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

\UseRawInputEncoding

A note on the top Lyapunov exponent of linear cooperative systems

Michel Benaïm Institut de Mathématiques, Université de Neuchâtel, Switzerland Claude Lobry C.R.H.I, Université Nice Sophia Antipolis, France Tewfik Sari ITAP, University of Montpellier, INRAE, Institut Agro, Montpellier, France Édouard Strickler Université de Lorraine, CNRS, Inria, IECL, Nancy, France
(March 4, 2025)
Abstract

In a recent paper [6], P. Carmona gives an asymptotic formulae for the top Lyapunov exponent of a linear TT-periodic cooperative differential equation, in the limit T.T\rightarrow\infty. This short note discusses and extends this result. The assumption that the system is TT-periodic is replaced by the more general assumption that it is driven by a continuous time uniquely ergodic Feller Markov process (ωt)t>0.(\omega_{t})_{t>0}. When ωt\omega_{t} is replaced by ωtT=ωt/T,\omega^{T}_{t}=\omega_{t/T}, asymptotic formulas for the top Lyapunov exponent in the fast (i.e TT\rightarrow\infty) and slow (T0T\rightarrow 0) regimes are given.

1 Notation and main results

Let d1d\geq 1 be an integer. Let {\cal M} denote the closed convex cone consisting of real d×dd\times d matrices having off diagonal nonnegative entries. Elements of {\cal M} are usually called Metzler matrices. As usual, a matrix MM\in{\cal M} is called irreducible if for all i,j{1,,d}i,j\in\{1,\ldots,d\} there exist nn\in{\mathbb{N}} and a sequence i1=i,i2,,in=ji_{1}=i,i_{2},\ldots,i_{n}=j such that Mil,il+1>0M_{i_{l},i_{l+1}}>0 for l=1,,n1.l=1,\ldots,n-1. Equivalently eMe^{M} has positive entries. Throughout, we let SS denote a compact metric space and

A:S,A:S\rightarrow{\cal M},

a continuous mapping. We consider the linear differential equation

dydt=A(ωt)y\frac{dy}{dt}=A(\omega_{t})y (1)

with initial condition y(0)=x+d{0},y(0)=x\in{\mathbb{R}}^{d}_{+}\setminus\{0\}, under the following assumptions:

(i)

The process (ωt)t0(\omega_{t})_{t\geq 0} is a continuous time Feller Markov process111The precise definition will be recalled in the beginning of Section 2 on SS and is uniquely ergodic. By this, we mean that (ωt)t0(\omega_{t})_{t\geq 0} has a unique invariant probability measure denoted μ.\mu.

(ii)

The average matrix A¯=SA(s)μ(ds)\bar{A}=\int_{S}A(s)\mu(ds) is irreducible.

Remark 1

A sufficient (but non necessary) condition ensuring that A¯\bar{A} is irreducible is that A(s)A(s) is irreducible for some ss in the topological support of μ.\mu. The (easy) proof is left to the reader.

The assumption that A(s)A(s) is Metzler for all sS,s\in S, makes the non-autonomous differential equation (1) cooperative in the sense that y˙iyj0\frac{\partial\dot{y}_{i}}{\partial y_{j}}\geq 0 for all iji\neq j (we refer the reader to [9] for a comprehensive introduction to the theory of deterministic cooperative systems). Systems of this form naturally occur in population dynamics where individuals can migrate between different patches (see e.g [3], [4] and references therein) or different states (see e.g  [12], [6]). They also occur as linearized systems of non-linear cooperative systems (for instance in certain epidemic models [5]). In all these settings, the process (ωt)t0(\omega_{t})_{t\geq 0} represents the time fluctuations of the environment. The top Lyapunov exponent of the system characterizes the population growth rate, and its sign determines whether the population persists or dies out.

The following examples illustrate the fact that the process (ω)t0(\omega)_{t\geq 0} can be deterministic (Examples 1 and 2), or stochastic (Examples 3 and 4).

Example 1 (Periodic case)

Suppose S=/S={\mathbb{R}}/{\mathbb{Z}} identified with the unit circle and

ωt=s+t(𝗆𝗈𝖽 1)\omega_{t}=s+t\,({\sf mod}\,1)

for some sS.s\in S. This is the case considered in [6]. Observe that here μ\mu is the Lebesgue normalized measure on S.S.

Example 2 (Quasi-periodic case)

A natural generalization of Example 1 is as follows. Suppose S=(/)nS=({\mathbb{R}}/{\mathbb{Z}})^{n} is the nn-torus and

ωt=(s1+ta1,s2+ta2,,sn+tan)(𝗆𝗈𝖽 1)\omega_{t}=(s_{1}+ta_{1},s_{2}+ta_{2},\ldots,s_{n}+ta_{n})\,({\sf mod}\,1)

for some s=(s1,,sn)Ss=(s_{1},\ldots,s_{n})\in S and (a1,,an)(a_{1},\ldots,a_{n}) rationally independent numbers. That is i=1nkiai0\sum_{i=1}^{n}k_{i}a_{i}\neq 0 for any integers k1,,knk_{1},\ldots,k_{n} such that (k1,,kn)(0,,0).(k_{1},\ldots,k_{n})\neq(0,\ldots,0). Again (ωt)t0(\omega_{t})_{t\geq 0} is uniquely ergodic with μ\mu the Lebesgue measure on S.S.

Example 3 (Switching)

Suppose S={1,,n}S=\{1,\ldots,n\} for some nn\in{\mathbb{N}}^{*} and (ωt)t0(\omega_{t})_{t\geq 0} is an irreducible continuous time Markov chain on S.S. In other words, the infinitesimal generator of (ωt)t0(\omega_{t})_{t\geq 0} writes

Lf(i)=j=1naij(f(j)f(i))Lf(i)=\sum_{j=1}^{n}a_{ij}(f(j)-f(i))

for all f:S,f:S\mapsto{\mathbb{R}}, where (aij)(a_{ij}) is an irreducible rate matrix. Then (ωt)t0(\omega_{t})_{t\geq 0} is uniquely ergodic and μ\mu is the unique probability vector solution to

j=1n(μjajiμiaij)=0\sum_{j=1}^{n}(\mu_{j}a_{ji}-\mu_{i}a_{ij})=0

for all i=1,,ni=1,\ldots,n. This situation has been considered in [5].

Example 4

Suppose SS is a compact connected Riemannian manifold and (ωt)t0(\omega_{t})_{t\geq 0} a Brownian motion (or an elliptic diffusion or more generally, the solution to a uniquely ergodic stochastic differential equation) on SS. Then (ωt)t0(\omega_{t})_{t\geq 0} is uniquely ergodic and μ\mu is the normalized volume on SS (or a measure absolutely continuous with respect to the volume, in the diffusion case).

We now pass to the analysis of the long term behavior of (1).

Let Δ:=Δd1={x+d:i=1dxi=1}\Delta:=\Delta^{d-1}=\{x\in{\mathbb{R}}^{d}_{+}:\>\sum_{i=1}^{d}x_{i}=1\} be the unit d1d-1 simplex. Every y+d{0}y\in{\mathbb{R}}^{d}_{+}\setminus\{0\} can be written as

y=ρθ,y=\rho\theta,

with ρ=y,𝟏=i=1dyi>0\rho=\langle y,{\bf 1}\rangle=\sum_{i=1}^{d}y_{i}>0 and θ=yy,𝟏Δ.\theta=\frac{y}{\langle y,{\bf 1}\rangle}\in\Delta. Here and throughout, 𝟏{\bf 1} stands for the vector (1,,1)t,(1,\ldots,1)^{t}, and ,\langle\cdot,\cdot\rangle is the usual Euclidean scalar product on d.{\mathbb{R}}^{d}.

Using this decomposition, the differential equation (1) rewrites

dρdt=ρA(ωt)θ,1\frac{d\rho}{dt}=\rho\langle A(\omega_{t})\theta,1\rangle (2)

and

dθdt=F(ωt,θ),\frac{d\theta}{dt}=F(\omega_{t},\theta), (3)

where for all (s,θ)S×Δ(s,\theta)\in S\times\Delta

F(s,θ)=A(s)θA(s)θ,𝟏θ.F(s,\theta)=A(s)\theta-\langle A(s)\theta,{\bf 1}\rangle\theta. (4)

The following proposition is proved in [5], Proposition 2.13, in the case corresponding to Example 3. It mainly relies on the Random Perron-Frobenius theorem as proved by Ruelle [17] and later by Arnold, Demetrius and Gundlach [1] (see also [15], [13], and the references therein). The proof given in [5] extends to the general situation considered here. Details are given in the next section.

