PII - Principles of Quantum Mechanics - Horgan (2014) 87pg
PII - Principles of Quantum Mechanics - Horgan (2014) 87pg
PII - Principles of Quantum Mechanics - Horgan (2014) 87pg
Michaelmas 2014
Prof. R.R. Horgan
Contents
1 Introduction 1
2 Dirac Formalism 2
2.1 States and Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.2 Observables and measurements . . . . . . . . . . . . . . . . . . . . . . 4
2.3 Time evolution and the Schrödinger Equation . . . . . . . . . . . . . . 7
2.4 Bases and Representations . . . . . . . . . . . . . . . . . . . . . . . . 9
2.5 Position and momentum basis – wavefunctions . . . . . . . . . . . . . . 11
2.6 Simultaneous Measurements and Complete Commuting Sets . . . . . . 15
6 Perturbation Theory 36
6.1 The non-degenerate case . . . . . . . . . . . . . . . . . . . . . . . . . . 36
i
CONTENTS ii
7 Angular Momentum 43
7.1 Recap of orbital angular momentum . . . . . . . . . . . . . . . . . . . 43
7.2 General analysis of angular momentum eigenstates . . . . . . . . . . . 45
7.3 Matrix representations . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
7.4 Some physical aspects of angular momentum and spin . . . . . . . . . . 50
7.5 Addition of angular momentum . . . . . . . . . . . . . . . . . . . . . . 52
10 Quantum Basics 77
10.1 Classical and Quantum Data . . . . . . . . . . . . . . . . . . . . . . . . 77
10.2 EPR experiment and Bell’s inequality . . . . . . . . . . . . . . . . . . . 79
10.3 Density operators and hidden sectors . . . . . . . . . . . . . . . . . . . 82
CONTENTS iii
BOOKS
‚ B.H. Bransden and C.J. Joachain Quantum Mechanics, 2nd edition. Pearson 2000
(£50-60 on Amazon)
‚ J. Binney and D. Skinner The Physics of Quantum Mechanics, 3rd edition. Cappella
Archive 2013 (£23.71)
‚ P.A.M. Dirac The Principles of Quantum Mechanics. Oxford University Press 1967
reprinted 1999 (£23.99).
Mot of these books are expensive new and there are a lot of pedagogic books on
quantum mechanics, so it’s good to look at those in the library since many books are
good in one area but poor in another. Other books may be recommeneded through
the course.
1 INTRODUCTION 1
1 Introduction
‹ Recall features of elementary (IB) quantum mechanics:
‹ wave-particle duality. Waves behaving like particles – e.g., light quanta, pho-
tons and vice-versa; interference of electrons passing through crystal grating and
electron microscope. To make this more precise need:
‹ wavefunction ψpxq for particle. Probability density |ψpxq|2 ; probability is in-
trinsic to the theory.
‹ 0bservables become (hermitian) operators on wavefunctions. Lack of
commutation limits simultaneous measurement – leads to precise version of un-
certainty principle.
‹ Schrödinger’s equation specifies dynamics (evolution in time) and determines
energy levels.
This is enough to understand e.g., the hydrogen atom and transcends classical
physics.
‹ Aim of this course:
‹ Will not dwell on applications in any detail, but will keep track of what the mathe-
matical formalism is for.
‹ Assume IB QM and IA Dynamics but no electromagnetism beyond Coulomb’s law
and intuitive ideas about magnetism.
Plan:
1. Dirac formalism.
2. Harmonic oscillator.
3. Pictures of quantization.
4. Composite systems and identical particles.
5. Perturbation theory.
6. Angular momentum.
7. Transformations and symmetries.
8. Time-dependent perturbation theory.
9. Quantum basics.
2 DIRAC FORMALISM 2
2 Dirac Formalism
2.1 States and Operators
A quantum state is described at each instant by a state |ψy which belongs to a complex
vector space V . Then
moon , loo|ψy
looxφ| moon ÞÑ xφ|ψy
loomoon or formally V:ˆV Ñ C, (2.1.2)
‘bra’ ‘ket’ ‘bra(c)ket’
with
´ ¯
xφ| α1 |ψ1 y ` α2 |ψ2 y “ α1 xφ|ψ1 y ` α2 xφ|ψ2 y ,
´ ¯
β1 xφ1 | ` β2 xφ2 | |ψy “ β1 xφ1 |ψy ` β2 xφ2 |ψy , (2.1.3)
V ÐÑ V :
with |ψy ÐÑ xψ| “ p |ψyq: (use same label for corresponding states)
and α|ψy ` β|φy ÐÑ α˚ xψ| ` β ˚ xφ| .
(2.1.4)
The inner product is
V ˆV Ñ C
|φy, |ψy ÞÑ xφ|ψy “ p|φyq: |ψy , (2.1.5)
This means that the inner product is positive semidefinite. Note that knowing xφ|ψy
for all xφ| determines |ψy uniquely and vice-versa.
The physical content of any state is unaltered by changing |ψy Ñ α|ψy pα ‰ 0q. We
shall usually normalize states by k|ψyk2 “ 1 but still have the freedom to change
|ψy Ñ eiθ |ψy. The absolute phase of a single state never has any physical significance,
2 DIRAC FORMALISM 3
but relative phases in combination such as α|φy ` β|ψy can be significant; for example,
for interference phenomena.
The space V is complete; we assume appropriate sequences or series converge. A
complete inner product space of this kind is a Hilbert space and this term is often
used for the space V in QM. V can be either finite or infinite dimensional and we shall
see examples of both.
An operator Q is a linear map on states, V Ñ V :
and, by definition ´ ¯
Q α|φy ` β|ψy “ αQ|φy ` βQ|ψy . (2.1.8)
The same operator can be regarded as acting ‘to the left’ on dual states, V : Ñ V : :
or, equivalently,
´ ¯:
xφ|Q: |ψy “ Q|φy |ψy
“ xψ|Q|φy˚ @ |ψy, |φy . (2.1.12)
´ ¯:
xψ|pABq: |φy “ pABq|ψy |φy defn of pABq:
´ ¯:
“ A|ψ 1 y |φy |ψ 1 y ” B|ψy
“ xψ 1 |A: |φy defn of A:
´ ¯: ´ ¯
“ B|ψy A: |φy
“ xψ|B : A: |φy defn of B : .
(2.1.14)
Q: “ Q . (2.2.1)
Such operators are called observables because they correspond to physical, measur-
able, quantities e.g., position, momentum, energy, angular momentum. Key results for
any hermitian Q:
(i)
Q|ψy “ λ|ψy
and xψ|Q: “ λ˚ xψ|
(2.2.2)
ñ xψ|Q “ λ˚ xψ| since Q is hermitian
ñ xψ|Q|ψy “ λxψ|ψy “ λ˚ xψ|ψy .
But k|ψyk2 “ xψ|ψy ‰ 0 p|ψy ‰ 0q and so deduce
λ “ λ˚ . (2.2.3)
2 DIRAC FORMALISM 5
(ii) Let |ny be eigenstates of Q with eigenvalues λ “ qn real, with n a discrete label
possibly of infinite range.
Q|ny “ qn |ny
and Q|my “ qm |my
(2.2.4)
or xm|Q “ qm xm|
ñ xm|Q|ny “ qn xm|ny “ qm xm|ny .
So qn ‰ qm ñ xm|ny “ 0.
‹ For any observable Q there is an orthonormal basis of eigenstates t|nyu for the space
of states V with
Q|ny “ qn |ny ,
xm|ny “ δmn . (2.2.5)
where αn “ xn|ψy.
k|ψyk2 “ xψ|ψy “ 1
´ÿ ¯´ ÿ ˘ ÿ
ðñ ˚
αm xm| αn |ny “ |αn |2 “ 1 . (2.2.7)
m n n
There might be several states with the same eigenvalue λ. Define the eigenspace for
a given eigenvalue by
Vλ “ t|ψy : Q|ψy “ λ|ψyu , (2.2.8)
which has the basis t|ny : qn “ λu.
The degeneracy of λ is the number of states in this basis, or dim Vλ . We say that λ
is non-degenerate if the degeneracy is 1.
r Note that passing from our three key results to the conclusion p‹q is achieved by
choosing an orthonormal basis for each Vλ :
(ii) ensures that these spaces are mutually orthogonal;
(iii) implies that the sum of all the eigenspaces is V , the entire space of states. u
Consider a measurement of Q when the system is in state |ψy immediately before.
Then
In this example we had degeneracy: two states with eigenvalue 0. However, often have
the case with λ non-degenerate with eigenstate |ny unique up to a phase. Then
In general, ÿ ÿ
ppλq “ |αn |2 “ 1 , (2.2.13)
λ n
r The process of measurement is still a source of some deep questions about the inter-
pretation of QM u.
Quantum mechanical behaviour arises from the fact that observables do not commute
in general. In any state |ψy
1
x∆Ayψ x∆Byψ ě |xrA, Bsyψ | , (2.2.17)
2
so rA, Bs ‰ 0 means we cannot expect to measure exact values for A and B simulta-
neously. This generalized Uncertainty Principle follows from
The LHS is a quadratic in λ and the condition implies that the discriminant is ď 0;
the stated Uncertainty Principle then follows.
Paradigm example: position, x̂, and momentum, p̂, in one dimension obey
rx̂, p̂s “ i~
~
ñ ∆x ∆p ě . (2.2.19)
2
In D “ 3, x̂i and p̂i obey
rx̂i , p̂j s “ i~δij , (2.2.20)
and so the uncertainty principle applies to components of position and momentum
which are not orthogonal.
so that the normalization of |ψptqy, and hence the probabilistic interpretation, is pre-
served in time.
H is an observable: the energy. Consider the eigenstates
Example. Consider system with two energy eigenstates |1y, |2y with energy eigenvalues
E1 , E2 , respectively. We are interested in measuring Q defined by
Let the initial state, the state at t “ 0, be |ψp0qy “ |`y. Then have
1 ´ ¯
|ψptqy “ ? e´iE1 t{~ |1y ` e´iE2 t{~ |2y . (2.3.10)
2
The probability of measuring Q at time t and getting ˘1 is
ˇ ´
ˇ1 ¯´ ¯ˇˇ2
2 ´iE1 t{~ ´iE2 t{~
|x˘|ψptqy| “ ˇ x1| ˘ x2| e
ˇ |1y ` e |2y ˇˇ
2
ˇ ´
ˇ 1 ´iE t{~ ¯ˇˇ2
´iE t{~
“ ˇˇ e 1 ˘ e 2 ˇ
2 ˇ
$ ´ ¯
2 pE1 ´E2 qt
& cos
’
’ 2~
“ ´ ¯ (2.3.11)
’
% sin 2 pE1 ´E2 qt
’
2~
where Amn “ xm|A|ny are the matrix elements of the complex matrix representing
the operator A in this basis. Note that the entries in this matrix depend on the basis; a
familiar result in linear algebra for any linear map. In contrast, the result of operating
with A on any state is independent of the basis. Check this result
ÿ
|φy “ A|ψy
loooooomoooooon ô β m “ Amn αn . (2.4.4)
n
basis-independent
loooooooooomoooooooooon
basis-dependent
Clearly, this representation is multiplication of a vector by a matrix: β “ Aα. Also,
have that the Hermitian conjugate has the familiar matrix form:
setting
f pQq|ny “ f pqn q|ny (2.4.9)
defines f pQq provided f pqn q is defined for all n; f pQq is defined on a basis and so is
defined on any state. This is certainly true if f is a polynomial or a power series that
converges for all qn . If qn ‰ 0 @ n then can define
1
Q´1 |ny “ |ny , (2.4.10)
qn
The notation is ´ ¯ ´ ¯
|nyxm| |ψy
loomoon loomoon “ loomoon loomoon .
|ny xm|ψy (2.4.12)
operator state state number
This is confirmed by applying each side to an arbitrary state
ÿ
|nyxn|ψy “ |ψy “ I|ψy . (2.4.13)
n
In the case where the eigenvalues are degenerate then we can define a projection oper-
ator onto the subspace of eigenstates with eigenvalue λ by
ÿ
Pλ “ |nyxn| . (2.4.15)
n: qn “λ
The bases considered so far may be infinite but have been assumed discrete which
includes countably infinite bases. However, we can extend the index n to be continuous.
