1 Introduction ..... 2
2 Path Integral in QM ..... 2
3 Path integral for free QFT ..... 5
4 Path integral for interacting QFT ..... 7
5 LSZ reduction ..... 12
6 Scattering Amplitudes and Feynman Rules ..... 14
7 Loops and Renormalisation ..... 17
8 Classical Maxwell Theory ..... 22
9 LSZ and Path Integral For Photons ..... 25
10 Dirac Equation ..... 28
11 Classical Solutions of the Dirac Equation ..... 32
12 Quantizing the Dirac field ..... 35
13 LSZ for spin half ..... 39
14 Fermionic Propagator ..... 41
15 Quantum Electrodynamics ..... 44
16 Scattering in QED ..... 47
17 Renormalisation of QED ..... 50
1 Introduction
References
Srednicki, QFT
Tong, lecture notes on QFT
Peskin and Schroeder, Intro to QFT
Kabat, Lecture notes on Particle Physics
Overview
Path integrals
Loops and Renormalization
Spin 1 (Maxwell)
Spin 1//21 / 2 (Dirac)
QED
2 Path Integral in QM
Based on Srednicki Chap 6. In the first term you quantised scalar field theory by expanding the field in Fourier modes and promoting the coefficients to creation/annihilation operators. You also learned how to compute correlation functions using Wick's theorem in the interaction picture.
In this term, we will describe a more fundamental approach based on path integrds. This is useful for a number of reasons:
provides a non-perturbative definition of QFT
simplest way to quantise gauge theory
naturally generalises QM
Indeed, from the point of view of path integrals, QM can be thought of as 1d QFT (where dd-dimensional QFT refers to a QFT living in dd space-time dimensions). Let us therefore review path integrals in QM.
Consider the following Hamiltonian for a non-relativistic particle:
where [Q,P]=ℏ[Q, P]=\hbar (set ℏ=1\hbar=1 from now on). Let us compute the probability amplitude for the particle to start at position q^(')q^{\prime} at time t^(')t^{\prime} and end at position q^('')q^{\prime \prime} at time t^('')t^{\prime \prime} :
where Q|q^('):)=q^(')|q^('):)Q\left|q^{\prime}\right\rangle=q^{\prime}\left|q^{\prime}\right\rangle, etc. To do so, let's divide the time interval into N+1N+1 pieces of duration
Then introduce NN complete sets af position eigenstates:
{:(1)(:q^('')|e^(-iH(t^('')-t^(')))|q^('):)=intprod_(j=1)^(N)dq_(j)(:q^('')|e^(-iH delta t)|q_(N):)(:q_(N)|e^(-iH delta t)|q_(N-1):)dots(:q_(1)|e^(-iH delta t)|q^('):):}\begin{equation*}
\left\langle q^{\prime \prime}\right| e^{-i H\left(t^{\prime \prime}-t^{\prime}\right)}\left|q^{\prime}\right\rangle=\int \prod_{j=1}^{N} d q_{j}\left\langle q^{\prime \prime}\right| e^{-i H \delta t}\left|q_{N}\right\rangle\left\langle q_{N}\right| e^{-i H \delta t}\left|q_{N-1}\right\rangle \ldots\left\langle q_{1}\right| e^{-i H \delta t}\left|q^{\prime}\right\rangle \tag{1}
\end{equation*}
Note thati in the limit delta t rarr0\delta t \rightarrow 0
e^(-iH delta t)=e^(-i delta t(P^(2)//2m))e^(-i delta tV(Q))+theta(deltat^(2))e^{-i H \delta t}=e^{-i \delta t\left(P^{2} / 2 m\right)} e^{-i \delta t V(Q)}+\theta\left(\delta t^{2}\right)
where we used exp(A+B)=exp(A)exp(B)exp(-(1)/(2)[A,B]dots)\exp (A+B)=\exp (A) \exp (B) \exp \left(-\frac{1}{2}[A, B] \ldots\right)
(Baker-Campbell-Hausdorf formula).
Now focus on one term in (1):
where we inserted a complete set of momentum eigenstates in the first line and performed the integral over p_(1)p_{1} by completing the square and using
{:(6)int_(-oo)^(oo)dxe^(-ax^(2))=sqrt(pi//a):}\begin{equation*}
\int_{-\infty}^{\infty} d x e^{-a x^{2}}=\sqrt{\pi / a} \tag{6}
\end{equation*}
Repeating this for all the factors in (1),
(:q^('')|e^(-iH(t^('')t^(')))|q^('):)=((m)/(2pi i delta t))^((N+1)/(2))intprod_(k=1)^(N)dq_(k)exp(sum_(j=0)^(N)i[(m)/(2)((q_(j+1)-q_(j))/(delta t))^(2)-V(q_(j))]delta t)\left\langle q^{\prime \prime}\right| e^{-i H\left(t^{\prime \prime} t^{\prime}\right)}\left|q^{\prime}\right\rangle=\left(\frac{m}{2 \pi i \delta t}\right)^{\frac{N+1}{2}} \int \prod_{k=1}^{N} d q_{k} \exp \left(\sum_{j=0}^{N} i\left[\frac{m}{2}\left(\frac{q_{j+1}-q_{j}}{\delta t}\right)^{2}-V\left(q_{j}\right)\right] \delta t\right)
In the continuum limit N rarr ooN \rightarrow \infty, this gives
(:q^('')|e^(-iH(t^('')t^(')))|q^('):)=intDq exp[iint_(t^('))^(t^(''))dtL((q^(˙))(t),q(t))]\left\langle q^{\prime \prime}\right| e^{-i H\left(t^{\prime \prime} t^{\prime}\right)}\left|q^{\prime}\right\rangle=\int \mathcal{D} q \exp \left[i \int_{t^{\prime}}^{t^{\prime \prime}} d t \mathcal{L}(\dot{q}(t), q(t))\right]
where L=(1)/(2)mq^(˙)^(2)-V(q)\mathcal{L}=\frac{1}{2} m \dot{q}^{2}-V(q) is the Lagrangian and
Dq=((m)/(2pi ist))^((N+1)/(2))prod_(k=1)^(N)dq_(k)\mathcal{D} q=\left(\frac{m}{2 \pi i s t}\right)^{\frac{N+1}{2}} \prod_{k=1}^{N} d q_{k}
Summary: the amplitude can be computed by integrating over al paths with fixed endpoints, weighted by e^(iS)e^{i S}, where SS is the action of the path!
Now let us consider inserting the operator QQ time t_(1)t_{1} :
where |q^('),t^('):)=e^(iHt^('))|q^('):)\left|q^{\prime}, t^{\prime}\right\rangle=e^{i H t^{\prime}}\left|q^{\prime}\right\rangle and Q(t_(1))=e^(iHt_(1))Qe^(-iHt_(1))Q\left(t_{1}\right)=e^{i H t_{1}} Q e^{-i H t_{1}} in the Heisenberg picture. Using similar manipulations as before, this will result in a path integral with an insertion of q(t_(1))q\left(t_{1}\right) (the location of the particle at time t_(1)t_{1} for each given path):
Finally, note that if we take t^('')rarr oo,t^(')rarr-oot^{\prime \prime} \rightarrow \infty, t^{\prime} \rightarrow-\infty, this will generically project the external states onto the ground state. To see this, note that
Hence, if t^(')rarr-oot^{\prime} \rightarrow-\infty, the ground state will dominate (assuming (:n∣q^('):)!=0\left\langle n \mid q^{\prime}\right\rangle \neq 0 ).
Hence the following object generates correlators with initial and final states given by the ground state:
Z(f)=intDq exp[iint_(-oo)^(oo)dt(L+fq)]Z(f)=\int \mathcal{D} q \exp \left[i \int_{-\infty}^{\infty} d t(\mathcal{L}+f q)\right]
Now the path integral is over all feld configurations on spacetime (which we take to be 4 d ):
Z(J)=intDphi exp i intd^(4)x[L(del_(mu)phi,phi)+J(x)phi(x)]Z(J)=\int \mathcal{D} \phi \exp i \int d^{4} x\left[\mathcal{L}\left(\partial_{\mu} \phi, \phi\right)+J(x) \phi(x)\right]
where xx is shorthand for x^(mu)=(t, vec(x)),Lx^{\mu}=(t, \vec{x}), \mathcal{L} is the Lagrangian, Dphi propprod_(x)d phi(x)\mathcal{D} \phi \propto \prod_{x} d \phi(x).
Let us first consider free theory:
Actually Delta(x-x^('))\Delta\left(x-x^{\prime}\right) is ill-defined because there is a pole in the denominator when k^(2)=-m^(2)k^{2}=-m^{2}. Need to specify a prescription for avoiding this pole:
where epsilon > 0\epsilon>0. This is known as the Feynman propagator. Other prescriptions are possible, but this gives rise to time ordered propagation, so this is the one we will use (since we are interested in time-ordered path integrals). To see this, perform k^(0)k^{0} integral:
where we have expnded in epsilon\epsilon and rescaled it by positive numbers. If t > t^(')t>t^{\prime}, then close in lower hall plane so that numentor decays as |k^(0)|rarr oo\left|k^{0}\right| \rightarrow \infty. This picks up
the residue at k^(0)=omega_(k)-i epsilonk^{0}=\omega_{k}-i \epsilon. Similarly if t < t^(')t<t^{\prime}, then close the contour in upper hall plane and pick up residue at k^(0)=-omega_(k)+i epsilonk^{0}=-\omega_{k}+i \epsilon. In the end, we obtain
Delta(x-x^('))=i theta(t-t^('))int widetilde(dk)e^(ik*(x-x))+i theta(t^(')-t)int widetilde(dk)e^(-ik*(x-x^(')))\Delta\left(x-x^{\prime}\right)=i \theta\left(t-t^{\prime}\right) \int \widetilde{d k} e^{i k \cdot(x-x)}+i \theta\left(t^{\prime}-t\right) \int \widetilde{d k} e^{-i k \cdot\left(x-x^{\prime}\right)}
where widetilde(dk)=(d^(3)k)/((2pi)^(3)2omega_(k))\widetilde{d k}=\frac{d^{3} k}{(2 \pi)^{3} 2 \omega_{k}} and k^(mu)=(omega_(k),( vec(k)))k^{\mu}=\left(\omega_{k}, \vec{k}\right). Hence, propagation is indeed time ordered.
where the brakcet is a Wick contraction. More generally odd-point correlators will vanish since there will always be a left-over JJ in the prefator after taking an odd number of functional derivatives, and even-point correlators will be given by sum over all Wick contractions. This is Wick's theorem, which you proved last term using oscillators. As another example, consider the 4-point correlator:
where we will not worry about normalisation for now.
We can visualise the terms in this sum using Feynman diagrams:
A term in the sum with a particular value of VV and PP corresponds to a diagram with VV vertices and PP propagators. The number of surviving sources after taking all the functional derivatives is
E=2P-4VE=2 P-4 V
Diagrams with E=0E=0 are known as vacuum bubbles:
E=0,V=1quad E=0,V=2E=0, V=1 \quad E=0, V=2
There are no diagrams with an odd number of external sources. Here are some examples for E=2E=2 :
E=2,V=0E=2, V=0
E=2,V=1E=2, V=1
And some examples for E=4E=4 : ++ E=4,V=1E=4, V=1 E=4,V=1E=4, V=1
ง E=4,V=0E=4, V=0
Hence, the sum in (8) can be expressed as a sum over Feynman diagrams. The examples given above are connected diagrams, but the sum also contains disconnected diagrams, which can be expressed as products of connected ones:
E=4,V=0E=4, V=0
E=4,V=1E=4, V=1
We will soon see that Z(J)Z(J) can be expressed in terms of the exponential of the sum of connected diagrams, so let's focus on those for now.
