Second Quantization
Second Quantization
quantization)
(Advanced quantum mechanics)
October 1, 2011
References:
• F. Schwabl, Advanced Quantum Mechanics.
1
Such a reordering of elements of a given finite set is called a permutation.
In order to move ahead with the quantum theory we first need to study some
basic properties of these operations2 .
which means that the element 4 takes the place of the element 1, the element
5 takes the place of the element 2 etc. A special case of a permutation is a
cycle, in which the elements are substituted “in a chain”. An example is P134
(also written as (134)), which denotes a permutation in which 1 is replaced by
3, 3 by 4 and 4 by 1,
1 2 3 4 5 ...
P= .
3 2 4 1 5 ...
Here the right hand side denotes a sequence of cycle permutations that are
performed one after another from right to left (although for disjoint cycles the
order does not matter).
A very special case of a permutation, which is important for our lecture, is
a two-element cycle, that is, a permutation that just exchanges two elements,
say i and j. Such a permutation is called a transposition and will be denoted
by Pij or (ij). Note that Pij2 = E, where E is the identity (a “do nothing”
permutation). An important fact is that any permutation can be represented
as a sequence of transpositions (not necessarily disjoint). This can be done in
many different ways. However, it can be shown that in any such decomposition
of a given permutation the number of transpositions will be either always
odd or always even. This allows us to assign a parity to every permutation:
a permutation is said to be even (or to have a parity of +1) if it can be
decomposed into an even number of transpositions and to be odd (parity −1)
otherwise. The parity of a permutation P is denoted as (−1)P .
It is obvious that a composition (or product) of two permutations (that is,
performing two permutations P1 and P2 one after the other) is equivalent to
2
A complete discussion and proofs of the facts quoted here can be found, e.g., in
F.W. Byron, R.F. Fuller, Mathematics of Classical and Quantum Physics or, in Polish, in
R. Gonczarek, Teoria grup w fizyce, http://www.dbc.wroc.pl/Content/992/gonczarek.pdf.
2
performing some other single permutation P3 , which is written as P3 = P2 P1 .
Moreover, combining the permutations is associative, that is, performing the
product of P1 and P2 , and then P3 is the same as performing P1 , and then the
product of P2 and P3 ,
P3 (P2 P1 ) = (P3 P2 )P1 .
In addition, the set of all permutations contains a neutral element (identity) E
and for each permutation P one can find its inverse (that is, one that restores
the initial order of elements). The inverse to P is denoted by P −1 and is for-
mally defined by P −1 P = PP −1 = E. Taken together, these properties mean
that the set of permutations has the formal structure of a group. The group
of permutations (called also “symmetric group”) of N elements is denoted by
SN and the number of its elements (the order of the group) is N !.
An important theorem states that if {P1 , P2 , . . . , PN ! } are all the elements
of the group SN and each appears only once in this sequence than for any
permutation P ∈ SN , the sequence {PP1 , PP2 , . . . , PPN ! } also contains all
the elements of the group and each of them appears only once. The same is
true for the sequence {P1 P, P2 P, . . . , PN ! P}.
3
Some more mathematics is needed to make this statement absolutely strict. See the
texts on group theory recommended above.
3
From the theorem quoted at the end of Sec. 1.2 it easily follows that
P S+ = S+ P = S+ and P S− = S− P = (−1)P S− . (1)
In the proof of the second fact, we note that P Pi has the same parity as Pi if
P is even and opposite parity if P is odd. An obvious special case of this is
Pij S− = −S− .
We will now discuss a few properties of the permutation operator acting
on the quantum states of the N -particle system in the way defined above.
The permutation operator conserves the scalar product, that is, for any
states |Ψi, |Φi,
hΨ|Φi = hP Ψ|P Φi.
In view of the linearity of the permutation operator and of the scalar prod-
uct it is sufficient to verify this fact for two product configurations |Ψi =
|i1 . . . iα . . . iβ . . . iN i and |Φi = |i01 . . . i0α . . . i0β . . . i0N i. Moreover, as every per-
mutation can be composed of transpositions, one only has to show this equality
for transpositions. One has
Pαβ |Ψi = |i1 . . . iβ . . . iα . . . iN i, Pαβ |Φi = |i01 . . . i0β . . . i0α . . . i0N i,
hence
hP Ψ|P Φi = δi1 i01 . . . δiβ i0β . . . δiα i0α . . . δiN i0N = hΨ|Φi.
