5 Fermions and Photons

The scalar particles we have studied so far had spin 0. What about higher spins? One can show a fundamental particle must be in the fundamental representation of the Lorentz group SO+(1,3)\mathrm{SO}^{+}(1,3) which limits possible spins to

  • j=0j=0: Higgs boson but also e.g. pion, Helium-4, Carbon-12;

  • j=1/2j=1/2: quarks and leptons but also e.g. proton, neutron;

  • j=1j=1: vector bosons like photons, gluons, ZZ and WW, but also deuteron, Nitrogen-14;

  • j=3/2j=3/2: called a Rarita-Schwinger particle, no fundamental example has been discovered but composite particles like Lithium-7 or the Δ++\Delta^{++} exist;

  • j=2j=2: any massless j=2j=2 particle can be shown to be a graviton of which we unfortunately do not have a consistent theory.

5.1 Fermions

Since fermions are particles of matter, let us consider them first. The main problem with the KG equation are its negative energy solutions. If viewed not in the context of a QFT but as a non-relativistic quantum theory, this would mean that no ground state could exist since it could always have less energy. These solutions appear because the KG is quadratic in t\partial_{t}. This in turn was a consequence of the on-shell relation p2=m2p^{2}=m^{2}. Dirac’s approach was then to write the KG operator as a product

(2+m2)=(iγνν+m)(iγμμm).\displaystyle-(\partial^{2}+m^{2})=(i\gamma^{\nu}\partial_{\nu}+m)(i\gamma^{% \mu}\partial_{\mu}-m)\,. (263)

If we therefore choose our differential equation to be

(iγμμm)ψ=0,\displaystyle(i\gamma^{\mu}\partial_{\mu}-m)\psi=0\,, (264)

it is by construction linear in t\partial_{t}, manifestly Lorentz invariant and fulfils the on-shell condition. Unfortunately, the γμ\gamma^{\mu} cannot be a mere number since we require

γμγνμν=ημνμν.\displaystyle\gamma^{\mu}\gamma^{\nu}\partial_{\mu}\partial_{\nu}=\eta^{\mu\nu% }\partial_{\mu}\partial_{\nu}\,. (265)

Note that this does not mean that γμγν=ημν\gamma^{\mu}\gamma^{\nu}=\eta^{\mu\nu} since the tensor μν\partial_{\mu}\partial_{\nu} is symmetric under μ\mu-ν\nu exchange. Instead, we can write with the anti-commutator {A,B}=AB+BA\{A,B\}=AB+BA

γμγνμν=12(γμγν+γνγμ)μν=12{γμ,γν}μν.\displaystyle\gamma^{\mu}\gamma^{\nu}\partial_{\mu}\partial_{\nu}=\frac{1}{2}% \Big{(}\gamma^{\mu}\gamma^{\nu}+\gamma^{\nu}\gamma^{\mu}\Big{)}\partial_{\mu}% \partial_{\nu}=\frac{1}{2}\{\gamma^{\mu},\gamma^{\nu}\}\partial_{\mu}\partial_% {\nu}\,. (266)

This is now the defining property of the γ\gamma matrices

{γμ,γν}=2ημν.\displaystyle\{\gamma^{\mu},\gamma^{\nu}\}=2\eta^{\mu\nu}\,. (267a)
To ensure that the Hamiltonian is self-adjoint, we also require the following normalisation
(γ0)2=1and(γk)2=1withk=1,2,3.\displaystyle\big{(}\gamma^{0}\big{)}^{2}=1\qquad\text{and}\qquad\big{(}\gamma% ^{k}\big{)}^{2}=-1\quad\text{with}\quad k=1,2,3\,. (267b)

These properties are actually sufficient for anything we may want to use γ\gamma matrices for, even without writing them down as explicit 4×44\times 4 objects.

Note that γμ\gamma^{\mu} is a Lorentz vector, i.e. a list of four 4×44\times 4 matrices. This can be made a bit clearer when using indices for this spinor space. For example with the identity matrix in spinor space II

(264)\displaystyle\eqref{eq:dirac} =b=14(iγabμmIab)ψb=0,\displaystyle=\sum_{b=1}^{4}(i\gamma^{\mu}_{ab}-m\,I_{ab})\psi_{b}=0\,, (268)
(271a)\displaystyle\eqref{eq:gammaident:2} =μ=03b=14γabμγμ,bc=4Iac,\displaystyle=\sum_{\mu=0}^{3}\sum_{b=1}^{4}\gamma^{\mu}_{ab}\gamma_{\mu,bc}=4% \,I_{ac}\,, (269)
(271b)\displaystyle\eqref{eq:gammaident:3} =μ=03b,c=14γabμγbcνγμ,cd=2γadν.\displaystyle=\sum_{\mu=0}^{3}\sum_{b,c=1}^{4}\gamma^{\mu}_{ab}\gamma^{\nu}_{% bc}\gamma_{\mu,cd}=-2\gamma^{\nu}_{ad}\,. (270)

We will usually not write spinor indices and reserve the \cdot operator for a Lorentz or three-vector product.

Suggested Exercise

Show the following identities

γμγμ\displaystyle\gamma^{\mu}\gamma_{\mu} =4,\displaystyle=4\,, (271a)
γμγνγμ\displaystyle\gamma^{\mu}\gamma^{\nu}\gamma_{\mu} =2γν,\displaystyle=-2\gamma^{\nu}\,, (271b)
γμγνγργμ\displaystyle\gamma^{\mu}\gamma^{\nu}\gamma^{\rho}\gamma_{\mu} =4ηνρ,\displaystyle=4\eta^{\nu\rho}\,, (271c)
γμγνγργσγμ\displaystyle\gamma^{\mu}\gamma^{\nu}\gamma^{\rho}\gamma^{\sigma}\gamma_{\mu} =2γσγργν.\displaystyle=-2\gamma^{\sigma}\gamma^{\rho}\gamma^{\nu}\,. (271d)

For example

γμγμ=ημνγμγν=12ημν(γμγν+γνγμ)=ημνημν=4,\displaystyle\gamma^{\mu}\gamma_{\mu}=\eta_{\mu\nu}\gamma^{\mu}\gamma^{\nu}% \stackrel{{\scriptstyle*}}{{=}}\frac{1}{2}\eta_{\mu\nu}(\gamma^{\mu}\gamma^{% \nu}+\gamma^{\nu}\gamma^{\mu})=\eta_{\mu\nu}\eta^{\mu\nu}=4\,, (272)

where we have used at * that the η\eta tensor is symmetric. For the next relation, we write using the anti-commutator and the previous result

γμγνγμ=(2ημνγνγμ)γμ=2γν4γν=2γν.\displaystyle\gamma^{\mu}\gamma^{\nu}\gamma_{\mu}=(2\eta^{\mu\nu}-\gamma^{\nu}% \gamma^{\mu})\gamma_{\mu}=2\gamma^{\nu}-4\gamma^{\nu}=-2\gamma^{\nu}\,. (273)

The others follow exactly the same way.

