Вы находитесь на странице: 1из 19

Bell’s Theorem

Quantum Physics - December 2016

Joan Ariño, Marta Florido, Adam Teixidó

1
Contents

1 The way to Bell’s dilemma 3


1.1 EPR Paradox: A case against quantum mechanics . . . . . . . . . . . . . . 4
1.2 Bell’s approach: On the Einstein-Podolsky-Rosen Paradox . . . . . . . . . . 5

2 Bell’s inequalities 6
2.1 Bell’s inequalities hypotheses . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2 Preskill’s version of Bell’s inequality: a beautiful and simple derivation . . . 8
2.3 The original Bell’s inequality . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.4 The Clauser-Horne-Shimony-Holt inequality: a more general result . . . . . 11

3 Quantum violations of Bell’s inequality and Bell states 12


3.1 Quantum state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.2 Entanglement and Bell states . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.3 Bell’s original inequality violation . . . . . . . . . . . . . . . . . . . . . . . . 14

4 Loopholes in experimental tests of Bell’s inequality 16

5 Bell’s theorem consequences 18

6 Bibliography 19

2
1 The way to Bell’s dilemma

Quantum mechanics appearance challenged many assumptions which physicists had always
been tacitly making. Realism, the existence of a reality independent of an observer, is
probably the most important one. This idea was dismissed by some proponents of quantum
theory, who defended that there are certain physical quantities that don’t take any value
or, equivalently, that don’t exist before their measurement is performed.
Those physical quantities, such as momentum, position, energy, time, spin, were classi-
cally thought as ones that defined a system and dictated its evolution. However, quantum
theory brings to us the principle of complementarity: observable quantities (spin-x axis,
spin-y axis) and behaviors (particle/wave) come in pairs which can’t be observed simul-
taneously, and they are also tied by an uncertainty relation; then, the more precisely we
know one of them, the more information we will lose about the other one.
This impossibility of simultaneous determination is not a simple quirk of the theory.
In fact, it is regarded as a real thing and not as a measurement deficiency because it is an
intrinsic property of the wave-function itself.

Interpretations of the quantum quirks

Pascual Jordan, a key physicist involved in the birth of quantum mechanics, declared
with emphasis that observations not only disturb what has to be measured, but they also
produce it. In a measurement of position, for example, as performed with the gamma
ray microscope, the electron is forced to a decision. We compel it to assume a definite
position; previously it was, in general, neither here nor there; it had not yet made its
decision for a definite position [...] If by another experiment the velocity of the electron
is being measured, this means: the electron is compelled to decide itself for some exactly
defined value of the velocity [...] We ourselves produce the results of measurement.1
This interpretation was deepened due to the inexplicably results of several experimental
set-ups from the point of view of classical dynamics. For example, in the case of spin on
the Stern-Gerlach experiment, it is not possible to make up an explanation about why the
electrons get deflected into exactly two opposite beams when trying to assign a classical
spin to the electron.
The constant inconsistencies between the existing theory and the experimental results
generated some even more extreme views, as that of Bohr, who stated that particles were
not even really there:
Bohr once declared when asked whether the quantum mechanical algorithm could be
considered as somehow mirroring an underlying quantum reality: There is no quantum
world. There is only an abstract quantum mechanical description. It is wrong to think
that the task of physics is to find out how nature is. Physics concerns what we can say
about nature.”2
An opposite interpretation of quantum mechanics, the one that Einstein supported, is
that quantum mechanics had indeed a great statistical predictive power, but that it wasn’t
a complete model. Since for him the wave function could not be the complete picture, there
should exist a possibility that some complementary quantities that the wave function does
not determine exactly, exist independently of it, and that the wave function only reflects
1
M. Jammer, ibid, p.161, quoting E. Zilsel, “P. Jordans Versuch, den Vitalismus quanten mechanish zu
retten”, Erkenntnis 5 (1935) 56 - 64
2
M. Jammer, The Philosophy of Quantum Mechanics, John Wiley (1974), p.204, quoting A. Petersen,
Bulletin of the Atomic Scientist 19 (1963)

3
our lack or impossibility of knowledge.
In spite of all the controversy and being probabilistic, the accuracy of quantum me-
chanics predictions, and its explanatory capability of all the novel and bizarre experimental
results, generated a fast acceptance of it.
While the flourishing theory was on test, it wasn’t still known the broadness of im-
plications differing from classical mechanics and even from special relativity. The task of
clarifying them created many discussions (which will be addressed in short), Moreover, its
implications were more fundamental than was initially thought, as Bell’s Theorem proved.

