1 Natural units Almost all units in nature are derived. For instance, in the cgs system of units one chooses the unit of length to be cm, the unit of mass to be g and the unit of time to be s. All other units can be expressed in terms of these three. For instance, the unit of force can be found from Newton’s law to be ML T2 where M, L and T represent the particular units chosen for mass, length and time respectively. Similarly, the unit of charge can be found from Coulomb’s law to be M 1 2 L 3 2 T . We see that sometimes odd-looking fractional powers appear which is probably one reason why people have invented new names for these units. Anyway, the choice of units which are regarded as fundamental is by no means unique. For instance, instead of choosing M, L and T, we can choose as fundamental, the unit of energy E, the unit of velocity V and the unit of “action” A. The last unit is somewhat unusual. It is the unit that the action functional in mechanics or quantum mechanics carries. Since the action is dt (K − V ) we see that the unit is energy·time, the same as the constant ¯h carries. (It would also be possible to view A as the unit of angular momentum.) If we choose the units for velocity such that the speed of light, c = 1 and the unit of action such that ¯h = 1, then we have something that is called “natural units”. It only remains to choose a unit for energy, which we usually will choose to be eV . (1eV = 1.602 · 10−19 J.) To be able to convert between these two systems of units, we express the new units in terms of the old ones. It is not difficult to find that    V = L T E = ML2 T2 A = ML2 T or    M = E V 2 T = A E L = AV E (1) Using these relations we can make a small table Quantity CGS Natural Force ML T2 E2 AV Charge M 1 2 L 3 2 T (AV ) 1 2 Magnetic moment M 1 2 L 5 2 T (AV ) 3 2 E From the table we see that for instance charge is dimensionless in natural units (it is given in eV 0 ). Force has dimensions eV 2 while length, time and magnetic moment all have the same dimension eV −1 . If you are given a 1 number in natural units and you want to change it into CGS units you just have to insert the proper powers of ¯h and c to restore the dimensions. Here is an example: the charge of the electron in natural units is en = 8.543 · 10−2 eV 0 (2) Converting this to CGS we have to consult the table to see that we need to multiply this with √ ¯hc to restore the dimensions. Note here that the units you choose for ¯h and c will also give you the units of the final result. Since we are interested in CGS units we have to give ¯h and c in cm, g and s instead of the more usual m, kg and s. For reference I give them here ¯h = 1.05459 · 10−27 g cm2 s c = 2.9979 · 1010 cm s (3) Multiplying together gives ecgs = en √ ¯hc = 4.803 · 10−10 esu = 3.336 · 10−10 C. (4) Another example is the expression for the Bohr radius in natural units (a0)n = 1 mne2 n eV −1 (5) To transform this to an expression of dimension length we multiply with ¯hc to get (a0)cgs = ¯hc mne2 n (6) The formula still depends on the values of e and m in natural units though. We can convert these also by using the relation between the charges derived above and the relation between the masses as mn = mcgsc2 . This gives (a0)cgs = ¯h2 c2 mcgsc2e2 cgs = ¯h2 mcgse2 cgs (7) Comment 1: The definition and even the unit of charge differs between different systems of units. This comes about because there are different conventions about how to write Coulomb’s law. We have ecgs = ehl √ 4π = eSI √ 4πε0 (8) 2 ehl represents the charge in the Heaviside-Lorentz system of units. It clearly has the same dimension as the CGS charge. However, in the SI system there is an additional constant ε0 which has the dimension of inverse velocity squared. Charge in the SI system therefore has a different dimension than in CGS. Comment 2: There exist other systems of units with only one basic unit. For instance, in relativistic gravitational physics it is often advantageous to choose a system of units where the speed of light c = 1 and Newton’s gravitational constant G = 1. The remaining unit is the unit of length which can be chosen arbitrarily (light-year, m, cm etc.). This system of units is called geometrical units. In this system, for example, the mass of the earth is approximately 0.44 cm. 2 The Dirac equation There is a curious way to “derive” the Schr¨odinger equation. Namely, take the relation for the energy in classical physics E = p2 2m + V. (9) One gets the Schr¨odinger equation by making the replacement E → i¯h ∂ ∂t , pi → −i¯h ∂ ∂xi , (10) and then letting the relation (9) “act” on a wavefunction one gets i¯h ∂ ∂t ψ = − ¯h2 2m ∂2 x + V ψ. (11) This derivation inspired many people to try to derive a relativistic analog of the Schr¨odinger equation by starting with the relativistic energy relation E2 = p2 c2 +m2 c4 (or E2 = p2 +m2 in natural units) instead of starting with (9). Making the same substitution (10) as before we get a relativistic wave equation −∂2 t φ = − ∂2 x + ∂2 y + ∂2 z φ + m2 φ. (12) 3 This can be written in a more relativistic fashion by introducing a metric gµν = diag(1, −1, −1, −1) as gµν ∂µ∂νφ + m2 φ = 0, (13) an equation which is known as the Klein-Gordon equation. To find out more about its properties, we now go on to find solutions to the Klein-Gordon equation. For instance, there is a complete set of planewave solutions as we will now show. First we make the ansatz φ = e−ikµxµ . Acting on this with a four-derivative ∂µ gives us ∂µe−ikν xν = −ikµe−ikν xν . (14) Using this result twice we may insert the ansatz into the Klein-Gordon equation to get ∂µ∂µ φ + m2 φ = (−kµkµ + m2 )φ. (15) We see that for φ to be a solution to the Klein-Gordon equation we need the four momentum kµ to satisfy the relation kµkµ = m2 , (16) and rewriting the four momentum kµ in terms of its components kµ = (E, k) where k is the ordinary three momentum, we recover the relativistic energy relation E2 = k2 +m2 . Let us recapitulate; the Klein-Gordon equation has a complete set of plane wave solutions φ(x) = eik·x where the four momentum has to satisfy the relativistic energy condition k · k = m2 . Any solution can then be written as a linear combination of these plane waves. There is however a funny new feature of these solutions. If the four vector kµ = (E, k) gives a solution, then the four vector kµ = (−E, k) with negative energy is also a solution! Thus, for every solution with positive energy, there is a solution with negative energy which seems physically unacceptable since it would lead to an unstable theory (there would be no state with lowest energy = vacuum state). Dirac identified the root of this problem in the fact that the Klein-Gordon equation is quadratic in the time derivative whereas the Schr¨odinger equation is linear. He tried to get around this by introducing an equation which would be linear in time derivatives. To achieve this he used some interesting properties of the Pauli matrices σ1 = 0 1 1 0 , σ2 = 0 −i i 0 , σ3 = 1 0 0 −1 , (17) 4 which fulfill the relation σi σk = i ikl σl +δik 1. This made it possible for Dirac to write kµkµ = E2 − ki ki = E1 − ki σi E1 + kl σl . (18) That is, by writing the equation in terms of two by two matrices, he was able to split it into factors linear in energy. The price he had to pay was that the wave functions now become two dimension column vectors (or spinors as they are more commonly known). Thus our second attempt for a relativistic wave equation looks like this 1i∂t − σi i∂i 1i∂t + σl i∂l φA = m2 φA, (19) where φA = φ1 φ2 is a two dimensional column vector. By introducing a second two dimensional column vector mφB = 1i∂t + σl i∂l φA, (20) we can write an equation (well, really a system of equations) which is linear in time derivatives mφB = 1i∂t + σl i∂l φA, mφA = 1i∂t − σi i∂i φB. (21) For purely conventional reasons one often redefines the column vectors as φ± = φA ± φB which makes it possible to write the above equation as mφ+ = 1i∂tφ+ + σl i∂lφ− mφ− = −1i∂tφ− − σl i∂lφ+, (22) or, defining a four component column vector ψ = φ+ φ− and four by four matrices γ0 = 1 0 0 −1 ; γi = 0 σi −σi 0 , (23) we may write the resulting equations in a very compact form as γµ i∂µψ = mψ. (24) 5 Notice that this is a matrix equation (it is really four equations written in a very nice and compact form using matrices) and that it is linear in time derivatives which is exactly what Dirac wanted to achieve. This equation is known as the Dirac equation. To make the comparison to the ordinary Schr¨odinger equation more prominent, we can rewrite it as γ0 i∂tψ = −γl i∂lψ + mψ, (25) and using that γ0 γ0 = 1 we find i∂tψ = −γ0 γl i∂lψ + mγ0 ψ. (26) We thus see that the Hamiltonian operator that we get from the Dirac equation is H = −γ0 γl i∂l + mγ0 . Again, to get a feeling for the physics we can try to solve the equation. Since the wavefunction is a four component column vector we make the ansatz for a plane wave ψ = u(p)e−ip·x , (27) where u(p) is a four component column vector possibly dependent on p. Inserting this into the Dirac equation we get (iγµ ∂µ − m)ψ = (γµ pµ − m)u(p)e−ip·x , (28) so we see that for this to be a solution of the Dirac equation we need the four column vector u to satisfy the matrix equation (γµ pµ − m)u(p) = 0. (29) Using the expressions for the gamma matrices found earlier we can rewrite this in an even more explicit form      E − m 0 −p3 −p− 0 E − m −p+ p3 p3 p− −E − m 0 p+ −p3 0 −E − m           u1 u2 u3 u4      = 0, (30) where we have defined the complex combinations p± = p1±ip2. This equation has four independent solutions. We will find one of them, but I recommend that you similarly try to find the other three. Actually, for this equation to 6 be solvable we need the determinant of the matrix to be zero. We can easily evaluate it to be (E2 − p2 − m2 )2 so we see that a necessary condition for this equation to have solutions is that the “old” relativistic energy condition is satisfied. Unfortunately this means that we did not get rid of the solutions with negative energy. Therefore we first need to assume that the condition holds, then we can go on and try to find a solution. To make it a little bit simpler, let us first try it in the case where p = 0. Then the equation looks like      0 0 0 0 0 0 0 0 0 0 −2m 0 0 0 0 −2m           u1 u2 u3 u4      = 0, (31) for the case of positive energy, i.e. when E = +m and in the case where the energy is negative, i.e. when E = −m, it looks like      −2m 0 0 0 0 −2m 0 0 0 0 0 0 0 0 0 0           u1 u2 u3 u4      = 0. (32) In the positive energy case we have the two independent solutions      1 0 0 0      ,      0 1 0 0      , (33) and in the negative energy case the solutions look like      0 0 1 0      ,      0 0 0 1      . (34) Turning on the three momentum p we have to solve the full equations (30) but we can expect that the solutions should not differ too much from the zero p solutions, at least when p is small. Then we should be able to find a 7 solution of the form      1 0 a b      , (35) where a and b are small of order p (or possibly smaller). Inserting this ansatz into the equation immediately gives us that a = p3 E+m and b = p+ E+m . For reasons to be explained later we choose the normalization to be u† u = 2E which leads to the final answer ψ = √ E + m       1 0 p3 E+m p+ E+m       e−ip·x . (36) 3 The non-relativistic limit of the Dirac equa- tion One check that one should always do is to see how the new physics one is investigating reduces in known situations. In the case at hand this means that we should try to see how the physics of the Dirac equation looks in a non-relativistic situation. To do this, let us have a look at it in the form given in (22) but in momentum space. The equation looks like (E − m)φ+ = σl plφ−, (E + m)φ− = σl plφ+. (37) The non-relativistic limit means the limit where p m. This in turn implies that E = √ p2 + m2 = m + p2 2m + . . . = m + E(NR) where E(NR) is the nonrelativistic energy. This immediately tells us that the quantity E(NR) = E−m is small (of order m p m 2 or v2 c2 mc2 in ordinary units) while the quantity E+m is large (of order m). A look at the equations now tells us that φ+ is of order one while φ− is of order p m so it goes to zero in the non-relativistic limit. We can now solve for the “small” component φ− to get an equation for φ+ only since φ+ is what is left in the non-relativistic limit. Solving for φ− gives us φ− = 1 2m + E(NR) p · σφ+, (38) 8 which, when inserted back into the equation gives us E(NR) φ+ = p · σ 1 2m + E(NR) p · σφ+. (39) In the non-relativistic limit m E(NR) so we can expand the denominator to get E(NR) φ+ = p · σ 1 2m (1 − E(NR) 2m + . . .)p · σφ+, (40) and to lowest order we get back the non-relativistic Schr¨odinger equation E(NR) φ+ = p2 2m φ+. (41) This is maybe not a very exciting result but it is gratifying to see that we get the correct non-relativistic limit of our equation. A slightly more interesting result we get if we include a potential from an external electromagnetic field. This is done in a relativistically covariant fashion in the Dirac equation, introducing the relativistic electromagnetic vector potential Aµ = (ϕ, A), by replacing i∂µ → i∂µ − eAµ. This changes the Dirac equation to γµ (i∂µ − eAµ) ψ = mψ, (42) or, if we Fourier transform as γµ (pµ − eAµ) u(p) = mu(p). (43) When rewriting this in terms of the large and small components we get (E − eϕ − m)u+ = σ · (p − eA)u−, (E − eϕ + m)u− = σ · (p − eA)u+, (44) and solving for the small component we get u− = 1 2m + E(NR) − eϕ σ · (p − eA)u+, E(NR) u+|! = eϕ + σ · (p − eA) 1 2m + E(NR) − eϕ σ · (p − eA) u+.