Processing math: 96%
Research article Special Issues

Continuous dependence of an invariant measure on the jump rate of a piecewise-deterministic Markov process

  • We investigate a piecewise-deterministic Markov process, evolving on a Polish metric space, whose deterministic behaviour between random jumps is governed by some semi-flow, and any state right after the jump is attained by a randomly selected continuous transformation. It is assumed that the jumps appear at random moments, which coincide with the jump times of a Poisson process with intensity λ. The model of this type, although in a more general version, was examined in our previous papers, where we have shown, among others, that the Markov process under consideration possesses a unique invariant probability measure, say νλ. The aim of this paper is to prove that the map λνλ is continuous (in the topology of weak convergence of probability measures). The studied dynamical system is inspired by certain stochastic models for cell division and gene expression.

    Citation: Dawid Czapla, Sander C. Hille, Katarzyna Horbacz, Hanna Wojewódka-Ściążko. Continuous dependence of an invariant measure on the jump rate of a piecewise-deterministic Markov process[J]. Mathematical Biosciences and Engineering, 2020, 17(2): 1059-1073. doi: 10.3934/mbe.2020056

    Related Papers:

    [1] Jiongxuan Zheng, Joseph D. Skufca, Erik M. Bollt . Heart rate variability as determinism with jump stochastic parameters. Mathematical Biosciences and Engineering, 2013, 10(4): 1253-1264. doi: 10.3934/mbe.2013.10.1253
    [2] Simone Göttlich, Stephan Knapp . Modeling random traffic accidents by conservation laws. Mathematical Biosciences and Engineering, 2020, 17(2): 1677-1701. doi: 10.3934/mbe.2020088
    [3] Patrice Bertail, Stéphan Clémençon, Jessica Tressou . A storage model with random release rate for modeling exposure to food contaminants. Mathematical Biosciences and Engineering, 2008, 5(1): 35-60. doi: 10.3934/mbe.2008.5.35
    [4] Abdon ATANGANA, Seda İĞRET ARAZ . Deterministic-Stochastic modeling: A new direction in modeling real world problems with crossover effect. Mathematical Biosciences and Engineering, 2022, 19(4): 3526-3563. doi: 10.3934/mbe.2022163
    [5] Manuel González-Navarrete . Type-dependent stochastic Ising model describing the dynamics of a non-symmetric feedback module. Mathematical Biosciences and Engineering, 2016, 13(5): 981-998. doi: 10.3934/mbe.2016026
    [6] Damilola Olabode, Jordan Culp, Allison Fisher, Angela Tower, Dylan Hull-Nye, Xueying Wang . Deterministic and stochastic models for the epidemic dynamics of COVID-19 in Wuhan, China. Mathematical Biosciences and Engineering, 2021, 18(1): 950-967. doi: 10.3934/mbe.2021050
    [7] Carlos Ramos . Animal movement: symbolic dynamics and topological classification. Mathematical Biosciences and Engineering, 2019, 16(5): 5464-5489. doi: 10.3934/mbe.2019272
    [8] Linda J. S. Allen, Vrushali A. Bokil . Stochastic models for competing species with a shared pathogen. Mathematical Biosciences and Engineering, 2012, 9(3): 461-485. doi: 10.3934/mbe.2012.9.461
    [9] Virginia Giorno, Serena Spina . On the return process with refractoriness for a non-homogeneous Ornstein-Uhlenbeck neuronal model. Mathematical Biosciences and Engineering, 2014, 11(2): 285-302. doi: 10.3934/mbe.2014.11.285
    [10] Tingting Yu, Sanling Yuan . Dynamics of a stochastic turbidostat model with sampled and delayed measurements. Mathematical Biosciences and Engineering, 2023, 20(4): 6215-6236. doi: 10.3934/mbe.2023268
  • We investigate a piecewise-deterministic Markov process, evolving on a Polish metric space, whose deterministic behaviour between random jumps is governed by some semi-flow, and any state right after the jump is attained by a randomly selected continuous transformation. It is assumed that the jumps appear at random moments, which coincide with the jump times of a Poisson process with intensity λ. The model of this type, although in a more general version, was examined in our previous papers, where we have shown, among others, that the Markov process under consideration possesses a unique invariant probability measure, say νλ. The aim of this paper is to prove that the map λνλ is continuous (in the topology of weak convergence of probability measures). The studied dynamical system is inspired by certain stochastic models for cell division and gene expression.


    Piecewise-deterministic Markov processes (PDMPs) originate with M.H.A. Davis [1]. They constitute an important class of Markov processes that is complementary to those defined by stochastic differential equations. PDMPs are encountered as suitable mathematical models for processes in the physical world around us, e.g., in resource allocation and service provisioning (queing, cf. [1]) or biology: as stochastic models for gene expression and autoregulation [2,3], cell division [4], excitable membranes [5] or population dynamics [6,7].

    Mathematical research on PDMPs has been conducted over the years in various directions. Applications in control and optimization have been just one direction. The fundamentals of existence and uniqueness of invariant probability measures for Markov operators and semigroups of Markov operators associated with PDMPs, as well as their asymptotic properties, have attracted much attention. See e.g., [8,9,10,11], where the considered underlying state space is locally compact. The theory for the general case of non-locally compact Polish state space is less developed yet. It is considered e.g., in [3,5,12,13,14,15]. Another direction is that of establishing the validity of the Strong Law of Large Numbers (SLLN), the Central Limit Theorem (CLT) and the Law of the Interated Logarithm (LIL) for non-stationary PDMPs (cf. [16,17,18,19]), which has interest in itself for non-stationary processes in general [20].

