Sie sind auf Seite 1von 60

March

21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

THE ZERO-POINT FIELD AND THE EMERGENCE OF THE


QUANTUM

L. DE LA PENA,
A. M. CETTO and A. VALDES-HERNANDEZ
Instituto de Fsica, Universidad Nacional Aut
onoma de M
exico, Apartado Postal 20-364,
M
exico, DF, Mexico
luis@fisica.unam.mx,ana@fisica.unam.mx,andreavh@fisica.unam.mx
Received Day Month Year
Revised Day Month Year
A new way of arriving at the quantum formalism is presented, based on the recognition
of the reality of the random zero-point radiation field (zpf). The quantization of both
matter and radiation field is shown to emerge as a result of the permanent interaction
of matter with the zpf. Quantum mechanics is obtained both in its Schr
odinger and
its Heisenberg version, under certain well-defined conditions and approximations. The
theory provides for an explanation of the origin of entanglement. Further, the same
physical elements and hypotheses allow us to cross the doorway and go beyond quantum
mechanics, to the realm of (nonrelativistic) quantum electrodynamics.
Keywords: Foundations of quantum mechanics; zero-point field; radiative corrections
PACS numbers: 03.65.-w, 31.30.J-, 42.50.Lc, 44.40.+a

1. Introduction
In this paper we present a new way of arriving at the quantum formalism, based
on the recognition of the reality of the random zero-point radiation field (zpf).
The advantage of this approach lies in that one sees into the quantum world from
outside, which affords a perspective wider than the one reachable from within the
quantum theory proper. Many of the usual conceptual problems that characterize
present-day quantum mechanics (qm) are thus dissolved, and new physical elements
are integrated that help to clarify its physical and conceptual contents. At the same
time, the fresh perspective proposed invites us to cross the doorway and go beyond
the strictly quantum-mechanical realm.
When referring to the problems of quantum mechanics we have in mind basically those conceptual puzzles frequently bordering philosophy of science, to the
annoyance of some physicists, although their nature is physical that have been
under scrutiny and debate since the early days of the theory, but remain basically as
unsolved now as they were eighty years ago. We recall as examples: the irreducible
or unexplained indeterminism characteristic of the theory; the fact that it predicts
probabilities, not outcomes; that it then requires a measurement theory for its accomplishment, which means opening the door to observers and their subjectivism,
1

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

and giving birth to undefined boundaries between the quantum and the classical
world. Add to this the loss of realism; two opposed laws of evolution (Schrodingers
equation and the collapse of the wave function); nonlocality and, for many, superluminal influences, and so on. There are also some strictly technical questions, such
as the lack of a clear physical explanation of the mechanism that stabilizes the atom
and leads to quantized states (rather than a mere description of the phenomenon),
or of the mechanism that entangles two identical noninteracting particles. The list
is not short. All these are real questions, of actual interest to physics and to which
hundreds of books, papers, conferences and meetings have been devoted.
The paper is structured into four parts, according to the lectures of the course for
which they were prepared. The first part (section 2) presents Plancks distribution
for the blackbody radiation field as a consequence of the existence of the zpf without
any quantum hypothesis, and discusses the implications of this result. The second
part (section 3) presents Schr
odingers equation as a consequence of the action of the
same zpf on matter, and discusses the conditions under which quantization emerges.
The third part (section 4) shows that the Heisenberg formalism of quantum theory
can be derived on the basis of the same principles, and exhibits the conditions
needed to arrive at this result as well as some of its implications, including the
emergence of entanglement. The last part (section 5) shows that also the first-order
nonrelativistic radiative corrections of qed are correctly obtained, thus confirming
that the theory presented goes beyond quantum mechanics.
2. Plancks Distribution and the Zero-Point Energy
Some very fundamental properties of the equilibrium radiation field can be derived
from the mere consideration of the pervasive presence of its zero-point contribution.
For this purpose it is convenient to start by reviewing the thermodynamics of a
harmonic oscillator of frequency , which can be taken to represent a mode of the
field of that frequency. A careful thermo-statistical analysis of an oscillator system
in thermodynamic equilibrium reveals the far-reaching implications of the existence
of a fluctuating zero-point (nonthermal) energy term.
For earlier literature and details on the material presented in the following sections, see Refs. 1-10.
2.1. The zero-point energy and energy equipartition
Let us consider the radiation field as made of independent modes of oscillation of
frequency , in thermal equilibrium at a given temperature T. Then according to
Wiens law, the mean energy of every mode is given by an expression of the form11
U = f (/T ) = f (z).

(1)

Since no specific details about the oscillator are needed to arrive at this result, the
function f should have a universal character. This law will be at the base of our
considerations below.

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

In the limit T 0, Eq. (1) gives for the mean energy


E0 U (0, ) = f () = A,

(2)

so the zero-point energy (zpe) E0 the energy of the oscillators at absolute temperature T = 0 is determined by the value that f attains at infinity. In the usual
thermodynamic analysis the value of the constant f () = A is arbitrarily chosen
as zero, so there is no zpe. However, the more general (and more natural) solution
corresponds to a non-null value of A. In the case of the radiation field, this represents a physically more reasonable choice than a vacuum that is completely devoid
of electromagnetic phenomena. By taking A to be nonzero we attest the existence
of a zpe that fills the whole space and is proportional to ,
E0 = A = 21 ~.

(3)

The value of A (with dimensions of action) must be universal because it determines


the (universal, according to Kirchhoff) equilibrium distribution at T = 0; we have
put it equal to ~/2 in order to establish contact with present day knowledge.
A value of A different from zero means a violation of energy equipartition among
the oscillators, since the equilibrium energy becomes a function of the oscillator frequency. This holds at any temperature, at least because the zpe is part of the equilibrium energy. Hence the ensuing physics necessarily transcends classical physics.
Let us note that the selection made for A assigns an interesting meaning to
the variable z, which so far has been written just as z = /T. Since z should be
dimensionless (as follows from Eq. (1)), given the parameters at hand it becomes
naturally expressed as the ratio of two energies, z = ~/kB T.
2.2. General thermodynamic equilibrium distribution
Our aim now is to find the mean equilibrium energy U of a system of oscillators
as a function of the temperature. For this purpose we revisit in the following three
subsections some general thermo-statistical results, which are not exclusive of the
harmonic oscillator, but hold for any physical system that is in equilibrium at temperature T .
Following a standard procedure in thermodynamics, the probability that the
energy attains a value between E and E + dE for a fixed temperature can be written
in the general form for a canonical ensemble
1
g(E)eE dE,
Zg ()
Z
Zg () = g(E)eE dE,

Wg (E)dE =

(4a)
(4b)

where = 1/(kB T ) is the inverse


R temperature, Zg () is the partition function
that normalizes Wg (E) to unity, 0 Wg (E)dE = 1, and g(E) is a weight function
representing the intrinsic probability of the states with energy E, known as the

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

structure function. The mean value hf (E)i of any function f (E) is of course
Z
Wg (E)f (E)dE.
hf (E)i =

(5)

For the particular case f (E) = E, (5) gives the average value hEi = U by definition.
Equation (4a) constitutes an extension of the usual Boltzmann distribution to
the general case in which the states with energy E can have an intrinsic probability
that depends on E. The (classical) Boltzmann distribution is obtained from (4a) in
the particular case in which g(E) does not depend on E and can therefore be written
in the form (with due account of the dimensions)12
1
,
(6)
s
where s is a constant with dimensions of action, so g has the dimension of
(energy)1 . In this case one gets from the above equations:
gcl (E) =

eE
Wcl (E) = Wgcl (E) = R E ;
e
dE
0
1
hEi = U = = kB T.

(7a)
(7b)

From the last equation it follows that E0 = hE(T = 0)i = 0. This means that to
allow for the a nonthermal energy of the system a form for g(E) different from that
given by Eq. (6) must be used. Finding this g(E) becomes an important task in
what follows.
2.3. Thermal fluctuations of the energy
Equations (4) and (5) lead to a series of important results. With f (E) = E r , r a
positive integer, it follows that (the prime indicates derivative with respect to )
0

hE r i =


Zg0 r
hE i E r+1 ,
Zg

(8)

and further, from (4b),


1
hEi = U =
Zg

Z
0

Eg(E)eE dE =

Zg0
.
Zg

These two expressions combined give the recurrence relation

r+1
0
E
= U hE r i hE r i .
With r = 1 this gives a most important expression for the energy variance,


dU
,
E2 E 2 U 2 =
d
which can be rewritten as the well-known relation11


U
dU
2
2
E =
= kB T
= kB T 2 C
d
T

(9)

(10)

(11)

(12)

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

in terms of the specific heat (or heat capacity) C . Because C (=CV ) is surely
finite at low temperatures, the right-hand side of this expression is zero at T = 0,
whence
E2 (T = 0) = 0,

(13)

indicating that the present description does not allow for the dispersion of the
energy at zero temperature. Of course, this result refers to thermal fluctuations,
since the description provided by the distribution Wg is of thermodynamic nature:
temperature-independent fluctuations find no place in Wg . In the particular case of
a system of harmonic oscillators this becomes an important shortcoming, since the
zpf, being part of the field in the cavity, shares with it the presence of unavoidable
fluctuations.
2.4. General statistical equilibrium distribution
To go ahead it is required to derive an expression for E2 that allows for the nonthermal fluctuations excluded by the distribution Wg , and is expressed in terms of
the mean equilibrium energy U (). This can be achieved by paying attention to the
statistical distribution of the energy as a function of the mean energy, in contrast
to the thermodynamic description studied in section 2.2. In this form we are able to
bypass the problem presented by the still unknown function g(E). For this purpose
we look for a distribution Ws (E) that, according to a well established principle,
maximizes the entropy Ss , defined (up to an arbitrary additive constant) as12,13
Z
Ss = kB Ws (E) ln [cWs (E)] dE,
(14)
where c is an appropriate constant with dimensions of energy. In contrast to the
thermodynamic entropy S which is defined in the phase space of the particles, with
w(p, q) the distribution in such space,
Z
S = kB w(p, q) ln w(p, q)dpdq,
(15)
the statistical entropy Ss is normally interpreted as a measure of the disorder present
in the system. Thus the demand of maximal entropy implies maximum disorder,
which is considered the natural order under equilibrium. In the case of interest
here the system consists of an immense (practically infinite) number of independent
modes of the field that for each frequency interfere among themselves, so a final
state of maximal disorder is to be naturally expected.
The maximum-entropy formalism is designed to determine a distribution Ws (E)
that maximizes the function Ss subject to a set of auxiliary conditions (or constraints), which in the present case take the form (the subscript s denotes averages
with respect to Ws , to be distinguished from quantities calculated with Wg )
Z
Z
Ws (E)dE =1,
EWs (E)dE =U.
(16)

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

Of course, at the moment the mean value U () is still an unknown function of ,


to be determined. By applying the method of the Lagrange multipliers one arrives
thus at16
Ws (E) =

1 E/U
e
.
U

(17)

Note that the specific choice U = 1 (corresponding to E0 = 0) results in the usual


canonical distribution Eq. (7b).
Equation (17) gives the general statistical result
hE r is = r!U r ,


from which it follows, in particular, that E 2 s = 2U 2 ,whence
(E2 )s = U 2 .
We see that Ws indeed allows for nonthermal fluctuations, since at T = 0

(E2 )s 0 = U 2 (T = 0) = E02 .

(18)

(19)

(20)

To reproduce the thermodynamic condition (13) we must therefore subtract from


(E2 )s the quantity E02 . This gives E2 = (E2 )s E02 , or
E2 = U 2 E02 .

(21)

In combination with Eq. (11), this result leads to

dU
= E2 = U 2 E02 .
d

(22)

Now we are in possession of a differential equation for the mean energy U, which
allows to determine the function U (). An integration of the expression

d =

dU
dU
= 2
E2 (U )
U E02

(23)

subject to the appropriate condition at infinity (U as T ) yields


(
=

1
U
1
E0

for E0 = 0,
coth1 EU0

for E0 6= 0.

(24)

Although the case E0 = 0 can be treated as a limit of the case E0 6= 0, it is more


illustrative to treat each case separately. The functions in Eq. (24) can be inverted
to obtain
(
1
,
for E0 = 0;
U () =
(25)
E0 coth E0 , for E0 6= 0.

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

2.5. Mean equilibrium distribution of the oscillators


As Eq. (25) shows, the functional form of the mean energy depends critically on the
presence of E0 . For E0 = 0 the classical energy equipartition is recovered,
Ucl = 1 ,

(26)

whereas for E0 6= 0, a more complicated expression for U () is obtained. In particular, for a system of harmonic oscillators, with E0 = ~/2, Plancks law is obtained,
UPlanck = 21 ~ coth 12 ~.

(27)

The zpe is of course included, as can be seen by taking the limit T 0,


UPlanck ( ) = 21 ~ = E0 .

(28)

This establishes Plancks law as a physical result whose ultimate meaning or


cause is the existence of a fluctuating zpe, whereas its absence leads to the
equipartition of energy.
It is important to stress that Plancks law has been obtained without introducing any explicit quantum or discontinuity requirement. Equation (23), which is the
result of the recurrence relation (10) resulting in its turn from the general distribution Wg (E) and the quadratic dependence of the variance in U together with
Wiens law (which establishes the frequency dependence of the zero-point energy,
zpe), suffice to obtain Plancks law. The fact that the latter is the one that opens
the door to the zpe leads to the conclusion that Wiens law (with A 6= 0) is crucial
to obtain Plancks law and its quantum consequences. We are thus compelled to say
that Wiens law (with A 6= 0) is an extension of classical physics that enters into
the quantum domain; strictly speaking, as a precursor of Plancks distribution law
it should be considered to represent historically the first quantum law. It should
be enphazised that Eq. (27) and the ensuing consequences are of general validity,
regardless of the nature of the oscillators, provided they have a nonzero energy at
T = 0. The corroboration that the law that gave rise to quantum theory stems from
the existence of a fluctuating zpe brings to the fore the crucial importance of this
nonthermal energy for the understanding of qm and more generally, of quantum
theory.
A comment on the fluctuations of the zero-point energy is in place. We have
seen that for E0 6= 0 the thermal energy dispersion is given by
E2 (U ) = U 2 E0 2

(U = UPlanck ),

(29)

whereas in the classical case (E0 = 0),


E2 (U ) = U 2

(U = Ucl ).

(30)

Whilst in the latter case the thermal fluctuations of the oscillators energy depend
solely on its thermal mean energy, Ucl , in the former case Eq. (29) relates the thermal
fluctuations with the total mean energy UPlanck , which includes the temperatureindependent contribution. This nonthermal contribution cannot be derived from a

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

purely thermodynamic analysis as the one afforded by the distribution Wg . Therefore a statistical treatment was necessary, to arrive at the nonthermal fluctuations
given by equation (20).
2.6. Planck, Einstein and the zero-point energy
The previous discussion suggests separating the average energy UPlanck (which as
of now will be denoted simply by U ) into a thermal contribution UT and the
temperature-independent part E0 ,
U = UT + E0 .

(31)

The first term in this equation,


UT = E0 coth E0 E0 =

2E0
,
e2E0 1

(32)

is Plancks law without zpe.14, 15 At sufficiently low temperatures it takes the form
UT = 2E0 e2E0 .

(33)

This is the (approximate) distribution that was suggested by Wien at the end of the
19th century and considered for some time to be the exact law for the blackbody
distribution. In terms of (the correct) UT Eq. (29) reads
E2 = UT2 + 2E0 UT .

(34)

Equations (33) and (34) represent the germ of quantum theory, since it is precisely
on their basis that Planck and Einstein advanced the notion of the quantum (for the
material oscillators and for the radiation field, respectively), by putting 2E0 = ~.
The following comments contain a discussion of their respective points of view and
of the relation with our present notions based on the reality of the zpe. A remarkable
relationship will thus be disclosed.
It is important to note that no zpe was considered by neither Planck nor Einstein
in their analysis of Eq. (34). Instead, Planck interpreted the term 2E0 UT as a result
of the discontinuities in the processes of interchange of energy between matter and
field (more specifically, in the emissions, from 1912 onwards16 ). As for Einstein, he
interpreted this term as a manifestation of a corpuscular structure of the radiation
field, and thus pointed to it as the key to Plancks law.17 Now, from the point of view
proposed here the consideration of the zpe gives rise to an alternative understanding
of Eq. (34), namely an interpretation of the term 2E0 UT that does not depend on the
notion of quanta. The elucidation of the term UT2 as a result of the interference of
the modes of frequency of the thermal field18 suggests to interpret the term 2E0 UT
as due to additional interferences, now between the thermal field and a zero-point
radiation field that is present at all temperatures including T = 0, and whose mean
energy is just E0 . As is by now clear, in Eq. (34) there is no extra term E02 due to the
interference among the modes of the zpf themselves, because the thermodynamic
analysis made by Planck and Einstein had no room for the nonthermal fluctuations.

