Sie sind auf Seite 1von 6

The Electroweak Interaction: Part I

Dr Emily Nurse
2017

1 Introduction
In this module we will discuss the electroweak interaction, which is a merged theory describing both the
theory of electromagnetic interactions and the weak interaction of nuclear decay. We will learn that the
separate weak and electromagnetic interactions are low energy manifestations of the electroweak theory. We
will go on to discuss the Higgs mechanism that provides mass to the fundamental particles, and the associated
Higgs boson, with particular emphasis on its recent discovery at the LHC.
Before all this we will learn about some unexpected asymmetries (parity and charge conjugation violation)
that were observed in the weak interaction in the 1950s, which were eventually explained by the spin structure
in the weak interaction. This will lead to a discussion of another subtle asymmetry in the weak interaction,
that is still being understood today (CP violation).

2 Symmetries and conservation laws


As described by the famous theorem of mathematician Emmy Noether, conservation laws in physics come
from symmetries seen in nature. For example, conservation of momentum comes from a symmetry of the
laws of physics under a spatial translation. Or put another way, from the fact that the laws of physics do not
change if we perform a spatial translation (e.g. you would get the same results if you performed an experiment
in London or in Chicago). Similarly, the conservation of angular momentum comes from symmetry of physical
laws under a rotation and the conservation of energy comes from symmetry of physical laws under a translation
in time (you would get the same results if you performed an experiment today or this time next year). These
are remarkable results that can be derived from Quantum Mechanics (see e.g. Section 4 in “Particle Physics”,
B. R. Martin and G. Shaw.) Richard Feynman discusses this principle in his usual accesible and enlightening
way in the Feynman lectures: http:www.feynmanlectures.caltech.edu/I_52.html

2.1 Parity
There is a conserved quantity known as parity that comes from a symmetry under space-inversion1 . This
means that the physical laws of a system do not change if the spatial co-ordinates are inverted: x ! x,
y ! y and z ! z, which is equivalent to a mirror reflection (x ! x and y ! y) together with a
rotation through 180 (z ! z). We can establish the e↵ect of space-inversion on the following kinematic
properties of a system:
1 Often “parity” is used to mean the symmetry as well as the conserved quantity. I prefer to call the symmetry space-inversion

to avoid this confusion (note that in the previous examples this would be like e.g. energy and time having the same name).

1
PHAS3224: Nuclear and Particle Physics 2017

1. The position vector ~r changes to ~r.


2. The momentum vector p~ changes to p~.
~ (= ~r ⇥ p~) remains unchanged as both ~r and p~ change sign.
3. The angular momentum vector L
Parity is a discrete conserved quantum number and is equal to the eigenvalue of any wavefunction that is an
eigenstate of the space-inversion transformation, P̂ . We define the intrinsic parity of a particle by considering
the space-inversion operator acting on a particle at rest:
P̂ a = Pa a, (1)
where Pa is the particle’s intrinsic parity. Two successive space-inversion transformations take us back to
where we started:
P̂ P̂ a = P̂ Pa a = Pa2 a = a , (2)
so we can deduce that Pa = ±1. Intrinsic parities of particles are generally determined by experiment,
assuming that parity is conserved in interactions, together with the fact that a fermion and its antiparticle
must have opposite parities and a boson and its antiparticle must have the same parity.2 By considering a
multiparticle state with a wavefunction that is the product of single-particle wavefunctions3 , it is clear that
parity must be a multiplicative quantum number and the total parity of a system of particles a, b, ... with
L
orbital angular momentum is given by: Pa Pb ... ( 1) . The contribution to parity from the orbital angular
momentum comes from the consideration of the form of a wavefunction of a particle with definite angular
momentum and noting the fact that in spherical polar coordinates a space-inversion transformation implies:
r ! r, ✓ ! ⇡ ✓, ! ⇡ + .4
Conservation of parity in an interaction can be considered in one of two ways, which amount to the same
question:
1. Is the interaction invariant under the space-inversion transformation?
2. Is the total parity before the interaction equal to the total parity after the interaction?
Until the late 1950s it was believed, as we might intuitively expect, that all physical laws were invariant
under space-inversion and therefore conserved parity.

