Está en la página 1de 79

Universit`a degli Studi di Perugia

Facolt`a di Scienze Matematiche, Fisiche e Naturali


Corso di Laurea Triennale in Fisica

Tesi di Laurea Triennale

EPR Paradox
and
Bells Theorem
Candidato:
Danny Laghi

Relatore:
Prof. Gianluca Grignani

Anno Accademico 2012-2013


Sessione di Laurea 30 Settembre 2013

Alla mia famiglia, senza della


quale questo (piccolo) traguardo
non sarebbe stato raggiunto

Entanglement is not one, but rather the characteristic


trait of Quantum Mechanics.
dinger
Erwin Schro

Contents
1 Introduction

2 The EPR Paradox

11

2.1

Definition of entanglement . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.2

Reality, Locality and Completeness . . . . . . . . . . . . . . . . . . . . . . . 15

2.3

EPR Paradox . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.4

Bohm-Aharonovs formulation of the paradox . . . . . . . . . . . . . . . . . 23

3 Bells Theorem

28

3.1

Bells Inequalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3.2

CHSH Inequalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

3.3

Optical version of Bells Gedankenexperiment . . . . . . . . . . . . . . . . . 38

3.4

A. Aspects experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.5

3.4.1

Experiment with single-channel polarizers . . . . . . . . . . . . . . . 47

3.4.2

Experiment with two-channel polarizers . . . . . . . . . . . . . . . . 48

3.4.3

Experiment with time-varying polarizers . . . . . . . . . . . . . . . . 49

3.4.4

Towards the impossibility of FTL communications . . . . . . . . . . 51

No-Communication Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.5.1

Some outlines about the FLASH Project . . . . . . . . . . . . . . 58

4 Conclusions on Quantum Mechanics


4.1

4.2

61

Interpretations of Quantum Mechanics: Copenhagen, Bohm . . . . . . . . . 63


4.1.1

Copenhagen interpretation . . . . . . . . . . . . . . . . . . . . . . . 63

4.1.2

Bohms interpretation . . . . . . . . . . . . . . . . . . . . . . . . . . 64

Open questions on Quantum Mechanics . . . . . . . . . . . . . . . . . . . . 68

A Some expectation values of Quantum Mechanics

71

Bibliography

77
6

Chapter 1

Introduction
Quantum Mechanics, as its own peculiarity, introduces some ideas that depart from everyday experience and common sense.
Until the end of the 19th century, the interpretation of macroscopic physical phenomena was based on Newtons Laws, which explained mechanical, acoustic and thermal
phenomena, and on Maxwells Equations, which governed the electric, magnetic and optical ones. These two large classes of physical phenomena yielded a distinction between the
wave-like behavior (or the electromagnetic field radiation) and the corpuscolar character
of matter, but from the beginning of the last century, physicists realized that there were
some phenomena that could not be organized within the system of classic laws, and some
of them voided the distinction between particle and field: for instance, the atom structure
and the mechanism of emission and absorption of matter radiation. Gradually, there was
the necessity of introducing a theory, without vanishing the excellent results obtained with
classical mechanics until then, which was capable to extend the latter in order to describe
even those phenomena that were remaining not understood.
There were too facts for whom there were not any explanations, according to classical
laws and statistics which were incredibly losing their own descriptive and predictive power
for certain physical phenomena, like the black body radiation spectrum and the spectral
lines of hydrogen atom.

It is in this scenary that the first quantum theory was born,

laying its foundations with the introduction of the quantum of energy, introduced in 1900
by Max Planck in order to give a plausible explanation for the spectrum of black body radiation. Planck postulated (and then showed) that the energy exchanged between matter
and radiation does not show itself in a continuous manner, but by discrete and indivisible
quantities, or quanta of energy, that must be proportional to the frequency of the radiation; he obtained an expression for the spectrum, in accordance with the experimental
distribution, by properly adjusting the constant of proportionality. So the energy quanti7

zation was the first step toward a quantum theory able to give reason for the experimental
data, incompatible with the classical theory.
The second important step was taken in 1913 by Niels Bohr, who, basing his works
on the emission spectrum of hydrogenoid atoms, succeeded in generalizing the well-known
Balmers empirical formula for the spectrum of hydrogen atom, obtaining a first, fundamental, model for hydrogenoid atoms. This model was based on some postulates, the first
of them imposing a new discretization, this time for another physical quantity that was
always been considered like a continuous variable: the angular momentum. When the
quantization of the angular momentum is considered, some orbits were imposed to the
electron (second postulate): in these orbits (called stationary levels) the electron could
not emit energy quanta.
It will be only with the introduction of two formalisms, the matrix mechanics and the
wave mechanics, introduced by Werner Heisenberg (1925) and Erwin Schrodinger (1926)
respectively, that began to emerge a real quantum theory, which was able to clear up in
a more complete way the experiments, one over all, the double-slit experiment that
Richard Feynman was used to say it holds the essential mystery, rather, the only mystery
of Quantum Mechanics.
This theory shows a double nature wave-like and corpuscular of elementary
particles, confirms the presence of discrete energy levels as assumed by Planck, and gives
an inferior limit about how much precisely it is possible to investigate on physical world,
because of the well-known Uncertainty Principle.
But undoubtedly the most peculiar feature of Quantum Mechanics is the introduction
of a wave function , solution of the Schrodinger equation and representation of the
physical state of a quantum system, for which it is the complete description.
And it is the completeness of the wave function, that is, the quantum theory behind it,
that was called into question in 1935 by Albert Einstein, Boris Podolsky and Nathan Rosen
(EPR in the following), in their famous paper Can Quantum-Mechanical Description of
Physical Reality Be Considered Complete?, where they showed through a Gedankenexperiment their so famous argument, later baptized as EPR paradox, pointing out how the
quantum theory couldnt simultaneously satisfy three principles: reality, locality and completeness. Being assumed the first and the second as obvious and incontrovertible (it was
indeed the impossibility to accept an istantaneous spooky action at a distance one of
their key point), they inferred that it was the third principle to be abandoned, that is, the
wave function could not contain all of the possible information about quantum system.
Many controversies arose from this paper, mostly of them contesting the hypotheses
on the basis of their demonstration, so that the paradox became one of the principal

subject of discussion. Two main tendencies of opposite thought standed from that: on
one hand, those belonging to the Copenhagen school, which supported a probabilistic and
undetermined vision of the reality and thought that within this interpretation the paradox
was based on inconsistent hypotheses; on the other hand, there were those which, according to the EPR paradox conclusions, were convinced of the incompleteness of Quantum
Mechanics, so that the latter had to be completed with the introduction of additional
parameters, that could explain the queer correlations that, like in EPR demonstration,
seemed to emerge between remote systems.
In this way came up the hypothesis of the introduction of additional degrees of freedom
with respect to those considered in Quantum Mechanics, which allowed to make it a
deterministic and complete theory.
Nevertheless in 1964 an Irish physicist, John Stewart Bell, demonstrated that considering local theories with additional hidden variables, there would arise some inequalities
which generally are violated by Quantum Mechanics.
The importance of this work is that the experiments have the possibility to determine
if such a theory is possible or not, according to the results stemming from laboratory;
thus the possibility of choosing what interpretation to assume for Quantum Mechanics,
either a local deterministic theory, or a nonlocal probabilistic one, is no longer arbitrary,
but it is told by the physics itself; as we shall see, by means of the experimental results
obtained and comparated with Bells predictions, Nature has showed an exclusion about
the possibility of being local.
Since 1972 several experiments, more refined as they were executed, have been conceived and conducted to verify the effectiveness of Bells inequalities (in particular, as we
shall see, a series of decisive experiments has been performed in 1980-82 by Alain Aspect
et al.).
The experiments, within an acceptable range of experimental errors, proved that:
(a) nonlocal correlations predicted by Quantum Mechanics exist;
(b) Bells inequalities are violated.

Generally the arguments we are going to deal with are still not treated in a complete
uniform manner by Quantum Mechanics books, mainly because of the subtlety of the questions and, of course, the difficulty to verify them experimentally that arise out of
these subjects and because these arguments, though some sure points have been reached,

maybe have not been completely understood in their many facets yet 1 .
The following work just wants to run through the fundamental stages in the development of EPR problem, pointing out and describing as well the several possible solutions
for this problematic argument, and analyzing the main significative experiments that have
been made in order to verify some possible theoretic models, based on alternative principles
instead of orthodox Quantum Mechanics.
It is a sure thing that the concept that Einstein involved for questioning about the
completeness of quantum theory, i.e. the entanglement of two particles far apart from
each other, is a real physical phenomenon, whatever the meaning one could assume for the
latter terms.
The life and work of Einstein, hard critic of quantum theory, just because he was not
able to accept that Nature expresses itself in a probabilistic way (God doesnt play dice
with the world

), showed that curiously the German physicist was right even when he

thought he was wrong (e.g. about the cosmological constant).


And as for the quantum world, Einsteins paper of 1935, was actually the seed for one
of the most important discoveries in physics in the twentieth century: the actual discovery
of entanglement through physical experiments.
Bells contribution, whose works have been of fundamental importance in the development of this subject, as we shall see, set out to prove that the Einstein-Podolsky-Rosen
thought experiment was far from being an absurd idea to be used just to invalidate the
completeness of the quantum theory, but rather the description of a real phenomenon. The
existence of this phenomenon had to provide a new proof in favour of Quantum Mechanics,
while against a limiting view of reality.
We shall see how the experiments have actually shown that.
1

For instance, see [10].

10

Chapter 2

The EPR Paradox


Quantum Mechanics has been founded on radical revisions of many classical concepts.
For example, in order to take into account the wave-particle duality, it had to give up the
concept of classical trajectory, because of Heisenbergs Uncertainty Principle: choosing as
canonically conjugate observables the position and momentum of a particle, it describes
quantitatively the impossibility of defining precisely and simultaneously its position and
its velocity.
According to that, the thing which Einstein thought was that quantum formalism was
incomplete. The paper of 1935 by A. Einstein, B. Podolsky and N. Rosen [2] wants to
demonstrate the non-completeness of the wave function, thought as description of reality. What EPR show is substantially a thought experiment (Gedankenexperiment is the
Deutch term) that proofs how, assuming two principles indicated as reality and locality
principle, Quantum Mechanics formalism leads to a contradiction, unless one introduces
the existence of hidden variables.
EPR show that quantum formalism can allow for the existence of states describing
two particles, for which one is able to expect strong correlations either for the velocity or
the position, even if the particles are widely separated and no longer interacting to each
other. The position measurements would always give symmetric values with respect to
the origin, so that a measure of one particle would allow one to know with certainty the
value of the position of the other.
Similarly, measurements of the momentum of the two particles would lead to opposite
values, so that the measurement on one of them would be sufficient for having knowledge
of the value of the other with certainty.
Undoubtedly, one has to choose between a precise measurement of position or momentum for one particle, because of Heisenbergs Uncertainty Principle. But the measurement
on the first particle cannot disturb the second (distant) particle, and it is here that the
11

principle of locality is introduced, by means of which EPR conclude that the second particle must have had, long before the measurement, predetermined values of position and
momentum. Since for Quantum Mechanics it is not possible to know precisely and simultaneously the values of these quantitities, Einstein and his co-authors inferred that the
theory is incomplete.
Following up the publication of this paper, E. Schrodinger coined the term entanglement, just to describe the impossibility of factorizing a quantum state like the one
assumed by EPR.
But before looking in detail the EPR paper and the paradox that arises from it, it is
best to define what we intend for entangled states and, in general, for completeness, reality
and locality.

2.1

Definition of entanglement

Entanglement is a purely quantum phenomenon, i.e. it is not supplied with a classical


counterpart; it is in this way that the quantum state of two or more physical systems
depends on the states of everyone of the systems that compose it. A first definition of
entanglement that we propose here is just the first absolute definition of the term, due to
E. Schrodinger, who conied it himself following up EPR paper. But in order to get the
Schrodingers definition, it is primarily necessary to define the important concepts of pure
states and mixed states.
Definition 1. Given a vector space X , a finite linear combination
P
vex if i [0, 1] and
i i = 1 .

i i xi

is called con-

Moreover, C X is called convex if for any pair x, y C, x + (1 )y C,


[0, 1] (and thus any convex combination of elements in C belongs to C).
Definition 2. If C is convex, an element e C is called extreme if it cannot be written
as e = x + (1 )y, with [0, 1], x, y C \{e}.
Definition 3. Let H be a separable Hilbert space ( H is separable H has finite dimension). Let S(H) be a convex subset of H.
Then the extreme elements in S(H) are called pure states. Non extreme-states are
mixed states, or nonpure states.

12

Schrodinger defined as entangled those quantum pure states |i, from an ensamble of
systems, that cannot be represented in a form of simple tensorial products of eigenstates
of the systems, that is,
|i =
6 |1 i |2 i |n i ,
where denotes the tensorial product and |i i are some vectors that represent the states
of some Hilbert spaces. The states of composed systems, that can be instead represented
as tensorial products of the subset states, constitute the complement set of pure states,
known as product states.
It is easy to determine whether a pure state of a system, composed of only two subsystems, is an entangled state making use of Schmidts decomposition theorem [1], that is
always valid for these systems 1 . Indeed any bifactorized pure state |i can be written as
a sum of bi-orthogonal terms: it is always possible to write |i in a form
|i =

ci |ui i |vi i ,

with ci C, where the sets of vectors {|ui i} and {|vi i} consist of orthonormal unit vectors
spanning the space of possible-state vectors for the system, the index i running up to the
smaller of the dimensions of the two subsystem Hilbert spaces.
On the basis of our previous considerations, it is easy to get a second more formal
definition:
Definition 4. Let A and B be two non-interacting systems with their respective Hilbert
spaces HA and HB , and let |iA and |iB be two states in which there are the first and
the second system respectively. Then the state of the composed system, belonging to the
Hilbert space given by the tensorial product HA HB , is |iA |iB .
The states of the composed system that can be written in this way are called product
states or separable states.
Generally, not all of the states of a composed system are product states; let {|ui iA }
and {|vi iB } be some bases in their respective Hilbert spaces HA and HB . According to
1

In the simplePcase of finite dimensional spaces, Schmidts decomposition


theorem asserts that any
P
double sum = kl XP
a

, by means of unitary
kl xm yn , can be converted into a single sum =

P
transformations =
k Uk xk and =
l Vl yl . If {xk } and {yl } are two orthonormal bases for
two distinct vector spaces, then { } and { } are also two, possibly complete, orthonormal vector bases
for these spaces. The absolute values |a | are called the singular values of the matrix X, and are easily
calculated by noting that |a |2 are the nonvanishing eigenvalues of the Hermitian matrices XX and X X.
The corresponding sets of eigenvectors are { } and { }, respectively.

13

Schmidts decomposition, any state in the composed system can readily be written as
|iAB =

ci |ui iA |vi iB .

(2.1)

Then the state (2.1) is a product state, or a non-entangled state, if there is one and
only one i so that ci 6= 0, and all of the other cj = 0, j 6= i; that is, a state is non-entangled
if and only if
|iAB = ci |ui iA |vi iB ,

i fixed.

(2.2)

If instead there are more than one ci 6= 0 for the state 2.1, then such a state is called
entangled state.
If i = 1, . . . , N and if ci =

1 , i,
N

the state (2.1) is called to be maximally entangled.

