23.05.2014 Views

0.1 Klein-Gordon Equation 0.2 Dirac Equation

0.1 Klein-Gordon Equation 0.2 Dirac Equation

0.1 Klein-Gordon Equation 0.2 Dirac Equation

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

[revised Oct 15, 2010]<br />

<strong>0.1</strong> <strong>Klein</strong>-<strong>Gordon</strong> <strong>Equation</strong><br />

The relativistic Schrödinger equation<br />

∂ µ ∂ ν ψ − m 2 ψ = 0<br />

became known as the <strong>Klein</strong>-<strong>Gordon</strong> equation after it was rejected by Schrödinger<br />

as a relativistic one particle generalization of the NR Schrödinger equation.<br />

To list three of its defects (1.) It has solutions of unbounded negative energy<br />

so that it predicts continuous radiative transitions, (2.) The only conserved<br />

current 4-vector ψ † ∂ ↔ µ ψ can have negative j 0 and cannot therefore be interpreted<br />

as a probability density, and (3.) When generalized to include<br />

interaction with the electromagnetic field, its predictions do not agree with<br />

the observed spectrum of the H-atom. (1.) is easily fixed in a context where<br />

ψ is viewed as a quantized field. (<strong>Dirac</strong> had a solution in a one particle<br />

context.) (2.) can be fixed by interpreting the current as a charge-current<br />

4-vector. (<strong>Dirac</strong>’s equation fixed it for spin 1 2 particles.) (3.) was fixed by<br />

<strong>Dirac</strong>. (It can also be used for spin 0 particles.)<br />

<strong>0.2</strong> <strong>Dirac</strong> <strong>Equation</strong><br />

<strong>Dirac</strong>’s solution to the lack of a positive definite probability density was to<br />

make the equation linear in the space and time derivatives. In doing so he<br />

found an equation that included the electron spin and accounted correctly for<br />

its magnetic moment. His idea was to express the energy as a linear form in<br />

the momentum and mass by introducing non-commuting coefficients, realized<br />

as matrices, and allowing ψ to become a multicomponent wave function.<br />

Now requiring E 2 = p 2 + m 2 gives<br />

E = β · p + αm.<br />

β i β j + β j β i = 2g ij<br />

αβ i + β i α = 0<br />

α 2 = 1, (1)<br />

1


i.e., α and β i are a set of 4 anticommuting matrices whose square is unity.<br />

In terms of derivatives the <strong>Dirac</strong> equation is:<br />

i∂ 0 ψ = −iβ i ∂ i ψ + αmψ<br />

To give it a covariant appearence (deceptive because the γ’s are numerical<br />

matrices that do not change under LT, but see below), introduce γ i = αβ i<br />

and γ 0 = α, multiply the equation by α and obtain<br />

The γ µ satisfy<br />

−iγ µ ∂ µ ψ + mψ = 0 (2)<br />

{γ µ , γ ν } = −2g µν (3)<br />

where {} means anticommutator. A possible explicit form is<br />

( )<br />

( )<br />

1 0<br />

0 σ<br />

γ 0 =<br />

γ i i<br />

=<br />

0 −1<br />

−σ i 0<br />

Each entry is a 2 by 2 matrix, making γ 4 by 4. An equivalent explicit form<br />

is<br />

( )<br />

( )<br />

0 1<br />

0 σ<br />

γ 0 =<br />

γ i i<br />

=<br />

1 0<br />

−σ i (5)<br />

0<br />

The first form is the standard form, good for taking the NR limit. The<br />

second form is the chiral form, good for neutrinos and things. Changing<br />

involves reshuffling components of ψ. Both forms can be represented as<br />

direct products of two sets of Pauli matrices τ i for the skeleton and σ i for<br />

the entries: the first set is (τ 3 , iτ 2 σ i ) and the second is (τ 1 , iτ 2 σ i ). Both<br />

forms (and others you can easily invent) satisfy Eq.(3). In both forms, the<br />

hermitian conjugate of γ 0 is γ 0 and of γ i is −γ i . A very useful relation valid<br />

for both forms is<br />

γ µ† = γ 0 γ µ γ 0 .<br />

0.3 Free Particle Solutions<br />

Look for solution in the form of an eigenstate of the energy-momentum vector:<br />

