Quantum Lecture Notes
Quantum Lecture Notes
ON QUANTUM MECHANICS
Department of Physics
September 2004
CT Contents
1 Fundamental Concepts 1
ii
CONTENTS – MANUSCRIPT
1.6.1 Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
1.6.4 S and S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
1.6.7 Observable . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
1.7.3 Diagonalization . . . . . . . . . . . . . . . . . . . . . . . . . . 41
iii
CONTENTS – MANUSCRIPT
1.8.4 Translation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
2 Quantum Dynamics 62
iv
CONTENTS – MANUSCRIPT
v
CONTENTS – MANUSCRIPT
vi
CONTENTS – MANUSCRIPT
vii
CONTENTS – MANUSCRIPT
viii
CONTENTS – MANUSCRIPT
ix
CONTENTS – MANUSCRIPT
x
CONTENTS – MANUSCRIPT
xi
CONTENTS – MANUSCRIPT
angular momentum theory, creation and annihilation operators (the second quan-
• Co-requisite: Nil
xii
CONTENTS – MANUSCRIPT
xiii
CONTENTS – MANUSCRIPT
xiv
CONTENTS – MANUSCRIPT
xv
CONTENTS – MANUSCRIPT
xvi
CN Chapter 1
CT Fundamental Concepts
At the present stage of human knowledge, quantum mechanics can be regarded as the
are derived from physical events that almost entirely beyond the range of direct human
perception. It is not surprising that the theory embodies physical concepts that are
The most traditional way to introduce the quantum mechanics is to follow the
historical development of theory and experiment — Planck’s radiation law, the Einstein-
Debye’s theory of specific heat, the Bohr atom, de Broglie’s matter wave and so forth
of light, the Compton effect, and Franck-Hertz effect. In this way we can enjoy the
experience of physicists of last century to establish the theory. In this course we do not
1
CHAPTER 1 – MANUSCRIPT
follow the historical approach. Instead, we start with an example that illustrates the
Schiff’s book has a good introduction to this theory. Here we just list several experi-
tric effect (Einstein, 1904); Compton effect (1923); Combination Principle (Rita-
Rydberg, 1908)
• Discrete values for physical quantities: Specific heat (Einstein 1907, Debye 1912);
2
CHAPTER 1 – MANUSCRIPT
• Planck’s theory for black body radiation: = ~: Electromagnetic quanta (1900)
(1926)
The phenomenon in which charge particles are released from a material when it absorbs
electrons from the surface of a metal plate when light falls on it. In the broad sense,
however, the phenomenon can take place when the radiant energy is in the region of
visible or ultraviolet light, X rays, or gamma rays; when the material is a solid, liquid, or
gas; and when the particles released are electrons or ions (charged atoms or molecules).
rich Rudolf Hertz, who observed that ultraviolet light changes the lowest voltage at
which sparking takes place between given metallic electrodes. At the close of the 19th
3
CHAPTER 1 – MANUSCRIPT
Figure 1.1:
elementary negative charge. In 1900 Philipp Lenard, a German physicist, studying the
electrical charges liberated from a metal surface when it was illuminated, concluded
that these charges were identical to the electrons observed in cathode rays. It was fur-
ther discovered that the current (given the name photoelectric because it was caused
by light rays), made up of electrons released from the metal, is proportional to the
intensity of the light causing it for any fixed wavelength of light that is used. In 1902 it
was proved that the maximum kinetic energy of an electron in the photoelectric effect
is independent of the intensity of the light ray and depends on its frequency.
The observations that (1) the number of electrons released in the photoelec-
tric effect is proportional to the intensity of the light and that (2) the frequency, or
4
CHAPTER 1 – MANUSCRIPT
wavelength, of light determines the maximum kinetic energy of the electrons indicated
a kind of interaction between light and matter that could not be explained in terms of
classical physics. The search for an explanation led in 1905 to Albert Einstein’s funda-
mental theory that light, long thought to be wavelike, can be regarded alternatively as
penetrate matter, where it would collide with an atom. Since all atoms have electrons,
an electron would be ejected from the atom by the energy of the photon, with great
velocity. The kinetic energy of the electron, as it moved through the atoms of the
matter, would be diminished at each encounter. Should it reach the surface of the
material, the kinetic energy of the electron would be further reduced as the electron
overcame and escaped the attraction of the surface atoms. This loss in kinetic energy is
called the work function, symbolized by omega (). According to Einstein, each light
constant () and the frequency of the light (indicated by the Greek ). Einstein’s theory
of the photoelectric effect postulates that the maximum kinetic energy of the electrons
ejected from a material is equal to the frequency of the incident light times Planck’s
constant, less the work function. The resulting photoelectric equation of Einstein can
5
CHAPTER 1 – MANUSCRIPT
electric field and measuring the potential or voltage difference (indicated as V) required
to reduce its velocity to zero. This energy is equal to the product of the potential
thus, = .
and found to be correct but not complete. In particular, it failed to account for the fact
that the emitted electron’s energy is influenced by the temperature of the solid. The
remedy to this defect was first formulated in 1931 by a British mathematician, Ralph
Howard Fowler, who, on the assumption that all electrons with energies greater than
the work function would escape, established a relationship between the photoelectric
current and the temperature: the current is proportional to the product of the square
of the temperature and a function of the incident photon’s energy. The equation is
= 2 (), in which is the photoelectric current, and are constants, and ()
is an exponential series, whose numerical values have been tabulated; the dimensionless
value equals the kinetic energy of the emitted electrons divided by the product of the
temperature and the Boltzmann constant of the kinetic theory: = ( − ) ; in
6
CHAPTER 1 – MANUSCRIPT
atoms, performed (1914) by the German-born physicists James Franck and Gustav
Hertz .
electron tube. As the energy of the electrons was slowly increased, a certain critical
electron energy was reached at which the electron stream made a change from almost
undisturbed passage through the gas to nearly complete stoppage. The gas atoms were
able to absorb the energy of the electrons only when it reached a certain critical value,
indicating that within the gas atoms themselves the atomic electrons make an abrupt
transition to a discrete higher energy level. As long as the bombarding electrons have
less than this discrete amount of energy, no transition is possible and no energy is
absorbed from the stream of electrons. When they have this precise energy, they lose it
all at once in collisions to atomic electrons, which store the energy by being promoted
have been elastically scattered by electrons; it is a principal way in which radiant energy
is absorbed in matter. The effect has proved to be one of the cornerstones of quantum
7
CHAPTER 1 – MANUSCRIPT
Figure 1.2:
8
CHAPTER 1 – MANUSCRIPT
mechanics, which accounts for both wave and particle properties of radiation as well as
of matter.
and momentum just as material particles do; they also have wave characteristics, such
lower frequencies and longer wavelengths. In the Compton effect, individual photons
collide with single electrons that are free or quite loosely bound in the atoms of matter.
Colliding photons transfer some of their energy and momentum to the electrons, which
in turn recoil. In the instant of the collision, new photons of less energy and momentum
are produced that scatter at angles the size of which depends on the amount of energy
Because of the relation between energy and wavelength, the scattered photons
have a longer wavelength that also depends on the size of the angle through which the X
rays were diverted. The increase in wavelength or Compton shift does not depend on the
wavelength of the incident photon. The Compton effect was discovered independently
∆ =
(1 − cos )
9
CHAPTER 1 – MANUSCRIPT
Figure 1.3:
10
CHAPTER 1 – MANUSCRIPT
0 =
= 242631058 × 10−12 m = 0024̊
We start with the Stern-Gerlach experiment to introduce some basic concepts of quan-
tum mechanics. The two-state problem can be regarded as the most quantum. A lot of
important discoveries are related to it. It is worthy studying very carefully. We shall
11
CHAPTER 1 – MANUSCRIPT
Oven: silver atoms (Ag) are heated in the oven. The oven has a small hole through
Ag: (Electron configuration: 1s2 2s2 2p6 3s2 3p6 3d10 4s2 4p6 4d10 5s1 ). The outer
shell has only ONE electron (5s1 ) . The atom has an angular momentum., which is due
Shaped Magnet: N and S are north and south poles of a magnet. The knife-
edge of S results in a much stronger magnetic field at the point P than Q . i.e. the
Role of the inhomogeneous field: to change the direction of the Ag beam. The
= ( · B) ≈
If the magnetic field is uniform , i.e.
= 0, the Ag beam will not change its direction.
In the field different magnetic moments experience different forces, and the atoms with
different magnetic moments will change different angles after the beams pass through
the shaped magnet. Suppose the length of the shaped magnet . It takes a time
12
CHAPTER 1 – MANUSCRIPT
Figure 1.5: Beam from the Stern-Gerlach apparatus: (a) is expected from classical
∆
= = 2
= 2 ∝
spin ∝ )
Consequence: the spin of electron has two discrete values along the magnetic
13
CHAPTER 1 – MANUSCRIPT
field
⎧
⎪
⎪
⎨ ~2
=
⎪
⎪
⎩ −~2
It should be noted that the constant cannot be determined accurately from this exper-
iment.
SGẑ stands for an apparatus with inhomogeneous magnetic field in z direction, and
SGx in x direction
Case(a): no surprising!
14
CHAPTER 1 – MANUSCRIPT
15
CHAPTER 1 – MANUSCRIPT
say that the selection of + beam by SGx̂ completely destroys any previous
information about .
It is very helpful if you compare the situation with the analogy of the polarized light
through a Polaroid filter. The polarized light wave is described by a complex function,
E = 0 x̂ cos( − )
0 h (−) (−)
i
E = 12 x̂e ± ye
2
The Stern-Gerlach experiment shows that the spin space is not simply a 3-D vector
space and lead to consider a complex vector space. There exists a good analogy with
polarization of light. (Refer to Sakurai’s book). In this section we formulate the basic
16
CHAPTER 1 – MANUSCRIPT
Figure 1.7:
17
CHAPTER 1 – MANUSCRIPT
algebra has been known to mathematician before the birth of quantum mechanics, but
the Dirac notation has many advantages. At the early time of quantum mechanics
this notation was used to unify Heisenberg’s matrix mechanics and Schrodinger’s wave
atom with definite spin orientation, is represented by a state vector in a complex vector
space, denoted by |i, a ket. The state ket is postulated to contain complete informa-
tion about physical state. The dimensionality of a complex vector space is specified
A(|i) = A |i
A(|i) = |i
Two kets can be added to form a new ket: |i + |i = |i
One of the postulates is that |i and |i with the number 6= 0 represent
18
CHAPTER 1 – MANUSCRIPT
sponding to every ket there exist a bra. The names come from the word “bracket” →
“bra-c-ket”.
1. There exists a one-to one correspondence between a ket space and a bra space.
|i ⇐⇒ h|
|i ⇐⇒ ∗ h|
h|i = (h|i)∗
h|i ≥ 0
19
CHAPTER 1 – MANUSCRIPT
where the equality sign holds only if |i is a null ket. From a physicist’s point
Orthogonality:
h|i = 0
Operator: an operator acts on a ket from the left side, x |i and the result-
ing product is another ket. An operator acts on a bra from the right side, h| xand
— Hermitian operator: x = x†
— x |i ←→ h| x†
Multiplication
• — Noncommutative: xy 6= yx
20
CHAPTER 1 – MANUSCRIPT
y |i ⇒ h| y†
¡ ¢
xy |i = x (y |i) ⇒ h| y† x† = h| y† x†
as we are dealing with “legal” multiplications among kets, bras, and operators.
¯ ¯ ®∗
3. h |x| i = h| · (x |i) = {(h| x† ) · |i}∗ ≡ ¯x† ¯
The operators for observables must be Hermitian as physical quantities must be real.
21
CHAPTER 1 – MANUSCRIPT
If A = A† =⇒ = ∗
orthogonal.
G Proof.
We have
It is conventional to normalize
⎧ all eigenkets |0 i of A so that {|0 i} form an orthogonal
⎪
⎪
⎨ 1 if 0 = 00
0
set: h”| i = 0 00 =
⎪
⎪
⎩ 0 otherwise
22
CHAPTER 1 – MANUSCRIPT
2. An arbitrary ket in the ket space can be expanded in terms of the eigenkets of A:
P ¯ ® ®
|i = ¯() =⇒ () | =
µ ¶
P¯ ® ® P ¯¯ ® () ¯
|i = ¯() () | = ()
¯ |i
P ¯¯ ® () ¯
() ¯=1
(1 means the identity operator. For an arbitary ket |i 1 |i = |i)
P¯ ® ¯ P¯ ® ¯
( ¯() () ¯)( ¯() () ¯)
P ¯¯ ® () ¯ P¯ ® ¯ P¯ ® ¯
= () ¯ =⇒ ( ¯() () ¯)( ¯() () ¯ − 1) = 0
This relation is very essential to develop the general formalism of quantum me-
chanics.
¯ ® ¯
4. Projection operator: = ¯() () ¯
¯ ® ¯ ¯ ®
|i = ¯() () ¯ i = ¯()
23
CHAPTER 1 – MANUSCRIPT
¯ ®
selects that portion of the ket |i parallel to ¯() : · = such that
¯ ® ¯ ¯ ® ¯
· = (¯() () ¯)(¯() () ¯)
¯ ® ® ¯
= ¯() () |() () ¯
¯ ® ¯
= ¯() () ¯ =
Having specified the base kets, we show how to represent an operator, say X, by a
square matrix.
P¯ ® ¯ P¯ ® ¯
X = ( ¯() () ¯)X( ¯() () ¯)
24
CHAPTER 1 – MANUSCRIPT
and ⎛ ⎞
⎜ 0 ⎟
⎜ ⎟
⎜ .. ⎟
⎜ . ⎟
⎜ ⎟
⎜ ⎟
¯ () ® ⎜ ⎟
¯ =⎜ 0 ⎟
⎜ ⎟
⎜ ⎟
⎜ ⎟
⎜ 1 ⎟
⎜ ⎟
⎜ ⎟
⎝ .. ⎠
.
¯ ¯ ®
Thus the outer product of () ¯ and ¯() gives a new operator in the form of square
matrix, ⎛ ⎞
⎜ 0 ··· 0 ··· ⎟
⎜ ⎟
⎜ .. . . .. .. ⎟
⎜ . . . . ⎟
¯ () ® () ¯ ⎜ ⎟
¯ ¯ =⎜ ⎟
⎜ ⎟
⎜ 0 ··· 1 ··· ⎟
⎜ ⎟
⎜ ⎟
⎝ .. .. .. . . ⎠
. . . .
¯ ®
The inner product () ¯ () = 1 Thus X can be expressed in a matrix form,
⎛ ⎞
(1) ¯ ¯ (1) ® (1) ¯ ¯ (2) ®
⎜ ¯ X ¯ ¯ X ¯ ··· ⎟
⎜ ⎟
⎜ ¯ ¯ ® ¯ ¯ ® ⎟
⎜
X = ⎜ ¯ X ¯ ¯ ¯ ··· ⎟
(2) (1) (2)
X (2)
⎟
⎜ ⎟
⎝ .. .. ... ⎠
. .
¯ ®
For an operator A which eigenstates are ¯()
⎛ ⎞
(1)
⎜ 0 ··· ⎟
⎜ ⎟
⎜ ⎟
A=⎜
⎜ 0 (2) ··· ⎟
⎟
⎜ ⎟
⎝ . .. ... ⎠
.. .
25
CHAPTER 1 – MANUSCRIPT
P
4. Matrix-matrix multiplication: C = AB =⇒ = ·
=1
Relation
B 1.6.1 Measurements
procedure, and is quite different from that in classical physics. In the classic case the
measurement reflects the state of an object, such as position, velocity, and energy.
The process of measurement does not change the state of the object. For instance we
measure your body height, which does not change your body height. In the quantum
case the process of measurement changes the state of the object. If we want to measure
the spin state of electron we have to let electron go through the Stern-Gerlach apparatus
(of course we have other methods.). The magnetic field changes the spin state, and
select one of the eigenstates along the direction of the magnetic field. This is not a
particularly easy subject for a beginner. Let us first read the words of the great master,
26
CHAPTER 1 – MANUSCRIPT
–P.A.M. Dirac
P¯ ® ¯ P ¯ ®
|i = ( ¯() () ¯ i = ¯()
2. When the measurement is performed, the system is “thrown into” one of the
¯ ®
eigenstates, say ¯(1) of the observable
¯ (1) ®
|i A measurement ¯
−−−−−−−−−−−→
3. A measurement usually changes the state. When the measurement causes |i to
¯ ®
change into ¯(1) it is said that A is measured to be (1) !
Before the measurement the system is in |i ; after a measurement the system
¯ ®
is in ¯(1)
We are ready to derive the spin 1/2 operator we encountered in the Stern-Gerlach
27
CHAPTER 1 – MANUSCRIPT
|+i and |−i form a complete (?) and orthonormal (?) set of basis. From the complete-
~
S (|+i h+| + |−i h−|) = (|+i h+| − |−i h−|)
2
~
S = (|+i h+| − |−i h−|)
2
In general any operator for an observable can be expressed in terms of its eigenvalues
and eigenstates,
A | i = | i
X
A = | i h |
S+ = ~ |+i h−|
S− = ~ |−i h+|
28
CHAPTER 1 – MANUSCRIPT
⎧
⎪
⎪
⎨ S− |+i = ~ |−i h+|+i = ~ |−i
⎪
⎪
⎩ S− |−i = ~ |−i h+|−i = 0
Let ⎛ ⎞ ⎛ ⎞
⎜ 1 ⎟ ⎜ 0 ⎟
|+i ≡ ⎜
⎝ ⎠
⎟ ; |−i ≡ ⎜ ⎟
⎝ ⎠
0 1
¯ ® ¯ ®¯2
the probability for ¯() : = ¯ () | ¯
P ¯¯ ® () ®
provided that |i is normalized (h|i = 1). (|i = () | ).
X ¯ ® ® X
h|i = h ¯() () | = 1 =
29
CHAPTER 1 – MANUSCRIPT
P () ® () ® ®
hAi ≡ h |A| i = | |A| () () |
P ¯ ®¯2 P
= () × ¯ |() ¯ = ()
() ®
|A| () =
¯ ®
where is the probability in the state ¯() .
the eigenket and rejects all others. Mathematically, we can say that such a selective
B 1.6.4 S and S
We are now in the position to determine the eigenkets of S and S and the operators
eigenstates of S Similar to S , assume the two eigenstates of S are | +i and | −i
~
S = (| +i h +| − | −i h −|)
2
The beam of S + goes through SGz and splits into two beams with the same intensity,
30
CHAPTER 1 – MANUSCRIPT
h + | −i = 0
1 1 0
h+|+i + h−|−i −1 +1 = 0 ⇒ 01 = + 1
2 2
~
S = (| +i h +| − | −i h −|)
2
~ −1
= ( |+i h−| + 1 |−i h+|)
2
Likewise,
~ −2
S = ( |+i h−| + 2 |−i h+|)
2
=⇒ 1 − 2 = ±2
1 = 0 and 2 = 2
31
CHAPTER 1 – MANUSCRIPT
Therefore,
1
| ±i = (|+i ± |−i);
212
1
| ±i = (|+i ± |−i)
212
and
~
S = (|+i h−| + |−i h+|)
2
~
S = (− |+i h−| + |−i h+|)
2
S+ = S + S
S− = S − S
Define
~
S ≡
2
The eigenvalues of these three matrixes are ±1 (Please check it after class!)
32
CHAPTER 1 – MANUSCRIPT
Eample:
[S S ] = S · S − S · S
~ ~
= (|+i h−| + |−i h+|) (− |+i h−| + |−i h+|)
2 2
~ ~
− (− |+i h−| + |−i h+|) (|+i h−| + |−i h+|)
2 2
~2 ~2
= ( |+i h+| − |−i h−|) − (− |+i h+| + |−i h−|)
4 4
= ~S ;
1
{S S } = ~2
2
33
CHAPTER 1 – MANUSCRIPT
[S− S+ ] = −2S
= +
B 1.6.7 Observable
[A B] = 0
34
CHAPTER 1 – MANUSCRIPT
[A B] 6= 0
A having the same eigenvalues then the eigenvalues of the two eigenkets are said to be
degenerate.
the ket |i is a simultaneous eigenket of A and B. We can denote it by |i = | i e.g.
[S2 S ] = 0
3 2 ~
S2 |±i = ~ |±i ; S |±i = ± |±i
4 2
3 2
S2 ( |+i + |−i) = ~ ( |+i + |−i)
4
~
S ( |+i + |−i) = ( |+i − |−i)
2
are nondegenerate. Then the matrix elements h” |B| 0 i are all diagonal. (A |i = |i
¯ ®
G Proof. Suppose A is diagonal on the basis ket {¯() }
¯ ® ¯ ®
A ¯( = () ¯()
35
CHAPTER 1 – MANUSCRIPT
() ®
|[A B]| ()
() ®
= |[AB − BA]| ()
®
= (() − () ) () |B| () = 0
() ®
=⇒ |B| () = 0 if 6=
If A and B are compatible, we can always diagonalize them on one set of basis
1. Nondegenerate
|i −→ | i
2. Degenerate
P ¯ ®
|i −→ () ¯0 ()
¯ ®
−→ ¯0 ()
Incompatible Observables
36
CHAPTER 1 – MANUSCRIPT
G Proof. Proof: Assume A and B are incompatible and {|0 0 i} is a complete set of
=⇒ [A B] = 0
B. For instance consider a state with the angular momentum = 0. Although the
EXP Example 1
37
CHAPTER 1 – MANUSCRIPT
~
S = (| +i h +| − | −i h −|)
2
~
S = (| +i h −| − | −i h +|)
2
~
S = (− | +i h −| + | −i h +|)
2
~
S = (| +i h +| − | −i h −|)
2
~
S = (| +i h −| − | −i h +|)
2
~
S = (− | +i h −| + | −i h +|)
2
1 1
| ±i = √ | +i ± √ | −i
2 2
⎛ ⎞ ⎛ ⎞⎛ ⎞
⎜ | +i ⎟ ⎜ √1 √1 ⎟ ⎜ | +i ⎟
⎜ ⎟ = ⎜ 2 2 ⎟⎜ ⎟
⎝ ⎠ ⎝ ⎠⎝ ⎠
| −i √1 − √12 | −i
2
Suppose we have two incompatible observables A and B. The ket space can be viewed
as being spanned either by the set {|0 i} or by the set {|0 i}. i.e. A representation or
38
CHAPTER 1 – MANUSCRIPT
Theorem 2 Given two complete and orthogonal sets of base kets, there exist a unitary
¯ ® ¯ ®
operator U such that ¯() = U ¯()
P ¯¯ ® () ¯
U= () ¯
1)
¯ ® P¯ ® ® ¯ ®
U ¯() = ¯() () |() = ¯()
| {z }
2)
P ¯¯ ® () ¯
= () ¯=1
In the matrix representation, the operator U can be expressed in the form of a matrix
¯ ®
in the base ket {¯() }
® ®
U = { () |U| () } = { () |() }
39
CHAPTER 1 – MANUSCRIPT
A:
P ¯¯ () ® () ®
|i = |
B:
P ¯¯ () ® () ®
|i = |
() ®
|X| ()
() ¯ † ¯ ®
= ¯ U X U ¯()
=⇒ X = U† X U
− − − − − − A similarity transformation
40
CHAPTER 1 – MANUSCRIPT
P () ¯¯ ¯¯ () ®
Tr(X) ≡ X
P () ®
= |X| ()
B 1.7.3 Diagonalization
When we know the matrix elements of B in the base {|0 i} how do we obtain the
¯ ® ¯ ®
B ¯() = ¯()
41
CHAPTER 1 – MANUSCRIPT
¯ ®
In the base ket {¯() } the operator is expressed as
¯ ® ¯ ¯ ® ¯
B ¯() () ¯ = ¯() () ¯
X¯ ® ¯ X ¯ ® ¯
B ¯() () ¯ = ¯() () ¯
X ¯ ® ¯
B = ¯() () ¯
P () ® ® ®
|B| () () |() = () |()
where
®
= () |B| ()
and
() ®
= () |()
det(B − 1) = 0
EXP Example 2
42
CHAPTER 1 – MANUSCRIPT
⎛ ⎞
~⎜ 0 1 ⎟
S = ⎜ ⎟
2⎝ ⎠
1 0
⎛ ⎞
~
⎜ − 2 ⎟
det(S − 1) = 0 ⇒ det ⎜
⎝
⎟=0
⎠
~
2
−
~2 ~
2 − = 0⇒=±
4 2
Eigenvectors?
