02.07.2013 Views

Elementary Particle Physics - SS 04 Chapter 8

Elementary Particle Physics - SS 04 Chapter 8

Elementary Particle Physics - SS 04 Chapter 8

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Contents<br />

<strong>Elementary</strong> <strong>Particle</strong> <strong>Physics</strong> - <strong>SS</strong> <strong>04</strong><br />

<strong>Chapter</strong> 8<br />

Ian C. Brock<br />

30th July 20<strong>04</strong> – 17 : 26<br />

8 From Dirac to Renormalisation 1<br />

8.1 The Dirac Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />

8.2 Helicity and Helicity Conservation . . . . . . . . . . . . . . . . . . . . . . . 4<br />

8.3 Adjoint Spinors and Current Conservation . . . . . . . . . . . . . . . . . . 5<br />

8.4 Decay Rates and Cross-Sections . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

8.4.1 Decays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

8.4.2 Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

8.5 Feynman Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

8.6 Higher Orders and Renormalisation . . . . . . . . . . . . . . . . . . . . . . 14<br />

8.7 (g − 2) of Electron and Muon . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

8.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

8 From Dirac to Renormalisation<br />

The Dirac equation was the real starting point of relativistic quantum mechanics. It is the<br />

equation that provides the wavefunctions of spin 1/2 particles and leads to the introduction<br />

of antiparticles as equivalent partners to particles.<br />

Using the wavefunctions which are the solutions of the Dirac equation one can calculate<br />

scattering processes of fundamental particles. In the first part of the chapter I will introduce<br />

the Dirac equation as well as “Fermi’s Golden Rule”, which one needs when one wants to go<br />

from the transition amplitude or matrix elements, which one can calculate from Feynman<br />

graphs to decay rates or cross-sections.<br />

I have already introduced Feynman graphs as pictures which represent these calculations<br />

in perturbation theory. What I have not mentioned so far is that when one starts to<br />

actually calculate higher order graphs one very often gets answers which are infinite! This<br />

held up the development of Quantum electrodynamics (the quantum field theory of the<br />

electromagnetic interaction) for about 20 years! In the second half of this chapter I will<br />

discuss the problem and the solution.<br />

Lecture 16, 23/06/20<strong>04</strong>


8.1: The Dirac Equation 2<br />

8.1 The Dirac Equation<br />

The Klein-Gordon equation was introduced in <strong>Chapter</strong> 1, when I discussed antiparticles.<br />

A particle in free space is described by the de Broglie wave function:<br />

Ψ(x, t) = N exp[i(p · x − Et)] = Ne −ipx<br />

Starting from the relativistic energy-momentum relationship and substituting operators<br />

gives the Klein-Gordon equation:<br />

∂ 2 Ψ<br />

∂ 2 t = (∇2 − m 2 )Ψ<br />

The equation is suitable for describing spinless bosons, as no spin variable has been introduced.<br />

The Klein-Gordon equation is second order in the derivatives, which is the reason<br />

for the negative energy solutions. The Klein-Gordon equation also has another problem,<br />

in that it does not guarantee a positive-definite probability density for position, due to<br />

the 2nd order time derivative. Although we now know the implications of negative energy<br />

solutions, it provided the motivation for Dirac to try to find an equation that was linear<br />

in the derivatives and Lorentz invariant, in the hope that the probability problem and the<br />

one of of negative energy solutions could be solved.<br />

The simplest form that can be written down is that for massless particles, in the form of<br />

the Weyl equations:<br />

∂Ψ<br />

∂t<br />

<br />

<br />

∂Ψ ∂Ψ ∂Ψ<br />

= ± σ1 + σ2 + σ3<br />

∂x ∂y ∂z<br />

If this equation is valid it should also satisfy the Klein-Gordon equation. We square and<br />

compare coefficients to find:<br />

σ 2 1 = σ 2 2 = σ 2 3 = 1<br />

σ1σ2 + σ2σ1 = 0 etc.<br />

m = 0<br />

These results hold for both signs on the right-hand side of the Weyl equation and so must<br />

both be considered. The σ’s cannot be numbers, since they do not commute. In fact<br />

the equations define the 2 × 2 Pauli matrices which you should already know from atomic<br />

physics:<br />

σ1 =<br />

0 1<br />

1 0<br />

<br />

0 −i<br />

, σ2 =<br />

i 0<br />

<br />

1 0<br />

, σ3 =<br />

0 −1<br />

The equations can also be written in terms of the energy and momentum operators:<br />

Eχ = −σ · pχ<br />

Eφ = +σ · pφ<br />

From Dirac to Renormalisation Lecture 16, 23/06/20<strong>04</strong><br />

<br />

.