Question: What is the coefficient of a given connected diagram appearing in the sum in (8)? First note that
In general, the ceofficient of a connected diagram corresponds to 1//S1 / S, where SS is the symmetry factor of the diagram. The two-point diagram
has symmetry factor S=2S=2 because it is invariant under exchange of the two sources. Moreover, the vacuum diagram
has a symmetry factor 2 coming from the reflection symmetry of each bubble and another factor of 2 coming the freedom to exchange the two bubbles, giving a total symmetry factor of S=2^(3)=8S=2^{3}=8.
Using similar reasoning you can read of the following symmetry factors:
If in doubt, you can always compute S^(-1)S^{-1} by brute force by computing functional derivatives.
Let us return to the sum over all diaghams in Z(J)Z(J). Let C_(I)C_{I} stand for a particular connected diagram II, including its inverse symmetry factor. A general diagram DD can then be expressed as
where I//S_(D)I / S_{D} is an additional symmetry factor and the product is over distinct connected diagrams . If there are n_(I)n_{I} copies of diagram II, then there is a symmerry factor of n_(I)n_{I} ! associated with exchanging all the copies. Hence
where the sum over {n_(I)}\left\{n_{I}\right\} denotes the sum over sets of diagrams with n_(I)n_{I} copies of each distinct connected diagram II.
In summary, Z(J)Z(J) is proportional to the sum of connected diagrams. It follows that Z(0)Z(0) is proportional to the sum of connected vacuum diagrams (those with no sources). Hence,
where iW(J)i W(J) is the sum of connected non-vacuum diagrams including inverse symmetry factors. In practice, iW(J)i W(J) is the object of interest for computing scattering amplitudes. Before we explain how to compute scattering amplitudes using Feynman diagrams, let's review how to extract them from correlators.
5 LSZ reduction
This is based on Srednick Chapter 5). Let us review how to extract scattering amplitudes from correlation functions. First recall that
phi(x)=int widetilde(dk)(a(( vec(k)))e^(ik*x)+a^(†)(( vec(k)))e^(-ik*x))\phi(x)=\int \widetilde{d k}\left(a(\vec{k}) e^{i k \cdot x}+a^{\dagger}(\vec{k}) e^{-i k \cdot x}\right)
where k*x=-omega_(k)t+ vec(k)* vec(x),omega_(k)=sqrt( vec(k)^(2)+m^(2))k \cdot x=-\omega_{k} t+\vec{k} \cdot \vec{x}, \omega_{k}=\sqrt{\vec{k}^{2}+m^{2}}, and widetilde(dk)=(d^(3)k)/((2pi)^(3)2omega_(k))\widetilde{d k}=\frac{d^{3} k}{(2 \pi)^{3} 2 \omega_{k}}. It follows that
{:[ intd^(3)xe^(-ik*x)del_(t)phi(x)=(1)/(2omega_(k))((-iomega_(k))a(( vec(k)))+(iomega_(k))e^(2iomega_(k)t)a^(†)(-( vec(k))))],[=-(i)/(2)a( vec(k))+(i)/(2)e^(2iomega_(k)t)a^(†)(- vec(k))]:}\begin{aligned}
& \int d^{3} x e^{-i k \cdot x} \partial_{t} \phi(x)=\frac{1}{2 \omega_{k}}\left(\left(-i \omega_{k}\right) a(\vec{k})+\left(i \omega_{k}\right) e^{2 i \omega_{k} t} a^{\dagger}(-\vec{k})\right) \\
& =-\frac{i}{2} a(\vec{k})+\frac{i}{2} e^{2 i \omega_{k} t} a^{\dagger}(-\vec{k})
\end{aligned}
Hence,
{:[a( vec(k))=intd^(3)xe^(-ikx)(idel_(t)phi(x)+omega_(k)phi(x))],[=i intd^(3)xe^(-ik*x)del^(harr)_(t)phi(x)","quad fdel^(harr)_(mu)g=fdel_(mu)g-del_(mu)fg],[ rarra^(†)( vec(k))=-i intd^(3)xe^(ik*x)del^(harr)_(t)phi(x)]:}\begin{aligned}
a(\vec{k}) & =\int d^{3} x e^{-i k x}\left(i \partial_{t} \phi(x)+\omega_{k} \phi(x)\right) \\
& =i \int d^{3} x e^{-i k \cdot x} \stackrel{\leftrightarrow}{\partial}_{t} \phi(x), \quad f \stackrel{\leftrightarrow}{\partial}_{\mu} g=f \partial_{\mu} g-\partial_{\mu} f g \\
& \rightarrow a^{\dagger}(\vec{k})=-i \int d^{3} x e^{i k \cdot x} \stackrel{\leftrightarrow}{\partial}_{t} \phi(x)
\end{aligned}
In the interacting theory, (a,a^(†))\left(a, a^{\dagger}\right) become time-dependent. Assume that particles are non-interacting at t=+-oot= \pm \infty. Suppose we have two incoming and two outgoing particles:
Incoming state: |i:)=lim_(t rarr-oo)a^(†)( vec(k)_(1),t)a^(†)( vec(k)_(2),t)|0:)|i\rangle=\lim _{t \rightarrow-\infty} a^{\dagger}\left(\vec{k}_{1}, t\right) a^{\dagger}\left(\vec{k}_{2}, t\right)|0\rangle
Final state: |f:)=lim_(t rarr+oo)a^(†)( vec(k)_(3),t)a^(†)( vec(k)_(4),t)|0:)|f\rangle=\lim _{t \rightarrow+\infty} a^{\dagger}\left(\vec{k}_{3}, t\right) a^{\dagger}\left(\vec{k}_{4}, t\right)|0\rangle
The scattering amplitude is then given by:
{:[a^(†)( vec(k)","oo)-a^(†)( vec(k)","-oo)=int_(-oo)^(oo)dtdel_(t)a^(†)( vec(k)","t)],[=-i intd^(4)xdel_(t)(e^(ik*x)del^(harr)_(t)phi(x))],[=-i intd^(4)xe^(ik*x)(del_(t)^(2)+w_(k)^(2))phi(x)],[=-i intd^(4)xe^(ik*x)(del_(t)^(2)-grad ^(larr)^(2)+m^(2))phi(x)],[=-i intd^(4)xc^(ik*x)(-del^(2)+m^(2))phi(x)","quaddel^(2)=-del_(t)^(2)+ vec(grad)^(2)],[:.quada^(†)( vec(k)","-oo)=a^(†)( vec(k)","+oo)+i intd^(4)xe^(ik*x)(-del^(2)+m^(2))phi(x)],[a( vec(k)","+oo)=a( vec(k)","-oo)+i intd^(4)xe^(-ik*x)(-del^(2)+m^(2))phi(x)]:}\begin{aligned}
& a^{\dagger}(\vec{k}, \infty)-a^{\dagger}(\vec{k},-\infty)=\int_{-\infty}^{\infty} d t \partial_{t} a^{\dagger}(\vec{k}, t) \\
& =-i \int d^{4} x \partial_{t}\left(e^{i k \cdot x} \stackrel{\leftrightarrow}{\partial}_{t} \phi(x)\right) \\
& =-i \int d^{4} x e^{i k \cdot x}\left(\partial_{t}^{2}+w_{k}^{2}\right) \phi(x) \\
& =-i \int d^{4} x e^{i k \cdot x}\left(\partial_{t}^{2}-\overleftarrow{\nabla}^{2}+m^{2}\right) \phi(x) \\
& =-i \int d^{4} x c^{i k \cdot x}\left(-\partial^{2}+m^{2}\right) \phi(x), \quad \partial^{2}=-\partial_{t}^{2}+\vec{\nabla}^{2} \\
& \therefore \quad a^{\dagger}(\vec{k},-\infty)=a^{\dagger}(\vec{k},+\infty)+i \int d^{4} x e^{i k \cdot x}\left(-\partial^{2}+m^{2}\right) \phi(x) \\
& a(\vec{k},+\infty)=a(\vec{k},-\infty)+i \int d^{4} x e^{-i k \cdot x}\left(-\partial^{2}+m^{2}\right) \phi(x)
\end{aligned}
If we now plug these formulae into scattering amplitude, we see that time ordering moves a^(†)(k,+oo)a^{\dagger}(k,+\infty) to the left annihilating (:0|\langle 0| and a( vec(k),-oo)a(\vec{k},-\infty) to the right
annihilating |0:)|0\rangle. Hence we obtain
the diff ops trivialise or "amputate" the external props of the Feynman diagrams which contribute to the amplitude. Moreover, the integrals Fourier transform the corrdator to momentum space, where external momenta are on-shell, i.e. k^(2)=-m^(2)k^{2}=-m^{2}.
Summary: Compute scattering amplitudes by Fourier transforming correlator to momentum space, amputating external legs, and putting them on-shell. In practice, one can also allow the legs to be off-shell (tor example when renormalising). We will refer to such quantities as amputated correlators.
6 Scattering Amplitudes and Feynman Rules
This is based on Srednicki Chapter 10 and Peskin section 4.6. Let us apply the formalism developed so far to compute some scattering amplitudes. The first step is to compute a correlation function by taking functional derivatives of the generating functional. In general, this will give connected diagrams as well as disconnected diagrams (i.e. products of connected ones), but we only want to consider connected diagrams. To isolate the connected diagrams recall that
(Z(J))/(Z(0))=exp(iW(J))\frac{Z(J)}{Z(0)}=\exp (i W(J))
where iW(J)i W(J) is the sum of al connected diagrams with external sources (no vacuum bubbles). Hence, we will focus on connected correlators:
(:0|T phi(x_(1))dots phi(x_(n))|0:)_(c)=(1)/(i)(delta)/(delta J(x_(1)))cdots(1)/(i)(delta)/(delta J(x_(n)))iW(J)|_(J=0)\langle 0| T \phi\left(x_{1}\right) \ldots \phi\left(x_{n}\right)|0\rangle_{c}=\left.\frac{1}{i} \frac{\delta}{\delta J\left(x_{1}\right)} \cdots \frac{1}{i} \frac{\delta}{\delta J\left(x_{n}\right)} i W(J)\right|_{J=0}
We then apply the LSZ reduction formula to extract the amplitude.
As an example, let's compute a 3 -point scattering amplitude in phi^(3)\phi^{3} theory:
where the factor of 3 ! arises from the functional derivatives.
Now suppose that legs 1 and 2 are incoming and leg 3 is outgoing. LSZ then gives (:3∣1,2:)=\langle 3 \mid 1,2\rangle=
In general we will always get a delta function imposing momentum conservation so we will drop it:
{:[quad(:f∣i:)=(2pi)^(4)g^(4)(k_("in ")-k_("out "))iM],[ rarr iM_(3)=-ig]:}\begin{aligned}
& \quad\langle f \mid i\rangle=(2 \pi)^{4} g^{4}\left(k_{\text {in }}-k_{\text {out }}\right) i \mathcal{M} \\
& \rightarrow i \mathcal{M}_{3}=-i g
\end{aligned}
Note that this can be obtained from the following Feynman rule:
Now consider a 4-point amplitude. First we must compute
(:0|T phi(x_(1))dots phi(x_(4))|0:)_(c)=(1)/(i)(delta)/(delta J(x_(1)))cdots(1)/(i)(delta)/(delta J(x_(4)))iW(J)|_(J=0)\langle 0| T \phi\left(x_{1}\right) \ldots \phi\left(x_{4}\right)|0\rangle_{c}=\left.\frac{1}{i} \frac{\delta}{\delta J\left(x_{1}\right)} \cdots \frac{1}{i} \frac{\delta}{\delta J\left(x_{4}\right)} i W(J)\right|_{J=0}
At leading order, this will come from acting with functional derivatives on the following connected diagram with four external sources:
iW(J)=(1)/(8)i W(J)=\frac{1}{8}
There are 4 ! ways to act with the 4delta//delta J4 \delta / \delta J on the 4 sources. These 24 diagrams can be collected into 3 groups of 8 identical diagrams.
