Lecture Notes for MATH4017 Quantum Field Theory

\(\newcommand{\footnotename}{footnote}\) \(\def \LWRfootnote {1}\) \(\newcommand {\footnote }[2][\LWRfootnote ]{{}^{\mathrm {#1}}}\) \(\newcommand {\footnotemark }[1][\LWRfootnote ]{{}^{\mathrm {#1}}}\) \(\let \LWRorighspace \hspace \) \(\renewcommand {\hspace }{\ifstar \LWRorighspace \LWRorighspace }\) \(\newcommand {\mathnormal }[1]{{#1}}\) \(\newcommand \ensuremath [1]{#1}\) \(\newcommand {\LWRframebox }[2][]{\fbox {#2}} \newcommand {\framebox }[1][]{\LWRframebox } \) \(\newcommand {\setlength }[2]{}\) \(\newcommand {\addtolength }[2]{}\) \(\newcommand {\setcounter }[2]{}\) \(\newcommand {\addtocounter }[2]{}\) \(\newcommand {\arabic }[1]{}\) \(\newcommand {\number }[1]{}\) \(\newcommand {\noalign }[1]{\text {#1}\notag \\}\) \(\newcommand {\cline }[1]{}\) \(\newcommand {\directlua }[1]{\text {(directlua)}}\) \(\newcommand {\luatexdirectlua }[1]{\text {(directlua)}}\) \(\newcommand {\protect }{}\) \(\def \LWRabsorbnumber #1 {}\) \(\def \LWRabsorbquotenumber "#1 {}\) \(\newcommand {\LWRabsorboption }[1][]{}\) \(\newcommand {\LWRabsorbtwooptions }[1][]{\LWRabsorboption }\) \(\def \mathchar {\ifnextchar "\LWRabsorbquotenumber \LWRabsorbnumber }\) \(\def \mathcode #1={\mathchar }\) \(\let \delcode \mathcode \) \(\let \delimiter \mathchar \) \(\def \oe {\unicode {x0153}}\) \(\def \OE {\unicode {x0152}}\) \(\def \ae {\unicode {x00E6}}\) \(\def \AE {\unicode {x00C6}}\) \(\def \aa {\unicode {x00E5}}\) \(\def \AA {\unicode {x00C5}}\) \(\def \o {\unicode {x00F8}}\) \(\def \O {\unicode {x00D8}}\) \(\def \l {\unicode {x0142}}\) \(\def \L {\unicode {x0141}}\) \(\def \ss {\unicode {x00DF}}\) \(\def \SS {\unicode {x1E9E}}\) \(\def \dag {\unicode {x2020}}\) \(\def \ddag {\unicode {x2021}}\) \(\def \P {\unicode {x00B6}}\) \(\def \copyright {\unicode {x00A9}}\) \(\def \pounds {\unicode {x00A3}}\) \(\let \LWRref \ref \) \(\renewcommand {\ref }{\ifstar \LWRref \LWRref }\) \( \newcommand {\multicolumn }[3]{#3}\) \(\require {textcomp}\) \(\newcommand {\intertext }[1]{\text {#1}\notag \\}\) \(\let \Hat \hat \) \(\let \Check \check \) \(\let \Tilde \tilde \) \(\let \Acute \acute \) \(\let \Grave \grave \) \(\let \Dot \dot \) \(\let \Ddot \ddot \) \(\let \Breve \breve \) \(\let \Bar \bar \) \(\let \Vec \vec \) \(\newcommand {\slashed }[1]{\cancel {#1}}\) \(\newcommand {\tcbset }[1]{}\) \(\newcommand {\tcbsetforeverylayer }[1]{}\) \(\newcommand {\tcbox }[2][]{\boxed {\text {#2}}}\) \(\newcommand {\tcboxfit }[2][]{\boxed {#2}}\) \(\newcommand {\tcblower }{}\) \(\newcommand {\tcbline }{}\) \(\newcommand {\tcbtitle }{}\) \(\newcommand {\tcbsubtitle [2][]{\mathrm {#2}}}\) \(\newcommand {\tcboxmath }[2][]{\boxed {#2}}\) \(\newcommand {\tcbhighmath }[2][]{\boxed {#2}}\) \(\newcommand {\bbR }{\mathbb {R}}\) \(\newcommand {\bbC }{\mathbb {C}}\) \(\newcommand {\bbZ }{\mathbb {Z}}\) \(\newcommand {\sk }{\vspace {2mm}}\) \(\newcommand {\dd }{\mathrm {d}}\) \(\newcommand {\ii }{\mathrm {i}}\) \(\newcommand {\nn }{\nonumber }\) \(\newcommand {\HH }{\mathcal {H}}\) \(\newcommand {\LL }{\mathcal {L}}\) \(\newcommand {\TO }{\mathsf {T}}\) \(\newcommand {\Ker }{\mathrm {Ker}}\) \(\newcommand {\Imm }{\mathrm {Im}}\) \(\newcommand {\g }{\mathfrak {g}}\) \(\newcommand {\U }{\mathsf {U}}\) \(\newcommand {\SU }{\mathsf {SU}}\) \(\newcommand {\ket }[1]{\vert #1\rangle }\) \(\newcommand {\bra }[1]{\langle #1\vert }\) \(\newcommand {\braket }[2]{\langle #1\vert #2 \rangle }\) \(\newcommand {\expect }[3]{\langle #1\vert #2 \vert #3 \rangle }\) \(\newcommand {\noor }[1]{{:} #1 {:}}\)

