Sie sind auf Seite 1von 35

Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

Foundations of Quantum Mechanics

Dr. H. Osborn1

Michlmas 1997

1 A
LT
EXed by Paul Metcalfe comments and corrections to pdm23@cam.ac.uk.
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

T HANKS FOR YOUR PATIENCE


T O C LAIRE
Contents
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

Introduction v

1 Basics 1
1.1 Review of earlier work . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 The Dirac Formalism . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.1 Continuum basis . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.2 Action of operators on wavefunctions . . . . . . . . . . . . . 5
1.2.3 Momentum space . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.4 Commuting operators . . . . . . . . . . . . . . . . . . . . . 7
1.2.5 Unitary Operators . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.6 Time dependence . . . . . . . . . . . . . . . . . . . . . . . . 8

2 The Harmonic Oscillator 9


2.1 Relation to wavefunctions . . . . . . . . . . . . . . . . . . . . . . . 10
2.2 More comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

3 Multiparticle Systems 13
3.1 Combination of physical systems . . . . . . . . . . . . . . . . . . . . 13
3.2 Multiparticle Systems . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.2.1 Identical particles . . . . . . . . . . . . . . . . . . . . . . . . 14
3.2.2 Spinless bosons . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2.3 Spin 21 fermions . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3 Two particle states and centre of mass . . . . . . . . . . . . . . . . . 17
3.4 Observation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

4 Perturbation Expansions 19
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.2 Non-degenerate perturbation theory . . . . . . . . . . . . . . . . . . 19
4.3 Degeneracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

5 General theory of angular momentum 23


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
5.1.1 Spin 21 particles . . . . . . . . . . . . . . . . . . . . . . . . . 24
5.1.2 Spin 1 particles . . . . . . . . . . . . . . . . . . . . . . . . . 25
5.1.3 Electrons . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5.2 Addition of angular momentum . . . . . . . . . . . . . . . . . . . . . 25
5.3 The meaning of quantum mechanics . . . . . . . . . . . . . . . . . . 26

iii
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

iv
CONTENTS
Introduction
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

These notes are based on the course Foundations of Quantum Mechanics given by
Dr. H. Osborn in Cambridge in the Michlmas Term 1997. These typeset notes have
been produced mainly for my own benefit but seem to be officially supported. Recom-
mended books are discussed in the bibliography at the back.
A word or two about the philosophy of these notes seem in order. They are based in
content on the lectures given, but I have felt free to expand and contract various details,
as well as to clarify explanations and improve the narrative flow. Errors in content are
(hopefully) mine and mine alone but I accept no responsibility for your use of these
notes.
Other sets of notes are available for different courses. At the time of typing, these
courses were:
Probability Discrete Mathematics
Analysis Further Analysis
Quantum Mechanics Fluid Dynamics 1
Quadratic Mathematics Geometry
Dynamics of D.E.s Foundations of QM
Electrodynamics Methods of Math. Phys
Fluid Dynamics 2 Waves (etc.)
Applications of QM Dynamical Systems
Statistical Physics

They may be downloaded from


http://pdm23.trin.cam.ac.uk/pdm23/maths/ or
http://www.damtp.cam.ac.uk/
or you can email me on pdm23@cam.ac.uk to get a copy of the sets you require.
Even if you download them please email me to let me know, so that I can keep you up
to date with the errata and new note sets. The other people who have contributed time
and effort to these note sets are:
Richard Cameron Analysis Hugh Osborn Proofreading
Claire Gough Proofreading Malcolm Perry Accomodation
Kate Metcalfe Probability David Sanders Proofreading

Although these notes are free of charge anyone who wishes to express their thanks
could send a couple of bottles of interesting beer to Y1 Burrells Field, Grange Road.
Paul Metcalfe
15th December 1997

v
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

vi
INTRODUCTION
Chapter 1
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

The Basics of Quantum


Mechanics

Quantum mechanics is viewed as the most remarkable development in 20th century


physics. Its point of view is completely different from classical physics. Its predictions
are often probabilistic.
We will develop the mathematical formalism and some applications. We will em-
phasize vector spaces (to which wavefunctions belong). These vector spaces are some-
times finite-dimensional, but more often infinite dimensional. The pure mathematical
basis for these is in Hilbert Spaces but (fortunately!) no knowledge of this area is
required for this course.

1.1 Review of earlier work


This is a brief review of the salient points of the 1B Quantum Mechanics course. If
you anything here is unfamiliar it is as well to read up on the 1B Quantum Mechanics
course. This section can be omitted by the brave.
A wavefunction (x) : R3 7 C is associated with a single particle in three di-
mensions. represents the state of a physical system for a single particle. If is
normalised, that is Z
2 2
kk d3 x || = 1
2
then we say that d3 x || is the probability of finding the particle in the infinitesimal
region d3 x (at x).

Superposition Principle
If 1 and 2 are two wavefunctions representing states of a particle, then so is the
linear combination a1 1 + a2 2 (a1 , a2 C). This is obviously the statement that
wavefunctions live in a vector space. If = a (with a 6= 0) then and represent
the same physical state. If and are both normalised then a = e . We write
e to show that they represent the same physical state.
For two wavefunctions and we can define a scalar product
Z
(, ) d3 x C.

1
2 CHAPTER 1. BASICS

This has various properties which you can investigate at your leisure.

Interpretative Postulate
Given a particle in a state represented by a wavefunction (henceforth in a state
2
) then the probability of finding the particle in state is P = |(, )| and if the
wavefunctions are normalised then 0 P 1. P = 1 if .
We wish to define (linear) operators on our vector space do the obvious thing.
In finite dimensions we can choose a basis and replace an operator with a matrix.
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

For a complex vector space we can define the Hermitian conjugate of the operator A
to be the operator A satisfying (, A) = (A , ). If A = A then A is Hermitian.
Note that if A is linear then so is A .
In quantum mechanics dynamical variables (such as energy, momentum or angular
momentum) are represented by (linear) Hermitian operators, the values of the dynam-
ical variables being given by the eigenvalues. For wavefunctions (x), A is usually
a differential operator. For a single particle moving in a potential V (x) we get the
~2
Hamiltonian H = 2m 2 + V (x). Operators may have either a continuous or dis-
crete spectrum.
If A is Hermitian then the eigenfunctions corresponding to different eigenvalues
are orthogonal. We assume completeness that any wavefunction can be expanded
as a linear combination of eigenfunctions.
The expectation value for A in a state with wavefunction is hAi , defined to be
P 2 2 2
i i |ai | = (, A). We define the square deviation A to be h(A hAi ) i
which is in general nonzero.

