Home A family of commuting contraction semigroups on l 1 ( N ) and l ∞ ( N )
Article Open Access

A family of commuting contraction semigroups on l 1 ( N ) and l ( N )

  • Ernest Nieznaj EMAIL logo
Published/Copyright: July 10, 2025

Abstract

A family of commuting contraction semigroups ( P n ( t ) ) n N , defined on l 1 ( N ) , is presented. For this family, the product semigroup n = 1 P n ( t ) exists and has bounded generator. The infinite product of the corresponding family of adjoint semigroups ( P n ( t ) ) n N , defined on l ( N ) , also exists and its generator is bounded. Explicit formulae for these generators are also given.

MSC 2010: 47D03; 60J27; 60J35

1 Introduction

It is well known that if B is a bounded linear operator in a Banach space ( X , ) , then the family of operators ( P ( t ) ) t 0 given by

P ( t ) = e t B = k = 0 ( t B ) k k ! , t 0 ,

forms a strongly continuous semigroup of bounded operators in X [1, pp. 67, 251]. This means that it satisfies P ( t + s ) = P ( t ) P ( s ) , for all t , s 0 , and lim t 0 + P ( t ) x = x , for every x X . P ( 0 ) denotes I X , the identity operator in X . Moreover, the semigroup so defined is in fact uniformly continuous, that is lim t 0 + P ( t ) I X = 0 , and B is its infinitesimal generator, see p. 251, ibid.

In this article, we provide a family of bounded linear operators ( B n ) n N , defined on l 1 ( N ) , such that the corresponding family of semigroups ( P n ( t ) ) n N , where P n ( t ) = e t B n , satisfies the following conditions:

(1.1) P n ( t ) 1 , P m ( t ) P n ( s ) = P n ( s ) P m ( t ) ,

for any m , n N and t , s 0 . Throughout we use the convention that N = { 1 , 2 , 3 , } .

In other words, ( P n ( t ) ) n N is a sequence of commuting contraction semigroups on l 1 ( N ) . The reason for considering such semigroups is that then the product n = 1 N P n ( t ) , for any N 2 , is also a semigroup of contractions and its generator equals n = 1 N B n [2, p. 24]. We also give conditions under which the infinite product of those semigroups exists and its generator A = n = 1 B n is bounded, i.e.,

(1.2) n = 1 P n ( t ) lim N n = 1 N P n ( t ) = e t A , t 0 .

The limit in (1.2) is in the strong topology. An explicit formula for A is given in Theorem 3.3, which follows from a general convergence theorem for semigroups satisfying (1.1) proved in [3]. Some results, based on this theorem, have been obtained recently, see [4,5] and [6, p. 85].

This article is organized as follows. In Section 2, we summarize the previous results and define operators ( B n ) n N . Two specific examples are also provided. Main results of this article are contained in Section 3. Section 3.1 is devoted to the proof of (1.1). The Lumer-Phillips theorem is used to show that the B n s generate semigroups of contractions, see Lemma 3.1. In Section 3.2, we prove (1.2), and in Section 3.3, an analogous result is obtained for ( P n ( t ) ) n N , where P n ( t ) denotes the adjoint semigroup to P n ( t ) . An auxiliary lemma is proved in Section A.

2 A family of operators ( B n ) n N

2.1 Motivation from the theory of Markov chains

Recall that l 1 ( N ) is the Banach space of all absolutely summable sequences. This means that x = ( ξ i ) i N is an element of l 1 ( N ) if and only if i N ξ i < and then x l 1 ( N ) = i N ξ i . It is a separable space and any x l 1 ( N ) can be written as i N ξ i e i , where { e i } i N is the standard Schauder basis in this space, i.e., e i = ( , 0 , 1 , 0 , ) with 1 in the i th coordinate. For more details about l 1 ( N ) , see [7, Chapter 7].

In [4,5], the sequence of operators ( B n ) n N , defined in l 1 ( N ) , was considered, where

(2.1) B n x = ( η i ) i N = β n ξ i + α n ξ i + 2 n 1 , if i mod 2 n S n 1 , α n ξ i + β n ξ i 2 n 1 , if i mod 2 n S n 2 .

The sets S n 1 , S n 2 in (2.1) partition the set { 0 , 1 , 2 , , 2 n 1 } and are given by

(2.2) S n 1 = { 1 , 2 , 3 , , 2 n 1 } , S n 2 = { 0 , 2 n 1 + 1 , 2 n 1 + 2 , , 2 n 1 } .

The sequences ( β n ) n N and ( α n ) n N are assumed to be sequences of real positive numbers. The operators ( B n ) n N are in fact isomorphic images of the operators from [6, p. 83] and are generators of the transition semigroups associated with two-state Markov chains. These were used by Blackwell [8] to construct a Markov chain whose all states are instantaneous. To be more precise, he considered a Markov process ( X ( t ) ) t 0 defined as follows:

(2.3) X ( t ) = ( X 1 ( t ) , X 2 ( t ) , X 3 ( t ) , ) , t 0 ,

where X 1 ( t ) , X 2 ( t ) , X 3 ( t ) , is an infinite sequence of mutually independent, two-state Markov chains. The set of states of every component of X ( t ) is { 0 , 1 } and the transition semigroup e t B n , associated with X n ( t ) , is thus characterized by β n and α n . An explicit formula for this semigroup exists [6, p. 83].

Let S denote the state space of X ( t ) . Then S { 0 , 1 } , and this last set is uncountable by the known Cantor’s diagonal argument. So in general S may be uncountable. However, if we assume that the sequences ( β n ) n N , ( α n ) n N satisfy (the first condition by Blackwell)

(2.4) n = 1 β n α n + β n < + ,

then, with probability one, elements of S are sequences with only a finite number of 1’s. Therefore, S is countable and a bijection between S and N exists, and it is given in [5]. Moreover, the condition (2.4) also guarantees that the transition semigroup P ( t ) , associated with X ( t ) , is well defined. This means that it is a strongly continuous semigroup of contractions, and it is given by

(2.5) P ( t ) x = lim N n = 1 N e t B n x , x l 1 ( N ) , t 0 .

