<<

Hope not without despair, despair not without hope! Seneca (4 BC–65) 4 Free Relativistic Particles and Fields

Having learned how the many-particle Schr¨odinger theory can be reformulated as a quantum field theory, we shall now try to find possible field theories for the descrip- tion of relativistic many-particle systems. This will first be done classically. The fields will be quantized in Chapter 7.

4.1 Relativistic Particles

The nonrelativistic energy-momentum relation used in the Schr¨odinger theory p2 ε(p)= (4.1) 2M is valid only for massive particles which move much slower than the of light [2] c =2.99792458 1010cm/sec. (4.2) × If particles are accelerated to large close to c this condition is no longer fulfilled. Instead of (4.1), the energy follows the relativistic law

ε(p)= c2p2 + c4M 2. (4.3) q In particular, the light particles themselves, the photons, follow this law with the M = 0. It will be convenient to replace the energy by the new variable

p0 ε(p)/c. (4.4) ≡ Then the relation (29.15) can be expressed as

p02 p2 = M 2c2. (4.5) − Thus, energy and momentum of a particle of mass are always such that the four- vector pµ =(p0,pi) (4.6) is situated on the upper hyperboloid with p0 > 0 in a four-dimensional energy- momentum space. This is called the mass shell of the particle of mass M. If the

240 4.1 Relativistic Particles 241 particles are massless, the hyperboloid degenerates into a cone, the so-called light cone. Since a free particle remains free when seen from any rotated, or uniformly moving, coordinate frame, energy and momentum transform in a way that keeps them always on the same mass shell. For a simple rotation of the frame this is obvious. The energy remains the same while the momentum p changes only its direction. For example, p may appear rotated around the z-axis by a transformation

′i i j p = R3(ϕ) jp , (4.7) where R3(ϕ) is the

cos ϕ sin ϕ 0 − R3(ϕ)  sin ϕ cos ϕ 0  . (4.8) ≡ 0 01     The angle ϕ is defined in such a way that, in the rotated frame, the momenta of the same particles appear rotated in the anticlockwise direction in the xy-plane, i.e., the coordinate axes are rotated clockwise with respect to the original frame. We speak of a passive rotation of the system. The effect is the same as if the observer had remained in the same frame but the experimental apparatus had been rotated in the anticlockwise sense, and with it all particle orbits. The transformations defined in this way are called active transformations. There are two equivalent ways of formulating all invariance principles, one based on the active and one on the passive way. In this text we shall use the passive way. The reader should be aware that different texts use different conventions and the formulas calculated in one cannot always be compared directly with those in the other, but may require changes which fortunately are rather straightforward. For a general rotation by an angle ϕ with an axis pointing in the direction of the 1 unit vector ³ˆ, the transformation has the matrix form

′i i j

p = R³ˆ (ϕ) jp . (4.9)

We shall also write, with a slightly shorter notation for the rotation matrix,

′i i j p = R j(³) p . (4.10)

Explicitly, this transformation reads

p = cos ϕ p + sin ϕ (³ˆ p)+ p . (4.11) ⊥ × ||

Here p||, p⊥ are the projections parallel and orthogonal to the rotation axis ³ˆ: ³ p (p ³ˆ) ˆ, p p p , (4.12) || ≡ · ⊥ ≡ − || 1Hats on vectors in this section denote unit vectors, not Schr¨odinger operators. 242 4 Free Relativistic Particles and Fields respectively. The set of all rotations form a group called the rotation group. Consider now another set of transformations in which the second frame moves with velocity v into the z-direction of the first. In the new frame, the z momentum − of the particle will appear increased. The particle appears boosted in the z-direction with respect to the original observer. The momenta in x- and y-directions are unaffected. Since the total four-momentum still satisfies the mass shell condition (4.3), the combination p02 p32 has to remain invariant. This implies that there must be a hyperbolic transformation− mixing p0 and p3 which may be parametrized by a hyperbolic angle ζ, called rapidity:

p′ 0 = cosh ζp0 + sinh ζp3, p′ 3 = sinh ζp0 + cosh ζp3. (4.13)

This is called a pure . We may write this transformation in a 4 4 -matrix form as × cosh ζ 0 0 sinh ζ µ 0 10 0 p′µ =   pν B (ζ)µ pν. (4.14) 0 01 0 ≡ 3 ν    sinh ζ 0 0 cosh ζ    ν  

The subscript 3 of B3 indicates that the particle is boosted into the z-direction. A similar matrix can be written down for x and y-directions. In an arbitrary direction ˆ , the matrix elements are

cosh ζ ζˆi sinh ζ

Bˆ(ζ) B( )= . (4.15)  ≡  ζˆi sinh ζ δij + ζˆiζˆj(cosh ζ 1)   −  The spatial velocity of a particle is given by

v ∂ε(p)/∂p. (4.16) ≡ In Schr¨odinger theory this is the velocity of a packet. In terms of v v , one defines the Einstein parameter ≡| |

1 γ = cosh ζ. (4.17) ≡ 1 v2/c2 − q With these , we can rewrite (4.15) as

γ γvi/c

B()= , (4.18)  γvi/c δij +(γ 1)vivj/v2   −  where (γ 1)vivj/v2 is equal to γ2vivj/c2(γ 1). − − 4.1 Relativistic Particles 243

By combining rotations and boosts, one obtains a 6-parameter manifold of ma-

trices

³  ³ Λ(, )= B( )R( ). (4.19) These are called proper Lorentz transformations. For all these, the combination

p′02 p′2 = p02 p2 = M 2c2 (4.20) − − is invariant. These matrices form a group, the proper . We can easily see that the Lorentz group allows reaching every momentum pµ on the mass shell µ by applying an appropriate group element to some fixed reference momentum pR. µ For example, if the particle has a mass M we may choose for pR the so-called rest momentum µ pR =(Mc, 0, 0, 0), (4.21)

and apply the boost in the pˆ-direction  Λ()= B( ), (4.22) with the rapidity given by

p0 p cosh ζ = , sinh ζ = | | . (4.23) Mc Mc With this, we can rewrite the general boost matrix (4.15) in the pure momentum form p0/M p pi/M 2c2

B()= | | . (4.24)  pi p /M 2c2 δij +ˆpipˆj (p0/M 1)   | | −  Instead of (4.22), we may use as a boost in the pˆ-direction the more general

expressions ³ Λ(p)= B()R( ), (4.25)

µ where R(³) is an arbitrary rotation. Also these leave the rest momentum pR invari- ant. In fact, the rotations form the largest subgroup of all proper Lorentz trans- µ formations which leaves the rest momentum pR invariant. It is referred to as the little group or Wigner group of a massive particle. It has an important physical significance since it serves to specify the intrinsic rotational degrees of freedom of the particle. If the particle is at rest it carries no orbital . If it happens that its quantum mechanical state remains completely invariant under the little group R, the particle must also have zero intrinsic angular momentum or zero . Besides this trivial representation, the little group being a rotation group can 1 3 have representations of any angular momentum s = 2 , 1, 2 ,... . In these cases, the state at rest has 2s + 1 components which are linearly recombined with each other upon rotations. 244 4 Free Relativistic Particles and Fields

The situation is quite different in the case of massless particles. They move with the of light and pµ cannot be brought to rest by a Lorentz transformation from the light cone. There is, however, another standard reference momentum from which one can generate all other momenta on the light cone. It is given by

pµ = (1, 0, 0, 1) p , (4.26) R | | with an arbitrary size of the spatial momentum p . It remains invariant under | | a different little group, which is again a three-parameter subgroup of the Lorentz group. The little groups will be discussed in detail in Section 4.15.3. It is useful to write the invariant expression (4.20) as a square of a four-vector pµ formed with the metric

1 1 g =   , (4.27) µν − 1  −   1     −  namely 2 µ ν p = gµνp p . (4.28) In general, we define a product between any two vectors as

pp′ g pµp′ν = p0p′0 pp′. (4.29) ≡ µν − Following Einstein’s summation convention, repeated greek indices are summed from zero to 3 [recall (2.101)]. A space with this scalar product is called . It is useful to introduce the covariant components of any vector vµ as

v g vν. (4.30) µ ≡ µν Then the scalar product can also be written as

′ ′µ pp = pµp . (4.31)

With this notation, the mass shell properties (4.20) for a particle before and after a Lorentz transformation simply reads

p′2 = p2 = M 2c2. (4.32)

Note that, apart from the minus signs in the metric (4.27), the mass shell condition p2 = p02 p12 p22 p32 = M 2c2 which is invariant under Lorentz transformations, is − − − 2 2 2 2 completely analogous to the spherical condition p4 + p1 + p2 + p3 = M 2c2 which is invariant under rotations in a four-dimensional euclidean space. Both groups are parametrized by six parameters associated with linear transformations in the six planes: the six Lorentz transformations in the planes 12, 23, 31;10, 20, 30, and the six rotation angles in the planes 12, 23, 31;14, 24, 34. In the case of the four- dimensional euclidean space these are all rotations forming the group of special 4.1 Relativistic Particles 245 orthogonal matrices called SO(4). The first letter S indicates the property “special”. A group of matrices is called special if all matrices have a unit determinant. By analogy, the proper Lorentz group is denoted by SO(1,3). The numbers in (1,3) indicate that, in the Minkowski metric (4.27), one diagonal element is equal to +1 and three are equal to 1. − The fact that all group elements are special follows from a direct calculation of the determinant of the matrices in (4.9) and (4.14). How do we have to describe the quantum mechanics of a free relativistic particle in Minkowski space? Energy and momenta p0 and pi must be related to the time and space derivatives of particle in the usual way ∂ ∂ ∂ pˆ0 = ih¯ ih¯ , pˆi = ih¯ . (4.33) ∂ct ≡ ∂x0 − ∂xi In relativistic notation these read ∂ pˆ = ih¯ . (4.34) µ ∂xµ Together with the coordinates, they satisfy the canonical commutation rules

[ˆpµ, pˆν] = 0, [xµ, xν] = 0, [ˆpµ, xν] = ihg¯ µν. (4.35)

We expect a spinless free particle with momenta pi to be described by a field φ(x) of the plane-wave type that is, analogous to a nonrelativistic wave (2.212):

−i(p0x0−pixi)/¯h −ipx/¯h φp(x)= e = e , (4.36) N N where is some normalization factor. Since the zeroth component p0 is fixed by N the mass shell condition (4.5), only the spatial momentum needs to be specified, just as in the nonrelativistic plane wave solutions (2.212). However, in contrast to those, there are now two solutions for each momentum p, one with energy p0 = ε(p)= √c2p2 + c4M 2, and one with p0 = ε(p)= √c2p2 + c4M 2. Thus we have two plane-wave solutions − −

(+) −i(ε(p)x0−pixi)/¯h −ipx/¯h φp (x)= e = e fp(x), N 0 i i N ≡N φ(−)(x)= e i(ε(p)x +p x )/¯h = f ∗ (x), (4.37) p N N −p of positive and negative energy, respectively. For later convenience, we have (+) introduced the notation fp(x) for the positive-energy solution φp (x), so that (−) ∗ φp (x)= f−p(x). ∗ 0 The energies of fp(x) and f−p(x) are seen by applyingp ˆ to these wave functions:

∗ ∗ i∂ fp(x)= ε(p)fp(x), i∂ f (x)= ε(p)f (x). (4.38) 0 0 −p − −p 246 4 Free Relativistic Particles and Fields

∗ ∗ The latter equation holds, of course, also for fp(x). The solutions fp(x) and fp(x) will also be called positive- and negative-frequency wave functions, respectively. If not stated differently, the zeroth component p0 will, from now on, always be identified with the positive energy ε(p). At this point we do not yet normalize the wave functions since we must first find a proper scalar product for calculating physical observables from these wave functions. This scalar product will be given in (4.177). We have stated previously that permissible energy-momentum states of a free particle can be realized by considering one and the same particle in different co- ordinate frames connected by a transformation Λ. Suppose that we change the coordinates of the same point as follows:

x x′ =Λx. (4.39) → Under this transformation the scalar product of any two vectors remains invariant:

x′y′ = xy. (4.40)

For rotations, this is obvious since xy = x0y0 x y. For Lorentz transformations the invariance is a direct consequence of the fact− that· the boost matrix (4.14) satisfies the relation µ µ′ gµµ′ B3(ζ) νB3(ζ) λ = gµλ, (4.41) or in matrix notation T B3 (ζ)gB3(ζ)= g. (4.42) The same relation holds obviously for the arbitrary boost matrix (4.15), and after a combination with all rotations for the general Lorentz transformation (4.19):

ΛT gΛ= g, (4.43) or µ ν gµν Λ λΛ κ = gλκ. (4.44) The invariance (4.40) of the scalar product is then verified in matrix notation as follows: x′y′ x′T gy′ = (Λx)T g(Λy)= xT ΛT gΛy = xT gy = xy. (4.45) ≡ This holds also for scalar products between momentum and coordinate vectors

p′x′ = px. (4.46)

If the metric were euclidean, this would be the definition of orthogonal matrices. In fact, in the notation (4.45) of scalar products in which the metric is suppressed, we may write (Λp)(Λx)= pΛ−1Λx = px, (4.47) so that there is no difference between the manipulation of orthogonal and Lorentz matrices. 4.2 Differential Operators for Lorentz Transformations 247

When changing spacetime coordinates from x to x′ =Λx, the plane of a particle behaves like

−ipΛ−1x′/¯h −i(Λp)x′/¯h ′ φp(x)= e = e = φp′ (x ). (4.48) N N This shows that in the new coordinates the same particle appears with a different momentum and energy: p′ =Λp. (4.49) Consider now an observable field φ(x) describing a particle which does not possess any intrinsic orientational degree of freedom, i.e., no spin. The field can be an arbitrary superposition of different plane wave functions. After a coordinate trans- formation it will still have the same value at the same spacetime point. Thus φ′(x′), as seen in the new frame, must be equal to φ(x) in the old frame:

φ′(x′)= φ(x). (4.50)

A field with this property is called a scalar field or, for historical reasons, a Klein- Gordon field [4].

4.2 Differential Operators for Lorentz Transformations

Equation (4.50) contains the same point of the physical system on both sides, labeled by different coordinates x and x′. For the derivation of consequences of symmetries (see Chapter 8), it is preferable to formulate the property (4.50) in the form of a transformation law at the same spacetime coordinates x (corresponding to different points of the physical system). Thus we shall express the transformation property (4.50) of a scalar field in the following form:

Λ φ(x) φ′ (x)= φ(Λ−1x). (4.51) −−−→ Λ For clarity, we have marked by a subscript Λ the transformation producing φ′(x). It is useful to realize that the inverse Lorentz transformation of the coordinates inside the field argument can also be achieved with the help of a differential . To find it we observe that the finite transformation matrices (4.9) and (4.18) can all be written in a convenient exponential form. We begin with the rotations. Consider the four-dimensional expression for the rotation (4.7) of the coordinate frame by an angle ϕ clockwise around the z-axis. It moves a point with the coordinates x to a point with the new coordinates x′µ = R (ϕ)µ xν, where R (ϕ) denotes the 4 4 3 ν 3 × -matrix 10 00 0 cos ϕ sin ϕ 0 R (ϕ)=   . (4.52) 3 0 sin ϕ −cos ϕ 0      00 01    248 4 Free Relativistic Particles and Fields

This can be written in the exponential form 00 00

 0 0 1 0  −iL3ϕ R3(ϕ) = exp  −  ϕ e . (4.53)  01 00  ≡       00 00     The matrix   0 000 0 010 L = i   (4.54) 3 − 0 1 0 0  −   0 000    is called the generator of this rotation within the Lorentz group. There are similar generators for rotations around x- and y-directions: 00 00 00 00 L = i   , (4.55) 1 − 00 01      0 0 1 0   −  000 0  0 0 0 1  L2 = i − . (4.56) − 000 0      010 0    For all three cases we may write the generators as 0 0 Li i , (4.57) ≡ − 0 ǫijk ! where ǫijk is the completely antisymmetric Levi-Civita with ǫ123 = 1 (see [12]). The pure rotation matrix (4.9) is given by the exponential

Λ= e−i³·L, (4.58) as can also be verified by expanding the exponential in a series. Let us now find the generators of the pure Lorentz transformations: First in the z-direction where we see, from (4.14), that the boost matrix is 0001  0000  B3(ζ) = exp   ζ  0000      1000    −iM3ζ   = e  ,  (4.59) with the generator 0001 0000 M = i   . (4.60) 3 0000      1000    4.2 Differential Operators for Lorentz Transformations 249

Similarly we have in the other directions

0100 1000 M = i   , (4.61) 1 0000      0000    0010 0000 M = i   . (4.62) 2 1000      0000    The general Lorentz transformation matrix (4.15) is given by the exponential

Λ= e−i·M, (4.63) as can be verified by expanding the exponential in a power series. The full Lorentz group is therefore generated by the six matrices Li, Mi, to be collectively denoted by Ga(a =1,..., 6). Every element of the group can be written

as  Λ= e−i(³·L+ ·M) e−iαaGa . (4.64)

≡ ³ If the exponential is expanded in a power series, one reobtains for  = 0 or = 0 the general transformation matrices (4.9) or (4.15), respectively. There exists a Lorentz-covariant way of specifying the generators of the Lorentz group. We introduce the 4 4-matrices × (Lµν )λκ = i(gµλgνκ gµκgνλ), (4.65) − labeled by the antisymmetric pair of indices µν, i.e.,

Lµν = Lνµ. (4.66) − There are 6 independent matrices which coincide with the generators of rotations and boosts as follows: 1 L = ǫ Ljk, (4.67) i 2 ijk 0i Mi = L . (4.68)

With the generators (4.65), we can write every element (4.459) of the Lorentz group as follows: −i 1 ω Lµν Λ= e 2 µν , (4.69) where the antisymmetric angular matrix ω = ω collects both, rotation angles µν − νµ and rapidities: k ωij = ǫijkϕ , (4.70) i ω0i = ζ . (4.71) 250 4 Free Relativistic Particles and Fields

Summarizing the notation we have set

1 i jk i 0i 1 ij 0i 1 µν  −i(³·L+ ·M) −i( ϕ ǫ L +ζ L ) −i( ω L +ω L ) −i ω L Λ= e = e 2 ijk = e 2 ij 0i = e 2 µν . (4.72)

Note that if the metric is euclidean, in which case it has the form

1 1 g = δ =   , (4.73) µν µν 1    1   µν   the situation is well familiar from basic matrix theorems in 4 dimensions. Then Λ would comprise all real orthogonal 4 4 -matrices which could be written as an exponential of all real antisymmetric 4× 4 -matrices. In our case only the iLs are × antisymmetric while the iMs are symmetric, a consequence of the minus signs in the Minkowski metric (4.27). The reason for writing the group elements in this exponential form in terms of six generators is that thereby the multiplication rules of infinitely many group elements can be completely reduced to the knowledge of the commutation rules among the six generators Ga = (Li, Mi) of rotations and boosts. This is a consequence of the Baker-Campbell-Hausdorff formula written in the form (see Appendix 4A)

A B A+B+ 1 [A,B]+ 1 [A−B,[A,B]]+... e e = e 2 12 . (4.74)

From this formula we find the multiplication rule

1 2 −iαaGa −iα Gb Λ1Λ2 = e e b 1 2 1 1 2 = exp iαaGa iαb Gb + [ iαaGa, iαb Gb] − − 2 − − 1 1 2 1 2 + [ i(αc αc )Gc, [ iαaGa, iαb Gb]] + ... . (4.75) 12 − − − − 

The exponent involves only commutators among Ga’s. For the Lorentz group these can be calculated from the explicit 4 4 -matrices (4.54)–(4.56) and (4.60)–(4.62). × The result is

[Li, Lj] = iǫijkLk, (4.76)

[Li, Mj] = iǫijkMk, (4.77) [M , M ] = iǫ L . (4.78) i j − ijk k This algebra of generators is called the Lie algebra of the group. The number of linearly independent matrices Ga (here 6) is called the rank r (here r = 6) of the Lie algebra. In the notation with the generators Ga, the algebra reads

[Ga,Gb]= ifabcGc. (4.79) 4.2 Differential Operators for Lorentz Transformations 251

The commutator of two generators is a linear combination of generators. The co- efficients fabc are called structure constants. They are completely antisymmetric in a, b, c, and satisfy the relation

fabdfdcf + fbcdfdaf + fcadfdbf =0. (4.80) This guarantees that the generators obey the Jacobi identity

[[Ga,Gb],Gc] + [[Gb,Gc],Ga] + [[Gc,Ga],Gb]=0, (4.81) which is the law of associativity for Lie Algebras. The relation (4.80) can easily be verified for the structure constants (4.76)–(4.78) using the identity for the ǫ-tensor

ǫijlǫlkm + ǫjklǫlim + ǫkilǫljm =0. (4.82) The Jacobi identity implies that the r matrices with r r elements × (F ) if (4.83) c ab ≡ − cab satisfy the commutation rules (4.79). They form the so-called adjoint representation of the Lie algebra. The matrix (4.57) for Li is precisely of this type. In terms of the matrices Fa of the adjoint representation, the commutation rules can also be written as [G ,G ]= (F ) G . (4.84) a b − c ab c Continuing the expansion in terms of commutators in the exponent of (4.75), the commutators can be executed successively, and one remains at the end with an expression 12 1 2 −iαa (α ,α )Ga Λ12 = e , (4.85) 12 1 2 with the parameters of the product αa being completely determined by αa and αa for any given structure constants fabc. µν In the tensor notation L for Li, Mi of Eqs. (4.67), (4.68), and with multipli- cation performed covariantly, so that products Lµν Lλκ have the matrix elements µν λκ τ (L )στ (L ) δ, the commutators read [Lµν , Lλκ]= i(gµλLνκ gµκLνλ + gνκLµλ gνλLµκ). (4.86) − − − Due to the antisymmetry in µ ν and λ κ it is sufficient to specify only the simpler commutators ↔ ↔ [Lµν , Lµλ]= igµµLνλ, no sum over µ, (4.87) − thereby omitting vanishing components in (4.86) in which none of the indices µν is equal to one of the indices λκ. After these preparations we are ready to derive a differential operator which achieves the transformation of the spacetime argument in (4.51). First we consider infinitesimal Lorentz matrices 1 Λ 1 i ω Lµν . (4.88) ≡ − 2 µν 252 4 Free Relativistic Particles and Fields

The transformation Λ x x′ =Λx (4.89) −−−→ can be written as an infinitesimal symmetry transformation 1 δ x = x′ x = i ω Lµν x. (4.90) s − − 2 µν Another way of expressing this coordinate transformation is 1 δ x = x′ x = i ω Lˆµν x, (4.91) s − − 2 µν where Lˆµν are the differential operators

Lˆµν i(xµ∂ν xν ∂µ). (4.92) ≡ − Inserting the 4 4 matrix generators (4.65) into (4.90), or the corresponding differ- × ential operators (4.92) into (4.91), the infinitesimal coordinate transformations are explicitly µ µ ν δsx = ω νx . (4.93) The associated general field transformation law (4.51) takes the infinitesimal form

Λ φ(x) φ(x)+ δ φ(x), with δ φ(x)= ωµ xν ∂ φ(x). (4.94) −−−→ s s − ν µ

By analogy with Eq. (4.91), we can rewrite δsφ(x) as 1 δ φ(x)= i ω Lˆµν φ(x), (4.95) s − 2 µν where Lˆµν are the differential operators

Lˆµν i(xµ∂ν xν ∂µ). (4.96) ≡ − In terms of the quantum mechanical momentum operators (4.34), these are equal to 1/h¯ times the operators of the four-dimensional angular momentum: 1 Lˆµν (xµpˆν xν pˆµ). (4.97) ≡ h¯ − They satisfy the same commutation relations (4.86) as the 4 4 -matrix generators Lµν of the Lorentz group. They form a representation of the× Lie algebra in terms of differential operators generating the Lorentz transformations. In working out the commutation rules among the differential operators Lˆµν , one conveniently uses the commutation rules between Lˆµν and xλ as well asp ˆλ:

[Lˆµν , xλ] = i(gµλxν gνλxµ)= (Lµν )λ xκ, (4.98) − − − κ [Lˆµν , pˆλ] = i(gµλpˆν gνλpˆµ)= (Lµν )λ pˆκ. (4.99) − − − κ 4.2 Differential Operators for Lorentz Transformations 253

These commutation rules identify xλ andp ˆλ as vector operators [recall the definition in (2.113)]. In general, an operator tλ1,···,λn is said to be a tensor operator of rank n if it is transformed by Lµν like xµ orp ˆµ in each tensor index: [Lˆµν , tˆλ1,...,λn ] = i[(gµλ1 tˆν,...,λn gνλ1 tˆµ,...,λn )+ ... +(gµλn tˆλ1,...,ν gνλn tˆλ1,...,µ)] − − − = (Lµν )λ1 tˆκλ2,...,λn (Lµν )λ2 tˆλ1κ,...,λn (Lµν )λn tˆλ1λ2,...,κ. (4.100) − κ − κ − κ The simplest examples for such tensor operators are tˆλ1,...,λn = xλ1 xλn or ··· tˆλ1,...,λn =p ˆλ1 pˆλn . Note that··· the commutators (4.86) of the generators among each other imply that they, themselves, are tensor operators of rank 2. It is worth observing that the differential operators Lˆµν can also be expressed in terms of the 4 4 matrix generators (4.65) as × i i Lˆµν = (Lµν ) xλpˆκ = xT Lµν pˆ = ipˆT Lµν x. (4.101) h¯ λκ h¯ − In this form, the operators Lˆµν follow the same algebraic construction rules as the † operatorsa ˆ Miaˆ in Section 2.5. There we showed that sandwich constructions be- tween creation and annihilation operatorsa ˆ†Lµν aˆ carry the commutation rules be- tween the matrices Lµν into a larger Hilbert space without changing their algebra. µ † Since ipˆµ and x commute in the same way asa ˆ anda ˆ, the same argument applies to the− sandwich construction (4.101), where the matrix generators Lµν of Eq. (4.65) between ipˆµ and xν produces an infinite-dimensional representation in terms of − differential operators acting on the Hilbert space of square-integrable functions. The commutation relations between the group generators (4.97) on the one side and vector and tensor operators on the other side can be used to calculate the effect of finite group transformations upon these operators. They can also be used to express the transformation property (4.50) in another way. A finite Lorentz transformation of a scalar field is obtained by exponentiating the generators just as in the 4 4 -representations (4.69) and (4.72). Thus we define the differential × operator representation of finite group elements (4.69) as

−i 1 ω Lˆµν Dˆ(Λ) e 2 µν . (4.102) ≡ Since the generators Lˆµν commute in the same way among each other as the 4 4 -matrix generators (4.65), the operators Dˆ(Λ) obey the same group multiplication× rules as the 4 4 -matrices Λ. This follows directly from the expansion (4.75) of the product in terms× of commutators. Let us apply such a finite transformation to the vector xµ and form Dˆ(Λ)xλDˆ −1(Λ). (4.103) We do this separately for rotations and Lorentz transformations, first for rotations. An arbitrary three-vector (x1, x2, x3) is rotated around the 3-axis by the operator ˆ Dˆ(R (ϕ)) = e−iϕL3 with Lˆ = i(x1∂ x2∂ ) by the operation 3 3 − 2 − 1 i −1 −iϕLˆ3 i iϕLˆ3 Dˆ(R3(ϕ))x Dˆ (R3(ϕ)) = e x e . (4.104) 254 4 Free Relativistic Particles and Fields

The right-hand side is evaluated with the help of Lie’s expansion formula:

i2 e−iAˆ Bˆ eiA =1 i[A,ˆ Bˆ]+ [A,ˆ [A,ˆ Bˆ]] + .... (4.105) − 2!

3 Since Lˆ3 commutes with x , this component is unchanged by the operation (4.113):

3 −1 −iϕLˆ3 3 iϕLˆ3 3 Dˆ(R3(ϕ))x Dˆ (R3(ϕ)) = e x e = x . (4.106)

For x1 and x2, the Lie expansion of (4.103) contains the commutators

i[L , x1]= x2, i[L , x2]= x1. (4.107) − 3 − 3 − Thus, the first-order expansion term transforms the two-dimensional vector (x1, x2) into (x2, x1). The second-order terms are obtained by commuting the operator − 1 2 iLˆ3 with these components, yielding (x , x ). To third-order, this is again trans- −formed into (x2, x1), and so on. Obviously,− all even orders reproduce the initial two-dimensional− vector− (x1, x2) with an alternating sign, while all odd powers are proportional to (x2, x1): − ˆ ˆ 1 1 e−iϕL3 (x1, x2)eiϕL3 = 1 ϕ2 + ϕ4 + ... (x1, x2)  − 2! 4!  1 1 + ϕ ϕ3 + ϕ5 + ... (x2, x1). (4.108)  − 3! 5!  − The even and odd powers can be summed up, and we obtain

ˆ ˆ e−iϕL3 (x1, x2)eiϕL3 = cos ϕ (x1, x2) + sin ϕ (x2, x1). (4.109) − Together with (4.106), the right-hand side corresponds precisely to the inverse of the rotation (4.52). Thus

ˆ ˆ i ˆ i ˆ −1 −iϕL3 i iϕL3 iϕL3 j −1 i j D(R3(ϕ))x D (R3(ϕ)) = e x e = e jx = R3 (ϕ) jx . (4.110)   By performing successively rotations around the three axes we can generate, in this way, any rotation:

ˆ ˆ i

³ ³

ˆ i ˆ −1 −i³·L i i ·L i ·L j −1 i j ³ D R³ ϕ x D R ϕ e x e e x R ϕ x , ( ( )) ( ( )) = = j = ³ ( ) j (4.111)   this being the finite rotation form of the commutation relation for the vector operator xi: [Lˆi, xk]= xj(Li)jk. (4.112) This holds for any vector operatorv ˆi instead of xi [recall again the definition in (2.113)]. 0 The time component x is obviously unchanged by a rotation since Lˆ3 commutes with x0. 4.2 Differential Operators for Lorentz Transformations 255

A similar calculation may be done for pure Lorentz transformations. Here we −iζMˆ 3 03 first consider a boost in the 3-direction B3(ζ) = e generated by Mˆ 3 = Lˆ = 0 3 i(x ∂3 + x ∂0) [recall (4.68), (4.72), and (4.92)]. Note the positive relative sign of − 03 i the two terms in the generator Lˆ that are caused by the fact that ∂i = ∂ , in 0 − contrast to ∂0 = ∂ . Thus we form

i −1 −iζMˆ 3 i iζMˆ 3 Dˆ(B3(ζ))x Dˆ (B3(ζ)) = e x e . (4.113)

The Lie expansion of the right-hand side involves the commutators

i[M , x0]= x3, i[M , x3]= x0, i[M , x1]=0, i[M , x2]=0. (4.114) − 3 − − 3 − − 3 − 3 Thus the two-vector (x1, x2) is unchanged, while the two-vector (x0, x3) is trans- formed into (x3, x0). In the second expansion term, the latter becomes (x0, x3), − and so on, yielding

ˆ ˆ 1 1 e−iζM3 (x0, x3)eiζM3 = 1+ ζ2 + ζ4 + ... (x0, x3)  2! 4!  1 1 ζ + ζ3 + ζ5 + ... (x3, x0), (4.115) −  3! 5!  which can be summed up as

ˆ ˆ e−iζM3 (x0, x3)eiζM3 = cosh ζ (x0, x3) sinh ζ (x3, x0). (4.116) − Together with the invariance of (x1, x2), this corresponds precisely to the inverse of the boost transformation (4.14). Hence we have

ˆ ˆ λ ˆ λ ˆ −1 −iζM3 λ iζM3 iζM3 κ −1 λ κ D(B3(ζ))x D (B3(ζ)) = e x e = e κx = B3 (ζ) κx . (4.117)   By performing successively rotations and boosts in all directions, we find for the entire Lorentz group:

1 µν 1 µν 1 µν ′ ′ λ −1 −i ωµν Lˆ λ i ωµν Lˆ i ωµν L λ λ −1 λ λ Dˆ(Λ)x Dˆ (Λ) = e 2 x e 2 =(e 2 ) λ′ x = (Λ ) λ′ x , (4.118) where ωµν are the parameters (4.70) and (4.71). In the last term on the right-hand side we have expressed the 4 4 -matrix Λ as an exponential of its generators as well, to emphasize the one-to-one× correspondence between the generators Lµν and their differential-operator representation Lˆµν . At first it may seem surprising that the group transformations appearing as a left-hand factor of the two sides of these equations are inverse to each other. However, we may easily convince ourselves that this is necessary to guarantee the correct group multiplication law. Indeed, if we perform two transformations after each other they appear in opposite order on the right- and left-hand sides:

λ −1 λ −1 −1 Dˆ(Λ2Λ1)x Dˆ (Λ2Λ1)= Dˆ(Λ2)Dˆ(Λ1)x Dˆ (Λ1)Dˆ (Λ2) ′ ′ ′′ ′ −1 λ ′ ˆ λ ˆ −1 −1 λ ′ −1 λ ′′ λ −1 λ ′ λ = (Λ1 ) λ D(Λ2)x D (Λ2)=(Λ1 ) λ (Λ2 ) λ x = [(Λ2Λ1) ] λ x . (4.119) 256 4 Free Relativistic Particles and Fields

If the right-hand side of (4.118) would contain Λ instead of Λ−1, the order of the factors in Λ2Λ1 on the right-hand side of (4.119) would be opposite to the order in Dˆ(Λ2Λ1) on the left-hand side. A straightforward extension of the operation (4.118) yields the transformation law for a tensor tˆλ1,...,λn = xλ1 xλn : ··· λ ,...,λ −1 −i 1 ω Lˆµν λ ,...,λ i 1 ω Lˆµν Dˆ(Λ)tˆ 1 n Dˆ (Λ) = e 2 µν tˆ 1 n e 2 µν −1 λ −1 λ λ′ ,...,λ′ = (Λ ) 1 ′ (Λ ) n ′ tˆ 1 n λ1 λn 1 µν ··· 1 µν ′ ′ i ωµν L λ1 i ωµν L λn λ ,...,λn = (e 2 ) λ′ (e 2 ) λ′ tˆ 1 . (4.120) 1 ··· n This follows directly by inserting an auxiliary unit factor 1 = Dˆ(Λ)Dˆ −1(Λ) = −i 1 ω Lˆµν i 1 ω Lˆµν λ e 2 µν e 2 µν between each pair of neighboring coordinates x i in the product tˆλ1,...,λn = xλ1 xλn , and by performing the operation (4.120) on each of them. The last term in (4.120)··· can also be written as

i 1 ω (Lµν ×1×1···×1 + ... + 1×Lµν ×1···×1) λ1...λn λ′ ...λ′ e 2 µν ′ ′ t 1 n . λ1...λn h i Since the commutation relations (4.100) determine the result completely, the trans- formation formula (4.120) is true for any tensor operator tˆλ1,...,λn , and not only for those composed from a product of vectors xλi . For a field φ(x) which can be expanded into a power series, the transformation law (4.120) generalizes immediately to

−1 −i 1 ω Lˆµν i 1 ω Lˆµν −1 i 1 ω Lµν Dˆ(Λ)φ(x)Dˆ (Λ) = e 2 µν φ(x)e 2 µν = φ(Λ x)= φ(e 2 µν x). (4.121)

The finite transformation law (4.51) of a scalar field can therefore be expressed with −i 1 ω Lˆµν the help of the differential operator Dˆ(Λ) = e 2 µν as

Λ ′ −1 −1 −i 1 ω Lˆµν i 1 ω Lˆµν φ(x) φ (x)= φ(Λ x)= Dˆ(Λ)φ(x)Dˆ (Λ) e 2 µν φ(x)e 2 µν .(4.122) −−−→ Λ ≡ The last factor on the right-hand side can, of course, be omitted if there are no x-dependent functions behind it. If a particle has spin degrees of freedom, its field transforms differently from (4.122). Then the wave function has several components to account for the spin orientations. The transformation law must be such that the spin orientation in space remains the same at the same space point. This implies that the field components, which specify the size and orientation with respect to the different coordinate axes, will have to be transformed by certain matrices. How this is done for relativistic fields has first been understood for electromagnetic and gravitational fields which exhibit vector and tensor transformation properties, respectively. These will be recalled in Sections 4.6 and 4.9, before generalizing them in Sections 4.10 and 4.13 to spin 1/2, and in Section 4.18 to spins of arbitrary . 4.3 Space Inversion and Time Reversal 257

4.3 Space Inversion and Time Reversal

In addition to the continuous Lorentz transformations, there are also two important µ discrete transformations which leave scalar products pµx invariant. First there is the space inversion, also called space reflection or transformation,

1 1 P =   , (4.123) − 1    −   1   −  which reverses the direction of the spatial vectors, x x. Note that a space inversion differs from a mirror reflection by a rotation.→ The − space inversion maps the generators Li, Mi of the Lorentz group into parity transformed generators

P P L LP P L P −1 = L , M M P P M P −1 = M . (4.124) i −−−→ i ≡ i i i −−−→ i ≡ i − i

This behavior is obvious in the tensor form of the generators (Li, Mi) = ( 1 ǫ Lik, L0i). Each spacelike index gives rise to a factor 1. The transformation 2 ijk − preserves the commutation rules (4.77)–(4.78):

P P P [Li , Lj ] = iǫijkLk , (4.125) P P P [Li , Mj ] = iǫijkMk , (4.126) [M P , M P ] = iǫ LP . (4.127) i j − ijk k In general, a mapping of the generators into linear combinations of generators that have the same commutation rules is called an automorphism of the Lie algebra. Second, there is the time inversion or time reversal transformation

1 − 1 T =   , (4.128) 1    1      which changes the sign of x0. When applied to the generators of the defining re- presentation (Li, Mi), the time reversal transformation produces the same auto- morphism of the Lie algebra as the parity transformation (4.124). This, however, is a special feature of the reality of the Lorentz transformation matrices Λ which makes the 4 4 -matrices of the generators (4.54)–(4.56) and (4.60)–(4.62) purely × imaginary. Physically, a process is invariant under time reversal if we are unable to judge whether a movie of the process runs forward or backward. Running it backwards amounts to changing momentum and angular momentum Li. Since momentum is generated by boost transformations, time reversal must change the direction of the generators Mi. For Hermitian matrices it is only necessary to change the eigenvalues, 258 4 Free Relativistic Particles and Fields

∗ such that we can also require that Li goes into Li . In fact, there is a natural automorphism of the Lie algebra (4.77)–(4.78) in which− we simply take the complex conjugate of the commutation rules, bringing them to

[ L∗, L∗] = iǫ ( L∗), (4.129) − i − j ijk − k [ L∗, M ∗] = iǫ ( M ∗), (4.130) − i − j ijk − k [ M ∗, M ∗] = iǫ ( L∗ ). (4.131) − i − j − ijk − k As we shall see in detail when discussing the time reversal properties of the various fields, this automorphism has precisely the desired observational consequences which we would like to associate with a time reversal transformation. Explicitly, time reversal transforms the generators as follows:

T T L LT T L T −1 = L∗, M M T T M T −1 = M ∗. (4.132) i −−−→ i ≡ i − i i −−−→ i ≡ i − i When the operations P and T are incorporated into the special Lorentz group SO(1,3), one speaks of the full Lorentz group. Note that the determinant of (4.123) and (4.128) is negative, so that the special Lorentz group SO(1, 3) no longer deserves the letter S in its name, and is called O(1, 3).

4.4 Free Relativistic Scalar Fields

The question now arises as to how the nonrelativistic free-field

2 3 ∗ h¯ 2 = dtd x ψ (x, t) ih∂¯ t + ∂x ψ(x, t), (4.133) A Z " 2M # introduced in (2.202), has to be modified in order to permit a quantum mechanical description of arbitrary relativistic n-particle states. According to the definition in (2.161), this is a local action. A field theory based on a local action is called local field theory. All field theories which explain successfully the properties of elementary particles have so far turned out to be local. The locality property seems to be extremely fundamental. Many fundamental have historically been discovered as action-at-a- forces. In the total field action, these correspond to nonlocal terms. Eventually, however, they have been shown to be the result of local actions that involve extra fields mediating the interaction. The initial idea for doing this came from Maxwell in the context of electromagnetism. Remember that the original Coulomb interaction was described by an action-at-a-distance which corresponds to a bilocal term in the action (see Section 2.8). Maxwell discovered that it can be reexpressed in terms of a local interaction by introducing an extra field called a potential field. In the relativistic generalization of the theory, this can further be viewed as a a zeroth component of a four-dimensional vector potential. In the quantized version of the theory, the vector potential is associated with particles of light called photons (see Chapter 12). The same holds for gravitational forces 4.4 Free Relativistic Scalar Fields 259 and their quanta, which represent particles called gravitons. In the description of nuclear forces, the same locality principle has led Yukawa to the discovery of the fundamental particle called π-meson (see Section 24.3, in particular Eqs. (24.31) and ′ −µ|x−x′| ′ (24.32) for the two-body potential V2(x x ) e / x x , and the associated wave equation obeyed by the π-field). − ∝ | − | In order to accommodate the kinematic features discussed in the last section we shall require the action to be invariant under Lorentz transformations of coordinates (4.39) and fields [for example (4.50)]. Hence, space and time derivatives have to appear on equal footing, i.e., both must appear linearly or quadratically if we want to maintain the usual principle of classical mechanics in which all differential equations are of second order in time. Depending on the possible internal spin degrees of freedom there are different ways of making the action relativistic. These will now be discussed separately. Consider first a field associated with a relativistic point particle which carries no spin degree of freedom, thus avoiding a nontrivial behavior under space rotations. Such a field was introduced in Eq. (4.50) as a scalar field and denoted by φ(x). As in the nonrelativistic case, the action of this field must contain the square of 2 the spatial derivatives ∂i to guarantee rotational invariance. Since there must be Lorentz symmetry between space and time derivatives, we are led to a classical local action

= dx0L = dx0d3xφ∗(x, t) c h¯2(∂2 ∂2) c φ(x, t) A − 1 0 − x − 2 Z Z h i = d4xφ∗(x) c h¯2∂µ∂ c φ(x), (4.134) − 1 µ − 2 Z h i 2 where c1h¯ ,c2 are two arbitrary real constants. It is easy to see that this action is indeed Lorentz invariant: Under the transfor- mation (4.39), the four-volume element does not change:

dx0d3x d4x d4x′ = d4x. (4.135) ≡ → This follows directly from Eq. (4.43) which implies that the determinant of the matrices Λ has the values det Λ = 1. If we therefore take the action in the new ± frame = d4x′ φ∗′(x′) c h¯2∂′µ∂′ c φ′(x′), (4.136) A − 1 µ − 2 Z h i we can use (4.50) and (4.134) to rewrite

= d4xφ∗(x) c h¯2∂′µ∂′ c φ(x). (4.137) A − 1 µ − 2 Z h i But since

′ ν ′µ µ ν ∂µ =Λµ ∂ν , ∂ =Λ ν∂ (4.138) with Λ ν g gνκΛλ , (4.139) µ ≡ µλ κ 260 4 Free Relativistic Particles and Fields we see that

∂′2 = ∂2, (4.140) and the transformed action coincides with the original action (4.134). As in (2.161), we call the integrand of the action a Lagrangian density:

(x)= φ∗(x) c h¯2∂µ∂ c φ(x). (4.141) L − 1 µ − 2 h i Then the invariance of the action under Lorentz transformations is a direct conse- quence of the Lagrangian density being a scalar field satisfying the transformation law (4.50): ′(x′)= (x). (4.142) L L This follows directly from the invariance (4.140) and φ′(x′)= φ(x). The free-field equation of motion is derived by varying the action (4.134) with respect to the fields φ(x),φ∗(x) independently. The independence of these variables is expressed by the functional differentiation rules δφ(x) δφ∗(x) = δ(4)(x x′), = δ(4)(x x′) (4.143) δφ(x′) − δφ∗(x′) − δφ(x) δφ∗(x) = 0, =0. (4.144) δφ∗(x′) δφ(x′)

Applying these rules to (4.134) we obtain directly

δ 4 ′ (4) ′ 2 ′2 ′ ∗A = d x δ (x x)( c1h¯ ∂ + c2)φ(x ) δφ (x) Z − − = ( c h¯2∂2 + c )φ(x)=0. (4.145) − 1 2 Similarly,

δ 4 ′ ∗ ′ 2 ′2 (4) ′ A = d x φ (x )( c1h¯ ∂ + c2)δ (x x) δφ(x) Z − − ← = φ∗(x)( c h¯2 ∂2 +c )=0, (4.146) − 1 2 where the arrow on top of the last derivative indicates that it acts on the field to the left. The second equation is just the complex conjugate of the previous one. The field equations (4.145) and (4.146) can be derived directly from the La- grangian density (4.141) by forming ordinary partial derivatives of with respect to all fields and their derivatives. Indeed, a functional derivative ofL a local action can be expanded in terms of derivatives of the Lagrangian density according to the general rule δ ∂ (x) ∂ (x) ∂ (x) A = L ∂µ L + ∂µ∂ν L + ..., (4.147) δφ(x) ∂φ(x) − ∂ [∂µφ(x)] ∂ [∂µ∂ν φ(x)] 4.4 Free Relativistic Scalar Fields 261 and a similar expansion holds for the derivative with respect to φ∗(x). These ex- pansions follow directly from the defining relations (4.143). Applying them to the Lagrangian density (4.141), the field equation for φ(x) is particularly simple:

δ ∂ (x) A = L =( c h¯2∂2 + c )φ(x)=0. (4.148) δφ∗(x) ∂φ∗(x) − 1 2

For φ∗(x), on the other hand, all derivatives written out in (4.147) have to be evaluated to obtain

δ ∂ (x) ∂ (x) ∂ (x) 2 2 ∗ A = L ∂µ L + ∂µ∂ν L =( c1h¯ ∂ + c2)φ (x)=0. δφ(x) ∂φ(x) − ∂ [∂µφ(x)] ∂ [∂µ∂ν φ(x)] − (4.149) The equation

∂ (x) ∂ (x) ∂ (x) L ∂µ L + ∂µ∂ν L + ... = 0 (4.150) ∂φ(x) − ∂∂µφ(x) ∂∂µ∂ν φ(x) is the Euler-Lagrange equation of a general local field theory. This expression is invariant under partial integrations within the local action = d4x (x). Take for A L example a Lagrangian density which is equivalent to (4.141) by aR partial integration in the action (4.137):

= c h¯2∂φ∗(x)∂φ(x) c φ∗(x)φ(x). (4.151) L 1 − 2 If this is inserted into (4.150), it produces once more the same field equations. The field equations (4.148) and (4.149) are solved by the quantum mechanical plane waves (4.36) and (4.37) of positive and negative energies, respectively:

−ipx/¯h ∗ ipx/¯h fp(x)= e , f (x)= e . (4.152) N p N These form a complete set of plane-wave solutions. The field equations (4.148) and (4.149) require the four-momenta to satisfy the condition c pµp c =0. (4.153) 1 µ − 2 This has precisely the form of the mass shell relation (29.25) if we choose c 2 = M 2c2. (4.154) c1

A positive sign of c1 is necessary for the field fluctuations to be stable. The size can be brought to unity by a multiplicative renormalization of the field. This makes the field normalization different from the nonrelativistic one in the action (4.133). After this, the mass shell condition fixes the free-field action to the standard form = d4xφ∗(x) h¯2∂µ∂ M 2c2 φ(x). (4.155) A − µ − Z h i 262 4 Free Relativistic Particles and Fields

The nonrelativistic limit of the action (4.155) is obtained by removing, from the positive frequency part of the field φ(x), a fast trivial oscillating factor corresponding to the rest energy Mc2, replacing

2 1 φ(x) e−iMc t/¯h ψ(x, t). (4.156) → √2M

For a plane wave fp(x) of Eq. (4.152), the field ψ(x, t) becomes ψp(x, t) = √2M e−i(p0c−Mc2)t/¯heipx/¯h. In the limit of large c, the first exponential becomes −ip2t/N2M e , such that the field ψp(x, t) coincides with the nonrelativistic plane wave (2.212) which extremizes the nonrelativistic action (4.133). The negative-frequency ∗ plane wave fp(x) in (4.152), on the other hand, does not contribute in this limit since it is equal to √2M ei(p0c+Mc2)t/¯heipx/¯h. This contains a temporal prefactor 2 N e2iMc t/¯h that oscillates with infinite frequency for c , and is therefore equivalent to zero by the Riemann-Lebesgue Lemma [13]. →∞ The appearance of the constantsh ¯ and c in all future formulas can be avoided if we from now on with new fundamental units l0, m0, t0, and E0 different from the ordinary CGS units. They are chosen to giveh ¯ and c the value 1. Expressed in terms of the conventional length, time, mass, and energy, these new natural units are given by h¯ h¯ l = , t = , (4.157) 0 Mc 0 Mc2 2 m0 = M, E0 = Mc . (4.158)

If, for example, the particle is a proton with mass mp, these units are

−11 l0 = 2.103138 10 cm (4.159) = Compton× wavelength of proton, −22 t0 = l0/c =7.0153141 10 sec (4.160) = time it takes light× to cross the Compton wavelength, m = m =1.6726141 10−24g, (4.161) 0 p × E0 = 938.2592 MeV. (4.162)

For any other mass, they can easily be rescaled. With these natural units we can drop c andh ¯ in all formulas and write the action simply as = d4xφ∗(x)( ∂2 M 2)φ(x). (4.163) A Z − − The Lagrange density of the complex scalar field may be taken either as

(x)= φ∗(x)( ∂2 M 2)φ(x), (4.164) L − − or (x)= ∂ φ∗(x)∂µφ(x) M 2φ∗(x)φ(x), (4.165) L µ − 4.4 Free Relativistic Scalar Fields 263 with an obvious modification for real fields φ(x). The surface term, by which the associated actions differ from each other after a partial integration, does not change the Euler-Lagrange equation (4.150). Actually, since we are dealing with relativistic particles there is no fundamental reason to assume φ(x) to be a complex field. In the nonrelativistic theory this was necessary in order to construct a term linear in the time derivative:

∗ dtφ i∂tφ. (4.166) Z For a real field φ(x) this would have been a pure surface term, that does not influence the dynamics of the system. For second-order time derivatives, as in (4.163), this is no longer necessary. Thus we shall also study the real scalar field with an action 1 = d4xφ(x)( ∂2 M 2)φ(x). (4.167) A 2 Z − − In this case it is customary to use a prefactor 1/2 to normalize the field. Here the Lagrange density may be taken either as 1 (x)= φ∗(x)( ∂2 M 2)φ(x), (4.168) L 2 − − or as 1 (x)= [∂φ(x)]2 M 2φ2(x) . (4.169) L 2{ − } For either field we obtain the Klein-Gordon equation

( ∂2 M 2)φ(x)=0. (4.170) − − For a complex field, there exists an important local conservation law, which generalizes Eq. (1.109) of the Schr¨odinger theory to relativistic fields. We define the four-vector of the probability :

↔ ∗ jµ(x)= iφ ∂µ φ, (4.171) which describes the probability flow of the Klein-Gordon particle. It is easy to verify that, due to (4.170), this satisfies the current conservation law

µ ∂µj (x)=0. (4.172)

This conservation law will permit us, in Chapter 17, to couple electromagnetism to the field and identify jµ(x) as the electromagnetic current (if we choose natural units in which the electric e is equal to unity). The deeper reason for the existence of a conserved current will be understood in Subsection 8.11.1, where we shall see that it is intimately connected with an 264 4 Free Relativistic Particles and Fields invariance of the action (4.163) of a free complex scalar field under arbitrary changes of the phase of the field φ(x) e−iαφ(x). (4.173) → It is this invariance which gives rise to a conserved current density [see (8.270), also (17.68)]. The zeroth component of jµ(x),

j0(x)= cρ(x) (4.174) describes the particle density ρ(x). The spatial integral over ρ(x):

1 Q(t)= d3x j0(x). (4.175) c Z is the total charge in natural units: Because of the local conservation law (4.172), the charge does not depend on time. This is seen by rewriting

3 0 3 µ 3 i 3 i Q˙ (t)= d x ∂0j (x)= d x ∂µj (x) d x ∂ij (x)= d x ∂ij (x) (4.176) Z Z − Z − Z and applying, to the right-hand side, Gauss’s theorem as in (1.110), assuming that currents vanish at spatial infinity. By removing, from the positive-frequency solutions of the Klein-Gordon field φ(x), the fast oscillation as in (4.156), we can take the nonrelativistic limit and find that the nonrelativistic limit of the spatial part of the current density (4.171) satisfies the local conservation law (1.109) of the Schr¨odinger theory. Since the current conservation law (4.172) is the direct relativistic generalization of the nonrelativistic probability conservation law (1.109), it is suggestive to define the matrix elements of the charge Q(t) as the scalar product between relativistic wave functions such as the plane waves (4.152). For states of momenta p and p′ we define the scalar product

↔ ′ 3 ∗ (fp , fp)t d x fp′ (x, t)i ∂0 fp(x, t). (4.177) ≡ Z It is formed as a spatial integral at any fixed time, that actually does not need to be recorded in the notation, since the result does not depend on t due to charge conservation. Analogous scalar products exist between positive- and negative-frequency so- ∗ ∗ lutions fp(x) and fp(x), and between two negative-frequency solutions fp(x) of ∗ different momenta. Both sets of wave functions fp(x) and fp(x) are needed to span the space of all solutions of the Klein-Gordon equation. Within the scalar product (4.177), we choose to normalize the plane wave functions so that they satisfy the orthogonality relations

∗ ∗ ∗ (fp′ , fp) = δp′ p, (f ′ , f ) = δp′ p, (f ′ , fp) =0. (4.178) t , p p t − , p t 4.5 Other Symmetries of Scalar Action 265

The spatial integrals ensure that the spatial momenta are equal or opposite to each other. Then the energies p0 are equal to each other, so that the time derivative in ∗ the scalar products produces either zero [between fp(t) and f−p(t)], or a positive value between equal wave functions. In a finite volume V , these are the norms of the ∗ wave functions fp(x, t) or fp(x, t). They coincide with the matrix elements of the charge (4.175), if the appropriate plane wave is inserted for the field φ(x) in (4.171). ∗ The charge of the plane waves with negative frequency fp(x) is negative: ↔ ∗ ∗ 3 ′ ∗ (fp, fp)t d x fp (x, t)i ∂0 fp(x, t) < 0, (4.179) ≡ Z so that the set of all scalar products is not positive-definite. Historically, this was an obstacle for the Klein-Gordon theory to represent a direct generalization of the Schr¨odinger theory to relativistic particles; rightfully so, as we shall see in Chapter 7. In a finite total spatial volume V , the properly normalized wave functions (4.152) are explicitly

1 −ipx ∗ 1 ipx fp(x, t)= e , f (x, t)= e , (4.180) √2Vp0 p √2Vp0

0 0 2 2 where p is the particle energy p = ωp = √p + M . The norms are 1. ± In an infinite volume, a convenient normalization is

−ipx ∗ ipx fp(x, t)= e , fp(x, t)= e , (4.181) and the orthonormality relations become 0 -(3) ′ (fp′ , fp) = 2p δ (p p), t − ∗ ∗ 0 (3) ′ (f ′ , f ) = 2p δ- (p p), p p t − − ∗ (fp′ , fp)t = 0, (4.182) where δ-(3)(p′ p)=(2π)3δ(3)(p′ p), as defined in Eq. (1.196). The convenience in − − having a factor 2p0 accompany the δ--function is that this combination has pleasant transformation properties under the Lorentz group. It yields unity when integrated over the Lorentz-invariant volume element in momentum space: d3p 3 0 . (4.183) Z (2π) 2p The Lorentz invariance is obvious by rewriting this as

4 d p 0 2 2 3 Θ(p )δ(p M ). (4.184) Z (2π) − 4.5 Other Symmetries of Scalar Action

The actions (4.163) and (4.167) of a real or complex scalar field are invariant under more than just the Lorentz group. 266 4 Free Relativistic Particles and Fields

4.5.1 Translations of Scalar First, the actions are invariant under space as well as time translations of the coor- dinate system: x′µ = xµ + aµ. (4.185) Recall that under Lorentz transformations, a scalar field at the same spacetime point remains unchanged by the change of coordinates: Λ φ(x) φ′ (x)= φ(Λ−1x). (4.186) −−−→ Λ The same is true for translations: a φ(x) φ′ (x)= φ(x a). (4.187) −−−→ a − Inserting this into the Lagrangian density (4.164), we see that it transforms like a scalar field: ′(x)= (x a). (4.188) L L − Together with the trivial translational invariance of the volume integral, the action is indeed invariant. The combinations of Lorentz transformations and translations, ′µ µ ν µ x =Λ νx + a , (4.189) form a group called the inhomogeneous Lorentz group or Poincar´egroup. Under it, the scalar field transforms as φ(x) φ′(x)= φ(Λ−1(x a)). (4.190) −−−→ − Thus, a free scalar field theory is not only Lorentz-invariant but also Poincar´e- invariant. This holds also for real and complex scalar fields φ(x). Translations can be generated by a differential operator in just the same way as Lorentz transformations in Eq. (4.122). Obviously we can write the translation (4.187) as µ φ(x) φ′ (x)= φ(x a)= Dˆ(a)φ(x) eia pˆµ/¯hφ(x), (4.191) −−−→ a − ≡ wherep ˆµ = ih∂¯ µ is the differential operator of momentum (4.34). This is proved by applying Lie’s expansion formula (4.105) to the coordinates xµ: µ µ µ µ eia pˆµ/¯hxλeia pˆµ/¯h = e−a ∂µ xλea ∂µ = xλ aλ, (4.192) − Poincar´etransformations are then obtained from operations µ 1 ˆµν ′ −1 ˆ ˆ ia pˆµ/¯h −i 2 ωµν L φ(x) φa(x)= φ(Λ (x a)) = D(a)D(Λ)φ(x) e e φ(x), −−−→ − ≡ (4.193) with the parameters ωµν specified in (4.70) and (4.71). This follows from the beha- vior of the coordinate vector: iaµpˆ /¯h −i 1 ω Lˆµν λ i 1 ω Lˆµν −iaµpˆ /¯h λ −1 µ ν e µ e 2 µν x e 2 µν e µ x = (Λ ) (x a) , (4.194) ν − thus extending (4.118) to the Poincar´egroup. The last equation states in a global way the vector properties of xµ under Poincar´etransformations. 4.5 Other Symmetries of Scalar Action 267

4.5.2 Space Inversion of Second, the scalar actions (4.163) and (4.167) are invariant under the operation of space inversion [see (4.123)], under which the coordinates go into

P x x′ =x ˜ (x0, xi), (4.195) −−−→ P ≡ − whereas the scalar field is transformed as follows:

P φ(x) φ′ (x)= φ(˜x). (4.196) −−−→ P Note that this transformation behavior is not the only possible one. Since the parity operation is not related continuously to the identity, there is no reason why the field φ(x) should transform into itself as it does for the continuous group of Lorentz transformations. The parity operation forms, together with the unit element, a group called the group of space reflections. The group multiplication table reads as follows: 1 P 1 1 P (4.197) P P 1 . The only requirement to be satisfied by the field is to be consistent with the group multiplication law in the table. This is assured if the successive application of two parity operations, which result in the identity operation, leads back to the original field (cyclicity of order 2). It is therefore possible to choose any transformation law

P φ(x) φ′ (x)= η φ(˜x), (4.198) −−−→ P P 2 as long as ηP satisfies ηP = 1. This allows for two solutions, the above trivial one in (4.198) with η = 1, and the alternating one with η = 1, i.e., P P − η = 1. (4.199) P ± Thus the scalar field could also pick up a negative sign upon space reflection. If the interactions to which a particle is subjected are invariant under the parity operation, the value of ηP is a characteristic property of the particle. It is called the intrinsic parity of the particle. States with positive or negative intrinsic parities are familiar in quantum mechanics where they appear as bound states with even or odd orbital angular momentum, respectively. Only a particle with ηP = 1 is called a proper scalar particle, while η = 1 is called a particle. P − The most important fundamental particles of odd parity are the π-mesons which are the source of the long-range part of nuclear interactions. The pseudoscalar nature is most simply seen in the decay of π0 into two photons, which is the main reason for the finite lifetime τ = (8.4 0.6) 10−17sec of this particle (branching ratio of two-photon with respect to all decay± channels× is 98.798%). In the rest frame of the pion, the two final photons emerge in opposite directions and are polarized 268 4 Free Relativistic Particles and Fields parallel to each other. Under a space inversion, the two-photon state is transformed into itself, but with reversed polarization directions. This corresponds to a negative parity. The negative parity of a charged pion π−, whose lifetime is much longer [τ = (2.6030 0.0024) 10−8sec], can be deduced from the existence of the absorption of ± × a π−-meson at rest by a deuteron. The deuteron is a bound state of a neutron and a proton in an s-wave with parallel spins, so that the total angular momentum of a deuteron at rest is J = 1. For an s-wave, the parity of the orbital wave function is positive. An additional pion at rest does not change J. The final state consists of two neutrons flying apart in opposite directions. By the Pauli principle, their wave function has to be antisymmetric. Thus it can only be in spin-singlet states for even orbital angular momenta l, or in spin-triplet states for odd l. Since the final total angular momentum must be J = 1, only the spin-triplet l = 1 state is allowed. This, however, has a negative parity, which can only be caused by the additional π−-meson being a pseudoscalar particle. Note that the intrinsic parities of proton and neutron do not matter in this argument since these particles are present before and after the absorption process. If particles with definite intrinsic parity interact with each other and the interac- tions are invariant under space reflection, the intrinsic parity supplies characteristic selection rules in scattering and decay processes. In quantum mechanics, for exam- ple, decays of atomic states in the dipole approximation have to change the parity of the state since the dipole operator itself has a negative parity.