Suggested Exercise

Using the fact that tr(aA+bB)=atr(A)+btr(B){\rm tr}(aA+bB)=a{\rm tr}(A)+b{\rm tr}(B) and that tr(ABCD)=tr(DABC){\rm tr}(A\cdot B\cdots C\cdot D)={\rm tr}(D\cdot A\cdot B\cdots C), show that

tr(γμ)\displaystyle{\rm tr}(\gamma^{\mu}) =0,\displaystyle=0\,, (274a)
tr(γμ1γμkodd)\displaystyle{\rm tr}(\underbrace{\gamma^{\mu_{1}}\cdots\gamma^{\mu_{k}}}_{% \text{odd}}) =0,\displaystyle=0\,, (274b)
tr(γμγν)\displaystyle{\rm tr}(\gamma^{\mu}\gamma^{\nu}) =4ημν,\displaystyle=4\eta^{\mu\nu}\,, (274c)
tr(γμγνγργσ)\displaystyle{\rm tr}\Big{(}\gamma^{\mu}\gamma^{\nu}\gamma^{\rho}\gamma^{% \sigma}\Big{)} =4(ημνηρσημρηνσ+ημσηνρ).\displaystyle=4\Big{(}\eta^{\mu\nu}\eta^{\rho\sigma}-\eta^{\mu\rho}\eta^{\nu% \sigma}+\eta^{\mu\sigma}\eta^{\nu\rho}\Big{)}\,. (274d)

We begin with

tr(γμ)=12tr(γνγμγν)=12tr(γνγνγμ)=42tr(γμ)\displaystyle{\rm tr}(\gamma^{\mu})=\frac{1}{-2}{\rm tr}(\gamma^{\nu}\gamma^{% \mu}\gamma_{\nu})=\frac{1}{-2}{\rm tr}(\gamma_{\nu}\gamma^{\nu}\gamma^{\mu})=% \frac{4}{-2}{\rm tr}(\gamma^{\mu}) (275)

which can only be satisfied if the trace is zero. Similarly, we can show γ5\gamma^{5} (cf. next section)

tr(γμ1γμ2γμn)\displaystyle{\rm tr}(\gamma^{\mu_{1}}\gamma^{\mu_{2}}\cdots\gamma^{\mu_{n}}) =tr(γμ1γμ2γμnγ5γ5)=tr(γ5γμ1γμ2γμnγ5)=tr(γμ1γ5γμ2γμnγ5)\displaystyle={\rm tr}(\gamma^{\mu_{1}}\gamma^{\mu_{2}}\cdots\gamma^{\mu_{n}}% \gamma^{5}\gamma^{5})={\rm tr}(\gamma^{5}\gamma^{\mu_{1}}\gamma^{\mu_{2}}% \cdots\gamma^{\mu_{n}}\gamma^{5})=-{\rm tr}(\gamma^{\mu_{1}}\gamma^{5}\gamma^{% \mu_{2}}\cdots\gamma^{\mu_{n}}\gamma^{5})
=+tr(γμ1γμ2γ5γμnγ5)=(1)ntr(γμ1γμ2γμnγ5γ5).\displaystyle=+{\rm tr}(\gamma^{\mu_{1}}\gamma^{\mu_{2}}\gamma^{5}\cdots\gamma% ^{\mu_{n}}\gamma^{5})=(-1)^{n}{\rm tr}(\gamma^{\mu_{1}}\gamma^{\mu_{2}}\cdots% \gamma^{\mu_{n}}\gamma^{5}\gamma^{5})\,. (276)

If nn is odd, this means tr()=tr(){\rm tr}(\cdots)=-{\rm tr}(\cdots) which is only satisfied if the trace vanishes. Next,

tr(γμγν)=12(tr(γμγν)+tr(γνγμ))=12tr({γμ,γν})=ημνtr(1),\displaystyle{\rm tr}(\gamma^{\mu}\gamma^{\nu})=\frac{1}{2}\Big{(}{\rm tr}(% \gamma^{\mu}\gamma^{\nu})+{\rm tr}(\gamma^{\nu}\gamma^{\mu})\Big{)}=\frac{1}{2% }{\rm tr}\Big{(}\{\gamma^{\mu},\gamma^{\nu}\}\Big{)}=\eta^{\mu\nu}{\rm tr}(1)\,, (277)

with tr(1)=4{\rm tr}(1)=4.

Suggested Exercise

Finally, show that

(γμ)=γ0γμγ0.\displaystyle(\gamma^{\mu})^{\dagger}=\gamma^{0}\gamma^{\mu}\gamma^{0}\,. (278)

We can write the Dirac equation (264) using a Hamiltonian

iψt=(iγ0γ+γ0m)Hψ.\displaystyle i\frac{\partial\psi}{\partial t}=\underbrace{\Big{(}-i\gamma^{0}% \vec{\gamma}\cdot\vec{\nabla}+\gamma^{0}m\Big{)}}_{H}\psi\,. (279)

Since we want H=HH=H^{\dagger}, we need (γ0γ)(\gamma^{0}\vec{\gamma}) and (γ0)(\gamma^{0}) the be self-adjoint as well, justifying (267b).

5.1.1 Pauli’s fundamental theorem and basis of γ\gamma matrices

The γ\gamma matrices are fully defined through (267), i.e. we may use any set of matrices that fulfil these requirements. This means that if we have a different set of matrices (γ)μ(\gamma^{\prime})^{\mu} that also fulfil (267), they must be related to γμ\gamma^{\mu} through a constant invertible matrix SS

(γ)μ=S1γμS.\displaystyle(\gamma^{\prime})^{\mu}=S^{-1}\gamma^{\mu}S\,. (280)

This is called Pauli’s fundamental theorem. Its proof is not that important but it makes use of an important fact: we can write any product of γ\gamma matrices using a basis of 16=4×416=4\times 4 elements.

To pick these, it is customary to define a fifth γ\gamma matrix

γ5=iγ0γ1γ2γ3=i4!εμνρσγμγνγργσ,\displaystyle\gamma^{5}=i\gamma^{0}\gamma^{1}\gamma^{2}\gamma^{3}=-\frac{i}{4!% }\varepsilon^{\mu\nu\rho\sigma}\gamma_{\mu}\gamma_{\nu}\gamma_{\rho}\gamma_{% \sigma}\,, (281)

with the totally anti-symmetric tensor ϵμνρσ\epsilon^{\mu\nu\rho\sigma}, defined to be ε0123=1\varepsilon^{0123}=1.