2.2 Symmetries and Noether’s theorem

Many classical field theories admit symmetries, by which one means transformations \(T : \phi (x) \mapsto T\phi (x)\) on the set of fields that leave invariant the action functional, i.e.

\begin{equation} S[T\phi ] = S[\phi ] \quad . \end{equation}

For instance, all examples presented in Section 2.1 are invariant under Poincaré transformations because their Lagrangian densities are constructed via the tensor calculus on Minkowski spacetime. Let us illustrate this in detail for the example of the real Klein-Gordon field from Example 2.1.

  • Example 2.4 (Poincaré symmetries of the real Klein-Gordon field). Recall our conventions for Poincaré transformations \(x^{\prime \mu } =\Lambda ^{\mu }_{~~\nu } \,x^\nu + b^\mu \) from Section 1.4. A real scalar field transforms under Poincaré transformations as \(\Phi ^\prime (x^\prime ) = \Phi (x)\), i.e. the value of the transformed field \(\Phi ^\prime \) at the transformed point \(x^\prime \) is equal to the value of \(\Phi \) at \(x\). This point of view on Poincaré transformations is called passive since the field \(\Phi \) does not get transformed to a new field, but it is only expressed in a different set of coordinates \(x^\prime \). The active point of view on Poincaré transformations consists of evaluating the transformed field \(\Phi ^\prime \) at the old point \(x\) and interpreting this as a transformation \(T : \Phi (x) \mapsto T\Phi (x) = \Phi ^\prime (x)\) on the set of fields. Inverting the Poincaré transformation \(x^\prime =\Lambda x + b\) gives \(x = \Lambda ^{-1}(x^\prime - b)\), which when combined with the transformation law \(\Phi ^\prime (x^\prime ) = \Phi (x)\) yields

    \begin{flalign} \label {eqn:Poincarescalarfield} T : \Phi (x) \longmapsto T\Phi (x) = \Phi ^\prime (x) = \Phi \big (\Lambda ^{-1}(x - b)\big )\quad . \end{flalign} In index notation, the argument of \(\Phi \) reads as \((\Lambda ^{-1}(x - b))^\nu = (x^\rho - b^\rho )\,\Lambda _\rho ^{~~\nu }\). Using the chain rule for partial differentiation, we find that the partial derivatives of the transformed field \(T\Phi \) are given by