This factor of 8 cancels the symmetry factor. We then obtain
Now plug in Delta(y-z)=int(d^(4)k)/((2)^(4))(e^(ik*(y-z)))/(k^(2)+m^(2))\Delta(y-z)=\int \frac{d^{4} k}{(2)^{4}} \frac{e^{i k \cdot(y-z)}}{k^{2}+m^{2}}. Then we obtain
{:[(:f∣i:)=(-ig)^(2)(2pi)^(4)delta^(4)(k_(1)+k_(2)-k_(3)-k_(4))[(-i)/((k_(1)+k_(2))^(2)+m^(2))+(-i)/((k_(1)-k_(4))^(2)+m^(2))+(-i)/((k_(1)-k_(3))^(2)+m^(2))]],[ rarr iM_(4)=(-ig)^(2)[(-i)/((k_(1)+k_(2))^(2)+m^(2))+(-i)/((k_(1)-k_(4))^(2)+m^(2))+(-i)/((k_(1)-k_(3))^(2)+m^(2))]]:}\begin{aligned}
& \langle f \mid i\rangle=(-i g)^{2}(2 \pi)^{4} \delta^{4}\left(k_{1}+k_{2}-k_{3}-k_{4}\right)\left[\frac{-i}{\left(k_{1}+k_{2}\right)^{2}+m^{2}}+\frac{-i}{\left(k_{1}-k_{4}\right)^{2}+m^{2}}+\frac{-i}{\left(k_{1}-k_{3}\right)^{2}+m^{2}}\right] \\
& \rightarrow i \mathcal{M}_{4}=(-i g)^{2}\left[\frac{-i}{\left(k_{1}+k_{2}\right)^{2}+m^{2}}+\frac{-i}{\left(k_{1}-k_{4}\right)^{2}+m^{2}}+\frac{-i}{\left(k_{1}-k_{3}\right)^{2}+m^{2}}\right]
\end{aligned}
This could have been directly calculated from the following Feynman diagrams in momentum space:
where arrows indicate momentum flow, we associate -ig-i g to each vertex, and use the following Feynman rule for internal lines:
For external lines, we simply have a factor of 1 because of LSZ. We also impose momentum conservation at every vertex.
In general, we can compute any scattering amplitude by drawing all connected Fexnman diagrams contributing up to a given order in the coupling and evaluating each diagram using the rules above. In practice we can deduce the Feynman rules more directly using Wick's theorem in the interaction picture:
where we expanded e^(-iH_("int "))e^{-i H_{\text {int }}} to O(g)\mathcal{O}(g) and ignore spacetime location of the fields. We also introduced the rule
which follows from LSZ. This corresponds to an external line of a Feynman diagram. Contractions of phi\phi 's give internal lines. We first contracted (:3|\langle 3| with one of the three phi\phi 's giving a factor of 3 , then contracted |2:)|2\rangle with one of the 2 ramaining phi\phi 's giving a factor of 2 , and then contracted |1:)|1\rangle with the last remaing phi\phi. For derivative interactions, take del_(mu)rarr+-ik_(mu)\partial_{\mu} \rightarrow \pm i k_{\mu} when acting on an incoming/outgoing leg with momertun k_(mu)k_{\mu}. Contractions giving internal lines are replaced by the momentum space propagator given above.
7 Loops and Renormalisation
This is based on Srednicki chapters 14, 29 and Kabat section 7.2.1. So far we have only talked about tree-level Feynman diagrams. Now let's discuss looplevel diagrams, ie. diagrams with closed loops. Tree-level diagrams encode classical physics, while loop-level diagrams encode quantum corrections (this
can be seen by restoring ℏ\hbar ).
To compute loop-level diagrams, we can use the Feynman rues described before, along with two more rules:
for each closed loop, there will be an unfixed internal momentum l^(mu)l^{\mu} flowing through the loop. Integrate each loop momatum with measure d^(4)l//(2pi)^(4)d^{4} l /(2 \pi)^{4}.
Divide by the symmetry factor associated with exchanging only internal lines.
Let's illustrate how this works at 1-loop 4-point in phi^(4)\phi^{4} theory:
L=-(1)/(2)del_(mu)phidel^(u)phi-(1)/(2)m^(2)phi^(2)-(1)/(4!)gphi^(4)\mathcal{L}=-\frac{1}{2} \partial_{\mu} \phi \partial^{u} \phi-\frac{1}{2} m^{2} \phi^{2}-\frac{1}{4!} g \phi^{4}
Take legs 1,2 incoming and legs 3,4 outgoing:
We will refer to the the three diagrams on the right-hand-side as the s , t , and u channels, respectively. Let's compute the first diagram using Wick contractions:
where we've expanded exp(-iH_("int "))\exp \left(-i H_{\text {int }}\right) to O(g^(2))\mathcal{O}\left(g^{2}\right) and the subscripts x,yx, y simply remind us that the vertices are at different locations. Performing the Wick contractions gives
where we first contracted leg 1 with one the four phi_(y)\phi_{y} 's, giving a factor of 4 as well as a factor of 2 because we could have also contracted it onto one of the phi_(x)\phi_{x} 's. Then we contracted leg 2 with one of the three remaining phi_(y)\phi_{y} 's giving a factor of 3 (if we contracted leg 2 onto a different vertex than leg 1 , this would give the second or third diagram). Then we contracted leg 4 with one of the four phi_(x)\phi_{x} 's giving a factor of 4 , and leg 3 with one of the remaining 3phi_(x)3 \phi_{x} 's giving a factor of 3 . Finally, we contracted a phi_(x)\phi_{x} with one of the two remaining phi_(y)\phi_{y} 's giving
a factor of 2 , and contracted the remaining phi_(x)\phi_{x} with the remaing phi_(y)\phi_{y}. The two internal contractions give two internal propagators which form a closed loop.
where the factor of (1)/(2)\frac{1}{2} is a symmetry factor associated with exchanging internal lines. Note that the integral is divergent. This can be seen by taking the l rarr ool \rightarrow \infty limit of the integrand. In general if a loop integral goes like
intd^(4)ll^(D-4)\int d^{4} l l^{D-4}
and D >= 0D \geq 0, then the "superficial degree of divergence" is DD. For example, if D=2D=2 then the integral will be quadratically divergent. In the present example D=0D=0, which corresponds to a logarithmic divergence. These divergences must be regulated. A standard regulator is dimensional regularisation where we take spacetime dimension dd to be d=4-epsilond=4-\epsilon. In this course, we will use a cut-off on the loop momentum for pedagogical reasons. Let's see how this works.
First introduce "Feynman parameters":
(1)/(AB)=int_(0)^(1)dx(1)/((xA+(1-x)B)^(2))\frac{1}{A B}=\int_{0}^{1} d x \frac{1}{(x A+(1-x) B)^{2}}
This can be used to write the loop integral as follows:
{:[int(d^(4)l)/((l^(2)+m^(2))((l+k_(1)+k_(2))^(2)+m^(2)))=intd^(4)lint_(0)^(1)dx[x((l+k_(1)+k_(2))^(2)+m^(2))+(1-x)(l^(2)+m^(2))]^(-2)],[=int_(0)^(1)dx intd^(4)q(q^(2)+M^(2))^(-2)]:}\begin{aligned}
\int \frac{d^{4} l}{\left(l^{2}+m^{2}\right)\left(\left(l+k_{1}+k_{2}\right)^{2}+m^{2}\right)} & =\int d^{4} l \int_{0}^{1} d x\left[x\left(\left(l+k_{1}+k_{2}\right)^{2}+m^{2}\right)+(1-x)\left(l^{2}+m^{2}\right)\right]^{-2} \\
& =\int_{0}^{1} d x \int d^{4} q\left(q^{2}+M^{2}\right)^{-2}
\end{aligned}
where q=l+x(k_(1)+k_(2))q=l+x\left(k_{1}+k_{2}\right) and M^(2)=x(1-x)(k_(1)+k_(2))^(2)+m^(2)M^{2}=x(1-x)\left(k_{1}+k_{2}\right)^{2}+m^{2}. Next, Wick rotate to Euclidean signature. Let q^(mu)=(i bar(q)^(0), bar(q)^(1), bar(q)^(2), bar(q)^(3))rarrq^(2)=( bar(q)^(0))^(2)+( bar(q))^(2)+( bar(q)^(2))^(2)+q^{\mu}=\left(i \bar{q}^{0}, \bar{q}^{1}, \bar{q}^{2}, \bar{q}^{3}\right) \rightarrow q^{2}=\left(\bar{q}^{0}\right)^{2}+(\bar{q})^{2}+\left(\bar{q}^{2}\right)^{2}+( bar(q)^(3))^(2)\left(\bar{q}^{3}\right)^{2}. The the loop integral can be written as
iint_(0)^(1)dx intd^(4) bar(q)( bar(q)^(2)+M^(2))^(-2)i \int_{0}^{1} d x \int d^{4} \bar{q}\left(\bar{q}^{2}+M^{2}\right)^{-2}
We can then foliate R^(4)\mathbb{R}^{4} with 3 -spheres and write d^(4) bar(q)=2pi^(2) bar(q)^(3)d bar(q)d^{4} \bar{q}=2 \pi^{2} \bar{q}^{3} d \bar{q}, where 2pi^(2)2 \pi^{2} is the area of a unit 3 -sphere and bar(q)\bar{q} is the radius. In summary, we find that
int(d^(4)l)/((l^(2)+m^(2))((l+k_(1)+k_(2))^(2)+m^(2)))=2pi^(2)iint_(0)^(1)dxint_(0)^(Lambda)d bar(q) bar(q)^(3)( bar(q)^(2)+M^(2))^(-2)\int \frac{d^{4} l}{\left(l^{2}+m^{2}\right)\left(\left(l+k_{1}+k_{2}\right)^{2}+m^{2}\right)}=2 \pi^{2} i \int_{0}^{1} d x \int_{0}^{\Lambda} d \bar{q} \bar{q}^{3}\left(\bar{q}^{2}+M^{2}\right)^{-2}
where we have introduced a cutoff Lambda\Lambda. The integral over bar(q)\bar{q} is given by int_(0)^(Lambda)d bar(q) bar(q)^(3)( bar(q)^(2)+M^(2))^(-2)=(1)/(2)[ln((Lambda^(2)+M^(2))/(M^(2)))-(Lambda^(2))/(Lambda^(2)+M^(2))]∼(1)/(2)(ln((Lambda^(2))/(M^(2)))-1)\int_{0}^{\Lambda} d \bar{q} \bar{q}^{3}\left(\bar{q}^{2}+M^{2}\right)^{-2}=\frac{1}{2}\left[\ln \left(\frac{\Lambda^{2}+M^{2}}{M^{2}}\right)-\frac{\Lambda^{2}}{\Lambda^{2}+M^{2}}\right] \sim \frac{1}{2}\left(\ln \left(\frac{\Lambda^{2}}{M^{2}}\right)-1\right), for Lambda^(2)≫M^(2)\Lambda^{2} \gg M^{2}
Including the tree-level contribution, the total 4-point amplitude is depicted below
and is given by
where s,t,us, t, u are known as Mandelstam variables.
