3.1. Canonical Quantization on a Hilbert Space
Consider a linear classical system (with a finite or infinite number of degrees of freedom), described by a symplectic vector space, that we will call . The set of classical observables will hereby be denoted by . Roughly speaking, by quantization we will understand the passage from a classical description of a system to a quantum mechanical description. In contrast to the situation in the classical theory, where states live in the phase space and observables are real-valued functions on , in the quantum theory states belong to a Hilbert space , whereas observables are self-adjoint operators on . The basic PB, that equip the space of classical observables with an algebraic structure, are replaced at the quantum level with the canonical commutation relations (CCRs), that define an algebraic structure on the space of quantum observables. Thus, in very broad terms, the output of the quantization should be a Hilbert space of quantum states, and quantum observables represented on as self-adjoint operators, obeying the algebraic structure arising from the CCRs. For linear systems, the process of canonical quantization on a Hilbert space consists of (and it is accomplished by) three main steps:
- (i)
A selection of basic (elementary, or fundamental) classical observables .
- (ii)
The construction of an abstract quantum algebra of observables from , with the following two properties: (iia) for each basic observable there must be one, and only one, abstract quantum basic operator (observable) , and (iib) basic operators must satisfy the Dirac quantization condition, relating their commutators with the corresponding PBs.
- (iii)
The specification of a Hilbert space and a representation of the abstract basic observables as self-adjoint operators on .
For more details, we refer the reader, e.g., to Ref. [
61].
These rules are far from determining a unique quantum description. Indeed, the process entails ambiguities at different stages, and a series of choices must be made in order to accomplish the quantization and arrive to a, hopefully, well specified description. In fact, one has to face ambiguities from the very beginning of the process by making “a judicious selection” of fundamental observables
. This set of basic observables
is typically required to be a vector subspace of
, closed under PB, and such that every regular function on phase space can be obtained by (possibly a limit of) sums of products of its elements [
61,
62]. These requirements are intended to achieve that observables in
will be appropriately promoted to quantum operators satisfying the CCRs, allowing to avoid ambiguities like e.g., the well-known problem of factor ordering. However, it is not uncommon that various distinct basic sets can be found for the same system. So, in general there is not a unique canonical choice of elementary observables
, a fact which can give rise to non-equivalent quantum descriptions. This ambiguity is usually addressed by arguing “naturalness and simplicity” in favor of a particular classical canonical representation.
Once the set of fundamental observables is specified, the next step in the quantization is to construct an abstract quantum algebra of observables from the vector space . The algebra is constructed as follows. Let be the free associative algebra over the complex numbers generated by , i.e., the free associative complex algebra corresponding to . Thus, every has a representative in , where is a linear mapping. Next, the algebra is equipped with an involution operation ∗ which captures the complex conjugation; so, the representative of the real, basic observable is invariant under the involution operation, . More generally, if and only if , where . Then, the algebraic structure on the space of classical observables, provided by the PB, is carried to an analogous algebraic structure on quantum observables. For this, one takes the ∗-ideal of generated by elements of the form . This is precisely the Dirac quantization condition. The algebra of abstract quantum observables is the quotient algebra of by the ideal . The associative algebras and are related by the homomorphism , . Let ∧ be the mapping , and let us define . Thus, in particular, we have that for each , there is one and only one ∗-invariant operator . Given F, G, and in , their (abstract) ∗-invariant, basic operator counterparts , , and in satisfy the CCRs . By construction, any can be expressed as a sum of products of elementary operators.
The third and final step in the process is to find a Hilbert space supporting a representation of the (abstract) fundamental quantum observables as self-adjoint operators. This representation, however, turns out to be not unique in general. There exist, typically, different (i.e., not unitarily equivalent) Hilbert space representations of the CCRs. So, one generally has to deal with the problem of determining a preferred representation. It should be noted that, in contrast with the ambiguity in the choice of basic observables, which affects both linear mechanical and linear field theory systems, the lack of uniqueness of the representation is mainly an issue for field (i.e., infinite dimensional) systems. In fact, for linear, finite dimensional systems (i.e., linear mechanical systems), the specification of a unique preferred representation of the CCRs can be consistently and unambiguously established under certain requirements. Indeed, in view of the Stone-von Neumann uniqueness theorem, we can restrict our attention just to a single representation of the CCRs, namely the ordinary Schrödinger representation of quantum mechanics. However, the situation is quite different for linear field theories. There are infinitely many inequivalent Hilbert space representations of the basic quantum observables as self-adjoint operators, and no analogue of the Stone-von Neumann theorem exists to confront the uniqueness issue. To handle this ambiguity in the representation of the CCRs, the usual procedure is to appeal to the spacetime symmetries of the field system and look for symmetry invariant representations. Though this strategy leads to a unique quantum theory for a certain class of field systems (for instance, linear field theories in both Minkowski and stationary spacetimes), it should be stressed that in more general cases (like e.g., non-stationary settings) symmetries will simply not be enough to pick out a preferred representation and, therefore, extra criteria must be imposed in order to set a unique quantum theory.
