2nd Quantization
2nd Quantization
87
88 CHAPTER 3. FERMIONS AND BOSONS
Obviously, Pkl† = Pkl , so the eigenvalue λkl is real. Then, from Eqs. (3.5) and
(3.6) immediately follows
with the two possible solutions: λ = 1 and λ = −1. From Eq. (3.6) it follows
that, for λ = 1, the wave function |Ψi is symmetric under particle exchange
whereas, for λ = −1, it changes sign (i.e., it is “anti-symmetric”).
This result was obtained for an arbitrary pair of particles and we may
expect that it is straightforwardly extended to systems with more than two
particles. Experience shows that, in nature, there exist only two classes of
microparticles – one which has a totally symmetric wave function with re-
spect to exchange of any particle pair whereas, for the other, the wave func-
tion is antisymmetric. The first case describes particles with Bose-Einstein
statistics (“bosons”) and the second, particles obeying Fermi-Dirac statistics
(“fermions”)3 .
The one-to-one correspondence of (anti-)symmetric states with bosons (fer-
mions) is the content of the spin-statistics theorem. It was first proven by Fierz
[Fie39] and Pauli [Pau40] within relativistic quantum field theory. Require-
ments include 1.) Lorentz invariance and relativistic causality, 2.) positivity
of the energies of all particles and 3.) positive definiteness of the norm of all
states.
|Ψj1 ,j2 i± = C12 {|Ψj1 ,j2 i + A12 P12 |Ψj1 ,j2 i} , (3.8)
1
|Ψj1 ,j2 i± = √ {|Ψj1 ,j2 i ± P12 |Ψj1 ,j2 i} ≡ Λ±
12 |Ψj1 ,j2 i (3.10)
2
where,
1
Λ±12 = √ {1 ± P12 }, (3.11)
2
denotes the (anti-)symmetrization operator of two particles which is a linear
combination of the identity operator and the pair permutation operator.
The extension of this result to 3 fermions or bosons is straightforward. For
3 particles (1, 2, 3) there exist 6 = 3! permutations: three pair permutations,
(2, 1, 3), (3, 2, 1), (1, 3, 2), that are obtained by acting with the permuation op-
erators P12 , P13 , P23 , respectively on the initial configuration. Further, there
are two permutations involving all three particles, i.e. (3, 1, 2), (2, 3, 1), which
are obtained by applying the operators P13 P12 and P23 P12 , respectively. Thus,
the three-particle (anti-)symmetrization operator has the form
1
Λ±
123 = √ {1 ± P12 ± P13 ± P23 + P13 P12 + P23 P12 }, (3.12)
3!
where we took into account the necessary sign change in the case of fermions
resulting for any pair permutation.
This result is generalized to N particles where there exists a total of N !
permutations, according to4
|Ψ{j} i± = Λ±
1...N |Ψ{j} i,
(3.13)
4
This result applies only to fermions. For bosons the prefactor has to be corrected, cf.
Eq. (3.25).
3.2. N -PARTICLE WAVE FUNCTIONS 91
1 X
Λ±
1...N = √ sign(P )P̂ (3.14)
N ! P ǫSN
where the sum is over all possible permutations P̂ which are elements of the
permutation group SN . Each permutation P has the parity, sign(P ) = (±1)Np ,
which is equal to the number Np of successive pair permuations into which P̂
can be decomposed (cf. the example N = 3 above). Below we will construct
the (anti-)symmetric state |Ψ{j} i± explicitly. But before this we consider an
alternative and very efficient notation which is based on the occupation number
formalism.
The properties of the (anti-)symmetrization operators Λ± 1...N are analyzed
in Problem 1, see Sec. 3.8.
Here {n} denotes the total set of occupation numbers of all single-particle
orbitals. Since this is the complete information about the N -particle system
these states form a complete system which is orthonormal by construction of
the (anti-)symmetrization operators,
F + = H0 ∪ H1+ ∪ H2+ ∪ . . . ,
F − = H0 ∪ H1− ∪ H2− ∪ . . . . (3.17)
Here, we included the vacuum state |0i = |0, 0, . . . i which is the state without
particles which belongs to both Fock spaces.
and all hamiltonians commute, [ĥi , ĥj ] = 0, for all i and j. Then all par-
ticles have common eigenstates, and the total wave function (prior to anti-
symmetrization) has the form of a product
where the argument of the orbitals denotes the number (index) of the particle
that occupies this orbital. As we have just seen, for fermions, all orbitals have
to be different. Now, the anti-symmetrization of this state can be performed
immediately, by applying the operator Λ− 1...N given by Eq. (3.14). For two
particles, we obtain
1
|Ψj1 ,j2 i− = √ {|φj1 (1)i|φj2 (2)i − |φj1 (2)i|φj2 (1)i} =
2!
= |0, 0, . . . , 1, . . . , 1, . . . i. (3.19)
In the last line, we used the occupation number representation, which has
everywhere zeroes (unoccupied orbitals) except for the two orbitals with num-
bers j1 and j2 . Obviously, the combination of orbitals in the first line can be
written as a determinant which allows for a compact notation of the general
wave function of N fermions as a Slater determinant,
|φj1 (1)i |φj1 (2)i ... |φj1 (N )i
− 1 |φj2 (1)i |φj2 (2)i ... |φj2 (N )i
|Ψj1 ,j2 ,...jN i = √ =
N! ... ... ... ...
