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
 
Quantum Physics
Some Background

From the earliest days of quantum theory, when Max Planck in 1900 hypothesized an abstract "quantum of action" and Albert Einstein in 1905 hypothesized that energy comes in discrete physical quanta, there have been disagreements about "interpretations," misunderstandings about the underlying "reality" of the external world that could account for the apparent agreement between quantum theory and the observed experimental facts, without abandoning the classical physics ideas of continuity, causality, and determinism.

For example, Planck, the inventor of the quantum of action, used his constant h as a heuristic device to calculate the probabilities of various virtual oscillators (distributing them among energy states using Boltzmann's statistical mechanics ideas, the partition function, etc.). He quantized these mechanical oscillators, but not the radiation field itself. In 1913, Bohr similarly quantized the oscillators (electrons) in the "old quantum theory" and his planetary model of the electrons orbiting the Rutherford nucleus. Bohr's electrons "jump" discontinuously from orbit to orbit, emitting or absorbing discrete amounts of energy En - Em where n and m are orbital "quantum numbers." But Bohr insisted that the energy radiated in a quantum jump was continuous, ignoring, even rejecting, Einstein's light-quantum hypothesis.

In 1905, Einstein wrote, "the energy of a light ray spreading out from a point source is not continuously distributed over an increasing space but consists of a finite number of energy quanta which are localized at points in space, which move without dividing, and which can only be produced and absorbed as whole units." It was Einstein, not Planck, who made quantum mechanics discrete and indeterministic.
By comparison to Planck, Einstein had already in 1905 quantized the continuous electromagnetic radiation field as light quanta (today's photons). Planck denied the physical "reality" of any quanta (including his own virtual oscillators) until 1910 at the earliest. And Bohr did not accept discrete photons as being emitted and absorbed during quantum jumps until twenty years after Einstein proposed them - if then. Photons are now universally accepted, of course, and (sadly) standard quantum mechanics says they are emitted and absorbed during Bohr's "quantum jumps" of the electrons.

Einstein saw clearly that if the radiation emitted by an atom were to spread out diffusely as a classical wave into a large volume of space, how could the energy collect itself together again instantly to be absorbed by another atom - without having that energy travel faster than light speed as it gathered itself together in the absorbing atom? He clearly saw that a discrete, discontinuous "jump" was involved, something denied by many of the modern "interpretations" of quantum mechanics.

Einstein feared that if energy spread out in space had to move instantly to a given point, it would violate his principle of relativity. Thus began the conflict between quantum mechanics and relativity

It was not Born who first interpreted the wave function as giving us probabilities. It was Einstein.

He also saw that the wave that filled space moments before the detection of the whole quantum of energy must disappear instantly as all the energy in the quantum is absorbed by a single atom in a particular location. This was seen as a 'collapse" of a light wave twenty years before there was a "wave function" and Erwin Schrōdinger's wave equation! Later Einstein interpreted the wave at a point as the probability of light quanta at that point, many years before the so-called "Born Rule." Max Born said many times that his "statistical interpretation" of the wave function was based on Einstein's original suggestion.

The idea of something (later called the wave function) associated with the particle led to the problem of wave-particle duality, described first by Einstein in 1909. In 1927, he expressed concern that what came to be called nonlocality violates his special theory of relativity. To this day, it is the idea that quantum physics cannot be reconciled with relativity.

The nadir of interpretation was probably the most famous interpretation of all, the one developed in Copenhagen, the one Niels Bohr's assistant Leon Rosenfeld said was not an interpretation at all, but simply the "standard orthodox theory" of quantum mechanics.

It was the nadir of interpretation because Copenhagen wanted to put a stop to "interpretation" in the sense of understanding or "visualizing" an underlying reality. The Copenhageners said we should not try to "visualize" what is going on behind the collection of observable experimental data. Just as Kant said we could never know anything about the "thing in itself," the Ding-an-sich, so the positivist philosophy of Comte, Mach, Russell, and Carnap and the British empiricists Locke and Hume claim that knowledge stops at the "secondary" sense data or perceptions of phenomena, preventing access to the primary "objects."

Einstein's views on quantum mechanics have been seriously distorted (and his early work largely forgotten), perhaps because of his famous criticisms of Born's "statistical interpretation" and Werner Heisenberg's claim that quantum mechanics was "complete" without describing what particles are doing from moment to moment.

