Sie sind auf Seite 1von 35

Bell's Theorem

First published Wed Jul 21, 2004; substantive revision Sun Aug 22, 2004

Bell's Theorem is the collective name for a family of results, all showing the
impossibility of a Local Realistic interpretation of quantum mechanics. There are variants
of the Theorem with different meanings of “Local Realistic.” In John S. Bell's pioneering
paper of 1964 the realism consisted in postulating in addition to the quantum state a
“complete state”, which determines the results of measurements on the system, either by
assigning a value to the measured quantity that is revealed by the measurement regardless
of the details of the measurement procedure, or by enabling the system to elicit a definite
response whenever it is measured, but a response which may depend on the macroscopic
features of the experimental arrangement or even on the complete state of the system
together with that arrangement. Locality is a condition on composite systems with
spatially separated constituents, requiring an operator which is the product of operators
associated with the individual constituents to be assigned a value which is the product of
the values assigned to the factors, and requiring the value assigned to an operator
associated with an individual constitutent to be independent of what is measured on any
other constitutent. From his assumptions Bell proved an inequality (the prototype of
“Bell's Inequality”) which is violated by the Quantum Mechanical predictions made from
an entangled state of the composite system. In other variants the complete state assigns
probabilities to the possible results of measurements of the operators rather than
determining which result will be obtained, and nevertheless inequalities are derivable;
and still other variants dispense with inequalities. The incompatibility of Local Realistic
Theories with Quantum Mechanics permits adjudication by experiments, some of which
are described here. Most of the dozens of experiments performed so far have favored
Quantum Mechanics, but not decisively because of the “detection loophole” or the
“communication loophole.” The latter has been nearly decisively blocked by a recent
experiment and there is a good prospect for blocking the former. The refutation of the
family of Local Realistic Theories would imply that certain peculiarities of Quantum
Mechanics will remain part of our physical worldview: notably, the objective
indefiniteness of properties, the indeterminacy of measurement results, and the tension
between quantum nonlocality and the locality of Relativity Theory.

• 1. Introduction
• 2. Proof of a Theorem of Bell's Type
• 3. Experimental Tests of Bell's Inequalities
• 4. The Detection Loophole and its Remedy
• 5. The Communication Loophole and its Remedy
• 6. Variants of Bell's Theorem
• 7. Philosophical Comments
• Bibliography
• Other Internet Resources
• Related Entries
1. Introduction
In 1964 John S. Bell, a native of Northern Ireland and a staff member of CERN
(European Organisation for Nuclear Research) whose primary research concerned
theoretical high energy physics, published a paper in the short-lived journal Physics
which transformed the study of the foundation of Quantum Mechanics (Bell 1964). The
paper showed (under conditions which were relaxed in later work by Bell (1971, 1985,
1987) himself and by his followers (Clauser et al. 1969, Clauser and Horne 1974,
Mermin 1986, Aspect 1983)) that no physical theory which is realistic and also local in a
specified sense can agree with all of the statistical implications of Quantum Mechanics.
Many different versions and cases, with family resemblances, were inspired by the 1964
paper and are subsumed under the italicized statement, “Bell's Theorem” being the
collective name for the entire family.

One line of investigation in the prehistory of Bell's Theorem concerned the conjecture
that the Quantum Mechanical state of a system needs to be supplemented by further
“elements of reality” or “hidden variables” or “complete states” in order to provide a
complete description, the incompleteness of the quantum state being the explanation for
the statistical character of Quantum Mechanical predictions concerning the system. There
are actually two main classes of hidden-variables theories. In one, which is usually called
“non-contextual”, the complete state of the system determines the value of a quantity
(equivalently, an eigenvalue of the operator representing that quantity) that will be
obtained by any standard measuring procedure of that quantity, regardless of what other
quantities are simultaneously measured or what the complete state of the system and the
measuring apparatus may be. The hidden-variables theories of Kochen and Specker
(1967) are explicitly of this type. In the other, which is usually called “contextual”, the
value obtained depends upon what quantities are simultaneously measured and/or on the
details of the complete state of the measuring apparatus. This distinction was first
explicitly pointed out by Bell (1966) but without using the terms “contextual” and “non-
contextual”. There actually are two quite different versions of contextual hidden-
variables theories, depending upon the character of the context: an “algebraic context” is
one which specifies the quantities (or the operators representing them) which are
measured jointly with the quantity (or operator) of primary interest, whereas an
“environmental context” is a specification of the physical characteristics of the measuring
apparatus whereby it simultaneously measures several distinct co-measurable quantities.
In Bohm's hidden-variables theory (1952) the context is environmental, whereas in those
of Bell (1966) and Gudder (1970) the context is algebraic. A pioneering version of a
“hidden variables theory” was proposed by Louis de Broglie in 1926-7 (de Broglie 1927,
1928) and a more complete version by David Bohm in 1952 (Bohm 1952; see also the
entry on Bohmian mechanics). In these theories the entity supplementing the quantum
state (which is a wave function in the position representation) is typically a classical
entity, located in a classical phase space and therefore characterized by both position and
momentum variables. The classical dynamics of this entity is modified by a contribution
from the wave function

(1) ψ(x, t ) = R(x, t )exp[iS(x, t )/ ],


whose temporal evolution is governed by the Schrödinger Equation. Both de Broglie and
Bohm assert that the velocity of the particle satisfies the “guidance equation”

(2) v = grad(S/m),

whereby the wave function ψ acts upon the particles as a “guiding wave”.

De Broglie (1928) and the school of “Bohmian mechanics” (notably Dürr, Goldstein, and
Zanghì (1992)) postulate the guidance equation without an attempt to derive it from a
more fundamental principle. Bohm (1952), however, proposes a deeper justification of
the guidance equation. He postulates a modified version of Newton's second law of
motion:

(3) m d²x/dt ² = −grad[V(x,t ) + U(x,t )],

where V(x, t ) is the standard classical potential and U(x, t ) is a new entity, the “quantum
potential”,

(4) U(x,t ) = −( ²/2m) grad²R(x,t )/R(x, t ),

and he proves that if the guidance equation holds at an initial time t0, then it follows from
Eqs (2), (3), (4) and the time dependent Schrödinger Equation that it holds for all time.
Although Bohm deserves credit for attempting to justify the guidance equation, there is in
fact tension between that equation and the modified Newtonian equation (3), which has
been analyzed by Baublitz and Shimony (1996). Eq. (3) is a second order differential
equation in time and does not determine a definite solution for all t without two initial
conditions — x and v at t0.

Since v at t0 is a contingency, the validity of the guidance equation at t0 (and hence at all
other times) is contingent. Bohm recognizes this gap in his theory and discusses possible
solutions (Bohm 1952, p. 179), without reaching a definite proposal. If, however, this
difficulty is set aside, a solution is provided to the measurement problem of standard
quantum mechanics, i.e., the problem of accounting for the occurrence of a definite
outcome when the system of interest is prepared in a superposition of eigenstates of the
operator which is subjected to measurement. Furthermore, the guidance equation ensures
agreement with the statistical predictions of standard quantum mechanics. The hidden
variables model using the guidance equation inspired Bell to take seriously the hidden
variables interpretation of Quantum Mechanics, and the nonlocality of this model
suggested his theorem.

Another approach to the hidden variables conjecture has been to investigate the
consistency of the algebraic structure of the physical quantities characterized by Quantum
Mechanics with a hidden variables interpretation. Standard Quantum Mechanics assumes
that the “propositions” concerning a physical system are isomorphic to the lattice L(H) of
closed linear subspaces of a Hilbert space H (equivalently, to the lattice of projection
operators on H) with the following conditions: (1) the proposition whose truth value is
necessarily ‘true’ is matched with the entire space H; (2) the proposition whose truth
value is necessarily ‘false’ is matched with the empty subspace 0; (3) if a subspace S is
matched with a proposition q, then the orthogonal complement of S is matched with the
negation of q; (4) the proposition q, whose truth value is ‘true’ if the truth-value of either
q1 or q2 is ‘true’ and is ‘false’ if the answers to both q1 and q2 are ‘false’, is matched with
the closure of the set theoretical union of the spaces S1 and S2 respectively matched to q1
and q2 , the closure being the set of all vectors which can be expressed as the sum of a
vector in S1 and a vector in S2. It should be emphasized that this matching does not
presuppose that a proposition is necessarily either true or false and hence is compatible
with the quantum mechanical indefiniteness of a truth value, which in turn underlies the
feature of quantum mechanics that a physical quantity may be indefinite in value. The
type of hidden variables interpretation which has been most extensively treated in the
literature (often called a “non-contextual hidden variables interpretation” for a reason
which will soon be apparent) is a mapping m of the lattice L into the pair {1,0}, where
m(S)=1 intuitively means that the proposition matched with S is true and m(S)=0 means
intuitively that the proposition matched with S is false. A mathematical question of
importance is whether there exist such mappings for which these intuitive interpretations
are maintained and conditions (1) – (4) are satisfied. A negative answer to this question
for all L(H) where the Hilbert space H has dimensionality greater than 2 is implied by a
deep theorem of Gleason (1957) (which does more, by providing a complete catalogue of
possible probability functions on L(H)). The same negative answer is provided much
more simply by John Bell (1966) (but without the complete catalogue of probability
functions achieved by Gleason), who also provides a positive answer to the question in
the case of dimensionality 2 ; and independently these results were also achieved by
Kochen and Specker (1967). It should be added that in the case of dimensionality 2 the
statistical predictions of any quantum state can be recovered by an appropriate mixture of
the mappings m. (See also the entry on the Kochen-Specker theorem.)

In Bell (1966), after presenting a strong case against the hidden variables program
(except for the special case of dimensionality 2) Bell performs a dramatic reversal by
introducing a new type of hidden variables interpretation – one in which the truth value
which m assigns to a subspace S depends upon the context C of propositions measured in
tandem with the one associated with S. In the new type of hidden variables interpretation
the truth value into which m maps S depends upon the context C. These interpretations
are commonly referred to as “contextual hidden variables interpretations”, whereas those
in which there is no dependence upon the context are called “non-contextual.” Bell
proves the consistency of contextual hidden variables interpretations with the algebraic
structure of the lattice L(H) for two examples of H with dimension greater than 2. (His
proposal has been systematized by Gudder (1970), who takes a context C to be a maximal
Boolean subalgebra of the lattice L(H) of subspaces.)

Another line leading to Bell's Theorem was the investigation of Quantum Mechanically
entangled states, that is, quantum states of a composite system that cannot be expressed
as direct products of quantum states of the individual components. That Quantum
Mechanics admits of such entangled states was discovered by Erwin Schrödinger (1926)
in one of his pioneering papers, but the significance of this discovery was not emphasized
until the paper of Einstein, Podolsky, and Rosen (1935). They examined correlations
between the positions and the linear momenta of two well separated spinless particles and
concluded that in order to avoid an appeal to nonlocality these correlations could only be
explained by “elements of physical reality” in each particle — specifically, both definite
position and definite momentum — and since this description is richer than permitted by
the uncertainty principle of Quantum Mechanics their conclusion is effectively an
argument for a hidden variables interpretation. (It should be emphasized that their
argument does not depend upon counter-factual reasoning, that is reasoning about what
would be observed if a quantity were measured other than the one that was in fact
measured; instead their argument can be reformulated entirely in the ordinary inductive
logic, as emphasized by d'Espagnat (1976) and Shimony (2001). This reformulation is
important because it diminishes the force of Bohr's (1935) rebuttal of Einstein, Podolsky,
and Rosen on the grounds that one is not entitled to draw conclusions about the existence
of elements of physical reality from considerations of what would be seen if a
measurement other than the actual one had been performed. Bell was skeptical of Bohr's
rebuttal for other reasons, essentially that he regarded it to be anthropocentric.[1] See also
the entry on the Einstein-Podolsky-Rosen paradox.)

