<<

The in the Low Energy Limit

Bachelor of Science Degree in Physics

Bj¨ornLinder

Supervisor Michael Bradley

Department of Physics Ume˚aUniversity June 2016 Abstract

The purpose of this work is to give an introduction to relativistic for the . First a review of the Dirac equation and its ap- plication to a free electron and an electron in an electromagnetic field will be given. In particular the atom is studied. The low energy limit and the lowest order corrections to non-relativistic are inves- tigated. These corrections include, among others, the of the electron and -orbit coupling in the . The low energy limit will be studied both in the usual Pauli-Dirac representation and by us- ing the Foldy-Wouthuysen transformation. The Pauli-Dirac representation is used to obtain the correction terms for the energy levels in the hydrogen atom. The Foldy-Wouthuysen transformation is applied to obtain the corresponding Hamiltonian corrections to order α4.

Sammanfattning

Syftet med det h¨ar arbetet ¨ar att ge en introduktion till relativistisk kvant- mekanik till¨ampad p˚aelektronen. F¨orst redog¨ors Dirac-ekvationen och dess till¨ampning hos en fri elektron och en elektron i ett elektromagnetiskt f¨alt. S¨arskilt studeras ekvationens till¨ampning p˚av¨ateatomen. Den ickerelativistiska gr¨ansen och de l¨agsta st¨orningstermerna behandlas. Dessa st¨orningstermer in- kluderar bland annat det gyromagnetiska kvoten hos elektronen och spin-orbit coupling i v¨ateatomen. Den icke-relativistiska gr¨ansen analyseras b˚adei Pauli- Dirac representationen och genom att anv¨anda Foldy-Wouthuysen transforma- tionen. Pauli-Dirac representationen anv¨ands f¨or att hitta st¨orningstermerna till energiniv˚aernai v¨ateatomen. Foldy-Wouthuysen transformationen till¨ampas f¨or att hitta de motsvarande st¨orningstermerna av ordning α4 i Hamiltonianen. Contents

1 Introduction 3

2 Background 4 Klein-Gordon Equation ...... 4

3 Dirac Equation 6 Deriving the Dirac Equation ...... 6 Conservation Equation ...... 8

4 Solutions for the Free 9 Stationary Particle ...... 9 Particle in ...... 10

5 Spin and Spin Orientation 13

6 Solutions for the Hydrogen Atom 16 Obtaining the Radial Equations ...... 16 Solving the Radial Equations ...... 18

7 Foldy-Wouthuysen Transformation 22 ...... 23 Electromagnetic ...... 26

8 Conclusions 30

Appendix A 32

Appendix B 34

2 1 Introduction

At the heart of quantum mechanics lies Erwin Schr¨odinger’sfamous differential equation ∂ ih Ψ rr, t Hˆ Ψ rr, t . ∂t The solutions obtained by solving̵ ( the) = equation( ) successfully predict the be- haviour of many quantum mechanical systems. A case which is of great im- portance and studied in any undergraduate course in quantum mechanics is the case of an electron moving in the Coulumb potential, more specifically the po- tential produced by a hydrogen nucleus. The solutions for this potential can be obtained exactly but the results fall short of describing the physical world. Experimental data show that there must be more to the story than what is obtained from Schr¨odinger’sequation with a Hamiltonian for an electromag- netic field. The way around this difficulty is to introduce correction terms to the Hamiltonian, corrections that would also adjust the predicted energies. The correction terms of order α4 (α being the fine structure constant) needed to explain fine structure splittings of the hydrogen spectrum are

1 p2 2 H , relativistic correction r 8 m3c2 (e ) 1 ∂Φ H = − S LL, ( spin-orbit correction) so 2m2c2 r ∂r eh2 H = 2Φ r ,⋅ ( Darwin term) d 8m2c2 ̵ where p, S and L =are the∇ operators( ) for ,( spin and) the azimuthal , respectively. h is Planck’s reduced constant, e is the elemen- tary charge and Φ is the electric field potential. These correction terms are often introduced with some hand-waving̵ arguments and might not be very easy to motivate. The first correction term can be motivated by the relativistic effects that the electron is subjected to and the second two terms can be motivated 1 by the fact that the electron is a spin- 2 particle. Spin is a property that the Schr¨odingerequation fail to describe and to do so one would turn to an ex- tension of the equation, the Pauli-Schr¨odingerequation. The Pauli-Schr¨odinger equation introduces 2-component solutions and is completely compatible with 1 the spin property of all spin- 2 . In the 1920’s a search for an equation unifying and quantum mechanics was taking place. An equa- tion from which, it would turn out, the spin property and all the corrections terms of order α4 came out naturally.

3 2 Background Klein-Gordon Equation The relativistic analog to Schr¨odinger’sequation would, unlike the Schr¨odinger equation, meet all the requirements of Einstein’s theory of special relativity. An equation today known as the Klein-Gordon equation was proposed as a candi- date to by O. Klein, W. Gordon and E. Schr¨odinger[9, p.115]. Even though the Klein Gordon equation fail to describe the electron, it is a good way to introduce the Dirac Equation.

The energy of a free particle in special relativity is given by the relation

E c2p2 m2c4. (2.1) » Searching for a relativistic analog= ± to the Schr¨odingerequation+ we could replace E and p with the corresponding operators from quantum mechanics, ∂ E ih , p ih . (2.2) ∂t Applying this energy → on̵ a function→ψ−gives̵∇ a differential equation, ∂ψ ih h2c2 2 m2c4ψ. (2.3) ∂t » However, the Taylor expansion̵ = of± the−̵ spacial∇ + derivative give derivatives to all orders up to infinity. The derivatives of space and time are hence of different orders and they cannot be symmetric, which would be required in special rela- tivity [9, p.118].

If one instead takes the square of equation (2.1) and then substitute the quan- tum mechanical operators of (2.2) we get

∂2ψ h2 h2c2 2ψ m2c4ψ. (2.4) ∂t2 This is known as the Klein-Gordon−̵ = −̵ equation∇ + for a free particle and it can take on a slightly more compact form by introducing the d’Alembertian ( 1 2 2 2 c2 ∂ ∂t ), 2 m2c2 ≡ ψ ψ. (2.5) ( ~ ) − ∇ h2 2 The solutions to the Klein Gordon equation for a free particle are the plane − = ̵ solutions ̵ ψ Aei(Et−p⋅x)~h, (2.6) where the energy E can have both positive and negative values as described by = eq. (2.1) and hence allows for two separate solutions [9, p.120]. The theory does, however, not include spin which and hence cannot fully describe the electron. However, the major problem with the Klein-Gordon equation appears when

4 studying the analog to the conservation equation of non-relativistic quantum mechanics [8, p.468]. Gathering the terms in eq. (2.5) and multiplying with the complex conjugate ψ∗ gives

m2c2 ψ∗ ψ ψ∗ ψ 0. (2.7) h2 2 By subtracting eq. (2.7) by its complex conjugate we get + ̵ = ψ∗ ψ ψ ψ∗ 0, (2.8) 2 2 which is equivalent to − = ∂ 1 ∂ψ ∂ψ∗ ψ∗ ψ ψ ψ∗ ψ∗ ψ 0. (2.9) ∂t c2 ∂t ∂t This is of the sameŒ formŒ as the− continuity‘‘ + ∇ equation ⋅ Œ ∇ of− non-relativistic∇ ‘ = quantum mechanics, ∂ρ j 0. ∂t If multiplied with a factor h 2mi the+ second ∇ ⋅ = term of eq. (2.9) is identical to the current density j in the classical case but the analog to ρ, which classically can be interpreted as the position̵~ probability, is not the same. Multiplying with the missing factor we get

ih ∂ψ ∂ψ∗ P xx, t ψ∗ ψ , (2.10) 2mc2 ∂t ∂t ̵ which allows for negative( probabilities.) = Œ This is− a fundamental‘ error and comes from the fact that the Klein-Gordon is a second order differential equation, with respect to time, in which the initial conditions for ψ and ∂ψ ∂t may be chosen independently [9, p.120]. It was mainly because of this feature the Klein-Gordon equation was rejected. ~

5 3 Dirac Equation Deriving the Dirac Equation Dirac wanted to find a first order equation describing the problem, a relativistic analog in which the wave functions would still satisfy the energy-momentum relation E2 p2c2 m2c4 of the Klein-Gordon equation. In Dirac’s exact words ”The general interpretation of non-relativity quantum mechanics is based on the transformation= + theory, and is made possible by the wave equation of the form H E ψ 0, i.e. being linear in E or ∂ ∂t, so that the wave function at any time determines the wave function at any later time” [4, p.612]. ( − ) = ~ ∂ Using the definition of the d’Alembertian and the short-hand notation ∂t ∂t equation (2.5) can be written as = 1 m2c2 2 ∂2 ψ ψ. (3.1) c2 t h2

