Quantum Physics
Quantum Physics
With the study of quantum dynamics, we move from the world of spacecrafts and
rushing galaxies to that of subatomic elements: "Quantum theory arose from an attempt to
explain phenomena that lay beyond the scope of conventional classical physics. A central
failure of classical mechanics was its inability to account for the structure of atoms.”1
Until the end of the 19th century is was thought that the atom was the smallest
indivisible unit of matter. This view was challenged by J.J. Thompson’s discovery in 1897 of
the existence of electrons within the atom. While doing experiments with cathode rays
Thompson discovered that these rays were composed of very small, negatively charged
particles, which he called ‘electrons’. On this basis, he imagined the atom to consist of a
number of electrons swimming in a sea of positive electric charge, so that the whole looked
something like a raisin pudding. In a further development, Ernst Rutherford in 1911 identified
the atom’s nucleus. He did so by bombarding an atom with heavy alpha particles, and from
the manner in which the incident particles were deflected from the atom he deduced “that the
bulk of the mass in the atom was concentrated in a tiny ‘nucleus’ of positive electric charge,
with the electrons of negative charge distributed around it at assorted distances.”2 He termed
the massive particles that composed the nucleus ‘protons’, and also reckoned with the
existence, in the nucleus, of hypothetical particles of similar mass to the proton but
possessing no electric charge. These particles, the neutrons, were discovered by James
Chadwick in 1932.
Rutherford’s model of the atom was that of a solar system in miniature. But whereas in
Newton’s concept the planets orbited the sun under the impact of gravity, in Rutherford’s
model the electrons rotate around their nucleus on the basis of electrostatic forces. Unlike the
solar system in which all the planets orbit the Sun on the same horizontal plane, each of the
orbiting electrons has its own plane of circulation: horizontal, vertical and various
intermediate directions. This view, however, proved to be too simplistic. Indeed, in his study
of electromagnetism, Maxwell had shown that “any electrically charged matter that would
move non-uniformly must emit radiation, thereby decreasing its own energy.” 3 So, as the
radiating energy of a revolving electron continuously decreased, the electron should follow a
spiral path and fall into the nucleus. This atomic model could, thus, not account for the
stability of the atom.
In 1913 the Danish physicist Niels Bohr (1885-1962), who, as a young doctoral
student had worked in association with Rutherford in Great Britain, offered a solution to the
problem by integrating Max Planck’s new insight into the packet character of light radiation
into the study of the subatomic particles.
In the 1890s Max Planck (1858-1947) was engaged in the study of thermal radiation,
i.e., of the light radiation emitted from the surface of an object due to the object’s
temperature. The ideal thermal emitter is known as a black body, and the radiation it emits,
when heated, is called black body radiation. When Planck began his study of black body
1
Jonathan Halliwell, “Quantum Cosmology and the Creation of the Universe,” in Scientific American, December
7, 1991.
2
John Barrow, The World Within the World (Oxford/New York: Oxford University Press, 1990),171.
3
Mendel Sachs, Einstein versus Bohr: The Continuing Controversies in Physics (La Salle, Illinois: Open Court,
1988), 71.
99
radiation, it was already known that the amount of radiation did not depend on the material of
the heated object but only on the latter’s temperature. Yet, it was still unclear as to how the
frequency of radiation related to the temperature and hence to the total energy of the black
body: “In 1900 Max Planck suggested that light, X rays, and other waves could not be emitted
at an arbitrary rate, but only in certain packets, called quanta. Moreover, each quantum had a
certain amount of energy that was greater the higher the frequency of the waves.” 4 This lead
to the equation that would make him famous: E = hv ; E stands for the total energy of the light
source, v is the frequency of the light radiation, and h is a mathematical constant, called
‘Planck’s constant’. Planck’s constant h expresses the packet character of the emitted
radiation. The radiating energy of a system cannot take on any value. Rather, it can take on
only particular discrete values. It is “as if nature allowed one to drink a whole pint of beer or
no beer at all, but nothing in between.”5
This principle sets the tone for Niels Bohr’s atomic model and the restrictions it
imposed: orbiting electrons cannot be located at any distance from the nucleus; they can only
travel in special orbits: at a certain discrete set of distances from the nucleus with specific
energies. “Bohr’s rules required that the angular momentum of electrons in orbit about the
nucleus can occur only in integer multiples of h/2π,6 for which Dirac introduced the
convenient symbol ħ (pronounced h bar); that is: ħ = h/2π. Thus the only allowed values of
angular momentum (about any one axis) are 0, ħ , 2ħ, 3ħ, 4ħ etc.”7
Bohr was apparently impressed by the way in which Planck ‘quantized’ the light
radiation with the help of his constant h: only discrete energy packets (quanta) can be emitted
from the black body. So, too, Bohr emphasized that only specific distances from the nucleus
qualify as orbits for the rotating electrons. Only at those distances is it possible for electrons
to develop the exact energy and velocity that permanently keeps them on their orbit.
4
Stephen Hawking, A Brief History of Time. From the Big Bang to Black Holes, (London: Bantam Books, 1988
[1989]), 58.
5
Brian Greene, The Elegant Universe (New York: Vintage Books, 2000), 94.
6
Recall: π is 3. 1415927… ‘Pi’ obtains if one divides the circumference of a circle by its diameter.
7
Roger Penrose, The Emperor's New Mind. Concerning Computers, Minds, and the Laws of Physics (Oxford:
Oxford University Press, 1990), 231.
100
In a new move Bohr elaborated the shell-structure of the atom. The number of
electrons in an atom is determined by the number of protons in its nucleus: A nitrogen atom,
for example, has a nucleus made up of seven protons and seven neutrons; it is surrounded by
seven electrons. For Bohr, the number of electrons an atom possesses is spread over various
shells or energy levels. Bohr admitted of seven successive shells or energy levels, those being
closest to the nucleus (but at a considerable distance from it) having the lowest energy level.
Each shell has an upper limit to the number of electrons it can accommodate: the first shell is
complete with two electrons (resulting in a helium atom), the second is complete with eight
electrons, and the other successive shells have still larger accommodation capacities. As a
rule, a lower level must be totally filled before a higher shell can begin harboring electrons.
Take the nitrogen atom with its seven electrons. Two of these electrons are in the first shell,
whereas the remaining five are in the second shell. This rule of repartition holds for the first
18 elements in the periodic table of Mendeljev, but starting from the 19th element (the atom
with 19 electrons) certain levels can be left incomplete. It is important to notice that the ‘last’
electrons – those in the outer shell – determine the chemical properties of an atom. This is e.g.
the case with uranium which has 92 electrons.
There is still one important element to be mentioned: the electrons can gain or
lose energy by jumping from one allowed orbit to another, absorbing or emitting an
electromagnetic radiation with a frequency ν determined by the energy difference of the two
levels. This is technically called a quantum leap. When a nitrogen atom, e.g., is bombarded
with photons (i.e., quanta of electromagnetic radiation), one of its electrons may absorb the
photons’ energy and jump from its original energy level up to a higher energy level. This
jump leaves an empty space in the lower energy level, which renders the atom unstable: the
atom is then in an excited state. Yet, since the electrons always seek to get back to their most
stable state in the atom, namely their ground state, they will eventually relinquish the newly
acquired energy: the excited electron will drop back to its ground state by emitting a photon.
It is this phenomenon of ‘ejection’ that Einstein came across with his study of the
photoelectric effect, although in that case an electron was knocked out of the atom.
