TP3-Notes Lenz
TP3-Notes Lenz
TP3-Notes Lenz
PHYS3661
Quantum Mechanics
Epiphany 2015
Alexander Lenz
IPPP, Durham University
April 8, 2015
1
2
Contents
1 Scattering experiments and cross sections 8
1.1 Introduction and Motivation . . . . . . . . . . . . . . . . . . . 8
1.2 Basics in scattering . . . . . . . . . . . . . . . . . . . . . . . . 10
3
7 Advanced Topics in Scattering 37
7.1 Form factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
7.2 Electron scattering on nucleon . . . . . . . . . . . . . . . . . . 37
7.3 Collisions between identical particles . . . . . . . . . . . . . . 39
7.3.1 Two spin-less bosons . . . . . . . . . . . . . . . . . . . 40
7.3.2 Two spin 1/2 fermions . . . . . . . . . . . . . . . . . . 40
7.4 Introduction to multichannel scattering . . . . . . . . . . . . . 40
7.4.1 Elastic Scattering . . . . . . . . . . . . . . . . . . . . . 40
7.4.2 Inelastic Scattering . . . . . . . . . . . . . . . . . . . . 40
4
14 Solutions of the Dirac equation 74
14.1 Plane wave solutions of the Dirac equation . . . . . . . . . . . 74
14.2 Non-relativistic limit of the Dirac-equation . . . . . . . . . . . 75
17 Revision Lecture 89
17.1 Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
17.2 Quantum statistical mechanics . . . . . . . . . . . . . . . . . . 89
17.3 Relativistic Quantum mechanics . . . . . . . . . . . . . . . . . 89
17.4 Interpretation of Quantum mechanics . . . . . . . . . . . . . . 89
5
6
Some Comments:
Prerequisites:
Foundations of Physics 2A
Theoretical Physics 2
Mathematical Methods in Physics or Analysis in many variables
Foundations of Physics 3A
Theoretical Physics 3 - Electrodynamics
7
1 Scattering experiments and cross sections
Chapter [13.1] of [1]
Popular information
Advanced information
The LHC - is in principle a huge microscope, which tells us about the fun-
damental constituents of matter.
Atoms are built of electrons and the nuclei. The electrons are elementary ac-
cording to the current experimental precision. The nuclei consists of nucleons,
i.e. protons and neutrons. Protons and neutrons are to a first approximation
built out of three quarks, which are also elementary according to the current
8
experimental precision.
Currently we know the following fundamental particles and forces:
Matter constituents
Quarks Leptons
u c t q = +2/3 q = 0
e
d s b q = 1/3 e q = 1
Creation of mass
Masses of the elementary particles are in contradiction to some math-
ematical properties of the theory. The Higgs mechanism is a trick to
solve this problem, for the price of introducing a new particle: the
Higgs boson with spin 0.
This is the so-called standard model of particle physics, which has been
experimentally verified by hundreds of experiments - sometimes with an in-
credible precision.
Most of this knowledge comes from scattering-experiments, like the Ruther-
ford experiment (Geiger and Marsden 1909, Rutherford 1911) or the LHC.
The corresponding scattering events take place on a microscopic level. Thus
an understanding of the quantum mechanical foundations of scattering is the
basis for a fundamental understanding of the building blocks of the universe.
9
1.2 Basics in scattering
The basic idea of scattering is simple. We have a target and some incident
particles, that are flying towards and interacting with the target. In the case
of Rutherford scattering the target was a gold foil and the incident particles
were alpha particles (He++ ), in the case of LHC both target and incident
particles are protons.
10
The incident particle is described by its energy E and the impact parameter
b. If the incident particles and the target stay intact during the scattering
process, we speak about elastic scattering. In that case the position of the
incoming particles after scattering is given by the scattering angle (, ) and
a radial variable r.
For simplicity we assume besides elastic scattering that the target is infinitely
heavy and that there is an azimuthal symmetry of the interaction.
Now we denote by a small area of the incoming particle flux. All particle
within will be scattered into the solid angle element . If we normalise
the solid angle element to an unit element d, then the corresponding area of
the incoming particles reads d. Having an azimuthal symmetry, the problem
looks like:
11
Integrating over the full solid angle d, will then give the effective area of
the scattering target, called total cross section . Defining in addition the
differential cross-section d/d we get
Z2 Z
d d
Z
tot = d = d d sin . (1)
d d
0 0
d R2
= , (2)
d 4
= R2 , (3)
d 1 dN
:= , (5)
d L d
12
which is many times used as the initial definition of the differential cross
section. An elastic scattering event at LHC looks like
p+pp+p. (6)
The total cross-section for this scattering at LHC with a proton energy of 7
TeV reads [17]
The above concepts can also be generalised to inelastic processes, like the
production of a Higgs particle
p+pX +H . (9)
The cross section for this process is given by [19] (see e.g.[18] for a part of
the state-of-the-art calculation)
20pb . (10)
The number of events for an inelastic process can be determined from the
integrated luminosity
N = L , (11)
Z
N = Ldt . (12)
13
The LHC achieved until now an integrated luminosity of 20(f b)1 per exper-
iment (ATLAS and CMS), thus the following number of Higgs particles was
produced at each experiment
pb
N = 20f b1 20pb = 400 = 4 105 . (13)
103pb
To obtain the number of a certain Higgs decay like H one has to
multiply the overall number of produced Higgses with the corresponding
branching ratio, in that case
The value of the differential cross section for two colliding particles A and B
depends on the reference frame. Two commonly used ones are the laboratory
frame (L), where particle B is in the beginning at rest and centre of mass
frame (CM). One finds
sin
tan L = , (15)
cos +
3
(1 + 2 + 2 cos ) 2 d
d
= , (16)
d L |1 + cos | d CM
with = mA /mB .
14
2 Potential scattering (general features)
Chapter [13.2] of [1]
~2 ~ 2
i~ (~r, t) = + V (~r) (~r, t) , (17)
t 2m
where ~r describes the distance of the particle from the origin of the potential.
The scattering of particle A with the mass mA with a particle B with the mass
mB due to a force that is described by the potential is given by
2
~ ~2
i~ (~r, t) = + V (~r) (~r, t) , (18)
t 2
where is the reduced mass and ~r = ~rA ~rB is the relative distance between
the two particles. This is e.g. the case at LHC, where two protons scatter, but
with relativistic energies. But here we concentrate first on the fixed potential
case.
If the potential does not depend on time - as in our case - we can separate
the time dependence and we look for stationary solutions:
E
(~r, t) =: (~r)ei ~ t . (19)
p~2 ~2~k 2 1
E =: =: =: m~v 2 , (21)
2m 2m 2
with
15
Now we can rewrite Eq.(20) as
h i
~ 2 + ~k 2 U(~r) (~r) = 0 , (22)
Now we describe the incoming particle with a plane wave, i.e. all the incom-
ing particles have the same momentum p~ = ~~k and travel all in the same
direction, which choose to be the z-direction.
~
inc. (~r) = eik~r = eikz . (26)
This fulfils Eq.(25). With this normalisation we have for the particle density
= |inc. (~r)|2 = 1, i.e. one particle per unit volume. Thus the incident flux
reads
p
F = v= , (27)
m
1 m 1
2
= . (28)
ms s m3
The outgoing particle we want to describe by an outgoing wave
eikr
scat. (~r) = f (k, , ) , (29)
r
16
with the scattering amplitude f (k, , ). The factor 1/r is needed for the
conservation of probability. This wave function also fulfils Eq.(25) for very
large values of r.