ψ = ψ 0 e ipx ,<br />

(4)<br />

2


where ψ 0 is a constant 4-component column vector (<strong>Dirac</strong> spinor). ψ 0 must<br />

satisfy (γ µ p µ + m)ψ 0 = 0. If ψ 0 has upper components χ and lower components<br />

χ ′ (both ordinary 2-component spinors), then the equation takes the<br />

form ( ) ( )<br />

−E + m ⃗σ · ⃗p χ<br />

−⃗σ · ⃗p E + m χ ′ = 0 (6)<br />

using the standard form for the γ-matrices and replacing p 0 by E. There is<br />

a non-trivial solution if the determinant is zero. To evaluate things with σ<br />

use the identity σ i σ j = δ ij + iɛ ijk σ k . The determinant is (−(E) 2 + ⃗p 2 + m 2 ) 2<br />

which has double roots at E = ± ω with ω = + √ ⃗p 2 + m 2 , implying two<br />

positive energy states and two negative energy states for each value of ⃗ k.<br />

Negative energy states are still present, but <strong>Dirac</strong> eliminated them by saying<br />

that they were full, so the exclusion principle forbade transitions to them.<br />

A hole in the negative energy ’sea’ behaves like positive particle of the same<br />

mass: positron.<br />

DISCUSS: motion of holes.<br />

A solution of Eq.(6) is obtained by taking χ to be an arbitrary twocomponent<br />

spinor. The equation then requires that χ ′ = ⃗σ · ⃗pχ/(E + m).<br />

Note that the equation implies that χ ′ ≈ vχ in the NR limit. (χ called large<br />

components and χ ′ small components.) We will return to these states when<br />

we treat the <strong>Dirac</strong> equation as a field equation.<br />

The physical meaning of states and operators clearer if we multiply the<br />

<strong>Dirac</strong> equation by γ 0 and identify p 0 with the Hamiltonian:<br />

H = p 0 = γ 0 γ i p i + γ 0 m,<br />

where p i can be interpreted either as the momentum operator or its eigenvalue.<br />

Clearly the spatial momentum commutes with H, but what about the<br />

angular momentum? Calculation shows that the orbital angular momentum<br />

⃗r × ⃗p does not commute with H. By inspecting the commutator, we can find<br />

that adding a term gives an operator that does commute with H:<br />

J i = ɛ ijk r j p k + i 4 ɛ ijkγ j γ k<br />

The explicit form of the second term is<br />

( )<br />

S i = 1 σi 0<br />

2<br />

0 σ i<br />

3


which can be identified with the spin. The total angular momentum J i =<br />

L i + S i commutes with H.<br />

As in NR QM, The plane wave solutions above are not eigenstates of<br />

any of the angular momentum operators. (For eigenstates of ⃗ J see the H-<br />

atom below.) There is however an important operator that commutes with<br />

H and can be well-defined, the helicity h = ⃗σ · ⃗p/|⃗p|. It has eigenvalues ±1,<br />

and by choosing χ to be one of its eigenstates, the full ψ will also be an<br />

eigenstate. The helicity is not Lorentz invariant, however, because a boost<br />

in the direction of the momentum can bring the particle to rest and reverse<br />

its momentum without changing its spin.<br />

0.4 Transformation<br />

If we do an LT, the components of ψ get mixed like the components of a<br />

3-vector under rotation. The new wave function satisfies<br />

ψ ′ (x ′ ) = S(Λ)ψ(Λ −1 x ′ )<br />

by analogy with the similar equation for rotations. Note first that S must<br />

be a representation of the Lorentz group up to a phase. Starting with the<br />

<strong>Dirac</strong> equation in x, change variables to x ′ = Λx and multiply by S:<br />

