• Aucun résultat trouvé

A SUFFICIENT CONDITION FOR THE CONVERGENCE OF A FINITE MARKOV CHAIN

N/A
N/A
Protected

Academic year: 2022

Partager "A SUFFICIENT CONDITION FOR THE CONVERGENCE OF A FINITE MARKOV CHAIN"

Copied!
15
0
0

Texte intégral

(1)

THE CONVERGENCE OF A FINITE MARKOV CHAIN

ALINA NICOLAIE

We give a sufficient condition for the convergence of a finite state nonhomogeneous Markov chain in terms of convergence of a certain series.

AMS 2000 Subject Classification: 60J10.

Key words: simulated annealing, finite Markov chain.

1. PRELIMINARIES

Consider a finite Markov chain with state space S = {1, . . . , r} and transition matrices (Pn)n≥1. We shall refer to it as the finite Markov chain (Pn)n≥1. For all integersm≥0,n > m, define

Pm,n =Pm+1Pm+2· · ·Pn= ((Pm,n)i,j)i,j∈S. Assume that the limit

(1.1) lim

n→∞Pn=P

exists and that the limit matrixP hasp≥1 irreducible aperiodic closed classes and, perhaps transient states, so that it has the form

(1.2)

S1 0 · · · 0 0

0 S2 · · · 0 0

· · · ·

0 0 · · · Sp 0

L1 L2 · · · Lp T

 ,

where Si are ri×ri transition matrices, i = 1, p, associated with the p irre- ducible aperiodic closed classes,T concerns the transitions of the chain as long as it stays in the r−Pp

t=1rt transient states and the Li concern transitions from the transient states into the ergodic sets corresponding to Si,i= 1, p.

Such a class of Markov chains was, e.g., proposed as the mathematical model for simulated annealing, a stochastic algorithm for global optimization.

We refer to van Laarhoven and Aarts [8] for a general exposition and historical background.

MATH. REPORTS10(60),1 (2008), 57–71

(2)

Definition 1.1. A probability distribution µ= (µ1, . . . , µr) is said to be invariant with respect to an r×r stochastic matrix P ifµP =µ.

We shall need the following result

Theorem1.2. Consider a finite homogeneous Markov chain with state space S and transition matrix P of the form (1.2). Then

(1.3) lim

n→∞Pn=

Γ1 0 · · · 0 0

0 Γ2 · · · 0 0

· · · ·

0 0 · · · Γp 0

12 · · · Ωp 0

 ,

where

Γi=

µ(i)1 · · · µ(i)ri

· · · · µ(i)1 · · · µ(i)ri

, i= 1, p,

are strictly positive ri×ri matrices; each row of the matrixΓi is the invariant probability vector µ(i) = (µ(i)1 , . . . , µ(i)ri) with respect to the matrix Si, i= 1, p, and

i =

µ(i)1 zr1+r2+···+rp+1,i · · · µ(i)rizr1+r2+···+rp+1,i

· · · ·

µ(i)1 zr,i · · · µ(i)rizr,i

are (r −Pp

t=1rt) ×ri matrices, where zj,i= probability that the chain will enter and thus, will be absorbed in the set corresponding to Si given that the initial state is j from the set corresponding to T,j=Pp

t=0rt, r, i= 1, p [with convention r0 = 1].

Proof. For the form of Γi, i = 1, p, see, e.g., [5, p. 123] and for Ωi, i= 1, p, see, e.g., [7, p. 91].

Remark 1.3. Clearly,

(1.4) zj,i ≥0, ∀j=

p

X

t=0

rt, r, ∀i= 1, p, and

(1.5)

p

X

i=1

zj,i = 1, ∀j=

p

X

t=0

rt, r.

Theorem 1.4 (see, e.g., [6]). If P is the transition matrix of a finite Markov chain, then the multiplicity of the eigenvalue 1 is equal to the number of irreducible closed subsets of the chain.

(3)

Proof. See, e.g., [6, p. 126].

A vector x∈Cn will be understood as a column vector and x0 denotes the transpose of x.

