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

The leftmost column of ordered Chinese Restaurant Process up-down chains: intertwining and convergence

Kelvin Rivera-Lopez This work was supported in part by NSF grant DMS-1855568.Université de Lorraine, CNRS, IECL, F-54000 Nancy, France, [email protected]    Douglas Rizzolo University of Delaware, [email protected]
Abstract

Recently there has been significant interest in constructing ordered analogues of Petrov’s two-parameter extension of Ethier and Kurtz’s infinitely-many-neutral-alleles diffusion model. One method for constructing these processes goes through taking an appropriate diffusive limit of Markov chains on integer compositions called ordered Chinese Restaurant Process up-down chains. The resulting processes are diffusions whose state space is the set of open subsets of the open unit interval. In this paper we begin to study nontrivial aspects of the order structure of these diffusions. In particular, for a certain choice of parameters, we take the diffusive limit of the size of the first component of ordered Chinese Restaurant Process up-down chains and describe the generator of the limiting process. We then relate this to the size of the leftmost maximal open subset of the open-set valued diffusions. This is challenging because the function taking an open set to the size of its leftmost maximal open subset is discontinuous. Our methods are based on establishing intertwining relations between the processes we study.

Introduction

The construction and analysis of ordered analogues of Petrov’s [15] two-parameter extension of Ethier and Kurtz’s [3] infinitely-many-neutral-alleles diffusion model has recently attracted significant interest in the literature [7, 8, 20, 21, 23]. Recall that the 𝙴𝙺𝙿(α,θ){\tt EKP}(\alpha,\theta) diffusions constructed in [15] are a family of Feller diffusions on the closure of the Kingman simplex

¯={𝐱=(x1,x2,):x1x20,i1xi1}\overline{\nabla}_{\infty}=\left\{\mathbf{x}=(x_{1},x_{2},\dots)\ :\ x_{1}\geq x_{2}\geq\cdots\geq 0,\sum_{i\geq 1}x_{i}\leq 1\right\}

whose generator acts on the unital algebra generated by ϕm(𝐱)=i1xim\phi_{m}(\mathbf{x})=\sum_{i\geq 1}x_{i}^{m}, m2m\geq 2 by

𝒢=12(i=1xi2xi2i,j=1xixj2xixji=1(θxi+α)xi).\mathcal{G}=\frac{1}{2}\left(\sum_{i=1}^{\infty}x_{i}\frac{\partial^{2}}{\partial x_{i}^{2}}-\sum_{i,j=1}^{\infty}x_{i}x_{j}\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}-\sum_{i=1}^{\infty}(\theta x_{i}+\alpha)\frac{\partial}{\partial x_{i}}\right).

In [20], for each θ0\theta\geq 0, 0α<10\leq\alpha<1, and α+θ>0\alpha+\theta>0, we constructed a Feller diffusion 𝐗(α,θ)\mathbf{X}^{(\alpha,\theta)} whose state space 𝒰\mathcal{U} is the set of open subsets of (0,1)(0,1) such that the ranked sequence of lengths of maximal open intervals in 𝐗(α,θ)\mathbf{X}^{(\alpha,\theta)} is an 𝙴𝙺𝙿(α,θ){\tt EKP}(\alpha,\theta) diffusion. This was done by considering the scaling limit of up-down chains associated to the ordered Chinese Restaurant Process.

While many interesting properties of 𝐗(α,θ)\mathbf{X}^{(\alpha,\theta)} can be obtained from the corresponding properties for 𝙴𝙺𝙿(α,θ){\tt EKP}(\alpha,\theta) diffusions, properties that depend on the order structure cannot be. In this paper we begin to study nontrivial aspects of the order structure of these diffusions. Motivated by [6, Theorem 2 and Theorem 19] and [5, Theorem 5], which consider similar properties in closely related tree-valued processes, we consider the evolution of the left-most maximal open interval of 𝐗(α,0)\mathbf{X}^{(\alpha,0)} in running in its (α,0)(\alpha,0)-Poisson-Dirichlet interval partition stationarity distribution. Recall that the (α,0)(\alpha,0)-Poisson-Dirichlet interval partition is the distribution of {t(0,1):V1t>0}\{t\in(0,1):V_{1-t}>0\} where VtV_{t} is a (22α)(2-2\alpha)-dimensional Bessel process started from 0. We prove the following result.

Theorem 1.1.

Define ξ:𝒰[0,1]\xi\colon\mathcal{U}\to[0,1] by ξ(u)=inf{s>0:s[0,1]u}\xi(u)=\inf\{s>0:s\in[0,1]\setminus u\}. If 𝐗(α,0)\mathbf{X}^{(\alpha,0)} is running in its (α,0)(\alpha,0)-Poisson-Dirichlet interval partition stationarity distribution, then ξ(𝐗(α,0))\xi(\mathbf{X}^{(\alpha,0)}) is a Feller process. Moreover, the generator of its semigroup :𝒟C[0,1]C[0,1]\mathcal{L}\colon\mathcal{D}\subseteq C[0,1]\to C[0,1] is given by

f(x)=x(1x)f′′(x)αf(x)\mathcal{L}f(x)=x(1-x)f^{\prime\prime}(x)-\alpha f^{\prime}(x)

for x(0,1)x\in(0,1), where the domain 𝒟\mathcal{D} of \mathcal{L} consists of functions ff satisfying

  1. (D1)

    fC2(0,1)f\in C^{2}(0,1) and ζ(x)=x(1x)f′′(x)αf(x)\zeta(x)=x(1-x)f^{\prime\prime}(x)-\alpha f^{\prime}(x) extends continuously to [0,1][0,1],

  2. (D2)

    01(f(x)f(0))xα1(1x)α1𝑑x=0,\int_{0}^{1}(f(x)-f(0))x^{-\alpha-1}(1-x)^{\alpha-1}\,dx=0, and

  3. (D3)

    f(x)(1x)α0f^{\prime}(x)(1-x)^{\alpha}\to 0 as x1x\to 1.

We consider only the (α,0)(\alpha,0) case because the known stationary distribution of 𝐗(α,θ)\mathbf{X}^{(\alpha,\theta)} is an (α,θ)(\alpha,\theta)-Poisson-Dirichlet interval partition and, except in the (α,0)(\alpha,0) case, with probability 1 interval partitions with these distributions do not have left-most maximal open intervals. We remark that our theorem statement could be slightly simpler if we knew that 𝐗(α,θ)\mathbf{X}^{(\alpha,\theta)} had a unique stationary distribution, but this is currently an open problem.

Our proof is based on taking the scaling limit of the left-most coordinate in an up-down chain on compositions based on the ordered Chinese Restaurant Process, which are the same chains that were used in [20] to construct 𝐗(α,0)\mathbf{X}^{(\alpha,0)}.

Definition 1.1.

For n1n\geq 1, a composition of nn is a tuple σ=(σ1,,σk)\sigma=(\sigma_{1},...,\sigma_{k}) of positive integers that sum to nn. The composition of n=0n=0 is the empty tuple, which we denote by \varnothing. We write |σ|=n|\sigma|=n and (σ)=k\ell(\sigma)=k when σ\sigma is a composition of nn with kk components. We denote the set of all compositions of nn by 𝒞n\mathcal{C}_{n} and their union by 𝒞=n0𝒞n\mathcal{C}=\cup_{n\geq 0}\,\mathcal{C}_{n}.

An up-down chain on 𝒞n\mathcal{C}_{n} is a Markov chain whose steps can be factored into two parts: 1) an up-step from 𝒞n\mathcal{C}_{n} to 𝒞n+1\mathcal{C}_{n+1} according to a kernel pp^{\uparrow} followed by 2) a down-step from 𝒞n+1\mathcal{C}_{n+1} to 𝒞n\mathcal{C}_{n} according to a kernel pp^{\downarrow}. The probability Tn(σ,σ)T_{n}(\sigma,\sigma^{\prime}) of transitioning from σ\sigma to σ\sigma^{\prime} can then be written as

Tn(σ,σ)=τ𝒞n+1p(σ,τ)p(τ,σ).T_{n}(\sigma,\sigma^{\prime})=\sum_{\tau\in\mathcal{C}_{n+1}}p^{\uparrow}(\sigma,\tau)p^{\downarrow}(\tau,\sigma^{\prime}). (1)

Up-down chains on compositions, and more generally, on graded sets, have been studied in a variety of contexts [2, 6, 9, 10, 11, 15, 16], often in connection with their nice algebraic and combinatorial properties.

In the up-down chains we considered, the up-step kernel p(α,θ)p^{\uparrow}_{(\alpha,\theta)} is given by an (α,θ)(\alpha,\theta)-ordered Chinese Restaurant Process growth step [18]. In the Chinese Restaurant Process analogy, we view τ=(τ1,,τk)𝒞n\tau=(\tau_{1},\dots,\tau_{k})\in\mathcal{C}_{n} as an ordered list of the number of customers at kk occupied tables in a restaurant, so that τi\tau_{i} is the number of customers at the ithi^{th} table on the list. During an up-step, a new customer enters the restaurant and chooses a table to sit at according to the following rules:

  • The new customer joins table ii with probability (τiα)/(n+θ)(\tau_{i}-\alpha)/(n+\theta), resulting in a step from τ\tau to (τ1,,τi1,τi+1,τi+1,,τk)(\tau_{1},\dots,\tau_{i-1},\tau_{i}+1,\tau_{i+1},\dots,\tau_{k}).

  • The new customer starts a new table directly after table ii with probability α/(n+θ)\alpha/(n+\theta), resulting in a step from τ\tau to (τ1,,τi1,τi,1,τi+1,,τk)(\tau_{1},\dots,\tau_{i-1},\tau_{i},1,\tau_{i+1},\dots,\tau_{k}).

  • The new customer starts a new table at the start of the list with probability θ/(n+θ)\theta/(n+\theta), resulting in a step from τ\tau to (1,τ1,τ2,τk)(1,\tau_{1},\tau_{2}\dots,\tau_{k}).

We note that, for consistency with [7, 8], this up-step is the left-to-right reversal of the growth step in [18].

The down-step kernel pp^{\downarrow} we consider can also be thought of in terms of the restaurant analogy. During a down-step, a seated customer gets up and exits the restaurant according to the following rule:

  • The seated customer is chosen uniformly at random, resulting in a step from τ\tau to (τ1,,τi1,τi1,τi+1,,τk)(\tau_{1},\ldots,\tau_{i-1},\tau_{i}-1,\tau_{i+1},\ldots,\tau_{k}) with probability τi/n\tau_{i}/n (the ithi^{th} coordinate is to be contracted away if τi1=0\tau_{i}-1=0, or if the ithi^{th} table is no longer occupied).

Note that, in contrast to the up-step, the down-step does not depend on (α,θ)(\alpha,\theta).

Let (𝐗n(α,θ)(k))k0(\mathbf{X}^{(\alpha,\theta)}_{n}(k))_{k\geq 0} be a Markov chain on 𝒞n\mathcal{C}_{n} with transition kernel Tn(α,θ)T^{(\alpha,\theta)}_{n} defined as in Equation (1) using the p(α,θ)p^{\uparrow}_{(\alpha,\theta)} and pp^{\downarrow} just described. A Poissonized version of this chain was considered in [21, 23]. It can be shown that 𝐗n(α,θ)\mathbf{X}^{(\alpha,\theta)}_{n} is an aperiodic, irreducible chain. We denote its unique stationary distribution by Mn(α,θ)M^{(\alpha,\theta)}_{n} and note that this is the left-to-right reversal of the (α,θ)(\alpha,\theta)-regenerative composition structures introduced in [12].

The projection ϕ(σ)=σ1\phi(\sigma)=\sigma_{1} for σ\sigma\neq\varnothing gives rise to the leftmost column processes, defined by Yn(α,θ)=ϕ(𝐗n(α,θ))Y^{(\alpha,\theta)}_{n}=\phi(\mathbf{X}^{(\alpha,\theta)}_{n}). Let νn(α,θ)=Mn(α,θ)ϕ1\nu_{n}^{(\alpha,\theta)}=M^{(\alpha,\theta)}_{n}\circ\phi^{-1}, the distribution of the leftmost column when the up-down chain is in stationarity. The following result, interesting in its own right, is a key step in our proof of Theorem 1.1.

Theorem 1.2.

For n1n\geq 1, let μn\mu_{n} be a distribution on {1,,n}\{1,\ldots,n\}. Then, for all nn, the up-down chain 𝐗n(α,0)\mathbf{X}^{(\alpha,0)}_{n} can be initialized so that Yn(α,θ)Y^{(\alpha,\theta)}_{n} is a Markov chain with initial distribution μn\mu_{n}. Moreover, for any such sequence of initial conditions for 𝐗n(α,0)\mathbf{X}^{(\alpha,0)}_{n}, if the sequence {n1Yn(α,0)(0)}n1\{n^{-1}Y^{(\alpha,0)}_{n}(0)\}_{n\geq 1} has a limiting distribution μ\mu, then we have the convergence

(n1Yn(α,0)(n2t))t0(F(t))t0\left(n^{-1}Y^{(\alpha,0)}_{n}(\lfloor n^{2}t\rfloor)\right)_{t\geq 0}\Longrightarrow(F(t))_{t\geq 0}

in the Skorokhod space D([0,),[0,1])D([0,\infty),[0,1]), where FF is a Feller process with generator \mathcal{L} (as in Theorem 1.1) and initial distribution μ\mu.

