You are on page 1of 151

PHYS2222

Quantum Physics
Reading list
 For general reference the second-year course book: Introduction to the Structure
of Matter by J.J. Brehm and W.J. Mullin (Wiley, 1989, ISBN 0-471-60531-X).
10 copies in UCL library, available from P&A Department. Referred to as B&M
in these notes.
Advantages: suitable for most 2 nd-year Physics courses, good integration of
quantum physics with atomic physics.
Disadvantages: weak on more formal aspects of quantum mechanics.
 As a reasonably priced introduction: Quantum Mechanics by A.I.M. Rae (4th
edition, Institute of Physics Publishing, 2002, ISBN 0 7503 0839 7).
9 copies in UCL library, available from P&A Department.
Advantages: cheap, well suited to level of course, covers essentially all the
material at roughly the right level.
Disadvantages: not so useful for other courses.
 As a more advanced book that is also recommended for the third-year quantum
mechanics course: Quantum Mechanics by B.H. Bransden and C.J. Joachain (2nd
edition, Prentice Hall, 2000, ISBN 0582-35691-1).
10 copies in UCL library, available from P&A Department. Referred to as B&J
in these notes.
Advantages: material for 2222 is mostly presented at the start of the book.
Contains additional material going well beyond 2B22 for further reading. Useful
for both 3rd -year and 4th-year courses.
Disdavantages: coverage of some material (notably spin and emission/absorption
of radiation) is at a more advanced level than 2222 and is not so useful for this
course. Relatively expensive (£41 on Amazon).

2222 Quantum Physics 2006-7 1


Prerequisites
 First-year Mathematics for Physics and Astronomy (1B45
and 1B46) or equivalent
 Material from the second-year maths course (2246) will be
used after it has been covered in that course
 Basic relativistic kinematics (from 1B46) will be assumed,
and basic electromagnetism (field and potential of point
charge, interaction of magnetic dipole with magnetic field)
will be used as it is covered in 2201

2222 and other courses


 Some limited overlap with 1B23 Modern Physics,
Astronomy and Cosmology (bur different approach – 2222
is less descriptive and more rigorous). Areas covered by
both courses:
o Wave-particle duality (photoelectric effect, double-
slit experiment)
o Time-independent Schrődinger equation
o Significance of wave function and Heisenberg’s
Uncertainty Principle
1B23 is not a prerequisite for 2222!
2222 Quantum Physics 2006-7 2
Syllabus
1. The failure of classical mechanics [3 lectures] 
Photoelectric  effect,  Einstein’s  equation,  electron  diffraction  and  de  Broglie  relation.  
Compton scattering.  Wave­particle duality, Uncertainty principle (Bohr microscope).   
2. Steps towards wave mechanics [3 lectures] 
Time­dependent and  time­independent  Schrödinger equations.  The wave function and its 
interpretation. 
3. One­dimensional time­independent problems [7 lectures] 
Infinite square well potential.  Finite square well.  Probability flux and the potential barrier 
and  step.    Reflection  and  transmission.    Tunnelling  and  examples  in  physics  and 
astronomy.  Wavepackets.  The simple harmonic oscillator. 
4. The formal basis of quantum mechanics [5 lectures] 
The  postulates  of  quantum  mechanics  –  operators,  observables,  eigenvalues  and 
eigenfunctions.  Hermitian operators and the Expansion Postulate.   
5. Angular momentum in quantum mechanics [2 lectures] 
Operators, eigenvalues and eigenfunctions of  L  ˆ   and  L 
ˆ 2 . 
z

2222 Quantum Physics 2006-7 3


Syllabus (contd)

6. Three dimensional problems and the hydrogen atom [4 lectures] 
Separation of variables for a three­dimensional rectangular well.  Separation of space and time 
parts of the 3D Schrödinger equation for a central field.  The radial Schrödinger equation, and 
casting  it  in  a  form  suitable  for  solution  by  series  method.    Degeneracy  and  spectroscopic 
notation. 
7. Electron spin and total angular momentum [3 lectures] 
Magnetic moment of electron due to orbital motion.  The Stern­Gerlach experiment.  Electron 
spin  and  complete  set  of  quantum  numbers  for  the  hydrogen  atom.    Rules  for  addition  of 
angular  momentum  quantum  numbers.    Total  spin  and  orbital  angular  momentum  quantum 
numbers S, L, J.  Construct J from S and L.  

2222 Quantum Physics 2006-7 4


Photo-electric effect, Compton Davisson-Germer experiment, double-
scattering E  h slit experiment

h
p
Particle nature of light in
quantum mechanics
 Wave nature of matter in quantum
mechanics

Wave-particle duality

Postulates:
Time-dependent Schrödinger equation,
Born interpretation Operators,eigenvalues and
eigenfunctions, expansions in
Separation of variables
2246 Maths complete sets, commutators,
Methods III Time-independent Schrödinger expectation values, time
Frobenius equation evolution
method
Quantum simple Legendre
harmonic oscillator Hydrogenic atom 1D problems equation 2246

 
E n = n
1
2
ℏ 0
Angular momentum
operators
Radial solution Angular solution
2 Lˆz , Lˆ2
1Z Yl m ( ,  )
Rnl , E  
2 n2
2222 Quantum Physics 2006-7 5
Lecture style
• Experience (and feedback) suggests the biggest problems found by
students in lectures are:
– Pacing of lectures
– Presentation and retention of mathematically complex material
• Our solution for 2222:
– Use powerpoint presentation via data projector or printed OHP for written
material and diagrams
– Use whiteboard or handwritten OHP for equations in all mathematically
complex parts of the syllabus
– Student copies of notes will require annotation with these mathematical
details
– Notes (un-annotated) will be available for download via website or (for a
small charge) from the Physics & Astronomy Office
• Headings for sections relating to key concepts are marked with
asterisks (***)

2222 Quantum Physics 2006-7 6


Hertz J.J. Thomson
B&M §2.5; Rae §1.1;
B&J §1.2
1.1 Photoelectric effect
Metal plate in a vacuum, irradiated by ultraviolet light, emits
charged particles (Hertz 1887), which were subsequently shown
to be electrons by J.J. Thomson (1899).
Classical expectations
Light, frequency ν Electric field E of light exerts force F=-
Vacuum eE on electrons. As intensity of light
chamber Collecting increases, force increases, so KE of
Metal plate ejected electrons should increase.
plate

Electrons should be emitted whatever


the frequency ν of the light, so long as E
is sufficiently large

I
For very low intensities, expect a time lag
Ammeter between light exposure and emission,
while electrons absorb enough energy to
Potentiostat escape from material

2222 Quantum Physics 2006-7 7


Einstein

Photoelectric effect (contd)***


Actual results: Einstein’s
Maximum KE of ejected electrons is interpretation (1905):
independent of intensity, but light is emitted and
dependent on ν absorbed in packets
(quanta) of energy
For ν<ν0 (i.e. for frequencies below Millikan
a cut-off frequency) no electrons E  h (1.1)
are emitted An electron absorbs a
There is no time lag. However, rate single quantum in
of ejection of electrons depends on order to leave the
light intensity. material

The maximum KE of an emitted electron is then predicted to be:

K max  h  W (1.2)
Verified in detail
Work function: minimum energy through
Planck constant:
needed for electron to escape subsequent
universal constant of
from metal (depends on experiments by
nature
material, but usually 2-5eV) Millikan
h=6.63×10−34 Js 2222 Quantum Physics 2006-7 8
Photoemission experiments today
Modern successor to original photoelectric
effect experiments is ARPES (Angle-
Resolved Photoemission Spectroscopy)

Emitted electrons give information on


distribution of electrons within a material as
a function of energy and momentum

2222 Quantum Physics 2006-7 9


Frequency and wavelength for
light***
Relativistic relationship between a
particle’s momentum and energy: E 2  p 2 c 2  m0 2 c 4

For massless particles


propagating at the speed of light,
E 2  p 2c 2
becomes E =c∣ p∣

h=c∣ p∣

Hence find relationship h


∣ p∣=
between momentum p and c
wavelength λ:
But:
 =c

h
∣ p∣= (For light.)

2222 Quantum Physics 2006-7 10


Compton
B&M §2.7; Rae §1.2;
B&J §1.3
1.2 Compton scattering
Compton (1923) measured scattered intensity of X-rays (with well-defined wavelength)
from solid target, as function of wavelength for different angles.

X-ray source
Collimator Crystal
(selects angle) (selects
wavelength)

θ
Target

Result: peak in the wavelength


distribution of scattered radiation shifts to
longer wavelength than source, by an Detector
amount that depends on the scattering
angle θ (but not on the target material)
A.H. Compton, Phys.
Rev. 22 409 (1923)
2222 Quantum Physics 2006-7 11
Compton scattering (contd)
Classical picture: oscillating electromagnetic field would cause oscillations
in positions of charged particles, re-radiation in all directions at same
frequency and wavelength as incident radiation
p’
Photon
Incoming photon
Before p
After θ
Frequency  Electron
φ
pe
Compton’s explanation: “billiard ball” collisions between
X-ray photons and electrons in the material

Conservation of energy: 1 Conservation of momentum:


2 2 2 2 4 2
h m e c =h '  p e c me c  In electrons initial rest frame:
Write in terms of momentum: 1
p= p ' p e
2 2 2
pcme c = p ' c p e c me c 2 4 2 p e= p− p '

2 2 2 2
1
4 2
∣ pe 2∣ = p− p ' .  p− p ' =∣ p2∣∣ p ' 2∣ −2∣ p∣∣ p '∣ cos
 p− p '  cm e c = p e c m e c  ∣ pe 2∣= p− p ' 2 2 1−cos  pp '
2 2 2 4 3 2 2 4
 p− p'  c me c 2 p− p '  me c = p c me c
2
 p− p '  2 p− p '  me c= pe
2 2222 Quantum Physics 2006-7 12
Compton scattering (contd)
Equate above for p e 2 :
2 2
 p− p '  2 p− p '  m e c= p− p '  21−cos  pp '
⇒ p− p '  m e c=1−cos  pp'
1 1
⇒ me c − =1−cos 
p' p

Assuming photon
h
momentum related to p (1.3)
wavelength: 
h
me c  '   (1  cos  ) (1.4)
⇒ ' −=1−cos  me c
h
‘Compton wavelength’
of electron (0.0243 Å)

2222 Quantum Physics 2006-7 13


Puzzle

What is the origin of the component of


the scattered radiation that is not
wavelength-shifted?

ℏ2 2 x x ℏ2 2 2 2mE
Substitute: −  e =Ee ⇒ −  =E ⇒  =− 2
2m 2m ℏ
⇒ Assuming E0 RHS is -ve  is imaginary.

Second peak at original  due


2mE ℏ2 k 2
=±ikwherek 2= 2 or =E kis DeBroglie wave-number.p=ℏk
ℏ 2m

to nuclei collisions.

2222 Quantum Physics 2006-7 14


Wave-particle duality for light***
“ There are therefore now two theories of light, both indispensable, and - as one
must admit today despite twenty years of tremendous effort on the part of
theoretical physicists - without any logical connection.” A. Einstein (1924)

•Light exhibits diffraction and interference phenomena that


are only explicable in terms of wave properties
•Light is always detected as packets (photons); if we look, we
never observe half a photon
•Number of photons proportional to energy density (i.e. to
square of electromagnetic field strength)

2222 Quantum Physics 2006-7 15


De Broglie
B&M §4.1-2; Rae §1.4;
B&J §1.6
1.3 Matter waves***
“As in my conversations with my brother we always arrived at the conclusion that in
the case of X-rays one had both waves and corpuscles, thus suddenly - ... it was
certain in the course of summer 1923 - I got the idea that one had to extend this
duality to material particles, especially to electrons. And I realised that, on the one
hand, the Hamilton-Jacobi theory pointed somewhat in that direction, for it can be
applied to particles and, in addition, it represents a geometrical optics; on the other
hand, in quantum phenomena one obtains quantum numbers, which are rarely found
in mechanics but occur very frequently in wave phenomena and in all problems
dealing with wave motion.” L. de Broglie

Proposal: dual wave-particle nature of radiation


also applies to matter. Any object having
momentum p has an associated wave whose Prediction: crystals
wavelength λ obeys (already used for X-
ray diffraction) might
h
p= =ℏ k

k
2

(wavenumber)
 ℏ=
h
2  also diffract particles

2222 Quantum Physics 2006-7 16


Electron diffraction from crystals
Davisson G.P. Thomson
The Davisson-Germer experiment
θi (1927): scattering a beam of
electrons from a Ni crystal

θr
At fixed angle, find sharp peaks in intensity
as a function of electron energy

Davisson, C. J.,
"Are Electrons
At fixed accelerating voltage (i.e. fixed Waves?," Franklin
electron energy) find a pattern of pencil- Institute Journal
sharp reflected beams from the crystal 205, 597 (1928)

G.P. Thomson performed similar interference


experiments with thin-film samples

2222 Quantum Physics 2006-7 17


Electron diffraction from crystals
(contd)
Interpretation used similar ideas to those pioneered for scattering
of X-rays from crystals by William and Lawrence Bragg

Path difference:
William Bragg θi a cos r −a cos  i
(Quain Professor a cos  i Lawrence
of Physics, UCL, a cos  r −cos i  Bragg
1915-1923) Constructive interference when
θr
a cos  r −cos i =n
a
Modern Low Energy
Electron scattering Electron Diffraction
dominated by surface (LEED): this pattern of
layers “spots” shows the beams
Note θi and θr not of electrons produced by
surface scattering from
necessarily equal a cos  r complex (7×7)
Note difference from usual “Bragg’s reconstruction of a silicon
Law” geometry: the identical scattering surface
planes are oriented perpendicular to
the surface 2222 Quantum Physics 2006-7 18
The double-slit interference
experiment
Originally performed by Young (1801) with light. Subsequently also performed
with many types of matter particle (see references).