Theorem1.5. Let A= −Ir+P with P of the form (1.2). Then there exists a nonsingular complex r×r matrix Q such that

(1.6) A=QJ Q−1,

where J is a Jordan r×r matrix, and Qreads as

Q=

1 0 · · · 0 · · ·

· · · ·

1 0 · · · 0 · · ·

0 1 · · · 0 · · ·

· · · ·

0 1 · · · 0 · · ·

0 0 · · · 0 · · ·

· · · ·

0 0 · · · 0 · · ·

0 0 · · · 1 · · ·

· · · ·

0 0 · · · 1 · · ·

zr1+r2+···+rp+1,1 zr1+r2+···+rp+1,2 · · · zr1+r2+···+rp+1,p · · ·

· · · ·

zr,1 zr,2 · · · zr,p · · ·

 ,

where the first column contains 1 in the 1, r1 rows, the next p−1 columns contains1in theri−1+ 1, rirows,i= 2, p, and the lastr−pcolumns comprise complex numbers. For zj,i, j=Pp

t=0rt, r, i= 1, p, we have the meaning given in Theorem 1.2. The inverse Q−1 has the form

Q−1=

µ(1)1 · · · µ(1)r1 0 · · · 0 0 · · · 0 0 · · · 0

· · · ·

0 · · · 0 0 · · · 0 µ(p)1 · · · µ(p)rp 0 · · · 0

qp+1,1 · · · qp+1,r

· · · · qr,1 · · · qr,r

 ,

where µ(i) are the invariant probability vectors with respect toSi, i= 1, p, and the last r−p rows comprise complex numbers.

Proof. For the existence ofQsee, e.g., [4, p. 126]. See also [1]. We shall need some spectral properties of A. We have

(1.7) λ1 = 0

(4)

is an eigenvalue of A whose algebraic multiplicity is equal to the geometric multiplicity and equal to p. All other distinct eigenvalue λ2, . . . , λl+s of A satisfy

(1.8) |λi+ 1|<1, i= 2, l+s, (1.9) Reλi <0, i= 2, l+s.

Indeed, by Theorem 1.4, the characteristic polynomial of P has the form det(P −ϑIr) = (ϑ−1)pR(ϑ),

where grad R=r−p and R(1)6= 0. Replacing P =Ir+A we obtain det(A−λIr) =λpR(λ+ 1),

where λ=ϑ−1, and so (1.7) follows.

On the other hand, if we set ϑi all other distinct eigenvalues different from 1 of P, with left eigenvectors ϕi,i= 2, l+s, becauseP is stochastic we have |ϑi|<1,i= 2, l+s, and

ϕiP =ϑiϕi, i= 2, l+s.

By subtracting ϕi from both members we obtain ϕiA= (ϑi−1)ϕi, i= 2, l+s.

Hence λi = ϑi −1, i = 2, l+s, are eigenvalues of A. Then (1.8) and (1.9) follow. From (1.7) (see [4, pp. 129–131]) we have

J =

J1 0 · · · 0 0 J2 · · · 0

· · · ·

0 0 · · · Jl 0 · · · ·

0 0 · · · 0 Jl+1 · · · ·

· · · ·

0 0 · · · 0 0 · · · Jl+s

 ,

where J1 = 0p, i.e., the p ×p zero matrix, Jk is a diagonal mk×mk ma- trix with entries the eigenvalues λk, k = 2, l, whose algebraic and geometric multiplicities are equals, and

Jl+i=

λl+i ε(i)1 0 · · · 0

0 λl+i ε(i)2 · · · 0 0

· · · · 0 0 · · · λl+i ε(i)m

l+i−1

0 0 · · · 0 λl+i

(5)

are ml+i×ml+i matrices corresponding to eigenvalues whose geometric mul- tiplicities are smaller then their algebraic multiplicities and ε(i)t ∈ {0,1}, t= 1, ml+i−1,i= 1, s. Clearly, p+m2+· · ·+ml+s=r.

A key step in obtaining the firstpcolumns ofQis the remark that writing (1.6) as AQ =QJ we see that A applied to the tth column ofQ yields zero vector, ∀t= 1, p, hence the vectors corresponding to the first p columns ofQ are right eigenvectors associated with the eigenvalue λ1 = 0 ofA.

Now, we shall determine the right eigenspace corresponding to λ1 by solving the linear system

(1.10) (A−0Ir)X= 0r

which is equivalent toP X =X, whereX= (x1, . . . , xr).

The firstPp

t=1rt rows of the system (1.10) are equivalent to (Si−Iri)·(xr0+r1+···+ri−1, . . . , xr1+···+ri−1+ri) = 0, i= 1, p, with the general solution

(1.11) (xr0+r1+···+ri−1, . . . , xr1+···+ri) =αi(1,1, . . . ,1), αi∈C, i= 1, p.