While there are many ways to prove a result like Theorem 1.2, we take an approach based on the algebraic properties of the ordered Chinese Restaurant Process up-down chains. In particular, our proof is based on the following surprising intertwining result. For a positive integer ii and composition σ\sigma, we use the notation (i,σ)(i,\sigma) as a shorthand for the composition (i,σ1,σ2,,σ(σ))(i,\sigma_{1},\sigma_{2},\ldots,\sigma_{\ell(\sigma)}).

Theorem 1.3.

For n1n\geq 1, let Λn\Lambda_{n} be the transition kernel from {1,,n}\{1,\dots,n\} to 𝒞n\mathcal{C}_{n} given by

Λn(i,(i,σ))=Mni(α,α)(σ),\Lambda_{n}(i,(i,\sigma))=M^{(\alpha,\alpha)}_{n-i}(\sigma),

and let KnK_{n} be the transition kernel from [0,1][0,1] to {1,,n}\{1,\dots,n\} given by

Kn(x,i)=(ni)xi(1x)ni+νn(α,α)(i)(1x)n.K_{n}(x,i)=\binom{n}{i}x^{i}(1-x)^{n-i}+\nu^{(\alpha,\alpha)}_{n}(i)(1-x)^{n}.

If the initial distribution of 𝐗n(α,0)\mathbf{X}_{n}^{(\alpha,0)} is of the form μΛn\mu\Lambda_{n} for some distribution μ\mu on {1,,n}\{1,\dots,n\}, then the process Yn(α,0)Y_{n}^{(\alpha,0)} is Markovian. In this case, the following intertwining relations hold:

  1. (i)

    ΛnTn(α,0)=Qn(α,0)Λn,\Lambda_{n}T^{(\alpha,0)}_{n}=Q^{(\alpha,0)}_{n}\Lambda_{n}, where Qn(α,0)Q_{n}^{(\alpha,0)} is the transition kernel of Yn(α,0)Y_{n}^{(\alpha,0)}, and

  2. (ii)

    Knetn(n+1)(Qn(α,0)𝟏)=UtKnK_{n}e^{tn(n+1)(Q_{n}^{(\alpha,0)}-\mathbf{1})}=U_{t}K_{n} for t0t\geq 0, where UtU_{t} is the semigroup generated by the operator \mathcal{L} defined in Theorem 1.1 and 𝟏\mathbf{1} denotes the identity operator.

This paper is organized as follows. In Section 2, we show that the (α,0)(\alpha,0) leftmost column process is intertwined with its corresponding up-down chain and describe its transition kernel explicitly. This establishes part of Theorem 1.3. In Section 3, we state a condition under which the convergence of Markov processes can be obtained from some commutation relations involving generators. In Section 4, we analyze the generator of the limiting process. In Section 5, we show that our generators satisfy the commutation relations appearing in the result of Section 3. In Section 6, we verify the convergence condition appearing in the result in Section 3. In Section 7, we provide general conditions under which commutation relations involving generators lead to the corresponding relations for their semigroups. Finally in Section 8, we prove Theorems 1.1, 1.2, and 1.3.

The following will be used throughout this paper. For a compact topological space XX, we denote by C(X)C(X) the space of continuous functions from XX to \mathbb{R} equipped with the supremum norm. Finite topological spaces will always be equipped with the discrete topology. Any sum or product over an empty index set will be regarded as a zero or one, respectively. The set of positive integers {1,,k}\{1,...,k\} will be denoted by [k][k]. The falling factorial will be denoted using factorial exponents – that is, xb=x(x1)(xb+1)x^{\downarrow b}=x(x-1)\cdot\ldots\cdot(x-b+1) for a real number xx and nonnegative integer bb, and 00=10^{\downarrow 0}=1 by convention. The rising factorial will be denoted by (x)b=x(x+1)(x+b1)(x)_{b}=x(x+1)\cdots(x+b-1). We denote the gamma function by Γ(x)\Gamma(x). Multinomial coefficients will be denoted using the shorthand

(|σ|σ)={(|σ|σ1,,σ(σ)),σ,1,σ=.\binom{|\sigma|}{\sigma}=\begin{cases}\displaystyle\binom{|\sigma|}{\sigma_{1},...,\sigma_{\ell(\sigma)}},&\sigma\neq\varnothing,\\ 1,&\sigma=\varnothing.\end{cases}

The Leftmost Column Process

Our study of the leftmost column process will be mainly focused on the θ=0\theta=0 case. However, it will be useful to study the distribution of the (α,α)(\alpha,\alpha) leftmost column process when the up-down chain is in stationarity. As we will see, this distribution has a role in the evolution of the (α,0)(\alpha,0) process.

Proposition 2.1.

The stationary distribution of 𝐗n(α,α)\mathbf{X}^{(\alpha,\alpha)}_{n} is given by

Mn(α,α)(σ)=(nσ)1(α)nj=1(σ)α(1α)σj1,σ𝒞n,n0.M^{(\alpha,\alpha)}_{n}(\sigma)=\binom{n}{\sigma}\frac{1}{(\alpha)_{n}}\prod_{j=1}^{\ell(\sigma)}\alpha\,(1-\alpha)_{\sigma_{j}-1},\qquad\sigma\in\mathcal{C}_{n},\,n\geq 0.

Moreover, the following consistency conditions hold:

Mn(α,α)=Mn1(α,α)p(α,α)=Mn+1(α,α)p,n1.M^{(\alpha,\alpha)}_{n}=M^{(\alpha,\alpha)}_{n-1}p^{\uparrow}_{(\alpha,\alpha)}=M^{(\alpha,\alpha)}_{n+1}p^{\downarrow},\qquad n\geq 1. (2)
Proof.

The stationary distribution of 𝐗n(α,θ)\mathbf{X}^{(\alpha,\theta)}_{n} is identified in [20, Theorem 1.1] and the formula in the special case α=θ\alpha=\theta follows from [12, Formula 48]. The consistency conditions follows from [18, Proposition 6]. ∎

Proposition 2.2.

If 𝐗n(α,α)\mathbf{X}^{(\alpha,\alpha)}_{n} has distribution Mn(α,α)M^{(\alpha,\alpha)}_{n}, then Yn(α,α)Y^{(\alpha,\alpha)}_{n} has distribution

νn(α,α)(i)=(ni)α(1α)i1(ni+α)i 1(1in),i0,n1.\nu^{(\alpha,\alpha)}_{n}(i)=\binom{n}{i}\frac{\alpha\,(1-\alpha)_{i-1}}{(n-i+\alpha)_{i}}\,\mathbbm{1}(1\leq i\leq n),\qquad i\geq 0,\,n\geq 1.
Proof.

Let 1in1\leq i\leq n and σ𝒞ni\sigma\in\mathcal{C}_{n-i}. It can be verified that

Mn(α,α)(i,σ)=νn(α,α)(i)Mni(α,α)(σ).M^{(\alpha,\alpha)}_{n}(i,\sigma)=\nu^{(\alpha,\alpha)}_{n}(i)M^{(\alpha,\alpha)}_{n-i}(\sigma). (3)

Summing over σ\sigma concludes the proof. ∎

Let ni1n\geq i\geq 1 and σ𝒞ni\sigma\in\mathcal{C}_{n-i}. Consider taking an (α,0)(\alpha,0) up-step from (i,σ)(i,\sigma) followed by a down-step. Let UU be the event in which this up-step stacks a box on the first column of (i,σ)(i,\sigma), and let DD be the event in which the down-step removes a box from the first column of a composition. Then, ri,i+1=(UDc)r_{i,i+1}=\mathbb{P}(U\cap D^{c}), ri,i1=(UcD)r_{i,i-1}=\mathbb{P}(U^{c}\cap D), ri,i(1)=(UcDc)r^{(1)}_{i,i}=\mathbb{P}(U^{c}\cap D^{c}), ri,i(2)=(UD)r^{(2)}_{i,i}=\mathbb{P}(U\cap D), and ri,i=ri,i(1)+ri,i(2)r_{i,i}=r^{(1)}_{i,i}+r^{(2)}_{i,i} do not depend on σ\sigma. Indeed, we have the formulas

ri,i1=i(ni+α)n(n+1),ri,i(1)=(ni+1)(ni+α)n(n+1),ri,i+1=(iα)(ni)n(n+1),ri,i(2)=(iα)(i+1)n(n+1).\begin{matrix}r_{i,i-1}&=&\frac{i(n-i+\alpha)}{n(n+1)},\quad&r^{(1)}_{i,i}&=&\frac{(n-i+1)(n-i+\alpha)}{n(n+1)},\\ \vspace{-2mm}\\ r_{i,i+1}&=&\frac{(i-\alpha)(n-i)}{n(n+1)},\quad&r^{(2)}_{i,i}&=&\frac{(i-\alpha)(i+1)}{n(n+1)}.\end{matrix} (4)

We use these formulas to define r0,1,r_{0,-1}, r0,1,r_{0,1}, r0,0(1),r^{(1)}_{0,0}, r0,0(2),r^{(2)}_{0,0}, and r0,0=r0,0(1)+r0,0(2).r_{0,0}=r^{(1)}_{0,0}+r^{(2)}_{0,0}. Moreover, we extend ri,jr_{i,j} to be zero for all other integer arguments ii and jj.

The following is a useful identity relating the transition kernels of the (α,0)(\alpha,0) and (α,α)(\alpha,\alpha) chains.

Proposition 2.3.

For n1n\geq 1 and (i,σ),(j,σ)𝒞n(i,\sigma),(j,\sigma^{\prime})\in\mathcal{C}_{n}, we have the identity

Tn(α,0)((i,σ),(j,σ))\displaystyle T^{(\alpha,0)}_{n}\Big{(}(i,\sigma),(j,\sigma^{\prime})\Big{)} =ri,jp(α,α)(σ,σ)𝟙(j=i1)+ri,jp(σ,σ)𝟙(j=i+1)\displaystyle=r_{i,j}p^{\uparrow}_{(\alpha,\alpha)}(\sigma,\sigma^{\prime})\mathbbm{1}(j=i-1)+r_{i,j}p^{\downarrow}(\sigma,\sigma^{\prime})\mathbbm{1}(j=i+1)
+(ri,i(1)Tni(α,α)(σ,σ)+ri,i(2)𝟙(σ=σ))𝟙(j=i)\displaystyle\quad+(r^{(1)}_{i,i}T^{(\alpha,\alpha)}_{n-i}(\sigma,\sigma^{\prime})+r^{(2)}_{i,i}\mathbbm{1}(\sigma=\sigma^{\prime}))\mathbbm{1}(j=i)
+r1,0p(α,α)(σ,(j,σ))𝟙(i=1).\displaystyle\quad+r_{1,0}p^{\uparrow}_{(\alpha,\alpha)}(\sigma,(j,\sigma^{\prime}))\mathbbm{1}(i=1).
Proof.

Fix (i,σ)(i,\sigma) and (j,σ)(j,\sigma^{\prime}) in 𝒞n\mathcal{C}_{n}. Let 𝐂\mathbf{C}^{\uparrow} be the composition obtained by performing an (α,0)(\alpha,0) up-step from (i,σ)(i,\sigma) and 𝐂\mathbf{C}^{\downarrow} be the composition obtained by performing a down-step from 𝐂\mathbf{C}^{\uparrow}. As before, let UU be the event in which the up-step adds to the first column of a composition and DD be the event in which the down-step removes from the first column of a composition. Then, we have that

U={𝐂=(i+1,σ)},Uc={𝐂1=i},Dc{𝐂1=𝐂1},U=\big{\{}\mathbf{C}^{\uparrow}=(i+1,\sigma)\big{\}},\quad U^{c}=\{\mathbf{C}^{\uparrow}_{1}=i\},\quad D^{c}\subset\{\mathbf{C}^{\downarrow}_{1}=\mathbf{C}^{\uparrow}_{1}\},

and

D\displaystyle D {𝐂1>1,𝐂=(𝐂11,(𝐂)2(𝐂))}{𝐂1=1,𝐂=(𝐂)2(𝐂)}.\displaystyle\subset\Big{\{}\mathbf{C}^{\uparrow}_{1}>1,\,\mathbf{C}^{\downarrow}=(\mathbf{C}^{\uparrow}_{1}-1,(\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})})\Big{\}}\cup\Big{\{}\mathbf{C}^{\uparrow}_{1}=1,\,\mathbf{C}^{\downarrow}=(\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})}\Big{\}}.

