2016 Bookmatter TheQuantumHandshake
2016 Bookmatter TheQuantumHandshake
At physics meetings at which talks are given, it is conventional to leave some time
at the end of the presentation for questions about ideas that need clarification or
that might not have been adequately covered during the formal presentation. Werner
Heisenberg remarked [1] that no one who gave a talk at Niels Bohr’s Institute in
Copenhagen ever finished the presentation, because Bohr always asked too many
questions.
So, in the spirit of Bohr’s inquisitiveness, here are some questions that may have
been raised by material in this book, along with short answers. We will start with a
few basic questions.
Q: What is the meaning of the angular frequency ω and wave number k of waves?
A: Light waves have a characteristic frequency f indicating how many times per
second the electric field of the light wave oscillates. The angular frequency ω is just
the same characteristic expressed in radians of phase per second instead of oscillations
per second, so ω = 2π f . Light waves also have a characteristic wavelength λ as
they move through space, which is the spatial distance between one electric-field
maximum and the next. The wave number k is a way of looking at the reciprocal of
that characteristic, so that k = 2π/λ. The speed of light c is related to these quantities
as c = f λ = ω/k. When we deal with particle-waves, for example using the de
Broglie wavelength, we also characterize them in terms of ω and k.
Q: What is quantization?
A: It’s the idea that there are minimum size chunks for certain quantities like
energy and angular momentum. The minimum energy chunk for light of frequency
f is E = h f = ω, where h is Planck’s constant and is h divided by 2π . We call
the particle of light carrying this minimum-size energy chunk ω a photon.
Q: What is the difference between orbital angular momentum and spin angular
momentum?
A: In an atom, two distinct kinds of angular momentum are present: the angular
momentum created by the electrons orbiting the nucleus and the intrinsic spin angu-
lar momentum of the electrons themselves. The orbital angular momentum always
Appendix A: Frequently Asked Questions About Quantum Mechanics … 171
comes in units of , while the electron spins are each /2. The total angular momen-
tum is the vector sum of these components, and it can be quite complicated because of
the variety of ways in which the component angular momentum vectors can couple.
The parity (see below) of the system wave function depends on the orbital angular
momentum, with odd values (1, 3, 5, …) giving odd parity and even values (0, 2, 4, …)
giving even parity.
Q: What is parity?
A: A parity transformation means that the three spatial coordinates (x, y, and z)
of a quantum system are reversed in sign and direction. If the quantum wave function
remains the same under such a transformation, the system is said to have positive
parity. Example: cos(x) = cos(−x). If the quantum wave function changes sign
under such a transformation, the system is said to have negative parity. Example:
sin(x) = − sin(−x). Each energy state of an atom or nucleus has a definite parity,
either positive or negative. In a “quantum jump” atomic transition, if the parity
changes the properties of the emitted photon will be different than if the parity
remains the same. Parity is related to mirror-image symmetry.
Q: What is a fermion?
A: Fundamental particles called fermions, including electrons, muons, neutrinos,
and quarks, have an intrinsic angular momentum or “spin” of 1/2 , i.e. half an unit.
Composite fermions, which are made up of three quarks, include neutrons, protons,
and heavier baryons. Fermions obey Fermi-Dirac statistics and the Pauli exclusion
principle, the requirement that only one fermion can occupy any particular quantum
state. This leads to the electron-shell structure in atoms, to the neutron and proton
shell structures in nuclei, and to valence behavior and chemical bonding in chemistry.
In quantum wave mechanics, fermions are described using the Dirac wave equation.
The half-integer spin of fermions means that one must rotate such particles through
two full revolutions or 720◦ before they return to their original state. These peculiar
172 Appendix A: Frequently Asked Questions About Quantum Mechanics …
fermion particles also come with antimatter twins (e.g., positrons for electrons and
anti-quarks for quarks) of opposite parity (opposite mirror-symmetry).1
Q: What is a boson?
A: Fundamental particles called bosons, including photons, gluons, the Z 0 and
±
W weak-interaction particles, and the Higgs particle, the mediating particles for the
fundamental forces, all have an intrinsic spin that is either zero or an integer multiple
of . Composite bosons include alpha-particles, many atoms and nuclei, and π - and
K -mesons. Bosons obey Bose–Einstein statistics and can form a “condensate” in
which a large number of bosons all have identical wave functions and occupy the same
quantum state. Lasers are an example of the operation of Bose–Einstein statistics for
photons. In quantum mechanics, bosons are described by the electromagnetic wave
equation for light and the Klein–Gordon equation for massive particles.
1I have always suspected that there must be a connections between the peculiar fractional spin of
fermions and their antimatter and parity behavior, but no one has ever been able to explain it to me,
including prominent string theorists who claim to be masters of a “theory of everything”.
Appendix A: Frequently Asked Questions About Quantum Mechanics … 173
Attempting to restrict the wave to a small region causes it to “diffract” and spread
out as it propagates further.
Interference comes from the combination of two or more waves of the same
wavelength and frequency that overlap with some definite phase relation between
them. If two waves have a phase difference of 0 or some integer multiple of 2π ,
the interference will be constructive, and the resulting wave will have twice the
amplitude of its components. If two waves have a phase difference that is some odd-
integer multiple of π , the interference will be destructive, and the resulting wave will
cancel and have zero amplitude. A good example of this kind of behavior is two-slit
interference as discussed in Sect. 6.1.
conservation, can only be preserved if the particles have values of that quantity that
are correlated. See Sect. 3.3 for a discussion of entanglement.
Q: Can quantum nonlocality be used to send nonlocal signals faster than light and
backwards in time?
A: This question is addressed in some detail in Chap. 7. The short answer is that
nonlocal signaling is impossible unless the standard formalism of quantum mechan-
ics is an approximation to a more general theory that might allow such signaling.
Several philosophers of science [2–7] have demonstrated that the formulators of the
standard quantum formalism used the impossibility of nonlocal signals as a guide
in constructing the theory and thereby built that impossibility into the formalism.
It is not clear if there might be a more general quantum formalism in which this
built-in prohibition is eliminated. Thus, the doorway to nonlocal signaling is open
just a crack, but it is a very small crack.
they fail to explain quantum phenomena,2 or (b) they are found to be inconsistent with
the QM mathematics. See Sects. 6.15 and 6.20 for examples of the latter situation.