... ... ... ...
= |0, 0, . . . , 1, . . . , 1, . . . , 1, . . . , 1, . . . i. (3.20)
In the last line, the 1’s are at the positions of the occupied orbitals. This
becomes obvious if the system is in the ground state, then the N energetically
lowest orbitals are occupied, j1 = 1, j2 = 2, . . . jN = N , and the state has
the simple notation |1, 1, . . . 1, 0, 0 . . . i with N subsequent 1’s. Obviously, the
anti-symmetric wave function is normalized to one.
As discussed in Sec. 3.2.1, the (anti-)symmetric states form an orthonormal
basis in Fock space. For fermions, the restriction of the occupation numbers
leads to a slight modification of the completeness relation which we, therefore,
repeat:
h{n}|{n′ }i = δn1 ,n′1 δn2 ,n′2 . . . ,
1 X
X 1
. . . |{n}ih{n}| = 1. (3.21)
n1 =0 n2 =0
the main difference to the fermions is the plus sign. Thus, this wave function
is not represented by a determinant, but this combination of products with
positive sign is called a permanent.
The plus sign in the wave function (3.22) has the immediate consequence
that the situation j1 = j2 now leads to a physical state, i.e., for bosons, there
is no restriction on the occupation numbers, except for their normalization
∞
X
np = N, np = 0, 1, 2, . . . N, ∀p. (3.23)
p=1
Thus, the two single-particle orbitals |φj1 i and |φj2 i occuring in Eq. (3.22) can
accomodate an arbitrary number of bosons. If, for example, the two particles
are both in the state |φj i, the symmetric wave function becomes
|Ψj,j i+ = |0, 0, . . . , 2, . . . i =
1
= C(nj ) √ |φj (1)i|φj (2)i + |φj (2)i|φj (1)i , (3.24)
2!
where the coefficient C(nj ) is introduced to assure the normalization condition
+
hΨj,j |Ψj,j i+ = 1. Since the two terms in (3.24) are identical √ the normalization
gives 1 = 4|C(nj )|2 /2, i.e. we obtain C(nj = 2) = 1/ 2. Repeating this
analysis for a state with an arbitrary occupation number nj , there will be nj !
√
identical terms, and we obtain the general result C(nj ) = 1/ P nj . Finally, if
there are several states with occupation numbers n1 , n2 , . . . with ∞ p=1 np = N ,
−1/2
the normalization constant becomes C(n1 , n2 , ...) = (n1 ! n2 ! . . . ) . Thus, for
+
the case of bosons action of the symmetrization operator Λ1...N , Eq. (3.14), on
the state |Ψj1 ,j2 ,...jN i will not yield a normalized state. A normalized symmetric
state is obtained by the following prescription,
1
|Ψj1 ,j2 ,...jN i+ = √ Λ+ |Ψj1 ,j2 ,...jN i (3.25)
n1 !n2 !... 1...N
1 X
Λ+
1...N = √ P̂ . (3.26)
N ! P ǫSN
Hence the symmetric state contains the prefactor (N !n1 !n2 !...)−1/2 in front of
the permanent.
An example of the wave function of N bosons is
k
X
+
|Ψj1 ,j2 ,...jN i = |n1 n2 . . . nk , 0, 0, . . . i, np = N, (3.27)
p=1
3.2. N -PARTICLE WAVE FUNCTIONS 95
where np 6= 0, for all p ≤ k, whereas all orbitals with the number p > k are
empty. In particular, the energetically lowest state of N non-interacting bosons
(ground state) is the state where all particles occupy the lowest orbital |φ1 i,
i.e. |Ψj1 ,j2 ,...jN i+
GS = |N 0 . . . 0i. This effect of a macroscopic population which
is possible only for particles with Bose statistics is called Bose-Einstein con-
densation. Note, however, that in the case of interaction between the particles,
a permanent constructed from the free single-particle orbitals will not be an
eigenstate of the system. In that case, in a Bose condensate a finite fraction of
particles will occupy excited orbitals (“condensate depletion”). The construc-
tion of the N-particle state for interacting bosons and fermions is subject of
the next section.
and the N -particle wave function will, in general, deviate from a product of
single-particle orbitals. Moreover, there is no reason why interacting particles
should occupy single-particle orbitals |φp i of a non-interacting system.
The solution to this problem is based on the fact that the (anti-)symmetric
states, |Ψ{j} i± = |{n}i, form a complete orthonormal set in the N -particle
Hilbert space, cf. Eq. (3.16). This means, any symmetric or antisymmetric
state can be represented as a superposition of N -particle permanents or deter-
minants, respectively,
X
±
|Ψ{j} i± = C{n} |{n}i (3.29)
{n}, N =const
The effect of the interaction between the particles on the ground state wave
function is to “add” contributions from determinants (permanents) involving
higher lying orbitals to the ideal wave function, i.e. the interacting ground
state includes contributions from (non-interacting) excited states. For weak
interaction, we may expect that energetically low-lying orbitals will give the
dominating contribution to the wave function. For example, for two fermions,
the dominating states in the expansion (3.29) will be |1, 1, 0, . . . i, |1, 0, 1, . . . i,
|1, 0, 0, 1 . . . i, |0, 1, 1, , 0 . . . i and so on. The computation of the ground state of
an interacting many-particle system is thus transformed into the computation
96 CHAPTER 3. FERMIONS AND BOSONS
±
of the expansion coefficients C{n} . This is the basis of the exact diagonalization
method or configuration interaction (CI). It is obvious that, if we would have
obtained the eigenfunctions of the interacting hamiltonian, it would be repre-
sented by a diagonal matrix in this basis whith the eigenvalues populating the
diagonal.6
The main problem of CI-type methods is the exponential scaling with the
number of particles which dramatically limits the class of solvable problems.