2.1.1 The ⌧ ✓ puzzle


In 1949 Powell discovered two weakly decaying particles in cosmic ray showers that were known as the ✓+
and ⌧ + mesons. They had the same mass, charge and spin, so they looked much like the same particle. They
decayed weakly to two di↵erent final states: ✓+ ! ⇡ + ⇡ 0 and ⌧ + ! ⇡ + ⇡ 0 ⇡ 0 . While it is perfectly fine for one
particle to decay via di↵erent decay modes, these two final states have di↵erent parities. In order to see this
we use the experimentally determined fact that the intrinsic parity of both the charged and neutral pions
is 1, together with the fact that the orbital angular momentum in both cases is zero.5 The two particles
cannot then be the same particle if parity is conserved in the decay.
Inspired by this puzzle, in 1956 Lee and Yang surveyed the experimental evidence and suggested that
the weak interaction does not necessarily conserve parity (even though it had been a firm assumption by
physicists until this time).
2 If interested, a proof of why this is in e.g. “Modern Particle Physics”, M. Thomson.
3 This is only true if the particles are distinguishable.
4 For details see section 1.3.1 in “Nuclear and Particle Physics”, B. R. Martin.
5 If interested you can see a discussion of why this is in e.g. Section 6.6.1 of “Nuclear and Particle Physics”, B. R. Martin.

2
PHAS3224: Nuclear and Particle Physics 2017

2.1.2 Parity violation in decay


The first evidence for parity violation in the weak interaction came one year later in 1957 by Wu and
collaborators. She studied the decay of polarised Cobalt-60: 60 Co ! 60 Ni* + e + ⌫¯e . In this reaction
a neutron from within the 60 Co nucleus is converted into a proton and an electron and antineutrino are
emitted. Polarised means that the 60 Co spins are all aligned. Wu measured the direction of the emitted
electrons with respect to the 60 Co spin direction and found that they were more likely to be travelling in
the direction opposite the nuclear spin. What does this tell us? Lets consider what a space-inversion (or
parity transformation) of this reaction would do. It would leave the spin of the nucleus unchanged (since
spin is an angular momentum), but it would reverse the direction of the emitted electrons. Space-inversion
symmetry (and consequently parity conversion) of this process would then imply that the electrons should be
emitted with equal probability aligned and anti-aligned with the nuclear spin. It is clear then that the results
of Wu’s experiments proved that parity violation exists in the weak interaction. This was a monumental
breakthrough in the understanding of the weak interaction.
These results solved the ⌧ ✓ puzzle. What we now know is that the ✓+ and ⌧ + mesons are in fact the
same particle (known as the charged kaon: K + ) which can decay to either ⇡ + ⇡ 0 or ⇡ + ⇡ 0 ⇡ 0 since parity is
violated in these weak decays. Note that the strong and electromagnetic interactions do conserve parity.

2.2 Charge conjugation


Charge conjugation is another important discrete symmetry. A transformation under the charge conjugation
operator, Ĉ, changes a particle, a, into an antiparticle, ā, together with a phase factor6 , Ca , specific to that
particle:
Ĉ a = Ca ā . (3)
We can see from Equation 3 that a is only a eigenstate of Ĉ if the particle is its own antiparticle: a = ā,
e.g. for the photon and the neutral pion. In these cases the eigenvalue, Ca , is known as the intrinsic C-parity
of the particle. If we apply the operator twice we should be back where we began, so if a = ā we obtain:
Ĉ Ĉ a = ĈCa a = Ca2 a = a leading to Ca = ±1. For interactions involving particles that are eigenstates
of the charge conjugation operator, the total C-parity before and after the interaction must be conserved
if the interaction is symmetric under a charge conjugation transformation. Said another way: the total C-
parity of an interaction is conserved if the interaction is symmetric under particle $ antiparticle exchange.
The C-parity of particles can be measured by assuming charge conjugation conservation in interactions. For
particles that are not their own antiparticles eigenstates of charge conjugation can only be formed by linear
combinations of a particle–antiparticle pair, e.g. p12 K 0 + K¯0 where K 0 = ds̄ and K¯0 = sd. ¯ It is worth
noting that in the case that a 6= ā the phase factor Ca in Equation 3 cannot be measured and can therefore
be set to 1 (see problem sheet 3). Just like parity, C-parity is a multiplicative quantum number. It turns
out, as we shall see in the next section, that the weak interaction is not invariant under a charge conjugation
transformation, however both the electromagnetic and strong interactions are.
6 A phase factor in quantum mechanics is defined as ei✓ , where ✓ is known as the phase. The absolute value of a phase factor

is 1. A phase factor therefore does not have any a↵ect on probabilities.

