Y if and only if both XX and YY 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 FF, 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 F:(0,)2(0,)2F:(0,\infty)^{2}\rightarrow(0,\infty)^{2} be the involution given by

F(a,b)=(1a+b,1a1a+b).F(a,b)=\left(\frac{1}{a+b},\frac{1}{a}-\frac{1}{a+b}\right).

Let XX and YY be non-trivial (0,)(0,\infty)-valued independent random variables. It is then the case that (U,V):=F(X,Y)(U,V):=F(X,Y) are independent if and only if there exist λ,c1,c2>0\lambda,c_{1},c_{2}>0 such that

XGIG(λ,c1,c2),YGam(λ,c1),X\sim\mathrm{GIG}(\lambda,c_{1},c_{2}),\qquad Y\sim\mathrm{Gam}(\lambda,c_{1}),

and in this case, UGIG(λ,c2,c1)U\sim\mathrm{GIG}(\lambda,c_{2},c_{1}) and VGam(λ,c2)V\sim\mathrm{Gam}(\lambda,c_{2}). Hence, if moreover (U,V)(U,V) has the same distribution as (X,Y)(X,Y), then XGIG(λ,c,c)X\sim\mathrm{GIG}(\lambda,c,c) and YGam(λ,c)Y\sim\mathrm{Gam}(\lambda,c) for some λ,c>0\lambda,c>0.

As a direct corollary, by making the change of variables (a,b)(a,b1)(a,b)\to(a,b^{-1}), one can check a similar result for the involution F:(0,)2(0,)2F:(0,\infty)^{2}\rightarrow(0,\infty)^{2} given by

(8.1) F(a,b)=(b1+ab,a(1+ab)).F(a,b)=\left(\frac{b}{1+ab},a(1+ab)\right).

In this case, the random variables XX and UU have the same distribution as in Theorem 8.5, but YIG(λ,c1)Y\sim\mathrm{IG}(\lambda,c_{1}) and VIG(λ,c2)V\sim\mathrm{IG}(\lambda,c_{2}). Now, the above map is precisely FdK(1,0)F_{dK}^{(1,0)}, 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 FdK(α,β)F_{dK}^{(\alpha,\beta)} with αβ=0\alpha\beta=0. In light of the conclusion of Proposition 3.9, we conjecture that for general α,β0\alpha,\beta\geq 0, a similar result holds.

Conjecture 8.6.

Let α,β0\alpha,\beta\geq 0 with αβ\alpha\neq\beta, and recall the definition of FdK(α,β)F_{dK}^{(\alpha,\beta)} from (dKdV). Let XX and YY be non-trivial (0,)(0,\infty)-valued independent random variables. It is then the case that (U,V):=FdK(α,β)(X,Y)(U,V):=F_{dK}^{(\alpha,\beta)}(X,Y) are independent if and only if there exist λ,c1,c2>0\lambda,c_{1},c_{2}>0 such that

XGIG(λ,c1α,c2),YGIG(λ,c2β,c1),X\sim\mathrm{GIG}(\lambda,c_{1}\alpha,c_{2}),\qquad Y\sim\mathrm{GIG}(\lambda,c_{2}\beta,c_{1}),

and in this case UGIG(λ,c2α,c1)U\sim\mathrm{GIG}(\lambda,c_{2}\alpha,c_{1}) and VGIG(λ,c1β,c2)V\sim\mathrm{GIG}(\lambda,c_{1}\beta,c_{2}). Hence, if moreover (U,V)(U,V) has the same distribution as (X,Y)(X,Y), then XGIG(λ,cα,c)X\sim\mathrm{GIG}(\lambda,c\alpha,c), YGIG(λ,cβ,c)Y\sim\mathrm{GIG}(\lambda,c\beta,c) for some λ,c>0\lambda,c>0.

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 F:(0,)2(0,)×(0,1)F:(0,\infty)^{2}\rightarrow(0,\infty)\times(0,1) be the bijection given by

F(a,b)=(a+b,aa+b).F(a,b)=\left({a+b},\frac{a}{a+b}\right).