Though its foremost critic, Einstein frequently said that quantum mechanics was a most successful theory, the very best theory so far at explaining microscopic phenomena, but that he hoped his ideas for a continuous field theory would someday add to the discrete particle theory and its "non-local" phenomena. It would allow us to get a deeper understanding of underlying reality, though at the end he despaired for his continuous field theory compared to particle theories.

Many of the "interpretations" of quantum mechanics deny a central element of quantum theory, one that Einstein himself established in 1916, namely the role of indeterminism, or "chance," to use its traditional name, as Einstein did in physics (in German, Zufall) and as William James did in philosophy in the 1880's. These interpretations hope to restore the determinism of classical mechanics. Einstein hoped for a return to deterministic physics, but even more important for him was a physics based on continuous fields, rather than discrete discontinuous particles.

We can therefore classify various interpretations by whether they accept or deny chance, especially in the form of the so-called "collapse" of the wave function, also known as the "reduction" of the wave packet or what Paul Dirac called the "projection postulate." Most "no-collapse" theories are deterministic. "Collapses" in standard quantum mechanics are irreducibly indeterministic.

Many interpretations are attempts to wrestle with still another problem that Einstein saw as early as 1905, in "non-local" events something appears to be moving faster than light and thus violating his special theory of relativity (which he formulated in 1905).

So we can classify interpretations by whether they accept the instantaneous nature of the collapse, especially the collapse of the two-particle wave function of "entangled" systems, where two particles appear instantly in widely separated places, with correlated properties that conserve energy, momentum, angular momentum, spin, etc. These interpretations are concerned about nonlocality - the idea that "reality" is "nonlocal" with simultaneous events in widely separated places correlated perfectly - a sort of "action-at-a-distance."

Many interpretations prefer wave mechanics to quantum mechanics, seeing wave theories as continuous field theories. They like to regard the wave function as a real entity rather than an abstract possibilities function. De Broglie's pilot-wave theory and its variations (e.g., Bohmian mechanics, Schrödinger's view) hoped to represent the particle as a "wave packet" composed of many waves of different frequencies, such that the packet has non-zero values in a small volume of space. Schrödinger and others found such a wave packet rapidly disperses .

Finally, we may also classify interpretations by their definitions of what constitutes a "measurement," and particularly what they see as the famous "problem of measurement." Niels Bohr, Werner Heisenberg, and John von Neumann had a special role for the "conscious observer"in a measurement. Eugene Wigner claimed that the observer's conscious mind caused the wave function to collapse in a measurement.

So we have three major characterizations - indeterministic-discrete-discontinuous "collapse" vs. deterministic-continuous "no-collapse" theories, nonlocality-faster-than-light vs. local "elements of reality" in "realistic theories, and the role of the observer.

Another way to look at an interpretation is to ask which basic element (or elements) of standard quantum mechanics does the interpretation question or just deny? For example, some interpretations deny the existence of particles. They admit only waves that evolve unitarily under the Schrōdinger equation.

We can begin by describing those elements, using the formulation of quantum mechanics that Einstein thought most perfect, that of P. A. M. Dirac.


A Brief Introduction to Basic Quantum Mechanics
Einstein said of Dirac in 1930, "Dirac, to whom, in my opinion, we owe the most perfect exposition, logically, of this
[quantum] theory"
All of quantum mechanics rests on the Schrōdinger equation of motion that deterministically describes the time evolution of the probabilistic wave function, plus three basic assumptions, the principle of superposition (of wave functions), the axiom of measurement (of expectation values for observables), and the projection postulate (which describes the collapse of the wave function that introduces indeterminism or chance during interactions).

Dirac's "transformation theory" then allows us to "represent" the initial wave function (before an interaction) in terms of a "basis set" of "eigenfunctions" appropriate for the possible quantum states of our measuring instruments that will describe the interaction.

Elements in the "transformation matrix" immediately give us the probabilities of measuring the system and finding it in one of the possible quantum states or "eigenstates," each eigenstate corresponding to an "eigenvalue" for a dynamical operator like the energy, momentum, angular momentum, spin, polarization, etc.

Diagonal (n, n) elements in the transformation matrix give us the eigenvalues for observables in quantum state n. Off-diagonal (n, m) matrix elements give us transition probabilities between quantum states n and m.

Notice the sequence - possibilities > probabilities > actuality: the wave function gives us the possibilities, for which we can calculate probabilities. Each experiment gives us one actuality. A very large number of identical experiments confirms our probabilistic predictions to thirteen significant figures (decimal places), the most accurate physical theory ever discovered.