Thus we get for the full asymptotic wave function:
i~k~r eikr
~k e + f (k, , ) . (30)
r r
This ansatz can now be used to calculate the current density ~j
~j(~r) = ~ ()
h i
~ ~
( ) (31)
2mi
and one gets
~k |f (k, , )|2
jr = . (32)
m r2
jr represents the number of particles crossing a unit area per unit time and
the detector presents a cross-sectional area r 2 d to the scattered beam. The
number of particles entering the detector per unit time, dN is
~k
dN = jr r 2 d = |f (k, , )|2 d (33)
m
= v |f (k, , )|2 d (34)
F
= |f (k, , )|2 d . (35)
dN can also be described by the incoming flux F and the area d these
particles are crossing before the scattering:
d
dN = F d = F d . (36)
d
Equating the two expressions and using = 1 we get our master formula
d
= |f (k, , )|2 . (37)
d
The differential cross section is measured in the experiment, while the scat-
tering amplitude is calculated in theory, by solving the Schrodinger equation.
17
3 Spherical Bessel functions
Chapter [7.3/4] of [1]
Before continuing with the scattering problems we make some mathematics
interlude and introduce special functions that will be very helpful for the
solution of the Schrodinger equation.
2
~ 2 1 2 1 1
= 2 r + 2 sin + 2 2 (39)
r r r r sin r sin 2
1
~2
L
= 2 r2 2 2 , (40)
r r r ~r
with the angular momentum operator L. ~ Next we separate the radial and
the angular dependence of the solution
X
(~r) = cElm Elm(~r) (41)
E,l,m
Elm (~r) = REl (r)Ylm (, ) , (42)
with the radial function REl that has to be determined and the well-known
spherical harmonics Ylm (TP2, Edynamics).
The spherical harmonics can be expressed in terms of associated Legendre
functions and these in terms of Legendre polynomials - see also the
corresponding Wikipedia articles.
18
The following orthogonality relations hold
Z
Yl m Yl m
d = ll mm , (46)
Z1
2(l + m)!
Pkm (x)Plk (x)dx = ml , (47)
(2l + 1)(l m)!
1
Z1
2
Pm (x)Pn (x)dx = mn . (48)
2n + 1
1
Having a cylinder symmetry the solutions will be even more simple and there
is no dependence on m, which can then be set to 0 and thus the spherical
functions are simply replaced by Legendre polynomials:
Ylm (, ) Pl (cos ) . (49)
Coming back to the general case and using
~ 2 Ylm (, ) = l(l + 1)~2 Ylm (, ) ,
L (50)
we get for the time-independent Schrodinger equation in Eq.(22)
h i
~ 2 U(r) + ~k 2 Elm (~r) =
0 (51)
(" # )
1
L~2
r2 2 2 U(r) + ~k 2 REl (r)Ylm (, ) = 0 (52)
r 2 r r ~r
1 2 l(l + 1)
r U(r) + ~k REl (r) =
2
0 (53)
r 2 r r r2
2
2 l(l + 1)
+ U(r) + ~k REl (r) =
2
0 . (54)
r 2 r r r2
Eq.(54) is now the defining equation for the radial part of the Schrodinger
equation - the only missing part.
19
Taking in addition l = 0 we get
2
~ 2
+ k uE0(r) = 0 , (57)
r 2
which is solved by
sin(kr) cos(kr)
RE0 = A +B . (59)
r r
The cosine part gives a divergent function REl , at the origin, to avoid that
one simply can set B = 0.
2
2 l(l + 1)
+ + 1 Rl () = 0 , (60)
2 2
sin() cos()
R0 () = Ak + Bk . (61)
20
What equation holds for1
1 l ()
() := ? (63)
3 2(l + 1) 2
2(l + 1)
+ + l () = 0 (64)
3 2 2
2
2(l + 1) 2(l + 1)
+ + 1 (()) = 0 (65)
2 2
2
2(l + 2)
+ + 1 () = 0 . (66)
2
The last line is simply the defining equation for l+1 (). Thus we have
the desired recursion relation
1 l ()
l+1 () = (67)
leading to
l
1
l () = 0 () . (68)
With 0 () given in Eq.(59) we get two fundamental solutions of Eq.(60):
the spherical Bessel functions (or Spherical Bessel functions of
the first kind) jl () and the spherical Neumann functions (or
spherical Bessel functions of the second kind) nl ().2
l
1
l l sin
Rl () = jl () := (1) , (69)
l
l l 1 cos
Rl () = nl () := (1) . (70)
Eq.(55) is then solved by REl (r) = jl (kr) and REl (r) = nl (kr).
1
Clearly, quite some trial and error was necessary to get this idea!
2
The MATHEMATICA notation reads jl (x) = SphericalBesselJ[l, x] and nl (x) =
SphericalBesselY [l, x].
21
3.3.3 Properties of the solutions
sin
j0 () = (71)
sin cos
j1 () = (72)
2
3 1 3
j2 () = 3
sin 2 cos (73)
and
cos
n0 () = (74)
cos sin
n1 () = 2 (75)
3 1 3
n2 () = 3
cos 2 sin (76)
1.0
0.8
0.6
0.4
0.2
5 10 15
- 0.2
- 0.4
22
1.0
0.5
5 10 15
- 0.5
- 1.0
21
nl () = (1)l+1 Jl 1 () (82)
2 2
23
(1,2)
3.3.5 Spherical Hankel functions h
These functions are just linear combinations of the spherical Bessel and Neu-
mann functions
h(1)
() = j () + in () (83)
h(2) h(1)
() = j () in () = () (84)
24
4 The method of partial waves
After our mathematics excursion we turn back to the ansatz for the general
scattering solution, given in Eq.(30).
~ eikr
~k = eik~r + f (k, , ) . (90)
r
X
~k = Rl (k, r)Pl (cos ) , (91)
l=0
X
f (k, ) = fl (k)Pl (cos ) . (92)
l=0
Each term in the sum in Eq.(91) is called a partial wave and it is simultane-
ously an eigenfunction of L ~ 2 and Lz with eigenvalues l(l + 1)~2 and 0. Using
the spectroscopic notation we denote the l = 0, 1, 2, 3, , .. wave by s, p, d, f, ...
wave.
