1. Introduction
The problem of recovering the laws of classical physics from the quantum level is beset by conceptual problems. The exploration of the difference in the logical structure of the classical and quantum level was already started a long time ago by pointing out that the logic of quantum states is non-distributive as opposed to the Boolean logic of classical physics [
1,
2]. This structure was generalized later [
3,
4] but it remains to be seen how such an approach can help us to better understand the problem. The solution may even lie beyond quantum theory and be related to the usual separation of physical laws and initial conditions [
5,
6]. A slightly different trial starts with the construction of hidden parameter theories [
7] and seems to end at contextuality [
8,
9,
10] without reconciliation of the two regimes in sight.
Maybe the most disturbing qualitative difference of the quantum and classical descriptions is the apparent non-deterministic nature of the choice of the spectrum element of an observable, which is realized in a measurement. One is advised to rely on a new dynamical mechanism [
11], such as spontaneous localization [
12,
13,
14] or quantum stochastic processes [
15], as a quick fix up of this selection process. In the meantime, the phase transition, in particular, the spontaneous symmetry breaking [
16,
17,
18,
19,
20,
21,
22] has been proposed as an important ingredient of the collapse of the wave function, the selection of the observed spectrum element, without evoking new principles. A more systematic approach to deal with logical incompatibilities is the decoherent histories program [
23,
24,
25,
26], which restricts attention only to the completely decohered sequences of observations.
One can make a step toward a better understanding of these problems with the help of a peculiar feature of physical laws, namely, the measured physical values depend on the scales of the measurement. Furthermore, the quantitative dependence leads to the dependence of the physical laws on the scales of observation. For instance, one extracts different physical laws when nature is observed with different spatial resolutions. Thus, the quantum–classical crossover can, in principle, be systematically discussed in the framework of the renormalization group, a scheme to deal with the scale dependence of physical laws.
However, the scale dependence of physical laws is supposed to be smooth. How can the logic and determinism change continuously between the quantum to the classical regimes? A hint to the answer might be found in more recent developments. The need for increased accuracy of monitoring the motion of macroscopic bodies during the detection of gravitational waves has led to non-demolition measurements [
27], weak measurements [
28], a more careful treatment of quantum noise [
29], and to optomechanical devices [
30]. These examples show that classical concepts can be brought in agreement with quantum effects in a consistent and systematic manner. The idea of continuous measurement [
31] is another possibility to fit the quantum and the classical domains. The common elements of these ideas is the introduction of a flexible length or time scale, which suppresses the uncertainties in the measurements, thereby approaching classical physics from the quantum level. Another development, which narrows the qualitative differences of the quantum and the classical level is the possibility that the quantum jump, assumed to make the choice of the measured value, is actually a dynamical process enfolding in finite time [
32,
33].
The proposition is put forward in the present work that the generalization of the central limit theorem to quantum mechanics provides a simple and generic framework to approach classical physics in the macroscopic limit. The relation between the renormalization group and the central limit theorem was already discussed in the context of the generalized central limit theorem and the UV fixed points [
34,
35,
36,
37,
38]. The central limit theorem is considered here from the point of view of quantum mechanics without the systematic methodology of the renormalization group. The traditional classical laws are approached but never reached exactly, according to this scenario, in a manner similar to the way thermodynamics emerges from statistical physics. The surprisingly large value of Avogadro’s number renders the uncertainty of the average over a macroscopic sample negligible for all practical purposes without getting bogged down into the intricacy of contradicting logical structures. Such an approach of classical physics is not a new idea; it is the backbone of almost all attempts to arrive at the well-known classical domain from a microscopic level. The modest steps are presented below, simply to draw attention to analogies of the way that the central limit theorem functions in classical probability theory.