    In this paper, we are concerned with a special case of the PDMP described in [13,14], whose deterministic motion between jumps depends on a single continuous semi-flow, and any post-jump location is attained by a continuous transformation of the pre-jump state, randomly selected (with a place-dependent probability) among all possible ones. The jumps in this model occur at random time points according to a homogeneous Poisson process. The random dynamical systems of this type constitute a mathematical framework for certain particular biological models, such as those for gene expression [2] or cell division [4].

    The aim of the paper is to establish the continuous (in the Fortet-Mourier distance, cf. [21,Section 8.3]) dependence of the invariant measure on the rate of a Poisson process determining the frequency of jumps. While the SLLN and the CLT provide the theoretical foundation for successful approximation of the invariant measure by means of observing or simulating (many) sample trajectories of the process, this result asserts the stability of this procedure, at least locally in parameter space. It is a prerequisite for the development of a bifurcation theory. Moreover, even stronger regularity of this dependence on parameter (i.e., differentiability in a suitable norm on the space of measures) would be needed for applications in control theory or for parameter estimation (see e.g., [22]).

    The outline of the paper is as follows. In Section 2, several facts on integrating measure-valued functions and basic definitions from the theory of Markov operators have been compiled. Section 3 deals with the structure and assumptions of the model under study. In Section 4, we establish certain auxiliary results on the transition operator of the Markov chain given by the post-jump locations. More specifically, we show that the operator is jointly continuous (in the topology of weak convergence of measures) as a function of measure and the jump-rate parameter, and that the weak convergence of the distributions of the chain to its unique stationary distribution must be uniform. Section 5 is the essential part of the paper. Here, we establish the announced results on the continuous dependence of the invariant measure on the jump frequency for both, the discrete-time system, constituted by the post jump-locations, and for the PDMP itself.

    Let X be a closed subset of some separable Banach space (H,), endowed with the σ-field BX consisting of its Borel subsets. Further, let (BM(X),) stand for the Banach space of all bounded Borel-measurable functions f:XR with the supremum norm f:=supxX|f(x)|. By BC(X) and BL(X) we shall denote the subspaces of BM(X) consisting of all continuous and all Lipschitz continuous functions, respectively. Let us further introduce

    fBL:=max{f,|f|Lip}for anyfBL(X),

    where

    |f|Lip:=sup{|f(x)f(y)|xy:x,yX,xy}.

    It is well-known (cf. [23,Proposition 1.6.2]) that BL defines a norm in BL(X), for which it is a Banach space.

    In what follows, we will write (Msig(X),TV) for the Banach space of all finite, countably additive functions (signed measures) on BX, endowed with the total variation norm TV, which can be expressed as

    μTV:=|μ|(X)=sup{|f,μ|:fBM(X),f1}forμMsig(X),

    where

    f,μ:=Xf(x)μ(dx)

    and |μ| stands for the absolute variation of μ (cf. e.g., [24]). The symbols M+(X) and M1(X) will be used to denote the subsets of Msig(X), consisting of all non-negative and all probability measures on BX, respectively. Moreover, we will write M1,1(X) for the set of all measures μM1(X) with finite first moment, i.e., satisfying ,μ<.

    Let us now define, for any μMsig(X), the linear functional Iμ:BL(X)R given by

    Iμ(f)=f,μforfBL(X).

    It easy to show that IμBL(X) for every μMsig(X), where BL(X) stands for the dual space of (BL(X),BL) with the operator norm BL given by

    φBL:=sup{|φ(f)|:fBL(X),fBL1}for anyφBL(X).

    Moreover, we have IμBLμTV for any μMsig(X).

    Furthermore, it is well known (see [25,Lemma 6]), that the mapping

    Msig(X)μIμBL(X)

    is injective, and thus the space (Msig(X),TV) may be embedded into (BL(X),BL). This enables us to identify each measure μMsig(X) with the functional IμBL(X). Note that BL induces a norm on Msig(X), called the Fortet-Mourier (or bounded Lipschitz, cf. e.g., [26,27]) norm and denoted by FM. Consequently, we can write

    μFM:=IμBL=sup{|f,μ|:fBL(X),fBL1}for anyμMsig(X).

    As we have already seen, generally μFM=IμBLμTV for any μMsig(X). However, for positive measures the norms coincide, i.e., μFM=μ(X)=μTV for all μM+(X) (cf. [25]).

    Let us now write D(X) and D+(X) for the linear space and the convex cone, respectively, generated by the set {δx:xX}BL(X) of functionals of the form

    δx(f):=f(x)for anyfBL(X),xX,

    which can be also viewed as Dirac measures. It is not hard to check that the BL-closure of D(X) is a separable Banach subspace of BL(X). Moreover, one can show \hbox(cf. [27,Theorems 2.3.8-2.3.19]) that M+(X)=clD+(X) (using the completeness of X), which in turn implies that Msig(X) is a BL-dense subspace of clD(X), i.e., clMsig(X)=clD(X). The key idea underlying the proof of this result is to show that every measure μM+(X) can be represented by the Bochner integral (for definition see e.g., [28) of the continuous map XxδxclD(X),i.e.,

    μ=Xδxμ(dx)clD+(X).

    In particular,it follows that (clMsig(X),BL|clD(X)) is a separable Banach space.

