You are on page 1of 239

2

Contents
1 Preliminaries 9
1.1 Vector Spaces, Deal Spaces, and Scalar Products . . . . . . . . . . . . . . . . . . . . 9
1.1.1 Vector Spaces and kets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.1.2 Dual Spaces and Bras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.1.3 Scalar Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.1.4 Tensor Products of 2 Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . 12
1.2 Linear Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.2.1 Products of Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.2.2 Matrix Representation of Operators . . . . . . . . . . . . . . . . . . . . . . . 13
1.2.3 Hermitian and Unitary Operators . . . . . . . . . . . . . . . . . . . . . . . . 15
1.2.4 Transformations (Unitary) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.2.5 The Eigenvalue Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.2.6 Functions of Operators, Generalizations to Innite Dimensions, and More . . 18
2 Classical Mechanics Review 23
2.1 Lagranges and Hamiltons Formulations of Classical Theory . . . . . . . . . . . . . . 23
2.1.1 Lagrangian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.1.2 Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.1.3 Cyclic Coordinates, Poisson Brackets, and Canonical Transformations . . . . 25
2.1.4 Symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3 Postulates of QM 27
3.1 Postulates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.2 Compatability, Incompatibility and Uncertainty . . . . . . . . . . . . . . . . . . . . . 29
3.2.1 Compatible Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.2.2 Incompatible Variables and Uncertainty Relationships . . . . . . . . . . . . . 30
3.3 Pure States and Mixtures: Density Matrices . . . . . . . . . . . . . . . . . . . . . . . 33
3.3.1 Density Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.3.2 Density Matrix of a Pure State . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.3.3 Density Matrix of Mixed States . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.3.4 Time Evolution of the Density Matrix . . . . . . . . . . . . . . . . . . . . . . 37
3.4 Schrodinger Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.4.1 Solving the Schrodinger Equation . . . . . . . . . . . . . . . . . . . . . . . . . 39
3
4 CONTENTS
4 Simple Problems in 1 Dimension 41
4.1 Free Particle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.1.1 Solving the Schrodinger Equation . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.1.2 Propagator for a Free Particle in the Position Basis . . . . . . . . . . . . . . . 42
4.2 Particle in a Box . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.3 Step Potentials: Reection and Transmission . . . . . . . . . . . . . . . . . . . . . . 46
4.4 Finite Square Well: Discrete Energies and Resonances . . . . . . . . . . . . . . . . . 49
4.5 Square Potential Barrier and Tunneling . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.6 General Properties of 1-D Schrodinger Equation . . . . . . . . . . . . . . . . . . . . 59
4.6.1 Asymptotic Behavior of Solutions . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.6.2 Nature of Energy Eigenvalues . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5 Wave Packets 63
5.1 Plane Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.2 Wave Packets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.2.1 Spacial Extent of Wave Packets . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.2.2 Motion of the Wave Packet . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.2.3 Delay of Wave Packets at Resonance . . . . . . . . . . . . . . . . . . . . . . . 70
6 Harmonic Oscillator and Second Quantization 73
6.1 Harmonic Oscillator in Position Representation . . . . . . . . . . . . . . . . . . . . . 73
6.2 Second Quantization: Creation and Annihilation Operators . . . . . . . . . . . . . . 77
6.2.1 Harmonic Oscillator in Energy Eigenbasis . . . . . . . . . . . . . . . . . . . . 77
6.2.2 Eigenfunctions in the Position Representation . . . . . . . . . . . . . . . . . . 82
6.3 Coherent States: Minimum Uncertainty Wave Packets . . . . . . . . . . . . . . . . . 83
6.3.1 Uncertainty Relations in Energy Eigenstates . . . . . . . . . . . . . . . . . . 83
6.3.2 Minimum Uncertainty States . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
6.3.3 Physical Meaning of the Coherent State . . . . . . . . . . . . . . . . . . . . . 85
7 Systems with N Degrees of Freedom 93
7.1 Tensor Products (Direct Products) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
7.2 Identical Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.2.1 Symmetric and Anti-symmetric States: Bosons and Fermions . . . . . . . . . 95
7.2.2 Distinguishing Fermions and Bosons . . . . . . . . . . . . . . . . . . . . . . . 97
8 Classical Limit and WKB Approximation 99
8.1 Ehrenfests Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
8.2 Classical Limit and Wavepacket Spreading . . . . . . . . . . . . . . . . . . . . . . . . 102
8.3 Classical Limit of S-Eqn and WKB Approximation . . . . . . . . . . . . . . . . . . . 104
8.3.1 Quantum Mechanical Probability Current . . . . . . . . . . . . . . . . . . . . 105
8.3.2 Connect Probability Current in Classical Mechanics to Quantum Mechanics . 106
8.3.3 WKB Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
CONTENTS 5
9 Symmetries 111
9.1 Translations, Translational Invariance, and Cons. of Mom. . . . . . . . . . . . . . . . 111
9.1.1 Active Translations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
9.1.2 Passive Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
9.1.3 Translations For a Many Particle System . . . . . . . . . . . . . . . . . . . . 115
9.2 Time Translational Invariance and Energy Conservation . . . . . . . . . . . . . . . 116
9.3 Discrete Symmetries: Parity and Time Reversal . . . . . . . . . . . . . . . . . . . . 117
9.3.1 Parity Invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
9.3.2 Time-Reversal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
9.4 Rotations, Rotational Invariance, and Cons. of Ang. Mom. . . . . . . . . . . . . . . 120
9.4.1 Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
9.4.1.1 Rotations in 2D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
9.4.1.2 Rotations in 3D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
9.4.2 Eigenvalues of Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . 126
9.4.2.1 Eigenvalues of

L
z
. . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
9.4.2.2 Solutions to Problems with Azimuthal Symmetry . . . . . . . . . . 127
9.4.2.3 Eigenvalues of

L
2
and

L
z
. . . . . . . . . . . . . . . . . . . . . . . . 128
9.4.3 Matrix Representation of Angular Momentum . . . . . . . . . . . . . . . . . 131
9.4.3.1 Matrix Elements of

J
2
and

J
z
. . . . . . . . . . . . . . . . . . . . . 131
9.4.3.2 Matrix Elements of

J
x
and

J
y
. . . . . . . . . . . . . . . . . . . . . 132
9.4.4 Finite Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
10 Central Potential 137
10.1 Hamiltonian in Spherical Coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . 137
10.2 Solution of Spherically Symmetric Schrodinger Equation . . . . . . . . . . . . . . . . 138
10.2.1 Solution of the Angular Part . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
10.2.2 Solution to the Radial Part . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
10.2.3 Examples: Free Particles and Central Square Well Potentials . . . . . . . . . 144
10.2.3.1 Free Particles and Spherical Bessel Functions . . . . . . . . . . . . . 144
10.2.3.2 Central Square Well . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
10.2.3.3 The Coulomb Potential: Hydrogen and H-like Atoms (Ions) . . . . . 149
11 Angular Momentum: Spin 155
11.1 Brief Review of Properties of Angular Momentum Operators . . . . . . . . . . . . . 155
11.2 Spin: Evidence in Atomic Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . 156
11.3 Spin Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
11.3.1 Spin Operators, Spinors, and Eigenvalues . . . . . . . . . . . . . . . . . . . . 158
11.3.2 Pauli Spin Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
11.3.3 Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
11.4 Spin Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
11.4.1 Spin
1
2
and the Bloch Sphere . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
11.4.2 Rotations on the Bloch Sphere . . . . . . . . . . . . . . . . . . . . . . . . . . 166
11.5 Particles with Spin 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
11.5.1 Rotations of a Vector Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
6 CONTENTS
12 Addition of Angular Momentum 173
12.1 Addition of 2 Angular Momenta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
12.1.1 Finding Allowed Values of J and M . . . . . . . . . . . . . . . . . . . . . . . 174
12.1.2 Example of Addition of Two S=
1
2
States . . . . . . . . . . . . . . . . . . . . . 176
12.2 Two Nucleon System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
12.3 Clebsch-Gordan Coecients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
12.4 Irreducable Tensor Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
12.4.1 Irreducable Invariant Subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . 181
12.4.2 Irreducable Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
12.4.3 Wigner-Eckart Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
13 Perturbation Theory 185
13.1 Perturbation Theory for a Non-Degenerate, Discrete, Energy Level . . . . . . . . . . 185
13.1.1 First Order Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
13.1.2 Example: Harmonic Oscillator in an Electric Field . . . . . . . . . . . . . . . 188
13.1.3 Example: Stark Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
13.2 Degenerate Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192
13.2.1 Example: Coupled Harmonic Oscillator . . . . . . . . . . . . . . . . . . . . . 194
13.3 Time Dependent Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . 197
13.3.1 Transition Properties Between Energy States . . . . . . . . . . . . . . . . . . 200
13.3.2 Time-Independent Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . 202
13.3.3 Transition to a Continuum of States . . . . . . . . . . . . . . . . . . . . . . . 204
13.3.4 Periodic Perturbations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
13.3.5 Sudden and Adiabatic Perturbations . . . . . . . . . . . . . . . . . . . . . . . 207
14 Scattering 209
14.1 Scattering Cross Sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
14.2 Waves and Phase Shifts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
14.3 Example: Scattering from a Hard Sphere . . . . . . . . . . . . . . . . . . . . . . . . 215
14.4 Scattering from a Square Well Resonances . . . . . . . . . . . . . . . . . . . . . . . 217
15 Relativistic Quantum Mechanics: The Dirac Equation 221
15.1 Spinless Free Particles - The Klein-Gordan Equation . . . . . . . . . . . . . . . . . . 221
15.2 Dirac Equation for Free Particle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
15.3 Electromagnetic Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
15.3.1 Electron Spin and Magnetic Moment . . . . . . . . . . . . . . . . . . . . . . . 223
15.4 The Dirac Equation for the Hydrogen Atom . . . . . . . . . . . . . . . . . . . . . . . 225
15.4.1 Fine Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
15.5 Negative Energy Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
15.5.1 Dirac Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
15.5.2 Feynmanns Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
CONTENTS 7
16 Feynmann Path Integral 235
16.1 Path Integrals - Concepts and the Classical Action . . . . . . . . . . . . . . . . . . . 235
16.2