NB. F1(a,b)=(ab,a(1b))F^{-1}(a,b)=(ab,a(1-b)). Let XX and YY be non-trivial (0,)(0,\infty)-valued independent random variables. It is then the case that (U,V):=F(X,Y)(U,V):=F(X,Y) are independent if and only if there exist λ,c1,c2>0\lambda,c_{1},c_{2}>0 such that

XGam(λ1,c),YGam(λ2,c),X\sim\mathrm{Gam}(\lambda_{1},c),\qquad Y\sim\mathrm{Gam}(\lambda_{2},c),

and in this case, UGam(λ1+λ2,c)U\sim\mathrm{Gam}(\lambda_{1}+\lambda_{2},c) and VBe(λ1,λ2)V\sim\mathrm{Be}(\lambda_{1},\lambda_{2}).

Corollary 8.8.

Let F:(0,)×(0,1)(0,)2F:(0,\infty)\times(0,1)\rightarrow(0,\infty)^{2} be the bijection given by

F(a,b)=(ab,a(1b)).F(a,b)=\left(ab,a(1-b)\right).

Let XX and YY be non-trivial (0,)(0,\infty)-valued and (0,1)(0,1)-valued, respectively, independent random variables. It is then the case that (U,V):=F(X,Y)(U,V):=F(X,Y) are independent if and only if there exist λ,c1,c2>0\lambda,c_{1},c_{2}>0 such that

XGam(λ1+λ2,c),YBe(λ1,λ2),X\sim\mathrm{Gam}(\lambda_{1}+\lambda_{2},c),\qquad Y\sim\mathrm{Be}(\lambda_{1},\lambda_{2}),

and in this case, UGam(λ1,c)U\sim\mathrm{Gam}(\lambda_{1},c) and VGam(λ2,c)V\sim\mathrm{Gam}(\lambda_{2},c).

Theorem 8.9 ([35]).

Let F:(0,1)2(0,1)2F:(0,1)^{2}\rightarrow(0,1)^{2} be the involution given by

F(a,b)=(1b1ab,1ab).F(a,b)=\left(\frac{1-b}{1-ab},1-ab\right).

Let XX and YY be non-trivial (0,1)(0,1)-valued independent random variables. It is then the case that (U,V):=F(X,Y)(U,V):=F(X,Y) are independent if and only if there exist p,q,r>0p,q,r>0 such that

XBe(p,q),YBe(p+q,r),X\sim\mathrm{Be}(p,q),\qquad Y\sim\mathrm{Be}(p+q,r),

and in this case, UBe(r,q)U\sim\mathrm{Be}(r,q) and VBe(q+r,p)V\sim\mathrm{Be}(q+r,p). Hence, if moreover (U,V)(U,V) has the same distribution as (X,Y)(X,Y), then XBe(p,q)X\sim\mathrm{Be}(p,q), YBe(p+q,p)Y\sim\mathrm{Be}(p+q,p).

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 (0,1)(0,1) to (0,)(0,\infty) via the bijection x1x11x\mapsto\frac{1}{x^{-1}-1} (the inverse of which is x11+x1x\mapsto\frac{1}{1+x^{-1}}). The ultra-discretization procedure is then given by applying the limit

F(a,b)limε0(ιεlogF(1)(eιaε1,eιbε1),ιεlogF(2)(eιaε1,eιbε1)),F(a,b)\mapsto\lim_{\varepsilon\to 0}\left(\iota\varepsilon\log F^{(1)}\left(e^{\iota a\varepsilon^{-1}},e^{\iota b\varepsilon^{-1}}\right),\iota\varepsilon\log F^{(2)}\left(e^{\iota a\varepsilon^{-1}},e^{\iota b\varepsilon^{-1}}\right)\right),

where we take ι=1\iota=1 for Conjectures 8.10 and 8.11, and ι=1\iota=-1 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 F(a,b)=FudK(0,)(a,b)F(a,b)=F^{(0,\infty)}_{udK}(a,b), then F:22F:{\mathbb{R}}^{2}\to{\mathbb{R}}^{2} is an involution. For any c>0c>0, F:[c,c]×[c,)[c,c]×[c,)F:[-c,c]\times[-c,\infty)\to[-c,c]\times[-c,\infty) is an involution, and for any c1,c2>0c_{1},c_{2}>0, F:[c1,c2]×[c2,)[c2,c1]×[c1,)F:[-c_{1},c_{2}]\times[-c_{2},\infty)\to[-c_{2},c_{1}]\times[-c_{1},\infty) is a bijection. Let XX and YY be absolutely continuous \mathbb{R}-valued independent random variables satisfying P(X>0)P(X<0)0P(X>0)P(X<0)\neq 0. It is then the case that (U,V):=F(X,Y)(U,V):=F(X,Y) are independent if and only if there exist λ,c1,c2>0\lambda,c_{1},c_{2}>0 such that