Therefore, in recent years a large variety of approximate methods has been
developed. Here we mention multiconfiguration (MC) approaches such as
MC Hartree or MC Hartree-Fock which exist also in time-dependent variants
(MCTDH and MCTDHF), e.g. [MMC90] and are now frequently applied to
interacting Bose and Fermi systems. In this method not only the coefficients
C ± (t) are optimized but also the orbitals are adapted in a time-dependent
fashion. The main advantage is the reduction of the basis size, as sompared to
CI. A recent time-dependent application to the photoionization of helium can
be found in Ref. [HB11]. Another very general approach consists in subdivid-
ing the N -particle state in various classes with different properties. This has
been termed “Generalized Active Space” (or restricted active space) approach
and is very promising due to its generality [HB12, HB13]. An overview on first
results is given in Ref. [HHB14].
6
This N -particle state can be constructed from interacting single-particle orbitals as
well. These are called “natural orbitals” and are the eigenvalues of the reduced one-particle
density matrix. For a discussion see [SvL13].
3.3. SECOND QUANTIZATION FOR BOSONS 97
yielding the same explicit definition that is familiar from the harmonic os-
cillator7 : the adjoint of â†i is indeed an annihilation operator reducing the
7
See our results for coupled harmonic oscillators in section 2.3.2.
98 CHAPTER 3. FERMIONS AND BOSONS
occupation of orbital |φi i by one. In the third line of Eq. (3.31) we introduced
the modified Kronecker symbol in which the occupation number of orbital i is
missing,
i
δ{n},{n ′} = δn1 ,n′1 . . . δni−1 ,n′i−1 δni+1 ,n′i+1 . . . . (3.32)
ik
δ{n},{n ′} = δn1 ,n′1 . . . δni−1 ,n′i−1 δni+1 ,n′i+1 . . . .δnk−1 ,n′k−1 δnk+1 ,n′k+1 . . . . (3.33)
which proves the statement since no restrictions with respect to i and k were
made. Analogously one proves the relations (3.34), see problem 18 . We now
consider the second quantization representation of important operators.
the creation operator(s). For example, single and two-particle states with the
proper normalization are obtained via
where, in the second (third) line, the 1 (2) stands on position i, whereas in
the last line the 1’s are at positions i and j. This is readily generalized to an
arbitrary symmetric N -particle state according to9 .
1 n 1 n 2
|n1 , n2 , . . . i = √ â†1 â†2 . . . |0i (3.36)
n1 !n2 ! . . .
whereas, for ni = 0, â†i âi |{n}i = 0. Thus, the symmetric state |{n}i is an
eigenstate of n̂i with the eigenvalue coinciding with the occupation number
ni of this state. In other words: all n̂i have common eigenfunctions with the
hamiltonian and commute, [n̂i , H] = 0.
The total particle number operator is defined as
∞
X ∞
X
N̂ = n̂i = â†i âi , (3.38)
i=1 i=1
because
P∞ its action yields the total number of particles in the system: N̂ |{n}i =
n
i=1 i |{n}i = N |{n}i. Thus, also N̂ commutes with the hamiltonian and
has the same eigenfunctions.
9
The origin of the prefactors was discussed in Sec. 3.2.4 and is also analogous to the case
of the harmonic oscillator Sec. 2.3.
100 CHAPTER 3. FERMIONS AND BOSONS
Single-particle operators
Consider now a general single-particle operator10 defined as
N
X
B̂1 = b̂α , (3.39)
α=1
where b̂α acts only on the variables associated with particle with number “α”.
We will now transform this operator into second quantization representation.
To this end we define the matrix element with respect to the single-particle
orbitals
bij = hi|b̂|ji, (3.40)
and the generalized projection operator11
N
X
Π̂ij = |iiα hj|α , (3.41)
α=1
Proof:
We first expand b̂, for an arbitrary particle α into a basis of single-particle
orbitals |ii = |φi i,
∞
X ∞
X
b̂ = |iihi|b̂|jihj| = bij |iihj|,
i,j=1 i,j=1
where we used the definition (3.40) of the matrix element. With this result we
can transform the total operator (3.39), using the definition (3.41),
N X
X ∞ ∞
X
B̂1 = bij |iiα hj|α = bij Π̂ij , (3.43)
α=1 i,j=1 i,j=1
10
Examples are the total momentum, total kinetic energy, angular momentum or potential
energy of the system.
11
For i = j this definition contains the standard projection operator on state |ii, i.e. |iihi|,
whereas for i 6= j this operator projects onto a transition, i.e. transfers an arbitrary particle
from state |ji to state |ii.
3.3. SECOND QUANTIZATION FOR BOSONS 101
N
1 X
Π̂ij |{n}i = √ Λ+ |iiα hj|α · |j1 i|j2 i . . . |jN i. (3.44)
n1 !n2 ! . . . 1...N α=1
The product state is constructed from all orbitals including the orbitals |ii and
|ji. Among the N factors there are, in general, ni factors |ii and nj factors
|ji. Let us consider two cases.