This requires some modifications in all relevant formulas:
$ ş
& |ψy “ ş dn αn |ny
ÿ ż
Ñ dn I “ ş dn |nyxn|
Q “ dn qpnq|nyxn|
n
%
xn|my “ δnm Ñ δpn ´ mq (2.4.16)
with |αn |2 “ |xn|ψy|2 . There is no longer a probability for discrete outcomes but a
probability density for the continuous range of n. We will see this below for position
and momentum operators.
2 DIRAC FORMALISM 11
So, in particular, ż
2
k|ψyk “ xψ|ψy “ dx |ψpxq|2 “ 1 (2.5.7)
It is very important that the eigenstates of x̂ and p̂ can be chosen so that they are
related by
1
xx|py “ ? eipx{~ , (2.5.10)
2π~
1
ñ xp|xy “ ? e´ipx{~ . (2.5.11)
2π~
2 DIRAC FORMALISM 12
We justify this later after deducing some consequences. First find action of x̂ and p̂ in
terms of position wavefunctions:
x̂|ψy wavefunction : xx|x̂|ψy “ xxx|ψy “ xψpxq
p̂|ψy wavefunction : xx|p̂|ψy
ż
“ dp xx|p̂|pyxp|ψy [resolution of identity using p-states]
ż
“ dp pxx|pyxp|ψy
ż
B ´ ¯
“ dp ´ i~ xx|py xp|ψy
Bx
ż
B
“ ´i~ dp xx|pyxp|ψy
Bx
B B
“ ´i~ xx|ψy “ ´ i~ ψpxq . (2.5.12)
Bx Bx
So
B
xx|p̂|ψy “ ´ i~ xx|ψy, (2.5.13)
Bx
and have recover familiar results. However, also have new possibility. Can expand
states in momentum basis instead:
ż
|ψy “ dp ψ̃ppq|py ,
N.B. „ ˆ ˙ ˆ ˙
B B
xx|rx̂, p̂s|ψy “ x ´i~ ` i~ x ψ “ i~ψ , (2.5.19)
Bx Bx
which verifies the commutation relation.
r The transforms between x and p space are familiar but here we are deriving all the
results, including the transform inversion theorem, on the assumption that t|xyu and
t|pyu are bases. u
The corresponding representations of the Hamiltonian
p̂2
Hpx̂, p̂q “ ` V px̂q (2.5.20)
2m
are
~2 B 2
on ψpxq : H ÝÑ ´ ` V pxq ,
2m Bx2
ˆ ˙
p2 B
on ψ̃ppq : H ÝÑ ` V i~ . (2.5.21)
2m Bp
It may be easy to interpret the potential term in momentum space. E.g., V pxq “
λxn ñ ˆ ˙
B Bn
V i~ “ λpi~qn n , (2.5.22)
Bp Bp
but more generally need to use first principles.
ż
xp|V px̂q|ψy “ dx xp|V px̂q|xyxx|ψy
ż ż
“ dx V pxqxp|xy dp1 xx|p1 yxp1 |ψy
ż ˆ ż ˙
1 1 ´ipp´p1 qx{~
“ dp dx V pxq e ψ̃pp1 q
2π~
ż
1
“ ? dp1 Ṽ pp ´ p1 qψ̃pp1 q . (2.5.23)
2π~
Thus H|ψy “ E|ψy becomes
~2 B 2 ψ
´ ` V pxq ψpxq “ E ψpxq in position space ,
2m Bx2
ż
p2 ? 1 1 1 1
2m ψ̃ppq ` 2π~ dp Ṽ pp ´ p q ψ̃pp q “ E ψ̃ppq in momentum space .
(2.5.24)
r Note that the convolution theorem derived here. u
Now return to the key condition in Eq. (2.5.10) and justify it:
1
xx|py “ ? eipx{~ . (2.5.25)
2π~
2 DIRAC FORMALISM 14
The point is that eigenstates are only ever unique up to a phase, even if normalized,
so we need to show there is a way to choose |xy and |py which makes this result true.
Doing this will involve an approach to translations to which we return later. Claim
that
|x0 ` ay “ e´iap̂{~ |x0 y , (2.5.26)
which involves the translation operator
8 ˆ ˙n
ÿ 1 ´ia
U paq ” e ´iap̂{~
“ p̂n , (2.5.27)
n“0
n! ~
defines position eigenstates |xy @ x given one with x “ x0 . To check this first note
that
rx̂, p̂s “ i~ ñ rx̂, p̂n s “ i~n p̂n´1 . (2.5.28)
r Note that x̂ acts like “ i~ d{dp̂ ” inside a commutator. u Thus find
8 ˆ ˙n
ÿ 1 ´ia
rx̂, U paqs “ i~n p̂n´1 “ aU paq . (2.5.29)
n“1
n! ~
So
Similarly,
|p0 ` by “ eibx̂{~ |p0 y , (2.5.31)
defines momentum eigenstates |py @ p given one with p “ p0 . Then
‚ Since t|xyu is a basis we cannot have xx|p0 y “ 0 for every x, and then Eq. (2.5.26)
implies xx0 |p0 y ‰ 0, the required result, since
‚ Now, the phase of xx0 |p0 y is a matter of convention but the modulus must be con-
sistent with
xp|p1 y “ δpp ´ p1 q , (2.5.35)
which is the desired normalization for the t|pyu basis. To check:
ż
1
xp|p y “ dx xp|xyxx|p1 y
ż
1 ipp1 ´pqx{~
“ dx e “ δpp ´ p1 q as required. (2.5.36)
2π~
‚ Similarly,
ż
1
xx|x y “ dp xx|pyxp|x1 y
ż
1 ippx´x1 q{~
“ dp e “ δpx ´ x1 q as expected. (2.5.37)
2π~
Note that the operator U paq implements translation by a on the position states.
holds for the restriction |ψy, |φy P Vλ . Hence, D a basis for Vλ consisting of
eigenstates of Q1 . Call these |λ, λ1 y.
‚ Doing this for each Vλ gives a basis of such joint eigenstates for V .
|λ1 , λ2 , λ3 , . . .y . (2.6.6)
A “ f pQ1 , Q2 , Q3 , . . .q (2.6.7)
An example is the generalization from one to three dimensions of the position and
momentum operators px̂, p̂q. These obey the commutation relations defined in terms
of their Cartesian component operators in usual notation
There are other possibilities such as tx̂1 , x̂2 , p̂3 u leading to mixed position and momen-
tum space wavefunctions.
Also have
mω 2 1 i
aa: “ x̂ ` p̂2 ´ px̂p̂ ´ p̂x̂q
2~ 2mω~ 2~
1 1
“ H ` ,
~ω 2
and similarly
1 1
a: a “ H ´ (opposite sign for commutator),
~ω 2
(3.1.4)
Thus,
λ ě 0 all eigenvalues non-negative
(3.1.8)
“ 0 iffi a|λy “ 0 .
Next consider commutators
rN, a: s “ ra: a, a: s “ a: ra, a: s “ a: ,
(3.1.9)
rN, as “ ra: a, as “ ra: , asa “ ´a .
These relations imply that a: and a act on eigenstates by respectively raising and
lowering the eigenvalues by 1, provided the new states are non-zero and so actually are
eigenstates.
To find whether the new states are non-zero we compute their norms.
which is never zero since λ ě 0. Because of these properties a: and a are called,
respectively, creation and annihilation operators.
Suppose there is an eigenstate |λy with λ not an integer. Then
From calculations of norms above, we can choose normalized eigenstates |ny, xn|ny “ 1
which are then related by
? *
a: |ny “ ?n ` 1 |n ` 1y
ladder operators (3.1.17)
a|ny “ n |n ´ 1y
2
a|0y “ 0 , (3.1.18)
a a
1
have normalized eigenstates
0
1
|ny “ ? pa: qn |0y . (3.1.19)
n! 0 state
3 THE HARMONIC OSCILLATOR 20
In the absence of any internal structure can take tx̂u or tp̂u or tN u as a complete
commuting set. Then the energy levels are non-degenerate (eigenvalues of N label
them uniquely) and, in particular, |0y is completely specified by
a|0y “ 0 . (3.1.20)
If there is some internal structure then all states can carry an additional label i as-
sociated with some observable Q (or its eigenvalues) commuting with x̂, p̂, a, a: , N .
All energy levels have the same degeneracy with states |n; iy related by a, a: without
affecting i.
The analysis above is convenient for finding wavefunctions. In the position represen-
tation
|0y ù ψ0 pxq “ xx|0y
´ ¯1{2 ´ ¯ ´ ¯1{2 ´ ¯
a “ mω i p̂
x̂ ` mω ù mω ~ B
x ` mω
2~ 2~ Bx
´ ¯ (3.1.21)
a|0y “ 0 ù ~ B ψ pxq “ 0
x ` mω 0
Bx
2
´ mωx
ùñ ψ0 pxq “ N e 2~
¯1{4
´
with normalization factor N “ π~ mω .
Can also find wavefunctions for higher energy states by using Eq. (3.1.19). E.g.,
´ ¯1{2 ´ ¯
|1y “ a: |0y ù ψ1 pxq “ mω ~ B ψ pxq
x ´ mω 0
2~ Bx
(3.1.22)
´ ¯1{2
“ 2mω xψ0 pxq .
~
The correct normalization is guaranteed.
‚ The oscillator is the simplest QM model beyond steps, wells etc. that can be solved
exactly; the hydrogen atom with a Coulomb potential is also special in this respect. It
is a very useful example to use as test case for new ideas, approaches and techniques.
‚ Consider a smooth potential V pxq with equilibrium point x0 (V 1 px0 q “ 0). For
displacements x from equilibrium
1
V px0 ` xq “ V px0 q ` V 2 px0 qx2 ` Opx3 q , (3.2.1)
2
and so if the displacements are not too large neglecting the Opx3 q contribution may
be a good approximation. Indeed, can include the effects of these anharmonic cor-
rections systematically using perturbation theory (see later). The point is that we
start with a soluble model. E.g., diatomic molecules where the quantization of vibra-
tional energies is important in understanding the internal energy and hence the heat
capacity of the gas – has macroscopic consequences. In other systems this approach
can breakdown, though.