XstExp(λ,c1,c2),YsExp(λ,c2),X\sim\mathrm{stExp}(\lambda,-c_{1},c_{2}),\qquad Y\sim\mathrm{sExp}(\lambda,-c_{2}),

and in this case, UstExp(λ,c2,c1)U\sim\mathrm{stExp}(\lambda,-c_{2},c_{1}), VsExp(λ,c1)V\sim\mathrm{sExp}(\lambda,-c_{1}). Hence, if moreover (U,V)(U,V) has the same distribution as (X,Y)(X,Y), then XstExp(λ,c,c)X\sim\mathrm{stExp}(\lambda,-c,c), YsExp(λ,c)Y\sim\mathrm{sExp}(\lambda,-c) for some c>0c>0.

Conjecture 8.11.

If F(a,b)=FudK(J,K)(a,b)F(a,b)=F^{(J,K)}_{udK}(a,b) for some <J,K<-\infty<J,K<\infty, then F:22F:{\mathbb{R}}^{2}\to{\mathbb{R}}^{2} is an involution. Also, for any c<min{J2,K2}c<\min\{\frac{J}{2},\frac{K}{2}\}, F:[c,Jc]×[c,Kc][c,Jc]×[c,Kc]F:[c,J-c]\times[c,K-c]\to[c,J-c]\times[c,K-c] is an involution, and for any c1,c2<min{J2,K2}c_{1},c_{2}<\min\{\frac{J}{2},\frac{K}{2}\}, F:[c1,Jc2]×[c2,Kc1][c2,Jc1]×[c1,Kc2]F:[c_{1},J-c_{2}]\times[c_{2},K-c_{1}]\to[c_{2},J-c_{1}]\times[c_{1},K-c_{2}] is a bijection. Let XX and YY be absolutely continuous \mathbb{R}-valued independent random variables satisfying P(X>J2)P(X<J2)P(Y>K2)P(Y<K2)0P(X>\frac{J}{2})P(X<\frac{J}{2})P(Y>\frac{K}{2})P(Y<\frac{K}{2})\neq 0. It is then the case that (U,V):=F(X,Y)(U,V):=F(X,Y) are independent if and only if there exist λ>0\lambda>0 and c1,c2<min{J2,K2}c_{1},c_{2}<\min\{\frac{J}{2},\frac{K}{2}\} such that

XstExp(λ,c1,Jc2),YstExp(λ,c2,Kc1),X\sim\mathrm{stExp}(\lambda,c_{1},J-c_{2}),\qquad Y\sim\mathrm{stExp}(\lambda,c_{2},K-c_{1}),

and in this case, UstExp(λ,c2,Jc1)U\sim\mathrm{stExp}(\lambda,c_{2},J-c_{1}), VstExp(λ,c1,Kc2)V\sim\mathrm{stExp}(\lambda,c_{1},K-c_{2}). Hence, if moreover (U,V)(U,V) has the same distribution as (X,Y)(X,Y), then XstExp(λ,c,Jc)X\sim\mathrm{stExp}(\lambda,c,J-c), YstExp(λ,c,Kc)Y\sim\mathrm{stExp}(\lambda,c,K-c) for some c<min{J2,K2}c<\min\{\frac{J}{2},\frac{K}{2}\}.

Remark 8.12.

It is also possible to write down discrete versions of the previous two conjectures, replacing the stExp\mathrm{stExp} distribution with the sstbGeo\mathrm{sstbGeo} 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 sdAL\mathrm{sdAL} distribution.

Theorem 8.13 ([8]).

Let F:22F:\mathbb{R}^{2}\rightarrow\mathbb{R}^{2} be the bijection given by

F(a,b)=(min{a,b},ab).F(a,b)=\left(\min\{a,b\},a-b\right).

