Sie sind auf Seite 1von 104

School of Physical Sciences

PH 502
Wavemechanics & Quantum Physics

COURSE SCRIPT

Dr. Peter Blmler

Dr. P. Blmler: PH 502 Wave Mechanics

Syllabus

Syllabus for Part II (Dr. Blmler)


7. The Foundations and Postulates of Quantum Mechanics:
7.1
7.2

Philosophy and Concept of Axiomatic Physics


Operators and Eigenfunctions:
7.2.1
Properties of Operators
7.2.2
Eigenfunctions and Eigenvalues
7.2.3
Linear Combinations
7.2.4
Orthogonality, Normalisation and Orthonormality
7.2.5
Hermiticity
7.2.6
Commutators
7.2.7
List of Important Operators

7.3

Postulates (Axioms) of Quantum Mechanics


7.3.1
7.3.2
7.3.3
7.3.4

8.

Wavefunction
Physical Systems
Schrdingers Equation
Probability and Expectation Value

7.4

Uncertainty Principle (principle of indeterminacy)

7.4

Dirac Notation

Confined Particles
8.1

Recapitulation: 1D-Problem particle in a box


E.1

9.

Excursus: Separation of variables

8.2

2D-Problem: Particle in a Square Well


8.2.1 Example Quantum Corrals

8.3

3D-Problem particle in a real box, Degeneracy

The Harmonic Oscillator


9.1
9.2

Classical Description
1D Harmonic Oscillator in QM
9.2.1
Solution of Hermites Differential Equation
9.2.2
Correspondence Principle

9.3

The 2D Harmonic Oscillator

9.4

The 3D Harmonic Oscillator

10. Rotational Motion


10.1

Classical Rotation in a 2D Plane

10.2

Some Quantum-Mechanical Considerations


E2: Excursus: Co-ordinate Transforms of the Laplacian Operator
E2.1
Method 1: Explicit Solution for 2D Circular Co-ordinates
E2.2
Method 2: Curvilinear Co-ordinate Transforms

10.3

Exact Solution for the Rigid Rotator

10.4

Particle Rotating on a Sphere

10.5

Angular Momentum

11. The Hydrogenic Atom


11.1

Motion in a Coulomb field

11.2

Solution of the Radial Differential Equation

11.3

The Complete Solution: Atomic Orbitals

11.4

Linear Combinations: Hybrid Orbitals

11.5

Quantum numbers, Orbital Shapes and Degeneracy


11.5.1 The Periodic Table of Elements

11.6

Electron Spin
11.5.2 Selection Rules and Atomic Spectra

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

page: 1

Dr. P. Blmler: PH 502 Wave Mechanics

Useful physical constants

Useful constants
Planck constant:

6.62607610-34 Js = 4.13566910-15 eV s

h
h=

h
2

1.05457310-34 Js = 6.58212210-16 eV s

Speed of light (in vac.):

2.99792458108 m/s

Electron mass:

me

9.10939010-31 kg
510.9991 keV/ c2

Proton mass:

mp

1.67262310-27 kg
938.2723 MeV/ c2

Neutron mass:

mn

1.67492910-27 kg
939.5656 MeV/ c2

Fundamental charge:

1.60217710-19 C

Compton wavelength:

C =

Boltzmann constant:

kB

1.38065810-23 J/K

Vacuum permittivity

8 .854188 10 12 C2 /(Jm )

4 0

1 .112650 10 10 C2 /(Jm )

h
me c

2.4263105810-12 m

combinations:
hc = 1.986410-25 Jm = 1239.8 eV nm
h c = 3.161510-26 Jm = 197.33 eV nm

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

page: 2

Greek Alphabet/Units

Dr. P. Blmler: PH 502 Wave Mechanics

page: 3

The Greek alphabet


A

alpha

nu

beta

xi

delta

omicron

gamma

pi

epsilon

rho

zeta

s, V

eta

tau

q, J

theta

upsilon

iota

kappa

chi

lambda

psi

mu

omega

sigma

f, phi

Units
Prefix

Exponent Symbol

deci
centi
milli
micro
nano
pico
femto
atto
zepto
yocto

-1
-2
-3
-6
-9
-12
-15
-18
-21
-24

d
c
m

n
p
f
a
z
y

Prefix

Exponent Symbol

deca
hecto
kilo
mega
giga
tera
peta
exa
zetta
yotta

1 = 10-10 m

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

1
2
3
6
9
12
15
18
21
24

da
h
k
M
G
T
P
E
Z
Y

Dr. P. Blmler: PH 502 Wave Mechanics

Course Summary / Spine

page: 4

COURSE SPINE
What is that?
This section is designed to give you a coarse outline of the learning outcomes of each section. You
might find this useful for exam preparations and revision.
From the PDF-File you can always link to
Blue colour provides links to the particular section!
The following symbols mean:

Link to detailed discussion

Recap of material from previous courses

Mathematical detail, method or solution

Illustration in a MAPLE program

Workshop or exercise

Recap pervious knowledge!


Classical wave mechanics (PH 301)

Qualitative description of quantum mechanics (PH 301)

Partial differential equations (PH 501)

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

7.

Course Summary / Spine

The Foundations and Postulates


of Quantum Mechanics:

7.2 Operators and Eigenfunctions


Definitions and examples

To know:
1. Properties of operators, notation
2. Eigenfunctions and eigenvalues, degeneracy, physical meaning
3. Linear combination
4. Orthogonality, normalisation and orthonormality
5. Hermiticity, real eigenvalues
6. Commutators
7. Important operators: x , y , z , p x ,p y , p z , L x , L y , L z , H , K , U

Exercises and workshops

7.3/7.4

The Postulates of Quantum Mechanics:


Definitions and examples

To know:
1. Wavefunction, probability interpretation, normalisation
2. Hermitian operators and physical observables
3. Expectation values and deductions
4. Schdinger equations
5. Uncertainty principle (as in (7.4.1))

Exercises and workshops

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

page: 5

7.2 Operators

Dr. P. Blmler: PH 502 Wave Mechanics

page: 6

7.2. Operators and Eigenfunctions:


What is an operator?
An operator is a symbol that tells you to do something to whatever follows the symbol.
More mathematically it can be understood as a symbol that represents a rule to transform a following
function into another function.
and an expression like A
f will be referred to as A
The notation in this script for operators is A
representing a vectoroperating on f . For rigorousity we will very often use symbols like A
operator. The underline shall simply remind you that this operator may act differently on individual
co-ordinates (or vector components).

Comparison: Scalars, Functions, Vectors, Matrices, Operators


Scalars and Functions:
A linear function converts a scalar into another:
b = f(a)

Vectors and Matrices:


A matrix is a linear vector function, thus it converts a vector into another:
b=Ma
b1
m11

example: b = b2 = M a = mik ak = m21


b
m
3
31

m12
m22
m32

m13 a1 m11 a1 + m12 a 2 + m13 a3


m23 a2 = m21 a1 + m22 a2 + m23 a3


m33 a3 m31 a1 + m32 a2 + m33 a3

Functions and Operators:


An operator is a function of higher order, thus it generally converts a function into another function:

a( x )
b( x) = M
In quantum mechanics, physical observables (e. g. energy, position, linear and angular momentum)
are represented mathematically by operators. For example the total energy (kinetic energy K plus
, where H is called Hamiltonpotential energy U) is represented by an operator H = K + U
operator or simply Hamiltonian. It is defined as:

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

7.2 Operators

Dr. P. Blmler: PH 502 Wave Mechanics

page: 7

h2 2

H=
+U
2m
We will later derive this formula. The link to experimentally assessable values of the energy (its
average value) of a system described by H and in a state described by a wavefunction (r) is given
.
by its expectation value H

= * (r) H
(r ) dr
H

7.2.1 Properties of Operators:


Most of the properties of operators are obvious. Nevertheless, they are summarised for
completeness:
and B :
Sum and Difference: of two operators A

(A + B ) f = A f
(A B ) f = A f

The product of two operators is defined by:

f
+ B
f
B

[ ]

B
f =A
B
f
A

(7.2.1)

(7.2.2)

on f then operate with A


on the result of B
f.
which means: First operate with B
f =B
f
Equality: Two operators are equal if A

holds for all functions f.

(7.2.3)

Identity: The identity operator I does nothing (multiplies with 1 or adds 0)

(7.2.4)

n is defined as n successive applications of A


, e.g.
The nth power of an operator A

[ ]

2f = A
A
f =A
A
f
A

(7.2.5)

) is defined by a Taylor expansion:


The exponential of an operator exp( A
2
3
) = e A = I + A
+ A + A +K
exp( A
2!
3!
The associative law holds for operators:
Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(7.2.6)

Dr. P. Blmler: PH 502 Wave Mechanics

7.2 Operators

page: 8

( ) ( )

B
C
f = A
B
C f
A

(7.2.7)

The commutative law does not generally hold for operators!

B
f B A
f
A

(7.2.8)

see commutators!

Examples:
f ( x) = f ( x) + c
1) Addition of a constant: A
d
2) Derivative (wrt x): D f ( x) =
f ( x ) = f ( x ) (could be named prime operator or dot
dx
operator for derivatives wrt time)
3) For D operating on f(x) =exp(ax):
d
D f(x) = D exp (ax ) =
exp (ax ) = a exp (ax ) = a f(x)
dx
4) For D 2 operating on f(x) =exp(ax 2):
exp ax 2 = D
d exp ax 2 = D
2 a x exp ax 2
D 2 f(x) = D D
dx

d
= 2a
x exp ax 2 = 2a exp ax 2 + ax 2 exp ax 2 = 2a 1 + ax 2 f(x)
dx
d2
D 2 corresponds to the second derivative D 2 =
dx 2

[ ( )]

( ) [
( ( )) ( ( )

( )]
( ) (

HINT: To calculate pure operator expressions it is good advice to include a virtual function
(e.g. f) to remind you that you have to operate!
d
Example: For the two operators D
and x x (meaning just multiply with x) calculate
dx
:
D x - x D
[x f ] = D
[xf ] = f + xf
D
f = x [ f ] = xf
x D
x - x D
f = f and D
x - x D = 1
hence : D

D x :
x D :

[ ]
(

and not 0! This illustrates that operators do not generally commute.

Do not forget to divide by f after your operations!!!

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

7.2 Operators

Dr. P. Blmler: PH 502 Wave Mechanics

page: 9

Linear Operators:
Almost all operators encountered in quantum mechanics are linear operators.
Definition: A linear operator has to satisfy the following two conditions:
( f + g)= A
f +A
g
A
(af ) = aA
f
A

(7.2.9)

where a is a constant and f and g are functions.


Examples:
1) D is a linear operator:
d
df ( x) dg ( x)
g (x )
D ( f ( x) + g ( x ) ) = ( f ( x) + g ( x ) ) =
+
= D f ( x) + D
dx
dx
dx
(af ( x ) ) = d (af ( x ) ) = a df ( x ) = a D f ( x)
and D
dx
dx
2)

S ( )2 -an operator that squares- is not linear:


S ( f ( x) + g ( x ) ) = ( f ( x ) + g ( x) )2 = f 2 ( x) + 2 f ( x) g ( x) + g 2 ( x ) f 2 ( x ) + g 2 ( x)
(af ( x ) ) = (af ( x) )2 = a 2 f 2 ( x) a S f ( x)
and S

7.2.2 Eigenfunctions and Eigenvalues:


(German: eigen = own, characteristic, individual, belonging to, specific, inherent!)
f = af the functions f are called eigenfunctions of the operator A
, which scale the functions
For A
by the eigenvalues a.

f = af
A
operator

eigenfunction

eigenvalue

(7.2.10)
eigenfunction

is a function f such that the application of A


on f
In other words: An eigenfunction of an operator A
gives f again, times a constant, the eigenvalue, a.
is a linear operator with an eigenfunction g (and eigenvalue b) , then any multiple of g is also an
If A
.
eigenfunction of A
(7.2.11)
g = bg then A
(cg )
Proof: A

( 7. 2. 9 )

g = cb g = dg
cA

Hence cg is an eigenfunction with the eigenvalue d = cb


Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

7.2 Operators

Dr. P. Blmler: PH 502 Wave Mechanics

page: 10

d ax
Example: exp( ax ) = e ax is eigenfunction of D because
e = a e ax
dx
2
2
d ax 2
( e ax is not an eigenfunction, because
e
= 2ax e ax ).
dx

Degeneracy: If there is more than one (n = 1,2,..m) eigenfunction, n , to an eigenvalue, a,


they are called m-fold degenerate.

(7.2.12)

Deduction:
Any linear combination of a set of degenerate eigenfunctions is also an eigenfunction.

Proof:

1 , 2 , 3 ,... m are all eigenfunctions of A


having the
= a for n = 1, 2, ...m . Hence, this state is m-fold degenerate.
A
n

same

eigenvalue

a.

The linear combination of a set of degenerate eigenfunctions is given by


m

cn n
n =1

g =A

c n n =

n =1
n =1

= a
cn A
n
{
a n

cn n = ag

n =1

qed!

Physical meaning:
If a system is in an eigenstate of observable A (i.e., when the wavefunction is an eigenfunction of the
) then the expectation value of A is the eigenvalue of the wavefunction.
operator A

7.2.3 Linear Combinations:


A function g (which is no eigenfunction) can be expanded in terms of eigenfunctions n of an
(A
= a ):
operator A
n

g=

c
n

(7.2.13)

The new function g is a linear combination of eigenfunctions with the combination coefficients c n
(which might be complex!). (in mathematical terms: The eigenfunctions of an operator form a
complete set, from which other functions can be build up)
Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

7.2 Operators

Dr. P. Blmler: PH 502 Wave Mechanics

page: 11

on a (general) function g, which is not one of


From this we can deduce the effect of an operator A
the eigenfunctions n :
g
A

( 7 .2 .13 )

c = c A = c{a = b
n

n n
bn

= g~

Hence this generates another function g~ , which also is a linear combination of the eigenfunctions;
however with different coefficients.
A special case is given we use linear combinations of degenerate eigenfunctions (see chapter
11.4):

7.2.4 Orthogonality, Normalisation and Orthonormality:


Definitions:
1) Orthogonality:
Two functions n and m are orthogonal if the integral over all space (V) is zero.

*
m n

dV = 0

(7.2.14)

2) Normalisation:
A function n is normalised if the integral over all space (V) is one.

*n n dV = 1

(7.2.15)

3) Orthonormality:
If both properties are fulfilled, the functions are called orthonormal.

0 if n m
*m n dV = nm =
1 if n = m

where nm is called Kronecker-delta.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(7.2.16)

7.2 Operators

Dr. P. Blmler: PH 502 Wave Mechanics

page: 12

7.2.5 Hermiticity:
Definition of Hermiticity:
is called hermitian, if it fulfils
An operator A

dV = * A

*m A
n
n m dV =

* * dV =
nA
m

* dV
n A
m

(7.2.17)

Deduction:
A hermitian operator has real eigenvalues

(7.2.18)

Proof of Deduction (7.2.18):


is assumed to be hermitian. 1) A
= a
A
* * = a * *
2) A
On 1) we operate with * from the left side and on 2) with from the left,too:
= * a = a *
1) * A
* * = a * * = a * * = a * *
2) A
we can integrate over all space or simply recognise, that the left sides of 1) and 2) are identical
was assumed to be hermitian. Hence the right sides must be
(definition of hermiticity) because A
identical, which they are only when a * = a , which is only fulfilled when a is real (qed!).

Physical importance:
All observables (real numbers) correspond to hermitian operators, but -of course- not all hermitians
to observables. (see also expectation values)

(7.2.19)

Deduction:
Eigenfunctions corresponding to different eigenvalues of hermitian operators are orthogonal.

Proof of Deduction (7.2.19):


=a
Given: A
n
n n

A
2) A

1)

*
m

*
n

and

dV = a

=a
A
m
m m

dV = a

with a n am and a n , am

dV = *m a n n dV = an *m n dV
*
n m m

*
n

dV

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

IR

Dr. P. Blmler: PH 502 Wave Mechanics

7.2 Operators

page: 13

Subtract the complex conjugate of 2) from 1):

*
*
*
dV * A

*m A
n
n m dV = a n m n dV am n m dV

to be hermitian, the two terms on the left side are identical, hence:
Because we have assumed A

a n *m n dV a*m n *m dV = an a *m

) *m n dV

= 0

is hermitian, the eigenvalues an and am are real hence a * = a a , hence the


Because A
m
m
n
difference of the eigenvalues is always different from zero and it follows
was to prove (qed!).