8.1: The Dirac Equation 3<br />

where E and p are the operators. χ and φ are two-component wavefunctions called spinors.<br />

They are separate solutions of the two Weyl equations. As I will discuss in a few minutes<br />

the equations have in total four solutions corresponding to particle and antiparticle states<br />

with two spin substates of each.<br />

If the fermion mass is included we need to enlarge the equations by including a mass term,<br />

which leads to the Dirac equation:<br />

EΨ = (α · p + βm)Ψ<br />

The matrices α and β are 4×4 matrices operating on four-component spinor wavefunctions.<br />

They are given by:<br />

<br />

0<br />

α =<br />

σ<br />

<br />

σ<br />

,<br />

0<br />

<br />

I<br />

β =<br />

0<br />

<br />

0<br />

,<br />

−I<br />

where each element denotes a 2 × 2 matrix and ‘1’ denotes the unit 2 × 2 matrix and σ<br />

are the Pauli matrices. The matrix α has three components. This representation is the<br />

so-called Dirac-Pauli representation of the matrices – other representations are possible.<br />

A more common way to write the Dirac equation is in its covariant form, replacing the<br />

operators by derivatives:<br />

<br />

<br />

µ ∂<br />

iγ − m Ψ = 0<br />

∂x µ<br />

(iγ µ ∂µ − m) Ψ = 0<br />

where the γµ (with µ = 0, 1, 2, 3) are related to the above matrices via:<br />

γ µ ≡ (β, βα)<br />

Note that Ψ is a four-element column matrix. However, it is not a 4-vector. I will return<br />

to combinations that are 4-vectors a bit later. Writing out the matrices explicitly we have:<br />

γ 0 <br />

I 0<br />

0 σ<br />

=<br />

, γ =<br />

, γ<br />

0 −I<br />

−σ 0<br />

5 <br />

0 I<br />

= .<br />

I 0<br />

It is useful to define the product matrix:<br />

γ 5 ≡ iσ 0 σ 1 σ 2 σ 3 =<br />

0 I<br />

I 0<br />

It is also very common to write down the Dirac equation in momentum space. Before and<br />

after an interaction with particle propagate as plane waves in free space. The fermion field<br />

can then be written as:<br />

Ψ(x) = u(p)e −ipµxµ<br />

From Dirac to Renormalisation Lecture 16, 23/06/20<strong>04</strong>


8.2: Helicity and Helicity Conservation 4<br />

where u now only depends on energy and momentum and not on the position of the particle.<br />

u is then a solution of:<br />

(γ µ pµ − m)u(pµ) = 0<br />

Although this equation does not look very different it is. The Dirac equation contains<br />

derivatives. The above equation is a pure matrix equation and as such is much easier to<br />

solve. One can construct four independent solutions to the Dirac equation – two of them<br />

have positive energy and two of them have negative energy/ The positive energy four-spinor<br />

solutions are:<br />

u (1,2) = √ <br />

E + m<br />

χ (1,2)<br />

p·σ<br />

E+m χ(1,2)<br />

<br />

χ (1) =<br />

1<br />

0<br />

Rewriting the negative energy solutions such that E > 0 we have:<br />

v (1) (p) = u (4) (−p) = √ E + m<br />

p·σ<br />

E+m χ(2)<br />

<br />

, χ (2) <br />

0<br />

=<br />

1<br />

χ (2)<br />

v (2) (p) = −u (3) (−p) = √ p·σ<br />

E + m E+mχ(1) χ (1)<br />

where E is the energy of the particle or antiparticle defined by E = p 2 + m 2 .<br />

The normalisation I have adopted is that from Halzen and Martin of 2E particles per unit<br />

volume.<br />

Using the definition of v given above, they satisfy the Dirac equation with the sign of pµ<br />

reversed:<br />

(γ µ pµ + m)v(pµ) = 0<br />

8.2 Helicity and Helicity Conservation<br />

For a massless particle with positive energy we have E = |p|. Inserting this in the first<br />

equation for the spinors we have:<br />

The quantity<br />

σ · p<br />

χ = −χ<br />

|p|<br />

H ≡<br />

σ · p<br />

|p|<br />

= −1<br />

is called the helicity or handedness of a particle. It measures the sign of the component of<br />

the spin of the particle, jz = ± 1 in the direction of motion (z-direction). The z-component<br />

2<br />

of spin and the momentum vector p together define a screw sense. H = +1 corresponds<br />

to a right-handed screw while H = −1 corresponds to a left-handed screw. The solution<br />

From Dirac to Renormalisation Lecture 16, 23/06/20<strong>04</strong>


8.3: Adjoint Spinors and Current Conservation 5<br />

of the equation given above represents a left-handed, positive energy particle. It can also<br />

represent a particle with negative energy −E and momentum −p. In this case:<br />

−Eχ = −σ · (−p)χ<br />

H =<br />

σ · (−p)<br />

= +1<br />

|p|<br />

Thus the equation represents either a left-handed particle or a right-handed antiparticle<br />

state.<br />

From the definition you can see that helicity is a well-defined Lorentz-invariant quantity for<br />

a massless particle. Such particles travel with the velocity of light. If one makes a Lorentz<br />

transformation to another reference frame with velocity v < c, it is therefore impossible to<br />

reverse the helicity. Neutrinos have a very small mass and for all practical purposes are<br />

well described by one of the two Weyl equations. Solutions of the Dirac equation are not<br />

pure helicity eigenstates, due to the mass term. Rather they are an admixture of the LH<br />

and RH functions. However, at high energies, when the particles are highly relativistic,<br />

the Weyl equations provide a good description also for massive particles.<br />

An important consequence of this is that for interactions involving vector or axial vector<br />

fields (i.e. those mediated by vector or axial vector bosons), helicity is conserved in the<br />

relativistic limit. As the mediators of the strong, electromagnetic and weak interactions<br />

are vector bosons, whose coupling is either vector or a mixture of vector and axial vector this<br />

constraint applies to all fundamental interactions. The reason for helicity conservation is<br />

that such interactions do not mix the LH and RH solutions of the Weyl equations. Writing<br />

these as ΨL ≡ χ and ΨR ≡ φ, consider as an illustration the momentum operator U = p<br />

operating on a mixture of LH and RH helicity states, i.e. Ψ ∗ UΨ, where Ψ = aΨL + bΨR,<br />

a and b are constants and Ψ ∗ is the complex transpose of the two-component spinor Ψ. I<br />

have omitted the space-time dependence e −ipx , as it cancels in the product. Making use of<br />

the Weyl equations one finds that the cross terms are zero since Ψ ∗ L ΨR − Ψ ∗ R ΨL = 0. If U<br />

were a scalar operator the cross terms would be proportional to Ψ ∗ L ΨR + Ψ ∗ R ΨL = 0 and<br />

would not disappear.<br />

In scattering processes at high energy a LH particle remains LH and RH remains RH.<br />

