2.1. Euler Rescaling as Canonical Transformations on the Extended Phase Space
In order to deal with a constrained system with a fractional power of a Hamiltonian, we work in the extended phase space. As discussed in [
23], coherent states for a constrained system will in general have some restriction on their label in order to ensure that their labels are consistent with the constraints of the system. As mentioned above, we restrict our discussion to fractional powers of the harmonic oscillator Hamiltonian here. However, the strategy can be carried over to more complicated systems, if the constraint associated with temporal diffeomorphisms (for general relativity, this is the Hamiltonian constraint) can be written in a deparametrised
1 form at the classical level, and the set of coherent states that one wants to use for the computations has good semiclassical properties as far as integer powers of the Hamiltonian are considered. We discuss this aspect in more detail in our conclusions in
Section 4. To explain how the Euler rescaling can be useful in this context, let us consider the following set up: we examine a Hamiltonian that is given by some fractional power of the harmonic oscillator in one dimension formulated on the phase space
with elementary variables
. We denote the Hamiltonian as
where
is a rational number
with
and
is the Hamiltonian of the harmonic oscillator that is given by
. In order to map this dynamical system into a constrained system with a deparametrised constraint, we work in the extended phase space
in which the temporal coordinate is also treated as a canonical variable with coordinates
. The constraint of the system in the extended phase space has the form
where
k is some arbitrary real and non-zero number and all remaining Poisson brackets vanish. Let us briefly comment on the units of the involved quantities. From the constraint
C, we find that
. Furthermore, we find that
and
. Here, deparametrisation means that the constraint can be written linearly in the temporal momentum and the remaining part of the constraint does not include
t. In the extended phase space, as shown for instance in [
39], we can write down a set of first-order Hamilton’s equations with respect to an evolution parameter that we denote by
s
Now, we are interested, in a constrained system, in the dynamics of the observables, which are phase space functions in the so-called reduced phase space. The reduced phase space can be obtained by a symplectic reduction with respect to
C and can be coordinated by the corresponding elementary observables associated with
. These observables are quantities that are required to commute with the constraint
C. From now on, let us consider the choice
. In this case, the physical Hamiltonian, which generates the evolution of the observables, is then given by the function
evaluated at the observables of
q and
p. Let us denote the observables of
by
and
, then the classical Hamilton’s equation in the reduced phase space reads as follows:
where we denote the evolution parameter in the reduced phase space by
to match with our later notation at the end of
Section 2. We realise that, for the choice
and under the identification
,
and
, the Hamilton’s equations in (
1) and (
2) agree for this subset of variables. In this sense, we can cast any classical Hamiltonian system with a given Hamiltonian
H into a constrained system with constraint
in the extended phase space that is written linearly in the temporal momentum. Looking at the equations of motion in (
1) with
, we realise that the outer derivative of
involves the first-order equations for
, and likewise,
can be absorbed into a redefinition of the temporal coordinate and with respect to the transformed time the Hamiltonian is simply linear in the harmonic oscillator Hamiltonian
. In the extended phase space, this can be formulated as a canonical transformation of the form
The variables
have the units
and
. This transformation—often denoted as Euler rescaling—was discussed in a more general context for instance in [
39]. Note that in our case, this is rather a kind of dual Euler rescaling, since here the new temporal momentum
is a function of
only, whereas the new temporal coordinate
T is a function of
. In contrast to the Euler rescaling in [
39], the new temporal variable
T is a function of
t only and
a function of
. Furthermore, in [
39], the transformation to
t involves an integral. Assuming that
, we can multiply the entire constraint
C by
and obtain
where we used the weak ≈ equivalence of quantities on the constraint surface
and
since
. In this sense, the new constraint
implies
on the constraint surface, which requires
and thus
, leading to
. An important property of the above-defined canonical transformation is that
, and thus also any function of it, is a constant of motion which on the reduced phase space can be identified with the energy of the physical system. As a consequence, when we use the rewritten and equivalent version of the constraint in (
4) in the next subsection to construct coherent states in constrained systems, we have to take this into account and consider that not
is the energy of our original system that we start from but
and thus
, where
denotes the energy of the system and can be determined once the phase space variables are given. Keeping track of the original definition of the energy of the system before the dual Euler rescaling has been applied goes in the same direction as the idea of a kind of reference metric suggested by Klauder in [
46] in order to be able to have a consistent interpretation of the dynamical operator even if a transformation of the phase space variables has been applied.