In other words, a state is entangled if and only if it cannot be factorized.


As an explanatory example of this definition, lets consider two bases {|0iA , |1iA } HA
and {|0iB , |1iB } HB .
Then an entangled state of the composed system HA HB is represented by
1
(|0iA |1iB |1iA |0iB ) ,
2
that is, omitting tensorial product symbols,
1
(|0iA |1iB |1iA |0iB ) ,
2
where it is evident the impossibility of attributing separately a quantum state either to
the system A or to the system B without involving the other.
A similar state to the one above is a singlet state for a pair of electrons. Thus if we
knew, by experimental measurements, that for example the system A would be in the state
|1iA , then, unless of coefficients that we can neglect, we will able to know automatically
that the system B will be in the state |0iB .
How to obtain entangled states
There are some different ways for achieving entangled states of two systems.

14

For instance, in the decay of a spin 0 particle, the final products are two spin 1/2
particles in a singlet state (like 0 + e+ + e ).
Another example is in the e+ e annihilation, that can produce a meson, that can
0 in a spin singlet state.
decay, for instance, in a couple of K 0 K
Similarly it is possible to create photons entangled in polarization as result of positronium annihilation (e+ + e + ).
One more example of entangled state is the above mentioned singlet spin of two electrons near an atomic nucleus.
Entanglement has numerous applications beyond its use in the paradox topic as well;
for instance an application of such states is in Quantum Information Theory, in Quantum
Cryptography, where in the one time pad transmission, a couple of entangled particles is
used to exchange a random coding key between sender and receiver, in a completely secure
way from eavesdropping attacks.

2.2

Reality, Locality and Completeness

As we already said, EPR paper relies substantially upon the assumption of some principles,
i.e. the completeness, the reality and the locality. So it is good to find for these concepts
some definitions applicable to our contest [2].
Principle of reality
The elements of physical reality cannot be determined by a priori philosophical considerations, but they are to be found using measurements and experimental results. Far from
being an exhaustive definition, EPR consider as satisfactory the following definition for
an element of physical reality.
If, without in any way disturbing a system, we can predict with certainty (i.e., with
probability equal to unity) the value of a physical quantity, then there exist an element of
physical reality corresponding to this physical quantity, that is, an objective property of
the system, independent of any eventual external observer.
Thus, all of the possible physical observables must have some preexistent values for
any possible measurement, before of carrying out that measurement.

15

Principle of locality
Accordingly to the existence of an upper limit for the velocity of propagation of a signal,
coincident with the speed light c in vacuum, an object is affected only directly by his
immediate neighbourhood. This lead us to a definition of locality.
Given two physical systems, lets suppose that during a certain time interval they
remain isolated from each other. Then the temporal evolution of the physical properties
of one of them during that time interval cannot be influenced by operations executed on
the other (Einsteins locality).
Condition of completeness of a theory
According to the definition adopted by EPR in their paper (called by them condition of
completeness):
In order for a theory to be complete, it is necessary that any elements of physical
reality must have a counterpart in a physical theory.
Whether an element of physical reality has no counterpart in the theory, then we shall
say that the theory is not complete.

2.3

EPR Paradox

Given two noncommuting physical quantities, because of the Uncertainty Principle, the
exact knowledge of one of them precludes the exact knowledge of the other. In fact, any
attempt of determining a quantity by means of measurement changes the system state, so
that the value of the other quantity is undetermined.
In other words, two noncommuting physical quantities cannot have both expected
values.
Starting from this assumption, EPR state that:
1. either the Quantum Mechanics description of reality given by the wave function is
not complete;
2. or the physical quantities associated with noncommuting operators cannot have simultaneous reality.

16

It can be easily shown that one of the above conclusions has to be necessarily true, for
there are no other possible situations.
Actually, if we supposed ab absurdum that both the assumptions were wrong, we
could consider two noncommuting physical quantities A, B, that have simultaneous reality
(negation of 2.). Then, for the principle of reality, both the solutions should have some
defined values that, by means of the completeness condition, would be part of the complete
description of reality. But then, if the wave function was able to provide such complete
description of the reality (negation of 1.), it would include those values, which would
be therefore predictable. But this is an absurd, for it would contradict the Uncertainty
Principle.
In this way, we have showed that 1. and 2. are complementary: whether we assume
the first as false, then the second must be necessarily true, and vice versa.
EPR create the paradox assuming the completeness of the wave function and, at the
same time, deducing the physical reality for two quantities, whose associated operators do
not commute, that is, inferring that both 1. and 2. are false, thus obtaining an absurd,
as we have seen above.
So they conclude that, in order to solve the paradox, one of the three assumptions has
to be dropped out, that is one amongst:
the principle of reality
the principle of locality
the completeness of Quantum Mechanics.
Since it seemed that the first two hypotheses could not be debated, unless of having a
senseless theory (but actually it is just the debate on these assumptions to open to other
possible solutions of the paradox), EPR drop out the validity of the remaining assumption,
that is, Quantum Mechanics is an incomplete theory.
We now shall see what are the arguments of EPR paper and how the paradox arises
from it, i.e. how to get both the negation of 1. and 2..

Paradox formulation
It is helpful to run through the sheer introductory passages of the EPR paper [2].
Lets consider a particle having a single degree of freedom, i.e. in a unidimensional
space. The fundamental concept of the theory is the concept of state, which is supposed
17

to be completely characterized by the wave function , which is a function of the variables


chosen to describe the particles behaviour. Corresponding to each physically observable
quantity A there is an operator, which may be designated by the same letter (with a hat).
that is, if:
If is an eigenfunction of the operator A,
= a,
A

(2.3)

where a R, then the physical quantity A has with certainty the value a whenever the
particle is in the state described by . In accordance with the principle of reality, for this
particle there is an element of physical reality corresponding to the physical quantity A.
For instance, lets consider,
i

= e ~ p0 x ,

(2.4)

where p0 is a numerical constant and x the independent variable. Taking the operator
corresponding to the linear momentum of a particle,
p = i~

,
x

(2.5)

we obtain
p = i~

= p0 ,
x

(2.6)

and therefore, in the state (2.4), the momentum assume with certainty the value p0 . The
momentum of the particle therefore has physical reality.
Now lets consider the case in which the eigenvalue equation (2.3) is not valid. In this
case we can speak no longer of the quantity A as if it had a definite value. Such a case
is that, for example, of the position of the particle. Its corresponding operator, x
, is the
operator of multiplication by the independent variable x. Hence,
x
= x 6= a.

(2.7)

In accordance to Quantum Mechanics, it is just possible to say that the relative probability of a coordinate measurement giving a result lying between a and b is
Z
P (a, b) =

dx = b a.

dx =

(2.8)

Since this probability is independent of a, but depends only upon the difference b a,

18

we see that all of the values for the coordinate are equally probable. From another point
of view, we could have obtained such a conclusion without any computation, noting that
the state (2.4) is representative of a unidimensional free wave of definite momentum p0 ,
so it is not possible to localize exactly a free wave, which extends as far as infinity, for it
has an equivalent probability of being in any point of the (unidimensional) space.
So a defined value of the position of a particle in a state given by (2.4) is unpredictable,
but it can be reached only by direct measurement. In accordance with Quantum Mechanics, this action alters the particle state, that becomes a position eigentstate (and no longer
of momentum). Therefore, after the determination of the position, the particle will be no
longer in the state given by Eq. (2.4).
We can conclude that whenever the momentum of a particle is known, its position has
no physical reality.
Now, supposing that the wave function of a certain physical system is complete (negation of 1.), yet Einstein, Podolsky and Rosen found that it is possible to have two simultaneous real physical quantities in spite of the fact that their respective operators do not
commute (negation of 2.), that is an absurd.
In order to deduce such an absurd, EPR built the following reasoning.
Lets consider two systems I and II (which could be punctiform or hard rigid bodies, or
whatever other generic physical systems), described by the variables x1 and x2 respectively.
We suppose that they interact with each other from the time t = 0 to the time t = T .
We suppose further that the states of the two systems before t = 0 are known. Then, by
means of the wave function , it is possible to describe the state of the entangled system
I + II for any successive time t > T using the Schrodingers equation, even if, after the
interaction, it is no longer possible to work out the states of I or II, for they are no longer
factorable states.
Now lets consider the physical quantity A, relative to the system I, which has some
eigenvalues a1 , a2 , a3 , . . . and eigenfunctions u1 (x1 ), u2 (x1 ), u3 (x1 ), . . . .
The wave function , considered as function of x1 , at a certain fixed time t > T , can
then be expressed as:
(x1 , x2 ) =

n (x2 )un (x1 ),

(2.9)

n=1

where x2 stands for the variables used to describe the second system. Here n (x2 ) are to
be regarded merely as the coefficients of the expansion of into a series of orthogonal
functions un (x1 ), i.e. as Fouriers coefficients (un (x1 ), ).

19

Now we want to measure A; suppose that it is found that it has the value ak .
According to the precipitation postulate, after a measurement, the state of the system
is given by 0 = P , where P is a projection operator on the eigenspace relative to
the found eigenvalue. Since the system I has fallen in the state described by the wave
function uk (x1 ), we get:
(x1 , x2 ) = k (x2 )uk (x1 ).

(2.10)

Since for systems formed by independent subsystems the wave function is the product
of the wave functions of the single subsystems, from Eq. (2.10) we deduce that, after the
measurement, the second system will collapse into a definite state too, which is described
by the wave function k (x2 ).
So the wave packet given by the infinite series (2.9) is reduced to a single term given
by (2.10).
We now propose to consider, instead of the physical quantity A, another physical quantity, say B, relative to the system I as well, with eigenvalues b1 , b2 , b3 , . . . and eigenfunctions
v1 (x1 ), v2 (x1 ), v3 (x1 ), . . . .
In this way, analogously to Eq. (2.9), we can write
(x1 , x2 ) =

m (x2 )vm (x1 ),

(2.11)

m=1

where m (x2 ) are the new coefficients.


In analogy with the previous passages, we measure B and suppose we find it in a
value bj ; then the system collapses into the state described by the wave function vj (x1 );
consequently we obtain
(x1 , x2 ) = j (x2 )vj (x1 ),

(2.12)

from which we deduce that system II has collapsed into the state described by the wave
function j (x2 ).
We see therefore that, as a consequence of the two different measurements performed
upon the first system, the second system may be left in states with two different wave
functions: k (x2 ) and j (x2 ), and, since at the time of measurement the two systems I,
II no longer interact, no real change can take place in the second system in consequence
of anything that may be done in the first system.

20

Thus, it is possible to assign two different wave functions, k and j , to the same
reality (the second system after the interaction with the first). Therefore, if we were able
to choose two functions for the eigenfunctions belonging to two noncommuting operators,
relative to two, non compatible, physical quantities P and Q, which assume the values pk
and qj respectively, then we would automatically obtain the proof that two noncommuting
quantities are simultaneously real (negation of 2.), that is exactly our aim.
To show this, lets consider two generic systems, for example two point-like particles.
We suppose that the wave function of the total system is:
+

Z
(x1 , x2 ) =

e ~ (x1 x2 +x0 )p dp,

(2.13)

where x0 is a constant.
Choosing as physical quantity A the observable linear momentum P , to which is naturally associated the operator P of the form

P = i~ ,
x
and measuring that quantity for the particle I, it will assume a value p; consequently the
particle will fall into a state described by the eigenfunction
i

up (x1 ) = e ~ x1 p .

(2.14)

From now on, we consider the case of a continuous spectrum; hence Eq.(2.9), representative of the wave function of the total system, is
Z

(x1 , x2 ) =

p (x2 )up (x1 )dp;

(2.15)

the projection on up (x1 ) will naturally lead to


i

p (x2 ) = e ~ (x2 x0 )p ,

(2.16)

that is nothing but the eigenfunction of the operator P = i~ x 2 , relative to the eigenvalue p for the momentum of the second particle.
On the other hand, if we choose as physical quantity B the observable position Q,
of multiplication by x, and performing a measurement
whose operator is the operator Q
on the particle I, the latter will assume a value x, and consequently it will fall into the

21

state described by the eigenfunction


vx (x1 ) = (x1 x),

(2.17)

where (x1 x) is the well-known Dirac delta-function.


Eq. (2.11), representative of the wave function of the total system, now becomes
Z

e ~ (xx2 +x0 )p dp;

(2.18)

e ~ (xx2 +x0 )p dp = h(x x2 + x0 ),

(2.19)

vx (x1 )dx1

(x1 , x2 ) =

so we clearly have
Z

x (x2 ) =

= x2 , relative to the eigenvalue x + x0 for the


that is the eigenfunction of the operator Q
position of the second particle.
are
At this point the paradox is obtained. In fact it is immediate to verify that P and Q
a pair of quantum operators associated with a couple of canonically conjugate variables,
which do not commute:
h

i
= i~.
P , Q

(2.20)

Finally, we have obtained two wave functions, k and j , associated with two non
compatible quantities, that is, two noncommuting quantities P and Q, belonging to the
second system, that became simultaneously real: thus it follows the paradox.
Starting from the assumption that the wave function gave a real complete description
of physical reality, EPR came to the conclusion that two physical quantities, with noncommuting operators, could have simultaneous reality. Consequently the negation of 1. leads
to the negation of the only alternative 2.. Thus we conclude that the quantum mechanical
description of physical reality given by the wave function is not complete.
EPR conclude their paper saying that one could object them that the criterion of
reality they have chosen is not sufficiently restrictive. Indeed, we would not arrive to
a paradox if we just insisted on the point that two or more physical quantities could be
viewed as simultaneous elements of reality only whether they could be measured or predicted
simultaneously. From this point of view, since either one or the other, but not both
simultaneously of the quantitites P and Q could be expected, they are not simultaneously
real and the paradox therefore would not arise.
Nevertheless, this procedure is such as to make the reality of P and Q depending on
22

the measuring process that is executed on the first system, which does not interfere in
any way with the second one. And it is here that the principle of locality makes its part,
preventing from taking into account that possibility. With EPR words:
definition of reality could be expected to permit this

No reasonable

[2].

The paper ends opening to the search of new possible methods capable to describe
completely the physical reality, i.e. seeking for a new theory comprehensive of the lacking
information, so that, with the addition of additional parameters it would be able to completely describe the physical reality.

2.4

Bohm-Aharonovs formulation of the paradox

Now we shall examine another formulation of the EPR paradox, following the example
ideated by David Bohm and Yakir Aharonov [4]. In this model, a quantum system is considered from the point of view of its own spin variables. This formulation of the paradox
is more handy and comfortable from a mathematical point of view, since, working with
spins, the operators and the eigenfunctions will become matrices and vectors respectively,
reducing all of the problem to a simple matricial calculus; moreover, the theoretical formulation of the problem is nearer to the experimental situations where it has been really
examinated.
Lets consider a diatomic molecule of spin 0, whose atoms (I and II) are our two
systems taken into account, having spin 1/2. It is straightforward that the wave function
of the composed system will be given by a singlet state of spin. We suppose that the
couple of atoms interact for a certain time and then become entangled; after that time
they are separated. From the time of separation on, there is no longer any interaction
between them.
We can certainly define the wave function of the composed system, even if we do
not know the functions that describe the state of each single atom separately; omitting
the orbital part, which is quite inessential for the description of the problem, the state
describing the composed system is a singlet state of spin:
1
= [+ (I) (II) (I)+ (II)] ,
2

(2.21)

where (I) e (II) are the vectors that represent the state of the first and the second
particle respectively, with spin ~2 on the direction of the axis z. In that state, for the
conservation of the angular momentum, the total spin always must vanish; whether one
23

of the two atoms assumes a positive value, the other will be certainly negative.
Given the description of Bohm-Aharonovs model, it is easy to see that the quantities
to be measured are two of the three components of the spin between one of the two atoms,
that we know from Quantum Mechanics to be incompatible; that is, the relative operators
generally obey to the relation:
[Si , Sj ] = i~ijk Sk .
In order to see this, we maintain an analogy with EPR steps, choosing as physical
quantity the Sx component of the spin for I; measuring it, we shall obtain two possible
values: + ~2 and ~2 .
In the first case, as we know from the knowledge of the eigenvectors of Sx , the atom I
collapses into the state given by
1

1
+ (I) =
2

!
,

eigenvector of the operator Sx , relative to the assumed eigenvalue + ~2 .


