1. Introduction
Throughout this paper, our studies are directed toward the category of Hilbert spaces. Its objects are all Hilbert spaces, i.e., real or complex vector spaces endowed with hermitian positive definite sesquilinear forms on (inner products) such that the associated norms are complete. The morphisms of this category are all the maps C between Hilbert spaces and that are linear (i.e., for all scalars and ) and bounded (i.e., ).
The space of all bounded linear maps between Hilbert spaces and is actually a Banach space with the pointwise operations and the operator norm defined above. The adjoint of an operator is the operator , defined uniquely by the formula . The space has the structure of a algebra with the usual composition of operators and with the involution . Certain special operators are involved in our research: positive operators (i.e., ), contractions (i.e., ), (linear) isometries (i.e., ; equivalently, , where denotes the identity operator on ), coisometries (i.e., is isometric), and unitary (i.e., A is both isometric and coisometric) and normal operators (i.e., ). and are said to be unitarily equivalent if there exists a unitary operator such that .
If and are two Hilbert spaces, then the Cartesian product can be naturally endowed with a Hilbert space structure using the addition and multiplication by components and the inner product . The new structure is usually denoted by and called the direct sum between Hilbert spaces and . If and , then their operator direct sum is the map given by .
A subspace of a Hilbert space is a closed linear manifold of . Its orthogonal complement is the subspace of all vectors , which are orthogonal to (i.e., for every ; the usual notation for this orthogonality is ). The orthogonal complement is also a direct complement, i.e., and a representation of an element into the form is uniquely determined by the conditions and . Thus, we are entitled to define the orthogonal projection of onto , which is a bounded linear operator between and (or ), by . The (Hilbert) dimension of , denoted by , is the number of elements in an orthonormal basis (i.e., a maximal set of unit norm pairwise orthogonal vectors in ). For two subspaces and of , we use the notation for the usual sum indicating the fact that and are orthogonal (orthogonal sum). For a family of subspaces, their intersection is also a subspace. This property is not inherited by their union. In this case, one can consider their closed linear span , i.e., the smallest subspace of , which contains all the subspaces . If are pairwise orthogonal, we use the notation for the closed span . If , then its kernel is a subspace of , while its range is generally only a linear manifold of . A subspace of is invariant under if . is reducing for A if it is invariant under both A and .
These basic facts on the category of Hilbert spaces can be found in any introductory book on functional analysis or linear operator theory. The author’s recommendations are [
1,
2,
3].
The famous dilation theorem of Béla Sz.-Nagy [
4] shows that every contraction
T that acts on a Hilbert space
can be (power) dilated by an isometric (equivalently, a unitary) operator
V on a bigger space
:
Nowadays, dilations represent an important instrument of study in many disciplines such as invariant subspace theory, interpolation theory, operator algebras, dynamical systems, control, prediction theory, and so on (see the excellent survey article of Orr Moshe Shalit [
5] or the books by Sz.-Nagy, Foiaș, Bercovici and Kérchy [
6], Foiaș and Frazho [
7], Foiaș, Frazho, Gohberg and Kaashoeck [
8], Rosenblum and Rovnyak [
9], Kakihara [
10], and so on).
If
is a coisometric extension of
T, i.e.,
is invariant under
and
, then
V is an isometric dilation of
. In minimality conditions, i.e.,
, the two concepts are actually equivalent. In addition to the numerous applications in dilation theory, the theory of extensions was proved by P.S. Muhly and B. Solel to be useful in the representation theory of tensor algebras associated with
correspondences or in the study of
dynamical systems [
11,
12]. A good reference in this context is the Ph.D. thesis of T. Wolf [
13].
In our approach, a coisometric extension on a Hilbert space
of the contraction
T is a pair
, where
and
are isometric operators such that
(
T is a quasi-affine transform of
, with quasi-affinity
, according to the terminology introduced in [
14]). In fact, in this setting,
is invariant under
and
is unitarily equivalent with
T. Given the operator
V, it is our aim in this paper to characterize
T such that there exists
, which intertwines
T and
. We also describe and classify the isometric operators
. This type of extension has been considered recently, e.g., in [
15,
16].