*
m n

dV = 0 , which

7.2.6 Commutators:
The key feature of operators -and the reason why they are used in quantum mechanics- is the fact
that they do not generally commute (as we will later see with the uncertainty principle). Another
approach -originally proposed by Heisenberg- used matrices, which do not commute either.
Because of this it is useful to introduce a quantity which expresses this essential property. It is called
the commutator operator [ , ] and defined as follows:
Definition of commutator operator:

,B ] A
B
B
A

[A

(7.2.20)

Notation:
B ] = 0 it is said that A
and B commute!
If [ A,
B ] 0 it is said that A
and B do not commute!
If [ A,
( example above)

Commutator algebra:

for linear operators (see 7.2.9):

,A
] = 0
[A

(7.2.21)

,B
] = [B
,A
]
[A

(7.2.22)

, bB ] = b [ A
,B
]
[A

(7.2.23)

,B
C ] = A
B
C
B
C
A
=A
B
C
B C
A
B
A
C
+B
A
C

[A
B
C
B
A
C
+B
A
C
B
C
A

=A

)
(7.2.24)

+B
]
,B
]C
[A
,C
= [A
B , C
] = [A
,C
]B
+A
[B , C
]
[A

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(7.2.25)

7.2 Operators

Dr. P. Blmler: PH 502 Wave Mechanics

) (

page: 14

,B
+C
]= A
B + C
B + C A
=A
B + A
C
B
A
C
A = [ A
,B
] + [A
,C
]
[A

(7.2.26)

+ B , C
] = [A
,C
] + [A
,B
]
[A

(7.2.27)

7.2.7 List of important operators:


The following table gives a list of typical observables in their classical definition and the
corresponding operators

observable

classical

QM-operator

r = ( x, y , z )

r = (x , y , z ) = ( x, y, z )

x
px
t

y
p = p y = mr& = m
t
p
z

z
t

p x
x

h
=
p = p y = ih =
y
i
p

z
z

position r = ( x, y , z )
linear momentum

p = px , p y , pz

angular momentum
L = Lx , Ly , Lz

see also below

Lx

L = Ly = r p
L
z

1
p 2 + p 2y + p z2
2m x

kinetic energy, K

K=

potential energy, U

U(x,y,z), e.g. U ( x) =

total energy, E

L x

L = L y =

Lz

E=K+U

k 2
x
2

y z
y p z z p y
z
y

h
z p x x p z = z x
i

z
x p y p
x y
x
y
x
y

h2 2
h 2 2
2
2
K =
=
+
+
2m
2m x 2 y 2 z 2
= U ( r ) e.g. U = k r 2
U
2
2

h 2
H =
+U
2m

In section 10.5 we will derive


sin cot cos
L x

h
h

L = L y = i cos cot sin


where L z = i is to remember!

Lz

continue to next section (7.3)


Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

7.3

7.3 Postulates

page: 15

Postulates (Axioms) of Quantum Mechanics:

The postulates cannot be proven. They are axioms of quantum mechanics. However, they are
plausible in the context of defining a fixed framework (mathematical rigorous and without
contradiction) for the description of quantum systems.
Note that different authors very often use different formulations and orders of these postulates!

7.3.1 The wavefunction


POSTULATE 1:
1a) Wavefunction:
The state of a physical system is fully described by the wavefunction (=state function):

( r1 , r 2 ,...; t )
with r1, r2,... as the co-ordinates of particle 1, 2, ... and time t.
In this lecture we will deal exclusively with single particles and time-independent problems,
hence the wavefunction simplifies to:

(r )

(7.3.1)

The wavefunction must be single-valued, continuous (differentiable) and finite.


1b) There is a probabilistic interpretation (Born-Hypothesis) to the wavefunction:
The amplitude of (x) is related to the probability of finding the particle at x. The amplitude is
given by *, thus the probability P(r)dr of finding the particle in the volume dr is

probability density function:

P(r ) d r *(r ) ( r ) d r

(7.3.2)

and is consequently normalised:

Normalisation condition:

* ( r ) ( r ) d r 1

This can be interpreted as finding the particle somewhere is one.


Comment: However it is customary to also normalise many-particle wavefunctions to 1.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(7.3.3)

7.3 Postulates

Dr. P. Blmler: PH 502 Wave Mechanics

page: 16

7.3.2 Physical Systems


POSTULATE 2:
2a) A specific state of a system, in which a physical observable A has one specific value a,
, of which then
must be described by (at least) one eigenfunction n to the operator A
an is an eigenvalue.

= a
A
n
n

(7.3.4)

) the
Hence, in any physical measurement of the observable A (associated with operator A
only values that will ever be observed are the eigenvalues a, which satisfy eq. (7.3.4).
A particular and extremely important case is the Schrdinger equation, which
corresponds to the total energy as the observable. (cf. heuristic derivation)
in
2b) Every (classical) physical observable corresponds to a linear hermitian operator A
quantum mechanics, because it must yield real eigenvalues!
(7.3.5)
(see 7.2.5 and 7.2.7)

This postulate captures the central point of quantum mechanics: The values of dynamical variables
can be quantized!
with eigenvalue a, then any measurement of the
If the system is a state, which is an eigenstate of A
quantity A will result in value a.
Although measurements must always yield an eigenvalue, the state doesnt have to be an eigenstate
.
of A
(see
An arbitrary state g can be expanded in a complete set of eigenfunctions (or eigenvectors) of A
linear combinations in 7.2.3).
=a
Hence,
i with A
i
i i
n

and with eq. (7.2.13)

g=

c
i

(7.3.6)

where n may go to infinity. In this case we only know that the measurement of A will yield one of the
values ai, but we do not know which one in particular. However, we do know the probability that
eigenvalue am will occur. It is the squared magnitude of the coefficient c m

(prove see next part)

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

7.3 Postulates

Dr. P. Blmler: PH 502 Wave Mechanics

page: 17

7.3.3 Expectation Values:


POSTULATE 3:
When a system is described by a wavefunction , the mean value of the observable A is equal to
of the corresponding operator A
. The expectation value is defined as
the expectation value A

A dV
*
dV
*


or for normalised wavefunctions: A

(7.3.7)

A dV
*

(7.3.8)

Deductions:

(corresponding to the observable of interest


When m is an eigenfunction of the operator A
A), the determination of A always yields one result, the eigenvalue am.
(7.3.9)
, a single measurement of A yields a single result, which is
When is not an eigenfunction of A
. The probability that a particular eigenvalue am is measured is
one of the eigenvalues am of A
2

equal to c m , where cm is the coefficient of the eigenfunction m in the expansion of .(7.3.10)


Proof of Deduction (7.3.9):
(A
= a ) and is normalised then
If is eigenfunction to A
=
A

A dV = a dV = a dV = a
*

Hence a series of experiments to determine A will have the average value a.

Proof of Deduction (7.3.10):


Now the system is considered to be in no eigenstate, but rather in a state expressed by a linear
= a
combination of eigenfunctions: =
c n n with A
n
n n

=
A
=

c A
c dV =
dV
c *m cn *m A
m
m
n
n
n

m n
m
n

c*m cn *m a n n dV
m

c*m cn an *m n dV
m

as hermitian, because we want to calculate the


In a way it only makes sense to consider A
expectation value of a (physical, real) observable. Then the eigenfunctions m and n are
Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

7.3 Postulates

Dr. P. Blmler: PH 502 Wave Mechanics

page: 18

orthogonal (deduction 2b) and cancel for all n m . In other words, only the terms with n = m
remain. If we further assume normalised wavefunctions, we yield

c*m cm am *m m dV = c*m cm am = cm 2 am

=
A

This is a sum of eigenvalues for that state, weighted by the square magnitude of the combination
coefficients of the probing wavefunction.
Example: Assume a wavefunction which is a linear combination of two eigenstates
with the (real) eigenvalues a and b. Hence,
(eigenfunctions) of the linear, hermitian operator A
= a and A
= b . Calculate the expectation value of A:
= c a a + cb b where A
a
a
b
b

[c + c ] A [c + c ]dV
+c A
] dV = [c + c ] [c a + c b ] dV
= [c + c ] [c A

= ac c dV + bc c dV + ac c dV + bc c dV
14243
142
14243
14243
4 43
4

= *A
dV =
A
a a

*
a a

*
a

b b

*
a b

=1 (normal.)

*
a b

*
b a

= 0 ( 7 .2 .19 )

*
b

= 0 ( 7. 2. 19 )

*
b b

*
b b

=1 (normal.)

= a ca 2 + b cb 2
The average (or expectation) value of the observable A is a weighted average of eigenvalues, where
the weights correspond to the square of the coefficients of the eigenfunctions the wavefunction is
composed from.

Tools:
=
If the mean value of A is defined by A

A dV
*

2
Consequently the mean value of A2 is defined by A

A dV
*

It is also useful to define a mean square deviation:

(A A )

(A A ) dV = (A A )(A A ) dV

= (A
A )(A A ) dV = A A A A A + A

= A
2 A A + A dV
= A
dV 2 A A dV + A dV
2

dV

(7.3.11)

( )

2 2 A
2+ A
2 = A
2 A
2 A
A
2
= A

is also called deviation operator and A


rootThe last definitions are common and useful ! A
mean-square-deviation operator or rms-operator.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

7.3 Postulates

page: 19

7.3.4 The Schrdinger-Equation:


The central equation of quantum mechanics has to be accepted as a postulate as well! However, you
can find some plausible and heuristic arguments in an attachment

POSTULATE 4:
The wavefunction or state function of a system evolves in time according to the

time-dependent Schrdinger-equation:

(r , t ) = ih (r , t )
H
t

(7.3.12)

However, in the following chapters we will exclusively work with time-independent systems, for
which the time-independent Schrdinger-equation has been already introduced in postulate 2 eq.
(7.3.4):

time-independent Schrdinger-equation:

2
h
(r , t ) =
( r , t ) = E ( r , t )
2 ( r , t ) + U
H

2m

(7.3.13)

Comment:
These are the postulates to an extend necessary to understand the course. Other postulates e.g.
concerning the symmetry of the wavefunction or alternatively the Pauli-exclusion principle may be
defined.

continue to next section (7.4)

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Heuristics: Schrdingers equation

Dr. P. Blmler: PH 502 Wave Mechanics

page: 20

Excursus: The Time-Independent Schrdinger Equation


We start with the one-dimensional classical wave equation,
2u
x 2

1 2u

(1)

v 2 t 2

By introducing the separation of variables u ( x, t ) = ( x ) f (t )


we obtain

2 ( x)

f (t )

x 2

( x) 2 f ( t )

v2

t 2

(2)

If we introduce one of the standard wave equation solutions for f (t ) such as exp( it ) (the constant
can be taken care of later in the normalisation), we obtain
d 2 ( x )
dx 2

v2

( x)

(3)

Now we have an ordinary differential equation describing the spatial amplitude of the matter wave as
a function of position. The energy of a particle is the sum of kinetic and potential parts
E =

p2
+ U (x )
2m

(4)

which can be solved for the momentum, p, to obtain


p = 2 m( E U ( x ) )

(5)

Now we can use de Broglies formula to get an expression for the wavelength
=

h
=
p

h
2m( E U ( x) )

(6)

The term 2 /v 2 in equation (3) can be rewritten in terms of if we recall that = 2 and
= v :
2
v

4 2 2
v

4 2
2

2m(E U ( x) )
h2

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(7)

Heuristics: Schrdingers equation

Dr. P. Blmler: PH 502 Wave Mechanics

page: 21

When this result is substituted into equation (3) we obtain the famous time-independent
Schrdinger equation
d 2 ( x )
dx 2

2m
h2

(E U ( x) ) ( x)

= 0

(8)

which is almost always written in the form


h 2 d 2 ( x )

+ U (x ) (x ) = E (x )
2m dx 2

(9)

This single-particle one-dimensional equation can easily be extended to the case of three dimensions,
where it becomes

h2 2
(r ) + U ( r ) ( r ) = E ( r )
2m

(10)

A two-body problem can also be treated by this equation if the mass m is replaced with a reduced
mass .
It is important to point out that this analogy with the classical wave equation only goes so far. We
cannot, for instance, derive the time-dependent Schrdinger equation in an analogous fashion
(for instance, that equation involves the partial first derivative with respect to time instead of the
partial second derivative). In fact, Schrdinger presented his time-independent equation first, and
then went back and postulated the more general time-dependent equation.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Heuristics: Schrdingers equation

Dr. P. Blmler: PH 502 Wave Mechanics

page: 22

The Time-Dependent Schrdinger Equation


We are now ready to consider the time-dependent Schrdinger equation. Although we were able to
derive the single-particle time-independent Schrdinger equation starting from the classical wave
equation and the de Broglie relation, the time-dependent Schrdinger equation cannot be derived
using elementary methods and is generally given as a postulate of quantum mechanics. It is possible
to show that the time-dependent equation is at least reasonable if not derivable, but the arguments
are rather involved
The single-particle three-dimensional time-dependent Schrdinger equation is
ih

2 ( r , t )
t 2

h2 2
=
( r , t ) + U (r ) ( r , t )
2m

(11)

where U is assumed to be a real function and represents the potential energy of the system. Wave
Mechanics is the branch of quantum mechanics with eq. (11) as its dynamical law. Note that
equation (11) does not yet account for spin or relativistic effects.
Of course the time-dependent equation can be used to derive the time-independent equation. If we
write the wavefunction as a product of spatial and temporal terms, (r , t ) = ( r ) f (t ) , then eq.
(11) becomes
h2 2

= f (t )
+ U ( r ) ( r )
2m

(12)

ih 2 f ( t )
1 h2 2
=

+
U
(
r
)

( r )
f (t ) t 2
(r ) 2m

(13)

(r ) i h

or

2 f (t )
t 2

Since the left-hand side is a function of f(t) only and the right hand side is a function of r only, the
two sides must equal a constant. If we tentatively designate this constant E (since the right-hand side
clearly must have the dimensions of energy), then we extract two ordinary differential equations,
namely
1 d 2 f (t )
iE
=
2
f (t ) dt
h
and

h2 2
(r ) + U ( r ) ( r ) = E ( r )
2m

(14)

(15)

The latter equation is once again the time-independent Schrdinger equation. The former
equation is easily solved to yield

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

Heuristics: Schrdingers equation

page: 23

Et
f (t ) = exp i
h

(16)

The Hamiltonian in eq. (15) is a Hermitian operator, and the eigenvalues of a Hermitian operator
must be real, so E is real. This means that the solutions f(t) are purely oscillatory, since f(t) never
changes in magnitude (recall Euler's formula e i = cos i sin ). Thus if
Et
(r , t ) = ( r ) exp i
h

(17)

then the total wave function (r , t ) differs from (r ) only by a phase factor of constant magnitude.
There are some interesting consequences of this. First of all, the quantity (r , t )

is time

independent, as we can easily show:


Et
Et
(r , t ) 2 = * (r , t ) ( r , t ) = * ( r ) exp i (r ) exp i = * ( r ) ( r )
h
h

(18)

Secondly, the expectation value for any time-independent operator is also time-independent, if
(r , t ) satisfies eq. (17). By the same reasoning applied above,
=
A

( r , t ) dV =
*( r , t) A

(r ) dV
* (r ) A

(19)

For these reasons, wave functions of the form (17) are called stationary states. The state (r , t ) is
stationary, but the particle it describes is not!
Of course eq. (17) represents a particular solution to eq. (11). The general solution to eq. (11) will
be a linear combination of these particular solutions, i.e.
(r , t ) =

E jt

c j j ( r ) exp i

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(20)

Dr. P. Blmler: PH 502 Wave Mechanics

7.4 Uncertainty Principle

page: 24

7.4 The Uncertainty Principle:


The uncertainty principle (or better: principle of indeterminacy) is NOT strictly a postulate!
A typical (but weak) definition is:

Uncertainty principle (principle of indeterminacy):


Denies the ability to determine simultaneously and with arbitrarily high precision, the values of
particular pairs of observables, named complementary variables. (e.g. position and linear
momentum):
h
p x x
(7.4.1)
2
=
where A

2 A
2 defines the uncertainty of the observables (specifically the root mean
A

square deviations from the mean.


Why is this a weak definition? We still do not know what observables are complementary and why
their uncertainty is h / 2 .
Can one measure an eigenvalue/observable precisely? What are the conditions for this?
state a:
when A is measured, one determines precisely a.
state b:
when B is measured, one determines precisely b.

At the moment we claim:


Deduction:

(7.4.2)

If two eigenstates are to have simultaneously precisely defined values, their corresponding operators
B
= B A
or using definition (7.2.20): [ A
,B
]= 0
must commute. That is A

Proof of deduction (7.4.2):


= a and B = b or is simultaneously eigenfunction to A
and B .
eigenstates: A
B
= A
b = bA
= ba = ab = aB = B
a = B
A

hence, A
B
= B A
or A
B
B A
=0
and A
But what we have deduced so far is only the necessary condition. We have not proven the
B
B A
= 0 implies that A
and B have the same eigenfunction. Or if
sufficient part. This is A
= a is also eigenfunction to B for A
B
B A
=0?
A
= a , then B A
= B a = aB

Given: A

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

7.4 Uncertainty Principle

Dr. P. Blmler: PH 502 Wave Mechanics

page: 25

B
= B A
we can rewrite:
because of A
= A
B
= aB

B A

( ) ( )

B
= a B
A

= a it directly follows, that B because both equations must be


when compared with A
fulfilled. If we name the proportionality constant b or B = b , we have proven the claim.
Since we have not made any specific reference in our proof, this is true for all linear operators!
Note: For simplification we have been not really mathematically rigorous. The exact proof it is necessary that

is member

are also eigenfunctions that fulfil the


of a complete set and non-degenerate. Otherwise linear combinations of B
condition.