−iSγ µ S −1 (∂ µ x ′ν )∂ ′ νSψ(Λ −1 x) + mSψ(Λ −1 x) = 0,<br />

where we used the chain rule and inserted S −1 S remembering that S acts in<br />

the same space as the γ’s and may not commute with them. The result is<br />

the <strong>Dirac</strong> equation in the new frame with the new wave function identified<br />

as above, provided S(Λ) satisfies the condition<br />

S −1 γ µ S = Λ µ ν γν (7)<br />

Since S(Λ) is a representation of the Lorentz Group, we can find it by finding<br />

its infinitesimal elements. Putting S = 1 + ɛ ab Σ ab ,<br />

(1 − ɛ ab Σ ab )γ µ (1 + ɛ ab Σ ab ) = (δ µ ν + ɛ abg µν G ab<br />

µν )γν<br />

By ansatz, the solution is found to be Σ ab = − 1 2γ a γ b . It is interesting to note<br />

the explicit forms of Σ for rotations<br />

( ) ⃗σ<br />

Σ ij = i k<br />

0<br />

2<br />

0 ⃗σ k .<br />

4


and for boosts<br />

Σ 0i = − 1 2<br />

( 0 ⃗σ<br />

k<br />

⃗σ k 0<br />

The i for rotations but not for boosts means that S is unitary for rotations<br />

but not for boosts. Good because spatial volume is preserved by rotations<br />

but not by boosts because of Lorentz contraction, so ψ † a ψ a is preserved by<br />

rotations but changed by boosts.<br />

Thus ψ † ψ is not a scalar. However, it is easy to show using γ µ† = γ 0 γ µ γ 0<br />

that ψ † γ 0 ψ does not change under an infinitesimal LT, and therefore under<br />

any LT. It therefore satisfies the condition S ′ (x) = S(Λ −1 x) that defines it<br />

as a scalar. This is important enough to warrant special notation:<br />

ψ = ψ † γ 0 .<br />

)<br />

.<br />

Using Eq.(7), you should be able to prove the following<br />

ψψ is a scalar.<br />

ψγ µ ψ is a vector.<br />

ψγ µ γ ν ψ is a second rank tensor.<br />

Other covariants can be expressed with the aid of the matrix<br />

γ 5 = iγ 0 γ 1 γ 2 γ<br />

( )<br />

3<br />

0 1<br />

=<br />

1 0<br />

( ) −1 0<br />

=<br />

0 1<br />

in standard form<br />

in chiral form<br />

You can prove<br />

ψγ 5 ψ is a pseudoscalar<br />

ψγ µ γ 5 ψ is a pseudovector (axial vector).<br />

It is also true that covariants with 3 γ’s can be expressed in terms of the<br />

others.<br />

5


0.5 NR Limit<br />

Nothing better shows the physics buried in the <strong>Dirac</strong> equation than taking<br />

the NR limit. To understand, we must add the coupling of the <strong>Dirac</strong> electron<br />

to the electromagnetic field (discussion below in <strong>Dirac</strong> <strong>Equation</strong> as a Field<br />

<strong>Equation</strong>).<br />

γ µ (−i∂ µ + eA µ ) ψ + mψ = 0 (8)<br />

or in Hamiltonian form<br />

H = −eA 0 ψ + γ 0 γ i (p i + eA i )ψ + mγ 0 ψ (9)<br />

Thus Hψ = Eψ in two-component form with E moved to left is<br />

( m − eA 0 − E ⃗σ · (⃗p + eA)<br />

⃗ ) ( ) χ<br />

⃗σ · (⃗p + eA)<br />

⃗ −m − eA 0 − E χ ′ = 0.<br />

Eliminate χ ′ in favor of χ, and A 0 = φ:<br />

(<br />

(m − E − eφ) + ⃗σ · ⃗Π 1<br />

m + E + eφ<br />

)<br />

⃗σ · ⃗Π χ = 0 (10)<br />

where we used the abbreviation ⃗ Π = ⃗p+e ⃗ A. To get to the NR limit, subtract<br />

the rest energy from E to get the NR energy W = E − m and write the<br />

denominator<br />

1<br />

2m + (W + eφ) ≈ 1<br />

2m − 1<br />

1<br />

(W + eφ) +<br />

4m2 8m (W + 3 eφ)2 + . . .<br />

First term in () in Eq.(10) is O(mv 2 ); second is O(mv 2 ) plus m times higher<br />

powers of v 2 . Using the commutator [p i + eA i , p j + eA j ] = −ieF ij and the<br />

identity σ i σ j = g ij + iɛ ijk σ k , and keep terms up to order mv 2 , this becomes<br />