3.2. Linear Scalar Field Theory: Quantization
Let us now review the quantization of the scalar field theory introduced in
Section 2.2. We first consider the covariant phase space approach. As we have seen, linear functionals
provide a natural set of observables on
with non-trivial PB that are proportional to the unit function [see Equation (
12)]. Since observables on
can be obtained by taking linear combinations of products of natural observables
and the unit function
(which provides the constant functions on
S), the subspace
of
qualifies as an admissible set of basic observables. (Here,
denotes the vector space given by the set
over the field
.) This, together with the “naturalness and simplicity” of our choice, leads us to select the commented subspace as the set
of fundamental (basic, or elementary) classical observables.
By equipping the phase space
with a compatible complex structure
J, the field
can be decomposed into the “positive and negative frequency” parts,
and
, defined by
J. In addition, the completion of the inner product space
in the norm
defines the “one-particle” Hilbert space
. Notice that
is an inner product not only for
, but also for
; in fact,
and
are orthogonal subspaces with respect to this product. Thus, the Cauchy completion of
gives a complex Hilbert space
H, which decomposes into the orthogonal
-eigenspaces of
J, with the
-eigenspace being precisely the so-called one-particle Hilbert space
, whereas the
-eigenspace is the complex conjugate of
,
. Let
and
be the orthogonal projections arising from the inner product
. The restrictions of
and
to
S are nothing but the real-linear bijections from
S to
and
, respectively. In terms of the restrictions of
and
to
S, the field decomposition defined by
J reads
. Thus, basic observables can be written in the form
, where
are, respectively, the annihilation and creation-like variables associated with the complex structure
J. By using complex linearity and continuity, we get that
where
and
, with
.
The next step in the process of quantization is to specify
, the algebra of abstract quantum observables. According to the discussion in
Section 3.1, this algebra is constructed from the complexification of
,
. However, notice that we have an enlarged vector space
, perfectly valid to construct the algebra. So, we will take
to specify
. Since every (complex) elementary variable
, with
, can be uniquely expressed in the form (
24), the vector space
can be naturally rewritten as
The only non-zero PB between the complex elementary variables (
25) are
where
[i.e., the Hermitian inner product (
9)]. The quantum algebra
is defined starting with the complex vector space
(
25), exactly as we have explained in
Section 3.1 (with
replaced by
). As a result of the construction, we get operators
and
, satisfying
and obeying the commutation relations
for all
. The abstract quantum counterparts of the basic observables
are given by the ∗-invariant elementary operators
that fulfill the CCRs
In order to accomplish the quantization, we need to specify a Hilbert space supporting a representation of the fundamental quantum observables
as self-adjoint operators. Note, however, that a Hilbert space structure has been already chosen from the introduction of a complex structure: the one-particle Hilbert space
. It is from
that the Hilbert space of the quantum theory is constructed. Concretely, the one-particle Hilbert space defines the symmetric Fock space
that is the desired Hilbert space. The structure of
allows for a natural representation of
and
, subject to the commutation relations (
27), as the annihilation and creation operators
and
on
. Thus, the fundamental observables
are represented on the Fock space by the self-adjoint operators defined by Equation (
28), obeying the CCRs (
29). This is the standard procedure, in the covariant approach, for the Fock quantization of a linear scalar field
given a complex structure
J. Since
J can be any compatible complex structure, what we really have is a family of Fock representations of the CCRs parameterized by the set
. This set splits naturally into equivalence classes
of complex structures that lead to unitarily equivalent Fock representations, and it is well known that
is formed by an infinite number of them. That is, there are infinitely many inequivalent Fock representations of the CCRs. Therefore, in order to specify a unique quantum description, up to unitarity, a preferred complex structure
(or, more generally, an equivalence class
) must be chosen.
