A Review of Quantum Mechanics
Chun Wa Wong
Department of Physics and Astronomy, University of California, Los Angeles, CA 90095-1547
(Dated: September 21, 2006)
Review notes prepared for students of an undergraduate course in quantum mechanics at UCLA,
Fall, 2005 Spring, 2006. Copyright
c
2006 Chun Wa Wong
I. The Rise of Quantum Mechanics
1.1 Physics & quantum physics:
Physics = objective description of recurrent physi-
cal phenomena
Quantum physics = “quantized” systems with both
wave and particle properties
1.2 Scientific discoveries:
Facts Theory/Postulates Facts
1.3 Planck’s black body radiation:
Failure of classical physics:
Emissivity =
1
4
uc, λ
max
T = const,
energy density = u = ρ
ν
hEi,
hEi
exp
6= k
B
T, (1)
(ρ
ν
= density of states in frequency space).
Planck hypothesis of energy quantization (E
n
=
nhν) explains observation:
hEi =
P
n=0
ne
nx
P
n=0
e
nx
=
e
/k
B
T
1
, (2)
where x = /(k
B
T ).
1.4 Einstein’s photoelectric effect:
Photoelectrons ejected from a metal by absorbed
light of frequency ν have maximum energy:
E = W (work function) (3)
that depends on the frequency ν of light and not
its intensity.
1.5 Bohr’s atomic model:
Electron of mass m in a classically stable circular
orbit of radius r = a and velocity v around the
atomic nucleus:
mv
2
r
=
e
2
G
r
2
, or T =
U
2
, (4)
Electronic address: cwong@physics.ucla.edu
(e
2
G
= e
2
/4πǫ
0
in Gaussian units). Hence
E = T + U =
U
2
= T. (5)
Bohr’s two quantum postulates (L quantization and
quantum jump on photon emission):
L = mva = n~
~
~, n = 1, 2, ...,
= E
n
i
E
n
f
. (6)
For ground state:
E
1
=
U
1
2
=
e
2
G
2a
= 13.6 eV
= T
1
=
~
2
2ma
2
,
a =
~
2
me
2
G
= 0.53
˚
A (Bohr radius), (7)
For excited states:
r
n
= n
2
a, E
n
=
E
1
n
2
. (8)
1.6 Compton scattering:
Energy-momentum conservation for a photon of lab
energy E = hc scattered from an electron (of
mass m) initially at rest in the lab (before=LHS,
after=RHS; 1=photon, 2= recoiling e):
p = p
1
+ p
2
,
E + mc
2
= E
1
+
q
m
2
c
4
+ p
2
2
c
2
. (9)
Photon scattered into lab angle θ has longer wave-
length λ
1
:
λ
1
λ = λ
e
(1 cos θ), where
λ
e
=
h
mc
=
hc
mc
2
= 2.4 pm (10)
is the electron Compton wavelength.
1.7 de Broglie’s matter wave:
Massive matter and massless light satisfy the same
energy-momentum and momentum-wavelength re-
lations
E
2
= p
2
c
2
+ m
2
c
4
,
p = ~k =
h
λ
. (11)
2
Thus moving matter is postulated to be a wave with
a motional de Broglie wavelength λ = h/p, leading
to the diffraction maxima for both matter and light
of
2-sided formula: 2d sin θ = ,
1-sided formula: D sin θ = .
II. Waves and quantum waves
Classical and quantum waves have identical mathemat-
ical properties.
2.1 Waves in physics:
Waves rise and fall, travel, have coherence in space
and in time, interfere, and diffract.
Two classes of waves:
(a) Inertial waves in massive media: ocean waves,
sound waves, vibrations of violin strings;
(b) Noninertial waves in vacuum: EM waves,
matter waves.
2.2 Mathematical description of waves:
The 1-dimensional wave equation (1DWEq)
2
x
2
1
v
2
2
t
2
Ψ(x, t) = 0 (12)
is an equation of state that gives a wave function
Ψ(x, t) describing an unfolding event in spacetime.
The wave equation is a partial differential equation
(PDE) because it depends on more than one vari-
able, here the two variables x and t.
In contrast, the solution x(t) of an equation of mo-
tion such as a Newton equation describes how a
single point x(t) on an object evolves in time. Sim-
ilarly, a snapshot of an object is not necessarily a
wave. A wave function Ψ(x, t) is needed to describe
a wave’s coherent structure in both space and time.
However, a wave (any wave) can have particle prop-
erties if a point x can be defined on it such that one
can describe how its position changes in time.
The wave equation (12) describes linear waves sat-
isfying the superposition principle that a sum
of solutions is also a solution :
Ψ = a
1
Ψ
1
+ a
2
Ψ
2
. (13)
Traveling waves:
(a) f (xvt)= a wave traveling to the right (+x),
(b) g(x + vt)= a wave traveling to the left (-x),
(c) the composite wave a
1
f(x vt) + a
2
g(x + vt)
is also a wave.
LDEs (linear DEs) with constant coefficients
have exponential solutions:
d
dx
e
ikx
= ike
ikx
,
d
dt
e
iωt
= iωe
iωt
. (14)
These DE are called eigenvalue equations be-
cause the constant ik is called an eigenvalue of the
differential operator d/dx, while the solution e
ikx
is
called its eigenfunction belonging to its eigenvalue
ik.
The 1DWEq (12) has four distinct (i.e., linearly
independent) eigenfunctions:
Ψ(x, t) = e
±i(kx±ωt)
. (15)
Energy in both classical and quantum waves is car-
ried by the intensity or energy flux
I |Ψ|
2
. (16)
Its quadratic dependence on Ψ is responsible for the
interference between two traveling waves Ψ
i
=
A
i
e
i
with real A
i
:
|Ψ
1
+ Ψ
2
|
2
= |Ψ
1
|
2
+ |Ψ
2
|
2
+ 2Re(Ψ
1
Ψ
2
)
= A
2
1
+ A
2
2
+ 2A
1
A
2
cos(∆s),
where s = k
x
ω
k
t
. (17)
In the limit k 0,
dk
= v
g
= group velocity
6= v =
ω
k
= wave velocity. (18)
Examples: Certain waves satisfy the relation ω =
Ak
p
. Then v
g
/v = p. Specific examples are p = 2
for transverse vibrations of a thick bar, p = 3/2
for short ripples moving under surface tension, and
p = 1/2 for long waves on deep sea.
The result (17) is similar in structure to the squared
length of a sum of two vectors in space that inter-
sect at an angle θ:
|A
1
+ A
2
|
2
= A
2
1
+ A
2
2
+ 2A
1
A
2
cos θ, (19)
Two waves of the same amplitude traveling in op-
posite directions interfere form a standing wave:
Ψ = A[cos(kx ωt) + cos(kx + ωt)]
= 2A cos(kx) cos(ωt) (20)
with factorized space and time dependences.
3
2.3 Matter-wave quantization:
Light wave of momentum p and energy E = pc is
described by the wave function:
e
i(kxωt)
= e
i[(p/~)x(E/~)t]
. (21)
de Broglie and Schr¨odinger suggested that the RHS
expression holds also for m atter waves so that:
~
i
x
e
i(p/~)x
= ˆp e
i(p/~)x
= p e
i(p/~)x
,
i~
t
e
i(E/~)t
=
ˆ
H e
i(E/~)t
= E e
i(E/~)t
.(22)
The differential operators that appear are called
the momentum operator ˆp and the Hamiltonian (or
energy) operator
ˆ
H, respectively.
Commutation relations: The rules of differen-
tial calculus dictates that a differential operator
does not commute with its own variable:
x
x = 1 + x
x
, or
x
, x
= 1;
t
t = 1 + t
t
, or
t
, t
= 1. (23)
These results give the fundamental commuta-
tors of wave mechanics
[ˆp, x] =
~
i
, [
ˆ
H, t] = i~. (24)
2.4 Wave packet and the uncertainty principle:
A wave packet is a superposition of waves of the
type (21)
f(x, t) =
1
2π
Z
−∞
g(k)e
i[kxω(k)t]
dk, (25)
with a linear superposition function g(k) sometimes
called a spectral amplitude.
Expectation value (= mean value) of a property
A(x, t) of the wave:
hA(x, t)i
Z
−∞
ρ
P
(x, t)A(x, t)dx, where
ρ
P
(x, t) =
|f(x, t)|
2
R
−∞
|f(x, t)|
2
dx
(26)
is the normalized probablity density of finding the
wave at the spacetime point (x, t).
The expectation value of a time-independent prop-
erty B(k) of the wave can be calculated by using
the spectral amplitude g(k):
hB(k)i
Z
−∞
˜ρ
P
(k)B(k)dk, where
˜ρ
P
(k) =
|g(k)|
2
R
−∞
|g(k)|
2
dx
(27)
is the normalized probablity density of finding the
wave in k space”. (We shall not consider more
complicated properties that also depend on the
time t or the energy E.)
Uncertainty relations: A wave of any kind sat-
isfies the uncertainty relation
xk
1
2
, where
(∆x)
2
= h(x ¯x)
2
i, ¯x = hxi,
(∆k)
2
= h(k
¯
k)
2
i,
¯
k = hki. (28)
The expression for matter waves is usually written
as the Heisenberg uncertainty principle:
xp
~
2
. (29)
Spreading wave packets: A Gaussian wave
packet has the spectral amplitude
g(k) = e
αk
2
, with k =
1
2
α
. (30)
It gives rise to a wave function at time t = 0 of
f(x, t = 0) e
x
2
/(4α)
, with x =
α. (31)
Hence the wave packet at time t = 0 has the mini-
mal uncertainty product of
xk =
1
2
. (32)
For t 6= 0, however, the uncertainty product could
be greater than the minimal value of 1/2. This hap-
pens if the system is dispersive, meaning a nonzero
second Taylor coefficient β in the following expan-
sion:
ω(k) ω(k = 0) + k
dk
0
+
1
2
k
2
d
2
ω
dk
2
0
= ω(0) + kv
g
+
1
2
k
2
β. (33)
The resulting approximate wave function at any
time t can then be evaluated to the closed form
f(x, t) e
x
2
t
/(4α
t
)
, with
x
t
= x v
g
t, α
t
= α +
2
t. (34)
Thus the position x of the wave moves in time with
the group velocity
v
g
=
dk
0
. (35)
Its position uncertainty increases in time to
x(t) =
|α
t
|
α
=
α
"
1 +
βt
2α
2
#
1/2
. (36)
4
For a massive particle in free space E
2
= p
2
c
2
+
m
2
c
4
. Hence
β = ~
d
2
E
dp
2
= ~
m
2
c
6
E
3
6= 0. (37)
Thus the spatial width of the Gaussian packet of
a massive particle spreads out in both positive and
negative times.
III. Wave mechanics
Wave mechanics describes physics by using the
Schodinger wave eq. and its wave functions.
3.1 Schr¨odinger wave equation (SchWEq) :
First quantization: The classical E-p relation
E
2
= p
2
c
2
+m
2
c
4
is quantized into a wave equation
by replacing E, p by the differential operators
E
ˆ
H = i~
t
p ˆp =
~
i
x
: (38)
i~
t
2
Ψ(x, t) =
"
~
i
x
2
c
2
+ m
2
c
4
#
Ψ, (39)
where the wave function Ψ(x, t) has to be added for
the operators to operate on. In this way, a classical
state of motion “spreads out” into a wave structure.
Nonrelativistic limit: Likewise, the NR energy
E = T + V = p
2
/2m + V (not containing the rest
energy mc
2
) is quantized into the SchWEq:
i~
t
Ψ(x, t) =
"
1
2m
~
i
x
2
+ V (x)
#
Ψ, (40)
Commutation relations: A differential operator
does not commute with its variable. Hence
[ˆp, x] =
~
i
, [
ˆ
H, t] = i~. (41)
Group velocity: E
2
= p
2
c
2
+ m
2
c
4
EdE =
c
2
pdp. Hence
vv
g
=
E
p
dE
dp
= c
2
. (42)
3.2 Born’s interpretation of |Ψ|
2
:
The squared wave function gives the probabililty
density of finding system at spacetime location
(x, t):
|Ψ(x, t)|
2
= ρ
P
(x, t) (43)
in the sense that the probability for finding the
state somewhere in space at time t is
P (t) =
Z
−∞
ρ
P
(x, t)dx
=
Z
−∞
|Ψ(x, t)|
2
dx = 1. (44)
By choosing P (t) = 1, the wave function is said to
be normalized.
3.3 Conservation of |Ψ|
2
:
The probability density can be moved around, but
cannot be created or destroyed in the absence of
explicit creative or destructive physical processes:
0 =
d
dt
ρ
P
(x, t) =
t
+
dx
dt
x
ρ
P
(x, t)
=
t
ρ
P
(x, t) +
x
j
P
(x, t). (45)
This is called the continuity equation for the con-
served probability current density
j
P
(x, t) = v ρ
P
(x, t), (46)
where v = dx/dt.
IV. Wave mechanics in one spatial di-
mension
Wave mechanics is applied to simple one-dimensional
problems to illustrate how the SchWEq is solved and how
its wave character affects the properties of a system.
4.1 Solving the Schr¨odinger wave equation
(SchWEq) :
Separation of varables: Since solutions are
unique up to a superposition, we may look for a
solution of the factorized form
Ψ(x, t) = ψ(x)φ(t). (47)
Then
ψ(x)
ˆ
Hφ(t)
ψ(x) φ(t)
=
φ(t)
ˆp
2
2m
+ V
ψ(x)
φ(t) ψ(x)
= E. (48)
On simplification, the first expression is clearly a
function of t only, while the middle expression is a
function of x only. Hence the separation constant
E must be a constant independent of t or x.
