0% found this document useful (0 votes)
6 views

LinearResponse

Chapter 5 discusses the treatment of time evolution in quantum mechanics using three representations: the Schrödinger picture, the Heisenberg picture, and the interaction picture. Each picture offers a different approach to handle time dependence in quantum systems, particularly focusing on how operators and state vectors evolve over time. The chapter also introduces the time evolution operator and its significance in perturbation theory and linear response theory.
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
6 views

LinearResponse

Chapter 5 discusses the treatment of time evolution in quantum mechanics using three representations: the Schrödinger picture, the Heisenberg picture, and the interaction picture. Each picture offers a different approach to handle time dependence in quantum systems, particularly focusing on how operators and state vectors evolve over time. The chapter also introduces the time evolution operator and its significance in perturbation theory and linear response theory.
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 14

Chapter 5

Time evolution pictures

Using the second quantization procedure, we have so far only treated energy eigenstates
with a trivial time dependence eiω t , instant processes at a single time t = 0, and systems
where interactions are approximated by time-independent mean field theory. But how
does one then treat the general case of time dependence in second quantization? That
question will be addressed in this chapter, where time evolution is discussed using three
representations, or “pictures”: the Schrödinger picture, the Heisenberg picture, and the
interaction picture. These representations are used in the following chapters to develop
general methods for treating many-particle systems.

5.1 The Schrödinger picture


The Schrödinger picture is useful when dealing with a time-independent Hamiltonian H,
i.e. ∂t H = 0. Any other operator A may or may not depend on time. The state vectors
|ψ(t)� does depend on time, and their time evolution is governed by Schrödinger’s equation.
The time-independence of H leads to a simple formal solution:
i
i�∂t |ψ(t)� = H |ψ(t)� ⇒ |ψ(t)� = e− Ht
|ψ0 �. (5.1)
In the following we will measure the energy in units of frequency, such that � drops out
of the time-evolution equations: ε/� → ε and H/� → H. At the end of the calculations
one can easily convert frequencies back to energies. With this notation we can summarize
the Schrödinger picture with its states |ψ(t)� and operators A as:

 states :
 |ψ(t)� = e−iHt |ψ0 �,
The Schrödinger picture operators : A, may or may not depend on time. (5.2)


H, does not depend on time.

To interpret the operator e−iHt we recall that a function f (B) of any operator B is defined
by the Taylor expansion of f ,

� f (n) (0)
f (B) = Bn. (5.3)
n!
n=0

87
88 CHAPTER 5. TIME EVOLUTION PICTURES

While the Schrödinger picture is quite useful for time-independent operators A, it may
sometimes be preferable to collect all time dependencies in the operators and work with
time-independent state vectors. We can do that using the Heisenberg picture.

5.2 The Heisenberg picture


The central idea behind the Heisenberg picture is to obtain a representation where all the
time dependence is transferred to the operators, A(t), leaving the state vectors |ψ0 � time
independent. The Hamiltonian H remains time-independent in the Heisenberg picture.
If the matrix elements of any operator between any two states are identical in the two
representations, then the two representations are fully equivalent. By using Eq. (5.2) we
obtain the identity

�ψ � (t)|A|ψ(t)� = �ψ0� |eiHt Ae−iHt |ψ0 � ≡ �ψ0� |A(t)|ψ0 �. (5.4)

Thus we see that the correspondence between the Heisenberg picture with time-independent
state vectors |ψ0 �, but time-dependent operators A(t), and the Schrödinger picture is given
by the unitary transformation operator exp(iHt),

 states :
 |ψ0 � ≡ eiHt |ψ(t)�,
The Heisenberg picture operators : A(t) ≡ eiHt A e−iHt . (5.5)


H does not depend on time.