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

From this new perspective the notion of intrinsic discontinuities in the energy
interchange or in the field itself is unnecessary to explain either Plancks law or
the linear term in Eq. (34). By contrast, it is the existence of a fluctuating zpf
what accounts for the equilibrium spectrum. This could of course not be Plancks
or Einsteins interpretation because the zpe was still unknown at that time, even
though their results were consistent with its existence.
What is important to stress is that the interpretation made here of the linear
term 2E0 UT implies the existence of a fluctuating zero-point radiation field the
vacuum (radiation) field with a mean energy per mode of frequency equal to
E0 = ~/2. As will be remarked in section 2.8, the existence of fluctuations of this
energy is required to recover several other important characteristics of the quantum
description.
2.7. Continuous versus discrete
We have just seen how three alternative approaches provide three quite different
readings of the same quantity UT2 + 2E0 UT . In these approaches, either the (continuous) zpf or the quantization is identified as the notion underlying the Planck
distribution. Therefore the next logical step is to inquire about the relation between
the zpe and quantization. Is quantization inevitably linked to Plancks law, or is it
merely the result of a point of view, of a voluntary choice?
An answer to this question is found from an analysis of the partition function
obtained from (25). It follows by an integration of the equation (d ln Zg /d) =
U () that the partition function is
Zg =

C
.
sinh E0

(35)

The value of the constant C is determined by requiring the classical result Zg = 1


to be recovered in the limit T . This leads to
E0
Zg () =
.
(36)
s sinh E0
The constant s can be determined from the entropy, which is, up to an additive
constant and with S = Sg ,
~
kB ln(2 sinh E0 ) + kB U.
s
In the zero-temperature limit this reduces to
Sg = kB ln

(37)

~
~
kB E0 + kB E0 = kB ln .
(38)
s
s
Setting the origin of the entropy at T = 0 leads to s = ~.The partition function
takes thus the form
1
Zg () =
.
(39)
2 sinh E0
Sg (T 0) = kB ln

March
21,
Brazil*de*la*pena*et*al

10

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

2.7.1. The origin of discreteness


We now discuss the discontinuities characteristic of the quantum description, which
are hidden in the continuous description given by the distribution Wg (see Refs.
19-21 for related discussions). To this end we expand Eq. (36) and write
Zg =

X
1
eE0
=
eE0 (2n+1) ,
=
2 sinh E0
1 e2E0
n=0

Zg =

eEn ,

(40)
(41)

n=0

with
En (2n + 1)E0 = ~n + 21 ~.
The function g(E) can now be determined by means of (4b),
Z
Z X

X
Zg () =
eEn =
(E E n )eE dE,
g(E)eE dE =
0

n=0

(42)

(43)

n=0

whence
g(E) =

(E En ).

(44)

n=0

The introduction of (43) and (44) into Eq. (4a) finally determines the probability
density Wg (E),
Wg (E) =

1 X
(E En )eE .
Zg n=0

This distribution gives for the mean value of a function f (E)


Z

X
1 X
En
hf (E)i =
Wg (E)f (E)dE =
f (En )e
=
wn f (En ),
Zg n=0
0
n=0

(45)

(46)

with the canonical weights given by


wn =

eEn
eEn
= P E .
n
Z
n=0 e

(47)

The result (46) shows that the mean value of a function of the continuous variable E
calculated with the distribution Wg (E), can be obtained equivalently by averaging
over the set of discrete indices (or states) n, with respective weights wn . These
weights correspond to those of a canonical ensemble, which suggests identifying En
with the discrete energy levels of the quantum oscillators (including the zpe), as
follows from (41). Equation (46) can be recognized as the description afforded by
the density matrix for the canonical ensemble with weights wn (see e.g. Ref. 22).
Although the two averages in Eq. (46) (calculated using Wg or wn ) are formally
equivalent, their descriptions are essentially different, the first one referring to an

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

11

average over the continuous variable E, the second one to a summation over discrete
states (levels) with energies En . Because the En are completely characterized by
the states, it is natural to interpret the right-hand side of (46) as a manifestation
of the discrete nature of the energy. The mechanism leading to this discreteness,
seemingly excluding all other values of the energy, is due to the highly pathological
distribution g(E), Eq. (44). It is important to bear in mind that nevertheless the
existence of fluctuations leading to the natural linewidth23 and other processes,
effectively dilutes the discrete distribution of energies into a somewhat smoothenedout distribution acquiring a more continuous shape. Thus g(E) should be seen as a
theoretical limiting distribution, an explanation for which will be found in section
3 below.
From the present point of view, the quantization of the radiation field does not
follow from some intrinsic property, but arises as a property acquired by the field
through its interaction with matter in equilibrium. In other words, quantization is
here exhibited as an emergent property of matter and field in interaction, an idea
that is closely examined all along the present work.
2.8. A quantum statistical distribution
As was discussed in sections 2.4 and 2.5, the existence of a fluctuating zpe requires a
more general distribution than Wg , able to account for all fluctuations of the energy,
including the nonthermal contribution. Such distribution was shown to be Ws , and it
led to results that are equivalent to the quantum description, in which temperatureindependent fluctuations appear as a characteristic trait. A closer study of this
problem will help us to establish contact with one of the most frequently used
distributions in quantum statistics.
We recall that the statistical distribution appropriate to include all fluctuations
is given by Eq. (17), namely
1 E/U
e
.
(48)
U
Ws (E) maximizes the statistical entropy Ss , whereas Wg (E) maximizes the thermodynamic entropy S defined in the phase space of the particles. The crucial point
that guarantees that both distributions describe the same physical system is that
in both instances the energy distribution corresponds to maximal entropy.
Ws (E) yields for the variance of the energy at all temperatures (including T = 0)
the expression
Ws (E) =

(E2 )s = U 2 .

(49)

This result goes back to Lorentz (see Ref. 18) when applied to the fluctuations
of the thermal field. It can also be obtained by demanding that it should hold as
well for the field at zero temperature, as follows by considering that thermal and
zero-point field are part of the same creature. It is reinforced by the observation
that the fluctuations of the field arise as a result of the interferences among the

March
21,
Brazil*de*la*pena*et*al

12

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

immense number of independent modes of a given frequency, so that the centrallimit theorem24,25 applies, which means that the moments (and thus the variance)
correspond to those of a normal distribution.
As has been remarked, the specific choice U = 1 (E0 = 0) results in the usual
canonical distribution and leads to the classical expression (30). But in presence
of the nonthermal energy E0 , U is given by Plancks law and the resulting total
fluctuations are indeed (with UT given by (32))
(E2 )s = U 2 = (UT + E0 )2 = UT2 + 2E0 UT + E02 .

(50)

Therefore, the energy does not have a fixed value at T = 0, but is allowed to
fluctuate with variance E02 . As has been stressed, this term represents the nonthermal
contribution to the fluctuations of the energy.
In conformity with the present discussion the total energy can be written as
consisting of two fluctuating parts,
E = ET + E0 ,

(51)

ET and E0 being the thermal and nonthermal energies, respectively. The total energy
fluctuations are then given by the sum of three terms,
(E2 )s = E2T + E20 + 2(ET , E0 ),

(52)

where (ET , E0 ) is the covariance of the variables indicated by its arguments,


(ET , E0 ) hET E0 i hET i hE0 i.

(53)

Comparing Eq. (52) with Eq. (50), we arrive at (ET , E0 ) = 0. This shows that the
fluctuations of ET and E0 are statistically independent, as is expected due to the
independence of their sources.
The statistical entropy is given, as follows from (48), by
Z
Ss = kB Ws ln Ws dE = kB ln U + kB ,
(54)
from where it follows that (Ss /U ) = (kB /U ). A comparison with the thermodynamic entropy, which satisfies (Sg /U ) = (1/T ), shows that these two entropies
coincide only when U = kB T, i.e., for E0 = 0.
Let us now investigate how the nonthermal fluctuations become manifest in the
statistical properties of the ensemble of oscillators. The usual expression for the
energy of the harmonic oscillator (again with m = 1)
E = 21 (p2 + 2 q 2 ),

(55)

can be used as a starting point to effect a transformation from the energy distribution given by Eq. (48) to a distribution ws (p, q) defined in the oscillators phase
space (p, q). To this end we introduce the pair of variables (E, ) related to the
couple (p, q) by the extended canonical transformation26

p = 2E cos ,
(56a)
r
2E
q=
sin ,
(56b)
2

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

13

so that ws (p, q) is given by25




(E, )
,
ws (p, q) = Ws (E(p, q), (p, q))
(p, q)
the Jacobian of the transformation being


(E, ) 1
(p, q)
= 1.
=
(E, )
(p, q)

(57)

(58)

Now W (E) is a marginal probability density that can be obtained from Ws (E, ) by
integrating over the variable , so that
Z2
W (E) = Ws (E, )d.
(59)
0

For a system of harmonic oscillators in equilibrium, the trajectories (in general,


the surfaces) of constant energy do not depend on , so all values of are equally
probable, which means that
1
W (E).
(60)
Ws (E, ) =
2
Using Eqs. (55) and (57) we thus obtain for the distribution in phase space
 p2 + 2 q 2 

ws (p, q) =
W (E(p, q)) =
exp
.
(61)
2
2U
2U
This expression, which is known in quantum theory as the Wigner function for the
harmonic oscillator whenever U corresponds to Plancks law,27 can be factorized as
a product of two normal distributions,
2
2
2
2
1
1
ws (p, q) = wp (p)wq (q) = q
ep /2p q
eq /2q ,
(62)
2p2
2q2
where p2 = U and q2 = U/ 2 . The product of these dispersions gives
E2T
U2
E02
E02
~2
=
+

=
,
(63)
2
2
2
2
4
where Eq. (29) (with E2 written appropriately as E2T ) was used to write the second
equality and the value E0 = ~/2 was introduced into the last one.
Equation (63) points to the fluctuating zpe as the ultimate (and irreducible)
source of the Heisenberg inequalities. The magnitude of q2 p2 is bounded from below because of the nonthermal energy fluctuations; the minimum value ~2 /4 is
reached when all thermal fluctuations have been suppressed. Therefore, descriptions afforded by purely thermal distributions such as Wg cannot account for the
meaning of these inequalities. This result stresses again the fact that once a zpe
has been introduced into the theory, new distributions (specifically statistical rather
than thermodynamic) are needed to include its fluctuations and to obtain the corresponding quantum statistical properties. Note that the Heisenberg inequalities
should be understood as referring to ensemble averages, due to the statistical nature of (63).
q2 p2 =

March
21,
Brazil*de*la*pena*et*al

14

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

2.9. Comments on the reality of the zero-point fluctuations


The concept of a zero-point energy of the radiation field entered into scene as early
as 1912, with Plancks second derivation of the blackbody spectrum.16 Yet further to
the frustrated attempt by Einstein and Stern,28 and despite the suggestive proposal
made by Nernst in 191629 to consider this field as responsible for atomic stability,
little or no attention was paid to its existence as a real physical entity that could have
a role in the newly developing qm. Interestingly, it was the crystallographers and
the physical chemists who through fine spectroscopic analysis verified the existence
of the zpe linked however to matter, not to the field.3032
Today it is well accepted that the fluctuations of the electromagnetic vacuum are
responsible for important observable physical phenomena. Perhaps their best known
manifestations, within the atomic domain, are the Lamb shift of energy levels and
their contribution to the spontaneous transitions of the excited states to the ground
state, as will be seen in section 5. By far the most accepted evidence of the reality of
the zpf is the Casimir effect, that is, the force between two parallel neutral metallic
plates resulting from the modification of the field by the boundaries (see e.g. Refs.
33, 34 and section 5). Thus the existence of the zpf can be considered a reasonably
well established physical fact. In the following chapters we will have occasion to
study in depth the essential role played more broadly by the fluctuating zpf in its
interaction with matter at the atomic level.
3. Emergence of Quantum Mechanics
The above discussion led us to an important conclusion: the radiation field in equilibrium with matter acquires a discrete energy distribution, by virtue of its zero-point
component. This was interpreted to mean that the quantization of the field comes
about from its interaction with matter. Then, what about matter?
As our journey progresses it will become clear that also matter is so strongly
influenced by its interaction with the background field that it ends up acquiring
its quantum properties. Once again, quantization is revealed as a phenomenon that
emerges as a result of the matter-field interaction.
The material presented here is mostly based on previous work; detailed derivations and additional references can be found in Refs. 10 and 35-38.
3.1. Embarking on our journey towards the Schr
odinger equation
The main actor in this section is a charged particle immersed in the zero-point field
(zpf), performing a bounded motion under the action of an external conservative
force. In addition, the particle may be subject to some external radiation field. What
is important, however, is that the zpf is always present. To keep the exposition as
simple as possible, a one-dimensional motion is normally considered.
We start from the Abraham-Lorentz equation of motion
e
...
(64)
m
x = f (x) + m x + eE(x, t) + v B(x, t).
c

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

15

...
The term m x , with = 2e2 /3mc3 , represents the radiation-reaction force on the
particle. For an electron this force is normally small ( 1023 s). The fields E(x, t)
and B(x, t) must be represented by stochastic variables.
A number of simplifications and approximations will be introduced to go ahead.
First we simplify the Lorentz force considering that in the nonrelativistic regime
the magnetic term becomes negligible compared with the electric force, so
...
m
x = f (x) + m x + eE(x, t).

(65)

As a second step we use the long-wavelength approximation, considering that in the


region of space occupied by the particle during its motion, the electric field does
not vary appreciably. The x-dependence of E(x, t) can then be neglected, so that
one can write
...
m
x = f (x) + m x + eE(t).

(66)
E

E(t) is a stochastic variable with zero mean value, E(t) = 0. The spectral
energy density () of the field corresponds to a mean energy ~/2 per mode of
frequency , hence (see e.g. Ref. 23)
1
2 1
~ 3
() = modes () ~ = 2 3 ~ =
,
2
c 2
2 2 c3

(67)

which represents a highly colored noise. This result can also be expressed in terms
of the correlation of the Fourier transform,
Z +
1

E()
=
E(t)eit dt,
(68)
2
3
E

( 0 ) = 2~ ( 0 ).
E()
E
3c3

(69)

Equation (69) means that the Fourier components of the (stationary) zpf that
pertain to different frequencies are statistically independent.
3.2. Fokker-Planck-type equation in phase space
We write
mx = p,

...
p = f (x) + m x + eE(t),

(70)

and start from the continuity equation for the total system in terms of the density
R(x , p , t) of states in the whole phase space, where {x , p } denotes the set of
variables of both particle and field,


R

+
(x R) +
(p R) = 0.
(71)
t
x
p

March
21,
Brazil*de*la*pena*et*al

16

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

The ensuing Fokker-Planck-type equation for Q(x, p, t), which is the average of R
over the field variables, is a complicated integro-differential equation, or, equivalently, a differential equation of infinite order. A good part of the complication is
due to the memory developed as the system evolves in time. This equation reads35,39
Q
1


...
+
pQ +
(f + m x ) Q = e2 DQ,
t
m x
p
p

(72)

is
where the diffusion operator D


 2k
X

E
1 P E
Q.
D(x,
p, t)Q(x, p, t) = P E G
eG
p
p

(73)

k=0

P is the smoothing (projection) operator, which averages over the stochastic field
variables, so that


E
Q = P R = R , Q = 1 P R, . R = Q + Q.
is the operator
G

GA(x,
p) =

eL(tt ) A(
x,
p)dt0 ,

is the Liouvillian operator of the particle,a


L
= 1 p + (f + m ...
x),
L
m x
p

(74)

and
x,
p under the integral denote the values that the phase-space variables must
have at time t0 so as to evolve towards x, p at time t.
Note that, by virtue of (73), the right-hand side of the Fokker-Planck-type Eq.
(72) contains integro-differential terms of increasing order in e2 and in the energy
density of the zpf, which makes it virtually impossible to find an exact solution for
this equation.
3.3. Some relations for average values
From Eq. (72) it is a straightforward matter to derive equations for the average
values of dynamical quantities. Consider the average value of a general phase function G(x, p) that has no explicit time dependence. By multiplying (72) by G and
integrating by parts it follows that

 






d
1
G
G
G
... G
2
hGi =
p
+ f
+ m x
e
D(t) .
(75)
dt
m
x
p
p
p
a Strictly

speaking Eq. (74) is not a Liouvillian, due to the radiation reaction term, that here is
taken as an external force.