̂ |i = |i
43
CHAPTER 1 – MANUSCRIPT
We generalize of a vector space with finite dimension to that with infinite dimension
discrete continuous
2. The discrete sum over the eigenvalues is replaced by an integral over a continuous
variable.
2. () = (−)
3. 0 () = − 0 (−)
4. () = 0
6. () = 1 (−)
44
CHAPTER 1 – MANUSCRIPT
R
9. ( − )( − ) = ( − )
sin
() = lim
→∞
1 2 2
() = lim −
→0 12
() = ·
R
+∞ R
+∞
∗ () () = 2 −( − )
−∞ −∞
R
+ sin( − )
= 2 lim −( − ) = 2 2 lim
→∞ − →+∞ ( − )
= 2 2( − ) = ( − )
√
⇒ 2 2 = 1 =⇒ = 1 2
continues spectra, we consider the position operator in one dimension. The eigenkets
x |0 i = 0 |0 i
45
CHAPTER 1 – MANUSCRIPT
are postulated to form a complete set. The state ket for an arbitrary physical states
Suppose we place a very tiny detector that clicks only when the particle is precisely
at 0 and nowhere else after the detector clicks, we can say the state in question is
represented by |0 i. In practice the best the detector can do is to locate the particle
∆ ∆
within a narrow range (0 − 2
0 + 2
) (∆ is very small). When the detector clicks,
Assume that h0 |i does not change appreciably within the narrow interval, the prob-
|h |i|2 ∆
Obviously,
R
+∞
|h|i|2 = 1 = h|i
−∞
In the wave mechanics, the physical state is represented by a wave function. Thus h|i
is the wave function in the position space for the state |i
R
Wave function in position space: h0 |i in |i = 0 |0 i h0 |i is the wave
function in the position space. We assume that the position eigenkets |0 i are complete.
46
CHAPTER 1 – MANUSCRIPT
It can be three dimensional, i.e. 0 in |0 i stand for three components of the position
B 1.8.4 Translation
Suppose we start with a state that is well located around 0 , and introduce an translation
T() |i = | + i
h| T()† = h + |
¯ ¯ ®
¯T()† T()¯ = h + | + i = h|i
47
CHAPTER 1 – MANUSCRIPT
1. = 0 :
T( = 0) = 1
3. T(−)T() = 1
4. T()† T() = 1
48
CHAPTER 1 – MANUSCRIPT
=⇒ [x K] =
Z Z
|i → T() |i = T() |i h|i = | + i h|i
Z
= |i h − |i + →
Z Z
= |i (h|i − h|i ) = |i (1 − ) h|i
Z Z
also =⇒ |i h| T() |i = |i h| (1 − K) |i
Z
= |i (h|i − K h|i ) =⇒ K = −
49
CHAPTER 1 – MANUSCRIPT
1
h|i = 12
+x~
(2)
2
=
~
de Broglie wave is also called matter wave, any aspect of the behavior or
with the mathematical equations that describe waves. By analogy with the
wave and particle behavior of light that had already been established exper-
imentally, the French physicist Louis de Broglie suggested (1924) that par-
years later the wave nature of electrons was detected experimentally. Ob-
smaller than that of electrons, so their wave properties have never been de-
tected; familiar objects show only particle behavior. De Broglie waves play
50
CHAPTER 1 – MANUSCRIPT
p = −~
[x p] = ~
[x p] () = (xp − px) () = −~ () + ~ (())
= −~ () + ~ () + ~ ()
Therefore
51
CHAPTER 1 – MANUSCRIPT
x =~
The commutation relation [x p] = ~ shows that the position and the momentum are
® ®
(∆x)2 (∆p)2 ≥ ~2 4
® ®
(∆A)2 (∆B)2 ≥ |h[A B]i|2 4
The uncertainty relation can be understood from the postulate of measurement. For
an eigenstate of A,
®
(∆A)2 = 0
® ®
(∆A)2 (∆B)2 = 0
first.
52
CHAPTER 1 – MANUSCRIPT
EXP Example 3
1
0 () = exp[0 ~]
(2~)12
¯ ¯2
The probability to find x is a constant in the whole space, ¯0 ()¯ =
1
2~
= 0
2®
hi → 0; → +∞
EXP Example 4
A particle at = 0 :
Z
(−0 )~
=
2~
() = 0 ~
53
CHAPTER 1 – MANUSCRIPT
= 0
2®
hi → 0; → +∞
EXP Example 5
1
() = 12
exp[−2 220 ]
14 0
space,
Z
1
() = ()−~
(2)12
Z
1 1
= 12
exp[−2 220 − ~]
(2)12 14 0
Z µ ¶2
1 1 1 0 1 ³ 0 ´2
= 12
exp[− + ] exp[− ]
(2)12 14 0 2 0 ~ 2 ~
√ 14 12 Z µ ¶2
2 0 1 1 0 1 ³ 0 ´2
= √ exp[− + ] exp[− ]
(2)12 12 20 2 0 ~ 2 ~
0
= 12
exp[−2 20 2~2 ]
14 0
∼ ~
54
CHAPTER 1 – MANUSCRIPT
G Proof.
® ® ®
(∆A)2 = (A − hAi)2 = A2 − 2A hAi + hAi2
2®
= A − hAi2 ≥ 0
To simplify the problem, we can take h|i = h|i = 1 and |h|i|2 ≤ 1 Let
|i = ∆A |i
|i = ∆B |i
® ®
=⇒ (∆A)2 (∆B)2 ≥ |h∆A∆Bi|2
where we have used the hermiticity of A and B since they are observables,
∆A = (∆A)†
55
CHAPTER 1 – MANUSCRIPT
2)
1
∆A∆B = (∆A∆B − ∆B∆A)
2
1
+ (∆A∆B + ∆B∆A)
2
1 1
= [∆A ∆B] + {∆A ∆B}
2 2
1 1
= [A B] + {∆A ∆B}
2 2
= (AB)† − (BA)†
= B† A† − A† B†
= −[A B]
56
CHAPTER 1 – MANUSCRIPT
⎧
⎪
⎪
⎨ h |H| i =
=⇒ =⇒ = ∗
⎪
⎪ ¯ ¯ ®
⎩ ¯H† ¯ = ∗
Q† = −Q
¯ ¯ ®
h |Q| i = ¯Q† ¯ = ∗
=⇒ = − ∗
Therefore
1 1
h∆A∆Bi = h[A B]i + (h{∆A ∆B}i)
2 2
= + Im h∆A∆Bi + Re h∆A∆Bi
1 1
|h∆A∆Bi|2 = |h[A B]i|2 + |h{∆A ∆B}i|2
4 4
1
≥ |h[A B]i|2
4
® ® 1
(∆A)2 (∆B)2 ≥ (h[A B]i)2
4
57
CHAPTER 1 – MANUSCRIPT
[x p] = ~
® ®
(∆x)2 (∆p)2 ≥ ~2 4
Since the position and momentum operators are incompatible, [x p] 6= 0 we cannot
find a complete set of simultaneous eigenkets for x and p such that the two observables
can be measured at the same time. For instance for an eigenstate of x, x |i = |i, it
EXP Example 6
which will give off a photon that has the frequency 0 if the atom is at
rest. Because of the Doppler effect, the motion of the atom toward
given approximately by
≈ 0 (1 + ) =⇒ = ( − 1)
0
58
CHAPTER 1 – MANUSCRIPT
~
of the atom uncertain by the amount ∆ ≈
· Since the later
the photon is emitted, the longer the atom has the higher velocity
and the farther it will have traveled. This position uncertainty arises
we do not know when the velocity changed and hence where the
∆ ≈ ∆ ≈ ( − 1)
0
≈ ∆ ≈
0 0
Therefore
∆ · ∆ ∼ ~
59
CHAPTER 1 – MANUSCRIPT
EXP Example 7
= 0 0 + ·
⎛ ⎞ ⎛ ⎞
⎜ 0 0 ⎟ ⎜ 0 ⎟
= ⎜
⎝
⎟+⎜
⎠ ⎝
⎟
⎠
0 0 0
⎛ ⎞ ⎛ ⎞
⎜ 0 − ⎟ ⎜ 0 ⎟
+⎜
⎝
⎟+⎜
⎠ ⎝
⎟
⎠
0 0 −
⎛ ⎞
⎜ 0 + − ⎟
= ⎜
⎝
⎟
⎠
+ 0 −
Eigenvalues of X:
det( − ) = 0
60
CHAPTER 1 – MANUSCRIPT
⎛ ⎞
⎜ 0 + − − ⎟
det ⎜
⎝
⎟
⎠
+ 0 − −
¡ ¢12
=⇒ ± = 0 ± 2 + 2 + 2
61
CN Chapter 2
CT Quantum Dynamics
This chapter is devoted exclusively to the dynamic development of state kets and/or
observables. In other words, we are concerned here with the quantum mechanical
We now discuss how a physical state evolves with time. Suppose we have a physical
system whose state ket at 0 is represented by |i. Let us denote the ket corresponding
| 0 : i
62
CHAPTER 2 – MANUSCRIPT
As in the case of translation, we assume that the two kets are connected by the time-
evolution operator
Several properties of U:
1. The identity
U(0 0 ) = 1
2. The unitarity
U† ( 0 )U( 0 ) = 1
3. The composition:
evolution operator
As
63
CHAPTER 2 – MANUSCRIPT
U(0 + 0 ) = U(0 0 ) + U( 0 )|=0 +
= 1 − Ω̂ +
Just like the operator K in the translation operator T(), Ω̂ is a Hermitian operator.
⇓ ˜˜˜ ⇓
1
®
(2)12
(−(−0 )) ˜˜˜˜˜˜˜ (2)112 |−Ω(−0 ) |
→ 0 + =⇒ Ω →
= ~ (Planck-Einstein Relation)
In the classical mechanics, the energy is related to the Hamiltonian operator H , which
Hamiltonian operator H:
Ω̂ = H~
Thus
64
CHAPTER 2 – MANUSCRIPT
H
(0 + 0 ) = 1 − +
~
~ in Ω̂ = H~ and K = p~: The two constants must be the same which can be checked
by comparing the quantum mechanical equation of motion with the classical equation
of motion, say
x p
=
We are now in the position to derive the fundamental differential equation for the time-
evolution operator U( 0 ) and a physical state |i From the composition property of
U
Take → 0+ we obtain
~ U( 0 ) = HU( 0 )
65
CHAPTER 2 – MANUSCRIPT
~ U( 0 ) |i = HU( 0 ) |i
~ | 0 ; i = H | 0 ; i
Here we list several formal solutions to the Schrodinger equation for the time-evolution
operator U( 0 )
~ U( 0 ) = HU( 0 )
U( 0 )
~ = H
U( 0 )
~ ln U( 0 ) = H
Z
ln U( 0 ) − ln U(0 0 ) = − H
~ 0
U( 0 ) = exp[− H( − 0 )]
~
Strictly speaking, the present derivation is not correct completely since U is an operator,
commute
Z Z
ln U( 0 ) = − H()
0 ~ 0
66
CHAPTER 2 – MANUSCRIPT
∙ Z ¸
U( 0 ) = exp − H()
~ 0
[H(1 ) H(2 )] 6= 0
Z Z Z 1
(−)2
U( 0 ) = 1 − 1 H(1 ) + 1 2 H(1 )H(2 ) + · · ·
~ 0 ~2 0 0
Z Z 1 Z −1
P∞ −
= 1+ ( ) 1 2 × H(1 )H(2 ) · · · H( )
=1 ~ 0 0 0
Energy Eigenkets:
|i, we consider a special base ket used in expanding |i Assume that we know all
¯ ® ¯ ®
H ¯() = ¯()
¯ ®
{¯() } forms a complete and orthogonal set of base kets. Therefore
P ¯ () ® () ¯
= − ~
(−0 ) ¯ ¯
¯ ®
Once the expansion of the initial ket |i in terms of {¯() } is known,
P ¯¯ ® () ¯ P ¯ ®
|i = () ¯ i = ¯()
67
CHAPTER 2 – MANUSCRIPT
H
| 0 ; i = − ~ (−0 ) |i
P ¯ () ®
= − ~
(−0 ) ¯
In other words
( = 0 ) = → () = − ~
(−0 )
The phase of () changes with time, but its modulus remains unchanged. If all
eigenvalues are not degenerate, the relative phases of do vary with time and the
A special case: if the initial state happens to be one of energy eigenkets, say
¯ ®
|i = ¯() we have
¯ () ®
| 0 ; i = − ~
(−0 ) ¯
The state remains unchanged. Thus the basic task in quantum dynamics is reduced
to find a complete set of energy eigenket and eigenvalues. It is very helpful if we can
find a maximal set of commuting observables A, B, C,...which also commute with the
Hamiltonian
68
CHAPTER 2 – MANUSCRIPT
In the case, the energy eigenket can be characterized by the indices { } ≡ {}
For example, ⎧
⎪
⎪
⎪
⎪ A |i = |i
⎪
⎪
⎪
⎪
⎪
⎪
⎨ B |i = |i
⎪
⎪
⎪
⎪ ···
⎪
⎪
⎪
⎪
⎪
⎪
⎩ H |i = |i
sion.
function of time. To end this we treat an example here: spin precession. We start with
=− S·B=− S = S
As ⎛ ⎞
~⎜ 1 0 ⎟
S = ⎜ ⎟
2⎝ ⎠
0 −1
69
CHAPTER 2 – MANUSCRIPT
~
± = ± (= )
2
Since the Hamiltonian is time-independent, the time-evolution operator for the state
ket
1
=⇒ |±i = ± ~ |±i .
2
i.e., |±i are also the energy eigenkets with eigenvalues ±~2 Suppose at = 0
At a later time t,
= + − 2 |+i + − + 2 |−i
70
CHAPTER 2 – MANUSCRIPT
Specifically, (1) Let us suppose the initial state |i is the spin up state |+i i,e. + = 1
and − = 0 then
| 0 : i = − 2 |+i
This is a stationery state ! (2) Suppose the initial state |i is the spin state | +i :
1 1
| +i = |+i + |−i
212 212
At a later time t
1 1
| 0 : i = − 2 |+i + 2 |−i
212 212
The probability for the system to be found in the states | ±i = √1 (|+i ± |−i):
2
h + | 0 : i
1 1
= (h+| + h−|) (− 2 |+i + 2 |−i)
212 212
1 −
= ( 2 + 2 ) = cos
2 2
h − | 0: i = − sin
2
|h + | 0: i|2 = cos2 ;
2
|h − | 0: i|2 = sin2
2
71
CHAPTER 2 – MANUSCRIPT
~ ~
= |h + | 0: i|2 − |h − | 0: i|2
2 2
~
= cos
2
Similarly, we have
~
hS i = sin ;
2
hS i = 0
Unitary operators are used for many different purposes in quantum mechanics. It
satisfies that U† U = 1 Under the unitary transformation that changes the state kets
it is important to keep in mind that the inner product of a state bra and a state ket
remains unchanged,
®
h|i = |U† U|
72
CHAPTER 2 – MANUSCRIPT
¡ ¢ ¡ ¢
h| U† X (U |i) = h| U† XU |i
that follows from the associative axiom of multiplication. This identity suggests two
two approaches:
Approach I:
¯ ¯ ®
hXi = h 0 : | X | 0 : i = ¯U† XU¯
Approach II:
|i → |i
X → U† XU ≡ X()
¯ ¯ ®
hXi = h |X()| i = ¯U† XU¯
hXi = hXi
73
CHAPTER 2 – MANUSCRIPT
Approach I (The Schrodinger Picture): The state ket varies with time while
|i → U |i X → X
Recall the Schrodinger equation for the wave function, i.e., the state.
Approach II (The Heisenberg Picture): The operators vary with time while
We denote the operator X() in the Heisenberg picture. U() = U( 0 ) The
74
CHAPTER 2 – MANUSCRIPT
We now derive the fundamental equation of motion in the Heisenberg picture. Assume
() £ † ¤
X = U ()X() U()
µ ¶
†
= U () X() U() + U† ()X() U()
1 † 1
= − U ()HX() U() + U† ()X() HU()
~ ~
1 † 1
= + U ()X() U()U† ()HU() − U† ()HU()U† ()X() U()
~ ~
1 ()
= [X H() ]
~
()
~ X = [X() H() ]
H() = U† HU = H
Thus
()
~ X = [X() H]
75
CHAPTER 2 – MANUSCRIPT
of the same form as in classical physics; we merely replace the classical 0 and
assumption we can reproduce the correct classical equation in the classical limit.
2. When the physical system in question has no classical analogues, we can only
guess the structure of the Hamiltonian. We try various forms until we get the
[x (p)] = ~
and
[p (x)] = −~
where F and G are functions that can be expanded in powers of 0 and 0 respectively.
EXP Example 1
76
CHAPTER 2 – MANUSCRIPT
mechanics:
p2 1
H= = (p2 + p2 + p2 )
2 2
bers)
1
p = [p H] = 0
~
1 p p (0)
x = [x H] = =
~
So
p (0)
x () = x (0) +
We know
but
~
[x () x (0)] = −
EXP Example 2
A particle in a potential.
p2
H= + V()
2
77
CHAPTER 2 – MANUSCRIPT
p 1
= [p V()] = − V()
~
x 1 p
= [x H] =
~
Furthermore,
2 x 1 x 1 p 1 p
2
= [ H] = [ H] ≡
~ ~
2 x
= −∇ ()
2
Newtonian Equation?
tonian is
2
= + 2 2
2 2
78
CHAPTER 2 – MANUSCRIPT
where and are Hermitian since they are physically dynamic variables. The quantum
condition is
[ ] = − = ~
As the first step to solve the problem, we evaluate the commutator basket of and
with
1 ~
[ ] = [ 2 ] =
2
2
[ ] = [ 2 ] = −~ 2
2
~
[ + ] = ~ 2 +
1 1
= ~ 2 ( + )
2 2
1
=±
We have
[ ± ] = ±~( ± )
As
2~
[ + − ]=
we take
³ ´12 ³ ´12
= ( + ); † = ( − )
2~ 2~
79
CHAPTER 2 – MANUSCRIPT
such that
[ † ] = 1
Furthermore,
1
† = ( − )( + )= −
2~ ~ 2
Thus
1
= ~( + )
2
where
= †
Obviously,
[ ] = 0
eigenvalues n. So
|i = |i
and
1
|i = ( + )~ |i
2
What’s ?
80
CHAPTER 2 – MANUSCRIPT
[† ] = −†
As a result,
© ª
† |i = [ † ] + † |i
= ( + 1)† |i
Likewise,
operator and † the increasing operator. From the point of view of energy, the increase
energy. Hence the terms “creation and annihilation operator” for † and are deemed
|i = | − 1i
¯ + ¯ ®
¯ ¯ = = | |2 ≥ 0
81
CHAPTER 2 – MANUSCRIPT
|i = 12 | − 1i
Likewise,
† |i = ( + 1)12 | + 1i
the sequence shall not terminate and leads to eigenkets with negative eigenvalues. This
1
=0 = ~
2
1
|0i 2
~
3
|1i = † |0i 2
~
1 5
|2i = 212
(† )2 |0i 2
~
1 7
|3i = (3!)12
(† )3 |0i 2
~
.. ..
. .
1
|i = (!)12
(† ) |0i ( + 12 )~
82
CHAPTER 2 – MANUSCRIPT
0 ¯ †¯ ®
¯ ¯ = [ + 1]12 0 +1
83
CHAPTER 2 – MANUSCRIPT
|0i = 0
=⇒ h || 0i = 0
Z
=⇒ 0 h || 0 i h0 |0i = 0
Recall
³ ´12
= ( + )
2~
h || 0 i = ( − 0 )
h || 0 i = −~( − 0 )
~
=⇒ ( + ) h|0i = 0
84
CHAPTER 2 – MANUSCRIPT
Figure 2.1: Energy eigenfunctions for the first six states of the harmonic oscillator.
1 1 2
h|0i = 1 12
exp(− 2 )
4 0 2 0
In general
* ¯ ¯ + * ¯ ¯ +
¯ († ) ¯ ¯ († ) ¯
¯ ¯ ¯ ¯
h|i = ¯ ¯0 = ¯ ¯ h|0i
¯ (!) ¯
12 ¯ (!) ¯
12
− +1 2
0 2
2 − 220
= 1 ( − 0 )
4 (2 !)12
So far we have not discussed the time evolution of oscillator state ket or observables
like and . Everything we did is supposed to hold at some instant of time, say = 0.
85
CHAPTER 2 – MANUSCRIPT
Figure 2.2: The position probability density for the state n=10 of a harmonic oscillator
(solid curve) and for a classical oscillator of the same total energy (dashed line).
Now we come to see the time evolution of and . In the Heisenberg picture,
1
= [ ] = − 2 ;
~
1
= [ ] =
~
2 1 1
2
= [ ] = [−2 ] = − 2 ;
~ ~
2 1 1
2
= [ ] = [ ] = −2 ;
~ ~
Equivalently,
= − =⇒ () = (0)−
†
= † =⇒ † () = † (0)
86
CHAPTER 2 – MANUSCRIPT
(0) −
() + () = ((0) + )
(0)
() − () = ((0) − )
1
() = (0) cos + (0) sin
we have
∙ ¸12
~ ∗ ∗ −
h |()| i = ( 1 + 0 1 )
2 0
It is noted that the average value of the position () in an energy eigenstate is always
zero.