(45) 9 Notice that we have to be careful in which order we write things since ϕ and A depend on x and thus do not commute with p. Using the same approximations as before we get an equation for u+ E(NR) u+ = eϕ + 1 2m σ · (p − eA)σ · (p − eA) u+. (46) To evaluate this we again need to use the properties of the Pauli matrices to be able to write σ · (p − eA)σ · (p − eA) = σi σk (p − eA)i(p − eA)k = (δik + i ikl σl )(p − eA)i(p − eA)k = (47) (p − eA) · (p − eA) + iσ · (p − eA) × (p − eA). The cross product can be evaluated as lik (p − eA)i(p − eA)k = 1 2 lik [(p − eA)i, (p − eA)k] = 1 2 lik (−e [pi, Ak] − e [Ai, pk]) = − lik e [pi, Ak] = (48) ie lik ∂iAk = ieBl , (49) which gives us the non-relativistic equation (Pauli equation) E(NR) u+ = (p − eA)2 2m + eϕ − e 2m σ · B u+. (50) This is exactly the Schr¨odinger equation for a non-relativistic spin half particle with an intrinsic magnetic moment µ = e m s where s = σ 2 is the spin operator. This is a very interesting result. We see that only from the requirement that the theory should be relativistically invariant, we find that particles carry an intrinsic magnetic moment. This is not something that we can turn off or change in any way. It is fundamentally built into the theory and comes from the relativistic invariance. Furthermore, it cannot be understood in any classical sense as “something charged going around in circles”. In fact, you can easily verify by yourself that if we have some charged particle moving in a circle of radius R it produces a magnetic moment which is µ = e 2m L and what we get out of our equation is twice this value. We say that the electron has a gyromagnetic ratio of 2. In fact this is not completely true and this value receives quantum corrections which can 10 be computed with great accuracy (moreover, you will in principle be able to do it yourself using what you learn in this course). One can go on and keep higher order corrections to this result. This will result in extra terms in the Hamiltonian. The calculation is slightly more involved since now it will not be justified to neglect φ− any more. Anyway, it is still possible to write a non-relativistic Hamiltonian for a two component spinor. If one puts A = 0 (no magnetic field) the Schr¨odinger equation becomes p2 2m + eϕ − p4 8m3 − eσ · (E × p) 4m2 − e 8m2 · E ψ = E(NR) ψ. (51) The first two terms are the lowest order terms which we have already derived (remember that we put A = 0). The next three terms are higher order corrections. If we for instance apply this Hamiltonian to the hydrogen atom they will give small corrections to the spectrum (known as fine structure). The third term is simply the first non-trivial correction to the non-relativistic energy (from expanding √ p2 + m2 −m). The fourth term is called the Thomas term and it has the interpretation as an interaction between the spin of the electron and the effective magnetic field it sees when moving through the electric field. It can be rewritten as a spin-orbit interaction (proportional to S · L). The last term is known as the Darwin term. It represents an interaction with the charge density that produces the electric field. In the hydrogen atom it gives a shift in energy of the s-states. There is also something called hyperfine structure of the hydrogen spectrum. It comes from the interaction of the magnetic moments of the proton and the electron but is a much smaller effect than is the fine structure. I would like to point out once again that all these terms one gets automatically from the Dirac equation when going to the non-relativistic limit. There are no additional assumptions involved. Quite a nice little equation! 4 Transformation properties of the Dirac equa- tion You are familiar with how covariant and contravariant vectors transform when we change coordinate systems (we also say “when we do Lorentz rotations” or “boost” the coordinate system) in special relativity. The typical 11 contravariant vector is the coordinate vector xµ itself. When we do a Lorentz boost it transforms into x µ = Λµ νxν where, if we for instance boost to a coordinate system which is moving with speed v in the x direction we have Λµ ν =       1√ 1−v2 − v√ 1−v2 0 0 − v√ 1−v2 1√ 1−v2 0 0 0 0 1 0 0 0 0 1       (52) We may define the matrix Λ ν µ = gµρΛρ σgσν and we can check that Λρ µΛ ν ρ = δν µ. All covariant quantities (for example the momentum vector pµ or a vector field Aµ or the ordinary derivative operator ∂µ transform as Aµ = Λ ν µ Aν. Therefore the scalar product is invariant x µ pµ = xµ pµ. Using this information it is easy to see that for a scalar field φ (a scalar field is defined by the property that it does not transform at all under Lorentz transformations) the Klein-Gordon equation is invariant under Lorentz transformations ∂µ∂µ φ + m2 φ = 0 (53) A spinor is not invariant under Lorentz transformations but transforms as ψa = Sabψb for some matrix S which we will not need the exact form of. The Dirac equation itself transforms as i/∂ψ − mψ = 0 → iγµ Λ σ µ ∂σ (Sψ) − mSψ = 0 (54) or iS−1 γµ SΛ σ µ ∂σψ − mψ = 0 (55) We see that for the Dirac equation to be invariant we need that S−1 γµ S = Λµ σγσ (56) Taking the hermitian conjugate of this equation and using that we know from the explicit representation of the gamma matrices that (γµ )† = γ0 γµ γ0 we get γ0 S† γ0 = S−1 (57) Having this formula we may investigate how for instance ψ† transforms under Lorentz transformations. We get ψ † = ψ† γ0 S−1 γ0 (58) 12 So the hermitian conjugate does not transform as the inverse of the original object. However, if we check how ¯ψ ≡ ψ† γ0 transforms we find ¯ψ = ¯ψS−1 (59) which is indeed “nicer” since we can form objects with simple Lorentz transformation properties from it, for instance ¯ψ ψ = ¯ψψ (scalar) ¯ψ γµ ψ = ¯ψS−1 γµ Sψ = Λµ ν ¯ψγν ψ (vector) (60) 5 Field quantization (“second quantization”) To be able to describe quantum systems where the number of particles is able to change (for instance, an electron and a positron annihilates into two photons) we use a formalism called “second quantization”. Notice that the name second quantization is rather badly chosen since it is not a question about “quantizing again”. It is simply yet an alternative formalism for describing the states we have in the quantum world. It is not only used in relativistic quantum mechanics, but also in for instance solid state physics or anywhere were our quantum system consists of many types of particles which can also change into each other. As a technical detail to simplify computations, let us imagine that our universe is a box with side length L. Then the universe has finite volume V = L3 and if we impose periodical boundary conditions, the allowed momenta form a countable set. In this universe the allowed momenta can be written as k = 2π L (n1, n2, n3) , (61) for any integers n1, n2, n3. In the end of each calculation we may let L → ∞ (if we have done the calculation correctly, nothing should depend on L). Imagining that we have ordered the allowed momenta in some particular way, we may write them as k1, k2, . . . , ki, . . .. This gives us the possibility to write an arbitrary state of the system as |nk1 , nk2 , . . . , nki , . . . , (62) 13 which we interpret as meaning: there are nk1 particles with momenta k1 (that means plane waves), there are nk2 particles with momenta k2 and so on. These states in fact form a complete basis, so any state can be written as a linear combination of these basis states. To be able to write down how operators act on these states we consider the “basic” operators ak and a† k satisfying the commutation relations ap, a† k = δp,k. That is, if p and k are different then ap and a† k commute but if they are the same they satisfy the usual harmonic oscillator algebra. Then, remembering the harmonic oscillator, we have that aki |nk1 , nk2 , . . . , nki , . . . , = √ nki |nk1 , nk2 , . . . , nki − 1, . . . , a† ki |nk1 , nk2 , . . . , nki , . . . , = nki + 1|nk1 , nk2 , . . . , nki + 1, . . . . (63) Thus, the a, a† operators describe the basic operations which take us between different states. For instance, a process where a particle with momentum k1 is scattered into a particle with momentum k2 would be accomplished by the operator a† k2 ak1 acting on the state |1k1 , 0, . . . . Explicitly, using (63), we would have a† k2 ak1 |1k1 , 0, . . . = a† k2 |0, 0, . . . = |0, 1k2 , . . . . (64) As in the case of the harmonic oscillator, any state in the basis can be constructed by acting with the a† operators on the vacuum |nk1 , nk2 , . . . , nki , . . . = a† k1 nk1 nk1 ! a† k2 nk2 nk2 ! . . . a† ki nki nki ! . . . |0 . (65) Let us now look at the coordinate representation of these states. Since we know the total number of particles in each state we know how many coordinates we need, i.e. one particle has three coordinates x, two particles have six coordinates x1, x2 etc. Therefore we have x|1k = φk(x) x1, x2|1k1 , 1k2 = φk1 (x1)φk2 (x2), (66) where we have denoted the coordinate representation of the state with momentum k as φk = 1√ V eik·x . The factor 1√ V is a normalization factor. 