Let us recall that, in general, there are no representations of the CCRs by bounded operators [
1,
63]. For the KG field, the quantum fundamental observables
turn out to be all unbounded operators [except
]. So, questions concerning the domains of definition should be treated carefully. For instance, a proper definition of the elements of the quantum algebra on the Hilbert space becomes an intricate task, because
contains polynomials. In order to avoid this unwieldy situation, the usual procedure is to consider the exponentiated version of
, namely
. Formally, the CCRs are replaced with the Weyl relations
together with the adjoint relations
By equipping the vector space spanned by all finite, complex linear combinations of the
’s with the product (
31)—extended by linearity to the vector space—and the involution operation (
32), we get a complex associative ∗-algebra [with unit element
]. Given a Hilbert space representation of the CCRs (
29), the ∗-algebra generated by the
’s becomes a subalgebra,
, of the
-algebra of all bounded linear operators on the Hilbert space. The closure of
thus leads to a
-(sub)algebra
, which is known as the Weyl algebra. Although one might think that the Weyl algebra defined in this way would be a representation-dependent algebra, actually this is not the case:
is fully independent of the particular representation used [
64,
65]. Thus, in order to avoid domain problems, one can consider the Weyl algebra
and look for a unique preferred representation of the relations (
31) and (
32).
From the definition of
and Equation (
28), we can write the Weyl generators in terms of the annihilation and creation operators on
,
By using the commutation relations (
27), the relationship
, and the Baker-Campbell-Hausdorff (BCH) formula, it is not difficult to see that the generators in Equation (
33) satisfy indeed the relations (
31) and (
32). In this way, we can construct the (concrete) Weyl algebra
, which is a subalgebra of
, the
-algebra of all bounded linear operators on
. Let us now consider the vacuum state
, i.e., the unique normalized state
that is annihilated by all the annihilation operators
. By using the BCH formula and Equation (
27), a direct calculation shows that the vacuum expectation value of
in
is given by
Here,
is the norm of
with the real inner product
defined by
[see Equation (
7)].
The relationship (
34) defines a quasi-free algebraic state
. The triple
is, in fact, the same that would be obtained by employing
on the (abstract) Weyl algebra
in the so-called Gelfand-Naimark-Segal (GNS) construction [
66,
67]. The representation
of the Weyl algebra
, defined by the complex structure
J, is moreover irreducible, which is tantamount to saying that the state
is pure. Conversely, pure quasi-free states of the Weyl algebra are associated with complex structures, and give rise to Fock representations as above [
1,
68].
Let us consider now the quantization in the canonical phase space approach. Our choice of a natural set of elementary classical observables on
leads to the real vector space
equipped with the PB (
10). In addition, let us introduce a complex structure
. The abstract algebra
is constructed from the complex vector space
, where
is the Cauchy completion of
with respect to
. The fundamental quantum operators
and
in
satisfy the CCRs:
. Schrödinger-like representations of the CCRs are naturally available in the canonical approach, as follows. The CCRs are represented on a Hilbert space
of wave functionals on a quantum configuration space
, with the basic operators of configuration and momentum,
and
, acting on the wave functionals by multiplication and by derivation plus multiplication, respectively. The measure
, that is of Gaussian type, is determined by the complex structure
j. However, it does not encode the full information about the complex structure, in general. Apart from a derivative, the momentum operator
contains, in general, two multiplicative terms, namely a factor associated with the Gaussian character of
, and possibly another (non-trivial) multiplicative term that contains further information about the complex structure
j. To be more specific, the general form of a complex structure
is given by
, where
, and
are linear operators satisfying
and
for all unit weight scalar densities
and scalars
. Relationships (
35) come from the condition
, whereas restrictions (
36) follow from requiring that
be a symmetric and positive definite bilinear form. The measure and the basic operators of configuration and momentum are [
36,
69]
Note that the representation defined by Equation (
38), in the Hilbert space
, is a representation of the Fock type, i.e., corresponds to a pure quasi-free state of the Weyl algebra.To see this explicitly, let us introduce the Weyl operators
. For an initial reference time
, the map
, where
, naturally induces a bijection between the algebra generated by the objects
and the corresponding one generated by the operators
, that we will call
. Consider now the unit constant functional
. It can again be shown that the expectation values of the Weyl generators read
,where
is the norm associated with the real inner product
on the phase space
. Using the bijection
, one concludes that
indeed defines a pure quasi-free state of the Weyl algebra, associated with the complex structure
. Since a state (in fact the evaluation of a state on the generators) uniquely characterizes a unitary equivalence class of the Weyl algebra, it follows that
and
are just different realizations of the same representation of the Weyl relations, i.e., there is a unitary map
, with
, that intertwines
with
. To make this relationship fully explicit, let us display the form of the annihilation and creation operators on
, that are readily seen to be
where
,
and
is the “positive frequency” part of the Cauchy data
, that is to say
. (For a comprehensive discussion on the Schrödinger representation for a linear scalar field in flat and curved spacetime, including the relationship between the covariant and the canonical approaches to quantization, as well as measure theoretical aspects, see Refs. [
36,
69,
70]).
