You are on page 1of 34

Andrei Tokmakoff, MIT Department of Chemistry, 3/8/2007 4-1

4. INTERACTION OF LIGHT WITH MATTER

One of the most important topics in time-dependent quantum mechanics for chemists is the
description of spectroscopy, which refers to the study of matter through its interaction with light
fields (electromagnetic radiation). Classically, light-matter interactions are a result of an
oscillating electromagnetic field resonantly interacting with charged particles. Quantum
mechanically, light fields will act to couple quantum states of the matter, as we have discussed
earlier.

Like every other problem, our starting point is to derive a Hamiltonian for the light-matter
interaction, which in the most general sense would be of the form

H = H M + H L + H LM . (4.1)

The Hamiltonian for the matter H M is generally (although not necessarily) time independent,

whereas the electromagnetic field H L and its interaction with the matter H LM are time-dependent.
A quantum mechanical treatment of the light would describe the light in terms of photons for
different modes of electromagnetic radiation, which we will describe later. However, we will start
with a common semiclassical treatment of the problem. For this approach we treat the matter
quantum mechanically, and treat the field classically. For the field we assume that he light only
presents a time-dependent interaction potential that acts on the matter, but the matter doesn’t
influence the light. (Quantum mechanical energy conservation says that we expect that the change
in the matter to raise the quantum state of the system and annihilate a photon from the field. We
won’t deal with this right now). In that case, we can really largely ignore H L , and we have a

Hamiltonian that can be solved in the interaction picture representation:

H ! H M + H LM t() (4.2)
()
= H0 + V t

Here, we’ll derive the Hamiltonian for the light-matter interaction, the Electric Dipole
Hamiltonian. It is obtained by developing a classical Hamiltonian for the interaction of charged
particles with an electromagnetic field, and then substituting quantum operators for the matter.
4-2

p ! "i!#̂
(4.3)
x ! x̂

In order to get the classical Hamiltonian, we need to work through two steps: (1) We need
to describe electromagnetic fields, specifically in terms of a vector potential, and (2) we need to
describe how the electromagnetic field interacts with charged particles.

Brief summary of electrodynamics

Let’s summarize the description of electromagnetic fields that we will use. A derivation of the
plane wave solutions to the electric and magnetic fields and vector potential is described in the
appendix. Also, it is helpful to review this material in Jackson1 or Cohen-Tannoudji, et al.2

(
> Maxwell’s Equations describe electric and magnetic fields E, B . )
> To construct a Hamiltonian, we must describe the time-dependent interaction potential (rather
than a field).

( )
> To construct a potential representation of E and B , you need a vector potential A r ,t and a

( )
scalar potential ! r ,t . For electrostatics we normally think of the field being related to the

electrostatic potential through E = !"# , but for a field that varied in time and in space, the

electrodynamic potential must be expressed in terms of both A and ! .


> In general an electromagnetic wave written in terms of the electric and magnetic fields requires
6 variables (the x,y, and z components of E and B). This is an overdetermined problem;
Maxwell’s equations constrain these. The potential representation has four variables
( Ax , Ay , Az and ! ), but these are still not uniquely determined. We choose a constraint – a

representation or guage – that allows us to uniquely describe the wave. Choosing a gauge such
that ! = 0 (Coulomb gauge) leads to a plane-wave description of E and B :

1 # A r ,t
2
( )
2
( )
!" A r ,t + 2
c #t 2
=0 (4.4)

!" A = 0 (4.5)
4-3

This wave equation allows the vector potential to be written as a set of plane waves:

( ) + A*!ˆ e#i( k "r #$ t ) .


( )
A r ,t = A0!ˆ e
i k "r # $ t
0
(4.6)

This describes the wave oscillating in time at an angular frequency ω and propagating in space

in the direction along the wavevector k , with a spatial period ! = 2" k . The wave has an

amplitude A0 which is directed along the polarization unit vector !ˆ . Since ! " A = 0 , we see

that k ! "ˆ = 0 or k ! "ˆ . From the vector potential we can obtain E and B

"A
E=!
"t (4.7)
i ( k %r ! # t ) !i ( k %r ! # t ) (
= i# A0 $̂ & e !e
' )

B = !" A
( ) % e%i( k $r %& t ) ) (4.8)
( ()
= i k " #̂ A0 ' e
i k $r % & t
*

If we define a unit vector along the magnetic field polarization as b̂ = k ! "ˆ ( ) k = k̂ ! "ˆ , we

see that the wavevector, the electric field polarization


and magnetic field polarization are mutually
orthogonal k̂ ! "ˆ ! b̂ .
Also, by comparing eq. (4.6) and (4.7) we see that
the vector potential oscillates as cos ωt, whereas the
field oscillates as sin ωt. If we define

1
E = i! A0 (4.9)
2 0

1
B = i k A0 (4.10)
2 0
then,

( )
E r ,t = E0 !ˆ sin k " r # $ t ( ) (4.11)
4-4

( ) ( )
B r ,t = B0 b̂sin k ! r " # t . (4.12)