One way to solve the problemŒ∇ of− the second‘ = order̵ derivatives would be to fac- torise the d’Alembertian operator into a square of some operator and apply this ∂ new operator once to the wave function. Introducing the notation ∂x ∂x, ∂ ∂ ∂y and ∂t, the factorisation ∂y ∂t = 2 = = 1 i 2 ∂2 A∂ B∂ C∂ D∂ , (3.2) c2 t x y z c t can work if the rightŒ∇ − requirements‘ = Œ on+ the coefficients+ + A,‘ B, C, D are chosen. They have to be anti-commutative so that the cross derivatives disappear and they also need to be their own inverse to preserve the second order derivatives. In other words we need that

AB BA 0,BC CB 0, etc. and A2 B2 C2 D2 1. (3.3)

The requirements+ = can+ be met= if the coefficients are= matrices.= = In fact= these matrices are closely related to the Dirac γ-matrices that Dirac defined in his slightly different derivation of the equation,

A iγ1,B iγ2,C iγ3,D γ0. (3.4) 1 The are 4= 4 hermitian= matrices= built= up of the of non-relativistic quantum mechanics, × 0 σ I 0 γk k and γ0 , (3.5) σk 0 0 I

0 1=  0 −i =1 0 1Defined as Œσ =  − σ =  σ =  −.‘ 1 1 0 2 i 0 3 0 −1

6 where I is the 2 2 identity . The 0’s in the matrices are to be interpreted as the 2 2 zero matrices. Some properties of the Dirac matrices are × ν µ µν × γ , γ + 2g I, (3.6a) γ0 2 I, γk 2 I, (3.6b) [ ] = γ0 † γ0, γk † γk, (3.6c) ( ) = ( ) = − ν µ ν µ µ ν µν where γ , γ + denotes the( anti-) = ( ) γ= −γ γ γ and g is the metric from special relativity. Introducing Einstein’s sum convention and the space- time notation[ ] from special relativity, xµ x0, x1, x+2, x3 ct, x, y, z , we can rewrite the factorised operator.2 We get that = ( ) = ( ) 2 2 i ∂ A∂ B∂ C∂ D∂ iγµ . (3.7) x y z c t ∂xµ Œ + + + ‘ = Œ ‘ Applying this factor once to a wave function should in accordance with (3.1) give the eigenvalue equation

∂ mc iγµ ψ ψ, (3.8) ∂xµ h Œ ‘ = which is the Dirac Equation for a free particle̵ in its manifest covariant form.3 There is another common way of expressing the Dirac Equation and to do so we notice that the gamma matrices are related by a common matrix factor β γ0. Hence they can be written as = 1 2 3 0 γ βα1, γ βα2, γ βα3, γ β. (3.9)

Substituting (3.9) into= the Dirac= Equation (3.8)= we get =

i mc iβα ∂ iβα ∂ iβα ∂ β∂ ψ ψ. (3.10) 1 x 2 y 3 z c t h

Multiplying withŒchβ on the+ left and+ rearranging+ the‘ terms= ̵ leaves ̵ mc hc iα ∂ iα ∂ iα ∂ β ψ ih∂ ψ, (3.11) 1 x 2 y 3 z h t ̵ −̵ where we have used− thatŒ β is+ an invertible+ matrix− ̵ with‘ =β 1 β. We can simplify further to get 2 cαα ih mc β ψ ih∂tψ = (3.12) or equivalently ( ⋅ (− ̵∇) + ) = ̵ cαα p mc2β ψ Eψ. (3.13)

2 See Appendix A for a short introduction( ⋅ + to the) metric= and the notation of special relativity. 3There are different ways of representing the gamma matrices, which causes the Dirac Equation to take on different forms. This representation is based on notation by Bjorken, Drell and Schweber.

7 This equation is of the same form as the Schr¨odingerequation, Hψˆ Eψ, with the Hamiltonian given by Hˆ cαα p mc2β. = (3.14)

The hermitian matrices β and α are= ⋅ +

0 I 0 0 σk 0 k β γ , αk γ γ k 1, 2, 3 , (3.15) 0 I σk 0 = =  =  = ( = ) and satisfy −

α2 α2 α2 β2 1, 1 2 3 (3.16) α , β 0, β2 1, α , α 2δ , k + = =k l =+ =kl where αi, αj + is[ the notation] = for the= anticommutator[ ] =αiαj αjαi and δkl is the Kronecker delta function. [ ] + Conservation Equation To obtain a conservation equation we first multiply (3.12) on the left with the ψ†

† 2 ψ ih∂tψ cαα ih ψ mc βψ 0. (3.17)

Introducing the hermitian‰ ̵ adjoint+ ⋅ version( ̵∇ ) of− (3.12) andŽ = multiplying it on the right with ψ we get

† † 2 † ih∂tψ cαα ih ψ mc βψ ψ 0, (3.18) where we have used‰ − that̵ the− matrices⋅ ( ̵∇ α) −and β areŽ hermitian.= Taking the difference of them and setting it equal to zero gives us the equation

† † † † ψ ∂tψ ∂tψ ψ c ψ αk∂xk ψ αk ∂xk ψ ψ 0, (3.19)

Œ + ( ) ‘ + Œ + ( ) ‘ = which can be interpreted as a conservation equation if ρ ψ†ψ and the current † density jk ψ αk∂kψ [8, p.477]. = =

8 4 Solutions for the Free Particle

Looking at the Hamiltonian Hˆ cαα p mc2β, which is a 4 4 matrix, it must be so that any wave-function ψ satisfying the Dirac equation must be a four element row vector, = ⋅ + × ψ1 xx, t ψ2 xx, t Ψ xx, t ⎡ ⎤ . (4.1) ⎢ψ3(xx, t)⎥ ⎢ ⎥ ⎢ (x )⎥ ( ) = ⎢ψ4 xx, t ⎥ ⎢ ( )⎥ This 4-element vector solution is called⎢ a bispinnor⎥ or a Dirac . Solutions ⎢ ( )⎥ for a particle not subjected to any potential⎣ can⎦ be found by making the ansatz

̵ Ψ xx, t u p ei[(p⋅x~h)−(Et~h)], (4.2) p where u p is a 4 1 column( vector.) = ( ) Stationary( ) Particle× The most simple solution to the Dirac equation is found in the case of a free, stationary particle with p 0. The Hamiltonian then reduces to Hˆ mc2β and the equation becomes = ∂ψ = ih mc2βψ. (4.3) ∂t ̵ −imc2t~h̵ 2 The plane-wave ansatz reduces down to= uj 0 e (if E mc ) and plug- ging it into the Dirac equation we get the system of equations ( ) = + mc I 0 u mc u A A , (4.4) h 0 I uB h uB   =  where the uj’s are split into̵ the− two 2 1 column̵ vectors uA and uB. This requires that uB (u3 and u4) is zero. However, if instead the was chosen uB (u1 and u2) has to vanish.× For either of these cases the solutions for uA and uB are the Pauli

1 0 χ or χ . (4.5) 1 0 2 1 =  =  Hence there are four solutions, with two corresponding to positive energy and two corresponding to negative energy [7, p.89-90]. The solutions are

1 0 0 0 0 −imc2t~h̵ 1 −imc2t~h̵ 0 imc2t~h̵ 0 imc2t~h̵ ⎡ ⎤ e , ⎡ ⎤ e , ⎡ ⎤ e , ⎡ ⎤ e . (4.6) ⎢0⎥ ⎢0⎥ ⎢1⎥ ⎢0⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢0⎥ ⎢0⎥ ⎢0⎥ ⎢1⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎣ ⎦ ⎣ ⎦ ⎣ ⎦ ⎣ ⎦

9 Particle in Motion We could now start looking for solutions for the free particle with p 0. The following derivation is based on a derivation made by Franz Schwabl in his book Advanced Quantum Mechanics [9, p.147-149]. We start by introducing> the 4-position vector x and 4-momentum vector k.

x x0, x1, x2, x3 ct, x, y, z ct,xx (4.7a) k k0, k1, k2, k3 E c, p , p , p E c,kk (4.7b) ≡ ( ) = x( y z ) = ( ) Assuming that the≡ solutions( will be) = in( accordance~ with) = the( ~ stationary) particle we will try to find independent solutions corresponding to the positive and negative energies. The four 4-component solutions are assumed to be in the form