In the second half of the 1920s, the Copenhagen school became renowned for its
method of statistical measurement. This method laid bare some unexpected features: the
phenomena in the quantum world seemed to behave in a random, a-causal fashion, so that
they ought to be described “in terms of a new probability calculus, called ‘quantum
mechanics’”8 More particularly, the subatomic particles sometimes behaved as wave
undulations and sometimes as packets of quanta (the so called “wave-particle duality”). The
fact that light (photon) has both wave-like and particle-like properties was still easy to
understand, but there were strong indications that matter particles, too, such as electrons, also
had wave-like properties.9
The wave character of particles is clearly shown in the double-slit experiment that
was originally performed by Thomas Young in 1801: “Consider a partition with two parallel
narrow slits in it. On one side of the partition one places a source of light of a particular color
(that is, of a particular wavelength). Most of the light will hit the partition, but a small amount
will go through the slits. Now suppose one places a screen on the far side of the partition from
the light. Any point on the screen will receive waves from the two slits. However, in general,
the distance the light has to travel from the source to the screen via the two slits will be
8
Mendel Sachs, Einstein versus Bohr, 82.
9
This extension had been proposed by the French physicist Louis de Broglie in 1924.
101
different. This will mean that the waves from the slits will not be in phase with each other
when they arrive at the screen: in some places the waves will cancel each other out, and in
others they will reinforce each other. The result is a characteristic pattern of light and dark
fringes.”10 The light and dark fringes are the result of interference of waves: waves whose
crests and troughs run parallel fortify each other, whereas waves whose crests and troughs do
not run parallel, cancel each other. Yet, other experiments reveal the weird character of the
world of the quanta. If one replaces the source of light (the bulb) with a bunch of electrons,
the result will be really mind-boggling. In that case electrons sent one after the other to the
partition with the two slits would behave as if one single electron went through both of the
slits at once so as to produce the light and dark fringes, typical of interference. In other words,
a single electron fired at a time behaves as if it was interfering with itself in order to yield the
interference on the screen.
10
Stephen Hawking, A Brief History of Time, 63.
11
Paul Davis, God and the New Physics (London: Penguin Books, 1983), 103.
102
place their own accents. Bohr launched the complementarity principle according to which an
experiment can demonstrate the particle-like properties of matter, or its wave-like properties,
but not both at the same time, whereas Heisenberg made headlines with his uncertainty
principle. He formulated this principle for the first time in 1926: if one measures a particle's
position, then its velocity becomes uncertain; if one measures its velocity, its position
becomes unpredictable. For Heisenberg, this uncertainty does not flow from some inaccuracy
in our measure instruments; “it is truly intrinsic to nature."12
Indeed, in order to determine the behavior of an orbiting electron about its
nucleus, one will have "to measure its present position and velocity accurately."13 Now, the
obvious way to do this is to shower the particle with an amount of energy, which can be done
experimentally with light photons. When some of the light waves are scattered, this indicates
that the electron got hit, and its position is detected. One, however, also observes that, owing
to the added amount of energy, the particle abruptly changed its velocity. So, no prediction
about its future behavior can be made in spite of, or precisely due to, the measurement.
Moreover, if one would try to measure the particle's position more accurately, one would need
to bombard it with light of shorter wavelength (the particle is detected then between the
shorter wave crests). But this procedure still aggravates the situation: "[for] the more
accurately one measures the position, the shorter the wavelength of the light that one needs
and hence the higher the energy of a single quantum [that is going to hit the electron]. So the
velocity of the particle will be disturbed by a larger amount. In other words, the more
accurately you try to measure the position, the less accurately you can measure its speed, and
vice versa."14 Measurements usher in the randomness that is so typical of the quantum world:
Observed particles just pop up, reveal part of their being, and disappear again by changing
their velocity and wavelength.
When looking back at his earlier achievement, Bohr must have been shocked to
have to admit that, from the standpoint of the new quantum theory, it becomes impossible to
simultaneously determine an electron’s location and its rotational velocity (its momentum).
Uncertainty replaces precise prediction. That is the reason why, in the new quantum
mechanical approach, instead of using the term ‘orbits’, one prefers to speak of ‘orbitals’, or
‘clouds’ of possible locations of the electrons.
The only real information one gets from the sub-atomic ghost world is through
experiments. Yet, it is impossible to accurately predict what one is going to observe. This
creates the need for engaging in probability calculus. "In general, quantum mechanics does
not predict a single definite result for an observation. Instead, it predicts a number of
different possible outcomes and tells us how likely each of these is. That is to say, if one
made the same measurement on a large number of similar systems, each of which started off
in the same way, one would find that the result of the measurement would be A in a certain
number of cases, B in a different number, and so on. One could predict the approximate
number of times that the result would be A or B, but one could not predict the specific result
of an individual measurement."15 This probability calculus is practiced in the various schools
of quantum mechanics, including the Copenhagen school. But in 1926 the Austrian physicist
Erwin Schrödinger developed a model of particle wave function that, in a rather classical
12
Ibid., 102.
13
Stephen Hawking, A Brief History of Time, 58.
14
Ibid.,59.
15
Ibid.,60.
103
16
Roger Penrose, The Emperor's New Mind. 227.
17
Ibid., 243.
18
Brian Greene, The Fabric of the Cosmos. Space, Time and the Texture of Reality (London/New York: Penguin
Books, 2004), 208-209.
19
Roger Penrose, The Emperor's New Mind, 251.
20
Ibid., 250.
104
merely a calculational procedure, and does not attempt to describe the world as it actually ‘is’.
[Yet, Schrödinger] attributes objective physical reality to the quantum description – the
quantum state.” 21 It is this ‘reality-geared’ focus that motivated Schrödinger to underline the
deterministic character of a particle’s wave function in a probabilistic context.
The ‘collapse of the wave function’ is a case in point of what is at stake in the
controversy. In the wave function collapse, the initial superposition of several different states
appears to be reduced to a single one of those states after interaction with an observer. Indeed,
whenever one makes a ‘measurement’ one magnifies the quantum effects to the classical level
of observation: one jumps from the level of probability amplitudes, in which the mysterious
interference happens, to the magnified level of observation, where the interference stops and
the wave function collapses. Penrose illustrates this with the help of a thought experiment that
slightly modifies the double-slit scenario by placing a particle detector at one of the slits:
“Since when it is observed”, he writes, “a photon – or any other particle –always appears as a
single whole and not as some fraction of a whole, it must be the case that our detector detects
either a whole photon or no photon at all. However, when the detector is present at one of the
slits, so that an observer can tell which slit the photon went through, the wavy interference
pattern at the screen disappears. In order for the interference to take place, there must
apparently be a ‘lack of knowledge’ as to which slit the particle ‘actually’ went through.”22
This again shows that one ought to regard the particle/wave as being spread out spatially,
rather than being always concentrated at single points. Only in this spatially ‘spread-out’
view does it make sense to speak of the evolution-in-time of a wave function; only then are
we able to grasp that a ‘particle’ can pass through the two slits at once, or that pairs of
particles may simultaneously interact the one with the other even when they are light years
apart.