As before the original operator A may be time dependent. The important equation of
motion governing the time evolution of A(t) is easily established. Since H is time inde-
pendent, the total time derivative of A in the Heisenberg picture is denoted with a dot,
Ȧ, while the explicit time derivative of the original Schrödinger operator is denoted ∂t A:
� � � �
Ȧ(t) = eiHt iHA − iAH + ∂t A e−iHt ⇒ Ȧ(t) = i H, A(t) + (∂t A)(t), (5.6)

where X(t) always means eiHt Xe−iHt for any symbol X, in particular for X = ∂t A. In
this way an explicit time-dependence of A is taken into account. Note how carefully the
order of the operators is kept during the calculation.
Both the Schrödinger and the Heisenberg picture require a time-independent Hamil-
tonian. In the general case of time-dependent Hamiltonians, we have to switch to the
interaction picture.

5.3 The interaction picture


The third and last representation, the interaction picture, is introduced to deal with the
situation where a system described by a time-independent Hamiltonian H0 , with known
energy eigenstates |n0 �, is perturbed by some, possibly time-dependent, interaction V (t),

H = H0 + V (t), with H0 |n0 � = εn0 |n0 �. (5.7)


5.3. THE INTERACTION PICTURE 89

The key idea behind the interaction picture is to separate the trivial time evolution due
to H0 from the intricate one due to V (t). This is obtained by using only H0 , not the full
H, in the unitary transformation Eq. (5.5). As a result, in the interaction picture both
the state vectors |ψ̂ (t)� and the operators Â(t) depend on time. The defining equations
for the interaction picture are

 states :
 |ψ̂ (t)� ≡ eiH0 t |ψ(t)�,
The interaction picture operators : Â(t) ≡ eiH0 t A e−iH0 t . (5.8)


H0 does not depend on time.

The interaction picture and the Heisenberg picture coincide when V = 0; i.e., in the non-
perturbed case. If V (t) is a weak perturbation, then one can think of Eq. (5.8) as a way
to pull out the fast, but trivial, time dependence due to H0 , leaving states that vary only
slowly in time due to V (t).
The first hint of the usefulness of the interaction picture comes from calculating the
time derivative of |ψ̂ (t)� using the definition Eq. (5.8):
� � � �
i∂t |ψ̂ (t)� = i∂t eiH0 t |ψ(t)� + eiH0 t i∂t |ψ(t)� = eiH0 t (−H0 + H)|ψ(t)�, (5.9)

which by Eq. (5.8) is reduced to

i∂t |ψ̂ (t)� = V̂ (t) |ψ̂ (t)�. (5.10)

The resulting Schrödinger equation for |ψ̂ (t)� thus contains explicit reference only to the
interaction part V̂ (t) of the full Hamiltonian H. This means that in the interaction picture
the time evolution of a state |ψ̂ (t0 )� from time t0 to t must be given in terms of a unitary
operator Û (t, t0 ) which also only depends on V̂ (t). Û (t, t0 ) is completely determined by

|ψ̂ (t)� = Û (t, t0 ) |ψ̂ (t0 )�. (5.11)

When V and thus H are time-independent, an explicit form for Û (t, t0 ) is obtained by
inserting |ψ̂ (t)� = eiH0 t |ψ(t)� = eiH0 t e−iHt |ψ0 � and |ψ̂ (t0 )� = eiH0 t0 e−iHt0 |ψ0 � into
Eq. (5.11),

eiH0 t e−iHt |ψ0 � = Û (t, t0 ) eiH0 t0 e−iHt0 |ψ0 � Û (t, t0 ) = eiH0 t e−iH(t−t0 ) e−iH0 t0 .

(5.12)
From this we observe that Û −1 = Û † , i.e. Û is indeed a unitary operator.
In the general case with a time-dependent V̂ (t) we must rely on the differential equation
appearing when Eq. (5.11) is inserted in Eq. (5.10). We remark that Eq. (5.11) naturally
implies the boundary condition Û (t0 , t0 ) = 1, and we obtain:

i∂t Û (t, t0 ) = V̂ (t) Û (t, t0 ), Û (t0 , t0 ) = 1. (5.13)

