-isometric composition operators on discrete spaces
Abstract.
In this paper we study composition operators on discrete spaces . We establish the classification of underlying graphs of such operators. We obtain the characterization of -isometric composition operators such that the underlying graph has one cycle. The paper contains also the solution of the Cauchy dual subnormality problem in the class of such composition operators.
Key words and phrases:
weighted shifts on graphs, m-isometries, composition operators, completely hyperexpansive operators, Cauchy dual subnormality problem1991 Mathematics Subject Classification:
Primary 47B33; Secondary 47B20, 47B371. Introduction
The composition operators form the important class of operators. The study of such a class of operators is motivated by the Banach-Stone theorem (see [9]), which states that a linear surjective isometry between spaces of continuous functions on compact spaces is in fact a weighted composition operators. The composition operators appear also naturally in ergodic theory (so-called Koopman operator, which is a composition operator induced by measure-preserving transformation). The class of composition operators has been studied extensively for a long time (see e.g. [22], [27]). In [12] Budzyński et.al. presented an example of unbounded non-hyponormal composition operators having graph with one cycle, which generates the Stieltjes moment sequence. In [24] Pietrzycki obtained an analytic model for a weighted composition operator on discrete space under the assumption that the underlying graph of this operator has finite branching index.
-isometric operators were introduced by Agler in [2] and investigated in details in [3] (as well as its subsequent parts). From that time on there have been published many papers devoted to study properties of -isometries (see e.g. [11], [8], [29], [17], [19]). In [1] and [10] -isometric weighted shifts were studied and, eventually, the characterization of weights of -isometric shift by real monic polynomials of degree at most was obtained (see [1, Theorem 1]). In [18] Jabłoński and Kośmider investigated -isometricity of composition operator on discrete space such that the graph associated to this operator has one cycle and one branching point.
The Cauchy dual of a left-invertible operator was introduced by Shimorin in [26] on the occasion of his study of Wold-type decompositions. From [7, Proposition 6] it follows that the Cauchy dual of a completely hyperexpansive weighted shift is always a subnormal contraction. In [13] there was posed a question whether it is true for every completely hyperexpansive operator. In [5] Chavan et. al. were considering this problem in the more restricted class of 2-isometries and obtained a negative answer. They also provided several sufficient conditions for 2-isometry to have subnormal Cauchy dual. The Cauchy dual subnormality problem was studied also in [6] and [14].
The present paper is devoted to study -isometric composition operators on discrete spaces in a more general setting than in [18]. It is organized as follows. In Section 2 we introduce the notation and recall basic definitions and some standard results needed in further parts of the paper. The aim of Section 3 is to fully classify self-maps of a countable set. If we associate a certain graph to such a map, it turns out that if this graph is assumed to be connected, then there are only two posibilities for its geometry: either this graph is a rootless directed tree or it consists of rooted directed trees connected by a cycle (see Theorem 3.5). In Section 5 we investigate the composition operators (viewed as certain weighted shifts) connected to graphs of the second type, that is, the ones having a cycle. We generalize the results presented in [18], where the very particular case of graph with a cycle and the only one branching point was studied. The assumption that the graph has a cycle allows us to obtain certain recursive formulas (see Lemma 4.4), which are key tools in further considerations. The main result of Section 5 is Theorem 5.5, which gives us the characterization of -isometricity in terms of the rank of certain matrices. We also prove that the notions of complete hyperexpansivity and 2-isometricity coincide in the class of composition operators having graphs with one cycle. Section 6 is devoted to study the Cauchy dual subnormality problem for 2-isometric composition operators on graphs with one cycle. In Theorem 6.2 the surprising characterization of composition operators having subnormal Cauchy dual is presented. It turns out that if the Cauchy dual is subnormal, then for every vertex lying on the cycle the measure representing the sequence is two-atomic. Using this result, we provide an example of a composition operator on graph with one cycle, for which the Cauchy dual is not subnormal.
2. Preliminaries
Denote by and the set of non-negative integers and integers, respectively and by and the field of real and complex numbers, respectively. For we set
If , then stands for the set of all polynomials of degree at most with coefficients in . From the binomial formula we can derive the following simple fact.
Observation 2.1.
If for some and , then
The next lemma will play a crucial role in our considerations.
