Lecture 10 (Spinors and The Dirac Equation) PDF
Lecture 10 (Spinors and The Dirac Equation) PDF
Fall 2008
Matthew Schwartz
Lecture 10:
Spinors and the Dirac Equation
1 Introduction
1
From non-relativistic quantum mechanics, we already know that the electron has spin 2
. We
usually write it as a doublet
↑
|ψ i = (1)
↓
You might also remember the Pauli matrices
0 1 0 −i 1 0
σ1 = , σ2 = , σ3 = (2)
1 0 i 0 0 −1
which simplify the treatment of spin.
For example, the interaction of the electron with a magnetic field appears in what is some-
times called the Schrodinger-Pauli equation:
K · LK 12×2 − 2µBBK · Kσ ψ
2
p
i∂tψ = Hψ = + V (r) − µBB (3)
2m
e
where µB = 2m is the “Bohr magneton” (the size of the electron’s orbital magnetic moment) and
K = xK × Kp is ethe angular momentum operator. We have also written Kσ = (σ1, σ2, σ3) to call
L
attention to the fact that these σi′s transform as the components of a vector, just like the mag-
netic field Bi. Thus (σ K · BK )ψ is rotationally invariant. This is non-trivial, and only works
because
[σi , σ j ] = 2iεijkσk (4)
which are the same algebraic relations satisfied by infinitesimal rotations (we will review this
shortly).
Keep in mind that σi do not change under rotations – they are always given by Eq. (2) in
any frame. ψ is changing and Bi is changing, and these changes cancel when we write (σ K · BK )ψ.
We could also have written down a rotationally invariant equation of motion for ψ
1 · ∂tψ − ∂iσiψ = 0 (5)
Since ∂i transforms like a 3-vector and so does σiψ, this equation is rotationally invariant. It
turns out it is Lorentz invariant too. In fact, this is just the Dirac equation!
If we write
σ µ = (12×2, σ1, σ2, σ3) (6)
Then it is
σ µ∂ µψ = 0 (7)
which is nice and simple looking. Actually, this is the Dirac equation for a Weyl spinor, which is
not exactly the same as the equation commonly called the Dirac equation.
By the way, it does not follow that this equation is Lorentz invariant just because we’ve
written it as σ µ∂ µ. For example,
(σ µ∂ µ + m)ψ = 0 (8)
is not Lorentz invariant. To understand these enigmatic transformation properties, we have to
know how to construct the Lorentz group out of the σ µ’s and see how those act on ψ. But
instead of just guessing the answer (or looking it up), we’re going to do something a little more
general, and motivate what kinds of transformation properties are possible.
1
2 Section 2
Where Λ is a combination of rotations and boosts. We saw that we could write a Lorentz trans-
formation as the product of 3 rotations and 3 boosts:
1 0 0 0 1 1
0 cos θxy − sin θxy cos θ xz − sin θ xz
1
Λ =
0 sin θx y cosθx y 0
(11)
1 cosθx z − sin θx z
0 0 0 1 sin θ xz cosθ xz sin θxz cos θxz
coshβx sinhβx coshβ y sinhβ y coshβz sinhβz
sinhβx coshβx 1 1
× (12)
1 sinhβ y coshβ y 1
1 1 sinhβz coshβz
This is a particular representation. of the Lorentz group. That is, it is one embedding of the
group into a set of matrices.
The group itself is a mathematical object independent of any particular representation. To
extract the group away from its representations, it is easiest to look at infinitesimal transforma-
tions. We can always write these group elements as
Λ = exp(iθiλi) = 1 + iθiλi + (13)
General representations 3
where θi are 6 numbers, corresponding to the 6 angles θxy , θxz , θ yz , βx y , βx z , β yz and Λi are
called generators. For the regular, vector, representation above, the generators are
0 0 0
0 −1 0 −1 0
λ1 = V12 = i , λ2 = V13 = i , λ3 = V23 = i
1 0 0 0 −1
0 1 0 1 0
0 1 0 1 0 1
1 0 0 0
λ4 = V01 = i , λ5 = V02 = i , λ6 = V03 = i (14)
0 1 0 0
0 0 1 0
The λ ′s and the V ′s are just different notations for the same objects. The generators satisfy:
So now we can define the Lorentz group as the set of transformations generated by these genera-
tors.