By integration of this differential equation we get the integral equation



1 t �
Û (t, t0 ) = 1 + dt V̂ (t� ) Û (t� , t0 ), (5.14)
i t0
90 CHAPTER 5. TIME EVOLUTION PICTURES

which we can solve iteratively for Û (t, t0 ) starting from Û (t� , t0 ) = 1. The solution is
� � �
1 t
1 t t1
Û (t, t0 ) = 1 + dt1 V̂ (t1 ) + dt1 V̂ (t1 ) dt2 V̂ (t2 ) + . . . (5.15)
i t0 i2 t0 t0

Note that in the iteration the ordering of all operators is carefully kept. A more compact
form is obtained by the following rewriting. Consider for example the second order term,
paying special attention to the dummy variables t1 and t2 :
� t � t1
dt1 V̂ (t1 ) dt2 V̂ (t2 )
t0 t0
� � � �
1 t t1
1 t t2
= dt1 V̂ (t1 ) dt2 V̂ (t2 ) + dt2 V̂ (t2 ) dt1 V̂ (t1 )
2 t0 t0 2 t0 t0
� t � � �
1 t
1 t t
= dt1 dt2 V̂ (t1 )V̂ (t2 )θ(t1 − t2 ) + dt2 dt1 V̂ (t2 )V̂ (t1 )θ(t2 − t1 )
2 t0 t0 2 t0 t0
� t � t � �
1
= dt1 dt2 V̂ (t1 )V̂ (t2 )θ(t1 − t2 ) + V̂ (t2 )V̂ (t1 )θ(t2 − t1 )
2 t0 t0
� �
1 t t
≡ dt1 dt2 Tt [V̂ (t1 )V̂ (t2 )], (5.16)
2 t0 t0

where we have introduced the time ordering operator Tt . Time ordering is easily general-
ized to higher order terms. In n-th order, where n factors V̂ (tj ) appear, all n! permutations
p ∈ Sn of the n times tj are involved, and we define1

Tt [V̂ (t1 )V̂ (t2 ) . . . V̂ (tn )] ≡ V̂ (tp(1) )V̂ (tp(2) ) . . . V̂ (tp(n) ) × (5.17)
p∈Sn
θ(tp(1) − tp(2) ) θ(tp(2) − tp(3) ) . . . θ(tp(n−1) − tp(n) ).

Using the time ordering operator, we obtain the final compact form (see also Exercise 5.2):

� � t
1 � 1 �n t � � � �

� t
−i � V̂ (t� )
Û (t, t0 ) = dt1 . . . dtn Tt V̂ (t1 ) . . . V̂ (tn ) = Tt e t0 dt . (5.18)
n! i t0 t0
n=0

Note the similarity with a usual time evolution factor e−iε t . This expression for Û (t, t0 ) is
the starting point for infinite order perturbation theory and for introducing the concept
of Feynman diagrams; it is therefore one of the central equations in quantum field theory.
A graphical sketch of the contents of the formula is given in Fig. 5.1.

1
For n = 3 we have Tt [V̂ (t1 )V̂ (t2 )V̂ (t3 )] =
V̂ (t1 )V̂ (t2 )V̂ (t3 )θ(t1 −t2 )θ(t2 −t3 )+V̂ (t1 )V̂ (t3 )V̂ (t2 )θ(t1 −t3 )θ(t3 −t2 )+V̂ (t2 )V̂ (t3 )V̂ (t1 )θ(t2 −t3 )θ(t3 −t1 )+
V̂ (t2 )V̂ (t1 )V̂ (t3 )θ(t2 −t1 )θ(t1 −t3 )+ V̂ (t3 )V̂ (t1 )V̂ (t2 )θ(t3 −t1 )θ(t1 −t2 )+ V̂ (t3 )V̂ (t2 )V̂ (t1 )θ(t3 −t2 )θ(t2 −t1 ).
5.4. TIME-EVOLUTION IN LINEAR RESPONSE 91

Figure 5.1: The time evolution operator Û (t, t0 ) can be viewed as the sum of additional
phase factors due to V̂ on top of the trivial phase factors arising from H0 . The sum
contains contributions from processes with 0, 1, 2, 3, . . . scattering events V̂ , which happen
during the evolution from time t0 to time t.