This fact together with the conservation of angular momentum determines the angular<br />

distributions in many interactions. It also implies that if one makes a beam of RH electrons<br />

these will only interact with LH positrons and not with RH positrons. For studies which<br />

one wants to make with polarised particles this is a great help, as one only needs to polarise<br />

one of the beams.<br />

8.3 Adjoint Spinors and Current Conservation<br />

The adjoint spinor is defined as:<br />

¯Ψ = Ψ † γ 0 = (Ψ ∗ 1, Ψ ∗ 2, −Ψ ∗ 3, −Ψ ∗ 4)<br />

********************* Begin skip *********************<br />

From Dirac to Renormalisation Lecture 16, 23/06/20<strong>04</strong>


8.3: Adjoint Spinors and Current Conservation 6<br />

From the conservation of probability, the rate of decrease of the number of particles in a<br />

given volume is equal to the total flux of particles out of that volume. Thus:<br />

− ∂<br />

<br />

<br />

ρ dV = j ·<br />

∂t<br />

<br />

ˆn dS = ∇ · j dV<br />

V<br />

S<br />

where I have used Green’s (or Gauss’) theorem. The probability and flux density are<br />

therefore related by the “continuity” equation:<br />

∂ρ<br />

∂t + ∇ · j = 0<br />

Expressing this in terms of the 4-vector j µ = (ρ,j) we have<br />

∂µj µ = 0<br />

I have not shown that the quantity so defined is a 4-vector, but one can.<br />

Returning to the Dirac equation:<br />

(iγ µ ∂µ − m)Ψ = 0<br />

we can also write it in terms of the adjoint spinor as:<br />

i∂µ ¯ Ψγ µ + m ¯ Ψ = 0<br />

Multiplying the Dirac equation from the left by ¯ Ψ and the adjoint equation from the right<br />

by Ψ and adding we obtain:<br />

¯Ψγ µ ∂µΨ + ∂µ ¯ Ψ γ µ <br />

Ψ = ∂µ<br />

¯Ψγ µ<br />

Ψ = 0<br />

Thus<br />

j µ = ¯ Ψγ µ Ψ<br />

satisfies the continuity equation and suggests that we should identify j µ with the probability<br />

and flux densities. In order to avoid problems with negative probability densities Pauli and<br />

Weisskopf inserted the charge of the electron −e into j µ , so that ρ is the charge density<br />

instead of the probability density:<br />

j µ = −e ¯ Ψγ µ Ψ<br />

j µ is the electron four-vector density, often called the electron current. To complete the<br />

identification one must show that j µ transforms as a 4-vector, which again can be done.<br />

********************* End skip *********************<br />

From Dirac to Renormalisation Lecture 16, 23/06/20<strong>04</strong><br />

V


8.4: Decay Rates and Cross-Sections 7<br />

The quantity<br />

j µ = ¯ Ψγ µ Ψ<br />

is the electron four-vector density, often called the electron current. It can be shown to<br />

transform as a 4-vector.<br />

For future reference it is convenient to write down the different forms of currents consistent<br />

with Lorentz covariance. Ψ and ¯ Ψ are not 4-vectors, but one can form the following bilinear<br />

quantities:<br />

¯ΨΨ Scalar +under P<br />

¯Ψγ µ Ψ Vector −under P (space components)<br />

¯Ψγ 5 Ψ Pseudoscalar −under P<br />

¯Ψγ 5 γ µ Ψ Axial Vector +under P (space components)<br />

I have left out the tensor, as it is very rarely needed.<br />

8.4 Decay Rates and Cross-Sections<br />

In order to calculate the decay rate or scattering cross-section one needs 2 ingredients:<br />

• The amplitude (M) or matrix element for the process and<br />

• The phase space or density of final states available.<br />

The amplitude contains all the dynamical information and we calculate it by evaluating the<br />

relevant Feynman diagrams using the Feynman rules appropriate to the interaction we are<br />

considering. The phase space factor contains only kinematical information; it depends on<br />

the masses, energies and momenta of the participants and reflects the fact that a process<br />

is more likely to occur if there is a lot of “room to manoeuvre” in the final state. For<br />

example the decay of a heavy particle into many light secondaries has a large phase space<br />

factor, as there are many different ways to apportion the available energy. In the β-decay<br />

of the neutron there is almost no extra mass to spare and the phase space factor is very<br />

small.<br />

The transition rate for a given process is determined using Fermi’s “Golden Rule” given<br />

by:<br />

transition rate = 2π<br />

¯h |M|2 × (phase space)<br />

A non-relativistic derivation of the Golden Rule can be found most books on quantum<br />

mechanics. For a relativistic version you have to consult a quantum field theory book. End of<br />