Lemma 2.2.
Let . Suppose and for every write . Assume that for every the series
is absolutely convergent. Then the series
(2.1) |
is absolutely convergent for every and . Moreover,
Proof.
If and , then
Hence, the series (2.1) is absolutely convergent. Set , . For every we have
Thus, for . Consequently, . ∎
For a topological space we denote by the -algebra of Borel subsets of ; if , then by we denote the Dirac delta measure. For , stands for characteristic function of ; we abbreviate for . If is a complex Hilbert space, then stands for the -algebra of all linear and bounded operators on .
By a discrete measure space we mean a pair , where is an at most countable set and is a positive measure on such that for every ; for simplicity we write for . If is a discrete measure space and is a transformation on , then is absolutely continuous with respect to , where is a positive measure on defined by
Denote by the Radon-Nikodym derivative of with respect to ; we can easily compute that
(2.2) |
Assuming additionally that , we define the composition operator as follows:
For more details about general theory of composition operators we refer to [22].
We say that is normal if ; is called subnormal if there exists a Hilbert space containing (an isometric embedding of) and a normal operator such that . Recall that is completely hyperexpansive if
We say that is -isometric (), if
For an at most countable set we define
this is a Hilbert space with the norm given by:
The vectors () defined as follows:
form an orthonormal basis of .
In the subsequent part of this section we introduce some basic notions of graph theory (for more details see [23] and [4]). By a directed graph we mean a pair , where is a non-empty at most countable set and ; elements of are called vertices and elements of are called edges. Every pair such that and is called a subgraph of . For a vertex we define its in-degree, out-degree and degree as follows:
If , then a path from to in is a sequence such that , and for every . A path is called a cycle if ; the cycle is said to be simple if the vertices () are distinct.
Graph is said to be connected if for every two distinct vertices there exists a sequence such that , and for , where
If is a directed graph, then its subgraph is called a connected component if is connected and for each connected subgraph of satisfying and we have . In other words, connected components of are maximal connected subgraphs of .
A connected graph is called a directed tree if has no cycles and for every we have . If is a directed tree, then there is at most one vertex such that ; if it exists, we call this vertex the root of (see [16, Proposition 2.1.1]). The reader can verify that for rooted directed trees the following induction principle holds.
Lemma 2.3.
Let be a directed tree with root . Let satisfy the following conditions:
-
(i)
,
-
(ii)
for every we have .
Then .
We say that a graph is strongly connected if for any two distinct vertices there is a path from to and a path from to (in other words, and lie on a common cycle). For a directed graph we define a relation in the following way: if and only if either or there exists a path from to in and a path from to in . It is a matter of routine to verify that is an equivalence relation on . Each equivalence class of induces a subgraph , where , which is called a strongly connected component of .
3. Classification of graphs of self-maps of countable sets
In [15, Lemma 4.3.1] there was proved that weighted shifts on rootless directed trees with positive weights are unitarily equivalent to composition operators on certain discrete spaces (see also [12, Lemma 3.1.4]). In [12, Theorem 3.2.1] the authors classified composition operators on discrete spaces, for which the associated graph is connected and has at most one vertex of out-degree greater than 1. In [28, Proposition 2.4] composition operators on graphs with the property that every vertex has out-degree 1 are characterized. In this section, using graph-theoretical approach, we give the full classification of graphs of self-maps of at most countable set. This classification was mentioned in [24, p.898]; however, the proof was not provided there.
Suppose is an at most countable set and is a transformation on X; for we abbreviate: . Consider the graph , where and . In the next few results we investigate the geometry of graph .
Observation 3.1.
Suppose is an at most countable set and . Then, for every we have . Moreover, if for , , there exists a path from to , then there is the unique path from to consisting of distinct vertices.
Proof.
The first part can be easily derived from the definition of . Let us prove the ”moreover” part. Obviously, we can always find a path from to consisting of distinct vertices, so we only have to prove uniqueness. Let and be paths connecting and such that for every , and , for every ; without loss of generality we can assume . Since , we have that . Proceeding by induction, we can show that for all . In particular, we have . This, by the choice of , implies that and for all . ∎
The proceeding results reveal the structure of strongly connected components of . Lemma 3.2 shows that every connected component of may contain at most one non-trivial strongly connected component.
