Robert Sang - QM
Robert Sang - QM
R.T. Sang
SCE3337 Quantum Mechanics III
(Quantum Mechanics and Quantum Optics)
Teaching Team
Dr Robert Sang
Robert Sang can be found in Science II Level 0 Room 0.15, non-personal contact is
possible via email at the following address: [email protected] or via the telephone on
3875 3848 .
Lecture notes/Problem Sheets can be found on the web page:
http://www.sct.gu.edu.au/~sctsang
It is your responsibility to ensure that you are up-to-date with them and
should be downloaded prior to each lecture.
1 . 0 Introduction
This course has been divided into two parts; The first part deals with the introduction of
operators and techniques that are applied on a regular basis in dealing with problems that
require the use of Quantum Mechanics these include:
The second half of the course deals with Quantum Optics which includes
There will be eight lectures that deal with the development of the tools of Quantum
Mechanics which will also involve three tutorials.
Following the exam will be a further nine lectures in quantum optics which will also
have two tutorials.
There will be a final examination at the conclusion of the course which will cover all
aspects of the course.
Useful Texts
Modern Physics and Quantum Mechanics- Anderson
Quantum Mechanics- Cohen-Tannoudji, Diu, Laloë
The Quantum Theory of Light- Loudon
Optical Resonance and Two Level Atoms-Allen and Eberly
SCE3337 Quantum Mechanics III 2
R.T. Sang
2 . 0 Basics of One Particle Wave Function Space
The quantum state of a particle is defined at any given instant by a wavefunction ψ(r,t).
Recall from QMII that the wave function of a particle ψ(r,t): |ψ(r,t)|2 d 3 r represents the
probability of finding, at time t, the particle in volume d3 r =dxdydz about the point r.
Therefore the total probability of finding the particle over all space is equal to 1,
mathematically this is expressed as:
∫ (r,t) d 3r = 1
2
where the integration extends over all space. The above integral must converge (the sum of
the probabilities must yield 1) and this type of integral is called a square integrable
function. Mathematicians call this set of functions L2 . In QM this set of functions is to
wide in scope since |ψ(r,t)|2 has an actual physical meaning therefore we can only keep
functions ψ(r,t) which are everywhere defined (we can’t have particles disappearing!),
continuous and differentiable. It is also possible to define wavefunctions that have a
bounded domain, this allows us to be certain that a particle can be found in a finite region
of space. For example we can define a space inside the laboratory in which our particle can
be found. We shall call this group of functions that satisfy our conditions the F set of
wavefunctions which is a subset of the L2 set. This statistical interpretation of Quantum
Mechanics is due to Born, Heisenberg and Bohr.
a) F as a vector Space
It is easy to show that F satisfies all the criteria of a vector space, for example if ψ1 (r) and
ψ2 (r) exists in F then the sum of the two vectors should yield another vector that belongs to
the vector space F:
where λ 1 and λ 2 are two arbitrary complex numbers. Now in order to show that ψ(r)
belongs to the wavefunction space F then we need to demonstrate that it is square
integrable. Squaring ψ(r) gives
The last two terms of this expression have the same modulus which has the upper limit
when ψ1 (r)= ψ2 (r) then the final terms equal
1 | 2 | (r)| +| (r)| ]
|ψ(r)|2 is therefore equal to or smaller than the maximum function whose integral must
converges since ψ1 (r) and ψ2 (r) are square integrable then ψ(r) must also be square
integrable and hence exists in the vector space F.
SCE3337 Quantum Mechanics III 3
R.T. Sang
b) The Scalar Product
( , )=∫ *( r) (r)d 3r
The integral always converges if ϕ(r) and ψ(r) belong to F. The scalar product has the
following properties
( )=( )*
( 1 + 2 )= 1 )+ 2 )
( 1 + 2 )= 1 * )+ 2 * )
The scalar product is said to be linear in the second relationship and anti-linear with
respect to the third relationship. If (ϕ,ψ)=0 then the two functions are said to be
orthogonal.
( , )=∫ 2
(r,t) d 3 r
This number is always real and positive and can only equal zero if ψ(r) =0. (ψ,ψ) is the
norm of ψ(r).
c) Linear Operators
(r)'=A (r)
A (r) = A 1 (r)+ 2 (r)] = A 1 (r)+ 2 (r)
∂ψ(x,y,z)
The differential operator Dx which differentiates wrt x: Dx ψ(x,y,z) =
∂x
SCE3337 Quantum Mechanics III 4
R.T. Sang
d) Products of Operators
We say that B operates on ψ(r) which produces a new function ϕ(r)=Bψ(r), A then
operates on the new function ϕ(r).
In general AB BA i.e. they do not commute. The commutator of the operators A and
B is written as [A,B] and has the following definition
[ A, B] = AB - BA
Hence [X, Dx ] = -1. Since the commutator is non-zero X does not commute with Dx .
Exercise 1: Evaluate the commutation [Dx , X 2 ] where the operator X2 =x2 . Is it equal to
[X2 , Dx ]?
SCE3337 Quantum Mechanics III 5
R.T. Sang
(u (r), u (r))
i j = ∫ u (r )*u (r)d r =
i j
3
ij
where δij is the Kronecker delta function which takes the following values
This set of functions is said to constitute a basis if every function (r) that
exists in F can be expanded in one and only one way in terms of u i(r):
(r ) = ∑ ci ui (r)
i
The components of the wave function in the ui(r) basis my be found by multiplying both
sides of this equation by uj(r) and integrating over all space.
Proof:
(u j (r ), (r)) = (u j (r), ∑ ci ui (r))
i
= ∑ ci ji
i
= cj
The coefficient cj of the wave function ψ(r) on uj(r) is equal to the scalar product of wave
function ψ(r) with uj(r). Once the basis set {u i(r)} is chosen, it is completely equivalent to
express the wave function ψ(r) in terms of the ci coefficients with respect to their basis
functions. The set of numbers c j is said to represent (r) in the {u i(r)} basis.
SCE3337 Quantum Mechanics III 6
R.T. Sang
f) Expression for the Scalar Product in Terms of the Components
Let ϕ(r) and ψ(r) be two wave functions which have the following expansions in the basis
{ui(r)}:
(r) = ∑ bi ui (r)
i
(r ) = ∑ c j u j (r )
j
( , ) = (∑ biui (r ), ∑ c ju j (r))
i j
= ∑ bi * c j ij
i, j
= ∑ bi * ci
i
Therefore
( , ) = ∑| c | i
2
The scalar product of two wave functions can be expressed in terms of the components of
the functions in the basis {ui(r)}.
SCE3337 Quantum Mechanics III 7
R.T. Sang
3 . 0 Dirac Notation and the Postulates of Quantum Mechanics
We now look at using a new notation which is simpler to write down that the notation that
has been used until present (saves us having to write all the messy integrals in QM). This
new notation is called Dirac Notation as it was first introduced by Dirac. We will assume
that each quantum state can be represented by a state vector which belongs to an abstract
space we will call E which is called the state space of the particle. E is a subspace of Hilbert
Space∗ .
Postulate 1: The quantum state of any physical system is characterised by a state vector,
belonging to a space E which is that state space of the system.
3 . 1 Dirac Notation
Any element or vector of E space is called ket therefore by postulate 1 we can define any
wavefunction which describes a state α by a ket:
For every ket there is a bra and vice versa. We will now consider some more of the
postulates of QM using this notation.
Postulate 2: For every physical observable the is an associated operator Q such that
Q|α> = q α|α>
Note that operators act on kets from the left and bras from the right:
<α|Q = qα*<α|
∗
Hilbert Space you may recall from second year is an infinite dimensional complex space. A Hilbert space
is a vector space in which there is a well defined scalar product and a norm in terms of this scalar product
and the vector space is complete and all function defined within this space must be square integrable.
SCE3337 Quantum Mechanics III 8
R.T. Sang
Postulate 3: The set of eigenfunctions {|i>} associated with an operator Q form a
complete orthogonal set of wavefunctions and satisfy the following eigenvalue equation;
Q|i> = qi|i>
∫ψ i * ψ jdτ = δij
Another way to think of this is that there is no overlap of the wavefunctions in all space
hence the dot product must yield zero. Therefore no eigenfunction depends on any either
eigenfunction for its definition and as such is independent.
The expansion coefficients are in general complex. Note that N does not have to be finite as
there can also be an infinite number i.e. N→∞ of eigenfunctions to provide a full and
complete description of an interaction process (eg: electron-atom collisions) although it is
often that only a few of these eigenfunctions are necessary for a accurate calculation.
It should also be noted that the expansion may require an integration over continuous
states, as in the case where the interaction takes place in continuum. This case can occur for
an electron-atom collision in which the atom is ionised after the interaction and hence the
wavefunction describing the system has to take into account that the incident, scattered and
ionised electrons are free of a constraining potential of the atom.
<ψ|ψ> = 1
= ∑ c i *c j < i | j >= ∑ c i *c jδij
i,j i,j
= ∑ ci *ci = ∑ | ci | = 1
2
i i
As we would expect for a normalised wavefunction the sum of the coefficients must be
unity since the probability must be conserved.
SCE3337 Quantum Mechanics III 9
R.T. Sang
Postulate 5: The expectation of an operator is described in this notation by <Q> =
< |Q| >.
∫ * Q d 3r
∫ ψ *Q ψd r
3
<ψ|Q|ψ> = = for normalised wavefunctions
∫ 3
* d r
Q |ψ> = Q ∑ c j | j >
j
= ∑c q | j > j j
j
= ∑ ∑ c *c q i j j < i| j > = ∑ ∑ c *c q δ
i j j ij
i j i j
= ∑| c | q 2
i i
i
This equation yields a weighted average of the possible eigenvalues of Q. The probability
of a single measurement yielding qi is therefore given by |ci|2 . The probability of finding a
system in the state |i> is |ci|2 .
Q = Q†
Given that Q|i> = qi|i> and that Q = Q† then we need to show that qi are real then
Now the operator Q has been defined to be Hermitian i.e. <i| Q† |i> = <i| Q|i>
Recall that for a complex number q that q = Re(q) + Im(q) and q* = Re(q) - Im(q)
Therefore equating the two outcomes above reveals that the only way that qi * = qi is if they
have no imaginary terms so qi must be real.
Suppose that we have a complete orthonormal set of wavefunctions {|a>}. We defined the
Projection Operator# such that
Pa = |a><a|
Thus the projection operator projects a wavefunction onto a particular basis function.
Summing over all projection operators in the basis {|a>} gives
∑ P | ψ >=∑ c
a a |a>
a a
The term ∑c a | a > is just the wavefunction written in the {|a>} basis set hence
a
#
Note the difference between the projection operator given by |a><a| and the integral that we introduced
earlier <a|a>.
SCE3337 Quantum Mechanics III 11
R.T. Sang
∑ P | ψ >=I| ψ >
a
a
where I is the Identity Operator which maps a wavefunction back onto itself. The complete
sum of projector operators therefore yields
∑P a = I or ∑ | a >< a | = I
a a
SCE3337: QMIII 1
R.T. Sang
Lecture 2
3 . 3 Matrix Representation of Operators
This is a typical quantum mechanical process where we say that Q acts on |ψa> to produce a
new wavefunction |ψb >. We now apply postulate 4 and let |ψb > be represented by a
superposition of basis functions {|k>} such that
|ψ > = ∑ a | j >
a j
j
We now multiply the expression by the bra-vector <i| which yields for the LHS:
bi = ∑a Q j ij
j
Qij is called the matrix element of the operator Q and is given by Qij = <i|Q|j>
It is easy to see how this process defined by the expression |ψb > = Q|ψa> can be represent
by the following matrix equation
SCE3337: QMIII 2
R.T. Sang
Wavefunction kets are represented by column vectors whereas their complex conjugate
(bras) are represented by row vectors (b1 *, b 2 *, b 3 *, ..., b n *).
3 . 4 Matrix Inversions
A very useful property of the matrix representation is a technique that allows matrix
inversions:
where Q-1 is the inverse of the matrix Q. This expression is valid provided that the
determinate of the matrix |Q| ≠ 0, i.e. Q is non-singular. Recall from first year maths that
Q Q-1 = I
where I is the Identity Matrix (in QM we call it the Identity Operator since I|ψb > = |ψb >).
This matrix is given by:
1 0 0 . . . 0
0 1 0 . . . 0
0 0 1 . . . 0
I= . . . . . . .
. . . . . . .
. . . . . . .
0 0 0 . . . 1
If Q is non singular a very simple technique that can be used to find the inverse is the
cofactor technique:*
1 T
Q-1 = C
|Q |
*
It is also possible to use the Gaussian elimination or row reduction techniques to find the inverse.
SCE3337: QMIII 3
R.T. Sang
CT is the transpose1 of the cofactor matrix which is a matrix of cofactors, the elements of
which are cofactors of the original matrix Q. The individual elements are called cofactors
cij.
To find the inverse, one simply finds the cofactor matrix, then take the transpose (exchange
rows and columns), and then multiply each matrix element by 1/|Q|
Cij = (-1)i+j|qij|
|qij| is the minor of the element Qij. It is a scalar value given by the (N-1) x (N-1)
determinant remaining in the original matrix when the ith row and jth column are struck
out.
a 0 − b
A = 0 1 0
b c a
1 0 0 0 0 1
First Row: −(1)1+1 = a, −(1)1+ 2 = 0, −(1)1+ 3 = −b
c a b a b c
0 −b a −b a 0
Second Row: −(1)2+1 = −bc , −(1)2+ 2 = a 2 + b2 , −(1)2+ 3 = −ac
c a b a b c
0 −b a −b a 0
Third Row: −(1)3+1 = b, −(1)3+2 = 0, −(1)3+3 =a
1 0 0 0 0 1
a 0 −b
C = − bc a + b2 −ac
2
b 0 a
Now to calculate the inverse matrix A-1 we need the transpose of the matrix:
a −bc b
C = 0 a + b2 0
T 2
− b −ac a
1
To find the transpose just swap the rows with the columns, ie CijT = Cji
SCE3337: QMIII 4
R.T. Sang
1 0 0 0 0 1
|A| = +(a) - (0) + (-b)
c a b a b c
= a2 +b2
a −bc b
1
-1
A = 2 0 a + b2
2
0
a +b
2
− b −ac a
Q|ψ> = λ|ψ>
Expanding |ψ> with respect to the complete orthonormal basis set of functions {|j>} yields
N
|ψ> = ∑c j | j>
j
∑ c jQ | j > = λ ∑ c j | j >
j j
N N
∑ c jQij = λ∑ c jδ ij
j j
∑ c (Q j ij − λδ ij ) = 0
j
The only non-trivial solution of this expression (cj≠0) is when the determinate
The determinant yields a polynomial of order N where N is the dimension of the matrix.
The roots of the polynomial are the eigenvalues of the matrix Q.
Note that if Q was a diagonal matrix then the secular equation would be
Q11 − λ 0 0 .
0 Q22 − λ 0 .
|Qij-λδ ij| = =0
0 0 Q33 − λ .
. . . .
Hence the eigenvalues of a diagonal matrix are just equal to the diagonal elements Q11 ,
Q22 ,... Q NN. It can be shown that there is a transformation, called a unitary transformation,
such that any Hermitian matrix may be written in diagonal form. This is a very useful
property when Q is very large as it enables on to find the eigenvalues of a large system very
simply using a computer program.
1 0 −2
Q= 0 0 0
−2 0 4
Find the eigenvalues and hence the eigenvectors for this operator.
Q|a> = λ|a>
∑ c (Q j ij − ij )=0
j =1
1− λ 0 −2
0 −λ 0 =0
−2 0 4 −λ
−λ 0 0 0 0 −λ
⇒ (1− λ ) − (0) + (−2) =0
0 4 −λ −2 4− λ −2 0
⇒ (1- λ){ - 4λ + λ2 } + 4 λ = 0
⇒ -4 λ + λ2 + 4 λ2 - λ 3 + 4 λ = 0
⇒ λ{ 5λ - λ2 } = 0
⇒ λ{ λ ( 5- λ ) } = 0
This equation is satisfied when λ = 0,0,5 and hence these are the eigenvalues λ 1 =0, λ 2 =0
λ 3 =5.