As a consequence, P is a unitary operator. Indeed, using the previous fact,
we have
hΦ|P Ψi = hP −1 Φ|P −1 P Ψi = hP −1 Φ|Ψi.
This means that P −1 = P † , that is, P is unitary.
If the particles are identical (indistinguishable) then, classically, the system
with particles i and j having the positions ri , rj and the momenta pi , pj must
have the same properties as that with the particle i having the position and
momentum rj , pj and the particle j having the position and momentum ri , pi .
Therefore, the physical quantities cannot change if we exchange two particles,
say, i and j. The same holds true for any other permutation of particles. We
expect that also in quantum mechanics any observable describing a measurable
property of a system of many identical particles must be invariant under any
permutation of particles. This can be stated formally by noting that if |Ψi
is an arbitrary state of the system of identical particles and |Φi = P |Ψi is
the state with particles reordered by a permutation P than the average value
of any observable A must be the same in the two states (because permuting
identical particles cannot have a physical effect). That is,
hΨ|A|Ψi = hΦ|A|Φi = hΨ|P † AP |Ψi.
As this must be true for any state |Ψi we can conclude that
A = P † AP for any P.
This can also be written in the form [P, A] = 0, which is a formal statement
of the fact that A must be symmetric with respect to any permutation.
4
As a consequence of this fact, if |Ψi is an eigenstate of an observable A
then P |Ψi is an eigenstate as well, belonging to the same eigenvalue. Indeed, if
A|Ψi = a|Ψi then A(P |Ψi) = P A|Ψi = P a|Ψi = a(P |Ψi). If one allowed an
arbitrary vector |Ψi to be a valid representation of a physical state then P |Ψi
would, in general, be a different vector for each P ∈ SN and every eigenstate
of the observable A would be N ! times degenerate. This enormous degeneracy
(think of 1024 electrons in a piece of a metal) would be completely undetectable
because it affects every physical observable. It turns out, however, that this is
not the case.
5
i1 , i2 , . . . and insist that the state labels i1 , i2,... appear in the configuration used
for antisymmetrization in the prescribed order. Once this sign ambiguity is
solved, a state is fully characterized by the number of particles in each single-
particle state i, denoted by ni . Therefore, the states of N fermions which are
obtained by antisymmerization of product configurtions can be denoted as
X
|n1 n2 . . . ni . . .i = S− |i1 . . . iα . . . iN i, ni = 0, 1, ni = N, (2)
i
where ni is the number of times i appears among the labels iα on the right-hand
side. This representation of a many-particle state is called the occupation
number representation or the second quantization representation.
For bosons, symmetrization yields ni ! copies of each different configuration
if the original configuration contained the state i ni times. This produces a
factor of ni ! in the norm. Therefore, in order to get a normalized completely
symmetric state one has to cancel these factors and the symmetrized state of
bosons in the occupation number representation is
1
|n1 n2 . . . ni . . .i = √ S+ |i1 . . . iα . . . iN i,
n1 !n2 ! . . . ni ! . . .
X
ni = 0, 1, . . . ni = N.
i
hn01 n02 . . . n0i . . . |n1 n2 . . . ni . . .i = δn1 n01 δn2 n02 . . . δni n0i . . . , (3)
6
Hence,
1 1 X
|Ψi = √ S± |Ψi = √ ci1 ...iN S± |i1 i . . . |iN i.
N! N ! i1 ...iN
This shows that, indeed, any completely (anti)symmetric state is a linear com-
bination of (anti)symmetrized product configurations, that is, these states form
a complete set in HN . In the occupation number representation, this com-
pleteness relation reads
X
|n1 n2 . . . ni . . .ihn1 n2 . . . ni . . . | = IN , (4)
n1 +n2 +...+ni +...=N
Formally, the Fock space is a direct sum of the N -particle spaces for all N ,
F = H0 ⊕ H1 ⊕ . . . ⊕ HN ⊕ . . . .