At the conclusion of Bell (1966), in which Bell gives a new lease on life to the hidden
variables program by introducing the notion of contextual hidden variables, he performs
another dramatic reversal by raising a question about a composite system consisting of
two well separated particles 1 and 2: suppose a proposition associated with a subspace S1
of the Hilbert space of particle 1 is subjected to measurement, and a contextual hidden
variables theory assigns the truth value m(S1/C) to this proposition. What physically
reasonable conditions can be imposed upon the context C? Bell suggests that C should
consist only of propositions concerning particle 1, for otherwise the outcome of the
measurement upon 1 will depend upon what operations are performed upon a remote
particle 2, and that would be non-locality. This condition raises the question: can the
statistical predictions of Quantum Mechanics concerning the entangled state be
duplicated by a contextual hidden variables theory in which the context C is localized? It
is interesting to note that this question also arises from a consideration of the de Broglie-
Bohm model: when Bohm derives the statistical predictions of a Quantum Mechanically
entangled system whose constituents are well separated, the outcome of a measurement
made on one constituent depends upon the action of the “guiding wave” upon
constituents that are far off, which in general will depend on the measurement
arrangement on that side. Bell was thus heuristically led to ask whether a lapse of locality
is necessary for the recovery of Quantum Mechanical statistics.

2. Proof of a Theorem of Bell's Type


Bell (1964) gives a pioneering proof of the theorem that bears his name, by first making
explicit a conceptual framework within which expectation values can be calculated for
the spin components of a pair of spin-half particles, then showing that, regardless of the
choices that are made for certain unspecified functions that occur in the framework, the
expectation values obey a certain inequality which has come to be called “Bell's
Inequality.” That term is now commonly used to denote collectively a family of
Inequalities derived in conceptual frameworks similar to but more general than the
original one of Bell. Sometimes these Inequalities are referred to as “Inequalities of Bell's
type.” Each of these conceptual frameworks incorporates some type of hidden variables
theory and obeys a locality assumption. The name “Local Realistic Theory” is also
appropriate and will be used throughout this article because of its generality. Bell
calculates the expectation values for certain products of the form (σ1·â)(σ2·ê), where σ1 is
the vectorial Pauli spin operator for particle 1 and σ2 is the vectorial Pauli spin operator
for particle 2 ( both particles having spin ½), and â and ê are unit vectors in three-space,
and then shows that these Quantum Mechanical expectation values violate Bell's
Inequality. This violation constitutes a special case of Bell's Theorem, stated in generic
form in Section 1, for it shows that no Local Realistic Theory subsumed under the
framework of Bell's 1964 paper can agree with all of the statistical predictions of
quantum theory.

In the present section the pattern of Bell's 1964 paper will be followed: formulation of a
framework, derivation of an Inequality, demonstration of a discrepancy between certain
quantum mechanical expectation values and this Inequality. However, a more general
conceptual framework than his will be assumed and a somewhat more general Inequality
will be derived, thus yielding a more general theorem than the one derived by Bell in
1964, but with the same strategy and in the same spirit. Papers which took the steps from
Bell's 1964 demonstration to the one given here are Clauser et al. (1969), Bell (1971),
Clauser and Horne (1974), Aspect (1983) and Mermin (1986).[2] Other strategies for
deriving Bell-type theorems will be mentioned in Section 6, but with less emphasis
because they have, at least so far, been less important for experimental tests.

The conceptual framework in which a Bell-type Inequality will be demonstrated first of


all postulates an ensemble of pairs of systems, the individual systems in each pair being
labeled as 1 and 2. Each pair of systems is characterized by a “complete state” m which
contains the entirety of the properties of the pair at the moment of generation. The
complete state m may differ from pair to pair, but the mode of generation of the pairs
establishes a probability distribution ρ which is independent of the adventures of each of
the two systems after they separate. Different experiments can be performed on each
system, those on 1 designated by a, a′, etc. and those on 2 by b, b′, etc. One can in
principle let a, a′, etc. also include characteristics of the apparatus used for the
measurement, but since the dependence of the result upon the microscopic features of the
apparatus is not determinable experimentally, only the macroscopic features of the
apparatus (such as orientations of the polarization analyzers) in their incompletely
controllable environment need be admitted in practice in the descriptions a, a′, etc. , and
likewise for b, b′, etc. The remarkably good agreement — which will be presented in
Section 3 — between the experimental measurements of correlations in Bell-type
experiments and the quantum mechanical predictions of these correlations justifies
restricting attention in practice to macroscopic features of the apparatus. The result of an
experiment on 1 is labeled by s, which can take on any of a discrete set of real numbers in
the interval [−1, 1]. Likewise the result of an experiment on 2 is labeled by t, which can
take on any of a discrete set of real numbers in [−1, 1]. (Bell's own version of his theorem
assumed that s and t are both bivalent, either −1 or 1, but other ranges are assumed in
other variants of the theorem.)

The following probabilities are assumed to be well defined:

(5) p1m (s |a, b, t ) = the probability that the outcome of the measurement performed on 1
is s when m is the complete state, the measurements performed on 1 and 2 respectively
are a and b, and the result of the experiment on 2 is t;

(6) p2m (t |a, b, s ) = the probability that the outcome of the measurement performed on 2
is t when the complete state is m, the measurements performed on 1 and 2 are
respectively a and b, and the result of the measurement a is s.

(7) pm (s, t |a, b ) = the probability that the results of the joint measurements a and b,
when the complete state is m, are respectively s and t.

The probability function p will be assumed to be non-negative and to sum to unity when
the summation is taken over all allowed values of s and t. (Note that the hidden variables
theories considered in Section 1 can be subsumed under this conceptual framework by
restricting the values of the probability function p to 1 and 0, the former being identified
with the truth value “true” and the latter to the truth value “false”. )

A further feature of the conceptual framework is locality, which is understood as the


conjunction of the following Independence Conditions:

Remote Outcome Independence (this name is a neologism, but an appropriate one, for
what is commonly called outcome independence)

(8a) p1m (s |a, b, t ) ≡ p1m (s |a, b ) is independent of t,

(8b) p2m (t |a, b, s ) ≡ p2m (t |a, b ) is independent of s;

(Note that Eqs. (8a) and (8b) do not preclude correlations of the results of the experiment
a on 1 and the experiment b on 2; they say rather that if the complete state m is given, the
outcome s of the experiment on 1 provides no additional information regarding the
outcome of the experiment on 2, and conversely.)

Remote Context Independence (this also is a neologism, but an appropriate one, for what
is commonly called parameter independence):

(9a) p1m (s |a, b ) ≡ p1m (s | a ) is independent of b,

(9b) p 2m (t |a, b ) ≡ p2m (t | b ) is independent of a.

Jarrett (1984) and Bell (1990) demonstrated the equivalence of the conjunction of (8a,b)
and (9a,b) to the factorization condition:
(10) pm (s, t |a, b ) = p1m(s |a ) p2m(t |b ),

and likewise for (a, b′), (a′, b ) , and (a′,b′) substituted for (a, b ).

The factorizability condition Eq. (10) is also often referred to as Bell locality. It should be
emphasized that at the present stage of exposition, however, Bell locality is merely a
mathematical condition within a conceptual framework, to which no physical
significance has been attached — in particular no connection to the locality of Special
Relativity Theory, although such a connection will be made later when experimental
applications of Bell's theorem will be discussed.

Bell's Inequality is derivable from his locality condition by means of a simple lemma:

(11) If q, q′, r , r′ all belong to the closed interval [−1,1], then S ≡ qr + qr′ + q′r − q′r′
belongs to the closed interval [-2,2].

Proof: Since S is linear in all four variables q, q′, r, r ′ it must take on its maximum and
minimum values at the corners of the domain of this quadruple of variables, that is, where
each of q, q′, r, r′ is +1 or −1. Hence at these corners S can only be an integer between -4
and +4. But S can be rewritten as (q + q′)(r + r′) – 2q′r′, and the two quantities in
parentheses can only be 0, 2, or -2, while the last term can only be -2 or +2, so that S
cannot equal +-3, +3, -4, or +4 at the corners. Q.E.D.

Now define the expectation value of the product st of outcomes:

(12) Em (a, b ) ≡ Σs Σt pm(s,t |a, b )(st ), the summation being taken over all the allowed
values of s and t.

and likewise with (a, b′). (a′, b ). and (a′, b′) substituted for (a, b ). Also take the
quantities q, q′, r, r′ of the above lemma (11) to be the single expectation values:

(13a) q = Σs sp1 m(s|a),

(13b) q′ = Σs sp1 m(s|a′),

(13c) r = Σt tp2 m(t |b),

(13d) r′ = Σt tp2 m(t |b′).

Then the lemma, together with Eq. (12), factorization condition Eq. (10), and the bounds
on s and t stated prior to Eq. (5), implies:

(14) -2 ≤ Em(a,b ) + Em(a,b′) + Em(a′,b ) − Em(a′,b′) ≤ 2.

Finally, return to the fact that the ensemble of interest consists of pairs of systems, each
of which is governed by a mapping m, but m is chosen stochastically from a space M of
mappings governed by a standard probability function ρ — that is, for every Borel subset
B of M, ρ(B ) is a non-negative real number, ρ(M ) = 1, and ρ(Uj Bj) = Σj ρ(Bj) where the
Bj's are disjoint Borel subsets of M and Uj Bj is the set-theoretical union of the Bj's. If we
define

(15a) pρ(s,t |a,b ) ≡ ∫M pm(s,t |a,b ) dρ

(15b) Eρ(a,b ) ≡ ∫M Em(a,b )dρ = Σs Σt ∫M pm(s,t |a, b )(st ) dρ

and likewise when (a, b′), (a′, b ), and (a′, b′) are substituted for (a, b ), then (14), (15a,b),
and the properties of ρ imply

(16) -2 ≤ E ρ(a,b ) + E ρ(a,b′) + E ρ(a′,b ) − Eρ(a′,b′) ≤ 2.

Ineq. (16) is an Inequality of Bell's type, henceforth called the “Bell-Clauser-Horne-


Shimony-Holt (BCHSH) Inequality.”

The third step in the derivation of a theorem of Bell's type is to exhibit a system, a
quantum mechanical state, and a set of quantities for which the statistical predictions
violate Inequality (16). Let the system consist of a pair of photons 1 and 2 propagating
respectively in the z and −z directions. The two kets |x>j and |y>j constitute a polarization
basis for photon j (j =1, 2), the former representing (in Dirac's notation) a state in which
the photon 1 is linearly polarized in the x-direction and the latter a state in which it is
linearly polarized in the y-direction. For the two-photon system the four product kets |x>1
|x>2, |x>1 |y>2, |y>1 |x>2, and |y>1 |y>2 constitute a polarization basis. Each two-photon
polarization state can be expressed as a linear combination of these four basis states with
complex coefficients. Of particular interest are the entangled quantum states, which in no
way can be expressed as |φ>1|ξ>2, with |φ> and |ξ> single-photon states, an example being

(17) | Φ > = (1/√2)[ |x>1 |x>2 + |y>1 |y>2 ],

which has the useful property of being invariant under rotation of the x and y axes in the
plane perpendicular to z. Neither photon 1 nor photon 2 is in a definite polarization state
when the pair is in the state |ψ>, but their potentialities (in the terminology of Heisenberg
1958) are correlated: if by measurement or some other process the potentiality of photon
1 to be polarized along the x-direction or along the y-direction is actualized, then the
same will be true of photon 2, and conversely. Suppose now that photons 1 and 2 impinge
respectively on the faces of birefringent crystal polarization analyzers I and II, with the
entrance face of each analyzer perpendicular to z. Each analyzer has the property of
separating light incident upon its face into two outgoing non-parallel rays, the ordinary
ray and the extraordinary ray. The transmission axis of the analyzer is a direction with
the property that a photon polarized along it will emerge in the ordinary ray (with
certainty if the crystals are assumed to be ideal), while a photon polarized in a direction
perpendicular to z and to the transmission axis will emerge in the extraordinary ray. See
Figure 1:
Figure 1
(reprinted with permission)

Photon pairs are emitted from the source, each pair quantum mechanically described by |
Φ> of Eq. (17), and by a complete state m if a Local Realistic Theory is assumed. I and II
are polarization analyzers, with outcome s=1 and t=1 designating emergence in the
ordinary ray, while s = −1 and t = −1 designate emergence in the extraordinary ray.