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

17

By taking successively G = x, x2 , p, p2 , xp, one obtains


d
hxi =
dt
d
2
x =
dt

1
hpi ;
(76a)
m
2
hxpi ;
(76b)
m
D
E
d
...

hpi = hf i + m h x i e2 D(t)
;
(76c)
dt
D
E
d
2
...

p = 2 hf pi + 2m h x pi 2e2 pD(t)
;
(76d)
dt
D
E
d
1
2
...

p + hxf i + m hx x i e2 xD(t)
,
(76e)
hxpi =
dt
m
whence for H = (p2 /2m) + V , the mechanical Hamiltonian function,
d
e2 D E
...
hHi = h x pi
pD(t) .
(77)
dt
m
Let us consider that a stationary state is eventually reached by the system.
Under stationarity the time derivative of the mean value of any variable is zero.
It follows that hpi = 0 and hxpi = 0, which means that x and p end up being
uncorrelated. Further, Eq. (76e) reads
D
E
1
2
...

p + hxf i + m hx x i = e2 xD(t)
.
(78)
m

2
1
Neglecting the terms of order e2 this reduces to m
p + hxf i = 0, which is the
virial theorem. Thus, Eq. (76e) is a time-dependent form of the virial theorem for
the present problem. Also under stationarity, (77) reduces to
D
E
...

m h x pi = e2 pD(t)
,
(79)
which constitutes the energy-balance condition. This is an important result that will
be used later.
D
E

From (76c) we see that any net diffusion giving rise to a term e2 D(t)
develops a net force acting on the particle; this is similar to the osmotic force in the
case of Brownian diffusion. As follows from (76d), this D
force conveys
to the particle
E

2

a net kinetic energy per unit of time of value e /m pD(t)


, in agreement with
Eq. (77).
3.4. Transition to the configuration space
At this point it is convenient to reduce the present description to the configuration
space of the particle, in order to make contact with the Schrodinger description
of quantum mechanics. The transition to configuration space can be performed
in a systematic way with the help of the characteristic function (or momentum
e
generating function) Q,
Z
e z, t) = Q(x, p, t)eipz dp.
Q(x,
(80)

March
21,
Brazil*de*la*pena*et*al

18

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

The probability density (x, t) in configuration space is a marginal probability,



Z
Z

e 0, t),
= Q(x,
(81a)
(x, t) = Q(x, p, t)dp =
Q(x, p, t)eipz dp
z=0

and the local moments of p are given by


Z
1
n
n
hp i (x) hp ix =
pn Qdp = (i)n
(x)

e
1 nQ
n
e
Q z

(81b)

z=0

Note that the moments hpn i (x) represent partially averaged


quantities, for a given
R
value of x. The fully averaged quantities are hpn i = hpn ix (x)dx.
By taking the Fourier transform of (72) and assuming that all surface terms
appearing along the integrations vanish at infinity, one gets
e
e
e
Q
1 2Q
g
e f 0 z Q = ie2 z(DQ).

(82)
i
izf (x)Q
t
m xz
m
z
It is important to observe that instead of making the transition from the phase
space to the configuration space, we could have made the transit to the momentum
space. In such case one gets
P (k, p, t) =
Pe
p
1
i k Pe +
t
m
2 p

Q(x, p, t)eikx dx,

g
k 0 , p)Pe(k 0 , p)dk 0 = e2 (DP
)(k, p),
K(k
p

(83)

(84)

0
f (x)p.
(85)
m
Things become more complicated in the p-description, since instead of a differential
equation one gets an integro-differential equation (one with long memory). In what
follows we shall continue to analyse the reduced description in the configuration
space.
A convenient procedure to get the most of Eq. (82) consists in expanding it into
a power series around z = 0, and separating the coefficients of z k (k = 0, 1, 2, . . .).
The first three equations thus obtained are
with

K(x, p) = f (x) +

1
+
(hpix ) = 0;
t
m x

(86a)

1
2

^
(hpix ) +
( p x ) f f 0 hpix = e2 (DQ)
;
(86b)
t
m x
m
z=0


2 
1
3


^
p x +
( p x ) f hpix f 0 p2 x = 2e2 (DpQ)
; (86c)
t
m x
m
z=0

The subsequent equations (corresponding


to higher powers of z) are connected to
the above by the same elements , hpix , p2 x , . . . in addition to contributions

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

19

^ The entire set of these equations constitutes thus


deriving from the term z (DQ).
an infinite hierarchy of coupled nonlinear equations.
The first member of the hierarchy is the continuity equation, which describes the
transfer of matter in configuration space. It follows that the mean flux of particles
is j(x) = (x) hpix /m, hence the local mean velocity is given by
v(x) =

j(x)
1
=
hpix .
(x)
m

(87)

The function v(x) refers to a locally averaged quantity; it differs essentially from
the instantaneous velocity of one (specific) particle that visits the neighborhood of
x at time t, and varies randomly from one particle to another. To avoid confusion
we call v(x) local mean velocity or flux (flow ) velocity.
The second equation describes the transfer of momentum, or equivalently, the
evolution of the current density j(x,

t)
= v(x, t)(x, t). It contains, in addition to
and hpix , the second moment p2 x , whose value is determined by the third
equation, the one describing the transfer of kinetic energy (up to a factor 1/2m).
These moments are
!


Z
e

1 Q

1
e

,
(88a)
pQdp = i
=
i
ln
Q
hpix =


e z
(x)
z
Q
z=0
z=0
!

 2
2

e

1 2Q

e
e
=

.
(88b)
p x=
ln
Q
ln
Q



2
2
e z


z
z
Q
z=0
z=0

z=0

Combining these expressions gives

2
2
p x hpix =



2
e
.
ln Q
z 2
z=0

(89)

With the change of variables (x, z) (z+ , z ),


z+ = x + z,

z = x z,

(90)

where is a parameter with dimensions of action, whose value remains to be fixed,


the above expressions transform into


e
hpix = i + ) ln Q
,
z=0

2
2
p x hpix = 2





2
2
e
e
ln
Q
+
4

ln
Q
.
+


x2
z=0
z=0

(91)

e in the general form


Writing Q
e + , z , t) = q+ (z+ , t)q (z , t)(z+ , z , t)
Q(z

(92)

March
21,
Brazil*de*la*pena*et*al

20

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

e + , z ), and taking into


where represents the nonfactorizable contribution to Q(z
account (81a), one gets
hpix = i
with

q+ (x, t)
ln
+ g,
x q (x, t)

g = i ( + ) ln |z=0 ,

(93a)
(93b)

and

2
2
2
p x hpix = 2 2 ln + ,
x
2
with = 4 (+ ln )|z=0 .

(94a)
(94b)

The first two equations of the hierarchy, generalized to three dimensions, become
thus (a sum over repeated indices is understood)

+
(vj ) = 0,
t
xj



2
1
2

(vi vj )
ln +
ij fi =

m (vi )+m
t
xj
m xj xi xj
m xj


fi
g
i
= vi
e2 (DQ)
.

xj

(95a)
(95b)

z=0

These equations will be analysed in the following sections.


e (x, z, t) = Q(x,
e z, t), so that from (92) (we
Note that according to Eq. (80), Q
come back to a three-dimensional description for convenience),

q+ (z , t) = q
(z , t),

q (z+ , t) = q+
(z+ , t),

(z+ , z , t) = (z , z+ , t). (96)

e can be rewritten therefore as


Q
e + , z , t) = q(z+ , t)q (z , t)(z+ , z , t),
Q(z
where

q(z+ , t) q+ (z+ , t),

q (z , t) q+
(z , t).

(97)

(98)

Further, from (81a) and (97) it follows that one may write
e 0, t) = q (x, t)q(x, t)0 (x, t),
(x, t) = Q(x,

(99)

with 0 (x, t) = |z=0 a real function that can be taken as a constant without loss
of generality, absorbing its possible time and space dependence into the functions
q(x, t), q (x, t). We therefore write
0 (x, t) = 1,

(x, t) = q (x, t)q(x, t).

(100)

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

21

3.5. The Schr


odinger equation
The Fokker-Planck-type Eq. (72) or its equivalent in (x, z)-space, Eq. (82)
contains essentially two kinds of terms: the first three originate in the Newtonian
part of the Liouvillian, and the last two originate in the matter-field interaction and
describe the dissipative and stochastic (diffusive) behavior of the particle, respectively. The latter gave rise to the terms on the right-hand side of Eq. (95b). The
first one, proportional to (hence to e2 ), is due to the radiation reaction and has
a dissipative effect over the motion. The second one (also proportional to e2 , plus
higher-order terms) describes a permanent fluctuating action over the motion.
We assume that the systems of present interest reach in the time-asymptotic
limit, when the irreversible processes involving the radiation field have disappeared
a state of balance between the mean absorbed and radiated powers, as expressed
in Eq. (79). This is to be expected for all bound systems, but we leave this point
open to further consideration. Under such condition the two radiative contributions
essentially cancel each other in the mean. The fact that both terms are proportional
to e2 means that their remaining contribution should represent radiative corrections
and can therefore be neglected in a first approximation. We describe here this situation of energy balance in the radiationless approximation, and leave the detailed
discussion about the energy-balance condition for subsection 3.5.2. The first two
equations of the hierarchy reduce then to

+
(vj ) = 0,
(101a)
t
xj



2
1
2
m (vi )+m
(vi vj )
ln +
ij fi = 0. (101b)

t
xj
m xj xi xj
m xj
A series of algebraic manipulations allows us to recast them as


g
1

Mq +
v ( g) +
[ ()]i = 0

xi
t
m
i

(102)

and its complex conjugate, with


q 1 (2i + g)2 q + V q 2i q .
M
2m
t

(103)

The functions gi and ij are defined through Eqs. (93b) and (94b). Since according
acts over a probto Eq. (100), q (x, t)q(x, t) = (x, t), we see that the operator M
q = 0, Eq. (103) becomes the Schrodinger equation for
ability amplitude. When M
q(x, t) (x, t), containing the as yet undetermined parameter and the vector
function g(x,t),
1

2
(2i + g) + V = 2i
,
2m
t

(104)

(x, t) = (x, t) (x, t)

(105)

and

March
21,
Brazil*de*la*pena*et*al

22

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

corresponds to the Born rule. Equation (104) requires that the remaining terms in
Eq. (102) cancel each other, i.e.,


1
g
v ( g) = [ ()] .
(106)
t
m
In other words, Eq. (104) is the Schrodinger equation with minimal coupling (with
g = ec A), provided that: a) the parameter has the value }/2, and b) condition
(106) holds.
The determination of the parameter is made below, where it is shown that
indeed, its value is given correctly by }/2. As to the functions gi and ij , we note
from (93b) and (94b) that they depend on the first and second derivatives, respectively, of the (nonfactorizable) function (z+ , z ), evaluated at z = 0. The function (z+ , z , t) cannot be found exactly without solving the complete hierarchy of
equations in configuration space. For the time being we shall make some reasonable
assumptions that allow us to move ahead; once we have learned the meaning of
condition (106) we may attempt to study the general (unconditioned) case.
3.5.1. Factorizable case
e + , z , t) is
Let us consider the case in which the coupling function g = 0, i.e., Q(z
factorizable to first order in z. Then it follows that
() = 0.

(107)

Equation (104) reduces, with (x, t) = q(x, t), to

2 2 2
+ V = 2i
,
m
t

(108)

and the local momentum is given by


hpix = mv(x) = i ln [(x, t)/ (x, t)] .

(109)

A general solution to Eq. (107) is


=

K
,

(110)

where K is a divergencefree second-rank tensor. From Eq. (106) we verify that the
quantity determining the function that enters into the phase-space distribution
(Eq. (97)), is related to radiative corrections. On the contrary, the quantity in
Eq. (110) is not necessarily a small radiative correction. Thus,

it can be 2taken to
correct, when necessary, the negative value of the quantity p2 x hpix , which
occurs in some quantum mechanical problems. Of course, this last requirement is
not
part
of the usual formal baggage of quantum mechanics because functions such
as p2 x do not belong to the Hilbert-space formalism, and (both parts of) the
function are unknown to that theory.

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

23

Notice from Eqs. (101) that with = 0 we are left with a couple of nonlinear
equations for (x, t) and v(x, t),which is uncoupled from the rest of the hierarchy
and is entirely equivalent to Schrodingers equation:

+
(vj ) = 0,
t
xj



2
2

(vi vj )

ln fi = 0.
m (vi )+m
t
xj
m xj xi xj

(111a)
(111b)

The term that contains the ln is due to the momentum fluctuations (transcribed to
configuration space by the reduction process). It is nonlocal in nature even in the
single-particle case due to its dependence on the distribution of the particles in
the entire space. It encapsulates, as will become clear below, the quantum behavior
of matter, including quantum fluctuations and the characteristic apparent quantum
non-local effects. Since the source of the momentum fluctuations is the zpf, it follows
that the quantum fluctuations are conventional fluctuations with a causal origin.
3.5.2. Detailed energy balance: Determining
We now focus on the energy-balance condition (79)
D
E
...

m h x pi = e2 pD(t)

(112)

and use it to determine the value of . According to this equation, energy balance is reached when the average power dissipated by the particle along its orbital
motion (the left-hand side) is compensated by the average power absorbed by the
fluctuations impressed upon the particle by the random field along the mean orbit.
The statistical description of the system is now given by Eq. (108), so we use
it to carry out the calculations, to lowest order in e2 . The mean values on both
sides of Eq. (112) are calculated for the particle in its ground state (0 ) and the
background field also in its ground state, with spectral energy density given by (67)
0 () =

~ 3
.
2 2 c3

(113)

The left-hand side of (112) gives


...
2
4
m h x pi0 = mk 0k
|x0k | ,
(114)
R
with 0k = (E0 Ek ) /2 and x0k = 0 xk dx, where Ek are the energy eigenvalues
and k are the corresponding eigenfunctions of Eq. (108). To calculate the righthand side we write to lowest order in e2
Z
0
4 2

e2 D(t)Q
=
e d dt0 0 () cos (t t0 )eL(tt ) Q(t0 ).
(115)
p
3
p
After multiplying by p/m and integrating over the phase space we obtain
Z t
e2 D E }
pD =
d 3
dt0 cos (t t0 ) I(t t0 )
m
0
0

(116)

March
21,
Brazil*de*la*pena*et*al

24

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

with
0

I(t t ) = dx

dp p 0 Q(x0 , p0 , t0 ) =
p
0

p
p0


,

(117)

where p0 = p(t0 ) evolves towards p(t) and the subindex 0 indicates that the mean
value of the propagator p/p0 is to be calculated in the ground state. Using the
solutions of Eq. (108) we thus write


p
1
mX
2
0
=
h[
x
,
p

]i
=
k0 |x0k | cos k0 (t t0 ),
(118)
0
p0 0
2i
2
k

which inserted in Eq. (116) gives after integrating over time, in the time-asymptotic
limit (t ), the result
}m X 4
e2 D E
2
0k |x0k | .
pD =
m
2
0

(119)

Equating this with (114), we obtain for the value


=

}
.
2

With this result for , Eq. (108) coincides exactly with Schrodingers equation
i~

}2 2

=
+ V .
t
2m

(120)

This is, then, the door through which Plancks constant enters into the quantum description. Note that the balance condition is seen to be satisfied not only
globally, but frequency by frequency, for every 0k , so that it reflects a condition
of detailed energy balance between particle and field, which means that the spectral density of the vacuum field at equilibrium remains unaffected. In fact the zpf
with spectrum 3 is the single one that guarantees detailed balance and leads
to equilibrium with the ground state of the material subsystem. This radically departs from the classical situation, in which detailed equilibrium is reached with the
Rayleigh-Jeans spectrum, proportional to 2 , as was established by van Vleck almost a century ago.40,41 The structure of the Rayleigh-Jeans spectrum is tightly
linked to the Maxwell-Boltzmann distribution, so the above results confirm that the
quantum systems obey a nonclassical statistics.
It is clear that, in general, the equality (112) can hold only for a selected set
of mean stationary motions. This discloses the mechanism responsible for quantization: the atomic stationary states are those for which the equality holds. Such
demanding energy equilibrium can hold only for certain orbital motions. Equation
(112) explains thus why atoms reach and maintain their stability. Thanks to the existence of the zpf it becomes possible to explain this stability: the electrons radiate
to the field, but at the same time they absorb energy from it.

