Bell's Theorem (3) - Proof of A Theorem of Bell's Type

You might also like

Download as txt, pdf, or txt
Download as txt, pdf, or txt
You are on page 1of 6

Proof of a Theorem of Bell's Type

Bell (1964) gives a pioneering proof of the theorem that bears his name, by firs t making explicit a conceptual framework within which expectation values can be calculated for the spin components of a pair of spin-half particles, then showin g that, regardless of the choices that are made for certain unspecified function s 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 d enote collectively a family of Inequalities derived in conceptual frameworks sim ilar to but more general than the original one of Bell. Sometimes these Inequali ties are referred to as Inequalities of Bell's type. Each of these conceptual fram eworks incorporates some type of hidden variables theory and obeys a locality as sumption. The name Local Realistic Theory is also appropriate and will be used thr oughout this article because of its generality. Bell calculates the expectation values for certain products of the form (1)(2), where 1 is the vectorial Pauli spin op rator for particle 1 and 2 is the vectorial Pauli spin operator for particle 2 ( both particles having spin ), and and are unit vectors in three-space, and then s hows 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 th e framework of Bell's 1964 paper can agree with all of the statistical predictio ns of quantum theory. In the present section the pattern of Bell's 1964 paper will be followed: formul ation of a framework, derivation of an Inequality, demonstration of a discrepanc y between certain quantum mechanical expectation values and this Inequality. How ever, a more general conceptual framework than his will be assumed and a somewha t 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 g iven here are Clauser et al. (1969), Bell (1971), Clauser and Horne (1974), Aspe ct (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 le ast so far, been less important for experimental tests. The conceptual framework in which a Bell-type Inequality will be demonstrated fi rst 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 c omplete 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 which is independent of the adventures of each of the two systems after they separate. D ifferent 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 inc lude characteristics of the apparatus used for the measurement, but since the de pendence of the result upon the microscopic features of the apparatus is not det erminable experimentally, only the macroscopic features of the apparatus (such a s orientations of the polarization analyzers) in their incompletely controllable environment need be admitted in practice in the descriptions a, a, etc. , and li kewise for b, b, etc. The remarkably good agreement which will be presented in Se ction 3 between the experimental measurements of correlations in Bell-type exper iments and the quantum mechanical predictions of these correlations justifies re stricting attention in practice to macroscopic features of the apparatus. The re sult 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 bivalen t, 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 per formed on 1 is s when m is the complete state, the measurements performed on 1 a nd 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 per formed on 2 is t when the complete state is m, the measurements performed on 1 a nd 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 measuremen ts 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 unit y 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 co nceptual 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 a s the conjunction of the following Independence Conditions: Remote Outcome Independence (this name is a neologism, but an appropriate on e, 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 co mplete state m is given, the outcome s of the experiment on 1 provides no additi onal information regarding the outcome of the experiment on 2, and conversely.) Remote Context Independence (this also is a neologism, but an appropriate on e, 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 o f (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 localit y. It should be emphasized that at the present stage of exposition, however, Bel l 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 ma de 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 + qr qr 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 va riables, 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) 2qr, a nd 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 ) t pm(s,t |a,b )(st ), the summation being taken over all the allow d values of s and t. and likewise with (a, b). (a, b ). and (a, b) substituted for (a, b ). Also take the qua ntities q, q, r, r of the above lemma (11) to be the single expectation values: (13a) q = p1m(|a), (13b) q = p1m(|a), (13c) r = t tp2m(t |b), (13d) r = t tp2m(t |b). Then the lemma, together with Eq. (12), factorization condition Eq. (10), and th e 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 s ystems, 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 that is, for every Borel subset B of M, (B ) is a non-negative real number, (M ) = 1, and (Uj B j) = j (Bj) where the Bj's are disjoint Borel subsets of M and Uj Bj is the set-th eoretical union of the Bj's. If we define (15a) p(,t |a,b ) M pm(s,t |a,b ) d (15b) E(a,b ) M Em(a,b )d = t M pm(s,t |a,b )(st ) d and likewise when (a, b), (a, b ), and (a, b) are substituted for (a, b ), then (14), (1 5a,b), and the properties of imply (16) 2 E(a,b ) + E(a,b) + E(a,b ) E(a,b) 2. Ineq. (16) is an Inequality of Bell's type, henceforth called the Bell-Clauser-Ho rne-Shimony-Holt (BCHSH) Inequality. The third step in the derivation of a theorem of Bell's type is to exhibit a sys tem, a quantum mechanical state, and a set of quantities for which the statistic al predictions violate Inequality (16). Let the system consist of a pair of phot ons 1 and 2 propagating in the z-direction. The two kets |x>j and |y>j constitut e 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-direct ion 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 st ate can be expressed as a linear combination of these four basis states with com plex coefficients. Of particular interest are the entangled quantum states, whic h in no way can be expressed as |>1|>2, with |> and |> single-photon states, an exam