(+) (+) −ik⋅x~h̵ (−) (−) ik⋅x~h̵ Ψ ur k e , Ψ ur k e , r 1, 2 . (4.8)

We will also introduce= ( ) the Feynman slash= notation( ) ( = )

µ ν γ ν γ νµ γ0ν0 γ νν. (4.9)

The Dirac Equation eq. (3.8)~ ≡ can⋅ = then be= written− in⋅ the compact form

mc i∂ ψ 0. (4.10) h Œ ~ − ‘ = (+) (−) ̵ Plugging ψr and ψr into the Dirac Equation (4.10) we get the two equations

(+) (−) k mc ur k 0 k mc ur k 0. (4.11)

The two factors are(~ − each) others( ) conjugate,= (~ + ) ( ) = k mc k mc kk mc 2, (4.12) and if kk is investigated( more~ − closely)(~ + we) find= ~ ~ that− ( )

~ ~ µ ν 1 µ ν µν kk k γ k γ k k γ , γ + k k g . (4.13) µ ν µ ν 2 µ ν This is just the invariant~ ~ = inner product= of[ the 4-momentum,] = E2 k kµ k2 m2c2. (4.14) µ c2 Substituting (4.14) into eq. (4.12)= we see− that= it is equal to zero. Hence equa- (+) (−) tions (4.11) are fulfilled if the coefficients ur and ur contain the corresponding conjugate factor. The solutions are simply the coefficients from the free particle at rest times these factors

(+) (−) ur k N k mc ur 0 ur k N k mc ur+2 0 (4.15)

( ) = (~ + ) ( ) ( ) = (~ − ) ( ) 10 Once again introducing the Pauli spinors we can write the equations for the coefficients as

χr 0 ur k N k mc and vr k N k mc , (4.16) 0 χr ( ) = (~ + )  ( ) = (~ − )  where the 0’s are abbreviations for the 2 1 zero matrices. The calculations are straightforward, × χ I 0 0 σ χ k r k0 ki i r 0 0 I σi 0 0

~  = Œ 0  i −  E ‘  = k I k−σi χr − c χr i 0 k σi k I 0 σ kkχr − =   =  0 I 0 0 ⋅ σ 0 k k0 ki i χr 0 I σi 0 χr − −~  = Œ 0  i +  ‘  = k I k σ+i 0 − k σσχr i 0 E . k σi k I χr c χr − ⋅ =   =  This give the coefficients − E k σ (+) c mc χr (−) k σσχr ur k N and ur k N E . (4.17) σ kkχr c mc χr ( + ) ⋅ ( ) =  ( ) =  Using the fact that ⋅ ( − ) p p ip σ k z x y (4.18) px ipy pz − ⋅ =  we get the solutions + − E mc2 c 0 2 (+) 0 (+) E mc c u1 N ⎡ ⎤ , u2 N ⎡ ⎤ , (4.19a) ⎢( +pz )~ ⎥ ⎢ px ipy ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢( + )~ ⎥ = ⎢ px ipy ⎥ = ⎢ pz ⎥ ⎢ ⎥ ⎢ ( − ) ⎥ ⎢ p ⎥ ⎢ p ip ⎥ ⎢ ( +z ) ⎥ ⎢ x− y ⎥ (−) ⎣ px ipy ⎦ (−) ⎣ pz ⎦ u N ⎡ 2 ⎤ , u N ⎡ ⎤ . (4.19b) 1 ⎢ E mc c⎥ ⎢ −( 0− ) ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ( + ) ⎥ ⎢ − 2 ⎥ = ⎢ 0 ⎥ = ⎢ E mc c⎥ ⎢ ⎥ ⎢ ⎥ ⎢( − )~ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ The solutions are more⎣ commonly⎦ written as ⎣( − )~ ⎦

11 1 0 ′ 0 ′ 1 u1 N ⎡ 2 ⎤ , u2 N ⎡ 2 ⎤ , ⎢ cpz E mc ⎥ ⎢c px ipy E mc ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ 2 ⎥ ⎢ 2 ⎥ = ⎢c px ipy E mc ⎥ = ⎢ cpz E mc ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ~( + ) ⎥ ⎢ ( − )~( + (4.20a))⎥ ⎢ ⎥ ⎢ ⎥ ⎣ ( + )~( + 2 )⎦ ⎣ − ~( + ) 2⎦ cpz E mc px ipy E mc 2 2 ′ c px ipy E mc ′ cpz E mc v1 N ⎡ ⎤ , v2 N ⎡ ⎤ . ⎢ − ~(S 1S + ) ⎥ ⎢−( − )~(S0 S + )⎥ ⎢ ⎥ ⎢ ⎥ ⎢− ( + )~(S S + )⎥ ⎢ ~(S S + ) ⎥ = ⎢ 0 ⎥ = ⎢ 1 ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ (4.20b)⎥ ⎢ ⎥ ⎢ ⎥ ⎣ ⎦ ⎣ ⎦ (−) We have used that the energy in the ur solutions have a negative sign relative (+) (−) (+) to the energies in ur . It is straightforward to show that the u ’s and u ’s are orthogonal to each other for a given p as well as

† ′ ur p ur′ p 0 for r r (4.21)

When it comes to the normalisation( ) ( there) = are two≠ different approaches [7, p.93]. The first is that † ur p ur p 1 (4.22) which implies that ( ) ( ) = N ′ E mc2 2 E . (4.23) » The second convention is that = S S + )~ S S † 2 ur p ur p E mc , (4.24) which gives ( ) ( ) = S S~ N ′ E mc2 2mc2. (4.25) » = (S S + )~

12 5 Spin and Spin Orientation

For either sign of the energy there are two different solutions. We introduce the 4 4 spin operator h h σ 0 S Σ , (5.1) × 2 2 0 σ ̵ ̵ where σ σ1, σ2, σ3 are the Pauli≡ spin≡ matrices. In this sense it is really the ˆ 4 4 extension of the classical quantum mechanic spin operator S c h 2 σ. For a particle= ( with p) 0, which was the first of the two cases we solved in the× previous section, the different solutions correspond to different spin= ( states,̵~ ) as they are eigenspinors= to the operator Sˆ3. Applying the operator to these solutions we get the eigenvalues h h Sˆ u 1 u , Sˆ u 1 u , (5.2) 3 1 2 1 3 2 2 2 ̵ ̵ h h Sˆ v = (+1) v , Sˆ v = (−1) v . (5.3) 3 1 2 1 3 2 2 2 ̵ ̵ The first solutions for both= signs(+ ) of the energy= hence(− ) correspond to spin 1 and the second solutions of each set correspond to spin 1. For the free particle with p 0 this is not true and applying the operator Sˆ3 to the solutions+ will not give an equivalent eigenvalue [7, p.92-93]. It will− be possible if we assume that the≠ momentum is parallel to the spin, which can be seen as following. If the momentum is along the z-axis we get

1 0 0 0 1 ˆ h 0 1 0 0 ′ 0 h S3u1 p pz ⎡ ⎤ N ⎡ 2 ⎤ 1 u1, 2 ⎢0 0 1 0 ⎥ ⎢cpz E mc ⎥ 2 ⎢ ⎥ ⎢ ⎥ ̵ ⎢ − ⎥ ⎢ ⎥ ̵ ( = ) = ⎢0 0 0 1⎥ ⎢ 0 ⎥ = (+ )( ) ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ~( + )⎥ ⎢ ⎥ ⎢ ⎥ ⎣ − ⎦ ⎣ ⎦ 1 0 0 0 0 ˆ h 0 1 0 0 ′ 1 h S3u2 p pz ⎡ ⎤ N ⎡ ⎤ 1 u2. 2 ⎢0 0 1 0 ⎥ ⎢ 0 ⎥ 2 ⎢ ⎥ ⎢ ⎥ ̵ ⎢ − ⎥ ⎢ 2 ⎥ ̵ ( = ) = ⎢0 0 0 1⎥ ⎢ cpz E+ mc ⎥ = (− )( ) ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ Likewise the eigenvalues⎣ are h 2 and− ⎦ h 2⎣− for v~(1 and+v2 respectively.)⎦ Hence we can define the helicity operator ̵~ −̵~ p Λ Σ , (5.4) p ≡ ⋅ of which the for a free particleS S in motion is an eigenfunction [7, p. 93]. In other words the spinors are eigenspinors to the spin operator in the direction of the momentum. The states corresponding to 1 eigenvalues as right-handed states and the states corresponding to 1 eigenvalues as left- handed states. The right-handed states are those with motion(+ ) and spin in the (− ) 13 same direction and the left-handed states are those with motion and spin in opposite directions.