(<br />

−W − eφ + 1 Π<br />

2m ⃗ 2 + e )<br />

⃗σ<br />

m 2 · ⃗B χ = 0. (11)<br />

This is the Pauli equation for a spinning NR particle. The last term is<br />

the interaction of the spin magnetic moment with the magnetic field. The<br />

magnetic moment is (e/m) S ⃗ rather than the orbital (e/2m) L, ⃗ indicating the<br />

g-factor of 2 put in by hand in pre-<strong>Dirac</strong> times and justified a postiori as the<br />

Thomas precession.<br />

6


In higher order, the equivalence of DE to Pauli equation is approximate<br />

because the energy eigenvalue appears in the expansion of the second term<br />

in Eq.(10) so that is does not have the same form as the Pauli equation.<br />

Nonetheless, we can obtain an approximate Pauli equation from the next<br />

term in the expansion above as follows:<br />

− 1<br />

4m 2 (⃗σ · ⃗Π)(W + eφ)(⃗σ · ⃗Π) = − 1<br />

4m 2 (⃗σ · ⃗Π) 2 (W + eφ) − 1<br />

4m 2 (⃗σ · ⃗Π)[eφ, ⃗σ · ⃗Π].<br />

It is correct to order v 4 to replace W + eφ in the first term on the right by<br />

(⃗σ · ⃗Π) 2 according to Eq.(11). In the absence of a magnetic field ⃗ A = 0, the<br />

result is the order v 4 relativistic kinetic energy correction −p 4 /8m 3 . Taking<br />

the commutator in the second term and using the σ i σ j identity yields<br />

second =<br />

ie<br />

4m (δ 2 ij + iɛ ijk σ k )Π i ∂ j φ<br />

in which the first term is known as the Darwin term and the second is the<br />

spin-orbit term which was added phonomenologically to the Pauli equation<br />

in the olden days. To appreciate the physical significance of these terms, let’s<br />

specialize to A ⃗ = 0, whence the Darwin term becomes<br />

Darwin = − ie<br />

4m 2 (∇2 φ − ⃗ E · ⃗∇).<br />

This spin independent term represents a relativistic correction to the effective<br />

potential which has obscure experimental consequences. For the pure<br />

coulomb field, the first term is a delta function which shifts the energy of H-<br />

atom s-states. The significance of the spin-orbit term appears if we specialize<br />

to central potentials φ(r) where it becomes<br />

SO = −<br />

e 1 ∂φ<br />

⃗σ · ⃗L.<br />

4m 2 r ∂r<br />

COMMENT<br />

A more systematic and elegant way of approaching the NR approximation<br />

is the Foldy-Wouthuysen (FW) transformation. The idea is to make a unitary<br />

transformation of the states and operators, ψ = Uψ and O F W = UOU † with<br />

UU † = 1, such that the new Hamiltonian is block diagonal, decoupling the<br />

upper and lower components and giving separate two-component equations<br />

for each. For a free particle, this is accomplished by<br />

√<br />

m + ω<br />

U =<br />

2ω + γ i p<br />

√ i<br />

,<br />

2ω(ω + m)<br />

7


in which ω is the operator + √ ⃗p 2 + m 2 which commutes with ⃗p. It is easily<br />

verified that UU † = 1, and a lengthy calculation shows that the FW<br />

Hamiltonian reduces to<br />

( )<br />

ω 0<br />

H F W =<br />

.<br />

0 −ω<br />

Thus the upper components represent the positive energy solutions and the<br />

lower components represent the negative energy solutions.<br />

Although the momentum operator is not altered by the FW transformation,<br />

since U commutes with ⃗p, the position operator is modified and<br />

becomes a non-local operator in a way that is difficult to calculate in this<br />

context. This is an indication of the fact that it is not possible to form a<br />

state that is localized to a single point using positive energy solutions only.<br />

