Citation for this page in APA citation style.           Close


Philosophers

Mortimer Adler
Rogers Albritton
Alexander of Aphrodisias
Samuel Alexander
William Alston
Anaximander
G.E.M.Anscombe
Anselm
Louise Antony
Thomas Aquinas
Aristotle
David Armstrong
Harald Atmanspacher
Robert Audi
Augustine
J.L.Austin
A.J.Ayer
Alexander Bain
Mark Balaguer
Jeffrey Barrett
William Barrett
William Belsham
Henri Bergson
George Berkeley
Isaiah Berlin
Richard J. Bernstein
Bernard Berofsky
Robert Bishop
Max Black
Susanne Bobzien
Emil du Bois-Reymond
Hilary Bok
Laurence BonJour
George Boole
Émile Boutroux
Daniel Boyd
F.H.Bradley
C.D.Broad
Michael Burke
Lawrence Cahoone
C.A.Campbell
Joseph Keim Campbell
Rudolf Carnap
Carneades
Nancy Cartwright
Gregg Caruso
Ernst Cassirer
David Chalmers
Roderick Chisholm
Chrysippus
Cicero
Randolph Clarke
Samuel Clarke
Anthony Collins
Antonella Corradini
Diodorus Cronus
Jonathan Dancy
Donald Davidson
Mario De Caro
Democritus
Daniel Dennett
Jacques Derrida
René Descartes
Richard Double
Fred Dretske
John Dupré
John Earman
Laura Waddell Ekstrom
Epictetus
Epicurus
Austin Farrer
Herbert Feigl
Arthur Fine
John Martin Fischer
Frederic Fitch
Owen Flanagan
Luciano Floridi
Philippa Foot
Alfred Fouilleé
Harry Frankfurt
Richard L. Franklin
Bas van Fraassen
Michael Frede
Gottlob Frege
Peter Geach
Edmund Gettier
Carl Ginet
Alvin Goldman
Gorgias
Nicholas St. John Green
H.Paul Grice
Ian Hacking
Ishtiyaque Haji
Stuart Hampshire
W.F.R.Hardie
Sam Harris
William Hasker
R.M.Hare
Georg W.F. Hegel
Martin Heidegger
Heraclitus
R.E.Hobart
Thomas Hobbes
David Hodgson
Shadsworth Hodgson
Baron d'Holbach
Ted Honderich
Pamela Huby
David Hume
Ferenc Huoranszki
Frank Jackson
William James
Lord Kames
Robert Kane
Immanuel Kant
Tomis Kapitan
Walter Kaufmann
Jaegwon Kim
William King
Hilary Kornblith
Christine Korsgaard
Saul Kripke
Thomas Kuhn
Andrea Lavazza
Christoph Lehner
Keith Lehrer
Gottfried Leibniz
Jules Lequyer
Leucippus
Michael Levin
Joseph Levine
George Henry Lewes
C.I.Lewis
David Lewis
Peter Lipton
C. Lloyd Morgan
John Locke
Michael Lockwood
Arthur O. Lovejoy
E. Jonathan Lowe
John R. Lucas
Lucretius
Alasdair MacIntyre
Ruth Barcan Marcus
Tim Maudlin
James Martineau
Nicholas Maxwell
Storrs McCall
Hugh McCann
Colin McGinn
Michael McKenna
Brian McLaughlin
John McTaggart
Paul E. Meehl
Uwe Meixner
Alfred Mele
Trenton Merricks
John Stuart Mill
Dickinson Miller
G.E.Moore
Thomas Nagel
Otto Neurath
Friedrich Nietzsche
John Norton
P.H.Nowell-Smith
Robert Nozick
William of Ockham
Timothy O'Connor
Parmenides
David F. Pears
Charles Sanders Peirce
Derk Pereboom
Steven Pinker
Plato
Karl Popper
Porphyry
Huw Price
H.A.Prichard
Protagoras
Hilary Putnam
Willard van Orman Quine
Frank Ramsey
Ayn Rand
Michael Rea
Thomas Reid
Charles Renouvier
Nicholas Rescher
C.W.Rietdijk
Richard Rorty
Josiah Royce
Bertrand Russell
Paul Russell
Gilbert Ryle
Jean-Paul Sartre
Kenneth Sayre
T.M.Scanlon
Moritz Schlick
Arthur Schopenhauer
John Searle
Wilfrid Sellars
Alan Sidelle
Ted Sider
Henry Sidgwick
Walter Sinnott-Armstrong
J.J.C.Smart
Saul Smilansky
Michael Smith
Baruch Spinoza
L. Susan Stebbing
Isabelle Stengers
George F. Stout
Galen Strawson
Peter Strawson
Eleonore Stump
Francisco Suárez
Richard Taylor
Kevin Timpe
Mark Twain
Peter Unger
Peter van Inwagen
Manuel Vargas
John Venn
Kadri Vihvelin
Voltaire
G.H. von Wright
David Foster Wallace
R. Jay Wallace
W.G.Ward
Ted Warfield
Roy Weatherford
C.F. von Weizsäcker
William Whewell
Alfred North Whitehead
David Widerker
David Wiggins
Bernard Williams
Timothy Williamson
Ludwig Wittgenstein
Susan Wolf