‚ More complicated systems can be analyzed in terms of normal modes: each mode is
a coherent motion in which all degrees of freedom oscillate with common frequency ω.
This is common classically and can now quantize this motion. The general solution
for the classical oscillator is
‚ Benzene ring with 6 CH units which oscillate around the “clock face” of the ring.
They are treated as if joined by identical springs. Actually, analyzed by discrete
group theory based on the symmetries of the ring.
‚ Crystal with # atoms N „ 1023 . The forces between the atoms are approximately
elastic and in 3D there are N “ 3N independent coordinates. Each of the 3N modes
is a collective motion of the atoms and if the approximation of elastic forces is good
then interaction between normal modes is small. If you excite just a single mode then
no other mode starts up – no energy transfer between modes; they are effectively
independent oscillators.
In fact,
ω
upxq “ eik¨x with polarization K k and |k| “ . (3.2.5)
c
This gives a wave solution with behaviour e˘ipk¨x´ωtq . General solution is a linear
combination of normal modes for various ω, , k – exact for EM field.
4 PICTURES AND OPERATORS 22
U U : “ U :U “ 1 or U : “ U ´1 . (4.1.1)
|ψy Ñ
Þ |ψ 1 y “ U |ψy
xψ| Ñ Þ xψ 1 | “ xψ|U : , (4.1.2)
and on operators
A ÞÑ A1 “ U AU : , (4.1.3)
under which all physical properties are unchanged:
Furthermore
C “ AB ÞÑ C 1 “ A1 B 1
C “ rA, Bs ÞÑ C 1 “ rA1 , B 1 s for any operators.
General results for unitary operators (compare with those for hermitian operators)
(i)
U |ψy “ λ|ψy
ñ xψ|U : “ λ˚ xψ|
(4.1.7)
ñ xψ|U : U |ψy “ k|ψyk2 “ |λ|2 k|ψyk2 ,
and hence |λ| “ 1 p|ψy ‰ 0q .
(ii)
U |ny “ λn |ny
and U |my “ λm |my
(4.1.8)
or xm|U : “ λ˚m xm| “ λ´1
m xm|
ñ xm|U : U |ny “ xm|ny “ λn λ´1m xm|ny .
So λn ‰ λm ñ xm|ny “ 0.
Thus far we have worked in the Schrödinger picture where states depend on time
and operators do not. We can use U ptq to pass to the Heisenberg picture where the
time dependence is shifted from states to operators as follows (subscript denotes the
picture)
Schrödinger Heisenberg
all physical predictions are the same in either picture. Note that HH “ HS “ H.
The Heisenberg picture makes QM look a little more like classical mechanics where
position, momentum etc. are the variables that evolve in time. To specify the dynamics
in the H-picture we now need an equation to tell us how operators evolve in time. In
the S-picture the Schrödinger equation tells us how states evolve. Now
d d ` itH{~ ˘
AH ptq “ e AS e´itH{~
dt dt
iH itH{~ iH
“ e AS e´itH{~ ´ eitH{~ AS e´itH{~
~ ~
i
“ rH, AH ptqs . (4.2.8)
~
or
E.g., a particle in one dimension x̂ptq, p̂ptq in Heisenberg picture (drop H subscripts).
We have that
rx̂ptq, p̂ptqs “ i~. (4.2.9)
4 PICTURES AND OPERATORS 25
d x̂ptq “ 1 p̂ptq
, d2 x̂ ` ω 2 x̂ “ 0
dt m dt2
/
.
ñ (4.2.13)
d p̂ptq “ ´mω 2 x̂ptq / 2
- d p̂ ` ω 2 p̂ “ 0 .
dt dt2
The solution is
p̂p0q
x̂ptq “ x̂p0q cos ωt ` sin ωt
mω
The final step in Dirac’s systematic approach to QM: have seen how to incorporate
position and momentum wavefunctions and S and H pictures in a single logical frame-
work. But how do we pass from general classical system to its quantum version? In
particular, what are the fundamental quantum commutation relations between observ-
ables ; why rx̂, p̂s “ i~?
Any classical system can be described by a set of generalized positions xi ptq and mo-
menta pi ptq with 1 ď i ď N (may include angles, angular momentum etc.) and a
Hamiltonian Hpxi , pi q.
In classical dynamics a fundamental idea is that of the Poisson bracket of any two
functions f pxi , pi q and gpxi , pi q, say, which is defined to be
ÿ ˆ Bf Bg Bg Bf
˙
tf, gu “ ´ , (4.3.1)
i
Bxi Bpi Bxi Bpi
which is a new function of xi and pi . (pxi , pi q are coordinates on phase space and PB
is a symplectic structure.) In particular,
‚
classical f, g ÝÑ quantum fˆ, ĝ (4.3.4)
functions operators
In particular, get
rx̂i , p̂j s “ i~δij (4.3.6)
which are the canonical commutation relations.
For those taking IIC Classical Dynamics this relationship between classical and quan-
tum mechanics should be mentioned near the end of the course.
All this provides a sound basis for understanding classical mechanics as a limit of quan-
tum mechanics with ~ Ñ 0. Going the other way, turning ~ “on” is more problematic
and not guaranteed to be either unique or, in some cases, even consistent. For exam-
ple, if we carry out the procedure above it is correct to Op~q but there may be Op~2 q
ambiguities related to how operators are ordered in defining functions like f pxi , pj q:
does xi multiply pi on left or right?
Alternative approach to quantization is to use path integrals which are sums of contri-
butions from all possible trajectories or paths between initial and final configurations
in phase space.
One of these is the classical trajectory or path. It is derived from an action principle:
the path that minimizes the action associated with the path. However, the quantum
amplitude involves contributions from all trajectories. This approach has its advan-
tages but, in principle, is equivalent to canonical quantization. In general need both,
especially for complicated systems where there are constraints amongst the variables.
V “ V1 b V2 (5.1.2)
5 COMPOSITE SYSTEMS AND IDENTICAL PARTICLES 28
consists of all linear combinations of tensor product states |ψy b |φy (duals xψ| b xφ|)
subject to
´ ¯
|ψy ` |ψ 1 y b |φy “ |ψy b |φy ` |ψ 1 y b |φy
´ ¯
1
|ψy b |φy ` |φ y “ |ψy b |φy ` |ψy b |φ1 y
´ ¯ ´ ¯ ´ ¯
α|ψy b |φy “ |ψy b α|φy “ α |ψy b |φy , (5.1.3)
A ÐÑ A b I acting just on V1
B ÐÑ I b B acting just on V2 . (5.1.6)
Consider a particle in two dimensions with position operators x̂1 , x̂2 . Basis of joint
eigenstates can be constructed as
This is the V “ V1 b V2 tensor product of states for two one-dimensional particles. The
wavefunction for |ψy b |φy is
´ ¯´ ¯
xx1 | b xx2 | |ψy b |φy “ xx1 |ψyxx2 |φy
“ ψpx1 qφpx2 q . (5.1.8)
5 COMPOSITE SYSTEMS AND IDENTICAL PARTICLES 29
5.2 Spin
Experiment shows that particles generally carry an internal degree of freedom called
spin or intrinsic angular momentum. Even if the particle appears ‘elementary’ or
pointlike, its space of states will be of the form V “ Vspace b Vspin with basis
|x, ry “ |xy|ry , (5.2.1)
where r takes a finite set of values: the quantum numbers associated with spin. The
particle is not ‘structureless’: the position operators, x̂, are not a complete commut-
ing set by themselves – there are additional observables Q acting just on Vspin with
rx̂i , Qs “ 0. We will understand these operators later in the study of angular momen-
tum but for now concentrate on the states.
Each kind of particle has a definite total spin S which is a half-integer 0, 21 , 1, 32 , . . .;
this is a basic characteristic like its mass or charge. For a spin S particle of non-zero
mass there are 2S ` 1 basis states in Vspin labelled by convention r “ S, S ´ 1, . . . , ´S.
E.g.,
S basis states
0 |0y
1
| 12 y, | ´ 12 y also written Òyon, loo|mo (5.2.2)
2 loo|mo Óyon
up down
1 |1y, |0y, | ´ 1y
The existence of spin states is revealed by e.g. the Stern-Gerlach experiment
5 COMPOSITE SYSTEMS AND IDENTICAL PARTICLES 30
beam of
splits beam in two:
s = 1/2 in this case
particles
constructed from single particle states. If the particles are identical, Va u V , some-
thing interesting can be added.
Consider the simplest case N “ 2. Define an operator W which exchanges particles by
its action on basis states:
When the two particles are identical its action on a general 2-particle state is
because |Ψy and W |Ψy must be physically equivalent if the particles are indistin-
guishable. But, given its action on the basis states
W 2 “ 1 ñ η 2 “ 1 or η “ ˘ 1 . (5.3.4)
5 COMPOSITE SYSTEMS AND IDENTICAL PARTICLES 31
Similarly, for multiparticle states with N ě 2 we can define Wpa,bq which exchanges
pxa , ra q Ø pxb , rb q by this action on the basis states. Then for a general N -particle
state
Wpa,bq |Ψy “ ηpa,bq |Ψy (5.3.6)
2
with, again, ηpa,bq “ ˘1 because Wpa,bq “ 1.
For any permutation π of t1, 2, . . . , N u define
Wπ |x1 , r1 ; x2 , r2 ;. . . ; xN , rN y
“ |xπp1q , rπp1q ; xπp2q , rπp2q ; . . . ; xπpN q , rπpN q y (5.3.7)
on the basis states. On a general state
Wπ |Ψy “ ηπ |Ψy for some ηπ . (5.3.8)
But algebra of swaps or transpositions implies ηpa,bq “ ˘ 1 with the same value for
all pairs pa, bq since any two swaps are conjugate. This makes physical sense since the
particle are identical and the initial choice for the labelling is not unique. Then, since
any π can be obtained as a sequence of swaps, we have alternative outcomes
#
1
ηπ “ (5.3.9)
sgnpπq “ p´1qp# swaps needed for πq ,
with the same alternative for all π. These correspond to two inequivalent 1-D repre-
sentations of the permutation group.
Hence, there are two fundamentally different kinds of particles:
Note that this applies only to identical particles. Indistinguishability has a different
character in QM from classical physics. It is the consequence of saying that you cannot
attach a label to a given particle and uniquely identify it from any other. You can no
longer follow individual particles because of the uncertainly principle.
In addition have the remarkable ‹ Spin-statistics relation. ‹
5 COMPOSITE SYSTEMS AND IDENTICAL PARTICLES 32
‚ Most common elementary particles are fermions: electrons, protons, neutrons, neu-
trinos, quarks, muons, τ – all spin 12 .
‚ Particles associated with forces are bosons: photons (EM), W ˘ , Z (weak nuclear),
gluons (strong nuclear) – all spin 1.
‚ Other particles such as mesons are bosons e.g., π, K are spin 0, the ρ is spin 1, and
many more have been observed with higher spin.