    What is more,according to [27,Theorem 2.3.22], the dual space of clMsig(X)=clD(X) with the operator norm

    κclD:=sup{|κ(φ)|:φclD(X),φBL1},κ[clD(X)],

    is isometrically isomorphic with the space (BL(X),BL), and each functional κ[clD(X)] can be represented by some fBL(X), in the sense that κ(φ)=φ(f) for φclD(X). In particular, we then have κ(μ)=Iμ(f)=f,μ whenever μMsig(X) (by identifying μ with Iμ).

    In view of the above observations, the norm BL is convenient for integrating (in the Bochner sense) measure-valued functions p:EMsig(X), where E is an arbitrary measure space. The Pettis measurability theorem (see e.g., [28,Chapter Ⅱ,Theorem 2]), together with the separability of clMsig(X), ensures that p is strongly measurable as a map with values in clMsig(X) (i.e., it is a pointwise a.e. limit of simple functions) if and only if, for any fBL(X), the functional Etf,p(t)R is measurable. Moreover, we have at our disposal the following result (see [27,Propositions 3.2.3-3.2.5] or [29,Proposition C.2]), which provides a tractable condition guaranteeing the integrability of p and ensuring that the integral is an element of Msig(X):

    Theorem 2.1. Let (E,Σ) be a measurable space endowed with a σ-finite measure ν, and let p:EMsig(X) be a strongly measurable function. Suppose that there exists a real-valued function gL1(E,Σ,ν) such that

    p(t)TVg(t)for a.e.tE.

    Then the following conditions hold:

    (i) The function p is Bochner ν-integrable as a map acting from (E,Σ) to (clMsig(X),BL|clD(X)). Moreover, we have

    Ep(t)ν(dt)TVEp(t)TVν(dt).

    (ii) The Bochner integral Ep(t)ν(dt)clMsig(X) belongs to Msig(X) and satisfies

    (Ep(t)ν(dt))(A)=Ep(t)(A)ν(dt)for anyABX.

    Another crucial observation is that the restriction of the weak topology on Msig(X), generated by BC(X), to M+(X) equals to the topology induced by the norm FM|M+(X) (cf. [25,Theorem 18] or [21,Theorem 8.3.2]). In particular, the following holds:

    Theorem 2.2. Let μn,μM+(X) for every nN. Then limnμnμFM=0 if and only if μnwμ, as n, that is,

    limnf,μn=f,μfor anyfBC(X).

    Let us now recall several basic definitions concerning Markov operators acting on measures. First of all, a function P:X×BX[0,1] is called a stochastic kernel if, for any fixed ABX, xP(x,A) is a Borel-measurable map on X, and, for any fixed xX, AP(x,A) is a probability Borel measure on BX. We can consider two operators corresponding to a stochastic kernel P, namely

    μP(A)=XP(x,A)μ(dx)forμMsig(X),ABX (2.1)

    and

    Pf(x)=Xf(y)P(x,dy)forfBM(X),xX. (2.2)

    The operator ()P:Msig(X)Msig(X), given by (2.1), is called a regular Markov operator. It is easy to check that

    f,μP=Pf,μfor anyfBM(X),μMsig(X),

    and, therefore, P():BM(X)BM(X), defined by (2.2), is said to be the dual operator of ()P.

    A regular Markov operator ()P is said to be Feller if its dual operator P() preserves continuity, that is, PfBC(X) for every fBC(X). A measure μM+(X) is called an invariant measure for ()P whenever μP=μ.

    We will say that the operator ()P is exponentially ergodic in the Fortet-Mourier distance if there exists a unique invariant measure μM1(X) of ()P, for which there is q[0,1) such that, for any μM1,1(X) and some constant C(μ), we have

    μPnμFMC(μ)qnfor anynN.

    The measure μ is then usually called exponentially attracting.

    A regular Markov semigroup (P(t))tR+ is a family of regular Markov operators ()P(t):Msig(X)Msig(X), tR+:=[0,), which form a semigroup (under composition) with the identity transformation ()P(0) as the unity element. Provided that ()P(t) is a Markov-Feller operator for every tR+, the semigroup (P(t))tR+ is said to be Markov-Feller, too. If, for some νMsig(X), νP(t)=ν for every tR+, then ν is called an invariant measure of (P(t))tR+.

    Recall that X is a closed subset of some separable Banach space (H,), and let (Θ,BΘ,ϑ) be a topological measure space with a σ-finite Borel measure ϑ. With a slight abuse of notation, we will further write dθ only, instead of ϑ(dθ).

    Let us consider a PDMP (X(t))tR+, evolving on the space X through random jumps occuring at the jump times τn, nN, of a homogeneous Poisson process with intensity λ>0. The state right after the jump is attained by a transformation wθ:XX, randomly selected from the set {wθ:θΘ}. The probability of choosing wθ is determined by a place-dependent density function θp(x,θ), where x describes the state of the process just before the jump. It is required that the maps (x,θ)p(x,θ) and (x,θ)wθ(x) are continuous. Between the jumps, the process is deterministically driven by a continuous (with respect to each variable) semi-flow S:R+×XX. The flow property means, as usual, that S(0,x)=x and S(s+t,x)=S(s,S(t,x)) for any xX and any s,tR+.

    Let us now move on to the formal description of the model. Given λ>0 and μM1(X), on some suitable probability space, we first define a discrete-time stochastic process (Xn)nN0 with initial destribution μ, so that

    Xn+1=wθn+1(S(Δτn+1,Xn))for everynN0,

    with Δτn+1:=τn+1τn, where (τn)nN0 and (θn)nN are sequences of random variables with values in R+ and Θ, respectively, defined in such a way that τ0=0, τn Pμ-a.s., as n, and

    Pμ(Δτn+1t|Wn)=1eλtfor anytR+,nN0,Pμ(θn+1B|S(Δτn+1,Xn)=x,Wn)=Bp(x,θ)dθfor anyxX,BBΘ,nN0,

    with W0:=X0 and Wn:=(W0,τ1,,τn,θ1,,θn) for nN. We also demand that, for any nN0, the variables Δτn+1 and θn+1 are conditionally independent given Wn.