1. The Schrōdinger Equation.

The fundamental equation of motion in quantum mechanics is Erwin Schrōdinger's famous wave equation that describes the evolution in time of his wave function ψ.

iℏ δψ / δt = H ψ         (1)

Max Born interpreted the square of the absolute value of Schrōdinger's wave function |ψn |2 (or < ψn | ψn > in Dirac notation) as providing the probability of finding a quantum system in a particular state n. As long as this absolute value (in Dirac bra-ket notation) is finite,

< ψn | ψn > ≡ ψ* (q) ψ (q) dq < ,         (2)

then ψ can be normalized, so that the probability of finding a particle somewhere < ψ | ψ > = 1, which is necessary for its interpretation as a probability. The normalized wave function can then be used to calculate "observables" like the energy, momentum, etc. For example, the probable or expectation value for the position r of the system, in con figuration space q, is

< ψ | r | ψ > = ψ* (q) r ψ (q) dq.         (3)

2. The Principle of Superposition.

The Schrōdinger equation (1) is a linear equation. It has no quadratic or higher power terms, and this introduces a profound - and for many scientists and philosophers a disturbing - feature of quantum mechanics, one that is impossible in classical physics, namely the principle of superposition of quantum states. If ψa and ψb are both solutions of equation (1), then an arbitrary linear combination of these,

| ψ > = ca | ψa > + cb | ψb >,         (4)

with complex coefficients ca and cb, is also a solution.

Together with Born's probabilistic (statistical) interpretation of the wave function, the principle of superposition accounts for the major mysteries of quantum theory, some of which we hope to resolve, or at least reduce, with an objective (observer-independent) explanation of irreversible information creation during quantum processes.

Observable information is critically necessary for measurements, though observers can come along anytime after the information comes into existence as a consequence of the interaction of a quantum system and a measuring apparatus.

The quantum (discrete) nature of physical systems results from there generally being a large number of solutions ψn (called eigenfunctions) of equation (1) in its time independent form, with energy eigenvalues En.

H ψn = En ψn,         (5)

The discrete spectrum energy eigenvalues En limit interactions (for example, with photons) to specifi c energy diff erences En - Em.

In the old quantum theory, Bohr postulated that electrons in atoms would be in "stationary states" of energy En, and that energy differences would be of the form En - Em = , where ν is the frequency of the observed spectral line.

Einstein, in 1916, derived these two Bohr postulates from basic physical principles in his paper on the emission and absorption processes of atoms. What for Bohr were assumptions, Einstein grounded in quantum physics, though virtually no one appreciated his foundational work at the time, and few appreciate it today, his work eclipsed by the Copenhagen physicists.

The eigenfunctions ψn are orthogonal to each other

< ψn | ψm > = δnm         (6)

where the "delta function"

δnm = 1, if n = m, and = 0, if n ≠ m.         (7)

Once they are normalized, the ψn form an orthonormal set of functions (or vectors) which can serve as a basis for the expansion of an arbitrary wave function φ 

| φ > = n = 0 n = ∞ cn | ψn >.         (8)

The expansion coefficients are

cn = < ψn | φ >.         (9)

In the abstract Hilbert space, < ψn | φ > is the "projection" of the vector φ onto the orthogonal axes ψn of the ψn "basis" vector set.

2.1 An example of superposition.

Dirac tells us that a diagonally polarized photon can be represented as a superposition of vertical and horizontal states, with complex number coefficients that represent "probability amplitudes." Horizontal and vertical polarization eigenstates are the only "possibilities," if the measurement apparatus is designed to measure for horizontal or vertical polarization.

Thus,

| d > = ( 1/√2) | v > + ( 1/√2) | h >          (10)

The vectors (wave functions) v and h are the appropriate choice of basis vectors, the vector lengths are normalized to unity, and the sum of the squares of the probability amplitudes is also unity. This is the orthonormality condition needed to interpret the (squares of the) wave functions as probabilities.

When these (in general complex) number coefficients (1/√2) are squared (actually when they are multiplied by their complex conjugates to produce positive real numbers), the numbers (1/2) represent the probabilities of finding the photon in one or the other state, should a measurement be made on an initial state that is diagonally polarized.

Note that if the initial state of the photon had been vertical, its projection along the vertical basis vector would be unity, its projection along the horizontal vector would be zero. Our probability predictions then would be - vertical = 1 (certainty), and horizontal = 0 (also certainty). Quantum physics is not always uncertain, despite its reputation.