Proposition 1

Let (ρt,θt)(\rho_{t},\theta_{t}) be solution to ((2), (3)). The process (ωt,θt)t0(\omega_{t},\theta_{t})_{t\geq 0} is a Feller Markov process uniquely ergodic on S×Δ.S\times\Delta.

Let π\pi denote its (unique) invariant probability and let

Λ=S×ΔA(s)θ,𝟏π(dsdθ).\Lambda=\int_{S\times\Delta}\langle A(s)\theta,{\bf 1}\rangle\pi(dsd\theta).

Then, for every initial conditions ρ(0)>0,θ(0)Δ\rho(0)>0,\theta(0)\in\Delta and ω0=s,\omega_{0}=s, with probability one,

limtlog(ρt)t=Λ.\lim_{t\rightarrow\infty}\frac{\log(\rho_{t})}{t}=\Lambda.

For further notice, we call Λ\Lambda the top Lyapunov exponent222see the remark 2 for a justification of this terminology of the system given by (1). For periodic linear differential equations it corresponds to what is sometimes called the principal Lyapunov exponent [13], or the largest Floquet multiplier [6]. That is, the Floquet exponent with the largest real part. For further details we refer the reader to the Section II.2 of the excellent survey [13] by Mierczyński.

The following corollary easily follows from Proposition 1. It provides simple estimates of Λ.\Lambda. Other estimates, mainly for periodic systems, can be found in [13] and in [14] for more general systems.

Corollary 2

The following inequalities hold true:

(i)
S[mini=1,,dj=1dAji(s)]μ(ds)ΛS[maxi=1,,dj=1dAji(s)]μ(ds);\int_{S}[\min_{i=1,\ldots,d}\sum_{j=1}^{d}A_{ji}(s)]\mu(ds)\leq\Lambda\leq\int_{S}[\max_{i=1,\ldots,d}\sum_{j=1}^{d}A_{ji}(s)]\mu(ds);
(ii)
Sλmin(A(s)+A(s)t2)μ(ds)ΛSλmax(A(s)+A(s)t2)μ(ds),\int_{S}\lambda_{min}(\frac{A(s)+A(s)^{t}}{2})\mu(ds)\leq\Lambda\leq\int_{S}\lambda_{max}(\frac{A(s)+A(s)^{t}}{2})\mu(ds),

where λmin\lambda_{min} (respectively λmax\lambda_{max}) stands for the smallest (largest) eigenvalue.

Proof:    (i).(i). For all θΔ\theta\in\Delta and sSs\in S

mini=1,,dj=1dAji(s)A(s)θ,1maxi=1,,dj=1dAji(s),\min_{i=1,\ldots,d}\sum_{j=1}^{d}A_{ji}(s)\leq\langle A(s)\theta,1\rangle\leq\max_{i=1,\ldots,d}\sum_{j=1}^{d}A_{ji}(s),

and the result follows from the integral representation of Λ.\Lambda.

(ii).(ii). Let y2=y,y.\|y\|_{2}=\sqrt{\langle y,y\rangle}. Then,

ddtlog(yt2)=y,A(ωt)ytyt22=θt,A(ωt)θtθt22.\frac{d}{dt}\log(\|y_{t}\|_{2})=\frac{\langle y,A(\omega_{t})y_{t}\rangle}{\|y_{t}\|_{2}^{2}}=\frac{\langle\theta_{t},A(\omega_{t})\theta_{t}\rangle}{\|\theta_{t}\|_{2}^{2}}.

Unique ergodicity of (ωt,θt)t0,(\omega_{t},\theta_{t})_{t\geq 0}, then implies (see e.g  [2], Propositions 7.1 and 4.58) that for every initial condition y0+d{0},y_{0}\in{\mathbb{R}}^{d}_{+}\setminus\{0\},

Λ=limtlog(yt2)t=S×Δθ,A(s)θθ22π(dsdθ)\Lambda=\lim_{t\rightarrow\infty}\frac{\log(\|y_{t}\|_{2})}{t}=\int_{S\times\Delta}\frac{\langle\theta,A(s)\theta\rangle}{\|\theta\|_{2}^{2}}\pi(dsd\theta)

almost surely. Now, for all u+du\in{\mathbb{R}}^{d}_{+} such that u2=1,\|u\|_{2}=1,

λmin(A(s)+A(s)t2)u,A(s)uλmax(A(s)+A(s)t2).\lambda_{min}(\frac{A(s)+A(s)^{t}}{2})\leq\langle u,A(s)u\rangle\leq\lambda_{max}(\frac{A(s)+A(s)^{t}}{2}).

This proves the result. \Box
In the particular case of a periodic system (Example 1), more can be said.

Proposition 3

Suppose S=/[0,1[S={\mathbb{R}}/{\mathbb{Z}}\backsim[0,1[ as in Example 1. There exists a continuous 11-periodic function tθ(t)Δ,t\in{\mathbb{R}}\rightarrow\theta^{*}(t)\in\Delta, such that: For all sSs\in S and ωt=s+t(𝗆𝗈𝖽 1),\omega_{t}=s+t\,({\sf mod}\,1), tθ(s+t)t\rightarrow\theta^{*}(s+t) is the unique 11-periodic solution to (3). It is globally asymptotically stable in the sense that

limtθ(t)θ(s+t)=0\lim_{t\rightarrow\infty}\|\theta(t)-\theta^{*}(s+t)\|=0

for every solution (θ(t))t0(\theta(t))_{t\geq 0} to (3) with ωt=s+t(𝗆𝗈𝖽 1).\omega_{t}=s+t\,({\sf mod}\,1). In particular,

π(dsdθ)=dsδθ(s)(dθ)\pi(dsd\theta)=ds\delta_{\theta^{*}(s)}(d\theta)

and

Λ=01A(s)θ(s),𝟏𝑑s.\Lambda=\int_{0}^{1}\langle A(s)\theta^{*}(s),{\bf 1}\rangle ds.

1.1 Slow and fast regimes

For all T>0,T>0, let ωtT=ωt/T.\omega^{T}_{t}=\omega_{t/T}. Like (ωt)t0(\omega_{t})_{t\geq 0}, (ωtT)t0(\omega^{T}_{t})_{t\geq 0} is a Feller Markov process on S,S, uniquely ergodic with invariant probability μ.\mu. The parameter 1/T1/T can be understood as a velocity parameter. For instance, in the context of Example 1, (ωtT)t0(\omega^{T}_{t})_{t\geq 0} is a TT-periodic signal. In the context of Example 3, its mean sojourn time in each state iSi\in S is proportional to T.T.

Consider the differential equation (1) with (ωt)t0(\omega_{t})_{t\geq 0} replaced by (ωtT)t0.(\omega^{T}_{t})_{t\geq 0}. We let πT\pi^{T} and ΛT\Lambda^{T} denote the corresponding invariant probabilities on S×ΔS\times\Delta and top Lyapunov exponent as defined in Proposition 1. This section considers the fast and slow regimes obtained as T0T\rightarrow 0 and T.T\rightarrow\infty.

For a d×dd\times d real matrix M,M, we let λmax(M)\lambda_{max}(M) denote the largest real part of its eigenvalues (sometimes called the spectral abscissa of MM; see e.g [6]).

For r>0r>0 sufficiently large, A¯+rI\bar{A}+rI has nonnegative entries and is irreducible. Hence, by Perron-Frobenius theorem (applied to A¯+rI\bar{A}+rI), λmax(A¯)\lambda_{max}(\bar{A}) is an eigenvalue and there exists a unique vector, the Perron-Frobenius vector of A¯,\bar{A}, θΔ,\theta^{*}\in\Delta, such that

A¯θ=λmax(A¯)θ.\bar{A}\theta^{*}=\lambda_{max}(\bar{A})\theta^{*}.
Proposition 4 (Fast regime)
limT0πT=μδθ\lim_{T\rightarrow 0}\pi^{T}=\mu\otimes\delta_{\theta^{*}}

(for the weak* topology) and

limT0ΛT=λmax(A¯).\lim_{T\rightarrow 0}\Lambda^{T}=\lambda_{max}(\overline{A}).