    A standard computation shows that (Xn)nN0 is a time-homogeneous Markov chain with transition law Pλ:X×BX[0,1] given by

    Pλ(x,A)=0λeλtΘp(S(t,x),θ)1A(wθ(S(t,x)))dθdtforxX,ABX, (3.1)

    that is,

    Pλ(x,A)=P(Xn+1A|Xn=x)for anyxX,ABX,nN0.

    On the same probability space, we now define a Markov process (X(t))tR+, as an iterpolation of the chain (Xn)nN0, namely

    X(t)=S(tτn,Xn)fort[τn,τn+1),nN0.

    By (Pλ(t))tR+ we shall denote the Markov semigroup associated with the process (X(t))tR+, so that, for any tR+, Pλ(t) is the Markov operator corresponding to the stochastic kernel satisfying

    Pλ(t)(x,A)=Pμ(X(s+t)A|X(s)=x)for anyABX,xX,sR+. (3.2)

    We further assume that there exist a point ˉxX, a Borel measurable function J:X[0,) and constants αR, L,Lw,Lp,λmin,λmax,¯p>0, such that

    LLw+αλ<1for eachλ[λmin,λmax], (3.3)

    and, for any x,yX, the following conditions hold:

    ϰ:=supxX0eλmintΘp(S(t,x),θ)wθ(S(t,ˉx))dθdt<, (3.4)
    S(t,x)S(t,y)LeαtxyfortR+, (3.5)
    S(t,x)S(s,x)(ts)emax{αs,αt}J(x)for0st, (3.6)
    Θp(x,θ)wθ(x)wθ(y)dθLwxy, (3.7)
    Θ|p(x,θ)p(y,θ)|dθLpxy, (3.8)
    Θ(x,y)min{p(x,θ),p(y,θ)}dθ¯p,whereΘ(x,y):={θΘ:wθ(x)wθ(y)Lwxy}. (3.9)

    Note that, upon assuming (3.3), we have λ>max{0,α} for any λ[λmin,λmax]. In what follows, we will write shortly

    ˉα:=max{0,α}. (3.10)

    Moreover, let us introduce

    Msig,J(X)={μMsig(X):J,|μ|<},

    where J is given in (3.6).

    Note that, if (H,|) is a Hilbert space and A:XH is an α-dissipative operator with α0, i.e.,

    AxAy|xyαxy2for anyx,yX,

    which additionally satisfies the so-called range condition, that is, for some T>0,

    XRange(idXtA)fort(0,T),

    then, for any xX, the Cauchy problem of the form

    {y(t)=A(y(t))y(0)=x

    has a unique solution tS(t,x) such that the semi-flow S enjoys conditions (3.5), with L=1, and (3.6), with J(x)=Ax (cf. [30,Theorem 5.3 and Corollary 5.4], as well as [13,Section 3]).

    Moreover, upon assuming compactness of Θ, condition (3.4) can be derived from the conjunction of (3.6) and (3.7) at least in two cases: whenever p does not depend on the pre-jump state, i.e., p(y,θ)=ˉp(θ) for some continuous density function ˉp:ΘR+, or if all the transformations wθ, θΘ, are Lipschitz continuous with a common Lipschitz constant Lw (see [13,Corollary 3.4] for the proof).

    Furthermore, note that conditions formulated in a manner similar to (3.7)–(3.9) are commonly required while examining the asymptotic properties of random iterated function systems (see [26,31,32]), which are covered by the discrete-time model discussed here (in the case where S(t,x)=x). In this connection, it is also worth mentioning that the example described in [33] indicates that the condition of type (3.8) cannot be omitted even in the simplest cases. More precisely, the system {(w1,p),(w2,1p)}, consisting of two contractive maps w1, w2 and a positive continuous probability function p, may admit more than one invariant probability measure (unless at least the Dini continuity of p is assumed).

    Finally, let us indicate that conditions (3.3)–(3.9) are naturally satisfied by a few particular biological models, such as e.g., the model for gene expression [2] (cf. also [13,Section 5]), the model of autoregulated gene expression [3] or the one for cell division [4,15] (see also [34]).

    Consider the abstract model introduced in Section 3. In order to simplify notation, for any tR+, let us introduce the function Π(t):X×BX[0,1] given by

    Π(t)(x,A):=Θp(S(t,x),θ)1A(wθ(S(t,x)))dθforxX,ABX. (4.1)

    Note that Π(t) is a stochastic kernel, and that the corresponding Markov operator is Feller, due to the continuity of p(,θ), S(t,) and wθ, θΘ. Moreover, observe that, for an arbitrary λ>0, we have

    μPλ(A)=X0λeλtΠ(t)(x,A)dtμ(dx)=0λeλtXΠ(t)(x,A)μ(dx)dt=0λeλtμΠ(t)(A)dtfor anyμMsig(X),ABX. (4.2)

    Lemma 4.1. Suppose that conditions (3.6)–(3.8) hold. Then, for any λ>0 and any μMsig,J(X), the function teλtμΠ(t) is Bochner integrable as a map from R+ to (clMsig(X),BL|clMsig(X)), and we have

    μPλ=0λeλtμΠ(t)dt.