The crystals are also idealized by assuming that no incident photon is absorbed, but each
emerges in either the ordinary or the extraordinary ray. Quantum mechanics provides an
algorithm for computing the probabilities that photons 1 and 2 will emerge from these
idealized analyzers in specified rays, as functions of the orientations a and b of the
analyzers, a being the angle between the transmission axis of analyzer I and an arbitrary
fixed direction in the x-y plane, and b having the analogous meaning for analyzer II:

(18a) probΦ(s,t |a,b ) = | <Φ|θs>1 |φt>2 |2 .

Here s is a quantum number associated with the ray into which photon 1 emerges, +1
indicating emergence in the ordinary ray and −1 emergence in the extraordinary ray when
a is given, while t is the analogous quantum number for photon 2 when b is given; and |θs
>1 |φt >2 is the ket representing the quantum state of photons 1 and 2 with the respective
quantum numbers s and t. Calculation of the probabilities of interest from Eq. (18a) can
be simplified by using the invariance noted after Eq. (17) and rewriting |Φ > as

(19) |Φ> = (1/√2)[ |θ1>1 |θ1>2 + |θ−1>1 |θ−1>2 ].

Eq. (19) results from Eq. (17) by substituting the transmission axis of analyzer I for x and
the direction perpendicular to both z and this transmission axis for y.

Since |θ−1>1 is orthogonal to |θ1>1, only the first term of Eq. (19) contributes to the inner
product in Eq. (18a) if s=t=1; and since the inner product of | θ1 >1 with itself is unity
because of normalization, Eq. (18a) reduces for s = t = 1 to

(18b) probΦ(1,1|a,b ) = (½)| 2<θ1|φ1>2 |2.

Finally, the expression on the right hand side of Eq. (18b) is evaluated by using the law
of Malus, which is preserved in the quantum mechanical treatment of polarization states:
that the probability for a photon polarized in a direction n to pass through an ideal
polarization analyzer with axis of transmission n′ equals the squared cosine of the angle
between n and n′. Hence
(20a) probΦ(1,1|a,b ) = (½)cos2σ,

where σ is b-a. Likewise,

(20b) probΦ(−1,−1|a,b ) = (½) cos2σ, and

(20c) probΦ(1,−1|a,b ) = probΦ(−1,1|a,b ) = (½)sin2σ.

The expectation value of the product of the results s and t of the polarization analyses of
photons 1 and 2 by their respective analyzers is

(21) EΦ(a,b ) = probΦ(1,1|a,b ) + probΦ(−1,−1|a,b ) − probΦ(1,−1| a,b ) − probΦ(−1,1|a,b )


= cos2σ − sin2σ = cos2σ.

Now choose as the orientation angles of the transmission axes

(22) a = π/4, a′ = 0, b = π/8, b′ = 3 π/8 .

Then

(23a) EΦ(a,b ) = cos2(-π/8) = 0.707,

(23b) EΦ(a,b′) = cos2(π/8) = 0.707,

(23c) EΦ(α′,b ) = cos2(π/8) = 0.707,

(23d) EΦ(a′,b′) = cos2(3π/8) = - 0.707.

Therefore

(24) SΦ ≡ EΦ(a,b ) + EΦ(a,b′) + EΦ(a′,b ) − EΦ(a′,b′) = 2.828.

Eq. (24) shows that there are situations where the Quantum Mechanical calculations
violate the BCHSH Inequality, thereby completing the proof of a version of Bell's
Theorem. It is important to note, however, that all entangled quantum states yield
predictions in violation of Ineq. (16), as Gisin (1991) and Popescu and Rohrlich (1992)
have independently demonstrated. Popescu and Rohrlich (1992) also show that the
maximum amount of violation is achieved with a quantum state of maximum degree of
entanglement, exemplified by |Φ > of Eq. (17).

In Section 3 experimental tests of Bell's Inequality and their implications will be


discussed. At this point, however, it is important to discuss the significance of Bell's
Theorem from a purely theoretical standpoint. What Bell's Theorem shows is that
Quantum Mechanics has a structure that is incompatible with the conceptual framework
within which Bell's Inequality was demonstrated: a framework in which a composite
system with two subsystems 1 and 2 is described by a complete state assigning a
probability to each of the possible results of every joint experiment on 1 and 2, with the
probability functions satisfying the two Independence Conditions (8a,b) and (9a,b), and
furthermore allowing mixtures governed by arbitrary probability functions on the space
of complete states. An ‘experiment’ on a system can be understood to include the context
within which a physical property of the system is measured, but the two Independence
Conditions require the context to be local – that is, if a property of 1 is measured only
properties of 1 co-measurable with it can be part of its context, and similarly for a
property of 2. Therefore the incompatibility of Quantum Mechanics with this conceptual
framework does not preclude the contextual hidden variables models proposed by Bell in
(1966), an example of which is the de Broglie-Bohm model, but it does preclude models
in which the contexts are required to be local. The most striking implication of Bell's
Theorem is the light that it throws upon the EPR argument. That argument examines an
entangled quantum state and shows that a necessary condition for avoiding action-at-a-
distance between measurement outcomes of correlated properties of the two subsystems
— e.g., position in both or linear momentum in both — is the ascription of “elements of
physical reality” corresponding to the correlated properties to each subsystem without
reference to the other. Bell's Theorem shows that such an ascription will have statistical
implications in disagreement with those of quantum mechanics. A penetrating feature of
Bell's analysis, when compared with that of EPR, is his examination of different
properties in the two subsystems, such as linear polarizations along different directions in
1 and 2, rather than restricting his attention to correlations of identical properties in the
two subsystems.

3. Experimental Tests of Bell's Inequalities


When Bell published his pioneering paper in 1964 he did not urge an experimental
resolution of the conflict between Quantum Mechanics and Local Realistic Theories,
probably because the former had been confirmed often and precisely in many branches of
physics.

It was doubtful, however, that any of these many confirmations occurred in situations of
conflict between Quantum Mechanics and Local Realistic Theories, and therefore a
reliable experimental adjudication was desirable. A proposal to this effect was made by
Clauser, Horne, Shimony, and Holt (1969), henceforth CHSH, who suggested that the
pairs 1 and 2 be the photons produced in an atomic cascade from an initial atomic state
with total angular momentum J = 0 to an intermediate atomic state with J = 1 to a final
atomic state J = 0, as in an experiment performed with calcium vapor for other purposes
by Kocher and Commins (1967). This arrangement has several advantages: first, by
conservation of angular momentum the photon pair emitted in the cascade has total
angular momentum 0, and if the photons are collected in cones of small aperture along
the z and -z directions the total orbital angular momentum is small, with the consequence
that the total spin (or polarization) angular momentum is close to 0 and therefore the
polarizations of the two photons are tightly correlated; second, the photons are in the
visible frequency range and hence susceptible to quite accurate polarization analysis with
standard polarization analyzers; and third, the stochastic interval between the time of
emission of the first photon and the time of emission of the second is in the range of 10
nsec., which is small compared to the average time between two productions of pairs, and
therefore the associated photons 1 and 2 of a pair are almost unequivocally matched. A
disadvantage of this arrangement, however, is that photo-detectors in the relevant
frequency range are not very efficient — less than 20% efficency for single photons and
hence less that 4% efficient for detection of a pair — and therefore an auxiliary
assumption is needed in order to make inferences from the statistics of the subensemble
of the pairs that is counted to the entire ensemble of pairs emitted during the period of
observation. (This disadvantage causes a “detection loophole” which prevents the
experiment, and others like it, from being decisive, but procedures for blocking this
loophole are at present being investigated actively and will be discussed in Section 4).

In the experiment proposed by CHSH the measurements are polarization analyses with
the transmission axis of analyzer I oriented at angles a and a′, and the transmission axis
of analyzer II oriented at angles b and b′. The results s = 1 and s = −1 respectively
designate passage and non-passage of photon 1 through analyzer I, and t = 1 and t = −1
respectively designate passage and non-passage of photon 2 through analyzer II. Non-
passage through the analyzer is thus substituted for passage into the extraordinary ray.
This simplification of the apparatus causes an obvious problem: that it is impossible to
discriminate directly between a photon that fails to pass through the analyzer and one
which does pass through the analyzer but is not detected because of the inefficiency of
the photo-detectors. CHSH dealt with this problem in two steps. First, they expressed the
probabilities pρ(s,t |a, b ), where either s or t (or both) is −1 as follows:

(25a) pρ(1,−1|a,b ) = pρ(1,1|a,∞) − pρ(1,1|a,b ),

where ∞ replacing b designates the removal of the analyzer from the path of 2. Likewise

(25b) pρ(−1,1|a,b ) = pρ(1,1|∞,b ) − pρ(1,1|a,b ),

where analyzer II is oriented at angle b and ∞ replacing a designates removal of the


analyzer I from the path of photon 1; and finally

(25c) pρ(−1,−1|a,b ) = 1 − pρ(1,1|a,b ) − pρ(1,−1|a,b ) − pρ(−1,1|a,b ) = 1 − pρ(1,1|a, ∞) −


pρ(1,1| ∞,b ) + pρ(1,1|a,b ),

Their second step is to make the “fair sampling assumption”: given that a pair of photons
enters the pair of rays associated with passage through the polarization analyzers, the
probability of their joint detection is independent of the orientation of the analyzers
(including the quasi-orientation ∞ which designates removal). With this assumption,
together with the assumptions that the local realistic expressions pρ(s,t |a, b ) correctly
evaluate the probabilities of the results of polarization analyses of the photon pair (1,2),
we can express these probabilities in terms of detection rates:

(26a) pρ(1,1|a,b ) = D(a,b )/D0


where D(a,b ) is the counting rate of pairs when the transmission axes of analyzers I and
II are oriented respectively at angles a and b, and D0 is the detection rate when both
analyzers are removed from the paths of photons 1 and 2;

(26b) pρ(1,1|a,∞) = D1(a,∞)/D0,

where D1(a,∞) is the counting rate of pairs when analyzer I is oriented at angle a while
analyzer II is removed; and

(26c) pρ(1,1|∞,b ) = D2(∞,b )/D0,

where D2(∞,b ) is the counting rate when analyzer II is oriented at angle b while analyzer
I is removed. When Ineq. (16) is combined with Eq. (15b) and with Eqs (26a,b,c) relating
probabilities to detection rates, the result is an inequality governing detection rates,

(27) −1 ≤ D(a,b )/D0 + D(a,b′)/D0 + D(a′,b )/D0 − D(a′,b′)/D0 − [D1(a)/D 0 + D2(b )/D0 ] ≤


0.

(In (27), D1(a) is an abbreviation for D(a,∞), and D2(b) is an abbreviation for D(∞,b).)

If the following symmetry conditions are satisfied by the experiment:

(28) D(a,b ) = D(|b−a |),

(29) D1(a ) = D1, independent of a;

(30) D2(b ) = D2, independent of b,

one can rewrite Ineq. (27) as

(31) −1 ≤ D(|b-a |)/D0 + D(|b′−a|)/D0 + D(|b-a′|)/D0 − D(|b′-a′|)/D0 − [D1/D0 + D2/D0] ≤ 0.

The first experimental test of Bell's Inequality, performed by Freedman and Clauser
(1972), proceeded by making two applications of Ineq. (31), one to the angles a = π/4, a′
= 0, b = π/8, b′ = 3π/8, yielding

(32a) −1 ≤ [3D(π/8)/D0 − D(3π/8)/D0 ] − D1/D0 − D2/D0 ≤ 0,

and another to the angles a = 3π/4, a′ = 0, b = 3π/8, b′ = 9π/8, yielding

(32b) −1 ≤ [3D(3π/8)/D0 − D(π/8)/D0 ] − D1/D0 − D2/D0 ≤ 0.