Another representation or the Lorentz generators is given
These are the classical generators of angular momentum generalized to include time. You can
check that J µν satisfy the commutation relations of the Lorentz algebra.
Technically, we say the generators make up the Lorentz Algebra so(1,3) which generates the
Lorentz Group SO(1, 3). A common convention is to use lowercase letters for the names of alge-
bras and uppercase letters for groups. As another technicality, we note that it is possible for two
different groups to have the same algebra. For example, the Proper Orthonchronous Lorentz
group and the Lorentz Group have the same algebra, but the Lorentz group has in addition the
discrete symmetries time reversal and parity reversal. It’s a small difference, but it’s a difference
and the groups are not identical.
3 General representations
There is a very nice way to study the representations of the Lorentz group. Start with the rota-
tion generators Ji and the boost generators K j . For example, you could take J1 = V23 , K1 = V01 ,
etc. as a particular representation of this basis. These satisfy
[Ji , J j ] = iǫijkJk (18)
where ǫijk is defined by ε123 = 1 and being antisymmetric in all of its indices (it is called the
totally antisymmetric tensor .) As is probably clear to you already, [Ji , J j ] = iǫijkJk is the
algebra for rotations, so(3) so the Ji generate the subgroup of rotations in 3D.
Now take the linear combinations
1
Ji+ = (Ji + iKi) (21)
2
1
Ji− = (Ji − iKi) (22)
2
4 Section 3
which satisfy
[Ji+, J j+] = iǫijkJk+ (23)
So we have found 2 commuting subgroups of the Lorentz group. The algebra of the J ′s is the
3D rotation algebra, so(3), more commonly called so(2). So we have shown that
(technically, it should really be sl(2, R) = so(1, 1) not su(2), but these algebras are the same up
to some i’s, and physicists usually just say su(2).).
The names so(n) and su(n) stand for special orthogonal algebra and special unitary algebra.
Orthogonal means it preserves a norm defined with transpose: V TV is invariant under SO(n).
Unitary means it preserves a norm defined with adjoint V †V is invariant under SU(n). Special
means the determinant is 1, which just normalizes the elements and gets rid of phases.
The decomposition so(3, 1) = su(2) × su(2) makes studying the representations very easy. We
already know what the representations are of su(2), since this is the algebra of Pauli matrices,
which generates the 3D rotation group SO(3) (so(3) = su(2)). The representations are character-
ized by a quantum number j, and have 2j + 1 elements, labeled by m = jz, So representations of
the Lorentz group are characterized by two numbers A and B. The (A, B) representation has
(2A + 1)(2B + 1) degrees of freedom.
The regular rotation generators are J K = JK + + JK −, where I now use the vector superscript to
call attention the fact that the spins must be added vectorially (remember Clebsch-Gordon coef-
ficients?). The bottom line is that each (A, B) representation contains particles of spin J = A +
B , A + B − 1, , |A − B |. For example,
So we see that the regular 4D vector representation A µ contains spins 1,0 so it corresponds to
1 1
the ( 2 , 2 ) representation of the Lorentz group.
Since the group element is the exponential of the generator Λ = eiλ, unitarity requires that λ † =
λ, that is, that λ be Hermetian. Looking back at our explicit 4D representation for the Lorentz
generators, we see that the rotations
0 0 0
0 −1 0 −1 0
J1 = i , J2 = i , J3 = i
1 0 0 0 −1
0 1 0 1 0
1
Spin 2
representation 5
3.2 summary
In summary, the irreducible representations of the Lorentz group are characterized by two spins
(A, B). Physical particles are characterized by mass, m and spin j. Fields generally describe
particles of many spins, and one has to make sure that a physical theory only has causally prop-
agating particles of the spins we want.