Lemma 3.2.
Suppose is an at most countable set and is such that is connected. Then contains at most one strongly connected component with non-empty set of edges (that is, , where ).
Proof.
Suppose to the contrary that and are two different strongly connected components of with non-empty sets of edges. Obviously, . Let and . We will show that there exists a path from to . Since is connected, there exists a sequence () such that , and for every . If , then from the fact that is strongly connected, and we obtain that . This implies that there exists a path from to in . Hence, possibly after extending a sequence , without loss of generality we can assume that . If for every , then the sequence is the path from to . Suppose that it is not the case; in particular, . Set
Since , and , it follows that . If , then we have . Thus, is the path from to . If , then the sequence
have the property that for every , so the above procedure can be applied to this sequence. Proceeding this way, we eventually obtain the path from to . By symmetry, there exists also a path from to . Since and are strongly connected, we have and , which is a contradiction. ∎
The next result shows that non-trivial strongly connected components of are of very special type.
Lemma 3.3.
Suppose is an at most countable set and is such that is connected and has a unique strongly connected component with non-empty set of edges. Then is a simple cycle.
Proof.
If , then the conclusion obviously holds. Assume . Let , . By the strong connectedness of , and lie on the common cycle; we will prove that this cycle is actually the whole graph . By Observation 3.1 there exist paths from to and to consisting of distinct vertices; call by the vertices of such a path from to and by the vertices of such a path from to . Set . It is enough to show that . Suppose to the contrary that it is not the case and let . Let be the path connecting and and consisting of distinct vertices. If , then the reasoning as in the proof of Observation 3.1 gives us that for , which implies that . If , then the same reasoning gives us that . Since it necessarily holds that , we have . In both cases we get a contradiction. ∎
The following lemma says that after removing non-trivial strongly connected component from we actually obtain a family of directed trees.
Lemma 3.4.
Suppose is an at most countable set and is such that is connected and has a unique strongly connected component with non-empty set of edges. Set , where . If is a connected component of , then is a rooted directed tree.
Proof.
From Lemma 3.2 it follows that has no cycles. This, together with Observation 3.1, implies that is a directed tree. It remains to prove that has a root. Let and . Proceeding as in the proof of Lemma 3.2 we obtain that there exists a path from to in . Let be the path from to consisting of distinct vertices. Set . From the fact that and that is the only edge coming to it follows that has in-degree 0 in . This implies that is the root of . ∎
Now we are in the position to state the main results of this section, which classifies graphs of self-maps .
Theorem 3.5.
Suppose is an at most countable set and is such that is connected. Then exactly one of the following holds:
-
(i)
is a rootless directed tree,
-
(ii)
satisfies the conditions below:
-
(a)
there exist and distinct vertices such that for and
-
(b)
there exists a partition of such that for every the graph , where , is a rooted directed tree with the root having the property that there exists exactly one satisfying .
-
(a)
Proof.
First, suppose has no non-trivial strongly connected components. This implies that has no cycles. By Observation 3.1, is a rootless directed tree, so in this case (i) holds. Assuming now that has a non-trivial strongly connected component, by calling Lemmata 3.3 and 3.4, we derive that satisfies the conditions in (ii). ∎
In Figure 1 we present the example of graph satisfying Theorem 3.5.(ii); for the convenience, we denote for , where is the cardinality of the set on the right hand side.

4. Composition operators as weighted shifts on directed graphs
In this section we present how to transform a composition operator on discrete space into a weighted shift on a directed graph; we prove several routine results about this weighted shift. Suppose that is a discrete measure space and is a transformation on such that . Define the family of weights by
(4.1) |
From (2.2) it can be easily derived that
Since we assume that , the above allows us to define the operator as follows:
It is a matter of routine to prove the following lemma.
Lemma 4.1.
Let be a discrete measure space. Suppose is such that . Then the operator given by
makes and unitarily equivalent.
Proof.
First, we show that is unitary. For we have
If , then ; if , then
Hence,
which implies that . Since is the orthonormal basis of , it follows that for , that is, is isometry. In turn, , so . Therefore, is unitary. Next, if , then
and
Hence, , what completes the proof. ∎
The lemma below states that connected components of induce reducing subspaces of , so we can restrict our investigation to the case when the graph is connected.