1.1 EPR Paradox: A case against quantum mechanics

The paper written by Einstein Podolsky and Rosen started by considering mutually non
defined variables such as the ones that the Heisenberg principle states, position and mo-
mentum.
We will simplify their argument by saying that we have two particles A and B which
started with same momentum and position. When we measure the momentum of particle
A we will immediately know the momentum of particle B. We could also measure particle
B position, and through some calculations, recover A’s position. However, now we would
exactly know both properties, which is strictly forbidden by Heisenberg’s principle. Then,
somehow, when we measure the momentum of one particle, immediately the position of
the other must become indefinite to forbid us from knowing more than allowed. Thus it
seems that, if the wave-function is the only one that explains all measurements and there
is nothing more (in other words, that the quantum mechanical treatment is complete) the
particles must be communicating somehow information about they being measured and
the result they showed up; and strikingly they must do so instantaneously.
At this point, they considered that reality is local, which means: there can not happen
any interaction between systems faster that a speed limit, the speed of light. The conse-
quences of not imposing this assumption, as will be discussed in the following section, can
be disastrous or, in the best case, go against the spirit of special relativity. By applying
this concept of locality, they argued that reality must be separable, that we must be able
to describe both systems independently from one another.
To accomplish this on the A and B particle system, they considered that certain
variables existed previously to measurement and determined exactly its result. In this
way they would account for the apparent transmission of information between the two (in
the future they will be called entangled) particles, because all the possible measurement
results would be already contained locally in each particle.
As the wave function doesn’t take into account those variables, it cannot be the whole
picture. Therefore they reached the conclusion that has been given before as the point
of view of Einstein: that the requirement of locality implies that quantum mechanics is
not a complete theory. So they proposed that there may exist a more complete theory
of reality, which allows us to maintain locality by adding variables. And this could be a
non-probabilistic theory, thereby recovering the characteristics of classical mechanics.
On their own words: While we have thus shown that the wave function does not provide
a complete description of the physical reality, we left open the question of whether or not
such a description exists. We believe, however, that such a theory is possible..3
As we said, their shown is only true when assuming a reasonable definition of real-
3
EPR original paper: A. Einstein et al. ”Can quantum mechanical description of physical reality be
considered complete?”, Physical Review, (1935)

4
ity, which among other unspecified assumptions is that measurements on Q can’t have
instantaneous effects on the reality of another system P.
Bell’s categorical answer to the question that had been left open is no.
He demonstrated that such a proposed theory is not compatible neither with quantum
mechanics predictions nor reality, after the necessary experimentation to prove it was done,
as we shall explain later.

1.2 Bell’s approach: On the Einstein-Podolsky-Rosen Paradox

Bell continued developing the argument proposed by the EPR Paradox, and by using
simple probabilistic arguments, he arrived to the conclusion that the theory envisioned
by the paper’s writers could not be achieved (given that quantum mechanics was right,
because in the days that he proposed his theorem, the experimental set-ups weren’t precise
and adequately enough done to ensure the result of his Theorem).
As Bell said about Einstein’s proposed modifications to Quantum mechanics, he adopted
the more rational view. However, his idea wasn’t the one that ruled reality.
Contrary to what we might think, and following what we could infer from his quote,
he wasn’t trying to justify quantum mechanics; he instead started on the side of Einstein.
Moreover, his proof is not an ultimatum on the certainty of the orthodox quantum me-
chanical interpretation; on the contrary, after accepting the result of his Theorem, he even
thought that the opposite could be true and supported a theory of hidden variables: for
instance, he was specially impressed by the David Bohm’s theory of pilot waves.
However, as has been said before, this theory differed from the ones proposed by Ein-
stein because it does not preserve locality. He thought that the easiest solution to recover
the Lorentz invariance of the theory, which is what is needed in order to reconcile it with
special relativity, was to add an aether. The same aether theory that was previously
considered, during years, and finally dismissed and ruled out by, among other experi-
ments, Michelson Moreley’s one, and by Maxwell’s electromagnetism, and supplanted by
the impressive Einstein’s special theory of relativity, which seemed to solve elegantly the
situation. This aether could be the unique medium through which instantaneous action
between different points of spacetime could be transmitted. However, it would be unde-
tectable, such as the ones that appear in some proposed solutions to Michelson’s Moreley
observations of non relative motion to the luminiferous aether.
The reason why Bell opposed to the main interpretation of Quantum Mechanics, and
started considering all this possibilities, is that he thought it to be a non clean theory.
For example, he thought that observer and observation were very ill-defined concepts,
which only induced an artificial distinction, so in his later years he refused to talk about
observables, and instead replaced them with new entities referred to as beables (which
refer to the elements that the theory in question considers to be real).

5
2 Bell’s inequalities

Bell’s theorem essentially states that any physical theory of local hidden variables is in-
compatible with the empirical predictions of quantum mechanics.
Its proof involves the violation of Bell-type inequalities. In this section, the hypotheses
underlying Bell’s inequalities will be explained, proceeding then to derive some of them,
which can be shown to be violated by quantum mechanical predictions for specific systems
(therefore proving Bell’s theorem).
In particular, the derivation of Preskill’s version of Bell’s inequality will be given
which is rather simple and beautiful. Then, Bell’s original inequality will be obtained.
And finally, a slightly modified version of Clauser-Horne-Shimony-Holt inequality will be
presented as an improved version of Bell’s original inequality since it will be observed
that its generalized hypothesis include it as a result and they are easier to control in
experimental setups.

2.1 Bell’s inequalities hypotheses

All Bell-type inequalities are based in the following three assumptions:

1. Hidden variable model:


Specifically, Bell’s Theorem imposes restrictions on the so called theories of hidden
variables, the ones which assume an independent reality for all measurements.
In a hidden variable model, a parameter λ is used to describe the state of a physical
system. The probability that the system is in the Rstate λ in a specific experiment
is given by the probability distribution ρ(λ), and ρ(λ) dλ = 1 (λ might denote
a single variable or a set which can be discrete or continuous, or even a set of
functions. Since for the purpose of this section, the nature of λ will not be relevant,
it will be simply written as if it were a single continuous parameter). In any case,
it is worth noticing that hidden variable models consider that every particle exists
with a given speed, position, spin... In opposition to the Copenhagen interpretation
which considers that those properties just become to exist when measured.
Some hidden variable theories such as Pilot Wave theory or its more modern coun-
terpart Bohmian Mechanics have appeared. They try to give the same predictions
as Quantum Mechanics but being both deterministic and offering a description of
reality independent of an observer.
However, they were dismissed due to being clearly non-local theories, considering a
wave-function that can be modified by and in turn can modify instantaneously every
particle in the universe. Another reason for its lack of acceptance was due to the
higher simplicity of Quantum Mechanics mathematics and formalism.