Note, E0 B0 = ! k = c .
4-5

Hamiltonian for radiation field interacting with charged particle

Now, let’s find a classical Hamiltonian that describes charged particles in field in terms of the
vector potential. Start with Lorentz force on a particle with charge q

(1) (
F = q E+v ! B . ) (4.13)

Here v is the velocity of the particle. Writing this for one direction (x) in terms of the Cartesian
components of E , v and B , we have:

(2) (
Fx = q Ex + v y Bz ! vz By . ) (4.14)

In Lagrangian mechanics, this force can be expressed in terms of the total potential energy U as
"U d # "U &
(3) Fx = ! + (4.15)
"x dt %$ "vx ('

Using the relationships that describe E and B in terms of A and ! , inserting into eq. (4.14) and
working it into the form of eq. (4.15), we can show that:

(4) U = q! " qv # A (4.16)


This is derived in CTDL,3 and you can confirm by plugging it into eq. (4.15).
Now we can write a Lagrangian in terms of the kinetic and potential energy of the particle

L = T !U (4.17)

1
(5) L= mv 2 + qv ! A " q# (4.18)
2
The classical Hamiltonian is related to the Lagrangian as
H = p!v " L
(6) (4.19)
= p!v " 12 mv 2 " q v ! A " q#

!L
Recognizing p= = mv + qA (4.20)
!v

we write v=m (
1 p ! qA .
) (4.21)

Now substituting (4.21) into (4.19), we have:


4-6

( ) ( p " qA) ( p " qA)! A + q#


2 q
1 1
H= m p! p " qA " 2m " m (4.22)

1
( ) ( )
2
H= " p ! qA r ,t $ + q& r ,t (4.23)
2m # %

This is the classical Hamiltonian for a particle in an electromagnetic field. In the Coulomb

( )
gauge ! = 0 , the last term is dropped.

We can write a Hamiltonian for a collection of particles in the absence of a external field

! pi2 $
H0 = ' # + V0 ri & . ( ) (4.24)
i " 2mi %
and in the presence of the EM field:
" 1 %
( ( )) ( )
2
H = ($ pi ! qi A ri + V0 ri ' . (4.25)
i # 2mi &

qi qi
H = H0 ! # ( ) # 2m
2
Expanding: p " A+ A" pi + A (4.26)
i 2mi i i i

Generally the last term is considered small. That is the energy of particles high relative to
amplitude of potential. (This term should be considered for extremely high field strength, which
significantly distorts the potential binding molecules together). For weak fields we have:

H = H0 + V t () (4.27)

qi
()
V t =" (
p ! A + A! pi
2mi i
) (4.28)
i

Now we are in a position to substitute the quantum mechanical momentum for the classical. Here
the vector potential remains classical, and only modulates the interaction strength.

p = ! i!" (4.29)

()
V t =#
i
i!
(
q ! " A + A"! i
2mi i i
) (4.30)
4-7

( )
We can show that !" A = A"! . Notice !" A = !" A + A"! (chain rule). For instance if we are
operating on a wavefunction !" A # = ( !" A ) # + A"( ! # ) , but we are working in the

( )
Coulomb gauge !" A = 0 . Now we have:

i!qi
()
V t =#
mi
A!" i
i
(4.31)
q
= $ # i A! pi
i mi

For a single charge particle our interaction Hamiltonian is


q
()
V t =!
m
A" p
(4.32)
q i ( k "r ! $ t )
= ! %' A0#̂ " p e + c.c.(*
m& )

Electric Dipole Approximation

Under most circumstances, we can neglect the wavevector dependence of the interaction potential.

If the wavelength of the field is much larger than the molecular dimension ! " # ( ) ( k " 0) ,
then eik !r "1 .
We do retain the spatial dependence for certain types of light matter interactions. In that
case we define r0 as the center of mass of a molecule and expand

eik !ri = eik !r0 e ( i 0 )


ik ! r "r

(4.33)
= e 0 #$1 + ik !( ri " r0 ) + … %&
ik !r

For interactions, with UV, visible, and infrared (but not X-ray) radiation, k ri ! r0 <<1 , and setting

r0 = 0 means that eik !r " 1 . We retain the second term for quadrupole transitions: charge

distribution interacting with gradient of electric field and magnetic dipole.


4-8

4.2 Electric Dipole Hamiltonian


Now, using A0 = iE0 2! ,

!iqE0
()
V t =
2m" &
%#̂ $ p e!i" t ! #̂ $ p e+i" t '
( (4.34)

!qE0
()
V t =
m"
( )
#̂ $ p sin " t
(4.35)
!q
=
m"
(
E(t)$ p )
or for a collection of charged particles (molecules):
% q (E
() ( )
V t = ! ' $ i "ˆ # pi * 0 sin + t
& i mi )+
(4.36)

This is known as the electric dipole Hamiltonian (EDH).