At best, the paritcle can be localized to within a Compton wavelength m −1 .<br />

It is possible to do a FW transformation for particle subject to potentials,<br />

but it must be done sucessively for each higher order in v 2 . Omit it here.<br />

0.6 H Atom<br />

0.6.1 Hamiltonian<br />

The relativistic treatment of the H-atom is a major success of the <strong>Dirac</strong><br />

theory. More generally, we treat a particle in a central field in the absence of<br />

a magnetic field. In particular, we neglect the (weak) magnetic field due to<br />

the nuclear magnetic moment. With the <strong>Dirac</strong> equation γ 0 (−i∂ 0 + eA 0 )ψ +<br />

γ i ∂ i ψ + mψ = 0 as a starting point, we rewrite in the form of a Schrödinger<br />

equation i∂ t ψ = Hψ with the Hamiltonian<br />

H = γ 0 γ i p i − α r + γ0 m<br />

( )<br />

m −<br />

α<br />

⃗σ · ⃗p<br />

=<br />

r<br />

⃗σ · ⃗p −m − α (12)<br />

r<br />

where α = e 2 /4π is the fine structure constant, value approximately 1/137.<br />

−α/r can be replaced by any central potential −eφ(r) or V (r).<br />

0.6.2 Conserved Quantities<br />

As noted earlier, the total angular momentum ⃗ J is conserved. Therefore we<br />

can have simultaneous eigenstates of J 2 , J z , and H. in addition, the parity<br />

8


is conserved. The operator P defined by P ψ(⃗r) = ψ(−⃗r) does not commute<br />

with the Hamiltonian because it changes the sign of the momentum operator.<br />

However P = γ 0 P does commute. Note that P commutes with the spin<br />

operator S i = i 4 ɛ ijkγ j γ k , so it does not change its value. This is proper<br />

because the spin, like the orbital angular momentum, should be a good axial<br />

vector and not change sign under space reflection. According to NRQM,<br />

operators J 2 , L 2 , and J z have eigenstates, generally denoted by Ylj m. j can<br />

be either l + 1 2 or l − 1 2 . We introduce the quantity ω with values ±1 such<br />

that j = l + 1 2ω. The parity P of the eigenstate is (−1) j− ω 2 so that states<br />

with different values of ω have different parity. We denote the eigenstates by<br />

Yωj m and they satisfy<br />

J 2 Y m ωj = j(j + 1)Y m ωj<br />

J z Y m ωj = mY m ωj<br />

L 2 Y m ωj = (j − ω 2 )(j − ω 2 + 1)Ym ωj<br />

PY m ωj = (−1) j− ω 2 Y<br />

m<br />

ωj<br />

Don’t need to know their explicit forms.<br />

0.6.3 Eigenstates<br />

Look for solutions that are simultaneous eigenstates of H, P, and angular<br />

momentum. Since P = γ 0 P looks like<br />

( )<br />

P 0<br />

P =<br />

,<br />

0 −P<br />

the parity of the upper components is opposite to the parity of the lower<br />

components. Our eigenstate must look like<br />

ψ = 1 ( ) f(r)Y<br />

m<br />

ωj<br />

r ig(r)Y−ωj<br />

m , (13)<br />

where the i and 1/r are put in for convenience. The following identities ease<br />

the calculation along. Try to prove them, but proofs will be provided in class.<br />