To obtain the identity, we note that

Tn(α,0)((i,σ),(j,σ))={𝐂=(j,σ)},T^{(\alpha,0)}_{n}((i,\sigma),(j,\sigma^{\prime}))=\mathbb{P}\{\mathbf{C}^{\downarrow}=(j,\sigma^{\prime})\},

and rewrite this probability by conditioning on the above sets. Of particular importance will be the following observations: the conditional distribution of (𝐂)2(𝐂)(\mathbf{C}^{\downarrow})_{2}^{\ell(\mathbf{C}^{\downarrow})} given (𝐂)2(𝐂)(\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})} and DcD^{c} is p((𝐂)2(𝐂),),p^{\downarrow}((\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})},\cdot\,), and conditionally given UcU^{c}, (𝐂)2(𝐂)(\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})} is independent of DD and has distribution p(α,α)(σ,).p^{\uparrow}_{(\alpha,\alpha)}(\sigma,\cdot\,). We also make use of the fact that the events {𝐂=(n+1|ρ|,ρ)}\{\mathbf{C}^{\uparrow}=(n+1-|\rho|,\rho)\} and {(𝐂)2(𝐂)=ρ}\{(\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})}=\rho\} are identical, since the size of 𝐂\mathbf{C}^{\uparrow} is known to be n+1n+1.

Our first conditional probability is given by

(𝐂=(j,σ)|U,D)\displaystyle\mathbb{P}(\mathbf{C}^{\downarrow}=(j,\sigma^{\prime})|U,D) =(𝐂=(j,σ)|𝐂=(i+1,σ),D)\displaystyle=\mathbb{P}(\mathbf{C}^{\downarrow}=(j,\sigma^{\prime})|\mathbf{C}^{\uparrow}=(i+1,\sigma),D)
=((i,σ)=(j,σ)|𝐂=(i+1,σ),D)\displaystyle=\mathbb{P}((i,\sigma)=(j,\sigma^{\prime})|\mathbf{C}^{\uparrow}=(i+1,\sigma),D)
=𝟙((j,σ)=(i,σ)).\displaystyle=\mathbbm{1}((j,\sigma^{\prime})=(i,\sigma)).

Next, we will condition on UDcU\cap D^{c}. Notice that this is a null set when i=ni=n. When i<ni<n, we have

(𝐂=(j,σ)|U,Dc)\displaystyle\mathbb{P}(\mathbf{C}^{\downarrow}=(j,\sigma^{\prime})|U,D^{c}) =(𝐂1=j,(𝐂)2(𝐂)=σ|𝐂=(i+1,σ),Dc)\displaystyle=\mathbb{P}(\mathbf{C}^{\uparrow}_{1}=j,(\mathbf{C}^{\downarrow})_{2}^{\ell(\mathbf{C}^{\downarrow})}=\sigma^{\prime}|\mathbf{C}^{\uparrow}=(i+1,\sigma),D^{c})
=𝟙(j=i+1)((𝐂)2(𝐂)=σ|(𝐂)2(𝐂)=σ,Dc)\displaystyle=\mathbbm{1}(j=i+1)\mathbb{P}((\mathbf{C}^{\downarrow})_{2}^{\ell(\mathbf{C}^{\downarrow})}=\sigma^{\prime}|(\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})}=\sigma,D^{c})
=𝟙(j=i+1)p(σ,σ).\displaystyle=\mathbbm{1}(j=i+1)p^{\downarrow}(\sigma,\sigma^{\prime}).

Conditioning on UcDU^{c}\cap D will require two cases. For i>1i>1, we have

(𝐂=(j,σ)|Uc,D)\displaystyle\mathbb{P}(\mathbf{C}^{\downarrow}=(j,\sigma^{\prime})|U^{c},D) =(𝐂=(j,σ)|𝐂1=i,D)\displaystyle=\mathbb{P}(\mathbf{C}^{\downarrow}=(j,\sigma^{\prime})|\mathbf{C}^{\uparrow}_{1}=i,D)
=((i1,(𝐂)2(𝐂))=(j,σ)|𝐂1=i,D)\displaystyle=\mathbb{P}((i-1,(\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})})=(j,\sigma^{\prime})|\mathbf{C}^{\uparrow}_{1}=i,D)
=𝟙(j=i1)((𝐂)2(𝐂)=σ|Uc,D)\displaystyle=\mathbbm{1}(j=i-1)\mathbb{P}((\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})}=\sigma^{\prime}|U^{c},D)
=𝟙(j=i1)((𝐂)2(𝐂)=σ|Uc)\displaystyle=\mathbbm{1}(j=i-1)\mathbb{P}((\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})}=\sigma^{\prime}|U^{c})
=𝟙(j=i1)p(α,α)(σ,σ),\displaystyle=\mathbbm{1}(j=i-1)p^{\uparrow}_{(\alpha,\alpha)}(\sigma,\sigma^{\prime}),

and for i=1i=1, we have

(𝐂=(j,σ)|Uc,D)\displaystyle\mathbb{P}(\mathbf{C}^{\downarrow}=(j,\sigma^{\prime})|U^{c},D) =(𝐂=(j,σ)|𝐂1=1,D)\displaystyle=\mathbb{P}(\mathbf{C}^{\downarrow}=(j,\sigma^{\prime})|\mathbf{C}^{\uparrow}_{1}=1,D)
=((𝐂)2(𝐂)=(j,σ)|Uc,D)\displaystyle=\mathbb{P}((\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})}=(j,\sigma^{\prime})|U^{c},D)
=((𝐂)2(𝐂)=(j,σ)|Uc)\displaystyle=\mathbb{P}((\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})}=(j,\sigma^{\prime})|U^{c})
=p(α,α)(σ,(j,σ)).\displaystyle=p^{\uparrow}_{(\alpha,\alpha)}(\sigma,(j,\sigma^{\prime})).

Finally, we condition on UcDcU^{c}\cap D^{c}. We have that

(𝐂=(j,σ)|Uc,Dc)\displaystyle\mathbb{P}\big{(}\mathbf{C}^{\downarrow}=(j,\sigma^{\prime})|U^{c},D^{c}\big{)}
=(𝐂1=j,(𝐂)2(𝐂)=σ|𝐂1=i,Dc)\displaystyle=\mathbb{P}\Big{(}\mathbf{C}^{\uparrow}_{1}=j,(\mathbf{C}^{\downarrow})_{2}^{\ell(\mathbf{C}^{\downarrow})}=\sigma^{\prime}|\mathbf{C}^{\uparrow}_{1}=i,D^{c}\Big{)}
=𝟙(j=i)((𝐂)2(𝐂)=σ|Uc,Dc)\displaystyle=\mathbbm{1}(j=i)\,\mathbb{P}\Big{(}(\mathbf{C}^{\downarrow})_{2}^{\ell(\mathbf{C}^{\downarrow})}=\sigma^{\prime}|U^{c},D^{c}\Big{)}
=𝟙(j=i)τ𝒞n+1i((𝐂)2(𝐂)=τ|Uc,Dc)((𝐂)2(𝐂)=σ|𝐂=(i,τ),Dc)\displaystyle=\mathbbm{1}(j=i)\!\!\!\!\sum_{\tau\in\mathcal{C}_{n+1-i}}\!\!\!\!\!\mathbb{P}\Big{(}(\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})}=\tau|U^{c},D^{c}\Big{)}\,\mathbb{P}\Big{(}(\mathbf{C}^{\downarrow})_{2}^{\ell(\mathbf{C}^{\downarrow})}=\sigma^{\prime}|\mathbf{C}^{\uparrow}=(i,\tau),D^{c}\Big{)}
=𝟙(j=i)τ𝒞n+1i((𝐂)2(𝐂)=τ|Uc)((𝐂)2(𝐂)=σ|(𝐂)2(𝐂)=τ,Dc)\displaystyle=\mathbbm{1}(j=i)\!\!\!\!\sum_{\tau\in\mathcal{C}_{n+1-i}}\!\!\!\!\!\mathbb{P}\Big{(}(\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})}=\tau|U^{c}\Big{)}\,\mathbb{P}\Big{(}(\mathbf{C}^{\downarrow})_{2}^{\ell(\mathbf{C}^{\downarrow})}=\sigma^{\prime}|(\mathbf{C}^{\uparrow})_{2}^{\ell(\mathbf{C}^{\uparrow})}=\tau,D^{c}\Big{)}
=𝟙(j=i)τ𝒞n+1ip(α,α)(σ,τ)p(τ,σ)\displaystyle=\mathbbm{1}(j=i)\!\!\!\sum_{\tau\in\mathcal{C}_{n+1-i}}\!\!\!p^{\uparrow}_{(\alpha,\alpha)}(\sigma,\tau)p^{\downarrow}(\tau,\sigma^{\prime})
=𝟙(j=i)Tni(α,α)(σ,σ).\displaystyle=\mathbbm{1}(j=i)\,T^{(\alpha,\alpha)}_{n-i}(\sigma,\sigma^{\prime}).

Collecting the terms above with the appropriate terms in (4) establishes the result. ∎

Let n1n\geq 1. We define a transition kernel Λn\Lambda_{n} from [n][n] to 𝒞n\mathcal{C}_{n} by

Λn(i,(i,σ))=Mni(α,α)(σ),\Lambda_{n}(i,(i,\sigma))=M^{(\alpha,\alpha)}_{n-i}(\sigma),

and a transition kernel Φn\Phi_{n} from 𝒞n\mathcal{C}_{n} to [n][n] by

Φ(σ,i)=𝟙(σ1=i).\Phi(\sigma,i)=\mathbbm{1}(\sigma_{1}=i).
Proposition 2.4.

For n1n\geq 1, the transition kernel Qn(α,0)=ΛnTn(α,0)ΦnQ^{(\alpha,0)}_{n}=\Lambda_{n}T^{(\alpha,0)}_{n}\Phi_{n} satisfies

ΛnTn(α,0)=Qn(α,0)Λn.\Lambda_{n}T^{(\alpha,0)}_{n}=Q^{(\alpha,0)}_{n}\Lambda_{n}. (5)

Consequently, if the initial distribution of 𝐗n(α,0)\mathbf{X}_{n}^{(\alpha,0)} is of the form μΛn\mu\Lambda_{n}, then Yn(α,0)Y^{(\alpha,0)}_{n} is a time-homogeneous Markov chain with transition kernel Qn(α,0)Q^{(\alpha,0)}_{n}. Moreover, the transition kernel Qn(α,0)Q^{(\alpha,0)}_{n} is given explicitly by

Qn(α,0)(i,j)=ri,j+r1,0νn(α,α)(j)𝟙(i=1).Q^{(\alpha,0)}_{n}(i,j)=r_{i,j}+r_{1,0}\nu^{(\alpha,\alpha)}_{n}(j)\mathbbm{1}(i=1).
Proof.

Let CnC_{n} be the kernel on [n][n] defined by the right side of the above equation. Fix i,j[n]i,j\in[n] and σ𝒞nj\sigma^{\prime}\in\mathcal{C}_{n-j}. Using Proposition 2.3 and the identities (2) and (3), we compute