Therefore, for the conservation of angular momentum, it is necessarily
= + (I) (II),

(2.22)

from which we also deduce the state of the second atom, which will have a wave function
given by
1
(II) =
2

eigenvector of the operator Sx , relative to the eigenvalue ~2 of the spin for atom II.
In the second case all is analogous, but with inverted eigenvalues and eigenvectors for
the two atoms.
Choosing as physical quantity B the component Sz of the spin for I, and measuring
it, we will get the same two possible values of the previous case.
So we see that whether the observable assumes a value + ~2 , then the atom I collapses
into the state given by
+ (II) =

24

1
0

!
,

eigenvector of the operator Sz , relative to the assumed value + ~2 .


Then the state of the total system will be
= + (I) (II),
from which we can work out a state for the second atom, that will be equal to
(II) =

0
1

!
,

eigenvector of the operator Sz , relative to the eigenvalue ~2 of spin for atom II.
In accordance with what we saw for Sx , when the measure of the observable B assumes
the second of the two possible values, it is all analogous, but with inverted eigenvalues and
eigenvectors for the atoms.
In this way Bohm and Aharonov demonstrate that the generic quantities A and B
are reduced to a pair of observables (in our example: Sx and Sz ), whose operators do not
commute. Then the two wave functions, called by Einstein k and j , represent two states
of simultaneous reality for these operators, related to the second atom (in the example:
+/ (II) for the Sx -spin and +/ (II) for the Sz -spin).
The result obtained is the same as that obtained by Einstein, Podolsky and Rosen:
the steps followed are equivalent. Indeed both start considering a pair of incompatible
quantities, and they successively measure these observables on one of the two considered
systems. Both of them finally demonstrate that, measuring these quantities on one of the
two systems, they are able to determine these values on the other system too, but without
altering its state; that is, both demonstrate that the wave functions for the quantitites
of the second systems, could correspond with the eigenfunctions of two noncommuting
operators, related to two incompatible physical quantities.

In order to solve the contradiction of both the examined formulation, one can drop
one of the three hypotheses assumed, that we list here again:
i) Principle of reality,
ii) Principle of locality,
iii) Completeness of Quantum Mechanics.

25

According to EPR, the hypothesis to be abandoned is nothing but the completeness


of Quantum Mechanics, since, as they have demonstrated, the quantum description of
reality, provided by the wave function, comes out to be inconsistent with the assumed
principles. As well, since completeness of Quantum Mechanics means indeterminism,
they are sure of the need to complete the latter by means of a more fundamental theory,
in which the incompleteness is overwhelmed introducing additional dynamical parameters,
the so-called hidden variables: in this way, in addition to the solution of the paradox,
we would get back an essentially deterministic vision of the world, in which it is possible
to associate with certainty a definite value for every quantity, i.e. a definite element of
reality. As a matter of fact, the weak point of their proof is precisely to take for granted
the assumptions i) and ii), as we already said, but this is not to be considered as a gross
mistake of the authors, but rather a historical limit of their work.
It is better to stress again the importance of one of the critical point of the argumentation of Einstein and collaborators. Expressing the principle of reality for the Gedankenexperiment of Bohm-Aharonov, it states that, if one can execute an operation, that allows
him to predict with certainty the value of a measurement for a quantity without disturbing
the measured spin, then the measurement has a definite value, apart from the fact that this
operation is actually executed or not.
In accordance with the Copenhagen interpretation that is the same of Bohr who,
by the way, replied with an article [3] to that of EPR as follows the concept of reality
cannot be legitimately applied to a property unless there is a device able to measure it,
whereas Einstein intended this vision as anthropocentric, assering that instead physical
systems have intrinsic properties, apart from the fact that they are observed or not.
So it is here that arises the most large difference between the two currents of thought,
each of them defending their own position with their arguments.
But it is important to point out at once the concept of simultaneity, used more than
once by EPR, since this word was a surprising expression for people who knew very
well that this term was undefined in the teory of relativity. Let us examine this issue
with Bohms singlet model. [5] One observer, conventionally called Alice, measures the
z-component of the spin for her particle and find +~/2. Then she immediately knows
that if another distant observer, Bob, measures (or has measured, or will measure) the
z-component of the spin for his particle, the result is certainly ~/2. One could then ask
when Bobs particle acquires the state with sz = ~/2. This question is meaningless, but
it has a definite answer: Bobs particle acquires this state instantaneously in the Lorentz

26

frame that we arbitrarily choose to perform our calculations. Of course, Lorentz frames
are not material objects: they exist only in our imagination. When Alice measures her
spin, the information she gets is localized at her position, and will remain so until she
decides to broadcast it. Absolutely nothing happens at Bobs location. From Bobs point
of view, any spin directions are equally probable, as can be verified experimentally by
repeating the experiment many times with a large number of singlets without taking in
consideration Alices results. Thus, after each one of her measurements, Alice assigns a
definite pure state to Bobs particle, while from Bobs point of view the state is completely
random. It is only if and when Alice informs Bob of the result she got (by mail, telephone,
radio, or by means of any other material carrier, which is naturally restricted to the speed
of light) that Bob realizes that his particle has a definite pure state. Until then, the two
observers can legitimately ascribe different quantum states to the same system. For Bob,
the state of his particle suddenly changes, not because anything happens to that particle,
but because Bob receives information about a distant event. So reality might differ for
different observers.
Anyway, the conclusions reached by EPR raised a discussion that nowadays is still
in other ways object of heated debates.
Is it really possible to complete Quantum Mechanics?
In other terms: is it possible to consider quantum states as averages over those states
for which the results of any possible measures are, in principle, completely determined?

27

Chapter 3

Bells Theorem
There were two principal currents of thought that wanted to give an explanation to the
question opened with the paper of Einstein, Podolsky and Rosen: on one hand there
were the determinists, i.e. those people who believed that Quantum Mechanics should
be extended in a theory with local hidden variables 1 ; on the other hand, there were
the indeterminists, that instead did not consider as right the formulation of the paradox
because of its wrong assumptions.
In 1964 the Irish physicist John Stewart Bell published the paper On the EinsteinPodolsky-Rosen Paradox [6], where he provided an elegant solution that could be verified
experimentally, in order to consolidate the more or less validity of EPR argumentations.
He proved, through his inequalities, that any possible theory with hidden variables,
with the requirement of locality, is in contradiction with the statistical previsions of Quantum Mechanics, e.g. the former cannot be considered as the completion of the latter, but
they are to be set on different planes. As we shall emphasize several times, it is best to
notice from now on that a possible interpretation with nonlocal hidden variables is not
forbidden; in such a case, the conflict does not arise.
We can illustrate and summarize the Bells Theorem as follows:
(Bells) Theorem. There is no local hidden variable theory that can reproduce (all) the
predictions of Quantum Mechanics.
Bell demonstrated that for certain phenomena local reality implies some constraints
i.e. the inequalities , which are not required but violated by Quantum Mechanics.
The experiments that have been carried so far have shown an evident violation of Bells
1

Here we assume local in the sense of Einstein, that is, remote systems do not interact with each others
and they behave as if they were independent.

28

inequalities; so we get an empiric evidence against local reality, showing that some of the
spooky actions at a distance expected by EPR, actually occur.
Nevertheless, the principles of special relativity are not violated. Indeed, it has been
proved theoretically and experimentally that, because of the No-Communication Theorem,
it is impossible for one experimenter to use these special quantum effects to communicate
information to another with a velocity faster than light.
As a first point, Bell was interested in formulating Einsteins theory within a mathematical frame. In order to do that, he based his work on Bohms formulation of the
paradox, considering a couple of spin-1/2 particles in a singlet spin state. The EPR additional variables should have had the aim of making the theory causal and local. What
Bell was trying to (and finally) show is that such an operation is incompatible with the
statistical previsions of Quantum Mechanics.

3.1

Bells Inequalities

We consider a couple of 12 -spin particles, moving freely in opposite directions and forming a
system in a singlet spin state, of the form (2.21). It is possible to make measurements, e.g.
by Stern-Gerlach magnets, on selected components of the spins ~1 and ~2 . From Quantum
, where a
is a
Mechanics we know that, whether a measurement of the component ~1 a
will give with certainty value 1
unit vector, yields value +1, then measurement of ~2 a
and vice versa.
We now introduce the principle of locality seen in the last chapter, that establishes that
whether the measurements are made at places remote from one another, the orientation of
one magnet does not influence the result obtained with the other. Since we can predict in
advance the result of measuring any chosen component of ~2 after a measurement of the
same component of ~1 , it follows that the result of any such measurements must actually
be predetermined.
Since the wave function describing the initial state does not determine the result of
an individual measurement, this predetermination implies the possibility of adding some
new variables that would represent some properties intrinsic to any considered pair of
particles, and that are not considered by the wave function just because of their difference
from couple to couple. These additional parameters are the so-called hidden variables,
that Bell labels with the letter . It is the same argument whether stands for a single
variable or a set of variables, or even a set of functions, or if the variables are discrete or
continuous. We consider for simplicity as a single continuous parameter.

29

is determined by the
With such addition, the result A of the measurement of ~1 a
is in a
and by the parameter , and the result B of the measurement of ~2 b
vector a
and , that is
similar manner determined by b

A(
a, ) = 1;
) = 1.
B(b,

(3.1)

The principle of locality, the crucial point in the argumentation of Einstein, Podolsky
and Rosen, implies that the result B for particle 2 does not depend from the arrangement

of the magnet for particle 1, neither A from b.


a
Defining by () the probability distribution of , then the expectation value of the
is
and ~2 b
product of the two components ~1 a
=
P (
a, b)

).
d()A(
a, )B(b,

(3.2)

This value should be equal to the quantum mechanical expectation value, which for a
singlet spin state is

D
E
=

~2 b
~1 a
a b.

(3.3)

Actually, as will be shown below, these two expressions are not equal and they are
mutually exclusive.
Since () is a probability distribution, it is normalized,
Z
d() = 1,

(3.4)

and because the relations (3.1) are valid, (3.2) cannot be less than 1. It could reach the
if
=b
value 1 only for a
),
A(
a, ) = B(b,

(3.5)

except for a set of points of zero probability. Thus (3.2) becomes


=
P (
a, b)

).
d()A(
a, )B(b,

, another unit vector; therefore, with the help of (3.1),


Now define c
2

For an explicit evaluation, see Appendix A.

30

(3.6)

Z
h
i

) A(
) = d() A(
P (
a, b) P (
a, c
a, )A(b,
a, )A(
c, )
Z
h
i
) A(b,
)A(
=
d()A(
a, )A(b,
c, ) 1 ,

(3.7)

from which

Z
h
i

)A(
) d() 1 A(b,
a, b) P (
a, c
c, ) .
P (

(3.8)

c
), then
For (3.6), the second term of the RHS of the last Eq. is equal to P (b,




c
P (
) P (
) .
1 + P (b,
a, b)
a, c

(3.9)

The inequality (3.9) is the first of a family of inequalities that are collectively called
Bells Inequalities.
This inequality sets a constraint on the expectation values with hidden variables, valuated along the three considered directions.
c
c
| for small |b
|. Consequently, the value
Indeed, generally the RHS is of order |b

) cannot be stationary at the minimum value (-1 for b = c


) and therefore cannot
P (b, c
be equal to the quantum mechanical result (3.3) in these conditions.
Moreover neither quantum mechanical correlation (3.3) can be approximated arbitrarily by (3.2).
Just to prove it, we now suppose to make more repeated experiments. The directions
c
, b,
will not be strictly determined, but will lie into right circular cones
of unit vectors a
of small width since physically the directions are always affected by certain errors.
Accordingly, instead of (3.2) and (3.3) we consider their averaged functions

P (
a, b)

0
b

b,

a b,

(3.10)

0 ) and
0 over vectors a
0 and
where the bar denotes independent average of P (
a0 , b
a0 b
We suppose that for any unit vector a
and b.
and
within specified small angles of a

P (
a, b) approximates quantum mechanical results well, that is, the difference between

the averages is bounded on the upper side by a small quantity :





P
(
a
,
b)
+
a

b

.
31

(3.11)

Our aim is to show that cannot be made arbitrarily small, making the impossibility
of (3.2) to be equal to (3.3).

and b
Lets consider that for all a



;
ba
b
a

(3.12)






+a
P (
+a
+ a
a
+ .
b
b
b
b
a, b)
a, b)
P (

(3.13)

then

From (3.2) we can define


=
P (
a, b)

) 1,
d()A(
a, )B(b,

(3.14)

where


A(
a, ) 1




B(b, ) 1.

(3.15)

we get
= b,
From (3.13) and (3.14), taking a
Z

h
i
)B(b,
) + 1 + .
d() A(b,

(3.16)

Now, from definition (3.14) it turns out that

h
i
) A(
d() A(
a, )B(b,
a, )B(
c, )
Z
h
i
) 1 + A(b,
)B(
a, )B(b,
c, )
=
d()A(
Z
h
i
)B(b,
) .

d()A(
a, )B(
c, ) 1 + A(b,

P (
) =
P (
a, b)
a, c

With the help of the two conditions (3.15),

Z


h
i

)B(
)
a, b) P (
a, c
d() 1 + A(b,
c, )
P (
Z
h
i
)B(b,
) .
+
d() 1 + A(b,
Making use of (3.14) and (3.16),
32

(3.17)

c
) 1 + P (b,
) + + .
a, b) P (
a, c
P (

(3.18)

Finally, using inequalities (3.13) we have for the RHS,

c
c
c
c
) + + = 1 + P (b,
) + b
b
++
1 + P (b,




c
) + b
b
++
1 + P (b,
c
c
c
+ 2( + ),
1b

(3.19)

while for the LHS,

P
(
a
,
b)

P
(
a
,
c
)
=
P
(
a
,
b)
+
a

P
(
a
,
c
)
+
a

c









+a
P (
P (
b
) + a
c

c
a
b
a
a, b)
a, c



2( + ),
c
a
b
a
(3.20)
from which, for (3.19) and (3.20), (3.18) becomes



2( + ) 1 b
c
c
a
b
+ 2( + ),
a

(3.21)




+ b
c
c
a
b
1.
4( + ) a

(3.22)

that is

The interesting result here is that this inequality is not always verified for small values
of (as on the contrary it should be in order to make equal the expectation values (3.2)
=b
c
c
= 0 and a
b
= 12 , then
and (3.3)). Actually, if we consider, for instance, a
from (3.22) we obtain
4( + )

2 1,

(3.23)

from which we conclude that, for little , cannot be made arbitrarily small. This is
the proof that the difference between the expectation values (3.2) and (3.3) is necessarily
finite, so that their corresponding theories are not compatible with each other.
Therefore, it could be reasonable to debate the principle of locality, or even to state
that if a theory with additional parameters wants to explain in a right manner quantum
mechanical rules, it necessarily has to be nonlocal, e.g. it should be in such a way that it
33

could allow for any measurement made on a system to influence any further measurement
made on other systems at any distance. Finally it is now justified the statement of Bells
Theorem given at the beginning of this chapter.
It is important to emphasize that Bell does not exclude local realistic theories, but
simply demonstrates that these latter would be in contradiction with quantum previsions,
so these cannot be considered as a complement of a quantum mechanical theory. Admitting
the existence of hidden variables automatically cancels the hypothesis of locality, but this
does not mean that there would be absolutely no possibility of making a different model
of additional parameters which could resolve the paradox and, at the same time, observe
quantum mechanical predictions.
For the moment, the problem raised by Bell is only the need to choose one or the other,
otherwise we have to renounce the concept so intuitive and taken for granted of
locality, the price to pay for allowing the introduction of a new theory with hidden variables
that would be consistent with Quantum Mechanics.