The PDE can thus be broken up into two ordinary
DEs:
ˆ
Hφ(t) =
i~
t
φ(t) = Eφ(t),
"
1
2m
~
i
x
2
+ V (x)
#
ψ(x) = Eψ(x), (49)
5
x = 0
x = 0
E
V
0
- V
0
0
k
2
q
2
q
2
BCs: Ψ
= Ψ
Ψ '
= Ψ '
FIG. 1: Wave numbers and wave functions at a potential step.
where E appears in both ODEs as the energy
eigenvalue. The rst ODE gives the time-
dependent factor
φ
E
(t) = e
i(E/~)t
, (50)
but the second equation is a second-order ODE
whose solution requires the imposition of two
Boundary conditions at a suitable boundary x =
x
0
:
ψ
<
(x
0
) = ψ
>
(x
0
),
ψ
<
(x
0
)
d
dx
ψ
<
(x)
x
0
= ψ
>
(x
0
). (51)
Note that if the first BC is violated, ψ
(x
0
) becomes
infinite and the DE (differential eq.) cannot be sat-
isfied. If the second BC is violated, ψ
′′
(x
0
) becomes
infinite. The DE also cannot be satisfied, except for
the special case where V (x
0
) has a compensating
infinity at x
0
.
4.2 Particle in a box of length L:
With the two BCs ψ(0) = ψ(L) = 0, the solutions
turn out to be the same as those describing classical
vibrations of a violin string of length L:
ψ
n
(x) =
r
2
L
sin
L
x
(52)
for a normalized wave function.
4.3 Potential step:
Figure 1 shows a particle of mass m incident from
the left with energy E. The wave numbers k, q are
defined as follows:
E =
~
2
k
2
2m
, E V =
~
2
q
2
2m
; (53)
or k
2
= q
2
k
2
0
, where k
2
0
= 2mV
0
/~
2
. The parti-
cle’s matter wave is partially reflected at the bound-
ary x = 0, and partially transmitted across it (with
no further reflection downstream):
x < 0 : ψ
<
(x) = e
ikx
+ Be
ikx
,
x > 0 : ψ
>
(x) = A
T
e
iqx
. (54)
x = 0 x = L
E
V
0
Κ
2
0
k
2
q
2
FIG. 2: Wave numbers for a rectangular potential barrier.
Its probability current density is conserved across
the boundary:
j
<
=
~k
m
1 |B|
2
= j
>
=
~q
m
|A
T
|
2
. (55)
Its reflection and transmission coefficients at the
boundary are:
R = |B|
2
, T = 1 R. (56)
4.4 Potential barrier:
A particle of incident energy E higher than the
height V
0
of a rectangular potential barrier is par-
tially reflected and partially transmitted by a potet-
nial barrier. The wave functions for x < 0 before
the barrier and for x > L after the barrier remain
the same as Eq.(54), but the presence of the sec-
ond boundary at x = L causes reflected waves to
be present in the barrier region
0 < x < L : ψ
<
(x) = Ce
ikx
+ De
ikx
. (57)
The two boundary conditions at the second bound-
ary x = L are enough to determine the two extra
coefficients C and D. The reflection coefficient re-
mains R = |B|
2
, but its value is changed. The
transmission coefficient is
T = 1 R =
1
1 +
k
2
q
2
2kq
2
sin
2
(qL)
. (58)
Quantum tunneling: The formula for T works
even when the energy E falls below the height V
0
of
the potential barrier, as shown in Fig. 2. Then q
2
=
κ
2
< 0, and q = becomes purely imaginary.
The function [sin(qL)/q]
2
in the formula is analyti-
cally continued to the function [sinh(κL)]
2
. The
resulting transmission coefficient T across the finite
barrier is shown in Fig. 3. The fact that T does not
vanish for E < E
0
, as it does in classical mechanics,
is called quantum tunneling. Note that tunneling
can occur only when the barrier thickness is finite.
6
1 2 3 4 5
EV
0
0.2
0.4
0.6
0.8
1
T
FIG. 3: Transmission coefficient across a rectangular potential
barrier.
Alpha
emission Fusion
FIG. 4: Tunneling across the Coulomb barrier in α decay and
in nuclear fusion.
4.5 Applications of quantum tunneli ng:
Alpha decay: An α particle inside a nucleus feels
an attractive nuclear potential. As the α particle
leaves the nucleus in alpha decay, it sees the repul-
sive Coulomb potential outside the nucleus. These
two potentials combine to form a potential barrier
that the α particle must tunnel through on its way
out, as illustrated in Fig. 4. The decay probablility
is thus fT , the product of the frequency f of the
α particle’s hitting the inside wall of the potential
barrier and the transmission coefficient T through
the potential barrier (that gives the tunneling prob-
ability per attempt).
Nuclear fusion: is just the time reversed situation
of emission. Here a charged particle, say a proton,
is trying to enter a nucleus from the outside by tun-
neling through its very high and thick Coulomb po-
tential barrier. The transmission coefficient is typi-
cally very small, but increases very rapidly with the
energy E. If a number of atomic nuclei are exposed
to a gas of protons of increasing temperature, the
high-energy tail of the Boltzmann energy distribu-
tion of the proton gas will allow more and more
protons to get into atomic nuclei to initiate nuclear
fusion reactions. For example, the pp reaction
p + p d + e
+
+ ν (59)
(where e
+
is the positron and ν is a particle called
HaL
-W
Sample
External
field
HbL
-W
Sample
Scanning
tip
FIG. 5: Tunneling through a finite potential barrier (a) in
cold emission, and (b ) in a scanning tunneling microscope.
the electron neutrino) powers the conversion of
mass into solar energy in the interior of the Sun.
The reaction rate depends on the temperature T
as T
4
.
The luminosity of a star usually increases as M
3
of
its mass M, or T
24
of its central temperature. Such
an unusually strong dependence led Bethe in 1938
to look for another source of hydrogen burning. He
discovered that this alternative route is a certain
CNO cycle that involves certain carbon, nitrogen
and oxgen isotopes in intermediate steps.
Cold emission: Electrons are bound in metals by
their work function W . However, they can be in-
duced to tunnel out of the metal by applying an
external electric field to change the potential to
a triangular penetrable potential, as illustrated in
Fig. 5(a).
In s canning tunneling microscopes, the tunnel-
ing is induced by the proximity of the scanning tip,
as shown in Fig. 5(b). Since the tunneling current
increases greatly as the thickness of the potential
barrier decreases, one has a very sensitive device for
mapping the geometrical shape of a surface down
to atomic dimensions.
4.6 Bound states in a potential well:
If V (x) constant as x ±∞, the lower of these
two contant values is usually taken to be the zero
of the energy scale. Two situations can then be
realized:
(a) A system with E > 0 has continuous energy
spectrum with wave functions that are not lo-
calized or square-integrable, while
(b) a system with E < 0 has a discrete energy
spectrum with localized wave functions that
are square-integrable. The resulting physical
states are called bound states in a potential
well.
The wave numbers associated with a bound state in
an attractive one-dimensional square-well potential
7
x = -a x = a
E
-V
0
Κ
2
V=0
q
2
FIG. 6: Wave numbers associated with a bound state in an
attractive squ are-well potential.
are summarized in Fig. 6, where
|E| = E =
~
2
κ
2
2m
, E V =
~
2
q
2
2m
. (60)
Such an attractive one-dimensional potential has
at least one bound state, the ground state of
even parity whose wave function is proportional to
cos qx. This is because such a wave function al-
ready “bends” over at the boundaries, and can al-
ways be matched to a decreasing exponential func-
tion outside the potential well.
V. The mathematical structure of quan-
tum mechanics
Physical states in quantum mechanics are linear waves.
Both quantum states and classical linear waves can be
superposed to form other waves, in the same way that
ordinary vectors in space can be added to form other
vectors.
5.1 Fourier transform (FT):
The superposition principle for linear wave func-
tions (solutions of linear wave equations) is formal-
ized in the methods of Fourier series and Fourier
transforms (FT). Both methods and their general-
izations are used in quantum mechanics.
The FT of a wave function f(x) in one spatial vari-
able x is defined by the integral
F{f(x)} g(k) =
1
2π
Z
−∞
f(x)e
ikx
dx. (61)
For example,
F
d
dx
f(x)
= ikg(k)
ip
~
g(k). (62)
Thus the momentum operator ˆp = (~/i)/∂x yields
the momentum value p = ~k associated with the
spectral amplitude g(k).
Fo uri er inversion formula:
F
1
{g(k)} f(x) =
1
2π
Z
−∞
g(k)e
ikx
dk. (63)
This is how a wave packet is constructed, namely
by taking a linear superposition of waves. Also
F
1
d
dk
g(k)
= ixf(x). (64)
Thus a position operator ˆx = i∂/∂k gives the value
of the position x associated with the wave function
f(x).
The FT shows that waves can be described either
by wave functions f (x) in space, or by their spectral
amplitudes g(k) in k-space. Functions of one space
can be constructed from those in the other space.
Hence wave properties are completely specified ei-
ther in ordinary space or in k-space, a simplifica-
tion” that follows from a wave’s coherent structure.
Table I gives some simple examples of FTs.
Dirac δ-function: Since
F
1
{F{f(x)}} = f(x), (65)
we find (by writing out the two integral transforms
on the LHS in two steps) that
Z
−∞
δ(x x
)f(x
)dx
= f(x), (66)
where
δ(x x
) =
1
2π
Z
−∞
e
ik(xx
)
dk (67)
TABLE I: Some Fourier t ransforms.
Property If f(x) is then F{f(x)} is
f(x) g(k)
e
αx
2
1
2α
e
k
2
4α
unit step function
a
1
2π
2
k
sin ka
Derivatives (
d
dx
)
n
f(x) (ik)
n
g(k)
Derivatives (ix)
n
f(x) (
d
dk
)
n
g(k)
Translation f(x a) e
ika
g(k)
Attenuation f(x)e
αx
g(k )
CC
b
Real g
(k ) = g(k)
Real and even g
(k ) = g(k) = g(k)
Real and odd g
(k ) = g(k) = g(k)
a
from a to a
b
Complex conjugation
8
is called a Dirac δ-function.
Dirac has given a simple intuitive construction of
his δ-function as a rectangular function of width ǫ
and height 1 with a unit area under it:
δ(x x
) = lim
ǫ0
1/ǫ, when |x x
| ǫ/2,
0, otherwise.
(68)
5.2 Operators and their eigenfunctions:
When a differential operator operates on its eigen-
function, it gives a function that is proportional
to the eigenfunction itself:
ˆp e
ikx
=
~
i
x
e
ikx
= ~k e
ikx
,
ˆx e
i(x/~)p
=
i~
p
e
ixp/~
= x e
ixp/~
. (69)
The proportionality constant is called its eigen-
value.
Two functions or eigenfunctions can be combined
into an inner product:
(φ, ψ)
Z
−∞
φ
(x)ψ(x)dx
=
Z
−∞
g
φ
(k)g
ψ
(k)dk.
Dirac denotes this inner product by the bracket
symbol hφ|ψi.
5.3 Dirac notation:
Functions in function space and states of a physi-
cal system in quantum mechanics are mathemati-
cally similar to vectors in space. Dirac emphasizes
this similarity using a mathematical notation that
works for vectors, functions and quantum states.
A vector (function, state) can be expanded in terms
of basis vectors (functions, states). Dirac distin-
guishes between two kinds of vectors (functions,
states) column vectors (functions, ket states)
and row vectors (complex-conjugated functions,
bra states), where the name bra and ket comes from
the two parts of the word “bracket”:
|ψi = ket = ψ
1
|1i + ψ
2
|2i
=
ψ
1
ψ
2
= column vector
hφ| = bra = φ
1
h1| + φ
2
h2|
= (φ
1
, φ
2
) = row vector
= |φi
, (70)
For simple vectors, the basis states |1i and |2i are
just the unit basis vectors usually denoted e
1
and
e
2
, respectively. The
symbol is called an adjoint
operation. It combines the matrix operation of
transposition with the complex conjugation of all
“matrix elements”.
The Dirac bracket hφ|ψi is a scalar product of
vectors or an inner product of functions
hφ|ψi = (φ
1
, φ
2
)
ψ
1
ψ
2
= φ
1
ψ
1
+ φ
2
ψ
2
, (71)
where the complex conjugation on the bra com-
ponents is needed to keep the self scalar product
hψ|ψi nonnegative. For two orthonormal vectors
or states:
hi|ji = δ
ij
. (72)
Eigenbras and ei genkets: One goes beyond sim-
ple vector spaces by using the eigenfunctions of op-
erators such as position ˆx and momentum ˆp. In
fact, we immediately associate their eigenfunctions
with abstract eigenstates in the following eigen-
bra or eigenket form:
ˆx|xi = x|xi, and hx|ˆx = x
hx| = xhx|,
ˆp|pi = p|pi, and hp|ˆp = p
hp| = php|, (73)
where the eigenvalues x and p are real, as is true for
all physical observables, i.e., properties that can be
measured experimentally. The eigenvalues x and p
normally range continuously from to .
Projection and quantum destructive mea-
surement: A simple 2-dimensional vector |Ai can
be decomposed into its components A
i
= hi|Ai with
respect to the basis |ii:
|Ai = |1iA
1
+ |2iA
2
=
N
X
i=1
|iihi|
!
|Ai. (74)
In a similar way, an N-dimensional vector space is
made up of the N one-dimensional subspaces pro-
jected out by the projection operators
P
i
= M
i
= |iihi| (75)
in the sense that
P
i
|Ai = A
i
|ii (76)
leaves only that part of the vector that is in this
ith subspace. P
i
is also called a measurement sym-
bol (hence M
i
). This measurement is not a classi-
cal nondestructive measurement but a quantum de-
structive measurement that destroys all other com-
ponents of the original vector |Ai. The projection
collapses the state |Ai into just the part A
i
|ii in
the ith subspace.
Completeness relations: The original vector |Ai
can be restored by adding up all its projections
9
A
i
|ii. Hence an identity operator can be con-
structed for this linear vector space
1 =
N
X
i=1
|iihi| (77)
for discrete basis states that satisfy the orthonor-
mality relation hi|ji = δ
ij
.
The idea of completeness or closure holds also for
eigenstates of operators such as ˆx and ˆp that have
continuous eigenvalues:
1 =
Z
−∞
|xihx|dx
=
Z
−∞
|pihp|dp. (78)
The resulting basis states |xi or |pi are said to be
continuous basis states.