Now we can appreciate the importance of operators, the fact that they do not generally commute
and why a commutator was defined!!
For a more general description of the uncertainty principle, we want to investigate now the case,
that the two operators do not commute and suppose this property in the most general form:
,B
] = iC

[A
Furthermore, the system is in an arbitrary state described by , which is not necessarily an
or B .
eigenfunction ofA
When we measure the system, we get
=
dV
A
* A

=
B

and

B dV
*

but we are more interested in the spread of values around each of these mean values (or in other
words, how broad the values are distributed, or how certain the measurement is). Hence, we use
another deviation operator
B

A
and b = B
a A
and calculate its commutator:
A
,B
B
]= A
A
B B
B B

[ a , b ] = [ A

B B
A
A
B
+ A

= A

)( ) ( )(A A )
A
A
B
B A
+ A
B
)
B ) (B

B
B
A
= [A
,B
] = iC
=A
and B be hermitian, hence a and b
Because we are only interested in real outcomes, we let A
are also hermitian and we define an integral
I=

(a ib )

) (

2
dV where a i b = a * +i b * * a ib

with as a real but arbitrary parameter. This integral relates the two deviations and measures how
much a is different from b (via ) and it is always greater then zero (because of the square).
I=

(a ib )

dV =

(a * *)(a ib )dV + i (b * *)(a ib ) dV 0

now we use the hermiticity of a and b :

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

7.4 Uncertainty Principle

Dr. P. Blmler: PH 502 Wave Mechanics

page: 26

( *a )(a ib )dV + i ( * b )(a ib )dV


= * (a + ib )(a ib )dV = * [ a + b ia b + i b a ] dV

I=

= 2 a 2 + b 2 i [ a , b ] = 2 a 2 + b 2 + C

This can be rearranged to


2


2
C
C
+ b 2
I = a 2 +
0

2

2
2

a
4

This statement is true for all real , but we are searching for the minimum deviation (minimal
uncertainty). Because only scales the first term, the minimum is for this term to vanish or

C
=
.
2

2 a
2
The minimum deviation is hence found for b

2
C
4 a 2

a 2 b 2
A

Because a A

1 2
C
4

( )

A
2 = A
2 A
2 A
A
2
, a 2 = A

we can rewrite the equation using rms-operators:

(A )2 (B )2

1 2
C
4

B
1 C

A
2
This is a more general form of the uncertainty principle!

Uncertainty principle (principle of indeterminacy):


and B is different from zero,
When the commutator of two operators A

,B
] = iC

[A

(7.4.3)

the observables to these operators are complementary. and hence cannot be determined with
arbitrary precision. The highest precision that can be achieved is then given by:

B
1 C

A
2

=
where A

2 A
2 defines the uncertainty of the observables.
A

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(7.4.4)

Dr. P. Blmler: PH 502 Wave Mechanics

7.4 Uncertainty Principle

page: 27

Examples:
1) Calculate the uncertainty of space and associated linear momentum (e.g. in x-direction):
h
h
h
h h
h
[ x , p x ] = x

x =x
x
= = ih
i x
i x
i x
i x i
i
h
= h and we directly get eq. (7.4.1): p x
hence C
x
2
2) Now calculate the uncertainty of spatial observable y and linear momentum in
h
h
h
h h
x-direction: [ y ,p x ] = y

y =y
y
=0
i x
i x
i x
i x i
Hence, we can determine both simultaneously and precise. The two operators commute and are
therefore not complementary!

continue to next section (7.5)

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

7.5 Dirac Notation

Dr. P. Blmler: PH 502 Wave Mechanics

page: 28

7.5 Dirac Bracket Notation:


In this optional chapter a very elegant and simple notation is introduced. However, it will be NOT
USED in following chapters to avoid confusion. This part should simply provide you with the
necessary definitions, so that you are enabled to read quantum mechanical literature which often uses
this type of notation.
Diracs bracket notation is an abbreviation for the typical integrals in quantum mechanics and we will
see that may postulates can be expressed in a more handy and short style.
The basic definition is:
n
mA

dV
*m A
n

(7.5.1)

where the symbol n is called a ket and denotes the state expressed by wavefunction n . The
complex conjugate of the wavefunction *n is denoted by the bra n .
When bras and kets are strung together with an operator between them, like as in the bracket
n the integral of eq. (7.5.1) is to be understood. When the operator is simply multiplication
mA
by 1 (or I ) the operator is omitted from the bracket.
Hence the following conditions from the previous postulates become in bracket notation:
normalisation condition:

n n =1

(7.5.2)

orthonormality condition:

n m = nm

(7.5.3)

hermiticity:

n = nA
m*
mA

(7.5.4)

continue to next section (8.1)

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Course Summary / Spine

Dr. P. Blmler: PH 502 Wave Mechanics

page: 29

Confined particles:
Solutions

Excursus 1: Separation of variables

To know:
1. Solutions for particle in 1D box

n =

2
a

nx
sin
,
a

En =

n2 h2
8ma

, n = 1, 2, 3, K

2. Zero-point energy, uncertainty principle


3. Separation of variables (approach to 2D/3D problems)
4. Degeneracy

Exercises and workshops

To know:
General recipe for solving QM-problems
1. Determine potential energy
2. Determine boundary conditions
3. Find suitable co-ordinate system
4. Try to separate Schrdinger equation
5. Solve differential equation
6. Determine constants/quantum numbers via boundary
conditions, normalisation, zero-point energy.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

8 Confined Particles

Dr. P. Blmler: PH 502 Wave Mechanics

8.

page: 30

Confined Particles

8.1 Recapitulation: 1D-Problem particle in a box


Mechanical analogy: Vibrating string, see [1] page 554ff.

U(x)

Fig. 8.1: Co-ordinate system for the 1D problem.

The Hamiltonian for a particle in the infinitely deep well is:


h2 d2
H = + U (x )
2m d x 2

(8.1.1a)

0 for 0 x a
U (x ) =
(8.1.1b)
for x < 0 and x > a
The particle cannot penetrate the infinitely high potentials at x = 0 and x = a. Therefore, the
wavefunction is only non-zero for the region defined by U(x) = 0, resulting in the Hamiltonian we
already know for free translational motion. The Schrdinger equation is then
with:

d2
dx

2mE
h2

( x ) = A cos( kx) + B sin( kx), k 2mEh 2

with the general solution:

(8.1.2)
(8.1.3)

Considering the boundary conditions (0) = (a) = 0 one gets (0) = A = 0 and
(a) = B sin(ka) = 0. The only non-trivial (particle actually exists) solution is ka being an integer
multiple of .
k=

n
, n 1, 2, 3, K
a

(8.1.4)

This implies a quantized energy of the particle:


En =

n2 h 2 2
2ma 2

n2 h2
8ma 2

, n = 1, 2, 3, K

(8.1.5)

To complete the problem, only B remains to be determined. This is done by normalising the
problem. Because the probability to find the particle in the box must be unity, one simply has to solve
the following equation:

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

8 Confined Particles

Dr. P. Blmler: PH 502 Wave Mechanics


a

*n ( x ) n ( x)

dx =

sin

( )d x = [

2 nx
a

B2 x
2

( )]

a
a
sin 2 n x
4n
a
0

page: 31
B2a
=
2

(8.1.6)

with B = 2 a the complete solution of (8.1.1) is:

Normalised solution for a particle confined in a 1D potential well:

En [8ma 2 /h 2 ]

n =

2
a

sin

nx
,
a

En =

n2 h2
8ma 2

, n = 1, 2, 3, K

(8.1.7)

yn*(x ) yn (x )

yn(x)

n=4

n=3

n=2

n=1

Fig. 8.2: 1D-confined particle: Graphical representation of the wavefunctions (left) and the probability
densities (right) versus the energy of the state for n = 1, 2, 3 and 4.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

8 Confined Particles

Dr. P. Blmler: PH 502 Wave Mechanics

page: 32

8.2 2D-Problem: Particle in a Square Well


The particle is now confined in a rectangular well with linear dimensions a1 and a2 as depicted in Fig.
8.3. The mechanical analogy to this problem would be a vibrating rectangular drum or sheet, whose
edges are fixed.

U(x,y)

a1

a2
0
0

Fig. 8.3: Geometry of the rectangular well. Note the graphical necessity to give infinity a value!

The Hamiltonian for this problem is:


h2
H =
2m

2
2

+
x 2 y 2

+ U ( x, y)

0 for 0 x a1 and 0 y a 2
U ( x, y ) =
for all other
Hence the Schrdinger equation is (cf. argument above (8.1.4)):
with:

2 ( x, y )
x 2

2 ( x , y )
y 2

2mE
h2

( x, y)

(8.2.1a)

(8.2.1b)

(8.2.2)

This partial differential equation can only be solved if we can separate the variables ( see
excursus E1). We now substitute (8.2.2) with y(x,y) = X(x)Y(y) and divide by XY as shown in
excursus E1:
X X + Y Y = 2mE h 2

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(8.2.3)

8 Confined Particles

Dr. P. Blmler: PH 502 Wave Mechanics

page: 33

Each term must be a constant summing up to the energy term, which we therefore separate as well
E = EX + EY. So (8.2.3) can be separated into two similar parts:
X ( x) =

2mE X
h

X ( x)

and

2mEY

Y ( y ) =

Y ( y)

with E X + E Y E

(8.2.4)

Each of these equations has the form of (8.1.2) for the 1D-problem and even the boundary
conditions are the same! So we can readily write the solution as:

Normalised solution for a particle confined in a 2D potential well:


n x n y
2
sin 1 sin 2
a1 a 2
a1 a 2

n1 , n 2 = X n 1 ( x) Yn 2 ( y) =

n12

a2
1

h
En 1 , n 2 = EnX + E Yn =
1
2
8m

a)

n 22
a 22

(8.2.5)

b)

n2

n1

1
2

c)

y
a2

d)
x

a1

Fig. 8.4: Wavefunctions for a 2D well with a 2 = 2 a 1: a) y11, b) y12, c) y11, d) matrix with grey
shaded pictures for n 1,n 2 = 1,2,3,4. Dark colours represent negative values, bright colours
positive values.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

8 Confined Particles

Dr. P. Blmler: PH 502 Wave Mechanics

page: 34

A typical feature of 2D/3D systems becomes apparent when the box is geometrically square (cube
for 3D, see Fig. 8.6), that is a1 = a2 = a. Then the energies are
En 1 , n 2 =

h2
8m a

(n12 + n 22 )

(8.2.6)

That means that yab and yba have exactly the same energy, although the wavefunctions are different.
The property of different states having the same energy is called degeneracy, and the corresponding
states to this energy are said to be degenerate.
The wavefunctions of degenerate states typically can be transformed by simple geometrical
transformation (like 90 rotations). Thus, degeneracy can always be traced to an aspect of the
systems symmetry.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

8 Confined Particles

page: 35

8.2.1 Example Quantum Corrals


Don Eigler and co-workers [2] used the tip of a scanning tunnelling microscope (STM) to place
iron atoms on a copper surface (cf. Fig. 8.5a). After finishing the construction of such quantum
corrals they used the same instrument in a different mode to scan the surface, which exhibited
probability distributions of a trapped electron.

a)

b)

c)
Fig. 8.5: STM measurements of quantum corrals: a) Different stages in the assembly of a round
corral, b) the finished round quantum corral with a diameter of 143 , c) a rectangular corral.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

8 Confined Particles

Dr. P. Blmler: PH 502 Wave Mechanics

page: 36

8.3 3D-problem particle in a real box


We have now all tools to solve the 3D Schrdinger equation

h2 2
( x, y, z ) + U ( x , y , z ) ( x, y , z ) = E ( x, y, z )
2m

(8.3.1)

for a particle in a 3D box with the following boundary conditions:


0 for 0 x a1 and
U ( x, y, z ) =
for all other

0 y a2

and

0 z a3

(8.3.2)

Derive the following solution of the problem:


Normalised solution for a particle confined in a 3D potential well:
n1 , n 2 , n 3 ( x, y , z ) =

En1 , n 2 , n 3

n x n y n z
8
sin 1 sin 2 sin 3
a1a 2 a3
a1 a 2 a3

h 2 n12 n22 n32


=
+
+
8m a12 a22 a32

(8.3.3)

with n1 , n2 , n3 = 1, 2, 3, K

Problem 8.1: Find degenerate states in Fig. 8.4d. Can you tell by looking at the displayed
probability density? Find other states for n1,n2 > 4!
Problem 8.2: Assuming the rectangular corral in Fig. 8.5.c was scanned at 4.2K and contains a
single electron, can you estimate the distance of the iron atoms using eq. (8.2.5)?
Discuss how realistic your result is and what assumptions might be wrong!
Thermal energy E = k BT.
Comparing the thermal and QM energy one can work out the distance of the STM
tip, assuming a particle in a 3D box.
Problem 8.3: How degenerate is (5,6,7) for a particle in a cubic box? (It is not 6)

References:
1.
2.

Mary L. Boas: Mathematical Methods in the Physical Sciences 2nd ed., Wiley,
New York 1983.
M. F. Crommie, C. P. Lutz, D. M. Eigler: Confinement of electrons to quantum
corrals on a metal surface, Science 262(5131) (1993) 218-220; M. F. Crommie, C.
P. Lutz, D. M. Eigler and E. J. Heller, Physica D 83 (1995) 98-108.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

8 Confined Particles

page: 37

Fig. 8.6: The first energy levels of a partic le in a cubical box with a 1 = a 2 = a 3 = a together with the
quantum numbers n 1, n 2, n 3 of the corresponding wavefunctions and the degeneracy of each
energy level. The ground state energy is E0 = E111 = 3h 2/(8ma2).

continue to next section (9.1)

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Excursus 1

Dr. P. Blmler: PH 502 Wave Mechanics

E1

page: 38

Excursus: Separation of variables:

The problem satisfies Laplaces equation


2
2
2
2 ( x , y , z ) = 0 with the Laplacian Operator
2 + +

x 2 y 2 z 2

(E1.1)

However, we want to restrict it to two dimensions for the moment and try a solution of independent
variables, that is:
( x, y ) = X ( x ) Y ( y)

(E1.2)

Substituting (E1.2) into (E1.1), we get


Y ( y)

d 2 X (x )
dx 2

+ X (x )

d 2Y ( y )
dy 2

=0

(E1.3)

Note that we can write this now in ordinary rather than partial derivatives, because X depends only
on x and Y on y!
We now divide (E1.3) by XY to separate the terms:
1 d 2 X (x )
2
X (x )
x4
1442d4
3

depends on x only

1 d 2Y ( y )
Y ( y ) dy 2
14
4244
3

=0

(E1.4)

depends on y only

Non-trivial solutions must therefore fulfil the following condition:


1 d 2 X (x )
1 d 2Y ( y )
=
= k2
2
2
X ( x ) dx
Y ( y ) dy
or

X = k 2 X

with k 0

and Y = k 2 Y

(E1.5)
(E1.6)

The constant k 2 is called the separation constant and (E1.6) has the following solutions:
sin( kx)
X ( x) =
cos( kx)

and

exp( ky)
Y ( y) =
exp( ky)

(E1.7)

The solutions (or linear combinations) of the initial problem are therefore:
e ky sin kx
ky
e
sin kx
( x, y ) = X ( x )Y ( y ) = ky
e cos kx
e ky cos kx

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(E1.8)

Course Summary / Spine

Dr. P. Blmler: PH 502 Wave Mechanics

9.

The Harmonic Oscillator

9.1

Classical description
Detail

page: 39

Classical Description

To know:
1. Potential energy: U = k x 2 with k = m2 = 22 m 2 x2
2

2. Classical forbidden regime


9.2

1D Harmonic Oscillator in QM
Detail (solution of differential equation)

Detail (discussion of solution, graphs)

MAPLE exercise

To know:

1. Schrdinger-Equation:

h2 d2
m2 2
( x) +
x ( x) = E ( x)
2m dx 2
2

2. Solutions: Energy and quantum numbers

E = s + 1 h = s + 1 h
2

with s = 0, 1, 2,K

Wavefunction = Gaussian times Hermite-polynomial


2
s ( ) = N s H s ( ) e 2

with = 2 m h x = m h x

3. Zero-point energy s = 0 E0 = h2
4. Potential tunneling (QM Classical)
5. Correspondence principle
Exercises and workshops

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Course Summary / Spine

Dr. P. Blmler: PH 502 Wave Mechanics

9.3

page: 40

The 2D and 3D Harmonic Oscillator

Detail (separation of variables)

Detail (discussion of solution, graphs)

MAPLE exercise

To know:
h2 2
1. Schrdinger-Equation:
(r ) + 22 m 2 r 2 ( r ) = E ( r )
2m
with r = ( x, y, z )
2. Solutions: Energy and quantum numbers*

(sx +

s y + s z + 3 h
2
with s x , s y , s z = 0, 1, 2, 3, K

Es x s y s z =

Wavefunction = Gaussian times Hermite-polynomial


2
s ( ) = N s H s ( ) e 2

with = 2 m h x = m h x

3. Degeneracy
*for identical oscillation frequencies in all 3 dimensions.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

9 Harmonic Oscillator

Dr. P. Blmler: PH 502 Wave Mechanics

9.

page: 41

The Harmonic Oscillator

9.1 Classical description

Fig. 9.1 The one-dimensional harmonic oscillator. Displacement from equilibrium (x = 0) is denoted by
x.