1), j 6= i: Since the single-particle orbitals form an orthonormal basis, hj|ji =
1, multiplication with hj|α reduces the number of occurences of orbital j in the
product state by one, whereas multiplication with |iiα increases the number of
orbitals i by one. The occurence of nj such orbitals (occupied by nj particles)
in the product state gives rise to an overall factor of nj because nj terms of
the sum will yield a non-vanishing contribution.
Finally, we compare this result to the properly symmetrized state which
follows from |{n}i by increasing ni by one and decreasing nj by one will be
denoted by
{n}ij = |n1 , n2 . . . ni + 1 . . . nj − 1 . . . i
1
=p Λ+
1...N · |j1 i|j2 i . . . |jN i. (3.45)
n1 ! . . . (ni + 1)! . . . (nj − 1)! . . .
It contains the same particle number N as the state |{n}i but, due to the differ-
√ √
ent orbital occupations, the prefactor in front of Λ+1...N differs by n j / n i + 1,
compared to the one in Eq. (3.44) which we, therefore, can rewrite as
√
ni + 1
{n}ij
Π̂ij |{n}i = nj √
nj
= â†i âj |{n}i. (3.46)
Thus, the results (3.46) and (3.47) can be combined to the operator identity
N
X
|iiα hj|α = â†i âj (3.48)
α=1
where bi are the eigenvalues of b̂. Equation (3.49) naturally generalizes the
familiar spectral representation of quantum mechanical operators to the case
of many-body systems with arbitrary variable particle number.
Two-particle operators
A two-particle operator is of the form
N
1 X
B̂2 = b̂α,β , (3.50)
2! α6=β=1
Proof:
We expand b̂ for an arbitrary pair α, β into a basis of two-particle orbitals
13
See problem 2.
3.3. SECOND QUANTIZATION FOR BOSONS 103
The second sum is readily transformed, using the property (3.48) of the sigle-
particle states. We first extend the summation over the particles to include
α = β,
N
X N
X N
X N
X
|iiα |jiβ hk|α hl|β = |iiα hk|α |jiβ hl|β − δk,j |iiα hl|α
α6=β=1 α=1 β=1 α=1
14
Note that the order of the creation operators and of the annihilators, respectively, is
arbitrary. In Eq. (3.52) we have chosen an ascending order of the creators (same order as the
indices of the matrix element) and a descending order of the annihilators, since this leads
to an expression which is the same for Bose and Fermi statistics, cf. Sec. 3.4.1.
104 CHAPTER 3. FERMIONS AND BOSONS
where we used the general matrix elements with respect to k-particle product
states, bj1 ...jk m1 ...mk = hj1 . . . jk |b̂|m1 . . . mk i. Note again the inverse ordering of
the annihilation operators. Obviously, the result (3.55) includes the previous
examples of single and two-particle operators as special cases.
Due to the Pauli principle we expect that there will be no additional prefactors
resulting from multiple occupations of orbitals, as in the case of bosons15 . So
far we do not know how these operators look like explicitly. Their definition
has to make sure that the N -particle states have the correct anti-symmetry
and that application of any creator more than once will return zero.
To solve this problem, consider the example of two fermions which can
occupy the orbitals k or l. The two-particle state has the symmetry |kli =
−|lki, upon particle exchange. The anti-symmetrized state is constructed of
the product state of particle 1 in state k and particle 2 in state l and has the
properties
† † † †
Λ− −
1...N |kli = âl âk |0i = |11i = −Λ1...N |lki = −âk âl |0i, (3.57)
i.e., it changes sign upon exchange of the particles (third equality). This
indicates that the state depends on the order in which the orbitals are filled,
i.e., on the order of action of the two creation operators. One possible choice
is used in the above equation and immediately implies that
Figure 3.2: Illustration of the phase factor α in the fermionic creation and
annihiliation operators. The single-particle orbitals are assumed to be in a
definite order (e.g. with respect to the energy eigenvalues). The position of
the particle that is added to (removed from) orbital φk is characterized by the
number αk of particles occupying orbitals to the left, i.e. with energies smaller
than ǫk .
16
where we have introduced the
2 anti-commutator . In the special case, k = l,
we immediately obtain â†k = 0, for an arbitrary state, in agreement with
the Pauli principle. Calculating the hermitean adjoint of Eq. (3.58) we obtain
that the anti-commutator of two annihilators vanishes as well,
We expect that this property holds for any two orbitals k, l and for any N -
particle state that involves these orbitals.
Now we can introduce an explicit definition of the fermionic creation oper-
ator which has all these properties. The operator creating a fermion in orbital
k of a general many-body state is defined as17
X
â†k | . . . , nk , . . . i = (1 − nk )(−1)αk | . . . , nk + 1, . . . i, αk = nl (3.60)
l<k
where the prefactor explicitly enforces the Pauli principle, and the sign factor
takes into account the position of the orbital k in the many-fermion state and
the number of fermions standing “to the left” of the “newly created” particle,
cf. Fig. 3.2. In other words, with αk pair exchanges (anti-commutations) the
particle would move from the leftmost place to the position (e.g. according
to an ordering with respect to the orbital energies Ek ) of orbital k in the N -
particle state. We now derive the annihilation operator by inserting a complete
16
This was introduced by P. Jordan and E. Wigner in 1927.
17
There can be other conventions which differ from ours by the choice of the exponent αk
which, however, is irrelevant for physical observables.