By the Wold–Halmos decomposition theorem (Theorem 1), any isometric operator on
can be represented as the direct sum between a unitary operator and a unilateral shift. Unitary operators are well understood. They have a spectral representation and an associated functional calculus. On the other hand, unilateral shifts possess a simple geometrical structure. In fact, a
unilateral shift (or, simply, a shift) is just an operator that is unitarily equivalent to the operator
of multiplication by the independent variable,
, on the
-valued Hardy space
for a certain complex Hilbert space
. In fact, an isometric operator
S on
is a shift if and only if there exists a subspace
of
, which is
wandering, i.e.,
and
generating, i.e.,
The subspace
is called the
defect of
S and can be computed in terms of
S (in fact,
). The dimension of
, i.e., the
multiplicity of
S determines
S up to a unitary equivalence. Let us also note that
S is a shift if and only if it is completely non-unitary (i.e., there are no non-null subspaces reducing
S to a unitary operator) if and only if
S is of class
(or,
is strongly stable). As introduced in [
6] (Chapter II.4), a contraction
T on
is of
class if
tends strongly to 0 as
(i.e.,
for every
).
T is of
class if
belongs to the class
. A very good introductory material on shifts is [
17]. The included terminology has been extracted from [
6] (Chapter I).
Let A and B be bounded linear operators acting on complex Hilbert spaces and , respectively. Let also be a linear isometry.
Definition 1. The pair is said to:
The paper starts with a general discussion on the link between extensions and dilations. Their main properties are emphasized here. We show that an extension is always a dilation. However, if is a dilation of A, then is also an extension of A if and only if is invariant under B. We also prove that for the property of a pair for a dilation, respectively, an extension is preserved when we restrict B to one of its invariant subspaces, which contains .
Definition 2. A dilation of A is said to be minimal
if is the smallest subspace that is invariant under B and contains , i.e., We observe that, for a given minimal dilation
of
A, its adjoint
extends
. We also indicate necessary and sufficient conditions for two minimal dilations of
A in order to be equivalent. The novelty of the results presented in this section is that they are formulated in full generality, while in books on dilation theory, the study is restricted to the case when
A is a contraction,
B is isometric or unitary,
is a subspace of
, while
is just the inclusion map (see, e.g., [
6]).
In
Section 3, we present necessary and sufficient conditions on an isometric operator
V on
in order to be
-extended by a given unitary operator
U on
. This description is given only by numerical invariants, i.e., multiplicities of certain isometric operators associated with
U. In order to formulate this result precisely, we include some definitions for increased readability. A complex spectral measure is a map
E on the Borel
-algebra
on
into
such that
,
for every Borel set
and
for every pairwise disjoint family
of Borel sets. We sketch the construction of Halmos [
18] (Chapter III) for the multiplicity function
u attached to the spectral measure
E. For a given nonzero finite measure
on
, its multiplicity
is the minimum value between the powers of maximal orthogonal systems of type
for nonzero measures
, which are absolutely continuous with respect to
, where, for
,
is the measure
. We show that
extends
V if and only if
the spectral measures of the unitary part of
V and the restriction of
U to one of its reducing subspaces
possess identical multiplicity functions and
the shift part of
V and the shift part of
have identical multiplicities, where
is a subspace of
, which contains
and it is invariant under
U.
An important result of Ciprian Foiaș [
19] describes the structure of contractions of class
: they can be represented as restrictions of backward shifts (i.e., adjoints of shifts). We prove, in the following section, that there exists an isometric operator
such that
extends
T if and only if
T is a contraction of class
and
. Any solution
of the equation
has the form
, where:
where
is an arbitrary isometric operator. Here, according to the B. Sz.-Nagy and C. Foiaș terminology [
6] (Chapter I.3),
is the
defect operator of
T,
is the corresponding
defect space, and
is the
defect index of
T.