5.4 Time-evolution in linear response


In many applications the perturbation V̂ (t) is weak compared to H0 . It can therefore be
justified to approximate Û (t, t0 ) by the first order approximation

1 t
Û (t, t0 ) ≈ 1 + dt� V̂ (t� ). (5.19)
i t0

This simple time evolution operator forms the basis for the Kubo formula in linear response
theory, which, as we shall see in the following chapters, is applicable to a wide range of
physical problems.

5.5 Time dependent creation and annihilation operators


It is of fundamental interest to study how the basic creation and annihilation operators
a†ν and aν evolve in time given some set of basis states {|ν�} for a time-independent
Hamiltonian H. As in Sec. 1.3.4 these operators can be taken to be either bosonic or
fermionic. Let us first apply the definition of the Heisenberg picture, Eq. (5.5):

a†ν (t) ≡ eiHt a†ν e−iHt , (5.20a)


−iHt
aν (t) ≡ e iHt
aν e . (5.20b)

In the case of a general time-independent Hamiltonian with complicated interaction terms,


the commutators [H, a†ν ] and [H, aν ] are not simple, and consequently the fundamental
(anti-)commutator [aν (t1 ), a†ν (t2 )]F,B involving two different times t1 and t2 cannot be
given in a simple closed form:

[aν1 (t1 ), a†ν2 (t2 )]F,B =


(5.21)
eiHt1 aν1 e−iH(t1 −t2 ) a†ν2 e−iHt2 ± eiHt2 a†ν2 e−iH(t2 −t1 ) aν1 e−iHt1 = ??
92 CHAPTER 5. TIME EVOLUTION PICTURES

No further reduction is possible in the general case. In fact, as we shall see in the following
chapters, calculating (anti-)commutators like Eq. (5.21) is the problem in many-particle
physics.
But let us investigate some simple cases to get a grasp of the time evolution pictures.
Consider first a time-independent Hamiltonian H which is diagonal in the |ν�-basis,

H= εν a†ν aν . (5.22)
ν

The equation of motion, Eq. (5.6), is straightforward:2

ȧν (t) = i[H, aν (t)] = ieiHt [H, aν ]e−iHt


� � † � � � �
= ieiHt εν � aν � aν � , aν e−iHt = ieiHt εν � −δν,ν � aν � e−iHt
ν� ν�
−iHt
= −iεν e iHt
aν e = −iεν aν (t). (5.23)

By integration we obtain
aν (t) = e−iεν t aν , (5.24)
which by Hermitian conjugation leads to

a†ν (t) = e+iεν t a†ν . (5.25)

In this very simple case the basic (anti-)commutator Eq. (5.21) can be evaluated directly:

[aν1 (t1 ), a†ν2 (t2 )]F,B = e−iεν1 (t1 −t2 ) δν1 ,ν2 . (5.26)

For the diagonal Hamiltonian the time evolution is thus seen to be given by trivial phase
factors e±iε t .
We can also gain some insight into the interaction picture by a trivial extension of the
simple model. Assume that

H = H0 + γH0 , γ � 1, (5.27)

where H0 is diagonalized in the basis {|ν�} with the eigenenergies εν . Obviously, the full
Hamiltonian H is also diagonalized in the same basis, but with the eigenenergies (1 + γ)ε.
Let us however try to treat γH0 as a perturbation V to H0 , and then use the interaction
picture of Sec. 5.3. From Eq. (5.8) we then obtain

|ν̂(t)� = eiεν t |ν(t)�. (5.28)

But we actually know the time evolution of the Schrödinger state on the right-hand side
of the equation, so

|ν̂(t)� = eiεν t e−i(1+γ)εν t |ν� = e−iγεν t |ν�. (5.29)


2
We are using the identities [AB, C] = A[B, C] + [A, C]B and [AB, C] = A{B, C} − {A, C}B, which
are valid for any set of operators. Note that the first identity is particularly useful for bosonic operators
and the second for fermionic operators (see Exercise 5.4).
5.6. SUMMARY AND OUTLOOK 93