Combining Ineq. (32a) and Ineq. (32b) yields

(33) -(¼) ≤ [D(π/8 )/D0 − D(3π/8)/D0 ] ≤ ¼ .


The Quantum Mechanical prediction for this arrangement, taking into account the
uncertainties about the polarization analyzers and the angle from the source subtended by
the analyzers, is

(34) [D(π/8)/D0 − D(3π/8)/D0 ]QM = (0.401+/-0.005) − (0.100+/-0.005) = 0.301+/-0.007,

The experimental result obtained by Freedman and Clauser was

(35) [D(π/8)/D0 − D(3π/8)/D0 ]expt = 0.300 +/- 0.008,

which is 6.5 sd from the limit allowed by Ineq. (33) but in good agreement with the
quantum mechanical prediction Eq. (34). This was a difficult experiment, requiring 200
hours of running time, much longer than in most later tests of Bell's Inequality, which
were able to use lasers for exciting the sources of photon pairs.

Several dozen experiments have been performed to test Bell's Inequalities. References
will now be given to some of the most noteworthy of these, along with references to
survey articles which provide information about others. A discussion of those actual or
proposed experiments which were designed to close two serious loopholes in the early
Bell experiments, the “detection loophole” and the “communication loophole”, will be
reserved for Section 4 and Section 5.

Holt and Pipkin completed in 1973 (Holt 1973) an experiment very much like that of
Freedman and Clauser, but examining photon pairs produced in the 91P1 →73S1→63P 0
cascade in the zero nuclear-spin isotope of mercury-198 after using electron
bombardment to pump the atoms to the first state in this cascade. The result was contrary
to that of Freedman and Clauser: fairly good agreement with Ineq. (33), which is a
consequence of the BCHSH Inequality, and a disagreement with the quantum mechanical
prediction by nearly 4 sd. Because of the discrepancy between these two early
experiments Clauser (1976) repeated the Holt-Pipkin experiment, using the same cascade
and excitation method but a different spin-0 isotope of mercury, and his results agreed
well with the quantum mechanical predictions but violated the consequence of Bell's
Inequality. Clauser also suggested a possible explanation for the anomalous result of
Holt-Pipkin: that the glass of the Pyrex bulb containing the mercury vapor was under
stress and hence was optically active, thereby giving rise to erroneous determinations of
the polarizations of the cascade photons.

Fry and Thompson (1976) also performed a variant of the Holt-Pipkin experiment, using
a different isotope of mercury and a different cascade and exciting the atoms by radiation
from a narrow-bandwith tunable dye laser. Their results also agreed well with the
quantum mechanical predictions and disagreed sharply with Bell's Inequality. They
gathered data in only 80 minutes, as a result of the high excitation rate achieved by the
laser.

Four experiments in the 1970s — by Kasday-Ullman-Wu, Faraci-Gutkowski-Notarigo-


Pennisi, Wilson-Lowe-Butt, and Bruno-d’Agostino-Maroni used photon pairs produced
in positronium annihilation instead of cascade photons. Of these, all but that of Faraci et
al. gave results in good agreement with the quantum mechanical predictions and in
disagreement with Bell's Inequalities. A discussion of these experiments is given in the
review article by Clauser and Shimony (1978), who regard them as less convincing than
those using cascade photons, because they rely upon stronger auxiliary assumptions.

The first experiment using polarization analyzers with two exit channels, thus realizing
the theoretical scheme in the third step of the argument for Bell's Theorem in Section 2,
was performed in the early 1980s with cascade photons from laser-excited calcium atoms
by Aspect, Grangier, and Roger (1982). The outcome confirmed the predictions of
quantum mechanics over those of local realistic theories more dramatically than any of its
predecessors — with the experimental result deviating from the upper limit in a Bell's
Inequality by 40 sd. An experiment soon afterwards by Aspect, Dalibard, and Roger
(1982), which aimed at closing the communication loophole, will be discussed in Section
5. The historical article by Aspect (1992) reviews these experiments and also surveys
experiments performed by Shih and Alley, by Ou and Mandel, by Rarity and Tapster, and
by others, using photon pairs with correlated linear momenta produced by down-
conversion in non-linear crystals. Some even more recent Bell tests are reported in an
article on experiments and the foundations of quantum physics by Zeilinger (1999).

Pairs of photons have been the most common physical systems in Bell tests because they
are relatively easy to produce and analyze, but there have been experiments using other
systems. Lamehi-Rachti and Mittig (1976) measured spin correlations in proton pairs
prepared by low-energy scattering. Their results agreed well with the quantum
mechanical prediction and violated Bell's Inequality, but strong auxiliary assumptions
had to be made like those in the positronium annihilation experiments. In 2003 a Bell test
was performed at CERN by A. Go (Go 2003) with B-mesons, and again the results
favored the quantum mechanical predictions.

The outcomes of the Bell tests provide dramatic confirmations of the prima facie
entanglement of many quantum states of systems consisting of 2 or more constituents,
and hence of the existence of holism in physics at a fundamental level. Actually, the first
confirmation of entanglement and holism antedated Bell's work, since Bohm and
Aharonov (1957) demonstrated that the results of Wu and Shaknov (1950), Compton
scattering of the photon pairs produced in positronium annihilation, already showed the
entanglement of the photon pairs.

4. The Detection Loophole and its Remedy


The summary in Section 3 of the pioneering experiment by Freedman and Clauser noted
that their symmetry assumptions, Eqs. (28), (29), and (30), are susceptible to
experimental check, and furthermore could have been dispensed with by measuring the
detection rates with additional orientations of the analyzers. The fair sampling
assumption, on the other hand, is essential in all the optical Bell tests performed so far for
linking the results of polarization or direction analysis, which are not directly observable,
with detection rates, which are observable. The absence of an experimental confirmation
of the fair sampling assumption, together with the difficulty of testing Bell's Inequality
without this assumption or another one equally remote from confirmation is known as the
“detection loophole” in the refutation of Local Realistic Theories, and is the source of
skepticism about the definitiveness of the experiments.

The seriousness of the detection loophole was increased by a model of Clauser and Horne
(CH) (1974), in which the rates at which the photon pairs pass through the polarization
analyzers with various orientations are consistent with an Inequality of Bell's type, but
the hidden variables provide instructions to the photons and the apparatus not only
regarding passage through the analyzers but also regarding detection, thereby violating
the fair sampling assumption. Detection or non-detection is selective in the model in such
a way that the detection rates violate the Bell-type Inequality and agree with the quantum
mechanical predictions. Other models were constructed later by Fine (1982a) and
corrected by Maudlin (1994) (the “Prism Model”) and by C.H.Thompson (1996) (the
“Chaotic Ball model”). Although all these models are ad hoc and lack physical
plausibility, they constitute existence proofs that Local Realistic Theories can be
consistent with the quantum mechanical predictions provided that the detectors are
properly selective. The detection loophole could in principle be blocked by a test of the
BCHSH Inequality provided that the detectors for the 1 and 2 particles were sufficiently
efficient and that there were a reliable way of determining how many pairs impinge upon
the analyzer-detector assemblies. The first of these conditions can very likely be fulfilled
if atoms from the dissociation of dimers are used as the particle pairs, as in the proposed
experiment of Fry and Walther (Fry & Walther 1997, 2002, Walther & Fry 1997), to be
discussed below, but the second condition does not at present seem feasible.
Consequently a different strategy is needed.

Tools which are promising for blocking the detection loophole are two Inequalities
derived by CH (Clauser & Horne 1974), henceforth called BCH Inequalities. Both differ
from the BCHSH Inequality of Section 2 by involving ratios of probabilities. The two
BCH Inequalities differ from each other in two respects: the first involves not only the
probabilities of coincident counting of the 1 and 2 systems but also probabilities of single
counting of 1 and 2 without reference to the other, and it makes no auxiliary assumption
regarding the dependence of detection upon the placement of the analyzer; the second
involves only probabilities of coincident detection and uses a weak, but still non-trivial,
auxiliary assumption called “no enhancement.”

The experiment proposed by Fry and Walther intends to make use of the first BCH
Inequality, dispensing with auxiliary assumptions, but the second BCH Inequality will
also be reviewed here, because of its indispensability in case the optimism of Fry and
Walther regarding the achievability of certain desirable experimental conditions turns out
to be disappointed.

The conceptual framework of both BCH Inequalities takes an analyzer/detector assembly


as a unit, instead of considering the separate operation of two parts, and it places only one
analyzer/detector assembly in the 1 arm and one in the 2 arm of the experiment; in other
words, for both analyzers I and II the two exit channels are detection and non-detection.
Let N be the number of pairs of systems impinging on the two analyzer/detector
assemblies, N1(a) the number detected by 1's analyzer/detector assembly, N2(b) the
number detected by 2's analyzer/detector assembly, and N12(a,b ) the number of pairs
detected by both, where a and b are the respective settings of the analyzers. Let N1 (m,a ),
N2(m,b ), and N12(m,a,b ) be the corresponding quantities predicted by the Local Realistic
Theory with complete state m. Then the respective probabilities of detection by the
analyzer/detector assemblies for 1 and 2 separately and by both together predicted by the
local realistic theory with complete state m are

(36a) p1(m,a ) = N1(m,a )/N,

(36b) p2(m,b ) = N2(m,b )/N

(36c) p(m,a,b ) = N12(m,a,b )/N.

Note that the total number N of pairs appears in the denominators on the right hand side,
but the difficulty of determining this quantity experimentally is circumvented by CH,
who derive an inequality concerning the ratios of the probabilities in Eqs. (36a,b,c), so
that the denominator N cancels. The analogues of the Independence Conditions of (8a,b)
and (9a,b) are taken as part of the conceptual framework of a Local Realistic Theory,
which implies the factorization of p(m,a,b ):

(37) p(m,a,b ) = p1(m,a )p2( m,b ).

CH then prove a lemma similar to the lemma in Section 2 : if q,q′ are real numbers such
that q and q′ fall in the closed interval [0,X], and r,r′,are real numbers such that r and r′
fall in the closed interval [0,Y], then

(38) −1 ≤ qr + qr′ + q′r − q′r′ - qY − rX ≤ 0.

Then making the substitution

(39) q = p1(m,a ), q′ = p1(m,a′), r = p2(m,b ), r′ = p2 (m,b′),

and using Eq. (37) we have

(40) −1 ≤ p(m,a,b ) + p(m,a,b′) + p(m,a′,b ) - p(m,a′,b′) − p1(m,a )Y − p2(m,b )X ≤ 0.

When the Local Realistic Theory describes an ensemble of pairs by a probability


distribution ρ over the space of complete states M then the probabilities corresponding to
those of (36a,b,c) are

(41a) pρ1(a) = ∫M p1(m,a )dρ,


(41b) pρ2(b ) = ∫M p2(m,b )dρ,
(41c) pρ(a,b ) = ∫M p(m,a,b )dρ,
and when X and Y are taken to be 1, as they certainly can be from general properties of
probability distributions, then

(42) −1 ≤ pρ(a,b ) + pρ (a,b′) + pρ(a′,b ) - pρ(a′,b′) - pρ1(a ) - pρ2(b ) ≤ 0 ,

Since CH are seeking an Inequality involving only the ratios of probabilities they
disregard the lower limit and rewrite the right hand Inequality as

(43) [pρ(a,b ) + pρ(a,b′) + pρ(a′,b ) - pρ(a′,b′)]/[p ρ1(a ) + pρ2(b )] ≤ 1.