Scientists

David Albert
Michael Arbib
Walter Baade
Bernard Baars
Jeffrey Bada
Leslie Ballentine
Marcello Barbieri
Gregory Bateson
Horace Barlow
John S. Bell
Mara Beller
Charles Bennett
Ludwig von Bertalanffy
Susan Blackmore
Margaret Boden
David Bohm
Niels Bohr
Ludwig Boltzmann
Emile Borel
Max Born
Satyendra Nath Bose
Walther Bothe
Jean Bricmont
Hans Briegel
Leon Brillouin
Stephen Brush
Henry Thomas Buckle
S. H. Burbury
Melvin Calvin
Donald Campbell
Sadi Carnot
Anthony Cashmore
Eric Chaisson
Gregory Chaitin
Jean-Pierre Changeux
Rudolf Clausius
Arthur Holly Compton
John Conway
Jerry Coyne
John Cramer
Francis Crick
E. P. Culverwell
Antonio Damasio
Olivier Darrigol
Charles Darwin
Richard Dawkins
Terrence Deacon
Lüder Deecke
Richard Dedekind
Louis de Broglie
Stanislas Dehaene
Max Delbrück
Abraham de Moivre
Bernard d'Espagnat
Paul Dirac
Hans Driesch
John Eccles
Arthur Stanley Eddington
Gerald Edelman
Paul Ehrenfest
Manfred Eigen
Albert Einstein
George F. R. Ellis
Hugh Everett, III
Franz Exner
Richard Feynman
R. A. Fisher
David Foster
Joseph Fourier
Philipp Frank
Steven Frautschi
Edward Fredkin
Benjamin Gal-Or
Howard Gardner
Lila Gatlin
Michael Gazzaniga
Nicholas Georgescu-Roegen
GianCarlo Ghirardi
J. Willard Gibbs
James J. Gibson
Nicolas Gisin
Paul Glimcher
Thomas Gold
A. O. Gomes
Brian Goodwin
Joshua Greene
Dirk ter Haar
Jacques Hadamard
Mark Hadley
Patrick Haggard
J. B. S. Haldane
Stuart Hameroff
Augustin Hamon
Sam Harris
Ralph Hartley
Hyman Hartman
Jeff Hawkins
John-Dylan Haynes
Donald Hebb
Martin Heisenberg
Werner Heisenberg
John Herschel
Basil Hiley
Art Hobson
Jesper Hoffmeyer
Don Howard
John H. Jackson
William Stanley Jevons
Roman Jakobson
E. T. Jaynes
Pascual Jordan
Eric Kandel
Ruth E. Kastner
Stuart Kauffman
Martin J. Klein
William R. Klemm
Christof Koch
Simon Kochen
Hans Kornhuber
Stephen Kosslyn
Daniel Koshland
Ladislav Kovàč
Leopold Kronecker
Rolf Landauer
Alfred Landé
Pierre-Simon Laplace
Karl Lashley
David Layzer
Joseph LeDoux
Gerald Lettvin
Gilbert Lewis
Benjamin Libet
David Lindley
Seth Lloyd
Hendrik Lorentz
Werner Loewenstein
Josef Loschmidt
Ernst Mach
Donald MacKay
Henry Margenau
Owen Maroney
David Marr
Humberto Maturana
James Clerk Maxwell
Ernst Mayr
John McCarthy
Warren McCulloch
N. David Mermin
George Miller
Stanley Miller
Ulrich Mohrhoff
Jacques Monod
Vernon Mountcastle
Emmy Noether
Donald Norman
Alexander Oparin
Abraham Pais
Howard Pattee
Wolfgang Pauli
Massimo Pauri
Wilder Penfield
Roger Penrose
Steven Pinker
Colin Pittendrigh
Walter Pitts
Max Planck
Susan Pockett
Henri Poincaré
Daniel Pollen
Ilya Prigogine
Hans Primas
Zenon Pylyshyn
Henry Quastler
Adolphe Quételet
Pasco Rakic
Nicolas Rashevsky
Lord Rayleigh
Frederick Reif
Jürgen Renn
Giacomo Rizzolati
Emil Roduner
Juan Roederer
Jerome Rothstein
David Ruelle
David Rumelhart
Tilman Sauer
Ferdinand de Saussure
Jürgen Schmidhuber
Erwin Schrödinger
Aaron Schurger
Sebastian Seung
Thomas Sebeok
Franco Selleri
Claude Shannon
Charles Sherrington
David Shiang
Abner Shimony
Herbert Simon
Dean Keith Simonton
Edmund Sinnott
B. F. Skinner
Lee Smolin
Ray Solomonoff
Roger Sperry
John Stachel
Henry Stapp
Tom Stonier
Antoine Suarez
Leo Szilard
Max Tegmark
Teilhard de Chardin
Libb Thims
William Thomson (Kelvin)
Richard Tolman
Giulio Tononi
Peter Tse
Alan Turing
Francisco Varela
Vlatko Vedral
Mikhail Volkenstein
Heinz von Foerster
Richard von Mises
John von Neumann
Jakob von Uexküll
C. S. Unnikrishnan
C. H. Waddington
John B. Watson
Daniel Wegner
Steven Weinberg
Paul A. Weiss
Herman Weyl
John Wheeler
Wilhelm Wien
Norbert Wiener
Eugene Wigner
E. O. Wilson
Günther Witzany
Stephen Wolfram
H. Dieter Zeh
Semir Zeki
Ernst Zermelo
Wojciech Zurek
Konrad Zuse
Fritz Zwicky