3. The Axiom of Measurement.

The axiom of measurement depends on the idea of "observables," physical quantities that can be measured in experiments. A physical observable is represented as an operator A that is "Hermitean" (one that is "self-adjoint" - equal to its complex conjugate, A* = A). The diagonal n, n elements of the operator's matrix,

< ψn | A | ψn > = ∫ ∫ ψ* (q) A (q) ψ (q) dq,         (11)

are interpreted as giving the expectation value for An (when we make a measurement).

The molecule suffers a recoil in the amount of hν/c during this elementary process of emission of radiation; the direction of the recoil is, at the present state of theory, determined by "chance"...

The weakness of the theory is, on the one hand, that it does not bring us closer to a link-up with the wave theory; on the other hand, it also leaves time of occurrence and direction of the elementary processes a matter of "chance."

It speaks in favor of the theory that the statistical law assumed for [spontaneous] emission is nothing but the Rutherford law of radioactive decay.
Albert Einstein, 1916

The off -diagonal n, m elements describe the uniquely quantum property of interference between wave functions and provide a measure of the probabilities for transitions between states n and m.

It is the intrinsic quantum probabilities that provide the ultimate source of indeterminism, and consequently of irreducible irreversibility, as we shall see.

Transitions between states are irreducibly random, like the decay of a radioactive nucleus (discovered by Rutherford in 1901) or the emission of a photon by an electron transitioning to a lower energy level in an atom (explained by Einstein in 1916).

The axiom of measurement is the formalization of Bohr's 1913 postulate that atomic electrons will be found in stationary states with energies En. In 1913, Bohr visualized them as orbiting the nucleus. Later, he said they could not be visualized, but chemists routinely visualize them as clouds of probability amplitude with easily calculated shapes that correctly predict chemical bonding.

The off-diagonal transition probabilities are the formalism of Bohr's "quantum jumps" between his stationary states, emitting or absorbing energy = En - Em. Einstein explained clearly in 1916 that the jumps are accompanied by his discrete light quanta (photons), but Bohr continued to insist that the radiation was classical for another ten years, deliberately ignoring Einstein's foundational efforts in what Bohr might have felt was his area of expertise (quantum mechanics).

The axiom of measurement asserts that a large number of measurements of the observable A, known to have eigenvalues An, will result in the number of measurements with value An that is proportional to the probability of fi nding the system in eigenstate ψn.

Quantum mechanics is a probabilistic and statistical theory. The probabilities are theories about what experiments will show. Experiments provide the statistics (the frequency of outcomes) that confirm the predictions of quantum theory - with the highest accuracy of any theory ever discovered!

4. The Projection Postulate.

The third novel idea of quantum theory is often considered the most radical. It has certainly produced some of the most radical ideas ever to appear in physics, in attempts by various "interpretations" to deny it.

The projection postulate is actually very simple, and arguably intuitive as well. It says that when a measurement is made, the system of interest will be found in (will instantly "collapse" into) one of the possible eigenstates of the measured observable.

We have several possibilities for eigenvalues. We can calculate the probabilities for each eigenvalue. Measurement simply makes one of these actual, and it does so, said Max Born, in proportion to the absolute square of the probability amplitude wave function ψn.

Note that Einstein saw the chance in quantum theory at least ten years before Born

In this way, ontological chance enters physics, and it is partly this fact of quantum randomness that bothered Einstein ("God does not play dice") and Schrōdinger (whose equation of motion for the probability-amplitude wave function is deterministic).

The projection postulate, or collapse of the wave function, is the element of quantum mechanics most often denied by various "interpretations." The sudden discrete and discontinuous "quantum jumps" are considered so non-intuitive that interpreters have replaced them with the most outlandish (literally) alternatives. The famous "many-worlds interpretation" substitutes a "splitting" of the entire universe into two equally large universes, massively violating the most fundamental conservation principles of physics, rather than allow a diagonal photon arriving at a polarizer to suddenly "collapse" into a horizontal or vertical state.

4.1 An example of projection.

Given a quantum system in an initial state | φ >, we can expand it in a linear combination of the eigenstates of our measurement apparatus, the | ψn >.

| φ > = n = 0 n = ∞ cn | ψn >.         (8)

In the case of Dirac's polarized photons, the diagonal state | d > is a linear combination of the horizontal and vertical states of the measurement apparatus, | v > and | h >. When we square the (1/√2) coefficients, we see there is a 50% chance of measuring the photon as either horizontal or vertically polarized.

| d > = ( 1/√2) | v > + ( 1/√2) | h >          (10)

4.2 Visualizing projection.

When a photon is prepared in a vertically polarized state | v >, its interaction with a vertical polarizer is easy to visualize. We can picture the state vector of the whole photon simply passing through the polarizer unchanged.