Lecture<br />

16<br />

From Dirac to Renormalisation Lecture 17, 25/06/20<strong>04</strong>


8.4: Decay Rates and Cross-Sections 8<br />

8.4.1 Decays<br />

Suppose particle 1 decays into several other particles 2, 3, 4, ..., n. The decay rate is given<br />

by the formula:<br />

3<br />

2 S c d p2<br />

dΓ = |M|<br />

2¯hm1 (2π) 3 3 c d p3<br />

2E2 (2π) 3 3 c d pn<br />

· · ·<br />

2E3 (2π) 3 <br />

2En<br />

×(2π) 4 δ 4 (p1 − p2 − p3 · · · − pn)<br />

where pi = (Ei/c, pi) is the 4-momentum of the ith particle. The delta function enforces<br />

the conservation of energy and momentum. We assume that the decaying particle is at<br />

rest: p1 = (m1c,0). S is the product of statistical factors: 1/j! for each group of j identical<br />

particles in the final state. The above formula gives the differential rate for a decay where<br />

the momentum of particle 2 lies in the range d3p2 about the value p2, etc. Normally we are<br />

not interested in individual momenta and so we integrate over the outgoing momenta to<br />

get the total decay rate Γ for the mode we are considering. If there are only two particles<br />

in the final state the result is:<br />

Γ = S<br />

<br />

c<br />

<br />

2 2 1 |M|<br />

δ<br />

¯hm1 4π 2 E2E3<br />

4 (p1 − p2 − p3) d 3 p2 d 3 p3<br />

If we have a 2-body decay into massless secondaries π 0 → γγ then one can show that<br />

(exercise?):<br />

Γ =<br />

S<br />

|M|<br />

16π¯hm1<br />

2<br />

i.e. the integration can be carried out independently of the form of |M| 2 (strictly true<br />

in the case of spin 0 particles and also if we want the spin averaged amplitude). M is<br />

evaluated at the momenta dictated by the conservation laws: p3 = −p2 and |p2| = mc/2.<br />

One can consider the more general case of a 2-body decay in which the outgoing particles<br />

have masses m2 and m3. In this case we have:<br />

Γ =<br />

S|p|<br />

8π¯hm 2 1c |M|2<br />

where p is the magnitude of either outgoing momentum. It is given in terms of the 3 masses<br />

involved:<br />

<br />

c<br />

|p| = m<br />

2m1<br />

4 1 + m4 2 + m4 3 − 2m2 1m2 2 − 2m2 1m2 3 − 2m2 2m2 3<br />

If both decay particles are massless then |p| = m1c/2, as we had above. The spin factor is<br />

S = 1<br />

2 for π0 → γγ, whereas it is S = 1 for π 0 → ν¯ν The final formula for 2-body decay is<br />

very simple and general. If there are three or more particles in the final state the integrals<br />

cannot be done until the form of M is known.<br />

From Dirac to Renormalisation Lecture 17, 25/06/20<strong>04</strong>


8.5: Feynman Rules 9<br />

8.4.2 Scattering<br />

Suppose particles 1 and 2 collide, producing particles 3, 4, ..., n. The cross-section is given<br />

by:<br />

dσ = |M| 2 ¯h 2 S<br />

4 (p1 · p2) 2 − (m1m2c2 ) 2<br />

2 c d p3<br />

(2π) 3 2 c d p4<br />

2E3 (2π) 3 2 c d pn<br />

· · ·<br />

2E4 (2π) 3 <br />

2En<br />

×(2π) 4 δ 4 (p1 + p2 − p3 − p4 · · · − pn)<br />

This equation determines the cross-section for a process in which the momentum of particle<br />

3 lies in the range d 3 p3 about the value p3, that of particle 4 lies in the range d 3 p4 about<br />

the value p4 and so on. We often want to consider the case where we look at the angle<br />

at which particle 3 emerges. We then integrate over all the other momenta and over the<br />

magnitude of p3; what’s left gives us dσ/dΩ, the differential cross-section for the scattering<br />

of particle 3 into solid angle dΩ.<br />

By far the most common process that we will encounter is:<br />

1 + 2 → 3 + 4<br />

We start in the CM frame. In this frame p2 = −p1. Hence p1 · p2 = E1E2/c 2 + p 2 1. One can<br />

then show that:<br />

(p1 · p2) 2 − (m1m2c 2 ) 2 = (E1 + E2)|p1|/c<br />

Thus<br />

dσ =<br />

2 ¯hc S|M|<br />

8π<br />

2c (E1 + E2)|p1|<br />

d 3 p3 d 3 p4<br />

E3E4<br />

δ 4 (p1 + p2 − p3 − p4)<br />

Again one can carry out the integration without knowing the explicit form of M and we<br />

obtain:<br />

dσ<br />

dΩ =<br />

2 ¯hc S|M|<br />

8π<br />

2<br />

(E1 + E2) 2<br />

|pf|<br />

|pi|<br />

where pf is the magnitude of either outgoing momentum and pi is the magnitude of either<br />

incoming momentum. This equation applies to almost all scattering processes that we will<br />

consider.<br />

8.5 Feynman Rules<br />

If we are given a matrix element we now know how to calculate a cross-section or a decay<br />

rate. How do we calculate the matrix element? To start with I will discuss QED and restrict<br />

myself to electrons, positrons and photons. Extending to the other leptons is trivial (they<br />

are just a repeat of electrons and positrons). Extension to quarks is also straightforward<br />