This is the first BCH Inequality. In principle this Inequality could be used to adjudicate
between the family of Local Realistic Theories and Quantum Mechanics, provided that
the detectors are sufficiently efficient and also provided that the single detection counts
are not spoiled by counting systems that do not belong to the pairs in the ensemble of
interest. In experiments using photon pairs from a cascade, as most of the early Bell tests
were, it can happen that the second transition occurs without the first step in the cascade,
thus producing a single photon without a partner. To cope with this difficulty CH make
the “no enhancement assumption”, which is considerably weaker than the “fair sampling”
assumption used in Section 3 : that if an analyzer is removed from the path of either 1 or
2 — an operation designated symbolically by letting ∞ replace the parameter a or b of
the respective analyzer — the resulting probability of detection is at least as great as
when a finite parameter is used, i.e.,

(44a) p1(m,a ) ≤ p1(m,∞),

(44b) p2(m,b ) ≤ p2(m,∞)

Now let the right hand side of (44a) be X and the right hand side of (44b) be Y and insert
these values into Inequality (40), and furthermore use Eq. (37) twice to obtain

(45) −1 ≤ p(m,a,b ) + p(m,a,b′) + p(m,a′,b ) − p(m,a′,b′) − p1(m,a,∞) − p2(m,∞,b ) ≤ 0.

Integrate Inequality (45) using the distribution ρ and then retain only the right hand
Inequality in order to obtain an expression involving ratios only of probabilities of joint
detections:

(46) [pρ(a,b ) + pρ(a,b′) + pρ(a′,b ) − pρ(a′,b′)] / [pρ(a,∞) + p ρ(∞,b )] ≤ 1

This is the second BCH Inequality.

In the experiment initiated by Fry and Walther (1997), but not yet complete, dimers of
199
Hg are generated in a supersonic jet expansion and then photo-dissociated by two
photons from appropriate laser beams. Each 199Hg atom has nuclear spin ½, and the total
spin F (electronic plus nuclear) of the dimer is 0 because of the symmetry rules for the
total wave function of a homonuclear diatomic molecule consisting of two fermions. (The
argument for this conclusion is fairly intricate but well presented in Walther and Fry
1997). Because the dissociation is a two-body process, momentum conservation
guarantees that the directions of the two atoms after dissociation are strictly correlated, so
that when the two analyzer/detector assemblies are optimally placed the entrance of one
199
Hg atom into its analyzer/detector will almost certainly be accompanied by the
entrance of the other atom into its assembly; the primary reason for the occasional failure
of this coordination is the non-zero volume of the region in which the dissociation occurs.
Since each of the 199Hg atoms (with 80 electrons) is in the electronic ground state it will
have zero electronic spin, and therefore the total spin F of each atom (which then equals
its nuclear spin ½) will be F = ½. Given any choice of an axis, the only possible values of
the magnetic quantum number FM relative to this axis are FM=½ and -½. The directions of
this axis, θ1 for atom 1 and θ2 for atom 2, are the quantities to be used for the parameters
a and b in the BCH Inequalities. The angles θ1 and θ2 are physically fixed by the
directions of two left-circularly polarized 253.7 nm laser beams propagating in parallel
planes each perpendicular to the plane in which the atoms 1 and 2 travel from the
dissociation region. See Figure 2:

Figure 2
(reprinted with permission)

This is a schematic of the experiment showing the direction of the mercury dimer beam,
together with a pair of the dissociated atoms and their respective detection planes. The
relative directions of the various laser beams are also shown.

The left-circular polarization ensures that if a photon is absorbed by atom 1 then FM of


atom 1 will decrease by one, which is possible only if FM is initially ½, and likewise for
absorption by atom 2. Thus each laser beam is selective and acts as an analyzer by taking
only an FM=½ atom, 1 or 2 as the case may be, into a specific excited state. Detection is
achieved in several steps. The first step is the impingement on atom 1 of a 197.3 nm laser
beam, timed to arrive within the lifetime of the excited state produced by the 253.7 nm
laser beam; absorption of a photon from this beam by 1 will cause ionization (and
likewise an excited 2 is ionized). The second step is the detection of either the resulting
ion or the associated photoelectron or both — the detection being a highly efficient
process, since both of these products of ionization are charged. The first BCH Inequality
predicts that the coincident detection rates and the single detection rates in this
experiment satisfy
(47) [ D(θ1, θ2) + D(θ1, θ′2) + D(θ′1, θ2) - D(θ′1, θ′2)] / [D1(θ1) + D2(θ2)] ≤ 1.

Fry and Walther calculate that with proper choices of the four angles, large enough
values of detector efficiencies, large enough probability of 1 entering the region of
analysis, large enough probability of 2 entering its region of analysis conditional upon 1
doing so, and small enough probability of mistaken analysis (e.g., mistaking an FM = -½
for an FM = ½ because of rare processes) the quantum mechanical predictions will violate
(47). If these predictions are fulfilled, no auxiliary assumption like “no enhancement”
will be needed to disprove Inequality (43) experimentally. The detection loophole in the
refutation of the family of Local Realistic Theories will thereby be closed. Fry and
Walther express some warning, however, against excessive optimism about detecting a
sufficiently large percentage of the ions and the electrons produced by the ionization of
the Hg atoms, together with a low rate of errors due to background or noise counts: “the
Hg partial pressure, as well as the partial pressure of all other residual gases, must be kept
as low as possible. An ultra high vacuum of better than 10-9 Torr is required and the
detector must be cooled to liquid nitrogen temperatures to freeze out background Hg
atoms… Equally important is the suppression of photoelectrons produced by scattered
photons.” (Fry & Walther 1997, p. 67). If these desiderata are not achieved, it would be
necessary to resort to BCH's second Inequality, which required the ”no enhancement”
assumption.

5. The Communication Loophole and its Remedy


The derivations of all the variants of Bell's Inequality depend upon the two Independence
Conditions (8a,b) and (9a,b). Experimental data that disagree with a Bell's Inequality are
not a refutation unless these Conditions are satisfied by the experimental arrangement. In
the early tests of Bell's Inequalities it was plausible that these Conditions were satisfied
just because the 1 and the 2 arms of experiment were spatially well separated in the
laboratory frame of reference. This satisfaction, however, is a mere contingency not
guaranteed by any law of physics, and hence it is physically possible that the setting of
the analyzer of 1 and its detection or non-detection could influence the outcome of
analysis and the detection or non-detection of 2, and conversely. This is the
“communication loophole” in the early Bell tests. If the process of analysis and detection
of 1 were an event with space-like separation from the event consisting of the analysis
and detection of 2 then the satisfaction of the Independence Conditions would be a
consequence of the Special Theory of Relativity, according to which no causal influences
can propagate with a velocity greater than the velocity of light in vacuo. Several
experiments of increasing sophistication between 1982 and the present have attempted to
block the Communication Loophole in this way.

Aspect, Dalibard, and Roger (1982) published the results of an experiment in which the
choices of the orientations of the analyzers of photons 1 and 2 were performed so rapidly
that they were events with space-like separation. No physical modification was made of
the analyzers themselves. Instead, switches consisting of vials of water in which standing
waves were excited ultrasonically were placed in the paths of the photons 1 and 2. When
the wave is switched off, the photon propagates in the zeroth order of diffraction to
polarization analyzers respectively oriented at angles a and b, and when it is switched on
the photons propagate in the first order of diffraction to polarization analyzers
respectively oriented at angles a′ and b′. The complete choices of orientation require time
intervals 6.7 ns and 13.37 ns respectively, much smaller than the 43 ns required for a
signal to travel between the switches in obedience to Special Relativity Theory. Prima
facie it is reasonable that the Independence Conditions are satisfied, and therefore that the
coincidence counting rates agreeing with the quantum mechanical predictions constitute a
refutation of Bell's Inequality and hence of the family of Local Realistic Theories. There
are, however, several imperfections in the experiment. First of all, the choices of
orientations of the analyzers are not random, but are governed by quasiperiodic
establishment and removal of the standing acoustical waves in each switch. A scenario
can be invented according to which the clever hidden variables of each analyzer can
inductively infer the choice made by the switch controlling the other analyzer and adjust
accordingly its decision to transmit or to block an incident photon. Also coincident count
technology is employed for detecting joint transmission of 1 and 2 through their
respective analyzers, and this technology establishes an electronic link which could
influence detection rates. And because of the finite size of the apertures of the switches
there is a spread of the angles of incidence about the Bragg angles, resulting in a loss of
control of the directions of a non-negligible percentage of the outgoing photons.

The experiment of Tittel, Brendel, Zbinden, and Gisin (1998) did not directly address the
communication loophole but threw some light indirectly on this question and also
provided the most dramatic evidence so far concerning the maintenance of entanglement
between particles of a pair that are well separated. Pairs of photons were generated in
Geneva and transmitted via cables with very small probability per unit length of losing
the photons to two analyzing stations in suburbs of Geneva, located 10.9 kilometers apart
on a great circle. The counting rates agreed well with the predictions of Quantum
Mechanics and violated one of Bell's Inequalities. No precautions were taken to ensure
that the choices of orientations of the two analyzers were events with space-like
separation. The great distance between the two analyzing stations makes it difficult to
conceive a plausible scenario for a conspiracy that would violate Bell's Independence
Conditions. Furthermore — and this is the feature which seems most to have captured the
imagination of physicists — this experiment achieved much greater separation of the
analyzers than ever before, thereby providing the best reply to date to a conjecture by
Schrödinger (1935) that entanglement is a property that may dwindle with spatial
separation.

The experiment that comes closest so far to closing the Communication Loophole is that
of Weihs, Jennenwein, Simon, Weinfurter, and Zeilinger (1998). The pairs of systems
used to test a Bell's Inequality are photon pairs in the entangled polarization state

(48) |Ψ> = 1/√2 (|H>1 |V>2 − |V>1 |H>2),

where the ket |H> represents horizontal polarization and |V> represents vertical
polarization. Each photon pair is produced from a photon of a laser beam by the down-
conversion process in a nonlinear crystal. The momenta, and therefore the directions, of
the daughter photons are strictly correlated, which ensures that a non-negligible
proportion of the pairs jointly enter the apertures (very small) of two optical fibers, as
was also achieved in the experiment of Tittel et al. The two stations to which the photon
pairs are delivered are 400 m apart, a distance which light in vacuo traverses in 1.3 μs.
Each photon emerging from an optical fiber enters a fixed two-channel polarizer (i.e., its
exit channels are the ordinary ray and the extraordinary ray). Upstream from each
polarizer is an electro-optic modulator, which causes a rotation of the polarization of a
traversing photon by an angle proportional to the voltage applied to the modulator. Each
modulator is controlled by amplification from a very rapid generator, which randomly
causes one of two rotations of the polarization of the traversing photon. An essential
feature of the experimental arrangement is that the generators applied to photons 1 and 2
are electronically independent. The rotations of the polarizations of 1 and 2 are
effectively the same as randomly and rapidly rotating the polarizer entered by 1 between
two possible orientations a and a′ and the polarizer entered by 2 between two possible
orientations b and b′. The output from each of the two exit channels of each polarizer
goes to a separate detector, and a “time tag” is attached to each detected photon by means
of an atomic clock. Coincidence counting is done after all the detections are collected by
comparing the time tags and retaining for the experimental statistics only those pairs
whose tags are sufficiently close to each other to indicate a common origin in a single
down-conversion process. Accidental coincidences will also enter, but these are
calculated to be relatively infrequent. This procedure of coincidence counting eliminates
the electronic connection between the detector of 1 and the detector of 2 while detection
is taking place, which conceivably could cause an error-generating transfer of
information between the two stations. The total time for all the electronic and optical
processes in the path of each photon, including the random generator, the electro-optic
modulator, and the detector, is conservatively calculated to be smaller than 100 ns, which
is much less than the 1.3 μs required for a light signal between the two stations. With the
choice made in Eq. (22) of the angles a, a′, b, and b′ and with imperfections in the
detectors taken into account, the Quantum Mechanical prediction is

(49) Sψ ≡ Eψ(a,b ) + Eψ(a,b′) + Eψ(a′,b ) − Eψ(a′,b′) = 2.82,

which is 0.82 greater than the upper limit allowed by the BCHSH Ineq. (16). The
experimental result in the experiment of Weihs et al. is 2.73 +/- 0.02, in good agreement
with the Quantum Mechanical prediction of Eq. (49), and it is 30 sd away from the upper
limit of Ineq. (16). Aspect, who designed the first experimental test of a Bell Inequality
with rapidly switched analyzers (Aspect, Dalibard, Roger 1982) appreciatively
summarized the import of this result:

I suggest we take the point of view of an external observer, who collects the data from the
two distant stations at the end of the experiment, and compares the two series of results.
This is what the Innsbruck team has done. Looking at the data a posteriori, they found
that the correlation immediately changed as soon as one of the polarizers was switched,
without any delay allowing for signal propagation: this reflects quantum non-separability.
(Aspect 1999)
The experiment of Weihs et al. does not completely block the detection loophole, and
even if the experiment proposed by Fry and Walther is successfully completed, it will
still be the case that the detection loophole and the communication loophole will have
been blocked in two different experiments. It is therefore conceivable — though with
difficulty, in the subjective judgment of the present writer — that both experiments are
erroneous, because Nature took advantage of a separate loophole in each case. For this
reason Fry and Walther suggest that their experiment using dissociated mercury dimers
can in principle be refined by using electro-optic modulators (EOM), so as to block both
loopholes: “Specifically, the EOM together with a beam splitting polarizer can, in a
couple of nanoseconds, change the propagation direction of the excitation laser beam and
hence the component of nuclear spin angular momentum being observed. A separation
between our detectors of approximately 12 m will be necessary in order to close the
locality loophole” (Fry & Walther 2002) [See Fig. 2 and also note that “locality
loophole” is their term for the communication loophole.]