The coefficients fl (k) in the expansion of the scattering amplitude are called
partial wave amplitudes. According to our master formula for the cross
25
section, we get
d
= |f (k, )|2 (93)
d
X X
= fl (k)fl (k)Pl (cos )Pl (cos ) . (94)
l =0 l=0
Z+1
d d
Z
= d = 2 d cos (95)
d d
1
X
X Z+1
= 2 fl (k)fl (k) Pl (x)Pl (x)dx (96)
l =0 l=0 1
X 1 X
= 4 |fl (k)|2 = l . (97)
l=0
2l + 1 l=0
Thus a knowledge of the partial wave amplitudes gives immediately the total
cross section. The relation between the partial wave amplitudes and the
radial function Rl (k, r) in the partial wave can be derived as follows:
~ eikr
~k = eik~r + f (k, , ) (98)
r
X X X eikr
Rl (k, r)Pl (cos ) = (2l + 1)il jl (kr)Pl (cos ) + fl (k)Pl (cos )
l=0 l=0 l=0
r
(99)
eikr
Rl (k, r) = (2l + 1)il jl (kr) + fl (k) (100)
r
26
Bessel and Neumann functions
1 l l
Rl (k, r) = Bl (k) sin kr Cl (k) cos kr . (102)
kr 2 2
By defining
p
Al (k) := Bl (k)2 + Cl (k)2 , (103)
Cl (k) Cl (k) Bl (k)
tan l (k) := , sin l (k) := , cos l (k) := , (104)
Bl (k) Al (k) Al (k)
we get
Al (k) l l
Rl (k, r) = cos l (k) sin kr + sin l (k) cos kr
kr 2 2
(105)
Al (k) l
= sin kr + l (k) . (106)
kr 2
The l (k) are called phase shifts and they denote the strength of the scat-
tering in the l-th partial wave.
If there is no potential at all acting, then the solution in Eq.(102) has to hold
everywhere and in particular for r = 0. Since nl (r) is singular at r = 0, the
corresponding coefficient has to be set to 0, Cl (k) = 0. In that case we get
also a vanishing phase shift l (k) = 0.
Combing the asymptotic form of the radial function with our result from
Eq.(100) we get
l
l sin kr 2 eikr
Al (k) l
sin kr + l (k) = (2l + 1)i + fl (k) ,
kr 2 kr r
(107)
which can be solved by
Al (k) = (2l + 1)il eil (k) , (108)
2l + 1 2il (k)
fl (k) = e 1 . (109)
2ik
With this we get some final expressions for the scattering amplitude and for
the differential cross section:
1 X
(2l + 1) e2il (k) 1 Pl (cos ) ,
f (k, ) = (110)
2ik
l=0
4
l = |fl (k)|2 , (111)
2l + 1
4(2l + 1) 2
= sin l (k) . (112)
k2
27
Therefore the knowledge of the scattering phase is sufficient to solve a scat-
tering problem.
Setting = 0 in Eq.(110) we get
il (k)
eil (k)
1X il (k) e
[f (k, 0)] = (2l + 1) e (113)
k 2i
l=0
1X
= (2l + 1) sin2 l (k) (114)
k l=0
k
= . (115)
4
This is the famous Optical Theorem: the total cross section is given by
the forward ( = 0) scattering amplitude.
tan 0 (k)
ls := lim , (116)
k0 k
one finds
f k0
ls , (117)
d
k0
ls2 , (118)
d
k0
4ls2 . (119)
28
4.5 Resonances
According to Eq.(112) the l-th partial cross section reads
4 1
l = (2l + 1) . (120)
k 2 1 + cot2 l (k)
The energy where cot l (k) becomes zero is called resonance energy ER .
At the resonance one can write
ER E
cot l (k) = , (121)
(ER )/2
with the resonance width (ER ). This leads to the Breit-Wigner reso-
nance formula
4 2 (ER )/4
l = (2l + 1) , (122)
k2 (E ER )2 + 2 (ER )/4
which is valid in the vicinity of a resonance. For M ER it looks like:
29
4.6 Hard-sphere potential
-
To repeat this lecture: Chapter [13.3/4] of [1]
30
5 The integral equation of potential scatter-
ing
Chapter [13.5] of [1]
31
Now we have a solution for the Fourier transform of the Greens function, so
we are left with transforming G0 (k, ~q) back to G0 (k, ~r).
d3 q ei~q~r
Z
G0 (k, ~r) = (132)
(2)3 k 2 q 2
Z
Angular integr. 1 qeiqr
= 2 dq (133)
4 ir q2 k2
Z
eiqr eiqr
PFD 1 1
= + dq (134)
4r 2i q+k qk
ikr
Cauchy e
= . (135)
4r
This is the final result for the Greens function!
32
This is now the promised derivation of Eq.(30).
d m2 2
= 2 4 T~k~k . (142)
d 4 ~
33
6 The Born approximation
Chapter [13.6] of [1]
(144)
In the next step we replace ~k (~r ) by the whole integral solution and so on.
In the end we get an expansion like
~
0 (~r) = eik~r , (145)
Z
~ ~
1 (~r) = eik~r + G0 (k, ~r ~r )U(~r )eik~r d~r , (146)
Z
i~k~ ~
2 (~r) = e r
G0 (k, ~r ~r )U(~r )eik~r d~r
+
Z Z
~
+ G0 (k, ~r ~r )U(~r ) G0 (k, ~r ~r )U(~r )eik~r d~r d~r (147)
,
... = ...
If the potential U gives rise to a small perturbation of the system, then such
an expansion might converge and the real result could be approximated by
the first terms.
34
~
with the momentum transfer ~.
~ := ~k ~k .
(150)
~ 2 = 2k 2 2~k ~k
2 = ||
= 2k 2 (1 cos )
= 4k 2 sin2 ,
2
= 2k sin . (151)
2
Therefore the Born scattering amplitude for a central potential reads
Z Z Z2
B 1 2
f (k, , ) = drr U(r) d sin deir cos (152)
4
0 0 0
Z Z+1
1
= drr U(r) d cos eir cos
2
(153)
2
0 1
Z
1
= drr sin(r)U(r) . (154)
0
er
UY (r) = U0 , (155)
r
we get
U0
fYB = (156)
2 + 2
and thus
d B U02
= . (157)
d (2 + 2 )2
35
The Coulomb potential can be considered as a special case of the Yukawa
potential
1
UC (r) = U0 (158)
r
and we get
U0
fCB = , (159)
2
leading to the famous Rutherford scattering formula
d B U02
= (160)
d 4
2
qA qB 1
= . (161)
40 16E sin4
2
2
6.4 Propagator
The Greens function in
Z
i~k~ ~
n (~r) = e + G0 (k, ~r ~r )U(~r )eik~r d~r
r
Z Z
~
+ G0 (k, ~r ~r )U(~r ) G0 (k, ~r ~r )U(~r )eik~r d~rd~r
+... (162)
36
7 Advanced Topics in Scattering
7.1 Form factor
Assume that the potential is given by an extended distribution of electric
charge. Considering the interaction of this distribution with another charge
e we have the following potential
Ze2
Z
V (~r) = (r )dr~ , (163)
s
The Fourier transform of the charge distribution, , is called the form factor
of the extended object. This is a very general result: the cross-section of an
extended object is given by the cross-section of the corresponding point like
object times the form factor squared.
dExtended dP ointlike
= ||2 . (169)
d d
The Fourier transformation of the charge distribution can now be obtained
by comparing the measurement of the differential cross section with the cal-
culation of the corresponding point-like cross section.
37
d d
= .
d d Rutherf ord
d d 2 2
= 1 sin ,
d M ott d Rutherf ord 2
with = v/c.
E
d d
= .
d M ott d M ott E
4EE
Four momentum transfer: q 2 = (p p )2 c2
sin2 2 ; Q2 := q 2 .
d d 2
= 1 + 2 tan ,
d s=1/2P oint d M ott 2
with = Q2 /(4M 2 c2 ).
38
One would expect the following structure:
d d 2 2 2
= F (Q ) 1 + 2 tan ,
d d M ott 2
with F (0) = g/2.