ˆr is the unit radial vector.<br />

i.)<br />

⃗σ · ⃗p = −i⃗σ · ˆr∂ r +<br />

9<br />

i⃗σ · ˆr<br />

⃗σ ·<br />

r<br />

⃗L


ii.) ⃗σ · ˆr commutes with ⃗ J and changes parity, so you can prove<br />

⃗σ · ˆrY m ωj = −Y m −ωj<br />

iii.) Since ⃗σ · ⃗L = J 2 − L 2 − 1 4 ⃗σ2 , it is easy to get<br />

⃗σ · ⃗L Y m ωj = (ωj + 1 2 ω − 1) Ym ωj<br />

Putting the Hamiltonian and the wave function together, we obtain the<br />

coupled equations<br />

0.6.4 Solution<br />

f ′ − ω(j + 1 2 )<br />

r<br />

g ′ + ω(j + 1 2 ) g −<br />

r<br />

At r = ∞ both f and g satisfy the equation<br />

(<br />

f − m + E + α )<br />

g = 0 (14)<br />

r<br />

(<br />

m − E − α )<br />

f = 0 (15)<br />

r<br />

f ′′ − (m 2 − E 2 )f = 0<br />

which has decaying solution f = e −√ m 2 −E 2r , suggesting the new dimensionless<br />

radial variable<br />

ρ = √ m 2 − E 2 r<br />

Introducing this variable and new independent variables such that f = F e −ρ<br />

and g = Ge −ρ brings the equations to the form<br />

dF<br />

dρ − F − βF ρ<br />

− 1 ν G − αG ρ<br />

dG<br />

dρ − G + βG ρ<br />

− νF + αF ρ<br />

= 0<br />

= 0<br />

(16)<br />

in which ν stands for √ (m − E)/(m + E) and β stands for ω(j+ 1 2 ). Examination<br />

of the equations near ρ = 0 shows F and G must behave as ρ s where<br />

s = √ β 2 − α 2 , a power which is less than |β| (≥1) by something of the order<br />

of α 2 . Introducing<br />

F (ρ) = ∑ a p ρ s+p<br />

p=0<br />

10


and a similar series for G with coefficients b p leads to the following recursion<br />

relation<br />

(s + p + 1)a p+1 − a p − βa p+1 − 1 ν b p − αb p+1 = 0<br />

(s + p + 1)b p+1 − b p + βb p+1 − νa p + αa p+1 = 0<br />

The series must terminate to avoid generating the exponentially increasing<br />

solution. If p is the index of the highest power of ρ before termination, then<br />

we can deduce that a p = −νb p from the fact that a p+1 and b p+1 are both zero<br />

while a p and b p are non-zero. If we use this relation in the recursion relations<br />

above with p → p − 1, we find that all the coefficients can be eliminated and<br />

a quadratic equation for ν obtained:<br />

The appropriate solution is positive:<br />

αν 2 + 2(p + s)ν − α = 0.<br />

ν = − p + s<br />

α<br />

√ (p ) 2 + s<br />

+ + 1<br />

α<br />

It is then a matter of algebra to deduce from this that<br />

E =<br />

m<br />

√<br />

1 + α2<br />

(p+s) 2<br />

To make sense of this and to compare it to the NR result, we can expand it in<br />

powers of α = e 2 /4π ≈ 1/137, noting the dependence of s = √ (j+ 1 2) 2 − α 2<br />

on α. The final result is<br />

E = m<br />

(1 − α2<br />

2n 2 − α4<br />

2n 4 ( n<br />

j+ 1 2<br />

− 3 4<br />

)<br />

)<br />

+ O(α 6 ) ,<br />

in which n stands for p + j + 1 2. After the rest energy, the α 2 term is the<br />

Rydberg energy calculated in NR QM and the α 4 term is the relativistic<br />

correction. In all orders the energy depends only on j and n, so some but<br />

not all of the ’accidental’ degeneracy of the H atom remains. DISCUSSION<br />

11


0.7 <strong>Dirac</strong> <strong>Equation</strong> as a Field <strong>Equation</strong><br />

