Lecture Notes
Lecture Notes
Y. X. Zhao
Department of Physics, The University of Hong Kong, Hong Kong, China
E-mail: [email protected]
1 Group cohomology 1
1.1 The vertex viewpoint 1
1.2 The bond viewpoint 2
1.3 The long exact sequence 3
2 Space groups 4
2.1 H 1 (P, Rd /L) 4
2.2 H 2 (P, L) 5
2.3 An example with mirror symmetry 5
1 Group cohomology
–1–
Let C n (G, A) be the set of equivariant functions from Gn+1 to A. A function
f : Gn+1 → A is called equivariant if
δ n+1 ◦ δ n = 0. (1.5)
where
Z n (G, A) := ker δ n , B n (G, A) := imδ n−1 . (1.8)
An immediate result is that
H 0 (G, A) = AG . (1.9)
Here, AG is the subset of A, where each element is invariant under the action of all
g ∈ G.
–2–
Then,
δf (1, h1 , h1 h2 , · · · , h1 · · · hn+1 )
=f (h1 , h1 h2 , · · · , h1 h2 · · · hn+1 ) − f (1, h1 h2 , · · · , h1 h2 · · · hn+1 ) + · · ·
(1.12)
=ρh1 f (1, h2 , · · · , h2 h3 · · · hn+1 ) − f (1, h1 h2 , · · · , h1 h2 · · · hn+1 ) + · · ·
=ρh f¯(h2 , h3 , · · · , hn+1 ) − f¯(h1 h2 , h3 , · · · , hn+1 ) + · · · .
1
Hence, we obtain
dn f¯(h1 , · · · , hn+1 ) =ρh1 f¯(h2 , h3 , · · · , hn+1 ) + (−1)n+1 f¯(h1 , h2 , · · · , hn )
X n
(1.13)
+ f¯(h1 , · · · , hi−1 , hi hi+1 , hi+2 , · · · , hn+1 )
i=1
–3–
2 Space groups
Given a point group P and a lattice L, all possible space groups can be classified by
H 2 (P, L) ∼ = H 1 (P, Rd /L), which is explained in the following.
Here, L is spanned by a set of primitive lattice vectors ei ∈ Rd with i =
1, 2, · · · , d, i.e.,
L = {ni ei , ni ∈ Z}. (2.1)
Clearly, L is an Abelian group under the addition of vectors. The action of P on L
is inherited from the natural action of O(d) on Rd , since P ⊂ O(d) and L ⊂ Rd . For
each R ∈ P , the action may be explicitly written by
d
X
Rei = ej R̂ji (2.2)
j=1
with R̂ji ∈ Z.
For a space group Γ associated with the lattice L and the point group P , L is a
normal subgroup of Γ, and the quotient group Γ/L is isomorphic to P , i.e., we have
the short exact sequence,
0 → L → Γ → P → 1. (2.3)
Compared with (τR1 R2 , R1 R2 ), the difference of τR1 R2 from τR1 + R1 τR2 should be
an element of L, as written by
This is just the 1-cocycle equation for H 1 (P, Rd /L), or alternatively τ ∈ Z 1 (P, Rd /L).
There are two centers, the coordinate origin and the center respected by O(d). If
we shift the displacement of the two centers by r, then τR for each R ∈ P is changed
by
drR = Rr − r. (2.6)
It is noticed that dr satisfies the 1-cocycle equation above, and dr ∈ B 1 (P, Rd /L). It
is clear that changing the displacement of two centers does not change the isomorphic
class of the space group or the space group for short, and hence each space group
corresponds to an element of H 1 (P, Rd /L).
–4–
2.2 H 2 (P, L)
H 1 (P, Rd /L) is isomorphic to H 2 (P, L), which can be explained as follows in the
context of space groups. Each τ corresponds to a function ω : P × P → L by
Then, it is straightforward to check that ω satisfies the 2-cocycle equation for H 2 (P, L),
as given by
and therefore we have ω ∈ Z 2 (P, L). For the lifting of τR , τR + ℓ(R) is equally well
for an arbitrary ℓ(R) ∈ L, since both (τR , R) and (τR + ℓ(R), R) are mapped to R
in the short exact sequence for Γ. Accordingly, ω(R1 , R2 ) is changed by
Here, we observe that dℓ ∈ B 2 (P, L). Hence, each space group corresponds to an
element of H 2 (P, L).
and L is spanned by
e1 = [1, 0]T , e2 = [0, 1]T . (2.11)
We only need to find all possible τM = [a, b]T . There is only one nontrivial equation,
τM + M τM = 0 mod L. The left-hand side is simplified by
−1 0 a a 0
+ = . (2.12)
0 1 b b 2b
0 0
τM = ′
, τM = , (2.14)
0 1/2
τM and τM
′
correspond to wallpaper groups P m and P g, respectively.
–5–
3 Projective representations in quantum mechanics
Let H be a Hilbert space representing a symmetry group G. Each group element
g ∈ G is represented by a unitary or anti-unitary operator Lg . Thus, in general we
assume a G-action c on U (1), i.e.,
cg (u) = u or u∗ , (3.1)
for all g1 , g2 ∈ G.
Since for all g ∈ G Lg are unitary or anti-unitary operators, the multiplication
of Lg satisfies the associativity, i.e.,
–6–
Considering the associativity of three AR ,
we find that
ν(R, R) = [ν(R, R)]∗ . (3.8)
Hence, there are two solutions,
and accordingly,
Z 2,c (Z2 , U (1)) ∼
= Z2 . (3.10)
In fact, B 2,c (Z2 , U (1)) = 0. This is because A2R = ν(R, R) is invariant under any
modification of the phase of AR , i.e.,
T 2 = (−1)2s 1. (3.14)
UT H ∗ UT† = H. (3.17)
–7–
But if C is preserved by the system, then the Hamiltonian satisfies the following
identity:
UC H ∗ UC† = −H. (3.18)
C is called the particle-hole symmetry because the above equation requires that the
spectrum of H be symmetric with respect to the zero energy.
In the presence of both C and T , we may consider the projective representations
of ZT2 × ZC2 with both C and T represented by anti-unitary operators AT /C . Then,
for a given multiplier ν, the commutation relation of AT and AC is modified as
AT AC = eiα AC AT (3.19)
with eiα = ν(C, T )/ν(T, C). We can always modify the phases of the operators as
Then, for the equivalent representation ÃC , ÃT and ÃC commute with each other:
Then,
(ÃT C )2 = 1 (3.23)
T 2 = ηT , C 2 =ηC , S 2 = 1, i2 = −1,
(3.26)
[T , C] = 0, {T , i} = 0, {C, i} = 0.
–8–
Non-chiral classes Chiral classes
A AI D AII C AIII BDI DIII CII CI
T 0 +1 0 −1 0 0 +1 −1 −1 +1
C 0 0 +1 0 −1 0 +1 +1 −1 −1
S 0 0 0 0 0 1 1 1 1 1
Table 1. Tenfold symmetry classes. The signs ±1 denote the square of the symmetry
operators, and 0 denotes the absence of the symmetry.
G = N ⋊ P, (3.27)
g = np (3.28)
with n ∈ N and p ∈ P . On the other hand, varying n over N and p over P , the
product np exhausts G. Since N is a normal subgroup of G, the conjugation by
elements of P is a P -action on N , which is denoted as
Here, σ and α are just the restrictions of ν on N and P , respectively, namely the
restricted multipliers. Moreover, for µ is a multiplier, σ and g satisfy the following
consistency relations:
–9–
• There is the canonical action of P on the set of multipliers Z 2 (N, A) of N ,
which is induced from the conjugation action of P on N . Explicitly, under
the action, p sends σ to σ ◦ (cp × cp ). The P -action on Z 2 (N, A) preserves
cohomological classes, i.e., it is trivial on H 2 (N, A). This is ensured by the
first consistency relation, which implies that σ and σ ◦ (cp × cp ) are similar to
each other through g(∗, p).
for g̃
λ(n)/λ(cp−1 n) (3.35)
χ(np) = λ(n)
λ(n1 Cp1 (n2 )) λ(n1 cp1 (n2 )) λ(cp1 (n2 ))
= (3.36)
λ(n1 )λ(n2 ) λ(n1 )λ(cp1 n2 ) λ(n2 )
We can learn much from the factorized form about the possibilities of multipliers
of some important groups in physics. In the following subsections, let discuss some
of them.
– 10 –
is a group homomorphism from P to Z 1 (N, A). Formally, this can be expressed as
χ ∈ Z 1 (P, Z 1 (N, A)).
Then, the factorization form becomes
• N =R Recall that if N = R, R
b = R. (q, p/ℏ) ∈ R × R b can be interpreted
as a point in the phase space R × R
b of a particle on a line. Accordingly, the an
α-representation W of R × R satisfies
W (q)W (p) =eipq/ℏ W (p)W (q),
(3.43)
W (q1 )W (q2 ) = W (q2 )W (q1 ), W (p1 )W (p2 ) = W (p2 )W (p1 ).
– 11 –
3.2.3 Translation groups
Let us start with considering the translation group Rd of a d-dimensional flat space.
We denote all elements of Rd as x = (x1 , x2 , · · · , xd )T with every entry xi ∈ R. Here,
g satisfies We note that every multiplier of R is trivial. Then, every multiplier of
R2 = R × R is equivalent to
and
g(x + y, z) = g(x, z)g(y, z), g(z, x + y) = g(z, x)g(z, y). (3.46)
for all x, y, z ∈ R. Hence,
1 2
σ(x1 , x2 ) = eiF12 x1 x2 . (3.47)
Inspired from the simple case of d = 2, we claim the following form of multipliers
Fij xi1 xj2
Pd
σ(x1 , x2 ) = ei i<j , (3.48)
A = F − FT. (3.50)
Since every coboundary is symmetric, the d(d − 1)/2 independent entries of A rep-
resent distinct cohomological classes, i.e.,
– 12 –
The constraint for the restricted multiplier ν is analyzed as follows. For d = 2,
the left-hand side is always equal to 1. But for d ≥ 3, an immediate consequence is
that A = 0, since RT AR = A for all R ∈ SO(d).
For all d ≥ 2, we see ln g(x, R) is additive in x, and therefore g(x, R) takes the
form
g(x, R) = eibR ·x , (3.54)
where bR is a vector determined by R. Then, the identity g(x, R1 R2 ) = g(R2 x, R1 )g(x, R2 )
leads to
bR1 R2 = R2T bR1 + bR2 . (3.55)
Let us introduce Z
b̄ = dµ(R) bR (3.56)
SO(d)
bR = b̄ − RT b̄. (3.57)
eib̄·x2
g(x2 , R1 ) = . (3.58)
eib̄·R1 x2
It is significant to observe that g can always be eliminated by the transformation
e−ib̄·x .
Thus, the equivalence classes of multipliers of E 2 is determined by νR for the
translation subgroup and ω for SO(2). Since H 2 (SO(2), U (1)) = 0, the second
cohomological group is given by
has a natural interpretation. The entry Fij can be interpreted as the gauge flux
through the unit area specified by the ith and jth directions, which can be seen from
Li Lj L−1 −1
i Lj = e
iFij
(3.61)
with Li the translation for a unit length along the ith direction. A uniform magnetic
field Bij obviously respect the translations invariance. However, the gauge connection
– 13 –
Ai adopted to describe B is not invariant under translations. Lb transforms Ai (x)
to Ai (x − b). Since Ai (x) and Ai (x − b) corresponds to the same magnetic field Bij ,
they are related by a gauge transformation, i.e.,
or equivalently Z x+b
ϕb (x) = Ai (y)dy i . (3.63)
x
Then, the translation operator is defined by
R x+b
Ai (y)dy i
Lb ψ(x) = eiϕb (x) ψ(x + b) = ei x ψ(x + b), (3.64)
B = F − F T = 2A. (3.65)
It is significant to note that for translation groups the gauge invariance in quantum
mechanics is now consistent with the coboundary invariance of cohomological classes.
Then, we further consider rotations R ∈ SO(d). In two dimensions, no rotation
can change B12 , i.e., any uniform magnetic field can preserve the rotation symmetry
group SO(2). But in more than two dimensions, Bij can be rotated to −Bij . For
instance, B12 can be rotated for π through the x direction to −B12 . Hence, any
nonzero Bij breaks the symmetry group SO(d) with d ≥ 3. Consistently, the compo-
nent of translation subgroup is present in H 2 (E 2 , U (1)), but not in H 2 (E d , U (1)) for
d ≥ 3. Moreover, the trivial and nontrivial classes of H 2 (SO(d), U (1)) with d ≥ 3
can be realized by integral spins and half-integral spins, respectively.
(x, v) ∈ R3 × R3 . (3.66)
The quotient group G/N = SO(3) × R consists of rotations R and time translations
t, which denoted by
(R, t) ∈ SO(3) × R. (3.67)
The action of SO(3) × R on R3 × R3 is given by
– 14 –
Hence, the group multiplication is given by
• Similar to the Euclidean group, for the Galilean group, g can also be trivialized.
The justification can be obtained from generalizing the argument for E d . Since
g(∗, (R, t)) corresponds to a homomorphism from SO(3) × R to R c3 × R c3 ∼=
R × R . Hence,
3 3
– 15 –
• For the multiplier α of SO(3)×R, we observe that SO(3)×R is a direct product
group. Hence, we can apply the factorization form to α. It turns out that α is
always similar to a multiplier of SO(3). Hence, we set
λ
ν(((x1 , v1 ), (R1 , t1 )), ((x2 , v2 ), (R2 , t2 ))) = exp[i (x1 ·v2 −x2 ·v1 )]ω(R1 , R2 ). (3.80)
2
Later when discussing the σ-representations of the Galilean group, we shall see ℏλ is
the mass of a particle, and [ω] ∈ Z2 = {±1} corresponds to integral and half-integral
spins, respectively.
G = Z2 × Z2 × · · · × Z2 , (3.82)
– 16 –
Here, a†i and aj are the particle creation and annihilation operators at sites i and
j, respectively. Hij represents the hopping amplitudes tij from site j to i if i ̸= j
and the onsite energy ϵi at site i if i = j. H is a Hermitian matrix and called the
one-particle Hamiltonian of the tight-binding model.
Each hopping amplitude tij may have a phase eiϕij (such that tij = |tij |eiϕij ),
which is called the gauge connection of the lattice model. For each closed loop C
formed by successive hoppings, one can compute the product WC of the phases of all
the hopping amplitudes involved,
Y
WC = eiϕij (4.2)
⟨ij⟩∈C
Here, WC is called the Wilson loop operator of the loop C, and the gauge flux ΦC
through the loop C is given by
WC = e−iΦC . (4.3)
For each site i, we may change the phase of a†i for each i by an arbitrary eiθ .
i
Accordingly, the hopping amplitudes are transformed as tij 7→ eiθ tij e−iθ , which lead
i j
A′ = RART . (4.5)
where the phase assigned to the ith site is eiθR . Then, the physical symmetry operator
i
ρ(R) = GR R. (4.6)
– 17 –
That is, after the spatial transformation R, the gauge transformation GR is needed
to recover the original gauge connection configuration A. Notably, it is ρ(R) = GR R
that commutes with the Hamiltonian H, i.e.,
[ρ(R), H] = 0. (4.7)
where GR (i) = eiθR , namely the phase assigned to site i, and R(i) is the site trans-
i
formed from i by R.
Then, we consider the successive action of two spatial symmetries, ρ(R1 ) =
GR1 R1 and ρ(R2 ) = GR2 R2 . Their product is given by
with
∆G (R1 , R2 ) = GR1 R1 GR2 R1−1 /GR1 R2 . (4.10)
– 18 –
4.2 The classification with time-reversal invariance
Let us consider the symmetry group:
G × ZT2 . (4.14)
Here, G is a symmetry group whose representations are unitary, and ZT2 is the two-
element group generated by time reversal T . T should be represented by an anti-
unitary operator.
Then, each group element can be written as gT a with g ∈ G and a = 0, 1.
Suppose that under the projective representation ρ, the phase factor λ arises through
We shall prove that by appropriately modifying the phase of each operator ρ(gT a ),
we can always transform λ(g1 T a1 , g2 T a2 ) into the form,
H 2,c (G × ZT2 ) ∼
= H 2 (G, Z2 ) × Z2 . (4.17)
Note that ρ(T ) is an anti-unitary operator, i.e., ρ(T )c = c∗ ρ(T ) for c ∈ C. Hence,
We further modify the operators for the other half of group elements as
(4.21)
p
ρ̃(gT ) := λ(g, T )λ(T, g)ρ(gT ),
for all g ∈ G. Note that ρ̃(T ) = ρ(T ). Then, one observes that
– 19 –
Let λ̃ denote the phase factor for ρ̃. Restricting on G, λ̃ satisfies
for all g1 , g2 ∈ G. The left-hand side commutes with ρ̃(T ), so does the right-hand
side. Hence, ν := λ̃|G×G ∈ Z2 = {±1}. On the other hand, λ̃(T, T ) appears in
Above, we have repeatedly used the relations: ρ̃(g)ρ̃(T ) = ρ̃(gT ) and ρ̃(g)ρ̃(T ) =
ρ̃(T )ρ̃(g).
ΓP ∼
= L ⋊D P. (4.26)
and
σ(RT t1 , RT t2 ) γ(t1 + t2 , R)
= , (4.29)
σ(t1 , t2 ) γ(t1 , R)γ(t2 , R)
γ(t, R1 R2 ) = γ(t, R1 )γ(R1T t, R2 ). (4.30)
Let us consider a particular case with both σ and α trivial. That is, the only
nontrivial components of the multiplier correspond to the phase factors modifying the
– 20 –
algebraic relations between translations and point-group elements. Then, equation
(4.29) is simplified to be
which is equivalent to
Here, Lb is the reciprocal lattice dual to L. The similarity between (2.5) and (4.37)
motivates us to interpret κR as momentum-space fractional translations on the re-
ciprocal lattice.
The momentum k corresponds to the representation ρ(t, 1) = eik·t of transla-
tions. Then, the operation of (t′ , R) on k is given by
i.e.,
T k−RT κ )·t
eik·t 7→ ei(R R
(4.36)
Since
κR + RκRT = 0 mod L,
b (4.37)
we obtain
T k+κ
eik·t 7→ ei(R RT
)·t
. (4.38)
Thus, the operation of R on momentum space is given by
R : k 7→ Rk + κR . (4.39)
– 21 –