The same is true of a photon prepared in a horizontally polarized state | h > going through a horizontal polarizer. And the interaction of a horizontal photon with a vertical polarizer is easy to understand. The vertical polarizer will absorb the horizontal photon completely.

The diagonally polarized photon | d >, however, fully reveals the non-intuitive nature of quantum physics. We can visualize quantum indeterminacy, its statistical nature, and we can dramatically visualize the process of collapse, as a state vector aligned in one direction must rotate instantaneously into another vector direction.

As we saw above (Figure 2.1), the vector projection of | d > onto | v >, with length (1/√2), gives us the probability 1/2 for photons to emerge from the vertical polarizer. But this is only a statistical statement about the expected probability for large numbers of identically prepared photons.

When we have only one photon at a time, we never get one-half of a photon coming through the polarizer. Critics of standard quantum theory sometimes say that it tells us nothing about individual particles, only ensembles of identical experiments. There is truth in this, but nothing stops us from imagining the strange process of a single diagonally polarized photon interacting with the vertical polarizer.

There are two possibilities. We either get a whole photon coming through (which means that it "collapsed" or the diagonal vector was "reduced to" a vertical vector) or we get no photon at all. This is the entire meaning of "collapse." It is the same as an atom "jumping" discontinuously and suddenly from one energy level to another. It is the same as the photon in a two-slit experiment suddenly appearing at one spot on the photographic plate, where an instant earlier it might have appeared anywhere.

We can even visualize what happens when no photon appears. We can imagine that the diagonal photon was reduced to a horizontally polarized photon and was completely absorbed.

Why can we see the statistical nature and the indeterminacy? First, statistically, in the case of many identical photons, we can say that half will pass through and half will be absorbed.

The indeterminacy is that in the case of one photon, we have no ability to know which it will be. This is just as we cannot predict the time when a radioactive nucleus will decay, or the time and direction of an atom emitting a photon.

This indeterminacy is a consequence of our diagonal photon state vector being "represented" (transformed) into a linear superposition of vertical and horizontal photon state vectors. Thus the principle of superposition together with the projection postulate provides us with indeterminacy, statistics, and a way to "visualize" the collapse of a superposition of quantum states into one of the basis states.

Problems in Quantum Physics
We have identified several problems in quantum physics that have new and plausible solutions when analyzed in terms of information.

The Connection to Information Philosophy
The Schrödinger equation that describes the time evolution of the wave function is linear, continuous, time-reversible, and deterministic. It predicts probabilities for all possible locations, energy eigenvalues, and other observable quantities for a quantum system. But it does not predict or describe the so-called "collapse" of the wave function.

The "collapse" of the wave function actualizes one of those possibilities.

Without the indeterministic "collapse" of the wave function, no new information could ever be created in the universe. In a universe described by a wave function that never collapses, nothing ever happens.

The "collapse" is discontinuous, irreversible, and indeterministic, involving ontological chance, as first clearly seen by Albert Einstein in 1916. Ernest Rutherford saw that the time of radioactive nuclear decay is random in 1902, and he insightfully asked Niels Bohr in 1913 "How does the electron know which of your orbits to jump to?" Bohr could not say. But it was Einstein who first saw that the time and the directions of matter and light particles are fundamentally random when light and matter interact. He called it a "weakness in the theory."

Properly understanding quantum physics is thus central to understanding information philosophy. While this will present a challenge for some philosophers, especially those philosophers of science who have spent their careers challenging the standard interpretation of quantum mechanics, our goal is to provide vivid explanations of standard quantum mechanics, and the dozen or so problems above, with new illustrative diagrams and animations for the canonical experiments.

In a deterministic world there is only one possible future. The information in such a world is constant (conserved like matter and energy).

As Claude Shannon proved in his Theory of the Communication of Information, there must be alternative possibilities for new information to be generated. If there are two possibilities, an experiment (or a message) yields one bit of information. If four possibilities, two bits, etc. Since there is just one possible future in a deterministic universe, no new information is created. Many philosophers and physicists think that information is a conserved quantity.

The Information Philosopher proposes to show that everything created since the origin of the universe over thirteen billion years ago has involved just two fundamental physical processes that combine to form the core of all creative processes. These two steps occur whenever even a single bit of new information is created and comes into the universe.