2. Special Relativity: locality requirement Special relativity imposes heavy re-


strictions on a possible non-local theory. For example, a faster than light trans-
mission of information would appear as a signal traveling past in time for the right
observer. So, through the use of an intermediary it would be possible to affect your
own past and then generate paradoxes (see: Figure 1).
However, a number of non-local theories have emerged that avoid those paradoxes.
One of those is choosing one single frame at which ultra-luminic signaling is permit-
ted. This way, instantaneous effects are allowed but it is no longer possible to affect
your past light cone. It is significant to point out that this comes at a cost: different

6
inertial reference frames do not share indistinguishable nature laws, a key idea of
mechanics.
In addition although traditional conception of special relativity does not allow any
kind of causation it is worth noticing it that neither EPR nor Bell theorem offer
the possibility of transmitting information. As a result, this causation, although
classical forbidden cannot result on a paradox such as the detailed on Figure 1

Figure 1: If faster than light communication was allowed in every inertial system of ref-
erence the following paradox would be possible: you would be allowed to hire a distant,
fast-moving relativistic hitman outside your light cone. Then the killer could shoot a
tachyon, a particle moving faster than light in his reference frame, which could cause
the death of your grandmother. Notice how the paradox is avoided if faster than light
communication is allowed in one unique reference frame: if going faster than light just
works in your frame, the shot tachyon wouldn’t be allowed, as it is going to your past. If
the hitman is in the chosen frame, the hiring call is impossible as it is a killing request
originated on the future going to the past which is neither allowed.

3. “Free variables” assumption: it is also called freedom of choice or no super-


determinism, and it is basically assuming that the measurement settings of the ex-
periments are chosen randomly or freely, which is equivalent to requiring that the
probability distribution ρ(λ) is independent of the measurement settings. However,
it is important to notice that, if the hidden variable λ posited by the candidate
physical theory in consideration includes also the description of the state of the
measurement apparatus, then this assumption will be necessarily violated. These
theories were described as being superdeterministic by John Bell. Although it is pos-
sible to think that one can never be sure of whether the same factors λ that influence
the outcomes of the experiments will also influence the measurement settings, John
Bell dismissed this way of thinking since, if free variables didn’t exist, this skepticism
would lead to a conspiratorially entangled world where, for example, the results of
scientific experimentation should be disregarded, as well as to strange consequences
as those owned by causal chains that go faster than light. Clauser, Horne, Shimony
and Holt also found it reasonable not to worry about the possibility of conspiracies
if no specific causal linkage between λ and the apparatus settings was proposed by

7
the candidate physical theory.

2.2 Preskill’s version of Bell’s inequality: a beautiful and simple deriva-


tion

In the derivation of the Bell inequality proposed by Preskill, following Mermin’s suggestion,
a system formed by two objects is studied. Let’s take into consideration three of their
properties x̂, ŷ, ẑ that can only take two values: −1 and 1. The hypotheses that will be
used in the derivation of the inequality will be: the two objects are identical (in other
words, if the same property of the two objects is measured, the same outcome will be
obtained), counterfactual definiteness, locality and free variables, which have already been
explained in detail.

Derivation of Preskill’s version of Bell’s inequality

The notation Psame (x̂, ŷ) will be used to express the probability that when property x̂
of the first object is measured and property ŷ of the second object is measured, the
same outcome is obtained (for example, since the two objects are identical, Psame (x̂, x̂) =
Psame (ŷ, ŷ) = Psame (ẑ, ẑ) = 1).
Preskill’s version of Bell’s inequality holds that, under these conditions,

Psame (x̂, ŷ) + Psame (x̂, ẑ) + Psame (ŷ, ẑ) ≥ 1 .

Let’s consider first its derivation from the probability diagram in figure 2.
The gray area represents Psame (x̂, ẑ) (that is, the
probability that the outcome when measuring property
x̂ of the first object will be the same as the outcome ob-
tained when measuring property ẑ of the second object).
In the same way, the dashed area represents Psame (x̂, ŷ).
Since the dotted area is complementary to the two re-
gions just mentioned, it will represent Pdif f (x̂, ŷ and ẑ),
which is the probability that the outcome when measur-
ing x̂ of the first object is different from the outcome
obtained when measuring ŷ or ẑ of the second object.
Figure 2: Probability diagram. We have assumed that the properties can only have
two possible outcomes when they are measured, so this
implies that in the dotted region, the outcomes for the
measurement of ŷ and ẑ of the second object will be the same. Then, since the two
objects have identical properties, Psame (ŷ, ẑ) (which is the probability that the outcome
of measuring ŷ of the first object and ẑ of the second one) will be larger than or equal to
the dotted region.
Therefore, and due to the fact that the sum of the three regions is larger than or equal
to the full circle of area 1, the aforementioned Preskill’s version of Bell’s inequality will
hold. It can also be seen in a more intuitive way that it is valid: the sum of the three
probabilities will be larger than or equal to 1 since all the possible combinations are taken
into account, but some of them might be counted more than once. This inequality can be
shown to be violated by quantum mechanical predictions.
Notice that, if counterfactual definiteness wasn’t assumed, it could not be inferred that

8
in the dotted region, the outcomes for properties ŷ and ẑ of the second object should be
the same, since only one measurement at a time can be done in the same object and the
properties wouldn’t exist before they were measured. In the other hand, if locality wasn’t
assumed, no deductions could be done because the measurement of a property of the first
object could modify the outcome of measuring another property in the second one.