Harmonic Perturbation: Matrix Elements


To better interpret this, let’s evaluate the transition rates induced by V(t). For a perturbation

()
V t = V0 sin ! t the rate of transitions induced by field is

!
( ) ( )
2
wk! = Vk! %&" Ek # E! # "$ + " Ek # E! + "$ '( (4.37)
2"
Let’s look at the matrix elements for the EDH.

!qE0
Vk! = k V0 ! = k #̂ $ p ! (4.38)
m"

We can evaluate the matrix element k p ! using

i!p
!" r , H 0 #$ = (4.39)
m
which gives
4-9

m
k p! = k r H0 ! H0 r !
i"
=
m
i"
( k r ! E! ! Ek k r ! ) (4.40)

= im" !k k r !

So we have
" !k
Vk! = ! iqE0 k #̂ $r ! (4.41)
"
or for a collection of particles

" !k & )
Vk! = ! iE0 k #̂ $ ( % qi ri + !
" ' i *
" !k
= ! iE0 k #̂ $ µ ! (4.42)
"
"
= !iE0 !k µ kl
"

µ is the dipole operator. The dipole moment can be expressed more generally as the distribution
of charge in the molecule:

µ = ! dr r " r ( ) (4.43)

So ()
V t = ! µ " E(t) (4.44)

Then the rate of transitions between quantum states induced by the electric field is
! 2"
2

( ) ( )
2
wk! = E0 k!
µ %# Ek $ E! $ "" + Ek $ E! + "" '
2" " 2 kl & (
(4.45)
!
(" ) ( )
2 2
= 2 E0 µ kl %&# k!
$ " + # " k! + " '(
2"
This is an expression for the absorption spectrum since the rate of transitions can be related to
the power absorbed from the field. More generally we would express the absorption spectrum in
terms of a sum over all initial and final states, the eigenstates of H0:

!
( ) ( )
2
w fi = "
2
E0 µ fi &# $ fi % $ + # $ fi + $ ( (4.46)
i, f ! 2 ' )
4-10

The strength of interaction between light and matter is given by the matrix

element µ fi ! f µ " #̂ i . The scalar part f µi says that you need a change of charge

distribution between f and i to get effective absorption. This matrix element is the basis of

selection rules based on the symmetry of the states. The vector part says that the light field must
project onto the dipole moment. This allows information to be obtained on the orientation of
molecules, and forms the basis of rotational transitions.
4-11

Relaxation Leads to Line-broadening

Let’s combine the results from the last two lectures, and describe absorption to a state that is
coupled to a continuum. What happens to the probability of absorption if the excited state decays
exponentially?

k relaxes exponentially
... for instance by coupling to continuum

Pk !exp #$ "wnk t %&

We can start with the first-order expression:


!i t
bk =
! #t0
d" k V t (4.47)

! i i# t
or equivalently
!t
bk = " e k" Vk" t
!
() (4.48)

We can add irreversible relaxation to the description of bk , following our early approach:

! i i# t w
!t
bk = " e k" Vk" t " nk bk
! 2
() (4.49)

()
Or using V t = !iE0 µ k! sin " t

! "i i# t w
!t
bk = e k" sin # t Vk" " nk bk t
! 2
()
(4.50)
E0 # k" $ i(# k" +# ) i(# k" "# )t ' w
=
2i!# % & e "e )
(
µ k" " nk bk t
2
()

The solution to the differential equation


y! + ay = b ei! t (4.51)

bei" t
is ()
y t = Ae ! at
+
a + i"
. (4.52)
4-12

E µ # ei(" k! +" )t ei(" k! !" )t &


()
bk t = A e
! wnk t / 2
+ 0 k! % ! ( (4.53)
2i" (
%$ wnk / 2 + i " k! + " )
wnk / 2 + i " k! ! " (' ( )
Let’s look at absorption only, in the long time limit:
# e ( k! ) &
i ! "! t
E0 µ k!
()
bk t = % ( (4.54)
2" %$ ! k! " ! " iwnk / 2 ('

For which the probability of transition to k is


2
2 E02 µ k! 1
Pk = bk = (4.55)
(! )
2 2
4" "! + wnk / 4
2
k!

The frequency dependence of the transition probability has a Lorentzian form:

The linewidth is related to the relaxation rate from k into the continuum n. Also the linewidth is
related to the system rather than the manner in which we introduced the perturbation.

Andrei Tokmakoff, MIT Department of Chemistry,


3/8/2007

Appendix: Review of Free Electromagnetic Field

Maxwell’s Equations (SI):

(1) !" B = 0

(2) "# E = ! / $0

!B
(3) "# E = $
!t
4-13

!E
(4) "# B = µ0 J +$0 µ0
!t

E : electric field; B : magnetic field; J : current density; ! : charge density; !0 :


electrical permittivity; µ0 : magnetic permittivity

We are interested in describing E and B in terms of a scalar and vector potential. This is
required for our interaction Hamiltonian.