    Proof. Let λ>0 and μMsig(X). Note that condition (3.6) implies that

    S(t,x)S(s,x)J(x)eˉα(t+s)|ts|for anys,tR+,xX,

    where ˉα is given by (3.10). Hence, applying (3.7) and (3.8), we see that, for every fBL(X),

    |f,μΠ(t)f,μΠ(s)|=|Π(t)fΠ(s)f,μ|XΘp(S(t,x),θ)|f(wθ(S(t,x)))f(wθ(S(s,x)))|dθ|μ|(dx)+XΘ|p(S(t,x),θ)p(S(s,x),θ)||f(wθ(S(s,x)))|dθ|μ|(dx)(|f|LipLw+fLp)XS(t,x)S(s,x)|μ|(dx)fBL(Lw+Lp)J,|μ|eˉα(t+s)|ts|for anys,tR+.

    This shows that the map tf,eλtμΠ(t) is continuous for any fBL(X), and thus it is Borel measurable. Consequently, it now follows from the Pettis measurability theorem (cf. [28) that the map teλtμΠ(t) is strongly measurable. Furthermore,we have

    eλtμΠ(t)TVμTVeλtfor anytR+,

    which,due to Theorem 2.1,yields that teλtμΠ(t)clMsig(X) is Bochner integrable (with respect to the Lebesgue measure) on R+,and that the integral is a measure in Msig(X),which satisfies

    (0λeλtμΠ(t)dt)(A)=0λeλtμΠ(t)(A)dtfor anyABX.

    The assertion of the lemma now follows from (4.2).

    Lemma 4.2. Let fBL(X). Upon assuming (3.5),(3.7) and (3.8),we have

    μΠ(t)FM(1+(Lw+Lp)Leαt)μFMfor anyμMsig(X),tR+.

    Proof. Let fBL(X) be such that fBL1. Obviously,Π(t)f1 for every tR+. Moreover,from conditions (3.5),(3.7),(3.8) it follows that Π(t)fBL(X),and

    |Π(t)f|Lip(Lw+Lp)Leαtfor anytR+,

    since

    |Π(t)f(x)Π(t)f(y)|=|Θp(S(t,x),θ)f(wθ(S(t,x)))dθΘp(S(t,y),θ)f(wθ(S(t,y)))dθ|(Lw+Lp)S(t,x)S(t,y)(Lw+Lp)Leαtxyfor allx,yX,tR+.

    Therefore,for any μMsig(X) and any tR+,we obtain

    |f,μΠ(t)|=|Π(t)f,μ|Π(t)fBLμFM,

    which gives the desired conclusion.

    Lemma 4.3. For any λ1,λ2>0,we have

    0|λ1eλ1tλ2eλ2t|dt|λ1λ2|(1λ1+1λ2).

    Proof. Without loss of generality,we may assume that λ1<λ2. Since 1exx for every xR,we obtain

    0|λ1eλ1tλ2eλ2t|dtλ10|eλ1teλ2t|dt+(λ2λ1)0eλ2tdt=λ10eλ1t(1e(λ2λ1)t)dt+λ2λ1λ2λ1(λ2λ1)0eλ1ttdt+(λ2λ1)λ2=|λ1λ2|(1λ1+1λ2),

    which completes the proof.

    Lemma 4.4. Let Msig(X) be endowed with the topology induced by the norm FM,and suppose that conditions (3.5)–(3.8) hold. Then,the map (ˉα,)×Msig,J(X)(λ,μ)μPλMsig(X), where ˉα is given by (3.10), is jointly continuous.

    Proof. Let λ1,λ2>ˉα and μ1,μ2Msig,J(X). Note that,due to Lemma 4.1,we have

    μ1Pλ1μ2Pλ2FM=0(λ1eλ1tμ1Π(t)λ2eλ2tμ2Π(t))dtFMμ1TV0|λ1eλ1tλ2eλ2t|dt+0λ2eλ2tμ1Π(t)μ2Π(t)FMdt,

    where the inequality follows from statement (i) of Theorem 2.1 and the fact that μ1Π(t)TVμ1TV. Further,applying Lemmas 4.2 and 4.3,we obtain

    μ1Pλ1μ2Pλ2FMμ1TV|λ1λ2|(1λ1+1λ2)+μ1μ2FM(1+(Lw+Lp)Lλ2λ2α).

    We now see that μ1Pλ1μ2Pλ2FM0,as |λ1λ2|0 and μ1μ2FM0,which completes the proof.

    Suppose that conditions (3.4),(3.5) and (3.7)–(3.9) hold. Then,according to [13,Theorem 4.1] (or [14,Theorem 4.1]), for any λ[λmin,λmax] satisfying LLw+αλ1<1, there exist a unique invariant measure μλM1,1(X) for Pλ and constants qλ(0,1), CλR+ such that

    μPnλμλFMqnλCλ(1+V,μ+V,μλ)for anyμM1,1(X)and anynN, (4.3)

    where V:X[0,) is given by V(x)=xˉx.

    Following the proof of [13,Theorem 4.1], we may conclude that qλ and Cλ depend only on the jump rate od the PDMP and other constants appearing in conditions (3.3)–(3.5) and (3.7)–(3.9) (note that they do not depend on the structure of the model, that is the definitions of S, wθ and p).

    Upon assuming (3.3)–(3.5) and (3.7)–(3.9), there exists C0>0 such that

    V,μλC0for anyλ[λmin,λmax]. (4.4)

    Indeed, let us first define

    a:=λmaxLLwλminαandb:=λmaxϰ,

    where ϰ is given in (3.4), and observe that a<1, due to (3.3). Proceeding similarly as in Step Ⅰ of the proof of [13,Theorem 4.1], we see that conditions (3.5) and (3.7) imply the following:

    PλV(x)aV(x)+bfor anyxXand anyλ[λmin,λmax],

    which further gives

    PnλV(x)anV(x)+b1afor anynNand anyλ[λmin,λmax].