The fact that n = 1 e t B n is the transition semigroup of X ( t ) follows from the independence of its components. The second condition introduced by Blackwell requires that

(2.6) n = 1 β n = + ,

and it implies that all states of X ( t ) are instantaneous, i.e., the Q -matrix of the process has −∞ in all the diagonal entries. Incidentally, the condition (2.6), together with (2.4), implies also that n = 1 α n = + .

From these two conditions, it follows the generator of P ( t ) , denoted as A gen , is densely defined and unbounded. In [5], it was found, by the application of [3, Proposition 2.7], that A gen = A ¯ , where

(2.7) A x lim N n = 1 N B n x , x D ( A ) l 1 ( N ) .

Here, D ( A ) denotes the domain of A . Because A is not bounded, D ( A ) l 1 ( N ) . For example, any x with a finite number of non-zero components does not belong to D ( A ) . It should be added that in this case, the operator A defined as the limit of n = 1 N B n , when N , is not densely defined in l ( N ) . Thus, we cannot apply the proposition from [3] and conclude the continuity of the adjoint semigroup P ( t ) .

Blackwell was interested in the case in which all states of (2.3) are instantaneous, and his example is one of the few such known, see [6, p. 82] and the references contained therein. This chain is so interesting because some explicit calculations, just described earlier, can be carried out that are related to it.

If instead of (2.6) we assume that ( β n ) n N and ( α n ) n N satisfy (in addition to (2.4))

(2.8) n = 1 ( α n + β n ) < + ,

then all states of X ( t ) are stable, which means the Q -matrix of the process has finite values in all its diagonal entries. Moreover, the operator A determined by (2.7) is then bounded, D ( A ) = l 1 ( N ) , and it generates the transition semigroup of X ( t ) [4]. This also follows from [3, Proposition 2.7], and the semigroup is defined by (2.5). Let us also emphasize that the condition (2.8) (or (2.6)) does not contradict (2.4).

The application of [3, Proposition 2.7] in the stable case also shows that (2.8) overrides (2.4). In other words, the condition (2.8) is sufficient for A to be bounded. Recall that (2.4) ensures that the set of states of X ( t ) is countable, i.e., it is a Markov chain.

Therefore, when the condition (2.8) holds true, we can omit (2.4). The operator A and the semigroup P ( t ) are still well defined and P ( t ) = e t A . However, because we no longer assume (2.4), the connection between P ( t ) and Markov chains is lost. The state space of (2.3) could be uncountable, and X ( t ) would be a Markov process. It is not clear what P ( t ) then describes. Currently, I am not able to provide any interpretation. Similarly, the meaning of B n from definition 1, as well as A in Theorem 3.3, is unclear. Nevertheless, from the point of view of semigroups, these objects are well defined.

2.2 Definition of the B n ’s

This special form of the operators given by (2.1) suggests how to generalize them so that they still commute and generate semigroups of contractive operators. The goal of this article is to prove that this generalization works and the construction is based on a finite collection of sequences of numbers, so we begin with them.

Let an integer d 2 be fixed and suppose that there are given d ( d 1 ) sequences of positive numbers, denoted by ( β i , j n ) n N , i.e.,

(2.9) β i , j n > 0 , n 1 , i , j = 1 , 2 , , d and i j .

These numbers, for fixed n , can be thought of as off-diagonal entries in a d × d matrix, see Examples 1, 2. Based on (2.9), define ( β i , i n ) n N as follows:

(2.10) β i , i n j = 1 , j i d β i , j n , i = 1 , 2 , , d .

The numbers β i , i n are diagonal entries, in the matrix analogy. It should be added that the matrix analogy is only perfect for n = 1 . From (2.10), it follows that if n is fixed, then j = 1 d β i , j n = 0 , for every i = 1 , 2 , , d .

The sets S n 1 , S n 2 , determined by (2.2), partition the set Z 2 n 1 = { 0 , 1 , 2 , , 2 n 1 } into two parts of equal size. In a similar way, we introduce a partition of Z d n 1 = { 0 , 1 , 2 , , d n 1 } , denoted as S n 1 , S n 2 , , S n d , into d parts of equal size. Namely, define

(2.11) S n k = { ( k 1 ) d n 1 + 1 , ( k 1 ) d n 1 + 2 , , k d n 1 } , k = 1 , 2 , , d 1 , S n d = { 0 , ( d 1 ) d n 1 + 1 , ( d 1 ) d n 1 + 2 , , d n 1 } .

In other words, S n k = d n 1 for k = 1 , 2 , , d and

Z d n 1 = k = 1 d S n k and S n k 1 S n k 2 = , k 1 k 2 .

The sets S n 1 , S n 2 , …, S n d can also be defined recursively and in order to do this let us first introduce some notation. If S is a subset of Z , where Z stands for the set of integers, and a Z , then S + a denotes the set { s + a : s S } . So { 1 , 2 } + 3 = { 4 , 5 } . With a little abuse of notation, we also introduce

[ a , b ] { a , a + 1 , a + 2 , , b 1 , b } , a b , a , b Z .

This should not lead to misunderstandings, since all indices in a vector ( ξ i ) or in a sum j = j 1 j 2 ξ j are assumed to be integers. With this notation, we write

( ξ i ) i [ 1 , d n ] = ( ξ 1 , ξ 2 , , ξ d n ) , S n k = [ ( k 1 ) d n 1 + 1 , k d n 1 ] .

The recursive definition of (2.11) would be to assume S n 1 = { 1 , 2 , 3 , , d n 1 } and

S n k = ( S n 1 + ( k 1 ) d n 1 ) mod d n , k = 2 , 3 , , d ,

where the sum taken modulus d n ensures that 0 S n d . We can now give the following definition.