Note that Proposition 4 has been proven for Example 3 in ([5], Corollary 2.15). The next result generalizes [6] beyond Example 1. Let 𝗌𝗎𝗉𝗉(μ)\mathsf{supp}(\mu) be the topological support of μ.\mu. Assume that for all s𝗌𝗎𝗉𝗉(μ),s\in\mathsf{supp}(\mu), A(s)A(s) is irreducible. Then, under this assumption, there exists for all s𝗌𝗎𝗉𝗉(μ)s\in\mathsf{supp}(\mu) a unique Perron-Frobenius vector for A(s),θ(s)ΔA(s),\theta^{*}(s)\in\Delta characterized by

A(s)θ(s)=λmax(A(s))θ(s).A(s)\theta^{*}(s)=\lambda_{max}(A(s))\theta^{*}(s).
Proposition 5 (Slow regime)

Assume that for all s𝗌𝗎𝗉𝗉(μ),s\in\mathsf{supp}(\mu), A(s)A(s) is irreducible. Then

limTπT=μ(ds)δθ(s)\lim_{T\rightarrow\infty}\pi^{T}=\mu(ds)\delta_{\theta^{*}(s)}

(for the weak* topology) and

limTΛT=Sλmax(A(s))μ(ds).\lim_{T\rightarrow\infty}\Lambda^{T}=\int_{S}\lambda_{max}(A(s))\mu(ds).

2 Proofs

Notation and Background

If XX is a metric space (such as S,Δ,S×ΔS,\Delta,S\times\Delta) we let B(X)B(X) denote the space of real valued Borel bounded functions on XX and C(X)B(X)C(X)\subset B(X) the subspace of bounded continuous functions. For all fB(X)f\in B(X) we let f=supxX|f(x)|.\|f\|_{\infty}=\sup_{x\in X}|f(x)|. If ν\nu is a probability on XX and fB(X)f\in B(X) we write ν(f)\nu(f) for Xf𝑑ν.\int_{X}fd\nu.

Our main assumption that (ωt)t0(\omega_{t})_{t\geq 0} is a Feller Markov process on S,S, means, as usual, that (ωt)t0(\omega_{t})_{t\geq 0} is a Markov process whose transition semigroup (Pt)t0(P_{t})_{t\geq 0} is Feller. That is:

(a)

Pt(C(S))C(S);P_{t}(C(S))\subset C(S);

(b)

limt0Ptf(s)=f(s)\lim_{t\rightarrow 0}P_{t}f(s)=f(s) for all fC(S)f\in C(S) and sS.s\in S.

It turns out (see e.g [10], Theorem 19.6) that (a)(a) and (b)(b) make (Pt)t0(P_{t})_{t\geq 0} strongly continuous in the sense that limt0Ptff=0\lim_{t\rightarrow 0}\|P_{t}f-f\|_{\infty}=0 for all fC(S).f\in C(S).

An invariant probability for (ωt)t0(\omega_{t})_{t\geq 0} (or (Pt)t0(P_{t})_{t\geq 0}) is a probability μ\mu on SS such that for all t0,t\geq 0, μPt=μ\mu P_{t}=\mu (i.e μ(Ptf)=μ(f)\mu(P_{t}f)=\mu(f) for all fB(S)f\in B(S)). Feller continuity and compactness of SS imply that such a μ\mu always exists (see e.g [2], Corollary 4.21). Our assumption that (ωt)t0(\omega_{t})_{t\geq 0} is uniquely ergodic means that μ\mu is unique.

A useful consequence of Feller continuity is that we can assume without loss of generality that (ωt)t0(\omega_{t})_{t\geq 0} is defined on the space Ω\Omega consisting of càdlàg (right-continuous, left limit) paths ω:+S\omega:{\mathbb{R}}_{+}\rightarrow S equipped with the Skorohod topology and associated Borel sigma field (see e.g [10], Theorem 19.15). As usual, for all sSs\in S we let s\mathbb{P}_{s} denote the law of (ωt)t0(\omega_{t})_{t\geq 0} starting from ω0=s\omega_{0}=s and μ=Ssμ(ds).\mathbb{P}_{\mu}=\int_{S}\mathbb{P}_{s}\mu(ds). The associated expectations are denoted 𝔼s\mathbb{E}_{s} and 𝔼μ.\mathbb{E}_{\mu}.

For all ωΩ\omega\in\Omega and t0t\geq 0 we let 𝚯t(ω)\mathbf{\Theta}_{t}(\omega) denote the shifted path defined as 𝚯t(ω)(s)=ω(t+s).\mathbf{\Theta}_{t}(\omega)(s)=\omega(t+s). Ergodicity of μ\mu for the Markov process (ωt)t0(\omega_{t})_{t\geq 0} makes μ\mathbb{P}_{\mu} ergodic (but not uniquely ergodic) for the dynamical system (𝚯t)t0(\mathbf{\Theta}_{t})_{t\geq 0} on Ω\Omega (see e.g [2], Proposition 4.49).

2.1 Proof of Propositions 1 and 3

For ωΩ\omega\in\Omega the solution to (1) writes y(t)=Φ(t,ω)xy(t)=\Phi(t,\omega)x where (Φ(t,ω))t0(\Phi(t,\omega))_{t\geq 0} is solution to the matrix valued differential equation

dMdt=A(ωt)M,M(0)=Id.\frac{dM}{dt}=A(\omega_{t})M,M(0)=Id.

Let +\cal{M}_{+}\subset{\cal M} denote the set of d×dd\times d Metzler matrices having positive diagonal entries and +++\cal{M}_{++}\subset{\cal M}_{+} the set of matrices having positive entries. Observe that

Φ(t,ω)+\Phi(t,\omega)\in\cal{M}_{+}

for all t0.t\geq 0. Indeed, for rr large enough and all sS,A(s)+2rIdrIds\in S,\,A(s)+2rId\geq rId so that e2rtΦ(t,ω)eRtIde^{2rt}\Phi(t,\omega)\geq e^{Rt}Id (componentwise).

For all θΔ\theta\in\Delta, the solution to (3) with initial condition θ(0)=θ,\theta(0)=\theta, writes

θ(t)=Ψ(t,ω)θ:=Φ(t,ω)θΦ(t,ω)θ,𝟏.\theta(t)=\Psi(t,\omega)\theta:=\frac{\Phi(t,\omega)\theta}{\langle\Phi(t,\omega)\theta,\bf{1}\rangle}.
Lemma 6

For μ\mathbb{P}_{\mu} almost all ωΩ:\omega\in\Omega:

(i)

There exists NN\in{\mathbb{N}} such that Φ(t,ω)++\Phi(t,\omega)\in\cal{M}_{++} for all tN;t\geq N;

(ii)

For all θ,θΔ\theta,\theta^{\prime}\in\Delta

limtΨ(t,ω)θΨ(t,ω)θ=0.\lim_{t\rightarrow\infty}\|\Psi(t,\omega)\theta-\Psi(t,\omega)\theta^{\prime}\|=0.

Proof:    (i).(i). First observe that Φ(t,ω)++ertΦ(t,ω)++\Phi(t,\omega)\in{\cal M}_{++}\Leftrightarrow e^{rt}\Phi(t,\omega)\in{\cal M}_{++} for all r>0.r>0. Therefore, replacing A(s)A(s) by A(s)+rIdA(s)+rId for r>A,r>\|A\|_{\infty}, we can assume without loss of generality that A(s)+A(s)\in{\cal M}_{+} for all sS.s\in S.

Let x(t)=Φ(t,ω)xx(t)=\Phi(t,\omega)x with x+d{0}.x\in{\mathbb{R}}^{d}_{+}\setminus\{0\}. Suppose xi(0)>0.x_{i}(0)>0. Then xi(t)>0x_{i}(t)>0 because x˙i(t)Aii(ωt)xi(t)0.\dot{x}_{i}(t)\geq A_{ii}(\omega_{t})x_{i}(t)\geq 0. By irreducibility of A¯\bar{A}, for all jij\neq i there exists a sequence i0=i,i1,,in=ji_{0}=i,i_{1},\ldots,i_{n}=j such that A¯ikik1>0\bar{A}_{i_{k}i_{k-1}}>0 for k=1,n.k=1,\ldots n. By ergodicity, there exists a Borel set Ω~Ω\tilde{\Omega}\subset\Omega with μ(Ω~)=1\mathbb{P}_{\mu}(\tilde{\Omega})=1 such that for all ωΩ~\omega\in\tilde{\Omega}

1t0tA(ωu)𝑑uA¯.\frac{1}{t}\int_{0}^{t}A(\omega_{u})du\rightarrow\bar{A}.

Therefore, for all ωΩ~,\omega\in\tilde{\Omega}, there exists a sequence t1>t2>>tnt_{1}>t_{2}>\ldots>t_{n} with

Aikik1(ωtk)>0.A_{i_{k}i_{k-1}}(\omega_{t_{k}})>0.