In the face of the spectacular experimental achievement of Weihs et al. and the
anticipated result of the experiment of Fry and Walther there is little that a determined
advocate of local realistic theories can say except that, despite the spacelike separation of
the analysis-detection events involving particles 1 and 2, the backward light-cones of
these two events overlap, and it is conceivable that some controlling factor in the overlap
region is responsible for a conspiracy affecting their outcomes. There is so little physical
detail in this supposition that a discussion of it is best delayed until a methodological
discussion in Section 7.

6. Some Variants of Bell's Theorem


This section will discuss in some detail two variants of Bell's Theorem which depart in
some respect from the conceptual framework presented in Section 2. Both are less
general than the version in Section 2, because they apply only to a deterministic local
realistic theory — that is a theory in which a complete state m assigns only probabilities 1
or 0 (‘yes’ or ‘no’) to the outcomes of the experimental tests performed on the systems of
interest. By contrast, the Local Realistic Theories studied in Section 2 are allowed to be
stochastic, in the sense that a complete state can assign other probabilities between 0 and
1 to the possible outcomes. At the end of the section two other variants will be mentioned
briefly but not summarized in detail.

The first variant is due independently to Kochen,[3] Heywood and Redhead (1983) and
Stairs (1983). Its ensemble of interest consists pairs of spin-1 particles in the entangled
state

(50) |Φ> = 1/√3 [ |z,1>1 |z,−1>2 - |z,0>1 |z,0>2 + |z,−1>1 |z,1>2 ],

where |z,i>1, with i = −1 or 0 or 1 is the spin state of particle 1 with component of spin i
along the axis z , and |z,i>2 has an analogous meaning for particle 2. If x,y,z is a triple of
orthogonal axes in 3-space then the components sx, sy, sz of the spin operator along these
axes do not pairwise commute; but it is a peculiarity of the spin-1 system that the squares
of these operators — sx2, sy2, sz2 — do commute, and therefore, in view of the
considerations of Section 1, any two of them can constitute a context in the measurement
of the third. If the operator of interest is sz2, the axes x and y can be any pair of orthogonal
axes in the plane perpendicular to z, thus offering an infinite family of contexts for the
measurement of sz2. As noted in Section 1 Bell exhibited the possibility of a contextual
hidden variables theory for a quantum system whose Hilbert space has dimension 3 or
greater even though the Bell-Kochen-Specker theorem showed the impossibility of a
non-contextual hidden variables theory for such a system. The strategy of Kochen and of
Heywood-Redhead is to use the entangled state of Eq. (50) to predict the outcome of
measuring sz2 for particle 2 (for any choice of z ) by measuring its counterpart on particle
1. A specific complete state m would determine whether sz2 of 1, measured together with
a context in 1, is 0 or 1. Agreement with the quantum mechanical prediction of the
entangled state of Eq. (50) implies that sz2 of 2 has the same value 0 or 1. But if the
Locality Conditions (8a,b) and (9a,b) are assumed, then the result of measuring sz2 on 2
must be independent of the remote context, that is, independent of the choice of sx2 and sy2
of 1, hence of 2 because of correlation, for any pair of orthogonal directions x and y in the
plane perpendicular to z. It follows that the Local Realistic Theory which supplies the
complete state m is not contextual after all, but maps the set of operators sz2 of 2, for any
direction z, noncontextually into the pair of values (0, 1). But that is impossible in view of
the Bell-Kochen-Specker theorem. The conclusion is that no deterministic Local Realistic
Theory is consistent with the Quantum Mechanical predictions of the entangled state
(50). An alternative proof is thus provided for an important special case of Bell's
theorem, which was the case dealt with in Bell's pioneering paper of 1964: that no
deterministic local realistic theory can agree with all the predictions of quantum
mechanics. An objection may be raised that sz2 of 1 is in fact measured together with only
a single context — e.g., sx2 and sy2 — while other contexts are not measured, and
“unperformed experiments have no results” (a famous remark of Peres 1978). It may be
that this remark is a correct epitome of the Copenhagen interpretation of quantum
mechanics, but it certainly is not a valid statement in a deterministic version of a Local
Realistic interpretation of Quantum Mechanics, because a deterministic complete state is
just what is needed as the ground for a valid counterfactual conditional. We have good
evidence to this effect in classical physics: for example, the charge of a particle, which is
a quantity inferred from the actual acceleration of the particle when it is subjected to an
actual electric field, provides in conjunction with a well-confirmed force law the basis for
a counterfactual proposition about the acceleration of the particle if it were subjected to
an electric field different from the actual one.

A simpler proof of Bell's Theorem, also relying upon counterfactual reasoning and based
upon a deterministic local realistic theory, is that of Hardy (1993), here presented in
Laloë's (2001) formulation. Consider an ensemble of pairs 1 and 2 of spin-½ particles, the
spin of 1 measured along directions in the xz-plane making angles a=θ/2 and a′=0 with
the z-axis, and angles b and b′ having analogous significance for 2. The quantum states
for particle 1 with spins +½ and −½ relative to direction a′ are respectively |a′,+>1 and |a′,
−>1, and relative to direction a are respectively

(51a) |a,+>1 = cosθ|a′,+>1 + sinθ|a′,−>1


and

(51b) |a,−>1 = -sinθ|a′,+>1 + cosθ|a′,−>1;

the spin states for 2 are analogous. The ensemble of interest is prepared in the entangled
quantum state

(52) |Ψ> = -cosθ|a′,+>1 |b′,−>2 − cosθ|a′,−>1 |b′,+>2 + sinθ|a′,+>1 |b′,+>2

(unnormalized, because normalization is not needed for the following argument). Then
for the specified a, a′, b, and b′ the following quantum mechanical predictions hold:

(53) <Ψ|a,+>1 |b′,+>2 = 0;


(54) <Ψ|a′,+>1 |b,+>2 = 0 ;
(55) <Ψ|a′,−>1 |b′,−>2 = 0;

and for almost all values of the θ of Eq. (52)

(56) <Ψ|a,+>1 |b,+>2 ≠ 0 ,

with the maximum occurring around θ = 9o. Inequality (56) asserts that for the specified
angles there is a non-empty subensemble E′ of pairs for which the results for a spin
measurement along a for 1 and along b for 2 are both +. Eq. (53) implies the
counterfactual proposition that if the spin of a 2 in E′ had been measured along b′ then
with certainty the result would have been -; and likewise Eq. (54) implies the
counterfactual proposition that if the spin of a 1 in E′ had been measured along a′ then
with certainty the result would have been -. It is in this step that counterfactual reasoning
is used in the argument, and, as in the argument of Kochen-Heywood-Redhead-Stairs in
the previous paragraph, the reasoning is based upon the deterministic Local Realistic
Theory. Since the subensemble E′ is non-empty, we have reached a contradiction with
Eq. (55), which asserts that if the spin of 1 is measured along a′ and that of 2 is measured
along b′ then it is impossible that both results are -. The incompatibility of a deterministic
Local Realistic Theory with Quantum Mechanics is thereby demonstrated.

An attempt was made by Stapp (1997) to demonstrate a strengthened version of Bell's


theorem which dispenses with the conceptual framework of a Local Realistic Theory and
to use instead the logic of counterfactual conditionals. His intricate argument has been the
subject of a criticism by Shimony and Stein (2001, 2003), who are critical of certain
counterfactual conditionals that are asserted by Stapp by means of a “possible worlds”
analysis without a grounding on a deterministic Local Realistic Theory, and a response
by Stapp (2001) himself, who defends his argument with some modifications.

The three variants of Bell's Theorem considered so far in this section concern ensembles
of pairs of particles. An entirely new domain of variants is opened by studying ensembles
of n-tuples of particles with n≥3. The prototype of this kind of theorem was demonstrated
by Greenberger, Horne, and Zeilinger (1989) (GHZ) for n=4 and modified to n=3 by
Mermin (1990) and by Greenberger, Horne, Shimony, and Zeilinger (1990) (GHSZ). In
the theorem of GHZ an entangled quantum state was written for four spin ½ particles and
the expectation value of the product of certain binary measurements performed on the
individual particles was calculated. They then showed that the attempt to duplicate this
expectation value subject to the constraints of a Local Realistic Theory produces a
contradiction. A similar result was obtained by Mermin for a state of 3 spin-½ particles
and by GHSZ for a state of 3 photons entangled in direction. Because of the length of
these arguments and limitations of space in the present article the details will not be
summarized here. Furthermore, for the philosophically crucial purpose of demonstrating
experimentally the validity of Quantum Mechanical predictions and the violation of the
corresponding predictions of Local Realistic Theories the examples using pairs of
particles in Section 3 are more promising than n-tuple experiments with n≥3. In
particular, it is evident that the detection loophole is more difficult to block in an
experiment performed with n-tuples of particles, n≥3 , than in an experiment using pairs,
because the net efficiency of detecting n-tuples is proportional to the product of the
efficiencies of the detectors of the individual particles.

7. Philosophical Comments
Two different classes of philosophical questions are raised by reflection upon the
theoretical and experimental investigations concerning Bell's Theorem. Questions of one
class are logical and methodological: whether one can legitimately infer from these
investigations that quantum mechanics is non-local,and whether the experimental data
definitively prove that Bell's Inequalities are violated. Questions of the other class are
metaphysical: upon assumption that the logical and methodological questions are
answered positively, what conclusions can be drawn about the structure and constitution
of the physical world, in particular is nature non-local despite the remarkable success of
relativity theory?