4.5.3 Time Reversal of Scalar Field As a second extension of the Lorentz invariance of the scalar actions (4.163) or (4.167) we can reverse the sign of the time axis via the time reversal transformation

T x x′ = x,˜ (4.200) −−−→ T − which has the same multiplication table as the parity transformation:

1 T 1 1 T (4.201) T T 1 . The field transforms like

T φ(x) φ′ (x)= η φ(x ), (4.202) −−−→ T T T where again η = 1. (4.203) T ± Note that this transformation law holds for both real and complex fields. The field transformation law (4.202) should not be confused with the corresponding 4.5 Other Symmetries of Scalar Action 269 transformation law of wave functions. In order to clarify this point consider non- relativistic Schr¨odinger theory. There, given a solution ψ(x, t) of the free-particle Schr¨odinger equation 2 h¯ 2 i∂t + ∂x ψ(x, t)=0, (4.204) 2M ! the wave function ψ∗(x, t) represents also a solution with the same energy. The presence of only a single− time derivative necessitates the complex conjugation. A plane-wave solution of momentum p

−iEt/¯h+ipx/¯h ψp(x, t)= e , (4.205) transforms like T ′ ∗ ψp(x, t) ψ (x, t)= η ψ (x, t). (4.206) −−−→ p T T p − This satisfies the Schr¨odinger equation

2 h¯ 2 ′ i∂t + ∂x ψp T (x, t)=0. (4.207) 2M ! From (4.205) we see that the right-hand side of (4.206) is equal to

ηT ψ−p(x, t). (4.208)

The particle momentum in the transformed wave function is reversed, so that the particle appears to run backwards in time. Since the transformation law (4.206) involves complex conjugation, scalar prod- ucts between two arbitrary wave functions ψ (t) ψ (t) = d3x ψ∗(x, t)ψ (x, t) go h 2 | 1 i 2 1 over into their complex conjugates at the negative time: R

T ψ (t) ψ (t) ψ′ (t) ψ′ (t) = ψ ( t) ψ ( t) ∗ = ψ ( t) ψ ( t) . (4.209) h 2 | 1 i −−−→ T h 2 | 1 iT h 2 − | 1 − i h 1 − | 2 − i This property guarantees the preservation of probabilities under this transformation. In general, any transformation which carries all scalar products into their Hermitian conjugates is referred to as antiunitary. Antiunitarity implies that the time reversal operation is necessarily antilinear. A transformation is antilinear if the coefficients of a linear combination of wave functions go over into their complex conjugates. At the level of Schr¨odinger differential operators, antiunitarity produces a sign reversal in the transformation properties of energy and momentum. The defining representation T −1tT = t, T −1xT = x (4.210) − implies that −1 −1 T i∂ T = i∂ , T i∂xT = i∂x. (4.211) t − t The antiunitary representation of this operation is

−1 −1 i∂ = i∂ , i∂x = i∂x. (4.212) T tT t T T − 270 4 Free Relativistic Particles and Fields

It leaves the energy invariant, while reversing the direction of the momentum:

T T E E, p p. (4.213) −−−→ −−−→− Thus particles keep their positive energy but run backwards in time. A unitary representation would have the opposite effect and produce a state which cannot be found in nature. In contrast to the wave functions, the Schr¨odinger field operator, which was the result of second quantization in Chapter 2, transforms under time reversal like

T ψˆ(x, t) ψˆ′ (x, t)= η ψˆ(x, t). (4.214) −−−→ T T − At the operator level, the effects of complex conjugation are brought about by the antilinearity of the time reversal operator in the second-quantized Hilbert space. T In fact, the transformation law (4.206) for the wave functions can be derived from (4.214). In terms of , the transformation (4.214) reads T T ψˆ(x, t) −1ψˆ(x, t) = η ψˆ(x, t). (4.215) −−−→ T T T − Expanding the field operator terms of creation and annihilation operators as in Eq. (2.215), -3 ψˆ(x, t)= d p ψp(x, t)ˆap, (4.216) Z and using the antilinearity of , we obtain T −1 ˆ -3 ∗ −1 -3 iEt/¯h−ipx/¯h −1 ψ(x, t) = d p ψp(x, t) aˆp = d p e aˆp . (4.217) T T Z T T Z T T According to (4.215), the right-hand side has to be equal to

-3 iEt/¯h+ipx/¯h ηT d p e aˆp, (4.218) Z which implies that the annihilation operators transform like

−1 aˆp = η aˆ p. (4.219) T T T − A particle of momentum p goes over into a particle of momentum p, which is the − correct transformation law. Thus, in spite of their contradictory appearance, the two transformation laws (4.206) for wave functions and (4.214) for Schr¨odinger fields, with their antiunitary operator implementation (4.215) after field quantization, are completely consistent with each other. Thus we need no longer to be astonished about the absence of a complex conjugation on the right-hand side of the field transformation law (4.202). Another feature of antiunitarity, which has to be kept in mind, is that the time- reversed field operator does not satisfy the Schr¨odinger equation of the original field

2 h¯ 2 ih∂¯ t + ∂x ψˆ(x, t)=0, (4.220) 2M ! 4.5 Other Symmetries of Scalar Action 271 but the transformed equation 2 ∗ h¯ 2 ˆ′ ih∂¯ t + ∂x ψT (x, t)=0. (4.221) 2M ! This follows immediately by multiplying (4.220) with the operator from the left and passing on to the right-hand side of the differential operator.T The antilin- T earity causes the complex conjugation of the differential operator in (4.222). This is necessary to produce the correct Schr¨odinger equation (4.207) for the time-reversed wave functions. At first sight the field equation appears to be in contradiction with the correspon- dence principle from which one might expect equations for operators to go directly over into those for classical objects in the limit of smallh ¯. Properly, however, this limit must be taken on equations for measurable amplitudes, not the opera- tors themselves, and these do follow the Schr¨odinger equation.2 Take for instance the single-particle amplitude of an arbitrary state Ψ in the Heisenberg picture 0 ψˆ(x, t) Ψ , which satisfies | i h | | i 2 h¯ 2 ih∂¯ t + ∂x 0 ψˆ(x, t) Ψ =0. (4.222) 2M ! h | | i For the time-reversed field operator (4.214), the equation is

2 ∗ h¯ 2 ih∂¯ t + ∂x 0 ψˆ(x, t) Ψ =0. (4.223) 2M ! h | − | i The time-reversed amplitude on the right-hand side is the amplitude for the time reversed state Ψ (since 0 −1 = 0 ): T | i h |T h | 0 ψˆ(x, t) Ψ = 0 −1ψˆ(x, t) Ψ = 0 ψˆ(x, t) Ψ . (4.224) h | − | i h |T T | i h | T | i This obeys the Schr¨odinger equation 2 h¯ 2 ih∂¯ t + ∂x 0 ψˆ(x, t) Ψ , (4.225) 2M ! h | − T | i as it should. The operator implementation of the time reversal transformation will be dis- cussed in detail in Chapter 7 for fields and particles of different spin (see Subsec- tions 7.1.6, 7.4.4, 7.5.4, and 7.7.1). An important observation should, however, be made right here: As a consequence of antiunitarity, the phase factor ηT appearing in the time reversal transformation of a complex field is arbitrary. It cannot be fixed in the usual way by applying the transformations T twice. The antilinearity ∗ will change ηT , that arises in the first transformation, into ηT , so that the combined phase factor is η η∗ = 1. This is fulfilled for any phase factor η = eiγ , not just 1. T T T ± Since the phase factor ηT is arbitrary, it may be chosen arbitrarily, for instance η 1. (4.226) T ≡ 2This is a manifestation of Ehrenfest’s theorem for the semiclassical limit of field equations. 272 4 Free Relativistic Particles and Fields

4.5.4 Charge Conjugation of Scalar Field At the level of a relativistic scalar field φ(x) there is one further discrete symmetry. We can change φ(x) into φ∗(x) without changing the action (4.163). This operation is called charge conjugation denoted by C. Since C2 = 1, a complex field can transform in two possible ways

C φ(x) φ′ (x)= η φ∗(x), (4.227) −−−→ C C where the phase factor can take the values

η = 1. (4.228) C ± For a real field with the action (4.167), we may simply drop the complex conjugation on the right-hand side of (4.227). An explanation is necessary for the name of this operation. In Eq. (4.171) we have seen that there exists a conserved current jµ(x) which can be used to couple the complex scalar field to electromagnetism. If this is done, jµ(x) becomes an electromagnetic current density, and the integral Q(t) = d3x j0(x) is the charge of the field system. Now, under the transformation (4.227)R this electromagnetic current density changes its sign:

C jµ(x) jµ ′(x)= jµ(x). (4.229) −−−→ C − Thus the transformation reverses the charge of the field, and this is why the discrete operation C is called charge conjugation. As we shall discuss later in Chapter 24, electromagnetic and strong interactions are invariant under charge conjugation, implying that the phase factor ηC is a fixed measurable property of the particle, called charge parity. Take for instance the field of the neutral meson π0. As mentioned above, the particle decays with a lifetime of (8.4 0.6) 10−17sec, mostly into two photons. Since these are neutral particles, they have± a charge× parity on their own. Whatever it is (we shall see in Subsection 4.7.2 that it is negative), the two-photon state must have a positive charge parity, and this must consequently be the charge parity of the π0-meson.

4.6 Electromagnetic Field

Electromagnetic fields propagate with light velocity, and their field equations have no mass term [3]. They exist with two polarization degrees of freedom (right and left linear or circular polarizations), and are described by the usual electromagnetic action. Historically, this was the very first example of a relativistic classical field theory, and it could also have served as a guideline for the previous construction of the action (4.167) of a real scalar field φ(x). 4.6 Electromagnetic Field 273

4.6.1 Action and Field Equations

The action may be given in terms of a real auxiliary four-vector potential Aµ(x) from which the physical electric and magnetic fields can be derived as follows:

Ei = (∂0Ai ∂iA0)= ∂ Ai ∂ A0, (4.230) − − − t − i 1 1 Bi = ǫ (∂j Ak ∂kAi)= ǫ (∂ Ak ∂ Aj). (4.231) −2 ijk − 2 ijk j − k It is useful to introduce the so-called four-curl of the vector potential, the tensor

F = ∂ A ∂ A . (4.232) µν µ ν − ν µ Its six components are directly the electric and magnetic field strengths

F = F 0i = ∂0Ai + ∂iA0 = ∂ Ai ∂ A0 = Ei, (4.233) 0i − − − 0 − i F = F ij = ∂iAj ∂jAi = ∂ Aj + ∂ Ai = ǫ Bk, (4.234) ij − − i j − ijk or, in a more conventional notation, 1 E(x) = A˙ (x) ∇A0(x), (4.235) − c − B(x) = ∇ A(x). (4.236) × µν For this reason the tensor Fµν is also called the field tensor. Note that F is related to the fields Bi and Ei in the same way as the generators Lµν of the Lorentz group were related to Li and Mi in Eq. (4.67). The electromagnetic action reads

4 1 4 µν 4 1 2 2 = d x = d x Fµν F = d x (E B ). (4.237) A Z L −4 Z Z 2 −

The four-curl Fµν satisfies a so-called Bianchi identity for any smooth Aµ:

µν ∂µF˜ =0, (4.238) where 1 F˜µν = ǫµνλκF (4.239) 2 λκ is the dual field tensor, with ǫµνλκ being the four-dimensional Levi-Civita tensor with ǫ0123 = 1. Note that F F˜µν =4E B (4.240) µν · is a pseudoscalar. Equation (4.238) can be rewritten as

1 1 ǫµνλκ∂ F = ǫµνλκ(∂ ∂ ∂ ∂ )A (x)=0. (4.241) 2 µ λκ 2 µ λ − λ µ κ 274 4 Free Relativistic Particles and Fields

Multiplying this by another ǫ-tensor and using the tensor identity

µν′λ′κ′ ν′ λ′ κ′ λ′ κ′ ν′ κ′ ν′ λ′ ǫµνλκǫ = δν δλ δκ + δν δλ δκ + δν δλ δκ ′ ′ ′ ′ ′ ′ ′ ′ ′ δ λ δ ν δ κ δ ν δ κ δ λ δ κ δ λ δ ν , (4.242) − ν λ κ − ν λ κ − ν λ κ we obtain the integrability condition of a lemma attributed to H.A. Schwarz [11], according to which the derivatives of an integrable function will always commute:

(∂ ∂ ∂ ∂ )A (x)=0. (4.243) µ ν − ν µ λ The which extremize the action are

δ ∂ (x) 1 µν µA = ∂µ L = ∂µF (x)=0, (4.244) δA (x) − ∂[∂µAν (x)] 2 or more explicitly (gµν∂2 ∂µ∂ν )A (x)=0. (4.245) − ν Separating the equations (4.238) and (4.244) into space and time components they are seen to coincide with the four Maxwell’s equations in empty space:

∂ F˜µν = 0 : ∇ B =0, ∇ E + ∂ B =0, (4.246) µ · × t ∂ F µν = 0 : ∇ E =0, ∇ B ∂ E =0. (4.247) µ · × − t The first equation in (4.246) states that there can be no magnetic monopoles. The second equation is Faraday’s law of induction. The first equation in (4.247) is Coulomb’s law in the absence of charges, the second is Amp`ere’s law in the absence of currents (including, however, Maxwell’s displacement current caused by the time derivative of the electric field). In terms of the vector field Aµ(x), the action reads explicitly

4 1 4 µ ν ν µ = d x (x) = d x [∂ A (x)∂µAν(x) ∂ Aν(x)∂ Aµ(x)] A Z L −2 Z − 1 4 µν 2 µ ν = d x Aµ(x)(g ∂ ∂ ∂ )Aν (x). (4.248) 2 Z − The latter form is very similar to the scalar action (4.167). The first piece is the same as in (4.167) for each of the spatial components A1, A2, A3. The time component A0, however, appears with an opposite sign. A field with this property is called a ghost field. When trying to quantize such a field, the associated particle states turn out to have a negative norm. In a consistent physical theory, such states must never appear in any scattering process. In comparison with the scalar field action (4.134), the second gradient term ν µ ∂ Aν∂ Aµ in the action (4.248) is novel. It exists here as an additional possible Lorentz invariant since Aµ is a vector field under Lorentz transformations. 4.6 Electromagnetic Field 275

It is instructive to insert the individual components A0 and A = (A1, A2, A3) into the action and find 1 = d4x A0(x)( ∇2)A0(x) 2A0(x)∂ iAi(x) A 2 − − 0∇ Z h A(x)(∂2 ∇2)A(x) Ai(x) i iAj(x) . (4.249) − 0 − − ∇ ∇ i This shows that the field component A0 appears without a time derivative. As a consequence, the component A0 remains classical when going over to quantum field theory in Chapter 7. It will be fully determined by the classical field equation.

4.6.2 Gauge Invariance The field tensor (4.232) is invariant under local gauge transformations

A (x) A′ (x)= A (x)+ ∂ Λ(x), (4.250) µ −−−→ µ µ µ where Λ(x) is any smooth field which satisfies the integrability condition

(∂ ∂ ∂ ∂ )Λ(x)=0. (4.251) µ ν − ν µ Gauge invariance implies that one scalar field degree of freedom in Aµ(x) does not contribute to the physically observable electromagnetic fields E(x) and B(x). This degree of freedom can be removed by fixing a gauge. One way to do that is to require the vector potential to satisfy the Lorenz gauge condition, which means that the field Aµ has a vanishing four divergence [1]:

µ ∂µA (x)=0. (4.252)

For a vector field satisfying this condition, the field equations (4.245) decouple end become simply four massless Klein-Gordon equations:

∂2A (x)=0. (4.253) − ν If a vector potential Aµ(x) does not satisfy this condition, one may always per- form a gauge transformation (4.250) to a new field A′µ(x) that has no four diver- gence. We merely have to choose a gauge function Λ(x) solving the differential equation ∂2Λ(x)= ∂ Aµ(x), (4.254) − µ ′µ ′µ and A (x) will satisfy ∂µA (x) = 0. There are infinitely many solutions to equation (4.254). Given one solution Λ(x) which leads to the Lorenz gauge, one can add any solution of the homogeneous Klein- Gordon equation without changing the four-divergence of Aµ(x). The associated gauge transformation

A (x) A (x)+ ∂ Λ′(x), ∂2Λ′(x)=0, (4.255) µ −−−→ µ µ 276 4 Free Relativistic Particles and Fields are called restricted gauge transformations or gauge transformations of the second kind, or on-shell gauge transformations. If a vector potential Aµ(x) in the Lorenz gauge solves the field equations (4.245), the gauge transformations of the second kind can be used to remove the spatial divergence ∇ A(x, t). Under (4.255), the components A0(x, t) and A(x, t) go over into ·

A0(x) A′0(x, t)= A0(x, t)+ ∂ Λ′(x, t), → 0 A(x) A′(x, t)= A(x, t) ∇Λ′(x, t). (4.256) → − Thus, if we choose the gauge function 1 Λ′(x, t)= d3x′ ∇ A(x′, t), (4.257) − 4π x x′ · Z | − | then ∇2Λ′(x, t)= ∇ A(x, t) (4.258) · and the gauge-transformed field A′(x, t) has no spatial divergence, being completely transverse: ∇ A′(x, t)= ∇ [A(x, t) ∇Λ(x, t)]=0. (4.259) · · − This condition ∇ A(x, t) = 0 (4.260) · is known as the Coulomb gauge- or radiation gauge of the vector potential A(x, t).

The solution (4.257) to the differential equation (4.258) is undetermined up to an arbitrary solution Λ′(x) of the homogeneous Poisson equation

∇2Λ′(x, t)=0. (4.261)

Together with the property ∂2Λ′(x, t) = 0 from (4.255), one also has

2 ′ ∂t Λ (x, t)=0. (4.262) This leaves only trivial linear functions Λ′(x, t) of x and t which do not describe propagating waves. Another possible gauge is obtained by removing the zeroth component of the vector potential Aµ(x) to satisfy the field equations (4.245). We form again

A′µ(x)= Aµ(x)+ ∂µΛ(x), (4.263) but now with a gauge function

t ′ ′ Λ(x, t)= dt A0(x, t ). (4.264) − Z Then A′µ(x) will indeed satisfy (4.245), while having the property

A′0(x)=0. (4.265) 4.6 Electromagnetic Field 277

This is called the axial gauge. The solutions of Eqs. (4.264) are determined up to a trivial constant, leaving no more gauge freedom of the second kind, as in (4.255). For free fields, the Coulomb gauge and the axial gauge coincide. This is a consequence of Coulomb’s law ∇ E = 0 in Eq. (4.247). By expressing E(x) explicitly in terms of the spatial and· time-like components of the vector potential,

E(x)= ∂ A(x) ∇A0(x), (4.266) − 0 − Coulomb’s law reads ∇2A0(x, t)= ∇ A˙ (x, t). (4.267) − · This shows that if ∇ A(x) = 0, also A0(x) = 0, and vice versa. The differential equation· (4.267) can be integrated to 1 1 A0(x, t)= d3x′ (∇ A˙ )(x′, t). (4.268) 4π x′ x · Z | − | In an infinite volume with asymptotically vanishing fields there is no freedom of adding to the left-hand side a nontrivial solution of the homogeneous Poisson equa- tion ∇2A0(x, t)=0, (4.269) which in principle would be possible. In the presence of charges, Coulomb’s law will have a source term and read [see Eq. (12.52)] ∇ E(x, t)= ρ(x, t), (4.270) · where ρ(x, t) is the electric . Now the divergence of (4.266) yields the equation ∇2A0(x, t)= ∇ A˙ (x, t) ρ(x, t), (4.271) − · − which is solved by 1 1 A0(x, t)= d3x′ ρ + ∇ A˙ (x′, t). (4.272) 4π x′ x · Z | − |   In contrast to the previous (4.268), the vanishing of ∇ A(x, t) no longer implies · A0(x, t) 0, but determines it to be the instantaneous Coulomb potential around the charge≡ distribution ρ(x′, t): 1 1 A0(x, t)= d3x′ ρ(x′, t). (4.273) 4π x′ x Z | − | Remarkably, there is no retardation. This is an apparent violation of the relativity principle. The contradiction will be resolved due to gauge invariance in Chapters 7 and 12. Note that the fields Aµ can still be modified by adding ∂µΛ and one has the possibility of choosing Λ(x, t) either to satisfy the Coulomb gauge

∇ A(x, t) 0, (4.274) · ≡ 278 4 Free Relativistic Particles and Fields or any other gauge, such as the axial gauge

A0(x, t) 0. (4.275) ≡ Only for free fields the two gauges coincide. The gauge properties of the free-field action (4.248) can be made more explicit with the help of the transverse projection operator [compare (4G.1)]

∂µ∂ν P µν gµν , (4.276) t ≡ − ∂2 which has the property

µν λκ νκ Pt gνλPt = Pt . (4.277)

Then the action (4.248) contains only the transverse part

Aµ P µνA . (4.278) t ≡ t ν of the vector potential:

1 4 µν 2 1 4 µ 2 ν = d x Aµ(x)Pt ∂ Aν(x)= d x gµν At (x)∂ At (x). (4.279) A 2 Z 2 Z A gauge transformation (4.264) changes the field Aµ(x) only by a gradient of a scalar field Λ(x), and this contributes only to the longitudinal part of the vector field

Aµ P µνA , (4.280) l ≡ l ν where [compare (4G.1)]

∂µ∂ν P µν = gµν P µν. (4.281) l ≡ ∂2 − t µ It therefore leaves the transverse part Al invariant due to the orthogonality relation

µν λκ Pt gνλPl =0. (4.282)

µν λκ µ The completeness relation Pl + Pt = 1 ensures that any vector field A (x) can µ µ µ be decomposed into a sum A (x)= At (x)+ Al (x).

4.6.3 Lorentz Transformation Properties of Electromagnetic Fields The Lorentz transformation properties of the electromagnetic fields were understood a long time ago within classical electrodynamics. They are the origin of the famous Lorentz acting on charged particles in motion. The experimentally observed electric and magnetic forces can be derived by going from the laboratory frame with 4.6 Electromagnetic Field 279

fields E, B to the reference frame of the moving particle with fields E′, B′, via the transformation

′ ′ v E|| = E|| , E⊥ = γ E⊥ + B , (4.283)  c ×  ′ ′ v B|| = B|| , B⊥ = γ B⊥ E , (4.284)  − c ×  with v being the velocity of the particle and γ 1/ 1 v2/c2 the Einstein parame- ≡ − ter (4.17). These equations can also be written withoutq separating transverse and longitudinal components as v γ2 v v E′ = γ E + B E , (4.285)  c ×  − γ +1 c  c ·  v γ2 v v B′ = γ B E B . (4.286)  − c ×  − γ +1 c  c ·  The transformed fields exert the observed electric and magnetic forces eE′ +gB′. The subscripts || and ⊥ indicate projections of the fields parallel and orthogonal to v. From this transformation one may derive the transformation law of the vector field Aµ under Lorentz transformations. Let the frame, in which the moving particle is at rest, be related to the laboratory frame by

x = B()x, (4.287)

where B() is a boost into the v-direction with a rapidity ζ determined by the velocity v via the hyperbolic relations v v cosh ζ = γ =1/ 1 v2/c2, sinh ζ = γ , tanh ζ = . (4.288) − c c q Then the transformation law (4.284) is equivalent to

′µ ′ µ ν A (x )= B ν()A (x), (4.289) apart from an arbitrary gauge transformation. An analogous transformation law holds for rotations, so that the transformations (4.284) and their rotated forms correspond to the Lorentz transformations:

′µ ′ µ ν A (x )=Λ νA (x), (4.290) plus possible gauge transformations. In the notation (4.51), we write the transfor- mation law as Λ Aµ(x) A′ µ(x)=Λµ Aν(Λ−1x). (4.291) −−−→ Λ ν By analogy with Eq. (4.122) for the scalar field, and recalling (4.69), this transfor- mation can be generated as follows:

µ Λ ′µ µ ν −i 1 ω Jˆλκ µ A (x) A (x)= Dˆ(Λ)Λ A (x) e 2 λκ A (x), (4.292) −−−→ ν ≡ 280 4 Free Relativistic Particles and Fields

with the parameters ωµν specified in (4.70) and (4.71), and the combined operator being Jˆλκ Lλκ 1+1ˆ Lˆλκ (4.293) ≡ × × being the generators of Lorentz transformations of both spacetime vector indices and field arguments. The former are generated by the 4 4 matrices Lµν of Eq. (4.65), the latter by the differential operators (4.92). The combined× generators (4.293) are called the generators of the total four-dimensional angular momentum. The two generators on the right-hand side of Eq. (4.293) act upon different spaces, transforming once the vector index and once the spacetime coordinate x. Often one therefore writes, shorter and somewhat sloppily,

Jˆµν Lµν + Lˆµν , (4.294) ≡ with the tacit understanding that the right-hand side abbreviated the direct sum (4.293) of spin and orbital generators. Since the two terms in (4.293) act independently on space and spin indices, the operators Jˆλκ satisfy the same commutation rules as Lλκ and Lˆλκ:

[Jˆµν , Jˆλκ]= i(gµλJˆνκ gµκJˆνλ + gνκJˆµλ gνλJˆµκ). (4.295) − − − The transformation laws (4.290) and (4.292) differ from those of a scalar field in Eqs. (4.50) and (4.122) in the way discussed above for particles with nonzero intrinsic angular momentum. The field has several components. It points in the same spatial direction before and after the change of coordinates. This is ensured by its components changing in the same way as the coordinates of the point xµ. µ Note that the four-divergence ∂ Aµ(x) is a scalar field in the sense defined in (4.50). Indeed

′µ ′ ′ µ ν λ ν ∂ Aµ(x )=(Λ ν∂ )Λµ Aλ(x)= ∂ Aν(x). (4.296)

For this reason the second term in the action (4.248) is Lorentz-invariant, just as the mass term in (4.167). The invariance of the first term is shown similarly:

′ν ′ ′2 ′ ′ ν λ ′2 κ ′ ν ′2 ν 2 A (x )∂ Aν(x )=Λ λA (x)∂ Λν Aκ(x )= A (x)∂ Aν(x)= A (x)∂ Aν(x). (4.297)

Hence the action (4.248) does not change under Lorentz transformations, as it should.

4.7 Other Symmetries of Electromagnetic Action

Just as the scalar action, also the electromagnetic action (4.237) is invariant under more symmetry transformations than those of the Lorentz group. 4.7 Other Symmetries of Electromagnetic Action 281

4.7.1 Translations of the Under spacetime translations (4.185) of the coordinates, the vector field transforms like A′µ(x)= Aµ(x a). (4.298) − The combination of these with Lorentz transformations forms the Poincar´egroup (4.189), ′µ µ ν µ x =Λ ν x + a , (4.299) under which the field Aµ(x) transforms like

A′µ(x)=Λµ Aν(Λ−1(x a)), (4.300) ν − leaving the action (4.248) invariant. As in the scalar equation (4.193), we can generate all Poincar´etransformations of the vector potential Aµ(x) with the help of differential operators such as

µ ′µ µ ν iaµpˆ /¯h −i 1 ω Jˆλκ µ A (x) A (x)= Dˆ(a)Dˆ(Λ)Λ A (x) e µ e 2 λκ A (x), (4.301) −−−→ ν ≡ with the parameters ωµν specified in (4.70), (4.71).

4.7.2 Space Inversion, Time Reversal, and Charge Conjugation of the Vector Field Under space inversion, the four-vector Aµ(x) behaves as follows:

P Aµ(x) A′µ (x)= A˜µ(˜x). (4.302) −−−→ P Under time reversal one has

T Aµ(x) A′µ (x)= A˜µ( x˜), (4.303) −−−→ T − where A˜µ =(A0, Ai). (4.304) − In principle, there is the possibility of a vector field V µ(x) transforming like

P V µ(x) V ′µ (x)= η V˜ µ(˜x), (4.305) −−−→ P P µ with ηP = 1. For ηP = 1 the field V (x) is called an axial vector field. The electromagnetic± gauge field −Aµ(x), however, is definitely a proper vector field. This follows from the vector nature of the electric field and the axial vector nature of the magnetic field, which are observed in the laboratory. Similarly, the behavior of a physically observable vector field V µ with respect to time reversal is given by

T V µ(x) V ′µ (x)= η V˜ µ∗(x ), (4.306) −−−→ T T T 282 4 Free Relativistic Particles and Fields

with an arbitrary phase factor ηT . If the vector field is real and physically observable, then the only alternatives are ηT = 1. For the electromagnetic vector potential µ ± A (x), the phase factor ηT is as specified in (4.303). It reflects the fact that under time reversal, all spatial currents change their directions whereas the zeroth compo- nent stays the same. This reverses the direction of the B-field but has no influence on the E-field generated by flowing charges. The complex conjugation on the right-hand side of (4.306) has the same origin as in the transformation law (4.227) of the complex scalar field. The operation of charge conjugation is performed by exchanging the sign of all charges without changing their direction of flow. Then both E and B change their directions. Hence

C Aµ(x) A′µ (x)= Aµ(x). (4.307) −−−→ C − In general, the vector field could be transformed as

C Aµ(x) A′µ (x)= η Aµ(x), (4.308) −−−→ C C with η = 1. The fact that the electromagnetic field has η = 1 means that it C ± C − is odd under charge conjugation.

4.8 Plane-Wave Solutions of Maxwell’s Equations

The plane-wave solutions of the field equations (4.245) are direct extensions of Eqs. (4.180) and (4.181):

µ 1 µ −ikx µ∗ 1 µ∗ ikx fk (x, t)= ǫ (k,λ)e , fk = ǫ (k,λ)e , (4.309) √2Vk0 √2Vk0 or µ µ −ikx µ∗ µ∗ ikx fk (x, t)= ǫ (k,λ)e , fk = ǫ (k,λ)e , (4.310) with the momentum on the mass shell with M = 0, the so-called light cone. The momentum-dependent four-vectors ǫµ(k,λ) specify the polarization of the plane elec- tromagnetic wave. The label λ counts the different polarization states. In the Lorenz gauge, the vector potential must have a vanishing four divergence (4.252), implying the condition µ kµǫ (k,λ)=0. (4.311) Being solutions of the wave equations, there is a further restricted gauge freedom µ µ (4.255). We may add to the solutions fk (x, t) or fk (x, t) the total gradient of a function Λ(x) which is itself a plane wave Λ(x)= e−ikx with k2=0, thus solving also the Klein-Gordon equation. The total gradient adds to the polarization vector a term proportional to the four-momentum kµ:

ǫµ(k,λ) ǫ′µ(k,λ)= ǫµ(k,λ)+ kµΛ(k,λ). (4.312) → 4.8 Plane-Wave Solutions of Maxwell’s Equations 283

By choosing 1

Λ(k,λ)= k ¯(k,λ), (4.313) −k2 · the spatial part of the polarization vector ǫ′µ(k,λ) acquires the property

k ¯(k,λ)=0, (4.314) · which is the Coulomb gauge (4.260) for the polarization vector. We can also choose 1 Λ(k,λ)= ǫ0(k,λ), (4.315) −k0 and ǫ′µ(k,λ) will satisfy ǫ0(k,λ)=0, (4.316) which is the axial gauge (4.265) for the polarization vector. In Section 4.6.2 we showed that the two gauges coincide for free fields. Here we can see this once more explicitly. The Lorenz gauge (4.311) implies that spatial and time-like components of ǫµ(k,λ) are related by 0 0

k ǫ (k,λ)= k ¯(k,λ), (4.317) · so that the two conditions (4.314) and (4.316) are indeed the same, and

µ 1 ¯ k k ¯ k k ǫ ( ,λ)= 0 ( ,λ), ( ,λ) . (4.318) k ·  Since the four-component vectors ǫ0(k,λ) are restricted by two conditions, only two of them can be independent. These will be labeled by λ = 1 and are con- structed as follows: ± In the axial gauge with (4.316), we set

µ

ǫ (k,λ) (0, ¯(k,λ)), (4.319) ≡ and impose on the spatial part the Coulomb gauge property (4.314). This equation is solved by two polarization vectors orthogonal to the spatial momentum k. These

are defined uniquely by the following consideration: If k points in the z-direction, ¯ then the two vectors ¯(k, 1) coincide with the eigenvectors ( 1) of the 3 3 - ± ± × matrix L of the rotation group in Eq. (4.54), with eigenvalues 1. There are three 3 ± eigenvectors ¯(λ) (λ =1, 0, 1) which are determined by the equations

− ¯ L ¯(λ)= λ (λ), λ =1, 0, 1. (4.320) 3 − The result is

1 0

1 1 ¯ ¯( 1) = i , (0) = 0 . (4.321) √  ±  √   ± ∓ 2 0 − 2 1         284 4 Free Relativistic Particles and Fields

The opposite signs of ¯( 1) are chosen to comply with the so-called Condon-Shortley phase convention3 to be± discussed in detail in Subsection 4.18.3 (see p. 356 and Fig. 4.3). They ensure that the 3 3 raising and lowering matrices formed from the 3 3 spatial submatrices of the generators× (4.55) and (4.56), × 0 0 1 ∓ L± = L1 iL2 =  0 0 i  , (4.322) ± 1 i −0    ± 

have the positive matrix elements

¯ ¯ ¯ L ¯( 1) = √2 (0), L (0) = √2 (+1), (4.323)

+ − +

¯ ¯ ¯ L ¯(+1) = √2 (0), L (0) = √2 ( 1). (4.324)

− − − ¯ The vectors ¯( 1) and (0) are the so-called spherical components of the three- dimensional unit± vectors:

1 0 0

¯ ¯ ¯1 =  0  , 2 =  1  , 3 =  0  , (4.325) 0 0 1             related by

1

¯ ¯ ¯( 1) ( 1 i 2). (4.326) ± ≡ ∓√2 ±

Together with the unit vector ¯ ¯ (0) = 3, (4.327) they form a basis of the unitary spin-1 representation of the rotation group.

In order to obtain the polarization vectors ¯(k, 1) for momenta in an arbitrary ± ˆ direction, we must rotate ¯( 1) into the direction k. We shall use the rotation ± matrix4 cos θ cos φ sin φ sin θ cos φ −iφL3 −iθL2 − R(θ,φ)= e e =  cos θ sin φ cos φ sin θ sin φ  , (4.328) sin θ 0 cos θ    −  with the spherical angles θ [0, π), φ [0, 2π), (4.329) ∈ ∈ to arrive at a momentum direction sin θ cos φ ˆ k =  sin θ sin φ  , (4.330) cos θ     3E.U. Condon and G.H. Shortley, Theory of Atomic Spectra, Cambridge University Press, New York, 1935. 4Some authors prefer to use the rotation matrix R(θ, φ)= e−iφL3 e−iθL2 eiφL3 . 4.8 Plane-Wave Solutions of Maxwell’s Equations 285 with the polarization vectors

cos θ cos φ i sin φ 1 ∓

¯ (k, 1) = cos θ sin φ i cos φ . (4.331) √   ± ∓ 2 sin±θ    −  Together with the third vector ˆ ¯ (k, 0) k, (4.332) ≡ for which there is no electromagnetic plane wave, they form a representation of spin 1 that diagonalizes L2 and the helicity matrix H = kˆ L formed from the matrices · L of Eq. (4.57). The labels λ = 1 specify the two helicities of a light wave running along kˆ. They are observed in the± form of right and left circularly polarized light. Since the spatial polarization vectors (4.331) are orthogonal to the momentum vector k, they are also referred to as transverse polarization vectors, and the Coulomb gauge condition (4.314) is also called transverse gauge condition.

The spatial polarization vectors ¯(k,λ) are orthonormal:

′ ∗ ′ ′ ¯ ¯(k,λ) (k,λ ) = δ , λ,λ =0, 1, (4.333) · λλ ± and they are transversely complete:

i j i j ∗ ij ij k k

ǫ (k,λ)¯ (k,λ) = P (k) δ . (4.334) T ≡ − k2 λ=X−1,1 ij The matrix PT (k) is a projection into a direction transverse to k. It satisfies the defining property of a projection matrix

ij ik ik PT (k)PT (k)= PT (k). (4.335) The contribution of λ = 0 to the polarization sum (4.334) is the longitudinal pro- jection

i j

i j ∗ ij i j ∗ k k ¯ ǫ (k,λ)¯ (k,λ) = P (k) ǫ (k, 0) (k, 0) = . (4.336) L ≡ k2 λX=0 The orthonormality (4.333) goes over to the four-dimensional polarization vectors in the axial gauge (4.319) as follows:

µ ′ ∗ ′ ǫ (k,λ)ǫ (k,λ ) = δ ′ , λ,λ =0, 1. (4.337) µ − λλ ± These vectors have the reflection property

ǫµ (k,λ)= ǫµ (k, λ)∗, (4.338) − which follows from the fact that k k corresponds to θ π θ, φ φ + → − → − → π (mod 2π). 286 4 Free Relativistic Particles and Fields

In four dimensions, the polarization sum over the two helicities in (4.334) leads to the following 4 4 polarization tensor: × µν µ ν ∗ PT (k) = ǫ (k,λ)ǫ (k,λ) . (4.339) λX=±1 In the axial gauge where the polarization vectors have the form (4.319), this reads

µν 0 0 µν µ ν ∗ PT (k) = ǫ (k,λ)ǫ (k,λ) =  0 P ij(k)  λX=±1 T µν µν 0 0  1 0 = = gµν + . (4.340)  0 δij kikj/k2  −  0 kikj/k2  − −     This is a 4 4 -matrix projecting into a purely spatial two-dimensional subspace transverse to× the vector k. It contains only purely spatial nonzero components ij PT (k). There is also an associate longitudinal projection tensor which reads in the Lorenz gauge (4.317)

µν µν 1 0 1 0 P µν(k) = ǫµ(k, 0)ǫν(k, 0)∗ = = . (4.341) L  ij   i j 2  1 PL (k) 1 k k /k     The projection is obviously noncovariant since it lends a special significance to the zeroth component of the electromagnetic vector field. To exhibit the noncovariance, it is useful to introduce a fixed timelike unit vector 1 0 ηµ   . (4.342) ≡ 0      0    We also define a purely spacelike unit vector orthogonal to it, pointing along the direction of k: 0 k¯µ . (4.343) ≡ kˆ ! This can be expressed in terms of ηµ and the momentum vector kµ as follows: kµ (kη)ηµ k¯µ − . (4.344) ≡ (kη)2 k2 − q The fixed vector eliminates the zeroth component of kµ, no matter whether it is on-shell or off-shell. We readily show that the polarization sum (4.340) can be rewritten as

P µν(k) gµν + ηµην k¯µk¯ν = gµν + ηµην + P µν(k) T ≡ − − − L kµkν kµην + kνηµ ηµην = gµν + kη k2 . (4.345) − − (kη)2 k2 (kη)2 k2 − (kη)2 k2 − − − 4.9 Gravitational Field 287

In the Lorenz gauge it is useful to introduce, in addition to the three four- dimensional polarization vectors (4.318), also an extra fourth vector that points parallel to the four-momentum of the particle. It will be called scalar polarization vector, and denoted by µ µ ǫ (k,s) k =(ωk, k). (4.346) ≡ The associated vector field corresponds to a pure gauge degree of freedom, since in x-space it has the form ∂µΛ. As such, it transforms under an extra indepen- dent and irreducible representation of the Lorentz group describing a scalar particle degree of freedom. Thus, it certainly does not contribute to the gauge-invariant electromagnetic action and is no longer part of the vector particle. One may define scalar products by complete analogy with those for scalar field in Eqs. (4.177), except that they contain an extra contraction of the polarization indices: ↔ µ∗ ′ 3 (fp , fp) d x fp′ (x, t)i ∂0 fµp(x, t), (4.347) ≡ Z with obvious definitions between positive- and negative-frequency solutions, fp′ (t) ∗ ∗ and fp′ (t), and between two negative-frequency solutions fp′ (t).

4.9 Gravitational Field

The gravitational field is carried by a varying metric gµν(x) in spacetime. Its pres- ence manifests itself in a local dependence of the invariant distance between events:

2 µ ν (ds) = gµν(x)dx dx . (4.348) The in such a spacetime no longer satisfy the axioms of Minkowski geo- metry since the spacetime can have a local curvature. According to the equivalence principle of general relativity, the motion of point particles is independent of their mass. All point particles follow the lines of shortest distance in this geometry, the so-called geodesics. The generation of the gravitational field is governed by a complicated nonlinear theory and deserves a detailed treatment on its own, not to be elaborated in this text. In the present context we merely state a few relevant facts about a very weak gravitational field running through empty spacetime. It may be described by a small deviation of the metric gµν (x) from the Minkowski metric (4.27), which in this context will be denoted by ηµν, for better distinction. The deviation is h (x) g (x) η . (4.349) µν ≡ µν − µν

Note that, while the metric depends linearly on hµν (x), g (x) η (x)+ h , (4.350) µν ≡ µν µν µν the tensor g , being the inverse of gµν , has a nonlinear expansion

gµν(x) ηµν(x) hµν + hµ hλν hµ hλ hκν + (h3). (4.351) ≡ − λ − λ κ O 288 4 Free Relativistic Particles and Fields

4.9.1 Action and Field Equations

The action of the field hµν (x) is obtained from the famous Einstein-Hilbert action of the gravitational field

f 1 = d4x√ gR, (4.352) A −2κ Z − where R is the curvature scalar of spacetime, and κ is related to the famous Newton gravitational constant

G 6.673 10−8 cm3g−1s−2 (4.353) N ≈ · by 1 c3 = . (4.354) κ 8πGN A natural length scale of gravitational physics is the Planck length, which can be formed from a combination of Newton’s gravitational constant (4.353), the light velocity c 3 1010 cm/s, and Planck’s constanth ¯ 1.05459 10−27: ≈ × ≈ × 3 −1/2 c −33 lP = 1.615 10 cm. (4.355) GN h¯ ! ≈ × This is the Compton wavelength l h/m¯ c associated with the Planck mass P ≡ P defined by

1/2 ch¯ −5 22 2 mP = 2.177 10 g=1.22 10 MeV/c . (4.356) GN ! ≈ × ×

The constant 1/κ in the action (4.352) can be expressed in terms of the Planck length as 1 h¯ = 2 . (4.357) κ 8πlP

κ The curvature scalar R is formed from the Riemann curvature tensor Rµνλ by the contraction R = R µ gµνR with the Ricci tensor R R µ. The symbol µ ≡ µν νλ ≡ µνλ g denotes the determinant of the metric tensor gµν which makes the volume element √ gd4x invariant under coordinate transformations. The curvature tensor may be written− as a covariant curl [10]

R κ (∂ Γ ∂ Γ ) κ [Γ , Γ ] κ (4.358) µνλ ≡ µ ν − ν µ λ − µ ν λ λ of the matrices (Γµ)ν formed from the Christoffel connection 1 (Γ ) λ Γ λ = gλκ(∂ g + ∂ g ∂ g ), (4.359) µ ν ≡ µν 2 µ νκ µ µκ − κ µν 4.9 Gravitational Field 289 in terms of which it reads

R κ ∂ Γ κ ∂ Γ κ Γ δΓ κ +Γ δΓ κ. (4.360) µνλ ≡ µ νλ − ν µλ − µλ νδ µλ νδ Using this the integrand in (4.352) can be replaced by

√ gR = ∂ (gµν√ g) Γ λ δ λΓ κ + √ ggµν Γ κ Γ λ Γ λ Γ κ . (4.361) − λ − µν − µ νκ − µλ νκ − µν λκ h  i   The first term is a pure divergence, and it may be omitted in the action (4.352), µν keeping only the second term. After setting gµν = ηµν + hµν , and inserting g from (4.351), we calculate

µ ν µ √ g = e(1/2)tr log(−ηµν −hµν ) = e(1/2)tr log(−ηµν )e(1/2)hµ −(1/4)hµ hν +... − 1 1 1 = 1+ h µ h ν h µ + (h µ)2 + .... (4.362) 2 µ − 4 µ ν 8 µ

We also expand the Christoffel symbols to linear order in hµν , 1 Γ λ γ λ ∂ h λ + ∂ h λ ∂λh . (4.363) µν ≈ µν ≡ 2 µ ν ν µ − µν   Inserting this into (4.360), we find the linear contributions to the curvature tensor (4.360): 1 R [∂ ∂ h ∂ ∂ h (µ ν)] + .... (4.364) µνλκ ≈ 2 µ λ νκ − ν κ µλ − ↔ For the Ricci tensor defined by the contraction

R κ gνλR κ (4.365) µ ≡ µνλ this amounts to 1 R (∂2h ∂ ∂ h ∂ ∂ h + ∂ ∂ h)+ .... (4.366) µκ ≈ −2 µκ − µ λ λκ − κ λ λµ µ κ The ensuing scalar curvature R R µ starts out like5 ≡ µ R ∂2h ∂ ∂ hµν + ... ∂2hs + ..., (4.367) ≈ − µ ν ≡ s s where h is the result of applying the scalar projection operator Pµν,λκ of Eq. (4G.6) to hλκ and taking the trace

hs ∂2h ∂ ∂ hµν = ηµνP s hλκ. (4.368) ≡ − µ ν µν,λκ The quadratic part of the action is found by inserting (4.363) into Eq. (4.361), which contributes to R in (4.367) a quadratic term

∆(2)R = gµν Γ κ Γ λ Γ λ Γ κ . (4.369) − µλ νκ − µν λκ   5 The omitted quadratic parts in hµν have been given in Appendix B of Ref. [16]. 290 4 Free Relativistic Particles and Fields

The corresponding quadratic action governing free gravitational waves is

f 1 (2) = d4x ∆(2)R. (4.370) A −2κ Z Explicitly this reads

f (2) 1 4 λ λ λ ν µ µν µ ν d x (∂µhν + ∂ν hµ ∂ hµν )(∂ hλ + ∂λh ∂ h λ) A ≈ −8κ − − Z n (∂ hµλ + ∂ hµλ ∂λh µ)(∂ h ν + ∂ h ν ∂ν h ) . (4.371) − µ µ − µ ν λ λ ν − νλ o Using the symmetry of hµν , the right-hand side can be rearranged to

f (2) 1 4 νλ 2 µ νλ µν 2 = d x(h ∂ hνλ +2∂ hµν∂λh 2h∂µ∂νh h∂ h), (4.372) A −8κ Z − − λ where h is defined to be the trace of the tensor hµν, i.e., h hλ . An alternative way of writing this is ≡

f 1 4 λµκσ ντδ = d x hµν ǫ ǫλ ∂κ∂τ hσδ. (4.373) A −8κ Z The equivalence can be verified with the help of the identity

λµκσ ντδ µν κτ σδ µτ κδ σν µδ κν στ µν κδ στ µτ κν σδ µδ κτ σν ǫ ǫλ = η η η η η η η η η + η η η + η η η + η η η . − − − (4.374) A further useful way of writing the field action (4.373) is obtained by using the Einstein tensor 1 Gµν Rµν gµνR, (4.375) ≡ − 2 whose linear approximation reads

1 Gµκ= Rµκ gµκR − 2 1 1 (∂2hµκ ∂µ∂ hλκ ∂κ∂ hλµ + ∂µ∂κh)+ ηµκ(∂2h ∂ ∂ hνλ), (4.376) ≈ −2 − λ − λ 2 − ν λ with a trace G G µ = R = (∂2h ∂ ∂ hµν ). This may be written as a ≡ µ − − − µ ν four-dimensional double curl: 1 Gµν ǫλµκσǫ ντδ∂ ∂ h , (4.377) ≈ −4 λ κ τ σδ from which we see that the action (4.373) becomes simply

f 4 1 4 µν = d x (x)= d x hµν G . (4.378) A Z L 4κ Z 4.9 Gravitational Field 291

The Einstein tensor plays a similar role in as the dual field tensor F˜µν does in electromagnetism [recall (4.238)]. First, being a double-curl (4.377), it trivially satisfies a Bianchi identity µν ∂µG = 0 (4.379) for any smooth single-valued field hµν (x), i.e., any field hµν (x) which satisfies the integrability condition (∂ ∂ ∂ ∂ )h (x) = 0. Second, it is invariant under local λ κ − κ λ µν gauge transformations [just as (4.239) is under (4.250)]:

h (x) h (x)+ ∂ Λ (x)+ ∂ Λ (x). (4.380) µν → µν µ ν ν µ These are the linearized versions of Einstein’s general coordinate transformations:

xµ xµ +Λµ(x). (4.381) →

As a consequence of this invariance, the symmetric tensor hµν (x) carries only 6 instead of 10 independent physical components. f The free-field equations are obtained by variation of with respect to hµν(x): f A δ 1 A = Gµν (x)=0. (4.382) δhµν (x) 2κ Thus, for a free gravitational field, the Einstein tensor vanishes, and so does the Ricci tensor R = G 1 g G: µν µν − 2 µν

Rµν (x)=0. (4.383)

In the presence of , the field equation (4.382) will be modified. After adding m f an action of the matter fields to the field action , the right-hand side will be m m A µνA −1 shown in Eq. (5.71) to become equal to the tensor T (x)= 2√ g δ /δgµν(x). − − This is the total symmetric energy-momentum tensor of the material particlesA to be derived in (5.66). The vanishing of the variation of the total action

f m tot = + = 0 (4.384) A A A extends therefore the field equation (4.382) by a source term due to matter:

m µν Gµν (x)= κ T (x) (4.385)

4.9.2 Lorentz Transformation Properties of Gravitational Field

Under Lorentz transformations, hµν (x) behaves of course like a tensor. In a straight- forward generalization of the transformation law (4.290) we may immediately write

′µν µ ν λκ −1 h (x)=Λ λΛ κh (Λ x). (4.386) 292 4 Free Relativistic Particles and Fields

This transformation can be generated by analogy with (4.292) as follows:

µν Λ ′µν µ ν λκ −i 1 ω Jˆλκ µν h (x) h (x)= Dˆ(Λ)Λ Λ h (x) e 2 λκ h (x), (4.387) −−−→ λ κ ≡ with the parameters ωµν specified in (4.70) and (4.71), and the operator

Jˆλκ Lλκ 1 1+1ˆ Lλκ 1+1ˆ 1 Lˆλκ. (4.388) ≡ × × × × × × This is a direct generalization of the total angular momentum operator (4.293). The commutation rules between the generators Jˆµν are of course given by (4.295), as in the case of the vector potential.