    \begin{flalign} \nn (\partial _\mu T\Phi )(x) &= \frac {\partial }{\partial x^\mu } \Phi \big (\Lambda ^{-1}(x - b)\big )\\ \nn &= \frac {\partial \big ((x^\rho - b^\rho )\,\Lambda _\rho ^{~~\nu }\big )}{\partial x^\mu }~~(\partial _\nu \Phi )\big (\Lambda ^{-1}(x - b)\big )\\ &=\Lambda _{\mu }^{~~\nu }\,(\partial _\nu \Phi )\big (\Lambda ^{-1}(x - b)\big )\quad , \end{flalign} which we recognize as the transformation law of a covector field under (active) Poincaré transformations. With these preparations, we can now confirm Poincaré invariance of the Klein-Gordon action (2.8) by a simple calculation

    \begin{flalign} \nn S_{\mathrm {KG}}[T\Phi ] &= \int _{\bbR ^d}-\frac {1}{2}\,\bigg (\eta ^{\mu \nu }\,(\partial _\mu T\Phi )(x) \, (\partial _{\nu }T\Phi )(x) + m^2\,(T\Phi )^2(x)\bigg )\,\dd x\\ \nn &= \int _{\bbR ^d}-\frac {1}{2}\,\bigg (\eta ^{\mu \nu }\,\Lambda _{\mu }^{~~\rho }\,\Lambda _\nu ^{~~\sigma }\, (\partial _\rho \Phi )\big (\Lambda ^{-1}(x - b)\big ) \,(\partial _\sigma \Phi )\big (\Lambda ^{-1}(x - b)\big ) + m^2\,\Phi ^2\big (\Lambda ^{-1}(x - b)\big )\bigg )\,\dd x\\ \nn &= \int _{\bbR ^d}-\frac {1}{2}\,\bigg (\eta ^{\rho \sigma }\, (\partial _\rho \Phi )\big (\Lambda ^{-1}(x - b)\big ) \,(\partial _\sigma \Phi )\big (\Lambda ^{-1}(x - b)\big ) + m^2\,\Phi ^2\big (\Lambda ^{-1}(x - b)\big )\bigg )\,\dd x\\ &= \int _{\bbR ^d}-\frac {1}{2}\,\bigg (\eta ^{\rho \sigma }\, (\partial _\rho \Phi )(y) \,(\partial _\sigma \Phi )(y) + m^2\,\Phi ^2(y)\bigg )\,\dd y = S_{\mathrm {KG}}[\Phi ]\quad . \end{flalign} The third step uses the property (1.19) of the inverse Minkowski metric. In the fourth step we have changed the integration variables to \(y:= \Lambda ^{-1}(x- b)\) and used the change of variables formula \(\dd y = \vert \det (\Lambda ^{-1})\vert \, \dd x = \dd x\) for the volume element, together with the fact that \(\det (\Lambda ) = \pm 1\) which follows by taking determinants on both sides of (1.17). (Since as the default we consider proper and orthochronous Poincaré transformations, we even have \(\det (\Lambda ) = + 1\).)

Poincaré symmetries are an example of so-called spacetime symmetries, which are by definition transformations \(T : \phi (x) \mapsto T\phi (x)\) such that the value of the transformed field \(T\phi (x)\) at the point \(x\) is determined from the value of \(\phi \) at a different point. Compare this with (2.22), which shows that the Poincaré transformed scalar field \(T\Phi \) at \(x\) is determined from the field \(\Phi \) at \(\Lambda ^{-1}(x - b)\). Another type of symmetries is given by the so-called internal symmetries for which the transformed field \(T\phi (x)\) at \(x\) is determined from the field \(\phi (x)\) at the same point \(x\). It is again best to illustrate this by an example.