During his whole life time Albert Einstein refused to accept the premises of the
Copenhagen school. For him, God ‘does not play dice’; there must be a more stringent order
of cause and effect in the universe than Bohr and Heisenberg were willing to admit. In order
to refute Bohr’s interpretation that the specific behavior of the quantum world did not exist
before it was observed, he set up with his associates Boris Podolsky and Nathan Rosen the
paradox that is known after their initials: the EPR-paradox; it was published in 1935 with the
title “Can Quantummechanical Description of Physical Reality Be Considered Complete?”
“The essence of the EPR experiment is to examine the decay of an unstable
elementary particle (an electrically neutral one, for example) into two photons. When the
decay occurs, the photons will move off in opposite directions with equal momenta [i.e. with
the same velocity] in order that the total momentum is conserved. These photons will also
both possess an intrinsic spin,23 and if one spins clockwise then the other must spin
anticlockwise in order to conserve the total spin of the system during the decay process. By
considering the complementary attribute to this spin, EPR were able to produce what they
regarded as a paradox that demonstrated that quantum theory could not be a complete
description of what was occurring. When the particle decays we cannot predict which residual
photon will be found spinning in the clockwise direction. There is an equal probability for
either the clockwise or the anticlockwise direction. But even if we let them fly apart to
opposite sides of the universe, as soon as we measure the spin of one of them we should
21
Ibid., 226.
22
Ibid., 236.
23
For a detailed analysis of the spin notion, see the next section.
105
know, without any shadow of doubt, that the spin of the other will be found to be equal and
opposite to the one we have just measured. This we will know without having measured it.
Thus, EPR claims that the unmeasured spin of the other photon must be regarded as real
according to their definition of physical reality, because it is predictable. It must, therefore, be
related to some element of an observer-independent reality.”24
EPR made a point. It is possible to determine (without measuring it) the spin
value of photon B moving with the same velocity in the opposite direction by only measuring
the spin value of photon A; for on the basis of the conservation of the total spin of the system
one can tell that when photon A has a clockwise spin, photon B must have an anticlockwise
spin, and vice versa. So, there is something objective in the quantum world, which is
observer-independent. Yet, EPR also points to an inherent problem: how can photon B know
that photon A took on a clockwise spin during measurement and accordingly adopt a
counterclockwise spin? How can photon B seemingly simultaneously react to the new event,
even when it finds itself at the other side of the universe? Such an instantaneous interaction
clashes with the relativity principle that nothing can travel faster than the speed of light,
300,000 km/sec.
In the meantime, however, this instantaneous communication, technically called
‘quantum nonlocality’ has been confirmed: “It looks as if spatial distances do not exist.
Einstein and his associates claimed that such a nonsensical result demonstrates the fact that
quantum theory is not a complete theory. Yet, what in Einstein’s time could only be a thought
experiment, nowadays, with our present technology, has been many times performed. And it
has turned out that the ‘crazy’ prediction of quantum mechanics is correct. ‘Quantum
nonlocality’, as it is now called, is an empirical fact, and physicists have to live with it […] It
is only our intuition that cannot easily accept that two elementary particles can interact with
each other with no mediation of space and time distances separating them.”25
When commenting on the EPR paradox, Penrose comes to the same conclusion.
For him too, it is evident that “the measurement of an entangled system [i.e. of photon A and
its twin photon B] sits extremely uncomfortable with the requirements of relativity, since a
measurement of one part of an entangled pair would seem to have to affect the other
simultaneously, which is not a notion that we ought to countenance if we are to remain true to
the relativistic principles.”26 What makes quantum physics incompatible with Einstein’s
theory of special relativity is precisely the phenomenon of entanglement. Quantum mechanics
conceives of two photons moving outwards in space as “photon pairs, acting as a single unit.
Neither photon individually has an objective state: the quantum state applies only to the two
together.”27 It is this entanglement that enables the twin pair to instantaneously interact when
one of them is ‘forced’ by a measurement to take on a certain spin value. All this happens as
if the photon pair, even when they are light years apart, was not in the least encumbered by
the classical notion of space-time. Entanglement of twin photons is an element that Einstein
could not have foreseen.
24
John Barrow, The World within the World, 146.
25
Michael Heller, “Where Physics Meets Metaphysics,” in Shahn Majid (ed.), On Space and Time (Cambridge:
Cambridge University Press, 2008), 260.
26
Roger Penrose, The Road to Reality. A Complete Guide to the Laws of the Universe (London: Jonathan Cape,
2004), 593.
27
Roger Penrose, The Emperor's New Mind, 287.
106
Up till now we have mentioned only the elementary particles electron, proton,
neutron, and photon. Yet, the more new particles were discovered, the stronger the need to
classify them. The earliest classification was that of leptons (leptos in Greek means ‘light’,
not heavy) and baryons (barus in Greek means heavy). The electron is a lepton, whereas
proton and neutron fall under the category of the baryons. Leptons cannot further be divided
in subcomponents, but baryons can. As we will see below, proton and neutron (which are
baryons) are made up of quarks. That is the reason why recently one has begun to refer only
to two types of basic constituents of mater: leptons and quarks. Ever since the discovery of
quarks each composite particle made up of quarks and held together by the strong force is
called a hadron (hadros in Greek means bulky). Hadrons are categorized into two families:
baryons (see above) made up of three quarks, and mesons (from the Greek mesos: medium)
made up of one quark and one antiquark. The best-known baryons are protons and neutrons;
the best-known mesons are the pions and kaons, which were discovered in cosmic ray
experiments in the late 1940s and early 1950s.
The above classifications mainly deal with matter particles, with the neglect of
energy particles. Leptons, quarks, baryons and mesons, are matter particles, also termed
fermions, named after Enrico Fermi. But besides them, there also exist energy particles or
particles that are carriers of forces (the electromagnetic force, the strong and the weak nuclear
force, and gravity). The force carrying particles are termed bosons, derived from the name of
Satyendra Nath Bose. All the fundamental particles in nature fall into these two categories:
they are either fermions or bosons. The table below lists the differences.
A look at the table shows that the basic difference between fermions and bosons
consists in their specific spin. Spin is a measure of a particle’s rotation about its axis; it is an
intrinsic property of the particle (i.e. not arising from some orbital motion about a center).
Fermions have half-integral spin, whereas bosons have integral spin, that is: fermions,
whatever they are doing, perform time and again half a rotation about their axis, whereas
bosons perform time and again a whole rotation about their axis.
There is still another striking feature. A particle’s spin either takes on a ‘spin up’
value (rotating upward anti-clockwise) or a ‘spin down’ value (rotating downward clockwise).
We have seen this phenomenon at work in the EPR paradox: there it became evident that in
the case of two photons moving off in opposite directions one had to have a clockwise spin
(‘spin down’) and the other an anticlockwise spin (‘spin up‘). A differentiation like this not
only occurs with bosons (a photon is a boson); it is also typical of the fermions, for which the
divide between ‘spin down’ and ‘spin up’ is, in a sense, required to make the Pauli Exclusion
Principle work.
The Pauli Exclusion Principle was formulated by the Austrian physicist Wolfgang
Pauli in 1925. It states that no two identical fermions (with half-integral spin) may occupy
the same quantum state simultaneously. The insight into this principle flows from statistic
data. Bosons have the tendency to clump into the same quantum state, thus fortifying the
107
strength of their common effect, as this is the case with laser beams, whereas fermions behave
as if they were forbidden from sharing identical quantum states, and thus, from blending.
Pauli used the Exclusion Principle to explain the arrangement of electrons in an
atom. In the quantummechanical approach to atoms, the space surrounding the dense nucleus
is thought of as consisting of orbitals, or regions in which it is most probable to find electrons.