We have seen that an energy eigenstates does not behave like the classical oscillator —
in the sense of oscillating expectation values for and — no matter how large n may
87
CHAPTER 2 – MANUSCRIPT
be.
imitates the classical oscillator? In wave-function language, we want a wave packet that
bounces back and forth without spreading in shape. It turns out that a coherent state
A coherent state:
|i = |i
Properties of |i
P∞
1. The coherent state |i = =0 () |i
̄
|()|2 = ( exp(−̄))
!
2. It can be obtained by translating the oscillator ground state by some finite dis-
tance.
® ®
(∆())2 (∆ ())2 = ~2 4
4. = (̄)12
88
CHAPTER 2 – MANUSCRIPT
Oscillator
We take
~
20 =
2
Ψ → − 2
We assume
2
Ψ → ()− 2
Then
00 0
− 2 + ( − 1) = 0
89
CHAPTER 2 – MANUSCRIPT
00 0
− 2 + 2 = 0
such that
1
= 2 + 1; = ( + )~ = 0 1 2
2
µ ¶
2 2
− 2 + Ψ = Ψ
∙µ ¶µ ¶ µ ¶µ ¶¸
1
+ − + − + Ψ = Ψ
2
µ ¶
1
= √ + ;
2
µ ¶
† 1
= √ − ;
2
[ † ] = 1
¡ † ¢
2 + 1 Ψ = Ψ
Ψ0 = 0
90
CHAPTER 2 – MANUSCRIPT
expand the initial ket in terms of the eigenkets of an observable that commute with H.
| 0 i
= exp(− ( − 0 )) | 0 i
~
P ¯ ®
= exp(− ( − 0 )) |0 i 0 ¯ ̇0
{0 } ~
P
= exp(− 0 ( − 0 )) |0 i h0 | 0 i
{0 } ~
h| 0 : i
P
= exp(− 0 ( − 0 )) h|0 i h0 | i
0 ~
Z
P
= 0 exp(− 0 ( − 0 )) h|0 i h0 |0 i h0 | 0 i
0 ~
Z
= 0 ( 0 00 ) h0 | 0 i
where
P
( 0 00 ) = h|0 i h0 |0 i − ~ 0 (−0 )
0
91
CHAPTER 2 – MANUSCRIPT
( ) ≡ h | 0 : i
we have
Z
( ) ≡ 0 ( 0 00 ) (0 )
We call the kernel of the integral operator as the propagator in the wave mechanics.
In any given problem the propagator depends only on the potential and is independent
of the initial and final wave-function. It can be constructed once the energy eigenvalues
and eigen functions are given. Clearly the time evolution of the wave function is com-
pletely predicted if is known and (0 ) is given initially. The only peculiar feature,
if any, is that when a measurement intervenes, the wave function changes abruptly in
an uncontrollable way into one of the eigen functions of observable being measured.
Two Properties of :
~ ( 0 00 ) = ( 0 )
G Proof.
P
( 0 00 ) = h|0 i h0 |0 i − ~ 0 (−0 )
0
92
CHAPTER 2 – MANUSCRIPT
P
~ ( 0 00 ) = h|0 i h0 |0 i 0 − ~ 0 (−0 )
0
P
= h| 0 − ~ 0 (−0 ) |0 i h0 |0 i
0
P
= h| − ~ (−0 ) |0 i h0 |0 i
0
= h| − ~ (−0 ) |0 i
= ()( 0 )
1.
G Proof.
P 0
lim ( 0 00 ) = lim h|0 i h0 |0 i − ~
(−0 )
→0 →0 0
P
= lim h|0 i h0 |0 i = h|0 i = ( − 0 )
→0 0
The propagator K is simply the Green’s function for the time-dependent wave equation
satisfying
~2 2
(− ∇ + () − ~ )( 0 00 ) = −~( − 0 )( − 0 ) (2.1)
2
93
CHAPTER 2 – MANUSCRIPT
The -function ( − 0 ) is needed on the right hand side of the Eq.(2.1) because K
varies discontinuously at = 0
EXP Example 3
A free particle in 1D
P2
H=
2
P |i = |i
2
H |i = |i
2
94
CHAPTER 2 – MANUSCRIPT
Thus
( 0 00 )
∙ ¸
P 0 2
= h|i h| i exp − ( − 0 )
2~
Z ∙ ¸
0 2
= exp − − ( − 0 )
2~ ~ ~ 2~
Z+∞ ∙ ¸
0 − 0 ( − 0 ) 2 ( − 0 )2
= exp − ( − ) +
2~ 2~ − 0 2~( − 0 )
−∞
Z+∞ ∙ ¸
− 0 ( − 0 ) 2 ( − 0 )2
= lim exp − || − ( − ) +
→0+ 2~ 2~ − 0 2~( − 0 )
−∞
∙ ¸12
( − 0 )2
= exp( )
2~( − 0 ) 2~( − 0 )
95
CHAPTER 2 – MANUSCRIPT
(1) = 0
Z
() = ( 0)
Z
P
= h|i h|i − ~
Z
P
= h|i h|i − ~
P
= h|i − ~
= Tr(−~ )
Z∞
̃() ≡ − ()~−~ ( → 0+ )
0
Z∞ X
= − −(− +)~
0
X 1
= ( → 0+ )
− +
EXP Example 4
96
CHAPTER 2 – MANUSCRIPT
r
00 0
( ; 0 ) =
2~ sin[( − 0 )]
∙ ¸
{(002 + 02 ) cos ( − 0 ) − 200 0 }
× exp
2~ sin[( − 0 )]
To gain further insight into the physical meaning of the propagator, we rewrite it in an
P
( 0 0 ) = h|i h|0 i − (−0 )~
X D ¯¯ ¯¯ E D ¯¯ 0 ¯¯ E
= ¯− ~ ¯ ¯ ~ ¯ 0
X D ¯¯ ¯¯ E D ¯¯ 0 ¯¯ E
= ¯− ~ ¯ ¯ ~ ¯ 0
= h |0 0 i
Where | i and |0 0 i are to be understood as an eigenket and of the position operators
in the Heisenberg picture. Roughly speaking, h |0 0 i is the amplitude for the particle
97
CHAPTER 2 – MANUSCRIPT
− 1
− −1 = ∆ = ( = 1 )
−1
h |1 1 i
Z Z Z
= −1 −2 2
98
CHAPTER 2 – MANUSCRIPT
Before proceeding further, it is profitable to review here how paths appear in classical
mechanics. Suppose we have a particle subjected to a force field, say gravitional field.
() ⇔
(1 1 ) ⇔ ( 0)
µ ¶12
2
( ) ⇔ (0 )
1
( ̇) = ̇2 − ()
2
action
Z
( ̇) = 0
1
Lagraingian equation:
− =0
̇
The basic difference between classical mechanics and quantum mechanics should be
particle’s motion; in contrast, in quantum mechanics all possible paths must play roles
99
CHAPTER 2 – MANUSCRIPT
including those which do not bear any resemblance to the classical path. Yet the
1
h |−1 −1 i = exp [( − 1)~]
(∆)
where
Z
( − 1) = ( ̇)
−1
Because at any given time the position kets in the Heisenberg picture form a complete
Z
® ®
= 0 |0 0 h0 0 | i |
=
and so on. If we somehow guess the form of h2 2 |1 1 i for a finite time interval by
compounding the appropriate transition amplitude for infinitesimal time interval. This
100
CHAPTER 2 – MANUSCRIPT
Z
1
( − 1) = ( ̇2 − ())
2
−1
1 − −1 2 − −1
= ∆( ( ) −( ))
2 ∆ 2
we obtain
1 ³ ´12
=
(∆) 2~∆
and
³ ´12 2
lim 2~ ∆ = ()
∆→0 2~∆
To summarize, as ∆ → 0 we obtain
³ ´12 (−1)
h |−1 −1 i = ~
2~∆
101
CHAPTER 2 – MANUSCRIPT
Z Y
−1 (−1)
h |1 1 i = lim ( ) 2 −1 −2 2 ~
→+∞ 2~∆
=2
Z
h |1 1 i ≡ D((1 )) ∫1 (̇)~
1
chanics based on the concept of paths. The ideas we borrowed from the conventional
form of the quantum mechanics are (1) the superposition principle (used in summing
the contribution from various alternate paths), (2) the composition property of the
Function
In classical mechanics, it is well known that the zero point of potential energy is of no
physical significance. The force that appears in Newton’s second law depends only on
102
CHAPTER 2 – MANUSCRIPT
−
= :
− 5 − 5
when → + 0
= → =
Let us look at the time evolution of a Schrodinger picture state ket subject to some
potential
() =⇒ () + 0
2 2
= + () =⇒ = + () + 0
2 2
^
| 0 : i =⇒ | 0 : i =?
| 0 : i = − ~ (−0 ) | 0 i
Thus
0 (−0 )
^
| 0 : i = − ~ | 0 : i
−(−0 )~
103
CHAPTER 2 – MANUSCRIPT
Even though the choice of the absolute scale of the potential is arbitrary,
potential difference are of non-trivial physical significance and in fact, can be detected
in a very striking way. A beam of charged particles is split into two parts, each of
which enters a metallic cage. A particle in the beam can be visualized as a wave packet
whose dimension is much smaller than the dimension of the cage. Inside the cage the
potential is spatially uniform. Hence the particle in the cage experience no force. If we
desire, a finite potential difference between the two cages can be maintained by turning
on a switch. The final state which the two beams reach at interference region is
1 1
| i = | 0 : i + | 0 : i
212 212
Thus
104
CHAPTER 2 – MANUSCRIPT
This is an observable effect! This is a purely quantum mechanical effect and has no
classical analogue.
D = E + 4P
H = B − 4M
with
1
5·A+ =0
105
CHAPTER 2 – MANUSCRIPT
with
1 2
52 Λ − Λ=0
2 2
the Maxwell’s equations keep unchanged! In quantum mechanics, the Hamiltonian for
1
= (p − A)2 +
2
1 2
= (p2 − p · A − A · p + 2 A2 ) +
2
1 2 2
6= (p2 − p · A + 2 A2 ) +
2
1 1
= [ ] = ( − )
~
where = Define
x
≡ =p− A
106
CHAPTER 2 – MANUSCRIPT
~
[ ] = − ( − )
~
= − (5 × A)
2
= +
2
We now study the Schrodinger’s equation with and in the position space.
D ¯ ¯ E
¯ 2¯
¯(p − A()) ¯ 0 :
Z Z D ¯ ¯ ED ¯ ¯ E
¯ ¯ ¯ ¯
= 1 2 ¯(p − A())¯ 1 1 ¯(p − A()) |2 i h2 ¯ 0 :
= (−~ 5 − A())2 h| 0 : i
1
[ (−~ 5 − ())2 + ] h| 0 : i = ~ h| 0 : i
2
Denote
( ) ≡ h| 0 : i
and
= ∗ ()()
107
CHAPTER 2 – MANUSCRIPT
= ( ∗ ) + ∗ = − 5 ·j
+5·j=0
where
~
j= Im(∗ 5 ) − A ||2
For
= 12 ~
j = (5 − A)
B = ̂ = ( − )̂
108
CHAPTER 2 – MANUSCRIPT
(2). = − = 0
A → A + 5( )
2
Let us denote by |i the state ket in the presence of A, and |̃i the state ket in the
presence of A + 5(
2
) The Schrodinger’s equations for these two state kets are
1
[ (p − A)2 + ] | 0 : i = ~ | 0 : i
2
1 ^ ^
[ (p − A − 5 Λ)2 + ]| 0 : i = ~ | 0 : i
2
^
What’s the relation of | 0 : i and | 0 : i ?
|̃i = G |i
with GG † =1
1
G †[ (p − A − 5 Λ)2 + ]G | 0 : i = ~ | 0 : i
2
We require
G † G =
109
CHAPTER 2 – MANUSCRIPT
® D E
(2). |p − A| = ̃|p − Ã|̃ :
³ ´ ³ ´
G† p − A − 5 Λ G = p − A
G † [p G] = + 5 Λ
G † (−~∇G) = + 5 Λ
−~∇ ln G = + 5 Λ
h i
G = exp Λ
~
Therefore, if we take
|̃i = ~ Λ |i
then
1
− ~ Λ ( (p − A − 5 Λ)2 + ) ~ Λ
2
1
= (p − A)2 +
2
̃( ) = ~ Λ ( )
110
CHAPTER 2 – MANUSCRIPT
Figure 2.5:
A → A + ∇Λ
→
→ + Λ
Consider a particle of charge e going above or below a very long impenetrable cylinder,
as shown in Fig.
Inside the cylinder is a magnetic field parallel to the cylinder axis, taken to be
normal to the plane of Fig. So the particle paths above and below enclose a magnetic
flux.
111
CHAPTER 2 – MANUSCRIPT
Hence the particle does not experience the Lorentz force outside cylinder.
⎧
⎪
⎪
⎨ (0 1 0), if
2
A=
⎪
⎪
⎩ (0 1 2 0), if
2
Our object is to study how the probability of finding the particle in the interference
region B depends on the magnetic flux. For pedagogical reason we prefer to use the
1 1 2
(0) 2
= ̇ → ̇ + ̇ · A
2 2
R
(0) ( − 1) → (0) ( − 1) + ̇ · A
−1
R
= (0) ( − 1) + · A
−1
where is the differential line element along the path segment. So when we consider
Q
exp( (0) ( − 1)~)
Q R
=⇒ exp( (0) ( − 1)~) exp( A · )
~ −1
112
CHAPTER 2 – MANUSCRIPT
The probability for finding the particle in the interference region B depends on the
modulus square of the entire transition amplitude and hence on the phase difference
between the contribution from the paths going above and below. i.e.
∝ 1 + cos(1 − 2 )
where
( )
R R
1 − 2 = A · − A ·
~ 1 1
I
= A ·
~
ZZ
= ∇ × A ·
~
=
~
where stands for the magnetic flux inside the impenetrable cylinder. This means
that as we change the magnetic field strength, there is a sinusoidal component in the
probability for observing the particle in region B with a period given by a fundamental
2~
= 4135 × 10−7 Gauss-cm2
||
We emphasize that the interference effect discussed here is a purely quantum mechani-
cal. Classically the motion of a charged particle is determined by the Newton’s second
law supplemented by the force law of Lorentz. In this problem the magnetic field is
113
CHAPTER 2 – MANUSCRIPT
confined in the cylinder, and the particle experiences no force. However the vector
potential exists in the outside of cylinder. It seems that the vector potential A is more
fundamental than the magnetic field B. It is to be noted that the observable effect
depends on the magnetic flux not A directly. This effect has been observed experimen-
tally.
B= r̂
2
∇ · B =4 (r)
B = ∇×A
∙ ¸
1
= r̂ ( sin ) −
sin
∙ ¸
1 1
+θ̂ − ( )
sin
∙ ¸
1
+φ̂ ( ) −
∇ · (∇ × A) = 0
114
CHAPTER 2 – MANUSCRIPT
The solution I
∙ ¸
(1 − cos )
A= φ̂
sin
for − , and
∙ ¸
() (1 + cos )
A =− φ̂
sin
for We can get a solution for B, but A is doubly valued. The difference of the
vector potential is
2
A() − A() = − φ̂
sin
2
∇Λ = ∇(−2 ) = − φ̂
sin
The the wave functions in two different vector potential will differ by
µ ¶
() −2
Ψ = exp Ψ()
~
The wave function must be single valued because once we choose particular gauge,
the expansion of the state ket in terms of the position eigenkets must be unique. Let
115
CHAPTER 2 – MANUSCRIPT
us examine the behavior of the wave function Ψ() , on the equator = 2 with a
definite radius r, which is a constant. When we increase the azimuthal angle along
the equator from 0 to 2, Ψ() as well as Ψ() must return to its original value because
2
= ±
~
~ ~
= ± ±2 ···
2 || 2 ||
We should emphasize that quantum mechanics does not require the existence
of magnetic monopole.
Since
Z Z
| 0 : i = |i h| 0 : i = |i Ψ ( )
116
CHAPTER 2 – MANUSCRIPT
is defined as the probability density. So ( ) is the probability that the particle
appears in a narrow range round at time . This is the so-called the probabilistic
( )
∙ ¸ ∙ ¸
∗ ∗
= Ψ ( ) Ψ( ) + Ψ ( ) Ψ( )
1 1
= − ( ∗ Ψ∗ )Ψ + Ψ∗ Ψ
~ ~
Assume
~2 2
= − ∇ +
2
1 1
( ) = ~(Ψ∗ (∇2 Ψ) − Ψ(∇2 Ψ∗ )) + ( − ∗ )Ψ∗ Ψ
2 ~
117
CHAPTER 2 – MANUSCRIPT
As
~ 1
= ∇(Ψ∗ ∇Ψ − Ψ∇Ψ∗ ) + Im( ) ·
2 ~
1
+∇·j = Im( ) ·
~
~ ∗ ~
j = − (Ψ ∇Ψ − Ψ∇Ψ∗ ) = Im(Ψ∗ ∇Ψ)
2
+∇·j=0
ten used for nuclear reaction for particles absorbed by nuclei, it is accounted for the
disappearance of particles.
Except for the probability density, the wave function contains more physics
118
CHAPTER 2 – MANUSCRIPT
Ψ∗ ∇Ψ = 12 ∇12 + ∇
~
=⇒ j = ∇
The spatial variation of the phase of the wave function characterizes the probability
flux; the stronger the phase variation, the more intense the flux. The direction of j at
some point is seen to be normal to the surface of a constant that goes through that
j = (∇) ≡ v
~2 2 1
− (∇2 12 + ∇12 · ∇) − 2 12 |∇|2 + ( )12 ∇2 + 12
2 ~ ~ ~
12
= ~( + 12 )
~
Taking ~ → 0 , we obtain
1
|∇( )|2 + () + ( ) = 0
2
119
CHAPTER 2 – MANUSCRIPT
Figure 2.6: One-dimensional square well potential with (a) perfectly rigid walls (b)
finite potential.
This is the Hamilton-Jacobi equation in classical mechanics. In this sense, the Schrodinger
equation goes back to the classical mechanics in the limit ~ → 0 In the classical me-
chanics ( ) stands for Hamiltonian principal function. In a stationary state with
( ) = () −
120
CHAPTER 2 – MANUSCRIPT
A 2.8 Examples
Since the potential is infinite at || the wave function must vanish at the points
~2 2
− =
2 2
2 22
2 = − 2
= −2
~
sin + cos = 0
− sin + cos = 0
121
CHAPTER 2 – MANUSCRIPT
1
= ( + ) (2.2)
2
(2 + 1)
= cos = cos (2.3)
2
~2 2 (2 + 1)2 2 ~2
= = (2.4)
2 2 82
= (2.5)
= sin = sin (2.6)
~2 2 2 2 ~2
= = (2.7)
2 2 22
~2 2
− + 0 =
2 2
~2 2
− = ( − 0 )
2 2
122
CHAPTER 2 – MANUSCRIPT
For 0
⎧
⎪
⎪
⎨ 1 sin + 1 cos
() =
⎪
⎪
⎩ 2 sin + 2 cos
The wave equation at || is the same as that in the rigid wall potential, and the
where = (2~2 )12 We now impose on the solutions (1) and (2) the requirements
123
CHAPTER 2 – MANUSCRIPT
2 sin = ( − )−
2 cos = ( + )−
(1) = 0 and =
tan =
(2) = 0 and = −
cot = −
C Parity
~2 2 ()
− + ()() = ()
2 2
~2 2 (−)
− + ()(−) = (−)
2 2
124
CHAPTER 2 – MANUSCRIPT
The () and (−) are solutions of the same wave equation with the same eigenvalues
E. Addition or subtraction these two equations gives () ± (−) the solution of the
Schrodinger equation. To simplify the notation, () = ±(−) Such wave function
cos = cos(−)
sin = − sin(−)
Consider a charged particle in a uniform magnetic field. Assume the particle is confined
in a two-dimensional plane, and the magnetic field is perpendicular to the plane, say
1
= ( − )2
2
with
= ̂ = ( − )̂
(1). = − = 0
125
CHAPTER 2 – MANUSCRIPT
1 h 2 2
i
= ( − ) + ( − )
2
∙ ¸
1 2 2
= ( − ) + ( )
2
Notice that [ ] = 0 i.e. p is a good quantum number. The general form of the
wavefunction is
Ψ = ()0
∙ ¸
1 2 2
(0 − ) + ( ) () = ()
2
Alternatively
∙ ¸
2 2 2 2
+ ( − 0 ) () = ()
2 22
0
with 0 =
Thus the problem is reduced to the simple harmonic oscillator.