Definition 1

For n 1 define B n : l 1 ( N ) l 1 ( N ) as follows: for x = ( ξ i ) i N let B n x ( η i ) i N , where

(2.12) η i = j = 1 k 1 β k j , k n ξ i j d n 1 + β k , k n ξ i + j = 1 d k β k + j , k n ξ i + j d n 1 , if i mod d n S n k ,

where the set S n k is determined by (2.11).□

In this article, we use the convention that if in a sum j = j 1 j 2 ξ j we have j 2 < j 1 , then the sum equals zero. For example, taking i = 1 in (2.12), since S n 1 = { 1 , 2 , , d n 1 } , we obtain

η 1 = β 1,1 n ξ 1 + j = 1 d 1 β 1 + j , 1 n ξ 1 + j d n 1 .

Furthermore, it is worth noting that if i [ 1 , d n ] with i mod d n S n k , then both i j d n 1 [ 1 , d n ] , for j = 1 , 2 , , k 1 , as well as i + j d n 1 [ 1 , d n ] , for j = 1 , 2 , , d k . This follows from the fact that i lies between ( k 1 ) d n 1 + 1 and k d n 1 . Therefore, in the first case, we have

i j d n 1 ( k 1 ) d n 1 + 1 ( k 1 ) d n 1 = 1

and, in the second case,

i + j d n 1 k d n 1 + ( d k ) d n 1 = d n .

In a similar way, if i S l , where S l = [ 1 , d n ] + l d n , for some integer l 1 , and i mod d n S n k , then i j d n 1 S l , for j = 1 , 2 , , k 1 and i + j d n 1 S l , for j = 1 , 2 , , d k .

We draw an important conclusion from these observations. Namely, the one that the (infinite) matrix of B n , denoted by M ( B n ) , can be written in the following way:

(2.13) M ( B n ) = M ( B ˜ n ) 0 0 0 M ( B ˜ n ) 0 0 0 M ( B ˜ n )

where 0 denotes the d n × d n matrix with all entries zero and M ( B ˜ n ) is the d n × d n matrix of B n truncated to the first d n coordinates. In other words, it is the matrix of the map B ˜ n : R d n R d n defined as follows:

B ˜ n x ( η ˜ i ) i [ 1 , d n ] with η ˜ i = η i , i [ 1 , d n ] ,

where η i is given by (2.12). In this case, x = ( ξ i ) i [ 1 , d n ] . The fact that M ( B n ) is the matrix of B n simply means

B n x = x M ( B n ) , x l 1 ( N ) .

So it is clear that B n is a bounded linear operator on l 1 ( N ) , and using the triangle inequality, we obtain an estimate

(2.14) B n x l 1 ( N ) c n x l 1 ( N ) ,

where c n is given by

c n = d i , j = 1 ; i j d β i , j n .

Example 1

For d = 2 , the B n ’s are given by (2.1), and two independent sequences ( β 1 , 2 n ) n N , ( β 2,1 n ) n N , introduced in (2.9), are denoted by ( β n ) n N and ( α n ) n N , respectively. This notation is used in [4, 5] and in [6, p. 83]. Rewriting (2.1) in terms of ( β 1 , 2 n ) n N and ( β 2,1 n ) n N gives

B n x = ( η i ) i N = β 1 , 2 n ξ i + β 2,1 n ξ i + 2 n 1 , if i mod 2 n S n 1 , β 2,1 n ξ i + β 1 , 2 n ξ i 2 n 1 , if i mod 2 n S n 2 ,

where the sets S n 1 , S n 2 are given by (2.2). In this case, β 1,1 n = β 1 , 2 n , β 2,2 n = β 2,1 n , for n 1 . In particular, S 1 1 = { 1 } , S 1 2 = { 0 } , S 2 1 = { 1 , 2 } , S 2 2 = { 0 , 3 } and

M ( B ˜ 1 ) = β 1 , 2 1 β 1 , 2 1 β 2,1 1 β 2,1 1 M ( B ˜ 2 ) = β 1 , 2 2 0 β 1 , 2 2 0 0 β 1 , 2 2 0 β 1 , 2 2 β 2,1 2 0 β 2,1 2 0 0 β 2,1 2 0 β 2,1 2 .

For n = 3 , we have S 3 1 = { 1 , 2 , 3 , 4 } , S 3 2 = { 0 , 5 , 6 , 7 } , and

M ( B ˜ 3 ) = β 1 , 2 3 0 0 0 β 1 , 2 3 0 0 0 0 β 1 , 2 3 0 0 0 β 1 , 2 3 0 0 0 0 β 1 , 2 3 0 0 0 β 1 , 2 3 0 0 0 0 β 1 , 2 3 0 0 0 β 1 , 2 3 β 2,1 3 0 0 0 β 2,1 3 0 0 0 0 β 2,1 3 0 0 0 β 2,1 3 0 0 0 0 β 2,1 3 0 0 0 β 2,1 3 0 0 0 0 β 2,1 3 0 0 0 β 2,1 3 .

Example 2

Consider d = 3 . In this case, there are six independent sequences: ( β 1 , 2 n ) n N , ( β 1,3 n ) n N , ( β 2,1 n ) n N , ( β 2 , 3 n ) n N , ( β 3,1 n ) n N , ( β 3,2 n ) n N . Then

β 1,1 n = ( β 1 , 2 n + β 1,3 n ) , β 2,2 n = ( β 2,1 n + β 2 , 3 n ) , β 3,3 n = ( β 3,1 n + β 3,2 n ) , n 1 ,

and (2.12) takes the form

B n x = ( η i ) i N = β 1,1 n ξ i + β 2,1 n ξ i + 3 n 1 + β 3,1 n ξ i + 2 3 n 1 , if i mod 3 n S n 1 , β 1 , 2 n ξ i 3 n 1 + β 2,2 n ξ i + β 3,2 n ξ i + 3 n 1 , if i mod 3 n S n 2 , β 1,3 n ξ i 2 3 n 1 + β 2 , 3 n ξ i 3 n 1 + β 3,3 n ξ i , if i mod 3 n S n 3 ,

where the sets S n 1 , S n 2 , S n 3 , for n 1 , are given by

S n 1 = { 1 , 2 , 3 , , 3 n 1 } S n 2 = { 3 n 1 + 1 , 3 n 1 + 2 , , 2 3 n 1 } S n 3 = { 0 , 2 3 n 1 + 1 , 2 3 n 1 + 2 , , 3 n 1 } .