3
PHAS3224: Nuclear and Particle Physics 2017

2.2.1 Parity and charge conjugation violation in muon decay


In 1957 various experiments measured the angular distribution of electrons (positrons) from the decay of
polarised (anti)muons: µ ! e ⌫¯e ⌫µ or µ+ ! e+ ⌫e ⌫¯µ . These measurements showed it to be of the form:
✓ ◆
1 1
Re± (✓) = ± 1 ± cos ✓ (4)
2 µ 3
where Re (✓) is the rate of electrons from muon decay as a function of the angle between the muon spin
and the electron direction, ✓, Re is the same thing for positrons in antimuon decay, µ is the muon decay
rate constant and µ+ is the antimuon decay rate constant. The first thing to note from Equation 4 is the
di↵erence between the form of Re (✓) and Re+ (✓), seen immediately from the sign in front of the 13 cos ✓
term (positive for e+ and negative for e ). If charge conjugation were conserved, we would expect the form
of the two to be identical (as a charge conjugation transformation would just change the µ decay into a µ+
decay and vice versa). By inspecting Equation 4 we can also see that parity is violated in this decay. The
parity transformation implies that ✓ ! ⇡ ✓ and ! ⇡ + in spherical polar coordinates. If we consider
the muon spin to be aligned with the z-axis so that for a parity transformation, ✓ ! ⇡ ✓ in Equation 4,
such that 1 + 13 cos ✓ ! 1 13 cos ✓ and vice versa. We therefore conclude that parity is also violated in muon
and antimuon decay.
However, an important thing to note is that if we consider a charge conjugation transformation (C)
together with a parity transformation (P), we are back where we started, as the sign in front of 13 cos ✓
changes twice. This is telling us that even though the weak interaction violates C and P symmetries, it is
invariant under a combined CP transformation. While this is true to a good approximation, we will see later
that CP violation does exist at a small level in the weak interaction, and this is intimately linked to the
dominance of matter over antimatter in our Universe. We should also reiterate here that the individual C
and P symmetries are conserved in the electromagnetic and strong interactions.

2.3 The spin structure of weak interactions


The C and P violation that was observed experimentally in the 1950s is now explained within the Standard
Model by the spin structure of weak interactions. In order to understand this we will first introduce helicity,
which is defined as the projection of a particle’s spin on its direction of motion: H = ~s|~p.~p| , where ~s is the
particle spin and p~ is the particle momentum. Recall that the projection of a fermion’s spin on any chosen axis
can take two possible values: + 12 or 12 . Therefore, a fermion can either have a positive helicity: H = + 12 or
a negative helicity: H = 12 . A negative helicity means that the particle is left-handed. A positive helicity
means it is right-handed. It turns out that only left-handed neutrinos and right-handed antineutrinos interact
weakly. This is the origin of the observed parity violations in weak interactions, as we shall see. And in fact,
since neutrinos only interact via the weak interaction, it turns out that only left-handed neutrinos (⌫L ) and
right-handed antineutrinos (¯ ⌫R ) are observed in nature! This was proved in 1958 when Goldhaber measured
the helicity of the neutrino.
For any particle with non-negligible mass, such that v < c, it is always possible for an observer to
travel faster and overtake the particle. A left-handed particle would then appear right-handed. Therefore,
for fermions other than neutrinos we cannot say that they are 100% left or right-handed. For relativistic
⇣ 2 ⌘2
fermions with mass, the forbidden helicity states are suppressed by an approximate factor: ⇠ mc E , where
m is the fermion mass and E is the fermion energy. This is clearly a small number for relativistic fermions.
We can see that a parity transformation on a left-handed neutrino brings us to a right-handed neutrino
(the momentum changes direction and the spin remains unchanged). Since right-handed neutrinos do not

4
PHAS3224: Nuclear and Particle Physics 2017

exist it is clear that parity is violated. Similarly, charge conjugation transformation on a left-handed neutrino
turns it into a left-handed antineutrino, which also does not exist, demonstrating that charge conjugation
is also violated. However a combined CP transformation would change a left-handed neutrino into a right-
handed antineutrino, which does exist, so it appears CP can be conserved. We will see in the next sections
explicitly how the lack of the existence of right-handed neutrinos and left-handed antineutrinos led to the
observations in -decay and muon-decay that we have discussed.

2.3.1 decay revisited


We now revisit the parity violation observed in decay for Cobalt-60, where we saw that electrons were
preferentially emitted in the direction opposite the direction of the nuclear spin. 60 Co has a spin of 5
and 60 Ni* has a spin of 4. In order to conserve angular momentum, the spin of the produced electron and
antineutrino (both spin-1/2) must both be aligned with the nuclear spin. In order to conserve momentum
the electron and antineutrino will be travelling in opposite directions. In the case that the antineutrino is
in the direction of the nuclear spin and the electron is in the opposite direction, the antineutrino will be
right-handed and the electron will be left-handed in the massless limit, which is perfectly allowed. However,
the case where the electron is in the direction of the nuclear spin, and the antineutrino is in the opposite
direction, we are dealing with a left-handed antineutrino and a right-handed electron in the massless limit.
This configuration is forbidden by the spin structure of the weak interaction and leads to the observed
experimental results, where the electron is preferentially emitted in the direction opposite the nuclear spin.