2 In fact, many “interpretations” that fail to explain quantum nonlocality, particularly those that, like
the decoherence interpretation, focus exclusively on wave function collapse, remain alive and well
and continue to attract adherents.
3 As a personal note, my own work on the Transactional Interpretation was done well after I had
been granted tenure at the University of Washington and probably has had a net negative impact
on my academic career, since my DOE-funded primary research area, experimental nuclear and
relativistic heavy-ion physics, is very far removed from quantum interpretations and the philosophy
of science.
176 Appendix A: Frequently Asked Questions About Quantum Mechanics …
… however far the [quantum] phenomena transcend the scope of classical physical
explanation, the account of all evidence must be expressed in classical terms. The argument
is simply that by the word “experiment” we refer to a situation where we can tell others what
we have done and what we have learned, and that, therefore, the account of the experimental
arrangements and of the results of the observations must be expressed in unambiguous
language with suitable application of the terminology of classical physics.
This crucial point … implies the impossibility of any sharp separation between the behavior
of atomic objects and the interaction with the measuring instruments which serve to define the
conditions under which the phenomena appear. … Consequently, evidence obtained under
different experimental conditions cannot be comprehended within a single picture, but must
be regarded as complementary in the sense that only the totality of the phenomena exhausts
the possible information about the objects.
4I was on the second row of a large lecture hall when Niels Bohr, near the end of his life, gave a
Physics Colloquium at Rice University in 1960. He recounted for us his solution of the Einstein
Clock paradox at the 8th Solvay Conference in 1930, and he spoke at length on the power of the
principle of complementarity. His Danish-accented English was somewhat difficult to understand
at first, but it was a very interesting talk.
178 Appendix A: Frequently Asked Questions About Quantum Mechanics …
Q: That is not time-symmetric; why is it the emitter rather than the absorber that
selects the transaction”
A: There is a quantum mechanical arrow of time that favors the emitter over the
absorber. It is related to the electromagnetic arrow of time that favors retarded over
advanced electromagnetic waves. This is discussed further in Sect. 9.1.
Q: How can the Transactional Interpretation account for the interaction of two
wave functions, for example, those involving only spins?
A: This question is one for the quantum formalism, not the interpretation of it. The
wave functions that the TI describes are generated from a wave equation containing
a Hamiltonian that characterizes the interactions present in the system that is being
Appendix A: Frequently Asked Questions About Quantum Mechanics … 179
described mathematically. Thus, the wave functions, as offer and confirmation waves,
have the interactions already built into them by the formalism that they describe.
Q: Why should one accept an interpretation that has waves going backwards in
time, when nothing in the real world behaves that way?
A: Two reasons: (1) it seems to be the only way of providing any visualizable
explanation of quantum entanglement and nonlocality, and (2) the advanced waves
that go backwards in time, as complex-conjugated wave functions ψ ∗ , are present in
the standard formalism of quantum wave mechanics for all to see, if they just open
their eyes and their mind.
180 Appendix A: Frequently Asked Questions About Quantum Mechanics …
These are questions taken from Chap. 3, which discusses quantum nonlocality:
Q: Can the quantum wave functions of entangled systems be objects that exist in
normal three-dimensional space?
A: Yes. The wave functions are objects in normal three-dimensional space, and
the entanglement and nonlocal correlations between objects are arranged by the
application of conservation laws at the vertices of the final transaction.
Q: What are the true roles of the observers and measurements in quantum
processes that involve several separated measurements on entangled subsystems?
A: The observer and measurement can be end-points of a transaction, but they have
no special status. Giving “observer knowledge” a special role amounts to confusing
the effect with the cause in wave function collapse.
Q: What is wave function collapse (or state vector reduction) and how does it
occur, particularly for entangled systems?
A: Wave function collapse and state vector reduction can be understood as the
formation of a transaction that projects out one possible outcome from among the
possibilities implicitly or explicitly contained in the initial offer wave function. For
entangled systems a multi-vertex transaction forms in which conservation laws are
enforced at all vertices together to enable entanglement and nonlocal correlations.
Q: What are the underlying physical processes that make quantum nonlocality
possible?
A: The underlying physical process that makes quantum nonlocality possible is
the transactional handshake between retarded and advanced waves across space-time
that connects entangled and separated parts of an overall system back to the event
at which they separated, so that conservation laws appropriate to the system can be
enforced.
Appendix A: Frequently Asked Questions About Quantum Mechanics … 181
These are questions taken from various “challenges to the TI” extracted from Internet-
based discussions of the Transactional Interpretation.
Q: “Can the Transactional Interpretation make new predictions that are experi-
mentally testable, so that it can be verified or falsified?”
A: No consistent interpretation of quantum mechanics can be tested experimen-
tally, because each is an interpretation of the same quantum mechanical formalism,
and the formalism makes the predictions. The Transactional Interpretation is an
exact interpretation of the QM formalism. Like the Many-Worlds and the Copen-
hagen interpretations, the TI is a “pure” interpretation that does not add anything
ad hoc, but does provide a physical referent for a part of the formalism that has
lacked one (e.g., the advanced wave functions appearing in the Born probability rule
and amplitude calculations). Thus the demand for new predictions or testability from
an interpretation is based on a conceptual error by the questioner that misconstrues
an interpretation as a modification of quantum theory. According to Occam’s Razor,
the hypothesis that introduces the fewest independent assumptions is to be preferred.
The TI offers this advantage over its rivals, in that the Born probability rule is a result
rather than an independent assumption.
These questions are Nick Herbert’s “Quantum Mysteries”, taken from his online
blog “Quantum Tantra: Investigating New Doorways into Nature”. It is available
online at http://quantumtantra.blogspot.com.
Q: Quantum Mystery # 1a: What does the quantum wave function really represent?
A: The wave function represents an offer to form a transaction that constitutes
an interaction and an exchange of conserved quantities. It is a retarded wave that
propagates at the natural velocity of the particle it describes through normal three-
184 Appendix A: Frequently Asked Questions About Quantum Mechanics …
dimensional space, but it is not, in itself, observable; only the transaction it may form
is observable.
Q: Quantum Mystery # 1b: What is really happening in the world before any
measurements are made?
