Week 14
Week 14
mentum
21.1 Historical motivation from physics
Arguably the best motivation for this chapter, is figuring out the mystery of
spin, which first appeared experimentally in non-relativistic quantum physics
as a mysterious kind of angular momentum without a satisfactory classical
mechanics correspondence. This is explained in Section 21.6, after establishing
some terminology.
Nowadays, the motivation is often illustrated by a puzzle in special relativis-
tic quantum mechanics. For C-valued “scalar” wavefunctions on Rd , the stan-
ℏ2
dard Schrödinger equation has kinetic energy term being 2m times the Lapla-
29
P d 2
cian ∆ = − j=1 ∂j . This has no chance of being strictly compatible with
ℏ2
relativity, since the Schrödinger equation iℏ∂t = 2m ∆ has time-derivative and
spatial-derivatives of different orders. Considering the massless case for sim-
plicity (and taking the speed of light c = 1), the relativistic energy-momentum
relation is E 2 = |p|2 (see (4.3)). So under the quantum mechanical correspon-
∂
dence pj ∼ −iℏ∂j and E ∼ H, theP abstract Schrödinger equation iℏ ∂t =H
2 2 2 2 2 2
should have H ∼ E = |p| ∼ −ℏ j ∂j = ℏ ∆. Thus we seek a self-adjoint
2
“first-order square root of ℏ ∆” to serve as the “relativistic Hamiltonian” H.
At first, this seems like an impossible task: ( j −iℏ∂j )2 contains mixed
P
derivatives ∂j ∂k which are not present in the Laplacian. But in 1928, P.A.M.
Dirac found a beautiful trick to get rid of ∂j ∂k . It should be stressed that Dirac
was inspired by a matrix-angular-momentum construction of Pauli in 1927,
which was itself motivated by puzzling experimental results in the properties
of quantum electrons. For (3 + 1)-dimensional Minkowski spacetime, Dirac
formulated his famous first-order Dirac operator D, / which acted on C4 -valued
“functions” ψ. The use of quotation marks is deliberate. As we will learn, ψ
is a “spinor” object, fundamentally distinct from a naı̈ve collection of scalar
functions, much like how a tangent vector field is not merely a collection of d
scalar functions.
The key algebraic idea behind Dirac’s construction can be illustrated from
29
In view of Section 3.5, we should rather view ∆ as a connection Laplacian (Section 23.2)
for a flat connection on a trivializable Hermitian line bundle over Rd .
218
the case of Euclidean R2 , where the Dirac operator is
0 ∂x − i∂y 0 ∂
D/ = −i ≡ −2i ¯ , (21.1)
∂x + i∂y 0 ∂ 0
219
By design, v 2 = −q(v) holds in Cl(V, q), so we have the anticommutation
relations in Cl(V, q),
F : T (V ) → A , v1 ⊗ · · · ⊗ vn 7→ f (v1 ) ⊗ · · · ⊗ f (vn ).
fˆ : T (V )/Iq ≡ Cl(V, q) → A .
220
Proposition 21.2. Cl(V, q) has dimension nk=0 nk = 2n , with a basis given
P
by eI with I running over all multi-indices. Each of Cl0 (V, q) and Cl1 (V, q) has
dimension 2n−1 , except for the case n = 0, where Cl(V, q) = Cl0 (V, q) = K.
Note that Cl0 (V, q) is a subalgebra of Cl(V, q), while Cl1 (V, q) is a comple-
mentary linear subspace.
Cl1 . A linear basis for Cl1 is {1, e1 }, where e21 = −1. So e1 is a square root
of −1, and Cl1 ∼
=C∼ = C as R-algebras.
221
Cl2 . A linear basis for Cl2 is {1, e1 , e2 , e1 e2 }. The elements e1 , e2 , e1 e2 are
anticommuting square roots of −1. So Cl2 is isomorphic to the quaternions
H = spanR {1, I, J, K} as R-algebras, which have multiplication rule
I 2 = J 2 = K 2 = IJK = −1.
222
Cl0 . This is C as a C-algebra.
Oy : v 7→ v − 2⟨y|v⟩y, v ∈ Rn .
223
(Optional.) Let R ∈ O(n) and d = dim ker(R − 1n ). By induction on n − d,
we will show that R is a product of n − d reflections.