We postulate that the states with different total number of particles are or-
thogonal. This means that Eq. (3) can be rewritten for the Fock space
hn01 n02 . . . n0i . . . |n1 n2 . . . ni . . .i = δn1 n01 δn2 n02 . . . δni n0i . . . (orthogonality).
(5)
7
By construction, the states |n1 n2 . . . ni . . .i span the Fock space, therefore the
relation generalizing Eq. (4) is
X
|n1 n2 . . . ni . . .ihn1 n2 . . . ni . . . | = I (completeness), (6)
n1 n2 ...ni ,...
This means that any state can be constructed by acting on the vacuum state
with appropriate creation operators,
1 n1 n2
† †
ni
|n1 n2 . . . ni . . .i = √ a1 a2 . . . a†i . . . |0i.
n1 !n2 ! . . .
The action of its adjoint (Hermitian conjugate), ai , is found by conjugating
the definition given in Eq. (7),
p
hn01 n02 . . . n0i . . . |ai = n0i + 1hn01 n02 . . . n0i + 1 . . . |,
and multiplying both sides from the right by |n1 n2 . . . ni . . .i,
p Y
hn01 n02 . . . n0i . . . |ai |n1 n2 . . . ni . . .i = n0i + 1δn0i +1,ni δn0j ,nj , (8)
j6=i
where we used the orthogonality relation (5). Next, one uses the completeness
relation, Eq. (6), to write
X
ai |n1 . . . ni . . .i = |n01 . . . n0i . . .ihn01 . . . n0i . . . |ai |n1 . . . ni . . .i
n01 ...n0i ...
X p Y
= |n01 . . . n0i . . .i n0i + 1δn0i +1,ni δn0j ,nj ,
n01 ...n0i ... j6=i
8
where we used Eq. (8). Hence, one finds for the action of the annihilation
operator,
√
ni |n1 . . . ni − 1 . . .i, ni > 0,
ai |n1 . . . ni . . .i = (9)
0, ni = 0.
aj ai |n1 , . . . , ni , . . .i = ai aj |n1 . . . ni . . .i
and
a†j a†i |n1 . . . ni . . .i = a†i a†j |n1 . . . ni . . .i
(note that the cases i = j and i 6= j must be considered separately). For a
subsequent action of a creation and annihilation operator one finds
and
√
ai a†i |n1 . . . ni . . .i = ni + 1ai |n1 , . . . , ni + 1, . . .i = (ni + 1)|n1 . . . ni . . .i
Moreover, Eq. (10) shows that the states |n1 . . . ni . . .i are eigenstates of the
operators a†i ai and the corresponding eigenvalue is the number of particles in
the state i. Therefore, n̂i = a†i ai is called the number operator for the state i.
The total particle number operator is therefore,
X X
N̂ = n̂i = a†i ai . (11)
i i
The states |n1 . . . ni . . .i are therefore particle number eigenstates, that is,
states with a well-defined number of particles in each state. They are com-
monly called number states or Fock states. Keep in mind that these states
form an orthonormal basis in the Fock space.
9
2.2 Creation and annihilation operators for fermions
We would like now to define the creation and annihilation operators for fermions
in the same way as we did for bosons. In analogy with bosons, we want to
define creation operators in such a way that
As already pointed out in Sec. 1.4, the basis states for the Hilbert space
HN of an N -fermion system must be
X
|n1 n2 . . . ni . . .i = S− |i1 i2 . . . iN i = a†i1 a†i2 . . . a†iN |0i, ni = N,
i
where the creation operators for different states are arranged according to the
fixed ordering of the states. Next, we combine the states for different N to
form the Fock space, as previously. Now, however, upon acting with a creation
operator a†i on an N -particle state we not necessarily get a correct (N + 1)-
particle basis state because the operators may not be ordered properly. Let
in < i < in+1 for a certain n. Then
where the phase factor originates from flipping the creation operators until
a†i gets onto its place. If the occupation number representation is introduced
according to Eq. (2) then the above identity can be written as
10
earlier) from this definition. Note that, consistently with this definition, the
Fock state can be expressed in the form
n1 n2 ni
† †
|n1 n2 . . . ni , . . .i = a1 a2 . . . a†i . . . |0i, ni = 0, 1.