Presentations

Biosemiotics
Free Will
Mental Causation
James Symposium
 
Common Cause

We explore the idea that simultaneous measurement outcomes at two entangled particles widely separated in space can be correlated as the result of causes or conditions in the past light cone of the events, for example, symmetries and conservation laws at their entanglement state preparation.

This idea opposes the standard view, developed over five decades by Albert Einstein, that some kind of faster-than-light instantaneous interaction between a particle and its wave function, or between two particles, is needed to explain nonlocality and entanglement.

In his later years, Einstein famously called two-particle entanglement "spooky action at a distance" (spukhafte Fernwirkung) in a March 3, 1947 letter to Max Born.

We shall critically examine the details of entanglement state preparations and the experimental data from particle measurements in six different kinds of entanglement experiments.

The first three (from the 1930's to 1960's) were thought ("gedanken") experiments (really hypothetical theories) involving material particles, atoms and electrons.

The first was Einstein's theoretical work starting in 1905 and culminating in the 1935 EPR paper, his most cited work and the touchstone for all subsequent research on entanglement. David Bohm's proposal in the 1950's that local hidden variables could explain entanglement is the second kind of thought experiment. The third is the great work of John Stewart Bell in the 1960's, including his Bell Theorem and his claim that if hidden variables exist they must be non-local.