Lemma 4.2.
Suppose is a discrete measure space and is a transformation on such that . Let be a connected component of . Then the space is a reducing subspace for .
Proof.
Let . Since is a connected component of we have that for every satisfying , and . Hence, , which implies that . Next, we will prove that is also invariant for . It can be easily seen that for every and every . From this it follows that for every . This implies that , what completes the proof. ∎
It turns out that that the operator preserves the orthogonality of the standard orthonormal basis of .
Observation 4.3.
Suppose is a discrete measure space and is a transformation on such that and that is connected. Then for every vertices , .
Proof.
Let be distinct vertices. By Observation 3.1, we have
Since for every two distinct vertices , we deduce from the above that
Hence, . ∎
Set111We stick to the convention that .
For the further use we emphasize several formulas.
Lemma 4.4.
Suppose is a discrete space and is such that and that is connected and satisfies Theorem 3.5.(ii).
-
(i)
For every , .
-
(ii)
For every and every we have
(4.2) -
(iii)
For every and every the following recursive formula is satisfied:
(4.3)
5. -isometric composition operators: graphs with one cycle
In [18] the authors studied composition operators with the property that the associated graphs have one cycle and one branching point. In this paper we also limit our considerations to operators such that related graphs have one cycle, but with no further assumptions on degrees of vertices. From now on we fix a discrete space and a self-map such that , so that (and consequently ) is bounded. Since, by Observation 4.1, the composition operator is unitarily equivalent to , we investigate -isometricity of . Whenever we assume that the associated graph defined in Section 3 satisfies Theorem 3.5.(ii), we stick to the notation introduced there.
Let us start from the general result concerning -isometric composition operators on discrete spaces.
Lemma 5.1 (cf. [18, Proposition 2.5]).
Suppose is such that is connected and such that . For the following conditions are equivalent:
-
(i)
is -isometric,
-
(ii)
for every there exists a polynomial satisfying
-
(iii)
for every there exists a polynomial satisfying
(5.1)
Before we state the proof we need the following observation.
Observation 5.2.
Suppose is such that is connected and such that . Let . Let and suppose that is polynomial satisfying (5.1). Then there exists a constant depending only on and such that for .
Proof.
First, note that the coefficient form the unique solution of the following system of linear equations222We stick to the convention that .:
Applying Cramer’s rule (see [21, Theorem 3.36.(7)]) to the above system of linear equations, we obtain that there exist constants () depending on (not on the operator ) satisfying
Then, for every
Hence, the constant
satisfies . Obviously, depends only on and . ∎
Proof of Lemma 5.1.
The equivalence of (i) and (ii) follows from remarks in [3, p. 389]; the implication (ii)(iii) obviously holds. It remains to prove that (iii) implies (ii). Enumerate , where . Let . Then, by Observation 4.3,
(5.2) |
For let be the polynomial satisfying (5.1). Let be the constant given by Observation 5.2. Then, for every ,
Hence, the series is absolutely convergent for every . By Lemma 2.2,
In the next results we characterize -isometric composition operators such that the graph has one cycle. Let us begin with the following lemma.
Lemma 5.3.
Proof.
Let . Enumerate , where is the cardinality of the set on the left hand side. Write , , . We will check that the series
(5.4) |
is absolutely convergent for every . Since, by Observation 4.3, for every ,
it follows that the series is convergent. Using Observation 5.2, we obtain that the series is absolutely convergent. By Lemma 2.2, . Since for every , the leading coefficient of is non-negative for every . Hence, (5.3) holds. This completes the proof. ∎
It turns out that if is -isometric, then on ’tree parts’ of it is actually -isometric. The similar result is presented in [18, Theorem 2.10] for composition operators such that the graph has one cycle and one branching point; however, there is no clear way how to adjust the proof in a way to obtain the same conclusion in a more general situation. In our proof we use a different approach to prove this result in full generality.
Lemma 5.4.
Proof.