In particular, for n = 1 , we have S 1 1 = { 1 } , S 1 2 = { 2 } , S 1 3 = { 0 } and M ( B ˜ 1 ) is as follows:

M ( B ˜ 1 ) = β 1,1 1 β 1 , 2 1 β 1,3 1 β 2,1 1 β 2,2 1 β 2 , 3 1 β 3,1 1 β 3,2 1 β 3,3 1 = ( β 1 , 2 1 + β 1,3 1 ) β 1 , 2 1 β 1,3 1 β 2,1 1 ( β 2,1 1 + β 2 , 3 1 ) β 2 , 3 1 β 3,1 1 β 3,2 1 ( β 3,1 1 + β 3,2 1 ) .

For n = 2 , we have S 2 1 = { 1 , 2 , 3 } , S 2 2 = { 4 , 5 , 6 } , S 2 3 = { 0 , 7 , 8 } and M ( B ˜ 2 ) can be written as follows:

M ( B ˜ 2 ) = β 1,1 2 0 0 β 1 , 2 2 0 0 β 1,3 2 0 0 0 β 1,1 2 0 0 β 1 , 2 2 0 0 β 1,3 2 0 0 0 β 1,1 2 0 0 β 1 , 2 2 0 0 β 1,3 2 β 2,1 2 0 0 β 2,2 2 0 0 β 2 , 3 2 0 0 0 β 2,1 2 0 0 β 2,2 2 0 0 β 2 , 3 2 0 0 0 β 2,1 2 0 0 β 2,2 2 0 0 β 2 , 3 2 β 3,1 2 0 0 β 3,2 2 0 0 β 3,3 2 0 0 0 β 3,1 2 0 0 β 3,2 2 0 0 β 3,3 2 0 0 0 β 3,1 2 0 0 β 3,2 2 0 0 β 3,3 2 ,

where β 1,1 2 = ( β 12 2 + β 13 2 ) , β 2,2 2 = ( β 21 2 + β 23 2 ) , and β 3,3 2 = ( β 31 2 + β 32 2 ) .

3 Main results

3.1 Proofs of (1.1)

Lemma 3.1

The operator B n , given by (2.12), generates a semigroup of contractions, i.e.,

(3.1) P n ( t ) x l 1 ( N ) = e t B n x l 1 ( N ) x l 1 ( N )

for every t 0 and x l 1 ( N ) .

Proof

We prove (3.1) using the Lumer-Phillips theorem, see [9, Theorem 2.1]. This theorem states that a bounded linear operator B , defined in a Banach space ( X , ) , generates a semigroup of contractive operators if and only if it is dissipative, which means that it satisfies the condition ( λ B ) x λ x , for all λ > 0 [10, p. 75].

Dissipativity can be checked by somewhat simpler condition. Namely, if X is a real Banach space, then B is dissipative if and only if for every x X there exists j ( x ) J ( x ) such that

(3.2) B x , j ( x ) 0 ,

where

J ( x ) { x X : x , x = x 2 = x 2 } ,

see [10, Proposition 3.23]. It is worth to mention that J ( x ) is always nonempty by the Hahn-Banach theorem, see [7, p. 181]. Recall that X stands for the dual space of X consisting of the bounded linear functionals on X and x , x denotes x ( x ) . It is also a Banach space, see [7, p. 180].

We show (3.2) for B = B n . In our case, X = l 1 ( N ) . It is well known, see [7, p. 207], that X can be identified with the Banach space of all bounded sequences, denoted by l ( N ) . If x = ( ξ i ) i N is an element of l ( N ) , then x l ( N ) = sup i N ξ i .

Let x = ( ξ i ) i N be any non-zero element of l 1 ( N ) and define x l ( N ) as follows:

(3.3) x x l 1 ( N ) ( sgn ( ξ i ) ) i N ,

where sgn ( 0 ) = 0 and sgn ( a ) = a a for a non-zero a R . Then x J ( x ) , due to

x , x = x l 1 ( N ) i = 1 ξ i sgn ( ξ i ) = x l 1 ( N ) 2 .

The main tool used in proving B n x , x 0 is the following elementary inequality

(3.4) a sgn ( b ) a sgn ( a ) = a , a , b R .

By (2.12) and (3.3), we need to prove

(3.5) B n x , x = x l 1 ( N ) j J i = j + 1 j + d n η i sgn ( ξ i ) 0 ,

where J = { m d n : m 0 , m Z } . The outer sum in (3.5) is simply from j = 0 to + , where j increases by a multiple of d n . From (2.13), it is clear that to prove (3.5), it is enough to show

(3.6) Ω i = 1 d n η i sgn ( ξ i ) 0 .

By substituting for η i in (3.6), we can write Ω = Σ 1 + Σ 2 , where

Σ 1 = i = 1 d n σ i ξ i ,

with

σ i = j = 1 k 1 β k , k j n sgn ( ξ i j d n 1 ) + j = 1 d k β k , k + j n sgn ( ξ i + j d n 1 ) , if i mod d n S n k ,

and

Σ 2 = i = 1 d n τ i ξ i sgn ( ξ i ) = i = 1 d n τ i ξ i ,

with

τ i = β k , k n , if i mod d n S n k .