Time dependence
This is governed by the Schrodinger equation


~ = H,
t
where H is the Hamiltonian. H must be Hermitian for the consistency of quantum
mechanics:

~ (, ) = (, H) (H, ) = 0
t
if H is Hermitian. Thus we can impose the condition (, ) = 1 for all time (if is
normalisable).
If we consider eigenfunctions i of H with eigenvalues Ei we can expand a general
wavefunction as X Ei
(x, t) = ai e ~ t i (x).
i
2
If is normalised then the probability of finding the system with energy Ei is |ai | .

1.2 The Dirac Formalism


This is where we take off into the wild blue yonder, or at least a more abstract form of
quantum mechanics than that previously discussed. The essential structure of quantum
1.2. THE DIRAC FORMALISM 3

mechanics is based on operators acting on vectors in some vector space. A wavefunc-


tion corresponds to some abstract vector |i, a ket vector. |i represents the state of
some physical system described by the vector space.
If |1 i and |2 i are ket vectors then |i = a1 |1 i + a2 |2 i is a possible ket vector
describing a state this is the superposition principle again.
We define a dual space of bra vectors h| and a scalar product h|i, a complex
number.1 For any |i there corresponds a unique h| and we require h|i = h|i .
We require the scalar product to be linear such that |i = a1 |1 i + a2 |2 i implies
h|i = a1 h|1 i + a2 h|2 i. We see that h|i = a1 h1 |i + a2 h2 |i and so
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

h| = a1 h1 | + a2 h2 |.
We introduce linear operators A|i = | i and we define operators acting on bra
vectors to the left h|A = h | by requiring h |i = h|A|i for all . In general, in
h|A|i, A can act either to the right or the left. We define the adjoint A of A such
that if A|i = | i then h|A = h |. A is said to be Hermitian if A = A .
If A = a1 A1 + a2 A2 then A = a1 A1 + a2 A2 , which can be seen by appealing to
the definitions. We also find the adjoint of B A as follows:
Let B A|i = B| i = | i. Then h | = h |B = h|A B and the result
follows. Also, if h|A = h | then | i = A |i.
We have eigenvectors A|i = |i and it can be seen in the usual manner that the
eigenvalues of a Hermitian operator are real and the eigenvectors corresponding to two
different eigenvalues are orthogonal.
We assume completeness
P that is any |i can be expanded in terms of the basis
ket vectors, |i = i ai | i i where A|i i = i |i i and ai = hi |i. If |i is
normalised h|i = 1 then the expected value of A is hAi = h|A|i, which
is real if A is Hermitian.
P
The completeness relation for eigenvectors of A can be written as 1 = i |i ihi |,
which gives (as before)
X
|i = 1|i = |i ihi |i.
i

P
We can also rewrite A = i |i ii hi | and if j 6= 0 j then we can define
A1 = i |i i1
P
i h i |.
We now choose P orthonormal basis {|ni} with hn|mi
an P = nm and the complete-
ness relation 1 = n |nihn|. We can thus expand |i = n an |ni with an = hn|i.
We now consider a linear operator A, and then A|i = n an A|ni = m am |mi,
P P

with am = hm|A|i = n an hm|A|ni. Further, putting Amn = hm|A|ni we get


P

am =
P
n Amn an and therefore solving A|i = |i is equivalent to solving the
equation Aa = a. AmnPis called the matrix representation of A. We also have
matrix P
h| = n an hn|, with an = m am Amn , where Amn = Anm gives the Hermitian
conjugate matrix. This is the matrix representation of A .

1.2.1 Continuum basis


In the above we have assumed discrete eigenvalues i and normalisable eigenvectors
|i i. However, in general, in quantum mechanics operators often have continuous
1 bra ket. Who said that mathematicians have no sense of humour?
4 CHAPTER 1. BASICS

spectrum for instance the position operator x in 3 dimensions. x must have eigen-
values x for any point x R3 . There exist eigenvectors |xi such that x|xi = x|xi for
any x R3 .
As x must be Hermitian we have hx|x = xhx|. We define the vector space required
in the Dirac formalism as that spanned by |xi.
For any state |i we can define a wavefunction (x) = hx|i.
We also need to find some normalisation criterion, which uses the 3 dimensional
Dirac delta function to get hx|x i = 3 (x x ). Completeness gives
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

Z
d3 x|xihx| = 1.

We can also recover the ket vector from the wavefunction by


Z
|i = 1|i = d3 x|xi(x).

Also hx|x|i = x(x); the action of the operator x on a wavefunction is multipli-


cation by x.
Something else reassuring is
Z
h|i = h|1|i = d3 xh|xihx|i
Z
2
= d3 x |(x)| .

The momentum operator p is also expected to have continuum eigenvalues. We


can similarly define states |pi which satisfy p|pi = p|pi. We can relate x and p using
the commutator, which for two operators A and B is defined by
h i
A, B = AB B A.

The relationship between x and p is [xi , pj ] = ~ij . In one dimension [x, p] = ~.


We have a useful rule for calculating commutators, that is:
h i h i h i
A, B C = A, B C + B A, C .

This can be easily proved simply by expanding the right hand side out. We can use this
to calculate x, p2 .

x, p2 = [x, p] p + p [x, p]

= 2~p.

It is easy to show by induction that [x, pn ] = n~pn1 .


We can define an exponential by
n
a~p
X 1 ap
e = .
n=0
n! ~
1.2. THE DIRAC FORMALISM 5
h ap
i
We can evaluate x, e ~ by


" n #
h
a~p
i X 1 ap
x, e = x,
n=0
n! ~

X 1 n
ap
= x,
n=0
n! ~

X 1 a n
= [x, pn ]
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

n=0
n! ~

X 1 a n
= ~pn1
n=1
(n 1)! ~

X a n1 n1
=a p
n=1
~
ap
= ae ~

and by rearranging this we get that


ap ap
xe ~ = e ~ (x + a)
ap
and it follows that e ~ |xi is an eigenvalue of x with eigenvalue x + a. Thus we
ap
see e ~ |xi = |x + ai. We can do the same to the bra vectors with the Hermi-
ap ap
tian conjugate e ~ to get hx + a| = hx|e ~ . Then we also have the normalisation
hx + a|x + ai = hx |xi.
ap ap
We now wish to consider hx + a|pi = hx|e ~ |pi = e ~ hx|pi. Setting x = 0 gives
ap
ha|pi = e ~ N , where N = h0|pi is independent of x. We can determine N from the
normalisation of |pi.

Z

(p p) = hp |pi = da hp |aiha|pi
Z
a(pp )
2
= |N | da e ~
2
= |N | 2~ (p p)
1
12
So, because we are free to choose the phase of N , we can set N = 2~ and
1 12 xp
thus hx|pi = 2~ e ~ . We could define |pi by
12 Z
1
Z
xp
|pi = dx |xihx|pi = dx |xie ~ ,
2~

but we then have to check things like completeness.