The electric charge and the magnetic current have, however, different
distributions, thus we get two different form factors. Due to historic
reasons one writes:
Rosenbluth-formula (1955)
In the end one gets e.g. information about the structure and the size of a
proton: it turns out that a good fit for the charge density of the nucleon is
given by a Fermi distribution
1
(r) = 0 , (170)
1 + exp rR
a
39
the angular coordinates ( , + ). Thus interference can occur and
the corresponding scattering amplitudes have to be added. We get for the
differential cross section (for finding a particle at (, )
d
= |f (, ) + f ( , + )|2 (171)
d
= |f (, )|2 + |f ( , + )|2 + 2 (f (, )f ( , + )) ,
(172)
40
8 The density matrix
Chapter [14.1/2/3] of [1]
8.1 Definition
Quantum statistical mechanics is a combination of statistical mechanics
and quantum mechanics. In this case we have two sources for the necessity
of a statistical description:
A practical necessity because of a huge number of systems or degrees
of freedom. Having an infinite amount of computing power, we sim-
ply could overcome this problem and we would not need a statistical
description.
h|i = 1 . (174)
41
Choose a complete set of basis vectors |ni, i.e. orthogonal eigen vectors
of some complete set of operators:
or
hn|n i = (n n ) , (176)
if n is a continuous variable.
Completeness tells us X
1= |nihn| . (177)
n
with
with W being the statistical weight of the pure state |i, i.e. the proba-
bility of finding the whole system in the state |i:
0 W 1 , (185)
XN
W = 1 . (186)
=1
42
All in all we can now write
N XX
X
hAi = hn|iW h|n ihn |A|ni (187)
=1 n n
XX
= hn||n ihn |A|ni (188)
n n
X
= hn|A|ni (189)
n
= Tr(A) , (190)
Remarks:
Taking A to be the unit operator and keeping in mind that the pure
states are normalised to one, we have
Tr() = 1 . (192)
Tr(A)
hAi = . (193)
Tr()
Definition:
nn = hn||n i (194)
XN
= hn|iWh|n i (195)
=1
N
()
X
= W cn c()
n . (196)
=1
43
Diagonal elements: the diagonal elements of the density matrix
N
X 2
nn = hn|ni = W c()
n
(197)
=1
kk = kk kk . (200)
Now let us assume that the whole system is in a pure state |i,
meaning W = . So the density operator reads
= |ih| , (203)
( )2 = . (204)
44
This is now a new criterion for deciding whether a state was pure
or not!
kk = k k . (209)
8.2 Polarisation
8.3 Von-Neumann equation
In the Schrodinger-picture we have at the time t0
N
X
(t0 ) = W |(t0 )ih(t0 )| . (210)
=1
with
H
U(t, t0 ) = ei ~ (tt0 ) (212)
for a time-independent Hamiltonian. Thus the density operator at time t is
given as
N
X
(t) = W |(t)ih(t)| (213)
=1
= U(t, t0 )(t0 )U (t, t0 ) (214)
i H (tt0 ) +i H (tt0 )
= e ~ (t0 )e ~ . (215)
45
This last equation is equivalent to
i~ (t) = [H, (t)] , (216)
t
which is called the von-Neumann-equation or sometimes also .Liouville-
von-Neumann-equation
Remarks:
46
9 Quantum mechanical ensembles
Chapter [14.4/5] of [1]
=0, (224)
t
which is equivalent to H having no explicit time dependence and
[H, ] = 0 . (225)
Now we can choose a basis |ni in which both H and are diagonal
= (H) , (228)
a fixed volume V;
E0 E E0 + .
Notation:
47
: the total number of distinct pure states in which a member of the
ensemble can be found
= (N , V, E0 , ) .
S = kB log , (230)
The definition agrees with the usual thermodynamical one, but now
is calculated within using quantum mechanics.
48
3. The temperature T of the system is defined as
1 S
= . (231)
T E
The systems S and R are in contact and can thus exchange energy -
the interaction is supposed to be so weak, that at any time S and R
are in definite energy eigenstates with energies ER and ES .
E = ER + ES , (232)
E0 E E0 + with E0 .
49
The probability of finding the combined system with energies in [E, E +
dE] is
Cd(E) E0 E E0 +
dP = , (233)
0 else
where C is a constant describing the probability of finding one state.
Denoting the corresponding numbers in the system S and R by dS
and dR , we have
d = dS dR , (234)
dP = CdS (ES )dR (ER )(E ES ER ) , (235)
where the -function expresses the energy conservation.
dPn is the probability that the system S has the energy En without
regard to the state of the reservoir (i.e. we are summing over all the
states in the reservoir that yield the same En )
Z
dPn = CdS (En ) (E En ER )dR (ER ) (236)
= CdS (En )R (E En ) , (237)
with R (E En ) being the total number of states in the reservoir,
subject to the conservation of energy.
The corresponding entropy of the reservoir is given as
SR = kB log R , (238)
SR
R = e kB
. (239)
50
This is already the desired result: the probability of finding the system
En
S in the energy state with En is given by Ae kB T .
nm = AeEn nm , (247)
with = 1/(kB T ).
Remarks:
1. The coefficient A can be extracted from the condition Tr() = 1.
X
1 = nn (248)
n
X
= AeEn (249)
n
X
QNS := A1 = eEn = Tr eH .
(250)
n
2. The sum runs over all distinct states, which is not necessarily the same
as over all energy eigenstates.
eH
= . (251)
QNS
51
9.3 Grand canonical Ensemble
This is similar the canonical ensemble, but now we allow in addition the
exchange of particles between the system S and the reservoir R. The total
number of particles
N = NR + NS (256)
is constant.
All derivations are similar up to difference in the Taylor expansion of the
entropy, where now the change of particle numbers is taken into account.
SR (E En , N NS ) = SR (E, N )
SR (E, N ) SR (E, N )
En (NS ) NS + ... .
E N
En (NS )
= SR (E, N ) + NS + ... .
T T
(257)
The density matrix gets now the form
nm = Be[En (NS )NS ] nm , (258)
with B being a constant and = (SR /N /kB ). = / is called
chemical potential.
In order to give an operator form of the density operator, we introduce the
number operator S , that is counting the number of particles in the system
S. Its eigenvalues are N = 0, 1, 2, . . . . Thus we get for the density operator
= BeH(S )S . (259)
The constant B can again be determined from Tr() = 1 and we get
B1 := Z(, T ) = Tr eH(S )S .
(260)
Z is called the grand partition function and we can write
XX
Z(, T ) = eEn (N )N (261)
N n
" #
X X
= eEn (N ) eN (262)
N n
" #
X X N
eEn (N ) e
= (263)
N n
X
= QN (T )z N , (264)
N
52
9.4 Spin 1/2 particle in a magnetic field
9.5 Average of a particle in a box
53
10 Systems of non-interacting particles
Chapter [14.6] of [1]
where the sum runs over all energy eigenstates and nj is the number of
particles having the same energy Ej . The total number of particles in the
system is X
NS = N = nj . (266)
j
This leads to
N!