2.3 The original Bell’s inequality

Bell’s original inequality first appeared in November 1964, in Bell’s paper “On the Einstein
Podolski Rosen paradox”, and it opened the line of research about Bell’s theorem. It is
currently a very active and relevant topic for the scientific community, since it establishes
an incompatibility between the two fundamental pillars of modern physics: relativity and
quantum mechanics.

Hypotheses and notation

In its derivation, it is assumed that a hidden variable model (with hidden variable λ and
probability distribution ρ(λ)) describes the system, which is basically formed by two spin
one-half particles which are moving freely in opposite directions, and with a perfectly
anti-correlated spin (that is, if the measurement of the spin of the first particle along
a unit vector yields and outcome of +1, then the other particle must yield the opposite
outcome, −1). The measurements on their spins are made by Stern-Gerlach magnets along
directions defined by unit vectors â, b̂, ĉ . . . .
Apart from the hidden variable model assumption, there are additional hypotheses
that must be done: locality (which means that the events happening in the first particle
(the choice of the measurement settings and the act of measurement) will be irrelevant
for the outcome of the second one and the other way around), and free variables (in other
words, the measurement settings, for example x̂ and ŷ, are free variables in the sense that
they do not influence the state λ of the system, which means that ρ(λ) = ρ(λ|x̂, ŷ)).
In the derivation of Bell’s inequality, x̂ will denote the property that is measured in
the first particle, whose outcome will have a value X which will depend only on x̂ and
λ. In the same way, ŷ will denote the property measured in the second particle whose
outcome will be Y . Therefore, X(x̂, λ) will be the outcome obtained when measuring x̂
in the first particle, and Y (ŷ, λ), the outcome obtained when measuring ŷ in the second
particle. Now, let E(x̂, ŷ) denote the expectation value of the product of the outcomes
of measuring property x̂ in the first particle and property ŷ in the second one. It will be
given by the following expression:
Z
E(x̂, ŷ) = X(x̂, λ)Y (ŷ, λ)ρ(λ) dλ ,

where the integral does not have a strict meaning, since λ could be any discrete or con-
tinuous parameter or set of functions, for example, but rather emphasizes that an average
over the possible states of the system described by λ is being calculated.

Derivation of Bell’s original inequality

Bell’s inequality can be stated as follows:

1 + E(ŷ, ẑ) ≥ |E(x̂, ŷ) − E(x̂, ẑ)| .

9
But where does it come from? John Bell was focused in finding out whether a theory
of local hidden variables was compatible with quantum mechanics predictions, as Einstein
had proposed. Although violations of Bell’s inequalities by quantum mechanics will be
discussed further in the next section, in order to be able to fully understand the origin
of Bell’s inequality
R it is worth mentioning here that quantum mechanics predicts that
E(x̂, ŷ) = X(x̂, λ)Y (ŷ, λ)ρ(λ) dλ must be equal to the value −x̂ · ŷ, where x̂ and ŷ are
the unit vectors that define the measurement settings of the experiments, whose minimum
is reached at x̂ = ŷ, with the value −1.
From here, he observed that E(x̂, ŷ) can only be −1 at x̂ = ŷ as is predicted by
quantum mechanics, if and only if X(x̂, λ) = −Y (ŷ, λ) except at a set of points λ of zero
probability, for all the possible directions x̂ = ŷ. Then, he used this fact to rewrite E(x̂, ŷ)
as follows:
Z
E(x̂, ŷ) = − X(x̂, λ)X(ŷ, λ)ρ(λ) dλ .

Then, he introduced another unit vector ẑ and obtained that


Z
E(x̂, ŷ) − E(x̂, ẑ) = − (X(x̂, λ)X(ŷ, λ) − X(x̂, λ)X(ẑ, λ))ρ(λ) dλ =
Z
= (X(x̂, λ)X(ŷ, λ)(X(ŷ, λ)X(ẑ, λ) − 1)ρ(λ) dλ
R R
Again, using that |X| ≤ 1 and | f | ≤ |f |, the expression above can be rewritten as
follows: Z
|E(x̂, ŷ) − E(x̂, ẑ)| ≤ (1 − X(ŷ, λ)X(ẑ, λ))ρ(λ) dλ
R
Since the second term of the integral is actually E(ŷ, ẑ) and ρ(λ)dλ = 1,

1 + E(ŷ, ẑ) ≥ |E(x̂, ŷ) − E(x̂, ẑ)| .

And this is Bell’s original inequality! If it is violated by the predictions of quantum


mechanics, quantum mechanics will be incompatible with a local hidden variable model,
for the inequality Bell obtained follows from the assumption that X(x̂, λ) = −Y (ŷ, λ) if
x̂ = ŷ, which is the only way in which the condition E(x̂, ŷ) = −1 can be satisfied if x̂ = ŷ,
and therefore it will also be a requirement that must be met by any local hidden variable
model that completes quantum mechanics. However, it will be shown in section X that
quantum mechanics predictions indeed violate the inequality. So there will not be any
local hidden variable model compatible with it, which is what Bell’s theorem states.