1 h i
= ( − )2 + ( − )2
2
∙ ¸
1 2 2
= ( − ) + ( + )
2 2 2
126
CHAPTER 2 – MANUSCRIPT
In this case both p and p are not good quantum numbers. We cannot solve this
= − ;
2
= +
2
~
[ ] = −2~ = −
2
1 £ 2 ¤
= + 2
2
1
= [( + )( − ) + ( − )( + )]
4
2~
[ − + ] =
Take
r
= ( − )
2~
r
†
= ( + )
2~
Thus [ † ] = 1
~ ¡ † ¢
= + †
2
µ ¶ µ ¶
~ † 1 † 1
= + ⇔ ~ +
2 2
127
CN Chapter 3
This chapter is concerned with a systematic treatment of angular momentum and re-
lated topics.
a ket |i or the wave function (r) We describe a rotation by a linear operator ,
which is so defined that any vector r is rotated into the new vector · r. The rotation
changes the ket |i into the new ket |0 i or change the wave function (r) into the
hr|i = hr|i
128
CHAPTER 3 – MANUSCRIPT
¯ + ¯ ®
r ¯D ()D()¯ = hr|i
hr|0 i = hr|i
( ) ( ) = ( ) ( ) = ( + ) = ( )
6 3 3 6 3 6 2
( ) ( ) 6= ( ) ( )
2 2 2 2
the vector, we obtain a new vector, V with and . These two vectors are
connected by a 3 × 3 matrix
⎛ ⎞ ⎛ ⎞⎛ ⎞
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ ⎟ ⎜ ⎟⎜ ⎟
⎜ ⎟ ⎜ ⎟⎜ ⎟
⎜ ⎟=⎜ ⎟⎜ ⎟
⎜ ⎟ ⎜ ⎟⎜ ⎟
⎜ ⎟ ⎜ ⎟⎜ ⎟
⎝ ⎠ ⎝ ⎠⎝ ⎠
129
CHAPTER 3 – MANUSCRIPT
The requirement that the rotated vector be real when the vector are real means
that the elements of are real. The length of the vector and do not change,
which means
⎛ ⎞
⎜ ⎟
⎜ ⎟
⎜ ⎟
( ) ⎜
⎜
⎟ = 2 + 2 + 2
⎟
⎜ ⎟
⎝ ⎠
= =
= = 1
where “T” stands for a transpose of a matrix: = To be definite we consider a
130
CHAPTER 3 – MANUSCRIPT
=
Similarly ⎛ ⎞
⎜ 1 0 0 ⎟
⎜ ⎟
⎜ ⎟
() = ⎜
⎜ 0 cos − sin
⎟
⎟
⎜ ⎟
⎝ ⎠
0 sin cos
⎛ ⎞
⎜ cos 0 sin ⎟
⎜ ⎟
⎜ ⎟
() = ⎜
⎜ 0 1 0 ⎟
⎟
⎜ ⎟
⎝ ⎠
− sin 0 cos
We are particularly interested in an infinitesimal form, from which a great deal can be
131
CHAPTER 3 – MANUSCRIPT
2
learned about the structure of . Take sin ≈ cos ≈ 1 − 2 we have
⎛ ⎞
⎜ 1 0 0 ⎟
⎜ ⎟
⎜ ⎟
() = ⎜
⎜ 0 1 − 2 2 − ⎟
⎟
⎜ ⎟
⎝ ⎠
0 1 − 2 2
= 1 + () + (2 ) + · · ·
⎛ ⎞
⎜ 1 − 2 2 0 ⎟
⎜ ⎟
⎜ ⎟
() = ⎜
⎜ 0 1 0 ⎟
⎟
⎜ ⎟
⎝ ⎠
− 0 1 − 2 2
⎛ ⎞
⎜ 1 − 2 2 − 0 ⎟
⎜ ⎟
⎜ ⎟
() = ⎜
⎜ 1 − 2 2 0 ⎟
⎟
⎜ ⎟
⎝ ⎠
0 0 1
⎛ ⎞⎛ ⎞ ⎛ ⎞⎛
⎜ 1 0 0 ⎟ ⎜ 1 − 2 2 0 ⎟ ⎜ 1 − 2 2 0
⎟⎜ 1 0
⎜ ⎟⎜ ⎟ ⎜ ⎟⎜
⎜ ⎟⎜ ⎟ ⎜ ⎟⎜
⎜ 0 1 − 2 2 − ⎟⎜ 0 1 ⎟−⎜
0 0 1 0 ⎟ ⎜ 0 1 − 2 2
⎜ ⎟⎜ ⎟ ⎜ ⎟⎜
⎜ ⎟⎜ ⎟ ⎜ ⎟⎜
⎝ ⎠⎝ ⎠ ⎝ ⎠⎝
0 1 − 2 2 − 0 1 − 2 2 − 0 1 − 2 2 0 1
⎛ ⎞
¡ ¢
⎜ 0 −2 − − 12 2 + 1 ⎟
⎜ ⎟
⎜ ¡ ¢ ⎟
⎜ 2 0 − − 12 2 + 1 ⎟
⎜ ⎟
⎜ ⎟
⎝ ¡ 1 2 ¢ ¡1 2 ¢ ¡ 1 2 ¢2 ¡ 1 2 ¢ ¡1 2 ¢ ⎠
− −2 + 1 + 2 − 1 −2 + 1 + −2 + 1 2 − 1
132
CHAPTER 3 – MANUSCRIPT
⎛ ⎞
2
⎜ 0 − 0 ⎟
⎜ ⎟
⎜ ⎟
= ⎜ 2 0 0 ⎟
⎜
⎟
⎜ ⎟
⎝ ⎠
0 0 0
Elementary matrix manipulation leads to
If we just remain the quantities of first order in , any two rotations about different
In quantum mechanics
rotated system is expected to look different from the state ket corresponding to the
associated with an operator D() in the appropriate ket space such that
where |i and |i stand for the kets of the rotated and unrotated systems, respectively.
To construct the rotation operator, it is again fruitful to examine first its properties
J·n
D(n) = 1 −
~
133
CHAPTER 3 – MANUSCRIPT
D () = lim [D ( )]
→+∞
= lim (1 − · )
→+∞ ~
∙ ¸− ~
= lim (1 − · )~(− )
→+∞ ~
= exp(− )
~
1
Definition: lim (1 + ) ≡
→+∞
more concepts. For every , there exists a rotation operator D() in the appropriate
ket space
=⇒ D()
1. Identity:
2. Closure:
3. Inverse:
134
CHAPTER 3 – MANUSCRIPT
4.Associativity:
As
we examine
2 2 2 2
= (1 − − )(1 − − )
~ 2~2 ~ 2~2
2 2 2 2
−(1 − − )(1 − − )
~ 2~2 ~ 2~2
2
D (2 ) = 1−
~
=⇒ [ ] = ~
Similarly, we obtain
[ ] = ~
135
CHAPTER 3 – MANUSCRIPT
or
J × J = ~J
v × v = 0
If the vector is of infinitesimal length and only quantities of first order in are
= + ×
where ⎛ ⎞
⎜ 1 − ⎟
⎜ ⎟
⎜ ⎟
=⎜
⎜ 1 − ⎟
⎟
⎜ ⎟
⎝ ⎠
− 1
136
CHAPTER 3 – MANUSCRIPT
To derive the formula we have used the infinitesimal rotation, for example,
⎛ ⎞
⎜ cos − sin 0 ⎟
⎜ ⎟
⎜ ⎟
() = ⎜
⎜ sin cos 0 ⎟
⎟
⎜ ⎟
⎝ ⎠
0 0 1
⎛ ⎞
⎜ 0 − 0 ⎟
⎜ ⎟
⎜ ⎟
= 1+⎜
⎜ 0 0 ⎟
⎟
⎜ ⎟
⎝ ⎠
0 0 0
Now we wish to find a transformation |i that changes the ket |i into the
ket |i
h |i = h |i
®
=⇒ h |i = −1 |
137
CHAPTER 3 – MANUSCRIPT
Thus we have
≈ ( − × )
=⇒ D() = 1 − · + · · ·
~
where
L=r×P
= − (= −~( − ))
= − (= −~( − ))
= − (= −~( − ))
[ ] = ~
[ ] = ~
138
CHAPTER 3 – MANUSCRIPT
Consider a rotation by a finite angle about the z-axis. If the ket of a spin 1/2 system
with
D () = − ~
~
= ~ − ~
= ~ (|+i h−| + |−i h+|) − ~
2
~ ³
´
= 2 |+i h−| 2 + − 2 |−i h+| − 2
2
~
= [(|+i h−| + |−i h+|) cos + (|+i h−| − |−i h+|) sin ]
2
= cos − sin
⎛ ⎞ ⎛ ⎞
⎜ ⎟ ⎜ cos − sin ⎟
⎜ ⎟ ⎜ ⎟
⎜ ⎟ ⎜ ⎟
D () ⎜ ⎟ D () = ⎜
† ⎜ ⎟
⎜ sin + cos
⎟
⎟
⎜ ⎟ ⎜ ⎟
⎝ ⎠ ⎝ ⎠
⎛ ⎞⎛ ⎞
⎜ cos − sin 0 ⎟ ⎜ ⎟
⎜ ⎟⎜ ⎟
⎜ ⎟⎜ ⎟
⎜ ⎟ ⎜
= ⎜ sin cos 0 ⎟ ⎜ ⎟ ⎟
⎜ ⎟⎜ ⎟
⎝ ⎠⎝ ⎠
0 0 1
139
CHAPTER 3 – MANUSCRIPT
Compare with
=
The Hamiltonian is
= − S·B
= S
140
CHAPTER 3 – MANUSCRIPT
Conclusion: the period for the state ket is twice as long as the period for spin precession:
precession = 2;
The branch of mathematics that is appropriate for a full treatment of symmetry is the
theory of groups. Here we give a few basic definitions: A set of objects , , ,... form
a group if a process can be defined that enables us to combine any two of the objects,
such as and , to form an object , and if the following conditions are satisfied:
2. The group contains an identity or unit member that has the properties =
3. Each member has an inverse −1 also in the group, such that −1 = −1 = .
() = ()
141
CHAPTER 3 – MANUSCRIPT
The members of the group are called elements. Though we frequently refer to
the combination as “multiplication”, this does not mean ordinary multiplication. For
example, the set of integers, positive, negative, and zero, form a group if the law of
Two groups are said to be isomorphic to each other if there is a unique one-
to-one correspondence between elements of the two groups such that products of cor-
Examples:
(1). The elements are 1 and −1 and the law of combination is multiplication.
We discussed rotations of a vector. The rotated vector and unrotated vector are con-
=
142
CHAPTER 3 – MANUSCRIPT
2. Identity: = 1
This group is named SO(3), where S stands for special, O stands for orthog-
onal, 3 for three dimension. If the vector is n-dimensional, the R form a SO(n) group.
B 3.2.3 “Special”?
Consider ⎛ ⎞
⎜ cos − sin 0 ⎟
⎜ ⎟
⎜ ⎟
() = ⎜
⎜ sin cos 0 ⎟
⎟
⎜ ⎟
⎝ ⎠
0 0 1
Its determinant is
det( ()) = 1
In fact,
det() = 1 =⇒ ””
−1 is not one of elements in SO(N). We postulate that D() has the same group
143
CHAPTER 3 – MANUSCRIPT
since
det[ ] = 0
Another rotation we discussed is for the spin 1/2 system, where the rotation matrices
where a and b are complex satisfying ||2 + ||2 = 1 All these matrices form a group.
(1) Closure:
where
(2) Identity: ⎛ ⎞
⎜ 1 0 ⎟
(1 0) = ⎜
⎝
⎟
⎠
0 1
(3) Inverse:
† ( )( ) = 1
144
CHAPTER 3 – MANUSCRIPT
If we set
Re() = cos
2
Re() = − sin
2
Im() = − sin
2
Im() = − sin
2
·
( ) =⇒ − 2
As det[( )] = 1, this group is called SU(2) group. SU(2) and SO(3) have a two-to-one
In SO(3) , a rotation 2 produces 1. More generally, ( ) and (− −) correspond
To describe a rotation, usually we need three parameters: two parameters for the
rotation axis and one parameter for the rotation angle. In classical mechanics, an
arbitrary rotation of a rigid body can be accomplished in three steps, known as Euler
rotations.
145
CHAPTER 3 – MANUSCRIPT
146
CHAPTER 3 – MANUSCRIPT
Three steps:
(a) : R () y → y0
In terms of 3×3 orthogonal matrices the product of the three operations can be written
as
There appear both rotations about body axes and about the space-fixed axes. This
Therefore,
3 2 3
= − 2
− 2
− 2
147
CHAPTER 3 – MANUSCRIPT
Recall that
·
exp(− ) = cos − · sin
2 2 2
tum
[ ] = ~
[ ] = ~
[ ] = ~
148
CHAPTER 3 – MANUSCRIPT
× = ~
In this section we work out the eigenvalues and eigenkets of angular momentum. To
(1). 2 = + +
[ 2 ] = 0
2 | i = | i
| i = | i
To determine the values of and , it is convenient to work with the ladder operator,
± .
[ ] = ~
149
CHAPTER 3 – MANUSCRIPT
[ ± ] = ±~±
[+ − ] = 2~
[± 2 ] = 0
0 0 ¯ 2¯ ®
¯ ¯ = 0 0 ;
Then
= ( ± ~)± | i
150
CHAPTER 3 – MANUSCRIPT
1
2 = (+ − + − + ) + 2
2
= − + + 2 + ~ = + − + 2 − ~
If + ( ) or − ( ) is not equal to zero, | ± ~i is also one of eigenkets. Suppose
that we keep on applying ± to both side of the equation above. We can obtain
numerical eigenkets with smaller and smaller or larger and larger ± ~ until the
151
CHAPTER 3 – MANUSCRIPT
So
=⇒ max = −min
Clearly, we must be able to reach | max i by applying + successively to | min i a
We obtain
max = min + ~
As a result,
max = ~
2
and
= ( + 1)~2
2 2
It is conventional to define
max = −min = ~ = ~
2
152
CHAPTER 3 – MANUSCRIPT
and
| i =⇒ | i
such that
| i = ~ | i
The eigenkets {| i} form a basis for angular momentum operator
h | 0 0 i = 0 0
As a result,
153
CHAPTER 3 – MANUSCRIPT
¿ ¯ ¯ À
0 0
¯ ¯
0 0 ¯ + + − ¯
h | | i = ¯ ¯
2
~
= [( + 1) − ( + 1)]12 0 0 +1
2
~
+ [( + 1) − ( − 1)]12 0 0 −1
2
h 0 0 | | i
~
= [( + 1) − ( + 1)]12 0 0 +1
2
~
− [( + 1) − ( − 1)]12 0 0 −1
2
0 ¯ 2¯ ®
¯ ¯ = ( + 1)~2 0
Our choice of a representation in which 2 and are diagonal has led to discrete
sequences of values for the corresponding labels j and m. The infinite matrices thus
obtained are most conveniently handled by breaking them up into an infinite set of
154
CHAPTER 3 – MANUSCRIPT
(2) = 1 = ±1 0
⎛ ⎞ ⎛ ⎞
⎜ 0 1 0 ⎟ ⎜ 0 − 0 ⎟
⎜ ⎟ ⎜ ⎟
⎜ ⎟ ⎜ ⎟
= 212 ⎜ 1 0 1 ⎟
~ ⎜
⎟ =
~
212
⎜ − 0 − ⎟
⎜ ⎟
⎜ ⎟ ⎜ ⎟
⎝ ⎠ ⎝ ⎠
0 1 0 0 0
⎛ ⎞ ⎛ ⎞
⎜ 1 0 0 ⎟ ⎜ 1 0 0 ⎟
⎜ ⎟ ⎜ ⎟
⎜ ⎟ ⎜ ⎟
= ~ ⎜
⎜ 0 0 0
⎟
⎟ = 2~ ⎜ 0 1 0 ⎟
2 2⎜
⎟
⎜ ⎟ ⎜ ⎟
⎝ ⎠ ⎝ ⎠
0 0 −1 0 0 1
(3) = 32 = ± 32 ± 12
⎛ ⎞
12
⎜ 0 3 0 0 ⎟
⎜ ⎟
⎜ ⎟
⎜ 312 0 2 0 ⎟
~⎜ ⎟
= ⎜ ⎟
2⎜
⎜ 0
⎟
⎟
⎜ 2 0 312 ⎟
⎜ ⎟
⎝ ⎠
0 0 312 0
155
CHAPTER 3 – MANUSCRIPT
⎛ ⎞
12
⎜ 0 −3 0 0 ⎟
⎜ ⎟
⎜ ⎟
⎜ −312 0 −2 0 ⎟
~⎜ ⎟
= ⎜ ⎟
2⎜
⎜
⎟
⎟
⎜ 0 2 0 −312 ⎟
⎜ ⎟
⎝ ⎠
0 0 312 0
⎛ ⎞
⎜ 3 0 0 0 ⎟
⎜ ⎟
⎜ ⎟
⎜ 0 −1 0 0 ⎟
~⎜ ⎟
= ⎜
⎜
⎟
⎟
2⎜ ⎟
⎜ 0 0 −1 0 ⎟
⎜ ⎟
⎝ ⎠
0 0 0 −3
⎛ ⎞
⎜ 1 0 0 0 ⎟
⎜ ⎟
⎜ ⎟
⎜ 0 1 0 0 ⎟
15 2 ⎜ ⎟
= ~ ⎜
2 ⎟
4 ⎜ ⎜ 0 0 1
⎟
⎜ 0 ⎟
⎟
⎜ ⎟
⎝ ⎠
0 0 0 1
Have obtained the matrix elements of and ± , we are now in position to study the
matrix elements of rotation operator D()
¯ ¯ ®
D0 () = 0 ¯−·̂~ ¯
()
Since [ 2 ] = 0 we have
[D() 2 ] = 0
156
CHAPTER 3 – MANUSCRIPT
So D() | i is still an eigenket of 2 with the same eigenvalue ( + 1)~2
We just consider the matrix elements of D() with the same . For every there are
ket
X X ()
() |i = |0 i h0 |()| i = 0 () |0 i
0 0
As is well known, the Euler angles may be used to characterize the most general rotation.
We have
D ¯ ¯ E
() ¯ ¯
0 ( ) = 0 ¯− ~ − ~ − ~ ¯
D ¯ ¯ E
0 ¯ ¯
= ( −) 0 ¯− ~ ¯
For = 12 ⎛ ⎞
~⎜ 0 − ⎟
= ⎜ ⎟
2⎝ ⎠
0
⎛ ⎞
nD ¯ ¯ Eo ⎜ cos
− sin
⎟
0 ¯ − ¯ 2 2
¯ ~ ¯ = ⎜
⎝
⎟
⎠
sin 2 cos 2
= cos 0 − sin
2 2
157
CHAPTER 3 – MANUSCRIPT
To derive the matrix we can make use of the identity: ( })2 = 14
For = 1, ⎛ ⎞
⎜ 0 − 0 ⎟
⎜ ⎟
~ ⎜ ⎟
= 12 ⎜ 0 − ⎟
⎜
⎟
2 ⎜ ⎟
⎝ ⎠
0 0
⎛ ⎞
⎜ cos2 2
1
− 212 sin sin2 2 ⎟
D ¯ ¯ E ⎜ ⎟
0 ¯ − ¯
⎜ ⎟
{ ¯ ~ ¯ } = ⎜ 1
⎜ 212 sin cos 1
− 212 sin ⎟
⎟
⎜ ⎟
⎝ ⎠
sin2 2 1
212
sin cos2 2
To derive the matrix we can make use of the identity: ( })3 = }
There exists a very interesting connection between the algebra of angular momentum
and the algebra of two independent simple harmonic oscillators. Let us consider two
types of oscillator: plus and minus type. The creation and annihilation operator are
158
CHAPTER 3 – MANUSCRIPT
+ = †+ + ; − = †− −
We have
such that
à !à !
(†+ )+ (†− )−
|+ − i = |0 0i
(+ !)12 (− !)12
with
± |0 0i = 0
We define
+ ≡ ~†+ − ; − ≡ ~†− +
~ † 1 ~
≡ (+ + − †− − ) = [+ − ] = (+ − − )
2 2~ 2
159
CHAPTER 3 – MANUSCRIPT
We can prove that these operators satisfy the angular momentum relations
[ ± ] = ±~±
= + + −
we can prove
1
2 = 2 + (+ − + − + )
2
~2 ~2
= (+ − − )2 + (†+ − †− + + †− + †+ − )
4 2
~2 2
= [ + −2 − 2+ − + 2+ (1 + − ) + 2− (1 + + )]
4 +
= ( + 1)~2
2 2
Therefore, we have
+ + − + + −
2 |+ − i = ~2 ( + 1) |+ − i
2 2
~
|+ − i = (+ − − ) |+ − i
2
160
CHAPTER 3 – MANUSCRIPT
These expression can be reduced to the familiar forms for angular momentum provided
that
+ =⇒ +
− =⇒ −
So we can check
| i = ~ | i
We define the spin 1/2 operators ± and in terms of the creation and annihilation
↑ + ↓ = 2 = 1
161
CHAPTER 3 – MANUSCRIPT
+ = ~†↑ ↓
− = ~†↓ ↑
~ +
= ( ↑ − +
↓ ↓ )
2 ↑
Since [↑ †↑ ] = [↓ †↓ ] = 1, we usually call and † the boson operators. In the case
of = 12 2 = († )2 = 0 So they are called hard-core bosons since no two bosons
n o
↓ †↓ = 1
n o
↓ †↑ = 0
162
CHAPTER 3 – MANUSCRIPT
S = S1 + S2
S = S1 + S2
We first examine the two kets: |1+ 2+i and |1− 2−i.
163
CHAPTER 3 – MANUSCRIPT
We denote
Similarly,
Recall that
1 1
| = 1 = 0i = S− | = 1 = 1i = (†1↑ †2↓ + †1↓ †2↑ ) |0i
212 ~ 212
We have obtained three eigenkets for 2 and . There should exist another eigenket
since there are four possible configurations for two spins. The fourth eigenket must be
It is easy to check
|i = 0
h = 1 = 0 | i = 0
1
= − =⇒ = − =
212
164
CHAPTER 3 – MANUSCRIPT
Furthermore,
1
S2 |i = 0 =⇒ |i = | = 0 = 0i = (†1↑ †2↓ − †1↓ †2↑ ) |0i
212
1
| = 0 = 0i = (†1↑ †2↓ − †1↓ †2↑ ) |0i
212
Schwinger’s scheme can be used to derive in a very simple way, a closed formula for
rotation matrices. We apply the rotation operator D() to |i. In the Euler angle
rotation the only non-trivial rotation in the second one about y- axis. So
D() |i
[()†+ ()
−1 +
] [()†− ()
−1 −
]
= D() |0i
[( + )!]12 [( − )!]12
1 1
= (+ − − ) = (†+ − − †− + )
2 2
We have
|0i = 0
165
CHAPTER 3 – MANUSCRIPT
and
We come to calculate
± () = D()†± D−1 () = − ~ †± + ~
1
~
+ () = − − ~ [ +
+ ]
~
1 − † 1
= ~ − ~ = − ()
2 2
Similarly
1
− () = − + ()
2
2 1 2
2 ± () + ( ) ± () = 0
2
We have
± () = cos †± ± sin †∓
2 2
166
CHAPTER 3 – MANUSCRIPT
⎛ ⎞ ⎛ ⎞⎛ ⎞
⎜ †+⎟ −1 ⎜ cos sin ⎟ ⎜
2 ⎟
2
†+
D() ⎜
⎝
⎟ D () = ⎜
⎠ ⎝
⎟⎜
⎠⎝
⎟
⎠
† †
− − sin 2 cos 2 −
Recalling the binomial theorem,
X
!
( + ) = −
=0
!( − )!
()
X 0 0 12
0 [( + )!( − )!( + )!( − )!]