Suggested Exercise

Show using the anti-commutation relations and the definition of γ5\gamma^{5}

(γ5)\displaystyle(\gamma^{5})^{\dagger} =γ5,\displaystyle=\gamma^{5}\,, (282a)
(γ5)2\displaystyle(\gamma^{5})^{2} =1,\displaystyle=1\,, (282b)
{γ5,γμ}\displaystyle\{\gamma^{5},\gamma^{\mu}\} =0.\displaystyle=0\,. (282c)

For example, for the anti-commutator

{γ5,γλ}=i4!εμνρσ{γμγνγργσ,γλ},\displaystyle\{\gamma^{5},\gamma^{\lambda}\}=-\frac{i}{4!}\varepsilon_{\mu\nu% \rho\sigma}\{\gamma^{\mu}\gamma^{\nu}\gamma^{\rho}\gamma^{\sigma},\gamma^{% \lambda}\}\,, (283)

we write

γμγνγργσγλ\displaystyle\gamma^{\mu}\gamma^{\nu}\gamma^{\rho}\gamma^{\sigma}\gamma^{\lambda} =γμγνγρ2ηλσγμγνγσ2ηλρ+γμγργσ2ηνλγνγργσ2ημλ+γλγμγνγργσ.\displaystyle=\gamma^{\mu}\gamma^{\nu}\gamma^{\rho}2\eta^{\lambda\sigma}-% \gamma^{\mu}\gamma^{\nu}\gamma^{\sigma}2\eta^{\lambda\rho}+\gamma^{\mu}\gamma^% {\rho}\gamma^{\sigma}2\eta^{\nu\lambda}-\gamma^{\nu}\gamma^{\rho}\gamma^{% \sigma}2\eta^{\mu\lambda}+\gamma^{\lambda}\gamma^{\mu}\gamma^{\nu}\gamma^{\rho% }\gamma^{\sigma}\,. (284)

Therefore

{γ5,γλ}=2i4!εμνρσ(γμγνγρηλσγμγνγσηλρ+γμγργσηνλγνγργσημλ+γλγμγνγργσ)\displaystyle\{\gamma^{5},\gamma^{\lambda}\}=-\frac{2i}{4!}\varepsilon_{\mu\nu% \rho\sigma}\Big{(}\gamma^{\mu}\gamma^{\nu}\gamma^{\rho}\eta^{\lambda\sigma}-% \gamma^{\mu}\gamma^{\nu}\gamma^{\sigma}\eta^{\lambda\rho}+\gamma^{\mu}\gamma^{% \rho}\gamma^{\sigma}\eta^{\nu\lambda}-\gamma^{\nu}\gamma^{\rho}\gamma^{\sigma}% \eta^{\mu\lambda}+\gamma^{\lambda}\gamma^{\mu}\gamma^{\nu}\gamma^{\rho}\gamma^% {\sigma}\Big{)} (285)

We can write any product of γ\gamma matrices using the following basis

Γ={11,γμ4,i2[γμ,γν]6,γ51,γμγ54}.\displaystyle\Gamma=\Big{\{}\underbrace{1}_{1},\underbrace{\gamma^{\mu}}_{4},% \underbrace{\tfrac{i}{2}[\gamma^{\mu},\gamma^{\nu}]}_{6},\underbrace{\gamma^{5% }}_{1},\underbrace{\gamma^{\mu}\gamma^{5}}_{4}\Big{\}}\,. (286)

The numbers indicate that the number of basis elements of this form. We refer to these as scalar, vector, tensor, pseudo-scalar and pseudo-vector respectively.

If we simultaneously transform the spinor ψ\psi as ψS1ψ\psi\to S^{-1}\psi, the Dirac equation transforms to

(iγm)ψ(iγm)ψ=(i(S1γS)m)S1ψ=(iγm)\displaystyle(i\gamma\cdot\partial-m)\psi\to(i\gamma^{\prime}\cdot\partial^{% \prime}-m)\psi^{\prime}=(i(S^{-1}\gamma S)\cdot\partial-m)S^{-1}\psi=(i\gamma% \cdot\partial-m) (287)

This means physics will always be invariant under basis change.

Even if the exact form of the γ\gamma matrices does not matter, it is sometimes helpful to have one

γ0=(12×202×202×212×2),γi=(02×2σiσi02×2),\displaystyle\gamma^{0}=\begin{pmatrix}\phantom{-}1_{2\times 2}&\phantom{-}0_{% 2\times 2}\\ \phantom{-}0_{2\times 2}&-1_{2\times 2}\end{pmatrix}\,,\qquad\gamma^{i}=\begin% {pmatrix}0_{2\times 2}&\sigma^{i}\\ -\sigma^{i}&0_{2\times 2}\end{pmatrix}\,, (288)

with the Pauli matrices σi\sigma^{i}

σ1=(0110),σ2=(0ii0),σ3=(1001).\displaystyle\sigma^{1}=\begin{pmatrix}0&1\\ 1&0\end{pmatrix}\,,\quad\sigma^{2}=\begin{pmatrix}0&-i\\ i&0\end{pmatrix}\,,\quad\sigma^{3}=\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}\,. (289)

5.1.2 Transformations of the Dirac equation

If we want to view the spinor ψ\psi as physically meaningful, we need to understand how it transforms in different frames. Consider therefore two frames described by xx and xx^{\prime} with (x)μ=Λνμxν(x^{\prime})^{\mu}=\Lambda^{\mu}_{\phantom{\mu}\nu}x^{\nu}. If ψ(x)\psi(x) is a solution in the xx frame and ψ(x)\psi^{\prime}(x^{\prime}) is a solution in the xx^{\prime} frame, we must have

(iγμxμm)ψ(x)=(iγμ(x)μm)ψ(x)=0.\displaystyle\bigg{(}i\gamma^{\mu}\frac{\partial}{\partial x^{\mu}}-m\bigg{)}% \psi(x)=\bigg{(}i\gamma^{\mu}\frac{\partial}{\partial(x^{\prime})^{\mu}}-m% \bigg{)}\psi^{\prime}(x^{\prime})=0\,. (290)

We can use (27) to transform the derivative

0=(iγμxμm)ψ(x)=(iγμΛμν(γ)ν(x)νm)ψ(Λ1x).\displaystyle 0=\bigg{(}i\gamma^{\mu}\frac{\partial}{\partial x^{\mu}}-m\bigg{% )}\psi(x)=\bigg{(}i\underbrace{\gamma^{\mu}\Lambda^{\nu}_{\phantom{\nu}\mu}}_{% (\gamma^{\prime})^{\nu}}\frac{\partial}{\partial(x^{\prime})^{\nu}}-m\bigg{)}% \psi(\Lambda^{-1}x^{\prime})\,. (291)

The new γ\gamma matrices (γ)μ=Λνμγν(\gamma^{\prime})^{\mu}=\Lambda^{\mu}_{\phantom{\mu}\nu}\gamma^{\nu} still need to fulfil (267a)

{(γ)α,(γ)β}={Λμαγμ,Λνβγν}=ΛμαΛνβ{γμ,γν}=ΛμαΛνβ2ημν=2ηαβ,\displaystyle\{(\gamma^{\prime})^{\alpha},(\gamma^{\prime})^{\beta}\}=\{% \Lambda^{\alpha}_{\phantom{\alpha}\mu}\gamma^{\mu},\Lambda^{\beta}_{\phantom{% \beta}\nu}\gamma^{\nu}\}=\Lambda^{\alpha}_{\phantom{\alpha}\mu}\Lambda^{\beta}% _{\phantom{\beta}\nu}\{\gamma^{\mu},\gamma^{\nu}\}=\Lambda^{\alpha}_{\phantom{% \alpha}\mu}\Lambda^{\beta}_{\phantom{\beta}\nu}2\eta^{\mu\nu}=2\eta^{\alpha% \beta}\,, (292)

since (23) guarantees ημνΛρμΛσν=ησρ\eta_{\mu\nu}\Lambda^{\mu}_{\phantom{\mu}\rho}\Lambda^{\nu}_{\phantom{\nu}% \sigma}=\eta_{\sigma\rho}. Pauli’s fundamental theorem proves the existence of a spinor matrix S(Λ)S(\Lambda) such that