  • Example 2.5 (Internal symmetries of the real and complex Klein-Gordon field). For a real scalar field \(\Phi (x)\in \bbR \), we can consider the transformation \(T : \Phi (x) \mapsto T\Phi (x) = -\Phi (x)\) that flips the sign. This transformation is a symmetry of the Klein-Gordon action (2.8), as one can easily check

    \begin{flalign} \nn S_{\mathrm {KG}}[T\Phi ] &= \int _{\bbR ^d}-\frac {1}{2}\,\bigg (\eta ^{\mu \nu }\,(\partial _\mu T\Phi )(x) \, (\partial _{\nu }T\Phi )(x) + m^2\,(T\Phi )^2(x)\bigg )\,\dd x\\ &=(-1)^2 \int _{\bbR ^d}-\frac {1}{2}\,\bigg (\eta ^{\mu \nu }\,(\partial _\mu \Phi )(x) \, (\partial _{\nu }\Phi )(x) + m^2\,\Phi ^2(x)\bigg )\,\dd x = S_{\mathrm {KG}}[\Phi ]\quad . \end{flalign} Note that this transformation is also a symmetry of the action functional (2.10) with interaction terms, provided that the potential \(V(\Phi )\) is an even function, i.e. \(V(-\Phi ) = V(\Phi )\). Since \(T\Phi (x)\) at \(x\) is determined from \(\Phi (x)\) at the same point \(x\), this is an example of an internal symmetry.

    The case of a complex scalar field \(\Phi (x)\in \bbC \) is richer, in the sense that there is not only one but an infinite family of internal symmetries that is labeled by a continuous parameter. Consider the transformation \(T : \Phi (x)\mapsto T\Phi (x) = e^{-\ii \,\alpha }\,\Phi (x)\) that rotates the field \(\Phi \) by any constant complex phase \(e^{-\ii \, \alpha }\in \mathsf {U}(1)\subseteq \bbC \), i.e. \(\alpha \) is taken to be independent of \(x\). The transformation of the complex conjugate field \(\Phi ^\ast \) is then given by \(T: \Phi ^\ast (x)\mapsto e^{\ii \,\alpha }\,\Phi ^\ast (x)\). The action (2.17) and its special case (2.19) are clearly invariant under this transformation, because every summand involves an equal power of \(\Phi \) and \(\Phi ^\ast \), such that the complex phases cancel out. For example, for the quadratic terms we have that \(T\Phi ^\ast (x) \, T\Phi (x) = e^{\ii \,\alpha }\, \Phi ^\ast (x)\, e^{-\ii \,\alpha }\, \Phi (x) = \Phi ^\ast (x)\,\Phi (x)\). Since there are infinitely many complex phases \(e^{-\ii \,\alpha }\in \mathsf {U}(1)\subseteq \bbC \), corresponding to choosing \(\alpha \in [0,2\pi )\), we found a continuous family of internal symmetries for the complex scalar field. It is worthwhile to note that constancy of the phase \(e^{-\ii \,\alpha }\) is crucial for having a symmetry: If \(e^{-\ii \,\alpha (x)}\) would depend on \(x\), then the partial derivatives of the transformed field receive an extra term \((\partial _\mu T\Phi )(x) = \partial _\mu (e^{-\ii \,\alpha (x)}\,\Phi (x)) = e^{-\ii \,\alpha (x)}\, (\partial _\mu \Phi )(x) -\ii \,(\partial _\mu \alpha )(x)\, e^{\ii \,\alpha (x)}\,\Phi (x)\), which would spoil the invariance of the kinetic term in the actions (2.17) and (2.19). Such local, in the sense of \(x\)-dependent, internal transformations will play an important role later in this module when we will discuss gauge theories.

You might now ask: What’s the point about having field theories with symmetries? What do I gain from this? These are indeed very good questions that can be answered in different ways. Let me present you two particularly important answers.

  • 1. Field theories that admit Poincaré symmetries are compatible with the laws of special relativity. This means that they will be particularly useful for applications to high-energy physics, which involves very fast particles and hence is the natural habitat of special relativity.

  • 2. To each continuous family of symmetries there is an associated conserved current, which in particular yields a time-independent quantity called the conserved charge. This holds true independently of whether one has internal or spacetime symmetries, and it is know as Noether’s theorem.