U(t) for a free particle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
8 CONTENTS
Chapter 1
Preliminaries
Mathematical Preliminaries
A fundamental principle of QM is the superposition principle of dynamic states: The possible dy-
namical states of a given quantized system may be superposed linearly and therefore may be
represented by vectors of a certain linear space (to be dened). In this space, dynamic variables
are associated with linear operators.
1.1 Vector Spaces, Deal Spaces, and Scalar Products
1.1.1 Vector Spaces and kets
Denition 1: A vector space, or linear space:
V = [1) , [2) , . . . , [a) , [b) , . . .
has the following properties:
1. Is a group under addition
closure ([a) +[b) V )
exists identity: ([a) +[0) = [a) [a))
inverse: ([a) +[a) = [0))
associative: ([a) + ([b) +[c)) = ([a) +[b)) +[c))
2. Commutative under addition
3. We can also dene a multiplication of elements of V with num-
bers , , , . . . so that [a) , [b) V :
[a) V
( [a)) = () [a)
( +) [a) = [a) + [a)
([a) +[b)) = [a) +[b)
1 [a) = [a) [a)
9
10 CHAPTER 1. PRELIMINARIES
Remark 1: [a) , [b) , . . . are vectors
, , . . . are scalars that dene a eld over which the vector space is dened
If , , . . . R, then we have a real vector space
If , , . . . C, then we have a complex vector space
Remark 2: [a) is a ket in Dirac notation
It is an abstraction of the more familiar column vector
Denition 2: The vectors [i) for i = 1 . . . n are linearly independent if:
n

i=1
a
i
[i) = 0 a
i
= 0 i
Denition 3: Let n be the number of linearly independent vectors in V. Then n =
the dimension of the vector space, denoted V
n
, and the linearly independent
vectors form a basis.
1.1.2 Dual Spaces and Bras
Denition 1: The adjoint of a vector [a) in the vector space denes a bra a[, which
is a member of the dual space. The dual space is antilinear.
If [c) = = [a) + [b) , then c[ =

a[ +

b[
Remark 1: bras are abstractions of row vectors
Remark 2: The adjoint operator takes the ket from the vector space to the corre-
sponding bra in dual-space and vice versa
[a)

= a[ and a[

= [a)
Remark 3: If there any type of scalar multipliers, they are replaced by complex
conjugates. Adjoints of complicated expressions are written in reverse order.
Ex: [[a) b[ [c)]

= c[

[b) a[

1.1.3 Scalar Products


Denition 1: For any vectors [a) , [b) V
n
, we denote a[b) to be the inner product,
or scalar product. a[b) C, and has the properties:
a[b) = b[a)

1.1. VECTOR SPACES, DEAL SPACES, AND SCALAR PRODUCTS 11


a[a) 0, = 0 [a) = [0)
Linear in ket and antilinear in bra
Remark 1: The properties of the scalar product imply the Schwartz inequality:
[a[b)[
2
< a[a) b[b)
Postulate: The space of ket vectors (and the dual space of bra vectors) form a
Hilbert Space. It is complete and separable. Here, complete refers to the
existence of a basis in which any vector can be expressed as a superposition of
the basis vectors.
Denition 2: [a) , [b) V
n
are orthogonal i a[b) = 0
Denition 3:
_
a[a) = [a[ is the norm of the vector [a)
Denition 4: The set of basis vectors, [1) , . . . , [n), that have i[j) =
ij
, i, j, is an
orthonormal basis.
Theorem (Gram-Schmidt): An orthonormal basis may be formed from a linear
combination of basis vectors.
Remark 2: We will write vectors as expansions of orthonormal basis so that
[a) =
n

i=1
a
i
[i) and [b) =
n

j=1
b
j
[j)
a[b) =
n

i,j=1
a

i
b
j
i[j)
..

ij
=
n

i
a

i
b
i
Denition 5: A subspace is a subset of the vector space V that is also a vector space.
Denition 6:
1
and
2
are orthogonal subspaces if every vector in
1
is orthogonal
to every vector in
2
.
Denition 7: The vectors orthogonal to
1
form a subspace
x
1
which is the complementary subspace
to
1
.
12 CHAPTER 1. PRELIMINARIES
Theorem: Any vector [a) V may be written
[a) = [a
1
) +[a
x
1
)
where [a
1
)
1
and [a
x
1
)
x
1
and
1
,
x
1
V .
Denition 8: [a
1
) is the projection of [a) on the subspace of
1
.
1.1.4 Tensor Products of 2 Vector Spaces
To represent a system described by products of wave functions in the bra-ket notation,
we use Tensor Products.
Example: Let

a
1
_
represent the state of particle 1. Let

a
2
_
represent the state of
particle 2. Then the 2 particle system is represented by the tensor product:

a
1
_

a
2
_

a
1
_

a
2
_

a
1
a
2
_
Remark 1: An operator

1
(

2
) only operates on

a
1
_
(

a
2
_
), so that, if:

a
1
_
=

b
1
_

a
2
_
=

b
2
_
Then

a
1
a
2
_
=

b
1
a
2
_

a
1
a
2
_
=

a
1
b
2
_
Furthermore, every operator

1
commutes with

2
.
Ex:

a
1
a
2
_
=

a
1
a
2
_
1.2 Linear Operators
1.2.1 Products of Operators
Consider an operator

that transforms one vector [a) into another [a

):

[a) =

_
Denition 1: a linear operator obeys the following:


[a) =

[a)
1.2. LINEAR OPERATORS 13


[[a) + [b)] =

[a) +

[b)
a[

= a[

[a[ +b[ ]

= a[

+ b[

Remark 1: the products of 2 operators are just 2 operations carried out in sequence:
(

) [a) =

(

[a))
In general, the product does not commute. Dene the commutator:

,

]
Denition 2: The inverse of a operator

is denoted

1
, and

1
= I =
identity operator.
Remark 2: The inverse of a product of operators is:
(

)
1
=

1
1.2.2 Matrix Representation of Operators
Consider a vector [a) and operator

such that

[a) = [b)
The matrix form of

has elements
ij
= i[

[j), so that for:


[a) =

i
a
i
[i)
[b) =

j
b
j
[j)
Then
b
j
= j[b) = j[

[a)
= j[

i
a
i
[i)
=

i
a
i
j[

[i)
=

ij
a
i
14 CHAPTER 1. PRELIMINARIES
And, in matrix and vector representation:
_

_
b
1
b
2
.
.
.
b
n
_

_
=
_

_
1[

[1) 1[

[2) 1[

[n)
2[

[1)
.
.
.
.
.
.
.
.
.
n[

[1) n[

[n)
_

_
_

_
a
1
a
2
.
.
.
a
n
_

_
Examples:
The identity operator

I
I
ij
= i[

I [j) = i[j) =
ij
The Projection operator

P: Consider a set of basis vectors
[1) , [2) , . . . , [n). The projection of any vector onto the i
th
basis
vector is obtained by:

P
j
[a) = [i) i[a) = i[a) [i)
So:

P
i
= [i) i[ projector
Remark 1: Because the basis vectors form a complete set (in Hilbert Space), we
have
[a) =

i
a
i
[i) =

i
i[a)
..
a
i
[i)
=

i
[i) i[a)
= (

i
[i) i[
. .

P
i
) [a)
Since this is true for all [a):

I =

i
[i) i[ a completeness relation
Remark 2: Projectors have the property that:

P
2
=

P
P
i
P
j
= [i) i[j) j[ =
ji
[i) j[ =
ij

P
i
1.2. LINEAR OPERATORS 15
In fact, any Hermitian operator (dened below) that satises the above is a
projector.
Remark 3: The projector is the rst example of an operator expressed as an outer product
of vectors: [i) i[. The jk element of this operator in matrix form is:
j[i) i[k) =
ij

ik
In matrix representation it is:
[i) i[ =
_

_
0
0
.
.
.
1(i
th
)
.
.
.
0
_

_
_
0, 0, 0, , 1(i
th
), , 0

=
_

_
0 i
th
0
.
.
.
.
.
.
.
.
.
i
th
1
.
.
.
.
.
.
.
.
.
.
.
.
0 0
_

_
This concept can be generalized to other operators.
1.2.3 Hermitian and Unitary Operators
Denition 1: Consider an operator

.

Hermitian

Antihermitian
Remark 1: Any operator

may be decomposed into its Hermitian and Antihermi-
tian parts:

H
+

A
Where

H
=
1
2
(

A
=
1
2
(

)
Remark 2: A hermitian operator is positive-denite:
a[

[a) 0 [a)
Denition 2: An operator

U is unitary if:

U

U

=

U

U =

I (

=

U
1
)
16 CHAPTER 1. PRELIMINARIES
Remark 3: If

U
1
and

U
2
are unitary, then so is the product of

U
1

U
2
.
Theorem: The inner product of vectors [a) , [b) is invariant under a unitary transfor-
mation.

U [a) =

U [b) =

_
Then:

_
= a[

U

U [b) = a[b)
1.2.4 Transformations (Unitary)
Consider a transformation

T.
Denition 1: An active transformation transforms all vectors in a subspace:
[a)

_
=

T [a)
Denition 2: A passive transformation transforms all operators on that space, leav-
ing the vectors unchanged.

=

T

T
1
Remark 1: In general,

T does not conseve Hermitian conjugation. For this, we
require that if:

=

T

T
1
and (

=

T

T
1
Then:
(

= (

T
1
)

= (

T
1
)

must = (

T
1
)

=

T
1
or

T

T
1
=

T
1

T =

I


T is unitary
Remark 2: If we have unitary transformations, active and passive transformations
are equivalent:
a[

[b)

_
= a[

U

U [b) where

_
=

U [a)
Remark 3: In addition of Hermitian conjugation, unitary transformations conserve:
1.2. LINEAR OPERATORS 17
Trace
Determinant
Algebraic equations between matrices and vectors
Remark 4: Any Hermitian matrix may be diagonalized by a unitary transformation.

D
where

D
is diagonal.
The diagonal elements are called eigenvalues of

. They are real, and they are
solutions of the secular or characteristic equation.
det(

I) = 0 (See below)
Remark 5: A single unitary matrix may diagonalize 2 (or more) Hermitian matrices

,

if: [

,

] = 0.
Remark 6: If the vector space is real, then

U
1
=

U



U
1
=

U

orthogonal transformation
which is a rotation.
1.2.5 The Eigenvalue Problem
Denition 1: Consider a linear operator

A and [u). Then C is an eigenvalue of

A if:

A[u) = [u)
and [u) is an eigenvector or eigenket.
Remark 1: If [u) is an eigenket of

A, then any multiple of [u) is also an eigenket be-
longing to the eigenvalue . These eigenkets form a subspace of the eigenvalue .
If n equals the dimension of the subspace, then:
n = 1 single or non-degenerate eigenvalues
n > 1 degenerate eigenvalues, where n equals the order of
degeneracy
Remark 2: If v[

A =

v[, then v[ is an eigenbra of



A. If

A is Hermitian, then:
The eigenvalues for bras and kets are the same
The eigenvalues are real
18 CHAPTER 1. PRELIMINARIES
The subspace of eigenbras is dual to the subspace of eigenkets
for the same eigenvalue.
Remark 3: Eigenkets make up the columns of the unitary matrix that diagonalizes
a Hermitian matrix.
Remark 4: In some cases, the unitary matrix is time dependent, and gives the
time evolution of a vector in some initial state.
[x(t)) =

U(t) [x(0))
In these cases, it is called the time-evolution operator, or propagator.
1.2.6 Functions of Operators, Generalizations to Innite Dimensions,
and More
In some cases, we will encounter functions of operators. (In fact, the propagator can
be expressed as e
i

Ht/
with

H the Hamiltonian, see Section 2)
In order to evaluate the action of these on vectors, we expand functions in a Taylor
Series.
f(

) =

n=0
a
n

n
Remark 1: Expressing f(

) as a Taylor Series in ne so long as the series converges.