Generally: A vector field F assigns a vector to each point in space. The divergence of the
field

!Fx !Fy !Fz


(5) "# F = + +
!x !y !z

is a scalar. For a scalar field ! , the gradient

"! "! "!


(6) #! = xˆ + yˆ + zˆ
"x "y "z

is a vector for the rate of change at on point in space. Here xˆ 2 + yˆ 2 + zˆ 2 = rˆ 2 are unit
vectors.

Also, the curl

xˆ yˆ zˆ
(7) "# F = ! ! !
!x !y !z
Fx Fy Fz

is a vector whose x, y, and z components are the circulation of the field about that
component.

Some useful identities from vector calculus are:

(8) ! " (!# F ) = 0


4-14

(9) "# ("! ) = 0

(10) !" (!" F ) = ! (!# F ) $ ! 2 F

We now introduce a vector potential A (r , t ) and a scalar potential ! (r , t ) , which we will

relate to E and B

Since !" B = 0 and ! (!" A ) = 0 :

(11) B = !" A

Using (3), we have:

!A
"# E = $ "#
!t

or

! #A "
(12) $% & E + ' = 0
( #t )

From (9), we see that a scalar product exists with:

"A
(13) E+ = # $! (r , t )
"t
convention
or

"A
(14) E= # $!
"t

So we see that the potentials A and ! determine the fields B and E :


(15) B (r , t ) = !" A (r , t )

"
(16) E (r , t ) = # $! (r , t ) # A (r , t )
"t

We are interested in determining the wave equation for A and ! . Using (15) and
differentiating (16) and substituting into (4):
4-15

" $2 A $! #
(17) %& (%& A ) + '0 µ0 ( 2 + % ) = µ0 J
* $t $t +
Using (10):

" 2 $2 A # % $! &
(18) - '( A + )0 µ 0 2 .
+ ( + (* A + )0 µ0 , = µ0 J
1 $t 2 / $t 0

From (14), we have:

"# $ A
#$ E = % % # 2!
"t

and using (2):

#$V % A
(19) # & 2! = " / '0
$t

Notice from (15) and (16) that we only need to specify four field components
( Ax , Ay , Az , ! ) to determine all six E and B components. But E and B do not uniquely

determine A and ! . So, we can construct A and ! in any number of ways without

changing E and B . Notice that if we change A by adding "! where ! is any function

of r and t , this won’t change B (!" (! # B ) = 0 ). It will change E by (# ""t $! ), but we

#!
can change ! to " $ = " % . Then E and B will both be unchanged. This property of
#t
changing representation (gauge) without changing E and B is gauge invariance. We can
transform between gauges with:

(20) A" (r , t ) = A (r , t )+ # $ ! (r , t )
gauge
transformation
#
(21) ! $ (r , t ) = ! (r , t ) % " (r , t )
#t

Up to this point, A! and Q are undetermined. Let’s choose a ! such that:

"!
(22) #$ A + %0 µ0 =0 Lorentz condition
"t
4-16

then from (17):

!2 A
(23) "# 2 A + $0 µ0 = µ0 J
!t 2

The RHS can be set to zero for no currents.

From (19), we have:

# 2! "
(24) $0 µ0 2 % & 2! =
#t $0

Eqns. (23) and (24) are wave equations for A and ! . Within the Lorentz gauge, we can
still arbitrarily add another ! (it must only satisfy 22). If we substitute (20) and (21) into
(24), we see:

"2 !
(25) # 2 ! $%0 µ0 =0
"t 2

So we can make further choices/constraints on A and ! as long as it obeys (25).

For a field far from charges and currents, J = 0 and ! = 0 .

!2 A
(26) "# 2 A +$0 µ0 =0
!t 2

" 2!
(27) #$ ! +%0 µ0 2 = 0
2

"t

We now choose ! = 0 (Coulomb gauge), and from (22) we see:

(28) !" A = 0

So, the wave equation for our vector potential is:

!2 A
(29) "# A +$0 µ0 2 = 0
2

!t

The solutions to this equation are plane waves.


4-17

(30) A = A0 sin (!t # k $ r + " )


! : phase
(31) = A0 cos (!t $ k % r + " # )

k is the wave vector which points along the direction of propagation and has a magnitude:

(32) k 2 = ! 2 µ0 "0 = ! 2 / c 2

Since (28) ! " A = 0

#k $ A0 cos (!t # k $ r + " )= 0

(33) ! k " A0 = 0 k ! A0

A0 is the direction of the potential → polarization. From (15) and (16), we see that for
! =0:

#A
E=$ = $ ! A0 cos (!t $ k % r + " )
#t

B = &' A = $ (k ' A0 )cos (!t $ k % r + " )

( k )E)B
4-18
Andrei Tokmakoff, MIT Department of Chemistry, 5/19/2005

4.4 Rate of Absorption and Stimulated Emission

The rate of absorption induced by the field is

!
wk ! (" ) = E0 (" ) k $% # ("k ! & " )
2 2
2
ˆ µ ! (4.1)
2"

The rate is clearly dependent on the strength of the field. The variable that you can most easily
measure is the intensity I (energy flux through a unit area), which is the time-averaged value of the
Poynting vector, S

c
S=
4"
(E ! B) (4.2)

c c 2
I= S = E2 = E0 (4.3)
4! 8!