Step 3: We now determine the three eigenfunctions which must be orthogonal. Going
back to the eigenvalue equation and substituting in the value for λ 1 :
3 3 3
∑ c (Q j ij − ij ) = ∑ c j (Qij − 0. ij ) = ∑ c jQij = 0
j =1 j =1 j =1
1 0 −2 c1
⇒ 0 0 0 • c 2 = 0
−2 0 4 c 3
c1 - 2c3 =0
0 =0
-2c1 +4c3 = 0
2
Hence the eigen vector for the eigen value λ 1 is: |a1 > = c3 0
1
At this point c 3 is arbitrary but since it is usual to normalise the eigenfunctions we can
solve for this scalar using the normalisation condition:
2 2 2 2 2
c1 + c2 + c3 = 2.c 3 + c 3 = 1
1
⇒ c3 =
5
2
1
|a1 > = 0
5
1
Now for the second eigenvalue λ2 =0 we would obtain the same result from our secular
equation ie c2 would be arbitrary and c1 = 2c 3 . But you recall that the solutions for each
eigenvalue must have eigenvectors that are orthogonal. Hence |a2 > must be orthogonal to
|a1 >. This is simple to evaluate since.
c1 2c 3
|a2 > = c 2 = c 2 (this uses the fact that c1 = 2c3 )
c c
3 3
Thus
2c 3
<a1 |a2 > = [ 2 0 1] • c2 = 0
c
3
0
|a2 > = c 2
0
For the normalised eigenfunction c2 can be determined from the normalisation condition.
2 2 2 2
c1 + c2 + c3 = c 2 = 1 ⇒ c2 = 1
Thus
0
|a2 > = 1
0
Now we determine the final eigenfunction for the eigenvalue λ 3 =5. Applying the identical
procedure as before
1 − 5 0 −2 c1
0 −5 0 c 2 = 0
−2 0 4 − 5 c 3
−4 0 −2 c1
⇒ 0 −5 0 c 2 = 0
−2 0 −1 c 3
This yields
-4c1 - 2c3 =0
-5 c2 =0
-2c1 - c3 =0
Thus
1
c1 = − c and -5 c2 = 0 ⇒ c2 = 0
2 3
The final eigenfunction in its general form is
1
|a3 > = c3 0
−2
2 2 2 2 2 1
Applying normalisation gives c1 + c2 + c3 = c 3 + −2c 3 = 1 ⇒ c3 =
5
SCE3337: QMIII 9
R.T. Sang
1
1
|a3 > = 0
5
−2
Exercise: Check that the eigenvector |a3 > is orthogonal to |a1 > and |a2 >
SCE3337: QMIII 1
R.T. Sang
Lecture 3
4 . 0 The Simple Harmonic Oscillator
x m
F=-kx
Consider the simple harmonic oscillator in the figure above. When the mass m is pulled
back and released it undergoes Simple Harmonic Motion (SHM) which is
expressed by the equation:
d 2x
m + kx = 0
dt 2
k is known as the stiffness of the spring or the spring constant for the restoring force
shown in the figure. This equation is well known and can be rewritten in the following
form
d2 x
=− 2
x
dt 2
k
where = is the resonant frequency of the SHO. The equation of motion has the
m
solution (this is very simple just try trial solution eqt and sub into equation above);
x = x0 cos( t)
The simple harmonic oscillator plays a large role in many areas of both classical and
quantum physics. It provides a model for systems diverse as
• Atoms
• Molecules
• Electromagnetic Radiation
• Atom Traps
The classical equations of motion for a SHO are given by
px 2
T= (Kinetic Energy)
2m
1
V = kx 2 (Potential Energy)
2
px 2 1 2
Recall that the Hamiltonian is H = T + V = + kx
2m 2
SCE3337: QMIII 2
R.T. Sang
We now use the principle of correspondence to determine the quantum mechanical
Hamiltonian for the SHO. By direct analogy we exchange the linear momentum
operator px by
∂
px −ih
∂x
Thus
−h2 2 1
H= + m 2
x2
2m x 2 2
Let | > be an eigenfunction of the operator H. Recall that the Hamiltonian operating on
a wavefunction yields energy eigenvalues and we now week solutions to the eigenvalue
equation:
−h 2 2 1
+ m >= E | >
2
( x 2 )|
2m x 2 2
−h 2 2 1
⇒ > +( m x 2 − E) | >= 0
2
|
2m x 2 2
2
For convenience we multiply through by which yields
h
−h 2 m 2E
| > +( x2 − )| >= 0
m x2 h h
This equation may be solved by a power technique developed by Dirac which is quite
often used in solving equations involving non-commuting operators. It also forms the
basis of much advanced theoretical work in quantum mechanics.
m
Let = then the equation above can be recast:
h
−1 2
2E
2 | >+ x2 | >= | >
x h
we now make the substitutions
q= x ⇒ q 2 = x 2 and
−i 1 ∂2
p= ⇒p =−
2
x ∂x 2
(p 2 + q 2 )| >= | >
2E
where is a dimensionless energy term given by =
h
SCE3337: QMIII 3
R.T. Sang
4 . 1 The Quantum Mechanical Properties of p and q
Proof:
Therefore [p,q] = -i
Since p and q do not commute we can not simply factor the expression p 2 + q 2 into
(p+iq)(p-iq) in the eigenvalue equation. Instead consider the following linear forms:
1
⇒ p2 + q2 =
2
{(q + ip )(q − ip ) + (q − ip )(q + ip )}
We have now transformed or operator form of the eigenvalue equation into an equation
that contains linear terms of the operators p and q. This effectively transforms a second
order differential equation into two consecutive first order differential equations.
We now introduce two new operators called the raising and lowering operators as
1
a= (q + ip) The lowering operator which is commonly called the
2
annihilation operator
1
a† = (q − ip) The raising operator also called the creation operator
2
SCE3337: QMIII 4
R.T. Sang
The Hamiltonian expression can now be written in terms of these operators:
1 1 1 1
p2 + q2 = { (q + ip) (q − ip) + (q − ip) (q + ip)}
2 2 2 2
p 2 + q 2 = {aa † + a† a}
We can further rewrite the operators p and q in terms of the operators a and a† :
1 1
a + a† = (q + ip) + (q − ip) = 2q
2 2
( a + a †)
1
⇒ q=
2
And
1 1
a − a† = (q + ip) − (q − ip) = 2ip
2 2
( a − a† )
i
⇒ p=−
2
[ a ,a ] = a a
† †
− a† a =
⇒
1 1 1 1 1
{ (q + ip) (q − ip) − (q − ip) (q + ip)} = {(q2 + p2 + 1)− (q 2 + p2 − 1)}
2 2 2 2 2
⇒ [ a ,a ] = a a
† †
− a† a =1
a a † = 1+ a† a
⇒ Important result!
a † a = a a† -1
It is also useful to note that using the commutation relation we can re-write the
Hamiltonan in terms of these operators as
H = aa † + a † a
= 1 + 2a †a
= 2aa † − 1
The multiplication of the two operators a† a is called the number operator n, the reason
for this will become more clear later.
SCE3337: QMIII 5
R.T. Sang
4 . 3 Eigenvectors and Eigenvalues of a† a and a a† :
Using the results derived in the last page we can now evaluate our eigenvalue equation.
( − 1)
⇒ a †a | >= nˆ | >= | >
2
Hence the number operator nˆ has the same eigenfunctions as the Hamiltonian.
( − 1)
The eigenvalue of the number operator nˆ = a† a is given by
2
Consider a further operation in which we operate with a† a:
( − 1)
a †a | >= | >
2
( − 1)
(aa † − 1)| >= | >
2
( − 1)
aa † | >= | > + 1| >
2
( + 1)
⇒ aa † | >= | >
2
Once again we have shown that the operator aa† has the same eigenfunctions as the
Hamiltonian.
( + 1)
The eigenvalue of aa† is given by
2
SCE3337: QMIII 6
R.T. Sang
4 . 4 Eigenvectors and Eigenvalues of the Annihilation and Creation
Operators a and a† :
We now consider the application of the raising and lowering operators to the
Hamiltonian eigenvalue equation.
H|n> = n |n>
we want to factor out a† the reason will become apparent further along, to accomplish
this we use the commutation relation and replace a † a with aa † − 1
This is exactly the same form of the energy eigenvalue equation given above.
This shows that given a state |n> of dimensionless energy n there exists another state
which has two units of energy more than this state. The upper state has been defined as
|n+1>.
One can repeat this procedure which yields a ladder of levels increase to infinity.
From this we can conclude that
aH|n> = n a|n>
This is the same as the eigenvalue equation. This equation demonstrates that given a
state |n> of dimensionless energy n there exists a state which has 2 units of lower
energy than this state. The lower state is defined as |n-1>.
( + 1)
aa † | n >= n
| n > and (1)
2
( − 1)
a † a | n >= n
|n> (2)
2
Summing equations (1) and (2) yields:
( + 1) ( n − 1)
(aa + a† a) | n >= n + |n >
†
2 2
( + 1) ( n − 1)
⇒ H | n >= n + |n >
2 2
If now operate with a† on equation (1) we can determine a higher order eigenfunction of H:
( + 1) †
a † a[a† | n >] = n
[a | n >] (4)
2
( + 1)
a † a | n + 1 >= n
| n + 1>
2
Replacing a† a with aa† -1 (commutation relation) in equation (4) reveals
( + 1) †
(aa † − 1)[a † | n >] = n
[a | n >]
2
( + 1) †
⇒ aa † [a† | n >] − 1[a † | n >] = n
[a | n >]
2
( + 1) †
⇒ aa † [a† | n >] = n
[a | n >] + 1[a† | n >]
2
( + 3) †
⇒ aa † [a† | n >] = n
[a | n >] (5)
2
We now sum equations (4) and (5) giving
SCE3337: QMIII 2
R.T. Sang
( + 3) ( n + 1) †
(aa + a† a)[a† | n >] = n + [a | n >]
†
2 2
( + 3) † †
a † aa † [a † | n >] = n
a [a | n >]
2
( + 3) † 2
⇒ a † a(a† )2 | n >= n
(a ) | n >
2
( + 3) † 2
⇒ a † a[(a† )2 | n >] = n
[(a ) | n >] (7)
2
( + 3)
⇒ a † a | n + 2 >= n
|n+2>
2
We now replace a† a in equation (7) by aa† -1 :
+ 3) † 2
(aa − 1)[(a† ) 2 | n >] =
(
[(a ) | n >]
† n
2
( + 3) † 2
⇒ aa † [(a† )2 | n >]− 1[(a† )2 | n >] = n
[(a ) | n >]
2
( + 5) † 2
⇒ aa † [(a† )2 | n >] = n
[(a ) | n >] (8)
2
( + 3) ( n + 5) † 2
(a a + aa )[(a ) | n >] = n + [(a ) | n >]
† † † 2
2 2
Therefore
If we do this procedure m times then equations (3), (6) and (9) allow us by induction to
deduce the general result:
It is clear that physically we must have a lower limit when applying the lowering operator
since we have postulated that the energy of the simple harmonic oscillator can not be
negative. Thus there exist a lowest positive energy state which we define to be the ground
state. This means that the measurement of energy, or the expectation value of this energy
after a measurement has been obtained must be non-negative. The expectation value is
given by:
Hence for
Now we define the wavefunction |n> = |0> as the ground state (the state with lowest
energy). Clearly acting on this wavefunction with the lowering operator must result in a
condition of no energy (otherwise the state |0> would not be the lowest state). Hence
a|0> = 0
−1
a † a | n >= n
|n>
2
Hence for the ground state
−1
a † a | 0 >= 0
|0> (2)
2
So assuming that |0> exists, for equation (2) to be consistent with equation (1) requires that
−1
a † a | 0 >= 0
| 0 >= 0
2
⇒ 0 =1
Recall from last lecture that n was dimensionless energy term given by
2En
n = putting n = 0:
h
2E0
0 =
h
2E0
⇒ 1=
h
1
⇒ E0 = h
2
This says that the lowest energy state is non-zero which is strikingly different to the
classical SHO which has a lowest energy state equal to zero.
1
E0 = h
2
= 1 + h
1 1
⇒ E1 = h + h
2 2
= 2 + h
1 1
⇒ E2 = 2h + h
2 2
Therefore by induction
En = n + h
1
where n=0,1,2,.......
2
Energy
3
En +1 = n + h
1
2
a†
En = n + h
2
1 a
En − 1 = n − h
2
3
E1 = h
2
1
E0 = h
2
E=0
This enables us to diagrammatically represent the energy states of the quantum SHO:
Note that the dimensionless energy term is given by
2En 2 1
= = n+ h = 2n + 1
h 2
n
h
The meaning of the operator a† a as the number operator nˆ is now clear since:
−1
a † a | n >= nˆ | n >= n
|n>
2
2n + 1 − 1
nˆ | n >= | n >= n | n >
2
Thus
nˆ | n >= n | n > i.e. n is the nth energy state of the harmonic oscillator.
SCE3337: QMIII 6
R.T. Sang
4 . 8 Normalisation of Eigenfunctions
It should be noted that the derivations that we have used so far for the wavefunctions of
quantum mechanical SHO namely |n+1> = a† |n> have not been normalised. To achieve this
we require that
<n-1|n-1> = <n|n> = 1
As a result we need to introduce a numerical factor into the expressions which have been
derived above to account for the normalisation. Therefore we should write the expressions
as
a n = Bn−1 n − 1
a † n = Bn+1 n + 1
where B m are coefficients that are yet to be determined. As an example we will determine
the normalised ground state wavefunction.
Recall
1
a= (q + ip)
2
1
a0 = (q + ip) 0 (1)
2
−i
Also recall that q = x and p =
x
then
q q
= ⇒ x= sub ( ∂ x) in for p gives
x
−i −i
p= = = −i
x q q
1
a0 = (q + ) 0 =0
2 q
⇒ (q + ) 0 =0
q
⇒ 0 = −q 0
q
SCE3337: QMIII 7
R.T. Sang
⇒ ∫ =− ∫ q q
0
q2
⇒ ln 0 = − + c0
2
Thus
q2
−
2
0 = C0 e
< 0| 0 >= 1
∞ ∞
∫ ∫e
− q2
⇒ < 0| 0 >= dq =| C0 |
* 2
0 0 dq
−∞ −∞
∫ e −q dq =
2
where the integral
−∞
1 =| C0 | 2
Problem Sheet 2 will be used to determine the normalisation coefficients for an arbitrary
wavefunction |n>.
SCE3337: QMIII 1
R.T. Sang
Lecture 5
5 . 0 Approximation Techniques in Quantum Mechanics
We have already solve the Schrödinger equation for the time independent case for the
simple harmonic oscillator. In principle, the physical system is described by
either of these equations depending whether we are interested in behaviour
or a system that is time dependent or time independent.
As it turns out real nature is not simple and there are few exactly solvable problems. Some
examples of exact solutions are:
Actually even in reality the hydrogen atom is not exactly solvable even though the
wavefunction can be written down exactly for this system. Small perturbations such as spin
orbit effects have to be considered to allow for real experimental observations.
If a system consists of more than one particle then interactions between the particles also
has to be taken into consideration. As an example the He atom has two electrons that not
only interact with the ionic core but also interact with each other. It is impossible to solve
such a three body problem exactly.
Early quantum physicists did not however completely give up on solving these problems
they made allowances for difficult problems by introducing approximation techniques.
SCE3337: QMIII 2
R.T. Sang
The basic idea of time independent perturbation theory is as follows. Suppose we have a
system which has a Hamiltonian H0 and we apply a small perturbation, h, to the system
such that the system Hamiltonian is:
H = H0 + h
Here H0 has a much greater influence over the system than h does. We also assume that the
Schrödinger equation can be solved exactly:
H0 | i >= Ei | i >
|i> are the associated eigenkets of the Hamiltonian H0 and Ei are the corresponding
eigenvalues. The eigenkets form a complete orthonormal set as we have shown in previous
lectures. Then any ket vector can be written as a linear superposition of the eigenkets |i>
with
H = H0 + h
where is a free parameter defined in the interval 0 1. This allows us to turn the
perturbation on and off.
Where |i'> are the eigenkets of the perturbed system and as such are not the
same eigenkets as |i> and as such the eigenvalues Ei Ei'.
We further assume that the sets { |i'> } and { |i> } are non-degenerate (i.e. their
eigenvalues are unique). This point will be important in the following discussion as it
allows us to get around the problem of a division by zero that will come up later.
Perturbation theory covering degenerate states will be covered later in the course. We will
also say in the limit that as 0
lim Ei'
→ Ei
→0
We now let the new eigenkets and eigenvalues be represented by the following power
series:
It is assumed that successive terms of this power series gets smaller and as a result the
series converges. Substitution of the new eigenkets and eigenvalues into H | i' >= Ei' | i' >
gives;
(H0 + h)| i' >= Ei' | i' > Substitution of the power series for |i'> and Ei' yield
(
= Ei + Ei1 + 2
Ei2 + .....+ n
)
Ein + ... {| i > + | i1 > + 2
| i2 > + ....+ n
| in > +..}
644 40 −Order
7444 8 6444444 474444444
1st− Order
8
{
{H0 | i > − Ei | i >} + H0 | i1 > +h | i > −Ei | i1 > − Ei1 | i > + }
2
144
{
H0 | i2 > + h | i1 > − Ei2 | i > −Ei1 | i1 > −Ei | i2 > +
4444444 424444444444 3
} 3
{...} + ......= 0
2nd − Order
For the above equation to be valid each of the terms in brackets must separately equate to
zero. The first and second terms give us the 0th order and 1st order terms respectively:
The first order equation can be solved by noting that the eigenket | i1 > can be expressed as
a linear superposition of the unperturbed eigenkets |i>:
| i1 >= ∑ a1 j | j >
j
H0 ∑ a1 j | j > + h | i >= Ei ∑ a1 j | j > + Ei1 | i >
j j
The first order energy correction term is therefore given by the matrix element <i|h|i> taken
between the unperturbed eigenkets |i>.