Next, using this fact and the completeness relation (6), we find
X
ai |n1 . . . ni . . .i = |n01 . . . n0i . . .ihn01 . . . n0i . . . |ai |n1 . . . ni . . .i
n01 ...n0i ,...
P
= ni (2 − ni )(−1) j<i nj |n1 . . . ni − 1 . . .i
P
= ni (−1) j<i nj |n1 . . . ni − 1 . . .i. (13)
aj ai |n1 , . . . , nj , . . . , ni , . . .i
P
= aj ni (−1) i0 <i ni0 |n1 . . . nj . . . ni − 1 . . .i
P P
= ni nj (−1) i0 <i ni0 (−1) j0 <j nj0 |n1 . . . nj − 1 . . . ni − 1 . . .i
but
ai aj |n1 . . . nj . . . ni . . .i
P
= ai nj (−1) j0 <j nj0 |n1 . . . nj − 1 . . . ni . . .i
P P
= ni nj (−1) i0 <i ni0 −1 (−1) j0 <j nj0 |n1 . . . nj − 1 . . . ni − 1 . . .i,
where the difference in the second phase factor results from the fact that the
particle from the state j, preceding the state i, has already been removed by
the first operator. It is clear that the two results have an opposite sign. Hence,
we conclude that ai aj = −aj ai . Note that this is also true for i = j as ai ai = 0
By conjugation, we get a similar relation for the creation operators. It can be
checked in the same way that ai a†j = −a†j ai for i 6= j. Finally, we have
P P
= (1 + ni )(1 − ni )(−1) i0 <i ni0 (−1) i0 <i ni0 |n1 . . . ni . . .i
= (1 − n2i )|n1 . . . ni . . .i
and
P P
= n2i (−1) i0 <i ni0 (−1) i0 <i ni0 |n1 . . . ni . . .i. (14)
11
This means that ai a†i + a†i ai = 1. Altogether, we find that the fermionic
operators satisfy the anticommutation relations
n o n o
{ai , aj } = a†i , a†j = 0, ai , a†j = δij , (fermions)
As ni are the eigenvalues of the number operator and the number states form
a basis in the Fock space, the energy operator (Hamiltonian) for a system of
non-interacting bosons must have the form
X X
H= i n̂i = i a†i ai . (non-interacting particles)
i i
4
This specific choice is not important for the general formalism. However, choosing the
basis states to be the eigensattes of a single-particle Hamiltonian is very common in practical
applications of the formalism.
12
In the matrix element the index α has been suppressed because all the particles
are identical, hence all the operators Gα are identical. The single-particle
operator G can now be written in the form
X X
G= hi|G|ji |iiα hj|α .
ij α
P
Let us focus on the operator α |iiα hj|α . It is convenient to treat the fermion
and boson cases separately.
For bosons, one has
X 1 X
|iiα hj|α | . . . ni . . . nj . . .i = √ |iiα hj|α S+ |i1 . . . iN i
α
n 1 !n 2 ! . . . α
1 X
= √ S+ |iiα hj|α |i1 . . . iN i,
n1 !n2 ! . . . α
P
where we used the fact thatPS+ commutes with the symmetric operator α |iiα hj|α .
Now, consider the object α |iiα hj|α |i1 . . . iN i. Each time iα = j for a certain
α, it is replaced by i. After symmetrization, each such configuration yields the
vector q
. . . (ni + 1)! . . . (nj − 1)! . . . | . . . ni + 1 . . . nj − 1 . . .i.
The same result will appear for each occurrence of j, that is, nj times. Hence,
X
S+ |iiα hj|α |i1 . . . iN i
α
q
= nj . . . (ni + 1)! . . . (nj − 1)! . . . | . . . ni + 1 . . . , nj − 1 . . .i
and
X √ √
|iiα hj|α | . . . ni . . . nj . . .i = nj ni + 1| . . . ni + 1 . . . nj − 1 . . .i
α
= a†i aj | . . . ni . . . nj . . .i,
where the final equality follows from Eqs. (7) and (9).