‚ The recently discovered Higgs boson (LHC experiments) has almost certainly spin 0
although this is still to be confirmed.
‚ The graviton has spin 2 but is yet to be observed - not likely in the near future (if
ever).
‚ The spin-statistics theorem applies even if the particles are not ‘elementary’. Indeed,
nucleons (proton etc) and mesons are made of quarks. Atoms obey the theorem, too.
with the wavefunctions for the basis of energy eigenfunctions for Vspace b Vspace
Spin 0: Vspin is just one state so can ignore. Spin-statistics ñ the particles are
bosons, and states are in pV b V qS which in this case is pVspace b Vspace qS . The
allowed states are then
*
Ground state ΨS0 px1 , x2 q 2E0 both non-
degenerate (5.5.9)
1st excited state ΨS1 px1 , x2 q E0 ` E1
Note that only one of the two possible 1st excited states is allowed.
Spin 21 : Vspin has basis t| Òy, | Óyu for each particle. Vspin b Vspin therefore has
the basis
t| Òy| Òy, | Òy| Óy, | Óy| Òy, | Óy| Óyu , (5.5.10)
or normalized combinations with definite symmetry are
| Òy| Òy
´ ¯ ´ ¯
?1 | Òy| Óy ` | Óy| Òy ?1 | Òy| Óy ´ | Óy| Òy
2 2
(5.5.11)
| Óy| Óy
The allowed states are (in hybrid notation, see Eq. (5.2.4))
5 COMPOSITE SYSTEMS AND IDENTICAL PARTICLES 34
There are four states in all made up of 1 ` 3 “ 4. The spin quantum numbers
associated with these are S “ 0 and S “ 1, respectively.
We can see that representations of the two-particle permutation group, S2 , are used
extensively in the analysis. The general analysis for N identical particles is based on
the representation theory of SN , the permutation group for N particles.
1 2 Ze2
Hpx̂, p̂q “ p̂ ´ . (5.6.2)
2m 4π0 |x̂|
Single electron states in Vspace are similar to hydrogen atom states (for which Z “ 1)
$
’
’ evals
H En
&
|n, l, my which are joint eigenstates of 2 2 (5.6.3)
’
’ L ~ lpl ` 1q
L3
%
~m
5 COMPOSITE SYSTEMS AND IDENTICAL PARTICLES 35
where
ˆ ˙2
m Ze2 1 2 Z
2
En “ ´ 2
“ ´ α 2
mc2 ,
2 4π0 ~ n 2n
n “ 1, 2, . . . principal quantum number ,
l “ 0, 1, . . . , n ´ 1 m “ 0, ˘1, ˘2, . . . , ˘l ,
looooooooomooooooooon (5.6.4)
2l ` 1 states
where α “ e2 {p4π0 ~cq « 1{137 is the fine-structure constant. So degeneracy (excluding
spin!) is
1 ` 3 ` . . . ` p2n ´ 1q “ n2 for level En . (5.6.5)
Now, including Vspin basis t| Òy, | Óyu, each electron has 2n2 states at energy En .
Structure of electrically neutral atoms with N “ Z:
‚ Fill up energy levels, starting with the lowest, using Pauli Principle.
‚ This gives a rough picture of the periodic table with some qualitative insights into
chemical properties.
‚ The states that belong to a given energy En is called a shell. Atoms with completely
filled shells are unreactive/stable elements chemically. E.g.
E1 filled for Z “ 2 Ñ He
E1 & E2 filled for Z “ 10 Ñ Ne (5.6.6)
atomic n 1 2 3 4
no. element s s p s p d s p d f
1 H 1
2 He 2
6 C 2 2 2
9 F 2 2 5
10 Ne 2 2 6
11 Na 2 2 6 1
19 K 2 2 6 2 6 1
20 Ca 2 2 6 2 6 2
26 Fe 2 2 6 2 6 6 2
28 Ni 2 2 6 2 6 8 2
29 Cu 2 2 6 2 6 10 2
6 PERTURBATION THEORY 36
6 Perturbation Theory
Few quantum mechanical systems can be solved exactly. In perturbation theory we
start from a known, soluble, system
with t|nyu an orthonormal basis of energy eigenstates, and calculate the energies and
eigenstates for a new perturbed Hamiltonian
States t|nyu are still a basis when µ ‰ 0, and so we can always write
ÿ
|ψy “ α|ry ` βj |jy
j‰r
ÿ
“ αp|ry ` γj |jyq , (6.1.2)
j‰r
α Ñ 1, βj , γj Ñ 0 as µ Ñ 0 . (6.1.3)
6 PERTURBATION THEORY 37
and, so far, this is still exact (all orders in µ). Substituting in the series expansions
from Eq. (6.1.4) and keeping terms to Opµ2 q gives
ÿ
µErp1q ` µ2 Erp2q ` . . . “ µxr|V |ry ` µ2 cj1 xr|V |jy ` . . . . (6.1.8)
j‰r
Again this is exact but the 2nd term on RHS is Opµ2 q, and so to leading order in µ
µ ci1 pEr ´ Ei q “ µxi|V |ry
xi|V |ry
ñ ci1 “ (note Ei ‰ Er since states non-degenerate)
Er ´ Ei
(6.1.11)
and so substituting in Eq. (6.1.8) we find
ÿ xr|V |jyxj|V |ry ÿ |xj|V |ry|2
Erp2q “ “ . (6.1.12)
j‰r
Er ´ Ej j‰r
Er ´ Ej
In summary,
ÿ |xj|V |ry|2
2
E “ Er ` µxr|V |ry ` µ ` ... , (6.1.13)
j‰r
Er ´ Ej
˜ ¸
ÿ xj|V |ry
|ψy “ α |ry ` µ |jy ` ... , (6.1.14)
j‰r
Er ´ Ej
6 PERTURBATION THEORY 38
where α is chosen so that |ψy has unit norm. This is second order perturbation theory.
Example:
p̂2
´ ¯
1 2 2
H “ 2m ` 2 mω x̂ “ ~ω a a ` 2 : 1
´ ¯ (6.1.15)
states |ny, En “ ~ω n ` 2 . 1
~ 2
Perturb with V “ mω 2 x̂2 “
2
mω pa2 ` a: ` 2a: a ` 1q. (6.1.16)
2
m
ω
Have
1
xn|V |ny “ 2
~ωp2n ` 1q
1
? ?
xn ` 2|V |ny “ 2
~ω n ` 1 n `2
1
? ?
xn ´ 2|V |ny “ 2
~ω n n ´ 1
xm|V |ny “ 0 all other m (6.1.17)
Indeed, it may be that the series is only asymptotic and more sophisticated methods
are needed to estimate the energy shifts within a given accuracy.
Example. Ground state energy for Helium.
The unperturbed problem is two non-interacting electrons, charge ´e orbiting nucleus
with charge `2e. The Hamiltonian is
p̂2 2e2
Hpx̂1 , p̂1 q ` Hpx̂2 , p̂2 q with Hpx̂, p̂q “ ´ . (6.1.23)
2m 4π0 |x̂|
Single electron states and energies are
ˆ ˙2
|n l my 1 2e2 1 2α2 2
En “ ´ m ” ´ mc , (6.1.24)
wavefn ψnlm pxq 2 4π0 ~ n2 n2
2
where α “ 4πe ~c is the dimensionless fine structure constant: α « 1{137. The lowest
0
energy eigenstate for two electrons is
|Ψy “ |1 0 0y b |1 0 0y b |χy
1 ´ ¯
with |χy “ ? | Òy| Óy ´ | Óy| Òy the spin state . (6.1.25)
2
This state is totally antisymmetric since constrained by Fermi statistics. The two
electron wavefunction is then
Ψpx1 , x2 q “ ψ100 px1 qψ100 px2 q
ˆ ˙ 32
1 2
ψ100 pxq “ ? e´|x|{a2
π a2
ˆ ˙
1 1 ~
with pZ “ 2q a2 “ 2 “ 12 ˚ pBohr radiusq (6.1.26)
α mc
The two-electron unperturbed energy is
2E1 “ ´ 4α2 mc2 « ´ 108.8 eV . (6.1.27)
Compare with -13.6 eV for the hydrogen atom. (Note: mc2 « 500 KeV.)
Experimentally, the ground state for He is -79.0 eV. However, have neglected the
electron-electron interaction:
e2 1
. (6.1.28)
4π0 |x̂1 ´ x̂2 |
6 PERTURBATION THEORY 40
~c
µ “ α, V px̂1 , x̂2 q “ . (6.1.29)
|x̂1 ´ x̂2 |
ż ż
~c
“ α d x1 d3 x2 |ψ100 px1 q|2 |ψ100 px2 q|2
3
|x1 ´ x2 |
5 2 2
“ α mc « 34.0 eV . (6.1.30)
4
The corrected ground state energy is then ´108.8 eV ` 34.0 eV « ´74.8 eV; in much
better agreement with experiment. Note that the variational principle is more efficient
(see AQM) at this level but does not help with higher-order corrections.
We might naively expect that the perturbation series is an expansion in α but we see
p1q
that E1 and the first-order correction E1 are both 9 α2 . In fact, all corrections are
9 α2 , so what is the expansion parameter? Including the 2nd-order correction the
energy can be written 2
ˆ ˙ˇ
2 2 2 51 25 1 ˇ
E “ ´ Z α mc 1 ´ ` ` . . . ˇ . (6.1.31)
8 Z 256 Z 2 ˇ
Z“2
So the expansion is in 1{Z for Z “ 2. The series “looks” convergent and gives an
answer close to experiment but to my knowledge it is not known if it actually converges.
However, treated as an asymptotic series it does give a believable answer. This is typical
of many problems in bound state systems.
In scattering theory (Quantum Electrodynamics) the expansion parameter is α but
convergence is still not provable.
E1 “ . . . “ Er “ Es “ . . . “ EN “ λ degenerate states
(6.2.1)
|1y, . . . |ry, |sy, . . . |N y reserve r, s to label these only
2
This is the correct expression for all atoms with two electrons. For example, Li` which has Z “ 3
and Be2` which has Z “ 4.
6 PERTURBATION THEORY 41
Denote other states, not in this degenerate set, by |jy and reserve j to label these only.
Then, for general µ, expand the eigenfunction of the perturbed Hamiltonian as
E “ λ ` µE p1q ` µ2 E p2q ` . . . ,
ÿ ÿ
|ψy “ αr |ry ` βj |jy .
r j
with αr “ ar ` Opµq, βj “ Opµq . (6.2.3)
Note that for µ Ñ 0, |ψy Ñ |ψ0 y P Vλ but we do not yet know the values of the ar
which determine it uniquely in this limit. We shall see that the possible choices for the
ar are determined by the perturbation itself. Substitute into Eq. (6.2) and find
˜ ¸
ÿ ÿ ÿ ÿ
αr pEr ` µV q|ry ` βj pEj ` µV q|jy “ E αr |ry ` βj |jy . (6.2.4)
r j r j
Rearranging gives
ÿ ÿ ÿ ÿ
αr pE ´ Er q|ry ` βj pE ´ Ej q|jy “ µ αr V |ry ` µ βj V |jy . p˚˚q
r j
loooooooooooooooooooooooomoooooooooooooooooooooooon r j
loooooooooooooooomoooooooooooooooon
terms containing energy shifts terms containing V
(6.2.5)
Look for energy shift E “ λ ` µE p1q ` Opµ2 q and note that the correction is no longer
associated with a single state of the unperturbed problem but with all N states in Vλ .