ple being (17) | > = (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 a xes in the plane perpendicular to z. The total quantum state of the pair of phot ons 1 and 2 is invariant under the exchange of the two photons, as required by t he fact that photons are integral spin particles. Neither photon 1 nor photon 2 is in a definite polarization state when the pair is in the state |>, but their p otentialities (in the terminology of Heisenberg 1958) are correlated: if by meas urement 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 t rue of photon 2, and conversely. Suppose now that photons 1 and 2 impinge respec tively on the faces of birefringent crystal polarization analyzers I and II, wit h the entrance face of each analyzer perpendicular to z. Each analyzer has the p roperty of separating light incident upon its face into two outgoing non-paralle l 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 id eal), while a photon polarized in a direction perpendicular to z and to the tran smission axis will emerge in the extraordinary ray. See Figure 1: Image 1 Figure 1 (reprinted with permission) Photon pairs are emitted from the source, each pair quantum mechanically describ ed by |> of Eq. (17), and by a complete state m if a Local Realistic Theory is as sumed. 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 e xtraordinary 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 mecha nics provides an algorithm for computing the probabilities that photons 1 and 2 will emerge from these idealized analyzers in specified rays, as functions of th e orientations a and b of the analyzers, a being the angle between the transmiss ion axis of analyzer I and an arbitrary fixed direction in the x-y plane, and b having the analogous meaning for analyzer II: (18a) prob(,t |a,b ) = | <|>1 |t>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 | >1 |t >2 is the ket representing the quantum state of photons 1 and 2 with the respective quantum numbers s and t. Calculation of the probabilit ies of interest from Eq. (18a) can be simplified by using the invariance noted a fter Eq. (17) and rewriting | > as (19) |> = (1/2)[ |1>1 |1>2 + |1>1 |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 fo r y. Since |1>1 is orthogonal to |1>1, only the first term of Eq. (19) contributes to th e inner product in Eq. (18a) if s=t=1; and since the inner product of | 1 >1 with itself is unity because of normalization, Eq. (18a) reduces for s = t = 1 to

(18b) prob(1,1|a,b ) = ()| 2<1|1>2 |2. Finally, the expression on the right hand side of Eq. (18b) is evaluated by usin g the law of Malus, which is preserved in the quantum mechanical treatment of po larization 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) prob(1,1|a,b ) = ()cos2, whee is ba. Likewise, (20b) prob(1,1|a,b ) = () cos2, and (20c) prob(1,1|a,b ) = prob(1,1|a,b ) = ()sin2. 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 s2. Now choose as the orientation angles of the transmission axes (22) a = /4, a = 0, b = /8, b = 3 /8 . Then (23a) E(a,b ) = cos2(-/8) = 0.707, (23b) E(a,b) = cos2(/8) = 0.707, (23c) E(,b ) = cos2(/8) = 0.707, (23d) E(,b) = cos2(3/8) = 0.707. Therefore (24) S E(,b ) + E(,b) + E(,b ) E(,b) = 2.828. Eq. (24) shows that there are situations where the Quantum Mechanical calculatio ns violate the BCHSH Inequality, thereby completing the proof of a version of Be ll's Theorem. It is important to note, however, that all entangled quantum state s yield predictions in violation of Ineq. (16), as Gisin (1991) and Popescu and Rohrlich (1992) have independently demonstrated. Popescu and Rohrlich (1992) als o show that the maximum amount of violation is achieved with a quantum state of maximum degree of entanglement, exemplified by | > 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 significan ce of Bell's Theorem from a purely theoretical standpoint. What Bell's Theorem s hows is that Quantum Mechanics has a structure that is incompatible with the con ceptual framework within which Bell's Inequality was demonstrated: a framework i n which a composite system with two subsystems 1 and 2 is described by a complet e state assigning a probability to each of the possible results of every joint e xperiment on 1 and 2, with the probability functions satisfying the two Independ ence 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 prope rty of the system is measured, but the two Independence Conditions require the c

(21) E(a,b ) = prob(1,1|a,b ) + prob(1,1|a,b ) prob(1,1| a,b ) prob(1,1|a,b

ontext 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 o f 2. Therefore the incompatibility of Quantum Mechanics with this conceptual fra mework 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 preclu de models in which the contexts are required to be local. The most striking impl ication of Bell's Theorem is the light that it throws upon the EPR argument. Tha t argument examines an entangled quantum state and shows that a necessary condit ion 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 b oth is the ascription of elements of physical reality corresponding to the correla ted 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 tw o 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.

You might also like