For fermions, the derivation is quite similar (there are no additional nor-
malization factors but one has to be careful with the sign). We have
X X X
|iiα hj|α | . . . ni . . . nj . . .i = |iiα hj|α S− |i1 . . . iN i = S− |iiα hj|α |i1 . . . iN i.
α α α
As we deal with fermions, the state j can appear at most once among iα . If
it does appear it will be changed into i. A non-zero result will be obtained
from antisimmetrization only if i was not present initially. Thus, if ni = 0 and
nj = 1 we obtain the initial configuration with j replaced by i, which we write
in the form,
X
S− |iiα hj|α |i1 . . . iα . . . iN i = nj (1 − ni )S− |i1 . . . iN ij→i .
α
13
Note, however, that by changing j into i we get the state labels in the wrong
order. The correct order is restored by moving the label i to the beginning
of the list
Pand thanPback to the place, where it shouldP be. If i < j then this
requires k<j nk + k<i nk flips. P If i > j than we need k<j nk transpositions
to move to the beginning and k<i nk −1 to move to the end, where “-1” stands
for the label j which is counted in the sum but shouldn’t be as it has just been
moved from its place. As a result, one finds
X
|iiα hj|α | . . . ni . . . nj . . .i
α
P P
nj (1 − ni )(−1)Pk<j nk +Pk<i nk | . . . ni + 1 . . . nj − 1 . . .i,
i < j,
= n + n −1
nj (1 − ni )(−1) k<j k k<i k
| . . . nj − 1 . . . ni + 1 . . .i, i > j.
Using the definitions of a†i and aj it is easy to verify that exactly the same
result is produced by a†i aj | . . . ni . . . nj . . .i. It can be checked that the two
results are also identical for i = j. Thus, we get
X
|iiα hj|α | . . . ni . . . nj . . .i = a†i aj | . . . ni . . . nj . . .i,
α
where fαβ is an operator acting on the states of the particles α and β (for
instance, a typical interaction energy depends only on the coordinates of the
two particles). Again, the form of this term cannot depend on which pair
of particles we choose. The condition that α 6= β corresponds to the fact
that a particle does not interact with itself. The factor 1/2 assures that each
pair is taken into account only once. In order to find the second quantization
representation of such an operator we note that
X X X
|iiα |jiβ hk|α hl|β = |iiα hk|α |jiβ hl|β − |iiα hk|α |jiα hl|α
α6=β α,β α
X X X
= |iiα hk|α |jiβ hl|β − |iiα hl|α δkj .
α β α
14
In the first term, we can use the representation in terms of creation and anni-
hilation operators found above,
X X
|iiα hk|α = a†i ak , |jiα hl|α = a†j al .
α α
and use the (anti)commutation relations δkj = ak a†j ± a†j ak , where the upper
sign is for fermions and the lower one is for bosons. As a result we get
X
|iiα |jiβhk|α hl|β = a†i ak a†j al − a†i ak a†j al ± a†i a†j ak al = a†i a†j al ak ,
α6=β
15
Their physical interpretation is that they create or annihilate a particle at a
point r.
Since Ψ(r) is a linear combination of annihilation operators which commute
or anticommute (for bosons and fermions, respectively) they obviously also
commute or anticommute. For two operators of different kind, one finds
X h i X
† 0 ∗ 0 †
ψi (r)ψj∗ (r 0 )δij
Ψ(r), Ψ (r ) ± = ψi (r)ψj (r ) ai , aj =
±
ij ij
X
= ψi (r)ψi∗ (r 0 ),
i
~2 ~2
Z Z
†
T =− 3
d rΨ (r)∆Ψ(r) = d3 r∇Ψ† (r)∇Ψ(r). (17)
2m 2m
and the single-particle potential energy U has the form
Z
U = d3 rΨ† (r)U (r)Ψ(r) (18)
16
An important example is the two-particle interaction energy
Z
1
V = d3 rd3 r 0 Ψ† (r)Ψ† (r 0 )V (r, r 0 )Ψ(r 0 )Ψ(r). (19)
2
[A(t), B(t)]± = eiHt/~ Ae−iHt/~ eiHt/~ Be−iHt/~ ± eiHt/~ Be−iHt/~ eiHt/~ Ae−iHt/~
= eiHt/~ ABe−iHt/~ ± eiHt/~ BAe−iHt/~ = eiHt/~ [A, B]± e−iHt/~
= eiHt/~ Ce−iHt/~ = C(t).