SCE3337: QMIII 5
R.T. Sang
We can now evaluate what the perturbed ket vector |i'> to first order will be. The first order
correction to the eigenket |i> is given by
| i1 >= ∑ a1 j | j >
j
substitution back into the first order term yields
The coefficients a1 j are evaluated from equation (1) derived in the previous section, hence
for the case when k i (i.e. non-degenerate case) then:
67=08
< k | h |i >
⇒ a1k =
( Ei − Ek )
k is a dummy variable hence we can call it anything so make it j. Substitution of this
expression into equation(2) gives;
< j |h |i >
| i' >=| i > + ∑ (E − E ) |j>
j≠i i j
States that are close together will have a large amount of mixing. We can see this
mathematically as the denominator in the above expression will get large. As the states
1
become separated the mixing will become weaker since ≈ 0.
( Ei − E j )
#
It is valid to call this effect an interference as both the amplitudes and the phases of the matrix elements
play a role.
SCE3337: QMIII 6
R.T. Sang
The extension to second order is relatively straight forward by following the same
procedure as for the first order corrections.
H0 ∑ a2 m | m > + h ∑ a1 j | j > = Ei ∑ a2 m | m > + Ei 2 | i > + Ei1 ∑ a1 j | j >
m j m j
∑a 2m < k |H0 | m > + ∑ a1 j < k | h | j >= Ei ∑ a2 m < k | m > + Ei1 ∑ a1 j < k | j > +Ei 2 < k | i >
m j m j
Recall that the first energy correction term is given by Ei1 =< i | h | i > thus
= ∑ a1 j <i | h | j >
j≠i
SCE3337: QMIII 7
R.T. Sang
< k | h |i >
Recall that a1k = . Upon substitution into the above equation yields:
( Ei − Ek )
< j | h | i >< i | h | j >
Ei2 = ∑
j≠i ( Ei − E j )
Therefore the second order correction to the energy is
Ei2 = ∑
j≠i ( Ei − E j )
The second order correction to the eigenket is messy to derive (see problem sheet 3) but
straight forward using the same procedure as the first order correction and is given by
| i' >=| i > + | i1 > + | i2 > Expanding i1 and i2 in basis sets yields
Therefore
Exercise: Derive this expression for the corrected eigenvector to second order.
Beyond second order, perturbation theory is seldom used as it becomes very messy
quickly.
SCE3337: QMIII 8
R.T. Sang
You may recall from second year that if we apply an external uniform magnetic field B to
an atom with a magnetic moment then it will experience a perturbation. The magnetic
moment is given by:
= − gJ B
J
J
h
If the magnetic field is weak (<104 Gauss) the total angular momentum J=L+S is a good
quantum number the perturbation to the system Hamiltonian is therefore given by
h=− • B = gJ B
J•B
J
h
The first order energy shift is then
Ei1 = gJ B
B < JmJ | J Z | JmJ >
h
Ei1 = gJ B
BmJ h < Jm J | JmJ >
h
⇒ Ei1 = gJ B BmJ
m J = -1 mJ = 0 mJ = 1
3
6 P1
Ei
253.7nm Radiation
1
6 S0
No Perturbing Field
SCE3337: QMIII 9
R.T. Sang
If we now turn on a weak magnetic field we can calculate the perturbation of the energy
levels of the 61 S0 and 6 3 P1 states due to the field. To do this we need to determine the
Landè g factor which is given by
3
For the 61 S0 state gJ=1 and for the 63 P1 state g J = . Now in the ground state the only mJ
2
value is 0 hence this level is not perturbed since the perturbed energy is proportional to mJ.
In the case of the 63 P1 state the degenerate energy levels will be split into three
nondegenerate states by the amount
3
Ei1 = B B
2
Thus the perturbed energy levels of the 63 P1 state will be
Ei' = Ei + Ei1
mJ = 0 Ei1
3
6 P1
m J = -1 Ei1
Ei
253.7nm Radiation
1
6 S0
B-field ON
It should be noted that we have used the non-degenerate perturbation theory even though in
the 63 P1 state the energy levels were degenerate in mJ. In this special case it is ok to do so
as the operator JZ has a definite value whether there is a perturbation or not as a result there
is little mixing of these states.
SCE3337: QMIII 1
R.T. Sang
Lecture 6
6 . 0 Perturbation Theory for Degenerate States
In previous lectures we had assumed that there was no degeneracy among the perturbed
states. However in practice it is often encountered that there are several
states that have the same energy, that is they are degenerate.
The effect of a perturbation is usually to lift the degeneracy (this was seen in the
Zeeman effect example last lecture) in the first order correction.
There are problems associated with the application of time independent perturbation
theory if only part of the degeneracy is lifted in the first order correction. This is seen
when one tries to apply second order corrections to degenerate states, recalling the
equation to second order:
< k | h | i >= 0
Therefore the second order energy correction term yields,
Consider two states |i> and |k> that are nearly degenerate, with all other states well
removed from these states. We also suppose that the matrix element <k|h|i> 0. The first
order correction to the ket |i> is given by
We now substitute in to the eigenvalue equation (H0 + h)| n' >= E'n | n' > :
We now write the perturbed energy term as E'n = En + which is just the sum of the
unperturbed energy plus the correction term;
⇒ ∑C [< k | h | j > −
nj kj ]= 0
j
h11 − h12
=0
h21 h22 −
1
(h + h22 ) ± ( h11 − h22 ) − 4h12 h21
2
=
2 11
Therefore this system with two degenerate levels, the correction term takes on two
values. As an example let h11 = h22 = A and h12 = h21 = B , then from the above equation
= A ± B . We can determine the coefficients Cnj by reconsidering the equation
∑C [< k | h | j > −
nj kj ] = 0 then
j
A− B C1 0
= using the solution =A+B we get
B A − C2 0
− B B C1 0
=
B − B C2 0
− BC1 + BC2 = 0
⇒
BC1 − BC2 = 0
⇒ C1 = C2
1
⇒ | >= C1
1
From normalisation ∑C i
2
= 1 thus
i
C1
(C1 C2 ) = C1 + C2 = 1
2 2
C2
For C1 = C2 then
1
C12 + C12 = 1 ⇒ C1 =
2
SCE3337: QMIII 4
R.T. Sang
1
| >= (|1 > + | 2 >) Called a symmetric wavefunction
2
1
| >= (|1 > − |2 >)
2
As an example of one of the uses of degenerate perturbation theory we will calculate the
electric field Stark splitting of the n=2 level of hydrogen by a constant DC electric field
E which we will assume points in the Z direction.
The n=2 level has four degenerate levels which all have the form
nlm = Rnl (r)Θlm ( )Φ m ( ) :
1 r r
= 3
2 − exp −
4 2 a0 a0 2a0
200
1 r r
210 = exp − cos
3
4 2 a0 a0 2a0
1 r r ±i
= exp− sin e
2a0
21±1 3
4 2 a0 a0
where (x,y,z) = (rsin cos ,rsin sin ,rcos ) transform into the usual spherical
coordinates.
The perturbation Hamiltonian is given by the dot product between the electric dipole of
the atom and the electric field vector E:
h = −E• D
D is the dipole moment of the atom and is given by D=qd, where q is the charge (in our
case q=-e) and d = zkˆ = r cos kˆ . Hence our perturbation operator h is given by
For example the matrix element < 2l' ml' | z |2 lml > must have odd parity since
changing sign of the coordinate makes z -z.
Recall that the definite integral over symmetric limits of an odd function = 0 (c.f.
a
∫ xdx = 0 ) therefore the diagonal matrix elements must be equal to zero since the
−a
integral < 2lml | z | 2lml > is over symmetric limits.
The parity of the wavefunction is defined by the factor (-1)l. Therefore is follows that
for non-zero matrix elements:
Furthermore ml=ml' for a non-zero integral. This is seen when considering the form of
the wavefunction above, where the m component is explicit in the function
Φ m ( ) = e im .
< 2l' ml' | r cos | 2lml >= ∫ Rnl* (r)r 3 Rnl' dr∫ Θ*lm l ( )sin Θl' ml ' d ∫ e e d
r 0
{∫ .......drd } ∫ e
2
−i(m l − ml ' )
= d
0
644=0,
47 444
always
8
{∫ .......drd } ∫ cos(m
2 2
= d3 − i ∫ sin(m − m ) d
4−
2m )4
144 44 l l' l l'
0 =1⇒ ml = ml ' 0
=0⇒ ml ≠ ml '
There are four solutions to this equation 1,2 =±3a0 eE, 3,4 =0. Therefore the
degeneracy is partially lifted by the electric field.
We now evaluate the perturbed wavefunctions which require solutions to the equation:
− −3a0 eE 0 0 C1
−3a0 eE − 0 0 C2
= 0
0 0 − 0 C3
0 0 0 − C4
−3a0 eE −3a0 eE 0 0 C1
−3a0 eE −3a0 eE 0 0 C2
= 0
0 0 −3a0 eE 0 C3
0 0 0 −3a0eE C4
1 1 0 0 C1
1 1 0 0 C2
⇒ −3a0 eE = 0
0 0 1 0 C3
0 0 0 1 C4
1 1 0 0 C1
1 1 0 0 C2
⇒ = 0 multiplying out gives
0 0 1 0 C3
0 0 0 1 C4
C1 + C2 = 0
C1 + C2 = 0
C3 = 0
C4 = 0
2 2
C1 + C2 = 1 so
1 1
C1 = and C2 = −
2 2
1 1
| 1 >= | 200 > − | 210 >
2 2
Similarly one can show using the same procedure that for =-3a0 eE
1 1
| 2 >= | 200 > + | 210 >
2 2
The eigenvalues and eigenvectors for the n=2 Stark shifted states of hydrogen are given
by
=+3a0 eE 1 1
| 1 >= | 200 > − | 210 >
2 2
=-3a0 eE 1 1
| 2 >= | 200 > + | 210 >
2 2
=0 | 3 >=|211 >
=0 | 4 >=| 21− 1 >
SCE3337: QMIII 8
R.T. Sang
Energy
E
|200> |210> |211> |21-1>
Before Perturbation
1 1
| 1 >= | 200 > − | 210 >
2 2
E + 3a0 eE
3a0 eE
E
|200> |210> 3a0 eE
E − 3a0eE
1 1
| 2 >= | 200 > + | 210 >
2 2
With perturbing electric field
SCE3337: QMIII 1
R.T. Sang
Lecture 7
7 . 0 Time Dependent Perturbation Theory
In lectures so far we have covered perturbations to the Hamiltonian that are time
independent. We now develop an approximation technique to allow us to handle
perturbations that have a time dependence. We write the time dependent perturbation
Hamiltonian as
H(t)=H0 +h(t)
Where H0 is the time independent unperturbed Hamiltonian and h(t) is a perturbation which
is time dependent.
h(t) << H0
H0 | i >= Ei | i >
This equation leads to stationary states (time independent) which have eigenvalues given by
Ei. It should be noted that this equation is formally a solution of the Schrödinger equation
when the potential is time independent.
h2 2
H=− ∇ + V(r)
2m r
Thus
h2 2 ∂
− ∇ r | Ψi (r,t) > + V(r) | Ψi (r,t) >= ih | Ψi (r,t) >
2m ∂t
SCE3337: QMIII 2
R.T. Sang
We now try separation of variables by splitting the wavefunction into time independent and
time dependent part such that
h2 2
− ∇ | (r) >| i (t) > +V(r)| (r) >| i (t) >= ih | (r) >| i (t) >
2m r i i i
t
h2
⇒ − | i (t) > ∇ 2r | i (r) > +V(r) | i (r) >| i (t) >=| i (r) > ih | i (t) >
2m t
1
we now multiply both sides of the equation by which gives
| i (r) >| i (t) >
1 h2 2 1
− ∇ | (r) > + V(r)| (r) > = ih | i (t) >
| i (r) > 2m r | i (t) >
i i
t
Since both sides are only dependent on either | i (r) > or | i (t) > both sides of the
equation are equal to constants:
1 h2 2 1
⇒ − ∇ | (r) > + V(r)| (r) > = C = ih | i (t) >
| i (r) > 2m r | i (t) >
i i
t
h2 2
− ∇ | (r) > +V(r)| (r) >= C | (r) >
2m r i i i
We have recovered the time-independent SE and is just the usual time independent
eigenvalue equation, hence
C = Ei
SCE3337: QMIII 3
R.T. Sang
1
ih | i (t) >= C
| i (t) > t
1
⇒ ih | i (t) >= Ei
| i (t) > t
−i
⇒ | i (t) >= E | (t ) >
t h i i
d | i (t) > −i
⇒ = Ei dt
| i (t) > h
Integrating reveals
d | i (t) > −i
∫ | i (t) >
= Ei ∫ dt
h
−i
⇒ ln (| i (t ) > ) = E t + C'
h i
−i E t −i E t
h i h i
⇒ | i (t) >= Ae =e
The wavefunction for the stationary states (time independent potential) is therefore given by
− i E t
h i
| Ψi (r,t) >=| i (r) >| i (t) >=| i (r) > e
Consider the wavefunctions of the full Hamiltonian H(t). Since | Ψi > form a complete
orthonormal set thus we can expand the wavefunctions as
⇒ ∑ a (t)( Hn 0 + h(t)) | Ψn (r,t) >= ih ∑ an (t) | Ψn (r,t) >
t n
n
⇒ ∑ a (t)( Hn 0 + h(t)) | Ψn (r,t) >= ih∑ a˙n (t) | Ψn (r,t) > +ih ∑ an (t)
t
| Ψn (r,t) >
n n n
∑ a (t) H
| Ψn (r,t) > − ih | Ψn (r,t) > = ih ∑ a˙n (t) | Ψn (r,t) > − ∑ an (t)h(t)| Ψn (r,t) >
t 44443
n 0
n 1444444244 n n
This term = 0 for the unperturbed wavefunction since
H 0 |Ψ n (r,t )>=ih |Ψn (r ,t )>
t
∑ a (t) < Ψ
n k (r,t)| h(t)| Ψn (r,t) > −ih∑ a˙ n (t) < Ψk (r,t)| Ψn (r,t) >= 0
n n
We now separate the spatial and time dependence in the wavefunction so that the bra and
ket vectors above are defined as
Ek t
i
< Ψk (r,t)| = e h
< ψ k (r)|
En t
− i
| Ψn (r,t) >= e h
| ψ n (r) >
∑ a (t) <
− ih∑ a˙n (t)e
h h
n k (r)|h(t)| n (r) > e kn =0
n n
[ Ek − E n ]t
i
∑ a (t) <
h
⇒ n k (r)|h(t)| n (r) > e = iha˙ k (t)
n
∑ a (t)h (t)e( )
= ih hkn (t) =< (r) >
i kn t
n kn ak (t) where k (r)| h(t)| n
n t
SCE3337: QMIII 5
R.T. Sang
This is a set of coupled first order differential equations, one for each ak(t) which determine
the an (t) coefficients. These equations are not able to be solved exactly since all
of the an (t) coefficients are related to only the derivative of the kth coefficient. Notice that if
the perturbation is zero then a˙ k (t) = 0 ,therefore ak (t) must be a constant. This suggests that
provided the perturbation is small, the coefficients change slowly. As a first approximation
we will assume that the coefficients an (t) on the LHS of the equation are constant.
Suppose at t=0, the system is in some state | Ψj (r,t = 0) > this requires that aj(0)=1 and
an (0)=0 for n j. This can be interpreted as the system is totally in the j state at the start.