Here we clearly see that the fast Schrödinger time dependence given by the phase factor
eiεν t , is replaced in the interaction picture by the slow phase factor eiγεν t . The reader can
try to obtain Eq. (5.29) directly from Eq. (5.18).
Finally, we briefly point to the complications that arise when the interaction is given
by a time-independent operator V not diagonal in the same basis as H0 . Consider for
example the Coulomb-like interaction written symbolically as
� ��
H = H0 + V = εν � a†ν � aν � + Vq a†ν1 +q a†ν2 −q aν2 aν1 . (5.30)
ν� ν1 ν2 q

The equation of motion for aν (t) is:


� � �
ȧν (t) = i[H, aν (t)] = −iεν aν (t) + i Vq a†ν1 +q (t) a†ν2 −q (t), aν (t) aν2 (t) aν1 (t)
ν1 ν2 q

= −iεν aν (t) + i (Vν2 −ν − Vν−ν1 )a†ν1 +ν2 −ν (t) aν2 (t) aν1 (t). (5.31)
ν1 ν2

The problem in this more general case is evident. The equation of motion for the single
operator aν (t) contains terms with both one and three operators, and we do not know
the time evolution of the three-operator product a†ν1 +ν2 −ν (t) aν2 (t) aν1 (t). If we write
down the equation of motion for this three-operator product we discover that terms are
generated involving five operator products. This feature is then repeated over and over
again generating a never-ending sequence of products containing seven, nine, eleven, etc.
operators. In the following chapters we will learn various approximate methods to deal
with this problem.

5.6 Summary and outlook


In this chapter we have introduced the fundamental representations used in the descrip-
tion of time evolution in many-particle systems: the Schrödinger picture, Eq. (5.2), the
Heisenberg picture, Eq. (5.5), and the interaction picture, Eq. (5.8). The first two pictures
rely on a time-independent Hamiltonian H, while the interaction picture involves a time-
dependent Hamiltonian H of the form H = H0 + V (t), where H0 is a time-independent
Hamiltonian with known eigenstates. Which picture to use depends on the problem at
hand.
We have derived an explicit expression, Eq. (5.18), for the time evolution operator
Û (t, t0 ) describing the evolution of an interaction picture state |ψ̂ (t0 )� at time t0 to |ψ̂ (t)�
at time t. We shall see in the following chapters how the operator Û (t, t0 ) plays an
important role in the formulation of infinite order perturbation theory and the introduction
of Feynman diagrams, and how its linearized form Eq. (5.19) forms the basis of the widely
used linear response theory and the associated Kubo formalism.
Finally, by studying the basic creation and annihilation operators we have gotten a
first glimpse of the problems we are facing, when we are trying to study the full time
dependence, or equivalently the full dynamics, of interacting many-particle systems.
94 CHAPTER 5. TIME EVOLUTION PICTURES
Chapter 6

Linear response theory

Linear response theory is an extremely widely used concept in all branches of physics.
It simply states that the response to a weak external perturbation is proportional to the
perturbation, and therefore all one needs to understand is the proportionality constant.
Below we derive the general formula for the linear response of a quantum system exerted
by a perturbation. The physical question we ask is thus: supposing some perturbation H � ,
what is the measured consequence for an observable quantity, A. In other words, what is
�A� to linear order in H � ?
Among the numerous applications of the linear response formula, one can mention
charge and spin susceptibilities of e.g. electron systems due to external electric or magnetic
fields. Responses to external mechanical forces or vibrations can also be calculated using
the very same formula. Here we utilize the formalism to derive a general expression for
the electrical conductivity and briefly mention other applications.