(γ)μ=Λνμγν=S(Λ)1γμS(Λ).\displaystyle(\gamma^{\prime})^{\mu}=\Lambda^{\mu}_{\phantom{\mu}\nu}\gamma^{% \nu}=S(\Lambda)^{-1}\gamma^{\mu}S(\Lambda)\,. (293)

Therefore,

0=(iγμxμm)ψ(x)=S(Λ)1(iγμ(x)νm)S(Λ)ψ(Λ1x).\displaystyle 0=\bigg{(}i\gamma^{\mu}\frac{\partial}{\partial x^{\mu}}-m\bigg{% )}\psi(x)=S(\Lambda)^{-1}\bigg{(}i\gamma^{\mu}\frac{\partial}{\partial(x^{% \prime})^{\nu}}-m\bigg{)}S(\Lambda)\psi(\Lambda^{-1}x^{\prime})\,. (294)

If we left-multiply with S(Λ)S(\Lambda) and identify ψ(x)=S(Λ)ψ(Λ1x)\psi^{\prime}(x)^{\prime}=S(\Lambda)\psi(\Lambda^{-1}x^{\prime}), we arrive at

0=(iγμ(x)νm)ψ(x),\displaystyle 0=\bigg{(}i\gamma^{\mu}\frac{\partial}{\partial(x^{\prime})^{\nu% }}-m\bigg{)}\psi^{\prime}(x^{\prime})\,, (295)

the transformed Dirac equation (290)

To study S(Λ)S(\Lambda) consider an infinitesimal transformation Λ\Lambda which should also be infinitesimal in SS, i.e. we have

Λνμ\displaystyle\Lambda^{\mu}_{\phantom{\mu}\nu} =δνμ+ωνμ,\displaystyle=\delta^{\mu}_{\phantom{\mu}\nu}+\omega^{\mu}_{\phantom{\mu}\nu}\,, (296)
S(Λ)\displaystyle S(\Lambda) =1+iωμνΣμν.\displaystyle=1+i\omega^{\mu\nu}\Sigma_{\mu\nu}\,. (297)

Since ωμν+ωνμ=0\omega_{\mu\nu}+\omega_{\nu\mu}=0, we have the same anti-symmetry for Σ\Sigma. Let us now calculate what happens to γ\gamma

Λνμγν\displaystyle\Lambda^{\mu}_{\phantom{\mu}\nu}\gamma^{\nu} =!S(Λ)1γμS(Λ),\displaystyle\stackrel{{\scriptstyle!}}{{=}}S(\Lambda)^{-1}\gamma^{\mu}S(% \Lambda)\,, (298)
γμ+ωνμγν+𝒪(ω2)\displaystyle\Leftrightarrow\qquad\qquad\gamma^{\mu}+\omega^{\mu}_{\phantom{% \mu}\nu}\gamma^{\nu}+\mathcal{O}(\omega^{2}) =(1iωλρΣλρ)γμ(1+iωλρΣλρ)+𝒪(ω2),\displaystyle=\Big{(}1-i\omega^{\lambda\rho}\Sigma_{\lambda\rho}\Big{)}\gamma^% {\mu}\Big{(}1+i\omega^{\lambda\rho}\Sigma_{\lambda\rho}\Big{)}+\mathcal{O}(% \omega^{2})\,, (299)
12ωλρ(δλμγρδρμγλ)=ωνμγν\displaystyle\Leftrightarrow\qquad\qquad\frac{1}{2}\omega^{\lambda\rho}\Big{(}% \delta^{\mu}_{\phantom{\mu}\lambda}\gamma_{\rho}-\delta^{\mu}_{\phantom{\mu}% \rho}\gamma_{\lambda}\Big{)}=\omega^{\mu}_{\phantom{\mu}\nu}\gamma^{\nu} =iωλρ(γμΣλρΣλργμ)=iωλρ[γμ,Σλρ]\displaystyle=i\omega^{\lambda\rho}\Big{(}\gamma^{\mu}\Sigma_{\lambda\rho}-% \Sigma_{\lambda\rho}\gamma^{\mu}\Big{)}=i\omega^{\lambda\rho}[\gamma^{\mu},% \Sigma_{\lambda\rho}] (300)
12(δλμγρδρμγλ)\displaystyle\Leftrightarrow\qquad\qquad\frac{1}{2}\Big{(}\delta^{\mu}_{% \phantom{\mu}\lambda}\gamma_{\rho}-\delta^{\mu}_{\phantom{\mu}\rho}\gamma_{% \lambda}\Big{)} =i[γμ,Σλρ].\displaystyle=i[\gamma^{\mu},\Sigma_{\lambda\rho}]\,. (301)

This is satisfied by

Σλρ=i8[γλ,γρ]\displaystyle\Sigma_{\lambda\rho}=-\frac{i}{8}[\gamma_{\lambda},\gamma_{\rho}] (302)

which can be integrated to

S(Λ)=exp(i8ωμν[γμ,γν])forΛSO+(1,3).\displaystyle S(\Lambda)=\exp\Big{(}-\frac{i}{8}\omega^{\mu\nu}[\gamma_{\mu},% \gamma_{\nu}]\Big{)}\qquad\text{for}\quad\Lambda\in\mathrm{SO}^{+}(1,3)\,. (303)
Suggested Exercise

Show that

[γμ,γλγρ]=2(δλμγρδρμγλ),\displaystyle[\gamma^{\mu},\gamma_{\lambda}\gamma_{\rho}]=2(\delta^{\mu}_{% \phantom{\mu}\lambda}\gamma_{\rho}-\delta^{\mu}_{\phantom{\mu}\rho}\gamma_{% \lambda})\,, (304)

by using the anticommutator. Use this to show that (302) is a solution of (301).

The above discussion is only valid for proper transformations that have detΛ=+1\det\Lambda=+1. To also cover improper transformations, consider

P=(1000010000100001).\displaystyle P=\begin{pmatrix}1&0&0&0\\ 0&-1&0&0\\ 0&0&-1&0\\ 0&0&0&-1\end{pmatrix}\,. (305)

This transformation flips the spatial components, i.e. it is a parity transformation. We need to fulfil

S(P)1γ0S(P)=γ0andS(P)1γiS(P)=γi.\displaystyle S(P)^{-1}\gamma^{0}S(P)=\gamma^{0}\qquad\text{and}\qquad S(P)^{-% 1}\gamma^{i}S(P)=-\gamma^{i}\,. (306)

There are two possible choices for S(P)S(P)

S(P)=S(P)1=±γ0\displaystyle S(P)=S(P)^{-1}=\pm\gamma^{0} (307)

correctly transforms γμ\gamma^{\mu}. When acting on the wavefunction, we can have two solutions as well

ψ(t,x)ψ(t,x)=±γ0ψ(t,x).\displaystyle\psi(t,\vec{x})\to\psi^{\prime}(t,\vec{x})=\pm\gamma^{0}\psi(t,-% \vec{x})\,. (308)

The sign is called intrinsic parity and only starts to matter once we consider system with changing numbers of particles.