A: Retarded offer waves propagate from the source to potential sites for trans-
action formation, advanced confirmation waves propagate back to the source, the
source selects from among the advanced/retarded echos it receives, and a transaction
(an advanced/retarded standing wave), possibly one involving a measurement, may
form. The transaction is what is “really happening”.
Q: Quantum Mystery # 2b: How does a quantum possibility decide to turn into
an actuality?
A: The offer wave represents a “quantum possibility”. It travels to potential sites
of transaction formation, where it stimulates them to generate advanced confirmation
waves that travel back to the source. The source randomly and hierarchly selects one
confirmation for transaction formation, and a transaction forms, turning possibility
into actuality.
Q: Quantum Mystery # 3a: What (if anything) is actually exchanged between two
distant entangled quantum subsystems?
A: Nothing is directly exchanged between distant entangled quantum subsystems.
Each subsystem exchanges retarded and advanced waves with the system producing
the entangled subsystems, and the boundary conditions at each transaction vertex
enforce the conservation laws that are the basis of the entanglement and the resulting
correlation.
Q: Quantum Mystery # 3b: When will physicists get smart enough to be able to
tell their kids a believable story about what’s really going on between Alice and Bob?
A: After they read this book and understand the Transactional Interpretation.
These are questions taken from Tammaro, reference [23], Sect. 9, which criticizes
the Transactional Interpretation and a number of other QM interpretations. Many of
the questions about the TI that Tammaro raises are answered in Chaps. 5 and 6, but
here are a few that perhaps require special attention.
Appendix A: Frequently Asked Questions About Quantum Mechanics … 185
Q: How can the probability relation used for the Dirac Equation, P = ψψ † , be
reconciled with the Transactional Interpretation?
A: For the relativistic Dirac wave equation, which is used for fractional-spin
fermion particles subject to Fermi-Dirac statistics, the probability is P = ψψ † ,
where † indicates that the conjugate-transpose of a column matrix must be taken
and a matrix multiplication performed. Is this different from the Born rule discussed
above?
No. The use of column and row matrices is a book-keeping technique used in
Dirac algebra to keep track of four-momentum components, spin projections, etc.,
and the transpose operation simply ensures that appropriate pairs of elements in the
matrices will be properly paired with and multiplied by their complex conjugates,
as the Born rule indicates. The Dirac probability procedure is just the Born rule,
generalized for Dirac algebra.
Q: How can the probability relation used for the Klein–Gordon Equation, P =
i
2mc2
(ψ ∗ dtd ψ− ψ dtd ψ ∗ ), be reconciled with the Transactional Interpretation and its
use of the Born probability rule?
A: The problem with the Klein–Gordon (K-G) equation (if it is a problem) is that it
has both retarded and advanced wave functions as solutions, and the latter have energy
eigenvalues that are negative. The result is that the Born probability ψψ ∗ as applied
to general Klein–Gordon wave functions does not satisfy the continuity equation,
because it associates positive probabilities with the negative energy solutions. The
probability needs to be negative for advanced waves that when emitted carry negative
energy backwards in time to a previous point where they may form a transaction by
handshaking with a retarded wave. The probability needs to be negative because
energy disappears at the “emission” point and moves in the negative time direction,
while ψψ ∗ always gives a positive probability value.
∗ d ∗
As stated above, the K-G probability relation P = 2mc i
2 (ψ dt ψ − ψ dt ψ ) must
d
be used in the general case. Referring to Appendix B.2, the energy operator that
extracts the energy from a wave function is E = i dtd , i.e., the time derivative
∗ d ∗
multiplied by i. Therefore, the expression P = 2mc i
2 (ψ dt ψ −ψ dt ψ ) is equivalent
d
∗ ∗
to (ψ Eψ − ψEψ )/(2mc ). The energy operator E extracts the energy from the
2
wave function or its complex conjugate, subtracts them (which is equivalent to adding,
since they will have opposite signs), divides the added eigen-energies by 2mc2 , and
leaves behind the ψψ ∗ Born product. Since both retarded waves with positive energy
eigenvalues and advanced waves with negative energy eigenvalues are solutions to
the K-G equation, both kinds of waves can be the represented by ψ in the probability
equation, and the latter will have negative energy eigenvalues. If ψ is a retarded
wave, i.e., contains exp(−i Et/), the energy eigenvalue E = mc2 will be positive,
so the procedure will give a net positive probability. If ψ is an advanced wave, i.e.,
contains exp(+i Et/), the energy eigenvalue E = −mc2 will be negative, so the
procedure will give a net negative probability. Thus the K-G probability expression
is just a generalization of the Born rule that includes a book-keeping mechanism
for insuring that the probability for both advanced and retarded wave solutions to
186 Appendix A: Frequently Asked Questions About Quantum Mechanics …
the Klein–Gordon equation has the proper sign. It is completely consistent with the
Transactional Interpretation.
We note that some textbooks attempt to associate the negative-energy K-G solu-
tions with antimatter rather than with advanced waves. This perhaps works for the
special case of fermion-composite particles like π -mesons, but the Klein–Gordon
equation describes particles obeying Bose–Einstein statistics or bosons, and funda-
mental point-like bosons do not have antimatter twins. It is the fermion particles,
described by the Dirac equation and obeying Fermi-Dirac statistics and the Pauli
principle, that have the peculiarities of half-integer spins and antimatter twins with
opposite parity. Further, the association of K-G wave function solutions with “charge
currents”, i.e., multiplying probabilities by particle charge to eliminate negative prob-
abilities, is questionable because important point-like bosons, e.g., photons, gluons,
Z 0 bosons, Higgs bosons, and perhaps axions, are electrically uncharged.5
These are questions taken from the “Interpretations of Quantum Mechanics” article
by Peter J. Lewis published in the online Internet Encyclopedia of Philosophy [27].
The article, despite its general title, seems to be a puff-piece promoting the Everett-
Wheeler Many-Worlds Interpretation.