Alternative method
of detection: scan
y a detector across
the plane and
record arrivals at
d each point
θ
Incoming beam of Detecting
particles (or light) d sin 
screen
(scintillators or
particle
D detectors)

2222 Quantum Physics 2006-7 19


Results
Neutrons, A
Zeilinger et al. 1988
Reviews of Modern
Physics 60 1067-
1073

He atoms: O Carnal and J Mlynek


1991 Physical Review Letters 66
2689-2692

Fringe visibility C60 molecules: M


decreases as Arndt et al. 1999
molecules are Nature 401 680-
heated. L. 682
Hackermüller et
al. 2004 Nature With multiple-
427 711-714 slit grating

Without grating
2222 Quantum Physics 2006-7 20
Double-slit experiment: interpretation
Interpretation: maxima and minima arise from alternating constructive and
destructive interference between the waves from the two slits

Spacing between maxima:


Constructive interference:
d sin =n  n D
⇒ Maximum at y n = D tan  n≈ D  n ≈
⇒ n th maximum occurs at: d
d sin  n =n 
For small  ,sin ≈tan  D
⇒ Spacing at
n d
 n≈
d
Example: He atoms at a temperature
of 83K, with d=8μm and D=64cm
p2 3 h
Expect: = K T ⇒ p=4.8×10−24 Ns⇒ = =1.03 Å
2m 2 B p
D −6
⇒ Spacing = =8.24×10 m
d
2222 Quantum Physics 2006-7 21
Double-slit experiment: bibliography
Some key papers in the development of the double-slit experiment during the
20th century:
• Performed with a light source so faint that only one photon exists in the
apparatus at any one time (G I Taylor 1909 Proceedings of the Cambridge
Philosophical Society 15 114-115)
• Performed with electrons (C Jönsson 1961 Zeitschrift für Physik 161 454-
474, translated 1974 American Journal of Physics 42 4-11)
• Performed with single electrons (A Tonomura et al. 1989 American Journal
of Physics 57 117-120)
• Performed with neutrons (A Zeilinger et al. 1988 Reviews of Modern
Physics 60 1067-1073)
• Performed with He atoms (O Carnal and J Mlynek 1991 Physical Review
Letters 66 2689-2692)
• Performed with C60 molecules (M Arndt et al. 1999 Nature 401 680-682)
• Performed with C70 molecules, showing reduction in fringe visibility as
temperature rises so molecules “give away” their position by emitting
photons (L. Hackermüller et al. 2004 Nature 427 711-714)
An excellent summary is available in Physics World (September 2002 issue,
page 15) and at http://physicsweb.org/ (readers voted the double-slit
experiment “the most beautiful in physics”).
2222 Quantum Physics 2006-7 22
Matter waves: key points***
• Interference occurs even when only a single particle (e.g. photon or electron) in
apparatus, so wave is a property of a single particle
– A particle can “interfere with itself”
• Wavelength unconnected with internal lengthscales of object, determined by
momentum
• Attempt to find out which slit particle moves through causes collapse of
interference pattern (see later…)

Wave-particle duality for matter particles

•Particles exhibit diffraction and interference phenomena that


are only explicable in terms of wave properties
•Particles always detected individually; if we look, we never
observe half an electron
•Number of particles proportional to….???

2222 Quantum Physics 2006-7 23


B&M §4.5; Rae §1.5; B&J §2.5 (first part only)
1.4 Heisenberg’s gamma-ray microscope and a first look at the
Uncertainty Principle
The combination of wave and particle pictures, and in particular the significance of
the ‘wave function’ in quantum mechanics (see also §2), involves uncertainty: we
only know the probability that the particle will be found near a particular location.

Screen forming
image of particle

Particle
θ/2

y
Light source,
wavelength λ

Lens, having angular


diameter θ

 Heisenberg
Resolving power of lens:  y

2222 Quantum Physics 2006-7 24
Heisenberg’s gamma-ray microscope and the Uncertainty
Principle***
Range of y-momenta of photons after
scattering, if they have initial momentum p: p

Py
Y-component of  momentum θ/2
after scattering.
px
− psin 
2
 p y  psin

2 p
(Assuming magnitude of
p unchanged i.e.
⇒ Partice momentum must lie Neglecting Compton
in the same range: effect.)

 p y =2psin 

2
≃ p (Small angle/small lens)

Remember resolution power of  y  p y h


lens from classical optics:
 Heisenberg’s Uncertainty
 y Principle

2222 Quantum Physics 2006-7 25
  y  p y h
Rae §2.1, B&J §3.1, B&M §5.1

2.1 An equation for the matter waves: the time-


dependent Schrődinger equation***
Classical wave equation (in
one dimension):  ( x, t )  wave displacement (in 1d)
2  1  2 
= 2
x 2
 t2 v  wave velocity

 ( x, t )
e.g. Transverse waves on
a string:

Can we use this to describe the


matter waves in free space?

2222 Quantum Physics 2006-7 26


An equation for the matter waves (2)
Seem to need an equation that   2 
Try:  =
involves the first derivative in  t  x2
time, but the second derivative in
space  ( x, t ) is "wave function" associated with matter wave

Put:  x , t = Ae i kx− t  So, wave equation is:


d2 d 2
 =−k 2
Ae i kx− t 
; =−i  Ae i  kx− t 
2m d  d 
dx
2
dt = 2
i ℏ dt dx
⇒−i  =−k 2

 
2 2
p2 ℏ k  −ℏ
We want the energy to be ⇒ ℏ = Multiplying both sides by
2m 2m 2m
So choose  so this is true.
ℏk2
⇒−i  =−k 2
2m  ℏ 2 2 
2m iℏ =−
⇒ = t 2m  x 2
iℏ
(for matter waves in free space)
2222 Quantum Physics 2006-7 27
An equation for the matter waves (3)
p2
For particle with potential energy V(x,t), E  V ( x, t )
2m
need to modify the relationship between
energy and momentum: Total energy = kinetic energy + potential energy

Suggests corresponding modification to


Schrődinger equation: d  −ℏ d 
2 2
We had: i ℏ =
dt 2m dx 2
LHS: Gives ℏ   for a plane wave.
ℏ k 2
RHS: Gives  for a plane wave.
2m
⇒Try adding a term V  x , t  on RHS.

 ℏ 2 2 
iℏ =− 2
V  x ,t 
t 2m  x
Time-dependent Schrődinger equation
2222 Quantum Physics 2006-7 28
Schrődinger
The Schrődinger equation: notes
•This was a plausibility argument, not a derivation. We believe the Schrődinger
equation to be true not because of this argument, but because its predictions agree
with experiment.
•There are limits to its validity. In this form it applies to
•A single particle, moving in one direction, that is
•Non-relativistic (i.e. has non-zero rest mass and velocity very much below c)
•The Schrődinger equation is a partial differential equation in x and t (like classical
wave equation)
•The Schrődinger equation contains the complex number i. Therefore its solutions
are essentially complex (unlike classical waves, where the use of complex numbers
is just a mathematical convenience)
•Positive i is merely a convention.

2222 Quantum Physics 2006-7 29


The Hamiltonian operator
 ℏ 2 2 
iℏ =− V  x ,t 
t 2m  x 2

Can think of the RHS of the Schrődinger equation as a


differential operator that represents the energy of the
particle.

This operator is called the Hamiltonian of the

[ ]
2 2
particle, and usually given the symbol Ĥ ℏ d
− 2
 V  x , t ≡ H 
2m d x

Hence there is an alternative (shorthand) form Kinetic Potential


for time-dependent Schrődinger equation: energy energy
operator operator
 
iℏ =H 
t
2222 Quantum Physics 2006-7 30
Rae §2.1, B&J §2.2, B&M §5.2

2.2 The significance of the wave function***


Ψ is a complex quantity, so what can be its significance for the results
of real physical measurements on a system?

Remember photons: number of photons per unit volume is


proportional to the electromagnetic energy per unit volume,
hence to square of electromagnetic field strength.

Postulate (Born interpretation): probability of finding particle in a small


length δx at position x and time t is equal to 2
 ( x, t )  x (2.6)
2
   *

Note: |Ψ(x,t)|2 is real, so probability is also real, as required.

|Ψ|2 δx
Total probability of finding particle between
positions a and b is
b b

∑ ∣  x , t ∣  x ∫∣  x , t ∣ dx
2 2
 x0
x =a a x
Born 2222 Quantum Physics 2006-7 a b 31
Example
Suppose that at some instant of time a particle’s wavefunction is

{
 x , t  2x
0
(For  0 x0.909 )
(Otherwise) }
What is: Probability: ∣ ∣  x
2

2
(a) The probability of finding With: x=0.5, =1, ⇒ ∣∣ =1
the particle between x=0.5 and ⇒ Probability =1 x 10
−4

x=0.5001?
2
(b) The probability per unit Probability for unit length: ∣∣
length of finding the particle = 1.44
at x=0.6?
0.5

Total probability =∫∣∣  x


2
(c) The probability of finding 0
the particle between x=0.0 and 0.5
x=0.5? = ∫ 4x 2  x=0.167
0

2222 Quantum Physics 2006-7 32


Normalization
Total probability for particle to be anywhere should be one (at
any time): ∞

∫ ∣  x , t ∣ dx=1
2
Normalization condition
−∞

Suppose we have a solution to the ∞

N = ∫ ∣∣  x
2
Schrődinger equation that is not
normalized, Then we can −∞
1
•Calculate the normalization integral −
2
  x , t  N
 x ,t
•Re-scale the wave function as ∞ ∞ 2
∣ ∣

So: ∫ ∣ new∣  x=∫
2
(This works because any solution to the x
S.E., multiplied by a constant, remains a −∞ −∞ N
solution, because the equation is linear and N
homogeneous) = =1
N
Alternatively: solution to Schrödinger equation contains an arbitrary
constant, which can be fixed by imposing the condition (2.7)
2222 Quantum Physics 2006-7 33
Normalizing a wavefunction - example
Suppose that at some time we have the following form to the wave-function:


 x , t 
{  a 2− x 2 for −a xa
0 otherwise }
Normalisation integral:
∞ a

N = ∫ ∣  x , t ∣  x=∫   x , t 
2 2

−∞ −a

[ ]
−a a x a 3 a
x
= ∫ a − x  x= a x−
2 2 2

−a 3 −a

=
2a 3
3
− −
3
=  
2a 3 4a 3
3
This is not, in general, equal to


1. Therefore  is not correctly 3
3a
normalised. To get a correctly   x , t =
4
 a 2− x 2
normalised wave-function, take:

2222 Quantum Physics 2006-7 34


Rae §2.3, B&J §3.1

2.3 Boundary conditions for the wave-function


Unacceptable 


The wavefunction must:
Multivalued.
1. Be a continuous and single-valued function
of both x and t (in order that the probability x
density be uniquely defined) 

Discontinuous.

 x
2. Have a continuous first derivative (unless the Discontinuous.
potential goes to infinity) d
Change in only
dx
allowed if potential ±∞
x
3. Have a finite normalization integral. ∞

0 ∫ ∣∣  x∞
2

−∞
2222 Quantum Physics 2006-7 35
Rae §2.2, B&J §3.5, B&M §5.3
2.4 Time-independent Schrődinger
equation***
Suppose potential V(x,t) (and hence force on  ℏ   2 2
iℏ =− V  x 
particle) is independent of time t: t 2m  x 2

LHS involves only RHS involves only variation of Ψ with


variation of Ψ with t x (i.e. Hamiltonian operator does not
depend on t)
Look for a solution in which the time and
space dependence of Ψ are separated:   x , t= x T t 

Substitute: dT ℏ2 d2
i ℏ   x ,t  =− T t  2 V  x x T t 
dt 2m dx
Divide by   xT t : Equal to constant
i ℏ dT ℏ2 d 2  E as both sides are
=− V  x independent of
T dt 2m  dx 2
each other.
2 2
i ℏ dT ℏ d 
=− V  x = E
T dt 2m  dx 2
Independent Independent of
2222 Quantum Physics 2006-7 36
of position x. time t.
Time-independent Schrődinger equation
Solving the time equation: i ℏ
(contd)
dT dT
=E ⇒ i ℏ =ET
T dt dt
Linear ordinary differential equation with constant coefficients.
t
Try: T = Ae ⇒ Substitute.
iE
i ℏ =E ⇒ =−

iEt


Solution is: T t =e
The space equation becomes:
ℏ2 d 2 
− V  x= E
2m  dx2
Now multiply by  to get the final version.
2 2
ℏ d   =E 
− 2
V  x=E  or H
2m d x
Time-independent Schrődinger equation
2222 Quantum Physics 2006-7 37
Notes
• In one space dimension, the time-independent Schrődinger equation is an
ordinary differential equation (not a partial differential equation)
• The sign of i in the time evolution is determined by the choice of the sign of i in
the time-dependent Schrődinger equation
• The time-independent Schrődinger equation can be thought of as an eigenvalue
equation for the Hamiltonian operator:
Operator × function = number × function
Ĥ  E
(Compare Matrix × vector = number × vector) [See 2246]
• We will consistently use uppercase Ψ(x,t) for the full wavefunction (time-
dependent Schrődinger equation), and lowercase ψ(x) for the spatial part of the
wavefunction when time and space have been separated (time-independent
Schrődinger equation)
• Probability distribution of particle is now independent of time (“stationary
state”):

∣  x∣
−i E t 2
2
∣ x ,t ∣ = e ℏ

=∣e ∣ ∣ x ∣
−i E t 2
For a stationary state we can use ℏ 2
either ψ(x) or Ψ(x,t) to compute 2
probabilities; we will get the same =∣ x∣
result. 2222 Quantum Physics 2006-7 38
Rae §3.1, B&J §3.1, B&M §5.1

2.6 SE in three dimensions


To apply the Schrődinger equation in the real (three-dimensional)  
world we keep the same basic structure: iℏ =H  Ĥ  E
t

BUT
  x , t     x , y , z , tor  r , t
Wavefunction and potential energy are now  x    x , y , zor r 
functions of three spatial coordinates: V  x  V  x , y , z ,t or V r 
Kinetic energy now involves three 2 2 2
p
2
p x  p y  p z
components of momentum 
2m 2m

[ ]
2 2 2 2 2 2 2
ℏ  ℏ    ℏ 2
⇒ −  −   =− ∇
2m  x 2 2m  x 2  y 2  z 2 2m
Interpretation of wavefunction:
∣  x , y , z , t ∣2 V = Probability of finding particle ⇒∣∣2 = Probability per unit
in a small volume dV around volume.
(x, y, z)

2222 Quantum Physics 2006-7 39


Puzzle
The requirement that a plane wave
 ( x, t )  exp[i (kx  t )]
plus the energy-momentum relationship for free-non-relativistic particles
p2
E
2m
led us to the free-particle Schrődinger equation.