– one just has to take into account the different charge of the quarks and remember that<br />

they come in 3 colours.<br />

Free electrons and positrons are represented by the wave functions:<br />

From Dirac to Renormalisation Lecture 17, 25/06/20<strong>04</strong>


8.5: Feynman Rules 10<br />

ψ(x) = ae−ip·xu (s) (p) ψ(x) = aeip·xv (s) (p)<br />

with s = 1, 2 for the 2 spin states<br />

The spinors satisfy the Dirac equations in momentum space<br />

(γ µ pµ − m)u = 0 (γ µ pµ + m)v = 0<br />

and their adjoints satisfy<br />

ū(γ µ pµ − m) = 0 ¯v(γ µ pµ + m) = 0<br />

The solutions are orthogonal<br />

ū (1) u (2) = 0 ¯v (1) v (2) = 0<br />

normalised<br />

ūu = 2m ¯vv = −2m<br />

and<br />

<br />

complete<br />

s=1,2 u(s) ū (s) = (γ µ pµ + m) <br />

s=1,2 v(s) ¯v (s) = (γ µ pµ − m)<br />

I gave a set of solutions in the last lecture.<br />

⇒ Transparency A complete set of spinors (spinor.tex)<br />

We will normally (but not always) average over the electron and positron spins, and in<br />

that case it does not matter that the solutions are not pure spin up and spin down. End of<br />

Lecture<br />

********************* Begin skip *********************<br />

17<br />

A free photon is represented by the wave function:<br />

A µ (x) = ae −ip·x ɛ µ<br />

(s)<br />

where s = 1, 2 for the two spin states (or polarisations) of the photon. The polarisation<br />

vectors satisfy the momentum space Lorentz condition The Lorentz condition in the<br />

Maxwell equations is given by ∂µA µ = 0. The Maxwell equations in terms of the 4-vector<br />

A µ = (V, A) are:<br />

[∂µ (∂ µ A ν − ∂ ν A µ )] = 4π ν<br />

J<br />

c<br />

The equation is invariant under a gauge transformation A µ → A µ + ∂ µ λ, where λ is an<br />

arbitrary function of space and time. This gauge freedom can be exploited to impose an<br />

extra constraint on the potential ∂µA µ = 0. One then has:<br />

[✷A ν ≡ ∂µ∂ µ A ν ] = 4π ν<br />

J<br />

c<br />

The potential still has a further gauge freedom provided that ✷λ = 0. I will not discuss<br />

the details here. Most often the Coulomb gauge will be used, where A 0 = 0, which implies<br />

that ∇ · A = 0. Thus the polarisation vectors satisfy:<br />

ɛ µ pµ = 0<br />

From Dirac to Renormalisation Lecture 18, 01/07/20<strong>04</strong>


8.5: Feynman Rules 11<br />

The solutions are orthogonal:<br />

and normalised:<br />

In the Coulomb gauge we have:<br />

µ ∗<br />

ɛ (1) ɛµ (2) = 0<br />

ɛ µ ∗ ɛµ = 1<br />

ɛ 0 = 0 ɛ · p = 0<br />

one can then show that the polarisation vectors obey the completeness relation:<br />

∗<br />

ɛ(s) ɛ i (s) j = δij − ˆpiˆpj<br />

s=1,2<br />

A convenient pair (in the Coulomb gauge, where ɛ · p = 0 and p points in the z direction)<br />

is:<br />

ɛ(1) = (1, 0, 0) ɛ(2) = (0, 1, 0)<br />

A massless particle has only two possible spin states.<br />

********************* End skip *********************<br />

The procedure to calculate the amplitude M for a particular Feynman graph is sketched<br />

in the transparency and given in somewhat more detail below:<br />

⇒ Transparency Feynman rules (chapter08 figs.sxi)<br />

1. Notation. Label the incoming and outgoing 4-momenta p1, p2, ..., pn and the corresponding<br />

spins s1, s2, ..., sn; label the internal 4-momenta q1, q2, ..., qn. Assign arrows<br />

to the lines: arrows on external fermion lines indicate whether it is an electron or a<br />

positron; arrows on internal fermion lines are assigned so that the direction of flow<br />

is preserved (i.e. every vertex must have one arrow entering and one arrow leaving).<br />

2. External lines. External lines contribute the following factors:<br />

• Incoming electrons: u<br />

• Outgoing electrons: ū<br />

• Incoming positrons: ¯v<br />

• Outgoing positrons: v<br />

• Incoming photons: ɛ µ<br />

µ ∗<br />

• Outgoing photons: ɛ<br />

3. Vertex factors. Each vertex contributes a factor<br />

igeγ µ<br />

The dimensionless coupling constant ge is related to the charge of the positron via<br />

ge = e = √ 4πα. For up-type quarks we must use 2<br />

√<br />

4πα and for down-type quarks<br />

√ 3<br />

4πα.<br />

we use 1<br />

3<br />

From Dirac to Renormalisation Lecture 18, 01/07/20<strong>04</strong>


8.5: Feynman Rules 12<br />

4. Propagators. Each internal line contributes a factor:<br />

Electrons and positrons:<br />

Photons:<br />

i(γ µ qµ+m)<br />

q2−m2 −igµν<br />

q2 5. Conservation of energy and momentum. For each vertex write a delta function of<br />