Representations: A discrete representation is a
specification of a vector (or state) |Ai by its com-
ponents A
i
in a discrete basis. A continuous rep-
resentation uses the continuous basis states |xi to
give the components A(x) = hx|Ai in the expansion
|Ai =
Z
−∞
dx |xiA(x). (79)
Wave functions: The (position) wave function of
a state or wave |ψi is just its component ψ(x) =
hx|ψi in the position representation. Its compo-
nent ψ(p) = hp|ψi in the momentum representa-
tion is called its momentum wave function. A wave
function can be expanded into an arbitrary discrete
representation of basis states |ii
hx|ψi ψ(x) =
N
X
i=1
hx|iihi|ψi
=
N
X
i=1
φ
i
(x)ψ
i
, (80)
where φ
i
(x) = hx|ii is the wave function of the basis
state |ii. Incidentally, N must be infinite to match
the number of points on the straight line.
A wave function can also be expanded in another
continuous representation, say the momentum rep-
resentation:
hx|ψi = ψ(x) =
Z
−∞
dp hx|pihp|ψi
=
1
2π
Z
−∞
dp
~
e
i(p/~)x
ψ(p). (81)
This is just the inverse Fourier transform relation.
Dirac δ-function: The equivalence
ψ(x) =
Z
−∞
δ(x x
)ψ(x
)dx
= hx|ψi =
Z
−∞
hx|x
ihx
|ψidx
(82)
shows that
hx|x
i = δ(x x
). (83)
It immediately follows that the Dirac δ-function
can be represented in infinitely many ways. For
example,
hx|x
i =
N
X
i=1
hx|iihi|x
i =
N
X
i=1
φ
i
(x)φ
i
(x
). (84)
Quantum measurement and wave-function
collapse: Electrons emitted from a source pass
through a single slit and fall on a distant observing
screen on the other side of the slit: As a coherent
matter wave, each electron forms a diffraction pat-
tern on the observing screen in the direction y of
the diffraction pattern. (Diffraction = spreading of
a wave into the geometrical shadow regions.)
A detector of acceptance y placed at position
y = d on the observing screen detects electrons
from signals of their passage through the detec-
tor. If ψ(y
) = hy
|ψi is the wave function of self-
diffracting electrons normalized in some way, the
part that enters the detector is
ψ(d) =
Z
d+∆y/2
dy/2
dy
|y
ihy
|ψi. (85)
The probability of an electron entering the detector
is
P (d) =
Z
d+∆y/2
dy/2
dy
|ψ(y
)|
2
. (86)
Note that the electron described by the wave func-
tion ψ(y) can be anywhere on the screen before the
detection. However, once it is known to have gone
into the detector, e.g., by the detector signal it gen-
erates, its wave function has collapsed into ψ(d).
This is how a quantum measurement can modify
the state itself.
5.4 Operators and their matrix elem ents:
Matrix mechanics
Operators as matrices: Operators can be visu-
alized as matrices because they satisfy matrix op-
erations. Hence quantum mechanics is also called
matrix mechanics. For example, an operator
ˆ
A
acting on a state always gives another state:
ˆ
A|ψi = c|φi (87)
10
in much the same way that a square matrix multi-
plying a column vector always gives another column
vector. It can happen occasionally that the result
is actually proportional to the original state |ψi
ˆ
A|ai = a|ai. (88)
These special states are called the eigenstates of
ˆ
A. The proportionality constants a are called their
eigenvalues. Note how the eigenvalue a is used to
label the eigenstate |ai itself. This operator equa-
tion is called an eigenvalue equation.
Noncommuting operators: Operators, like ma-
trices, do not generally commute. For example,
[ˆp, ˆx] =
~
i
. (89)
In matrix mechanics, commutators like this specify
the fundamental wave nature of the matter waves
described by these operators. It is the starting
point from which other results can be derived. In
contrast, the wave property of matter is described
in wave mechanics by postulating the quantization
rule
ˆp =
~
i
x
, (90)
and the wave function ψ(x) on which ˆp operates.
Operator matrix elements: Operators like ma-
trices have matrix elements. The matrix elements
of a matrix or operator
ˆ
M can be expressed conve-
niently in the Dirac notation as
M
ij
= hi|
ˆ
M|ji, (91)
a two-sided representation in terms of a suitable
basis |ii. A 2 × 2 matrix or operator can thus be
written in operator form as
ˆ
M = M
11
|1ih1| + M
12
|1ih2|
+ M
21
|2ih1| + M
22
|2ih2|. (92)
The ket-bra combinations on the RHS are them-
self operators or square matrices. The off-diagonal
operators are the transition operators
|ψihφ| =
ψ
1
ψ
2
(φ
1
φ
2
) =
ψ
1
φ
1
ψ
1
φ
2
ψ
2
φ
1
ψ
2
φ
2
, (93)
while the diagonal operators (with φ = ψ) are pro-
jection operators.
An operator in the representation of its own eigen-
states is a diagonal matrix. Thus
hx|ˆx|x
i = hx|x
ix
= x
δ(x x
)
hx|ˆp|x
i = hx|x
i
~
i
d
dx
= δ(x x
)
~
i
d
dx
.(94)
Note how the operator ˆp becomes a diagonal differ-
ential operator in the x-representation, and how it
comes out to the right of the Dirac bracket to avoid
operating on it. The Hamiltonian operator too has
interesting diagonal representations:
hE|
ˆ
H|E
i =
hE|E
iE
= δ(E E
)E
, if continuous
hn|n
iE
n
= δ
nn
E
n
, if discrete.
(95)
Eigenvalue eq. as wave eq.: Matrix mechanics
can be shown to be equivalent to wave mechanics by
writing the operator eigenvalue equation
ˆ
H|Ψ
E
i =
E|Ψ
E
i as a Schr¨odinger wave equation:
Ehx|Ψ
E
i = hx|
ˆ
H|Ψ
E
i
=
Z
−∞
dx
hx|
ˆp
2
2m
+ V (ˆx)
x
iΨ
E
(x
)
=
Z
−∞
dx
δ(x x
)
"
1
2m
~
i
d
dx
2
+ V (x
)
#
× Ψ
E
(x
)
=
"
1
2m
~
i
d
dx
2
+ V (x)
#
Ψ
E
(x). (96)
5.5 Hermitian operators and physical observ-
ables:
Hermitian operator: is one for which
ˆ
H
=
ˆ
H
T
=
ˆ
H (itself). (97)
It has the important properties that
(a) Eigenvalues are real.
(b) Eigenstates can be orthonormalized.
Physical observables can have only real values.
Hence they are described by Hermitian operators
in quantum mechanics.
Dirac notation: For any operator
ˆ
A:
hφ|
ˆ
i =
Z
−∞
φ
(x)[
ˆ
](x)dx
=
Z
−∞
{[
ˆ
A
φ](x)}
ψ(x)dx
= h
ˆ
A
φ|ψi. (98)
The Hermiticity of an operator
ˆ
H is confirmed by
the equality hφ|
ˆ
H|ψi = h
ˆ
Hφ|ψi of the two distinct
integrals involved. Note that if φ(x) is a column
wave function (called a spinor), the transposition
part of the adjoint operator will come into play to
change it into a row spinor.
11
5.6 One-dimensional harmonic oscillator:
The 1DHO problem can be solved elegantly in ma-
trix mechanics. First write its Hamiltonian
ˆ
H =
ˆp
2
2m
+
1
2
2
ˆx
2
(99)
in the compact form
ˆ
h =
ˆ
H
~ω
=
1
2
ˆq
2
+ ˆy
2
,
where ˆy =
ˆx
x
0
,
ˆq =
x
0
~
ˆp =
1
i
d
dy
. (100)
Here the oscillator length x
0
=
p
~/mω is defined
by the equation
~ω =
2
x
2
0
. (101)
The dimensionless position and momentum opera-
tors do not commute:
[ˆq, ˆy] = 1/i. (102)
The dimensionless Hamiltonian
ˆ
h can be factorized:
ˆ
h = ˆa
ˆa +
1
2
,
where ˆa =
1
2
(ˆy + iˆq)
ˆa
=
1
2
(ˆy iˆq) (103)
are called the step-down and step-up operators, re-
spectively. They do not commute
ˆa, ˆa
= 1. (104)
Number operator and its eigenstates: The
number operator
ˆ
N = ˆa
ˆa is Hermitian. It has real
eigenvalues n that can be used to label its eigen-
states
ˆ
N|ni = n|ni. (105)
It then follows that |ni is also an energy eigenstate:
ˆ
H|ni = (n +
1
2
)~ω|ni. (106)
The eigenvalue n of the number operator is a non-
negative number or integer 0, 1, 2, ... This im-
portant quantization property follows from non-
negative nature of its Hamiltonian (meaning that
h
ˆ
Hi 0) and from the nonzero commutators
[
ˆ
N, ˆa
] = ˆa
, [
ˆ
N, ˆa] = ˆa. (107)
The commutators show that ˆa
and ˆa change the
state |ni as follows:
ˆa
|ni =
n + 1|n + 1i, ˆa|ni =
n|n 1i. (108)
Oscillator quanta: The quantum number n (qu.
no. = any number used to label a quantum state)
gives the number of oscillator energy quanta, each
of energy ~ω, that can be emitted from that state.
Thus ˆa
increases the number of energy quanta in
the state by 1, and is called a creation operator,
while the destruction operator ˆa decreases n by
1. Physical processes that causes the absorption
or emission of energy quanta can be expressed in
terms of these operators.
The ground state which by definition cannot emit
any energy has the qu. no. n = 0. It has a
zero-point energy ~ω/2 that comes from the wave
spreading of its wave function from the classical
point x = 0 of stable equilibrium.
Wave functions: The impossibility of energy
emission from the ground state is stated by the op-
erator equation
ˆa|0i = 0. (109)
Its y-representaion gives the first-order differential
equation for the ground-state (GS) wave function:
y +
d
dy
φ
0
(y) = 0. (110)
The resulting normalized GS wave function is
φ
0
(y) = α
0
e
y
2
/2
, where α
0
=
1
π
1/4
. (111)
Excited states can be created by adding energy
quanta to the GS:
|ni =
a
)
n
n!
|0i. (112)
Consequently, its wave function can be constructed
by successive differentiations of the GS wave func-
tion:
φ
n
(y) = hy|
a
)
n
n!
|0i
=
α
0
2
n
n!
y
d
dy
n
e
y
2
/2
. (113)
Operators in the number representation: In
the number representation |ni, the only nonzero
matrix elements of ˆa and ˆa
lie along the diagonal
one line away from the main diagonal of the matrix:
ˆa =
0
1 0 .
0 0
2 .
0 0 0 .
. . . .
, ˆa
=
0 0 0 .
1 0 0 .
0
2 0 .
. . . .
. (114)
12
Hence the dimensionless position and momentum
operators are the square matrices
ˆy =
1
2
a + ˆa
) =
1
2
0
1 0 .
1 0
2 .
0
2 0 .
. . . .
,
ˆq =
1
i
2
a ˆa
) =
1
i
2
0
1 0 .
1 0
2 .
0
2 0 .
. . . .
.
(115)
5.7 Uncertainty principle:
The quintessential wave property described by the
Heisenberg uncertainty principle takes on a sur-
prisingly elegant form in matrix mechanics: If the
square uncertainty of the expectation value of a
physical observable
ˆ
A in the quantum state |ψi is
defined as
(∆A)
2
= h(
ˆ
A
¯
A)
2
i
ψ
, (116)
then the product of uncertainties of two physical
observables
ˆ
A and
ˆ
B in the same quantum state
satisfies the inequality
(∆A)(∆B)
1
2
h[
ˆ
A,
ˆ
B] i
ψ
. (117)
Simultaneous eigenstates: |a, bi exist for com-
muting operators
ˆ
A and
ˆ
B:
ˆ
A|a, bi = a|a, bi,
ˆ
B|a, bi = b|a, bi. (118)
Then A = 0 = ∆B.
5.8 Classical correspondence:
Bohr’s correspondence principle: Quantum
mechanics reproduces classical mechanics in the
limit of large quantum numbers. Example: The
probability density of quantum state |ni of the
1DHO agrees with the classical value:
lim
n→∞
ψ
QM
n
(x)
2
ρ
CM
(x). (119)
For a periodic system of period T , the classical
probability density can be defined in terms of the
probablity dP (x) of finding the classical point mass
within a width dx of the position x during a mo-
ment dt in time t:
dP (x) = ρ
CM
(x)dx =
2
T
dt, or
ρ
CM
(x) =
2
v(x)T
, (120)
where v(x) = dx/dt is the velocity at position x.
See Fig. 4-18, 19 of Bransden/Joachain for a com-
parison between quantum and classical ρs for the
1DHO.
Ehrenfest’s theorem: The expectation value of
a physical observable satisfies a classical equation
of motion. The expectation value of an operator
ˆ
A(t) in the time-dependent state
|Ψ(t)i = e
i
ˆ
Ht/~
|Ψ(0)i is (121)
h
ˆ
A(t)i
t
hΨ(t)|
ˆ
A(t)|Ψ(t)i
= hΨ(0)|e
i
ˆ
Ht/~
ˆ
A(t)e
i
ˆ
Ht/~
|Ψ(0)i. (122)
A time differentiation of the three time factors gen-
erates three terms:
d
dt
h
ˆ
A(t)i
t
= h
ˆ
A(t)
t
i
t
+
i
~
h[
ˆ
H,
ˆ
A(t)] i
t
. (123)
Examples:
d
dt
hˆxi
t
=
i
~
h[
ˆ
H, ˆx] i
t
=
1
m
hˆpi
t
;
d
dt
hˆpi
t
=
i
~
h[
ˆ
H, ˆp] i
t
= −h
d
ˆ
V
dˆx
i
t
. (124)
VI. Quantum mechanics in three spatial
dimensions
Different directions in space are independent of one
another in quantum mechanics in the sense that
[ˆx
i
, ˆx
j
] = [ˆp
i
, ˆp
j
] = [ˆx
i
, ˆp
j
] = 0, if i 6= j. (125)
QM problems in three-dimensional space can be solved
by using methods of both wave and matrix mechanics.