(i )
aj (t) = ∑ an (t)h jn (t)e jn t
ih
t n
After t=0, since the perturbation is weak aj(t) 1 still and all other coefficients an (t) ≈0
hence only this term will contribute to the sum therefore the right hand side of the equation
reduces the sum to
(i jj t )
RHS= a j (t)hjj (t)e ≈ h jj (t)
{
=1
Therefore
ih aj (t) = h jj (t)
t
−i
a j (t) = h (t)
t h jj
Integrating gives
−i
t t
d
∫ dt
a j (t)dt = ∫ hjj (t)dt
h 0
0
−i
t
t
i
⇒ a j (t) = 1 − ∫ h jj (t)dt
h0
SCE3337: QMIII 6
R.T. Sang
d
ak (t) = ∑ an (t)hkn (t)e(
i kn t )
ih
dt n
But since all an (t) 0 except aj(t) then only one term in the sum of the RHS significantly
contributes
d (i kj t )
ih ak (t) = a j (t)hkj (t)e
dt {
≈1
d i (i kj t )
⇒ ak (t) ≈ − hkj (t)e
dt h
Integrating this expression:
t t
d i (i )
∫ ak (t)dt ≈ − ∫ hkj (t)e kj t
dt
0
dt h0
t
i (i )
3 ≈ − h ∫ hkj (t)e
kj t
ak (t) − a1
k2(0) dt
≈0 0
t
i (i )
ak (t) = − ∫ hkj (t)e kj t
dt
h0
SCE3337: QMIII 7
R.T. Sang
As an example we will consider a step function perturbation which is turned on at t=0 and
remains constant thereafter. The matrix elements of the perturbation are therefore constants
and can be taken outside the integral:
i
a j (t) = 1 − h jjt
h
The other coefficients are
t
i (i )
ak (t) = − hkj ∫ e kj t
dt
h 0
t
i −i ( i kj t )
⇒ ak (t) = − hkj e
h kj 0
i −i ( i )
− 1
kj t
⇒ ak (t) = − hkj e
h kj
hkj ( i )
− 1
kj t
⇒ ak (t) = − e
h kj
The probability of finding the system in the state | Ψk > with the system starting at time t=0
2
in state | Ψj > is ak (t) . Then for k j we find
2
h e ( i ) (i )
*
− 1 • e − 1
2 kj t kj t
ak (t) = kj
h kj
2
h e (i ) (− i )
− 1 • e − 1
kj t kj t
= kj
h kj
2
h e (i ) (− i ) (i ) ( −i )
+ 1
kj t kj t kj t kj t
= kj e −e −e
h kj
{ }
2
h 2 − e( i kj t ) (− i kj t )
= kj +e
h kj
SCE3337: QMIII 8
R.T. Sang
[ { }]
h
= kj 2 − 2cos( kj t)
h kj
t
now cos(2A ) = 1 − 2sin 2 ( A) ⇒ cos( t ) = 1 − 2sin 2 kj
kj
2
2
h 2 kjt
= kj 2 − 21 − 2sin
h kj 2
2 kjt
sin
2 2
2
⇒ ak (t) = 2hkj
h2 kj2
Thus the probability of finding the system in state | Ψk > oscillates at an angular frequency
which corresponds to the transition frequency as shown in the figure below:
0.8
0.6
0.4
0.2
x
-6 -4 -2 0 2 4 6
0
SCE3337: QMIII 1
R.T. Sang
Lecture 8
7 . 2 The Harmonic Perturbation and Fermi's First Golden Rule
We now look at one of the more important time dependant perturbations which is an
oscillatory perturbation. These types of perturbations are important as nearly all matter-
electromagnetic radiation interactions can be thought of as an oscillatory perturbations
produced by an electric or a magnetic field. We have already seen the effect of DC electric
and magnetic fields we now consider the effect of AC fields.
h(t) = V cos( t )
ei t + e− i t
h(t) = V cos( t ) = V
2
)e( )
iVkj
∫ (e + e −i
i kj t
⇒ ak (t) = − i t t
dt
2h 0
iVkj t i ( + )t ( kj − )t
2h ∫0
i
⇒ ak (t) = − +e
kj
e dt
e i( kj + ) t e ( kj )
t
i − t
iV
⇒ ak (t) = − kj +
2h i( kj + ) i ( kj − )
0
iV ei ( kj + )t
−1 e(
i kj − )t
− 1
⇒ ak (t) = − kj +
2h i ( kj + ) i( kj − )
We notice that if kj ≈ then the term with the negative sign dominates all the other terms.
This process corresponds to absorption of a photon from the perturbing field with energy
h kj exciting the system from the lower state | Ψj > to the higher energy state | Ψk > , the
transition has the frequency:
SCE3337: QMIII 2
R.T. Sang
Ek − E j
= kj =
h
In this case the second term in the brackets dominates such that
Vkj e i( kj − )t
− 1
ak (t) ≈ −
2h ( kj − )
Vkj e i ( ∆ )t − 1
⇒ ak (t) = −
2h ( ∆ )
∆ i ∆2 t ∆
−i
t
i −e 2
V t
2 e
= − kj e
2h (∆ )
i ∆ t ∆
−i
t
Vkj 2i i ∆2 t e 2 − e 2
=− e
2h ∆ 2i
∆ t
2
2 Vkj 2 2 sin 2
ak (t) = 2 t ∆ t
4h
2
One can show that a similar expression can be found for the stimulated emission term. You
sin x 2
may recognise that the factor in the brackets has the form which is identical to the
x
intensity profile of the single slit diffraction pattern as shown below:
0.8
0.6
0.4
0.2
x
-8 -6 -4 -2 0 2 4 6 8
0
∆ t
− ≤ ≤
2
2h∆ t
−2h ≤ ≤ 2h ⇒ −h ≤ ∆Et ≤ h
2
Note that h is Planck's constant. This sets an upper limit on the energy difference
multiplied by the time t which has elapsed since the field was turned on:
∆Et ≤ h
This is just the uncertainty principle! In effect it says that the longer that the perturbation is
on, the more nearly ω =ω kj .
SCE3337: QMIII 4
R.T. Sang
This is an interesting result, since it says that for near resonant absorption or emission the
probability of the process is proportional to the time squared, however experimentally
it is observed that the probability is directly proportional to the time.
The reason for this difference is due to the fact that even under ideal conditions there is an
intrinsic width to the energy levels for an excited atom (due to the uncertainty principle)
which causes the probability to be proportional to t and not t2 .
It is assumed that even for a perfectly monochromatic source of excitation with a single
frequency such that ∆ = − kj means that will have a range of values. Then to
get the probability of getting a transition from state | Ψj > to | Ψk > is given by the
integration over (large range):
∆ t
2
Vkj 2 2 ∞ sin 2
W(t) = 2 t ∫ ∆ t d(∆ )
4h ∆ =−∞
2
∆ t 2
Taking t to be fixed we let X = , hence d(∆ ) = dX and substituting into the above
2 t
equation gives
Vkj 2 2 2 ∞ sin( X ) 2
t X∫−∞ X
W(t) = 2 t dX
4h
1442443
Vkj 2
⇒ W(t) = 2 t
2h
Lecture 9
8 . 0 Selection Rules
The selection rules allow us to determine which optical transitions are allowed and which
are forbidden. They can be determined by considering the matrix element of the electric
dipole operator.
V = −E• D
where D is dipole operator and is given by D=-er. Thus the matrix element is
The radial matrix element can be expanded in terms of spherical polar coordinates, since the
relation between the Cartesian coordinates and the spherical polar coordinates are defined
by
θ
z r
φ
x
x = r sin cos
y = r sin sin
z = r cos
Where we have separated the time dependence of the wavefunctions. We now separate the
wavefunction as done previously:
j (r) = Rn j l j (r )Θ lj m j ( )Φm j ( )
Then
∞ 2
∫Φ ∫e
−imk
cos Φ m j d =
* im j
mk cos e d
0 0
2
i( m j −m k )
[e + e − i ]d
1
= ∫e
i
2 0
2
1 i( m j −m k +1) i ( m j − mk −1 )
=
2 ∫e +e d
0
Therefore the only non-zero matrix element for the z component is for m=0.
Thus the selection rules for the magnetic projection quantum numbers for dipole allowed
transitions are:
∆m = 0,±1
SCE3337: QMIII 3
R.T. Sang
Consider the spin of the electron. The total wavefunction may be written as the product of
the spatial and spin terms:
= nlml s
The electric dipole operator does not act on the spin wavefunction (it does with magnetic
fields, but very weakly with time varying electromagnetic waves) so we can write:
< >= e ∫ * *
k (r) sk | er | j (r) sj k (r) sk r j (r) sj d
= e∫ ∫ (r)d = e ∫
* * *
sk sj d k (r)r j k (r)r j (r)d sk s j
Therefore
∆s = 0
We can use the parity of the wavefunction to determine the L selection rule. Consider the
effect of changing the sign of the x,y,z coordinates of the one electron atom wavefunctions.
In spherical polar coordinates changing the sign is equivalent to:
r → r, → − , → +
Thus
Under Parity
ψ nlml (r,θ,φ) → ψ nlm l (r, π − θ,π + φ) = (−1)l ψ nlml (r,θ,φ)
Transformation
The parity of the wavefunction is determined on whether l is even or odd. This result is
applicable for all bound or unbound eigenfunctions for any potential that is spherical in
form.
We have shown previously that the integration or r which is an odd function requires that
the wavefunctions < k (r)| and | j (r) > must have different parity so as to yield a
symmetric definite integral of an even function (see the section on the DC Stark shift in
hydrogen). If < k (r)| has even parity then lk must also be even. If | j (r) > has odd
parity then l j must be odd. Clearly then lk − l j = ∆ l = ±1,±3,±5 etc and ∆l = ±2,±4,±6.....
must be excluded. Thus l can only change by odd integer values.
SCE3337: QMIII 4
R.T. Sang
Consider the case when the atom starts of in the j=1 and mj=-1 states and decays via a
photon to the j=0, mj=0 then if the photon which is emitted travels along the z direction it is
Right Hand Circularly (RHC) polarised and carries away one unit of angular
momentum - h .
The maximum amount of angular momentum that a photon can carry is h , so this means
that the total change in angular momentum is ∆j = ± 1. We can see that photons can have
two angular momentum states, it is also possible to form a superposition state, the
wavefunction of which is given by:
1 1
LHC + RHC
2 2
This corresponds to linearly polarised light. As such an atom emitting such a photon must
exhibit no change in angular momentum, thus ∆j = 0
Hence the selection rules for the total angular momentum j are
∆j = ± 1,0
Clearly if ∆j = ± 1,0 then only ∆l = ±1 are allowed for single photon emission since s=0.
Thus the selection rule for l is:
∆l = ±1
Also note that j=0 j=0 transitions are not possible since no angular momentum will be
transferred.
∆j = ± 1,0
∆l = ±1
∆ml = ±1,0
∆s = 0
Lecture 10
9.0 The Einstein A and B coefficients
The basic model of the interaction process between electromagnetic radiation and atoms can
treated by a model first introduced by Albert Einstein. This theory was
phenomenonolgical in nature and makes no explicit use of quantum mechanics except
that the energy levels of the atoms are assumed to be quantised and it is convenient to
regard the electromagnetic field as photons.
Suppose that we have N identical atoms in a gas and each atom has only two energy levels
E1 and E2 (we will label the states |1> and |2>) with E1 < E2 as shown below:
N2 E g 2
2
|2>
B 21W A21
-hω Spontaneous Emission
B 12W
Absorbtion |1>
N1 E g 1
1
Stimulated Emission
• Absorption
Suppose that the energy density of the radiation passing through the gas is W( ) where
W( ) is the average energy density, that is the energy per unit volume per unit bandwidth.
An atom in the lower energy state |1> can make an up going transition to the higher energy
state |2> by absorbing a photon of energy h = E2 − E1 . We assume in this case that the
transition of this type is proportional to the energy density with a constant of
proportionality of B 12 . The upward transition probability is B12 W .
• Stimulated Emission
If an atom is in the state |2> and another photon that is resonant with the transition passes
this atom, the presence of this photon can induce the atom to emit an identical type of
photon which will then take the atom back to state |1>. This rate is proportional to the
energy density of the incident radiation, with a constant of proportionality of B 21 . Thus the
stimulated emission probability from state |2> to |1> is B21 W .
QMIII: SCE3337 2
R.T. Sang
• Spontaneous Emission
Consider an atom in the higher energy state |2>. There is a finite probability given by A 21
that the atom will pass from this state to the state |1>. In doing so it will emit a
photon which has a random direction and polarisation with energy hω
(hence the term spontaneous emission). Actually this process was a major problem for
quantum theory, the explanation of which had to wait until the advent of Quantum
Electrodynamics (QED). This process can be explained as follows: the states |2> and |1>
are eigenstates of the Hamilitonian and for an isolated system, once it is in this energy
eigenstate, if it is unperturbed should remain there forever. QED relies on the basis that
even when there are no photons around, the electromagnetic field still has a zero point
energy (recall your lectures on the SHO!) and it is this field which induces the atom to
decay. The so called vacuum fluctuations are an infinite virtual supply of zero point
energy photons which cover all of the frequency range of the EM spectrum, which induces
the atom to emit photons (like stimulated emission). You might actually think this is a crazy
idea since that means the vacuum field has an infinite energy and this still remains one of
the unsolved problems of QED. But there have been experiments such as the Casmir Effect
which shows conclusively that this zero point energy exists.
We now consider the influence of these three processes on the energy level populations, N 1
and N 2 .
The total number of atoms is given by N which is the sum of the populations:
N = N1 + N2
dN1 dN
=− 2
dt dt
Now the rate of change of the population in |1> is:
dN1
= (rate at which atoms enter the lower state) - (rate at which they leave)
dt
= (Stimulated Emission Rate + Spontaneous emission rate) - (Absorption
Rate)
= N2 B21 W + N2 A21 − N1 B12 W
This rate equation holds for the general case of electromagnetic radiation interacting with
these two level atoms. Now consider the special case of thermal equilibrium. In this case,
the population levels are constant, thus our rate equation becomes:
Hence
For thermal equilibrium with no external radiation field on the gas, the relative number of
atoms in various energy states is given by the ratio of the Maxwell-Boltzman distributions
for each level:
E
− 1
kT
g1 kT
h
N1 g1e
= E
= e
N2 − 2
kT g2
g2 e
A21
W=
g1 h
kT
g e B12 − B21
2
1
W=
g1 B12 h
kT B21
g A e −
2 21 A21
Now this expression for the energy density must be consistent with the Planck's law for
the radiative energy distribution of a body in thermal equilibrium which is given by :
h 3
d 1
Wd = For single ⇒ W=
c h
c kT
h
2 3 2 3
kT
e − 1 e − 1
h 3
Equating the our derived expression for the energy density with the Planck relation gives:
1 1
=
g1 B12 h
kT B c kT
2 3 h
g A e − 21 e − 1
2 21 A21 h 3
Also
2 3
g1 B12 c
=
g2 A21 h 3
2 3
c h 3
But B21 = 3 A21 ⇒ A21 = 2 3 B21
h c
Substitution into our expression yields
B12 g2
=
B21 g1
As shown the three Einstein coefficients are inter-related. It can be seen that
without introducing the stimulated emission process, consistency between the Planck
formula and the Einstein expression could not be achieved.
Furthermore it must be stressed that although the relationship between the Einstein
coefficients have been derived for the thermal equilibrium case, they hold generally since
the coefficients are independent of the magnitude of the energy density or the temperature.
It should also be noted that for thermal equilibrium, the radiative energy density W is
distributed isotropically in space. This is of course not so with a light beam. The
relationships do remain valid in systems such as a gas or a fluid in which the atoms or
molecules have random orientations so that the interaction within the gas as a whole is
isotropic. However in solids, the constituent atoms or molecules may be locked in a
common orientation. In this case, the bulk material may have quite anisotropic optical
properties.
One point of interest is the ratio of power emitted in the spontaneous emission process
compared to the stimulated process:
A21
W = Ratio of spontaneous to stimulated emission
B21
We need to look in detail at the form of the energy density W which is dependent on the
frequency of the radiation. Recall for the case of thermal equilibrium that the energy density
is given by Planck's Law for a frequency interval of → + d is:
h 3
d
Wd =
c h
2 3
kT
e − 1
h
For room temperature at approximately 300K then the ratio ≈ 1, the corresponding
k BT
frequency is around 6x1012 Hz with 50µm which is in the infrared part of the spectrum.
QMIII: SCE3337 5
R.T. Sang
h
For frequencies where << 1 (that is for longer wavelengths such as microwave,
k BT
radiowave etc) then the energy density is large since the exponent in the denominator is
close to 1 as such
This means that for frequencies less than 6x10 12 Hz, the thermally induced
spontaneous rate is much less than the stimulates emission process.
h
In the case where frequencies are larger than 6x1012 Hz i.e. >> 1 then the exponential
k BT
in the term for the energy density dominates and as such the energy density is small thus
As such for frequencies that are greater than 6x10 12 Hz, the thermally
induced spontaneous emission rate is much larger than the stimulated
process.
So far in this section of the course, we have treated the Einstein A and B coefficients as
parameters that are experimentally determined. But, in fact, the absorption and emission of
the radiation can be calculated by using Time Dependent Perturbation Theory.