Limitations of Bell’s original inequality

Bell’s inequality, however, has fall into disuse because the perfect anticorrelation condition
that the properties of the two components of the system must meet is very difficult to
reproduce experimentally.
In contrast, Clauser-Horne-Shimony-Holt inequality, which was developed in 1969, 5
years after John Bell’s original inequality was first published, and then slightly modified by
Bell himself in 1971, is an improved version of it and has been used much more commonly
in Bell test experiments. The essential reason is that the perfect anticorrelation condition
is no longer required, so its assumptions are weaker than the ones used in Bell’s original
inequality.

10
2.4 The Clauser-Horne-Shimony-Holt inequality: a more general result

The Clauser-Horne-Shimony-Holt (CHSH) inequality is derived based on the study of a


system formed by two objects. x̂ still denotes the property measured in the first object
whose outcome is X, and ŷ, the property of the second object that is being measured
whose outcome is Y . However, in CHSH inequality the perfect anticorrelation condition
is not required at all, and the only restriction imposed on X and Y is that |X|, |Y | ≤ 1.
It will be supposed that the property x̂ to be measured in the first object can be chosen
in x̂ ∈ {x̂1 , x̂2 }. In the same way, ŷ ∈ {ŷ1 , ŷ2 }.
The usual hypotheses will be used: hidden variable model (with hidden variable λ),
local causality and free variables. However, in this case, stochastic hidden variables are
also taken into consideration in the derivation of the inequality (that is, theories that
for each possible state of the system described by λ do not predict specific outcomes, but
probabilities distribution for the outcomes), by defining E(x̂, ŷ) (also called the correlation
between the outcomes X and Y ) in the following way:

Z X Z
E(x̂, ŷ) = XY P (X|x̂, λ)P (Y |ŷ, λ)ρ(λ) = X̄(x̂, λ)Ȳ (ŷ, λ)ρ(λ) dλ ,
X,Y

where X̄ is thePexpected value of X for the state determined by a specific λ and x̂ (that
is, X̄(x̂, λ) = X XP (X|x̂, λ)) and hence |X̄| ≤ 1 . Ȳ is defined analogously.
The CHSH is obtained from considering combinations of different correlations. Firstly,
Z
E(x̂1 , ŷ1 ) ± E(x̂1 , ŷ2 ) = X̄(x̂1 , λ)(Ȳ (ŷ1 , λ) ± Ȳ (ŷ2 , λ))ρ(λ) dλ
Z
E(x̂2 , ŷ1 ) ∓ E(x̂2 , ŷ2 ) = X̄(x̂2 , λ)(Ȳ (ŷ1 , λ) ∓ Ȳ (ŷ2 , λ))ρ(λ) dλ

Now, the absolute value


R ofR these quantities is taken into consideration and, using the
fact that |Ā| ≤ 1 and | f | ≤ |f |, the following inequalities are obtained:
Z
|E(x̂1 , ŷ1 ) ± E(x̂1 , ŷ2 )| ≤ |Ȳ (ŷ1 , λ) ± Ȳ (ŷ2 , λ)|ρ(λ) dλ
Z
|E(x̂2 , ŷ1 ) ∓ E(x̂2 , ŷ2 )| ≤ |Ȳ (ŷ1 , λ) ∓ Ȳ (ŷ2 , λ)|ρ(λ) dλ

A final step to complete the derivation of the CHSH inequality is just noting that
|x ± y| + |x ∓ y| can take the values 2x (if |x| ≥ |y| and x ≥ 0), −2x (if |x| ≥ |y| and x ≤ 0),
2y (if |y| ≥ |x| and y ≥ 0) and −2y (if |y| ≥ |x| and y ≤ 0). Due to the fact that |Ȳ | ≤ 1,

|E(x̂1 , ŷ1 ) ± E(x̂1 , ŷ2 )| + |E(x̂2 , ŷ1 ) ∓ E(x̂2 , ŷ2 )| ≤ 2 .

This is Clauser-Horne-Shimony-Holt inequality which, as has already been said, repre-


sents a generalization of Bell’s original inequality, since it takes into consideration stochas-
tic local hidden variables, and it doesn’t impose a perfect anticorrelation condition on the
components of the system.
In the next section, it will be explained how quantum mechanics violates Bell-type
inequalities, and therefore Bell’s theorem will be finally proven, thus paving the way to
the reflection on Bell’s theorem deep consequences in physics and its role in the scientific
world nowadays.

11
3 Quantum violations of Bell’s inequality and Bell states

To understand why quantum mechanics gives us a prediction of probabilities that violates


the Bell inequality, we will have to introduce how it describes the world and deduces its
predictions.

3.1 Quantum state

Imagine that we have an isolated quantum system. A vector in a complex space, which
we will call ket and denote as |ψi, describes a state of the quantum system, or in short, a
quantum state.
An example of quantum state is the spin: if we choose an axis of measurement, |↑i
means that the spin is upwards in that axis, and |↓i means that the spin is downwards.
Now we need to know how to obtain the probability of a measurement result. To do
so, we have to introduce another vector called bra hφ|, that acts upon a state |ψi. The
result will tell us the probability that we measure the system described by |ψi to be on
the state |φi.
To obtain that probability we follow the Born’s Rule:

p(φ) = | hφ|ψi |2 (1)

Where, in the end, the composition of the bra and the ket, called bracket, is the dot or
scalar product (in fact, it is a generalization of it to complex spaces, called inner product,
but we will keep the previous name for simplicity and because it is computed in the same
way, except that the second vector is conjugated). It is also important to note that every
ket has one associated bra, and that its scalar product must be: hψ|ψi = 1, then in this
way the probability is normalized to one.
Even a wave function ψ can be considered as a quantum state |ψi. Then, the last
mentioned condition, turns out to be the usual normalization condition, if we use that the
”scalar” product in the function space is defined as:
Z ∞
hf, gi = f (t)g(t)∗ dt (2)
−∞