Representations of the type , determined by a complex structure J in the covariant phase space, will hereafter be called J-Fock representations, whereas the corresponding representations of the form , constructed from the canonical perspective, will be called j-Fock representations.
In the rest of our discussion, the domain of definition of the different quantum observables will not play a relevant role. Hence, in what follows we will consider representations of the CCRs only.
3.3. Bogoliubov Transformations and Unitary Implementability
Let us take two compatible complex structures on the phase space
, say
and
, and assume that their associated inner products,
for
, define equivalent norms on
. Then, the corresponding Hilbert spaces
and
may be identified, and can be viewed as two distinct splittings of the same Hilbert space
H [
1]. Consider also the orthogonal projections
and
defined by the inner product
on
H. Then, let
and
be the restrictions of
and
, respectively, to
. Similarly, let
and
be the respective restrictions of
and
to
. In this setting, it can be shown that [
1]
and
For an element
of
H, let
and
be the components of
with respect to the splitting
of
H, and
and
their components with respect to the splitting
of
H. In short,
and
. We know, in particular, that
. Since the orthogonal projections of
and
onto
are
and
, we get that
is the orthogonal projection of
onto
. A similar calculation shows that the orthogonal projection of
onto
is given by
. So, we have
This transformation, with
C and
D satisfying relationships (
41), is known as a Bogoliubov transformation. Note that, from Equations (
41) and (
42), the inverse of (
43) is
Associated to each of the complex structures
and
, there is a set of elementary variables [see Equation (
25)],
Nonetheless, the vector spaces
and
are, in fact, the same vector space
[recall that
, so that
and
are simply two different decompositions of the linear space
]. Let us be more precise. It follows from Equation (
24) that the linear space
is decomposed by a complex structure
J into the direct sum of
and
. Hence, the complex structures
and
decompose the vector space
as
and
, respectively. Since
, we get that
and
are nothing but two different decompositions of
, as we had commented. The explicit relationship between the annihilation and creation-like variables associated with
and
are
In order to get these identities, we have used the definition of the annihilation and creation-like variables (
23), the action of the complex structures
and
on their corresponding eigenvectors, and the linearity of
, as well as the decomposition of
and
with respect to
, namely
and
. Relationships (
46) give the form in
of the annihilation-like variables
and the creation-like variables
, defined by the complex structure
. It is not difficult to see that the PB between the variables
and
, given in Equation (
46), satisfy indeed Equation (
26).
We emphasize that different complex structures (compatible with
and that give rise to equivalent norms on
S) provide different generators for the (same) abstract quantum algebra
(this is so because different complex structures just introduce different splittings in
, the space from which
is constructed). Let us denote the abstract algebra
by
in the basis provided by the annihilation and creation-like variables defined by the complex structure
; i.e.,
with
, where
and
. The abstract quantum counterparts of
and
are the operators
and
, satisfying
and the CCRs (
27). When
and
are replaced, respectively, with
and
(for
) in Equation (
46) we get expressions for
and
in
. According to the discussion in
Section 3.2, the algebra
(for
) is then represented on the Fock space
(constructed from the one-particle Hilbert space
) by declaring (representing)
and
as the annihilation and creation operators on
, renaming then
and
. Thus, in spite of the
J-independence of
H and
, the representation of the algebra on the Hilbert space turns out to be a decomposition-dependent process: every complex structure (or, equivalently, decomposition) gives rise to a different Fock space representation of
. The annihilation and creation operators on the Fock space
,
and
, are represented on
as
Hence, in general,
does not annihilate the
-vacuum state
. A direct calculation shows that the
-number operator
, represented on
, has the following expectation value in the vacuum state
:
The vacuum state
in the Fock representation
corresponds to a state
satisfying
Actually, provided that
and
define equivalent norms, it can be shown [
1] that the necessary and sufficient condition for the unitary equivalence of the Fock representations
and
is that
B fulfills the Hilbert-Schmidt condition
In that case, there exists a unitary map
such that
with
and
given by relationships (
47), and
solving Equation (
49). We also note that the requirement (
50) on
B is equivalent to impose the Hilbert-Schmidt condition on
. Indeed, since
, it follows that
for all
. Similarly, we have that
for all
. Thus, the two complex structures lead to unitary equivalent representations of the CCRs if and only if
defines a Hilbert-Schmidt operator, either on
or on
.