Another representation of the amplitude of the field is the energy density

I 1 2
U= = E0 (for a monochromatic field) (4.4)
c 8!

Using this we can write

4! 2
w k! = 2 U (") k #$µ % ("k! & ")
2
ˆ ! (4.5)
"

1 2
or for an isotropic field where E 0 ! xˆ = E 0 ! yˆ = E 0 ! zˆ = E0
3

4! 2
U (") µ k! # ("k! $ ")
2
w k! = 2
(4.6)
3"

or more commonly

w k! = Bk! U (!k! ) (4.7)

4! 2 2
Bk! = 2 µ k! Einstein B coefficient (4.8)
3"
4-19

(this is sometimes written as Bk! = (2! 3" 2 ) µ k!


2
when the energy density is in ν).

U can also be written in a quantum form, by writing it in terms of the number of photons N

E 02 ! !3
N!! = U=N (4.9)
8" " 2 c3

B is independent of the properties of the field. It can be related to the absorption cross-section, σA.

total energy absorbed / unit time


!A =
total incident intensity (energy / unit time / area )
"" # w k! "" # Bk! U ("k! )
= =
I c U ("k! ) (4.10)

""
!A = Bk!
c

More generally you may have a frequency dependent absorption coefficient


!A (") # Bk! (") = Bk! g (") where g(ω) is a lineshape function.

The golden rule rate for absorption also gives the same rate for stimulated emission. We find for
two levels m and n :

wnm = wmn

Bnm U(! nm ) = Bnm U(! nm ) since U(! nm ) = U(! mn ) (4.11)

Bnm = Bmn

The absorption probability per unit time equals the stimulated emission probability per unit time.

Also, the cross-section for absorption is equal to an equivalent cross-section for stimulated
emission, (!A )nm = (!SE )mn .
4-20

Now let’s calculate the change in the intensity of incident light, due to absorption/stimulated
emission passing through sample (length L) where the levels are thermally populated.

dI
= ! N n "A dx + N m "SE dx (4.12)
I

dI
= ! (N n ! N m )"a dx (4.13)
I

N n , N m These are population of the upper and lower states, but expressed as a
population densities. If N is the molecule density,

# e!"En $
Nn = N % & (4.14)
' Z (

!N=N n " N m is the thermal population difference between states.

Integrating over a pathlength L:

I
= e!"N#a L for high freq. !N " N (4.15)
I0

# e! N"a L N : cm !3 "n : cm 2 L :cm

or written as Beer’s Law:

I
A = ! log =C"L C : mol / liter !:liter / mol cm (4.16)
I0

! = 2303 N "A
4-21

4.5 SPONTANEOUS EMISSION

What doesn’t come naturally out of semi-classical treatments is spontaneous emission—transitions


when the field isn’t present.

To treat it properly requires a quantum mechanical treatment of the field, where energy is
conserved, such that annihilation of a quantum leads to creation of a photon with the same energy.
We need to treat the particles and photons both as quantized objects.

You can deduce the rates for spontaneous emission from statistical arguments (Einstein).

For a sample with a large number of molecules, we will consider transitions between two states
m and n with Em > En .

The Boltzmann distribution gives us the number of molecules in each state.

Nm / Nn = e !!" mn / kT (4.17)

For the system to be at equilibrium, the time-averaged transitions up Wmn must equal those down
Wnm . In the presence of a field, we would want to write for an ensemble

?
N m Bnm U (!mn ) = N n Bmn U (!mn ) (4.18)

but clearly this can’t hold for finite temperature, where Nm < Nn , so there must be another type of
emission independent of the field.

So we write
Wnm = Wmn
(4.19)
N m (A nm + Bnm U (!mn )) = N n Bmn U (!mn )
4-22

If we substitute the Boltzmann equation into this and use Bmn = Bnm , we can solve for Anm :

( )
Anm = Bnm U (! mn ) e !! mn /kT " 1 (4.20)

For the energy density we will use Planck’s blackbody radiation distribution:

"!3 1
U (!) = 2 3 "!mn / kT (4.21)
"c e
! #1
#$%$&
U! N!

U ! is the energy density per photon of frequency ω.


N ! is the mean number of photons at a frequency ω.

!!3
" A nm = Bnm Einstein A coefficient (4.22)
# 2 c3

The total rate of emission from the excited state is

!!3
w nm = Bnm U (!nm ) + A nm using U (!nm ) = N (4.23)
" 2 C3

! !3
= 2 3 Bnm (N + 1) (4.24)
" c

Notice, even when the field vanishes ( N ! 0) , we still have emission.