1.2.2 Action of operators on wavefunctions


We recall the definition of the wavefunction as (x) = hx|i. We wish to see what
operators (the position and momentum operators discussed) do to wavefunctions.
6 CHAPTER 1. BASICS

Now hx|x|i = xhx|i = x(x), so the position operator acts on wavefunctions


by multiplication. As for the momentum operator,

Z
hx|p|i = dp hx|p|pihp|i
Z
= dp phx|pihp|i
12 Z

1 xp
= dp pe ~ hp|i
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

2~
d
Z
= ~ dp hx|pihp|i
dx
d d
= ~ hx|i = ~ (x).
dx dx
d

The commutation relation [x, p] = ~ corresponds to x, ~ dx = ~ (acting on
(x)).

1.2.3 Momentum space


|xi 7 (x) = hx|i defines a particular representation of the vector space. It is
sometimes useful to use a momentum representation, (p) = hp|i. We observe that
Z
(p) = dx hp|xihx|i
21 Z
1 xp
= dx e ~ (x).
2~

In momentum space, the operators act differently on wavefunctions. It is easy to


d
see that hp|p|i = p(p) and hp|x|i = ~ dp (p).
We convert the Schrodinger equation into momentum space. We have the operator
p2
equation H = 2m + V (x) and we just need to calculate how the potential operates on
the wavefunction.

Z
hp|V (x)|i = dx hp|V (x)|xihx|i
12 Z
1 xp
= dx e ~ V (x)hx|i
2~
1
ZZ
x(p p)
= dx dp V (x)(p )e ~
2~
Z
= dp V (p p )(p ),

xp
1
dp e
R
where V (p) = 2~
~ V (x). Thus in momentum space,

p2
Z
Hp (p) = (p) + dp V (p p )(p ).
2m
1.2. THE DIRAC FORMALISM 7

1.2.4 Commuting operators


h i
Suppose A and B are Hermitian and A, B = 0. Then A and B have simultaneous
eigenvectors.

Proof. Suppose A|i = |i and the vector subspace V is the span of the eigenvec-
tors of A with eigenvalue . (If dim V > 1 then is said to be degenerate.)
As A and B commute we know that B|i = AB|i and so B|i V . If is
non-degenerate then B|i = |i for some . Otherwise we have that B : V 7 V
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

and we can therefore find eigenvectors of B which lie entirely inside V . We can label
these as |, i, and we know that

A|, i = |, i
B|, i = |, i.

These may still be degenerate. However we can in principle remove this degener-
acy by adding more commuting operators until each state is uniquely labeled by the
eigenvalues of each common eigenvector. This set of operators is called a complete
commuting set.
This isnt so odd: for a single particle in 3 dimensions we have the operators x1 , x2
and x3 . These all commute, so for a single particle with no other degrees of freedom
we can label states uniquely by |xi. We also note from this example that a complete
commuting set is not unique, we might just as easily have taken the momentum opera-
tors and labeled states by |pi. To ram the point in more, we could also have taken some
weird combination like x1 , x2 and p3 .
For our single particle in 3 dimensions, a natural set of commuting operators in-
volves the angular momentum operator, L = x p, or Li = ijk xj pk .
We can find commutation relations between Li and the other operators we know.
These are summarised here, proof is straightforward.
h i
Li , xl = ~ilj xj
h i
Li , x2 = 0
h i
Li , pm = ~imk pk
h i
Li , p2 = 0
h i
Li , Lj = ~ijk Lk
h i
Li , L2 = 0
h i
p2
If we have a Hamiltonian H = 2m + V (|x|) then we can also see that L, H = 0.
We choose as a commuting set H, L2 and L3 and label states |E, l, mi, where the
eigenvalue of L2 is l(l + 1) and the eigenvalue of L3 is m.
8 CHAPTER 1. BASICS

1.2.5 Unitary Operators


An operator U is said to be unitary if U U = 1, or equivalently U 1 = U .
Suppose U is unitary and U |i = | i, U |i = | i. Then h | = h|U and

h | i = h|i. Thus the scalar product, which is the probability amplitude of finding
the state |i given the state |i, is invariant under unitary transformations of states.
For any operator A we can define A = U AU . Then h |A | i = h|A|i and
matrix elements are unchanged under unitary transformations. We also note that if
C = AB then C = A B .
The quantum mechanics for the |i, |i, A, B etc. is the same as for | i, | i, A ,
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.


B and so on. A unitary transform in quantum mechanics is analogous to a canonical
transformation in dynamics.

Note that if O is Hermitian then U = eO is unitary, as U = eO = eO .

1.2.6 Time dependence


This is governed by the Schrodinger equation,


~ |(t)i = H|(t)i.
t

H is the Hamiltonian and we require it to be Hermitian. We can get an explicit


solution of this if H does not depend explicitly on t. We set |(t)i = U (t)|(0)i,
Ht
where U (t) = e ~ . As U (t) is unitary, h(t)|(t)i = h(0)|(0)i.
If we measure the expectation of A at time t we get h(t)|A|(t)i = a(t). This
description is called the Schrodinger picture. Alternatively we can absorb the time de-
pendence into the operator A to get the Heisenberg picture, a(t) = h|U (t)AU (t)|i.
We write AH (t) = U (t)AU (t). In this description the operators are time dependent
(as opposed to the states). AH (t) is the Heisenberg picture time dependent operator.
Its evolution is governed by

h i
~ AH (t) = AH (t), H ,
t
which is easily proven.
1 2
For a Hamiltonian H = 2m p(t) + V (x(t)) we can get the Heisenberg equations
for the operators xH and pH

d 1
xH (t) = pH (t)
dt m
d
pH (t) = V (xH (t)).
dt
These ought to remind you of something.
Chapter 2
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

The Harmonic Oscillator

In quantum mechanics there are two basic solvable systems, the harmonic oscillator
and the hydrogen atom. We will examine the quantum harmonic oscillator using al-
gebraic methods. In quantum mechanics the harmonic oscillator is governed by the
Hamiltonian
1 2 1
H = p + 2 m 2 x2 ,
2m

with the condition that [x, p] = ~. We wish to solve H|i = E|i to find the energy
eigenvalues.
We define a new operator a.

m 12 p

a = x +
2~ m
m 12
p


a = x .
2~ m

a and a are respectively called the annihilation and creation operators. We can
easily obtain the commutation relation a, a = 1. It is easy to show that, in terms

H = 21 ~

aa i + a a ,

of the annihilation and creation operators, the Hamiltonian
h i h
which reduces to ~ a a + 12 . Let N = a a. Then a, N = a and a , N = a .