Consider now an idealised atom consisting of two energy levels |1> and |2> with transition
frequency ω12 . We now apply a harmonic perturbation on this system with radiation that
has frequency ω and is near resonant with the transition frequency as shown below.
|2> E 2
-hω
ω12
|1> E 1
The transition probability is given by Fermi's first Golden Rule (see section 7.2 of the
lecture notes):
Vkj2
W(t) = 2 t
2h
QMIII: SCE3337 6
R.T. Sang
Recall that in the harmonic perturbation case the perturbation had the form h = V cos t .
This term more accurately should be written as h = V cos( t + k • r) which means that
there is a spatial dependence to the perturbation (without this term we have just assumed
that the perturbation is at the origin). This perturbation Hamiltonian describes the
interaction between an atom and the electric field of the light. Consider the electromagnetic
wave to be polarised along the x direction and propagates along the z direction. The atom
consists of a nucleus and is surrounded by n electrons, over the dimensions of the atom,
which is in the order of 10-10 m.
For optical frequencies > 1015 Hz then k • r =kz <<1. Thus the spatial variations of the
electric and magnetic fields are virtually constant over the atomic dimension and hence the
electrons in the atom experience the same electromagnetic fields. This approximation is
known as the Dipole Approximation.
h = −E • D
= −E 0 • D cos t (we have dropped the spatially dependent term)
Vkj = − < k | E0 • D | j >= −E0 • < k | D | j > with E0 parallel to the x-direction.
Substituting this back into Fermi's First Golden Rule gives the transition probability of
going from state |1> to state |2> is
< 1| Dx | 2 >
2
2
W12 (t) = E0 t
2h 2
Now the energy density of an electromagnetic wave is given by
1 2
W= 0 E0
2
Hence the probability of transition can be related to the energy density such that;
2
W12 (t) = < 1| Dx | 2 > Wt
0h
2
It follows therefore if we divide this equation by t we will get the transition rate:
2
W12 (t)/ t = 2 < 1| Dx | 2 > W
0h
QMIII: SCE3337 7
R.T. Sang
2
B12 W = < 1| Dx | 2 > W
0h
2
2
B12 = < 1| Dx | 2 >
0h
2
What about a gas or molecules or atoms? In this case all of the electric dipoles will be
orientated in different directions at any particular time in space in random directions.
Given that θ is the angle between the dipole and the E field then
2
< 1| Dx | 2 > = < 1| D |2 > cos2
2
Thus we need the average value of cos 2 integrated over the solid angle of a sphere which
is given by
1
cos 2 =
3
Exercise: Show this
B12 g2 g1 g1
B21 = 2 < 1| D | 2 >
2
= ⇒ B21 = B ⇒
B21 g1 g2 12 g2 3 0 h
QMIII: SCE3337 8
R.T. Sang
This approach does not yield the Einstein A coefficient directly but we can get this from the
relationship between the A and B coefficients derived using the rate equation model:
h 3
A21 = 2 3 B21
c
Therefore upon substitution into the relationship for the A coefficient reveals;
3
g1
A21 = < 1| D | 2 >
2
g2 3 hc 3 0
QMIII: SCE3337 1
R.T. Sang
Lecture 11
1 0 . 0 Optical Excitation of Atoms
Let the beam be switched on at time t=0 and all atoms are in their ground state. We want to
find the number of atoms in their excited states at some time t later. Recall the rate equation
for a two level system (lecture 10) was given by
dN1 dN
= − 2 = N2 A + ( N2 − N1 )BW
dt dt
If the total population is N then
N = N1 + N2 ⇒ N1 = N − N2
dN2
− = N2 A + (2 N2 − N )BW
dt
= N2 ( A + 2BW ) − NBW
dN2
⇒ −
dt
This differential equation can be solved by separating the variables.
dN 2
− = dt
N2 ( A + 2BW ) − NBW
let
c1 = A + 2BW
c2 = NBW
QMIII: SCE3337 2
R.T. Sang
Integrating yields
N2 t
dN2 dx 1
−∫
N2 c1 − c2 ∫0
= dt recall the standard integral ∫ ax + b = a ln (ax + b)
0
Hence
N
1 2
− ln( N c
2 1 − c )
2 = [t ]0
t
c1 0
ln ( N2 c1 − c2 ) − ln(−c2 ) = −c1 t
N c −c
ln 2 1 2 = −c1 t
−c2
N c − c
− 2 1 2 = e( 1 )
−c t
c2
Therefore
N2 =
c2
c1
{1 − e( 1 )
−c t
}
Substitution for c2 and c1 gives N 2 as
N2 =
NBW
A + 2BW
1−e (
− A + 2BW t )
{ }
Graphically the solution looks like:
0.7
0.6
0.5
Population
0.4
N2
0.3
0.2
0.1
0
0 20 40 60 80 100
time
QMIII: SCE3337 3
R.T. Sang
It is interesting to look at this population for different ranges of the exponential in this
expression. For ( A + 2BW )t << 1. In this case we can Taylor expand the exponential:
e( ≈ 1− ( A + 2BW )t + H.O.T .
− A +2 BWt )
N2 =
NBW
A + 2BW
{ [
1 − 1− ( A + 2BW )t ]}
( A + 2BW )t
NBW
⇒ N2 =
A + 2BW
⇒ N2 = NBWt
We can determine the long time dependence of the population in the excited state quite
easily as the exponential term becomes very small and as such negligible:
NBW
N2 ≈
( A + 2BW )
We could also obtain this solution for long times by solving the d.e. for N 2 at steady state.
Recall that in the visible spectrum and for an ordinary light source (such as a fluorescent
lamp) that the spontaneous emission rate is greater that then stimulated rate i.e. A >> BW
as such from the above equation we deduce that
N2
<< 1 This means that most of the atoms are in the ground state (N 1 ).
N
If we used a much more powerful light source such as a laser beam then BW >> A
(stimulated emission processes are much greater that the spontaneous processes) using this
condition on our steady state solution gives
NBW N
N2 = =
2BW 2
The interpretation of this expression is that for powerful excitation, the limiting population
in the excited state is half of the total population, so no matter how many more extra
photons we put into the system the population in the excited state can not exceed 50% of
the total population. The effect is called Saturation and can be seen in the graph above. It
is simple to explain this effect as in the high light intensity case the simulated emission
effects balance the absorption effects (where we have assumed that spontaneous emission
is negligible in high intensity fields).
QMIII: SCE3337 4
R.T. Sang
If we now turned the incident light beam off, the excited atoms will return to their ground
state via spontaneous emission of photons. Then our rate equation for the excited
population becomes with W set to zero
= N2 ( A + 2BW ) − NBW = N2 A
dN2
−
dt
dN 2
⇒ = − N2 A
dt
This equation is separable and easily solved with the boundary condition that at time t=0
there are N20 steady state atoms in the excited state then
N t
dN
∫N 0 N22 = ∫0 −Adt
2
Therefore
N2 = N20 e − At
0.5
0.4
Population
0.3
N2
0.2
0.1
0
0 20 40 60 80 100
time
The radiation emitted in this process is called fluorescent radiation and the measurement of
this provides another experimental technique of measuring the Einstein A coefficient.
1
r =
A
QMIII: SCE3337 5
R.T. Sang
You have already encountered in your electromagnetism courses, the macroscopic theory of
absorption and dispersion of electromagnetic radiation passing through a medium. In this
lecture we will establish the connection between the Einstein A and B coefficients and this
macroscopic theory. This is important as we need to be able to make the quantum theory
converge with the classical macroscopic theory on a large scale.
Let us first re-visit some of the aspects of the macroscopic theory. Consider the picture
below:
Incident
Light Transmitted
Light
GAS
We have a light beam incident on a gas and this medium we consider as a dielectric material
where the polarisability is proportional to the applied electric field:
P= 0 E
n2 = 1 +
For a plane, monochromatic wave passing through a medium, the electric field vector is
given by
E(z,t) = E0 e− i( t − nkz )
E(z,t) = E0 e− i( t − kz )
• e i( n −1 ) kz = E0 (z,t)e i( n−1 )kz
The effect of the medium on the electromagnetic wave is given by the second exponential
term. For small susceptibilities i.e. <<1 then
1
1
n = (1+ )2 ≈1 + (this is found by Taylor expanding the function about the
2
point = 0)
is in general complex so it may be separated into its real and imaginary components such
that
= +i
QMIII: SCE3337 6
R.T. Sang
The first exponential is related to the refraction of the wave (Snell's Law) as it enters the
medium and the second term is related to the absorption of the radiation.
1
I=
2
0 c E(z,t)
2
Thus the intensity of the wave after passing through the medium is then
1
c E0 (z,t) e −
2
I= kz
0
2
Consider a gas of two level atoms once again. We know that if we shine near resonant light
on the medium that three process can occur which will effect the transmission intensity of
the incident light as depicted above.
QMIII: SCE3337 7
R.T. Sang
1) Absorption will reduce the transmission intensity the rate of which is proportional to
∝ BW
2) The stimulated emission process can be thought of a producing a carbon copy on
the incident photon which travels in the same direction with identical polarisation etc and
adds in phase to the incident light beam. The rate of this process is proportional to
∝ BW .
3) The spontaneous emission process is due to the vacuum electromagnetic field which
has all frequencies, polarisations etc stimulating the excited atom to emit a photon and as
such the atom emits into a 4π steradian solid angle thus reducing the intensity of the
transmitted beam. This rate of emission is proportional to ∝ A .
Suppose that now that we have our atoms in steady state and for simplicity we assume the
medium has identical degeneracies: g1 =g2 then our rate equation for the ground state
population is
dN1
= N 2 BW + N2 A − N1 BW = 0
dt
Thus
N2 A = ( N1 − N2 ) BW
This tells us that the rate at which photons are scattered out of the beam due to spontaneous
emission is balanced by the difference due to absorption and stimulated emission
processes.
Up to this point we have assumed that the frequencies of photons absorb/emitted by the
atoms have been a single frequency but in reality the emitted photons have a spread of
frequencies over which the atoms can absorb of emit photons. The spread of frequencies
under these gas slice conditions is mainly due to the Doppler effect (atoms moving at
different velocities will see the photons with different frequencies) as well as collisions of
the atoms in the gas. To account for this frequency spread we introduce a function F( )
which is the fraction of atoms with frequency :
∞
∫ F( )d =1
−∞
Therefore over some range of frequencies say → + d , the net rate at which light
energy is absorbed by N atoms is
( N1 − N2 )BWh F ( )d
QMIII: SCE3337 8
R.T. Sang
The energy contained in the slice of the medium of cross sectional area a and length dz in
the frequency range of → + d is
F
F=∑ dx
i xi i
Hence
I
I(z + dz) = I(z) + dz
z
The expression for the energy in the medium becomes:
I I
Energy in Medium = a I(z)d − I(z) + dzd = −a dzd
z z
That is the energy in the medium is equal to the change in energy of the incident light beam
over the volume V. Recall the rate of energy transferred to a medium by the rate equation
approach was
( N1 − N2 )BWh F ( )d
N 1 is the number of atoms in the ground state and N 2 the number atoms in the excited state.
In a slice of gas which has the volume V;
QMIII: SCE3337 9
R.T. Sang
adz
Number of ground state atoms = N1 ×
V
adz
Number of excited state atoms = N2 ×
V
Therefore our expression becomes
adz
Rate of energy change in volume V = ( N1 − N 2 ) BWh F ( )d
V
Equating this to the other expression for the energy in the medium gives
I adz
−a dzd = ( N1 − N2 ) BWh F( )d
z V
Therefore
I Bh
= -( N1 − N2 ) F( )W
z V
From the rate equation steady state expression
NBW
N2 =
( A + 2BW )
Then
NA
( N1 − N2 ) = N − 2N2 =
( A + 2BW )
Hence
I NA Bh
=- F( )W
z ( A + 2BW ) V
I
We also note that W = , thus our expression becomes
c
F ( )
I NA Bh I
=-
z A + 2B I V c
c
1 + 2 BI dI = - NBh F( )dz
Ac I Vc
QMIII: SCE3337 10
R.T. Sang
BI B
2 = 2 W << 1 and as such is negligible compared to the 1 term on the RHS of
Ac A
the equation.
Therefore
dI NBh
=- F ( )dz
I Vc
We now make the substitution
NBh
K= F( )
Vc
Hence
dI
=- Kdz
I
Integrating
I z
dI
∫ I ∫0
= -Kdz
I0
Thus
I(z) = I0 e − Kz
This is just an exponential fall off in intensity and provided that the atom density and the
spectral profile are known we can then obtain the absorption coefficient B. Essentially we
have recovered Beer's Law.
K( ) =
N Bh
F( )
V c
1 c N
BF( ) = K ( ) where is the density: =
h V
∫ F( )d =1
−∞
Hence
∞
c 1
B=
h ∫ K ( )d
−∞
1 1
Usually the frequency spread is small and the term can be factored out as yielding
0
for the Einstein B coefficient
∞
c
B=
h ∫ K(
0 −∞
)d
Case 2: If the absorption or the stimulated emission rate >> spontaneous decay rate (eg:
strong laser excitation), thus the term
BI
2 >> 1
Ac
Therefore
BI dI Bh
2 =- F( )dz
Ac I c
Ah
⇒ dI =- F( )dz
2 c
Integrating
I z
Ah
∫I dI = - 2 c F ( )∫0 dz
0
Thus
Ah
I(z) = I0 − F( ) z
2 c
We get the interesting result that the intensity falls off linearly with distance rather than
exponentially. We also notice that the intensity is proportional to the Einstein A coefficient
unlike our low intensity case which had a rapid exponential decay dependence on the B
coefficient.
SCE3337: QMII 1
R.T. Sang
Consider our two level atom once again being irradiated with near resonant light as shown
below:
The usual three processes will occur: absorption, spontaneous emission and stimulated
emission. If the number of excited atoms, N2 is greater than N1 , then instead of absorption
of the incident light beam, amplification will take place. It can be thought of as if the gas
had a negative absorption coefficient. The condition in which N2 >N 1 is known as
population inversion. This is impossible for a two level atom as we have seen in
previous lectures
We now consider a method of creating a population inversion via the use of three energy
levels. Consider a system that consists of three non-degenerate energy levels, level |0> is a
ground state and states |1> and |2> are excited states as shown below:
We have a light beam with energy density W p which is used to excite population from the
|0> to the |2> transition. A light beam used in this manner is called a pump beam. The
SCE3337: QMII 2
R.T. Sang
1
fraction of the total number of atoms pumped up is usually small and in the order of .
10 6
An atom in level |2> can emit light by making a downward transition by either |2>→|1> or
|2>→|0> and an atom in level |1> can decay to the ground state (|0>). We will soon see via
rate equation that it is possible to achieve a population inversion between levels |2> and |1>
i.e. N2 >N 1 .
We have some radiation with a frequency ω that corresponds to the transition frequency
between levels |2> and |1> that is incident on the system i.e.
( E2 − E1 )
=
h
This radiation has energy density W and hence will induce transitions from |1>→|2>
(absorption) to as well as |2>→|1> stimulated emission. Now since there is a population
inversion the number of simulated photons is greater than the absorbed photons and hence
the incident radiation will be amplified.
We assume that only the three levels in our diagram participate hence the total population is
N = N1 + N2 + N 0
The three rate equations that describe the change in population of the three levels are given
by
dN 2
= N0 B02 W p − N2 B02 W p − N2 A20 − N2 B21 W + N1 B21 W − N2 A21
dt
dN1
= N 2 B21W − N1 B21 W + N2 A21 − N1 A10
dt
dN 0
= − N0 B02 W p + N 2 B02 W p + N2 A20 + N1 A10
dt
These equations must balance so that the sum of the rates is equal to zero
In the steady state condition all the equation above are set to zero (i.e. the population in all
levels stay the same). Under this condition, equations may be found for the level
populations N0 , N 1 , N 2 in terms of N, W and W p . The terms for this are rather lengthy and
SCE3337: QMII 3
R.T. Sang
the main features of the steady state solution may be appreciated more easily by the
following procedure.