The classical equation of motion of a particle with mass m is given by Hookes law (cf. Fig. 9.1):
m&x& = kx
or

(9.1.1)

&x& + 20 x = 0

with k = m02 as the spring constant and 0 the natural frequency. The energy of the system is then E = K + U =
In the turning points x 0 the particle comes to a rest (cf. Fig. 9.2) and thus the kinetic energy K = 0
and the energy is entirely potential

E = U = k x02
2

(9.1.3)

For x 2 > x02 , the kinetic energy is negative, which is classically forbidden.

Fig. 9.2: The potential U(x) of the harmonic oscillator. The turning points of the oscillation are at
x = x0, where U(x0) = E.

For the classical case. the probability PCL(x) of finding the particle in the interval dx around the point
x at any time is
Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

9 Harmonic Oscillator

Dr. P. Blmler: PH 502 Wave Mechanics


P CL ( x ) d x =

page: 42

dt

= 0 dt
T0
2

(9.1.4)

( )

2
with T0 as the period. Combining eqn. (9.1.2) and (9.1.3) gives dd xt = 20 x02 x 2 and thus

P CL ( x ) =

0 d t
1
=
2 d x
2 x 02 x 2

(9.1.5)

The probability must be normalised (N as a normalisation constant) according to


x0

x0

Thus N = 2 and

CL

N
( x) d x =
2

x0

x 02

1
x2 2

x0

P CL ( x ) =

N
dx =
2

x 0
N
arcsin
= 1

x0 x
2

x 02

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(9.1.6)

(9.1.7)

9 Harmonic Oscillator

Dr. P. Blmler: PH 502 Wave Mechanics

page: 43

9.2 1D Harmonic Oscillator in QM


U ( x ) = 1 kx 2 = 1 m2 x 2 = 2 2 m 2 x 2

Since the potential energy is:

(9.2.1)

with as the angular frequency or = 2, the Schrdinger equation is derived straight-forwardly:


h2 d 2

( x) + U ( x) ( x) = E ( x )
2m dx 2
d

dx

(x ) +

8 m
h

[E 2 m x ] ( x) = 0
2

2 2

2E
m
and 2
x
h
h

(9.2.3)

d2

() = ( )
d 2

(9.2.4)

We substitute and simplify with


thus (9.2.2) becomes:

(9.2.2)

Note: The presented solution uses operators. Brehm and Mullin provide you with an
alternative way using series expansion (check their pages 244-250).
The left side of equation (9.2.4) naively reminds you of equation a2-b2 = (a+b) (a-b), but be
careful because we are dealing with operators here. So we try:
d
d

+
=
+ ( ) = + 2
d
d

d2

= 2 2 1
d

(9.2.5a)

and analogously
d2

d
d


+ = 2 2 + 1
d

d
d

(9.2.5b)

To simplify further we introduce the following operators as abbreviations:


d

a
+
d

and

a +

d

(9.2.6)

So eqn. (9.2.5) now look like this:

d2

() = a a + + 1 ()
d 2

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(9.2.7a)

Dr. P. Blmler: PH 502 Wave Mechanics

9 Harmonic Oscillator

d2

( ) = a + a 1 ( )
d 2

page: 44

(9.2.7b)

From this equation or the corresponding formulation of the Schrdinger equation in (9.2.4), which is
then (where l() denotes the wavefunction to the energy term )
a a + ( ) = ( + 1) ( )

(9.2.8a)

a +a ( ) = ( 1) ()

(9.2.8b)

we see that that for all wavefunctions

a +a a a + = 2

(9.2.9)

Now we take the Schrdinger equation of (9.2.8b) and operate on both sides with a
a a + a () = ( 1) a ( )

(9.2.10)

and combine it with eq. (9.2.9):

(a +a 2)a () = ( 1) a ()
a + a a () = ( 1) a ( ) + 2a ()

(9.2.11)

a + a a () = ( 3) a ( )
This corresponds to the Schrdinger equation (9.2.8) for the energy (eigenvalue) -2, or
a + a ( 2 ) ( ) = ( 3) ( 2 ) ()

(9.2.12)

Therefore, a ( ) must be proportional to ( 2 ) . If we repeat the procedure of (9.2.10 -11), the


functions corresponding to energy levels (-4), (-6), .... can be created. For this reason the
operator a is called annihilation operator, because operated on it generates the wavefunction
that corresponds to the next lower energy level. Analogously a + is called creation operator,
because it increases the energy level. The pair a and a + are also called ladder operators.
However, annihilation is more useful at the moment, because we know that the decrease of energy
cannot be continued indefinitely, since kinetic and mean potential energy are positive it cannot
become negative. Therefore, there must be a minimum energy assigned by min. and a applied to
( min ) consequently must give 0: a ( min ) = 0 .
(9.2.13)
We operate on both sides with a + and use eq. (9.2.8b) to set it in relation to the Schrdinger
equation:
0 = a + a ( min ) = ( min 1) ( min )

(9.2.14)

thus, min.= 1. As demonstrated above, can take integer values differing by 2, so we conclude that
= 1, 3, 5, 7, ... or with 2s + 1, where s = 0, 1, 2, 3,... Now we
re-substitute using the definition in (9.2.3):

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

9 Harmonic Oscillator

Dr. P. Blmler: PH 502 Wave Mechanics


E =

h
= s + 1 h = s + 1 h
2
2
2

with s = 0, 1, 2,K

page: 45

(9.2.15)

What is left is the determination of the wavefunction, which is now rather simple. Therefore, we take
the wavefunction corresponding to the lowest energy level and all we have to do is to operate on it
with a+ to get the next higher wavefunction and continuation of this procedure gives us any s we
like. (Note that now s rather then is used as a label.)

( )

s ( ) = N s a +

0 ()

= N s dd dd dd K dd
()


( s times) 0

(9.2.16)

with Ns as a normalisation constant. So all left to determine is 0, from which we know it must
satisfy (9.2.13) or
d

d 0
a 0 = + 0 = 0 or
= d ln 0 = d
0
d

2
0 ( ) = N 0 e 2

which can readily be solved to be a Gauian:

(9.2.17)
(9.2.18)

1, 2, 3, etc. can then be found using eq. (9.2.16) and are


2
2
d

1 () exp = 2 exp
2
2
d

2
= N1 [2] exp

2 ( ) = N 2 4 2 2 exp
2

2
3 ( ) = N 3 8 3 12 exp
2

(9.2.19)

the polynomials in squared brackets are known as the Hermite polynomials H1, H2, H3,... which
are summarised in Table 9.1 together with the normalisation constants Ns, which are obtained by

the requirement

() ( ) d = 1 .

Normalised solution for the 1D harmonic oscillator:

E = s + 12 h = s + 12 h
2
s ( ) = N s H s () e 2

with s = 0, 1, 2, K

with = 2 m h x = m h x

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(9.2.15)
(9.2.20)

Dr. P. Blmler: PH 502 Wave Mechanics

9 Harmonic Oscillator

page: 46

Table 9.1: Energy levels Es, quadratic normalisation constants N s2 and Hermite polynomials H
for various s
s
Es [J]
H s(x)
N2

h
2

3h
2

1 4m
2
h

5h
2

1 4m
8
h

42 2

7h
2

1 4 m
48
h

83 12

9 h
2

1 4 m
384
h

164 48 2 + 12

11h
2

1
4 m
3840
h

325 1603 + 120

4m
h
2 s s!

2 ds
2
( 1) s e
e
d s

....
s

4 m
h

(s + 12 )h

It can be shown that:

Exercise 9.1:

H s( y ) 2 y H s (y) + 2 s H s ( y) = 0

(9.2.21)

H s +1 ( y ) + 2 s H s 1 ( y ) 2 y H s ( y ) = 0

(9.2.22)

H s ( y ) = 2s H s 1 ( y )

(9.2.23)

Show that the normalisation constants in Tab. 9.1 are correct! (Use:

exp( a
0

2 2

x ) dx =

for a > 0)
2a

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

9 Harmonic Oscillator

E [hn]

Dr. P. Blmler: PH 502 Wave Mechanics

11/2

9/2

7/2

5/2

3/2

1/2

y(x)

page: 47

y*(x ) y(x )

Fig. 9.3: Graphs of the wavefunctions (left) and probability density functions (right) of the 1D
harmonic oscillator for the first 5 energy levels versus dislocation y. The red curves on the
right represent the probability density functions of the classical harmonic oscillator (cf. eq.
(9.1.7)).

On the right side of Fig. 9.3 we observe, that with increasing energy or s the quantum mechanical
probability density PsQM () is becoming more and more similar to the classical probability density
PCL () (red curves in Fig. 9.3). This observation is further illustrated in Fig. 9.4, where the two
functions are compared for E20 = 21h/2.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

9 Harmonic Oscillator

page: 48

P(x)

Dr. P. Blmler: PH 502 Wave Mechanics

Fig. 9.4: Illustration of the correspondence principle: Black curve: Quantum mechanical probability
density function of a high energy state (20())2. Red curve: Classical probability density
function from (9.1.7).

The principle that for large quantum numbers (e.g. energies) the quantum mechanical probability
distribution function PQM resembles the classical probability distribution function PCL is called
correspondence principle (see Fig. 9.4), it describes the limiting regime where quantum mechanics
meets classical mechanics (c.f. de Broglie relation).
correspondence principle:

lim PsQM ( r ) = P CL ( r )

dt
dr

In other words, the system behaves only classically, when there is sufficient energy.
Exercise 9.2: Demonstrate the correspondence principle of a particle in a 1D-box.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(9.2.24)

9 Harmonic Oscillator

Dr. P. Blmler: PH 502 Wave Mechanics

page: 49

9.3 The 2D Harmonic Oscillator

Fig. 9.5: Classical picture of a 2D harmonic oscillator.

The solution of the 2D harmonic oscillator is omitted (see eq. (8.2.3) and the following treatment
of the 3D harmonic oscillator). The results are:

Normalised solution for the 2D harmonic oscillator:

Es x s y = s x + 1 h x + s y + 1 h y = s x + 1 h x + s y + 1 h y
2

with s x , s y = 0, 1, 2,K
2
2
s x s y ( , ) = N s x s y H s x () e 2 H s y () e 2

(9.3.1)

(9.3.2)

with = 2 m x h x = m x h x
and = 2 m y h y = m y h y
N s2 s =
x y

4m
(s +s )
2 x y sx ! s y ! h
1

x +y

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(9.3.3)

Dr. P. Blmler: PH 502 Wave Mechanics

quantum numbers
( s x , sy )

degeneracy

(04) (13) (22) (31) (40)

(03) (12) (21) (30)

(02) (11) (20)

(01) (10)

(00)

E/hn

energy
levels

9 Harmonic Oscillator

page: 50

0
Fig. 9.6: Energy diagram for a symmetrical (Ux = Uy = U) two-dimensional harmonic oscillator.
Compare with Fig. 9.7.

s2

s1

0
1
2
3
Fig. 9.7: Wavefunctions for a symmetrical (Ux = Uy = U) two-dimensional harmonic oscillator. Black
for negative, white for positive values. Compare the pattern for the degenerate states from
Fig. 9.6.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

9 Harmonic Oscillator

page: 51

Fig. 9.8: The first four wavefunctions with identical quantum numbers (00, 11, 22, 33) for a
symmetric potential. (Note: The states with mixed quantum numbers are not shown and not
both potentials are drawn. The red line is an diagonal average).

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

9 Harmonic Oscillator

Dr. P. Blmler: PH 502 Wave Mechanics

page: 52

9.4 The 3D Harmonic Oscillator


U (r ) =

The potential shall be given by

k 2
r = 22 m 2 r 2
2

(9.4.1)

r 2 = x 2 + y2 + z 2

with

(9.4.2)

The Schrdinger equation is then:

h2 2
( r ) + 2 2 m 2 r 2 ( r ) = E ( r )
2m

(9.4.3)

Completely analogously we substitute as in the 1D case (cf. eq. (9.2.3)):

and:

2E
2
=
E X + E Y + E Z X + Y + Z
h h

m
x,
h

m
y,
h

m
z
h

(9.4.4)

(9.4.5)

which yields the Schrdinger equation in the following form:

2
2
2

+
+
(, , ) 2 + 2 + 2 (, , ) = ( , , )
2
2
2

(9.4.6)

As in the 2D and 3D treatment of a confined particle in zero potential (cf. chapter 8.2 and 8.3) we
assume the solution to be a product of functions X,Y,Z, which depend on one spatial co-ordinate
only:
(, , ) = X ( ) Y ( ) Z ( )

(9.4.7)

we substitute into (9.4.6) and divide by XYZ:


1 dX ( )

2 + X
X ( ) d 2

1 dY ( )
+
2 + Y
Y ( ) d 2

1 dZ ( )
2
Z
+

=0

Z ( ) d 2

(9.4.8)

When each term becomes zero the equation is fulfilled, for instance for the first term:
1 dX ( )
2 + X = 0
2
X ( ) d
dX ()
d

2 X () + X X () = 0
dX ( )
d

(9.4.9)

2 X () = X X ( )

This the same expression as eq. (9.2.4), of which we know the results (eq. (9.2.15) and (9.2.16)).
Hence, the solution of the Schrdinger equation for the 3D harmonic oscillator is given by:
Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

9 Harmonic Oscillator

Dr. P. Blmler: PH 502 Wave Mechanics

page: 53

Normalised solution for the 3D harmonic oscillator:


(for a symmetrical potential*)

Es x s y s z = s x + 1 h + s y + 1 h + s z + 1 h = s x + s y + s z + 3 h
2
2
2
2
with s x , s y , s z = 0, 1, 2, 3, K
2
2
2
s x s y s z (, , ) = N s x s y s z H s x ( ) e 2 H s y () e 2 H s z ( ) e 2

with = 2 m h x , = 2 m h y , and = 2 m h z

N s2 s s =
x y z

1
2

(s x + s y + s z )

3
4m 2

sx ! s y !sz !

* in eq. (9.4.1) we have assumed a symmetrical potential, represented by a single k or n.

Fig. 9.9: Energy diagram for the first 4 states of a symmetrical 3D harmonic oscillator.

continue to next section (10.1)

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(9.4.10)

(9.4.11)
(9.4.12)

(9.4.13)

Course Summary / Spine

Dr. P. Blmler: PH 502 Wave Mechanics

page: 54

10. Rotational Motion


10.1/2

Classical Rotation / QM Considerations


Detail

Classical Description

To know:
1. Linear / Rotational Motion:
2. Reduced Mass
3. Boundary Conditions

10.3 Exact Solution for the Rigid Rotator


Detail (Co-ordinate transform)

Detail (discussion of solution, graphs)

To know:
1. Schrdinger-Equation in circular co-ordinates
2. Solutions and quantum numbers

1
e im
2

m ( ) =
Em = m

2 h

2I

and

with m = 0, 1, 2, 3, K
Lm
z = mh

3. Uncertainty and m = 0
4. Shape of wavefunctions, auxiliary functions, Lz

Exercises and workshops

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

Course Summary / Spine

page: 55

10.4 Particle Rotating on a Sphere:


Detail (Co-ordinate transform)

Detail (Solution of differential eq.)

Detail (discussion of solution, graphs)

MAPLE exercise

To know:
1. Schrdinger-Equation in spherical co-ordinates.
2. Integration limits for spherical co-ordinates.
3. Solutions (spherical harmonics)*

2 Yml (, ) = l(l + 1) Yml (, )


Schrdinger-eq.
l
l
im
wavefunctions Ym (, ) = N lm Pm (cos ) e

h2
energy El = l (l + 1)
2I
4.
5.
6.
7.

Quantization conditions l = 0, 1, 2, K and l m l


Orbital Shapes (simple ones)*, dependence on l, m
Probability density, ()-dependence
Degeneracy

* not in detail, but their composition and dependence (the most simple you could remember!)