3.2

CHSH Inequalities

John Clauser, Michael Horne, Abner Shimony and Richard Holt (in the following CHSH),
in a paper of 1969 [7] generalized Bells inequalities in a way so that these could be applied
to realizable experiments. Successively (1971), a generalization of Bellinequalities would
be given by Bell himself as well [6].
The CHSH inequalities are based upon a combination of four correlation coefficients
measured on four different orientations of the polarizers a definition for coefficient of
polarization will be given in the next section as well.
Lets consider a set of couple of entangled particles that move one in the opposite
direction of the other, so that one enters apparatus Ia and the other apparatus IIb , where
a and b are adjustable apparatus parameters. In each apparatus, a particle must select
one of the two channels labeled +1 and 1 respectively. Let the results of selections be
represented by A(a) and B(b), where a and b represent the adjustable parameters of the
two measurement apparatus. So A(a) and B(b) can assume values 1 according as the
first or second channel is selected.
Suppose now that there is a statistical correlation of A(a) and B(b), due to information
carried by and localized within each particle, and that at some time in the past the particles
constituting one pair were in contact and communication regarding this information. The

34

information, which emphatically is not quantum mechanical, is part of the content of a


set of hidden variables, denoted collectively by .
The results of the two selections must be deterministic functions, denoted by A(a, )
and B(b, ).
A reasonable request of locality requires A(a, ) to be independent of the parameter
b, and B(b, ) to be likewise independent of a, since the two selections may occur at an
arbitrarily great distance from each other.
Finally, since the pair of particles is generally emitted by a source in a manner physically independent of the adjustable parameters a and b, we assume that the normalized
probability distribution, (), characterizing the ensamble is independent of a and b.
We can now define a correlation function as
Z
P (a, b) =

d()A(a, )B(b, ),

(3.24)

where is the total space of the hidden variables .


Using the definition (3.24) it follows that

Z
|P (a, b) P (a, c)|

d() |A(a, )B(b, ) A(a, )B(c, )|

Z
d() |A(a, )B(b, )| [1 B(b, )B(c, )]

=
Z

d() [1 B(b, )B(c, )]


Z
= 1 d()B(b, )B(c, ).

(3.25)

Lets suppose that for some parameters b0 and b we have P (b0 , b) = 1 , where
0 1. Experimentally interesting cases will have close to but not equal to zero.
Note as well that in this point of the argumentation we are avoiding the experimentally
unrealistic restriction of Bell that for some couple of b0 and b there is perfect correlation
(i.e., = 0).
Dividing the region into two regions + and such that = { | A(b0 , ) = B(b, )},
from (3.24) we have

35

P (b , b) =

d()A(b0 , )B(b, )

d()A(b0 , )B(b, )
d()A(b , )B(b, ) +

Z
Z +
d() [B(b, )]2
d() [B(b, )]2
=

+
Z
Z
d()
d()
=

+
Z
Z
Z
= 1 =
d() =
d() +
d() ,
0

from which
Z

d() = .
2

We can therefore write the integral of RHS of (3.25) as

d()B(b, )B(c, ) =

d()B(b, )B(c, ) +
+

d()B(b, )B(c, )

d()A(b , )B(c, )

d()A(b0 , )B(c, )

d()A(b0 , )B(c, ) +

=
+

d()A(b0 , )B(c, ) +

Z
2
d()A(b0 , )B(c, )

Z
Z
0
=
d()A(b , )B(c, ) 2
d()A(b0 , )B(c, )

Z


P (b0 , c) 2
d() A(b0 , )B(c, )

= P (b , c) = P (b0 , c) + P (b0 , b) 1,

(3.26)

therefore from (3.25) and (3.26),


|P (a, b) P (a, c)| 2 P (b0 , b) P (b0 , c);

(3.27)

this is an expression for CHSH inequalities.


In principle entire measuring devices, each consisting of a filter followed by a detector,

36

could be used for Ia and IIb , and the values 1 of A(a) and B(b) would denote detection or
nondetection of the particles. Inequalities (3.27) would then apply directly to experimental
counting rates.
Nevertheless, if photons are used, this manner could not make to a very check of
(3.27), for the photomultipliers have a little efficiency; so from now on we shall interpret
A(a) = 1 and B(b) = 1 as the coming out or less of photons from their respective
filters, that for example could be linear polarizers having orientations defined by a and b.
At this point it is worthwhile to introduce an exceptional value, , taken from
parameters a and b for representing polarizer remotion; clearly, we will necessarily have
A() = B() = +1.
Since P (a, b) is a correlation function of emission (of photons), it has to be made a
further assumption in order to derive an experimental prediction: that if a pair of photons
emerges from Ia and IIb , the probability of their joint detection is independent of a and b.
Then if the flux into Ia and IIb is a constant independent of a and b, the rate of coincidence
detection R(a, b) will be proportional to w [A(a)+ , B(b)+ ], where w [A(a) , B(b) ] is the
probability that A(a) = 1 and B(b) = 1.
Letting
R0 = R(, ),

R1 (a) = R(a, ),

R2 (b) = R(, b),

and making use of the evident formulas

P (a, b) = w [A(a)+ , B(b)+ ] w [A(a)+ , B(b) ] w [A(a) , B(b)+ ] + w [A(a) , B(b) ] ,


and

w [A(a)+ , B()+ ] = w [A(a)+ , B(b)+ ] + w [A(a)+ , B(b) ] ,


w [A()+ , B(+)+ ] = w [A(a)+ , B(b)+ ] + w [A(a) , B(b)+ ] ,
w [A()+ , B()+ ] = P (a, b) + 2w [A(a)+ , B(b) ] + 2w [A(a) , B(b)+ ] ,
we get
P (a, b) =

4R(a, b) 2R1 (a) 2R2 (b)

+ 1.
R0
R0
R0

37

(3.28)

Finally, we can now express (3.27) in terms of experimental quantities, namely coincidence rates with both polarizers in, and with one and then the other removed. Supposing
that R1 (a) and R2 (b) are constants equal to R1 and R2 experimentally determined, CHSH
inequalities are
|R(a, b) R(a, c)| + R(b0 , b) + R(b0 , c) R1 R2 0.

(3.29)

This is the generalized formulation of Bells inequalities by CHSH, in a favourable


manner to be applied to experiments with photons. In the next section we shall see how
to rewrite these in a more explicit manner, thus applying them on the optical version of
Bells Gedankenexperiment.

3.3

Optical version of Bells Gedankenexperiment

As far as we have seen until now, Bells Theorem states that local realistic theories are in
disagreement with Quantum Mechanics, but there is nothing which tells us that these theories are to be rejected. On the contrary, the disagreement between the theories illustrated
by Bell arises from rather unusual situations.
It is just for this reason that an empirical testing by means of appropriately designed
experiments is required to test those critical regions in which conflicts have origin.
The more significative experiments, as well as complete, are those by Alain Aspect,
who was interested in the individuation of critical regions and in the analysis of photon
behavior in situations built up ad hoc. Section 3.4 will deal with the experimental aspect
and the results obtained by Aspect.
What we wish to study now is a reformulation of Bells Theorem [9], adopted by Aspect himself as starting point for his whole experimental examinations: a rewriting of
the inequalities in terms of correlation coefficients between the directions of the photons,
making more clear the importance of the hypothesis of locality within the theorem.
In the optical version of Gedankenexperiment of Einstein, Podolsky and Rosen, due to
Bohm (see Fig. 3.1), a source S emits pairs of photons with different frequencies, 1 and
2 , which propagate along opposite directions O~z, in a way similar to that of a pair of
1
2 -spin

particles in a singlet spin state.

Lets suppose that the state entangled which describes polarization of the two photons
is given by the ket
38

Figure 3.1: The Einstein, Podolsky, Rosen and Bohm Gedankenexperiment.

1
|(1 , 2 )i = (|x, xi + |y, yi) ,
2

(3.30)

where |xi and |yi represent states describing two different orthogonal directions of linear
polarization. This state is not a product state, therefore it is not possible to fix any single
photons with arbitrary polarization, instead we have to see the system in its whole entirety.
Two linear polarizers I and II, placed at the sizes of the source and oriented in the
have the aim to analyse photons and see their
and b,
directions given by unit vectors a
tracks. So it is possible to make measurements on linear polarization of two photons
through the analysis of two detectors (represented by + and in the Figure) placed next
to each of the polarizers. Both will give result +1 or 1 depending on whether the photon
polarization occurs in a direction parallel or perpendicular to the one of the polarizer itself.
that one
Labeled P (
a) the probability of obtaining the result 1 for 1 , and P (b)
of obtaining 1 for 2 , for these measurements quantum mechanical predictions for single
detection are

1
P+ (
a) = P (
a) = ,
2
= P (b)
= 1,
P+ (b)
2

(3.31)

according to the fact that no polarization cannot to be established for single photons;
therefore any measurement will give a random result.
to be the probability of combined detection of 1 in channel
Defining now P (
a, b)
quantum mechanical previsions
), and 2 in channel of II (b),
of I (oriented along a
for combined detection are
3
4

For an explicit evaluation, see Appendix A.


See note 3.

39

= P (
= 1 cos2 ab ,
P++ (
a, b)
a, b)
2
= P+ (
= 1 sin2 ab ,
P+ (
a, b)
a, b)
2

(3.32)

and b.
where ab is the angle between a
In particular, we can see that, in the case for ab = 0, we get

= P (
= 1,
P++ (
a, b)
a, b)
2

P+ (
a, b) = P+ (
a, b) = 0,

(3.33)

that is, whether the photon 1 gives result +1 (whose probability is 50%), then photon 2
as well will give certainly result +1 (analogously for -1), that it means total correlation.
This prevision is perfectly in accordance with all we have discussed so far, since measuring the polarization along the same direction of two photons emicted by the same
source at the same time, we obtain two results equal and opposite, as it happens with the
spin of two atoms belonging to the same molecule.
In this contest it is very useful to define the following quantity.
Definition (Correlation coefficient of polarization). It is called correlation coefficient of polarization the quantity
= P++ (
+ P (
P+ (
P+ (

E(
a, b)
a, b)
a, b)
a, b)
a, b).

(3.34)

Hence, substituting (3.32), in the quantum mechanical case this coefficient becomes:
= cos (2ab ) .
EM Q (
a, b)

(3.35)

For ab = 0 we have instead


EM Q (0) = 1,
i.e. total correlation.
It is evident that the value obtained effectively explains strong correlations that relate
measurement results of 1 and 2 . Generally, the correlation coefficient provides a quantitative criterion in order to quantify the correlation between random results obtained from
any individual measurement.
40

So we have to admit that there are some properties typical for any pair of photons
Einstein called them elements of physical reality that explain correlations, according
to the type of correlated results one obtains. These properties, different from pair to pair,
are not taken in consideration by the vector state |(1 , 2 )i, that instead is the same for
any pair of photons.
Here it is evident again the reason that guided Einstein to conclude that Quantum
Mechanics was not complete. And here there is the need of giving account of these
properties introducing some additional parameters, or hidden variables.
We can finally explain such interations through a classical description, and we can
hope of finding again quantum previsions by taking the average of the expectation values
over the hidden variables (procedure that brought Bell to his inequalities)
Now we shall try to work out Bells inequalities using the model described above.
the results re) and B(, b)
Labelling the set of additional parameters and A(, a
, and analyser II oriented
spectively obtained from analyser I oriented respectively along a

)]
along b, these quantities can only assume values 1, hence the quantity 1 [1 + A(, a
2

could assume only values +1 (in case of result +) and 0 (otherwise), and analogously, the
)] could assume only values +1 (in case of result ) and 0 (otherquantity 21 [1 A(, a
wise). Hence, given the probability distribution of , that is (), the expectation values
for single detection are found to be:

Z
1
)] ,
P (
a) =
d() [1 A(, a
2
Z
h
i
= 1 d() 1 B(, b)
,
P (b)
2
whereas for combined detection:

Z
1
4
Z
=1
P (
a, b)
4
Z
1

P+ (
a, b) =
4
Z
1

P+ (
a, b) =
4
=
P++ (
a, b)

h
i
,
)] 1 + B(, b)
d() [1 + A(, a
h
i
,
)] 1 B(, b)
d() [1 A(, a
h
i

d() [1 + A(, a)] 1 B(, b) ,


h
i
.
)] 1 + B(, b)
d() [1 A(, a

Substituting these quantities in definition (3.34), after some straightforward passages,


one finds that the correlation coefficient, averaged over the distribution of , is given by:
41

=
E(
a, b)

)B(, b).
d()A(, a

(3.36)

Now we define a new quantity, that allows us to write the inequalities representing 
in
0
0

, a
, b, b :
explicit form the correlation coefficient, quantity which can be denoted by s , a



b
0
, a
0 , b,
s , a

def

A(, a
0 ) + A(, a
+ A(, a
0)
)B(, b)
)B(, b
0 )B(, b)
0 )B(, b
A(, a
h
i
h
i
B(, b
0 ) + A(, a
+ B(, b
0 ) . (3.37)
) B(, b)
0 ) B(, b)
A(, a

Since A and B can assume only the values 1, then




b
0 = 2,
, a
0 , b,
s , a
from which, averaging over the distribution of one gets
Z
2



b
0 +2,
, a
0 , b,
d() s , a

(3.38)

that is, defining


E(
0 ) + E(
+ E(
0 ),
S := E(
a, b)
a, b
a0 , b)
a0 , b

(3.39)

we obtain the inequalities


b
0 ) +2.
0 , b,
2 S(
a, a

(3.40)

(3.40) are well-known as BCHSH inequalities, i.e. inequalities of Bell generalized by


Clauser, Horne, Shimony, Holt. These inequalities are based on a combination of four
correlation coefficients of polarization, measured along four orientations of the polarizers.
Then S is a measurable quantity.
However, BCHSH inequalities in some particular situations (which will be specified
shortly) are in conflict with Quantum Mechanics. Indeed if we put the system in the
configuration illustrated by Figure 3.2, with

,
8
+ a0 b0 ,

ab = ba0 = a0 b0 =
ab0 = ab + ba0

42

substituting the quantities E in (3.39) with their quantum mechanical values (3.35), we
get

SM Q = 2 2.
This quantum mechanical prevision deeply violates the upper limit of inequalities
(3.40).