Can you use a similar argument to suggest an equation for free relativistic
particles, with energy-momentum relationship:
E 2  p 2 c 2  m0 2 c 4
Einstein: E =ℏ  DeBroglie: p=ℏ k
2 2  2 2 2  2
E =− 2
ℏ p =− 2

t x

2 2
2   2 2   2 4
⇒ Try: −ℏ 2
=−ℏ c 2
m 0 c  (Klein-Gordon equation.)
t x

2222 Quantum Physics 2006-7 40


3.1 A Free Particle
Free particle: experiences no forces so potential energy independent Linear ODE with
of position (take as zero) constant coefficients so
try
Time-independent Schrődinger
equation: ℏ2 d 2    exp( x)
− 2
=E 
2m d x

ℏ2 2  x x ℏ2 2 2 2mE
Substitute: −  e =Ee ⇒ −  =E ⇒  =− 2
2m 2m ℏ
⇒ Assuming E 0 RHS is -ve   is imaginary.
2 2
2mE ℏ k
=±ik where k 2 = 2 or =E k is DeBroglie wave-number. p=ℏ k
ℏ 2m

General solution:
ikx −ikx
  x=Ce  De ( C and D are constants.)
= C cos kx i sin kxD coskx−i sin kx
= A coskxB sin kx where: A=C D , B=iC −D
2222 Quantum Physics 2006-7 41
3.1 A Free Particle (Cont.)
Combine with time dependence to
get full wave function:
iEt Et
i
ℏ ikx ℏ
 x , t = x e =Ce De−ikx e
Et Et
i kx−  i kx− 
ℏ ℏ
 = Ce  De

Travelling wave Travelling wave


moving to the right. moving to the left.
t  x t  −x

2222 Quantum Physics 2006-7 42


Notes
• Plane wave is a solution (just as well, since our plausibility
argument for the Schrődinger equation was based on this
being so)
• Note signs:
– Sign of time term (-iωt) is fixed by sign adopted in time-dependent
Schrődinger Equation
– Sign of position term (±ikx) depends on propagation direction of
wave
• There is no restriction on the allowed energies, so there is a
continuum of states

2222 Quantum Physics 2006-7 43


Rae §2.4, B&J §4.5,
B&M §5.4
3.2 Infinite Square Well
Consider a particle confined to a finite V(x)
length –a<x<a by an infinitely high
potential barrier V =∞ V 0 V =∞

No solution in barrier region (particle would x


have infinite potential energy). -a a
2 2
In well region: ℏ  
− =E  As for free particle.
2m  x 2
2mE
⇒  x= A cos kxB sin kx With: k 2 = 2

Boundary conditions:
Continuity of ψ at x=a:  a=0 ⇒ A cos kaB sin  ka=0 Note discontinuity in dψ/dx
allowable, since potential
Continuity of ψ at x=-a: −a=0 ⇒ A cos −kaB sin −ka=0 is infinite
or A cos ka−B sin  ka=0

 Since: cos−=cos sin −=−sin 


Even function. Odd function. 
2222 Quantum Physics 2006-7 44
Infinite square well (2)
Add and subtract these
conditions: Adding gives: 2A cos  ka=0
Subtracting gives: 2B sin  ka=0
Either, a) Put B=0 and cos  ka=0
n Even solution: ψ(x)=ψ(-x)
⇒ ka=  n=1, 3,5, 7....
2
n
⇒k= ,  x= A cos kx
2a
Or, b) Put A=0 and sin  ka=0
n Odd solution: ψ(x)=-ψ(-x)
⇒ ka= n=2, 4, 6,8. ...
2
n
⇒k= ,  x=B sin  kx
2a
Energy:
Same Schrödinger equation for particle in box as for particle in free space.
→ Same relation between E and k.
ℏ 2 k 2 n2 2 ℏ 2 n2 2 ℏ 2
⇒ E= = E=
2m 2m 2a 2 8ma 2
2222 Quantum Physics 2006-7 45
Infinite well – normalization and notes
Normalization: Need to choose constants so  is normalised. (i.e: Total over integral is 1).
Even soln. n
a a k= ( n odd.)⇒sin±2ka=sin ±n =0
21 2a
∫∣∣  x=1⇒ ∫ ∣ A∣ 2 [1cos2kx] x
2
2 2 1
−a −a ⇒ RHS=∣ A∣ xa ⇒ Choose A so...∣ A∣ =
a
[ ]
a
∣ A∣2 1
⇒ x sin 2kx =1 1
2 2k −a  A=
a
Notes on the solution:
• Energy quantized to particular values (characteristic of bound-state problems in quantum
mechanics, where a particle is localized in a finite region of space.
• Potential is even under reflection; stationary state wavefunctions may be even or odd (we say
they have even or odd parity)
• Compare notation in 1B23 and in books:
– 1B23: well extended from x=0 to x=b
– Rae and B&J: well extends from x=-a to x=+a (as here)
– B&M: well extends from x=-a/2 to x=+a/2
(with corresponding differences in wavefunction)

2222 Quantum Physics 2006-7 46


The infinite well and the Uncertainty
Principle
Position uncertainty in well:

 x≃2a (We know the particle is inside the box.)

Momentum uncertainty in lowest state from classical argument (agrees


with fully quantum mechanical result, as we will see in §4)
State with n = 1 corresponds to a particle bouncing back and forth in the well.
Momentum will be alternately: ℏ
±ℏ k or ±
2a
ℏ h
⇒ Momentum uncertainty:  p=2 ℏ k = =
a 2a
Compare with Uncertainty
Principle:
h
So:  x  p=2a =h
2a Ground state close to
2222 Quantum Physics 2006-7 minimum uncertanty 47
Rae §2.4, B&J §4.6
3.3 Finite square well
V(x)
I II III
Now make the potential well
more realistic by making the
barriers a finite height V0 V0

Assume:  0E V 0 x
-a a
Region I: Region II: Region III:
2 2
ℏ   ℏ 2 2  ℏ2 2 
− V 0 =E  − =E  − V 0 =E 
2m  x 2 2m  x 2
x
2m  x 2
Try:  x=e x x
2 2 ⇒ =C ' e D ' e
ℏ  Free particle solutions apply:
− V 0 =E  (  same as in Region 1.)
2m
2 2
ℏ  = A cos  kxB sin  kx Decaying term forbidden.
=V 0− E 0
2m Growing term allowed.
2mE
 is real. k 2= 2 − x
ℏ = D ' e
2 2m V 0 −E 
=±  = 2

x − x
⇒=C e D e
2222 Quantum Physics 2006-7 48
D is not normalisable so decaying
term C is only term allowed. D = 0
Finite square well (2)
Match value and derivative of
wavefunction at region boundaries:
x  a x  a
Match ψ:
− a − a
A cos ka −B sin ka=C e A coska B sin  ka=D e
Match dψ/dx:
− a − a
kA sin ka kB coska = C e −kAsinkakB coska=− D e

Add and subtract:


− a
2A cos ka =C D e 2 B cos ka=C− De
− a

− a − a
2Bsinka= D−C e 2 Asin ka=C De

2222 Quantum Physics 2006-7 49


Finite square well (3)
Divide equations:
k tan  ka= (Unless A=0 and C=−D )
k cot  ka=−k (Unless B=0 and C= D )
Must be satisfied simultaneously:
Either: k tan ka= and B=0 , C= D (Even solution)
Or: k cot  ka=− and A=0 , C=−D (Odd solution)
Cannot be solved algebraically. Convenient form for graphical solution:
k &  are related by the requirement:
ℏ2 2 2
 k  =V 0
2m
2 2 2 2 2mV0
⇒ k  =k 0 Where:  k 0 =
ℏ2
k tan ka =   k 0 −k 
2 2
 Even solution 
 Odd solution  k cot  ka=−   k 0 2 −k 2 
⇒ Search for intersections of  y=k tan  ka or  y=k cot  ka  with a circle  y=±   k 0 2 −k 2 
2222 Quantum Physics 2006-7 50
Graphical solution for finite well
5
4
3
2 ktan(ka)
1
sqrt(k0^2-k^2)
0
-1 0 kcot(ka)
5 10
-2 -sqrt(k0^2-k^2)
-3
The radius of the circle
-4 corresponds to the
-5 depth of the well.

k0=3, a=1
2222 Quantum Physics 2006-7 51
Notes
• Penetration of particle into “forbidden” region where V>E
(particle cannot exist here classically)
• Number of bound states depends on depth of potential well,
but there is always at least one (even) state
• Potential is even function, wavefunctions may be even or odd
(we say they have even or odd parity)
• Limit as V0→∞:

k 0  ∞ , circle becomes very large.
n
⇒ Solutions at  ka=  (as for an infinite well.)
2

2222 Quantum Physics 2006-7 52


Example: the quantum well
Quantum well is a “sandwich” made of two different semiconductors in which the energy
of the electrons is different, and whose atomic spacings are so similar that they can be
grown together without an appreciable density of defects:

Material A
(e.g. AlGaAs) Material B (e.g. GaAs)

Electron potential
energy

Position

Now used in many electronic devices (some


transistors, diodes, solid-state lasers) Kroemer Esaki

2222 Quantum Physics 2006-7 53


Rae §9.1; B&M
§5.2, B&J §3.2 3.4 Particle Flux
In order to analyse problems involving ∞ ∞
scattering of free particles, need to understand i k x − t  2
normalization of free-particle plane-wave ∫ ∣A e ∣ dx= ∫ ∣A∣2 dx=∞
−∞ −∞
solutions.
Conclude that if we try to
normalize so that

∫ ∣∣ dx =1
2

−∞
will get A=0.

This problem is related to Uncertainty Principle:


Position completely undefined;
single particle can be anywhere
Momentum is completely defined from -∞ to ∞, so probability of
finding it in any finite region is
zero

2222 Quantum Physics 2006-7 54


Particle Flux (2)
More generally: what is rate of change of
probability that a particle exists in some region
(say, between x=a and x=b)?
a b x

[ ]
b b ∗
 ∗  

t a
 ∗
 dx=∫ 
t

t
dx
a
Use time-dependent Schrődinger equation:

[ ] [ ]
2 2 ∗ 2 2
 ℏ   ℏ  ∗
iℏ =− 2
V  ; iℏ =− V 
t 2m  x t 2m  x 2

{ [ ] [ ] }
b
1 ∗ ℏ2 2 1 ℏ 2 2
⇒∫

 − 2
V −  − 2
V  dx
a iℏ 2m  x iℏ 2m  x

[ ]
b
1 ℏ2 2
∗   2  ∗
= − ∫   x2 −  x 2 dx
i ℏ 2m a

[ ]
b 2
iℏ ∗   2  ∗
= ∫
2m a

x 2
−
x 2
dx

2222 Quantum Physics 2006-7 55


Particle Flux (3)
Integrate by parts:

[ ] { }
b ∗ b b
d iℏ ∗ d d iℏ d ∗ d  d  d ∗

dt a

  dx=
2m

dx
−
dx
− ∫
2m a dx dx

dx dx
dx
a

   
∗ ∗
iℏ d d ∣ − i ℏ  ∗ d  − d  ∣
= ∗ −
2m dx dx ∣ x=b 2m dx dx ∣ x=a

Flow of probability leaving at x = b Flow of probability entering at x = a

Interpretation:
Particle flux at position  x : Flux entering Flux leaving
at x=a
[ ]
∗ at x=b
iℏ ∗  
  x=−  −
2m x x

 or    x = Im  ∗
m

x [ ] a b x

Note: a wavefunction that is


real carries no current
Note: for a stationary state can use either ψ(x) or Ψ(x,t)
2222 Quantum Physics 2006-7 56
Particle
Sanity check: apply to free-particle
Flux (4)
plane wave.   x , t = A ei k x − t  ;  ∗ = A ∗ e−i  k x− t 

  
=ikA ei  k x− t  ; =−ikA ∗ e−i k x − t 
x x
iℏ
⇒ Flux   x =−
2m
[  A ∗ e−i k x − t  ikA ei k x − t  − A ei  k x− t  −ikA∗ e−i  k x − t   ]

iℏ iℏ 2
=− [2ikAA ∗ ]=− [2ik ∣A∣ ]
2m 2m
2
∣A∣ ℏ k
=
m
Makes sense:
# particles passing x per unit time = # particles per unit length × velocity
Wavefunction describes a “beam” of particles.
Two common normalising methods:
1) One particle per unit length. 2
We will use this method.
⇒∣A∣ =1
2
2)One particle per unit time. ∣A∣ ℏ k
⇒ =1
m

2222 Quantum Physics 2006-7 57


Rae §9.1; B&J §4.3
3.5 Potential Step
Consider a potential which rises suddenly V(x)
at x=0:

Case 1 V0
x
0
Boundary condition: particles only
Case 1: E<V0 (below step) incident from left

x<0 (But E > 0 ) x>0

Free particle S.E.: Same as Regions 1 & 3


2 2
ℏ 2 2  region 2 of ℏ   of finite well.
− =E  finite well. − 2
V 0 =E 
2m  x 2 2m  x
2 2
ℏ k 2 2 x − x ℏ 
⇒ = Ae ikx Be−ikx With: =E ⇒ =Ce  De With: =V 0 −E
2m 2m
Choose 1 particle per unit length  The  Ce x  term is not normalisable.
2222 Quantum Physics 2006-7 58
in incoming beam ⇒ A=1
Potential Step (2)
Continuity of ψ at x=0: 1 B= D

d
Continuity of at x  0 : i k 1−B =− D
dx

Solve for reflection and transmission:

i k × equation 1  equation 2:
2ik
2 i k =i k − D ⇒ D=
i k −

× equation 1  equation 2:


i k 
i k −i k  B=0 ⇒ B=−
−i k 

2222 Quantum Physics 2006-7 59


Transmission and reflection coefficients
ℏk
Incident particle flux from left=
m
What is the flux reflected back to the left? : Transmitted flux: = D e− x 
iℏ
[ ] ∣D∣2 [ e− x − e− x −e − x − e − x  ]

iℏ ∗   −
−  − = B e−i k x  2m
2m x x =0
i ℏ 2 ik x
=− ∣B∣ [ e −i k e −i k x −e i k x i k e i k x  ] i.e.: Probability of transmission T =0
2m
ℏk 2
=− ∣B∣
m

Hence the probability of reflection:


Reflected flux 2 i k −i k
R= =∣B∣ =BB ∗ = =1
Incident flux −i k i k

So all particles are reflected.