The macroscopic limit of quantum systems can be realized in different manners. The typical rearrangement corresponds to the measurement of a microscopic property by a macroscopic measuring apparatus. To comply with the probabilistic predictions of quantum mechanics, one considers a large set of equivalent and independent microscopic systems. Another possibility is to look for macroscopic quantum effects, where a single observed system of macroscopic size is used [
39]. Such macroscopic quantum phenomena form a wide and colorful set. The simplest is the indistinguishability of identical elementary particles, leading to the solution of Gibbs mixing paradox. It is perhaps the most universal macroscopic quantum effect, owing to its scale independence. Superconductivity or superfluidity are related to spontaneous symmetry breaking and are clear realizations of macroscopic phenomena, treated by an approximation valid only in the thermodynamical limit. In general, the semi-classical approximation can be used to discover a wide class of macroscopic quantum phenomena as saddle point effects [
40]. The non-unique definition of the actual number of degrees of freedom [
41,
42,
43,
44,
45] and the non-triviality of their distribution over the possible states [
46] indicate the richness of macroscopic quantum phenomena. A common feature of the macroscopic limit in different systems is the decoherence of the macroscopic (collective) variables [
47,
48,
49]. It is important to realize that the macroscopic variable may have an intrinsic environment within the system without a tensor product structure of the Hilbert space [
50]. We restrict our attention in this work to the first case, where the measured phenomenon is microscopic and only the measuring apparatus amplifies it to a macroscopic size.
We show that the average of an observable over a set of independent systems behaves classically up to quantum fluctuations and such a reduction of the quantum fluctuations takes place independently of the measurement process. The actual measurement apparatus can be treated in a similar fashion to establish the almost classical nature of its macroscopic pointer variable. Hence, the measurement of an average over a large set of microscopic systems involves a correlation between two classical variables. However, the measurement of a single microscopic event reveals the usual difficulty of the measurement theory, the dynamics of the amplification of a microscopic signal to a macroscopic level.
We start with the extension of the central limit theorem over expectation values in
Section 2 by calculating the second cumulant of the distribution of the average of an observable over of a set of
independent microscopic systems. The decrease of the width of the fluctuations is compatible with a reduction of the Planck constant
. The determination of the effective dynamics of the average over the microscopic system observables requires a scheme that can handle open interaction channels and remains valid at both sides of the quantum–classical crossover.
Section 3 contains a brief motivation of the Closed Time Path (CTP) formalism used in this work. The dynamical generalization of the central limit theorem for the average of the expectation value of a continuous coordinate is given in
Section 4. The role of the cumulants is taken over by the connected Green functions and they are shown to become suppressed as
. The result of the continuous observation is spoiled by the mass-shell singularities. To avoid this problem, the observations over a discrete set of time are considered. Furthermore, the condition of the applicability of the central limit theorem is given for an interactive set of microscopic systems. The clarification of the role of a macroscopic measuring apparatus in the measuring process is taken up in
Section 5. It is shown that a harmonic measuring apparatus realizes a linear amplification, and Wilson’s cloud chamber is briefly discussed, where the collapse of the wave function takes place as a dynamical process in a strongly coupled environment. Finally, the summary is presented in
Section 6. Two appendices are included with some technical details. The basic equations of the CTP propagator are collected in
Appendix A. A harmonic measuring apparatus is discussed in
Appendix B.
2. Central Limit Theorem for an Observable
We consider the measurement of an observable A on a closed microscopic system whose pure states belong to the Hilbert space . The measurement is repeated on the ensemble of copies of the system; the pure states of the ensemble form the direct product . The observable A represented on the n-th system, , acts on the n-th factor of the direct product space, and stands for the average observable over the system ensemble. If the systems are indistinguishable, then one has to (anti-)symmetrize their state in but this does not change the expectation value of where the elements of the ensemble are not distinguished. We allow some variation of the preparation of the systems within the ensemble, namely, we assume that systems are placed into the state characterized by some reduced density matrix whose detailed specification is unnecessary at the moment; all we need is the normalization .
The limit can be realized by two different scenarios: (i) The textbook example consists of an ensemble of identically prepared, independent microscopical systems, accessed one-by-one, and the expectation value of is identified with the average of the measurements. (ii) Another, more realistic alternative is to consider a set of microscopic systems, where we measure directly , e.g., the center of mass of a mirror in a gravitation radiation detector or a total electric dipole moment of a solid in a linear response study.