Remember, for the semiclassical treatment, the total rate of stimulated emission was

!!3
w nm = Bnm (N ) (4.25)
" 2 c3

If we use the statistical analysis to calculate rates of absorption we have

!!3
w mn = Bmn N (4.26)
" 2 c3

The A coefficient gives the rate of emission in the absence of a field, and thus is the inverse of the
radiative lifetime:
1
!rad = (4.27)
A
Andrei Tokmakoff, MIT Department of Chemistry, 5/19/05
4-23

4.6 Quantized Radiation Field

Background

Our treatment of the vector potential has drawn on the monochromatic plane-wave solution to the
wave-equation for A. The quantum treatment of light as a particle describes the energy of the light
source as proportional to the frequency !! , and the photon of this frequency is associated with a
cavity mode with wavevector k = ! / c that describes the number of oscillations that the wave can
make in a cube with length L. For a very large cavity you have a continuous range of allowed k.
The cavity is important for considering the energy density of a light field, since the
electromagnetic field energy per unit volume will clearly depend on the wavelength λ = 2π/|k| of
the light.

Boltzmann used a description of the light radiated from a blackbody source of finite volume at
constant temperature in terms of a superposition of cavity modes to come up with the statistics for
photons. The classical treatment of this problem says that the energy density (modes per unit
volume) increases rapidly with increasing wavelength. For an equilibrium body, the energy
absorbed has to equal the energy radiated, but clearly as frequency increases, the energy of the
radiated light should diverge. Boltzmann used the detailed balance condition to show that the
particles that made up light must obey Bose-Einstein statistics. That is the equilibrium probability
of finding a photon in a particular cavity mode is given by

1
f (!) = !! / kT
(4.1)
e "1

From our perspective (in retrospect), this should be expected, because the quantum treatment of
any particle has to follow either Bose-Einstein statistics or Fermi-Dirac statistics, and clearly light
energy is something that we want to be able to increase arbitrarily. That is, we want to be able to
add mode and more photons into a given cavity mode. By summing over the number of cavity
modes in a cubical box (using periodic boundary conditions) we can determine that the density of
cavity modes (a photon density of states),
!2
g (!) = 2 3 (4.2)
"c
Using the energy of a photon, the energy density per mode is
!!3
!! g (!) = (4.3)
" 2 c3
and so the probability distribution that describes the quantum frequency dependent energy density
is

!!3 1
u (!) = !! g (!) f (!) = 2 3 !! / kT (4.4)
"c e #1
4-24

4.6 The Quantum Vector Potential


So, for a quantized field, the field will be described by a photon number Nk j , which represents the

number of photons in a particular mode ( k , j ) with frequency ! = ck in a cavity of volume v. For


light of a particular frequency, the energy of the light will be N kj!! . So, the state of the

electromagnetic field can be written:

!EM = N k , j , N k , Nk ,… (4.5)
1 1 2 , j2 3 , j3

If my matter absorbs a photon from mode k 2 , then the state of my system would be

"!EM = N k , j , N k # 1, N k , j , … (4.6)
1 2 , j2 3

What I want to do is to write a quantum mechanical Hamiltonian that includes both the matter and
the field, and then use first order perturbation theory to tell me about the rates of absorption and
stimulated emission. So, I am going to partition my Hamiltonian as a sum of a contribution from
the matter and the field:

H 0 = H EM + H M (4.7)

If the matter is described by !M , then the total state of the E.M. field and matter can be

expressed as product states:

! = !EM !M (4.8)

And we have eigenenergies

E = E EM + E M (4.9)
! N i !1
4-25
k Ni
Now, if I am watching transitions from an initial state ! to a final state k , then I can express the
initial and final states as:

matter field
!I = !; N1 , N 2 , N 3 … , N i , …

!F = k; N1 , N 2 , N 3 , … , N i ± 1, …

! + : emission "
# $
& % : absorption '

(4.10), (4.11)

Where I have abbreviated Ni ! Nki , ji , the energies of these two states are:

EI = E! + " N j "! j ( ) ! j = ck j (4.12)


j
(
EF = Ek + " N j !! j ± !! i )
j

!k $
So looking at absorption # & , we can write the Golden Rule Rate for transitions between
"! %
states as:

2! 2
w k! = " (E k # E ! # "$) %F V (t ) %I (4.13)
"

Now, let’s compare this to the absorption rate in terms of the classical vector potential:

2! q2 2 2
w k! =
v "2
'
kj
" (#k! $ #) 2 A k, j k %
m
ˆ j &p ! (4.14)

If these are to be the same, then clearly V (t ) must have part that looks like (!
ˆ "p ) that acts on the
matter, but it will also need another part that acts to lower and raise the photons in the field. Based
on analogy with our electric dipole Hamiltonian, we write:
4-26