NB. F1(a,b)=(a+max{b,0},amin{b,0})F^{-1}(a,b)=(a+\max\{b,0\},a-\min\{b,0\}). Let XX and YY be non-trivial \mathbb{R}-valued independent random variables. It is then the case that (U,V):=F(X,Y)(U,V):=F(X,Y) are independent if and only if there exist λ1,λ2,c>0\lambda_{1},\lambda_{2},c>0 such that

XsExp(λ1,c),YsExp(λ2,c)X\sim\mathrm{sExp}(\lambda_{1},c),\qquad Y\sim\mathrm{sExp}(\lambda_{2},c)

or θ1,θ2(0,1),m>0,M\theta_{1},\theta_{2}\in(0,1),m>0,M\in{\mathbb{Z}} such that

XssGeo(1θ1,M,m),YssGeo(1θ2,M,m),X\sim\mathrm{ssGeo}(1-\theta_{1},M,m),\qquad Y\sim\mathrm{ssGeo}(1-\theta_{2},M,m),

and in this case UstExp(λ1+λ2,c)U\sim\mathrm{stExp}(\lambda_{1}+\lambda_{2},c), VAL(λ1,λ2)V\sim\mathrm{AL}(\lambda_{1},\lambda_{2}), or UssGeo(1θ1θ2,M,m)U\sim\mathrm{ssGeo}(1-\theta_{1}\theta_{2},M,m), VsdAL(1θ1,1θ2,m)V\sim\mathrm{sdAL}(1-\theta_{1},1-\theta_{2},m), respectively.

Corollary 8.14.

Let F:22F:\mathbb{R}^{2}\rightarrow\mathbb{R}^{2} be the bijection given by

F(a,b)=(a+max{b,0},amin{b,0}).F(a,b)=\left(a+\max\{b,0\},a-\min\{b,0\}\right).

Let XX and YY be non-trivial \mathbb{R}-valued independent random variables. It is then the case that (U,V):=F(X,Y)(U,V):=F(X,Y) are independent if and only if there exist λ1,λ2,c>0\lambda_{1},\lambda_{2},c>0 such that

XsExp(λ1+λ2,c),YAL(λ1,λ2),X\sim\mathrm{sExp(\lambda_{1}+\lambda_{2},c)},\qquad Y\sim\mathrm{AL}(\lambda_{1},\lambda_{2}),

or θ1,θ2(0,1),m>0,M\theta_{1},\theta_{2}\in(0,1),m>0,M\in{\mathbb{Z}} such that

XssGeo(1θ1θ2,M,m),YsdAL(1θ1,1θ2,m),X\sim\mathrm{ssGeo}(1-\theta_{1}\theta_{2},M,m),\qquad Y\sim\mathrm{sdAL}(1-\theta_{1},1-\theta_{2},m),

and in this case UsExp(λ1,c)U\sim\mathrm{sExp}(\lambda_{1},c), VsExp(λ2,c)V\sim\mathrm{sExp}(\lambda_{2},c), or UssGeo(1θ1,M,m)U\sim\mathrm{ssGeo}(1-\theta_{1},M,m), VssGeo(1θ2,M,m)V\sim\mathrm{ssGeo}(1-\theta_{2},M,m), respectively.

Conjecture 8.15.

Let F:22F:\mathbb{R}^{2}\rightarrow\mathbb{R}^{2} be the involution given by

F(a,b)=(min{a,0}b,min{a,b,0}ab).F(a,b)=\left(\min\{a,0\}-b,\min\{a,b,0\}-a-b\right).

Let XX and YY be absolutely continuous \mathbb{R}-valued independent random variables. It is then the case that (U,V):=F(X,Y)(U,V):=F(X,Y) are independent if and only if there exist p,q,r>0p,q,r>0 such that

XAL(p,q),YAL(p+q,r),X\sim\mathrm{AL}(p,q),\qquad Y\sim\mathrm{AL}(p+q,r),

and in this case, UAL(r,q)U\sim\mathrm{AL}(r,q), VAL(q+r,p)V\sim\mathrm{AL}(q+r,p). Hence, if moreover (U,V)(U,V) has the same distribution as (X,Y)(X,Y), then XAL(p,q)X\sim\mathrm{AL}(p,q), YAL(p+q,p)Y\sim\mathrm{AL}(p+q,p).

Remark 8.16.

Since this article was completed, some of the above conjectures have been addressed in [2]. In particular, under technical conditions, Theorems 1.1, 1.2 and 1.3 of [2] confirm Conjectures 8.6, 8.15 and 8.10, respectively. It remains to check discrete versions of the latter two claims.