5.1.3 Solutions of the Dirac equation

To eventually construct a field theory, we will need a basis of solutions to the wave equation (264). As always, we begin with a Fourier-transformation of ψ(x)\psi(x)

ψ(x)=d3p(2π)32Ep(u(p)eipx+v(p)eipx).\displaystyle\psi(x)=\int\frac{{\rm d}^{3}\vec{p}}{(2\pi)^{3}\sqrt{2E_{\vec{p}% }}}\Big{(}u(p){\rm e}^{-ip\cdot x}+v(p){\rm e}^{ip\cdot x}\Big{)}\,. (309)

Here the uu and vv objects are vectors in spinor space. They fulfil the momentum-space Dirac equation

(γpm)u(p)=(γp+m)v(p)=0.\displaystyle(\gamma\cdot p-m)u(p)=(\gamma\cdot p+m)v(p)=0\,. (310)

In the restframe of the particle where p=(E,0,0,0)p=(E,0,0,0), we have

(γ01)u(0)=(γ0+1)v(0)=0.\displaystyle(\gamma^{0}-1)u(0)=(\gamma^{0}+1)v(0)=0\,. (311)

We can find explicit answers for the spinors using the explicit representation of (288)

u1(0)=(1000),u2(0)=(0100),v1(0)=(0010),v2(0)=(0001).\displaystyle u_{1}(0)=\begin{pmatrix}1\\ 0\\ 0\\ 0\end{pmatrix}\,,\quad u_{2}(0)=\begin{pmatrix}0\\ 1\\ 0\\ 0\end{pmatrix}\,,\quad v_{1}(0)=\begin{pmatrix}0\\ 0\\ 1\\ 0\end{pmatrix}\,,\quad v_{2}(0)=\begin{pmatrix}0\\ 0\\ 0\\ 1\end{pmatrix}\,. (312)

Note that we have not one but two solution for each direction of pp. One can show that these correspond to the two spin directions.

To turn these into solutions for u(p)u(p), we could perform a Lorentz boost. Alternatively, we can note that

(γpm)(γp+m)=(p2m2),\displaystyle(\gamma\cdot p-m)(\gamma\cdot p+m)=(p^{2}-m^{2})\,, (313)

to write

ur(p)(γp+m)ur(0),vr(p)(γp+m)vr(0).\displaystyle u_{r}(p)\propto(\gamma\cdot p+m)u_{r}(0)\,,\qquad v_{r}(p)% \propto(-\gamma\cdot p+m)v_{r}(0)\,. (314)
Suggested Exercise

(313) relies on the fact that (γa)(γa)=a2(\gamma\cdot a)(\gamma\cdot a)=a^{2}. Proof this.

As we will see, a suitable normalisation is

ur(p)γ0us(p)=δrs,vr(p)γ0vs(p)=δrs,ur(p)γ0vs(p)=vs(p)γ0ur(p)=0.\displaystyle u_{r}(p)^{\dagger}\gamma^{0}u_{s}(p)=\delta_{rs}\,,\qquad v_{r}(% p)^{\dagger}\gamma^{0}v_{s}(p)=-\delta_{rs}\,,\qquad u_{r}(p)^{\dagger}\gamma^% {0}v_{s}(p)=v_{s}(p)^{\dagger}\gamma^{0}u_{r}(p)=0\,. (315)

This leads to

ur(p)=γp+m2m(m+Ep)ur(0),vr(p)=γp+m2m(m+Ep)vr(0).\displaystyle u_{r}(p)=\frac{\gamma\cdot p+m}{\sqrt{2m(m+E_{\vec{p}})}}u_{r}(0% )\,,\qquad v_{r}(p)=\frac{-\gamma\cdot p+m}{\sqrt{2m(m+E_{\vec{p}})}}v_{r}(0)\,. (316)

Further, we can show that (completeness relation)

r=1,2ur(p)ur(p)γ0=γp+mandr=1,2vr(p)vr(p)γ0=γpm.\displaystyle\sum_{r=1,2}u_{r}(p)u_{r}(p)^{\dagger}\gamma^{0}=\gamma\cdot p+m% \qquad\text{and}\qquad\sum_{r=1,2}v_{r}(p)v_{r}(p)^{\dagger}\gamma^{0}=\gamma% \cdot p-m\,. (317)

5.1.4 Quantisation of the free Dirac field

To define the Lagrangian of the free Dirac field, it is helpful to first define the adjoint spinor

ψ¯=ψγ0.\displaystyle\bar{\psi}=\psi^{\dagger}\gamma^{0}\,. (318)

With this, the Lagrangian can be written as

=ψ¯(iγμμm)ψ,\displaystyle\mathcal{L}=\bar{\psi}(i\gamma^{\mu}\partial_{\mu}-m)\psi\,, (319)

since it results in the correct Euler-Lagrange equation for ψ¯\bar{\psi}. To see this, we calculate

(μψ)=iψ¯γμandψ=ψ¯m,\displaystyle\frac{\partial\mathcal{L}}{\partial(\partial_{\mu}\psi)}=i\bar{% \psi}\gamma^{\mu}\quad\text{and}\quad\frac{\partial\mathcal{L}}{\partial\psi}=% -\bar{\psi}m\,, (320)

and write

0=iψ¯μγμmψ¯.\displaystyle 0=-i\bar{\psi}\overleftarrow{\partial_{\mu}}\gamma^{\mu}-m\bar{% \psi}\,. (321)

Here we have used the notation μ\overleftarrow{\partial_{\mu}} to indicate that the derivative is acting to the left rather than the right. We can Hermitian-conjugate this to arrive at (264).

The assosciated Hamiltonian is

H=d3x=d3xψ¯(iγ+m)ψ.\displaystyle H=\int{\rm d}^{3}x\ \mathcal{H}=\int{\rm d}^{3}x\ \bar{\psi}(-i% \vec{\gamma}\cdot\vec{\nabla}+m)\psi\,. (322)

Since uu and vv are vectors, we will extend our Fourier-decomposition of the fermion field slightly to split the creation and annihilation operators from the spinor vectors, i.e. we write

ψ(x)=d3p(2π)32Epr=1,2(ar(p)ur(p)eipx+br(p)vr(p)e+ipx).\displaystyle\psi(x)=\int\frac{{\rm d}^{3}\vec{p}}{(2\pi)^{3}\sqrt{2E_{\vec{p}% }}}\sum_{r=1,2}\Big{(}a_{r}(\vec{p})u_{r}(p){\rm e}^{-ip\cdot x}+b_{r}(\vec{p}% )v_{r}(p){\rm e}^{+ip\cdot x}\Big{)}\,. (323)