1
4 Spin 2
representation
Now that we know that the irreducible representations of the Lorentz group are characterized by
1
two spins, (A, B), how do we find the explicit representations? In particular, what do the ( 2 , 0)
1
and (0, 2 ) representations look like? We need to find 2x2 matrices that satisfy
[Ji+, J j+] = iǫijkJk+ (31)
1
Similarly, the (0, 2 ) representation is
1 σi
(0, ): Ji+ = 0, Ji− = (38)
2 2
K = JK + + JK −
What does this mean for actual Lorentz transformations? Well, the rotations are J
K
and the boosts are K = i(J− − J+ ) so
1 σi σi
( , 0): Ji = , Ki = − i (39)
2 2 2
1 σi σi
(0, ): Ji = , Ki = i (40)
2 2 2
Since the Pauli matrices are Hermetian σi† = σi we see that the rotations are Hermetian but the
boosts are anti-Hermetian. This is the same as what we found for the vector representation.
Also notice that these two representations are complex conjugates of each other. In contrast, the
vector representation was real .
1
Explicitly, if ψL is a ( 2 , 0) spinor, known also as a left-handed Weyl spinor, then under rota-
tions angles θi and boost angles βi
1 1
ψL = (1 + iθiσi + βiσi + )ψL
1
(iθiσi +βiσi)
ψL → e 2 (41)
2 2
Similarly,
1 1
ψR → eiθiσi − βiσiψR = (1 + iθiσi − βiσi + )ψR (42)
2 2
Infinitesimally,
1
δψL = (iθi + βi)σiψL (43)
2
1
δψR = (iθi − βi)σiψR (44)
2
Note again the angles θi and βi are real numbers. Although we mapped Ji− or Ji+ to 0, we still
have non-trivial action of all the Lorentz generators. So these are faithful irreducible representa-
tions of the Lorentz group. Similarly
1
δψL† = ( − iθi + βi)ψL† σi (45)
2
† 1
δψR = ( − iθi − βi)ψL† σi (46)
2
Which is great. However, this term is not real. But then we can just add the Hermetian conju-
gate, so
†
m ψR ψL + ψL† ψR (51)
is ok.
What about kinetic terms? We could try
†
ψR ψL + ψL† ψR (52)
which is both Lorentz invariant and real. But this is actually
not a very interesting Lagrangian.
ψ1
We can always split up our field into components ψL = ψ , where ψ1 and ψ2 are just regular
2
fields. Then we see that this is just the Lagrangian for a couple of scalars. So it’s not enough to
declare the Lorentz transformation properties of something, the Lagrangian has to force those
transformation properties. In the same way, a vector field is just 4 scalars until we contract it
with ∂ µ in the Lagrangian.
To proceed, let’s look at
ψL† σ jψL (53)
This transforms as
1 1
δψL† σ jψL = ψL† σ j [(iθi + βi)σiψL] + [ψL† ( − iθi + βi)σi]σ jψL (54)
2 2
iθi † β
= ψL(σ jσi − σ jσi)ψL + i ψL† (σiσ j + σ jσi)ψL (55)
2 2
= θiεi jkψL† σkψL + β jψL† ψL (56)
Thus we have found that
βi †
δψL† ψL = ψ σiψL (57)
2 L
δψL† σ jψL = θiεijkψL† σkψL + β jψL† ψL (58)
Similarly,
† βi †
δψR ψR = − ψ σiψL (59)
2 L
†
δψR σ jψR = θiεijkψL† σkψL + − β jψL† ψL (60)
Now, recall how a vector transforms
δV0 = βiV0 (61)
δV j = β jV0 + θiε ij kVk (62)
This is just the transformation of the vector V µL = (ψL† ψL , ψL† σ jψL). Thus
5 Dirac matrices
Expanding them out, the Dirac matrices are
1 0 σi
γ0 = γi = (70)
1 − σi 0
Or,
0 1 0 1
0 1 0 1
γ0 =
1
, γ1 = (71)
0 −1 0
1 0 −1 0
0 −i 0 1
0 i 0 −1
γ2 = , γ3 = (72)
i 0 −1 0
−i 0 1 0
They satisfy
{γ µ , γν } = 2η µν (73)
In the same way that the algebra of the Lorentz group is more fundamental than any particular
representation, the algebra of the γ ′s is more fundamental than any particular representation of
them. We say the γ ′s form the Dirac algebra, which is a special case of a Clifford algebra. This
particular form of the Dirac matrices is known as the Weyl representation.