Figure 3.2: Orientations with ab = ba0 = a0 b0 = 8 .


In this particular situation it turns out that quantum previsions cannot be obtained
from theories with hidden variables, therefore we can conclude that a local deterministic
theory does not exist, according to the very general model showed in this Section,
which reproduces all quantum mechanical previsions.
It is reasonable to ask what are precisely the critical regions and for which angles one
could have the maximum conflict. Deriving S with respect to the three independent angles
ab , ba0 , a0 b0 , we have that SM Q is extreme in correspondence with
ab = ba0 = a0 b0 = ,

(3.41)

SM Q () = 3 cos(2) cos(6).

(3.42)

and for (3.35),

Finally, deriving this quantity with respect to and letting equal to zero,
dSM Q
= 6 sin(6) 6 sin(2) = 0,
d
we obtain the angles for which SM Q has its maximum and minimum values, that are
respectively
43


max
SM
= 2 2
Q

min
SM
Q = 2 2

,
8
3
per =
,
8

per =

(situations illustrated in Figure 3.2 and 3.3 respectively).

Figure 3.3: Orientations with ab = ba0 = a0 b0 =

3
8 .

At last, with a brief study of function, it is possible to plot a graphic representing the
behaviour of S varying with the angle , as shown in Figure 3.4.

In conclusion, Bells Theorem brings out a conflict between theories with hidden variables and certain quantum mechanical previsions (according to the model expounded in
this section) and provides a quantitative criterion to clarify this conflict.
Discussing the hypotheses, the fundamental assumptions on the bases of the described
model are three: the existence of hidden variables, determinism and the locality condition. These hypotheses, if simultaneously assumed, give rise to an incoherence between
the created theory and Quantum Mechanics. The first two of them are not to be under
discussion since they are parts of the theory itself. The third instead is a handy and
obvious assumption that it seems absurd not to accept. And yet, whether in the case
that the first two hypotheses do not want to be dropped out, if one wish to complete
44

Figure 3.4: S() = 3 cos(2) cos(6) as expected by Quantum Mechanics for pairs in
the state (3.30). The conflict arises in the hatched zone.
Quantum Mechanics by a theory with hidden variables, then this theory will have to be
necessarily nonlocal. Indeed if the condition of locality is abandoned, it can be easily
or (, a
the demonstration that brings to
, b)
, b),
shown that with quantities like A(, a
BCHSH inequalities drops, thus voiding their result.

3.4

A. Aspects experiments

Between 1980 and 1982 Alain Aspect and his equipe tried to create more complex experimental apparatuses, compared with those used until that time, to verify Bells inequalities,
and they made three experiments to study the validity of these constraints using atomic
cascades [9].
The first experiment was been organized as a direct check for inequalities. Nevertheless,
it had some limits, e.g. single-channel polarizers were used , so that only photons polarized
could pass, while those polarized orthogonally were blocked. Therefore
(or b)
parallel to a
only the results + could be revealed, and measurements of coincidence could only provide

an evaluation for P++ (


a, b).
So one could not know whether the missed measure of
correlation between a pair of photons was due to a real absence of correlation (i.e. to the
action of the polarizer), or to a scarce efficiency of the detector.
For this reason a second experiment was performed, in this case preserving a configuration like that of the Bohm-Aharonovs thought experiment. This experiment implemented
45

two-channel polarizers that made possible to detect the correlation of every photons that

reached the polarizer, thus even allowing for obtain a valuation for P (
a, b).
Finally, the third experiment, which made use of polarizers with orientation variable
in time, can be considered as the more accurate and complete experiment, to which the
others can be reconducted. Moreover this experiment proves experimentally the impossibility of faster-than-light communications.
As common source for each of the three equipments, it has been used an atomic cascade
(J = 0) (J = 1) (J = 0). The main feature of these experiments with respect to the
previous was the usage of a very powerful and stable source, that allowed to cut the data
recording times from several hours to a few minutes. The source exploited the cascade
4p2 1 S 0 4s4p 1 P 1 4s2 1 S 0 of

40

Ca.

Figure 3.5: Relevant energy levels for a 40 Ca cascade. The atoms, excited by absorption
of two photons K and D , emit visible photons 1 and 2 correlated in polarization.
This cascade, represented in Fig. 3.5, produces two visible photons 1 and 2 correlated in polarization. The atoms of

40

Ca are thus excited from the ground state to the

upper energetic level by absorption of two photons, K and D , generated by two laser
beams. The first one (K = 406.7 nm) is generated by ions of Kryptons, as the second
was a dye laser, brought to resonance for the two-photon process (D = 581 nm); lasers
have parallel polarizations. First of all, there is a feedback cycle that checks on the dye
laser wavelenght, in order to have the maximum signal of fluorescence by the cascade.
Then a second feedback cycle checks on the power emitted by the Krypton laser in order
to guarantee a constant emission from the cascade. Strictly speaking, as a result of the
atom excitation, one electron of each atoms is led to jump up to two energy levels beyond
its ground state. When the electron falls from two energy levels, it sometimes emits a
pair of entangled photons. In this way, using a power of 40 mW for each laser, the typical
46

emission efficiency of the cascade is of 4 107 photons per second.

3.4.1

Experiment with single-channel polarizers

In their first experiment, Aspect and collegues have mounted a pile of two polarizers made
up of ten glass plates to the Brewsters angle; in front of each of these polarizer, a linear
polarizer has been placed, in order to transmit photons polarized in a way parallel with
the axes of the polarizer, and to stop those polarized perpendicularly.
In this type of experiment, one talks about single-channel polarizers because only the
value +1 is measured for each photon belonging to the couples of photons emitted by the
source. Therefore, the use of single-channel polarizers allows one to determine only the

5 , since it cannot be established if the result 1 for a photon


result R(
a+, b+)
= R(
a, b)
was due just to an orthogonal polarization with respect to the polarizer axes, or to a scarce
efficiency of the counting system. Accordingly, auxiliary recordings with one or both the
polarizers removed were needed, thus obtaining the following quantities:
R(, ) = R0

R(
a+, ) = R1 (
a)

R(, b+)
= R2 (b),

from which new BCHSH inequalities are obtained:


1 S 0 0,

(3.43)

i
1 h
+ R(
+ R(
0 ) R(
0 ) R1 (
.
R(
a, b)
a0 , b)
a0 , b
a, b
a0 ) R2 (b)
R0

(3.44)

with

S0 =

Experimental test of Bells inequalities has provided


0
Sexp
= 0.126 0.014,

(3.45)

that violates inequalities (3.43) by 9 standard deviations, and is on good agreement with
quantum mechanical previsions
SM Q = 0.118 0.005,

(3.46)

With analogy to the notation used in Section 3.2, by R(


a+, b+)
we intend the counting result of the

and b.
detection of coincidences parallel to the single-channel polarizers (+) oriented with respect to a

47

(in this case the error accounts for the uncertainty in the measurements of the polarizer
efficiences).

3.4.2

Experiment with two-channel polarizers

When a photon is stopped by the polarizer in the single-channel experiment, it is lost and
there is no way to establish whether and how it was correlated to another photon. This is
the reason why, in the second experiment made by Aspect, two channels have been used.
Indeed, if a photon is stopped by the polarizer, then the photon is reflected by it and so
could be still detected. In this way the coincidence rate was much greater and leaded to
a more precise experiment.
With two-channel polarizers, the experiment implemented is more similar to that of
the scheme in Figure 3.1. The polarizers used were polarizing cubes that transmit a
respectively) and reflect the orthogonal one. This
, or to b
polarization (parallel to a
polarization separator and its corresponding photomultipliers were mounted on a rotating
device. The latter (i.e. a polarimeter) supplies the results + and for linear polarization
respectively). Strictly speaking, it is an optical analogous of a
(b
measurements along a
Stern-Gerlach filter for 1/2 spin particles.
and with a four-coincidence
and b,
Setting the polarimeters I and II oriented along a
counting system, it has been possible to measure, in a single data recording period, the
The correlation coefficient for measurement along a
and
four coincidence rates R (
a, b).
is directly
b
= R++ + R R+ R+ .
E(
a, b)
R++ + R + R+ + R+

(3.47)

So it has been sufficient to repeat the same measurements for the other three orientations: in that way, BCHSH inequalities (3.40) could be directly verified.
This procedure is correct if the measured values (3.47) are like the correlation coefficients of polarization previously defined by (3.34), i.e. if one assumes that a set of pair
actually revealed is a significant sample of all of the emitted pairs. But, because of the
simmetry of the experimental apparatus, this assumption is revealed very reasonable, since
the measurements +1 and 1 are treated in equal manner, i.e. the efficiencies of revelation
for both channels of a polarimeter are equal.

Thus, after verification of the fact that, nevertheless each of the four rates R (
a, b)
changed drastically, their total sum was the same varying the orientations, it has been
made the counting experiment in agreement with the orientations of Fig. 3.2 and 3.3, i.e.

48

the values for which the maximum conflict is expected.


The value obtained is
Sexp = 2.697 0.015,

(3.48)

that violates inequalities (3.40) by more than 40 standard deviations, a result in a very
excellent agreement with Quantum Mechanics predictions (with the adopted polarizers
and solid angles):
SM Q = 2.70 0.05,
whose uncertainty accounts for a small lack of symmetry for both channels of each polarizers (1%). This asymmetric effect has been calculated and cannot however cause a
variation more than 2% for SM Q .

3.4.3

Experiment with time-varying polarizers

Maybe the most interesting experiment conducted by Aspect is the one in which timevarying polarizers have been mounted. This because the assumption of the principle of
locality is reasonable, but it is not prescribed by any physical law. So, accordingly to
the results obtained so far, we could imagine that maybe some fixed analysers can be
placed along their respective directions sufficiently in advance, in order to allow them to
communicate one each other by means of an exchange of signals faster than or equal to
c 6 . If such interactions existed, Bells inequalities would no longer be valid, since they
have as their hypothesis the independence of the measure made on a system with respect
to any disposition of the other apparatus.

In the experimental apparatus (see Fig. 3.6), each polarizer is substituted by a system
composed of a commutation device (C1 and C2 ) followed by two polarizers in two different
and b
0 by side II. The optical switch is able to lead
and a
0 by side I, b
orientations: a
rapidly the incident flux from a polarizer to the other one. Every system is thus equal
to a variable polarizer switched between two orientations. The distance between the two
commutators is L = 12 m.
The characteristic point of this Aspects experiment that will be considered the
very new feature that yields a definitive evidence of nonlocality is the light switching,
6

From now on, by c we mean the speed of light in the vacuum: c = 2.9979 108 ms1 .

49

Figure 3.6: Time-varying experiment with optical commutators C1 and C2 . Time of


commutation: 10 ns.
which is made through acousto-optic interaction of light rays by an ultrasonic stationary
wave in the water.
Thus, when the wave changes in the transparent water container, the light beam,
hitting the liquid, is deflected modifying the regulation of the polarizer. The incident angle
(Bragg angle) and the acoustic power are regulated to obtain a complete commutation.
At an acoustic frequency of 25 MHz, the switching frequency is 50 MHz. Accordingly,
every 10 ns there is a change in orientation of the equivalent variable polarizer. Since this
time interval (10 ns) and the mean life of the intermediate level of the atomic cascade (5
ns) are small with respect to

L
c

(40 ns), the detection of an event on one hand, and the

corresponding change of orientation on the other, are separated by a space-like interval 7 .


Nevertheless, by means of the large beams used, the commutation was not complete
since the incidence angle was not exactly the Bragg angle. So, in order to obtain a better
commutation, the divergence of the beams was reduced: such a thing led to a reduction
of the coincidence rates detected with respect to those ones of previous experiments.
According to this, the only consequence was a longer data-recording time.
With a data-recording of 8000 s, the polarizers configured in the extreme configurations
of Fig. 3.2 and 3.3, plus additional 8000 s, necessary for ausiliary calibration measurements
with the removal of half of all the polarizers, and taking into account for systematic errors,
the test for inequalities (3.43) led to
0
Sexp
= 0.101 0.020,
7

(3.49)

Two events are said to have a space-like separation if they do not influence between each other by
signals that propagate at a speed less than or equal to the speed light. For a definition of space-like
interval, see note 10.

50

that violates inequality (3.43) by 5 standard deviations, and is in good agreement with
quantum mechanical previsions:
SM Q = 0.113 0.005.

(3.50)

Therefore, even this experiment shows the reasonableness of Quantum Mechanics to


disadvantage of a theory of hidden variables which obey to Einsteins locality.
At this point, accordingly to the strong evidence of correlations, between the acceptation of local additional parameters that transmit themselves information faster-thanlight, or the rejection of an explanation that uses these additional parameters, we have to
choose the second position. An exam of experimental results let us conclude that Quantum Mechanics describes correctly the reality, whereas local hidden variable theories give
previsions proved wrong by the experiments.

3.4.4

Towards the impossibility of FTL communications

Actually, the results obtained by Aspects experiment with time-varying polarizers leads
to a further outcome, as we anticipated: that is, the impossibility of faster-than-light
communication, or FTL.
To better understand how this is not possible, suppose that the observer I wants to
transmit some information to the observer II using spin correlation measurements for two
particles spatially separated and in opposite directions. Suppose that I and II decide to
measure the Sz component. Then, without asking for anything, the observer II will know
exactly the result of the measure for the first observer. But this does not mean that the
two observers are communicating one to each other as observed in Section 2.4, since the
observer II obtains from his measure just a sequence of positive or negative signs without
having any useful information.
In order to get any kind of information, it would be necessary that the observers had a
decoding key, by means of that one is able to establish what meaning (i.e. information) to
give to the sequence of symbols + and , according to the order and the kind of spin that
are obtained from measurements, in order to let the observers obtain useful information
from a certain sequence of Sz relevations. But this necessarily involves an exchange of
a decoding key between the observers, that can occur only by classical channels, at a
velocity less than or equal to c. The reason of this is that entanglement, being a markedly
quantum (i.e. probabilistic) phenomenon, does not allow to the observer (that is, to the
sender of the possible communication) to choose how the spin revelations should be made

51

by who measures, making impossible any kind of key exchange.


In other words: only after matching their results (using a conventional method of
communication, that cannot broadcast signals faster or equal-than-light), the observers I
and II could notice the coincidence of their results.
Thus we have seen at a quick first analysis how it is not possible to utilize entangled
states for a direct exchange of information through FTL communications.