6.1 Separation of variables in rectangular coor-
dinates:
If a Hamiltonian
ˆ
H =
ˆ
H
1
+
ˆ
H
2
+
ˆ
H
3
is separable
into terms referring to different dimensions in rect-
angular coordinates, its energy eigenstates can be
factorized into the product states |n = n
1
n
2
n
3
i =
|n
1
i|n
2
i|n
3
i, one in each spatial dimension:
ˆ
H|n
1
n
2
n
3
i = (E
1
+ E
2
+ E
3
)|n
1
n
2
n
3
i. (126)
The method works for other operators too:
ˆp
2
|p
1
p
2
p
3
i = (p
2
1
+ p
2
2
+ p
2
3
)|p
1
p
2
p
3
i. (127)
Examples:
(a) 3-dim HO: E
n
= ~ω(n+3/2), n = n
1
+n
2
+n
3
.
(b) Free particle: p
2
= p
2
1
+ p
2
2
+ p
2
3
in the state
|p
1
p
2
p
3
i.
(c) Particle in a box of sides L
i
: The discrete mo-
mentum spectrum is defined by the quantum
numbers n
i
= k
i
L
i
= 1, 2, ....
13
6.2 Separation in spherical coordinates:
Kinetic energy: In the spherical coordinates r =
(r, θ, φ),
ˆ
p
2
separates into a radial part and an an-
gular part:
ˆ
p
2
= ˆp
2
r
+
ˆ
L
2
r
2
,
where
ˆ
L =
ˆ
r ×
ˆ
p =
e
x
e
y
e
z
ˆx ˆy ˆz
ˆp
x
ˆp
y
ˆp
z
=
ˆ
L
x
e
x
+
ˆ
L
y
e
y
+
ˆ
L
z
e
z
(128)
is the orbital angular momentum operator. Note
that the position and momentum operators in a
typical term of
ˆ
L such as e
x
ˆyˆp
z
commute because
they refer to different directions in space. However,
ˆ
L is quite a different type of operator from
ˆ
r or
ˆ
p,
because its components
ˆ
L
i
do not commute with
one another:
[
ˆ
L
i
,
ˆ
L
j
] = i~
ˆ
L
k
, (i, j, k in RH order). (129)
On the other hand,
[
ˆ
L
2
,
ˆ
L
i
] = 0, for any component i. (130)
The two commuting operators
ˆ
L
2
,
ˆ
L
z
can be used
to define the simultaneous eigenstate |ℓmi:
ˆ
L
2
|ℓmi = ~
2
( + 1)|ℓmi,
ˆ
L
z
|ℓmi = ~m|ℓmi, (131)
where and m are called the orbital (angular mo-
mentum) and magnetic quantum numbers, respec-
tively. This eigenstate describes completely an an-
gular state in spherical coordinates. The resulting
angular wave function
hθφ|ℓmi = Y
ℓm
(θ, φ) (132)
is called a spherical harmonic.
In the position represention, the
ˆ
L
z
equation can
be written as the differential equation
~
i
φ
Y
ℓm
(θ, φ) = mY
ℓm
(θ, φ). (133)
The solutions in the variable φ are
Y
ℓm
(θ, φ) = f(θ)e
imφ
. (134)
These solutions are periodic in φ with a period of
2π:
e
im(φ+2π)
= e
imφ
. (135)
Hence m must be an integer (0, ± 1, ±2, ...). It will
turn out that is a nonnegative integer bounded
from below by |m|.
The kinetic energy operator in spherical coordi-
nates is
ˆ
T =
ˆ
T
r
+
ˆ
T
L
=
1
2m
ˆp
2
r
+
ˆ
L
2
r
2
!
, (136)
Central po te ntial: Central potentials are spheri-
cally symmetric and depend only on the radial dis-
tance from the center of potential (or force):
V = V (r) 6= f (θ, φ). (137)
For central potentials, the Hamiltonian separates
in the radial and angular coordinates
ˆ
H = (
ˆ
T
r
+
ˆ
V ) +
ˆ
T
L
. (138)
As a result, the wave function of definite energy E
factorizes as follows:
ψ
Eℓm
(r) = hr|Eℓmi = R
E
(r)Y
ℓm
(θ, φ). (139)
The radial factor satisfies the Schr¨odinger equation
1
2m
ˆp
2
r
+
1
r
2
~
2
( + 1)
+ V (r)
R
E
(r)
= ER
E
(r), (140)
where
ˆp
2
r
= ~
2
2
r
= ~
2
1
r
2
d
dr
r
2
d
dr
= ~
2
1
r
d
2
dr
2
r. (141)
Note that the wave function R
E
(r) depends on
as well as the energy eigenvalue E. An en-
ergy eigenstate in three-dimensional space has only
three state labels.
Laplace equation: of electrostatics is also im-
portant in QM because its solutions describe the
quantum states of a static or resting free particle
(V = 0, p = 0 and E = 0):
2
ψ(r) = 0. (142)
The solutions regular at the origin r = 0 are the
solid spherical harmonics (or harmonic polynomi-
als)
Y
ℓm
(r) = r
Y
ℓm
(θ, φ). (143)
Solutions of the Laplace equations are sometimes
called harmonic functions.
The solutions in rectangular coordinates are sim-
pler. They are all rectangular polynomials with
terms of the form x
n
1
y
n
2
z
n
3
having the same de-
gree n = n
1
+ n
2
+ n
3
. All rectangular polynomials
14
TABLE II: Rectangular polyn omials and solid spherical har-
monics.
n Rect. Poly. Harmonic? Number Y
ℓm
?
0 1 Y 1 Y
00
1 z Y 3 Y
1m=0
x ±iy Y m = ±1
2 2z
2
x
2
y
2
Y 5 Y
2m=0
z(x ± iy) Y m = ±1
(x ± iy)
2
Y m = ±2
x
2
+ y
2
+ z
2
N 1
can be written in the spherical form with definite
ℓ, m:
r
n
Y
ℓm
(θ, φ) =
harmonic, if = n
non harmonic, if < n
, (144)
but only those where the power n of r is equal to
the degree of Y
ℓm
are solutions of the Laplace
equation.
Because
2
is simple in rectangular coordinates,
the result
2
x
n
1
=
d
2
dx
2
x
n
1
= n
1
(n
1
1)x
n
1
2
(145)
can be used to determine if any rectangular poly-
nomial is a harmonic function. Examples are given
in Table II. Functions with definite m include:
m = ±1 : x ± iy = r sin θe
±
m = 0 : z, x
2
+ y
2
. (146)
All solid spherical harmonics can be built from
these two classes of functions.
6.3 Three-dimensional harmonic oscillator:
Rectangular coordinates: The Hamiltonian of
the 3DHO is separable in rectangular coordinates:
ˆ
H =
ˆ
p
2
2m
+
1
2
2
ˆ
r
2
=
ˆ
H
1
+
ˆ
H
2
+
ˆ
H
3
. (147)
Consequently, its energy eigenstates are factorable:
|n
1
n
2
n
3
i =
"
a
1
)
n
1
n
1
!
#"
a
2
)
n
2
n
2
!
#"
a
3
)
n
3
n
3
!
#
|000i. (148)
All possible rectangular polynomials made up of
the components ˆa
i
of the vector operator
ˆ
a
can
create a 3DHO state. Their energy eigenvalues are
E
n
= (n +
3
2
)~ω, where
n = n
1
+ n
2
+ n
3
= 0, 1, 2... (149)
Spherical coordinates: The energy eigenstates
are
|nℓmi (
ˆ
a
)
2ν
Y
=n2ν,m
(
ˆ
a
)|000i, ν = 0, 1, ...(150)
All the states shown in Table II with x replaced by
the operator ˆa
1
, etc., are valid 3DHO states. The
degeneracy of these states is thus
D(n) =
1
2
(n + 1)(n + 2) =
n
X
=even or odd
(2 + 1). (151)
6.4 Plane wave in free space:
Three-dimensional plane waves have the product
wave functions
e
ik·r
= e
i(k
x
x+k
y
y+k
z
z)
=
X
=0
A
0
R
E
(r)Y
0
(θ, φ)
=
X
=0
i
(2 + 1)j
(kr)P
(cos θ) (152)
where the choice k = ke
3
makes explicit the axial
symmetry of the wave function about the propaga-
tion direction e
k
.
The spherical Bessel function j
(ρ = kr) that ap-
pears is the solution of the radial equation
1
ρ
2
d
ρ
2
d
+
( + 1)
ρ
2
1
j
(ρ) = 0. (153)
It has simple properties when ρ 0 and . The
Legendre polynomials P
(cos θ) are those used in
electrostatics.
6.5 Hydrogen atom:
Similarity with the Bohr model: In wave me-
chanics, the energy expectation values will turn out
to satisfy the Bohr-model relations
E = T + U =
U
2
= T,
E
n
=
E
1
n
2
,
E
1
=
e
2
G
2a
, T
1
=
~
2
2ma
2
, (154)
where n is the principal (or Bohr) quantum num-
ber, and a is the Bohr radius.
Differences from the Bohr model: The single
circular orbit of radius a
n
= n
2
a of the Bohr model
spreads out into states of quantum numbers n, ℓ, m,
with 0 < n. This result is consistent with the
spherical symmetry of the electrostatic attraction
between the electron and the proton. The degen-
eracy of states is then
d(n) =
n1
X
=0
(2 + 1) = n
2
. (155)
15
Wave functions: are
ψ
nℓm
(r) = R
nℓ
(r)Y
ℓm
(θ, φ). (156)
The radial functions R
nℓ
are functions of
ρ = κr, where
~
2
κ
2
2m
= E. (157)
Then
R
nℓ
(ρ) =
u
nℓ
(ρ)
ρ
= ρ
e
ρ
v(ρ) (158)
can be written in two other forms with each func-
tion satisfying a different differential equation:
2
r
R
nℓ
(r) =
( + 1)
r
2
2m
~
2
e
2
G
r
+ κ
2
R
nℓ
,
u
′′
nℓ
(ρ) =
( + 1)
ρ
2
ρ
0
ρ
+ 1
u
nℓ
,
ρv
′′
(ρ) + 2( + 1 ρ)v
+ (ρ
0
2 2)v = 0. (159)
where ρ
0
= 2/κa and v
(ρ) = dv(ρ)/dρ. All solu-
tions have
ρ
0
= 2n, giving E
n
=
E
1
n
2
, (160)
in agreement with the Bohr model.
Polynomials v
ν
(ρ): The factor v(ρ) in Eq.(158)
is a polynomial of ρ:
(a) v
0
(ρ) = 1 gives solutions with = n 1,
(b) v
1
(ρ) = ρ c(a constant) gives solutions with
= n 2,
(c) polynomials of degree ν, namely
v
ν
(ρ) =
ν
X
j=0
c
j
ρ
j
, (161)
give solutions with = n 1 ν.
Hydrogen states: The properties of the states of
the hydrogen atom are summarized in Table III.
The spectroscopy notation used is:
[] = s, p, d, f, ... for = 0, 1, 2, 3, ... (162)
The spectroscopic symbol for a quantum state is
2s+1
[]
j
, giving the spin degeneracy 2s + 1, the
orbital angular momentum [] and the total spin
quantum number j.
VII. Spin and statistics
The study of orbital angular momentum states leads
us to intrinsic spins and quantum statistics.
TABLE III: States of the hydrogen atom. The radial wave
functions are expressed in terms of the dimensionsless distance
˜r = r/a.
n E
n
m Spect.Symbol R
nℓ
(˜r)
1 E
1
0 0 1s e
˜r
2 E
1
/4 1 0, ±1 2p ˜re
˜r/2
0 0 2s (1 ˜r/2)e
˜r /2
n E
1
/n
2
max
= n 1 n[
max
] ˜r
max
e
˜r/n
n[]
7.1 Angular momentum:
The orbital angular momentum operator
ˆ
L =
ˆ
r ×
ˆ
p =
e
x
e
y
e
z
ˆx ˆy ˆz
ˆp
x
ˆp
y
ˆp
z
=
ˆ
L
x
e
x
+
ˆ
L
y
e
y
+
ˆ
L
z
e
z
(163)
is very different from
ˆ
x and
ˆ
p because its compo-
nents do not commute with one another:
[
ˆ
L
x
,
ˆ
L
y
] = i~
ˆ
L
z
, or
ˆ
L ×
ˆ
L = i~
ˆ
L. (164)
However, every component
ˆ
L
i
commutes with
ˆ
L
2
=
ˆ
L
2
x
+
ˆ
L
2
y
+
ˆ
L
2
z
. (165)
The angular momentum states |ℓmi are the simul-
taneous eigenstate of the two commuting operators
ˆ
L
2
,
ˆ
L
z
:
ˆ
L
2
|ℓmi = ~
2
( + 1)|ℓmi,
ˆ
L
z
|ℓmi = ~m|ℓmi, (166)
The nature of these states can be studied by using
the
Ladder o perators: for angular momentum states
ˆ
L
±
=
ˆ
L
x
± i
ˆ
L
y
. (167)
These operators have the structure of
x ± iy = r sin θe
±
(168)
that in wave functions have the magnetic quan-
tum numbers m = ±1. Hence
ˆ
L
±
can be called
the m = ±1 components of
ˆ
L in spherical coor-
dinates. Operators and states are both matrices,
though states are not square matrices. They can
be multiplied together. In spherical coordinates,
the m value of a product of functions or matrices
is the sum of the m values of the individual func-
tions or matrices. Hence we expect that
ˆ
L
±
|ℓmi
to be a state of magnetic quantum number m ± 1,
respectively.