4.9.3 Other Symmetries of Gravitational Action Just as the scalar and electromagnetic actions, the gravitational action (4.378) is invariant under extensions of the Lorentz group.

4.9.4 Translations of Gravitational Field Under spacetime translations (4.185), the gravitational field transforms like a scalar field in (4.187) and a vector field in (4.298):

h′µν(x)= hµν (x a). (4.389) − The combinations of translations and Lorentz transformations

′µ µ ν µ x =Λ νx + a (4.390) form the Poincar´egroup (4.189). Under these the fields transform like

h′µν (x)=Λµ Λν hλκ(Λ−1(x a)), (4.391) λ κ − leaving the action (4.378) invariant. As in the scalar and vector cases (4.193) and (4.301), we can of course generate all Poincar´etransformations on the field with the help of differential operators:

µ 1 λκ µν ′µν µ ν λκ ia pˆµ/¯h −i ωλκJˆ µν h (x) h (x)= Dˆ(a)Dˆ(Λ)Λ λΛ κh (x) e e 2 h (x), −−−→ ≡ (4.392) with the parameters ωµν specified in (4.70), (4.71).

4.9.5 Space Inversion, Time Reversal, and Charge Conjugation of Gravitational Field

Since hµν (x) determines the invariant distances in space via (4.348) and (4.350), it transforms like dxµdxν under space inversion and time reversal. Under charge conjugation, it is invariant. 4.9 Gravitational Field 293

Thus we have, under space inversion,

P hµν (x) h′µν (x)= h˜µν (˜x), (4.393) −−−→ P where the tilde reverses the sign of hµν (x) for each spatial index, whereas under time reversal, the field hµν transforms like

T hµν (x) h′µν(x)= h˜µν (x ). (4.394) −−−→ T T

In principle, an arbitrary tensor field tµν(x) has two possible transformation behaviors under space inversion:

P tµν (x) t′µν (x)= η t˜µν (˜x), (4.395) −−−→ P P µν with ηP = 1, where in the case ηP = 1 the field t (x) is called a pseudotensor field. The gravitational± field hµν (x), however,− is definitely a tensor field. This follows from its metric nature and the distance definition in (4.348), (4.350). Similarly, the phase factor of a tensor field tµν (x) arising from time reversal could in principle be T tµν (x) t′µν(x)= η t˜µν (x ), (4.396) −−−→ T T T with an arbitrary phase factor ηT . For an observable real field, however, only ηT = 1 are admissible. The gravitational field has η = 1 to preserve the definition of ± T the distance in (4.348), (4.350) under space inversion. Under charge conjugation, the gravitational interactions are invariant, so that

C hµν (x) h′µν (x)= hµν (x). (4.397) −−−→ C

In general, a tensor field tµν(x) could transform like

C tµν (x) t′µν (x)= η tµν (x), (4.398) −−−→ C C with η = 1. C ± 4.9.6 Gravitational Plane Waves To discuss the properties of gravitational waves we remove the gauge freedom (4.380) by fixing a specific gauge, the so-called Hilbert gauge: 1 ∂ hµν (x)= ∂ν h λ(x). (4.399) µ 2 λ It corresponds to the Lorenz gauge of electromagnetism [recall (4.252)]. The Hilbert gauge can always be achieved by a gauge transformation. If hµν (x) is not in this gauge, we simply perform the transformation (4.380) and determine Λµ(x) from the differential equation

∂2Λν(x)= ∂ hµν (x) 1 ∂ν h(x). (4.400) − µ − 2 294 4 Free Relativistic Particles and Fields

The gauge transformation (4.380) can be used to fix another property of hµν. By applying it to the trace h(x) h µ(x), using the special gauge function Λ = ∂ Λ, ≡ µ ν ν we find

h(x) h′(x)= h(x)+2∂2Λ(x). (4.401) → ′ This can be used to arrive at a traceless field hµν (x). If h(x) is nonzero, h (x) vanishes if we choose

Λ(x)= (1/2∂2)h(x). (4.402) − It is useful to introduce the field 1 φ (x) h (x) η h(x). (4.403) µν ≡ µν − 2 µν With this, the Einstein tensor (4.376) reads 1 1 Gµκ (∂2φµκ ∂µ∂ φλκ ∂κ∂ φλµ) ηµν ∂ ∂ φλκ (4.404) ≈ −2 − λ − λ − 2 λ κ with the trace G = G µ = 1 (∂2φ +2∂ ∂ φµν). Under the gauge transformations µ − 2 µ ν (4.380), the field φµν(x) changes like

φµν (x) φµν (x)+ ∂µΛν(x)+ ∂ν Λµ(x) ηµν ∂ Λκ(x), (4.405) → − κ and the Einstein tensor is invariant. Imposing now the Hilbert gauge condition (4.399), we see that

ν ∂ φµν(x)=0, (4.406) and the Einstein tensor (4.404) reduces to 1 G = ∂2φ . (4.407) µν −2 µν The free-field equation of motion (4.382) implies a massless Klein-Gordon equation for each field component:6 ∂2φ (x)=0. (4.408) − µν This must be solved in the Hilbert gauge (4.406). Since the graviton field obeys the zero-mass Klein-Gordon equation (4.408), the plane waves in the field φµν(x) are proportional to e−ikx and eikx, with k0 lying 0 2 on the light cone k = ωk = √k . These waves are accompanied by symmetric polarization ǫµν (k,λ) [compare (4.309)]:

µν 1 µν −ikx µν∗ 1 µν∗ ikx fk (x, t)= ǫ (k,λ)e , fk = ǫ (k,λ)e . (4.409) √2Vk0 √2Vk0 6Compare Eq. (4.253) for the electromagnetic field in the Lorenz gauge. 4.9 Gravitational Field 295

In the Hilbert gauge, the polarization tensors ǫµν (k,λ) satisfy the transversality condition µν kµǫ (k,λ)=0. (4.410) For a plane wave, we can further perform a gauge transformation (4.405) with 2 functions Λµ(x) which satisfy the zero-mass Klein-Gordon equation ∂ Λµ(x) = 0. These are the gauge transformations of the second type that are completely analog- ous to their electromagnetic versions (4.255). The functions Λµ(x) can be chosen to have the plane-wave form

±ikx 2 Λµ(x)=Λµe , with k =0. (4.411)

µν The vector Λµ is a constant that makes the transformed polarization tensor ǫ (k,λ) traceless:

µ ǫµ (k,λ)=0. (4.412) It is also chosen to make it axial in the sense analogous to (4.319): ǫµ0(k,λ)=0. (4.413) Under a gauge transformation of the second kind, the Fourier transforms φµν(k) of the field φµν (x) receive an additional term φµν(k) φµν(k)+ kµΛν + kν Λµ ηµνk Λκ, (4.414) → − κ which does not contribute to any observable quantities. In principle, there exists ten possibilities of forming symmetric transverse polar- ization tensors ǫµν (k,λ) from the transverse polarization vectors (4.318). For their construction we recall the extra vector (4.346). This is the scalar polarization vector that points parallel to the four-momentum: ǫµ(k,s) kµ. (4.415) ≡ This polarization is unphysical since it corresponds to a pure gauge transformation. The gauge condition (4.410) eliminates four of these combinations. The remaining six polarization tensors are constructed from symmetrized tensor products of the transverse polarization vectors (4.331) by forming the combinations: ǫµν (k, 2) ǫµ (k, 1)ǫν (k, 1), H ≡ H H 1 ǫµν (k, 1) [ǫµ (k, 1)ǫν (k, 0)+(µ ν)], H ≡ √2 H H ↔ 1 2 ǫµν (k, 0) [ǫµ (k, 1)ǫν (k, 1)+(µ ν)] + ǫµ (k, 0)ǫν (k, 0), H ≡ √6 H H − ↔ √6 H H 1 ǫµν (k, 1) [ǫµ (k, 1)ǫν (k, 0)+(µ ν)], H − ≡ √2 H − H ↔ ǫµν (k, 2) ǫµ (k, 1)ǫν (k, 1), H − ≡ H − H − 1 1 ǫµν (k,s) [ǫµ (k, 1)ǫν (k, 1)+(µ ν)] ǫµ (k, 0)ǫν (k, 0). (4.416) H ≡ √3 H H − ↔ − √3 H H 296 4 Free Relativistic Particles and Fields

These combinations are formed with the help of the Clebsch-Gordan coefficients calculated in Appendix 4E and listed in Table 4.2. They couple two spin-one objects symmetrically to five components of spin 2 and one component of spin zero.

2 2 = 1 1 1 1 , | i | i| i 1 2 1 = 1 1 1 0 + 1 0 1 1 , | i √2 | i| i | i| i   1 2 2 0 = 1 1 1 1 + 1 1 1 1 + 1 0 1 0 , | i √6 | i| − i | − i| i √6| i| i   1 2 1 = 1 1 1 0 + 1 0 1 1 , | − i √2 | − i| i | i| − i   2 2 = 1 1 1 1 , | − i | − i| − i 1 1 0 0 = 1 1 1 1 + 1 1 1 1 1 0 1 0 . (4.417) | i √3 | i| − i | − i| i − √3| i| i   µ Since the polarization vectors of the three spin-one states satisfy kµǫH (k,λ) = 0, the resulting tensors satisfy automatically the Hilbert gauge condition (4.410). In addition, they are traceless by construction:

µ ǫH µ (k,λ)=0, (4.418) as follows directly from the explicit four-vectors in the Lorenz gauge (4.318):

µ 0 µ 1 ǫH (k, 1) = , ǫH (k, 0) = . (4.419) ± ¯(k, 1) kˆ ± ! ! µ They satisfy the orthogonality properties ǫH (k, 1)ǫH µ(k, +1) = 0 and ǫµ (k, 1)ǫ (k, 0) = 0. The tracelessness of the polarization− tensors is, of course, H ± H µ a consequence of the invariance under the gauge transformation (4.401). The gauge invariance of the second type (4.414) reduces the six degrees of free- dom allowed by the gauge condition (4.410) to only two physical degrees. µ µ µ µ Setting Λ = ǫH (k, 1), ǫH (k, 1), and ǫH (k,s), and using (4.410), we see that − µ the three combinations (4.414) become precisely the polarization tensors ǫH (k, 1) µν ± and ǫH (k,s), respectively. µ Note that not only the scalar polarization vector ǫH (k,s), but also the polar- µ ization vector ǫH (k, 0) is unphysical. This is seen by introducing, in addition to the longitudinal polarization vector (4.346), a further linearly independent four- component object ǫµ(k, s¯)=(k0, k). (4.420) − This four-component object is not a vector as can be seen by forming the product with the polarization vector ǫµ(k,s) yielding:

µ 02 2 2 ǫµ(k,s)ǫ (k, s¯)= k + k =2k , 4.9 Gravitational Field 297 which is not Lorentz-invariant. We shall call the object ǫµ(k, s¯) an antiscalar. The four objects ǫµ (k, 1), ǫµ(k,s), ǫµ(k, s¯) form a complete basis in the space of po- H ± larization vectors. This is expressed in the completeness relation 1 ǫµ (k, 1)ǫν (k, 1)+ǫµ (k, 1)ǫν (k, 1)+ [ǫµ(k,s)ǫν(k, s¯)+(µ ν)]= ηµν.(4.421) H H − H − H 2k2 ↔ − Using (4.339), this can be rewritten as 1 P µν(k)+ [ǫµ(k,s)ǫν (k, s¯)+(µ ν)]= ηµν. (4.422) T 2k2 ↔ − The first two terms on the left-hand side can obviously be replaced by the polariza- µν tion tensor ǫH (k, 0), so that (4.421) may be rewritten as 1 ǫµν (k, 0) = ηµν [ǫµ(k,s)ǫν(k, s¯)+(µ ν)]. (4.423) H − − 2k2 ↔ µ µ Inserting on the right-hand side the explicit form ǫH (k,s) = k for the scalar po- µ 2 µ larization vector (4.415), and setting ǫH (k, s¯)/2k = Λ , the right-hand side can µ ν ν µ µν κ be rewritten as k Λ k Λ + η kκΛ , thus demonstrating that the polarization µν − − tensor ǫH (k, 0) is of the pure gauge form (4.414), and thus unphysical. Hence we remain with only two physical polarization tensors ǫµν (k, 2). These H ± describe gravitational waves with helicities λ = 2. An analysis of the temporal µν ± behavior of the fields shows that the tensors ǫH (k, 2) give the gravitational waves a circular polarization, anticlockwise or clockwise around± the momentum direction, respectively. These are analogs of the circularly polarized light waves whose polar- ization vectors ǫµ (k, 1) were discussed in Section 4.8. H ± In electromagnetism, one often describes plane waves with the help of real po- larization vectors ǫµ (k, 1) = [ǫµ (k, +1) + ǫµ (k, 1)], ǫµ (k, 2) = i[ǫµ (k, +1) ǫµ (k, 1)].(4.424) H − H H − H H − H − These describe linearly polarized light waves whose field vectors oscillate along the directions orthogonal to the momentum k. By analogy, we introduce the real com- binations of the two physical polarization tensors 1 ǫµν (k) [ǫµν (k, +2) + ǫµν (k, 2)], (4.425) H+ ≡ √2 H H − 1 ǫµν (k) = [ǫµν (k, +2) ǫµν (k, 2)]. (4.426) H× √2i H − H −

The motion of a circular ring of mass points in a plane gravitational wave hµν (x), with the polarization tensors (4.425) or (4.426), reveals the physical properties of the associated waves. They distort the circular ring periodically into an ellipsoidal one, in which the principal axes point in the directions 1 and 2. A wave carrying the µν polarization tensor ǫH×(k) has the same effect with the axes rotated by 45 degrees. For more details see Section 5.5. 298 4 Free Relativistic Particles and Fields

The spatial components of the two polarization tensors satisfy the completeness relation ǫij (k)ǫkl (k)+ ǫij (k)ǫkl (k) = ǫij (k, 2)[ǫkl (k, 2)]∗ + ǫij (k, 2)[ǫkl (k, 2)]∗ H+ H+ H× H× H H+ H − H× − ij,kl ˆ = PT T (k), (4.427) ij,kl ˆ where PT T (k) is the projection matrix 1 1 P ij,kl(kˆ) P ik(kˆ)P jl(kˆ)+ P il(kˆ)P jk(kˆ) P ij(kˆ)P kl(kˆ), (4.428) T T ≡ 2 T T T T − 2 T T h i formed from products of the transverse projection matrices (4.334) of the electro- magnetic waves. It is easy to verify the projection property of (4.428): ij ˆ st,kl ˆ ij,kl ˆ PT T st(k)PT T (k)= PT T (k). (4.429) Let us apply this projection to the purely spatial part of a gravitational plane wave hλκ(x) in the Hilbert gauge (4.399), which by a further gauge transformation of the ν second kind (4.414) has been made traceless and axial. In this way, the field φµ has µ µ been transformed to satisfy φµ = 0, φ0 = 0. The result is a transverse-traceless field: ij ij,kl hT T (x)= PT T hkl(x). (4.430) Its modes briefly called T T -waves. The explicit form of the nonzero purely spatial µν,λκ ˆ components of PT T (k) is P ij,lm(kˆ)= δilδjm 1 δijδlm 2δilkˆjkˆm + 1 δijkˆlkˆm + 1 δlmkˆikˆj + 1 kˆikˆjkˆlkˆm. (4.431) T T − 2 − 2 2 2 If this is applied to any symmetric tensor hlm(x), the result is transverse l ∂ hml =0, (4.432) ij,lm ˆ since kiPT T (k) = 0, and it is traceless since ii,lm ˆ ij,ll ˆ PT T (k)=0, PT T (k)=0. (4.433) The latter is a reflection of the gauge property (4.418) that was derived from the ij,lm ˆ particular gauge invariance under (4.401). If PT T (k) is applied to a symmetric traceless tensor, it can be simplified to P ij,lm(kˆ)= δilδjm 2δilkˆjkˆm + 1 kˆikˆjkˆlkˆm. (4.434) T T − 2 After the on-shell gauge fixing of the second type of the fields hµν in the Hilbert µ µ gauge, the field components have the gauge properties h0 =0, hµ = 0. The full set of projection matrices in the space of symmetric tensor fields, which are not restricted by any gauge condition, is summarized in Appendix 4G. There we derive a completeness relation (4G.17) which permits decomposing any symmetric tensor field hµν into its irreducible components under transformations of the Lorentz group: (2) l s hµν = hµν + hµν + hµν , (4.435) where h(2) P (2) hλκ, hl P l hλκ, hs P s hλκ. (4.436) µν ≡ µν,λκ µν ≡ µν,λκ µν ≡ µν,λκ 4.10 Free Relativistic Fermi Fields 299

4.10 Free Relativistic Fermi Fields

For Fermi fields, the situation is technically more involved. Experimentally, fermions always have an even number of spin degrees of freedom. We shall denote the asso- ciated field by ψa, where the index a labels the different spin components. Under rotations, these spin components are transformed into each other, as observed ex- perimentally in the Stern-Gerlach experiment. We shall see below that Lorentz transformations also lead to certain well-defined mixtures of different spin compo- nents. The question arises as to whether we can construct a Lorentz-invariant action involving (2s + 1) field components. To see the basic construction principle we use as a guide the known transformation law (4.290) for the four-vector field Aµ. For an arbitrary spinor field we postulate the transformation law

Λ ψ (x) ψ′ (x′)= D (Λ)ψ (x), (4.437) a −−−→ a ab b with an appropriate (2s +1) (2s + 1) spinor transformation matrix D (Λ) which × ab we have to construct. This can be done by purely mathematical arguments. The construction is the subject of the so-called group . First of all, we perform two successive Lorentz transformations,

′′ ′ x =Λx =Λ2x =Λ2Λ1x. (4.438)

Since the Lorentz transformations Λ , Λ are elements of a group, the product Λ 1 2 ≡ Λ2Λ1 is again a Lorentz transformation. Under the individual factors Λ2 and Λ1, the field transforms as

Λ1 ψ(x) ψ′(x′) = D(Λ )ψ(x), −−−→ 1 Λ2 ψ′(x′) ψ′′(x′′) = D(Λ )ψ′(x′), (4.439) −−−→ 2 so that under Λ = Λ2Λ1 the spinor goes over into

Λ2Λ1 ψ(x) ψ′′(x′′)= D(Λ )D(Λ )ψ(x). (4.440) −−−→ 2 1 For the combined Λ, the transformation matrix is D(Λ), so that

′′ ′′ ψ (x )= D(Λ2Λ1)ψ(x). (4.441)

Comparing this with (4.440) shows that the matrices D(Λ) mixing the spinor field components under the Lorentz group must follow a group multiplication law. Their products must be the same for the group elements itself. Thus the mapping

Λ D(Λ) (4.442) → is a homomorphism, and the various D(Λ)’s form a matrix representation of the group. 300 4 Free Relativistic Particles and Fields

Note that the transformation law (4.290) for the vector field Aµ(x) follows the same rule with −i 1 ω Lµν D(Λ) Λ= e 2 µν (4.443) ≡ being the defining 4 4 -representation of the Lorentz group. The parameters ω × µν are specified in (4.70) and (4.71). The scalar field transformation law (4.50) follows trivially the same rule, where D(Λ) 1 is the identity representation. ≡ The group laws for Λ and D(Λ) are sufficiently stringent to allow only for a countable set of fundamental 7 finite-dimensional transformation matrices D(Λ). We shall see below that these are characterized by two quantum numbers, s1 and s2, with 1 3 either one of these taking all possible half-integer or integer values 0, 2 , 1, 2 ,... . The representation spaces associated with D(s1,s2)(Λ) will turn out to harbor particles of spins s s to s + s . Hence, particles with a single fixed spin s can only follow | 1 − 2| 1 2 the D(s,0)(Λ) or D(0,s)(Λ) transformation laws. In Sections 4.18 and 4.19 we shall learn how the representation matrices D(s1,s2)(Λ) are most efficiently constructed. We have to determine all possible sets of six matrices which satisfy the commutation relations (4.86). Any set of such matrices forms a representation of the Lie algebra defined by (4.86). In this general framework, the scalar particles studied so far transform with the trivial representation D(0,0)(Λ) 1 of the Lorentz group. They are said to have spin ≡ zero. Let us now study the smallest nontrivial representation.

4.11 Spin-1/2 Fields

The smallest matrices satisfying the subalgebra (4.76) associated with the rotation subgroup of the Lorentz group are

σi L = , (4.444) i 2 where σi are the

0 1 0 i 1 0 σ1 = , σ2 = − , σ3 = . (4.445) 1 0 i 0 0 1 ! ! − ! 2 The two-component basis on which L and L3 are diagonal, with eigenvalues 3/4 and 1/2, respectively, are the Pauli ±

1 1 1 0 χ( 2 )= , χ(− 2 )= . (4.446) 0 ! 1 !

7Mathematically, fundamental means that the representation is irreducible. Any arbitrary re- presentation is equivalent to a direct sum of irreducible ones. 4.11 Spin-1/2 Fields 301

The full Lie algebra (4.76)–(4.78) can be satisfied in two inequivalent ways, either by the choice M = iσi/2, (4.447) i − or by i Mi = iσ /2. (4.448) The first amounts to the representation

σi σi (Li, Mi)= , i (4.449) 2 − 2 !

1 that is also denoted by ( 2 , 0), the second to

σi σi (Li, Mi)= , i (4.450) 2 2 !

1 denoted by (0, 2 ). Exponentiating these generators we obtain the global represen- tations of the Lorentz group:

1

   ( ,0) −i( ³· /2−i · /2) D 2 (Λ) = e , (4.451)

1

   (0, ) −i( ³· /2+i · /2) D 2 (Λ) = e . (4.452) An alternative decomposition that will later be useful is a factorization into rotations and subsequent boosts:

1

 ³  ( ,0) −· /2 −i · /2 D 2 (Λ) = e e , (4.453)

1

 ³  (0, ) · /2 −i · /2 D 2 (Λ) = e e . (4.454) The right-hand sides can easily be calculated explicitly by expanding them in a Taylor series. The expansion terms separate naturally into even and odd powers,

2 ˆ 2

³  due to the normalization property of the -matrices (ˆ ) =1and( ) = 1, where

ˆ  ³ˆ, are the directions of rotation axis and Lorentz boost, respectively. For rotations,

2k 2k 2k+1 2k+1

³ ³ this implies for integer k the power laws (³) = ϕ , ( ) = (ˆ ) ϕ , so that

∞ n n ∞ k 2k ∞ k 2k+1

( i) ³ˆ ( 1) ϕ ( 1) ϕ  −i³· /2 e = − = − i³ˆ − . (4.455) n! 2 ! (2k)! 2 − (2k + 1)! 2 nX=0 kX=0   kX=0   Summing up separately even and odd powers of the parameters ϕ, this becomes

ϕ ϕ 

−i³· /2

 ³

R³ (ϕ)= e = cos i ˆ sin . (4.456) ˆ 2 − · 2 Similarly we find that

ζ ζ 

∓ · /2 ˆ

  B()= e = cosh sinh . (4.457) 2 ∓ · 2 302 4 Free Relativistic Particles and Fields

When applying these representations to a particle of mass M at rest, energy and momentum are boosted from pµ =(M, 0) to µ 0 ˆ p =(p , p)= M(cosh ζ,  sinh ζ). (4.458) Using the relations

ζ 1 p + M cosh = (cosh ζ +1)= 0 , 2 s2 s 2M ζ 1 p M sinh = (cosh ζ 1) = 0 − , (4.459) 2 s2 − s 2M 1 we can express the pure Lorentz transformations of the ( 2 , 0)-representation in terms of energy and momentum of the boosted particle as follows:

1 

−· /2  B( )= e = (p0 + M p). (4.460) 2M(p0 + M) − · q We may also use the Einstein parameter γ to write this as

1  B()= (γ +1 γ v/c). (4.461) 2(γ + 1) − · q It is convenient to introduce an extra 2 2 “Pauli matrix”: × 1 0 σ0 = , (4.462) 0 1 ! and define a four-vector of 2 2 matrices: × σµ (σ0, σi). (4.463) ≡ Note that the four Pauli matrices satisfy a multiplicative algebra: σ0σ0 = σ0, σiσ0 = σ0σi = σi, σiσj = iǫijkσk + δijσ0. (4.464) With σµ, we can form Lorentz-covariant matrices: pσ p σµ = pµσ . (4.465) ≡ µ µ 1 This notation allows us to write the boosts in the ( 2 , 0)-representation of the Lorentz group as

M + pσ  e−· /2 = . (4.466) 2M(p0 + M) q 1 For a corresponding expression in the (0, 2 )-representation, we define, by analogy withx ˜ (x0, xi) in Eq. (4.195), the space inverted vectors: ≡ − p˜µ (p0, pi), ≡ − σ˜µ = (σ0, σi). (4.467) − 4.11 Spin-1/2 Fields 303

Then

M +˜pσ M + pσ˜  e · /2 = = . (4.468) 2M(p0 + M) 2M(p0 + M) q q For many explicit calculations to follow it is useful to realize that the boost matrices (4.466) and (4.468) may be considered as the square-root of the same

expression with twice the rapidity: 

∓· ˆ  e = cosh ζ  sinh ζ. (4.469) ∓ · Because of (4.458), the right-hand sides have simple momentum representations

pσ pσ˜

   e−· = , e · = . (4.470) M M Using these, we may write the boost matrices shorter as

pσ pσ˜

   e−· /2 = , e · /2 = , (4.471) r M s M a notation which will be convenient in later calculations. Having thus succeeded in finding the smallest dimensional representations of the α˙ Lorentz group, we define fields ξα(x) and η (x) to have the corresponding transfor- mation laws:

Λ ′ ′ ( 1 ,0) β ξ (x) ξ (x )= D 2 (Λ) ξ (x), α −−−→ α α β Λ α˙ ′α˙ ′ (0, 1 ) α˙ β˙ η (x) η (x )= D 2 (Λ) ˙ η (x). (4.472) −−−→ β The different transformation behavior of the two kinds of spinors is exhibited by the lower undotted and upper dotted indices. These spinors were introduced in 1929 1 1 by Hermann Weyl [6] and are referred to as Weyl spinors of type ( 2 , 0) and (0, 2 ), respectively. Let us now see whether we can construct a Lorentz-invariant free-field action from Weyl spinors which can contain only quadratic terms in the fields and their first derivatives. First we look for invariant quadratic terms without spacetime derivatives, which are needed to describe particles with a mass. Let us begin by looking for suitable combinations of ξ and η, which are invariant only under the ( 1 ,0) (0, 1 )

rotation subgroup. For this, the representation matrices D 2 and D 2 are both  equal to the same unitary matrix U = e−i· /2 to be applied to both ξ and η. Due to the unitarity of U, all quadratic expressions ξ†ξ, η†η, η†ξ, ξ†η are rotationally invariant quadratic field combinations, for example,

ξ′†(x′)ξ′(x′) = ξ†(x)U †Uξ(x) = ξ†(x)ξ(x). (4.473) 304 4 Free Relativistic Particles and Fields

1 1  ( ,0) −· /2 (0, )

Consider now pure Lorentz transformations. Then D 2 = e and D 2 =  e· /2 are both nonunitary but Hermitian. The quadratic expressions ξ†ξ and η†η are no longer invariant. However, since the two representation matrices are inverse to each other, ( 1 ,0)† ( 1 ,0) (0, 1 )−1 D 2 = D 2 = D 2 , (4.474) the mixed quadratic expressions η†ξ and ξ†η are invariant field combinations. Thus we conclude: If the action should contain a quadratic field combination without 1˙ 2˙ spacetime derivatives, then both spinors ξ, η, i.e., four field components ξ1,ξ2, η , η , are needed to form an invariant. Let us now construct a Lorentz invariant term involving spacetime derivatives ∂µ. It is necessary if an action is supposed to describe a particle which can move through spacetime. Consider first the spatial derivatives. It is easy to see that

† i † i ξ σ ∂iξ, η σ ∂iη, † i † i ξ σ ∂iη, η σ ∂iξ (4.475) are all rotationally invariant. Take for example

†′ ′ i ′ ′ ′ † † i ′ ξ (x )σ ∂iξ (x )= ξ (x)U σ ∂i Uξ(x). (4.476)

From the commutation rules among the Pauli matrices σi it is easy to derive the transformation law −1 i i j

U σ U = R³ˆ (ϕ) jσ . (4.477) The proof of this proceeds in the same way as in the finite transformation of the spatial vector xi in (4.106) with the help of Lie’s expansion formula (4.105). The commutation rules between iσ3/2 and σi are precisely the same as those between i − Lˆ3 and x , so that we find for a rotation around the 3-axis,

3 3 eiσ ϕ/2σ3e−iσ ϕ/2 = σ3, 3 3 eiσ ϕ/2σ1e−iσ /2ϕ = cos ϕ σ1 sin ϕ σ2, 3 3 − eiσ ϕ/2σ2e−iσ ϕ/2 = sin ϕ σ1 + cos ϕ σ2, (4.478) corresponding precisely to (4.477) with the matrix (4.7). The 2 2 representation × has the advantage, used already in (4.456), that the exponentials can be expanded into linear combinations of σ0 and σi. If we do this on the left-hand sides, we can calculate the right-hand sides also using products of σi rather than commutators as in Lie’s expansion formula (4.105). Now, since derivatives transform like a vector,

R ′

∂ ∂ = R³ (ϕ) ∂ , (4.479) i −−−→ i ˆ ij j the expression (4.476) is indeed a rotationally invariant field combination. For the other terms in (4.475), the proof is the same. 4.11 Spin-1/2 Fields 305

How can we extend the expressions (4.475) to form relativistic invariants? For this we remember that in the boost matrices (4.471), the four-vector generalizations µ µ σ , σ˜ of the Pauli matrices appeared naturally contracted with pµ. This suggests studying

† µ † µ ξ σ ∂µξ, η σ ∂µη, † µ † µ ξ σ ∂µη, η σ ∂µξ, (4.480) and once more the same combinations, but with σµ replaced byσ ˜µ. The additional time derivatives in (4.480) are trivially invariant under rotations and thus do not destroy the rotational invariance of the spatial parts shown in (4.476). Consequently we have to study only the behavior under pure Lorentz transformations in, say, the z-direction. Under these, the x- and y-components do not change. It is easy to verify that

3 3 3 e−σ ζ/2σ0e−σ ζ/2 = e−σ ζ = cosh ζ σ0 + sinh ζ ( σ3), 3 3 3 − e−σ ζ/2( σ3)e−σ ζ/2 = e−σ ζ( σ3) = sinh ζ σ0 + cosh ζ ( σ3), −3 3 3 − − eσ ζ/2σ0eσ ζ/2 = eσ ζ = cosh ζ σ0 + sinh ζ σ3, 3 3 3 eσ ζ/2σ3eσ ζ/2 = eσ ζσ3 = sinh ζ σ0 + cosh ζ σ3. (4.481) In contrast to the transformation laws (4.478), this cannot be derived with the help of Lie’s expansion formula (4.105), since the exponentials on the left-hand sides have the same exponents. However, as before in the calculation of (4.457), we may use the multiplicative algebra of σi-matrices rather than the commutation rules between them to expand the exponentials into linear combinations of σ0 and σi. The first two lines in (4.481) show that the matricesσ ˜µ transform like a four- 1 µ vector under ( 2 , 0)-boosts, the remaining lines σ show the corresponding behavior 1 under (0, 2 )-boosts. In combination with rotations, we thus have proved the Lorentz transformation behavior:

( 1 ,0)† µ ( 1 ,0) µ ν D 2 (Λ)˜σ D 2 (Λ) = Λ νσ˜ , (0, 1 )† µ (0, 1 ) µ ν D 2 (Λ)σ D 2 (Λ) = Λ νσ . (4.482) This allows us to conclude that the quadratic field terms with a derivative

† µ † µ ξ σ˜ ∂µξ, η σ ∂µη (4.483) are Lorentz-invariant. For instance

1 1 ′† ′ µ ′ ′ ′ † (0, 2 )† µ ′ (0, 2 ) † µ ν ′ η (x )σ ∂µη (x ) = η (x)D (Λ)σ ∂µD (Λ)η(x)= η (x)Λ ν σ ∂µη(x) † ν = η (x)σ ∂ν η(x). (4.484) It is easy to see that the other quadratic combinations in (4.480) are not invariant. If we allow only for these lowest-order in the derivative terms, the most general Lorentz-invariant action reads

4 4 † µ † µ † † = d x (x)= d x (ξ σ˜ i∂µξ + η σ i∂µη M1ξ η M2η ξ). (4.485) A Z L Z − − 306 4 Free Relativistic Particles and Fields

Observe that this expression involves necessarily both two-component spinors ξ and η. Only for zero parameters M1,2, a single species of two-component spinors, ξ or η, possesses an invariant action. The equations of motion are obviously

µ iσ˜ ∂µξ(x) = M1η(x), µ iσ ∂µη(x) = M2ξ(x). (4.486)

Combining the second equation with the first and vice versa we find the two second- order field equations

( ∂2 M M )ξ(x) = 0, (4.487) − − 1 2 ( ∂2 M M )η(x) = 0. (4.488) − − 1 2 In deriving these, we have used the relation

σµσ˜ν + σνσ˜µ =2gµν, (4.489) by which µ ν 2 σ˜ ∂µσ ∂ν = ∂ , (4.490) which is easily shown by direct evaluation. In momentum space, equations (4.487) are solved by particles of mass

M = M1M2. (4.491) q For M = 0, it is useful to combine the two spinors ξ and η into a single four- 6 component object, called or :

ξ(x) ψ(x)= . (4.492) η(x) ! In terms of this, the action may be written as

M 0 = d4x (x)= d4x ψ¯(x) iγµ∂ 2 ψ(x), (4.493) µ 0 M A Z L Z " − 1 !# where γµ are the 4 4 Dirac matrices: × µ µ 0 σ γ = µ , (4.494) σ˜ 0 ! and ψ¯(x) is the adjoint Dirac spinor defined by

ψ¯(x) ψ†(x)γ0 =(η†(x),ξ†(x)). (4.495) ≡ The Dirac matrices satisfy the anticommutation rules

γµ,γν =2gµν. (4.496) { } 4.11 Spin-1/2 Fields 307

It has become customary to abbreviate the contraction of γµ with any vector vµ by /v γµv . (4.497) ≡ µ µ In this notation, the derivative terms γ ∂µ in (4.493) become simply /∂ . For the Dirac spinor, the equations of motion (4.486) take the form

M 0 i/∂ 2 ψ(x)=0. (4.498) " − 0 M1 !# This is almost, but not quite, the wave equation postulated first by Dirac for the . He assumed a diagonal 4 4 -mass matrix × M 0 1 0 2 = M , (4.499) 0 M1 ! 0 1 ! and proposed the (i/∂ M) ψ(x) = 0 (4.500) − corresponding to an action

= d4x (x)= d4x ψ¯(x)(i/∂ M) ψ(x). (4.501) A Z L Z − Just as in the case of a complex scalar field, there exists a four-vector jµ(x)= ψ¯(x)γµψ(x), (4.502) which by virtue of the Dirac equation satisfies a local conservation law [recall (4.171) and (4.172)]: µ ∂µj (x)=0. (4.503) This four-vector will be used in Chapter 12 to couple Dirac fields to electromag- netism, thus becoming the electromagnetic current density of the Dirac field. Later in Subsection 8.11.1 we shall see that the local conservation law (4.503) is a conse- quence of the invariance of the Dirac action (4.501) under arbitrary changes of the phase of the field: ψ(x) e−iαψ(x). (4.504) → It is this invariance which gives rise to a conserved current density [see (8.272), also (12.48)]. By construction, the actions (4.493) and (4.501) are invariant under the bispinor or Dirac-Lorentz transformations Λ ψ (x) ψ′ (x′)= D b(Λ)ψ (x), (4.505) a −−−→ a a b where the 4 4 -matrices D(Λ) consist of rotations and pure Lorentz transformations:

×

   e−i³· /2 0 e− · /2 0

D R ,D B .    ( )= −i³· /2 ( )= · /2 (4.506) 0 e ! 0 e ! 308 4 Free Relativistic Particles and Fields

With the 2 2 matrices (4.453) and (4.454) we can write D(Λ) in the form × ( 1 ,0) D 2 (Λ) 0 D(Λ) = (0, 1 ) , (4.507) 0 D 2 (Λ) ! The invariance can be seen most directly by combining the transformation laws (4.482) with the 4 4 -relation × µ µ ν D¯(Λ)γ D(Λ) = Λ νγ , (4.508) where we define

( 1 ,0) † ¯ D 2 (Λ) 0 −1 D(Λ) (0, 1 ) † = D (Λ). (4.509) ≡ 0 D 2 (Λ) !

From (4.508), the invariance of the derivative term in (4.501) follows at once, using (4.138), (4.139), and (4.44):

¯′ ′ µ ′ ′ ′ ¯ ¯ µ ′ ¯ µ ν ′ iψ (x )γ ∂µψ (x ) = iψ(x)D(Λ)γ D(Λ)∂µψ(x)= iψ(x)Λ νγ ∂µψ(x) ν = iψ¯(x)γ ∂ν ψ(x). (4.510)

In terms of Dirac matrices (4.494), the pure Lorentz transformations can also be written as M + /pγ0 D(B)= . (4.511) 2M(p0 + M) q The representation matrices of all Lorentz transformations may be expressed in terms of covariant generators as

−i 1 ω Sµν D(Λ) = e 2 µν , (4.512) where ωµν is the same antisymmetric matrix as in (4.69). They contain both the rotation and boost parameters as specified in (4.70) and (4.71). Taking the matrices

µν  (4.506) to the limit of small ³ and , we identify the 4 4 -matrix generators S as ×

k i ij 1 σ 0 0i i σ 0 S = ǫijk k , S = − i . (4.513) 2 0 σ ! 2 0 σ ! The 4 4 -generator of the rotation group on the left-hand side contains the direct × 4 4 -extension of the Pauli matrices: ×

 0

¦ . (4.514) ≡ 0  ! The spin matrix 1

S ¦ (4.515) ≡ 2 4.11 Spin-1/2 Fields 309

i 1 jk has the components S = 2 ǫijkS . The generator of the pure Lorentz transformations on the right-hand side of (4.513) is also written as S0i = iαi/2 with the matrix

 0

« = − . (4.516) 0  ! It is customary to introduce the matrices i σµν [γµ,γν]. (4.517) ≡ 2 In terms of these, equations (4.513) can be summarized as 1 Sµν σµν . (4.518) ≡ 2 It is easy to check that the matrices Sµν satisfy the same commutation rules as the defining generators of the Lorentz group in (4.87): [Sµν ,Sµλ]= igµµSνλ, no sum over µ. (4.519) − Thus they are generators of a new 4 4 -representation of the Lorentz group. The × 4 4 -Dirac representation matrices D(Λ) transforming are mathematically inequivalent× to the defining 4 4 -representation Λ of the Lorentz group transforming × vectors. With the , the generators Sµν have the following commutation rules: [Sµν,γλ]= (Lµν )λ γκ = i(gµλγν gνλγµ), (4.520) − κ − − which state that γµ is a vector operator in bispinor space [recall (4.99)]. The trans- formation law (4.508) is the global consequence of these rules. In terms of the generators Sµν, we can write the field transformation law (4.505) more explicitly as Λ ′ −1 −i 1 ω Sµν −1 ψ(x) ψ (x)= D(Λ)ψ(Λ x)= e 2 µν ψ(Λ x). (4.521) −−−→ Λ As in the case of the scalar field [recall (4.122)], it is useful to perform the trans- formation of the spacetime argument on the right-hand side in terms of the differ- ential operator of four-dimensional angular momentum. Thus we rewrite (4.521) as Λ ′ −i 1 ω Jˆµν ψ(x) ψ (x)= Dˆ(Λ)D(Λ)ψ(x)= e 2 µν ψ(x), (4.522) −−−→ Λ where Jˆλκ Sλκ 1+1ˆ Lˆλκ (4.523) ≡ × × are the generators of the total four-dimensional angular momentum of a Dirac field, by analogy with (4.293). The tensor products exhibits the two separate represen- tation spaces associated with the Dirac index and the spacetime coordinates. The commutation rules between the generators Jˆµν are of course the same as in the case of the vector potential (4.295), and in fact for any spin. 310 4 Free Relativistic Particles and Fields

4.12 Other Symmetries of the Dirac Action

As in the scalar case, the spin-1/2 action (4.501) is invariant under more than just the Lorentz group.

4.12.1 Translations and Poincar´eGroup First, it is automatically invariant under translations (4.185) for which

µ ψ(x) ψ′ (x)= ψ(x a). (4.524) −−−→ a − Together with the Lorentz transformations, these form the inhomogeneous Lorentz group or the Poincar´egroup (4.189). Under it, the spinor transforms as

Λ,a ψ(x) ψ′(x)= ψ(Λ−1(x a)). (4.525) −−−→ − Extending (4.522), we can generate all Poincar´etransformations by the operations

Λ,a ′ iaµpˆ /¯h −i 1 ω Jˆµν ψ(x) ψ (x)= Dˆ(Λ)D(Λ)ψ(x)= e µ e 2 µν ψ(x). (4.526) −−−→ 4.12.2 Space Inversion In contrast to scalar fields, the Poincar´einvariance does not automatically imply invariance under parity transformations. In the quantum theory of and photons called (QED), it is an additional requirement confirmed by all experimental data. The action of free electrons must therefore be invariant under parity transformations. To achieve this, we first have to define an appropriate way to transform a bispinor under space reflections P . The bispinor must form a representation of the Lorentz group extended by P ,

P ψ(x) ψ′ (x)= D(P )ψ(˜x), (4.527) −−−→ P ′ where xP =x ˜ as in (4.195). The representation matrices D(P ) must combine with the representation matrices of the Lorentz group D(Λ) in the same way as the 4 4 -matrix of space reflections × 1 1 P =   (4.528) − 1  −   1     −  combines with the Lorentz transformations Λ. From the explicit matrices (4.54)– (4.56) and (4.60)–(4.62) we find

P −1LP = L, P −1MP = M. (4.529) − 4.12 Other Symmetries of Dirac Action 311

Thus rotations commute with reflections, which is intuitively obvious since x and x rotate both with the same 3 3 matrices. Pure Lorentz transformations, on the − × other hand, are space-inverted to boost into the opposite direction. Since upper and lower components in a Dirac spinor contain the boost matrices (4.471) in opposite directions, we can immediately write down the transformation law for a Dirac spinor under space inversion as

P ′ 0 1 ψ(x) ψP (x)= ηP ψ(˜x). (4.530) −−−→ 1 0 !

The phase ηP is the intrinsic parity of the field ψ(x). The representation matrix of the reflection P in Dirac space is denoted by

0 1 D(P )= ηP = ηP γ0. (4.531) 1 0 !

The property P 2 = 1 must be reproduced by the representation matrix D(P ). By applying two successive space inversions on ψ(x) we can conclude that the intrinsic parity can only have the values η = 1. P ± We easily check that the matrix D(P ) transforms the 4 4 -bispinor represen- × tation (4.513) of the generators Li and Mi as in the defining representation (4.529):

D−1(P )D(L )D(P )= D(L ),D−1(P )D(M )D(P )= D( M ). (4.532) i i i − i We now postulate Lorentz invariance of the action (4.493) under space reflection. Since M 0 M 0 D−1(P ) 2 D(P )= 1 , (4.533) 0 M1 ! 0 M2 ! this is only possible if M1 = M2, so that the action takes the Dirac form (4.501). Also, since space inversion transforms ξ(x) into η(x), a parity-invariant theory nec- essarily contains both fields and thus the full bispinor ψ(x). Having set M1 = M2, the mass term of the Dirac action is parity-invariant. In order to ensure the invariance of the derivative term, we observe that the represen- tation matrix (4.531) of P satisfies

D−1(P )γµD(P )=˜γµ. (4.534)

Hence we calculate

¯′ ′ µ ′ ′ ′ ¯′ µ ˜ ′ ¯ −1 µ ˜ ψ (x )iγ ∂µψ (x ) = ψ (˜x)iγ ∂µψ (˜x)= ψ(x)D (P )iγ ∂µD(P )ψ(x) ¯ µ = ψ(x)iγ ∂µψ(x), (4.535) which proves the invariance of the derivative term, and thus of the full Dirac action. The Dirac equation (iγµ∂ M)ψ(x) = 0 (4.536) µ − 312 4 Free Relativistic Particles and Fields can trivially be rewritten as (iγµ∂˜ M)ψ(˜x)=0. (4.537) µ − ′ Using (4.530) and (4.531), we can replace ψ(˜x) by ηP γ0ψP (x), and take the matrix γ0 to the left of the Dirac operator with the help of (4.534). The result is (iγµ∂ M)ψ′ (x)=0, (4.538) µ − P i.e., the Dirac equation for the mirror-reflected bispinor. As we shall see in Chapter 12, the interactions of electrons with electromagnetism are described with extreme accuracy by the parity-invariant theory called quantum electrodynamics (QED). The electrons in this theory are described by the Dirac action (4.501). A theory based only on a single two-component spinor field ξ or η is neces- sarily massless and violates parity. Such a theory was used successfully to de- scribe neutrino processes. There exist several neutrinos in nature, one associated − − − with every charged lepton e , µ , τ whose masses are me = 0.510 MeV, mµ = 105.66 MeV, mτ = 1777.03 0.30 MeV, the latter two having finite lifetimes τ = (2.19703 0.00004) 10±−6 sec and τ = (290.6 1.1) 10−15 sec. The cor- µ ± × τ ± × responding neutrinos are denoted by νe, νµ, ντ . The six leptons seem to exist in nature by complete analogy to six which are the elementary building blocks of strongly interacting particles. The analogous configurations are illustrated in Fig. 4.1.

e− µ− τ − u c t νe νµ ντ ! d s b ! Figure 4.1 Six leptons and quarks.

The electron-neutrino was postulated in 1931 by Pauli in order to explain an apparent violation of energy conservation in the final state of the β-decay of the neutron. Energy conservation would have been violated if only the observed particles proton and electron emerged from the decay. From the energy spectrum of the electron one can deduce that the mass of the electron neutrino is extremely small, less than 2 eV/c2. The most precise value is expected from an ongoing experiment performed in Karlsruhe. There one studies the beta radiation of tritium nuclei which decay in 12.3 years with a total energy release of 18.6 keV shared by an electron and a neutrino. The energy of the electron is measured in an experiment called KATRIN (acronym for the Karlsruhe Tritium Neutrino Experiment) by a giant electrostatic spectrometer of diameter 7 m and length 20 m. After this the particle identity is confirmed in a semiconducting detector. The experiment is sensitive to electron neutrino masses 0.4 eV/c2. The masses of the other two neutrinos have presently ≥ the bounds mνµ < 0.19 MeV and mντ < 0.18.3 MeV. In summary, in a mirror-symmetric Lorentz-invariant theory of particles with the lowest nontrivial spin 1/2, the action is given by Dirac’s expression (4.501). If 4.12 Other Symmetries of Dirac Action 313 parity is allowed to be violated, there are two simpler Lorentz-invariant actions of massless particles: 4 4 † µ = d x (x)= d xξ (x)iσ˜ ∂µξ(x), (4.539) A Z L Z or

4 4 † µ = d x (x)= d x η (x)iσ ∂µη(x). (4.540) A Z L Z The parity-invariant action (4.501) is the correct one for electrons, while the action (4.539) describes neutrinos. For neutrinos, it has become customary to work also with four-component bispinors ψ(x), but making only use of the two upper compo- nents. The upper and lower components are extracted from ψ(x) by the projection matrices 1 1 0 0 0 Pu,l (1 γ5) , , (4.541) ≡ 2 ∓ ≡ 0 0 ! 0 1 ! where γ denotes the 4 4 -matrix 5 ×

5 0 1 2 3 1 0 5 γ5 γ = iγ γ γ γ = − γ . (4.542) ≡ 0 1 ! ≡

This is a Lorentz-invariant matrix, since it may also be expressed in the contracted form i γ = ǫ γµγν γλγκ γ5, (4.543) 5 4! µνλκ ≡ where ǫµνλκ is the completely with ǫ0123 = 1 (see [12]). Using γ5, the actions (4.539) and (4.540) can be written as

4 1 4 µ = d x (x)= d x ψ¯(x)iγ ∂µ(1 γ5)ψ(x). (4.544) A Z L 2 Z ∓

From (4.534) we see that under space inversion, γ5 transforms as follows:

D−1(P )γ D(P )= γ , (4.545) 5 − 5 thereby interchanging the two actions (4.544) with each other. The parity-violating actions (4.539) and (4.540), or (4.544), have an interesting history. After having been proposed by Weyl in 1929 to describe massless spin-1/2 particles [6] it was initially rejected on theoretical grounds, since at that time all interactions were firmly believed to be invariant under space reflections. Electro- magnetic and nuclear interactions had definitely displayed this property, and it was suggestive to assume that nature should follow the same principle in all its interac- tions. In 1956, however, Lee and Yang suggested that a violation of parity can be deduced from the existence of the two decay modes of the heavy mesons K0 and K+. The first decays into π0π0 with relative s-waves, the second into π+π+π− with both π+π+ and π−π+ in relative s-waves. Since the pion has negative parity the 314 4 Free Relativistic Particles and Fields violation is manifest. In 1957, the above authors pointed out the relevance of mea- suring the β-decay from a polarized nucleus [7]. If parity was an invariance of weak interactions, the distribution of electrons would have to be symmetric with respect to the direction of spin. Indeed, since the scalar product between spin, which is an axial vector, and the momentum vector is a pseudoscalar operator, its expectation value should vanish. In 1957, Madame Wu and collaborators [8] performed the historic experiment observing a nonzero up-down asymmetry in the distribution of 60 8 electrons coming from polarized 27Co (see Fig. 4.2).