By right continuity of (ωt)(\omega_{t}) we also have Aikik1(ωt)>0A_{i_{k}i_{k-1}}(\omega_{t})>0 for tkttk+εt_{k}\leq t\leq t_{k}+\varepsilon for some ε>0.\varepsilon>0. It follows that x˙i1(t)Ai1,i(ωt)x1(t)>0\dot{x}_{i_{1}}(t)\geq A_{i_{1},i}(\omega_{t})x_{1}(t)>0 for all t1tt1+ε.t_{1}\leq t\leq t_{1}+\varepsilon. Hence xi1(t)>0x_{i_{1}}(t)>0 for all t>t1.t>t_{1}. Similarly xi2(t)>0x_{i_{2}}(t)>0 for all t>t2t>t_{2} and, by recursion, xj(t)>0x_{j}(t)>0 for all t>tn.t>t_{n}. In summary, we have shown that for all i,j{1,,d}i,j\in\{1,\ldots,d\} and ωΩ~,\omega\in\tilde{\Omega}, there exists a time tnt_{n} depending on i,j,ωi,j,\omega such that for all ttnt\geq t_{n} xj(t)>0x_{j}(t)>0 whenever xi(0)>0.x_{i}(0)>0. This proves (i).(i).

(ii).(ii). Let ++d={xd:xi>0, for all i=1d}{\mathbb{R}}^{d}_{++}=\{x\in{\mathbb{R}}^{d}\>:x_{i}>0,\mbox{ for all }i=1\ldots d\} and Δ˙=Δ++d\dot{\Delta}=\Delta\cap{\mathbb{R}}^{d}_{++} be the relative interior of Δ.\Delta. The projective or Hilbert metric dHd_{H} on ++d{\mathbb{R}}^{d}_{++} (see Seneta [18]) is defined by

dH(x,y)=logmax1idxi/yimin1idxi/yi.d_{H}(x,y)=\log\frac{\max_{1\leq i\leq d}x_{i}/y_{i}}{\min_{1\leq i\leq d}x_{i}/y_{i}}.

Note that for all α,β>0,\alpha,\beta>0, dH(αx,βy)=dH(x,y)d_{H}(\alpha x,\beta y)=d_{H}(x,y) so that dHd_{H} is not a distance on ++d.{\mathbb{R}}^{d}_{++}. However its restriction to Δ˙\dot{\Delta} is. Furthermore, for all θ\theta, θΔ˙\theta^{\prime}\in\dot{\Delta},

max1id|θiθi|edH(θ,θ)1.\max_{1\leq i\leq d}|\theta_{i}-\theta^{\prime}_{i}|\leq e^{d_{H}(\theta,\theta^{\prime})}-1. (5)

By a theorem of Birkhoff (see e.g [18], Section 3.4), for all M+,M\in{\cal M}_{+},

sup{x,y++ddH(x,y)>0}dH(Mx,My)dH(x,y)=τ[M]\sup_{\{x,y\in{\mathbb{R}}^{d}_{++}\>d_{H}(x,y)>0\}}\frac{d_{H}(Mx,My)}{d_{H}(x,y)}=\tau[M] (6)

where 0τ(M)10\leq\tau(M)\leq 1 is the number defined as τ(M)=1r(M)1+r(M)\tau(M)=\frac{1-\sqrt{r(M)}}{1+\sqrt{r(M)}} with r(M)=mini,j,k,lminMikMjlMjkMilr(M)=\min_{i,j,k,l}\min\frac{M_{ik}M_{jl}}{M_{jk}M_{il}} if M++M\in{\cal M}_{++} and r(M)=0r(M)=0 if M+++.M\in{\cal M}_{+}\setminus{\cal M}_{++}. In particular, for M+M\in{\cal M}_{+}, τ(M)<1\tau(M)<1 if and only if M++M\in{\cal M}_{++}.

For all 0st,0\leq s\leq t, let

Fs,t(ω)=max{log(τ[Φ(ts,𝚯s(ω))]),st}[st,0].F_{s,t}(\omega)=\max\{\log(\tau[\Phi(t-s,\mathbf{\Theta}_{s}(\omega))]),s-t\}\in[s-t,0].

We claim that (Fs,t)0st(F_{s,t})_{0\leq s\leq t} is a sub-additive process. That is:

(i)

Fs,t𝚯v=Fs+v,t+v,F_{s,t}\circ\mathbf{\Theta}_{v}=F_{s+v,t+v}, and

(ii)

Fs,uFs,t+Ft,u,F_{s,u}\leq F_{s,t}+F_{t,u},

for all stus\leq t\leq u and v0.v\geq 0.

The first assertion is immediate because 𝚯s𝚯v=𝚯s+v.\mathbf{\Theta}_{s}\circ\mathbf{\Theta}_{v}=\mathbf{\Theta}_{s+v}. For the second, by the cocycle property

Φ(us,𝚯s(ω))=Φ(ut,𝚯t(ω))Φ(ts,𝚯s(ω)).\Phi(u-s,\mathbf{\Theta}_{s}(\omega))=\Phi(u-t,\mathbf{\Theta}_{t}(\omega))\circ\Phi(t-s,\mathbf{\Theta}_{s}(\omega)).

Thus,

log(τ[Φ(us,𝚯s(ω))]log(τ[Φ(ut,𝚯t(ω))]+log(τ[Φ(ts,𝚯s(ω))].\log(\tau[\Phi(u-s,\mathbf{\Theta}_{s}(\omega))]\leq\log(\tau[\Phi(u-t,\mathbf{\Theta}_{t}(\omega))]+\log(\tau[\Phi(t-s,\mathbf{\Theta}_{s}(\omega))].

This proves (ii)(ii).

Note also that t,sFs,t(ω)t,s\rightarrow F_{s,t}(\omega) is continuous and that sup0st1|Fs,t|1,\displaystyle\sup_{0\leq s\leq t\leq 1}|F_{s,t}|\leq 1, so that the integrability conditions required for the continuous time version of Kingman’s subadditive ergodic theorem (as stated in [11], Theorem 5.6) are satisfied. Therefore, by this theorem,

lim suptlog(τ[Φ(t,ω)])tlimtF0,t(ω)t=γ,\limsup_{t\rightarrow\infty}\frac{\log(\tau[\Phi(t,\omega)])}{t}\leq\lim_{t\rightarrow\infty}\frac{F_{0,t}(\omega)}{t}=\gamma,

μ\mathbb{P}_{\mu} almost surely, where

γ=inft>0𝔼μF0,tt.\gamma=\inf_{t>0}\mathbb{E}_{\mu}\frac{F_{0,t}}{t}.

Clearly γ<0.\gamma<0. For otherwise we would have that τ[Φ(n,ω)]=1Φ(n,ω)+++\tau[\Phi(n,\omega)]=1\Leftrightarrow\Phi(n,\omega)\in{\cal M}_{+}\setminus{\cal M}_{++} for all n,μn\in{\mathbb{N}},\,\mathbb{P}_{\mu} almost surely, in contradiction with (i).(i).

Let NN be like in assertion (i)(i) of the Lemma. Then, by what precedes, μ\mathbb{P}_{\mu} almost surely,

lim suptlog(dH(Ψ(t+N,ω)θ,Ψ(t+N,ω)θ))t\limsup_{t\rightarrow\infty}\frac{\log(d_{H}(\Psi(t+N,\omega)\theta,\Psi(t+N,\omega)\theta^{\prime}))}{t}
lim suptlog(τ[Φ(t,𝚯N(ω))])t+lim suptlog(dH(Ψ(N,ω)θ,Ψ(N,ω)θ))t=γ.\leq\limsup_{t\rightarrow\infty}\frac{\log(\tau[\Phi(t,\mathbf{\Theta}_{N}(\omega))])}{t}+\limsup_{t\rightarrow\infty}\frac{\log(d_{H}(\Psi(N,\omega)\theta,\Psi(N,\omega)\theta^{\prime}))}{t}=\gamma.

By inequality (5), this concludes the proof. \Box
Let (Qt)t0(Q_{t})_{t\geq 0} denote the semigroup of the process (ωt,θt)t0.(\omega_{t},\theta_{t})_{t\geq 0}. Then, for all fB(S×Δ)f\in B(S\times\Delta) (s,θ)S×Δ,(s,\theta)\in S\times\Delta,

Qtf(s,θ)=𝔼s[f(ωt,Ψ(t,ω)(θ))].Q_{t}f(s,\theta)=\mathbb{E}_{s}[f(\omega_{t},\Psi(t,\omega)(\theta))].
Lemma 7

The semigroup (Qt)t0(Q_{t})_{t\geq 0} is Feller.

Proof:    We need to show that (a)(a) Qt(C(S×Δ)C(S×Δ)Q_{t}(C(S\times\Delta)\subset C(S\times\Delta) and (b)(b) limt0Qtf(s,θ)=f(s,θ)\lim_{t\rightarrow 0}Q_{t}f(s,\theta)=f(s,\theta) for all fC(S×Δ).f\in C(S\times\Delta).