16
Before demonstrating this result, let us first note
however that the ladder operators, like their con-
stituent rectangular components of
ˆ
L
x
,
ˆ
L
y
, com-
mute with
ˆ
L
2
itself. Hence they do not change the
orbital quantum number when acting on the state
|ℓmi. However, they too do not commute with
ˆ
L
z
:
[
ˆ
L
z
,
ˆ
L
±
] = ±~
ˆ
L
±
. (169)
These commutators can be used to show explicitly
that
ˆ
L
+
is the step-up operator for m, increasing
its value by 1, while the step-down operator
ˆ
L
decreases m by 1:
|φ
±
i =
ˆ
L
±
|ℓmi = A
ℓm
|ℓm ± 1i. (170)
A
ℓm
can be found by calculating the normalization
hφ
±
|φ
±
i using
ˆ
L
2
ˆ
L
2
z
=
ˆ
L
2
x
+
ˆ
L
2
y
=
ˆ
L
ˆ
L
±
± ~
ˆ
L
z
. (171)
The result is
ˆ
L
±
|ℓmi = ~
p
( + 1) m(m ± 1) |m ± 1i. (172)
Possible ℓm values: m
2
is bounded because
h
ˆ
L
2
z
i < h
ˆ
L
2
i, or m
2
< ( + 1). (173)
Hence m
min
m m
max
. From Eq.(172):
ˆ
L
+
|ℓm
max
i = 0 m
max
= ℓ,
ˆ
L
|ℓm
min
i = 0 m
min
= ℓ. (174)
So the total number of steps N on the ladder is
m
max
m
min
= 2 = N 1. (175)
Assuming that ladders with any number of steps
exist in nature, one finds
Odd N = 2 + 1 =
N 1
2
= integer,
Even N = 2s + 1 s =
N 1
2
= half integer.
(176)
The orbital quantum numbers s are integers be-
cause they are also the integer degrees of the har-
monic polynomials Y
ℓm
. The half-integer s’s are
the spin quantum numbers of an intrinsic spin an-
gular momentum
ˆ
S. The word “spin” is also used
for any angular momentum.
Angular-momentum representations:
ˆ
L
2
,
ˆ
L
z
and therefore also
ˆ
L
±
,
ˆ
L
x
,
ˆ
L
y
are N × N matrices
for those states with N steps in their ladder.
Example: For spin s = 1/2, the ladder has only
two steps, with spin states “up” and “down”.
Then the spin operators are 2 × 2 matrices:
ˆ
S
2
=
(3/4)~
2
diag(1, 1). The Pauli spin matrices are
the components of
ˆ
σ = 2
ˆ
S/~:
σ
z
=
1 0
0 1
, σ
x
=
0 1
1 0
, σ
y
=
0 i
i 0
. (177)
7.2 Addition of angular momenta:
Product states: Two angular momentum opera-
tors
ˆ
L and
ˆ
S are independent if [
ˆ
L
i
,
ˆ
S
j
] = 0. The
simultaneous eigenstates of the commuting opera-
tors
ˆ
L
2
,
ˆ
L
z
,
ˆ
S
2
,
ˆ
S
z
are the product states
|ℓm
sm
s
i = |ℓm
i|sm
s
i. (178)
They have the degeneracy
d(ℓ, s) = (2 + 1)(2s + 1). (179)
Coupled states: Two independent angular mo-
menta
ˆ
L and
ˆ
S can be added to the total angular
momentum:
ˆ
J =
ˆ
L +
ˆ
S. (180)
The four operators
ˆ
L
2
,
ˆ
S
2
,
ˆ
J
2
,
ˆ
J
z
can be shown to
commute among themselves. Hence their simulta-
neous eigenstates |ℓsjmi exist:
ˆ
L
2
|ℓsjmi = ~
2
( + 1)|ℓsjmi,
ˆ
S
2
|ℓsjmi = ~
2
s(s + 1)|ℓsjmi,
ˆ
J
2
|ℓsjmi = ~
2
j(j + 1)|ℓsjmi,
ˆ
J
z
|ℓsjmi = ~m|ℓsjmi. (181)
Clebsch-Gordan coefficients: The two distinct
ways (product or coupled) of specifying the states
of two angular momenta are related by a change of
representation:
|ℓsjmi =
X
m
m
s
|ℓm
sm
s
ihℓm
sm
s
|
!
|ℓsjmi,
|ℓm
sm
s
i =
X
j
|ℓsjmihℓsjm|
|ℓm
sm
s
i, (182)
where each sum involves only one summation in-
dex because m
+ m
s
= m. The transformation
(Clebsch-Gordan) coefficients are chosen real, and
therefore
hℓsjm|ℓm
sm
s
i = hℓm
sm
s
|ℓsjmi. (183)
Examples: Two spin 1/2 states can be coupled to
a total spin of S = 1 or 0:
|
1
2
1
2
SM = 11i = |↑↑i, |
1
2
1
2
1 1i = |↓↓i,
|
1
2
1
2
10i =
1
2
|↑↓+↓↑i;
|
1
2
1
2
00i =
1
2
|↑↓−↓↑i, (184)
where () denotes the spin up (down) magnetic
state of the spin 1/2 system.
17
Vector model: The length J of the vector sum
J = L + S satisfies a triangle inequality:
|L S| J L + S, but
j
min
= | s| j j
max
= + s, (185)
for quantum angular momenta involves the quan-
tum numbers.
Construction of total angular momentum
states: The jth ladder of states has 2j + 1 steps.
The topmost state m = + s of the longest ladder
j
max
= + s is simply
|ℓs + s + si = |ℓssi. (186)
Then step down to find the remaining states on the
ladder:
ˆ
J
|ℓsjmi = ~
p
j(j + 1) m(m 1) |ℓsjm 1i
= (
ˆ
L
+
ˆ
S
)
X
m
m
s
|ℓm
sm
s
i
× hℓm
sm
s
|ℓsjmi, (187)
where the RHS is a sum over two terms involving
ˆ
L
|ℓm
i and
ˆ
S
|sm
s
i, respectively.
The topmost state of the next ladder with j =
+ s 1 is the remaining or unconstructed state
with m = + s 1. It is constructed by orthonor-
malization from |s + s + s 1i, using the prod-
uct representation. By convention, its overall phase
is taken to be the Wigner (aka Condon-Shortley)
phases. The other states on the ladder are found
by stepping down. This construction method works
for all the remaining j ladders.
7.3 Identical particles in quantum mechanics:
Spin-statistics theorem: Identical particles of
integer (half integer) spins satisfy the Bose-Einstein
(Fermi-Dirac) statistics. They are called bosons
(fermions).
Two identical particles are indistinguishable in
their probability densities
ρ(2, 1) = ρ(1, 2), (188)
but their wave functions may differ by a negative
sign
Ψ(2, 1) = ±Ψ(1, 2) for
bosons
fermions
. (189)
Boson (Fermion) wave functions are symmetric (an-
tisymmetric) in the particle labels.
Examples: If α, β are single-particle (s.p.) states:
Ψ
αβ
(1, 2) [ψ
α
(1)ψ
β
(2) ± ψ
α
(2)ψ
β
(1)] . (190)
TABLE IV: electronic structure of the elements. The quan-
tum numbers of some atomic ground states are also given as
the spectroscopic symbol
2S+1
[L]
J
.
Z Element Spatial Configuration Ground State
1 H 1s
2
S
1/2
2 He (1s)
2 1
S
0
3 Li (He)2s
2
S
1/2
4 Be (He)(2s)
2 1
S
0
5 B (He)(2s)
2
2p
2
P
1/2
6 C (He)( 2s)
2
(2p)
2
7 N (He)(2s)
2
(2p)
3
8 O (He)(2s)
2
(2p)
4
9 F (He)(2s)
2
(2p)
5 2
P
3/2
10 Ne (He)(2s)
2
(2p)
6 1
S
0
If β = α is the same s.p. state:
Bosons : Ψ
αα
(1, 2) = ψ
α
(1)ψ
α
(2) 6= 0, but
Fermions : Ψ
αα
(1, 2) = 0. (191)
These important results are called:
Bose-Eins tein condensation: The ground state
of a system of identical bosons is one where every
boson is in the lowest energy single-particle state.
Pauli exclusion principle: No two identical
fermions can populate the same single-particle
state.
Many identical particles: The wave function of
N identical bosons is symmetric under any per-
mutation of the N particle labels. For identical
fermions, the wave function changes sign under an
odd permutation of the N particle labels, but re-
mains unchanged under an even permutation.
7.4 The Periodic Table:
Electronic structure of the elements: is sum-
marized in Table IV:
Helium atom: Its two electrons with spins point-
ing and can populate the same 1s spatial state.
They complete the principal qu. no. n = 1 shell.
The shell n can accomodate 2n
2
electrons.
To get the lowest energy, both electrons in the He
ground state (GS) occupy the lowest s.p. spatial
state nℓm = 100. The total wave function
Ψ
0
(1, 2) = ψ
100
(r
1
)ψ
100
(r
2
)χ
SM=00
(1, 2)
= Ψ
0
(2, 1) (192)
must be antisymmetric in the two fermion labels 1
and 2. Hence the spin “function” χ (not a “state”)
for the total intrinsic spin must be the antisymmet-
ric function with S = 0.
18
The atomic Hamiltonian
ˆ
H(1, 2) =
Ze
2
G
r
1
Ze
2
G
r
2
+
e
2
G
r
12
(193)
gives the GS energy (with Z = 2)
E
0
(He) Z
3
E
R
+ E
12
109 eV + 30 eV. (194)
where E
R
= 13.6 eV is the Rydberg energy.
Closed subshells: The m
s
value of the spin
state |sm
s
i changes when the quantization axiz e
z
changes direction. However, if all m
s
states are
filled by 2s + 1 particles, they will remain filled
when the arbitrarily chosen direction e
z
changes.
Hence the total M
S
is 0, and this unique state of
2s + 1 particles has S = 0. Similarly, the complete
filling of the 2 + 1 states of different m
gives a
unique spherical symmetric spatial system of 2 + 1
particles that has L = M
L
= 0. This is why the
ground states of He, Be and Ne are all
1
S
0
states.
Spin-orbit interaction: The atomic Hamilto-
nian contains a relativistic term A
LS
ˆ
L ·
ˆ
S, A
LS
>
0. It gives a negative (attractive) energy of
(A
LS
/2)(+1) for the s.p. state with the total an-
gular momentum j = 1/2. For the j = + 1/2
state on the other hand, the energy is repulsive,
namely (A
LS
/2). This explains why the B ground
state in the spatial 2p configuration has the quan-
tum numbers
2
P
1/2
of the lowest-energy single elec-
tron state outside the closed n = 1 shell and the
closed 2s subshell.
The F ground state is one electron short of the
closed n = 2 shell. The lowest energy is obtained by
putting the missing electron or hole in the higher
j = 3/2 state to give a
2
P
3/2
ground state, i.e.,
leaving a higher state unpopulated by populating
all the lower-energy states in the 2p subshell.
7.5 Free Fermi gas:
Box norm al ization: For a particle extending over
all space, the plane-wave wave functions e
ik·r
are
not square-integrable. This feature causes awk-
wardness including the appearance of Dirac δ-
functions. The problem can be circumvented by
subdividing space into identical cubes of sides L
satisfying a periodic boundary condition (BC). The
wave functions in each identical cube can then be
normalized.
One-dimensional space: The periodic BC
φ
k
x
(x + L) = e
ik
x
(x+L)
= φ
k
x
(x) = e
ik
x
x
(195)
gives the discrete spectrum
k
x
L = 2πn
x
, with n
x
= 0, ±1, ± 2, ... (196)
and the normalized wave functions
φ
k
x
(x) =
1
L
e
ik
x
x
. (197)
The corresponding states |n
x
i can be labeled by the
qu. no. n
x
. They form a discrete basis for wave
functions in the representative 1D box.
Three-dimensional s pace: The normalized 3D
wave functions in the representative cube are
hr|ni =
1
L
3/2
e
ik·r
, where k
i
=
2π
L
n
i
, (198)
and n = n
1
n
2
n
3
.
Fe rm i gas: For identical fermions, each spatial
state can be populated at most singly (the spin
degeneracy being treated separately). Then n is
called a (vector) occupation number. A free parti-
cle of mass m has energy E
k
=
p
(~k)
2
c
2
+ m
2
c
4
.
So the lowest energy state of a free Fermi gas of
N fermions in the cube is one where the states of
lowest k or occupation number n = |n| are popu-
lated. With spatial isotropy (spherical symmetry),
the occupation pattern is that of a sphere in n or
k space of radius k = k
F
:
N =
Z
dN =
Z
d
3
n =
L
2π
3
Z
d
3
k
=
Z
k
F
0
dN
dk
dk =
L
2π
3
4π
3
k
3
F
. (199)
Here
dN
dk
=
L
2π
3
4πk
2
(200)
is called the density of states in k-space. The radius
of the Fermi sphere in k-space
k
F
= (6π
2
ρ
#
)
1/3
, where ρ
#
=
N
L
3
, (201)
is called the Fermi momentum. It does not depend
on the size of the cube. For electrons with their spin
degeneracy d(s) = 2, the total electron number is
N
e
= 2N.
The energy per particle in a free Fermi gas is
E
N
= hE
k
i
gas
=
R
dN
dk
E
k
dk
R
dN
dk
dk
=
(3/5)ǫ
F
, if nonrelativistic;
(3/4)~ck
F
, if very relativistic.
(202)
The nonrelativistic energy
ǫ
F
=
~
2
k
2
F
2m
(203)
at the surface k = k
F
of the Fermi sphere is called
the (NR) Fermi energy.
Chandrasekhar limit: A white dwarf star con-
tains Z protons, Z electrons and A Z neutrons.
The main contributions to its energy are the repul-
sive kinetic energy of the highly relativistic elec-
tron gas and the attractive gravitational potential
19
energy of its mass M . Both energies are inversely
proportional to R, the radius of the star, but they
have different M dependences:
E
Tot
= b
M
4/3
R
c
M
2
R
> 0, unbound,
= 0, critical,
< 0, collapse.
(204)
The white dwarf will collapse gravitationally if its
mass exceeds the critical mass
M
c
=
b
c
3/2
. (205)
VIII. Approximation methods for time-
independent problems
Few problems in quantum mechanics are exactly solv-
able. We first consider some approximation methods for
Hamiltonians that are time-independent.
8.1 Time-independe nt perturbation theory:
2 × 2 matrices: can be diagonalized exactly. The
Hermitian matrix
H =
0 b
b a
, (206)
where both a and b are real, has the eigenvalues
and orthonormal eigenvectors
E
1,2
=
1
2
(a E),
E =
p
a
2
+ 4b
2
a + 2
b
2
a
;
ψ
1
=
cos θ
sin θ
, ψ
2
=
sin θ
cos θ
,
tan θ =
E
1
b
b
a
. (207)
Many problems of physical interest involving only
two dominant states can be solved approximately
by this method.