Let us define a pumping rate r which describes the population in the N2 level due to the
pump:
rN = N0 B02 W p − N2 B02 W p
= ( N0 − N2 )B02 W p
rN is the net rate at which atoms are excited into the state |2> by the pump. Consider the
steady state condition for the second rate equation (2) then,
Now from the third rate equation we have under steady state conditions
6 44 47444 8
From above this term = Nr
⇒ 0 = N2 A20 + N1 A10 − Nr
N2 ( A10 + B21 W )
=
N1 ( A21 + B21 W )
Thus in order to get a population inversion between |2> and |1> we must have
(A10 + B21 W )
> 1 which will occur for A10 > A21
(A21 + B21 W )
This tells us that atoms that are excited to state |2> must decay relatively slowly into state
|1> from which they drop rapidly into the ground state. We can solve the above equations
to give
N ( A21 + B21W )r
N1 =
A10 ( A21 + B21 W ) + A20 ( A10 + B21 W )
N ( A10 + B21W )r
N2 =
A10 ( A21 + B21 W ) + A20 ( A10 + B21 W )
SCE3337: QMII 4
R.T. Sang
1 1 . 3 Lasing Action
It is all very well to get a population inversion in a gas, it is another problem to obtain the
so called lasing action. In essence, a laser consists of some medium, in which
population inversion occurs as well as an enclosed cavity (sometimes it is also referred to
as a resonator) with reflecting walls. A generic simple laser is shown below:
Suppose that we have created a population inversion in a gas medium which has an energy
level structure similar to the three level system we considered above. The population
inversion may be created by a flashlamp or another laser. A spontaneously emitted photon
with a frequency that matches the transition energy will also stimulate the emission of other
identical photons. Surrounding the medium by two mirrors allows us to redirect photons in
one dimension back and forth an in doing so stimulate more photons. Only photons
travelling at right angles to the mirrors will be significantly amplified and of course will be
useful to the laser beam. As the photons travel back and forth through the medium, the
number of photons identical with the first photon will build up provided that the
amplification or the gain of the medium is greater than the cavity losses due to imperfect
mirrors. The point at which the gain equals the losses is called threshold. One of the
mirrors in the laser cavity is made to be slightly transparent and allows a fraction of the
photons to escape, these escaped photons constitute the laser beam.
The mirrors enforce boundary conditions (i.e. the electric field must vanish at the surface of
the mirrors) on the photons propagating within the cavity and hence only standing
waves will be formed in this cavity (remember that the photons are also
electromagnetic waves). As such the cavity will only support wavelengths with n =L
2
where L is the length of the laser cavity.
Recall that
c
c=f ⇒ f =
2L
=
n
Thus
nc
fn = = nf1
2L
where f 1 is the fundamental mode. This means that the possible output frequencies are
c
spaced by . For a 1 meter laser the frequency spacing is f=150MHz.
2L
SCE3337: QMII 6
R.T. Sang
c
2L
The figure above displays the output from the laser which consists of a number of discrete
lines within a Gaussian envelope of the Doppler broadened line (essentially this broadening
is due to the motion of the atoms, since we have some thermal distribution of atoms (i.e.
different speeds) then they atoms will see the stimulated photons with different frequencies
hence the Gaussian envelope). The discrete lines are called the longitudinal modes of the
laser and such operation is multimode behaviour.
It is possible to run in the so called single mode by filtering out undesired frequencies by
using frequency selecting devices within the laser cavity such as an etalon.
1 1 . 5 Transverse Modes
A laser also behaves like a waveguide and as such each longitudinal mode also has a
transverse mode structure (TEM). Essentially this comes from the fact that our waves are
three dimensional. The transverse profile of the electric field within the laser resonator can
be found by using Huygen's principle. It is possible to show that the transverse profile of a
spherical wave is Gaussian in nature which is what we expect to see from the output of the
laser. Many laser resonators use spherical mirrors rather than planar mirrors as it is simpler
to align these systems. These system of mirrors also support higher order Gaussian beam
modes. If we consider our laser longitudinal modes in the z direction then the TEM mode
structure is denoted as TEMmn where m and n define the number of nodes (ie the number of
0 intensity points) in the x and y directions respectively. Some of the lower order modes
are shown below:
TEM 00 TEM 01
TEM 11 TEM 20
SCE3337: QMII 7
R.T. Sang
Optical Excitation
One of the obvious techniques that we can use to create our population inversion is to use a
high intensity light source on our medium. One such source is known as a flash lamp
which produces a high intensity burst of incoherent light. The ruby laser is one such laser
that works on this principle and is surrounded by a flashlamp as shown below
The ruby (Al2 O3 ) is doped with chromium atoms which creates Cr3+ ions in the ruby (it is
the chromium atoms in ruby that give them the characteristic redness). The presence of the
ruby with the Cr3+ ions modifies the ions energy level structure as shown below
SCE3337: QMII 8
R.T. Sang
The flashlamp provides the pump beam with radiation at approximately 550nm in which the
ground state Cr3+ ions absorb the radiation and are promoted to the 4 F2 energy level. This
state decays almost entirely to the 2 E state via rapid non-radiative decay. The 2E state is a
metastable state with a long lifetime in the order of 5msec and as such it is possible to
obtain a population inversion between the ground state and this metastable state provide that
the pump is strong. The transition from this state to the ground state is the lasing transition.
One of the problems associated with this laser as it requires a great deal of pump energy
and hence a low efficiency. This however was the first optical laser which was created by
Maiman. It was interesting to see that Maiman publish this paper in 1960 in Nature# in a
communication that was only 300 words long!
Other optical excitation is also possible such as the use of another laser such as in the case
of a dye laser which use another laser for their pump.
Another possibility is a diode laser array which produces intense pump beams for high
laser power such as those used in Nd:YAG lasers. These types of lasers are also solid state
and the host crystal in this case is Y3 AL5 O12 (called YAG, an acronym for yttrium
aluminium garnet) in which Y3+ ions are replaced (doped) with Nd3+ ions (Neodymium). It
is also possible to use flash lamps for population inversion in this medium.
Transfer of Excitation
This technique is used in many gas lasers such as the helium-neon laser (He-Ne).
Essentially this laser consists of a gas of He and Ne which has a high voltage place
between two electrodes which creates a gas discharge. Electrons from the gas discharge
inelastically collide with ground state He atoms exciting them to the 23 S metastable state.
The figure below shows the relevant energy levels for He and Ne. Thermalising collisions
occur between the excited He and the ground state Ne which can promote a transfer of
energy from the excited He atoms to the ground state neon atoms leaving the Ne atoms in
the 2s state which has comparable energy to the 23 S state of He (naturally the He atoms go
back to the ground state if they lose their energy). The 2s state actually consists of 5
different energy levels separated by 0.15eV. A population inversion exists between the 2s
states and the lower energy 2p states which consists of 10 closely spaced energy levels.
The lifetime of the 2p states is about half as long as the 2S states and it is possible to get
lasing action on thirty of the allowed 2s→2p transitions. The excited 2p Ne atoms can then
decay via photon allowed transitions to lower energy levels and finally to the ground state
for repumping.
#
T.H. Maiman, Nature 187, 493 (1960).
SCE3337: QMII 9
R.T. Sang
SCE3337: QMIII 1
R.T. Sang
Lecture 13
1 . 0 Radiation Pressure
Photons that make up a travelling electromagnetic wave with wavevector k each carry
momentum. Their behaviour in this respect is like any other particle in quantum mechanics,
being governed by the de Broglie relation between particle wavelength and momentum:
h 2
P= but k = (1.1)
hence
P = hk (1.2)
As with any interacting particles the interaction of photons with any other particles must
conserve momentum in the interaction. The effect of photon momentum on electrons
was seen in the early experiment of Compton in 1922 where he observed the conservation
of momentum in the scattering of high energy X-ray photons by electrons.
Low energy photons (including the visible spectrum) can also have interesting momentum
exchanges with relatively heavy atoms. Let's consider the effects of the exchange of photon
momentum on the three possible processes that can occur with the basic three atom-photon
interactions that we have previously considered (absorption, stimulated emission and
spontaneous emission). The effect of the three process on a two level atom is shown on the
following page.
In the absorption process, hk of momentum is transferred from the light beam to the
hk
atom which will acquire an additional velocity of v = in the direction of the incident
M
light beam where M has been used to denote the atomic mass.
If the excited atom subsequently is induced to emit a stimulated photon then the excited
atom will produce an identical photon as the light beam in its decay to the ground state and
in the process will loose the velocity it previously gained due the absorption as such the
hk
change in velocity of the atom will be v = − due to the recoil (conservation of
M
momentum see figure).
If the atom decays due to the vacuum field (spontaneous emission) then the photon
from the decay can be emitted in any of the 4π radian direction of solid angle, hence the
change in momentum of the atom from the decay process can be in any of these emission
directions, thus the recoil of the atoms due to this process is in a random direction with
each direction being equally possible. Hence on average, there will be no cancellation of the
previously acquired momentum from the original absorption process.
It follows that after many absorption/spontaneous emission processes the atoms will be
pushed by the photons in the direction of the light beam by on average by hk of momentum
per photon . The transfer of momentum to the atoms is equivalent to a pressure exerted on a
gas by the radiation.
SCE3337: QMIII 2
R.T. Sang
Suppose that the number of atoms is sufficiently large to produce smooth time
dependencies in the atomic populations. The rate of change of the atomic momentum (all
atoms) P atom in the presence of radiative energy density W with frequency ω that is in
resonance with out two level atom above, is given by the difference between the absorption
and stimulated emission rates (just the spontaneous emission rate):
dPatom
= hk( N1 − N2 ) BW (1.3)
dt
where N 1 is the population in the ground state and N 2 the population in the excited state, B
is the Eintein B coefficient. For a two level atom this quantity will always be positive since
N 1 >N2 , it is however possible to get negative momentum transfers for three levels atoms
where population inversions may occur but we will restrict ourselves to the two level atom.
Under steady state conditions the rate equations for a two level system yield:
( N1 − N2 )BW = N2 A (1.4)
dPatom
= hkN 2 A (1.5)
dt
Recall for a two level system that the rate equations under steady state conditions yields a
population in the excited state given by
NBW
N2 = (1.6)
( A + 2BW )
Thus
dPatom BW
= hkNA (1.7)
dt ( A + 2BW )
For high light powers reaching a saturation level BW >> A then
dPatom hkNA
= (1.8)
dt 2
This is an upper limit of momentum change due to the light beam. Any increase in the light
beam energy density produces little effect on the momentum transfer rate once saturation is
reach (actually this is only true for the rate equation model in reality there can be greater
momentum transfers at higher laser intensities when various polarisations are used, this
comes out in a full quantum mechanical investigation and will be investigated later in the
course).
0.5
0.3
0.2
0.1
0
0 10 20 30 40 50
(BW/A)
Hence the light beam induces a steady state average force on each atom the magnitude of
which is given by the above expression which is shown graphically above as a function of
scaled energy density. At saturation the force reaches a maximum which is
A
F (1) atom = hk (1.10)
2
This force is often referred to the spontaneous force due to its relationship with the
spontaneous decay rate from the excited state.
To get an idea of the magnitude of the spontaneous emission force we can look at a real
atom such as sodium interacting with a single photon that is in resonance with the 32 S1/2
ground state to the 32 P3/2 excited state which has a wavelength of 589nm . The Einstein A
coefficient (decay rate) is equal to 6.25x107 sec-1. Then for one photon the force is
A hA
F (1) atom = hk = = 3.5X10 −20 (N) (1.11)
2 2
At first glance this would appear to be a relatively small force but this is deceptive, it is
more instructive to look at the acceleration due to a single photon:
A hA
a (1)atom = hk = (1.12)
2M 2 M
The mass of sodium is 23x1.661x10-27 kg
If your car could decelerate at this rate it would be able to stop travelling from an initial
velocity of 100km in 2.8x10-5 secs with a stoping distance of only 0.5mm!
A constant stream of photons could therefore decelerate an atomic beam if they could
absorb photons over a period of time. For a typical atom at room temperature, the velocity
is around 500m/s which means that you would need an interaction time in the order of
5x10-4 s. The real difficulty in the process of slowing the atom down using the light force,
is that in the atom's frame of reference it would see a changing frequency of the slowing
laser light due to the Doppler shift so after a few absorptions the atoms would no longer be
in resonance with the slowing light and hence could not continue to absorb photons and
decelerate. There are techniques to get around this problem which will be considered in the
next lecture.
Let's look at an experiment involving the deflection of atoms via light. Consider the
experiment below in which a well collimated beam of ground state sodium atoms interacts
at right angles to a laser beam resonant with 32 S1/2(F=2)→32 P3/2(F'=3) hyperfine transition
in sodium which as a wavelength of 589nm.
The figures below display a profile of the deflected atomic beam. Notice that two peaks are
seen. The peak labelled F=1 are ground state atoms in the F=1 hyperfine level, these
atoms will not be deflected due to the fact that the laser beam is not in resonance
with this state and no momentum exchange can occur since photons will not be absorbed.
In this case this beam of atoms act as a reference point.
Atoms in the F=2 ground state will however be deflected since they are in resonance
with the incident laser beam and will absorb photons. The deflection will be in the direction
of propagation of the laser beam as we have learnt from our analysis above.
SCE3337: QMIII 5
R.T. Sang
F'=3
60
Light tuned to this transtion
40
20
F=2
0 3 2S Ground State
8 9 10 11 12 13 14 15 16
1/2
Dist (mm) F=1
The deflection of atoms from atomic beams via optical forces also has practical applications
as they can be applied to isotope selection provided that the isotopes have transitions that
are resolvable by the deflecting light beam.
Radiation pressure due to light is now routinely used in laboratories around the world for
the manipulation of atoms. This force has been used to slow atoms down and trap atoms
(using three mutually perpendicular counter propagating laser beams). A new state of
matter has been created called a Bose-Einstein Condensate (BEC). In a BEC the atoms are
slowed down so much (cooled) their deBroglie wavelength increases in size to a point
where near neighbour atoms in the gas have overlapping wavefunctions. In this case a
classical description of this matter no longer holds and a more appropriate description is
that of a macroscopic wavefunction. BEC has been used in recent times to create the
atomic equivalent to a laser i.e. an atom laser.
SCE3337: QMIII 1
R.T. Sang
Lecture 14
1 3 . 0 The Width and Shape of Spectral Lines
Up to this point we have just touched on the fact that spectral lines have a finite spread of
frequencies that is we say they have a finite width. There are three mechanisms which are
mainly responsible for the spread in frequency of emitted light.
The first of theses process is due to the decay process itself, and the width is called the
natural line width. Suppose that an atom is in a state |k> and can make downward
transitions to a number of states |j>. The lifetime of the state |k> is related to various
Einstein A coefficients by the following. If k is the lifetime, the downward transition
probability is
1
= sum of the emission probabilities = ∑A kj
k j
This is the transition probability per unit time. Notice that we are only considering the
spontaneous emission rate, not the stimulated emission rate.
An estimate of the natural width can be obtained from the Uncertainty Principle relating
energy and time:
∆E∆t ≈ h
Consider a two level atom labelled |k> and |j>. State |k> is higher in energy than |j> as
shown below. We can determine the natural spread of frequencies of the radiation emitted
from state |k> to |i> by the following:
E k |k> . ∆Ek
hω
Ej |j> ∆E
j
If the lifetime of the state |k> is k then the uncertainty in the time t is characterised by the
lifetime of the state:
∆t = k
⇒ ∆E k = h
⇒ h∆ k =h
1
⇒∆ =
k
SCE3337: QMIII 2
R.T. Sang
where the width of the line is the sum of the two energy level widths involved:
∆ =
[∆E j + ∆E k ]
h
Often the lower state is the ground state which has an infinitely long lifetime provided that
the atom is not subject to a perturbation. The frequency spread is then only that for the
upper state only. Obviously the longer the lifetime of the state, the more narrow the energy
and hence frequency spread.
This is useful to know but it doesn't tell us anything about the spectra profile (or the line
shape of intensity Vs frequency) of the emitted radiation. We can gain some further insight
into this shape of the spectral profile by considering a classical model of an excited atom
which was due to Lorentz.
ω
x 0
y k
We assume that the excited atom is represented by a classical dipole oscillator consisting of
an electron vibrating up and down in simple harmonic motion as shown above. Suppose
that the frequency of oscillation is 0 which corresponds to the transition frequency. We
also let the atom be subject to electromagnetic radiation which is propagating in the x
direction which is plane polarised in the z direction (i.e. the electric field vector points in the
z direction) with oscillation frequency such that:
E = E0 e − i t
The electric field will induce the electron to oscillate in the direction of the electric field (up
and down in the z direction) due to the force caused by the interaction between the charge
and the electric field. This is known as the radiation reaction force and is given by
F = qE
Therefore
F = (−e)E 0e − i t
(1)
Recall that the equation of motion for a simple harmonic oscillator is given by
d 2x
m + kx = 0
dt 2
SCE3337: QMIII 3
R.T. Sang
d2 x
m 2 + x = 0
2
(2)
dt 0
k
where 0 =
m
If we now drive the simple harmonic oscillator with the radiation above we can equate
equations (1) and (2):
d2 x
m 2 + x = (−e)E0 e −i
2 t
dt 0
d2 x dx
m 2 +Γ + x = (−e)E 0e − i
2 t
dt dt 0
eE0 −i t
−
Γ
t
e
x = Ae 2 −i ' t
e − m
0 − −i Γ
2 2
where
Γ 2
'= −
2
0
2
The first term is a transient term which decays off exponentially with time and represents
spontaneous emission. The second term is a steady state term which oscillates at the driving
frequency of the atom.