Quantum Cryptography
Instead, it is possible to exploit entangled states (for instance, entangled photons) for
an absolutely secure exchange of a cryptographic key clearly at velocity less than c
, by means of which to communicate successively by classical channels. The safety is
guaranteed by the fact that if a third observer, say III, measured a certain value of the
polarization component of one or more photons making part of the series that I and II
are mutually exchanging i.e. of the cryptographic key , then, after that measurement, the polarization state of the photon would alter and, because of the Uncertainty
Principle, some errors would be introduced into the measurement of the observer II even
for measurements that would be correct; when, at the end of the key exchange (that is,
the series of single polarized photons), after some procedures, observers I and II match
their keys, they will find inevitably that these are different for some values, which proves
that an external observer (III) has cut off some photons and hence that the key is not
secure.
This is what one does in Quantum Cryptography, where one of the fundamental problem is that of generating a sequence of random numbers, as long as the message to send,
in order to use the One Time Pad communication

well-known as Vernam cipher

, the unique cryptographic system whose security is mathematically demonstrated [11]


9

(for this reason it is also called perfect cipher). Once obtained a key as described,

100% safe, it is possible then to establish a secure communication between sender and
receiver. This is an example of how the entanglement is utilized practically, in this case in
a communication contest. On June of 2004 in Cambridge this has been implemented for
Quantum Net, the first net with more than two quantum cryptographic nodes. In Qnet
data flow by usual cables in optic fibre, but the peculiarity is that they are codified using coding keys obtained by the exchange of an entangled series of single polarized photons.
8

Cryptographic method in which the key lenght, composed of random characters, is equal to the lenght
of the message to code.
9
Firstly showed in 1949 by Claude Shannon.

52

3.5

No-Communication Theorem

In the light of what we said, one question could arise:


How does system I communicate to system II the outcome of the measurement of a
certain observable A, in time to produce the correlations we know of, and without violating
the special relativity limits?
First note that the question is ill posed, because it understates that the outcome of
measuring A causes the outcome of B. Relativistically speaking, in spacetime regions
where the two measurements are taken (or can be taken) the latter are, in relativistic
language, causally disjoint: there is no future-directed space-like or time-like path
10

in the cone of spacetime joining them. This means that the cronological order of

the events is conventional, and depends on the choise of inertial frame. Indeed, special
relativity states that it is possible to choose an inertial frame in which As measurement
precedes in time Bs, or another frame in which the situation is opposed: that is, Bs
measurement precedes in time As.
So it makes no sense to say that the outcome of the experiment on B is the consequence,
or the cause, of the outcome on A

11 .

Nevertheless, one could resort to the partial conventionality of Einsteins synchronisation procedure in order to dismiss that problem. But despite the conventional choises
underpinning special relativity, it is known that the correlations between causally disjoint
events are dangerous in relativistic theories, for they can produce causal paradoxes: with
a chain of causally disjoint events we can put events in the history of a given system in
any chronological order whatsoever. If it were possible to use the correlations of causally
disjoint events to transfer information either way, we would be able to communicate with
the past (inside the light cone) and thereby obtain causal paradoxes.
It can be proved, by accepting the standard formulation of Quantum Mechanics for
systems made by entangled states, that no piece of information can be transmitted from
event X, where part of the system is measured, to event Y , where the other part is
measured, by measuring any arbitrary pairs of quantities and exploiting the quantum
correlations between the readings. Not only that, but observing the outcomes on one part
of the system we cannot establish whether on the other part some measurements have been
10
Given two distinct simultaneous events, their interval is space-like: t = 0 in fact implies s2 < 0,
being s2 = c2 t2 x2 . Instead if two events occur in the same space point, their interval is time-like:
x = 0, thereby s2 > 0.
11
We already noted that in Section 2.4, where we have talked about the concept of simultaneity used by
EPR.

53

taken, if they are being taken as we speak, nor if they will be taken in the future.
We shall consider in the following two ways of transferring information from X to Y
via EPR correlations.
Given a quantum system S composed of two subsystem A and B, the Hilbert space
of S will be given by HS = HA HB with obvious notation. Lets consider also a pair of
observables GA and GB , with discrete spectrum, defined in A and B respectively. Such a
pair of observables with discrete spectrum can be relative to internal degrees of freedom
of their respective systems in which they are defined.
First possibility
Consider the single pairs of measurements on A and B of the observables GA and GB
respectively, which we know have correlated outcome. We cannot pass information from
X to Y using the correlation, because the outcome, albeit correlated, is completely accidental. Making a practical example, it is like having two coins, A, B with the remarkable
property that each time one shows heads, the other one gives tails, independent of the
fact they are tossed far away, rapidly, and that A is tossed before or after B in some frame.
The coins, though, have a quantum character and it is physically impossible to force one
to give a certain result: the outcome of the toss is determined in a probabilistic way and
whatever our wish is. Thus the two coins, i.e. our quantum system made by parts A and
B, cannot be used as a sort of Morse telegraph to transmit information between X and Y .

Second possibility
Consider not the single measurements of GA and GB , but a large number thereof, and
study the statistical features of the outcome distributions. The statistics of the measurements of GA might be different according to whether we measure GB as well, or whether
we measure a new quantity G0B . In this way, by measuring or not measuring GB (and
measuring G0B or measuring nothing at all) in Y , we can send an elementary signal to X,
of the type yes or no, that we recover by checking esperimentally the statistics of A.
It is possible to demonstrate that, even with such procedure, it is not possible to
transmit any information, since the statistics relative to GA is exactly the same in case
we also measure GB (or any other G0B ) or we do not measure GB .
But before this study, we recall some properties of the density matrix .

54

A few notions about the density matrix


We recall that, given a dynamical quantum system of physical quantities, it is possible
to associate with any dynamical variable of the system a definite statistical distribution
of the possible values assumed by it. The dynamical state of a quantum system can be
completely known if it is possible to determine precisely the variables of one of the possible
complete sets of compatible variables associated with the system.
When the knowledge of the system is not complete, one says that the system has a
certain probability p1 , p2 , . . . , pm , . . . , of being in dynamical states represented by kets
|1i , |2i , . . . , |mi , . . . , respectively . In other words, the dynamical state of a quantum
system cannot be more represented by a unique vector, instead by a statistical mixture of
vectors.
A statistical mixture can be described by means of the density operator or statistical
operator
=

|mi pm hm| ,

where the vectors |mi are normalized to unity (but not necessarily orthogonal) and the
quantities pm have the characteristic properties of statistical weights, namely
pm 0,

pm = 1.

The average value of the observable A is equal to


hAi = T r A.
Indeed it can be easily shown that
T r A =

pm T r (|mi hm| A).

So, having knowledge of , it is possible to derive a statistical distribution of the results


for measurements of A.
Very generally, if PD is the projector upon the subspace spanned by the eigenvectors
of A belonging to the eigenvalues located in a certain domain D of the spectrum of A,
whether an observation made on the system shows that it is an eigenstate of A belonging
to D, the density operator after the measurement is, without a normalization constant,
the projection PD PD of the operator , which represents the statistical mixture before

55

the measurement. The constant is determined by the constraint that this operator must
have trace equal to unity. We therefore obtain that
T r PD PD = T r PD ;
then the evolution in time of density operator during the act of measuring leads to

PD PD
.
T r PD

For a more complete study of the properties of the density operator, we refer the reader
to [21], [26].

We now can begin by discussing the second possibility.


We take the state I(HA HB ) for the system composed of A and B. Suppose
that GA = G(A) IB , where G(A) is a self-adjoint operator in HA having a discrete and
(A)

(A)

(A)

finite spectrum: {g1 , g2 , . . . , gn }, associated with its respective eigenspaces Hg(A)


k

(GA )

HA HB , targets of orthogonal projectors Pk

:= PkG

(A)

IB .

Analogously GB = IA G(B) , where G(B) is a self-adjoint operator in HB with discrete


(B)

(B)

(B)

and finite spectrum: {g1 , g2 , . . . , gm }, associated with its correspondent eigenspaces


(GB )

Hg(B) HA HB , targets of orthogonal projections Pk


k

:= IA PkG

(B)

(B)

Now, if we measure GB on state reading gk , the post-measurement state is:


1
Tr

(G )
(G )
Pk B Pk B

 Pk(GB ) Pk(GB ) .

Considering all of the possible readings of B, if we measure in some frame first B


and successively A, the system we are intending to test on A is the quantum mixture:
0 =

m
X
k=1

Tr

pk
 Pk(GB ) Pk(GB ) ,
(GB )
(GB )
Pk
Pk



(G )
(G )
(B)
where pk = T r Pk B Pk B is the probability of reading gk for the measure of B.
Thus:
0 =

m
X

(GB )

Pk

k=1

56

(GB )

Pk

(A)

Hence the probability of getting gh

for A, when B has been measured (irrespective

of the latters outcome) is:


(A)
P(gh |B)

= Tr

(G )
0 Ph A

m
X

= Tr

!
(G )
(G ) (G )
Pk B Pk B Ph A

k=1

At this point, with the help of the property of linearity and ciclicity of trace, we have:

(A)
P(gh |B)

m
X



(G )
(G ) (G )
T r Pk B Pk B Ph A

k=1

=
=

m
X
k=1
m
X



(G ) (G ) (G )
T r Pk B Ph A Pk B


(G ) (G ) (G )
T r Pk B Pk B Ph A ,

k=1
(GB )

where in the last step we have used Pk


(GB )

projectors. On the other hand Pk


theorem

(GA )

Ph

(GB )

(GA )

= Ph
(GB )

Pk

= Pk

(G )

Pk B from the structure of the


P (GB )
= I by the spectral
e
k Pk

12 .

Therefore:

(A)
P(gh |B)

m
X



(G ) (G )
T r Pk B Ph A

k=1

= Tr


m
X

!
(G ) (G )
Pk B Ph A

k=1

(G )
Ph A

= Tr


(A)
= P gh
.

(A)

Thus we have obtained that: the probability of obtaining gh

from A when the quan-

tity B has been measured ( with any possible outcome), coincides with the probability of
(A)

obtaining gh

from A without measuring B.

12
For finite-dimensional
Hilbert spaces, the spectral theorem states that each Hermitean operator O can
P
be written O = i oi P (|oi i), where {oi } is the set of eigenvalues comprising the eigenvalue spectrum of O
and P (|oi i) is the projector onto the finite Hilbert subspace spanned by |oi i. This provides the spectral
decomposition (also called eigenvalue expansion) of the operator O. When the state of the system is instead
mixed, by necessity being described by a statistical operator that is not a projector, the expectation value
is hOi = T r (O).

57

So, measuring part B of the system, the presence or the absence of the correlation is
completely irrelevant if we observe only part A of the system. From this follows that, in
conclusion, even if we consider the statistics of the results for the measures of A, there is
no way to transmit information using EPR correlations.
There are several generalizations of the result obtained above. For instance, we refer the reader to a paper of G.C. Ghirardi, A. Rimini and T. Weber [12], entitled A
General Argument Against Superluminal Transmission Through The Quantum Mechanical Measurement Process , where it is proved, by more general characters than ours, the
impossibility of transmission of faster-than-light signals.

3.5.1

Some outlines about the FLASH Project

In 1981 the American physicist N. Herbert proposed a Gedankenexperiment in order to


use quantum nonlocality to obtain a FTL communication

13

[13].

Herberts apparatus was an idealized laser gain tube which would have macroscopically
distinguishable outputs when the input was a single arbitrarily polarized photon. Indeed,
the word LASER is an acronym for Light Amplification by Stimulated Emission of Radiation. However, besides the stimulated emission, there is also spontaneous emission,
which results in noise. Herberts claim was that the noise would not prevent identifying
the polarization of the incoming photon, at least statistically. For this, he used the notion
of quantum compounds and many properties of laser physics.
This apparatus has been realized for the first time in 2007 by a group of Italian
research workers (T. De Angelis, F. De Martini, E. Nagali, F. Sciarrino) [15], using the
optical parametric amplification 14 of a single photon forming an EPR-entangled pair into
an output field involving 5 103 photons. The experimental apparatus is represented in
Figure 3.7.
Two space-like distant observers, say Alice and Bob, share two polarization entangled
photons generated by a common EPR source, represented in the figure by Crystal 1.
Alice detects by two phototubes a certain polarization of her photon in a two orthogonal
measurement basis, that with no lack of generalization are chosen as linear polarization
and circular polarization with respect to a classical Cartesian (tridimensional) frame.
13
Actually, Nick Herberts FLASH paper was wrong, but anyway it was published. For more details
about this story and its contribution to the no-cloning theorem, see [14].
14
An optical parametric amplifier (OPA) is a particular kind of amplifier that allows to directly amplify
a weak optical signal through a highly non-linear medium, without converting it into an electric signal.

58

Figure 3.7: Configuration of the quantum injected optical parametric amplifier. The
polarization entangled photon is generated by Crystal 1.
The choise of the basis is the only coding method accessible to Alice in order to establish
a meaningful communication with Bob.
Well, if Bob could guess the coding basis chosen by Alice, then a FTL signaling process
would be established; nevertheless, since the detection of an unknown single particle cannot
carry any information on the coding basis (as the No-Communication Theorem states),
the purpose by Herbert is been that of letting Bob make a new kind of measurement on
the photon through the amplification by a polarization independent amplifier, in order to
split the amplified beam by symmetric beam-splitter (BS), so that Bob could perform a
measurement on half of the amplified particles in the linearly polarized basis, and on the
other half in the circularly polarized basis.
Actually this proposal cannot be carried out, because of the quantum no-cloning theorem, that asserts that for the Quantum Mechanics postulates, no quantum amplifier can
duplicate accurately two or more nonorthogonal quantum states.
Herbert, aware of the impossibility to produce perfect clones of any input qubit
where a qubit is a quantum information unity, which in our specific case is represented
by the photon , because of the noisy contributions from spontaneous emission, proposed
to measure the mean values of two physical quantities, depending on the number of pair
of electronic signals detected by two pairs of photodetectors, that should have had a
dependance on the coding bases chosen by Alice.
Omitting some technical and numerical details of the experiment (for more details we
refer to the original paper of the experiment listed in Bibliography), the final result of the
experiment, based on an accurate theoretical and experimental analysis, is that the mean
59

values measured have not shown any kind of dependence on the coding bases chosen by
Alice. In other words, whatever measurement setup Bob intends to choose for making
measurements, he will not be able to identify, among all of the possible setups, the one
adopted by Alice.
This result is general and prevents from any type of superluminar communication based
on quantum nonlocality.
For completeness, we mention that actually, superluminar group velocities has been
observed in barrier tunnelling in condensed matter [16] [17]. However this is not a problem
because special relativity does not forbid the group velocity to exceed c. It is the front
velocity of a wave packet that is the relevant criterion for signal transmission, and the
front velocity never exceeds c.
So, the importance of the FLASH experiment is that it explained, in a general and
definitive manner, the exact reason of the failure of FLASH project, besides any similar
FTL proposal.