We have further flipped the propagation direction of the vv spinors to ensure positive energy. Recall how we used CCRs to quantise the KG field

[a^(p),a^(q)]\displaystyle[\hat{a}(\vec{p}),\hat{a}^{\dagger}(\vec{q})] =(2π)3δ(3)(pq),\displaystyle=(2\pi)^{3}\delta^{(3)}(\vec{p}-\vec{q})\,, (83a)
[a^(p),a^(q)]\displaystyle[\hat{a}(\vec{p}),\hat{a}(\vec{q})] =[a^(p),a^(q)]=0.\displaystyle=[\hat{a}^{\dagger}(\vec{p}),\hat{a}^{\dagger}(\vec{q})]=0\,. (83b)

Here we would have

[ar(p),as(q)]=[br(p),bs(q)]=(2π)3δ(3)(pq)δrsand all other commutators zero.\displaystyle[a_{r}(\vec{p}),a_{s}^{\dagger}(\vec{q})]=[b_{r}(\vec{p}),b_{s}^{% \dagger}(\vec{q})]=(2\pi)^{3}\delta^{(3)}(\vec{p}-\vec{q})\delta_{rs}\qquad% \text{and all other commutators zero}\,. (324)

However, this will lead to a contradiction. We can show that the Hamiltonian of this theory is

H=d3p(2π)3Epr(ar(p)ar(p)br(p)br(p)).\displaystyle H=\int\frac{{\rm d}^{3}p}{(2\pi)^{3}}E_{\vec{p}}\sum_{r}\Big{(}a% _{r}^{\dagger}(\vec{p})a_{r}(\vec{p})-b_{r}^{\dagger}(\vec{p})b_{r}(\vec{p})% \Big{)}\,. (325)

Since we can view arara_{r}^{\dagger}a_{r} and brbrb_{r}^{\dagger}b_{r} as particle numbers for particles of type aa and bb, this would mean that creating more bb-type particles decreases the energy of the system.

If we instead swap bb and bb^{\dagger}

ψ(x)\displaystyle\psi(x) =d3p(2π)32Epr=1,2(ar(p)ur(p)eipx+br(p)vr(p)e+ipx),\displaystyle=\int\frac{{\rm d}^{3}\vec{p}}{(2\pi)^{3}\sqrt{2E_{\vec{p}}}}\sum% _{r=1,2}\Big{(}a_{r}(\vec{p})u_{r}(p){\rm e}^{-ip\cdot x}+b^{\dagger}_{r}(\vec% {p})v_{r}(p){\rm e}^{+ip\cdot x}\Big{)}\,, (326)
ψ¯(x)\displaystyle\bar{\psi}(x) =d3p(2π)32Epr=1,2(ar(p)u¯r(p)e+ipx+br(p)v¯r(p)eipx),\displaystyle=\int\frac{{\rm d}^{3}\vec{p}}{(2\pi)^{3}\sqrt{2E_{\vec{p}}}}\sum% _{r=1,2}\Big{(}a_{r}^{\dagger}(\vec{p})\bar{u}_{r}(p){\rm e}^{+ip\cdot x}+b_{r% }(\vec{p})\bar{v}_{r}(p){\rm e}^{-ip\cdot x}\Big{)}\,, (327)

and chose anitcommutation relations

{ar(p),as(q)}={br(p),bs(q)}=(2π)3δ(3)(pq)δrsand all other anticommutators zero,\displaystyle\{a_{r}(\vec{p}),a_{s}^{\dagger}(\vec{q})\}=\{b_{r}(\vec{p}),b_{s% }^{\dagger}(\vec{q})\}=(2\pi)^{3}\delta^{(3)}(\vec{p}-\vec{q})\delta_{rs}% \qquad\text{and all other anticommutators zero}\,, (328)

we would find

H=d3p(2π)3Epr(ar(p)ar(q)+br(p)br(p)).\displaystyle H=\int\frac{{\rm d}^{3}p}{(2\pi)^{3}}E_{\vec{p}}\sum_{r}\Big{(}a% _{r}^{\dagger}(\vec{p})a_{r}(\vec{q})+b_{r}^{\dagger}(\vec{p})b_{r}(\vec{p})% \Big{)}\,. (329)

which means that both aa^{\dagger} and bb^{\dagger} create a particle of mass m2=p2m^{2}=p^{2}.

Suggested Exercise

Show (325) and (329).

Suggested Exercise

You may find the following quantum mechanics problem instructive. We normally consider the bosonic harmonic oscillator defined as

HB=ω2(aa+aa)with[a,a]=1,[a,a]=[a,a]=0.\displaystyle H_{B}=\frac{\omega}{2}(a^{\dagger}a+aa^{\dagger})\qquad\text{% with}\qquad[a,a^{\dagger}]=1\,,\quad[a,a]=[a^{\dagger},a^{\dagger}]=0\,. (330)

Now define the fermionic oscillator with

HF=ω2(bbbb)with{b,b}=1,{b,b}={b,b}=0.\displaystyle H_{F}=\frac{\omega}{2}(b^{\dagger}b-bb^{\dagger})\qquad\text{% with}\qquad\{b,b^{\dagger}\}=1\,,\quad\{b,b\}=\{b^{\dagger},b^{\dagger}\}=0\,. (331)

Write HFH_{F} and HBH_{B} in terms of number operators NF=bbN_{F}=b^{\dagger}b and NB=aaN_{B}=a^{\dagger}a. What are the allowed eigenvalues of NFN_{F} and NBN_{B}?

You can also define a combined system H=HB+HFH=H_{B}+H_{F} with |n=|nB|nF|nB,nF|{n}\rangle=|{n_{B}}\rangle\otimes|{n_{F}}\rangle\equiv|{n_{B},n_{F}}\rangle. This system treats bosons and fermions the same and is therefore supersymmetric. Show that the supercharge operator Q=abQ=ab^{\dagger} fulfils

{Q,Q}={Q,Q}=0,ω{Q,Q}=H,[Q,H]=[Q,H]=0.\displaystyle\{Q,Q\}=\{Q^{\dagger},Q^{\dagger}\}=0\,,\quad\omega\{Q,Q^{\dagger% }\}=H\,,\quad[Q,H]=[Q^{\dagger},H]=0\,. (332)

Therefore, QQ is a conserved quantity. Finally, apply Q|nB,nFQ|{n_{B},n_{F}}\rangle and explain what the operator does to a fermion or boson.

We can now define a few states. As before, we define |0|{0}\rangle as the state destroyed by aa and bb

ar(p)|0=br(p)|0=0.\displaystyle a_{r}(\vec{p})|{0}\rangle=b_{r}(\vec{p})|{0}\rangle=0\,. (333)

We can also define two different one-particles states

|p,s,+=2Epas(p)|0and|p,s,=2Epbs(p)|0.\displaystyle|{\vec{p},s,+}\rangle=\sqrt{2E_{\vec{p}}}a_{s}^{\dagger}(\vec{p})% |{0}\rangle\qquad\text{and}\qquad|{\vec{p},s,-}\rangle=\sqrt{2E_{\vec{p}}}b_{s% }^{\dagger}(\vec{p})|{0}\rangle\,. (334)

These states are properly normalised such that

p,s,±|q,r,±=2Ep(2π)3δ(3)(pq)δrs.\displaystyle\langle{\vec{p},s,\pm}|{\vec{q},r,\pm}\rangle=2E_{\vec{p}}\,(2\pi% )^{3}\delta^{(3)}(\vec{p}-\vec{q})\delta_{rs}\,. (335)

5.1.5 Charge of the Dirac field and bilinear forms

The Dirac Lagrangian (319) has a symmetry ψeiαψ\psi\to{\rm e}^{i\alpha}\psi which means that there must be conserved current.