the form:<br />

(2π) 4 δ 4 (k1 + k2 + k3)<br />

where the k’s are the 4-momenta coming into the vertex. If the arrow for a particle<br />

leads out of the vertex, then k is minus the 4-momentum of the line. If the arrow for<br />

an antiparticle comes into the vertex (i.e. we have an outgoing antiparticle) then k<br />

is also minus the 4-momentum of the line.<br />

6. Integrate over internal momenta. For each internal momentum q write a factor<br />

and integrate.<br />

d 4 q<br />

(2π) 4<br />

7. Cancel the delta function. The result at the end will include a factor<br />

(2π) 4 δ 4 (p1 + p2 + ... − pn)<br />

corresponding to overall energy-momentum conservation. Cancel this factor and<br />

what remains is −iM.<br />

The procedure is then to write down all diagrams contributing to the process in question,<br />

calculate the amplitude for each and add them up to get the total amplitude, which is then<br />

inserted into the appropriate formula for the cross-section or lifetime.<br />

One extra factor must not be forgotten. Fermions have antisymmetric wavefunctions.<br />

Therefore we have to insert a minus sign in combining amplitudes that differ only in the<br />

interchange of two identical external fermions. As the total amplitude will eventually be<br />

squared, it does not matter which diagram you associate the minus sign with.<br />

8. Antisymmetrisation. Include a minus sign between diagrams that differ only in the<br />

interchange of two incoming (or outgoing) electrons (or positrons), or an incoming<br />

electron with an outgoing positron (or vice versa).<br />

Fermion loops need some special attention which I will say a bit about later.<br />

The propagator has a form consistent with what one may naively expect. The exchange of<br />

particles of high mass is suppressed by a factor 1/m 2 , as are large 4-momentum transfers.<br />

I will not say any more in these lectures on how one can extract the explicit form.<br />

********************* Begin skip *********************<br />

From Dirac to Renormalisation Lecture 18, 01/07/20<strong>04</strong>


8.5: Feynman Rules 13<br />

The explicit form of the propagator can be extracted from the Lagrangian. In general the<br />

Lagrangian density contains a part Lfree for each participating field and an interaction term<br />

Lint. As discussed at the beginning of the lectures, the form of the interaction term is given<br />

by imposing local gauge invariance. The propagator is obtained by applying the Euler-<br />

Lagrange equation to the spin-specific free Lagrangian. This yields the free field equation<br />

for the particle, which is then written in momentum space using the usual prescription.<br />

The propagator is i multiplied by the inverse of the field corresponding to the motion<br />

operator.<br />

• Spin 0 is the Klein-Gordon operator. In Hilbert space this is:<br />

∂µ∂ µ + m 2 φ = 0<br />

In momentum space using i∂µ → pµ this becomes:<br />

p 2 − m 2 φ = 0<br />

Hence the propagator is given by:<br />

DF (spin 0; p) =<br />

i<br />

p 2 − m 2<br />

• Spin 1/2 is the Dirac operator. In Hilbert space this is:<br />

In momentum space we have<br />

[iγ µ ∂µ − m] ψ = 0<br />

[p − m] ψ = 0<br />

where p = pµγ µ . Hence the propagator is given by:<br />

DF (spin 1/2; p) =<br />

i<br />

p − m<br />

= i p + m<br />

p 2 − m 2<br />

• Spin 1 with mass is the Proca operator. In Hilbert space we have:<br />

∂µ (∂ µ A ν − ∂ ν A µ ) + m 2 A ν = 0<br />

In momentum space this is given by:<br />

2 2<br />

−p + m ν<br />

gµν + pµpν A = 0<br />

where the metric gµν is given by<br />

gµν =<br />

The propagator is then given by:<br />

⎛<br />

⎜<br />

⎝<br />

1 0 0 0<br />

0 −1 0 0<br />

0 0 −1 0<br />

0 0 0 −1<br />

⎞<br />

⎟<br />

⎠<br />

DF (Spin 1; m = 0; p) = −i gµν − pµpν/m 2<br />

p 2 − m 2<br />

From Dirac to Renormalisation Lecture 18, 01/07/20<strong>04</strong>


8.6: Higher Orders and Renormalisation 14<br />

• Massless spin 1 is the Maxwell operator. In Hilbert space we have:<br />

[∂µ (∂ µ A ν − ∂ ν A µ )] = 0<br />

With the Lorentz condition ∂µA µ = 0, we have in momentum space:<br />

2 ν<br />

−p gµν A = 0<br />

which implies that the propagator is as I gave above for the photon:<br />

DF (Spin 1; m = 0; p) = −i gµν<br />

p 2<br />

Note that one cannot simply take the limit of the Proca operator as m → 0 to<br />

obtain the propagator for a massless particle. Current conservation ensures that<br />

pµpν/m 2 → 0 for massless particles.<br />

********************* End skip *********************<br />

The propagators as defined above are only valid for stable particles. For the Z and W ±<br />

boson one has to include the Breit-Wigner form of the resonance.<br />

8.6 Higher Orders and Renormalisation<br />

While the basics for QED were laid down in the 30’s it took until the 50’s before it was<br />

really established as a proper theory. The reason is that if you consider higher order<br />

Feynman graphs and try to calculate them you typically get infinity! For example if you<br />

consider electron-muon scattering, one diagram that one must consider is where the virtual<br />

photon splits into an electron-positron pair for a short time (called vacuum polarisation).<br />