Applying (3.4) in Σ 1 gives

Ω = Σ 1 + Σ 2 i = 1 d n σ ˜ i ξ i + i = 1 d n τ i ξ i ,

where

σ ˜ i = j = 1 k 1 β k , k j n + j = 1 d k β k , k + j n , if i mod d n S n k .

The final step is to notice that σ ˜ i = τ i , see (2.10). Therefore, we have

Ω i = 1 d n ( σ ˜ i + τ i ) ξ i = 0 ,

which completes the proof of (3.6) and (3.5).□

Lemma 3.2

Let P n ( t ) = e t B n , for n 1 , where B n is given by (2.12). Then

(3.7) P m ( t ) P n ( s ) x = P n ( s ) P m ( t ) x ,

for any m , n N , t , s 0 , and x l 1 ( N ) .

Proof

Because the B n s are bounded, the condition (3.7) is equivalent to

(3.8) B m B n x = B n B m x ,

see e.g. [10, p. 19]. So we prove (3.8) and it suffices to consider the case for m < n . A consequence of this assumption is that m n 1 . Let x = ( ξ i ) i N l 1 ( N ) be fixed and denote

( η i ) i N = B m ( ξ i ) i N , ( ζ i ) i N = B n ( η i ) i N , ( η i ) i N = B n ( ξ i ) i N , ( ζ i ) i N = B m ( η i ) i N .

It can be seen from (2.13) that in order to prove (3.8), it is enough to show

(3.9) ζ i = ζ i , for i [ 1 , d n ] .

Fix i [ 1 , d n ] . Then, for some k 1 , k 2 [ 1 , d ] , we have

(3.10) i mod d m S m k 1 , i mod d n S n k 2 .

As a result, for some l { 0 , 1 , , d n 1 m } we also have

i S l = S m k 1 + l d m + ( k 2 1 ) d n 1 .

We begin to calculate ζ i , i.e., the i th coordinate of B n B m x . From (2.12) and (3.10) we obtain

(3.11) ζ i = j 2 = 1 k 2 1 β k 2 j 2 , k 2 n η i j 2 d n 1 + β k 2 , k 2 n η i + j 2 = 1 d k 2 β k 2 + j 2 , k 2 n η i + j 2 d n 1 ,

where

η i = j 1 = 1 k 1 1 β k 1 j 1 , k 1 m ξ i j 1 d m 1 + β k 1 , k 1 m ξ i + j 1 = 1 d k 1 β k 1 + j 1 , k 1 m ξ i + j 1 d m 1 .

What remains to be calculated in (3.11) is η i + j 2 d n 1 and η i j 2 d n 1 . These should be expressed in terms of k 1 and k 2 . We achieve this from m n 1 . Namely,

i mod d m S m k 1 ( i + j d n 1 ) mod d m S m k 1 ,

where j is an integer. Therefore,

η i + j 2 d n 1 = j 1 = 1 k 1 1 β k 1 j 1 , k 1 m ξ i + j 2 d n 1 j 1 d m 1 + β k 1 , k 1 m ξ i + j 2 d n 1 + j 1 = 1 d k 1 β k 1 + j 1 , k 1 m ξ i + j 2 d n 1 + j 1 d m 1 ,

and, in a similar way,

η i j 2 d n 1 = j 1 = 1 k 1 1 β k 1 j 1 , k 1 m ξ i j 2 d n 1 j 1 d m 1 + β k 1 , k 1 m ξ i j 2 d n 1 + j 1 = 1 d k 1 β k 1 + j 1 , k 1 m ξ i j 2 d n 1 + j 1 d m 1 .

So (3.11) can be written as sum of nine terms, i.e.,

(3.12) ζ i = j = 1 9 T j ,

where T 1 , T 2 , , T 9 are given by

T 1 = j 2 = 1 k 2 1 j 1 = 1 k 1 1 β k 2 j 2 , k 2 n β k 1 j 1 , k 1 m ξ i j 2 d n 1 j 1 d m 1 , T 2 = β k 1 , k 1 m j 2 = 1 k 2 1 β k 2 j 2 , k 2 n ξ i j 2 d n 1 , T 3 = j 2 = 1 k 2 1 j 1 = 1 d k 1 β k 2 j 2 , k 2 n β k 1 + j 1 , k 1 m ξ i j 2 d n 1 + j 1 d m 1 , T 4 = β k 2 , k 2 n j 1 = 1 k 1 1 β k 1 j 1 , k 1 m ξ i j 1 d m 1 , T 5 = β k 2 , k 2 n β k 1 , k 1 m ξ i , T 6 = β k 2 , k 2 n j 1 = 1 d k 1 β k 1 + j 1 , k 1 m ξ i + j 1 d m 1 , T 7 = j 2 = 1 d k 2 j 1 = 1 k 1 1 β k 2 + j 2 , k 2 n β k 1 j 1 , k 1 m ξ i + j 2 d n 1 j 1 d m 1 , T 8 = β k 1 , k 1 m j 2 = 1 d k 2 β k 2 + j 2 , k 2 n ξ i + j 2 d n 1 , T 9 = j 2 = 1 d k 2 j 1 = 1 d k 1 β k 2 + j 2 , k 2 n β k 1 + j 1 , k 1 m ξ i + j 2 d n 1 + j 1 d m 1 .

Now we calculate ζ i , i.e., the i th coordinate of B m B n x . From (2.12) and (3.10), we have

(3.13) ζ i = j 1 = 1 k 1 1 β k 1 j 1 , k 1 m η i j 1 d m 1 + β k 1 , k 1 m η i + j 1 = 1 d k 1 β k 1 + j 1 , k 1 m η i + j 1 d m 1 ,

where

η i = j 2 = 1 k 2 1 β k 2 j 2 , k 2 n ξ i j 2 d n 1 + β k 2 , k 2 n ξ i + j 2 = 1 d k 2 β k 2 + j 2 , k 2 n ξ i + j 2 d n 1 .