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 λ,c1,c2\lambda,c_{1},c_{2}\in\mathbb{R} with c1<c2c_{1}<c_{2}, the shifted truncated exponential distribution with parameters (λ,c1,c2)(\lambda,c_{1},c_{2}), which we denote stExp(λ,c1,c2)\mathrm{stExp}(\lambda,c_{1},c_{2}), has density

1Zeλx𝟏[c1,c2](x),x,\frac{1}{Z}e^{-\lambda x}\mathbf{1}_{[c_{1},c_{2}]}(x),\qquad x\in\mathbb{R},

where ZZ is a normalizing constant.

Shifted exponential distribution:

For λ>0\lambda>0, cc\in{\mathbb{R}}, the shifted exponential distribution with parameters (λ,c)(\lambda,c), which we denote sExp(λ,c)\mathrm{sExp}(\lambda,c), has density

1Zeλx𝟏[c,)(x),x,\frac{1}{Z}e^{-\lambda x}\mathbf{1}_{[c,\infty)}(x),\qquad x\in{\mathbb{R}},

where ZZ is a normalizing constant. We use the convention that stExp(λ,c,)=sExp(λ,c)\mathrm{stExp}(\lambda,c,\infty)=\mathrm{sExp}(\lambda,c) when λ>0\lambda>0.

Shifted scaled (truncated bipartite) geometric distribution:

For θ>0\theta>0, MM\in\mathbb{Z}, N{}N\in\mathbb{Z}\cup\{\infty\} such that MNM\leq N, κ>0\kappa>0 and m(0,)m\in(0,\infty), we say a random variable XX has shifted scaled truncated bipartite geometric distribution with parameters 1θ1-\theta, MM, NN, κ\kappa and mm if

𝐏(X=mx)=1Zθxκι(x),x{M,M+1,,N},\mathbf{P}\left(X=mx\right)=\frac{1}{Z}\theta^{x}\kappa^{\iota(x)},\qquad x\in\{M,M+1,\dots,N\},

where ι(2x)=0,ι(2x+1)=1\iota(2x)=0,\iota(2x+1)=1 and ZZ is a normalising constant; in this case we write XsstbGeo(1θ,M,N,κ,m)X\sim\mathrm{sstbGeo}(1-\theta,M,N,\kappa,m). Note that, if N=N=\infty, then we require that θ<1\theta<1 for the distribution to be defined. We observe that sstbGeo(1θ,0,N,1,1)\mathrm{sstbGeo}(1-\theta,0,N,1,1) is simply the distribution of the usual parameter 1θ1-\theta geometric distribution conditioned to take a value in {0,1,,N}\{0,1,\dots,N\}. In the special case when θ<1\theta<1, N=N=\infty, κ=1\kappa=1, we say that XX has shifted scaled geometric distribution with parameters 1θ1-\theta, MM and mm, and write XssGeo(1θ,M,m)X\sim\mathrm{ssGeo}(1-\theta,M,m).

Asymmetric Laplace distribution:

For λ1,λ2(0,)\lambda_{1},\lambda_{2}\in(0,\infty), the asymmetric Laplace distribution with parameters (λ1,λ2)(\lambda_{1},\lambda_{2}), which we denote AL(λ1,λ2)\mathrm{AL}(\lambda_{1},\lambda_{2}), has density

1Z(eλ1x𝟏(0,)(x)+eλ2x𝟏(,0)(x)),x,\frac{1}{Z}\left(e^{-\lambda_{1}x}\mathbf{1}_{(0,\infty)}(x)+e^{\lambda_{2}x}\mathbf{1}_{(-\infty,0)}(x)\right),\qquad x\in{\mathbb{R}},

where ZZ is a normalizing constant.

Scaled discrete asymmetric Laplace distribution:

For θ1,θ2(0,1)\theta_{1},\theta_{2}\in(0,1) and m(0,)m\in(0,\infty), we say a random variable XX has scaled discrete asymmetric Laplace distribution with parameters (1θ1,1θ2,m)(1-\theta_{1},1-\theta_{2},m) if

𝐏(X=mx)={1Zθ1x,x{0,1,2,},1Zθ2x,x{,2,1},\mathbf{P}\left(X=mx\right)=\left\{\begin{array}[]{ll}\frac{1}{Z}\theta_{1}^{x},&x\in\{0,1,2,\dots\},\\ \frac{1}{Z}\theta_{2}^{-x},&x\in\{\dots,-2,-1\},\end{array}\right.

where ZZ is a normalizing constant; in this case we write XsdAL(1θ1,1θ2,m)X\sim\mathrm{sdAL}(1-\theta_{1},1-\theta_{2},m).