For the next components xr1+···+rp+1, . . . , xr we note that P X =X⇒PkX =X, ∀k≥1.

For the last r−Pp

t=1rt components, the above equation amounts to

r

X

j=1

(Pk)r1+···+rp+i,jxj =xr1+···+rp+i, i= 1, r−

p

X

t=1

rt, k≥1, Equivalently,

p−1

X

t=0

r1+···+rt+1

X

j=r0+···+rt

(Pk)r1+···+rp+i,jxj+

r

X

j=r1+···+rp+1

(Pk)r1+···+rp+i,jxj =xr1+···+rp+i. For j=Pp

t=0rt, rand i= 1, r−Pp

t=1rtwe have [see 5, p. 91]

(1.12) lim

k→∞(Pk)r1+···+rp+i,j = 0 and for i= 1, r−Pp

t=1rt from representation (1.3) we get

k→∞lim(Pk)r1+···+rp+i,r0+···+rs−1+t−1(s)t zr1+···+rp+i,s, s= 1, p, t= 1, rs. So, letting k → ∞ in the last sum and using the fact that µ(i), i= 1, p, are probability vectors, we obtain

p

X

j=1

zr1+r2+···+rp+i,jαj =xr1+···+rp+i, i= 1, r−

p

X

t=1

rt.

(6)

We thus obtain the general solution

p

X

j=1

αj(0, . . . ,0,1, . . . ,1,0, . . . ,0, zr1+···+rp+1,j, zr1+···+rp+2,j, . . . , zr,j)0, of (1.10), with αj ∈ C, j = 1, p, where the components of the row vectors corresponding to r0+· · ·+rj−1, r1+· · ·+rj are 1 and those corresponding to the remaining ones from 1,Pp

t=1rt, are 0,j = 1, p.

Next, we pick particular values for the parametersαi,i= 1, p, to obtain plinear independent vectors; choosingα1= 1,α2 =· · ·=αp = 0, thenα2= 1, α13 =· · ·=αp= 0 etc., we obtain the first pcolumns of the matrix Q.

A key step in obtaining the firstprows ofQ−1 is the remark that writing (1.6) as Q−1A=J Q−1 which means that thetth row ofQ−1, applied toA is equal to the tth rows of J applied to Q−1,∀t= 1, p. Since the matrix J Q−1 has the first p rows equal to the zero vector, we conclude that the tth row of Q−1 is a left eigenvector of A associated withλ1= 0, ∀t= 1, p.

Next, we shall determine the left eigenspace corresponding toλ1 by solv- ing the linear system

(1.13) Y0(A−0Ir) = 00r

which is equivalent to Y0P = Y0, where Y0 = (y1, . . . , yr). The first Pp t=1rt equations of system (1.13) are equivalent to

(yr0+r1+···+ri−1, . . . , yr1+···+ri)0·(Si−Iri) = 0, i= 1, p, with the general solution

(yr0+r1+···+ri−1, . . . , yr1+···+ri) =βi(i)1 , . . . , µ(i)ri), βi ∈C, i= 1, p.

For the next components yr1+···+rp+1, . . . , yr we note that Y0P =Y0 ⇒Y0Pk=Y0, ∀k≥1.

For the last r−Pp

t=1rt components, the above equation amounts to

r

X

j=1

yj(Pk)j,r1+···+rp+i =yr1+···+rp+i, i= 1, r−

p

X

t=1

rt, k ≥1, Equivalently,

p−1

X

t=0

r1+···+rt+1

X

j=r0+···+rt

yj(Pk)j,r1+···+rp+i+

r

X

j=r1+···+rp+1

yj(Pk)j,r1+···+rp+i =yr1+···+rp+i. By letting k→ ∞in the last sum, and using (1.12) and (1.3) we get

yr1+···+rp+i= 0, i= 1, r−

p

X

t=1

rt.

(7)

We thus obtain the general solution

p

X

j=1

βj(0, . . . , µ(j)1 , . . . , µ(j)rj,0, . . . ,0)0,

of (1.13), with βj ∈ C, j = 1, p, where the components of the vector corre- sponding to the subscripts r0+· · ·+rj−1, r1+· · ·+rj are the components of the invariant vector associated with Sj and the others are equals to 0.