14 Now consider the operator φ(x) = k akφk(x). (67) When its hermitian conjugate acts on the vacuum, it creates a state φ† (x0)|0 = k 1 √ V e−ik·x0 a† k|0 = k |1k 1 √ V e−ik·x0 . (68) In the coordinate representation, the resulting state looks like x|φ† (x0)|0 = k x|1k 1 √ V e−ik·x0 = 1 V k eik·(x−x0) . (69) The final expression may seem a little bit strange but it is really a delta function. If x = x0 the exponential oscillates for each k and on average it is zero. For x = x0 however, all exponentials are 1, independently of k, so the sum diverges. The integral over x of this function gives zero for all terms with k = 0 and V for the k = 0 term. The factor 1 V ensures that the final result of the integration is 1. Thus we see that the operator φ† (x0) creates a wavefunction which in coordinate representation is a delta function located at x0 or in other words, the operator φ† (x0) creates a particle completely localized at x0 complementary to a† k which creates a particle with fixed momentum k. In the same way, the operator φ(x0) annihilates a particle located at x0. Now consider any operator on the full system that can be thought of as being composed of operations on single particles. In formulas we would write O = a Oa, (70) where the sum is over all particles in the system. This is a very general expression. Many operators of physical interest are of this type. For instance, the total energy is the sum of the energy of the single particles. The total momentum is the sum of the momenta of the single particles etc. The operator Oa is a “normal” one particle operator. Its action on a one particle state can be expanded into a linear combination of one particle states Oxφa(x) = b φb(x)fba, (71) 15 where, as usual fba = d3 x φ∗ b(x)Oxφa(x). (72) Thus the action of each of the one particle operators in (70) can be seen as a reshuffling of the particles. The total number of particles is not changed, but they are moved between different states. If we want to write how the operator (70) acts on the basis (62) we know that since it does not change the total number of particles but rather shuffles them around, it has to be written as a linear combination of the operators a† aab since this operator first annihilates a particle in state b but then immediately creates a particle in state a. Explicitly we write O = ab faba† aab, (73) and you should check that the coefficients fab are really the same as in (72) by for instance check how O acts on one particle states. Thus we may write O = ab d3 x φ∗ b(x)Oxφa(x)a† aab = (74) d3 x a aaφa(x) † Ox b abφb(x) = d3 x φ† (x)Oxφ(x), where we have used the operator φ(x) defined in (67). From this we read off the procedure for writing operators (this works only for operators that can be thought of as being composed of operations on the single particles) in the second quantized formalism. Take the one particle operator (here written Ox) and compute what looks like an expectation value but instead of a wave function we use the operator φ(x). Since we know what the operator φ(x) does we know how to interpret this expression intuitively. First the operator φ annihilates a particle located at x (if there is a particle there, otherwise the result is zero), then the operator Ox computes whatever it should compute (the energy, momentum or something else) and finally the particle is created again by φ† . The integral means that this process is repeated for each point in space and then summed. Notice also that here is the origin of the awkward term “second quantization”. It comes from the fact that the operator φ looks like an arbitrary 16 wavefunction but with the coefficients in the expansion replaced by the annihilation operators aa. Thus it looks like the wave function is “quantized again” which of course is not true and is the source of much confusion. Second quantization is just a formalism, within the framework of ordinary quantum theory, to describe systems with many particles and in particular where the types of particles may change. There is a pleasant surprise incorporated in the second quantized formalism. Since the creation operators a† a commute with themselves, any wavefunction is automatically symmetric with respect to interchange of these particles as should be the case for bosons. This naturally leads to the question what one should do if one would like to instead describe fermions since in that case the wavefunctions should be anti-symmetric with respect to interchange of particles. The natural thing to try is to use operators which do not commute, but anti-commute. That is, operators ba, b† a which satisfy ba, b† b = bab† b + b† bba = δab, {ba, bb} = b† a, b† b = 0. (75) In this case, since the creation operators anti-commute, we get an extra minus sign when we interchange particles b† ab† b = −b† bb† a giving us the required behavior under interchange of particles. Furthermore we see that if we try to put more than one particle in each state |2 = b† b† |0 = 1 2 b† , b† |0 = 0! (76) This means that the Pauli principle is automatically incorporated when we use anti-commuting creation/annihilation operators. 6 Dyson’s method - the interaction picture Since we have found the time dependent plane wave solutions of the Dirac equation, we completely know the time evolution of any state, if there are no interactions (the theory without interactions we call the free theory). Just Fourier expand the full wavefunction at any given time and then let the individual plane waves evolve in time. The problem comes when we consider an interacting theory. Then the Hamiltonian can be written as a sum of two operators H = H0 + HI where H0 is the free Hamiltonian which is 17 responsible for the (almost trivial) time evolution of the free theory (i.e. the plane waves are eigenstates of H0) and HI is the interaction Hamiltonian which does not necessarily commute with H0 or even with itself at different times. This makes the time evolution problem quite involved. However, Dyson has invented a nice little trick which “hides” the (almost trivial) time evolution of the free theory so that we may concentrate on the (slightly more complicated) time evolution given by the interaction Hamiltonian HI. We want to solve the “Schr¨odinger” equation (H0 + HI) |ψ = i ∂ ∂t |ψ . (77) Now define a new state |ψ = e−iH0t |χ . Inserting this we get (H0 + HI) e−iH0t |χ = i ∂ ∂t e−iH0t |χ = e−iH0t i ∂ ∂t + H0 |χ . (78) Multiplying from the left with eiH0t and using that H0 commutes with itself we get eiH0t HIe−iH0t |χ = i ∂ ∂t |χ . (79) If we define a time dependent interaction Hamiltonian HI(t) = eiH0t HIe−iH0t this equation takes a very simple form HI(t)|χ = i ∂ ∂t |χ , (80) i.e. it looks just like the Schr¨odinger equation, but with the Hamiltonian HI(t). The “trivial” time dependence generated by H0 is taken care of by making the operators in HI evolve in time like in the free theory. Notice that since |χ is a solution to the time dependent Schr¨odinger equation it depends on time as in the so called Schr¨odinger representation while the operator HI(t) = eiH0t HIe−iH0t depends on time as an operator in the Heisenberg representation! This is a funny mix of representations known as the interaction representation. Anyway, from our studies of Quantum Mechanics we know how to solve the time dependence of |χ . The solution is given as |χ(t) = U(t, t0)|χ(t0) , (81) 18 where we have introduced the time evolution operator U(t, t0) = T e −i t t0 HI (t)dt , (82) where T represents the time ordering operator. One may worry that it is really the wave function |ψ which is the “correct” wave function which one should use to calculate probability amplitudes but it is not difficult to show using what you know about the time evolution operator (exercise) that the expressions ψf|T e−i (H0+HI )dt |ψi , (83) and χf|T e−i HI (t)dt |χi , (84) are equal. 7 The quantized Dirac field We would now like to write the equivalent of the field operators (67) for the Dirac field. Since we know that electrons are fermions, we know that we should use anti-commuting creation/annihilation operators rather than commuting ones. With this in mind we may immediately write down a candidate for the field operators ψ(x) = 1 √ V k 4 r=1 1 2 |E| bk,ru(r) (k)eik·x . (85) Here r ∈ {1, 2, 3, 4} is an index which runs over the four independent spinor solutions. Since we are using the interaction representation, we now need to make this operator transform in time according to the free theory. That is, ψ(t) = eiH0t ψe−iH0t . If one then uses the formula eA Be−A = B + [A, B] + 1 2! [A, [A, B]] + . . . and the fact that H0, b† k,r = E(k)b† k,r 1 , one finds that 1 This follows from the fact that H0|1k = E(k)|1k and that H0|0 = 0. Alternatively one may explicitly evaluate the (second quantized) Hamiltonian which turns out to be H0 = k,r Eb† k,rbk,r 19 bk,r(t) = bk,re−iEt which tells us that ψ(x, t) = 1 √ V k 2 r=1 1 √ 2E bk,ru(r) (k)e−iE(k)t+ik·x + 4 r=3 1 2 |E| bk,ru(r) (k)e+i|E(k)|t+ik·x   . (86) Notice that we had to separate the positive energy solutions (r = 1, 2) from the negative energy solutions (r = 3, 4) since they will have a time dependence of different type! In fact, the time dependence of the negative energy annihilation operators looks more like the time dependence of a creation operator. Loosely one can reason as follows; if the operator O(t) creates something the final state should have bigger energy than the initial state. Thus we have Ef|O(t)|Ei = Ef|eiH0t Oe−iH0t |Ei = Ef|Oei(Ef −Ei)t |Ei , (87) so we see that a creation operator should have the time dependence e+iEt . If, on the other hand, the operator O would annihilate something, then Ef < Ei and the time dependence going together with annihilation should thus be of the type e−iEt . This gives us a hint on how to treat the problematic negative energy solutions which seem to be inherent in any relativistic theory. Namely we define d† k,1 = −b−k,4, d† k,2 = b−k,3. (88) The annihilation of a negative energy electron is thus reinterpreted as the creation of a new positive energy particle, a positron. The new particle has exactly the same properties as the electron (mass etc) except that the charge is opposite. It is known as the anti-particle of the electron. To go together with this we also redefine the spinors v(1) (k) = −u(4) (−k), v(2) (k) = u(3) (−k). (89) With these new definitions we may rewrite the field operator in the Heisenberg representation as ψ(x, t) = 1 √ V k 2 r=1 1 √ 2E bk,ru(r) (k)e−iE(k)t+ik·x 20 +d† k,rv(r) (k)e+iE(k)t−ik·x . (90) According to what we know about second quantization, this operator will annihilate an electron at point x and create a positron at x. 8 Scattering of electrons in the field of a nucleus (Rutherford/Mott scattering) We will consider the scattering of an electron in the field of a positively charged heavy particle, typically in the field of a nucleus, so we will take the potential to be: φ = A0 = Ze r (91) The incoming state will of course be a state describing the incoming electron with momentum p and spin r. That is, it will be described by the state |i = b† p,r|0 (92) since the b† operator creates a state of one electron with the specified momentum and spin. The outgoing, final state is also a state with only one electron, but since the electron has scattered it has a different momentum k and a (possibly) different spin s. This is given by the state |f = b† k,s|0 (93) The probability amplitude for this process (scattering of an electron with momentum p and spin r to an electron with the momentum k and spin s we get by taking the initial state |i and evolve it with the time evolution operator and finally taking the overlap with the state f|. The probability amplitude is therefore given by the expression M = f|T e−i HI |i (94) and the probability is of course the absolute square of the probability ampli- tude. We know that HI is small (since e is a small number) so we can evaluate the probability amplitude in perturbation theory f|T e−i HI |i = f|i − ie f| d4 xˆ¯ψ/A ˆψ|i + . . . (95) 21 Notice here that the time ordering operator T is trivial since all operators are at the same time. Only in the higher order terms is the T operator important. If k = p which means that scattering has taken place, the first term is zero. Inserting the information we have we can compute M = − i V q1,t1 q2,t2 d4 x Ze2 r 0|bk,s   1 2Eq1 dq1,t1 ¯vq1,t1 e−iq1·x + 1 2Eq1 b† q1,t1 ¯uq1,t1 eiq1·x   γ0   1 2Eq2 d† q2,t2 vq2,t2 eiq2·x + 1 2Eq2 bq2,t2 uq2,t2 e−iq2·x   b† p,r|0 (96) There are in principle four different terms but clearly only terms with the same number of creation and annihilation operators will survive. There are two such terms, the first being the one where we select the positron creation/annihilation operators from the parenthesis. The operators squeezed between the vacuum states in that case are 0|bk,sdq1,t1 d† q2,t2 b† p,r|0 (97) and using the anti-commutation relations we can transform this into δp,kδr,sδq1,q2 δt1,t2 (98) We see that this term does not give anything unless p = k which means that no scattering is taking place. It would represent a process where the electron is just passing by when a positron is created and annihilated out of the vacuum. This clearly have no effect on the scattering process and we will therefore drop this term. The second term is the one with only electron creation/annihilation operators which, writing only the operators, gives us 0|bk,sb† q1,t1 bq2,t2 b† p,r|0 = δk,q1 δs,t1 δq2,pδt2,r (99) This is the type of term we expect. The interpretation is that the incoming electron gets annihilated and there is a new electron (with new momentum 22 and spin) created, i.e. the electron gets “scattered”. We thus get the formula M = − i V q1,t1 q2,t2 d4 x Ze2 r ei(q1−q2)·x 4Eq1 Eq2 ¯uq1,t1 γ0 uq2,t2 δk,q1 δs,t1 δq2,pδt2,r (100) We can use the Kronecker deltas to get rid of the summations M = − i V d4 x Ze2 r ei(k−p)·x 4EpEk ¯uk,sγ0 up,r = −i U(k − p) V 4EpEk (¯uk,sγ0 up,r) T/2 −T/2 dt ei(Ek−Ep)t (101) where we have introduced the 3-dimensional Fourier transform of the coulomb potential U(k) ≡ d3 x Ze2 |x| e−ik·x = 4πZe2 |k|2 (102) When we let the interaction time T go to infinity, the last integral in (101) is just (2π times) a delta function of the energy telling us that the energy is conserved in the scattering process. Now the actual probability, let us call it P, is given by the absolute square of the probability amplitude P = |M|2 = |U|2 4V 2EpEk ¯uk,sγ0 up,r 2 |2πδ(Ek − Ep)|2 (103) Here the last term, the delta function square, may seem a little bit odd, but we can treat it using a trick, namely, we may write it as |2πδ(Ek − Ep)|2 = lim T→∞ 2πδ(Ek − Ep) T/2 −T/2 dt ei(Ek−Ep)t (104) Due to the presence of the delta function, the Ek − Ep in the integrand can be replaced by 0. This means that the integrand can be replaced by 1 and thus the integral is equal to T. We therefore get the result P = lim T→∞ |U|2 4V 2EpEk ¯uk,sγ0 up,r 2 T2πδ(Ek − Ep) (105) 23 which gives us an expression for the probability per unit time w ≡ P T = |U|2 4V 2EpEk ¯uk,sγ0 up,r 2 2πδ(Ek − Ep) (106) Since we cannot experimentally separate scattering into final energy and momenta which are close to each other, we need to sum over all these probabilities to get a total probability for scattering into a state with final energy Ek or into a state with energy close to it. This we do by multiplying the probability with the density of states ρ(Ek) and then integrate over energy wtot = dEk |U|2 4V 2EpEk ¯uk,sγ0 up,r 2 2πδ(Ek − Ep)ρ(Ek) = |U|2 4V 2E2 p ¯uk,sγ0 up,r 2 2πρ(Ep) (107) This expression now depends only on the energy of the incoming electron Ep which we will hereafter denote by just E. Note however that the spinors u still depend on the 3-dimensional final momentum k. Because of energy conservation |k| = |p|, but the direction can still be different. The density of states function ρ(E) we find in the following way. We have assumed that the universe is a (large but finite) box with length, width and hight L. In such a box the allowed 3-momenta are not arbitrary but rather discrete points momentum space. Only momenta with values ki = 2π L ni i = 1, 2, 3, where ni are integers are allowed. This gives a density of states in momentum space as ρ(k) = L3 (2π)3 = V (2π)3 . Since the energy is a function of the 3-momenta we can write ρ(E)dE = ρ(k)d3 k = V (2π)3 k2 d |k| dΩk (108) where we have introduced spherical coordinates in momentum space. From this it follows that ρ(E) = V (2π)3 k2 d |k| dE dΩk (109) and since dE d|k| = |k| E we have ρ(E) = V (2π)3 E |k| dΩk (110) 24 We thus have the number w = |U|2 4V 2E2 ¯uk,sγ0 up,r 2 V (2π)2 E |k| dΩk (111) representing the probability per unit time that a particle gets scattered into the space angle dΩk. More precisely, since the incoming wave-function is extended in all space and is normalized to one, which means that there is only one particle in the whole universe, we have calculated the probability for scattering if we have an incoming flux (= the number of particles per unit time and unit area) of v V (where v = |k| E is the speed of the incoming particle). Since we would like to get a number which is not dependent on the particular incoming flux that we have chosen, we divide the probability w by the flux and get a number called the (differential) cross section. This number characterizes the physical process and is not dependent on any particular choice of flux used in the experiment. It is given by dσ = |U(k − p)|2 4(2π)2 ¯uk,sγ0 up,r 2 dΩk (112) To get the actual number of scattered particles per unit time that we will measure in our detector, we have to multiply this number with the incoming flux we are using in the experiment. If we are interested here in the cross section when the incoming particle has some particular spin and the outgoing particle also has some fixed spin we just insert their corresponding spinors u and ¯u in the expression for the cross-section above and we are done. However, if we assume that the initial state is unpolarized which means that half of the particles have spin up and the other half have spin down, but that the relative phases of the particles are totally random, then the resulting probability (cross section) is given by averaging over the spin of the initial wave-function. In this case that means summing the final result over r and multiplying by 1 2 . If we also do not measure the spin of the outgoing particle we have to sum the final probability (cross section) over the separate probabilities to measure an outgoing particle with spin up and an outgoing particle with spin down. This gives us dσ = |U(k − p)|2 4(2π)2 1 2 r s ¯uk,sγ0 up,r 2 dΩk (113) The sums over the different spins can be written r s ¯uk,sγ0 up,r ¯up,rγ0 uk,s (114) 25 or, writing out the matrix indexes explicitly s (uk,s)a (¯uk,s)b γ0 bc r (up,r)c (¯up,r)d γ0 da (115) which, defining the matrix Mab(k) ≡ s (uk,s)a (¯uk,s)b, can be written Tr M(k)γ0 M(p)γ0 (116) Using the explicit representation of the spinors one can find that M(k) = /k + m (117) so we have dσ = |U(k − p)|2 8(2π)2 Tr (/k + m)γ0 (/p + m)γ0 dΩk (118) Using the gamma-matrix anti-commutation relations we can compute Tr (/k + m)γ0 (/p + m)γ0 = 4 m2 + EkEp + k · p (119) Finally using that Ek = Ep and |k| = |p| we can choose a coordinate system so that p is along the z-axis and k is pointing in the (θ, φ) direction. Inserting this we get the final formula for the relativistically corrected Rutherford formula also called the Mott cross section. dσ = Z2 e4 4 sin4 θ 2 E2 |k|4 1 − v2 sin2 θ 2 dΩ (120) 9 Pair creation What is the probability that an electron/positron pair is created in the po- tential Aµ = 0, 0, 0, √ 4πa cos (ωt) (121) This 4-potential represents an electric field directed in the 3 direction and oscillating with frequency ω. In this case the initial and final states are of course given by |i = |0 (122) |f = b† k1,r1 d† k2,r2 |0 26 representing the fact that initially we do not have any particles at all but we will end up with both an electron with momentum k1 and a positron with momentum k2. As usual the probability amplitude is given by M = f|T e−i HI |i (123) which, to lowest order in the expansion parameter e can be written as M = −ie f| d4 xˆ¯ψ/A ˆψ|0 = −ie d4 xA3 0|bk1,r1 dk2,r2 ˆ¯ψγ3 ˆψ|0 (124) It is quite clear that the only piece that will survive is the piece containing the operator b† from ˆ¯ψ and the operator d† from ˆψ. What remains is − ie V q1,s1 q2,s2 d4 xA3 ei(q1+q2)·x 4Eq1 Eq2 ¯uq1,s1 γ3 vq2,s2 0|bk1,r1 dk2,r2 b† q1,s1 d† q2,s2 |0 (125) which, using the anti-commutation relations simplifies to ie V d4 xA3 ei(k1+k2)·x 4Ek1 Ek2 ¯uk1,r1 γ3 vk2,r2 (126) Inserting the expression for the potential we write iea √ 4π V 4Ek1 Ek2 ¯uk1,r1 γ3 vk2,r2 d4 x cos(ωt)ei(k1+k2)·x (127) The integral can be performed by rewriting cos(ωt) in terms of exponentials as d4 x cos(ωt)ei(k1+k2)·x = d3 xe−i(k1+k2)·x dtei(Ek1 +Ek2 )t eiωt + e−iωt 2 = (2π)4 2 δ3 (k1 + k2) (δ (Ek1 + Ek2 + ω) + δ (Ek1 + Ek2 − ω)) (128) The term containing δ (Ek1 + Ek2 + ω) will clearly not give any contribution since Ek1 , Ek2 and ω are all positive. Therefore we have for the probability amplitude M = i(2π)4 ea √ 4π 2V 4Ek1 Ek2 ¯uk1,r1 γ3 vk2,r2 δ3 (k1 + k2) δ (Ek1 + Ek2 − ω) (129) 27 Again we see that the delta functions express energy conservation ω = Ek1 + Ek2 and momentum conservation k1 + k2 = 0. Namely, the frequency of the electric field has to represent an energy which precisely matches the energy of the created electron/positron pair. Also, the electron/positron has to come out back-to-back so that momentum is conserved. This also means that Ek1 = Ek2 ≡ E. The probability is the absolute square of the probability amplitude P = πe2 a2 4V 2E2 ¯uk1,r1 γ3 vk2,r2 2 (2π)3 δ3 (k1 + k2) (2π)δ (2E − ω) 2 (130) By a similar trick as in the last section we evaluate the square of the delta functions to be (2π)3 δ3 (k1 + k2) (2π)δ (2E − ω) 2 = V T(2π)4 δ3 (k1 + k2) δ (2E − ω)(131) giving us P = T πe2 a2 (2π)4 4V E2 ¯uk1,r1 γ3 vk2,r2 2 δ3 (k1 + k2) δ (2E − ω) (132) As in the previous example, we cannot separate final states which are too close in phase space. Therefore we have to sum over these probabilities to get a total “effective” probability. As in the previous example, this means including a factor V (2π)3 d3 k for each final particle giving us P = T πe2 a2 (2π)4 4V E2 ¯uk1,r1 γ3 vk2,r2 2 δ3 (k1 + k2) δ (2E − ω) V (2π)3 d3 k1 V (2π)3 d3 k2 (133) One of the integrals is easily performed using the first delta function giving us that k1 = −k2 ≡ k and the second integral we can perform, as in the previous case after rewriting d3 k = |k|2 d |k| dΩk = |k|2 d|k| dE dEdΩk = |k| EdEdΩk. The result is (noticing that δ(2E − ω) = 1 2 δ(E − ω 2 )) P = V T e2 a2 |k| 16πω ¯uk1,r1 γ3 vk2,r2 2 dΩk (134) Notice that the probability is proportional to the volume (of the region with the electric field) and the time we let the field act, in accordance to physical 28 expectations. This makes it more useful to speak about the probability per unit volume and unit time, P V T . If we are not observing the spins of the final particles we have to sum over the different probabilities of observing the different possible spins. Then we will get the formula P V T = e2 a2 |k| 16πω dΩk r1,r2 ¯uk1,r1 γ3 vk2,r2 2 (135) Using the trick of the last section, this can be rewritten as P V T = e2 a2 |k| 16πω dΩkTr (/k1 + m)γ3 (/k2 − m)γ3 (136) or, using the anti commutation relations of the gamma matrices and remembering that k1 = (E, k), k2 = (E, −k) P V T = e2 a2 |k| 16πω dΩk2ω2 1 − v2 cos2 (θ) = e2 a2 ω2 16π 1 − 4m2 ω2 1 − v2 cos2 (θ) dΩk (137) which gives the probability of an electron/positron pair with a momentum with angle θ towards the electric field. If we are interested in the total probability, irrespective of the angle, we have to integrate over dΩ to get P V T = e2 a2 ω2 6 1 − 4m2 ω2 1 + 2m2 ω2 (138) Notice that there is a “threshold” in the energy. The probability is zero for ω ≤ 2m ≤ 2E, i.e. the energy of the photons in the field must be larger than the mass of the electron/positron pair to be able to create it. In contrast the probability is non-zero for arbitrary small amplitude a of the field. 10 The quantized electro-magnetic field In the two previous examples the electro-magnetic field was treated classically as an external field. In order to incorporate photons into the theory we need to quantize also the electro-magnetic field. We will do this in a relativistically 29 covariant fashion, so let us start by recapitulating some notation. Remember that we may use a scalar potential φ and a vector potential A to describe the electric and magnetic fields E = − φ − ∂tA, B = × A. (139) Introducing the four vector Aµ = (φ, A) we may write the electric and the magnetic field in a compact form Fµν = ∂µ Aν − ∂ν Aµ =      0 −Ex −Ey −Ez Ex 0 −Bz By Ey Bz 0 −Bx Ez −By Bx 0      . (140) Using Fµν , Maxwell’s equations can also be written covariantly as ∂µFµν = 0, ∂µ Fνσ + ∂σ Fµν + ∂ν Fσµ = 0. (141) Here we have used Einstein’s summation convention (sum over repeated indexes). It is interesting to observe that Aµ is not uniquely specified by the electric and magnetic field. Namely, if we define a new vector potential by Aµ new = Aµ old + ∂µ χ, (142) for any function χ(x), the field strength Fµν , and thus the electric and magnetic fields, remain unchanged. This we can use to simplify the form of Maxwell’s equations. If we choose χ so that ∂µAµ new = 0, i.e. so that ∂µ∂µ χ + ∂µAµ old = 0, then we have ∂µFµν new = ∂µ (∂µ Aν new − ∂ν Aµ new) = ∂µ∂µ Aν new = 0, (143) that is, each component of the vector potential has to satisfy the (massless) Klein-Gordon equation (which we have already solved!). Thus, the vector potential we will use will have to satisfy two equations ∂µAµ = 0, ∂µ∂µ Aν = 0. (144) 30 However, this still does not completely specify Aµ . We may still shift it as Aµ fin = Aµ new + ∂µ Λ with a Λ satisfying ∂µ∂µ Λ = 0 since this leaves the two equations (144) invariant. This additional invariance can be used to choose A0 fin = 0. We have thus seen that we can always choose a vector potential which satisfies the following three equations A0 = 0, ∂µAµ = 0, (145) ∂µ∂µ Aν = 0. This choice of the form of the vector potential (or choice of gauge as the jargon goes) is known as the Coulomb gauge. Using this information we may now immediately write down the quantized electromagnetic field Aµ = 1 √ V k,α 1 √ 2ω ak,α (α) µ e−ikx + a† k,α (α) µ eikx . (146) From the third equation in (145) we find that kµkµ = ω2 − k2 = 0. The second equation tells us that k · (α) = 0 while the first equation tells us that (α) 0 = 0. We thus find that (α) µ is a four vector with zero time component and orthogonal to the four momentum. Thus, out of the four orthonormal four vectors (0) µ = (1, 0), (1) µ = (0, ¯(1) ), (2) µ = (0, ¯(2) ), (147) (3) µ = (0, k |k| ), only (1) µ and (2) µ are admissible. Thus we see that the sum over α in (146) is restricted to α = 1, 2 in order for Aµ to satisfy all three equations in (145). The (α) µ vectors are known as the polarization vectors of the photon. As usual, when we second quantize, the ak,α, a† k,α become annihilation and creation operators which annihilate/create a photon with momentum k and polarization (α) µ . 