2222 Quantum Physics 2006-7 60


Potential Step (3)
Case 2: E>V0 (above step)

Solution for x>0 is now


ik ' x ℏ2 k 2
−i k ' x
 x= F e G e With: =E −V 0
2m
No incoming particles from right ⇒ G=0
Matching conditions:
 continuous. 1 B=F

continuous. i k 1− B=i k ' F
x
k −k '
i k × equation 1equation 2 ⇒ 2 i k =i  k k '  F ⇒ B=
k k '
k −k '
i k ' × equation 1equation 2 ⇒ i  k ' −k i k ' k  B=0 ⇒ B=
k k '
Transmission and reflection coefficients:
By our argument, reflection probability:
 k −k ' 2
2
R=∣B∣ = 2
 No longer 1 if k =k '
 k k ' 
2222 Quantum Physics 2006-7 61
Summary of transmission through potential
step
ℏ k ' ℏ k ' 4k2
2
Transmitted flux: =∣F∣ =
m m k k ' 2
Transmitted flux 4k k '
 Transmission probability: = =
Incident flux k k ' 2
Notes:
•Some penetration of particles into forbidden region even for energies
below step height (case 1, E<V0);
•No transmitted particle flux, 100% reflection (case 1, E<V0);
•Reflection probability does not fall to zero for energies above barrier (case
2, E>V0).
•Contrast classical expectations:
100% reflection for E<V0, with no penetration into barrier;
100% transmission for E>V0
2222 Quantum Physics 2006-7 62
Summary of transmission through potential
step (Cont.)

 k −k ' 2 4k k ' k 2 −2 k k 'k ' 2 4 k k '


Check RT = 2
 2
= 2
 k k '   k k '   kk ' 
=1 as required.

2222 Quantum Physics 2006-7 63


Rae §2.5; B&J §4.4; B&M §5.9

3.6 Rectangular Potential


V(x)
Barrier
I II III
Now consider a potential barrier of
finite thickness:
V0
x
0 a
Boundary condition: particles only
Assume 0 EV 0
incident from left

Region I: Region II: Region III:

 x= Ae i k x  B e−i k x =C e x D e− x  x=F e i k x G e−i k x

 With
ℏ2 k 2
2m
=E  With
ℏ 2 2
2m
=V 0 −E  (With k as in region I.)
Exclude G term as this term
Take A=1 (i.e. 1 particle per N.B: No reason to exclude either represents particles travelling
unit length.) solution since region II is finite. from the right.
2222 Quantum Physics 2006-7
⇒ G=0 64
Rectangular Barrier (2)
Match value and derivative of
wavefunction at region boundaries:

x0 x  a
Match ψ: 1 B=C D [1] C e a  D e− a = F e i k a [ 2]
Match dψ/dx: i k 1− B=C− D [3] C e a − D e− a =i k F e i k a [4]
Eliminate wavefunction in central
region:
Eliminate B :
i k [1][3]: 2 i k =i k C i k − D
a ik a
[2][4]: 2  C e =i k  F e
[2]−[4]: 2  D e− a =−i k  F e i k a
Eliminate C & D in the same way...
4 i k  e−i k a
To get F =
i k 2 e− a −−i k 2 e  a
2222 Quantum Physics 2006-7 65
Rectangular Barrier (3)
Transmission and reflection coefficients:
2
∣F∣ ∣F∣
2
16 k 2 2
Transmission probability: T= 2= =
∣A∣ 1 ∣i k 2 e− a−−i k 2 e a∣
2

2
Reflection probability: R=∣B∣ =1−T

For very thick or high barrier:  a ≫1 ⇒ e a ≫e− a

16 k 2  2 16 k 2  2 −2  a
T≃ = e
22
∣−i k  ∣ e 2 a −i k 2

16 E V 0 −E 
= e− a
V 02
Non-zero transmission (“tunnelling”) through classically
forbidden barrier region:
2222 Quantum Physics 2006-7 66
Examples of tunnelling
Tunnelling occurs in many situations in physics and astronomy:
1. Nuclear fusion (in stars 2. Alpha-decay
and fusion reactors)
V V
Coulomb
interaction
(repulsive)
Incident
particles

Internuclear Distance x of α-
Strong nuclear distance x particle from nucleus
force Initial α-particle
(attractive) energy

( Ze) 2 V
Barrier height ~ ~ MeV Distance x of
4 0 rnucleus electron from
? thermal energies (~keV) Work function surface
W
3. Field emission of Material
electrons from surfaces Vacuum 67
(e.g. in plasma displays)
Rae §2.6; B&M §5.5; B&J §4.7

3.7 Simple Harmonic Oscillator


Mass m Force F   kx
Example: particle on a
spring, Hooke’s law k
Angular frequency 0 =
restoring force with spring m
constant k: Potential energy V ( x )  1 kx 2  1 m0 2 x 2
2 2
x
Time-independent
Schrődinger equation: ℏ2 d 2  1 2
−  m  0 x =E 
2m d x 2
2
Problem: still a linear differential equation
but coefficients are not constant.

   
1 1
m 0 2 d dx d ℏ −
2 d
Simplify: change variable to y= x ⇒ = =
ℏ d y d y d x m 0 dx

ℏ m 0 d 2  1 2
Substitute: −  ℏ  0 y = E y
2m ℏ d y 2 2
2
d  2E
⇒ − y 0 =0 =
d y2 ℏ 0
2222 Quantum Physics 2006-7 68
Simple Harmonic Oscillator (2)
Asymptotic solution in the

 
limit of very large y: 2
n y
 y= y exp ±
2

Check:

 
   
2 2
d  n−1 y n1 y S.E. for large y
=n y exp ± ±y exp ±
dy 2 2 d2 2
− y =0
 
2
d  y2 dy 2
= [ nn−1 y ] exp ±
n−2  n  n2
⇒ ±2n−1 y  y
dy 2
2 is satisfied.

≃y
n2
 
exp ±
y2
2
2
= y  (For large y .) Only -ve part gives
normalisable solution.

 
2
y
Substitute:  y= H  y exp ±
2

2222 Quantum Physics 2006-7 69


Simple Harmonic Oscillator (2) (Cont.)
H is the correction function to find all solutions.

Equation for H:
d
dy
= [ H '  y− y H  y ] exp −
y2
2  
d2
 
2
y
[ H ' '  y −2 y H '  y − H  y y 2
H  y  ] exp −
d y2 2

Substitute in S.E:
2
 
[ H ' '  y −2 y H '  y y −1 H  y ] exp − y2
2
2
 − y  H  y exp −
y2
2  
= 0

⇒ H ' '  y−2 y H '  y−1 H  y=0

2222 Quantum Physics 2006-7 70


Simple Harmonic Oscillator (3)
Must solve this ODE by the power-series ∞
H  y= ∑ a p y
p
method (Frobenius method); this is done as
an example in 2246. p=0

We find:
•The series for H(y) must terminate in order to obtain a normalisable solution
•Can make this happen after n terms for either even or odd terms in series (but not both)
by choosing:
=2n1 (For some integer n .) ⇒
1
 
E =ℏ 0 n
2
If n is even, we must also choose a1 =0 , so all odd terms in the series vanish giving an even solution.
If n is odd, we must also chose a 0=0 , so all even terms vanish to give an odd solution.

Label resulting functions of H by Hn is known as the nth Hermite


the values of n that we choose. polynomial.

2222 Quantum Physics 2006-7 71


The Hermite polynomials
For reference, first few Hermite polynomials are:

H 0  y   1;
H1  y   2 y;
H 2  y   4 y 2  2;
H 3  y   8 y 3  12 y;
H 4  y   16 y 4  48 y 2  12.
NOTE:
Hn contains yn as the highest power.
Each H is either an odd or an even function, according to whether n
is even or odd.

2222 Quantum Physics 2006-7 72


Simple Harmonic Oscillator (4)
Transforming back to the original variable x, the
wavefunction becomes:

0  x=C 0 exp 
−m0 x 2
2ℏ  , 1  x=C 1 x exp 
−m 0 x 2
ℏ  , etc.

C 0 , C 1 , etc. are normalising constants that satisfy:∫∣ n  x∣ d x=1


2

Probability per unit length of finding the particle is:

2
∣C n  x∣ when the system is in state n .
2 2
E.g.:∣C 0  x∣ =∣C 0∣ exp

−m 0 x 2
ℏ 
2222 Quantum Physics 2006-7 73
Simple Harmonic Oscillator (4) (Cont.)
Compare classical result: probability of finding particle in a length δx is proportional to
the time δt spent in that region:

dx
p classical  x d x  dt 
v

For a classical particle with total energy E, velocity is given by

1
2
m v 2V  x = E
1
⇒ V=

2 E−V  x 
m
1
⇒ pclassical  =


 E −V  x E − 1 m  x 2
2 0

2222 Quantum Physics 2006-7 74


Notes
 
E= n
1
2
ℏ 0 in nth state.

• “Zero-point energy”: Ground state energy: E 0 = 1/ 2 ℏ 0


• “Quanta” of energy: Can only add or subtract multiples of ℏ 0
• Even and odd solutions
• Applies to any simple harmonic oscillator, including
– Molecular vibrations
– Vibrations in a solid (hence phonons)
– Electromagnetic field modes (hence photons), even though this field does not
obey exactly the same Schrődinger equation
• You will do another, more elegant, solution method (no series or
Hermite polynomials!) next year
• For high-energy states, probability density peaks at classical turning
points (correspondence principle)
2222 Quantum Physics 2006-7 75
4 Postulates of QM
This section puts quantum mechanics onto a more formal mathematical
footing by specifying those postulates of the theory which cannot be
derived from classical physics.
Main ingredients:
3. The wave function (to represent the state of the system);
4. Hermitian operators (to represent observable quantities);
5. A recipe for identifying the operator associated with a given
observable;
6. A description of the measurement process, and for predicting the
distribution of outcomes of a measurement;
7. A prescription for evolving the wavefunction in time (the time-
dependent Schrődinger equation)

2222 Quantum Physics 2006-7 76


4.1 The wave function
Postulate 4.1: There exists a wavefunction Ψ that is a continuous,
square-integrable, single-valued function of the coordinates of all the
particles and of time, and from which all possible predictions about the
physical properties of the system can be obtained.

Examples of the meaning of “The coordinates of all the particles”:

For a single particle moving in


one dimension: x
For a single particle moving in
three dimensions:
 x , y , z  r ,  , etc.

For two particles moving in three


dimensions:  x1 , y1 , z 1 , x2 , y2 , z 2  etc.
The modulus squared of Ψ for any value of the coordinates is the probability
density (per unit length, or volume) that the system is found with that particular
2222 Quantum Physics 2006-7 77
coordinate value (Born interpretation).
4.2 Observables and operators
Postulate 4.2.1: to each observable quantity is associated a linear, Hermitian
operator (LHO).

An operator L is linear if, and only if:  1 f 1 c 2 f 2 ]=c 1 L[


L[c  f 1 ]c 2 L[
 f 2]

Examples: which of the operators


For arbitrary functions f 1 & f 2
and constants c 1 & c 2 . 
defined by the following equations are
linear?
Note: the operators
Lˆ1[ f ] f 2 NOT linear as L1 [c1 f 1 c 2 f 2 ]=c1 f 1 c 2 f 2 2 involved may or may
not be differential
Lˆ [ f ]
2 xf Linear as L 2 [c 1 f 1 c 2 f 2 ]=c 1 x f 1c 2 x f 2 operators (i.e. may or
Lˆ3 [ f ] f NOT linear as L 3 [c 1 f 1 c 2 f 2 ]=  c 1 f 1 c 2 f 2
may not involve
differentiating the
df d f1 d f2 wavefunction).
Lˆ4 [ f ] Linear as 
L 4 [c 1 f 1 c 2 f 2 ]=c 1 c 2
dx dx dx
2222 Quantum Physics 2006-7 78
Hermitian operators
An operator O is Hermitian if and only if:

[ ]
∞ ∞ ∗

∫ f ∗  O g  dx= ∫ g ∗  O f  dx
−∞ −∞

for all functions f,g vanishing at infinity.

Compare the definition of a Hermitian matrix M:



M i j =[ M j i ]

Analogous if we identify a matrix element with an integral:


M ij ∫ f i ∗ O f j  dx
−∞

2222 Quantum Physics 2006-7


(see 3226 course for more 79
detail…)
Hermitian operators: examples
The operator x is Hermitian

[ ]
∞ ∞ ∞ ∗

∫ f  x g dx= ∫ g x f dx=
∗ ∗
∫ g ∗ x f dx (Since x is real.)
−∞ −∞ −∞

d
The operator is not Hermitian.

dx ∞ ∗
dg ∞ df
∫ f∗
dx= [ f g ]−∞ − ∫ g

dx [ f ∗ g ]=0 as f & g vanish at ∞
−∞ dx −∞ dx

[ ]
∞ ∗

=− ∫ g ∗ df
dx
dx ∴ Not Hermitian because of -ve sign. (This is 'Anti-Hermitian.')
−∞

d2
The operator 2
is Hermitian.
dx

[ ]
∞ 2 ∞ ∞ ∗
d g dg dg df
∫ f ∗

dx 2
dx = f ∗
dx
−∫
dx dx
dx
−∞ −∞ −∞

[ ]
∞ ∞ ∞
df ∗ d2 f ∗ 2
∗ d f
=− g ∫ 2 g dx= ∫ g 2 dx
dx −∞ −∞ dx −∞ dx
2222 Quantum Physics 2006-7 80
Eigenvectors and eigenfunctions
Postulate 4.2.2: the eigenvalues of the operator represent the possible
results of carrying out a measurement of the corresponding quantity.