The Lorentz generators are
i
S µν = [γ µ , γν ] (74)
4
They satisfy the Lorentz algebra for any γ ′s satisfying the Clifford algebra. That is, you can
derive from {γ µ , γν } = 2η µν that
[S µν , S ρσ ] = i(ηνρS µσ − η µρSνσ − ηνσS µρ + η µσSνρ) (75)
In the Weyl representation, the Lorentz generators are
1 σk i σi
Si j = εij k , Ki = S0i = − (76)
2 σk 2 − σi
Or, very explicitly
1 0 1 0 1
1 −1 i − , S23 = 1 1 0
1 0
S12 = , S13 = (77)
2 1 2 0 1 2 0 1
−1 −1 0 1 0
0 −1 0 −1 −1
i −1 0 , S02 = 1 1 0 , S03 = i 1
S01 = (78)
2 0 1 2 0 1 2 1
1 0 −1 0 −1
1 1
These are block diagonal. These are the same generators we used for the ( 2 , 0) and (0, 2 ) repre-
sentations above. It makes it clear that the Dirac representation is reducible, it is the sum of a
left-handed and a right-handed spinor representation.
Rotations 9
6 Rotations
Now let’s see what happens when we rotate by an angle θ in the xy plane. We use
Λ = exp(iθx yV12 ) (80)
How do we exponentiate a matrix? The standard trick is to first diagonalizing it with a unitary
transformation, do the rotation, then transform back. This unitary transformation is like
choosing a direction, except it is purely mathematical as the direction must be complex!
First, for the vector representation
0 0
0 1
= U −1 −1
V12 = i U (81)
−1 0 1
0 0
So,
0
exp( − iθx y)
Λ(θxy) = exp(iθx yV12 ) = U −1 U (82)
exp(iθxy)
0
Λ(2π) = 1 (83)
That is, we rotate 360 degrees and we’re back to where we started.
For the spinor representation
Λs = exp(iθ µνS µν ) (84)
the 12 rotation is already diagonal:
1
1 −1
S12 = (85)
2 1
−1
So,
i
exp( 2 θxy )
i
− exp( 2 θxy )
Λ(θxy) = exp(iθxyS12 ) = (86)
i
exp( 2 θxy )
i
− exp( 2 θxy )
−1
1
Λ(2π) = (87)
−1
1
10 Section 7
Thus a 2π rotation does not bring us back where we started! If we rotate by 4π it would. So we
1
say spinors are spin 2 . What does this mean physically? I have no idea. There are plenty of
1
physical consequences of spinors being spin 2 , but this business of rotating by 2π is not actually
a physical thing you can do.
As an aside, note that the factor of 2 is determined by the normalization of the matrices,
which is set by the Lie Algebra. For each representation by itself, this would be arbitrary, but it
is important for expressions which combine the representations to be invariant.
7 Lorentz Invariants
The γ matrices themselves transform nicely under the Lorentz group.
where µ refers to which γ matrix, and α and β index the elements of that matrix.
Then the equation
{γ µ , γν } = 2η µν (90)
really means
γ µαγγνγβ + γναγγ µγβ = 2η µνδ αβ (91)
And the equation
i
S µν = [γ µ , γν ] (92)
4
Should really be written as
αβ i αγ γβ
S µν = γ µ γν − γναγγ µγβ (93)
4
For an expression like
V µV µ = V µη µνVν = V µ {γ µ , γν }Vν (94)
To be invariant, it must be that µ transforms in the vector representation.
Next consider
ψ † ψ → ψ †Λs† (Λsψ) (95)
For this to be invariant, we would need Λs† = Λs−1, that is, for the representation of the Lorentz
group to be unitary. The spinor representation, like any other representation is not unitary,
because the boost generators are anti-Hermetian.
It is useful to study the properties of the Lorentz generators from the Dirac algebra itself,
without needing to choose a particular basis for the γ µ. First note that
Again, we see that the rotations are unitary and the boosts are not. You can see this from the
explicit representations above. But because we showed it algebraically, it is true in ANY repre-
sentation.