If we are to work with the constraint
in (
4), then it will seem that we have not gained much, since we have simply moved the fractional power from the Hamiltonian to the momentum
. However, as shown in [
35], using Kummer functions, fractional powers of the momentum operator can be well approximated by the standard harmonic oscillator coherent states, and we use those results here to obtain appropriate coherent states on the kinematical Hilbert space which approximate the quantum constraint well semiclassically.
Now, given the constraint in the desired form, we can proceed in two directions. Either we consider Dirac quantisation and solve the constraint in the quantum theory, or we derive the reduced phase at the classical level and apply reduced phase space quantisation. At this stage, both are equally justified. In the context of coherent states, this carries over to the situation that when applying Dirac quantisation, those coherent states are usually constructed on the kinematical Hilbert space. However, in order to actually compute relevant semiclassical expectation values, one would like to use physical coherent states that encode some information about the constraints in the system. As mentioned at the beginning of this section, a strategy to obtain physical coherent states from a given set of kinematical coherent states was presented in [
36] and applied to a couple of examples there. We follow this strategy in the next subsection and apply it to fractional Hamiltonians combined with the Euler rescaling discussed above, where this technique is still based on using the standard harmonic oscillator coherent states. At the end of the next subsection, we also show that, in this case, reduced phase space quantisation and Dirac quantisation will yield the same set of physical coherent states.
2.2. Physical Coherent States for Constraints with Fractional Hamiltonians
We want to apply the techniques introduced in [
36] to our fractional powers
of the harmonic oscillator Hamiltonian, which we shortly refer to as fractional Hamiltonians. Instead of considering the fractional Hamiltonians directly, we deal with the extended phase space as described in
Section 2.1 and consider a constraint of the form
where
is the canonical conjugate momentum to a new time variable
t and a constant of motion with respect to the fractional Hamiltonian in consideration. For fixed phase space coordinates
, the temporal momentum
corresponds to the negative energy of the system; that is,
with
. Notice that for
, this reduces to the case for the harmonic oscillator. Because of the general fractional power
of the harmonic oscillator Hamiltonian, the constraint in general might be difficult to handle. Therefore, we transform the constraint using the dual Euler rescaling to obtain an equivalent constraint as displayed in (
4) in
Section 2.1, which reads
for
. The kinematical Hilbert space of this model is
, where we use for both Hilbert spaces the standard Schrödinger representation; i.e., for the first one, the occupation number representation, and for the second one, the momentum representation. The kinematical inner product for two kinematical states
and
has the following form:
The constraint operator is then simply given by
As a first step, we define kinematical coherent states whose expectation value of
reproduces to the lowest order in
ℏ the classical constraint. These kinematical coherent states can be obtained from a tensor product of the standard harmonic oscillator coherent states as follows:
where
and
are classical labels associated with the extended phase space. The explicit form of these states is given by
and
with
carrying units
such that the arguments of all exponentials are dimensionless. If we define a similar dimensionless label
also for the temporal phase space coordinates, then
will enter as
. The coherent state
is already normalised, and we choose
such that
is also normalised and thus
as well. The semiclassical expectation value of the constraint operator
can be computed as
Using the techniques presented in [
35], we can express the second semiclassical expectation value in terms of Kummer functions and obtain
where
with
denotes the Kummer function of the first kind, also called the confluent hypergeometric function of the first kind. For more details on Kummer functions and particularly on how their Fourier transform can be used to obtain the above semiclassical expectation value, we refer the reader to the work in [