(ΛnTn(α,0))(i,(j,σ))\displaystyle(\Lambda_{n}T^{(\alpha,0)}_{n})(i,(j,\sigma^{\prime}))
=σ𝒞niΛn(i,(i,σ))Tn(α,0)((i,σ),(j,σ))\displaystyle\hskip 28.45274pt=\sum_{\sigma\in\mathcal{C}_{n-i}}\Lambda_{n}(i,(i,\sigma))T^{(\alpha,0)}_{n}((i,\sigma),(j,\sigma^{\prime}))
=ri,jσ𝒞niMni(α,α)(σ)(p(α,α)(σ,σ)𝟙(j=i1)+p(σ,σ)𝟙(j=i+1))\displaystyle\hskip 28.45274pt=r_{i,j}\!\!\sum_{\sigma\in\mathcal{C}_{n-i}}\!\!M^{(\alpha,\alpha)}_{n-i}(\sigma)\left(p^{\uparrow}_{(\alpha,\alpha)}(\sigma,\sigma^{\prime})\mathbbm{1}(j=i-1)+p^{\downarrow}(\sigma,\sigma^{\prime})\mathbbm{1}(j=i+1)\right)
+𝟙(j=i)σ𝒞niMni(α,α)(σ)(Tni(α,α)(σ,σ)ri,i(1)+𝟙(σ=σ)ri,i(2))\displaystyle\hskip 28.45274pt\quad+\mathbbm{1}(j=i)\!\!\sum_{\sigma\in\mathcal{C}_{n-i}}\!\!M^{(\alpha,\alpha)}_{n-i}(\sigma)\left(T^{(\alpha,\alpha)}_{n-i}(\sigma,\sigma^{\prime})r^{(1)}_{i,i}+\mathbbm{1}(\sigma=\sigma^{\prime})r^{(2)}_{i,i}\right)
+r1,0𝟙(i=1)σ𝒞niMni(α,α)(σ)p(α,α)(σ,(j,σ))\displaystyle\hskip 28.45274pt\quad+r_{1,0}\mathbbm{1}(i=1)\!\!\sum_{\sigma\in\mathcal{C}_{n-i}}\!\!M^{(\alpha,\alpha)}_{n-i}(\sigma)p^{\uparrow}_{(\alpha,\alpha)}(\sigma,(j,\sigma^{\prime}))
=ri,j(Mnj(α,α)(σ)𝟙(j=i1)+Mnj(α,α)(σ)𝟙(j=i+1))\displaystyle\hskip 28.45274pt=r_{i,j}\left(M^{(\alpha,\alpha)}_{n-j}(\sigma^{\prime})\mathbbm{1}(j=i-1)+M^{(\alpha,\alpha)}_{n-j}(\sigma^{\prime})\mathbbm{1}(j=i+1)\right)
+𝟙(j=i)(Mnj(α,α)(σ)ri,i(1)+Mnj(α,α)(σ)ri,i(2))+r1,0𝟙(i=1)Mn(α,α)(j,σ)\displaystyle\hskip 28.45274pt\quad+\mathbbm{1}(j=i)\left(M^{(\alpha,\alpha)}_{n-j}(\sigma^{\prime})r^{(1)}_{i,i}+M^{(\alpha,\alpha)}_{n-j}(\sigma^{\prime})r^{(2)}_{i,i}\right)+r_{1,0}\mathbbm{1}(i=1)M^{(\alpha,\alpha)}_{n}(j,\sigma^{\prime})
=ri,jMnj(α,α)(σ)+r1,0𝟙(i=1)νn(α,α)(j)Mnj(α,α)(σ)\displaystyle\hskip 28.45274pt=r_{i,j}M^{(\alpha,\alpha)}_{n-j}(\sigma^{\prime})+r_{1,0}\mathbbm{1}(i=1)\nu^{(\alpha,\alpha)}_{n}(j)M^{(\alpha,\alpha)}_{n-j}(\sigma^{\prime})
=Cn(i,j)Λn(j,(j,σ))\displaystyle\hskip 28.45274pt=C_{n}(i,j)\Lambda_{n}(j,(j,\sigma^{\prime}))
=(CnΛn)(i,(j,σ)).\displaystyle\hskip 28.45274pt=(C_{n}\Lambda_{n})(i,(j,\sigma^{\prime})).

The final equality follows from the fact that Λn(j,)\Lambda_{n}(j,\,\cdot\,) is supported on {σ𝒞n:σ1=j}\{\sigma\in\mathcal{C}_{n}:\sigma_{1}=j\}. This establishes the identity ΛnTn(α,0)=CnΛn.\Lambda_{n}T^{(\alpha,0)}_{n}=C_{n}\Lambda_{n}. Observing that ΛnΦn\Lambda_{n}\Phi_{n} is the identity kernel on [n][n], we find that

Qn(α,0)=ΛnTn(α,0)Φn=CnΛnΦn=Cn,Q^{(\alpha,0)}_{n}=\Lambda_{n}T^{(\alpha,0)}_{n}\Phi_{n}=C_{n}\Lambda_{n}\Phi_{n}=C_{n},

from which we obtain (5) and the explicit description of Qn(α,0)Q^{(\alpha,0)}_{n}. The final claim follows from applying Theorem 2 in [22]. ∎

Convergence from Commutation Relations

In this section, we provide a condition under which commutation relations between operators implies the convergence of those operators in an appropriate sense. In the interest of generality, we first state this condition in the setting of Banach spaces, but we then reformulate it in the context of Markov processes to suit our purposes. The general setting is as follows.

Let V,V1,V2,V,V_{1},V_{2},\ldots be Banach spaces and π1,π2,\pi_{1},\pi_{2},\ldots be uniformly bounded linear operators with πn:VVn\pi_{n}\colon V\to V_{n}. These spaces will be equipped with the following mode of convergence.

Definition 3.1.

A sequence {fn}n1\{f_{n}\}_{n\geq 1} with fnVnf_{n}\in V_{n} converges to an element fVf\in V (and we write fnff_{n}\to f) if

fnπnfn0,\|f_{n}-\pi_{n}f\|\xrightarrow[n\to\infty]{}0,

where for convenience, we denote every norm by the same symbol \|\cdot\|.

Proposition 3.1.

For n1n\geq 1, let Ln:DnVVnL_{n}\colon D_{n}\subset V\to V_{n} and An:VnVnA_{n}\colon V_{n}\to V_{n} be linear operators in addition to A:DVDA\colon D\subset V\to D. Suppose that for every fDf\in D,

  1. (i)

    AnLnf=LnAfA_{n}L_{n}f=L_{n}Af for large nn, and

  2. (ii)

    (Lnπn)f0(L_{n}-\pi_{n})f\longrightarrow 0 as nn\to\infty (the sequence need only be defined for large nn).

Then for fDf\in D, the sequence fn=Lnff_{n}=L_{n}f (defined for large nn) satisfies

fnfandAnfnAf.f_{n}\longrightarrow f\qquad\text{and}\qquad A_{n}f_{n}\longrightarrow Af.
Proof.

Let fDf\in D and nn be large enough so that (i) holds. In particular, we can define fn=Lnff_{n}=L_{n}f. Writing

fnπnf=Lnfπnf,\|f_{n}-\pi_{n}f\|=\|L_{n}f-\pi_{n}f\|,

it is clear that fnff_{n}\to f. Writing

AnfnπnAf\displaystyle\|A_{n}f_{n}-\pi_{n}Af\| =AnLnfπnAf\displaystyle=\|A_{n}L_{n}f-\pi_{n}Af\|
=LnAfπnAf\displaystyle=\|L_{n}Af-\pi_{n}Af\|
=(Lnπn)Af\displaystyle=\|(L_{n}-\pi_{n})Af\|

and noting that AfDAf\in D, we obtain the other convergence. ∎

In the probabilistic context, the above result has some additional consequences.

Theorem 3.1.

Let EE be a compact, separable metric space, AA be the generator of the Feller semigroup S(t)S(t) on C(E)C(E), and DD be a core for AA that is invariant under AA. For each n1n\geq 1, let EnE_{n} be a finite set endowed with the discrete topology, ZnZ_{n} be a Markov chain on EnE_{n}, γn:EnE\gamma_{n}\colon E_{n}\to E be any function, and Ln:DnC(E)C(En)L_{n}\colon D_{n}\subset C(E)\to C(E_{n}) be a linear operator. Denote the transition operator of ZnZ_{n} by SnS_{n} and the projection ffγnf\mapsto f\circ\gamma_{n} by πn:C(E)C(En)\pi_{n}\colon C(E)\to C(E_{n}). Let {δn}n1\{\delta_{n}\}_{n\geq 1} and {εn}n1\{\varepsilon_{n}\}_{n\geq 1} be positive sequences converging to zero such that εn1δn1.\varepsilon_{n}^{-1}\delta_{n}\to 1. Suppose that for fDf\in D, the following statements hold:

  1. (a)

    δn1(Sn𝟏)Lnf=LnAf\delta_{n}^{-1}(S_{n}-\mathbf{1})L_{n}f=L_{n}Af for large nn, and

  2. (b)

    (Lnπn)f0(L_{n}-\pi_{n})f\longrightarrow 0 as nn\to\infty (the sequence need only be defined for large nn).

Then,

  1. (i)

    the discrete semigroups {1,Sn,Sn2,}n1\{1,S_{n},S_{n}^{2},...\}_{n\geq 1} converge to {S(t)}t0\{S(t)\}_{t\geq 0} in the following sense: for all fC(E)f\in C(E) and t0t\geq 0,

    Snt/εnπnfnS(t)fS_{n}^{\lfloor t/\varepsilon_{n}\rfloor}\pi_{n}f\xrightarrow[n\to\infty]{}S(t)f
  2. (ii)

    the above convergence is uniform in tt on bounded intervals, and

  3. (iii)

    if AA is conservative and the distributions of γn(Zn(0))\gamma_{n}(Z_{n}(0)) converge, say to μ\mu, then we have the convergence of paths

    γn(Znt/εn)F(t)\gamma_{n}(Z_{n}\lfloor t/\varepsilon_{n}\rfloor)\Longrightarrow F(t)

    in the Skorokhod space D([0,),E)D([0,\infty),E), where F(t)F(t) is a Feller process with initial distribution μ\mu and generator AA.

Proof.

This is a combination of Proposition 3.1 and standard convergence results. In particular, for fDf\in D, we can define the sequence fn=Lnff_{n}=L_{n}f for large nn and obtain the convergence

fnfandδn1(Sn𝟏)fnAf.f_{n}\longrightarrow f\quad\qquad\text{and}\quad\qquad\delta_{n}^{-1}(S_{n}-\mathbf{1})f_{n}\longrightarrow Af.

Recalling that εn1δn1,\varepsilon_{n}^{-1}\delta_{n}\to 1, we then obtain the convergence εn1(Sn𝟏)fnAf.\varepsilon_{n}^{-1}(S_{n}-\mathbf{1})f_{n}\to Af. Applying Chapter 1 Theorem 6.5 in [4] then yields the convergence of semigroups in (i) and (ii). Applying Chapter 4 Theorem 2.12 in [4] yields the path convergence in (iii). ∎

The Limiting Generator

In this section, we introduce the generator of a Feller process on [0,1][0,1] that will be identified as the limiting process. We describe this generator both on a core of polynomials and on its full domain. However, the core description is sufficient for the analysis that will follow.

Let 𝒫\mathcal{P} denote the space of polynomials on [0,1][0,1] equipped with the supremum norm. We will study the operator :𝒫𝒫\mathcal{B}\colon\mathcal{P}\to\mathcal{P} and the functional η:𝒫\eta\colon\mathcal{P}\to\mathbb{R} given by

(f)(x)=x(1x)f′′(x)αf(x),x[0,1],(\mathcal{B}f)(x)=x(1-x)f^{\prime\prime}(x)-\alpha f^{\prime}(x),\quad x\in[0,1],

and

η(f)\displaystyle\eta(f) :=01(f(x)f(0))xα1(1x)α1𝑑x\displaystyle\hskip 1.0pt\raisebox{0.4pt}{:}=\int_{0}^{1}(f(x)-f(0))x^{-\alpha-1}(1-x)^{\alpha-1}\,dx
=01f(x)xα(1x)αα1𝑑x.\displaystyle=\int_{0}^{1}f^{\prime}(x)x^{-\alpha}(1-x)^{\alpha}\alpha^{-1}\,dx. (6)

Letting ={0,1,2,},\mathbb{N}=\{0,1,2,\ldots\}, we define a family of polynomials {hn}n{1}\{h_{n}\}_{n\in\mathbb{N}\setminus\{1\}} by

hn(x)=s=0nxs(1)ns(n1)ss!(sα)ns(ns)!,x[0,1].h_{n}(x)=\sum_{s=0}^{n}\,x^{s}(-1)^{n-s}\frac{(n-1)_{s}}{s!}\frac{(s-\alpha)_{n-s}}{(n-s)!},\qquad x\in[0,1].

Note that h01h_{0}\equiv 1 and hnh_{n} has degree nn. Moreover, these polynomials are related to the Jacobi polynomials Pn(a,b)P_{n}^{(a,b)} and the shifted Jacobi polynomials Jn(a,b)J_{n}^{(a,b)} [19, 24] by the identity

hn(x)=Jn(α1,α1)(x)=Pn(α1,α1)(2x1),x[0,1].h_{n}(x)=J_{n}^{(\alpha-1,-\alpha-1)}(x)=P_{n}^{(\alpha-1,-\alpha-1)}(2x-1),\qquad x\in[0,1].
Proposition 4.1.

Let =kerη\mathcal{H}=\ker\eta and ωn=n(n1)\omega_{n}=-n(n-1) for n{1}.n\in\mathbb{N}\setminus\{1\}. The following statements hold:

  1. (i)

    hn=ωnhn\mathcal{B}h_{n}=\omega_{n}h_{n} for all n{1}n\in\mathbb{N}\setminus\{1\},

  2. (ii)

    the family {hn}n{1}\{h_{n}\}_{n\in\mathbb{N}\setminus\{1\}} is a Hamel basis for \mathcal{H}, and

  3. (iii)

    \mathcal{H} is a dense subspace of C[0,1]C[0,1].

Proof.

The claim in (i) can be obtained from the classical theory of Jacobi polynomials (e.g. (4.1.3), (4.21.2), and (4.21.4) in [24]).