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

25

3.5.3. Local velocities


From the discussion in sections 3.1 and 3.2 it becomes clear that the particles follow
(stochastic) trajectories, so the study of such trajectories becomes an important
subject for the theory, at least because their knowledge should help to get a better
inkling on the quantum behavior of matter and its motions. As a complement to
the above discussion, notice that in parallel to the (local) current velocity v =
hpix /m one may introduce a velocity u associated with the mean local value of the
momentum fluctuations (cf. Eq. (94a)),


~ ln
~
1
1
u=
=
+
.
(121)
2m x
2m x
x
This stochastic velocity (it disappears if the zpf is disconnected) has some place
in usual qm, although it goes largely unnoticed. Indeed, from combining Eqs. (87)
and (121) it follows that

= p = m(v iu).
(122)
x
We see that an application of the quantum operator i~/x gives more than just
the expected current velocity. It also carries with it information about the diffusion
taking place in the momentum subspace (projected onto the configuration space),
in the form of an additional effective momentum mu (affected by the imaginary
unit, i). By taking into
that the mean values satisfy hv 0 i = (2m/~) huvi

2account

0
and hu i = (2m/~) u we arrive at


p /2m = 21 m v 2 + u2 .
(123)
i~

This result shows that the two velocities contribute symmetrically to the mean value
of the kinetic energy. From this point of view they are equally important, despite
the concealed presence (or patent absence) of u in usual quantum theory. Being
associated with diffusion, the velocity u is the counterpart of the osmotic velocity
characteristic of Brownian processes.
Consider
a stationary state with v = 0, so its mean kinetic energy is given by

2
1
m
u
.
Since
from (121) one gets
2

2

2
~2

~2
3
1
m
u
=

d
x
=

2 ln ,
(124)
2
8m

8m
the total mean energy of the system is given in this case by


Z
}2 2
hHi = (x)
ln

+
V
(x)
dx.
8m x2

(125)

Under the demand that Rthis mean energy acquire an extremum value under conservation of probability, (x)dx = 1, with = 2 , this variational problem has as
solution the Euler-Lagrange equation42

}2 2
+ V = E,
2m x2

(126)

March
21,
Brazil*de*la*pena*et*al

26

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

where E is a Lagrange multiplier. This result emphasizes the remarkable role played
by the averaged local dispersion of the momentum, Eq. (124): it leads directly to the
(stationary) Schr
odinger equation and guarantees that the stationary probability
distribution of particles corresponds to an extremum (normally a minimum) of the
mean energy of the system. These extrema of the mean energy correspond to the
quantized solutions.
The observation that the energy of the stationary states is a local minimum is
very suggestive. Of course every particle finds eventually its own specific (stochastic)
trajectory, with a certain mean energy (averaged over the trajectory) and more or
less stable. The energy over the ensemble of such trajectories acquires a minimum
mean value, which corresponds to the set of most robust motions and generates a
certain spatial (probability) distribution of particles. A first image that comes to
mind is that the orbits are trapped once they are close enough to the stationarity
condition. This picture will become more transparent below, with the demonstration
that the stationary states correspond to situations that satisfy an ergodic condition.
The trapping of the orbits is a normal occurrence, at least in the long run (i.e. the
time-asymptotic limit), because the stochastic field compels the particle to explore
the neighboring phase-space regions; as long as the particle is not trapped, it will
continue probing the phase space. This means that the stationary orbits are qualitatively analogous to limit cycles, although here the attraction basin is formed by the
lowest average energy. Similar possibilities have been suggested earlier by several
authors.4345
3.6. Quantum potential and nonlocality in the single-particle case
By writing the wave function in its polar form

= eiS/~ .

(127)

with S(x, t) real and S/~ dimensionless, the Schrodinger Eq. (120) gives, after some
elementary simplifications (it is advantageous to use here 3-D vector notation)
S
i~
1
2
+
=
(S) + V +
t
2 t
2m
~2
~2 2
i~
i~ 2
2
+
()

S
S.
(128)
2
4m
2m
2m
2m
To disentangle this expression we write the continuity equation in the form

1
1
= (v) = 2 S S .
(129)
t
m
m
Equation (128) reduces then to

S
1
~2
~2 2
2
2
=
(S) + V +
,
()
2
t
2m
4m
2m

S
1
~2 2
2
or
+
(S) + V
= 0.

t
2m
2m

(130)

(131)

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

27

The point with this result is that if one defines an effective potential acting on the
particles by means of

~2 2
Veff = V + VQ , VQ =
(132)
,
2m

where VQ is known as the quantum potential (see e.g. Ref. 46; some authors call it
Bohms potential), then (131) takes the form of a Hamilton-Jacobi equation for the
principal function S,
S
1
2
+
(S) + Veff = 0.
(133)
t
2m
This procedure is effectively followed in certain occasions, particularly in the causal
de Broglie-Bohm interpretation of qm.46 Whereas the rest of the terms in Eq. (131)
are local, the function VQ (x) bears information (through ) about the distribution
of all the members of the ensemble in the entire configuration space. Therefore Eq.
(133) differs essentially from a real Hamilton-Jacobi equation, which, by definition,
describes the motion of a congruency of (single) particles acted on by local potentials. (A congruency refers to a single-valued trajectory field.) With the addition of
VQ , (133) refers to an ensemble of similar particles; it acquires a statistical meaning described in configuration space by (x). By becoming nonlocal, its meaning
becomes obscure. Assume for instance that each member of the ensemble is a single particle, separated by any other member by hours, or days, and that the same
experiment is performed a huge number of times until the ensemble is well approximated. Under such circumstances there is clearly no opportunity for any nonlocal
physical effect to take place among the members of the ensemble. This nonlocality
is not ontological, but rather a semblance, an artifact of the reduced statistical description that vanishes by going back to the full phase-space statistical description,
which is as local as any true statistical description can be.
According to (121) and (132), the mean value of VQ is equal to


(134)
hVQ i = 21 m2 u2 .
Therefore another form of expressing the mean energy is, using Eq. (123),
D E

2
1
2
= 1 p
H
+ hV i =
v + hVQ + V i .
(135)
2m
2m


Here the kinetic energy of diffusion m2 u2 /2 becomes reinterpreted as a quantum
potential energy hVQ i . This is a usual translation in the literature, clearly legitimate
for the mean values, although somewhat misleading for the functions VQ and mu2 /2
themselves.
The fact that the most extensive line of research on quantum nonlocalities during
the last decades is related to the Bell inequalities, has led to the widespread conviction that nonlocality is a property exclusive of multipartite quantum systems. It
should be stressed that independently of interpretation, the Schrodinger equation
contains the quantum potential even if disguised and hence the associated

March
21,
Brazil*de*la*pena*et*al

28

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

quantum nonlocalities. These are present in all cases, including the single-particle
one. As for multiparticle systems, extra nonlocalities arise due to the correlations
among variables that pertain to different particles, particularly for entangled states.
These extra nonlocalities are responsible for the violation of the Bell inequalities.
The mechanism that gives rise to the entanglement between the components of a
pair of noninteracting particles within the present theory is discussed in section 4.5
3.7. Phase-space distribution; Wigners function
Our starting point for the derivation of the Schrodinger equation in configuration
space was the phase-space Fokker-Planck-type Eq. (72). We have also at our disposal
the momentum representation of the Schrodinger equation. Is it then possible to
proceed in the opposite sense, starting from the usual quantum description provided
separately in configuration or momentum space, and recover a unique full phasespace description? That this question cannot be answered in the positive is a wellestablished fact. Let us briefly look into the matter from the present perspective
and disclose the reason for this difficulty.
e z, t) can be inverted and combined with Eqs. (90) and
Equation (80) for Q(x,
(92) to obtain
Z
1
e z, t)eipz dz =
Q(x,
(136)
Q(x, p, t) =
2
Z
1
=
q+ (x + z, t)q (x z, t)(x + z, x z, t) eipz dz.
(137)
2
By construction Q(x, p, t) furnishes a true (Kolmogorovian) probability density in
phase space. This means that if the exact solutions for q+ , q and were known
for all values of z, we would have a full phase-space description for the particle.
However, all we can construct is an approximate form W (x, p, t) obtained by taking
= 1 and allowing z to remain as a Fourier variable, which gives
Z
1
(x + y, t) (x y, t)ei2py/~ dy,
(138)
W (x, p, t) =
~
with y = z = ~z/2. This is the well-known Wigner phase-space function.47,48
As a result, we cannot guarantee W to be a true Kolmogorovian probability. And
indeed, despite its recognized value, it is not, since as is well known it can take on
negative values in some regions of phase space for almost all states and systems
(the exception being the Gaussian states). The right solution to this long-standing
problem is of course to recognize the intrinsic limitation of W that ensues from its
approximate nature, and to revert to the full distribution Q(x, p, t).
As was shown above, the assumptions and approximations made to arrive at
the Schr
odinger equation imply an enormous simplification of the problem at hand,
since only the first two equations of the infinite hierarchy are used. But this comes
with a high price: the loss of a true phase-space description. We conclude that a
most important problem that remains open is the investigation of the possibilities

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

29

offered by the full phase-space probability (136). It is clear that its study opens a
wide door to some new physics.
4. Heisenberg Quantum Mechanics
The findings presented in the previous section suggest to explore the Heisenberg
formulation of quantum mechanics from a new perspective. We achieve this by
introducing the principle of ergodicity of the stationary solutions of the particlezeropoint field problem, which ultimately leads to the matrix description. The response of the particle is found to be always linear in the field components, regardless
of the nonlinearities of the external force. This is the essence of linear stochastic
electrodynamics (lsed).3637,4951
The path followed here is perhaps less intuitive than the one presented in the
preceding sections, but at the same time it is more revealing and illustrative of
several of the intricacies of qm. It serves to disclose different and to some extent
complementary features of the physics underlying the quantization process, that
remain hidden in the Schr
odinger formalism. Moreover, the extension of the theory
developed to the case of two particles allows for a clear physical understanding of
the mechanism leading to quantum entanglement.
4.1. Resonant solutions in the stationary regime
As before, we start from the approximate (non relativistic, long-wavelength) equation of motion
...
m
x = f (x) + m x + eE(t).
(139)
The stationary solutions of this equation, xstat (t), can be decomposed into a timet
independent contribution, which coincides with the time average (() ) defined as
Z
t
1 T stat
xstat (t) = lim
x (t)dt,
(140)
T T 0
plus an oscillatory contribution that averages to zero. We thus express xstat (t) in
the form
X

t
xstat (t) = xstat (t) +
x
k ak eik t + c.c. ,
(141)
k6=0

where we have excluded from the sum the term corresponding to the null frequency,
0 = 0, and x
k represents an amplitude (stochastic in principle) associated with
the frequency k . We now define
x(t) as
X

x(t) =
x
k ak eik t + c.c.
(142)
k
stat

With this definition


x(t) and x (t) differ only in their time-independent term,
t
t
with
x(t) = x
0 a0 +c.c.= 2xstat (t) .

March
21,
Brazil*de*la*pena*et*al

30

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

We assume that the external force f (x) can be expanded as a power series of x
in the form f (x(t)) =n=1 cn xn (t), and that once the stationary state is reached it
can be decomposed in a form analogous to (141)

X
t
fk ak eik t + c.c. .
(143)
f stat (t) = f stat (t) +
k6=0

Let us now define


f =

fk ak eik t + c.c.,

(144)

whose relation to f stat (t) is analogous to that previously defined between


x(t) and
xstat (t). The quantities as fk or x
k depend in general on the set of variables {aj }
for non-linear forces. In other words, neither Eq. (142) nor (144) are explicit developments in the field variables ak ; yet the time dependence in both equations is fully
expressed through the factors eik t . Introducing Eqs. (142) and (144) into (139)
leads, after separating the terms that oscillate with frequency k ,to
k eik t ,
mk2 x
k eik t = fk eik t im k3 x
k eik t + eE

(145)

where we used the expansion


E(t) =

k ak eik t + c.c.
E

(146)

for the component of the electric field in the direction of motion. Therefore,
x
k =

k
e E
,
m k

with

k k2 i k3 +

fk
.
m
xk

(147)

Thus the important contributions to


x(t) come from the poles of x
k , i.e., they
correspond to those frequencies that satisfy
k2

fk
.
m
xk

(148)

For frequencies in the atomic range, the resonances are extremely sharp due to the
small value of (recall that 1023 s for electrons). We will denote the set of
solutions of Eq. (148) as {k }res and refer to its elements as resonance frequencies.
Because of the stochastic nature of the background field, the solutions of Eq.
(i)
(139) constitute a stochastic process xstat (t), where the index (i) signals the dependence of xstat (t) on the field realization (i). When the set {i} of all the realizations of
the field is considered, an ensemble of single-particle solutions is determined. Thus
the statistical set can be reproduced by considering an ensemble of particles, each
of which is subject to a different realization of the field. In other words, the averages
(i)

over the ensemble of realizations of the field (denoted as () ) can alternatively be


determined by averaging over the ensemble of particles. Now, according to the above
discussion, in the long run the particles reach stationary states that can be labeled
with an index (which in its turn, being a one-dimensional problem, is in direct

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

31

correspondence with the mechanical energy E ). To distinguish among these different stationary states accessible to the mechanical subsystem, we decompose the
ensemble {i} into subensembles, {i} = {i } , such that the energy corresponding
to those particles subject to the field realization i {i } is precisely E .
In what follows we focus our attention on a particular subensemble {i } , and
construct the appropriate expansions for the dynamical variables corresponding to
such subensemble. The frequencies that play an important role in the dynamics of
the particle (to be determined by the theory itself) constitute then an -dependent
subset of all the resonance frequencies {k }res . We refer to these as relevant frequencies and denote them by , where serves to enumerate the different frequencies
of the set, so that when varies the coincide with the different (resonance)
frequencies of the solutions
x(t) in a state characterized by .
Now, the fact that the amplitudes x
k , fk (or more generally Ak ) and the variables
ak in expansions such as (142) and (144) are associated with the frequency k , leads
us to introduce the same couple of indices ( and ) with the meaning already
, x
explained for the frequencies in such quantities, so that we write E
,
k , x
f and a instead of E
k , fk , and ak , respectively. In this way, the transit
from a description referred to the complete ensemble ({i}) to one restricted to
A
(t),
the subensemble {i } is achieved by performing the substitutions: A(t)
k , and ak a in expansions of the type (142). Thus, for the subsystem
that has attained the stationary state we write,
X

x (t) =
x
a ei t + c.c.,
(149)

. The can acquire positive or negative values, so


and similarly for f and E
the present expansions, at variance with those of the form (142), are not necessarily
in terms of positive and negative frequencies. The transition from k to requires
therefore a reordering of the terms in the sum.
For the subensemble characterized by the (mean) energy E , we are led to rewrite
Eq. (139) as
m

d3
x
d2
x
.
= f + m
+ eE
2
dt
dt3

(150)

are introduced, Eq. (145) rewritten for the


When the expansions for
x , f and E
state becomes

X
X
2
3
a ei t + c.c.
m
x
a ei t + c.c.=
f im
x
+ eE

(151)
Assuming that detailed balance is satisfied for each relevant frequency , we are
led to
m

3
d2 x
(t)
(t)
(t) + m d x
(t),
=
f
+ eE
dt2
dt3

(152)

March
21,
Brazil*de*la*pena*et*al

32

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

with
A (t) A ei t .

(153)

As for the terms oscillating with the factor ei t , they give rise to an equation
that is the complex conjugate of (152). From this latter we obtain
x
=

e E
,
m

2
3
=
i
+

f
.
m
x

(154a)

This means that the mechanical system responds resonantly to those frequencies that solve, in analogy with Eq. (148), the (approximate) system of equations
f

mx .The relevant frequencies that satisfy such equations are precisely the
resonance frequencies associated with the subensemble , which constitute a subset
of {k }res . As before, quantities such as x
, f , , etc. depend in principle on
the set of stochastic amplitudes {a} .
The fact that Eq. (150) decomposes into two equations, namely (152) and its
complex conjugate, allows us to restrict the study to the solutions of (152) only.
This latter is but the detailed (term by term) form of the equation
...
m
x = f + m x + eE ,
(155)

where we have defined generically the complex quantities


X
A (t) =
A a ei t .