Finally, we pick particular values for the parametersβi,i= 1, p, to obtain plinear independent vectors; choosingβ1= 1,β2 =· · ·=βp = 0, thenβ2= 1, β13 =· · ·=βp = 0, etc., we obtain the firstprows of the matrixQ−1. If A = (Aij) is an m ×n matrix, then (see [4, p. 295]) define the matrix norm

|||A|||= max

i=1,m n

X

j=1

|Aij|.

For M ⊆ {1, . . . , m},N ⊆ {1, . . . , n},M, N 6=∅we also define AM×N = (Aij)i∈M,j∈N.

Theorem1.6 (see, e.g., [6]). Let (an)n≥0 be a sequence of real numbers convergent to 0 and P

n=0bn an absolute convergent series. Then

n→∞lim

n

X

i=0

aibn−i = 0.

Proof. See, e.g., [6, pp. 34–36].

Theorem1.7 (see, e.g., [3]). Let ||| · |||be a matrix norm. Let (Un)n≥1

be a sequence of matrices such that |||Un||| ≤ 1, ∀n≥ 1. Then the following statements are equivalent.

(i)Q

n=rUn converges, ∀r≥1.

(ii)Q

n=1[Un+An]converges for all sequences (An)n≥1, when P

n=1|||An|||<∞.

Proof. See, e.g., [3, p. 101].

Theorem 1.8 (see, e.g., [4]). Let A be an n×n matrix. Then A is invertible if there exists a matrix norm ||| · ||| such that |||In−A||| < 1. If so, then

A−1 =

X

k=0

(In−A)k. Proof. See, e.g., [4, p. 301].

(8)

Theorem1.9 (see, e.g., [4]). (i) If ||| · ||| is an arbitrary matrix norm, then |||I||| ≥1.

(ii) Let A be an n×n matrix. If ||| · ||| is a matrix norm such that

|||A|||<1, then

|||(In−A)−1||| ≤ |||In||| −(|||In||| −1)|||A|||

1− |||A||| .

Proof. See, e.g., [4, pp. 301–302].

2. A CONVERGENCE RESULT

In this section we give our main result following an idea of a theorem of Gidas [1]. Our result is a sufficient condition for the convergence of a finite state nonhomogeneous Markov chain in terms of convergence of a certain series.

Definition 2.1 ([9]). A finite Markov chain (Pn)n≥1 is said to be conver- gent if ∀m≥0 the sequence (Pm,n)n>m converges.

Definition 2.2 ([9]). Two Markov chains (Pn)n≥1 and (Pn0)n≥1 are said to be equivalent if

X

n≥1

|||Pn−Pn0|||<∞.

Theorem 2.3 ([2]). For two equivalent chains (Pn)n≥1 and (Pn0)n≥1

we have

∃ lim

n→∞Pm,n if and only if ∃ lim

n→∞Pm,n0 , ∀m≥1.

Consequently, two equivalent stochastic chains converge or diverge simultane- ously.

Proof. See [2].

The general results from Theorem 2.4 below are new. For the special case p= 1 see also [5, p. 226].

Theorem 2.4. Let (Pn)n≥1 be a nonhomogeneous Markov chain with state space S such thatPn→P as n→ ∞. Suppose that P has exactly p≥1 recurrent aperiodic classes Si, i= 1, p, and, possibly, transient states, i.e., P is of the form (1.2). Let µ(i) be the invariant probability vectors with respect to Si, i = 1, p and zj,i, j =Pp

t=1rt+ 1, r, i = 1, p, as in Theorem 1.2. Let Vn=Pn−P, ∀n≥1, where lim

n→∞Vn= 0. LetQ and Q−1 as in Theorem 1.5.

(9)

Then

n→∞lim(Pm,n)i,j = 0, i∈S, j=

p

X

t=0

rt, r, ∀m≥0.

If, moreover,

(2.1)

X

n=1

|||(Q−1VnQ)S×M|||<∞, where M ={1, . . . , p}, then the chain (Pn)n≥1 is convergent.

Proof. By the Chapman-Kolmogorov equation we have Pm,n =Pm,n−1Pn, 0≤m < n.

(with the convention Pn,n = Ir). Subtracting Pm,n−1 from both members, we obtain

(2.2) Pm,n−Pm,n−1 =Pm,n−1[−Ir+Pn], 0≤m < n.