MB
gE = j (nj !) Maxwell-Boltzmann
gE = g BE = 1 Bose-Einstein . (269)
EF D
gE = 0, 1 Fermi-Dirac
Remarks:
ad MB: If all energies Ej , Ek , ..., El are different, then each energy level
is N!-fold degenerate. If, however, nj particles with nj > 1 have the
same energy, then an interchange among them does not change the
wave function.
54
ad BE: there is only one total symmetric wave function.
Now the partition function reads for all the three cases:
" #
X X
QN (T ) = gE exp nj Ej , (270)
{nj } j
P
where denotes a sum over every possible set of numbers nj that is com-
{nj }
patible with the total number N.
The average value of the energy hEi of the system can be calculated if the
average value hni i of each occupation number ni is known.
X
hEi = Ei hni i . (271)
i
which is equivalent to
log QN (T )
hni i = . (273)
(Ei )
For the case of MB the sum can be performed explicitly to obtain
" #N
X
QN (T ) = exp (Ej ) , (274)
j
leading to
hni i = exp [ (Ei + )] , (275)
with
P
exp (Ej )
j
= log . (276)
N
55
For the case of BE and FD it is easier to start with the grand partition
function
X
( )
X X X
Z = . . . gE exp [nj Ej (N)] nj (277)
n1 =0 n2 =0 j j
X X Y
= . . . gE exp {nj [Ej (N) + ]} (278)
n1 =0 n2 =0 j
YX
= gE exp {nj [Ej (N) + ]} . (279)
j nj =0
log Z
= . (283)
[Ei (N)]
Now all three cases (MB, BE and FD) can be expressed with one formula:
1
hni i = (284)
exp [Ei (N) + ] +
with
0 MB
= 1 BE . (285)
+1 FD
The function
1
hn(E)i = (286)
exp [E + ] +
is called the distribution function. Plotted as a function of E + one
gets
56
4
-4 -2 2 4
Remarks:
57
11 The Klein-Gordon equation
Chapter [15.1] of [1]
11.1 Derivation
Using the correspondence principle
E i~ , (288)
t
~~
~p , (289)
i
the Schrodinger equation can be derived from the non-relativistic energy
momentum relation:
p~2
E = , (290)
2m
~2
~2
i~ (t, ~x) = (t, ~x) . (291)
t 2m
It is a priori clear, that the Schrodinger equation cannot be applied to rela-
tivistic energies.
If we want to discuss highly energetic phenomena, like particle accelerators,
early universe, black holes,.. we have to start from the relativistic energy mo-
mentum relation. This will immediately lead to the Klein-Gordon equa-
tion, which was derived by Schrodinger (1925), Gordon (1926) and Klein
(1927).
E 2 = p~2 c2 + m2 c4 (292)
2
~ 2 + m2 c4 )(t, ~x) .
~2 2
(t, ~x) = (~2 c2 (293)
t
The last equation can be rewritten as
1 2
mc 2
~ +
2
=0 . (294)
c2 t2 ~
Remarks:
The differential operator appears also e.g. in the wave equation in
electro dynamics. It is called DAlembert operator and it is Lorentz
invariant. This can be seen by writing
1 2 ~2,
= = = (295)
x x c2 t2
58
~ and
with x = (ct, ~x), x = (ct, ~x), = /x = (1/c /t, )
~
= /x = (1/c /t, ).
Thus (m can be written as p p and is assumed to be a scalar) the
whole Klein-Gordon equation is Lorentz invariant.
E E q , (296)
q~
p~ ~p A . (297)
c
Hence the energy-momentum relating changes to
q~ 2 2
(E q)2 = (~p A) c + m2 c4 , (298)
c
leading to
2 " 2 #
~~ q~
i~ q (t, ~x) = A c2 + m2 c4 (t, ~x) . (299)
t i c
59
11.3 Stationary solutions
Now we look for stationary solutions
E
(~x, t) =: (~x)ei ~ t , (301)
of the Klein-Gordon equation, given in Eq.(299).
Assuming that A~ and are time independent, we get
" 2 #
~ q
~ A ~ c2 + m2 c4 (~x) .
(E q)2 (~x) = (302)
i c
60
This gives
2
d 2 d l(l + 1) m
+ + 2 2 (E q(r)) REl (r) = 0 ,
dr 2 r dr r2 ~
(310)
11.4 Interpretation
Looking at the Klein-Gordon equation a little more in detail one immediately
finds some seemingly inconsistencies.
1. In the same way as it was done for the Schrodinger-equation one can
also derive a continuity equation for the case of the Klein-Gordon equa-
tion.
~ ~
+j =0. (313)
t
In this case the probability density (~r, t) and the probability current
density ~j(~r, t) reads
i~
(~r, t) = , (314)
2mc2 t t
~j(~r, t) = ~ ()
~ ~ ) .
( (315)
2mi
Here one easily sees that the probability density can also have negative
values, prohibiting a simple interpretation of this quantity.
2. The free Klein-Gordon equation in Eq.(294) can be solved by plane
waves
~
(~r, t) = Aei(k~rt) . (316)
Inserting this ansatz into Eq.(294) one gets
~2 2 = m2 c4 + ~2 c2 k 2 (317)
E = m2 c4 + ~2 c2 k 2 . (318)
61
Now we have in principle a positive and a negative energy value; a
result that has to be understood somehow.
The source of all these problems lies actually in the fact, that we tried to
derive a relativistic single-particle wave function. Such an endeavour has,
however, to fail. Having access to very large energies, we can always have
pair creation. If we want to describe single electrons |e i interacting with
photons, we always will have the possibility of e.g. an electron sending out
a photon, which again splits up into an electron-positron pair. Thus the 3
particle state |e e+ e i has also to be taken into account, as well as states with
even more particle. This can only be done in the framework of Quantum
Field Theory, the topic of one of the Level 4 courses.
We will here , however, follow the historic path, where first another single
particle wave function was introduced.
62
12 The Dirac equation
Chapter [15.2] of [1]
12.1 Introduction
12.2 Derivation
~c k 2
i~ = k + mc = H . (319)
t i
63
This equation should of course fulfil the relativistic energy-momentum rela-
tion. To test this requirement we look at H 2 :
2
2 ~c i 2 ~c j 2
~ = i + mc j + mc (320)
t2 i i
~mc3 i
2 2 i j i 2 4 2
= ~ c i j + ( + )i + m c (321)
i
i j j i
~mc3 i
2 2 + i 2 4 2
= ~ c i j + ( + )i + m c
2 i
(322)
!
2 2 ij 2 4
=
~ c i j + m c . (323)
i j + j i = 2 ij , (324)
i + i = 0 , (325)
2 = 1 . (326)
This clearly can not be fulfilled by simple numbers, but matrices might do
the job. i and are called Dirac-matrices.
The following form of the Dirac- matrices is called Dirac representation;
it fulfils Eq.(324-326):
0 k
k I2 0
= , = , (327)
k 0 0 I2
( i )2 = I2 , (331)
i j = iijk k for i 6= j , (332)
Tr( i ) = 0, (333)
det( i ) = 1 . (334)
64
With this definitions Eq.(324-326) can be easily checked.
Remarks:
= , k = k . (335)
The form of the Dirac matrices is not unique! Every non-singular trans-
formation of Eq.(327) Mi M 1 , MM 1 fulfils also Eq.(324-326).