(156)

In what follows we work with Eq. (155) which is in direct correspondence with
the original equation of motion and with expansions of the form (156).
4.2. The principle of ergodicity
Being the particle subject to the permanent interchange of energy with the random
field, it appears natural to consider that once in the stationary regime, the system
satisfies an ergodic principle, so that the time average of a function A(i) (t) coincides
with its ensemble average. In this and the following sections we shall explore the
consequences of introducing the demand of ergodicity.
We decompose x (t) into its time-independent contribution (which coincides
t
with x (t) ) plus an oscillating part that averages to zero,
X
x (t) = x
a +
x
a ei t ,
(157)
6=

with = 0.The second term in this expression corresponds to the deviations of


x (t) from its mean value, and its modulus allows us to calculate the variance x2
defined as

t
t 2

x2 = x x (t) .
(158)

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

33

Direct calculation gives


x2 =

|
x a | .

(159)

6=

According to the discussion in subsection 4.1, x (t) depends on the specific


realization (i) {i } of the field. Therefore, x
, the set { }, and clearly a
also depend on the field realization. We therefore add the superindex (i) to these
stochastic quantities in order to stress their dependence on the field realization.
Thus, x2 is expressed in an explicit stochastic form as
X (i) (i) 2
x a .
x2(i)
=
(160)

6=

x2

By construction,
is time-independent, although it depends on the specific realization of the field (i). On the other hand, the variance obtained by averaging over
the ensemble of realizations,
x2 (t)


(i)
(i) 2

= x x (t) ,

(161)

may depend in general on t, but not on i, by definition. At this point we introduce the
ergodic hypothesis. This means that both expressions for x2 , Eqs. (160) and (161),
coincide. Under this condition the right-hand side of (160) must be independent of
the realization; that is, the ergodic condition implies that
X (i) (i) 2
x2 =
x a is independent of (i).
(162)

(6=)

Since the right-hand side of this equation is a sum of statistically independent terms,
each one being a non-negative quantity, the only possible way for condition (162) to
be satisfied in general is that each term of the sum is independent of the realization,
so that



(i) 2 (i) 2
is independent of (i) (for 6= ),
(163)
x a

(i)

(i)

whence the moduli of both x


and a are independent of (i), and their polar
(i)

(i)

forms become x
= ei ,

(i)

(i)

(i)

a = r ei . Since x
always appears next to

(i)

a , any random contribution contained in the phase of x


can be transferred to
(i)

the (random) phase of a , which allows us to redefine


(i)

x
= x
,

0(i)

(i)

a r ei

( 6= ).

(164)

0(i)

The new a represents the value acquired by the stochastic field variable through
its interaction with matter. As for the term with = , we revert to the demand
of ergodicity on the trajectory x (t), which requires that
(i)
x t = x
(i)
a = x (i) .
a = x

(165)

March
21,
Brazil*de*la*pena*et*al

34

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es
(i)

Hence the above equations are also valid for = , with = = 0. In


(i)
0(i)
summary, our results are (with a instead of a , for simplicity)
(i)

(i)

a = r ei ,

= 0

(i)

x
= x
.

(166a)
(166b)

Now the variance of x takes the form


X (i) (i) 2 X
2 2
x a =
|
x | r
.
x2 =

(167)

6=

6=

Following a similar procedure one can calculate the variance of the momentum
(i)
p . This allows to conclude that = , so that all the stochasticity has been
(i)

absorbed into the phases of the a . Neither x


nor depend now on (i), which
means that also
f is independent of (i).

(168)

This is a most remarkable outcome of the principle of ergodicity, whose meaning


and implications are examined below.
The fact that the x
, f and are nonstochastic variables in the present
are also independent of
approximation means that the quantities and E
(i). Hence the Fourier developments above are in fact explicit developments in the
(i)
variables a , and further,
x(i)
(t) =

(i)
e XE
a ei t + c.c.
m

(169)

a(i) .
has become a linear function of the stochastic components of the field, E

For this reason, the theory that ensues as a result of the condition of ergodicity is
termed Linear Stochastic Electrodynamics (lsed).
4.2.1. Quantization as an outcome of ergodicity
To expand the force f (or more generally, any dynamical variable A (x)) in terms
of the amplitudes x
that appear in the expression for x ,
X
x (t) =
x
a ei t ,
(170)

we must construct the expansions that correspond to the different powers of x for
a given state . For
for the quadratic case, the most immediate option is
 example,
2
2
to write f = x = (x ) . In this case we get


2 X X
(i)
(i)

f(i) = x(i)
=
x
0 x
00 a 0 a 00 ei(0 +00 )t .
(171)

00

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

35

(i)

On the other hand, according to Eq. (143), f has the form


(i) X
f2 0 a(i) 0 ei0 t ,
x
f(i) = x2 =

(172)

f2 0 is still to be determined. A comparison between


where the quantity f 0 = x
Eqs. (171) and (172) leads to
X
(i) (i)
f2 0 a(i) 0 ei0 t =
x
0 x
x
00 a 0 a 00 ei(0 +00 )t

00
(i)

(i)
= x
(t)
x 0 a 0 ei0 t ,

(173)

f2 0 = x(i)
which means that one should make the identification x
x 0 .This ex (t)
f
pression is inconsistent with Eq. (168), which indicates
that neither x2 0 nor x
0

2
2
depend on i. The result shows that the election x = (x ) , in spite of being the
most natural, is inconsistent with the implications of the ergodic principle. Consequently the problem of determining the expansion of a given power of x requires a
more careful analysis. The rather lengthy procedure of such analysis can be seen in
Ref. 37, where it is shown in detail that in order to guarantee consistency with the
ergodic principle one must write
X
X

f2 a ei t =
x2 =
x
x
0 x
0 a 0 a 0 ei(0 +0 )t .
(174)

, 0

Equation (174) goes hand in hand with the following conditions on the stochastic
variables and the relevant frequencies (no summation over repeated indices!)
a 0 a 0 = a ,

(175)

0 + 0 = .

(176)

The relation (175) is easily generalized to any number of factors by a successive


(chained) application of it,
a 0 a 0 00 a 00 000 a (n1) = (a 0 a 0 00 ) a 00 000 a (n1)
= [(a 00 ) a 00 000 ] a (n1)
= [a 000 ] a (n1)
= a .

(177)

With each a (n) (m) written in polar form according to (166a), Eq. (177) can be
broken down into the couple of equations
0 + 0 00 + ... + (n1) = ,

(i)

(i)

(i)

(i)

(178a)

r 0 r 0 00 r 00 000 r (n1) = r .

(178b)

Since the number of factors on the left-hand side of this last equation is unrestrained,
its only (nontrivial) solution is
r = 1,

, .

(179)

March
21,
Brazil*de*la*pena*et*al

36

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

In its turn, the general solution to Eq. (178a) is


(i)

(i)

= (i)
,

(180)

where each of the represents a random phase. Combining Eqs. (179) and (180)
with (166a), we get
a = ei = ei( ) ,

(181)

whence a = a . Similarly, Eq. (176) can be generalized to an arbitrary number


of terms,
0 + 0 00 + ... + (n1) = ,

(182)

in analogy with (178a), and the general solution is therefore of the form
= .

(183)

The relations (177) and (182), which are fundamental for the theory, constitute
the chain rule. The frequencies appearing in the chain rule are the relevant frequencies, which are either frequencies of resonance or linear (chained) combinations of
them, according to (182). With regard to the final phases of the field, Eq. (178a)
tells us that those pertaining to the relevant modes become partially correlated,
which confirms that not only the material part, but also the near background field
is affected during the evolution of the complete system towards equilibrium.
4.2.2. The matrix rule
We now use the chain rule to recast (174) into the form

X
X X
f2 a ei t =

x
x
0 x
0 a ei t ,

(184)

2 is given by the sum


which shows that xg

2 =
xg

x
0 x
0 ,

(185)

embodying the rule for matrix multiplication applied to the amplitudes x


. Thus
x
can be identified with the element of a square matrix x
such that

2 = x
xg
2 ,
(186)

which leads, together with Eq. (185), to recast Eq. (172) into

X X
X



x2 =
x
0 x
0 a ei t =
x
2 a ei t .

(187)

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

37

It is now easy to verify that Eqs. (177) and (182) allow us to write for an
arbitrary power of x under the ergodic condition
X
(xn ) =
(
xn ) a ei t ,
(188)

with (
x ) given by the element of the corresponding matrix product. Thus,
every dynamical variable A(t) that can be expressed as a power series of x or x
or, more generally, as a power series of the form h(x) + g(x)

can be expanded,
whenever the particle has reached a state , as Eq. (156), and has a square matrix
A associated with it, with elements given by the amplitudes A . Note that the
condition of Hermiticity of the matrix x
, x
= x
, is consistent with the property

x
(k ) = x
(k ) satisfied by the amplitudes x
k = x
(k ) of the original expansion
(142).
4.3. Consequences of the ergodic principle for the dynamics
By reducing a product of the amplitudes a to a linear function of such variables,
the chain rule has as a direct consequence the emergence of a matrix algebra for
the amplitudes A . Moreover, application of Eq. (182) shows that also the fundamental oscillators of the form (153) satisfy a matrix algebra. In line with the
customary notation, in what follows A represents the matrix whose elements are
A , in contrast to subsection 4.2.2, where A denoted the corresponding matrix.
With this notation we can rewrite (152) as
d3 x
(t)
d2 x
(t)

= f(t) + m
+ eE(t).
(189)
2
dt
dt3
It is important to stress that this is much more than a new form of expressing
the equation of motion. As a consequence of the ergodic condition, neither x

(
x) nor f (f ) depend on the coefficients a , and hence all reference to the
stochastic variables has vanished from (189). That is, the original stochastic Eq.
(139), written in terms of c-(stochastic-)numbers, has been transformed into a nonstochastic matrix equation (q-numbers) as a result of the ergodic principle.
Equation (189) is precisely the equation of motion of non-relativistic qed. This
is a most important outcome, pointing to the convergence of the present theory and
(non-relativistic, spinless) qed. It should be clear, however, that the equivalence
between the two descriptions (qed and lsed) refers to their formal features. Although they share the same equations, they are conceptually distinct, there being
important differences in their physical outlook.
The fact that the oscillators A satisfy a matrix algebra allows us to determine
the evolution law for every dynamical variable A that can be represented in the
form (156). Indeed, by considering the time derivative of A (t) and using Eq. (183),
one arrives at
dA (t) X h i
=
i , A
a ei t ,
(190)
dt

March
21,
Brazil*de*la*pena*et*al

38

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

e
is a diagonal matrix with elements
= . It follows that A
where

=
h
i
A(t)

i ,
, or, in closed matrix notation,

h
i

dA(t)

.
= A(t),

dt

(191)

This result is a direct consequence of the structure of the expansion A (t) and of
the antisymmetry of the frequencies . This antisymmetry is thus at the root
of the description of the evolution of the dynamical operators by an algebra of
is central in determining
commutators. Equation (191) shows that the matrix
the evolution of the mechanical subsystem. As this evolution is controlled by the
should be related with the
Hamiltonian of the particle, it follows that the matrix

matrix H.
4.3.1. Radiationless regime. Contact with quantum mechanics
Once the quantum regime has been attained that is, once the system has reached
the stationary and ergodic limit, thanks to the combined effect of radiation reaction
and the zpf the radiative terms can be neglected in a first (radiationless) approximation, as discussed in section (3.5). Under this approximation the Hamiltonian
matrix (of the particle) reduces to
2
= p + V (
x),
H
2m

(192a)

where x
and p evolve according to
p = m

d
x
,
dt

d
p
f =
.
dt

(193)

commutes with the


The stationarity condition implies, through Eq. (191), that H
and hence is also diagonal, with elements
diagonal matrix ,
= E .
H

(194)

and H
are related (diagonal) matrices, we may write
as a product of the
Since

form H G(H)
i

h
i

dA(t)

G
.
= A(t),
H
dt

(195)

H)
to preserve
An inspection of Eqs. (192a) and (193) shows that for the function G(
and
the linear correspondence between V (
x) and f(
x), it must be independent of H

hence constant; we therefore write = C H, with C independent of H, i.e., of the


specific problem, and therefore
i

h
i

dA(t)

= C A(t),
H .
dt

(196)

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

39

Applied to A = x
, this equation leads to the commutator
1
I.
(197)
C
This is an important result, as it indicates that the value of the basic commutator
fixes the scale of the time evolution in the quantum regime, according to Eq. (196).
Hence the commutator of x
and p and the equation that governs the evolution of
the dynamical variables (both in the radiationless approximation) are intimately
related, which endows the commutator with a dynamical meaning.
Given the universality of Eq. (197), one may use the problem of a harmonic
oscillator of frequency 0 to determine the value of the constant C. For this purpose,
recall from section 2.8 that the minimum value of the product of the dispersions of
x and p for the oscillator is
[
x, p] =

x2 p2


min

~2
E02
=
.
02
4

(198)

On the other hand the variances x2 and p2 satisfy the Robertson-Schrodinger inequality,

2
2
x2 p2 14 |h[
x, p]i| + 21 {
x, p} h
xi h
pi ,
(199)
with {
x, p} = x
p + px
. From this (strictly algebraic) expression and Eq. (197), it
follows that

2
x, p]i| = 4C1 2 ,
(200)
x2 p2 min = 14 |h[
whence by comparing with (198) one obtains C = ~1 , i.e.,
[
x, p] = i~I.

(201)

The present derivation of this well-known quantum relation is important because


it shows that the value of the commutator [
x, p] is determined exclusively by the
properties of the zpf; more specifically, by the constant that fixes the mean energy
of each of its modes. It follows that as a result of the action of the zpf every
quantum system acquires unavoidable fluctuations, determined in general by the
same universal constant. It is thus confirmed that the so-called quantum fluctuations
are real and causal.
4.4. The Heisenberg equation and energy eigenvalues. Transition
to the Hilbert-space description
With C = ~1 introduced in Eq. (196) we get
i~

dA h i
= A, H ,
dt

(202)

which is the Heisenberg equation for the dynamical operator A(t).


On the other
hand, from (191) and (194) = E /~, except for an additive constant, whence

March
21,
Brazil*de*la*pena*et*al

40

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

(183) can be identified with Bohrs transition rule,


~ = E E ,

(203)

and the resonance frequencies with the quantum transition frequencies. This identification shows that the transitions occur during resonant responses of the mechanical
system driven by the zpf. This result explains how it is that the electron knows in
advance the energy of the state where it will land when realizing a transition, as the
energy difference is determined precisely by the resonance. The resonant response
of the particle to a selected set of frequencies of the background field constitutes the
physical mechanism responsible for the transition of the particle to one among the
collection of accessible stationary states. Which transition will effectively take place
in a given case is a question of chance, since it depends on the precise conditions of
the atom and the mode of the field at the moment of the transition.This resonant
phenomenon, along with the fact that the quantities featuring in Eq. (203) become
non-stochastic, point to the ergodic condition as a quantization principle.
was used in deriving the evolution equation.
The diagonal form of the matrix H
However, this latter continues in force after a unitary change of basis is performed
into a non-diagonal matrix H
0 = U HU
. In the latter case the
that transforms H
0
becomes a fluctuating quantity, and the states cease
energy that corresponds to H
to be stationary. Thus Eq. (202) gives the general law of evolution in the quantum
regime; it includes stationary and time-dependent states.
To complete the present description it becomes necessary to introduce the vectors of an appropriate Hilbert space that represent the states of the quantum system.
We start by noting that the matrix A associated to any dynamical variable that
can be written in the form (156) and whose elements are given by the elementary
oscillators A (t) = A ei t can be expanded as
X
X
A =
A (t) |e i he | =
A ei t |e i he | ,
(204)
,

in terms of a basis {|e i he |} of operators constructed from the vectors of a complete


orthonormal basis {|e i}.
The time dependence of A which lies at the core of the Heisenberg
representation can be transferred from the matrix elements to the vectors of
a new basis obtained by means of the unitary transformation
|e i |(t)i = ei(E /~)t |e i ,

|e i = |(0)i ,

(205)

so that Eq. (204) takes the form (we used Eq. (203))
X
A =
A ei t |(t)i h(t)| =
,

X
,

A |(0)i h(0)| = A(0).