Where the star means that we take its conjugate. By plugging in the wave function,
we get what we expected:
Z ∞
1 = hψ, ψi = ψ(t)ψ(t)∗ dt (3)
−∞

If we take a look again to the case of spins, now we know: h↑|↑i = 1


Instead, if we compute (taking into consideration that spin up and spin down exclude
one each other): h↑|↓i = 0
Since this is a scalar product, it means that up spin and down spin are orthogonal. In
fact, they are a basis for the spin of a particle such as an electron, that has spin 1/2.
Quantum states have the weird property that a linear combination of them (with
complex coefficients, because we are in a complex space) is also a quantum state. This is
what is usually called a superposition, and means that a quantum state can be anywhere
in between two (or any number, or even continuum, of) different states.

12
For example:
|ψi = α |↑i + β |↓i α, β ∈ C (4)

That would mean that the spin of the particle is in any possible direction. Because,
as we said, they form a basis, or in other words, we can get to any possible spin state
through linear combinations of them.
Here we have described the bra-ket notation, in a practical way, and now we will use
it to break a Bell inequality.

3.2 Entanglement and Bell states

Let’s move to a system with two particles.


If the first particle is described by: |ψ1 i = √1 |↑i + √1 |↓i
2 2
(Which is a superposition state with equal probability of its spin being measured down
or up).
And the second particle: |ψ2 i = |↓i
Then the state of the whole combined system (S) can be expressed as their product,
concisely, their tensor product, that has the same properties as the usual product:
 
1 1 1 1
|ψS i = |ψ1 ψ2 i = |ψ1 i |ψ2 i = √ |↑i + √ |↓i |↓i = √ |↑↓i + √ |↓↓i (5)
2 2 2 2

And if we take as an example the expression |↑↓i, it doesn’t mean anything more than
that the first particle spin is up and the second particle spin is down.
In this case, every particle had its own description, and by multiplying them what we
are doing is simply writing their states one next to the other, in the same ket.
However, now an interesting phenomenon appears, we can get states which can not be
separated as the product of two states, one for each particle.
One of such inseparable states is:

1  
|ψi = √ |↑↓i + |↓↑i (6)
2
That can be interpreted as: the particles always have opposite spins when measured
in the same axis, and we could get this state after a disintegration where two outgoing
particles must have total spin 0.
This is what is called an entanglement between the two particles, because they are tied
in some way, and one of them cannot be described without the other.
We can prove it by taking the most general expression for every particle and multiplying
them:

(c1 |↑i + c2 |↓i)(c3 |↑i + c4 |↓i) = c1 c3 |↑↑i + c1 c4 |↑↓i + c2 c3 |↓↑i + c2 c4 |↓↓i (7)

By equating it to our entangled state |ψi, we ask it to have coefficients c1 c3 and c2 c4 as


zero. In turn, this implies that either c1 c4 or c2 c3 are equal to zero, making impossible to
make our proposed state equal to |ψi because those coefficients must be both non-zero.
The ”simplest” maximally entangled quantum states are called Bell states, because
they are the ones studied in Bell inequalities and violate them the most. In the case that

13
we are considering (two half-spin particles) there are only four possible Bell states, and
the |ψi presented before is one of them.

3.3 Bell’s original inequality violation

Remember that the original Bell’s inequality was

1 + E(ŷ, ẑ) ≥ |E(x̂, ŷ) − E(x̂, ẑ)|

Where the x̂, ŷ, ẑ are arbitrary axis (we take them as unitary). E(x̂, ŷ) gives a measure
of the correlations between spin measurements when performed in the axis x̂ for the first
electron and in the axis ŷ for the second.
To derive that specific inequality, we require perfectly anti-correlated spin measure-
ments, if we do them on the same axis.
If we consider the previous entangled state:
1  
|ψi = √ |↑↓i + |↓↑i
2

When the first electron is measured to be |↑i in a certain axis x̂, then the wave-function
”collapses” into a single state, coherent with our measurements:

|ψi = |↑↓i

Now, if we measure the second electron, its spin must be |↓i in the x̂ axis. If we define,
for electron i: (
+1 if spin is |↑i
Si
−1 if spin is |↓i
Then we can evaluate:

E(x̂, x̂) = S1 (x̂, |↑↓i) · S2 (x̂, |↑↓i) = +1 · (−1) = −1

And so we founded the desired perfect anti-correlation. However, if instead of measuring


the spin of the second electron in the x̂ axis, we measure it in an arbitrary ŷ, the probability
that the result is either an |↑i or an |↓i state can still be calculated, because we know from
the first measurement that the second electron’s spin is |↓i in the x̂ direction.
Now, if the axis ŷ of measurement differs by an angle θ from the axis x̂ (its angle is
given by the expression: cos θ = x̂ · ŷ), we can express |↓ix̂ (which is a short notation for
the spin being down in axis x̂ ) in the new basis that is on the same direction as the axis
ŷ, |↑iŷ , |↓iŷ , by making the corresponding projections on the complex space:

θ θ
|↓ix̂ = sin |↑iŷ + cos |↓iŷ
2 2

Now, the probabilities can be obtained by using the Born’s Rule. First, we apply the
bras correspondent to the wanted measurements in ŷ axis:

θ
θ
θ
↓ŷ ↓x̂ = sin ↓ŷ ↑ŷ + cos ↓ŷ ↓ŷ = cos
2 2 2


θ
θ
θ
↑ŷ ↓x̂ = sin ↑ŷ ↑ŷ + cos ↑ ↓ŷ = sin
2 2 ŷ 2

14
Then we obtain the probabilities by squaring:
θ
| ↓ŷ ↓x̂ |2 = cos2


2

θ
| ↓ŷ ↑x̂ |2 = sin2


2
Now we can calculate the expected value of the outcomes, by assigning, again a +1 to
a spin up and a −1 to a spin down:
θ θ
S2 (ŷ, |↑↓ix̂ ) = sin2 − cos2 = − cos θ
2 2
Where we used the well known formula of the cosine of the double angle to simplify.
Finally, we can compute:

E(x̂, ŷ) = S1 (x̂, |↑↓ix̂ ) · S2 (ŷ, |↑↓ix̂ ) = +1 · (− cos θ) = −x̂ · ŷ

It was already mentioned that quantum mechanics predicts that E(x̂, ŷ) = −x̂ · ŷ,
and now we have proven that result. Now, after being able to make quantum mechanical
predictions about spin, we are ready to break Bell’s original inequality.
π π
If we take x̂ as a reference, and ŷ, ẑ vectors that are in angles 2 and 4 respectively to
it, then: π 
E(x̂, ŷ) = −x̂ · ŷ = − cos =0
2

π  √
2
E(x̂, ẑ) = −x̂ · ẑ = − cos =−
4 2

π √
π 2
E(ŷ, ẑ) = −ŷ · ẑ = − cos − =−
4 2 2
.
By substituting on the initial inequality:
√ √
2 2
1− ≥ 0 +
2 2


2
Since 2 ≈ 0.707 :
0.293 ≥ 0.707

Violating the inequality. Therefore we have founded the desired contradiction.

15
4 Loopholes in experimental tests of Bell’s inequality

It has been already shown that quantum mechanics is incompatible with physical theories
of local hidden variables, just as Bell’s theorem states. However, this does not imply that
nature doesn’t admit a theory of local hidden variables, since quantum mechanics is just
a candidate theory and its predictions about nature could be wrong in sufficiently critical
situations.
Since John Bell originally published his paper in 1964, “On the Einstein Podolski Rosen
paradox”, a lot of research concerning this apparent incompatibility has been done, and
one of the main questions stemming from it, which is still looking for a widely accepted
answer, is the following: is nature incompatible with local hidden variable theories?
Up to this time, quite a lot of experiments whose aim was to recreate the common
scenario used in the derivation of Bell’s inequalities have been performed, and most of
them have resulted in their violation, in agreement with quantum mechanical predictions.
However, these experimental tests of Bell’s inequalities must not be based only on the local
hidden variable model hypothesis, but also on other additional assumptions: the “freedom
of choice” that has already been explained as well as the “fair-sampling” assumption,
which has not been mentioned until now since it did not concern the theoretical approach
to Bell’s theorem that has been presented here.
The fair sampling assumption was stated in a paper published in 1969 by Clauser,
Horne, Shimony and Holt as: “if a pair of photons emerges from a polarizer, the proba-
bility of their joint detection is independent of the polarizer orientations.”. The essential
idea that lies behind this statement is that while an entire ensemble of particles might not
violate Bell’s inequality, it is conceivable under local hidden variable theories that certain
sub-ensembles of it do violate it; the fair sampling assumption means, in fact, assuming
that any sub-ensemble of particles that is detected will accurately represent the whole
ensemble of particles. This assumption becomes a real difficulty when performing experi-
ments with photons, since single-photon detectors are not very efficient and many photons
are easily absorbed or lost before being detected. Here, we use the word efficiency as the
probability that if a pair of photons emerges from a polarizer, they are jointly detected.
As a result, the efficiency will be 100% if the probability that a pair of emitted photons
from the source is detected is 1.
All of these assumptions are not easily applied to experimental setups, and this is the
reason why the empirical violations of Bell’s inequalities that have been observed until very
recently, admit explanations compatible with local hidden variable theories. Therefore,
they cannot be taken as the proof of the incompatibility of nature with them.
These problems with the experimental design are usually called loopholes. Below are
three of the main loopholes that experiments must close in order to result in valid conclu-
sions:

1. Locality or communication loophole: it is open if the setting choice or the


measurement result of one of the objects can be communicated to the second object
in time to influence the measurement that is performed there. How can this loophole
be closed? By simply creating a space-like separation of each local measurement on
one object from the setting choice and the measurement in the other one. And how is
this done experimentally? The measurement settings of each object must be chosen
so quickly that no physical signal (limited by the speed of light) can communicate
one object with the other, so that it is impossible that the chosen setting or the
measurement result of one of them affects the other.

16
2. Freedom-of-choice loophole: this is basically the “free variables” assumption
that has already been discussed. That is, the probability distribution ρ(λ) of the
possible states of the system λ must be independent of the experimental settings. It
can be experimentally closed if, assuming that λ is determined in the source where
the system is prepared before the measurement, the experimental settings are chosen
with a space-like separation from the creation of the system, so that the two events
cannot influence each other.