Therefore N a = a N 1 and N a = a N + 1 .
Suppose |i is an eigenvector of N with eigenvalue . Then the commutation rela-
tions give that N a|i = ( 1) a|i and therefore unless a|i = 0 it is an eigenvalue
of N with eigenvalue 1. Similarly N a |i = ( + 1) a |i.
But for any |i, h|N |i 0 and equals 0 iff a|i = 0. Now suppose we have
an eigenvalue / {0, 1, 2, . . . }. Then n such that an |i is an eigenvector of N
with eigenvalue n < 0 and so we must have {0, 1, 2, . . . }. Returning to the
Hamiltonian we get energy eigenvalues En = ~ n + 12 , the same result as using the
Schrodinger equation for wavefunctions, but with much less effort.
We define |ni = Cn an |0i, where Cn is such as to make hn|ni = 1. We can take
Cn R, and evaluate h0|an an |0i to find Cn .

9
10 CHAPTER 2. THE HARMONIC OSCILLATOR

1 = hn|ni
= Cn2 h0|an an |0i
= Cn2 h0|an1 aa an1 |0i
Cn2
= 2 hn 1|aa |n 1i
Cn1
Cn2
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

= 2 hn 1|N + 1|n 1i
Cn1
Cn2
= 2 (n 1 + 1)hn 1|n 1i
Cn1
Cn2
=n 2 .
Cn1

1
We thus require Cn = Cn1 / n and as C0 = 1 we get Cn = (n!) 2 and so we
have the normalised eigenstate (of N ) |ni = 1n! a |0i (with eigenvalue n). |ni is also
an eigenvector of H with eigenvalue ~ n + 21 . The space of states for the harmonic

oscillator is spanned by {|ni}.


We also need to ask if there exists a non-zero state |i such that a |i = 0. Then

0 = h|aa |i = h|i + h|a a|i h|i > 0.

So there exist no non-zero states |i such that a |i = 0.

2.1 Relation to wavefunctions


We evaluate
m 12 ~ d

0 = hx|a|0i = x+ hx|0i
2~ m dx
and we see that 0 (x) = hx|0i satisfies the differential equation

d m
+ x 0 (x) = 0.
dx ~
1 m 2
This (obviously) has solution 0 (x) = N e 2 ~ x for some normalisation constant
N . This is the ground state wavefunction which has energy 12 ~.
For 1 (x) = hx|1i = hx|a |0i we find
m 12
1 (x) = hx| x p |0i
2~ m
m 12
~ d
= x 0 (x)
2~ m dx
1
2m 2
= x0 (x).
~
2.2. MORE COMMENTS 11

2.2 More comments


Many harmonic oscillator problems are simplified using the creation and annihilation
operators.1 For example

12
~
hm|x|ni = hm|a + a |ni
2m
12


~
= n hm|n 1i + n + 1 hm|n + 1i
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

2m
12


~
= n m,n1 + n + 1 m,n+1 .
2m

This is non-zero only if m = n 1. We note that xr contains terms as ars , where


0 s r and so hm|xr |ni can be non-zero only if n r m n + r.
Ht Ht
It is easy to see that in the Heisenberg picture aH (t) = e ~ ae ~ = et a.
Then using the equations for xH (t) and pH (t), we see that
1
xH (t) = x cos t + m p sin t.

Also, HaH (t) = aH (t)(H + ~), so if |i is an energy eigenstate with eigenvalue


E then aH (t)|i is an energy eigenstate with eigenvalue E + ~.

1 And such problems always occur in Tripos papers. You have been warned.
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

12
CHAPTER 2. THE HARMONIC OSCILLATOR
Chapter 3
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

Multiparticle Systems

3.1 Combination of physical systems


In quantum mechanics each physical system has its own vector space of physical states
and operators, which if Hermitian represent observed quantities.
If we consider two vector spaces V1 and V2 with bases {|ri1 } and {|si2 } with
r = 1 . . . dim V1 and s = 1 . . . dim V2 . We define the tensor product V1 V2 as the
vector space spanned by pairs of vectors

{|ri1 |si2 : r = 1 . . . dim V1 , s = 1 . . . dim V2 }.

We see that dim(V1 V2 ) = dim V1 dim V2 . We also write the basis vectors of
V1 V2 as |r, si. We can define a scalar product on V1 V2 in terms of the basis vectors:
hr , s |r, si = hr |ri1 hs |si2 . We can see that if {|ri1 } and {|si2 } are orthonormal
bases for their respective vector spaces then {|r, si} is an orthonormal basis for V1 V2 .
Suppose A1 is an operator on V1 and B2 is an operator on V2 we can define an
operator A1 B2 on V1 V2 by its operation on the basis vectors:

A1 B2 |ri1 |si2 = A1 |ri1 B2 |si2 .

We write A1 B2 as A1 B2 .

Two harmonic oscillators


We illustrate these comments by example. Suppose

p2i
Hi = + 1 mx2i i = 1, 2.
2m 2
We have two independent vector spaces Vi with bases |nii where n = 0, 1, . . . and
ai and ai are creation and annihilation operators on Vi , and

Hi |nii = ~ n + 12 |nii .

For the combined system


P we form the tensor product V1 V2 with basis |n1 , n2 i
and Hamiltonian H = i Hi , so H|n1 , n2 i = ~ (n1 + n2 + 1) |n1 , n2 i. There are
N + 1 ket vectors in the N th excited state.

13
14 CHAPTER 3. MULTIPARTICLE SYSTEMS

The three dimensional harmonic oscillator follows similarly. In general if H1 and


H2 are two independent Hamiltonians which act on V1 and V2 respectively then the
Hamiltonian for the combined system is H = H1 + H2 acting on V1 V2 . If {|r i}
and {|s i} are eigenbases for V1 and V2 with energy eigenvalues {Er1 } and {Es2 }
respectively then the basis vectors {|ir,s } for V1 V2 have energies Er,s = Er1 + Es2 .

3.2 Multiparticle Systems


We have considered single particle systems with states |i and wavefunctions (x) =
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

hx|i. The states belong to a space H.


Consider an N particle system. We say the states belong to Hn = H1 HN
and define a basis of states |r1 i1 |r2 i2 . . . |rN iN where {|ri ii } is a basis for Hi .
A general state |i is a linear combination of basis vectors and we can define the
N particle wavefunction as (x1 , x2 , . . . , xN ) = hx1 , x2 , . . . , xN |i.
The normalisation condition is
Z
2
h|i = d3 x1 . . . d3 xN |(x1 , x2 , . . . , xN )| = 1 if normalised.

2
We can interpret d3 x1 . . . d3 xN |(x1 , x2 , . . . , xN )| as the probability density
that particle i is in the volume element d3 xi at xi . We can obtain the probability
density for one particle by integrating out all the other xj s.