(206)

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

41

The vectors {|(t)i} are directly related to the energy values {E }, and evolve in
time according to Eq. (205). This allows us to establish contact with the Schrodinger
representation of qm, as is usually done.
The statistical meaning of the quantities appearing in this description can be
recovered as follows. From Eq. (206) one gets the expected relation between the
matrix elements of A and the corresponding state vectors,
|e i = h(t)| A |(t)i h| A |i .
A = he | A(t)

(207)

This result, along with Eq. (202), leads to the standard quantum-mechanical formalism. For A Hermitean, Eq. (165) (which holds in general for every dynamical
variable of the form (156)) allows us to write
t
A = hA i = A = h| A |i ,

(208)

whence h| A |i can legitimately be called an expectation (or mean) value. Further,


Eq. (207) applied to A = x together with Eqs. (167) and (179) leads to
X
X
2
2
x2 =
|h| x
|i| =
h| x
|i h| x
|i = h| x
2 |i h| x
|i . (209)
(6=)

(6=)

Since the variance is obtained by calculating averages either over time or over the
realizations of the field, the first equality confirms that the quantities h| x
|i possess a statistical connotation. An equation similar to (209) holds for the variance
2
of any dynamical variable A, so that although no trace of stochasticity remains
A
in the quantities (207), their statistical nature has not been lost. Moreover, the
last equality in (209) indicates that the quantity as calculated within the standard
quantum formalism should be interpreted just as the statistical variance. From this
it follows, in particular, that the Heisenberg inequality actually involves statistical
variances, so that no reference to observations or measurements is required for its
interpretation: it ensues as a direct consequence of the persistent action of the zpf
once the ergodic regime has been reached, as mentioned earlier.
Finally, with the elements at hand it is a simple matter to make the transition
to the Schr
odinger equation in terms of wave functions of the form (x, t) =
hx| (t)i = eiE t/~ (x).
4.5. Bipartite systems. Emergence of entanglement
The theory just presented can be generalized to the case of two noninteracting
particles. Here we briefly sketch some of our main results (details can be seen in
Refs. 52, 53). The starting equations of motion are
1
...
...
m1 x
1 = f1 (x1 ) + m1 1 x 1 + e1 [E1 (t) + m2 2 x 2 ],
e2
1
...
...
m2 x
2 = f2 (x2 ) + m2 2 x 2 + e2 [E2 (t) + m1 1 x 1 ].
e1

(210a)
(210b)

March
21,
Brazil*de*la*pena*et*al

42

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

These equations are entirely similar to the one studied in the single-particle case
(Eq. (139)), the crucial difference being the presence of the last (radiative) terms,
which couple the particles. This coupling is fundamental, and shows that each particle modifies the field acting on the other one, so that the partners cease to be
independent, a fact that ultimately leads to nonclassical correlations between both
particles, as will be seen below.
The radiative coupling terms in Eqs. (210) superpose on the respective background field Ei (t) in the vicinity of the particle i(= 1, 2), giving rise to effective
...
external fields Eieff = Ei (t) + (mj j /ej ) x j . Since the particle located at x1 reaches
a stationary state when the mean power radiated by it balances the mean power
absorbed from the effective field E1eff , it follows that the solution x1 (assuming
the particle has reached a stationary state characterized by ) bears information
regarding the state 0 reached by the second particle. As a result of this coupling,
the solutions of Eqs. (210) must now be labeled by a compound index A = (, 0 )
that is in direct (though not univocal because of the possible energy degeneracy)
correspondence with the total mechanical energy EA = E + E0 . Moreover, the stationary solutions depend on the stochastic variables of the background field E1 (t)
({a }) and also on those of E2 (t) ({b0 0 }). This ultimately leads to the substitution xi (a ) xiA = xi0 (a , b0 0 ) for the solutions of (210), and similarly for
the expansions of the dynamical variables.
Now, if F (or G) represents a dynamical variable that belongs to particle 1 (or
2) in state (or 0 ), the expansion for the product variable F G in the radiationless
approximation can be written as

X
0 0 +
0 0 a b0 0
(F G)A = F G
F G
+ O,
(211)
, 0

=0 0 6=0

where O contains the set of all time-dependent (oscillatory terms) that average to
t
t
0 0 , the covariance (F G) (F G) t FA t GA t
zero. Since FA = F and GA = G
A
A
becomes

X
0 0 a b0 0
(F G)A =
F G
.
(212)
, 0

=0 0 6=0

Given that (or 0 0 ) represents a relevant frequency of particle 1 (or 2), this
equation shows that the existence of nontrivial (nonzero) common relevant frequencies is crucial for the existence of correlation between the (corresponding dynamical variables of the) particles. The common frequencies satisfy the condition
= 0 0 , or equivalently EA = EB , as follows from Eq. (203). Thus, when
the total energy EA of the actual state (, 0 ) has no degeneracies, the particles are
uncorrelated.
4.6. Entangled states
Let us focus on two-particle states A whose energy is degenerate (with EA = EK )
and consider the following expansion, describing the variable F G not in state A,

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

43

but rather in a state of (degenerate) energy EA :


(F G)EA =EK =

1
2

[(F G)A + (F G)K ]



X
0 0 a b0 0 + F G
0 0 a b0 0 ei( +0 0 )t
= 12
F G
, 0

1
2

X
, 0



0 0 + a a0 0 F G
0 0 a b0 0
F G

+O
=0 0

(213)
Accordingly, in what follows we consider the covariance
(F G)E

A =EK

(F G)EA =EK FEA =EK GEA =EK .

(214)

As was done in the one-particle case, we can now perform a unitary transfor H1 H2 (with
mation that transfers the evolution from the matrices F (t) G(t)
i(
+
0 0 )t

elements F G0 0 e
) to vectors of the expanded Hilbert space. In the
radiationless approximation, such evolving vectors are
|A(t)i = |(t)i |0 (t)i = ei(EA /~)t |eA i .

(215)

in terms of the basis {|A(t)i hB(t)|} we obtain the usual expression


Expanding F G
0 0 = hA| F G
|Bi .
F G

(216)

An essential difference with respect to the one-particle problem arises, since now
the energy EA is not always univocally related to a single vector of the transformed
basis (215). More specifically, an analysis of the spectral decomposition of (F G)A
allows us to conclude that for nondegenerate EA , the factorizable vector |i |0 i is
univocally related to EA , whereas for degenerate EA a new type of vector, that is
not an element of the basis {|A(t)i} , must be introduced. Whenever there exist
common relevant frequencies = 0 0 , the vector representing the state has the
structure
|i |0 i + AK |i | 0 i ,

(217)

where AK is a non-stochastic phase factor given by


DG a b0 0 = ei0 0 = ei ,

for

= 0 0 .

(218)

Thus, whenever the two particles share a common resonance frequency, a new class
of state vector arises in the transition to a Hilbert space description, which is nonfactorizable and gives rise to entanglement.
Since the entangled vectors are the suitable ones for describing the mechanical system in the degenerate case, they must be associated with the expansions
(F G)EA =EK in the form


(219)
(F G)EA =EK 12 |i |0 i + AK |i |0 i .

The factor 1/ 2 here originates in the factor 1/2, representing a (balanced) statistical weight, which appears in the first line of Eq. (213).

March
21,
Brazil*de*la*pena*et*al

44

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

The need for entangled vectors ultimately implies nonzero correlations between
(certain variables of) the particles, as stated after Eq. (212). From here it follows
that interaction with a common background field, emergence of correlations, and
entanglement are tightly linked notions. On the other hand, the covariance qm
FG
calculated in accordance with the quantum-mechanical rule,
D
E D ED E
G
F G
h| F G
|i h| F |i h| G
|i
qm
=
F
(220)
FG
agrees with the expressions obtained for (F G)A and (F G)E =E whenever |i is
A
K
given by |i |0 i and by (219), respectively. This exhibits the need for quantum
mechanics to resort to entangled states in order to properly describe these bipartite
correlations. The agreement between our results and those obtained with the usual
quantum methods confirms that also in the two-particle case, the present theory
correctly leads to quantum mechanics once the stationary, ergodic and radiationless
limit is taken and a description in terms of vectors in a Hilbert space is adopted.
Note that the correlation due to the structure of the vector (219) depends in
a most direct way on the phase factor AK , whose origin goes back to the field
variables, since DG can be also written as DG = ha( )b(0 0 )i. In this regard
it is important to point out that it is thanks to Eq. (218) which allows to write
AK as a nonstochastic phase factor that the background field leaves its footprint
in the Hilbert-space description, even in the radiationless limit and when all explicit
reference to the stochastic field variables has disappeared. The entanglement phase
factor AK appears thus as a vestige of the zero-point field, reminding us of its
active role as a member of the entire system.
4.6.1. Identical particles: unavoidable entanglement
As stated above, the existence of a common relevant frequency is sufficient for the
particles to get entangled, regardless of their nature. However, when the particles
are identical and are subject to the same external potential, their entanglement
becomes unavoidable, since in this case all the relevant frequencies (or equivalently,
the sets of accesible states) are common to both particles.
The double degeneracy of all the energies E(,0 ) = E(0 ,) leads us to associate
(F G)E(,0 ) =E(0 ,) with the vector (c.f. Eq. (219))


0
0
1
|i
|
i
+

|
i
|i
A a0 b0 , ( 6= 0 ) .
(221)
A
1
2
1
2 ,
2
Notice the need to introduce here the indices 1, 2 in order to avoid ambiguities with
regard to the particle that is in the given state or 0 .
A natural transformation in this kind of systems is the exchange of particles, Ip .
According to the present approach, exchanging the particles amounts to exchange
their actual accessible states (operation Is : 0 ), plus the fields in which they
are immersed (operation If : a b). Direct application of Ip = Is If shows that the
phase factor A is invariant under the exchange of particles,
Ip A = Ip a0 b0 = b0 a0 = A .

(222)

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

45

However, when the field variables are eliminated from the description and the transition to the product Hilbert space formalism is made, the exchange of particles
reduces to the substitution 0 . The exchange of particles is then represented in
this reduced description by Ipqm = Is , where the superindex qm distinguishes this
operator from the transformation Ip . Moreover, consistency with the radiationless
approximation requires writing A = a0 b0 in a form that does not make explicit
reference to the field variables. The required form turns out to be
A = exp(i0 ),

(223)

and this is the expression over which Ipqm acts. Since A must maintain its invariance
properties irrespective of the specific form we use to express it (whether (221) or
(223)), in addition to (222) we must have Ipqm A = A , with A given by (223).
From here it follows that A = A , and consequently, A being a phase factor, we
arrive at
A = 1,

6= 0 .

(224)

This result tells us that a system of two identical, non-interacting particles subject
to the same external potential, is described by maximally entangled, hence totally
(anti)symmetric states.
5. Contact with QED. Radiative Corrections
In this last part we extend our journey from quantum mechanics to a domain usually
considered to be the exclusive province of quantum electrodynamics: the radiative
transitions and corrections to the atomic energy levels.
We achieve this by drawing further consequences from the energy-balance equation, and more generally by focusing on the effects of the radiative terms that were
neglected in the transition to the Schrodinger formalism carried out above. The
results obtained coincide in every case with those of nonrelativistic spinless qed.
This confirms that the present theory goes beyond quantum mechanics. Previous
related work can be seen in Refs. 10 and 54.
5.1. Absence of detailed energy balance
We have obtained already an important result from the energy-balance condition
for a system in a stationary state, Eq. (112),
e2 D E
...
pD .
m hx x i =
m

(225)

Applied to the ground state it ensures the correct value for the Planck constant
in the Schr
odinger equation. Now instead of considering the particle in its ground
state, we assume that it is in an excited state n, the background field still being
in its ground state. Then both sides of the detailed-balance condition must be

March
21,
Brazil*de*la*pena*et*al

46

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

recalculated, in the time-asymptotic limit. For the left-hand side one obtains
X
...
2
4
m hx x in = m
nk
|xnk | , n > 0.
(226)
k

For the right-hand side one gets


X
e2 D E
2
4
pD = m
nk
|xnk | signkn .
m
n

(227)

Equation (227) contains now a mixture of positive and negative terms, whilst in
(226) all contributions have the same sign. As a result, according to Eq. (77) there
is a net loss of mean energy,
X
d
2
4
hHin = m
nk
|xnk | (1 signkn )
dt
k
X
2
4
= 2m
nk
|xnk | ,
(228)
k<n

indicating that there cannot be detailed balance between the zpf and the particle in
an excited state as was to be expected, since the zpf is the background radiation
field in its ground state. This confirms that only the ground state of the particle
(n = 0) is stable in the sole presence of the zpf.
Let us investigate whether there is any background field () = 0 ()g() with
which a mechanical system in its excited state can be in equilibrium, where g()
represents either an excitation of the background field or the contribution of an
external field. To respond to this question we observe that the expressions for the
mean power radiated by the particle, Eq. (226), and for the mean power provided
by the field, Eq. (227), contain in general mixtures of terms of different frequencies
(nk for different values of k), but with different signs, so that there is no way that
detailed balance is satisfied in general. Only for the particular case in which all
values of nk coincide, the possibility of detailed balance exists.
We shall explore this possibility for the case of the harmonic oscillator, in which
all |nk | that contribute are equal and coincide with the oscillator frequency 0 .
2
2
2
With |xnn+1 | = a(n+1), |xnn1 | = an, |xnk | = 0 for k 6= n1, and a = }/2m0 ,
Eq. (226) gives
1
...
(229)
m hx x in = m 04 a(2n + 1) = } 03 (2n + 1).
2
From (227), on the other hand, one gets, with () = 0 ()gn () where gn () is a
function to be determined,
e2 D E
1
pD = m gn (0 )04 a(n + 1 n) = } 03 gn (0 ).
(230)
m
2
n
A comparison of these two expressions shows that indeed, detailed balance exists
between a harmonic oscillator in its excited state n and a background field with
spectral energy density
() = 0 ()(2n + 1).

(231)

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

47

Also this result should not come as a surprise, since this field has precisely an energy
per normal mode 21 }0 (2n + 1) equal to the energy of the mechanical oscillator
with which it is in equilibrium.
5.2. Radiative transitions and atomic lifetimes: Einsteins A and
B coefficients
Now we investigate some important implications of the absence of detailed balance.
This can be done by using again Eq. (77) to obtain the average energy lost (or
gained) by the mechanical system due to the difference between the terms on the
right-hand side. According to Eqs. (226) and (227) (but now with () = 0 ()g();
() = (||)), this difference is given by
X
dHn
2
4
= m
nk
|xnk | [1 g(|nk |)signkn ] .
dt
k

It is convenient to rewrite the right-hand side by introducing g() = 1 + ga (), in


order to separate the contribution coming from the external (or additional) background field
a () = 0 ()ga (),

(232)

X
dHn
2
4
= m
nk
|xnk | [1 (1 + ga (kn ))signkn ] =
dt
k
X
2
4
= m
nk
|xnk | [(ga )kn >0 (2 + ga )kn <0 ] .