First, we prove that for every , the polynomial satisfying (5.1) with is of degree at most . By Lemma 5.1, there exist and () satisfying (5.1). Combining (5.1) and (4.3) we obtain that for every ,
(5.5) |
From Observation 2.1 we deduce that the polynomial is of degree less than or equal to . By Lemma 5.3, the formula
defines a polynomial of degree at most . We infer from (5.5) that and, consequently, . From Lemma 5.3 it follows that . Next, if and is such that , then proceeding as in the proof of Lemma 5.3, we obtain that
is the polynomial of degree at most . For every ,
Hence, . Since for every the leading coefficient of is non-negative, it follows that . Therefore, . The application of Lemma 2.3 completes the proof. ∎
Let us introduce the following notation: for and we set
where , , , and if ,
Now we are in the position to state the main result of this section.
Theorem 5.5.
Suppose is such that is connected and satisfies Theorem 3.5.(ii). Assume . For the following conditions are equivalent:
-
(i)
is -isometric,
- (ii)
Moreover, if (ii) holds, then , where is the solution of
(5.7) |
Proof.
(i) (ii). By Lemma 5.1, for every there exists satisfying (5.1). By Lemma 5.4, for every and every , . Next, writing and using the binomial formula, we have
Since, by Lemma 5.3, the formula
defines a polynomial of degree at most satisfying, by (4.3),
writing we obtain
(5.8) |
Hence, satisfies (5.7). Moreover, since the determinant of is non-zero, we infer from Cramer’s rule that is the only solution of this system. By the Kronecker-Capelli theorem (see [21, Theorem 3.36.(1)-(3)]) this implies that . Assume now . From the equality () we get that is also the solution of the following system of linear equations:
the matrix representation of this system takes the form:
(5.9) |
Again, by the Kronecker-Capelli theorem, , which gives (ii).
(ii) (i). By Lemma 5.1, it is enough to find the polynomials () satisfying (5.1). First, observe that for every the polynomial is uniquely determined by and (). Indeed, assuming we have satisfying (5.1) and using (4.2), we obtain
By Lemma 5.3,
is a polynomial of degree at most , which satisfies (5.1). Hence, we have to find only the polynomial . By Cramer’s rule, has only one solution; call this unique solution and define for . By (ii) and the Kronecker-Capelli theorem, has to satisfy also (5.9). Therefore,
Since satisfies (5.8), we get that . Thus, from (4.3) we obtain that for every . The ’moreover’ part easily follows from the above reasoning. ∎
In [18] the authors proved that in the class of composition operators such that the graph has one cycle and one branching point every completely hyperexpansive operator is 2-isometric. Using different approach, with the aid of Theorem 5.5, we are able to prove this result in full generality.
Theorem 5.6 (cf. [18, Corollary 2.15]).
Suppose is such that is connected and satisfies Theorem 3.5.(ii). Suppose . The following conditions are equivalent:
-
(i)
is 2-isometric,
-
(ii)
is completely hyperexpansive.
Proof.
(i)(i). By Lemma 5.1, if is 2-isometric, then it is -isometric for every . Combining this with [3, Proposition 1.5] we obtain that is completely hyperexpansive.
(ii)(i). From (ii), by [7, Remark 2], it follows that for every there exists the measure satisfying
(5.10) |
Inserting (5.10) into (4.3) we obtain that for every ,
(5.11) |
If , then for every we have
By (5.10) and Observation 4.3, for every ,
Thus, the series
is convergent for every and . By (5.10) we have that
Using (5) with replaced with , we get that for every ,
which implies that
Since for every , the integral on the left hand side is non-positive. On the other hand, we have for every , so each term in the sum on the right hand side is non-negative. Therefore, we get that for all , ,
(5.12) |
and for all ,
(5.13) |
By (5.13), . This implies that for all , with some . In turn, by (5.12), . Hence, for all , with some constants . Inserting these into (5) we obtain that for every and every ,
Hence,
(5.14) |
If , then, using (4.3) with and (5.10) with , we get that
if , by (4.2) applied with , we have
Combining the above with (5.14), we obtain
which implies that
Since for all , we have that for . On the other hand, . Therefore, for all , and, consequently, for every . Hence,
and
By (5.10), this implies that for and that for . It remains to show that
which implies that
Since we have already proved this for , , by Lemmata 3.4 and 2.3, it is enough to show that if for some , then for every satisfying . Assume is such that . Since
it follows from (5.10) that
(5.15) |
If we replace with in the above equality, we obtain
Hence,
which implies that for every satisfying we have . Thus, with some . Then (5.15) takes the form
But
Combining these two equalities we obtain that for every satisfying . ∎
As another corollary of the equality (4.3) we obtain that in our setting graphs of isometric composition operators consist only of cycles.