This current is
jVμ=ψ¯(x)γμψ(x).\displaystyle j_{V}^{\mu}=\bar{\psi}(x)\gamma^{\mu}\psi(x)\,. (336a)
It is customary to also define
jS\displaystyle j_{S} =ψ¯(x)ψ(x),\displaystyle=\bar{\psi}(x)\psi(x)\,, (336b)
j5\displaystyle j_{5} =ψ¯(x)γ5ψ(x),\displaystyle=\bar{\psi}(x)\gamma^{5}\psi(x)\,, (336c)
j5Vμ\displaystyle j_{5V}^{\mu} =ψ¯(x)γ5γμψ(x).\displaystyle=\bar{\psi}(x)\gamma^{5}\gamma^{\mu}\psi(x)\,. (336d)

It is easy to see that jVμj_{V}^{\mu} and j5Vμj_{5V}^{\mu} are conserved

μjVμ=(μψ¯)γμψ+ψ¯γμ(μψ)=(imψ¯)ψ+ψ¯(imψ)=0,\displaystyle\partial_{\mu}j_{V}^{\mu}=\big{(}\partial_{\mu}\bar{\psi}\big{)}% \gamma^{\mu}\psi+\bar{\psi}\gamma^{\mu}\big{(}\partial_{\mu}\psi\big{)}=(im% \bar{\psi})\psi+\bar{\psi}(-im\psi)=0\,, (337)
Suggested Exercise

Show that j5Vj_{5V} is conserved as well as long as m=0m=0.

Let us also see how jj transforms under Lorentz transformation. For example,

jSψ¯(x)ψ(x)=ψ¯(x)S(Λ)1S(Λ)ψ(x)=ψ¯(x)ψ(x)=jS(x).\displaystyle j_{S}\to\bar{\psi}^{\prime}(x^{\prime})\psi^{\prime}(x^{\prime})% =\bar{\psi}(x)S(\Lambda)^{-1}S(\Lambda)\psi(x)=\bar{\psi}(x)\psi(x)=j_{S}(x)\,. (338)
Suggested Exercise

Use the proof to show that

jVμ(x)(jV)μ(x)=ΛνμjVν(x),\displaystyle j_{V}^{\mu}(x)\to(j_{V}^{\prime})^{\mu}(x^{\prime})=\Lambda^{\mu% }_{\phantom{\mu}\nu}j_{V}^{\nu}(x)\,, (339)

and similarly for j5j_{5} and j5Vμj_{5V}^{\mu}.

The effect of parity is slightly more interesting. For example,

jS\displaystyle j_{S} ψ¯(x)ψ(x)=ψ¯(x)S(P)1S(P)ψ(x)=ψ¯(x)ψ(x)=jS,\displaystyle\to\bar{\psi}^{\prime}(x^{\prime})\psi^{\prime}(x^{\prime})=\bar{% \psi}(x)S(P)^{-1}S(P)\psi(x)=\bar{\psi}(x)\psi(x)=j_{S}\,, (340)
j5\displaystyle j_{5} ψ¯(x)(γ)5ψ(x)=ψ¯(x)S(P)1γ5S(P)ψ(x)=ψ¯(x)γ5ψ(x)=j5,\displaystyle\to\bar{\psi}^{\prime}(x^{\prime})(\gamma^{\prime})^{5}\psi^{% \prime}(x^{\prime})=\bar{\psi}(x)S(P)^{-1}\gamma^{5}S(P)\psi(x)=-\bar{\psi}(x)% \gamma^{5}\psi(x)=-j_{5}\,, (341)

and similarly for the vector currents. We have used that

S(P)1γ5S(P)=γ0γ5γ0=γ0γ0γ5=γ5.\displaystyle S(P)^{-1}\gamma^{5}S(P)=\gamma^{0}\gamma^{5}\gamma^{0}=-\gamma^{% 0}\gamma^{0}\gamma^{5}=-\gamma^{5}\,. (342)

This is the original of the labels we have used for the different basis elements in (286): since jSj_{S} (jVj_{V}) transforms like a Lorentz scalar (Lorentz vector) we call it a scalar (vector) current. The “pseudo” prefix indicates that the current picks up a sign under parity conservation, the same way that e.g. the angular momentum L=x×p\vec{L}=\vec{x}\times\vec{p} does.

Classically, the vector current jVj_{V} corresponds to the electromagnetic current with the charge density as ρ=jV0\rho=j_{V}^{0}. Let us calculate this current for our QFT

Q\displaystyle Q =d3xρ(x)=d3xψ(x)ψ(x)\displaystyle=\int{\rm d}^{3}x\ \rho(x)=\int{\rm d}^{3}x\ \psi^{\dagger}(x)% \psi(x) (343)
=d3xd3qd3p(2π)62Ep2Eqr,s(ar(p)ur(p)e+ipx+br(p)vr(p)eipx)(ar(p)ur(p)eipx+br(p)vr(p)e+ipx)\displaystyle=\int\frac{{\rm d}^{3}x\ {\rm d}^{3}\vec{q}\ {\rm d}^{3}\vec{p}}{% (2\pi)^{6}\sqrt{2E_{\vec{p}}2E_{\vec{q}}}}\sum_{r,s}\Big{(}a_{r}^{\dagger}(% \vec{p})u_{r}(p)^{\dagger}{\rm e}^{+ip\cdot x}+b_{r}(\vec{p})v_{r}(p)^{\dagger% }{\rm e}^{-ip\cdot x}\Big{)}\Big{(}a_{r}(\vec{p})u_{r}(p){\rm e}^{-ip\cdot x}+% b_{r}^{\dagger}(\vec{p})v_{r}(p){\rm e}^{+ip\cdot x}\Big{)} (344)
=d3p(2π)3r(ar(p)ar(p)+br(p)br(p))=d3p(2π)3r(ar(p)ar(p)br(p)br(p)),\displaystyle=\int\frac{{\rm d}^{3}\vec{p}}{(2\pi)^{3}}\sum_{r}\Big{(}a_{r}^{% \dagger}(\vec{p})a_{r}(\vec{p})+b_{r}(\vec{p})b_{r}^{\dagger}(\vec{p})\Big{)}=% \int\frac{{\rm d}^{3}\vec{p}}{(2\pi)^{3}}\sum_{r}\Big{(}a_{r}^{\dagger}(\vec{p% })a_{r}(\vec{p})-b_{r}^{\dagger}(\vec{p})b_{r}(\vec{p})\Big{)}\,, (345)

where we have dropped yet another d3p1\int{\rm d}^{3}p1 constant in the last step. This proves that the particles created by aa^{\dagger} have charge Q=+1Q=+1 and the those created by bb^{\dagger} have Q=1Q=-1 Therefore, we call the former particles and the latter antiparticles.