Let us now discuss the issue of dynamics. Consider a compatible complex structure
J on phase space
. As we have seen in
Section 2.2,
J evolves according to
[see Equation (
14)], where
is the linear symplectic transformation corresponding to the time evolution from
to
t. (Here,
J plays the role of an initial complex structure. Accordingly, all objects defined by
J, such as the annihilation and creation operators or the associated Hilbert space, will be labelled with a subscript, or superscript,
.) Thus, every
belongs to
and, consequently, we get a family of real inner products,
, on
S. Assume that, for each time
t, the linear symplectic bijections
and
are both continuous mappings on
(the Hilbert completion of
S with respect to the norm
defined by the inner product
). Then,
and
define equivalent norms for all
. (Let
be the Cauchy completion of
S with respect to
. Suppose that the linear symplectomorphism
and (its inverse)
are continuous. Then,
R and
are bounded in the norm
. Hence, using that
, it follows that
and
, with
, define equivalent norms.) The annihilation and creation operators induced by time evolution
on
, namely
and
, are given by Bogoliubov transformations of the form (
47),
for each
. Here, both
and
are in
, whereas the orthogonal projections (with respect to the
-decomposition)
and
satisfy relationships (
40). So,
and
fulfill the CCRs (
27). Clearly,
and
are the identity and the zero maps, respectively.
Since classical observables evolve according to , we have that and . On the other hand, the symplectic transformations induce a two-parameter family of ∗-automorphisms on , . Thus, the time evolution of the (abstract) elementary quantum observables is given by . In particular, we have that and , on .
The question of unitary implementability of the dynamics in the
J-Fock representation is whether or not there exist unitary operators
such that
If such operators exist, it also means that, within the Heisenberg picture, the evolution expressed by Equations (
52) is unitary, i.e.,
It follows from the discussion above that the unitary operators
exist, i.e., the classical dynamics dictated by
is unitarily implementable in the
J-Fock representation, if and only if
satisfies the Hilbert-Schmidt condition (
50) for all
t. This is tantamount to requiring that
be Hilbert-Schmidt on
for all
t, as we have seen. Another way to formulate this condition is to say that the antilinear part of
must be Hilbert-Schmidt on
for all
t (indeed, a symplectic transformation
R is unitarily implementable on a Fock space
if and only if its antilinear part with respect to the complex structure
J, namely
, is Hilbert-Schmidt on the one-particle space
defined by
J [
71,
72]).
Turning to the more algebraic perspective, the ∗-automorphisms
of the algebra
define ∗-automorphisms
of the Weyl algebra
via
, i.e.,
. A simple calculation shows that
; that is to say, the time evolution of observables in the Heisenberg picture is represented by the inverse of the automorphisms
, related to the inverse of
, of course (for details about symplectic transformations and automorphisms in the Weyl algebra see, for instance, Ref. [
73]). Again, the family of automorphisms
of the abstract Weyl algebra corresponds to unitary transformations in the
J-Fock representation if and only if
is Hilbert-Schmidt for all
t. Note also that the relation
between the algebraic states can be interpreted as the time evolution of the “initial” algebraic state
. So, in the Schrödinger picture, the issue of a unitary quantum dynamics becomes the question of whether or not the family of algebraic states
provide unitary equivalent representations of the (adjoint and) Weyl relations (
31) and (
32) [or, equivalently, of the CCRs (
29)].
Let us now focus in particular on the case of a free scalar field propagating in a spatially compact spacetime. Because of spatial compactness, every
and
can be written as
Here,
and
are orthonormal bases with respect to the Hermitian inner product
for, respectively,
and
, whereas
and
are complex constant numbers. Clearly,
is an orthonormal basis for the
-decomposition of
H, namely
. As before, let us denote by
and
the restriction of the Hermitian inner product
to, respectively,
and
. The projection of
onto
gives the vector
. Similarly, the projection of
onto
, gives the vector
. Hence, we have that
with
By performing an analogous calculation, one gets
and
[in fact, we can obtain them by simply switching the
t and
parameters in Equations (
57) and (
58)]. Since
and
[see Equation (
42)], we thus get that
where we have used
The first relation in Equation (
40), together with Equations (
57) and (
59), implies that the Bogoliubov coefficients
and
satisfy
From
and Equation (
57), it follows that the bases
and
of
H are related by
where
and
are given by Equation (
58), and satisfy the relation (
61). Equations (
62) and (
63) are the Bogoliubov transformations between basis vectors.