0 = (−1)−+
( + − )!!( − − 0 )!( − + 0 )
0 0
×(cos )2−2+− (sin )2−+
2 2
()
Notice that 0 () = h0 | D() |i
Gordan Coefficients
One of the important problems in quantum theory is the combination of the angular
momenta associated with two parts of a system (such as the spin and orbital angular
momenta of one electron) to form the angular momentum of the whole system. The
167
CHAPTER 3 – MANUSCRIPT
addition of two vectors 1 and 2 forms a triangle. The third vector = 1 + 2 . The
maximal value of is 1 + 2 and the minimal value of is |1 − 2 |. We start with two
with all components of J2 . The orthonormal eigenkets of J21 and J1 are |1 1 i and
J2 has no effect on them. Similarly, |2 2 i are the eigenkets of J22 and J2 and J1 has
we have
J1 · J2 · J1 ⊗ 2 + J2 ⊗ 1
(1 − ) ⊗ (2 − ) = 1 ⊗ 2 − ·
~ ~ ~
where 1 and 2 are two identity matrices, respectively. So we define the total angular
momentum
J = J1 ⊗ 2 + 1 ⊗ J2 ≡ J1 + J2
168
CHAPTER 3 – MANUSCRIPT
It is noted that
J × J = ~J
Since J2 J21 J22 and J are compatible, we denote their simultaneous eigenkets by
|1 2 i ≡ |i
Recall that there are (21 +1) eigenkets |1 1 i and (22 +1) eigenkets |2 2 i Therefore
1 2
X X
|1 2 i h1 2 | = 1
1 =−1 2 =−2
X
|i = |1 2 i h1 2 | i
1 2
( − 1 − 2 ) h1 2 | i = 0
It is apparent that h1 2 | i is zero unless = 1 + 2. The largest value of m
169
CHAPTER 3 – MANUSCRIPT
|1 + 2 1 + 2 i = |1 2 i
| = 1 + 2 = 1 + 2 i = |1 = 1 2 = 2 i
The next largest value of m is 1 + 2 − 1 and this occurs twice: when (1) 1 = 1 and
state with = 1 + 2 and = 1 + 2 − 1 Since for each value there must be values
1 + 2 − 1 By an extension of this argument we can see that each j value, ranging from
1 + 2 to |1 − 2 | by integer steps, appears just once. Each value is associated with
1 +2
X
(2 + 1) = (21 + 1)(22 + 1)
=|1 −2 |
The Clebsch-Gordan coefficients form a unitary matrix. Furthermore the matrix ele-
170
CHAPTER 3 – MANUSCRIPT
X
h1 2 | i h |01 02 i = 1 01 2 02
P
1 P
2
h |1 2 i h1 2 | 0 0 i = 0 0
1 =−1 2 =−2
Recursion Relations
2 2
X X
+ |i = (1+ + 2+ ) |01 02 i h01 02 | i
01 =−1 02 =−2
P 12
= [1 (1 + 1) − 01 (01 + 1)] |01 + 1 02 i h01 02 | i
P
+ [2 (2 + 1) − 2 (2 + 1)]12 |01 02 + 1i h01 02 | i
171
CHAPTER 3 – MANUSCRIPT
Similarly,
Construction Procedure
can leads to all Clebsch-Gordan coefficients . We start with the sector with the largest
value of m :
(1) = 1 + 2 = 1 + 2
h1 = 1 2 = 2 | = 1 + 2 = 1 + 2 i = 1
(2) = 1 + 2 − 1
(a) 1 = 1 2 = 2 − 1
(b) 1 = 1 − 1 2 = 2
= 1 + 2
= 1 + 2 − 1
| 1 + 2 1 + 2 i = |1 2 i
172
CHAPTER 3 – MANUSCRIPT
= [1 (1 + 1) − 1 (1 − 1)]12 |1 − 1 2 i + [2 (2 + 1) − 2 (2 − 1)]12 |1 2 − 1i
µ ¶12
1
h1 − 1 2 |1 + 2 , 1 + 2 − 1i =
1 + 2
µ ¶12
2
h1 2 − 1|1 + 2 , 1 + 2 − 1i =
1 + 2
h1 2 |1 + 2 − 1 1 + 2 − 1i
These two coefficients are determined by the orthogonality conditions h − 1| − 1 − 1i =
= 0
µ ¶12
2
h1 2 − 1|1 + 2 − 1 1 + 2 − 1i
1 + 2
µ ¶12
1
+ h1 − 1,2 |1 + 2 − 1 1 + 2 − 1i = 0
1 + 2
µ ¶12
1
=
1 + 2
µ ¶12
2
=
1 + 2
173
CHAPTER 3 – MANUSCRIPT
⎧
⎪
⎪
⎨ h1 + 2 − 1 1 + 2 − 1|1 + 2 − 1 1 + 2 − 1i = 1
⎪
⎪
⎩ 2 + 2 = 1
⎧
⎪ ³ ´12
⎪
⎨ h1 2 − 1|1 + 2 − 1 1 + 2 − 1i = + 1
1 +2
⎪ ³ ´12
⎪
⎩ h1 − 1 2 |1 + 2 − 1 1 + 2 − 1i = − 2
1 +2
The convention here is that the first matrix element, which has the form h1 − 1 | i,
(3) = 1 + 2 − 2
(a) 1 = 1 2 = 2 − 2
(b) 1 = 1 − 1 2 = 2 − 1
(c) 1 = 1 − 2 2 = 2
= 1 + 2
= 1 + 2 − 1
= 1 + 2 − 2
h1 2 |1 + 2 i and h1 2 |1 + 2 − 1 i are determined by the recursion re-
lations and the coefficients in (2). The coefficients h1 2 |1 + 2 − 2 i are deter-
174
CHAPTER 3 – MANUSCRIPT
In this construction procedure, the only difficult part is the use of orthogonally,
which becomes progressively more complicated as the rank of the submatrix increases.
However it needs be employed only once for each submatrix, and it is easier to work
out an example with particular numerical values for 1 and 2 than the general case
just considered.
(a) J− |1 + 2 1 + 2 − 1i
∙ ¸ 12
2 (22 − 1)
h1 2 − 2|1 + 2 1 + 2 − 2i =
(1 + 2 )(21 + 22 − 1)
∙ ¸ 12
41 2
h1 − 12 − 1|1 + 2 1 + 2 − 2i =
(1 + 2 )(21 + 22 − 1)
∙ ¸ 12
1 (21 − 1)
h1 − 22 |1 + 2 1 + 2 − 2i =
(1 + 2 )(21 + 22 − 1)
(b) J− |1 + 2 − 1 1 + 2 − 1i
∙ ¸ 12
1 (22 − 1)
h1 2 − 2|1 + 2 − 1 1 + 2 − 2i =
(1 + 2 )(1 + 2 − 1)
1 − 2
h1 − 12 − 1|1 + 2 − 1 1 + 2 − 2i = 1
[(1 + 2 )(1 + 2 − 1)] 2
∙ ¸ 12
2 (21 − 1)
h1 − 22 |1 + 2 − 1 1 + 2 − 2i = −
(1 + 2 )(1 + 2 − 1)
175
CHAPTER 3 – MANUSCRIPT
equality
Anybody who is not shocked by quantum theory has not understand it. – Niels Bohr
another measurement on B.
√
| = 0i = (|+; −i − |−; +i) 2
To sum up:
176
CHAPTER 3 – MANUSCRIPT
Many physicists have felt uncomfortable with the preceding interpretation of spin-
correlation measure. Their feelings reflect in Einstein’s locality principle: "But on one
supposition we should in my oppinion, absolutely hold fast: The real factual situation
of the system S2 is independent of what is done with the system S1 , which is spatially
A particle satisfying this property is said to belong to type (+ −). Noting
that we are not asserting that we can simultaneously measyre S and S to be + and -.
177
CHAPTER 3 – MANUSCRIPT
Particle 1 Particle 2
with equal populations, that is 25% each. Suppose a particular pair belong to type
(+ −) and observer A decides to measure S of particle 1; then he or she neces-
this sense that Einstein’s locality principle is incorportaed in this model: A’s results is
178
CHAPTER 3 – MANUSCRIPT
1 2
(+; +) = sin
2 2
179
CHAPTER 3 – MANUSCRIPT
where the factor 1/2 arises from the probability of initially obtaining S1 · a. From Bell’s
equality,
sin2 ≤ sin2 + sin2
2 2 2
If we take = 2 = 2 = 2, the inequality is then violated for
0 2
Nature has had the last laugh on EPR. Nearly thirty years after the EPR
paper was published, an experimental test was proposed that could be used to check
whether or not the picture of the world which EPR were hoping to force a return to
is valid or not. It turns out that Nature experimentally invalidates that point of view,
A 3.7 PROBLEM
⎛ ⎞
⎜ 0 − ⎟
1. Find the eigenvalues and eigenvectors of = ⎜
⎝
⎟. Suppose an electron
⎠
0
⎛ ⎞
⎜ ⎟
is in the spin state ⎜ ⎟
⎝ ⎠ If is measured, what is the probability of the result
~2?
2. Let n be a unit vector in a direction specified by the polar angles ( ). Show
180
CHAPTER 3 – MANUSCRIPT
1 ¡ ¢
= sin − L+ + L− + cos L
2
in the form of a 2 × 2 matrix when S is the spin operator with = 12 Let the
unit vector n have the polar angles and Show explicitly that your matrix for
(12)( )
³ ´ µ ¶ ³ ´
= exp − 3 exp − 2 exp − 3
2 2 2
⎛ ⎞
⎜ −(+)2 cos 2 −(−)2 sin 2 ⎟
= ⎜
⎝
⎟
⎠
−(−)2 sin 2 (+)2 cos 2
181
CHAPTER 3 – MANUSCRIPT
2 = 12 According to the matrix, write explicitly the eigenstates of the total
6. Use the raising or lowering operator to find the following Clebsch-Gordan coeffi-
cients
¯ À ¯ À ¯ À
¯3 1 ¯1 1 ¯1 1
¯ ¯ ¯
¯ 2 2 = |1 1i ⊗ ¯ 2 − 2 + |1 0i ⊗ ¯ 2 2
182
CN Chapter 4
CT Symmetries in Physics
to study particular aspects of physical systems by themselves. For example, the as-
conclusion that the linear momentum of a closed isolated system does not change as the
system moves. This makes it possible to study separately the motion of the center of
mass and the internal motion of the system. A systematic treatment of the symmetry
properties and the conservation law are useful in solving more complicated problems.
with the displacements of a physical system in space and time, with its rotation and
inversion in space, and with the reversal of the sense of progression of time.
183
CHAPTER 4 – MANUSCRIPT
L L
− =0
̇
L=T −V
where
V = V(1 2 )
1 1
L = ̇2 − 2
2 2
̈ = −
This is the Newtonian equation of motion! When the potential is related to velocity or
generalized velocity, the Lagrangian is constructed in a different way. For instance, the
L = T − + ·
184
CHAPTER 4 – MANUSCRIPT
→ +
i.e.,
L
= 0
=0
L
=
̇
In other words, the canonical momentum is conserved, i.e. the law of conservation.
operation like translation and rotation. It has become customary to call U a symmetry
operator regardless of whether the physical system itself possesses the symmetry to U.
When we say that a system possesses that symmetry to U, which implies that
† =
185
CHAPTER 4 – MANUSCRIPT
operator,
=1−
~
[ ] = 0
= 0
is also an eigenstate of G.
G Proof.
= ( 0 ) |i
= ( 0 ) |i
186
CHAPTER 4 – MANUSCRIPT
B 4.1.3 Degeneracy
Suppose that
[ ] = 0
for the symmetry operator and |i is an energy eigenket with eigenvalue . Then,
G Proof.
Suppose |i and |i represent two different kets. Then these two states with the
Remark 3 The proof is very trivial, but the concept plays a far more important role
in quantum mechanics.
If |i is non-degenerated, |i and |i must represent the same physical state although
In the case, we say that the state |i also possesses the symmetry. If
187
CHAPTER 4 – MANUSCRIPT
we say that the symmetry is spontaneously broken in the state although the Hamil-
tonian possesses the symmetry. The spontaneous symmetry breaking occurs when the
state is degenerated. The spontaneous broken symmetry is one of the most impor-
tant concepts in modern physics, from elementary particle physics to condensed matter
physics.
In 1960 Nambu was the first one to introduce the concept of spontaneous
ductivity. Then he applied the concept to elementary particles, which becomes one of
the corner stones of the physics of elementary particles, the standard model. Nambu
EXP Example 1
1 2
= + ()
2
[() ] = 0
X ()
() | i = 0 | 0 i
0
188
CHAPTER 4 – MANUSCRIPT
= − − + 1
EXP Example 2
Two-spin Problem
= S1 · S2
£ ¤
| i = (S1 + S2 )2 − S21 − S22 | i
2
= [ ( + 1) − 2( + 1)] | i
2
189
CHAPTER 4 – MANUSCRIPT
= −( + 1)
= 2
operator
(S1 +S2 )·
= ~
As
[ S] = 0
we have
† =
(1) 0, = 0
| = 0 = 0i = | = 0 = 0i
190
CHAPTER 4 – MANUSCRIPT
(2) 0, = 2
| = 2 = 0i 6= | = 2 = 0i
spontaneously broken.
C Space translation
P
(ρ) = exp[− · ρ]
~
[P ] = 0
191
CHAPTER 4 – MANUSCRIPT
C Time evolution
† ( ) ( ) = +
( ) = exp[− ]
~
† ( ) ( ) =
C Rotation
0 =
D† ()D() =
[ ] = 0
C Gauge Invariance
Ψ = ~ Ψ
This symmetry is related to the conservation of “charge”, i.e., the number of the par-
ticles. To sovle the Schrodinger equation, we find that there is a degree of freedom to
add a trivial phase factor before the wave function. This originates from the symmetry.
192
CHAPTER 4 – MANUSCRIPT
In this section we consider two symmetry operators that can be considered to be discrete
B 4.2.1 Parity
Figure 4.1:
changes a right-handed (RH) system into a left-handed (LH) system. Let us denote the
Π† Π = 1− Π† = Π−1 = Π
193
CHAPTER 4 – MANUSCRIPT
0 =
where ⎛ ⎞
⎜ −1 0 0 ⎟
⎜ ⎟
⎜ ⎟
=⎜
⎜ 0 −1 0
⎟
⎟
⎜ ⎟
⎝ ⎠
0 0 −1
How does an eigenket of the position operator transform under parity?Assume
̂ |i = |i
̂ (Π |i) = − (Π |i)
Π2 ̂ = Π (−̂Π) = ̂Π2
and
194
CHAPTER 4 – MANUSCRIPT
Thus
Π2 = 1
= |− − i
= T (−) |−i
= T (−)Π |i
Since
T () = 1 − +
~
it follows that
Π = − Π
or
{Π } = 0
195
CHAPTER 4 – MANUSCRIPT
Define
L = ×
then
ΠL = LΠ
where ⎛ ⎞
⎜ −1 0 0 ⎟
⎜ ⎟
⎜ ⎟
() =⎜
⎜ 0 −1 0
⎟
⎟
⎜ ⎟
⎝ ⎠
0 0 −1
Π() = ()Π
() = 1 − · } + · · ·
196
CHAPTER 4 – MANUSCRIPT
Π = Π
Π = Π
Π = Π
B 4.2.4 Parity
Let us now look at the parity property of wave function whose state ket is |i :
Ψ() = h|i
Then
Ψ(−) = Ψ()
EXP Example 3
197
CHAPTER 4 – MANUSCRIPT
~2 2 ()
− = ()
2 2
~2 2 (−)
− = (−)
2 2
Thus if () is one solution, (−) must be also one solution. The
solutions are
() = ±·}
Thus the wave functions are sin ~ and cos ~ respec-
tively.
EXP Example 4
2 1
= + 2 2
2 2
We have
Π† Π =
198
CHAPTER 4 – MANUSCRIPT
and ⎫ ⎧
⎪ ⎪
Π† xΠ = −x ⎪
⎬ ⎪
⎨ Π† † Π = −†
⇒
⎪ ⎪
Π pΠ = −p ⎪
† ⎭ ⎪
⎩ Π† Π = −
1 ¡ † ¢
|i = |0i
!
So
h|0i = h−|0i
EXP Example 5
199
CHAPTER 4 – MANUSCRIPT
Polar vector: vectors that are odd under parity such as x, p etc
|i = |i
then (a) if |i is nondegenerate, Π |i is also a parity eigenket; (b) if |i is degenerate,
1
(1 ± Π) |i
2
1 1¡ ¢
Π (1 ± Π) |i = Π ± Π2 |i
2 2
1
= ± (1 ± Π) |i
2
1
where the identity Π2 = 1 has been used. Furthermore the states 2
(1 ± Π) |i and |i
must represent the same state otherwise there would be two states with the same energy.
200
CHAPTER 4 – MANUSCRIPT
FIGURE (a) Periodic potential in one-dimension with periodicity a. (b) The periodic
potential when the barrier height between two adjacent lattice sites becomes infinite.
( ± ) = ()
† () () = + ;
† () () =
As is unitary, we have
=
201
CHAPTER 4 – MANUSCRIPT
is unitary, but not hermitian. † () 6= () We expect the eigenvalue to be a complex
number of modulus 1.
to see a special case of the periodic potential when the barrier height between the two
adjacent lattice sites is made to go to infinite as shown in Figure. Denote |i a state
ket in which the particle is localized at the site. When is applied to it
() |i = | + 1i
X
+∞
|i = |i
=−∞
X
+∞ X
+∞
() |i = () |i = | + 1i
=−∞ =−∞
X
+∞
−
= (+1) | + 1i
=−∞
Assume that the barrier potential is still very large, but finite. The overlap of two states
at nearest neighbor sites is very small, and all other overlaps can be ignored. That is
202
CHAPTER 4 – MANUSCRIPT
h| |i = 0
h| | ± 1i = −
|i = 0 |i − | + 1i − | − 1i
X
= {0 |i h| − | + 1i h| − | − 1i h|}
X
|i = {0 |i h| − | + 1i h| − | − 1i h|} |i
X
|i = |i
=0
X
= | + i
=0
X
= − (+) | + i
=0
203
CHAPTER 4 – MANUSCRIPT
otherwise they could not be identical in all respects. In classical mechanics the existence
distinguish between particles that are identical except for their path since each particle
can be followed during the course of an experiment. In quantum mechanics, the finite
size and the spreading of the wave packets that can describe individual particles often
Two identical particles are indistinguishable if the wave functions for the two particles
C Permutation Symmetry
1 2 1 2
= 1 + + (|1 − 2 |) + (1 ) + (2 )
2 2 2
204
CHAPTER 4 – MANUSCRIPT
is invariant
⇐⇒
†
12 (1 2 )12 = (2 1 )
Clearly
12 = 21
2
12 = 1
|i1 ⊗ |0 i2
1
[(|i1 ⊗ |0 i2 ± |0 i1 ⊗ |i2 )]
2
particles
205
CHAPTER 4 – MANUSCRIPT
Clearly
2 = 1
just as before, and the allowed eigenvalues of are ±1 It is important to note that,
in general
[ ] 6= 0
Although the allowed values of are ±1, system containing N identical
particles are either totally symmetrical or antisymmetric under the interchange of any
pair.
Boson:
Fermions
206
CHAPTER 4 – MANUSCRIPT
An immediate consequence of the electron being a fermion is that the electrons must
satisfy the Pauli exclusion principle, which states that no two electrons can occupy the
same state.
The product of N one-particle eigen functions are the energy eigen wave function of
total system..
= ( + + · · · + )
207
CHAPTER 4 – MANUSCRIPT
(1 2 · · · )
⎛ ⎞
⎜ (1) (2) · · · () ⎟
⎜ ⎟
⎜ ⎟
⎜ (1) (2) · · · () ⎟
⎜ ⎟
= det ⎜
⎜ .
⎟
⎟
⎜ .. .. ... .. ⎟
⎜ . . ⎟
⎜ ⎟
⎝ ⎠
(1) (2) · · · ()
This function is antisymmetric and satisfies the Schrodinger equation with energy
+ + · · · +
(1 2 · · · ) = 0
|0i = 0
† + † = 1
208
CHAPTER 4 – MANUSCRIPT
for fermions.
† − † = 1
with
EXP Example 6
Two-particle problem
Hamiltonian
= 1 + 2
where
1 2 1
1 = + 1 21
21 1 2
1 2 1
2 = 2 + 2 22
22 2
(1) 1 6= 2 , 1 6= 2
(2) 1 = 2 , 1 = 2
209
CHAPTER 4 – MANUSCRIPT
In this section we study a discrete symmetry, called time reversal. The terminology
Let us first look at the classical case: a motion of particle subjected to a certain force.
2 r
= −∇ ()
2
If r() is the solution of the equation, then r(−) is also the solution of the equation.
of motion keeps unchanged. Of course we should notice the change of the boundary
F = (E + v × B)
210
CHAPTER 4 – MANUSCRIPT
v → −v
j → −j
→
B → −B
E → E
1
∇·D =
0
∇·B = 0
D
∇×H− =
B
∇×E+ = 0
Ψ( ) ~2 2
~ = (− ∇ + )Ψ( )
2
Ψ( −) is not a solution of the equation because of the appearance of the first order
time derivative and the imaginary sign at the left hand side. However, Ψ∗ ( −) is a
solution.
211
CHAPTER 4 – MANUSCRIPT
¯ E
¯
|i → |̃i = |i |i → ¯̃ = |i
is said to be antiunitary if
D E
̃|̃ = h|i∗ ;
=
|i → Θ |i
where Θ |i is the time reversed state. More appropriately, Θ |i should be called the
a possible phase. Θ is an anti-unitary operator. We can see this properties from the
212
CHAPTER 4 – MANUSCRIPT
~ Ψ( ) = Ψ( )
−1
Θ~ Θ ΘΨ( ) = ΘΘ−1 ΘΨ( )
−1
Θ~ Θ = ~ ; ΘΘ−1 =
provided that ΘΘ−1 = − and Θ Θ−1 = (−)
ΘpΘ−1 = −p
ΘxΘ−1 = x
ΘJΘ−1 = −J (= x × p)
£ ¤
Θ x p Θ−1 = ~
213
CHAPTER 4 – MANUSCRIPT
Θ =
Θ−1 = −Θ = −
Θ | +i = − ~ − ~ Θ |+i = − ~ − ~ |−i
= | −i
~
|+i = |+i
2
~
Θ Θ−1 Θ |+i = Θ |+i
2
~
− Θ |+i = Θ |+i
2
Θ |+i = |−i
214
CHAPTER 4 – MANUSCRIPT
| −i = − ~ − (+)~ |+i = − ~ − ~ − ~ |+i
Θ = − ~ =
Θ |+i = + |−i
Θ |−i = − |+i
Θ2 = −1
In general
Θ = −
Θ2 = (−1)2
Kramers degeneracy: the ground state for the odd number of electrons in a
215
CHAPTER 4 – MANUSCRIPT
2
2 = −2 +
216
CN Chapter 5
States
As in the case of classical mechanics, there are relatively few physically interesting prob-
lems in the quantum mechanics which can be solved exactly. Approximation methods
are therefore very important in nearly all the applications of the theory. In this chapter
217
CHAPTER 5 – MANUSCRIPT
The variation method can be used for the approximate determination of the lowest or
ground-state energy level of a system when there is no closely related problem that is
capable of exact solution, so that the perturbation method is inapplicable. The most
famous achievements of the variational method in modern physics are the BCS theory
for superconductivity and Laughlin’s theory for fractional quantum Hall effect.
The third law of thermodynamics: "As a system approaches absolute zero, all
processes cease and the entropy of the system approaches a minimum value."
trial wave function. To this end we first prove a theorem of a great of importance. We
Theorem 5 The expectation value of a trial state ket is always not less than the ground
state energy
̄ ≥ 0
218
CHAPTER 5 – MANUSCRIPT
G Proof. We can always imagine that any trial state can be expanded as
¯ ® X ®
¯0̃ = |i |0̃
The variation method does not tell us what kind of trial state ket are to
estimate the ground state energy. Quite often we must appeal to physical intuition. In
practice, we characterize the trial state ket by one or more parameters can compute
The result of the variation method is an upper limit for the ground state energy of the
system, which is likely to be close if the form of the trial state ket or the trial wave
function resembles that of the eigenfunction. Thus it is important to make use of any
trial function.