Suggested Exercise

Show the above.

5.1.6 Dirac propagator

Finally, we should calculate the propagator of a fermion. To do this, we simply write

0|ψa(x)ψ¯b(y)|0\displaystyle\langle{0}|\psi_{a}(x)\bar{\psi}_{b}(y)|{0}\rangle =d3p(2π)32Epeip(xy)rur(p)au¯r(p)b(γp+m)ab,\displaystyle=\int\frac{{\rm d}^{3}p}{(2\pi)^{3}2E_{\vec{p}}}{\rm e}^{ip\cdot(% x-y)}\underbrace{\sum_{r}u_{r}(p)_{a}\bar{u}_{r}(p)_{b}}_{(\gamma\cdot p+m)_{% ab}}\,, (346)
0|ψ¯b(y)ψa(x)|0\displaystyle\langle{0}|\bar{\psi}_{b}(y)\psi_{a}(x)|{0}\rangle =d3p(2π)32Epeip(yy)rvr(p)av¯r(p)b(γpm)ab.\displaystyle=\int\frac{{\rm d}^{3}p}{(2\pi)^{3}2E_{\vec{p}}}{\rm e}^{ip\cdot(% y-y)}\underbrace{\sum_{r}v_{r}(p)_{a}\bar{v}_{r}(p)_{b}}_{(\gamma\cdot p-m)_{% ab}}\,. (347)

Keep in mind that ψ\psi and ψ¯\bar{\psi} are operator-valued vector fields and therefore have spinor indices aa and bb, not to be confused with the operators aa and bb. By writing pμiμp_{\mu}\to i\partial_{\mu}, we can pull the Dirac structure out and are left with D(xy)D(x-y) of (116)

0|ψa(x)ψ¯b(y)|0\displaystyle\langle{0}|\psi_{a}(x)\bar{\psi}_{b}(y)|{0}\rangle =+(iγx+m)abD(xy),\displaystyle=+(i\gamma\cdot\partial_{x}+m)_{ab}D(x-y)\,, (348)
0|ψ¯b(y)ψa(x)|0\displaystyle\langle{0}|\bar{\psi}_{b}(y)\psi_{a}(x)|{0}\rangle =(iγxm)abD(yx).\displaystyle=-(i\gamma\cdot\partial_{x}-m)_{ab}D(y-x)\,. (349)

Similarly, the delayed Green’s function can be written as

SRab(xy)=θ(x0y0)0|{ψa(x),ψ¯b(y)}|0=(iγx+m)abDR(xy).\displaystyle S_{R}^{ab}(x-y)=\theta(x^{0}-y^{0})\langle{0}|\{\psi_{a}(x),\bar% {\psi}_{b}(y)\}|{0}\rangle=(i\gamma\cdot\partial_{x}+m)_{ab}D_{R}(x-y)\,. (350)
Suggested Exercise

Verify that SRS_{R} is indeed a Green’s function of the Dirac operator imi\partial-m.

For the Fourier-transformed Green’s function, we find

S~R(xy)=i(γp+m)p2m2=iγpm.\displaystyle\tilde{S}_{R}(x-y)=\frac{i(\gamma\cdot p+m)}{p^{2}-m^{2}}=\frac{i% }{\gamma\cdot p-m}\,. (351)

We can use the same constructions to define the Feynman propagator

SF(xy)=d4p(2π)4iγpm+iϵeip(xy)=0|Tψ(x)ψ¯(y)|0.\displaystyle S_{F}(x-y)=\int\frac{{\rm d}^{4}p}{(2\pi)^{4}}\frac{i}{\gamma% \cdot p-m+i\epsilon}{\rm e}^{-ip\cdot(x-y)}=\langle{0}|T\psi(x)\bar{\psi}(y)|{% 0}\rangle\,. (352)

5.2 Vector fields

The last particle we will consider is the photon, i.e. the particle of the electromagnetic field. The classical Lagrangian is just

=14FμνFμνwithFμν=μAννAμ.\displaystyle\mathcal{L}=-\frac{1}{4}F^{\mu\nu}F_{\mu\nu}\quad\text{with}\quad F% ^{\mu\nu}=\partial^{\mu}A^{\nu}-\partial^{\nu}A^{\mu}\,. (353)

FμνF^{\mu\nu} is called the field-strength tensor and AμA^{\mu} the vector potential. We will not discuss how to quantise this field because AA is a gauge field which makes canonical quantisation much more complicated. This is because the conjugate momentum to A0A^{0} is zero as \mathcal{L} does not contain A˙0\dot{A}^{0} (the term 0A0\partial^{0}A^{0} would be in F00=0A00A0=0F^{00}=\partial^{0}A^{0}-\partial^{0}A^{0}=0). The way to circumvent this problem is add a gauge-fixing term to \mathcal{L} that forces a specific gauge, e.g. Lorentz gauge, i.e. μAμ=0\partial_{\mu}A^{\mu}=0, in which we can write down CCRs

[Aμ(x),A˙ν(y)]=iημνδ(xy).\displaystyle[A^{\mu}(x),\dot{A}^{\nu}(y)]=-i\eta^{\mu\nu}\delta(x-y)\,. (354)

This is very similar the KG field and we can almost proceed along the same lines88 8 There is one more problem related to the fact that states created by a0a^{0{\dagger}} have negative norm.. The photon field can be written as

Aμ(x)=d3p(2π)32Epλϵλμ(p)[ap,λeipx+ap,λeipx].\displaystyle A^{\mu}(x)=\int\frac{{\rm d}^{3}p}{(2\pi)^{3}\sqrt{2E_{\vec{p}}}% }\sum_{\lambda}\epsilon^{\mu}_{\lambda}(p)\Big{[}a_{p,\lambda}{\rm e}^{ip\cdot x% }+a_{p,\lambda}^{\dagger}{\rm e}^{-ip\cdot x}\Big{]}\,. (355)

Here we use λ\lambda to sum over polarisations and ϵ\epsilon to denote the polarisation vector itself. Naively we would expect two polarisations. However, the gauge fixing leads to two unphysical polarisations that we also need to sum over. Like the uu and vv spinors, the polarisation vector ϵ\epsilon has a complenetess relation

λϵλμ(p)ϵλν(p)=ημν.\displaystyle\sum_{\lambda}\epsilon_{\lambda}^{\mu}(p)\epsilon_{\lambda}^{*\nu% }(p)=-\eta^{\mu\nu}\,. (356)

The Feynman propagator of this theory is

DFμν(xy)=d4p(2π)4ημνp2+iϵeip(xy).\displaystyle D_{F}^{\mu\nu}(x-y)=\int\frac{{\rm d}^{4}p}{(2\pi)^{4}}\frac{% \eta^{\mu\nu}}{p^{2}+i\epsilon}{\rm e}^{-ip\cdot(x-y)}\,. (357)