Exercises and workshops

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

Course Summary / Spine

page: 56

10.5 Angular Momentum


Detail

Detail (discussion of solution, graphs)

To know:
1. Linear definitions, eg. . L x = y p z z p y = hi y z z y

2
2 2
2. L z = i and L = h
applications to spherical harmonics

3. Spatial quantization: L = h l (l + 1)

Exercises and workshops

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

10 Rotational Motion

Dr. P. Blmler: PH 502 Wave Mechanics

page: 57

10. Rotational Motion


10.1 Classical Rotation in a 2D Plane
We remember the classical expressions for linear and rotational motion, as listed in Table 10.1:
Tab.10.1: Classical expressions for linear and rotational motion in a plane
linear motion
rotational motion
co-ordinates

Cartesian: r = x,y

circular: r, f

mass

moment of inertia:
I = mi ri2
i

first derivative wrt t

linear velocity:
v x = x& , v y = y&

angular velocity:
v rv v
=
, = 2
r2

v = r&
momentum
force:
kinetic energy:

linear momentum:
p = mv

L = r p , L = I

linear force:
F = p& = ma

torque:
T = r F = I &

K=

angular momentum:

1 2 p2
mv =
2
2m

K=

1 2 L2
I =
2
2I

The typical case, where we wish to apply the following treatment, is the rotation of a (at least)
diatomic molecule around its centre of mass (Fig. 10.1).

z
L

r0

m
a)

b)

Fig. 10.1: a) Rotation of a diatomic molecule around its centre of mass (CM). b) Without loss of
generality the system can be described by the rotation of a reduced mass m.
Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

10 Rotational Motion

Dr. P. Blmler: PH 502 Wave Mechanics

page: 58

First we want to reduce the problem to a single mass problem. Since the rotation is around the
centre of mass. In good agreement with nature we can assume point-masses:
m1 r1 = m2 r2
m1 r1 = m2 ( r r1 ) or r1 =

hence:

and r0 = r1 + r2
m2
m1
r0 and r2 =
r
m1 + m2
m1 + m2 0
I = m1r12 + m2 r22

The moment of inertia of the molecule is:

(10.1.1)
(10.1.2)
(10.1.3)

Substitution of (10.1.2) in (10.1.3) gives:


I = m1

m12
m m (m + m1 ) 2
m1 m2
r02 + m2
r02 = 1 2 2
r0 =
r02
2
2
2
(
m
+
m
)
(m1 + m2 )
(m1 + m2 )
(m1 + m2 )
1
2

I = r02

m22

with =

(10.1.4)

m1 m2
as the ' reduced mass '
(m1 + m2 )

Analogously we can reduce the masses of a many-body problem, without loss of generality!

10.2 Some Quantum-Mechanical Considerations


For a simple start, the particle shall be confined on a circular path (e.g. a bead on a circular wire).
Thus moving in zero-potential, and the total energy is kinetic only.
E=K =

p2
L2
L2
=
=
2m
2I
2r02

(10.2.1)

We can relate this equation to de Broglies law:


p=

h
hr
or L =

(10.2.2)

The consequence of (10.2.2) is that the angular momentum is inversely proportional to the
wavelength of the particle. Thus, the shorter the wavelength the higher the angular momentum, the
higher the energy. If can take arbitrary values, after one round the particle-wave will have a
different amplitude at the same position. Therefore, the corresponding wavefunction will have
different values at the same position, which is a violation of the first QM-postulate. Hence, only such
are quantum mechanically allowed, which form standing waves on the orbit (see Fig. 10.2). In
other words has to be an integer multiple of the path length:
2r0 = m

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(10.2.3)

Dr. P. Blmler: PH 502 Wave Mechanics

10 Rotational Motion

page: 59

Otherwise the particle-waves could interfere destructively


and finally vanish. We havent specified the quantisation
number m yet. Because the particle can move clock- and
counter-clock wise we allow positive and negative integers.
Like in the case of a particle in a box, we might not allow
m = 0, because this would correspond to zero energy, but
we might reconsider this after a more complete quantum
mechanical treatment.

Fig. 10.2: a) Unacceptable and


b) acceptable wavefunctions

With the cyclic boundary conditions and principle thoughts,


we have already found expressions for the quantized energy
of the particle on a ring, the rigid rotator, or
unrestricted motion on a circular path. All left to do is to

combine eqs.(10.2.1)-(10.2.3):
Em =

L2 h 2 r02 h 2 r02 m2
=
=
=
2 I 2 I 2
82 I r02

h2 m 2
with m = 1, 2, 3, K
2I

(10.2.4)

Note: Careful that you do not confuse m with mass. At the moment we might call it the azimuthal
quantum number !!

10.3 Exact Solution for the Rigid Rotator


We do not have to consider a potential, so the Schrdinger equation for the rigid rotator is
formally the same as for the free particle in two-dimensional Cartesian co-ordinates with the
boundary conditions of being constrained to a circular path:
Schrdinger equation for a free particle in 2D
2
2
2 ( x , y ) = ( x , y ) + ( x, y ) = 2mE ( x , y )

x 2
y 2
h2

(10.3.1)

However, Cartesian co-ordinates are not a suitable co-ordinate frame. 2D-polar or circular coordinates are better suited for the problem, because r = r0 is a constant. Therefore we transform eq.
(10.3.1) into circular co-ordinates (r,) (see Excursus 2: eq. (E2.16)) and replace the mass m by
the reduced mass .

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

10 Rotational Motion

Dr. P. Blmler: PH 502 Wave Mechanics

page: 60

The Schrdinger equation for the rigid rotator in circular co-ordinates is therefore:
1 2
1
2E
(r , ) =
(r , )
r ( r, ) + 2
2
r r r
r
h2

2 (r , ) =

(10.3.2)

This looks more complicated then (10.3.1), but since r is not a variable but a constant r0 (supplied
by the rigidity of the rotators axis), it simplifies to (using eq. (10.2.4) to express E as L2 and m,
eq. (10.1.4) for I):
1 d2
r02
d

() =

() =

2E
h2

2Er02
2
h

()

(10.3.3)

() =

L
h

() = m 2 ()

We acknowledge that the partial differentials became ordinary, and that this equation has been
solved already for the free one-dimensional particle (see part 5c of John Dores part). At the
moment we still consider m as unrestricted in value despite our previous considerations.
The solutions to (10.3.3) are (for generality we allow complex solutions):
() = Ae im + Be im

(10.3.4)

The boundary conditions may be expressed as:


() = ( 2 + )

(10.3.5)

Aeim + Be im = Ae im ( 2 + ) + Be im ( 2 + )
Ae

im

+ Be

im

= Ae

im 2

im

+ Be

im 2

im

(10.3.6)

This can only be fulfilled if m is an integer (because e im 2 = 1 for m = 0, 1, 2, ...), which also
satisfies our previous considerations, and the wavefunction reproduces itself on the next circuit.
Energy and angular momentum are quantized.
Since we allow m to have negative and positive integer values, let us compare the cases m and -m in
eq. (10.3.4). Obviously they correspond to the same state when A and B are swapped. The
corresponding energies (Em m 2 ) are the same, so the states are degenerate. However, the
angular momentum L is pointing in opposite directions (+z and -z). Hence, the two states correspond
to rotation in opposite directions. Because the same considerations hold for the wavefunction with
either A = 0 or B = 0, we want to eliminate this unnecessary ambiguity and rewrite eq. (10.3.4) as:
m () = Aeim

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(10.3.7)

10 Rotational Motion

Dr. P. Blmler: PH 502 Wave Mechanics

page: 61

This also allows the determination of A by normalisation (Note the integration limits for ):
2

* () () d =
0

Hence, A = 1

im

im

d = A

1 d = 2 A

=1

(10.3.8)

and we could readily write down the solution. However, there is still the unsolved

case m = 0! For the discussion of a particle in a box we have used two arguments, as to why the
corresponding state doesnt exist: 1) It wouldnt agree with Heisenbergs uncertainty principle
(since the particle was confined, it has had a finite position range, so the momentum/energy must not
become zero). 2) The corresponding wavefunction (cf. eq. (8.1.7)) would become zero, which
would violate the first postulate.
Lets now consider the same arguments for the case of the rigid rotator. The second can be solved
directly and 0 ( ) = 1 0 . It is constant but not zero, so it exists! The first argument based on
2

the fact, that the particle is confined. Actually it is not, it moves free on its circular path in zero
potential. The idea that it is constrained to the ring originates from the viewpoint of its path in 2DCartesian co-ordinates, but we have solved the problem in circular co-ordinates, and we will find
expressions for its wavefunction for any . (The Schrdinger equation is identical with that of an
unconstrained particle in 1D Cartesian co-ordinates.)
Therefore, we allow m = 0 as an eigenvalue of the Schrdinger equation1.
Normalised solution of the Schrdinger equation for the rigid rotator
m () =

1
e im
2

Em = m 2

h2
2I

with m = 0, 1, 2, 3, K

(10.3.9)

Lmz = mh

(10.3.10)

and

From the solutions we directly observe, that each state is doubly degenerate due to rotation in
opposite directions (Lz pointing up- or downwards) but with the same speed (energy). Except for
m = 0 where Lz is zero and the question of direction doesnt arise.
There are several ways to depict the wavefunctions. A problem arises, because they are complex
e im = cos( m) + i sin( m) . We get somewhat around the problem by plotting the real part only
(since the imaginary is just 90 phase shifted) as done in Fig. 10.3.

The treatment in eq. (10.2.4) might still be treated as correct, because we didnt explicitly consider circular coordinates at that point. However, it has no particular meaning, because we didnt solve the Schrdinger equation
for 2D-Cartesian co-ordinates.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

10 Rotational Motion

page: 62

Fig. 10.3:
Graphical representation of the
wavefunction (real part) for the rigid
rotator
(m = 0,1,2,3).
The
wavefunctions for m = -1, -2, -3
look identical, but the direction of
rotation alters (small arrow) and the
angular momentum (bold arrow)
points down.

However, this gives a wrong picture, because it suggests, that the particle is distributed non-uniformly
over the circle. As shown in eq. (10.3.8) the probability distribution is independent of and hence
uniform. Hence, when the particle is in a state of definite angular momentum its distribution is
completely uniform!
A better representation is shown in Fig. 10.4, where an auxiliary function is constructed by
connecting each point of the perimeter to the centre and to mark on this radius a length proportional
to the amplitude of the wavefunction (see Fig. 10.4a). In other words, this is the plot of the amplitude
in polar co-ordinates. Hence, for 0 we would obtain a circle.
This procedure allows us to project the wavefunction into a plane and the overlay of its real and
imaginary part (see Fig. 10.4c and d). The sum of their magnitudes than somehow resembles the
probability distribution (note overlapping parts in Fig. 10.4c/d, they correspond to the gaps from a
circle).

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

10 Rotational Motion

Dr. P. Blmler: PH 502 Wave Mechanics

b)

a)

+
_

+ _
+
_

_
c)

page: 63

d)

Fig. 10.4: a) Construction of the auxiliary function from the real part of 1. b) Same for 2. c)
The auxiliary function from a). Dark colour denotes the real part, brighter grey the
imaginary part. d) Same as c) for 2 or b) respectively.

10.4 Particle Rotating on a Sphere:


The assumption, that we can describe rotational motion of molecules
by moving on a circular path is quite naive. Because how should the
molecules in a macroscopic sample know which plane is designated.
Typically there is no preferred direction. Therefore, a description by
rotation on the surface of a sphere is a more likely approximation.
We still assume the molecule to be completely rigid along its bond (r
= r0 = const.). From our experience with other two-dimensional
quantum systems, we expect 2 quantum numbers necessary to
Fig. 10.5:
describe the system, because there are two degrees of freedom in
Schematic representation of
rotation (see Fig. 10.5).
a particle moving on the
surface of a rigid sphere. The
arrows indicate the two
degrees of freedom in
rotation.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

10 Rotational Motion

page: 64

Hence, we have to express the Schrdinger equation in 3D polar (spherical) co-ordinates with a
wavefunction that depends on both spherical angles (and the radius), y(r, , ). Fig. 10.6 shows
their interdependence.

Fig. 10.6: Relation between Cartesian and spherical co-ordinates and their range.
In the excursus to this chapter it is shown, how the Laplacian operator is converted into spherical coordinates (see eq. (E2.2.24))
2 =

1 2
1

1
2
r
+
sin +

r 2 sin 2 2
r 2 r r
r 2 sin

Following the same assumptions as in section 10.3, we can directly derive the Schrdinger
equation (r = r0 = const., mass: m ):


1
2
2E
sin ( , ) +
( , ) =
(, ) (10.4.1)


r02 sin
r02 sin 2 2
h2
1

After multiplication with r02 sin 2 and setting r02 = I the equation becomes:


2
2 IE
sin sin (, ) +
( , ) =
sin 2 ( , )
2
2

(10.4.2)

We try a separation of variables by expressing the wavefunction by a product of functions that


depend on one variable only:
(, ) = ( ) ( )

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(10.4.3)

Dr. P. Blmler: PH 502 Wave Mechanics

10 Rotational Motion

page: 65

We substitute in (10.4.2) and divide by :


sin d
d
1 d2
2 IE
( ) =
sin 2
sin ( ) +
2
() d
d
() d 2
h

(10.4.4)

We rearrange the terms that depend on or on either side of the equation and substitute the
energy term by 2IEh 2 :
sin d
d
1 d2
2
( )
sin ( ) + sin =
() d
d
( ) d 2

(10.4.5)

To be solvable each side must correspond to a constant! We take a closer look at the right term and
recognise it as identical to eq. (10.3.3) when choosing the separation constant to be m2. So we
have to solve two equations depending on and :
d2
d
and

( ) = m2 ( )

sin d
d
2
2
sin ( ) + sin = m
() d
d

(10.4.6)

(10.4.7)

The first we have already solved in the previous section and can copy its solution from (10.3.9):
m ( ) =

1
e im with m = 0, 1, 2, K
2

(10.4.8)

Equation (10.4.7) however is new and more complicated. First we multiply with ( ) / sin 2 and
rearrange:

1 d
d
m 2
( ) = 0
sin ( ) +
2
sin d
d
sin

(10.4.9)

Now we want to get rid of all the sin terms. Because we have to take its derivative we choose for
substitution a new function
P( ) = cos

(10.4.10)

and with d = sin d and the Pythagoras relation 2 = 1 sin 2 equation (10.4.9) becomes:

d
dP ( )
m2
1 2
+
P ( ) = 0

2
d
d
1

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(10.4.11)

10 Rotational Motion

Dr. P. Blmler: PH 502 Wave Mechanics

(1 2 )P () 2P () + 1 m 2 P ( ) = 0
2

or

page: 66

(10.4.12)

This type of equation is called associated Legendre differential equation (see e.g. [1] page 504)
with eigenvalues = l ( l + 1) and has solutions Pl m () for l m l and l as an integer called
the associated Legendre functions. They have the following general form of (Legendre) polynomials
of degree l and order m:

( 2)
P m ( ) = ( 1) m 1

m
2

2l l !

l
(
2 1)
d l + m

dl + m

(10.4.13)

and can be normalised by the following integral over the entire range (N as the normalisation
constant):
1

(N P

2
2 ( l + m )!
(

)
d = N 2
=1
l
2l + 1 (l m )!
m

Nlm =

(2l + 1)(l m )!

(10.4.14)

2(l + m )!

Note that = cos has a range from -1 to 1!


(We might derive this solution either by series expansion or by expressing them by the mth derivative
of the Legendre functions Pl(), which can be generated by ladder operators of the type
b + = (1 2 ) d l and b = (1 2 ) d + l , completely analogous to the previous treatment of
d

the harmonic oscillator. However the solution is quite lengthy due to the interdependence of m and l)
We have seen that the dependent part of the Laplacian operator in spherical co-ordinates
produces differential equations of the Legendre type. It is common practice and useful to denote the
part with the angular derivatives of the Laplacian by a new operator, which we will call Legendre
operator (or Legendrian)2:
2 =

1

1 2
sin +
sin

sin 2 2

(10.4.15)

The Legendrian measures the curvature of a function relative to the surface of a sphere. Thus its application to a
wavefunction that represents a sphere is zero, because a sphere in spherical co-ordinates is flat.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

10 Rotational Motion

Dr. P. Blmler: PH 502 Wave Mechanics

page: 67

With these tools we rewrite the Schrdinger equation in (10.4.1):


2 () () = () ()
We have seen that m ( ) =

with

2IE
h2

( , ) = ( ) ( )

and

(10.4.16)

1
eim with m = 0, 1, 2, K
2

Re-substituting = cos the solutions of eq. (10.4.9) are:


( ) = Pml (cos ) with l m l

(10.4.17)

Because m is limited by l, it is often referred to as ml. As eigenvalues for the associated Legendre
equation we found with the condition that l is an integer:
=

2IE
h2

= l (l + 1) with l = 0, 1, 2, K

(10.4.18)

So the complete wavefunction is given by (using the normalisation condition in (10.4.14))


(, ) = ( ) ( ) =
with l = 0, 1, 2, K and

(2l + 1)(l m )!
4(l + m )!

Pml (cos ) e im Yml (, )

(10.4.19)

l m l

This lengthy expression is typically abbreviated by (again) new functions Yml ( , ) called spherical
harmonics. We summarise the mathematical treatment:
Particle rotating on a sphere:
Schrdinger equation:

2 Yml (, ) = l ( l + 1) Yml ( , )

Wavefunction (spherical harmonics):

Yml ( , ) =

Energy:

El = l (l + 1)

Quantization conditions:

l = 0, 1, 2, K and

(10.4.20)

(2l + 1)(l m )!
4 ( l + m )!