Take the inner product of xs| with p˚˚q to get (remember Er “ . . . “ Es . . . “ EN “ λ)
ÿ ÿ
pE ´ Es qαs “ µ αr xs|V |ry ` µ βj xs|V |jy , (6.2.6)
r j
but the second term on RHS is Opµ2 q and αr “ ar ` Opµq, and so hence to leading
order ÿ
xs|V |ry ar “ E p1q as . (6.2.7)
r
p1q
Thus, E is an eigenvalue of the matrix xs|V |ry and the ar are the components of
the corresponding eigenvector. Since there are N degenerate states this is an N ˆ N
matrix and so have N solutions for E p1q with associated eigenvectors giving the ar in
each case.
We should expect something like this:
‚ Raising the degeneracy in this way is important in many physical phenomena. For
example, band structure for the electron levels in a crystal giving rise to delocalization
of electrons originally bound in each atom and so leading to electrical conduction
(see AQM).
with
ża ża
xy
α “ dx dy 2 |ψ12 px, yq|2
0 0 a
ˆ ˙2 ż a ˙2
πx ¯2 a
ż ˆ
2 1 ´ 2πy
“ dx x sin dy y sin
a a2 0 a 0 a
1
“ . (6.2.12)
4
7 ANGULAR MOMENTUM 43
ża ża
xy ˚
β “ dx dy 2 ψ12 px, yqψ21 px, yq
0 0 a
ˆ ˙2 ˆ ż a ˙2
2 1 πx 2πx
“ dx x sin sin
a a2 0 a a
ˆ ˙2
16
“ . (6.2.13)
9π 2
(6.2.14)
„ż a ża ˆ ˙
πx 2πx 1 πx 3πx 8a2
dx x sin sin “ dx x cos ´ cos “´ 2 .
0 a a 0 2 a a 9π
Collecting results:
2 2 µ
‚ New ground state energy ~ π2 ` 4 .
ma
‚ Next two levels
5~2 π 2 µ 256
energies 2 ` ˘µ
2ma 4 81π 4
1 ´ ¯
states |ψy “ ? |12y ˘ |21y . (6.2.16)
2
Note that the state is unperturbed at first order in µ but the perturbation does
determine the choice of basis in Vλ in order that the perturbation can be carried
out systematically. This basis is the eigenstates of the matrix with elements xs|V |ry.
7 Angular Momentum
7.1 Recap of orbital angular momentum
Mainly to set the analysis to follow in some sort of context but also a few points of
special importance.
Consider the action of all these operators on wavefunctions. Using spherical polar
coordinates pr, θ, φq, the operators Li only involve angular derivatives and
1 B2 1
∇2 “ 2
r ´ 2 ~2
L2 . (7.1.4)
r Br
loomoon r
loomoon
radial angular
The joint eigenstates of L2 and L3 are the spherical harmonics Ylm pθ, φq:
where ˆ ˙
~2 1 B 2 ~2
´ prRl q ` lpl ` 1q ` V prq Rl “ ERl . (7.1.8)
2m r Br2 2mr2
Of particular importance is the behaviour of such solutions under parity: x ÞÑ ´x.
This is equivalent to
r ÞÑ r, θ ÞÑ π ´ θ, φ ÞÑ φ ` π . (7.1.9)
Then, under parity map have
p̂21 p̂2
H “ ` 2 ` V p|x̂1 ´ x̂2 |q
2m1 2m2
2
P̂ p̂2
“ ` ` V p|x̂|q , (7.1.11)
2M 2m
7 ANGULAR MOMENTUM 45
where
X̂ “ m1 x̂1 M
` m2 x̂2 x̂ “ x̂1 ´ x̂2
m2 p̂1 ´ m1 p̂2
P̂ “ p̂1 ` p̂2 p̂ “ M (7.1.12)
M “ m1 ` m2 m “ mM
1 m2 reduced mass
CoM dynamics essentially trivial and relative motion is governed by a spherically sym-
metric potential. The total wavefunction is then
I.e., plane wave solution in the CoM variables and effective single particle dynamics in
potential V p|x̂|q with reduced mass m.
Under x ÞÑ ´x have Ψ ÞÑ p´1ql Ψ. This is particularly important if the two particles
are identical.
J 2 |ψy “ λ|ψy ñ
2
λ “ xψ|J |ψy “ kJ1 |ψyk2 ` kJ2 |ψyk2 ` kJ3 |ψyk2 ě 0 . (7.2.2)
So far all we know is that m and j ě 0 are real numbers. To analyze the allowed
eigenvalues we define
J˘ “ J1 ˘ iJ2 , J˘: “ J¯ , (7.2.4)
and work with these new combinations. It is easy to check that
rJ3 , J˘ s “ ˘~J˘
rJ` , J´ s “ 2~J3
rJ 2 , J˘ s “ 0 . (7.2.5)
7 ANGULAR MOMENTUM 46
Furthermore, we find
J` J´ “ J 2 ´ J32 ` ~J3
J´ J` “ J 2 ´ J32 ´ ~J3 (7.2.6)
and
j ě m ě ´pj ` 1q
(7.2.12)
j ` 1 ě m ě ´j
and the states vanish only when equality occurs. Hence, from these bounds, we deduce
j ě m ě ´j
J` |j, my “ 0 iffi m “ j , (7.2.13)
J´ |j, my “ 0 iffi m “ ´j .
7 ANGULAR MOMENTUM 47
These results tell us all we need to know about the possible values of both j and m.
Remember, so far only know that m and j ě 0 are real.
Start from any given state |j, my.
Thus,
j “m`k and j “ ´m ` k 1 , k, k 1 integer
ñ 2j “ k ` k 1 , an integer. (7.2.16)
Furthermore, for a given value of j we have states
|j, my with m “ j, j ´ 1, . . . , ´j ` 1, ´j , a total of 2j ` 1 states. (7.2.17)
This general analysis has revealed two possibilities:
j integral m “ 0 , ˘1, ˘2, . . . , ˘j odd # states,
(7.2.18)
j half-integral m “ ˘ 21 , ˘ 23 , . . . , ˘j even # states.
j integral. This possibility is realized in orbital angular momentum. The states |j, my
correspond to wavefunctions ψjm pxq or Yjm pθ, φq.
j half-integral. This possibility cannot arise for orbital angular momentum since
there are no solutions of the differential equations which are well-behaved in this case.
Such states must correspond to intrinsic angular momentum or spin as introduced
earlier.
We usually write J “ S for spin. Our analysis shows that we must have j “ S integral
or half-integral for spin. Previously, we wrote |ry with S ě r ě ´S for spin-states.
Now we see that by this was meant
|ry ” |j, my with j “ S, m “ r. (7.2.19)
7 ANGULAR MOMENTUM 48
Analysis reveals mathematically that spin is possible with these quantum numbers but
still need to give (very brief) indication of experimental verification.
The set of states t|j, myu for fixed j is often called an angular momentum multiplet
or representation. From the analysis above (Eqns. (7.2.10, 7.2.11)) it is clear we can
choose normalized states |j, my with
a
J` |j, my “ ~ pj ´ mqpj ` m ` 1q|j, m ` 1y , (7.2.20)
a
J´ |j, my “ ~ pj ` mqpj ´ m ` 1q|j, m ´ 1y , (7.2.21)
which are key relations between the states. The whole multiplet can be defined by
‚ taking the top state, the one with maximum J3 eigenvalue: J` |j, jy “ 0,
‚ and applying J´n :
|j, j ´ ny “ Cjn J´n |j, jy , (7.2.22)
where Cjn is a constant computable using Eq. (7.2.21).
‚ Alternatively, can start with the bottom state, |j, ´jy and determine the others by
applying J`n .
Note that the choice of J3 as the member of the commuting set along with J 2 is a
convention. We could have chosen n¨J instead. n¨J has the same possible eigenvalues
as J3 but the eigenstates are linear combinations of the basis states t|j, myu. This will
be relevant when we discuss rotations.
This is used very widely for the case of spin- 12 (now use S rather than J ) with just two
states.
ˆ ˙ ˆ ˙
1 1 1 1 1 0
|2, 2y ù , |2, ´2y ù
0 1 (7.3.4)
previously: | Òy | Óy
ˆ ˙ ˆ ˙ ˆ ˙
1 0 0 1 0 0
S3 ù 12 ~ , S` ù ~ , S´ ù ~ . (7.3.5)
0 ´1 0 0 1 0
Moreover, we also write Si ù 12 ~σi where
ˆ ˙ ˆ ˙ ˆ ˙
0 1 0 ´i 1 0
σ1 ù , σ2 ù , σ3 ù , (7.3.6)
1 0 i 0 0 ´1
are called the Pauli matrices. Other combinations are
ˆ ˙ ˆ ˙
0 2 0 0
σ` “ σ1 ` iσ2 “ , σ´ “ . (7.3.7)
0 0 2 0
The Pauli matrices are hermitian, traceless matrices obeying
σ12 “ σ22 “ σ32 “ I , (7.3.8)
σ1 σ2 “ ´σ2 σ1 “ iσ3 and cyclic. (7.3.9)
These properties are conveniently summarized by
σi σj “ δij ` iεijk σk . (7.3.10)
Note that the antisymmetric part (in i, j) of this equation is
rσi , σj s “ 2iεijk σk , (7.3.11)
and corresponds to the fundamental commutation relation
rSi , Sj s “ i~εijk Sk , (7.3.12)
but the remaining, symmetric part, is special to spin- 21 .
The Pauli matrices are components of a vector
1
S ù 2
~σ , σ “ pσ1 , σ2 , σ3 q . (7.3.13)
If a and b are constant vectors (or at most operators which commute with S) then we
can contract ai and bj with both sides of Eqn. (7.3.10) to obtain
pa ¨ σqpb ¨ σq “ pa ¨ bqI ` ipa ^ bq ¨ σ . (7.3.14)
As a special case
pn ¨ σq2 “ I , (7.3.15)
2 1 2
where n is any unit vector. Note that this is equivalent to pn ¨ Sq “ which
4
~ I,
agrees with the fact that the eigenstates of n ¨ S are ˘ 21 ~; these are the only possible
results for measurement of spin along some direction n.
One last example of matrix properties corresponding to known properties of operators:
σ 2 “ σ12 ` σ22 ` σ32 “ 3 I , (7.3.16)
to be compared with
`1˘ `1 ˘ ~2
S 2 “ ~2 2 2
`1 I “ 3 I (7.3.17)
4
~2
which has eigenvalue 3 on any state.
4
7 ANGULAR MOMENTUM 50
J “ L`S . (7.4.2)
together imply
rJi , Jj s “ i~εijk Jk . (7.4.4)
How do we know the world works this way? Results of many experiments confirm it.
Here just mention the theoretical ideas underlying a few of them. The key idea is how
spin and angular momentum degrees of freedom enter into the Hamiltonian.