5 As a practical matter, I have had much experience doing calculations in ultra-relativistic heavy-ion
physics in which we numerically solved the Klein–Gordon equation and then used the numerical
wave function solutions produced to predict the behavior and correlations of π mesons [24–26]
from 200 GeV/nucleon ion-ion collisions. In such calculations, we did not use the K-G probability
relation quoted in the question above. Instead, we calculated the retarded-wave positive-energy K-G
wave-function solutions and used these with the ψψ ∗ Born relation to evaluate the probabilities
needed in the calculations. This works very well, as long as one does not use the advanced K-G
solutions with negative energies, which are easy to recognize and avoid.
Appendix A: Frequently Asked Questions About Quantum Mechanics … 187
Q: Does the story involving forwards and backwards waves constitute a genuine
explanation of transaction formation?
A: The question begs for a definition of “genuine”. What is required for a plausible
description and explanation (that is visible in the quantum formalism itself) to become
a “genuine explanation”? If mathematics is needed, in Sect. 5.6 we discuss Carver
Mead’s mixed-state calculation [15] using the standard quantum wave mechanics
formalism. That provides a mathematical description of the “quantum-jump” transfer
of a photon between atoms, a transaction that forms through the exchange of advanced
and retarded waves between the atoms.
Q: Isn’t the Everett interpretation the only interpretation that follows directly from
a literal reading of the standard theory of quantum mechanics?
A: That is a very peculiar assertion in view to the problems that the Everett
interpretation has with accommodating the Born Probability rule, entanglement, and
nonlocality [28], all parts of standard quantum mechanics. Everett tried to evade
the nonlocality issue by calling EPR nonlocality a “false paradox”. Further, the
Afshar experiment [29] (see Sect. 6.17) demonstrates the experimental presence of
interference in a situation in which the Everett interference rule (no interference
between distinguishable “worlds”) predicts no interference. We would say that the
Transactional Interpretation is the only interpretation that follows directly from a
literal reading of standard quantum mechanics.
Appendix B
A Brief Overview of the Quantum Formalism
Here we will provide a brief summary of the formalism of quantum wave mechanics,
keeping at a minimum the need for the reader to have a mathematical background. In
this discussion, we will use only one space dimension x. The formalism, of course,
is normally applied to waves in 3D space, but this requires the use of vectors and
gets in the way of explaining what is going on, so we will stick to 1D.
We will start with the concept of traveling waves. A useful example is a wave/kink
traveling down a clothes line. If you hit a taut clothes line with a stick, a downward
“kink” travels down the line until it hits the end, where it reflects and an upward
kink comes back. If you arrange to wiggle the end of the clothes line up and down
continuously and smoothly, ripples in the shape of sine waves travel down the line.
Ripples in a pond, ocean waves, earthquake waves, sound waves, and light waves all
behave in this same way. The wave has a fairly constant shape, but that shape moves
along in the medium (e.g., the clothes line, etc.) with some velocity.
We can focus on just those traveling waves that are sinusoidal (i. e., have the shape
of a sine wave). If you take a snapshot of a sinusoidal traveling wave on a clothes
line, you can identify its wavelength λ as the distance from one maximum to the next.
If you observe at one point along its travel, that section of rope will rise and fall and
you can identify its period T as the time it takes, starting from a height maximum, to
fall to a minimum height and then rise to the next maximum. You can also identify
its amplitude A as half the distance between a maximum point and a minimum point
as it oscillates. We can use these quantities to define some additional properties of
the wave. Its wave number k is given by k = 2π/λ and its angular frequency ω is
given by ω = 2π/T . The speed v of the wave as it travels in the medium is given by
v = λ/T = ω/k.
Using this information, we can write the wave function of the traveling wave as:
where x is a position in space, t is the time, and {} means the real part of a complex
function (see Chap. 4). The ωt term here has a minus sign in order to make the wave
move along the +x axis, because at later and later times it requires going to a larger
and larger x-value to keep the argument of the cosine function at the same value,
e.g., at a maximum or minimum.
The wave functions of quantum mechanics differ in only one way from the wave
functions for sound waves or ripples on a clothes line: we must use the entire complex
function instead of using just the real part:
In the quantum world, we will use wave functions for describing massless photons
of light, and we will also use wave functions for describing two classes of massive
particles, bosons and fermions. For the purposes of this discussion, all of these particle
types share the same form for their wave functions, but the quantum wave equations
to which these wave functions are solutions may be very different.
The wave function ψ(x, t) = A exp[i(kx − ωt)] of Eq. B.2 is what is called a “plane
wave”. It assumes that the amplitude A is a constant and describes traveling waves
with a definite momentum p = k and energy E = ω, but with no definite locations
in time or space. However, what do we do if we wish to describe a “wave packet” that
is localized in space? The answer is that we must make the amplitude A a function of
position x, which has the effect of making k a variable with a distribution of values.
This makes the wave more likely to be in the place where A is large than in the place
where A is small.
If we wish to localize a particle at a position x0 within a position uncertainty of
σx we can write the wave function as
√
ψ(x, t) = (σx 2π )−1 exp[−(x − x0 )2 /2σx2 ] exp[i(kx − ωt)] (B.3)
This makes the amplitude A a unit-area Gaussian distribution centered about x0 with
a variance of σx , as shown in Fig. B.1.
What is the effect of this localization on the variable k which is complementary
to x? We can answer this question, as discussed in Sect. 2.4, by calculating φ(k), the
momentum-space wave function of ψ(x), which is the Fourier transform of ψ(x). It is
well known that the Fourier transform of a Gaussian distribution is another Gaussian
distribution, with the widths of the two Gaussians in a see-saw relation: The Fourier
Appendix B: A Brief Overview of the Quantum Formalism 191
Fig. B.1 A quantum wave function ψ(x) of Eq. B.3 localized at x0 = 0 with a width σx = 1 and
k = 10. The real part of ψ(x) is shown as red/solid and the imaginary part as blue/dashed
Now that we have a wave function, we can ask what it describes and how we can
access the information that it contains. Let us start by focusing on light waves.
Light is a combination of a sinusoidal electric field and a sinusoidal magnetic field,
vibrating in phase but at right angles to each other and moving at the speed of light
in the direction perpendicular to both fields. Figure B.2 shows a vertically polarized
classical light wave, as predicted by Maxwell’s equations.
The quantum mechanical version of light differs from this picture because
the energy and momentum are quantized into photons. Each photon carries an
energy of E = ω and a momentum of p = k, where is Planck’s constant
(6.62606957 × 10−34 m2 kg/s) divided by 2π . By substituting in the energy and
momentum expressions, we can write the wave function ψγ that is the offer wave
for a photon as: ψγ (x, t) = A exp[ i ( px − Et)].