Definition of an eigenvalue for a general linear operator:


  n=qn  n
Q n is an eigenfunction.
q n is an eigenvalue.
Compare definition of an eigenvalue of a matrix:
M v= n v n v n is an eigenvector.
 n is an eigenvalue.
Example: the time-independent Schrődinger
equation: H =E 
The energy in the T.I.S.E. is an eigenvalue of the Hamiltonian operator, and  is an eigenfunction.
Interpretation of Eigenfunction: It is a state of the system in which there is a definite value
of the quantity concerned.
2222 Quantum Physics 2006-7 81
Important fact: The eigenvalues of a Hermitian operator are real (like the
eigenvalues of a Hermitian matrix).
Proof: Let   m=q m  m
Q (Where Q is Hermitian.)

[ ][ ]
∞ ∞ ∗

Then ∫ m ∗  Q m  dx = ∫ m ∗  Q m  dx (Using f = g = m )
−∞ −∞

[ ]
∞ ∞

⇒ ∫ m ∗   m dx = qm ∫ ∣m∣ dx
Q
2
(Which is real.)
−∞ −∞

But ∫ ∣m∣2 dx is real. ∴ qm is real.


−∞
Postulate 4.2.3: immediately after making a measurement, the wavefunction
is identical to an eigenfunction of the operator corresponding to the
eigenvalue just obtained as the measurement result.

Start with wave-function  and then measure quantity Q ⇒ Obtain result qm ( one of the
eigenvalues of Q ) ⇒ Leave system with corresponding wave-function  m

Ensures that we get the same result if we immediately re-measure the


same quantity.
2222 Quantum Physics 2006-7 82
4.3 Identifying the operators
Postulate 4.3: the operators representing the position and
momentum of a particle are

x = x
i.e. x = x 
p x =−i ℏ
x
d
r =r
[
p =−i ℏ i

x
j

y
k

z]=−i ℏ ∇

i.e. p x =−i ℏ
(one dimension) dx (three dimensions)
Other operators may be obtained from the corresponding
classical quantities by making these replacements.
Examples:
The Hamiltonian (representing the total energy as a
function of the coordinates and momenta)
2
p −ℏ 2
−ℏ 2  2
H =
2
V  r = ∇ V  r  or V  x
2m 2m 2 m  x2
Angular momentum:
 r × p=−i ℏ r×∇
L= ⇒ L =−i ℏ r× ∇ 
2222 Quantum Physics 2006-7 83
Eigenfunctions of momentum
The momentum operator is Hermitian, as
required:

[   ] [   ]
 
∞ ∞ ∞ ∗
dg df Factor of −i makes p
∫ f ∗  p g  dx= −i ℏ ∫ f ∗
dx
dx =−i ℏ −∫ g ∗
dx
dx
Hermitian, even though
−∞ −∞ −∞

[ ] [ ]
∞ ∗ ∞ ∗
df isn't.
= −i ℏ ∫ g ∗ dx = ∫  p f  dx
g ∗
x
−∞ dx −∞

Its eigenfunctions are plane waves:

de i k x
ik x ik x
p x e =−i ℏ =−i ℏ i k e
dx
=ℏ k e i k x
e i k x is indeed an eigenfunction of p x with eigenvalue ℏ k
This corresponds to a state having a definite momentum ℏ k (in agreement with DeBroglie.)

2222 Quantum Physics 2006-7 84


Orthogonality of eigenfunctions
The eigenfunctions of a Hermitian operator belonging to
different eigenvalues are orthogonal.

If Q  n=q n  n ;   m=qm  m
Q with q n≠qm then ∫ n ∗ m dx=0
−∞

Proof:
Use Hermitian definition, taking : f = n , g =m

[ ]
∞ ∞ ∗

⇒ ∫ n∗  Q m  dx= ∫ m ∗  Q n  dx
−∞ −∞

[ ]
∞ ∗ ∞

RHS= q n ∫ m n dx =q n ∫ m  n ∗ dx

(Since q n is real.)
−∞ −∞

LHS=q m ∫  n ∗ m dx
−∞

We chose qm ≠q n ⇒ ∫ n∗ m dx=0


−∞

RHS=LHS=q m−qn  ∫  n ∗ m dx=0


−∞

2222 Quantum Physics 2006-7 85


Orthonormality of eigenfunctions
What if two eigenfunctions have the same eigenvalue? (In this
case the eigenvalue is said to be degenerate.)

Any linear combination of these eigenfunctions is also an


eigenfunction with the same eigenvalue:
  c1 1 c 2  2 =c1 Q 1 c 2 Q  2
Q
=c1 q 1 c 2 q 2
=q c 1  1c 2 2 
 Q 1 1 =q 1
Q 2 2 =q  2 
So we are free to choose as the eigenfunctions two linear
combinations that are orthogonal.
We can choose to have all the eigenfunctions orthogonal regardless of
whether the eigenvalues are the same or different.


If the eigenfunctions are all orthogonal and normalized,
they are said to be orthonormal. ∫ n∗ m dx= n m
−∞

2222 Quantum Physics 2006-7 86


Orthonormality of eigenfunctions: example
Consider the solutions of the time-independent Schrődinger equation
(energy eigenfunctions) for an infinite square well:  Eigenfunctions of 
H

{  
}
1 n x
cos , for odd n a

 n=  a
2a
For −a xa , otherwise 0 ∫∣n∣2 dx=1
1
a
sin  
n x
2a
, for even n
−a

We chose the constants so that normalization is correct:


Consider 2 different values n , m :
[1] 2 odd values:

   
a a
1 n x m x 1
∫ n m ∫a
 ∗
 dx= cos
2a
cos
2a
dx cos  cos = [ cos   −  cos     ]
2
−a −a

[    ]
a
1  nm x  n−m x
= ∫ cos cos dx
2a −a 2a 2a

[    ]
a
1 2a nm  x 2a  n−m x
= sin sin =0
2a  nm 2a n−m  2a −a

2222 Quantum Physics 2006-7 87


Orthonormality of eigenfunctions: example
(Cont.)
[2] 2 even values:
1
Similarly =0 sin  sin =
2
[ cos −−cos  ]

[3] Even & odd values:


a

∫ n∗ m dx=0 by symmetry.


−a
a

⇒ ∫ n∗ m dx=n m
−a

2222 Quantum Physics 2006-7 88


Complete sets of functions
The eigenfunctions φn of a Hermitian operator  x = ∑ an n  x 
form a complete set, meaning that any other n
function satisfying the same boundary conditions Where:  is an arbitrary function,
can be expanded as an are numbers, and n are eigenfunctions.

If the eigenfunctions are chosen to be orthonormal, the


coefficients an can be determined as follows:
∞ ∞

Consider: ∫ m ∗ dx=∑ an ∫ m ∗ n dx
−∞ n −∞

=∑ a n m n
n
=a m
⇒ Find number a m , multiply  by  m ∗ and integrate.

We will see the significance of such expansions when we


come to look at the measurement
2222 Quantum Physicsprocess.
2006-7 89
Normalization and expansions in complete
sets
The condition for normalizing the wavefunction is now
∞ ∞ ∞

[∑ am m  x ] [ ∑ a n n  X ] dx=∑ ∑ am

1= ∫ ∣ x∣ dx= ∫ an ∫  m ∗  x n ∗  x dx
2 ∗

−∞ −∞ m n m n −∞

∫ m ∗  x n∗  x dx=m n= {10 m=n


otherwise }
−∞

If the eigenfunctions φn are orthonormal, this becomes


1= ∑ ∑ am a n  m n= ∑ ∣a n∣
∗ 2

m n n

Natural interpretation: the probability of finding the system in2 the state
φn(x) (as opposed to any of the other eigenfunctions) is an

2222 Quantum Physics 2006-7 90


Expansion in complete sets: example
Consider an infinite square well, with a particle confined to −a xa
Hamiltonian has eigenfunctions:

{{ } }
1 n x
cos  ( n odd)
a 2a
For −a xa
n  x= 1 n x
sin   ( n even)
a 2a
0 otherwise
So any function f  x  satisfying some boundary conditions
(i.e zero outside well) can be represented as:
1 nx 1 n x
f  x= ∑ a n cos  ∑ sin  
n odd a 2 a n even  a 2 a
Eigenfunctions are orthonormal (see previous working)

{ }
a
1 n x
∫ a cos
2 a
 f  x dx ( n odd)
⇒ an = −a
a
1 n x
∫ a sin
2a
 f  x dx ( n even)
−a

This is a Fourier series representation of f  x


2222 Quantum Physics 2006-7 91
4.4 Eigenfunctions and measurement
Postulate 4.4: suppose a measurement of the quantity Q is made, and that
the (normalized) wavefunction can be expanded in terms of the
(normalized) eigenfunctions φn of the corresponding operator as

  x =∑ an  n  x 
n

Then the probability of obtaining the corresponding eigenvalue qn as the


2
measurement result is an

Corollary: if a system is definitely in eigenstate φn, the result measuring


Q is definitely the corresponding eigenvalue qn.

What is the meaning of these “probabilities” in


discussing the properties of a single system?

Still a matter for debate, but usual interpretation is that


the probability of a particular result determines the
frequency of occurrence of that result in measurements
2222 Quantum Physics 2006-7 92
on an ensemble of similar systems.
Commutators
In general operators do not commute: that is to say, the order in
which we allow operators to act on functions matters:
Q R ≠ R Q  (In general)
For example, for position and momentum operators:
(In x direction)
d
x  p = x X −i ℏ ⇒ x p − p x =i ℏ 
dx
p  x =−i ℏ

x
 x =−i ℏ  x 
d
dx  ⇒ [ x , p ]= x p− p x =i ℏ

We define the commutator as the


difference between the two orderings: [ Q , R ]=Q R−
 RQ

So, for position and momentum: Two operators commute if, and only if, their
commutator is zero. Note: The commutator
[ x , p x ]=i ℏ of any operator with itself is zero!
But: [ x , p y ]=0=[ x , p z ] [ Q , Q ]=Q Q −Q Q = 0
2222 Quantum Physics 2006-7 93
Compatible operators
Two observables are compatible if their operators share the same
eigenfunctions (but not necessarily the same eigenvalues).
Consequence: two compatible observables can have precisely-defined
values simultaneously.
Measure Re-measure Q,
Wavefunction of Measure observable R, Wavefunction of
observable Q, definitely
system is definitely obtain result system is still
obtain result qm obtain result qm
corresponding rm (the corresponding corresponding
(an eigenvalue of once again
eigenfunction φm eigenvalue of R) eigenfunction φm
Q)

Compatible operators commute with one another: Expansion in terms of


Consider a general wavefunction   x =∑ a n  n  x  joint eigenfunctions of
n
both operators

[
Q  R  = Q ∑ an r n  n  x 
n ] ( r n eigenvalues of R )
=∑ an qn r n n  x  ( Using linearity of operator)
n
And: R  Q =∑ an r n q n  n  x 
 
⇒ Q R = R Q  for any 
n
⇒ [ Q , R ]=0
Uncertainty of measurement is proportional to commutator.
Commuator here is zero, so no uncertainty.
Can also show the converse: any two 2222 Quantum Physics 2006-7 94
commuting operators are compatible.
Example: measurement of position
Eigenfunctions of position operator x would have to be states of definite position.
'Dirac  - functions'.
For now, consider approximate eigenstates: Suppose we have a series of detectors (in one dimension)
each sensitive to the presence of a particle in length 
Detectors are n , n1, n 2 etc. Starting at x n , x n 1 etc. on the x - axis.

{ }
1

2 th
Corresponding eigenfunctions:  n  x =  when x in n region
0 otherwise n  x 
∞ x 
2 n
1 1
Check normalisation: ∫ ∣ n∣ dx = ∫ dx = = 1
∞− x 
n  1

2


Check orthogonality: ∫  n m dx =0

(If n≠ m )
− ∞
Since at least one of  n and  m is always zero.
xn xn  1

2222 Quantum Physics 2006-7 95


Example: measurement of position (2)
Now take system with general wavefunction as   x =∑ an  n  x  (Becomes extinct as   0 )
n

 
∞ ∞ 1 1
− −
2 2
Where: an= ∫  n  x    x  dx = ∫   x  dx 

≃    xn 
− ∞ − ∞
2 2 2
Probability that n detector 'fires' =∣a n∣ =∣  x n ∣ ⇒ Consistent with Born interpretation that ∣∣
th

is a probability density. (Probability per unit length.)

Schematically:

≃ Measure position

x x ` x

n th detector fires.
2222 Quantum Physics 2006-7 96
Expectation values
The average (mean) value of measurements of the quantity Q is therefore the sum of
the possible measurement results times the corresponding probabilities:

〈 Q 〉 =∑ ∣a n∣ q n 2

n
We can also write this as:

∞ ∞

∫  ∑ an n  x   ∑ a m Q m  x  dx

∫   x  Q  x  dx =

−∞ −∞ n m

 
∞ ∞

= ∑ ∑ q m an a m ∫  n m dx
∗ ∗   m=qm m
Since Q ∫ n ∗ m dx=m n
n m −∞ −∞

= ∑ q n∣an∣ 〈 Q 〉
2
=
n

2222 Quantum Physics 2006-7 97


4.5 Evolution of the system
Postulate 4.5: Between measurements (i.e. when it is not disturbed by
external influences) the wave-function evolves with time according to the
time-dependent Schrődinger equation.

   is the Hamiltonian operator.


iℏ =H  Where H
t
This is a linear, homogeneous differential equation, so the linear combination of any two
solutions is also a solution: the superposition principle.