Consider for the moment the transient term. This term oscillates at almost exactly the
2
resonant frequency. The power radiated is proportional to x . From classical dipole theory
I(t) ∝ e −Γ t that is the damping constant is associated with the spontaneous emission
process.
What is the distribution of the emitted energy or the intensity distribution as a function of
frequency for this transient oscillation? We can use Fourier Analysis to gain some insight
into this. For example, we have the time dependence of the oscillator:
f(t)=0 t<0
SCE3337: QMIII 4
R.T. Sang
Γ
− t
2 −i ' t
f (t) = Ae e t ≥0
The Fourier transform allows the determination of the energy distribution in the frequency
domain:
∞
f (t) = ∫ F( )e −i t d
−∞
∞
1
F( ) =
2 ∫ f (t)e i tdt
−∞
The first expression for the Fourier transform says that a time dependent function can be
thought of as a summation of sinusoidal functions of amplitude F( ). The energy
associated with each harmonic is just the modulus square of F( ). Ignoring proportionality
constants,
I( ) ∝ F( )F * ( )
0 Γ ∞
1 1 − t
⇒ = ∫ 0 •e dt + ∫ A • e 2 • e − i ' t •e i t dt
i t
2 −∞ 2 0
144 2443 144 444244444 3
t <0 t≥0
∞ Γ
A − + i( ' − ) t
∫e 2
⇒ = dt
2 0
∞
−A 1 − 2 +i( Γ
' − ) t
⇒ = e
2 Γ + i ( ' − ) 0
2
−A 1
⇒ = [0 − 1]
2 Γ + i( ' − )
2
Thus
A 1
F( ) = Γ
2 + i( ' − )
2
Taking the modulus square gives the energy associated with each harmonic:
SCE3337: QMIII 5
R.T. Sang
A2 1 1
I( ) ∝ F( )F ( ) = •Γ
*
Γ
2 + i( ' − )
− i ( ' − )
2 2
A2 1
⇒ F( )F ( ) =
*
2 Γ
2
+ ( ' − )2
2
1
I( ) ∝
Γ + ( ' − ) 2
2
2
−1
0 ≈ 10 Rads and Γ ≈ 10 9 Rads −1 therefore to a good
15
For optical frequencies
approximation ' ≈ 0 . Thus the energy distribution is
1
I( ) ∝
Γ +
2
( − )
2
2 0
Lorentzian Profile
1
0.8
0.6 Γ
Ι(ω)
0.4 FWHM
0.2
0
−3Γ/2 −Γ −Γ/2 0 Γ/2 Γ 3Γ/2
(ω 0 −ω)
SCE3337: QMIII 6
R.T. Sang
The full width half maximum can be calculated as follows: At the maximum we have
( 0 − ) = 0 therefore
1
I( ) max =
Γ
2
2
1
I( ) half − max =
Γ
2
2
2
1 1
2 =
Γ Γ +
2
2
2 (
− )
2
2 0
Γ + ( Γ
2 2
) = 2
2
⇒ −
2 0
2
Γ 2 Γ 2
( ) = 2 −
2
⇒ 0 −
2 2
Γ 2 Γ
( ) = ⇒
2
⇒ 0 − 0 − =
2 2
This is the half width half maximum, the full width half maximum = 2xHWHM thus
FWHM = Γ
The full width half maximum of the Lorentzian profile is Γ, the damping constant which
was equal to the spontaneous emission rate. The so-called natural line width of a spectral
Γ
line is therefore quite narrow 1:106 and is solely determined by the spontaneous
0
emission rate.
SCE3337: QMIII 7
R.T. Sang
1 3 . 1 Doppler Broadening
Let us suppose that an atom of mass M is in an excited level E2 and has velocity v 2 . The
atom emits a photon of energy hω and falls to a lower energy level E1 . The photon has
momentum P = hk and causes the atom to recoil to a new velocity v 1 (recall the lecture on
radiation pressure) as show below:
Before After
P=Mv
2 P = Mv1 + hk
Atomic Level Structure
Atom
Atom
Ground State
Excited state
By conservation of momentum
Mv2 = Mv1 + P
By conservation of energy
1 1
E2 + Mv 2 2 = E1 + Mv12 + h (2)
2 2
Let ω0 be the frequency of the light emitted if the atom initially had zero velocity before and
after the collision, i.e.
h 0 = E2 − E1 (3)
M(v1 2 − v 2 2 )
1
( E2 − E1 ) − h = (4)
2
We can eliminate E2 , E 1 by equation (3) and From equation (1) we can eliminate v 1
hk
v1 = v 2 − (5)
M
Thus
SCE3337: QMIII 8
R.T. Sang
2h h2 k 2
v1 • v1 = v2 2 − v2 • k + 2 (6)
M M
Hence using equations (3) and (6) we can rewrite equation (4) as
1 2 2h h 2k 2
h −h = M v 2 − v2 • k + 2 − v2 2
2
0
M M
h 2k 2
⇒ h 0 −h = −hv 2 • k +
2M
Let us define the direction of the emitted photon to be in the z direction. Then
v2 cos h 2
0 − =− +
c 2Mc 2
where we have used k = and θ is the angle between the direction of the atom's velocity
c
and the emission direction of the photon. For the present case we will look at the maximum
effect which occurs for θ=0o then our expression reduces to:
1− v 2 + h 0
=
0 c 2Mc 2
We can now look at the relative orders of magnitude of the three terms on the RHS of this
equation. For a typical atom at room temperature with a transition at optical frequencies we
have the following parameters: v 2 ≈ 103 ms −1 , ω ≈ 1015 Rads −1 , M ≈ 10−27 kg.
v2
≈ 10−5
c
h
2
≈ 10 −9
2Mc
As such the third term in the equation on the RHS is negligible and can be ignored reducing
the expression for the transition frequency as
1− v 2
=
0 c
Taylor expanding the bracket term to first order about the zero point [ i.e. (1− x)−1 ≈ 1 + x ]
reveals
SCE3337: QMIII 9
R.T. Sang
1+ v 2
= 0 c
Therefore the frequency of the emitted radiation experiences a conventional Doppler shift
which is determined by the velocity component of the atom in the direction of emission.
v 2 v2 2
= 0 1+ +
c c
M −( Mv z 2 )
P(v z )dv z = e dv z
2
where
1
=
kB T
Using our expression for the Doppler shifted frequency we can write the z component of
the velocity as
c( − ) c
vz = ⇒ dv z =
0
d
0 0
defining
∆ = 2 0 = 2 0
2 2k B T
c M c M
This expression for the frequency distribution is known as a Gaussian lineshape. The peak
of the Gaussian is at ω=ω0 and the line has an intensity profile given by
2 ( − )
2
− 0
∆
I( )d = I0 e d
Gaussian Distribution
1
0.8
(Arb. Units)
0.6
Intensity
0.4
FWHM = ∆ ln(2)
0.2
0
−1.5 −1 −0.5 0 0.5 1 1.5
ω−ω 0
2kB T Kelvin
FWHM = ∆ ln(2) = 2 0
ln(2) ≈ 7.16x10−7 0 per
c M amu
As an example of the size of the broadening due to this effect consider a helium atom with
amu =4 at 300K. In this case
FWHM 300
= 7.16x10−7= 6.2x10 −6
0 4
Typical values for optical frequencies ω 0 ≈ 1015 Rads−1 then the FWHM is approximately
1010 Rads -1. Comparing this ration to the natural line width shows that only the atomic
transitions with the shortest lifetimes of 10-9 sec would a natural line width approaching that
of the Doppler width hence this effect, in general is more dominant than the natural
linewidth broadening.
SCE3337: QMIII 1
R.T. Sang
Lecture 15
1 4 . 0 Collision Broadening
So far we have considered two broadening mechanism that are responsible for the width
and the shape of spectral lines. We now look at a third mechanism for line broadening
which is due to the collisions of atoms which is called collision broadening. This
process can be quite complicated and we will only treat this in a rudimentary fashion.
Up to now we have treated atoms in a gas as if they to do interact but in any real gas of
atoms, atoms will be subject to the interaction forces of nearest neighbour atoms, ions or
molecules. This will of course perturb the state of any radiating atoms and as such will lead
to a broadening of the lineshape which is quite often larger than the natural linewidth. The
increase in the linewidth is a function of the density of the perturbing species and is
therefore also known as pressure broadening.
Consider the influence of a single perturber at a distance r as shown in the figure below:
Energy
Unperturbed Frequency
|k>
hω ki
|i>
h(ω k + ∆ω ki )
Perturbed Frequency
r
Distance between atoms
If ∆Vk (r) and ∆Vi (r ) are the changes in the energy levels of the states |k> and |i>
respectively for a two level transition, then the instantaneous change to the transition
frequency is given by
1 4 . 1 Interatomic Forces
It is often possible to represent the long range interaction between an excited atom and a
perturber by a potential of the form:
Cnk
∆Vk (r) =
rn
where Cnk is a constant which depends on the excited level involved as well as the
perturbing species. n is an integer an corresponds to various forces some of the well know
cases are:
• n=2: This applies to the case of hydrogen and hydrogenic ions in the electric fields
produced by other ions or electrons. These effectively give rise to a Stark shift of the
energy levels which depends linearly on the field strength.
• n=4: The describes the Stark broadening in helium and other systems where the splitting
is a quadratic function of the electric field of the perturbers.
• n=3: This applies to the case of the resonance dipole-dipole interaction. This interaction
is finite only when the excited atom interacts with an identical atom in the ground state
and when a strong allowed dipole transition, usually the resonance line of the atom,
connects the two levels.
• n=6: This is the usual long range attractive van der Waals dipole-dipole interaction
which always exists between any two atoms.
• The first two cases are only important when dealing with highly ionised gases, for
unionised gases, the most important interactions are the dipole-dipole interactions as
well as the van der Waals interaction. It should be noted that we have not discussed
short range forces as their effects are difficult to calculate and can not be expressed by
any general formula.
Recall from Fourier transform theory that we used in the last lecture that the shape of the
line profile at a frequency separation of = 0 - from the line centre is determined by the
1
radiation emitted during the time interval t, where ∆t ≈ . If we imagine the atom
∆
emitting radiation until perturbed by a strong collision, then the time of interest will be the
mean time between collisions, Tc. This must be compared with the duration of one
collision, tc.
We consider a collision regime where the duration of the collision time is less than the
change in frequency from the central frequency of a transition i.e.
1
tc << ≈ Tc where ∆ = −
∆ 0
In terms of the Lorentz approach, it is assumed that the an atom excited at t=0 would
continue to radiate according to the radiation from a classical dipole which from last lecture
was given by:
f (t)= 0 t <0
Γ
− t
2 −i
f (t) = Ae t ≥0
t
e 0
This emission of radiation would occur until a collision would occur and at such time the
emission was terminated, t=Ti. Using Fourier transform theory we get that the emission of
radiation is given by
I( ) ∝ F( )F * ( )
where
∞
1
F( ) =
2 ∫ f (t)e i tdt
−∞
In our case
Ti
1
F( ) = ∫ f (t)e
i t
dt
2 o
I( ) ∝ (1)
Γ2
( − 0) +
2
4
Exercise: Show this.
For =0 this distribution is identical to the Fraunhofer diffraction pattern of a single slit.
This spectrum is for only one collision time but in reality there will be other collision times
and as such the spectrum must be averaged over this.
The probability that an atom will undergo a collision in a distance dl after having travelled a
distance l without a collision is
l
− dl
P(l)dl = e
where is the mean free path (the average distance atoms travel prior to a collision) for the
excited atom. The time collision time is
Hence the mean number of collisions (mean frequency) per unit time is
1 v
=
Tc
The spectral distribution E(ω) can then be found by averaging the expression give by
equation (1) over the distribution given by equation (2) yielding:
Γ 1
+ T
2 2 1
I( ) ∝ A c
2
Γ 1
( − ) + +
2
2 Tc
0
2
FWHM = Γ +
Tc
Hence we see that the line is broadened by the collisions. As the density increases, the
collision time decreases and hence the line becomes broader. This expression can be related
1
to the collision cross section since is the mean number of collisions per unit time
Tc
FWHM = Γ + 2N v
= 2
is the collision radius and is taken as the distance between centres of the colliding atoms,
or alternatively as the sum of the radii of the colliding particles. The mean relative velocity
is given by
1
8k T ( M1 + M2 ) 2
v= B
M1 M2
SCE3337: QMIII 5
R.T. Sang
In order to compare the line broadening mechanisms discussed so far let us look at an
example atoms that has a mass of 100amu with an excited level that has a lifetime of 10 -8s
and radiates at 500nm. Let us assume that the gas is at 300K and we take a typical optical
collision cross section with is σ = 5x10−15 cm2 , we will assume that the gas pressure is
10Torr (1Torr=133.3 Pa). With these parameters it is possible to summarise the widths in
the table below:
We can see that under these conditions that the dominating factor to the linewidth is the
Doppler Broadening.
1 4 . 4 Voigt Profile
We have so far considered all of the line broadening effects separately but of course in any
real situation we would have a combination of broadening effects that will be acting
simultaneously. As a result the observed line shape is neither Lorentzian or Gaussian in
profile. To investigate the possibility of a combination of effects consider a moving atom
whose resonance frequency is observed to be at 0 ' . Due to collisions or to the natural
linewidth the emitted radiation consists of a distribution of frequencies about the central
frequency 0 ' which is given by the Lorentzian function (recall that both, collision
broadening and natural line broadening give spectral profiles that are Lorentzian):
Γ'
L( − ' , Γ') = 2
0
Γ' 2
( − 0' ) +
2
where ' is the full width of the distribution at half the intensity. The spectral profile of the
radiation emitted by the moving atoms is found by averaging the above equation over the
thermal distribution of the Doppler shifted atoms which is given by
2( − )
2
− 0
2
∆
G( − ' , ∆) = e
0
∆
Γ' 2( − )
2
∞ − 0
2 ∆
' , Γ' / ∆) = ∫ 2
I( − • e d '
0
Γ'2
∆ 0
( − ') +
2
0
0
4
The resulting profile is known as the Voigt Profile. The shape of the profile is
Γ'
determined uniquely by the ratio of . Unfortunately it is not possible to express this
∆
profile in an analytic form however it can be evaluated numerically. The figure below
compares the profiles of the normalised Lorentzian, Gaussian and Voigt profiles for the
identical full width half maximum.
SCE3337: QMIII 1
R.T. Sang
Lecture 16
1 5 . 0 The Optical Bloch Equations
It is impossible the discuss the interaction of collections of atoms with light exactly. It is
even impossible to treat even one atom's interaction with light exactly. In the next lectures
we will discuss a new model of matter light interactions via various approximations. The
principle approximation is that the optical radiation field is very nearly perfectly
monochromatic and it coincides in frequency with the transition frequency of the atom
which has only two energy levels; the two level atom.
A two level atom interacting with a single mode electromagnetic field is conceptually the
same kind of object as the spin half particle in a magnetic field. The basic dynamical
equations which follow from the Schrödinger equation and govern the evolution of the two
level atom variables are almost the same as those appropriate to spins. As such we use the
approach take by Bloch which was developed for magnetic resonance and apply it to optical
resonance in the two level atom.
To facilitate this application, we follow the custom of Bloch and define a fictitious electric
spin vector, or a pseudo spin vector, whose components are related to the atom's atomic
dipole and inversion. The equations obeyed by the components of this vector are frequently
called the Optical Bloch Equations.
Consider an atom with two energy levels labelled |e> and |g> with energies Ee and Eg
respectively with E e > E g as shown below
The atom is now subject to the interaction with a monochromatic optical oscillating electric
field which has a frequency that is near resonant with the transition. The atomic
unperturbed Hamiltonian is given by Ha with the unperturbed energy eigenvalues given by
the eigenvalue equation:
h
Ha | g >= − 0 |g>
2
h
Ha | e >= 0 |e>
2
SCE3337: QMIII 2
R.T. Sang
h
0
< j | Ha | i >= 2
0
h
0 −
2 0
Recall from previous lectures that the Interaction Hamiltonian for an atom interacting with
an electric field is given by
h(t) = −d • E(t )
E(t ) = E cos L t
Assume that the atomic dipole moment is aligned parallel to E(t) then
The matrix elements of h(t) are found by taking the expectation of h(t):
Because |g> and |e> are optically connected, their parity must be opposite (see Lecture 9)
such that
0 d
deg =< j | d | i >=
d 0
SCE3337: QMIII 3
R.T. Sang
1
It is possible to represent both the atomic and interaction Hamiltonians by spin matrices
2
which are known as the Pauli spin matrices. For a electron with total spin vector S then the
components of the vector are given by:
h 0 1 1
Sx = =
2 1 0 2 x
h 0 −i 1
Sy = =
2i 0 2 y
h 1 0 1
Sz = =
2 0 −1 2 z
Going back to the matrix form of the interaction Hamiltonian it is possible to rewrite it in
terms of the matrix S x:
0 1
= −E cos td
L
1 0
dE h 0 1
= −2 cos Lt
h 2 1 0
= −2Ω cos LtS x
dE
Ω=
h
The Rabi frequency gives the strength of coupling between the light field and the atom.