(a).(a). It is easy to verify that there exist constants c1,c20c_{1},c_{2}\geq 0 such that for all s,sS,θ,θΔs,s^{\prime}\in S,\theta,\theta^{\prime}\in\Delta

F(s,θ)F(s,θ)\displaystyle\|F(s,\theta)-F(s,\theta^{\prime})\| \displaystyle\leq c1θθ\displaystyle c_{1}\|\theta-\theta^{\prime}\| (7)
F(s,θ)F(s,θ)\displaystyle\|F(s,\theta)-F(s^{\prime},\theta)\| \displaystyle\leq c2A(s)A(s)\displaystyle c_{2}\|A(s)-A(s^{\prime})\|

where FF is defined by (4). Fix ε>0\varepsilon>0 and let ω~\tilde{\omega} be the path defined as ω~u=ωkε\tilde{\omega}_{u}=\omega_{k\varepsilon} for all kεu<(k+1)ε.k\varepsilon\leq u<(k+1)\varepsilon. Then, by Gronwall’s lemma,

Ψ(t,ω)(θ)Ψ(t,ω~)(θ)ct0tA(ω(u))A(ω~(u))𝑑u\|\Psi(t,\omega)(\theta)-\Psi(t,\tilde{\omega})(\theta)\|\leq c_{t}\int_{0}^{t}\|A(\omega(u))-A(\tilde{\omega}(u))\|du (8)

where ct=ec1tc2.c_{t}=e^{c_{1}t}c_{2}. Thus, by Jensen inequality,

𝔼s(Ψ(t,ω)θΨ(t,ω~))θ))2𝔼s(Ψ(t,ω)(θ)Ψ(t,ω~)(θ)2)\mathbb{E}_{s}(\|\Psi(t,\omega)\theta-\Psi(t,\tilde{\omega}))\theta)\|)^{2}\leq\mathbb{E}_{s}(\|\Psi(t,\omega)(\theta)-\Psi(t,\tilde{\omega})(\theta)\|^{2})
ct2t0t𝔼s(A(ωu))A(ω~u)2)du\leq c_{t}^{2}t\int_{0}^{t}\mathbb{E}_{s}(\|A(\omega_{u}))-A(\tilde{\omega}_{u})\|^{2})du

The choice of the norm being arbitrary we can assume that the norm on the right hand side of the preceding inequality is the Euclidean on d2.{\mathbb{R}}^{d^{2}}. Then, for all kεu<(k+1)ε,k\varepsilon\leq u<(k+1)\varepsilon,

𝔼s(A(ωu))A(ω~u)2)=𝔼s(𝔼(A(ωu))A(ω~u)2)|kε))\mathbb{E}_{s}(\|A(\omega_{u}))-A(\tilde{\omega}_{u})\|^{2})=\mathbb{E}_{s}\left(\mathbb{E}(\|A(\omega_{u}))-A(\tilde{\omega}_{u})\|^{2})|{\cal F}_{k\varepsilon})\right)
=𝔼s(Pukε(A2)(ωkε)2A(ωkε),Pukε(A)(ωkε)+A(ωkε)2)=\mathbb{E}_{s}(P_{u-k\varepsilon}(\|A\|^{2})(\omega_{k\varepsilon})-2\langle A(\omega_{k\varepsilon}),P_{u-k\varepsilon}(A)(\omega_{k\varepsilon})\rangle+\|A(\omega_{k\varepsilon})\|^{2})
sup0hεPh(A2)A2)+2APhAA:=δ(ε).\leq\sup_{0\leq h\leq\varepsilon}\|P_{h}(\|A\|^{2})-\|A\|^{2})\|_{\infty}+2\|A\|_{\infty}\|P_{h}A-A\|_{\infty}:=\delta(\varepsilon).

Observe that δ(ε)0\delta(\varepsilon)\rightarrow 0 as ε0\varepsilon\rightarrow 0 by strong continuity of (Pt)t0.(P_{t})_{t\geq 0}. Combining the two last inequalities, we get

𝔼s(Ψ(t,ω)(θ)Ψ(t,ω~)(θ)2)ct2t2δ(ε)\mathbb{E}_{s}(\|\Psi(t,\omega)(\theta)-\Psi(t,\tilde{\omega})(\theta)\|^{2})\leq c_{t}^{2}t^{2}\delta(\varepsilon) (9)

Let now fC(S×Δ).f\in C(S\times\Delta). Then, for every δ>0\delta>0 there exists α>0,\alpha>0, such that

|f(s,θ)f(s,θ)|δ+2f𝟏θθα|f(s,\theta)-f(s,\theta^{\prime})|\leq\delta+2\|f\|{\bf 1}_{\|\theta-\theta^{\prime}\|\geq\alpha}

Thus

|Qtf(s,θ)𝔼s(f(ωt,Ψ(t,ω~)(θ))||Q_{t}f(s,\theta)-\mathbb{E}_{s}(f(\omega_{t},\Psi(t,\tilde{\omega})(\theta))|
𝔼s(|f(ωt,Ψ(t,ω)(θ))f(ωt,Ψ(t,ω~)(θ))|)δ+2fα2ct2t2δ(ε).\leq\mathbb{E}_{s}\left(|f(\omega_{t},\Psi(t,\omega)(\theta))-f(\omega_{t},\Psi(t,\tilde{\omega})(\theta))|\right)\leq\delta+2\frac{\|f\|}{\alpha^{2}}c_{t}^{2}t^{2}\delta(\varepsilon).

This shows that the left hand term goes to 0 uniformly in (s,θ)S×Δ(s,\theta)\in S\times\Delta as ε0.\varepsilon\rightarrow 0.

In order to conclude it suffices to show that (s,θ)𝔼s(f(ωt,Ψ(t,ω~)(θ)))(s,\theta)\rightarrow\mathbb{E}_{s}\left(f(\omega_{t},\Psi(t,\tilde{\omega})(\theta))\right) is continuous. For all sS,s\in S, let (Ψts)t0(\Psi^{s}_{t})_{t\geq 0} denote the semi-flow on Δ\Delta induced by the autonomous differential equation dθdt=F(s,θ)\frac{d\theta}{dt}=F(s,\theta) Then for kεt<(k+1)εk\varepsilon\leq t<(k+1)\varepsilon

f(ωt,Ψ(t,ω~)(θ))=f(ωt,ΨtkεωkεΨεωεΨεω0(θ))f(\omega_{t},\Psi(t,\tilde{\omega})(\theta))=f(\omega_{t},\Psi^{\omega_{k\varepsilon}}_{t-k\varepsilon}\circ...\Psi^{\omega_{\varepsilon}}_{\varepsilon}\circ\Psi^{\omega_{0}}_{\varepsilon}(\theta))

Now, for every hC(Sk+2×Δ),h\in C(S^{k+2}\times\Delta), Feller continuity of (Pt)t0(P_{t})_{t\geq 0}, makes the map (s,θ)𝔼s(h(ωt,ωkε,,ω0,θ)(s,\theta)\rightarrow\mathbb{E}_{s}(h(\omega_{t},\omega_{k\varepsilon},\ldots,\omega_{0},\theta) continuous. This is immediate to verify when hh is a product function (i.e  h(sk+1,,s0,θ)=hk+1(sk+1)h0(s0)g(θ)h(s_{k+1},\ldots,s_{0},\theta)=h_{k+1}(s_{k+1})\cdot h_{0}(s_{0})g(\theta)) and the general case follows by the density in C(Sk+2×Δ)C(S^{k+2}\times\Delta) of the vector space span by product functions. This concludes the proof of (a).(a).