For time evolution we get the equation ~ t |i = H|i, where H is an operator
N
on H .
If the particles do not interact then
N
X
H = Hi
i=1

where Hi acts on Hi but leaves Hj alone for j 6= i. We have energy eigenstates in each
Hi such that Hi |r ii = Er |r ii and so |i = |r1 i1 |r2 i2 . . . |rN iN is an energy
eigenstate with energy Er1 + + ErN .

3.2.1 Identical particles


There are many such cases, for instance multielectron atoms. We will concentrate on
two identical particles.
Identical means that physical quantities are be invariant under interchange of
particles. For instance if we have H = H(x1 , p1 , x2 , p2 ) then this must equal the
permuted Hamiltonian H(x2 , p2 , x1 , p1 ) if we have identical particles. We introduce
U such that

U x1 U 1 = x2 U x2 U 1 = x1
U p1 U 1 = p2 U p2 U 1 = p1 .

We should also have U H U 1 = H and more generally if A1 is an operator on


particle 1 then U A1 U 1 is the corresponding operator on particle 2 (and vice versa).
3.2. MULTIPARTICLE SYSTEMS 15

Note that if |i is an energy eigenstate of H then so is U |i. Clearly U 2 = 1 and we


require U to be unitary, which implies that U is Hermitian.
In quantum mechanics we require |i and U |i to be the same states (for identical
particles). This implies that U |i = |i and the requirement U 2 = 1 gives that
= 1. In terms of wavefunctions this means that (x1 , x2 ) = (x2 , x1 ). If we
have a plus sign then the particles are bosons (which have integral spin) and if a minus
sign then the particles are fermions (which have spin 12 , 23 , . . . ).1
The generalisation to N identical particles is reasonably obvious. Let Uij inter-
1
change particles i and j. Then Uij H Uij = H for all pairs (i, j).
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

The same physical requirement as before gives us that Uij |i = |i for all pairs
(i, j).
If we have bosons (plus sign) then in terms of wavefunctions we must have

(x1 , . . . , xN ) = (xp1 , . . . , xpN ),

where (p1 , . . . , pN ) is a permutation of (1, . . . , N ). If we have fermions then

(x1 , . . . , xN ) = (xp1 , . . . , xpN ),

where = +1 if we have an even permutation of (1, . . . , N ) and 1 if we have an odd


permutation.
Remark for pure mathematicians. 1 and {1} are the two possible one dimensional
representations of the permutation group.

3.2.2 Spinless bosons


(Which means that the only variables for a single particle are x and p.) Suppose
we have two identical non-interacting bosons. Then H = Hi + H2 and we have
H1 |r ii = Er |r ii . The general space with two particles is H1 H2 which has
a basis {|r i1 |s i2 }, but as the particles are identical the two particle state space is
(H1 H2 )S where we restrict to symmetric combinations of the basis vectors. That
is, a basis for this in terms of the bases of H1 and H2 is
n o
|r i1 |r i2 ; 12 (|r i1 |s i2 + |s i1 |r i2 ) , r 6= s .

The corresponding wavefunctions are

r (x1 )r (x2 ) and 1 (r (x1 )s (x2 ) + s (x1 )r (x2 ))


2

1
and the corresponding eigenvalues are 2Er and Er +Es . The factor of 2 2 just ensures
normalisation and

1 (1hr |2hs | + 1hs |2hr |) 12 (|r i1 |s i2 + |s i1 |r i2 )


2

evaluates to rr ss + rs r s . P
For N spinless bosons with H = Hi the appropriate completely symmetric
states are

1 | i
r1 1 . . . | i
rN N + permutations thereof if ri 6= rj
N!

1 Spin will be studied later in the course.


16 CHAPTER 3. MULTIPARTICLE SYSTEMS

1
3.2.3 Spin 2
fermions
In this case (which covers electrons, for example) a single particle state (or wavefunc-
tion) depends on an additional discrete variable s. The wavefunctions are (x, s) or
s (x). The space of states for a single electron H = L2 (R3 ) C2 has a basis of the
form |xi|si |x, si and the wavefunctions can be written s (x) = hx, s|i. A basis
of wavefunctions is {r (x, s) = r (x) (s)}, where r and are labels for the basis.
takes two values and it will later be seen to be natural
to take = 12 .
(1)
We can also think of the vector = , in which case two possible basis
(2)
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.


1 0
vectors are and . Note that = .
0 1
The scalar product is defined in the obvious way: hr |r i = hr |r ih | i,
which equals rr if the initial basis states are orthonormal.
The two electron wavefunction is (x1 , s1 ; x2 , s2 ) and under the particle exchange
operator U we must have (x1 , s1 ; x2 , s2 ) 7 (x2 , s2 ; x1 , s1 ). The two particle
states belong to the antisymmetric combination (H1 H2 )A .
For N electrons the obvious thing can be done.

Basis for symmetric or antisymmetric 2 particle spin states


There is only one antisymmetric basis state
1
A (s1 , s2 ) = 1 (s1 ) 1 (s2 ) 1 (s1 ) 1 (s2 ) ,
2 2 2 2 2

and three symmetric possibilities:



1 (s ) 1 (s )
2 1 2 2


S (s1 , s2 ) = 12 1 (s1 ) 1 (s2 ) + 1 (s1 ) 1 (s2 ) s1 6= s2
2 2 2 2

(s ) (s ).
21 1 12 2

We can now examine two non-interacting electrons, with H = H1 + H2 and take


Hi independent of spin. The single particle states are |i i|s i.
The two electron states live in (H1 H2 )A , which has a basis

|r i1 |r i2 |A i;
1
(|r i1 |s i2 + |s i1 |r i2 ) |A i; r 6= s
2
1
(|r i1 |s i2 |s i1 |r i2 ) |S i; r 6= s,
2
with energy levels 2Er (one spin state) and Er + Es (one antisymmetric spin state and
three symmetric spin states).
We thus obtain the Pauli exclusion principle: no two electrons can occupy the same
state (taking account of spin).
As an example we can take the helium atom with Hamiltonian

p21 p2 2e2 2e2 e2


H = + 2 + .
2m 2m 40 |x1 | 40 |x2 | 40 |x1 x2 |
3.3. TWO PARTICLE STATES AND CENTRE OF MASS 17

If we neglect the interaction term we can analyse this as two hydrogen atoms and glue
the results back together as above. The hydrogen atom (with a nuclear charge 2e) has
2e2
En = 8 0n
2 , so we get a ground state for the helium atom with energy 2E1 with

no degeneracy and a first excited state with energy E1 + E2 with a degeneracy of four.
Hopefully these bear some relation to the results obtained by taking the interaction into
account.