h̵ It can be shown that neither of the operators L x p and S 2 Σ commute with the Hamiltonian so that in Heisenberg’s picture they are not constants of 4 ≡ × ≡ h̵ motion and hence not conserved. The total J x p 2 Σ is however a constant of motion in the presence of a spherically symmetrical po- tential and in the case of no potential [7, p.122]. In = × we+ can specify the orientation of the spin relative to the total angular momentum with the operator σ J . In solving the Dirac equation for a hydrogen potential it will be of interest to have a relativistic analog to such an operator. We introduce the operator βΣ⋅ J which is quite similar to the non-relativistic operator. Its commutation relation with the free Hamiltonian (equation 3.14) is ⋅ h H, βΣ J 2cβαα p (5.5) 2 ̵ [7, p.122]. We also have the[ following⋅ ] = commutation(− ⋅ ) relation, which is easier to verify, H, β cαα ppβ βcαα p 2cβαα pp. (5.6)

Based on these two equation[ ] = a new⋅ operator− ⋅ K= −that commutes⋅ with H can be introduced, h K βΣ J β β Σ L h . (5.7) 2 ̵ K does not only commute≡ with ⋅H −but it= also( commutes⋅ + ̵) with J , since it com- mutes with the two matrix terms of J , β and Σ L. Because of this we can 2 construct functions that are eigenfunctions of H,K,JJ and Jz. Further impor- tant relations between the eigenvalues k and j j ⋅1 can be derived. We have that ( + ) K2 β Σ L h β Σ L h Σ L h 2 (5.8) Σ L 2 2 Σ L h h2. = ( ⋅ + ̵) ( ⋅ + ̵) = ( ⋅ + ̵) To progress from here we will introduce= ( ⋅ the) fifth+ ( gamma-matrix⋅ )̵ + ̵

0 I γ5 iγ0γ1γ2γ3, (5.9a) I 0 ≡  = which anti-commutes with the other gamma-matrices

5 µ γ , γ + 0. (5.10a)

4 See Appendix B for more on Heisenberg’s[ picture] = and the derivation of this.

14 The components of Σ can then be expressed as (inspired by [2], Solution of the Dirac Equation for Hydrogen)

5 5 0 1 2 3 Σ1 γ α1 γ γ γ i γ γ , Σ γ5α γ5 γ0γ2 i γ3γ1 , (5.11) 2 = 2 = ( ) = − ( ) Σ γ5α γ5 γ0γ3 i γ1γ2 . 3 = 3 = ( ) = − ( ) This is verified by using the= associative= ( property) = − of( matrix) multiplication, taking into consideration that the gamma-matrices are anti-commutative and ’flipping two over’ at a cost of changing sign. Because of (5.11) we can write

ΣiΣi 1, (5.12a)

ΣiΣj iijkΣk. (5.12b) = where  is the Levi-Civita symbol.5 Using= this and that L L ihLL [7, p.123] we can rewrite the first term of eq. (5.8) as × = ̵ 2 Σ L ΣiLiΣjLj δij iijkΣk L2 iΣ L L ( ⋅ ) = ( + ) L2 hΣ LL. = + ⋅ ( × ) The operator can now be written as = − ̵ ⋅ K2 L2 hΣ L h2. (5.14)

We will now make use of that the= equation+ ̵ ⋅ + above̵ is very similar to another operator, namely J 2.

J 2 L2 hΣ L 3h 4 (5.15) K2 h2 4 = + ̵ ⋅ + ̵~ and in other words the eigenvalues must= be connected− ̵ ~ as 2 h 1 2 k2h2 j j 1 h2 j h2. (5.16) 4 2 ̵ The eigenvalues of the K̵ operator= ( + must)̵ + therefore= ‰ + beŽ nonzero̵ 1 k j . (5.17) 2 The sign of these eigenvalues determine= ±‰ whether+ Ž the spin of the particle is par- allel or anti-parallel to the angular momentum. A negative sign implies an anti-parallel state and a positive sign a parallel state [7, p.123].

5The Levi-Civita symbol takes on the values +1, −1 or 0. It will be +1 if the sub-indexes are in an cyclic order, for example 123 or 312, and −1 for the odd permutations of 123. Levi-Civita therefor has the anti-symmetric property that interchanging two indexes changes it’s sign. If any two indexes are the same it takes on the value zero.

15 6 Solutions for the Hydrogen Atom

The following derivation is based mainly on a derivation made by J.J. Sakurai in his book Advanced Quantum Mechanics [7, p.122-131]. The Hydrogen atom is the only atomic potential for which the Dirac equation can be solved exactly.

In the presence of a spherically symmetrical electromagnetic potential (with no magnetic vector potential, A 0) the Hamiltonian takes the form

2 H cα=α p βmc V r (6.1) where the potential that an electron= ⋅ + is subject+ ( to) in a hydrogen atom is well known to be e2 V r . 4π0r − ( ) = Obtaining the Radial Equations

Splitting the Dirac spinor Ψ into two 2 1 spinors ψA and ψB and applying the operator σ× L h 0 K β Σ L h (6.2) 0 σ L h ⋅ + ̵ introduced in the previous= ( section⋅ + ̵) gives=  two equations − ⋅ + ̵

σ L h ψA khψA, σ L h ψB khψB. (6.3)

At the same time,( ⋅ applying+ ̵) the= − operator̵ J(2 of⋅ which+ ̵) ψ is= also̵ an eigenfunction gives the equations

2 2 2 1 3 J ψA,B L hσσ 2 ψA,B j j 1 h ψA,B , j 2 , 2 , ... (6.4) From this we can= ( conclude+ ̵ ~ ) that= while( +ψ)is̵ not an eigenfunction= of L2, the components ψA and ψB can be, as can be seen by expanding the middle term in (6.4). Using the unitary property of the Pauli matrices, σiσi I, we can through (6.4) also obtain that when L2 acts on a two-component state vector 2 2 2 it is equivalent to J hσσ L 3h 4. Applying the operator L to=ψA,B must therefore give the relations − ̵ ⋅ − ̵ ~ 1 1 l l 1 j j 1 k and l l 1 j j 1 k , (6.5) A A 4 B B 4 if the eigenvalues( + ) = of( L+2,)l +l 1+, are assumed to( have+ ) a= factor( + h)2−. From+ equations (6.5) and (5.17) it is possible to determine the two possible values of lA and lB for a given k. ( + ) ̵ 1 1 1 k j lA j , lB j . 2 2 2 (6.6) 1 1 1 k =j + Ô⇒ l = j + , l = j − . 2 A 2 B 2 = −( + ) Ô⇒ = − = + 16 In other words the value of lA and lB for a given j differ with a value of 1. We make an ansatz for the components of the Dirac spinor

g r Y mj ψA jlA Ψ mj (6.7) ψB if r Yjl , ( ) B =  =  where the i is there to make the solution( real) and the functions Y are linear combinations of the spherical harmonics

1 1 l mj l mj Y mj 2 Y mj −1~2χ 2 Y mj +1~2χ jl ¾ 2l 1 l 1 ¾ 2l 1 l 2 (6.8) + + − + 1 = + for j l 2 , + + 1 ‰ 1 = + Ž l mj l mj Y mj 2 Y mj −1~2χ 2 Y mj +1~2χ jl ¾ 2l 1 l 1 ¾ 2l 1 l 2 (6.9) − + + + 1 = + for j l 2 , + + where χ1 and χ2 are the Pauli spinors as defined in (4.5).‰ The= − superindexŽ mj is the eigenvalue of the operator Jz calculated from h J ψ L σ ψ m hψ , m j, ..., j. (6.10) z A,B 3 2 3 A,B j A,B j ̵ =m(j + ) = ̵ = 2− 2 2 The functions Yjl are normalized eigenfunctions of J ,LL ,Jz and S as re- quired. To continue we look at the of the functions we have introduced. For a given k the orbital parities of Y mj and Y mj are opposite. The spherical jlA jlB harmonics transform under a parity inversion as [9, p.170]

mj l mj Yjl x 1 Yjl x (6.11) and hence, looking at (6.6), we can(− ) see= ( that− ) ψA and( ) ψB correspond to opposite σ⋅x parities. The operator r flips the parity of the spherical harmonics and hence it changes l. It turns out that σ x Y mj Y mj . (6.12) r jlA jlB ⋅ The square of the operator (i.e. applying= the− operator twice) reduce down to 1. In other words the operator will change l between lA and lB, which will prove useful.