Setting

(2.3) x(i)n =x(i)n (m) = ((Pm,n)i,1, . . . ,(Pm,n)i,r), 0≤m < n, and

x(i)m =ei, i∈S,

where (ei)i=1,r is the canonical basis of the linear spaceRr, then equation (2.2) reads as

(2.4) x(i)n −x(i)n−1 =x(i)n−1[−Ir+Pn], 0≤m < n, i∈S.

We remark that the x(i)n defined in (2.3) are solutions of equations of the type (2.5) xn−xn−1 =xn−1[−Ir+Pn], n≥1,

with

xn= (xn,1, . . . , xn,r), n≥0, under the conditions

(2.6) xn,i≥0, i= 1, r,

r

X

i=1

xn,i= 1, n≥0.

We are interested in the asymptotic behaviour of (2.5) under conditions (2.6).

Setting

A=−Ir+P

we can benefit of the result given in Theorem 1.5. Also, setting

(2.7) yn=xnQ, n≥0,

and

(2.8) Ven=Q−1VnQ, n≥1,

(10)

equation (2.5) becomes

(2.9) yn−yn−1 =yn−1J+yn−1Ven, n≥1.

From (2.7) we have xn=ynQ−1, hence (2.10) xn,i(1)i yn,1+

r

X

j=p+1

yn,jqji, i= 1, r1,

(2.11) xn,r1+···+rt−1+i(t)i yn,t+

r

X

j=p+1

yn,jqj,r1+···+rt−1+i, t= 2, p, i= 1, rt,

(2.12) xn,i=

r−p

X

j=1

yn,p+jqp+j,i, i=

p

X

t=0

rt, r.

Next, equations (2.9) amount to yn,i−yn−1,i=

p

X

j=1

yn−1,j(Ven)j,i+

r

X

j=p+1

yn−1,j(Ven)j,i, i= 1, p,

yn,i−yn−1,i2yn−1,i+

r

X

j=1

yn−1,i(Ven)j,i, i=p+ 1, p+m2,

yn,i−yn−1,ityn−1,i+

r

X

j=1

yn−1,j(Ven)j,i, i=p+

t−1

X

s=2

ms+ 1, p+

t

X

s=2

ms, t= 3, l, yn,p+Pl

s=2ms+1−yn−1,p+Pl

s=2ms+1l+1yn−1,p+Pl

s=2ms+1+ +

r

X

j=1

yn−1,j(Ven)j,p+Pl

s=2ms+1, yn,i−yn−1,il+1yn−1,i(1)k yn−1,i−1+ +

r

X

j=1

yn−1,j(Ven)j,i, i=p+

l

X

t=2

mt+ 2, p+

l+1

X

t=2

mt, k= 1, ml+1−1, and similar equations for i=p+Pl+1

t=2mt+ 1, r.

Next, we shall study the asymptotic behaviour ofyn,i,i=p+ 1, r. The boundedness of yn, n≥0, and the fact that lim

n→∞Vn= 0, allow one to assert that

(2.13) Wn,i:=

r

X

j=1

yn−1,j(Ven)j,i→0 as n→ ∞, i=p+ 1, r.

(11)

In the above system we shall be concerned with the equations corresponding toi∈ {p+ 1, . . . , p+Pl

t=2mt+ 1}. Equivalently, we can write yn,i= (λ2+ 1)yn−1,i+Wn,i, n≥1, i=p+ 1, p+m2, respectively,

yn,i= (λt+ 1)yn−1,i+Wn,i, n≥1, t= 3, l, i=p+

t−1

X

s=2

ms+ 1, p+

t

X

s=2

ms, Therefore,

(2.14) |yn,i| ≤ |y0,i| |λ2+ 1|n+

n

X

s=1

|Ws,i| |λ2+ 1|n−s, n≥1, i=p+ 1, p+m2. and similar equations do hold for i=p+Pt−1

s=2ms+ 1, p+Pt

s=2ms,t= 3, l.

By Theorem 1.6, using P

n=02+ 1|n<∞(see (1.8)) and (2.13), the second term on the right side of (2.14) converges to zero asn→ ∞. Using also (1.8), we get

n→∞lim yn,i= 0, i=p+ 1, p+m2, regardless of the initial data y0 and, similarly,

(2.15) lim

n→∞yn,i= 0, i=p+m2+ 1, p+

l

X

t=2

mt+ 1.