According to the Pauli Exclusion Principle, no two electrons with the same quantum state can
exist within the same orbital. This implies that each electron in an orbital has a unique
quantum state, described by a unique set of quantum numbers. The principle quantum number
indicates the energy level of the orbital, a second number represents the shape of the orbital; a
third number indicates the orbital angular momentum, and a fourth indicates the particle’s
spin direction (‘up’ or ‘down’).
Two electrons with the same energy level and orbital angular momentum can, to
be sure, exist in the same orbital, but it will be necessary that they have opposite spins: one
with ‘spin up’ the other with ‘spin down’. The opposite spin differentiates them, even when
for the rest they share the same properties. In this way, the Pauli Exclusion Principle keeps the
distinct electron waves apart, not allowing them to merge. If such a merger were to happen,
the atom in question would degenerate and fall apart. The Exclusion Principle applies to all
fermions, and thus also to quarks: “If the world had been created without the exclusion
principle, quarks would not form separate, well-defined protons and neutrons. Nor would
these, together with electrons, form separate well-defined atoms. They would all collapse to
form a roughly uniform, dense ‘soup’.”28 In other words, without the Exclusion Principle no
distinct forms of matter would ever have come into existence; the universe would still be an
undifferentiated ‘soup’ as it most probably was ‘in the beginning’, right after the Big Bang.
We cannot conclude this section without expounding on the nature and properties of
the force-carrying particles: the bosons with spin 0, spin 1 and spin 2. They all regulate the
interactions between matter particles, and do not obey the Pauli Exclusion Principle. So, they
can pile up and increase their strength. “Force-carrying particles can be grouped into four
categories according to the strength of the force that they carry and the particles with which
they interact.”29
(i)The gravitational force, which is located in the gravitational field, where it acts
between matter particles. This force is rather weak, “but it can act over large distances, and it
is always attractive […] it is pictured as being carried by a particle of spin 2, called the
graviton;” the existence of the graviton, however, is not yet experimentally verified. (ii) The
electromagnetic force, which interacts with electrically charged particles. It “is pictured as
being carried by massless particles of spin-1, called photons.” (iii) The weak nuclear force,
which is responsible for radioactivity; it is carried “by three spin-1 particles, known
collectively as massive vector bosons […] These bosons are called W+ (pronounced W plus),
W- (pronounced W minus), and Zo (pronounced Z naught).” Because these particles are so
massive their action radius is extremely short. (iv) The strong nuclear force, finally, which
holds the quarks together in the proton and neutron and holds the proton and the neutron
together in the nucleus, is of a comparably short range: “It is believed that this force is carried
by another spin-1 particle, called the gluon, which interacts only with itself and with the
quarks.”30 This is why the more one would try to separate the quarks, in a neutron e.g., the
more forcefully the gluon would react in keeping the quarks together.
28
Stephen Hawking, A Brief History of Time, 72.
29
Ibid., 73-74.
30
Ibid., 74-77.
108
The prediction that each particle would be accompanied by an anti-particle was made
by the British physicist Paul Dirac (1902-1984). In 1928 he elaborated an extension of the
Schrödinger equation, a new wave equation that is consistent with special relativity.
( i ) In retrospect, it is easy to understand that the existence of antiparticles logically
flows from Einstein’s formula E= mc2. Nowadays we know that if a particle is hit by its
antiparticle, they annihilate each other so that energy is released in the form of photons.
Conversely, “if sufficient energy is introduced into a system, localized in a suitably small
region, then there arises the strong possibility that this energy might serve to create some
particle together with its antiparticle.”31 Yet, the E= mc2 formula alone cannot give us a clue
as to why for every kind of particle there is a corresponding antiparticle with opposite electric
charge. It is here that quantum physics comes into play. According to this theory, a particle-
antiparticle pair can only be created on condition that there is no violation of charge
conservation: when a particle has a specific quantity of negative electric charge, e.g., its
antiparticle must have the same quantity of positive electric charge.
(ii) I have given this retrospective view in order to show how difficult it is to
harmonize quantum mechanics with Einstein’s special relativity. Dirac had to wrestle a great
deal with such intrinsic difficulties. In his attempt at bringing the Schrödinger equation in line
with special relativity, he met with an unsettling problem. The new equation made it clear that
for each quantum state with positive energy E, there was a corresponding state with energy
‘minus E’. In Einstein’s approach, there is no place for negative energy, that is: for an energy
state below zero. Yet, “in quantum mechanics, one has to consider that the various possible
things that ‘might’ happen, in a physical situation, can all contribute to the quantum state […]
So, even an ‘unphysical negative energy’ has to be considered as a ‘physical possibility’.”32
Facing up to this difficulty, Dirac set out to imagine a huge pool of negative-energy
electron states, and then wondered what would happen if a positive-energy electron33 would
tumble into this pool. Such a positive-energy electron would without doubt be absorbed into
the ‘sea’ of negative-energy electron states, and the more it would immerse in it, the more it
would shed its positive energy by emitting photons. Yet, if an unlimited number of electrons
were to do the same, a catastrophic instability would arise as a result of the limitless release of
energy. But apparently such instability has never occurred. To prevent this situation from
31
Roger Penrose, The Road to Reality, 610-611.
32
Ibid., 615.
33
Note that an electron with negative electric charge has positive energy.
109
happening, Dirac hypothesized that a ‘sea’ of negative-energy electrons fills the universe in
such a way that all of the lower-energy states were occupied. So, in line with the Pauli
Exclusion Principle, no other electron could fall into the ‘sea’. What remained possible,
however is, that one of the negative-energy particles could be lifted out of the ‘sea’, thus
leaving behind a hole that would act like a positive-energy electron with a reversed charge.
If then, Dirac continued, a negatively-charged electron would fall into the ‘hole’, the
effect of it would be flabbergasting, “it would result in the ‘hole’ and the electron annihilating
one another, in the manner that we now understand as a particle and its antiparticle
undergoing mutual annihilation,”34 leading to a bursting forth of energy in the form of
photons. Dirac’s ‘hole’ is indeed the electron’s antiparticle.
Initially Dirac thought that the ‘hole’ could be a proton. When he was conducting his
research, it was commonly held that protons were the only particles thought to have a positive
electric charge. This guess turned out to be erroneous, since the mass of the proton is about
1836 times larger than that of the electron, and the mass of a particle and an antiparticle,
respectively, had to be equal. That’s why Dirac in 1931 termed the predicted particle an anti-
electron, an unknown particle that we now call a positron, the existence of which was, the
very next year, effectively discovered by Carl Anderson.
In the previous pages our primary focus has been on Quantum Mechanics. This branch
of quantum physics refers to a system in which the number of particles is fixed and the fields,
such as the electromagnetic field, are continuous classical entities. Quantum Field Theories
(QFTs), however, move beyond this fixed view in that they analyze how particles are
incessantly being created and annihilated
Dirac can be regarded as a precursor of QFTs because he already reckoned with the
impact of electric fields on the formation of quanta. Yet, it is only a reformulation of his
equation that brings us into the proper domain of QFTs. Dirac still saw his equation as
depicting the evolution-in-time of one single electron, whereas, in fact, it should be read as
the evolution-in-time of a whole quantum field in which operators cause the creation and
annihilation of particles. Operators are mathematical entities that are defined by how they act
on something; they bring about transformations in the states of the field, for example, excite it
in a particular mode at a particular point x. Depending on these excitations particles can be
created or annihilated.