The fourth and the sixth kind of experiments were actual physical experiments done in laboratories with entangled light particles (photons). In the 1970's John Clauser and Stuart Freedman, and in the 1980's Alain Aspect used atomic cascades in calcium atoms to entangle photons. In the 1990's Anton Zeilinger and his colleagues in Vienna developed the spontaneous parametric down conversion(SPDC) techniques to generate entangled photons now used in most entanglement experiments. Clauser, Aspect, and Zeilinger received the 2022 Nobel Prize in Physics for their work.

Our fifth kind of entanglement study was the very popular thought experiment of David Mermin in the 1980's.

Although our focus is on experiments and their data correlations, we will also describe the mathematical quantum theory, the wave functions, superpositions or linear combinations of product states that perfectly predict the puzzling correlations.

We will also indicate the initial symmetry, physical condition, or conservation laws in the entangled state preparations that may be a "common cause" for the subsequent measurements of correlated events.

We cannot provide any physical mechanism or interactions between two particles that maintains the total angular momentum from moment to moment as they travel from state preparation to distant measurements, just as quantum mechanics cannot observe and describe the motions and interactions of electrons in and between different shells or orbitals in atoms and molecules.

But we might say that the property of total electron spin zero is "local" in the sense that it is traveling along with each particle just as David Bohm's "hidden variables" or David Mermin's "instruction sets" were thought to do.

And we can criticize the claim that the three components of spin angular momentum must exist and be defined in all spatial directions, x, y, and z (which is impossible), in order for experiments to find the spins opposite in all directions when measured.

Finally, we will trace the history in quantum mechanics of the most powerful common cause of all, the idea that every physical event can be traced back to a chain of causes and effects that goes back to the creation of the universe.

1) Einstein and the EPR paradox.

Einstein first described two particles exhibiting nonlocal behavior in a conversation with Leon Rosenfeld at a meeting in Berlin in 1933. Before this, nonlocality was between a single light quantum and its light wave.

In 1933, shortly before he left Germany to emigrate to America, Einstein attended a lecture on quantum electrodynamics by Léon Rosenfeld. Keep in mind that Rosenfeld was perhaps the most dogged defender of the Copenhagen Interpretation. After the talk, Einstein asked Rosenfeld,

“What do you think of this situation?” Suppose two particles are set in motion towards each other with the same, very large, momentum, and they interact with each other for a very short time when they pass at known positions. Consider now an observer who gets hold of one of the particles, far away from the region of interaction, and measures its momentum: then, from the conditions of the experiment, he will obviously be able to deduce the momentum of the other particle. If, however, he chooses to measure the position of the first particle, he will be able tell where the other particle is.

It is most unfortunate that Einstein did not explain that measuring the momentum of the first particle allows us to deduce the momentum of the second particle because of the conservation of linear momentum.

The same conservation principle explains, as Einstein says, "If, however, he chooses to measure the position of the first particle, he will be able to tell where the other particle is."

If Einstein had called this ability "knowledge (information) at a distance," instead of "spooky action at a distance," entanglement might never have been thought "spooky". It would be just a correlation of physical properties resulting from a common cause and a conservation law.

Einstein and colleagues Boris Podolsky and Nathan Rosen, proposed in 1935 a paradox (known by their initials as EPR or as the Einstein-Podolsy-Rosen paradox) to exhibit internal contradictions in the new quantum physics. They hoped to show that quantum theory could not describe certain intuitive "elements of reality" and thus was incomplete. They said that, as far as it goes, quantum mechanics is correct, just not "complete."

Einstein was correct that quantum theory is "incomplete" relative to classical physics, which has twice as many dynamical variables that can be known with arbitrary precision. But half of this information is missing in quantum physics, due to the indeterminacy principle which allows only one of each pair of non-commuting observables (for example momentum or position) to be known with arbitrary accuracy. Even more important, an individual particle, cannot be said to have a known position before a measurement, since evolution described by the unitary and deterministic Schrödinger equation provides us only probabilities.