106 CHAPTER 3. FERMIONS AND BOSONS
= (2 − nk )(−1)αk | . . . , nk − 1, . . . i
≡ nk (−1)αk | . . . , nk − 1, . . . i
where, in the third line, we used definition (3.32). Also, αk′ = αk because the
sum involves only occupation numbers that are not altered. Note that the
factor 2 − nk = 1, for nk = 1. However, for nk = 0 the present result is not
correct, as it should return zero. To this end, in the last line we have added
the factor nk that takes care of this case. At the same time this factor does not
alter the result for nk = 1. Thus, the factor 2 − nk can be skipped entirely, and
we obtain the expression for the fermionic annihilation operator of a particle
in orbital k
Using the definitions (3.60) and (3.61) one readily proves the anti-commutation
relations given by the
Theorem: The creation and annihilation operators defined by Eqs. (3.60) and
(3.61) obey the relations
and thus the anti-commutator vanishes as well. Consider now the case18 i < k:
P
âi âk |{n}i = âi nk (−1) l<k nl |n1 . . . nk − 1 . . . i =
P P
= ni nk (−1) l<k nl (−1) l<i nl |n1 . . . ni − 1 . . . nk − 1 . . . i.
Now we compute the result of the action of the exchanged operator pair
P
âk âi |{n}i = âk ni (−1) l<i nl |n1 . . . ni − 1 . . . i =
P P
= ni nk (−1) l<i nl (−1) l<k nl −1 |n1 . . . ni − 1 . . . nk − 1 . . . i,
The only difference compared to the first result is in the additional −1 in the
second exponent. It arises because, upon action of âk after âi , the number of
particles to the left of k is already reduced by one. Thus, the two expressions
differ just by a minus sign, which proves vanishing of the anti-commutator.
Single-particle operators
Consider now again a single-particle operator
N
X
B̂1 = b̂α , (3.67)
α=1
Proof:
As for bosons, cf. Eq. (3.43), we have
N X
X ∞ ∞
X
B̂1 = bij |iiα hj|α = bij Π̂ij , (3.69)
α=1 i,j=1 i,j=1
where Π̂ij was defined by (3.41), and it remains to show that Π̂ij = â†i âj , for
fermions as well. To this end we consider action of Π̂ij on an anti-symmetric
state, taking into accont that Π̂ij commutes with the anti-symmetrization op-
erator Λ−1...N , Eq. (3.14),
N
1 XX
Π̂ij |{n}i = √ sign(π)|iiα hj|α · |j1 iπ(1) |j2 iπ(2) . . . |jN iπ(N ) . (3.70)
N ! α=1 πǫSN
If the product state does not contain the orbital |ji expression (3.70) vanishes,
due to the orthogonality of the orbitals. Otherwise, let jk = j. Then hj|jk i = 1,
and the orbital |jk i will be replaced by |ii, unless the state |ii is already present,
then we again obtain zero due to the Pauli principle, i.e.
where we used the notation (3.45). What remains is to figure out the sign
change due to the removal of a particle from the i-th orbital and creation of
one in the k-th orbital. To this end we first P
“move” the (empty) orbital |ji
past all orbitals to the left occupied by αj = p<j np particles, requiring just
αj pair permutations P and sign changes. Next we move the “new” particle to
orbital “i” past αi = p<i np particles occupying the orbitals with an energy
3.4. SECOND QUANTIZATION FOR FERMIONS 109
lower then Ei leading to αi pair exchanges and sign changes20 . Taking into
account the definitions (3.60) and (3.61) we obtain21
which, together with the equation (3.69), proves the theorem. Thus, the second
quantization representation of single-particle operators is the same for bosons
and fermions.
Two-particle operators
As for bosons, we now derive the second quantization representation of a two-
particle operator B̂2 .
Theorem: The second quantization representation of a two-particle operator
is given by
∞
1 X
B̂2 = bijkl â†i â†j âl âk (3.73)
2! i,j,k,l=1
Proof:
As for bosons, we expand B̂ into a basis of two-particle orbitals |iji = |φi i|φj i,
∞ N
1 X X
B̂2 = bijkl |iiα |jiβ hk|α hl|β , (3.74)
2! i,j,k,l=1 α6=β=1
In the third line we have used the anti-commutation relation (3.63). After
exchanging the order of the two annihilators, which now leads to a sign change,
and inserting this expression into Eq. (3.74), we obtain the final result (3.73).
where we used the general matrix elements with respect to k-particle product
states, bj1 ...jk m1 ...mk = hj1 . . . jk |b̂|m1 . . . mk i. Note again the inverse ordering
of the annihilation operators which exactly agrees with the expression for a
bosonic system. Obviously, the result (3.75) includes the previous examples of
single and two-particle operators as special cases.
The fact that the fermionic occupation numbers have only two possible values
is very similar to the two spin projections of spin 1/2 particles and allows
for a very intuitive description in terms of spinors. Thus, an empty or singly
occupied orbital can be written as a column
1
|0i → − empty state, (3.76)
0
0
|1i → − occupied state, (3.77)
1
3.4. SECOND QUANTIZATION FOR FERMIONS 111
where the matrix of âk is the transposed of that of â†k and we introduced the
short-hand notation Ak for the matrix δnk ,0 δn′k ,1 in the space of single-particle
orbitals |φk i 23 .