219
CHAPTER 5 – MANUSCRIPT
1 ³ ´
h|0i = cos
12 2
µ 2 ¶µ 2 ¶
~
0 =
2 42
But suppose we did not know these. Evidently the wave function must vanish at
= ± Further the ground state should have even parity. The simplest analytic wave
h|0i = 2 − 2
10
= 0 ≈ 101320
2
A much better result can be obtained if we use a more general trial function
h|0i = − ||
220
CHAPTER 5 – MANUSCRIPT
( + 1) (2 + 1) ~2
̃ =
2 − 1 42
This gives
5 + 2612
̃ = 0 ≈ 1002980
2
h|0i = ( − )
We use the variation method with a simple trial function to obtain an upper limit for
the energy of the ground state of the helium atom. The helium atom consists of a
nucleus of charge +2 surrounded by two electrons, we find that its Hamiltonian is
~2 ¡ 2 ¢
= − ∇1 + ∇22
2
µ ¶
2 1 1 2
−2 + +
1 2 12
= 1 + 2 + 12
where r1 and r2 are two position vectors of two electrons with respect to the nucleus
2
as origin, and 12 = |r1 − r2 | If we neglect the interaction energy 12
between the two
221
CHAPTER 5 – MANUSCRIPT
electrons, the ground state wave function of H would be product of two normalized
Why is the wave function symmetric in the above equation? Spin singlet in
Why = 2?
We shall use Eq.(5.1) as the trial wave function and permit Z to be the vari-
∞ µ ¶
1 1 X 1
= (cos ) 2 1
12 2 =0 2
222
CHAPTER 5 – MANUSCRIPT
where is the spherical harmonics and is the angle of these two vectors.
2 2 42 52
̃ = − +
0 0 80
µ ¶
2 2 27
= −
0 8
"µ ¶2 µ ¶2 #
2 27 27
= − −
0 16 16
= 2716 ≈ 169
and
µ ¶2
27 2 2
̃ = − = −285
16 0 0
The experimental value for a helium atom is 29042 0 so our estimation is about
1.9% higher.
ate Case
analyses the modifications of discrete energy levels and the corresponding eigenfunctions
223
CHAPTER 5 – MANUSCRIPT
= 0 +
The perturbation theory is based on the assumption that when = 0, both the exact
¯ ®
eigenvalues and the exact energy eigenkets ¯(0) of 0 are know
(0)
¯ ® ¯ ®
0 ¯(0) = (0) ¯(0)
We are required to find approximate eigenkets and eigenvalues for the full Hamiltonian
problem
In general, is not the full potential operator. We expect that the is not very
“large” and should not change the eigenvalues and eigenkets of 0 very much.
unperturbed problem and the case of = 1 corresponds to the full strength problem
we want to solve.
224
CHAPTER 5 – MANUSCRIPT
Before we embark on a systematic presentation of the basic method, let us see how the
expansion in might indeed be valid in the exactly soluble two-state problem we have
encountered many times already. Suppose we have a Hamiltonian that can be written
as
= 0 +
¯ ® ¯ ¯ ® ¯
= 1 ¯1(0) 1(0) ¯ + 2 ¯2(0) 2(0) ¯
(0) (0)
¯ ® ¯ ¯ ® ¯
+12 ¯1(0) 2(0) ¯ + 21 ¯2(0) 1(0) ¯
¯ ®
where and ¯(0) are the exact eigenvalues and eigenstates for 0 . The index “0”
(0)
stands for the unperturbed problem. We can also express the Hamiltonian by a square
matrix ⎛ ⎞
(0)
⎜ 1 12 ⎟
=⎜
⎝
⎟
⎠
(0)
21 2
What is the eigenvalues of ?
⎡Ã ! ⎤12
(0) (0) (0) (0) 2
1 + 2 1 − 2
1 = +⎣ + 2 12 21 ⎦ ;
2 2
⎡Ã ! ⎤12
(0) (0) (0) (0) 2
1 + 2 1 − 2
2 = −⎣ + 2 12 21 ⎦
2 2
So this problem can be solved exactly. Let us suppose is small compared with
the relevant energy scale, the difference of the energy eigenvalues of the unperturbed
225
CHAPTER 5 – MANUSCRIPT
problem:
¯ ¯
¯ (0) (0) ¯
|12 | ¿ ¯1 − 2 ¯
As the expansion in Eq.(5.2) holds for 1, the expansion is available in the case of
small .
We are dealing with the perturbation of a bound state. The assumption that is small
compared with the energy scale suggests that we expand the perturbed eigenvalues and
the parameter such that the zero, first, etc., powers of correspond to the zero, first,
226
CHAPTER 5 – MANUSCRIPT
¯ ® ¯ ® ¯ ®
|i = ¯(0) + ¯(1) + 2 ¯(2) + · · · ;
¡¯ ® ¯ ® ¯ ® ¢
(0 + ) ¯(0) + ¯(1) + 2 ¯(2) + · · ·
¡ (0) ¢ ¡¯ ® ¯ ® ¯ ® ¢
= + (1) + 2 (2) + · · · × ¯(0) + ¯(1) + 2 ¯(2) + · · ·
Since the Equation is supposed to be valid for a continuous range of , we can equate
the coefficients of equal powers of on the both sides to obtain a series of equations
¡ ¢¯ ®
0 − (0) ¯(0) = 0 (5.3)
¡ ¢¯ ® ¯ ®
0 − (0) ¯(1) = ((1) − ) ¯(0) (5.4)
¡ ¢¯ ® ¯ ® ¯ ®
0 − (0) ¯(2) = ((1) − ) ¯(1) + (2) ¯(0) (5.5)
¡ ¢¯ ® ¯ ® ¯ ® ¯ ®
0 − (0) ¯(3) = ((1) − ) ¯(2) + (2) ¯(1) + (3) ¯(0) (5.6)
..
.
Zero-order perturbation
227
CHAPTER 5 – MANUSCRIPT
¯ ®
The Eq. (5.3) means that ¯(0) is one of the unperturbed eigenkets, as
(0)
expected. We can obtain it directly by taking = 0. The eigenvalue is Since we
¯ ®
are dealing with the perturbation of a bound state, the state ket ¯(0) is discrete. It
First-order perturbation
Multiplying an eigen bra of the unperturbed Hamiltonian from the left hand
(0) ¯ ¡ ¢¯ ® ¯ ¯ ®
¯ 0 − (0) ¯(1) = (0) ¯ ((1) − ) ¯(0) (5.7)
(0) ¯ ¡ ¢¯ ® ¯ ¯ ®
¯ 0 − (0) ¯(1) = (0) ¯ ((1) − ) ¯(0) (5.8)
¯ ¯ ®
(1) = (0) ¯ ¯(0) =
¯ ®
It is convenient to calculate ¯(1) by expanding it in terms of the unperturbed eigenkets
where the summation runs over all possible eigenkets of 0 . To simplify the notation,
we define
¯ ¯ ®
= (0) ¯ ¯(0)
228
CHAPTER 5 – MANUSCRIPT
¡ ¢ X (1) ¯ (0) ® ¯ ®
0 − (0) ¯ = ((1) − ) ¯(0)
X ¡ (0) ¢¯ ® ¯ ®
(0) ¯ (0)
(1)
− = ((1) − ) ¯(0)
¯
multiplying (0) ¯ from the left hand side, we obtain
(1)
= (0) (0)
−
Second-order perturbation
¯ ¯ ® X
(2) = (0) ¯ ¯(1) = (0) (0)
6= −
¯ ®
Similarly, we expand ¯(2) in terms of the unperturbed eigenkets
obtain
(0) ¯ ¡ ¢¯ ® ¯ ¯ ® ¯ ¯ ®
¯ 0 − (0) ¯(2) = (0) ¯ ((1) − ) ¯(1) + (0) ¯ (2) ¯(0)
That is
1 © (0) ¯ (1) ¯ ® ¯ ¯ ®ª
(2)
= (0) (0)
¯ ( − ) ¯(1) + (0) ¯ (2) ¯(0)
−
Summary
229
CHAPTER 5 – MANUSCRIPT
X
= (0) + + 2 (0) (0)
+ ···
6= −
¯ ®
|i = ¯(0)
X ¯ (0) ®
+ ¯
(0) (0)
6= −
XX ln ¯ (0) ®
+2 ¯
(0) (0) (0) (0)
6= 6= − −
X ¯ (0) ®
−2 ¯
(0) (0) (0) (0)
6= − −
+···
Remark 6 The state ket |i is not normalized. A normalized state ket should be
1
|i
h|i
To illustrate the perturbation method we developed in the last lecture, let us look
at several examples. All examples can be solved analytically. We can compare the
230
CHAPTER 5 – MANUSCRIPT
The first two examples concern a simple harmonic oscillator whose unperturbed Hamil-
where
³ ´12 ³ ´
† = −
2~
³ ´12 ³ ´
= +
2~
EXP Example 1
0 = 2 2 2
= 0 + 0
2 1
= + (1 + ) 2 2
2 2
231
CHAPTER 5 – MANUSCRIPT
1
1 = h0| 0 |0i = h0| 2 2 |0i = ~
2 4
EXP Example 2
0 =
2 1
= 0 + 0 = + 2 2 +
2 2
2 1 1
= + 2 ( − 0 )2 − 2 20
2 2 2
232
CHAPTER 5 – MANUSCRIPT
where
0 = −
2
~2 2 1 2 2 1
=− 2
+ − 2 20
2 2 2
1 = h0| 0 |0i = 0
X |h| 0 |0i|2
2 =
6=0
0 −
|h1| 0 |0i|2
=
0 − 1
2
= −
2 2
~2 2 2
(− ∇ − )Φ = Φ
2
233
CHAPTER 5 – MANUSCRIPT
Φ = () ( )
where ( ) is a spherical harmonic. (See Schiff, P.76 or Sakurai, P. 454). It is the
2 4 2 2
= − = −
2~2 2 20 2
234
CHAPTER 5 – MANUSCRIPT
Degeneracy
= 1 and = −1 0 +1
EXP Example 3
~2 2 2
0 = − ∇ −
2
235
CHAPTER 5 – MANUSCRIPT
= −
where
µ ¶12
1
00 ( ) =
4
(1)
1 = h0| 0 |0i
Z
= r |00 ( )10 ()|2 ()
Z Z 2 Z ∞
= sin 2 |00 ( )10 ()|2 ()
0 0 0
Z ∞ µ ¶
2 4 −0
= 3 − = −
0 0 0
(0) 2
1 = −
20
236
CHAPTER 5 – MANUSCRIPT
Case
Until now we have assumed that the perturbed state differs slightly from an unperturbed
bound state. The perturbation method for a nondegenerate case we developed fails
237
CHAPTER 5 – MANUSCRIPT
(0) (0)
we consider the special case of 1 = 2 , that is, the two states are energy degenerate.
In fact, this problem can be solved exactly; the energy eigenvalues are,
(0) 1
1 = 1 + (12 21 ) 2 ; (5.9)
(0) 1
2 = 1 − (12 21 ) 2 (5.10)
However we cannot get this result by using the nondegenerate perturbation theory.
difficulty because
of the nondegenerate case to accommodate such a situation. One way to avoid the
catastrophe is to choose our base kets in such a way that has no off-diagonal matrix
¯ (0) ® ¯ ® ¯ ®
¯̃ = ¯1(0) + ¯2(0)
238
CHAPTER 5 – MANUSCRIPT
| |2 + | |2 = 1
choose the perturbation potential is diagonal when the unperturbed state is degenerate.
We see from the two-state example that the degeneracy can be removed by choosing a
new set of base kets. Here we present a more detailed approach. Suppose now that there
239
CHAPTER 5 – MANUSCRIPT
are two states, |i and |i, that have the same unperturbed energy. Then we cannot
= 0. We first consider the case in which 6= 0 the initial state is not specified
by its unperturbed energy; the state may be |i or |i or any linear combination of
them. Let us suppose that the perturbation removes the degeneracy in some order,
so that for finite there are two states that have different energy. We start from the
¡ ¢¯ ®
0 − (0) ¯(0) = 0 (5.11)
¡ ¢¯ ® ¯ ®
0 − (0) ¯(1) = ((1) − ) ¯(0) (5.12)
Out of infinite number of combinations of the two states we choose the particular pair
¯ ® ¯ ®
|i = ¯(0) + ¯(0) ;
(0) =
(0)
= (0)
Substitute the combined state into the Eq.(5.12) and take the inner product of this
¯ ® ¯ ®
equation successively with ¯(0) and ¯(0) , we obtain
¡ ¢
− (1) + = 0
¡ ¢
− (1) + = 0
240
CHAPTER 5 – MANUSCRIPT
1 1£ ¤12
(1) = ( + ) ± ( − )2 + 4 | |2
2 2
=
= 0
In this case we say the degeneracy is not removed in the first order perturbation and
we have to consider the second order perturbation. On the other hand if either or both
(1)
of two equations are not satisfied, the two valued of are distinct, and each can be
used to calculate and . In this case, we can use the nondegenerate perturbation
(1)
theory to determine the higher order perturbation. The coefficients are determined
by
(1) (0) (0)
( − ) = − − ln
(1) (1)
for 6= if we assume that = = 0
241
CHAPTER 5 – MANUSCRIPT
(3) Identify the roots of secular equation with the first-order energy shifts; the
base kets that diagonalize the matrix are the correct zero-order kets to which the
(4) For higher orders use the formulas of the corresponding nondegenerate
perturbation theory except in the summations, where we exclude all contributions from
The change in the energy levels of an atom caused by a uniform external magnetic
field is called the Zeeman effect. We now consider the change of first order in the field
1
A= B×r
2
since B = 5×A and the gauge is ∇·A = 0 The energy related to the vector potential
is
1
= (p − A)2
2
~ 2
4 = A · 5+ A2
2
2
= − B·L+ 2 2 sin2
2 8 2
242
CHAPTER 5 – MANUSCRIPT
=− B · L
2
The energy eigenfunctions of the unperturbed hydrogen atom are usually chosen to
be eigenstates of with the eigenvalues ~ We choose the magnetic field is in the
(0)
1 = h| |i = ~
2
− · = − S·B
= − B · (L + 2S)
2
243
CHAPTER 5 – MANUSCRIPT
1
E = − ∇ ()
netic field,
v
= − ×E
The interaction energy related to the electron spin and the effective magnetic field
= − · B
µ ¶
S p 1
= · × ∇ ()
−
µ ¶
S p 1 r ()
= · ×
−
1 1 ()
= L·S
2 2
Consider the relativistic quantum correction, the spin-orbit coupling is only half of the
above term.
The change in the energy levels of an atom caused by a uniform external electric field
of strength E is called the Stark effect. The perturbation is now the extra energy of
244
CHAPTER 5 – MANUSCRIPT
the nucleus and electron in the external field and is ready shown to be
= cos
where the polar axis and E are in the direction of positive and is again a positive
quantity. In a hydrogen atom the Coulomb potential is rotationally invariant. The wave
function for any spherically symmetric ponetial energy, when expressed in spherical
harmonics, have even or odd parity according as the azimuthal quantum number is
even or odd. Since the perturbation is odd parity with respect to inversion, the ground
state has even parity and has no first-order Stark effect. The first excited state ( = 2)
of hydrogen is four fold degenerate; the quantum numbers and have the value
(0 0), (1 −1), (1 0), (1 1). Since the perturbation commutes with the z-component
of angular momentum,
[ ] = 0
we obtain
245
CHAPTER 5 – MANUSCRIPT
Z
= r∗210 (r) cos 200 ( r)
Z +1 Z ∞
2 −0
= cos cos (2 − )
1640 −1 0 0
= −30
so that half of the four fold degeneracy is removed in the first order.
246
CHAPTER 5 – MANUSCRIPT
imation
Shrodinger wave equation that shows its connection with the Bohr-Sommerfeld quan-
When the Hamiltonian depends on the time, there are no stationary solution of the
Schrödinger equation. Thus our identification of a bound state with a discrete energy
= 0 + ()
247
CHAPTER 5 – MANUSCRIPT
where 0 does not contain time explicitly. The problem of 0 or () = 0 is assumed
0 |i = |i
X
|i = (0) |i
In the system of 0 , the state ket at a later time will evolve into
X
|()i = (0)− ~ |i
Our question is how the state ket of (not 0 ) changes as time goes on.
~ |()i = (0 + ()) |()i (5.13)
X
|()i = ()− ~ |i (5.14)
248
CHAPTER 5 – MANUSCRIPT
X∙ ¸ X ∙
− ~
¸
− ~
= ~ () |i + () ~ |i
X∙ ¸ X
= ~ () − ~ |i + () − ~ |i
" #
X
= (0 + ()) ()− ~ |i
X X
= () − ~ |i + () ()− ~ |i
we obtain
X∙ ¸ X
~ () − ~ |i = () ()− ~ |i
X X
~ = h| |i ~ = h| |i (5.15)
−
=
~
Interaction picture
Two kinds of changes have been made in going from Eq. (5.13). First, we
have changed the representation from specified in terms of the coordinates to being
249
CHAPTER 5 – MANUSCRIPT
~ | 0 i = () | 0 i
~ = [ 0 ]
Exact soluble problem of time-dependent potential are rather rare. However, a two
~ 1 () = 4 2 () (5.16)
~ 2 () = −4 1 () (5.17)
where 4 = + 1 − 2
250
CHAPTER 5 – MANUSCRIPT
2 () = (4 )−1 ~ 1 ()
2 − ∆ = 2 ~2 ;
1³ p ´
= ∆ ± ∆2 + 2 ~2
2
1 = + 2 (4+ ) + − 2 (4− )
~ ~
2 = − (4 + )+ − 2 (4− ) − (4 − )− − 2 (4+ )
2 2
251
CHAPTER 5 – MANUSCRIPT
then the probability for being found in each of the two states is given by
µ ¶
2 2 1
|2 ()| = 2 2 2
sin2
+ ( + 12 ) ~ 4 2
The two-state problem has many physical applications: Spin magnetic reso-
nance, maser etc. Consider a spin 1/2 system subjected to a t-independent uniform
magnetic field in the z-direction and, in addition, a t-dependent magnetic field rotating
in the xy plane.
We treat the effect of the uniform t-independent field as 0 and the effect of
~0
0 = − (|+i h+| − |−i h−|)
2
~1
() = − [cos (|+i h−| + |−i h+|) + sin (− |+i h−| + |−i h+|)]
2
~1 £ − ¤
= − |+i h−| + + |−i h+|
2
252
CHAPTER 5 – MANUSCRIPT
Figure 5.1:
Figure 5.2:
253
CHAPTER 5 – MANUSCRIPT
Four Nobel Prize winners who took advantage of resonance in two-level sys-
tems
• Bloch and Purcell (1952): on B field in atomic nuclei and nuclear magnetic mo-
ments;
• Townes, Basov, and Prochorov (1964): on masers, lasers, and quantum optics;
We now return to Eq. (5.3), replace by , and express the as the power series
in
2 (2)
= (0) (1)
+ + + · · ·
(0)
= 0
(+1) X ()
~ = h| |i + ~
254
CHAPTER 5 – MANUSCRIPT
The zero-order coefficient are constant in time. We shall assume that all except one of
(0)
the , say = are zero, so that the system is initially in a definite unperturbed
energy state. Thus the solutions to the zero order coefficients are
(0)
= h|i =
(1)
The integral constant has been chosen in such a way that vanish at = 0 To first
→ , (that is, the probability that the system, initially in the state , be found at
¯ ¯2
= ¯(1)
¯ 6=
It is also worth noting that for 0 the coefficient of the state is given to first
Z
−1
≈ 1 + (~) h| () |i
−∞
∙ Z ¸
≈ exp − h| () |i
~ −∞
So the principal effect of the perturbation on the initial state is to change the phase.
255
CHAPTER 5 – MANUSCRIPT
The result takes particular simple form if the perturbation is independent of time,
We then obtain
(1)
= − h| |i
~
1 − ~
(1)
= h| |i
sin( 2~)
= −2 2~ h| |i
Hence
Z
− ~ −1 − ~
|i = (1 + (~) h| () |i ) |i
−∞
This is in agreement with the result using the stationary perturbation theory. The first
256
CHAPTER 5 – MANUSCRIPT
the probability is
1
(1)
= h| |i2 2
~2
Eq. (5.18) takes a particularly simple form if the perturbation depends harmonically
on the time except for being turned on at one time and off at a latter time. We shall
call these two times 0 and 0 , respectively, and assume that we can write
where h| |i is time independent. The first order amplitude at any time at or after
0
µ ¶
h| |i ( +)0 − 1 ( −)0 − 1
(1)
=− −
~ + −
The structure of Eq. (??) suggests that the amplitude is appreciable only when the
denominator of one or the other of the two terms is particularly zero. The first term
the present, we specialize to a situation in which the initial state is a discrete bound
257
CHAPTER 5 – MANUSCRIPT
Figure 5.3:
state and the final state is one of a continuous set of dissociated states. The first order
probability of finding the system in the state |i after the perturbation is removed is
¯ (1) ¯ 2 2
¯ ( 0 )¯2 = 4 |h| |i| × sin [( − ) 0 2] (5.19)
~2 ( − )2
The factor sin2 [( − ) 0 2] ( − )2 is plotted in Figure. The height of the
main peak increases in proportional to 20 . Thus if there is a group of states |i that
have energy nearly equal to + ~, and for which |h| |i| is roughly independent
to 0 .
258
CHAPTER 5 – MANUSCRIPT
X ¯ ¯2 Z
¯ ¯2
(1)
= ¯ (1) ¯
⇒ ( ) ¯(1)
¯
≈
where () is the number of final states with energies between and + .
the concept of an energy density () of final state is sensible, since we are considering
the case in which the transition is to one or another of a continuous set of dissociated
states.