The most that can be said is that the particle can be found anywhere the probability amplitude is non-zero. This was the core idea of Einstein's claim of "incompleteness." For Bohr to deny this and call quantum mechanics "complete" was just to play word games, which infuriated Einstein.

Einstein was also correct that indeterminacy makes quantum theory an irreducibly discontinuous and statistical theory. Its predictions and highly accurate experimental results are statistical in that they depend on an ensemble of identical experiments, not on any individual experiment. Einstein wanted physics to be a continuous field theory, in which all physical variables are completely and locally determined by the four-dimensional field of space-time in his theory of relativity.

Einstein and his colleagues Erwin Schrödinger, Max Planck, (later David Bohm), and others hoped for a return to deterministic physics, and the elimination of mysterious quantum phenomena like the superposition of states, the mysterious "collapse" of the wave function, and Schrödinger's famous cat. EPR continues to fascinate determinist philosophers of science who hope to prove that quantum indeterminacy does not exist.

What happens according to the information interpretation of quantum mechanics is an instantaneous change in the information about probabilities (actually complex probability amplitudes). Nothing physical (matter or energy) is moving anywhere.
As we've seen, Einstein had been bothered by "nonlocal" phenomena between a light quantum and its light wave (function) since his 1905 photoelectric paper. But this phenomenon was even more clearly exhibited in EPR experiments as the apparent transfer of something physical, an "action," from one particle to another particle faster than the speed of light.

The 1935 paper was based on Einstein's 1933 question to Leon Rosenfeld about two material particles fired in opposite directions from a central source with equal velocities. He imagined them starting at t0 some distance apart and approaching one another with high velocities. Then for a short time interval from t1 to t1 + Δt the particles are in contact with one another.

After the particles are measured and become entangled at t1, quantum mechanics describes them with a single two-particle wave function Ψ12 that is not the simple product of two one-particle wave functions Ψ1 and Ψ2.

Einstein said that at a later time t2, a measurement of one particle's position would instantly establish the position of the other particle - without measuring it explicitly. And this is correct, just as after the collision of two billiard balls, measurement of one ball tells us exactly where the other one is due to conservation of momentum. But this is not "action at a distance." It is more nearly "knowledge at a distance."

Note that Einstein is implicitly using conservation of linear momentum to know the position of the second particle. Although conservation laws are rarely cited as the explanation, they may be the physical reason that entangled particles always produce correlated results. If the results were not always correlated, the implied violation of a fundamental conservation law would be a much bigger story than mysterious entanglement itself, as interesting as that is.

This idea of something measured in one place "influencing" measurements far away challenged what Einstein thought of as "local reality." Einstein thought that when the particles moved far enough apart they could be treated as separated, with independent wave functions Ψ1 and Ψ2.

But Erwin Schrödinger quickly replied to the EPR paper, telling Einstein that his "separation principle" (Trennungsprinzip) was not correct. The particles can not be separated into a product of independent wave functions, for example, particle A in state 1, particle B in state 2,

Ψ12 ≠ Ψ1A Ψ2B

Instead, particles A and B are each randomly found in state 1 or 2. But Particle B is certain to be in state 2 if particle A is measured in state 1.

Ψ12 = 1/√2 (Ψ1A Ψ2B) + 1/√2 (Ψ2A Ψ1B)

Quantum mechanics says that the particles are not in "pure" quantum states, but a "mixture" of two states, which maintain the coherent phase relations that allow them to interfere with one another.

Schrödinger used Paul Dirac's 1926 principle of superposition and John von Neumann's 1932 motion of "mixed states" in the "density matrix" to create two of the most popular and controversial ideas in quantum mechanics.

First, Schrödinger described the two particles as "entangled" at their first encounter. He called it verschränkt in German. Verschränkt means something like cross-linked. It describes someone standing with arms crossed, where each arm reaches out to touch the other. Today EPR is the classic example of entanglement.

Second, Schrödinger introduced his famous cat, claiming it is in a mixed state or a superposition of live and dead cats!