A logical question has been raised by Fine. In a paper (Fine 1999), which analyzes the
construction of Hardy discussed in Section 6, Fine concludes with a philosophical thesis:
“That means that the Hardy theorem, like other variants on Bell, is not a ‘proof of
nonlocality’. It is a proof that locality cannot be married to the assignment of determinate
values in the recommended way. That is interesting and significant. It is not, however, a
demonstration that quantum mechanics is nonlocal, much less (as some proclaim) that
nature is.” Fine's analysis of Hardy's construction relies upon his earlier paper (Fine
1982b) which contains the following theorem (which is a combination of his Theorem 4,
p. 1309, and footnote 5 on p. 1310):

For a correlation experiment with observables A1, A2, B1, B2 and with exactly the four
pairs Ai, Bj (i = 1,2; j= 1,2) commuting, the following statements are equivalent: (I) the
Bell/CH inequalities hold for the single and double probabilities of the the experiment;
(II) there is a joint distribution P(A1,A2,B 1,B2) compatible with the observed single and
double distributions; (III) there is a deterministic hidden variables theory for A1, A2, B1, B2
returning the observed single and double distributions; (IV) there is a well-defined joint
distribution (for the noncommuting pairs B1, B2) and joint distributions P(A1,B1,B 2),each
of the latter compatible with B1, B2 and with the observed single and double distributions;
(V) there exists a factorizable (so-called “local”) stochastic hidden variables theory for
A1, A2, B1, B2 returning the observed single and double distributions.[4]

Proposition (I) is the one of the five propositions in this theorem which is amenable to
direct experimental confirmation or disconfirmation. I shall accept for the present the
experimental disconfirmation of some of these Inequalities, leaving a consideration of
methodological doubts about this disconfirmation for discussion below. With this proviso
it follows that each of the propositions (II), (III), (IV), and (V) is disconfirmed by modus
tollens. The disconfirmation of (V) logically implies the falsity of the conjunction of all
the premisses from which the Bell/CH (called “BCH” in Section 4) Inequalities are
inferred: namely, the assumptions (5), (6), and (7) in Section 2 about the existence of
well-defined single and double probabilities, and the Independence Conditions (8a,b) and
(9a,b), respectively called “Remote Outcome Independence” and “Remote Context
Independence.” Usually the assumptions (5), (6), and (7) are not doubted, for two
reasons: first, they are implicit in the pervasive assumption in hidden-variables
investigations that the phenomenological assertions of quantum mechanics are correct —
an assumption which even permits us to use the concept of a “complete state”, denoted by
m, which is the quantum state itself if there are no hidden variables, or the complete
specification of the hidden variables if such entities exist; second, there is overwhelming
experimental confirmation of these assumptions by the practical success of quantum
mechanics, not just in experimentation regarding hidden-variables hypotheses. (In spite
of these weighty considerations there is one important program which attempts to weaken
or replace assumptions (5), (6), and (7), namely that of Stapp, briefly discussed in Section
6.) Consequently, the falsity of the conjunction of the premisses from which the BCH
Inequality is derived implies the falsity of one or both of the Independence Conditions
(8a,b) and (9a,b). Since the failure of either of these Conditions is prima facie in
contradiction with relativistic locality, it is not important for the present concern to
investigate which of the two Independence Conditions is weaker — a question that will
be taken up later in Section 7. The conclusion at this stage in the argument, when
propositions (II), (III), and (IV) of Fine's theorem have been neglected, is that the
experimental disconfirmation of the BCH Inequalities does imply the occurrence of non-
locality in natural phenomena and since the quantum mechanical analysis of pairs of
systems in entangled states anticipates non-local phenomena, quantum mechanics itself is
a a non-local theory. This pair of conclusions is what Fine claims, in the passage from
(Fine 1999) quoted above, has not been demonstrated. What he allows, in virtue of his
theorem is the weaker conclusion that “locality cannot be married to the assignment of
determinate values in the recommended way.”

But is this retrenchment to a weaker conclusion logically justified? The strong conclusion
that quantum mechanics and nature are non-local has been derived from one part of the
theorem — that (V) implies (I) — together with some auxiliary analysis of premisses (5),
(6), and (7). The other parts of the theorem — the equivalences of (II), (III), (IV) to each
other and to (I) and (V) — provide information supplementary to that provided by the
equivalence of (I) and (V), but as a matter of logic do not diminish the information given
by the last equivalence.
The last resort of a dedicated adherent of local realistic theories, influenced perhaps by
Einstein's advocacy of this point of view, is to conjecture that apparent violations of
locality are the result of conspiracy plotted in the overlap of the backward light cones of
the analysis-detection events in the 1 and 2 arms of the experiment. These backward light
cones always do overlap in the Einstein-Minkowski space-time of Special Relativity
Theory (SRT) — a framework which can accommodate infinitely many processes
besides the standard ones of relativistic field theory. Elimination of any finite set of
concrete scenarios to account for the conspiracy leaves infinitely many more as
unexamined and indeed unarticulated possibilities. What attitude should a reasonable
scientist take towards these infinite riches of possible scenarios? We should certainly be
warned by the power of Hume's skepticism concerning induction not to expect a solution
that would be as convincing as a deductive demonstration and not to expect that the
inductive support of induction itself can fill the gap formed by the shortcoming of a
deductive justification of induction (Hume 1748, Sect. 4). One solution to this problem is
a Bayesian strategy that attempts to navigate between dogmatism and excessive
skepticism (Shimony 1993, Shimony 1994). To avoid the latter one should keep the mind
open to a concrete and testable proposal regarding the mechanism of the suspected
conspiracy in the overlap of the backward light cones, giving such a proposal a high
enough prior probability to allow the possibility that its posterior probability after testing
will warrant acceptance. To avoid the former one should not give the broad and
unspecific proposal that a conspiracy exists such high prior probability that the concrete
hypothesis of the correctness of Quantum Mechanics is debarred effectively from
acquiring a sufficiently high posterior probability to warrant acceptance. This strategy
actually is implicit in ordinary scientific method. It does not guarantee that in any
investigation the scientific method is sure to find a good approximation to the truth, but it
is a procedure for following the great methodological maxim: “Do not block the way of
inquiry” (Peirce 1931).[5]

A second solution, which can be used in tandem with the first, is to pursue theoretical
understanding of a baffling conceptual problem that at present confronts us: that the
prima facie nonlocality of Quantum Mechanics will remain a permanent part of our
physical world view, in spite of its apparent tension with Relativistic locality. This
solution opens the second type of philosophical questions mentioned in the initial
paragraph of the present section, that is, metaphysical questions about the structure and
constitution of the physical world. Among the proposals for a solution of the second kind
are the following.

1. The tension between Quantum Mechanical nonlocality and Relativistic locality is


not serious, and there is indeed a kind of “peaceful coexistence” (if one adopts a
famous political maxim of Khrushchev) between the two branches of physics. It is
indeed true that the measurements in regions of space-like separation of correlated
quantities in two systems that are Quantum Mechanically entangled have
correlated outcomes that cannot be accounted for by hidden variables, and such
correlations are a kind of causality, unprecedented in pre-quantum physics. And
yet it has been shown by several investigators (Eberhard 1978, Ghirardi, Rimini &
Weber 1980, Page 1982) that this kind of causality cannot be used to send
messages superluminally between the stations of the two measurements, which is
the principal prohibition of Relativistic locality. Bohm's nonlocal model
peacefully coexists with relativistic locality for another reason: that the width of
the effective wave function which is employed in the guidance equation is not
sufficiently controllable to ensure a desired result of measurement of the quantity
required to transmit a message. The proposal of “peaceful coexistence” was in
fact espoused at one time by the present author (Shimony 1978, Section V), but he
was dissuaded from it by a powerful anti-anthropocentric argument of John Bell:

Do we then have to fall back on ‘no signaling faster than light’ as the expression
of the fundamental causal structure of contemporary theoretical physics? That is
hard for me to accept. For one thing we have lost the idea that correlations can be
explained, or at least this idea awaits reformulation. More importantly, the ‘no
signaling …’ notion rests on concepts that are desperately vague, or vaguely
applicable. The assertion that ‘we cannot signal faster than light’ immediately
provokes the question:

Who do we think we are?

We who make ‘measurements,’ we who can manipulate ‘external fields,’ we who


can ‘signal’ at all, even if not faster than light. Do we include chemists, or only
physicists, plants, or only animals, pocket calculators, or only mainframe
computers? (Bell 1990, Sec. 6.12)

2. There may indeed be “peaceful coexistence” between Quantum nonlocality and


Relativistic locality, but it may have less to do with signaling than with the
ontology of the quantum state. Heisenberg's view of the mode of reality of the
quantum state was briefly mentioned in Section 2 — that it is potentiality as
contrasted with actuality. This distinction is successful in making a number of
features of quantum mechanics intuitively plausible — indefiniteness of
properties, complementarity, indeterminacy of measurement outcomes, and
objective probability. But now something can be added, at least as a conjecture:
that the domain governed by Relativistic locality is the domain of actuality, while
potentialities have careers in space-time (if that word is appropriate) which
modify and even violate the restrictions that space-time structure imposes upon
actual events. The peculiar kind of causality exhibited when measurements at
stations with space-like separation are correlated is a symptom of the slipperiness
of the space-time behavior of potentialities. This is the point of view tentatively
espoused by the present writer, but admittedly without full understanding. What is
crucially missing is a rational account of the relation between potentialities and
actualities — just how the wave function probabilistically controls the occurrence
of outcomes. In other words, a real understanding of the position tentatively
espoused depends upon a solution to another great problem in the foundations of
quantum mechanics − the problem of reduction of the wave packet.
3. Yes, something is communicated superluminally when measurements are made
upon sysems characterized by an entangled state, but that something is
information, and there is no Relativistic locality principle which constrains its
velocity. There are many expressions of this point of view, an eloquent one being
the following by Zeilinger:

The quantum state is exactly that representation of our knowledge of the complete
situation which enables the maximal set of (probabilistic) predictions of any
possible future observation. What comes new in quantum mechanics is that,
instead of just listing the various experimental possibilities with the individual
probabilities, we have to represent our knowledge of the situation by the quantum
state using complex amplitudes. If we accept that the quantum state is no more
than a representation of the information we have, then the spontaneous change of
the state upon observation, the so-called collapse or reduction of the wave packet,
is just a very natural consequence of the fact that, upon observation, our
information changes and therefore we have to change our representation of the
information, that is, the quantum state. (1999, p. S291).

This point of view is very successful at accounting for the arbitrarily fast
connection between the outcomes of correlated measurements, but it scants the
objective features of the quantum state. Especially it scants the fact that the
quantum state probabilistically controls the occurrence of actual events.

4. A radical idea concerning the structure and constitution of the physical world,
which would throw new light upon quantum nonlocality, is the conjecture of
Heller and Sasin (1999) about the nature of space-time in the very small,
specifically at distances below the Planck length (about 10-33 cm). Quantum
uncertainties in this domain have the consequence of making ill-defined the
metric structure of General Relativity Theory. As a result, according to them,
basic geometric concepts like point and neighborhood are ill-defined, and non-
locality is pervasive rather than exceptional as in atomic, nuclear, and elementary
particle physics. Our ordinary physics, at the level of elementary particles and
above, is (in principle, though the details are obscure) recoverable as the
correspondence limit of the physics below the Planck length. What is most
relevant to Bell's Theorem is that the non-locality which it makes explicit in
Quantum Mechanics is a small indication of pervasive ultramicroscopic
nonlocality. If this conjecture is taken seriously, then the baffling tension between
Quantum nonlocality and Relativistic locality is a clue to physics in the small.
Regrettably we not longer have John Bell, with his incomparable analytic powers,
to comment on this radical proposal.