Now notice that

( i j d m 1 ) mod d n S n k 2 , for j [ 1 , k 1 1 ] ,

and, similarly,

( i + j d m 1 ) mod d n S n k 2 , for j [ 1 , d k 1 ] .

This allows to express η i j 1 d m 1 and η i + j 1 d m 1 in terms of k 1 and k 2 , see (3.10), and write ζ i in a similar way as ζ i , i.e., as the sum of nine terms, cf. (3.12). The final form of (3.13) is such that

ζ i = j = 1 9 T j = j = 1 9 T j = ζ i .

In other words, T 1 , , T 9 and T 1 , , T 9 may only differ in order. This completes the proof.□

3.2 Convergence of n = 1 N P n ( t )

Suppose that i N . Then, there exists n 0 such that for every n n 0

i mod d n S n 1 = { 1 , 2 , , d n 1 } .

For the proof, from i d n 1 we obtain n log d ( i ) + 1 and n 0 can be written explicitly using the ceiling function, i.e.,

(3.14) n 0 = log d ( i ) + 1 .

To recall, x = min { n Z : n x } .

In formula (2.12), which defines B n , n was fixed, and it was not necessary to denote η i as η i ( n ) . However, in what follows, we sum these terms over n , see (3.15) and (3.19), so the latter notation is used.

As mentioned earlier, the product n = 1 N P n ( t ) , for any N 2 , is a semigroup of contractions and its generator A N equals n = 1 N B n [2, p. 24]. In our case, it means that

(3.15) A N x = ( ζ i ) i N , ζ i = n = 1 N η i ( n ) ,

where x = ( ξ i ) i N l 1 ( N ) and η i ( n ) is given by (2.12). It turns out that A N still has a manageable form even if N = + , cf. [4, Theorem 3.1].

Theorem 3.3

Suppose that the sequences ( β i , j n ) n N satisfy

(3.16) i , j = 1 ; i j d n = 1 β i , j n < ,

and let B n be given by (2.12). Then the strong limit of n = 1 N e t B n , denoted as ( T ( t ) ) t 0 , i.e.,

(3.17) T ( t ) x lim N n = 1 N e t B n x , x l 1 ( N ) ,

is a semigroup of contractions, and its generator A is bounded and

A x = lim N n = 1 N B n x , x l 1 ( N ) .

Furthermore, denote ( ζ i ) i N = A x . Then

(3.18) ζ 1 = n = 1 ( β 1,1 n ξ 1 + j = 1 d 1 β 1 + j , 1 n ξ 1 + j d n 1 ) ,

and for i 2 , we have

(3.19) ζ i = n = 1 n 0 1 η i ( n ) + n = n 0 ( β 1,1 n ξ i + j = 1 d 1 β 1 + j , 1 n ξ i + j d n 1 ) ,

where η i ( n ) and n 0 are given by (2.12) and (3.14), respectively.

Proof

We use [3, Proposition 2.7]. This proposition says that if ( e t B n ) n N is a family of commuting semigroups of contractions, in a Banach space ( X , ) , and the set

D 1 = x n = 1 D ( B n ) : n = 1 + B n x <

is dense in X , then the semigroup given by (3.17) is well defined. Moreover, its generator is the closure of A , where A = lim N n = 1 N B n , with D ( A ) being D 1 . In this proposition, the generators may not be bounded, so D ( B n ) denotes the domain of B n .

In our case, X = l 1 ( N ) and D ( B n ) = X , since B n is bounded, see (2.14). Furthermore, D 1 = l 1 ( N ) , because by (2.14) and (3.16), we have

n = 1 + B n x l 1 ( N ) d x l 1 ( N ) i , j = 1 ; i j d n = 1 β i , j n < ,

for every x l 1 ( N ) . In particular, A is bounded and

A d i , j = 1 ; i j d n = 1 β i , j n .

Since the norm convergence in l 1 ( N ) implies convergence in coordinates, components of A x are limits of components of A N x , where A N is given by (3.15). Thus, (3.18) and (3.19) follow, and this completes the proof.□

3.3 Convergence of n = 1 N P n ( t )

As mentioned earlier, the dual space of l 1 ( N ) can be identified with l ( N ) [7, p. 207]. Since B n is bounded, it induces a linear map B n : l ( N ) l ( N ) , called the adjoint of B n [11, p. 15]. Moreover, B n = B n , so if B n is a contraction, then B n is also a contraction, see also (A1).

Let ( η i ) i N = B n x , where x = ( ξ i ) i N l 1 ( N ) . In our case, we have

(3.20) η i = j = 1 k 1 β k , k j n ξ i j d n 1 + β k , k n ξ i + j = 1 d k β k , k + j n ξ i + j d n 1 , if i mod d n S n k .

This formula simply says that M ( B n ) = M T ( B n ) , i.e., the matrix of B n is equal to the transpose of M ( B n ) , see (2.13). Therefore, it can be written as follows:

M ( B n ) = M ( B ˜ n ) 0 0 0 M ( B ˜ n ) 0 0 0 M ( B ˜ n ) = M T ( B ˜ n ) 0 0 0 M T ( B ˜ n ) 0 0 0 M T ( B ˜ n )

where M ( B ˜ n ) denotes the matrix of B n truncated to the first d n coordinates. Similar estimate to (2.14) can be obtained for B n . Namely, we have

(3.21) B n x l ( N ) ( d 1 ) i , j = 1 ; i j d β i , j n x l ( N ) .

Moreover, notice that the equality ( B n B m ) = B m B n , together with (3.8), implies that the B n ’s also commute. In Lemma A.1, we prove that ( e B ) = e B . This allows for a correct definition

(3.22) P n ( t ) e t B n = ( e t B n ) = ( P n ( t ) ) , n 1 .