It was at this point in quantum history that the most controversial two-particle equation above appeared that combines the ontological chance that Einstein discovered in 1916 with the idea that one quantum state can be described as in a linear combination or superposition of two other states which Dirac introduced in his transformation theory of quantum mechanics in 1927.

In Dirac's theory, the squared coefficients of the two states give us the probabilities of finding the particles in the one or the other state.

The equation combines quantum randomness with quantum interference. Interference was made famous in the two-slit experiment, which Richard Feynman almost thirty years after EPR called the “one mystery,” the only mystery at the heart of quantum mechanics.

Note that the equation does not describe two material particles interacting with one another but their abstract immaterial wave functions Ψ1 and Ψ2 interfering with one another.

These quantum mechanical wave functions, solutions to Schrödinger’s equations of motion (which replace Newton's equations of motion in classical mechanics), were thought by Schrödinger to be describing matter or energy, photons for light waves, mass and perhaps electric charge for electrons.

But in the quantum mechanics of Heisenberg, Jordan, Born, and Dirac the wave functions became “probability amplitudes,” whose absolute squares predict the probability of finding the values of observable quantum properties. And those predictions have been confirmed with extraordinary accuracy by countless experiments.

Let’s look at the equation in its simplest form that describes the superposition state of Schrödinger’s cat.

| Cat > = ( 1/√2) | Live > + ( 1/√2) | Dead >

Although this equation predicts interference between the cat states, such interference is never seen in cats, though it has been measured in surprisingly large macroscopic objects.

Nevertheless, squaring the coefficients 1/√2 tells us that there is 50% chance of finding such a cat in either the live or dead state, i.e., which is confirmed in principle.

Cats = (1/2) Live + (1/2) Dead.

Let's see how this simple equation also describes the two-slit experiment.

Ψ = ( 1/√2) | Left > + ( 1/√2) | Right >

The wave function past the two slits is a linear combination or superposition of the wave function passing the left slit | Left > with the wave function passing the right slit | Right >.

Note that whichever slit the particle passes through (and it must go through just one, because a quantum particle cannot become two, violating conservation of mass and/or energy), the probabilities of finding it on the screen are determined by the two-slit superposition. If a particle was detected passing through the left slit, or if the right slit were closed, the interference pattern would depend only on that slit's wave function | Left >.

Given that the double-slit interference appears even if only one particle at a time is incident on the two slits, we see why many say that the particle interferes with itself. But it is the wave function alone that is interfering with itself. Whichever slit the particle goes through, it is the probability amplitude ψ, whose squared modulus |ψ|2 gives us the probability of finding a particle somewhere, the interference pattern. It is what it is because the two slits are open.

This is the deepest metaphysical mystery in quantum mechanics. How can an abstract immaterial probability wave influence the particle paths to show interference when large numbers of independent particles are collected?

Why interference patterns show up when both slits are open, even when particles go through just one slit, though we cannot know which slit or we lose the interference
When there is only one slit open (here the left slit), the probabilities pattern has one large maximum (directly behind the slit) and small side fringes. If only the right slit were open, this pattern would move behind the right slit.

If we add up the results of some experiments with the left slit open and others with the right open we don't see the multiple fringes that appear with two slits open.

When both slits are open, the maximum is now at the center between the two slits, there are more interference fringes, and these probabilities apply whichever slit the particle enters. The solution of the Schrödinger equation depends on the boundary conditions - different when two holes are open. The "one mystery" remains - how these "probabilities" can exercise causal control (statistically) over matter or energy particles.

Feynman's path integral formulation of quantum mechanics suggests the answer. His "virtual particles" explore all space (the "sum over paths") as they determine the variational minimum for least action, thus the resulting probability amplitude wave function can be said to "know" which holes are open.

Einstein criticized the collapse of the wave function as "instantaneous-action-at-a-distance." This criticism resembles the criticisms of Newton's theory of gravitation. Newton's opponents charged that his theory was "action at a distance" and instantaneous. Einstein's own field theory of general relativity shows that gravitational influences travel at the speed of light and are mediated by a gravitational field that shows up as curved space-time.