3.3 Two particle states and centre of mass


Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

p2 p2
Suppose we have a Hamiltonian H = 2m1 + 2m2 + V (x1 x2 ) defined on H2 . We can
separate out the centre of mass motion by letting

1
P = p1 + p2 p = 2 (p1 p2 )
1
X = (x1 + x2 )
2 x = x1 x2 .
h i
Then Xi , Pj = ~ij , [xi , pj ] = ~ij and X, P and x, p commute respectively.
2 2
We can rewrite the Hamiltonian as H = 2M P
+ h, h = pm + V (x), where M = 2m and
2
we can decompose H into HCM Hint . HCM is acted on by X and P and has wave-
functions (X). Hint is acted on by x, p and any spin operators. It has wavefunctions
(x, s1 , s2 ). We take wavefunctions (x1 , s1 ; x2 , s2 ) = (X)(x, s1 , s2 ) in H2 .
P.X
This simplifies the Schrodinger equation, we can just have (X) = e ~ and then
2
P
E = 2M + Eint . We thus need only to solve the one particle equation h = Eint .
Under the particle exchange operator U we have

(x, s1 , s2 ) 7 (x, s2 , s1 ) = (x, s1 , s2 ),

with a plus sign for bosons and a minus sign for fermions. In the spinless case then
(x) = (x).
If we have a potential V (|x|) then we may separate variables to get

x
(x, s1 , s2 ) = Yl R(|x|)(s1 , s2 )
|x|

x x
with Yl |x| = (1)l Yl |x| . For spinless bosons we therefore require l to be even.

3.4 Observation
Consider the tensor product of two systems H1 and H2 . A general state |i in H1 H2
can be written as X
|i = aij |i i1 |j i2
i,j

with |i i1 H1 and |j i2 H2 assumed orthonormal bases for their respective vector


spaces.
Suppose we make a measurement on the first system leaving the second P system
unchanged, and find the first system in a state |i i1 . Then 1hi |i = j aij |j i2 ,
which we write as Ai |i2 , where |i2 is a normalised state of the second system. We
18 CHAPTER 3. MULTIPARTICLE SYSTEMS

2
interpret |Ai | as the probability of finding system 1 in state |i i1 . After measurement
system 2 is in a state |i2 .
If aij = i ij (no summation) then Ai = i and measurement of system 1 as |i i1
determines system 2 to be in state |i i2 .
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.
Chapter 4
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

Perturbation Expansions

4.1 Introduction
Most problems in quantum mechanics are not exactly solvable and it it necessary to
find approximate answers to them. The simplest method is a perturbation expansion.
We write H = H0 + H where H0 describes a solvable system with known eigenvalues
and eigenvectors, and H is in some sense small.
We write H() = H0 + H and expand the eigenvalues and eigenvectors in
powers of . Finally we set = 1 to get the result. Note that we do not necessarily
have to introduce ; the problem may have some small parameter which we can use.
This theory can be applied to the time dependent problem but here we will only discuss
the time independent Schrodinger equation.

4.2 Non-degenerate perturbation theory


Suppose that H0 |ni = n |ni for n = 0, 1, . . . . We thus assume discrete energy levels
and we assume further that the energy levels are non-degenerate. We also require H
to be sufficiently non-singular to make a power series expansion possible.
We have the equation H()|n ()i = En ()|n ()i. We suppose that En ()
tends to n as 0 and |n ()i |ni as 0. We pose the power series
expansions

En () = n + En(1) + 2 En(2) + . . .
|n ()i = N |ni + |n(1) i + . . . ,

substitute into the Schrodinger equation and require it to be satisfied at each power
of . The normalisation constant N is easily seen to be 1 + O(2 ). The O(1) equation
is automatically satisfied and the O() equation is

H0 |n(1) i + H |ni = En(1) |ni + n |n(1) i.

(1) (1)
Note that we can always replace |n i with |n i + |ni and leave this equation
(1)
unchanged. We can therefore impose the condition hn|n i = 0. If we apply hn| to

19
20 CHAPTER 4. PERTURBATION EXPANSIONS

(1)
this equation we get En = hn|H |ni the first order perturbation in energy. If we
apply hr| where r 6= n we see that

hr|H |ni
hr|n(1) i =
r n

and therefore
X |rihr|H |ni
|n(1) i = .
r n
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

r6=n

Note that we are justified in these divisions as we have assumed that the eigenvalues
are non-degenerate. On doing the same thing to the O(2 ) equation we see that

En(2) = hn|H |n(1) i


2
X hr|H |ni

= .
r n
r6=n

This procedure is valid if r n is not very small when hr|H |ni =


6 0.
d
Using these results we can see that d En () = hn ()|H |n ()i and

X 1
|n ()i = |r ()ihr ()|H |n ()i.
Er () En ()
r6=n

H
Also = H and so

2
En () = 2hn ()|H |n ()i.
2

Example: harmonic oscillator


2
p
Consider H = 2m + 21 m 2 x2 + m 2 x2 , which can be viewed as H0 + H , where
H0 is the plain vanilla quantum harmonic oscillator Hamiltonian.
Calculating the matrix elements hr|x2 |ni required is an extended exercise in ma-
nipulations of the annihilation and creation operators and is omitted. The results are

En(1) = ~ n + 12

En(2) = 21 ~ n + 21 .

We thus get the perturbation expansion for En

En = ~ n + 21 1 + 12 2 + O(3 ) .


This system can also be solved exactly to give En = ~ n + 12

1 + 2 which
agrees with the perturbation expansion.
4.3. DEGENERACY 21

4.3 Degeneracy
The method given here breaks down if r = n for r 6= n. Perturbation theory can be
extended to the degenerate case, but we will consider only the first order shift in r .
We suppose that the states |n, si, s = 1 . . . Nn have the same energy n . Ns is the
degeneracy of this energy level.
As before we pose a Hamiltonian H = H0 + H such that H0 |n, si = n |n, si
and look for states |()i with energy E() n as 0.
The difference with the previous method is that we expand |()i as a power series
in in the basis of eigenvectors of H0 . That is
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

X
|()i = |n, sias + | (1) i.
s

As the as are arbitrary we can impose the conditions hn, s| (1) i = 0 for each s and
n. We thus have to solve H|()i = E()|()i with E() = n + E (1) . If we take
the O() equation and apply hn, r| to it we get
X
as hn, r|H |n, si = ar Er(1)
s

which is a matrix eigenvalue problem. Thus the first order perturbations in n are
the eigenvalues of the matrix hn, r|H |n, si. If all the eigenvalues are distinct then
the perturbation lifts the degeneracy. It is convenient for the purpose of calculation
to choose a basis for the space spanned by the degenerate eigenvectors in which this
matrix is as diagonal as possible.1

1 Dont ask...
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

22
CHAPTER 4. PERTURBATION EXPANSIONS
Chapter 5
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

General theory of angular


momentum

For a particle with position and momentum operators x and p with the commutation
relations [xi , pj ] =h ~ij iwe define L = x p. It can be seen that L is Hermitian and
it is easy to show Li , Lj = ~ijk Lk .