35]. As far as the semiclassical computations are concerned, we are interested the sector in which
ℏ is small compared to one, which allows us to express the semiclassical expectation value as an expansion in (fractional) powers of
ℏ. The classical limit can then be obtained in the limit where we send
. Consequently, in the case of the Kummer function, we can use its asymptotic behaviour for large arguments, which is well known. As shown in [
35], the relevant asymptotic expansion for the semiclassical expectation value is given by
where
denotes the Pochhammer symbols, also called raising factorials with
and
. Given these asymptotics of the Kummer function, we obtain for the semiclassical expectation value of
where we use
. Hence, in the semiclassical limit
, we recover the classical constraint
The point that we obtain in the limit
the correct classical expression confirms the theorem in [
12] based on the Hamburger momentum problem by explicit computations in our toy model
2. Due to use of the techniques introduced in [
35], we can also explicitly compute the higher than leading order terms. In this sense, our results extend those in [
12] concerning the formalism for non-polynomial operators. Note that for the special case that
with
, we have
, and then the first argument of the Kummer function is
, and in this case it can be expressed in terms of Hermite polynomials yielding for instance the expected semiclassical expectation value for
for the choice of
. The rather unusual powers of
ℏ involving
are due to the fact that, in our case, the unit of
is
, whereas for the spatial coordinates, one uses the characteristic length of the harmonic oscillator
to introduce dimensionless quantities and
is linear in
ℏ. For odd integers, we find that
can also become negative, but then even at the classical level, due to the fact that
, the constraint
has no solutions, and that is why we work with
here. Note that this is similar to the situation for the reference matter models, where one usually also restricts this to certain parts of the full phase space by restricting the sign of the clock momentum; see for instance the discussion in [
25,
26,
28,
35,
47,
48,
49].
The discussion so far was completely at the kinematical level; therefore, we apply the group averaging procedure to obtain physical coherent states along the lines of [
36]. In our case, the group averaging operator is given by
where
commutes with
,
is rewritten in terms of the number operator
, and the constraint is rescaled by
in order to obtain a dimensionless quantity. Now, we calculate the action of the unitary operator involved in the group averaging
on the kinematical coherent state
leading to
where
denotes, as before, the standard coherent state in the momentum representation. Next, we apply the group averaging to obtain physical coherent states, which in our case will not be elements of
but a distribution on a dense subset
, following closely the formalism in [
36]. In addition, we introduce a projection operator
which projects on the negative part of the spectrum of
to ensure that the classical condition
which requires
is also fulfilled at the quantum level. This projection operator can be implemented via
, where
denotes the usual Heaviside function that vanishes if
. Then, we obtain the physically constrained coherent states as follows:
where we interchange the order of the integration over
with the summation and integration over
, use the definition of the Fourier transform of the delta function, and define
in the last step. Here,
K is a real constant whose value can be chosen such that the resulting physical coherent states are normalised as done in (
21) below. Because the spectrum of
is the entire real line, we find, even if we project to its negative part, that
, and thus one obtains a non-trivial distribution after group averaging. Similar to the example of the linear constraint in [
36], where also a distributional physical coherent state is obtained, the result of the group averaging can be understood as the restriction of the kinematical coherent state to the constraint surface with an additional modification in the measure. The physical inner product can be explicitly computed and reads
The norm of
then becomes
where for the absolute value we have
, and in the last line we define
. That the norm is finite is ensured by the fact that