For Einstein, fields like gravitation and electromagnetism are "ponderable." A disturbance of the field at one place is propagated at some finite velocity to other parts of the field. But mathematical probability is not a ponderable field in this sense.

When a probability function collapses to unity in one place and zero elsewhere, nothing physical, neither matter nor energy, is moving from one place to the other. Only information changes.

For a detailed history of Einstein's concerns about single-particle nonlocality over the thirty years before EPR, see this page.

2) David Bohm and hidden variables.

In our second kind of "thought experiment," David Bohm replaced Einstein's separating particles with a hydrogen molecule disassociating into two hydrogen atoms, each with ℏ/2 of spin angular momentum.

Instead of measuring linear momentum, Bohm proposed using two hydrogen atoms that are prepared in an initial state of known total spin angular momentum zero (the H2 molecule). Momentum and position are continuous variables. Spin is discrete. Bohm argued that measurements of discrete variables would be more precise. Bohm also proposed local "hidden variables" might be needed to explain the correlations. Here is his description. Note that it includes the two-particle wave function describing the superposition of mixed states.

We consider a molecule of total spin zero consisting of two atoms, each of spin one-half. The wave function of the system is therefore

ψ = (1/√2) [ ψ+ (1) ψ- (2) - ψ- (1) ψ+ (2) ]

where ψ+ (1) refers to the wave function of the atomic state in which one particle (A) has spin +ℏ/2, etc. The two atoms are then separated by a method that does not influence the total spin. After they have separated enough so that they cease to interact, any desired component of the spin of the first particle (A) is measured. Then, because the total spin is still zero, it can immediately be concluded that the same component of the spin of the other particle (B) is opposite to that of A.

Note that when Bohm says "because the total spin is still zero, it can immediately be concluded that the same component of the spin of the other particle (B) is opposite to that of A," he is implicitly using the conservation of total spin angular momentum.

Note also that our superposition equation for the two particles predicts a 50% chance that the first particle will be spin up (ψ+ (1)) and the second down (ψ- (2)) and a 50% chance of the reverse, that the first particle will be spin down (ψ- (1)) and the second will be up (ψ+ (2)). In either case the total spin is always certain to be conserved as zero.

Next note that while the total spin is certain to be zero, the outcome for each particle is completely random, half the outcomes are found up and the other half outcomes down.

Finally note that these amazing predictions of outcomes individually random but together perfectly correlated, as confirmed by numerous experiments, provide us with no mechanisms, no interactions between the particles that produce these perfectly correlated outcomes. It is simply that as Bohm says, "because the total spin is still zero, it can immediately be concluded that the same component of the spin of the other particle (B) is opposite to that of A."

We can ask ourselves whether our first thought experiment (EPR) really needed some mechanism, some interaction, as Einstein feared, to keep the particles moving symmetrically away from their center? In the absence of an external asymmetric force, their motions are mirror images, preserving their original symmetry and their conservation of total linear momentum.

If linear momentum can be conserved (by symmetry) without instantaneous interactions, isn't conservation of spin angular momentum a much more plausible explanation than impossible faster-than-light interactions?

In 1964, John Bell made a study of EPR and David Bohm's suggestion of local hidden variables that could provide a mechanism to explain entanglement.

Bell proposed an experiment using photons and polarizers that measures the angular dependence of the falloff in perfect correlations when experimenters at A and B (usually called Alice and Bob) don't set their polarizers at the same angle (which we argue is needed to preserve the symmetry of the initial entanglement and the conservation of critical properties like spin angular momentum).

Correlations are perfect when they measure at the same (pre-agreed-upon) angle. When their polarizer angles differ by ninety degrees, all correlations are lost.

At intermediate angle differences θ, correlations diminish proportional to the square of the cosine of their angle difference - cos2θ.

3) John Bell and His Inequalities.

4) John Clauser, Stuart Freedman, Alain Aspect, and the atomic cascade.

5) David Mermin and Instruction Sets.
4) Anton Zeilinger and spontaneous parametric down conversion (SPDC).

Normal | Teacher | Scholar