Thus, the evolution of operators in the Heisenberg picture conserves the com-
mutation and anticommutation relations. We will also make use of the relation
17
Let us find the evolution equation for the fermionic field operator Ψ(r)
(the calculation for bosonic operators is very similar). We start the calculation
of the commutator appearing in the Eq. (21) with the single-particle energy
contribution. We will not write the time dependence explicitly. One finds
Z
3 0 † 0 0 0
−[U, Ψ(r)] = − d r Ψ (r )U (r )Ψ(r ), Ψ(r)
Z
= − d3 r 0 U (r 0 ) Ψ† (r 0 )Ψ(r 0 )Ψ(r) − Ψ(r)Ψ† (r 0 )Ψ(r 0 )
Z
= − d3 r 0 U (r 0 ) −Ψ† (r 0 )Ψ(r)Ψ(r 0 )
= U (r)Ψ(r).
~2
Z
d3 r 0 Ψ† (r 0 )∆0 Ψ(r 0 )Ψ(r) − Ψ(r)Ψ† (r 0 )∆0 Ψ(r 0 )
−[T, Ψ(r)] =
2m
~2
Z
d3 r 0 −Ψ† (r 0 )Ψ(r)∆0 Ψ(r 0 )
=
2m
− δ(r − r 0 ) − Ψ† (r 0 )Ψ(r) ∆0 Ψ(r 0 )
~2
= − ∆Ψ(r),
2m
where ∆0 denotes the Laplasian with respect to the coordinate r 0 and we used
the fact that {∆Ψ(r 0 ), Ψ(r)} = 0 since Ψ(r 0 ) anticommutes with Ψ(r) for any
r0 .
Finally, for the interaction term, we have
Z Z
1 3 0
d3 r 00 V (r 0 , r 00 ) Ψ† (r 0 )Ψ† (r 00 )Ψ(r)
−[V, Ψ(r)] = − dr
2
−Ψ(r)Ψ† (r 0 )Ψ† (r 00 ) Ψ(r 00 )Ψ(r 0 )
Z Z
1 3 0
d3 r 00 V (r 0 , r 00 ) Ψ† (r 0 ) δ(r − r 00 ) − Ψ(r)Ψ† (r 00 )
= − dr
2
− δ(r − r 0 ) − Ψ† (r 0 )Ψ(r) Ψ† (r 00 ) Ψ(r 00 )Ψ(r 0 )
Z
1
= − d3 r 0 Ψ† (r 0 )V (r 0 , r)Ψ(r)Ψ(r 0 )
2
Z
1
− d3 r 00 Ψ† (r 00 )V (r, r 00 )Ψ(r 00 )Ψ(r)
2
Z
= d3 r 0 Ψ† (r 0 )V (r 0 , r)Ψ(r 0 )Ψ(r),
where we use the fact that V (r, r 0 ) = V (r 0 , r) and changed the order of anni-
hilation operators in the last step.
18
Hence, the field equation, that is, the evolution equation for the operator
Ψ(r) is
~2
i~Ψ̇(r, t) = − ∆Ψ(r, t) + U (r)Ψ(r, t)
2m
Z
+ d3 r 0 Ψ† (r 0 , t)V (r 0 , r)Ψ(r, t)Ψ(r 0 , t).
As the density should satisfy the continuity equation ρ̇(r, t) = −∇ · j(r, t),
where j is the current density, we identify the operator
~ †
Ψ (r, t)∇Ψ(r, t) − ∇Ψ† (r, t) Ψ(r, t)
j(r, t) =
2im
as the current operator. In fact, it can be shown directly that the current
operator in the representation of field operators has exactly this form.