60

Chapter 4

Conclusions on Quantum
Mechanics
We are able now to draw some conclusions about the foundations of Quantum Mechanics.
In Chapter II we have seen how the inductive reasoning used by Einstein, Podolsky and
Rosen yields the conclusion that the principle of locality, the principle of reality and the
completeness of quantum mechanical description are three postulates incompatible with
each others. Indeed, assuming these three hypotheses, a paradox comes up, solvable only
rejecting at least one of these latter.
EPR discovered that Quantum Mechanics predicts strong correlations between measurements on two particles in an entangled state. They interpreted these correlations as
the result of shared properties determined at the time of their initial interaction and carried along by each particle. Theories completed in such a way, implement a view of the
physical world called local realism, because individual physical properties are attributed
to each of the separated partners. This type of interpretation was favoured by Einstein.
On the basis of this interpretation stands the idea that Quantum Mechanics is a sort
of phenomenological theory, that does not describe the world as it really is, but otherwise
gives its behaviour in average. A more detailed description of physical reality could so be
given by a more fundamental theory, in principle deterministic, that accounts for additional
dynamical parameters i.e. the hidden variables. In this way, besides surmounting the
paradox, it is possible to obtain a completely deterministic vision of the world.
By the terms theory with hidden variables we intend a theory characterized by:
- a set of parameters, called hidden variables, whose knowledge allow to expect, in a
deterministic manner, the evolution of the system;
- a statistical distribution of hidden variables;
61

- a rule that establishes the correspondence between the value of the hidden variable
and that of the measured observable;
- a rule that fixes the distribution of the hidden variables after the measurement.
But the intepretation of Einstein was strongly opposed by another great quantum
physicist, Niels Bohr, because of the violation of his complementarity principle.
John Bells formulation of his celebrated inequalities, as we have seen in Section 3.1,
fixed a limit to the correlations predicted by local realistic theories and made it possible to
settle the debate by performing an experiment to test the inequalities. For a well-designed
experiment, Quantum Mechanics predicts a violation of Bells inequalities. There were
thus two possibilities, both interesting: either the experimental results would obey Bells
inequalities, thus exhibiting a failure of Quantum Mechanics, or they would violate Bells
inequalities, forcing us to renounce the Einsteins local realistic world vision.
Starting with the pioneering work of John Clauser et al., that we discussed in Section
3.2, a series of more and more refined experiments has brought overwhelming evidence that
the actual degree of correlation experimentally obtained, indeed violates Bells inequalities.
The clear violation of Bells inequalities leads to the conclusive rejection of theories that
are simultaneously realistic and local: Quantum Mechanics cannot be completed by local
hidden variable theories.
Therefore, unless one denies the validity of all of the experiments conducted, independently by the fact that one agrees or not with the standard formulation of Quantum
Mechanics, we have to be in agreement with the existence of the correlation anticipated
by Quantum Mechanics and with the fixing of the expected values at the moment of the
measure.
The solution to the EPR paradox proposed by Einstein et al. is not clearly the only
solution: abandoning one of the two principles introduced in the EPR reasoning, we can
obtain a different vision of the world.
Rejection of the principle of reality

If we abandon the principle of reality we admit

that the wave function is not a feature of the physical system, independent from the
observer, but a mathematical expedient that synthesizes all of the possible information
that the observer can have about the considered system. In this sense, Quantum Mechanics
is a theory able to predict the temporal evolution of the knowledge of the observer about
the system, but without being able to demand it for a description of the system itself.
Hence, the measurements made on some systems alter the knowledge that one can obtain
about them. By this point of view, also the principle of locality is saved by the fact that

62

observations on systems modify only the knowledge of the observers about them, without
referring to their physical properties. This is the basic idea of the vision of Copenhagen
school.
Rejection of the principle of locality The other solution consists in abandoning the
principle of locality, so agreeing with the instantaneous variation at distance of the physical properties of a system and accepting an interpretation in which the wave function is a
property of the system itself. Therefore, renouncing to locality, the physical system would
be implicitly considered as an indivisible whole.
It is necessary to precise that the debate about a possible deterministic completion of
Quantum Mechanics is not concluded, since, as we have pointed out more than once in
this work, it is not excluded the possibility of nonlocal hidden variable theories equivalent
to Quantum Mechanics, for Bells inequalities stop to be valid with a hypothesis of nonlocality.

4.1

Interpretations of Quantum Mechanics: Copenhagen,


Bohm

In this manner we arrive to some possible interpretations of Quantum Mechanics, each


based on their own principles that tell us how the experimental results must be interpreted.

4.1.1

Copenhagen interpretation

Most of the scientific community tends to support the Copenhagen interpretation,


defined as well standard interpretation, that rests its roots in the renouncing to the
principle of reality, as seen above. It relies upon two assumptions:
1. A particle, for instance an electron, has only a potential existence until the moment
it is observed. Once observed, by means of a detector, its potentiality collapses
into a concrete and objective existence of the particle;
2. When not manifested as a particle, the potential state is represented by a wave
function which has to be intended just as a mathematical formula: there is no real
wave behind the numbers. Substantially, according to this intepretation, an objective
subatomic reality does not exist.

63

As we said in the last Section, this intepretation puts at the center of the question the
observer, which interacting with the quantum system, perturbs it and makes it collapse
necessarily into one of the possible admitted states. It is the process of observation that
makes possible for the quantum system to reveal itself to the observer or, it is better
to say, to the instrumental apparatus.
Lets consider for instance the Youngs double-slit experiment. The potential existence
of an electron is represented by a mathematical expression called wave function. As the
particle hits the double-slit screen, the wave function splits into two wave functions. In
absence of an observer, on a quantum level, nothing of objective is present yet. Beyond
the slits, the two wave functions interfere among each other producing wavefronts. As
well the interference process and the creation of wavefronts correspond to nothing of real:
they are just the result of mathematical elaborations.
At this point, suppose that an observer wants to determine the position of the electron. In order to make that, a detecting plate is placed at a certain distance from the
slits, capable of producing a flashing at the point hit by the electron. According to this
interpretation, when the wave front of the function comes to contact with the detecting
plate, it collapses. The potential existence of the electron thus becomes real and on the
detecting plate a short flashing allows the observer to detect the electron position.
The interpretation of such results, for Copenhagen school, is that it is possible just to
establish in a probabilistic manner the point of the plate where any particle will hit on,
without being able to exactly predict where a definite particle is going to hit; this is not
because of the unknowledge of some initial conditions, maybe included by some hidden
variables, but simply because the wave function, describing one or more particles, has a
meaning of probability amplitude for the system, and it would make no sense to request
deterministic information from it.

4.1.2

Bohms interpretation

Among all of the interpretations alternative to the standard one, Bohms interpretation
is one of the well-known, maybe the best complete nonlocal hidden variable theory, directed
to interpret and resolve the several paradoxes that came up from the assumptions needed
by standard interpretation.
This theory takes its name by David Bohm, who invented it in 1952 starting from the
idea of the pilot wave introduced by L. V. De Broglie in the Solvay Congress on 1927 1 .
1
This idea consists in considering a particle associated with a wave function which guides the particle
in motion whence the term pilot wave. Such pilot wave is mathematically described by the classical

64

Since it is a nonlocal hidden variable theory, it gives an objective and deterministic


description of the reality, in which, even when not observed or measured, subatomic world
is populated by real objects, like an electron and the fields surrounding it. Moreover, the
particle movements are not casual, but causal, i.e. determined by the field. So we see
that adopting this interpretation there is no reason to suppose the collapse of the wave
function, because of its objective reality that is, the particle and their respective fields
are always real quantities, not created just according to the action of the observer.
Hence, the so-called Bohmian mechanics is about point particles in motion. In a
Bohmian universe everything is made out of particles. Their motion is guided by the wave
function. That is why the wave function is there: that is its role. Moreover, Bohmian
mechanics happens to be deterministic. For instance, a substantial success of Bohmian
mechanics is the explanation of quantum randomness or Borns statistical law, on the
basis of Boltzmanns principles of statistical mechanics, i.e., Borns law is not an axiom
but a theorem in Bohmian mechanics.
The quantity |t |2 , with t a solution of Schrodingers equation, satisfies a continuity
equation, the so-called quantum flux equation. The particles in Bohmian mechanics move
along the flow lines of the quantum flux. In other words the quantum flux equation is the
continuity equation for transport of probability along the Bohmian trajectories.
Bohmian mechanics is defined by two equations: one is the Schrodinger equation for
the guiding field t , while the other is the equation for the positions of the particles. It
is nonlocal in the sense of Bells inequalities and therefore, according to the experimental
tests of Bells inequalities, concurs with the basic requirements that any correct theory of
nature must fulfill.
As well, the interpretation of Bohms theory is causal/ontological and is founded on
the following assumptions:
1. The wave function, , is not just a mathematical symbol, but rather a real field,
that can be called -field, existing objectively;
2. In addition to the -field, there exists a real particle, mathematically represented
by a set of coordinates, always well-defined, which change in a well-defined manner;
3. A particle tends to move according to the classical laws of motion with a velocity
inversely proportional to its mass and directly proportional to the classical potential
(V ) belonging to the -field;
quantum mechanical wave function, modified by a factor which accounts for the influence on the particle
motion.

65

4. The particle is not subjected just to the classical potential, but also to an additional
quantum potential (Q);
5. The -field is characterized by a very rapid and chaotic fluctuation. The quantum
mechanical values for this particle are an average over these random fluctuations on
a typical time interval;
6. The fluctuation of the -field is determined by a deeper sub-quantum level more
or less, as the fluctuation of pollen in Brownian motion is determined by a deeper
atomic level.
The absolutely new thing introduced by Bohm with respect to De Broglie is the idea
of quantum potential (Q), a force that acts on the particle with nonlocal influence. Introducing this nonlocal hidden variable, Bohms theory is able to expect the same results as
the orthodox quantum theory as well.
Now turning again to the double-slit experiment, in Bohms interpretation the electron
is an object characterized by position and velocity. It is always surrounded by another
object, a force field, the -field, comparable to an electromagnetic field. This field satisfies
Schrodinger equation, more or less as Maxwells equations are satisfied by the electromagnetic field. So this field is causally determined too. Moreover, it reflects even the hidden
variables of sub-quantum level, that determine in detail its oscillations and fluctuations
near a mean value, equivalent to that obtained resolving the Schrodingers equation.
Particle and wave are going together to hit the double-slit plate. Obviously, the particle
can pass through only one of the two slits; instead, the -field pass through both of the
slits. While passing through the slits, the -field is refracted in a way similar to any
other field for instance, electromagnetic field. The interference produces a quantum
potential highly complex beyond the slits that, generally, does not diminish in distance.
The particle is therefore deflected in its movement by the quantum potential Q. In other
words, the information contained in the quantum potential determines the trajectory of
the particle, i.e. the experiment outcome.
In a statistical sample of systems, having all the same initial wave function, the fluctuation of the -field yields an interference pattern equal to that predicted by Copenhagen
interpretation.
At the end of such considerations, it would seem that Bohms theory is able to confute
the Uncertainty Principle, which has its own general and mathematical validity. Actually,
it turns out that Bohms theory does not at all refuse the Uncertainty Principle.
First of all, note that the causality that Bohm talks about acts in the relation between
sub-quantum and quantum levels. That is, the fluctuation of the field on a quantum level
66

is causally determined by hidden variables of the sub-quantum level. In no way one has to
think about a reintroduction of the causality as a return to determinism on a quantum level.
At this level, that is inaccessible through quantum experiments, the Uncertainty Principle
remains strictly valid. Since the -field, as seen before, fluctuates casually with large
velocity, it is not possible to take its behaviour in a certain time. What one could make
is just to evaluate the mean variations of the field over a little region of spacetime, and
implicitly, to introduce a certain variation among the mean value evaluated (on the time
interval ). This deviation corresponds completely to Heisenberg Uncertainty Principle.
Therefore Bohmian mechanics is a deterministic theory, and a realist one, i.e. it
removes quite the role and the necessity of an observer for the quantum world. But it
is nonlocal: this means that two events that happen in a certain point of the space can
have immediate consequences in the farest point from it. So we have the so-called spooky
action at a distance.
On the other hand, Bohms theory has a very greater level of difficulty in calculation
with respect to classical Quantum Mechanics, but this not necessarily means that it is
wrong. Practically, it cannot be falsified. Nevertheless, there are still several concepts
that need to be clarified in this model.
One of the reasons that generally make such theory not supported by the scientific
community is that the quantum potential changes instantaneously, varying with experimental conditions, thus implying considerable quantities of effective calculus with the
variation of the experimental apparatus, even if, in the light of Bohms theory, such an argument should not be so unusual since, because of the indivisibility between the observer,
the measurement apparatus and the particle under study, there should be something
that is, the quantum potential that changes when the properties of an experiment are
modified.
A more valid reason of criticism is based on the difficulties, at an experimental level, of
studying and checking the validity of this theory, that are not yet at a level acceptable to
get any conclusion about its effective power of prediction, since in realizable experiments
Bohmian mechanics is in accordance with the quantum probabilistic phenomenology; so,
by a practical point of view, it is not so essential to make further tests for Bohms theory.
However this is an interesting theory, whose complexity and problems are plainly superior to the level of this work. For deeper details, we recommend the reader to specialized
texts (for instance [22], [23], [32]).
John Bell, speaking about the double-slit experiment according to the Bohm-De Broglie
interpretation, commented:

67

While the founding fathers [of Quantum Mechanics] agonized over the question particle

or wave, De Broglie in 1925 proposed the obvious answer particle and wave. Is it
not clear from the smallness of the scintillation on the screen that we have to do with
a particle? And is it not clear, from the diffraction and interference patterns, that the
motion of the particle is directed by a wave? De Broglie showed in detail how the motion
of a particle, passing through just one of two holes in screen, could be influenced by waves
propagating through both holes. And so influenced that the particle does not go where the
waves cancel out, but is attracted to where they cooperate. This idea seems to me so natural
and simple, to resolve the wave-particle dilemma in such a clear and ordinary way, that it
is a great mystery to me that it was so generally ignored [27].

4.2

Open questions on Quantum Mechanics

We have been seeing till now how different are some of the possible interpretations for
explaining quantum phenomenology. We have seen also how different are the problems
that involve the standard interpretation, though it is the most adopted.
One of the main problem for instance, as such it has been judged by J. Bell,
P. Dirac and A. Einstein is that of the ontological statute of the measure, that is,
the acceptance of the action of the measure generating the reality of what one observes.
This point pretends to put the role of the observer in the foreground, without whom
an observable would not be defined. With regard to this, it is notorious an anecdote
concerning the question that Einstein ironically proposed to his colleague A. Pais:

Do

you really believe the moon is not there when you are not looking at it? [30].
Another absolutely open question is the micro-macro aporia. The things of the
quantum world, both in Copenhagen and Bohms theory, have a different behaviour with
respect to the things of the macroscopic one. At the present, there is no way to set a limit
yet, neither physical nor logical, to split the microscopic from the macroscopic.
Among the many proposals in the field, one of these is the dynamical reduction model
proposed in 1985 by the Italian Gian Carlo Ghirardi, Alberto Rimini and Tullio Weber
authors of important papers about the Non-Communication theorem [12] and the NoCloning theorem, in addition, as well, to the well-known Ghirardi-Rimini-Weber theory,
which includes the above-mentioned topic, within the problem of interpretation of measure
of quantum systems [18][19]. This mathematical model starts from the idea of modifying
Schrodingers dynamics. In fact the evolution of the wave function is not linear, as expected
by the well-known equation proposed by the Austrian physicist, but is subjected to the
influence of statistical influences. In other words, the wave function is subjected, on

68

random intervals of time, to spontaneous processes which correspond to the localization


in space of the microconstituents of any physical system. The frequence of these processes
is very little on an atomic scale, so the localization of a single particle remains indefinite.
That is, nothing changes in the overlapping of states, where there is a single not observed
i.e. not perturbed electron. This frequence increases with the number of constituents
of a physical system. Macroscopic objects are therefore constituted by a large number of
quantum particles, so their organization is very high, and the overlapping of quantum
states of the macroscopic object system is virtually cancelled, even when the system is
not perturbed, i.e. observed.
Anyway the question still remains open, until all the controversial aspects will be explained and some experimental evidences will agree with any proposed theoretical models.
Finally, how not to pay any attention to a phenomenon like entanglement, totally
confirmed on an experimental level, but maybe not completely understood on a phenomenological level yet.
Abner Shimony has referred to entanglement as passion at a distance, in an effort
to avoid the trap of assuming that one can somehow use entanglement to send a message
faster than light. Shimony believes that entanglement still allows for Quantum Mechanics
and relativity theory to enjoy a peaceful coexistence, in the sense that entanglement
does not violate special relativity in a strict sense (no messages can travel faster than
light and we have seen how every experimental attempt just failed on this purpose).
Other physicists, however, that we could define as realists `a la Einstein, believe that
the spirit of relativity theory is still violated by entanglement, because something
(whatever it may be) does travel faster than light (in fact, infinitely fast) between two
entangled particles. The late John Bell was of this belief [28].
Possibly a way to understand entanglement is to avoid looking at relativity theory
altogether, and not to think of two entangled entities as particles sending a message from
one to the other. In a paper entitled Quantum Entanglement: from Poppers experiment
to quantum eraser [20], Yanhua Shih argues that because two entangled particles are (in
some sense) not separate entities, there is even no apparent violation of the Uncertainty
Principle, as EPR had suggested.
Entangled particles transcend space. The two or three entangled entities are really
parts of one system, and that system is unaffected by physical distance between its components. The system acts as a single entity.