From the point of view of the path integral, we can think of Lambda\Lambda as a cutoff above which we integrate out all the higher energy modes of the field (you will explore this in the HW). Hence, we should think of this as describing an effective field theory of modes with energy below the cut-off. For low energy observers, the amplitude must be independent of the cutoff. If we allow gg to depend on Lambda\Lambda, we then have
where we neglect ℏubrace(dg//d ln Lambdaubrace)_(O(ℏ))=O(ℏ^(2))\hbar \underbrace{d g / d \ln \Lambda}_{\mathcal{O}(\hbar)}=\mathcal{O}\left(\hbar^{2}\right). From this we compute the beta function
where 1//g(mu)1 / g(\mu) is an integration constant, g(Lambda)g(\Lambda) is the bare coupling, g(mu)g(\mu) is the renormalised coupling (measured by an experiment), and mu\mu is the renormalisation scale (in practice, this is the energy scale of the experiment). This formula tells us how the coupling g(Lambda)g(\Lambda) has to change in order to compensate for a change in Lambda\Lambda. For each value of g(mu)g(\mu), there is a family of EFT's related by "renormalisation group flow" whose physical consequences are the same:
Note that g(Lambda)g(\Lambda) diverges at Lambda_(max)=mu exp(16pi^(2)//3g(mu))\Lambda_{\max }=\mu \exp \left(16 \pi^{2} / 3 g(\mu)\right). This is known as a "Landau pole". Some new physics has to kick in before the scale Lambda_(max)\Lambda_{\max }. Equivalently, it we hold Lambda\Lambda and g(Lambda)g(\Lambda) fixed, then we see that renormalised coupling g(mu)g(\mu) must change as we vary renormalisation scale mu\mu. This is referred to as a "running coupling." At high energy, the theory becomes strongly coupled so perturbation theory not reliable.
Let us express the amplitude in terms of g(mu)g(\mu). First note that
Plugging this back into amplitude and dropping terms of O(ℏ^(2))\mathcal{O}\left(\hbar^{2}\right) then gives iM_(4)=-ig(mu)+(iℏg(mu)^(2))/(32pi^(2))int_(0)^(1)dx[ln((mu^(2))/(M^(2)(s)))+ln((mu^(2))/(M^(2)(t)))+ln((mu^(2))/(M^(2)(mu)))-3]i \mathcal{M}_{4}=-i g(\mu)+\frac{i \hbar g(\mu)^{2}}{32 \pi^{2}} \int_{0}^{1} d x\left[\ln \left(\frac{\mu^{2}}{M^{2}(s)}\right)+\ln \left(\frac{\mu^{2}}{M^{2}(t)}\right)+\ln \left(\frac{\mu^{2}}{M^{2}(\mu)}\right)-3\right]
Hence, the Lambda\Lambda-dependence cancels out and the amplitude is explicitly finite. It has been "renormalised."
Let us now consider more general beta\beta function:
Hence, coupling goes to zero at high energies and becomes strong at low energies. This is known as "asymptotic freedom."
Now suppose that beta(g_(**))=0\beta\left(g_{*}\right)=0 for some particular value g_(**)g_{*}. Then g_(**)g_{*} is a "fixed point" because if the coupling starts at g_(**)g_{*}, it will stay there as we vary Lambda\Lambda since
If beta^(')(g_(**)) < 0\beta^{\prime}\left(g_{*}\right)<0, then gg is driven to g_(**)g_{*} as Lambda\Lambda increases, correspponding to "ultraviolet (UV) fixed point." If beta^(')(g_(**)) > 0\beta^{\prime}\left(g_{*}\right)>0, then gg is driven away from g_(**)g_{*} as Lambda\Lambda increases, corresponding to an "infrared (IR) fixed point" (see HW). Such fixed points can describe phase transitions and quantum gravity via AdS/CFT.
8 Classical Maxwell Theory
This is based on Srednicki Chapters 54, 55. Let us recall Maxwell's equations:
where epsilon^(ijk)=-epsilon^(jk)=-epsilon^(ikj),epsilon^(123)=1\epsilon^{i j k}=-\epsilon^{j k}=-\epsilon^{i k j}, \epsilon^{123}=1 and indices are raised and lowered using eta_(mu nu)=diag{-1,1,1,1}\eta_{\mu \nu}=\operatorname{diag}\{-1,1,1,1\}. Now Maxwell's equations take the following form:
Maxwell's equations are invariant under A_(mu)rarrA_(mu)-del_(mu)GammaquadA_{\mu} \rightarrow A_{\mu}-\partial_{\mu} \Gamma \quad (gauge transformations). Indeed, under this transformation
Using deltaA_(mu)=-del_(mu)Gamma\delta A_{\mu}=-\partial_{\mu} \Gamma, we can impose Lorenz gauge: del_(mu)A^(mu)=0\partial_{\mu} A^{\mu}=0. Setting J^(mu)=0J^{\mu}=0, the EOM reduce to
These are just the EOM of 4 massless scalars! On the other hand, there are fewer that 4 degrees of freedom due to gauge-invariace. To see this, let's Fourier transform to momentum space:
A_(mu)(x)=int(d^(4)k)/((2pi)^(4))e^(ik*x)A_(mu)(k)A_{\mu}(x)=\int \frac{d^{4} k}{(2 \pi)^{4}} e^{i k \cdot x} A_{\mu}(k)
where delta(k^(2))\delta\left(k^{2}\right) enforces the equations of motion and epsilon_(mu)\epsilon_{\mu} is a polarisation vector. After doing so, the gauge condition corresponds to k*epsilon(k)=0k \cdot \epsilon(k)=0 and the residual gauge transformation corresponds to epsilon_(mu)(k)rarrepsilon_(mu)(k)+i omega(k)k_(mu)\epsilon_{\mu}(k) \rightarrow \epsilon_{\mu}(k)+i \omega(k) k_{\mu}. Hence epsilon_(mu)\epsilon_{\mu} and epsilon_(mu)+i alphak_(mu)\epsilon_{\mu}+i \alpha k_{\mu} are physically equivalent, where alpha\alpha is arbitrary and k^(2)=0k^{2}=0.
Since epsilon\epsilon satisfies epsilon*k=0\epsilon \cdot k=0 and has a residual gauge symmetry, it has two physical degrees of freedom, epsilon_(+-)\epsilon_{ \pm}. To see this more explicitly, choose a frame where
Then epsilon*k=0rarrepsilon^(mu)=(0,b,c,0)+a(1,0,0,1)\epsilon \cdot k=0 \rightarrow \epsilon^{\mu}=(0, b, c, 0)+a(1,0,0,1). Using the residual gauge symmetry, we may then take
The two independent solutions are then given by epsilon_(+-)^(mu)=(1)/(sqrt2)(0,1,∓i,0)\epsilon_{ \pm}^{\mu}=\frac{1}{\sqrt{2}}(0,1, \mp i, 0), where we have fixed the normalisation. These correspond to positive and negative helicity. In summary, the polarisation vectors satisfy
where we assume that k^(2)=0k^{2}=0. These properties were derived in a particular frame, but they are Lorentz-invarlient so hold in all frames.
Hence, we find the following solution to the EOM:
A^(mu)(x)=int(d^(3)k)/((2pi)^(3))int dk^(0)delta(k^(2))theta(k^(0))sum_(h=+-)[a_(h)(( vec(k)))epsilon_(-h)^(mu)(( vec(k)))e^(ik*x)+a_(h)^(**)(( vec(k)))epsilon_(h)^(mu)(( vec(k)))e^(-ik*x)]A^{\mu}(x)=\int \frac{d^{3} k}{(2 \pi)^{3}} \int d k^{0} \delta\left(k^{2}\right) \theta\left(k^{0}\right) \sum_{h= \pm}\left[a_{h}(\vec{k}) \epsilon_{-h}^{\mu}(\vec{k}) e^{i k \cdot x}+a_{h}^{*}(\vec{k}) \epsilon_{h}^{\mu}(\vec{k}) e^{-i k \cdot x}\right]
where k^(mu)=(omega_(k),( vec(k)))k^{\mu}=\left(\omega_{k}, \vec{k}\right). Note that A^(mu)=(A^(mu))^(**)A^{\mu}=\left(A^{\mu}\right)^{*} since epsilon_(h)^(**)=epsilon_(-h)\epsilon_{h}^{*}=\epsilon_{-h}. Using
int dx delta(f(x))=sum_(i)|1//f^(')(x_(i))|,quad f(x_(i))=0\int d x \delta(f(x))=\sum_{i}\left|1 / f^{\prime}\left(x_{i}\right)\right|, \quad f\left(x_{i}\right)=0
we then perform the k^(0)k^{0} integral to obtain
int widetilde(dk)sum_(h=+-)[epsilon_(-h)^(mu)(( vec(k)))a_(h)(( vec(k)))e^(ik*x)+epsilon_(h)^(mu)(( vec(k)))a_(h)^(**)(( vec(k)))e^(-ik*x)]\int \widetilde{d k} \sum_{h= \pm}\left[\epsilon_{-h}^{\mu}(\vec{k}) a_{h}(\vec{k}) e^{i k \cdot x}+\epsilon_{h}^{\mu}(\vec{k}) a_{h}^{*}(\vec{k}) e^{-i k \cdot x}\right]
Note that this is essatially the mode expansion for two scalar fields dressed with polarisations. In particular, we can quantise by promoting the Fourier coefficients to creation and annihilation operators, as we did for the scalar field.
9 LSZ and Path Integral For Photons
This is based on Srednicki Chapters 56,57,6256,57,62. We previously found that the general solution to Maxwell theory in the absence of sources is given by
where a_(h)^(†)( vec(k))//a_(h)( vec(k))a_{h}^{\dagger}(\vec{k}) / a_{h}(\vec{k}) creates/annihilates a photon with momentum vec(k)\vec{k} and polarisation epsilon_(h)^(mu)\epsilon_{h}^{\mu}. Moreover, we can obtain a formula for creation and annihilation operators in exactly the same way we did for scalar fields:
{:[a_(h)^(†)( vec(k))=-iepsilon_(-h)^(mu)( vec(k))intd^(3)xe^(ik*x)del^(harr)_(t)A^(mu)(x)],[a_(h)( vec(k))=+iepsilon_(h)( vec(k))intd^(3)xe^(-ik*x)del^(harr)_(t)A^(mu)(x)]:}\begin{aligned}
& a_{h}^{\dagger}(\vec{k})=-i \epsilon_{-h}^{\mu}(\vec{k}) \int d^{3} x e^{i k \cdot x} \stackrel{\leftrightarrow}{\partial}_{t} A^{\mu}(x) \\
& a_{h}(\vec{k})=+i \epsilon_{h}(\vec{k}) \int d^{3} x e^{-i k \cdot x} \stackrel{\leftrightarrow}{\partial}_{t} A^{\mu}(x)
\end{aligned}
Following the same steps used to derive LSZ for scalars, we can make the following replacements when computing photon amplitudes:
{:[a_(h)^(†)( vec(k)","-oo) rarr iepsilon_(-h)^(mu)(k^(`)^('))intd^(4)xe^(ik*x)(-del^(2))A_(mu)(x)],[a_(h)(ℏ_(k),+oo) rarr iepsilon_(h)^(mu)(ℏ_(k))intd^(4)xe^(-ik*x)(-del^(2))A_(mu)(x)]:}\begin{aligned}
a_{h}^{\dagger}(\vec{k},-\infty) & \rightarrow i \epsilon_{-h}^{\mu}\left(\grave{k}^{\prime}\right) \int d^{4} x e^{i k \cdot x}\left(-\partial^{2}\right) A_{\mu}(x) \\
a_{h}\left(\hbar_{k},+\infty\right) & \rightarrow i \epsilon_{h}^{\mu}\left(\hbar_{k}\right) \int d^{4} x e^{-i k \cdot x}\left(-\partial^{2}\right) A_{\mu}(x)
\end{aligned}
where the a^(†)a^{\dagger} creates incoming photons and the aa annihilates outgoing photons.