Pml (cos ) e im

h2
2I
l m l

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(10.4.21)

(10.4.22)
(10.4.23)

10 Rotational Motion

Dr. P. Blmler: PH 502 Wave Mechanics

page: 68

Table 10.2: Composition of the spherical harmonics Yml ( , ) for the quantum numbers
l 3 and -l m +l
l

El *

Pml () **

Pml (cos )

l
Nm
***

cos

1
2

m 1 2

m sin

1
4

1
2

1
2

m 3 1 2

m 3 cos sin

1
12

30

31 2

3 sin 2

1
24

30

12

1
2

12

12

12

(3 2 1)
(

1
2

(3 cos 2 1)

(5 3 3 )

1
( 2 5 sin 2 ) cos
2

m 32 5 cos2 1 sin

1
12

21

15 sin 2 cos

1
120

210

m 15 sin 3

1
120

35

( )
32
m 15(1 2 )
15 1 2

1
4

1
2

m 3 5 2 1 1 2

Yml ( , )

3
4

cos

m 83 sin e i
5
16

(3 cos 2 1)

m 815 cos sin e i


15
32

sin 2 e i 2

7
( 2 5 sin 2
16

21
64

) cos

(5 cos 2 1)sin ei

105
sin 2 cos e i 2
32

35
sin 3
64

e i 3

* in multiples of h 2 (2 I ) = l(l+1)
** from eq. (10.4.13)
***from eq. (10.4.14)
So far for math, but what do the spherical harmonics look like? We have solved a problem for a
particle on a sphere, and the result is a (sometimes) complex wavefunction on the surface of the
sphere. How can one visualise this? The problem is identical with projecting the globe of the earth
onto a 2D-map! We are used to projections of the earth onto a cylinder, so that the poles become
as large as the equator, although they are points (pseudo Mercator-projectons). In our case the
latitudes correspond to from 0 (south pole) to p (north pole) and the longitudes to from 0
(Greenwich) to 2 (again Greenwich, so UK is at the left and right edge). Although this is a rather
contorted view of reality, you might find it instructive (see Fig. 10.7).
These wavefunctions projected on a sphere are more realistic, but you are restricted to the view on a
single hemisphere (see Fig. 10.8).
Another possibility is to plot auxiliary functions (like in the 2D case, cf. Fig. 10.4), but this time we
will plot the amplitude of the wavefunctions in spherical co-ordinates (see Fig. 10.9). Like in the
2D-case this disables us to distinguish the sign of , but it is a more realistic representation of the
wavefunction. (To understand the relation between Fig. 10.7, 10.8 and 10.9, find the nodes -now
nodal planes- in Fig. 10.7 and mark them on the projections onto a globe, as in Fig. 10.8.
Realise that is spanning the entire diameter, so that nodes separated by -or half the image widths
Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

10 Rotational Motion

Dr. P. Blmler: PH 502 Wave Mechanics

page: 69

in Fig. 10.7- correspond to the same nodal plane in Fig. 10.8 and 10.9. For a better
comparison switch to graph 10.10, where all three representations are compared next to each
other.)
Finally we want to calculate the probability density function:
QM
Plm
( , ) = Yml * (, ) Yml (, )

lm* () *m ( ) lm ( ) m ( )

(10.4.24)

Consulting eq. (10.4.8) We directly see that *m ( ) m ( ) = 1 and from eq. (10.4.13) that
lm* ( ) = lm ( ) , because lm () is always real. Hence, (10.4.24) becomes:

QM
Plm
( ) = lm ( )

(10.4.25)

We have just proven that the probability density function is (like in the 1D case) homogeneous with
QM
respect to . In other words, Plm
() is not homogeneous with respect to .
QM
Fig. 10.10 shows Plm
() in the three different visualisation modes we have used (map

corresponds to Fig. 10.7, globe to 10.8, aux. to 10.9). Note the similarity of the probability
density functions along the diagonal of Fig. 10.10! The auxiliary functions look almost alike (e.g.
QM
QM
P1QM
1 () P2 2 () P3 3 () ), however they are not identical! They have the same symmetry,

but since they correspond to different energy levels, their curvature along must increase with the
quantum numbers, as it can be clearly seen form the map-graphs! Also note that the nodal planes
now correspond to black colour and since we have taken the square of the wavefunction, we might
no longer consider them as planes but nodal cones (planes are then cones for =/2 and 0)!
The last question we are going to consider is the energy term scheme (see Fig. 10.11) for the
particle on a sphere and the degeneracy of its levels. As shown by eq. (10.4.22), the energy of the
system doesnt depend on m, although m is always spanning a range -l m +l resulting in the
same energy. Hence the degeneracy of each energy level is 2 times l (each m is covering the range of
l twice) +1 (for the zero):
Degeneracy of El

is

2l+1

for l = 0, 1, 2, ....

(10.4.26)

Adjacent levels have a energy difference of :


h2
h2
E = El +1 El = [(l + 1)( l + 2) l (l + 1)]
= (l + 1)
2I
I

(10.4.27)

Note: Some authors also introduce a so-called rotational constant B. It is defined as


B=

h
4cI

hence El = l (l + 1) 2hcB = l (l + 1) hcB

with c the speed of light.


Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(10.4.28)

10 Rotational Motion

Dr. P. Blmler: PH 502 Wave Mechanics

Yl m ( , )

m=0
Real

m = +1
Real

Imag

m = -1
Real

Imag

page: 70

m = +2
Real

Imag

m = -2
Real

Imag

m = +3
Real

Imag

m = -3
Real

Imaginar
y

l=0

l=1

l=2

l=3

Fig. 10.7:

Cylindrical projections of the wavefunctions Yl m ( , ) of a particle on a sphere (pseudo Mercator-projections). The vertical axis is from 0 to
, the horizontal axis is from 0 to 2. White colours depict positive and black negative amplitudes. Nodes are in mid-grey! (Note that this is a
rather distorted view of the wavefunctions).

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

10 Rotational Motion

Dr. P. Blmler: PH 502 Wave Mechanics

Yl m ( , )

m=0
Real

l=0

m = +1
Real

Imag

m = -1
Real

Imag

page: 71

m = +2
Real

Imag

m = -2
Real

Imag

m = +3
Real

Imag

m = -3
Real

Imaginar
y

f=0
q=0

l=1

l=2

l=3
Fig. 10.8:

Projections of the wavefunctions from Fig. 10.7 on a sphere. The view is from = 30 and = 270 from 0 to 2 (on the globe of the earth this
would correspond to somewhere over the southern Pacific). White colours depict positive and black negative amplitudes. Nodes are in mid-grey!

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

Yl m ( , )

m=0
Real

10 Rotational Motion
m = 1

Real

page: 72
m = 2

Imaginary

Real

m = 3
Imaginary

Real

Imaginary

l=0

l=1

l=2

l=3

Fig. 10.9:

Projection of the magnitude of the wavefunctions Yl m ( , ) in spherical co-ordinates, generating a 3D-version of the auxiliary functions as
described in Fig. 10.4.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics


m=0

QM
Plm
( )

map

l=0

10 Rotational Motion

globe

page: 73

m = 1
aux.

map

globe

m = 2
aux.

map

globe

m = 3
aux.

map

globe

aux.

f=0
q=0

l=1

l=2

l=3

Fig. 10.10: The probability density functions of the particle on a sphere. As shown in eq. (10.4.25) it is only a function of . The three different
visualisations are directly compared. Map corresponds to the pseudo Mercator-projection of Fig. 10.7. Globe is a plot of the wavefunction on
the sphere as in Fig. 10.8, and aux. is the plot of the amplitude in spherical co-ordinates (cf. Fig. 10.9). Note that the levels gray scale now
from 0 to max, so the nodal planes/cones are black!
Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

10 Rotational Motion

Dr. P. Blmler: PH 502 Wave Mechanics


energy
levels

quantum
number

degeneracy

l=4

12

l=3

l=2

2
0

l=1
l=0

3
1

E
2
[2I/ h ]
20

Fig. 10.11:

l (+1)
l

page: 74

2l+1

Energy diagram for a particle on a sphere.

10.5 Angular Momentum


Before we finish the chapter, we want to investigate the angular momentum of unrestricted motion on
a sphere, or free 3D rotation. The angular momentum is defined as (see Tab. 10.1)
ex
Lx
x px


L = Ly = r p = y p y = x
L
z p
px
z
z

ey
y
py

yp z zp y
ez

z = zp x xpz

pz
xp y yp x

(10.5.1)

Also from Tab. 10.1 we see that the kinetic energy can be expressed in terms of angular momentum.
Since we didnt have to consider potential energy in section 10.4, we can express the total energy
as kinetic energy.
E =K=
with

L2
2I

(10.5.2)

L2 = L L = L2x + L2y + L2z

(10.5.3)

In the context of eigen-equations and the entire context of quantum-mechanics, it is more useful to
express angular momentum Lx, Ly, Lz and L2as operators. The straightforward way is to rewrite
(10.5.1) for each component of L by the operators of position and linear momentum:
L x = y p z z p y = hi y z z y

(10.5.4)

(10.5.5)

L z = x p y y p x = hi x y y x

(10.5.6)

L y = z p x x p z = hi z x x z

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

10 Rotational Motion

page: 75

As before we rewrite these equations in polar notation, which is more suitable to the problem. From
Fig. 10.6 we get the expressions for the Cartesian co-ordinates:
x = r sin cos
y = r sin sin
z = r cos

(10.5.7)

Hence,
L x = h r sin sin r cos
i
z
y
L = h r cos r sin cos
y

(10.5.8)

L z = hi r sin cos y r sin sin x

Now, we express the partial derivatives of the polar co-ordinates by the Cartesian (you might use
the recipe of E2.2 - the other way around) or work it out:

=
r

x y z

+
+
= sin cos + sin sin + cos
r x r y r z
x
y
z
x y z

+
+
= r cos cos + r cos sin r sin
x y z
x
y
z
x y z

+
+
= r sin sin + r sin cos +
x y z
x
y

(10.5.9)

After some reformulation (for which we do not need r , because L must not depend on r), we
get:
L x = h sin cot cos
i


L y = h cos cot sin
i


L = h
z

(10.5.10)

Check the result by substituting the expressions of (10.5.9) into (10.5.10) and try to derive (10.5.8).
This is easier then the other way around ( cot = cos / sin ).
We are left with deriving an equation for L 2 = L (L ) = L x (L x ) + L y (L y ) + L z (L z ) . When we do
by using the expressions in (10.5.10), this becomes quite lengthy, but try it. The approach here is
different by looking at eq. (10.5.2). If the system has no potential energy (like in section 10.4),
the total energy is purely kinetic and expressed by L2. In operator notation we conclude, that the
following eigen-equation must be valid for L 2 :
L 2 (, ) = 2 IE ( , )
Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(10.5.11)

Dr. P. Blmler: PH 502 Wave Mechanics

10 Rotational Motion

page: 76

When we compare this with eq. (10.4.16) -substituting the terms2 IE


2 ( , ) =
(, )
h2

(10.5.12)

we find (cf. (10.4.15), the same result we would get using the explicit expressions in (10.5.10)) :
2

1 2
h 1
L 2 = h 2 2 =
sin


i sin
sin 2 2

(10.5.13)

Since we have solved eq. (10.5.12) in section 10.4 (see (10.4.20)), all we have to do is to
substitute:
L 2 Yml (, ) = h 2 2 Yml (, ) = h 2 l (l + 1) Yml ( , )

with l = 0, 1, 2, K (10.5.14)

So the eigenvalues of L 2 with respect to the spherical-harmonics are simply h 2 l (l + 1) .


A further look to eq. (10.5.10) reveals another term, we came across in solving the Schrdingerequation for the rotators. It is L z = h and when applied to the -part of the sphericali

harmonics (see eq. (10.4.8)) we directly get:


h
h
() = Nlm ( )

i
i
h
= Nlm ( )
m e im = hm Yml ( , )
2

L z Yml (, ) =

N lm ( )

e im
2

with

(10.5.15)
l m l

We have just seen that the magnitude of the total angular momentum L is quantized having
eigenvalues h l (l + 1) and that additionally L z is quantized by eigenvalues hm .
As long as we do not assign the problem to specific Cartesian direction - for instance by applying an
additional (magnetic) field- there is nothing specific about the z co-ordinate. The simple dependence
on occurs because we had chosen it to be perpendicular to z. We will later see, that this choice
makes the x and y component of the angular momentum non-measurable, as their contribution to the
probability density function vanishes (see eq. (10.4.25)). However, if we would have chosen x or
y to lie perpendicular to the angular momentum relative to them would turn out to be quantized,
and the remaining other would be non-determinable.
These results are really important, because we can now conclude, that any system with a spherically
symmetric potential U(r) -hence a central force problem, which can be separated from the angular
terms- has solutions for its angular terms as above.
Graphically we can summarise these considerations. As an example we consider l = 2 and hence
L = h l (l + 1) = h 6 . If a preferred direction exists (as in the presence of an external field) or
we simply chose it to be z, L z = hm with m = -2, -1, 0, 1, 2. Thus 5 different directions of the
angular momentum vector exist relative to the z-axis, as visualised in Fig. 10.12. Note that a
Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

10 Rotational Motion

Dr. P. Blmler: PH 502 Wave Mechanics

page: 77

completely parallel or anti-parallel alignment to the z-axis is not possible, since its maximum
projection on the preferred axis is lh , while having a length of h l (l + 1) > hl . (Only for very large
l ( l + 1) l , so that the object is able to rotate solely

values of l -as for macroscopic objects- is


around a preferred axis).

z
m = +2

2h

m = +1

h
h 6

0
h

m=0

m = -1

2h

m = -2

Fig. 10.12: Vector model of angular momentum:


Possible alignments of L with respect to the z-axis.
As discussed above (and proven later) if we select z as a preferred axis, x- and y-component of the
angular momentum are uncertain or have equal probability to lie on cones (see Fig. 10.13) with halfapex-angles determined by = arccos m / l ( l + 1) . (Do not think of the angular momentum
vector sweeping out one of the cones, but simply lying at some position on the cones).

Fig. 10.13: Cones of location for the angular


momentum vector

continue to next section (11.1)


Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Excursus 2

Dr. P. Blmler: PH 502 Wave Mechanics

Excursus 2:

page: 78

Co-ordinate Transforms of the Laplacian


Operator:

E2.1 Method 1: Explicit solution for 2D circular co-ordinates.


2
2
2

=
+
x 2
y 2

The Laplacian is

(E2.1)

We want to express it in circular co-ordinates (r,) (see Fig. 10.1b)


x = r cos

with r 2 = x 2 + y 2

(E2.2)

r = x2 + y2

or

y
sin
=
= tan
x
cos

and

y = r sin

and

or

= arctan

(E2.3)
y
x

(E2.4)

The total differential of an arbitrary function f is:


f
f
dr +
d
r

(E2.5)

f
f r
f
=
+
x
r x
x

(E2.6)

df =

hence

and

f
f r
f
=
+
y
r y
y

(E2.7)

so we have to find expressions for (using eq. (E2.2) to (E2.4))

12
1 2
r
2
1
=
x + y2
= 2x x 2 + y 2
=
x
x
2

x
2

x +y

x
= cos
r

y
y
1
y
y
sin
=
arctan
=
=
=
=
2
x
x
x
r
x2 1 + y
x2 + y 2
r2
2

(E2.8)

(E2.9)

r
=
y

y
= = sin
r
x2 + y2

1
=
y
x

1
1+

x2

x
x2 + y2

(E2.10)
cos
r

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(E2.11)

Dr. P. Blmler: PH 502 Wave Mechanics

Excursus 2

page: 79

Therefore, we can express the first partial derivatives of f by substituting (E2.8/9) into (E2.6) and
(E2.10/11) into (E2.7):
f
f
f sin
f
sin f
fx
=
cos
= cos

x
r
r
r
r

(E2.12)

f
f
cos f
f y
= sin
+
y
r
r

(E2.13)

To find the second partial derivative we substitute f in (E2.6) by f x and in (E2.7) by f y :


2 f
x 2
2 f
y 2

f x
f x r
f x
=
+
r
r x
x

(E2.14)

f y
f y r
f y
=
+
y
r y
y

(E2.15)

We have to be careful, because f x and f y depend on r and ! Substituting (E2.12) and (E2.8/9)
into (E2.14):
2 f
x 2

f sin f
cos
r
r

f sin f

cos
cos

r
r

2 f
sin f
sin 2 f
= cos
+

r r
r 2
r 2

cos

f
2 f sin cos f
sin 2 f
sin + cos

r
r
r

r 2

= cos 2

2 f
r 2

sin

sin

sin 2 f sin cos 2 f sin cos 2 f 2 sin cos f sin 2 2 f

+
+
+
r r
r
r
r
r

r2
r2
2

and analogously by substituting (E2.13) and (E2.10/11) into (E2.15):


2 f
y 2


f
cos f

f
cos f cos
sin
+
sin +
sin
+

r
r
r

r
r r

= sin 2

2 f
r 2

cos 2 f sin cos 2 f sin cos 2 f 2 sin cos f cos 2 2 f


+
+

+
+
r r
r
r
r
r

r2
r2
2

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

Excursus 2

page: 80

The Laplacian is then given by the sum of the last two equations:

2
2
2
2
2
2
2
2
2 f = f + f = sin 2 + cos 2 f + sin + cos f + sin + cos + f

r
r
x 2
y 2
r 2
r2
2

and using Pythagoras: sin 2 + cos 2 = 1 and getting rid of f


2
2
2 = +

x 2 y 2

r 2

1
1 2
+
r r r 2 2

1 2
1
r + 2 2
r r r
r

(E2.16)

where we used a nice abbreviation for the two terms that depend on r.

