1. Introduction
Various theories of quantum gravity (such as string theory) predict the existence of a minimum measurable length scale, usually on the order of the Planck length [
1,
2,
3,
4]. Black hole physics also suggests the existence of a fundamental limit with which we can measure distances [
5,
6,
7]. This is because the energy needed to probe spacetime below the Planck length scale exceeds the energy needed to produce a black hole in that region of spacetime. On the other hand, Doubly Special Relativity theories predict maximum observable momenta [
8,
9]. All of these theories give rise to the modified commutation relations between position coordinates and momenta, which in turn give rise to the Generalized Uncertainty Principle.
Motivated by these results, we examine a quantum theory exhibiting a minimum measurable time scale. Our starting point is the Page–Wootters formalism of time evolution in quantum mechanics. It is based on the idea that time evolution arises as a result of correlations between the “clock” and the rest of the system. Next, we introduce the minimal time scale into the theory by modifying the commutation relations between the time operator and the operator conjugate to it (the frequency operator ). Such a modification causes the time observable to not be represented by a self-adjoint operator but only by a symmetric operator. This is expected as it is no longer possible to consider states of the system at a particular instance of time. As a result we use states of the “clock” system maximally localized around instances of time to construct continuous time representation of the system and we arrive at a modified Schrödinger equation describing time evolution of the system. A minimal time scale also allows us to introduce a discrete time representation of the system and a corresponding discrete Schrödinger equation describing time evolution on a lattice. Interestingly, the received discrete and continuous time representations are equivalent with each other and describe the same time evolution of the system.
According to the received Schrödinger Equations (
56) and (
57), time evolution is governed by an effective Hamiltonian different than the Hamiltonian of the system. The reason for this is that in the case without a minimal time scale, the isolated system possesses time-translation symmetry. The Hamiltonian is the generator of this symmetry. However, when there exists a minimum measurable time scale, the time-translation symmetry breaks down and the Hamiltonian no longer generates translations in time.
Other approaches to time evolution with a minimal time scale in quantum mechanics can be found in the literature. Worth noting is the paper [
10], where the authors deform the Heisenberg algebra of spacetime variables receiving a modified Schrödinger equation
where
is the deformation parameter. Interesting results were also received in [
11], where a deformation of the Wheeler–DeWitt equation led to a discretization of time. The paper [
12] by P. Caldirola should also be mentioned, where the author postulates the existence of a universal interval
of proper time. The existence of such a universal interval causes the reaction of a particle to the applied external force to not be continuous. As a result, positions of the particle along its world line are discretized.
Breaking of time-translation symmetry is a basic requirement for the creation of time crystals [
13,
14,
15]. Therefore, minimal time scales can lead to a behavior of the system undergoing time evolution similar to that of a time crystal [
10,
11].
This paper is structured in the following way. In
Section 2, we review the Page–Wootters formalism on which the developed theory will be based. In
Section 3, we introduce a minimal time scale into the theory by modifying the commutation relations between the time and frequency operators.
Section 4 contains the analysis of the time evolution of a couple simple quantum systems when a minimal time scale is present. The systems of one, two and three spins-
in an external magnetic field are investigated, as well as a free particle and a harmonic oscillator. The conclusions and a discussion of the received results is given in
Section 5.
2. Page–Wootters Formalism
In a standard description of quantum mechanics, time is treated as a background structure with respect to which quantization is performed. In the Schrödinger equation, time is treated as a classical parameter. Physically, it represents the time shown by a “classical” clock in the laboratory. Moreover, in relativistic mechanics, standard quantization techniques are imposing the canonical commutation relations on constant-time hypersurfaces. Furthermore, the Wheeler–DeWitt equation [
16] says
where
is the Hamiltonian constraint in quantized general relativity and
stands for the wave function of the universe. The universe as a whole is static and does not evolve.
Therefore, it is important to give time a fully quantum description so that time becomes a quantum degree of freedom. In the following section, we will review the Page–Wootters formalism [
17,
18,
19,
20], which accomplishes this and which will be particularly suited for the considerations presented in this paper. Other proposals of quantum descriptions of time can be found in the literature [
21,
22,
23,
24,
25,
26,
27].