Mi M 1 Mj M 1 + Mj M 1 Mi M 1 = 2 ij , (336)
Mi M 1 MM 1 + MM 1 Mi M 1 = 0 , (337)
MM 1 MM 1 = 1 . (338)
This derivation also implies that the wave function is now a four
component object. It is called the Dirac-spinor.
We multiply the Dirac-equation from Eq.(319)b with from the left and
get
~c
i~ = k k + mc2 . (342)
t i
65
The adjoint relation to that reads
~c
i~ = k k + mc2 . (343)
t i
The difference of these two equations, Eq.(342) - Eq.(343) - gives
~c k
= k + k k + mc2 . (344)
i~
t i
Taking into account that k and are hermitian, i.e. k = k and =
we get
~c
= k k
i~ (345)
t i
k
+ ck = 0 (346)
t
~ ~j = 0 .
+ (347)
t
In the last line we defined the probability density to be
= = |1 |2 + + |4 |2 , (348)
jk = c k . (349)
j = 0 , (350)
with
j 0 := c , (351)
j := (j 0 , j k ) . (352)
66
With 0 = /(ct) and the definitions for the -matrices
0 := , (354)
k := k , (355)
this becomes
mc
i + = 0. (356)
~
Remarks:
Hermiticity: 0 is hermitian and k is anti-hermitian.
0 = = = 0 , (357)
k = k = (k ) = k = k = k . (358)
( 0 )2 = ()2 = 12 , (359)
2
k = k k = k k = 12 , (360)
0 k + k 0 = k + k = 0 , (361)
k l + l k = k l + l k = 0 (l 6= k) . (362)
{ , } = + = 2g 14 . (363)
Co-Variant -matrices:
= g . (364)
B ~ ~ ,
/ := B = B = B 0 0 B (365)
68
13 Lorentz-invariance of the Dirac equation
Chapter [15.3] of [1]
69
These transformations consist of translations (a ) and the homogeneous
Lorentz-transformations ( ), which can again be split up in Rotations
(in space), Lorentz-boosts and some discrete transformations like parity
(space-reflection) or time-reflection5
Remarks:
= = g (378)
= g (379)
!
=
g , (380)
g = g gt = g . (381)
The last equation actually defines the homogeneous Lorentz group and
it gives
det = 1 , (382)
0 0 1 or 0 0 1 . (383)
70
and of Lorentz-boosts.
An example for a space rotation is e.g.
1 0 0 0
0 cos sin 0
R12 = 0 sin cos 0
,
(384)
0 0 0 1
1 0 0
cosh sinh 0 0 1 2 1 2
sinh cosh 0 0 1 0 0
L01 = = 1 2 1 2 ,
0 0 1 0
0 0 1 0
0 0 0 1
0 0 0 1
(385)
with = v/c and tanh = . The representation of the full Lorentz-
group is described in detail in [5].
x = x + a , (387)
(x ) = S()(x) , (388)
= , (390)
(x) = S()1 (x ) , (391)
71
to get
mc
i + S()1 (x ) = 0 (392)
~
mc
iS() S()1 + (x ) = 0 . (393)
~
In the second line we have multiplied everything with S() from the r.h.s. .
The Dirac equation is now form invariant, if
S() S()1 = (394)
= S()1 S() . (395)
Remarks:
Any four component that transforms as given by Eq.(388) is called a
Dirac-spinor.
The transformation matrix S is defined by Eq.(395).
A detailed construction of S is described in detail e.g. in [5]. Here two
examples.
1. A space rotation around the z-axis:
1 0 0 0
0 cos sin 0
R12 = 0 sin cos 0
0 0 0 1
cos 2 + i sin 2
0 0 0
0
cos 2 i sin 2 0 0
SR12 = .
0 0 cos 2 + i sin 2 0
0 0 0 cos 2
i sin 2
(396)
Note that the matrix R12 has a periodicity of 2, while SR12 has
a periodicity of 4.
2. A Lorentz-boost along the x-axis
cosh sinh 0 0
sinh cosh 0 0
L01 =
0 0 1 0
0 0 0 1
sinh 2
cosh 2 0
0 cosh 2 sinh 2 0
SL01 = 0
. (397)
sinh 2 cosh 2 0
sinh 2 0 0 cosh 2
72
13.3 Bilinear forms
The adjoint spinor is defined as
= 0 . (398)
It transforms as
= S = S (399)
= S 0 . (400)
One can show (e.g. [5]) that
S 0 = b 0 S 1 , (401)
with
+1 for 00 +1
b= . (402)
1 for 00 1
Thus we get for the transformation of the adjoint spinor
= S 1 , (403)
if we only consider transformations that do not change the direction of time.
Now we can derive the transformation properties of the following bilinears
= ,
= , (404)
= ,
with
i
:=
[ , ] . (405)
2
Thus we get the following transformation properties
1. transforms as a Lorentz-scalar.
2. transforms as a Lorentz-vector. alone is not a Lorentz-
vector! It is always given by the Definition in Eq.(363).
3. transforms as a Lorentz-tensor.
73
14 Solutions of the Dirac equation
Chapter [15.4/6] of [1]
14.1 Plane wave solutions of the Dirac equation
We are looking now for solutions of the free Dirac equation of the form
~~
p r Et p x
(~r, t) = A u ei ~ = A u ei ~ . (406)
A is a constant, u is a four component spinor being independent from the
space-time coordinate x . Inserting this ansatz into Eq.(319) we get
Eu = ck pk + mc2 u .
(407)
To get a first idea about the structure of this solution we consider the case
of a particle at rest, i.e. ~p = 0. We get
mc2 I2
2 0
Eu(0) = mc u(0) = u(0) . (408)
0 mc2 I2
This equation has two different eigenvalues E+ = +mc2 and E = mc2 ,
that occur each twice and the following four eigenvectors
1 0 0 0
0 1 , u(3) (0) = 0 , u(4) (0) = 0 .
u(1) (0) = (2)
0 , u (0) = 0 1 0
0 0 0 1
(409)
The first two spinors describe a spin 1/2 particle and the second two a spin
1/2 anti-particle.
The positron, the anti-particle of the electron was discovered in 1932, about
4 years afterwards the Dirac-equation was postulated. This created also a
whole new field of science-fiction literature, see e.g. [23].
If we have a solution of the form c1 u(1) (0) + c2 u(2) (0) + c3 u(3) (0) + c4 u(4) (0)
in the system, where the particle is at rest ~p = 0, then we get in the system
that has a finite momentum ~p in x-direction
~
r E t p x
i p~ i
(~r , t ) = A u e ~ = Au e ~ , (410)
with
cosh 2 sinh 2
0 0
0 cosh +c3 sinh 2 +c4 0 .
u = Su = c1
+c2 2
0 sinh cosh 0
2 2
sinh 2 0 0 cosh 2
(411)
74
14.2 Non-relativistic limit of the Dirac-equation
To discuss the non-relativistic limit of the Dirac equation we start from the
free Dirac-equation and couple the electro-magnetic field via minimal cou-
pling.
h k e i
i~ = c pk Ak + mc2 + e . (412)
t c
Next we decompose the four-component Dirac-spinor into two two-component
spinors and .