Noting that hnh_{n} has degree nn shows that the family {hn}n{1}\{h_{n}\}_{n\in\mathbb{N}\setminus\{1\}} is independent. Since h01h_{0}\equiv 1, it clearly lies in \mathcal{H}. To see that the other hnh_{n} also lie in \mathcal{H}, we use (i) to identify them as elements in the range of \mathcal{B} and observe that this range lies in \mathcal{H}. Indeed, this can be verified using (4): for f𝒫f\in\mathcal{P}, we have that

η(f)\displaystyle\eta(\mathcal{B}f) =01(x(1x)f′′(x)αf(x)+αf(0))xα1(1x)α1𝑑x\displaystyle=\int_{0}^{1}(x(1-x)f^{\prime\prime}(x)-\alpha f^{\prime}(x)+\alpha f^{\prime}(0))x^{-\alpha-1}(1-x)^{\alpha-1}\,dx
=01f′′(x)xα(1x)α𝑑xα01(f(x)f(0))xα1(1x)α1𝑑x\displaystyle=\int_{0}^{1}f^{\prime\prime}(x)x^{-\alpha}(1-x)^{\alpha}\,dx-\alpha\int_{0}^{1}(f^{\prime}(x)-f^{\prime}(0))x^{-\alpha-1}(1-x)^{\alpha-1}\,dx
=αη(f)αη(f)\displaystyle=\alpha\,\eta(f^{\prime})-\alpha\,\eta(f^{\prime})
=0.\displaystyle=0.

To obtain equality from the containment span{hn}n{1},\text{span}\{h_{n}\}_{n\in\mathbb{N}\setminus\{1\}}\subset\mathcal{H}, we observe that the former space is a maximal subspace of 𝒫\mathcal{P} (it has codimension one) while the latter is a proper subspace of 𝒫\mathcal{P}.

The claim in (iii) will follow from showing that η\eta is not continuous (see Chapter 3 Theorem 2 in [1]). To see that this holds, notice that the functions fj(x)=(1x)j,j1,f_{j}(x)=(1-x)^{j},\,j\geq 1, have norm 1 but their images under η\eta are unbounded:

η(fj)\displaystyle\eta(f_{j}) =01jxα(1x)j1+αα1𝑑x\displaystyle=-\int_{0}^{1}jx^{-\alpha}(1-x)^{j-1+\alpha}\alpha^{-1}\,dx
=Γ(1α)Γ(j+α)αΓ(j).\displaystyle=-\frac{\Gamma(1-\alpha)\Gamma(j+\alpha)}{\alpha\,\Gamma(j)}.\qed
Proposition 4.2.

The operator |\mathcal{B}|_{\mathcal{H}} is closable and its closure, |¯\overline{\mathcal{B}|_{\mathcal{H}}}, is the generator of a Feller semigroup on C[0,1]C[0,1].

Proof.

We show that |\mathcal{B}|_{\mathcal{H}} satisfies the conditions of the Hille-Yosida Theorem. For λ>0\lambda>0, Proposition 4.1(i)-(ii) show that the range of λ|\lambda-\mathcal{B}|_{\mathcal{H}} is exactly \mathcal{H}. Proposition 4.1(iii) then tells us that this range, as well as the domain of |\mathcal{B}|_{\mathcal{H}}, is dense in C[0,1]C[0,1].

To establish the positive-maximum principle, suppose that ff\in\mathcal{H} has a nonnegative maximum at y[0,1]y\in[0,1]. If y0y\neq 0, the tools of differential calculus show that (|f)(y)0,(\mathcal{B}\big{|}_{\!{}_{\mathcal{H}}}f)(y)\leq 0, as desired. When y=0y=0, consider the element FL1[0,1]F\in L^{1}[0,1] given by

F(x)=(f(x)f(0))xα1(1x)α1F(x)=(f(x)-f(0))x^{-\alpha-1}(1-x)^{\alpha-1}

almost everywhere. Since f(x)f(0)f(x)\leq f(0) on [0,1][0,1], the norm of FF is given by

F1\displaystyle\|F\|_{1} =01|f(x)f(0)|xα1(1x)α1𝑑x\displaystyle=\int_{0}^{1}|f(x)-f(0)|x^{-\alpha-1}(1-x)^{\alpha-1}\,dx
=01(f(x)f(0))xα1(1x)α1𝑑x\displaystyle=-\int_{0}^{1}(f(x)-f(0))x^{-\alpha-1}(1-x)^{\alpha-1}\,dx
=η(f).\displaystyle=-\eta(f).

Recalling that f=kerηf\in\mathcal{H}=\ker\eta, it follows that F=0F=0 almost everywhere. Together with the continuity of ff, this implies that ff(0),f\equiv f(0), and consequently, (|f)(y)0.(\mathcal{B}\big{|}_{\!{}_{\mathcal{H}}}f)(y)\leq 0.

The final result in this section is the explicit description of the generator |¯\overline{\mathcal{B}|_{\mathcal{H}}} and its domain Dom(|¯)\text{Dom}(\overline{\mathcal{B}|_{\mathcal{H}}}).

To begin, we define an operator ^:C[0,1]C2(0,1)C(0,1)\hat{\mathcal{L}}\colon C[0,1]\cap C^{2}(0,1)\to C(0,1) by

^f(x)=x(1x)f′′(x)αf(x).\hat{\mathcal{L}}f(x)=x(1-x)f^{\prime\prime}(x)-\alpha f^{\prime}(x).

We will write ^fC[0,1]\hat{\mathcal{L}}f\in C[0,1] whenever ^f\hat{\mathcal{L}}f can be continuously extended to [0,1][0,1]. Recalling the definition of \mathcal{L} and 𝒟\mathcal{D} from Theorem 1.1, we see that \mathcal{L} is the restriction of ^\hat{\mathcal{L}} to 𝒟\mathcal{D}. We also define functions m:(0,1]m\colon(0,1]\to\mathbb{R} and s:(0,1]s\colon(0,1]\to\mathbb{R} by

m(x)=1xt1α(1t)α1𝑑t=α1xα(1x)αm(x)=\int_{1}^{x}t^{-1-\alpha}(1-t)^{\alpha-1}\,dt=-\alpha^{-1}x^{-\alpha}(1-x)^{\alpha}

and

s(x)=1xtα(1t)α𝑑t.s(x)=\int_{1}^{x}t^{\alpha}(1-t)^{-\alpha}\,dt.

Note that ^\hat{\mathcal{L}} admits the factorization

^f=1m(fs),\hat{\mathcal{L}}f=\frac{1}{m^{\prime}}\left(\frac{f^{\prime}}{s^{\prime}}\right)^{\prime},

from which we obtain the formula

f(x)f(c)=f(c)s(c)(s(x)s(c))+cxcy^f(z)m(z)𝑑zs(y)𝑑y,x,c(0,1).f(x)-f(c)=\frac{f^{\prime}(c)}{s^{\prime}(c)}(s(x)-s(c))+\int_{c}^{x}\int_{c}^{y}\hat{\mathcal{L}}f(z)m^{\prime}(z)dz\,s^{\prime}(y)dy,\quad x,c\in(0,1). (7)

Another identity that will be useful is

1ym(z)𝑑zs(y)=m(y)s(y)=α1,y(0,1).\int_{1}^{y}m^{\prime}(z)dz\,s^{\prime}(y)=m(y)\,s^{\prime}(y)=-\alpha^{-1},\qquad y\in(0,1). (8)
Proposition 4.3.

The identity |¯=\overline{\mathcal{B}|_{\mathcal{H}}}=\mathcal{L} holds, where \mathcal{L} is as defined in Theorem 1.1.

Proof.

We begin by showing that the following holds:

f(x)f(1)=1x1yf(z)m(z)𝑑zs(y)𝑑y,f𝒟,x[0,1].f(x)-f(1)=\int_{1}^{x}\int_{1}^{y}\mathcal{L}f(z)m^{\prime}(z)dz\,s^{\prime}(y)dy,\quad f\in\mathcal{D},\,x\in[0,1]. (9)

To do this, we will take limits in (7). First we take the limit c1c\to 1. The term f(c)s(c)\frac{f^{\prime}(c)}{s^{\prime}(c)} converges to zero due to (D3) (see Theorem 1.1). The limit of the integral is handled by the dominated convergence theorem – a suitable bound follows from the boundedness of f\mathcal{L}f and (8). This establishes the formula for x(0,1)x\in(0,1). Taking now the limit x0x\to 0 (the dominated convergence theorem can be applied as before) establishes the x=0x=0 case. The x=1x=1 case is trivial.

Now we show that Dom(|¯)𝒟.\text{Dom}(\overline{\mathcal{B}|_{\mathcal{H}}})\subset\mathcal{D}. Fixing fDom(|¯)f\in\text{Dom}(\overline{\mathcal{B}|_{\mathcal{H}}}), there exists a sequence {fn}n1\{f_{n}\}_{n\geq 1} of functions in \mathcal{H} such that

fnfandfn|¯f.f_{n}\longrightarrow f\qquad\text{and}\qquad\mathcal{B}f_{n}\longrightarrow\overline{\mathcal{B}|_{\mathcal{H}}}f. (10)

Noting that fn𝒟f_{n}\in\mathcal{D} for all nn, we can apply (9). In this case, the identity fn=fn\mathcal{B}f_{n}=\mathcal{L}f_{n} yields

fn(x)fn(1)=1x1yfn(z)m(z)𝑑zs(y)𝑑y,x[0,1].f_{n}(x)-f_{n}(1)=\int_{1}^{x}\int_{1}^{y}\mathcal{B}f_{n}(z)m^{\prime}(z)dz\,s^{\prime}(y)dy,\quad x\in[0,1]. (11)

Using (10) and the dominated convergence theorem, we can take the limit nn\to\infty above. A suitable bound follows from the boundedness of the sequence {fn}\{\mathcal{B}f_{n}\} and (8). We obtain

f(x)f(1)=1x1y|¯f(z)m(z)𝑑zs(y)𝑑y,x[0,1].f(x)-f(1)=\int_{1}^{x}\int_{1}^{y}\overline{\mathcal{B}|_{\mathcal{H}}}f(z)m^{\prime}(z)dz\,s^{\prime}(y)dy,\quad x\in[0,1]. (12)

Together with the fact that |¯fC(0,1),\overline{\mathcal{B}|_{\mathcal{H}}}f\in C(0,1), mC1(0,1)m\in C^{1}(0,1) and sC2(0,1),s\in C^{2}(0,1), this expression implies that fC2(0,1)f\in C^{2}(0,1). Differentiating the expression yields the identity

|¯f=1m(fs)=^f on (0,1).\overline{\mathcal{B}|_{\mathcal{H}}}f=\frac{1}{m^{\prime}}\left(\frac{f^{\prime}}{s^{\prime}}\right)^{\prime}=\hat{\mathcal{L}}f\quad\text{ on }(0,1). (13)

This shows that ff satisfies (D1). To obtain (D2), we recall that

01(fn(x)fn(0))xα1(1x)α1𝑑x=0\int_{0}^{1}(f_{n}(x)-f_{n}(0))x^{-\alpha-1}(1-x)^{\alpha-1}\,dx=0

for all nn and extend this to ff by taking the limit nn\to\infty. Once again, we apply the dominated convergence theorem. A preliminary bound can be obtained from (8) and (11):

|x1(fn(x)fn(0))|\displaystyle\left|x^{-1}(f_{n}(x)-f_{n}(0))\right| =x1|0x1yfn(z)m(z)𝑑zs(y)𝑑y|\displaystyle=x^{-1}\left|\int_{0}^{x}\int_{1}^{y}\mathcal{B}f_{n}(z)m^{\prime}(z)dz\,s^{\prime}(y)dy\right|
x1fn0xy1m(z)𝑑zs(y)𝑑y\displaystyle\leq x^{-1}\|\mathcal{B}f_{n}\|\int_{0}^{x}\int_{y}^{1}m^{\prime}(z)dz\,s^{\prime}(y)dy
=fnα1.\displaystyle=\|\mathcal{B}f_{n}\|\alpha^{-1}.

The boundedness of the sequence {fn}\{\mathcal{B}f_{n}\} then provides a suitable bound.

To obtain (D3), we differentiate (12) and compute

|f(x)s(x)|\displaystyle\left|\frac{f^{\prime}(x)}{s^{\prime}(x)}\right| =|1x|¯f(z)m(z)𝑑z|\displaystyle=\left|\int_{1}^{x}\overline{\mathcal{B}|_{\mathcal{H}}}f(z)m^{\prime}(z)dz\,\right|
|¯fx1m(z)𝑑z\displaystyle\leq\|\overline{\mathcal{B}|_{\mathcal{H}}}f\|\int_{x}^{1}m^{\prime}(z)dz\,
=|¯f(m(x))\displaystyle=\|\overline{\mathcal{B}|_{\mathcal{H}}}f\|(-m(x))
x10.\displaystyle\xrightarrow[x\to 1]{}0.