⇒ Transparency Electron-muon scattering higher order (emuscat.mnf)<br />

Bubble looks fairly harmless. However, when you calculate the graph the momentum of<br />

the electron or positron is not defined and so you have to integrate over all possible values.<br />

When you do so you find that if k is the 4-momentum of the electron (and q − k is the<br />

momentum of the positron), the integral diverges as ln |k|!<br />

Such infinities appear very many loop diagrams. How to get out of the problem? For large<br />

values of q2 , i.e. hard scattering the integral has a term that goes as:<br />

I(q 2 ) = α<br />

3π ln<br />

2 M<br />

where M 2 is a large cut-off value rather than the true infinite value. I said that at each<br />

electromagnetic vertex one has to write the factor √ 4πα = e. If, however, instead of e you<br />

write:<br />

<br />

eR ≡ e<br />

−q 2<br />

1 − e2<br />

2 M<br />

ln<br />

12π2 m2 From Dirac to Renormalisation Lecture 18, 01/07/20<strong>04</strong><br />

1/2


8.6: Higher Orders and Renormalisation 15<br />

and if eR is finite then remarkably all such graphs give you a finite correction. Now you<br />

have to think what is the e that one inserts in such graphs? The charge you measure is<br />

defined in terms of electron scattering at long range. Thus the charge you measure also<br />

includes the effect of such higher order graphs. If this is in fact eR and this is finite, then<br />

I should calculate the graphs with e (which may well be infinite), but at the end I should<br />

write the result in terms of eR that subtracts an infinite cut-off term and leaves me with a<br />

finite answer! Or instead I just use eR everywhere and as such have automatically included<br />

such higher order corrections (be careful there are a number of technical details in the<br />

calculations that have to clarified when you do this).<br />

One also finds that masses have to be corrected in a similar way. The measured mass is<br />

given by an (infinite) naked mass minus an (infinite) correction. In order for renormalisation<br />

to work it is necessary that all the divergent terms can all be absorbed into the masses<br />

and coupling strengths.<br />

If you think this is strange, I don’t blame you. The proof of the pudding is first of all<br />

whether it works! In fact it works extremely well. As you may of heard the Lamb shift<br />

between the S and P states in hydrogen probes exactly such effects as you are looking<br />

at different distances between the electron and the nucleus. Close to the nucleus there is<br />

a cloud of e + e − pairs that are polarised due to the charge of the nucleus. This cloud is<br />

exactly the sort of correction graph that I just drew.<br />

An intuitive explanation of this effect can be seen by considering the screening of a charge<br />

q by a dielectric medium. The negative end of the dipole is attracted towards q and the<br />

positive end is repelled. The “halo” of negative charge partially cancels the field due to q.<br />

⇒ Transparency Charge screening (screen1.jpg)<br />

⇒ Transparency Vacuum fluctuations and coupling strengths vs. distance (screen2.jpg)<br />

In QCD, i.e. due to the self-interaction of the gluons, this in fact works the other way round!<br />

While the vacuum polarisation diagrams that contain fermion-antifermion pairs result in a<br />

reduction of the effective colour charge, those with gluon loops lead to an increase at large<br />

distances. Which of the 2 wins is given by the factor:<br />

a ≡ 2f − 11n<br />

where f = 6 is the number of quark flavours and n = 3 is the number of colours. So in<br />

our world, the QCD coupling decreases at short distances, which leads to the effect known<br />

as asymptotic freedom, i.e. even though quarks interact strongly with each other, at short<br />

distances we can regard them as free particles and as such measure their properties and<br />

even “observe” them.<br />

As a result of the renormalisation the coupling strengths are no longer constants, but vary<br />

with q 2 . For the electromagnetic coupling strength αQED, while the value at q 2 = 0 is<br />

≈ 1/137 at MZ = 91 GeV its value is ≈ 1/128. As we have already seen, the variation of<br />

the strong coupling strength is stronger:<br />

From Dirac to Renormalisation Lecture 18, 01/07/20<strong>04</strong>


8.7: (g − 2) of Electron and Muon 16<br />

⇒ Transparency αs vs. q 2 (cernlep8 5-<strong>04</strong>.jpg)<br />

The plot shows the same trend as the one in the previous chapter, which was from the<br />

PDG. They show that LEP and PETRA data are very consistent when treated in a similar<br />

way. In addition the variation of αS with q 2 within a single experiment is clearly seen. A<br />

similar trend is seen in the measurements made at HERA:<br />

⇒ Transparency αs vs. q 2 (20<strong>04</strong> LLouise Poeschl6.jpg)<br />

In most of these measurements the energy of the measured jets sets the scale, i.e. the<br />

value of αS that one measures. These are not all the same (as they are in e + e − ), as the<br />

proton is a composite object and the struck quark can therefore carry a variable fraction<br />

of the proton momentum. In addition, ep interactions are a scattering process and not an<br />

annihilation process; hence the q 2 varies from one event to the next. In e + e − q 2 is (mostly)<br />

just given by the c.m. energy √ s.<br />

The masses of particles also depend on the q 2 of the process that measures them. First<br />

evidence for such a variation of the b quark mass has been shown (albeit model dependent)<br />

by the LEP experiments.<br />

⇒ Transparency Running b quark mass (cernlep10 5-<strong>04</strong>.jpg)<br />

8.7 (g − 2) of Electron and Muon<br />

Turning back to QED, another effect of the vacuum polarisation is that it leads to a<br />

slight modification of the magnetic dipole momentum of elementary fermions. One of the<br />

great triumphs of the Dirac equation is that it explained why the magnetic moment of the<br />

electron is:<br />

with S = 1σ<br />

and the gyromagnetic ratio<br />

2<br />

µ = −g e<br />

2m S<br />

g = 2<br />

There is however a small correction to this number, which can be both precisely calculated<br />

and measured. The calculations have been made to several orders of perturbation theory:<br />