Main example for us is interaction with a background magnetic field. From classical
EM (no previous knowledge required) a moving distribution of electric charge interacts
with magnetic field Bpxq to produce energy
´µ ¨ B , (7.4.5)
This final result is all that is important for us. Now pass to quantum theory and
consider the effect of including such a term in the Hamiltonian for the electron in the
hydrogen atom (works for any spherically symmetric potential).
(i) Before we turn on any magnetic field have ψnlm pxq joint eigenstates of H, L2 , L3 with
degenerate energies E independent of m (as discussed in section 7.1).
7 ANGULAR MOMENTUM 51
H Ñ H ´ γBL3 , (7.4.8)
but ψnlm pxq are still eigenstates and energies now split
E Ñ E ´ γB~m ´l ďmďl .
looooomooooon (7.4.9)
2l ` 1 levels
This is observed: spectral lines split into distinct lines. This is the Zeeman effect.
(iii) An electron in the atom has both an orbital magnetic moment µL 9L and a spin mag-
netic moment µS 9S. Like two bar magnets near each other there is an interaction
energy
µL ¨ µS 9 L ¨ S . (7.4.12)
This is called spin-orbit coupling and when included in the atomic Hamiltonian it
leads to splitting of spectral lines into doublets because of the spin- 21 of the electron.
The two yellow sodium D lines are a famous outcome. This is called atomic fine
structure.
(iv) Unlike orbital angular momentum, it is sometimes appropriate to consider spin to-
tally divorced from any space degrees of freedom. E.g., electron somehow confined
to one atomic site in a crystal. Then
H “ ´ γS ¨ B (7.4.13)
can be the complete Hamiltonian. If allow Bpzq “ p0, 0, Bpzqq to vary in space
rather than just separating energies can instead physically separate atoms according
to whether their spin states are Ò or Ó. This is due to
d dB ~ dB
force in z-dirn “ ´ p´γS ¨ Bpzqq “ ´ γS3 “ ¯γ (7.4.14)
dz dz 2 dz
This is the Stern-Gerlach experiment.
7 ANGULAR MOMENTUM 52
(v) The proton is spin- 21 and has a magnetic moment. Even if l “ 0 for the electron in
the H atom and no external field there is a weak interaction between the proton spin
I and electron spin S 9I ¨ S. There are two narrowly spaced levels and transitions
give radiation with wavelength « 21cm which is observed from interstellar hydrogen.
(vi) In water the proton magnetic moments of the hydrogen nuclei interact with an
external B-field, 9 ´ I ¨ B to produce two levels where the transition absorbs or
radiates microwaves. This forms the basis of MRI scanners.
Essential idea is that spin or angular momentum behaves like a magnet/dipole. Actu-
ally, an iron bar magnet is magnetized because of the sum of the microscopic electron
spin magnetic dipoles in the atoms which, below the Curie temperature (TC ), prefer to
align and produce the macroscopic magnetic dipole. (For T ą TC the magnetization
vanishes; at T “ TC there is a second-order phase transition.)
The existence of an atomic magnetic moment in iron and also the preference for them
to align and produce a macroscopic magnetization are due to the Pauli Exclusion
Principle (spin-statistics theorem) – a quantum effect. Look up Hund’s rule.
Consider two independent systems with angular momentum operators J p1q , J p2q acting
on spaces of states V p1q , V p2q each with standard basis t|ji , mi yu consisting of joint
piq
eigenstates of pJ piq q2 , J3 for i “ 1, 2.
Construct space of states for the combined system as V “ V p1q b V p2q with basis
(shorthand)
The key idea is to find the top state for a given total angular momentum with
J` |J, M y “ 0 ô J “ M . (7.5.10)
Given M , which is easy to determine, we therefore know J for this state. Then the
others with the same J are found by applying J´ .
7 ANGULAR MOMENTUM 54
J M
| j1 ` j2 , j1 ` j2 y “ |j1 y|j2 y , (7.5.11)
This is called the highest weight state since it is the state with the largest value
of M possible for given J. (This terminology arises in the group theory approach.)
The phase of the state on RHS (P V p1q b V p2q ) is chosen by convention to be `1.
p1q p2q
‚ Apply J´ “ J´ ` J´ to this state on LHS and RHS, respectively. We get
p1q p2q
J´ J´ ` J´
a ? ?
2pj1 ` j2 q |j1 ` j2 , j1 ` j2 ´ 1y “ 2j1 |j1 ´ 1y|j2 y ` 2j2 |j1 y|j2 ´ 1y
(7.5.12)
using standard formulas for V on LHS and for V p1q , V p2q on RHS. The normalized
state is then
d d
j1 j2
|j1 ` j2 , j1 ` j2 ´ 1y “ |j1 ´ 1y|j2 y ` |j1 y|j2 ´ 1y . (7.5.13)
j1 ` j2 j1 ` j2
‚ But we have not yet found all |J, M y states. We started from a unique state with
M “ j1 ` j2 . At the next level down, with M “ j1 ` j2 ´ 1, there are two states:
j1+j2-1
j1+j2-2
-(j1+j2)+1
-(j1+j2)
where
e top states Ó apply J´
unique state K others (7.5.17)
u other states Ñ
with same M value
Can check that top states |ψy with J3 eigenvalue M are annihilated by J` directly.
This is also guaranteed by
Thus,
xJ, M ` 1|J` |ψy “ 0 @ J ą M ô J` |ψy “ 0 . (7.5.19)
The whole process stops with J “ |j1 ´ j2 | by counting. The number of states in the
alternative bases must be the same.
j1ÿ
`j2
p2j ` 1q “ p2j1 ` 1qp2j2 ` 1q . (7.5.20)
j“|j1 ´j2 |
The range of J values matches bounds for addition of classical vectors J “ J p1q `J p2q
with lengths of J, j1 , j2 .
(ii) Clebsch-Gordan coefficients are found by explicit calculation of states for given j1
and j2 .
J 1 0
M Compare this results with the combination
1 of spin states in Eqns. (5.5.14) and (5.5.15)
with | 21 y “ | Òy, |´ 12 y “ | Óy which were
0 found previously by demanding definite sym-
metry; they are precisely the same.
-1
Have
3 (triplet) J “ 1 states symmetric
(7.5.25)
1 (singlet) J “ 0 state antisymmetric
Indeed, the permutation group and its representations are often central to constructing
multiplets in this way.
Tables for CG coefficients can be found in the Particle Data Group (PDG) Tables
( http://pdg.lbl.gov/2014/reviews/rpp2014-rev-clebsch-gordan-coefs.pdf). The PDG
collate all reviews and tables for properties of elementary particles.
The table for 2 b 1 giving states with M “ 1 is
8 TRANSFORMATIONS AND SYMMETRIES 57
J 3 2 1
m1 m2 M +1 +1 +1
2 -1 1/15 1/3 3/5
1 0 8/15 1/6 -3/10
0 1 2/5 -1/2 1/10
? a
Take but keep the sign. E.g., ´3{10 Ñ ´ 3{10.
Now consider a group G and transformations of a QM system U pgq for each g P G with
In any given case U pgq is a representation of G. Our aim is to find the unitary
operators U pgq when G is a group of translations, rotations or reflections. In these
cases we know how G acts geometrically
x ÞÑ gpxq , (8.1.6)
and on operators
U pgq: x̂ U pgq “ gpx̂q . (8.1.8)
These statements are equivalent. E.g., assuming the second one
´ ¯
x̂ U pgq|xy “ U pgqgpx̂q|xy “ gpxq U pgq|xy , (8.1.9)
x ÞÑ ga pxq “ x ` a (8.1.13)
U paq|xy “ |x ` ay
:
U paq x̂ U paq “ x̂ ` a , (8.1.14)
has precisely these properties. The effect of translation by a on ψpxq should be ψpx´aq
which is illustrated by the picture below.
R(x) R(x-a)
x x
0 a 0 a
Confirmation: on wavefunctions
ˆ ˙
B B
p̂ Ñ ´ i~ , U paq Ñ exp ´a (8.1.16)
Bx Bx
8 TRANSFORMATIONS AND SYMMETRIES 59
and so
ˆ ˙
B
U paqψpxq “ exp ´a ψpxq
Bx
1 2 2
“ ψpxq ´ a ψ 1 pxq ` a ψ pxq ´ . . .
2!
“ ψpx ´ aq Taylor’s theorem, (8.1.17)
as expected. Because of its special role in U paq the momentum p̂ is called the generator
of translations.
Then
Also have
U pα ` δαq “ U pαqU pδαq “ U pδαqU pαq , (8.2.10)
so deduce
rQ, U pαqs “ 0 . (8.2.11)
From above have
BU pαq i ´ α ¯
“ ´ Q U pαq ñ U pαq “ exp ´i Q given U p0q “ 1. (8.2.12)
Bα ~ ~
Q is the generator of this family of transformations, the one-parameter subgroup,
within G. To find Q we equate
with
ˆ ˙ ˆ ˙
: i i
U pαq A U pαq “ I ` αQ ` ... A I ´ αQ ` ...
~ ~
i
“ A ` αrQ, As ` Opα2 q (8.2.14)
~
to obtain
rQ, As “ ´i~f pAq . (8.2.15)
Knowing the RHS for any A determines Q. Conversely, the behaviour of any quantity
under the transformation is fixed by its commutation relation with the generator.
Now suppose the continuous family gα corresponds to a symmetry of the quantum
system
rU pαq, Hs “ 0 ô rQ, Hs “ 0 . (8.2.16)
The second equation says that the observable Q is a conserved quantity:
For a spinless particle this is what we seek. Note that L transforms as a vector:
For a particle with spin the generator above needs to be modified to ensure that S
also transforms as a vector
rQ, Ss “ ´i~ n ^ S , (8.3.9)
and this is achieved by taking
Note that L and S are not separately conserved unless U “ 0. However, L2 and S 2
are conserved as is J 2 . So we can write (J “ L ` S)
J 2 ´ L2 ´ S 2 1
L¨S “ “ ~2 pjpj ` 1q ´ lpl ` 1q ´ sps ` 1qq (8.3.14)
2 2
for eigenstates of J 2 , L2 , S 2 with eigenvalues pj, l, sq, respectively.
We now have alternative definitions of a scalar operator A and a vector operator v
in terms of commutation relations:
scalar: rJi , As “ 0
vector: rJi , vj s “ i~ εijk vk . (8.3.15)
‚ On angular momentum states |j, my, fixed j, the angular momentum generators Ji
can be represented as p2j ` 1q ˆ p2j ` 1q matrices (see section 7.3).
‚ Then ˆ ˙
i
U pθq “ exp ´ θn ¨ J (8.3.16)
~
can be computed easily in some cases.