60 27Co

Figure 4.2 Asymmetry observed in the distribution of electrons from the β-decay of 60 polarized 27Co.

The polarization of the sample was done by placing it into a strong magnetic field. By going to an extremely low , a sufficient population difference be- tween spin down and spin up was achieved that made the experiment display a clear violation of parity. In later experiments it was found that the violation is even maximal, in the sense that the unobserved neutrino emitted in the decay process can only have one polarization along its momentum direction, the other being com- pletely forbidden. A massless neutrino possesses no mirror image in nature and can 1 be described by a pure Weyl action with only a 2 (1 γ5)ψ field. We shall see in − 1 Chapter 7 that also a massless antineutrino is described by the field 2 (1 γ5)ψ. 60 − The initial 27Co-state has a spin s = 5 and intrinsic parity ηP = +1, thus P + 60 being a s = 5 -state. The β-decay transforms it into an excited state of 28Ni with spin-parity sP = 4+. As such, it is a so-called Gamow-Teller transition. In this transition it can be shown that only combinations of tensor and axialvector couplings contribute (T-A). The details will be explained in Chapter 27. A year later a crucial hypothesis was made by several authors [19] that the weak interactions are mediated by a specific combination of vector and axial vector couplings. This is the famous V-A hypothesis which eventually led to the present of weak and electromagnetic interactions (see Chapter 27). The bispinors 1 (1 γ5)ψ are eigenvalues of the matrix γ5. Their eigenvalues are 2 ∓ 1 or +1. They are called states of left or right , respectively. − There exist, of course, many equivalent representations of the Lorentz group extended by the discrete transformation of space inversion on spin-1/2 fields. Instead

8 A In nuclear physics the customary notation for a nucleus X is Z XN , where A is the nucleon number, Z the number of protons or the atomic number (also the charge number ), and N the number of neutrons. The last label is not really necessary since the name of the nucleus is specified uniquely by A and Z. 4.12 Other Symmetries of Dirac Action 315 of the bispinors ψ(x) transforming with the 4 4 -matrices (4.507), in which parity exchanges upper and lower two-component spinors× in ψ(x) via the matrix D(P ) of (4.531), consider symmetric and antisymmetric combinations ψ (x) S ψ(x), (4.546) D ≡ D where SD is the similarity transformation matrix 1 1 1 SD . (4.547) ≡ √2 1 1 ! − In the bispinor ψD(x), upper and lower components are eigenstates of opposite parity. These fields transform according to the 4 4 representation × −1 DD(Λ) = SDD(Λ)SD 1 ( 1 ,0) (0, 1 ) 1 ( 1 ,0) (0, 1 ) D 2 + D 2 (Λ) D 2 D 2 (Λ) 2 − 2 − =   1 1   1 1   . (4.548) 1 ( 2 ,0) (0, 2 ) 1 ( 2 ,0) (0, 2 ) 2 D D (Λ) 2 D + D (Λ)  − −        When boosting a massive particle from rest to momentum pµ this matrix becomes

explicitly, with (4.471) and (4.468),

      1 e−· /2 + e · /2 1 e− /2 e · /2 2 − 2 −

DD(B()) =      

       1 e−· /2 e · /2 1 e− · /2 + e · /2  − 2 − 2        1 pσ/M + pσ/M˜ 1 pσ/M pσ/M˜ = 2 − 2 −  1 qpσ/M qpσ/M˜  1 qpσ/M + qpσ/M˜   − 2 − 2  q q  q q   1 M + p p  = 0 · . (4.549) p  M + p0 ! 2M(M + p0) · q The Dirac matrices which ensure in this case the invariance of the action (4.493) are now

µ µ −1 1 0 0  γD = SDγ SD = , , (4.550) 0 1  0 ( − ! − !) 5 5 −1 0 1 2 3 0 1 γD = SDγ SD = iγDγDγDγD = . (4.551) 1 0 ! In terms of these, the boost transformation (4.549) takes the form 0 M + /p DγD DD(B)= , (4.552) 2M(p0 + M)

q −1 which is the same as the similarity-transformed SDD(B)SD of the boost matrix (4.511) in the chiral representation. The generators are

k i ij ij −1 1 σ 0 0i 0i −1 i 0 σ SD = SDS SD = ǫijk k , SD = SDS SD = i ,(4.553) 2 0 σ ! 2 σ 0 ! 316 4 Free Relativistic Particles and Fields which are equal to i σµν S σµν S−1 = [γµ ,γν ], (4.554) D ≡ D D 2 D D ij as in (4.517). The generators of rotations SD are the same as in the chiral represen- tation (4.513). Indeed, writing Sij = ǫ 1 Σi , we see that the 4 4 -generalization D ijk 2 D × of the Pauli matrices (4.514) is invariant under the similarity transformation SD:

−1  0

¦ ¦ ¦D = SD SD = = , (4.555) 0  !

ij 1 i so that we can write SD = ǫijk 2 Σ , as before in Eq. (4.515). For small momenta, the boost matrix (4.549) has the limit

1 p /2

DD(B( )) · . (4.556) ≈ p  /2 1 · ! This shows that the spinors in Dirac’s representation of the gamma matrices have small lower (or upper) components for slow particles (or antiparticles). The Dirac representation is therefore useful for studying the nonrelativistic limit of Dirac par- ticles. For such calculations it is advantageous to state the Dirac matrices in a direct- product form

0 3 2  γ = σ 1, ­ = iσ , (4.557) D × D × in which γ5 = σ1 1 and the generators of the Lorentz group (4.554) take the form D × σij = ǫ Σk = ǫ 1 σk, σ0i = iσ1 σi. (4.558) D ijk ijk × D × In each case, the first matrix mixes upper and lower components, whereas the second matrix acts on the up and down components of the spin. Actually, it is this representation of the Dirac matrices which was stated in his original paper [5] and later in many textbooks. This is why it is referred to as the standard representation. In it, the chirality matrix γ5 is not diagonal, as in the representation (4.542) of Section 4.11. To emphasize this property, the 4 4 - matrices (4.494) of Section 4.11 are referred to as chiral representation of the gamma× matrices (also called Weyl representation).

4.12.3 Dirac’s Original Derivation Note that Dirac did not find his matrices from group-theoretic considerations. In- stead of searching for a relativistic Schr¨odinger equation for an electron which, in contrast to the Klein-Gordon equation, contains only a single time derivative, so that there would be no negative-energy solutions [5], he looked for a time-independent electron field that satisfies the wave equation

Hψˆ (x)= pˆ2 + M 2ψ(x)= Eψ(x) (4.559) q 4.12 Other Symmetries of Dirac Action 317

with only the positive square-root. Since in relativistic theories energy and momen- tum appear on equal footing, he searched for a way to take an explicit square-root. For this he allowed ψ(x, t) to consist of several components, which would somehow represent the spin degrees of freedom of the electron. So he made the ansatz i i HˆDψ(x)=(α pˆ + βM)ψ(x)= Eψ(x), (4.560)

with αi, β being unknown matrices. Then he required that applying HˆD twice to 2 2 2 ψ(x) should give HˆDψ(x)=(pˆ +M )ψ(x)= E ψ(x). This led him to the algebraic relations i j α ,α = δij, n o αi, β = 0, (4.561) n βo2 = 1. He solved these anticommutators with the matrices

0 1 0 0 0 

« ­ β = βD γD = , « = D γD D = . (4.562) ≡ 0 1 ≡  0 − ! ! He could, of course, have solved them just as well in the chiral representation by

0 0  0 ­ β γ , « γ = − . (4.563) ≡ ≡ 0  ! By multiplying Eq. (4.560) with β and going over to a time-dependent equation by replacing E byp ˆ0 = i∂x0 , he obtained the Dirac equation in the form (γµ pˆ M)ψ(x) = 0 (4.564) D µ − with the matrices γ0 β, γi βαi. (4.565) D ≡ D ≡ The anticommutation relations (4.561) go over into the anticommutation relations (4.496) for the Dirac matrices γµ ,γν =2gµν. (4.566) { D D} Inserting the bispinor form (4.492) of ψ(x) into the Dirac equation (4.564), we find for upper and lower components ξ(x) and η(x) the equations

i∂ ξ(x)+ i ∇η(x)= Mξ(x), t ·

i∂ η(x)+ i ∇ξ(x)= M η(x). (4.567) t · − They can be combined to a single bispinor equation

(∂ + « ∇ + iMβ )ψ(x)=0. (4.568) t D · D The equations (4.567) should be contrasted with their chiral versions, where they are given by the two lines in (4.486) for M1 = M2 = M:

i∂ ξ(x) i ∇ξ(x) = Mξ(x), t − ·

i∂ η(x)+ i ∇η(x) = M η(x). (4.569) t · 318 4 Free Relativistic Particles and Fields

4.12.4 Maxwell Equations Written `ala Dirac It is interesting to note that the Maxwell equations (4.246) and (4.247) can be

brought to a similar form using a spin-1 version of the matrix «D:

0 L

«M , (4.570) ≡ L 0 ! where (Li)jk = iǫijk are the generators (4.57) of rotation for a vector field. The analog of the Dirac− bispinor is the “Maxwell bivector”:

E ψM . (4.571) ≡ iB !

The analogs of the two separate spinor equations (4.567) are the equations:

∂ E + L ∇(iB) = 0, t · ∂ (iB)+ L ∇ E = 0. (4.572) t · These coincide with the Maxwell equations (4.246) and (4.247). Note, however, that the bivectors (4.571) cannot be used to set up an action analog to Dirac’s (4.493) for zero mass. That must involve the local vector potential Aµ. A formulation which incorporates the dual symmetry between and magnetism is nevertheless possible by defining the two bivector components 1 = E, B 2 = B, and an associated pair of vector potentials a (a = 1, 2), whose two- dimensionalB curls are these fields: A

a =(∇ )a. (4.573) B ×A

The curls are formed with the Levi-Civita tensor ǫab = ǫba and ǫ12 = 1. Then we can write the Maxwell action as − 1 = d4x a ǫ ∂ b δ b . (4.574) A 2 B ab tA − abB Z   Indeed, a variation of this action does yield the Maxwell equations since

δ = d4x δ b ǫ ∂ b δ b , (4.575) A B ab tA − abB Z   which, after an integration by parts, becomes

δ = d4x δ a ǫ ∂ b δ ∇ b . (4.576) A A ab tB − ab × B Z   The vanishing of the variation δ produces the equations A ∂ B + ∇ E =0, ∂ E ∇ B =0, (4.577) t × t − × which are precisely the Maxwell equations (4.246) and (4.247) [20]. 4.12 Other Symmetries of Dirac Action 319

These are invariant under the duality transformation δE = αB, δB = αE, (4.578) − and so is the action (4.576). The second of the duality transformations (4.578) corresponds to a nonlocal transformation of the vector potential: δA = α∇ (∇2)−1E. (4.579) × When calculating the small-momentum limit (4.556) we have noted that the Dirac representation is most convenient for studying the nonrelativistic limit. This limit, in which Mc2 , corresponds in natural units to letting M . The energies of slowly moving→ ∞ particles are very close to M, so that (i∂ →M ∞)χ(x) is t − much smaller than (i∂t +M)η(x), which can be approximated by 2Mη(x). The lower equation in (4.567) can therefore be solved approximately by the relation between lower and upper spinor

 ∇ η(x) i · ξ(x). (4.580) ≈ − 2M We can also remove the fast temporal oscillations as in (4.156) and replace

ξ(x) 2 1 Ψ(x) e−iMc t/¯h . (4.581) η(x) ! → √2M Φ(x) ! If we solve again the lower equation by a relation like (4.580), the upper equation reduces to the Schr¨odinger equation for each spinor component:

1 2

i∂ Ψ(x)= ( ∇) Ψ(x). (4.582) t −2M · Using 2 i j ij k 2

( ∇) = σ σ = δ + iǫ σ = ∇ , (4.583) · ∇i∇j ijk ∇i∇j this becomes h i 1 i∂ Ψ(x)= ∇2Ψ(x). (4.584) t −2M In the presence of electromagnetic interactions, the last step is nontrivial, yielding the nonrelativistic (6.114) with the correct magnetic moment of a Dirac particle. In both representations, we can insert one equation into the other and find that ξ(x) and η(x) satisfy the Klein-Gordon equations (4.487) with M1 = M2 = M: ( ∂2 M 2)ξ(x) = 0, − − ( ∂2 M 2)η(x) = 0. (4.585) − − This follows simply from the Dirac equation (4.500) upon multiplication by (i/∂ + M) and working out

(i/∂ + M)(i/∂ M) ψ(x)= /∂ 2 M 2 ψ(x)= ∂2 M 2 ψ(x). (4.586) − − − − −     320 4 Free Relativistic Particles and Fields

In the massless case, the Dirac equations (4.567) have a very similar structure to Maxwell’s equations (4.246) and (4.247):

∂ B + ∇ E = 0, t × ∂ E ∇ B = 0. (4.587) t − × To see this similarity we rewrite the cross product with the help of the 3 3 - generators (4.57) of the rotation group, (L ) = iǫ , where they read × i jk − ijk i∂ E + i(L ∇)(iB) = 0, t · i∂ (iB)+ i(L ∇) E = 0, (4.588) t · thus becoming quite similar to the Dirac equations (4.567) derived from Dirac’s representation of γ -matrices. The reader is encouraged to discuss the analogy between the transformation properties of (ξ, η) in Dirac’s representation and (E, iB)

and the generators  and L.

4.12.5 Pauli-Villars Equation for the Klein-Gordon Field It is worth mentioning that Dirac’s procedure of deducing a matrix version of the relativistic Schr¨odinger equation (4.559) has other solutions that is not linear in the momentum as HˆD in (4.560), for example:

pˆ2 Hˆ = (σ + iσ )+ Mσ . (4.589) FV 2M 3 2 3 The subscript FV indicates that this Hamiltonian was first proposed by Feshbach and Villars [17]. Here Pauli matrices have no relation to spin. They are merely employed to specify the 2 2 -matrix HˆFV. By using the multiplication rules (4.464), × ˆ 2 ˆ 2 2 2 it is easy to verify that the 2 2 -matrix HFV has the same square HFV = pˆ + M as the 4 4 Dirac matrix H× of Eq. (4.560). Thus the solutions of the equation × D HˆFVψ(x) = Eψ(x) have again the proper relativistic energy-momentum relation. However, contrary to the solutions of the Dirac equation (4.560), they carry no spin. In fact, a field theory based on the Lagrangian density

= ψ∗(x, t)(i∂ Hˆ )ψ(x, t) (4.590) L t − FV is completely equivalent to the Klein-Gordon theory of scalar particles.

4.12.6 Charge Conjugation In Section 4.5.4 we observed that the action of a scalar field was invariant under an extra discrete symmetry not related to the Lorentz group, namely charge conjuga- tion. It consisted of a simple exchange of the scalar field by its complex conjugate. A similar invariance can be found for the action of the Dirac field. There is only one complication: We must make sure that this operation commutes with the Lorentz 4.12 Other Symmetries of Dirac Action 321 group. Thus we must form linear combinations of the components of the conjugate bispinor ψ∗(x) which transform again like the original bispinor ψ(x). Let us call this new bispinor ψc(x) Cψ¯T (x), (4.591) ≡ where the superscript T on the right-hand side indicates a transposition of the row vector ψ¯, which makes ψc a column vector. The operation of charge conjugation is then defined by C ψ(x) ψ′ (x)= η ψc(x), (4.592) −−−→ C C with a phase η = 1. (4.593) C ± ′ The matrix C is determined by the requirement that ψC (x) must satisfy the Dirac equation: (i/∂ M)ψ′ (x)=0. (4.594) − C Inserting the right-hand side of (4.592), this reads (i/∂ M)Cψ¯T (x)=0, (4.595) − or [iC−1γµC∂ M]ψ¯T (x)=0. (4.596) µ − Its transposed form is ← ψ¯(x)[i(C−1γµC)T ∂ M]=0. (4.597) µ − Consider, on the other hand, the adjoint of the Dirac equation (iγµ∂ M)ψ(x)=0, (4.598) µ − which is ← ψ†(x)( i㵆 ∂ M)=0. (4.599) − µ − Multiplying this by γ0 from the right and using the fact that

−1 µ† 0 µ (γ0) γ γ = γ , (4.600) we see that ← ψ¯(x)( iγµ ∂ M)=0. (4.601) − µ − ′ Comparing this with (4.597) we conclude that ψT (x) satisfies the Dirac equation if the matrix C fulfills the identity C−1γµC = γµT . (4.602) − In both the chiral and the Dirac representation, the transposition of γµ changes only the sign of γ2. A matrix C with this property in the chiral representation is given by c 0 C = , (4.603) 0 c − ! 322 4 Free Relativistic Particles and Fields where c is the 2 2 matrix × 0 1 c = iσ2 = − . (4.604) − 1 0 ! This matrix is the two-dimensional representation of rotation around the 2-axis by an angle π: 2 c = e−iπσ /2, (4.605) as can easily be verified by using (4.456) [or by a direct power series expansion as in (4.455)]. From this rotation property it follows directly [or via Lie’s expansion formula (4.105) as in (4.478)] that

σ1 σ1 −1 2 − 2 c  σ  c =  σ  , (4.606) σ3 σ3        −  and we find c−1 0 0 σµ c 0 0 c−1σµc = − 0 c−1 σ˜µ 0 0 c c−1σ˜µc 0 − ! ! − ! − ! = ( γ0,γ1, γ2,γ3)= γµT , (4.607) − − − so that (4.602) is fulfilled. Note that the 2 2 -matrix c satisfies the identities × c = cT = c−1 = c†, (4.608) − − − which also hold for the 4 4 -matrices C: × C = C∗ = CT = C−1 = C†. (4.609) − − − Using these properties, we find that the conjugate Dirac field behaves under the transformation (4.592) as

C ψ¯(x) ψ¯′ (x)= η∗ ψ¯ c(x), (4.610) −−−→ C − C with ψ¯ c(x) ψT (x)C. (4.611) ≡ This follows from the simple calculation:

ψ¯ = ψ∗T γ0 (Cψ¯T )∗T γ0 =(Cγ0T ψ∗)∗T γ0 = ψT γ0 CT γ0 = ψT C = ψ¯ c. (4.612) → − − Note that ψ¯ c(x)= ψc(x), (4.613) since

ψ¯ c = ψc†γ0 =(Cψ¯T )†γ0 =(ψ¯T )†C†γ0 =(ψ†γ0)T †C†γ0 = ψT γ0T †C†γ0 = ψT C. (4.614) 4.12 Other Symmetries of Dirac Action 323

The minus sign on the right-hand side of Eq. (4.610) will be seen in Chapter 7 to have the important consequence that antiparticles have the opposite intrinsic parity of particles. By writing the charge conjugation matrix (4.603) as

C = iγ0γ2, (4.615) we can take the result directly to the Dirac representation (4.550) where9

c 0 0 c C iγ0 γ2 = S S−1 = . (4.616) D ≡ − D D − D 0 c D c 0 − ! ! The reason for the name charge conjugation is the same as for the scalar field in (4.227). In contrast to the scalar case, however, this cannot simply be seen by studying the effect of charge conjugation upon the conserved particle current. In contrast to Eq. (4.229) which shows that the current reverses its sign under charge conjugation, the operation (4.591) with C satisfying (4.602) leaves the current density (4.502) unchanged:

C jµ(x) jµ′(x)= jµ(x). (4.617) −−−→ This follows directly from

C jµ(x) = ψ¯(x)γµψ(x) ψT (x)C−1γµCψ¯T (x) −−−→− = ψT (x)γµT ψ¯T (x)= ψ¯(x)γµψ(x)= jµ(x). (4.618)

The proper physical effect will only be reached after field quantization. This turns the fields into anticommuting fermion operators which produce a sign change in the last step of the transformation (4.618), thus justifying the name charge conjugation for the operation (4.591). It is possible to imitate this effect of quantization at the classical level by imag- ining the classical fields to be anticommuting or Grassmann variables. Such fields will be introduced in Chapter 14 and used in Chapter 25.

4.12.7 Time Reversal Let us now see how time reversal acts upon the Dirac field. Under time reversal, the direction of a particle momentum and angular momentum are both reversed, and the generators of the Lorentz group are subject to an automorphism (4.132). The same automorphism is now applied to the 4 4 bispinor representation D(T ). × Writing D(T )= ηT DT , we must have

D−1D(L )D = D(L )∗,D−1D(M )D = D(M )∗. (4.619) T i T − i T i T − i 9 The minus sign is added to agree with Dirac’s sign convention for CD. 324 4 Free Relativistic Particles and Fields

The explicit form of the transformation matrix DT is now determined by the ′ requirement that the time-reversed field ψT (x) defined by

T ψ(x) ψ′ (x)= D(T )ψ∗(x ), (4.620) −−−→ T T with x = x˜ has to satisfy the Dirac equation T − (iγµ∂ M)ψ′ ∗(x)=0. (4.621) µ − T The reason for the complex conjugation of the field on the right-hand side of (4.620) was discussed in Subsection 4.5.3, where it was shown that the Schr¨odinger equa- tion for the time-reversed Schr¨odinger operator carries a complex conjugation [see Eq. (4.222)]. This is needed to keep the energy in the time-dependent phase factor eip0t positive for t t. Inserting (4.620)→ into − (4.621) we obtain

D−1(T )(iγµ∂ M)D(T )ψ∗( x˜)=0. (4.622) µ − − From the original Dirac equation we know that

( iγ˜µ∂ M)ψ( x˜)=0, (4.623) − µ − − or (iγ˜µ∗∂ M)ψ∗( x˜)=0. (4.624) µ − − To be compatible with (4.622), the matrix D(T ) has to satisfy

D−1(T )γµ∗D(T )=˜γµ. (4.625)

In both the chiral and the Dirac representation, the γµ-matrices have the property

γµ∗ =γ ˜µT . (4.626)

Using the property (4.602) of the matrix C, we can substitute

γµT = CγµC−1, (4.627) − and the condition (4.625) becomes

D−1(T )Cγ˜µC−1D(T )= γ˜µ. (4.628) − This is satisfied by D(T )= ηT Cγ5. (4.629) It is easy to verify that this matrix transforms the generators of the Lorentz group for Dirac spinors (4.518) in the way required by (4.619):

D−1(T )SµνD(T )= Sµν∗. (4.630) − 4.12 Other Symmetries of Dirac Action 325

4.12.8 Transformation Properties of Currents An important role in interacting field theory is played by bilinear combinations of the Dirac field formed with 16 combinations of Dirac matrices, collectively called Γ, which are all selfadjoint under the Dirac conjugation (4.495):

† Γ=¯ γ0Γ γ0. (4.631)

These are the scalar, vector, tensor, axialvector, and pseudoscalar matrices: i Γ 1, Γµ γµ, Γµν σµν = [γµ,γν], Γµ γµγ , Γ iγ , (4.632) S ≡ V ≡ T ≡ 2 A ≡ 5 P ≡ 5 which form a so-called Clifford algebra. They are used to define corresponding current densities. The most important of these is the vector current density

µ ¯ µ jV (x)= ψ(x)γ ψ(x), (4.633) which is the source of electromagnetism (see Chapter 12). By sandwiching the other Γ-matrices between two Dirac fields, one obtains fields which transform under the Lorentz group as scalar, tensor, axialvector, and pseudoscalar fields. For instance,

µ ¯ µ jA(x)= ψ(x)γ γ5ψ(x) (4.634) is an axial vector current which, together with the vector current, is responsible for weak interactions. The combination i jµν(x)= ψ¯(x)[γµ,γν]ψ(x) (4.635) T 2 is a tensor current related to the current spin density to be introduced in Sec- tion 8.6.2. The different possible current densities are shown in Table 4.12.8, which also lists the behavior of these currents under the discrete transformations T,C,P , and their various combinations.

¯ µ Table 4.1 Transformation properties of various composite fields jS = ψψ, jV = ¯ µ µν ¯ i µ ν µ ¯ µ ¯ ψγ ψ, jT = ψ 2 [γ , γ ]ψ, jA = ψγ γ5ψ, P = ψiγ5ψ. The wiggles on vectors and tensors denote the parity transformed objects for each index. For the charge-conjugated composites we have inserted the minus-sign arising after second quantization explained after Eq. (4.618). µ jS(x) jV (x) jT (x) jA(x) jP (x) ¯ ¯ µ ¯ i µ ν ¯ µ ¯ ψψ ψγ ψ ψ 2 [γ ,γ ]ψ ψγ γ5ψ ψiγ5ψ P jS(˜x) ˜jV (˜x) ˜jT (˜x) ˜jA(˜x) jP (˜x) C j (x) j (x) j (x) −j (x) −j (x) S − V − T A P T jS( x˜) ˜jV (˜x) ˜jT ( x˜) ˜jA( x˜) jP ( x˜) P CT j (−x) j ( x) −j ( −x) j −( x) −j ( −x) S − − V − T − − A − P − 326 4 Free Relativistic Particles and Fields

4.13 Majorana Fields

In the chiral and Dirac representations of γµ-matrices used so far, the bispinor fields ψ(x) are necessarily complex since only σ2, and thus γ2, is imaginary, whereas σ0, σ1, σ3 and thus γ0,γ1,γ3 are real. One may then wonder whether the Dirac equa- µν tion (i/∂ M)ψ(x) = 0, and thus the Lorentz transformations e−iωµν S , necessarily mix real− and imaginary parts of a spin-1/2 field. It can easily be seen that this is not so. The complex conjugate Dirac fields are transformed by the 4 4 -representation matrices ×

( 1 ,0)∗ ∗ D 2 0 D (Λ) = (0, 1 )∗ . (4.636) 0 D 2 !

( 1 ,0)∗ (0, 1 )∗ As far as rotations are concerned, D 2 and D 2 are equivalent to the original representations by a similarity transformation:

( 1 ,0)∗ −1 ( 1 ,0) D 2 = c D 2 c, (0, 1 )∗ −1 (0, 1 ) D 2 = c D 2 c, (4.637) with c = iσ2. This follows directly by writing the 2 2 rotation matrices in the − × explicit form

∗ ∗

 ³  e−i³· /2 = ei · /2. (4.638)   The complex conjugation reverses the 1- and 3-components in the exponent, since σ1, σ3 are real, while preserving the 1-component, since σ2 is imaginary. Using (4.606) we see that

−1 ∗  c c = , (4.639) − so that the right-hand side of (4.638) becomes

 ³  ei³· = c−1e−i · /2c, (4.640) which is the same as (4.637). Therefore, the charge-conjugated bispinor

T 0 c ψc(x) Cψ¯T (x)= Cγ0 ψ∗(x)= − ψ∗(x) (4.641) ≡ c 0 ! transforms under rotations just as ψ(x) itself. Consider now pure Lorentz transformations of the complex-conjugate bispinor ψ(x):

∗  B −· /2 ∗ ∗′ ′ e 0 ∗

ψ x ψ x ∗ ψ x .  ( ) ( )= · /2 ( ) (4.642) −−−→ 0 e !

With (4.639), the right-hand side becomes  c−1e· /2c 0

ψ∗ x .  −1 −· /2 ( ) (4.643) 0 c e c ! 4.13 Majorana Fields 327

Writing ψ(x) as in (4.492) we see that the upper complex conjugate components cξ∗(x) transform like the lower components η, whereas the lower components cη∗(x) transform like ξ(x). Hence also under Lorentz transformations, ψc(x) behaves like ψ, and we can write for the entire proper Lorentz group the transformation law

ψc′(Λx)= D(Λ)ψc(x), (4.644) with the transformation matrix D(Λ) satisfying the relation

C−1γ0−1D∗(Λ)γ0C = C−1D¯(Λ)C = D(Λ) (4.645)

[recalling the definition of Dirac-adjoint matrices (4.495)]. Since ψ and ψc both transform in the same way under D(Λ) we may form the combinations 1 χ (ψ + ψc), ≡ √2 1 χ′ (ψ ψc), (4.646) ≡ √2i − which are separately irreducible representations of the Lorentz group and eigenstates of charge conjugation with charge parity η . Since the original field had 4 complex ± C degrees of freedom, these combinations can only have half as many degrees of free- dom, i.e., four real degrees of freedom. Explicitly, the components of the bispinors (4.646) satisfy, in the chiral representation, the relations:

χ∗ = χ , χ∗ = χ , (4.647) 1 − 4 2 3 χ′ = χ′ ∗, χ′ = χ′ ∗. (4.648) 1 − 4 2 3 We may now ask whether there are γ-matrices which make these real degrees of freedom explicit. This would be the case if we would find a representation of the γ-matrices in which Cγ0 is the unit matrix. Then ψc would be equal to ψ∗ and the fields χ, χ′ would be purely real. Such a representation does indeed exist. It is given by the γ-matrices in the so-called Majorana representation:

2 3 0 0 σ 1 iσ 0 γM = 2 , γM = 3 , σ 0 ! 0 iσ ! 0 σ2 iσ1 0 γ2 = − , γ3 = − . (4.649) M σ2 0 M 0 iσ1 ! − ! They are obtained from γµ in the chiral representation (4.494) by a similarity trans- formation µ µ −1 γM = SMγ SM , (4.650) with the transformation matrix 1 1 σ2 1+ σ2 SM = − . (4.651) 4 1+ σ2 1+ σ2 − ! 328 4 Free Relativistic Particles and Fields

µ The action expressed with Majorana matrices γM is invariant under Lorentz trans- formations −1 DM(Λ) = SM D(Λ)SM . (4.652) In the Majorana representation (4.649), all γ-matrices are purely imaginary, so that the Dirac equation (iγµ ∂ M)χ(x) = 0 (4.653) M µ − is purely real. The complex conjugate field satisfies the same equation as χ itself:

(iγµ ∂ M)χ∗(x)=0. (4.654) M µ − A matrix C of complex-conjugation satisfying (4.602) is now given by

2 0 0 σ 0 c CM = γM = 2 = i . (4.655) σ 0 ! c 0 !

2 In contrast to the other two representations, the normalization is CM = 1 rather than C2 = 1, satisfying − C = CT = C−1 = C† (4.656) M − M M M rather than (4.609). This is more convenient here since we want two successive applications of the operations (4.641) to produce the identity operation (ψc)c = 0T 2 ψ. That requires (CM γ ) = 1. In the other two representations of the gamma matrices where C anticommutes with γ0, one has C2 = 1. In the Majorana 0 2 − representation where CM and γ commute, one has CM = 1. Note that up to a factor i, the matrix CM happens to coincide with CD of Eq. (4.616). It should be pointed out that CM is not related to C by a similarity transfor- −1 0 2 µ mation SMCSM = iγMγM, since γM does not have the same sign changes under µ µ µ T 0 1 2 3 transposition as γ and γD: Whereas γ = (γ , γ ,γ , γ ) holds also for the Dirac matrices γµ , the Majorana matrices satisfy γ−µ T =( −γ0,γ1,γ2,γ3). D M − According to (4.936), Cγ0 is equal to the unit matrix, so that

ψC = ψ∗. (4.657)

The bispinors (4.646) reduce to 1 1 χ (ψ + ψ∗), χ′ (ψ ψ∗), (4.658) ≡ √2 ≡ √2i − which are now real fields transforming irreducibly under the Lorentz group. They are called Majorana spinors. Under the operation of charge conjugation they transform into themselves

C χ(x) χ′ (x)= η χ(x), (4.659) −−−→ C C with a charge parity η = 1. C ± 4.13 Majorana Fields 329

µ Note that between Majorana spinors χ(x), the quadratic expressionsχγ ¯ Mχ and µ ν χγ¯ MγMχ are identically zero. The γ5-matrix (4.542) has now the form σ2 0 c 0 γ = iγ0 γ1 γ2 γ3 = = i γ5 . (4.660) 5M M M M M 0 σ2 0 c ≡ M − ! − ! In recent years solutions of the Majorana type have become relevant for describ- ing electrons in condensed matter and this has led to a resurgence of applications of this subject [9].

4.13.1 Plane-Wave Solutions of Dirac Equation By analogy with the scalar case we now seek for all plane-wave solutions of the Dirac equation (4.500): (i/∂ M)ψ(x)=0. (4.661) − We make an ansatz −ipx ipx e c e fp s3 (x) u(p,s3) , fp s (x) v(p,s3) , (4.662) ≡ Vp0/M 3 ≡ Vp0/M q q thereby distinguishing, as in (4.180), waves with positive and negative frequencies, and allowing for a spin orientation index s3. Due to the presence of Dirac indices, the solutions will no longer be merely the complex-conjugates of each other, as c in (4.152). The superscript of fp s3 (x) indicates the appropriate generalization of complex conjugation. If the wave functions in (4.662) are supposed to solve the Dirac equation (4.661), the bispinors u(p,s3) and v(p,s3) in momentum space have to satisfy the Dirac equations in momentum space (/p M)u(p,s )=0, (/p + M)v(p,s )=0. (4.663) − 3 3 The normalization of these wave functions will be chosen as in the scalar case by requiring the charge (4.175) of these solutions to be of unit size, with the charge density j0(x) of Eq. (4.502). In Section 4.4 we have introduced scalar products for solutions of the Klein- Gordon equation (4.177) with the help of the zeroth component of the conserved particle current. This is generalized to the Dirac case by introducing the scalar products

3 0 (f ′ ′ , f ) d x f¯ ′ ′ (x, t)γ f (x, t)= δ ′ δ ′ , p s3 p s3 p s3 p s3 p ,p s3,s3 ≡ Z c c 3 ¯c 0 c ′ ′ (f ′ ′ , f ) d x f ′ ′ (x, t)γ f (x, t)= δp ,pδs ,s , p s3 p s3 p s3 p s3 3 3 ≡ Z c 3 ¯c 0 (f ′ ′ , fp s ) d x f ′ ′ (x, t)γ fp s (x, t)=0, p s3 3 p s3 3 ≡ Z c 3 0 c (f ′ ′ , f ) d x f¯ ′ ′ (x, t)γ f (x, t)=0. (4.664) p s3 p s3 p s3 p s3 ≡ Z 330 4 Free Relativistic Particles and Fields

From these we deduce the orthonormality conditions for the bispinors:

0 0 ′ 0 p ′ 0 p u¯(p,s )γ u(p,s3) = δs′ ,s , v¯(p,s )γ v(p,s3) = δs′ ,s , 3 M 3 3 3 M 3 3 u¯(p,s′ )γ0v( p,s ) = 0, v¯( p,s′ )γ0u(p,s ) = 0. (4.665) 3 − 3 − 3 3

The reversal of the momentum in v( p,s3) appears in the second line since the spatial integrals in (4.665) enforce opposite− momenta in scalar products between solutions of positive and negative frequency. According to this, vanishing scalar products in bispinor space are necessary to produce orthogonality. In contrast to the scalar product (4.177) for Klein-Gordon wave functions, both positive- and negative-frequency solutions have now a positive charge, since for any spinor, ψ¯(x)γ0ψ(x)= ψ†(x)ψ(x) is positive definite. The explicit form of the bispinors u(p) and v(p) depends on the representation employed for the matrices γµ. The different cases will be discussed separately. In an infinite volume we use plane wave functions analogous to (4.181):

−ipx c ipx fp (x) u(p,s )e , f (x) v(p,s )e . (4.666) s3 ≡ 3 p s3 ≡ 3 They satisfy the Lorentz-invariant orthonormality conditions:

3 0 0 ′ (f ′ ′ , f ) d x¯f ′ ′ (x, t)γ f (x, t)=2p δ-(p p)δ ′ , p s3 p s3 p s3 p s3 s3,s3 ≡ Z − c c 3 ¯c 0 c 0 - ′ ′ (f ′ ′ , f ) d x f ′ ′ (x, t)γ f (x, t)=2p δ(p p)δs ,s , p s3 p s3 p s3 p s3 3 3 ≡ Z − c 3 ¯c 0 (f ′ ′ , fp s ) d x f ′ ′ (x, t)γ fp s (x, t)=0, p s3 3 p s3 3 ≡ Z c 3 0 c (f ′ ′ , f ) d x¯f ′ ′ (x, t)γ f (x, t)=0. (4.667) p s3 p s3 p s3 p s3 ≡ Z Spinors in Chiral Representation Using the chiral representation (4.494) for γµ, Eqs. (4.663) take the form

0 pσ 0 pσ u(p,s3)= Mu(p,s3), v(p,s3)= Mv(p,s3). (4.668) pσ˜ 0 ! pσ˜ 0 ! − We can immediately write down 4 2 -matrices solving these equations. The first × is solved by

1 pσ u(p)= M , (4.669) √  q pσ˜  2 M  q  and the second by

1 pσ v(p)= M . (4.670) √2  q pσ˜  − M  q  4.13 Majorana Fields 331

This follows from the matrix identities

pσ˜ pσ 2 pσ˜ pσ pσ pσ˜ pσ M = M (4.671) s M ≡ r M s M r M s M M and 1 pσpσ˜ = (p p σµσ˜ν + p p σ˜µσν )= p p = M 2, (4.672) 2 µ ν µ ν µ ν the latter being a direct consequence of (4.489). The 4 2 -matrices (4.669) and (4.670) can be multiplied by an arbitrary 2 2- matrix from× the right, and they will still solve the equations (4.668). There× are several convenient choices for such a matrix with different advantages, as we shall see below. The two-column vectors in the 4 2 -matrices form independent bispinor solutions of Eqs. (4.668). The projection into× these is accomplished by multiplication from the right with two unit spinors, the Pauli spinors (4.446)

1 1 1 0 χ( 2 )= , χ( 2 )= . (4.673) 0 ! − 1 ! By multiplying the 4 2 -matrices (4.669) with the unit spinors (4.673), we obtain × the canonical bispinors

1 pσ u(p,s )= M χ(s ). (4.674) 3 √  q pσ˜  3 2 M  q  3 The unit spinors (4.673) are eigenvectors of the spin-1/2 generator L3 = σ /2 of the rotation group: L χ( 1 )= 1 χ( 1 ), L χ( 1 )= 1 χ( 1 ). (4.675) 3 2 2 2 3 − 2 − 2 − 2 The associated bispinors at rest:

1 0 1 χ( 1 ) 1 0 1 χ(− 1 ) 1 1 1 2   1 2   u(0, 2 )= 1 = , u(0, 2 )= 1 = (4.676) √2 χ( 2 ) ! √2 1 − √2 χ(− 2 ) ! √2 0          0   1      are eigenstates of the 4 4 bispinor representation of the generator of rotations × around the z-axis [recall (4.514)–(4.518)]:

3 3 12 1 12 1 3 1 σ 0 S = S = σ = Σ = 3 . (4.677) 2 2 2 0 σ !

In order to construct explicit bispinors v(p,s3) we do not directly multiply them with the unit spinors (4.673) from the right-hand side, as we did to obtain u(p,s3) in (4.674), but we first use the above-observed freedom of multiplying (4.670) by an arbitrary 2 2 -matrix from the right. This is necessary to construct a spinor v(p,s ) × 3 332 4 Free Relativistic Particles and Fields with the physically most appropriate transformation properties under Lorentz trans- formations. From the 4 2 -matrices (4.670) it is possible to find directly the solutions v(p,s ) × 3 of the second equation in (4.668). We simply define v(p,s3) as the charge-conjugated spinor of v(p,s3) by an operation of the form (4.591), i.e.,

T v(p,s3)= Cu¯ (p,s3). (4.678)

It is easy to verify that this v(p,s3) solves (4.668). For a proof we take

† 0 1 † pσ˜ pσ u¯(p,s3)= u (p,s3)γ = χ (s3) , (4.679) √2 M M  q q  and form T pσ˜ T 1 M ∗ u¯ (p,s3)= T χ (s3). (4.680) √2  q pσ   M   q  Multiplying this by the charge-conjugation matrix C of (4.603) yields

T pσ˜ T 1 c M Cu¯ (p,s3)=  T  χ(s3), (4.681) √2 cq pσ  − M   q  with c = iσ2 of Eq. (4.604). At this place we realize that due to the hermiticity property of− the Pauli matrices (8.158) and (4.462), one has

σµ∗ = σµT , (4.682) such that relation (4.639) implies the four-component relation

cσµ∗c−1 = cσµT c−1 =σ ˜µ. (4.683)

c T With this, the charge-conjugated spinor u (p,s3) = Cu¯ (p,s3) goes directly over into the bispinor

pσ 1 M ∗ v(p,s3)= cχ (s3). (4.684) √2  q pσ˜  − M  q  Thus, while the 4 2 -solutions u(p) of Eq. (4.669) are multiplied by the Pauli spinors χ(s ) of Eq.× (4.673), the 4 2 -solutions v(p) of Eq. (4.670) are multiplied 3 × from the right by the spinors

χc(s ) cχ(s )= χ( s )( 1)s−s3 . (4.685) 3 ≡ 3 − 3 − These are called charge-conjugated Pauli spinors. Their explicit form is

c 0 1 χ (s3)= , − . (4.686) 1 ! 0 ! 4.13 Majorana Fields 333

This construction is necessary to ensure that the Pauli spinors v(p,s3) at rest have the same transformation behavior under rotations as the spinors u(p,s3) at rest.

Under rotation, the original basis spinors χ(s3) are multiplied by the 2 2 rotation   × matrix e−i · /2:

1/2

R

   ′ −i· /2 ′ −i · /2 χ(s3) χ (s3)= e χ(s3)= χ(s3) e ′ . (4.687) s3,s −−−→ s′ =−1/2 3 3 X  

The last step follows from the specific form (4.673) of the unit spinors. The same mixing occurs in the charge-conjugated spinors:

R ∗  χc(s ) χ′c(s )= cχ′∗(s )= cei· /2χ∗(s ). (4.688) 3 −−−→ 3 3 3 Using (4.639), we see that

   c−1e−i· /2c = ei · /2, (4.689) so that the right-hand side becomes

1/2

     −i· /2 ∗ −i · /2 c c ′ −i · /2 e cχ (s3)= e χ (s3)= χ (s3) e ′ . (4.690) s3,s s′ =−1/2 3 3 X  

c Thus χ (s3) is indeed rotated precisely like χ(s3). At rest, the 4 2 -matrices (4.669) and (4.670) become × 1 σ0 1 σ0 u(p)= 0 , v(p)= 0 . (4.691) √2 σ ! √2 σ ! −

Hence the rotation properties of the bispinors u(p,s3) = u(p)χ(s3) and v(p,s3) = c c v(p)χ (s3) at rest are the same as those of χ(s3) and χ (s3). Explicitly, the bispinors v(p,s3) at rest become

0 1 c c − 1 χ ( 1 ) 1 1 1 χ (− 1 ) 1 0 1 2   1 2   v(0, 2 )= c 1 = , v(0, 2 )= c 1 = , √2 χ ( 2 ) ! √2 0 − √2 χ (− 2 ) ! √2 1     −   −    1   0  −   (4.692) to be compared with (4.676) for u(0,s3). For the bispinors at rest in Eqs. (4.676) and (4.692), the Dirac equations in momentum space (4.663) take the simple form

M(γ0 1)u(0,s )=0, M(γ0 + 1)v(0,s )=0. (4.693) − 3 3 334 4 Free Relativistic Particles and Fields

By applying the boost matrix (4.511), we find the alternative expression for the bispinors with momentum p:

0 0 c M + /pγ 1 χ(s3) M + /pγ 1 χ (s3) u(p,s3)= , v(p,s3)= c . √2 χ(s3) ! √2 χ (s3)! 2M(p0 + M) 2M(p0 + M) − q q (4.694)

Since γ0 is a simple off-diagonal unit matrix, we can replace it by 1 in the left and ± right equation, respectively, and write just as well

c M + /p 1 χ(s3) M /p 1 χ (s3) u(p,s3)= , v(p,s3)= − c . √2 χ(s3) ! √2 χ (s3)! 2M(p0 + M) 2M(p0 + M) − q q (4.695)

The two sets of bispinors u(p,s3) and v( p,s3) satisfy the orthonormality con- ditions (4.665). Using (4.674) and (4.684), we− find pσ

† ′ 1 T pσ pσ˜ M ′ u (p,s3)u(p,s ) = χ (s3) ,  r  χ(s ) 3 2  M s M  pσ˜ 3 r    s     M    T 1 pσ pσ˜ ′ = χ (s3) + χ(s3) 2  M M  0 0 p T ′ p = χ (s3)χ(s )= δs ,s′ , (4.696) M 3 M 3 3 pσ˜ † ′ 1 c T pσ˜ pσ c ′ v ( p,s3)v( p,s ) = χ (s3) ,  s M  χ (s ) − − 3 2 s M − M  pσ 3 r        − r M  p0   = δs s′ , (4.697) M 3 3 pσ˜ † ′ 1 T pσ pσ˜ c ′ u (p,s3)v( p,s ) = χ (s3) ,  s M  χ (s ) − 3 2  M −s M  pσ 3 r        − r M  = 0,   (4.698) v†( p,s )u(p,s′ ) = 0. (4.699) − 3 3

The reason for the appearance of the negative momenta in the bispinors v( p,s3) c − is that the plane wave solutions fp s3 (x) in (4.662) carry negative momenta, so p c that states of a fixed momentum are associated with fp s3 (x) and f−p s3 (x). The momentum reversal in the conjugate wave functions goes along with the reversal of the spin orientation in the charge-conjugated Pauli spinors in Eq. (4.685). The 4.13 Majorana Fields 335 physical reason for these two reversals will be understood after field quantization in Section 7.4.3. Inserting a matrix γ0 between the bispinors in (4.696)–(4.699) we may also derive orthonormality relations between bispinors u(p,s3) and v(p,s3):

′ u¯(p,s )u(p,s ) = δ ′ , 3 3 s3,s3 ′ v¯(p,s )v(p,s ) = δ ′ , 3 3 s3,s3 ′ − u¯(p,s3)v(p,s3) = 0, ′ v¯(p,s3)u(p,s3) = 0. (4.700)

The two sets of spinors u(p,s ) and v( p,s ) span the spinor space at a fixed 3 − 3 momentum. This may be expressed by a completeness relation

0 † † p u(p,s3)u (p,s3)+ v( p,s3)v ( p,s3) = . (4.701) s − − M X3 h i To prove this 4 4 -matrix equation in Dirac space, we derive the separate polar- × ization sums10 for u- and v-spinors. These can be calculated directly from (4.674) and (4.684) as follows:

1 r M † pσ˜ pσ u(p,s3)¯u(p,s3) =   χ(s3)χ (s3) , 2 pσ˜ s M M  s3   s3 r X   X  s M     pσ 1 1 M /p + M =  pσ˜  = , (4.702) 2 1 2M  M   pσ 

1 r M c c† pσ˜ pσ v(p,s3)¯v(p,s3) =   χ (s3)χ (s3) , 2 pσ˜ −s M M  s3   s3 r X   X  − s M     pσ  1 1 − M /p M =  pσ˜  = − . (4.703) 2 1 2M  M −    Combining the two polarization sums for p and p, respectively, and multiplying − them by the Dirac matrix γ0 from the right, proves their completeness. Subtracting the two polarization sums from each other yields another relation that is the Dirac-adjoint version of the completeness relation (4.701) [recall (4.495)]:

[u(p,s3)¯u(p,s3) v(p,s3)¯v(p,s3)]=1. (4.704) s − X3 10They may be called semi-completeness relations. 336 4 Free Relativistic Particles and Fields

This contains a minus sign in the second sum which reflects the minus sign in the second orthogonality relation (4.700). This sign will be important in Section 7.10 to prove a famous theorem on the relation between spin and statistics of fundamental particles. Polarization sums will frequently be needed later, in particular in the process of field quantization. We introduce the sums

† † P (p) u(p,s3)u (p,s3), P¯(p) v(p,s3)v (p,s3), (4.705) ≡ s ≡ s X3 X3 0 0 defined only for p on the mass shell, p = ωp. They satisfy the relation P¯(p) = P ( p). Similar polarization sums exist for plane-wave solutions for any spin. In − − general, the polarization sums P (p) and P¯(p) of positive and negative energies of momenta p and p, respectively, fulfill the relation − P¯(p)= P ( p), (4.706) ± − where the upper sign holds for integer spin, and the lower for half-integer spin. The matrices P (p) and P ( p) are projection matrices onto solutions of momenta p and 0 − 0 p with energies p = ωp and p = ωp, respectively. As such they satisfy − − P ( p)2 = P ( p). (4.707) ± ± It is always possible to find a single covariant expression for P (p) defined for arbitrary 0 0 off-shell values of p which, for the on-shell values p = ωp, reduces to the above projections P (p) and P ( p). In the Dirac case, where P±(p)=(/p M)γ0/2M, we − 0 − verify that (4.707) is true for p = ωp. ± Spinors in Dirac Representation Let us also write down the bispinors in the Dirac representation (4.550) of the γ- matrices. The rest bispinors are solutions of equations (4.693), where

1 0 γ0 = . D 0 1 − ! Thus we have 1 0 χ( 1 ) 0 χ(− 1 ) 1 1 2   1 2   u(0, 2 )= = , u(0, − 2 )= = − , (4.708) 0 ! 0 0 ! 0          0   0      and 0 0 0 0 0 1 0 1   1   v(0, 2 )= c 1 = , v(0, 2 )= c 1 = . (4.709) χ ( 2 ) ! 0 − χ (− 2 ) ! √2 1        −   1   0      4.13 Majorana Fields 337

The bispinors at finite momentum are obtained from these by applying the 4 4 boost matrix (4.511), yielding ×

M + /p D χ(s3) M /p D 0 u(p,s3)= , v(p,s3)= − c . 0 χ (s3) 2M(p0 + M) ! 2M(p0 + M) ! q q (4.710)

0 Since γD has a simple diagonal form with eigenvalues 1 for upper and lower spinor 0 ± components [see Eq. (4.562)], we have replaced γD directly by its eigenvalues when going from (4.511) to (4.710) [as we did from (4.694) to (4.695)]. More explicitly, we can write

 p p0 + M ·  s 2M   2M(p0 + M)  c u(p,s3)= χ(s3), v(p,s3)= χ (s3).(4.711)  p q  ·   p0 + M       2M(p0 + M)   s     2M   q    In this representation, the bispinors u(p,s3) of slowly moving particles have large upper and small lower spinor components. The converse is true for the bispinors v(p,s3). This is what makes the original Dirac spinors useful for discussing the nonrelativistic limit of spin-1/2 particles, as observed before in the boost matrix (4.556) and in x-space equations (4.569). The Dirac spinors possess, of course, the same polarization sums (4.705) as in µ the chiral case, if the appropriate Dirac matrices γD are used on the right-hand side.

Helicity Spinors Sometimes, the choice of the spinors (4.673) with the particle spins quantized along the z-axis is not the most convenient basis in spinor space. Instead of the z-axis, one may choose any quantization direction, in particular, the direction of the momentum of the particle. This amounts to multiplying the 4 2 -matrix solutions (4.669) by a 2 2 rotation matrix from the right: × × 3 2 R(pˆ) e−iφσ /2e−iθσ /2, (4.712) ≡ where θ,φ are the spherical angles of the momentum p. In contrast to the notation of Eq. (4.9), the rotation matrix carries now an argument indicating that the z- direction is rotated into the momentum direction pˆ.11 Equivalently, we may choose in the bispinors (4.674), instead of χ(s3), a basis χh(pˆ,λ) with λ = 1/2, defined by ±

3 2 χ (pˆ,λ) R(pˆ) χ(s ) e−iφσ /2e−iθσ /2χ(λ). (4.713) h ≡ 3 ≡ 11As before in this sections, hats on vectors denote unit vectors, not Schr¨odinger operators. 338 4 Free Relativistic Particles and Fields

The explicit components are θ θ cos e−iφ/2 sin e−iφ/2 χ (pˆ, 1 )=  2  , χ (pˆ, 1 )=  − 2  . (4.714) h 2 θ h − 2 θ  iφ/2   iφ/2   sin e   cos e   2   2      They diagonalize the projection of the angular momentum in the rest frame along the direction of motion of the particle pˆ = (sin θ cos φ, sin θ sin φ, cos θ), the so-called helicity: 1 1 cos θ sin θ e−iφ

h(pˆ) L pˆ =  pˆ = . (4.715) ≡ · 2 · 2 sin θ eiφ cos θ − ! The eigenvalues are 1 h(p)χ(pˆ,λ)= λχ(pˆ, h), λ = . (4.716) ±2

The eigenstates χh(pˆ,λ) are called helicity spinors. The associated bispinors are, as in (4.674) and (4.684),

pσ pσ˜ 1 r M 1 s c uh(p,λ)=   χ(pˆ,λ), vh(p,λ)=  M  χ (pˆ,λ),(4.717) √2 pσ˜ √2 pσ          s M   − M     r  with the charge-conjugated helicity spinors χc(pˆ,λ)= cχ∗(pˆ,λ). (4.718) These diagonalize the 4 4 bispinor representation of the helicity: ×

1 1  0

H(pˆ) S pˆ = ¦ pˆ = pˆ. (4.719) ≡ · 2 · 2 0  ! · This is a direct consequence of the fact that H(pˆ) commutes with the boost matrix (4.511). Alternatively, we can obtain the helicity bispinors (4.717) by first boosting the bispinors at rest (4.713), (4.714) into the z-direction, and rotating them afterwards into the pˆ-direction:

p0σ0 p σ3 R(pˆ) −| | χ(h) 1  s M  uh(pˆ,λ) = , (4.720) 0 0 3 √2  p σ + p σ   R(pˆ) χ(h)   | |   s M    and

p0σ0 + p σ3 p c 1 R(ˆ) | | χ (h) v (pˆ,λ)=  s M  . (4.721) h 0 0 3 √2  p σ p σ   R(pˆ) χc(h)   −| |   − s M    4.13 Majorana Fields 339

The equality with (4.717) follows from the transformation law (4.477), according to which 3

R(pˆ) p σ = p R(pˆ). (4.722) | | · One of the important advantages of the helicity spinor is that it has a smooth limit as the particle mass M tends to zero. Indeed, by expanding M 2 p0 = p2 + M 2 = p + + ... (4.723) | | 2 p q | | we see that p0σ0 p σ3 M→0 2 0 0 −| | , M −−−→ M 0 p | | ! p0σ0 + p σ3 M→0 2 p 0 | | | | . (4.724) M −−−→ M 0 0 !