is already converging, and due to the absolute value of the Gaussian evaluated at
for large values of
n, the sum involved in the norm is even more strongly decreasing. We can obtain normalised physical coherent states by choosing
and using the states
If we compute the expectation value of the constraint
with respect to the non-normalised physical coherent states
setting
, we obtain
Next, we compute the expectation value of the Dirac observable
in the physical coherent states, and we obtain
As in [
36,
38], we assume that the coherent states are peaked on the constraint surface. If we further use that
, where
denotes the energy of the original system we started with and the 0-label is introduced because the energy is determined by
, then we obtain
, yielding
Although the zero point energy comes out exactly, in the case of the expectation value related to the classical energy
, this is not similar. Here, the corresponding contribution involves two sums: one from the norm and a second one from the expectation value, where the latter involves the coefficient
with an index shifted by one. This carries over to a shift in
and to the absolute value of the Gaussian; thus, we obtain
with
The sum of the squared norm in the denominator involves the same expression but with
n and not
in
. The label
is the classical label of the coherent states associated with the temporal momentum. As discussed, on the classical constraint surface, we can identify
with a classical energy equal to
; that is, the
-th fractional power of the energy of the classical harmonic oscillator. Due to the
ℏ in the denominator in the Gaussian, it is narrowly peaked around the value of the classical energy
. Hence, the peak of the Gaussian with its fractional argument will be located at
. Because the classical energy is assumed to be large compared to the eigenvalues
, a reasonable choice for
is a value that corresponds to large
n in
. Consequently, the peak and hence the main contribution of this Gaussian with fractional argument will be at large values for
n. Furthermore, the
in each summand has the additional effect that the summands are further decreasing strongly with increasing
n. Therefore, in the sum in the numerator, we can replace
by
, and the corrections due to this replacement are very small. The shift in
involved in
will be of minor order compared to the effects coming from the Gaussian and inverse factorial. If the absolute value of the Gaussian were absent, then to find a justification for why large values of
n will be most dominant would be difficult. Thus, we realise that this is a specific feature of the physical coherent states. Assuming that we choose reasonable values for the classical energy that are sufficiently large compared to the energy eigenvalues of the harmonic oscillator Hamiltonian, we obtain
Note that a similar strategy was considered in [
45]. There, the physical inner product still involves integrals, and therefore variables were introduced that encode the deviation from the value around which the Gaussians in the coherent states are peaked. The resulting semiclassical expectation values were then written as an expansion consisting of a classical momentum variable and the width of the Gaussian. Despite the fact that this seems to be more elaborate than in our case in the sense that they also include corrections around the classical value, the techniques they use cannot directly be carried over to our case, since we have no integrals involving Gaussians for the expectation value with respect to the remaining physical states. Furthermore, these corrections arise because functions in the integrand are Taylor expanded. More close to our case is the work done in [
36], where, among others, semiclassical expectation values of generic observables being quadratic in annihilation and creation operators for constraints involving the number operator were discussed, leading to a similar situation as in our case with two sums: one from the norm in the denominator and the second one from the expectation values in the numerator. They considered the asymptotic values for these observables and assumed that one of the classical labels
is very large and tends to infinity. Given this, they could show that these two sums will drastically simplify if the dominant contributions are considered, yielding the correct classical values of the quantum observables under these assumptions.