Writing the Hamiltonian (6.1) in its matrix form [9, p.172]

mc2 V r σ p Hˆ , (6.13) σ p mc2 V r − ( ) ⋅ =  ⋅ − − ( )

17 and substituting the Dirac spinor ψ into the Dirac equation we get the equations

2 c σ p ψB E V r mc ψA (6.14a) c σ p ψ E V r mc2 ψ (6.14b) ( ⋅ ) A = ( − ( ) − ) B σ⋅x The parity operator r (comes⋅ ) in= by( rewriting− ( ) + the scalar) product σ x σ x ∂ σ p σ x σ p ihr iσσ L . (6.15) r2 r2 ∂r ⋅ ⋅ ̵ Substituting this⋅ as= well as( the⋅ )( 2 ⋅1) component= Œ − vectors+ for⋅ψA‘ and ψB we can calculate the left-hand sides of equations (6.14), × df mj 1 k h mj σ p ψB h Y fY , (6.16a) dr jlA r jlA ̵ dg (1− k)h ⋅ ̵ mj mj (σ ⋅ p)ψA = i−h Y −i gY . (6.16b) dr jlB r jlB ( + )̵ If substituted back into( ⋅ (6.14)) = the̵ combinations+ of spherical harmonics can be removed, leaving just df 1 k hc ch f E V mc2 g, (6.17a) dr r ̵ ̵ dg (1− k)hc ch + g = −E( −V −mc2 f.) (6.17b) dr r ( + )̵ For convenience we introduce̵ + F r rf= r( and− G+ r ) rg r and arrive at the radial equations ( ) = ( ) ( ) = ( ) dF k hc F E V mc2 G, (6.18a) dr r ̵ ŒdG − k ‘ = −( − − ) hc G E V mc2 F. (6.18b) dr r ̵ Œ + ‘ = ( − + ) Solving the Radial Equations To solve for the radial functions we introduce the short-hand notation [9, p.172]

2 2 α1 mc E hc, α2 mc E hc, σ α1α2, (6.19) ρ rσ, γ Ze2 4πhc√ Zα, = ( + )~̵ = ( − )~̵ = and substitute into (6.18) to acquire = = ( ~ ̵ ) = d k α γ F 2 G 0, (6.20a) dρ ρ σ ρ Œ d + k ‘ − Œα − γ ‘ = G 1 F 0. (6.20b) dρ ρ σ ρ Œ − ‘ − Œ + ‘ = 18 This couple of equations can be solved by applying the differential operator d dρ in one of the equations and then substitute the result into the other [9, p.173]. It can then be found that for large ρ’s ~ d2F d2G F, G (6.21) dρ2 dρ2 In other words the solutions will= behave as exponential= functions e±ρ, with e−ρ being the normalisable solution. This motivates the power series

−ρ −ρ s m −ρ −ρ s m F e f ρ e ρ amρ ,G e g ρ e ρ bmρ , (6.22) m=0 m=0 which= when substituted( ) ≡ Q into (6.20) turn the= equations( ) ≡ into Q kf α γ f ′ f 2 g 0, (6.23a) ρ σ ρ − + kg − Œα − γ ‘ = g′ g 1 f 0. (6.23b) ρ σ ρ Further conclusions about the− − power− Œ series+ can‘ be= drawn by looking at the coefficients in front of ρs+ν−1 after the series have been inserted into (6.23). The coefficients are α2 s ν k b b − γa a − 0, (6.24a) ν ν 1 ν σ ν 1 α1 (s + ν + k)a − a − + γb − b − = 0. (6.24b) ν ν 1 ν σ ν 1 which gives a recursive( + relationship− ) − between− them.− Setting= ν 0 we get

s k b0 γa0 0, = (6.25) s k a0 γb0 0, ( + ) + = which has non-zero solutions if( the− ) − = is zero. Which is to say that s k2 γ2. (6.26) » To determine which sign is most= suitable± − we take into consideration that the functions need to be normalisable. This can only be achieved if we take the positive root of (6.26).

Using the recursive relationship between the coefficients it can be shown that the series converge to e2ρ [8, p.486]. Because of this the series need to be finite i.e. there must exist an n′ such that

an′+1 bn′+1 0, an′ 0, bn′ 0. (6.27) ′ Setting ν n 1 in (6.24) with= the= termination≠ assumption≠ in (6.27) gives the ratio = + a ′ α n 2 . (6.28) bn′ ¾α1 = − 19 If we multiply (6.24a) by σ and (6.24b) by α2, take the difference between the ′ two and set ν n we can obtain an equation involving just an′ and bn′ . This gives ′ = 2σ s n γ α1 α2 γE (6.29) and substituting back everything we defined in (6.19) we get ( + ) = ( − ) = mc2 2 E2 s n′ γE (6.30) » and finally ( ) − ( + ) = mc2 mc2 E . (6.31) 2 2 2 γ ¼ Z α 1 ′ 2 1 (s+n ) (n′+ (j+ 1 )2−Z2α2)2 = ¼ = ½ 2 This is the energy levels for+ an electron in+ an hydrogen atom solved exactly, no simplifying assumptions were made along the way. In this form it might be hard to draw any conclusions, but it is easily checked that there are no variables in (6.31) that we aren’t familiar with. We can make a definition for the principal quantum number, 1 n n′ k n′ j , (6.32) 2 and then write ≡ + S S = + + 2 −1~2 Zα E mc2 1 (6.33) 1 1 2 2 n j 2 j 2 Zα =  + Œ ¼ ‘ which makes it more clear that− ( +n corresponds) + ( + ) to− the( n) from non-relativistic quantum mechanics. Eq. (6.33) is the exact fine-structure formula for hydrogen and if expanded in terms of α to order α4 we obtain the energy levels with fine structure corrections as obtained from perturbation theory, E α2 n 1 3 E mc2 1 1 , (6.34) n2 n2 j 2 4 = −  + Œ + − ‘ where E1 is the energy of hydrogen as calculated with Schr¨odinger’s equation [6, p.275-276]. There is also the term corresponding to the rest energy, which is usually not considered in the non-relativistic case. Except for this rest energy term, this is of course the sum of the energy levels and correction terms obtained for the fine structure [6, p.270-274]

E En Efs, (6.35a) E 2 4n E n 3 = E + E , (6.35b) fs 2mc2 j 1 2 r so ( ) E= 2 n jŒ j − 1 l l‘ =1 3+ 4 E n + ~ , (6.35c) so mc2 l l 1 2 l 1 ( ) [ ( + ) − ( + ) − ~ ] =  E 2 4n E ( + n~ )( + ) 3 , (6.35d) r 2mc2 l 1 2 ( ) = −  − + ~ 20 2 where En is the energy levels E1 n . Eso is the correction from spin-orbit 6 coupling and Er is the relativistic correction. To obtain the corresponding cor- rection operators to the Hamiltonian− ~ we will have to take a different approach, this is done in the next section.

6One could also introduce the energy correction from the Darwin term to deal with the difficulties arising with l = 0 for Er and Eso. The sum of the two correction is however correct for any l so this will be omitted [6, p.274].

21 7 Foldy-Wouthuysen Transformation

The solutions to the Dirac spinor ψ we have obtained in the earlier sections are built up of components that correspond to positive energy solutions and to negative energy solutions. To get a meaningful classical limit of the Dirac 1 equation we could wish for it to converge to the Pauli theory of spin- 2 particles, where the states are described by two-component wave functions.7 One way of doing this is making use of the fact that in the non-relativistic limit two out of the four components of each spinor becomes negligible. If the small component equations are solved and substituted into the equations for the large components one approximately gets the Pauli spin equations [5, p.29]. However, according to Foldy and Wouthuysen, with this method there are problems. All four spinor components still need to be used when calculating expectation values to the v2 2 order c2 α and there are difficulties concerning the operators used in the two theories. In Dirac Pauli representation the velocity operator, x˙ cαα, has eigenvalues≈ c (since the components of α has eigenvalues 1).8 First of all this does not agree with the physical world. Secondly, it does not agree with= the ve- locity operator± of Pauli theory where the operator is p mc2±and has eigenvalues ranging from to [3, p.3]. In the Dirac theory the components of velocity do not even commute, while as in Pauli theory they do.~ −∞ ∞ Foldy and Wouthuysen showed in their famous paper [5] that for a particle behaving according to the Dirac equation there exist another representation in which the positive and negative energy states are described by separate two- component wave functions, true for both relativistic and non-relativistic ener- gies. For the free electron this representation can be obtained exactly and for the electromagnetic potential it can be done in a series to any order. Foldy and Wouthuysen also showed that in this representation there exists a position op- erator whose time derivative can be interpreted as a velocity operator which in the non-relativistic limit agrees with the one found in the Pauli representation [5, p.30]. Further, in representation that we have studied, the Dirac-Pauli rep- resentation, we’ve seen that L and S are not constants of motion. This required us to work with the total angular J . Foldy and Wouthuy- sen found that in their representation both the orbital angular momentum and the spin are constants of motion [3, p.3].