= (413)
Inserting this into Eq.(412) we get
~ ~ 2
i~ =c + e + mc , (414)
t ~ ~
Next we make in the second line of the equation above the following assump-
tion
i~t , e 2mc2 (418)
to get
~ ~ v
= =O . (419)
2mc c
Thus is denoted the large component of the Dirac-spinor and as the small
component. Next we insert Eq.(419) into the first line of Eq.(417) and we
get
1
i~ = (~ ~ )(~ ~ ) + e . (420)
t 2m
75
Using the identity
one gets
(~ ~ )(~ ~ ) = 2 + i~ (~ ~ ) (422)
e~ ~ .
= 2 ~ B (423)
c
So we get for the large component
1 e ~ 2 e~ ~
i~ = ~p A ~ B + e . (424)
t 2m c 2mc
This is the Pauli-equation that describes non-relativistic electrons with
spin 1/2. Moreover the Dirac equation gives automatically the correct Lande
factor g = 2 of the electron, which had to be introduced by hand to the Pauli-
equation in order to be compatible with experiment. We can introduce the
Lande factor in the above equation via ~ g~ /2.
~
Looking at the special case of a constant B-field
~ = 1B
A ~ ~x , (425)
2
~ =
B ~ A ~, (426)
we arrive at
2
e2 ~ 2
p~ e ~ ~ ~
i~ = (L + 2S) B + A + e , (427)
t 2m 2mc 2mc2
with S~ = ~~ /2 and L
~ = ~x p~. The factor 2 in front of S
~ in Eq.(427) repre-
sents the Lande factor g.
More precise measurements have shown, however, that the Lande-factor de-
viates slightly from 2, thus the anomalous magnetic moment of the electron
is introduced as
g2
a= . (428)
2
This quantity is measured very precisely by now. Theoretical deviations from
the prediction of the Dirac equation (a = 0) arise as soon as one is considering
the effects of Quantum electro dynamics (QED). The most precise values are
currently [24, 25]
76
This is an amazing agreement! This is the most precise tested prediction in
science.
To reach this, accuracy 12672 5-loop diagrams have to be calculated [25] in
QED.
Nowadays the above comparison is used to fix the numerical value of the
electro-magnetic coupling, e = e2 /(4).
77
Newtons law can be derived from the Lagrange function via the Euler-
Lagrange equations.
F = 0 (435)
D = ieA (438)
78
15 Solutions of the Dirac equation for a cen-
tral potential
Chapter [15.5] of [1]
Next we would like to separate the variables as in the case of the Schrodinger-
equation. There we had L ~ 2 and Lz as conserved quantum numbers and we
found solutions that are simultaneous eigenvectors of H, L ~ 2 and Lz . Since
the Dirac-equation always contains spin, we expect now that not the angular
momentum L ~ is conserved, but the total angular momentum J~ = L ~ + S.
~ An
explicit calculation gives
~ = i~c(~
[H, L] ~p) 6= 0 , (442)
~ = ~
S ~ , (443)
2
with
k 0
k = . (444)
0 k
Remarks: this definition gives
The angular momentum commutation relation is fulfilled
~ 2 read
The eigenvalues of S
~ 2 = 3 ~2 = s(s + 1)~2 ,
S (446)
4
with s = 1/2.
79
The commutator of the spin operator with the Hamiltonian reads
~ = +i~c(~
[H, S] p~) 6= 0 . (447)
Comparing Eq.(442) with Eq.(447) we get the conservation of the total an-
gular momentum
~ =0.
[H, J] (448)
One can also show that
[H, J~2 ] = 0 . (449)
Thus we can find wave functions that are simultaneous eigenfunctions of H,
J~2 and Jz , with the eigenvalues E, j(j + 1)~2 and mj ~.
We make the following ansatz for the Dirac-equation with a spherical poten-
tial !
iGlj (r) l
(, )
ljmj (r, , ) = r
iFlj (r) ~
x
jml
j
(450)
r
~
x
jmj
(, )
with the 2-component spinors
( (+) 1
jmj (, ) for j = l +
ljmj (, ) = ()
2
1
(451)
jmj (, ) for j = l 2
and q
l+mj + 12
Y 1 (, )
(+)
jmj (, ) = q 2l+1 1 l,mj 2 , (452)
lmj + 2
2l+1
Yl,mj + 1 (, )
2
q
lmj + 12
Y 1 (, )
()
jmj (, ) = q 2l+1 1 l,mj 2 . (453)
l+mj + 2
2l+1
Yl,mj + 1 (, )
2
Y are the spherical harmonics and Glj (r) and Flj (r) are some unknown func-
tions that describe the radial dependence. They obey the following differen-
tial equation
E m V (r) dFlj (r) 1 Flj (r)
Glj (r) = j+ (454)
~c dr 2 r
E + m V (r) dGlj (r) 1 Glj (r)
Flj (r) = + j+ (455)
~c dr 2 r
for j = l 1/2. The radial equation are solved by the ansatz
80
In the case a hydrogen-like problem we have V (r) = Z/r, with =
e2 /(40 ~c) = 1/137 and we get then the following energy eigenvalues
mc2
En,j = s 2 , (458)
1+ qZ
2
n(j+ 21 )+ (j+ 21 ) (Z)2
Z 2 2 (Z)4
2 1 3 6 6
En,j = mc 1 +O Z . (459)
2n2 2n3 j + 12 4n
The first term is just the rest mass, the second term is the result of the
Schrodinger theory for the hydrogen atom and the third term is the first
relativistic correction, that gives a fine structure splitting.
More details of this calculation can be found in [5].
81
16 Measurements and interpretation
Chapter [17.1-4] of [1]
16.1 Introduction
Quantum mechanics is one of the most successful concepts of physics. It is
the foundation of many technical applications like Computer, Laser,... The
theory is thoroughly tested by experiments and all the predictions agree
perfectly with the measurements. In that sense there is no doubt that
quantum mechanics is the correct language to describe microscopic
effects, at least in the energy range, that is currently accessible to us.
The inherent statistical nature of quantum mechanical predictions as well
as the quantum mechanical concept of measurements has, however, always
raised fundamental questions about the interpretation of the theory. To
some extend this is a philosophical question, depending also strongly on the
definition of the meaning of the word understanding. Nevertheless, we
would like to conclude this lecture course with a brief overview of different
interpretations of quantum mechanics.
Some reviews of this topic can also be found e.g. in [26, 27, 28].
82
2. States, particles,... are described by wave functions (states vectors).
A general wave function can be written as a linear combination of
the |ni: X
= cn |ni . (461)
n
Here two concepts arise, that are not obvious to understand: the inherent
probabilistic nature of quantum mechanics (point 4) and the nature of the
measurements (point 5), or in other words: what is actually the collapse of
the wave function?
To become more concrete, let us assume that our concern is to measure the
position of a particle.
Quantum mechanics gives us only the probability for finding the parti-
cle at a certain location. The probability of finding the particle at time
t0 at the location ~x0 is given by |(~x = ~x0 , t = t0 )|2 .
Now we know for sure that the particle was at the time t0 at the position
~x0 and we can ask ourselves next: where was the particle slightly before
the measurement?6 This question is not governed by our general rules of
quantum mechanics, some possible answers are [7]:
83
2. nowhere (orthodox position);
Remarks:
1. If the particle was actually at the position ~x0 and we just cannot de-
scribe this with quantum mechanics, then this is a kind a failure of
the current formulation. In that case there will be a more fundamental
theory, that contains all the results from quantum mechanics and in
addition knows, where the particle was before the measurement.