Corollary 5.7.
Suppose is such that is connected and satisfies Theorem 3.5.(ii). Suppose . If is isometric, then , that is, . In particular, is finite dimensional and is unitary.
Proof.
If is isometric, then so is for every . Thus, from (4.3) it follows that
Since for all , we have that
From this it follows that
Since for all , the only possibility for the above sum being zero is that the set of indices is empty, that is, . This implies that is finite dimensional. Hence, is invertible and, by [3, Proposition 1.23], unitary. ∎
In the end of this section we provide an example of use of Theorem 5.5. We characterize 3-isometric composition operators on graphs, in which the only branching points are the vertices on the cycle. Let and . Denote
Define by the formula
Then the graph takes the form
we present this graph in Figure 2

Theorem 5.8.
Let and be defined as above and let be a measure on such that for every . Assume that . The following conditions are equivalent:
-
(i)
is 3-isometric,
-
(ii)
for all ,
(5.16) and444Note that under the assumption we have that for every , ; similar argument shows that . By (5.16), we get that for , .
(5.17) where and
Proof.
(i)(ii). If is 3-isometric, then so is . By Theorem 5.5, this implies that for every and there exists a polynomial satisfying (5.1) with ; it is a matter of routine to verify that takes the form
(5.18) |
Since, by (4.1),
it follows that the polynomial satisfies the equality for every ; moreover, by (5.18) and (4.1), takes the form
Hence, (5.16) holds. Next, by (4.1) and (5.18), the polynomial given by (5.6) takes the form
By Theorem 5.5, the polynomial satisfying (5.1) with takes the form , where is the solution of the system
(5.19) |
Solving the above system of equations, we obtain that . From (4.2) and (4.1) it follows that for every ,
After rearrangement we get (5.17).
(ii)(i). Using (5.16), it can be verified that for every , the polynomial
satisfies (5.1) for . Next, from the equality
it follows that for every , and the polynomial
satisfies (5.1) with . Assume now . We check that . Set . As in the proof of converse implication, we obtain that the coefficients of satisfies (5.19). Combining (5.17) with (4.2) and (4.1) we deduce that satisfies
This implies that the coefficients of satisfies also the following system of equations:
where , . From the Kronecker-Capelli theorem it follows that . The application of Theorem 5.5 completes the proof. ∎
6. Subnormality of Cauchy dual of composition operator
In [18] the authors studied the Cauchy dual subnormality problem for 2-isometric composition operators with the property that the graph has one cycle and one branching point and obtained the solution in case of graphs with cycle of length 1. In this section we investigate the Cauchy dual subnormality problem for 2-isometric composition operators with graphs having one cycle in full generality. Again, we stick to the notation of Theorem 3.5.(ii). First, we are going to derive formulas for and .
Lemma 6.1.
Suppose is such that is connected and satisfies Theorem 3.5.(ii). Suppose . Then
-
(i)
for every , where is the only vertex satisfying , that is is the weighted shift on , where ,
-
(ii)
for every ,
-
(iii)
if is left-invertible, then
where , ; in particular, is the weighted shift on .
Proof.
(i). Let . Then
Hence, if and only if . By Observation 3.1, there is only one vertex satisfying ; for this vertex we have . (ii) and (iii) are simple consequences of (i). ∎
Note that if is assumed to be 2-isometric, then, by [25, Lemma 1], it is also an expansion, so that , .
Assume that is connected and satisfies Theorem 3.5.(ii). Suppose is -isometric. For further applications let us emphasize two formulas (cf. Lemma 4.4). For every and we have
and
(6.1) |
where
The above formulas are consequences of Lemma 6.1.(iii) and the fact that for 2-isometric every () is equal 1 (see Lemma 5.4). Note also that for every .
It is well-known (see e.g. [5]) that the Cauchy dual of 2-isometric operator is always a contraction (in our setting it can be easily derived from Lemma 6.1.(iii)). This observation will be crucial in our considerations about subnormality of the Cauchy dual of .