6.1 The general Kubo formula


Consider a quantum system described by the (time independent) Hamiltonian H0 in ther-
modynamic equilibrium. According to Sec. 1.5 this means that an expectation value of a
physical quantity, described by the operator A, can be evaluated as
1 1 �
�A� = Tr [ρ0 A] = �n|A|n�e−βEn , (6.1a)
Z0 Z0 n

ρ0 = e−βH0 = |n��n|e−βEn , (6.1b)
n

where ρ0 is the density operator and Z0 =Tr[ρ0 ] is the partition function. Here as in
Sec. 1.5, we write the density operator in terms of a complete set of eigenstates, {|n�}, of
the Hamiltonian, H0 , with eigenenergies {En }.
Suppose now that at some time, t = t0 , an external perturbation is applied to the
system, driving it out of equilibrium. The perturbation is described by an additional time
dependent term in the Hamiltonian
H(t) = H0 + H � (t)θ(t − t0 ). (6.2)

95
96 CHAPTER 6. LINEAR RESPONSE THEORY

Figure 6.1: Illustration of the linear response theory. At times before t0 the system is in
equilibrium, after which the perturbation is turned on. The system is now evolving accord-
ing to the new Hamiltonian and is in a non-equilibrium state. The Kubo formula relates
the expectation value δ�A�non−eq in the non-equilibrium state to a equilibrium expectation
value �· · · �eq of the more complicated time-dependent commutator [Â(t), Ĥ � (t� )].

We emphasize that H0 is the Hamiltonian describing the system before the perturbation
was applied, see Fig. 6.1 for an illustration. Now we wish to find the expectation value
of the operator A at times t greater than t0 . In order to do so we must find the time
evolution of the density matrix or equivalently the time evolution of the eigenstates of the
unperturbed Hamiltonian. Once we know the |n(t)�, we can obtain �A(t)� as

1 � 1
�A(t)� = �n(t)|A|n(t)�e−βEn = Tr [ρ(t)A] , (6.3a)
Z0 n Z0

ρ(t) = |n(t)��n(t)|e−βEn . (6.3b)
n

The philosophy behind this expression is as follows. The initial states of the system
are distributed according to the usual Boltzmann distribution e−βE0n /Z0 . At later times
the system is described by the same distribution of states but the states are now time-
dependent and they have evolved according to the new Hamiltonian. The time dependence
of the states |n(t)� is of course governed by the Schrödinger equation

i∂t |n(t)� = H(t)|n(t)�. (6.4)

Since H � is to be regarded as a small perturbation, it is convenient to utilize the inter-


action picture representation |n̂(t)� introduced in Sec. 5.3. The time dependence in this
representation is given by

|n(t)� = e−iH0 t |n̂(t)� = e−iH0 t Û (t, t0 )|n̂(t0 )�, (6.5)

where by definition |n̂(t0 )� = eiH0 t0 |n(t0 )� = |n�.


6.1. THE GENERAL KUBO FORMULA 97

�t
To linear order in H � , Eq. (5.19) states that Û (t, t0 ) = 1 − i t0 dt� Ĥ � (t� ). Inserting this
into (6.3a), one obtains the expectation value of A up to linear order in the perturbation
� t
1 � −βEn
�A(t)� = �A�0 − i dt� e �n(t0 )|Â(t)Ĥ � (t� ) − Ĥ � (t� )Â(t)|n(t0 )�
t0 Z0 n
� t
= �A�0 − i dt� �[Â(t), Ĥ � (t� )]�0 . (6.6)
t0

The brackets ��0 mean an equilibrium average with respect to the Hamiltonian H0 . This
is in fact a remarkable and very useful result, because the inherently non-equilibrium
quantity �A(t)� has been expressed as a correlation function of the system in equilibrium.
The physical reason for this is that the interaction between excitations created in the
non-equilibrium state is an effect to second order in the weak perturbation, and hence not
included in linear response.
The correlation function that appears in Eq. (6.6), is called a retarded correlation
function, and for later reference we rewrite the linear response result as
� ∞
� −η(t−t� )
δ�A(t)� ≡ �A(t)� − �A�0 = dt� CAH
R
� (t, t )e , (6.7)
t0

where �� ��
� �
R
CAH � (t, t ) = −iθ(t − t ) Â(t), Ĥ � (t� ) . (6.8)
0

This is the famous Kubo formula which expresses the linear response to a perturbation,

H �.We have added a very important detail here: the factor e−η(t−t ) , with an infinitesimal
positive parameter η, has been included to force the response at time t due to the influence
of H � at time t� to decay when t � t� . In the end of a calculation we must therefore take
the limit η → 0+ . For physical reasons the (retarded) effect of a perturbation must of
course decrease in time. You can think of the situation that one often has for differential
equations with two solutions: one which increases exponentially with time (physically

not acceptable) and one which decreases exponentially with time; the factor e−η(t−t ) is
there to pick out the physically relevant solution by introducing an artificial relaxation
mechanism.