Summary: To compute photon amplitudes, first compute time-ordered correlators of A_(mu)(x)A_{\mu}(x), Fourier transform to momentum space, amputate external propagators, and dress with polarisation vectors. In practice we will not distinguish between epsilon_(+)^(mu)\epsilon_{+}^{\mu} and epsilon_(-)^(mu)\epsilon_{-}^{\mu} and simply dress the amputated correlator with epsilon^(mu)( vec(k))\epsilon^{\mu}(\vec{k}). Due to gauge symmetry, the amplitude must be invariant if we shift an external polarisation by its momentum:
epsilon_(mu)(k)rarrepsilon_(mu)(k)+alpha k mu\epsilon_{\mu}(k) \rightarrow \epsilon_{\mu}(k)+\alpha k \mu
i.e. if we replace a polarisation with its momentum, then the amplitude vanishes. This is known as the "Ward identity."
Now we must understand how to compute time-ordered corrclators of A_(mu)(x)A_{\mu}(x). This is easiest to do using the path integral formalism. We have
The xx-integral gives (2pi)^(4)delta^(4)(k+k^('))(2 \pi)^{4} \delta^{4}\left(k+k^{\prime}\right). Performing k^(')k^{\prime} integral then gives iS_(0)=(i)/(2)int(d^(eta)k)/((2pi)^(4))[-A_(mu)(k)(k^(2)eta^(mu nu)-k^(mu)k^(nu))A_(nu)(-k)+J^(mu)(k)A_(mu)(-k)+J^(mu)(-k)A_(mu)(k)]i S_{0}=\frac{i}{2} \int \frac{d^{\eta} k}{(2 \pi)^{4}}\left[-A_{\mu}(k)\left(k^{2} \eta^{\mu \nu}-k^{\mu} k^{\nu}\right) A_{\nu}(-k)+J^{\mu}(k) A_{\mu}(-k)+J^{\mu}(-k) A_{\mu}(k)\right]
The next step is to complete the square but the 4xx44 \times 4 matrix k^(2)eta^(mu nu)-k^(mu)k^(nu)k^{2} \eta^{\mu \nu}-k^{\mu} k^{\nu} is not invertible. To see this, consider the matrix
Hence the remaining three eigenvalues are all 1. From this, we see that the component of A_(mu)(k)A_{\mu}(k) along k_(mu)k_{\mu} does not contribute to the quadratic term in the action. Moreover, it does not contribute to the linear term since del_(mu)J^(mu)(x)=\partial_{\mu} J^{\mu}(x)=0rarrk^(mu)J_(mu)(k)=00 \rightarrow k^{\mu} J_{\mu}(k)=0. Thus the path integral over this component just gives an oo\infty constant which we can discard.
The non-trivial part of the path integral is over the components of A_(mu)(k)A_{\mu}(k) in the space orthogonal to k_(mu)k_{\mu}. The matrix P^(mu nu)(k)P^{\mu \nu}(k) projects 4 -vectors into this subspace and in this subspace it is the identity matrix. Therefore with in this subspace the inverse of k^(2)P^(mu nu)(k)k^{2} P^{\mu \nu}(k) is 1//k^(2)P_(mu nu)(k)1 / k^{2} P_{\mu \nu}(k) and we can complete the square in the action to give
where xi=1\xi=1 is called "Feynman gauge" and xi=0\xi=0 is called "Lorenz gauge" or "Landau gauge." This form of the propagator can be derived by adding a "gauge-fixing" term to the Lagrangian:
After adding this term and going to momentum space, the 4xx44 \times 4 matrix in the qaudratic part of the action becomes invertable and the inverse is the propagator given above.
In this way, we can compute photon amplitudes by drawing Feynman diagrams. We can also compute them from Wick's theorem in the interaction picture replacing widehat(A)^(mu)A^(nu)\widehat{A}^{\mu} A^{\nu} with the propagator and
where Omega_(rho sigma)=-Omega_(sigma rho)\Omega_{\rho \sigma}=-\Omega_{\sigma \rho} are six numbers that pananeterise the Lorentz transformation ( 3 rotations, 3 boosts) and M^(rho sigma)\mathcal{M}^{\rho \sigma} are generators, whose representation
depends on the spin of the particle. For spin-1, the generator of rotations in the x^(1)-x^(2)x^{1}-x^{2} plane is
[M^(mu nu),M^(rho sigma)]=(eta^(mu sigma)mu^(nu rho)-mu harr nu)-rho harr sigma\left[M^{\mu \nu}, M^{\rho \sigma}\right]=\left(\eta^{\mu \sigma} \mu^{\nu \rho}-\mu \leftrightarrow \nu\right)-\rho \leftrightarrow \sigma
The genentors in any representation must obey this algebra.
For spin-s, a Lorentz transformation becomes trivial for a rotation of 2pi//s2 \pi / \mathrm{s}. For spin- 1 , consider a rotation by theta\theta in the x^(1)-x^(2)x^{1}-x^{2} plane:
Next we will look at spin- 1//21 / 2, which decribes electrons. In this case, we expect that a spatial rotation by 4pi4 \pi will be trivial but not by 2pi2 \pi, i.e. Lambda(4pi)=1\Lambda(4 \pi)=1 but Lambda(2pi)!=1\Lambda(2 \pi) \neq 1. Since Lambda(4pi)=(Lambda(2pi))^(2)rarr Lambda(2pi)=-1\Lambda(4 \pi)=(\Lambda(2 \pi))^{2} \rightarrow \Lambda(2 \pi)=-1. Let us postulate the following generators and verify that they imply this property:
This is known as the "spinor representation." Let's see what rotations and boosts look like. For a rotation by theta\theta in x^(1)-x^(2)x^{1}-x^{2} plane,
We will denote gamma^(mu)del_(mu)=del\gamma^{\mu} \partial_{\mu}=\not \partial. The choice of relative coefficients will become clear when we show the EOM rarr\rightarrow the Klein-Gordon equation.
For a scalar field the Hamiltonian density is H=(delL)/(del(del_(0)phi))del_(0)phi-L\mathcal{H}=\frac{\partial \mathcal{L}}{\partial\left(\partial_{0} \phi\right)} \partial_{0} \phi-\mathcal{L}. Similarly, for a Dirac field
del_(mu)j^(mu)=(del_(mu)( bar(psi)))gamma^(mu)psi+ bar(psi)del psi=im bar(psi)psi-im bar(psi)psi=0.\partial_{\mu} j^{\mu}=\left(\partial_{\mu} \bar{\psi}\right) \gamma^{\mu} \psi+\bar{\psi} \not \partial \psi=i m \bar{\psi} \psi-i m \bar{\psi} \psi=0 .
where we noted that
{:[del psi=-im psi rarr(del psi)^(†)=impsi^(†)],[ rarr im bar(psi)=(del psi)^(†)gamma^(0)=del_(mu)psi^(†)(gamma^(mu))^(†)gamma^(0)=del_(mu) bar(psi)gamma^(mu)]:}\begin{aligned}
\not \partial \psi=-i m \psi & \rightarrow(\not \partial \psi)^{\dagger}=i m \psi^{\dagger} \\
& \rightarrow i m \bar{\psi}=(\not \partial \psi)^{\dagger} \gamma^{0}=\partial_{\mu} \psi^{\dagger}\left(\gamma^{\mu}\right)^{\dagger} \gamma^{0}=\partial_{\mu} \bar{\psi} \gamma^{\mu}
\end{aligned}
The Noether charge corresponding to the current in (10) then given by
Q=intd^(3)xj^(0)=intd^(3)x bar(psi)gamma^(0)psiQ=\int d^{3} x j^{0}=\int d^{3} x \bar{\psi} \gamma^{0} \psi
11 Classical Solutions of the Dirac Equation
This is based on Srednicki pg 238-239 and Tong section 4.7. Recall that the Dirac equation implies Klein-Gordon equation. Hence, it admits plane-wave solutions:
psi(x)=u( vec(p))e^(ip*x)+v( vec(p))e^(-ip*x)\psi(x)=u(\vec{p}) e^{i p \cdot x}+v(\vec{p}) e^{-i p \cdot x}
where p^(0)=omega_(p)=sqrt( vec(p)^(2)+m^(2)),u( vec(p))p^{0}=\omega_{p}=\sqrt{\vec{p}^{2}+m^{2}}, u(\vec{p}) and v( vec(p))v(\vec{p}) are 4 component Dirac spinors. Plugging this into Dirac equation gives
Each of these equations has two linearly independent solutions that we will call u_(s)( vec(p))u_{s}(\vec{p}) and v_(s)( vec(p))v_{s}(\vec{p}), where s=1,2s=1,2. The general solution is then
psi(x)=sum_(s=1,2)int widetilde(dp)[b_(s)(( vec(p)))u_(s)(( vec(p)))e^(ip*x)+d_(s)^(†)(( vec(p)))v_(s)(( vec(p)))e^(-ip*x)]\psi(x)=\sum_{s=1,2} \int \widetilde{d p}\left[b_{s}(\vec{p}) u_{s}(\vec{p}) e^{i p \cdot x}+d_{s}^{\dagger}(\vec{p}) v_{s}(\vec{p}) e^{-i p \cdot x}\right]
where widetilde(dp)=(d^(3)p)/((2pi)^(3)2omega_(p))\widetilde{d p}=\frac{d^{3} p}{(2 \pi)^{3} 2 \omega_{p}}.
Let us now work out u( vec(p))u(\vec{p}) from its EOM:
{:[0=(p+m)u( vec(p))=([m,p*sigma],[p* bar(sigma),m])((u_(L)(( vec(p))))/(u_(R)(( vec(p)))))],[sigma^(mu)=(1"," vec(sigma))"," bar(sigma)^(mu)=(1","- vec(sigma))rarr p*sigma=-p^(0)+ vec(p)* vec(sigma)","p* bar(sigma)=-p^(0)- vec(p)* vec(sigma)]:}\begin{aligned}
& 0=(\not p+m) u(\vec{p})=\left(\begin{array}{cc}
m & p \cdot \sigma \\
p \cdot \bar{\sigma} & m
\end{array}\right)\binom{u_{L}(\vec{p})}{u_{R}(\vec{p})} \\
& \sigma^{\mu}=(\mathbb{1}, \vec{\sigma}), \bar{\sigma}^{\mu}=(\mathbb{1},-\vec{\sigma}) \rightarrow p \cdot \sigma=-p^{0}+\vec{p} \cdot \vec{\sigma}, p \cdot \bar{\sigma}=-p^{0}-\vec{p} \cdot \vec{\sigma}
\end{aligned}𝟙𝟙
u_(L//R)u_{L / R} are called "left/right-handed chiral spinors." We now have
where p_(i)p_(j)sigma^(i)sigma^(j)=(1)/(2){sigma^(i),sigma^(j)}p_(i)p_(j)=delta^(ij)p_(i)p_(j)= vec(p)^(2)p_{i} p_{j} \sigma^{i} \sigma^{j}=\frac{1}{2}\left\{\sigma^{i}, \sigma^{j}\right\} p_{i} p_{j}=\delta^{i j} p_{i} p_{j}=\vec{p}^{2}.