⇒ Transparency Some leading diagrams contributing to the magnetic moment (g-2-feyn.jpg)<br />

The theoretical expectation is:<br />

g − 2<br />

2<br />

theory<br />

e<br />

= 0.5<br />

<br />

α<br />

<br />

− 0.32848<br />

π<br />

<br />

α<br />

2 + 1.19<br />

π<br />

= (115 965 230 ± 10) × 10 −11<br />

<br />

α<br />

3 + · · ·<br />

π<br />

From Dirac to Renormalisation Lecture 18, 01/07/20<strong>04</strong>


8.7: (g − 2) of Electron and Muon 17<br />

while the measured value is:<br />

<br />

g − 2<br />

2<br />

exp<br />

e<br />

= (115 965 218 59 ± 38) × 10 −11<br />

where the uncertainty in the theory actually comes mainly from the uncertainty in the<br />

value of α. You see a remarkable agreement between experiment and data.<br />

Measurements have also been made of (g − 2) for the muon. Over the past few years a<br />

group at Brookhaven has been working extremely hard to improve the accuracy of the<br />

measurements. The idea of the experiment is illustrated in the next transparency:<br />

⇒ Transparency Sketch of the main idea behind a g − 2 experiment (g-2.jpg)<br />

In the experiment at BNL, polarised muons are injected into a superconducting storage<br />

ring. The ring provides a very uniform static dipole field, B. If the g factor of a muon<br />

were exactly equal to two, the spins and momenta of the muons would stay parallel during<br />

storage. This would mean that the cyclotron frequency, ωc, of the muon is equal to its spin<br />

precession frequency, ωa. However, due to the anomaly on the magnetic moment, the spin<br />

precesses faster than the momentum ωs = ωc(1 + aµγ).<br />

From the theoretical side, an idea of the Feynman graphs that have to be calculated and<br />

the size of the various contributions is given in the transparency:<br />

⇒ Transparency Prediction for muon g − 2 (pw eps03 talk.pdf, P.42)<br />

The results for µ + and µ − are summarised in the following transparency:<br />

⇒ Transparency Muon g − 2 measurements and predictions (g2results2.eps)<br />

For the record the numbers are:<br />

aµ + = 11 659 203(8) × 10−10<br />

aµ − = 11 659 214(9) × 10−10<br />

aµ = 11 659 208(6) × 10 −10<br />

a e+ e −<br />

µ = 11 659 181(8) × 10 −10<br />

a τ µ = 11 659 196(7) × 10 −10<br />

(0.7 ppm)<br />

(0.7 ppm)<br />

(0.5 ppm)<br />

(0.7 ppm)<br />

(0.6 ppm)<br />

The average measured value from aµ differs from the predictions made using the e + e − and<br />

the τ decay data by 2.7σ and 1.4σ respectively. The 2 different theoretical predictions<br />

come from using e + e − → hadrons data at lower energies or using τ decay data for this<br />

energy range. There appear to be some incompatible measurements between 850 MeV and<br />

1 GeV. Discrepancies at lower energies have now disappeared after a reevaluation of the<br />

e + e − data.<br />

It is unlikely that one will see much improvement in these measurements (the predictions<br />

may well improve), as the proposal to continue the experiment at BNL with the aim of<br />

From Dirac to Renormalisation Lecture 18, 01/07/20<strong>04</strong>


8.8: Summary 18<br />

achieving an accuracy of 0.4 ppm, which should be sensitive to electroweak corrections was<br />

unfortunately not approved.<br />

The difference between the 2 values is somewhat under 3 standard deviations. While<br />

this cannot yet be seen as overwhelming evidence that there is some physics beyond the<br />

standard model it does provide a hint. It may indicate that there are some new, as yet<br />

unobserved, particles that can also contribute to the vacuum polarisation. One favourite<br />

explanation is supersymmetry, that introduces integer spin partners of all fermions and<br />

half-integer partners of all bosons. There are good theoretical arguments for the existence<br />

of supersymmetry, but no direct experimental evidence has yet been seen.<br />

8.8 Summary<br />

Infinity - infinity give wonderful agreement between experiment and theory!<br />

In this chapter I gave a very brief overview of the Dirac equation. I showed in principle how<br />

one can calculate cross-sections and decay rates using the Fermi Golden Rule and Feynman<br />

graphs. I discussed what happens when one tries to calculate higher orders. Although the<br />

expected contributions are small, in fact the graphs often given infinite contributions.<br />

These are resolved using the procedure called renormalisation, which absorbs these infinite<br />

corrections into a redefinition of the masses and charges of the fundamental particles. While<br />

this procedure may seem to be very arbitrary, the proof of the pudding, is that it works<br />

wonderfully. I have illustrated the level of agreement between theory and experiment with<br />

the (g − 2) measurements. I will show more comparisons in <strong>Chapter</strong> 10. End of<br />

Lecture<br />

18<br />

From Dirac to Renormalisation Lecture 18, 01/07/20<strong>04</strong>

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!