A model for the minimal isometric dilation
V of
T (hence, the model also describes the coisometric extension
of
) has been proposed by Schäffer in [
20]. More precisely, the minimal isometric dilation of
T is given on the Hilbert space
by the formula:
We deduce that the operator
is a coisometric extension of
. This result was the starting point for developing a precise geometrical structure for such dilations, respectively extensions (cf. [
21,
22,
23]). In the model above, in the terminology introduced in Definition 1, the pair
extends
, where
is the embedding
. In our approach, our job is, given
V, to find necessary and sufficient conditions on
T in order to ensure the existence of an isometric operator
such that
. Therefore, we look for operators
that might be different from just an embedding. In order to solve this problem, we need to introduce a special positive contraction. Since
is a decreasing sequence of positive operators, it has a strong limit, which will be denoted by
and called the
asymptotic limit associated with
T. It has been used as a tool in the construction of the isometric dilation in [
6] (Chapter I.10), in various invariant subspace problems [
24], in similarity problems [
25], for Putnam–Fuglede-type results [
26], and so on (see, also, [
27,
28,
29,
30,
31]). It seems that this limit appeared for the first time in 1967, in the french edition of the book [
6]. Associated notions are the
asymptotic space of
T, i.e., the closure
of the range of
and the
asymptotic index of
T, i.e., the Hilbert dimension of
.
In the last section of the paper, we move on to the general case in which
V is written, by the Wold–Halmos decomposition theorem (Theorem 1), as the direct sum between a unitary operator
U and the shift
. We prove that, given complex Hilbert spaces
and a unitary operator
U on
, there exists an isometric operator
such that
extends
T if and only if
T is a contraction,
and
extends the isometric operator
for a certain isometric operator
Y that acts from
into
. Any solution
to this problem has the form:
where
is an isometric operator that intertwines
and
, while
has the form:
X being any isometric operator from
into
. We also characterize
-extensions in which the corresponding adjoint dilations are minimal. Several examples are given for illustrative purposes.
2. Extensions and Dilations
Throughout this section, the symbols and denote complex Hilbert spaces with , is a linear isometry, while A and B are bounded linear operators acting on and , respectively.
When
is a closed subspace of
, one can take
as the inclusion map of
into
. Then
is the orthogonal projection
of
onto
. Hence,
is the usual extension of
A:
is invariant under
B and
. At the same time, the pair
dilates
A if
which is exactly the concept introduced by Béla Sz.-Nagy in [
4].
Even under minimality conditions, a dilation might not be unique, up to a unitary equivalence. One may note that, in general, the structures of and B could become quite complicated.
It is an immediate consequence of Definition 1 that is a dilation of A if and only if is a dilation of . Such a property is not valid, in general, for extensions. Another easy consequence of the corresponding definitions is their transitive properties: if extends (dilates) A and extends (dilates) B, then extends (dilates) A. Here, C is a bounded linear operator that acts on the complex Hilbert space , and is a linear isometry.
Remark 1. If extends A, i.e., , then and, inductively, for every . This shows that dilates A.
The converse is, in general, false. However,
If dilates A, then also extends A if and only if is invariant under B. Indeed, for each , it holds:so Hence, extends A if and only if . Since , we deduce that, for every , must be in the range of π. It follows that , as required.
In particular, if dilates A, A is an isometry and B is a contraction, then extends A. Indeed, for , and . This forces in (2). If is invariant under B, then is a restriction of , i.e., extends A. Any restriction of has this form. We use the fact that is invariant under B if and only if , that is . Hence, is an extension of A. If is a restriction of , then If is a subspace of , which is invariant under B and contains , then one can consider the operators and defined by . Since , we immediately deduce that dilates (extends) A if and only if dilates (extends) A. This property can be extended to the case when is not necessarily a subspace of .
Proposition 1. Let be a complex Hilbert space and an isometric operator such that contains , and it is invariant under B. Then dilates (extends) A if and only if dilates (extends) A. In this case, extends , which is unitarily equivalent to , i.e., , where is the unitary operator Proof. As seen before,
is invariant under
B if and only if
. Observe, also, that
is an isometric operator:
We used, for the first equality above, the fact that contains and .