Now let us make the following ansatz: quadu_(L)=A(p*sigma)xi^(')\quad u_{L}=A(p \cdot \sigma) \xi^{\prime} for some 2component spinor xi^(')\xi^{\prime}. The second equation in (11) implies
To make the solution more symmetric, choose A=-1//m,xi^(')=sqrt(-p* bar(sigma))xiA=-1 / m, \xi^{\prime}=\sqrt{-p \cdot \bar{\sigma}} \xi, where square root of a matrix is defined by going to a basis where it is diaganal and taking square root of its eigenvalues. Note that in rest frame, p^(mu)=(m, vec(0))p^{\mu}=(m, \overrightarrow{0}) so
where we noted that p*sigmasqrt(-p* bar(sigma))=sqrt(-p*sigma)sqrt((-p*sigma)(-p* bar(sigma)))=-msqrt(-p*sigma)p \cdot \sigma \sqrt{-p \cdot \bar{\sigma}}=\sqrt{-p \cdot \sigma} \sqrt{(-p \cdot \sigma)(-p \cdot \bar{\sigma})}=-m \sqrt{-p \cdot \sigma}.
It's convenient to introduce a basis xi_(s)\xi_{s} and eta_(s),s=1,2\eta_{s}, s=1,2 such that
Note that we need anticommutators. Using commutators would lead to various inconsistencies such as negative-norm states or a Hamiltonian which is unbounded from below, as you will see in the HW. In general, integer-spin fields must be commuting and half-integer spin fields must be anti-commiting This known as "spin-statistics theorem".
Now recall the expansion:
psi(x)=sum_(s=1,2)int widetilde(dp)[b_(s)(( vec(p)))u_(s)(( vec(p)))e^(ip*x)+d_(s)^(†)(( vec(p)))v_(s)(( vec(p)))e^(-ip*x)]\psi(x)=\sum_{s=1,2} \int \widetilde{d p}\left[b_{s}(\vec{p}) u_{s}(\vec{p}) e^{i p \cdot x}+d_{s}^{\dagger}(\vec{p}) v_{s}(\vec{p}) e^{-i p \cdot x}\right]
Let us use this to express b,db, d in terms of psi,psi^(†)\psi, \psi^{\dagger} and then deduce the algebra of b,db, d. We have that
Similarly, computing intd^(3)xe^(ip*x)psi(x)\int d^{3} x e^{i p \cdot x} \psi(x), multiplying by v_(s)^(†)( vec(p))v_{s}^{\dagger}(\vec{p}), and using v_(s)^(†)( vec(p))v_(s^('))( vec(p))=v_{s}^{\dagger}(\vec{p}) v_{s^{\prime}}(\vec{p})=2omega_(p)delta_(ss^(')),v_(s)^(†)( vec(p))u_(s^('))(- vec(p))=02 \omega_{p} \delta_{s s^{\prime}}, v_{s}^{\dagger}(\vec{p}) u_{s^{\prime}}(-\vec{p})=0, we find
d_(s)^(†)( vec(p))=intd^(3)xe^(ip*x)v_(s)^(†)( vec(p))psi(x)d_{s}^{\dagger}(\vec{p})=\int d^{3} x e^{i p \cdot x} v_{s}^{\dagger}(\vec{p}) \psi(x)
Taking complex conjugates:
{:[b_(s)^(†)( vec(p))=intd^(3)xe^(ip*x)psi^(†)(x)u_(s)( vec(p))],[d_(s)( vec(p))=intd^(3)xe^(-ip*x)psi^(†)(x)v_(s)( vec(p))]:}\begin{aligned}
& b_{s}^{\dagger}(\vec{p})=\int d^{3} x e^{i p \cdot x} \psi^{\dagger}(x) u_{s}(\vec{p}) \\
& d_{s}(\vec{p})=\int d^{3} x e^{-i p \cdot x} \psi^{\dagger}(x) v_{s}(\vec{p})
\end{aligned}
where we noted that {d_(s)(( vec(p))),d_(s)^(†)(( vec(p)))}=2omega_(p)(2pi)^(3)delta^(3)( vec(0))=2omega_(p)V\left\{d_{s}(\vec{p}), d_{s}^{\dagger}(\vec{p})\right\}=2 \omega_{p}(2 \pi)^{3} \delta^{3}(\overrightarrow{0})=2 \omega_{p} V, where VV is the volume of space, and E_(0)=(1)/(2)int(d^(3)p)/((2pi)^(3))omega_(p)\mathcal{E}_{0}=\frac{1}{2} \int \frac{d^{3} p}{(2 \pi)^{3}} \omega_{p} is the vacuum energy. Hence, the Hamiltonion counts the number of b-type particles plus the number of d-type particles weighted by their energy.
The antisymmetry of the states under particle exchange is konwn as "FermiDirac statistics" and implies that rarr| vec(p),s, vec(p),s:)=0\rightarrow|\vec{p}, s, \vec{p}, s\rangle=0, which is known as "Pauli exclusion principle," i.e. two fermions cannot occupy the same state. In contrast, integer-spin fields satisfy bosonic (commuting) statistics, so multiple bosons can occupy the same state. As mentioned above, quantising spin- (1)/(2)\frac{1}{2} fields using commuting statistics would lead to inconsistencies. Hence, Dirac theory implies Pauli exclusion principle!
Electric charge:
Let us now right the electric charge in terms of creation and annihilation operators.
where we performed manipluations similar to the ones used to derive (12). In summary, the electric charge is equal to the number of b-type particles minus the number of d-type particles. We can think of b-type particles as electrons and d-type particles as positrons.
13 LSZ for spin half
This is based on Srednicki Chapter 41 and Peskin pg 118-119. Recall that
psi(x)=sum_(s)intd_(p)[b_(s)(( vec(p)))u_(s)(( vec(p)))e^(ip*x)+d_(s)^(†)(( vec(p)))v_(s)(( vec(p)))e^(-ip*x)]\psi(x)=\sum_{s} \int d_{p}\left[b_{s}(\vec{p}) u_{s}(\vec{p}) e^{i p \cdot x}+d_{s}^{\dagger}(\vec{p}) v_{s}(\vec{p}) e^{-i p \cdot x}\right]
where bb annihilates bb-type particles (ie. electrons) and d^(†)d^{\dagger} aretes dd-type particles (ie. positrons). Hence, psi\psi annihilates particles and creates antiparticles. Similarly, bar(psi)\bar{\psi} annihilates antiparticles and creates particles. We can use this fact to compute scattering amplitudes of Dirac particles. For example, the e^(-)e^(-)rarre^(-)e^(-)e^{-} e^{-} \rightarrow e^{-} e^{-}amplitude is given by
(:f∣i:)=(:0|Tb_(s_(4))( vec(p)_(4),+oo)b_(s_(3))( vec(p)_(3),+oo)b_(s_(2))^(†)( vec(p)_(2),-oo)b_(s_(1))^(†)( vec(p)_(1),-oo)|0:)\langle f \mid i\rangle=\langle 0| T b_{s_{4}}\left(\vec{p}_{4},+\infty\right) b_{s_{3}}\left(\vec{p}_{3},+\infty\right) b_{s_{2}}^{\dagger}\left(\vec{p}_{2},-\infty\right) b_{s_{1}}^{\dagger}\left(\vec{p}_{1},-\infty\right)|0\rangle
where TT denotes time ordering. Since the operators ate anti-commuting, any odd permutation gives a minus sign. We expect this to be related to the following time-ordered corrolator:
(:0|T psi(x_(4))psi(x_(3)) bar(psi)(x_(2)) bar(psi)(x_(1))|0:)\langle 0| T \psi\left(x_{4}\right) \psi\left(x_{3}\right) \bar{\psi}\left(x_{2}\right) \bar{\psi}\left(x_{1}\right)|0\rangle
where bar(Psi)\bar{\Psi} 's create the incoming electrons and Psi\Psi 's anmithate the two outgoing electrons. The precise relation to the scatting amplitude is given by the LSZ formula.