• For n − d = 0, R is the identity transformation.
Consider R′ = S ◦ R. Then
and also
R′ (v) = S(R(v)) = v ⇒ v ∈ ker(R′ − 1n ).
Thus
n − dim ker(R′ − 1n ) ≤ n − (d + 1),
and the induction hypothesis applies to R′ : it is the product of at most
n − dim ker(R′ − 1n ) ≤ n − 1 − d
The Cartan–Dieudonné result, Theorem 21.4, shows that any SO(n) ele-
ment is a product of an even number of reflections, Oy1 . . . Oy2k .
Now, observe that in the Clifford algebra, the anticommutation relation,
(21.2), leads to
224
Definition 93. The spin group is the group in Cln generated by products of
even numbers of unit vectors,
Proof. For y1 , y2 ∈ S n−1 , we have y2 y1 = (−y2 )(−y1 ) = y2−1 y1−1 = (y1 y2 )−1 .
Thus, by (21.4),
gv = vg, ∀ v ∈ Rn . (21.6)
225
We conclude that g1 = 0, so g = g0 has no e1 factor. The same argument with
v = e2 , . . . , en shows that g does not contain any e2 , . . . , en−1 , en factor, so g
can only be a scalar in Cln . Together with
g 2 = (y1 · · · y2k )2 = ±y12 . . . y2k
2
= ±1,
we conclude that g ∈ ker χ has to be either +1 or −1. Clearly −1 = e21 and
χ(−1) = 1n , so −1 ∈ ker χ. Thus ker χ = {±1}.
Finally, we sketch why Spin(n) is a Lie group. Let Cln× denote the group of
invertible elements of Cln ; it may be checked to be an open submanifold, thus
a Lie group (this is a general feature of finite-dimensional algebras regarded
as vector space manifolds). The adjoint representation,
Ad : Cln× → GL(Cln ), Ad(a) : b 7→ aba−1 ,
is a continuous map. We can regard χ as the restriction of Ad to the subgroup
Spin(n); here, χ(g) ∈ SO(n) is promoted to an automorphism of Cln . Since
SO(n) is closed, the preimage Spin(n) is a closed subgroup of Cln× , thus a Lie
group (by a general closed subgroup theorem in Lie theory).
Proposition 21.6. For n ≥ 2, the spin group Spin(n) is connected.
Proof. Write g ∈ Spin(n) as g = y1 , . . . y2k with yi ∈ S n−1 . Let γi (t) be a path
in S n−1 connecting yi to y1 , so Π2k
i=1 γi (t) is a path in Spin(n) connecting g to
y12k = (−1)k . This shows that g is connected to either +1 or −1 in Spin(n).
It remains to check that +1 and −1 are connected in Spin(n). For or-
thonormal e1 , e2 , the path
γ(t) = − cos(t) − sin(t)e1 e2 , t ∈ [0, π].
joins −1 to +1 in Cln . But we can factorize
γ(t) = (cos(t/2)e1 + sin(t/2)e2 ) (cos(t/2)e1 − sin(t/2)e2 ) ,
into the product of two paths of unit vectors. So γ is actually a path within
Spin(n).
Remark 6. For n ≥ 3, it may be shown that SO(n) has fundamental group
π1 (SO(n)) = Z2 . For instance,
cos t − sin t 0
γ : t 7→ sin t cos t 0 , t ∈ [0, 2π] (21.7)
0 0 1n−2
226
is a non-trivial loop in SO(n), but the doubled loop with t ∈ [0, 4π] is de-
formable to a constant loop. On the other hand, Spin(n) is simply-connected
(for n ≥ 3). So Spin(n) is also the universal cover of SO(n).
The existence of the spin double-cover has great physical significance. After
applying a full rotation of SO(n) reference frames, a vector in Rn is certainly
returned to itself. We know that quantum mechanics encodes physical config-
urations in complex√Hilbert spaces, so it is conceivable that nature uses some
intricate complex “ vector” which can distinguish between the application of
one full rotation versus two full ones.
Mathematically, we are seeking a quantity
Spin(2) = {a + b e1 e2 : a2 + b2 = 1},
with group inversion given by (a, b) 7→ (a, −b). It may be parametrized by the
1-parameter subgroup generated by e1 e2 ∈ spin(2),
227
So the double-covering map is
χ cos 2θ − sin 2θ 0 −1
Spin(2) ∋ exp(θe1 e2 ) 7→ = exp 2θ ∈ SO(2).
sin 2θ cos 2θ 1 0
(21.9)
ẽ1 = e2 e3 , ẽ2 = e3 e1 .
Thus
Cl30 ∼
= Cl2 = spanR {1, ẽ1 , ẽ2 , ẽ1 ẽ2 } = spanR {e ∼
2 e3 , e3 e1 , e1 e2 } = H. (21.10)
|{z} |{z} |{z}
I J K
||q||2 = a2 + b2 + c2 + d2 .
Cl30 ∋ e3 y = a1 e3 e1 + a2 e3 e2 − a3 ←→ a1 J − a2 I − a3 1 ∈ H,
228
Thus we have an injective homomorphism Spin(3) ,→ Sp(1) of Lie groups.
Since Sp(1) ∼
= S 3 is a connected Lie group of the same dimension30 as Spin(3),
general Lie theory (omitted) shows that Spin(3) ,→ Sp(1) is in fact an isomor-
phism.
There is a further accidental isomorphism Sp(1) ∼ = SU(2). One such iso-
morphism is given below. Represent the R-algebra H as an R-subalgebra of
M2 (C), via
a − id −c − ib
H ∋ a + bI + cJ + dK = . (21.11)
c − ib a + id
It is easy to check that the matrix on the right side belongs to SU(2) iff
a2 + b2 + c2 + d2 = 1 iff the left side is a unit quaternion, i.e., Sp(1) element.
Thus we have (non-canonical) identifications
(21.10) (21.11)
w −z̄
Spin(3) ∼ = Sp(1) ∼ = SU(2) = 2 2
: |w| + |z| = 1 . (21.12)
z w̄
For example,
0 −1 0 · · · 0
1 0
0 ··· 0
E12
0 0
= 0 ··· 0 .
.. .. .. .. ..
. . . . .
0 ··· ··· ··· 0
Next, let us analyze spin(n). The group Cln× is open in the vector space Cln
(exercise), so it is a Lie group whose Lie algebra cl×
n is identified with Cln , at
least as a vector space. Let us check that the Lie bracket on cl×n coincides with
the commutator on Cln . Note that Cln ⊂ End(Cln ) via left-multiplication.
30
SO(3) and Spin(3) are 3-dimensional.
229
So Cln× ⊂ GL(Cln ) is a matrix Lie group. Thus the Lie bracket on cl×
n = Cln
coincides with the commutator Lie bracket on End(Cln ) (restricted to the
subalgebra Cln ). By a further restriction, Spin(n) ⊂ Cln× has spin(n) ⊂ cl×
n
being a Lie subalgebra of Cln with commutator as Lie bracket.
Proposition 21.7. The double-covering homomorphism χ : Spin(n) → SO(n)
induces the Lie algebra isomorphism
χ∗ : spin(n) → so(n)
ei ej 7→ 2Eij , 1 ≤ i < j ≤ n.
Proof. For 1 ≤ i < j ≤ n, consider the curve
γ : t 7→ cos(t) + sin(t)ei ej , 1 ≤ i < j ≤ n,
which lies in Spin(n) by the proof of Prop. 21.6. It has
γ̇(0) = ei ej ∈ Te Spin(n) = spin(n).
The surjective Lie group homomorphism χ induces, by Exercise 18.1, a Lie
algebra map
χ∗ : spin(n) → so(n)
d(χ ◦ γ)
ei ej 7→ .
dt t=0
230
By dimension considerations,
spin(n) = spanR {ei ej : 1 ≤ i < j ≤ n},
so we have completely specified χ∗ , which is thus also an isomorphism.
d
into 2 independent components.
A typical example is a particle with mass m rotating in a circle of radius
r in the j-k plane with constant angular velocity ωjk , its angular momentum
2-form is
L = Ljk dxj ∧ dxk , Ljk = mr2 ωjk .
If the rotating particle has electric charge q (so it is a “current loop”), then it
possess a magnetic moment 2-form P µ, in the sense that when subjected to a
background magnetic 2-form B = j<k Bjk dxj ∧dxk , it has magnetic potential
energy X
−⟨µ, B⟩ = − µjk Bjk .
j<k
31
In physics, a 3D angular momentum is sometimes called a pseudovector, because it
changes sign simply by an unphysical choice of orientation.