(b).(b). Let fC(S×Δ)f\in C(S\times\Delta) and δ>0.\delta>0. Because Ψ(t,ω)(θ)θtF\|\Psi(t,\omega)(\theta)-\theta\|\leq t\|F\|_{\infty}, Qtf(s,θ)𝔼s(f(ωt,θ))δ\|Q_{t}f(s,\theta)-\mathbb{E}_{s}(f(\omega_{t},\theta))\|\leq\delta for all tt sufficiently small. By Feller continuity of (Pt)(P_{t}) limt0𝔼s(f(ωt,θ))=limt0Pt(f(,θ))(s)=f(s,θ).\lim_{t\rightarrow 0}\mathbb{E}_{s}(f(\omega_{t},\theta))=\lim_{t\rightarrow 0}P_{t}(f(\cdot,\theta))(s)=f(s,\theta). \Box
We can now conclude the proof of Proposition 1. It follows from Lemma 6 (ii)(ii) that for all f:S×Δf:S\times\Delta\mapsto{\mathbb{R}} continuous (hence uniformly continuous) and all θ,θΔ,\theta,\theta^{\prime}\in\Delta,

limt|f(ωt,Ψ(t,ω)θ)f(ωt,Ψ(t,ω)θ)|=0\lim_{t\rightarrow\infty}|f(\omega_{t},\Psi(t,\omega)\theta)-f(\omega_{t},\Psi(t,\omega)\theta^{\prime})|=0

s\mathbb{P}_{s} almost surely, for μ\mu almost all sS.s\in S. Hence, for all θ,θΔ,\theta,\theta^{\prime}\in\Delta,

limt|Qtf(s,θ)Qtf(s,θ)|=0,\lim_{t\rightarrow\infty}|Q_{t}f(s,\theta)-Q_{t}f(s,\theta^{\prime})|=0, (10)

for μ\mu almost all sS.s\in S. Let now π\pi be an invariant probability of (Qt)t0.(Q_{t})_{t\geq 0}. Such a π\pi always exist because (Qt)t0(Q_{t})_{t\geq 0} if Feller on S×ΔS\times\Delta compact. To prove that π\pi is unique, assume that π\pi^{\prime} is another invariant probability. Then, writing π(f)\pi(f) for S×Δf(s,θ)π(dsdθ),\int_{S\times\Delta}f(s,\theta)\pi(dsd\theta),

π(f)π(f)=π(Qtf)π(Qtf)\pi(f)-\pi^{\prime}(f)=\pi(Q_{t}f)-\pi^{\prime}(Q_{t}f)
=S[Δ×Δ(Qtf(s,θ)Qtf(s,θ))π(dθ|s)π(dθ|s)]μ(ds)=\int_{S}\left[\int_{\Delta\times\Delta}(Q_{t}f(s,\theta)-Q_{t}f(s,\theta^{\prime}))\pi(d\theta|s)\pi(d\theta^{\prime}|s)\right]\mu(ds)

where for each sS,s\in S, π(.|s)\pi(.|s) (respectively π(.|s)\pi^{\prime}(.|s)) is a conditional distribution of π\pi (respectively π\pi^{\prime}) (see [7], Section 10.2). It then follows from (10) and dominated convergence that π(f)=π(f).\pi(f)=\pi^{\prime}(f). Thus π=π.\pi=\pi^{\prime}. This proves unique ergodicity.

Now, unique ergodicity and Feller continuity of (ωs,θs)s0(\omega_{s},\theta_{s})_{s\geq 0} imply that for every continuous function g:S×Δg:S\times\Delta\rightarrow{\mathbb{R}}

limt1t0tg(ωs,θs)𝑑s=g𝑑π\lim_{t\rightarrow\infty}\frac{1}{t}\int_{0}^{t}g(\omega_{s},\theta_{s})ds=\int gd\pi

s,θ\mathbb{P}_{s,\theta} almost surely for all s,θS×Δ.s,\theta\in S\times\Delta. (see e.g  [2], Proposition 7.1 for discrete time chains combined with Proposition 4.58 to handle continuous time). This concludes the proof of Proposition 1 with g(s,θ)=A(s)θ,𝟏)g(s,\theta)=\langle A(s)\theta,{\bf 1}\rangle).

Remark 2

By the multiplicative ergodic theorem, there exist numbers Λ1<<Λr,rd,\Lambda_{1}<\ldots<\Lambda_{r},r\leq d, called Lyapunov exponents, such that for μ\mathbb{P}_{\mu} almost all ω\omega and all xd{0},x\in{\mathbb{R}}^{d}\setminus\{0\},

limtlogΦ(t,ω)xt:=Λ(x,ω){Λ1,,Λr}.\lim_{t\rightarrow\infty}\frac{\log{\|\Phi(t,\omega)x\|}}{t}:=\Lambda(x,\omega)\in\{\Lambda_{1},\ldots,\Lambda_{r}\}.

The set of xdx\in{\mathbb{R}}^{d} for which Λ(x,ω)<Λr\Lambda(x,\omega)<\Lambda_{r} is a vector space (depending on ω\omega) having nonzero codimension. On the other hand, by what precedes, Λ(x,ω)=Λ\Lambda(x,\omega)=\Lambda for all x+d{0}.x\in{\mathbb{R}}^{d}_{+}\setminus\{0\}. It follows that Λ=Λr.\Lambda=\Lambda_{r}.

Proof of Proposition 3

For sS=/,s\in S={\mathbb{R}}/{\mathbb{Z}}, let ω[s]Ω\omega[s]\in\Omega be the path defined as

ωt[s]=s+t(𝗆𝗈𝖽 1).\omega_{t}[s]=s+t\,({\sf mod}\,1).

By Brouwer fixed point theorem, the map Ψ(1,ω[0]):ΔΔ\Psi(1,\omega[0]):\Delta\mapsto\Delta has a fixed point θ.\theta^{*}. Set θ(t)=Ψ(t,ω[0])(θ).\theta^{*}(t)=\Psi(t,\omega[0])(\theta^{*}). Then

θ(t+1)=Ψ(t,ω[1])Ψ(1,ω[0])(θ)=θ(t)\theta^{*}(t+1)=\Psi(t,\omega[1])\circ\Psi(1,\omega[0])(\theta^{*})=\theta^{*}(t)

proving that tθ(t)t\rightarrow\theta^{*}(t) is 11-periodic.

For all sSs\in S and θΔ\theta\in\Delta

limtΨ(t,ω[s])(θ)θ(t+s)=limtΨ(t,ω[s])(θ)Ψ(t,ω[s])(θ(s))=0,\lim_{t\rightarrow\infty}\|\Psi(t,\omega[s])(\theta)-\theta^{*}(t+s)\|=\lim_{t\rightarrow\infty}\|\Psi(t,\omega[s])(\theta)-\Psi(t,\omega[s])(\theta^{*}(s))\|=0,

by Lemma 6 applied with ω=ω[s]\omega=\omega[s] and θ=θ[s].\theta^{\prime}=\theta^{*}[s]. Observe here, that the conclusions of Lemma 6 hold with ω=ω[s]\omega=\omega[s] for all sSs\in S simply because ω[s]\omega[s] is 11-periodic.

Remark 3

The proof given here can be re-interpreted in the classical framework of Floquet’s theory used in [6]. By Floquet’s theorem, every solution to dydt=A(ωt[0])y\frac{dy}{dt}=A(\omega_{t}[0])y writes y(t)=P(t)etBy(0),y(t)=P(t)e^{tB}y(0), where PP is a 11-periodic matrix such that P(0)=Id.P(0)=Id. The matrix eBe^{B} has nonnegative entries, hence, by Perron-Frobenius theorem, a eigenvector y+d{0}.y^{*}\in{\mathbb{R}}^{d}_{+}\setminus\{0\}. The point θ\theta^{*} in the proof above is the projection of yy^{*} on the simplex, θ=y/|y|.\theta^{*}=y^{*}/|y^{*}|.

3 Proof of Propositions 4 and 5

For all T>0,T>0, let (QtT)t0(Q^{T}_{t})_{t\geq 0} denote the semigroup of (ωtT,θt)t0(\omega_{t}^{T},\theta_{t})_{t\geq 0} with ωtT=ωt/T\omega_{t}^{T}=\omega_{t/T} and (θt)t0(\theta_{t})_{t\geq 0} is solution to (3) when ωt\omega_{t} is replaced by ωtT.\omega^{T}_{t}. Using the notation of the preceding section one sees that

QtT(f)(s,θ)=𝔼s[f(ωt/T,Ψ(t,ωT)(θ))]Q^{T}_{t}(f)(s,\theta)=\mathbb{E}_{s}\left[f(\omega_{t/T},\Psi(t,\omega^{T})(\theta))\right]

for all fB(S×Δ).f\in B(S\times\Delta).

Proof of Proposition 4

For all θΔ,\theta\in\Delta, let F¯(θ)=SF(s,θ)μ(ds),\bar{F}(\theta)=\int_{S}F(s,\theta)\mu(ds), where FF is defined by (4). Let (Ψ¯t)t0(\bar{\Psi}_{t})_{t\geq 0} denote the semi-flow on Δ\Delta induced by the differential equation θ˙=F¯(θ).\displaystyle\dot{\theta}=\bar{F}(\theta). The following lemma follows from the averaging principle as given in Freidlin and Wentzell [8](Theorem 2.1, Chapter 7).

Lemma 8

For all δ>0\delta>0 and t0,t\geq 0,

limT0μ(supθΔ,0utΨ(u,ωT)(θ)Ψ¯u(θ)δ)=0.\lim_{T\rightarrow 0}\mathbb{P}_{\mu}\left(\sup_{\theta\in\Delta,0\leq u\leq t}\|\Psi(u,\omega^{T})(\theta)-\bar{\Psi}_{u}(\theta)\|\geq\delta\right)=0.