Now, observe that one of the Dirac matrices is Hermetian, γ0 (γ0 is the only Hermetian
Dirac matrix because the metric signature is (1,-1,-1-,1)). Moreover
⇒ γ µ† = γ0 γµγ0 (101)
† ih † † i ih i i
⇒ γ0S µν γ0 = γ0 γ µ , γν γ0 = γ0 γ µ† γ0, γ0 γν† γ0 = [γ µ , γν ] = Sµν (102)
4 4 4
† †
(γ0Λsγ0) = γ0exp(iθ µνSµν ) † γ0 = exp( − iθ µνγ0S µνγ0) = exp( − iθ µνS µν ) = Λs−1 (103)
Then, finally,
ψ † γ0 ψ → ψ †Λs† γ0(Λsψ) = ψ † γ0Λs−1Λsψ = ψ † γ0 ψ (104)
ψ̄ ≡ ψ † γ0 (105)
The point is that ψ̄ transforms according to Λs−1. Thus ψ̄ψ is Lorentz invariant.
We can also construct objects like like
ψ̄ ( − i ∂ − m) = 0 (116)
where the derivative acts to the left. This makes the conjugate equation look more like the orig-
inal Dirac equation.
We can write our spinor as the left and right helicity parts
ψL
ψ= (131)
ψR
The Dirac equation is
− m iσ µDµ ψL
(132)
iσ̄ µD µ − m ψR
meaning
(iσ̄ µ∂ µ + eσ̄ µA µ)ψR = mψL (133)
(iσ µ∂ µ + eσ µA µ)ψL = mψR (134)
So the electron mass mixes the left and right handed states.
In the absence of a mass, this implies
0 = iσ µ∂ µψR = (E + Kσ · Kp )ψR (135)
0 = iσ̄ µ∂ µψL = (E − Kσ · Kp )ψL (136)
So the left and right handed states are eigenstates of the operator Kσ · Kp with opposite eigenvalue.
This operator projects the spin on the momentum direction. We call spin projected on the
direction of motion the helicity, so the left and right handed states have opposite helicity. Since
the mass mixes left and right handed states, massive fermions are not eigenstates of helicity.
This makes ssene – in the electron’s rest frame, there is no direction of motion so helicity is not
well definedl in the massless limit, there is no rest frame, and a left-handed electron can never
turn into a right-handed electron, and a right-handed electron can never turn into a left-handed
electron, even in the presence of electromagnetic fields.
The fact that projection of spin on the direction of momentum is a good quantum number
for massless particles works for massless particles of any spin. For any spin, we will always find
(E ± sJKKp )Ψs = 0, where JK are the rotation generators of spin s. For spin 1/2, JK = σK2 . For pho-
tons, the rotation generators are listed in section 2. For example, Jz = V23 has eigenvalues ± 1
with eigenstates (0, i, 1, 0) and (0, − i, 1, 0). These are the states of circularly polarized light in
the z direction. They are helicity eigenstates. So massless particles always have two helicity
states. It is true for spin 1/2 and spin 1, as we have seen, it is true for gravitons (spin 2),
Rarita-Schwing fields (spin 3/2) and spins s > 2 (although, as we have seen, it is impossible to
have interacting theories with massless fields of spin s > 2).
We have seen that the ψL and ψR states
• do not mix under Lorentz Transformation
• ψL and ψR each have two components on which the σ ′s act. These are the two spin
states of the electron – both left and right handed spinors have 2 spin states.
14 Section 10
1 + γ5 1 − γ5
0 1
PR = = , PL = = (142)
2 1 2 0
For the i = 1 and i = 3, σiTσ2 = σiσ2 = − σ2σ1. For i = 2, σiTσ2 = − σ2σ2 = − σ2σi. So
σiTσ2 = − σ2σi (150)
Then
T 1 T T 1 T
δ(ψL σ2) = (iθi + βi)ψL σi σ2 = ( − iθi − βi) ψL σ2 σi (151)
2 2
Majorana fermions 15
This is an example of the spin-statistics theorem. I think it’s the simplest way to see the
connection between anti-symmetrization and Lorentz invariance.
You know it from quantum mechanics. Fermions come in anti-symmetric states. It may look
more familiar if we use arrows for the ψ1 and ψ2 states:
1
|ψi = √ (|↑i|↓i − |↓i|↑i) (155)
2
We see that wavefunctions are antisymmetric to the exchange of particles. This is the Pauli
exclusion principle. Here we see that it follows simply from Lorentz invariance.