Our goal for the rest of this section is to derive Noether’s theorem and to provide examples of conserved charges. The terminology continuous family of symmetries means that we have a whole family of transformations \(T_{\beta } : \phi (x)\mapsto T_\beta \phi (x)\), labeled by one or many continuous parameters \(\beta =(\beta _1,\dots ,\beta _N)\in \bbR ^N\), that leaves invariant the action \(S[T_\beta \phi ] = S[\phi ]\). The Poincaré transformations from Example 2.4 are an example, where the continuous parameters label the different rotation angles, Lorentz boosts and translations of a Poincaré transformation, and also the phase rotations for a complex scalar field from Example 2.5 are an example, where the parameter labels the complex phase. On the other hand, the sign transformation \(\Phi (x) \mapsto -\Phi (x)\) for a real scalar field is not a continuous family of symmetries, because there is no parameter we can vary.

Assuming that the transformation \(T_{0}\phi (x) = \phi (x)\) corresponding to the parameter \(\beta =0\) is the identity transformation and that \(T_\beta \phi (x)\) is differentiable at \(\beta =0\), we can Taylor expand

\begin{flalign} \label {eqn:expand1storder} T_\beta \phi (x) = \phi (x) + \delta _\beta \phi (x) + \mathcal {O}(\beta ^2) \end{flalign} to first order in \(\beta \) to obtain the infinitesimal transformation \(\delta _\beta \phi \) of \(\phi \) under the family of transformations \(T_\beta \). Expanding also the invariance condition \(S[T_\beta \phi ] = S[\phi ]\) of the action functional to first order in \(\beta \) and considering the integrands, one obtains that the infinitesimal transformation of the Lagrangian density

\begin{flalign} \label {eqn:FNoether} \delta _\beta \LL (\phi ,\partial \phi ) :=\frac {\partial \LL (\phi ,\partial \phi )}{\partial \phi _a}\,(\delta _\beta \phi )_a + \frac {\partial \LL (\phi ,\partial \phi )}{\partial (\partial _\mu \phi _a)}\,\partial _\mu (\delta _\beta \phi )_a =\partial _\mu F^\mu \end{flalign} must be the divergence of some vector field \(F^\mu (x)\). This now allows us to formulate and prove Noether’s theorem.

  • Theorem 2.6 (Noether’s theorem). Let \(\delta _\beta \phi \) be an infinitesimal symmetry transformation, e.g. obtained by Taylor expanding a continuous family of symmetries (2.26). Then the vector field

    \begin{equation} J^\mu :=\frac {\partial \LL (\phi ,\partial \phi )}{\partial (\partial _\mu \phi _a)} \,(\delta _\beta \phi )_a - F^\mu \end{equation}

    is a conserved current in the sense that its divergence

    \begin{flalign} \partial _\mu J^\mu = 0 \end{flalign} vanishes whenever the field \(\phi \) satisfies the Euler-Lagrange equations (2.6).

  • Proof. We compute

    \begin{flalign} \nn \partial _\mu J^\mu &= \partial _\mu \bigg (\frac {\partial \LL (\phi ,\partial \phi )}{\partial (\partial _\mu \phi _a)}\,(\delta _\beta \phi )_a \bigg ) - \partial _\mu F^\mu \\ \nn &=\partial _\mu \bigg (\frac {\partial \LL (\phi ,\partial \phi )}{\partial (\partial _\mu \phi _a)}\bigg ) ~(\delta _\beta \phi )_a + \frac {\partial \LL (\phi ,\partial \phi )}{\partial (\partial _\mu \phi _a)}~\partial _\mu (\delta _\beta \phi )_a - \frac {\partial \LL (\phi ,\partial \phi )}{\partial \phi _a}\,(\delta _\beta \phi )_a - \frac {\partial \LL (\phi ,\partial \phi )}{\partial (\partial _\mu \phi _a)}\,\partial _\mu (\delta _\beta \phi )_a \\ &=\bigg ( \partial _\mu \bigg (\frac {\partial \LL (\phi ,\partial \phi )}{\partial (\partial _\mu \phi _a)}\bigg ) - \frac {\partial \LL (\phi ,\partial \phi )}{\partial \phi _a}\bigg )~(\delta _\beta \phi )_a =0\quad . \end{flalign} In the second step we have used the Leibniz/product rule for \(\partial _\mu \) and inserted (2.27). In the third step we have used that the 2nd and the 4th term cancel each other. The last step follows by using the Euler-Lagrange equations (2.6).