Thus the massless helicity spinors uh(p,λ) and vh(p,λ) have only two nonzero com- ponents. We shall normalize them to u† (p,λ)u(p,λ)=2p0 =2 p , v† (p,λ)v(p,λ)=2p0 =2 p , (4.725) h | | h | | as opposed the normalization to p0/M of the massive spinors (4.717). The explicit form is 0 0 u (p, 1 ) = p   u (p), h 2 | | 1 ≡ R q  R(pˆ)     0 !    0 R(pˆ) 1 u (p, 1 ) = p  !  u (p), (4.726) h − 2 | | 0 ≡ L q      0    and 0 0 v (p, 1 ) = p   = v (p) h 2 | | 0 R q  R(pˆ)   1   − !    1 R(pˆ) − 0 v (p, 1 ) = p  !  = v (p). (4.727) h − 2 | | 0 L q      0    The helicity bispinors uh(p,λ) and vh(p,λ) are eigenstates of the chirality matrix 1 0 γ5 = − , (4.728) 0 1 ! 340 4 Free Relativistic Particles and Fields with the eigenvalue 2λ. By applying the 4 4 projection matrix 1−γ5 to the bispinors × 2 uh(p,λ), forming

1 γ − 5 u (p,λ), (4.729) 2 h we obtain a negative helicity state. Such projected bispinors are used for the de- scription of neutrinos which only exist with negative helicity (left-handed neutrinos). As we shall see later in Section 27, weak interactions involve also the orthogonally projected bispinors

1+ γ 5 v (p,λ) (4.730) 2 h which describe antineutrinos. These exist only with positive helicity (right-handed antineutrinos).

4.14 Lorentz Transformation of Spinors

Let us study the behavior of the bispinors u(p,s3) and v(p,s3) under Lorentz trans- formations. This will be most straightforward in the chiral representation, where we may focus our attention upon the upper components only, which will be denoted by ξ(p,s3). The properties of the lower components, to be denoted by η(p,s3), can be obtained by a simple change in the direction of the momentum. The upper

components can be written explicitly as  −· /2 ξ(p,s3)= B()χ(s3)= e χ(s3). (4.731)

( 1 ,0) Applying to this a general Lorentz transformation D 2 (Λ), the momentum p is ′ µ ν changed to some other vector p , which is the spatial part of the four-vector Λ νp . The transformation can be done in three steps: First, deboost the particle by ap- plying a boost opposite to the particle’s momentum which brings it to rest, with the µ µ four-momentum p becoming pR = (M, 0). Second, perform a rotation, and third, boost the particle to its final four-momentum p′µ. Thus we can write the general Lorentz transformations as

1 ′

 «    ( 2 ,0) ′ ′ −1 − · /2 −i · /2 · /2 D (Λ) = B( )W (p , Λ,p)Bˆ (ζ) e e e . (4.732)  ≡ The rotation W (p′, Λ,p) in the middle is called a Wigner rotation. It is an element of the little group of the massive particle acting only in its rest frame [see the earlier short discussion on p. 35]. Let p′ be the momentum reached from the momentum p after a Lorentz trans- formation Λ. Then the spinor (4.731) changes as follows:

Λ 1 ′

 «  ′ ′ ( ,0) − · /2 −i · /2 ξ(p,s ) ξ (p ,s ) = D 2 (Λ)ξ(p,s )= e e χ(s ) 3 −−−→ 3 3 3 4.14 Lorentz Transformation of Spinors 341

1/2

 «  − · /2 ′ −i · /2 = e χ(s3) e ′ s ,s3 s′ =−1/2 3 3 X   1/2 ′ ′ ′ = ξ(p ,s )W ′ (p , Λ,p), (4.733) 3 s3,s3 s′ =−1/2 3 X

where we have used the rotation property (4.687) of the spinors χ(s3), which amounts here to 1/2 ′ ′ ′ W (p , Λ,p)χ(s )= χ(s )W ′ (p , Λ,p). (4.734) 3 3 s3,s3 s′ =−1/2 3 X

By analogy, the spinor η(p,s3) transforms like

1/2 Λ ′ ′ (0, 1 ) ′ ′ ′ η(p,s ) η (p ,s )= D 2 (Λ)η(p,s )= η(p ,s )W ′ (p , Λ,p), (4.735) 3 3 3 3 s3,s3 −−−→ s′ =−1/2 3 X

implying for the Dirac spinor u(p,s3) the transformation law

1/2 Λ ′ ′ ′ ′ ′ u(p,s ) u (p ,s )= D(Λ)u(p,s )= u(p ,s )W ′ (p , Λ,p). (4.736) 3 3 3 3 s3,s3 −−−→ s′ =−1/2 3 X The result can be expressed most compactly in terms of the 4 2 -matrix form (4.670) for the bispinor solutions as ×

Λ u(p) u′(p′)= D(Λ)u(p)= u(p′)W (p′, Λ,p). (4.737) −−−→ We are now prepared to understand the group-theoretic reason for the occur- rence of the rotation matrix c = e−iπσ2/2 in the charge-conjugated bispinor v(p) of Eq. (4.684). The 4 2 solutions v(p) of (4.670) transform in the same way as u(p) × of (4.669):

Λ v(p) v′(p′)= D(Λ)v(p)= v(p′)W (p′, Λ,p). (4.738) −−−→

The behavior of the Dirac spinors v(p,s3) is found by multiplying this equation from ∗ the right-hand side with cχ (s3) [recall (4.684)], leading to

Λ v(p,s ) v′(p′,s )= D(Λ)v(p,s )= v(p′)W (p′, Λ,p)cχ∗(s ). (4.739)

3 −−−→ 3 3 3  Now we use the fact that the 2 2 -Wigner rotation can be written as e−i«· /2, which satisfies the relation (4.689), so× that

c−1W (p′, Λ,p)c = W ∗(p′, Λ,p), (4.740)

to rewrite on the right-hand side

′ ∗ ∗ ′ ∗ ∗ ′ ∗ W (p , Λ,p)cχ (s3)= cW (p , Λ,p)χ (s3)= cχ (s )W ′ . (4.741) 3 s3s3 342 4 Free Relativistic Particles and Fields

We obtain the transformation law for the bispinors v(p,s3):

1/2 Λ ′ ′ ′ ′ ∗ ′ v(p,s3) v (p ,s3)= D(Λ)v(p,s3)= v(p ,s )W ′ (p , Λ,p). (4.742) 3 s3s3 −−−→ s′ =−1/2 3 X Thus we find that under Lorentz transformations, the spin orientations of the bispinors v(p,s3) are linearly recombined with each other by the complex-conjugate ∗ ′ Wigner rotations W ′ (p , Λ,p). This is a consequence of the presence of the matrix s3s3 c in the 4 2 -matrices (4.670) for v(p), which has reversed canonical spin indices. × Had we used the bispinors v(p,s3) in the form (4.684), the same result would have been obtained from the observation (4.688), that the spin indices of the charge- c conjugated Pauli spinors χ (s3) are linearly recombined with each other by the complex-conjugate rotation matrix. The transformation properties (4.737) and (4.738) can be verified most easily in an infinitesimal form for spinors at rest. If Λ is an infinitesimal rotation R with µ µ µ the 4 4 -matrix R ν = δ ν i³ L ν , the left-hand sides must be multiplied by

× − · i ¦ D(Λ) = R = 1 i³ /2, where Σ is given by (4.514). This produces the same

− ·  infinitesimal Wigner rotation 1 i³ /2 on the right-hand sides. Thus we have the − · relations

iΣiu(0)= u(0)iσi, iΣiv(0)= v(0)iσi. (4.743)

Using (4.669) and (4.670), these become explicitly

0 0 0 0 i σ 1 σ i i σ 1 σ i ∗ iΣ 0 = 0 iσ , iΣ 0 = 0 c( iσ ) . (4.744) σ ! √2 σ ! σ ! √2 σ ! − − − Let us also write down the Wigner rotations for the helicity spinors (4.717). Since they arise from the 4 2 -matrix solutions (4.669) by a multiplication from × the right with the 2 2 rotation matrix (4.712), a Lorentz transformation of uh(p,λ) yields obviously ×

1/2 Λ ′ ′ ′ ′ ′ uh(p,λ) uh (p ,λ)= D(Λ)uh(p,λ)= uh(p ,λ )Wh λ′λ(p , Λ,p), (4.745) −−−→ ′ λ =X−1/2 with the helicity form of the Wigner rotation

′ −1 ′ ′ Wh(p , Λ,p)= R (pˆ )W (p , Λ,p)R(pˆ). (4.746) Similarly,

1/2 Λ ′ ′ ′ ′ H∗ ′ vh(p,λ) vh (p ,λ)= D(Λ)vh(p,λ)= vh(p ,λ )Wλ′,λ(p , Λ,p), (4.747) −−−→ ′ λ =X−1/2 with W H∗(p′, Λ,p)= R−1 ∗(pˆ′)W ∗(p′, Λ,p)R∗(pˆ). (4.748) 4.15 Precession 343

4.15 Precession

The properties of relativistic spinors under Lorentz transformations are crucial for a phenomenon known in atomic physics as Thomas precession. The Thomas precession is a direct consequence of what may be called Wigner precession.

4.15.1 Wigner Precession Consider an electron moving around an atomic nucleus. In each time interval ∆t, it receives a small centripetal Lorentz boost changing its momentum. Let us see what happens to the upper two components of the canonical bispinors u(p,s3) which are

explicitly  −· /2

ξ(p,s )= B( )χ(s ) e χ(s ). (4.749) 3 3 ≡ 3 At an instance of time t, the electron moves with a certain velocity through space. Its state can be described by the two-component spinor ξ(p,s3) defined in Eq. (4.731). As the atomic force acts on the electron, it is accelerated towards the nucleus. Thus, after a small time interval dt, the electron will have a new ′ momentum and a spinor ξ(p ,s3) which can be obtained from the first by applying

a small Lorentz boost  −d¯· /2

B(d¯)= e (4.750) ′ to the spinor ξ(p,s3), which changes its momentum from p to p . The resulting transformation is split into three factors, as in (4.732). The first is a pure boost in the p direction, which brings the four-momentum p to its rest frame where it − is pR = (M, 0). The second factor is a rotation, and the third is a boost into the ′ final four-momentum p . In this process, the spin indices of the spinor ξ(p,s3) are linearly recombined with each other by a Wigner rotation according to Eq, (4.733).

′  Let us calculate this, taking advantage of the fact that d¯ is very small. Then

    differs very little from , say = + d , where d is another small rapidity of

the order of d¯. To indicate the smallness of the associated rotation vector in the

Wigner rotation we shall denote it by d«W. Its size is calculated from the equation

    ¯    e−id«W· /2 = e( +d )· /2e−d · /2e− · /2. (4.751)

Before calculating d«W exactly, let us estimate it for slowly moving particles where

  d¯, d , and are all of the same order. Then we may expand both sides of (4.751) up to the second order in all quantities as follows:

2

      1 id« /2 1+( + d ) /2+( + d ) /4 − W · ≈ ·

h 2 i 2

 ¯    1 d¯ /2+(d ) /4 1 /2+ /4 . (4.752) × − · − ·

h i h i ¯ In the product on the right-hand side, we set d d to cancel the first-order terms. The second-order terms decompose into Hermitian≈ and antihermitian parts. Since

we are interested only in d«, we must extract the antihermitian part. Using the

identity

  a  b = a b + i (a b) , (4.753) · · · × · 344 4 Free Relativistic Particles and Fields

we find

          1 id« /2 1 i [( + d ) d ] + [( + d ) ] [d ] /4

− W · ≈ − { × × − × }·

  1+ i (d ) /4, (4.754) ≈ × ·

so that we obtain the rotation vector for small :

1

  d« d . (4.755) W ≈ 2 × The spin matrix is rotated under a Wigner rotation as follows:

′ −1

   = W W. (4.756)

→  For an infinitesimal W 1 id« /2 this yields

≈ − W ·

«   «  d = i [d , ]= d . (4.757) W · W ×

An accelerated point particle receives a small boost d¯ in each small time interval

dt. In this time interval, the spin precesses at an angular velocity d«W/dt. In the limit dt 0, Eq. (4.755) implies an angular velocity of Wigner rotations →

1 ˙ 1

  ª v v˙ . (4.758) W ≈ 2 × ≈ 2c2 ×

4.15.2 Thomas Precession A relative of the Wigner precession is observable in atomic physics as a Thomas precession. In an atom, the small additional boost acts on the electron moving with momentum p in its rest frame. This implies that the small boost (4.750) has to be

replaced by

 ¯   

˜ −1 −· /2 −d · /2 · /2

 ¯  B(d¯)= B( )B(d )B ( )= e e e , (4.759)

and (4.751) becomes

      ¯  e−id« T· /2 = e( +d )· /2e− · /2e−d · /2. (4.760)

The small-velocity calculation (4.752) becomes now

2

      1 id« /2 1+( + d ) /2+( + d ) /4 − T · ≈ ·

h 2 i 2

  ¯  ¯ 1  /2+ /4 1 d /2+(d ) /4 , (4.761) × − · − ·

h i h i ¯ leading with d d to

          1 id« /2 1 i [( + d ) ] + [( + d ) d ] [ d ] /4

− R · ≈ − { × × − × }·

  1+ i ( d ) /4. (4.762) ≈ × · 4.15 Precession 345

The resulting small Thomas rotation vector:

1

  d« d (4.763) T ≈ −2 × is exactly the opposite of the Wigner rotation vector in Eq. (4.755). Of course, the same thing is true for the rate of the Thomas precession

1 ˙ 1

  ª v v˙ . (4.764) T ≈ −2 × ≈ 2c2 × For the spin vector S, which is the total angular momentum in the electron’s rest frame, this amounts to the equation of motion dS 1

= ª S (v v˙ ) S . (4.765) dt T × ≈ −2c2 × ×

For finite , this equation will acquire relativistic correction factors and become 1 γ2

ª = v v˙ . (4.766) T c2 γ +1 × The derivation of this expression is somewhat tedious and will therefore be given in Appendix 4B. The angular velocity of the Wigner rotation has observable consequences in atomic physics, where it is seen as a Thomas precession. This will be discussed in more detail in Subsection 6.1.3. It is a purely kinematic effect, caused entirely by the structure of the Lorentz group. Mathematically speaking, it is due to the fact that pure Lorentz transformations do not form a subgroup of the full Lorentz group. When performing pure Lorentz transformations one after another in such a way that the final frame is again at rest with respect to the initial one, the result is always a Wigner rotation.

4.15.3 Spin Four-Vector and Little Group The working of the Wigner rotations in the little group found in Section 4.14 can be understood independently of the particular spinors. For any massive elementary or composite physical system we introduce a quantity called total spin four-vector. µν It is a combination of the total angular momentum Jˆ = Lˆµν + Sµν with the total momentum operatorp ˆ which together form a vector 1 Sˆ = ǫ Jˆνλpˆκ, (4.767) µ 2 µνλκ

0123 where ǫµνλκ is the totally antisymmetric unit matrix with ǫ = 1 (see [12]). For massive elementary particles of momentum pµ, the time and space compo- nents of the spin four-vector become explicitly

Sˆ0 = p Jˆ, Sˆ = p0Jˆ p Kˆ , (4.768) · − × 346 4 Free Relativistic Particles and Fields where Jˆ = (Jˆ23, Jˆ31, Jˆ12) and K = (Jˆ01, Jˆ02, Jˆ03). Studying particles at fixed mo- menta, we have dropped operator hats on the momenta in (4.768), and the generators of orbital angular momentum Lˆµν become differential operators in momentum space, where they read explicitly [compare (4.92)]

Lˆ i(p ∂ p ∂ ), (4.769) µν ≡ µ ν − ν µ with ∂ ∂/∂pµ. µ ≡ Using the commutation relations (4.295) and (4.99), the components of Sˆµ in (4.767) can be shown to satisfy the commutation rules

µ ν µν λ κ [Sˆ , Sˆ ]= iǫ λκSˆ pˆ . (4.770) The proof makes use of the tensor identity (4.242) which, after taking advantage of ′ ′ the antisymmetry of Jˆλλ in the indices λ and λ′, and the symmetry of pκpκ in κ and κ′, leads to

µ ν′ λλ′ κ κ′ 1 νν′ µ λλ′ κ κ′ ǫ ǫ ′ ′ Jˆ p p = ǫ ǫ ′ ′ Jˆ p p , (4.771) µνλκ λ κ 2 µκ λλ κ and thus to the right-hand side of (4.770). The same result can of course be derived without the lengthy identity (4.242) by considering time and space components in (4.768) separately, using the commutators (4.76), (4.78), and (4.99). Then we find

[Lˆ ,p0]= 0, [Lˆ ,pj] = iǫ pk, i i ijk (4.772) [Mˆ ,p0] = ipi, [Mˆ ,pj] = iδ p0. i − i − ij For a free particle, pµ is independent of time, and so is Sˆµ. By definition, the spin four-vector Sˆµ is orthogonal to the four-momentum:

µ Sˆ pµ =0. (4.773)

The physical significance of Sˆµ becomes clear by going into the rest frame of a massive particle where the system has no velocity, so that [recall (4.21)]

λ pR = Mc(1, 0, 0, 0). (4.774) Then 1 Sˆ0 = 0, Sˆi = Mc ǫijkJˆik Mc Jˆ. (4.775) R R 2 ≡ i Removing an overall factor Mc, we define the operators of Wigner rotations

Wˆ i Sˆi /Mc, (4.776) ≡ R satisfying the commutation relations

i j k [Wˆ , Wˆ ]= iǫijkWˆ . (4.777) 4.15 Precession 347

Thus the total spin four-vector has the property that its spatial components coincide in the rest frame with the total angular momentum of the system. This is certainly time independent due to angular momentum conservation. Moreover, at zero momentum, the orbital part of Jˆi vanishes, so that only the i ˆi i spin part S of J survives, and we can drop the hats on top of SR which indicate the presence of differential operators. Then we obtain pure spin matrices for the operators Wˆ i: W i Si /Mc, (4.778) ≡ R satisfying the same commutation relations as the operators in (4.777). They will be called spin three-vectors. The relation between the spin three-vector and the spin four-vector is obtained by applying the pure Lorentz transformation matrix (4.18) to (4.775), yielding γ2 1 1 Sˆ = Sˆ + (Sˆ v)v, Sˆ0 = S v. (4.779) R γ +1 c2 R · c · The inverse relations are γ 1 γ 1 Sˆ = Sˆ Sˆ0 v = Sˆ (Sˆ v)v, Sˆ0 =0, (4.780) R − γ +1 c − γ +1 c2 · R as can be verified with the help of the relation v2 γ2 1 = − , (4.781) c2 γ2 which implies that γ/(γ 1)c2 =(γ 1)/γv2. Note that − − 1 γ Sˆ = Sˆ v = Sˆ v. (4.782) 0 c · c R · For massless particles, the Wigner rotations have quite different properties from those of massive particles. In the special reference frame in which the massless par- µ ticle runs along the z-axis with a reference momentum pR = (1, 0, 0, 1)p introduced in (4.26), these components become Sˆ0 = p Jˆ3, Sˆ1 = p (J 1 + K2), Sˆ3 = p (Jˆ2 Kˆ 1), Sˆ3 = p Jˆ3. (4.783) − The three independent components Wˆ 3 Sˆ0 = Sˆ3, W 1 Sˆ1, W 2 Sˆ1 (4.784) ≡ ≡ ≡ satisfy the commutation relations [Wˆ 3, Wˆ 1]= iWˆ 2, [Wˆ 3, Wˆ 2]= iWˆ 1, [Wˆ 1, Wˆ 2]=0. (4.785) − These generate a euclidean group in a plane. Recall the definition of this group. In D dimensions, it consists of D(D 1) generators L of the D-dimensional rotation − ij group:

[Lij, Lik]= iLjk, (4.786) 348 4 Free Relativistic Particles and Fields

and D generators of translation pi, which commute with each other and are vector operators under rotations:

[p ,p ]=0, [L ,p ]= i (δ p δ p ) . (4.787) i j ij k ik j − jk i These commutation rules can be obtained from those of the Lorentz group in D- dimensions [see Eqs. (4.76)–(4.78)] by setting pi = Mi/c and letting c go to infinity. This construction is called group contraction. The commuting generators Wˆ 1 and Wˆ 2 can be diagonalized simultaneously like commuting momentum operators in a plane with arbitrary continuous eigenvalues w1 and w2, respectively. The third generator Wˆ 3 generates rotations in this plane with discrete eigenvalues w3 = λ, where λ are azimuthal quantum numbers which can be equal to an integer of a half-integer number. In mirror-symmetric theories, both signs have to occur. In an arbitrary reference frame, Wˆ 0 is given, according to (4.768), by the operator Wˆ 0 = p J/p. This shows that the eigenvalues λ measure · the angular momentum around the momentum direction, i.e., the helicity of the particle. It turns out that in nature, all massless particles happen to follow a representa- tion of the Wigner algebra which have only trivial eigenvalues w1 = w2 = 0. They are characterized completely by the helicity, which is unchanged under Wigner ro- tations. It merely receives a pure phase factor multiplying the helicity spinors or massless polarization vectors. The occurrence of only such a subset of all possible zero-mass representations can be understood by a limiting process such as the one performed in the derivation of the massless spinors (4.726) and (4.727). We imagine for a moment that all massless particles carry a small mass which we let go to zero. It can then be verified that the limiting spinors change under a Lorentz transformation merely by a phase factor associated with helicity λ = 1/2: ± iω(p′,Λ,p)/2 −iω(p′,Λ,p)/2 D(Λ)uR(p)= uR(p)e ,D(Λ)uL(p)= uL(p)e , −iω(p′,Λ,p)/2 iω(p′,Λ,p)/2 D(Λ)vR(p)= uR(p)e ,D(Λ)uL(p)= uL(p)e . (4.788)

For the polarization vectors of electromagnetism ǫµ(k,λ) in (4.319), and the tensors ǫµν (k,λ) of gravity in (4.416) we have, similarly,

µ ν µ iλω(p′,Λ,p) µ µ′ νν′ µµ′ iλω(p′,Λ,p) Λ ν ǫ (p,λ)= ǫ (p,λ)e , Λ νΛ ν′ ǫ (p,λ)= ǫ (p,λ)e .(4.789)

The reader is invited to derive this directly from the explicit expressions for these objects.

4.16 Weyl Spinor Calculus

Weyl has devised a simple calculus for constructing spinor invariants of the Lorentz

group. It is very similar to the tensor calculus. The spinor in the upper two com-  ponents of the Dirac field, which transforms under boosts via the matrix e− · /2, 4.16 Weyl Spinor Calculus 349

was previously denoted by ξα, while the spinor in the lower components, transform-

˙   · /2 β ing via e , was previously denoted by η . Complex conjugation brings ξα into

∗  ∗ −· /2 a spinor (ξα) which transforms via e . Such a spinor is given a lower dotted index, i.e., we write (ξ )∗ ξ∗ . (4.790) α ≡ α˙

˙ ∗  Similarly, we define (ηβ)∗ which transforms via e· /2 as

˙ (ηβ)∗ η∗β. (4.791) ≡ From the earlier discussion in Section 4.11 we know that

α˙ ∗ ∗ β˙ (η ) ξα, (ξβ) η (4.792) are Lorentz invariants. With the above notation, these can be viewed as

∗α ∗ β˙ η ξα, ξ β˙ η . (4.793)

Thus the invariants arise by simple contractions of equal upper and lower indices. A further invariant can be constructed from two spinors which both have lower indices α and β, namely ′ βα ξβ c ξα, (4.794) with the 2 2 -charge conjugation matrix c = iσ2 of (4.604). Writing (4.794) in matrix notation× as ξ′T cξ, it goes under rotations− over into

T

 ³  ξ′T e−i³· /2ce−i · /2ξ. (4.795)

T ∗ ∗

   Since  = [see (8.158)] and c = c from (4.639), this is obviously invari- − ant. A similar manipulation shows invariance under boosts. Thus the matrix cαβ constitutes an antisymmetric (or symplectic) metric in spinor space. Accordingly, we define ξβ cβαξ . (4.796) ≡ α Then the invariant (4.794) arises by a contraction of equal upper and lower indices, just as in the notation in Minkowski space:

′ βα ′α ξβc ξα = ξ ξα. (4.797)

Similarly, we can form an invariant from two ηβ˙ -spinors:

′β˙ −1 α˙ η (c )β˙α˙ η , (4.798) defining −1 α˙ η ˙ (c ) ˙ η , (4.799) β ≡ βα˙ ′ β˙ which makes the contraction ηβ˙ η Lorentz-invariant. 350 4 Free Relativistic Particles and Fields

β˙ ( 1 ,0) The Lorentz transformation matrices associated with ξα and η are D 2 (Λ) 1 1 ′ 1 ˙ (0, 2 ) ( 2 ,0) α (0, 2 ) β and D (Λ), respectively. They carry Weyl indices D (Λ)α and D (Λ) β˙′ . It is possible to combine Weyl spinors to vectors rather than scalars with the help of the σµ-matrices (8.156) and ofσ ˜µ. They may be thought of as carrying Weyl labels µ (σ )αβ˙ , (4.800) and ˙ (˜σµ)βα. (4.801) Then the indices show directly which spinors are required to form vectors:

∗ µ ∗ µ βα˙ ∗ µ ∗α µ β˙ ξ σ˜ ξ = ξ β˙ (˜σ ) ξα, η σ η = η (σ )αβ˙ η . (4.802) The vector nature of these combinations is proved by rewriting the transformation law (4.508) in a 2 2 -form as × ( 1 ,0) −1 µ (0, 1 ) µ ν D 2 (Λ) σ D 2 (Λ) = Λ νσ , (0, 1 ) −1 µ ( 1 ,0) µ ν D 2 (Λ) σ˜ D 2 (Λ) = Λ νσ˜ . (4.803)

Written with Weyl indices, this reads

1 1 ˙ ( 2 ,0) −1 ′ α µ (0, 2 ) β µ [D (Λ) ]β (σ )αβ˙ D (Λ) β˙′ = (σ )ββ˙′ , (0, 1 ) −1 α˙ µ α˙ ′β ( 1 ,0) β′ µ αβ˙ ′ [D 2 (Λ) ] α˙ ′ (˜σ ) D 2 (Λ)β = (˜σ ) . (4.804)

In Weyl’s calculus, the Dirac equation reads

′ α µ ′ Mδα (iσ ∂µ)αβ˙′ ξα ′ (i/∂ M)ψ(x)= − µ βα˙ ′ β˙ β˙ =0. (4.805) − (iσ˜ ∂µ) Mδ ˙′ ! η ! − β 4.17 Massive Vector Fields

In order to understand weak interactions and some strongly interacting particles, we must also learn to describe massive vector fields. They can be electrically neutral or carry electric charges 1. ± 4.17.1 Action and Field Equations The action of a neutral massive vector field V µ(x) can be obtained by writing down an action like the electromagnetic one in (4.237), and simply adding a mass term:

4 4 1 µν 1 2 µ = d x (x)= d x Fµν F + M VµV , (4.806) A Z L Z −4 2  where the field tensor is now

F (x) ∂ V (x) ∂ V (x). (4.807) µν ≡ µ ν − ν µ 4.17 Massive Vector Fields 351

Charged vector fields are described by the action

4 4 1 ∗ µν 2 ∗ µ = d x (x)= d x Fµν F + M Vµ V . (4.808) A Z L Z −2  In either case, the equation of motion reads

µν 2 ν ∂µF + M V =0, (4.809) or more explicitly

[( ∂2 M 2)g + ∂ ∂ ]V µ(x)=0. (4.810) − − µν µ ν As for electromagnetic fields, the Euler-Lagrange equation for the zeroth component V 0(x) does not involve the time derivative of V 0(x) and is therefore not a dynam- ical equation, but it relates V 0(x) to the spatial components V i(x) and their time derivatives via 1 V 0(x)= ∂ F µ0/M 2 = ∇ V˙ (x)+ ∇2V 0(x) . (4.811) − µ M 2 · h i In the limit M 0, this gives rise to Coulomb’s law (4.267). Taking the four- → divergence of (4.809), we see that it vanishes:

µ ∂µV (x)=0. (4.812)

Physically, this eliminates any scalar content ∂µs(x) from the vector potential. In contrast to the electromagnetic vector field Aµ(x), the zero four-divergence is not a matter of choice as in the Lorenz condition (4.252), but it follows here from the Euler-Lagrange equations. Inserting (4.812) back into (4.810), we find that the four components of V µ(x) satisfy the Klein-Gordon equation

( ∂2 M 2)V µ(x)=0, (4.813) − − which is the massive version of the electromagnetic field equations (4.253) in the Lorenz gauge.

4.17.2 Plane Wave Solutions for Massive Vector Fields The plane-wave solutions look the same as in Eqs. (4.309) and (4.310) for the electro- magnetic vector potential. The mass of the vector field modifies only the possible polarization vectors ǫµ(k,λ). As in the electromagnetic case, the zero divergence property (4.812) eliminates one degree of freedom in the polarization vectors. It is, however, impossible to eliminate more, since there exists no gauge invariance, and thus no analog of the restricted gauge transformations (4.255). Adding to V µ(x) a gradient ∂µΛ(x) and inserting the new field into the field equation (4.809) produces 2 the condition M ∂µΛ(x) = 0, admitting only a trivial constant for Λ(x). The polar- ization vector has therefore three independent components. Physically this reflects 352 4 Free Relativistic Particles and Fields the fact that a massive vector particle can be studied in its rest frame. There the 2 third component of angular momentum L3, and the square L have three eigenstates corresponding to three linear combinations of the spatial vector components. With µ the restriction (4.812), the polarization vectors must satisfy kµǫ (k,s3) = 0. For particles at rest only three polarization vectors are allowed: 0 0 1 1 0 ǫµ(0, 1) =   , ǫµ(0, 0) =   . (4.814) ± ∓√2 i 0      ±     0   1      2 These are obviously eigenstates of the 4 4 -angular momentum matrices L3 and L in the defining representation (4.54)–(4.56).× The relative phases have been chosen as in (4.321) to comply with the Condon-Shortley convention as on p. 284. Recall that µ ′ this means that by applying L+ to ǫ (0,λ), one obtains the states with λ = λ 1, multiplied by positive matrix elements √2. The polarization vectors of momentum± k are obtained from those at rest by applying the boost matrices (4.24):

µ µ ǫ (k,s3)= B()ǫ (0,s3). (4.815) The zero-helicity polarization vector ǫµ(k, 0) is also called longitudinal polarization vector, that exists for vector particles if these have a mass. The three boosted polarization vectors satisfy the orthogonality relations

∗ µ ′ ǫµ(k,s3) ǫ (k,s ) = δs s′ . (4.816) 3 − 3 3 In order to find the completeness relations, we boost the polarization vectors at rest µ 3 (4.814) to a final momentum k (ωk, 0, 0,k ) in the z-direction using the matrix (4.15), and we obtain ≡ 0 k3 1 1 0 ǫµ (k3ˆz, 1) =   , ǫµ (k3ˆz, 0) =   , (4.817) H ± ∓ i H M 0  ±     0   ω     k      where zˆ is the unit vector in the z-direction. We can now calculate the completeness sum k32 0 0 0 1 0 100 ǫµ(k3ˆz,s )ǫν(k3ˆz,s )∗ =   . (4.818) 3 3 2 0 010 s3 M   X  2   0 00 ω   k    By rewriting the right-hand side in the form k0k0 00 0 1 0 10 0 gµν +   , (4.819) − M 2 0 01 0    3 3   0 00 k k    4.17 Massive Vector Fields 353 we recognize the general covariant form for any direction of the momentum k:

kµkν µ k ν k ∗ µν µν ǫ ( ,s3)ǫ ( ,s3) = P (k) g 2 . (4.820) s ≡ − − M ! X3 If the vectors are first boosted into the z-direction and then rotated into the direction of their momentum:

kˆ = (sin θ cos φ, sin θ sin φ, cos θ), (4.821) with the polar angles θ,φ, we obtain the polarization vectors in the helicity basis. The subsequent rotation into the direction k by the four-dimensional extension of the matrix (4.328) leads to the polarization vectors of helicities λ = 1 and 0: ±

µ 0 µ 1 k ǫH (k, 1) = , ǫH (k, 0) = | | , (4.822) ± ¯(k, 1) M ωkkˆ ± ! !

where ¯(k, 1) are the three-dimensional polarization vectors (4.331). The four- ± µ dimensional polarization vectors ǫH (k, 1) agree with (4.318), and the vector µ ± ǫH (k, 0) denotes the longitudinal polarization vector associated with the three- i i dimensional ǫH (k, 0) = ωkk /M. The covariant completeness sum (4.820) can be derived directly from the general polarization vectors (4.822). It reads explicitly

2 2 0 i µ ˆ ν ˆ ∗ k /M k k ǫ (k,s3)ǫ (k,s3) = | i | 0 2 ij i j 2 , (4.823) s k k /M δ + k k /M ! X3 which is the rotated version of (4.819). It will sometimes be convenient to view the photon as an M 0 -limit of a massive vector meson. For this purpose we have to add a gauge fixin→g term to the Lagrangian to allow for a proper limit. The extended action reads

4 4 1 µν 1 2 µ 1 µ 2 = d x (x)= d x FµνF + M VµV (∂µV ) , (4.824) A Z L Z −4 2 − 2α  resulting in the field equation 1 ∂ F µν + M 2V ν + ∂ν ∂ V µ =0, (4.825) µ α µ which reads more explicitly

2 2 1 µ ( ∂ M )gµν + 1 ∂µ∂ν V (x)=0. (4.826)  − −  − α  ν Multiplying (4.825) with ∂ν from the left gives, for the divergence ∂ν V , the Klein- Gordon equation 2 2 ν (∂ + αM )∂νV (x)=0, (4.827) from which the constraint (4.812) follows in the limit of large α. 354 4 Free Relativistic Particles and Fields

4.18 Higher-Spin Representations

Given the fundamental spin-1/2 field, it is very simple to generalize the transfor- mation matrices to higher spins. A system with two spin-1/2 particles can have spin 1 or 0. Similarly, n spin-1/2 particles can couple to spin n/2, n 1/2,... down to 1/2 or 0. Thus in order to build an arbitrary spin s, all we have− to do is put 2s spin-1/2 representations together in an appropriate fashion. The problem is completely analogous to the previous extension in Section 2.5 of the one-particle Schr¨odinger equation to an arbitrary n particle equation. Thus we shall construct − representations of arbitrary spin by a “second quantization” of spin, and further of the generators of the entire Lorentz group.

4.18.1 Rotations If the particle is at rest, spin is defined by the rotation subgroup. The 2 2 Hermitian generators × 1

L =  (4.828) 2 may be considered as the analog of the single-particle Schr¨odinger operator ¯h2 2 2M ∂x + V1(x; t) in Eq. (2.93), or as the matrix Mi in the commutation rules −(2.99). According to (2.101), the second quantized version of L reads 1 Lˆ = aˆ† σi a,ˆ (4.829) i 2

† wherea ˆα anda ˆα with α =1, 2 are two bosonic creation and annihilation operators. As proved in general in (2.102), these operators satisfy the same commutation rules as the Pauli matrices σi:

[Lˆi, Lˆj]= iǫijkLˆk. (4.830)

The states 1 , 1 aˆ† 0 , 1 , 1 aˆ† 0 (4.831) | 2 2 i ≡ 1| i | 2 − 2 i ≡ 2| i may be identified as the two basis states of the fundamental spin-1/2 representation. The effect of the three operators Lˆi is, as in the general Eq. (2.103):

† † 1 i Lˆ , aˆ =a ˆ ′ σ ′ , i α α 2 α α h i 1 i aˆ ′ , Lˆ = σ ′ aˆ . (4.832) α i 2 α α α h i On these states, the second-quantized operators (4.829) have eigenvalues

Lˆ s,s = s(s + 1) s,s , Lˆ s,s = s s,s . (4.833) | 3i | 3i 3| 3i 3| 3i In the present restriction to the rotation group we shall use only lower indices 1 and 2 rather than Weyl indices of the previous section. 4.18 Higher-Spin Representations 355

As in Section 2.5, we may now compose all higher representations of the rotation group by combining many of these fundamental representations, and forming states such as 2s (a† ) 0 . (4.834) αi | i iY=1 It is easy to see that linear combinations of such states with a fixed number of a† form an invariant representation space of the rotation group. The reason is that the three generators Lˆi commute with the ”particle” number operator N =ˆa†a.ˆ (4.835)

In fact, a little algebra shows that the Casimir operator of the rotation group Lˆ 2, which characterizes an irreducible representation, is equal to 1 1 1 Lˆ 2 = (ˆa† σi aˆ)2 = Nˆ Nˆ +1 , (4.836) 4 2 2  showing explicitly that under rotations, states built from a fixed number 2s of spin creation operators a† always maintain the same number 2s, and that the eigenvalue of the Casimir operator in the space of 2s particles is s(s + 1). Thus the space is indeed invariant under rotations. But it is also an irreducible representation of spin s. To see this take the complete set of basis states in the space

1 † n1 † n2 n1, n2) = (ˆa1) (ˆa2) 0 , (4.837) | √n1!n2! | i

† † which diagonalize the occupation numbers of the two operators a1 and a2, and which are normalized to unity. Applying to these the operator σ3 1 Lˆ =ˆa† aˆ = (ˆa† aˆ a† a ), (4.838) 3 2 2 1 1 − 2 2 we see that it measures the difference in the number of “spin up” and “spin down” † † particles created by a1 and a2, respectively: 1 L n , n )= (n n ) n , n ) m n , n ). (4.839) 3| 1 2 2 1 − 2 | 1 2 ≡ | 1 2

Thus the state has the azimuthal (“magnetic”) quantum number m = n1 n2. The operators − Lˆ Lˆ + iLˆ , Lˆ Lˆ iLˆ , (4.840) + ≡ 1 2 − ≡ 1 − 2 on the other hand, have the form

ˆ † 0 1 † L+ =a ˆ aˆ =ˆa1 aˆ2, 0 0 !

ˆ † 0 0 † L− =a ˆ aˆ =ˆa2 aˆ1. (4.841) 1 0 ! 356 4 Free Relativistic Particles and Fields

Among each other and with Lˆ3, they satisfy the commutation rules

[Lˆ−, Lˆ+] = 2Lˆ3,

[Lˆ3, Lˆ+] = Lˆ+, [Lˆ , Lˆ ] = Lˆ . (4.842) 3 − − − They remove a spin down while adding a spin up and vice versa. Their matrix elements are

Lˆ n , n ) = (n + 1)n n +1, n 1), +| 1 2 1 2| 1 2 − q Lˆ n , n ) = n (n + 1) n 1, n + 1). (4.843) −| 1 2 1 2 | 1 − 2 q It is now obvious that starting from an arbitrary state, say n, 0) with no number | of spin-down particles and eigenvalue n/2 of Lˆ3, we can reach every other state n k, k) with k = 1,...,n by repeated application of Lˆ . The process ends at | − − 0, n), where Lˆ n, 0) = 0. This proves the irreducibility of the representation. | −| Conventionally, the highest eigenvalue of Lˆ3 is referred to as the value s of angular momentum: s = n/2. (4.844) It appears in the eigenvalue (4.836) of Lˆ 2 as

Lˆ 2 n k, k)= s(s + 1) n k, k). (4.845) | − | − Hence the states may be reexpressed in terms of the numbers s and m as

1 † s+m † s−m s, m = s + m, s m)= (ˆa1) (ˆa2) 0 , (4.846) | i | − (s m)!(s + m)! | i − q and the matrix elements (4.839) and (4.843) read

Lˆ s, m = m s, m , (4.847) 3| i | i Lˆ s, m = (s m)(s + m + 1) s, m +1 , (4.848) +| i − | i q Lˆ s, m = (s + m)(s m + 1) s, m 1 . (4.849) −| i − | − i q

For smaller spin values the effect of the raising and lowering operators Lˆ+ and Lˆ− upon the states s, m is illustrated in Fig. 4.3. | i Note that we could have defined the states (4.846) with arbitrary phase factors iφ1 iφ2 † † e and e accompanying a1 and a2. Then the application of L+ and L− would produce phase factors e−i(φ1−φ2) and ei(φ1−φ2) in (4.848) and (4.849), respectively. It is easy to verify that these phases drop out in the commutation rules (4.842) of the rotation group, and thus in (4.830). The choice of a positive square-root in (4.848) and (4.849) without such extra phase factors is known as the Condon-Shortley phase convention. 4.18 Higher-Spin Representations 357

Figure 4.3 Effect of raising and lowering operators Lˆ and Lˆ upon the states s,m . + − | i 4.18.2 Extension to Lorentz Group It is now quite simple to extend this construction of spin representations of the rotations to the entire Lorentz group. For this we deduce from the commutation rules between the generators Li and Mi in Eq. (4.76)–(4.78) that the combinations Jˆ1 = (Lˆ + iMˆ )/2, Jˆ2 = (Lˆ iMˆ )/2 (4.850) − have the commutation rules ˆ1 ˆ2 Ji , Jj = 0, h ˆ1 ˆ1i ˆ1 Ji , Jj = iǫijkJk , (4.851) h ˆ2 ˆ2i ˆ2 Ji , Jj = iǫijkJk . h i Therefore they generate two independent sets of rotations. Extending the previous construction, we now form the second-quantized operators

Jˆ(1) =a ˆ†  a,ˆ 2 Jˆ(2) = ˆb†  ˆb, (4.852) 2

and obtain  Lˆ = Jˆ(1) + Jˆ(2) =ˆa†  aˆ + ˆb† ˆb, (4.853)

2 2  1 1  Mˆ = (Jˆ(1) Jˆ(2))= aˆ† aˆ ˆb† ˆb . (4.854) i − i  2 − 2  These operators correspond to forming the second-quantized operators

1 † ˆ† aˆ Lˆ = (ˆa , b ) ¦ , (4.855) 2 ˆb !

ˆ 1 † ˆ† aˆ M = (ˆa , b ) i« , (4.856) 2 ˆb ! 358 4 Free Relativistic Particles and Fields with the 4 4 -representation matrices (4.514) and (4.516) of the Lorentz group. For a clearer× display of the Lorentz transformation properties one may take advantage of the Weyl notation. From (4.853) and (4.854) we see that the operators † † ( 1 ,0) (0, 1 )

a and b transform according to the fundamental representations D 2 and D 2 ,  since the 2 2 -matrices between them are for rotations  /2, i /2 and for Lorentz

× − †  boosts /2, i /2, respectively. Thus we may write them in the Weyl notation as aα and b† β˙ . By applying to the operators Jˆ(1) and Jˆ(2) the same arguments as to the rotation group, we can now easily see that the set of states

1 † a † a ˙ b ˙ b a a b b n1 n2 ˆ† 1 n1 ˆ† 2 n2 n1, n2, n1, n2 = (ˆa1) (ˆa2) (b ) (b ) 0 (4.857) | i a a b b | i n1!n2!n1!n2! q a a a b b b for fixed numbers n = n1 + n2 and n = n1 + n1 are irreducible representation spaces of the whole Lorentz group. They are denoted by na nb (s1,s2) , . (4.858) ≡ 2 2 !

The states (4.857) live in the direct-product space of the two operators Jˆ(1) and Jˆ(2). They may be relabeled by the quantum numbers of the two rotation subgroups as in (4.846): 1 s1, m1; s2, m2 = | i (s m )!(s + m )!(s m )!(s + m )! 1 − 1 1 1 2 − 2 2 2 q ˙ ˙ (ˆa† )s1+m1 (ˆa† )ns1−m1 (ˆb† 1)s2+m2 (ˆb† 2)s2−m2 0 . (4.859) × 1 2 | i Since by (4.853) the operator of angular momentum Lˆ is the direct sum of those of the two rotation subgroups, an irreducible representation D(s1,s2)(Λ) of the Lorentz group contains different irreducible representations of the rotation subgroup gen- erated by Lˆ. They are obtained from the rules of addition of angular momenta. 2 The operator Lˆ and the third component Lˆ3 can be diagonalized with eigenvalues s(s + 1) and m by forming the linear combination

s, m = s1, m1; s2, m2 s1, m1; s2, m2 s, m , (4.860) | i m ,m | ih | i X1 2 where s , m ; s , m s, m are Clebsch-Gordan coefficients [18]. Their calculation h 1 1 2 2| i and properties are recalled in Appendix 4E. The values of total angular momentum s occurring in the decomposition (4.860) are

s s s s + s . (4.861) | 1 − 2| ≤ ≤ 1 2 The Clebsch-Gordan coefficients are orthogonal and complete, so that (4.860) can be inverted to

s1, m1; s2, m2 = s, m s, m s1, m1; s2, m2 . (4.862) | i s,m | ih | i X 4.18 Higher-Spin Representations 359

4.18.3 Finite Representation Matrices To complete this discussion let us calculate the finite representation matrices. Due to the decompositions (4.853) and (4.854) into generators of rotations, we only need those of the rotation group.

Rotation Group

We first observe that every 3 3 -rotation matrix R³ˆ (ϕ) in Eq. (4.9) can be decom- posed into Euler angles ×

−i³·L −iαL3 −iβL2 −iγL3

R³ (ϕ)= e = e e e R(α,β,γ), (4.863) ˆ ≡ and so can the general rotation operator

ˆ ³ ˆ ˆ ˆ e−i ·L = e−iαL3 e−iβL2 e−iγL3 Rˆ(α,β,γ). (4.864) ≡ ˆ −i³·L Since Lˆ3 is diagonal on the states jm , the finite rotation e acts on the states jm as | i | i j ′ ˆ Rˆ(α,β,γ) jm = jm′ e−i(m α+mγ) jm′ e−iβL2 jm | i ′ | i h | | i mX=−j j ′ j jm Dm′,m(α,β,γ), (4.865) ≡ ′ | i mX=−j so that the only nontrivial matrix elements are

j ′ −iβLˆ2 d ′ (β)= jm e jm . (4.866) m m h | | i For a single creation operatora ˆ†, we have from (4.832): 1 i [Lˆ , aˆ† ] = [ˆa†aˆ aˆ†aˆ , aˆ†]= aˆ† , 2 1 2i 1 2 − 2 1 1 2 2 1 i [Lˆ , aˆ† ] = [ˆa†aˆ aˆ†aˆ , aˆ†]= aˆ†, (4.867) 2 2 2i 1 2 − 2 1 2 −2 1 and therefore

ˆ ˆ β β e−iβL2 aˆ†eiβL2 =a ˆ† cos +ˆa† sin , 1 1 2 2 2 ˆ ˆ β β e−iβL2 aˆ†eiβL2 = aˆ† sin +ˆa† cos . (4.868) 2 − 1 2 2 2 Of course, this is just the statement thata ˆ† 0 anda ˆ† 0 are transformed according 1| i 2| i to the spin-1/2-representation of the rotation group [recall (4.456)]:

α′ cos β sin β −iβLˆ2 † iβLˆ2 † 2 2 e a e =a ˆ ′ − α α  β β  sin 2 cos 2 α  ′  † −iβσ2/2 α =a ˆα′ e α . (4.869)   360 4 Free Relativistic Particles and Fields

An arbitrary state (4.846) goes over into 1 e−βL2 jm = | i (j m)!(j + m)! − q j+m j−m † β † β † β † β aˆ1 cos +ˆa2 sin aˆ1 sin +ˆa2 cos 0 . (4.870) × 2 2 ! − 2 2 ! | i

We now expand the right-hand side into a sum of products of two creation operators. After ordering the terms, we rewrite the rotated state as

′ ′ −βL2 1 † j+m † j−m j e jm = (ˆa1) (ˆa2) 0 dm′m(β). (4.871) ′ ′ | i m′ (j m )!(j + m )! | i X − q This defines the matrix elements of the rotations around the second axis by an angle β:

′ ′ ∞ j (j + m )!(j m )! j + m j m d ′ (β) = m m v − ′ − u (j + m)!(j m)! j m k ! k ! u − kX=0 − − t ′ ′ β 2k+m +m β 2j−2k−m −m ( 1)j−m−k cos sin . (4.872) × − 2 ! 2 !

The sum can be expressed in terms of hypergeometric functions

ab a(a + 1) b(b + 1) z2 F (a, b, c; z) 1+ z + + ..., (4.873) ≡ c c(c + 1) 2! in terms of which they read

′ ′ ′ m −m ′ m +m m −m j ( 1) (j m)!(j + m )! β β d ′ (β) = − − cos sin m m ′ v ′ (m m)!u(j + m)!(j m )! 2 ! 2 ! − u − t β F j + m′, j + m′ + 1; m′ m + 1; sin2 . (4.874) × − − 2 !

This formula is directly applicable for m′ m, where the hypergeometric function is regular at the origin. For m′ < m we use≥ the property

j m′−m j m′−m j d ′ (β)=( 1) d ′ (β)=( 1) d ′ (β) (4.875) m m − mm − −m −m to exchange the order. Additional useful relations are

j j−m′ j d ′ (β π)=( 1) d ′ (β) (4.876) m m − − −m m and j j+m′ d ′ (π)=( 1) δ ′ . (4.877) m m − m ,−m 4.18 Higher-Spin Representations 361

¡The hypergeometric functions can also be expressed in terms of Jacobi polyno- mials: β F j + m′, j + m′ + 1; m′ m + 1; sin2 − − 2 !