In consequence of (3.22) and (1.1), we obtain

(3.23) P n ( t ) 1 , P m ( t ) P n ( s ) = P n ( s ) P m ( t ) ,

for any m , n N and t , s 0 . So we are in a similar situation as in Section 3.2. In other words, the product n = 1 N P n ( t ) , for any N 2 , is a semigroup of contractions and its generator A N equals n = 1 N B n , which means

(3.24) A N x = ( ζ i ) i N , ζ i = n = 1 N η i ( n ) ,

where x = ( ξ i ) i N l ( N ) and η i ( n ) is given by (3.20). We have the following theorem, cf. [4, Theorem 3.2].

Theorem 3.4

Suppose that the sequences ( β i , j n ) n N satisfy (3.16) and let B n be given by (3.20). Then the strong limit of n = 1 N e t B n , denoted as ( T ( t ) ) t 0 , i.e.

(3.25) T ( t ) x lim N n = 1 N e t B n x , x l ( N )

is a semigroup of contractions, and its generator A is bounded and

A x = lim N n = 1 N B n x , x l ( N ) .

Furthermore, denote ( ζ i ) i N = A x . Then

(3.26) ζ 1 = n = 1 ( β 1,1 n ξ 1 + j = 1 d 1 β 1,1 + j n ξ 1 + j d n 1 ) ,

and for i 2 , we have

(3.27) ζ i = n = 1 n 0 1 η i ( n ) + n = n 0 ( β 1,1 n ξ i + j = 1 d 1 β 1,1 + j n ξ i + j d n 1 ) ,

where η i ( n ) and n 0 are given by (3.20) and (3.14), respectively.

Proof

The proof is analogous to that of Theorem 3.3, i.e., we use [3, Proposition 2.7]. By (3.21), for every x l ( N ) , we have

n = 1 + B n x l ( N ) ( d 1 ) x l ( N ) i , j = 1 ; i j d n = 1 β i , j n < .

Thus, D 1 = l ( N ) , A is bounded and

A ( d 1 ) i , j = 1 ; i j d n = 1 β i , j n .

As in l 1 ( N ) , the norm convergence in l ( N ) implies the coordinate-wise convergence, so components of A x are limits of components of A N x , where A N is given by (3.24). Thus, (3.26) and (3.27) follow, and this completes the proof.□

Acknowledgments

I would like to thank the reviewers for their comments and remarks.

  1. Funding information: The author states no funding involved.

  2. Author contributions: The entire article was written by the author.

  3. Conflict of interest: The author states no conflict of interest.

  4. Data availability statement: Data sharing is not applicable to this article as no new data were created or analyzed in this study.

Appendix

It is well known [1, p. 163] that if B and C are bounded linear operators, given in a Banach space X , then

(A1) B = B , ( B + C ) = B + C , ( B C ) = C B ,

where B and C are adjoint operators to B and C , respectively. Recall also that ( X ) denotes the space of all bounded linear maps on X . To justify (3.22), we prove the following lemma.

Lemma A.1

Let X be a Banach space and suppose that B ( X ) . Then

(A2) ( e B ) = e B .

Proof

For N 1 denote

S N = n = 0 N B n n ! , S N = n = 0 N ( B ) n n ! .

These operators converge uniformly to e B and e B , respectively [1, p. 251]. This means

lim N e B S N = 0 , lim N e B S N = 0 .

By (A1), we have ( S N ) = S N , for N 1 , which implies

( e B ) S N = ( e B S N ) = e B S N .

Therefore, for any x X , we have

lim N ( e B ) x S N x lim N ( e B ) S N x = 0 .

In consequence,

( e B ) x = e B x , x X ,

and this concludes the proof of (A2).□

References

[1] A. Bobrowski, Functional Analysis for Probability and Stochastic Processes, Cambridge University Press, Cambridge, 2005. 10.1017/CBO9780511614583Search in Google Scholar

[2] R. Nagel (Ed.), One-parameter semigroups of positive operators, Lecture Notes in Mathematics, vol. 1184, Springer-Verlag, Berlin, 1986. Search in Google Scholar

[3] W. Arendt, A. Driouich, and O. El-Mennaoui, On the infinite product of C0 semigroups, J. Funct. Anal. 160 (1998), 524–542. 10.1006/jfan.1998.3341Search in Google Scholar

[4] E. Nieznaj, The infinite product of contraction semigroups on l1(N) and l∞(N), Arch. Math. 119 (2022), 593–600. 10.1007/s00013-022-01794-2Search in Google Scholar

[5] E. Nieznaj, On a certain operator related to Blackwell’s Markov chain, J. Theoret. Probab. 35 (2022), no. 3, 1501–1510. 10.1007/s10959-021-01110-8Search in Google Scholar

[6] A. Bobrowski, Generators of Markov Processes, Cambridge University Press, Cambridge, 2021. Search in Google Scholar

[7] S. Axler, Measure, Integration and Real Analysis, Springer, Cham, Switzerland, 2020. 10.1007/978-3-030-33143-6Search in Google Scholar

[8] D. Blackwell, Another countable Markov process with only instantaneous states, Ann. Math. Statist. 29 (1958), 313–316. 10.1214/aoms/1177706735Search in Google Scholar

[9] G. Lumer, R. S. Phillips, Dissipative operators in a Banach space, Pacific J. Math. 11 (1961), 679–698. 10.2140/pjm.1961.11.679Search in Google Scholar

[10] K. J. Engel, and R. Nagel, A Short Course on Operator Semigroups, Springer, New York, 2006. Search in Google Scholar

[11] N. L. Carothers, A Short Course on Banach Space Theory, Cambridge University Press, Cambridge, 2005. 10.1017/CBO9780511614057Search in Google Scholar

Received: 2024-05-25
Revised: 2025-04-22
Accepted: 2025-05-13
Published Online: 2025-07-10