Remark 2: It is often helpful to transform into an eigenbasis. For example, if
f(

) = e

1.2. LINEAR OPERATORS 19


Then
e

n=0

n
n!
=

n=0
1
n!
_
_
_
_
_

2
.
.
.

n
_
_
_
_
_
n
=

n=0
1
n!
_
_
_
_
_

n
1

n
2
.
.
.

n
n
_
_
_
_
_
=
_
_
_
_
_
e

1
e

2
.
.
.
e
n
_
_
_
_
_
Remark 3: A power series expansion is also useful for taking derivatives with respect
to a parameter. For instance, consider:
f(

) = e

Then
d
d
e

=
d
d

n=0

n
n!
=

n=0
n
n1

n
n!
=

n=1

n1

n
(n 1)!
=

n=0

n+1
n!
=

e

Remark 4: Here

behaves as if it was a complex number. The important distinction
is that multiple operators in general do not commute.
Generalization to Innite Vector Space
Most of the the formalism weve encountered deals with nite vector spaces.
20 CHAPTER 1. PRELIMINARIES
Sometimes we will need to consider innite vector spaces which contain vectors having
innite norm. The eigenvalues in this case are part of a continuous spectrum, and
most of the formalism still holds. (For instance, Section 1.2.5, remark 2 remains
valid.) Other generalizations are:
x[y) = (x y) Delta Dirac Function
where (x y) = 0 x ,= y
_
b
a
(x y) dy = 1 a x b
Also,

I =
_
b
a
[x) x[ dx
expresses completeness for a continuous spectrum.
Remark 5: One caveat concerning operators in a continuous vector space:

is not sucient to be Hermitian.


Example: Consider the operator:
p = i
d
dx
acting in real space representation. The matrix elements in this representation:
x[ p [y) = i(x y)
d
dy
where x, y are eigenvectors. Checking to see if p is Hermitian:
Is [ p [) = [ p [)

[ p [) =
_
b
a
_
b
a
[x) k[ p [y) y[) dxdy
=
_
b
a
_
b
a
[x) (i)(x y)
d
dy
y[) dxdy
=
_
b
a
[x) (i)
d
dx
x[) dx
=
_
b
a
x[)

(i)
d
dx
x[) dx
= (i) x[)

x[) [
b
a

_
b
a
(i) x[)
d
dx
x[)

dx
= (i) x[)

x[) [
b
a

_
b
a
_
[x) (i
d
dx
) x[)
_

dx
1.2. LINEAR OPERATORS 21
which will equal what we want if :
(i) x[)

x[) [
b
a
= 0
So we require surface terms to vanish for continuous Hermitian operators.
Remark 6: Restricting surface terms to vanish is equivalent to considering the
Physical Hilbert Space.
Remark 7: If (a, b) = (, ), then the typical functions
x[) (x)
xa,b
0 or e
ikx
End of Chapter Asides:
e
ikx
e
ik

= 0 if k = k

If k ,= k

, then:
lim
k
e
ikx
e
ik

x
= lim
L

_
L+
0
e
i(kk

)x
dx = 0
Some other properties:
p =

k
x[x) = x[x) or k

[k) = k [k)

k = i
d
dx
in position

k = i
d
dx
in momentum
1
2
_

e
ik(xy)
dk = (x y)
22 CHAPTER 1. PRELIMINARIES
Chapter 2
Classical Mechanics Review
2.1 Lagranges and Hamiltons Formulations of Classical Theory
2.1.1 Lagrangian
Consider a system with N degrees of freedom. In classical theory, the dynamical state
of the system is completely specied by the generalized coordinates q
1
, q
2
, . . . , q
N
and
generalized velocities q
1
, q
2
, . . . , q
N
.
Remark 1: The description gives the evolution of the system in conguration space.
Denition 1: The Lagrangian of the system:
L L(q
1
, q
2
, . . . , q
N
, q
1
, q
2
, . . . , q
N
)
is a characteristic function from which the equations of motion can be derived:
d
dt
_
L
q
i
_

L
q
i
= 0 for i = 1, . . . , N
These are called Lagranges Equations of Motion.
Remark 2:
p
i
=
L
q
i
These are called Lagrange conjugate momentum.
Remark 3: If all forces in the system are derivable from a static potential (the
situation we usually encounter), then the Lagrangian takes the form:
L = T
..
Kinetic
V
..
Potential
23
24 CHAPTER 2. CLASSICAL MECHANICS REVIEW
Remark 4: The Lagrange Equations of Motion may be derived by minimizing the
action of the system:
Action = S =
_
t
2
t
1
Ldt
S =
_
t
2
t
1
Ldt Principle of Least Action
2.1.2 Hamiltonian
An alternative but equivalent formulation of mechanics is in terms of the Hamiltonian.
Here, the dynamical state of the system is represented by a point in 2n dimensional
phase space, whose coordinates are the N position coordinates, q
1
, q
2
, . . . , q
N
, and the
N corresponding conjugate momenta p
1
, p
2
, . . . , p
N
.
Denition 1: The Hamiltonian is dened as the function H:
H H(q
1
, q
2
, . . . , q
N
, p
1
, p
2
, . . . , p
N
) =
N

i=
q
i
L
q
i
L
The equations of motion are:
q
i
=
H
p
i
, i = 1 . . . N
p
i
=
H
q
i
, i = 1 . . . N
Remark 1: If L = T V , then H = T + V = the total energy. In more general
situations, we still consider the Hamiltonian to be the total energy of the
system.
Remark 2:

H
dH
dt
=
H
t
This can be seen by using the chain rule and the Hamiltonian Equations of
Motion. This implies that if the Hamiltonian is not an explicit function of
time, then energy is conserved and we say that the system is conservative.
Remark 3: The Hamiltonian formulation is particularly important for QM since the
generalized coordinates and the conjugate momenta become operators in the
Hilbert Space.
2.1. LAGRANGES AND HAMILTONS FORMULATIONS OF CLASSICAL THEORY 25
2.1.3 Cyclic Coordinates, Poisson Brackets, and Canonical Transforma-
tions
Denition 1: A coordinate q
i
is cyclic if:
p
i
=
H
q
i
= 0
which is a statement of conservation of conjugate momentum.
Denition 2: The Poisson Bracket between two variable (p, q),(p, q) is given by
,

i
_

q
i

p
i


p
i

p
i
_
Remark 1: Consider (p, q). Then
d
dt
=
N

i
_

p
i
p
i
+

q
i
q
i
_
=
N

i
_

p
i
H
p
i


q
i
H
q
i
_
= , H
So, if a variables Poisson Bracket with the Hamiltonian vanishes, then that
variable is conserved.
Remark 2: Note the similarity between the Poisson bracket and the commutator.
Remark 3:
q
i
, q
j
= p
i
, p
j
= 0
q
i
, p
j
=
ij
Denition 3: A transformation of coordinates and momenta
q q

(q, p)
p p

(q, p)
that preserves Hamiltons equations of motion is canonical.
q

i
=
H
p

i
and p

i
=
H
q

i
26 CHAPTER 2. CLASSICAL MECHANICS REVIEW
Remark 4: Poisson brackets are invariant under canonical transformations.
2.1.4 Symmetries
Theorem (Nothers Theorem): Symmetry properties in the Hamiltonian imply
the existence of conserved quantities (conservation laws).
Proof: See Goldstein (p587, Section 13.7)
Remark 1: In classical mechanics, this is equivalent to nding a variable whose
Poisson bracket with the Hamiltonian vanishes. In QM, well see that an
operator whose commutator with the Hamiltonian vanishes, then this implies
that there exists a conserved quantity.
Examples:
Translational Invariance Conservation of Linear Momentum
Rotational Invariance Conservation of Angular Momentum
Time Translational Invariance Conservation of Energy
Some examples closely related to QM:
Reection Invariance Conservation of Parity
Invariance under Permutation Conservation of Symmetry
(symmetric, anti-symmetric)
Remark 2: Symmetries imply degeneracy in the eigenspectrum. Breaking the sym-
metry often lifts the degeneracy of the system. (Eg: Zeeman Eect)
Chapter 3
Postulates of QM
3.1 Postulates
I. The state of a particle is represented by a vector [(t)) in Hilbert Space.
II. Every observable in classical mechanics corresponds to a linear Hermitian
operator in QM. In particular:
x

X
p

P
(x, p)

(

X,

P)
_

T=

P
2
/2m

V =

V (

X)
The particular action of the operators depends on the choice of basis.
III. Any measurement of the observable associated with the operator

will
result only in values, , which satisfy

[) = [)
IV. For a system described by the normalized wave function [), the average
value of the observable corresponding to

is

) [

[)
This is called the expectation of

. Note: if [) is not normalized, one need
only divide the above equation by [).
V. The time evolution of the state vector is governed by the Schrodinger Equa-
tion:
i
d
dt
[(t)) =

H[(t))
where

H is the Hamiltonian.
27
28 CHAPTER 3. POSTULATES OF QM
Remark 1: In measurements, [) may not initially be an eigenstate of

(Postulate
III). But one can expand [) in an eigenbasis:
[) =

i
[
i
)
i
[)
The coecients in the expansion
i
[) give the probability of nding the
system in eigenstate [
i
) according to:
P() = [ [) [
2
Remark 2: Once an eigenvalue is measured, the system has probability of 1 of being
in the corresponding eigenstate. The measurement (even if idea) has changed
the state of the system; this is known as the collapse of the wave function.
Example:
[) =
1
2
[
1
) +
1
2
[
2
) +
1

2
[
3
)
The system has probability
_
1/4 of being in state
1
1/4 of being in state
2
1/2 of being in state
3
Say that a measurement gives the eigenvalues of
3
, then the state after the
measurement is
[) = [
3
)
Remark 3: Because coecients correspond to probabilities, expansions of normalized
wave functions must have

i
[
i
[) [
2
= 1
Remark 4: We must be careful with the consequences of Postulate II! Operators,
unlike observables, must obey commutation relations. In general, we will take
the symmetrized form (though this is not a universal prescription).
Eg: xp