7This was one of the problems with the Dirac Equation according to Foldy and Wouthuysen in their paper On the Dirac Theory of Spin 1/2 Particles and Its Non-Relativistic Limit [5]. 8A short derivation of this operator can be found in Appendix B.

22 Free particle Adapting to as in the original paper we set h c 1. The trans- formation is presented following [5, p.30-31]. The canonical (unitary) transfor- mation that achieves the required decoupling is ̵ = =

ψ′ eiSψ, (7.1) where S is not necessarily time-independent.= The transformation of the Hamil- tonian can be derived from the Dirac equation Eψ Hψ. Transforming the left-hand side we get = −iS ′ −iS ′ −iS ′ Eψ i∂tψ i∂t e ψ i ∂te ψ ie ∂tψ . (7.2)

Transforming the= right-hand= ( side we) get= ( ) + ( )

Hψ He−iSψ′ (7.3)

Setting (7.2) equal to (7.3), multiplying= on the left with eiS and rearranging we find that ′ iS −iS −iS ′ i∂tψ e He ∂te ψ (7.4)

The transformed Hamiltonian= can‰ then be− written( ) asŽ ′ iS −iS H e H i∂t e . (7.5)

Foldy and Wouthuysen proceed= to define( − ) βαα p S i 2 tan−1 p m . (7.6) p ⋅ Now the choice of S may seem= −( ~ mysterious,) especially( ~ ) with the trigonometric function. The explanation lies in the fact that the goal is to construct S in such a way that H′ contains no operators that mix the large and small components of the Dirac spinor. The matrix operators that do this will be referred to as odd operators. For example the α and γ matrices (except γ0) cause this mixing and are hence odd. Even matrices on the other hand do not cause this difficulty, like the β γ0 and Σ matrices [9, p.181]. The S defined in (7.6) does the job, it removes all the the odd operators of the Hamiltonian. = A slightly more explanatory approach is taken by Drell & Bjorken in Relativistic Quantum Mechanics [1, p.45-48]. They define

S iβαα ppθ eiS eβαα⋅ppθ. (7.7)

If we expand the exponential≡ − function⋅ in→ terms= of θ we obtain θ2 θ3 eβαα⋅ppθ 1 βαα p θ βαα p 2 βαα p 3 ..., (7.8) 2 6 = + ( ⋅ ) + ( ⋅ ) + ( ⋅ ) + 23 which can be simplified using the relations

α p 2 αiαjpipj p2, (7.9a) βαα p 2 β2 α p 2 p2. (7.9b) ( ⋅ ) = = We then get that ( ⋅ ) = − ( ⋅ ) = − θ2 θ3 θ4 eiS 1 βαα p θ p2 p2 βαα p p4 ... (7.10) 2 6 24 If we compare the= result+ ( of⋅ (7.10)) − with− the series( ⋅ of) the+ trigonometric+ functions it is straightforward to verify that we have βαα p eiS cos pθ sin pθ (7.11) p ⋅ Similarily we have for the inverse= ( ) + ( ) βαα p e−iS cos pθ sin pθ . (7.12) p ⋅ Substituting these results into= the( Hamiltonian) − transformation( ) (7.5) we obtain

βαα p βαα p H′ cos pθ sin pθ H cos pθ sin pθ , p p ⋅ ⋅ since S is time-independent= Œ ( ) + [9, p.182].( )‘ TheŒ free( ) particle− hamiltonian,( )‘ H α p mβ, anticommutes with the factor in front of the sin pθ term (see eq. (3.16)) so that = ⋅ + ( ) βαα p βαα p H′ cos pθ sin pθ H cos pθ sin pθ p p ⋅ ⋅ = Œ ( ) + ( )‘ Œ ( ) − H e−2βα(α⋅ppθ)‘ βαα p βαα p α p cos 2pθ sin 2pθ βm cos 2pθ = ‰sin 2pθ Ž p p ⋅ ⋅ Performing= ⋅ Œ the matrix( ) − multiplication( )‘ and+ onceŒ again( keeping) − in mind( the)‘ com- mutation relations α p H′ α p cos 2pθ βp sin 2pθ βm cos 2pθ m sin 2pθ p m p⋅ = α ⋅ p cos( 2pθ) + sin( 2pθ) + βm cos( 2)pθ− sin 2(pθ ). p m What is left= is choosing⋅ ( ( θ )in− such a( way)) that+ the( odd( operators) + of( the)) expression vanishes. This motivates tan 2pθ p m, or equivalently θ 1 2p tan−1 p m which turns S into the same one that Foldy and Wouthuysen defined in their paper (eq. 7.6). Substituting( this) we= ~ obtain the new Hamiltonian= ( ~ ) [1, p.48]( ~ )

H′ β m2 p2 1~2. (7.16)

= ( + ) 24 This Hamiltonian, unlike the one from the Dirac representation is in agreement with the one from classical physics [3, p.3]. Notice that from the definition of β, eq. (3.15), we get two uncoupled equations. Two with a positive sign in front and two with a negative sign.

25 Electromagnetic Field In the case of an electromagnetic field the transformation is more complicated. The Hamiltonian is then given by

H α p eAA βm eΦ (7.17) where A is the vector potential and Ψ is the scalar potential. To solve this we = ⋅ ( − ) + + write the Hamiltonian in an simplified manner as

H βm (7.18) where α p eAA is the even term and eΦ is the odd term [9, p.183-184]. = + E + O The terms carry the properties O = ⋅( − ) E = β β, (7.19a) β β. (7.19b) E = E Following the steps of Bjorken andO Drell= −O [1, p.48-49], we expand the trans- formation equation for the Hamiltonian with the help of the Baker-Hausdorff formula.9 The following expression is obtained to n’th order

′ iS −iS iS −iS iS −iS H e H i∂t e e He e Se˙ i2 i3 in H i S, H S, S, H S,= S, (S, H− ) ... = S, S, ...− S, H ... = 2! 6 n! i i3 in + [ ...] + S˙ [ [S, S˙ ]] + S,[ S,[ S,[ S˙ ]]]]...+ + S,[ S,[ ... S,[S˙ ...] ]]...+ 2 6 n! To get a better− feeling− [ for] this− expression[ [ [ we]]]] can− see− that[ for[ small[ p] m]], i.e.− in the non-relativitic limit, we have from eq. (7.6) that ~ βαα p βαα p iS 1 2 p m (7.21) p 2m ⋅ ⋅ ≈ ( ~ )1 ( ~ ) = which is the same to say as iS m and thus we interpret the long expression in 1 (7.20) as a series in m [9, p.184]. Following the approximation of (7.21), where S contains the odd term of the∝ free particle Hamiltonian, we make an ansatz that α p eAA S iβ iβ 2m, (7.22) 2m ⋅ ( − ) which also makes S an odd= − term. In the non-relativistic= − O~ limit we terminate the series so that it takes the form i2 i3 H′ H i S, H S, S, H S, S, S, H 2! 6 i4 i i3 S,= S,+S,[ S, H] + [S˙ [ S,]]S˙+ [ S,[ S,[ S, S]]]]˙ + 24 2 6 9 A −A = + [ ] + 1 [ [ ]] + + 1 [ [ [ ] ]] + e Be B [ A,[ B [ 2 [A, A,]]]] B − ... − n! [A, A,] ...− A,[ B ...[ [ ... [9,]]]] p.184].