  • Remark 2.7. The requirement that \(\phi \) satisfies the Euler-Lagrange equations (2.6) is in general necessary for the Noether current \(J^\mu \) to be conserved. Satisfying the Euler-Lagrange equations is often called on-shell in physics slang, which is why you will probably hear the terminology that \(J^\mu \) is an on-shell conserved current.

  • Remark 2.8. The reason why a conserved current \(J^\mu \) is so useful is that it has an associated conserved charge, which is defined by the integral

    \begin{flalign} Q(t) := \int _{\bbR ^{d-1}} J^0(t,\mathbf {x})\, \dd \mathbf {x} \end{flalign} of the \(0\)-component of \(J^\mu \) over the space coordinates \(\mathbf {x}\in \bbR ^{d-1}\). The charge is time-independent, hence the name conserved, as shown by the following calculation

    \begin{flalign} \frac {\dd Q(t)}{\dd t} = \int _{\bbR ^{d-1}} \frac {\partial J^0(t,\mathbf {x})}{\partial t}\, \dd \mathbf {x} = - \int _{\bbR ^{d-1}} \nabla \cdot \mathbf {J}(t,\mathbf {x})\, \dd \mathbf {x} = 0\quad , \end{flalign} where in the second step we have used current conservation \(0=\partial _\mu J^\mu = \frac {\partial J^0}{\partial t} + \nabla \cdot \mathbf {J}\) and in the third step Gauss’ integration theorem. (As usual, it is assumed that the fields fall off sufficiently fast at infinity so that there are no boundary terms.)

  • Example 2.9 (Relativistic momentum of the real Klein-Gordon field). Recall from Example 2.4 that translations of spacetime define a continuous family of symmetries \(T_b : \Phi (x)\mapsto T_b\Phi (x) = \Phi (x - b)\) for the real Klein-Gordon field. Taking the first order Taylor expansion as in (2.26) gives

    \begin{flalign} \delta _b\Phi = - b^\mu \,\partial _\mu \Phi \quad . \end{flalign} For the transformation of the Lagrangian density (2.27) one finds

    \begin{flalign} \delta _b\LL = \partial _\mu \big (-b^\mu \, \LL \big )\quad \Longrightarrow \quad F^\mu = -b^\mu \, \LL \quad . \end{flalign} The associated Noether current from Theorem 2.6 then reads as

    \begin{flalign} \nn J^\mu &= \partial ^\mu \Phi \, b^\nu \,\partial _\nu \Phi + b^\mu \LL \\ &= b_\nu ~\bigg (\partial ^\mu \Phi \,\partial ^\nu \Phi - \frac {1}{2}\,\eta ^{\mu \nu }\Big (\partial ^\rho \Phi \,\partial _\rho \Phi + m^2\,\Phi ^2\Big )\bigg )\quad . \end{flalign} The term in the parenthesis is called the energy-momentum tensor and it is often denoted by

    \begin{flalign} \label {eqn:EMtensorKG} T^{\mu \nu } := \partial ^\mu \Phi \,\partial ^\nu \Phi - \frac {1}{2}\,\eta ^{\mu \nu }\Big (\partial ^\rho \Phi \,\partial _\rho \Phi + m^2\,\Phi ^2\Big )\quad . \end{flalign} Note that current conservation \(\partial _\mu J^\mu = 0\), for all translation vectors \(b^\nu \), is equivalent to the conservation law \(\partial _\mu T^{\mu \nu } =0\) of the energy-momentum tensor that you have probably seen before in your relativity module. The conserved charges associated with the energy-momentum tensor define the relativistic momentum