Considering now the result in (
24), we can solve this for the classical energy, leading to an expression that involves fractional powers of the semiclassical expectation value of the harmonic oscillator Hamiltonian:
We realise that in the limit
, the
-th power of the expectation value of
with respect to the normalised physical coherent states
agrees with the classical energy
. If we had worked with the normal ordered Hamiltonian
, as for instance done in [
36], the
ℏ corrections due to the zero point energy would have even been absent. The reason why shifting the fractional power from the Hamiltonian to the temporal momentum works here is that the fractional power is reintroduced in the final result by requiring that, for physically coherent states, their labels are peaked on the constraint surface, which is a physically reasonable assumption and in this sense carries the fractional power of the operators over to the classical labels of the coherent states, where they can be handled in a simpler manner. Let us compare the situation at the physical and kinematical levels in this aspect. For this purpose, we consider the Dirac observable
, which at the physical level coincides with the harmonic oscillator Hamiltonian. In the two cases, we obtain for the semiclassical expectation values
for large
n and
Thus, even if we assume that the coherent state is peaked on the classical constraint surface, where we find that
and consider the expansion of the Kummer function for large arguments shown in (
14), we can observe that we obtain for the kinematical expectation values
ℏ corrections to
that are not caused by the zero point energy of the harmonic oscillator but due to the in general fractional power associated with the temporal momentum. The underlying reason for this is that in the case of the physical coherent states due to the involved delta function, the inner product is modified, and hence, the generally fractional powers of
need no longer be integrated against the Gaussian of the coherent state, leading exactly to Kummer’s function, as stated above in the kinematical case. In this sense, the coherent states intrinsically encode some dynamical properties via their labels and are beside the group averaging adapted to the constraint under consideration. Note that using coherent states that are peaked on the constraint surface was also crucial in [
36] in order to obtain good semiclassical results for the operators corresponding to the classical Dirac observables. In our example discussed so far, the coherent states are perfectly adapted to the Hamiltonian
. As a consequence the relation between the classical energy
and the semiclassical expectation value in (
24) and (
25) is very simple. For more complicated Hamiltonians, one obtains a more complicated function of the coherent state labels
, in which one then also replaces
by
. However, in order for the semiclassical states to be reasonable in lowest order in
ℏ, we expect to obtain
, plus possible further additional terms which then come with higher orders in
ℏ and can be interpreted as small corrections to the classical value. In order to test whether the semiclassical limit is correct, which corresponds to the limit
, the method here can be useful, but working with the possible corrections involved could become problematic because the final step involves solving for
, which requires that the inverse function of the right-hand side of (
25) exists. If we want to encode the fact that the coherent states are peaked on the constraint surface directly into their labels, we can achieve this by implementing the corresponding restriction on the labels
. In our case, we have
. Hence, we can label the coherent states with
. Then, following the computations conducted above, we also obtain the results in (
24) and (
25). Although the states are adapted to the fractional power
of the Hamiltonian by construction, the label involves the inverse power
, which requires us to solve for
. In the next section, we present a construction of coherent states for which such an inverse power is no longer present.
One could ask how the situations regarding the labels and the forms of the states might change if we apply reduced phase space quantisation instead of Dirac quantisation. As pointed out in [
36], the physical inner product can often be identified with the inner product on the reduced phase space, and we discuss the situation for this model here. If we perform a reduced phase space quantisation, we can identify the phase space variable
t with our clock. Since the constraint
is in a deparametrised form, we can construct Dirac observables for
by choosing a gauge fixing condition
and using the power series expansion introduced in [
50,
51,
52]. In this simple model, the power series can be written in closed form, and we obtain the following for the Dirac observables:
where
denotes the iterated Poisson bracket with
and
with
Reinserting this back into the observables in (
26), the closed form of these observables is given by
The algebra of these observables satisfies the standard canonical Poisson algebra; that is,
, and all remaining ones vanish. Given this explicit form of the observables, we can explicitly show that the physical
indeed generates their evolution. We have
and
The physical Hamiltonian is
and can be quantised using the standard Schrödinger representation, and hence the reduced phase space is simply
, where
acts by multiplication and
as a derivative operator. The quantisation of the Hamiltonian allows us to formulate the corresponding Heisenberg equations for
and
with the Hamiltonian operator
. Considering the Schrödinger picture, one obtains a standard Schrödinger-like equation with
as the involved Hamiltonian operator. Physical coherent states on the reduced physical Hilbert space can be constructed as
With respect to the inner product of
, these physical coherent states are normalised, as one can easily see. The physical coherent states obtained via group averaging can be isometrically embedded into
using the map
where we assume, as mentioned above, that the constant
was chosen such that the coherent states
in
were normalised. Using these rescaled states
in the reduced inner product yields the same result as that for the physical coherent states in the physical inner product. Because in the reduced phase space any function involving the variables
can be expressed as a function of
, only the expectation values for Dirac observables with respect to physical coherent states using group averaging and reduced phase space quantisation agree under the identification
.