Let us consider the expansion of the field in terms of the basis modes
associated with the initial complex structure
J,
It is straightforward to check that the annihilation and creation-like observables
and
evaluated at
give
and
. That is,
and
can be viewed as coordinate functions on
S (which can be identified with the space of coefficients
), so that we can write
and
. From Equation (
27) we get that
and
satisfy the standard CCRs
. The Fock space of quantum states
is generated by repeatedly applying the creation operators
on
, the state annihilated by all
. Employing Equation (
54) on the basis modes
(i.e.,
) and using Equation (
57), we get that —if it turns out to be unitary—the “evolution” of
from
t to
would be given by
The time evolution from the initial time
to an arbitrary final time
t is obtained simply by interchanging
t with
in the above equation. Note that
and that
. Thus, the evolution of the annihilation operator
associated with
, from
to
t, would be given by
Analogously, we obtain that the evolution of the creation operator
, from
to
t, would be dictated by the unitary transformation
A direct calculation shows that the unitarity condition, i.e., the Hilbert-Schimdt condition on
, turns out to be the requirement that the Bogoliubov coefficients
be square summable,
It is worth remarking that unitarity (or not) of is a basis-independent issue. Indeed, given any other orthonormal basis in , it is not difficult to see that is equal to , i.e., the result of does not depend on the specific choice of basis considered to perform the calculation.
By using the isomorphism
between the linear spaces
S and
, one can obtain the counterpart of the above quantization in the canonical approach. The configuration and momentum of the field
, expanded in the positive and negative frequency mode solutions associated with
J [see Equation (
64)], are given by
where
and
. In terms of the complex structure induced on
, i.e.,
, the annihilation and creation-like variables read
and
, respectively. The promotion of these variables to quantum operators corresponds to the annihilation and creation operators (
39) in the Schrödinger representation, with label
. The time evolved operators of annihilation and creation,
and
, are respectively given by the right-hand side of Equations (
66) and (
67), that define the mapping
in the current representation.
Let us conclude with the following remark concerning unitarity. It follows from Equations (
51), (
52), (
54) and (
55) that if
U is a unitary map, then so is
(and vice versa). For unitary
U, we have in particular that
Let us now suppose, without any further assumptions, that Equation (
70) is satisfied. Then, a calculation along the lines of Ref. [
74] shows that
Let
be the composition
. Thus, we obtain from Equation (
71) that
By applying Equation (
72) to the vacuum state
of
, we get that the state
in
must satisfy the relationship
Therefore, the maps are unitary mappings if and only if is Hilbert-Schimdt on . However, since is precisely the complex structure resulting from evolving J in time, we have that the Hilbert-Schimdt condition is trivially satisfied and, therefore, is always a unitary map for all and t. Note, nonetheless, that unitarity of does not imply that U (nor ) must be necessarily unitary.
Let us consider the above condition (
73) with
replaced with some
. Then, unitarity of
means that complex structures
differing from
can be consistently considered at time
t only if
is a Hilbert-Schmidt operator. More specifically, from the unitarity of
, it follows that
, where
and
is a normalizable state satisfying Equation (
73). Since the expectation value of
at final time is given by
(see for instance Ref. [
73]), we have that
, which certainly holds only if
is Hilbert-Schmidt either on
or on
. (The consistency condition that
be Hilbert-Schmidt was introduced and considered in Refs. [
74,
75], within the canonical space approach, as a general condition of unitary evolution.) For a thorough discussion on quantum unitary dynamics in cosmological scenarios see Ref. [
60].
3.4. The Scalar Field with Time Dependent Mass
As we pointed out in
Section 2.3, the 0-spin boson field
propagating in a spatially compact FLRW spacetime can be treated, after the time dependent scaling
, as a free scalar field with time dependent mass (or, equivalently, as a scalar field subject to a time dependent potential) propagating in a static background, obeying the equation of motion (
20). Here, we will consider the same class of system, but adding also the case of a background with one-dimensional spatial sections with the topology of a circle. Besides, the time dependent function
in the potential
will be considered (except for very mild conditions that will be specified below) as a general real function. Let us remark that for non-negative
, the function can be interpreted as a squared time dependent mass.
More concretely, we consider here a real scalar field
governed by the equation
in a static background
where
is the standard Riemannian metric of a spatial manifold
that we will allow to be either a circle
, a three-sphere
, or a three-dimensional torus
. Besides,
is the LB operator associated to
. According to our general discussion in
Section 2.2, the canonical phase space is the real linear space
equipped with the standard symplectic structure (
5). (With respect to Equation (
21), we now rename
and
to simplify our notation). The covariant phase space is the linear space
S of smooth solutions to Equation (
74) arising from initial data on
,
and
, equipped with the symplectic structure (
11) (with the identification
,
, or
). We recall that
stands for the fixed (but) arbitrary initial reference time. The PB between the canonically conjugate variables of configuration and momentum are given by
, where
x denotes abstractly the coordinates of a point on
.