Kubo formula in the frequency domain


It is often convenient to express the response to an external disturbance in the frequency
domain. Let us therefore write the perturbation in terms of its Fourier components

dω −iωt �
H � (t) = e Hω , (6.9)

such that CAH


R
� becomes

� ∞
� dω −iωt� R
R
CAH � (t, t ) = e CAHω� (t − t� ), (6.10)
−∞ 2π
98 CHAPTER 6. LINEAR RESPONSE THEORY

because �[Â(t), Ĥω� (t� )]�0 only depends on the difference between t and t� , which can easily
be proven using the definition of the expectation value. When inserted into the Kubo
formula, one gets (after setting t0 = −∞, because we are not interested in the transient
behavior)
� ∞ � ∞
dω −iωt −i(ω+iη)(t� −t) R
δ�A(t)� = dt� e e CAHω� (t − t� )

�−∞∞
−∞
dω −iωt R
= e CAHω� (ω), (6.11)
−∞ 2π

and therefore the final result reads in frequency domain

δ�Aω � = CAHR
� (ω), (6.12a)
� ∞ω
R
CAHω
� (ω) = dteiωt e−ηt CAH
R
ω
� (t). (6.12b)
−∞

Note again that the infinitesimal η is incorporated in order to ensure the correct physical
result, namely that the retarded response function decays at large times.

6.2 Kubo formula for conductivity


Consider a system of charged particles, electrons say, which is subjected to an external
electromagnetic field. The electromagnetic field induces a current, and the conductivity
is the linear response coefficient. In the general case the conductivity may be non-local in
both time and space, such that the electric current Je at some point r at time t depends
on the electric field at points r� at times t�
� � �
Jeα (r, t) = dt� dr� σ αβ (rt, r� t� ) E β (r� , t� ) (6.13)
β

where σ αβ (r, r� ; t, t� ) is the conductivity tensor which describes the current response in
direction êα to an applied electric field in direction êβ .
The electric field E is given by the electric potential φext and the vector potential Aext

E(r, t) = −∇r φext (r, t) − ∂t Aext (r, t). (6.14)

The current density operator of charged particles in the presence of an electromagnetic


field was given in Chap. 1. For simplicity we assume only one kind of particles, electrons
say, but generalization to more kinds of charge carrying particles is straightforward by
simple addition of more current components.1 For electrons Je = −e�J�. The perturbing
term in the Hamiltonian due to the external electromagnetic field is given by the coupling
1
With more carriers the operator for the electrical current becomes Je (r) = i qi Ji (r), where qi are
the charges of the different carriers. Note that in this case the currents of the individual species are not
necessarily independent.
6.2. KUBO FORMULA FOR CONDUCTIVITY 99

of the electrons to the scalar potential and the vector potential. To linear order in the
external potential
� �
Hext = −e dr ρ(r)φext (r, t) + e dr J(r) · Aext (r, t), (6.15)

where the latter term was explained in Sec. 1.4.3. Let A0 denote the vector potential
in the equilibrium, i.e. prior to the onset of the perturbation Aext , and let A denote the
total vector potential. Then we have

A = A0 +Aext , (6.16)

Again according to Sec. 1.4.3, the current operator has two components, the diamagnetic
term and the paramagnetic term
e
J(r) = J∇ (r) + A(r)ρ(r), (6.17)
m
In order to simplify the expressions, we can choose a gauge where the external electrical
potential is zero, φext = 0. This is always possible by a suitable choice of A(r, t) as you
can see in Eq. (6.14). The final result should of course not depend on the choice of
gauge. The conductivity is most easily expressed in the frequency domain, and therefore
we Fourier transform the perturbation. Since ∂t becomes −iω in the frequency domain we
have Aext (r, ω) = (1/iω)Eext (r, ω), and therefore the external perturbation in Eq. (6.15)
becomes in the Fourier domain

e
Hext,ω = dr J(r) · Eext (r, ω). (6.18)