4 Momentum representation
The momentum representation is just the second quantization representation
in terms of the creation and annihilation operators, where the single particle
basis is chosen as the basis of free-particle states, labeled by the wave number
vector k and represented in the position representation by the plane waves,
1
ψk (r) = √ eik·r ,
V
where we assumed that the system volume is V and that the periodic boundary
conditions are imposed. According to Eq. (15), the field operators may be
represented as
1 X −ik·r † 1 X ik·r
Ψ† (r) = √ e ak , Ψ(r) = √ e ak . (22)
V k V k
19
performing the derivation and using the orthonormality condition
Z
1 0
d3 rei(k−k )·r = δk,k0
V
one finds easily for the kinetic energy
X ~2 k 2 †
T = ak ak .
k
2m
where Z
Uq = d3 re−iq·r U (r).
For the two-particle interaction which depends only on the relative position of
the two particles, r − r 0 (as is usually the case), we have
Z Z
1 X 0 0 0 0
V = 2
d 3
r d3 0
r e−iq·r e−iq ·r V (r − r 0 )eik ·r eik·r a†q a†q0 ak0 ak
2V kk0 qq 0
Z Z
1 X
i(k+k0 −q−q 0 )·r 0 0 00
= 2
dr3
e d3 r00 V (r 00 )e−i(k −q )·r a†q a†q0 ak0 ak
2V kk0 qq 0
1 X 1 X
= δk−q,q0 −k Vk0 −q0 a†q a†q0 ak0 ak = Vq−k a†q a†k+k0 −q ak0 ak
2V kk0 qq0 2V kk0 q
1 X
= Vp a†k+p a†k0 −p ak0 ak .
2V pkk0
Another useful relation is the expression for the Fourier transform of the
particle density,
Z X
ρ(q) = d3 rρ(r)e−iq·r = a†p ap+q ,
p
which is easy to get by substituting the field operators from Eq. (22) into
Eq. (20).
20
5 Spin
In the formalism of creation and annihilation operators, spin is just another
quantum number that has to be included to label the state. For instance,
in the momentum representation, the creation and annihilation operators are
akσ , a†kσ , with the commutation relations
h i
[akσ , ak σ ]± = akσ , ak0 σ0 = 0, [akσ , a†k0 σ0 ]± = δkk0 δσσ0 .
0 0
† †
±
21
From the introductory physics course or from the solid state course one
(hopefully) remembers that in the ground state, noninteracting fermions fill
the states up to the value of momentum
~kF = ~(3π 2 n)1/3 , (25)
called the Fermi momentum, where n is the particle density. This value of
momentum corresponds to the Fermi energy
~2 kF2 ~2 (3π 2 n)2/3
F = = .
2m 2m
The average kinetic energy per particle, k = Ek /N must be proportional to
the Fermi energy (as confirmed by the calculation presented below). Note
that the average distance r between particles in a system with a density n is
proportional to 1/n1/3 , hence
1
∼ F ∼ .
r2
On the other hand, the Coulomb energy is inversely proportional to the inter-
particle distance,
1
C ∼ .
r
Hence, if the density is high (r is small) the kinetic energy dominates and the
Coulomb energy can be treated as a correction. This is the limit we are going
to study in this example.
The Hamiltonian of the system is given by Eq. (23) without the single
particle potential U and with
e2 1 0
V (r − r 0 ) = 0
e−µ|r−r | ,
4πε0 |r − r |
where we introduce an exponential cut-off factor in order to deal with various
divergences that will appear in the calculation. At the end, after taking the
thermodynamic limit V → ∞, we will put µ = 0. Using the relation
Z
1 4π
d3 re−iq·r e−µr = 2
r q + µ2
one finds the Fourier transform of the Coulomb potential with the cut-off
e2 1
Vq = .
ε0 q + µ 2
2
22
where we used the fact that the density of the background charge must be
equal to that of the electrons. The calculation is done by first shifting the
variables according to r 00 = r − r 0 and then performing the trivial integrals in
the spherical coordinates.
The interaction between the N electrons at the points ri and the uniform
background is given by
e2 X e2 1 N 2
Z
n
Ee−B = − d3 r e−µ|r−ri | = − ,
4πε0 i |r − ri | ε0 µ 2 V
e2 1 X † †
Vq=0 = a a 0 0 ak0 σ0 akσ .
2Vε0 µ2 kk0 σσ0 kσ k σ
Using the anticommutation relations and the definition of the number operator
we get
a†kσ a†k0 σ0 ak0 σ0 akσ = n̂kσ (n̂k0 σ0 − δkk0 δσσ0 ) .