In the Page–Wootters formalism, one assigns to the time degree of freedom a Hilbert space
. This Hilbert space describes a “clock”, which will measure the flow of time. The system undergoing the time evolution will be described by the Hilbert space
. The joint Hilbert space of the “clock” and the system is
We will assume that
is isomorphic to the Hilbert space of a particle on a line, i.e., the space
of complex-valued functions defined on
and square integrable with respect to the Lebesgue measure. It is worth noting that other choices are also possible [
28,
29]. On
, we introduce the time operator
corresponding to the measurements of time and the frequency operator
conjugate to
:
When
, then the operators
and
will be represented as the operators of multiplication and differentiation:
On
, we introduce the constraint operator of the model
with
the system Hamiltonian. Such constraint describes the simplest case of the “clock” non-interacting with the system and where the system Hamiltonian is time independent. It is, however, possible to consider more general cases. Since
and
are self-adjoint operators, the constraint operator
will be also self-adjoint. If we treat
as a total Hamiltonian, then in accordance with the Wheeler–DeWitt Equation (
2) we define physical states of the model as vectors
satisfying
We will use the double-ket notation to denote states
from the total Hilbert space
. Formula (
7) defines physical states as (generalized) eigenvectors of the operator
associated with the null eigenvalue. The physical states
provide a complete description of the temporal evolution of the system
by representing it in terms of correlations (entanglement) between the latter and the “clock” system.
The conventional state
of the system
at time
can be obtained via projection of
with a generalized eigenstate
of the time operator
:
where
is the generalized eigenvector of the operator
associated with the eigenvalue
:
By writing (
7) in the time representation in
and with the help of (
5):
we verify that the vectors
obey the Schrödinger equation.
By projecting
on the generalized eigenstates
of
:
where
is the generalized eigenvector of the operator
associated with the eigenvalue
:
we obtain the eigenvector equation of the operator
:
Specifically, for such that equals an element of the spectrum of , then is an eigenvector of at that eigenvalue; otherwise, .
When the total system is in a state
, the probability that the measurement of an observable
(a self-adjoint operator defined on the Hilbert space
) will give a value
from the spectrum of
given that the “clock” reads the time
is postulated as the conditional probability
where
is an eigenvector of the operator
associated with an eigenvalue
. We can see that the above postulate reproduces the standard Born rule of computing the probabilities of measurements.
Remark 1. The original formulation of the formalism by Page and Wootters was criticized as it seemed that it is not reproducing the correct formulas when calculating the probability of measuring an observable at time when one finds the system at an eigenstate of an observable at time . In other words we are not receiving the correct propagators of subsequent measurements of observables and . For the possible resolution of this problem, see [18,19,20]. Remark 2. The Page–Wootters formalism can be extended to incorporate a time-dependent Hamiltonians by replacing the constraint operator (6) withwhere is an operator on that explicitly depends on the time operator . Remark 3. The generalized eigenvectors of operators appearing in the Page–Wootters formalism are formally defined using rigged Hilbert spaces. A rigged Hilbert space consists of a Hilbert space , together with a dense subspace , such that is given a topological vector space structure with a finer topology than the one inherited from ; i.e., the inclusion map ı is continuous. By , we will denote the dual space to , i.e., the space of continuous linear functionals . Similarly, by we will denote the dual space to . The adjoint map ı* to ı defined byis injective by the density of in . Therefore, ı* provides us with an inclusion of in . By identifying with in accordance to the Riesz representation theorem we obtain the inclusion of in , i.e., By (16), the duality pairing between and is compatible with the inner product on : The space plays the role of the space of test functions and is the space of distributions (generalized vectors).
As an example let and be the Schwartz space of rapidly decreasing functions. Then, is the space of tempered distributions. The operators and given by (5) restricted to also take values in . They are also continuous with respect to the topology in . Therefore, they can be extended to operators on and we obtain the following eigenvector equations for these operators Both and are generalized vectors as they are not elements in .
3. Incorporation of a Minimal Time Scale
A minimal time scale can be incorporated into the formalism by modifying the commutation relations for the operators of time and frequency,
and
:
Here
is a positive constant describing the smallest possible resolution with which we can measure time. For
, we recover the standard commutation relations for
and
. The commutation relations of the form (
20) for the operators of position and momentum were considered in [
30] from the point of view of a minimal length scale in quantum mechanics. In the following subsection, we will review some of the results from this paper which will be crucial for our later considerations.