  • Step 1: A quantum process - the "collapse of a wave function."

    The formation of even a single bit of information that did not previously exist requires the equivalent of a "measurement." This "measurement" does not involve a "measurer," an experimenter or observer. It happens when the probabilistic wave function that describes the possible outcomes of a measurement "collapses" and a matter or energy particle is actually found somewhere.

  • Step 2: A thermodynamic process - local reduction, but cosmic increase, in the entropy.

    The second law of thermodynamics requires that the overall cosmic entropy always increases. When new information is created locally in step 1, some energy (with positive entropy greater than the negative entropy of the new information) must be transferred away from the location of the new bits or they will be destroyed, if local thermodynamical equilibrium is restored. This can only happen in a locality where flows of matter and energy with low entropy are passing through, keeping it far from equilibrium.

This two-step core creative process underlies the formation of microscopic objects like atoms and molecules, as well as macroscopic objects like galaxies, stars, and planets.

With the emergence of teleonomic (purposive) information in self-replicating systems, the same core process underlies all biological creation. But now some random changes in information structures are rejected by natural selection, while others reproduce successfully.

Finally, with the emergence of self-aware organisms and the creation of extra-biological information stored in the environment, the same information-generating core process underlies communication, consciousness, free will, and creativity.

The two physical processes in the creative process, quantum physics and thermodynamics, are somewhat daunting subjects for philosophers, and even for many scientists.

Quantum mechanics and thermodynamics are at the core of all creation
By creation we mean the emergence or coming into existence of recognizable information structures from a prior chaotic state in which there is no recognizable order or information.

Note there are three distinct kinds of information emergence:

  1. the order out of chaos when the matter in the universe forms information structures
  2. the order out of order when the material information structures form self-replicating biological information structures
  3. the immaterial information out of order when organisms with minds externalize information, communicating it to other minds and storing it in the environment

By information we mean a quantity that can be understood mathematically and physically. It corresponds to the common-sense meaning of information, in the sense of communicating or informing. It also corresponds to the information stored in books and computers. But it also measures the information in any physical object, like a recipe, blueprint, or production process, and the information in biological systems, including the genetic code and the cell structures.

Ultimately, the information we mean is the departure of a physical system from pure chaos, from "thermodynamic equilibrium." In equilibrium, there is only motion of the microscopic constituent particles ("the motion we call heat"). The existence of macroscopic structures, such as the stars and planets, and their motions, is a departure from thermodynamic equilibrium.

Information is mathematically related to the measure of disorder known as the entropy by Ludwig Boltzmann's famous formula S = k log W, where S is the entropy and W is the probability - the number of ways (or microstates) that the internal components (the matter and energy particles of the system) can be rearranged and still be the same system (in a particular observable macrostate).

The second law of thermodynamics says that the entropy (or disorder) of a closed physical system increases until it reaches a maximum, the state of thermodynamic equilibrium. It requires that the entropy of the universe is now and has always been increasing.

This established fact of increasing entropy led many scientists and philosophers to assume that the universe we have is "running down" to a "heat death." They think that means the universe began in a very high state of information, since the second law requires that any organization or order is susceptible to decay. The information that remains today, in their view, has always been here. There is nothing new under the sun.

But the universe is not a closed system. It is in a dynamic state of expansion that is moving away from thermodynamic equilibrium faster than entropic processes can keep up. The maximum possible entropy is increasing much faster than the actual increase in entropy. The difference between the maximum possible entropy and the actual entropy is potential information, as shown by David Layzer.

Creation of information structures means that in parts of the universe the local entropy is actually going down. Creation of a low entropy system is always accompanied by radiation of entropy away from the local structures to distant parts of the universe, into the night sky for example.

Information increases and we are co-creators of the universe

Creation of information structures means that today there is more information in the universe than at any earlier time. This fact of increasing information fits well with an undetermined universe that is still creating itself. In this universe, stars are still forming, biological systems are creating new species, and intelligent human beings are co-creators of the world we live in.

All this creation is the result of the one core creative process. Understanding this process is as close as we are likely to come to understanding the idea of an anthropomorphic creator of the universe, a still-present divine providence, the cosmic source of everything good and evil.

We will look next at the physics of quantum mechanics and thermodynamics in the creative process, then at information theory, on which we construct our information philosophy.

For Teachers
For Scholars

Part Six - Chance Part Eight - Afterword
Normal | Teacher | Scholar