We note that, for
and
, the following set of inequalities holds true:
According to (
3),
is a dilation of
A if and only if
is a dilation of
A. In view of (
4), if
A is extended by
, then
. Conversely, if
A is extended by
, then it follows by (
4) that
. Since both
and
are elements of
and
, we deduce that
. Therefore,
, as required.
extends followed by Remark 1. The unitary equivalence between and is a consequence of the definition of . □
We now start with a bounded linear operator A on , with a Hilbert space such that and with a linear isometry . Our next aim is to build and parametrize the class of all operators such that is an extension of A.
In order to obtain the equality,
is enough to define
B on
as follows:
One can take for
any bounded linear operator from
into
. Thus, according to the decomposition
,
B has the matrix representation:
where
and
.
We observe that, by an easy computation with matrices, B is isometric if and only if A is isometric, is orthogonal to , and is a row isometry (i.e., ). One can take and isometric operators in order to build an isometric extension of A.
Similar computations show that B is coisometric if and only if Y is coisometric, and are orthogonal and . Taking again , the extension is coisometric if and only if A and Y are coisometric operators.
We deduce that B is unitary if and only if A is isometric, Y is coisometric, , and .
As observed in Proposition 1, the dilation extension properties, respectively, are preserved if we restrict B to one of its invariant subspaces, which contains .
As we will see later in this section, the minimality condition does not always ensure the uniqueness, up to a unitary equivalence, of such a dilation.
As noted earlier,
is a dilation of
A if and only if
is a dilation of
. However, there is no reason to assume that the minimality of
implies the minimality of
. If
extends
A, then the smallest subspace of
that is invariant under
B and contains
is
In other words, if
is the unitary operator
and
, then
is an extension of
A, which is unitarily equivalent to
A:
Hence, this situation does not present any real interest.
A dilation is not, in general, an extension. However, under the minimality assumption, the dilation adjoint is actually an extension.
Remark 2. If
is a minimal dilation of
A, then
is an extension of
. Let and . Then since for . In view of the minimality condition (1), this implies that . Our final aim in this section is to express the necessary and sufficient conditions in order to ensure that a minimal dilation is unique.
Definition 3. Two minimal dilations and of A are said to be equivalent if there exists a unitary operator (here, and ) such that and .
If and are equivalent, then, according to Definition 1, extends and extends .
We are now in a position to present the conditions that describe the equivalence between two minimal dilations.
Proposition 2. Two minimal dilations and of A are equivalent if and only if Proof. The operator
U must be defined on the generators of
, which are, by minimality,
, where
and
. More precisely, in view of the axioms of Definition 3, we must have:
Formula (
6) correctly defines an isometric operator on
if and only if
for all sequences
and
of elements in
with finite support. We can write (
7) in equivalent form as:
By minimality, the operator
U, as defined in (
6), is also surjective. Hence,
U is a unitary operator. Moreover, by the same Formula (
6),
and
that is,
, according to the minimality of
. □
Minimal extensions are indeed equivalent: if
extends
A, then
it is independent of the choices of
and
B.
Examples of equivalent classes of dilations are the ones of selfadjoint operators and isometric operators, respectively.
3. Unitary Extensions
In this section, we provide necessary and sufficient conditions on a given unitary operator U in order to extend, in the sense of Definition 1, a given isometric operator V. We also discuss the structure of U.
The Norwegian statistician Herman Wold discovered in [
32] a remarkable decomposition for every stationary stochastic process. More precisely, his decomposition separates the deterministic part from the part corrupted by noises (in fact, the moving average of white noise). It was the cornerstone of prediction theory for such processes and has nowadays applications in many domains such as machine learning [
33], traffic flow prediction [
34], modeling of helicopter rotor aerodynamic noise [
35], or image processing [
36]. The Wold decomposition theorem has been formulated for the general case of isometric operators on Hilbert spaces (cf., e.g., [
17,
37,
38]):
Theorem 1. Let V be an isometric operator on the Hilbert space . Then there exists an orthogonal decomposition of the formuniquely determined by the conditions: - (a)
reduces V to a unitary operator ;
- (b)
reduces V to a shift .
More precisely, and .