E2.2 Method 2: Curvilinear Co-ordinate Transforms


It can be shown (see [1] page 431- 433) that for orthogonal co-ordinates

(
(

)
)


h h
x
r
1

y = h h h hr h
r (h h )
z
r

( )

( )

( )

( )

(E2.17)

2
2
2
y
hr2 = x + + z
r
r
r

with

2
2
2
y
h2 = x + + z

y
h2 = x + + z

(E2.18)
2

and
2 = 2 + 2 + 2 =

2
2
2
x

1
hr h h

h h
hr h
h h

+ r
+
r hr r h h

(E2.19)

The differentials can be derived by the Jacobian :


dx dy dz = J dr d d

with the Jacobian (determinant) J =

( x, y, z )
=
( r , , )

x
r
y
r
z
r

(E2.20)
x

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

= hr h h

(E2.21)

Excursus 2

Dr. P. Blmler: PH 502 Wave Mechanics

page: 81

Application to circular co-ordinates:


For the 2D example in E2.1, we derive from eq. (E2.2):

( )

2
2
y
hr2 = x + = (cos )2 + (sin )2 = 1
r
r
2

y
h2 = x + = ( r sin )2 + (r cos )2 = r 2

2
1 r r

2 = 1 r + 1 1 = 1 r + 1
and hence: = and
r
r r r r r r r r r 2 2

Application to spherical co-ordinates:


The relation between Cartesian and polar co-ordinates is (see Fig. 10.6):
x = r sin cos
y = r sin sin
z = r cos

(E2.22)

The scaling factors for curvilinear transforms are (see (E2.18))

( )2 + yr 2 + (zr )2 = (sin cos )2 + (sin sin )2 + cos2


= sin 2 (cos 2 + sin 2 ) + cos 2 = 1

hr2 = xr

( )

( )

2
2
2
y
h2 = x + + z = (r cos cos ) 2 + (r cos sin )2 + ( r sin )2 = r 2

y
h2 = x + + z = ( r sin sin )2 + (r sin cos )2 = r 2 sin 2



and hence (see (E2.17))

r 2 sin
sin r 2
r

r
1
1

= 2
(r sin ) = 2
r sin

r sin
r sin
(r )
r

and (see (E2.19)):

(E2.23)

2
1

1
2 =

r
sin

+
(
sin

)
+

sin
r 2 sin r
=

1 2
1


1

r
+ 2
sin + 2
2 r
2
r
r sin 2
r
r sin

(E2.24)

The differential expression is (see (E2.21)):


dx dy dz = r 2 sin dr d d
Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(E2.24)

Dr. P. Blmler: PH 502 Wave Mechanics

Excursion 2

page: 82

Tab. E2.1: You might find the following table generally useful:

co-ordinates:

Cartesian
x,y,z

hr
h
h

dV

2
x2

2D circular
x = r cos

3D-cylindrical
x = r cos

3D-spherical
x = r sin cos

y = r sin

y = r sin

y = r sin sin

z = z
1
h z = 1
r

z = r cos
1
r
rsinq

1
r


x

y

z

1 r r
r

r
r
1
r

z

sin r 2

r
1
r sin

r 2 sin

1 1 2
r +
r r r r 2 2

1 1 2
2
+
r +
r r r r 2 2 z 2

1 2
1


1
2
r
+
sin +

r 2 sin 2 2
r 2 r r
r 2 sin

r dr d dz

r 2 sin dr d d

2
2
+ 2+ 2

dx dy dz

r dr d

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

Course Summary / Spine

page: 83

11. The Hydrogen Atom


Detail (Solution of differential eq.)

Detail (discussion of solution, graphs)

MAPLE exercise

To know:
1. Coulomb potential, effective potential, angular momentum barrier, energy
spectrum, Schrdinger-equation for central force problem.
2. Bohr-model, -radius
3. Radial waveform (Laguerre polynomials)*
4. Solutions*

(r , , ) = N nlmR (r ) () ()
En =

Z 2 e 4
802h 2 n 2

with n = 1, 2, 3,... and


5.
6.
7.
8.

l n 1 and

l m l

Quantum numbers n,l,m and degeneracy.


Radial distribution function (construction, spherical Jacobian)
Orbitals*
Basic idea of hybrid orbitals

* not in detail, but their dependence (the most simple you could remember!)

Exercises and workshops

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

11 Hydrogenic Atom

Dr. P. Blmler: PH 502 Wave Mechanics

page: 84

11. The Hydrogenic Atom


11.1 Motion in a Coulomb field
The motion of a single electron in the Coulomb-field of charged nucleus is to be solved. We
remember that the force varies as F(r) 1/r2 and hence the potential energy (F = -dU/dr) as
U(r) 1/r. To be exact the Coulomb-potential between a nucleus of charge Ze and an electron of
charge -e is:
U (r ) =

Ze 2
40 r

(11.1.1)

with 0 = 8.854188 10 12 C 2 / (Jm ) the permittivity in the vacuum. As introduced in part 10.1
we can simplify the two-particle problem to a quasi one-particle problem using its reduced mass
m m
= e nucleus , where mnucleus is the mass of the nucleus.
me + mnucleus
me mp

For instance for 1H: Z = 1 and =

me + mp

Hence the Schrdinger-equation for this problem is:

h2 2
Ze 2

= E
2
40 r

(11.1.2)

Because the problem is similar to the rotation of a particle on the surface on the sphere, but now with
a radial potential, an expression of this equation in spherical co-ordinates, (r , , ) , appears to be
appropriate (using eqn. (E2.24)) :
1 2
1


1
2
2 r
+ 2
sin + 2 2
( r , , ) +

r sin
r sin 2
r r r
2
Ze 2
E +
( r , , )
4 0 r
h 2

(11.1.3)
=

As before, separation of variables is tried by :


(r , , ) = R( r ) ( ) ( )

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(11.1.4)

11 Hydrogenic Atom

Dr. P. Blmler: PH 502 Wave Mechanics

page: 85

Insertion into eq. (11.1.3) gives


d 2 dR
R d
d
R d 2
r
+
sin
+
+

d
r 2 dr dr
r 2 sin d
r 2 sin 2 d 2

(11.1.5)

2
Ze 2
E
+

R
40 r
h 2

r2
following the typical separation-procedure eq. (11.1.5) is multiplied by
:
R
1 d 2 dR
1
d
d
1
d 2
2
Ze 2 2
r
+
sin

+
+
E
+

r =
R dr dr
sin d
d
40 r
sin 2 d 2
h 2

0 (11.1.6)

and rearranged:
1
1 d 2 dR
2
Ze 2 2
d
d
1
d2
r
+
E
+
r
=

sin

(11.1.7)
2
2
R dr dr
40 r

sin

h 2

sin

Hence, the problem is separated into a radial part (left) and an angular part (right). The right side
(square brackets) we identify as the differential equation for the particle on a sphere (see eq.
(10.4.15) to (10.4.20)) and can readily insert the solutions (spherical harmonics):
1 d 2 dR
2
Ze 2 2 (10 .4 .15 )
1 2
=


r
+ 2 E +
r
R dr dr
40 r

h
(10 . 4. 19 )

Yml

2 Yml

(10 .4 .20 )

1 d 2 dR
2
Ze 2 2
r
+ 2 E +
r l ( l + 1) = 0
R dr dr
40 r
h
1 2 d2 R
dR
2
Ze 2 2
+ 2r +
r
E +
r l ( l + 1) = 0
R dr 2
dr
40 r
h 2
d 2 R( r )
dr 2

(11.1.8)
l (l + 1)

R
r2

(11.1.9)

2
2 dR( r )
Ze 2
l (l + 1)

+
+ 2 E+

R( r ) = 0
r dr
4 0 r
h
r 2

This is not surprising, because the problem can be understood as a particle moving on a collection of
spherical shells. However, what is left to find is its distribution over these shells, the radial part.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

11 Hydrogenic Atom

Dr. P. Blmler: PH 502 Wave Mechanics

page: 86

Now we want to go -in a way- some steps backwards, to understand our findings. Therefore, we
express equation (11.1.9) in the standard form of the Schrdinger-equation:

h 2 l ( l + 1)

h2 1 d 2 d
r
R
(
r
)
+
+
U
(
r
)

R( r )
2
2 r 2 dr dr
2r

= ER( r )

h 2
R( r ) + U eff ( r ) R (r ) = ER( r )
2

where an effective potential U eff ( r ) =

(11.1.10)

h 2 l (l + 1)

U e ff [10-18J]

U [10 -17J]

+ U ( r ) is introduced with U(r) as the


2r 2
Coulomb-potential from eq. (11.1.1). The first term appears due to the rotation of the electron,
which counter-acts the Coulomb-potential as a repulsive core for l > 0. It becomes infinitely high as
r 0, hence prevents the system to collapse, and is therefore called angular momentum
barrier. The nature of the two potentials and their sum is depicted in Fig. 11.1.
1.0
0.8

a)

0.6

b)

l=3

1
0.4
0.2

a.m.b.
r []

-0.2

10

U eff( r)
U( r)

r []
4

10

-1

-0.8
-1.0

l=1

-0.4
-0.6

l=2

l=0
-2

Fig. 11.1: Potentials for 1H: a) green: Coulomb-potential, U(r); blue: angular momentum barrier
(a.m.b.) and red curve: sum of a) and b), the effective potential Ueff(r). Calculations for
l = 1. b) Ueff(r) for l = 0, 1, 2 and 3.
In Fig. 11.1 we recognise the attracting nature of the Coulomb-potential (negative) and the repelling
character of the angular momentum barrier. If the energy becomes higher then the Coulomb-term
and the angular momentum term dominates, the particle is retracted from the nucleus and must
behave like a free (unbound) particle, hence without quantized energy states (energies form a
continuum, see Fig. 11.2). This process is called ionisation and defined by Ueff. Due to the wider
potential range for higher energies, we can also conclude that the energy levels for the bound states
must converge to that limit. Due to the finite value of the potential at a discrete bound energy state we
can finally expect the particle to tunnel beyond the classical limits (that is where the energy level
meets the potential).

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

11 Hydrogenic Atom

U eff [10-18J]

Dr. P. Blmler: PH 502 Wave Mechanics

page: 87

continuum of unbound,
free-particle states

r []

discrete,
bound
states

10

-1

Fig. 11.2: Sketch of the quantum mechanical energy spectrum as


to be expected from general considerations.

11.2 Solution of the Radial Differential Equation


The aim of this part is to solve eq. (11.1.9). This will be done in elementary steps, which requires
many substitutions to guide through the essential maths. We start with the last line of (11.1.9) and
rewrite it in a shorter notation:
d 2 R( r )
dr 2

2
2 dR( r )
Ze 2
l (l + 1)
+ 2 E +

R( r ) = 0
r dr
40 r
h
r 2
2
2
Ze 2
l (l + 1)
R + R + 2 E +

R = 0
r
4 0 r
h
r 2

(11.2.1)

the terms in the squared brackets can be simplified by suitable substitution:

E
h2

and

Ze 2
40 h 2

(11.2.2)

leading to:
R +

2
2
l (l + 1)
R + 2R +
R
R =0
r
r
r2

(11.2.3)

We are now testing the equation for r , meaning the electron leaves the nucleus to become a free
particle. In eq, (11.2.3) that causes the terms that depend on 1/r to vanish:
+ 2R = 0
R

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(11.2.4)

11 Hydrogenic Atom

Dr. P. Blmler: PH 502 Wave Mechanics

page: 88

The solution is the same as for the free (non-restricted/non-confined) particle:

R ( r ) = A exp i 2 r + B exp + i 2 r

(11.2.5)

where E represents the energy. This general (complex) solution can only be normalised (to
the real value 1), when we assume E or < 0, so that i 2 = i 1 2 = 2 with a
positive radicand. (Try these normalisations as examples or alternatively substitute with a negative,
see eq. (11.2.8)) This can also be suggested when looking at Fig. 11.2 (but now we have
mathematical proof)! The normalisation condition then reads as:

(A exp (

2 r + B exp 2 r

))2 dr = 1

(11.2.6)

The first term becomes infinite for r and is hence unacceptable, because we know that the
wavefunction must vanish, while the second (becoming zero) is acceptable for these conditions.
Therefore, A = 0 and

R ( r ) = B exp 2 r

(11.2.7)

or with changing our substitution from (11.2.2) to


2 =

2E
h

(11.2.8)

Now, we try a first approach for a general solution by incorporating the knowledge for r and
ignore B, knowing that the final normalisation of the wavefunction must recover it.
R( r ) = R f ( r ) = e r f ( r )

(11.2.9)

with f(r) as a new unknown function! Hence,


R (r ) = e r f ( r )
R( r ) = e r f ( r ) e r f ( r ) = e r ( f ( r ) f ( r ) )
R( r ) = e r ( f (r ) f (r ) ) e r ( f (r ) f ( r ) )

= e r f ( r ) 2f ( r ) + 2 f (r )

(11.2.10)

These results are used in the radial part of the Schrdinger-equation (eq. (11.2.3)):

e r f ( r ) 2 f (r ) + 2 f ( r ) +

2 r

2 l (l + 1) r
e ( f ( r ) f ( r ) ) + 2 +

f (r ) = 0
e
r
r

r2

Multiplication with exp(r) and rearranging yields

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

11 Hydrogenic Atom

2
2

2 l ( l + 1)
f ( r )
f (r ) + 2 +
2 f (r) = 0
r
r
r

2( + ) l (l + 1)
1
f ( r ) + 2 f ( r ) + 2 + 2 +

f (r) = 0
r
r

r2

page: 89

f ( r ) 2 f (r ) + 2 f (r ) +

(11.2.11)

and with 2 = 2 (see definition in (11.2.8))


2( + ) l ( l + 1)
1

f ( r ) + 2 f (r ) +

f (r) = 0
r
r

r2

(11.2.12)

So far things do not really look simplified. We try another approach for the evaluation of the
unknown part of this equation f(r) by expressing is as a power series (with coefficients b) and trying
to find an upper limit (quantum number) to meet the boundary conditions, which means that the series
is transformed into a polynomial of a degree given by this upper limit:
f (r )

bq r q

and hence R( r ) = e r

bq r q

(11.2.13)

q =0

q= 0

for substitution into eq. (11.2.12), we calculate the first and second derivative:
f (r ) =

bq r q
q

f ( r ) =

q bqr q 1

(11.2.14)

f ( r ) =

q(q 1) bq r q 2
q

Used in eq. (11.2.12) this gives:

[(q (q 1) bq r q 2 ) + (2qbq r q 2 2 qbq r q 1 ) + (2 2)bq r q 1 l(l + 1)bq r q 2 ]


q

[q (q 1) + 2q l (l + 1)]bq r

q 2

[(2 2 ) 2q ]bq r

q 1

=0

=0
(11.2.15)

In order to become zero the coefficients for each power of r have to cancel. Hence, we match the
indices for r q and set them to be zero:

[( q + 2)( q + 1)
bq + 2
bq + 1

+ 2(q + 2) l ( l + 1) ]bq + 2 2 [( q + 1) ( )] bq +1 = 0

2 [( q + 1) ( ) ]
2( q + 2 )
=
[(q + 2)( q + 1) + 2( q + 2) l (l + 1) ] ( q + 2)( q + 3) l (l + 1)

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(11.2.16)

Dr. P. Blmler: PH 502 Wave Mechanics

11 Hydrogenic Atom

page: 90

We further simplify this by reducing the index, q, by 1, which results in the following recursionformula:
bq +1 = bq

2((q 1) + 2 )
( q + 1)
= 2bq
( q + 2)( q + 1) l (l + 1)
(q + 2)( q + 1) l (l + 1)

(11.2.17)

The next step is to find a limiting case for q and hence dominating and l2, hence assuming q >
l:
2bq

(q + 1)
lim bq +1 2bq
=
(q + 2)( q + 1)
q

q+2

(11.2.18)

The purpose of a recursion-formula is to address the next higher coefficient by a lower. So we want
to try to express bq by b0 (the lowest coefficient relating to a constant). We start by calculating b1, b2
etc... to end up with a general expression for large q:
2
1
b1 b0 = (2 )b0
3
3
1
1
b2 (2)b1 =
(2)2 b0
4
3 4
1
1
b3 (2)b2 =
(2)3 b0
5
3 4 5
...
2
bq = (2)q b0
q!

(11.2.19)

This reminds us of the Taylor-expansion series for the exp-function, we recall:

ar

q =0

hence R( r )

( 4 .2 .13 )

(ar )q
q!

e r

bq r q

( 4 .2 .19 )

q =0

e r 2b0

q =0

(2r )q
q!

(11.2.20)

R( r ) = 2b0 e r e 2 r = 2b0 e r
Without loss of generality, we can deposit the 2 in the normalisation constant and get
R (r ) = b0 e r

(11.2.21)

However, this cannot be the final result, because we have to remember the boundary condition
lim R( r ) = 0 and from the definition in (11.2.8) we reassure that > 0 ! Therefore an upper
r

limit p has to be introduced to stop the series and convert it into a polynomial of degree p, e.g.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

11 Hydrogenic Atom

Dr. P. Blmler: PH 502 Wave Mechanics

page: 91

0 q p., which means that the series stops after p or b p +1 = 0 . Hence b p + 2 = 0 and all
following coefficients vanish due to eq. (11.2.17) and for b p 0 this gives:
b p +1 = 2b p

( p + 1)
= 0
( p + 2)( p + 1) l (l + 1)

(11.2.22)

The equation is fulfilled for a) the nominator = 0 and b) the denominator 0, or:
( p + 1) =

p +1 n

a)

(11.2.23)

where a new (simplified) quantum number, n = p + 1, is introduced.