Although the <strong>Dirac</strong> equation is somewhat sucessful as a one particle equation,<br />

the idea of a filled sea of negative energy states is bizarre, and is also<br />

not applicable to particles of integral spin that do not satisfy the exclusion<br />

principle. The integral spin particles also do not have a consistent quantum<br />

treatment with positive definite probability density. There were other difficulties<br />

such a the <strong>Klein</strong> paradox. Therefore the <strong>Dirac</strong> equation came to be<br />

regarded as a field equation to be subjected to a form of canonical quantization<br />

just like the scalar and electromagnetic fields. To realize this, we must<br />

have a Lagrangian for the <strong>Dirac</strong> equation. A Lagrangian density that serves<br />

is<br />

L = iψγ µ ∂ µ ψ − mψψ (17)<br />

It is easily verified that the equations of motion arising from this are the<br />

<strong>Dirac</strong> equation and its conjugate:<br />

i∂ µ ψγ µ + mψ = 0<br />

When we come to construct the Hamiltonian, however, there are pathological<br />

elements in the calculation. For example, the momentum conjugate to ψ is<br />

identically zero, and the momentum conjugater to ψ is<br />

Π = ∂L<br />

∂∂ 0 ψ<br />

= iψγ 0<br />

Since the momentum conjugate to ψ is iψ † , a coordinate, we are dealing with<br />

a constrained system. Though there are sophisticated methods to deal with<br />

such constraints, we will follow a simple practical method and consider only<br />

ψ to be a coordinate in the Lagrangian. When we construct the Hamiltonian<br />

as<br />

H = Π∂ 0 ψ − L,<br />

We find that the resulting Hamiltonian does not involve time derivatives,<br />

and that these could not be eliminated anyway as in the previous examples<br />

because the canonical momentum does not contain time derivatives. Despite<br />

these peculiarities, the Hamiltonian density derived above leads to the<br />

Hamiltonian<br />

∫<br />

H = d 3 xψ ( † iγ 0 γ i ∂ i + γ 0 m ) ψ,<br />

12


expressed in terms of the fields. Although this looks like the expectation<br />

value of the <strong>Dirac</strong> Hamiltonian in the single particle theory, it now signifies<br />

the field Hamiltonian with ψ and ψ interpreted as fields to be quantized later.<br />

0.7.1 <strong>Dirac</strong> Field in Fourier Space<br />

To formulate the theory in Fourier space, we begin by expressing ψ as a sum<br />

of free particle solutions of the free <strong>Dirac</strong> equation. Denote the two positive<br />

energy solutions by<br />

ψ = u s ( ⃗ k)e i⃗ k⃗x e −iωt<br />

and the two negative ones by<br />

ψ = ṽ s ( ⃗ k)e i⃗ k⃗x e iωt<br />

with the usual definition of ω and s = 1, 2. The tilde is used because we will<br />

later wish to redefine the negative energy states to make them more easily<br />

identifiable as antiparticle states. Because u s and ṽ s are eigenstates of the<br />

Hermitian H belonging to different eigenvalues, they satisfy u † s (⃗ k)ṽ s ′( ⃗ k) = 0.<br />

They can also be chosen orthogonal among themselves by appropriate choices<br />

of the spinor χ in Eq.(??). This can be done simply by choosing the χ s ’s<br />

to be orthogonal. However, unless the χ’s are eigenstates of the helicity<br />

h = σ · ˆp which is conserved by the free H, the resulting u’s or v’s will not<br />

be eigenstates of any component of the spin even though χ is always such an<br />

eigenstate. In fact, the components of the spin are not conserved by the free<br />

<strong>Dirac</strong> Hamiltonian unless they are in the direction of the momentum. We<br />

normalize the states to satisfy the orthonormality relations<br />

u † s( ⃗ k)u s ′( ⃗ k) = 2ωδ ss ′,<br />

ṽ † s (⃗ k)ṽ s ′( ⃗ k) = 2ωδ ss ′,<br />

u † s (⃗ k)ṽ s ′( ⃗ k) = 0.<br />

Beginning with the superposition<br />

∫ (<br />

ψ(x) = a s ( ⃗ k)u s ( ⃗ k)e i⃗k·⃗x + ˜b s ( ⃗ k)ṽ s ( ⃗ )<br />

k)e i⃗ k·⃗x<br />

˜dk ∑ s<br />

where the quantities with tildes will be modified in anticipation of their role<br />

as antiparticle amplitudes. Since an antiparticle associated with a negative<br />