According to the classical spectral theorem, to every normal operator
N on a complex Hilbert space
, it corresponds a unique compact, complex spectral measure
E such that
(cf., e.g., [
18], §44, Theorem 1). It follows immediately that
for any given unitary operator
U from a complex Hilbert space
onto
. In other words, two normal operators
and
are unitarily equivalent if and only if their spectral measures
and
are unitarily equivalent:
for a certain unitary operator
U and for all complex Borel sets
. According to the theory developed for nonseparable Hilbert spaces by Paul R. Halmos [
18] (Chapter III) and Arlen Brown [
39], the equivalence of two spectral measures is characterized by the equality of their multiplicity functions. Therefore, it holds the following description of the unitary equivalence between two normal operators.
Theorem 2. Two normal operators are unitarily equivalent if and only if their spectral measures possess identical multiplicity functions.
We now have all the ingredients to prove the main result of this section.
Theorem 3. Let V be an isometric operator on a complex Hilbert space with the Wold decomposition . Furthermore, let U be a unitary operator on another complex Hilbert space . Then, there exists an isometric operator such that extends V if and only if:
- (a)
There exists a subspace of , which reduces U such that the spectral measures of the unitary operators and possess identical multiplicity functions;
- (b)
There exists a subspace of , which contains , and it is invariant under U such that the shifts and possess identical multiplicities (here, ); equivalently, .
The operator π has the form:Here is a unitary operator that satisfies , while is the unitary operator defined for every (), bywhere φ is a unitary operator between and . Proof. We firstly show that, for a subspace
of
that is invariant under
U,
is wandering for
U and
Indeed, for
and
,
and
, that is
Furthermore, for any
,
The equality
is a consequence of (
10).
Let us now suppose that, for a certain isometric operator
,
extends
V. Then
is a closed subspace of
(as
is isometric), which reduces
U:
The unitary operator
intertwines
and
:
In view of Theorem 2, the spectral measures of
and
possess identical multiplicity functions. Finally, the subspace
is invariant under
U (indeed,
), it contains
and
Hence, and have identical Hilbert dimensions.
Conversely, let
be a reducing subspace for
U,
, a unitary operator that intertwines
and
,
, a subspace of
that is invariant under
U, contains
and such that there exists a unitary operator
(according to
and
). Let
be the operator defined as in (
9). Then, for every
(
),
This shows that
is an isometric operator. It is also surjective since for every
(
),
and
. In addition,
We deduce that .
The operator
, defined by (
8), is isometric:
since
and
are orthogonal subspaces:
due to the observation that
and
.
This proves that extends V. □
As observed in the first section,
dilates
. We can thus consider the smallest subspace of
, which is invariant under
and contains
:
We observe that
that is,
reduces
U. We next show that:
The sum in the right side of this equation is orthogonal:
If we apply
U to equation (
11), we obtain the relation:
which is obviously true. We proceed inductively to prove that:
Let
be the isometric operator defined by:
where
and
,
We deduce, by Proposition 1, that is a unitary extension of V, while is a unitary extension of .
We proved the following theorem.
Theorem 4. Let V be a linear isometry on the complex Hilbert space . Then is a unitary extension of V, where is the inclusion map of into , while W, which acts on , has the matrix representation: being the map of evaluation in 0: In addition, is a minimal unitary dilation of .
5. The General Case
At the beginning, we present two examples. The first one shows that the converse of the result presented in Remark 2 is not always valid.
Example 2. Let be non-null complex Hilbert spaces, S a shift on and acting on .
The operator V defined on by:is isometric:and its adjoint can be computed according to the formula If π is the embedding of into :then extends T: In order to prove that the isometric dilation of is not minimal, we compute the smallest subspace, which is invariant under V and contains : With this aim, we observe that, for every and , and, inductively, Obviously, for a given with , for every . This observation proves that is strictly contained in , so the dilation is not minimal.
The second example shows that the minimality condition is necessary in order to obtain the conclusion of Remark 2.
Example 3. We use the objects (spaces and operators) from the previous example. Let us firstly observe that reduces V since, on the one hand, and, on the other hand, for and , Let be a fixed element in of unit norm and . Then . Consider also a fixed function of norm 1. Let and . The operator W, defined on by:is isometric. Indeed, for any , , while g and are elements of . Then In this series of equalities, we used the fact that since and . The adjoint of W can be computed according to the formula: Then, for any , it holds: We deduce that cannot be an extension of T since, otherwise, , which contradicts its choice. We also observe that, by (17), , and, inductively, for every and . Consequently, Hence, is a dilation of .