(1
p4
+422
)( 4
p +4
1) l1
(l2
+3
1) 0
3

b)

>0

>0

( p + 2)( p + 1) >
n( n + 1) >
n >
n

l (l + 1)
l (l + 1)
l
l +1

(11.2.24)

As at the start of the derivation of this formula we (c.f. eq. (11.2.18)) we have to assume p+1 < l.
As expected, we found another quantum number n, related to l, which is related to m (cf. eq.
(10.4.23)). From condition a) we can directly calculate the energies associated with the radial part of
the wavefunction, by re-substituting (11.2.2) and (11.2.8):
2E
Ze 2
=
h
40 h 2 n
En =

Z 2 e 4
32 2 20 h 2 n 2

(11.2.25)

Z 2 e 4
8 20 h 2 n 2

for n l + 1 = 1, 2, 3,K

This is exactly the result Bohr found in his treatment of the hydrogen atom and we also identify our
parameter
=

1
rB

where rB is the Bohr-radius.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(11.2.26)

11 Hydrogenic Atom

Dr. P. Blmler: PH 502 Wave Mechanics

page: 92

The only thing left to do, is the exact determination of the wavefunction R(r). We summarise our
equipment for this task:
r n 1 q

Rn , l ( r ) = exp
bq r
nrB q = 0

1
1
, =
,
rB
nrB

for n l + 1 = 1, 2, 3,K
(q + 1)
coefficien ts : bq +1 = 2bq
( q + 2)( q + 1) l (l + 1)
bq +1 (q + 2)( q + 1) l (l + 1)
bq =
2
( q + 1)
and start with the lowest allowed state: n = 1, hence l = 0
r
R1,0 ( r ) = b0 exp
rB
R1,0 () = b0 e

(11.2.27)

r
with =
rB

where we introduced a dimensionless co-ordinate as the radius in multiples of the Bohr-radius.


Normalisation gives

R12, 0 ( ) d

=0

= b02

=0

b02
d =
=1
2

(11.2.28)

hence the normalised wavefunction for the ground state is:


R1,0 () =

2 e

(11.2.29)

Now the steps are repeated for n = 2, hence we have to distinguish l = 0, 1 or R2, 0 ( r ) and
R2,1 ( r ) . Generally this is determined by:
r
R2 ,l ( r ) = exp
nrB

1
r

(b0 + b1r ) (11.2.30)


bq r q = exp
q =0
nrB

where b1 can be expressed in terms of b0 by the recursion formula in (11.2.17):


a) case: n = 2, l = 0
b1 = 2b0

r
hence, R2, 0 ( ) = exp
2rB

b
b0 0
2rB

1
2 rB

b
= 0
2 1 0
2 rB
rB


= N 2 ,0 1 e 2
2

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(11.2.31)

11 Hydrogenic Atom

Dr. P. Blmler: PH 502 Wave Mechanics

page: 93

b) case: n = 2, l = 1
If we would proceed as in the previous case eq. (11.2.17) yields zero in the denominator.
Therefore, we use the formula backwards
bq +1 (q + 2)( q + 1) l (l + 1)
bq =
2
( q + 1)
therefore we determine b0. and use b1 for normalisation:
b
2 2
b0 = 1
= 0
1
2
1
2 rB

hence R2,1 () =

b1 rB e

= N2 ,1 e

rB

(11.2.32)

Note, that the final results in a) and b) contain the associated normalisation constants N 2 , 0 and
N 2 ,1 .
Continuation of this procedure gives the following table of the radial-parts of the wavefunction for the
hydrogen-like atom, which are depicted in Fig. 11.3. (Note: That Fig. 11.3 shows the radial part
of the wavefunction and not a 3D-projection of the real distribution of the electron, since the angular
terms have been neglected so far)
Table 11.1: The radial-parts Rn, l () of the wavefunction for the hydrogen-like atom as the function
of the radius in multiples of the Bohr-radius.
n

Rn, l ()
R1,0 =

2 e

1
R2, 0 =
(2 ) e 2
2

1
R2,1 =
e 2
2

2
R3, 0 =
27 18 + 22 e 3
27

R3,1 =

2
9 3

(6 ) e
2

2 2 3
R3, 2 =
e
27

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

11 Hydrogenic Atom

page: 94

Fig. 11.3: Left: The radial part of the wavefunction Rnl() for n=1,2 and 3. Right: The
corresponding probability density function Pnl().
Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

11 Hydrogenic Atom

Dr. P. Blmler: PH 502 Wave Mechanics

page: 95

The radial part of the eigen-functions of the hydrogenic atom are another ortho-normal class of
polynomials (like the Hermite and Legendre polynomials), called associated Laguerre polynomials:
They are solutions of the so-called associated Laguerre equation, in which eq. (11.2.1) can be
transformed.
xf ( x) + (k + 1 x ) f ( x) + j f ( x ) = 0

associated Laguerre equation :

solutions : f ( x) = L kj ( x)
associated Laguerre polynomial s :

L jk ( x ) = (1) k

(11.2.33)
j+ k

j + k xi
(1) i

dx k i = 0
i i!
dk

In this notation the radial wavefunction becomes:


Rnl () = N nl e

/ n

2l + 1
Ln l 1 ()
n

with n = 1, 2, 3,... and l n 1


(11.2.34)

2 ( n l 1)!

n 2n(n + l )!

N nl =

11.3 The Complete Solution: Atomic Orbitals


The complete solution for an electron in a Coulomb-potential (hydrogenic atom) was calculated by
separation of variables, where the radial part was solved in the previous part, after we have
recognised that the angular parts were already solved in chapter 10 with the spherical harmonics,

Yml ( , ) , as eigen-function and hl (l + 1) / 2 I as eigen-values (cf. eq. (10.4.20) to (10.4.23)).


Hence, the eigen-functions of the hydrogenic atom are given according to eq. (11.1.4):
Hydrogenic Atom:
(r , , ) = N nlm R( r ) ( ) ( )
l

with Rnl ( ) = e / n Ln2ll+11 ( )


n

and =

r
rB

( ) ( ) = Yml (, ) = Pml (cos ) e im


N nlm =
En =

Ze 2
40 h 2

3
2 ( n l 1)! (2l + 1)( l m )!

n 2n(n + l )! 4 ( l + m )!

Z 2 e 4

8 20 h 2 n 2
with n = 1, 2, 3,... and

l n 1 and

l m l

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

(11.3.1)

11 Hydrogenic Atom

Dr. P. Blmler: PH 502 Wave Mechanics

page: 96

As expected the three degrees of freedom in motion generate three quantisation conditions,
described by the three quantum numbers n, m and l. Different nomenclature is in use for them:
n

principal quantum number:


because n defines the shell

n = 1, 2, 3, 4,....
n = K, L, M, N, ....

orbital quantum number (sometimes also called azimuthal quantum number): l n + 1


l = 0, 1, 2, 3, 4, 5..., (n-1)
l = s, p, d, f, g, h

spectroscopic notation:
m

l m l

magnetic quantum number (sometimes also ml )

m = -l, (-l+1),....-3, -2, -1, 0, 1, 2, 3,... , (l-1), l


m = -l, (-l+1), ..., , , , , , , ,... , (l-1), l

spectroscopic notation:

The following table is illustrating the mutual dependence of the quantum numbers.
Tab. 11.2: Summary of the interdependence of the quantum numbers for the hydrogenic atom, for
n = 1, 2 and 3 including nomenclature and degeneracy. (Neglecting the spin of the
electron, which is done in the next lecture)
n
1
2
3
l
m

0
0

0
0

name
sp.n.*

1s

2s

1s

2s

-1

1
0

2p
2p

2p

0
0

-1

3s
2p

D(n)**
1
4
*spectroscopic name including m-notation.
** D(n) degeneracy of state.

3s

1
0

-2

-1

3p
3p

3p

2
0

3d

3d

3d
3p

3d

3d

3d

The energy is determined by n only (cf. eq. (11.2.25)). Hence we have 2l+1 degenerate states
(span of m), where the span of l is defined by n-1. Therefore the degeneracy of a state n is given by:
n 1

D( n) =

(2l + 1) = 1 + 3 + 5 + .... + (2n 3) + (2n 1) = 2 2n = n 2


n

(11.3.2)

l =0

Combining the equations in (11.3.1) gives the complete wavefunctions for the hydrogenic atom as
listed in Tab. 11.3:

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

11 Hydrogenic Atom

page: 97

Tab. 11.3: The complete wavefunctions, nlm ( , , ) , of the hydrogenic atom for
n = 1, 2 and 3.
Name

nlm ( , , )

1s

1
e

2s

1
(2 )e / 2
4 2

2p

-1

1
8

2p

2p

e / 2 sin e i

1
e / 2 cos
4 2
1
8

e / 2 sin e i

3s

1
27 18 + 2 2 e / 3
81 3

3p

-1

1
(6 ) e / 3 sin e i
81

3p

2
(6 ) e / 3 cos
81

3p

1
(6 ) e / 3 sin ei
81

3d

-2

1
2 e / 3 sin 2 e 2i
162

3d

-1

1
2 e / 3 sin cos e i
81

3d

1
81 6

2 e / 3 3 cos 2 1

3d

1
2 e / 3 sin cos ei
81

3d

1
2 e / 3 sin 2 e 2i
162

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

11 Hydrogenic Atom

Dr. P. Blmler: PH 502 Wave Mechanics

page: 98

For the calculation of the normalisation constant N nlm it is important to realise, that the integration is
(preferably) done in spherical co-ordinates (cf. Tab. E2.1) and the integrand can be further
simplified according to eq. (10.4.25), where we have learned that the -term behaves constant.
2
N nlm

2
N nlm

*( x, y, z)(x, y, z)dxdydz

( E 2. 1)

0 0 0
2

* ( r , , ) ( r , , ) r 2 sin dr dd

(10 . 4. 25 )

(11.3.3)

r = 0 =0 =0
2
N nlm
2

[rR(r )()]

sin drd = 1

r =0 =0

This procedure can be easily understood, when we have a look at the s-orbitals, which according to
l = 0 have spherical symmetry. Hence, we can replace the radial probability distribution function (e.g.
in Fig. 11.3/right) by an integration over spherical shells of thickness dr. Hence, we have to substitute
the probing volume (d) of the radial probability distribution function, Pn 0 () d = Rn20 () d ,
with the volume of the spherical shells (42 d) to get the projection of the three-dimensional
~
distribution function, Pn0 () (we call it the radial distribution function), of the electron on the
radius or
~
Pn 0 () d = 42 Rn20 () d

(11.3.4)

This principle is illustrated in Fig. 11.4 using the 1s-orbital as an example. From the plot of the
radial distribution function for the 1s, 2s and 3s-orbital (see Fig. 11.5) the Bohrs idea of shells, in
which the electron is located can be anticipated, but we also realise, that the nth shell has only the
maximum probability. The electron has a finite probability to be found in lower shells. (Closer
examination also uncovers that -except of 1s- that the maximum is not at integer multiples of rB) We
also realise that there is a small but non-zero probability to find the electron inside the nucleus,
assuming the nucleus has a finite radius. This is of great importance for nuclear reaction e.g. +decay via electron capture.
Finally we want to visualise the atomic orbitals, which is a difficult task, because we have a 3Dwavefunction which has a different values at the co-ordinates ,, . Therefore, we have to depict it
either as a cloud of varying density, or we have to pick a certain threshold of probability to which we
assign a surface (isosurface). Both ways of visualisation are shown in Tab. 11.4, where the different
signs of the lobes have been colour-coded (positive = red, negative = green).
For an on-line visualisation, you can download Orbital-Viewer (a program from D. Manthey) from
the webpage (http://wwwnmr.ukc.ac.uk/staff/pb/teach/PH502/Aos/ ao.html) of the
course.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

11 Hydrogenic Atom

Dr. P. Blmler: PH 502 Wave Mechanics

page: 99

P()

~
P ()

sampled volume 42d

d
1
r=rB

Fig. 11.4: Illustration for the construction of the radial probability density function for the 1s-orbital:
Upper: If an infinitesimal small and constant sampling volume d is scanned along the
radius, we will receive the radial-part of the electron density P(). (If additionally the
detector would be moved on a circle, for each the same value would be recorded. The
1s-orbital is isotropic.) Lower: Now the electron density is measured over a spherical
~
shell of thickness d, to determine the radial probability distribution function P () , which
has a maximum at the Bohr radius ( = 1).

16
14

2.5

~
P10( )

2
1.5

1
2

0.5
0

12
10
8

~
P20( )

10

15

20

25

30

10

15

20

25

30

P~30 ()

6
4
2
0

10

15

20

25

30

Fig. 11.5: The radial distribution function (not-normalised) for the first three s-states (l = 0).
Compare with the according states in Fig. 11.3.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

11 Hydrogenic Atom

Dr. P. Blmler: PH 502 Wave Mechanics

page: 100

Tab. 11.4: Atomic orbitals for hydrogenic atoms for n = 1, 2 and 3. On the left of each cell the orbital is shown as a probability plot using 10000 points. On
the right (for ns-orbitals in the centre) an isosurface (corresponding to a probability of 10-4) is shown. For the ns-orbitals additionally, an octant of
this isosurface has been cut away to show the inner shells. Red colour corresponds to positive lobes, green to negative.

0/s

0/s

-2

-1

1/p

0/s

1/p

2/d

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

m
0

+1

+2

Dr. P. Blmler: PH 502 Wave Mechanics

11 Hydrogenic Atom

page: 101

Fig. 11.6: Energy levels for hydrogen 1H including the l 5. Energy is measured in units of
h 2 /( 2rB2 ) .

11.4 Linear Combinations: Hybrid Orbitals


We have seen from Tab. 11.3 that the wavefunctions of the hydrogenic atom are partially real and
partially imaginary. The first case applies for m = 0 the latter for m 0. The there are n2 degenerated
eigen-functions belonging to the same n, we can create new functions by linear combination, which
are entirely real. This has practical applications in chemistry. Since the new functions are created
from the set of solutions, we call them hybrid orbitals. For instance
1
(2 p + 2 p ) = 1 e / 2 sin [cos + i sin ] + e / 2 sin [cos i sin ]

8 2

(E2.22)
x
2
1
/ 2
[
]
=
e
sin cos
=
e / 2 2px
8 2
4 2
rB

Analogously we get:
(E2.22)
y
1
1
(2 p 2 p ) = 1 e / 2 sin sin =
e / 2 2p y

4 2
4 2
rB

2 p =

(E2. 22 )
z
1 2 / 2
1
e
cos =
e / 2 2p z
8
4 2
rB

Hence, we receive three equivalent 2p-functions, which can be correlated to the three Cartesian coordinates. This is useful for geometrical considerations (e.g. chemical complexes).
Similarly we proceed with the other wavefunctions (see Tab. 11.5)
Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

11 Hydrogenic Atom

page: 102

Tab. 11.5: Hybrid orbitals for hydrogenic atoms for n = 1, 2 and 3.


Symbol

combination

hybrid wavefunction

1s

1s

1
e

2s

2s

1
(2 )e / 2
4 2

2pz

2p

1
e / 2 cos
4 2

2px

1
(2p + 2p )
2

1
e / 2 sin cos
4 2

2py

m1

1
(2p 2p )
2

1
e / 2 sin sin
4 2

3s

3s

1
27 18 + 2 2 e / 3
81 3

3pz

3p

2
(6 ) e / 3 cos
81

3px

1
(3p + 3p )
2

2
(6 ) e / 3 sin cos
81

3py

m1

1
(3p 3p)
2

2
(6 ) e / 3 sin sin
81

3d z 2

3d

3dxz

1
(3d + 3d )
2

2 2 / 3
e
sin cos cos
81

3dyz

m1

i
(3d 3d )
2

2 2 / 3
e
sin cos sin
81

3d x 2 y 2 3

1
(3d + 3d )
2

3dxy

1
81 6

m2

i
(3d 3d )
2

1
81 2
1
81 2

2 e / 3 3 cos 2 1

2 e / 3 sin 2 cos 2
2 e / 3 sin 2 sin 2

The hybrid orbitals do not look different from those in Tab. 11.4. They are just transferred back to
Cartesian co-ordinates (with slightly different normalisation constants).

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Dr. P. Blmler: PH 502 Wave Mechanics

Thanks!

page: 103

Acknowledgements:
I would like to thank Dr. C. Isenberg to provide access to his notes for a first layout of the course.
They have turned out to be very helpful and guiding for a new lecturer. I also want to thank Mr. Ian
Campbell for proof-reading the first draft of the manuscript.

Dr. P. Blmler, School of Physical Sciences, University of Canterbury, UK

Das könnte Ihnen auch gefallen