2.3.2 Muon decay revisited


We now revisit the parity violation observed in muon decay. This is a three body decay so the orientation
of the particles is not as simple, but if we look at the decays with the highest possible electron (or positron)
energies, then we are in a configuration where the electron (positron) is travelling back-to-back with the
neutrino and the antineutrino, which would then travel very close to each other. In this case, ensuring
momentum conservation, the energy of the electron/positron will be Ee ⇠ mµ /2. If we consider the two
possible cases where: (1) the electron travels in the direction opposite the spin of the decaying muon and
(2) the electron travels in the same direction of the decaying spin, we can convince ourselves (see lectures)
that there is only one possible spin alignment for each that does not involve either a right-handed neutrino
or a left-handed antineutrino. In order to conserve angular momentum, in case (1) the electron must have
negative helicity and in case (2) the electron must have positive helicity. In the relativistic limit case (2)
is suppressed. Since the mass of the electron is very small compared to the decaying muon, the relativistic
limit is a good approximation, and we come to the experimentally observed conclusion that electrons are
preferentially emitted in the direction opposite the muon spin. If we work through the same logic for antimuon
decay, we reach the opposite conclusion for the direction of the positrons, just as was observed experimentally.

2.3.3 Pion decay


Consider the charged pion decay process:
⇡ + ! `+ + ⌫ `
where `+ is a positron or an antimuon. Ignoring helicity suppression, we would expect these rates to be
similar, due to lepton universality (e.g. the coupling of the W + to the e+ ⌫e is the same as that to the µ+ ⌫µ ).
There are some expected di↵erences in the rates due to kinematic e↵ects. The smaller mass of the electron
with respect to the muon results in it having a larger momentum, and due to the density of states factor

5
PHAS3224: Nuclear and Particle Physics 2017

of the reaction rate, we would expect the rate to be just over two times that for the muons (see problem
sheet 3). Experimentally, however, the branching ratio to µ+ ⌫µ is (99.98770 ± 0.00004)% and that to e+ ⌫e is
(0.01230 ± 0.00004)%. This huge di↵erence in the branching ratios can be understood by helicity suppression
resulting from the spin structure of the weak interaction.
If you consider the process in the rest frame of the decaying ⇡ + , the `+ and the ⌫` would be produced back-
to-back, travelling in opposite directions in order to conserve momentum. Now, the ⇡ + is a spin-0 particle
and the `+ and the ⌫` fermions are both spin-1/2 particles. So, in order to conserve angular momentum, the
spins of the two particles must be anti-aligned. This gives two possible scenarios:
1. a left-handed ⌫` and a left-handed `+
2. a right-handed ⌫` and a right-handed `+

We know that 2. is forbidden, as right-handed neutrinos are not observed in nature. So the only way the
process can occur is via 1. But this involves a left-handed antiparticle, the `+ . If the `+ were massless,
this decay would not be possible. Since the charged leptons are not massless the decay is possible, but it is
⇣ 2 ⌘2
suppressed by the factor mc E , where m is the charged lepton mass and E is its energy. We can determine
the charged lepton’s energy in the rest frame of the ⇡ by conserving momentum and energy before and after
the decay to give: m⇡ c2 = E` + E⌫ and |p` |c = |p⌫ |c. The neutrino mass is negligible compared to the
2
energies giving: E⌫ = |p⌫ |c = |p` |c. Now write the expression: E`2 = p2` c2 + m2` c4 = m⇡ c2 E` + m2` c4 =
2 2 4 2 2 4
E` + m⇡ c 2m⇡ c E` + m` c and rearrange to give:

m2` + m2⇡ c2
E` = .
2m⇡
⇣ ⌘2
2m⇡ m`
This leads to a suppression factor of m2` +m2⇡
. Using m⇡ = 139.57 MeV/c2 , mµ = 105.7 MeV/c2 and
me = 0.51 MeV/c2 we can straightaway see that the muon is not highly relativistic in this decay, but the
✓ ◆2
me (m2µ +m2⇡ )
electron is. The ratio of the suppression factor for electrons to that for muons is mµ (m2 +m2 ) = 5.7⇥10 5 .
e ⇡

A detailed calculation, combined with the density of states factor mentioned above, leads to a ratio that is
in very good agreement with the experimentally determined branching ratios.

Das könnte Ihnen auch gefallen