In order to exploit the frequency domain formulation of linear response we want to write
the definition of the conductivity tensor in Eq. (6.13) in frequency domain. Because we are
only considering linear response the conductivity tensor is a property of the equilibrium
system and can thus only depend on time differences σ αβ (rt, r� t� ) = σ αβ (r, r� , t − t� ). The
frequency transform of Eq. (6.13) is therefore simply that of a convolution and hence
� �
Jeα (r, ω) = dr� σ αβ (r, r� , ω) E β (r� , ω). (6.19)
β

Now since Eq. (6.18) is already linear in the external potential Eext and since we are
only interested in the linear response, we can replace J in Eq. (6.18) by J0 = J∇ + m e
A0 ρ,
thus neglecting the term proportional to Eext · Aext . Eq. (6.18) is therefore replaced by

e
Hext,ω = dr J0 (r) · Eext (r, ω). (6.20)

To find the expectation value of the current we write
e
�J(r, ω)� = �J0 (r, ω)� + � Aext (r, ω)ρ(r)�. (6.21)
m
100 CHAPTER 6. LINEAR RESPONSE THEORY

For the last term in Eq. (6.21) we use that to linear order in Aext the expectation value
can be evaluated in the equilibrium state
e e e
� Aext (r, ω)ρ(r)� = Aext (r, ω)�ρ(r)�0 = Eext (r, ω)�ρ(r)�0 . (6.22)
m m iω
For the first term in Eq. (6.21) we use the general Kubo formula in Eq. (6.7). Since the
equilibrium state does not carry any current, i.e. �J0 �0 = 0, we conclude that �J0 � = δ�J0 �.
In frequency domain we should use the results Eq. (6.12a) and substitute J0 (r) for the
operator “A”, and Hext,ω for “Hω� ”, which leads to �J0 (r, ω)� = CRJ0 (r)Hext,ω (ω). Collecting
things we now have
e
�J(r, ω)� = CR
J0 (r)Hext,ω (ω) + �ρ(r)�0 Aext (r,ω). (6.23)
m
Writing out the first term
� � e β �
J0 (r)Hext,ω (ω)
CR = dr� CR
J (r)J β (r� )
(ω) E (r , ω). (6.24)
0 0 iω
β

Comparing with the definition of the non-local conductivity in Eq. (6.19), we can now col-
lect the two contributions to the conductivity tensor. The first term comes from Eq. (6.24)
and it is seen to of the same form as (6.12a), in particular the response is non-local in
space. In contrast, the second term in Eq. (6.22) stemming from the diamagnetic part
of the current operator is local in space. Now collecting the two terms and using that
Je = −e�J�, we finally arrive at the linear response formula for the conductivity tensor

ie2 R ie2 n(r)


σ αβ (r, r� , ω) = Παβ (r, r� , ω) + δ(r − r� )δαβ , (6.25)
ω iωm
where we have used the symbol ΠR = CJR0 J0 for the retarded current-current correlation
function. In the time domain it is given by
�� ��
ΠR � � � �
αβ (r, r , t − t ) = CJ α (r)J β (r� ) (t − t ) = −iθ(t − t )
R
Jˆ0α (r, t), Jˆ0β (r� , t� ) . (6.26)
0 0 0

Finding the conductivity of a given system has thus been reduced to finding the
retarded current-current correlation function. This formula will be used extensively in
Chap. 14.

6.3 Kubo formula for conductance


The conductivity σ is the proportionality coefficient between the electric field E and the
current density J, and it is an intrinsic property of a material. The conductance on the
other hand is the proportionality coefficient between the current I through a sample and
the voltage V applied to it, i.e. a sample specific quantity. The conductance G is defined
by the usual Ohm’s law
I = GV. (6.27)

You might also like