‚ j “ 0: single state |0, 0y with J |0, 0y “ 0. Hence
Consider for simplicity rotations about the 3 (or z) axis – J3 is diagonal in our
representation. Then
» ¨ ˛fi ¨ ´iθ ˛
1 0 0 e 0 0
i
U pθk̂q “ expp´ θJ3 q “ exp ´iθ 0 0 0
– ˝ ‚ fl “ ˝ 0 1 0‚. (8.3.19)
~
0 0 ´1 0 0 eiθ
So under this rotation states behave as
Deduce that the states behave as the (spherical) components of a vector. More
familiar if we compare with same rotation on coordinates in spherical basis (bear in
mind for later):
But note the appearance of θ{2 in final result. Under any rotation by 2π we have
We do not get exactly the same state vector after rotation by 2π, which may seem
surprising. However, |χy and ´|χy have same physical content. Change in sign can
be a real effect if e.g. have two particles and rotate just one; change in sign can make
the difference between constructive and destructive interference.
Under rotation by 4π get exactly the same state
Similar behaviour under 2π and 4π rotations for any states with half-integral spin.
P : x ÞÑ ´ x . (8.4.1)
This is not part of a continuous family of isometries connected to the identity (un-
like e.g., rotations). Any reflection in a plane can be achieved by composing P with
rotations.
The corresponding unitary operator acts on position eigenstates as
U |xy “ | ´ xy
so U2 “ 1
ñ U : “ U ´1 “ U . (8.4.2)
U x̂ U “ ´x̂
U p̂ U “ ´p̂
U LU “ L pL “ x̂ ^ p̂q
USU “ S
UJU “ J. (8.4.5)
8 TRANSFORMATIONS AND SYMMETRIES 65
UHU “ H, (8.4.6)
which are thus states of definite parity. Note that for two-particle system P inter-
changes the spatial positions and so its eigenvalue determines the spatial symmetry
of the state: symmetric ηψ “ 1, anti-symmetric ηψ “ ´1.
There is also a notion of intrinsic parity where a particle state changes as
As with spin this is not attributable to any internal structure; it is a property of the
particle concerned. Values for some particles can be chosen by convention but then
others are determined. In general, for a system of particles,
" * " * " *
total spatial product of all
“ ˆ (8.4.11)
parity parity intrinsic parities
x
a
p´1ql ηa ηb . (8.4.12)
Note that parity gives rise to a multiplicative conservation law in contrast to the
additive conservation of momentum or angular momentum.
8 TRANSFORMATIONS AND SYMMETRIES 66
MIRROR
p -p
‚ This is a strong interaction process – total angular momentum, J , and parity con-
served.
‚ π d system:
Note that l “ 0 is the dominant state for low energy scattering (See AQM course).
‚ n n system:
Orbital ang. mom. l unknown to begin with but symmetry of spatial wavefunction
p´1ql . Total spin S “ 0 or 1.
j “ l “ S “ 1. (8.4.19)
H “ H0 ` V ptq , (9.1)
where H0 is time-independent with known eigenstates and eigenvalues and V ptq small in
some sense (remember perturbation theory earlier). I have suppressed the dependence
of V ptq on space coordinates.
‹ Our aim: to calculate the effect of V ptq order-by-order (actually stop at first order)
and, in particular, calculate probabilities for transitions between eigenstates of H0 as
functions of time. This is a dynamical question even if V is time-independent.
The interaction picture is defined by moving the ‘known’ part of the time evolution,
which is due to H0 , from states to operators:
|ψptqy “ eiH0 t{~ |ψptqy
i t 1
ż
1 1
“ ´ dt xf | eiH0 t {~ V pt1 q e´iH0 t {~ |iy
~ 0
An important special case is when V pt1 q is constant in time (at least for 0 ď t1 ď tq.
The amplitude above is then
i t iωt1
ż
1
´ e xf |V |iy dt1 “ p1 ´ eiωt q xf |V |iy , (9.1.12)
~ 0 ~ω
t
4
tw2
ft(w)
2p 4p w
available states t t
Now
sin 2 x
ż8 ż8
lim ft pωq “ t , ft pωq dω “ 2 dx “ 2π . (9.2.2)
ωÑ0 ´8 ´8 x2
9 TIME-DEPENDENT PERTURBATION THEORY 70
As t increases ft pωq is concentrated more and more around ω “ 0 and P ptq will be
non-negligible only when
2π 2π
´ ÀωÀ . (9.2.3)
t t
So for t large enough we have
ρpEf q∆E “ # states with specified parameters and energy in range pEf , Ef ` ∆Eq.
Note, we are assuming that |xf |V |iy|2 only depends on Ef and does not depend on
other unspecified quantum numbers which distinguish the different states with energy
Ef . This assumption is easily relaxed by summing over them, too.
For t sufficiently large
and then
ż
t
P ptq “ 2 dEf ρpEf q 2π~ δpEf ´ Ei q |xf |V |iy|2
~ F
$
’
& 0 if Ef is not in F
ˇ
“ 2πt ˇ
2 ˇ (9.2.10)
% ~ ρpEi q|xf |V |iy| ˇ
’ if it is
Ef “Ei
‚ t not so large that 1st order perturbation theory breaks down: need P ptq ! 1.
‚ t large enough that the δ-function approximation is valid. This means changes in
ρpEf q|xf |V |iy|2 are small for changes in Ef of order ~{t (see diagram above).
P ptq 2π 2
t “ ~ ρpEi q |xf |V |iy| Fermi’s Golden Rule.
where Ei and Ef are now the energies of initial and final atomic states, respectively.
Compare λ with the size of the atom given by the Bohr radius
1 ~
a0 „ . (9.3.6)
α mc
Have |q| “ 2π{λ and x is confined to the atomic interior so |x| À a0 . Thus,
a0 1
q¨x À „ α „ . (9.3.7)
λ 137
To lowest order we therefore neglect q ¨ x and consider the interaction to be
e
VI ptq “ ´ p̂ ¨ ε cos pω0 tq . (9.3.8)
m
We now note that
p̂
rx̂, H0 s “ i~ , (9.3.9)
m
and so
ie
VI ptq “cos pω0 tqrx̂ ¨ ε, H0 s (9.3.10)
~
Now the first-order transition amplitude i Ñ f is
i t 1 ipEf ´Ei qt1 {~
ż
´ dt e xf |VI pt1 q|iy
~ 0
This is called the dipole approximation since the operator appearing in the matrix
element has the form ex̂ ¨ ε cos pω0 tq and may be interpreted as the interaction of the
electric dipole operator d̂ “ ´ex̂ with a classical electric field given by E 9 ε cos pω0 tq
of the form V ptq “ ´d̂ ¨ E. However, the derivation is from minimal substitution and
there are clearly corrections higher-order in q ¨ x. Can analyze formula in two parts.
(i) Time dependence. Separate cos into exponentials and integrate gives two terms
żt
1 1
dt1 eipEf ´Ei qt {~ e¯iω0 t . (9.3.12)
0
which occur with equal probability. These processes are absorption and emission,
respectively. Proceeding as in derivation of the Golden Rule could find precise for-
mula for transition rate. This would require computing the density of states ρpEf q
(see Statistical Physics for example of this).
labelled as usual by principal and ang. mom. quantum numbers. Use parity and
angular momentum properties of operators and states to determine when xf |x̂|iy ‰ 0.
‹ Parity.
Parity operator U satisfies U 2 “ 1 and have
‹ Angular momentum.
First show that x̂i |n l my are like product states for addition of ang. mom. 1 to l.
Intuitively, x̂i is a vector and we know that l “ 1 states also behave like a vector
under rotations. Now
and define
1
X1 “ ´ ? px̂1 ` ix̂2 q
2
X0 “ x̂3
1
X´1 “ ? px̂1 ´ ix̂2 q . (9.3.21)
2
It is easy to check
rL3 , Xq s “ ~qXq
a
rL˘ , Xq s “ ~ p1 ¯ qqp1 ˘ q ` 1qXq˘1 , (9.3.22)
9 TIME-DEPENDENT PERTURBATION THEORY 74
look back at (8.3.21). Compare with standard formulas in section 7.5 and Eqs.
(7.5.1) for action of L3 , L˘ on ang. mom. states |1, qy. These formulas match
exactly implying that with respect to angular momentum
Xq |n l my behave just like product states |1, qy|l, my . (9.3.23)
Another indication that this is an exact parallel is that in the position represen-
tation using spherical polars coordinates
c c
1 4π 4π
X˘1 “ ¯ ? r sin θ e˘iφ “ r Y1 ˘1 pθ, φq , X0 “ r cos θ “ r Y1 0 pθ, φq ,
2 3 3
(9.3.24)
which are angular momentum wavefunctions for l “ 1:
Ylm pθ, φq “ xθ, φ|l, my . (9.3.25)
What this means is that
ˆ ˙ ˆ ˙
L Xq |n l my “ rL, Xq s|n l my ` Xq L|n l my (9.3.26)
corresponds exactly to
ˆ ˙ ˆ ˙ ˆ ˙
L |1, qy|l, my “ L|1, qy |l, my ` |1, qy L|l, my , (9.3.27)
which is the usual action on product states for addition of angular momentum
for two subsystems. Use the same strategy as before:
˛ note that
ˆ ˙
L` X1 |n l ly “ 0
ˆ ˙ ˆ ˙
L3 X1 |n l ly “ ~pl ` 1q X1 |n l ly . (9.3.28)
˛ Find orthogonal state to |l ` 1, ly and take this as the next top state |l, ly.
˛ Repeat procedure as before.
Then we find that Xq |n l my has angular momentum q. numbers
l ` 1 or l or l ´ 1 for L2 and
(9.3.31)
m`q for L3
So for xn1 l1 m1 |Xq |n l my to be non-zero need
*
l1 P pl ` 1, l, l ´ 1q
second set of selection rules. (9.3.32)
m1 “ m ` q pq “ 0, ˘1q
But parity already forbids l1 “ l.
9 TIME-DEPENDENT PERTURBATION THEORY 75
Summary: the possible atomic transitions for dipole radiation are given by
l1 “ l ˘ 1
m1 “ m if ε3 ‰ 0 (9.3.33)
m1 “ m ˘ 1 if ε1 or ε2 ‰ 0
|iy “ |n l my|N y
|f y “ |n1 l1 m1 y|N 1 y (9.4.2)
where N, N 1 are the number of photons or the oscillator level in the quantum version
for the field intensity which is N ~ω0 (ignoring the oscillator zero-point energy). As
before have two possible processes corresponding to time-dependent factors
Thus for general atomic energy levels Ei ą Ej and EM radiation of frequency ω0 have
transitions when Ei ´ Ej “ ˘~ω0 .
Ei
absorption emission
Ej
loAomoon `
iÑj BiÑj ¨ pintensityq
looooooooomooooooooon or B jÑi ¨ pintensityq
looooooooomooooooooon
spontaneous stimulated spontaneous (9.4.5)
The quantities above are called the Einstein A and B coefficients. (There is a nice
original treatment using thermal equilibrium, the Boltzmann distribution for atomic
excitations and the Bose-Einstein distribution appropriate for indistinguishable pho-
tons.) Always have BiÑj “ BjÑi . Also, with intensity measured in natural units of
photon number then have
AiÑj “ BiÑj “ BjÑi . (9.4.6)
Stimulated emission is the basis of operation of the laser.