(233)

The term within the brackets in the second line of this equation, proportional to ga ,
represents the absorptions and the second one, proportional to 2+ga , the emissions.
It is clear from this result that there can be absorptions only when the background
field is excited or there is an external component, whilst the emissions can be either
spontaneous (in presence of just the zpf) or else stimulated by the additional field
(represented by ga ). The coefficients appearing in the various terms determine the
respective rates of energy gain and energy loss; therefore, they should be expected to
be directly related with Einsteins A and B coefficients. The coefficient A is defined
as the time rate for spontaneous emissions,
X
dHn = }
nk Ank dt,
(234)
k

and the coefficients B, which determine the rate of energy gain or loss due to
transitions induced (stimulated) by the external field, are defined through
X
dHn = }
|nk | Bnk,kn a (|nk |)dt.
(235)
k

March
21,
Brazil*de*la*pena*et*al

48

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

With the aid of these definitions Eq. (233) can be rewritten in the even more
transparent form
X
dHn
=
} |nk | [a (|nk |)Bkn ]
(236)
dt
k>n
X

} |nk | [Ank + a (|nk |)Bnk ].


k<n

whence, by comparison with (233),


3

Ank =

4e2 |nk |
2
|xnk | ,
3}c3

(237)

in agreement with the qed prediction.23,55 In its turn, the coefficient B is given by
the first term (in the case of absorptions) or the last one (for emissions) within the
square brackets in Eq. (233),
2

Bnk = Bkn =

4
m nk
|xnk | ga (nk )
4 2 e2
2
=
|xnk | ,
} |nk | a (nk )
3}2

(238)

a result that also coincides with the respective formula of qed (or qm).23 This
confirms the key role played by both radiative terms the radiation reaction and
the background field (which always contains the zpf but can include the additional
component) in determining the rates of transition between stationary states of
the mechanical system. The results also demonstrate the equivalence of the present
theory and nonrelativistic qed.
The expressions for the Einstein coefficients Ank , Bnk , and Bkn , involve each one
the single frequency |nk | , which means that the system as a whole reaches a state
of detailed equilibrium, i.e., equilibrium of matter with the field at each separate
frequency, as was already noticed. The theory has thus led us to a transition from
global equilibrium to detailed equilibrium in the quantum regime. We recall that
this demand was one of Einsteins hypotheses in his pioneering work where he
introduced the absorption and emission coefficients.56
It is pertinent to ask here at what point does quantization enter in Einsteins
paper so as to arrive at the Planck distribution, an inquiry that comes to surface not
infrequently. A current answer to this question is that quantization is introduced by
assuming discrete atomic levels. However, this is wrong, as Einstein and Ehrenfest
demonstrated some time after the initial paper by redoing the calculations with a
continuous distribution of atomic levels,57 recovering the old results. The correct
answer is that quantization enters through the introduction of a source that includes
at the outset the possibility of the zpf, able to generate spontaneous transitions.
This can be easily verified by redoing the Einsteinian calculation (see Eq. (240)),
but omitting any of the three terms that lead to matter-field equilibrium (stimulated absorptions and emissions, and spontaneous emissions). The absence of the
latter leads to absurd results (such as atomic coefficients that depend on the temperature). It is interesting to observe that the omission of the term that describes

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

49

the stimulated emissions in Eq. (233) (after introducing appropriate populations)


leads to the approximate expression for Plancks law proposed by Wien (Eq. (33)),
which correctly approximates it at low temperatures, so it already contains some
quantum principle (as corresponds to the consideration of the zpf). All this can be
easily seen in the present context by focusing on just two states n and k, with En
Ek = }nk > 0 and respective populations Nn , Nk . When the system is in thermal
equilibrium the relation
Nk /Nn = exp(En Ek )/kB T

(239)

holds (forgetting about possible but inconsequencial degeneracies). Since according


to Eq. (233) the number of emissions is proportional to Nn ga (nk ) and the number
of absorptions is proportional to Nk [2 + ga (nk )], from the equilibrium condition
Nn ga = Nk (2 + ga )

(240)

one obtains indeed Plancks law (for the thermal field)


ga (nk ) =

2
.
exp [(En Ek )/kB T ] 1

(241)

The ratio of the A to the B coefficients at any given temperature is


3

Ank
~ |nk |
= 20 (|nk |).
=
Bnk
2 c3

(242)

This relation and the equality Bnk = Bkn were predicted by Einstein on the basis of
his statistical considerations; here they follow from the theory, as is the case in qed.
Notice in particular the factor 2 in Eq. (242). Given the definition of the coefficients,
one could expect the ratio in this equation to correspond exactly to the spectral
density of the zpf, which would mean a factor of 1. The factor 2 seems to suggest
that the zpf has double the ability of the rest of the electromagnetic field to induce
transitions. The correct explanation is another: inspection of Eq. (233) shows that
one should actually write 20 = 0 + 0 . One of these two equal contributions to
spontaneous decay is due to the effect of the fluctuations impressed on the particle
by the field; the second one is the expected contribution due to radiation reaction,
that is, Larmor radiation. Not surprisingly, they turn out to be equal: it is precisely
their equality what leads to the exact balance of these two contributions when the
system is in its ground state, guaranteeing the stability of this state. Yet one can
frequently find in the literature that all the spontaneous decay is attributed to one
or the other of these two causes, more frequently to Larmor radiation. It is an
important result of both the present theory and qed that the two effects contribute
with equal shares. (Interesting related discussions can be seen in Refs. 23, 58-61.)
5.3. Radiative corrections to the energy levels
The derivation of the Schr
odinger equation presented in section 2 confirms that
the radiative terms (or corrections) give rise just to corrections to the solutions of

March
21,
Brazil*de*la*pena*et*al

50

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

the (unperturbed) Schr


odinger equation. The Einstein A and B coefficients for the
lifetimes of excited states, determined by the right-hand side of Eq. (233), pertain
to this category. A further important even if smaller radiative correction,
one that represents a major success of qed, is the shift of the atomic levels due
to another residual effect of the zpf. Indeed, the effective work realized by the
fluctuating motions of the bound particle gives rise to a tiny modification of the
mean kinetic energy that affects the energy levels, as is here shown by means of a
direct approach.
5.4. The Lamb shift
Let us go back to Eq. (76e),
D E
1
2
d
...
,
hxpi =
p + hxf i + m hx x i e2 xD
(243)
dt
m
and use it to calculate the radiative energy shift. As explained in section 3.2, Eq.
(243) is a time-dependent version of the virial theorem, with radiative corrections
included. It is interesting to observe that the average values are here taken over the
ensemble, instead of over time. This is but an example of the ergodic properties
acquired by the quantum states, a matter discussed at length in section 4.2.
In the stationary state, the two previously neglected terms in Eq. (243) represent
radiative corrections to the (kinetic) energy T ,
e2 D E
m
...
hx x in +
xD .
(244)
hT in =
2
2
n
For consistency, the contribution of these terms is again calculated to lowest
D Eorder
...
, in
in e2 , which means calculating the two average values, hx x in and xD
n
the radiationless limit. The first one gives, using the solutions of the Schrodinger
equation,
m

d
...
hx x in = hx f in =
hT in = 0,
(245)
2
2
2 dt
which means that the Larmor radiation does not contribute to the energy shift
in the mean, in a stationary state. The correction to the energy is therefore due
exclusively to the diffusion produced by the interaction of the particle with the
background field, represented by the second term in Eq. (244):
Z
2e2
3
e2 D E
2
xD =
|x
|

d 2
.
(246)
nk
kn
3
2
3c k
kn 2
n
0

The radiative correction to the mean energy is therefore given by (in three dimensions, for comparison purposes)
Z
e2 D E
2e2
3
2
En =
xD =
|x
|

d 2
.
(247)
nk
kn
3
2
3c k
kn 2
n
0
This result coincides with the formula derived by Power62 for the Lamb shift
on the basis of Feynmans argument.63 According to Feynman, the presence of

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

51

the atom creates a weak perturbation on the nearby field, thereby acting as a
refracting medium. The effect of this perturbation is to change the frequencies of
the background field from to /n(), n being the refractive index. The shift of
the zpf energy due to the presence of the atom is then23,62
En =k,

1 }k
1
1
k, }k ' k, [n(k ) 1] }k ,
2 n(k )
2
2

(248)

and the refractive index is given in this approximation by (Ref. 58, chapter 9)
2

n(k ) ' 1 +

4 |dmn | mn
,
2 2
3} m mn

(249)

where dmn = exmn is the transition dipole moment. After an integration over the
and summation over the polarizations = 1, 2, Power obtains in the
solid angle k
continuum limit for k the formula
Z
2
3
2
|d
|

,
(250)
En =
d
mn
mn
2 2
3c3 m
mn
0
which coincides with the previous result, Eq. (247).
The Lamb shift proper (called also observable Lamb shift) is obtained by subtracting from the total energy shift given by (247), the free-particle contribution
represented by this same expression in the limit of continuous electron energies
(when kn can be ignored compared with in the denominator),
Z
Z
2e2
e2 }
2
Efp =
|x
|

d .
(251)
d

=
nk
kn
3c3 k
mc3 0
0
This gives for the Lamb shift proper of level n
ELn = En Efp

2e2
2 3
|xnk | kn
=
3c3 k

d
0

2
kn

,
2

(252)

which agrees again with the nonrelativistic qed formula. The integral diverges logarithmically, so inserting the usual (non-relativistic) regularizing cutoff c = mc2 /},
one gets


mc2
2e2
2 3


|x
|
(253)
ELn =

ln
nk
kn
}kn .
3c3 k
This is the Bethes well-known expression. Note, however, that in the present approach (as in Powers) no mass renormalization was required; we come back to this
point below.
The interpretation of the Lamb shift as a change of the atomic energy levels due
to the interaction with the surrounding zpf is fully in line with the general approach
of the present theory. It constitutes one more manifestation of the influence of the
particle on the field, which is then fed back on the particle. An alternative way of
looking at this reciprocal influence is by considering the general relation between

March
21,
Brazil*de*la*pena*et*al

52

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

the atomic polarizability and the refractive index of the medium affected by it
(for n() ' 1),
n() = 1 + 2().

(254)

By comparing this expression with Eq. (249) one obtains


2

n () =

4 |dmn | mn
,
2 2
3} m mn

(255)

which is the Kramers-Heisenberg formula.58 This indicates that the Lamb shift can
also be viewed as a Stark shift associated with the dipole moment d() = ()E
induced by the electric component of the zpf on the atom,
ELn =

1
2

hd Ein .

(256)

For completeness, let us look at the radiative energy corrections as is proper (mutatis mutandis) of qed and also physically very suggestive. The (minimal) coupling
of the particle to the zpf gives the Hamiltonian
H = H0

e2
e
A p+
A2 ,
mc
2mc2

(257)

where H0 = (p2 /2m) + V is the original Hamiltonian and A represents the vector
potential associated with the zpf. Note that the Hamiltonian is now a stochastic
variable, whence the (mean) energy shift is given by the average of Eq. (257) over
the realizations of the field (denoted by the superscript E),
En =

E
e2
e
E
Ap +
A2 .
mc
2mc2

With the correlation of the field given by (see section 3.1)


Z
E
E(t)E(t0 ) = (4/3)
0 () cos (t t0 )d,

(258)

(259)

calculation of the second term is straightforward and yields


Z
E
e2 }
e2
2
A =
d ,
(260)
Efp =
2mc2
mc3 0
R
E
since A2 = (2}/c) d for the zpf and 0 () = ~ 3 /(2 2 c3 ). This result
reproduces Eq. (251), the free-particle correction to the energy, which does not
contribute to the observable Lamb shift. Thus the free-particle Lamb shift is just
E
the contribution due to the presence of the unperturbed zpf; the term A2 is of
universal value, independent of the system or its state, a testimony of the ubiquitous
presence of the zpf.
For the calculation of the first term one needs to consider the effect (to first order
in e, and returning to the 1-D notation) of the stochastic field A on the particle
momentum p. This is most easily done by rewriting the original equations of motion

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

53

...
in terms of the (canonical) variables (we continue however to treat the term m x
has an external force acting on the particle.), i.e.
...
mx = p (e/c)A, p = f (x) + m x
(261)
whence the new (stochastic) Liouvillian is
= 1 (p e A) + (f + m ...
x ).
L
m
c x p

(262)

With p(t) = eL(tt ) p(t0 ) one obtains to lowest order in e2 , after averaging
over the realizations of the field, taking the time-asymptotic limit and integrating
over the particle phase space,
 0
 e 2  2} 
e
p

hApin =
,
(263)
ddt0 cos (t t0 )
mc
mc
3c
x n
whence (returning to 3-D notation for comparison purposes)
Z
e
2e2

2 3
ELn =
hA pin =
|x
|

.
d 2
nk
kn
2
mc
3c3 k

0
kn

(264)

This result coincides with Eq. (252), thus confirming the consistency of the different
approaches.
It seems convenient to point out some differences between the procedures used
in sed and in qed to arrive at the formula for the Lamb shift. In the qed case,
second-order perturbation theory is used, with the interaction Hamiltonian given
int = (e/mc)A
p
by H
. But the energy derived from this term,23
Z
2e2

2 2

|xnk | kn
d
,
(265)
3c3 k

nk
0
still contains the (linearly divergent) free-particle contribution
Z

Z
4e2
2e2
1
2 2
2
|xnk | kn
d =
|pnk |
d

3c3 k
3c3 2m k
0
0

(266)

that must be subtracted to obtain the Lamb shift proper. Because the result is
proportional to the mean kinetic energy, the ensuing correction represents a mass
correction (mass renormalization),
Z
4e2
m =
d,
(267)
3c3 0
which with the cutoff c = mc2 /} becomes m = (4/3)m. On the other hand,
in the derivation presented here to obtain the formula for EL , Eq. (252), there was
no need to renormalize the mass. The result (267) is just the classical contribution
to the mass predicted by the Abraham-Lorentz equation (see Ref. 36, Eq. 3.114);
in the equations of motion (2.1) this contribution has been already subtracted, so
there is no more need of mass renormalization in the sed calculation. As has been
seen, however, the formula (252) (common to both sed and renormalized qed) still
has a logarithmic divergence, which calls for the introduction of the cutoff frequency

March
21,
Brazil*de*la*pena*et*al

54

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

c as was done by Bethe, thus leading to a very satisfactory result for the Lamb
shift. From the present perspective the problem of the divergence is closely linked
to the unsolved problem of the (unphysical) divergence at high frequencies of the
zpf energy density given by Eq. (67).
5.5. External effects on the radiative energy corrections
By now it is clear that some basic properties of the vacuum such as the intensity
of its fluctuations or its spectral distribution are directly reflected in the radiative
corrections. This means that a change in such properties can in principle lead to an
observable modification of these corrections. The background field can be altered,
for instance, by raising the temperature of the system, by adding external radiation,
or by introducing objects that alter the distribution of the normal modes of the field.
Such environmental effects have been studied for more than 60 years, normally
within the framework of quantum theory although some calculations have been
made also within the framework of sed, in particular for the harmonic oscillator,
leading to comparable results (see e.g. Refs. 64-66). Here we have the possibility of
applying the formulas derived in the previous sections to the general case, without
restricting the calculations to the harmonic oscillator. The task is facilitated and
becomes transparent by the use of the present theory because the presence (and
action) of the background radiation field is clear from the beginning.
In section 5.2 we have already come across one observable effect of a change in the
background field: according to Eq. (233) the rates of stimulated atomic transitions
are directly proportional to the spectral distribution of the external (or additional)
background field, be it a thermal field or otherwise. In the case of a thermal field
in particular, with ga (nk ) given by (241), the (induced) transition rate from state
n to state k becomes (using Eqs. (236) and (238))
dNnk
= 0 (nk )a (nk )Bnk =
dt
3
2
4e2 |nk | |xnk |
1
=
.
3}c3
e}|nk |/kB T 1

(268)

This result shows, as is well known, that no eigenstate is stable at T > 0, because
the thermal field induces both upward and downward transitions. For downward
transitions (nk > 0) we can rewrite (268) for comparison purposes in terms of Ank
as given by Eq. (237),
dNnk
Ank
= }| |/k T
,
B
nk
dt
e
1

(269)

which indicates that the effect of the thermal field on the decay rate is hardly noticeable at room temperature (kB T ' .025 eV), since for typical atomic frequencies
1
[exp (} |nk | /kB T ) 1] ranges between exp(40) and exp(400).
When the geometry or the spectral distribution of the field is modified by the
presence of conducting objects, the transition rates are affected accordingly. Assume,

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

55

for simplicity, that the modified field is isotropic, with the density of modes of a
given frequency nk reduced by a factor g(nk ) < 1. Then according to the results of
section 5.2 the corresponding spontaneous and induced transition rates are reduced
by this factor, since both A and B are proportional to the density of modes.
By enclosing the atoms in a high-quality cavity that excludes the modes of this
frequency one can therefore virtually inhibit the corresponding transition. For the
more general anisotropic case the calculations are somewhat more complicated,
without however leading to a substantial difference from a physical point of view.
These cavity effects have been the subject of a large number of experimental tests
since the early works of Kleppner,67 Goy et al.68 and others. Concerning the physical
implication of these effects, they provide a clear proof that the background field,
including the zpf, acts as a mediator between the atom and the cavity walls, so
as to directly influence the (spontaneous and induced) transition rates. How else
could the atom register the influence of its surroundings, even before the emission?
Moreover, it is clear that, rather than the energy levels of the initial and final atomic
states, it is the (resonance) frequencies that play an essential role in determining
the transition rates, as expressed in the formulas for A and B.
The changes in the energy shift produced by the addition of an (external or
thermal) background field can be calculated readily from Eqs. (251) and (252). Let
a be the spectral energy density of the additional field, so that = 0 + a . Then
the formulas for the variations of the (first-order) radiative corrections are obtained
by determining the shifts produced by the total field and subtracting the original
shifts produced by the zpf (0 ). The results are
Z
Z
e2 }
a
a
4e2
2
|xnk | kn
(270)
d 2 =
d ,
(Efp ) =
3
3} k

mc
0
0
0
Z
2e2
a

2 3
(ELn ) =
|x
|

d
(271)
nk
kn
2 2 ,
3c3 k

0
0
kn
for a homogeneous field. If, for instance, the additional field represents blackbody
radiation at temperature T , i.e. a (T ) = 20 /( 1) with  = exp(}/kT ), Eq.
(270) gives
Z
Z