    \begin{flalign} P^\mu := \int _{\bbR ^{d-1}} T^{0\mu }\,\dd \mathbf {x} \end{flalign} of the Klein-Gordon field. Explicitly, the \(\mu =0\)-component reads as

    \begin{flalign} E:= P^0 = \int _{\bbR ^{d-1}} T^{00}\,\dd \mathbf {x} = \int _{\bbR ^{d-1}} \bigg (\frac {1}{2}\dot {\Phi }^2 + \frac {1}{2}(\nabla \Phi )^2 +\frac {m^2}{2}\,\Phi ^2\bigg )\,\dd \mathbf {x}\quad , \end{flalign} where \(\dot {\Phi } := \frac {\partial \Phi }{\partial t}\) denotes the time derivative, and it describes the energy. (As we will see in Section 2.3, this agrees with the Hamiltonian.) The spatial components

    \begin{flalign} P^i = -\int _{\bbR ^{d-1}} \dot {\Phi }\,\partial ^i \Phi \,\dd \mathbf {x}\quad , \end{flalign} for \(i=1,\dots , d-1\), describe the momentum.

  • Example 2.10 (Relativistic angular momentum of the real Klein-Gordon field). Recalling again Example 2.4, we consider now the case of a Lorentz transformation \(x^\prime = \Lambda x\). To obtain a suitable continuous family of Lorentz transformations, we write \(\Lambda = e^{\omega }= \sum _{n=0}^\infty \frac {\omega ^n}{n!}\) as the exponential of a matrix \(\omega \). Expanding the condition (1.17) to first order in \(\omega \) yields

    \begin{flalign} \nn \eta _{\mu \nu }\,\Lambda ^{\mu }_{~~\rho }\,\Lambda ^\nu _{~~\sigma } &= \eta _{\mu \nu }\,\big (\delta ^{\mu }_{~~\rho } + \omega ^\mu _{~~\rho }\big )\,\big (\delta ^\nu _{~~\sigma } + \omega ^\nu _{~~\sigma }\big )+\mathcal {O}(\omega ^2) \\ &= \eta _{\rho \sigma } + \omega _{\rho \sigma } + \omega _{\sigma \rho } + \mathcal {O}(\omega ^2) = \eta _{\rho \sigma }\quad , \end{flalign} hence the index lowering \(\omega _{\mu \nu } := \eta _{\mu \rho }\,\omega ^{\rho }_{~~\nu }\) must be antisymmetric

    \begin{flalign} \label {eqn:omegaantisym} \omega _{\mu \nu } = - \omega _{\nu \mu }\quad . \end{flalign} One can check that the latter condition implies that \(\Lambda = e^{\omega } \) satisfies (1.17) to all orders in \(\omega \). Hence, we obtain a continuous family of symmetries \(T_\omega : \Phi (x)\mapsto T_\omega \Phi (x) = \Phi (e^{-\omega } x)\) for the real Klein-Gordon field. Taking the first order Taylor expansion as in (2.26) gives

    \begin{flalign} \delta _\omega \Phi = -\omega ^\mu _{~~\nu }\,x^\nu \,\partial _\mu \Phi \quad . \end{flalign} For the transformation of the Lagrangian density (2.27) one finds

    \begin{flalign} \delta _\omega \LL = -\omega ^{\mu }_{~~\nu }\,x^\nu \,\partial _\mu \LL = \partial _\mu \Big (-\omega ^\mu _{~~\nu } \,x^\nu \,\LL \Big )\quad \Longrightarrow \quad F^\mu = -\omega ^\mu _{~~\nu } \,x^\nu \,\LL \quad , \end{flalign} where in the second equality we have used that \(\omega ^{\mu }_{~~\mu }= \eta ^{\mu \nu }\omega _{\nu \mu }=0\) as a consequence of the antisymmetry condition (2.39). The associated Noether current from Theorem 2.6 then reads as