′ ′ ′ (j m )! ′ (m −m,m +m) = − (m m)!P ′ (cos β). (4.878) (j m)! − j−m − l The matrix elements d00(β) coincide with the Legendre polynomials,

l d00(β)= Pl(cos β), (4.879)

l iγ and the matrix elements dm0(β)e are proportional to the spherical harmonics Ylm(β,γ):

l iγ m l iγ 4π d (β)e =( 1) d (β)e = Ylm(β,γ). (4.880) m0 − 0m s2l +1 1 For j = 2 we reobtain the spinor representation of the rotation group

1/2 cos β/2 sin β/2 −iβσ2/2 dm′m(β)= − = e , (4.881) sin β/2 cos β/2 ! while for j = 1 we find the vector representation

1 √ 1 2 (1 + cos β) sin β/ 2 2 (1 cos β) 1 − − d ′ (β)= sin β/√2 cos β sin β/√2 . (4.882) m m  −  1 (1 cos β) sin β/√2 1 (1 + cos β)  2 2   −  The indices have the order +1/2, 1/2 and +1, 0, 1, respectively. The representation functions (4.865)− of all rotations−

j −i(mα+m′γ) j Dm,m′ (α,β,γ)= e dm,m′ (β) (4.883) have the following orthonormality properties:

2π π 2π 2j +1 j1 j2 ′ ′ ′ ′ dα dβ dγ Dm ,m (α,β,γ)Dm ,m (α,β,γ)= δj1,j2 δm1,m2 δm ,m . (4.884) 8π2 1 1 2 2 1 2 Z0 Z0 Z0

′ ′ At equal m1 = m2, these yield the integrals

2π π 2j +1 j1 j2 ′ ′ dα dβ Dm ,m (α, β, 0)Dm ,m (α, β, 0) = δj1,j2 δm1,m2 . (4.885) 4π 1 1 2 1 Z0 Z0

j The representation matrices Dm,m′ (α,β,γ) with j = 1 are related to the original 3 3 -rotations R(α,β,γ) of Eq. (4.863) by a similarity transformation. It is the same × transformation which relates the three spherical components "(λ) in Eq. (4.321) to 362 4 Free Relativistic Particles and Fields the unit vectors (4.325). The eigenvectors (4.321) supply us with the matrix elements of the desired similarity transformation. Identifying the scalar products i 1, m as h | i spherical components of a vector:

i 1, m ǫi(m), (4.886) h | i ≡ we can write, using the 3 3 matrices (4.54) × 1 (L ) = j 1, m m 1, m k , 3 jk h | i h | i mX=−1 (L ) = j 1, 1 √2 1, 0 k + j 1, 0 √2 1, 1 k . (4.887) ± jk h | ± i h | i h | i h ∓ | i In Dirac’s bracket notation, the original 3 3-matrices R(α,β,γ) in (4.863) may be considered as matrix elements of the general× rotation operator Rˆ(α,β,γ) in (4.864) between the basis states i : | i R (α,β,γ)= i Rˆ(α,β,γ) j . (4.888) ij h | | i From the manipulation rules of Dirac brackets it is then obvious that the matrix elements transform under finite rotations as

3 3 R j 1, m = i Rˆ(α,β,γ) j j 1, m ij h | i h | | ih | i jX=1 jX=1 3 1 = i 1, m′ 1, m′ Rˆ(α,β,γ) j j 1, m (4.889) ′ h | ih | | ih | i jX=1 mX=−1 1 1 ′ ′ ˆ ′ 1 = i 1, m 1, m R(α,β,γ) 1, m = i 1, m Dm′,m(R), ′ h | ih | | i ′ h | i mX=−1 mX=−1 which may also be written in a matrix form as12

1

′ 1 ¯ R ¯(m)= (m )Dm′,m(R). (4.890) ′ mX=−1 In Eq. (4.112) we stated the transformation law of a vector operator [see also Eq. (2.113)]: [Lˆi, vˆk]=ˆvj(Li)jk, (4.891) With the help of the above similarity transformation, we find the spherical compo- nents of the vector operatorv ˆi:

3 vˆ(m) vˆ i 1, m , (4.892) ≡ i h | i Xi=1 12Note that the spinor transformation laws (4.745) and (4.747) are a generalization of this relation. 4.18 Higher-Spin Representations 363 or, explicitly, 1 vˆ( 1) (ˆv1 ivˆ2), vˆ(0) vˆ3. (4.893) ± ≡ ∓√2 ± ≡ For these components, the commutation rules (4.891) become

1 ′ ′ [Lˆi, vˆ(m)] = vˆ(m ) 1, m Lˆi 1, m . (4.894) ′ h | | i mX=−1 They may be generalized to an arbitrary spherical tensor operator vˆ(j, m) of spin j:

1 ′ ′ [Lˆi, vˆ(j, m)] = vˆ(j, m ) j, m Lˆi j, m . (4.895) ′ h | | i mX=−1

For Lˆ3 and Lˆ±, these commutation relations become

[Lˆ , vˆ(j, m)]=v ˆ(j, m) m, [Lˆ , vˆ(j, m)]=v ˆ(j, m 1) (j m)(j m + 1). (4.896) 3 ± ± ∓ ± q They are in one-to-one correspondence with the relations (4.847), (4.848), and (4.849) for the states j, m . For finite rotations,| theyi give rise to the transformation behavior

ˆ ˆ−1 ′ j R vˆ(j, m)R =v ˆ(j, m )Dm′m(R). (4.897) The use of defining such spherical tensor operators lies in the fact that all their matrix elements are related to each other by Clebsch-Gordan coefficients (4.860) via the so-called Wigner-Eckart theorem. Applyingv ˆ(j, m) to a state j′, m′ , we obtain a statev ˆ(j, m) j′, m′ , which transforms by a direct product of the| representationi matrices (4.865)| andi (4.897) like a state j, m; j′, m′ . Its irreducible contents can be | i obtained with the help of the Clebsch-Gordan series (4.860). If we therefore expand

vˆ(j, m) j′, m′ = j′′, m′′ j′′, m′′ vˆ(j, m) j′, m′ , (4.898) | i ′′ ′′ | ih | | i j X,m the matrix elements j′′, m′′ vˆ(j, m) j′, m′ must be proportional to the associated h | | i Clebsch-Gordan coefficients:

j′′, m′′ vˆ(j, m) j′, m′ = j′′ v(j) j′ j′′, m′′ j, m; j′, m′ . (4.899) h | | i h || || ih | i The proportionality constants j′′ v(j) j′ are independent of the azimuthal quan- h || || i tum numbers m, m′, m′′. They are called the reduced matrix elements of the spherical tensor operatorv ˆ(j, m). They vanish if j′′ does not satisfy the vector coupling con- dition j j′ j′′ j + j′. | − | ≤ ≤ j −iσ2π/2 2 For j =1/2, the matrix dm′m(π) is equal to e = iσ . It is therefore the spin-j representation of the matrix c of (4.604), and will therefore− be denoted by

(j) j j+m′ c d ′ (c)=( 1) δ ′ . (4.900) ≡ m ,m − m ,−m 364 4 Free Relativistic Particles and Fields

j When applied to the representation matrix Dm′m(α,β,γ) as a similarity transfor- mation, we find a spin-j generalization of the important 2 2 -relation (4.689): ∗ × [c(j)]−1Dj(α,β,γ)c(j) = Dj (α,β,γ). (4.901) The matrix c(j) gives rise to an invariant bilinear form for any pair of spherical tensor operatorsv ˆ(j, m) andv ˆ′(j, m): j j ′ (j) ′ ′ j+m (ˆv, vˆ ) vˆ(j, m)cmm′ vˆ (j, m )= ( 1) vˆ(j, m)ˆv(j, m). (4.902) ≡ ′ − − m,mX=−j mX=−j This product remains invariant under rotations, since DjT (α,β,γ)c(j)Dj(α,β,γ)= c(j). (4.903) The invariance of (4.902) is a generalization to spin-j operators of the Weyl invari- ′ βα ance of the spinor product ξβ c ξα of (4.794). For j = 1, the invariant product (4.902) is equivalent to the ordinary scalar product. This is seen by replacing the spherical componentsv ˆ(1, m) by the cartesian ones on the right-hand side of (4.902) according to (4.893), yielding (ˆv, vˆ′)= δ vˆ vˆ′ . (4.904) − ij i j Also for j = 1, the spherical relation (4.903) is equivalent to the invariance of the Kronecker symbol δ under rotations in the 3 3 defining representation: ij × R ′ R ′ δ ′ ′ = δ . (4.905) i1i1 i2i2 i1i2 i1i2 The invariance of scalar products (4.902) formed with c(j) can be used to extend the Weyl calculus to spin-j objects as follows: The spherical tensor operatorv ˆ(j, m) j is written asv ˆ( m ), and a contravariant spherical tensor operator is introduced as follows: j m (j) j j+m j ′ vˆ( j ) cm,m vˆ( m′ ) =( 1) vˆ( −m ). (4.906) ≡ ′ − mX=−j Then the invariant form (4.902) can simply be written as

′ j ′ m (ˆv, vˆ )=ˆv( m ) vˆ ( j ), (4.907) with the convention that pairs of upper and lower indices m are assumed to be summed. The relation between the axis-angle representation and the Euler-angle form of the rotations on the two sides of (4.863) is easily found by comparing the explicit 2 2 -representations (4.469) of the two forms: × 3 2 3 e−iασ /2e−iβσ /2e−iγσ /2 α α β β γ γ = cos iσ3 sin cos iσ2 sin cos iσ3 sin  2 − 2  2 − 2 !  2 − 2  β α + γ β α γ = cos cos + sin − iσ1 2 2 2 2 β α γ β α + γ sin cos − iσ2 sin sin iσ3. (4.908) − 2 2 − 2 2 4.19 Higher Spin Fields 365

1 2 3 0 Comparing the coefficients of (σ , σ , σ )=  and σ = 1 gives ϕ α γ α γ α + γ β

³ˆ sin = sin − , cos − , sin sin , 2 − 2 2 2  2 ϕ β α + γ cos = cos cos . (4.909) 2 2 2 More details on the rotation group can be found in the textbook Ref. [18].

Lorentz Group To extend these results to the Lorentz group we make use of the fact that due to the decompositions (4.853) and (4.854), pure rotations can be decomposed as

ˆ ˆ(1) ˆ(2)

³ ³ e−i³·L = e−i ·J e−i ·J , (4.910) where J(1) and J(2) are the matrices (4.850). The pure Lorentz transformations are

ˆ ˆ(1) ˆ(2)

  e−i·M = e− ·J e ·J , (4.911)

ˆ(1) ˆ(2) 

with e− ·J e ·J having again matrix elements of rotations, as calculated above.  Thus, given the parameters ³ and of the Lorentz transformation in question, we merely have to find the corresponding Euler angles and take the corresponding rotation matrices from (4.865) and (4.874). Note that for pure Lorentz transformations the rotation parameters are imag- inary so that the trigonometric functions become hyperbolic. For pure Lorentz transformations with imaginary angles, the relation between the axis-angle and the Euler representations corresponding to the two sides of (4.863) is then given by rela- tions like (4.909), but with cosine and sine functions continued to the corresponding hyperbolic forms.

4.19 Higher Spin Fields

The construction of invariant actions can be generalized to fields of arbitrary spin. If we restrict ourselves to those representations which contain only one spin, the situation is very similar to the spin-1/2 case: There are two spinor fields of the β˙ (s,0) (0,s) Weyl-type ξα and η , transforming according to the D and D -representations, respectively

Λ ′ ′ (s,0) α ξ(x) ξ ′ (x ) = D (Λ) ′ ξ (x), −−−→ α α α Λ ′β˙′ ′ (0,s) β˙′ β˙ η(x) η (x ) = D (Λ) ˙ η (x). (4.912) −−−→ β Now, according to the last section, the matrices D(s,0)(Λ) are just the symmetrized ( 1 ,0) direct products of 2s representations D 2 (Λ). They satisfy the same relation as ( 1 ,0) D 2 itself: D(s,0)(Λ) = D(0,s)(Λ)−1. (4.913) 366 4 Free Relativistic Particles and Fields

Hence ξ†η , η†ξ (4.914) are the only Lorentz-invariant bilinear combinations of the spinor fields. What 1 about invariants involving derivatives? For this we recall that in the spin- 2 case, the expressions † µ † µ ξ iσ˜ ∂µξ , η iσ ∂µη (4.915) were invariant due to the property (4.482). The invariance remains true for a product of 2s factors whose right and left indices are symmetrized. Therefore

ξ†(i2sσ˜µ1 ... σ˜µ2s ∂ ∂ )ξ ξ†(˜σi∂){2s}ξ (4.916) × × µ1 ··· µ2s ≡ and

η†(i2sσµ1 ... σµ2s ∂ ∂ )η η†(σi∂){2s}η (4.917) × × µ1 ··· µ2s ≡ are invariants, where the curly brackets indicate the symmetrization of the indices. We may therefore write the action as

= d4x ξ†(˜σi∂){2s}ξ + η†(σi∂){2s}η M 2sξ†η M 2sη†ξ . (4.918) A − 1 − 2 Z n o In the absence of mass terms, each of the derivative pieces gives by itself an invari- ant action which maximally violates parity. This fact is essential for accommodating maximal parity violation into the weak interactions discussed on p. 314. The equa- tions of motion (4.487) and (4.488) become

{2s} 2s (˜σi∂) ξ(x) = M1 η(x), (4.919) {2s} 2s (σi∂) η(x) = M2 ξ(x). (4.920)

They can be inserted into each other to give

{2s} {2s} ξ(x) 2s 2s ξ(x) (˜σi∂) (σi∂) = M1 M2 . (4.921) ( η(x) ) ( η(x) )

The left-hand side contains a product of two symmetrized products. Since each 2s factors of the product are symmetric under simultaneous exchange of left and right indices, we can omit the symmetrization in the contracted indices, and use in each of them relation (4.490) to derive

[(˜σi∂)(σi∂)]{2s} =( ∂21){2s} =( ∂2)2s(1){2s}, (4.922) − − with (1){2s} being the unit matrix in the symmetrized subspace. In momentum space, this amounts to the mass shell relation

2 2s 2s 2s 2 2s (p ) = M1 M2 =(M ) . (4.923) 4.19 Higher Spin Fields 367

As in the spin-1/2 case, space inversion changes ∂ ∂˜ and the representation matrices → D(s,0)(Λ) D(0,s)(Λ). (4.924) → If one wants to have a representation space of the Lorentz group including space inversions, one must combine the two spinors ξ(x) and η(x) into a bispinor with 2 (2s + 1) components × ξ(x) ψ(x)= . (4.925) η(x) ! On this space, parity is represented as in (4.530) by

P ψ(x) ψ′ (x)= D(P )ψ(˜x), (4.926) −−−→ P with a representation matrix which looks like (4.531), but contains now four blocks of (2s + 1) (2s + 1) -matrices: × 0 1 D(P )= ηP = ηP γ0. (4.927) 1 0 !

It is obvious that this matrix changes the generators L and M of the Lorentz group in the spin-s representation as in Eqs. (4.529). Invariance under space inversion requires

2s M1 = M2 = M , (4.928) and thus the presence of both derivative terms in (4.918). The action (4.918) can now be reformulated in a Dirac-type form using generalized γ-matrices defined by

{2s} {2s} 0 (σi∂) (γi∂) = {2s} . (4.929) (˜σi∂) 0 !

With these we can write down a parity-invariant action for the bispinors ψ(x) as

= d4x ψ¯(x) (γi∂){2s} M 2s ψ(x), (4.930) A − Z h i where the conjugate bispinor ψ¯(x) reads

† ¯ η † 0 {2s} ψ(x) † = ψ (x)(γ ) . (4.931) ≡ ξ !

The field equation is (γi∂){2s} M 2s ψ(x)=0. (4.932) − h i 368 4 Free Relativistic Particles and Fields

4.19.1 Plane-Wave Solutions One can easily write down plane wave solutions of the spin-s wave equation (4.932): −ipx ipx e c e fp s3 (x) u(p,s3) , fp s (x) v(p,s3) , (4.933) ≡ Vp0/M 3 ≡ Vp0/M q q where u(p,s3) and v(p,s3) are the positive- and negative-energy solutions of momen- tum p and p, respectively, satisfying the generalized Dirac equations in momentum space − (/p {2s} M)u(p,s )=0, (/p {2s} + M)u(p,s )=0. (4.934) − 3 3 The second can be obtained from the first via a relation like (4.678): T v(p,s3)= Cu¯ (p,s3). (4.935) Here C is the charge conjugation matrix for arbitrary spin: {2s} {2s} 0 c 0 C = γ = {2s} . (4.936) 0 c ! The matrices c{2s}0 are equivalent to the matrices c(s) introduced in Eq. (4.900). They have the important property that c{2s} =( 1)2s. (4.937) − The rest spinors χ{2s}(s ) have symmetrized labels 1/2 and 1/2. These are uniquely 3 − specified by the number n1 of up-spins and n2 of down-spins, which are the labels of {2s} the basis vectors n1, n2) in Eq. (4.837). Thus we may write χ (s3) more explicitly {2s} | as χn1,n2 (s3). The label s3 specifies the eigenvalues of the third component of angular momentum, and corresponds to the label m of the basis vectors (4.846). Hence {2s} χn1,n2 (s3)= δs3,(n1−n2)/2, n1 + n2 =2s. (4.938) These spinors satisfy the obvious completeness relation {2s} {2s} ∗ ′ ′ χ (s3)χ (s3) = δn1,n2 . (4.939) n1,n2 n1,n2 s X3 {2s} Using this, we find that the spinors u (p,s3) have a polarization sum [compare (4.705)]

{2s} {2s} {2s} {2s} ∗ M + pσ P (p) u (p,s3)u (p,s3) = . (4.940) ≡ s 2M ! X3 It is a straightforward generalization of the Dirac case. The polarization sums for {2s} the spinors v (p,s3) can be calculated similarly. From the spinors w2s(p,s ) we form the mirror-reflected spinors w2s( p,s ) 3 − 3 {2s} {2s} {2s} {2s} pσ˜ w (p˜,s3)= B (ζ)χ (s3)= , (4.941) pˆ s M and after multiplication with the generalized charge-conjugation matrix c2s, we com- 2s 2s bine both spinors to bispinors u (p,s3) and v (p,s3) of particle and antiparticles of spin s. 4.20 Vector Field as a Higher-Spin Field 369

4.20 Vector Field as a Higher-Spin Field

Some remarks are useful concerning the field transformations under the representa- tion D(s1,s2)(Λ), with both s ,s = 0. They were omitted in the above discussion, 1 2 6 although the most prominent example is one of them. It is a spinor field with 1 β˙ s1 = s2 = 2 , which we denote by ξα (x). This field is equivalent to a vector field Aµ(x), which was discussed before in Section 4.6 for the massless case, and in Sec- tion 4.17 for a nonzero mass. To see this equivalence, we observe first of all that β˙ both representations have a spin content 0 and 1. For the spinor ξα (x) this fol- lows from the addition rule of angular momenta (4.861). In the vector field Aµ(x), the zeroth component transforms according to the spin-0 representation, the spatial components according to the spin-1 representation of the rotation group. There exists a simple relation between the two fields. The spinor field transforms 1 1 according to the 2 , 2 representation of the Lorentz group as follows:   Λ ′ ′ ′ ′ β˙ ( 1 ,0) α (0, 1 ) β˙ β˙ ξ ξ (x ) ′ = D 2 (Λ) ′ D 2 (Λ) ˙ ξ (x) −−−→ α α β α ( 1 ,0) (0, 1 )T β˙′ = D 2 (Λ)ξ(x)D 2 (Λ) . (4.942) α′ h i The 2 2 components of the spinor can be mapped into the four components of a × vector by forming

µ µ β˙′α β˙ µ ξ (x) c ˙ ˙′ σ˜ ξ (x) = tr[cσ˜ ξ(x)]. (4.943) ≡ ββ α Using the Lorentz transformation rules of Section 4.16 it is easy to verify that ξµ(x) transforms indeed like the vector field Aµ(x) in Eq. (4.290):

Λ µ ′µ ′ µ ( 1 ,0) (0, 1 )T ξ (x) ξ (x ) = tr cσ˜ D 2 (Λ)ξ(x)D 2 (Λ) −−−→ h −1 (0, 1 )T µ ( 1 ,0) i = tr c(c D 2 (Λ)c)˜σ D 2 (Λ)ξ(x) . (4.944) h i Now we make use of the relation (4.683) to set

−1 µT µ −1 (0, 1 )T ( 1 ,0)† c σ c = σ , c D 2 (Λ)c = D 2 (Λ), (4.945) − and to rewrite (4.944) as

′µ ′ ( 1 ,0)† µ ( 1 ,0) ξ (x ) = tr cD 2 (Λ)˜σ D 2 (Λ)ξ(x) . (4.946)   With the help of (4.482), we now obtain the vector property of the composite field ξm(x): ′µ ′ µ ν µ ν ξ (x )=Λ ν tr [cσ˜ ξ(x)]=Λ ν ξ (x), (4.947) so that ξµ(x) transforms indeed like Aµ(x) in (4.290). ′ A special feature of all representations D(s,s ) with s = s′ is that they are in- variant under space inversions since this interchanges s and s′. Thus no doubling of 1 1 fields is needed to accommodate space inversions. In the vector form of the 2 , 2 representation this was observed before in the transformation law (4.305).   370 4 Free Relativistic Particles and Fields

4.21 Rarita-Schwinger Field for Spin 3/2

Another frequently-encountered form of higher spin fields which is not of the (s, 0)+ (0,s) type is due to Rarita and Schwinger and describes spin-3/2 particles [14]. It combines vector and bispinor properties and is written as ψµa(x), thus transforming according to

( 1 ,0) Λ ′ ′ ν D 2 (Λ) 0 ψµa(x) ψµa(x )=Λµ 1 ψνb(x). (4.948) (0, 2 ) −−−→ 0 D (Λ) !ab

1 1 Group-theoretically speaking, this is a direct product of the representations 2 , 2 1 1   (for the indices µ, ν) and 2 , 0 + 0, 2 (for the indices a, b). We can employ the usual rules  for the addition of angular momentum and apply them to J(1) and J(2) in (4.853). Then the direct product of two representations

(s ,s ) (s′ ,s′ ) D 1 2 D 1 2 (4.949) × must have the following irreducible contents:

(|s −s |,|s′ −s′ |) (|s −s |,|s′ −s′ |+1) (|s −s |,s′ +s′ ) D 1 2 1 2 + D 1 2 1 2 + ... + D 1 2 1 2 |s −s |+1,|s′ −s′ |) (|s −s |+1,|s′ −s′ |+1) (|s −s |+1,s′ +s′ ) + D 1 2 1 2 + D 1 2 1 2 + ... + D 1 2 1 2 + ... (s +s ,|s′ −s′ |) (s +s ,|s′ −s′ |+1) (s +s ,s′ +s′ ) + D 1 2 1 2 + D 1 2 1 2 + ... + D 1 2 1 2 .

1,2 In this expansion, the spins s1 and s2 of J combine to all spins from s1 s2 to ′ ′ | −′ | ′ s1 +s2 [recall (4.861)]. Similarly, the spins s1 and s2 couple to all spins from s1 s2 ′ ′ (s1,s2) | − | to s1 + s2. Therefore ψµa is equivalent to a sum of D representations: 1 1 1 1 0, + 1, + , 0 + , 1 . (4.950)  2  2 2   2 

Remember that the symmetry with respect to the interchange s1 s2 is necessary for a parity-invariant Lagrangian. Now, if we want to describe↔ only a spin-3/2 1 1 particle, the representations (0, 2 )and( 2 , 0) are superfluous and have to be projected out. This can be done by a constraint analogous to the Lorentz condition for the electromagnetic field: µ ∂ ψµa(x)=0. (4.951) 1 1 Obviously, this derivative transforms like (0, 2 )+( 2 , 0) and has only a spin-1/2 content, which is therefore removed from (4.950). It remains to make sure that 1 1 the representation (1, 2 )+( 2 , 1) in ψµa describes only a spin-3/2 particle. This is achieved by another condition imposed on the field:

µ γ abψµb =0. (4.952)

1 1 The associated projection of the field transforms once more like (0, 2 )+( 2 , 0), and 1 setting it equal to zero eliminates one more spin- 2 degree of freedom, thus ensuring the survival of only the spin-3/2 content in ψµa. Appendix 4A Derivation of Baker-Campbell-Hausdorff Formula 371

Finally we construct an invariant action with the property that the equations of motion automatically satisfy the constraints (4.951) and (4.952). There are now several possible invariants which can be used. If we allow at most a single derivative, we may combine

µ ν µ ν ν µ µ ν µ ψ¯ iγ ∂νψµ, ψ¯µγ ∂ ψν , ψ¯µγ ∂ ψν , ψ¯µγ γ ψν, ψ¯ ψµ. (4.953)

The most general combination which leads to a Hermitian action of a pure spin-3/2 particle can be shown to be13

4 4 µν = d x (x)= d x ψ¯µ(x)L (i∂)ψν (x). (4.954) A Z L Z Here Lµν (i∂) is the differential operator

Lµν (i∂) = (i/∂ M)gµν + wγµi∂ν + w∗γν∂µ − 1 + (3ww∗ + w + w∗ + 1)γµi/∂γν 2 3 +M 3ww∗ + (w + w∗)+1 γµγν, (4.955)  2  and w is an arbitrary . The equations of motion are given by

µν L (i∂)ψν (x)=0. (4.956)

It can easily be verified that a field ψµa(x), which satisfies the constraints (4.951) and (4.952), solves (4.956) if and only if the Dirac equation is fulfilled separately for each vector index µ: a′ (i/∂ M) ψ ′ (x)=0. (4.957) − a µa The particle has obviously a mass M. Some algebra is necessary to deduce that the constraints (4.951) and (4.952) follow from (4.956). For this we go to momentum µν space and contract L (p)ψν (p) = 0 once with γµ and once with pµ, using the µ µ relations γµ /p = /pγµ +2pµ and γ γµ = 4. The two contractions yield γ ψµ = 0 µ − and p ψµ = 0, which are a direct consequence of the anticommutation rules (4.496).

Appendix 4A Derivation of Baker-Campbell-Hausdorff Formula

The standard Baker-Campbell-Hausdorff formula, from which our formula (4.74) can be derived, reads ˆ ˆ ˆ eAeB = eC , (4A.1) where

1 Cˆ = Bˆ + dtg(eadA teadB)[Aˆ]. (4A.2) Z0 13See Notes and References for literature. 372 4 Free Relativistic Particles and Fields

Here g(z) is the function ∞ log z (1 z)n g(z) = − (4A.3) ≡ z 1 n +1 n=0 − X and adB is the operator associated with Bˆ in the so-called adjoint representation, which is defined by adB[Aˆ] [B,ˆ Aˆ]. (4A.4) ≡ One also defines the trivial adjoint operator (adB)0[Aˆ] = 1[Aˆ] Aˆ. By expanding the exponentials in Eq. (4A.2) and using the power series (2A.3), one finds the≡ explicit formula

∞ n ˆ ˆ ˆ ( 1) 1 C = B + A + − n n +1 1+ pi n=1 i i i i i=1 X p ,q ;Xp +q ≥1 (adA)p1 (adB)q1 (adA)Ppn (adB)qn [Aˆ]. (4A.5) × p1! q1! ··· pn! qn! The lowest expansion terms are 1 Cˆ = Bˆ + Aˆ 1 adA + adB + 1 (adA)2 + 1 adA adB + 1 (adB)2 +... [Aˆ] − 2 2 6 2 2 1   + 1 (adA)2 + 1 adA adB + 1 adBadA + (adB)2 + ... [Aˆ] 3 3 2 2  1 1  = Aˆ + Bˆ [B,ˆ Aˆ]+ ([A,ˆ [A,ˆ Bˆ]] + [B,ˆ [B,ˆ Aˆ]]) + .... (4A.6) − 2 12 To prove formula (4A.2) and thus the expansion (4A.5), we proceed in a way similar to the derivation of the interaction formula (1.303). We derive and solve a differential equation for the operator function ˆ ˆ Cˆ(t) = log(eAteB). (4A.7) This determines the function Cˆ(t) from its value Cˆ(1) at t = 1. The starting point is the observation that for any Mˆ , ˆ ˆ eC(t)Meˆ −C(t) = eadC(t)[Mˆ ], (4A.8) by the definition of adC. The left-hand side can also be rewritten as

ˆ ˆ ˆ ˆ eAteBMeˆ −Be−At = eadA teadB[Mˆ ], (4A.9) so that we have eadC(t) = eadA teadB. (4A.10) Differentiation of (4A.7) shows that

ˆ d ˆ eC(t) e−C(t) = A.ˆ (4A.11) dt − The left-hand side, on the other hand, can be rewritten in general as

ˆ d ˆ ˙ eC(t) e−C(t) = f(adC(t))[Cˆ(t)], (4A.12) dt − where ez 1 f(z) − . (4A.13) ≡ z This will be verified below. It implies that ˙ f(adC(t))[Cˆ(t)] = A.ˆ (4A.14) Appendix 4B Wigner Rotations and Thomas Precession 373

We now define the function g(z) as in (4A.3) and see that it satisfies

g(ez)f(z) 1. (4A.15) ≡ We therefore have the trivial identity

˙ ˙ Cˆ(t)= g(eadC(t))f(adC(t))[Cˆ(t)]. (4A.16)

Using (4A.14) and (4A.10), this turns into the differential equation

˙ Cˆ(t)= g(eadC(t))[Aˆ]= eadA teadB[Aˆ], (4A.17) from which we find directly the result (4A.2). To complete the proof we must verify (4A.12). For this consider the operator

ˆ d ˆ Oˆ(s,t) eC(t)s e−C(t)s. (4A.18) ≡ dt Differentiating this with respect to s gives

ˆ d ˆ ˆ d ˆ ∂ Oˆ(s,t) = eC(t)sCˆ(t) e−C(t)s eC(t)s Cˆ(t)e−C(t)s s dt − dt ˆ ˙  ˆ    = eC(t)sCˆ(t)e−C(t)s − ˙ = eadC(t)s[Cˆ(t)]. (4A.19) − Hence

s ′ ′ Oˆ(s,t) Oˆ(0,t) = ds ∂ ′ Oˆ(s ,t) − s Z0 ∞ n+1 s n ˙ = (adC(t)) [Cˆ(t)], (4A.20) − (n + 1)! n=0 X from which we obtain ˆ d ˆ ˙ Oˆ(1,t)= eC(t) e−C(t) = f(adC(t))[Cˆ(t)], (4A.21) dt − which is what we wanted to prove. Note that the final form of the series for Cˆ in (4A.6) can be rearranged in many different ways, using the Jacobi identity for the commutators. It is a nontrivial task to find a form involving the smallest number of terms.14 The derivation is an excerpt of the textbook cited in Ref. [1] on p. 80.

Appendix 4B Wigner Rotations and Thomas Precession

Here we calculate the full rate of Wigner rotations and the related Thomas precession.

Wigner Rotations

For brevity, we denote the small rotation (4.751) by

   ¯   

−id«· dt/2 (t+dt)· /2 −d · /2 − · /2 −1

¯  R(t) e = e e e = B ( (t + dt))B(d )B( ). (4B.1) ≡ 14For a discussion see J.A. Oteo, J. Math. Phys. 32 , 419 (1991). 374 4 Free Relativistic Particles and Fields

The pure rotation character of the product on the right-hand side is obvious, since a particle in its rest frame is transformed by three boost transformations back to the rest frame. Being a small

rotation, the left-hand side has necessarily the form  R =1 iª dt/2, (4B.2) − W ·

where ªW is an angular velocity describing the Wigner precession rate with a Heisenberg equation

−1 −1

 ª  ˙ = U(R) U(R) = . (4B.3) W ×

The parameter d¯ of the infinitesimal Lorentz transformation in (4B.1) must be chosen such that  the final laboratory rapidity is  + d . As with every Lorentz transformation, the product

−1

¯  B ()B(d )B( ) (4B.4) can be decomposed into a product of a Lorentz transformation and rotation:

−1 ′ ′

¯  « ¯ B ( )B(d )B( ) R(d )B(d ), (4B.5) ≈

′ ′

¯  where both parameters d« and d are small. Then we can expand (4B.1) up to first order in d as follows:

−1 ′ ′

 « ¯ R = B ( (t + dt))B( )R(d )B(d )

˙ ′ ′  « ¯ = 1+ B( )B( )dt + [R(d ) 1] + [B(d ) 1]. (4B.6) − − − It is straightforward to calculate the second term in the notation (4.461), in natural units with c = 1, where

1  B( )= (γ +1 γv ) (4B.7) 2(γ + 1) − · and p

˙ γ˙ 1    B( )= B( )+ [γ ˙ (1 + v )+ γv˙ ] . (4B.8) − −2(γ + 1) − 2(γ + 1) · · ˙ After multiplying this with B() we obtain p

˙ γ˙ 1     B( )B( )= + [γ ˙ (1 + v )+ γv˙ ] (γ +1 γv ). (4B.9) − −2(γ + 1) 2(γ + 1) · · − ·

The sum of all terms without any factor  cancel each other since they are equal to γ˙ γ˙ 1 + γγ˙ v2 + γ2v v˙ , (4B.10) −2(γ + 1) 2 − 2(γ + 1) ·   and this can be shown to vanish, being equal to

γ˙ γ˙ γγ˙ 1 + γγ˙ v2 + =0, (4B.11) −2(γ + 1) 2 − 2(γ + 1) γ2   after using the trivial identities

γ2 1 γ˙ = v v˙ γ3, v2 = − . (4B.12) · γ2

With the help of Eq. (4.753), the remaining terms can be decomposed as follows:

˙ ˙  «  ¬  B( )B( )= i ˙ /2 /2, (4B.13) − − · − · Appendix 4B Wigner Rotations and Thomas Precession 375

where the angular velocity «˙ is γ2

«˙ = v v˙ , (4B.14) −(γ + 1) × ˙ whereas the parameter ¬ is found, via Eq. (4B.12), to be

˙ 1 2 ¬ = [ γ˙ v γ(γ + 1)v˙ ]= γ v˙ γv˙ . (4B.15) (γ + 1) − − − k − ⊥

The vectors v˙ k and v˙ ⊥ denote the projections of v˙ parallel and orthogonal to v, respectively. The second (Hermitian) term in (4B.13) corresponds in (4B.6) to a pure infinitesimal Lorentz transformation. Since the final result (4B.6) must be a pure rotation, this term must be canceled by the last term in (4B.6), which is also Hermitian. Thus we conclude that

′ 1 2  B(d¯ )=1 γ v˙ + γ v˙ dt. (4B.16) − 2 k ⊥ · The remaining antihermitian term is of the type (4B.2). It gives a first contribution to the angular velocity of Wigner rotations (after reinserting here the omitted light velocity c): 1 γ2

ª = (v v˙ ). (4B.17) W1 −c2 γ +1 × ′ In order to find the second contribution to the angular velocity we must calculate the term R(d« )

¯  1 in (4B.6). Thus we transform B(d¯)=1 d /2 according to (4B.4) and (4B.5) and obtain:

− ·

 ¯   

′ ′ · /2 −d · /2 − · /2 −1

¯  ¯ 

R(d« )B(d )= e e e = B ( )B(d )B( ) 

1 d¯ 1  =1 (γ +1+ γv ) · (γ +1 γv ). (4B.18) − 2(γ + 1) · 2 2(γ + 1) − ·

We now use the twop rules p

        [a , b ]=2i(a b) , a b c =a b c a cb + b c a + i(a b) c, (4B.19) · · × · · · · · · − · · · · × · to calculate

′ ′ 1 2 2 2 2

¯ ¯  ¯  ¯  R(d« )B(d ) 1= (γ + 1) + v γ d 2γ v d v +2iγ(v d ) . − −4(γ+1) · − · · × ·   (4B.20) 2 Expressing v via (4B.12), the bracket simplifies to 2γ(γ + 1). After separating d¯ into parallel and orthogonal projections with respect to v, we obtain

′ ′ 1 iγ

¯ ¯  ¯  ¯  R(d« )B(d ) 1= γ d + d + (v d ) . (4B.21) − −2 k · ⊥ · γ +1 × ·  

By comparison with (4B.16), we identify from the Hermitian terms

¯ ¯ d¯k = γv˙ kdt, d ⊥ = γv˙ ⊥dt, , d = γv˙ dt, (4B.22) ′ and find that the extra rotation R(d« ) in the decomposition (4B.6) is

ª  R(d« ) 1 i /2, (4B.23) ≈ − W2 × with an angular velocity ªW2 which is twice the negative of (4B.17): 2 γ2

ª = (v v˙ ). (4B.24) W2 c2 γ +1 × The total angular velocity of Wigner rotation is therefore

1 γ2

ª ª ª = + = (v v˙ ), (4B.25) W W1 W2 c2 γ +1 × which generalizes the small-velocity result in Eq. (4.758) to (4.766). 376 4 Free Relativistic Particles and Fields

Thomas Precession It is now easy to modify the calculation to obtain the corresponding generalization of the Thomas

frequency (4.764). We simply replace the small Lorentz transformation B(d¯) in (4B.1) by the small Lorentz transformation (4.759) in the rest frame of the moving electron. As a consequence, ˜ the transformation (4B.5) is simply B(d¯) and the only difference with respect to the previous ′

calculation is that the rotation R(d« ) is absent. For this reason the over-compensating rotation ª by ªW2 is absent and we find again that the rate of Thomas precession T is equal to the first contribution ªW1 in Eq. (4B.17) to the Wigner precession, and thus precisely equal to the opposite of the total rate of Wigner rotation:

1 γ2 ª ª = = (v v˙ ). (4B.26) T − W −c2 γ +1 ×

Appendix 4C Calculation in Four-Dimensional Representation

The above calculations can certainly also be performed in the 4 4 -representation of the Lorentz group. As an illustration, let us rederive the 4 4 -version of Eq.× (4B.13). We denote the 4 4 × ×

-representation B˙ ˆ( ζ)Bˆ(ζ) by Λ(˙ v)Λ(v). Differentiating (4.18) with respect to time we see   − − that vi v˙i γ˙ γ˙ γ − c − c Λ(˙ v)=   . (4C.1) − vi v˙ i γ(γ + 2) vivj γ2 v˙ ivj + viv˙ j γ˙ γ γ˙ +  2 2 2   − c − c (γ + 1) c γ +1 c    Multiplying this with Λ(v) from the right yields a first row [Λ(˙ v)Λ(v)]0 in the product: − i v2 vv˙ γ2 v2 vi v˙ i γ3 vv˙ vi γγ˙ 1 γ2 , γ˙ γ 1 γ . (4C.2) − c2 − c2 − − γ +1 c2 c − c − γ +1 c2 c       Using again the relations vv˙ /c2 =γ/γ ˙ 3, and v2/c2 = (γ2 1)/γ2 [compare (4B.12)], the first entry and theγ ˙ -terms in the second entry disappear, and we remain− with

v˙ i γ3 vv˙ vi [Λ(˙ v)Λ(v)]0 = 0, γ . (4C.3) − µ − c − γ +1 c2 c  

Introducing the components v˙ k and v˙ ⊥ of the acceleration parallel and orthogonal to v, 2 2 2 such that (vv˙ )v = v v˙ k = (γ 1)v˙ k/γ , this can be expressed in terms of the 4 4 - − ×0 matrices (4.60)–(4.62) generating pure Lorentz transformations. Their first rows are (M1) i = 0 0 i(0, 1, 0, 0), (M2) i = i(0, 0, 1, 0), (M3) i = i(0, 0, 0, 1), such that we can write with the vector notation M (M ,M ,M ): ≡ 1 2 3 i [Λ(˙ v)Λ(v)]0 = γ2v˙ + γv˙ M. (4C.4) − i c k ⊥ · 

If we replace the 4 4 -generators M in this equation by the 2 2 -generators i/2 of pure Lorentz transformations,× we obtain × −

1 2

γ v˙ + γv˙ , (4C.5) 2c k ⊥ ·  which agrees with the previous result (4B.16), apart from the factor 1/c omitted there. A third way of deriving this result makes use of the spin four-vector introduced in Eq. (4.767). Here the precession rate is calculated by comparing the spin at time t, where the velocity is v(t), with the spin at t+dt, where the velocity is v(t+dt). During this time interval, the spin four-vector Appendix 4D Hyperbolic Geometry 377 has changed from Sµ to S¯µ Sµ + S˙ µdt. The initial spin is obtained by bringing the electron to ≡ µ 2 its rest frame via a deboosting Lorentz transformation Λ ν ( v). Using Eq. (4.18) and v from (4B.12), we have −

vi γ2 vivj S i =Λi ( v)Sµ = γ S0 + Si + Sj . (4C.6) R µ − − c γ +1 c2 The final spin is obtained by a similar Lorentz transformation with a slightly different velocity v(t + dt). The result is

Si + dSi = Λi ( v(t + dt))(Sµ + S˙ µdt) R R µ − = Λi ( v)Sµ + [Λ˙ i ( v)Sµ +Λi ( v)S˙ µ]dt. (4C.7) µ − µ − µ − We now use the fact that for an acceleration by a pure boost, which does not change the total angular momentum, the change of the spin four-vector S˙ µ is parallel to the direction of uµ = i i µ (γ,γv ). This will be shown in Eq. (6.59). Moreover, we can easily verify that Λ µ( v)u =0 by substituting uµ for Sµ in (4C.6). Hence we obtain −

dSi S˙ i = R = Λ˙ i( v)Sµ R dt µ − vi v˙i γ(γ + 2) vivj γ2 v˙ivj + viv˙ j = γ˙ + γ S0 +γ ˙ Sj + Sj. (4C.8) − c c (γ + 1)2 c2 γ +1 c2   Expressing Si with the help of (4.779), and using vv˙ /c2 =γ/γ ˙ 3, the last term becomes

γ2 viv˙ j γ2 viv˙j γ2 vv˙ vivj γ2 viv˙ j γ vivj Sj = Sj + Sj = Sj +γ ˙ Sj . γ +1 c2 γ+1 c2 R γ+1 c2 c2 R γ+1 c2 R (γ+1)2 c2 R   i i i i 0 i i i i We now use (4.779) to substitute v S /c = γv SR/c and S = v S /c = γ v SR/c. Then all terms containingγ ˙ cancel each other and we arrive at the formula for the temporal change of the spin vector:

2 2 ˙ i 1 γ i j i j j 1 γ i S = (v v˙ v˙ v )S = [(v v˙ ) S ] ª S , (4C.9) R c2 γ +1 − R −c2 γ +1 × × R ≡ T × R with the vector of angular velocity (4B.17). As a rate of change of a three-vector, it corresponds to a pure rotation. With the help of the generators (Li)jk = iǫijk of the rotation group, we may also write − ˙ S = i(ª L)S . (4C.10) R − T · R Appendix 4D Hyperbolic Geometry

Such kinematic calculations can, incidentally, be done quite elegantly in a geometric approach, called here Geometric Calculus. One may exploit the fact that the four-velocities uµ pµ/M can be written as ≡ µ ˆ u = c(cosh ζ,  sinh ζ). (4D.1) This shows that up to a factor c they are vectors on a unit hyperbola. These are hyperbolic analogs of euclidean vectors on a unit sphere

µ uE = c(cos α, «ˆ sin α). (4D.2) As such, relative rapidities follow the hyperbolic version of spherical trigonometry, called

− /2 Lobachevski geometry [24]. The product of three pure Lorentz transformations B() = e

378 4 Free Relativistic Particles and Fields

  with rapidities a, b, c can be represented as a triangle in hyperbolic space. The angles of the tri-

ˆ ˆ

  angle γa,γb,γc indicate the relative angles between the corresponding -vectors, i.e., cos γa = b c, etc. (see Fig. 4.4). The angles and sides of the triangle are then related by the cosine and sine·

theorems  cosh ζ = cosh ζ cosh ζ sinh  sinh cos γ , (4D.3) a b c − a b c

cos γ = cos γ cos γ + sin γ sin γ cosh  , (4D.4) a − b c a b c and sinh ζ sinh ζ sinh ζ a = b = c , (4D.5) sin γa sin γb sin γc respectively.

Figure 4.4 Triangle formed by rapidities in a hyperbolic space. The sum of angles is smaller than 1800. The angular defect yields the angle of the Thomas precession.

Given two sides plus one of the three angles, say ζa, ζb,γc, we can use the Napier analogies,

1 1 sinh 2 (ζa ζb) tan 2 (γa γb) 1 − = 1 − , sin 2 (ζa + ζb) cot 2 γc 1 1 cosh 2 (ζa ζb) tan 2 (γa + γb) 1 − = 1 (4D.6) cosh 2 (ζa + ζb) cot 2 γc to calculate the other two angles γa,γb. After that, either one of the two analogous formulas

1 1 sin 2 (γa γb) tanh 2 (ζa ζb) 1 − = 1 − , sin 2 (γb + γb) tanh 2 ζc 1 1 cos 2 (γa γb) tanh 2 (ζa + ζb) 1 − = 1 (4D.7) cos 2 (γa + γb) tanh 2 ζc serves to calculate the third side ζc. Since the hyperboloid has a negative unit radius, the sum of the angles is less than π. The angular defect, also called excess,

E = π γ γ γ , (4D.8) − a − b − c determines the area A of the triangle. For a hyperbola of radius R, the area is

A = R2E. (4D.9) Appendix 4E Clebsch-Gordan Coefficients 379

In spherical geometry this formula is known as Girard’s theorem. The angular defect is given in terms of the three sides by the hyperbolic version of the L’Huillier’s formula in spherical trigono- metry [22] E s s ζ s ζ s ζ tan = tanh tanh − a tanh − b tanh − c , (4D.10) 4 r 2 2 2 2 where s = (ζa + ζb + ζc)/2. (4D.11) For R it reduces to Heron’s formula [23] → ∞ A = s(s a)(s b)(s c), s = (a + b + c)/2 = semiparameter. (4D.12) − − − Another formulap is E 1 + cosh ζa + cosh ζb + cosh ζc cos = 2 2 2 . (4D.13) 2 4 cosh (ζa/2) cosh (ζb/2) cosh (ζc/2) A pure Lorentz transformation of a particle is a parallel transport along one side of a triangle. When doing three successive parallel transports around a triangle, a particle which was initially at rest comes again to rest. Its spin, however, changes the direction by the angular defect which is determined by the area integral. Since the radius is here equal to 1, formula (4D.13) determines directly the total angle of the Thomas precession. − The reader is encouraged to derive the rate of the Thomas precession once more using the Geometric Calculus [24].

Appendix 4E Clebsch-Gordan Coefficients

A direct product of irreducible representation states s1,m1 and s2,m2 can be decomposed into a sum of irreducible representation states sm with total| angulari momentum| i s = s s , , (s + | i | 1 − 2| ··· 1 s2). This is done with the help of Clebsch-Gordan coefficients. For this we multiply any product state with the completeness relation of all irreducible representation states

s s,m s,m =1, (4E.1) | ih | s m=−s X X and obtain

s1+s2 s s ,m ; s ,m = s,m s,m s ,m ; s ,m . (4E.2) | 1 1 2 2i | ih | 1 1 2 2i s,m=−s s=|Xs1−s2| X The expansion coefficients on the right-hand side are the desired Clebsch-Gordan coefficients. The expansion (4E.2) can be inverted by means of a similar completeness relation in the product space: s1 s2 s ,m ; s ,m s ,m ; s ,m =1, (4E.3) | 1 1 2 2ih 1 1 2 2| m1=−s1 m2=−s2 X X yielding the expansion

s1 s2 s,m = s ,m ; s ,m s ,m ; s ,m s,m . (4E.4) | i | 1 1 2 2ih 1 1 2 2| i m1=−s1 m2=−s2 X X By subjecting these relations to an arbitrary rotation (4.863), and using (4.865), we find the transformation behavior of the Clebsch-Gordan coefficients:

s ′ ′ ′ s −1 s −1 Dm,m′ s,m s1,m1; s2,m2 (D ) ′ (D ) ′ = s,m s1,m1; s2,m2 , (4E.5) h | i m1,m1 m2,m2 h | i 380 4 Free Relativistic Particles and Fields or, because of unitarity of the representation matrices,

s s ∗ s ∗ ′ ′ ′ D ′ (D ′ ) (D ′ ) s,m s1,m ; s2,m = s,m s1,m1; s2,m2 . (4E.6) m,m m1,m1 m2,m2 h | 1 2i h | i Since the Clebsch-Gordan coefficients are real following the Condon-Shortley convention we also have s ∗ s s ′ ′ ′ (D ′ ) D ′ D ′ s,m s1,m ; s2,m = s,m s1,m1; s2,m2 . (4E.7) m,m m1,m1 m2,m2 h | 1 2i h | i The Clebsch-Gordan coefficients are related in a simple way to the more symmetric Wigner 3j-symbols defined as follows:

s1−s2−m3 1/2 s1 s2 s3 s3, m3 s1,m1; s2,m2 = ( 1) (2s3 + 1) . (4E.8) h − | i − m1 m2 m3   As a consequence of relation (4.901), this has the invariance property

s s s s1 s2 s3 s1 s2 s3 D ′ D ′ D ′ = . (4E.9) m1,m1 m2,m2 m3,m3 m′ m′ m′ m m m  1 2 3   1 2 3  The Levi-Civita symbol ǫijk is a cartesian version of the Wigner 3j-symbol for s1 = s2 = s3 = 1. It exhibits the invariance (4E.9) with respect to the 3 3 defining representation matrices of the rotation group: × ′ ′ ′ ′ ′ ′ Ri1i1 Ri2i2 Ri3i3 ǫi1i2i3 = ǫi1i2i3 . (4E.10) Under even permutations of columns, the 3j-symbols are invariant, whereas under odd permuta- tions, they pick up a phase factor ( 1)s1+s2+s3 . Note also the property − s s s s s s 1 2 3 = ( 1)s1+s2+s3 1 2 3 . (4E.11) m1 m2 m3 − m1 m2 m3    − − −  In Eq. (4.906) we introduced a contravariant notation for spin-j objects. This is also done in relation (4E.8), writing it as

s1−s2−s3 1/2 s1 s2 m3 s3,m3 s1,m1; s2,m2 ( 1) (2s3 + 1) . (4E.12) h | i≡ − m1 m2 s3   The simplest 3j-symbol is

j j 0 j−m −1/2 = ( 1) (2j + 1) δ ′ , (4E.13) m m′ 0 − m,−m   2j (j) this being also equal to ( 1) c ′ . In the contravariant notation, one has − m,m j m′ 0 ′ = ( 1)2j (2j + 1)−1/2δ m . (4E.14) m j 0 − m   In order to calculate the Clebsch-Gordan coefficients we observe that the state of maximal quantum numbers s ,s ; s ,s is a state s,m of the irreducible representation with the maxi- | 1 1 2 2i | i mal angular momentum s = m = s1 + s2. By repeatedly applying the lowering operator of angular momentum to it, following the general relation (4.849), we obtain the matrix elements (compare Fig. 4.3) L s,s = (2s) 1 s,s 1 , −| i · | − i L− s,s 1 = p(2s 1) 2 s,s 1 , | − i − · | − i . .p (4E.15) L s, s +2 = (2s 1) 2 s, s +1 , −| − i − · | − i L− s, s +1 = p(2s) 1 s, s . | − i · | − i p Appendix 4E Clebsch-Gordan Coefficients 381

Table 4.2 Lowest Clebsch-Gordan coefficients s,m s ,m ; s ,m . The table entries h | 1 1 2 2i CG are to be read as √CG. The coefficients are all real. For more symmetry properties ± ± see Eqs. (4E.22). Table is taken from the Particle Properties Data Booklet in Ref. [25].

s ,m ; s ,m s,m h 1 1 2 2| i = ( 1)s−s1−s2 s , m ; s , m s, m − h 2 − 2 1 − 1| − i = ( 1)s−s1−s2 s ,m ; s m s,m − h 2 2 1 1| i 382 4 Free Relativistic Particles and Fields

In the direct-product space, an application of the lowering operator L− 1+1 L− to the state s ,s ; s ,s yields, with the same rules as in (4E.15), × × | 1 1 2 2i (L 1+1 L ) s ,s ; s ,s = (2s ) 1 s ,s 1; s ,s − × × − | 1 1 2 2i 1 · | 1 1 − 2 2i + p(2s2) 1 s1,s1; s2,s2 1 . (4E.16) · | − i Continuing this with the help of the general relationp

(L 1+1 L ) s ,m ; s ,m = (s + m )(s m + 1) s ,m 1; s ,m − × × − | 1 1 2 2i 1 1 1 − 1 | 1 1 − 2 2i + p(s2 + m2)(s2 m2 + 1) s1,m1; s2,m2 1 , (4E.17) − | − i we find all other states s,m of the irreduciblep representation with s = s + s . | i 1 2 The state of the lower total angular momentum s1 + s2 1 with a maximal magnetic quantum number m = s is obtained from the orthogonal combination− of (4E.17):

s +s 1,s +s 1 = (2s ) 1 s ,s 1; s ,s (2s ) 1 s s ; s s 1 . (4E.18) | 1 2− 1 2 − i 1 · | 1 1 − 2 2i− 2 · | 1 1 2 2 − i p p This can be verified by applying to it the raising operator (L+ 1+1 L+), generalizing (4.849) to the direct-product space: × ×

(L 1+1 L ) s ,m ; s ,m = (s m )(s + m + 1) s ,m + 1; s ,m + × × + | 1 1 2 2i 1 − 1 1 1 | 1 1 2 2i + p(s2 m2)(s2 + m2 + 1) s1,m1; s2,m2 +1 , (4E.19) − | i and finding that it is annihilated if m1 or mp2 reach their highest possible values s1 or s2, respec- tively. By applying the lowering operator to the state (4E.18), we generate all states of the irreducible representation s1 + s2 1,m with m = s1 s2 +1,...,s1 + s2 1. Multiplying| (4E.17)− fromi the left by s,m− −and using the Hermitian− adjoint of relation (4.849), we obtain the recursion relation h |

(s + m)(s m + 1) s ,m ; s ,m s,m − h 1 1 2 2| i p = (s1 m1 + 1)(s1 + m1) s1,m1 1; s2,m2 s,m 1 − h − | − i + p(s2 m2 + 1)(s2 + m2) s1,m1; s2,m2 1 s,m 1 . (4E.20) − h − | − i Similarly we can take the raisingp operator relation (4E.19) in the direct-product space, go over to the adjoint, and multiply it by s,m from the left to find h | (s m)(s + m + 1) s ,m ; s ,m s,m − h 1 1 2 2| i p = (s1 + m1 + 1)(s1 m1) s1,m1 + 1; s2,m2 s,m +1 − h | i + p(s2 + m2 + 1)(s2 m2) s1,m1; s2,m2 +1 s,m +1 . (4E.21) − h | i The Clebsch-Gordan coefficientsp have the following important symmetry properties:

s,m s ,m ; s ,m = ( 1)j−s1−s2 s,m s ,m ; s ,m h | 1 1 2 2i − h | 2 2 1 1i = s, m s , m ; s , m h − | 2 − 2 1 − 1i = ( 1)j−s1−s2 s, m s , m ; s , m − h − | 1 − 1 2 − 2i 2s +1 = ( 1)s1−m1 s , m s ,m ; s, m − 2s +1h 2 − 2| 1 1 − i r 2 2s +1 = ( 1)s2+m2 s , m s, m; ,s , m . (4E.22) − 2s +1h 1 − 1| − 2 − 3i r 1 Some frequently-needed values are listed in Table 4.2. Appendix 4F Spherical Harmonics 383

Appendix 4F Spherical Harmonics

The spherical harmonics are defined as

2l +1 (l m)! 1/2 Y (θ, ϕ) ( 1)m − P m(cos θ)eimϕ, (4F.1) lm ≡ − 4π (l + m)! l   m where Pl (z) are the associated Legendre polynomials 1 dl+m P m(z)= (1 z2)m/2 (z2 1)l. (4F.2) l 2ll! − dxl+m − The spherical harmonics are orthonormal with respect to the rotation-invariant scalar product

π 2π ∗ ′ ′ ′ ′ dθ sin θ dϕ Ylm(θ, ϕ)Yl m (θ, ϕ)= δll δmm . (4F.3) Z0 Z0 Explicitly, they read for the lowest few angular momenta: 1 Y (θ, φ) = , 00 2 √π

3 1 iφ Y11(θ, φ) = sin θ e , − r4 π √2 3 Y10(θ, φ) = cos θ, r4 π

1 15 2 2 iφ Y22(θ, φ) = sin θe , 4r2 π 15 iφ Y21(θ, φ) = sin θ cos θe , −r8 π 5 3 1 Y (θ, φ) = cos2 θ , 20 4 π 2 − 2 r  

1 35 3 3 iφ Y33(θ, φ) = sin θe , − 4r4 π 1 105 2 2 iφ Y32(θ, φ) = sin θ cos θe , 4 r 2 π 1 21 2 iφ Y31(θ, φ) = sin θ 5cos θ 1 e , −4 r4 π − 7 5 3  Y (θ, φ) = cos3 θ cos θ . (4F.4) 30 4 π 2 − 2 r   The spherical harmonics with a negative magnetic quantum number m are obtained from the relation Y (θ, φ) = ( 1)mY ∗ (θ, φ). (4F.5) lm − l,−m − For m = 0, the spherical harmonics reduce to

2l +1 Ylm(θ, φ)= Pl(cos θ), (4F.6) r 4π where 1 dl P (z) P 0(z)= (z2 1)l (4F.7) l ≡ l 2ll! dzl − 384 4 Free Relativistic Particles and Fields are the Legendre polynomials. j l For integer j = l, the rotation functions dm,m′ (β) can be derived recursively from dm,0(β) = Pl(cos θ) with the help of the recursion relation

′ ′ j j−1 2 (j + m )(j + m 1)d ′ (β)= (j + m)(j + m 1)(1 + cos θ) d ′ (β) − m,m − m−1,m −1 2 2 j−1 j−1 +2p j m sin θ d ′ (β)+ (jp m)(j m 1)(1 cos θ) d ′ (β). − m,m −1 − − − − m+1,m −1 p p (4F.8) j For an iterative determination of the rotation functions dm,m′ (β) with half-integer j we use the recursion relation

′ ′ j j m j−1/2 β j + m j−1/2 β d ′ (β)= − d ′ cos + d ′ sin . (4F.9) m ,m j m m +1/2,m+1/2 2 j m m −1/2,m+1/2 2 s − s − Inserting m = 1/2 and using (4.876), we deduce −

j j + m j−1/2 β j m j−1/2 β dm 1/2(β)= dm−1/2,0 cos − dm+1/2,0 sin . (4F.10) sj +1/2 2 − sj +1/2 2

j For j = l +1/2, the right-hand side contains only Legendre polynomials. Starting from dm,1/2(β), we find all other rotation functions with half-integer j from the recursion relation (4F.8). For j =1/2 and j = 1, the explicit results were given in Eqs. (4.881) and (4.882). For j =3/2 we obtain 1 β d3/2 (β) = (1 + cos β)cos , 3/2,3/2 2 2 √3 β d3/2 (β) = (1 + cos β) sin , 3/2,1/2 − 2 2 √3 β d3/2 (β) = (1 cos β)cos , 3/2,−1/2 2 − 2 1 β d3/2 (β) = (1 cos β) sin , 3/2,−3/2 −2 − 2 1 β d3/2 (β) = (3 cos β 1)cos , 1/2,1/2 2 − 2 1 β d3/2 (β) = (3 cos β + 1) sin . (4F.11) 1/2,−1/2 2 2 The remaining matrix elements are obtained via the relation (4.876). Similarly we have for j =2 the matrix elements: 1 d2 (β) = (1 + cos β)2, 2,2 4 1 d2 (β) = (1 + cos β) sin β, 2,1 −2 √6 d2 (β) = sin2 β, 2,0 4 1 d2 (β) = (1 cos β) sin β 2,−1 − 2 − 1 d2 (β) = (1 cos β)2, 2,−2 4 − 1 d2 (β) = (1 + cos β)(2 cos β 1), 1,1 2 −

2 3 d1,0(β) = sin β cos β, −r2 Appendix 4G Projection Matrices for Symmetric Tensor Fields 385

1 d2 (β) = (1 cos β)(2 cos β + 1), 1,−1 2 − 1 d2 (β) = (3 cos2 β 1). (4F.12) 0,0 2 − Appendix 4G Projection Matrices for Symmetric Tensor Fields in D Dimension

These projection matrices can all be constructed from appropriate combinations of the longitudinal and transversal projection matrices whose D = 3 -dimensional versions where stated in Eqs. (4.336) and (4.334):

P l (k)= kˆ kˆ and P t (k)= η P l (k). (4G.1) µν µ ν µν µν − µν Recall that due to the identity

l t Pµν (k)+ Pµν (k)= ηµν , (4G.2) these permit decomposing any vector field into its longitudinal and transversal parts:

A = Al + At P l Aµ + P t Aµ. (4G.3) µ µ µ ≡ µν µν µ µ 2 Here ηµν is the D-dimensional generalization of the Minkowski metric (4.27), and kˆ k /√k . The spin-2 projection matrix reads ≡ 1 1 P (2) (k)= P t (k)P t (k)+ P t (k)P t (k) P t (k)P t (k), (4G.4) µν,λκ 2 µλ νκ µκ νλ − D 1 µν λκ   − which for D = 3 reduces to (4.428). The spin-1 projection matrix reads 1 P (1) (k) = P t (k)P l (k)+ P t (k)P l (k)+ P t (k)P l (k)+ P t (k)P l (k) µν,λκ 2 µλ νκ µκ νλ νκ µλ νλ µκ = 1 (kˆ kˆ η + kˆ kˆ η + kˆ kˆ η + kˆ kˆ η ) 2kˆ kˆ kˆ kˆ .  (4G.5) 2 µ λ νκ µ κ νλ ν λ νκ µ κ νλ − µ ν λ κ There are four further projections 1 P (0s) (k) = P t (k)P t (k), µν,λκ D 1 µν λκ 1− P (0w) (k) = P l (k)P l (k), µν,λκ D 1 µν λκ (0sw) t − l Pµν,λκ(k) = Pµν (k)Pλκ(k), (0ws) l t Pµν,λκ(k) = Pµν (k)Pλκ(k). (4G.6) The first two are commonly collected into a single spin-0 projection: 1 P (0) (k) P (0w) (k)+ P (0s) (k)= (η η kˆ kˆ η η kˆ kˆ + kˆ kˆ kˆ kˆ ). (4G.7) µν,λκ ≡ µν,λκ µν,λκ D 1 µν λκ − µ ν λκ − µν λ κ µ ν λ κ − The projections satisfy the completeness relation 1 P (2) (k)+ P (1) (k)+ P (0) (k)= [η η + η η ]= 1 . (4G.8) µν,λκ µν,λκ µν,λκ 2 µλ νκ µκ νλ µν,λκ By analogy with (31.15), the gravitational field can be decomposed into the three associated components

(2) (1) (0) hµν (k)= hµν (k)+ hµν (k)+ hµν (k), (4G.9) 386 4 Free Relativistic Particles and Fields where

h(2) P (2) hλκ P t P t 1 P t P t hλκ, (4G.10) µν ≡ µν,λκ ≡ µλ νκ − D−1 µν λκ  h(1) P (1) hλκ = h P t P t hλκ 1 P l P l hλκ, (4G.11) µν ≡ µν,λκ µν − µν λκ − D−1 µν λκ and

h(0) P (0) hλκ = 1 P t P t hλκ. (4G.12) µν ≡ µν,λκ D−1 µν λκ

Alternatively we may decompose the field hλκ into

(2) l s hµν (k)= hµν (k)+ hµν (k)+ hµν (k), (4G.13) where

h(2) P (2) hλκ P t P t hνκ 1 P t P t hλκ, (4G.14) µν ≡ µν,λκ ≡ µλ νκ − D−1 µν λκ  hl P L hλκ h P t P t hνκ (4G.15) µν ≡ µν,λκ ≡ µν − µλ νκ and

hs P s hλκ P 0s hλκ 1 P t P t hλκ (4G.16) µν ≡ µν,λκ ≡ µν,λκ ≡ D−1 µν λκ is the scalar part. The three projections together satisfy the completeness relation

(2) l s Pµν,λκ + Pµν,λκ + Pµν,λκ = 1µν,λκ. (4G.17)

Notes and References

For other introductions to the theory of fields can be found in the textbooks by S. Schweber, An Introduction to Relativistic , Harper and Row, New York, 1962; C. Itzykson and J.-B. Zuber, Quantum Field Theory, McGraw-Hill, New York, 1985; S. Weinberg, The Quantum Theory of Fields, Cambridge Univ. Press, London, 1996; M.E. Peskin and D.V. Schroeder, An Introduction to Quantum Field Theory, Sarat Book House, Kolkata, 2005. The particular citations in this chapter refer to:

[1] The Lorenz condition is named after the Dutch physicist Ludvig Lorenz, On the Identity of the Vibrations of Light with Electrical Currents, Philos. Mag. 34, 287 (1867). It is a Lorentz- invariant gauge condition and often falsely called ”Lorentz condition”, confusing the author with Hendrik Lorentz [H.A. Lorentz, Theory of Electrons, 2nd edition, (1915), Dover. N.Y. (1966)].