We pass now to the study of the structure of coisometric extensions of contraction operators T on a complex Hilbert space .
In the description of , the asymptotic limit associated with T plays an important role. Some of its main properties are collected in the following proposition.
Proposition 3 ([
40], Chapter 3).
Let be the asymptotic limit associated with a contraction . Then:- (a)
;
- (b)
;
- (c)
;
- (d)
.
In view of condition
, we can show that, for all
, the following equalities hold true:
This formula allows us to define the isometric operator
which will be called the
asymptotic isometry associated with
T.
According to the Wold–Halmos decomposition theorem (Theorem 1), the isometric operator V can be represented as the direct sum between a unitary operator and a shift. This is the reason why we study coisometric extensions , where V is the direct sum between a unitary operator U acting on a complex Hilbert space and the operator of multiplication by the independent variable on the Hardy space .
Theorem 7. Let be complex Hilbert spaces, and a unitary operator. Then there exists a linear isometry such that extends T if and only if , and there exists an isometric operator such that extends .
Any solution π of the equation has the form , where being any isometry with , while being an arbitrary isometric operator. Proof. Let
be an isometric operator such that:
If
has the matrix representation
, where
and
, then Equation (
20) can be translated by:
Furthermore,
due to the fact that
is isometric.
We proceed as in the proof of Theorem 5. Since
W has the form:
where
, we deduce immediately that
. Equation (
22) becomes:
We replace
h with
in (
23) and obtain that:
We finally arrive at the equality
by (
12) and the first condition of (
21). Equivalently, there exists an isometric operator
such that
. This also implies that
. We replace our findings in (
23) and use (
12) again to obtain that, for every
,
Since
tends strongly to
as
, it also tends weakly to the same limit. Thus,
Let
be an isometric operator such that
. The equation
can be rewritten as:
or, equivalently, as:
This means that
is a unitary extension of the isometric operator
. As a conclusion,
has the form
given by (
18).
Conversely, if
T is a contraction with its defect index
, there exists an isometric operator
. Let
be the isometric operator that intertwines
and
. The operator
, introduced by (
18) and (
19), is well-defined and isometric since
In addition,
which shows that
extends
T. □
Sometimes, we can numerically express this extendability problem, as we can see in the following example.
Example 4. Let be complex Hilbert spaces, a non-null subspace of , , a unitary operator and ( denotes the unit circle). If denotes the orthogonal projection of onto , then has unit norm. Obviously, and . Consequently, , , , and . We deduce immediately that . Therefore, extends for a certain given isometric operator if and only if reduces U and (equivalently, ). In other words, there exists such an isometric operator Y if and only if (here, the nullity of an operator is the Hilbert dimension of its kernel). Moreover, for every , With these computations, Theorem 7 takes the form:
“There exists a linear isometry such that extends if and only if and .
Any solution
of the equation
has the form
, where
while
and
being arbitrary isometric operators.”
Remark 3. The condition on the existence of an isometric operator with the property that is a unitary extension of can be characterized by Theorem 3 as follows:
- (a)
There exists a subspace of , which reduces U such that the spectral measures of the unitary operators and possess identical multiplicity functions;
- (b)
There exists a subspace of that contains , and it is invariant under such that the shifts and possess identical multiplicities.
We describe the necessary and sufficient conditions under which the dilation
of
is minimal:
Our aim is to compute the subspace
. Let
. Then
so
where
is the map:
One can apply, successively, the operator
to this equation in order to obtain that:
We deduce that .
We proceed by computing the powers
for every
and
. An inductive procedure shows that:
where
. Since
by our remarks above, it follows that
Finally, by a summation of subspaces,
The converse inclusion is obvious.
We conclude that if and only if and . Equivalently, X is a unitary operator and the dilation of is minimal.
With these findings, we can now reformulate Theorem 7 in order to obtain minimal isometric dilations.