Note that if , then is finite dimensional. In such a case every 2-isometric operator is automatically invertible and, by [3, Proposition 1.23], unitary. Since for every unitary operator , if and is 2-isometric, is obviously subnormal. In the further investigation we always assume that . Let us present the main result in this section.
Theorem 6.2.
Let be such that is connected, satisfies Theorem 3.5.(ii) with additional assumption that . Assume that . If is -isometric, then the following conditions are equivalent:
-
(i)
is subnormal,
-
(ii)
for every ,
with , where
Proof.
(i) (ii). From [20, Theorems 3.1 and 2.2] and the fact that is contractive it follows that for every there exists the measure such that
In turn, by Theorem 5.5 and Lemma 6.1, we have
which implies that for . From this and (6) we derive that for every ,
Hence,
Since is regular, it follows that
(6.2) |
Since , by Corollary 5.7 we have that . This implies that for all , . Thus,
(6.3) |
with some . In turn, for every we have
The above equality implies that .
(ii) (i). From Theorem 5.5 it follows that
which implies, by Lemma 6.1, that
We check that
(6.4) |
Since (6.4) holds for , it is enough to check that if (6.4) holds for some , then it also holds for . By (6) we have
Using [20, Theorems 3.1 and 2.2] we obtain subnormality of . ∎
In [18] the authors proved that the Cauchy dual of 2-isometric composition operator on graph with one cycle of length 1 and one branching point is subnormal. As a corollary, we get the same result in a more general version.
Corollary 6.3 (cf. [18, Theorem 4.6]).
Assume that is connected and satisfies Theorem 3.5.(ii) with additional assumptions that and . If is 2-isometric, then is subnormal.
In [5, Example 6.6] the authors gave an example of 2-isometry, for which the Cauchy dual is not subnormal. With the help of Theorems 5.5 and 6.2 we can give another example.
Example 6.4.
Let and
The graph is exactly the graph defined at the end of Section 5 taken with and (see Figure 2). Set as follows
where . It is easy to verify that ; moreover, for every and , so the polynomial satisfies (5.1) with . The polynomial given by the formula (5.6) takes the form
Hence, in view of Theorem 5.5, is 2-isometric if and only if . The latter holds if and only if the polynomial
satisfies the equalities and , which is equivalent to the following system of equations
(6.5) | ||||
(6.6) |
Assume is 2-isometric. By Theorem 6.2, if is subnormal, then there exists satisfying
which implies that
(6.7) |
Note that
and
To obtain 2-isometric operator , the Cauchy dual of which is not subnormal, it is enough to find , which satisfy (6.5) and (6.6) and does not satisfy (6.7); this holds if we take, for instance, , .
Statements and Declarations
Funding. The author of the publication received an incentive scholarship from the funds of the program Excellence Initiative - Research University at the Jagiellonian University in Kraków.
Competing interests. The authors have no conflicts of interest to declare that are relevant to the content of this article.
Acknowledgements
The author would like to thank Zenon Jabłoński for his helpful suggestions concerning the final version of this paper. We are also grateful to anonymous reviewer for the comments improving the exposition quality of the paper.
References
- [1] Belal Abdullah and Trieu Le. The structure of -isometric weighted shift operators. Oper. Matrices, 10(2):319–334, 2016.
- [2] Jim Agler. A disconjugacy theorem for Toeplitz operators. Amer. J. Math., 112(1):1–14, 1990.
- [3] Jim Agler and Mark Stankus. -isometric transformations of Hilbert space. I. Integral Equations Operator Theory, 21(4):383–429, 1995.
- [4] A.V. Aho, J.E. Hopcroft, and J.D. Ullman. Data Structures and Algorithms. Adison-Wesley, Reading, 1983.
- [5] Akash Anand, Sameer Chavan, Zenon Jan Jabłoński, and Jan Stochel. A solution to the Cauchy dual subnormality problem for 2-isometries. J. Funct. Anal., 277(12):108292, 51, 2019.
- [6] Akash Anand, Sameer Chavan, Zenon Jan Jabłoński, and Jan Stochel. The Cauchy dual subnormality problem for cyclic 2-isometries. Adv. Oper. Theory, 5(3):1061–1077, 2020.