    \begin{flalign} J^\mu = - \omega _{\nu \rho }\,x^\nu \,T^{\mu \rho } = \frac {1}{2}\,\omega _{\nu \rho }\,\Big (x^\rho \,T^{\mu \nu }-x^\nu \,T^{\mu \rho } \Big ) \quad , \end{flalign} where \(T^{\mu \rho }\) denotes the energy-momentum tensor (2.36) and in the last step we used that \(\omega _{\nu \rho } = -\omega _{\rho \nu }\) is antisymmetric. Denoting the term in the parenthesis by

    \begin{flalign} (\mathcal {J}^\mu )^{\rho \nu } := x^\rho \,T^{\mu \nu }-x^\nu \,T^{\mu \rho }\quad , \end{flalign} current conservation \(\partial _\mu J^\mu =0\), for all parameters \(\omega _{\nu \rho }\), is equivalent to the conservation law \(\partial _\mu (\mathcal {J}^\mu )^{\rho \nu }=0\). The conserved charges associated with \((\mathcal {J}^\mu )^{\rho \nu } \) define the relativistic angular momentum

    \begin{flalign} \label {eqn:KGangularmomentum} L^{\rho \nu } := \int _{\bbR ^{d-1}} (\mathcal {J}^0)^{\rho \nu } ~\dd \mathbf {x} = \int _{\bbR ^{d-1}} \Big (x^\rho \,T^{0\nu } - x^{\nu }\,T^{0\rho }\Big ) ~\dd \mathbf {x} \end{flalign} of the Klein-Gordon field. Since the angular momentum (2.44) contains only contributions arising from the dynamics of the field, we can conclude that the Klein-Gordon field does not carry any internal angular momentum, a.k.a. spin.

  • Example 2.11 (Electric charge of the complex Klein-Gordon field). Recall from Example 2.5 that the complex Klein-Gordon field admits a continuous family of internal symmetries \(T_\alpha : \Phi (x)\mapsto T_\alpha \Phi (x) = e^{-\ii \,\alpha }\,\Phi (x)\). Taking the first order Taylor expansion as in (2.26) gives

    \begin{flalign} \delta _\alpha \Phi = -\ii \,\alpha \,\Phi \quad ,\qquad \delta _{\alpha }\Phi ^\ast = \ii \,\alpha \,\Phi ^\ast \quad . \end{flalign} For the transformation of the Lagrangian density (2.27) one finds

    \begin{flalign} \delta _\alpha \LL = 0 \quad \Longrightarrow \quad F^\mu = 0 \end{flalign} because the Lagrangian density (2.19) is invariant under phase rotations. The associated Noether current from Theorem 2.6 then reads as

    \begin{flalign} J^\mu = \frac {\partial \LL }{\partial (\partial _\mu \Phi )}\,\delta _\alpha \Phi + \frac {\partial \LL }{\partial (\partial _\mu \Phi ^\ast )}\,\delta _\alpha \Phi ^\ast = \alpha \,\Big ( \ii \,\partial ^\mu \Phi ^\ast \, \Phi - \ii \, \partial ^\mu \Phi \, \Phi ^\ast \Big )\quad . \end{flalign} The term in the parenthesis is called the electromagnetic current of the complex Klein-Gordon field and it is often denoted by

    \begin{flalign} j^\mu := \ii \,\Big ( \Phi \,\partial ^\mu \Phi ^\ast -\Phi ^\ast \, \partial ^\mu \Phi \Big )\quad . \end{flalign} The conserved charge associated with \(j^\mu \) defines the electric charge

    \begin{flalign} Q:= \ii \, \int _{\bbR ^{d-1}} \Big (\Phi ^\ast \, \dot {\Phi } - \Phi \,\dot {\Phi }^\ast \Big )\,\dd \mathbf {x}\quad , \end{flalign} where \(\dot {\Phi } = \frac {\partial \Phi }{\partial t} = \partial _0 \Phi = -\partial ^0\Phi \) denotes the time derivative. (The minus sign in the last step is due to \(\eta ^{00}=-1\) in the Minkowski metric.)