!q 1
V (t ) = $
m v k, j
(
pk "# ˆ + p † "#
ˆjA k, j k
ˆ†
ˆ *j A k, j ) (4.15)

ˆ and A
where A ˆ † are lowering/raising operators for photons in mode k . These are operators in
k, j k, j

the field states, whereas pk remains only an operator in the matter states. So, we can write out the
matrix elements of V as

q 1
"F V (t ) "I = ! k pk #$
ˆ! ˆ …, N , …
… , N i ! 1,… A i i
m v
(4.16)
1
A (i
!)
= %k! k $#µ
ˆ !
v

Comparing with our Golden Rule expression for absorption,

! # 2k! 2
wk! = " ( # k! $ # ) 2 E0 µ k!
2
(4.17)
2" 2
#

We see that the matrix element

(! ) E 02 E 02
Ai = but = N!"
4v"2 8#
(4.18)
2#!
= N
v"

So we can write

2!!
 k, j = a
v " k, j
(4.19)
2!! †
 †k, j = a
v " k, j

where a, a † are lowering, raising operators. So


4-27

ˆ =
A &k, j
2$!
v#
ˆ j a kj e (
% (
i k ! r "#t )
+ a †kj e
" i (k ! r "#t )
)
So what we have here is a system where the light field looks like an infinite number of harmonic
oscillators, one per mode, and the field raises and lowers the number of quanta in the field while
the momentum operator lowers and raises the matter:

H = H EM + H M + V (t ) = H 0 + V (t )

H EM = , !#k a †kj a kj +
k, j
( 1
2 )

pi2
HM = , + Vi ( r , t )
i 2mi

!q
V (t ) = A"p
m

q 2$!
=,
m v #k
(ˆ j "p )%a k, j ei(k"r !#t ) + a †k, j e (
' (
*
! i k " r !#t ) &
)+
k, j

= V( ) + V(
! +)

Let’s look at the matrix elements for absorption (! k! > 0)

!q 2""
k, N i ! 1 V ( k, N i ! 1 (#$
ˆ p )a ! , N i
!)
!, N i =
m v%

!q 2""
= N i k #$
ˆ p!
m v%

2""%
= !i N i #$µ
ˆ k!
v
4-28

and for stimulated emission (! k! < 0)

!q 2""
k, N i + 1 V ( k, N i + 1 (#$
ˆ p )a † !, N i
+)
!, N i =
m v%

!q 2""
= N i + 1 k #$
ˆ pˆ !
m v%

2""%
= !i N i + 1 #$µ
ˆ k!
v

We have spontaneous emission! Even if there are no photons in the mode ( Nk = 0) , you can still
have transitions downward in the matter which creates a photon.

Let’s play this back into the summation-over-modes expression for the rates of
absorption/emission by isotropic field.

2! "2 2
w k" = ' d" # ("k" $ ") ' d% & k, N i + 1 V ( ) ", N i
+

# (2! )
2 2 3
j

2! "2 8!
(2!#")(Ni + 1)
2
= µ k"
# 2 (2!c )3 $3%&% '
$%&%' averageover polarization
number density per mode

4 (N i + 1)"3 2
= µ k"
3#c3

#"3
= Bk" (N i + 1)
! 2 c3
!
energy density
per mode

So we have the result we deduced before.


4-29

Appendix: Rates of Absorption and Stimulated Emission

Here are a couple of more detailed derivations:

Version 1:

Let’s look a little more carefully at the rate of absorption wk! induced by an isotropic, broadband
light source

wk ! = # wk ! (! ) " E (! )d !

where, for a monochromatic light source

!
wk ! (" ) = ( ) # ("k ! & " )
2 2
E0 " ˆ µ !
k $%
2" 2

For a broadband isotropic light source !(" ) d" represents a number density of electromagnetic
modes in a frequency range d! —this is the number of standing electromagnetic waves in a unit
volume.

For one frequency we wrote:

ˆ e i (k "r #$t ) + c.c.


A = A0 !

but more generally:

ˆje(
i k ! r "#t )
A = % A 0k $ + c.c.
k, j

where the sum is over the k modes and j is the polarization component.

By summing over wave vectors for a box of fixed volume, the number density of modes in a
frequency range d! radiated into a solid angle d! is

1 !2
dN = d! d"
(2# )
3
c3

and we get ρE by integrating over all !