We have shown that Dom(|¯)𝒟\text{Dom}(\overline{\mathcal{B}|_{\mathcal{H}}})\subset\mathcal{D} and |¯=\overline{\mathcal{B}|_{\mathcal{H}}}=\mathcal{L} on Dom(|¯)\text{Dom}(\overline{\mathcal{B}|_{\mathcal{H}}}) (see (13)). Therefore, it only remains to show that Dom(|¯)=𝒟.\text{Dom}(\overline{\mathcal{B}|_{\mathcal{H}}})=\mathcal{D}. From Lemma 19.12 in [13], it suffices to show that \mathcal{L} satisfies the positive maximum principle. To this end, suppose that f𝒟f\in\mathcal{D} has a nonnegative maximum at y[0,1]y\in[0,1]. If y1y\neq 1, then the desired inequality can be obtained as in Proposition 4.2. If y=1y=1, we use (D1), L’Hôpital’s rule, (D3), and (8) to establish the existence of limits

f(1)\displaystyle\mathcal{L}f(1) =limx1f(x)\displaystyle=\lim_{x\to 1}\mathcal{L}f(x)
=limx11m(x)(fs)(x)\displaystyle=\lim_{x\to 1}\frac{1}{m^{\prime}(x)}\left(\frac{f^{\prime}}{s^{\prime}}\right)^{\prime}\big{(}x\big{)}
=limx11m(x)f(x)s(x)\displaystyle=\lim_{x\to 1}\frac{1}{m(x)}\frac{f^{\prime}(x)}{s^{\prime}(x)}
=limx1αf(x)\displaystyle=\lim_{x\to 1}-\alpha f^{\prime}(x)
=limx1αf(x)f(1)x1\displaystyle=\lim_{x\to 1}-\alpha\,\frac{f(x)-f(1)}{x-1}
=αf(1).\displaystyle=-\alpha f^{\prime}(1).

Noticing that f(1)0f^{\prime}(1)\geq 0 concludes the proof. ∎

Generator Relations

In this section, we show that our generators satisfy the commutation relations appearing in Theorem 3.1. Here, we rely on an alternative description of the limiting generator in terms of Bernstein polynomials.

For k0k\geq 0, let 𝒫k\mathcal{P}_{k} be the subspace of 𝒫\mathcal{P} consisting of polynomials with degree at most kk. Similarly, define

k=𝒫k,k0.\mathcal{H}_{k}=\mathcal{H}\cap\mathcal{P}_{k},\qquad k\geq 0.

Recall the Bernstein polynomials

bi,k(x)=(ki)xi(1x)ki,i,k0.b_{i,k}(x)=\binom{k}{i}x^{i}(1-x)^{k-i},\quad i\in\mathbb{Z},\,k\geq 0.

Note that bi,k0b_{i,k}\equiv 0 whenever i<0i<0 or i>ki>k. For each k0k\geq 0, the collection {bi,k}i=0k\{b_{i,k}\}_{i=0}^{k} forms a basis of 𝒫k\mathcal{P}_{k} and a partition of unity – that is, i=0kbi,k1.\sum_{i=0}^{k}b_{i,k}\equiv 1. We also have the relations

bi,k\displaystyle b_{i,k}^{\prime} =k(bi1,k1bi,k1),\displaystyle=k(b_{i-1,k-1}-b_{i,k-1}), (14)
bi,k\displaystyle b_{i,k} =k+1ik+1bi,k+1+i+1k+1bi+1,k+1,\displaystyle=\tfrac{k+1-i}{k+1}\,b_{i,k+1}+\tfrac{i+1}{k+1}\,b_{i+1,k+1}, (15)

and

x(1x)bi,k=(i+1)(k+1i)(k+1)(k+2)bi+1,k+2,x(1-x)\,b_{i,k}=\tfrac{(i+1)(k+1-i)}{(k+1)(k+2)}\,b_{i+1,k+2}, (16)

which hold whenever the relevant quantities are defined.

For n1n\geq 1, we define a transition kernel from [0,1][0,1] to [n][n] by

Kn(x,i)=bi,n(x)+νn(α,α)(i)b0,n(x).K_{n}(x,i)=b_{i,n}(x)+\nu^{(\alpha,\alpha)}_{n}(i)b_{0,n}(x).
Proposition 5.1.

Let n1n\geq 1. As an operator from C([n])C([n]) to C[0,1]C[0,1], KnK_{n} is injective and

n\displaystyle\mathcal{H}_{n} ={j=0ncjbj,n:c0,,cn,c0=j=1nνn(α,α)(j)cj}\displaystyle=\Bigg{\{}\sum_{j=0}^{n}c_{j}b_{j,n}:c_{0},\ldots,c_{n}\in\mathbb{R},\,\,\,c_{0}=\sum_{j=1}^{n}\nu^{(\alpha,\alpha)}_{n}(j)c_{j}\Bigg{\}} (17)
=range Kn.\displaystyle=\text{range }K_{n}. (18)
Proof.

Let n1n\geq 1. From the independence of the Bernstein polynomials and the identity

range Kn=span{bi,n(x)+νn(α,α)(i)b0,n(x)}i=1n,\text{range }K_{n}=\text{span}\big{\{}b_{i,n}(x)+\nu^{(\alpha,\alpha)}_{n}(i)b_{0,n}(x)\big{\}}_{i=1}^{n},

it follows that the range of KnK_{n} is an nn-dimensional space. As a result, KnK_{n} is injective. Observing that the right hand side of (17) has dimension at most nn and contains the range of KnK_{n}, it follows that these two spaces are equal. Since n\mathcal{H}_{n} also has dimension nn (see Proposition 4.1(ii)), it only remains to show that the range of KnK_{n} is contained in n\mathcal{H}_{n}. The containment in 𝒫n\mathcal{P}_{n} is clear. For the containment in \mathcal{H}, we simply compute, for i[n]i\in[n],

η(bi,n(x)+νn(α,α)(i)b0,n(x))\displaystyle\eta(b_{i,n}(x)+\nu^{(\alpha,\alpha)}_{n}(i)b_{0,n}(x))
=(ni)01xiα1(1x)ni+α1𝑑xnα1νn(α,α)(i)01xα(1x)n1+α𝑑x\displaystyle\hskip 22.76219pt=\binom{n}{i}\int_{0}^{1}x^{i-\alpha-1}(1-x)^{n-i+\alpha-1}\,dx-n\alpha^{-1}\nu^{(\alpha,\alpha)}_{n}(i)\int_{0}^{1}x^{-\alpha}(1-x)^{n-1+\alpha}\,dx
=(ni)Γ(iα)Γ(ni+α)Γ(n)nα1νn(α,α)(i)Γ(1α)Γ(n+α)Γ(n+1)\displaystyle\hskip 22.76219pt=\binom{n}{i}\frac{\Gamma(i-\alpha)\Gamma(n-i+\alpha)}{\Gamma(n)}-n\alpha^{-1}\nu^{(\alpha,\alpha)}_{n}(i)\frac{\Gamma(1-\alpha)\Gamma(n+\alpha)}{\Gamma(n+1)}
=0.\displaystyle\hskip 22.76219pt=0.\qed
Proposition 5.2.

The action of \mathcal{B} on the Bernstein polynomials is given by

bi,n=n(n+1)k=0n(rk,i𝟙(k=i))bk,n,0in.\mathcal{B}b_{i,n}=n(n+1)\sum_{k=0}^{n}(r_{k,i}-\mathbbm{1}(k=i))\,b_{k,n},\qquad 0\leq i\leq n.
Proof.

Let n2n\geq 2 and 0in.0\leq i\leq n. Applying (14) twice, we see that

bi,n′′\displaystyle b_{i,n}^{\prime\prime} =n(bi1,n1bi,n1)\displaystyle=n(b_{i-1,n-1}^{\prime}-b_{i,n-1}^{\prime})
=n(n1)(bi2,n22bi1,n2+bi,n2).\displaystyle=n(n-1)(b_{i-2,n-2}-2b_{i-1,n-2}+b_{i,n-2}).

Applying now (16), we have that

x(1x)bi,n′′(x)=n(n1)((i1)(n+1i)(n1)nbi1,n(x)2i(ni)(n1)nbi,n(x)+(i+1)(n1i)(n1)nbi+1,n(x))=(i1)(n+1i)bi1,n(x)2i(ni)bi,n(x)+(i+1)(n1i)bi+1,n(x)\displaystyle\begin{split}&x(1-x)b_{i,n}^{\prime\prime}(x)\\ &\hskip 18.49428pt=n(n-1)\Big{(}\tfrac{(i-1)(n+1-i)}{(n-1)n}\,b_{i-1,n}(x)-\tfrac{2i(n-i)}{(n-1)n}\,b_{i,n}(x)+\tfrac{(i+1)(n-1-i)}{(n-1)n}\,b_{i+1,n}(x)\Big{)}\\ &\hskip 18.49428pt=(i-1)(n+1-i)\,b_{i-1,n}(x)-2i(n-i)\,b_{i,n}(x)+(i+1)(n-1-i)\,b_{i+1,n}(x)\end{split} (19)

Using (14) and (15), we find that

bi,n=n(bi1,n1bi,n1)=n(n+1inbi1,n+inbi,nninbi,ni+1nbi+1,n)=(n+1i)bi1,n+(2in)bi,n(i+1)bi+1,n.\displaystyle\begin{split}b_{i,n}^{\prime}&=n(b_{i-1,n-1}-b_{i,n-1})\\ &=n\Big{(}\tfrac{n+1-i}{n}\,b_{i-1,n}+\tfrac{i}{n}\,b_{i,n}-\tfrac{n-i}{n}\,b_{i,n}-\tfrac{i+1}{n}\,b_{i+1,n}\Big{)}\\ &=(n+1-i)\,b_{i-1,n}+(2i-n)\,b_{i,n}-(i+1)\,b_{i+1,n}.\end{split} (20)

As a result,

bi,n\displaystyle\mathcal{B}b_{i,n} =(i1α)(n+1i)bi1,n(α(2in)+2i(ni))bi,n\displaystyle=(i-1-\alpha)(n+1-i)\,b_{i-1,n}-(\alpha(2i-n)+2i(n-i))\,b_{i,n}
+(i+1)(n1i+α)bi+1,n\displaystyle\quad+(i+1)(n-1-i+\alpha)\,b_{i+1,n}
=n(n+1)(ri1,ibi1,n+(ri,i1)bi,n+ri+1,ibi+1,n)\displaystyle=n(n+1)\left(r_{i-1,i}\,b_{i-1,n}+(r_{i,i}-1)\,b_{i,n}+r_{i+1,i}\,b_{i+1,n}\right)
=n(n+1)k=i1i+1(rk,i𝟙(k=i))bk,n.\displaystyle=n(n+1)\sum_{k=i-1}^{i+1}(r_{k,i}-\mathbbm{1}(k=i))\,b_{k,n}.

Recalling that rk,i𝟙(k=i)r_{k,i}-\mathbbm{1}(k=i) is zero unless i1ki+1i-1\leq k\leq i+1 and bk,n0b_{k,n}\equiv 0 unless 0kn0\leq k\leq n, we can change the lower and upper limits of the sum to 0 and nn, respectively. This establishes the n2n\geq 2 case. When n=1n=1, we observe that (20) still holds and the first and last quantities of (LABEL:scaledBernsteinSecondDerivativeExpansion) are still equal. When n=0n=0, the claim is trivial. ∎

Proposition 5.3.

For n1n\geq 1, the following relation holds on C([n])C([n]):

Kn=Knn(n+1)(Qn(α,0)𝟏).\mathcal{B}K_{n}=K_{n}\,n(n+1)(Q_{n}^{(\alpha,0)}-\mathbf{1}).
Proof.

Let n1n\geq 1 and i[n]i\in[n]. Define ei:[n]e_{i}\colon[n]\to\mathbb{R} by ei=𝟙(i=).e_{i}=\mathbbm{1}(i=\cdot). From Proposition 5.2, we have that

n1(n+1)1Knei\displaystyle n^{-1}(n+1)^{-1}\mathcal{B}K_{n}e_{i}
=n1(n+1)1(bi,n+b0,nνn(α,α)(i))\displaystyle\qquad=n^{-1}(n+1)^{-1}\mathcal{B}(b_{i,n}+b_{0,n}\nu^{(\alpha,\alpha)}_{n}(i))
=k=0n(rk,i𝟙(k=i)+νn(α,α)(i)(rk,0𝟙(k=0)))bk,n\displaystyle\qquad=\sum_{k=0}^{n}(r_{k,i}-\mathbbm{1}(k=i)+\nu^{(\alpha,\alpha)}_{n}(i)(r_{k,0}-\mathbbm{1}(k=0)))\,b_{k,n}
=(r0,i+νn(α,α)(i)(r0,01))b0,n+k=1n(rk,i𝟙(k=i)+νn(α,α)(i)r1,0𝟙(k=1))bk,n.\displaystyle\qquad=(r_{0,i}+\nu^{(\alpha,\alpha)}_{n}(i)(r_{0,0}-1))\,b_{0,n}+\sum_{k=1}^{n}(r_{k,i}-\mathbbm{1}(k=i)+\nu^{(\alpha,\alpha)}_{n}(i)r_{1,0}\mathbbm{1}(k=1))\,b_{k,n}.