3. Fair-sampling or detection loophole: as has already been explained, this loop-


hole arises from the fact that it might happen that an experiment detects only a
sub-ensemble of the emitted particles and it is assumed to represent the entire en-
semble of particles. This loophole is closed by using Bell’s inequalities that do not
follow from the fair sampling assumption (such as Clauser and Horne (CH) inequal-
ity, derived in 1974, or Eberhard inequality which explicitly includes undetected
events, derived in 1993), and by using detectors of particles with adequate efficiency
(it has been also shown that each Bell’s inequality has a specific minimum efficiency
that the detectors used in the experiments must have, in order not to be violated by
the problems arising from the fair sampling assumption).

There have been a lot of experiments that have tried to close individual loopholes. For
instance:

• In 1982, Alain Aspect et al. was the first to use a fast switching of the measurement
settings in order to close the locality loophole.

• In 1998, Gregor Weihs et al. designed an improved experiment based on Aspect’s


setup using fast random switching.

• In 2001, Rowe et al. closed for the first time the fair-sampling loophole.

• In 2010, Thomas Scheidl et al. addressed both the freedom-of-choice loophole and
the locality loophole.

However, it was not until last year, in December 2015, in the paper “Significant
loophole-free test Bell’s theorem with entangled photons”, that these three loopholes were
simultaneously closed in a single experiment with high statistical significance, which con-
sisted basically of polarization-entangled photon pairs, high-efficiency detectors, and fast
random choices of the measurement settings which were space-like separated from the pho-
ton generation as well as from the detection of the other photon. It is a “loophole-free”
experiment in the sense that that other possible loopholes that were not taken into consid-
eration imply exotic consequences (such as superdeterminism) and closing them is usually
thought of as being an unattainable goal. Therefore, this experiment has been widely
accepted as the proof of nature being incompatible with local hidden variable theories.
It is worth mentioning that this year 2016, specifically on 30th November, an innovative
test of Bell’s inequality was performed: the project is called the BIG Bell Test (BBT) and
its special feature is that the measurement settings have been chosen by many people
around the world who have generated a unique sequence of 0 and 1 through a simple
browser game.

17
5 Bell’s theorem consequences

Bell theorem initially states:


No physical theory of local hidden variables can ever reproduce all of the predictions of
quantum mechanics.
As previously discussed, it is generally accepted for the scientific community that Bell
inequalities violations have been experimentally proved and that they are free of any free
variables wrong assumption. So this leads us to the following statement:
No physical theory of local hidden variables can reproduce nature.
Nowadays there is no consensual opinion on which is the solution this no-go theorem.
However we could simply the different positions by classifying on two categories:

1. Giving up on locality:
As previously explained locality is generally perceived as a necessary component to
any theory because it is considered that otherwise it would be contradictory to special
relativity. But as it has already been explained this may not be true, since some
theories offer faster than light communication and do not produce any paradoxes
when confronted to special relativity. Moreover, although some scientist as the very
same Bell refused to speak in these terms, quantum predictions, yet non-local with
a hidden variables model, does not offer faster than light communication between
two observers. So although not orthodox non-locality may be able to do not produce
paradoxes if mixed with relativity.

2. Giving up on hidden variables:


Hidden variable models offer a more classical prediction of reality. In other words,
what a closed box contains is always well defined: the cat is either dead or alive.
Giving up on those would force us to stop thinking of an independent reality. Re-
alism, a philosophical therm, would be no longer supported by physicians. Many
scientists as Bell or Einstein refused to do so. However this view supported by Bohr
offers a way out of both Bell theorem and the EPR paradox: variables which have
not been observed lose all sense. They are ill-defined and meaningless so no theorem
can be built on them.

18
6 Bibliography

• Bell, John S. (1964). “On the Einstein Podolsky Rosen paradox”.

• Bell, John S. (1980). “Bertlmann’s socks and the nature of reality”.

• Bohr, Niels (1935). “Quantum Mechanics and Physical Reality”.

• Einstein, A., Podolsky, B., Rosen, N. (1935). “Can Quantum-Mechanical Descrip-


tion of Physical Reality Be Considered Complete?”.

• Laudisa, Federico (2008). “Non-Local Realistic Theories and the Scope of the Bell
Theorem”.

• Aspect, Alain (2007). “Quantum mechanics: To be or not to be local”.

• Maccone, L. (2013). “A simple proof of Bell’s inequality”.

• Norsen, T. (2011). “John S. Bell’s concept of local causality”.

• Khrennikov, A., Ramelow, S., Ursin, R., Wittmann, B., Kofler, J., Basieva, I. (2014).
“On the equivalence of Clauser-Horne and Eberhard inequality based tests”.

• Giustina, M., Mech, A., Ramelow, S., Wittmann, B., Kofler, J., Beyer, J., Lita,
A., Calkins, B., Gerrits, T., Sae Woo Nam, Ursin, R., Zeilinger, A. (2012). “Bell
violation with entangled photons, free of the fair-sampling assumption”

• Giustina, M., Versteegh, M., Wengerwosky, S., Handsteiner, J., Hochrainer, A.,
Phelan K., Steinlechner, F., Kofler, J., Larsson, J., Abellán, C., Amaya, W., Pruneri,
V., Mithcell, M., Beyer, J., Gerrits, T., Lita, A., Shalm, L., Sae Woo Nam, Scheidl,
T., Ursin, R., Wittmann, B., Zeilinger, A. (2015). “Significant-loophole-free test of
Bell’s theorem with entangled photons”.

• http://www.unige.ch/gap/quantum/_media/publications:bib:gisin1991c.pdf

• http://scriptiesonline.uba.uva.nl/document/502639

• http://www.theory.caltech.edu/classes/ph125a/istmt.pdf

• https://plato.stanford.edu/entries/qt-epr/

19

Вам также может понравиться