We can see that this wave function contains the photon’s energy and momentum
as interior variables, but how can we get them out? The trick is to use calculus. The
192 Appendix B: A Brief Overview of the Quantum Formalism
Fig. B.2 A vertically polarized classical light wave showing the electric (red) and magnetic (green)
field directions, the wavelength λ, and the direction of wave motion (pink)
calculus operation of differentiation, ddx f (x), determines the rate at which a function
f (x) changes when x is changed. For example, if f (x) = x 2 , then ddx f (x) = 2x.
So what happens if we differentiate the function f (x) = exp(ax)? The answer is:
d
dx
exp(ax) = aexp(ax). In other words, differentiating the exponential with respect
to x leaves it unchanged, except that it pulls out the quantity that is multiplying x
inside the exponential.
Therefore, ddx ψγ (x, t) = ip ψγ (x, t). In other words, the operation of differenti-
ating the photon wave function with respect to x pulls the quantity i p/ out of the
wave function. This is good, but we would rather not have the i/ as extra baggage.
Therefore, we can define the momentum operator:
d
P≡ (B.4)
i dx
Pψγ (x, t) = pψγ (x, t). (B.5)
The momentum operator P extracts just the momentum p from the wave function.
Similarly, differentiation with respect to the time t will extract a quantity that includes
the energy E, the variable in the wave function that t multiplies. In particular, we can
define the energy operator as E ≡ i dtd . Here the i has moved from denominator to
numerator because the Et term has a minus sign, and i = −1/i. Thus, Eψγ (x, t) =
Eψγ (x, t) and energy operator extracts the photon energy E from the wave function.
In the language of quantum mechanics, we say that p and E are the momentum and
energy eigenvalues of the wave function.
If these operators operate twice on the wave function, they extract the square of
the eigenvalue. This allows us to define square operators that extract the momentum-
squared and the energy-squared. The momentum-squared operator is P2 ≡ PP ≡
2
−2 ddx 2 , so that P2 ψγ (x, t) = p 2 ψγ (x, t). Similarly, the energy-squared operator is
2
E2 ≡ EE ≡ −2 dtd 2 , so that E2 ψγ (x, t) = E 2 ψγ (x, t).
Appendix B: A Brief Overview of the Quantum Formalism 193
So far, we have used the wave functions without asking where they came from. Now
we will drop the other shoe. The wave equation is a differential equation that relates
the wave function to its derivatives, and the wave function is a solution of this wave
equation, i.e., a mathematical function that, when differentiated once or twice and
substituted into the wave equation will satisfy the indicated equality.
The electromagnetic wave equation is used in both classical mechanics and quan-
tum mechanics. It is derived by manipulating Maxwell’s equations, the set of equa-
tions that describe the classical behavior of electric and magnetic fields. The elec-
tromagnetic wave equation for the electromagnetic vector potential A of an electro-
2 2
magnetic wave is the differential equation dtd 2 A = c2 ddx 2 A. The quantum mechanical
version of the electromagnetic wave equation is:
d2 ω2 d 2 2 d
2
ψγ = ψγ = c ψγ . (B.6)
dt 2 k2 d x 2 dx2
We note that the operations performed on ψγ are similar to the momentum-squared
operator P2 and energy-squared operator E2 defined in Sect. B.2 above. In particular,
using these operators, we can rewrite the wave equation as:
Thus, in operator terms the electromagnetic wave equation is just the operator version
of the energy-momentum relation of a photon, i.e., E = cp or E 2 = c2 p 2 .
The energy-momentum relation for massive particles is different from the above
photon relation. In particular, if E is the kinetic energy and p the momentum of a
particle of mass m, then the non-relativistic relation between them is p 2 /2m = E.
This suggests that, using operators, the wave equation for a massive particle should
be similar, i.e., that:
(1/2m)P2 ψm = Eψm . (B.9)
−2 d 2 d
2
ψm = i ψm (B.10)
2m d x dt
or more simply;
i d 2 d
ψm = ψm . (B.11)
2m d x 2 dt
194 Appendix B: A Brief Overview of the Quantum Formalism
This is the Schrödinger wave equation, the fundamental non-relativistic wave equa-
tion of quantum mechanics.
Thus, by the straightforward use of energy-momentum relations and operators
we have derived that fundamental wave equations of quantum mechanics. The quan-
tum mechanical wave function given in Sect. B.1 above is the solution for both the
electromagnetic wave equation and the Schrödinger equation.
The use of operators to extract values of variables from the wave function, as
described above, is incomplete, because the extracted quantity is multiplied by the
wave function. The question of how to make a prediction of the outcome of a mea-
surement remains to be defined.
As indicated in Sect. B.2, we can construct operators that can extract quantities
like energy and momentum from the wave function. These can be used to calcu-
late a prediction of the “expectation value” of the quantity by making a “quantum
mechanical sandwich” with wave functions as the “bread” and an operator as the
“meat”. This is interpreted by the TI as follows: The offer wave ψ contains informa-
tion that can be extracted by some operator Q (perhaps a device that measures the
quantity q) so that Qψ = qψ. Multiplying this by the confirmation wave ψ ∗ gives
ψ ∗ (Qψ) = qψψ ∗ = ( pr ob)q. The latter is the probability of the variable of interest
having a value of q at the location of the wave function ψ. Since we are interested in
the overall value, this must be averaged over all space to get the “expectation value”
of q, which we represent as q .
For example, suppose that we want to calculate the expectation value E of the
energy of a system described by wave function ψ(x). The procedure to calculate it is:
∞
E = ψ ∗ (x)Eψ(x)d x. (B.12)
−∞
That’s how quantum wave mechanics calculations are done. The Transactional
Interpretation provides insights into why this mathematics describes Nature in action.
Appendix C
Quantum Dice and Poker—Nonlocal Games
of Chance
In this Appendix we want to describe two games of chance that are analogous to the
behavior of entangled particles. These may help in understanding how non-classical
the quantum game is.