Scalar functions on
can be expanded in terms of harmonics, i.e., in terms of solutions of the eigenvalue equation for the LB operator on
:
, where (1)
for
, with
, (2)
for
, with
, and (3)
for
, with
and
(
). The eigenfunctions
can be chosen as the complex exponential functions
and
for the
and the
cases [
denotes the integer
n and the triple
, respectively], whereas for the
case
stands for the (hyper)spherical harmonics
on
[here
denotes collectively the set of indices
, with
,
, and
]. (For a description of the harmonics in non-vanishing spatial curvature see, for instance, Ref. [
76].) The functions
are orthonormal with respect to the
-product on
, namely
, where
. The configuration and momentum of the field can be expressed as
where
and
are the complex Fourier coefficients of the expansion in the complete set
. Since the field is a real one, these Fourier coefficients satisfy the following reality conditions:
for the circle case,
for the three-torus case, and
for the three-sphere case, where
. The reality conditions are obtained by using that
and
, that the configuration
and the momentum
of the field are real functions, and by employing the specific relationship between
and its complex conjugate
:
on
,
on
, and
on
.
Incorporating the time dependence in our field, and recalling that it is real, we can decompose it in a Fourier expansion of the form
where the functions of time
are solutions to the second-order differential equations
This equation follows from the field Equation (
74) when the spatial part is evaluated in the harmonic
. We note that the Equation (
78) is real. Therefore if
provides a solution, so does its complex conjugate
. The relation between the functions
and the coefficients
above depend on the complex conjugation properties of the eigenfunctions
of the LB operator. For instance, in the
case we get that the Fourier coefficients of the configuration field are given by
, where
is the initial time. On the other hand, it is worth remarking that the dynamical Equation (
78) depends exclusively on the eigenvalue of the LB operator,
, rather than on the label of the harmonic,
. As a consequence, except for the dependence on
that the initial conditions determined by
may impose at
, the functions
vary only with the value of
. Indicating the dependence on this eigenvalue with a subscript
n, we can then rewrite the field (
77) in the following manner:
This field decomposition respects the symmetries of the field equations. Here,
is a set of arbitrary complex constants, and the functions
are conveniently normalized solutions to Equation (
78) (as we explain below). The subscript
n can be chosen to correspond to the absolute value of the harmonic label
n for the case of the circle, to the Euclidean norm of
for the three-torus, and to the first index in the set
for the three-sphere.
Most important for the quantization it is the fact that, given that Equation (
78) is real and of second-order, as we have commented, we can choose the complex solution
so that
is an independent solution. In this way, we obtain a splitting of the space of solutions between “positive and negative” frequency modes, namely
and
. According to the discussion in
Section 3.2, there is an associated complex structure
J, with corresponding annihilation-like variables given by
and creation-like variables provided by their complex conjugates. From the orthonormality of the field solutions
with respect to the Hermitian inner product
, and of the eigenfunctions
with respect to the
-product on
, it follows that
In this perspective,
J is ultimately defined by the functions
, and thus the choice of a complex structure is equivalent to the choice of a set of complex solutions
to Equation (
78) for every of the LB eigenspaces, satisfying Equation (
80) (see Ref. [
77] for details). The field decomposition (
79) is fully adapted to this perspective, and immediately gives an expression for the field operator in the Heisenberg picture, when the constants
and
are replaced with annihilation and creation operators, acting on the Hilbert space constructed from
J, as described in
Section 3.2.
Making contact with the canonical perspective, and since the solutions
are determined by the initial conditions, we have that, in terms of the Cauchy data at the initial reference time
, the annihilation-like variables are given by
where
is the initial complex structure on
induced by
J, namely
.
Clearly, the group of spatial symmetries of the metric
, say
, is a group of symmetries of
and, consequently, of the equation of motion (
74). In the same spirit of demanding invariance under such symmetries that we adopted above, we note that a
-invariant complex structure does not only allow for a unitary implementation of the spatial isometries corresponding to
, but furthermore for a
-invariant representation of the CCRs. A simple obvious choice, that ensures a
-invariant Fock representation, is the massless free field representation provided by the complex structure (
22),
This complex structure defines the annihilation-like variables
. (We exclude in principle the zero mode. This does not affect the field properties of the system. Besides, the zero mode can be quantized separately as a mechanical system.) Notice that
is determined by the initial conditions
and
for Equation (
78) (these can be checked to provide valid initial conditions; see, for instance, Refs. [
12,
13]). Indeed, substituting the complex structure (
82) into Equation (
81) we get
where we have used that
, as well as the orthonormality of the eigenfunctions
with respect to the
-product on
. Hence, for
and
, the annihilation-like variables
reproduce in fact the massless free annihilation-like variables
.