Bibliography
• Aspect, A. [1983], Trois tests expérimentaux des inégalités de Bell par mesure de
corrélation de polarization de photons (Thèse d’Etat, Orsay).
• Aspect, A. [1992], “Bell's Theorem: the naïve view of an experimentalist,” in
Quantum [Un]speakables, ed. R.A. Bertlmann and A. Zeilinger (Springer, Berlin-
Heidelberg-New York), 119-153.
• Aspect, A. [1999], “Bell's inequality test: more ideal than ever,” Nature 398, 189-
190.
• Aspect, A., Grangier, P., and Roger, G. [1982], “Experimental realization of
Einstein-Podolsky-Rosen-Bohm Gedankenexperiment: a new violation of Bell's
Inequalities,” Physical Review Letters 49, 91-94.
• Aspect, A., Dalibard, J., and Roger, G. [1982], “Experimental test of Bell's
Inequalities using time-varying analyzers,” Physical Review Letters 49, 1804-
1807.
• Baublitz, M., and Shimony, A. [1996], “Tension in Bohm's interpretation of
quantum mechanics,” in J. Cushing, A. Fine and S. Goldstein (eds), Bohmian
Mechanics and Quantum Mechanics: An Appraisal, Boston Studies in the
Philosophy of Science, Vol. 184 (Dordrecht: Kluwer Academic Publishers), 251-
264.
• Bell, J.S. [1964], “On the Einstein-Podolsky-Rosen paradox,” Physics 1, 195-200.
• Bell, J.S. [1966], “On the problem of hidden variables in quantum mechanics,”
Reviews of Modern Physics 38, 447-452.
• Bell, J.S. [1971], “Introduction to the hidden-variable question,” in Foundations
of Quantum Mechanics. Proceedings of the International School of Physics
‘Enrico Fermi’, course IL, ed. B. d’Espagnat (Academic, New York), 171-181.
• Bell, J.S. [1985] “The theory of local beables,” Dialectica 39, 97-102.
• Bell, J.S. [1987], Speakable and Unspeakable in Quantum Mechanics (Cambridge
University Press, Cambridge U.K.). Contains reprints of the preceding and related
papers.
• Bell, J.S. [1990], “La nouvelle cuisine,” in Between Science and Technology, ed.
A. Sarlemijn and P. Kroes (Elsevier, Amsterdam).
• Bohm, D. [1952], “A suggested interpretation of the quantum theory in terms of
‘hidden’ variables,” I. Physical Review 85, 166-179; II. Physical Review 85, 180-
193.
• Bohm, D. and Aharonov, Y. [1957], “Discussion of experimental proof for the
paradox of Einstein, Podolsky, and Rosen,” Physical Review 108, 1070-1076.
• Bohr, N. [1935], “Can quantum-mechanical description of physical reality be
considered complete?” Physical Review 48, 696-702.
• Clauser, J. F. [1976], “Experimental investigation of a polarization correlation
anomaly,” Physical Review Letters 36, 1223-1226.
• Clauser, J.F. and Horne, M.A. [1974], “Experimental consequences of objective
local theories,” Physical Review D 10, 526-535.
• Clauser, J.F., Horne, M.A., Shimony, A., and Holt, R.A. [1969], “Proposed
experiment to test local hidden-variable theories,” Physical Review Letters 23,
880-884..
• Clauser, J.F. and Shimony, A. [1978], “Bell's Theorem: experimental tests and
implications,” Reports on Progress in Physics 41, 1881-1927.
• de Broglie, L. [1927], “La mécanique ondulatoire et la structure atomique de la
matière et du rayonnement, ” Journal de Physique et du Radium 8, 225-241.
• de Broglie, L. [1928], “La nouvelle dynamique des quanta,” in H. Lorentz (ed.),
Rapports et Discussions du Cinquième Conseil de Physique Solvay (Paris:
Gauthier-Villars), 105-141.
• d'Espagnat, B. [1976], Conceptual Foundations of Quantum Mechanics, 2nd ed.
(Benjamin, Reading MA).
• d'Espagnat, B. [1984], “Nonseparability and the tentative descriptions of Reality,“
Physics Reports 110, 201-264.
• Dürr, D., Goldstein, S., and Zanghì, N. [1992], “Quantum equilibrium and the
origin of absolute uncertainty,” Journal of Statistical Physics 67, 843-907.
• Eberhard, P. [1978], “Bell's theorem and the different concepts of locality,”
Nuovo Cimento 46B, 392-419.
• Einstein, A., Podolsky, B., and Rosen, N. [1935], “Can quantum-mechanical
description of physical reality be considered complete?” Physical Review 47, 770-
780.
• Freedman, S.J. and Clauser, J.F. [1972], “Experimental test of local hidden
variable theories,” Physical Review Letters 28, 938-941.
• Fine, A. [1982a], “Some local models for correlation experiments,” Synthese 50,
279-294.
• Fine, A. [1982b], “Joint distributions, quantum correlations, and commuting
observables,” Journal of Mathematical Physics 23, 1306-1310.
• Fine, A. [1999], “Locality and the Hardy theorem,” in J. Butterfield and C.
Pagonis (eds), From Physics to Philosophy (Cambridge UK: Cambridge
University Press).
• Fry, E.S. and Thompson, R.C. [1976], “Experimental test of local hidden-variable
theories,” Physical Review Letters 37, 465-468.
• Fry, E.S., and Walther, T. [1997], “A Bell Inequality experiment based on
molecular dissociation — Extension of the Lo-Shimony proposal to 199Hg (nuclear
spin ½) dimers,” in Experimental Metaphysics, ed. R.S. Cohen, M. Horne, and J.
Stachel (Kluwer, Dordrecht-Boston-London), 61-71.
• Fry, E.S. and Walther, T. [2002], “Atom based tests of the Bell Inequalities – the
legacy of John Bell continues,” in Quantum [Un]speakables, ed. R.A. Bertlmann
and A. Zeilinger (Springer, Berlin-Heidelberg-New York), 103-117.
• Ghirardi, G.C., Rimini, A., and Weber, T. [1980], “A general argument against
superluminal transmission through the quantum mechanical measurement
process,” Lettere al Nuovo Cimento 27, 293-298.
• Gisin, N. [1991], “Bell's inequality holds for all non-product states,” Physics
Letters A 154, 201-202. [Note: the title of this paper is erroneous and should be
replaced by “Bell's inequality is violated by all non-product states.”]
• Gleason, A.M. [1957], “Measures on the closed subspaces of a Hilbert space,”
Journal of Mathematics and Mechanics 6, 885-893.
• Greenberger, D.M., Horne, M.A., and Zeilinger, A. [1989], “Going beyond Bell's
Theorem,” in Bell’s Theorem, Quantum Theory and Conceptions of the Universe,
ed. M. Kafatos (Kluwer, Dordrecht-Boston-London), 69-72.
• Greenberger, D.M., Horne, M.A., Shimony, A., Zeilinger, A. [1990],“Bell's
theorem without inequalities,” American Journal of Physics 58, 1131-1143.
• Gudder, S. [1970], “On hidden-variable theories,” Journal of Mathematical
Physics 11, 431-436.
• Hardy, L. [1993], “Nonlocality for two particles without inequalities for almost all
entangled states,” Foundations of Physics 13, 1665-1668.
• Heller, M. and Sasin, W. [1999], “Non-commutative unification of general
relativity and quantum mechanics,” International Journal of Theoretical Physics
38, 1619.
• Heisenberg, W. [1958], Physics and Philosophy (Harper, New York), 53.
• Heywood, P. and Redhead, M.L.G. [1983], “Nonlocality and the Kochen-Specker
paradox,” Foundations of Physics 13, 481-499.
• Holt, R.A. [1973], Ph.D. Thesis, Harvard University.
• Hume, D. [1748], Enquiry concerning Human Understanding (London).
• Jarrett, J.P. [1984], “On the physical significance of the locality conditions in the
Bell arguments,” Noûs 18, 569-589.
• Kochen, S. and Specker, E.P. [1967], “The problem of hidden variables in
quantum mechanics,” Journal of Mathematics and Mechanics 17, 59-87.
• Kocher, C.A. and Commins, E. [1967], “Polarization correlation of photons
emitted in an atomic cascade,” Physical Review Letters 18, 575-577.
• Laloë, F. [2001], “Do we really understand quantum mechanics? Strange
correlations, paradoxes, and theorems,” American Journal of Physics 69, 655-701.
• Lamehi-Rachti, M. and Mittig, W. [1976], “Quantum mechanics and hidden
variables: A test of Bell's inequality by the measurement of spin correlation in
low-energy proton-proton scattering,” Physical Review D 14, 2543-2555.
• Maudlin, T. [1994], Quantum Non-Locality and Relativity (Blackwell, Oxford
U.K. and Cambridge MA), 162-188.
• Mermin, N.D. [1986], “Generalizations of Bell's Theorem to higher spins and
higher correlations,” in Fundamental Questions in Quantum Mechanics, ed. L.M.
Roth and A. Inomato (Gordon and Breach), 7-20.
• Mermin, N.D. [1990], “Quantum mysteries revisited,” American Journal of
Physics, 58, 731-733.
• Page, D. [1982], “The Einstein-Podolsky-Rosen physical reality is completely
described by quantum mechanics,” Physics Letters 91A, 57-60.
• Peirce, C.S. [1931], “The first rule of reason,” Collected Papers, vol. 1, ed. C.
Hartshorne and P. Weiss (Harvard University Press, Cambridge MA), Parag. 135,
56.
• Peres, A. [1978], “Unperformed experiments have no results,” American Journal
of Physics 46, 745-747.
• Pitowsky, I. [1989], Quantum Probability, Quantum Logic (Springer, Berlin).
• Popescu, S. and Rohrlich, D. [1992], “Generic quantum nonlocality, Physics
Letters A 166, 293-297.
• Price, H. [1996], Time's Arrow and Archimedes' Point (Oxford University Press,
Oxford).
• Schrödinger, E. [1926]. “Quantisierung als Eigenwertproblem,” (4th
communication), Annalen der Physik 81, 109-139.
• Schrödinger, E. [1935], “Discussion of probability relations between separated
systems,” Proceedings of the Cambridge Philosophical Society 31, 555-563.
• Shimony, A., [1978], “Metaphysical problems in the foundations of quantum
mechanics,” International Philosophical Quarterly 8, 2-17.
• Shimony, A. [1993], Search for a Naturalistic World View, vol. 1 (Cambridge
University Press, Cambridge U.K.), 202-243, 274-298.
• Shimony, A., [1994] “Empirical and rational components in scientific
confirmation,” in PSA 1994, vol. 2, ed. D. Hull, M. Forbes, and R.M. Burian
(Philosophy of Science Association, East Lansing, MI).
• Shimony, A. [2001], “The logic of EPR,” Annales de la Fondation Louis de
Broglie 26 (no spécial 3/3), 399-410.
• Shimony, A. and Stein, H. [2001], “Comment on ‘Nonlocal character of quantum
theory,’ ” American Journal of Physics 69, 848-853.
• Shimony, A. and Stein, H. [2003], “On quantum non-locality, special relativity,
and counterfactual reasoning,” in Revisiting the Foundations of Relativisitc
Physics: Festschrift in Honor of John Stachel, ed. A. Ashtekar, R.S. Cohen, D.
Howard, J. Renn, S. Sarkar, and A. Shimony (Kluwer, Dordrecht-Boston-
London), 499-521.
• Stairs, A. [1983], “Quantum logic, realism, and value-definiteness,” Philosophy
of Science 50, 578-602.
• Stapp, H.P. [1997], “Nonlocal character of quantum theory,” American Journal of
Physics 65, 300-304.
• Stapp, H.P. [2001], “Response to ‘Comment on ‘Nonlocal character of quantum
theory’ by Abner Shimony and Howard Stein,’ ” American Journal of Physics 69,
854-859.
• Thompson, C.H. [1996], “The Chaotic Ball: an intuitive analogy for EPR
experiments,” Foundations of Physics Letters 9, 357.
• Tittel, W., Brendel, J., Zbinden, H., and Gisin N. [1998], “Violation of Bell's
inequalities by photons more than 10 km apart, Physical Review Letters 81, 3563-
3566.
• Walther, T., and Fry, E.S. [1997], “On some aspects of an Hg based EPR
experiment,” Proceedings of the workshop in Honor of E.C.G. Sudarshan,
Zeitschrift fűr Naturforschung 54a, 20-24.
• Weihs, G., Jennewein, T., Simon, C., Weinfurter, H., and Zeilinger, A. [1998]
“Violation of Bell's inequality under strict Einstein locality condition, “ Physical
Review Letters 81, 5039-5043.
• Wu, C.S., and Shaknov, I. [1950], Physical Review 77, 136.
• Zeilinger, A. [1999], “Experiment and the foundations of quantum physics,”
Reviews of Modern Physics 71, S288-S297.

Das könnte Ihnen auch gefallen