26 1 2 All we need to get our result are odd operators of order lower than m and 1 3 even operators of order lower than m , and hence all the other terms can be neglected [9, p.184]. It should be mentioned that the product of two( odd) and two even operators gives an odd operator,( ) while mixing even and odd gives an 1 2 odd operator [5, p.35]. In other words, if there are operators m4 , even, and 1 m3 , odd, they will be neglected. It may also be good to remind ourselves that that β is an even operator. What is left is now to evaluateO the terms of (7.23).EO From [1, p.50] we obtain β 1 i S, H , β 2, (7.24a) 2m m i2 β 2 1 1 S, S, H [ ] = −O + , [O, E] + O3, (7.24b) 2 2m 8m2 2m2 i3 O 3 1 [ [ ]] = − S, S,−S, H [O [O E]] − βO4, (7.24c) 3! 6m2 6m3 i4 O β 4 [ [ S,[ S, ]]]S, =S, H − β O (7.24d) 4! 24m3 O which also can be calculated using the[ properties[ [ [ of]]]] (7.19),= neglecting operators 1 with a factor m larger than sufficient. From these equations, though they look messy, two important conclusions can be drawn. First, since we have H′ H i S, H ... where H βm the transformation has removed the initial odd operator we had from the Hamiltonian. We have however, as the= equations+ [ above] + show, introduced= + severalE + O new odd operators. Even though it might not be completely obvious, there are in fact no odd operators of order 1 0 m left [4, p.35]. The final terms are [1, p.50] ( ) iβ ˙ S˙ i , (7.25a) 2m i i O S, S˙ = − , ˙ . (7.25b) 2 8m2 − [ ] = − [O O] 1 0 The crucial part is that we’ve removed the odd operator of order m and now 1 have odd operators of at most order m . We still want to remove these and the way to do this is to apply the same transformation in a recursive manner( ) with a new S, until they agree with the criteria under which we neglect odd operators. To do this we first gather the odd and even operators and define ′ and ′ [1, p.50] E O 2 4 1 1 ′ β , , , ˙ , (7.26a) 2m 8m3 8m2 8m2 O O E = E + Œ − ‘ − [Oβ [O E]] − 3 [OiβO˙] ′ , . (7.26b) 2m 3m2 2m O O which then leads us to define the newO = Hamiltonian[O E] − +

H′ βm ′ ′. (7.27)

≡ + E + O 27 1 0 In analogy with how we defined S to remove the operators of order m we will define iβ S′ ′, ( (7.28)) 2m 1 2 which now is of order m . We go through= O the same procedure again, neglecting terms not fulfilling the requirements. With this iteration we obtain [9, p.185] ( ) β iβO˙ ′ H′′ βm ′ ′, ′ (7.29) 2m 2m as our transformed Hamiltonian= + andE + this time[O E the] + odd operators are of at most 1 ′′ ′′ order m2 . Performing the transformation one more time with S iβ 2m we obtain H′′′ βm ′ = O(7.30)~ since now the odd operator ′′ is of order 1 3 and can be neglected [9, p.186]. = +mE What is left is to evaluate the even operators that remain O ( )

2 4 1 βm ′ βm β , , i ˙ , (7.31) 2m 8m3 8m2 O O + E = + E + Œ − ‘ − O [O E] + O which of course requires combining the terms in correct ways. Omitting the calculations, see for example [9, p.186], we obtain

p eAA 2 1 H′′′ β m p eAA 2 eΣ B 2 eΦ 2m 8m3 ( − ) e ie = Œ + − [( −βΣ )B − ⋅ Σ] curl‘ + E (7.32) 2m 8m2 e e Σ− E p⋅ eAA− ⋅ div EE. 4m2 8m2 This Hamiltonian completely− decouples⋅ the× ( upper− ) − two components from the lower two components of the since there are no odd operators. Notice for example that because of the β (diagonal matrix with first two diagonal components ( 1) and the two second ( 1)) the m term (mc2 in SI-units) will be negative for the first two components of the bispinor and negative for the second two components.+ This supports the interpretation− that the first components of the bispinor are positive energy solutions and the second two negative energy solutions. Based on [9, p.187] we let the Hamiltonian operate on a spinor ψ ψ1 and get for the component ψ1 ψ2 =  1 e p4 i∂ ψ m eΦ p eAA 2 σ B t 1 2m 2m 8m3 (7.33) =  + e + ( − ) − e ⋅ − Σ E p eAA div E ψ . 4m2 8m2 1 − ⋅ × ( − ) −

28 For the hydrogen atom we have a spherical constant potential Φ r , which leads 1 ∂Φ to E Φ r r ∂r r and A 0. This allows for the simplification [9, p.187] ( ) = −∇ ( ) = − =1 ∂Φ 1 ∂Φ Σ E p Σ r p Σ LL. (7.34) r ∂r r ∂r Substituting (7.34) into⋅ (7.33)× = − we can identify⋅ × = the− following⋅ operators.

H0 m rest energy

H1 eΦ r potential energy = ( ) 1 2 H3 = ( p)(eAA kinetic energy) 2m e Hµ = (σ −B ) coupling of magnetic( moment) 2m (7.35) p2 2 H = ⋅ ( relativistic correction) r 8m3 (e )1 ∂Φ H = − Σ L ( spin-orbit correction) so 4m2 r ∂r e H = 2Φ r ⋅ ( Darwin term) d 8m2 The fourth term,= − the coupling∇ ( )( of the , usually written) as µ e 1 B where µ g 2m 2σ [6, p.271-272], implies that the Land´eg-factor for the electron is g 2. This agrees with experiments and is a further indication that⋅ the Dirac equation≡ − is a powerful tool. =

29 8 Conclusions

We have, by applying two different methods, been able to obtain some of the most interesting corrections to the Schr¨odingertheory. The Taylor expansion of the exact energy expression gave us the energy correction terms and the trans- formation due to Foldy and Wouthuysen gave us the corresponding correction terms to the Hamiltonian. There are more corrections that can be found and are needed to explain phenomena such as hyperfine splitting. This would however require that the electron’s interaction with the nucleus is taken into considera- tion, for example introducing the spin of the nucleus. There is also , which is not predicted by the Dirac equation. The energy correction due to the Lamb shift is slightly smaller, being of order α5 which was also outside the scope of this text. The discovery of the Dirac equation is seen as one of the greatest achievements of modern physics. Dirac had his own idea of how to interpret the negative energy solutions. He presented his famous ’’ where all the negative energy states were occupied by , protected by Pauli’s exclu- sion principle. In this model the around us is a sea of infinite negative energy and the can be explained as a hole moving through the sea. Today we are more comfortable with the idea of anti-matter, which quantum field theory can explain in a more rigorous way.

30 References

[1] James. D. Bjorken and Sidney D. Drell. Relativistic Quantum Mechanics. McGraw-Hill, Inc., 1964. [2] Jim Branson. Dirac Equation. url: http://quantummechanics.ucsd. edu/ph130a/130_notes/node477.html. [3] John P. Costella and Bruce H. J. McKellar. “The Foldy-Wouthuysen trans- formation”. In: Am.J.Phys. 63. 1119. (1995). url: arXiv:hep-ph/9503416. [4] P. A. M. Dirac. “On the Dirac Theory of Spin 1/2 Particles and Its Non- Relativistic Limit”. In: Proc. R. Soc. 117.778 (1928), pp. 610–624. doi: 10.1098/rspa.1928.0023. [5] Leslie L. Foldy and Siegfried A. Wouthuysen. “On the Dirac Theory of Spin 1/2 Particles and Its Non-Relativistic Limit”. In: Physical Review 78.1 (1950), pp. 29–36. url: http://einstein.drexel.edu/~bob/Quantum_ Papers/Foldy-Wouthuysen.pdf. [6] David J. Griffiths. Introduction to Quantum Mechanics. Pearson Prentice Hall. 2nd Edition., 2005. isbn: 0-13-191175-9. [7] J.J. Sakurai. Advanced Quantum Mechanics. Addison-Wesley Publishing Company, Inc. 11th Printing., 1987. isbn: 0-201-06710-2. [8] Leonard I. Schiff. Quantum Mechanics. McGraw-Hill, Inc. 3rd Edition., 1968. isbn: 0-201-06710-2. [9] Franz Schwabl. Quantenmechanik f¨urFortgeschrittene (QM II). (German) [Advanced Quantum Mechanics]. Springer-Verlag Berlin Heidelberg. 3rd Edition., 2000. isbn: 3-540-67730-5.