The central limit theorem can easily be recovered for transition amplitudes or expectation values. We discuss the latter; the former can be recovered in similar manner. Let us start with the cumulant generator function of the observable
A, defined by the following:
We assume a continuous, unbound spectrum; the extension of our procedure for a bounded or discrete spectrum is straightforward. The probability distribution of
A is given by the following Fourier transform:
since
It is easy to see that the generator function of
,
assumes the following form:
where
is the average of the generator functions
. One uses quenched averaging because the choice of the initial state influences the system for an arbitrarily long time in closed dynamics. It is a matter of a trivial expansion in
to find the normal distribution around
of
width,
The decisive dynamical difference between scenarios (i) and (ii) is that while several measurements are performed in (i), only a single measurement is carried out in (ii). The central limit theorem corresponds to scenario (i), and the dynamical aspect of the narrowing of the peak of in scenario (ii) is that the disturbance of a single measurement is distributed over a large number of systems. Such a point of view is consistent with the effective reduction of Plank’s constant, , for canonically conjugate pairs, and . The suppression of the incompatibility of canonically conjugate pairs opens the possibility of recovering a unique trajectory for the average observable . Hence the single measurement of scenario (ii) is non-demolishing when . The mathematical equivalence of the two scenarios indicates that classical physics is reached by the averages in both cases.
The non-demolishing nature of the measurement of
can simply be established in the Heisenberg representation. Let us assume that the ensemble of systems of scenario (i) and the measuring apparatus follow a closed dynamics with Hamiltonian
, where
denotes the Hamiltonian of the apparatus,
and
represent the system Hamiltonian
and the system-apparatus interaction energy
, respectively, on the
n-th microscopic system,
B being an apparatus operator. The Hamiltonian of scenario (ii) is the same, except
. The commutator
in the Schrödinger representation implies
in the Heisenberg representation where
and the integration of the Heisenberg equation yields the following:
where the measurement is non-demolishing [
27] as
. This result is formal as it stands since possible singularities in time or frequency are ignored; a more careful discussion of this point follows below.
The difference between the two scenarios is shown clearly by the dynamics of the measuring apparatus, namely, an apparatus observable, , has the commutator with and 0 in the cases (i) and (ii), respectively. The disturbance of the apparatus by the measurement increases with the number of measurements, and both the measured system and the measuring apparatus should be renewed after each measurement.
3. CTP Formalism
The formalism to follow the macroscopical limit of quantum systems should be a CQCO scheme: it should handle classical (C) [
51,
52,
53,
54,
55] and quantum (Q) [
56,
57,
58,
59] dynamics because the macroscopical limit implies classical physics should further cover closed (C) and open (O) dynamics since a macroscopic system cannot be kept isolated on the microscopic scale. These possibilities are offered by the CTP formalism, initially proposed to deal with the perturbation expansion in the Heisenberg representation [
60] and with non-equilibrium phenomena in many-body systems [
61]. The dynamics is defined in this work by a CTP action functional, used with the variational principle in classical mechanics and with the path integral in the quantum case. The basic idea of this scheme is outlined below for a one-dimensional particle with coordinate
x within the time interval
.
3.1. Classical Closed Dynamics
The motion is followed in the CTP scheme in both directions of time [
62] along the continuous trajectory
,
in such a manner that the particle follows a second order equation of motion forward and backward in time for
and
, respectively. The auxiliary conditions on the trajectories are
,
, with the given initial coordinate and velocity,
and
, respectively. The auxiliary conditions usually make the velocity discontinuous at
, at the reversion of the time direction. The trajectory
is then broken up into two trajectories by introducing
and
for
. The set of varied trajectories is, therefore, given by the pair of trajectories
, which satisfy the initial conditions
,
and are closed at the final time,
owing to the continuity of
.