Next, we pay attention to the equations corresponding toifromp+Pl

t=2mt+2 to p+Pl+1

t=2mt. The case ε(i)k = 0 for some k = 1, ml+1−1 is similar to the preceding one. We are interested now in cases for whichε(i)k = 1. As in (2.14), we obtain

(2.16) |yn,i| ≤ |y0,i| |λl+1+ 1|n+

n

X

s=1

[|ys,i−1|+|Ws,i|]|λl+1+ 1|n−s. For i= p+Pl

t=2mt+ 2, p+Pl+1

t=2mt, the first term on the right hand side of (2.16) converges to 0 as n→ ∞. SinceP

n=0l+1+ 1|n <∞ (see (1.8)), by Theorem 1.6 (see (1.8), (2.13) and (2.15)), the second term from (2.16) converges to 0 as n→ ∞ fori=p+Pl

t=2mt+ 2. Hence

n→∞lim yn,p+m2+···+ml+2 = 0.

Similarly,

n→∞lim yn,i= 0, i=p+

l

X

t=2

mt+ 3, p+

l+1

X

t=2

mt.

(12)

We can thus conclude that

(2.17) lim

n→∞yn,i= 0, i=p+ 1, r, regardless of the initial data y0.

Setting

Rn,i=

r

X

j=p+1

yn,j(Ven+1)j,i, i= 1, p, n≥0.

equation (2.9) for the first p components becomes

(2.18) yn,i−yn−1,i=yn−1,1(Ven)1,i+yn−1,2(Ven)2,i+· · ·+yn−1,p(Ven)p,i+Rn−1,i, i= 1, p. Letm≥0 andi∈S be arbitrarily fixed. Let

xn=x(i)n , n > m,

where x(i)n are defined like in (2.3). It follows using (2.17) and (2.12), that

(2.19) lim

n→∞xn,k = 0, k=

p

X

t=0

rt, r.

and the first conclusion of theorem. In matrix notation, system (2.18) becomes

 yn,1

yn,2

· · · yn,p

0

=

 yn−1,1

yn−1,2

· · · yn−1,p

0

Cn+

Rn−1,1

Rn−1,2

· · · Rn−1,p

0

,

where

Cn=

1 + (Ven)1,1 (Ven)1,2 · · · (Ven)1,p (Ven)2,1 1 + (Ven)2,2 · · · (Ven)2,p

· · · ·

(Ven)p,1 (Ven)p,2 · · · 1 + (Ven)p,p

 .

Setting

Y(n) =

 yn,1

· · · yn,p

 and R(n) =

 Rn,1

· · · Rn,p

, n > m, the matrix equation above becomes

(Y(n))0 = (Y(n−1))0Cn+ (R(n−1))0, n > m.

The solution of this matrix equation is (2.20)

(Y(n))0 = (Y(m))0

n

Y

s=m+1

Cs+

(R(n−1))0+

n−2

X

s=m

(R(s))0

n

Y

t=s+2

Ct

, n > m.

(13)

Setting

Cn=Ip+Bn, n > m, we have

|||Bn|||= max

k=1,p p

X

j=1

|(Ven)k,j|, n > m.

The convergence ofQn

s=m+1Csfrom (2.20) asn→ ∞follows from Theo- rem 1.7. We shall now prove that the second term from (2.20) converges as n→ ∞, too. It follows from (2.1) that

∃n0 > m+ 2 such that|||Ip−Cn|||=|||Bn|||<1, n≥n0. So, by Theorem 1.8 we have

∃(Cn)−1, ∀n≥n0 and (Cn)−1=

X

t=0

[Ip−Cn]t=

X

t=0

[−Bn]t, ∀n≥n0, and we can thus write

n−2

X

s=m

(R(s))0

n

Y

t=s+2

Ct+ (R(n−1))0 =

n0−2

X

s=m

(R(s))0

n

Y

t=s+2

Ct+

+

n−1

X

s=n0−1

(R(s))0

s+1

Y

l=n0

(Cs+n0+1−l)−1

n

Y

t=n0

Ct, ∀n≥n0. To prove that the expression above converges amounts to show that

T(n) :=

n

X

s=n0−1

(R(s))0

s+1

Y

l=n0

(Cs+n0+1−l)−1, n≥n0, converges. From the boundedness of (yn)n and condition (2.1) we have (2.21)

X

n=1

|||Rn|||<∞.