 1   2 
iℏ = H 1 iℏ = H 2
t t


⇒ iℏ
t
 c 1  1 c 2  2 = c 1 H 1  c 2 H  2
= H  c 1  1 c 2  2 

(Since H is linear.)

2222 Quantum Physics 2006-7 98


Calculating time dependence using expansion
in energy eigenfunctions
Suppose the Hamiltonian is time-independent. In that case we know that solutions of the
time-dependent Schrődinger equation exist in the form:
 n  x , t =exp 

i En t
ℏ  n  x 
where the wavefunctions ψ(x) and the energy E correspond to one solution of
the time-independent Schrődinger equation:
H n = E n n
We know that all the functions ψn together form a complete set, so we can expand
  x , 0=∑ a n n  x 
n

Hence we can find the complete time dependence


(superposition principle):

  x , t =∑ a n   x ,t = ∑a e
 −
i En t
ℏ  x
n n
n n

2222 Quantum Physics 2006-7 99


Time-dependent behaviour: example
Suppose the state of a particle in an infinite square well at time t=0
is a `superposition’ of the n=1 and n=2 states
  x , 0=c1 1  xc 2 2  x

=c 1
1
     
a
cos
x
2a
Wave function at a subsequent time t
c 2
1
a
sin
x
a
i E1t i E2 t
− −
ℏ ℏ
  x , t=c1 1  x e c 2 2  x e

   } {  
i E1 t i  E 2−E 1  t
1 −
ℏ x x − ℏ
= e c1 cos c 2 sin e
a 2a a
Probability density

∣     ∣
i  E 2−E 1  t 2
2 1 x x −

∣ ∣  x ,t  = c 1 cos c 2 sin e
a 2a a

{  ∣∣       [ ]}
i  E 2 −E 1 t
1 2 x 2 x x x −
= ∣c 1∣ cos 2  c 2 sin 2 2 cos sin  c 1∗ c 2 e ℏ
a 2a a 2a a

Probability Probability Oscillates with an angular frequency:


distribution distribution E − E1
in state 1. in state 2. = 2
2222 Quantum Physics 2006-7 ℏ 100
Rate of change of expectation value
Consider the rate of change of the
expectation value of a quantity Q:
d 〈 Q 〉 d

= ∫ ∗  Q    dx
dt d t −∞

     
∞ ∞ ∞
d ∗  d 
Q ∗  d
=∫  Q   dx ∫   dx ∫  Q

dx
−∞ dt −∞ dt −∞ dt

〈 〉
∞ ∞
1 
= ∫  H    Q   dx d Q  1 ∫  ∗  Q

 H   dx
−i ℏ −∞ dt i ℏ −∞

[ ]
∞ ∞ ∗ ∞
∗ ∗
 are Hermitian ⇒
H and Q ∫  H    Q   dx= ∫  Q     dx = ∫  ∗ H  Q
H    dx
−∞ −∞ −∞

d 〈Q
〉
〈 〉

dQ 1
⇒ =  ∫   Q H − H Q   dx
∗  
dt dt i ℏ −∞

= 〈 〉

dQ
dt

1
iℏ
〈 [ Q , H ] 〉
Commutator: Comes from the
Comes from intrinsic time time dependence of the wave
dependence of operator. function.
2222 Quantum Physics 2006-7 101
Example 1: Conservation of probability
Rate of change of total probability that the
particle may be found at any point:
∞ ∞
 
t
∫ ∣∣
2
dx=
t
∫  ∗  1×  dx Total probability is the
−∞ −∞
“expectation value” of the
operator 1.
=
〈 〉 1
t
1
 〈 [ 1 , H ] 〉
iℏ

=0 since the commutator of 1 with any operator is zero: 〈 [ 1 , H ] 〉= H − H =0


⇒ ∫ ∣∣ dx=constant
2

−∞

Total probability conserved (related to existence of a well


defined probability flux – see §3.4)

2222 Quantum Physics 2006-7 102


Example 2: Conservation of energy
Consider the rate of change of the mean energy:

〈 〉 〈 〉

  
dH 1   d H
〈E〉 = ∫  ∗
  dx =
H  〈[ H , H ]〉 = 0
t  t −∞ dt iℏ dt

dH
If Hamiltonian is constant in time, i.e. if =0
dt

d 〈E 〉 Conservation of energy provided


⇒ =0
dt Hamiltonian is constant in time.

Even although the energy of a system may be uncertain (in the sense that measurements of the
energy made on many copies of the system may be give different results) the average energy
is always conserved with time.

2222 Quantum Physics 2006-7 103


Reading: Rae Chapter 5;
B&J§§6.1,6.3; B&M§§6.2-6.5

5.1 Angular momentum operators


Angular momentum is a very important quantity in three-
dimensional problems involving a central force (one that is Central
always directed towards or away from a central point). In that point
case it is classically a conserved quantity: r

dL d F
=  R× p  = ṙ× pr × ṗ
dt dt The origin of r is the same central

m
p
 
= × p   r× F  =0
point towards/away from which
the force is directed.

We can write down a quantum-mechanical operator for it by applying our usual rules:

∣ ∣
i j k
 r × p=r×  −i ℏ ∇ = x y z
L=
d d d
−i ℏ −i ℏ −i ℏ
dx dy dz

2222 Quantum Physics 2006-7 104


5.1 Angular momentum operators (cont.)
Individual components:


L x = y p z − z p y =−i ℏ y

z
−z

y 

L y = z p x − x p z =−i ℏ z

x
−x

z 

L z= x p y − y p x =−i ℏ x

y
−y

x 
2222 Quantum Physics 2006-7 105
5.2 Commutation relations***
Remember:
The different components of angular momentum do
[ x, p x ]  i h
not commute with one another.
[ y, p y ]  ih
[ L x , L y ]= L x L y − L y L x [ z , pz ]  i h
= y p z− z p y  z p x − x p z  − z p x − x p z  y p z − z p y 
= y p z z p x − z p x y p z  − y p z x p z  x p z y p z − z p y z p x  z p x z p y  z p y x p z − x p z z p y 
= y p z z p x − z p x y p z    z p y x p z− x p z z p y 
=−[ z , p z ] y p x  [ z , p z ] x p y
=i ℏ  x p y − y p x 
=i ℏ L z
By similar arguments get the cyclic permutations:

[ L x , L y ]=i ℏ L z [ L y , L z ]=i ℏ L x [ L z , L x ]=i ℏ L y

2222 Quantum Physics 2006-7 106


Commutation relations (2)
The different components of L do not commute with one another, but they do commute with
the (squared) magnitude of the angular momentum vector:
Note a useful formula:
L = L x  L y  L z
2 2 2 2
[ A2 , B ]= A2 B− A B A A B A− B A2
Consider: [ L 2 , L z ] =[ L x 2 , L z ] [ L y 2 , L z ]  [ L z , L z ] =a [ A , B ]  [ A , B ] A
and [ L x 2 , L z ] = L x [ L x , L z ] [ L x , L z ] L x
= L x −i ℏ L y  −i ℏ L y L x
=−i ℏ  L x L y  L y L x 
Similarly: [ L y , L z ]=i ℏ  L x L y  L y L x 
Important consequence: we cannot find
⇒ [ L , L z ]=[ L x 2 , L z ][ L y 2 , L z ] simultaneous eigenfunctions of all three
=0 components.
But we can find simultaneous eigenfunctions of one
component (conventionally the z component) and L2

2222 Quantum Physics 2006-7 107


5.3 Angular momentum in spherical polar
coordinates On this slide, hats refer
to unit vectors, not
Spherical polar coordinates are the natural coordinate system in which to
operators.
describe angular momentum. In these coordinates,
z
 1  1  r  reˆ r
∇ = er  e  e (see 2246)
 r r   r sin    θ
r y
So the full (vector) angular momentum operator

∣ ∣
can be written φ
e r e  e
x
L=−i ℏ r×∇ =−i ℏ r er ×∇ =−i ℏ r 0 0

 1  1  
 r r   r sin    k 2
z

L=−i ℏ − e
1 
sin   
 e


Note: Has no er component
  and doesn't depend on r .

To find z-component, note that unit vector k e



in z-direction satisfies
 e=−sin 
k⋅ 
⇒ 
L z = k⋅L=−i ℏ
 e=0
k⋅ 
2222 Quantum Physics 2006-7 108
L in spherical polar coordinates
2
On this slide, hats refer
L 2 = L⋅L=−ℏ 2  r ×∇ ⋅ r×∇  to unit vectors, not
operators.
=−ℏ 2 [ r 2 ∇ 2 −  r⋅∇  ∇⋅r  ]
2
=−ℏ r 2
r  [{  
1  2 
2

r
r
 r
 2
1
r sin  


sin 


  1
 2 2
2
r sin    2−r
 
 r  r
r
}  ]
   
2
2 1  2  1   1 
Since ∇ = 2 r  2 sin   2 2
r r  r r sin      r sin    2
2 2
and radial parts of r ∇ and  r⋅∇  ∇⋅r  cancel.

[   ]
2
2 2 1   1 
L =−ℏ sin   2
sin      sin    2

Depends only on angular behaviour of wavefunction. Closely


related to angular part of Laplacian (see 2246 and Section 6).

2222 Quantum Physics 2006-7


2 1  2 
∇ = 2
r r
r  −
L2
 r ℏ2 r 2
109
5.4 Eigenvalues and eigenfunctions
Look for simultaneous eigenfunctions of L2 and one
component of L (conventional to choose Lz)
Eigenvalues and eigenfunctions of L z :

−i ℏ = 

i 

⇒ = A e ℏ
Note: L z only depends on 
Physical boundary condition: wave-function
must be single-valued

2 =
Taking our eigenfunction we have:

 
I 2  i  i 2 i 2
ℏ ℏ ℏ ℏ
2 = A e =Ae e e =1 Quantization of angular
 momentum about z-axis (compare
⇒ =integer , =m ℏ where m is an integer. Bohr model)

= A e i m  2222 Quantum Physics 2006-7 110
Eigenvalues and eigenfunctions (2)
Now look for eigenfunctions of L2, in the form (ensures solutions remain eigenfunctions of Lz,
f  ,= =e i m   as we want)

Eigenvalue condition becomes

L2 [ e i m   ]= ℏ 2 e i m  Let eigenvalue = ℏ 2

[   ]
2
ei m     
⇒ −ℏ
2
sin   2  e
i m
 = ℏ
2 im
e 
sin      sin   2

Divide through by ℏ 2 e i m

⇒ −
1 
sin    
sin 


m2

 2 =
sin 

2222 Quantum Physics 2006-7 111


The Legendre equation
Make the substitution

=cos 

 d  
⇒ = =−sin  , and sin 2 =1−cos2 =1−2
  d   

⇒ Get:
d
d[ 2 d
1− 
d ][
 −
m2
1− 2
=0
]

This is exactly the Legendre equation, solved in


2246 using the Frobenius method.
2222 Quantum Physics 2006-7 112
Legendre polynomials and associated
Legendre functions
In order for solutions to exist that remain finite at μ=±1 (i.e. at θ=0
and θ=π) we require that the eigenvalue satisfies

=l l1 , where l=0,1, 2, 3, 4, ... (like SHO, where we found restrictions on
energy eigenvalue in order to produce
normalizable solutions)

The finite solutions are then the associated Legendre functions, which can be
written in terms of the Legendre polynomials:

 
∣∣ ∣∣
m 2 2 d
P =1− 
l P l 
d
where m is an integer constrained to lie between –l and +l.
Legendre polynomials:
P 0 =1 P 1 = ¿
1 1
P 2 = 3  2 −1 P 3 = 5 3−3 
2 2
etc.
2222 Quantum Physics 2006-7 113
Spherical harmonics
The full eigenfunctions can also be written as spherical harmonics:
Y ml  ,=c l m P ml cos  ei m

 c l m=−1m
 l−m! 2 l1
 lm! 4  
Because they are eigenfunctions of Hermitian operators with different eigenvalues, they are
automatically orthogonal when integrated over all angles (i.e. over the surface of the unit
sphere). The constants C are conventionally defined so the spherical harmonics obey the
following important normalization condition:
 2

] Y l '  ,= l ,l ' m , m' = 10 { if l=l ' and m=m'


}
∗ m'
∫ sin  d  ∫ [ l
d  Y m
 ,
otherwise
0 0

First few examples (see also 2246): Y00 ( ,  ) 


1
4
Remember 3 3 ( x  iy )
Y11 ( ,  )   sin  exp(i )  
x  r sin  cos  8 8 r
3 3 z
y  r sin  sin  Y10 ( ,  )  cos  
4 4 r
z  r cos  3 3 ( x  iy )
114
Y ( ,  ) 
1
sin  exp(i ) 
8 8
1
r
Shapes of the spherical harmonics
Y 0 Y11 Y10
z 0

l 1
x Re[Y11 ]
m0
l0 3
Yl  m
cos 
m0 l 1 4
1 m 1 Imaginary
Yl m 
4 3
Yl m   sin  exp(i )
8
Real
To read plots: distance from origin corresponds to magnitude (modulus) of
plotted quantity; colour corresponds to phase (argument).
115
(Images from http://odin.math.nau.edu/~jws/dpgraph/Yellm.html)
Shapes of spherical harmonics (2)
Y22 Y21 Y20
z
y

x Re[Y 2
2
] Re[Y21 ]
l2
m0
5
Yl m  (3cos 2   1)
l2 l2 16
m2 m 1 Imaginary
15 15
Yl m  sin 2  exp(2i ) Yl m   sin  cos  exp(i )
32 8
Real
To read plots: distance from origin corresponds to magnitude (modulus) of
plotted quantity; colour corresponds to phase (argument).
116
(Images from http://odin.math.nau.edu/~jws/dpgraph/Yellm.html)
5.5 The vector model for angular
momentum***
To summarize:

Eigenvalues of Lˆ2 are l (l  1) h 2 , with l  0,1, 2,K

l is known as the principal angular momentum quantum number: determines


the magnitude of the angular momentum

Eigenvalues of Lˆz are mh, with m  l ,K , 1, 0,1,K  l

m is known as the magnetic quantum number: determines the component of


angular momentum along a chosen axis (the z-axis)

These states do not correspond to well-defined values of Lx and Ly, since these operators
do not commute with Lz.