We previously showed that
b_(s)^(†)( vec(p))=intd^(3)xe^(ip*x) bar(psi)(x)gamma^(0)u_(s)( vec(p))b_{s}^{\dagger}(\vec{p})=\int d^{3} x e^{i p \cdot x} \bar{\psi}(x) \gamma^{0} u_{s}(\vec{p})
In an interacting theory, b_(s)^(†)( vec(p))b_{s}^{\dagger}(\vec{p}) can be time-dependent:
{:[b_(s)^(†)( vec(p)","-oo)-b_(s)^(†)( vec(p)","+oo)=-int_(-oo)^(oo)dtdel_(0)b^(†)( vec(p)","t)],[=-intd^(4)xdel_(0)(e^(ip*x)( bar(psi))(x)gamma^(0)u_(s)(( vec(p))))],[=-intd^(4)x bar(psi)(x)(gamma^(0)del ^(larr)ubrace(-igamma^(0)p^(0)ubrace))u_(s)( vec(p))e^(ip*x)],[-igamma^(i)p_(i)-im","" using "(p+m)u_(s)( vec(p))=0],[=-intd^(4)x bar(Psi)(x)(gamma^(0)del ^(larr)-gamma^(i) vec(del)_(i)-im)u_(s)( vec(p))e^(ip*x)],[=i intd^(4)x bar(psi)(x)(idel ^(larr)_(mu)gamma^(mu)+m)u_(s)( vec(p))e^(ip*x)]:}\begin{aligned}
& b_{s}^{\dagger}(\vec{p},-\infty)-b_{s}^{\dagger}(\vec{p},+\infty)=-\int_{-\infty}^{\infty} d t \partial_{0} b^{\dagger}(\vec{p}, t) \\
& =-\int d^{4} x \partial_{0}\left(e^{i p \cdot x} \bar{\psi}(x) \gamma^{0} u_{s}(\vec{p})\right) \\
& =-\int d^{4} x \bar{\psi}(x)(\gamma^{0} \overleftarrow{\partial} \underbrace{-i \gamma^{0} p^{0}}) u_{s}(\vec{p}) e^{i p \cdot x} \\
& -i \gamma^{i} p_{i}-i m, \text { using }(\not p+m) u_{s}(\vec{p})=0 \\
& =-\int d^{4} x \bar{\Psi}(x)\left(\gamma^{0} \overleftarrow{\partial}-\gamma^{i} \vec{\partial}_{i}-i m\right) u_{s}(\vec{p}) e^{i p \cdot x} \\
& =i \int d^{4} x \bar{\psi}(x)\left(i \overleftarrow{\partial}_{\mu} \gamma^{\mu}+m\right) u_{s}(\vec{p}) e^{i p \cdot x}
\end{aligned}
Note that in free theory bar(psi)(idel ^(larr)_(mu)gamma^(mu)+m)=0\bar{\psi}\left(i \overleftarrow{\partial}_{\mu} \gamma^{\mu}+m\right)=0 by the EOM, but in an interacting theory this does not vanish so b^(†)b^{\dagger} indeed changes over time. Hence,
b_(s)^(†)( vec(p),-oo)=b_(s)^(†)( vec(p),+oo)+i intd^(4)x bar(psi)(x)(idel ^(larr)_(mu)gamma^(mu)+m)u_(s)( vec(p))e^(ip*x)b_{s}^{\dagger}(\vec{p},-\infty)=b_{s}^{\dagger}(\vec{p},+\infty)+i \int d^{4} x \bar{\psi}(x)\left(i \overleftarrow{\partial}_{\mu} \gamma^{\mu}+m\right) u_{s}(\vec{p}) e^{i p \cdot x}
On the other hand, when we plug this into scattering amplitude, time-ordering moves b^(†)( vec(p),+oo)b^{\dagger}(\vec{p},+\infty) to the left and it annihilates (:0|\langle 0|. Hence, we can make the replacement
b_(s)^(†)( vec(p),-oo)rarr i intd^(4)x bar(psi)(x)(idel ^(larr)_(mu)gamma^(mu)+m)u_(s)( vec(p))e^(ip*x)b_{s}^{\dagger}(\vec{p},-\infty) \rightarrow i \int d^{4} x \bar{\psi}(x)\left(i \overleftarrow{\partial}_{\mu} \gamma^{\mu}+m\right) u_{s}(\vec{p}) e^{i p \cdot x}
Similarly we find
{:[b_(s)( vec(p)","+oo) rarr i intd^(4)xe^(-ip*x) bar(u)_(s)( vec(p))(-i del+m)psi(x)],[d_(s)^(†)( vec(p)","-oo) rarr-i intd^(4)xe^(ip*x) bar(v)_(s)( vec(p))(-i del+m)psi(x)],[d_(s)( vec(p)","+oo) rarr-i intd^(4)x bar(psi)(x)(idel ^(larr)_(mu)gamma^(mu)+m)v_(s)( vec(p))e^(-ip*x)]:}\begin{aligned}
b_{s}(\vec{p},+\infty) & \rightarrow i \int d^{4} x e^{-i p \cdot x} \bar{u}_{s}(\vec{p})(-i \not \partial+m) \psi(x) \\
d_{s}^{\dagger}(\vec{p},-\infty) & \rightarrow-i \int d^{4} x e^{i p \cdot x} \bar{v}_{s}(\vec{p})(-i \not \partial+m) \psi(x) \\
d_{s}(\vec{p},+\infty) & \rightarrow-i \int d^{4} x \bar{\psi}(x)\left(i \overleftarrow{\partial}{ }_{\mu} \gamma^{\mu}+m\right) v_{s}(\vec{p}) e^{-i p \cdot x}
\end{aligned}
Hence, to obtain a scattering amplitude, Fourier transform a time-ordered correlator to momentum space and act with the Dirac operator, amputating external propagators. Incoming particles/outgoing antipartides are created/annihilated by bar(psi)\bar{\psi}, while incoming antiparticles/outgoing particles are created/annihilated by psi\psi, as expected. Coming back to the example of e^(-)e^(-)rarre^{-} e^{-} \rightarrowe^(-)e^(-)e^{-} e^{-}scattering, LSZ gives
In terms of momentum space Feynman rules, we have
where arrows on lines indicate charge plow. Note that
In general we move fields around such that contractions are untangled and pick up factors of (-1)(-1) in the process. When doing so, keep the ordering of labels in (:dots|\langle\ldots| and |dots:)|\ldots\rangle. In terms of Feynman diagrams, this means that diagrams related by an odd permutation of external legs will have a relative (-1)(-1).
14 Fermionic Propagator
This is based on Srednicki Chapter 42. Recall that
First compute (:0|psi bar(psi)|0:)\langle 0| \psi \bar{\psi}|0\rangle. Note that the d^(†)d^{\dagger} in psi\psi and dd in bar(psi)\bar{\psi} can be dropped since they annihilate the vacuum on left and right, respectively. We are then left with
Similarly, in (:0| bar(psi)psi|0:)\langle 0| \bar{\psi} \psi|0\rangle car neglect b^(†)b^{\dagger} in bar(psi)\bar{\psi} and bb in psi\psi since they annihilate vacuum and we obtain:
{:[(:0|T psi(x) bar(psi)(y)|0:)=theta(x^(0)-y^(0))int widetilde(dp)e^(ip*(x-y))(-p+m)],[+theta(y^(0)-x^(0))int widetilde(dp)e^(-ip*(x-y))(p+m)]:}\begin{aligned}
\langle 0| T \psi(x) \bar{\psi}(y)|0\rangle= & \theta\left(x^{0}-y^{0}\right) \int \widetilde{d p} e^{i p \cdot(x-y)}(-\not p+m) \\
& +\theta\left(y^{0}-x^{0}\right) \int \widetilde{d p} e^{-i p \cdot(x-y)}(\not p+m)
\end{aligned}
Recall that
int(d^(4)p)/((2pi)^(4))(e^(ip*(x-y))f(p))/(p^(2)+m^(2)-i epsilon)=i theta(x^(0)-y^(0))int widetilde(dp)e^(ip*(x-y))f(p)+i theta(y^(0)-x^(0))int widetilde(dp)e^(-ip*(x-y))f(-p)\int \frac{d^{4} p}{(2 \pi)^{4}} \frac{e^{i p \cdot(x-y)} f(p)}{p^{2}+m^{2}-i \epsilon}=i \theta\left(x^{0}-y^{0}\right) \int \widetilde{d p} e^{i p \cdot(x-y)} f(p)+i \theta\left(y^{0}-x^{0}\right) \int \widetilde{d p} e^{-i p \cdot(x-y)} f(-p)
where f(p)f(p) is a polynomial in p^(mu)p^{\mu} and p^(0)=omega_(p)p^{0}=\omega_{p} on the right-hand-side. This can be derived by contour integration in complex p^(0)p^{0} plane. Puting everything together we find
(:0|T psi(x) bar(psi)(y)|0:)=(1)/(i)int(d^(4)p)/((2pi)^(4))e^(ip*(x-y))((-p+m))/(p^(2)+m^(2)-i epsilon)\langle 0| T \psi(x) \bar{\psi}(y)|0\rangle=\frac{1}{i} \int \frac{d^{4} p}{(2 \pi)^{4}} e^{i p \cdot(x-y)} \frac{(-\not p+m)}{p^{2}+m^{2}-i \epsilon}
we define this to be the Wick contraction of the fields:
(:0|T psi(x) bar(psi)(y)|0:)-=psi(x)(◻)/(psi)(y)\langle 0| T \psi(x) \bar{\psi}(y)|0\rangle \equiv \psi(x) \frac{\square}{\psi}(y)
Let S(x-y)=i(:0|T psi(x) bar(psi)(y)|0:)S(x-y)=i\langle 0| T \psi(x) \bar{\psi}(y)|0\rangle. This is the Green's function of the Dirac operator:
where we noted that -p^(2)=-p_(mu)p_(nu)gamma^(mu)gamma^(nu)=-p_(mu)p_(nu _)(1)/(2){gamma^(mu),gamma^(nu)}=p^(2)-\not p^{2}=-p_{\mu} p_{\nu} \gamma^{\mu} \gamma^{\nu}=-p_{\mu} p_{\underline{\nu}} \frac{1}{2}\left\{\gamma^{\mu}, \gamma^{\nu}\right\}=p^{2}. Moreover it is easy to see that (:0|T psi(x)psi(y)|0:)=(:0|T psi(x)psi(y)|0:)=0\langle 0| T \psi(x) \psi(y)|0\rangle=\langle 0| T \psi(x) \psi(y)|0\rangle=0 since there is no way pair up a bb with a b^(†)b^{\dagger} or a dd with a d^(†)d^{\dagger}. Hence the only nontrivial Wick contraction is between psi\psi and bar(psi)\bar{\psi}, similar to a complex scalar field.
We may now evaluate time-ordered correlators of Dirac fields using Wick contractions:
Hence, (:0|Tpsi_(alpha)(x_(1)) bar(psi)_(beta)(x_(2))psi_(gamma)(x_(3)) bar(psi)_(delta)(x_(4))|0:)=((1)/(i))^(2)[S(x_(1)-x_(2))_(alpha beta)S(x_(3)-x_(4))_(gamma delta)-S(x_(1)-x_(4))_(alpha delta)S(x_(3)-x_(2))_(gamma beta)]\langle 0| T \psi_{\alpha}\left(x_{1}\right) \bar{\psi}_{\beta}\left(x_{2}\right) \psi_{\gamma}\left(x_{3}\right) \bar{\psi}_{\delta}\left(x_{4}\right)|0\rangle=\left(\frac{1}{i}\right)^{2}\left[S\left(x_{1}-x_{2}\right)_{\alpha \beta} S\left(x_{3}-x_{4}\right)_{\gamma \delta}-S\left(x_{1}-x_{4}\right)_{\alpha \delta} S\left(x_{3}-x_{2}\right)_{\gamma \beta}\right]
In momentum space, we have the following Feynman rule for the Dirac propgator:
which arises from the Wick contraction of psi\psi with bar(psi)\bar{\psi}. The arrow on the line indices flow of charge. If momentum flows in the opposite direction of the charge
flow, take p rarr-pp \rightarrow-p in the Feynman rule. As we will see later, the direction of the arrows on internal lines are determined by the arrows on external lines (which come from LSZ) and charge conservation at the interaction vertices. One can then simply take the momentum on an internal line to point in the same direction as the charge flow and use the above Feynman rule.
15 Quantum Electrodynamics
This is based on Srednicki chapter 58 and Peskin pg 120. Recall Maxwell theory:
L=i bar(psi)p psi-m bar(psi)psi\mathcal{L}=i \bar{\psi} \not p \psi-m \bar{\psi} \psi
And recall the Noether current quadJ_(mu)=e bar(psi)gamma^(mu)psi\quad J_{\mu}=e \bar{\psi} \gamma^{\mu} \psi where ee is the charge of electron. A simple guess for how to couple electrons/positrons to photons is to identify Noether current with the source in Maxwell theory:
so psi\psi must transform somehow in order to cancel the gauge transformation of the last term. The following definition does the job:
psi(x)rarre^(-ie Gamma(x))psi(x)\psi(x) \rightarrow e^{-i e \Gamma(x)} \psi(x)
This can be obtained by promoting the U(1)U(1) global symmetry to a local one, where the U(1)U(1) transformation can vary from point to point in spacetime. This known as "gauging" the global symmetry. After gauging, the fermionic mass term remains invariant, but the derivative term does not.