Theorem 8. Let be complex Hilbert spaces, and a unitary operator. Then there exists a linear isometry such that is a minimal isometric dilation of T if and only if , and there exists an isometric operator such that is a minimal unitary dilation of .
Any solution π of this problem has the form , where being any isometry with , while being an arbitrary unitary operator. Remark 4. If is the unitary operator defined by:then is a unitary extension of , the dilation of is minimal and has the matrix representation: These observations are consequences of Theorem 4.
We also discuss the particular situation when U is the identity operator on . Then the condition that extends can be expressed by . Equivalently, or, as and have identical closures of their ranges, .
Corollary 1. Let be complex Hilbert spaces and . Then there exists a linear isometry such that extends T if and only if , , and .
Any solution π of the equation has the form , whereand and being arbitrary isometric operators. Corollary 2. Let be complex Hilbert spaces and . Then there exists a linear isometry such that is a minimal isometric dilation of T if and only if , , and .
Any solution π of this problem has the form , whereand and being arbitrary unitary operators. 6. Conclusions
Given a bounded linear operator T acting on a complex Hilbert space and an isometric operator V, which is defined on the complex Hilbert space , we completely describe the conditions that should be imposed on T in order to ensure the existence of an isometry such that . We provide, in addition, parametrizations for the set of all solutions of this problem. We also discuss the connection with the problem of finding linear isometries such that for every positive integer n. The two problems are actually equivalent in minimality conditions, i.e., when is the smallest subspace, which is invariant under V and contains . This means that a result on extensions can be translated into a result on dilations and the other way around. However, when the minimality assumption is not satisfied, dilations are more general than extensions. We formulated an example in which is a dilation of , but is not an extension of T. Another example shows that even if is an extension of T, the dilation of is not always minimal.
When is a subspace of and is the corresponding inclusion (so, is the orthogonal projection of onto ), the first equation becomes an extension problem, i.e., , while the second one is the classical dilation problem proposed by Sz.-Nagy, i.e., for every . In other words, we propose generalized versions of the classical results and provide some clarification on the structure of these dilations or extensions. Our approach is, in some sense, different from the one usually followed in dilation theory. We have the operator V, with all its structure, and we study the properties of T such that it can be -extended to . Finally, we precisely describe the operator .
The first particular situation that was taken into consideration in this study is when
V is a unitary operator. In this case,
T must be a linear isometry. Our characterization is exclusively numeric, and it involves the multiplicities of certain unilateral shifts and the multiplicity functions associated with the spectral measures of some unitary operators. It is also connected with the invariant subspace theory since it requires the existence of some subspaces that are reducing for
V or that are invariant under
V. Passing to the case when
V is a given shift, we obtain similar results as the one obtained by Foiaș in the classical case, i.e.,
T must be a contraction of class
. We present, in addition, the exact form of
. In full generality, by the Wold–Halmos theorem, a linear isometry
V can be written as the direct sum between a unitary operator
U on a Hilbert space
and a unilateral shift
(i.e., the multiplication by the independent variable on a certain
-valued Hardy space on the unit disc). We obtain that
T must be a contraction, its defect index is less than the Hilbert dimension of
and
is a unitary extension of the asymptotic isometry
associated with
T. For the last assertion, in order to obtain a more precise description, we are able to use the results of
Section 3. One can also note that the isometry
associated with the asymptotic limit
has been extensively studied in the literature, so we have a good chance of clarifying the connection between
T,
and
. We also gave an example in which
U has a particular form; namely, it is the identity operator
. In this case,
and
.
While the classical results were widely studied and are useful in many domains, e.g., in invariant subspace theory, interpolation theory, prediction, control, and so on, it is natural to expect that these generalizations would provide larger classes of applications. Some steps forward in this direction were already made recently (see, e.g., [
15,
16,
41,
42]). A natural way to extend these results is to pass to the multivariable case, i.e., to systems
of commuting contractions acting on the same Hilbert space
. The starting point in this direction is a result of Curto and Vasilescu [
43] (a refined version has been proposed in [
44]), who provided necessary and sufficient conditions on such a tuple in order to possess a backward multishift extension. This subject will be treated elsewhere.