The action of the trajectory pair is the following:
where
denotes the traditional action. The minus sign in front of the second term on the right hand side is induced by the reversed time direction for
and the infinitesimal imaginary term is added to split the degeneracy of the action, a necessary condition to define Green functions, cf.
Appendix A.
The CTP generalization of the traditional action is optional in the case of closed dynamics, the only advantage being the trading of the boundary conditions, required by the traditional variation principle, to initial conditions. However, this change may be important for non-integrable dynamics.
3.2. Classical Open Dynamics
The traditional action principle of classical mechanics is applicable for systems with holonomic forces, represented by a term
in the Lagrangian, which induces a conservative force, as follows:
in the Euler–Lagrange equation. This can be generalized to a semi-holonomic force [
63], as follows:
generated by a function
in the Lagrangian. We separate here two roles of the trajectory
, namely, the identification of the location of the particle and the variable on which the derivative acts. The proper bookkeeping of such a separation is the reduplication of the degree of freedom,
, introduced already above. The imposition of the equation of motion and identical initial conditions for
and
sets
and guarantees
,
in (
10).
The simplicity of the action (
8) such that the trajectories
are not coupled is a characteristic feature of closed dynamics. Let us now assume that the system and its environment together follow closed dynamics, given by the action
where
x and
y denote the system and the environment coordinate, respectively. It is easy to see that the elimination of the environment degrees of freedom by their equations of motion couples
and
. The form
of the action, obtained by separating the coupling between the trajectories in
, shows that the holonomic and the semi-holonomic forces are generated by
and
, respectively. One can see furthermore that the open forces arising in any subsystem of closed dynamics are semi-holonomic.
3.3. Closed Quantum Dynamics
The reduplication of the degrees of freedom, needed to handle classical semi-holonomic forces by the variational principle, arises from quantum mechanics, as well. In fact, let us consider the expectation value
of the observable
A in the pure normalized state
. This expression involves two representatives of the same state, a bra and a ket, progressing in opposite directions in time. The summation is over the quantum fluctuations in the basis
, and the factorizability of the coefficients
indicates that the quantum fluctuations in the bra and the ket are independent. The independent dynamics of the bra and ket sector can be represented by employing two Hilbert spaces [
64], giving rise to a reduplication of the physical operator set. This structure is optional for closed dynamics since the dynamics in the two Hilbert spaces are equivalent; they are related by Hermitian conjugation.
The dynamical role of the reduplication is particularly clear when the expectation value (
11) is considered at some time
t,
, where
stands for the density matrix. The path integral expression for the density matrix, obtained by breaking up the two time evolution operators into the product of infinitesimal time propagation, is as follows:
where one integrates over trajectories
, and the action is given by (
8). The two trajectories represent the independent quantum fluctuations in the bra and the ket factors.
3.4. Open Quantum Dynamics
The state of an open system is, in general, mixed, and its density matrix is not factorizable, as with a sum of at least two contributions. Hence, the bra and the ket fluctuations are correlated and inseparable, and the formal reduplication of the degrees of freedom, mentioned for closed dynamics, becomes unavoidable. It is advantageous to use the Keldysh variables, namely, the average and the difference , representing the physical coordinate and the quantum fluctuations, respectively.
The reduced system density matrix can be obtained by integrating out the environment degrees of freedom in the path integral, as follows:
where the integration is over the open system trajectories,
, and closed environment trajectories,
. The effective system action
to be used in (
12) for the reduced system density matrix now contains the influence action functional [
65] given by the following:
The relation
follows from the unitarity of the underlying closed full dynamics.
The CTP formalism offers a simple access to decoherence [
47,
48,
49]. The decoherence is a basis-dependent phenomenon, and its appearance in the coordinate representation can easily be seen by inspecting the path integral (
12). The coordinate decoherence is qualitatively the suppression of the strongly off-diagonal elements of the density matrix
. This is a dynamical process [
66] resulting from the system–environment interactions and can be recognized as a suppression of the contributions of trajectory pairs
with large
. The natural source of such a suppression is the increase of
, and the strongly decohered semi-classical dynamics of a macroscopic variable corresponds to a path integral where only trajectory pairs
contribute. Such a path integral is actually a simple realization of the decoherent histories view of quantum dynamics.