X

P +

P

X
2
. .
This also makes a Hermitian operator!
Remark 5: For degenerate eigenvalues, probabilities represent projection onto the
corresponding eigenspace. Probability is for measuring a certain eigenvalue.
3.2. COMPATABILITY, INCOMPATIBILITY AND UNCERTAINTY 29
Remark 6: For continuous eigenspectra, we require that
_
P() d =
_
[ [) [
2
d
=
_
[) [) d
= [) = 1
And P() is called the probability density. P() d is the probability that
the measurement will give a value in the range to +d.
Remark 7: Not all QM observables have classical counterparts. (Eg. Spin)
Remark 8: The probabilistic nature of QM is a fundamental aspect of the theory.
It is not the same as the classical probabilistic description of a many-particle
system where the probabilistic approach is for practical purposes and where,
in principle, know the exact q(0), p(0) and the nature of the interaction would
give exact time evolution solutions.
3.2 Compatability, Incompatibility and Uncertainty
3.2.1 Compatible Variables
Denition 1: Two variables are compatible if their corresponding operators com-
mute:
[

,

] = 0
Remark 1: If two (Hermitian) operators commute, one can nd a unitary transfor-
mation which simultaneously diagonalizes them. Also, they will have a simul-
taneous eigenbasis. (See, for example, HW 2, Problem 2c, Shankar 1.8.10)
Denition 2: Two variables are incompatible if their corresponding operators do not
commute.
[

,

] ,= 0
Remark 2: Incompatible variables do have have any simultaneous eigenkets.
Remark 3: Two variables do not have to be either compatible or incompatible; they
may have some simultaneous eigenkets, but still do not commute. Thus they
fall into neither classication.
30 CHAPTER 3. POSTULATES OF QM
Denition 3: For any number of variables that mutually commute, one can nd a si-
multaneous eigenbasis which simultaneously diagonalizes each observable. The
set of these commuting variables is called a complete set of commuting observables
or a complete set of compatible variables if the eigenbasis is unique (up to an
overall phase); ie: is degenerate.
Remark 4: Consider an observable

and state vector [). One can expand [) in
an eigenbasis of

. The basis of

is unique (up to an overall phase) if none of
its eigenvalues are degenerate. If there is a degenerate eigenvalue, and if there
is an observable

that is compatible with

, then the simultaneous eigenbasis
will be unique (up to an overall phase). If there are no longer any degenerate
eigenvalues, then

and

form a complete set of commuting observables.
Remark 5: Compatible variables may be measured simultaneously and precisely. A
simultaneous measurement of all compatible variables will completely dene
the state vector of the physical system (and furthermore, that state will be
unique). Also, the dynamical state of the system is completely specied by
the quantum numbers associated with the eigenbasis.
Remark 6: A simultaneous measurement of a complete set of compatible variables
is akin to preparing the system at the time t
0
, after which the system evolves
according to the Schrodinger Equation (See 3.1, Postulate V)
Denition 4: A pure state is one that has been prepared by measuring all values of
compatible variables so that the state vector is know exactly.
Remark 7: In practice, a complete preparation is rare, and the dynamical state is not
specied exactly. In this case, the system is in a mixed state, and is described
by a statistical measure of states.
Remark 8: Mixed states include the statistical description in a classical sense, as
well as the inherent probabilistic nature of QM. We will visit this shortly in
Section 3.3 below).
3.2.2 Incompatible Variables and Uncertainty Relationships
Recall Section 3.1, Postulate IV (

) = [

[)).
Denition 1: The uncertainty in

is given by:


_
(

_
)
2
_
1/2
=
_
[ (

_
)
2
[)
3.2. COMPATABILITY, INCOMPATIBILITY AND UNCERTAINTY 31
In a statistical description, this is also called the standard deviation or root-
mean-square, and is a measure of the uncertainty due to uctuations.
Note:

=
_
(

_
)
2
_
1/2
=
_

2
2

_
+
_

_
2
_
1/2
=
_
_

2
_

_
2
_
1/2
Remark 1: If the state [) is an eigenstate of

, then there is no uncertainty.
(

= 0)
Uncertainty Derivation: Consider two incompatible Hermitian operators

and

that have
[

,

] = i

where

is also Hermitian. Let [

,

]
+
=

denote the anticommutator,


and let [) be a normalized state so that:
_

_
= [

[)
_

_
= [

[)
Then
(

)
2
(

)
2
= [ (

_
)
2
[) [ (

_
)
2
[)
If we let

=


_

_
and

=


_

_
, then we have:
(

)
2
(

)
2
= [ (

)
2
[) [ (

)
2
[)
= [ (

)(

) [) [ (

)(

) [)
= [ (

) [) [ (

) [) since Hermitian.

[)

2
by Schwartz Inequality
=

[)

2
32 CHAPTER 3. POSTULATES OF QM
But now, we also have that
(

)(

) =

2
+

2
=
1
2
[

]
+
+
1
2
[

]
And so we get that:
(

)
2
(

)
2

[)

2
=

[
1
2
[

]
+
[) +
1
2
[

[)

1
4
[ [

]
+
[)
2
+
1
4
[

[)
2
This is the most general uncertainty relation between any two Hermitian op-
erators.
Remark 2: If

and

are canonically conjugate, then

I
And thus
(

)
2
(

)
2

1
4
[ [

]
+
[)
2
+

2
4


2
4



2
THE uncertainty relation
Remark 3: The equality holds for

[) = (constant)

[)
[ [

]
+
[) = 0
Remark 4: The most popular statements are
x p

2
holds for each component
Et

2
3.3. PURE STATES AND MIXTURES: DENSITY MATRICES 33
3.3 Pure States and Mixtures: Density Matrices
Recall from Section 3.2.1 that:
A pure state is one where we know the dynamical state of the system exactly
A mixed state is one where the exact state of the system is unknown.
Suppose that we want to measure the quantity

in a mixed system. Then there is a probability
P
m
of obtaining the average
_

_
m
= m[

[m), and
_

_
=

m
P
m
m[

[m)
where

indicates an average in the classical sense.


Remark 1: In the above equation, the [m)s are normalized but not necessarily orthogonal, and
P
m
0 and

m
P
m
= 1
Remark 2: There are two kinds of averaging in the equation above:
Quantum average m[

[m) for each system in [m)


Classical average over systems with dierent [m)
3.3.1 Density Matrix
Denition 1: A density matrix (or operator) is a convenient way to describe a mixed
state:

m
P
m
[m) m[
This is called the density operator.
Remark 1: The average value of

in a mixed state is given by
_

_
= Tr (

)
To see this:
Tr (

) =

n
n[

[n)
=

n,m
P
m
n[m)
. .
nm
m[

[n)
=

m
P
m
m[

[m)
=
_

_
34 CHAPTER 3. POSTULATES OF QM
Remark 2: It is sucient to know in order to calculate all measurable physical
quantities, and thus every quantized, statistical mixture is exactly and com-
pletely dened by its density operator.
Remark 3: = [) [ is a special case: it is the density matrix formulation of a
pure state (all P
m
= 0 except 1). In this case:
_

_
= [

[)
= Tr [

[)
= Tr ([) [

)
= Tr (

)
which is the usual quantum average.
Properties of the Density Matrix:
1. is positive denite: [ [) 0 H.
[ [) =

m
P
m
[m) m[) =

m
P
m
[ [m) [
2
0
2. is Hermitian

=
_

m
P
m
[m) m[
_

m
P
m
[m) m[ =
3. has unit trace
Tr ( ) = Tr
_

m
P
m
[m) m[
_
=

m
P
m
Tr ([m) m[)
. .
1
=

m
P
m
= 1
4. Tr (
2
) 1
Tr (
2
) =

n
P
m
P
n
[m) m[n)
. .
nm
m[ =

m
P
2
m
1
3.3.2 Density Matrix of a Pure State
If the density operator represents a pure state, ie. = [m) m[, then we also have

2
= (projector)
Tr (
2
) = 1 (see properties above)
3.3. PURE STATES AND MIXTURES: DENSITY MATRICES 35
Suppose the density operator is expressed in its matrix form in a particular represen-
tation:
_
_
_
_
_
_

11

12

1n

21
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

n1

nn
_
_
_
_
_
_
with
nm
= n[ [m).
Question: Given the explicit matrix form, how do we determine whether it represents
a pure state or a mixed state?
Answer: Since the pure state can be written = [) [, and
[) =
_

1
.
.
.

n
_

_
Then we have that:
[) [ =
_

1
.
.
.

n
_

_[

1
, ,

n
] =
_

_
[
1
[
2

2

1

1
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

1
[
n
[
2
_

_
Now, if satises a pure state, then its elements must satisfy:

ij

ji
=
ij

ij
= [
ij
[
2
=
ii

jj
If this holds for all i, j, then the density matrix is a pure state; otherwise it is a mixed
state. This can be seen by doing the matrix multiplication.
3.3.3 Density Matrix of Mixed States
Given that =

j
P
j
[
j
)
j
[ it is not required that the
j
s be orthogonal.
Example: Consider a two state system where [1) and [2) form the orthogonal basis
for the 2D Hilbert space. One possible density operator is:
=
1
2
[1) 1[ +
1
2
_
1

2
([1) +[2))
1

2
(1[ +2[)
_
50% mixture of state [1) and
1
2
([1) +[2)) which are not orthogonal.
36 CHAPTER 3. POSTULATES OF QM
Remark 1: The density operator in the above is a non-unique representation of
non-orthogonal mixtures.
Remark 2: It is always possible to write as a mixture of orthogonal states. For ex-
ample, consider the general density operator =

j
P
j
[
j
)
j
[. Let R be the
matrix representation of in the orthonormal basis dened by [1) , [2) , . . . , [N).
Then:
R =
_

11

12

1N

21
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

N1

NN
_

mn
= m[ [n) =

j
P
j
m[
j
)
j
[n)
Then we write:
= [[1) , [2) , . . . , [N)]R
_

_
1[
2[
.
.
.
N[
_

_
= [[1) , [2) , . . . , [N)]

. .
[|
1
,|
2
,...,|
N


UR

. .
2
6
6
6
4
p
1
.
.
.
p
N
3
7
7
7
5


U
_

_
1[
2[
.
.
.
N[
_

_
. .
2
6
6
6
4

1
[
.
.
.