Z Z
¯ ¯2 ( − ) | |2
( ) ¯(1)
¯ =4 ( ) sin 2
2~ | − |2
Follow from
1 sin2
lim = ()
→+∞ 2
( − )
sin2 2~
→ ( − )
| − |2 2~
2 ®
= | |2 ( ≈ )
~
2
= () |h| |i|2
~
This formula, first obtained by P. A. Dirac, was later dubbed by E. Fermi “The Golden
259
CN Chapter 6
CT Collision Theory
Problems for which the energy eigenvalues are continuously distributed usually arise in
connection with the collision of a particle with a force field. In a collision problem the
energy is specified in advance, the behavior of the wave function at great distance is
found in terms of it. This asymptotic behaviors can then be related to the amount of
scattering of the particle by the force. There are so few systems of physical interest for
which exact solutions can be found that approximation methods play an important part
in applications of the theory. Various methods that are useful in scattering problems
260
CHAPTER 6 – MANUSCRIPT
Figure 6.1:
We consider first the one-dimensional collision of a particle with the square potential
barrier. In this problem we are interested in a particle that approaches from the region
classical problem, the particle is always reflected if its energy is less than that of the
top of the barrier, and it is always transmitted if the energy is great. We shall see that,
in the quantum problem, both reflection and transmission occur with finite probability
Asymptotic behaviors
We are interested in representing a particle that approaches from the left with
energy 0 and may be turned back by the potential barrier or penetrate through
261
CHAPTER 6 – MANUSCRIPT
it. Thus the asymptotic behavior in the region where () = 0 is as follows: for 0
we want the wave function to represent a particle moving to the left (reflected particle)
as well as to the right (incident particle): for , we want the wave function to
~2 2
− =
2 2
where
µ ¶12
2
= ~ =
~2
Normalization
~
= [Ψ∗ ∇Ψ − (∇Ψ∗ ) Ψ]
2
~ ¡ 2 ¢
() = || − ||2 0
~
() = ||2
262
CHAPTER 6 – MANUSCRIPT
The two currents are independent of the position and should be equal due to the
Scattering coefficients
The character of the solution inside the potential barrier depends on whether
C The case 0
where
µ ¶12
2 ( − 0 )
=
~2
+ = +
( − ) = ( − )
At = ,
¡ ¢
+ − =
263
CHAPTER 6 – MANUSCRIPT
We can eliminate and and solve for the ratios and
(2 − 2 ) (1 − 2 )
=
( + )2 − ( − )2 2
¡ ¢
4 1 − (−)
=
( + )2 − ( − )2 2
C The case 0
264
CHAPTER 6 – MANUSCRIPT
particles per unit time and count the number of incident particles that emerge per unit
time in a small solid angle 4 0 (?) centered about a direction that has polar angles
0 and 0 with respect to the bombarding direction as polar axis. This number will
265
CHAPTER 6 – MANUSCRIPT
be proportional to , and 4 0 provided that the flux is small enough so that there
bombarded particles by their recoil out of the target region, and provided also that the
bombarded particles are far enough apart so that each collision process involves only
one of them.
The different scattering cross section 0 (0 0 ): the number of incident parti-
0 (0 0 )4 0
For a collision of a particle with a fixed scattering center, the definition of the differential
scattering cross section is equally valid in the laboratory and center-of-mass coordinate
systems, since a scattering center that is fixed has an infinite effective mass and so
that the center of mass of the system does not move. For a collision between two
particles of finite mass, however, the differential cross section applies in general only
to the laboratory coordinate system and to the observation of the scattered incident
particle. It does not describe in the observation of the recoil bombarded particle in the
266
CHAPTER 6 – MANUSCRIPT
Figure 6.3:
267
CHAPTER 6 – MANUSCRIPT
tems
The relation between the differential cross section and angles in the laboratory
system and in the center-of-mass system can be found by translating the laboratory
system in the direction of the incident particle with sufficient speed to bring the center
of mass to rest. Assume that a particle of mass 1 and initial speed strikes a particle
of mass 2 that is initially at rest. The center of mass moves with the speed
1
0 =
1 + 2
00
Thus in the center-of-mass system the speeds of two particles are 00 = − 0 and − .
If the collision is elastic, they evidently recede from the center of mass after the collision
= 0
sin
tan 0 =
+ cos
0 1
= 00
=
2
268
CHAPTER 6 – MANUSCRIPT
The relation between the cross section in the laboratory and center-of-mass
coordinate systems can be obtained from the definitions, which imply that the same
number of particles are scattered into differential solid angle 0 about 0 0 as are
12
(1 + 2 + 2 cos )
0 (0 0 ) = ( )
|1 + cos |
It should be noted that the total cross section is the same for both laboratory and
center-of-mass systems and also for both the outgoing particles, since the total number
of collision that takes place is independent of the mode of description of the process.
1
Dependence on = 2
For = 1,
sin
tan 0 = = tan
+ cos 2
Thus 0 = 2 and varies from 0 to 2 as varies from 0 to ; in this case
269
CHAPTER 6 – MANUSCRIPT
increases further to . In this case 0 (0 0 ) is usually infinite at the maximum value
of 0 , although this singularity gives a finite contribution to the total cross section; no
In this section, we assume that the potential is a function only of , and we find
the connection between the solution separated in spherical polar coordinate and the
asymptotic form.
Asymptotic behaviors
1
= [ + ( ) ] (6.1)
where = ~ and is the speed of the incident particle. The probability current
270
CHAPTER 6 – MANUSCRIPT
density J is given by
~ £ † ¡ ¢ ¤
J() = Ψ (∇Ψ) − ∇Ψ† Ψ
2
1 1
∇= r̂+ θ̂ + φ̂
sin
Since the second and third terms are small when is large, the current for large is in
The number of scattered particles entering the solid angle (or detector) per unit time
is
= 2
~
= | ( )|2
scattering cross section but cannot itself be found without solving the wave equation
271
CHAPTER 6 – MANUSCRIPT
The radial and angular parts of the solution can be separated by taking
where
is the spherical harmonic. The problem now possesses symmetry about the polar axis,
so that , and are independent of the angle . That is, we just consider the
case of = 0. In the case the =0 is reduced to the Legendre polynomials. Thus,
the general solution can then be written as a sum of products of radial function and
Legendre polynomials,
X
∞
= (2 + 1) () (cos )
=0
X
∞
= (2 + 1) −1 () (cos )
=0
272
CHAPTER 6 – MANUSCRIPT
where
µ ¶12
2
=
~2
2 ()
() = → 0 as → +∞
~2
µ ¶
1 2 1 2
=
2 2
∝ ± → sin( + 0 )
In some specialized problems, () can be neglected for greater than some distance
, and may be small enough so that the term in not negligible. The general form
−→ sin( − + ) → ∞ (6.2)
2
1 +1
→ cos( − ) → ∞
2
→ →0
(2 + 1)!!
273
CHAPTER 6 – MANUSCRIPT
1 +1
→ sin( − ) → ∞
2
(2 − 1)!!
→ − → 0
+1
X
∞
cos = (2 + 1) () (cos )
=0
X
∞
= (2 + 1) () (cos ) + −1 ( )
=0
X∞
= (2 + 1) ()−1 sin( − + ) (cos )
=0
2
are the coefficient for different , and should take the form
=
Thus it gives
1 X
∞
() = (2 + 1)(2 − 1) (cos )
2 =0
274
CHAPTER 6 – MANUSCRIPT
() = | ()|2
¯∞ ¯2
1 ¯X ¯
¯ 2 ¯
= 2 ¯ (2 + 1)( − 1) (cos )¯
¯ ¯
=0
Z
= 2 () sin
0
4 X
∞
= (2 + 1) sin2
2 =0
Z +1
2
cos (cos )0 (cos ) = 0
−1 2 + 1
The total cross section can also be related to (0). it follows from the generating
function for the Legendre polynomials that (1) = 1 for = 0 and all , so that
1 X
∞
¡ ¢
(0) = (2 + 1) 2 − 1
2 =0
2 4
= [ (0) − ∗ (0)] = Im (0)
275
CHAPTER 6 – MANUSCRIPT
from the incident beam, so that its intensity is smaller behind the scattering region
( ≈ 0) than in front of it. This can occur only by the interference between two terms
Phase shifts
The angle is called the phase shift of the l partial wave, since according to
Eq.(6.2) it is the difference in phase between the asymptotic forms of the actual radial
function () and the radial function () in the absence of the scattering potential.
the phase shifts completely determine the scattering, and the scattering cross section
A 6.4 Applications
In general, for a given potential, phase shifts are calculated from a numerical solution
for scattering by an attractive square well, for which the reduced potential () is
⎧
⎪
⎪
⎨ −0 ( 0) ;
() =
⎪
⎪
⎩ 0
276
CHAPTER 6 – MANUSCRIPT
At low energies, the scattering is dominated by the = 0 wave. Using the formula
sin
0 () =
cos
0 () = −
277
CHAPTER 6 – MANUSCRIPT
we have
tan − tan
tan 0 () =
+ tan tan
4 2
=0 = sin 0
2
µ ¶2
2 tan
≈ 4 1 −
Since the scattered particle cannot penetrate into the region , the wave function in
the exterior region must vanish at = . Since the scattered particle cannot penetrate
into the region the wave function in the exterior region, given by Eq.(6.3), must
278
CHAPTER 6 – MANUSCRIPT
()2+1
= −
(2 + 1)!!(2 − 1)!!
Hence |tan ()| quickly decreases as increases. As a result, the low-energy scattering
is always dominated by the s-wave. Since 0 () = sin and 0 () = − cos , we
have
0 = −
0 = 42
= −
2
so that
4 X
max
2
≈ 2
(2 + 1) sin2 ( − )
=0 2
≈ 22
It is interesting to compare this result with that obtained from classical mechanics.
279
CHAPTER 6 – MANUSCRIPT
of the fundamental differences between classical and quantum mechanics. In the polar
() = · + ()
− () = −· + ( − )
Symmetric or anti-symmetric:
Classical case
280
CHAPTER 6 – MANUSCRIPT
For spin 1/2 fermions, the wave functions contains two parts, spin and spatial
part. For spins, two spins can form one spin singlet = 0 (antisymmetric), and one
spin triplet = 1 (symmetric). The total wave function should be antisymmetric. Thus
one part of the wave function is symmetric and another part must be antisymmetric.
1
( ) = |( ) + ( − + )|2
4
3
+ | ( ) − ( − + )|2
4
field
1
= (p − A)2 + ()
2
1 2
= p + () − A·p
2
281
CHAPTER 6 – MANUSCRIPT
where the term A2 is omitted and the symmetric gauge is taken, ∇ · A = 0 For a
A = 20 ̂ cos( n · x − )
where ̂ · n = 0 We write
A = 0 ̂[exp( n · x − ) + exp(− n · x + )
= − A·p
0 £ ¤
= − ε̂ · p n·x− + − n·x+
where the first term is responsible for absorption, and the second term is for stimulated
emission.
0 n·x−
† = − ( ε̂ · p)
2 2 20 ¯¯
n·x−
¯2
¯ ( − − ~)
→ = h| ε̂ · p |i
~ 2 2
282
CHAPTER 6 – MANUSCRIPT
= (E × H)
4
1 2
with E = − 1 A and B = ∇ × A 2
|0 |2
~→
= 1 2
2
|0 |2
4 2 ~ 2 ¯¯ n·x−
¯2
¯ ( − − ~)
= ( ) h| ε̂ · p |i
2 ~
2 2
~ ∼ '
0
This leads to
~
=∼ 2
= 137
Then ∼ 137 ¿ 1 for light atams. In this case, it becomes a good approxi-
mations
n·x = 1 + n · x + · · ·
h| n·x− ε̂ · p |i → ε̂ · h| p |i
~
[ 0 ] =
283
CHAPTER 6 – MANUSCRIPT
h| p |i = h| [x 0 ] |i =
~
h| n·x− ε̂ · p |i → ε̂ · h| x |i
The photoelectric effectis the ejection of an electro when the atom is placed in the
radiation field. The basic process is considered to be the transition from an atomic
bound state to a continuum state ( 0). Therefore, |i the ket for an atomic state,
say, the ground state of hydrogen atom, while |i is the ket for a continuum state say
The basic task in this calculation is the to calculate the number of the
final state near the energy and + , or the density of state
() = lim
→0
Consider a cubic box of side L3 . The plane wave has the form,
k ·x
hx|k i =
32
where
= 2
284
CHAPTER 6 – MANUSCRIPT
The energy
~2 2 ~2 (2)2 2
= = 2
( + 2 + 2 )
2 2
= 2 Ω
= 2 Ω
() = 2
Ω
µ ¶3 µ ¶3
= 2
= (2 ~2 )12
2 ~ 2 ~2
The initial state is taken to be a ground state of hydrogen-like wave function except
Z " µ ¶32 #
−k ·x n·x−
= ε̂ · x 32 (−~∇) −0
0
In this integral, notice that ε̂ · ∇ n·x− = 0
Z " µ ¶32 #
n·x− −~∇−k ·x n·x− −0
hk | ε̂ · p |i = −ε̂ · x
32 0
Z " µ ¶32 #
~ (ε̂ · k ) −k ·x n·x− −0
= x
32 0
285
CHAPTER 6 – MANUSCRIPT
2 (ε̂ · )2 5 1
= 32
Ω 0 [(0 )2 + 2 ]4
5
where q = k − ()n
Let us begin with the time-independent formulation of scattering processes. The Hamil-
tonian is written as
2
= 0 + = +
2
0 stands for the kninetic energy. In the collision theory, the source of incident particles
and the detection location of the scattered particles are very far away from the scattering
center. 0 can describe the behaviors of incident and reflected particle very well as the
potential approaches to zero when the distance is much larger than the interaction
0 |i = |i
286
CHAPTER 6 – MANUSCRIPT
When → 0 |Ψi → |i We assume the two equations have the same energy eigen-
becomes
¯ (+) ® 1 ¯ ®
¯Ψ = |i + ¯Ψ(+) (6.4)
− 0 +
Z
¯ ® 1 ¯ ®
hx ¯Ψ(+) = hx |i + x0 hx| |x0 i hx0 | ¯Ψ(+)
− 0 +
Z
2 ¯ ®
= hx |i + 2 x0 + (x x0 ) hx0 | ¯Ψ(+)
~
287
CHAPTER 6 – MANUSCRIPT
0 ~2 1
+ (x x ) ≡ hx| |x0 i
2 − 0 +
Z
~2 1
= p hx| pi hp| |pi hp |x0 i
2 − 0 +
Z
~2 1
= p hx| pi hp| 2
|pi hp |x0 i
2 − 2 +
Z
~2 1
= p hx| pi 2
hp |x0 i
2 − 2 +
Z
~2 p exp[p · (x − x0 )~]
=
2 (2~)3 − 2 2 +
Z
0 q exp[q · (x − x0 )]
+ (x x ) =
(2)3 2 − 2 +
Z ∞ 2 Z 2 Z
exp[ |x − x0 | cos ]
= sin
0 (2)3 0 0 2 − 2 +
Z +∞
exp[ |x − x0 |] − exp[− |x − x0 |]
=
4 2 |x − x0 | 0 2 − 2 +
1 exp[ | − 0 |]
= −
4 | − 0 |
288
CHAPTER 6 – MANUSCRIPT
Figure 6.4:
I
= 2
− 0
I X
() = 2 −1 = 2 × (sum of enclosed residues).
Z +|x−x0 |
¯ ® 2 0
¯ ®
hx ¯Ψ(+) = hx |i − 2 hx0 | ¯Ψ(+)
~ |x − x0 |
Z +|x−x0 |
¯ ® 2 0
¯ ®
hx ¯Ψ(+) = hx |i − 2 (0 ) hx0 ¯Ψ(+)
~ |x − x0 |
289
CHAPTER 6 – MANUSCRIPT
|x| = À |x0 | = 0
0
= (1 − cos + )
0
+|x−x | −k0 ·r0
≈
|x − x0 |
So,
Z
¯ ® 2 ¯ ®
hx ¯Ψ(+) → hx |i − 2 x0−k ·r (0 ) hx0 ¯Ψ(+)
0 0
~
We have
Z
0 4 2 exp[−k0 · x0 ] ®
(k k) = − 2 x0 (x0
) hx0
| Ψ(+)
~ (2)32
4 2 0 ¯¯ (+) ®
= − 2 hk | Ψ
~
2
= |(k0 k)|
Ω
290
CHAPTER 6 – MANUSCRIPT
The formula contains the unknown wave function. We have to introduce the approxi-
mation.
¯ (+) ® 1 ¯ ®
¯Ψ = |i + ¯Ψ(+)
− 0 +
µ ¶
1 1 ¯ (+) ®
= |i + |i + ¯
Ψ
− 0 + − 0 +
1
= |i + |i +
− 0 +
®
Assume the effect of the scattering is not very strong, we replace hx0 | Ψ(+) by hx0 | i
in the integral,
®
hx0 | Ψ(+) → hx0 | i = exp[−k · x0 ]
Z
0 4 2 1 0 0
( ) = − 2 x0 (k−k )•x (x0 )
~ (2)3
Z Z 2 Z
1 2 2 2
= − sin cos ()
4 ~2 0 0 0
Z +∞
2
= − 2 () sin
~ 0
291
CHAPTER 6 – MANUSCRIPT
Define
¯ ®
¯Ψ(+) = |i
we have
¯ ® 1
|i = ¯Ψ(+) = |i + |i
− 0 +
1
= +
− 0 +
1
= +
− 0 +
1 1
+ + ···
− 0 + − 0 +
1 2
(k0 k) = − (2)3 hk| |k0 i
4 ~2
G Proof.
1 2
( = 0) = ( ) = − (2)2 hk| |ki
4 ~2
292
CHAPTER 6 – MANUSCRIPT
®
Im hk| |ki = Im k| |Ψ(+)
∙µ ¶ ¸
(+) ¯ (+) ¯ 1 ¯ (+) ®
= Im Ψ ¯ − Ψ ¯ ¯Ψ
− 0 −
£ ¯ ¯ ®¤
Im hk| |ki = Im − Ψ(+) ¯ ( − 0 ) ¯Ψ(+)
¯ ¯ ®
= − Ψ(+) ¯ ( − 0 ) ¯Ψ(+)
Z
¯ ¯ ®
= − k0 Ψ(+) ¯ |k0 i hk0 | ( − 0 ) |k0 i hk0 | ¯Ψ(+)
Z
¯ ¯ ® ~2 02
= − k0 Ψ(+) ¯ |k0 i hk0 | ¯Ψ(+) ( − )
2
Z
¯ ¯ ® ~2 02
= − 02 0 Ω0 Ψ(+) ¯ |k0 i hk0 | ¯Ψ(+) ( − )
2
Z
¯ ¯ ® 2 1
= − Ω0 Ψ(+) ¯ |k0 i hk0 | ¯Ψ(+) 02 2 0
~ 2
p
At the last step we take 0 = 2~2 .
Z
0 ¯ ¯ ¯2
Im hk| |ki = − 2 Ω0 ¯ Ψ(+) ¯ |k0 i¯
~
293
CHAPTER 6 – MANUSCRIPT
tial
−
() = 0
Z
0 sin 1
(k k) = − () = −0 2
+ 2
where
= |k0 − k|
Z Z +∞
0
() sin = − sin
0
Z +∞
0 − −
= −
0 2
µ ¶
0 1 1
= −
2 − +
0
=
2 + 2
We obtain
2 µ ¶2
(2 0 2 ) 1 0 2 1
() = =
~4
16 sin4
4
2
16 sin4
2
294
CHAPTER 6 – MANUSCRIPT
This is the Rutherford scattering cross section. From Rutherford formula we conclude
that the charge-charge interaction in atomic scale obey Coulomb law, i.e., 1. Actually
this formula does not contain the Planck constant, and can be drived from the classical
theory.
295
CN Chapter 7
CT Selected Topics
I shall briefly introduce the Dirac equation. Because of extensive interests in spintronics
and graphene, I shall focus on the spin-orbit coupling (electric manipulation of electron
~ Ψ = Ψ
296
CHAPTER 7 – MANUSCRIPT
2
where = 2
+.
→ → ~
→ −~∇
[ ] = ~
+ ∇ · = 0
where
= Ψ∗ Ψ
~ ∗
= − (Ψ ∇Ψ − Ψ∇Ψ∗ ) = Re(Ψ∗ Ψ)
2
tells us that
2 = 2 4 + 2 2
Ψ = exp[( · − )~]
297
CHAPTER 7 – MANUSCRIPT
p
= ± 2 4 + 2 2
+ ∇ · = 0
where
~
= [Ψ∗ Ψ − Ψ Ψ∗ ]
22
~ ∗
= − (Ψ ∇Ψ − Ψ∇Ψ∗ )
2
Difficulties:
1. Negative energy
2. Negative probability
( − ())2 Ψ = (2 4 − 2 ~2 ∇2 )Ψ
= · + 2
= 2 2 + 2 4
298
CHAPTER 7 – MANUSCRIPT
As a result,
+ = 2
+ = 0
2 = 1
In 2D, we have
=
=
=
In 3D, Dirac realized that and should be at least 4×4 matrices. One representation
is
⎛ ⎞
⎜ 0 ⎟
= ⎜
⎝
⎟ = ⊗
⎠
0
⎛ ⎞
⎜ 0 ⎟
= ⎜
⎝
⎟ = ⊗ (= 0 ⊗ )
⎠
0 −
299
CHAPTER 7 – MANUSCRIPT
= Ψ† Ψ
= Ψ† Ψ
300
CHAPTER 7 – MANUSCRIPT
The solution is
p p
= 2 2 + 2 4 = ± 2 2 + 2 4
·
0 =
+ 2 0
and ⎛ ⎞
⎜ 0 ⎟
+ = + ⎜
⎝
⎟
⎠
√2 ·
0
2 +2 4 +2
⎛ ⎞
√2 ·
⎜ 0 ⎟
− = − ⎜
⎝
− 2 +2 4 −2 ⎟
⎠
0
The rest energy of electron 2 = 051099906 MeV = 051 × 106 The electron energy
relativistic limit
⎛ ⎞
⎜ 0 ⎟
+ = + ⎜
⎝
⎟
⎠
0
301
CHAPTER 7 – MANUSCRIPT
⎛ ⎞
⎜ 0 ⎟
− = − ⎜
⎝
⎟
⎠
0
L=r×p
such that
[ ] = 0
302
CHAPTER 7 – MANUSCRIPT
[ · ( − ) + 2 + ]() = ()
⎛ ⎞
⎜ () ⎟
Denote () = ⎜
⎝
⎟
⎠
()
( − − 2 ) = · ( − )
( − + 2 ) = · ( − )
1
( − − 2 ) = · ( − ) 2
· ( − )
− +
1
( − ) = · ( − ) · ( − )
− + 22
1 h i2
≈ · ( − )
2
303
CHAPTER 7 – MANUSCRIPT
X
( · A) ( · B) =
X
= ( + )
= A · B + · (A × B)
h i2 ~
· ( − ) = ( − )2 + · (∇ × )
2
= " #12
2 2
1+ h √ i2
+12+ (+12)2 − 2 2
304
CHAPTER 7 – MANUSCRIPT
1
( − ) = · ( − ) 2
· ( − )
− + 2
1
( − ) = · ( − ) 2
· ( − )
2 −
1 −
( − ) = 2
· ( − )(1 − 2
) · ( − )
2 2
1 h i2
= · ( − )
2
1 h i h i
+ 2 2 · ( − ) ( − ) · ( − )
4
h i2 ~
· ( − ) = ( − )2 + · (∇ × )
h i h i
· ( − ) · ( − ) = ~ · (∇ × ) + · · ·
Final results
∙ ¸
1 2 ~
= ( − ) + + ·
2 2
~
+ · (∇ × p) + · · · · · ·
42 2
305
CHAPTER 7 – MANUSCRIPT
~
= · (∇ × p)
42 2
~ 1 1 1
= 2 2
· (r × p) = S·L
4 22 2
1
= [ ]
~
1
= [ ]
~
306
CHAPTER 7 – MANUSCRIPT
Figure 7.1:
307
CHAPTER 7 – MANUSCRIPT
Figure 7.2:
308
CHAPTER 7 – MANUSCRIPT
Figure 7.3:
309
CHAPTER 7 – MANUSCRIPT
Figure 7.4:
310
CHAPTER 7 – MANUSCRIPT
Figure 7.5:
311
CHAPTER 7 – MANUSCRIPT
the most intriguing systems in solid state physics today. The most popular description
of graphene band structure is the tight binding one, first done by Wallace. The band
structure exhibits very unique features: the first two bands do not have a gap between
them, and do not overlap either. In fact they intersect in two inequivalent points,
called Dirac points in the first BZ. resulting in four degenerate modes. The vicinity
of the dirac points is not parabolic (as in most crystals) but in fact conical. Hence,
the group velocity ≡ ∇k E() is independent of the energy. The Fermi level for
undoped graphene lies exactly at the intersection points, and graphene is a gapless
one can write a relativistic dynamic equation for the excitations. Such equation can be
derived from the tight binding model, and the resulting equations are infact the well
√
= 3
√
1 = ( 32 +12)
√
2 = ( 32 −12)
312
CHAPTER 7 – MANUSCRIPT
Figure 7.6:
√
Unit cell area is: = 2 sin(3) = 32.
triangular lattices A and B, with basis vectors a for lattice A and vectors δ with
= 1 2 3 connecting A to B. (b) the green hexagon is a Brillouin zone (BZ) and pink
diamond is the extended BZ for the honeycomb lattice. The reciprocal lattice vectors
are b .