5.1 Introduction
More generally we define Hermitian angular momentum operators J with the commu-
tation relation [Ji , Jj ] = ~ijk Jk .1 We ask on what space of states can this algebra of
operators be realised, or alternatively, what are the representations?
We choose one component of J whose eigenvalues label the states; in accordance
with convention we choose J3 . Then we have J3 |mi = m~|mi for m R (discrete)
and we also have hm |mi = m m .
By applying the commutation relation we get (easily) that J3 , J2 = 0, so as J2

is Hermitian we can choose a simultaneous eigenbasis. That is, J2 |mi = ~2 |mi. We


know that 0 since J2 is the sum of the squares of Hermitian operators.
At this stage we choose an alternate basis; that is we split J into J+ , J and J3 ,

with J = J1 J2 . Note that J = J . The commutation relations for Ji give that
[J3 , J ] = ~J and [J+ , J ] = 2~J3 . It will be useful later to note that

(
J J+ + J32 + ~J3
J2 = 12 (J+ J + J J+ ) + J32 = .
J+ J + J32 ~J3

Proof of this is immediate. We can now rewrite the [J3 , J ] commutation rela-
tion as J3 J = J (J3 ~) and so we see that J |mi is an eigenvector of J3 with

eigenvalue (m + 1)~ and so J |mi = ~Nm |m 1i, with some normalisation con-

stant Nm . By evaluating the norm of J |mi, and noting that hm|mi = 1 we see that
2
(Nm ) = m2 m.
We can now define states |m ni for n = 0, 1, 2, . . . . We can pin them down more
2
by noting that (Nm ) 0 for positive norms. However the formulae we have are,
1 Because we are now grown up we will omit the hats if they do not add to clarity.

23
24 CHAPTER 5. GENERAL THEORY OF ANGULAR MOMENTUM

given , negative for sufficiently large |m| and so to avoid this we must have mmax = j
2
such that J+ |ji = 0 and so Nj+ = j 2 j = 0 and so = j(j + 1).

We can perform a similar trick with J ; there must exist mmin = j such that
J | j i = 0; thus = j (j + 1). So j = j and as j = j = j n for some
n {0, 1, 2, . . . } we have j = 0, 12 , 1, 23 , . . . .
In summary the states can be labelled by |j mi such that

J2 |j mi = ~2 j(j + 1)|j mi
J3 |j mi = ~m|j mi
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

1
J |j mi = ~ ((j m) (j m + 1)) 2 |j m + 1i

with m {j, j + 1, . . . , j 1, j} and j {0, 12 , 1, 23 , . . . }. There are 2j + 1 states


with different m for the same j. |j mi is the standard basis of the angular momentum
states.
We have obtained a representation of the algebra labelled by j. If J = L = x p
we must have j an integer.

P
P A we can define a matrix A by A|i = | iA .
Recall that if we have

Note that (BA) = B A . Given j, we have (J3 )m m = ~m m m and


p
(J )m m = ~ (j m) (j m + 1) m ,m1 , giving us (2j + 1)(2j + 1) matrices
satisfying the three commutation relations [J3 , J ] = ~J and [J+ , J ] = 2~J3 .
If J are angular momentum operators which act on a vector space V and we have
|i V such that J3 |i = ~k|i and J+ |i = 0 then is a state with angular
n
momentum j = k. The other states are given by J |i, 1 n 2k. The conditions
2 2
also give J |i = ~ k (k + 1) |i

1
5.1.1 Spin 2
particles
This is the simplest non-trivial case. We have j = 12 and a two dimensional state space
with a basis | 21 12 i and | 12 21 i. We have the relations J3 | 12 21 i = 21 ~| 12 21 i and

J+ | 12
1
2i =0 J | 21
1
2i = ~| 21 12 i
J | 2 21 i
1
=0 J+ | 2 21 i
1
= ~| 21 1
2 i.

It is convenient to introduce explicit matrices such that


X
J| 12 mi = | 12 m i 21 ~ ()m m .
m

The matrices are 2 2 matrices (called the Pauli spin matrices). Explicitly, they
are

0 2 0 0 1 0
+ = = 3 =
0 0 2 0 0 1

1 0 1 0
1 = (+ + ) = 2 = (+ ) = .
2 1 0 2 0

Note that 12 = 22 = 32 = 1 and = . These satisfy the commutation relations


[i , j ] = 2ijk k (a slightly modified angular momentum commutation relation)
and we also have 2 3 = 1 (and the relations obtained by cyclic permutation), so
5.2. ADDITION OF ANGULAR MOMENTUM 25

2
i j + j i = 2ij 1. Thus if n is a unit vector we have (.n) = 1 and we see that
.n has eigenvalues 1.
1 2 3 2
matrices s =
We define the angular momentum 2 ~ and so s = 4 ~ 1.
1 0
The basis states are 12 = and 12 = .
0 1

5.1.2 Spin 1 particles


We apply the theory as above to get
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.


1 0 0 0 2 0
S3 = 0 0 0 S+ = 0 0 2
0 0 1 0 0 0
and S = S+ .

5.1.3 Electrons
Electrons are particles with intrinsic spin 21 . The angular momentum J = x p + s,
where s are the spin operators for spin 12 .
The basic operators for an electron are x, p and s. We can represent these operators
by their action on two component wavefunctions:
X
(x) = (x) .
= 21

In this basis x 7 x, p 7 ~ and s 7 21 ~. All other operators are constructed


in terms of these, for instance we may have a Hamiltonian

p2
H= + V (x) + U (x).L
2m
where L = x p.
If V and U depend only on |x| then [J, H] = 0.

5.2 Addition of angular momentum


Consider two independent angular momentum operators J(1) and J(2) with J(r) acting
on some space V (r) and V (r) having spin jr for r = 1, 2.
We now define an angular momentum J acting on V (1) V (2) by J = J(1) + J(2) .
Using the commutation relations for J(r) we can get [Ji , Jj ] = ~ijk Jk .
We want to construct states |J M i forming a standard angular momentum basis,
that is such that:

J3 |J M i = ~M |J M i

J |J M i = ~NJ,M |J M 1i


p
with NJ,M = (J M ) (J M + 1). We look first for states in V which satisfy
J+ |J Ji = 0 and J3 |J Ji = ~J|J Ji. The maximum value of J we can get is j1 + j2 ;
and |j1 +j2 j1 +j2 i = |j1 j1 i1 |j2 j2 i2 . Then J+ |j1 +j2 j1 +j2 i = 0. Similarly this is
26 CHAPTER 5. GENERAL THEORY OF ANGULAR MOMENTUM

an eigenvector of J3 with eigenvalue ~ (j1 + j2 ). We can now apply J repeatedly to


form all the |J M i states. Applying J we get

|J M 1i = |j1 j1 1i1 |j2 j2 i2 + |j1 j1 i1 |j2 j2 1i2 .