N
[
3
7
7
7
5
=

j
P
j
[
j
)
j
[
Remark 3: If [1) , [2) , . . . , [N) form a orthonormal basis, then so does [
1
) , [
2
) , . . . , [
N
)
because

U is unitary.
3.3. PURE STATES AND MIXTURES: DENSITY MATRICES 37
Recall Example:
=
1
2
[1) 1[ +
1
2
_
1

2
([1) +[2))
1

2
(1[ +2[)
_
=
3
4
[1) 1[ +
1
4
[1) 2[ +
1
4
[2) 1[ +
1
4
[2) 2[
= [[1) , [2)]
_
3/4 1/4
1/4 1/4
_ _
1[
2[
_
= [[1) , [2)]R
_
1[
2[
_
Now we would just diagonalize R to create a diagonalized in some basis.
Remark 4: In some cases, the non-uniqueness of goes even further.
= [[1) , [2)]
_
1/2 0
0 1/2
_ _
1[
2[
_
= [[1) , [2)]

U

U

_
1/2 0
0 1/2
_

U
. .
1
2

_
1[
2[
_
= [[1) , [2)]

U
. .
_
1/2 0
0 1/2
_

U

_
1[
2[
_
. .
= [[
1
) , [
2
)]
_
1/2 0
0 1/2
_ _

1
[

2
[
_
In this case, can be expressed in an innite number of forms depending on
the unitary transformation

U used.
3.3.4 Time Evolution of the Density Matrix
Regression: QM can be expressed in equivalent pictures depending on whether
the transformations which govern time evolution are passive or active. (Recall
1.2.4)
The Schrodinger picture (active) has state vectors evolve with
time while operators remain constant.
The Heisenberg picture (passive) has states remain constant
while operators evolve.
In addition, there is
Interaction Picture: Here, the non-interacting part of the Hamil-
tonian is in the Schrodinger picture while the interacting piece
has time-dependent operators.
38 CHAPTER 3. POSTULATES OF QM
Time Evolution of : Recall 3.1 - Postulate V i [(t)) =

H[(t)). What then
is the time evolution of ?

j
P
j
_
[
j
)
_
d
dt

j
[
_
+
_
d
dt
[
j
)
_

j
[
_
=

j
P
j
_

j
[
_
1
i

H[
j
)
_

+
1
i
[
j
)
j
[
_
=
1
i

j
P
j
_

H[
j
)
j
[ [
j
)
j
[

H
_
=
1
i
[

H, ]
Remark 1: This is the quantum version of the Liouville Equation from classical
statistical mechanics. (see 2.1.3, remark 1 with = the density of phase
space.)
Remark 2: Well see later that the Heisenberg equation of motion for an operator,

A
H
, is given by:

A
H
=
1
i
[

A
H
,

H]
Witness that the equation of motion for has the commutator in reverse order.
Remark 3: The density matrix is a special operator since it changes in time in the
Schrodinger picture.
3.4 Schrodinger Equation
Recall 3.1, Postulate V, which said that the Schrodinger equation is:
i
d
dt
[(t)) =

H[(t))
where

H is the Hamiltonian of the system.
Remark 1: In general, one may use Postulate II to construct the Hamiltonian

H from the classical
Hamiltonian. Modications to this are required if there is no classical analog (eg spin).
Remark 2: Though we can only postulate about the form of the time evolution of a vector in
Hilbert space, we do know that it must be linear and homogeneous (so that we can have
linear superposition). We also know that it must be a DE of the 1st order with respect to
time (so that the evolution is uniquely determined by the state at some initial time).
3.4. SCHRODINGER EQUATION 39
Remark 3: Predictions of Quantum Theory must coincide with classical mechanics in the regime
where classical mechanics is valid (so there is some formal analogy with classical mechanics).
This is called the correspondence principle.
3.4.1 Solving the Schrodinger Equation
For now, we will consider

H =

H( x, p), where the Hamiltonian is a function of po-
sition and momentum, but not explicitly time. In general, the time dependence (if
present) is treated perturbatively, so we will discuss Hamiltonians with explicit time
dependence later.
Solving the equation formally gives a solution of the form:
()
[) = e
i

Ht/
[(0))
Remark 1: One can write this as:
[(t)) =

U(t) [(0))
where

U(t) = e
i

Ht/
is called the propagator or time evolution operator.
Remark 2: Since

H is Hermitian,

U is unitary.
Remark 3: Time evolution may be thought of as a rotation of a state in Hilbert
space.
To evaluate (), it is useful to expand [(0)) in an energy eigenbasis. This will diago-
nalize the Hamiltonian and allows one to write:
e
i

Ht/
[E) =

n
1
n!
_
it

_
n
_
_
_
E
1
.
.
.
E
n
_
_
_[E)
=
_
_
_
e
iE
1
t/
.
.
.
e
iEnt/
_
_
_[E)
To determine the energy eigenbasis:
1. First solve the time-independent Schrodinger equation:

H[E) = E[E)
2. The expand the wavefunction in terms of the energy eigenbasis:
[(t)) =

E
[E) E[(t))
. .
a
E
(t)
=

E
a
E
(t) [E)
40 CHAPTER 3. POSTULATES OF QM
3. Now substitute this into i

[) =

H[) to obtain:
i
d
dt
a
E
(t) = Ea
E
a
E
(t) = a
E
(0)e
iEt/
= E[(0)) e
iEt/
[(t)) =

E
[E) E[(0)) e
iEt/
=

U(t) [(0))
This, by comparison, gives a propagator of the form:

U(t) =

E
[E) E[ e
iEt/
Remark 4: If the eigenspectrum is degenerate, an additional index is required to
specify the state:

U(t) =

E
[E, ) E, [ e
iEt/
If the eigenspace is continuous, then

E

_
dE
Remark 5: [E(t)) = [E) e
iEt/
are the normal modes called stationary states since
the probability distribution for any variable is time independent in these states.
Chapter 4
Simple Problems in 1 Dimension
In general, we will need to chose a basis to solve Schrodingers equation. Though one can choose
any basis, some choices are better than others depending on the form of the Hamiltonian. The
following are some examples of 1 dimensional problems.
Remark 1: The x basis (position basis) is the basis for the alternative form for QM based on the
wave equation. In this basis, x x,

P i
d
dx
, and

H i

t
.
4.1 Free Particle
4.1.1 Solving the Schrodinger Equation
In the free particle case, we have

H =
1
2m

P
2
Solving the time dependent Schrodinger equation:
()
i
d
dt
[(t)) =

P
2
2m
[)
which gives us general solutions of
[(t)) = [E) e
iEt/
Substituting this back into () gets us the time independent Schrodinger equation

P
2
2m
[E) = E [E)
41
42 CHAPTER 4. SIMPLE PROBLEMS IN 1 DIMENSION
Remark 1: Here, we may choose the momentum basis since [P) is also an eigenstate
of

P
2

P
2
2m
[P) = E[P)
_
p
2
2m
E
_
[P) = 0
p =

2mE
Thus the two orthogonal eigenstates are:
[E
+
) =

p =

2mE
_
[E

) =

p =

2mE
_
Remark 2: The energy eigenstate is degenerate since a particle can be moving to
the right or the left with the same energy. Using the compatible operator

P,
we can determine the possible outcomes so that a measurement of p, and thus
E, will give us a complete description of the dynamical state of the system.
Remark 3: The most general state is given by
[) = [E
+
) + [E

)
which is a superposition of the particle moving to the left and right.
Remark 4: Note that measuring the momentum gives us no information about the
position of the particle since p = 0 x = by Heisenbergs uncertainty
principle.
4.1.2 Propagator for a Free Particle in the Position Basis
The matrix elements for the propagator in the x-basis are
U(x, t; y
..
(x
0
)
) x[

U(t) [y) =
_

dp x[ e
ip
2
t/2m
[P) P[y)
Now consider:
x[y) = (x y) =
1
2
_

dke
ik(xy)
=
_

dp x[P) P[y) inserting a complete set


=
_

dk x[k = p/) k = p/[y)


4.2. PARTICLE IN A BOX 43
where the last step follows since we know p = k. Now, by comparison, we can see
that
x[k) k[y) =
1
2
e
ik(xy)
or
x[k = p/) =
1

2
e
ipx/
= x[P)
Now, using this to evaluate U(x, t; y), we get
U(x, t; y) =
1
2
_

e
ip(xy)/
e
ip
2
t/2m
dp
=
1
2
_

exp
_
it
2m
_
p
m(x y)
t
_
2
_
exp
_
im(x y)
2
/2t
_
. .
from completing the square
=
1
2
_
2m
it
e
im(xy)
2
/2t
=
_
m
2it
e
im(xy)
2
/2t
Remark 1: For a free particle whose state is known at some time t
0
, the state at
some t > t
0
is given by
(x, t) =
_
U(x, t; y, t
0
)(y, t
0
) dy
where
U(x, t; y, t
0
) = x[ U(t t
0
) [y)
Note that this depends only on t t
0
. This is because there is no explicit time
dependence in the Hamiltonian.
Remark 2: If we consider a free particle whose position is known at time t
0
:
(t
0
) = (x x
0
)
Then
(x, t) =
_
m
2it
_
1/2
e
im(xx
0
)
2
/2t
Physically, this corresponds to the delta function broadening as a function of
time.
44 CHAPTER 4. SIMPLE PROBLEMS IN 1 DIMENSION
Figure 4.1: Diagram of the Box Potential
4.2 Particle in a Box
Consider a particle under the inuence of a potential V (x) where:
V (x) =
_
0 |x|L/2
otherwise
We will choose the position basis for the time-independent Schrodinger equation:

H =
p
2
2m
+

V

H[) = E[)
x[

H[) = Ex[) = E(x)
x[
1
2m
_
i
d
dx
_
2
+

V [) =

2
2m
d
2
dx
2
(x) +

V (x)
Combining the last two, we have
d
2
dx
2
+
2m

2
(E V ) = 0
Now in Regions I and III,
I
=
III
= 0 (well see why in 4.4)
In Region II:
V = 0
d
2
dx
2
=
2mE

II
= Aexp
_
i
_
2mE/
2
x
_
+Bexp
_
i
_
2mE/
2
x
_
= Ae
ikx
+Be
ikx
where k =
_
2mE

2
Remark 1: We require that the wavefunction must be continuous

I
(L/2) =
II
(L/2) and
II
(L/2) =
III
(L/2)
4.2. PARTICLE IN A BOX 45
So then, applying the boundary conditions, we have that
()
0 = Ae
ikL/2
+Be
ikL/2
0 = Ae
ikL/2
+Be
ikL/2
Which implies
_
e
ikL/2
e
ikL/2
e
ikL/2
e
ikL/2
__
A
B
_
= [0)
The non-trivial solutions to this problem are found by nding the zeroes of the determinant.
e
ikL
e
ikL
= 0 = 2i sin(kL) k =
n
L
n = 0, 1, 2,
Now we substitute k back into () to get the allowed values of A and B:
Ae
in/2
+Be
in/2
= 0
A = Be
in
= B
_
1 if n is even
1 if n is odd
Thus we have that
(x) =
_
A(e
ikx
+e
ikx
) n odd
A(e
ikx
e
ikx
) n even
Remark 2: Choose A so that
_

dx = 1 (normalize):
(x) =
_
q
2
L
cos(
nx
L
) n is odd
q
2
L
sin(
nx
L
) n is even
Remark 3: If n = 0, then = 0 everywhere, which is not very exciting. Also, values of n are
physically equivalent, so we only deal with positive n to make life simpler.
Remark 4: Comparing our expressions for k:
_
2mE

2
=
n
L
We nd that the energy in the box in quantized
E
n
=
n
2

2
2mL
2
n = 1, 2, 3,
Remark 5: As L goes to innity, the energy spectrum becomes continuous.
Remark 6: The ground state, n = 1, has E =
2

2
/2mL
2
> 0 which in a necessary consequence
of the uncertainty principle.
x = L/2 p /2
46 CHAPTER 4. SIMPLE PROBLEMS IN 1 DIMENSION
since
p =
_
p
2
) p)
2
and p) = 0
(2mE
1
))
1/2
=

L
xp =
L
2

L
>

2
4.3 Step Potentials: Reection and Transmission
Consider the potential:
x = 0
V1
V2
Here, the time-independent Schrodinger equation is


2
2m
d
2
dx
2
= (E V
1
) , x < 0


2
2m
d
2
dx
2
= (E V
2
) , x > 0
(4.1)
Remark 1: We must consider two dierent energy regions: V
1
< E < V
2
and E > V
2
V
1
< E < V
2
:
We expect the wavefunctions to have the following form:
(x) =
_
e
i

2m(EV
1
)
, x<0
e

2m(V
2
E)
, x>0
(4.2)
Remark 2: The actual solution is constructed as linear combinations of the forms in
(4.2)
4.3. STEP POTENTIALS: REFLECTION AND TRANSMISSION 47
Remark 3: The solution oscillates in the region where E < V
2
, and decays (or grows)
exponentially when E > V
2
.
Remark 4: To satisfy the boundary conditions, we impose the following conditions:
1. (x) must remain nite as x to be physically realizable
2. (x) and

(x) must be continuous at the discontinuity at x = 0.