31 11 The electron propagator If we are interested in the next to lowest order corrections to the scattering in an external potential, we have to study the term − e2 2 f| d4 x1 d4 x2T ( ¯ψ/Aψ)(x1)( ¯ψ/Aψ)(x2) |i (148) Notice here that the T operator is non-trivial and important since the two ¯ψ/Aψ factors change place if t2 > t1. Let us first assume that t1 > t2. Then the ordering given above is the correct one and we can use the term as it stands. For each ψ factor there are two different choices for the operator, one associated to the electron and one associated with the positron. Since there are four ψ operators, we have in principle 16 different terms. However, out of these 16 terms, only 2 are non-trivial. All the others are either zero or they represent “non-connected” terms in the sense discussed before. For instance, there is one term which represents an electron getting scattered at x1 while at x2 a positron is created and annihilated but there is no contact between these two points. The first non-trivial term is schematically 0|bf b† 1b1b† 2b2b† i |0 = δf1δ12δ2i (149) representing the incoming electron being scattered first at x2 and then, later at x1. The second non-trivial term is 0|bf d1b1b† 2d† 2b† i |0 = −δ12δf2δ1i + δ12δ12δfi (150) the second term here is again “non-connected”, representing a process with no scattering because of the δfi but the first term is interesting and slightly hard to interpret. Since t1 > t2 we have to interpret it as the incoming electron flies past the point x2 where there is an electron/positron pair created and later annihilates with the positron just created at x2. The electron created at x1 is in fact the final outgoing electron. Graphically we have x2 x1 x2 x1 (1) (2) - 32 Similarly, when t2 > t1 we have graphically x2 x1 (4) x2 x1 (3) - + Using this knowledge we write diagram (1) and (3) (where the initial electron is annihilated in x2) as 0|bf ¯ψa(x1)|0 0|T ψb(x1) ¯ψc(x2) |0 0|ψd(x2)b† i |0 (151) and the diagrams (2) and (4) (where the initial electron is annihilated in x1) as 0|bf ¯ψc(x2)|0 0|T ψd(x2) ¯ψa(x1) |0 0|ψb(x1)b† i |0 (152) and the total amplitude is of course the sum of these two terms. We see that in this expression the object 0|T ψa(x1) ¯ψb(x2) |0 plays an important role. It represents the particle going between the points x1 and x2 and if t1 > t2 it is an electron but if t2 > t1 it is a positron. This object is therefore called the electron/positron propagator and we will now proceed to calculate it. Because of the time ordering operator T we have to consider two cases. Assume to begin with that t1 > t2. Then we pick out the electron creation/annihilation operators and the propagator can be written as k1,s1 k2,s2 e−ik1·x1+ik2·x2 2V √ E1E2 (u1)a(¯u2)b 0|b1b† 2|0 (153) Since 0|b1b† 2|0 = δk1,k2 δs1,s2 we can evaluate one of the sums “for free” k1,s1 eik1·(x2−x1) 2V E1 (u1)a(¯u1)b (154) Using the results of the previous section ( s u(s) a ¯u (s) b = (/k + m)ab) we can calculate the sum over the spin in the expression for the propagator. The result is k1 eik1·(x2−x1) 2V E1 (/k1 + m)ab (155) 33 For convenience we will here change the summation over momenta k into an integral. This we can do since the volume of the universe is large (so that the distribution of states in momentum space is almost continuous) which means that k = V (2π)3 d3 k. This gives us d3 k (2π)3 eiE(t2−t1) e−ik·(x2−x1) (/k + m)ab 2E (156) (where we have dropped the index 1 for convenience). Notice here that E is a function of k. The same analysis in the case where t2 > t1 gives − d3 k (2π)3 eiE(t1−t2) e−ik·(x1−x2) (/k − m)ab 2E (157) where the minus sign comes from the fact that the T operator has reordered two fermionic operators. We can rewrite the result in a more covariant form by using the integral − e−iE|t| 2E = lim ε→0 1 2πi ∞ −∞ dk0 e−ik0t k2 0 − E2 + iε (158) The first part of the propagator then becomes d3 k (2π)3 eiE(t2−t1) e−ik·(x2−x1) (/k + m)ab 2E = (−i/∂2 + m)ab d3 k (2π)3 eiE(t2−t1) 2E e−ik·(x2−x1) = (−i/∂2 + m)ab d3 k (2π)3 i dk0 2π eik0(t2−t1) k2 0 − E2 + iε e−ik·(x2−x1) (159) Since E2 = k2 + m2 the denominator of the integrand can be written k2 0 − E2 + iε = k2 0 − k2 − m2 + iε = k2 − m2 + iε giving us i (−i/∂2 + m)ab d4 k (2π)4 eik·(x2−x1) k2 − m2 + iε (160) and pushing the derivative operator back in, we get i d4 k (2π)4 (/k + m)ab k2 − m2 + iε eik·(x2−x1) (161) 34 The second part of the propagator (the one with t2 > t1) can similarly be rewritten i d4 k (2π)4 (−/k + m)ab k2 − m2 + iε eik·(x1−x2) (162) and, changing the integration variable from k to −k, we get i d4 k (2π)4 (/k + m)ab k2 − m2 + iε eik·(x2−x1) (163) which is exactly the same expression as for the part of the propagator with t1 > t2. We thus have a unique expression for the propagator G(x1 − x2) ≡ i d4 k (2π)4 (/k + m)ab k2 − m2 + iε e−ik·(x1−x2) (164) independent of whether t1 or t2 comes first. Notice that if we act with the Dirac operator i/∂ − m on the propagator we get (i/∂ − m)G(x) = i d4 k (2π)4 (/k − m)(/k + m) k2 − m2 + iε e−ik·x = i d4 k (2π)4 e−ik·x = iδ4 (x) (165) so that G(x) is the Green function of the Dirac operator in accordance with the usual interpretation of the propagator. 12 Compton scattering Compton scattering is scattering of a photon on an electron. In the initial state we therefore have a photon and an electron |i = b† p,sa† k,α|0 (166) Since the photon is physical the index α takes only the values 1, 2 corresponding to the two physical polarizations of the photon. The final state also contains a photon and an electron but with different spins and momenta |f = b† p ,s a† k ,α |0 (167) 35 As usual, the probability amplitude is given by M = f|T e−i HI |i = f|i − i f| HI|i − 1 2 f|T HI HI|i + . . . (168) The first term is non-zero only for the case when no scattering is taking place (p = p , k = k ). The second term is zero because it always involves exactly three photon creation/annihilation operators so the lowest non-trivial term is the third one. Separating the piece that has to do with electrons/positrons and the piece that has to do with photons we can write it as − e2 2 d4 x1d4 x2 0|bp ,s T ( ¯ψaψb)(x1)( ¯ψcψd)(x2) b† p,s|0 (γµ )ab (γν )cd 0|ak ,α T [Aµ(x1)Aν(x2)] a† k,α|0 (169) The piece associated with the electrons/positrons we have already calculated in the previous section. It is given by two terms corresponding to the cases where the electron first goes to the point x2, interacts, then goes to the point x1 where it is scattered to the final electron state and, oppositely when it goes first to x1 and then continues to x2. Graphically this can be represented as x1 x2x1 x2 + For the second process the expression is (¯up ,s )a √ 2V E eip ·x1 i d4 q (2π)4 (/q + m)bc q2 − m2 + iε e−iq·(x1−x2) (up,s)d √ 2V E e−ip·x2 (170) Here we recognize the middle part as the electron/positron propagator corresponding to the piece where the electron propagates from the point x2 to the point x1. The expression for the first term is similar. Evaluating the photon piece we get a sum of two terms. The first comes from choosing the creation operator in Aµ(x1) and the annihilation operator in Aν(x2), representing the case where the incoming photon is annihilated in x2 and the outgoing photon is created in x1, and the second term (where we also have to ignore a “non-connected” piece) comes from choosing the annihilation operator in 36 Aµ(x1) and the creation operator in Aν(x2), representing the case where the incoming photon is annihilated in x1 and the outgoing photon is created in x2. Explicitly we have 4π (εk ,α )µ √ 2V ω eik ·x1 (εk,α)ν √ 2V ω e−ik·x2 + 4π (εk ,α )ν √ 2V ω eik ·x2 (εk,α)µ √ 2V ω e−ik·x1 (171) Graphically we can write this as x2 x1 x1 x2 + Putting these two terms together we graphically get ++ ++ + Since we are integrating over x1 and x2 in the final expression these variables are really “dummy” variables, meaning that we can anywhere rename them as we wish. In particular we can interchange them x1 ↔ x2. From the pictures we see that two of the pictures change into the other two under this relabeling so we have really to calculate only two terms, graphically they look like this ++ 37 The second term we can write as π V 2 √ E Eω ω d4 x1 d4 x2i d4 q (2π)4 ¯up ,s /εk ,α (/q + m)/εk,αup,s q2 − m2 + iε eip ·x2 eik ·x2 e−ip·x1 e−ik·x1 e−iq·(x2−x1) (172) and doing the integrals over x1 and x2 we get π V 2 √ E Eω ω i d4 q (2π)4 ¯up ,s /εk ,α (/q + m)/εk,αup,s q2 − m2 + iε (2π)4 δ4 (p + k − q)(2π)4 δ4 (q − p − k) (173) Notice that the delta functions express momentum conservation at each of the vertexes. The expression for the first diagram is the same except that one has to switch positions for the polarization vectors ε and switch the place of k and k in the delta functions. We can get rid of one of the delta functions by performing the q integral which gives us − iπe2 (2π)4 δ4 (p + k − p − k) V 2 √ EE ωω ¯up ,s /εk ,α /p + /k + m (p + k)2 − m2 + iε /εk,αup,s+ ¯up ,s /εk,α /p − /k + m (p − k )2 − m2 + iε /εk ,α up,s (174) Here we notice that the delta function which is left just expresses the momentum conservation of the whole process. To simplify this expression further we can use the Dirac equation on the spinors /pup = mup (175) which, using the anti-commutation relations of the gamma matrices, leads to /p/εup = −/εmup + 2p · εup (176) and choosing a coordinate system where the initial electron is at rest p = (m, 0, 0, 0) so that p · ε = p · ε = 0 we get − iπe2 (2π)4 δ4 (p + k − p − k) V 2 √ EE ωω ¯u /ε /k/εu 2p · k + ¯u /ε/k /ε u 2p · k (177) Now we calculate the probability density by taking the absolute square of the amplitude. At the same time we say that we are not interested in the 38 polarization of the final electron, and the initial electron comes in a mixed state so that we have to include a sum 1 2 s,s . We then have the probability P = πe2 (2π)4 δ4 (p + k − p − k) 2V 2 √ EE ωω 2 1 2 s,s ¯u /ε /k/εu p · k + ¯u /ε/k /ε u p · k ¯u/ε/k/ε u p · k + ¯u/ε /k /εu p · k (178) where we have used that (γµ )† = γ0 γµ γ0 . Performing the spin sums and noticing that the first term in each parenthesis is equal to the second if we make the exchange ε ↔ ε and k ↔ −k we have P = πe2 (2π)4 δ4 (p + k − p − k) 2V 2 √ 2EE ωω 2 Tr /ε /k/ε(/p + m)/ε/k/ε (/p + m) (p · k)2 + /ε /k/ε(/p + m)/ε /k /ε(/p + m) (p · k)(p · k ) + ε ↔ ε k ↔ −k (179) The reason for the minus sign in the exchange of the photon momenta is that we do not want p or p to change, but since p = p + k − k we have to interchange k and k with an extra minus sign. Let us perform the trace over the first term explicitly. The fact that the trace over any odd number of gamma matrices is zero allows us to write it as Tr (/ε /k/ε/p/ε/k/ε /p ) + m2 Tr (/ε /k/ε/ε/k/ε ) (180) Using the gamma matrix algebra we know that /a/b = −/b/a + 2a · b which also implies that /a/a = a · a, and using the cyclicity of the trace, we can show