Suppose that you are given two dice that have somehow been entangled by a conser-
vation law that requires the number seven. You roll one of the dice and get a random
face between one and six. You can plot the results of many such rolls, and you find
a completely flat distribution, which each face equally probable.
Then you roll its entangled twin, and the face that appears, when added to that
from the first roll, is always seven. However, if you plot the distribution of faces from
many rolls, again there is a completely flat distribution. The dice are, at the same
time, completely random in the numbers they produce individually, and perfectly
correlated in their sum, in that together they always produce a seven.
In this way, the quantum dice behave in the same way as a pair of entangled par-
ticles, which give random but correlated results of selected measurements. Perhaps,
contrary to Einstein’s expectations, the Old One does play dice, and they are loaded
by entanglement.
Quantum poker is a “game” played with ordinary playing cards that illustrates the
nature of the nonlocal quantum correlations in EPR situations. Alice and Bob are
playing a game of quantum poker. The deck, handled only by the Dealer, consists of
eight cards, two copies each of four card types: [ A♠], [K ♠], [A♥], and [K ♥].
Each players chooses a “measurement” to make during the game and whispers
to the Dealer their choice of measurement. The measurement selected may be either
color (red or black) or hi/lo (ace or king). The Dealer randomly draws one card from
the deck, selects its duplicate from the deck, and gives one of these cards to Alice
and one to Bob. The Dealer then gives each player a second card, based on their
measurement choice. If the player is measuring color, the Dealer gives the player
the other card in the deck that is the same color; if the player is measuring hi/lo, the
Dealer gives the player the other high or low card from the deck.
Examples: (A1) Alice declares that she is measuring color, receives the [A♥],
and the Dealer also gives her the [K ♥]; (A2) Alice is measuring hi/lo, receives the
[A♥], and the Dealer also gives her the [ A♠]; (B1) Bob is measuring hi/lo, receives
the [K ♥], and the Dealer also gives him the [K ♠]; (B2) Bob is measuring color,
receives the [K ♥], and the Dealer also gives him the [A♥].
We can see that when Bob choses to do the same measurement as Alice, he always
receives the same resulting card pair. However, if he choses the other measurement,
he always gets a non-matching result. In any case, it should be clear that Alice cannot
send a message to Bob by using her measurement choice, and Bob cannot receive
such a message.
This game of quantum poker is exactly the same situation as if Alice and Bob
were receiving polarization-entangled photons, as in the Freedman-Clauser exper-
iment described in Sect. 6.8. They could, for example, choose to measure either
horizontal/vertical linear polarization or diagonal/antidiagonal 45◦ linear polariza-
tion for their received photon. If Alice and Bob choose the same measurement, they
will always get the same result. However, if they choose to do different measure-
ments, they will get uncorrelated results. In any case, Alice cannot send a message to
Bob using her choice of polarization measurements, even though their photons are
entangled.
One question remains: how does Nature arrange to play the role of the Dealer
in the card game of quantum entanglement, in which Alice and Bob may be very
far apart? The Transactional Interpretation provides an answer to this question, as
discussed in Sect. 6.6.
Appendix D
Detailed Analyses of Selected
Gedankenexperiments
| 1 ≡ | L : S1 : M1 : S2 : C (D.1)
| 2 ≡ | L : S1 : M1 : S2 : D (D.2)
| 3 ≡ | L : S1 : M2 : S2 : C (D.3)
| 4 ≡ | L : S1 : M2 : S2 : D (D.4)
| 5 ≡ | L : S1 : Z + . (D.5)
There are also “dead end” confirmation waves from C and D to the backside of the
atom that will cancel and vanish. These are:
Appendix D: Detailed Analyses of Selected Gedankenexperiments 201
11 | ≡ C : S2 : M2 : Z + | (D.11)
12 | ≡ D : S2 : M2 : Z + | (D.12)
The cancellation occurs because they were stimulated by equal-intensity offer waves
that were already 90◦ out of phase because of the reflection to D at S2 , and the
time-reversed confirmation wave from C receives an additional −90◦ phase shift on
reflection from S2 on its path to the atom. Thus, along the path to the atom the two
confirmation waves have equal amplitudes and a 180◦ phase difference, so that they
will cancel. We note that this subtlety was missed in reference [22] and was pointed
out in reference [30].
The offer wave components from the atom source X 0 to the atom detector
X ± are:
202 Appendix D: Detailed Analyses of Selected Gedankenexperiments
| a ≡ | X 0+ : Z + : X + (D.13)
| b ≡ | X 0+ : Z + : X − (D.14)
| c ≡ | X 0+ : Z − : X + (D.15)
| d ≡ | X 0+ : Z − : X − . (D.16)
If there had been no interaction between the photon and the atom (or if the Z +
box had not been in the photon’s path), the photon offer waves | 2 and | 4 and
atom offer waves | b and | d would be in destructive interference superpositions
and would cancel. This prevents detection of a photon at “dark” detector D and
prevents measurement of X-axis − 1/2 spin projection for the atom. The photon-
atom interaction with a 100 % interaction strength, as assumed by Hardy, means that
photon offer waves | 3 and | 4 may be completely blocked by the presence of
the atom in the Z + box. The atom offer waves | c and | d will also be absent
because absorption of the photon will place that atom in an excited state in box Z +,
which breaks the coherent superposition of the two Z-axis projections. The source
L has one photon to emit and three confirmation types to choose from: (1) 10 |:
no photon is detected at C or D because the photon was absorbed at the Z + box,
and the subsequent X-axis spin measurement on the atom had equal probabilities for
X -axis spin + 1/2 and − 1/2 , because the state superposition was broken by the atomic
excitation; (2) 6 | and 8 |: the photon is detected at C and the subsequent X-axis
spin measurement on the atom showed a preference for X -axis spin + 1/2 , with some
probability of − 1/2 ; and (3) 7 |: the photon is detected at D, the atom is present in
box Z + suppressing 9 |, and the subsequent X-axis spin measurement on the atom
had equal probabilities for X -axis spin + 1/2 and − 1/2 , i.e., transactions a | e and
b | f .