D E
{n} â†k {n′ } = (−1)αk δ{n},{n
k †
′ } Ak (3.86)
where the original prefactor 1 − n′k has been transformed into an additional
Kronecker delta for nk . The matrix of the annihilation operator is
23
We summarize the main properties of the matrices Ak and A†k which are a consequence
of the properties of â†k and âk and can be verified by direct matrix multiplication:
2
1. A2k = A†k = 0.
0 0
2. A†k Ak = = nk 1̂k , – a diagonal matrix with the eigenvalues of n̂k on the
0 1
diagonal, cf. Eq. (3.83).
1 0
3. Ak A†k = = (1 − nk )1̂k = 1 − A†k Ak , i.e. A†k and Ak anti-commute.
0 0
and (3.86) for the creator successively we obtain, for the case k 6= l,
D E XD E
{n} â†l âk {n′ } = {n} â†l {n̄} h{n̄} |âk | {n′ }i =
{n̄}
X
α′k
= (−1) (−1)ᾱl δn̄k ,0 δ{n̄},{n
k l
′ } δn̄k ,n′ −1 δnl ,1 δ{n̄},{n} δn̄l +1,nl
k
{n̄}
X
α′k
kl
= (−1) δnl ,1 δ{n},{n ′} (−1)ᾱl δn̄k ,0 δn̄k ,nk δn̄l ,n′l δn̄k ,n′k −1 δn̄l +1,nl
n̄,n̄k
†
X X
= kl
(−1)αk′ l δ{n},{n ′ } Al Ak , αk ′ l = n′m + nm , (3.88)
m<k m<l
which is a diagonal matrix in all orbitals except k and l whereas, with respect
to orbital k, it has the structure of the matrix (3.86) and, for orbital l, the form
of matrix (3.87). Note that the occupation numbers entering the exponent αk′ l
are restricted by the Kronecker symbols. For the case k = l we recover the
matrix of the particle number operator which is completely diagonal24
D E
{n} â†k âk {n′ } = h{n} |n̂k | {n′ }i = nk δ{n},{n′ } . (3.89)
With the results (3.88) and (3.89) we readily obtain the matrix represen-
tation of a single-particle operator, defined by Eq. (3.67),
D E X∞ D E
{n} B̂1 {n′ } = blk {n} â†l âk {n′ } (3.90)
k,l=1
First, for a diagonal operator B diag , blk = bk δkl , the result is simply
D E ∞
X N
X
{n} B̂1diag {n′ } = δ{n},{n′ } bk nk = δ{n},{n′ } bnk . (3.91)
k=1 k=1
where, in the last equality, we have simplified the summation by including only
the occupied orbitals which have the numbers n1 , n2 . . . nN .
24
ThisPresult is contained in expression (3.88). Indeed, in the special case k = l we obtain
αk′ l → m<k (n′m + nm ), δ{n},{n
kl k
′ } → δ{n},{n′ } and the matrix product in k-orbital space
yields A†k Ak → nk δnk ,n′k [cf. property 2 in footnote 23] which combines to the results (3.89).
114 CHAPTER 3. FERMIONS AND BOSONS
where αj̄k′ = p<j n̄p + p<k n′ etc. Performing the summation, with the help
P P
25
The non-diagonal matrix elements are transformed to summation over occupied orbitals
as
∞ D E N
{n} â†l âk {n′ } =
X X
bnl nk {n} â†nl ânk {n′ } ,
blk
k6=l=1 k6=l=1
where it remains to carry out the action oft the two operators. Note that the sign of the
result is different for nl < nk and nl > nk .
26
We first rewrite
X X
il
δ{n},{n̄ δ jk
p } {n̄p },{n′ }
= iljk
δ{n},{n ′ } δnj ,n̄j δnk ,n̄k δn̄i ,n′ δn̄l ,n′ ,
i l
{n̄p } n̄i n̄l n̄j n̄k
This is a general result which also contains the cases of equal index pairs.
Then, proceeding as in footnote 24, we obtain the results for the special cases.
jk †
i=l: (−1)αj′ k′ δ{n},{n ′ } ni Aj Ak
†
j=k: (−1)αil δ{n},{n
il
′ } nj Ai Al
†
ik
l=j: (−1)αik′ δ{n},{n ′ } (1 − nl )Ai Ak
1X
V̂ = wijkl â†i â†j âl âk , (3.96)
2 ijkl
where property (3.99) follows from the symmetry of the potential w(x, y) =
w(y, x). This allows us to eliminate double counting of pairs from the sum in
27
M. Heimsoth contributed to this section.
116 CHAPTER 3. FERMIONS AND BOSONS
Eq. (3.96)28
∞
X ∞
X
V̂ = −
wijkl â†j â†i âk âl , (3.100)
1≤i<j 1≤k<l
−
wijkl = wijkl − wijlk , (3.101)
Each of the two vectors contains N particles (the interaction does not change
the particle number), i.e. exactly N occupied orbitals which are all different.