On the other hand, Proposition 2.4 gives us that

Kn(Qn(α,0)𝟏)ei\displaystyle K_{n}(Q_{n}^{(\alpha,0)}-\mathbf{1})e_{i} =k=1n(bk,n+b0,nνn(α,α)(k))((Qn(α,0)𝟏)ei)(k)\displaystyle=\sum_{k=1}^{n}(b_{k,n}+b_{0,n}\nu^{(\alpha,\alpha)}_{n}(k))((Q_{n}^{(\alpha,0)}-\mathbf{1})e_{i})(k)
=k=1n(bk,n+b0,nνn(α,α)(k))(Qn(α,0)𝟏)(k,i)\displaystyle=\sum_{k=1}^{n}(b_{k,n}+b_{0,n}\nu^{(\alpha,\alpha)}_{n}(k))(Q_{n}^{(\alpha,0)}-\mathbf{1})(k,i)
=k=1n(bk,n+b0,nνn(α,α)(k))(rk,i𝟙(i=k)+νn(α,α)(i)r1,0𝟙(k=1)).\displaystyle=\sum_{k=1}^{n}(b_{k,n}+b_{0,n}\nu^{(\alpha,\alpha)}_{n}(k))(r_{k,i}-\mathbbm{1}(i=k)+\nu^{(\alpha,\alpha)}_{n}(i)r_{1,0}\mathbbm{1}(k=1)).

To show that the two expressions are equal, it will suffice to show that the coefficients of bk,nb_{k,n} are the same in each. For k1k\geq 1, this is immediate. For k=0k=0, we observe that each of the above functions lies in n\mathcal{H}_{n} (see Proposition 4.1 and (18)) and apply (17). ∎

The Convergence Argument

In this section, we verify the convergence condition appearing in Theorem 3.1. We rely on a description of the inverse of the transition operator KnK_{n} in terms of a variant of the Bernstein polynomials.

These variants fall into the class of degenerate Bernstein polynomials [14] and are given by

bi,k,n(x)=(ki)(nx)i(nnx)(ki)nk,0ikn.b_{i,k,n}^{*}(x)=\binom{k}{i}\frac{(nx)^{\downarrow i}(n-nx)^{\downarrow(k-i)}}{n^{\downarrow k}},\quad 0\leq i\leq k\leq n.
Proposition 6.1.

For ki0k\geq i\geq 0, we have the expansions

bi,k=j=0nbi,k,n(jn)bj,n,nk.b_{i,k}=\sum_{j=0}^{n}b_{i,k,n}^{*}\big{(}\tfrac{j}{n}\big{)}b_{j,n},\quad n\geq k.
Proof.

The expansions of a Bernstein polynomial in the Bernstein bases are given in Equation (2) in [19]. Let us verify that the coefficients in those expansions match the coefficients in the above expansions. Fix nki0n\geq k\geq i\geq 0. The coefficient of bj,nb_{j,n} in the above expansion is given by

bi,k,n(jn)=(ki)ji(nj)(ki)nk.b_{i,k,n}^{*}\left(\frac{j}{n}\right)=\binom{k}{i}\frac{j^{\downarrow i}(n-j)^{\downarrow(k-i)}}{n^{\downarrow k}}.

When j<ij<i or j>nk+ij>n-k+i, it is clear that this coefficient is zero. If instead ijnk+ii\leq j\leq n-k+i, this coefficient is reduces to

(ki)ji(nj)(ki)nk\displaystyle\binom{k}{i}\frac{j^{\downarrow i}(n-j)^{\downarrow(k-i)}}{n^{\downarrow k}} =(ki)j!(ji)!(nj)!(njk+i)!n!(nk)!\displaystyle=\binom{k}{i}\frac{\frac{j!}{(j-i)!}\frac{(n-j)!}{(n-j-k+i)!}}{\frac{n!}{(n-k)!}}
=(ki)(nk)!(ji)!(njk+i)!n!j!(nj)!\displaystyle=\binom{k}{i}\frac{\frac{(n-k)!}{(j-i)!(n-j-k+i)!}}{\frac{n!}{j!(n-j)!}}
=(ki)(nkji)(nj).\displaystyle=\binom{k}{i}\frac{\binom{n-k}{j-i}}{\binom{n}{j}}.

In either case, this coefficient agrees with the coefficient in [19]. ∎

Let ιn:[n][0,1]\iota_{n}\colon[n]\to[0,1] be defined by jjnj\mapsto\frac{j}{n} and ρn:C[0,1]C[n]\rho_{n}\colon C[0,1]\to C[n] be the associated projection, ffιn.f\mapsto f\circ\iota_{n}.

Proposition 6.2.

For nki1n\geq k\geq i\geq 1, we have the identity

Knρn(bi,k,n+νk(α,α)(i)b0,k,n)=bi,k+νk(α,α)(i)b0,k.K_{n}\rho_{n}(b_{i,k,n}^{*}+\nu^{(\alpha,\alpha)}_{k}(i)b_{0,k,n}^{*})=b_{i,k}+\nu^{(\alpha,\alpha)}_{k}(i)b_{0,k}.
Proof.

It follows from definition that

Knρn(bi,k,n+νk(α,α)(i)b0,k,n)\displaystyle K_{n}\rho_{n}(b_{i,k,n}^{*}+\nu^{(\alpha,\alpha)}_{k}(i)b_{0,k,n}^{*}) =j=1n(bj,n+νn(α,α)(j)b0,n)(bi,k,n(jn)+νk(α,α)(i)b0,k,n(jn)).\displaystyle=\sum_{j=1}^{n}(b_{j,n}+\nu^{(\alpha,\alpha)}_{n}(j)b_{0,n})\big{(}b_{i,k,n}^{*}\big{(}\tfrac{j}{n}\big{)}+\nu^{(\alpha,\alpha)}_{k}(i)b_{0,k,n}^{*}\big{(}\tfrac{j}{n}\big{)}\big{)}.

Meanwhile, Proposition 6.1 gives us the expansion

bi,k+νk(α,α)(i)b0,k=j=0n(bi,k,n(jn)+νk(α,α)(i)b0,k,n(jn))bj,n.b_{i,k}+\nu^{(\alpha,\alpha)}_{k}(i)b_{0,k}=\sum_{j=0}^{n}\big{(}b_{i,k,n}^{*}\big{(}\tfrac{j}{n}\big{)}+\nu^{(\alpha,\alpha)}_{k}(i)b_{0,k,n}^{*}\big{(}\tfrac{j}{n}\big{)}\big{)}b_{j,n}.

Upon comparison, we find that the coefficient of bj,nb_{j,n} is the same in both expressions whenever j1j\geq 1. Since both functions lie in n\mathcal{H}_{n}, the coefficients of b0,nb_{0,n} must agree as well (see (17)). As a result, the two functions are equal. ∎

Proposition 6.3.

For ki0k\geq i\geq 0, we have the convergence

bi,k,nnbi,k.b_{i,k,n}^{*}\xrightarrow[n\to\infty]{}b_{i,k}.
Proof.

We write

bi,k,n(x)\displaystyle b_{i,k,n}^{*}(x) =(ki)1nkr=0i1(nxr)s=0ki1(nnxs)\displaystyle=\binom{k}{i}\frac{1}{n^{\downarrow k}}\prod_{r=0}^{i-1}(nx-r)\prod_{s=0}^{k-i-1}(n-nx-s)
=(ki)nknkr=0i1(xrn)s=0ki1(1xsn),\displaystyle=\binom{k}{i}\frac{n^{k}}{n^{\downarrow k}}\prod_{r=0}^{i-1}\left(x-\frac{r}{n}\right)\prod_{s=0}^{k-i-1}\left(1-x-\frac{s}{n}\right),

and handle each factor separately. The constants nknk\frac{n^{k}}{n^{\downarrow k}} converge to 11 and each factor in a product converges to either u(x)=xu(x)=x or v(x)=1xv(x)=1-x. ∎

Proposition 6.4.

Let ff\in\mathcal{H} and fix m1m\geq 1 such that fmf\in\mathcal{H}_{m}. Then we have the convergence

(Kn1ρn)fnnm0(K_{n}^{-1}-\rho_{n})f\xrightarrow[n\to\infty]{n\geq m}0

in the sense of Definition 3.1.

Proof.

It suffices to consider the case when f=bi,k+νk(α,α)(i)b0,kf=b_{i,k}+\nu^{(\alpha,\alpha)}_{k}(i)b_{0,k} for some ii and kk satisfying 1ik1\leq i\leq k. Defining fn=bi,k,n+νk(α,α)(i)b0,k,nf_{n}=b_{i,k,n}^{*}+\nu^{(\alpha,\alpha)}_{k}(i)b_{0,k,n}^{*} for n1n\geq 1, it follows from Proposition 6.2 that

(Kn1ρn)f\displaystyle(K_{n}^{-1}-\rho_{n})f =ρn(fnf).\displaystyle=\rho_{n}(f_{n}-f).

Since the ρn\rho_{n} are uniformly bounded, the result follows from Proposition 6.3. ∎

Semigroup Relations from Generator Relations

In this section, we provide general conditions under which commutation relations involving generators lead to the corresponding relations for their semigroups.

Theorem 7.1.

Let AA and BB be the generators of the Feller semigroups VtV_{t} and WtW_{t}, respectively, and let \mathcal{E} and \mathcal{F} denote their respective domains. Suppose that there is a subspace EE\subset\mathcal{E}, a linear operator L:E¯¯L\colon\,\overline{\!E}\to\overline{\mathcal{F}}, and a set I(0,)I\subset(0,\infty) such that

  1. (i)

    LL is bounded,

  2. (ii)

    II is unbounded,

  3. (iii)

    E(λA)EE\subset(\lambda-A)E for λI\lambda\in I, and

  4. (iv)

    LA=BLLA=BL on EE.

Then LVt=WtLLV_{t}=W_{t}L on E¯\,\overline{\!E} for each t0t\geq 0.

Proof.

Fix λI\lambda\in I and let RλAR_{\lambda}^{A} and RλBR_{\lambda}^{B} be the resolvent operators corresponding to AA and BB respectively. It follows from (iii) that EE is invariant under RλAR_{\lambda}^{A}. Combining this with (iv), we obtain the following relation on EE:

RλBL\displaystyle R_{\lambda}^{B}L =RλBL(λA)RλA\displaystyle=R_{\lambda}^{B}L(\lambda-A)R_{\lambda}^{A}
=RλB(λB)LRλA\displaystyle=R_{\lambda}^{B}(\lambda-B)LR_{\lambda}^{A}
=LRλA.\displaystyle=LR_{\lambda}^{A}.

It then follows easily that

Lλ(λRλAI)=λ(λRλBI)Lon E,L\lambda(\lambda R_{\lambda}^{A}-I)=\lambda(\lambda R_{\lambda}^{B}-I)L\quad\text{on }E,

or equivalently, LAλ=BλLLA_{\lambda}=B_{\lambda}L on EE, where AλA_{\lambda} and BλB_{\lambda} are the Yosida approximations of AA and BB respectively. Noting that EE is invariant under AλA_{\lambda}, this extends to nonnegative integers kk:

LAλk=BλkLon E.LA_{\lambda}^{k}=B_{\lambda}^{k}L\quad\text{on }E.

Applying now (i), we have for fEf\in E and t0t\geq 0 the identity

LetAλf\displaystyle Le^{tA_{\lambda}}f =Lk=0tkk!(Aλkf)\displaystyle=L\sum_{k=0}^{\infty}\frac{t^{k}}{k!}(A_{\lambda}^{k}f)
=k=0tkk!(LAλkf)\displaystyle=\sum_{k=0}^{\infty}\frac{t^{k}}{k!}(LA_{\lambda}^{k}f)
=k=0tkk!(BλkLf)\displaystyle=\sum_{k=0}^{\infty}\frac{t^{k}}{k!}(B_{\lambda}^{k}Lf)
=etBλLf.\displaystyle=e^{tB_{\lambda}}Lf.

Letting λ\lambda become arbitrarily large (see (ii)) yields LVtf=WtLf.LV_{t}f=W_{t}Lf. This establishes the result on EE. The extension to E¯\,\overline{\!E} follows from the boundedness of LL. ∎

Corollary 7.1.

Let AA and BB be the generators of the Feller semigroups VtV_{t} and WtW_{t}, respectively, and let \mathcal{E} and \mathcal{F} denote their respective domains. Suppose that there is a subspace EE\subset\mathcal{E}, a linear operator L:E¯L\colon E\to\overline{\!\mathcal{F}}, and a filtration of EE by finite dimensional spaces {Ek}k1\{E_{k}\}_{k\geq 1} such that

  1. (i)

    AEkEkAE_{k}\subset E_{k} for all kk, and

  2. (ii)

    LA=BLLA=BL on EE.

Then LVt=WtLLV_{t}=W_{t}L on EE for each t0t\geq 0.

Proof.

Let k1k\geq 1. It follows from (i) that EkE_{k} is invariant under the injective operators {λA}λ>0\{\lambda-A\}_{\lambda>0}. Together with the fact that EkE_{k} is finite-dimensional, this implies that

(λA)Ek=Ek,λ>0.(\lambda-A)E_{k}=E_{k},\quad\lambda>0.