Semiclassical picture: each solution corresponds to a cone of angular momentum


vectors, all with the same magnitude and the same z-component.
2222 Quantum Physics 2006-7 117
The vector model (2)
Lz

Example: l=2 m=2 ⇒ L z =2 ℏ


⇒ Eigenvalue of L is l l 1 ℏ =6 ℏ
2 2 2

L Ly
Magnitude of angular momentum is
m=1 ⇒ L z =ℏ

 l l1 ℏ= 6 ℏ m=0 ⇒ L z =0


m=−1 ⇒ L z =−ℏ
Component of angular momentum in Lx
z direction can be
m=−2 ⇒ L z =−2 ℏ
2ℏ
ℏ Reason for restriction on values of m is
0 so L z does not exceed total angular
−ℏ momentum available.
−2 ℏ
2222 Quantum Physics 2006-7 118
Reading: Rae §3.2, B&J §7.4; B&M §5.11

6.1 The three-dimensional square well


z
Consider a particle which is free to move in three y
dimensions everywhere within a cubic box, which
extends from –a to +a in each direction. The particle
is prevented from leaving the box by infinitely high
potential barriers. x

Time-independent Schrödinger equation


within the box is free-particle like: V(x)


− ∇ 2  x , y , z =E  x , y , z  V  V 0 V 
2m

Separation of variables: take


x, or y,
-a a
 ( x, y, z )  X ( x)Y ( y ) Z ( z ) or z

with boundary conditions X ±a=Y ±a=Z ±a=0


2222 Quantum Physics 2006-7 119
Three-dimensional square well (2)
Substitute in Schrödinger equation:

{ }
2
ℏ d2 X d2Y d2 Z
− YZ 2
Z X 2
X Y 2
=E X Y Z
2m dx dy dz
Divide by XYZ:
ℏ2 d 2 X
− =Ex X
2 m d x2

{
ℏ2 1 d 2 X 1 d 2 Y 1 d 2 Z
 
2 m X d x 2 Y d y2 Z d z2
=E
} −
2
ℏ d2Y
2 m d y2
=EyY
2
ℏ d2 Z
With E= E x  E y  E z − =E z Z
2 m d z2

Three effective one-dimensional


Schrödinge equations.

2222 Quantum Physics 2006-7 120


Three-dimensional square well (3)
Wavefunctions and energy eigenvalues known from
solution to one-dimensional square well (see §3.2).
n x 2 2 ℏ 2
E x= 2
with n x =1, 2, 3, 4, ... Similarly for E y and E z
8ma
Note: We have 3 seperate quantum numbers n x , n y and nz
Total energy is
E= E x  E y  E z
2 ℏ 2
= 2 n x2n y 2n z 2 
8ma

This is an example of the power of separation of variables in a 3D problem. Now


we will use the same technique for the hydrogen atom.

2222 Quantum Physics 2006-7 121


Reading: Rae §§3.3-3.4, B&M Chapter 7, B&J §7.2 and §7.5

6.2 The Hamiltonian for a hydrogenic


atom***
For a hydrogenic atom or ion having nuclear charge +Ze and a
-e
single electron, the Hamiltonian is r
2
ℏ Z e 2
Note spherical symmetry –
H =− 2
∇ − potential depends only on r
2 me 4  0 r +Ze
Note: for greater accuracy we should use the reduced mass
corresponding to the relative motion of the electron and the nucleus
(since nucleus does not remain precisely fixed – see 1B2x):
me mN
= me = electron mass , m N = nuclear mass
me m N
The natural coordinate system to use is spherical polar coordinates. In this case the
Laplacian operator becomes (see 2246):

     
2
2 1  2  1   1  1  2  L2
∇ = 2 r  2 sin   2 2 = 2 r − 2 2
r  r  r r sin      r sin    2
r  r  r ℏ r
This means that the angular momentum about any axis, and also the d 〈 
L 2

total angular momentum, are conserved quantities: they commute [ H , L 2 ]=0 ⇔ =0
dt
with the Hamiltonian, and can have well-defined values in the
d 〈 L z 〉
energy eigenfunctions of the system. [ H , L z ]=0 ⇔ d t =0
2222 Quantum Physics 2006-7 122
6.3 Separating the variables
Write the time-independent Schrődinger equation as:
H r , , = E  r , ,

⇒ −
ℏ2 1 d 2 d 
2 m r2 d r
r  
dr

L 2
2mr
2
−
4
Z e2
  0 r
= E 

Now look for solutions in the form


 ( r ,  ,  )  R (r )Y ( ,  )
Substituting into the Schrődinger equation:
ℏ2
[  ]
2
1 d 2dR R 2 Ze
− Y  ,  2 r  2
L Y  ,− R Y =E R Y
2m r dr dr 2mr 4  0 r
1 L 2
{   }
2
2 ℏ 1 d 2 d R Z e2 r
⇒ Y  ,= E r − − r −
Y 2m 2m R d r dr 4  0
LHS depends only on  and , RHS depends only on r .
Both sides must equal some constant: 
2222 Quantum Physics 2006-7 123
The angular equation
We recognise that the angular equation is simply the eigenvalue condition for the
total angular momentum operator L2:

1 L 2
Y = ⇒ L 2 Y = 2 m e  Y
Y 2 me

Y is an eigenfunction of operator L 2

This means we already know the corresponding eigenvalues


and eigenfunctions (see §5):

We know L 2 Y ml   ,  = l  l  1 ℏ 2 Y ml   ,  

l  l 1  ℏ 2
⇒ Angular function Y is a spherical harmonic and =
2 me

Note: all2222
thisQuantum
would Physics
work for any spherically-symmetric
2006-7 124
potential V(r), not just for the Coulomb potential.
6.4 Solving the radial equation
Now the radial part of the Schrődinger equation becomes:


ℏ2

2 me R d r 
1 d 2dR
r
dr  2
 = E r 
Z e2
4  0
r


ℏ2

2 me r 2 d r
1 d 2d R
r
dr [

l  l 1  ℏ 2 Z e 2
2 me r
2 −
4  ]0 r
R= E R

Note that this depends on l, but not on m: it therefore involves the


magnitude of the angular momentum, but not its orientation.
Define a new unknown function χ by:  r 
R  r =
r
d R 1 d  dR d
⇒ = − 2 ⇒ r2 = r −
dr r dr r dr dr


d 2d R
dr
r
dr  
=r
d2 d  d 
dr
2
d r

dr

[ ]
2
ℏ2 d  l  l 1  ℏ 2 Z e 2
∴ above becomes: − 2 2 − = E 
2m d r 2 me r 4  0 r
2222 Quantum Physics 2006-7 125
The effective potential
This corresponds to one-dimensional motion with the effective potential

−Z e 2 l  l 1  ℏ 2
V eff  r  =  V(r)
4  0 r 2 m e r 2 l l1 ℏ
2

First term: 2 me r 2
Coulomb attraction:
2
dV −Z e
Force F =− =
d r 4  0 r 2
r
Second term:

'Centrifugal' repulsion: −Z e
2
2 2
d V l  l 1  ℏ 2 L L 4 0 r
Force F = − = = = = Rotational kinetic energy.
dr 2 me r 2 2 me r 2 2I

 me V r 2 m e V 2
Classically = 3
=
me r r 2222 Quantum Physics 2006-7 126
Atomic units***
Atomic units: there are a lot of physical constants in these expressions. It makes atomic
problems much more straightforward to adopt a system of units in which as many as
possible of these constants are one. In atomic units we set:
Planck constant ℏ = 1(Dimensions [ M L 2 T 1 ])

Electron mass me = 1 (Dimensions [ M ])


e2 3 2
Constant appearing in Coulomb's law = 1 (Dimensions [ M L T ])

4  0
It follows that:
Unit of length =
 
4  0 ℏ 2
e 2
m e
= 5.29177 × 10 11−
m = Bohr radius, a0

 
2
e 2
me 18
Unit of energy = = 4.35974 ×10 J = 27.21159 eV = Hartree, E h

4  0 ℏ 2

In this unit system, the radial equation becomes

[ ]
2
1 d  l  l 1  z
−  2 −  = E
2 d r2 2r r
2222 Quantum Physics 2006-7 127
Solution near the nucleus (small r)
For small values of r the second derivative and
centrifugal terms dominate over the others.
2
1 d  l  l 1 
−  2 = 0
2 d r2 2r
Try a solution to the differential equation in this limit as
  rk
k− 2 l  l 1  k 2
' ' = k  k −1  r and l l 1 r

⇒ 2 =   
r
So: − k  k −1l  l  1=0
⇒ k 2 − k − l  l 1= 0
⇒ [ k  l ][ k − l 1 ]=0 ⇒ k =−l or k =l 1
We want a solution such that R(r) remains finite as r→0,
so take
k = l 1 ⇒   rl 1
  
as r  0 R  rl

2222 Quantum Physics 2006-7 128


Asymptotic solution (large r)
Now consider the radial equation at very large distances from the nucleus, when both terms
in the effective potential can be neglected. We are looking for bound states of the atom,
where the electron does not have enough energy to escape to infinity:
2
Put: E =−
2
2 2 2
1d   d  2
− =−  ⇒ = −  
2 d r2 2 dr 2

x 2 2
Solutions: Try =e ⇒  = ⇒  =±
r r
Take  0 , general solution. = A e  B e

A e term is not normalisable ∴ A= 0


r

Inspired by this, let us rewrite the solution in terms of yet another unknown function,
F(r): − r
r  = F r  e

2222 Quantum Physics 2006-7 129


Differential equation for F
Can obtain a corresponding differential equation for F:
d  d F r
= e − F e  r
− −

dr dr
2
d  d2 F r d F r 2 r
e 2  e  F e
− − −
= − 
d r2 d r2 dr
Substituting in SE and cancelling factors of E  r gives:


1 d2 F
2 d r2

dF 1 2
dr 2
−  F
[ l  l 1  Z
2r
2 −
]
r
1 2
F =−  F ⇒
2
d2 F
dr
2 −
l  l 1 
r
2 F = 2 
dF Z
d r
−2
r
F

This equation is solved in 2246, using the F r  = r k ∑ a p r p = ∑ a p r p  k

Frobenius (power-series) method. p p

The indicial equation gives k  l or  l  1

regular solution behaves like F : r l 1 for small r.


2222 Quantum Physics 2006-7 130
Properties of the series solution
If the full series found in 2246 is allowed to continue up to an arbitrarily large number of
terms, the overall solution behaves like

F  r ≈ e 2  r (not normalizable)
⇒   r ≈ e 2  r e − r
= e r

Hence the series must terminate after a finite number of terms. This happens only if

Z
 n where n is an integer  l : n  l  1, l  2K

So the energy is


2
1 2 1 Z
E =−  =−
2 2 n

Note that once we have chosen n, the energy is independent of both m (a feature of all
spherically symmetric systems, and hence of all atoms) and l (a special feature of the
Coulomb potential, and hence just of hydrogenic atoms).
n is known as the principal quantum
2222number. It defines
Quantum Physics 2006-7 the “shell structure” of the 131
atom.
6.5 The hydrogen energy spectrum and wavefunctions***

 
2
Z
Each solution of the time-independent Schrődinger equation is E in units E
2 h
defined by the three quantum numbers n,l,m 1 0

16 1
− all
For each value of n=1,2,… we have a definite 9 all x7
energy: −
1 x5
4
Z2
En   2 (in atomic units) all
2n x3
For each value of n, we can have n possible values of
the total angular momentum quantum number l:

l=0,1,2,…,n-1 -1
l=0 l=1 l=2 l=3
For each value of l and n we can have 2l+1 values of
the magnetic quantum number m: Traditional nomenclature:
l=0: s states (from “sharp” spectral lines)
m  l , (l  1),K 0,K (l  1), l l=1: p states (“principal”)
l=2: d states (“diffuse”)
The total number of states (statistical weight) associated l=3: f states (“fine”)
with a given energy En is therefore …and so on alphabetically (g,h,i… etc)
n− 1
2222 Quantum Physics 2006-7 2p ⇒ n = 2 , l =1 132
∑  2 l 1 = n
2
l =0 3s ⇒ n =3 , l = 0
The radial wavefunctions
Radial wavefunctions Rnl depend on principal Full wavefunctions are:
quantum number n and angular momentum quantum n l m  r , , = Rn l  r  Y ml   , 
number l (but not on m)
Normalization chosen so that:
 
3 −Z r
Z 2 a0
R10  r =2 e ∞  2 2
a0 ∫ r R n l  r  dr ∫ sin    d  ∫ d  ∣Y l   ,  ∣ =1
2 2 m

   
3 −Z r 0 0 0
1 Z 2 Z r 2a  2
R 21 r = e 0
2
3 2 a0 a0 But: ∫ sin   d  ∫ d  ∣Y lm   , ∣ =1 for spherical harmonics.
0 0

  
3 −Z r
Z Zr ∞
2 2 a0 2 2
R 20 r =2 1− e ⇒ ∫ r R n l  r  dr =1
2 a0 2 a0 0

    
3 2 −Z r
4 Z 2 Zr 3a Note:
R32 r = e 0

27 10 3 a 0 a0 Probability of finding electron between radius r and r+dr is:

    
3 −Z r

R31  r =

4 2 Z 2
1−
Zr Zr
e
3 a0
r 2 Rn l 2  r =n l  r  since: n l = r R n l  r 
9 3 a0 6 a0 a0

  
3 2 2 −Z r
Z 2 2Z r 2Z r 3a i.e. r 2 R n l  r  is the probability per unit length of finding the
R30  r =2 1−  e 0

3 a0 3 a 0 27 a 02 electron at a radius r . NOT the probability per unit volume of


2
finding the particle at a given point in space: ∣n l m∣
Only s states (l=0) are finite at the
origin. 2222 Quantum Physics 2006-7 133

Radial functions have (n-l-1) zeros.