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Special Issue on Contemporary Developments in Graphs Defined on Algebraic Structures
  2. Forbidden subgraphs of TI-power graphs of finite groups
  3. Finite group with some c#-normal and S-quasinormally embedded subgroups
  4. Classifying cubic symmetric graphs of order 88p and 88p 2
  5. Simplicial complexes defined on groups
  6. Two-sided zero-divisor graphs of orientation-preserving and order-decreasing transformation semigroups
  7. Further results on permanents of Laplacian matrices of trees
  8. Special Issue on Convex Analysis and Applications - Part II
  9. A generalized fixed-point theorem for set-valued mappings in b-metric spaces
  10. Research Articles
  11. Dynamics of particulate emissions in the presence of autonomous vehicles
  12. The regularity of solutions to the Lp Gauss image problem
  13. Exploring homotopy with hyperspherical tracking to find complex roots with application to electrical circuits
  14. The ill-posedness of the (non-)periodic traveling wave solution for the deformed continuous Heisenberg spin equation
  15. Some results on value distribution concerning Hayman's alternative
  16. 𝕮-inverse of graphs and mixed graphs
  17. A note on the global existence and boundedness of an N-dimensional parabolic-elliptic predator-prey system with indirect pursuit-evasion interaction
  18. On a question of permutation groups acting on the power set
  19. Chebyshev polynomials of the first kind and the univariate Lommel function: Integral representations
  20. Blow-up of solutions for Euler-Bernoulli equation with nonlinear time delay
  21. Spectrum boundary domination of semiregularities in Banach algebras
  22. Statistical inference and data analysis of the record-based transmuted Burr X model
  23. A modified predictor–corrector scheme with graded mesh for numerical solutions of nonlinear Ψ-caputo fractional-order systems
  24. Dynamical properties of two-diffusion SIR epidemic model with Markovian switching
  25. Classes of modules closed under projective covers
  26. On the dimension of the algebraic sum of subspaces
  27. Periodic or homoclinic orbit bifurcated from a heteroclinic loop for high-dimensional systems
  28. On tangent bundles of Walker four-manifolds
  29. Regularity of weak solutions to the 3D stationary tropical climate model
  30. A new result for entire functions and their shifts with two shared values
  31. Freely quasiconformal and locally weakly quasisymmetric mappings in metric spaces
  32. On the spectral radius and energy of the degree distance matrix of a connected graph
  33. Solving the quartic by conics
  34. A topology related to implication and upsets on a bounded BCK-algebra
  35. On a subclass of multivalent functions defined by generalized multiplier transformation
  36. Local minimizers for the NLS equation with localized nonlinearity on noncompact metric graphs
  37. Approximate multi-Cauchy mappings on certain groupoids
  38. Multiple solutions for a class of fourth-order elliptic equations with critical growth
  39. A note on weighted measure-theoretic pressure
  40. Majorization-type inequalities for (m, M, ψ)-convex functions with applications
  41. Recurrence for probabilistic extension of Dowling polynomials
  42. Unraveling chaos: A topological analysis of simplicial homology groups and their foldings
  43. Global existence and blow-up of solutions to pseudo-parabolic equation for Baouendi-Grushin operator
  44. A characterization of the translational hull of a weakly type B semigroup with E-properties
  45. Some new bounds on resolvent energy of a graph
  46. Carmichael numbers composed of Piatetski-Shapiro primes in Beatty sequences
  47. The number of rational points of some classes of algebraic varieties over finite fields
  48. Singular direction of meromorphic functions with finite logarithmic order
  49. Pullback attractors for a class of second-order delay evolution equations with dispersive and dissipative terms on unbounded domain
  50. Eigenfunctions on an infinite Schrödinger network
  51. Boundedness of fractional sublinear operators on weighted grand Herz-Morrey spaces with variable exponents
  52. On SI2-convergence in T0-spaces
  53. Bubbles clustered inside for almost-critical problems
  54. Classification and irreducibility of a class of integer polynomials
  55. Existence and multiplicity of positive solutions for multiparameter periodic systems
  56. Averaging method in optimal control problems for integro-differential equations
  57. On superstability of derivations in Banach algebras
  58. Investigating the modified UO-iteration process in Banach spaces by a digraph
  59. The evaluation of a definite integral by the method of brackets illustrating its flexibility
  60. Existence of positive periodic solutions for evolution equations with delay in ordered Banach spaces
  61. Tilings, sub-tilings, and spectral sets on p-adic space
  62. The higher mapping cone axiom
  63. Continuity and essential norm of operators defined by infinite tridiagonal matrices in weighted Orlicz and l spaces
  64. A family of commuting contraction semigroups on l 1 ( N ) and l ( N )
  65. Pullback attractor of the 2D non-autonomous magneto-micropolar fluid equations
  66. Maximal function and generalized fractional integral operators on the weighted Orlicz-Lorentz-Morrey spaces
  67. On a nonlinear boundary value problems with impulse action
  68. Normalized ground-states for the Sobolev critical Kirchhoff equation with at least mass critical growth
  69. Decompositions of the extended Selberg class functions
  70. Subharmonic functions and associated measures in ℝn
  71. Some new Fejér type inequalities for (h, g; α - m)-convex functions
  72. The robust isolated calmness of spectral norm regularized convex matrix optimization problems
  73. Multiple positive solutions to a p-Kirchhoff equation with logarithmic terms and concave terms
  74. Joint approximation of analytic functions by the shifts of Hurwitz zeta-functions in short intervals
  75. Green's graphs of a semigroup
  76. Some new Hermite-Hadamard type inequalities for product of strongly h-convex functions on ellipsoids and balls
  77. Infinitely many solutions for a class of Kirchhoff-type equations
  78. On an uncertainty principle for small index subgroups of finite fields
Downloaded on 5.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/math-2025-0168/html?lang=en
Scroll to top button