[2] B.W. Petley, Nature 303, 373 (1983). Note that the light velocity c has, by definition, the value stated in Eq. (4.2). This has been so since 1983, when the previous meter has been redefined in the Conf´erence G´en´erale des Poids et Mesures in Paris to make this value exact. Notes and References 387

[3] Experimentally, the best upper limit for the mass term Mγ in the electromagnetic field equations can be deduced under terrestrial conditions from the shape of the Earth’s magnetic −48 field. The limit is Mγ < 4 10 g. This corresponds to a Compton wavelengthλ ¯γ = 10 · ¯h/Mγ c> 10 cm, which is larger than the diameter of the sun. Astrophysical considerations 16 (“wisps” in the crab nebula) giveλ ¯γ > 10 cm. If metagalactic magnetic fields would be discovered, the Compton wavelength would be larger than 1024 1025cm, quite close to the ultimate limit set by the horizon of the universe = c age of the− universe 1028cm. See G.V. Chibisov, Sov. Phys. Usp. 19, 624 (1976). × ∼ [4] O. Klein, Z. Phys. 37, 895 (1926); W. Gordon, ibid. 40, 117 (1926). V. Fock, ibid. 38, 242; 39, 226 (1926); Note that the name Klein-Gordon equation does injustice to Fock but more so to E. Schr¨odinger. He actually invented the Klein-Gordon equation first, from which he derived his famous nonrelativistic wave equation in the limit of large c, although his papers in Ann. Phys. 79, 361, 489; 80, 437; 81, 109 (1926), suggest the opposite order. This was pointed out by P.A.M. Dirac, The Development of Quantum Theory, Gordon and Breach, N.Y., 1971. See also Dirac’s popular articles in Nature 189, 335 (1961) and in Scientific American 208, 45 (1963). [5] P.A.M. Dirac, Proc. Roy. Soc. A 117, 610 (1928), A 118, 351 (1928). [6] H. Weyl, J. Phys. 50, 330 (1929). [7] T.D. Lee and C.N. Yang, Phys. Rev. 104, 254 (1956); 105, 167 (1957). [8] C.S. Wu, E. Ambler, R.W. Hayward, D.D. Hoppes, and R.P. Hudson, Phys. Rev. 105, 1413 (1957). [9] F. Wilczek, Majorana and Condensed Matter Physics, (arXiv:1404.0637); R. Jackiw, Emergent Fractional Charge and Multiple Majoranas, (arXiv:1404.6200). [10] This definition is completely analogous to the definition in nonabelian gauge theories [see (28.11)]. The relation to the standard Riemann tensor used in some gravitational textbooks κ κ (such as Weinberg’s [15]) is Rµνλ = R λµν . This makes our Ricci tensor (4.365) equal to the negative of the Ricci tensor appearing in those books, and the negative sign carries over to the Einstein tensor (4.376), leading to the Einstein equation (5.71). [11] H.A. Schwarz, Gesammelte Mathematische Abhandlungen, vol. 2, Springer, Berlin, 1890. [12] T. Levi-Civita, Absolute Differential Calculus, Blackie & Sons, London 1929. First published in Rome, 1925 (in Italian). [13] This statement holds in the sense of distributions. Statements about distributions must al- ways be integrated with an arbitrary smooth test function as a factor to test their validity. The Riemann-Lebesgue Lemma states that an integral over a fast oscillating function mul- tiplied by a smooth function yields zero. See Chapter 1 of E.C. Titchmarsh, Introduction to the Theory of the Fourier Integral, Oxford University Press, Oxford, 1937. [14] The Rarita-Schwinger action and its generalizations have been investigated in great detail in attempts to understand the pion nucleon scattering amplitude near the first resonance at around 1240 GeV. Among the many references see: R.D. Peccei, Phys. Rev. 176, 1812 (1968); L.S. Brown, W.J. Pardee, and R.D. Peccei, Phys. Rev. D 4, 2801 (1971); V. Bernard and U.G. Meissner, Phys. Lett. B 309, 421 (1993), Phys. Rev. C 52 2185 (1995); V. Bernard et al., Int. J. Mod. Phys. E 4, 193 (1995). 388 4 Free Relativistic Particles and Fields

Details and many references can be found in the comprehensive review by G. H¨ohler, Elastic and Charge Exchange Scattering of Elementary Particles, in Landolt- B¨ornstein, Vol. I, Springer, Berlin 1983. [15] S. Weinberg, Gravitation and Cosmology, J. Wiley and Sons, New York, 1972. [16] I. Antoniadis and N.C. Tsamis, Weyl Invariance and the Cosmological Constant, SLAC- Pub-3297. 1984. [17] H. Feshbach and F. Villars, Rev. Mod. Phys. 30, 24 (1958). [18] A thorough discussion of the rotation group is found in the textbook A.R. Edmonds, Angular Momentum in Quantum Mechanics, Princeton Univ. Press, Prince- ton, 1962. A short concise table of the Clebsch-Gordan coefficients can be found in the Particle Properties Data Booklet [25]. [19] E.C.G. Sudarshan and R.E. Marshak, Phys. Rev. 109, 1860 (1958); R.P. Feynman and M. Gell-Mann, ibid. 193 (1958); J.J. Sakurai, Nuovo Cimento 7, 649 (1958); W.R. Theis, Z. Phys. 150, 590 (1958); Fortschr. Physik 7, 559 (1959). [20] S. Deser and C. Teitelboim, Phys. Rev. D 13, 1592 (1976); C. Bunster and M. Henneaux, Phys. Rev. D 83, 045031 (2011), (arXiv:1011.5889); S. Deser, Class. Quant. Grav. 28, 085009 (2011) (arXiv:1012.5109). [21] H. Kleinert, Multivalued Fields in Condensed Matter, Electromagnetism, and Gravitation, World Scientific, Singapore 2009, pp. 1–497 (http://klnrt.de/b11). [22] D.D. Ballow and F.H. Steen, Plane and Spherical Trigonometry with Tables, Ginn, New York, 1943. [23] D.W. Mitchell, Mathematical Gazette 93, 108 (2009). [24] J.A. Smorodinskij, Fortschr. Phys. 13, 157 (1965); N.I. Lobachevski, La Theorie d´es Parall`eles, Albert Blanchard, Paris 1957 (http://galli- ca.bnf.fr/ark:/12148/bpt6k3942g); See also A. Sommerfeld, Electrodynamics, Academic, New York, 1949. [25] Go to this internet address: http://pdg.lbl.gov/2013/reviews/rpp2013-rev-clebsch- gordan-coefs.pdf. All power corrupts! Chinese Wisdom 5 Classical Radiation

Changes in the local or temporal charge distribution lead to changes in the elec- tromagnetic field created by them. If the change is sufficiently rapid, the electro- magnetic field will start propagating with light-velocity through spacetime. The standard example is the radiation emitted by an antenna, or the ultraviolet light emitted by an electron synchrotron. These and related phenomena will now be discussed.

5.1 Classical Electromagnetic Waves

In the presence of a classical source jµ(x), the electromagnetic action (4.237) contains an extra term

j 1 4 µ = d x A (x) jµ(x), (5.1) A − c Z so that the Euler-Lagrange equation for the electromagnetic field becomes 1 ∂ F µν = ∂2Aµ + ∂µ∂ Aν = jµ. (5.2) ν − ν − c Note that this equation is consistent with current conservation. Indeed, both sides vanish if multiplied by the derivative ∂µ. The right-hand side vanishes due to the conservation law (4.172), the left-hand side due to the antisymmetry of F µν which guarantees the validity of the so-called Bianchi identity ∂ ∂ F µν 0. (5.3) µ ν ≡ In order to solve the field equation (5.2), we have to choose a specific gauge. The Lorenz gauge is most convenient:

µ ∂µA (x)=0. (5.4) Then the equation reduces to the Klein-Gordon form (4.170) for each vector com- ponent [compare (4.253) and the related discussion of gauge transformations]: 1 ∂2Aµ = jµ. (5.5) c

389 390 5 Classical Radiation

5.1.1 Electromagnetic Field of a Moving Charge A point charge moving along the trajectory x¯(s) gives rise to a current

x¯˙ µ(s) µ ˙ µ (4) (3) x x j (x)= ec ds x¯ (s) δ (x x¯(s)) = ec 0 δ ( ¯(s)), (5.6) Z − dx¯ /ds − where s is an invariant length parameter, such that

dx¯0 v¯2(t) = 1 = γ(t). (5.7) ds s − c2 Here v¯(t) is the velocity along the trajectory x¯(t):

v¯(t) x¯˙ (t). (5.8) ≡ The current components are

j0(x)= ecδ(3)(x x¯(t)), j(x)= ev¯(t)δ(3)(x x¯(t)). (5.9) − − To find the electromagnetic field emerging from this current we solve the field equa- tion (5.5) by µ i 4 µ A (x)= d yGR(x y) j (y), (5.10) c Z − where GR(x) is the retarded Green function 1 G (x)= iΘ(x0)δ(x2)= iΘ(x0) δ(x0 R)+ δ(x0 R) , (5.11) R − − 4πR − − h i which satisfies the differential equation

∂2G (x)= iδ(x). (5.12) − R In momentum space, Eq. (5.5) can be solved by 1 Aµ(k)= jµ(k). (5.13) −ck2 Inserting (5.10) into the electromagnetic action consisting of a free part (4.237) plus a source term (5.1), we obtain the total action

1 4 µν 1 4 µ 1 4 4 ∗ µ BS = d xFµν F d xA jµ = 2 d xd yjµ(x)GR(x y)j (y). (5.14) A −4 Z − c Z −2c Z − This was first found in the nonrelativistic setting by Biot and Savart (see the text- book [1]). A similar action was found for fluctuating vortex lines in superfluids [2]. In energy-momentum space, the last term can be written as

4 1 d k ∗ 1 µ BS = 2 4 jµ(k) 2 j (k). (5.15) A 2c Z (2π) k 5.1 Classical Electromagnetic Waves 391

This is composed of a charge-charge and a current-current term

4 1 d k ∗ 1 ∗ 1 = ρ (k) ρ(k) j (k) j(k) + jj. (5.16) ABS 2 (2π)4 k2 − k2 ≡Aρρ A Z h i The current-current term may be decomposed further into a longitudinal and a transversal part. By inserting the projection matrices (4.334) and (4.336) between the currents, we find

4 4 ˆ ˆ ˆ ˆ 1 d k ∗ 1 1 d k ∗ kikj ∗ δij kikj jj= j (k) j(k)= ji (k) jj(k)+ji (k) − jj(k) ,(5.17) A −2 (2π)4 k2 −2 (2π)4" k2 k2 Z Z i where kˆi ki/ k is the unit vector pointing into the direction of the momentum. ≡ | | µ Now we make use of current conservation in momentum space kµj (k) = 0 to equate

kˆ j (k)=(k / k )j (k), (5.18) i i 0 | | 0 and the interaction (5.17) becomes

4 2 ˆ ˆ 1 d k k0 ∗ 1 ∗ δij kikj jj = j0 (k) j0(k)+ji (k) − jj(k) . (5.19) A −2 (2π)4 "k2 k2 k2 Z i Adding this to , we obtain the total interaction Aρρ 4 1 d k ∗ 1 ∗ 1 = ρ (k) ρ(k) j (k) j (k) = + j j . (5.20) ABS 2 (2π)4 − k2 − T k2 T ACoul A T T Z h i The first term is the instantaneous Coulomb interaction between the charges, the second is due to the transverse electromagnetic radiation between the currents. Let us write the vector potential (5.13) of the emitted radiation as 1 1 Aµ(x)= d3x′ jµ(x′) . (5.21) 4πc x x′ t′=t−|x′−x|/c Z | − | The Lorenz gauge may be exhibited more explicitly by multiplying Eq. (5.13) with the transverse projection operator for this gauge [see Eq.(4G.1)]:

t kµkν Pµν (i∂)= gµν . (5.22) − k2 ! Then (5.13) becomes

t 1 kµkν ν 1 t ν Aµ(k)= gµν j (k)= Pµν(k)j (k), (5.23) −ck2 − k2 ! −ck2

µ which fulfills explicitly the Lorenz gauge kµA (k) = 0. In x-space it reads 1 1 At (x)= d3x′ P t (i∂)jν (x′) . (5.24) µ 4πc x x′ µν t′=t−|x′−x|/c Z | − |

392 5 Classical Radiation

Introducing the distance vector R(t) x x¯(t) from the charge, the spacetime components (A0(x), A(x)) of the four-vector≡ A−µ(x) are the famous Li´enard-Wiechert potentials: 1 A0(x) = ec dt′ Θ(t t′) δ(t t′ R(t′)), (5.25) Z − 4πR − − 1 A(x) = e dt′ Θ(t t′) v¯(t′)δ(t t′ R(t′)). (5.26) Z − 4πR − − We now simplify the δ-functions as follows:

′ ′ 1 ′ 1 ′ δ(t t R(t )) = δ(t tR)= δ(t tR), ′ ′ ′ ′ − − d[t + R(t )]/dt t =tR − 1 n(tR) v¯(tR) − | | − · (5.27) where t t R(t )/c (5.28) R ≡ − R is the earlier time at which the field has to be emitted in order to arrive at time t at the observation point x. The unit vector n(t) denotes the direction of the emission:

R(t) n(t) . (5.29) ≡ R(t) | | Inserting R(t) into (5.25) and (5.26), we find the vector potential

1 e 1 ev¯/c A0(x)= , A(x)= . (5.30) 4π "(1 n v¯/c)R # 4π " (1 n v¯/c)R # − · ret − · ret The brackets with the subscript “ret” indicate that the time argument t inside the brackets is equal to the earlier time tR of emission as determined by (5.28). By forming the gauge-invariant combinations of derivatives of the vector potential we find the field strengths 1 E(x) = A˙ (x) ∇A0(x), (5.31) − c − B(x) = ∇ A(x), (5.32) × which, for the Li´enard-Wiechert potentials (5.30), have the values

e (n v¯/c) 1 n [(n v¯/c) v¯˙ ] 1 E(x, t)= − + × − × , (5.33) 4π "(1 n v¯/c)2 R2 (1 n v¯/c)3 R # − · − · ret B(x, t) = [n E] . (5.34) × ret The two terms in E(x, t) have different falloff-behaviors as functions of the dis- tance R from the source. The first is a velocity field which falls off like 1/R2. It is essentially the moving static field around the particle. The second term is an acceleration field, which has a slower fall-off proportional to 1/R. For this reason it can carry off radiation energy to infinity. Indeed, the energy flux through a solid 5.1 Classical Electromagnetic Waves 393 angle dΩ is given by the scalar product of the Poynting vector1 E B with the area element dS = r2dΩ: × 1 ˙ = dS (E B)= r2dΩr2(E B) kˆ = kˆ (kˆ k¨) 2 = r2dΩ cE2. (5.35) E · × × · 8πc3 | × × | For a radiating electron at small velocities near the coordinate origin, the accelera- tion field simplifies to e e E(x, t)= [xˆ (xˆ x¨)], B(x, t)= (xˆ x¨), (5.36) 4πrc2 × × 4πrc2 × with r = x . The radiated power per solid angle is then | | d ˙ e2 e2 E = (xˆ x¨)2 = x¨2 sin2 β, (5.37) dΩ (4π)2c3 × (4π)2c3 where β is the angle between the oscillating dipole ends and the direction of emission. By integrating over all solid angles, we obtain the total radiated power e2 2 e2 ˙ x x 2 x2 = 2 3 dΩ(ˆ ¨) = 3 ¨ . (5.38) E (4π) c Z × 3c 4π This is the famous Larmor formula of classical electrodynamics. For a harmonically oscillating charge at position

x(t)= x e−iωt + x∗eiωt =2 x cos(ωt + δ), (5.39) 0 0 | 0| equation (5.37) yields the temporal average power

d ˙ e2 ω4 e2 ω4 E = xˆ x 2 = x 2 sin2 β, (5.40) dΩ 8π2 c4 | × 0| 8π2 c4 | 0| and the total radiated power is given by the antenna formula e2 4 ω4 ˙ = x 2. (5.41) E 4π 3 c4 | 0| Note that, in comparison with standard textbooks on classical electrodynamics such as the one in Ref. [1] where the electromagnetic Lagrangian carries a prefactor 1/4π, the square of the charge carries here an extra factor 1/4π. Thus the factor e2 in (5.41) is related to the fine-structure constant α by e2 =4παhc¯ . Inserting (5.28) into (5.21) we obtain, for the radiated vector potential, the formula ′ 1 eiω|x−x |/c dω Aµ(x)= d3x′ jµ(x′,ω)e−iωt, (5.42) 4πc x x′ 2π Z | − | Z 1Note that the coincides with the components T 0i = (E B)i of the energy- µν 00 × 1 2 2 momentum tensor T of the electromagnetic field, whose component T = 2 (E + B ) is the energy density. For more details see the textbook in Ref. [1]. 394 5 Classical Radiation

where ∞ jµ(x′,ω)= dt eiωtjµ(x′, t) (5.43) Z−∞ are the temporal Fourier components of the current density. For large r = x , we may approximate | | iω|x−x′|/c iωr/c e e ′ e−iωnx /c, (5.44) x x′ ≈ r | − | leaving an x′-dependence only in the sensitive phase factor. This allows an exact splitting of (5.33) into a velocity and an acceleration field to carry off energy to infinity. At a point x far away from the source, the spherically radiated field (5.42) looks like a passing plane wave with eiωr/c eikx. Thus (5.42) becomes ≈ 1 1 dω Aµ(x, t)= e−iωtR jµ(k,ω), (5.45) 4πc r Z 2π where the time accounting for the retardation is

t = t r/ct, (5.46) R − and jµ(k,ω) is the momentum-space version of (5.43):

∞ jµ(k,ω)= dt d3x eiωt−ikxjµ(x, t). (5.47) Z−∞ Z At a fixed k of the outgoing wave, we can thus write 1 1 Aµ(x, t)= jµ(k, t ). (5.48) 4πc r R We now calculate the energy flux from formula (5.35). On the right-hand side of this equation, we express the electric field E(x) in terms of the vector potential via (5.31) and find

d2 ˙ 1 2 E = r2 c A˙ + ∇A0 . (5.49) dΩ  c  In momentum space, the Lorenz gauge (5.4) implies that

A0(k, t)= kˆ A(k, t), (5.50) · so that we can rewrite 1 1 1 A˙ (k, t)+ ikA0(k, t)= ∂ A(k, t) kˆA0(k, t) = ∂ A (k, t), (5.51) c c t − c t T h i where

A (k, t) P (kˆ)A(k, t) (5.52) T ≡ T 5.1 Classical Electromagnetic Waves 395 is the transverse part of the vector field A(k, t) defined by the projection matrix (4.334): P ij(kˆ) δij kˆikˆj. (5.53) T ≡ − We now express AT in terms of jT PT j using Eqs. (5.48) and (5.49), to find the radiated energy per unit time and solid≡ angle:

d ˙ 1 E = [∂ j (k, t )]∗ [∂ j (k, t )]. (5.54) dΩ 16π2c3 t T R · t T R

If the emission is from a periodically oscillating source with frequency ω = ck0, so that j(k, t)= j(k)e−iωt, we can replace this by

d ˙ ω2 E = j (k) 2. (5.55) dΩ 16π2c3 | T | Note that this radiated power was encountered before in the transverse part of the Biot-Savart interaction energy (5.20). An equivalent expression can be obtained by using current conservation in mo- mentum space (5.18) to write

k2 k2 j (k) 2 =jl(k)∗(δ kˆ kˆ )jm(k)= j(k) 2 0 j0(k) 2 = j (k)∗jµ(k) j0(k) 2. | T | lm − l m | | − k2 | | − µ − k2 | | (5.56) The momenta of the outgoing waves are real on-shell photons with k2 = 0, implying that (5.55) can be rewritten in the completely covariant form

d ˙ ω2 E = j (k)∗jµ(k). (5.57) dΩ −16π2c3 µ

5.1.2 Dipole Moment For long wavelengths, the spatial components of current density have a negligible dependence on k: ji(k, t) ji(k = 0, t) = d3x ji(x, t). With the help of an inte- gration by parts, the right-hand≈ side can further be rewritten as d3x xi∂ jk(x, t), R − k so that we have the approximate relation R

j(k, t) d3x x ∇j(x, t). (5.58) ≈ − Z We can now use the current conservation law (4.172) and find

3 0 3 j(k, t) ∂0 d x x j (x, t)= ∂t d x x ρ(x, t)= d˙ (t), (5.59) ≈ Z Z where d(t) d3x x ρ(x, t) (5.60) ≡ Z is the dipole of the charge distribution. 396 5 Classical Radiation

Let us perform the integral over all angles in (5.54). For this we use (5.56) and the angular averages 1 2 kˆ kˆ = δ , P T (kˆ) = δ , (5.61) h i ji 3 ij h ij i 3 ij to obtain, for long wavelengths,

1 2 2 ˙ 3 j x 3 ∂t d x ( , t) . (5.62) E ≈ 4πc 3  Z  Inserting here Eq. (5.59), we find the famous dipole formula for the total radiated power 2 1 ˙ = [d¨(t)]2. (5.63) E 3c4 4π For a single nonrelativistic point particle moving along the orbitx ¯(t), the spatial current density is j(x, t)= e x¯˙ (t) δ(3)(x x¯(t)), (5.64) − and Eq. (5.64) becomes

2 e2 ˙ = [x¯¨(t)]2, (5.65) E 3c3 4π in agreement with the Larmor formula (5.38).

5.2 Classical Gravitational Waves

By analogy with the generation of electromagnetic waves, changes in mass distri- butions lead to changes of the gravitational field. Since the adjustment to a new field configuration can propagate with the speed of light, the universe must be filled with gravitational waves. The collapse of stars, explosion of supernovas, birth of neutron stars, and similar dramatic events in the universe must all be accompanied by bursts of such waves whose general properties will now be studied.

5.2.1 Gravitational Field of Matter Source The gravitational field is determined by Einstein’s equation. That is derived by extending the Einstein-Hilbert action (4.352) for the gravitational field by the action m of all matter. Its action will be denoted by and consists of a sum of the actions of various matter fields plus those of worldlinesA of massive point particles. If we vary m m µν the metric gµν in , we find the energy-momentum tensor T (x) of all matter from the resulting variationA

m m 1 4 µν δ = d x √ gδgµν(x) T (x). (5.66) A −2 Z − 5.2 Classical Gravitational Waves 397

A corresponding variation of the field action (4.352) yields

f 1 4 1 µν µν µν δ = d x√ g gµνδg R + δg Rµν + g δRµν A −2 Z − −2  1 4 µν 1 µν = d x√ g δg (Rµν gµνR)+ g δRµν , (5.67) −2 Z −  − 2  where we have used the relation

δ√ g = 1 √ ggµνδg = 1 g δgµν. (5.68) − 2 − µν − 2 µν The last term in (5.67) vanishes in spaces without torsion2, so that we can express f δ in terms of the Einstein tensor Gµν (x) of Eq. (4.375) as A f 1 4 µν 1 4 µν δ = d x √ gδg (x)Gµν (x)= d x √ gδgµν (x)G (x). (5.69) A −2 Z − 2 Z − If we finally extremize the total gravitational action

grav f m = + (5.70) A A A with respect to δgµν , we obtain the Einstein equation for the gravitational field in the presence of matter:

m Gµν = κ T µν . (5.71)

This corresponds to

m µ µ µ Gµ = Rµ = R = κ T µ (x). (5.72) − − We have seen in Eq. (4.408) that, in the weak-field limit, the free gravitational µν µν 1 µν field equation is simplest if written down in terms of the field φ = h 2 η h. Since the Einstein tensor is given by (4.376), the linearized Einstein equation− (5.71) has the somewhat involved differential form

m µν 1 2 µκ µ λκ κ λµ µ κ 1 µκ 2 νλ µν G = (∂ h ∂ ∂λh ∂ ∂λh + ∂ ∂ h)+ η (∂ h ∂ν ∂λh )=κ T .(5.73) −2 − − 2 − Similar to the electromagnetic field equation (31.15) in the presence of sources, the sources on the right-hand side are consistent with the Bianchi identity of the free gravitational field (4.379). In the electromagnetic case, this is due to the cur- µ rent conservation law (2.226), whose relativistic formulation is ∂µj = 0. Here the Bianchi identity for Gµν is a consequence of the conservation law of the symmetric energy-momentum tensor:

m µν ∂ν T =0. (5.74)

2See Section 15.2 in the textbook Ref. [17]. 398 5 Classical Radiation

The symmetry basis of this conservation law will be discussed in Chapter 8 [see in particular Eq. (8.319)]. To solve the field equation (5.73) it is convenient to express the differential operator on the left-hand side in an easily invertible form. For this purpose we introduce the combination of products of transverse projection operators (4.276) [compare (4G.4)]: 1 1 P (2) (i∂) [P t (i∂)P t (i∂)+ P t (i∂)P t (i∂)] P t (i∂)P t (i∂). (5.75) µν,λκ ≡ 2 µλ νκ µκ νλ − 3 µν λκ

This projects the symmetric tensor field hµν (x) into the irreducible spin-2 subspace. We further introduce the spin-0 projection operator [compare (4G.6)]: 1 P s P t P t . (5.76) µν,λκ ≡ 3 µν λκ The two operators (5.75) and (5.76) can be used to express the left-hand side of (5.73) as follows:

G = 1 [∂2h ∂ ∂λh ∂ ∂λh + ∂ ∂ ∂ˆλ∂ˆκh +(∂ˆ ∂ˆ η )∂2h ] µν − 2 µν − µ λν − ν λµ µ ν λκ µ ν − µν s = 1 P (2) ∂2hλκ + 1 P t P t ∂2hλκ − 2 µν,λκ 3 µν λκ = 1 P (2) ∂2hλκ + P s ∂2hλκ. (5.77) − 2 µν,λκ µν,λκ

Using the decomposition (4.436) of the symmetric tensor field hµν , this equation can be written as

G (x)= 1 ∂2h(2)(x)+ hs (x), (5.78) µν − 2 µν µν and the field equation (5.73) takes the form

m 1 ∂2h(2)(x)+ ∂2hs (x)= κ T µν (x). (5.79) − 2 µν µν (2) s Multiplying this successively by Pµν,λκ and by Pµν,λκ, we find the two projected equations:

m m 1 2 (2) (2) λκ (2) 2 ∂ hµν (x) = κPµν,λκ T (x) κ T µν (x), (5.80) − m ≡ m ∂2hs (x) = κP s T λκ(x) κ T s (x). (5.81) µν µν,λκ ≡ µν The second equation determines the scalar part of the field hs h λ ∂ˆ ∂ˆ hµν from ≡ λ − µ ν m 2 s t λκ ∂ h (x)= κPλκ T (x). (5.82) Combining (5.80) and (5.81) we find

2κ (2) m h(2) + hs = P 1 P s T λκ (5.83) µν µν − ∂2 µν,λκ − 2 µν,λκ   m κ µλ νκ µκ νλ µν λκ = P P +P P P P T λκ. −∂2 t t t t − t t   5.2 Classical Gravitational Waves 399

At this place we can make use of the energy-momentum conservation law (5.74) on the right-hand side to rewrite this as

m m (2) s 2κ 1 λ hµν (x)+ hµν (x) = 2 T µν 2 ηµν T λ (x). (5.84) − ∂  −  Here it is useful to introduce the combination of energy-momentum tensors on the right-hand side as

m m m 1 λ T µν(x) T µν(x) ηµν T λ (x). (5.85) ≡ − 2 The corresponding field combination of the fields on the left-hand side is

h¯ (x) h(2)(x)+ hs (x). (5.86) µν ≡ µν µν This has (5.85) as its source, i.e.:

2κ m h¯ (x)= T (x). (5.87) µν − ∂2 µν This satisfies the Hilbert gauge condition (4.399). For the field φ (x)= h (x) 1 g h(x), this equation implies that µν µν − 2 µν 2κ m φµν(x)= T µν (x), (5.88) − ∂2

ν and the energy-momentum conservation law (5.74) ensures that ∂ φµν = 0, thus guaranteeing the Hilbert gauge [see (4.406)]. These are direct analogs of the elec- trodynamic field equation (5.4). Using (5.77), the quadratic part of the field action (4.378) takes the form

f 1 4 µν 1 4 (2) s 2 λκ = d x hµν G = d x hµν P 2P ∂ h , (5.89) A 4κ −8κ µν,λκ − µν,λκ Z Z   which can also be written as

f 1 = d4x h(2)µν ∂2h(2) 2hsλκ∂2hs . (5.90) A −8κ λκ − λκ Z h i m To this we have to add the coupling of matter coming from the linearized inter- action A m m int 1 4 µν = = d x hµν (x) T (x). (5.91) A A −2 Z This can be decomposed into spin-2 and scalar parts as

m 1 (2) m = int = d4x hµν (x) P + P s T µν(x), (5.92) A A −2 µν,λκ µν,λκ Z   400 5 Classical Radiation or

m 1 m = int = d4x h(2)(x)+ hs (x) T µν (x). (5.93) A A −2 µν µν Z h i The extremum of the sum of the gravitational actions (5.90) and (5.91)

grav f m = + (5.94) A A A lies at the field (5.86). This equation can be solved by analogy with (5.5) and (5.95), yielding

m 2κ 3 ′ 1 ′ hµν (x)= d x T µν (x ) . (5.95) −4πc x x′ t′=t−|x′−x|/c Z | − | m A similar equation holds for φµν expressed in terms of T µν(x′). Reinserting (5.86) into (5.94) yields the gravitational analog of the Biot-Savart law (5.20) in momentum space:

grav m m κ 4 µν 1 ∗ BS= d k T (k) 2 T µν (k) , (5.96) A − 2 Z k where

m m m m m m µν ∗ µν ∗ 1 µ ν ∗ T (k) T µν(k) = T (k) T µν (k) 2 T µ (k) T ν (k) m m − m m = T (2)µν (k) T (2)(k)∗ 1 T s µν (k) T s (k)∗. (5.97) µν − 2 µν The latter form follows directly from extremizing the sum of (5.90) and (5.93). It is a consequence of the propagator of the field hµν obtained from the free-field action (5.90):

P (2) (k) P s (k) h (k)h∗ (k) =4κ µν,λκ 1 µν,λκ , (5.98) h µν λκ i  k2 − 2 k2    after coupling it to the source via (5.93). For two static point charges of masses M and M ′ at the origin and x with the m m energy-momentum densities T 00(x′) = Mc2δ3(x) and T 00(x′) = M ′c2δ3(x′ x), this amounts to an interaction energy −

κc4MM ′ G MM ′ int = = N . (5.99) E − 4πr − 2r From this we derive the famous attractive Newton force between the mass points

G MM ′ F = N . (5.100) − r2 5.2 Classical Gravitational Waves 401

Evaluating the integral in Eq. (5.95) for x far away from the source, in the radiation zone, we may approximate 1/ x x′ 1/ x 1/r. Then φµν(x, t) | − | ≈ | | ≡ behaves locally like an outgoing plane wave. Thus it must be purely transverse, i.e., µν µν it is some linear combination of the polarization tensors ǫH+ and ǫH× of Eqs. (4.425) and (4.426). These have no 00-component and are traceless [recall (4.418)]. Thus only the spatially traceless part of φµν will contribute and we can approximate:

m m ij ij 2κ 3 ′ ij ′ ′ 1 ij k ′ ′ φ ϕ d x [T (x , t x x /c) 3 δ T k (x , t x x /c)]. (5.101) ≈ ≡−4πr Z −| − | − −| − | In a linear approximation, the energy-momentum term of the source is conserved `ala (5.74):

m µν ∂ν T =0+ .... (5.102)

The neglected terms come from the contribution of the gravitational field to the energy-momentum tensor. These are of order (κ). Neglecting them we have O m m i0 ij ∂0 T (x, t) = ∂j T (x, t), (5.103) m − m 00 0j ∂0 T (x, t) = ∂j T (x, t). (5.104) − From these equations we find that

m m 3 ′ 0j ′ 3 ′ ′i 0k ′ d x T (x , t) = d x x ∂k T (x , t) Z Z m 3 ′ ′i 00 ′ = ∂0 d x x T (x , t). (5.105) Z We see further that m m m 3 ′ ij ′ 3 ′ ′j ik ′ 3 ′ ′j i0 ′ d x T (x , t) = d x x ∂k T (x , t)= d x x ∂0 T (x , t) − Z Z m Z m 3 ′ i jk ′ 3 ′ ′i j0 ′ = d x x ∂k T (x , t)= d x x ∂0 T (x , t) − Z Z m m 1 3 ′ ′i j0 ′ ′j i0 ′ = ∂0 d x [x T (x , t)+ x T (x , t)] (5.106) 2 Z and that m m m 3 ′ ′j i0 ′ ′i j0 ′ 3 ′ ′i ′j k0 ′ d x [x T (x , t)+ x T (x , t)] = d x x x ∂k T (x , t) − Z Z m 3 ′ ′i ′i 00 ′ = ∂0 d x x x T (x , t). (5.107) Z Hence

m m 3 ′ ij ′ ′ 1 2 3 ′ ′i ′j 00 ′ ′ d x T (x , t )= ∂0 d x x x T (x , t ). (5.108) Z 2 Z Note the analogy with the electromagnetic equation (5.59). 402 5 Classical Radiation

5.2.2 Quadrupole Moment The equality (5.108) permits us to express the right-hand side of (5.101) in terms of the quadrupole moment

1 1 m Qij(t) d3x′ x′ix′j r2δij T 00(x′, t). (5.109) ≡ c Z  − 3  For the field emitted by a moving gravitational source we obtain therefore the simple equation κ 1 ϕij(x, t)= Q¨ij(t ), (5.110) 4πr c R where the retarded time is given by (5.46). Note the analogy with the electromag- netic formula (5.48). The field components φµν (x, t) determine the energy carried away by the wave. To quantify this, we need the analog of the Poynting vector for linearized gravity. This is supplied by the symmetric energy-momentum tensor. For weak fields, it can be derived from the quadratic action (4.372) by techniques to be developed later in Section 8. Anticipating the result we shall find, in Eq. (8.262), the symmetric µν energy-momentum tensor in the Hilbert gauge ∂µφ = 0:

f µν 1 µ λκ ν µ ν µν σκ λ 1 λ T = 2∂ φ ∂ φλκ ∂ φ∂ φ η ∂λφ ∂ φσκ ∂λφ∂ φ . (5.111) 8κ − − − 2  f   If the components T 0i are multiplied with the unit vector of the outgoing wave kˆ to form

f 2 2 0iˆi c 0 λκ i 0 i ˆi c T k = 2∂ φ ∂ φλκ ∂ φ∂ φ k , (5.112) 8κ − h i we obtain the energy current density along the kˆ-direction. Separating φµν into space and time parts, we can write

d ˙ f c2 E = c2 T 0ikˆi = 2∂0φkl∂iφkl 4∂0φk0∂iφk0 +∂0φ00∂iφ00 r2dΩ 8κ − h + ∂0φkk∂iφ00 + ∂0φ00∂iφkk ∂0φkk∂iφkk kˆi. (5.113) − i This has to be integrated over the surface of a large sphere with infinite radius. Before doing so we note that at large r we can forget all space derivatives of the prefactor 1/r since they give nonleading 1/r2, 1/r3,... contributions. We have to keep, however, derivatives with respect to r arising from the retarded time argument [recall the approximation (5.44)]. Hence we can approximate, to leading order in 1/r,

∂iφµν ∂0φµν∂ir = ∂0φµν kˆi. (5.114) ≈ − Applying this to the components φkl brings the first term in (5.113) to the form ˙kl2 i 00 00ˆi 2φ , while ∂ φ becomes ∂0φ k . It is furthermore possible to express the time 5.2 Classical Gravitational Waves 403 derivatives on the right-hand side of (5.113) in terms of time derivatives of the purely spatial field components of Eq. (5.88) in the Hilbert gauge (4.406):

∂ φk0 = ∂ φki = ∂0φkikˆi, (5.115) 0 − i implying that

∂ φ00 = ∂ φi0 = ∂ φi0kˆ = ∂ φijkˆ = φijkˆ kˆ . (5.116) 0 − i 0 i − j j i j Inserting on the right-hand side the asymptotic traceless expressions φij of Eq. (5.101), we finally obtain the radiated energy current, which becomes, to leading order in 1/r,

˙ f d 2 0iˆi 1 2 ˆlˆm 1 kl mr ˆ ˆ ˆ ˆ 2 E = c T k = ϕ˙ kl 2ϕ ˙ klϕ˙ kmk k + ϕ˙ ϕ˙ kkklkmkr . (5.117) r dΩ 4κ  − 2  The terms in brackets contain the transverse projection matrix (4.434). They can therefore be written as contractions with the polarization tensors (4.425), (4.426), and thus because of (4.427), as

ǫkl(kˆ)ϕ ˙ 2 + ǫkl(kˆ)ϕ ˙ 2 =ϕ ˙ P kl,mn(kˆ)ϕ ˙ . (5.118) | + kl| | × kl| kl T T mn Inserting here the field (5.110), we obtain the rate of energy emitted per unit solid angle dΩ:

d ˙ f 1 κ 1 E = r2c T 0ikˆi = Q˙¨ P kl,mn(kˆ)Q˙¨ . (5.119) dΩ 4 (4π)2 c2 kl T T mn This has a complicated angular dependence. However, the integral over all directions is easily found using the angular averages of products of all direction vectors 1 1 kˆ kˆ = δ , kˆ kˆ kˆ kˆ = δijδkl + δikδjl + δilδjk . (5.120) h i ji 3 ij h i j k mi 15   The tensor structure on the right-hand side follows directly from the rotational symmetry. The normalizations are found by contracting the indices, using nini = 1. Inserting (5.120) into (4.428) yields 2 1 1 P ij,kl(kˆ) = δikδjl + δilδjk δijδkl . (5.121) h T T i 5 2 − 3     Now the angular integral is straightforward, and we find, recalling κ from (4.357),

1 κ 1 2 kl G kl G 1 ˙ = (Q˙¨ )2 = (Q˙¨ )2, . (5.122) E 4 4π c2 5 5c5 c5 ≈ 3.6 1052 W × For comparison, recall the analogous electromagnetic radiation formula (5.35), where the direction-dependent energy loss due to dipole radiation is

d ˙ 1 E = kˆ (kˆ d¨) 2, (5.123) dΩ 8πc3 | × × | 404 5 Classical Radiation with the total radiated power: 1 ˙ = d¨2. (5.124) E 3c3 Since a quadrupole moment possesses a reference point, the reader may wonder about the translational invariance of the gravitational radiation formulas (5.122). Consider the second moment of the energy-momentum tensor after a translation m m 3 00 3 00 d x (xi ai)(xj aj) T (x, t)= d x xixj T (x, t) − − Z m Z m m 3 00 3 00 3 00 ai d x xj T (x, t) aj d x xi T (x, t)+ aiaj d x T (x, t). (5.125) − Z − Z Z We now observe that the last term is time-independent because of energy conserva- tion. The other two terms in the second line may be rewritten using the conservation µν law ∂µT (x, t) and a partial integration: m m m 2 3 00 3 i0 3 ik ∂ d x xj T (x, t)= ∂0 d x xj∂i T (x, t)= d x xj∂i∂k T (x, t) 0 − Z Z m Z m 3 jk 3 j0 = d x ∂k T (x, t)= ∂0 d x T (x, t)=0. (5.126) − Z − Z The last zero on the right-hand side follows from momentum conservation. Thus Qij(t) changes at most by a linear function of t, and formula (5.124) is indeed independent of the choice of the reference point when calculating the quadrupole moment.3 There exists another way of stating the radiation formula (5.119) that is more similar to the electromagnetic formula (5.55). We employ the relation (5.88) and the transverse traceless projection matrix (4.431) to replace (5.119) by

d2 ˙ κ m m k ∗ kl,mn k k E = 2 [∂t T kl( , t)] PT T ( )∂t T mn( , t) . (5.127) dΩ 16π   Now we make use of the energy-momentum conservation law (5.103) in Fourier m m m m ˆ iµ 0µ ˆ ˆ ij 00 space ki T (k,ω)= T (k,ω), kikj T (k,ω)= T (k,ω), and rewrite this in a fully covariant form analogous to (5.57) as

d ˙ κ m 2 m 2 ν k 1 λ k E = 2 ∂t T µ ( , t) 2 ∂t T λ ( , t) . (5.128) dΩ 16π ( − )

For long wavelengths, the spatial components of the energy-mom entum tensor m m m have a negligible dependence on k: T ij(k, t) T ij(k = 0, t) = d3x T ij(x, t). ≈ Then we can perform the integral over all angles in (5.128) using the angularR averages (5.120), yielding

κ 2 m 2 m 2 ˙ 3 j 1 3 i = ∂t d x T i (x, t) 3 ∂t d x T i (x, t) . (5.129) E 4π 5 ( Z  −  Z  ) 3For more subtle aspects of gravitational radiation see F.I. Cooperstock and P.H. Lim, Phys. Rev. Lett. 55, 265 (1985). 5.2 Classical Gravitational Waves 405

The right-hand side can be rewritten as

κ 2 m m m m ˙ 3 j 1 j k i 1 i k = d x∂t T i (x, t) 3 δi T k (x, t) ∂t T j (x, t) 3 δj T k (x, t) . (5.130) E 4π 5Z  −   −  As before we use (5.108) and (5.110) to replace

m m 1 3 ij 1 ij k ¨ij d x T (x, t) 3 δ T k (x, t) = Q , (5.131) Z  −  2c and obtain once more the flux equations (5.122).

5.2.3 Average Radiated Energy

If we express the time-dependent energy-momentum tensors in Eq. (5.128) in terms of their Fourier integrals, we obtain the radiated power per solid angle

˙ d κ ′ ′ −i(ω−ω′)t E = 2 dωdω ωω e dΩ 16π Z m m ∗ m m ν 1 ν λ µ ′ 1 µ λ ′ T µ (k,ω) 2 δµ T λ (k,ω) T ν (k,ω ) 2 δν T λ (k,ω ) . (5.132) ×  −   −  To understand the averaging process, consider the double frequency integral

′ f(t) dωdω′ωω′e−i(ω−ω )tA(ω)B(ω), (5.133) ≡ Z with smooth functions A(ω) and B(ω). Over a long time, the oscillation of the frequency difference ∆ω = ω ω′ cancel each other, and only the average rate of the integral −

t2 ′ dtei(ω−ω )t 2πδ(ω ω′) (5.134) Zt1 → − survives, implying that

2π f˙(t) dωω2A(ω)B(ω). (5.135) ≈ t t 2 − 1 Z

Integrating this over all times from t1 to t2 gives therefore the total radiated energy per unit solid angle:

m m ∗ m m d κ 2 ν 1 ν λ µ 1 µ λ E = dωω T µ (k,ω) 2 δµ T λ (k,ω) T ν (k,ω) 2 δν T λ (k,ω) . dΩ 8π  −   −  Z (5.136) 406 5 Classical Radiation

5.3 Simple Models for Sources of Gravitational Radiation

Let us calculate the radiated power for a few typical radiating systems. Following the textbook [3] we distinguish between oscillating and bursting systems. Consider a single nonrelativistic point particle moving along the orbitx ¯(t). Its energy-momentum tensor has the spatial components

m M i j T ij(x, t)= x¯˙ (t)x¯˙ (t) δ(3)(x x¯(t)). (5.137) c − Then Eq. (5.129) becomes

κ′ 2 M 2 1 2 ˙ ˙ i ˙ j ij ˙ k ˙ = 2 ∂t x¯ x¯ δ x¯ x¯k . (5.138) E 4π 5 c   − 3  The radiated energy decreases in time, and so does the total angular momentum of the gravitational system

m 3 j 0k Li(t)= ǫijk d x x T (x, t). (5.139) Z This happens at a rate

′ 2κ 1 il im 2G il im L˙ = ǫ Q˙¨ (t)Q˙¨ (t)= ǫ Q˙¨ (t)Q˙¨ (t). (5.140) k 8π 5c2 klm 5c5 klm For checking the dimensions of the above equations the following list of dimen- sions is useful:

h¯ h¯ h¯ ij h¯ [ ]=¯h, [T µν ]= , [E]= , [p]= , [Qij]=¯h sec, [ Q˙¨ ]= , A cm4 sec cm sec2 cm2 cm5 h¯ [κ]= , [G]= , [c5/G]= 3.6 1052 W. (5.141) h¯ h¯ sec3 sec2 ≈ ×

ij Note that Q˙¨ has the dimension of a power.

5.3.1 Vibrating Quadrupole Imagine two equal masses M oscillating at the ends of a spring (see Fig. 5.1). Their time-dependent distance is

d z = + a sin ωt . (5.142) ± 2 !

Assuming that a d, we may approximate ≪ d2 z2 + ad sin ωt. (5.143) ∼ 4 5.3 Simple Models for Sources of Gravitational Radiation 407

Figure 5.1 Two equal masses M oscillating at the ends of a spring as a source of gravi- tational radiation.

The quadrupole moment is therefore 4a Qij(t)= 1+ sin ωt Qij (5.144)  d  where 1 Md2 − Q (0) = 1 , (5.145) ij 6   − 2     so that 1 2Mad Q˙¨ (t)= Q˙¨ (t)= Q˙¨ (t)= ω3 sin ωt . (5.146) 11 22 −2 22 − 3 According to Eq. (5.119), the angular distribution of the quadrupole radiation in the direction n = kˆ is (see Fig. 5.1)

d ˙ 1 κ 2 2 2 E = Q˙¨ + Q˙¨ + Q˙¨ 2 (Q˙¨ n )2 +(Q˙¨ n )2 +(Q˙¨ n )2 dΩ 4 (4π)2c2 11 22 33 − 11 1 22 2 33 3 n h i 1 ˙¨ ˙¨ ˙¨ 2 + (Q11n1n1 + Q22n2n2 + Q33n3n3) . (5.147) 2  Introducing spherical angles

n1 = sin θ cos φ, n2 = sin θ sin φ, n3 = cos θ (5.148) for the direction n, and assuming that Q11 = Q22, the curly brackets in (5.147) become

˙¨2 ˙¨2 ˙¨2 2 ˙¨2 2 1 ˙¨ 2 ˙¨ 2 2 2Q11 + Q33 2(Q11 sin θ + Q33 cos θ)+ [Q11 sin θ + Q33 cos θ] . (5.149)  − 2  For Q = 2Q , this reduces to 33 − 11 2 9 Q˙¨ sin4 θ . (5.150) 11 2 408 5 Classical Radiation

The rate of energy radiation per solid angle dΩ is then

d ˙ 1 κ E = M 2[ω3ad sin ωt]2. (5.151) dΩ 2c2 (4π)2 The radiation is maximal in the direction of the equator, and vanishes in the pole directions of the oscillator. Integrating (5.151) over all angles gives the total emitted power 1 κ 32 ˙ = M 2[ω3ad sin ωt]2 π E 2c2 (4π)2 15 8G = M 2[ω3ad sin ωt]2, (5.152) 15c5 whose temporal average is 8G ˙ = M 2ω6a2d2. (5.153) E 15c5 The rate of radiation damping is defined as 1 1 dE γrad , (5.154) ≡ trad ≡ E dt

2 2 where trad is the damping time. Since the kinetic energy of each mass is Mω a /2, we obtain 46 γ = Md2ω4. (5.155) rad 15c5 The formula estimates the damping rate of any linearly oscillating system of two masses. The linear character of the oscillation is important since there is no gravi- tational radiation at all for a spherically symmetric pulsating star. Vibrational radiation may emerge from nova explosions at an early stage. These arise if a star circles around a white dwarf and transfers matter to him. After some time, the matter becomes large enough to explode. This explosion causes vibrations in the white dwarf with frequencies 0.01 to 1 Hz. The energy released in a nova explosion is typically 1045 erg, of which 10% could be deposited in vibrations, which send out gravitational radiation.