Comparison with Bohr model***
Bohr model Quantum mechanics

Angular momentum (about any axis) Angular momentum (about any axis) shown
assumed to be quantized in units of to be quantized in units of Planck’s
Planck’s constant: constant:

Lz  nh, n  1, 2,3,K Lz  mh, m  l ,K , l


Electron otherwise moves according Electron wavefunction spread over all
to classical mechanics and has a radii. Can show that the quantum
single well-defined orbit with radius mechanical expectation value of the
quantity 1/r satisfies
n 2 a0 1 Z
rn  , a0  Bohr radius  , a0  Bohr radius
Z r n 2 a0
Energy quantized and determined Energy quantized, but is determined
solely by angular momentum: solely by principal quantum number,
Z2 not by angular momentum:
En   2 Eh , Eh  Hartree
2n Z2
En   2 Eh , Eh  Hartree
2222 Quantum Physics 2006-7 2n 134
6.6 The remaining approximations
• This is still not an exact treatment of a real H atom, because we have
made several approximations.
– We have neglected the motion of the nucleus. To fix this we would need to
replace me by the reduced mass μ (see slide 1).
– We have used a non-relativistic treatment of the electron and in particular have
neglected its spin (see §7). Including these effects gives rise to
• “fine structure” (from the interaction of the electron’s orbital motion with its spin), and
• “hyperfine structure” (from the interaction of the electron’s spin with the spin of the
nucleus)
– We have neglected the fact that the electromagnetic field acting between the
nucleus and the electron is itself a quantum object. This leads to “quantum
electrodynamic” corrections, and in particular to a small “Lamb shift” of the
energy levels.

2222 Quantum Physics 2006-7 135


Reading: Rae Chapter 6; B&J §6.8, B&M Chapter 8 (all go further than 2B22)

7.1 Atoms in magnetic fields


Interaction of classically orbiting Orbit behaves like a current loop:
electron with magnetic field: ev

Loop current = ( -ve sign because chage =− e )
2 r
μ Magnetic moment = current × area

r
=
 
−e v

2r
 r 2=
−e
2 me
me v r =− B
L

v eℏ
Where  B = (The Bohr magneton)
2 me
In the presence of a magnetic field B, classical interaction energy is:
L
E =− ⋅B = B ⋅ B

Corresponding quantum mechanical expression (to a good approximation) involves the


angular momentum operator:
B
Contribution to Hamiltonian involving B =  H =  L ⋅ B

2222 Quantum Physics 2006-7 136
Splitting of atomic energy levels
Suppose field is in the z direction. The Hamiltonian operator is
 B
Hˆ  Hˆ 0  B z Lˆz
h
We chose energy eigenfunctions of the original atom that are eigenfunctions of Lz so these
same states are also eigenfunctions of the new H.

Hˆ 0 m  E0 m ;
  B B z 
⇒ H m = H 0 m  L z m
Lˆ   mh .
z m m ℏ
=  E 0m B B z  m

New
eigenvalue.

2222 Quantum Physics 2006-7 137


Splitting of atomic energy levels (2)

B0 B 0

m= l E = E 0 l  B B z
m= l −1 
E=E0

(2l+1) states with same


energy: m=-l,…+l m=−l E = E 0− l  B B z

(Hence the name


“magnetic quantum
number” for m.)
Predictions: should always get an odd number of
levels. An s state (such as the ground state of
2222 Quantum Physics 2006-7 138
hydrogen, n=1, l=0, m=0) should not be split.
7.2 The Stern-Gerlach experiment***
Produce a beam of atoms with a single electron in an s state (e.g.
hydrogen, sodium)

Study deflection of atoms in inhomogeneous magnetic field. Force on


atoms is
F =∇  ⋅ B 
N ⇒ if  is aligned with B , the atom is pushed towards high fields.

If  is anti-aligned with B atom is pushed towards low fields.


Results show two groups of atoms, deflected in
S opposite directions, with magnetic moments

 = ± B

Consistent neither with classical physics (which would predict a


continuous distribution of μ) nor with our quantum mechanics so
far (which always predicts an odd number of groups, and just one
2222 Quantum Physics 2006-7 139
for an s state). Gerlach
7.3 The concept of spin***
Try to understand these results by analogy with what we know about the ordinary (“orbital”)
angular momentum: must be due to some additional source of angular momentum that does
not require motion of the electron. Known as “spin”.
Introduce new operators to represent spin, assumed to have same
commutation relations as ordinary angular momentum:

[ S x , S y ]= i ℏ S z
etc. Where  S x , S y , S z  are components of spin angular momentum.
Define: S 2= S x 2  S y 2  S z 2

Corresponding eigenfunctions and eigenvalues:


S 2 = s = s  s1 ℏ 2  s
ms ms
ms = magnetic spin quantity
m m
S z s = m s ℏ s
s s

(will see in Y3 that these equations can be derived


directly from the commutation relations)
Find S does not have to be an integer, but can be an
integer or half integer and m s can vary from  S to − S 140
in integer steps. Goudsmit Uhlenbeck Pauli
Spin quantum numbers for an electron
From the Stern-Gerlach experiment, we know that electron spin along a
given axis has two possible values.

So, choose 1 1
S= ⇒ m s=− s..........  s −±
2 2
But we also know from Stern-Gerlach that magnetic moments associated with
the two possibilities are  = ±  B
Spin angular momentum is
So, have  =± 2 ms  B
twice as “effective” at
General interaction with magnetic field: producing magnetic moment as
B orbital angular momentum.
H = H 0  B ⋅  L ⋅ g S 

g = 2 (Dirac's relativistic theory)
g = 2.00231930437 (Quantum Electrodynamics)
g = g -factor (Measures how effective this particular form of angular momentum is at
producing a magnetic moment.)
2222 Quantum Physics 2006-7 141
A complete set of quantum numbers
Hence the complete set of quantum numbers for the electron in the H atom is: n,l,m,s,ms.
Corresponding to a full wavefunction

n , l , m , s , m  r ,  ,  = R n , l  r  Y l   ,    S
m ms
s

1
i.e. 2 states, with m s=± for each one we found before.
2
Note that the spin functions χ do not depend on the electron coordinates r,θ,φ; they
represent a purely internal degree of freedom.

H atom in magnetic field, with spin included, and field in the z direction:
B B z
H = H 0 

 L z  g S z 
⇒ Change in energy is  B B z  m g m s  ≃  B B z  m 2 m s 
For ground state  l =0, m=0  , get E =± B B z as observed.

2222 Quantum Physics 2006-7 142


7.4 Combining different angular momenta
So, an electron in an atom has two sources of angular momentum: Eigenvalues of J = j  j 1 ℏ 2
Eigenvalue of J z = m j ℏ
•Orbital angular momentum (arising from its motion through the
atom) S

•Spin angular momentum (an internal property of its own). J  LS


To think about the total angular momentum produced by L
combining the two, use the vector model once again:
Lz
|L-S|
m j =3/ 2
Vector addition between orbital angular S
momentum L (of magnitude L) and spin S (of
magnitude S): produces a resulting angular Ly
momentum vector J: quantum mechanics says its m j =1/2
magnitude lies somewhere between |L-S| and
L+S.(in integer steps). L
m j =−1/ 2
For a single electron, corresponding `total
angular momentum’ quantum numbers are L+S Lx
j  l  12 , l  12 Determines length of resultant m j =−3/ 2
angular momentum vector
m j   j ,K ,  j Determines orientation 3 1 143
e.g. l =1 , j= ,
2 2
Example: the 1s and 2p states of hydrogen
The 1s state:
1 1
l =0 , s= ⇒ 1 possibility j =
2 2
1
Purely spin angular momentum. m j =±
2
The 2p state:

 n= 2  , l =1 , S = 1
2
Now 2 possibilities: j =
1
2
,
3
2
; m j= −  1
2
,
1
2
, −
3
2
, −
1
2
,
1
2
,
3
2 
⇒ Get a doublet of states adding 2 different values of m j , and a quartet of states
adding 4 different values of m j

1 3
Even in no B - field. j = and j = states have different energies.
2 2
(Fine structure effects.)
2222 Quantum Physics 2006-7 144
Combining angular momenta (2)
The same rules apply to combining other angular momenta, from whatever source.
For example for two electrons in an excited state of He atom, one in 1s state and one in 2p state
(defines what is called the 1s2p configuration in atomic spectroscopy):
l1  0; s1  12 ; l2  1; s2  1
2
First construct combined orbital angular momentum L of both electrons:
L must be between ∣L1− L 2∣=1 and L 1 L 2= 1
⇒ S =0 , 1
Then construct combined spin S of both electrons:
∣S 1 S 2∣=0 and S 1  S 2 =1
⇒ S =0 , 1
Hence there are two possible terms (combinations of L and S):
…and four levels (possible ways of combining L and S to get different total angular
momentum quantum numbers)
L= 1 , S =0 ⇒ J =1
L= 1 , S =1 ⇒ J =0 ,1 , 2
2222 Quantum Physics 2006-7 145
Term notation
Spectroscopists use a special notation to describe terms and levels:

2 S 1
LJ
•The first (upper) symbol is a number giving the number of spin states
corresponding to the total spin S of the electrons
•The second (main) symbol is a letter encoding the total orbital angular momentum
L of the electrons:
•S denotes L=0
•P denotes L=1
•D denotes L=2 (and so on);
•The final (lower) symbol gives the total angular momentum J obtained from
combining the two.

Example: terms and levels from previous page would be:


1
L= 1 , S =0 ⇒ P1 (Singlet)
3 3
L= 1 , S =1 ⇒ P0 , P1 , 3 P2

2222 Quantum Physics 2006-7 146


7.5 Wavepackets and the Uncertainty Principle
revisited (belongs in §4 – non-examinable)
Can think of the Uncertainty Principle as arising from the structure of wavepackets. Consider a
normalized wavefunction for a particle located somewhere near (but not exactly at) position x0

[ ]
2
1
x x 0 2  ( x)
  x = 2   
2

4 − −
exp
4 2
= standard
deviation.
[ ]
1 2
x x
Probability density: ∣  x ∣ = 2  
2 2

2 − − 0 
exp 2
2
x0 x

Can also write this as a Fourier transform (see 2246):


% (k )
2

1 ∫∞
  x =   k  e i k x dk
2  ∞ −

(expansion in eigenstates of momentum)

2222 Quantum Physics 2006-7 147


k
7.5 Wavepackets and the Uncertainty Principle revisited
(belongs in §4 – non-examinable) (cont.)

Compare   x =∑ an  n  x 

1 ∫∞
then   k =   x  e i k x dx −

2 ∞ −


Compare an = ∫   x    x  dx ∗

− ∞

ei k x ⇒ n  x  (Eigenfunction)


 k  (Expansion coefficient)

2222 Quantum Physics 2006-7 148


Fourier transform of a Gaussian

1
k
 = ∫  x exp [−i k x ] dx
 2  −∞

[ ]
1 ∞ 2
1 −  x− x 0 
=  2  2  4 ∫ exp − exp [−i k x ] dx
 2 −∞ 4 
2

[ ]
1 ∞ 2
1 −  x− x 0 
=  2  2  4 exp [ −i k x0 ] ∫ exp − −i k  x − x0  dx
 2 −∞ 4 
2

[ ]
1 ∞ 2 2
1 − [ x− x 0 2 i k  ]
=  2  2  4 exp [ −i k x0 ] ∫ exp − 2 2
−k  dx
 2 −∞ 4
2

[ ]
− 1 ∞ 2
1
=  2   4 exp [ −i k x0 ] exp [−k 2  2 ] ∫ exp − x ' 2 dx
2
With: x '= x − x 02 i k 
2
, dx ' =dx
 2 −∞ 4
1
1 −
=  2   4 exp [ −i k x0 ] exp [−k 2  2 ]  2 
2

 2

 
−1

exp [−i k x 0 ] exp [ −k  ]
4 2 2
=
2 2

  [
1
 −
∣k
2
 ∣ = 2
2
exp −2 k 
2 2
]
2

[ ] [ ]  a 3

  
∞ 2 ∞ 2

  −
1
x x
2 dx =  a
2
∫ ∣k 2
 ∣ dk =
2
=1 ∫ exp − , ∫ x exp − dx =
a2 2
2 2
−∞ 2 2 ∞
− a −∞

2222 Quantum Physics 2006-7 149


Wavepackets and Uncertainty Principle (2)

〈
Mean-squared uncertainty in postion  x 2 = x−x 2 =
0 〉 ∫  0
−∞
x−x ∣
2
x ∣ dx= 22

2
Mean momentum: p=ℏ k ∣
 k ∣ is symmetric about k =0 ⇒ 〈 p 〉 =0

ℏ2
Mean-squared uncertainty in momentum:  p = 〈 p 〉 =ℏ 〈 k 〉 =
2 2 2 2
2
4
2 2 2 ℏ2
ℏ 2
⇒  x  p = 2
=
4 4
In fact, can show that this form of wavepacket (“Gaussian wavepacket”) minimizes the
product of Δx and Δp, so:

 x p  Rigorous statement of
2 uncertainty principle for
position and momentum.
2222 Quantum Physics 2006-7 150
Wavepackets and Uncertainty Principle (3)

Summary
Three ways of thinking of Uncertainty principle:
(3) Arising from the physics of the interaction of different types of measurement apparatus with the
system (e.g. in the gamma-ray microscope);
(4) Arising from the properties of Fourier transforms (narrower wavepackets need a wider range of
wavenumbers in their Fourier transforms);
(5) Arising from the fact that x and p are not compatible quantities (do not commute), so they
cannot simultaneously have precisely defined values.

General result (see third year, or Rae §4.5):

For general non-commuting operators Q  ,R



1  
 q  r  ∣〈 [ Q , R ] 〉∣
2
Where:  q and  r are RMS uncertainties in Q  and R

2222 Quantum Physics 2006-7 151

You might also like