So the sums over i, j and k, l, in fact, run over two (possibly different) sets of N
orbitals with the numbers (m1 , m2 . . . mN ) and (m′1 , m′2 . . . m′N ), respectively,29
Using the definitions of the creation and annihilation operators, Eqs. (3.60),
28
We summarize the main steps: First, using the anti-commutation relations of the anni-
hilators and perfoming an index change, we transform (the contribution k = l vanishes),
X X X
−
wijkl âl âk = (wijkl − wijlk )âl âk = wijkl âl âk .
kl k<l k<l
Extending this to the sum over i, j and using the symmetry properties (3.98), we obtain
(wijkl − wijlk )â†i â†j âl âk = (wijkl − wjikl − wijlk + wjilk )â†i â†j âl âk =
X X
ij,k<l i<j,k<l
29
by |{m}i = |{m}i(|{n}i) we will denote the subset of N occupied orbitals contained
in the state |{n}i. For example, a three-particle state |{n}i = |1, 0, 0, 1, 1i transforms into
|m1 m2 m3 i where the mi point to the original orbitals with numbers m1 = 1, m2 = 4, m3 = 5.
Note that the matrix h{n}|V̂ |{n′ }i is diagonal in all orbitals missing simultaneously in h{m}|
and |{m′ }i.
3.4. SECOND QUANTIZATION FOR FERMIONS 117
(3.61), and taking advantage of the operator order in (3.103)30 , the operators
can be evaluated, with the result
N
X N
X
′ −
h{m}|V̂ |{m }i = (−1)i+j+k+l wm ′
i mj m m
′
′ h{m}|mi ,mj |{m }im′ ,m′ ,
k l
k l
1≤i<j 1≤k<l
(3.104)
where the notation |{m′ }im′k ,m′l means that the single-particle orbitals with
number m′k and m′l are missing in the state |{m′ }i which now is a state of
N − 2 particles, and similarly for h{m}|mi ,mj . The scalar product of the two
anti-symmetric N −2-particle states in (3.104) is non-zero only if the two states
contain N − 2 identical orbitals. To simplify the analysis, in Eq. (3.104) we
have moved the missing orbitals to positions one and two in the states thereby
accumulating the total sign factor contained in this expression. Thus, the
remaining orbitals are not only identical but they also have identical numbers,
i.e. m3 = m′3 , m4 = m′4 , . . . .
Finally, expression (3.104) will be only non-zero if the missing orbitals fall
in one of three cases31 :
1. The two states are identical, {n} ≡ {n′ } and, consequently {m} ≡ {m′ }.
Then Eq. (3.104) yields
N
X
h{n}|V̂ |{n′ }i = δ{n},{n′ } −
wm i mj mi mj
. (3.105)
1≤i<j
2. The two states are identical except for one orbital: the orbital mp with
number p is present in state h{m}| but is missing in state |{m′ }i which,
instead, contains an orbital mr with number r missing in h{m}|. Then
the scalar product of the two N − 2 particle states is nonzero only if both
these states are annihilated and Eq. (3.104) yields32
N −1
mp m′r
X
′ † −
h{n}|V̂ |{n }i = δ{n},{n ′ } Amp Am′r (−1)p+r · Θ(p, r, i) wm i mp mi mr
′
1≤i,i6=p,r
(3.106)
Here Θ(p, r, i) = −1, if either mp < mi or m′r
< mi , otherwise Θ(p, r, i) =
1. This case describes single-particle excitations where |{n′ }i = |{n}rp i.
30
Since i < j and k < l, the signs produced by the first and second operators are inde-
pendent of each other.
31
Thereby we return to the full vectors (including the empty orbitals) and restore the
delta functions.
32
To obtain the correct sign we move the orbitals p and r to the last place in the product
in state h{n}| and in |{n′ }i, respectively and count the number of transpositions (difference).
118 CHAPTER 3. FERMIONS AND BOSONS
3. The two states are identical except for two orbitals with the numbers mp
and mq in h{m}| and m′r and m′s in |{m′ }i, respectively. Without loss of
generality we can use mp < mq and m′r < m′s . Then Eq. (3.104) yields
m m m′ m′s −
h{n}|V̂ |{n′ }i = δ{n},{n
p q r
A†mp Am′r A†mq Am′s (−1)p+q+r+s wm
′} p mq mr ms
′ ′
(3.107)
This case describes two-particle excitations where |{n′ }i = |{n}rspq i.
These results are known as Slater-Condon rules and were obtained by those
two authors in 1929 and 1930 [Sla29, Con30] and are of prime importance for
wave function based many-body methods such as configuration interaction
(CI) and Multiconfiguration Hartree-Fock (MCHF) and their time-dependent
extensions, e.g. [HHB14]. Similarly this representation is used in configura-
tion path integral Monte Carlo simulations of strongly correlated fermions,
e.g. [SBF+ 11] and references therein.
Figure 3.3: Illustration of the relation of the field operators to the second
quantization operators defined on a general basis {φi (x)}. The field operator
ψ̂ † (x) creates a particle at space point x (in spin state |σi) to which all single-
particle orbitals φi contribute. The orbitals are vertically shifted for clarity.
Here x = (r, σ), i.e. φi (x) is an eigenstate of the operator r̂, and the φi (x)
form a complete orthonormal set. Obviously, these operators have the desired
property to create (annihilate) a particle at space point r in spin state σ. From
the (anti-)symmetrization properties of the operators a and a† we immediately
obtain
h i
ψ̂(x), ψ̂(x′ ) = 0, (3.110)
h i∓
ψˆ† (x), ψˆ† (x′ ) = 0, (3.111)
∓
h i
ψ̂(x), ψˆ† (x′ ) = δ(x − x′ ). (3.112)
∓
h i ∞
X h i
† ′
ψ̂(x), ψ̂ (x ) = φi (x)φ∗j (x′ ) âi , â†j =
∓ ∓
i,j=1
X ∞
= φi (x)φ∗i (x′ ) = δ(x − x′ ) = δ(r − r′ )δσ,σ′ ,
i=1
where, in the last line, we used the representation of the delta function in
terms of a complete set of functions.