Define the covariant derivative:
D_(mu)psi=del_(mu)psi-ieA_(mu)psiD_{\mu} \psi=\partial_{\mu} \psi-i e A_{\mu} \psi
Under a gauge transformation we have:
D_(mu)psi rarrdel_(mu)(e^(-ie Gamma)psi)-ie(A_(mu)-del_(mu)Gamma)(e^(-ie Gamma)psi)=e^(-ieP)D_(mu)psiD_{\mu} \psi \rightarrow \partial_{\mu}\left(e^{-i e \Gamma} \psi\right)-i e\left(A_{\mu}-\partial_{\mu} \Gamma\right)\left(e^{-i e \Gamma} \psi\right)=e^{-i e P} D_{\mu} \psi
Hence,
bar(psi)D psi rarr(( bar(psi))e^(ie Gamma))(e^(-ie Gamma)D psi)= bar(psi)D psi\bar{\psi} D \psi \rightarrow\left(\bar{\psi} e^{i e \Gamma}\right)\left(e^{-i e \Gamma} D \psi\right)=\bar{\psi} \not D \psi
so this is gauge-invariant. Moreover,
{:[[D_(mu),D_(nu)]psi=],[=(del_(mu)-ieA_(mu))(del_(nu)psi-ieA_(nu)psi)-mu harr nu],[{:-ieA_(mu)del_(nu)psi-e^(2)A_(mu)A_(nu)psi)-mu harr nu],[=-ieF_(mu nu)psi]:}\begin{aligned}
{\left[D_{\mu}, D_{\nu}\right] } & \psi= \\
= & \left(\partial_{\mu}-i e A_{\mu}\right)\left(\partial_{\nu} \psi-i e A_{\nu} \psi\right)-\mu \leftrightarrow \nu \\
& \left.-i e A_{\mu} \partial_{\nu} \psi-e^{2} A_{\mu} A_{\nu} \psi\right)-\mu \leftrightarrow \nu \\
= & -i e F_{\mu \nu} \psi
\end{aligned}
More abstractly, we can write F_(mu nu)=(i)/(e)[D_(mu),D_(nu)]F_{\mu \nu}=\frac{i}{e}\left[D_{\mu}, D_{\nu}\right] and the auge transformation of D_(mu)psiD_{\mu} \psi can be written as
D_(mu)psi rarr(e^(-ie Gamma)D_(mu)e^(ie Gamma))(e^(-ie Gamma)psi)D_{\mu} \psi \rightarrow\left(e^{-i e \Gamma} D_{\mu} e^{i e \Gamma}\right)\left(e^{-i e \Gamma} \psi\right)
so D_(mu)rarre^(-ie Gamma)D_(mu)e^(ie Gamma)D_{\mu} \rightarrow e^{-i e \Gamma} D_{\mu} e^{i e \Gamma}. This provides another perspective on gauge -invariance of F_(mu nu)F_{\mu \nu} since under a gauge transformation we have
{:[F_(mu nu)=(i)/(e)[D_(mu),D_(nu)] rarr(i)/(e)[e^(-ie Gamma)D_(mu)e^(ie Gamma),e^(-ie Gamma)D_(nu)e^(ie Gamma)]],[=(i)/(e)e^(-ie Gamma)[D_(mu)D_(nu)]e^(ie Gamma)],[=e^(-ie Gamma)F_(mu nu)e^(ie Gamma)],[=F_(mu nu)]:}\begin{aligned}
F_{\mu \nu}=\frac{i}{e}\left[D_{\mu}, D_{\nu}\right] & \rightarrow \frac{i}{e}\left[e^{-i e \Gamma} D_{\mu} e^{i e \Gamma}, e^{-i e \Gamma} D_{\nu} e^{i e \Gamma}\right] \\
& =\frac{i}{e} e^{-i e \Gamma}\left[D_{\mu} D_{\nu}\right] e^{i e \Gamma} \\
& =e^{-i e \Gamma} F_{\mu \nu} e^{i e \Gamma} \\
& =F_{\mu \nu}
\end{aligned}
Photons are represented by wavy lines and fermions are represented by solid lines with arrows. We already derived the momentum space Feynman vales for external legs (from LSZ) and internal propagators. Now we just need to derive the interaction vertex coming from
Stripping off the external polarisation and spinors, the Feynman rule for the interaction vertex is given by
Note that one arrow always points towards vertex and one always points away from vertex due to charge conservation.
Summary:
Fermions described by solid lines with arrows, photons by wavy lines.
For incoming e^(-)e^{-}, arrow and p^(mu)p^{\mu} point toward vertex. For outgoing e^(-)e^{-}, arrow and p^(mu)p^{\mu} point away from vertex. For incoming e^(+)e^{+}, arrow points away from vertex and p^(mu)p^{\mu} points toward vertex. For outgoing e^(+)e^{+}, arrow points toward and vertex and p^(mu)p^{\mu} points away from vertex.
Arrows on internal solid lines are determined by charge conservation at vertices. It is conventional to choose p^(mu)p^{\mu} to point along internal arrows. Then conserve p^(mu)p^{\mu} at each vertex.
For each internal photon, use
For each internal fermion, use
Spinor indices are contracted in opposite direction of arrows along fermion lines starting with a bar(u)\bar{u} or bar(v)\bar{v} (corresponding to an outgoing e^(-)e^{-}or an incoming e^(+)e^{+}) and ending with uu or vv (corresponding to an in coming e^(-)e^{-}or outgoing e^(+)e^{+}).
Dress external photons with polarisations epsilon_(mu)\epsilon_{\mu}
The overall sign of a diagram is determined by drawing all fermian lines pointing in the same direction with a fixed ordering of the incoming arrows. If the ordering of the outgoing arrows is an odd permutation of some given ordering, then that diagram gets a ( -1 ).
For example, at tree-level e^(-)e^(-)rarre^(-)e^(-)e^{-} e^{-} \rightarrow e^{-} e^{-}is described by
so the second diagram comes with a ( -1 ).
Fermion loops:
To compute Feynman diagrams with fermion loops, we need two more rules:
Take the trace of the product of gamma\gamma matrices from the vertices and fermionic propagator numerators that arise when going backwards along fermion lines.
Multiply by (-1)(-1) for each fermion loop.
Example:
where we leave out the gamma\gamma matrices coming from the interaction vertices and the minus sign comes from anticommuting the psi\psi on the right all the way to the left.
16 Scattering in QED
This is based on Tong sections 6.6, 3.5.2, and Peskin pg 121,125, 126. Let us consider a few examples of tree-level scattering amplitudes.
e^(-)e^(-)rarre^(-)e^(-)e^{-} e^{-} \rightarrow e^{-} e^{-}
where the minus sign in the second term arises from the exchange of legs 3 and 4 , and we use shorthand: u_(i)=u_(s_(1))( vec(p)_(i))u_{i}=u_{s_{1}}\left(\vec{p}_{i}\right).
We will now show how the Coulomb potential arises from charged partides exchanging photons. To see this, let's take the non-relativistic limit of the following diagram which contributes to e^(-)e^(-)rarre^(-)e^(-)e^{-} e^{-} \rightarrow e^{-} e^{-}:
In non-relativistic QM, the amplitude for a particle to scatter from | vec(p)_(1):)\left|\vec{p}_{1}\right\rangle to | vec(p)_(4):)\left|\vec{p}_{4}\right\rangle due to a potential V( vec(x))V(\vec{x}) (in the Born approximation) is the Fourier transform of the potential:
which is the Coulomb potential. For details on how to compute integral on LHS, see Tong pg 67. Alternatively, it easy to see that both sides satisfy the same differential equation:
This is based on Kabat section 7.2.2 and Appendix C. The Coulomb potential receives quantum corrections from diagrams of the form
This has observable effects, for example the "Lamb shift" of the energy levels of
Hydrogen. Heuristically the virtual electron and positron in the loop screen the electron. This is known as "vacuum polarisation." This screening effect causes the electron charge to decrease at long distances, or low energies. Mathematically, this can be derived from renormalisation.
Consider QED with a cutoff Lambda\Lambda on Euclidean momentum:
L=-(1)/(4)Z_(A)F_(mu nu)F^(mu nu)+i bar(psi)D psi-m bar(psi)psi,quadD_(mu)=del_(mu)-ieA_(mu)\mathcal{L}=-\frac{1}{4} Z_{A} F_{\mu \nu} F^{\mu \nu}+i \bar{\psi} D \psi-m \bar{\psi} \psi, \quad D_{\mu}=\partial_{\mu}-i e A_{\mu}
where Z_(A)Z_{A} encodes the normalisation of the gauge field A_(mu)A_{\mu}. As we will see shortly, the normalisation will run with the scale Lambda\Lambda. This is known as "wavefunction renormalisation". Now imagine lowering the cutoff to Lambda^(')=Lambda-delta Lambda\Lambda^{\prime}=\Lambda-\delta \Lambda and write down a new theory
where Euclidean momentum is restricted to | vec(k)_(E)| < Lambda^(')\left|\vec{k}_{E}\right|<\Lambda^{\prime}. The two theories will agree provided we relate Z_(A)Z_{A} and Z_(A)^(')Z_{A}^{\prime} appropriately. For delta Lambda//Lambda\delta \Lambda / \Lambda small, write
The extra term can be thought of as a new 2-point interaction vertex with the following Feynman rule:
Note that Fourier transforming F_(mu nu)F^(mu nu)F_{\mu \nu} F^{\mu \nu} to momentum space gives 2(eta_(mu nu)k^(2)-k_(mu)k_(nu))2\left(\eta_{\mu \nu} k^{2}-k_{\mu} k_{\nu}\right), as we've seen before. An addifiond factor of 2 comes from the 2 possible ways of Wick contracting the two external states onto the vertex.
On the other hand in the unprimed theory, the photon propagator receives corrections from vacuum polarisation diagram given at the beginning of the chapter where the Euclidean loop momentum is restricted to Lambda^(') < |ℓ_(E)| < Lambda\Lambda^{\prime}<\left|\ell_{E}\right|<\Lambda and we neglect higher-loop corrections. In the HW, you will show that after amputating external legs this diagram evaluates to
=-4ie^(2)(eta_(mu nu)k^(2)-k_(mu)k_(nu))int_(0)^(1)dx int(d^(4)q)/((2pi)^(4))(2x(1-x))/((q^(2)+k^(2)x(1-x)+m^(2))^(2))+dots=-4 i e^{2}\left(\eta_{\mu \nu} k^{2}-k_{\mu} k_{\nu}\right) \int_{0}^{1} d x \int \frac{d^{4} q}{(2 \pi)^{4}} \frac{2 x(1-x)}{\left(q^{2}+k^{2} x(1-x)+m^{2}\right)^{2}}+\ldots
where qq is a Euclidean loop momentum with Lambda^(') < | vec(q)| < Lambda\Lambda^{\prime}<|\vec{q}|<\Lambda, and ... denotes terms that break gauge invariance. Such terms are an artifact of working with a cutoff and don't arise if we use a regulator that preserves gauge invariance like dimensional regularisation, so we ignore them.
Hence, for the two theories to agree, we need
(dZ_(A))/(d Lambda)delta Lambda=-4e^(2)int_(0)^(1)dx int(d^(4)q)/((2pi)^(4))(2x(1-x))/((q^(2)-k^(2)x(1-x)+m^(2))^(2))\frac{d Z_{A}}{d \Lambda} \delta \Lambda=-4 e^{2} \int_{0}^{1} d x \int \frac{d^{4} q}{(2 \pi)^{4}} \frac{2 x(1-x)}{\left(q^{2}-k^{2} x(1-x)+m^{2}\right)^{2}}
Assuming that k^(2)≪Lambda^(2)k^{2} \ll \Lambda^{2} and m^(2)≪Lambda^(2)m^{2} \ll \Lambda^{2}, we can set k^(2)=m^(2)=0k^{2}=m^{2}=0 :
Hence normalisation of Maxwell kinetic term depends on cutoff! If we rescale A_(mu)rarr1//sqrt(Z_(A))A_(mu)A_{\mu} \rightarrow 1 / \sqrt{Z_{A}} A_{\mu} to get a canonical kinetic term for the Maxwell field, then the interaction term becomes
where mu\mu is the renormalisation scale. Hence e_("phys ")^(2)e_{\text {phys }}^{2} increases as Lambda\Lambda increases, and there is a Landau pole at Lambda_(max)=mu exp(6pi^(2)//e_("phys ")^(2)(mu))\Lambda_{\max }=\mu \exp \left(6 \pi^{2} / e_{\text {phys }}^{2}(\mu)\right), like in the phi^(4)\phi^{4} theory.