By constructing the
-Fock representation, we get the annihilation and creation operators
and
defined by
[see Equation (
39)], where
and
. Then, introducing a Fourier decomposition, we obtain
and
, that are nothing but the result of promoting the observables (
81) and their complex conjugates [with
,
, and
] to quantum operators. Explicitly,
and
are given by
where the action of
and
on the Hilbert space is obtained from the Fourier decomposition of Equations (
37) and (
38), for the complex structure characterized by
,
, and
.
The time evolution of
is dictated by a Bogoliubov transformation of the form (
66), namely
with Bogoliubov coefficients
where
corresponds to
for the circle case,
for the three-torus, and
for the three-sphere. The coefficients
and
are given by
A straightforward calculation shows that .
Notice that, instead of considering a Fourier decomposition with respect to the set of complex functions
, one can decide to perform the expansion of the configuration and momentum of the field in terms of explicitly real functions. In that case, the Fourier coefficients become real as well, and no reality conditions need be imposed. Then, the corresponding Bogoliubov coefficients turn out to be of the form
For instance, in the
case, the non-zero modes of the system can be described in terms of real canonically conjugate variables
related to the complex variables
by
,
,
, and
, restricting now
n to be a positive integer,
. Since
, for
, the operators associated with
and
are respectively given by the self-adjoint operators
and
. We use this canonical transformation to recast the Schrödinger representation, with fundamental operators
and
, in terms of the self-adjoint operators
,
,
, and
. The
-annihilation operators are given by
and the creation operators are provided by their adjoints,
and
.
According to Equations (
85) and (
86), the time evolution of
(for all
and with
) is given by the Bogoliubov transformation
The expression for
is obtained by taking the adjoint of Equation (
90). It is not difficult to see that
and
are related to the annihilation and creation operators
and
by
for all positive integers
m (for negative
m,
and
can be found from the adjoint of the above relations). The time evolution of
and
can be determined by substituting Equation (
91) into Equation (
90),
Thus, in contrast with the expression (
90), where the modes
m and
are coupled, the evolution of the annihilation operators
and
is fully decoupled from the rest. Although the Bogoliubov coefficients
are the same ones as for
, the coefficients of the antilinear part are now given by
The -Fock representation is, by construction, invariant under the isometries of the spatial manifold (, , or , depending on the case). This property, however, turns out not to be enough to guarantee the uniqueness of the representation. Indeed, there are infinitely many complex structures which do not belong to the equivalence class of but are symmetry invariant. Thus, one has to look for extra requirements in order to select a unique preferred Fock representation. A natural requirement is to demand that the classical symplectic transformations associated with the time evolution are properly quantized as unitary operators (note that it is pointless to ask for time invariance, since time-translation symmetry is broken by the non-stationarity of the system). So, we restrict our attention to invariant Fock representations that admit, in addition, a unitary implementation of the dynamics.
In summary, we require that (1) the vacuum state be invariant under the (spatial) isometries of the manifold
, and that (2) the dynamics dictated by the field Equation (
74) be unitarily implementable. Remarkably, the
-Fock representation is the unique (up to unitary equivalence) symmetry invariant representation of the CCRs where a unitary implementation of the time evolution is available (i.e., it is the unique Fock representation satisfying the criteria of invariance and of unitarity). Furthermore, no canonical transformations (except for trivial ones) can lead to a field description from which an invariant Fock representation admitting a unitary implementation of the dynamics could be defined; i.e., the
-description is unique, up to trivial canonical transformations. The removal of the ambiguities in the quantization of scalar fields with time dependent mass is discussed in Ref. [
12] for the case of the circle topology, in Refs. [
13,
14,
15] for the case of the three-sphere topology, and in Refs. [
43,
44,
45] for the case of the three-torus topology. In all of these cases, it is sufficient (but not necessary) that the function
possesses a second derivative which is integrable in every compact subinterval of the time domain.
The rest of this work is an overview of these uniqueness results obtained within the context of cosmology; the arena in which the studies were motivated and developed. We will present a compilation of the uniqueness results attained for the quantization of Gowdy models, and of (test) scalar fields propagating in FLRW spacetimes, de Sitter spacetimes, and anisotropic Bianchi I universes.