1 "2 "2
#E (")d" = d" % d $ = d"
(2! ) 2! 2 c3
3
c3 !
4!

number density at ω
4-30

We can now write the total transition rate between two discrete levels summed over all
frequencies, direction, polarizations

! 1 "2
( ) ( ) ( ) d%
2 2
w k! = ) d" E " # " $ " k &
ˆ j 'µ !
(2! ) c3
0 k! 3
2" 2 j
#$$$%$$$&
8! 2
µ k!
3

E 0 ("k! ) "2
2
2
= µ k!
6! " c 2 3

We can write an energy density which is the number density in a range d! × # of polarization
components × energy density per mode.
rate of energy flow/c
!2 E2
U (!k! ) = 2 3 " 2 " 0
2# c 8#

w k! = Bk! U (!k! )

4# 2 2
Bk! = 2 µ k! is the Einstein B coefficient for the rate of absorption
3"

U is the energy density and can also be written in a quantum form, by writing it in terms of the
number of photons N

E 02 " !3
N"! = U (!k! ) = N 2 3
8" "c

The golden rule rate for absorption also gives the same rate for stimulated emission. We find for
two levels m and n :

wnm = wmn

Bnm U(! nm ) = Bnm U(! nm ) since U(! nm ) = U(! mn )

Bnm = Bmn

The absorption probability per unit time equals the stimulated emission probability per unit time.
4-31

Version 2:

Let’s calculate the rate of transitions induced by an isotropic broadband source—we’ll do it a bit
differently this time. The units are cgs.

The power transported through a surface is given by the Poynting vector and depends on k .

c c !2 A 02 ˆ !2 E 02
S= E"B = k=
4# 8# 2#

and the energy density for this single mode wave is the time average of S / c .

The vector potential for a single mode is

i (k "r #!t )
ˆe
A = A0 $ + c.c.

with ! = ck . More generally any wave can be expressed as a sum over Fourier components of the
wave vector:

i (k ! r "#t )
e
A = % Ak j $
ˆj + c.c.
k, j V

The factor of V normalizes for the energy density of the wave—which depends on k .

The interaction Hamiltonian for a single particle is:

!q
V (t ) = A"#
m

or for a collection of particles

qi
V (t ) = ! # A " pi
i mi

Now, the momentum depends on the position of particles, and we can express p in terms of an
integral over the distribution of particles:

p = $ d 3 r p (r ) p (r )= # pi ! (r " ri )
i

So if we assume that all particles have the same mass and charge—say electrons:

!q
V (t ) =
m # d 3r A (r ,t ) " p( r)
4-32

The rate of transitions induced by a single mode is:

2! q2 2 2
(w k! )k, j = 2
" (#k! $ # ) 2
ˆ j & p (r ) !
A k, j k %
V" m

And the total transition rate for an isotropic broadband source is:

w k! = ! (w k! )k, j
k, j

We can replace the sum over modes for a fixed volume with an integral over k :

1 d 3k dk k 2 d% d& & 2 d%
V
! "'
(2# )3 ' (2# )3
$ $ ' ( 2# C)3
k

So for the rate we have: d! = sin " d" dø

2! "2 q2 2 2
w k! = ) d" 2 # ("k! $ ") 2 ) d% ( k &'
ˆ p (r ) ! A k, j
" (2! c )3 m j

can be written as k

The matrix element can be evaluated in a manner similar to before:

q !q
m
ˆ j #p (r ) ! =
k"
m i
& k "#
ˆ pi $ ( r # ri ) !

!i
= ˆ k [r1 , H 0 ]$ ( r ! ri ) !
q & q "#
" i

= !i%k! & q "#


ˆ k r1 !
i

= !i%k! k "#µ
ˆ ! where µ = & q i ri
i

For the field


2
2 Ek E02
! Ak j
= !2" ij
=
4" 2
kij kij i
4-33

2! " 2 "k2! 2
( ) (
2
Wk ! = ) d " 2 # " $ " E ) j
d % k &
ˆ ' µ !
4" (2! c )3
k ! 0 j
"2
#$$$%$$$&
2
8! / 3 µk!
for isotropic

"2 2
= E0 µk !
6! " c
2 3

For a broadband source, the energy density of the light

I ! 2 E02
U= =
c 8" 3c3

4" 2
Wk ! = Bk ! U (!k ! )
2
Bk ! = µk !
3" 2

We can also write the incident energy density in terms of the quantum energy per photon. For N
photons in a single mode:

!! 3
N!! = Bk" N
" 2 c3

where Bk! has molecular quantities and no dependence or field. Note Bk! = B!k —ratio of S.E. =
absorption.

The ratio of absorption can be related to the absorption cross-section, ! A

P total energy absorbed/unit time


!A = =
I total intensity (energy/unit time/area)

P = "!" Wk! = "! Bk! U (!k! )


I = cU (!k! )
"!
#a = Bk!
c

or more generally, when you have a frequency-dependent absorption coefficient described by a


lineshape function g(! )

"!
"a (!) = Bk! g (!) units of cm 2
c
4-34

Readings
1
Jackson, J. D. Classical Electrodynamics (John Wiley and Sons, New York, 1975).
2
Cohen-Tannoudji, C., Diu, B. & Lalöe, F. Quantum Mechanics (Wiley-Interscience, Paris, 1977),
Appendix III.
3
Cohen-Tannoudji, et al. app. III, p. 1492.

You might also like