Let u∈Nand n≥n0. We have

|||T(n+u)−T(n)|||=

=

R(n+1)

n+2

Y

l=n0

(Cn+1+n0+1−l)−1+· · ·+R(n+u)

n+u+1

Y

l=n0

(Cn+u+n0+1−l)−1

≤ |||R(n+ 1)|||

n+2

Y

l=n0

|||(Cl)−1|||+· · ·+|||R(n+u)|||

n+u+1

Y

l=n0

|||(Cl)−1|||

(14)

expression which, by Theorem 1.9, is smaller then

|||R(n+ 1)||| n+2

Q

l=n0

[1− |||Bl|||]

+· · ·+ |||R(n+u)||| n+u+1

Q

l=n0

[1− |||Bl|||]

≤(|||R(n+ 1)|||+· · ·+|||R(n+u)|||) 1 cn0, where cn0 :=Q

l=n0

1− |||Bl|||

∈(0,1]. The inequalities above and (2.21) implies that (T(n))n is a Cauchy sequence. Therefore it is convergent. This implies that lim

n→∞yn,t,∀t= 1, p exists and, using also (2.10), (2.11) and (2.17), that lim

n→∞xn,k, k= 1,Pp

t=1rt exists. Sincei∈S and m ≥0 were arbitrarily chosen, it follows using also (2.19) that lim

n→∞Pm,n exists for allm≥0, i.e., the chain (Pn)n≥1 is convergent.

Example 2.5. Consider the chain (Pn)n≥1 given by

Pn=

1

2n1 12+ n+11 n(n+1)1 1−n+11 n+11 0

0 0 1

, ∀n≥2,

and P1 = I3. The chain (Pn)n≥1 is convergent because the conditions of Theorem 2.4 are fulfilled.

Acknowledgements. The author wishes to thank Dr. Udrea P˘aun for the interest he showed in this paper.

REFERENCES

[1] B. Gidas, Nonstationary Markov chains and convergence of the annealing algorithm.

J. Statist. Phys.39(1985), 73–131.

[2] J. Hajnal,Weak ergodicity in nonhomogeneous Markov chains.Proc. Cambridge Philos.

Soc.54(1958), 233–246.

[3] D.J. Hartfiel,Nonhomogeous Matrix Products.World Scientific, River Edge, 2002.

[4] R.A. Horn and C.R. Johnson,Matrix Analysis.Cambridge Univ. Press, New York, 1985.

[5] M. Iosifescu, Finite Markov Processes and Their Applications. Wiley, Chichester and Editura Tehnic˘a, Bucure¸sti, 1980; corrected republication by Dover, Mineola, N.Y., 2007.

[6] D.L. Isaacson and R.W. Madsen,Markov Chains:Theory and Applications. Wiley, New York, 1976.

[7] S. Karlin and H.M. Taylor,A First Course in Stochastic Processes.Academic Press, New York, 1975.

(15)

[8] P.J.M. van Laarhoven and E.H.L. Aarts,Simulated Annealing:Theory and Applications.

Reidel, Dordrecht, 1987.

[9] A. Mukherjea and R. Chaudhuri,Convergence of non-homogeneous stochastic chains, II.

Math. Proc. Cambridge Philos. Soc.90(1981), 167–182.

Received 16 March 2007 “Transilvania” University of Bra¸sov

Department of Mathematical Analysis and Probability Str. Iuliu Maniu 50

500157 Bra¸sov, Romania alinanicolae@unitbv.ro

Références

Documents relatifs

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

Rademacher, Uber partielle und totale Differenzierbarkeit von Functionen mehrerer Variabeln und uber die Transformation der Doppelintegrale.. Weaver,

To study the well definiteness of the Fr´echet mean on a manifold, two facts must be taken into account: non uniqueness of geodesics from one point to another (existence of a cut

However, there are few results concerning the convergence analysis of numerical methods for coagulation and fragmentation models (see [17] for quasi Monte-Carlo methods)....

The first theorem in [1] states that, if the Markov chain is strongly lumpable, then the eigenvalues of the probability transition matrix P b of agg(α, A, P ) (a partition P is

If the transition matrix of the Markov chain (X n ) is ϕ-observable, then the Rogers-Pitman condition is essentially a criterion for weak lumpability prop- erty to hold.... It has

Markov chain except that the probability measure is replaced by a complex amplitude measure and the transition probability matrix is replaced by a.. transition

The second procedure is of theoretical interest and leads to a general model selection theorem from which we derive rates of convergence over a very wide range of possibly