3.1. Minimal Time Scale Uncertainty Relation
The commutation relations (
20) imply the following generalized uncertainty principle for the uncertainties
and
of the measurements of time and frequency:
From (
21), it is not difficult to see that there exists a smallest possible value for the uncertainty
(see
Figure 1):
Every physical state cannot have the uncertainty in time smaller than this value. For states for which the expectation value of the frequency operator
is equal zero, we receive the absolutely smallest uncertainty in time equaling
It should be noted that there are states in the Hilbert space , like, for example, the eigenstates of the time operator , for which the uncertainty in time is smaller than . Such states are not considered physical.
The modification (
20) of the commutation relations for
and
is the simplest one leading to a minimal uncertainty in time. Other modifications are possible; however, we will not be dealing with such cases in the current work.
The introduction of a minimal uncertainty in time will cause the time operator
to not be essentially self-adjoint but only symmetric. A consequence of this is that eigenvectors of
are not physical states. It is no longer possible to consider states of the system
at particular instances of time. Therefore, we do not have at a disposal the representation of the Hilbert space
given by projecting states
onto eigenvectors
of
. However, the frequency operator
is still self-adjoint, and therefore we can use it to construct a frequency representation of the Hilbert space
. The Hilbert space
is represented as a space
of functions defined on
and square integrable with respect to the measure
. Such functions can be formally constructed by projecting states
onto eigenvectors
of
:
The scalar product in the Hilbert space
is given by the formula
and the frequency and time operators
,
are represented as appropriate multiplication and differentiation operators:
One can check with a direct calculation that the operators defined by (
26) satisfy commutation relations (
20).
Although eigenvectors of the time operator
are not physical states, it is possible to define states which are physical and closely resemble eigenstates of
. These are states of maximal localization around instances of time. A state
of maximal localization around an instance of time
is a state for which the expectation value of
is equal
and the uncertainty of
is equal to the smallest possible value
:
The conditions in (
27) uniquely determine the state
, which in frequency representation is given by the formula
Using the states of maximal localization, it is possible to define a representation of the Hilbert space
, which we will call a continuous time representation. This representation is given by projecting states
in
with states of maximal localization
The received wave functions
describe the probability amplitude for the system being maximally localized around the instance of time
. In the limit
, the wave function
of the ordinary time representation is recovered. From (
29), we receive the transformation of a state’s wave function in the frequency representation into its continuous time representation:
The inverse transformation is given by the formula
The operators
and
in the continuous time representation take the form
3.2. Discrete Time Representation
The family of states
for
forms an overcomplete set of vectors. We can choose from this set a smaller countable set forming a basis in the Hilbert space
. Using this basis, we can construct another representation of
which we will call a discrete time representation. For
, let us consider the following set of vectors:
The vectors composing this set are linearly independent and the set itself is complete. Thus, we receive a one-parameter family of bases of the Hilbert space
. These bases can be viewed as lattices with spacing
shifted by
. As we will see, the resulting discrete time representations will describe time evolution on lattices. The vectors from the set (
33) are not orthogonal. The scalar product of two states of maximal localization is equal:
From this, we can see that
so that the vectors from the set (
33) are orthonormal except for neighboring vectors.
Using a basis (
33), we can construct a discrete time representation of the Hilbert space
. Instead of expanding a state
in this basis and taking the sequence of coefficients of this expansion as the representation of the vector
, we will represent the state
with the following sequence:
From this formula, we receive the transformation of a state’s wave function in the frequency representation into its discrete time representation:
To calculate the inverse transform, let us rewrite the above formula by changing the variable under the integral sign:
From this, we can see that
are Fourier coefficients in the expansion of the function
into the Fourier series:
The function
is square integrable, which follows from the square integrability of
with respect to the measure
. Therefore, by virtue of Carleson’s theorem, the above Fourier series converges for almost all
, where the infinite sum
is viewed as the limit
. We can then change the variable again to receive the following inverse transform of (
37):
The transformation of a state’s wave function in the continuous time representation into its discrete time representation is the following:
To calculate the inverse transform, we combine transformations (
30) and (
41) receiving
Calculating the integral in the above formula, we receive the inverse transform of (
42):
where
. The transformations (
42) and (
44) might seem surprising because they allow us to reproduce the function
from the knowledge of its values at a countable set of points. This is, however, nothing strange because from (
30) after changing the variable under the integral, we can see that
is a Fourier transform of a square integrable function with compact support. Thus, by virtue of Paley–Wiener theorem,
extends to an entire function of exponential type and entire functions are completely determined by their values at a countable set of points.