Letting Lk:Ek¯L_{k}\colon E_{k}\to\,\overline{\!\mathcal{F}} denote the restriction of LL to EkE_{k}, it follows from (i) and (ii) that

LkA=BLkon Ek.L_{k}A=BL_{k}\quad\text{on }E_{k}.

Since EkE_{k} is finite-dimensional, LkL_{k} is bounded and Ek¯=Ek.\,\overline{\!E_{k}}=E_{k}. Applying Theorem 7.1, we find that LVt=WtLLV_{t}=W_{t}L on EkE_{k} for each t0t\geq 0. Taking a union over kk extends the identity to EE. ∎

Proofs of Main Results

Proof of Theorem 1.3.

The first claim was proved in Proposition 2.4. For the second claim, we appeal to Corollary 7.1. We take A=n(n+1)(Qn(α,0)𝟏)A=n(n+1)(Q_{n}^{(\alpha,0)}-\mathbf{1}), B=B=\mathcal{L}, L=KnL=K_{n}, and E=C([n])=EkE=C([n])=E_{k} for all kk. The containment AEkEkAE_{k}\subset E_{k} holds trivially and the identity LA=BLLA=BL was established in Proposition 5.3. Applying Corollary 7.1, we obtain the desired identity in terms of transition operators, which implies the same relation in terms of transition kernels. ∎

Proof of Theorem 1.2.

The claim about the existence of initial distributions for 𝐗n(α,0)\mathbf{X}^{(\alpha,0)}_{n} follows from Theorem 1.3. The second claim follows from applying Theorem 3.1 with E=[0,1]E=[0,1], A=A=\mathcal{L}, D=D=\mathcal{H}, En=[n]E_{n}=[n], Zn=Yn(α,θ)Z_{n}=Y^{(\alpha,\theta)}_{n}, γn(j)=jn\gamma_{n}(j)=\frac{j}{n}, Dn=nD_{n}=\mathcal{H}_{n}, Ln=Kn1L_{n}=K_{n}^{-1}, δn1=n(n+1)\delta_{n}^{-1}=n(n+1), and εn1=n2\varepsilon_{n}^{-1}=n^{2}. To verify that AA is the generator of a conservative Feller semigroup on C[0,1]C[0,1], DD is a core for AA, and DD is invariant under AA, we appeal to Propositions 4.3, 4.2, and 4.1. Condition (a) can be obtained from the identity in Proposition 5.3 by recalling that KnK_{n} is injective (see Proposition 5.1) and that each ff in D=D=\mathcal{H} lies in Dn=nD_{n}=\mathcal{H}_{n} for large nn. Condition (b) is exactly the result of Proposition 6.4. ∎

Proof of Theorem 1.1.

Define ι:𝒞𝒰\iota:\mathcal{C}\to\mathcal{U} by

ι(σ)=(0,σ1|σ|)(σ1|σ|,σ1+σ2|σ|)(|σ|σ(σ)|σ|,1).\iota(\sigma)=\bigg{(}0,\frac{\sigma_{1}}{|\sigma|}\bigg{)}\cup\bigg{(}\frac{\sigma_{1}}{|\sigma|},\frac{\sigma_{1}+\sigma_{2}}{|\sigma|}\bigg{)}\cup\ldots\cup\bigg{(}\frac{|\sigma|-\sigma_{\ell(\sigma)}}{|\sigma|},1\bigg{)}.

From [20, Theorem 1.3], we have that if

ι(𝐗n(α,θ)(0))𝐗(α,θ)(0),\iota(\mathbf{X}^{(\alpha,\theta)}_{n}(0))\Longrightarrow\mathbf{X}^{(\alpha,\theta)}(0),

then

(ι(𝐗n(α,θ)(n2t)))t0(𝐗(α,θ)(t))t0,\left(\iota(\mathbf{X}^{(\alpha,\theta)}_{n}(\lfloor n^{2}t\rfloor))\right)_{t\geq 0}\Longrightarrow\left(\mathbf{X}^{(\alpha,\theta)}(t)\right)_{t\geq 0},

where a\lfloor a\rfloor is the integer part of aa and the convergence is in distribution on the Skorokhod space D([0,),𝒰)D([0,\infty),\mathcal{U}), where the metric on 𝒰\mathcal{U} is given by the Hausdorff distance between the complements (complements being taken in [0,1][0,1]). If ξ\xi were continuous, the result would follow immediately, but ξ\xi is discontinuous. However, it is straightforward to show that if unuu_{n}\to u in 𝒰\mathcal{U} and ξ(un)c>0\xi(u_{n})\to c>0, then ξ(u)=c\xi(u)=c.

Assuming now that 𝐗n(α,0)\mathbf{X}^{(\alpha,0)}_{n} is running in stationarity, the fact that ι(𝐗n(α,0)(0))\iota(\mathbf{X}^{(\alpha,0)}_{n}(0)) converges in distribution to an (α,0)(\alpha,0) Poisson-Dirichlet interval partition distribution follows from [18] and the fact that ϕ(𝐗n(α,0))\phi(\mathbf{X}^{(\alpha,0)}_{n}) is a Markov chain follows from Theorem 1.3. Observe that (p(α,0))n1((1),)(p^{\uparrow}_{(\alpha,0)})^{n-1}((1),\cdot) is the stationary distribution of 𝐗n(α,0)\mathbf{X}^{(\alpha,0)}_{n} and, in the (α,0)(\alpha,0) ordered Chinese Restaurant Process growth step, no new table is ever created at the start of the list. Thus, for every kk, ϕ(𝐗n(α,0)(k))\phi(\mathbf{X}^{(\alpha,0)}_{n}(k)) is distributed like the size of the table containing 11 in the usual (α,0)(\alpha,0) Chinese Restaurant Process after nn customers are seated, see [17]. Consequently, since our chain is stationary, for each tt,

1nϕ(𝐗n(α,0)(n2t))=ξ(ι(𝐗n(α,0)(n2t)))=dξ(ι(𝐗n(α,0)(0)))W,\frac{1}{n}\phi(\mathbf{X}^{(\alpha,0)}_{n}(\lfloor n^{2}t\rfloor))=\xi(\iota(\mathbf{X}^{(\alpha,0)}_{n}(\lfloor n^{2}t\rfloor)))=_{d}\xi(\iota(\mathbf{X}^{(\alpha,0)}_{n}(0)))\Rightarrow W,

where WW has a Beta(1α,α)(1-\alpha,\alpha) distribution, see [17].

Therefore, from Theorem 1.2 with FF as defined there and F(0)=dWF(0)=_{d}W, passing to a subsequence if necessary, and using the Skorokhod representation theorem, we may assume that

((ι(𝐗n(α,0)(n2t)),ξ(ι(𝐗n(α,0)(n2s)))))t,s0a.s.((𝐗(α,0)(t),F(s)))t,s0\left(\left(\iota(\mathbf{X}^{(\alpha,0)}_{n}(\lfloor n^{2}t\rfloor)),\xi(\iota(\mathbf{X}^{(\alpha,0)}_{n}(\lfloor n^{2}s\rfloor)))\right)\right)_{t,s\geq 0}\overset{a.s.}{\longrightarrow}\left((\mathbf{X}^{(\alpha,0)}(t),F(s))\right)_{t,s\geq 0}

in D([0,),𝒰)×D([0,),[0,1])D([0,\infty),\mathcal{U})\times D([0,\infty),[0,1]). Fix t0t\geq 0. Since Feller processes have no fixed discontinuities, FF is almost surely continuous at tt and, therefore, since convergence in D([0,),𝒰)D([0,\infty),\mathcal{U}) implies convergence at continuity points,

ξ(ι(𝐗n(α,0)(n2t)))a.s.F(t).\xi(\iota(\mathbf{X}^{(\alpha,0)}_{n}(\lfloor n^{2}t\rfloor)))\overset{a.s.}{\longrightarrow}F(t).

Since F(t)=dWF(t)=_{d}W, (F(t)>0)=1\mathbb{P}(F(t)>0)=1 and, since

ι(𝐗n(α,0)(n2t))a.s.𝐗(α,0)(t),\iota(\mathbf{X}^{(\alpha,0)}_{n}(\lfloor n^{2}t\rfloor))\overset{a.s.}{\longrightarrow}\mathbf{X}^{(\alpha,0)}(t),

it follows that F(t)=a.s.ξ(𝐗(α,0)(t))F(t)=_{a.s.}\xi(\mathbf{X}^{(\alpha,0)}(t)). Consequently, F(t)F(t) is a modification of ξ(𝐗(α,0)(t))\xi(\mathbf{X}^{(\alpha,0)}(t)) and since FF has a Feller semigroup, so does ξ(𝐗(α,0))\xi(\mathbf{X}^{(\alpha,0)}). ∎

References

  • [1] Béla Bollobás. Linear Analysis: An Introductory Course. Cambridge University Press, 2 edition, 1999.
  • [2] Alexei Borodin and Grigori Olshanski. Infinite-dimensional diffusions as limits of random walks on partitions. Probab. Theory Related Fields, 144(1-2):281–318, 2009.
  • [3] S. N. Ethier and Thomas G. Kurtz. The infinitely-many-neutral-alleles diffusion model. Adv. in Appl. Probab., 13(3):429–452, 1981.
  • [4] Stewart N. Ethier and Thomas G. Kurtz. Markov processes: characterization and convergence. Wiley series in probability and mathematical statistics. J. Wiley & Sons, New York, Chichester, 2005.
  • [5] Noah Forman, Soumik Pal, Douglas Rizzolo, and Matthias Winkel. Interval partition evolutions with emigration related to the Aldous diffusion. arXiv preprint arXiv:1804.01205, 2018.
  • [6] Noah Forman, Soumik Pal, Douglas Rizzolo, and Matthias Winkel. Projections of the Aldous chain on binary trees: intertwining and consistency. Random Structures Algorithms, 57(3):745–769, 2020.
  • [7] Noah Forman, Soumik Pal, Douglas Rizzolo, and Matthias Winkel. Diffusions on a space of interval partitions: Poisson-Dirichlet stationary distributions. Ann. Probab., 49(2):793–831, 2021.
  • [8] Noah Forman, Douglas Rizzolo, Quan Shi, and Matthias Winkel. Diffusions on a space of interval partitions: The two-parameter model. arXiv:2008.02823, 2020.
  • [9] Jason Fulman. Commutation relations and Markov chains. Probab. Theory Related Fields, 144(1-2):99–136, 2009.
  • [10] Jason Fulman. Mixing time for a random walk on rooted trees. Electron. J. Combin., 16(1):Research Paper 139, 13, 2009.
  • [11] Han L. Gan and Nathan Ross. Stein’s method for the Poisson-Dirichlet distribution and the Ewens Sampling Formula, with applications to Wright-Fisher models. arXiv:1910.04976, 2020.
  • [12] Alexander Gnedin and Jim Pitman. Regenerative composition structures. Ann. Probab., 33(2):445–479, 2005.
  • [13] Olav Kallenberg. Foundations of Modern Probability. Probability and its Applications (New York). Springer-Verlag, New York, 2002.
  • [14] Taekyun Kim and Dae san Kim. Degenerate Bernstein polynomials, 2018.
  • [15] L. A. Petrov. A two-parameter family of infinite-dimensional diffusions on the Kingman simplex. Funktsional. Anal. i Prilozhen., 43(4):45–66, 2009.
  • [16] Leonid Petrov. 𝔰𝔩(2){\mathfrak{sl}}(2) operators and Markov processes on branching graphs. J. Algebraic Combin., 38(3):663–720, 2013.
  • [17] J. Pitman. Combinatorial stochastic processes, volume 1875 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 2006. Lectures from the 32nd Summer School on Probability Theory held in Saint-Flour, July 7–24, 2002.
  • [18] Jim Pitman and Matthias Winkel. Regenerative tree growth: binary self-similar continuum random trees and Poisson–Dirichlet compositions. Ann. Probab., 37(5):1999–2041, 2009.
  • [19] Abedallah Rababah. Jacobi-Bernstein basis transformation. Computational Methods in Applied Mathematics, 4:206–214, 06 2004.
  • [20] Kelvin Rivera-Lopez and Douglas Rizzolo. Diffusive limits of two-parameter ordered Chinese Restaurant Process up-down chains, 2020. arXiv:2011.06577.
  • [21] Dane Rogers and Matthias Winkel. A Ray-Knight representation of up-down Chinese restaurants. arXiv:2006.06334, 2020.
  • [22] L. C. G. Rogers and J. W. Pitman. Markov functions. The Annals of Probability, 9(4):573–582, 1981.
  • [23] Quan Shi and Matthias Winkel. Up-down ordered Chinese restaurant processes with two-sided immigration, emigration and diffusion limits. arXiv preprint arXiv:2012.15758, 2020.
  • [24] Gábor Szegö. Orthogonal Polynomials. American Mathematical Society, Providence, RI, 1975.