The dynamical origin of the reduplication of the degrees of freedom is similar in classical and quantum dynamics. The reduplication is optional for closed dynamics and becomes a necessity to handle semi-holonomic forces and the mixed components of the state. The reduplication leads to substantial complications in the equations, which may appear, at first sight, as unnecessary. However, these complications correspond to physical processes and their function is to enlarge the range of applicability of the formalism.
The CTP scheme is used in this work to find the effective dynamics of the average coordinate over a ensemble of particles and of the pointer variable of the measuring apparatus. The central limit theorem is easiest to state in terms of connected Green functions; it amounts to the claim that the limit
suppresses the
n-th order Green function to
. Few details about the second-order Green function, the propagator, are collected in
Appendix A.
6. Summary
The proposal is put forward in this work that one can avoid the logical conflict between quantum and classical physics when the latter is considered to be the macroscopic limit of the former. The quantum generalization of the central limit theorem is given by showing the narrowing of the probability distribution of the average observable over an independent identical system in the limit . A strong Gaussian suppression of the off-diagonal fluctuations is found, which removes the difference between the two CTP copies of the same physical degree of freedom, thereby restoring the unique classical trajectory, in agreement with the consistent history view of the macroscopic limit and the idea of the spontaneous localization. Therefore, the proposition is to regard classical physics as the limit of large but finite quantum systems, where the size of Avogadro’s number suppresses quantum fluctuations for all practical purposes. The conditions on the strength of the correlations among the microscopic systems are given to preserve the simplicity of the result. The observation over a discrete set of times reveals the discrete dynamics, characteristic at the time scale of the observations.
These results correspond to the effective dynamics of an observable, without considering the measurement—the process of the observation. When the central limit theorem is applied to the pointer variable of the measuring apparatus, it is advantageous to separate the case of measuring an average or a single quantity. The measurement of an average observable over a large ensemble of microscopic systems implies a classical correlation between two classical trajectories: that of the average measurement and the pointer. A harmonic apparatus yields a linear amplification of the measured signal within the realm of the central limit theorem. To achieve a non-linear amplification, we need correlations among the constituents of the apparatus beyond the limit the applicability of the quantum central limit. In the case of a single microscopic measurement, the linear amplification is obviously insufficient, and the measuring apparatus must be strongly correlated; the line of thought followed here is not applicable.
It is argued that Wilson’s cloud chamber represents a simple set of both weakly and strongly correlated constituents, where on the one hand, one can use the central limit theorem for the pointer variable and, on the other hand one, can retain the quantum fluctuations of the ionizing particle. The droplet formation in the chamber is a dynamical realization of the assumed collapse of the wave function. This view suggests that Neumann’s postulate of a sudden change of the quantum state can be avoided, and the impulsive nature of the quantum jump results from a large number in the dynamics, , which generates new time scales.
This idea is a natural extension of the view of the measurement as a spontaneous symmetry breaking, the breakdown of the initial rotational symmetry by the formation of the first droplet. The high degeneracy of the symmetrical ground state before the measurement is the dynamical origin of the strong correlation of the microscopic system and the measuring apparatus. In fact, the initial microscopic inhomogeneities of the gas exert a strong impact on the probability distribution of the location of the first droplet when the gas dynamics is taken into account by the (almost) degenerate perturbation expansion.
However, any further argument within the framework of quantum mechanics or quantum field theory can only push back the origin of the wave function collapse to earlier initial conditions without providing a deterministic origin. Therefore, the main problem of measurement theory, the choice of the observed spectrum value of an observable, remains untouched by these considerations.
The central limit theorem, used in this work, relies on a set of independent or weakly coupled microscopic systems. One may continue this line of thought and reach more realistic models by moving away from the vicinity of a Gaussian fixed point. This step implies the quantum renormalization group [
75] requiring the use of quantum field theoretical methods.