Quantum Field Theories inaugurate a new era of research in particle physics, as well
as establishing the basis for further rapprochement between quantum physics and special
relativity. This rapprochement can be seen, e.g., in the central position the notions ‘symmetry’
and ‘invariance’ occupy in the theory formation of QFTs: “Symmetries are perhaps the most
important objects in all of mathematics.”35 In geometry, an object is symmetric with respect to
a given operation, if this operation, when applied to the object (a geometrical shape), does not
appear to change it. There are various ‘symmetry operations’ that can be performed, such as
translation (which ‘slides’ an object from one area to another by a vector), or rotation around
a fixed central point x. In physics, however, symmetry has been generalized to mean
invariance – that is, lack of change — under any kind of transformation.
In QFTs the following symmetry operations are practiced: ( i ) C, which stands for
charge conjugation: “This operation replaces every particle by its antiparticle [... ] A physical
34
Roger Penrose, The Road to Reality, 625.
35
Shahn Majid, “Quantum Spacetime and Physical Reality,” in ID (ed.), On Space and Time (Cambridge:
Cambridge University Press, 2008), 80.
110
interaction that is invariant under the replacement of particles by their antiparticles (and vice
versa) is called C-invariant. (ii) P, which stands for parity: This operation replaces a particle
by its mirror image (reflection in a mirror). A physical interaction that is invariant under the
replacement of a particle by its mirror image is called P-invariant. (ii) T, which stands for time
reversal. “An interaction is invariant under T if it is unaltered if we view it from the
perspective of a time direction that is the reverse of normal”, 36 i.e., when time runs back. In
the three cases, one has to do with heuristic tools allowing the researchers to investigate
whether, in a particular domain, a certain basic property of the system is, in actual fact,
invariant. Moreover, the study of symmetry operations may reveal that under certain
circumstances a symmetry break must have taken place (see below: Electroweak Theory).
The original title Einstein planned for his special theory of relativity was ‘invariant
theory’. In this theory, the invariant par excellence is the speed of light (300,000 km/sec). The
speed of light is always the same, regardless of the time of measurement, regardless of the
speed of the light source, and regardless of the speed with which the platform moves from
which the speed of light is measured. Einstein’s special theory of relativity is based on the
implementation of symmetry operations. The type of symmetry he sought to establish is
referred to as ‘global symmetry’, that is: a symmetry in which the basic property (the speed of
light) is invariant under spatio-temporal transformations at large.
Quantum field theories, on the contrary, are ‘local symmetries’. Whereas global
symmetries “were invariances that existed when you changed things in the same way
everywhere, local gauge symmetry requires invariance after you have done different things in
different places at different times.”37 In short, local gauge symmetries are predicated on the
idea that symmetry transformations are performed locally in a particular domain of space-
time. This allows for diversification, but also for the imposition of strict conditions. In
Maxwell’s equations, e.g., the local gauge symmetry “forbids the existence of a photon
possessing a mass, and it dictates the precise way in which electrically charged particles
interact with light.”38
So, too, in Quantum Field Theories, “the local gauge symmetry dictates what
mediating forces must exist between the particles involved.”39 These mediating forces are all
bosons, particles with integer spin that are, with the exception of W+ and W-, themselves
their own antiparticles. In Quantum Electrodynamics electrically charged particles interact
with each other through an exchange of photons. For the weak nuclear force the mediating
bosons are W+, W-, and Z0, whereas in the strong nuclear force gluons mediate between the
quarks. For the force of gravity, finally, the graviton is postulated as the mediating boson.
Similarly special mathematical symmetry groups are operational in the various Field
Theories: the symmetry group U (1) in Quantum Electrodynamics; the symmetry group SU
(2) in the Weak Nuclear Force, and the symmetry group SU (3) in the Strong Nuclear Force.
Quantum Electrodynamics (QED) was the first success story of Quantum Field
Theory; it is an Abelian40 gauge theory that operates with the symmetry group U(1). QED
gives a highly satisfactory quantum description of the behavior of electrically charged
particles in an electromagnetic field, including the creation and mutual annihilation of
36
Roger Penrose, The Road to Reality, 638.
37
John Barrow, The World Within the World, 181.
38
Ibid.,181.
39
Ibid.,182.
40
Abelians are symmetry groups with the commutative property ab=ba.
111
particle-antiparticle pairs. It is based on the idea that electrically charged particles interact
with each other in the electromagnetic field by emitting or absorbing photons, the massless
particles of spin 1 that transmit the electromagnetic force.
A case in point of this interaction is “the electromagnetic attraction between
negatively charged electrons and positively charged protons in the nucleus, which causes the
electron to orbit the nucleus of the atom. The electromagnetic attraction is pictured as being
caused by the exchange of a large number of massless particles of spin 1, called photons.”41
These exchanged photons are virtual particles, which because of their confinement in atoms
can hardly be detected. They can, however, become real photons, when released from the
atom, which happens when an electron jumps from a higher allowed orbital to a lower one. In
that case it can freely travel though space.
But more spectacular is the way in which electron-positron pairs annihilate one
another by emitting two photons. In the diagram below, an electron (e-) and a positron (e+)
meet and annihilate each other, so as to give rise to two photons ( ɣ).
In 1948 the American physicist Richard Feynman brought the various interactions
between electrically charged particles and photons into picture with his Feynman diagrams.
These diagrams depict various constellations in which electrons and positrons travelling
independently through space and time may collide so as to annihilate and emit two photons:
conversely they may illustrate how a photon pair gives rise to the creation of an electron-
positron pair. The aim and purpose of these various diagrams is to offer a practical tool for
calculating the probability with which, in a particular constellation, the ‘photon exchange’
takes place, and with which intensity it may happen. In Feynman’s approach “a particle does
not just have a single history, as it would in a classical theory. Instead, it is supposed to follow
every possible path in space-time,”42 and to interact in every way available. The probability of
each final state is then obtained by summing over all the possible paths.
In the beginning QED was still plagued with infinities, which indicated that no
absolutely precise result was obtained. The difficulty was overcome in the late 1940s with the
development of a procedure called renormalization. “It consisted of the rather arbitrary
subtraction of certain infinite quantities to leave finite remainders. In the case of
electrodynamics it was necessary to make two such infinite subtractions, one for the mass and
one for the charge of the electron.”43 Ever since, renormalizability has become a prerequisite
41
Stephen Hawking, A Brief History of Time, 75.
42
Ibid, 141.
43
Stephen Hawking, “Is the End in Sight of Theoretical Physics?,” in John Boslough (ed.), Stephen Hawking’s
Universe (New York: Avon Books, 1980), 125-126.
112
for a theory’s recognition in the scientific community: “QED is a renormalizable theory. [...]
It is a common standpoint, among particle physicists, to take renormalizability as a selection
principle for proposed theories. Accordingly, any non-renormalizable theory would be
automatically rejected as inappropriate to Nature.”44
Quantum Electrodynamics lays at the basis of the Standard Model, which comprises
also the study of the weak and the strong nuclear force. The major success of the Standard
Model was the unification of the electromagnetic force and the weak nuclear force. Such
unification, dubbed Electroweak Theory, is not simple, for the two forces have on prima facie
consideration very little in common. First of all, the weak nuclear force acts across distances
smaller than the atomic nucleus, while the electromagnetic force can extend for great
distances in space-time (think of the light of stars reaching us from other galaxies). And
second, the weak force is some 1,000,000 times weaker than the electromagnetic force.