31 Appendix A Relativistic Notation The cornerstone on which special relativity stands on is the concept of space- time. In space-time we specify an event by 4 coordinates,

xµ x0, x1, x2, x3 ct, x, y, z ct,xx . (8.1) Observers moving relative to each other will not be able to agree on time and = ( ) = ( ) ≡ ( ) distance between two events but besides the there is one quantity they will agree on. Given that one observer measures the difference in space- time coordinates between two events to be ct, x, y, z and another observer that measures the same distance to be ct′, x′, y′, z′ they will both agree on the quantity ( ) ( ) s2 c2t2 x2 y2 z2 c2 t′2 x′ 2 y′ 2 z′ 2. (8.2)

If we are interested≡ in− infinitesimal− − = distances( ) − ( in) space-time− ( ) − ( we) might as well write the same expression as

ds2 dx0 2 dx1 2 dx2 2 dx3 2 c2dt2 dx2 dy2 dz2. (8.3)

Introducing= ( Einstein’s) − ( sum) − convention,( ) − ( where) = summation− − is performed− over in- dices appearing both as super-scripts and sub-scripts, and the symmetric metric of special relativity 1 0 0 0 0 1 0 0 gµν ⎡ ⎤ , (8.4) ⎢0 0 1 0 ⎥ ⎢ ⎥ ⎢ − ⎥ ≡ ⎢0 0 0 1⎥ ⎢ ⎥ ⎢ − ⎥ we can write ⎢ ⎥ ⎣2 µ ν− ⎦ ds gµν dx dx , (8.5) The metric g is its own inverse, so that multiplying it with a vector twice is µν = equal to just multiplying with the . For convenience we could denote the inverse as gµν to go along with Einstein’s sum convention. This will also be more clear if we define contravariant and covariant vectors. Given a position 4-vector xµ we have

µ 0 0 gµν x gµν x ,xx x , x xµ (8.6)

A vector denoted by a subscript= ( will be) = called( − a) covariant≡ vector while as a contravariant vector will be denoted by a superscript. Similarily we get

µν µν 0 0 µ g xµ g x , x x ,xx x . (8.7)

To get expressions for the energy= ( and− momentum) = ( ) of≡ special relativity we will imagine a situation where a particle is moving relative to a while it is at rest according to another frame of reference. The space-time

32 difference between observing the particle twice will in the first frame of reference lead to v2 ds2 c2dt2 dx2 dy2 dz2 cdt2 1 . (8.8) c2 In the second reference= frame,− where− the− particle= isŒ at rest,− ‘ we obtain

ds2 c2 dt′ 2. (8.9)

We call the time τ t′ the proper time,= ( it is) the time that the moving particle ’experiences’. The relationship between the different t’s is thus ≡ v2 dτ dt 1 . (8.10) ¾ c2 Using the proper time we define the= 4-velocity− to be

dxµ 1 dx dy dx uµ c, , , γ c,vv . (8.11) dτ v2 dt dt dt 1 c2 ≡ = ¼ Œ ‘ ≡ ( ) The 4-momentum is defined in the− straightforward way

pµ muµ γ mc, mvv γmc,pp . (8.12)

Kinetic energy in special relativity≡ = is( given by) ≡ ( )

mc2 E γmc2 (8.13) v2 1 c2 = = ¼ and hence 4-momentum can also be written as−

pµ E c,pp . (8.14)

= ( ~ )

33 Appendix B An alternative representation of Quantum Mechanics is the so called Heisen- berg picture. In Schr¨odinger’sformulation the state vectors carry with them a time dependence. In the Heisenberg picture the state vectors themselves do not depend on time while the operators do. This gives an alternate representation that can be very useful, especially when investigating the time dependencies of expectation values.

To start with we will look at the Schr¨odingerpicture where the Schr¨odinger equation can be written d ih ψ, t Hˆ ψ, t . (8.15) dt If there is no time-dependence̵ inS the⟩ = HamiltonianS ⟩ the time-dependent state vector ψ, t is correlated to an initial state by a time dependent factor [8, p.169],

̵ −iHˆ (t−t0)~h S ⟩ ψ, t e ψ, t0 Uˆ ψ, t0 . (8.16) ˆ The expectation valueS of some⟩ = operator SA can⟩ ≡ thenS be⟩ written † A ψ, t A ψ, t ψ, t0 Uˆ AUˆ ψ, t0 ̵ ̵ (8.17) iHˆ (t−t0)~h −iHˆ (t−t0)~h ⟨ ⟩ψ,= t⟨0 e S S ⟩ =Ae⟨ S ψ,S t0 ⟩, In the above expression= ⟨ we haveS used that [8, p.169]S ⟩ ˆ ̵ Uˆ † Uˆ −1 eiH(t−t0)~h, (8.18) which further makes Uˆ a unitary= operator= (Uˆ −1Uˆ Uˆ †Uˆ 1). In the Heisenberg picture the expectation values are calculated with the same expression (8.17) but instead state vectors and the operators are defined= differently.= To compare the two we will introduce a sub-indexes to keep track of the two pictures. The operator Aˆ in the two pictures is related by

† AˆH Uˆ AˆSUˆ (8.19) and the state vectors [7, p.113] = ˆ −1 ψ, t H U ψ, t S ψ, t0 S. (8.20a)

We can investigate the timeS dependence⟩ = S of⟩ the= S operator⟩ by differentiating. We obtain dAˆ ∂Uˆ † ∂Aˆ ∂Uˆ H Aˆ Uˆ Uˆ † S Uˆ Uˆ †Aˆ (8.21a) dt ∂t S ∂t S ∂t i i ∂Aˆ Hˆ Uˆ †=Aˆ Uˆ Uˆ †Aˆ+ Hˆ Uˆ +Uˆ † S U,ˆ (8.21b) h S S h ∂t − = ‰ ̵ Ž + ‰ ̵ Ž + 34 which can be simplified into

dAˆ i i H Hˆ Aˆ Aˆ H,ˆ (8.22a) dt h H h H = − since ∂AˆS ∂t 0 and H,ˆ Uˆ 0.̵ The commutator̵ can be motivated by the fact that Uˆ can be written in a series of Hˆ [8, p.169]. What is left is just the commutator( ~ of) A=ˆH and H[ˆ with] = an extra factor,

dAˆ 1 H Aˆ , Hˆ . (8.23) dt ih H With eq. (8.23) we can then investigate whether are conserved, = ̵ [ ] so called constants of motion. Based on the derivations made by J.J. Sakurai [7, p.113-114] we study the operators L and S. For a free particle we have the 2 Hamiltonian, simplified using Einsteins sum convention, H cαjpj βmc . The orbital angular momentum operator L x p is not a constant of motion since = + dLˆ 1 1 ≡ × 1 L , Hˆ x p x p , cα p βmc2 dt ih 1 ih 2 3 3 2 j j 1 = [ ] = [ − x p x p+, cα p ] ̵ ̵ ih 2 3 3 2 j j (8.24) 1 x p =, cα [p −x p , cα p ] ih 2 3 ̵j j 3 2 j j c α3p2 α2p3 = ̵ ‰[ ] − [ ]Ž and the same applies to the other components. In the above calculation we have = ( − ) used that pj ih ∂ ∂xj and pjpk acting on a state vector is 0 for all spacial combinations j, k. Hence we have ≡ − ̵( ~ ) dLL c α p 0. (8.25) dt h̵ We can show that the spin-operator=S( 2×Σ) is≠ not a constant of motion by using the identities of (5.11) and keeping in mind that swapping two gamma matrices of different index changes the sign of the≡ product ∂S 1 h 1 1 Σ , Hˆ Σ , cα p βmc2 ∂t ih 2 1 2i 1 j j ̵ 1 = [ ] = [ iγ2γ3, cγ+ 0γjp ] (8.26) ̵ 2i j 1 cp γ=0 γ[2−γ3γj γjγ2γ3 ]. 2 j If j 1 the result is zero and non-zero= otherwise.(− We+ are left) with ∂S 1 = 1 cγ0 2p γ3 2p γ2 c p γ0γ3 p γ0γ2 ∂t 2 2 3 2 3 (8.27) = ( − ) = ( c p2α−3 p3α2 ). = ( − ) 35 Just like the components of the orbital angular momentum operator the spin operator’s components are not constants of . The other components are calculated in the same way and leads to the conclusion that

dSS c α p . (8.28) dt The Heisenberg picture may also be= used− ( to× find) operators. We could find the velocity operator in the Dirac theory by differentiating the components of the position operator x. Doing this we get dx 1 1 x˙ x , cα p cα x p p x . (8.29) k dt ih k j j ih j k j j k and from non-relativistic= = quantum̵ [ mechanics] = ̵ we( know− that) the last factor is just the canonical commutation relation which is equal to ihδkj [6, p.133]. The result is simply x˙ cααα. ̵ (8.30)

=

36