In particular, for all fC(Δ)f\in C(\Delta) and t0,t\geq 0,

limT0𝔼μ[fΨ(t,ωT)fΨ¯t]=0.\lim_{T\rightarrow 0}\mathbb{E}_{\mu}\left[\|f\circ\Psi(t,\omega^{T})-f\circ\bar{\Psi}_{t}\|_{\infty}\right]=0.

Proof:    We claim that

limRsupt0μ(|1Rtt+RF(ωs,θ)𝑑sF¯(θ)|δ)=0.\lim_{R\rightarrow\infty}\sup_{t\geq 0}\mathbb{P}_{\mu}\left(\left|\frac{1}{R}\int_{t}^{t+R}F(\omega_{s},\theta)ds-\bar{F}(\theta)\right|\geq\delta\right)=0. (11)

Indeed, by stationarity (invariance of μ\mathbb{P}_{\mu} for (𝚯t)t0(\mathbf{\Theta}_{t})_{t\geq 0}),

μ(|1Rtt+RF(ωs,θ)𝑑sF¯(θ)|δ)=μ(|1R0RF(ωs,θ)𝑑sF¯(θ)|δ)\mathbb{P}_{\mu}\left(\left|\frac{1}{R}\int_{t}^{t+R}F(\omega_{s},\theta)ds-\bar{F}(\theta)\right|\geq\delta\right)=\mathbb{P}_{\mu}\left(\left|\frac{1}{R}\int_{0}^{R}F(\omega_{s},\theta)ds-\bar{F}(\theta)\right|\geq\delta\right)

for all t0;t\geq 0; and the right hand term goes to 0,0, as R,R\rightarrow\infty, by ergodicity of μ.\mu.

By the averaging theorem (Theorem 2.1, Chapter 7 in [8]), condition (11) implies that for all δ>0,t0\delta>0,t\geq 0 and θΔ,\theta\in\Delta,

limT0μ(sup0utΨ(u,ωT)(θ)Ψ¯u(θ)δ)=0.\lim_{T\rightarrow 0}\mathbb{P}_{\mu}\left(\sup_{0\leq u\leq t}\|\Psi(u,\omega^{T})(\theta)-\bar{\Psi}_{u}(\theta)\|\geq\delta\right)=0. (12)

By Lipschitz continuity (see (7)) and Gronwall’s lemma,

sup0utΨ(u,ωT)(θ)Ψ(u,ωT)(θ)+Ψ¯t(θ)Ψ¯t(θ)2ec1tθθ\sup_{0\leq u\leq t}\|\Psi(u,\omega^{T})(\theta)-\Psi(u,\omega^{T})(\theta^{\prime})\|+\|\bar{\Psi}_{t}(\theta)-\bar{\Psi}_{t}(\theta^{\prime})\|\leq 2e^{c_{1}t}\|\theta-\theta^{\prime}\|

for all θ,θΔ.\theta,\theta^{\prime}\in\Delta. Fix ε<δ4ec1t\varepsilon<\frac{\delta}{4}e^{-c_{1}t} and let {B(θi,ε),i=1,,N}\{B(\theta_{i},\varepsilon),i=1,\ldots,N\} be a finite covering of Δ\Delta by balls of radius ε.\varepsilon. Then

sup0ut,θΔΨ(u,ωT)(θ)Ψ¯u(θ)maxi=1,,Nsup0utΨ(u,ωT)(θi)Ψ¯u(θi)+δ/2.\sup_{0\leq u\leq t,\theta\in\Delta}\|\Psi(u,\omega^{T})(\theta)-\bar{\Psi}_{u}(\theta)\|\leq\max_{i=1,\ldots,N}\sup_{0\leq u\leq t}\|\Psi(u,\omega^{T})(\theta_{i})-\bar{\Psi}_{u}(\theta_{i})\|+\delta/2.

Hence

μ(sup0ut,θΔΨ(u,ωT)(θ)Ψ¯u(θ)δ)i=1Nμ(sup0utΨ(u,ωT)(θi)Ψ¯u(θi)δ/2).\mathbb{P}_{\mu}\left(\sup_{0\leq u\leq t,\theta\in\Delta}\|\Psi(u,\omega^{T})(\theta)-\bar{\Psi}_{u}(\theta)\|\geq\delta\right)\leq\sum_{i=1}^{N}\mathbb{P}_{\mu}\left(\sup_{0\leq u\leq t}\|\Psi(u,\omega^{T})(\theta_{i})-\bar{\Psi}_{u}(\theta_{i})\|\geq\delta/2\right).

The right hand term goes 0 as T0T\rightarrow 0 by (12). \Box
We now prove the proposition. Let πT\pi^{T} be the invariant measure of (QtT)t0(Q^{T}_{t})_{t\geq 0} and let π0\pi^{0} be a limit point of (πT)T>0(\pi^{T})_{T>0} for the weak* topology, as T0.T\rightarrow 0. That is: πTnfπ0f\pi^{T_{n}}f\rightarrow\pi^{0}f for some sequence Tn0T_{n}\rightarrow 0 and all fC(S×Δ).f\in C(S\times\Delta).

Let p:S×ΔΔp:S\times\Delta\rightarrow\Delta be the projection defined as p(s,θ)=θp(s,\theta)=\theta and let π2T=πTp1\pi_{2}^{T}=\pi^{T}\circ p^{-1} be the second marginal of πT.\pi^{T}. Similarly, set π20=π0p1.\pi^{0}_{2}=\pi^{0}\circ p^{-1}.

For all fC(Δ)f\in C(\Delta) and t0,t\geq 0,

π2Tf=πT(fp)=πTQtT(fp)=S×Δ𝔼s[f(Ψ(t,ωT)(θ))]πT(dsdθ).\pi_{2}^{T}f=\pi^{T}(f\circ p)=\pi^{T}Q_{t}^{T}(f\circ p)=\int_{S\times\Delta}\mathbb{E}_{s}[f(\Psi(t,\omega^{T})(\theta))]\pi^{T}(dsd\theta).

Thus,

|π2Tfπ2T(fΨ¯t)|=|S×Δ𝔼s[f(Ψ(t,ωT)(θ))f(Ψ¯t(θ))]πT(dsdθ)||\pi_{2}^{T}f-\pi_{2}^{T}(f\circ\bar{\Psi}_{t})|=\left|\int_{S\times\Delta}\mathbb{E}_{s}[f(\Psi(t,\omega^{T})(\theta))-f(\bar{\Psi}_{t}(\theta))]\pi^{T}(dsd\theta)\right|
S𝔼s[fΨ(t,ωT)fΨ¯t]μ(ds)=𝔼μ[fΨ(t,ωT)fΨ¯t].\leq\int_{S}\mathbb{E}_{s}[\|f\circ\Psi(t,\omega^{T})-f\circ\bar{\Psi}_{t}\|_{\infty}]\mu(ds)=\mathbb{E}_{\mu}[\|f\circ\Psi(t,\omega^{T})-f\circ\bar{\Psi}_{t}\|_{\infty}].

Here we have used the fact that the first marginal of πT\pi^{T} is μ.\mu. Using Lemma 8, it comes that

π20f=π20(fΨ¯t)\pi_{2}^{0}f=\pi_{2}^{0}(f\circ\bar{\Psi}_{t})

for all t0.t\geq 0. This proves that π20\pi_{2}^{0} is invariant for {Ψ¯t}t0,\{\bar{\Psi}_{t}\}_{t\geq 0}, but since {Ψ¯t}t0\{\bar{\Psi}_{t}\}_{t\geq 0} has θ\theta^{*} as globally asymptotically stable equilibrium, necessarily π20=δθ.\pi_{2}^{0}=\delta_{\theta^{*}}. On the other hand, the first marginal of π0\pi^{0} is μ.\mu. Thus π0=μδθ.\pi^{0}=\mu\otimes\delta_{\theta^{*}}. This concludes the proof.

Proof of Proposition 5

Recall (see the proof of Lemma 7) that for all sS,s\in S, we let (Ψts)t0(\Psi^{s}_{t})_{t\geq 0} denote the semi-flow on Δ\Delta induced by the differential equation θ˙=F(s,θ).\displaystyle\dot{\theta}=F(s,\theta).

Let (Qt)t0)(Q^{\infty}_{t})_{t\geq 0}) denote the Markov semigroup on S×ΔS\times\Delta defined as

Qtf(s,θ)=f(s,Ψts(θ))Q^{\infty}_{t}f(s,\theta)=f(s,\Psi^{s}_{t}(\theta))

for all fB(S×Δ).f\in B(S\times\Delta).