Hardy focused on outcome (3), and so we will examine it in detail. The photon is
detected at D because the destructive superposition of offer waves | 2 and | 4 has
been broken because offer wave | 4 is blocked by the atom, which must be in box
Z +. The photon transaction is confirmed only by 7 |, because 9 | is also blocked
by the atom. If this transaction forms, no other transaction may form because of the
one-photon boundary condition at source L. Similarly, the atom can have either of
two transactions, offer wave | a or | b confirmed by confirmation waves | e or
| f , respectively, so that measurement of X-axis spin projections + 1/2 and − 1/2
are equally probable. However, only one of these transactions may form, because of
the one-atom boundary condition at X 0.
Appendix D: Detailed Analyses of Selected Gedankenexperiments 203
Thus, photon detection at D means that the atom is definitely in box Z + and that
the atom, initially prepared at X 0 in X-axis spin state + 1/2 , can be found with equal
probability at X in either spin state, even though there has been no real interaction
between the photon and the atom. The Transactional Interpretation provides a way
of visualizing this non-classical result and understanding its origin.
Figure D.3 shows a polarization-entangled EPR experiment much like the Freedman-
Clauser expeeriment, except that we are using a variable-entanglement source like
the one described in Sect. 7.3. We initially set α = 0 for 100 % entanglement. When
θ is zero and the polarimeters are aligned, there will be a perfect anti-correlation
between the polarizations measured by Alice and by Bob. The random polarization
(H or V) that Alice measures will always be the opposite of that measured by Bob
(H A VB or V A H B ). However, when θ is increased, the perfect H A VB and V A H B anti-
correlations are degraded and correlated detections H A H B and V A VB , previously not
present, will begin to appear. Local theories require that for small θ rotations this
correlation degradation should increase linearly with θ , while quantum mechanics
predicts that it should increase as θ 2 , i.e., quadratically [31]. This is the basis of Bell’s
Inequalities discussed in Sect. 2.8.
The quantum mechanical analysis of this system is fairly simple because, assum-
ing that the entangled photons have a single spatial mode, their transport through
the system can be described by considering only the phase shifts and polarization
selections that the system elements create in the waves. We have used the formalism
of Horne, Shimony and Zeilinger [32] to perform such an analysis and to calculate
the joint wave functions for simultaneous detections at both detectors. These are:
Fig. D.3 A two-photon 4-detector EPR experiment using linear polarization with variable entan-
glement
204 Appendix D: Detailed Analyses of Selected Gedankenexperiments
√
Ψ H H (α, θ ) = [− sin(α) cos(θ ) + i cos(α) sin(θ )]/ 2 (D.21)
√
Ψ H V (α, θ ) = [− cos(α) cos(θ ) + i sin(α) sin(θ )]/ 2 (D.22)
√
ΨV H (α, θ ) = [cos(α) cos(θ ) − i sin(α) sin(θ )]/ 2 (D.23)
√
ΨV V (α, θ ) = [sin(α) cos(θ ) − i cos(α) sin(θ )]/ 2. (D.24)
Figure D.4 shows plots of these joint detection probabilities versus θ for the four
detector combinations with: α = 0 (100 % entangled), α = π/8 (71 % entangled),
Fig. D.4 Joint detection probabilities versus θ for the four detector combinations with: α = 0
(red/solid, 100 % entangled), α = π/8 (green/dashed, 71 % entangled), and α = π/4 (blue/dot-
dashed, 0 % entangled)
Appendix D: Detailed Analyses of Selected Gedankenexperiments 205
and α = π/4 (0 % entangled). One can see from these plots and equations that the
singles probabilities are: PH H + PH V = PV H + PV V = PH H + PV H = PH V + PV V =
1/2 , and so no signaling is possible.
Fig. D.6 Bob’s non-coincident singles detector probabilities PB1 (α, φ B ) and PB0 (α, φ B )
(Eqs. D.37 and D.38) for α = 0 (red/solid, 100 % entangled), α = π/8 (green/dash, 71 % entan-
gled), and α = π/4 (blue/dot-dash, 0 % entangled)
The non-coincident singles detector probabilities for Bob’s detectors are obtained
by summing over Alice’s detectors, which he does not observe. Thus
Note that these singles probabilities have no dependences on Alice’s phase φ A for
any value of α. Here again we see an example of Schrödinger steering, in that Alice
is manipulating the wave functions that arrive at Bob’s detectors, but not in such a
way that would permit signaling.
Figure D.6 shows plots of Bob’s non-coincident singles detector probabilities
PB1 (α, φ B ) and PB0 (α, φ B ) for the cases of α = 0 (100 % entangled), α = π/8
(71 % entangled), and α = π/4 (not entangled).
Appendix D: Detailed Analyses of Selected Gedankenexperiments 207
The non-coincident singles detector probabilities for Bob’s detectors are identical to
the singles detector probabilities of Eqs. D.37 and D.38 obtained when B S A was in
place.
We have performed this analysis of the experiment shown in Fig. D.7, tweaking the
mirror angles for maximum overlap of the waves on the two paths to detector D A .
The calculation gives large analytical expressions for joint detection probability as
a function of position on detector D A , but these must be integrated numerically to
208 Appendix D: Detailed Analyses of Selected Gedankenexperiments
Fig. D.7 A 3-detector wedge modification of the path-entangled dual-interferometer EPR experi-
ment with variable entanglement
Fig. D.8 Magnitudes of the wave functions Ψa1 (red/solid) and Ψa2 (blue/dotted) as functions of
position x on the face of detector D A . Oscillations are the result of Gaussian tail truncation by the
apex of wedge mirror W A
obtain the position-independent probabilities. Here Fig. D.8 shows the overlap of
the magnitudes of the wave functions for paths a1 and a2 versus position. The wave
functions have a basic Gaussian profile with oscillations arising from the truncation
of one Gaussian tail by W A .
Appendix D: Detailed Analyses of Selected Gedankenexperiments 209
Figure D.10 shows the corresponding probabilities for α = 0 (e.g., fully entan-
gled) of coincident photon pairs at Alice’s detector D A and at Bob’s detectors D B1
and D B0 . The probabilities are highly oscillatory because of the interference of the
two waves and the phase walk of the wave functions with angle, analogous to two-slit
interference.
To test the possibility of a nonlocal signal, we must integrate these probabilities
over the extent of the detector face and calculate difference functions from these
results and similar evaluations of Eqs. D.37 and D.38. We can expect some errors
in numerical integration due to the oscillation shown in Fig. D.9. The difference
functions as 2-D contour plots in φ B versus α are shown in Fig. D.10.