The study of the weak force aimed at developing a consistent gauge theory that could
withstand comparison with Quantum Electrodnamics. Weak interactions are responsible for
the decay of fundamental particles into other (lighter) particles: one observes the particle
vanishing and being replaced by two or more different particles. In beta minus decay, e.g., a
neutron decays into a proton, an electron, and an antineutrino.
During the 1960s Sheldon Glashow, Abdus Salam, and Steven Weinberg
independently of one another discovered that a gauge theory of the weak force could be
constructed, if the electromagnetic force was included in the theory formation. This inclusion
would provide the researchers with a platform from where to build an appropriate gauge
theory. A major obstacle, however, in developing such a theory was the fact that the W+, W-
and Z° bosons that regulate the weak interaction are extremely massive, while photons, the
bosons of the electromagnetic force, are massless. The heavy W+, W- and Z° particles are
accurately described by the mathematical symmetry group SU(2), whereas according to the
mathematical symmetry group U(1) that sets the tone in Quantum Electrodynamics the photon
must be masslesss. So, a new special symmetry device will be needed to unite the two
previous ones, namely the symmetry group SU(2) x U(1). It is this symmetry group that
supplies the underlying pattern for the way in which the bosons of the unified electroweak
force are fully interrelated: “All of W+, W-, Z0, and ɣ can, in a certain sense, be continuously
‘rotated into one another’,”45 so that all of them are, basically, on the same footing.
The above unification only works under special conditions. The key issue is, indeed,
that the four mediator particles not only have different electric charges – the photon ɣ, and Z0
have no electric charge, whereas W+ and W- have positive and negative charge – but also
different masses: the photon particle ɣ is massless, whereas W+, W-, and Z0 are extremely
massive (mass 80/90). So how can the theory claim that the electromagnetic and the weak
force behave as if they were basically the same? Facing up to this difficulty, Glashow, Salam,
and Weinberg assumed that the reconstructed ‘electroweak’ force is only workable under
extreme temperatures (or, high energy scales), as these existed in the early universe. So, the
asymmetry we observe today would be “the result of a spontaneous symmetry breaking that is
taken to have occurred in the early stages of the universe. Before that period, conditions were
very different from those holding today, and the standard electroweak theory asserts that in
the extremely high temperatures in the early universe the U(2) symmetry [that is: the SU(2) x
U(1) symmetry] held exactly,” so that W+, W-, Z0, and ɣ could be rotated into one another.
44
Roger Penrose, The Road to Reality, 678.
45
Ibid., 641.
113
“But, as the idea goes, when the temperature in the universe cooled (to below 1016 Ke,46 at
about 10-12 seconds after the Big Bang), the particles W+, W-, Z0, and ɣ were ‘frozen out’ by
this process of spontaneous symmetry breaking. […] Just three of them acquire mass, and are
referred to as the Ws and the Zo; the other one remains massless and is called the photon. In
the initial ‘pure’ unbroken version of the theory, when there was complete U(2) symmetry,
the Ws, Zo, and ɣ would all have to be effectively massless.” In order to explain the
mechanism of symmetry-breaking the three scientists had recourse to an additional interaction
with another particle/field, known as Higgs (particle): “The Higgs (field) is regarded as being
responsible for assigning mass to all these particles (including the Higgs particle itself) and
also to the quarks that compose other particles in the universe.”47
The massive character of the weak gauge bosons W+, W- , and Z0 can be derived from
the fact that the weak force is short range, which reveals that it is carried by massive particles.
In 1971 Gerard ’t Hooft proved that the unified electroweak theory proposed by Glashow,
Salam, and Weinberg was renormalizable, and the theory gained full respectability. In 1983
W+, W- , and Z0 were discovered at CERN (European Centre for Nuclear Research, near
Geneva, Switzerland), “with the correct predicted masses and other properties.”48 Glashow,
Salam and Weinberg received the Nobel Prize in physics in 1979 even before the predicted
particles were detected in the Geneva particle accelerator. The discovery of the Higgs
field/particle figured as a top priority on the agenda of the Large Hadron Collider that was
successfully put to use by CERN in March 2010. In July 2012 scientists at CERN discovered
traces of the Higgs field/particle. In 2013 François Englert, together with Peter Higgs,
received the Nobel Prize in physics for the discovery of the Higgs mechanism.
The successful reconstruction of the electroweak force was one of the major
achievements of the Standard Theory, soon to be followed in the 1970s by the elaboration of
Quantum Chromodynamics (QCD), the gauge theory that describes the working of the strong
nuclear force. According to this theory, the strong nuclear force is transmitted by bosons
dubbed gluons. Like photons, gluons are massless, have a spin of 1, and travel at the speed of
light. But they differ from photons in one important respect: they carry what is called ‘color’
charge, a property analogous to electric charge but which has almost no manifestation at
distances above the size of the atomic nucleus. QCD is modeled after Quantum
Electrodynamics. However, whereas in Quantum Electrodynamics the exchange of photons
took place between two particles with opposite electric charge (namely electron and positron),
QCD had to elaborate an exchange of gluons between three colored quarks (see detailed
analysis below). This change in perspective relates to the fact that it was discovered that
proton and neutron were not indivisible as was thought till then, but made up of three types
(flavors) of quarks.
9.7.3 Quantum Chromodynamics
For a long time attempts had been made to explain the way in which protons and
neutrons were held together in the atomic nucleus. In the 1930s it was suggested that the force
that bound them together was produced by a pion (a meson with a mass intermediate between
that of the electron and the proton). Yet, since the late 1940s a vast number of heavy particles
(hadrons) – Λ0, Σ0, Ω- etc. – had been discovered in cosmic rays and accelerators which all
fell into certain families, called multiplets. This discovery led to the insight that the heavy
46
One degree Kelvin (Ke) equals minus 273.15 degrees Celsius.
47
Ibid., 643.
48
Stephen Hawking, A Brief History of Time, 77.
114
particles,49 which feel the strong nuclear force, were made up of ‘families’ of smaller particles
for which the American physicist Murray Gell-Mann coined the name ‘quarks’. He took the
designation ‘quark’ from James Joyce’s novel Finnegans Wake, in which the whimsical
saying appears: ‘Three quarks for Muster Mark’. In his study of the strong nuclear force Gell-
Man worked with the SU(3) symmetry group and came to the conclusion that protons and
neutrons (which are hadrons) “were made up of three types of quarks referred to as three
flavors, rather unimaginatively called: ‘up’ , ‘down’, and ‘strange’. A mysterious feature of
quarks was that they have to possess fractional electric charge […], the up, down, and strange
quarks having respective charge values ⅔, -⅓, and -⅓.”50
A proton contains two ‘up’ quarks and one ‘down’ quark (+⅔ +⅔ -⅓) giving a total
electric charge of +1, whereas a neutron is made up of two ‘down’ quarks and one ‘up’ quark
(-⅓ -⅓ +⅔ ), giving a total electric charge of zero. So, the fractional electric charge poses no
problem at all, since the respective arrangements thereof result in a proton with a positive
electric charge, and a neutron with a zero electric charge, as this is required by the atomic
structure.