that the second term is m2 Tr (/k/ε/ε/k/ε /ε ) = m2 (ε · ε)(ε · ε )Tr (/k/k) = 0 (181) Thus it only remains to calculate the first term. Anti-commuting the /p/k in the middle and using that /k/k = k · k = 0 we have Tr (/ε /k/ε/p/ε/k/ε /p ) = 2(p · k)Tr (/ε /k/ε /p ) (182) 39 Then we can anti-commute /ε /k and use that ε · ε = −1 to get 2(k · p)Tr ((−/k/ε + 2(ε · k)) /ε /p ) = 8(k · p) (2(k · ε )(ε · p ) + (k · p )) (183) where we in the last step used that Tr (γµ γν ) = 4gµν . Similarly the second term in the full trace can be calculated to be −8(k · p)(k · p) + 16(ε · ε)2 (k · p)(k · p) − 8(k · ε )2 (k · p) + 8(k · ε)2 (k · p)(184) and the two last terms are given by the first two by the interchange above. Summing all the traces together there are a lot of cancellations and the final result is 8 k · p k · p + k · p k · p + 4(ε · ε)2 − 2 (185) or, using that k · p = mω and k · p = mω 8 ω ω + ω ω + 4(ε · ε)2 − 2 (186) Now let us return to the calculation of the probability, it can be written as P = πe2 (2π)4 δ4 (p + k − p − k) 2V 2 √ 2EE ωω 2 8 ω ω + ω ω + 4(ε · ε)2 − 2 = V T π2 e4 V 4EE ωω (2π)4 δ4 (p + k − p − k) ω ω + ω ω + 4(ε · ε)2 − 2 (187) Calculating the probability per unit time and summing over inseparable final states we have P T = π2 e4 V 3EE ωω ω ω + ω ω + 4(ε · ε)2 − 2 (2π)4 δ4 (p + k − p − k) V (2π)3 d3 p V (2π)3 d3 k (188) We get rid of the space-like part of the delta function if we perform one of the integrals, say the one over p , which leaves us with P T = π2 e4 (2π)2V EE ωω ω ω + ω ω + 4(ε · ε)2 − 2 δ(E + ω − E − ω)d3 k (189) 40 To get rid of the last delta function we can use that ω = |k | to rewrite d3 k = (ω )2 dω dΩk. What complicates things slightly is that since we already used the space-like delta function, E = |p |2 + m2 = |k − k |2 + m2 = ω2 + ω 2 − 2ωω cos(θ) + m2 is also a function of ω . To perform the integration we have to write δ(E + ω − E − ω)dω = δ(E + ω − E − ω) dω d(E + ω ) d(E + ω ) = dω d(E + ω ) (190) Now d(E + ω ) dω = ω − ω cos(θ) E + 1 (191) which, using that E = ω − ω + m, we can write as d(E + ω ) dω = ω(1 − cos(θ)) + m E = ωω (1 − cos(θ)) + mω E ω = k · k + p · k E ω = (k + p ) · k E ω = p · k E ω = p · k E ω = Eω E ω (192) where we have used that p + k = p + k which in turn implies (by squaring) that p · k = p · k . Finally, to get a number which is independent of how often we throw in photons (i.e. the cross section) we have to divide by the flux. Since the photon wave-function is normalized to one photon in the whole universe and since the speed of the photon is 1 (in natural units), the incoming flux is 1 V . This gives us an expression for the cross section as dσ = π2 e4 (2π)2EE ωω ω ω + ω ω + 4(ε · ε)2 − 2 E ω 3 Eω dΩ = e4 4m2 ω ω 2 ω ω + ω ω + 4(ε · ε)2 − 2 dΩ (193) which is the famous Klein-Nishina formula for the cross section of Compton scattering. 41 13 The photon propagator Remember that in a previous section we derived the electron/positron propagator which could be written as Gab(x1 − x2) = 0|T ψa(x1) ¯ψb(x2) |0 (194) As we will see in the next section, there is also a photon propagator corresponding to a photon propagating between two points x1 and x2. As one can guess, it can be written as Dµν(x1 − x2) = 0|T [Aµ(x1)Aν(x2)] |0 (195) Let us calculate it here so that we can use it when it appears in the next section. Inserting the expressions for the photon fields we have k1,α1 k2,α2 4π 2V √ ω1ω2 0|T (a† 1ε1µeik1·x1 + a1ε1µe−ik1·x1 ) (a† 2ε2νeik2·x2 + a2ε2νe−ik2·x2 ) |0 = k1,α1 k2,α2 4π 2V √ ω1ω2 θ (t1 − t2) ε1µε2νe−ik1·x1 eik2·x2 0|a1a† 2|0 +θ (t2 − t1) ε2νε1µe−ik2·x2 eik1·x1 0|a2a† 1|0 (196) Remembering that ak,α, a† k ,α = δk,k δα,α α = 1, 2 (197) (198) we find Dµν(x1 − x2) = k 4π 2V ω θ (t1 − t2) ε(1) µ ε(1) ν + ε(2) µ ε(2) ν e−ik·(x1−x2) +θ (t2 − t1) ε(1) ν ε(1) µ + ε(2) ν ε(2) µ eik·(x1−x2) . (199) This expression does not look too satisfying since the rather messy expression ε(1) ν ε(1) µ + ε(2) ν ε(2) µ appears. In fact, using the explicit expressions for the polarization vectors ε (0) k = (1, 0, 0, 0) 42 ε (1) k = (0, ¯ε1) ε (2) k = (0, ¯ε2) (200) ε (3) k = (0, k |k| ) where ¯ε1 and ¯ε2 are two three dimensional unit vectors which are orthogonal both to k and to each other, one may show that − ε(0) ν ε(0) µ + ε(1) ν ε(1) µ + ε(2) ν ε(2) µ + ε(3) ν ε(3) µ = −gµν. (201) One way to prove it would be to say that it is a symmetric matrix with eigenvalues (-1,1,1,1), therefore it is diagonalizable to −gµν using an orthogonal matrix M so that MT (−ε0εT 0 + ε1εT 1 + ε2εT 2 + ε3εT 3 )M = −g. But by multiplying on the left with M and on the right with MT we find that (−ε0εT 0 + ε1εT 1 + ε2εT 2 + ε3εT 3 ) = −MgMT = −g. Furthermore we may rewrite the − ε(0) ν ε(0) µ + ε(3) ν ε(3) µ piece using the four vectors kµ = (ω, −k) and aµ = (ω, k) as − ε(0) ν ε(0) µ + ε(3) ν ε(3) µ = − 1 2ω2 (aµkν + kµaν) . (202) We thus find that we may write the propagator Dµν(x1 − x2) = k −gµν + 1 2ω2 (aµkν + kµaν) × 4π 2V ω θ (t1 − t2) e−ik·(x1−x2) + θ (t2 − t1) e−ik·(x2−x1) (203) Again using the fact that for a high density of states we can switch the sum over k for an integral over the density of states k = V d3 k (2π)3 (204) and again using the rewriting of the energy dependence as − e−iω|t| 2ω = 1 2πi ∞ −∞ dk0 e−ik0t k2 0 − |k|2 + iε (205) 43 we can write the propagator as Dµν(x1 − x2) = d3 k (2π)3 4π gµν − 1 2ω2 (aµkν + kµaν) × θ(t1 − t2) 1 2πi dk0 k2 + iε eik0(t1−t2) eik·(x1−x2) + θ(t2 − t1) 1 2πi dk0 k2 + iε eik0(t1−t2) e−ik·(x1−x2) . (206) and changing the sign of the integration variable k in the second term, we see that both the term for t1 > t2 and the term for t2 > t1 have exactly the same form, just as for the electron/positron propagator, so we can write it in a compact form Dµν(x1 − x2) = −i4π d4 k (2π)4 gµν − 1 2ω2 (aµkν + kµaν) e−ik·(x1−x2) k2 + iε .(207) This is very nice except for the “ak + ka” piece which ruins the covariant form of the propagator. However, the fact that this piece contains explicit factors of kµ which is the same four momentum as the momentum which flows in the propagator saves us. To see this, let us consider a part of a Feynman diagram looking like p k p+k The electron line corresponds to an expression (up to normalization con- stants) ¯u(p + k)γµ u(p), (208) To the γµ is connected the momentum space propagator Dµν(k). The part of Dµν proportional to kµ together with (208) gives a contribution proportional 44 to ¯u(p + k)/ku(p) = ¯u(p + k) (/k + /p − /p) u(p) = {Dirac equation} = ¯u(p + k) (m − m) u(p) = 0. (209) Therefore the piece proportional to kµaν does not give any contribution to physical processes and can consequently be dropped. Similarly, the piece proportional to aµkν cancels in the other end of the propagator. The full proof of this fact is a little bit involved. One needs to check that it is true also when the photon propagator ends on an internal electron propagator and not on an external line as in the simple example above. If you believe me for now that this is true we can write the photon propagator simply as Dµν(x1 − x2) = −i4πgµν d4 k (2π)4 e−ik·(x1−x2) k2 + iε . (210) 14 Electron-electron scattering Let us now consider scattering of two electrons. Let us assume that initially they have momenta and spins p1, s1 and p2, s2. They are scattered into electrons with momenta and spins p3, s3 and p4.s4. We therefore take as the initial state |i = b† 1b† 2|0 (211) and as the final state we take |f = b† 4b† 3|0 (212) The probability amplitude is given by the usual expression M = f|T e−i HI |i (213) and the first non-trivial term is the term which is second order in HI. Since there are 4 b operators in the initial and final states the only combination which will give anything is when we choose the b operators from the 4 ψ fields. The expression is therefore 0|b3b4b† b b † b b† 1b† 2|0 = −δ1,p δ2,p δ3,p δ4,p + δ1,p δ2,p δ3,pδ4,p +δ1,p δ2,p δ3,p δ4,p − δ1,p δ2,p δ3,pδ4,p + non−connected pieces (214) 45 where we have used a shorthand notation in that each delta function also comes with a corresponding delta for the spin dependence so that δ1,p really means δ1,pδα1,α. These terms can be graphically represented as x1 x1 x1 x1 x2 x2 x2 x2 + 1 2 43 1 2 3 4 3 4 1 221 3 4 +- Again using the trick of changing the integration variables x1 and x2 we see that the two last diagrams are equal to the two first diagrams which leaves us with only two expressions to be calculated. The one corresponding to the first diagram we can write as e2 4V 2 √ E1E2E3E4 (¯u4γµ u2)(¯u3γν u1) d4 x1 d4 x2 ei(p4−p2)·x1 ei(p3−p1)·x2 0|T [Aµ(x1)Aν(x2)] |0 (215) where in the last expression we recognize the photon propagator Dµν(x1 −x2) which we calculated in the previous section. Inserting the expression we obtained we get e2 4V 2 √ E1E2E3E4 (¯u4γµ u2)(¯u3γν u1) d4 x1 d4 x2 d4 k (2π)4 ei(p4−p2−k)·x1 ei(p3−p1+k)·x2 4πgµν i(k2 + iε) (216) and performing the x1 and x2 integrals we get e2 4V 2 √ E1E2E3E4 (¯u4γµ u2)(¯u3γν u1) d4 k (2π)4 (2π)4 δ(p4 − p2 − k)(2π)4 δ(p3 − p1 + k) 4πgµν i(k2 + iε) (217) 46 We can get rid of one of the delta functions by doing the k integral e2 4V 2 √ E1E2E3E4 (¯u4γµ u2)(¯u3γν u1) (2π)4 δ(p1 + p2 − p3 − p4) 4πgµν i((p1 − p3)2 + iε) (218) Notice that the remaining delta function expresses the total conservation of momentum. The second diagram can be easily calculated when we realize that the only thing that differs between the second and the first diagram is that we have to switch 3 and 4 and also the overall sign. Totally we therefore have the probability amplitude M = e2 (2π)4 δ(p1 + p2 − p3 − p4) 4V 2 √ E1E2E3E4 (¯u4γµ u2) 4πgµν i((p1 − p3)2 + iε) (¯u3γν u1)− (¯u3γµ u2) 4πgµν i((p1 − p4)2 + iε) (¯u4γν u1)(219) If we assume that the incoming electrons are unpolarized so that we have to average over the incoming spins and that we do not observe the spin of the outgoing electrons so that we have to sum over the probabilities of observing different outgoing spin, we have to include a sum over 1 2 s1 1 2 s2 s3 s4 (220) Introducing the notation t = (p1 − p3)2 = (p4 − p2)2 u = (p1 − p4)2 = (p3 − p2)2 (221) we can write the total probability as P = T e4 (2π)6 δ(p1 + p2 − p3 − p4) 4V 3E1E2E3E4 1 4 s1,s2,s3,s4 (¯u4γµu2)(¯u3γµ u1) t − (¯u3γµu2)(¯u4γµ u1) u 2 (222) 47 which, using the expressions for the spin sums, can be written as P = T e4 (2π)6 δ(p1 + p2 − p3 − p4) 4V 3E1E2E3E4 1 4 Tr((/p4 + m)γµ(/p2 + m)γν)Tr((/p3 + m)γµ (/p1 + m)γν ) t2 + Tr((/p3 + m)γµ(/p2 + m)γν)Tr((/p4 + m)γµ (/p1 + m)γν ) u2 − Tr((/p4 + m)γµ(/p2 + m)γν(/p3 + m)γµ (/p1 + m)γν ) tu − Tr((/p3 + m)γµ(/p2 + m)γν(/p4 + m)γµ (/p1 + m)γν ) tu (223) Let us explicitly calculate the first trace and leave the other ones as an exercise. To do this we observe that Tr((/p + m)γµ(/q + m)γν) = Tr(/pγµ/qγν) + m2 Tr(γµγν) = 4 pµqν − (p · q)gµν + pνqµ + m2 gµν (224) giving the result for the first term 16 2(p1 · p2)(p3 · p4) + 2(p1 · p4)(p2 · p3) + 2(m2 − p2 · p4)(p1 · p3)+ 2(m2 − p1 · p3)(p2 · p4) + 4(m2 − p1 · p3)(m2 − p2 · p4) (225) and using that the relation p1 + p2 = p3 + p4 implies p1 · p2 = p3 · p4 p1 · p3 = p2 · p4 (226) p1 · p4 = p2 · p3 (227) we can write it as 32 (p1 · p2)2 + (p1 · p4)2 + 2m2 (m2 − p1 · p3) (228) Computing also the remaining traces we get the full answer P = T e4 (2π)6 δ(p1 + p2 − p3 − p4) 4V 3E1E2E3E4 48 8 (p1 · p2)2 + (p1 · p4)2 + 2m2 (m2 − p1 · p3) t2 + (p1 · p2)2 + (p1 · p3)2 + 2m2 (m2 − p1 · p4) u2 − 2(p1 · p2)(2m2 − p1 · p2) tu (229) Let us choose center of mass coordinates such that p1 = (E, p) p2 = (E, −p) p3 = (E, p ) (230) p4 = (E, −p ) and that p · p = |p|2 cos(θ). Then we see that we can write p1 · p2 = m2 + 2 |p|2 p1 · p3 = m2 + 2 |p|2 sin2 θ 2 (231) p1 · p4 = m2 + 2 |p|2 cos2 θ 2 To calculate the cross section we need to sum over the probabilities of observing final states which are close to each other in momentum space. This we do by including the factors V (2π)3 d3 p3 V (2π)3 d3 p4 (232) and dividing by the incoming flux v1+v2 V where v1 and v2 is the speed of the 1 and 2 particle respectively. In the center of mass system the speed of the two particles are equal and can be expressed as |p| E so that the flux is 2|p| EV . The integrals over the momenta can be taken care of in the usual way δ4 (p1 + p2 − p3 − p4)d3 