231
Here, it is assumed that Bjk are constant over the region of motion32 . Classical
electromagnetic theory gives
q
µ= L, (21.13)
2m
q
with the proportionality constant 2m called the gyromagnetic ratio. Thus
q
magnetic potential energy = − ⟨L, B⟩. (21.14)
2m
For a positively-charged particle, this energy is minimized/maximized when
its L is aligned/antialigned with B.
As one might anticipate, orbital angular momentum becomes tricky to
define for curved Riemannian manifolds, and we shall not attempt to do so.
What is more important is that nature turns out to have another “intrinsic”
source of magnetic moment, which does not arise from an orbiting charge
picture. It is supposed to come from a different corresponding intrinsic notion
of angular momentum, called spin angular momentum, carried by quantum
mechanical point particles (with electric charge). The role of electromagnetism
in this story will be deferred to Section 23.4.
Let us nail down mathematically, the concept of spin angular momentum.
First, we reexamine how orbital angular momentum in Euclidean R3 is pre-
sented as quantum mechanical operators. These are the L1 , L2 , L3 operators
given in Example 11.4. (A better labelling would be L23 , L31 , L12 .) They are
easily checked to satisfy the angular momentum commutation relations
so actually,
1
Ljk ↔ Ejk (21.16)
ih
is giving a representation of so(3) on L2 (R3 ). This generalizes to orbital an-
gular momentum operators representing so(d) on L2 (Rd ).
32
In practice, one assumes the current loop to be very small, e.g., an electron orbiting a
nucleus (which is, strictly speaking, a fictitious classical analogy). In fact, sometimes one
speaks of a magnetic dipole as an informal r → 0 limit of a current loop, keeping µ constant.
232
Orbital angular momentum operators are significant when studying Hamil-
tonians on L2 (Rd ) with SO(d) rotation symmetry where SO(d) acts on the
manifold Rd fixing a rotation centre. Basic examples are H = −∇2 + V with
V = V (r) a potential depending only on the radial coordinate. For such H,
any component of angular momentum, Ljk , commutes with H, and leads to a
conserved quantity in the quantum mechanics sense of Section 11.2. Namely,
there are simultaneous eigenfunctions of H and Ljk , describing describe quan-
tum states with sharply-defined and time-independent expectation values for
angular momentum in the j-k plane. (But beware that the Ljk do not commute
amongst themselves, so there are no quantum states with all components of
angular momentum being simultaneously sharply-defined!) Significantly, one
may deduce from the commutation relations themselves, that the eigenvalues
of Ljk are constrained to be integer multiples of ℏ (Exercise; see any physics
quantum mechanics textbook). Consequently, for an “orbiting” electron with
q = −e and mass me , the orbital angular momentum, thus also magnetic
moment, must be quantized,
eℏ
µjk ∈ integer × . (21.17)
2m
|{z}e
Bohr magneton
233
angular momenta to be the “infinitesimal spin rotation”
iℏ
Σjk := ej ek ∈ spin(d), 1 ≤ j < k ≤ d. (21.18)
2
There is an important extra factor of 21 compared to (21.16), inherited from the
Lie algebra isomorphism χ−1∗ : so(d) → spin(d), which is needed to maintain
the Lie bracket relations of angular momenta, Eq. (21.15). Soon, we will
provide the concrete “spin-representation” of Spin(d), which will realize these
Σij as concrete Hermitian spin angular momentum matrices.
A key point is that each manifold point carries its own intrinsic spin-
representation space, generalizing the tangent space, so spin orbital momen-
tum is a gauge-theoretic concept. The spin-representation is an intrinsic “con-
served” label for the quantum particle in question, independent of whatever
Hamiltonian is specifying the dynamics.
Even n = 2m. The algebra Cl2m ∼ = M2m (C) has only one irreducible repre-
m
sentation (over C) up to equivalence, namely, the defining one on C2 .
Odd n = 2m + 1. The chirality element splits in Cln into two chiral subal-
gebras,
1 ± ωC 1 ± ωC
Cln = Cln+ ⊕ Cln− , Cln± := Cln = Cln .