We can then perform the sums using the definition of the total particle number
operator to get
e2 1
VC,q=0 = (N̂ 2 − N̂ ).
2Vε0 µ2
Since we work with states with a fixed total number of particles we can replace
the total particle number operator N̂ by its eigenvalue N . Thus, the q = 0
term reduces to a number, corresponding to the (direct) Coulomb energy of
the electron gas,
e2 1
Ee = (N 2 − N ).
2V ε0 µ2
Note that the second term yields the contribution to the total energy propor-
tional to N/V = n. The corresponding energy per particle is proportional to
n/N = 1/V and vanishes in the limit V → ∞, n = const.
As we can see, neglecting the vanishing term in the last expression the
three Coulomb energies cancel, as expected.
Hence, our task is to calculate the approximate ground state energy of the
system described by the Hamiltonian
H = H0 + V,
where
X ~2 q 2
H0 = a†q aq
q
2m
is the Hamiltonian of non-interacting fermions and
1 e2 X 0 1 †
V = a 0 a† ak 0 ak
2V ε0 qkk0 q 2 k +q k−q
23
is the Coulomb energy which will be treated as a perturbation. Here the prime
means that the q = 0 term should be ommitted in the summation. We will
calculate the leading order correction in the first order of the perturbation
theory.
The unperturbed ground state of the Hamiltonian H0 is formed by creating
the electrons in the states with k ≤ kF ,
Y †
|ψ0 i = ak,σ |0i.
k≤kF ,σ
The unperturbed eigenvalue is just the kinetic energy of all these particles,
X ~2 k 2 Z kF
(0) ~2 V 4 3~2 kF2 3
E = = 2 3
4π k = N = N F ,
k≤k σ
2m 2m (2π) 0 10m 5
F
where we used Eq. (25) to write V = 3π 2 N/kF3 and changed the summation
into integration according to the usual rule
V
X Z
... → 3
d3 k . . . .
k
(2π)
One defines the radius r0 such that a sphere of radius r0 has the volume
equal to the volume per particle in the system,
1 4
= πr03 ,
n 3
and the dimensionless quantity rs = r0 /a0 , where
4πε0 ~2
a0 =
me2
is the Bohr radius. Expressed in terms of the new parameters, the kinetic
energy is
2/3
(0) e2 1 3 9π e2 2.21
E = N ≈ N.
8πε0 a0 rs2 5 4 8πε0 a0 rs2
The first order correction resulting from the Coulomb term is
e2 X 0 1
E (1) = 2
hψ0 |a†k+q,σ a†k0 −q,σ0 ak0 ,σ0 ak,σ |ψ0 i.
2ε0 V kk0 qσσ0 q
a non-zero term appears in the sum only if two electrons in the same state are
annihilated and created. As the term q = 0 is excluded, the only such term is
δσσ0 δk0 ,k+q a†k+q,σ a†k,σ ak+q,σ ak,σ = −δσσ0 δk0 ,k+q n̂k+q,σ n̂k,σ .
Since the state |ψ0 i is an eigenstate we can replace the number operators by
their values and the energy correction becomes
e2 X0 1
E (1) = − nk+q,σ nk,σ
2ε0 V kqσ q 2
e2 V2
Z Z
1
= − 2 6
d k d3 q 2 Θ(kF − |q + k|)Θ(kF − k).
3
2ε0 V (2π) q
24
Upon performing the integration and expressing the result by a0 and rs one
gets
1/3
(1) e2 9π 3N e2 0.916
E =− ≈− N.
8πa0 ε0 rs 4 2π 8πa0 ε0 rs
Hence the leading terms in the energy of the system are
e2
2.21 0.916
E= N − .
8πa0 ε0 rs2 rs ...
This expansion is valid for rs 1. The omitted terms are called correlation
corrections and physically correspond to the fact that the probability of finding
an electron very close to another one is low because of Coulomb repulsion (that
is, the electron positions are correlated). Calculation of these terms requires
more advanced techniques of the many-body theory.
Acknowledgment
I am grateful to Dominik Wójt for pointing out many mistakes in the previous
version of this text and the other parts of these lecture notes.
25