This was e.g. the position of Albert Einstein.
Since such a new theory is expected to contain more variables, they are
also called hidden variable theories. They were e.g. investigated by
Bohm [29, 30].
How does the fundamental theory look like?
0 e+ + e . (463)
7
This example actually stems from [32].
84
If the pion is at rest then the electron-positron pair is described by the
following wave function (spin = 0)
1
(| i | i) . (464)
2
Next we assume that in some distance xe of the former position of the pion,
the spin of the electron will be measured at the time te > t0 .
Let us assume that the outcome is spin-up. Thus it is clear that any future
measurement of the spin of the positron will give spin-down.
How can this be interpreted?
Remarks:
Somehow it seems that the information about the collapse of the wave
function is transported at the time te instantly from the location xe to
the position of the positron. Einstein and friends considered this to be
obviously impossible, because this would violate locality (spukhafte
Fernwirkung ghost-like long-distance action.
85
16.4 Bells equation
In 1964 Bell [33] was generalising the above experiment and has shown that
there is a measurable difference between position 1 and 2 (this excludes by
definition position 3). In addition he pointed out severe problems of theories
with hidden variables, giving thus evidence for the orthodox position, albeit
locality has to be given up to some extent and we are still left with a lack of
understanding of the measurement process.
We take the same experimental setup as in the EPR case, with the gen-
eralisation that now the spin-detectors can be moved in arbitrary directions.
In the EPR case both were measuring the z-component of the spin, while
the electron-spin component is measured relative to the vector ~a and the
positron-spin component is measured relative to the vector ~b. In the end the
relevant observable will be the quantum mechanical average of the product
of the two spins P (~a, ~b). The case ~a = ~b gives the result of the EPR-case. In
the EPR case quantum mechanics told us that the product will always give
1 1. Thus we have
P (~a, ~a) = 1 . (465)
If the vectors are anti-parallel ~a = ~b, we will obviously get
P (~a, ~a) = +1 . (466)
For a general direction of the vectors quantum mechanics will give
P (~a, ~b) = ~a ~b . (467)
It will turn out, however, that this simple result is incompatible with theo-
ries with hidden variables. To show that, we consider very general models
with hidden variables, assuming only that reality is given by the quantum
mechanical wave function and some hidden variable .
A(~a, ) gives the result of the measurement of the electron spin relative
to the vector ~a. It can take the values 1.
B(~b, ) gives the result of the measurement of the positron spin relative
to the vector ~b. It can take the values 1.
If ~a = ~b, we have
A(~a, ) = B(~a, ) . (468)
The average of the product of the two measurements is given by
Z
P (~a, b) = ()A(~a, )B(~b, )d ,
~ (469)
86
with () being the unknown probability distribution of the hidden variable.
We only demand
0 () 1 , (470)
Z
()d = 1 . (471)
to get
Z h i
|P (~a, ~b) P (~a, ~c)| () 1 A(~b, )A(~c, ) d (478)
This relation can easily be violated, assume e.g. ~a, ~b and ~c lie in a plane, ~a
and ~b are perpendicular and ~c lies in between them (with an angle of 45 ).
Then we get
|0 1/ 2| = 0.707 1 1/ 2 = 0.29 , (481)
87
which is obviously not correct.
Combing this result - hidden variable are not compatible with quantum me-
chanics8 - with the result from EPR have two possible resolutions:
All experimental tests have shown that Quantum mechanics is correct and
that Bells inequality is violated. It seems that locality does not hold and
that there is an instantaneous collapse of the wave function.
We are still left with lack of understanding of the collapse of the wave func-
tion. Scientist tried to establish the collapse of the wave-function experimen-
tally; this comes under the name of quantum Zeno effect, but there are
currently no convincing hints.
8
One can, however, construct some fancy hidden variable models for which Bells in-
equality is not applicable.
88
17 Revision Lecture
The course covered four topics
1. Scattering
17.1 Scattering
17.2 Quantum statistical mechanics
17.3 Relativistic Quantum mechanics
17.4 Interpretation of Quantum mechanics
89
90
References
- Reference text -
[1] Quantum Mechanics
Brian Harold Bransden, C. J. Joachain
Prentice Hall (28th January 2000)
ISBN-10: 0582356911
ISBN-13: 978-0582356917
- Recommended further reading -
91
ISBN-10: 0131911759
ISBN-13: 978-0131911758
[10] P. W. Higgs
Broken symmetries, massless particles and gauge fields
Phys. Lett. 12 (1964) 132
2962 citations counted in INSPIRE as of 20 Jan 2015
[11] P. W. Higgs
Broken Symmetries and the Masses of Gauge Bosons
Phys. Rev. Lett. 13 (1964) 508
3037 citations counted in INSPIRE as of 20 Jan 2015
[13] P. W. Higgs
Spontaneous Symmetry Breakdown without Massless Bosons
Phys. Rev. 145 (1966) 1156
2145 citations counted in INSPIRE as of 20 Jan 2015
[14] T. W. B. Kibble
Symmetry breaking in non abelian gauge theories
Phys. Rev. 155 (1967) 1554
1106 citations counted in INSPIRE as of 20 Jan 2015
92
[16] S. Chatrchyan et al. [CMS Collaboration]
Observation of a new boson at a mass of 125 GeV with the CMS exper-
iment at the LHC
Phys. Lett. B 716 (2012) 30 [arXiv:1207.7235 [hep-ex]]
3796 citations counted in INSPIRE as of 20 Jan 2015
[21] P. A. M. Dirac,
The Quantum theory of electron,
Proc. Roy. Soc. Lond. A 117 (1928) 610.
572 citations counted in INSPIRE as of 20 Jan 2015
93
Theoretical update of Bs Bs mixing,
JHEP 0706 (2007) 072 [hep-ph/0612167].
504 citations counted in INSPIRE as of 20 Jan 2015
[23] http://angelsanddemons.web.cern.ch/antimatter
[26] S. Weinberg,
Testing Quantum Mechanics,
Annals Phys. 194 (1989) 336.
166 citations counted in INSPIRE as of 20 Jan 2015
[27] R. Omnes,
Consistent interpretations of quantum mechanics,
Rev. Mod. Phys. 64 (1992) 339.
254 citations counted in INSPIRE as of 20 Jan 2015
[29] D. Bohm,
A Suggested interpretation of the quantum theory in terms of hidden
variables. 1.,
Phys. Rev. 85 (1952) 166.
637 citations counted in INSPIRE as of 20 Jan 2015
[30] D. Bohm,
A Suggested interpretation of the quantum theory in terms of hidden
variables. 2.,
Phys. Rev. 85 (1952) 180.
366 citations counted in INSPIRE as of 20 Jan 2015
94
complete?,
Phys. Rev. 47 (1935) 777.
1021 citations counted in INSPIRE as of 20 Jan 2015
[33] J. S. Bell,
On the Einstein-Podolsky-Rosen paradox,
Physics 1 (1964) 195.
555 citations counted in INSPIRE as of 20 Jan 2015
[34] H. Everett,
Relative state formulation of quantum mechanics,
Rev. Mod. Phys. 29 (1957) 454.
426 citations counted in INSPIRE as of 20 Jan 2015
95