We now consider rewriting the atomic Hamiltonian Ha in terms of the Pauli spin matrices,
recall that:
h 0
< j | Ha | i >= 2
0
h
0 −
2 0
h 1 0
= 0
2 0 −1
= 0 Sz
SCE3337: QMIII 4
R.T. Sang
Since we can put the Hamiltonians into these forms containing the Pauli spin matrices the
1
spin representation is therefore suitable for describing the interaction of monochromatic
2
light with a two-level atom. From second year you will have learnt that an electron in the
presence of a magnetic field will undergo Larmor precession. It is now possible to
develop the same analogy for the precession of our pseudo spin vector S about a pseudo
magnetic field in the following way:
Let us reconsider how spins precess in a magnetic field. This precession is known as the
Larmor precession. The magnetic moment of an electron is given by:
s = −gs B
S=− S
h
Recall that a magnetic dipole in the presence of a magnetic field B will experience a
torque (since the dipole will tend to want to align with the B field) which is given by
Γ= ×B
From this figure we can see that the is identical to the change in S , with respect to time,
hence we may write:
∆S dS
Γ= =
∆t dt
From the figure above we can see that the vector dS is just the arc length which is defined
as dl=rd thus in terms or our vectors
dS = S sin d
SCE3337: QMIII 5
R.T. Sang
dS d
Γ= = Ssin = Ssin (1)
dt dt
Equating equations (1) and (2) enables the precessional frequency to be defined such that
S sin = − S×B
⇒ Ssin = − SBsin
⇒ = B
The interaction energy of the magnetic moment in the magnetic field is given by
∆E = − S •B
∆E = − ( )
S z Bz
∆E = − (− Sz )Bz = BzSz
∆E = Sz
Ha = S = B0 Sz
0 z
This tell us that we can represent the Larmor precession of S , the pseudo spin vector about
a pseudo magnetic field B 0 which is parallel to the z direction.
SCE3337: QMIII 6
R.T. Sang
where
2Ω
B1 = − cos L t
The motion of the pseudo spin vector is quite complex as it precesses about B 0 and B 1 . It is
often convenient to re-express the pseudo magnetic field B 1 as
Ω
B1 = − (e Lt
+ e− Lt
)
To simplify our model, we now transform to a frame that rotates at the laser frequency - L
about OZ such that B1 → B1' where
Ω Ω
B1' = − (e Lt
+ e− Lt
)e − Lt
=− (1+ e −2 Lt
)
We now make the rotating wave approximation (RWA) which is an approximation
that allows us to discard high frequency terms (i.e. those rotating at twice the laser
frequency). The neglected high frequency terms give rise to the Bloch-Siegert shift. The
other term that is left is a constant in the (-)x direction and the magnitude is given by
Ω
B1' =
∆
B0 ' =
Ω
B1' = −
( B0 ' ) + ( B1' )
2 2
Beff =
Ω eff = Ω2 + ∆2
We can now investigate the precession of the pseudo spin vector S about B eff.
S =− S
From previous arguments we were able to demonstrate that the torque was
dS
Γ= = S × Beff
dt
Γ = − S × Beff
Γ = Beff × S = Ωeff × S
We can write eff and S in terms of components in the x', y', z' axis such that
Ω eff = (−Ω,0,∆)
ˆi ˆj kˆ
dSx dSy dSz
, , = Ω 0 ∆
dt dt dt
Sx Sy Sz
= i (−∆Sy )
dSx ˆ
dt
dSy ˆ
= j(ΩSz − ∆ Sx )
dt
= k( ΩSy )
dSz ˆ
dt
SCE3337: QMIII 8
R.T. Sang
u 1
Sx = ⇒ dSx = du
2 2
v 1
Sy = ⇒ dSy = dv
2 2
w 1
Sz = ⇒ dSz = dw
2 2
Thus the component equations become:
du
= −∆v
dt
dv
= ∆u +Ωw
dt
dw
=−Ωv
dt
Lecture 17
1 6 . 0 Interpretation of the Bloch Vector Components
Recall from early lectures that the state or the wavefunction of an atom may be represented
as a linear superposition of basis states:
n
| >= ∑c | i >
i
i =1
where cn are coefficients and {|i>} are a complete set of n orthonormal states. In the case of
the two level atom this sum is very simple and has only two terms which can be written as
where a and b are coefficients and are in general complex (see Lecture 1). For a normalised
wavefunction recall that the sum square of the coefficients is unity:
a + b =1
2 2
1
| e >=
0
0
| g >=
1
Now recall from the last lecture that we defined the w vector in terms of the S z component
of the pseudo spin vector:
w
Sz = ⇒ w = 2 Sz
2
The wavefunction can be written in terms of the column vectors as
0 1 b
| >= a + b =
1 0 a
< |= (b * a *)
Using the explicit expression for the matrix for S z the expectation of w is:
SCE3337: QMIII 2
R.T. Sang
h 1 0 b
< | w| >= 2 (b * a *)
2 0 −1 a
b
= h(b * a *)
− a
[
= h b2 − a2 ]
<w>
is the probability of finding the atom in the upper state - the probability of finding
h
the atom in the lower state, that is, w represents the population difference or as it is
sometimes referred to, the degree of atomic inversion or it is sometimes just called the
inversion.
Consider now the expectation of the dipole operator. Suppose that the matrix form of d was
0 d
d =
d * 0
0 d b
< |d| >= ( b* a*)
d * 0 a
da
= (b * a *)
d * b
= b* da + a * d * b
= 2Re( db * a)
u + iv
S + = Sx + iSy =
2
Putting in the expression for the matrices yields,
h 0 1 0 −i
2S+ = 2 + i
2 1 0 i 0
0 1
= 2h
0 0
SCE3337: QMIII 3
R.T. Sang
0 1 b
< | 2S + | >= 2h( b * a *)
0 0 a
a
= 2h(b * a *)
0
= 2hb * a
Recall from the previous page that the expectation of the dipole operator d was
Using the result for the expectation of 2S + enables this to be rewritten in terms of the
expectation of u and v as
1
< |d| >= 2Re [d(b * a)] = Re[d < | u + iv | >]
h
We now write the dipole operator in terms of its real and imaginary components such that,
d = dr + id i
1
< |d| >= [ d < u > − di < v >]
h r
This shows that u and -v represent the in-phase and in-quadrature (right angle) components
of the atomic dipole with the electric field {since in the Bloch equations (see last lecture) the
terms involving u are +u. and the v terms are -v. }.
The third Bloch equation that involves the time derivative of w, shows that the only
dependence on the other Bloch component vectors is v. Since this term only involves the
change of population and is coupled to the electric field that is producing the energy
changes we can conclude that v must be the absorptive component of the dipole and as
such u must be related to the dispersive part of the dipole.
u2 + v 2 + w 2 = a + b ( 2
)
2 2
=1
SCE3337: QMIII 4
R.T. Sang
We can define a new space where w, u and v are the unit vectors in this space and the block
vector is defined with respect to these unit vectors. The Bloch vector S rotates about the
dS
eff vector since =Ω eff × S .
dt
It is useful to note a couple of important cases:
(1) = 0: In this case the Bloch vector is confined to the w-v plane since Ωeff only has a
component in the u direction then eff points along this direction and as such the Bloch
vector precesses about the u axis and can not move into the u-v plane. If the atom is fully in
the ground state then w=-1, in the case where the atom is in the excited state w=1. The
direction of positive rotation is from the -w axis to the -v axis, the reason for this choice of
direction will be come clear in the next section.
Physically we see that for an oscillating electric field the population oscillates forever
between the ground and excited state while the electric field is on (a result that we saw in
time dependent perturbation theory). Of course in reality we must include spontaneous
emission which would dampen the amplitude of the oscillations. The Bloch vector
for the zero detuning case is shown below:
+1
u-v Plane
w-v Plane
Ω eff
v
u
S
-1
S(t) = u(t) + v(t) + w(t)
2 2 2
(2) 0: In this case the Bloch vector may rotate outside of the w-v plane and into the u-v
plane since eff is no longer constrained to the u axis as there is also a component with
magnitude ∆ along the w axis (see figure below). The magnitude of the unit vector must
remain normalised since [u2 + v 2 + w 2 = 1], to accommodate this condition, the magnitude
of w and v components must be reduced and as such complete population inversion is not
possible. The phase of the inversion (w), that is the position of the maxima and minima of
w, will also be different since S will precess differently to that of the of the zero detuning
case. The Bloch sphere for an arbitrary non-zero detuning case is shown below:
The simplest solution to the Bloch equations is to look at the on resonance case with ∆=0,
in this case the equations are reduced to
du
=0
dt
dv
=Ω w
dt
dw
=−Ωv
dt
We will assume that Ω is a constant (this is known as the actual Rabi solution, in the
general case when Ω =Ω(t) there is no analytic solution to the Bloch equations). Taking the
time derivative of the third equation reveals:
d2 w dv
2 = −Ω
dt dt
using equation (2) this can be rewritten as
d2 w
+Ω 2 w = 0
dt 2
This is a homogeneous, linear, second order differential equation, we will try solution
w =e t
Substitution of this solution into the second order de yields the following equation
e t +Ω 2 e t = 0
2
+Ω 2 = 0
2
Therefore
= ±iΩ
Hence we have the solution
w = AeiΩt + Be−iΩt
This can be also represented as the sum of two trigonometric functions:
SCE3337: QMIII 7
R.T. Sang
We can use some boundary conditions to find the solutions to the coefficients A and B. The
factor in the brackets {Ωt } is the angle that the Block vector makes with the w-v axis which
we will call θ then the w vector is then
w = Acos + Bsin
where is the angle that the Bloch vector makes with respect to the w axis in the w-v
plane.
Lets assume that at =0 that the atoms have a an initial inversion of w 0 then the block
vector is completely aligned in along the w axis and hence
w0 = A
w( ) = w0 cos
v ( ) = w0 sin
If all of the atoms are in the ground state at =0 (and hence t=0) then w 0 =-1 at t=0 the
solution for w as a function of is shown below.
π 3π 5π 7π 9π
0.5 Ω
1.5Ω
0
w
2Ω
-0.5
-1
0 400 800 1200 1600
Ωt (degree)
SCE3337: QMIII 8
R.T. Sang
When = , v( ) = −1 hence a positive is for a rotation about the u axis from the -w
2 2
axis to the -v axis. We can see from this graph that Ω , the Rabi frequency, determines the
rate at which transitions are coherently induced between the two atomic levels, the
frequency at which the light field is inducing these transition is known as the Rabi
flopping frequency.
π
16.5 and pulses:
2
The Rabi frequency gives the rate at which transitions are coherently induced between the
two atomic levels. An atom initially in the ground state has w=-1, if then at time t when
Ωt = then w=1, that is the population will be entirely in the upper state. Since we have
chosen Ω to be a constant, only the time varies which allows us to view this excitation
process as a coherent light in the form of a square pulse (as shown below) interacting with
the two level atom.
θ(t)
t
t1 t2
The square pulse, the area under the curve is called the pulse area
=Ω (t1 − t2 ) =
This interaction process will just invert a ground state atom. This type of pulse is termed a
π pulse. It is obvious that if we kept applying n π pulses, where n is an odd integer, the
population would be inverted n times. This property has become particularly useful in
atomic physics experiments.
π
In the case were the pulse is equal to then w= 0 and the atom will be in a superposition
2
state of |e> and |g>.
w w
w=1
Excited State
v=-1
θ=π Superposition state
-v v -v π v
θ=
2
Ground State Ground State
w=-1 w=-1
-w -w
π pulse
π pulse
2
The general case for the Rabi solution (i.e.∆≠0 and Ω(t)=Ω) for w is
∆ +Ω cos ∆ +Ω t ( )
[ )] [ ( )]
2 2 2 2
w = − u0 2
∆Ω
2 1 − cos
(∆ +Ω ) (
∆2 +Ω 2 t − v0 2
Ω
( ∆ +Ω 2 )
sin ∆ +Ω t + w0
2 2
(∆ 2
+Ω
2
)
This solution is somewhat tedious to derive but it is however useful to look at the solution
for the inversion in this case to demonstrate the effect on different detunings. u0 , v 0 , w 0 are
the initial values of the basis Bloch component vectors. It is easy to demonstrate that we
will recover the zero detuning case if we have ∆=0 , u0 = v 0 = 0 and w 0 =-1. The figure
below shows the effect of different detunings on w as a function of Ωt
π 3π 5π 7π 9π
∆=0
0.5
∆=Ω
0
w
∆=2Ω
-0.5
-1
0 400 800 1200 1600
Ωt (degree)
SCE3337: QMIII 10
R.T. Sang
The Bloch equations in the form that we have investigated so far, could not be used for a
real atom since there has been no decay representing spontaneous emission etc,
incorporated in the equations. The spontaneous emission process has implications on the
coherence or the phase of the system. As this process is random it introduces a loss of
phase in the system and hence a loss of coherence between the induced dipole of the atom
and the electric field of the light. We now introduce two damping constants that we label T1
and T2 that modify the optical Bloch equations in the following way:
du u
= −∆v −
dt T2
dv v
= −∆u +Ωw −
dt T2
dw w − weq
=−Ωv −
dt T1
The damping constants T1 and T2 are termed the longitudinal and transverse decay
constants. We see that T1 is related only to the population and hence it is related to
spontaneous emission. The T2 constant is related only to the u and v components and
hence is the dephasing factor between the atomic dipole and the electric field. T1 can be
determined, taking the following approach: Since T1 is related to the population only, then
w must decay exponentially. Assume that the we have applied a _ pulse to the system and
the light field is instantaneously turned off then the Bloch equation for the inversion is
dw w − weq
=−
dt T1
[
w(t) = wo − weq e ] T1
+ weq
Where w0 is the initial inversion. We are looking at times after the π pulse has been
initiated hence w0 = 1. When the light field has been turned off the population will decay
into the ground state hence weq = −1 ie the population is in the ground state hence
t
−
w(t) = [ 2]e −1
T1
Recall from earlier lectures that an excited state decays as e −Γ t hence we deduce that
1
=Γ
T1
b ∝ e −Γt
2
a is the constant for the ground state that does not decay hence:
Γ
− t
a*b ∝e 2
2
Hence T2 =
Γ
The solution to the Bloch equations with damping in the rotating frame can be found using
a Laplace transform technique and was first solved by Torrey for the magnetic resonance
case and as a result the solutions are often referred to as the Torrey solution. The general
solution to the equations is given by
where x(t) represents any of the vectors: w(t), u(t) and v(t). The constants , and s are
found from a secular equation of the Laplace equations and won't be quoted here as the
solution takes some time to derive (for those that are interested the book "Molecules and
Radiation" by J. Steinfeld outlines the solution). The constants A, B, C and D are found
via boundary conditions. Let's look at one well known case for ∆=0, in this case the
constants for general solution is
1 Γ
= =
T2 2
1 1 1 1 Γ 3
= + = Γ + = Γ
2 T1 T2 2 2 4
1 1 1 3Γ 2
s = Ω − + = Ω −
2 2
4 T1 T2 4
Assuming that we start with ground state atoms at t=0 then the solution to the equations are
u=0
3
− Γt
v =−e 4
sin(Ω' t )
3
− Γt
w =−e 4
cos(Ω' t )
3Γ 2
Ω' = Ω 2 −
4
SCE3337: QMIII 12
R.T. Sang
A graphical solution to the inversion with Ω >>Γ approx (Ω = 400Γ ) is shown below
0.5
0
w
-0.5
-1
0 400 800 1200 1600
Ωt
The population oscillates as a trig function since the field and the dipole are in phase and as
such is a coherent interaction, but has a exponential decay associated with it. It is clear
from this that the damping stops the oscillation of the population due to the spontaneous
emission process. As a result there is dephasing effect of the driving electric field with the
dipole and the oscillations are damped at an exponential rate. At large time the field and the
dipole are no longer coherent and as such we would expect that the oscillations would
completely die off. Since the driving field is strong, at long times, the population is driven
to is steady state population distribution, which for a two level atom under strong light
interaction, is half in the ground state and half in the excited state (see Lectures 9 and 10).
The oscillatory behaviour shown at early time is known as Optical Nutation and has
been observed experimentally at Griffith University in the Laser Atomic Physics
Laboratory.