Gamma distribution:

For λ,c(0,)\lambda,c\in(0,\infty), the gamma distribution with parameters (λ,c)(\lambda,c), which we denote Gam(λ,c)\mathrm{Gam}(\lambda,c), has density

1Zxλ1ecx𝟏(0,)(x),x,\frac{1}{Z}x^{\lambda-1}e^{-cx}\mathbf{1}_{(0,\infty)}(x),\qquad x\in\mathbb{R},

where ZZ is a normalizing constant.

Inverse gamma distribution:

For λ,c(0,)\lambda,c\in(0,\infty), the inverse gamma distribution with parameters (λ,c)(\lambda,c), which we denote IG(λ,c)\mathrm{IG}(\lambda,c), has density

1Zxλ1ecx1𝟏(0,)(x),x,\frac{1}{Z}x^{-\lambda-1}e^{-cx^{-1}}\mathbf{1}_{(0,\infty)}(x),\qquad x\in\mathbb{R},

where ZZ is a normalizing constant.

Generalized inverse Gaussian distribution:

For λ\lambda\in{\mathbb{R}}, c1,c2(0,)c_{1},c_{2}\in(0,\infty), the generalized inverse Gaussian distribution with parameters (λ,c1,c2)(\lambda,c_{1},c_{2}), which we denote GIG(λ,c1,c2)\mathrm{GIG}(\lambda,c_{1},c_{2}), has density

1Zxλ1ec1xc2x1𝟏(0,)(x),x,\frac{1}{Z}x^{-\lambda-1}e^{-c_{1}x-c_{2}x^{-1}}\mathbf{1}_{(0,\infty)}(x),\qquad x\in\mathbb{R},

where ZZ is a normalizing constant. We use the convention that GIG(λ,0,c)=IG(λ,c)\mathrm{GIG}(\lambda,0,c)=\mathrm{IG}(\lambda,c).

Beta distribution:

For λ1,λ2(0,)\lambda_{1},\lambda_{2}\in(0,\infty), the beta distribution with parameters (λ1,λ2)(\lambda_{1},\lambda_{2}), which we denote Be(λ1,λ2)\mathrm{Be}(\lambda_{1},\lambda_{2}), has density

1Zxλ11(1x)λ21𝟏(0,1)(x),x,\frac{1}{Z}x^{\lambda_{1}-1}(1-x)^{\lambda_{2}-1}\mathbf{1}_{(0,1)}(x),\qquad x\in\mathbb{R},

where ZZ is a normalizing constant.

qq-negative binomial distribution:

Fix q[0,1)q\in[0,1). For p,b[0,1)p,b\in[0,1) or p<0p<0, b=qLb=q^{-L} for some LL\in{\mathbb{Z}}, we say a random variable XX has qq-negative binomial distribution with parameters (p,b)(p,b) if

𝐏(X=n)=1Zpn(b;q)n(q;q)n,n{0,1,2,},\mathbf{P}\left(X=n\right)=\frac{1}{Z}p^{n}\frac{(b;q)_{n}}{(q;q)_{n}},\qquad n\in\{0,1,2,\dots\},

where (a;q)n:=(1a)(1aq)(1aqn1)(a;q)_{n}:=(1-a)(1-aq)\dots(1-aq^{n-1}) for n1n\geq 1, (a;q)0:=1(a;q)_{0}:=1, and ZZ is a normalising constant, which can be given explicitly as Z=(pb;q)(b;q)Z=\frac{(pb;q)_{\infty}}{(b;q)_{\infty}}; in this case we write XqNB(b,p)X\sim\mathrm{qNB}(b,p). Note that, if p,b[0,1)p,b\in[0,1), then the support of XX is +{\mathbb{Z}}_{+}, and if p<0p<0 and b=qLb=q^{-L} for some LL\in{\mathbb{Z}}, then the support of XX is {0,1,2,,L}\{0,1,2,\dots,L\}.

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 1+11+1 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 qq-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 2\mathbb{Z}^{2}, 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 M/M/1M/M/1 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 2MX2M-X 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.