    Now, let C0:=b(1a)1. Then, using the fact that μλ is an invariant measure of Pλ, we get

    V,μλ=V,μλPnλ=PnλV,μλanV,μλ+C0for anynNand anyλ[λmin,λmax].

    Going with n to infinity, we obtain the desired estimation (4.4). As a consequence, we may write (4.3) in the following form:

    μPnλμλFMqnλ˜Cλ(1+V,μ)for anyμM1,1(X)and anynN, (4.5)

    where ˜Cλ:=Cλ(1+C0).

    Lemma 4.5. Suppose that conditions (3.4), (3.5) and (3.7)–(3.9) hold with constants satisfying (3.3), and, for any λ[λmin,λmax], let μλ stand for the unique invariant probability measure of Pλ. Then, the convergence μPnλμλFM0 (as n) is uniform with respect to λ, whenever μM1,1(X).

    Proof. In view of [13,Theorem 4.1], it is sufficient to prove that the convergence is uniform with respect to λ.

    Let us consider the case where α0. Choose an arbitrary λ[λmin,λmax], and note that, by substituting t=λmaxλ1u, the operator Pλ can be expressed in the following form:

    μPλ(A)=X0λeλtΘp(S(t,x),θ)1A(wθ(S(t,x)))dθdtμ(dx)=X0λmaxeλmaxuΘp(Sλ(u,x),θ)1A(wθ(Sλ(u,x)))dθduμ(dx)

    for any μM1(X), ABX, where

    Sλ(u,x):=S(λmaxλu,x)foruR+,xX.

    Moreover, the semi-flow Sλ enjoys condition (3.5), since, for any tR+ and any x,yX, we have

    Sλ(t,x)Sλ(t,y)Leαλmaxλ1txyLeαtxy.

    Hence, we can write Pλ=˜Pλmax, where ˜Pλmax stands for the Markov operator corresponding to the instance of our system with the jump intensity λmax and the flow Sλ in place of S. Taking into account the above observation, it is evident that such a modified system still satisfies conditions (3.4)–(3.5) and (3.7)–(3.9) with constants determined by the primary model, which additionally satisfy LLw+αλ1max<1. Consequently, μλ is then an invariant measure of ˜Pλmax, and hence we can denote it by ˜μλmax. Finally, keeping in mind (4.5), we can conclude that there exist qλmax(0,1) and ˜CλmaxR+ such that

    μPnλμλFM=μ˜Pnλmax˜μλmaxFMqnλmax˜Cλmax(1+V,μ)for anyμM1,1(X),nN.

    In the case where α>0, the proof is similar to the one conducted above (except that this time we substitute t:=λminλ1u), so we omit it.

    Before we formulate and prove the main theorems of this paper, let us first quote the result provided in [35,Theorem 7.11].

    Lemma 5.1. Let (Y,ϱ) and (Z,d) be some metric spaces, and let E be an arbitrary subset of Y. Suppose that (fn)nN0 is a sequence of functions, defined on E, with values in Z, which converges uniformly on E to some function f:EZ. Further, let ˉy be a limit point of E, and assume that

    an:=limyˉyfn(y)

    exists and is finite for every nN0. Then, f has a finite limit at ˉy, and the sequence (an)nN0 converges to it, that is,

    limn(limyˉyfn(y))=limyˉy(limnfn(y)).

    We are now in a position to state the result concerning the continuous dependence of an invariant measure μλ of Pλ on the parameter λ. In the proof we will refer to the lemmas provided in Section 4, as well as to Lemma 5.1.

    Theorem 5.2. Suppose that conditions (3.4)–(3.9) hold with constants satisfying (3.3), and, for any λ[λmin,λmax], let μλ stand for the unique invariant probability measure of Pλ. Then, for every ˉλ[λmin,λmax], we have μλwμˉλ, as λˉλ.

    Proof. Let ˉλ[λmin,λmax]. Due to Lemma 4.5, we know that, for every μM1(X) and any λ[λmin,λmax], we have μPnλμλFM0, as n, and the convergence is uniform with respect to λ.

    Further, since M1(X)Msig,J(X), Lemma 4.4 yields that (ˉα,)×M1(X)(λ,μ)μPλM1(X) is jointly continuous, provided that M1(X) is equipped with the topology induced by the Fortet-Mourier norm. Hence, for any μM1(X) and any nN0, it follows that μPnλμPnˉλFM0, as λˉλ. Finally, according to Lemma 5.1, we get

    limλˉλμλ=limλˉλ(limnμPnλ)=limn(limλˉλμPnλ)=limnPnˉλμ=μˉλ,

    where the limits are taken in (Msig(X),FM). This, together with Theorem 2.2, gives the desired conclusion.

    In the final part of the paper we will study the properties of the Markov semigroup (Pλ(t))tR+, defined by (3.2). In order to apply the relevant results of [13], in what follows, we additionally assume that the measure ϑ, given on the set Θ, is finite. Then, according to [13,Theorem 4.4], for any λ>0, there is a one-to-one correspondence between invariant measures of the operator Pλ and those of the semigroup (Pλ(t))tR+. More precisely, if μλM1(X) is a unique invariant probability measure of Pλ, then νλ:=μλGλM1(X), where

    μGλ(A)=X0λeλt1A(S(t,x))dtμ(dx)for anyμM1(X),ABX,

    is a unique invariant probability measure of (Pλ(t))tR+.