2
y
2e2 }
2
d
=
(kT
)
dy
.
(272)
T (Efp ) =
mc3 0
1
mc2
exp
y
1
0
With

0 dy exp y1

2
6

this gives for the change of the free-particle energy

(kT )2 .
(273)
3mc2
The formula for the change of the Lamb shift is given according to Eq. (271) by


Z

4e2
1
2 3
(ELn ) =
|x
|
.
(274)

d
nk
kn
2 2
3c3 k
kn
exp(}/kT ) 1
0
T (Efp ) =

These results coincide with the qed predictions69,70 and the corresponding thermal
shifts have been experimentally observed (see e.g. Ref. 71). From the point of view

March
21,
Brazil*de*la*pena*et*al

56

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

of sed (or qed) their interpretation is clear: they represent additional contributions
to the kinetic energy impressed on the particle by the thermal field, according to
the discussion at the beginning of section 5.4.
6. What have we learned about quantum mechanics?
We have arrived along this work at an important substantiation: the quantum properties of both matter and field emerge quite naturally from the consideration of the
existence of a real, pervasive, ubiquitous zero-point radiation field. From the analysis in section 2 of the problem of the thermal radiation field in equilibrium we
concluded that in presence of its zero-point component the field becomes quantized.
We then made the corresponding analysis for matter, and found that also it becomes
quantized. The conclusion is that the zpf is the central piece that nature uses to
perform the miracle of quantization in general. This is another form of saying that
the quantum phenomenon, rather than being intrinsic to matter or to the radiation
field, emerges from the matter-field interaction.
We further learned that the description provided by the Schrodinger equation
ensues from a Fokker-Planck-type equation in phase space. The secret of qm lies
to a large extent in the fluctuations of the momentum transferred to configuration
space. This striking result can be understood by observing that the fluctuating zpf
is able to compensate for the associated dissipative effects of radiation reaction,
thus allowing the particle to reach a stable dynamic state.
The Schr
odinger description refers to a statistical ensemble, not to an individual
particle. This conclusion appears as inescapable, and marks a clear departure from
the usual (Copenhagen or orthodox) interpretation of qm, in favour of the less popular ensemble (or statistical) interpretation. Further, the theory affords a physical
cause for the fluctuations characteristic of qm. The so-called quantum fluctuations
appear as real, objective fluctuations impressed on the particle by the permanent
interaction of the atomic system (or whatever quantum structure is under study)
with the zpf. This puts the quantum fluctuations on a mundane perspective.
The transition from the original equation to the Schrodinger equation is an
irreversible formal procedure: relevant information is lost along the way. Thus, one
cannot transit back on purely logical steps. In particular, one cannot reconstruct the
true probability density in phase space from the (approximate) Wigner function.
This explains the large number of existing phase-space versions of qm, resulting
from the attempts to discover the real one (see, e.g., Ref. 72).
The theory is based on the notion of trajectories, which belong to subensembles
characterized by the local mean velocities v(x) and u(x). Due to its intrinsically
statistical nature, it cannot be applied in general to isolated events, so individual
trajectories appear as unknown.
Even though a true Fokker-Planck-type equation in phase space exists, the appropriate (radiationless) description in configuration space requires only a pair of
balance equations, one being the continuity equation, the other describing the bal-

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

57

ance of the average momentum. The latter resembles the classical Hamilton-Jacobi
equation, but contains a crucial term that originates in the fluctuations in momentum space and refers to an ensemble of particles, thus changing the meaning of the
equation.
The solutions of the Schr
odinger equation must be consistent with the demand of
energy balance, Eq. (79). In addition to fixing the scale of the quantum phenomenon
through the introduction of Plancks constant, this condition is also the guarantor
of the stationarity of the zpf itself, by being satisfied frequency by frequency.
Another cost of the simplifications made is the nonlocal nature of the quantum
description. This nonlocality, which is not ontological but the result of a cryptic description, has been the source of much quantum ado, and even of avowals bordering
on mysticism. An important point to stress is that the nonlocality of qm applies
even to single-particle systems. Thus nonlocality and entanglement (which requires
at least a couple of particles) are not the same thing.
The simplifications made have also had the effect of deleting from the final
description every explicit reference to the zpf the ultimate cause of the quantum
behavior! As a result, the reason for the stochasticity of the system becomes hidden
and the fluctuations become causeless. From this moment on, quantum mechanics
cannot be understood from within quantum mechanics. Precisely by exhibiting the
zpf as the source of stochasticity, our results explain the success of a number of
works within stochastic electrodynamics. From among those of relevance we recall
the important numerical simulations of Cole and Zou,73,74 leading to a correct
statistical prediction of the ground-state orbit for the H-atom.
Further, our results led to a highly interesting conclusion regarding the stationary states of qm, namely that they satisfy an ergodic principle. This condition is
at the core of the linear response of the mechanical system to the field: whatever
the external force f (x), when the system is in a quantum state , the (resonant
modes) of the background field drive it as if it were composed of a set of oscillators
of frequencies . The ergodic principle is thus central in defining a matrix algebra
for the dynamical description of the system, and by assigning a sure (nonstochastic)
value to the resonance frequencies it plays the role of a quantization principle: it
selects from among all the possible stationary solutions of Eq. (139), those that
are robust with respect to the field fluctuations, which are the quantum solutions.
Further, the results of section 4 show that also the (nearby) zpf acquires specific
properties, which are elegantly encapsulated (or rather concealed) in the Heisenberg
matrix formalism for the particle dynamics.
The same physical principles that allowed us to recover and reread the quantum
formalism in the one-particle case, also served to disclose the physical mechanism
behind the entanglement of two noninteracting particles embedded in the (common) background field. It is found that whenever the particles share one resonance
frequency, correlations arise between their motions, these being induced via the
background field. When the description is made in terms of state vectors of an appropriate Hilbert space, the entangled states emerge naturally as the only ones that

March
21,
Brazil*de*la*pena*et*al

58

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

can reproduce such correlations. Moreover, when the particles are identical the ensuing states are totally (anti) symmetric. With these results lsed reveals the physical
mechanism and origin both of them foreign to the usual quantum-mechanical
description of the quantum symmetrization postulate.
It is important to realize that although the zpf could appear at first as a sort of
collection of hidden variables introduced to complete the quantum description, this
is not the case. Quite the contrary: nothing is added to qm, but the latter emerges
from a more general theory that contains the zpf. The emerging description is
then naturally indeterministic, since in every case the specific realization of the
field is unknown. Quantum mechanics is thus exposed as a (handy, but incomplete)
description of the statistical behavior in configuration (or momentum) space of the
mechanical part of the particle-field system.
Finally, the radiative terms that were neglected in the transition to quantum
mechanics are identified by the theory as the source of all (nonrelativistic) radiative
corrections, including the elusive Einstein A coefficient, the Lamb shift, and the
cavity effects on such corrections. All this can be expressed succinctly by stating
that the complete, nonapproximate theory is equivalent to nonrelativistic qed.
Since, according to the present theory, present-day qm furnishes merely an
approximate, time-asymptotic, partially averaged description of the physical phenomenon, there exists plenty of room for further and deeper investigations. So far
nobody has explored, for instance, the consequences of using the density Q(x, p, t)
given by Eq. (136), or the behavior of the system before it reaches the state of energy
balance (the quantum regime) in which the approximations apply. What would it
look like? One should expect an entirely unknown behavior of matter that can neither be classical because the ~ due to the interaction with the field is already in the
picture, nor quantum-mechanical because the conditions to apply such description
have yet not been reached.
Undoubtedly an exploration into this realm of physics would represent a new
adventure in physics, with possibly very promising outcomes, including predictions
that are experimentally testable. This adds of course an important motivation to
undertake deeper and further investigations into the theory.
Acknowledgment. The authors gratefully acknowledge financial support provided
by DGAPA-UNAM through project IN106412.

March
21,
Brazil*de*la*pena*et*al

2013

11:40

WSPC/INSTRUCTION

FILE

ZPF and emergence of the quantum

59

References
1.
2.
3.
4.
5.
6.
7.
8.
9.

10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.

T. H. Boyer, Phys. Rev. 182, 1374 (1969).


T. H. Boyer, Phys. Rev. 186, 1304 (1969).
T. H. Boyer, Phys. Rev. D 27, 2906 (1983).
T. H. Boyer, Phys. Rev. D 29, 1096 (1984).
T. H. Boyer, Sci. American 253:2, 56 (1985).
T. H. Boyer, Am. J. Phys. 71, 866 (2003).
L. de la Pe
na and A. M. Cetto, Rev. Mex. Fs. 48, Suppl. 1, 1 (2002).
L. de la Pe
na, A. Valdes-Hern
andez, and A. M. Cetto, Am. J. Phys. 76, 947 (2008).
L. de la Pe
na, A. Valdes-Hern
andez, and A. M. Cetto, in New Trends in Statistical
Physics. Festschrift in Honor of Leopoldo Garca-Colins 80th Birthday, A. Macas and
L. Dagdug, eds. (World Scientific, Singapore, 2010).
A. M. Cetto, L. de la Pe
na and A. Valdes-Hern
andez, J. Phys. Conf. Series 361,
012013 (2012).
F. Mandl, Statistical Physics (John Wiley & Sons, New York, 1988).
R. K. Pathria, Statistical Mechanics (Butterworth-Heinemann, Oxford, 1996).
E. W. Montroll and M. F. Shlesinger, J. Stat. Phys. 32, 209 (1983).
M. Planck, Ann. d. Phys. 1, 719 (1900).
M. Planck, Ann. d. Phys. 4, 553 (1901).
M. Planck, Ann. d. Phys. 37, 642 (1912).
A. Einstein, Phys. Z. 10, 185 (1909).
V. Vedral, Modern Foundations of Quantum Optics (Imperial College Press, London,
2005).
E. Santos, Am. J. Phys. 43, 743 (1975).
O. Theimer, Am. J. Phys. 44, 183 (1976).
P. T. Landsberg, Eur. J. Phys. 2, 208 (1981).
C. Cohen-Tannoudji, B. Diu, and F. Laloe, Quantum Mechanics (John Wiley & Sons,
New York, 1977), Vol. I.
P. W. Milonni, The Quantum Vacuum (Academic Press, New York, 1994).
G. R. Grimmett and D. R. Stirzaker, Probability and Random Processes (Clarendon
Press, Oxfor, 1983), chapter 4.
A. Papoulis, Probability, Random Variables, and Stochastic Processes (McGraw-Hill,
Boston, 1991).
H. Goldstein, Classical Mechanics (Addison-Wesley, Reading, 1980).
M. Hillery, R. F. OConnell, M. O. Scully, and E. P. Wigner, Phys. Rep. 106, 121
(1984).
A. Einstein and O. Stern, Ann. d. Phys. 40, 551 (1913).
W. Nernst, Verh. Deutsch. Phys. Ges. 18, 83 (1916).
R. W. James, T. Waller, and D. R. Hartree, Proc. Roy. Soc. A 118, 334 (1928).
E. O. Wollan, Phys. Rev. 38, 15 (1931).
R. S. Mulliken 1933, J. Chem. Phys. 1, 492 (1933).
T. H. Boyer, in: Foundations of Radiation Theory and Quantum Electrodynamics, A.
O. Barut, ed. (Plenum, New York, 1980).
M. Bordag, G. L. Klimchitskaya and U. Mohideen, Advances in the Casimir Effect
(Oxford U. P. Oxford, 2009).
L. de la Pe
na and A. M. Cetto, J. Math. Phys. 18, 1612 (1977).
L. de la Pe
na and A. M. Cetto, The Quantum Dice. An Introduction to Stochastic
Electrodynamics (Kluwer Academic, Dordrecht, 1996).
L. de la Pe
na, A. Valdes-Hern
andez, and A.M. Cetto, Found. Phys. 39, 1240 (2009).
L. de la Pe
na, A. M. Cetto, A. Valdes-Hern
andez, and H. Franca, Phys. Lett. A 375

March
21,
Brazil*de*la*pena*et*al

60

2013

11:40

WSPC/INSTRUCTION

FILE

de la Pe
na, Cetto and Vald
es

(2011).
39. U. Frisch, in Probabilistic Methods in Applied Mathematics, Vol. I, A. T. BharuchaReid, ed. (Academic Press, New York, 1968).
40. J. H. van Vleck, Phys. Rev. 24, 347 (1924).
41. J. H. van Vleck and D. L. Huber, Rev. Mod. Phys. 49, 939 (1977).
42. S. Hassani, Mathematical Physics. A Modern Introduction to Its Foundations
(Springer, New York, 1999).
43. M. Surdin, Ann. Inst. Henri Poincare 13, 363 (1970).
44. G. tHooft, arXiv:quant-ph/0212095v1 (2002).
45. L. de la Pe
na and A. M. Cetto, in Fundamental Problems in Quantum Physics, M.
Ferrero and A. van der Merwe, eds. (Kluwer Academic, Dordrecht, 1995).
46. P. R. Holland, The Quantum Theory of Motion (Cambridge U. Press, Cambridge,
1993).
47. E. Wigner, Phys. Rev. 40, 749 (1932).
48. J. E. Moyal, Proc. Cambridge Phil. Soc. 45, 99 (1949).
49. L. de la Pe
na and A. M. Cetto, Found. Phys. 25, 573 (1995).
50. L. de la Pe
na and A. M. Cetto, Found. Phys. 31, 1703 (2001); arXiv//quant-ph:
050101v2.
51. L. de la Pe
na and A. M. Cetto, in: Quantum Theory: Reconsideration of Foundations 3, G. Adenier, A. Yu. Khrennikov, and Th. Nieuwenhizen, eds., AIP Conference
Proceedings 810 (2006).
52. A. Valdes-Hern
andez, Investigaci
on del origen del enredamiento cu
antico desde la perspectiva de la Electrodin
amica Estoc
astica Lineal, Ph. D. Thesis (Universidad Nacional
Aut
onoma de Mexico, Mexico, 2010).
53. A. Valdes-Hern
andez, L. de la Pe
na and A. M. Cetto, Found. Phys. 41, 843 (2011);
Physica E 42, 308 (2010).
54. A. M. Cetto and L. de la Pe
na, Phys. Scripta. T151, 014009 (2012).
55. L. E. Ballentine, Quantum Mechanics (Prentice-Hall, New York, 1989).
56. A. Einstein, Phys. Zeitschr. 18, 121 (1917).
57. A. Einstein and P. Ehrenfest, Phys. Zeitschr. 19, 301 (1923).
58. A. S. Davydov, Quantum Mechanics (Addison-Wesley, Reading, 1965).
59. V. M. Fain, Soviet Physics JETP 23, 882 (1976).
60. V. M. Fain, and Y. L. Khanin, Quantum Electronics, Vol. 1: Basic Theory (Pergamon,
Oxford, 1969).
61. J. Dalibard, J. Dupont-Roc, and C. Cohen-Tannoudji, J. Phys. (Paris) 43, 1617 (1982).
62. E. A. Power, Am. J. Phys. 34, 516 (1976).
63. R. P. Feynman, Proc. Solvay Institute (Interscience, New York, 1961).
64. A. M. Cetto and L. de la Pe
na, Phys. Rev. A 37, 1952 (1988).
65. A. M. Cetto and L. de la Pe
na, Phys. Rev. A 37, 1960 (1988).
66. A. M. Cetto and L. de la Pe
na, Phys. Scr. T21, 27 (1988).
67. D. Kleppner, Phys. Rev. Lett. 47, 233 (1981).
68. P. Goy, J. M. Raimond, M. Gross, and S. Haroche, Phys. Rev. Lett. 50, 1903 (1983).
69. P. L. Knight, J. Phys. A 5, 417 (1972).
70. W. Zhou and H. Yu, Phys. Rev. D 82 124067 (2010); arXiv:1012.4055v1 [hep-th] ,
2010.
71. L. Hollberg and J. L. Hall, Phys. Rev. Lett. 53, 230 (1984).
72. C. K. Zachos, D. B. Fairlie, and T. L. Curtright, eds. Quantum Mechanics in Phase
Space (World Scientific, Singapore, 2005).
73. D. C. Cole and Y. Zou, Physics Letters A 317, 14 (2003).
74. D. C. Cole and Y. Zou, Journal of Scientific Computing 20, 43, 379 (2004).

Das könnte Ihnen auch gefallen