69

In conclusion, it is necessary to say that beware of not to let down of importance the
question about which quantum mechanical interpretation to choose as correct in spite
of all these basic issues, in the daily practise the formalism of Quantum Mechanics works
incredibly well (for instance, lets just think about the exceptional accordance between
the value measured for the electron gyromagnetic g-ratio and the one predicted by QED
[33]).
To complete the topic about interpretations, we just have to consider the ironical Zero
Interpretation by David Mermin, summed up in the aphorism

Shut up and calculate!,

that interprets in the best way the thing that, however, Quantum Mechanics allows to
obtain better, regardless of the apparent paradoxes that can arise.
On the other hand, problems such as those discussed in this work, probably are not
yet sufficiently mature for a solution which, anyway, sooner or later, it will be necessary to
find if one day we wish to get a more profound knowledge of the foundations of Quantum
Mechanics.

70

Appendix A

Some expectation values of


Quantum Mechanics
associated with a
It is well known from Quantum Mechanics that given an operator A,
quantity A, and given a system C in the state
|i =

n
X

ck |k i ,

k=1

with n N, and some coefficients ci such that

|ci |2 = 1, the measure of the quantity A

on C, that is the quantum prevision of the result of such measurement, is given by


hAi = h| A |i =

n
X

|ck |2 hk | A |k i +

k=1

cm ck hm | A |k i ,

(A.1)

k6=m

where cm represents the complex conjugate of cm .


In particular, in the case of a system characterized by a 2-dimensional spectrum, with
wave function
|i = c1 |1 i + c2 |2 i ,
where |c1 |2 + |c2 |2 = 1, the mean value of A will be given by

hAi = h| A |i = |c1 |2 h1 | A |1 i + |c2 |2 h2 | A |2 i +


+ c1 c2 h1 | A |2 i + c2 c1 h2 | A |1 i .
This generic definition can be applied to any discrete operator.
71

(A.2)

1. Quantum expectation value for the simultaneous measurement of spins of

and b
two particles along the directions a
We propose to evaluate explicitly the value for the expression (3.3), that is
D
E
=

~2 b
~1 a
a b,
for a system in a state described by the wave function (2.21), rewritten for convenience
as:
1
= [|+i1 |i2 |i1 |+i2 ] .
2

(A.3)

Hence, for evaluating the mean value in (3.3), it is just necessary to apply the expression


,
= ~2 b
) and B
(A.2), substituting the operator A by the operators A = (~1 a

and of ~2 along the direction b,


associated with the measures of ~1 along the direction a
respectively.
will
For the definitions of the operators associated with the spin components, A and B
assume the form:

A = ax x + ay y =
= bx x + by y =
B

ax iay

ax + iay

bx iby

bx + iby

!
,
!
.

Then, taking care of the fact that operator A acts just on the first particle and operator
just on the second, applying (A.2) we have:
B

D
E

A B
=

1
|+i +
h+|A |+i1 2 h+|B
2
21
1
|+i
h+|A |i1 2 h|B
2
21

1
|i +
h|A |i1 2 h|B
2
21
1
|i .
h|A |+i1 2 h+|B
2
21

Executing some simple algebra, the last expression becomes

72

A B

=
=

1
[ (ax iay ) (bx + iby ) (ax + iay ) (bx iby )]
2
1
(2ax bx 2ay by ) = ax bx ay by ,
2

from which we work out the result of (3.3), that is


D

E D
E
=

= ~1 a
~2 b
A B
a b.

2. Quantum expectation value for the combined detection of polarization of


two photons that propagate along equal and opposite directions
Lets calculate expressions (3.31) and (3.32) for a state relative to a system composed of
direction, that is the
two photons, emitted by the same source, that propagate along the z
state described by Eq. (3.30),
1
|(1 , 2 )i = (|x, xi + |y, yi) ,
2
that can be rewritten as
1
|i = [|xi |xi + |yi |yi] .
2

(A.4)

In this representation of the state vector it is emphasized that the direction of polarization surely belongs to the plane xy, since it must necessarily be orthogonal to the direction
of propagation of the particle. In fact |xi and |yi represent the states corresponding the
and y
.
two orthogonal directions of polarization x
of the polarizers, and measuring the polarization of the
and b
Given two directions a
photons we will obtain for both the result +1 if the polarization of the photon is parallel
to the polarizer, 1 if it is orthogonal. What we are intending to evaluate are the single
and combined probabilities to obtain +1 and 1 for one or both the photons. In order
to do this, even in this case we can exploit (A.2), using the projection operator, whose
definition will be briefly given.
Consider two systems, 1 and 2, and some opportune Hilbert space associated with
them, H (1) and H (2) .
Definition (Projection operator). Consider an observable A whose eigenvalues and
eigenvectors are ak and |ak i respectively. The operator that projects the state of the total
system onto the subspace relative to the eigenvalue ai of A, measured only on the system
73

1 is given by
(1)

= |ai i hai | I (2) ,

where I (2) is the identity operator belonging to H (2) .


Now consider a second observable B with eigenvalues and eigenvectors bm and |bm i
respectively. The operator that projects the state of the total system onto the subspace
relative to the couple of eigenvalues {ai , bj }, associated respectively with the quantity A,
measured on system 1, and to the quantity B, measured on system 2, is given by
ij = |ai i hai | |bj i hbj | .
Therefore, calculating the probability that a measurement carried out on photon 1
gives result +1, is equivalent to projecting the state vector (A.4) onto the eigenspace
corresponding to that eigenvalue (analogously for 1), and the same argument can be
made for photon 2.
Lets consider, for instance, photon 1 and suppose that we want to determine its polarization by means of a measurement. For convenience we suppose that the eigenvector |xi
and the eigenvector |yi with the unit vector y
. Moreover,
coincides with the unit vector x
just to make the steps easier, since we obtain the result +1 in case of parallel polarization
and 1 in case of orthogonal polarization (to the direction of the polarizer, of course), we
readily see that the corresponding eigenvectors are
|+ia =

ax

!
= ak ,

ay

relative to the eigenvalue +1, and


|ia =

ax
ay

!
= a ,

relative to the eigenvalue 1.


Defining P+ (
a) the probability of photon 1 giving result +1 (i.e. the photon revealed
), we can calculate this value using the
when the polarizer is parallel to the unit vector a
projection operator on its relative subspace. In this case the operator assumes the form
(1)

+ = |+ia a h+| I,
thereby from (A.2) we get
74

(1)

P+ (
a) = h| + |i =
+


1
| hx | +ia |2 |x|2 + | hy | +ia |2 |y|2 +
2
1
[hx | +ia a h+ | yi hx | yi + hy | +ia a h+ | xi hy | xi] .
2

Making the opportune substitution and carrying out the calculations, one obtains
1
P+ (
a) = .
2
Analogously it is easy to calculate also the other probabilities, having
= P (b)
= 1.
P (
a) = P+ (b)
2
These are the probabilities of single detection for two photons.
As last result, we want to evaluate the quantum combined expectations. First of all,
calculating the probability for having both results +1 for the polarization measurements
is equivalent to projecting the state vector (A.4) on the eigenspace corresponding to this
pair of eigenvalues; an analogous procedure is valid for any other combination of the two
possible results.
Making use of the same previous conventions and adding only the eigenvalues related
to the two eigenvalues 1 for the second photon, it is easy to see that, for the same reasons
discussed in the case of photon 1, the eigenvalues for photon 2 are
|+ib =

bx

!
= bk ,

by

relative to the eigenvalue +1, and


|ib =

bx
by

!
= b ,

relative to the eigenvalue 1.


the combined probability of detecting +1 for photon 1 and +1 for
Defining P++ (
a, b)
photon 2, it is possible to evaluate this value using a projection operator, in this case
having the form
++ = |+ia a h+| |+ib b h+|.
Substituting in (A.2) we get

75

= h| ++ |i
P++ (
a, b)

1
=
| hx | +ia |2 | hx | +ib |2 + | hy | +ia |2 | hy | +ib |2 +
2
1
+
[hx | +ia a h+ | yi hx | +ib b h+ | yi + hy | +ia a h+ | xi hy | +ib b h+ | xi] .
2
Carrying out all the scalar products, we arrive at the final expression
=
P++ (
a, b)

1
1  2 1 2 2
b = |
(ax bx + ay by )2 =
a
a| |b| cos2 (ab ),
2
2
2

= 1, we have
from which, taking into account that |
a| = |b|
=
P++ (
a, b)

1
cos2 (ab ).
2

With a similar procedure it is straightforward to verify that


= 1 cos2 (ab )
P (
a, b)
2
and
= P+ (
= 1 sin2 (ab ).
P+ (
a, b)
a, b)
2
Finally, it is sufficient to observe that, because of the simmetry of the model, one has

= P (

P++ (
a, b)
a, b),
= P+ (

P+ (
a, b)
a, b).
Thus, since for the probability laws it must be
+ P (
+ P+ (
+ P+ (
= 1,
P++ (
a, b)
a, b)
a, b)
a, b)
we can conclude that

a, b)
1
= P+ (
= 1 2P++ (
P+ (
a, b)
a, b)
= sin2 (ab ).
2
2

76

Bibliography
[1] E. Schmidt, Math. Ann. 63 (1907) 443.
[2] A. Einstein, B. Podolsky, N. Rosen, Can Quantum-Mechanical Description of Physical
Reality Be Considered Complete?; Phys. Rev. 47 (1935) 777.
[3] N. Bohr, Can Quantum-Mechanical Description of Physical Reality Be Considered
Complete?; Phys. Rev. 48 (1935) 700.
[4] D. Bohm, Y. Aharonov, Discussion of Experimental Proof for the Paradox of Einstein,
Rosen, and Podolsky; Phys. Rev. 108 (1957) 1070.
[5] A. Peres, Einstein, Podolsky, Rosen, and Shannon; arXiv:/0310010v1 (2 Oct 2003).
[6] J. Bell, On the Einstein-Podolsky-Rosen Paradox; Physics 1 (1964) 195.
[7] J. F. Clauser, M. A. Horne, A. Shimony and R. A. Holt, Proposed Experiment to Test
Local Hidden-Variable Theories; Phys. Rev. Lett. 23 (1969) 880.
[8] J. Bell, Introduction to the hidden-variable question. Foundations of Quantum Mechanics; Proceedings of the International School of Physics Enrico Fermi, course IL,
New York, Academic (1971), pp 171-81.
[9] A. Aspect, Rendiconti della Ottava Conferenza Internazionale sulla Fisica Atomica
(2-6 Agosto 1982), G
oteborg, Svezia. Testo pubblicato su Atomic Physics 8. Edito
da I. Lindgren, A. Rosen e S. Svanberg.
[10] A. Aspect, To be or not to be local; Nature, Vol. 446 (19 April 2007), pp. 866-867.
[11] C. Shannon, Communication Theory of Secrecy Systems; Bell System Technical Journal, 28:4, October 1949, pp. 656-715.
[12] G. C. Ghirardi, A. Rimini, T. Weber, Lett. Nuovo Cimento, Soc. Ital. Fis. 27 (1980)
293.
77

[13] N. Herbert, FLASHA superluminal communicator based upon a new kind of quantum measurement; Foundations of Physics, 12 (1982) 12.
[14] A. Peres, How the No-Cloning Theorem Got its Name; arXiv:/0205076v1 (14 May
2002)
[15] T. De Angelis, F. De Martini, E. Nagali, F. Sciarrino, Experimental text of the NoSignaling Theorem; arXiv:/07051898v2 (17 May 2007).
[16] R. Y. Chiao and A. M. Steinberg, Progress in Optics XXXVII, ed. by E. Wolf (Elsevier, Amsterdam, 1997); Physica Scripta T76 (1998) 61.
[17] J. C. Garrison, M. W. Mitchell, R. Y. Chiao, E. L. Bolda, Phys. Letters 245A (1998)
19.
[18] G.C. Ghirardi, A. Rimini, T. Weber, A Model for a Unified Quantum Description
of Macroscopic and Microscopic Systems. Quantum Probability and Applications; L.
Accardi et al. (eds), Springer, 1985.
[19] G.C. Ghirardi, A. Rimini, T. Weber, Unified Dynamics for Microscopic and Macroscopic Systems; Phys. Rev. D 34 (1986) 470.
[20] Y. Shih, Y. Kim, Quantum Entanglement: from Poppers Experiment to Quantum
Eraser; Optics Communications, 179 (2000), pp. 357-369.
[21] U. Fano, Description of States in Quantum Mechanics by Density Matrix and Operator Techniques; Rev. Mod. Phys 29 (1957) 74.
[22] D. Bohm, Phys. Rev. 85 (1952) 166.
[23] D. Bohm, B. J. Hiley and P. N. Kaloyeru, Phys. Rep. 144 n.6 (1987) 321.
[24] Gregg Jaeger. Entanglement, Information, and the Interpretation of Quantum Mechanics. Springer, 2010.
[25] Valter Moretti. Spectral Theory and Quantum Mechanics With an Introduction to the
Algebraic Formulation. Springer, 2013.
[26] Albert Messiah. Quantum Mechanics. Dover Publications, 1999.
[27] John Stewart Bell. Six Possible Worlds of Quantum Mechanics. Proceedings of the
Nobel Symposium 65: Possible Worlds in Arts and Sciences. Stockholm, August 1115, 1986.
78

[28] John Stewart Bell. Speakable and Unspeakable in Quantum Mechanics. Cambridge
University Press, 2004.
[29] Oreste Nicrosini. Paradosso EPR e Teorema di Bell. Quaderni di Fisica Teorica,
Collana curata da Sigfrido Boffi, 1991.
[30] Abraham Pais. Subtle is the Lord: The science and the life of Albert Einstein.
Oxford University Press, 1982.
[31] Bernard DEspagnat. Conceptual Foundations Of Quantum Mechanics. Advanced
Book Classic, 1999.
[32] Detlef D
urr, Stefan Teufel, Bohmian Mechanics. The Physics and Mathematics of
Quantum Theory. Springer, 2009.
[33] B. Odom, D. Hanneke, B. DUrso, G. Gabrielse, New Measurement of the Electron
Magnetic Moment Using a One-Electron Quantum Cyclotron; Phys. Rev. Lett. 97
030801 (2006).

79

También podría gustarte