    The main result concerning the continuous-time model, which is formulated and proven below, ensures the continuity of the map λνλ.

    Theorem 5.3. Let ϑ be a finite Borel measure on Θ. Further, suppose that conditions (3.4)–(3.9) hold with constants satisfying (3.3), and, for any λ[λmin,λmax], let νλ stand for the unique invariant probability measure of (P_{\lambda}(t))_{t\in\mathbb{R}_+} . Then, for any \bar{\lambda}\in[\lambda_{min}, \lambda_{max}] , we have \nu_{\lambda}^* \stackrel{w}{\to} \nu_{\bar{\lambda}}^* , as \lambda\to\bar{\lambda} .

    Proof. Let \bar{\lambda}\in[\lambda_{\min}, \lambda_{\max}] , and let f\in BL(X) be such that \|f\|_{BL}\leq 1 . For any \lambda\in[\lambda_{\min}, \lambda_{\max}] , we have

    \begin{array}{l} \left\langle f, \nu_{\lambda}^*\right\rangle = \left\langle f, \mu_{\lambda}^*G_{\lambda}\right\rangle = \int_X\int_0^{\infty}\lambda e^{-\lambda t}f\left(S(t, x)\right)\, dt\, \mu_{\lambda}^*(dx), \end{array}

    whence

    \begin{array}{l} \begin{aligned} \left|\left\langle f, \nu_{\lambda}^*-\nu^*_{\bar{\lambda}}\right\rangle\right| \leq &\int_0^{\infty}\left|\lambda e^{-\lambda t}-\bar{\lambda}e^{-\bar{\lambda}t}\right|\, dt + \left|\int_0^{\infty}\bar{\lambda}e^{-\bar{\lambda}t}\left\langle f\circ S(t, \cdot), \mu^*_{\lambda}-\mu^*_{\bar{\lambda}}\right\rangle\, dt\right|. \end{aligned} \end{array}

    Note that, due to (3.5), f\circ S(t, \cdot)\in BL(X) and \|f\circ S(t, \cdot)\|_{BL}\leq 1+Le^{\alpha t} , and therefore

    \begin{array}{l} \begin{aligned} \left|\int_0^{\infty}\bar{\lambda}e^{-\bar{\lambda}t}\left\langle f\circ S(t, \cdot), \mu^*_{\lambda}-\mu^*_{\bar{\lambda}}\right\rangle\, dt\right| &\leq \left\|\mu_{\lambda}^*-\mu_{\bar{\lambda}}^*\right\|_{FM}\int_0^{\infty} \bar{\lambda} e^{-\overline{\lambda} t}\left(1+Le^{\alpha t}\right) \, dt \\ & = \left\|\mu_{\lambda}^*-\mu_{\bar{\lambda}}^*\right\|_{FM}\left( 1+\frac{L\bar{\lambda}}{\bar{\lambda}-\alpha}\right). \end{aligned} \end{array}

    Combining this and Lemma 4.3, finally gives

    \begin{array}{l} \left\|\nu_{\lambda}^*-\nu_{\bar{\lambda}}^*\right\|_{FM} \leq \left|\lambda-\bar{\lambda}\right| \left(\frac{1}{\lambda}+\frac{1}{\bar{\lambda}}\right)+c \left\|\mu_{\lambda}^*-\mu_{\bar{\lambda}}^*\right\|_{FM} \end{array}

    with c: = 1+L\bar{\lambda}(\bar{\lambda}-\alpha)^{-1} . Hence, referring to Theorems 5.2 and 2.2, we obtain

    \begin{array}{l} \lim\limits_{\lambda\to\bar{\lambda}}\left\|\nu_{\lambda}^*-\nu_{\bar{\lambda}}^*\right\|_{FM} = 0, \end{array}

    and the proof is completed.

    The work of Hanna Wojewódka-Ściążko has been supported by the National Science Centre of Poland, grant number 2018/02/X/ST1/01518.

    All authors declare no conflicts of interest in this paper.