- [7] Ameer Athavale. On completely hyperexpansive operators. Proc. Amer. Math. Soc., 124(12):3745–3752, 1996.
- [8] Catalin Badea and Laurian Suciu. The Cauchy dual and 2-isometric liftings of concave operators. J. Math. Anal. Appl., 472(2):1458–1474, 2019.
- [9] S. Banach. Théorie des opérations linéaires, volume 1 of Monogr. Mat., Warszawa. PWN - Panstwowe Wydawnictwo Naukowe, Warszawa, 1932.
- [10] Teresa Bermúdez, Antonio Martinón, and Emilio Negrín. Weighted shift operators which are -isometries. Integral Equations Operator Theory, 68(3):301–312, 2010.
- [11] Teresa Bermúdez, Antonio Martinón, and Juan Agustín Noda. Products of -isometries. Linear Algebra Appl., 438(1):80–86, 2013.
- [12] Piotr Budzyński, Zenon Jan Jabłoński, Il Bong Jung, and Jan Stochel. Subnormality of unbounded composition operators over one-circuit directed graphs: exotic examples. Adv. Math., 310:484–556, 2017.
- [13] Sameer Chavan. On operators Cauchy dual to 2-hyperexpansive operators. Proc. Edinb. Math. Soc. (2), 50(3):637–652, 2007.
- [14] Sameer Chavan, Soumitra Ghara, and Md. Ramiz Reza. The Cauchy dual subnormality problem via de Branges–Rovnyak spaces. Studia Math., 265(3):315–341, 2022.
- [15] Zenon Jan Jabłoński, Il Bong Jung, and Jan Stochel. A non-hyponormal operator generating Stieltjes moment sequences. J. Funct. Anal., 262(9):3946–3980, 2012.
- [16] Zenon Jan Jabłoński, Il Bong Jung, and Jan Stochel. Weighted shifts on directed trees. Mem. Amer. Math. Soc., 216(1017):viii+106, 2012.
- [17] Zenon Jan Jabłoński, Il Bong Jung, and Jan Stochel. -isometric operators and their local properties. Linear Algebra Appl., 596:49–70, 2020.
- [18] Zenon Jan Jabłoński and Jakub Kośmider. -isometric composition operators on directed graphs with one circuit. Integral Equations Operator Theory, 93(5):Paper No. 50, 26, 2021.
- [19] Eungil Ko and Jongrak Lee. On -isometric Toeplitz operators. Bull. Korean Math. Soc., 55(2):367–378, 2018.
- [20] Alan Lambert. Subnormality and weighted shifts. J. London Math. Soc. (2), 14(3):476–480, 1976.
- [21] Singh A. Nair, M. Thamban. Linear Algebra. Springer, 2018.
- [22] Eric A. Nordgren. Composition operators on Hilbert spaces. In Hilbert space operators (Proc. Conf., Calif. State Univ., Long Beach, Calif., 1977), volume 693 of Lecture Notes in Math., pages 37–63. Springer, Berlin, 1978.
- [23] O. Ore. Theory of Graphs, volume XXXVIII of Amer. Math. Soc. Colloq. Publ. American Mathematical Society, Providence, 1962.
- [24] PawełPietrzycki. A Shimorin-type analytic model on an annulus for left-invertible operators and applications. J. Math. Anal. Appl., 477(2):885–911, 2019.
- [25] Stefan Richter. Invariant subspaces of the Dirichlet shift. J. Reine Angew. Math., 386:205–220, 1988.
- [26] Sergei Shimorin. Wold-type decompositions and wandering subspaces for operators close to isometries. J. Reine Angew. Math., 531:147–189, 2001.
- [27] R. K. Singh and J. S. Manhas. Composition operators on weighted spaces of vector-valued continuous functions. Acta Sci. Math. (Szeged), 58(1-4):453–472, 1993.
- [28] J.B. Stochel. Weighted quasishifts, generalized commutation relation and subnormality. Ann. Univ. Sarav. Ser. Math., 3(3):i–iv and 109–128, 1990.
- [29] Laurian Suciu. Operators with expansive -isometric liftings. Monatsh. Math., 198(1):165–187, 2022.