‹ END OF NON-EXAMINABLE MATERIAL
10 QUANTUM BASICS 77
10 Quantum Basics
There are many aspects of QM which still seem rather mysterious, particularly in
relation to measurements and how they have been axiomatized. For example,
The main aim of this last section is to see how such questions may be addressed, and we
will also look briefly at some loosely related topics to do with information in classical
and quantum systems.
‚ In QM, s becomes an operator with eigenstates |0y and |1y corresponding to observed
values. The general normalized state of the system is
It would seem that a qubit contains vastly more information than a classical bit, but
we must beware: the information is not easily accessed.
‚ In QM, measuring s irrevocably changes |ψy since it is projected onto |0y or |1y in
the process. We could recover the values of |α|2 and |β|2 to some degree of accuracy
if we had a large number of copies of |ψy, by measuring s many times. This raises
the question of whether we can faithfully copy some given state |ψy in a quantum
system. It turns out that this is forbidden by the
10 QUANTUM BASICS 78
No-cloning Theorem.
We must be careful about what we mean. I can prepare an ensemble of systems each of
which is in the given state |ψy. For example, use a Stern-Gerlach apparatus to produce
a beam (the ensemble) of spin- 12 particles each with eigenvalue of σz “ 1: a “spin up”
state. Then clearly I have a lot of systems (i.e. the particles) in state
|ψy “ | Òy , (10.1.2)
¨ ˛ evals
cos 2θ
| Ò θy “ cos 2θ | Òy ` sin 2θ | Óy ù ˝ ‚ `1
θ
sin 2
(10.1.4)
¨ ˛
´ sin 2θ
| Ó θy “ ´ sin 2θ | Òy ` cos 2θ | Óy ù ˝ ‚ ´1
θ
cos 2
U
|ψy b |by ÝÑ |ψy b |ψy . (10.1.7)
U
|φy b |by ÝÑ |φy b |φy . (10.1.8)
xφ|ψyxb|by “ xφ|ψyxφ|ψy
ñ xφ|ψy “ 0 or 1 . (10.1.9)
This shows that copying can never work for general states in a space with dimen-
sion ą 1. This is a vital result for the success of the quantum encryption algorithm
where it is central to the idea that the encryption key for a classical algorithm can
be transmitted in such a way that a spy cannot intercept the message and decode it.
Interception irrevocably destroys the information and the interception is known to the
sender/receiver.
and components of spin σ A , σ B (leave out 12 ~ factor) of first and second particle are
subsequently measured by experimenters Alice (A) and Bob (B) in some order. How-
ever, they carry the particles far apart in space without altering the spin state |ψy
before measuring anything. Until the first measurement is made there is no definite
value for e.g. σzA or σzB .
Suppose Alice makes the first measurement of σzA and gets the result `1, say (prob-
ability 21 ). This projects |ψy onto the new state | Òy| Óy and so if Bob now measures
σzB he gets result ´1 with probability 1. He does not know this (unless Alice has sent
him a signal to tell him) but Alice knows with certainty what he will get! Einstein
was troubled by these kinds of non-local correlations but there is actually no loss of
causality here. Alice cannot use the collapse of the wave-packet
collapse
|ψy ÝÑ | Òy| Óy (Alice measured σzA “ `1q (10.2.2)
to send a message to Bob superluminally. Actually, Bob does not even know whether
or not Alice has made a measurement, since if the experiment is repeated many times
with an ensemble of pairs, Bob with get σzB “ `1 and ´1 with 50:50 probability
whether or not Alice has made a prior measurement.
10 QUANTUM BASICS 80
One way to explain such correlations and idea of non-locality is to ask whether the
spin components could actually take definite values, which are correlated, at the
instant the particles were created, i.e. at their point of origin. We still need a classical
probability distribution to describe how these values arise, but this ultimately might be
explained by some deterministic hidden variables. Can such a distribution reproduce
the predictions of QM. Or, if not, which possibility is supported by experiment?
A simple example is suppose I make two identical parcels one containing a green ball
and the other a red one. I give one to Alice and one to Bob who then travel far apart.
When Alice opens her parcel and sees, say, a green ball she knows instantly the Bob
has a red ball. This ”learning at a distance” is mundane and not controversial; there
is no perception of non-locality. Can repeat many times and, indeed, Bob has a 50:50
change of getting red and green. However, in this case each ball had a definite colour
before starting out – there is no analogue of ”collapse of the wavepacket”. Can the
spins in an entangled state be thought of as having definite spin components in a like
manner?
It is useful to think about measuring different components of σ A and σ B and use the
results above for eigenstates of σθ in Eqns. (10.1.4) and (10.1.6) to consider first what
happens to a single particle in conventional QM. Have
θ
|xÒ θ| Òy|2 “ |xÓ θ| Óy|2 “ cos 2
2
θ
|xÒ θ| Óy|2 “ |xÓ θ| Òy|2 “ sin 2 . (10.2.3)
2
These results clearly depend on the angle between k̂ and n defined earlier. Also, for a
component along a new direction φ we have
ˆ ˙
2 2 2 θ´φ
|xÒ θ| Ò φy| “ |xÓ θ| Ó φy| “ cos
2
ˆ ˙
2 2 2 θ´φ
|xÒ θ| Ó φy| “ |xÓ θ| Ò φy| “ sin . (10.2.4)
2
Now return to the two particle system |ψy and note that the spin-0 state is
ˆ ˙
1
|ψy “ ? | Òy| Óy ´ | Óy| Òy
2
ˆ ˙
1
“ ? | Ò θy| Ó θy ´ | Ó θy| Ò θy . (10.2.5)
2
This can be checked directly but also follows because spin-0 is a rotationally invariant
state – it cannot depend on the choice of the z-axis.
We want to consider probabilities for various outcomes when
Alice measures one of Bob measures one of
then (10.2.6)
σzA , σθA , σφA σzB , σθB , σφB
E.g.,
σzA “ 1 new state σθB “ ´1
then θ (10.2.7)
|ψy : prob: 1
| Òy| Óy prob: |xÓ θ| Óy|2 “ cos 2
2 2
10 QUANTUM BASICS 81
Total probabilities for this and some other measurement outcomes are similarly
θ
P pσzA “ `1, σθB “ ´1q “ 1
2
cos 2
2
φ
P pσzA “ ´1, σφB “ `1q “ 21 cos 2
2
ˆ ˙
θ´φ
P pσθA “ `1, σφB “ `1q “ 1
2
sin 2
. (10.2.8)
2
Also, know components along same direction are exactly anticorrelated. So, e.g.
sA A A
z , sθ , sφ and sB B B
z , sθ , sφ (10.2.10)
for each particle, all taking values ˘1 according to some probability distribution, with
values assigned as the particles are created (c.f. the classical red and green ball discus-
sion), yet still reproducing the QM results?
‚ Then, summing over variables that are not specified, we can write
P psA A
θ “ `1,sφ “ ´1q
“ P psA A
z “ `1, sθ “ `1q ` P psA A
z “ ´1, sφ “ ´1q (10.2.11)
‚ Specifying outcomes in terms of one measurement for A and one for B instead (using
exact anticorrelation) we get
P psA B A B A B
θ “ `1, sφ “ `1q ď P psz “ `1, sθ “ ´1q ` P psz “ ´1, sφ “ `1q . (10.2.12)
‚ Can such a probability distribution reproduce the results of QM? If so, then the
expressions derived for the QM probabilities above must satisfy Bell’s inequality.
This means ˆ ˙
2 θ´φ θ φ
sin ď cos 2 ` cos 2 (10.2.13)
2 2 2
10 QUANTUM BASICS 82
for any θ and φ. But this is false. E.g., θ “ 3π{4, φ “ 3π{2 with this inequality
would imply
ˆ ˙
2 3π 2 3π 3π
´ cos ´ sin ď cos 2
8 8
looooooooooooomooooooooooooon 4
cos 3π{4
1 1
ñ ? ď FALSE! (10.2.14)
2 2
The fact that this approach cannot reproduce QM means that we can distinguish these
alternatives experimentally, and it is QM which is correct. With these assumptions,
the indeterminate nature of the quantum state (no definite values of observables until
measurement is made) seems inescapable. There is a non-locality to the notion of
collapse of the wavefunction.
‚ Need a general formalism to handle this. Useful in e.g., statistical mechanics where
‘hidden’ degrees of freedom are all the fine details of the system which are not of
interest. I.e., we care about pressure, temperature etc. but not the dynamics of
individual particles.
V “ U b W (10.3.1)
where U is observed but W is the hidden/unobserved sector. The general state is then
ÿ
|Ψy “ αia |ψi y b |φa y , (10.3.2)
ia
with t|ψi yu and t|φa yu orthonormal bases for U and W , respectively. Assume
ÿ
k|Ψyk2 “ |αia |2 “ 1. (10.3.3)
ia
10 QUANTUM BASICS 83
What we really mean by distinction between U and W is that observables Q act just
on U , i.e. Q acts on V as the operator Q b I. Then
ÿÿ
˚
xQyΨ “ αjb αia xψj |Q|ψi y loomoon
xφb |φa y
ia j b
δab
ÿ
“ βij xψj |Q|ψi y . (10.3.4)
ij
ÿ
˚
The matrix βij “ αia αja is hermitian and positive-definite, and
a
ÿ ÿ
βii “ Trpβq “ |αia |2 “ 1 : trace or sum of eigenvalues . (10.3.5)
i ia
‚ Working just with U , the effect of the hidden sector W is to produce a probability
distribution tpi u for a particular set of states |χi y in the observed sector.
‚ If there is only one non-zero probability we say we have a pure state for U and |Ψy
can then be written as a single tensor product
|Ψy “ |χ1 y b |φ1 y ñ β “ diagp1, 0, . . . , 0q (10.3.9)
where I have chosen p1 “ 1, pi “ 0 i ą 1.
‚ If there is more than one non-zero probability we say we have a mixed state for U ,
and |Ψy cannot be written as a single tensor product. In this case we also say that
we have an entangled state of the subsystems U and W .
Density operators or matrices provide a powerful way of re-expressing results about
quantum measurements. Recall first that the trace of an operator A on a space V is
defined by
ÿ
Tr V pAq “ xn|A|ny , t|nyu any orthonormal basis for V. (10.3.10)
n
We can express all aspects of measurements in terms of traces of operators rather than
inner-products of states. So, with notation as before,
on V “ U b W
define ρ “ |ΨyxΨ| , the density operator corresponding to |Ψy .
(10.3.11)
10 QUANTUM BASICS 84
Then
ÿ ÿ ˆ ˙
xQyΨ “ xΨ|Q|Ψy “ xΨ|nyxn|Q|Ψy “ xn| Q|ΨyxΨ| |ny “ Tr V pQ ρq (10.3.12)
n n
To show this choose t|nyu “ t|χi yu as the basis to evaluate the trace. Then
ÿ ÿ
Tr U pQ ρ̄q “ xχj |Q|χi y looomooon
xχi |ρ̄|χj y “ pi x|Q|yχi “ xQy . (10.3.15)
i,j i
pi δij
‚ Provided the states |χi y obey the time-dependent Schrödinger equation with Hamil-
tonian H, the EoM for ρptq is (see Q4.8(a))
d
i~ ρ “ rH, ρs . (10.3.16)
dt