Perturbation theory: Let the Hamiltonian
ˆ
H =
ˆ
H
0
+
ˆ
H
contains a main part
ˆ
H
0
whose energy
eigenvalues E
(0)
n
, eigenfunctions φ
n
or eigenstates
|ni are already known. If a complicated but weak
perturbation
ˆ
H
is now added to the system, the
eigenvalue equation
(
ˆ
H
0
+ λ
ˆ
H
)ψ
n
= E
n
ψ
n
(208)
can be solved by a systematic expansion in powers
of λ that can be set back to its numerical value of
1 at the end of the expansion.
This λ expansion is done in both E
n
and ψ
n
:
(
ˆ
H
0
+ λ
ˆ
H
)
ψ
(0)
n
+ λψ
(1)
n
+ ...
=
E
(0)
n
+ λE
(1)
n
+ ...
ψ
(0)
n
+ λψ
(1)
n
+ ...
. (209)
The equation is solved by equating terms of the
same power of λ on the two sides of the equation:
λ
0
:
ˆ
H
0
ψ
(0)
n
= E
(0)
n
ψ
(0)
n
ψ
(0)
n
= φ
n
,
λ
1
: E
(1)
n
= H
nn
,
ψ
(1)
n
=
X
k6=n
H
kn
E
(0)
k
E
(0)
n
φ
k
,
where H
kn
= hk|
ˆ
H
|ni. (210)
8.2 Calculation of E
(1)
n
:
Many perturbations are made up of sums of fac-
torable terms of the type
ˆ
H
= Q(r)
ˆ
A (211)
involving a radial factor Q(r) and an angle-spin fac-
tor
ˆ
A for a system with intrinsic spin S. Then
E
(1)
n
= Q
nℓ
h
ˆ
Ai
α
,
where Q
nℓ
=
Z
0
|R
nℓ
(r)|
2
Q(r)r
2
dr,
α = ℓm
sm
s
or ℓsjm. (212)
Examples:
h
ˆ
L ·
ˆ
Si =
1
2
h
ˆ
J
2
ˆ
L
2
ˆ
S
2
i
ℓsjm
,
h
ˆ
L
z
+ 2
ˆ
S
z
i
ℓm
sm
s
= (m
+ 2m
s
)~,
h
ˆ
S
z
i
ℓsjm
=
~m
~
2
j(j + 1)
h
ˆ
J ·
ˆ
Si
ℓsjm
. (213)
8.3 Variational method:
Variational principle: For any wave function ψ:
E
ψ
= h
ˆ
Hi
ψ
hψ|
ˆ
H|ψi
hψ|ψi
E
0
(GS). (214)
The approximate wave function ψ gives a higher
energy when it contains excited-state components
of higher energies.
Variational wave function: is any wave func-
tion carrying a number of parameters b that can be
varied in order to minimize the energy
E(b) =
hψ(b)|
ˆ
H|ψ(b)i
hψ(b)|ψ(b)i
. (215)
The energy minimum E(b
0
) where dE(b)/db = 0 is
the best variational estimate because it is closest to
the true ground-state (GS) energy E
0
.
He atom: The ee repulsion between the two elec-
trons in the He atom can be eliminated approxi-
mately by reducing the actual nuclear charge Z = 2
to the effective charge z 27/16.
20
Hartree mean-field screening: In an atom with
Z electrons, the mutual ee interactions can be re-
placed by a mean field experienced by each electron
characterized by an effective nuclear charge of
z(r) =
Z, for r 0
1, for r
(216)
8.4 WKB approximation:
Since any complex number can be written in the
amplitude-phase form Ae
, one can look for a so-
lution of the 1D Schr¨odinger wave equation in this
amplitude-phase form. The WKB approximation
is a semi-classical result used between the two clas-
sical turning points x
1
and x
2
of the potential V (x)
where the local wave number
k(x) =
r
2m
~
2
[E V (x)] (217)
is real. It is obtained by ignoring the A
′′
(x) =
d
2
A(x)/dx
2
term. Then
ψ(x) = A(x)e
(x)
ψ
W KB
(x) =
C
p
k(x)
e
±i
R
x
x
1
k(x
)dx
. (218)
WKB quantization: At a turning point where
the potential has a finite slope, the wave func-
tion can penetrate into the classically forbidden re-
gion where the local wave number becomes purely
imaginary k . Under this analytic continua-
tion, the amplitude factor becomes complex, A
1/
k e
iπ/4
/
κ, thereby contributing an addi-
tional phase of π/4. The resulting WKB-modified
Bohr-Sommerfeld quantization condition is
Z
x
2
x
1
k(x)dx =
n
m
4
π, n = 1, 2, ..., (219)
where m is the number of turning points where the
potential is not an wall. The resulting WKB
energies are often quite good. They even agree
with the exact energies in the case of the one-
dimensional harmonic oscillator potentials.
IX. Approximation methods for time-
dependent problems
Time-dependent problems are more complicated partly
because one has to keep track of the time evolution of a
state, but mostly because the Hamiltonian itself might
be time-dependent.
9.1 The time-dependent two-state problem:
is exactly solvable for the time-independent Hamil-
tonian (206). Let a state be prepared initially as
|ψi |Ψ(t = 0)i = d
1
|ψ
1
i + d
2
|ψ
2
i, (220)
where |ψ
i
i are the time-independent eigenstates of
ˆ
H. Then it evolves at time t into
|Ψ(t)i = e
i
ˆ
Ht/~
|ψi = d
1
|Ψ
1
(t)i + d
2
|Ψ
2
(t)i
= d
1
e
iE
1
t/~
|ψ
1
i + d
2
e
iE
2
t/~
|ψ
2
i. (221)
Unperturbed basis: It is often useful to express
|Ψ(t)i in terms of the time-dependent unperturbed
basis states
|Φ
1
(t)i = e
iE
(0)
1
t/~
1
0
,
|Φ
2
(t)i = e
iE
(0)
2
t/~
0
1
(222)
of energies E
(0)
1
= 0 and E
(0)
2
= a:
|Ψ(t)i = c
1
(t)|Φ
1
(t)i + c
2
(t)|Φ
2
(t)i. (223)
The probabilities |c
i
(t)|
2
can be found by using the
energy eigenstates
ψ
1
=
cos θ
sin θ
, ψ
2
=
sin θ
cos θ
(224)
in Eq.(221):
|c
1
(t)|
2
=
d
1
cos θ d
2
e
iωt
sin θ
2
,
|c
2
(t)|
2
=
d
1
sin θ + d
2
e
iωt
cos θ
2
, (225)
where ω = (E
2
E
1
)/~. These probabilities are
in general functions of t. Indeed, their time depen-
dences betray the fact that the φ
i
are not energy
eigenstates, for otherwise they will be stationary
states with time-indepedent probabilites.
The expansion (223) is particularly interesting
when the basis states |φ
i
i are themselves physically
observable, as happens in
Neutrino oscillations: Neutrinos come in three
distinct varieties or “flavors”: ν
e
, ν
µ
, ν
τ
, associated
with the electron e, muon µ and tau τ, respec-
tively. Although they are physically distinct from
one another, they are not mass eigenstates (energy
eigenstates at zero momentum). The mass eigen-
states are linear combinations of these neutrino ba-
sis states of different flavors. For this reason, the
flavor probabilities are not constants of motion, but
instead they oscillate in time.
Solar neutrino oscillations: The neutrinos ini-
tially produced in the solar interior where masses
are converted to energy are ν
e
= φ
1
, and not
ν
µ
= φ
2
, i.e., c
2
(t = 0) = d
1
sin θ + d
2
cos θ = 0,
or d
1
= cos θ, d
2
= sin θ. Allowing only oscil-
lation into ν
µ
for simplicity, the probabilities for
finding ν
e
and ν
µ
at a later time t are
|c
1
(t)|
2
= 1 sin
2
2θ sin
2
(∆ωt/2),
|c
2
(t)|
2
= sin
2
2θ sin
2
(∆ωt/2). (226)
21
The neutrinos may have very small masses, so that
E
i
p
i
c +
m
2
i
c
3
2p
i
. (227)
The the probability for finding ν
e
of energy E at a
distance L = ct from the Sun is
|c
1
(t)|
2
= 1 sin
2
2θ sin
2
[A(∆m
2
c
4
)L/E];
m
2
= |m
2
1
m
2
2
|,
A =
1
4~c
= 1.27 GeV(eV)
2
km
1
. (228)
Experimental data for solar neutrinos of energies
broadly distributed around a mean value of E =
3 MeV give
(∆m
2
)c
4
= 7.3 × 10
5
(eV)
2
, tan
2
θ = 0.41. (229)
9.2 Time-dependent perturbation theory:
Expansion in exact eigenstates: Suppose
ˆ
H 6=
ˆ
H(t), and its energy eigenvalues E
j
and eigenstates
|ψ
j
i are known. Then any arbitrary quantum state
at time t can be expanded in terms of these energy
eigenstates:
|Ψ(t)i =
N
X
j=1
d
j
|Ψ
j
(t)i =
X
j
d
j
e
iE
j
t/~
|ψ
j
i. (230)
Expansion in basis states: This same state can
also be expanded in terms of a known set of basis
states |Φ
n
(t)i = e
iE
(0)
n
|φ
n
i that are eigenstates of
a known Hamiltonian
ˆ
H
0
of energies E
(0)
n
:
|Ψ(t)i =
X
n
c
n
(t)|Φ
n
(t)i =
X
n
c
n
(t)e
iE
(0)
n
t/~
|φ
n
i.(231)
It happens occasionally that both |ψi and |φi states
are physically observable. Then it is of interest to
express one observational probabilities in terms of
the other expansion coefficients. For example,
|c
n
(t)|
2
=
N
X
j=1
U
nj
d
j
e
iE
j
t/~
2
, (232)
is a generalization of what is done for neutrino os-
cillations to higher dimensions. Here U, with
U
nj
= hφ
n
|ψ
j
i, (233)
is an N ×N unitary matrix if both sets of states are
normalized. U is then made up of the eigenstates
|ψ
j
i stored columnwise.
TD perturbation: The expansion (231) can also
be used if
ˆ
H =
ˆ
H
0
+
ˆ
H
contains a time-dependent
(TD) perturbation
ˆ
H
. Unfortunately, the TD ex-
pansion coefficients c
n
(t) must be obtained more
laboriously by actually solving the TDSchEq
i~
t
|Ψ(t)i = (
ˆ
H
0
+ λ
ˆ
H
)|Ψ(t)i. (234)
The resulting differential equations (DEs) of mo-
tion for c
k
(t),
i~ ˙c
b
(t) = λ
X
k
H
bk
e
bk
t
c
k
(t), (235)
are a set of coupled first-order DEs that can be
solved easily in a PC. Here ω
bk
= (E
(0)
b
E
(0)
k
)/~.
Systematic TD perturbation theory: is con-
cerned with the Taylor expansion in λ (eventually
set to 1):
c
b
(t) = c
(0)
b
(t) + λc
(1)
b
+ ... (236)
Eq.(235) can then be separated term by term into
separate equations for c
(n)
b
:
λ
0
terms : i~ ˙c
(0)
b
(t) = 0 c
(0)
b
(t) = const,
λ
1
terms : i~ ˙c
(1)
b
(t) =
X
k
H
bk
e
bk
t
c
(0)
k
, ... (237)
These first-order DEs in time can be integrated di-
rectly, making it easy to calculate first-order wave
functions in TD perturbation theory.
Sinusoidal perturbation: Let the sinusoidal per-
turbation
ˆ
H
=
ˆ
V cos ωt, 0 t t
0
only, (238)
be turned on for a finite period of time. After the
perturbation has ceased, a state initially in |φ
a
i
will find itself in the state |φ
b
i, b 6= a, with the
probability
c
(1)
b
(t t
0
)
2
V
ba
2
2
j
2
0
(z)
t
2
0
~
2
, (239)
where z =
1
2
(ω
ba
± ω)t
0
,
j
0
(z) = sin(z)/z,
ω
ba
= [E
(0)
b
E
(0)
a
]/~. (240)
Thus most of the excitations are close to the two
states |φ
b
i with energies E
(0)
b
= E
(0)
a
± ~ω.
Golden rule: The transition rate during the in-
teraction lasting a time t
0
from a state a to a group
of states b is thus
W
ba
1
t
0
Z
−∞
|c
(1)
b
|
2
ρ
b
d(~ω
ba
)
2π
~
V
ba
2
2
ρ
b
, (241)
22
where ρ
b
= dN
b
/dE
b
is the density of final states.
Example: The Bohr transition of an atomic elec-
tron from a higher atomic state b to a lower state
a with the emission of a photon of energy ~ω =
E
(0)
b
E
(0)
a
is caused by the electric dipole interac-
tion
ˆ
H
=
ˆ
V cos(ωt). (242)
Here
ˆ
V = ǫ·
ˆ
DE
0
is the electric-dipole interaction
between the electric dipole operator
ˆ
D = q
ˆ
r of the
atomic electron and the electric field E = ǫE
0
of
the emitted photon. The latter can be obtained
from the photon energy density in a cube of side L:
~ω
L
3
=
ǫ
0
E
2
0
2
, (243)
where ǫ
0
is the permittivity of free space. When
ˆ
H
is used in the Golden Rule, the resulting transi-
tion rate is the Einstein coefficient for spontaneous
photon emission by the atom:
A = W
ba
=
4
3
ω
3
c
3
~
1
4πǫ
0
|D
ba
|
2
, (244)
9.3 Sudden, adiabatic and impulsive perturba-
tions:
In sudden perturbation, the Hamiltonian
changes suddenly by a finite amount to the final
value
ˆ
H
a
. If the change is so sudden that the state
|ψi at t = 0 has not changed, the state at times
t > 0 is just
|Ψ(t)i = e
i
ˆ
H
a
t/~
|ψi
=
X
n
e
iE
(a)
n
t/~
|φ
n
ihφ
n
|ψi
=
X
n
|Φ
n
(t)ihφ
n
|ψi. (245)
Adiabatic perturbation: When a time-
dependent Hamiltonian
ˆ
H(t) changes very slowly
in time, we expect the instantaneous eigenvalues
E
(t)
n
(t) and eigenstates |ψ
(t)
n
i defined by the eigen-
value equation
ˆ
H(t)|ψ
(t)
n
i = E
(t)
n
(t)|ψ
(t)
n
i (246)
to give a good approximation. That is
E
n
(t) E
(t)
n
(t), (247)
a result known as the adiabatic theorem.