Obviously if the fermion components commute this is just zero. So fermion components can’t
be regular numbers, they must be anticommuting numbers. Such things are called Grassman
numbers and satisfy a Grassman algebra. We will see that it’s easy to make the quantum fields
anti-commute, by demanding that the creation and annihilation operators anti-commute, a p† a q† =
− a q†a p† ,. And we’ll rederive spin-statistics from properties of the S-matrix. But here we are
seeing that if we are to make any sense of spinors classically, or in any way at all, they must
anticommute.
11 Majorana fermions
If we allow fermions to be Grassman numbers, then we can write down a Lagrangian for a single
Weyl spinor with a mass term
m
L = iψL† σ µ∂ µψL + i (ψL† σ2 ψL
⋆ T
− ψL σ2 ψL) (156)
2
These kinds of mass terms are called Majorana masses. Note that this mass term breaks the
symmetry under ψ → eiαψ, since
T T iα T
ψL σ2 ψL → ψLe σ2eiαψL = e2iαψL σ2 ψL (157)
So a particle with a Majorana mass cannot be charged.
The equation of motion for the Majorana fermion is
⋆
σ µ∂ µψL + mσ2 ψL =0 (158)
which follows from the Lagrangian above.
If we have a Majorana fermion, we can still use the Dirac algebra to describe it, but we have
to put it into a 4-component spinor
ψL
ψ= ⋆ (159)
− iσ2 ψL
⋆
This transforms like a Dirac spinor because σ2 ψL transforms like ψR.
Since (in the Weyl basis)
⋆
⋆ 0 σ2 ψL ⋆ iσ2 ψR
− iγ2 ψ = − i = ⋆ (160)
− σ2 0 ψR − iσ2 ψL
16 Section 12
So ψc has the oppisite charge from ψ, which is another reason that Majorana fermions can’t be
charged.
11.1 summary
In summary,
We have seen three types of spinors
• Dirac Spinors: massive, left AND right handed
• Weyl Spinors: massless, left OR right handed
• Majorana spinors: real constrained Dirac spinors.
We are not sure if the neutrino is a Majorana or a Dirac spinor. Weyl spinors play a very
important role in supersymmetry, and in the theory of the Weak interactions. But for QED, we
can just stick with Dirac spinors.
So we can just as easily interpret these solutions as anti-particles going forward in time. Obvi-
ously this interpretation is easier to swallow, but Feynman spent some time showing that there
really is no physically distinguishable difference.
Now back to spinors. The Dirac equation is
ψ̄ ( − i ∂ − eA − m) = 0 (175)
So ψ̄ is a particle with mass m and opposite charge to ψ: the positron.
Since the spinor also satisfies ( + m2)ψ = 0, it follows that p2 = m2. We already know that the
negative frequency solutions p0 = − m are not a problem since the antiparticles
where the square root of a matrix means diagonalize it and then take the square root. In prac-
tice, we will always pick p along the z axis, so we don’t really need to know how to make sense
of these formulas. But they are useful because they are explicitly Lorentz covariant so we can
manipulate them cleanly.
= 2mδss ′ (193)
Solving the Dirac equation 19
We taken ξs† ξs ′ = δss ′ by convention. This is the normalization for the inner product .
For m = 0 this way of writing the condition is not useful. Instead we can just set the normal-
ization in the frame of the particle;s direction by normalizing
√ ! √ !
† p · σ ξs † p · σ ξs ′
us us = √ √ = 2Eξs† ξs ′ = 2Eδss ′ (194)
p · σ̄ ξs p · σ̄ ξs ′
√
which has the same 2E factors we’ve been using.
We can also compute the outer product
2
X
us(p)ūs(p) = p + m (195)
s=1
1 0
where the sum is over s = 1, 2 corresponding to ξ = 0
and ξ = 1
. For the antiparticles,
2
X
vs(p)v̄s(p) = p − m (196)
s=1
It is also true that
2
X 2
X
usv̄s = ūsvs = 0 (197)
s=1 s=1
So when we sum Lorentz indices or internal spinor indices, we use an inner product and get a
number. When we sum over polarizations/spins, we get a matrix.