In what follows, we will derive the representation of the frequency operator
in the discrete time representation. Let us define a discrete derivative with the formula
From this we get
and consequently
where
is a function with an inverse
and we used the expansion of this function in a Taylor series
From (
48) and the fact that in the frequency representation the operator
is a multiplication operator (see (
26)) we receive the discrete time representation of the frequency operator
Using this result and Formulas (
26) and (
37) we can also receive the discrete time representation of the time operator
3.3. Time Evolution with a Minimal Time Scale
The introduction of a minimal time scale into the formalism will change the time evolution of the system. Since the time operator is no longer self-adjoint, its eigenvectors are not physical states and it is meaningless to talk about the system at particular instances of time. We can only consider the system localized around instances of time with a finite precision. For this reason, as indicators of time of the evolving system, we will take states of maximal localization around instances of time instead of eigenstates of the time operator .
The state
of the system
maximally localized around an instance of time
can be obtained via projection of the physical state
with the maximal localization state
:
To receive a modified Schrödinger equation governing the time evolution of states
, we write the constraint equation
in the continuous or discrete time representation in
. Before doing this, first let us use Theorem A1 from
Appendix A with the function
to rewrite the constraint equation
in a different form
In the continuous time representation, the constraint equation takes the form
from which and (
32) we receive the modified Schrödinger equation
Similarly, by writing the constraint equation in the discrete time representation and using (
50), we receive the discrete Schrödinger equation
The solution of (
56) subject to an initial condition
is of the form
and the solution of (
57) subject to an initial condition
is of the form
Equations (
56) and (
57) are equivalent and describe the same time evolution of the system. Knowing the solution of either of these equations, we can reconstruct the solution of the other equation by applying the transform (
42) or (
44).
5. Conclusions and Discussion
We developed a theory of non-relativistic quantum mechanics exhibiting a non-zero minimal uncertainty in time. The theory was based on the Page–Wootters formalism with a modified commutation relations between the time and frequency operators. We received a modified version of the Schrödinger equation in a continuous and discrete time representation. Interestingly, both representations are equivalent, so, in particular, by solving the discrete Schrödinger equation, we receive a time evolution on a lattice of a state vector, which we can then transform to the continuous time representation. It should be noted that even though we have a discrete time evolution, it does not mean that fundamentally time is discrete. It only means that, using a particular representation, we can describe time evolution on a lattice.
The developed formalism was used to describe time evolution of couple simple quantum systems. The received results could be tested experimentally, at least in principle. In particular, we found that in the system of spins- in an external magnetic field, the frequency of precession of the spins depends on the number of spins composing the system. Furthermore, the existence of a fundamental limit with which we can measure time causes the entanglement of the system of spins. These results are quite peculiar, especially that the spins do not need to explicitly interact with each other, but still there exists some kind of influence between them. The interpretation of these results can be difficult and may suggest that perhaps the considered modification of commutation relations is not physically correct and some different modification will be more suitable.
It should be noted that we have not explicitly assumed that the parameter describing the minimal time scale is dependent on the system undergoing time evolution. However, in view of the comment from the previous paragraph it might be possible that will depend on the system. Such a possibility will require further investigation.
In this paper, we only considered time-independent Hamiltonians. It is possible to generalize the theory to include time-dependent Hamiltonians by considering a constraint operator
as in (
15). However, in this case, it will not be possible to use Theorem A1 when deriving time evolution equations because operators
and
composing the constraint operator
do not commute.
Worth investigating will also be a relativistic quantum mechanics exhibiting a non-zero uncertainty in time and space. In the Page–Wootters formalism, time is treated as a quantum degree of freedom resulting from a correlation between the “clock” and the rest of the system. Interestingly the same approach can be performed to describe space degrees of freedom [
31]. Such formalism will have an advantage of treating time and space on an equal footing. Lorentz transformations can be defined as an appropriate unitary transformation. It is then possible to consider different modifications, commonly found in the literature, of the commutation relations between
and
operators, as well as position and momentum operators.