Thus, the differences between the probabilities predicted by of Eqs. D.37 and D.38
and the numerically-integrated probabilities of Fig. D.10 are on the order of a few
parts per million. Does this mean that there is a small residual nonlocal signal? No!
It means that any comparison involving numerical integration is subject to round-
off error and is not reliable beyond a few parts per million. The results shown in
Fig. D.10 demonstrate that no nonlocal signal is possible using the wedge-modified
configuration of Fig. D.7.
210 Appendix D: Detailed Analyses of Selected Gedankenexperiments
Fig. D.10 Difference between numerical singles probabilities and evaluations of Eqs. D.37 and
D.38. Here the regions labeled “A” reach minima of 5.7 × 10−7 , the regions labeled “B” reach
maxima of 6.08 × 10−6 , and the regions labeled “C” reach maxima of 5.51 × 10−6 . Small blotches
indicate regions in which numerical integration has produced larger errors
Appendix D: Detailed Analyses of Selected Gedankenexperiments 211
References
30. A.C. Elitzur, S. Dolev, Multiple interaction-free measurements as a challenge to the transac-
tional interpretation of quantum mechanics, in AIP Conference Proceedings Frontiers of Time:
Retrocausation—Experiment and Theory, vol. 863, ed. by D. Sheehan (2006), pp. 27–44
31. N. Herbert, Am. J. Phys. 43, 315 (1975)
32. M.A. Horne, A. Shimony, A. Zeilinger, in Sixty Two Years of Uncertainty, ed. by A.I. Miller
(Plenum Press, 1990)
33. A.F. Abouraddy, M.B. Nasr, B.E.A. Saleh, A.V. Sergienko, M.C. Teich, Phys. Rev. A 63,
063803 (2001)
34. H. Everett III, Rev. Mod. Phys. 29, 454 (1957). See also [36]
35. A.P. French, P.J. Kennedy (eds.), Niels Bohr, A Centernary Volume (Harvard University Press,
Cambridge, 1985)
36. J.A. Wheeler, Rev. Mod. Phys. 29, 463 (1957). See also [34]
Index
L
I Lamb shift, 52
I, 47 Left circular polarization, 86
Identity, 24 Leggett, Anthony J., 125
Indeterminism, 24 Leggett-Garg inequalities, 125
Indiana University, 3 Lewis, Gilbert N., 77
Institute for Advanced Studies, 28 Lewis, Peter, 68, 108, 186
Integer spin, 172 Light polarization, 30
Intensity interferometry, 92 Light waves, 10
Interaction-free measurement, 100, 117, LiIO3 crystal, 114
118, 120 Linear polarization, 30, 86, 171
Interference, 173 Local hidden-variable theories, 33
Interference complementarity, 142, 144 Localization, 190
Interference suppression, 78 Localized waves, 20
Internet Encyclopedia of Philosophy, 186 Logical positivism, 8, 17, 24, 58
Interpretation, 57, 66 Loopholes, 43
Interpretational problems, 6, 19, 24, 168 Low-intensity two-slit experiment, 76
216 Index
M Parity nonconservation, 3
Mach, Ludwig, 101 Parity transformation, 171
Magnetic flux quanta, 157 Particle in a box, 14
Malus’ law, 31, 34, 106, 125 Particle-like, 172
Matrix diagonalization, 17 Particle scattering, 18
Matrix element, 17 Path-entangled EPR, 139
Matrix inversion, 17 Path entanglement, 114
Matrix mechanics, 17, 18 Path label, 76
Maudlin gedankenexperiment, 67, 108, 182 Path labeling, 78
Maudlin, Tim, 67, 108, 182 Pauli exclusion principle, 50, 171
Maxwell, James Clerk, 10 Pauli, Wolfgang, 14, 41
Maxwell’s equations, 1 Peierls, Rudolf, 26
Mead calculation, 68 Penniston, Penny, 6
Mead, Carver, 66, 110, 179, 182, 186 Pflegor-Mandel experiment, 94
Measurement, 24 Philosophical economy, 65
Measurement problem, 24, 26, 70, 83, 186 Photoelectric effect, 10
Modern physics, 4 Photon, 10, 13, 19, 22, 27, 29, 30, 32, 40, 60,
Momentum entanglement, 114 67, 77, 78, 83, 86, 88, 91, 94, 96, 101,
Momentum operator, 192 170, 172
Muon, 171 Photon cascade, 32, 91
MWI, 177 Photon polarization, 37, 171
MZ interferometer, 101 Photon scattering, 40
Pilot-wave model, 5
Pion, 94
N Planck’s constant, 10, 170
Narlikar, J., 5 Planck, Max, 10
National Security Agency, 155 Podolsky, Boris, 28
Neutrino, 41, 171 Polarization basis, 30, 37
Newton’s 2nd Law, 57 Polarization correlations, 91
No self-interaction, 51 Polarization-entangled EPR, 138
No-signal theorems, 135 Polarization transformations, 87
Non-commuting polarizations, 86
Position-momentum complementarity, 22
Non-commuting variables, 85
Positivism, 8, 17, 24, 58, 70, 75, 131
Nonlocality, 5, 6, 24, 28, 29, 39, 42, 91, 152,
Prime numbers, 158
167, 174
Princeton University, 28
Nonlocal realistic theories, 124
Pseudo-time, 64, 178
Nonlocal signal, 37, 114, 115, 135, 174
Norman, Eric, 4
Q
O QM formalism, 173
Observables, 25 Quantization, 170
Observer-created reality, 27 Quantum computer, 155
Observers, 70 Quantum computing, 156
Occam’s Razor, 158, 181 Quantum dots, 157
Offer-confirmation echo, 63 Quantum eraser experiment, 98
Offer wave, 63 Quantum field theory, 54, 67, 135, 136, 179,
Operator, 173, 191 182
Quantum gates, 157
Quantum interpretations, 59, 174
P Quantum jumps, 13, 67
Paradox, 6, 58 Quantum liar paradox, 124
Parallel universes, 158, 177 Quantum mechanics, 6, 169
Parity, 3, 29, 31, 32, 50, 162, 171, 186 Quantum mystery, 75, 77
Index 217