If one watches the system for a sufficiently long
time, however, sooner or later additional features
will appear. They can be described conveniently
by using the time-dependent basis |ψ
(t)
k
i:
|Ψ(t)i =
X
k
c
k
(t)e
(i/~)
R
t
0
E
(t
)
k
(t
)dt
|ψ
(t)
k
i. (248)
The new features include the appearance of a ge-
ometrical phase ϕ
geo
n
and of transitions to other
states k 6= n.
Impulsive perturbation: An impulsive pertur-
bation contains a δ-function in time:
ˆ
H
(x, t) =
ˆ
A(x)δ(t). (249)
The TDSchEQ can be solved exactly to show that
the δ-function makes the state discontinuous at t =
0:
|Ψ(0
+
)i =
1
ˆ
A
2i~
!
1
1 +
ˆ
A
2i~
!
|Ψ(0
)i, (250)
X. Applications
In this chapter, we see how quantum problems of great
physical interest have been solved exactly or approxi-
mately.
10.1 Solving special m atrix eigenvalue problems:
Special matrices of arbitrarily large dimensions of
great physical interest can sometimes be solved an-
alytically by symmetry considerations.
To begin, it is easy to verify by inspection the eigen-
values and eigenvectors of the Pauli matrix σ
x
:
0 1
1 0
1
1
=
1
1
,
0 1
1 0
1
1
=
1
1
. (251)
Indeed, three simple rules often go a long way in
finding the eigenvectors and eigenvalues of certain
Hermitian matrices:
(a) The vector v
1
= (1, 1, ..., 1), where all compo-
nents are 1, is an eigenvector of a matrix where
the matrix elements of each row is a permu-
tation of those of the first row. Its eigenvalue
E
1
is just the sum of the matrix elements of
any row.
(b) Eigenvectors of a Hermitian matrix can be
made othogonal to one another.
(c) The sum of eigenvalues is equal to the trace of
the matrix.
Example 1: The vector v
1
= (1, 1, 1) is an eigen-
vector of the 3 × 3 matrix ((0,1,1),(1,0,1),(1,1,0)),
with eigenvalue E
1
= 2. The three vectors v
2
=
(2, 1, 1), v
3
= (1, 2, 1), v
4
= (1, 1, 2) are
all orthogonal to v
1
, but not linearly independent
of one another. Each has the eigenvalue -1. The
orthogonal eigenvectors can be taken to be v
1
, v
2
,
and v
5
= (v
3
v
4
)/3 = (0, 1, 1).
23
Example 2: The vectors v
1
, v
2
and v
5
are also
the orthogonal eigenvectors of the Hermitian ma-
trix ((1,1,1),(1,1,1),(1,1,1)), with the eigenvalues 3,
0, and 0, respectively. This is a special case of the
general result that two matrices M
1
and M
2
have
common eigenvectors if
M
2
= M
1
+ aI, (252)
where I is the identity matrix. Since the parame-
ter a can be complex, we see that non-Hermitian
matrices can also have orthogonal eigenvectors.
Example 3: The arbitrarily large N × N matrix
whose matrix elements are all 1’s has one “collec-
tive” eigenvector v
1
= (1, 1, ..., 1) with eigenvalue
N. The remaining eigenvalues are all 0. The or-
thogonal eigenvectors can be chosen to be v
2
=
(N 1, 1, 1, ..., 1), v
3
= (0, N 2, 1, ..., 1),
..., v
N
= (0, 0, ..., 0, 1, 1).
BCS theory of superconductivity: can be un-
derstood conceptually by using a large dimensional
N × N matrix Hamiltonian with the same small
negative matrix element ǫ everywhere. All eigen-
states have energy 0, except the collective eigen-
state v
1
= (1, 1, ..., 1) that has energy = ǫN .
The energy gap that separates the ground state
from the excited states makes it hard for the sys-
tem to be excited. If the ground state contains
current-carrying electrons, then at sufficiently low
temperatures (below a certain critical temperature
T
c
), these electrons cannot lose energy by inelastic
collisions with the crystal lattice in the conductor.
When this happens, both dissipation and resistiv-
ity vanish, and the medium becomes superconduct-
ing. (BCS = Bardeen, Cooper and Schrieffer who
discovered this fundamental theory of superconduc-
tivity.)
10.2 Stark effect in an external electric field:
A charge q moved a distance z against a constant
external electric field E = Ee
z
has the (dipole) in-
teraction energy
E
= q
Z
E.dr = qEz. (253)
The resulting perturbing Hamiltonian is
ˆ
H
= −E
ˆ
D
z
, where
ˆ
D = q
ˆ
r (254)
is called the electric dipole operator of the charge.
This operator is odd in parity, changing sign when
z z. Hence all first-order energies vanish:
E
(1)
n
= hn|
ˆ
H
|ni = 0. (255)
Since z = r cos θ is independent of the azimuthal
angle φ,
ˆ
D
z
has magnetic quantum number m = 0.
Hence
ˆ
H
has nonzero matrix elements only be-
tween two opposite-parity states i, j with the same
magnetic quantum number: m
i
= m
j
.
Example: In the n = 2 shell of the hydrogen atom,
there are four states: nℓm = 200; 210, 21 ± 1.
ˆ
H
connects only the two opposite-parity m = 0 states.
For these two states alone, it takes the form of a
2 × 2 matrix b((01), (1, 0)) whose eigenstates are
E
1,2
= b : ψ
1,2
=
1
2
(φ
200
± φ
210
) . (256)
Their energies E
1,2
thus move away from the undis-
turbed energy of the two states with m = ±1 by an
amount proportional to b, which is proportional to
E. [The state |ℓmi has parity (1)
.]
10.3 Aharonov-Bohm effect:
Gauge transformation: In classical electromag-
netism (EM) the electric and magnetic fields are
uniquely defined, but the EM potential (the scalar
potential ϕ and vector potential A) are not unique.
They can change by a gauge transformation
A A
= A + Λ,
ϕ ϕ
= ϕ
Λ
t
, (257)
where Λ is any scalar field, without changing the
EM fields
B = × A,
E = ϕ
A
t
. (258)
Relativistic electro dynamics: The basic dy-
namical variable in classical mechanics is the 4-
momentum (p, iE/c). To add EM, one needs an
EM 4-vector. There is a unique qualifying candi-
date: the EM 4-potential (A, iϕ/c). Hence electro-
dynamics for a particle of mass m and charge q can
be built up from the sum (p qA, i(E qϕ)/c).
This combination transforms like a 4-vector un-
der Lorentz transformations, thus guaranteeing the
correct result in different inertial frames. As a re-
sult, the EM 4-potential plays a more fundamental
role in quantum mechanics than the EM fields.
The energy-momentum relation for relativistic elec-
trodynamics is then
(E )
2
= (p qA)
2
c
2
+ m
2
c
4
. (259)
One can find from this the nonrelativistic kinetic
energy T
NR
:
T E mc
2
T
NR
= qϕ +
1
2m
(p qA)
2
. (260)
Quantum electrodynamics: We are now in a
position to quantize the mechanical terms in NR
24
electrodynamics, leaving the EM potential unquan-
tized for simplicity:
ˆ
H =
1
2m
(
ˆ
p qA)
2
+
ˆ
V + qϕ. (261)
Aharonov-Bohm phase: The EM interaction re-
sides primarily in the phase (or gauge) of quantum
wave functions. For the special case where ϕ = 0,
this important result can be demonstrated readily
by showing that the wave function ψ in the pres-
ence of the vector potential A differs from the wave
function ψ
without A by an A-dependent phase g:
ψ(r) = e
ig
ψ
(r). (262)
If true, the phase factor e
ig
serves the purpose of
removing the A dependence from the Schr¨odinger
equation when moved to the left of a “gauged” fac-
tor:
~
i
qA
e
ig
ψ
(r) = e
ig
~
i
ψ
(r). (263)
Since
~
i
e
ig
ψ
(r) = e
ig
~(g) +
~
i
ψ
(r), (264)
g has to satisfy the differential equation
~g = qA, or ~dg = qA · dr. (265)
Hence
g =
q
~
Z
r
r
0
A(r
) · dr
, (266)
where the line integral is in general path-dependent.
The integral is then said to be non-integrable. This
non-integrability can be made explicit by integrat-
ing around a closed circuit c once and simplifying
the result with the help of Stokes’s theorem:
g =
q
~
I
c
A(r
) · dr
=
q
~
Z
S
( × A) · dS =
q
~
Φ, (267)
where the total magnetic flux Φ across the surface
S enclosed by the circuit c depends on the magnetic
field ×A on and inside c. Such a dependence on
properties of the system outside the path c shows
why quantum mechanics in particular, and waves
more generally, describe nonlocal phenomena that
are spread out in space.
Magnetic flux quantization: Suppose the path
c is located in a region free of any EM field, then
the wave function must return to its original value
after one complete circuit. Hence its Aharonov-
Bohm phase can only be an integral multiple of 2π:
g = 2πn =
qΦ
~
, or Φ = nΦ
0
. (268)
As a result, the magnetic flux enclosed by c can
only exist in integral multiples of a quantized unit
flux Φ
0
= h/q. Experimental measurement deter-
mines that q = 2e in superconductors, thus show-
ing that the super current is carried by pairs of
electrons, now called Cooper pairs.
Gauge interactions: Other fundamental inter-
actions also work through the phase of quantum
wave functions. Hence they are called gauge inter-
actions. Indeed, the information contained in the
phase of a wave function is usually more important
than that contained in its amplitude. A picture of
Snow White can be made to appear in our eye as
a coherent wave. One can take the phase informa-
tion from Snow White’s wave function and use it
with the amplitude part of the wave function from
a picture of Happy, one of the seven dwarfs. A
composite wave like this has been constructed the-
oretically. Believe it or not, one sees Snow White
in the resulting hybrid picture, and not the dwarf.
10.4 Magnetic resonance:
Rabi measured the magnetic moments of atomic
states by finding the Zeeman splitting of the en-
ergies of different magnetic substates by resonance
matching. Atomic magnetic moments are tradition-
ally expressed as gµ
B
in units of the Bohr magneton
µ
B
= e~/2m, the theoretically value g being called
the Land`e g factor. The Hamiltonian involved is
ˆ
H =
ˆ
H
0
+
ˆ
H
, (269)
where
ˆ
H
0
=
ˆ
H
1
+
ˆ
H
LS
+
ˆ
H
2
contains the usual
NR Hamiltonian
ˆ
H
1
, the relativistic spin-orbit term
ˆ
H
LS
, and the Zeeman Hamiltonian
ˆ
H
2
=
ˆ
M · B
z
=
gµ
B
B
z
~
ˆ
J
z
= ω
0
ˆ
J
z
(270)
that gives the splitting between the magnetic sub-
states of the same j. Finally a small oscillatory
magnetic field is applied in the x-direction to give
a perturbing Hamiltonian
ˆ
H
= ω
x
ˆ
J
x
cos(ωt), where ~ω
x
= gµ
B
B
x
. (271)
This perturbation causes transitions between the
magnetic substates. Any quantum state
|Ψ(t)i =
X
m
c
m
(t)|Φ
m
(t)i
=
X
m
c
m
(t)e
m
t
|φ
m
i, (272)
25
where ω
m
= E
(0)
m
/~, can be expressed in term of the
unperturbed states |Φ
m
(t)i. The time-dependent
expansion coefficient c
m
(t) then satisfies the TD-
SchEq
i~ ˙c
m
(t) =
X
k
H
mk
e
mk
t
c
k
(t), (273)
where ω
mk
= ω
m
ω
k
.
Example: The problem can be solved approxi-
mately for c
m
(t) for the atomic n[]
j
= 2p
1/2
con-
figuration that has only two magnetic substates φ
±
of energies E
(0)
±
= ±~ω
0
/2. For example, if the sys-
tem is initially in the upper state (c
+
= 1, c
= 0),
then the probability of finding the system in the
upper and lower states at a later time t are respec-
tively:
P
++
= |c
+
|
2
= cos
2
(ω
R
t/2) + sin
2
χ sin
2
(ω
R
t/2),
P
+
= |c
|
2
= cos
2
χ sin
2
(ω
R
t/2), where (274)
ω
R
=
p
(∆ω)
2
+ (ω
x
/2)
2
, ω = ω
0
ω,
sin χ =
ω
ω
R
, cos χ =
ω
x
/2
ω
R
. (275)
Magnetic resonance: As the applied frequency
ω passes the Zeeman frequency ω
0
, the minimum
population of the φ
+
state falls sharply to 0 at ω
0
and then rises sharply back up to 1 again, thus
giving a clear signal for the measurement of ω
0
.
The third frequency ω
x
controls the period τ of
oscillation of the population between the two mag-
netic substates. At full resonance (ω = ω
0
), one
finds τ = 4π
x
.
XI. Scattering theory
Information about the dynamical properties of micro-
scopic systems like atoms, nuclei and particles can be
obtained by scattering beams of projectiles from targets
containing them, as first demonstrated by Rutherford,
Geiger and Marsden who elucidated atomic structure by
scattering α particles from a gold foil.
11.1 The scattering cross sections:
The differential cross section
d
=
1
L
dR
d
(276)
is the angular distribution dR/d of the reaction
rate R per unit luminosity L. L itself is the product
IN
T
, where I is the flu x or current density of the
incident beam (number of incoming particles per
second per cross sectional area of the beam, or
˙
N
i
/a
for a beam of uniform cross section a) and N
T
is
the number of target particles illuminated by this
beam. L depends on both beam and target. d
is the differential solid angle in a suitable inertial
frame.