Reciprocal lattice:
̄2 × ̂ √
1 = 2 = 2(1 3 1)
̄1 × ̂ √
2 = 2 = 2(1 3 −1)
313
CHAPTER 7 – MANUSCRIPT
X
0 = − †n n+δ + †n+δ n (7.1)
nδ
the hopping parameter, n ,n+δ are annihilation operators corresponding A,B sub-
lattices, and suppressing the spin degree of freedom which is irrelevant for optics. The
gain some more insight on the excitations in Fourier space. Expanding the operators
1 X k·a
a = (k) (7.3)
k
1 X k·(a+δ )
a+δ = (k) (7.4)
k
X 1 X X X (k−q)·n −q·δ
†n+δ n = 2
(q)† (k)
nδ
k q nδ
X X X
= (k)† (k) −k·δ ≡ (k)† (k)(k) (7.5)
k δ k
Performing the summation and plugging the results into (7.1) we obtain
X ¡ ¢
0 = Φ† (k)H0 Φ(k) Φ† (k) = (k)† (k)† (7.6)
k
314
CHAPTER 7 – MANUSCRIPT
⎛ ⎞
∗
⎜ 0 (k) ⎟
H0 (k) = ⎜
⎝
⎟
⎠ (7.7)
(k) 0
where
∙ √ √
µ ¶¸
3 − (2 3)
(k) = − + 2 cos (7.8)
2
√
∙ √
µ ¶¸
3 − 2 3
= − 1 + 2 cos (7.9)
2
Note that diagonalizing H0 , also diagonalize 0 . The energy bands are ob-
One can show that these points are connected by reciprocal lattice vectors, and only
two are inequivalent, denoted by K+ , K− . One can find the points by requiring = 0.
√
For this − 32
must be real, so = 0. Then = 23.
√ √
Alternatively, − 3
= −1, = ±2 3 and = ±23.
√
= − 32( ± ) (7.11)
315
CHAPTER 7 – MANUSCRIPT
One can cast the spectrum in the vicinity of the K± points in the manifestly
relativistic way. For this let us turn the coordinate axes by 90◦ so that → 0 and
√
= − 32(−0 ± 0 ) (7.13)
In the graphene, the effective velocity is about c/300 where c is the speed of light
→ − → ~ −
→ − → −~
For (+ + )
⎛ ⎞
⎜ 0 ~ −
+ ~ ⎟
() = ⎜
⎝
⎟
⎠
~ − − ~ 0
316
CHAPTER 7 – MANUSCRIPT
The communtator
[~ − + ~ ~ − − ~ ] = [− −~ ] + [~ − ]
~
= −2 = 2~2
2
0
= √ (~ − + ~ )
2~
† = √ (~ − − ~ )
2~
⎛ ⎞ ⎛ ⎞
√ ~ ⎜ 0 ⎟ ⎜ 0 ⎟
() = 2 ⎜ ⎟ = 0 ⎜ ⎟
⎝ † ⎠ ⎝ ⎠
0 † 0
0 =
0 † =
(0 )2 † = 2
† |i = |i
317
CHAPTER 7 – MANUSCRIPT
= h|i
√
= 0
0 0 √
= h||i = h| − 1i = h| − 1i
For = 0, =0 = 0.
Certain insulators have exotic metallic states surrounding arround the boundary or
surfaces. These states are formed by the strong spin-orbital coupling as a result of
The electrons in the surface states are well described by an effective Hamiltonoan of
= p · = ( + )
This system is different from graphene: there are double spin degeneracy and double
valley degeneracy of Dirac cone in graphene. The surface states in topological insulator
menatlly. Several families of topological insulator have been discovered. The materials
318
CHAPTER 7 – MANUSCRIPT
Figure 7.7:
may provide new route to generate novel particles and quantum states and to explore
Quantum statistical mechanics is the branch of physics dealing with systems in mixed
that statistics enters at two levels in quantum statistical mechanics: first, because of
the statistical interpretation of the wave function and second, because of our incomplete
319
CHAPTER 7 – MANUSCRIPT
for which, say = 60 are characterized by |i and the remaining = 40 are
which characterizes a state ket whose spin is pointing some definite direction. We
should not confuse the probability weight of the states |+i and |−i , ± and |± |2
Mixed ensembles:
¯ ®
1 ¯(1)
¯ ®
2 ¯(2)
¯ ®
3 ¯(3)
.. ..
. .
X
= 1
320
CHAPTER 7 – MANUSCRIPT
X
hi = h || i
X X
= h |i h| | i
X
= |h| i|2
Alternatively,
X
hi = h| i h |i h || i
X X
= h| { | i h | } |i
= ()
where
X
= | i h |
Two properties:
(a). is Hermitian;
() = 1
= | i h |
321
CHAPTER 7 – MANUSCRIPT
= | +i h +|
1 1
= (|+i + |−i) (h+| + h−|)
212 212
⎛ ⎞
1 1
⎜ 2 2 ⎟
= ⎜
⎝
⎟
⎠
1 1
2 2
322
CHAPTER 7 – MANUSCRIPT
We define a quantity
= − ( ln )
X
= − ln
=0
= ln
323
CHAPTER 7 – MANUSCRIPT
= ( ln )
To minimize :
= 0
i.e.,
(b). = 1
we obtain
= − −
1
= P −
−
324
CHAPTER 7 – MANUSCRIPT
X
= −
= (− )
1 −
=
= 1
= − ~
= 2 =
1 −
=
⎛ ⎞
1 −
⎜
0 ⎟
= ⎜
⎝
⎟
⎠
1 +
0
325
CHAPTER 7 – MANUSCRIPT
where
= − +
~ − −
h i =
2 − +
~
= − tanh()
2
Consider a system with two identical particles. Each particle has three non-degenerate
states: |1i |2i |3i with E1 , E2 , E3 . For classical identical particles: there 9 possible
configurations
|1 1i |1 2i |1 3i
326
CHAPTER 7 – MANUSCRIPT
objects which are equivalent and possess the same energy levels.
327
CHAPTER 7 – MANUSCRIPT
C Maxwell-Boltzmann Statistics
X
=
where the sum runs over all the eigenenergies of a single particle, and where n is the
X
=
nj can either be zero or any positive integer. If all the energies Ej are different, each
permutation of the particles results in a different wave function, thus each energy level
of the total system is N!-fold degenerate. However if nj1, the interchange between
these particles do not alter the wave function. Thus the number of distinct states
!
= Q
!
328
CHAPTER 7 – MANUSCRIPT
X X
= exp(− )
{ }
X ! X
= Q exp(− )
!
{ }
" #2
X
= exp(− )
1 X X
h i = exp(− )
{ }
ln
= −
−
= = − −
The wave function of many bosons is symmetric and the energy level of the system is
non-degenerate
1 X X X
h i = exp(− − )
{ }
329
CHAPTER 7 – MANUSCRIPT
where
X X X
= exp(− − )
{ }
⎛ ⎞
Y X
= ⎝ exp(− − )⎠
{ }
Y
= (1 − exp(− − ))−1
In the case,
ln
h i = −
1
=
exp( + ) − 1
The wave function of many bosons is antisymmetric, and the energy level of the system
is non-degenerate
= 1 for = 1 or 0
= 0 otherwise.
1 X X X
h i = exp(− − )
{ }
330
CHAPTER 7 – MANUSCRIPT
where
X X X
= exp(− − )
{ }
⎛ ⎞
Y X
1
= ⎝ exp(− − )⎠
=0
Y
= (1 + exp(− − ))
In the case,
ln
h i = −
1
=
exp( + ) + 1
One of the main features of boson system is that bosons tend to occupy the lowest energy
state at low temperatures. At high temperatures, both distribution laws for bosons and
for fermions become equal to that for the Maxwell-Boltzmann case approximately. The
density of particles on a certain state always tends to be zero. For a boson system, as
1
h i = → +∞
exp( + ) − 1
if both E and are equal to zero, the density of the particle at E=0 can be nonzero in
331
CHAPTER 7 – MANUSCRIPT
spectrum of energy is
~2 2
=
2
Z Z
1 3
() = ()()
(2)3
where
µ ¶32
1 2
() = 2 12
4 ~2
where
Z +∞
12
12 () =
0 + − 1
Since ≥ 0 the largest possible values of 12 () occurs when = 0 and numerically
12 ( = 0) = 2315
332
CHAPTER 7 – MANUSCRIPT
the function ( − 0 ) becomes less than unit, and the system increasingly condensed
Nath Bose at 1924 to 25. It was first observed in Helium-4 liquid. Recent years it was
observed in some systems with cooling atoms. It is one of the most active branches in
condensed matter physics. If you are interested in the recent development in this field,
http://www.aip.org/pt/webwatch/ww9703.html.
333
CHAPTER 7 – MANUSCRIPT
If the particle are confined to a cube of edge L and we require the wave function to be
Ψ ( + ) = Ψ ( + )
Ψ ( + ) = Ψ ( + )
Ψ ( + ) = Ψ ( + )
2 4 6
= 0 ± ± ± ···
Any component of k is of the form 2n, where n is a positive integer. The energy
spectrum with n is
~2 ¡ 2 ¢
= + 2 + 2
2
As two indistinguishable fermions cannot have all their quantum numbers identical,
each energy level (or state) can be occupied by at most one particle. Thus we start
filling the energy level from that with the lowest energy (i.e., k=0 here) until all N
The occupied energy levels may be represented as points inside a sphere in the
k-space. The energy at the surface of the sphere is the Fermi energy; the wave vectors
334
CHAPTER 7 – MANUSCRIPT
~2 2
=
2
1 4 3
= 3
(2) 3
Then
¡ ¢13
= 6 2
In short, a boson gas condensates into the lowest energy state, and a fermion
participating in the effect. The integer quantum Hall effect was observed by von Klitzing
et al in 1980. Simple theory suggests that the Hall conductivity at he plateaus should
be an integral multiple of 2 ~ and the experiments agree with that prediction to within
quantum Hall effect promises a method of providing very precise resistance standards
335
CHAPTER 7 – MANUSCRIPT
that are insensitive to the particular sample and the details of its fabrication. With
a more powerful magnetic field and lower temperature, Daniel C. Tsui and Horst L.
Stormer discovered that the plateaus of the conductivity is a fractional multiples of the
basic unit 2 ~ Within a year of their discovery, Rovert B. Laughlin has succeeded in
explaining their result. Through theoretical analysis he showed that the electrons in a
powerful magnetic field can condense to form a kind of quantum fluid related to the
What makes these fluids particularly important for researchers is that events
in a drop of quantum fluid can afford more profound insights into the general inner
structure and dynamics of matter. The Royal Swedish Academy of Sciences awarded
The 1998 Nobel Prize in Physics jointly to Tsui, Stomer and Laughlin “for their dis-
The Hall effect occurs when the charge carriers moving through a material experience a
potential difference across the side of the material which is transverse to the magnetic
336
CHAPTER 7 – MANUSCRIPT
Ohmic resistance is V/I . A magnetic field tin the positive z direction shifts positive
charge carriers in the negative y direction. This generates a Hall potential ( ) in
the y direction.
produced. Charges will accumulate on the transverse edges until a strong enough
condition is
=
337
CHAPTER 7 – MANUSCRIPT
=
where l is the thickness of material and d the width. Hence we obtain an expression
1
= · =
Therefore this effect can be used to determine the density of charge carriers in conduc-
tors and semi-conductors, and has become a standard tool in physics laboratories over
=
= ∝
In 1980 the German physicist Klaus von Klitzing discovered in a similar experiment that
the Hall resistance does not vary in linear fashion, but ”step-wise” with the strength
of the magnetic field. The steps occur at resistance values that do not depend on
the properties of the material but are given by a combination of fundamental physical
Hall resistance values, normal Ohmic resistance disappears and the material becomes
338
CHAPTER 7 – MANUSCRIPT
in a sense superconducting. For his discovery of what is termed the integer quantum
Hall effect von Klitzing was awarded the Nobel Prize in 1985.
In their refined experimental studies of the quantum Hall effect, using among
other things lower temperatures and more powerful magnetic field, Stomer, Tsui and
their co-workers found to their great surprise a new step in the Hall resistance which
was three times higher than von Klitzing’s highest. They subsequently found more and
more new steps, both above and between the integers. All the new step heights can be
expressed with the same constant as earlier but now divided by different fractions. For
this reason the new discovery is named the fractional quantum Hall effect..
339
CHAPTER 7 – MANUSCRIPT
A year after the discovery of the fractional quantum Hall effect, Laughlin offered a
theoretical explanation. According to his theory the low temperature and the powerful
magnetic field compel the electron gas to condense to form a new type of quantum fluid.
Since electrons are most reluctant to condense (They are what is termed fermions) they
first, in a sense, combine with the ”flux quanta” of magnetic field. Particularly for the
first step ( = 13) discovered by Stomer and Tsui, each of electrons captures three
flux quanta thus forming a kind of composite particle with no objection to condensing.
(They become what is termed bosons). Quantum fluids have earlier occurred at very
340
CHAPTER 7 – MANUSCRIPT
low temperatures in liquid helium and in superconductors. They have certain properties
in common, e.g. superfluidity, but they also show important differences in behaviors.
Apart from its superfluidity which explains the disappearance of Ohmic resistance at the
Hall resistance steps„ the new quantum fluid proposed by Laughlin has many unusual
properties. One of the most remarkable is that if one electron is added the fluid will
These quasiparticles are not particles in the normal sense but a result of the common
dance of electrons in the quantum fluid. Laughlin was the first to demonstrate that
the quasiparticles have precisely the correct fractional charge to explain the fractional
quantum Hall effect. Subsequent measurements have demonstrated more and more
fractional charged steps in the Hall effect, and Laughlin’s quantum fluid has proved
capable of explaining all the steps experimentally. The new quantum fluid strongly
Further reading
1. B. Davis, Splitting the electron, New Scientist, 31 Jan, 1998, p36
341
CHAPTER 7 – MANUSCRIPT
uniform magnetic field B. The mass of the particle is M and charge e. The Hamiltonian
is
1
= ( + )2
2
1
= [(−~ + )2 + (−~ + )2 ]
2
= (∇ × A)
1
A=− B×r
2
i.e.
1
= −
2
1
=
2
In the gauge
1
= (−~ − )2
2 2
1
+ (−~ + )2
2 2
342
CHAPTER 7 – MANUSCRIPT
Denote
Π = −~ −
2
Π = −~ +
2
1
= {(Π − Π )(Π + Π ) + [Π Π ]}
2
1 ~
= [(Π − Π )(Π + Π ) +
2
+ 1
= ~ [ + ]
2
where
³ ´12
+ = (Π − Π )
2~
³ ´12
= (−~ + − ~ + )
2~ 2 2
1
= 212 (− + )
̄ 4
1
= 212 (+ + ̄)
4
343
CHAPTER 7 – MANUSCRIPT
1
= ~ (+ + )
2
where
=
in a uniform magnetic field is equivalent to it. A set of the wave function has the form
= −~( − )
= ~( − ̄ )
̄
344
CHAPTER 7 – MANUSCRIPT
= = =
0
~
2 202
To make this degeneracy more apparent, we assume the system has the shape of a
1
+ = 212 ( − )
4
1
= 212 (− − )
4
([ + ] = 1)
[ ] = [+ ] = 0
Also , the operator annihilates the wave function ( ), just like the operator
Define
̄ = exp[20 ]
= exp[20 † ]
Thus, is an engenstate of ̄
̄ =
345
CHAPTER 7 – MANUSCRIPT
A complete set of eigenstates of the Landau level { }can be constructed ( =
1 )
= exp[2 ] ̄
1
(̄) = ~( + ) ( ̄)
2
The degeneracy of the Landau energy levels is −fold. The ratio of the number of
electrons to
=
is called the filling number of Landau level. When = 1it means that the first Landau
level is fully filled and when = 2, the second landau level is also filled. There is a
346
CHAPTER 7 – MANUSCRIPT
simple relation between the filling number and the quantized Hall conductivity. Suppose
=
J =
=
= =
2
= =
· ~
~
The integer quantum Hall effect occurs at = , (1,2,...) and the
thermal equilibrium can display no magnetic momentum, even in a magnetic field. The
347
CHAPTER 7 – MANUSCRIPT
magnetic momentum of a free atom has three principal sources: the spin with which
electrons are endowed, their orbital angular momentum about the nucleus; and the
348
CHAPTER 7 – MANUSCRIPT
align. Antiferromagnetism and spin density waves describe oscillatory ordering of mag-
netic moments. The classical dipolar interaction between the electron moments (which
is of order 105 ) is for too weak to explain the observed magnetic transition temper-
ature (which are of order 102 − 103 0 in transition metal and rare earth compounds)
The coupling mechanism that gives rise to magnetism derives from the fol-
• Coulomb repulsion
electrons in solids, we simply review some standard definitions and basic relations of
second quantization.
®
| =
349
CHAPTER 7 – MANUSCRIPT
The creation operator of state is † and its Hermitian conjugate is annihilation oper-
ator Both are defined with respect to the vacuum state |0i such that
|0i = 0
= †
For bosons,
h i
† = † − † =
{ } = 0
For electrons with spin = 12, we have to introduce a pair of operators, † where
{ } = 0
350
CHAPTER 7 – MANUSCRIPT
~ †
= ( ↑ − †↓ ↓ )
2 ↑
= S1 · S2
(3) [S1 + S2 ] = 0
Therefore S21 S22 and S2 = (S1 + S2 )2 and its z-component S are good
quantum numbers, but 1 2 are not. Hence we can denote the simultaneous eigenkets
351
CHAPTER 7 – MANUSCRIPT
such that
S21 | 1 2 i
= 1 (1 + 1) | 1 2 i
S22 | 1 2 i
= 2 (2 + 1) | 1 2 i
= ( + 1) | 1 2 i
= | 1 2 i
| 1 2 i = | 1 2 i
where
= [ ( + 1) − 1 (1 + 1) − 2 (2 + 1)]
2
352
CHAPTER 7 – MANUSCRIPT
and
Since the energy eigenvalues are independent of can be − the energy
eigenstates are (2 + 1)−fold degenerated. From the point of view of symmetry,
the degeneracy of the eigenstates originates from the invariance of H under the (2)
symmetry rotation,
† =
where
= exp[−S · n~]
The case of 0 :
= 1 − 2 (1 2 )
The two spins are antiparallel, which is called antiferromagnetic. The ground state
energy
= −(1 + 1)2
The case of 0
= 1 + 2
353
CHAPTER 7 – MANUSCRIPT
= − || 1 2
Ferromagnetic exchange coupling originates from the direct Coulomb interaction and
the Pauli exclusion principle. In the second quantized form, the two-body Coulomb
interaction is given by
Z
1
= ̃( )Ψ† ()Ψ†0 ()Ψ0 ()Ψ ()
2
X
= −
X
+ 0 (↑ + ↓ )(0 ↑ + 0 ↓ )
0
X
+ 0 † †0 0 0 0
0
+···
354
CHAPTER 7 – MANUSCRIPT
where
Z
1
0 = | ()|2 |0 ()|2 ̃( )
2
Z
1
0 = ̃( )†0 ()† ()0 () ()
2
X X 1
0 + +
0 0 0 0 = −2 0 ( · 0 + · 0 )
4
= ( − )
In the case,
Z
1
0 = | ()|2 |0 ()|2 0
2
2
̃ =
| − |
2
= 4 0
2
This is ferromagnetic!
355
CHAPTER 7 – MANUSCRIPT
Tunnelling between the two atoms (or states) is described by a hopping Hamiltonian
X
= − +
(1 +
2 + 2 1 )
X
= ↑ · ↓
For : there are six possible configurations which are eigenstates of H
(1) = 0, |1 ↑ 2 ↑i |1 ↑ 2 ↓i |1 ↓ 2 ↑i |1 ↓ 2 ↓i ;
(2) = , |1 ↑ 1 ↓i |2 ↓ 2 ↑i
Denote |i the unperturbed state with energy 0 and |i denote the two state
|i = + +
1 2 |0i
356
CHAPTER 7 – MANUSCRIPT
and
† †
|1i = 1↑ 1↓ |0i
† †
|2i = 2↑ 2↓ |0i
h | | 0 i = 0
¯ ¯ ® P h | | i h | | i
¯∆ (2) ¯ =
=12 h | | i − h | | i
1 X ¯¯ ¯¯ ® ¯¯ ¯¯ ®
= −
357
CHAPTER 7 – MANUSCRIPT
¯ ¯ ® ¯ ¯ 0®
¯ ¯ 1 1 ¯ ¯
† †
= h| (2↑ 1↑ + 2↓ 1↓ + ~)
† †
(2↑ 1↑ + 2↓ 1↓ + ~) |0 i
† †
= h| (2↑ 1↑ + 2↓ 1↓ )1↑ 1↓ (1 − 2↑ )(1 − 2↓ )
† †
×(1↑ 2↑ + 1↓ 2↓ ) |0 i
† † † †
= h| (−1↓ 1↑ 2↑ 1↓ − 1↑ 1↓ 2↓ 2↑ ) + 1↑ 2↓
1
= h| − 21 · 2 + (1↑ + 1↓ )(2↑ + 2↓ ) |0 i
2
42 1
∆ (2) = + (1 · 2 − )
4
two-site problem is spin singlet, i.e. = 0. Our discussion on the two-site problem
358
CHAPTER 7 – MANUSCRIPT
lattice
X
= ·
where and are the lattice sites and usually are of the nearest neighbour pair. can
X 42 X 1
= − † + ( · − )
4
hi
which is limited within the Hilbert space excluding double occupancy of electrons on
359