This gives a continuous probability in space (no source or sink
of particles).
Taking these considerations into account, the general solution will have the following
form:
(x) =
_
A
+
e
ikx
+A

e
ikx
, x < 0
B
+
e
x
+B

e
x
, x > 0
(4.3)
where
k =
_
2m(E V
1
) ; =
_
2m(V
2
E) (4.4)
Now, from Remark 4, we know that:
B
+
= 0 (from condition 1)
A
+
A

= B

(from condition 2, continuity of )


ik(A
+
A

) = B

(from condition 2, continuity of

)
Combining these and solving:
ik(A
+
A

) = (A
+
+A

)
(ik +)A
+
= (ik )A

(
2
+k
2
)e
i
= (
2
+k
2
)e
i
A

(4.5)
where
tan() =
k

= e
2i
A
+
Remark 5: A
+
and A

are not necessarily complex conjugates of one another, but


we can introduce an overall phase. We multiply (4.5) by i so that
A
+
=
A
2i
e
i
and A

=
A
2i
e
i
where A is some real constant.
Then, we have
(x) =
_
Asin(ki +), x < 0
Asin()e
x
, x > 0
(4.6)
48 CHAPTER 4. SIMPLE PROBLEMS IN 1 DIMENSION
Remark 6: In the region x < 0, we have a wave traveling toward the step potential
and a reected component of equal magnitude traveling away. We see this by
the time-dependent Schrodinger equation:
(x, t) = e
iEt/
Asin(kx +)
=
A
2i
e
i(Et/kx)
. .
Wave going right

A
2i
e
i(Et/+kx)
. .
Wave going left phase shifted by
Remark 7: Note that the wavefunction doesnt vanish for x > 0; there is some
barrier penetration that decays more rapidly for larger step sizes. If V
2
,
there is no barrier penetration. This is the justication for setting (x) = 0
outside of the innite square well in the previous section.
Remark 8: In wave phenomena, this penetration is a familiar part of reection and
transmission. In E&M, the penetrating part of light at an interface is called
an evanescent wave.
E > V
2
:
In this case, the solution of the Schrodinger equation is:
(x) =
_
A
+
e
ikx
+A

e
ikx
, x < 0
B
+
e
ik

x
+B

e
ik

x
, x > 0
(4.7)
with
k

=
_
2m(E V
2
) (4.8)
Imposing boundary conditions, we have
A
+
+A

= B
+
+B

(4.9)
ik(A
+
A

) = ik

(B
+
B

) (4.10)
Remark 9: Note that we only have 2 equations to specify 4 unknowns. This is
because there are 2 linearly independent solutions for each energy (energy is
2-fold degenerate, or doubly degenerate, or has a degeneracy of order 2.
The 2 possible scenarios corresponding to this degeneracy are:
So we see that the wave is launched from either the right or the left and is partially
reected (or transmitted) at the step boundary.
4.4. FINITE SQUARE WELL: DISCRETE ENERGIES AND RESONANCES 49
x = 0
V
1
V
2
x = 0
V
1
V
2
k
k
r
k
t
k
k
t
k
r
Case A
Case B
Consider now scenario A above, where the wave is launched from the left. In this case,
we have B

= 0, which, from Eq (4.9) above implies that


B
+
= A
+
+A

Using Eq (4.10) above, we also have that


ik(A
+
+A

) = ik

(A
+
+A

)
A
+
(k k

) = A

(k +k

) (4.11)
Thus we get that the solution for a wave originating from the left is
(x) =
_

_
A(e
ikx
+
k k

k +k

e
ikx
), x < 0
A
2k
k +k

e
ik

x
, x > 0
(4.12)
where A
+
= A for convenience. Now, including the time dependence for x < 0:
(x, t) = Ae
i(Et/kx)
+A
k k

k +k

e
i(Et/+kx)
(4.13)
Remark 10: As E , the amplitude
A(kk

)
k+k

of the reected wave 0 while the


amplitude
A(2k)
k+k

of the transmitted wave goes to the amplitude of the incident


wave (eg, the step became negligible). As the energy E V
2
, then k

0
and the amplitude of the reected wave goes to the amplitude of the incident
wave and the transmitted wave amplitude 0 since it is preparing to become
exponentially damped barrier penetration.
4.4 Finite Square Well: Discrete Energies and Resonances
Consider the potential
So we have
V (x) =
_
V
1
, |x|L/2
V
2
, |x|>L/2
From our previous two examples, we expect the following:
50 CHAPTER 4. SIMPLE PROBLEMS IN 1 DIMENSION
x
V
V2
V1
L
2

L
2
Discrete energies for V
1
< E < V
2
Barrier penetration for V
1
< E < V
2
where [x[ > L/2
For E > V
2
, transmission and reection at discontinuities in potential. We also expect
2-fold degeneracy since we can have waves coming from the left or the right.
Resonances as a function of energy in the transmission and reection of waves.
To actually solve the problem, we will again consider each energy range separately:
V
1
< E < V
2
:
The time independent Schrodinger equation is:

2
2m
d
2
dx
2
=
_
(EV
1
), |x|<L/2
(V
2
E), |x|>L/2
(1)
Then the solutions are
(x) =
_
A
+
e
ikx
+A

e
ikx
, [x[ L/2
B
+
e
x
+B

e
x
, [x[ > L/2
(2)
where
k =
_
2m(E V
1
) and =
_
2m(V
2
E) (3)
Applying boundary conditions, we have
(x ) <
B
+
= 0 for x > L/2
B

= 0 for x < L/2


(x) =
_

_
A
+
e
ikx
+A

e
ikx
, [x[ < L/2
B
+
e
x
, x < L/2
B

e
x
, x > L/2
(4)
Continuity of at x = L/2
A
+
e
ikL/2
+A

e
ikL/2
= B
+
e
L/2
(5)
A
+
e
ikL/2
+A

e
ikL/2
= B

e
L/2
(6)
4.4. FINITE SQUARE WELL: DISCRETE ENERGIES AND RESONANCES 51
Continuity of

at x = L/2
ik
_
A
+
e
ikL/2
A

e
ikL/2
_
= B
+
e
L/2
(7)
ik
_
A
+
e
ikL/2
A

e
ikL/2
_
= B

e
L/2
(8)
Now we solve the system:
1. take (5) (7)
( +ik)A
+
e
ikL/2
+ ( ik)A

e
ikL/2
= 2B
+
e
L/2
(9)
( ik)A
+
e
ikL/2
+ ( +ik)A

e
ikL/2
= 0 (10)
2. Similarly, take (6) (8)
( +ik)A
+
e
ikL/2
+ ( ki)A

e
ikL/2
= 0 (11)
( ik)A
+
e
ikL/2
+ ( +ki)A

e
ikL/2
= 2B

e
L/2
(12)
3. From (10) and (11), we have
A
2
+
= A
2

A
+
= A

(13)
Remark 1: The plus in (13) corresponds to an even parity solution (cosines in
box) the negative sign corresponds to an odd parity solution (sines in box).
Even Parity Solutions: Let
A
+
= A

=
A
2
(14)
Then from (10) or (11):
_

2
+k
2
cos
_
kL
2
+
_
= 0 (15)
where = arctan(k/).

kL
2
+ arctan
_
k

_
=
n
2
where n = odd (16)
Also, from (9) and (12), we get:
_

2
+k
2
Acos(kL/2 ) = 2B
+
e
L/2
_

2
+k
2
Acos(kL/2 ) = 2B

e
L/2
(17)
B
+
= B

=
A

2
+k
2
2
cos
_
kL
2

_
e
L/2
(18)
52 CHAPTER 4. SIMPLE PROBLEMS IN 1 DIMENSION
So the wave function for the even parity solutions is
(x) =
_

_
Acos(kx), [x[ < L/2
A

2
+k
2
2
cos
_
kL
2

_
e
L/2
e
x
, x < L/2
A

2
+k
2
2
cos
_
kL
2

_
e
L/2
e
x
, x > L/2
(19)
Odd Parity Solutions: Let
A
+
= A


A
2i
(20)
Then, following the same procedure as in the even case, we get the energy
constraints:
kL
2
+ arctan
_
k

_
=
n
2
(21)
And the coecients B
+
and B

are:
B
+
= B

= A

2
+k
2
2
sin
_
kL
2

_
e
L/2
(22)
This gives us odd parity solutions of:
(x) =
_

_
Asin(kx), [x[ < L/2
A

2
+k
2
2
sin
_
kL
2
+
_
e
L/2
(e
x
), x < L/2
A

2
+k
2
2
sin
_
kL
2
+
_
e
L/2
(e
x
), x > L/2
(23)
Remark 2: The general form of the solutions looks like:
Remark 3: The allowed energies are obtained from Eq (16) and (21) for the even
and odd parities (respectively). Consider the even parity energy values. It is
helpful to eliminate the so that all energy dependence is in k. As a reminder,
Eq (16) was:
n
2
=
kL
2
+ arctan
_
k

_
Looking at just the trig. component, we have the triangle:
4.4. FINITE SQUARE WELL: DISCRETE ENERGIES AND RESONANCES 53
E
V
(a) Even case: Note that the expo-
nential decay parts have the same
sign
E
V
(b) Odd case: Note that the ex-
ponential decay parts have opposite
signs

2
+k
2
arctan
_
k

_
Which means that we have:
arctan
_
k

_
= arcsin
_
k

k
2
+
2
_
= arcsin
_
k
2m(E V
1
) + 2m(V
2
E)
_
by Eq (3)
= arcsin
_
k
_
2m(V
2
V
1
)
_
(24)
Thus, Eq (16) becomes:
n
2

kL
2
= arcsin
_
k
_
2m(V
2
V
1
)
_
(25)
To nd the actual energy levels is very dicult, as the above is a transcendental
equation and thus must be solved either graphically or numerically. We can
look at it graphically by plotting the left and right hand sides of the equation
and looking for their intersections.
So solutions for n > 1 require:
n
2

1

_
2m(V
2
V 1)
L
2


2
where
1

_
2m(V
2
V 1)
L
2
54 CHAPTER 4. SIMPLE PROBLEMS IN 1 DIMENSION
k
k =
1
h
_
2m(V
2
V
1
)
lhs
rhs
is the largest k allowed by the rhs of Eq (25).