The coefficents and can be determined from the coefficents Na,b , and we must
2 2
have + = 1. If we choose |i a state orthogonal to this

|i = |j1 j1 1i1 |j2 j2 i2 + |j2 j2 i1 |j2 j2 1i2 .

J3 |i can be computed and it shows that |i is an eigenvector of J3 with eigenvalue


Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

~ (j1 + j2 1). Now

0 = h|j1 +j2 j1 +j2 1i h|J |j1 +j2 j1 +j2 i

and so h|J |i = 0 for all states |i in V . Thus J+ |i = 0 and hence we have |i =


|j1 +j2 1 j1 +j2 1i. We can then construct the states |j1 +j2 1 M i by repeatedly
applying J .
For each J such that |j1 j2 | J j1 + j2 we can construct a state |J Ji. We
define the Clebsch-Gordan coefficients hj1 m1 j2 m2 |J M i, and so
X
|J M i = hj1 m1 j2 m2 |J M i|j1 m1 i|j2 m2 i.
m1 ,m2

The Clebsch-Gordan coefficients are nonzero only when M = m1 + m2 .


We can check the number of states;
1 +j2
jX 1 +j2
jX n o
2
(2J + 1) = (J + 1) J 2 = (2j1 + 1) (2j2 + 1) .
J=|j1 j2 | J=|j1 j2 |

Electrons
Electrons have spin 21 and we can represent their spin states with 12 (s). Using this
notation we see that two electrons can form a symmetric spin 1 triplet

2 (s ) 1 (s )
1
1 2 2


1
m (s1 , s2 ) =
2
1 (s1 ) 1 (s2 ) + 1 (s1 ) 1 (s2 )
2 2 2 2


1 (s ) 1 (s )
2 1 2 2

and an antisymmetric spin 0 singlet;


1
0 (s1 , s2 ) = 12 (s1 ) 21 (s2 ) 12 (s1 ) 21 (s2 ) .
2

5.3 The meaning of quantum mechanics


Quantum mechanics deals in probabilities, whereas classical mechanics is determin-
istic if we have complete information. If we have incomplete information classical
mechanics is also probabilistic.
Inspired by this we ask if there can be hidden variables in quantum mechanics
such that the theory is deterministic. Assuming that local effects have local causes, this
is not possible.
5.3. THE MEANING OF QUANTUM MECHANICS 27

We will take a spin example to show this. Consider a spin 12 particle, with two
spin states |i and |i which are eigenvectors of S3 = 21 ~3 . If we choose to use two
component vectors we have

1 0
= = .
0 1

Suppose n = (sin , 0, cos ) (a unit vector) and let us find the eigenvectors of

cos sin
.n = .
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

sin cos
2
As (.n) = 1 we must have eigenvalues 1 and an inspired guess gives ,n and
,n as

,n = cos 2 + sin 2 and ,n = sin 2 + cos 2 .

Thus (reverting to ket vector notation) if an electron is in a state |i then the prob-
ability of finding it in a state |, ni is cos2 2 and the probability of finding it in a state
|, ni is sin2 2 .
Now, suppose we have two electrons in a spin 0 singlet state;
1
|i = {|i1 |i2 |i1 |i2 } .
2
Then the probability of finding electron 1 with spin up is 12 , and after making this
measurement electron 2 must be spin down. Similarly, if we find electron 1 with spin
down (probability 12 again) then electron 2 must have spin up. More generally, suppose
we measure electron 1s spin along direction n. Then we see that the probability for
electron 1 to have spin up in direction n (aligned) is 21 and then electron 2 must be in
the state |, ni2 .
If we have two electrons (say electron 1 and electron 2) in a spin 0 state we may
physically separate them and consider independent experiments on them.
We will consider three directions as sketched. For electron 1 there are three vari-
(1) (1)
ables which we may measure (in separate experiments); Sz = 1, Sn = 1 and
(1)
Sm = 1. We can also do this for electron 2.
(1) (2)
We see that if we find electron 1 has Sz = 1 then electron 2 has Sz = 1 (etc.).
If there exists an underlying deterministic theory then we could expect some prob-
ability distribution p for this set of experiments;

0 p Sz(1) , Sn(1) , Sm
(1)
, Sz(2) , Sn(2) , Sm
(2)
1
(1) (2)
which is nonzero only if Sdirn = Sdirn and
X
p({s}) = 1.
{s}

Bell inequality
Suppose we have a probability P distribution p(a, b, c) with a, b, c = 1. We define
partial probabilities pbc (b, c) = a p(a, b, c) and similarly for pac (a, c) and pab (a, b).
Then
28 CHAPTER 5. GENERAL THEORY OF ANGULAR MOMENTUM

pbc (1, 1) = p(1, 1, 1) + p(1, 1, 1)


p(1, 1, 1) + p(1, 1, 1) + p(1, 1, 1) + p(1, 1, 1)
pab (1, 1) + pac (1, 1).

Applying this to the two electron system we get


P Sn(1) = 1, Sm
(2)
= 1 P Sz(1) = 1, Sn(2) = 1 + P Sz(1) = 1, Sm
(2)
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

=1 .

We can calculate these probabilities from quantum mechanics



P Sz(1) = 1, Sn(2) = 1 = P Sz(1) = 1, Sn(1) = 1 = cos2
2

P Sz(1) = 1, Sm (2)
= 1 = cos2 +2 and

P Sn(1) = 1, Sm (2)
= 1 = sin2 2 .

The Bell inequality gives sin2


2 cos2
2 + cos2 +
2 which is not in general true.
References
Copyright 2004 University of Cambridge. Not to be quoted or reproduced without permission.

P.A.M. Dirac, The Principles of Quantum Mechanics, Fourth ed., OUP, 1958.
I enjoyed this book as a good read but it is also an eminently suitable textbook.

Feynman, Leighton, Sands, The Feynman Lectures on Physics Vol. 3, Addison-


Wesley, 1964.
Excellent reading but not a very appropriate textbook for this course. I read it as a
companion to the course and enjoyed every chapter.

E. Merzbacher, Quantum Mechanics, Wiley, 1970.


This is a recommended textbook for this course (according to the Schedules). I wasnt
particularly impressed but you may like it.

There must be a good, modern textbook for this course. If you know of one please
send me a brief review and I will include it if I think it is suitable. In any case, with
these marvellous notes you dont need a textbook, do you?

Related courses
There are courses on Statistical Physics and Applications of Quantum Mechanics in
Part 2B and courses on Quantum Physics and Symmetries and Groups in Quantum
Physics in Part 2A.

29

Das könnte Ihnen auch gefallen