p4d3 p3 = δ(E1 + E2 − E3 − E4) |p3|2 d |p3| dΩ (233) and since in the center of mass system we have that E4 = E3 = |p3|2 + m2 we can write δ(E1 + E2 − 2E3) |p3|2 d |p3| d(2E3) d(2E3)dΩ = |p3| E3 2 dΩ (234) 49 Putting everything together we have dσ = e4 (2π)6 4V 3E4 V 2 (2π)6 EV 2 |p| |p| E 2 8 (p1 · p2)2 + (p1 · p4)2 + 2m2 (m2 − p1 · p3) t2 + (p1 · p2)2 + (p1 · p3)2 + 2m2 (m2 − p1 · p4) u2 − 2(p1 · p2)(2m2 − p1 · p2) tu dΩ = e4 2E2 (p1 · p2)2 + (p1 · p4)2 + 2m2 (m2 − p1 · p3) t2 + (p1 · p2)2 + (p1 · p3)2 + 2m2 (m2 − p1 · p4) u2 − 2(p1 · p2)(2m2 − p1 · p2) tu dΩ (235) This cross section simplifies in the non-relativistic and the ultra-relativistic cases. In the non-relativistic case we have that |p| m so that to lowest order p1 · p2 ≈ m2 p1 · p3 ≈ m2 (236) p1 · p4 ≈ m2 t = (p1 − p3)2 = 2m2 − 2 m2 + 2 |p|2 sin2 θ 2 ≈ −4 |p|2 sin2 θ 2 u = 2m2 − 2 m2 + 2 |p|2 cos2 θ 2 ≈ −4 |p|2 cos2 θ 2 (237) which gives the expression dσ = e4 m2v4   1 sin4 θ 2 + 1 cos4 θ 2 − 1 sin2 θ 2 cos2 θ 2   dΩ (238) 50 where v is the velocity v = |p| m . Notice that the first term gives exactly the Rutherford cross-section. The two additional terms are of quantum mechanical origin and come about because the particles that scatter are quantum mechanically identical. This means that the cross section has to be invariant under θ → π − θ. The second term alone would be enough to achieve that. The third term is there however because scattering of identical Fermi particles is very much suppressed at θ = π 2 . This is essentially an effect of the Pauli principle which tells us that the total wave function has to be anti-symmetric. In the ultra-relativistic limit, |p| m, the cross section can similarly be written in a simple form (Møller) dσ = e4 8E2   1 + sin4 θ 2 cos4 θ 2 + 1 + cos4 θ 2 sin4 θ 2 + 2 sin2 θ 2 cos2 θ 2   dΩ (239) 15 Feynman rules, higher order processes Feynman graphs are very helpful to get an overview over the various contributions to probability amplitudes in higher orders. Basically in n-th order one has n vertices, and from the nature of the electromagnetic interaction Hamltonian follows that at each vertex two fermion lines and one photon line meet. There are a few general rules how to compose transition amplitudes at a certain order. 1. Draw all connected graphs with a given number of vertices. 2. Add a factor eγµ to every vertex and integrate over all of space. 3. Take a propagator G(x−y), or Dµν(x−y), respectively, for each fermion or photon line between the vertices at x and y. 4. Take free wave functions for external lines. 5. The exchange of any two fermions (in the construction of a graph from another one) gives a minus sign, as well as every closed fermion loop. 6. In the case of n ingoing positrons there is a relative factor (−1)n in comparison to n ingoing electrons. 51 As an example we take fourth-order diagrams for electron-positron scattering. Altogether there are 18 connected graphs, here we display five of them; x w y z i i’ f f’ (a) i xx w x z y f f’ i’(b) i (c) i’ z y x w f’f i (d) i’ z f f’ x y w i (e) i’ z x w f f’ y Their meaning is obviously the following: (a) Exchange of two photons, (b) annihilation, followed by creation of a new pair, (c) annihilation and creation, scattering of the two outgoing particles, (d) annihilation and creation, emission and absorption of a photon by one of the outgoing electrons, 52 (e) repeated annihilation and creation. The amplitudes for processes (a) and (b) can be calculated along the above rules, the calculation is lengthy but straightforward. In the cases (c), (d) and (e) however, new problems of a serious nature arise: One encounters divergent integrals, which require new technical tools and an interpretation the renormalization paradigm. In the following we study the renormalization procedure according to Pauli-Villars. 16 The vacuum polarization Let’s begin with diagram (e). Here in relation to the corresponding secondorder diagram a “fermion loop” is inserted into the photon propagator. Denote the ingoing electron’s momentum by p, the ingoing positron’s momentum by p , respectively. The photon four-momentum is p+p , which is on the other hand equal to the sum of the outgoing four-momenta. If the fermion four-momentum in one arch of the loop is k, then in the other arch it is p+p −k, where k is arbitrary. For a given momentum p+p =: q the photon propagator gµν q2+iε is replaced by 1 q2 + iε Iµν(q, m) 1 q2 + iε , (240) where Iµν contains a double fermion propagator Iµν(q, m) = −e2 d4 k (2π)4 Tr γµ 1 /k − m + iε γν 1 /k − /q − m + iε . (241) (The trace appears because of the closed loop.) For large values of k the integrand goes asymptotically like k−2 , so the momentum space integral Iµν(q) diverges quadratically, as arbitrarily large momenta may circulate in the loop. The resulting divergence is called an “ultraviolet catastrophe”. Technically a “cut-off” of high frequencies brings a remedy, but this means a “change of the rules in the course of the game” and needs a physical justification. Before modifying the theory in such a way it is convenient to carry out some formal transformations that reformulate the divergent integral. First we write the propagator in form of an integral 1 /k − m + iε = /k + m k2 − m2 + iε = −i(/k + m) ∞ 0 dz eiz(k2−m2+iε) . (242) 53 Insertion gives Iµν(q, m) = −4ie2 ∞ 0 dz1 ∞ 0 dz2 d4 k (2π)4 × (243) kµ(k − q)ν + kν(k − q)µ − gµν(k2 − k · q − m2 ) × exp iz1 k2 − m2 + iε + iz2 (k − q)2 − m2 + iε To carry out the k-integral, the part of the exponent containing k and q is rearranged in form of a complete square, exp i(z1 + z2) k − z2 q z1 + z2 2 − i (z2 q)2 z1 + z2 + iz2 q2 . With the definition l := k − z2 q z1 + z2 = k − q + z1 q z1 + z2 this becomes eil2(z1+z2) e i z1 z2 q2 z1+z2 , so that the k integral turns into three types of Gaußian integrals, namely dl (2π)4    1 lµ lµlν    eil2(z1+z2) = 1 16π2i(z1 + z2)2    1 0 igµν 2(z1+z2)    . Now Iµν(q, m) = α π ∞ 0 dz1 ∞ 0 dz2 (z1 + z2)2 e i q2 z1 z2 z1+z2 −(m2−iε)(z1+z2) × (244) 2(gµνq2 − qµqν) z1 z2 (z1 + z2)2 + gµν −i z1 + z2 − z1 z2 q2 (z1 + z2)2 + m2 (α = e2 4π is the fine structure constant.)) The part of the integrand in square brackets can be shown to contribute nothing. For this purpose we rescale the z’s, zi → λ zi. 54 Then this part becomes ∞ 0 ∞ 0 dz1 dz2 (z1 + z2)2 m2 − i λ(z1 + z2) − z1 z2 q2 (z1 + z2)2 e iλ z1 z2 q2 z1+z2 −(m2−iε)(z1+z2) = iλ ∂ ∂λ ∞ 0 ∞ 0 dz1 dz2 λ(z1 + z2)3 e iλ z1 z2 q2 z1+z2 −(m2−iε)(z1+z2) . Undoing the scaling transformation, λ zi → zi, makes the integral λ-independent, so the derivative is zero. One further transformation is done by inserting the identity 1 = ∞ 0 dλ λ δ 1 − z1 + z2 λ , (245) leading to the form Iµν(q, m) = 2iα π (qµqν − gµνq2 ) ∞ 0 ∞ 0 ∞ 0 dλ dz1 dz2 z1 z2 λ (z1 + z2)4 × δ 1 − z1 + z2 λ e i z1 z2 q2 z1+z2 −(m2−iε)(z1+z2) = (246) 2iα π (qµqν − ηµνq2 ) ∞ 0 ∞ 0 dz2 dz2 z1 z2 δ(1 − z1 − z2) ∞ 0 dλ λ eiλ(z1 z2 q2−m2+iε) . In the last line zi were again multiplied by λ. Evaluating the δ-function we finally find Iµν(q, m) = 2iα π (qµqν − gµνq2 ) 1 0 dz z(1 − z) ∞ 0 dλ λ eiλ(z1 z2 q2−m2+iε) , (247) where the original asymptotic divergence of the k-integral was replaced by a logarithmic divergence of the λ-integral. This is the divergence we want to cut off in the regularization procedure. It is removed by subtracting an analogous expression Iµν(q, M) with a fictive large mass M and the same behavior close to λ = 0, i. e. we consider ∞ 0 dλ λ eiλ(z1 z2 q2−m2+iε) − eiλ(z1 z2 q2−M2+iε) . 55 Integrals of this type can be calculated by introducing a further integration, ∞ 0 dλ λ e−aλ − e−bλ = ∞ 0 dλ b a dx e−λx = b a dx x = ln b a . (248) For large M the regularized expression for the fermion loop is approximately ¯Iµν(q) = Iµν(q, m) − Iµν(q, M) ≈ (249) 2iα π (qµqν − gµνq2 ) 1 0 dz z(1 − z) ln M2 m2 − q2z(1 − z) = iα 3π (qµqν − gµνq2 ) ln M2 m2 − 6 1 0 dz z(1 − z) ln 1 − q2 m2 z(1 − z) . When ¯Iµν is inserted into the photon propagator, the part with qµqν does not contribute to the amplitude for the same reasons as in the unrenormalized case. Taking the sum of the second-order and the fourth-order terms and neglecting iε in the denominator, we get, in first order in α, for the photon propagator Dµν → igµν q2 − i 1 q2 ¯Iµν 1 q2 , (250) explicitly igµν q2 1 − α 3π ln M2 m2 + 2α π 1 0 dz z(z − 1) ln 1 − q2 z(1 − z) m2 − iε . (251) In the limit q2 → 0 the renormalization amounts merely to a multiplication of the propagator by the factor Z3 = 1 − α 3π ln M2 m2 . (252) For a physical interpretation consider Coulomb scattering with small momentum transfer: The lowest-order expression e2 ¯uγ0 u/q2 is, in first order in α, replaced by e2 ¯uγ0 u q2 1 − α 3π ln M2 m2 =: e2 R ¯uγ0 u q2 . (253) eR = √ Z3 e is called the renormalized charge of the electron. 56 eR is the observed charge, which is measured as e2 R = 4π 137 in natural units. The parameter e in the Dirac equation would be the unobservable charge, if the electromagnetic interaction could be switched off. Accordingly e is called the “bare charge”, and eR is called the dressed charge. The interpretation of the difference between eR and e is that a charge is surrounded by a cloud of virtual photons, which, in turn, create short-lived electron-positron pairs, visualized by fermion loops in Feynman diagrams. As effective dipoles, these electron-positron pairs partially screen off the bare charge from distant observers, so that it appears smaller. The renormalized photon propagator (251) splits into two parts: the limit q2 → 0, which contains the cutoff parameter M and describes the static vacuum polarization, and a q-dependent part, which is physically meaningful as first-order correction in α. Fermion loops of the considered type are sometimes called “vacuum bubbles”. As an example of a first-order correction consider once more the Coulomb scattering amplitude: ie2 ¯uγ0u q2 1 − α 3π ln M2 m2 − α 15π q2 m2 ≈ ie2 R ¯uγ0u q2 1 − αR 15π q2 m2 + O(α2 R) . In position space the momentum space quantity q2 corresponds to the Laplacian operator, so that we get the following action on the electrostatic poten- tial 1 − αR 15πm2 e2 R 4πr = e2 R 4πr + αRe2 R 15πm2 δ(3) (x). (254) In first order in α there is a point-like potential that leads to a lowering of the energy levels of s-wave functions in atoms, which have their maximum at x = 0, ∆Enl = − Ze2 RαR 15πm2 |ψnl(0)|2 = − 1 2 Z2 α2 m 8Z2 α2 15πn3 δl0, (255) the Lamb shift. For hydrogen for example, a frequency shift by ν = ∆E/¯h = 27 MHz corresponds to the energy difference between the 2S1 2 and the 2P1 2 levels. For large momentum transfer, |q |2 ≈ −q2 m2 , on the other hand, the logarithm in the physical part of the propagator can be approximated in the following way ln 1 − q2 z(1 − z) m2 − iε ≈ ln |q|2 z(1 − z) m2 = ln |q |2 m2 + ln(z(1 − z)), 57 where the contribution of the last logarithm, when integrated with z(1 − z), is small, so that the unrenormalized propagator becomes − igµν q2 1 + α 3π ln |q |2 m2 − α 3π ln M2 m2 + O(α2 ). (256) Growing momentum transfer q partially compensates the renormalization. As large |q | means small impact parameter, particles coming very close to each other in a scattering process “dive” into their clouds of virtual dipoles and begin to feel the bare charge. What remains to do is to take care of free photon lines. Like propagators, they pick up a vacuum bubble in fourth order and a factor Z3 after renormalization. As for the propagator, √ Z3 is associated to the vertex, where the line begins or ends, for open ends, where there is no charge to renormalize, we have to divide the amplitude by √ Z3, when the renormalization is done. 17 Electron mass renormalization 17.1 Fourth-order correction to the fermion propaga- tor Diagram (d), where the electron emits and then absorbs a photon, is a contribution to the self-energy of an electric charge, which raises also a problem in classical electrostatics. The amplitude for the loop consisting of one electron and one photon propagator is iΣ(p) := (−ie)2 d4 k (2π)2 −i k2 − λ2 + i γµ i /p − /k − m − i γµ . (257) λ is a small photon mass introduced in order to avoid infrared divergences. (A physical argument could be a finite extension of the universe as a cutoff