5.3.2 Two Rotating Masses If the two masses in the previous example rotate around the z-axis (see Fig. 5.2) the quadrupole moment (5.145) in the xy-plane becomes

Md2 1 3cos2ωt 3 sin 2ωt Qij(t)= − − , (5.156) 4 3 sin 2ω 1+3cos2ωt − ! where M is the reduced mass M M M 1 2 , (5.157) ≡ M1 + M2 5.3 Simple Models for Sources of Gravitational Radiation 409

Figure 5.2 Two spherical masses in circular orbits around their center of mass. and ω is given by the third Kepler law:

G(M + M ) ω = 1 2 . (5.158) s r3 The third time derivatives of the quadrupole moments are therefore sin 2ωt cos2ωt ˙¨ 2 3 − − Qij(t)= 6Md ω  0 0  . (5.159) − cos2ωt sin 2ωt    −  Inserting these into (5.147), integrating over all angles, and averaging over all times yields the total emitted power 8G ˙ = M 2d4ω6. (5.160) E 5c5 Using ω of Eq. (5.158), this becomes 32G4 ˙ = (M M )2(M + M ). (5.161) E 5c5d5 1 2 1 2 The total energy of the binary system is 1 M M 1 GM M = 1 2 d2ω2 = 1 2 , (5.162) E 2 M1 + M2 2 d implying a rate of radiation loss 64G4 γ = M M (M + M ). (5.163) 5c5d4 1 2 1 2 Since E is inversely proportional to the distance d in (5.162), the relative decrease of the distance between the masses is d˙ 64G3 = M M (M + M ). (5.164) d −5c5d4 1 2 1 2 By Eq. (5.158) this implies that the frequency increases at a rate

ω˙ 3 d˙ = . (5.165) ω −2 d 410 5 Classical Radiation

Due to the smallness of the gravitational constant, the power radiated by planetary systems is extremely small. The Earth orbiting around the sun emits only 200 W, the Jupiter emits 5300 W. For neighboring double stars, the power can increase to 1030 W and more. For double neutron stars, d can be quite small, and the radiated power can easily reach values of 1045 W. Table 5.1 shows various astronomical objects and the gravitational amplitudes that can arrive from them here on Earth. See also the illustration on Fig. 5.3.

Table 5.1 Binary systems as sources of gravitational radiation [4]. The binary PSR 1913+16 emits radiation at multiples of 70 10−6 Hz due to the large eccentricity of × the orbit. System Masses Dist. Wave Luminosity at Amplitude frequ. atEarth Earth atEarth −6 30 −22 (M⊙) (pc) (10 Hz) (10 erg/s) (10 ) Eclipsing binaries ı Boo 1.0, 0.5 11.7 86 1.1 68.0 51 µ Sco 12,12 109 16 51 38.0 210 V Pup 16.5, 9.7 520 16 59 1.9 46 Cataclysmic binaries (novas) AMCVn 1.0,0.041 100 1900 300 240 5 WZSge 1,50.12 75 410 24 37 8 SSCyg 0.97,0.8330 84 2 20 30 Binary X-ray sources (black holes or neutron stars) Cyg X1 30,6 2500 4.1 1.0 1 4 PSR 1913+16 1.4, 1.4 5000 70 0.6 0.2 0.12 140 2.9 1.1 0.14 210 5.8 2.1 0.12

Formula (5.165) has been used as an indirect evidence for the existence of gravi- tational radiation. In 1974, Hulse and Taylor searched for pulsars (rotating neutron stars emitting radio pulses) with the Arecibo telescope and found an object whose emitted radio frequency is periodically modulated. The modulation is attributed to the Doppler shift caused by the orbital motion around an undetected companion. A careful analysis of the modulation allowed them to derive the eccentricity and the rate of the perihelion precession of the binary object (see Table 5.2 for details). The observed shift of the time of periastron passage is plotted in Fig. 5.4. The values of the masses were deduced from the perihel precession and the time delay of the signal passing the companion. They depend only on M1 + M2, M1, and M2 in different combinations. From the data, one deduces

M (1.4414 0.0002)M for pulsar, (5.166) 1 ∼ ± ⊙ M (1.3867 0.0002)M for companion. (5.167) 2 ∼ ± ⊙ 5.3 Simple Models for Sources of Gravitational Radiation 411

Figure 5.3 Gravitational amplitudes arriving on Earth from two possible sources. They are measured by LISA, which is a Laser Interferometer Space Antenna built by a joint three-spacecraft mission of ESA and NASA (see http://www.esa.int/ esaSC/120376 index 0 m.html). The other is masured by LIGO, a Laser Interferometer Gravitational Wave Observatory (see http://www.ligo.caltech.edu). Correlations must be monitored for distances between two objects lying 5 106 km apart. The letters × WDB in the box near (10−3, 10−22) denotes White Dwarf Binaries.

The properties of binary objects containing a pulsar can be studied so well that the approximation (5.161) is sensitive to several corrections [6]. Consider two masses orbiting around the common center-of-mass (see Figs. 5.5 and 5.6). If d denotes the distance of the two masses, the distances from the center-of-mass are d1 = M2d/(M1+M2), d2 = M1d/(M1+M2). Denoting the reduced mass M1M2/(M1+M2) as before by M, the components of the quadrupole moment are

2 2 cos ϕ sin ϕ cos ϕ Qij(t)= Md 2 . (5.168) sin ϕ cos ϕ sin ϕ ! The orbit of a Kepler ellipse has the general form a(1 e2) d = − , (5.169) 1+ e cos ϕ where a is the semi-major axis of the ellipse, e 1 b2/a2 is the eccentricity (b ≡ − denotes the semi-minor axis), and the angular velocityq is given by

ϕ˙(t)= d−2(t) (M + M )a(1 e2). (5.170) 1 2 − q From this we derive immediately ˙¨ 2 2 Q11 = P (1 + e cos ϕ) (2 sin 2ϕ +3e sin ϕ cos ϕ), (5.171) ˙¨ 2 2 Q22 = P (1 + e cos ϕ) [2 sin 2ϕ + e sin ϕ(1+3cos ϕ)], (5.172) Q˙¨ = P (1 + e cos ϕ)2[2 sin 2ϕ e sin ϕ(1 3 cos2 ϕ)] = Q˙¨ , (5.173) 12 − − − 21 412 5 Classical Radiation

Table 5.2 Some observed parameters of PSR 1913+16 (Table taken from Ref. [5]). Distance 21 000 ly Pulsar period (nominal) 59.02999792988 ms Semi-major axis 1 950 100 km Eccentricity 0.617131 0.000003 ± Orbital period 27907 0.00002 s Rateofprecession ofthe periastron (4.22263± 0.0003)o/y Amplitude of time-dilation factor 0.0044 ±0.0001 ± Rateofdecreaseoforbitalperiod 2.4184(9) 10−12 s/s=0.0000765s/y Rate of decrease of semimajor axis 3.5 m/y × Calculated lifetime (to final inspiral) 300 000 000 y Diameter of each neutron star 20 km Periastronseparation 746600km Apoastronseparation 3153600km Velocity of stars at periastron in CM frame 450 km/sec Velocity of stars at apoastron in CM frame 110 km/sec Rate of precession of spin axis ??

Figure 5.4 Shift of time of the periastron passage of PSR 1913+16 for each orbit caused by the shrinking of the Kepler orbits as a consequence of formula (5.179). Curve is from Einstein theory see Eq. (5.179) and Ref. [5] . { } where P is a power factor

G3M 2M 2(M + M ) P 2 1 2 1 2 . (5.174) v 5 2 5 ≡ u a (1 e ) u − t 5.3 Simple Models for Sources of Gravitational Radiation 413

Figure 5.5 Two pulsars orbiting around each other.

Figure 5.6 Two masses in a Keplerian orbit around the common center-of-mass.

Inserting (5.171)–(5.173) into the radiation power formula (5.147), and integrating over all emission angles dΩ yields the total emitted power as a function of time t:

8G4 M 2M 2(M + M ) ˙ = 1 2 1 2 [1+e cos ϕ(t)]4 12[1+e cos ϕ(t)]2 +e2 sin2 ϕ(t) . (5.175) E 15c5 a5(1 e2)5 { } − The time dependence of the semi-major axis a(t) follows the differential equation

a˙ 64G3 M M (M + M ) 73 37 = 1 2 1 2 1+ e2 + e4 . (5.176) a − 5c5 a4(1 e2)7/2 24 964 −   The excentricity e(t) changes in time according to

e˙ 304G3 M M (M + M ) 121 = 1 2 1 2 1+ e2 . (5.177) e − 15c5 a4(1 e2)5/2 304 −  

An arbitrary Kepler orbit will shrink to zero in the coalescence time τcoales. For a cicular orbit of semi-major axis a0 = a(0) and e0 = e(0) = 0, this is

3 1 4 64G M1M2(M1 + M2) = · 5 4 . (5.178) τ coales − 5c a0 414 5 Classical Radiation

1.

0.5

˙ 0 E .5 − 1. − 0t 0.05 Figure 5.7 Energy emitted by two point-masses on a circular orbit around each other.

For a circular orbit, the time dependence is shown in Fig. 5.7. Averaging over one period of the elliptical orbit yields 32G4 M 2M 2(M + M ) 73 37 ˙ = 1 2 1 2 1+ e2 + e4 . (5.179) E 5c5 a5(1 e2)7/2 24 96 D E −   With respect to a circular orbit of equal total energy, the power is enhanced by a factor 1+ 73 e2 + 37 e4 f = 24 96 , (5.180) (1 e2)7/2 − which grows rapidly from f(e)=1at e = 0 to infinity at e = 1. For a full multipole analysis of the radiation see [6]. Due to the shrinking of the Kepler orbits implied by formula (5.179), the orbital time shrinks according to the elliptic generalization of Eq. (5.165), implying that the periastron is reached a few seconds earlier for each orbit. The time shift was plotted in Fig. 5.4. The radiation properties of binary objects can be studied especially well if both stars are pulsars (recall Fig. 5.5). Such astronomical objects have recently been found. One of the two pulsars rotates with a period of 23 milliseconds (PSR J0737- 3039A) around its axis, the other with a period of 2.8 seconds (PSR J0737-3039B) [7].

5.3.3 Particle Falling into Star Among the bursting sources of gravitational radiation the simplest one consists of a mass falling into a star (see Fig. 5.8) emitting a spectrum shown in Fig. 5.9. If the mass starts at z = , its velocity is ∞ 1 GmM mz˙2 = , (5.181) 2 z 5.3 Simple Models for Sources of Gravitational Radiation 415 so that 1 GM (2GM)3/2 z˙ = (2GM)1/2, z¨ = , z¨˙ = . (5.182) −z1/2 − z2 − z7/2 The triple time derivative of the quadrupole moment

z2 0 0 − 2 Qij = m  0 z 0  (5.183) 0− 02z2     is 1 0 0 − Q˙¨ = m(6z ˙z¨ +2z z¨˙) 0 1 0 , (5.184) ij  −  0 0 2     and thus, because of (5.182),

1 0 0 (2GM)3/2 − Q˙¨ = m 0 1 0 . (5.185) ij z5/2   0− 0 2     Inserted into (5.122), this leads to an energy loss per second

2Gm2 ˙ = (6z ˙z¨ +2zz¨˙)2. (5.186) E 15c5 Combining this with (5.182) implies

d 1 2Gm2 E = (2GM)5/2. (5.187) dz z9/2 15c2

Figure 5.8 Particle falling radially towards a large mass. 416 5 Classical Radiation

Figure 5.9 Spectrum of the gravitational radiation emitted by a particle of mass m falling radially into a black hole of mass M. The quantity dE/dω gives the amount of energy radiated per unit frequency interval. The curve marked l = 2 corresponds to quadrupole radiation; the other curves (l = 3, l = 4) correspond to multipole radiation of higher order. Note that most of the radiation is emitted with frequencies below ω 0.5c3/GM [8]. ≃

The radiated energy from z = to z = R is ∞ 1 4Gm2 = (2GM)5/2. (5.188) E R7/2 105c5 Obviously, the radiated energy increases with decreasing R. Suppose the large object is a black hole. If we let the mass m fall down to the Schwarzschild radius: GM R =2 , (5.189) S c2 we obtain 2 m m = mc2 0.019mc2 . (5.190) E 105 M ≈ M If one takes relativistic effects into account which are due to the deviations of the Schwarzschild metric from flat space as the particle approaches Rs, the number 0.019 changes to 0.0104. For a black hole of mass M 10M⊙ with Schwarzschild Rs 30km, the total radiated energy is ∼ ∼ 2 1051erg. (5.191) E ∼ × The radiated energy is mostly emitted near the end of the process. See also Table 5.3 for a list of possible sources of astrophysical gravitational radiation. 5.3 Simple Models for Sources of Gravitational Radiation 417

Table 5.3 Typical astrophysical sources of gravitational radiation. Distances have been selected large enough to yield approximately three events per year [9]. Source Frequency Distance Amplitude (κA) Periodic sources Binaries 10−4 Hz 10pc 10−20 Nova 10−2 to1 500pc 10−22 Spinning neutron star (Crab) 60 2 kpc < 10−24 Bursting sources Coalescence of binary 10 to 103 100 MPC 10−21 −4 −21 Infall of star into 10M⊙ b.h. 10 10Mpc 10 Supernova 103 10kpc 10−18 4 −1 −19 Gravitational collapse of 10 M⊙ star 10 3Gpc 10

5.3.4 Cloud of Colliding Stars Let us treat the stars in a cloud approximately as point-like objects moving with velocities vn, where they have an energy-momentum tensor pµpν T µν = δ(3)(x x(t)). (5.192) n En − X Suppose the stars all collide at the origin at t = 0 and run away with changed velocities v¯n. Then the time-dependent energy-momentum tensor is

µ ν µ ν µν pnpn (3) p¯np¯n (3) T (x, t)= δ (x vnt)Θ( t) ¯ δ (x v¯nt)Θ(t), (5.193) n En − − − n En − X X where Θ(t) is the Heaviside function defined in Eq. (1.311). Representing this as a Fourier integral ∞ dω eiωt Θ(t)= i , (5.194) − −∞ 2π ω iη Z − and the δ-function similarly as

d3k (3) x v ikx δ ( )= 3 e , (5.195) − Z (2π) we see that the energy momentum tensor (5.193) has the Fourier components

µ ν µ ν µν pnpn i p¯np¯n i T (k,ω)= ¯ . (5.196) − n En ω vn k iη − En ω vn k + iη ! X − · − − · Now we observe that E (v k ω)= p k, (5.197) n × n · − n 418 5 Classical Radiation where k is the four-momentum of the emerging gravitational wave. Since the stars are nonrelativistic objects, we can further drop the iη’s. Thus

pµpν p¯µp¯ν T µν (k,ω)= i n n n n . (5.198) − n pnk − p¯nk ! X This energy-momentum tensor satisfies k T µν = (p p¯ ) = 0, as it should by µ n n − n momentum conservation. Inserting (5.198) intoP (5.132) yields the total radiated energy per solid angle and frequency interval dω:

′ d κ 2 σnσn 2 1 2 2 ′ ′ E = 2 ω (pnpn ) MnMn , (5.199) dΩdω 16π ′ (pnk)(pn′ k) − 2 Xnn   where the index n runs over the particles before and σn is equal to 1 before the collisions, and 1 afterwards. Integrating over all directions of k yields − n

2 ′ d κ 2 1+ βnn′ 1 1+ βnn E = ω σnσn′ MnMn′ log , (5.200) 2 2 1/2 ′ ′ dΩdω 16π ′ (1 βnn′ ) βnn 1 βnn ! Xnn − − where βnn′ is the relative velocity divided by c:

1/2 MnMn′ ′ βnn 1 2 . (5.201) ≡ " − (pnpn′ ) #

The integral over all frequencies diverges. This is due to the fact that we have assumed an instantaneous change of momenta during the collisions. In actual col- lisions, the change takes place over some finite collision time ∆t and the integral contains frequencies up to 1/∆t. For nonrelativistic two-body scattering, the radi- ated energy is d 2κ dω E = µ2 v4 sin2 θ, (5.202) dΩ 5π Z 2π where µ is the reduced mass, v are the relative velocities, and θ the scattering angles in the center-of-mass frame.

5.4 Orders of Magnitude of Different Radiation Sources

In order to have an idea of the amounts of energy that can be radiated in various processes, consider a massive steel rod of radius 1m, length d=20m, mass M 4.9 108g (=490 tons, using the steel density 7.89g/cm3). If the maximal quadrupole∼ × radiation can be obtained by rotating it around an axis orthogonal to the rod with an angular velocity ω, then formula (5.122) yields the total emitted power 2 ˙ = M 2l4ω6. (5.203) E 45c5 5.4 Orders of Magnitude of Different Radiation Sources 419

The angular velocity is limited by the tensile strength, which is t 3 10gdyn/cm2. Thus one can maximally use an angular velocity ≈ ×

8t 1/2 1 ω = 28 . (5.204) ρl2 ! ≈ sec

This gives a radiated power

˙ 10−23erg/sec. (5.205) E ≈ In order to have an idea how small this is we note that a single photon in the visible range has an energy

2π 2π hc¯ 10−271010 erg 1.5 10−12erg. (5.206) 4000A˚ ≈ 4 10−5 ≈ × × Thus the radiated power corresponds to the emission of one visible photon in 1011 seconds or 3000 years. ≈ ≈ If one wants to have any observable effects it is therefore necessary to look for radiation emitted by large stellar objects. Consider a bunch of stars of total mass M distributed over a region of size R. Their velocity is of the order R/T where T is the time it takes for the masses to move from one side to the diametrally opposite one. Their quadrupole moment is of the order of

Q R2M. (5.207) ∼ Hence we can estimate

2 2 ˙¨ R M R 1 Q 3 M . (5.208) ∼ T ∼  T  T The right-hand side has the dimension energy per time. It gives an estimate for the internal power flow in the system. The radiated power is equal to

1 R2M 2 ˙ . (5.209) E ∼ c5/G T 3 !

For the Sun-Jupiter system, this time is 2.5 1023 years. But for the binary system PSR 1913+16, the spiral time shrinks to≈ 3 ×108 years, as shown in Table 5.2. This × makes it observable, as we have seen in Fig. 5.4. For Kepler orbits of planets, on the other hand, the radiation damping is too small to be observed. For neutron stars a few thousand kilometers apart, the spiral times could shrink to the order of years or days, so it could be observed, at least in principle. Here the problem lies in the identification of the source. 420 5 Classical Radiation

5.5 Detection of Gravitational Waves

In order to observe more details of the gravitational waves, consider a test particle whose equation of motion reads

µ du µ 1 µ ν λ = ∂bhλ ∂ hbc u u . (5.210) dτ −  − 2  If it is initially at rest or moving very slowly, the right-hand side reduces to

µ 1 µ 2 ∂0h 0 ∂ h00 c . (5.211) −  − 2  Since the two physical polarization tensors have neither 00 nor 0i components, this vanishes. Thus two particles retain their relative positions as seen from the back- ground Minkowski frame. This does not mean that their physical distance remains unchanged. This dis- tance is evaluated not via the Minkowski metric ηab, but via the proper slightly distorted metric gab = ηab + hab. If a wave with a linear polarization ǫ+(k) runs along the z-axis

1 i(kz−ωt) hab = [ǫµ(1)ǫb(1) ǫµ(2)ǫb(2)] a+e + c.c. (5.212) √2 − and hits two particles at (0,d0/2, 0, 0) and (0, d0/2, 0, 0), which initially do not move due to the vanishing of duµ/dτ, their spatial− distance changes as follows:

0 b d d2 = (0,d , 0, 0)a(η + h )  0  − 0 ab ab 0    0      2 a+ = d0 1 cos(ωt δ) , (5.213) " − √2 − # where δ is the phase of the wave amplitude a+. We can best picture the change in the metric by imagining a circular necklace of mass points placed in the gravitational beam. If the momentum points into the orthogonal direction of the paper plane, the circle distorts into vertical and horizontal ellipses, as shown in Fig. 5.10. The distortions are of quadrupole character, and the µν area within the necklace remains invariant, due to the tracelessness of ǫ+ (k) (giving g = η + h a determinant √ g = 1+ (a2) leading to an invariant volume ab ab ab − O element d3x′ = √gd3x to lowest order in a). For a wave with polarization tensor ǫ (k ) = [ǫ (1)ǫ (2)+(a b)]/√2, the distance changes with time as follows: × zab µ b ↔

2 2 a d× = d0 1 sin(2φ) cos(ωt δ) . (5.214) " − √2 − # 5.5 Detection of Gravitational Waves 421

Figure 5.10 Distortions of a circular array of mass points caused by the passage of a µν gravitational quadrupole wave. The left is caused by a polarization tensor ǫ+ , the second µν by ǫ× .

The situation is the same as before, except for a rotation φ φ π/4. Thus the necklace undergoes the same quadruple distortions, except→ that− the principal axes of the ellipses lie along the diagonals as indicated by the subscript of the × polarization tensor. The rotation by 450 can also be displayed more directly by µν writing the polarization tensor ǫH× of Eq. (4.426) as a combination:

1 1 ǫµν = [ǫµ(1)ǫν(2)+(a b)] = [ǫµ( )ǫν ( ) ǫµ( )ǫν( )] , (5.215) H× √2 ↔ √2 ր ր − տ տ where 1 1 ǫµ( )= [ǫµ(1) + ǫµ(2)], ǫµ( )= [ǫµ(1) ǫµ(2)] (5.216) ր √2 տ √2 −

µν are the diagonal polarization vectors. Similarly we rewrite ǫH+(kz) of (4.425) as

µν 1 1 µ ν µ ν ǫ (kz)= [ǫµ(1)ǫb(1) ǫµ(2)ǫb(2)] = [ǫ ( )ǫ ( )+ ǫ ( )ǫ ( )] . (5.217) H+ √2 − √2 ր տ տ ր

µν µν The right-hand sides of (5.215) and (5.217) have the same forms as ǫH+ and ǫH× expressed in terms of ǫµ(1) and ǫb(2), but with ǫµ(1) and ǫb(2) exchanged by the 450-rotated diagonal polarization vectors ǫµ( ) and ǫµ( ). ր տ The acceleration of the distance

2 d d 1 a 2 cos2φ = d0 ω cos(ωt δ) (5.218) dt2 d √2 ( sin 2φ ) − implies the presence of tidal forces acting upon the necklace. Their field lines are shown in Fig. 5.11. If gravitational waves of helicity 2 hit the necklace, it is ± 422 5 Classical Radiation

Figure 5.11 Field lines of tidal forces of a gravitational wave runing along the z-direction ab ab with polarization tensor ǫ+ and ǫ× , respectively. The field lines change direction with a time dependence cos ωt. deformed into an ellipse with a fixed shape which rotates clockwise or counter- clockwise around the direction of the wave. The wave is circularly polarized. This is seen by taking again two particles at positions d d 0, cos φ, sin φ, 0 , (5.219) ± 2 2 ! and measuring their distances with the metric

i(kz−ωt) gµν = ηµν + ǫµ(+)ǫν (+)e + c.c. . (5.220) h i As a function of time t, the square distance of the two particles behaves like d2 = d 2 1 (cos2 φ sin2 φ)+ i2 cos φ sin φ ae−iωt + c.c. 0 − − = d 2 [1n ha cos(ωt 2φ δ)] , i o (5.221) 0 −| | − − implying the above-described rotations of the necklace, the azimuthal angle φ acting merely as a phase shift. The time-dependence of the length measured by the metric gµν = ηµν + hµν has a direct experimental equivalence. If we take a piezoelectric crystal, then it shows a pulsating voltage due to the distance changes between the atoms, even though its atoms remain at rest in the Minkowski coordinates. The distance is given by the minima of the interatomic potentials. Since the atomic interactions are due to electromagnetism which spreads through a space with the metric gµν, the distances of these minima change according to changes in gµν, and this gives rise to the piezoelectric voltage. How large are the distortions caused by a gravitational wave? If we assume a typical astrophysical source (for the emission mechanism see the next section) with an energy flux of 1010 erg/(cm2sec) at ω 104/sec, we calculate the distortion of the metric to be of≈ the order ∼ h 10−7. (5.222) µν ≈ 5.6 Inspiralling Plunge of One Black Hole into another 423

By analyzing seismometer data of the Earth’s vibrations, J. Weber estimated in 1967 [10] an upper limit for the flux of gravitational waves:

energy flux erg < 3 107 , (5.223) frequency × cm2 sec Hertz at a frequency 3.1 10−4 Hz. This would make the Earth-Moon distance oscillate by 10−7cm = 10×A˚ around the average distance 3.8 1010cm. This amplitude is ≈ × of the same order as the distance between the atoms in matter. Other possible observable effects are quadrupole vibrations which can be excited by incoming gravitational waves within the Earth or the Moon themselves. Their natural frequencies are 54 minutes or 15 minutes, respectively. In 1972, J. Weber built a gravitational detector consisting of a cylindrical aluminum block of 1.53 m length and 0.66 m diameter (weight 1.41 106g). The block has an eigenfrequency of 1.66 Hz. By a piezoelectric strain≈ transducer,× Weber measured length changes in the material of the block. Setting up one block at the University of Maryland and another one at Argonne National Laboratory and looking at coincidences be- tween the two, he eliminated random vibrations caused by the daily activities in the neighborhood of each detector [11]. In 1972, he observed two sudden simultaneous excitations which he interpreted as a possible response to a gravitational wave passing through. Unfortunately, his observation has not found any recurrence in spite of collective efforts of several laboratories [12]. A new approach was necessary to become sensitive to such small effects. This approach has finally led to the so-far undisputed discovery of the waves [13]. More details will be discussed in the forthcoming Section.

5.6 Inspiralling Plunge of One Black Hole into another

An extremely dramatic event which can release gigantic gravitational energies is the merger of a black hole with another in an inspiralling plunge. Such an event has been studied by many theorists using computer simulations of Einstein’s equation [18]. Analytically, it has been possible to develop an effective one-body approach to general-relativistic two-body problems that has made it possible to calculate the plunge of one body into a larger one as a powers series in the inverse distance [19]. If translated into an acoustical signal, the accelerating motion with the final merger should be hearable as a short chirp [20]. Such signals have recently been detected by the experimental collaboration LIGO (see the heading of Fig. 5.3), where more that 1000 researchers were linked via internet. The chirp was heard simultaneously in Hanford, Washington and in Livingston, Louisiana, and looks exactly [13] as expected from the theoretical computer simulations [18]. When the black holes are still far apart from each other, the signal has the general form calculated for two mass point in Fig. 31.15. As the inspiralling of the two bodies proceeds, the frequency increases, until one black hole merges with the other. 424 5 Classical Radiation

Figure 5.12 Two correlated chirps detected by the LIGO collaboration in Hanford, Washington and Livingston, Louisiana.

Appendix 5A Attractive Gravity versus Repulsive Electric Forces between Like Charges

The energy of gravitational waves gives a simple insight into why the fields lead to an attraction be- tween masses (which are always positive) while electromagnetism is repulsive between line charges. The physical components of gravitational waves are the purely spatial ones φij . Their energy has to be positive, and hence the field energy carries a plus sign when expressed in momentum space:

E k2(φij )2. (5A.1) grav ∝ In electromagnetism, the same is true for the spatial part of the electromagnetic field Ai:

2 E k2(Ai) . (5A.2) elm ∝ As a simple consequence of Lorentz invariance, the components φ00 and A0 have to appear with opposite signs in E:

2 002 Egrav k φ , ∝ 2 E k2A0 . (5A.3) em ∝ − But these components are the relevant ones coupling to the mass density T 00 or to the charge density j0, respectively, thereby giving rise to Newton’s or to Coulomb’s law. The opposite signs cause the forces to point in opposite directions. Appendix 5B Nonlinear Gravitational Waves

Plane gravitational waves are such simple phenomena that they can easily be found as solutions to the full nonlinear Einstein equations in the vacuum

Gµν =0. (5B.1)

If the wave runs in the z-direction we can make a two-dimensional ansatz for the invariant distance

(ds)2 = (dt)2 L2[e2β(dx)2 + e−2β(dy)2] (dz)2, (5B.2) − − Appendix 5B Nonlinear Gravitational Waves 425 where L and β are functions which depend only on z and t. It is useful to go to so-called light cone coordinates:

u = t z, − v = t + z, (5B.3) which sit on wave crests moving along the positive and negative z-directions. It can easily be verified that the only non-zero component of the Ricci tensor is

R = 2L−1(L′′ + β′2L), (5B.4) uu − so that the nonlinear wave equation in empty space reads

L′′ + β′2L =0. (5B.5)

Our previous linear waves correspond to the equation

L′′ 0, (5B.6) ≈ which can be solved by

L 1. (5B.7) ≈ In this limit, the square of the line elements are equal to

ds2 (1+2β)dx2 (1 2β)dy2 dz2, (5B.8) ≈− − − − corresponding to a small metric distortion

00 0 0 2β 0 h = , (5B.9) ab  2β −0 0   0 00      i.e., a wave with a polarization tensor ǫ+(kz)ab. Going back to the nonlinear case, we assume the space to be flat before the wave arrives, say at

t z = T, (5B.10) − − i.e., we take

β =0,L =1 for u< T. (5B.11) − We shall assume that the wave comes as a pulse of width 2T and that, after the pulse has passed, the final spacetime is left again flat. While the pulse passes, i.e., for T >u> T , we allow for an arbitrary β(u) but assume, for simplicity, that the pulse is not sharp: − 1 β′(u) . (5B.12) | | ≪ T Then we do not have to solve the full differential equation. While the pulse is passing, we have 1 β(u) = arbitrary, with β′ , (5B.13) | | ≪ T ′ u u L(u) = 1 du′ du′′[β′(u′′)]2 + ((β′T )4) for u ( T,T ). − O ∈ − Z−T Z−T 426 5 Classical Radiation

After it has passed, β(u) is again zero by assumption, and

u 1 L(u)=1 , with a + ((β′T )2) for u > T. (5B.14) − a ≡ T ′2 O −T β du

The change in L from 1 to u/a is physicallyR not observable. The space is flat after the pulse has passed. This is seen by going− to new coordinates X, Y and U, V defined by

X Y 1 X2 + Y 2 x = , y = and u = U, v = V + , (5B.15) 1 U/a 1 U/a a 1 U/a − − − which transforms the invariant distance to the Minkowski form:

ds2 = dUdV d2 (dY )2. (5B.16) − − Appendix 5C Nonexistence of Gravitational Waves in D=3 and D=2 Spacetime Dimensions

The counting procedure of physical degrees of freedom has an immediate consequence for the existence of gravitational waves in a hypothetical lower-dimensional world. In three dimensions, the symmetric tensor φµν has six independent components, three of which are eliminated by the µν Hilbert condition ∂µφ = 0. This leaves only three components. These, however, are just as many as there are gauge degrees of freedom:

φµν φµν + ∂µξν + ∂ν ξµ ηµν ∂ ξλ. (5C.1) → − λ In fact, all field degrees of freedom are gauge degrees of freedom. We see this by choosing k in the z-direction, where polarization vectors can be written as 1 1 1 ǫµ(l) = kµ = (1, 0, 1), √2 k √2 | | 1 ǫµ(l′) = (1, 0, 1), √2 − ǫµ(1) = (0, 1, 0). (5C.2)

They allow us to form the three symmetric polarization tensors which satisfy the Hilbert condition:

ǫµ(l)ǫν (l) ǫµ(1)ǫν (l) + (a b) (5C.3) ↔ ǫµ(1)ǫν (1).

The first is proportional to kµkν , i.e., to a pure gauge of the form (5C.1) with ξµ = kµ. The second is a pure gauge with ξµ = ǫµ(1). The third, finally, is equal to

ηµν [ǫµ(l)ǫν (l′)+ ǫµ(l′)ǫν (l)], (5C.4) − which has the pure gauge form (5C.1) with ξµ = ǫµ(l′)/√2 k . | | This implies that the three-dimensional gravitational field has no dynamical degrees of freedom. There exist no freely propagating gravitational waves in three spacetime dimensions. A three-dimensional Einstein theory would have a further disease. It would not even possess a Newtonian weak-field limit. To see this, we suppose to have found a field φµν from the equation

m ∂2φµν = 2κ T µν . (5C.5) − Appendix 5C Nonexistence of Gravitational Waves in D =3 and D =2 427

From the solution we calculate in D-dimensions 1 hµν = φµν ηµν φ. (5C.6) − D 2 − In order to guarantee Newton’s equation of motion in the weak-field limit, it is necessary that a massive point particle at the origin produces a field h00 satisfying the Poisson differential equation

∂2h00(x)= GMδ(3)(x). (5C.7)

Moreover, T µν has only a T 00-component appearing in (5C.7). Hence φµν has only a single non- vanishing component φ00. In three dimensions this implies that 1 D 3 h00 = φ00 η00φ = − φ00. (5C.8) − D 2 D 2 − − But φ00 must satisfy (5C.5)! Hence the Newton potential vanishes identically. Another equivalent place where this disease manifests itself is in the coupling of the gravita- tional field φµν to the energy-momentum tensor of a massive particle. Since this coupling is

µν H Tµν , (5C.9) we see that η η hµν T = φµν ab φ T = φµν T µν T λ . (5C.10) µν − D 2 µν µν − D 2 λ  −   −  With the particle being at rest and Tab having only a 00-component, the interaction becomes η φ00 T 00 T , (5C.11) 00 − D 2 00  −  which vanishes for D = 3. Only by a technical trick can we obtain a nonzero limit. The theory has to be defined for continuous dimensions D in the neighborhood of 3 with a coupling constant which diverges at D 3: − κ = κ /(D 3). 0 − The above-described diseases are not a consequence of the linear approximation to gravity. We have mentioned before that in three dimensions, the full curvature tensor Rµνλκ is completely determined in terms of the Ricci tensor 1 R = G g G, (5C.12) µν µν − D 2 µν − m and thus in terms of the Einstein tensor Gµν = κTµν . Due to Einstein’s equation Gµν = κ T µν , the Ricci tensor Rµν vanishes identically everywhere, except right at the mass point. This implies, in particular, that the empty-space field equation

Gµν =0, (5C.13) which we have used in four dimensions to find the Schwarzschild metric, can have only the trivial 4 solution gµν = ηµν , up to an irrelevant reparametrization of spacetime. It is curious to note that this disease makes it possible to develop a quantum theory of gravity, in contrast to four dimensions where such a theory does not yet exist. Thus it is possible to define

4For a detailed discussion see S. Giddings, J. Abbott, and K. Kuchar, Rel. Grav. 16, 751 (1986); see also S. Deser, R. Jackson, and S. Templeton, Am. Phys. (NY) 140, 372 (1982). 428 5 Classical Radiation a wave function for the universe. In the absence of matter, the entire Hilbert space consists only of one state, the vacuum 0 .5 Let us take a look at gravitational| i waves in two dimensions. There the only polarization tensor satisfying Hilbert’s constraint is 1 ǫµν (k)= kµkν , (5C.14) k 2 | | and this is obviously a pure gauge. As far as the equation (5C.6) is concerned, the situation is µν even worse than in three dimensions. It is impossible to recover from φ the metric gµν since the equation η h = φ µν φ (5C.15) µν µν − D 2 − is meaningless for D = 2. In order to see the origin of the problem let us choose a gauge different from Hilbert’s, and use the gauge freedom

h h +2∂ ξ , 00 → 00 0 0 h h + ∂ ξ + ∂ ξ , (5C.16) 01 → 01 0 1 1 0 to make h00 and h01 vanish. For the remaining component h11 h we find from (4.371) the Lagrangian density: ≡ f 1 2 2 2 2 = (∂h) 2(∂1h) + 2(∂1h) (∂h) =0, (5C.17) L 4 − − which vanishes identically. In fact, this property could have been an ticipated. It is well known in differential geometry that, in two dimensions, the Einstein action

f 1 = d2ξ√gR (5C.18) A 2κ Z is a pure surface term. By the Gauss-Bonnet theorem, it is entirely determined by the global topological properties of the space. For closed surfaces, it is equal to (4π/2κ)(1 h), where h is the number of handles of the surface. This makes it impossible to derive equations− of motion from such an action. For completeness, let us compare the situation with the electromagnetism case. In three dimensions, the vector potential has three components minus one, due to the Lorentz condition. This leaves two degrees of freedom. One of them is a pure gauge mode, the other is physical. Hence there exists a freely propagating photon in three dimensions. In two dimensions, the Lorenz gauge allows Aµ to be written in the form of a two-dimensional curl:

µ µν A = ǫ ∂ν φ. (5C.19)

In empty space, it satisfies the free field equation

∂2Aµ =0. (5C.20)

In terms of the field φ, the field strength is

F 01 = ∂0A1 ∂1A0 = ∂2φ. (5C.21) − − Note the relation with the theory of complex functions. For a zero field, φ is a harmonic function. It can therefore be considered as the real part of an analytic function

φ(x)=Re f(z), (5C.22)

5See H. Leutwyler, Phys. Rev. B 134, 1755 (1964). Appendix 5C Nonexistence of Gravitational Waves in D =3 and D =2 429 where

z = x + iy. (5C.23)

The vector Aµ is the real part of the gradient of this function

µ µν A = ǫ ∂ν Re f(z). (5C.24)

Now, the real and imaginary part of f satisfy the Cauchy-Riemann differential equations

∂1Re f = ∂2Im f, ∂ Im f = ∂ Re f. (5C.25) 1 − 2 Hence we can also write

Aµ = ∂µIm f. (5C.26)

This shows that Aµ is a pure gauge which, therefore, carries no electromagnetic field, in agreement with the assumption we started out from. It is important to point out that the nonexistence of propagating electromagnetic waves does not rule out Coulomb forces. They do not need the exchange of physical electromagnetic waves. The exchange of virtual electromagnetic fields with zero momentum is sufficient. The inhomogeneous equation 1 ∂2Aµ = jµ (5C.27) − c has, outside the charge distribution, only the solution ∂2Aµ = 0, and we have seen before that the only field satisfying this is a pure gauge field. Still, the Maxwell equation 1 ∂ F 01 = j0 (5C.28) 1 2 allows a nonvanishing constant field:

F 01 = const. (5C.29)

It is carried by the k = 0-components of Aµ to which the previous polarization discussions do not apply. It is this component which gives rise to the Coulomb force in two spacetime dimensions. Another way to see this is by considering the Maxwell action

1 1 = d2x F F µν + d2x jµA , (5C.30) A 2c ab c2 µ Z Z and by going to the Lorenz gauge,

µ µν A = ǫ ∂ν φ, (5C.31) in which it becomes (up to surface terms)

1 1 = d2x(∂2φ)2 + d2xǫµν (∂ j )φ. (5C.32) A 2c c2 µ b Z Z Extremizing this in the field φ gives the field equation 1 (∂2)2φ = ǫµν ∂ j . (5C.33) c µ ν 430 5 Classical Radiation

Reinserting this into the action, the extremum is found to be

′ ′ 1 2 µν 1 µ ν = d xǫ ∂ j ǫ ∂ ′ j ′ . (5C.34) Aextr c2 µ ν (∂2)2 µ ν Z Using

′ ′ ′ ′ ′ ′ ǫµν ǫµ ν = ηµµ ηνν ηµν ηνµ , (5C.35) − µ performing a partial integration, and taking advantage of the current conservation law ∂µj = 0, this becomes 1 1 = d2x j jµ, (5C.36) Aextr c2 µ ∂2 Z which is precisely the Biot-Savart interaction law between currents. Using once more current conservation in the form

2 2 2 1 d2x(j0 j1 ) = d2x j0 ∂ j1 ∂ j1 − − 1 ( ∂ )2 1 Z Z  − 1  1 1 = d2x ∂ j0 ∂ j0 ∂ j0 ∂ j0 (5C.37) 1 ( ∂ )2 1 − 0 ( ∂ )2 0 Z  − 1 − 1  ∂2 = d2xj0 j0, ( ∂ )2 Z − 1 we can rewrite (5C.36) as

1 1 = d2xj0 j0. (5C.38) Aextr 2c2 ( ∂ )2 Z − 1

This is an instantaneous linear potential between the charges carried by j0. It is due to the constant electric field F µ allowed in empty space. In this respect, the situation is quite different from the gravitational case in three dimensions. There the absence of Rµν outside a mass distribution implies a vanishing of the gauge invariant curvature tensor Rµνλκ and hence the vanishing of all tidal forces.

Appendix 5D Precession of Gyroscope in a Satellite Orbit

In a comoving frame, the tidal forces cause a precession of a rotating body proportional to its quadrupole moment Qkl. A rotating body without a quadrupole moment does not precess in free fall. However, if we observe such a gyroscope from another frame of reference, for example from a distance observer where the metric is asymptotically flat, the spin vector nµ does show precession.

Geodetic Precession In order to be specific, consider a gyroscope in a circular satellite orbit around the earth and let ni be the unit vector along which its spin points. Its equation of motion is most simple in a comoving frame of reference x′µ since there τ = t and dx′µ/dt = c(1, 0, 0, 0). In addition, the spin has only spatial components so that the convenient equation of motion Dnµ/dτ = 0 reduces to

dn′i = Γ¯ i ′ cn′j . (5D.1) dt′ − 0j It is possible to use this equation and calculate from it the precession rate of n′i with respect to the distant star. For this we go into a comoving frame which maintains a fixed orientation with Appendix 5D Precession of Gyroscope in a Satellite Orbit 431 respect to the distant stars. This is not a falling frame, so that its Christoffel symbols do not vanish, leading to a non-zero precession rate with respect to the distant stars. ′ Let us focus attention upon a fixed time t0 where the orbit is parallel to the x-axis in a coordinate frame anchored to the fixed star. Assume the earth to lie below the gyroscope on the z-axis, at a distance r. The gyroscope will experience an acceleration GM a = (5D.2) − r2 along the z-axis. Let the velocity with respect to the fixed star be v = vex. It is easy to calculate, ′ for the instant t0, the coordinate transformation to the comoving frame of reference. The motion is stopped by going to the new Lorentz frame with coordinates v x0 = x′0 + x′1, c v x1 = x′1 + x′0, c x2 = x′2, x3 = x′3. (5D.3)

We have assumed v c, for simplicity. The acceleration along the negative z-direction is removed by an additional transformation≪

x1 = x′1, x2 = x′2, a x3 = x′3 + (x′0)2. (5D.4) 2c2

′ ′ ′ It is useful to define the comoving frame x ,z such that the transformed metric g µν is Minkowskian. This is achieved by the following transformation of time: v a x0 = x′0 + x′1 + x′0x′3. (5D.5) c c2 The total transformation has the matrix form

′ 1+ ax 3/c2 v/c 0 ax′0/c2 a ∂x v/c 10 0 a =   = α b. (5D.6) ∂x′b 0 01 0  ax′0/c2 00 1      When subjecting to this transformation the almost Minkowskian metric gab = ηab + hab around the earth, we obtain

∂xc ∂xd g′ = g ab ∂x′a ∂x′b cd c d c d = α aα bηcd + α bα bhcd. (5D.7)

Neglecting quantities of the order a2 and aGM/r, this becomes

4v GM 0 0 0 − c c2r  4v GM v a  0 0 x′0 g′ = η + h + 2 2 . (5D.8) ab ab ab  − c c r c c   0 000     v a ′0   0 x 0 0   c c2    432 5 Classical Radiation

Adding the weak gravitational field of the earth, the metric reads GM 4v GM 1 2 0 0 − c2r − c c2r  4v GM 2GM v a  1 0 x′0 ′  − c c2r − − c2r c c2  gab =   . (5D.9)  2GM   0 0 1 0   2   − − c r   v a ′0 2GM   0 x 0 1   c c2 − − c2r    From this we calculate the Christoffel symbols v GM 1 v a Γ¯ 1′ = Γ¯ 3′ = 2 . (5D.10) 03 − 01 − c c2r2 − 2 c c2 All other components vanish. The second term has a purely kinematic origin related to the Thomas precession. Inserting (5D.2) we see that it removes 1/4 of the first term caused by the gravitational field, 3 v GM Γ¯ 1′ = Γ¯ 3′ = . (5D.11) 03 − 01 −2 c c2r2 Going back to (5D.1) we find that the rate of change of the spin direction in the comoving frame is given by dn′ 1 3 v GM = n′ 3, dt′ 2 c cr2 dn′ 2 = 0, (5D.12) dt′

y

S x

a

Figure 5.13 Gyroscope carrying a frame x′, z′ around the equator with a fixed orientation with respect to the fixed stars. In an elaborate experiment “Gravity Probe B” launched on April 20, 2004 from Vandenburg air base, gyroscopes were launched into polar orbits whose axes point towards IM Pegasi. If Einstein is correct the axis of rotation of the gyroscope should move by an angle which is as small as the diameter of a hair seen from 1/4 mile away [14]. Appendix 5D Precession of Gyroscope in a Satellite Orbit 433

dn′ 3 3 v GM = n′ 1. dt′ −2 c cr2 In vector notation, this reads ′ dn 3 GM ′ ′

= x v n ª n , (5D.13) dt′ 2 c2r3 × × ≡ G ×   where ªG is the angular velocity vector of the geodetic precession. The result can also be expressed in terms of the gravitational acceleration vector v˙ as follows dn′ 3 = (v v˙ ) n′. (5D.14) dt′ 2c2 × × In this form, the result was first obtained by de Sitter in 1916 [15]. Equation (5D.14) may also be interpreted in another way. Since the spin directions can be used to define the orientation of freely falling frames of reference, Eq. (5D.14) tells us how such a frame of reference is rotating with respect to the distant stars. Just as in atomic physics, we may interprete this result as being the consequence of a gravita- tional spin-orbit interaction energy 2GM 2 1 ∂U U = L S = , (5D.15) mc2r3 · mc2 r ∂r which must be corrected by the Thomas precession by replacing the number 2 by 2 1/2=3/2. Note that the existence of such a spin-orbit interaction causes spinning particles− to show an apparent violation of the equivalence principle: spinning electrons, protons, etc., do not move along geodesics. For a gyroscope 500 miles above the earth in a polar orbit with a spin orthogonal to the angular momentum (i.e., in the plane of the orbit), the rate of precession is maximal and has the value 3 GM GM 6.9′′ 2 2 . (5D.16) 2 c r r r ≈ year If the spin is parallel to the angular momentum, there is no precession.

Lense-Thirring or Frame-Dragging Precession These results are only true if we neglect the rotation of the earth around its axis. After the preparatory work in the last section it is easy to take also this effect into account. According to Eq. (5D.9), the gravitational field of the rotating earth has the additional matrix elements: G G h0i =2 (x S )i =2 ǫijkxj Sk , (5D.17) r3 × E r3 E where SE is the spin angular momentum of the earth. It contributes to the Christoffel symbol a term 1 xm Γ¯ i′ = (∂ k ∂ k )= Gǫ ǫ ∂ ǫnms Ss , (5D.18) 0j 2 j 0i − i 0j − ijk kln l r3 E which leads to the following contribution to dn′i/dt′: G 3xl Γ¯ i′n′j = ǫ Sl (S x) n′j . (5D.19) − 0j ijl r3 E − r2 E ·   Writing this as ∆Ω n′, we read off the additional precession rate × G 3x ∆Ω = S (x S ) . (5D.20) −r3 E − r2 · E   This is the frame-dragging or Lense-Thirring effect [16]. For a gyroscope in polar orbit with the spin axis parallel to the angular momentum, this gives the only contribution to the precession rate of 0.05′′ per year. For nonpolar orbits it is negligible compared with the much larger geodetic precession. 434 5 Classical Radiation

Notes and References

The classical radiation emanating from accelerating charged objects is discussed in many text- books, for example in Refs. [1] and [3], and in R.P. Feynman, F.B. Moringo, W.G. Wagner, Feynman Lectures on Gravitation, Addison-Wesley, Reading, Massachusetts, 1995; S. Weinberg, Gravitation and Cosmology, J. Wiley and Sons, New York, 1972; C.W. Misner, K.S. Thorne, and J.A. Wheeler, Gravitation, Freeman, San Francisco, 1973; C. Itzykson and J.-B. Zuber, Quantum Field Theory, McGraw-Hill, New York, 1985; M.H.P. van Putten, Gravitational Radiation, Luminous Black Holes and Gamma-Ray Burst Su- pernovae, Cambridge University Press, 2006; C.D. Ott, Classical and Quantum Gravity, 26, 06200 (2009); J.B. Hartle, Gravity: An Introduction to Einsteins General Relativity, Hartle. 2003; T.P. Cheng, Relativity, Gravitation and Cosmology, Oxford University Press, Oxford, U.K., Sec. ed., 2010. The individual citations refer to: [1] J.D. Jackson, Classical Electrodynamics, Wiley and Sons, New York, 1967. [2] A.L. Fetter, Phys. Rev. 152, 183. (1966); H. Kleinert, Gauge Fields in Condensed Matter, Vol. I World Scientific, Singapore, 1989. [3] A thorough discussion of various radiating systems can be found in the textbooks H. Ohanian and R. Ruffini, Gravitation and Spacetime, Cambridge University Press, 2012 (3rd edition) and H. Ohanian, Classical Electrodynamics, Allyn and Bacon, Boston, Mass., 1988. [4] D.H. Douglass and V.B. Braginsky, Gravitational-radiation Experiments, in S.W. Hawking and W. Israel (eds.), General Relativity, Cambridge University Press, Cambridge, 1979. [5] J.M. Weisberg and J.H. Taylor, Relativistic Binary Pulsar b1913+16: Thirty Years of Obser- vations and Analysis, in Proceedings of Aspen Winter Conference on Astrophysics: Binary Radio Pulsars, Aspen, Colorado, 11-17 Jan 2004 (astro-ph/0407149). [6] P.C. Peters and J. Mathews, Phys. Rev. 131, 435 (1963); P.C. Peters, Phys. Rev. 136, 1224 (1964); H.D. Wahlquist, Gen. Rel. Grav. 19, 1101 (1987). [7] M. Burgay, N. D’Amico, A. Possenti, R.N. Manchester, A.G. Lyne, B.C. Joshi, M.A. McLaughlin, M. Kramer, J.M. Sarkissian, F. Camilo, V. Kalogera, C. Kim, D.R. Lorimer, The Highly Relativistic Binary Pulsar PSR J0737-3039A: Discovery and Implications, (astro-ph/0405194), A.G. Lyne, M. Burgay, M. Kramer, A. Possenti, R.N. Manchester, F. Camilo, M.A. McLaughlin, D.R. Lorimer, N. D’Amico, B.C. Joshi, J. Reynolds, and P.C.C. Freire, A Double-Pulsar System - A Rare Laboratory for Relativistic Gravity and Plasma Physics, Science 8, January 2004. For animations see the internet page http://www.jb.man.ac.uk/news/doublepulsar. [8] The figure is taken from M. Davis, R. Ruffini, W.H. Press, and R.H. Price, Phys. Rev. Lett. 27, 1466 (1971). [9] K.S. Thorne, Gravitational Radiation, in S.W. Hawking, and W. Israel, eds., Three Hundred Years of Gravitation, Cambridge University Press, Cambridge, 1987. [10] J. Weber, Phys. Rev. Lett. 22, 1320 (1969). [11] The pioneering Weber bars which were hoped to detect graviatational waves (https://physics.aps.org/story/v16/st19). Notes and References 435

[12] It is emusing the read the historical discussion of Weber’s claim in the internet (https://en.wikipedia.org/wiki/Joseph Weber). [13] B.K. Abbot at al. Phys. Rev. Lett. 116, 061102 (2016). [14] For details see the Wikipedia page http://en.wikipedia.org/wiki/Gravity Probe B. [15] W. de Sitter, Mon. Not. Roy. Astron. Soc. 77, 155 (1915). [16] H. Pfister, General Relativity and Gravitation 39, 1735 (2007). [17] H. Kleinert, Multivalued Fields in Condensed Matter, Electromagnetism, and Gravitation, World Scientific, Singapore 2009, pp. 1–497 (http://klnrt.de/b11). [18] F. Pistorius, Phys. Rev. Lett. 95, 121101 (2005); B. Bruegmann, W. Tichy, and N. Jansen, Phys. Rev. Lett. 92, 211101 (2004); C. Gundlach, J.M. Martin-Garcia, G. Calabrese, and I. Hinder, Classical Quantum Gravity 22, 3767 (2005); M. Campanelli, C.O. Lousto, P. Marronetti, and Y. Zlochower, Phys. Rev. Lett. 96, 111101 (2006). [19] A. Buonanno and T. Damour, Phys. Rev. D 59, 084006 (1999). [20] https://www.youtube.com/watch?v=TWqhUANNFXw.