13


energy state of momentum ⃗ k has momentum − ⃗ k, we change to − ⃗ k in the<br />

integral over ˜b. Then we redefine ṽ s (− ⃗ k) to be v s ( ⃗ k) and ˜b s (− ⃗ k) to b † s (⃗ k). The<br />

latter takes account of the fact that after quantization a and a † annihilate<br />

and create particles respectively. Annihilation of a particle in a negative<br />

energy state creates an antiparticle there. so the roles of the operator and<br />

its conjugate are interchanged. The expansion for ψ now looks like<br />

∫ (<br />

ψ(x) = a s ( ⃗ k)u s ( ⃗ k)e i⃗k·⃗x + b † s( ⃗ k)v s ( ⃗ )<br />

k)e −i⃗ k·⃗x<br />

(18)<br />

˜dk ∑ s<br />

It is easy to substitute this expansion into the Hamiltonian to obtain<br />

∫ (<br />

H = ω a † s (⃗ k)a s ( ⃗ k) − b s ( ⃗ )<br />

k)b † s (⃗ k)<br />

˜dk ∑ s<br />

(19)<br />

using the fact that u s and v s are eigenstates of H D that satisfy the above<br />

orthonormality relations. It should be possible to prove that the Lagrangian<br />

has a form something like<br />

∫ ( i<br />

)<br />

L = ˜dk (a † ȧ + bḃ† − ȧ † a − ḃb† ) − 1<br />

2<br />

2ω(a † a + bb † ) .<br />

Perhaps you can prove it with the aid of the following identities involving the<br />

solutions u s ( ⃗ k) and v s ( ⃗ k), which we also list for future reference. Involving<br />

bars:<br />

Involving ± ⃗ k:<br />

u s ( ⃗ k)u s ′( ⃗ k) = 2mδ ss ′<br />

v s ( ⃗ k)v s ′( ⃗ k) = −2mδ ss ′<br />

u s ( ⃗ k)v s ′( ⃗ k) = 0<br />

Involving γ i k i :<br />

u † s (⃗ k)v s ′(− ⃗ k) = 0<br />

v † s (⃗ k)u s ′(− ⃗ k) = 0 (20)<br />

u s ( ⃗ k)γ i k i u s ′( ⃗ k) = 2 ⃗ k 2 δ ss ′<br />

v s ( ⃗ k)γ i k i v s ′( ⃗ k) = 2 ⃗ k 2 δ ss ′<br />

u s ( ⃗ k)γ i k i v s ′( ⃗ k) = 2ω√<br />

⃗k 2<br />

δ ss ′<br />

(21)<br />

14


This Lagrangian leads to solutions a( ⃗ k) = a 1 ( ⃗ k)e −iωt and b( ⃗ k) = b 1 ( ⃗ k)e −iωt<br />

where a 1 ( ⃗ k) and b 1 ( ⃗ k) are constants.<br />

The free <strong>Dirac</strong> field can also be written as a superposition of the time<br />

dependent solutions:<br />

∫<br />

ψ(x) =<br />

˜dk<br />

(a s ( ⃗ k)u s ( ⃗ k)e ikx + b † s (⃗ k)v s ( ⃗ )<br />

k)e −ikx<br />

with 4-dimensional k and x in analogy to the similar forms for the scalar and<br />

electromagnetic fields. The Hamiltonian has the same expression in terms of<br />

the a’s and b’s as in the previous time independent expansion, With the aid<br />

of the orthogonality relations between u’s and v’s, The expansion of ψ can<br />

be inverted to express a s ( ⃗ k) and b † s (⃗ k) in terms of ψ(x):<br />

∫<br />

a s ( ⃗ k) = d 3 xu † s( ⃗ k)e −ikx ψ(x)<br />

∫<br />

b † s (⃗ k) = d 3 xv s † (⃗ k)e ikx ψ(x)<br />

15

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!