    [1] M. H. A. Davis, Piecewise-deterministic Markov processes: a general class of non-diffusion stochastic models, J. Roy. Statist. Soc. Ser. B, 46 (1984), 353-388.
    [2] M. C. Mackey, M. Tyran-Kamińska, R. Yvinec, Dynamic behaviour of stochastic gene expression models in the presence of bursting, SIAM J. Appl. Math., 73 (2013), 1830-1852.
    [3] S. C. Hille, K. Horbacz, T. Szarek, Existence of a unique invariant measure for a class of equicontinuous Markov operators with application to a stochastic model for an autoregulated gene, Ann. Math. Blaise Pascal, 23 (2016), 171-217.
    [4] A. Lasota and M. C. Mackey, Cell division and the stability of cellular populations, J. Math. Biol., 38 (1999), 241-261.
    [5] M. G. Riedler, M. Thieullen, G. Wainrib, Limit theorems for infinite-dimensional piecewise deterministic Markov processes. Applications to stochastic excitable membrane models, Electron. J. Probab., 17 (2012), 1-48.
    [6] T. Alkurdi, S. C. Hille, O. Van Gaans, Persistence of stability for equilibria of map iterations in Banach spaces under small perturbations, Potential Anal., 42 (2015), 175-201.
    [7] M. Benaïm, C. Lobry, Lotka Volterra with randomly fluctuationg environments or 'how switching between benefcial environments can make survival harder', Ann. Appl. Probab., 26 (2016), 3754-3785.
    [8] M. Benaïm, S. Le Borgne, F. Malrieu, P. A. Zitt, Qualitative properties of certain piecewise deterministic Markov processes, Ann. Inst. Henri Poincar Probab., 51 (2014), 1040-1075.
    [9] M. Benaïm, S. Le Borgne, F. Malrieu, P. A. Zitt, Quantitative ergodicity for some switched dynamical systems, Electron. Commun. Probab., 17 (2012), 1-14.
    [10] F. Dufour, O. L. V. Costa, Stability of piecewise-deterministic Markov processes, SIAM J. Control Optim., 37 (1999), 1483-1502.
    [11] O. L. V. Costa, F. Dufour, Stability and ergodicity of piecewise deterministic Markov processes, SIAM J. Control Optim., 47 (2008), 1053-1077.
    [12] B. Cloez, M. Hairer, Exponential ergodicity for Markov processes with random switching, Bernoulli, 21 (2015), 505-536.
    [13] D. Czapla, K. Horbacz, H. Wojewódka-Ściążko, Ergodic properties of some piecewise-deterministic Markov process with application to gene expression modelling, Stochastic Process. Appl., 2019, doi: 10.1016/j.spa.2019.08.006.
    [14] D. Czapla, J. Kubieniec, Exponential ergodicity of some Markov dynamical systems with application to a Poisson driven stochastic differential equation, Dyn. Syst., 34 (2019), 130-156.
    [15] H. Wojewódka, Exponential rate of convergence for some Markov operators, Statist. Probab. Lett., 83 (2013), 2337-2347.
    [16] S. C. Hille, K. Horbacz, T. Szarek, H. Wojewódka, Limit theorems for some Markov chains, J. Math. Anal. Appl., 443 (2016), 385-408.
    [17] S. C. Hille, K. Horbacz, T. Szarek, H. Wojewódka, Law of the iterated logarithm for some Markov operators, Asymptotic Anal., 97 (2016), 91-112.
    [18] D. Czapla, K. Horbacz, H. Wojewódka-Ściążko, A useful version of the central limit theorem for a general class of Markov chains, preprint, arXiv:1804.09220v2.
    [19] D. Czapla, K. Horbacz, H. Wojewódka-Ściążko, The Strassen invariance principle for certain non-stationary Markov-Feller chains, preprint, arXiv:1810.07300v2.
    [20] T. Komorowski, C. Landim, S. Olla, Fluctuations in Markov processes. Time symmetry and martingale approximation, Springer-Verlag, Heidelberg, 2012.
    [21] V. I. Bogachev, Measure Theory, vol. II, Springer-Verlag, Berlin, 2007.
    [22] P. Gwiazda, S. C. Hille, K. Łyczek, A. Świerczewska-Gwiazda, Differentiability in perturbation parameter of measure solutions to perturbed transport equation, Kinet. Relat. Mod., (2019), in press, preprint arXiv:18l06.00357.
    [23] N. Weaver, Lipschitz Algebras, World Scientific Publishing Co. Pte Ltd., Singapore, 1999.
    [24] V. I. Bogachev, Measure Theory, vol. I, Springer-Verlag, Berlin, 2007.
    [25] R. M. Dudley, Convergence of Baire measures, Stud. Math., 27 (1966), 251-268.
    [26] T. Szarek, Invariant measures for Markov operators with application to function systems, Studia Math., 154 (2003), 207-222.
    [27] D. T. H. Worm, Semigroups on spaces of measures, Ph.D thesis, Leiden University, The Netherlands, 2010. Available from: www.math.leidenuniv.nl/nl/theses/PhD/.
    [28] J. Diestel, Jr. J. J. Uhl, Vector measures, American Mathematical Society, Providence, R.I., 1977.
    [29] J. Evers, S. C. Hille, A. Muntean, Mild solutions to a measure-valued mass evolution problem with flux boundary conditions, J. Diff. Equ., 259 (2015), 1068-1097.
    [30] M. Crandall, T. Ligget, Generation of semigroups of nonlinear transformations on general Banach spaces, Amer. J. Math., 93 (1971), 265-298.
    [31] R. Kapica, M. Ślęczka, Random iterations with place dependent probabilities, to appear in Probab. Math. Statist. (2019).
    [32] A. Lasota, J. A. Yorke, Lower bound technique for Markov operators and iterated function systems, Random Comput. Dyn., 2 (1994), 41-77.
    [33] O. Stenflo, A note on a theorem of Karlin, Stat. Probab. Lett., 54 (2001), 183-187.
    [34] J. J. Tyson, K. B. Hannsgen, Cell growth and division: a deterministic/probabilistic model of the cell cycle, J. Math. Biol., 23 (1986), 231-246.
    [35] W. Rudin, Principles of mathematical analysis, McGraw-Hill, Inc., New York, 1976.
  • This article has been cited by:

    1. Rafał Kapica, Dawid Komorek, Continuous Dependence in a Problem of Convergence of Random Iteration, 2023, 22, 1575-5460, 10.1007/s12346-023-00763-6
    2. Dawid Komorek, Continuous dependence of the weak limit of iterates of some random-valued vector functions, 2023, 97, 0001-9054, 753, 10.1007/s00010-023-00959-w
    3. Sander C. Hille, Esmée S. Theewis, Explicit expressions and computational methods for the Fortet–Mourier distance of positive measures to finite weighted sums of Dirac measures, 2023, 294, 00219045, 105947, 10.1016/j.jat.2023.105947
  • Reader Comments
  • © 2020 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(4382) PDF downloads(453) Cited by(3)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog