Вы находитесь на странице: 1из 23

Chap 1

Manybody wave function and Second

quantization
Ming-Che Chang
Department of Physics, National Taiwan Normal University, Taipei, Taiwan
(Dated: March 7, 2013)

I. MANYBODY WAVE FUNCTION


A. One particle

Hamiltonian and Schrodinger equation for a single particle:


H (1) = T + V ;
H (1) = .

(1)
(2)

The index can be a set of quantum numbers, e.g., = (n, l, m). Whenever required, we
will label, from low energy to high energy, = 1, 2, .
Orthogonality:
h | i = .
Completeness:

| ih | = 1.

(3)

(4)

The summation runs over all possible eigenstates, and 1 is the identity matrix in one-body
Hilbert space H(1) .

B. N particles, non-interacting

Hamiltonian and Schrodinger equation for non-interacting N particles:


H =

N
X

H (1) (ri );

(5)

i=1

H(r1 , r2 , ) = E(r1 , r2 , ).

(6)

(Note: For simplicity, unless necessary, we often write r as r.) Since there is no interaction
between particles, the Schrodinger equation is separable. Assume
(r1 , r2 , , rN ) = (r1 )(r2 ) (rN ),

(7)

H (1) i (ri ) = i i (ri ), i = 1, 2, N.

(8)

then

Manybody eigenstate is a product of 1-particle states,


1 ,2 , (r1 , r2 , , rN ) = 1 (r1 )2 (r2 ) N (rN )
(r1 , r2 , , rN |1 ,2 , ,N ),
2

(9)

with eigenvalues
E1 ,2 , = 1 + 2 + + N .

(10)

Notice that we have used a round bracket | ) to represent a product state. The angular
bracket | i is reserved for later use. Orthogonality:

Completeness:

01 ,02 , ,0N |1 ,2 , ,N = 01 1 02 2 0N N
X

|1 2 N )(1 2 N | = 1.

(11)

(12)

1 2 N

The summation runs over all possible manybody eigenstates {1 2 N }, and 1 is the
(1)
identity matrix in manybody Hilbert space H(N ) = H
H(1){z
H(1)}.
|
N times

C. Permutation symmetry for bosons and fermions

Definition of exchange operator:


Pij (r1 ri rj ) (r1 rj ri ).

(13)

For any boson state and fermion state, we have


Pij (r1 , r2 , ) = (r1 , r2 , ).

(14)

In order for the eigenstates to satisfy this permutation symmetry, we need to symmetrize
the state in Eq. (9).
For bosons,

1 X

P 1 (r1 )2 (r2 ) ,
B
(r
,
r
,

)
=
1
2
1 ,2 ,
N ! allP

(15)

where the summation runs over all possible permutations. Even though we have inserted a

factor 1/ N !, the state may still not be normalized (details later).


For fermions,
1 X
F1 ,2 , (r1 , r2 , ) =
(1)P P 1 (r1 )2 (r2 ) ,
N ! allP
where (1)P 1 for even/odd permutation.

(16)

Both summations in Eqs. (15) and (16) can be written in the following form,

1 (r1 ) 1 (r2 )

B/F
(r
,
r
,

)
=

2 (r1 ) 2 (r2 ) .
1 ,2 , 1 2

N!
..
...

(17)

For fermions, Eq. (16) is equal to the usual determinant | | , in which half of the terms
have negative signs. For bosons, Eq. (15) is equal to the so-called permanent | |+ , in
which all of the terms have positive signs. For fermions, F is equal to zero if i = j for
any two states. This is the Pauli exclusion principle.

D. Normalization

In a bracket notation, Eqs. (15) and (16) (now free from the coordinate representation)
are written as,
1 X
|1 , 2 , }
(1)P P |1 i|2 i
N ! allP
1 X
(1)P |P1 i|P2 i ,
=
N ! allP

(18)

where (P1 , P2 , ) is a permutation of (1 , 2 , ).


Bosons
{1 , 2 , |1 , 2 , } =

1 X
hP10 |P1 ihP20 |P2 i ,
N ! P,P 0

(19)

in which hP10 |P1 i = P10 P1 ... etc. If all of the particles live in different states, then after
using
P10 = P1

(20)

P20 = P2
..
. ,
we would have N ! terms (all of them being one) in the summation. Therefore,
{1 , 2 , |1 , 2 , } = 1.

(21)

However, if there are n1 particles live in state 1, n2 particles live in state 2, ... etc (recall
that 1, 2 are the label of one-particle eigenstate from low energy to high energy), then there
4

are n1 of the Pi states that are the same (= 1). Therefore, there are n1 ! different ways to
match (using Pi0 = Pi ) members in that n1 -group, n2 ! different ways to match members
in the n2 -group ... etc. As a result (pause and think),
{1 , 2 , |1 , 2 , } = n1 !n2 ! .

(22)

That is, in general this ket state is not normalized. The normalized state should be defined
as
1

|1 , 2 , },
n1 ! n2 !
h1 , 2 , |1 , 2 , i = 1.

|1 , 2 , i =

(23)

Fermions
Fermions do not have such a counting problem since no two particles can live in the same
state. We can identify the normalized state |1 , 2 , i as |1 , 2 , } and
h1 , 2 , |1 , 2 , i
1 X
0
(1)P +P hP10 |P1 ihP20 |P2 i ,
=
N ! P,P 0

(24)

= 1.
The factor (1)P

0 +P

must always be +1, since P 0 needs to be exactly the same as P , else

some of the bracket hPi0 |Pi i would be zero.


II. CREATION AND ANNIHILATION OPERATORS

In the formulation of second quantization, operators are written in creation and annihilation operators.

A. Occupation number representation

After symmetrization (for bosons) or antisymmetrization (for fermions), a N -particle


state becomes,
|1 i|2 i |N i |1 , 2 , , N }.

(25)

Similarly, a (N + 1)-particle state becomes,


|i|1 i|2 i |N i |, 1 , 2 , , N }.
5

(26)

The creation operator maps a N -particle state to a (N + 1)-particle state. It is defined


as follows,
a |1 , 2 , , N } = |, 1 , 2 , , N }

(27)

= (1)i1 |1 , 2 , , , , N }.
|{z}
ith

For annihilation operator a , if is the same as one of the i , but not the same as
others, then
ai |1 , 2 , , N }

(28)

= (1)i1 |1 , 2 , , (no i ), N }.
For bosons, if there are multiple coincidence of i s with , then
a |1 , 2 , , N }
N
X
=
h|i i |1 , 2 , , (no i ), N }.
| {z }
i=1

(29)

For fermions, |1 , 2 , , N } with multiple coincidence is zero, of course.


Bosons
In addition to the |1 , 2 , } notation, in which the slots are filled with quantum numbers
for particles at position r1 , r2 , , we introduce the occupation number representation:
|n1 , n2 , i, in which the slots are filled with occupation numbers for states 1, 2, . If the
then n` n` + 1 (see Eq. (23)),
state of the added particle = `,
| |{z}
, 1 , 2 , , N }
|{z} |{z}
|{z}
r
r1
r2
rN
p p
p
n1 ! n2 ! (n` + 1)!
=

(30)

| n1 , n2 , , n` + 1, i.
| {z }
|{z} |{z}

Notice that the number of slots for | i could be infinite, if there are infinite numbers of
single-particle eigen-states. Comparing Eqs. (27) and (30), we have
a` |n1 , n2 , , n` , i =

n` + 1|n1 , n2 , , n` + 1, i.

The square-root is the important amplification factor for lasers to work.


6

(31)

Similarly, one can have the annihilation operator that maps a N -particle state to a
(N 1)-particle state,
a` |n1 , n2 , , n` , i |n1 , n2 , , n` 1, i.

(32)

This follows naturally from Eq. (31):


a` |n1 , n2 , , n` , i
X
=
|{n0 }iha` {n0 }|n1 , n2 , i

(33)

{n0 }

Xp

n0` + 1|{n0 }ihn01 , , n0` + 1, |n1 , , n` i

{n0 }

n` |n1 , n2 , , n` 1, i.

The ket |{n0 }i is an abbreviation of |n01 , n02 , i. We have inserted a completeness relation,
and used a Hermitian conjugate operation to get the first equation.
By applying the creation operator repeatedly to the vacuum state |0, 0, i (abbreviated
as |0i), one can reach any of the manybody state |n1 , n2 , i as follows,
n1 n2
1
1
|n1 , n2 , i =
a1
a2
|0i.
n1 ! n2 !

(34)

Fermions
The rules are simpler for fermions since the occupation number ni ( i) can only be 0 or 1.
If the state ` is empty (n` = 0), then
a` |n1 , n2 , , 0, i = (1) |n1 , n2 , , 1, i,
where =

P`1
i=1

(35)

ni . If the state ` is occupied (n` = 1), then


a` |n1 , n2 , , 1, i = 0.

(36)

One cannot add two fermions to the same state because of the exclusion principle.
On the other hand, for the annihilation operator, one has
a` |n1 , n2 , , 1, i = (1) |n1 , n2 , , 0, i,
a` |n1 , n2 , , 0, i = 0,
where =

P`1
i=1

ni .
7

(37)

Also, a manybody state can be reached by applying the creation operator to the vacuum
state (compare with Eq. (34)),
n1 n2
|0i.
|n1 , n2 , i = a1
a2

(38)

Of course, the occupation numbers ni ( i) can only be 0 or 1.


For example, the electrons in an electron gas are labelled by quantum numbers (k, s).
The ground state of the electrons is a Fermi sphere in momentum space. The state of a
filled Fermi sphere with radius kF is
Y

|F Si =

ak ak |0i.

(39)

k<kF

B. Commutation relations

Recall that
a |1 , 2 , , N } = |, 1 , 2 , , N }.
Therefore,
a a |1 , 2 , , N } = |, , 1 , 2 , , N },

(41)

a a |1 , 2 , , N } = |, , 1 , 2 , , N }.
The two states on the RHS are identical for bosons, but differ by a sign for fermions. Since
such a connection applies to any manybody state |1 , 2 , , N }, we can remove the ket
and simply write, for bosons,
a a = a a

or

[a , a ] = 0

(42)

for any single-particle states , . Hermitian conjugate of this equation gives


[a , a ] = 0

(43)

For fermions, one has


a a = a a

or

{a , a } = 0

(44)

for any single-particle states , . Hermitian conjugate of this equation gives


{a , a } = 0
8

(45)

Notice that fermion creation operators do not commute, but anti-commute with each other.
Also, if = , then one has

= 0,

(46)

This is again a reflection of the exclusion principle.


For combined operation of creation and annihilation operators, it is more convenient to
use the occupation-number representation (Cf. Eq. (41)). One starts from fermions : From
Eq. (35), we have,
a` |n1 , n2 , , 0, i(1) |n1 , n2 , , 1, i,
where =

P`1
i=1

ni Also, from Eq. (37),


a` |n1 , n2 , , 1, i = (1) |n1 , n2 , , 0, i.

Combined operation gives ( = `)


a a |n1 , n2 , , 0, i = |n1 , n2 , , 0, i,

(48)

a a |n1 , n2 , , 0, i = 0.
One can also consider the possibility of n` = 1 and write down another two equations. Both
cases lead to a a + a a = 1, the identity operator. Its not difficult to show that, if the
annihilation and creation operators act on different states, and , then a a + a a = 0.
Therefore, in general, for fermions,
{a , a } = .

(49)

Its left as an exercise to show that, for bosons, similar argument leads to
[a , a ] = .

(50)

Finally, for both bosons and fermions,


a a |n1 , n2 , , n , i = n |n1 , n2 , , n , i.
Therefore, n
a a is also known as the occupation number operator.

(51)

C. Change of basis

Under a unitary transformation, the one-particle state |i changes to |


i. They are
related by the following unitary transformation,
|
i =

|ih|
i.

(52)

On such a new basis,


a |
1 ,
2, ,
N } = |
,
1,
2, ,
N }
X
h|
i|,
1,
2, ,
N }
=

(53)

1 ,
2, ,
N }.
h|
ia |

Therefore, after striping off the ket state, we have


a =

X
h|
ia .

(54)

That is, the creation operator transforms like |i. Also,


a =

X
h
|ia .

(55)

It can be shown that, if the new set {|


i} is also an orthonormal set, then,
[a , a] = for bosons

(56)

{a , a} = for fermions
That is, the canonical commutation relations remain invariant under a unitary transformation.

III. COORDINATE AND MOMENTUM REPRESENTATIONS

The ket states in previous Section are representation free. We will project such ket states
to specific basis, such as coordinate basis, or momentum basis. Recall that the single-particle
wave function is
(r) = hr|i.
10

(57)

We can change the -basis to r-basis using the unitary transformation (Eq. (54)), then
(r) =

h|ria =

(r)a .

We have rewritten ar as (r). Such an operator that creates (or annihilates) a particle at
a particular point in space is called a field operator. The inverse transformation is
Z
Z

a = d rhr|i (r) = d3 r (r) (r).

(59)

For example, the quantum state |i of a particle in an empty box with volume V0 can be

labelled by momentum k, and hr|ki = eikr / V0 . Therefore,


1 X ikr
e
ak .
(r) =
V0 k

(60)

This is nothing but the Fourier expansion. Its inverse transformation is,
Z

(k) =
d3 rhr|ki (r)
(61)
Z
1
=
d3 reikr (r).
V0

Solid state theorists prefer to move the 1/ V0 factor in Eq. (61) to Eq. (60). That is,
1 X ikr
e
ak ,
V0 k
Z

(k) =
d3 reikr (r).
(r) =

(62)

We have assumed the system is inside a box with a finite volume, so the momentum k is
quantized. If V0 approaches infinity, then one can replace the summation with an integral,
Z
1 X
d3 k

.
(63)
V0 k
(2)3
Consider another example: the hydrogen atom. The quantum numbers for a spinless
electron are = (n, l, m), and the field operator is
(r) =

nlm (r)anlm .

(64)

nlm

According to Eq. (56), one has


{(r), (r0 )} = (r r0 ).
11

(65)

With the help of Eq. (59), the particle-number operator in the coordinate representation
becomes
N =

a a

Z
d3 rd3 r0

(66)
X
hr|ih|r0 i (r)(r0 )

d3 r (r)(r).

We have used a completeness relation to remove the -summation, and used the orthogonality relation hr|r0 i = (r r0 ). Naturally, one can define the particle-density operator
as
(r) = (r)(r).

(67)

This looks like the usual particle density in quantum mechanics. But beware that this is an
operator relation, not a numerical relation.

IV. SECOND QUANTIZATION


A. One-body operator

To deal with manybody systems, we need an operator A that acts on N -body states
(1)

|1 i|2 i |N i. For example, if Ai

is the operator for a particle at position ri , then A is

the operator for all of the particles. The latter operates in a much larger, N -particle Hilbert
space H(N ) = H(1) H(1) H(1) .
(1)

The operator Ai

can be expanded using an explicit basis,


(1)

Ai

|ih|A(1) (ri )|ih|

(68)

(1)

A |ih|,

(1)

where A h|A(1) |i.


Before defining the operator A, it is convenient to rewrite, following Feynmans notation
(the below is not the usual direct product ),
|1 , 2 , , N } |1 i |2 i |N i.
12

(69)

(1)

Operator A for a manybody system is connected with Ai


A=

N
X

in the following way,

(1)

Ai .

(70)

i=1

More precisely, it is

A|1 i |2 i |N i = A(1) |1 i |2 i |N i

+ |1 i A(1) |2 i |N i

(71)

+ |1 i| |2 i A(1) |N i .
(1)

Lets first assume Ai = |ih|, then (see Eq. (29))


A|1 i |2 i |N i
N
X
=
h|i i |1 i |i |N i
| {z }
i=1

= a
=

(72)

N
X

(1)i1 i |1 i (no |i i) |N i

i=1

a a |1 i

|2 i |N i.

Notice that the summation here should be interpreted like the series in Eq. (72). The
P
(1)
sign (1) is for bosons/fermions. In general, when A = A |ih|, each term in the
summation can be identified as a a using the procedure above. As a result,
X (1)
A a a .
A=

(73)

It looks as simple as the operator for a single particle (Eq. 68).

1. Example: Density operator

Firstly, we need to know the first quantized form of density. For a single particle, it is
(1) (r) = (r r).

(74)

Notice that r is an operator, while r is just an ordinary vector. One can easily verify that
h|(1) (r)|i = (r)(r). The Fourier transform of the density operator is
Z
(1)
(q) =
dveiqr (1) (r)
= eiqr .
13

(75)

We have used a simpler notation dv for d3 r. For many particles (but still in first quantized
form),
(r) =

N
X

(ri r).

(76)

i=1

Its Fourier transform is


(q) =

eiqri .

(77)

To get the second quantized form, we need (see Eq. (73)),


=

(1)

a a .

(78)

a. Coordinate representation

In the coordinate representation, the matrix element in Eq. (78) is


(1)

r0 r00 = hr0 |(1) (r)|r00 i

(79)

= (r0 r)(r0 r00 ).


Replace the -summation in Eq. (78) by a r-integral, and rewrite ar as (r), then (following
Eq. (78))
Z
(r) =

Z
dv

dv 00 (r0 r)(r0 r00 ) (r0 )(r00 )

(80)

= (r)(r)
This is the same as Eq. (67), and is similar to the usual expression (r) = (r)(r). The
difference is that (r) is a one-particle wave function, while (r) is a field operator.

b. Momentum representation

In momentum representation, the matrix element is


(1)

k0 k00 = hk0 |(1) (q)|k00 i


Z
Z
0
dv 00 hk0 |r0 ihr0 |eiqr |r00 ihr00 |k00 i.
=
dv

(81)

00
The matrix element hr0 |eiqr |r00 i = eiqr (r0 r00 ). Also, hr|ki = eikr / V0 , so we get
(1)

k0 k00 = (k00 k0 q).


14

(82)

Therefore,
X

(q) =

(1)

k0 k00 ak0 ak00

(83)

k0 k00

ak0 ak0 +q .

k0

2. Example: Hamiltonian for non-interacting particles

Consider the following Hamiltonian for a single particle,


H (1) =

p2
+ V (1) (r).
2m

(84)

For a non-interacting manybody system, it becomes

N 2
X
pi
(1)
H=
+ V (ri ) .
2m
i=1

(85)

Inter-particle interaction V (ri rj ) will be discussed in the next Section.

a. Energy-eigenstate representation

Recall Eq. (1),


H (1) = .
The Hamiltonian matrix is diagonalized in the eigenstate basis,
(1)

H = .

(87)

Therefore, the second quantized form is quite simple,


H=

(1)

H a a =

a a .

(88)

b. Coordinate representation

The Hamiltonian matrix in the coordinate basis is


hr|H (1) |r0 i =

~2 2
(r r0 ) + V (1) (r)(r r0 ).
2m

15

(89)

Therefore, the second quantized Hamiltonian is


Z
H =
dvdv 0 hr|H (1) |r0 i (r)(r0 )

Z
~2 2
(1)

=
dv (r)
+ V (r) (r).
2m

(90)

Recall that the first quantized version is written as Eq. (85).


The Heisenberg equation for field operator is
[, H] = i~
which leads to

,
t

~2 2

(1)

+ V (r) (r, t) = i~ .
2m
t

(91)

(92)

Even though this looks exactly the same as the Schrodinger for a single particle, it is actually
an equation for the field operator. It is as if we have promoted the single-particle wave
function to a field operator. That is why the present formulation is called the second
quantization.

c. Momentum representation

The Hamiltonian matrix in the momentum basis is (recall that hr|ki = eikr / V0 )
Z
~2 k 2
(1) 0
0
hk|H |k i =
hk|k i + dvhk|riV (1) (r)hr|k0 i
2m
Z
~2 k 2
1
0
=
kk0 +
dvV (1) (r)ei(kk )r .
2m
V0

(93)

The second integral is actually the Fourier transform of V (1) (r): V (1) (k k0 ). The second
quantized Hamiltonian is
H =

XX
hk|H (1) |k0 iak ak0
k

k0

X ~2 k 2
k

2m

ak ak +

(94)

1 X X (1)
V (k k0 )ak ak0 .
V0 k k0

The potential term can also be written as (see Eq. (83))


1 X X (1)
1 X (1)
V (q)ak akq =
V (q)(q).
V0 k q
V0 q

(95)

Some useful operators in first and second quantized form are summarized in Table 1.

16

TABLE I One-body operator


1st quantization
P

particle density (r)

ri
i (

2nd quantization
(r)(r)

r)

iq
ri
ie

(q)
current density j(r)

1
2m

j(q)

1
2m

pi (ri
i [

iV

(1) (
ri

V (1) (q)

V (q)
magnetic moment density m(r)

r) + (ri r)
pi ]

P iqri
ie
i
+ eiqri p
i p

one-body potential V (r)

ri
i i (

dv (r)V (1) (r)(r).


k

V (1) (q)ak+q ak

(r)(r) [ is a spinor]

r)

iq
ri
i i e

m(q)

~ P (2k + q) a ak+q
k
k

2m

iq
ri
ie

~ (r)(r) (r) (r)


2mi

r)

ak ak+q

k k+q

B. Two-body operator

Two-body operators act on two particles at a time. A typical example is the interaction
potential V (ri rj ). In general, the operator for the whole system can be written as
A=

1 X (2)
A .
2 i6=j ij

(96)

One term in the summation,


(2)

Aij =

(2)

|i|0 ih|h0 |Aij |i| 0 ih|h 0 |

(97)

0 0

(2)

A0 0 |i|0 ih|h 0 |,

0 0

in which and label quantum states of particle i; 0 and 0 label quantum states of
particle j. Notice that we have switched the subscripts , 0 of A(2) . Some textbooks prefer
not to switch them.

17

(2)

Lets first assume Aij = |i|0 ih|h 0 |, then (for i < j)


(2)

Aij |1 i |N i
X
=
h|i ih 0 |j i|1 i |i |0 i |N i
i<j

= a a0

(98)

X
(1)i1 (1)j2 i 0 j
i6=j

|1 i (no |i i) (no |j i) |N i
= a a0 a 0 a |1 i |2 i |N i.
That is, |i|0 ih|h 0 | can be identified with a a0 a 0 a . The same is true if i > j. In
(2)

general, when Aij is a superposition of many terms, one has


A=

1 X (2)
A0 0 a a0 a 0 a .
2 0 0

(99)

1. Example: Inter-particle interaction

a. Coordinate representation

The matrix elements of V (2) (ri rj ) is


(2)

Vrr0 r00 r000 = hrr0 |V (2) (ri rj )|r000 r00 i

(100)

= V (2) (r r0 )(r r000 )(r0 r00 ).


Replace the summations in Eq. (100) by integrals over coordinates and write ar as (r), we
will get
1
V =
2

Z
dvdv 0 (r) (r0 )V (2) (r r0 )(r0 )(r).

(101)

The quartic operator can be written in terms of the density operator,


(r) (r0 )(r0 )(r)

(102)

= (r)(r) (r0 )(r0 ) (r)(r)(r r0 )


= (r)(r0 ) (r)(r r0 ).
The second term contributes to a self-interaction energy in Eq. (101). Therefore, the substraction helps removing the self-interaction in the first term.
18

b. Momentum representation

Expanding the field operator using plane waves,


1 X
(r) =
ak eikr ,
V0 k

(103)

then the inter-particle interaction becomes


X
1
ak4 ak3 ak2 ak1
2
2V0 k ,k ,k ,k
1 2 3 4
Z
0

dvdv 0 V (2) (r r0 )ei(k1 k4 )r ei(k2 k3 )r .

V =

(104)

Write the interaction potential as


V (2) (r r0 ) =

1 X (2)
0
V (q)eiq(rr ) ,
V0 q

(105)

then the space integral gives V0 k4 ,k1 +q k3 ,k2 q . Finally,


V =

1 X (2)
V (q)ak1 +q ak2 q ak2 ak1 .
2V0 k ,k ,q
1

(106)

One can visualize the item in the summation as follows: two incoming particles with momenta k1 and k2 are scattered by the potential and become outgoing particles with momenta
k1 + q and k2 q. The total momentum is conserved (elastic scattering), but there is a
transfer of momentum q from particle 2 to particle 1.

V. GENERAL DISCUSSION
A. Spin degree of freedom

From now on, we will focus only on systems of electrons (or fermions). Non-interacting
(or weakly interacting) electrons are referred to as electron gas. When interaction plays an
important role, we will call them as electron liquid. For spinful electrons in an empty box,
the Hamiltonian in coordinate representation is

XZ
~2 2

(1)
H =
dv s (r)
+ Vs (r) s (r)
2m
s
Z
1X
(2)
+
dvdv 0 s (r)s0 (r0 )Vss0 (r r0 )s0 (r0 )s (r),
2 s,s0
19

(107)

(1)

in which we have allowed for spin-dependent external potential Vs (r) and electron inter(2)

action Vss0 (r r0 ). The Hamiltonian in momentum representation is


H =

X ~2 k 2

aks aks +

1 X (1)
V (q)ak+q,s aks
V0 kqs s

2m
ks
1 X
(2)
V 0 (q)ak+q,s ak0 q,s0 ak0 s0 aks .
+
2V0 ks,k0 s0 ,q ss

(108)

The subscripts (k, s) are good quantum numbers for (non-interacting) electrons in an empty
box, which has free-particle energy 0k = ~2 k 2 /2m. For electrons in a lattice without spinorbit coupling, the proper label for quantum states is (n, k, s), where n is the band index,
and k is the quasi-momentum for Bloch electron. Also, the free-particle energy 0k has to be
replaced by Bloch energy nk .

B. Tight-binding model

In the tight-binding model, the electrons hop from one atom to another. The operator
anks for a Bloch state has to be replaced by al , where l contains information such as band
index n, lattice site R, and spin s. The Wannier function (sometimes atomic orbitals are
used instead) is
hr|li = hr|n, R, si = wn (r R)s ,
where s is a spinor,


1
+ = ,
0


0
= .
1

(109)

(110)

They have the following orthogonality and completeness relations:


X

s s0 = ss0 ,

s s = 122 .

(111)

The Hamiltonian for non-interacting electrons is


H0 =

(1)

Hlm als ams


,

(112)

lms

(1)
p2
+ V (1) (r)|mi,
assuming V (1) is a spin-independent
where l = (n, R), and Hlm = hl| 2m

potential. We have contracted the spin degree of freedom in the matrix elements.

20

If the interaction is also spin-independent, then


Vee =

1 X (2)
V
a a am0 al0 ,
2 lmm0 l0 lm m 0 l0 l m

(113)

where
(2)

Vlm m 0 l0
Z
=
dv1 dv2 wn l (r1 Rl )wn m (r2 Rm )V (2) (r1 r2 )

(114)

wnm0 (r2 Rm0 )wnl0 (r1 Rl0 ).


In reality, one may keep only the nearest-neighbor and the next-nearest-neighbor couplings. Furthermore, one might be able to use the one-band approximation, assuming
that the electrons only reside on one band n, then (n is omitted for simplicity), then
H = H0 + Vee
X (1)
=
HRR0 aRs aR0 s
RR0 s

1
2

(115)

(2)

VR1 R2 R0 R0 aR1 s aR2 s0 aR20 s0 aR10 s ,


2

R1 R2 R20 R10
ss0

in which we have assumed that the one-body and two-body potentials are spin-independent.
The simplest possible model (with interaction) is to keep only the nearest-neighbor hopping in the first term, and only the on-site energy in the second term. That is,
H = t

aRs aR0 s + U

<RR0 >s

= t

aR aR aR aR

(116)

aRs aR0 s + U

<RR0 >s

nR nR ,

where nRs is the occupation-number operator. Notice that there are only two parameters:
the hopping amplitude t and the on-site energy U . This is the famous Hubbard model
proposed by Hubbard more than 50 years ago to study narrow-band materials, such as
transition metal oxides. It is also considered as the underlying model for high temperature
superconductors. Despite enormous effort from numerous researchers, the phase diagrams
for the Hubbard model in dimension two and higher remain inconclusive. (see the article by
J. Quintanilla and C. Hooley in Phys. World, June, 2009)

21

C. What do we calculate

Most of the time we are interested in obtaining some of the following properties:
. Ground state energy
. Phase diagram (comparing lowest energies of different phases)
. Other ground state properties, such as charge order or spin order (correlation function),
density of states ... etc.
. Low-lying excitations (quasi-particles, collective excitations)
. Linear response function (conductivity, susceptibility ... etc)
. Other quantities of experimental interest.
Of course, whether the result is satisfactory or not depends on whether, at the first place,
the simplified model captures the essential ingredients of the phenomena we intend to study.

D. How do we calculate

The Hilbert space H(N ) of a manybody system is mind bogglingly big. In classical
mechanics, if the solution space of a particle has dimension d, then the solution space of the
whole system is N d. However, in quantum mechanics, the latter is dN . This causes major
problem for analytical and numerical calculations, and many different methods have been
proposed:
. Mean field approximation (MFA)
. Equation of motion (EOM) method
. Perturbation expansion using Greens function (diagram expansion)
. Variational method
. Density functional theory (DFT)
. Quantum Monte Carlo (QMC) method
. Density matrix renormalization group (DMRG)

22

. Tensor network renormalization group


.
The first 3 are perturbative calculations in essence. The last 4 are numerical methods.
Variation method first requires guess work, and then numerical computation.

Homework:
1. Under an unitary transformation, an ortho-normalized set {|i} is transformed to another ortho-normalized set {|
i}. Show that the anti-commutation relation (for fermions)
is invariant under the unitary transformation.
{a , a } = {a , a} = .
2. Show that, if

~2 2
(1)
+ V (r) (r),
dv (r)
2m

H=

then the field operator satisfies the following equation of motion,

~2 2

(1)

+ V (r) (r, t) = i~ .
2m
t
3. Start from
j(1) (r) =

1
[
p(r r) + (r r)
p] ,
2m

first find out its Fourier transform j(1) (q), then show that (see Table 1)
j(q) =

~ X
(2k + q) ak ak+q .
2m k

References
[1] Chap 6 of R.P. Feynman, Statistical Mechanics, a set of lectures, Addison-Wesley Publishing
Company, 1990.
[2] Chap 1 of H. Bruss and K. Flensberg, Many-body quantum theory in condensed matter physics,
Oxford University Press, 2004.

23

Вам также может понравиться