Yet, on close inspection, these electric charge arrangements create a problem. Quarks
are matter particles (fermions), so they all have half-integer spin, and consequently must obey
the Pauli Exclusion Principle that forbids identical fermions from sharing the same quantum
state. This principle, however, is not being respected: Indeed, in the configuration that forms
the proton (+⅔ +⅔ -⅓) we have two ‘up’ quarks with the same electric charge +⅔, whereas
the configuration of the neutron ( -⅓ -⅓ +⅔ ) includes two ‘down’ quarks with the same
electric charge -⅓. This boils down, in both cases, to a violation of the Pauli Exclusion
Principle. This raises the thorny question as to whether quarks are really existing entities or
just a matter of bookkeeping to ensure the correctness of the mathematical calculations
proper to the symmetry group SU(3). As a matter of fact, the bookkeeping only works if one
pretends that quarks are ‘bosons‘, to which the Pauli Exclusion Principle does not apply. Yet,
such a trick is, after all, not really convincing.
49
The same insight applies to mesons: see below.
50
Roger Penrose, The Road to Reality, 646.
115
In order to resolve this problem the researchers set out to postulate that, in line with
the triplet preference of the symmetry group SU(3), ”each flavor [i.e. type] of quark also
comes in three (so called) ‘colors’.”51 Color stands for ‘color charge’. The diversified quarks
are taken to have ‘red’, ‘green,’ or ‘blue’ color charge; these are combined in such a way as to
yield a colorless composite particle (red +green + blue = white). On the basis of this
procedure, the problem raised by the Pauli Exclusion Principle is resolved: in the arrangement
of the proton (+⅔ +⅔ -⅓ ), e.g., we have, to be sure, two ‘up’ quarks with the same electric
charge +⅔; yet, these two ‘up’ quarks are split up now into ‘up red’ and ‘up green’. In this
way, no two identical quarks share the same quantum state; in combination with a blue
‘down’ quark, with electric charge -⅓, they form together a colorless proton.
The theorem of quark colors has still a further advantage: it allows for a deeper insight
into the very nature of the gluons and their interaction with quarks. Researchers found out that
gluons carry with themselves color charge, in sharp contrast to the photons, the exchange
particles in Quantum Electrodynamics, which have no electric charge of their own. Gluons,
the bosons of the strong force, can, in other words, radiate an array of gluons, whereas
photons are not able to increase their number. This radiation generates a ‘sea‘ of virtual
gluons with the effect that the quarks in this ‘sea’ engage in a dynamic interaction both with
each other’s color charge and with the gluon’s intensified color charge. It is in this context
that one has to place the phenomenon of color change. When a gluon is transferred between
two quarks, both quarks change their color: a red quark that emits a red–antigreen gluon
becomes green, whereas a green quark that absorbs a red–antigreen gluon becomes red.
Besides the quarks’ color change, there is also the phenomenon of confinement. The
more quarks try to escape their confinement within the proton or the neutron, the more the
‘sea’ of virtual gluons grows in strength. At tiny distances from the center, quarks behave as if
they were nearly free to move around. However, when one begins to draw the quarks apart in
an attempt at knocking them out of a proton, the binding force grows stronger. This
phenomenon is in complete contrast with the electric and gravitational force between particles
whose effects become weaker with the square of the distance: “The strong force is more like
an elastic band, where the strength of the force increases in proportion to the distance of
stretch, and it drops to zero, when the distance becomes zero.” 52 It is this increase in force
that is held responsible for the fact that quarks cannot be individually pulled out of a proton or
neutron.
At the end of the Twentieth Century, particle accelerators were already strong enough
to test the results of particle physics. Colliding beam experiments resulted in the discovery of
the charm quark at SLAC (Stanford Linear Accelerator Center) in 1968. All six flavors of
quark have since been observed in particle accelerator experiments; the top quark, first
observed at Fermilab (Fermi National Accelerator Laboratory, Chicago) in 1995, was the last
to be discovered. The gluon was first discovered in 1979 at the Deutsches Elektronen-
Synchrotron (DESY) in Hamburg, Germany.
The strong force is not only at work in protons and neutrons (hadrons); it also plays a
decisive role in the constitution of mesons, particles with a mass intermediate between that of
the electron and the proton. Mesons, too, are made up of quarks, but for them the quarks come
in pairs, which makes them less stable (quarks and antiquarks may annihilate one another). A
‘pi’ particle, e.g is made up of one ‘down antiquark’ and one ‘up quark’ (+⅓ +⅔) resulting in
a total electric charge of +1, whereas a ‘kaon’, which was first discovered in 1947, contains
one ‘up quark’ and one ‘strange antiquark’ (+⅔ +⅓) giving an electric charge of +1. The
mesons also possess color charge. A meson may consist of a red quark and an anti-red quark,
51
Ibid., 648.
52
Roger Penrose, The Road to Reality, 679.
116
or a green quark and an antigreen quark, or a blue quark and an antiblue quark. The result of
all these combinations is, in each case, a colorless meson.
Figure 24- Forces merge at high energies. Credit: The Particle Adventure (www.particleadventure.org)
In the above scheme, gravity is not involved. An encompassing theory that would
include gravity –the so called ‘Theory of Everything’ (TE) –is not easy to obtain: “The main
difficulty in finding a theory that unifies gravity with the other forces is that general relativity
is a ‘classical theory’; that is, it does not incorporate the uncertainty principle of quantum
53
Stephen Hawking, A Brief History of Time, 79.
54
Ibid., 79.
117
mechanics. On the other hand, the other partial theories depend on quantum mechanics in an
essential way. A necessary first step, therefore, is to combine general relativity with the
uncertainty principle.”55 There are various attempts at integrating the uncertainty principle
into general relativity. One of them has come about as a result of the study of black holes.
The existence of black holes – ‘voids’ in space – flows from the general theory of
relativity, which predicts its own inability to determine what happens at a singularity. At a
singularity, the theory of relativity breaks down. “A singularity is defined to occur when the
path of a light ray, or that of a particle, comes to an end. If this happens, then on reaching the
end of its path the particle disappears from the universe because it runs out of space and
time.”56 The ‘big bang’, insofar as it is to be located before the formation of space-time, is a
case in point of a singularity. So, too, is a black hole that emerges in the expanding universe:
“If sufficient mass is attracted into a small enough region by the pull of gravity, then the
gravitational field that it creates can become so strong that nothing can escape –not even
light.”57 Any light or other signal that comes close to the black hole is dragged into it and
cannot escape to the outside world.