Lemma 9

For all fC(S×Δ)f\in C(S\times\Delta) and t0t\geq 0

limTQtTfQtf=0.\lim_{T\rightarrow\infty}\|Q^{T}_{t}f-Q^{\infty}_{t}f\|_{\infty}=0.

Proof:    Let fC(S×Δ).f\in C(S\times\Delta). By uniform continuity of f,f, for every t>0t>0 and δ>0\delta>0 there exists α>0\alpha>0 such that

|f(ωtT,Ψ(t,ωT)θ)f(s,Ψts(θ))|δ+2f𝟏{d(ωtT,s)+Ψ(t,ωT)(θ)Ψts(θ)α}.|f(\omega^{T}_{t},\Psi(t,\omega^{T})\theta)-f(s,\Psi^{s}_{t}(\theta))|\leq\delta+2\|f\|_{\infty}{\bf 1}_{\large\{d(\omega^{T}_{t},s)+\|\Psi(t,\omega^{T})(\theta)-\Psi^{s}_{t}(\theta)\|\geq\alpha\}}.

Thus

QtTf(s,θ)Qtf(s,θ)𝔼s(|f(ωtT,Ψ(t,ωT)(θ)f(s,Ψts(θ))|)\|Q^{T}_{t}f(s,\theta)-Q^{\infty}_{t}f(s,\theta)\|\leq\mathbb{E}_{s}\left(|f(\omega^{T}_{t},\Psi(t,\omega^{T})(\theta)-f(s,\Psi^{s}_{t}(\theta))|\right)
δ+2f𝔼s(Ψ(t,ωT)(θ)Ψts(θ))+Pt/T(d(.,s))(s)α.\leq\delta+2\|f\|_{\infty}\frac{\mathbb{E}_{s}(\|\Psi(t,\omega^{T})(\theta)-\Psi^{s}_{t}(\theta)\|)+P_{t/T}(d(.,s))(s)}{\alpha}.

By Feller continuity, Pt/T(d(.,s))(s)0P_{t/T}(d(.,s))(s)\rightarrow 0 uniformly in sSs\in S as T.T\rightarrow\infty. This follows for example from Lemma 19.3 (F3F_{3}) in [10]. Now the estimate (9) applied with ωT\omega^{T} in place of ω,\omega, (Pt/T)t0(P_{t/T})_{t\geq 0} in place of (Pt)t0(P_{t})_{t\geq 0} and ε>t\varepsilon>t gives

supsS𝔼s(Ψ(t,ωT)(θ)Ψts(θ))c2t2δ(ε/T).\sup_{s\in S}\mathbb{E}_{s}(\|\Psi(t,\omega^{T})(\theta)-\Psi^{s}_{t}(\theta)\|)\leq c^{2}t^{2}\delta(\varepsilon/T). (13)

with δ(ε/T)0\delta(\varepsilon/T)\rightarrow 0 as T.T\rightarrow\infty. This concludes the proof. \Box
We can now prove Proposition 5. Let πT\pi^{T} be the invariant measure of (ωtT,θt)(\omega^{T}_{t},\theta_{t}) and let π\pi^{\infty} be a limit point of (πT)T>0(\pi^{T})_{T>0} for the weak* topology. That is πTnfπf\pi^{T_{n}}f\rightarrow\pi^{\infty}f for some sequence TnT_{n}\rightarrow\infty and all fC(S×Δ).f\in C(S\times\Delta). Then,

|πT(f)πT(Qtf)|=|πT(QtT(f)Qt(f))|QtT(f)Qt(f).|\pi^{T}(f)-\pi^{T}(Q_{t}^{\infty}f)|=|\pi^{T}(Q_{t}^{T}(f)-Q_{t}^{\infty}(f))|\leq\|Q_{t}^{T}(f)-Q_{t}^{\infty}(f)\|_{\infty}.

Thus, by Lemma 9, (ii), π(f)=π(Qt(f)).\pi^{\infty}(f)=\pi^{\infty}(Q_{t}^{\infty}(f)). Now for all s𝗌𝗎𝗉𝗉(μ)s\in\mathsf{supp}(\mu)

limtQt(f)(s,θ)=limtf(s,Ψts(θ))=f(s,θ(s)).\lim_{t\rightarrow\infty}Q_{t}^{\infty}(f)(s,\theta)=\lim_{t\rightarrow\infty}f(s,\Psi^{s}_{t}(\theta))=f(s,\theta^{*}(s)).

Thus, since π(𝗌𝗎𝗉𝗉(μ)×Δ)=1,\pi^{\infty}(\mathsf{supp}(\mu)\times\Delta)=1, it comes that

π(f)=SΔf(s,θ(s))π(dsdθ)=Sf(s,θ(s))μ(ds).\pi^{\infty}(f)=\int_{S}\int_{\Delta}f(s,\theta^{*}(s))\pi^{\infty}(dsd\theta)=\int_{S}f(s,\theta^{*}(s))\mu(ds).

This proves the first part of Proposition 5. The second part follows directly from the first one.

4 Concluding remarks

The results and proofs given here all rely on the assumption that (ωt)t0(\omega_{t})_{t\geq 0} is a Markov process. In particular, they do not apply to the case where tωtt\rightarrow\omega_{t} is a deterministic periodic signal with discontinuities. This situation is investigated in the preprint [4]. The recent preprint [16] provides a first order expansion of ΛT\Lambda^{T} when TT goes to 0.

Acknowledgment

This research is supported by the Swiss National Foundation grants 200020 196999 and 200020 219913. We thank Philippe Carmona and an anonymous referee for their careful reading and useful comments.

References

  • [1] L. Arnold, V. M. Gundlach, and L. Demetrius. Evolutionary formalism for products of positive random matrices. Ann. Appl. Probab., 4(3):859–901, 1994.
  • [2] M. Benaïm and T. Hurth. Markov Chains on Metric Spaces, A Short Course, volume 99 of Universitext. Springer, Cham, 2022.
  • [3] M. Benaïm, C. Lobry, T. Sari, and E. Strickler. Untangling the role of temporal and spatial variations in persistence of populations. Theoretical Population Biology, 154:1–26, 2023.
  • [4] M. Benaïm, C. Lobry, T. Sari, and E. Strickler. When can a population spreading across sink habitats persist ? working paper, https://hal.inrae.fr/hal-04099082, May 2023.
  • [5] M. Benaïm and E. Strickler. Random switching between vector fields having a common zero. Ann. Appl. Probab., 29(1):326–375, 2019.
  • [6] P. Carmona. Asymptotic of the largest Floquet multiplier for cooperative matrices. Annales de la Faculté des sciences de Toulouse: Mathématiques, Ser. 6, 31(4):1213–1221, 2022.
  • [7] R. M. Dudley. Real analysis and probability, volume 74 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2002. Revised reprint of the 1989 original.
  • [8] M. Freidlin and A. Wentzell. Random perturbations of dynamical systems, volume 260 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer, Heidelberg, third edition, 2012. Translated from the 1979 Russian original by Joseph Szücs.
  • [9] M. Hirsch and H. Smith. Chapter 4 monotone dynamical systems. volume 2 of Handbook of Differential Equations: Ordinary Differential Equations, pages 239–357. North-Holland, 2006.
  • [10] O. Kallenberg. Foundations of modern probability, volume 99 of Probability Theory and Stochastic Modelling. Springer, Cham, 2021. Third edition of the 1997 original.
  • [11] U. Krengel. Ergodic Theorems, volume 99 of De Gruyter Series in Mathematics. De Gruyter, 1985.
  • [12] T. Malik and H. Smith. Does dormancy increase fitness of bacterial populations in time-varying environments? Bulletin of mathematical biology, 70:1140–62, 06 2008.
  • [13] J. Mierczyński. Estimates for principal lyapunov exponents: A survey. Nonautonomous Dynamical Systems, 1(1), 2014.
  • [14] J. Mierczyński. Lower estimates of top Lyapunov exponent for cooperative random systems of linear ODEs. Proc. Amer. Math. Soc., 143(3):1127–1135, 2015.
  • [15] J. Mierczyński and W. Shen. Principal Lyapunov exponents and principal Floquet spaces of positive random dynamical systems. II. Finite-dimensional systems. J. Math. Anal. Appl., 404(2):438–458, 2013.
  • [16] P. Monmarch and E. Strickler. Asymptotic expansion of the invariant measurefor markov-modulated odes at high frequency, 2023. arXiv 2309.16464.
  • [17] D. Ruelle. Analyticity properties of the characteristic exponents of random matrix. Adv. Math., 32(3):68–80, 1979.
  • [18] E. Seneta. Non-negative matrices and Markov chains. Springer Series in Statistics. Springer, New York, 2006. Revised reprint of the second (1981) edition.