Example: Two beams of the same cross sectional
area A collide head-on where their paths cross each
other. Each beam is made up of bunches of N
i
(i =
1, 2) particles per bunch. The target can be taken
to be one bunch of N
2
particles in beam 2. Then
the effective incident flux is I = fN
1
/A, where f is
the frequency of collision between bunches.
Rutherford cross section: For the scattering of
α-particles of charge Z
1
= 2 from an atomic nu-
cleus of charge Z
2
, Rutherford found from classical
mechanics that in the CM frame
d
=
d
4
2
1
sin
4
(θ/2)
,
where d =
Z
1
Z
2
e
2
G
T
CM
, (277)
is the distance of closest approach (or turning
point) for head-on collision at the NR kinetic en-
ergy T
CM
in the CM frame. Note that for backscat-
tering θ = π, the Rutherford differential cross sec-
tion d
2
/16 gives a direct measurement of the posi-
tion d of the classical turning point.
Total cross section: is proportional to the total
reaction rate:
σ =
Z
=
Z
d
dΩ =
Z
dR
L
=
R
L
. (278)
11.2 Quantum theory of scattering in a nutshell:
Partial-wave expansio n: of a plane wave into
spherical waves of good angular momentum
around the origin of coordinates has the Rayleigh
form:
ψ
k
(r) = e
ik·r
=
X
=0
i
(2 + 1)P
(cos θ)j
(kr),
r→∞
X
=0
i
(2 + 1)P
(cos θ)
i
2k
×
e
i(krℓπ/2)
r
e
i(krℓπ/2)
r
(279)
This is the time-independent wave function of a sta-
tionary scattering state. The time-dependent wave
function carries an additional time factor e
iωt
,
where ω = ~k
2
/2m. The additional time factor
shows that the wave function e
ikr
/r describes an
ingoing spherical wave collapsing towards the ori-
gin, while the wave function e
ikr
/r describes an
outgoing spherical wave expanding out from the ori-
gin.
Scattered wave: When a target particle is placed
at the origin of coordinates, the radial part j
of the
26
wave function (279) will be replaced by the func-
tion R
(r) from the SchEq with a potential in it. At
large distances r, the ingoing waves are still collaps-
ing towards the origin, and therefore do not know if
a potential is present there. In elastic scatterings,
an outgoing spherical wave has the same normal-
ization as before, and can be changed at most by a
phase, here taken to be 2δ
, that resides in a phase
factor S
= e
2
called an S-matrix element:
ψ(r)
r→∞
X
=0
i
(2 + 1)P
(cos θ)
i
2k
×
e
i(krℓπ/2)
r
S
e
i(krℓπ/2)
r
= ψ
k
(r) + ψ
sc
(r) where (280)
ψ
sc
(r)
r→∞
X
=0
i
(2 + 1)P
(cos θ)
×
S
1
2ik
e
i(krℓπ/2)
r
= f(θ)
e
ikr
r
. (281)
is the scattered wave function, the additional wave
function generated by the potential. Containing
only outgoing spherical waves, it can be written
asymptotically (r ) in terms of an angle-
dependent quantity called the scattering amplitude
f(θ) =
X
=0
(2 + 1)f
P
(cos θ),
where f
=
S
1
2ik
=
1
k
e
sin δ
(282)
if the scattering is elastic, meaning that the phase
shift δ
is real.
Flux: or probability current density is
j =
~
2im
(ψ
ψ ψψ
) j
k
+ j
sc
, (283)
where j
k
= ~k/m is the flux of the plane wave, and
j
sc
=
~k
m
e
r
|f(θ)|
2
r
2
(284)
is the flux of the scattered wave at large distances.
Differential cross se ction: A detector of area
dA = r
2
dΩ then measures the differential cross sec-
tion
d
=
1
L
k
dR
d
=
1
~k/m
(j
sc
· e
r
)
dA
d
= |f(θ)|
2
. (285)
Total cross section: is the result integrated over
the solid angle dΩ:
σ
tot
=
Z
d
dΩ =
Z
f
(θ)f (θ)d
=
X
σ
tot,ℓ
, (286)
where the two sums over , one from each f, has
been simplified to only one sum by using the or-
thogonality relation for Legendre polynomilas
Z
1
1
P
(cos θ)P
(cos θ)dΩ =
4π
2 + 1
δ
ℓℓ
, and (287)
σ
tot,ℓ
=
4π
k
2
(2 + 1) sin
2
δ
. (288)
Optical theorem: If δ
is real, one finds that
Imf(θ = 0) =
X
(2 + 1)Imf
=
X
(2 + 1)
sin
2
δ
k
=
k
4π
σ
tot
. (289)
This theorem holds even when the phase shift δ
=
α
+
becomes complex.
11.3 Breit-Wigner resonance:
The partial-wave total cross section σ
=
σ
ℓ, max
sin
2
θ
reaches a maximum value of σ
ℓ, max
=
(4π/k
2
)(2 + 1) whenever δ
(k) = π/2. This maxi-
mum value is called the unitarity limit.
If δ
(E) rises through δ
(E = E
R
) = π/2 sifficiently
rapidly, a Taylor expansion about E = E
R
needs
to be taken only to the terms
δ
π
2
+ (E E
R
)
dE
E
R
. (290)
Then tan δ
Γ/2
E E
R
, where
Γ
2
=
1
(
/dE)
E
R
> 0. (291)
Resonance: In the neighborhood of a resonance,
the partial-wave total cross section has the simple
E-dependence:
σ
= σ
ℓ, max
tan
2
δ
1 + tan
2
δ
= σ
ℓ, max
/2)
2
(E E
R
)
2
+ /2)
2
. (292)
This function has a sharp maximum at E = E
R
and
falls rapidly to half its maximal value at E E
R
=
±Γ/2. Hence Γ is called the resonance width (or
full width at half maximum, FWHM).
27
Decaying state: The partial-wave scattering am-
plitude near resonance has the simple form
f
=
1
k
e
sin δ
=
1
k
1
cot δ
i
=
1
k
Γ/2
(E E
R
) + iΓ/2
. (293)
At the complex energy E = E
R
iΓ/2, the time-
dependent probability density decays exponentially
|Ψ(r, t)|
2
= |ψ(r)e
i(E
R
iΓ/2)t/~
|
2
= |ψ|
2
e
Γt/~
. (294)
11.4 Yukawa’s theory of interactions:
Yukawa showed in 1935 that both the Coulomb
interaction mediated by the exchange of massless
photons between charges, and the strong or nuclear
interaction mediated by the exchange of massive
bosons called mesons, can be derived from quan-
tum mechanics.
Klein-Gordon equation: The Einstein energy-
momentum relation E
2
= p
2
c
2
+m
2
c
4
can be quan-
tized into the Klein-Gordon equation in free space:
ˆ
H
2
Φ(r, t) =
ˆ
p
2
c
2
+ m
2
c
4
Φ(r, t), (295)
where
ˆ
H = i~/∂t and
ˆ
p = (~/i). The result-
ing wave functions Φ oscillates in time as e
iEt/~
.
These free-space solutions are said to be on the
energy shell, because the E-p relation is satisfied
on a spherical shell in p-space.
Static solutions of the KG equation: The
KG equation does not have time-independent so-
lutions in free space, because these solutions have
E = 0 and therefore cannot satisfy Einstein’s
E-p relation. Yukawa showed that there are
static (hence energy-nonconserving) solutions near
a point charge q located at the origin that satisfy
the inhomogenious DE:
ˆ
p
2
c
2
+ m
2
c
4
φ(r) = 4π(~c)
2
qδ(r), (296)
For m = 0, the Yukawa equation simplifies to the
Poisson equation
2
φ(r) = 4πqδ(r). (297)
One can show with the help of Gauss’s theorem
(or Gauss’s law in electrostatics) that the resulting
static wave function of a massless photon around a
point charge that vanishes at r = is
φ(r) =
q
r
. (298)
This is just the Coulomb potential around the point
charge q (in Gaussian units, or q = q
SI
/
4πǫ
0
).
The photon in question cannot propagate freely. It
can only exist momentarily around its point source
(the point charge), and is then said to be a virtual
photon. In Feynman’s diagrammatic language, the
Coulomb interaction arises from the emission of a
virtual photon by a source (the q here) and its ab-
sorption by a test charge at distance r from it, or
vice versa.
The virtual particle in Yukawa’s theory of inter-
actions is a boson, otherwise it cannot be emitted
because of the conservation of fermion number. A
virtual boson that is massive satisfies the Yukawa
equation
(
2
µ
2
)φ(r) = 4πgδ(r), (299)
where the unit of charge is now denoted g, and µ =
mc/~ is the inverse reduced Compton wavelength.
The solution of this DE that vanishes at r = can
be shown to be the Yukawa potential
φ(r) =
g
r
e
µr
. (300)
Nuclear forces: were known in 1937 to have a
range of 1 1.4 fm. (fm = 10
15
m.) This
fact allowed Yukawa to predict that these forces
are caused by the exchange of bosons of mass
m = µ~/c 140 MeV/c
2
. Such strongly inter-
acting bosons, now called π mesons or pions, were
discovered in 1947.
11.5 Scattering at low energies:
Impact parameter: In classical mechanics, a par-
ticle impacting at a transverse distance b (the im-
pact parameter) from the center of a square-well
potential of range R will not “see” the potential if
b > R. If the incident momentum in the center-
of-mass frame is p, the interaction vanishes when
the angular momentum = r×p exceeds the value
max
= Max(kR/~), where Max(z) is the largest
integer in z.
The wave spreading in quantum mechanics makes
the connection between and b less sharply defined,
but it remains true that the scattering phase shift
δ
vanishes sharply when >
max
. Consequently,
f(θ)
max
X
=0
(2 + 1)f
P
(cos θ), (301)
Hence at sufficiently low energies, the S-wave ( =
0) term dominates:
σ
tot
4π
sin
2
δ
0
k
2
=
4π
k
2
+ k
2
cot
2
δ
0
. (302)
The fact that the leading Taylor term for = 0 is
independent of k comes from the properties of the
scattering wave function.
28
Effective-range expansion:
k cot δ
0
=
1
a(k)
=
1
a
0
+
1
2
r
0
k
2
+ ..., (303)
where a
0
is called the scattering length and r
0
is
called the effective range.
11.6 Phase shifts for finite-range potentials:
Radial wave function: The wave function in the
presence of a potential V is
ψ(r) =
X
=0
i
(2 + 1)P
(cos θ)
u
(r)
kr
, (304)
where the radial wave function u
satisfies the ra-
dial wave equation
d
2
u
dr
2
+
2µ
~
2
(E V )u
( + 1)
r
2
u
= 0. (305)
where µ is the reduced mass in the center-of-mass
system.
Attractive S- wave square-well potential: For
= 0 in an attractive square-well potential of depth
V
0
, the radial wave equation is
For r < R :
d
2
u
in
dr
2
+ κ
2
u
in
= 0,
u
in
= sin(κr); (306)
For r > R :
d
2
u
out
dr
2
+ k
2
u
out
= 0,
u
out
= sin(kr + δ
0
); (307)
k
2
=
2mE
~
2
, κ
2
0
=
2mV
0
~
2
, κ
2
= k
2
+ κ
2
0
. (308)
The phase shift δ
0
is determined by matching the
inside and outside logarithmic derivatives L
in
=
L
out
at r = R, where L = (du/dr)/u, to give
1
κ
tan κR =
1
k
tan(kR + δ
0
), or
δ
0
= tan
1
k
κ
tan κR
kR. (309)
Levinson’s theorem: If there are m bound states
inside the potential,
δ
(k = 0) = , (310)
except for an S-wave bound state at E = 0, which
contributes only π/2, half of the normal contribu-
tion.
11.7 Born approximation for two-particle scatter-
ings:
The Born (or first Born) approximation for the
scattering of two particles is a first-order TDPT
that can be obtained from the
Golden rule: A weak two-body interaction
ˆ
H
=
ˆ
V e
iωt
+
ˆ
V
e
iωt
(311)
causes the scattering from an initial two-particle
state i to a final two-particle state f within the solid
angle d
2
f
around the final relative momentum k
f
.
The transition rate is given, in first-order TDPT,
by the Golden Rule:
dw
fi
d
=
2π
~
|hf|
ˆ
V |ii|
2
ρ
f
(E)
4π
=
µ
f
(~k
f
)
(2π)
2
~
4
L
3
|hf|
ˆ
V |ii|
2
. (312)
Here µ
j
, k
j
are the reduced mass and relative mo-
mentum, respectively, in the state j (= i, f ) of the
two-body system in a cube of side L, and the den-
sity of final state factor is
ρ
f
(E)
4π
=
L
2π
3
k
2
dk
dE
. (313)
The scattering is elastic if k
f
= k
i
, and inelastic if
k
f
, k
i
.
The resulting differential scattering cross sec-
tion
fi
d
1
J
inc
dw
fi
d
(314)
is just the transition rate per unit incident flux (or
current density), which is
J
inc
=
v
i
L
3
=
~k
i
µ
i
L
3
(315)
in a cube of side L. Note that J
inc
has the expected
dimension of m
2
s
1
. The final result for the Born
approximation can be written compactly as
fi
d
f
=
µ
i
µ
f
(2π)
2
~
4
k
f
k
i
|
˜
V (q)|
2
, (316)
where q = k
f
k
i
, and
˜
V (q) L
3
hf|
ˆ
V |ii =
Z
e
iq·r
V (r)d
3
r (317)
is obtained by using the plane-wave wave functions
in the relative coordinate r normalized to one par-
ticle in a cube of side L:
hr|ii =
1
L
3/2
e
ik
i
·r
, hf|ri =
1
L
3/2
e
ik
f
·r
. (318)
29
Thus the cross section in the Born approximation
can be calculated directly from the Fourier trans-
form
˜
V (q) of the interaction potential V (r).
Examples of
˜
V (q) are those from the δ-shell and
the Yukawa potentials:
V (r) = A
0
δ(r R)
˜
V (q) = A
0
4πR
q
sin(qR);
V (r)
e
αr
r
˜
V (q)
4π
α
2
+ q
2
. (319)