L
_
2m(V
2
V
1
)

(n 1)
And thus there are a nite number of intersection points since the maximum
value allowed for k is

2m(V
2
V
1
)

(which corresponds to the kinetic energy


being equal to the depth of the well). Eg: in the gure below there are 7
allowed values (energies/intersection points).
V
1
V
2
E > V
2
:
Remark 4: In this range, there is 2-fold degeneracy. We will construct solutions for
the wave originating form .
4.4. FINITE SQUARE WELL: DISCRETE ENERGIES AND RESONANCES 55
Thus, we will have a solution of the form:
(x) =
_

_
A
+
e
ik

x
+A

e
ik

x
, x < L/2
B
+
e
ikx
+B

e
ikx
, [x[ < L/2
Ce
ik

x
, x > L/2
(1)
where
k =
_
2m(E V
1
) and k

=
_
2m(E V
2
) (2)
Solving the boundary conditions, we have continuity of at x = L/2:
A
+
e
ik

L/2
+A

e
ik

L/2
= B
+
e
ikL/2
+B

e
ikL/2
(3)
B
+
e
ikL/2
+B

e
ikL/2
= Ce
ik

L/2
(4)
And by the continuity of

at x = L/2:
ik

_
A
+
e
ik

L/2
A

e
ik

L/2
_
= ik
_
B
+
e
ikL/2
B

e
ikL/2
_
(5)
ik
_
B
+
e
ikL/2
B

e
ikL/2
_
= ik

Ce
ik

L/2
(6)
Solving these, if we go k

(4) (6), we get:


(k

+k)B
+
e
ikL/2
+ (k

k)B

e
ikL/2
= 2k

Ce
ik

L/2
(7)
(k

k)B
+
e
ikL/2
+ (k

+k)B

e
ikL/2
= 0 (8)
From (8), we have
B

= B
+
k k

k +k

e
ikL
(9)
Putting (9) (6), gives C in terms of B
+
k
_
B
+
e
ikL/2
B
+
k k

k +k

e
ikL/2
_
= k

Ce
ik

L/2
(10)
C =
2kB
+
e
i(kk

)L/2
k +k

(11)
Now, relate these to A
+
and A

. Take k

(3) + (5):
2k

A
+
e
ik

L/2
= (k

+k)B
+
e
ikL/2
+ (k

k)B

e
ikL/2
= B
+
e
ikL/2
_
(k +k

)e
ikL

(k k

)
2
(k +k

)
e
ikL
_
by (9)
=
B
+
e
ikL/2
(k +k

)
_
4kk

cos(kL) 2i(k
2
+ (k

)
2
) sin(kL)

56 CHAPTER 4. SIMPLE PROBLEMS IN 1 DIMENSION


B
+
=
A
+
e
i(k+k

)L/2
2k
k+k

cos(kL)
i(k
2
+(k

)
2
)
k

(k+k

)
sin(kL)
(12)
Remark 5: Note that
A+ B+ C
A B
where A

, B

, C are the amplitudes of the waves.


Remark 6: We may choose A
+
to be the overall scale factor of the wave, and thus we get the
other constants in terms of A
+
. The most important coecients are C and A

.
Now from (11) and (12):
C
A
+
=
e
ik

L
cos(kL)
i(k
2
+(k

)
2
)
2kk

sin(kL)
(13)
And we also have:
A

A
+
= e
ik

(k
2
(k

)
2
)
2kk


sin(kL)
cos(kL)
i(k
2
+(k

)
2
)
2kk

sin(kL)
(14)
Remark 7: One can show that

C
A
+

2
+

A
+

2
= 1 (15)
Physically, A
+
, as illustrated in Remark 5, represents an incoming beam of particles which
scatter o the square potential. Transmitted particles are represented by the beam with
amplitude C, and the reected particles are represented by the beam with amplitude A

.
So, by (15), the incident particle ux ( [A
+
[
2
) = the transmitted ux ux ( [C[
2
) + the
reected ux ( [A

[
2
).
Remark 8: From (13) and (14), we know that if:
k
n
L = n (16)
then the transmission coecient

C
A
+

2
= 1 and the reection coecient

A
+

2
= 0. This
gives a series of transmission resonances with resonance energies:
E
n
=

2
2m
_
n
L
_
2
(17)
4.4. FINITE SQUARE WELL: DISCRETE ENERGIES AND RESONANCES 57
At these energies, the entire wave is transmitted.
Remark 9: The energies in (17) are not discrete in the sense of the energies conned in the
well. All the energies in the neighborhood of the resonance energies are allowed, but the
transmitted wave amplitude is diminished if not at resonance.
Remark 10: The sharpness of resonances is determined by the ratio between k and k

. Consider
the case where
(
k

k
)
2
=
E V
2
E V
1
1 V
1
V
2
(a deep well) (18)
Then

C
A
+

1
cos
2
(kL) +
_
k
2k

_
2
sin
2
(kL)
(19)
In between resonances, where sin
2
(kL) 1, we have

C
A
+

2
4
_
k

k
_
2
1 (20)
This implies that transmission nearly vanishes between resonances (but not completely!).
k
n1
k
n
k
n+1
k k
T
k
To approximate resonance widths, lets expand the cos(kL) and sin(kL) in the vicinity of the
resonance.
cos(kL) = cos(k
n
L + k
n
L) = cos(k
n
L) cos(k
n
L) sin(k
n
L) sin(k
n
L)
(1) cos(k
n
L) + (0) sin(k
n
L)
1 for small k
n
sin(kL) = sin(k
n
L + k
n
L) = cos(k
n
L) sin(k
n
L) + cos(k
n
L) sin(k
n
L)
(1) sin(k
n
L) + cos(k
n
L)(0)
k
n
L for small k
n
58 CHAPTER 4. SIMPLE PROBLEMS IN 1 DIMENSION
With this in mind, in the vicinity of the resonance, (19) becomes:

C
A
+

1
1 +
_
kn
2k

n
_
2
(k
n
)
2
L
2
(21)
In order to get the width at the half-max point of the peak, we set the above to 1/2. This
implies that:
k
n
L =
2k

n
k
n
1 for deep wells
And so we can dene a half-width in k by
k
n
..
kkn
=
2k

n
k
n
L
(22)
Remark 11: We can convert to the half-width in energy by substituting (3) into (22):
_
2m(E V
1
) k
n
=
2k

n
k
n
L
_
2m(E V
1
) + 2m(E
n
V
1
) k
n
=
2k

n
k
n
L
_
2m(E
n
V
1
)
. .
kn

_
1 +
1
2
E E
n
E
n
V
1
_
k
n

2k

n
k
n
L
where the term in parenthesis is the rst terms of the Taylor expansion, which we can use
since we are in a deep well and the ratio is very small. Now, if we let E
n
= E E
n
, we
have:
E
n

4k

n
k
n
L
(E
n
V
1
)
k
n

4
_
2m(E
n
V
2
)(E
n
V
1
)
L 2m(E
n
V
1
)

_
8
2
(E
n
V
2
)
mL
2
(23)
Remark 12: Eq (23) is valid only for deep potentials.
4.5. SQUARE POTENTIAL BARRIER AND TUNNELING 59
4.5 Square Potential Barrier and Tunneling
Consider the following potential
V
0 L
V = V0
The methods from 2.2-2.4 can be applied here, though we dont do them explicitly here.
The new feature here is the tunneling eect, which can occur when 0 < E < V
0
. Classically,
an incoming particle will be reected at the barrier. Quantum Mechanicaly, a wave function can
experience barrier penetration. In some cases, this penetration is far enough that the wave can
emerge from the other side, as seen in Fig. 4.2.
Figure 4.2: An incident wave tunneling through a step barrier.
Thus, the quantum particle can tunnel through the potential.
Remark 1: The size of the transmitted wave is determined by the height of the potential (which
sets the decay rate for the exponential solution), and the width of the potential.
4.6 General Properties of 1-D Schrodinger Equation
Here, we summarize some of the results of 2.1-2.5.
Remark 1: Here we are considering the time independent Schrodinger equation in 1-D with one
restriction on the potential, V (x):
V (x) is a real function, bounded from below and piece-wise continuous over
the real line.
Remark 2: The following properties can be shown mathematically in Messiah. Here, we just state
the results. These should be MEMORIZED!
60 CHAPTER 4. SIMPLE PROBLEMS IN 1 DIMENSION
4.6.1 Asymptotic Behavior of Solutions
Here, lets state the results for x ; similar results hold for x . Choose x
0
so that either
1. E V (x) > 0 x > x
0
E
x0
V (x)
2. E V (x) < 0 x > x
0
E
x0
V (x)
Case 1: Solutions to the Schrodinger equation oscillate and are bounded as x .
If lim
x
V (x) = V
+
< , (or V (x) < ), and if it does so faster than
1
x
,
then the asymptotic solution of the equation is:

x
A
+
sin(kx +
+
), where k =
_
2m(E V
+
)
Case 2: For E < V
min
< V (x), and for x > x
0
, one solution decays faster than
e
x
, where =
_
2m(V
min
E)
All other solutions diverge at least as fast as e
x
, and are eliminated on the
basis of physical realizability. (2.3 - Remark 4)
4.6.2 Nature of Energy Eigenvalues
Assume that
V (x)
x
V
+
, V (x)
x
V

, V
+
> V

which will look something like:


V+
V
Then there are 3 energy ranges for which the following properties hold:
4.6. GENERAL PROPERTIES OF 1-D SCHRODINGER EQUATION 61
1. E > V
+
: Energy spectrum is continuous and degenerate (there can be left
or right travelling waves). The states themselves are unbounded.
2. V

< E < V
+
: Spectrum is continuous, but non-degenerate (waves can only
come in from the left and reect back to the left). States are unbounded.
3. E < V