Yet, in 1974 Stephen Hawking argued that the black hole must undergo the impact of
entropy and that, therefore, it ought to have a temperature and emit radiation. He corroborated
this view by showing that the gravitational field at the edge of the black hole (technically: at
its ‘event horizon’) must have quantum fluctuations with all that this involves: “One can think
of these fluctuations as pairs of particles of light or gravity that appear together at some time,
move apart and then come together again and annihilate each other.”58 The particles described
are virtual particles. Now, it may happen, Hawking goes on, that particles with negative
energy fall into the black hole and become real particles or antiparticles, while those with
positive energy escape from the black hole into infinity, in the form of radiating energy. This
baffling situation confronts us with a double unpredictability. If in quantum physics, in
normal circumstances, one cannot measure at the same time the position and the velocity
(momentum) of a particle, this uncertainty is doubled now by the fact that one is not able to
tell what happens with the twin particles that fell into the black hole: “An observer at a
distance from the black hole can measure only the outgoing particles, and he cannot correlate
them with those that fell into the hole because he cannot observe them.”59 Particles of light or
gravity can escape from the back hole, but what happens with their twin particles in the black
hole remains uncertain. In this way, the uncertainty principle of quantum mechanics is given a
place in the theory of relativity. Yet, this realization is only a modest step in the direction of a
real incorporation of quantum mechanics into the theory of relativity. Hawking acknowledges
that “we do not yet have a proper quantum theory of gravity, let alone one which unifies it
with the other physical interactions.”60
Another important area of research in view of the development of a Theory of
Everything (TE) is Supersymmetry (Susy). The term ‘supersymmetry’ refers to the
mathematical transformations that are needed to relate particles of integer spin (bosons) to
particles of half-integer spin (fermions). Supersymmetry posits that for every type of boson
there exists a corresponding type of fermion with the same mass and quantum number, and
vice versa, which suggests that originally they were practically the same. “Supersymmetry
demands […] that every fundamental particle in Nature has what is called a ‘superpartner’
55
Ibid., 164-165.
56
John Barrow, The World Within the World, 308.
57
Ibid., 310.
58
Stephen Hawking, A Brief History of Time,112.
59
Stephen Hawking, “Is the End in Sight of Theoretical Physics?,” in John Boslough (ed.), Stephen Hawking’s
Universe,133-134.
60
Ibid., 133.
118
with a spin that differs from that of the original particle by half a unit of spin. There needs to
be a 0-spin ‘selectron’ as partner to the electron, a 0-spin ‘squark’ to accompany each variety
of quark, a ½-spin ‘photino’ to partner the photon, a ½-spin ‘wino’ and ‘zino’ as respective
partners for the W and Z bosons, etc., etc.”61
This doubling serves a practical purpose: it allows the scientists to develop
mathematical constructs in which infinities are cancelled out by symmetries: “The virtual
particle/antiparticle pairs of spin ½ and 3/2 would have negative energy, and so would tend to
cancel out the positive energy of the spin 2, 1, and 0 virtual pairs. This would cause many of
the possible infinities to cancel out.”62 On the other hand, it became evident that the known
elementary particles were definitely not partners of each other under supersymmetry. So one
had to make the assumption that supersymmetry relates known bosons and fermions to
presently unknown fermions and bosons, as well as to explain why the ‘other half’ have not
yet been observed. For the proponents of the theory the answer to this query is simple:
supersymmetry would have been manifest at high energy scales; but must have been ‘broken’
at significantly lower temperatures. So, ‘super-partners’ would be observed only at
accelerators operating at these high energy scales.
Supersymmetry also became part of ‘string theory’. In the 1960s this theory was still
limited to the study of the working of the strong force. Its novelty consists in using 1-
dimensional oscillating lines (strings), instead of the classical point particles; these vibrating
lines can travel in space-time and form ‘world-sheets’ which interact with each other by
splitting and joining. Typical of strings is that they can vibrate in a specific manner, thus
giving the observed particles their proper flavor, charge, mass and spin. In 1984, string theory
incorporated supersymmetry; this enabled the string theorists to connect bosons and fermions,
as well as to reduce the number of space dimensions required by the theory from 26
dimensions to 10 or 11.
Yet, in the subsequent years, string theory ran into problems: The elaboration of
various versions of the equations led to the emergence of five major string theories, each of
them with a varying number of curled up dimensions and differing characteristics, such as
open loops and closed loops. Facing up to this difficulty, Edward Witten, the most famous
proponent of string theory, in 1994 opined that the five different versions might be describing
the same thing seen from various perspectives. In the words of Stephen Hawking: “String
theorists are now convinced that the five different string theories […] are just different
approximations to a more fundamental theory, each valid in different situations. That more
fundamental theory is called M-Theory [….] No one seems to know what the ‘M’ stands for,
but it may by ‘master’, ‘miracle’ or ‘mystery’. It seems to be all three.”63
M-theory is hailed by its adherents as the Theory of Everything (TE). Indeed, string
theory has no problem in bringing all the existing (and not yet discovered) particles into the
picture; so it has no difficulty in describing the specific properties of the spin-2 particle
graviton, the carrier of the gravitational force. Its theorists are convinced that they provided
the most impressive step towards a theory of quantum gravity. In a 1998 interview Edward
Witten declared: “String theory has the remarkable property of predicting gravity. […] The
fact that gravity is a consequence of string theory is one of the greatest theoretical insights
ever.”64
Not all scientists will share Witten’s enthusiasm. Roger Penrose, e.g,, is rather
skeptical about the achievements of string theory. For him, the string theorists are primarily
61
Roger Penrose, The Road to Reality, 875.
62
Stephen Hawking, A Brief History of Time,166.
63
Stephen Hawking & Leonard Mlodinow, The Grand Design (London: Bantam Press, 2010), 171.
64
Brian Greene, the Elegant Universe, 210 (Interview with Edward Witten, May 11, 1998).
119
fixated on cancelling out infinities (and thus on renormalizing their theory) with the help of
refined symmetries and dualities. But this purely and highly technically mathematical concern
threatens to make them loose track of the thorny problems that arise from the strict domain of
physics. He writes: “Those who come from the side of Quantum Field Theory [string theorists
being part of them] would tend to take renormalizability –or more exactly, finiteness –as the
primary aim of a quantum-gravity union. On the other hand, we from the relativity side would
take the deep conceptual conflicts between the principles of quantum mechanics and those of
general relativity to be the centrally important issues that needed to be resolved, and from
whose resolution we would expect to move forward to a new physics of the future.” 65 There
are apparently two different approaches in the search for the holy grail of quantum gravity:
one that seeks to integrate the findings of relativity into quantum physics (as this is done in
string theory), and one that seeks to incorporate quantum physics into general relativity (as
this is undertaken by Penrose).
The divergence between the two groups becomes manifest in their respective
elaboration of space-time. String theory operates with 10 or 11 space dimensions, 6 or 7 of
them being curled up at an infinitesimally small level, intertwined in such a way as to form
different types of 6-holded manifolds (technically called Calabi-Yau shapes). The aim and
purpose of these curled-up dimensions is to explain the immense variety of string vibrations
that account for the specific properties (mass, energy, charge, spin, etc.) of the specific
fermions and bosons. In this proliferation of particles, conditioned by the curled-up space
dimensions, the graviton is just a specifically oscillating string among other oscillating
strings; it is rather the result of a particular curled-up space formation, than that it comes to
bear on the specific ‘curvature of space-time, as is the case in general relativity.
Scientists, like Penrose, coming from the relativity perspective, on the contrary, focus
on the role played by gravity in the curvature of space time. If this is true, then the urgent
problem that needs to be tackled is how the specific geometry that is used to describe the
warping of space can be brought into rapprochement with string theory’s various curled-up
space dimensions. An equally urgent problem that needs to be resolved is that of the clashing
notions of time: instantaneous communication as this apparently happens in quantum physics
is at odds with the theory of relativity, according to which nothing can travel faster than the
speed of light. This shows, for Penrose, that radically new ideas about the nature of time are
required: “It is my opinion,” he writes, “that our present picture of physical reality,
particularly in relation to the nature of time, is due for a grand shake up –even greater perhaps
than that which has already been provided by present-day relativity and quantum
mechanics.”66
65
Roger Penrose, The Road to Reality, 893.
66
Roger Penrose, The Emperor's New Mind, 371.