120 CHAPTER 3. FERMIONS AND BOSONS
Single-particle operators
and obtain for the operator, taking into account the definitions (3.108) and
(3.109),
∞ Z
X
B̂1 = dxdx′ a†i φ∗i (x)hx|b̂|x′ iφj (x′ )aj =
i,j=1
Z
= dxdx′ ψ̂ † (x)hx|b̂|x′ iψ̂(x′ ). (3.115)
For the important case that the matrix is diagonal, hx|b̂|x′ i = b̂(x)δ(x − x′ ),
the final expression simplifies to
Z
B̂1 = dx ψ̂ † (x)b̂(x)ψ̂(x) (3.116)
Consider a few important examples. The first is again the density operator.
In first quantization the operator of the particle density for N particles follows
from quantizing the classical result for point particles,
N
X
n̂(x) = δ(x − xα ), (3.117)
α=1
3.5. FIELD OPERATORS 121
and the operator of the total density is the sum (integral) over all coordinate-
spin states Z Z
N̂ = dx n̂(x) = dx ψ̂ † (x)ψ̂(x), (3.120)
Two-particle operators
In similar manner we obtain the field operator representation of a general
two-particle operator
∞
1 X
B̂2 = hij|b̂|klia†i a†j al ak . (3.123)
2 i,j,k,l=1
33
This is the example of an (anti-)symmetrized pure state which is easily extended to
mixed states.
122 CHAPTER 3. FERMIONS AND BOSONS
Using again the defintion of the field operators the final result for a diagonal
two-particle operator in coordinate representation is
1
Z
B̂2 = dx1 dx2 ψ̂ † (x1 )ψ̂ † (x2 )b̂(x1 , x2 )ψ̂(x2 )ψ̂(x1 ) (3.125)
2
Note again the inverse ordering of the destruction operators which makes this
result universally applicable to fermions and bosons. The most important
example of this representation is the operator of the two-particle interaction,
Ŵ , which is obtained by replacing b̂(x1 , x2 ) → w(x1 − x2 ).
where we took into account that the integral over the finite volume V equals
zero for k 6= k′ and V otherwise.
3.6. MOMENTUM REPRESENTATION 123
Multiplication by φ∗k,σ (x) and integrating over x yields, with the help of con-
dition (3.127),
1
Z Z
ak,σ = dx φ∗k′ σ′ ψ̂(x) = 1/2 d3 r e−ik·r ψ̂(r, σ), (3.128)
V V
Single-particle operators
For a single-particle operator we have, according to our general result, Eq.
(3.69), and denoting x = (r, s), x′ = (r′ , s′ ),
XX †
B̂1 = akσ hkσ|b̂|k′ σ ′ i ak′ σ′
kσ k′ σ ′
XXZ
= dx dx′ a†kσ hkσ|xihx|b̂|x′ ihx′ |k′ σ ′ i ak′ σ′
kσ k′ σ ′
1 XX
Z
dx dx′ a†kσ e−ikr hx|b̂|x′ ieik r ak′ σ′ δσ,s δσ′ ,s′ , (3.130)
′ ′
=
V kσ k′ σ′
124 CHAPTER 3. FERMIONS AND BOSONS
1 X † 1
Z
′
= ak′ σ akσ d3 r ei(k−k )r e−iqr
V σkk′ V V
1X †
= a akσ . (3.133)
V σk k−q,σ
This shows that the Fourier component of the density operator, n̂q , describes
a density fluctuation corresponding to a transition of a particle from state
|φkσ i to state |φk−q,σ i, for arbitrary k. With this result we may rewrite the
single-particle potential (3.132) as
X
V̂ = V ṽq n̂−q . (3.134)
q
3.6. MOMENTUM REPRESENTATION 125
Two-particle operators
We now turn to two-particle operators. Rewriting the general result (3.73) for
a spin-momentum basis, we obtain
1 X X †
B̂2 = ak1 σ1 a†k2 σ2 hk1 σ1 k2 σ2 |b̂|k′1 σ1′ k′2 σ2′ i ak′2 σ2′ ak′1 σ1′ (3.135)
2! k σ k σ ′ ′ ′ ′
1 1 2 2 k 1 σ1 k 2 σ 2
We now apply this result to the interaction potential where the matrix element
in momentum representation was computed before, hk1 σ1 k2 σ2 |ŵ|k′1 σ1′ k′2 σ2′ i =
w̃(k1 − k′1 )δk1 +k2 −k′1 −k′2 δσ1 ,σ1′ δσ2 ,σ2′ , and w̃ denotes the Fourier transform of the
pair interaction, and the interaction does not change the spin of the involved
particles, see problem 6, Sec. 3.8. Inserting this into Eq. (3.135) and intro-
ducing the momentum transfer q = k′1 − k1 = k2 − k′2 , we obtain
1 X X
Ŵ = w̃(q)a†k1 σ1 a†k2 σ2 ak2 −q,σ2 ak1 +q,σ1 , (3.136)
2! k k q σ σ
1 2 1 2
5. Discuss what happened to the sum over α in the derivation of Eq. (3.48).