Clif
Clif
Spin Groups
Thomas Hjortgaard Danielsen
Contents
1 Cliord Algebras 5
1.1 Elementary Properties . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2 Classication of Cliord Algebras . . . . . . . . . . . . . . . . . . 11
1.3 Representation Theory . . . . . . . . . . . . . . . . . . . . . . . . 15
2 Spin Groups 19
2.1 The Cliord Group . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2 Pin and Spin Groups . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3 Double Coverings . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.4 Spin Group Representations . . . . . . . . . . . . . . . . . . . . . 28
2.5 Spin Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.6 The Dirac Operator . . . . . . . . . . . . . . . . . . . . . . . . . 38
3
4
Chapter 1
Cliord Algebras
1.1 Elementary Properties
In this chapter we will introduce the Cliord algebra and discuss some of its
elementary properties. The setting is the following:
Let V be a nite-dimensional vector space over the eld (predominantly R
K
or C) and ϕ : V × V −→ K V . ϕ is said to be
a symmetric bilinear form on
positive (resp. negative ) denite if for all 0 6= v ∈ V we have ϕ(v, v) > 0 (resp.
ϕ(v, v) < 0). ϕ is called non-degenerate if ϕ(v, w) = 0 for all v ∈ V implies
w = 0. From a bilinear form we construct a quadratic form Φ : V −→ K given
by Φ(v) := ϕ(v, v). We can recover the original bilinear form by the polarization
identity :
Φ(u + v) = ϕ(u + v, u + v) = Φ(u) + Φ(v) + 2ϕ(u, v),
hence
1
ϕ(u, v) = (Φ(u + v) − Φ(u) − Φ(v)). (1.1)
2
Thus we have a 1-1 correspondence between symmetric bilinear forms and
quadratic forms. Thus a quadratic form Φ is called positive denite /negative
denite /non-degenerate if ϕ is.
Denition 1.1. Let (V, Φ) be a vector space with a quadratic form Φ. The
associated Cliord algebra Cl(V, Φ) (abbreviated Cl(Φ)) is an associative, unital
algebra over K : V −→ Cl(Φ) obeying the relation i(v)2 =
with a linear map iΦ
Φ(v) · 1 (1 is the unit element of Cl(Φ)). Furthermore (Cl(Φ), iΦ ) should have
the property that for every unital algebra A, and every linear map f : V −→ A
2
satisfying f (v) = Φ(v) · 1 there exists a unique algebra homomorphism fb :
Two questions immediately arise: given (V, Φ), do such objects exist and if
they do, are they unique? Fortunately, the answer to both questions is yes:
Proposition 1.2. For any vector space V with a quadratic form Φ let = be the
two-sided ideal in the tensor algebra T (V ) spanned by all elements of the form
a ⊗ (v ⊗ v − Φ(v) · 1) ⊗ b (for a, b ∈ T (V ) and v ∈ V ). Then T (V )/=, with the
map iΦ : V −→ T (V )/= being π ◦ ι where ι : V −→ T (V ) is the injection of V
into T (V ), and π : T (V ) −→ T (V )/= is the quotient map, is a Cliord algebra,
and any other Cliord algebra over (V, Φ) is isomorphic to this one.
Proof. Uniqueness. Assume that Cl1 (Φ) and Cl2 (Φ) with linear maps i1 :
V −→ Cl1 (Φ) and i2 : V −→ Cl2 (Φ) are Cliord algebras. Since Cl1 (Φ) is
5
6 Chapter 1 Cliord Algebras
x FFF
V
xxx FFi2
x i1
FF
xxx FF
x| "
i2
/
b
Cl1 (Φ) o Cl2 (Φ)
i1
b
We see that
i2 = bi2 ◦ i1 = bi2 ◦ bi1 ◦ i2
and since bi2 ◦ bi1 is the unique map satisfying this, it must be idCl (Φ) . Likewise
2
bi1 ◦ bi2 = idCl (Φ) which means that the two Cliord algebras are isomorphic.
1
This proves uniqueness.
Existence. We now show that Cl(Φ) := T (V )/= is indeed a Cliord algebra. iΦ
2
is easily seen to satisfy iΦ (v) = Φ(v) · 1, where 1 ∈ Cl(Φ) is the coset containing
the unit element of T (V ). Now let f : V −→ A be linear with f (v)2 = Φ(v)·1. By
the universal property of the tensor algebra this map factorizes uniquely through
T (V ) to an algebra homomorphism f 0 : T (V ) −→ A, such that f = f 0 ◦ ι. f 0
0 2
inherits the property (f (v)) = Φ(v) · 1, and consequently
f 0 (v ⊗ v − Φ(v) · 1) = (f 0 (v))2 − Φ(v)f 0 (1) = (f 0 (v))2 − Φ(v) · 1
ε(v) : v1 ∧ · · · ∧ vk 7−→ v ∧ v1 ∧ · · · ∧ vk .
k
X
ι(v) : v1 ∧ · · · ∧ vk 7−→ (−1)j+1 ϕ(v, vj )v1 ∧ · · · ∧ vbj ∧ · · · ∧ vk .
j=1
2
c(v) = Φ(v) · 1
c(u)c(v) + c(v)c(u) = 2ϕ(u, v) · 1.
1.1 Elementary Properties 7
The map c : V −→ EndK (Λ∗ V ) is injective for if c(v) = 0 then c(v)2 = Φ(v)·1 =
0 i.e. v = 0 by non-degeneracy of Φ. By the universal property of Cliord alge-
bras there exists an algebra homomorphism b c : Cl(Φ) −→ EndK (Λ∗ V ) extending
c. By injectivity of c this map is an algebra isomorphism on the subalgebra of
EndK (Λ∗ V ) generated by the set {c(v) | c ∈ V }.
∗
Now we consider the subalgebra of EndK (Λ V ). Any element of the form
c(v1 ) · · · c(vm ) can, by linearity of c and by the relations above be written as a
linear combination of elements of the form c(ei1 ) · · · c(eik ) where i1 < · · · < ik .
Let B denote the set of these element. Hence, these elements span the subalge-
bra. To see that they are linearly independent, observe that c(ei1 ) · · · c(eik ) maps
1 to ei1 ∧ · · · ∧ eik . This evaluation map is linear and since the set {ei1 ∧ · · · ∧ eik }
∗
is a basis for Λ V , B is a linearly independent set, hence a basis. By the iso-
morphism b c the set
{i(ei1 ) · · · i(eik ) | i1 < · · · < ik }
is a basis for Cl(Φ).
If, furthermore, we have a vector space U with quadratic form Θ and linear
maps f : V −→ U and g : U −→ W satisfying Θ(f (v)) = Φ(v) and Ψ(g(u)) =
Θ(u), then g ◦ f = g ◦ f .
Exercise 1. Prove the preceding proposition.
The most interesting examples of Cliord algebras show up when the bilin-
ear form is non-degenerate. For the time being we consider only real vector
spaces and Cliord algebras. A prominent example of a vector space with non-
degenerate bilinear form is (Rp+q , ϕp,q ) where ϕp,q (ei , ej ) = 0 if i 6= j and
(
1, i≤p
ϕp,q (ei , ei ) =
−1, i > p
(e1 , . . . , ep+q is the standard basis of Rp+q ). The associated quadratic form is
denoted Φp,q and the corresponding Cliord algebra is denoted Clp,q . In this
case O(Φp,q ) and SO(Φp,q ) are the well-known orthogonal groups O(p, q) and
SO(p, q).
Example 1.5. 1) Let us consider the vector space R with the single basis
element e1 := 1 and the quadratic formΦ0,1 (x1 e1 ) = −x21 . By Proposition 1.3
{1, e1 } (where 1 is now the unit element of Cl0,1 ) is a basis for Cl0,1 . The fact
2
that e1 = Φ0,1 (e1 ) · 1 = −1 shows that the linear map Cl0,1 3 1 7−→ 1 ∈ C,
∼
e1 7−→ i ∈ C denes an algebra isomorphism Cl0,1 −−→ C and that the injection
R ,−→ Cl0,1 is given by x 7−→ ix, i.e. R sits inside Cl0,1 ∼= C as the imaginary
part. Thus, Cl0,1 is just the eld of complex numbers.
Exercise 2. Consider R2 with the standard basis {e1 , e2 } and the quadratic
2 2
form Φ0,2 (x1 e1 +x2 e2 ) = −x1 −x2 . Show that the corresponding Cliord algebra
can be identied with the algebra of quaternions H. Because of this, Cliord
algebras are sometimes called generalized quaternions.
Now, let (V, ϕ) be any real vector space with non-degenerate bilinear form ϕ,
and let {e1 , . . . , en } be a basis for V . We will consider the matrix (ϕij ) where
ϕij = ϕ(ei , ej ). Since ϕ is symmetric the matrix (ϕij ) is symmetric as well, i.e.
it can be diagonalized. Let λ1 , . . . , λn be the eigenvalues and f1 , . . . , fn a basis
of diagonalizing eigenvectors. This means that
Observe, that none of the eigenvalues are 0 (a 0 eigenvalue would violate non-
degeneracy of ϕ). Arrange the eigenvalues so that λ1 , . . . , λk are all strictly
positive and λk+1 , . . . , λn are all strictly negative. Dene
1
fei = p fi
|λi |
then we see that
(
1, i≤k
ϕ(fei , fej ) = 0 if i 6= j and ϕ(fei , fei ) = .
−1, i > k
A basis satisfying this is called a (real) orthonormal basis w.r.t. ϕ. Thus we have
proven
1
fei = √ fi .
λi
This basis satises ϕ(fei , fej ) = δij , and is thus called a (complex) orthonormal
basis . Thus we have shown
Theorem 1.7 (Classication of Complex Bilinear Forms). Let (V, ϕ) be
a complex vector space with non-degenerate bilinear form. Then there exists an
orthonormal basis for V and the map sending this basis to the standard basis
for Rn is an orthogonal isomorphism (V, ϕ) −→ (Cn , ϕn ).
Appealing to Proposition 1.4 we see that a complex Cliord algebra is iso-
morphic to Cl(Φn ) for some n.
Again we consider the general situation of Cliord algebras over K. Now we
want to equip the Cliord algebra with two involutions t and α which we will
need later in the construction of various subgroups of Cl(Φ).
Proposition 1.8. Each Cliord algebra Cl(Φ) admits a canonical anti-auto-
morphism, i.e. a linear map t : Cl(Φ) −→ Cl(Φ) that for all x, y ∈ Cl(Φ)
satises
t(x · y) = t(y) · t(x) , t ◦ t = idCl(Φ) , t|V = idV .
Proof. Consider the involution J of the tensor algebra given by v1 ⊗· · ·⊗vk 7−→
J
J(a ⊗ v ⊗ v ⊗ b − a ⊗ (Φ(v) · 1) ⊗ b)
= J(b) ⊗ v ⊗ v ⊗ J(a) − J(b) ⊗ (Φ(v) · 1) ⊗ J(a)
= J(b) ⊗ (v ⊗ v − Φ(v) · 1) ⊗ J(a) ∈ =.
graded algebra or a super algebra. Cl0 (Φ) is called the bosonic subalgebra (note
that it is actually a subalgebra), and Cl (Φ) is called the fermionic subspace .
1
b 0 = (A0 ⊗ B 0 ) ⊕ (A1 ⊗ B 1 )
(A⊗B)
b 1 = (A0 ⊗ B 1 ) ⊕ (A1 ⊗ B 0 ).
(A⊗B)
With this at hand we can accomplish out nal task of this section, showing how
Cl reacts to a direct sum of vector spaces. By an orthogonal decomposition of
(V, Φ), we understand a decomposition V = V1 ⊕ V2 such that if v = v1 + v2 we
have Φ(v) = Φ1 (v1 ) + Φ2 (v2 ) (equivalently if ϕ(V1 , V2 ) = 0).
g(v) = v1 ⊗ 1 + 1 ⊗ v2 .
A quick calculation shows that g(v)2 = Φ(v)(1 ⊗ 1) and thus by the univer-
sal property of Cl there exists an algebra homomorphism gb : Cl(V, Φ) −→
Cl(V1 , Φ1 )⊗
b Cl(V2 , Φ2 ) extending g . To see that this is indeed an isomorphism
choose a basis {e1 , . . . , em } for V1 and a basis {f1 , . . . , fn } for V2 and put them
together to a basis {e1 , . . . , fn } for V . A basis for Cl(V, Φ) consists of elements of
the form ei1 · · · eik fj1 · · · fjl where i1 < · · · < ik , k ≤ m and j1 < · · · < jl , l ≤ n.
Similarly a basis for Cl(V1 , Φ1 )⊗ b Cl(V2 , Φ2 ) is given by ei1 · · · eik ⊗ fj1 · · · fjl
(with the same restrictions on the indices as above). One can verify that
(where {e1 , . . . , ep+q } denotes the usual standard basis for Rp+q ). The associated
Cliord algebra was denoted Clp,q . As we saw in Example 1.5, we have
Cl0,1 ∼
=C and Cl0,2 ∼
= H. (1.4)
Cl1,0 ∼
=R⊕R and Cl2,0 ∼
= Cl1,1 ∼
= R(2) (1.5)
1 For a proof consult for instance [Lawson and Michelson], pp 26-27, Proposition 4.2.
12 Chapter 1 Cliord Algebras
f (ei ) · f (ej ) + f (ej ) · f (ei ) = (e0i ⊗ e001 e002 ) · (e0j ⊗ e001 e002 ) + (e0j ⊗ e001 e002 ) · (e0i ⊗ e001 e002 )
= (e0i e0j ) ⊗ (e001 e002 e001 e002 ) + (e0j e0i ) ⊗ (e001 e002 e001 e002 )
= −(e0i e0j + e0j e0i ) ⊗ (e001 e001 e002 e002 ) = −2δij 1 ⊗ 1
because e0i e0j + e0j e0i = 2δij · 1, as {e01 , . . . , e0n } is basis for Rn orthonormal w.r.t.
Φn,0 . For n + 1 ≤ i, j ≤ n + 2 we have
the Dimension Theorem from linear algebra tells us that fb is also injective. Thus
fb is the desired isomorphism.
The second isomorphism is proved in exactly the same way, and we avoid
repeating ourselves.
The proof of the third isomorphism is essentially the same as the two rst. Let
{e1 , . . . , ep+1 , ε1 , . . . , εq+1 } be an orthogonal basis for Rp+q+2 (i.e. ϕ(v, w) = 0
when v 6= w ) with the quadratic form Φp+1,q+1 , such that Φp+1,q+1 (ei ) = 1,
Φp+1,q+1 (εi ) = −1, and let {e01 , . . . , e0p , ε01 , . . . , ε0q } and {e001 , ε001 } be similar bases
p+q 1+1
for R and R (and thereby generators for the Cliord algebras Clp+1,q+1 ,
Clp,q and Cl1,1 respectively). We now dene a linear map f : Rp+q+2 −→
Clp,q ⊗ Cl1,1 by
( (
e0i ⊗ e001 ε001 if 1 ≤ i ≤ p ε0j ⊗ e001 ε001 if 1 ≤ j ≤ q
f (ei ) = , f (ε j ) = .
1 ⊗ e001 if i = p + 1 1 ⊗ ε001 if j = q + 1
Proof. Using the two rst isomorphisms from Lemma 1.12 a couple of times
yields
Cl0,n+8 ∼
= Cln+6,0 ⊗ Cl0,2 ∼
= ··· ∼= Cl0,n ⊗ Cl2,0 ⊗ Cl0,2 ⊗ Cl2,0 ⊗ Cl0,2
∼
= Cl0,n ⊗ Cl2,0 ⊗ Cl2,0 ⊗ Cl0,2 ⊗ Cl0,2
where the last isomorphism follows from the fact that for arbitrary real algebras
A and B we have A ⊗ B ∼ = B ⊗ A. From (1.4) we have Cl0,2 ∼
= H, and from (1.5)
that Cl2,0 = R(2). Thus, using H ⊗R H ∼
∼ = R(4) from Lemma 1.11, we get
Cl2,0 ⊗ Cl2,0 ⊗ Cl0,2 ⊗ Cl0,2 ∼
= R(2) ⊗ R(2) ⊗ (H ⊗R H) ∼
= R(4) ⊗ R(4)
∼
= R(16).
This completes the proof of the rst isomorphism. The proof of the second
isomorphism is identical to this.
Now it's evident that once we know the Cliord algebras Cl0,0 , Cl0,1 , . . . , Cl0,7
and Cl0,0 , Cl1,0 , . . . , Cl7,0 we know all of them: Consider Clp,q and assume that
p ≥ q . Then by the third isomorphism in Lemma 1.12 we have that Clp,q is iso-
⊗q ∼
morphic to Clp−q,0 ⊗(Cl1,1 ) = Clp−q,0 ⊗R(2q ), and Clp−q,0 can be expressed
as a tensor product of Clk,0 (with 0 ≤ k ≤ 7) and some copies of R(16). We can
do the same if q ≥ p. Using the isomorphisms from Lemma 1.11 one obtains
the table of Cliord algebras in Appendix A. From this table we see that any
Cliord algebra is either a matrix algebra or a sum of two such algebras, as we
pointed out in the beginning of this section.
Corollary 1.15. Clp,q is a direct sum of two matrix algebras exactly when
p−q ≡1 (mod 4) .
We conclude this treatment by proving
For the rest of this section we will consider complex Cliord algebras. It turns
out that complex Cliord algebras behave even nicer than their real counter-
parts. As we have already seen a complex Cliord algebra associated with a
non-degenerate bilinear form is isomorphic to Cl(Φn ). The fact that there is
only one index on Φ and not two as in the real case, indicates some sort of
simplication.
But rst we introduce a way of turning real vector spaces/algebras into com-
plex ones:
14 Chapter 1 Cliord Algebras
ΦC (v ⊗ λ) = λ2 Φ(v).
In the same way we dene the complexication of a real algebra A by A :=
C
Theorem 1.19 (Cartan-Bott II). We have the following 2- periodicity : ClCn+2 ∼
=
n ⊗C Cl2 ,
ClC and furthermore that ClC2 ∼
= C(2).
C
Proof. Invoking Lemma 1.12 and Lemma 1.18 we obtain the following chain
of isomorphisms:
ClC ∼ ∼
n+2 = Cl0,n+2 ⊗R C = Cln,0 ⊗R C ⊗R Cl0,2
∼
= Cln,0 ⊗R (C ⊗C C) ⊗R Cl0,2 ∼
= (Cln,0 ⊗R C) ⊗C (C ⊗R Cl0,2 )
∼
= ClC ⊗C ClC .
n 2
ρ0 (x) ◦ f = f ◦ ρ(x)
for all x ∈ A. Two representations are called equivalent if there exists an inter-
twiner between them which is also an isomorphism of vector spaces.
Just as we had complexication of a real algebra, we can complexify complex
representations: If ρ : A −→ End(V ) is a representation of a real algebra on a
complex vector space V , we dene the complexication ρC : AC −→ End(V ) by
ρC (x⊗λ) = λρ(v). The notions of invariant subspaces and irreducibility of a rep-
resentation and its complexied are closely related as the following proposition
shows
ρC (x ⊗ λ)W = λρ(x)W ⊆ W.
Proposition 1.23. The matrix algebra K(n) has only one irreducible K-repre-
sentation, namely the dening representation i.e. the natural isomorphism πn :
K(n) −−→ EndK (Kn ). The algebra K(n) ⊕ K(n) has exactly 2 inequivalent irre-
∼
We saw in the previous section that Clp,q is of the form K(n) ⊕ K(n) i
p−q ≡1 (mod 4) and a matrix algebra otherwise. This observation along side
with the preceding proposition yields the number of irreducible representations
of the real Cliord algebras.
16 Chapter 1 Cliord Algebras
But there is a slight problem here, in that the real Cliord algebra is not
always a real matrix algebra or a sum of real matrix algebras. Thus the irre-
ducible representations of Proposition 1.23 need not be real! For instance Cl1,4 ∼
=
H(2)⊕H(2), and Proposition 1.23 gives us two irreducible H-representations over
H2 ! But fortunately we can always turn a complex or quaternionic representa-
tion into a real representation if we just remember to adjust the dimension.
2
Without going further into details with keeping track of the dimensions we have
shown:
k
For n = 2k or n = 2k + 1 like above, dene ∆n := C2 . The elements of ∆n
are called Dirac spinors or complex n-spinors . The irreducible representations
of ClC
n are representations on ∆n .
In the case where n is odd we want to single out κ0n and dene κn := κ0n
which is just the composition
∼ π
κn = κ0n : ClC
n −−→ EndC (∆n ) ⊕ EndC (∆n ) −
−−1
→ EndC (∆n )
of the isomorphism with the projection π1 onto the rst component. Hence, for
each n we have an irreducible complex representation on ∆n called the complex
spin representation .
Finally, let's try to break up the action of the Cliord algebra into smaller
pieces and see how they act on the spinors. This requires introduction of the
so-called volume element. It is well-known that, has a unique volume
Λ∗ (Rn )
element Ω given unambiguously, in any orthonormal basis {e1 , . . . , en }, by Ω =
e1 ∧ · · · ∧ en . Applying the quantization map to this yields an element ω :=
Q(Ω) ∈ Cl0,n , also called the volume element , given by ω = e1 · · · en . For the
C
complex Cliord algebra Cln we dene the volume element by
n+1
ωC := ib 2 c ω.
2 k
In the case n = 2k we note that ω = (−1) and that ω commutes with every
element of Cl00,2k , while ω anti-commutes Cl10,2k , for instance we see that
ClC C C
2k = (Cl2k )+ ⊕ (Cl2k )−
where, in fact
1
(ClC C
2k )± = 2 (1 ± ωC ) Cl2k .
We can combine this splitting with the splitting given by the involution α to
obtain the spaces
0 1 C 0 1 1 C 1
(ClC
2k )± := 2 (1 ± ωC )(Cl2k ) and (ClC
2k )± := 2 (1 ± ωC )(Cl2k ) .
End(∆+
For the next proposition we will identify 2k ) as the subspace of End(∆2k )
+ −
consisting of maps ∆2k −→ ∆2k which map ∆2k to itself and which map ∆2k to
− ± ∓
0, and in a similar way we identify End(∆2k ) and Hom(∆2k , ∆2k ) as subspaces
of End(∆2k ).
i.e.κ2k (ξ)ψ ∈ ∆+ C 0
2k . For this result we only used that (ξ ∈ Cl2k ) , but to show
− 1 −
that κ2k (ξ) is 0 on ∆2k we need that ξ = (1 + ωC )ξ . Let ψ ∈ ∆2k , i.e. κ2k (1 +
2
ωC )ψ = 0, then
Spin Groups
2.1 The Cliord Group
Having introduced the Cliord algebra Cl(Φ), we proceed to dene its Cliord
group Γ(Φ). The point of doing this is that the Cliord group has two particu-
larly interesting subgroups, the pin and spin groups.
Let V be a nite-dimensional real vector space, and let Φ be a quadratic form
on V.
Denition 2.1. Let Cl∗ (Φ) denote the multiplicative group of invertible ele-
ments of Cl(Φ). The Cliord group (by some also called the Lipschitz group ) of
Cl(Φ) is the group
One mechanically veries that Γ(Φ) is truly a group. The group Cl∗ (Φ) is
an open subgroup of Cl(Φ), just as Aut(V ) is an open subgroup of End(V ) (at
least when V is nite-dimensional). In the latter case the Lie algebra of Aut(V )
is just End(V ) with the commutator bracket. In the same way the Lie algebra
cl∗ (Φ) of the group Cl∗ (Φ) is just Cl(Φ) with the commutator bracket.
It is very conspicuous from the denition that we are interested in a particular
representation Λ : Γ(Φ) −→ Aut(V ), namely Γ(Φ) 3 x 7−→ Λx where Λx : V −→
V is given by Λx (v) = α(x)vx−1 and called the twisted adjoint representation
(indeed, the form of Λx is reminiscent of the adjoint representation of a Lie
group).
One reason for considering Λx (v) = α(x)vx−1 instead of Adx (v) = xvx−1
is that the twisted adjoint representation keeps track of an otherwise annoying
sign. W.r.t. the bilinear form ϕ we dene, for x ∈ V with Φ(x) 6= 0 the reection
sx through the hyperplane orthogonal to x by
ϕ(v, x)
sx (v) := v − 2 x.
Φ(x)
Proposition 2.2. For any x ∈ V with Φ(x) 6= 0 we have x ∈ Γ(Φ), and the
map Λx : V −→ V given by Λx (v) = α(x)vx−1 is the reection through the
hyperplane orthogonal to x.
Proof. First, since x2 = Φ(x) · 1 6= 0, x is invertible in Cl(Φ) with inverse
19
20 Chapter 2 Spin Groups
x−1 = 1
Φ(x) x ∈V. Using this and Eq. (1.2) we see that
ϕ(v, x)
Λx v = v − 2 x = sx (v) ∈ V.
Φ(x)
vx0 = x0 v (2.1)
−vx1 = x1 v. (2.2)
e1 a0 + e21 a1 = a0 e1 + e1 a1 e1 = e1 a0 − e21 a1
where the last equality follows from a0 ∈ Cl0 (Φ) and a1 ∈ Cl1 (Φ). We deduce
2
0 = e1 a1 = Φ(e1 )a1 ; since Φ(e1 ) 6= 0, we have a1 = 0. So, the polynomial
expression for x0 does not contain e1 . Proceeding inductively, we realize that x0
does not contain any of the terms e1 , . . . , en and so must have the form x0 = t · 1
where t ∈ R.
We can apply an analogous argument to x1 to conclude that neither does
the polynomial expression for x1 contain any of the terms e1 , . . . , en . However,
x1 ∈ Cl1 (Φ), so x1 = 0.
Thus, x = x0 + x1 = t · 1. Since x 6= 0, we must have t ∈ R∗ . This shows
ker Λ ⊆ R∗ · 1; the reverse inclusion is obvious.
The assumption that Φ is non-degenerate is not redundant. Consider a real
vector space V dim V ≥ 2. If Φ ≡ 0, then Cl(V, Φ) = Λ∗ V, the exterior
with
algebra of V. Consider the element x = 1 + e1 e2 . Clearly, x−1 = 1 − e1 e2 , and
we have
α(1 + e1 e2 ) v (1 + e1 e2 )−1 = (1 + e1 e2 ) v (1 − e1 e2 ) = v,
Lemma 2.5. When Φ is non-degenerate, the norm possesses the following prop-
erties :
1) If v ∈ V then N (v) = Φ(v) · 1, i.e. N is an extension of Φ to the algebra.
t(α(z)) = z.
Therefore the norm becomes N (z) = zz = |z|2 , i.e. the square of the usual norm.
Exercise 4. Carry out a similar calculation for Cl0,2 ∼ = H.
For the next proposition recall the denition of the orthogonal group O(Φ)
(when Φ is non-degenerate!) as the endomorphisms f : V −→ V satisfying
Φ(f (v)) = Φ(v).
Proposition 2.7. For any x ∈ Γ(Φ), the map Λx is an orthogonal transforma-
tion of V. That is, Λ(Γ(Φ)) ⊆ O(Φ).
Proof. Let x ∈ Γ(Φ) and use the fact that N is a homomorphism:
The spin group consists of those elements of Pin(Φ) that are linear combinations
of even-degree elements:
It's not too dicult to verify that Pin(Φ) and Spin(Φ) really are groups. We
will write Pin(p, q) Spin(p, q) for the pin and spin groups associated with
and
Clp,q and likewise Pin(n) := Pin(0, n) and Spin(n) := Spin(0, n) (not to be
confused with the complex pin and spin groups Pin(Φn ) and Spin(Φn ) sitting
C
inside Cln !). Recall that the algebra structure in Cl0,n is that generated by the
2
relations v · v = −kvk · 1.
Since Φ is assumed to be non-degenerate, any real Cliord algebra is (up to
isomorphism) of the form Clp,q and thus we can in fact always assume the real
pin and spin groups to be of the form Pin(p, q) and Spin(p, q).
In addition to the complex pin and spin groups there are also complexica-
tions of the real pin/spin groups, namely let(V, Φ) be a real quadratic vector
space and dene Pinc (Φ) ⊆ Cl(Φ)C = Cl(Φ)⊗C to be the subgroup of invertible
c
elements in Cl(Φ) ⊗ C generated by Pin(Φ) ⊗ 1 and 1 ⊗ U(1). Similarly, Spin (Φ)
is dened as the subgroup generated by Spin(Φ) ⊗ 1 and 1 ⊗ U(1). These are
not to be confused with Spin(ΦC )!
Proposition 2.9. There is a Lie group isomorphism
∼
Spinc (Φ) −−→ Spin(Φ) ×±1 U(1)
i.e. the Spinc (Φ) is quotient of Spin(Φ) × U(1) where we identify (1, 1) and
(−1, −1). In particular Spinc (Φ) is connected if Spin(Φ) is connected.
Proof. We consider the smooth map Spin(Φ) × U(1) −→ Spinc (Φ) given by
(g, z) 7−→ gz . This is a surjective Lie group homomorphism. The kernel consists
−1
of elements (g, z) such that gz = 1 i.e. g = z ∈ U(1). But these elements are
rare, there are only ±1. Thus the kernel equals {±(1, 1)}, and the map above
descends to an isomorphism.
±1, i.e. u1 · · · up ∈ Pin(Φ). Hence T = Λu1 · · · Λup = Λu1 ···up which proves that
Λ|Pin(Φ) is surjective. The kernel is easily calculated:
To prove the analogous statement for Spin(Φ) we need rst to show that Λ maps
Spin(Φ) SO(Φ). Assume, for contradiction, that this is not the case, i.e. that
to
an element f ∈ O(Φ) \ SO(Φ) exists such that Λx = f for some x ∈ Spin(Φ). By
the Cartan-Dieudonne Theorem f can be written as an odd number of reections
f = s1 ◦ · · · ◦ s2k+1 , and to each such reection sj corresponds a vector uj so
that sj = Λuj . In other words we have
By Proposition 2.3 x−1 u1 · · · u2k+1 = λ·1 for some λ ∈ R∗ , i.e. x = λ−1 u1 · · · u2k+1
1
and this is a contradiction, since u1 · · · u2k+1 ∈ Cl (Φ). Thus Λ maps Spin(Φ)
to SO(Φ) and since {±1} ⊆ Spin(Φ) the kernel is still Z2 .
Example 2.11. Let's calculate Pin(1) and Spin(1). They are subgroups of
Cl0,1 ∼
= C and the vector space, from which the Cliord algebra originates, is
R which sits inside C as the imaginary line (cf. Example 1.5). In R we have
just two unit vectors, namely ±1. They sit in C as ±i and they generate Pin(1)
which are thus seen to be isomorphic to Z4 (the fourth roots of unity). Spin(1)
is then generated by products of two such unit vectors, i.e. Spin(1) = Z2 .
1, e1 , e2 , e3 , e1 e2 , e1 e3 , e2 e3 , e1 e2 e3
forms a basis for the Cliord algebra Cl0,3 . The group Spin(3) can be written
xvx−1 = u + λe1 e2 e3
24 Chapter 2 Spin Groups
But e1 e2 e3 = −e3 e2 e1 = e1 e2 e3 , so
Spin(3) × U(1) ∼
= SU(2) × U(1) −→ U(2)
by (A, z) 7−→ zA. It is well-known that this map is a surjective Lie group
homomorphism, and it is easily seen that the kernel is (I2 , 1) (I2 is just the 2×2
identity matrix). Thus a Lie group isomorphism is induced on the quotient, i.e.
∼
an isomorphism Spinc (3) −−→ U(2).
We can argue almost similarly to calculate Spinc (4). For brevity, put
nA 0
o
1
S(U(2) × U(2)) := | A1 , A2 ∈ U(2), det A1 det A2 = 1
0 A2
Dene a map
Spin(4) × U(1) ∼
= SU(2) × SU(2) × U(1) −→ S(U(2) × U(2))
zA1 0
(A1 , A2 , z) 7−→ .
0 zA2
This map is again a surjective Lie group homomorphism, as above, and the
∼
kernel is ±(I2 , I2 , 1), and thus it induces an isomorphism Spinc (4) −−→ S(U(2)×
U(2)).
Note that Pin(Φ) and Spin(Φ) are Lie groups. This is because the multiplica-
tive group Cl∗ (Φ) of invertible elements is an open subset of Cl(Φ) (this is a
general result for algebras) which is a nite-dimensional linear space, hence a
manifold. Thus, Cl∗ (Φ) is a manifold, and since multiplication and inversion are
∗
smooth maps, it is a Lie group. As Pin(Φ) is a closed subgroup of Cl (Φ) (since
N is continuous) and Spin(Φ) is a closed subgroup of Pin(Φ) (since Cl0 (Φ) is a
closed subspace of Cl(Φ)), they are Lie groups.
2.3 Double Coverings 25
is open, being the union of open sets g·O. Being a restriction of an open map p to
an open set V , p|V is an open map, and p|V is therefore a homeomorphism.
26 Chapter 2 Spin Groups
Corollary 2.17. The groups Pin(n) and Spin(n) are compact groups.
Proof. Since O(n) and SO(n) are compact and Pin(n) and Spin(n) are nite
coverings, the result follows from standard covering space theory.
This does not hold in general, for instance Spin(3, 1)0 = SL(2, C) which is
denitely not compact. If the identity component in non-compact, the entire
group must be non-compact.
∼
In Chapter 1 we constructed explicit isomorphisms Ψ : O(p, q) −−→ O(q, p)
∼
and Ψ : SO(p, q) −−→ SO(q, p). A natural question would be if these isomor-
phisms lift to isomorphisms on the level of pin and spin groups? For the pin
groups the answer is no, in general. But for the spin groups the answer is ar-
mative:
Putting p=0 in Theorem 2.16 and combining with Theorem 2.19 we get the
main result:
Corollary 2.20. For n ≥ 3, the group Spin(n) is the universal (double ) cover-
ing of SO(n).
G
It's a classical fact from dierential geometry that for connected Lie groups
and H, F : G −→ H is a covering map, then the induced map F∗ : g −→ h is
if
2
an isomorphism. Spin(n) is connected and simply connected (Theorem 2.19),
and SO(n) is connected. By Theorem 2.16, the homomorphism Λ : Spin(n) −→
SO(n) is a covering map, and thus we have an isomorphism
∼
Λ∗ : spin(n) −−→ so(n),
n(n−1)
in particular, dim spin(n) = dim so(n) = 2 .
Let's investigate this map a little further. Recall that the Lie algebra of Cl∗0,n
is just the Cliord algebra Cl0,n itself with the commutator bracket. Spin(n) is
a Lie subgroup of Cl∗0,n and hence the Lie algebra spin(n) is a Lie subalgebra of
Cl0,n .
Proposition 2.21. Let {e1 , . . . , en } be an orthonormal basis for Rn ,
then spin(n) ⊆ Cl0,n is spanned by elements of the form ei ej where 1 ≤ i <
j ≤ n. Furthermore Λ∗ maps ei ej to the matrix 2Bij ∈ so(n) where Bij is the
n × n-matrix which is −1 in its ij 'th entry and 1 in its ji'th entry.
Proof. Consider the curve
Spin(n) since it is the product of two unit vectors, and its value at
It is a curve in
t=0 1. Upon dierentiating at t = 0 we get 2ei ej , which
is the neutral element
is then an element of T1 Spin(n) ∼ = spin(n). They are all linearly independent i
Cl0,n hence also in spin(n), and there are exactly n(n−1)
2 of them, i.e. they span
spin(n).
Now, Λ is the restriction of the twisted adjoint representation to Spin(n), and
0 −1 n
since Spin(n) ⊆ Cl0,n we get Λ(g)v = gvg for g ∈ Spin(n) and v ∈ R . As for
the usual adjoint representation one can calculate
in particular we get
0,
6 i, j
k=
Λ∗ (ei ej )ek = ei ej ek − ek ei ej = 2ej , k=i
−2ei , k=j
We see that Λ∗ (ei ej ) acts in the same way on Rn as the matrix 2Bij , thus we
may identify Λ∗ (ei ej ) = 2Bij .
We can rephrase the rst part of this proposition by saying that under the
∼
symbol map σ : Cl0,n −−→ Λ∗ Rn , the Lie algebra spin(n) gets mapped to Λ2 Rn .
We end the section by a short description of some covering properties of
Spinc (Φ).
Proposition 2.22. The map Λc : Spinc (Φ) −→ SO(Φ)×U(1) given by [g, z] 7−→
(Λ(g), z 2 ) is a double covering.
Proof. Λ(−g) = Λ(g) and (−z)2 =
It is easy to see that it is well-dened (since
2
z ). It is a covering map because it is the quotient map of the even Z2 -action on
Spinc (Φ) given by (−1, [g, z]) 7−→ [−g, z] = [g, −z]. Thus it follows from Lemma
2.15.
We stress that we use the term spin representation for the irreducible Cliord
algebra representations and spinor representation for the associated spin group
representations.
Of course we can in a similar way dene the real spinor representation of
Spin(n) by restricting ρn to Spin(n), but we will not consider them here.
inside Spin(2k + 1).3 Denoting the injection ι the following diagram commutes:
Spin(2k)
κ2k
/ Aut(∆2k )
ι id
Spin(2k + 1) / Aut(∆2k+1 )
κ2k+1
Now, put H := ker κ2k+1 ⊆ Spin(2k + 1). The goal is to verify H = {1}, but
rst we show that H ∩ Spin(2k) = {1}. The inclusion ⊇ follows since 1 clearly
sits in ker κ2k+1 . Now, assume that h ∈ H ∩ Spin(2k). In particular, h ∈ H and
h = ι(e h) for some eh ∈ Spin(2k). Since h sits in H , κ2k+1 (h) = id∆2k+1 . From
the commutativity of the diagram it follows that κ2k (e h) = id∆2k+1 . But since
κ2k is injective, eh must be 1, and so must h. This shows H ∩ Spin(2k) ⊆ {1}.
Identifying elements A ∈ SO(n) with elements in SO(2k + 1) of the form
diag(A, 1), we obtain SO(2k) ⊆ SO(2k + 1) like the spin groups. Recall that
Λ : Spin(2k + 1) −→ SO(2k + 1) is a surjective homomorphism. Thus Λ(H) is a
normal subgroup of SO(2k + 1), since H as a kernel is normal in Spin(2k + 1).
Now we claim
Λ(H ∩ Spin(2k)) = Λ(H) ∩ SO(2k).
The inclusion ⊆ is obvious, and ⊇ follows from the surjectivity ofΛ. Hence
we have Λ(H) ∩ SO(2k) = {I} (here, I denotes the identity matrix). We want
to show that Λ(H) = {I}, so let A ∈ Λ(H) ⊆ SO(2k + 1). Its characteristic
polynomial is of odd degree, and it thus has a real root. As A ∈ SO(2k + 1),
all eigenvalues have modulus 1. Moreover, A preserves orientation, so this root
must be 1. Denote the corresponding eigenvector by v0 and choose an ordered,
2k+1
positively oriented orthonormal basis for R containing v0 as the last vector.
If B denotes the change-of-basis matrix, then we have the block diagonal matrix
BAB −1 = diag(A, e 1), where Ae ∈ SO(2k) and 1 is the unit of R. We can now
−1 −1
identify A with BAB
e . Hence, BAB ∈ SO(2k), and since Λ(H) was normal,
−1
we also have BAB ∈ Λ(H). All together we have BAB −1 ∈ Λ(H) ∩ SO(2k) =
{I} and so A = I .
Now we have Λ(H) = {I}. We have two possibilities: H = {1} or H = {±1}.
But −1 cannot be in the kernel of the spinor representation (because it's not in
the kernel of the spin representation, from which it came). Therefore, H = {1},
and κ2k+1 is injective.
This theorem is not as innocent as it might look. It actually tells us that the
spinor representations do not arise as lifts of SO(n)-representations, since a lift
of an SO(n)-representation necessarily contains {±1} in its kernel.
We now want to decompose the spinor representations into irreducible repre-
sentations. To this end we need:
Lemma 2.25. For any complex vector space V the endomorphism algebra
End(V ) is a simple algebra, i.e. the only ideals are the trivial ones. In par-
ticular if dim W < dim V , then any homomorphism ϕ : End(V ) −→ End(W ) is
trivial : ϕ ≡ 0.
Proof. Let n be the dimension of V V . Then we can think
and x a basis for
of End(V ) as the algebra of complex n × n-matrices. Now let I ⊆ End(V )
be any non-zero ideal, and let 6 a ∈ I . Then a has an eigenvalue λ 6= 0
0 =
(because C is algebraically closed). By a suitable basis transformation, given by
3 If Cl
0,2k is generated by {e1 , . . . , e2k } and Cl0,2k+1 is generated by {e1 , . . . , e2k+1 } then
0 0
we have a linear injection ι : Cl0,2k ,−→ Cl0,2k+1 by dening ι(ej ) = e0j . This restricts to an
injection ι : Spin(2k) ,−→ Spin(2k + 1).
30 Chapter 2 Spin Groups
ϕ : End(∆2k ) −→ End(W ).
W was a proper subspace of ∆2k+1 = ∆2k , so dim W < dim ∆2k . Lemma 2.25
now guarantees that ϕ ≡ 0. Since ϕ is an extension of κ2k+1 , this should also be
zero. As κ2k+1 is injective by Theorem 2.24 this is a contradiction, so W cannot
be invariant.
Example 2.27. Again we consider our favorite spin group Spin(3). What is
the spinor representation of Spin(3)? Recall that Spin(3) ∼
= SU(2) and that for
each n SU(2) has exactly one irreducible representation πn of dimension n + 1
on the space of homogenous n-degree polynomials in two variables. The spinor
representation κ3 is a 2-dimensional irreducible representation, thus it must be
equivalent to π1 .
representations κ±
2k on ∆2k such that κ2k = κ2k ⊕ κ2k .
± + −
so κ2k (g)ψ ∈ ∆+
2k . Likewise if ψ is a negative Weyl spinor.
0
ϕ : (ClC
2k ) −→ End(W ).
0 ∼ ∼
(ClC C
2k ) = Cl2k−1 = End(∆2k−1 ) ⊕ End(∆2k−1 ).
The covering space results in the previous section yields the following result
A close examination of the proofs above will reveal that nothing is used which
does not hold for Spinc (n) as well. We may therefore summarize the results above
c
in the following statement about the spin -representations:
Theorem 2.31. For each n the spinc -representation κcn is faithful. If n is odd,
the representation is irreducible and if n is even it splits in a direct sum of two
irreducible representations κcn = (κcn )+ ⊕ (κcn )− where (κcn )± are representations
on the space ∆± n.
Often, the vector bundle in question will be the tangent bundle of some man-
ifold M (provided of course that this manifold is oriented and has been given
a metric). Then we will write PSO (M ) and PSpin (M ) for the oriented frame
bundle resp. the spin bundle. If T M has a spin structure we say that M is a
spin manifold .
Let's try and see this from a local perspective. It is well-known that a principal
G-bundle over a manifold M can be (uniquely) described by the following data:
a cover (Uα )α∈A of open sets and for each pair (α, β) ∈ A × A a smooth map
(the transition function ) gαβ : Uαβ −→ G (where Uαβ := Uα ∩ Uβ ) satisfying
that gαα (x) = 1 for all x ∈ Uα and satisfying the cocycle condition ,
for all triples (α, β, γ) ∈ A × A × A and for all x ∈ Uαβγ . This collection of
data is called a gluing cocycle . SO(n)-bundle
So if we consider our principal
PSO (E), then there exists a cover (Uα ) and transition functions gαβ : Uαβ −→
SO(n) satisfying the cocycle condition. The existence of a spin structure is then
eαβ : Uαβ −→ Spin(n) over Λ satisfying the
equivalent to the existence of lifts g
cocycle condition. It is well-known that such a set of lifts exists if and only if the
second Stiefel-Whitney class is zero. In fact this follows more or less by denition
of the second Stiefel-Whitney class (in the setting of Cech cohomology). In the
armative case the possible spin structures on M are parametrized by elements
of the rst Cech cohomology group Ȟ 1 (M ; Z2 ) which is, of course, isomorphic to
1
the singular cohomology group H (M ; Z2 ) with coecients in Z2 . For instance
S for n ≥ 3 admits a spin structure, since H 2 (S n ; Z2 ) = 0, so the second Stiefel-
n
1 n
Whitney class can be nothing but 0. Since H (S ; Z2 ) = 0 the spin structure
must be unique.
The spin group Spin(n) has a distinguished complex representation, called the
spinor representation κn : Spin(n) −→ Aut(∆n ) where ∆n = C
2k
and where
k = b n2 c, n
the integer part of
2 . This is a faithful representation (hence does
not descend to a representation of SO(n)) and when n is odd it is irreducible.
For n even, it decomposes into a direct sum of two irreducible representations
−
κn = κ+ +
n ⊕ κn and the corresponding representation spaces are denoted ∆n and
−
∆n respectively.
Given a principal Spin(n)-bundle π : Q −→ M (originating from a spin
structure, say) we can form the associated complex vector bundle w.r.t. the
spinor representation, namely S := Q ×κn ∆n which is the quotient of the direct
product Q × ∆n under the equivalence relation (p, v) = (p · g, κn (g −1 )v) and
with projection q : S −→ M given by q([p, v]) = π(p). This is called the spinor
bundle and sections of this bundle are called spinors or Dirac spinors . If n is
even this bundle splits, in the same way as the representation, into two bundles
S = S+ ⊕ S− where in fact S± is the associated bundle Q ×κ± ∆±
n. Sections
Weyl spinors
n
of these vector bundles are called positive resp. negative or even
resp. odd chiral spinors . We will discuss these vector bundles and some of their
properties in more detail in the next section when we dene the Dirac operator.
Next, recall how the Lie group Spinc (n) is dened: It is the group inside
Cl0,n ⊗ C generated by Spin(n) ⊗ 1 and 1 ⊗ U(1). Equivalently, Spinc (n) =
Spin(n) ×±1 U(1), the quotient where we collapse the subgroup {±(1, 1)}. Thus,
in Spin(n) × U(1) we identify (g, z) with (−g, −z). The equivalence class con-
c
taining (g, z) will be denoted [g, z]. This is usually how we will view Spin (n).
c c
We can dene a Lie group homomorphism ρ : Spin (n) −→ SO(n) by
[g, z] 7−→ Λ(g) (again, Λ : Spin(n) −→ SO(n) is the double covering). This
is well-dened, since Λ(−g) = Λ(g), however, contrary to Λ, this is no longer a
covering map, since its bers are not discrete. Instead we may view this map as
an n-dimensional representation of Spinc (n).
2.5 Spin Structures 33
c
PSpin (E)1 / PSpin
c
(E)2
LLL rr
LLL r
L rrrc
Φc1 LLL rr
% yrr Φ 2
PSO (E)
c
The set of isomorphism classes of spin -structures on E is denoted Spinc (E). In
the case where E happens to be the tangent bundle of an oriented Riemannian
c
manifold a manifold equipped with a spin -structure is called a spinc -manifold .
c
The set of isomorphism classes of spin -structures on M is denoted Spinc (M ).
c
PSpin (E) := PSpin (E) ×±1 U(1)
more precisely we take the product PSpin (E) × U(1) and mod out by the equiv-
alence relation ∼ given by (p, z) ∼ (p0 , z 0 ) i π(p) = π(p0 ) and (p0 , z 0 ) = ±(p, z).
The space
c
PSpin (E) can then be equipped with a right Spinc (n) action
[p, z] · [g, z 0 ] = [p · g, zz 0 ]
c
which in combination with the projection map π
e : PSpin (E) −→ M given by
c c
π
e([p, z]) = π(p) turns PSpin (E) into a principal Spin (n)-bundle. Finally, dene
Φc : PSpin
c
(E) −→ PSO (E) by Φc ([p, z]) = Φ(p). We see that
The determinant line bundle can be given a hermitian ber metric by dening
h[p, w], [p, w0 ]i := hw, w0 iC = ww0 , we simply transfer the usual inner product
on C to L. This is well-dened since
sinceλ(g) ∈ U(1) and we may therefore form the unitary frame bundle L0 :=
PU (L), a principal U(1)-bundle over M .
Let's put a local perspective on this as we did for the spin structures. Our
starting point is again a principal SO(n)-bundle which is given in terms of a cover
(Uα ) and transition functions gαβ : Uαβ −→ SO(n). Picking a spinc -structure
eαβ : Uαβ −→ Spinc (n) along ρc
(if it exists) is then equivalent to picking lifts g
such that the cocycle condition is satised. We see that g eαβ must be of the form
[hαβ , zαβ ] where hαβ : Uαβ −→ Spin(n) and zαβ : Uαβ −→ U(1) are such that
Λ ◦ hαβ = gαβ and such that the pair
is either (−1, −1) or (1, 1), thus (hαβ ) and (zαβ ) need only satisfy the cocycle
condition up to a sign. However if the bundle admits a spin structure, we may
pickhαβ such that it does satisfy the cocycle condition, and then we can pick
c
zαβ = 1, this is the canonical spin -structure of a spin structure.
2
Given the families of maps (hαβ ) and (zαβ ) we can dene λαβ := zαβ which
maps Uαβ U(1). This family of maps satises the cocycle condition and thus
into
represents a principal U(1)-bundle. This bundle is nothing but the unitary frame
bundle of the determinant line bundle. By the remarks above we then conclude
c
that the determinant line bundle of the canonical spin -structure induced by a
spin structure has trivial determinant line bundle.
c c
Let M be a spin -manifold. How many dierent spin -structures does this
manifold have? To shed some light on this question, let Pic∞ (M ) denote the set
of complex Riemannian line bundles (i.e. bundles carrying a sesquilinear, conju-
gate symmetric, positive denit 2-form). This is a group under tensor product,
known as the Picard group . This group is in 1-1 correspondence with the set of
principal U(1)-bundles over M - the map from Pic∞ (M ) to the set of principal
U(1)-bundles is simply given by forming the unitary frame bundle. Recall that
the rst Chern class is a bijection
Moreover, at most nitely many spinc -structures have the same determinant line
bundle.
Proof. First, the action is free: if σ ⊗ L = σ , i.e. if [hαβ , zαβ ζαβ ] = [hαβ , zαβ ],
then we must have zαβ ζαβ = zαβ , i.e. ζαβ = 1 and since this is the gluing cocycle
for L, this bundle must be trivial.
c
The action is transitive: Assume we have two spin -structures σ1 and σ2
(i) (i) (1) (2)
given by gluing cocycles [hαβ , zαβ ]. Since Λ(hαβ ) = Λ(hαβ ) = gαβ we must have
(1) (2) (1) (2)
hαβ = ±hαβ . By a change of sign if necessary we can thus assume hαβ = hαβ .
(2) (1)
Now put ζαβ := zαβ /zαβ . Clearly ζαβ maps into U(1) and we see that
(2) i
(1) (1) zαβ
h
(2) (2) (1) (1)
[hαβ , zαβ ] = hαβ , zαβ (1) = [hαβ , zαβ ζαβ ]
zαβ
S c (E) := PSpin
c
(E) ×κcn ∆n .
vector bundles are isomorphic, giving us the following relations between the
spinor bundles and determinant line bundle L:
Λk S c (E) ∼
= L⊗k/2 . (2.5)
Λk/2 (S c (E)± ) ∼
= L⊗k/4 . (2.6)
The formula (2.5) also explains the name determinant line bundle, since the top
exterior product of a vector space or a vector bundle traditionally is called the
determinant.
We also need to discuss the notion of connections on vector bundles and prin-
cipal bundles. First we recall the denitions
a K-linear map
∇ : Γ(E) −→ Ω1 (M, E)
satisfying the Leibniz rule
∇(f s) = df ⊗ s + f ∇s (2.7)
τ∇ (X, Y ) := ∇X Y − ∇Y X − [X, Y ].
satisfying the following two axioms: (σg )∗ ω = Adg−1 ◦ω and ωp (A] (p)) = A for
allA ∈ g where A ]
is the fundamental vector eld on P determined by A.
Letting σp be the map G −→ P , g 7−→ p · g , then A] is given by
d
A] (p) = dσp (A) = (p · exp(tA)
dt t=0
= de
σg Hp S(P ),
and to see this put p := (Ψc )−1 (x, [g, z]). This is mapped to (Φc (p), [p, 1]) by Ξ
and Ψ maps the rst component to (x, Λ(g)) as it should. Note that
and the identity element is [p, 1]. It is easy to check that these operations are
well-dened. Since ξ 2 = −kξk2 · 1, we get
thus each ber is indeed a Cliord algebra of type (0, n). Thus Cl(E)x (the
ber in the Cliord bundle) is isomorphic to Cl(Ex ) (the Cliord algebra of the
vector space Ex ).
Observe that we have Rn ⊆ Cl0,n Rn
is a ρ-invariant subspace and
and that
that ρ(A)|Rn = A, so ρ restricted to this invariant subspace is just the dening
n
representation of SO(n) on R . We write it as id. But this means that we have
n ∼
the subbundle PSO (E) ×id R = E sitting inside Cl(E). Elements in E ⊆ Cl(E)
n
are characterized by being of the form [p, v] where v ∈ R . In particular we may
view Γ(E) as sitting inside Γ(Cl(E)).
For the purpose of studying spinor bundles, as we will do later in this section,
we need another description of the Cliord bundle. Assume that the oriented
Riemannian vector bundle E has a spin structure π : PSpin (E) −→ M with dou-
ble covering bundle mapΦ : PSpin (E) −→ PSO (E). Consider the representation
Ad : Spin(n) −→ Aut(Cl0,n ) given by
Ad(g)ξ = gξg −1
(recall that Spin(n) sits inside Cl0,n , so multiplication makes sense), then the
following diagram commutes (simply because Ad(g) is the unique extension of
Λ(g) to Cl0,n ):
Spin(n)
Ad / Aut(Cl0,n )
ll l5
llll
l
lll ρ
Λ
lll
SO(n)
From the principal Spin(n)-bundle PSpin (E) and the representation Ad, we can
form the associated bundle PSpin (E) ×Ad Cl0,n .
c c c
Analogously, dene Ad : Spin (n) −→ Aut(Cl0,n ) by Ad ([g, z]) = Ad(g).
This is well-dened since Ad(−g) = Ad(g).
Lemma 2.41. The map Ψ : PSpin (E) ×Ad Cl0,n −→ Cl(E) = PSO (E) ×ρ Cl0,n
given by [p, ξ] 7−→ [Φ(p), ξ] is a well-dened smooth algebra bundle isomorphism.
Similarly, the map Ψc : PSpin c
(E) ×Adc Cl0,n −→ Cl(E) given by [p, ξ] 7−→
[Φ (p), ξ] is a well-dened smooth algebra bundle isomorphism.
c
40 Chapter 2 Spin Groups
∇
e X (V · ψ) = (∇X V ) · ψ + V · (∇
e X ψ) (2.11)
The single most important example of a Dirac bundle is the spinor bundle
S(E) := PSpin (E) ×κn ∆n
as dened in the previous section. To show that it is a Dirac bundle we rst equip
it with an action of the Cliord bundle Cl(E) = PSpin (E) ×Ad Cl0,n . Consider
on the threefold product PSpin (E) × Cl0,n ×∆n the equivalence relation ∼
for any g ∈ Spin(n). Elements in the quotient space PSpin (E) × Cl0,n ×∆n / ∼
are denoted [p, ξ, v]. As in the proof of Lemma 2.41 one can show that the map
given by ([p, ξ], [p, v]) 7−→ [p, ξ, v] is a well-dened bundle isomorphism. This
allows us to dene the Cliord action in the following way: Dene
·g ·g
PSpin (E) × Cl0,n ×∆n / PSpin (E) × ∆n
µ
e
where the rst vertical map is (p, ξ, v) 7−→ (p · g −1 , Ad(g)ξ, κn (g)v) and the
−1
second is (p, v) 7−→ (p·g , κn (g)v). Thus µ e induces a map µ : Cl(E)×S(E) −→
S(E) given explicitly by the formula
This is the desired Cliord action, turning S(E) into a left Cl(E)-module.
Next we want to give S(E) a metric. Inside Cl0,n we have the nite group
(To make the notation in the following less cumbersome, we will simply write
Pn
the action of ρn (ξ) on v as ξ · v .) ξ=
i=1 ai ei is a unit vector in
If Rn ⊆ Cl0,n
then ρn (e) is a unitary operator as well: First we observe
n
DX n
X E
hξ · v, ξ · wi∆n = ai ei · v, aj ej · w
∆n
i=1 j=1
n
X X
= a2i hei · v, ei · wi∆n + ai aj hei · v, ej · wi∆n
i=1 i6=j
n
X X
a2i hv, wi∆n +
= ai aj hei · v, ej · wi∆n + hej · v, ei · wi∆n
i=1 i<j
= hv, wi∆n .
ξ ξ 1
hξ · v, wi∆n = h kξk · (ξ · v), kξk · wi∆n = hξ 2 · v, ξ · wi∆n = −hv, ξ · wi∆n .
kξk2
Note that this implies that ρn (ξ) is skew-adjoint, when ξ is in Cl10,n (the odd
0
part of Cl0,n ) and that ρn (ξ) is self-adjoint when ξ ∈ Cl0,n (the even part of
Cl0,n ). In particular, κn = ρn |Spin(n) is self-adjoint.
42 Chapter 2 Spin Groups
We can readily extend this inner product to a ber metric on S(E), simply
by dening
h[p, v], [p, w]i := hv, wi∆n .
This is well-dened by unitarity of κn (g):
eg∗ ω
σ e = (dΛ)−1 σg∗ Φ∗ ω = (dΛ)−1 Φ∗ σΛ(g)
∗
ω
= (dΛ)−1 Φ∗ Ad(Λ(g −1 )) ◦ ω = (dΛ)−1 ◦ Ad(Λ(g −1 ))(Φ∗ ω)
= Ad(g −1 ) ◦ (dΛ)−1 Φ∗ ω = Ad(g −1 ) ◦ ω
e.
This is the connection 1-form of the lifted connection, since one can easily check
that ker(e
ω )p = Hp PSpin (E).
There is a standard procedure for transforming connections on principal bun-
dles to connections on the associated bundles. In general let π : P −→ M be
a principal G-bundle, ρ : G −→ Aut(V ) a nite-dimensional representation of
G on V and E := P ×ρ V the associated vector bundle. Recall that there is a
1-1 correspondence between sections of E and functions f : P −→ V satisfying
f (p · g) = ρ(g −1 )f (p) (the so-called equivariant functions ). If ψ ∈ Γ(E) we write
ψb for the associated equivariant function.
The map ∇ : X(M ) × Γ(E) −→ Γ(E) given by (X, ψ) 7−→ ∇X ψ where ∇X ψ
where (Bi )i∈I is some basis for the Lie algebra so(n) where m = dim M .
Getting back to the spin bundle case, the connection ω
e onPSpin (E) induces
a connection ∇
e S(E). Let's try to unveil (2.13) in this
on the spinor bundle
particular setting. Lettα : Uα −→ PSpin (E) be a local section of the spin
bundle and put sα := Φ ◦ tα . This is a local section of the frame bundle PSO (E).
Let
eα := t∗α ω
A e and Aα := s∗α ω
denote the local gauge potentials of the connection ω, resp. the connection ω
e.
Then we have
eα = t∗α ω
A e = t∗α ((dΛ)−1 ◦ Φ∗ ω) = (dΛ)−1 ◦ t∗α Φ∗ ω
= (dΛ)−1 ◦ s∗α ω = (dΛ)−1 ◦ Aα ,
for x ∈ Uα (recall that the action of (κn )∗ is just ρn itself ). Now we pick the
usual basis (Bij )i<j for so(n) (where Bij is the n × n-matrix whose ij 'th entry
is −1, the ji'th entry is 1 and all other entries are 0) and write the so(n)-valued
α α α
P
1-form A in terms of this basis: A = i<j Aij Bij . Then (remembering that
dΛ : spin(n) −→ so(n) maps ei ej to 2Bij , where (ei ) is the standard basis for
Rn ) we get
X
−1 α
(∇X ψ)α (x) = Xx ψα − ρn (dΛ)
e Aij (Xx )Bij ψα (x)
i<j
1 X
= X x ψ α − ρn Aα
ij (Xx )ei ej ψα (x)
2 i<j
1X α
= X x ψα − A (Xx )ei ej · ψα (x). (2.15)
2 i<j ij
e X ψ(x), ψ 0 (x)i = Xx ψα − 1
D X E
0
h∇ Aα
ij (Xx )ei ej · ψα (x) , ψα (x)
2 i<j ∆n
1X α
= hXx ψ, ψ 0 (x)i∆n − A (Xx )hei ej · ψα (x), ψα0 (x)i∆n .
2 i<j ij
Note that
hei ej · ψα (x), ψα0 (x)i∆n = −hej · ψα (x), ei · ψα0 (x)i∆n = hψα (x), ej ei · ψα0 (x)i∆n
= −hψα (x), ei ej · ψα0 (x)i∆n ,
and therefore
N
X
Xx ψα = (Xx ψα,i )vi ∈ ∆n .
i=1
N
X N
X
hXx ψα , ψα0 (x)i∆n + hψα (x), Xx ψα0 i∆n = 0
(Xx ψα,i )ψα,i (x) + 0
ψα,i (x)Xx ψα,i
i=1 i=1
N
X
0
= Xx ψα,i ψα,i = Xx hψα , ψα0 i∆n .
i=1
Thus we have proved that the spin connection is compatible with the metric.
Finally we need to check condition 2 in the denition of a Dirac bundle, that
is
∇
e X (Y · ψ)(x) = (∇X Y )(x) + Yx · (∇
e X ψ(x)) (2.16)
for each x ∈ M. Again we use the local expressions, i.e. we consider a cover
(Uα ) which are domains of trivializations of both TM and E. Let Φα denote
the trivializations of the tangent bundle (we may assume it to preserve the
∼
metric on M, i.e. Φx : Tx M −−→ Rm is an isometry). Since ∇
eX is C-linear and
satises the Leibniz rule, it is sucient to verify the above condition for Y = Ek
−1
where Ek (x) = Φα (x, ek ) are local orthonormal vector elds.
But rst, recall the following formula for the dierential of the double covering
dΛ(X)v = Xv − vX
2.6 The Dirac Operator 45
Note that concatenation here means multiplication inside the Cliord algebra,
and not the Cliord action. Recall also formula (2.14) for the local form of
the Levi-Civita connection. It will be used in the following calculations (for
explanations see below):
1 X
(∇
e X (Ek · ψ))α (x) = Xx (ek · ψα ) − Aα
ij (Xx )ei ej ek · ψα (x)
2 i<j
1 X
= ek · (Xx ψα ) − ek Aα
ij (Xx )ei ej · ψα (x)
2 i<j
X
− (Aα
ij (Xx )Bij ek ) · ψα (x)
i<j
= ek · (∇
e X ψ)α (x) + (∇X Ek )α (x) · ψα (x)
and this is precisely the local form of the right-hand side of (2.16). For the rst
identity we used that (Ek · ψ)α = ek · ψα and in the second we used (2.17) as
well as the fact, that ek · is a linear map, and thus commutes with Xx . This
veries condition 2 and hence we have shown that the spinor bundle S(E) is a
Dirac bundle.
The second most important example of a Dirac bundle is the complex spinor
bundle S c (E). We can proceed in almost the same way we did before and we
begin by giving S c (E) a metric: On ∆n we can, as before, nd an inner product
h , i∆n such that κcn (g) is unitary for each g ∈ Spinc (n) and such that ρn (v) is
n
skew-adjoint for any v ∈ R ⊆ Cl0,n . We transfer this inner product to a ber
c
metric on S (E) in the usual way by dening
c c
Consider trivializations of PSpin (E) and Q over Uα , let tα : Uα −→ PSpin (E)
denote the corresponding local section and put sα := Ξ ◦ tα . If
denote the local gauge potentials (local connection 1-forms), then we have eα =
A
c −1 α α
(dΛ ) ◦ A . Note also that since A is an so(n) ⊕ iR-valued 1-form, we may
Aα = Aα
split it:
α α
ω ⊕ AA where Aω is the gauge potential for the connection ω on
PSO (E) and AA is the gauge potential for the connection A on L0 . From (2.18)
α
we get
(dΛc )−1 Aα = (dΛ−1 (Aα 1 α
ω ), 2 AA )
and the induced Lie algebra representation of κcn is just ρn restricted to the
1 α
Lie algebra. κn ( AA (Xx )) is just multiplication with the imaginary number
2
1 α −1 α
2 AA (Xx ) and (dΛ) Aω is already known to us from our discussion above.
Hence we get:
eA 1X α 1
(∇ X ψ)α (x) = Xx ψa − (A )ij (Xx )ei ej · ψα (x) + Aα (Xx )ψα (x). (2.19)
2 i<j ω 2 A
With this local expression at our disposal we can show that the requirements
of Denition 2.42 are satised, and hence that S c (E) is a Dirac bundle. The
arguments are identical to the ones above for the spin bundle and so we skip
them.
Lemma 2.44. The Dirac operator is a rst order dierential operator, and
given a local orthonormal frame {E1 , . . . , Em } for T M over U , the Dirac oper-
ator takes the local form
m
X
/
∂ψ|U = Ej · (∇
e E ψ).
j
(2.21)
j=1
Thus both ∂/ and ∂/2 (called the Dirac Laplacian) are elliptic.
Proposition 2.46. The Dirac operator is formally self-adjoint, i.e. for ψ1 and
ψ2 in Γc (S) (the set of sections of S with compact support ) we have
/ 1 |ψ2 ) = (ψ1 |∂ψ
(∂ψ / 2 ). (2.22)
m
X m
X
/ ψ)α (x) =
(D Ek · ∇
eE ψ
k
(x) = ek · (∇
e E ψ)α (x)
k
α
k=1 k=1
m n
X 1X
= ek · (Ek )x ψα − Aα
ij ((Ek )x )ei ej · ψα (x)
2 i<j
k=1
m
X 1X α
= ek · (Ek )x ψα − A ((Ek )x )ek ei ej · ψα (x) (2.23)
2 i<j ij
k=1
m
X 1X α
/ A ψ)α (x) =
(D ek · (Ek )x ψα − (A )ij ((Ek )x )ek ei ej · ψα
2 i<j ω
k=1
1
+ Aα ((E k )x )ek · ψ α (x) . (2.24)
2 A
where A is a choice of a connection on the determinant line bundle, and where
m is the dimension of the base manifold. If we replace the connection A on the
determinant line bundle by A + β for some β ∈ iΩ1 (M ) the local connection
1-forms change to from Aα α
A to AA + β and from the local expression for the
Dirac operator we immediately deduce
1
/ A+β ψ = D
D / A ψ + β · ψ. (2.25)
2
We will make extensively use of this result.
rst expression ωC denotes the volume section and in the two last expressions
it denotes the volume form ) that
1X α
[∇
e X (ωC · ψ)]α (x) = Xx (ωC · ψ)α − A (Xx )ei ej · (ωC · ψα (x))
2 i<j ij
1X α
= ωC · (Xx ψα ) − ωC · A (Xx )ei ej · ψα (x)
2 i<j ij
= ωC · (∇
e X ψ)α (x).
m
X m
X
/ C · ψ) =
D(ω Ei · ∇
e E (ωC · ψ) =
i
Ei · ω C · ∇
e E (ψ)
i
i=1 i=1
m
X
= −ωC · Ei · ∇ /
e E (ψ) = −ωC · Dψ
i
i=1
!
−
0 /
D
/=
D + .
/
D 0
c
Exactly the same holds true for the geometric Dirac operator. Also the spin -
c c
representation κ2k : Spin (2k) −→ Aut(∆2k ) decomposes into irreducible rep-
+ −
resentation spaces ∆2k = ∆2k ⊕ ∆2k and hence also the complex spinor bundle
S (E) exhibits a splitting S (E) = S c (E)+ ⊕ S c (E)− , relative to which the
c c
!
−
0 /A
D
/A =
D +
/A
D 0
±
where / A : Γ(S c (E)± ) −→ Γ(S c (E)∓ ).
D