of infinite wavelengths.) The integral is linearly divergent. Like in eq. (242) we introduce variables zi to rewrite the integrals in Σ, Σ(p) = α 2π ∞ 0 ∞ 0 dz1 dz2 (z1 − z2)2 2m − /p z1 z1 + z2 e i p2z1z2 z1+z2 −m2z2−λ2z1 , (258) and like in the case of vacuum polarization, we rescale zi by γzi and apply 1 = ∞ 0 dγ γ δ 1 − z1+z2 γ to obtain Σ(p) = α 2π 1 0 dz [2m − /p(1 − z)] ∞ 0 dγ γ eiγ[p2z(1−z)−m2z−λ2(1−z)+i ] . (259) 58 The integral J(p, m, λ) = ∞ 0 dγ γ eiγ[p2z(1−z)−m2z−λ2(1−z)+i ] (260) diverges logarithmically. Σ(p) is regularized by subtraction of an analogous integral with a large photon mass Λ, followed by application of (248), J(p, m, λ) − J(p, m, Λ) ≈ ln Λ2 (1 − z) m2z + λ2(1 − z) − p2z(1 − z) − i ≈ ln Λ2 (1 − z) m2z2 + ln m2 z2 m2z + λ2(1 − z) − p2z(1 − z) . Now λ can be dropped, then the last term is zero for p2 = m2 , i. e. for a free electron on the mass shell. This leads to the regularized expression ¯Σ(p) = α 2π 1 0 dz [2m − (1 − z)/p] ln Λ2 (1 − z) m2z2 + α 2π 1 0 dz [2m − (1 − z)/p] ln m2 z m2 − p2(1 − z) . The integrals in the first, cutoff-dependent term can be easily carried out, so that ¯Σ(p) = 3αm 4π ln Λ2 m2 − α 4π (/p − m) ln Λ2 m2 + (261) α 2π 1 0 dz [2m − (1 − z)/p] ln m2 z m2 − p2(1 − z) . The integral containing the physical corrections is approximately α 2π 1 0 dz[2m − (1 − z)/p] ln m2 z m2 − p2(1 − z) = αm π m2 − p2 p2 ln m2 − p2 m2 − (262) α 4π /p m2 − p2 p2 1 + m2 + p2 p2 ln m2 − p2 m2 . Close to the mass shell p2 ≈ m2 (p2 − m2 ≈ 2m(/p − m)), we find ¯Σ(p) 3α 4π m ln Λ2 m2 − α 4π (/p − m) ln Λ2 m2 + 4 ln m2 − p2 m2 . (263) 59 17.2 Propagator renormalization In the step from second to fourth order perturbation theory i /p−m is replaced by i /p − m + i /p − m (−iΣ(p)) i /p − m = i /p − m − Σ(p) + O(α2 ). (264) Here ¯Σ(p) is briefly written as Σ. This relation is shown in the following way 1 /p − m + 1 /p − m (−/p + m + Σ + /p − m) 1 /p − m = 1 /p − m + 1 /p − m (−/p + m + Σ) 1 − 1 /p − m − Σ (/p − m) 1 /p − m = 1 /p − m − 1 − 1 /p − m Σ 1 /p − m − 1 /p − m − Σ = 1 /p − m − Σ + 1 /p − m Σ 1 /p − m − 1 /p − m − Σ . The difference in the last parentheses is of order one in α like Σ, this proves relation (264). Now we write ¯Σ in the form ¯Σ(p) = δm − [Z−1 2 − 1 + C(p)](/p − m) (265) with δm = 3αm 4π ln Λ2 m2 . (266) The function C(p) is chosen such that C(p) = 0 when p = m, thus Z−1 2 − 1 = α 4π ln Λ2 m2 − 2 ln m2 λ2 . (265) is inserted into (264), i /p − m − ¯Σ = i /p − m − δm + [Z−1 2 − 1 + C(p)](/p − m) = i −δm + [Z−1 2 + C(p)](/p − m) = i Z2 (/p − m)[1 + Z2C(p)] − Z2δm . 60 As ¯Σ(p) is of order α, from (265) follows that 1 − Z2 + Z2 C(p) is of order α, thus Z2 = 1 + Z2 C(p) + O(α). δm being of order α, too, it follows that Z2 δm = (1 + Z2 C(p)) δm + O(α2 ), so that, up to order α2 , i /p − m − ¯Σ = i Z2 (/p − m − δm)[1 + C(p)] + O(α2 ). (267) With the definition of the renormalized physical mass, mph = m + δm = m 1 + 3α 4π ln Λ2 m2 (268) the cutoff constant Λ has disappeared. δm is interpreted as the electron’s mass increase coming from its electrostatic field. The unrenormalized mass m is unobservable. In the limit p = mph, when C(p) = 0, the propagator simply picks up a multiplicative factor Z2 in addition to the mass renormalization, i /p − m → i Z2 /p − mph . (269) Analogously to the case of charge renormalization, we could multiply the charges at the ends of the propagator by a factor √ Z2, but this factors will cancel in the end. For each free fermion line, however, we have to divide the amplitude by √ Z2. 18 Vertex correction 18.1 Vertices in fourth order The quantity corresponding to the loop in (c) is Λµ(p , p) = (270) (−ie)2 d4 k (2π)4 −i k2 − λ2 + i γν i /p − /k − m + i γµ i /p − /k − m + i γν . 61 Again λ is a small photon mass, the integral is logarithmically divergent. Consider the case of small energy-momentum transfer between an external source and a free fermion p ≈ p ≈ m (see figure in chapter 13). In this case Λµ is expressed in a simple way by the renormalization constant Z1, which is defined by ¯u(p) Λµ(p, p) u(p) = (Z−1 1 − 1) ¯u(p) γµ u(p). (271) Λµ(p, p) can be calculated from Σ(p), observing that Λµ(p, p) = − ∂Σ(p) ∂pµ , (272) this relation coming from ∂ ∂pµ 1 /p − m = − 1 /p − m γµ 1 /p − m . (273) (Compare (270) with (257).) With the aid of (272) Λµ(p, p ) can be split into a part with Z1, hiding the divergence, and a unique finite part Λc µ(p, p ), Λµ(p, p ) = (Z−1 1 − 1)γµ + Λc µ(p, p ). (274) Application of (272) to (265) yields ∂Σ(p) ∂pµ = −(Z−1 1 − 1) γµ, (note that C(p) = 0 for p = m), and in the sequel ¯u(p) Λµ(p, p) u(p) = (Z−1 2 − 1) ¯u(p) γµ u(p), which means Z1 = Z2, (275) so far up to order α. Like Z2, also Z1 could be plugged into a further charge renormalization. 62 18.2 Synopsis of renormalisations At a vertex of a Feynman diagram all the considered renormalisations meet. In the neighborhood there are the following diagrams, up to order e2 : (a) (b) (c) (d) (a) is the lowest (2nd order) graph, (b) - (d) show 4th order contributions. We consider the limit of small energy-momentum transfer, i. e. the limit of the photon four momentum q going to zero. The contributions corresponding to the graphs are the following: (a) −ieγµ (b) −ieγµ(Z−1 1 − 1) (c) 2ieγµ(Z−1 2 − 1) (d) −ieγµ(Z3 − 1) For the free photon line we divide by √ Z3, for each of the two free fermion lines by √ Z2. Altogether we obtain thus for the diagrams in the above figure the expression − ieγµ Z2 √ Z3 1 + (Z−1 1 − 1) − 2(Z−1 2 − 1) + Z3 − 1 . (276) The three expressions Z−1 1 − 1, Z−1 2 − 1, and Z3 − 1 are of order α. Now this is transformed, keeping only terms up to order α in every step. First we take out the factor 1 + (Z−1 1 − 1) from the square bracket: [1 + (Z−1 1 − 1)] 1 − 2(Z−1 2 − 1) − (Z3 − 1) 1 + (Z−1 1 − 1) . The counter of the fraction being already of order α, the denominator can be approximated by 1, leading to ≈ Z−1 1 [Z3 − 2(Z−1 2 − 1)]. 63 Then we take out the factor Z3, and because Z3 ≈ 1 + O(α), in first order (Z−1 2 − 1)/Z3 ≈ Z−1 2 − 1 : Z−1 1 Z3[1 − 2(Z−1 2 − 1)]. Up to order α the square bracket can be replaced by 1 [1 + (Z−1 2 − 1)]2 = Z2 2 . Collecting these terms we finally get −ie Z2 Z1 Z3 γµ (277) as the corrected vertex contribution. When Z1 = Z2, as we know it is in order α, then the renormalizations due to Z1 and Z2 cancel and all we get is −ieRγµ with eR = √ Z3 e, as it was obtained after handling the vacuum polarization. In the next chapter we will see that this is indeed the case in all orders. 19 The Ward-Takahashi identity In the last chapters we have studied the renormalization procedure in the lowest order, where divergences occur, that is in fourth order in e. In this order it was possible to hide all infinities in the electron’s/positron’s charge and mass. Particularly, two of three renormalization constants turned out to be equal. Generally, a theory is called renormalizable, if an approach of this kind works in all orders, more specifically, if a finite number of renormalization constants is sufficient. The appearance of new kinds of divergences that would require new renormalization constants in every order would spoil the predictive power of a theory. Concerning the relation Z1 = Z2 there is a general identity that extends its validity to all orders, the Ward-Takahashi identity. Furthermore it confirms that the term kµaν +kνaµ in the photon propagator does not contribute to amplitudes in the general case, when the photon does not necessarily couple immediately to free fermions. 64 To prove the Ward-Takahashi identity we consider arbitrary diagrams with at least one external photon with momentum k and denote the probability amplitude of the process by M(k). If we remove this photon, we get a simpler diagram, which contributes to a simpler amplitude M0. Inserting the photon somewhere else into the diagram gives a contribution to M(k), and summing over all diagrams that contribute to M0 and over all possible insertions in each of these diagrams gives the full amplitude M(k). The Ward-Takahashi identity applies for each diagram contributing to M0, once we sum over all insertion points. The external photon must attach either to a fermion line that runs out of the diagram to two external points, or to a closed fermion loop. 1) Fermion line with n vertices going to infinity. p pn−1 p3 p2 p1 p qn q3 q2 q1 The ingoing fermion momentum is p, the photon momenta are counted as ingoing, such that p1 = p + q1, . . . , pn = p = p + i qi. Now insert the photon with momentum k after the i-th vertex: qn qi+1 qi q1 k p + k pi+1 + k pi + k pi ppi−1pn−1 + k p1 At the vertex, where this photon couples to the fermion, we replace µ(k) by kµ, so that we obtain −iekµγµ = −ie[(/pi + /k − m) − (/pi − m)]. With this relation the expression for this vertex and the two adjacent fermion lines, represented by propagators, becomes i /pi + /k − m (−ie/k) i /pi − m = e i /pi − m − i /pi + /k − m . 65 Thus the diagram has a segment described by . . . i /pi+1 + /k − m γλi+1 i /pi − m − i /pi + /k − m γλi i /pi−1 − m γλi−1 . . . When we insert the photon at the position i−1, the corresponding expression is . . . i /pi+1 + /k − m γλi+1 i /pi + /k − m γλi i /pi−1 − m − i /pi−1 + /k − m γλi−1 . . . The first term in this expression cancels the second term of the previous one, and so on. In the sum over all possible insertions the unpaired terms at the ends survive. µ i kµ k q p = − q − k q p p + k e · The sum on the left-hand side is meant to be taken over all insertion points i. p + k has been relabeled as q. Obviously the right-hand side does not contribute to the transition amplitude M(k) for p → q. 2) Closed fermion loop. The left diagram of Fig. shows a closed fermion loop with n photons attached. In the right diagram a photon with momentum k is inserted between the positions i and i + 1. The momentum k exits at vertex 1 by convention. 66 pn +k p1 qn q2 p2 q1 qn pn q1 p1 q2 p2 q3p3 pi pi + k pi+1 + k k qi+1 pi−1 qi When we insert the photon between the vertices 1 and 2, we obtain the contribution −e d4 p1 (2π)4 Tr i /pn + /k − m γλn . . . i /p2 + /k − m γλ2 i /p1 − m − i /p1 + /k − m γλ1 . The first term will be canceled by one of the amplitudes coming from the insertion between 2 and 3, and so on. In the end two terms survive, −e d4 p1 (2π)4 Tr i /pn − m γλn i /pn−1 − m γλn−1 . . . i /p1 − m γλ1 − i /pn + /k − m γλn i /pn−1 + /k − m γλn−1 . . . i /p1 + /k − m γλ1 . After shifting the integration variable from p1 to p1 + k in the second term the two terms cancel. The diagrams with the photon inserted along a closed loop add up to zero. In the most general case there may be n ingoing and n outgoing fermions and an arbitrary number of further external photons. Graphically the WardTakahashi identity is shown in the following figure, 67 (q1 . . . qi . . . qn) µ k (p1 . . . pi . . . pn) Σ insertions = e Σi (q1 . . . qi − k . . . qn) (p1 . . . pi . . . pn) − (q1 . . . qi . . . qn) (p1 . . . pi + k . . . pn) in terms of a formula this is kµMµ (k, p1, . . . , pn, q1, . . . , qn) = e i [M0(p1, . . . , pn, q1, . . . , qi − k, . . . , qn) − M0(pi, . . . , pi + k, . . . , pn, q1, . . . , qn)] . The right-hand side does not contribute to the S-matrix. In the simplest case there is just one external fermion line, so that the left-hand side can be seen in fact as a full, renormalized vertex. µ i kµ k p + k p = − p p + k p p + k e · For full propagators S(p) = i /p−m−Σ(p) and a full vertex Γµ the diagram can be translated into S(p + k) [−iekµΓµ (p + k, p)] S(p) = e[S(p) − S(p + k)]. (278) 68 If we multiply by S−1 (inverse propagators = Dirac matrices) from the left and from the right, we obtain −ikµΓµ (p + k, p) = S−1 (p + k) − S−1 (p). (279) Sometimes this more special relation is called the Ward-Takahashi identity. From this we can find a relation between Z1 and Z2: In the limit k → 0 Γµ (p + k, p) → Z−1 1 γµ and S(p) → i Z2 /p − m . Expansion of (279) around k = 0 gives (recall the differential relation (272)) −ikµZ−1 1 γµ = −iZ−1 2 (/p + /k − m − /p + m), and from this follows Z1 = Z2 in all orders. In quantum electrodynamics no new types of divergences than the considered ones occur in higher orders; such theories are called renormalizable. 69