2 2
(Exercise.) Recall the parity automorphism α induced by v 7→ −v, which
+ −
swaps ω C for −ω C , thus Cl2m+1 for Cl2m+1 . In fact, chasing through the
periodicity isomorphism in Prop. 21.3, and the initial case Cl1 ∼
= C⊕C =
+ −
Cl1 ⊕ Cl1 , we obtain
Cl2m+1 ∼ +
= M2m (C) ⊕ M2m (C) = Cl2m+1 −
⊕ Cl2m+1 . (21.19)
234
Thus, there are two inequivalent representations of Cl2m+1 , according to whether
the second or first copy of M2m (C) is represented as 0; in turn, they are dis-
tinguished by whether the chirality element ω C is represented as +1 or as
−1.
We shall write S ± for the irreducible representation of Cl2m+1 which has
ω C acting as ±1, and refer to them as right-handed (+) and left-handed (−)
respectively33 .
Proof. In Cln , the even subalgebra Cln0 commutes with ω C . So the reduction
of c : Cln0 → End(S + ⊕ S − ) to S ± makes sense. Choose an isomorphism
Cln0 ∼
= Cln−1 , e.g., ẽi = ei en , i = 1, . . . , n − 1 (Exercise 21.1). Then the
chirality element for Cln0 ∼
= Cln−1 would be
n
C
ωn−1 = i 2 ẽ1 · · · ẽn−1 = (e1 en ) · · · (en−1 en )
n (n−1)(n−2)
= i 2 (−1) 2 (e1 · · · en−1 en )en−2
n
n
= i 2 e1 · · · en = ωnC ,
δn : Spin(n) → GL(S),
235
Proposition 21.9. For odd n, δn is an irreducible representation of Spin(n),
and is independent of which irreducible representation of Cln is used.
For even n, δn splits into δn+ ⊕ δn− acting on S + ⊕ S − , with δn+ and δn− being
inequivalent irreducible representations of Spin(n).
Proof. (Optional.) Odd n = 2m + 1. Recall the splitting of Cl2m+1 into chiral
subalgebras, Eq. (21.19), exchanged by the parity automorphism α. Since
0
Cl2m+1 is fixed under α, we have
0 + −
Cl2m+1 = {(φ, α(φ)) ∈ Cl2m+1 ⊕ Cl2m+1 }.
±
So the (inequivalent) irreducible representations c± : Cl2m+1 → Cl2m+1 →
End(S ± ) actually lead to equivalent irreducible representations
∼
=
c± : Cl2m+1
0 ±
→ Cl2m+1 → End(S ± ),
236
distinguished by whether c(ω C ) = c(−e1 e2 e3 ) = ±1. Either choice restricts to
the same representation for Cl30 = spanC {1, e1 e2 , e2 e3 , e3 e1 }, thus of Spin(3).
The standard basis for spin(3) is represented as
0 −i 0 −1 −i 0
c(e2 e3 ) = , c(e3 e1 ) = , c(e1 e2 ) = . (21.22)
−i 0 1 0 0 i
We had found this representation earlier, in Section 21.4.1, Eq. (21.10)–(21.11).
Recall (21.18), that the spin angular momenta are the spin(3) elements
iℏ iℏ iℏ
Σ23 = e2 e3 , Σ31 = e3 e1 , Σ12 = e1 e2 .
2 2 2
±
They are concretely represented on S as the spin angular momentum matrices
ℏ 0 1 ℏ 0 −i ℏ 1 0
S1 := c(Σ23 ) = , S2 := c(Σ31 ) = , S3 := c(Σ12 ) = ,
2 1 0 2 i 0 2 0 −1
(21.23)
first written down by Pauli in 1927.
Clifford multiplication
Proposition 21.10. The spin representation δn is compatible with Clifford
multiplication c(v), v ∈ Rn , in the sense that for all g ∈ Spin(n), v ∈ Rn ,
δn (g)c(v)δn (g −1 ) = c χ(g)(v) .
(21.24)
Proof. Since δn (·) is the restriction of c : Cln → End(S), we have
(21.5)
δn (g)c(v)δn (g −1 ) = c(g)c(v)c(g −1 ) = c(gvg −1 ) = c χ(g)(v) .
237