: If solutions exist, they have discrete energies and are bounded


( 0 as x ). Spectrum is non-degenerate.
Remark 1: Non-degeneracy is not a general property of bound states. In multi-
dimensions, we can have bound, degenerate states.
62 CHAPTER 4. SIMPLE PROBLEMS IN 1 DIMENSION
Chapter 5
Wave Packets
In the previous chapter, we looked at a variety of 1-D problems, and particularly at solutions with
waves extended from to . Physically, such waves dont exist(though reection, transmission,
barrier penetration and tunneling do occur). Instead, we have wave packets.
5.1 Plane Waves
The simplest wave with frequency and wavelength is a plane wave:
(r, t) e
i(tkr)
where = 2 and k =
2

.
Remark 1: The wave propagates in the direction of

k.
Denition 1: The phase velocity of this wave is:
v
phase
=

k

k =

k
This gives the speed at which planes of equal phase move in the direction

k. As a visual
conrmation, consider a normal set of plane waves and a phase shifted set, as shown below.
0 2 4
Normal:
0 2
Phase Shifted:
63
64 CHAPTER 5. WAVE PACKETS
We can see that the speed of the plane will be

T
= =

k
in both cases.
Denition 2: The period, T, is T =
1

.
Remark 2: We can not really associate plane waves with the motion of a particle (at least, not
in the classical sense). For this, we need to use wave packets.
5.2 Wave Packets
First, consider the motion of a classical particle. Here, the velocity is given by
v
class
=
p
m
p =
2E
p
; Since E =
p
2
2m
for free particles
Remark 1: Quantum mechanical particles, like light, possess wave-particle duality,
in which they can exhibit properties of both waves and particles, though the
properties can not be simultaneously observed.
For a quantum particle with energy E, we associate a wave-frequency by
E =
Remark 2: de Broglie developed the theory of matter waves based on the assumption
that particles have wave properties.
For a particle with momentum p, we assign a wave number k by
p = k = k

k
This is called the de Broglie relation. Since we can identify a wave with a particle, we
can also identify a phase speed:
v
phase
=

k

k =
E
p

k =
E
p
p
which diers from the classical case by a factor of 2.
To nd a relation between de Broglies matter waves and the speed of a classical
particle, consider
= E =
p
2
2m
=

2
k
2
2m
5.2. WAVE PACKETS 65
which gives
=
k
2
2m
=

2m
(k
2
x
+k
2
y
+k
2
z
) ()
And

k
=

k
x
x +

k
y
y +

k
z
z
=

m
(k
x
x +k
y
y +k
z
z)
=
p
m

k
= v
class
Or, in 1-D

k
=
k
m
=
p
m
= v
class
Remark 3: A dispersion relation is given by the frequency as a function of k: (k).
Remark 4: Here, we can obtain the classical velocity from the dispersion relation.
To associate a wave with a particle, we need waves of limited extent (not plane
waves, but wave packets). Wave packets can be constructed as a superposition
of plane waves and can be expressed as
(r, t) =
_

d
3
k

f(k

)e
i(

tk

r)
()
Remark 5: f(k) gives the weighting of the superimposed plane waves. It is gen-
erally complex, so it denes the amplitudes and the phases of the waves that
are added together.
Remark 6: Wave packets that are localized in space(like ()), move as a whole at
the group velocity.
v
group
=
k

=k
where k is the value of k

where f(k

) is peaked. The group velocity corresponds


to the classical velocity of a particle.
5.2.1 Spacial Extent of Wave Packets
For now, let us consider a 1-D wave packet.
(x, t) =
_

dk

f(k

)e
i(tk

x)
(1)
66 CHAPTER 5. WAVE PACKETS
At time t=0:
(x, 0) =
_

dk

f(k

)e
ik

x
(2)
A generic form of [f(k

)[ is:
k

[f(k

)[
k
k
If we let k

= k + k

,
(x, 0) = e
ikx
_

dk

f(k + k

)e
ik

x
(3)
The above is akin to shifting the peak to the origin in the k

integration.
Now, consider (3) for dierent values of x:
x=0:
(0, 0) =
_

dk

f(k + k

) ,= 0
If f(k

) is real, this is just the area under the curve, and thus ,= 0.
x =
100
k
: f(k

) is only signicantly dierent from 0 for the range of k

given by k < k

< k. This implies that the phase of the oscillary


term inside (3) varies over the range 100 < k

x < 100. This range


covers about 30 cycles of 2, so that the real part of f(k +k

)e
ik

x
looks
like:
The imaginary part looks similar with oscillating phase shifted by . When
the integral (3) is done over this function, it is very close to zero, which
results in a limited spacial extent of the packet.
5.2. WAVE PACKETS 67
Remark 1: x = 0 gives the largest contribution to the wave function ((x, 0)), and,
for x
1
k
, we get almost no contribution. For ranges

k
x

k
, there is
a small, but negligible contribution to (x, 0). Take x =

k
:
k

And integral (3) doesnt quite vanish because small values of x imply slower oscillations
which give bigger contributions, while large values of x imply faster oscillations,
which contribute nearly zero. This behavior gives a modulus wavefunction that
looks like:
x
[(x, 0)[
0
x

k
Remark 2: This acts as an envelope function for e
ikx
, multiplying the integral in (3).
So the wave function is a rapidly oscillating function with peaks at x = 0 and is
only signicantly dierent from zero in the region

k
x < x < x

k
.
Remark 3: For instance, a classical particle:
m = 1 kg, v = 10 m/s, x 10
15
m
where the value of x is approximately the size of an atomic nucleus. Now,
mv = k = 10 kg m /s
k = 10
35
m
1
=
2
k
10
24
m
The range, k of k needed to achieve this localization (x) is
k

x
10
15
m
1
68 CHAPTER 5. WAVE PACKETS
k 10
15
m
10
34
Number of Oscillations = 10
19

k
k
10
20
This implies that we need an accuracy to 21 decimal places if we want more
than a single wave number to contribute to the wave packet. And thus a
classical particle has a well dened position and momentum.
Remark 4: Now, for an electron:
m = 9 10
31
kg, E = 1 eV = 1.6 10
19
J
Consider k 10
10
m, which is about the size of an atom.
k =

2mE k 2 10
11
m
1
= =
2
k
= 3 10
11
m
Here, the range needed to achieve localization is k

x
3 10
10
m
1
.

k
k
0.1
And so a wave packet living in x 10
10
is made up of many wave numbers
and does not have a well dened momentum.
k 2 10
11
m
1
3 10
10
m
1
Figure 5.1: K Space representation of the wave packet.
5.2. WAVE PACKETS 69
10
34
x = 10
10
m
Figure 5.2: Real Space representation of the wave packet.
5.2.2 Motion of the Wave Packet
For a wave packet that represents a particle in the classical limit, the group velocity
(as dened in 5.2, Remark 6) corresponds to the classical velocity of a particle.
Consider (1) of the previous section.
(x, t) =
_

dk

f(k

)e
i(

tk

x)
with k

k + k

and

(x, t) = e
i(tkx)
_

dk

f(k + k

)e
i(

tk

x)
Following the same arguments, we can demonstrate that there is a localized wave
packet, but, if t ,= 0, the wave packet has moved.
Consider the phase =

t k

x. The position x is given as the position that


causes to change the slowest as k

varies over the relevant range.


d
dk

=0
= 0
0 =
d

dk

=0
t x
And this gives exactly the group velocity of the wave:
v
group
=
x
t
=
d

dk

=0
=
d

dk

=0
70 CHAPTER 5. WAVE PACKETS
5.2.3 Delay of Wave Packets at Resonance
Consider reection and transmission of a wave packet at a square potential for energies
E > V
0
:
0
V0
What is the delay in the appearance of the transmitted pulse?
The incident wave packet is:

i
(x, t) =
_
dk

A
+
(k

)e
i(

tk

x)
()
where A
+
(k

) is centered at k

n
resonance.
The transmitted wave packet is:

T
(x, t) =
_
dk

A
+
(k

)
C(k

)
A
+
(k

)
e
i(

tk

x)
From Ch 4, Section 2, Eqn (13):
C
A
+
=
e
ik

L
cos(kL)
i(k
2
+(k

)
2
)
2kk

sin(kL)

T
(x, t) =
_
dk

A
+
(k

)e
i(

tk

x)
cos
2
(kL) +
(k
2
+(k

)
2
)
2
(2kk

)
2
sin
2
(kL)
()
where
(k

) = k

L + arctan
_
k
2
+ (k

)
2
2kk

tan(kL)
_
(Here we just have written the complex number C/A
+
in its polar form). For sharp
resonances, k

/k 1, and in the vicinity of the resonance we have


(k

) k

n
L + arctan
_
k
n
2k

n
tan ((k k
n
)L)
_
5.2. WAVE PACKETS 71
k

n
=
Resonance Width
k

(k

)
To gure out the time delay, we note that the phase in the integrals () and () must be
minimized for the maximum contribution to the incident and transmitted waves. The
incident wave simply comes in at the group velocity (5.2.2), while the transmitted
wave has a delay.

d
dk

t k

x )

=k

n
=
d

dk

=k

n
t x
d
dk

=k

n
= 0
= v
group
t x
d
dk

=k

n
= 0
where
d
dk



Res. Width
by a rise over run argument. Then
x =
k
m
t +
d
dk

=k

n
=
k
m
(t +)
where is the time delay at resonance.
=
m
k
d
dk

=k

m
k

Res. Width
Remark 1: In order for this picture to hold, we assume that the envelope function
of the wave packet hold together so that there is a well dened peak across
the barrier. To ensure this, the wave must be broad enough (in x), so that the
delay is small compared to the motion of the packet:

_
k
m
_
1
k
72 CHAPTER 5. WAVE PACKETS
Chapter 6
Harmonic Oscillator and Second
Quantization
The harmonic oscillator (H.O.) is an important example in QM. A major application of H.O. is
QED, where elementary excitations (modes) of the electromagnetic eld are formally harmonic
oscillations. This also leads to the generalization of QED, which is quantum eld theory. Another
major application covers all areas where there are harmonic oscillations about some equilibrium
(like vibrations of molecules in crystals).
6.1 Harmonic Oscillator in Position Representation
The Hamiltonian for a single oscillator is

H =

T +

V =
p
2
2m
+
1
2
m
2
x
2
(1)
where =
_
k/m is the classical frequency.
Remark 1: A particle that is experiencing small oscillations about a stable equilibrium of potential
V(x) may be approximated by (1). If x
0
is a stable equilibrium about V(x), then
V (x) V (x
0
) +
dV
dx

x
0
. .
0
(x x
0
) +
1
2
d
2
V
dx
2

x
0
(x x
0
)
2
+
For small oscillators,
V (x)
1
2
d
2
V
dx
2

x=x
0
(x x
0
)
2

1
2
d
2
V
dx
2

0
x
2
=
1
2
m
2
x
2
with m
2
=
d
2
V
dx
2
[
0
.
73
74 CHAPTER 6. HARMONIC OSCILLATOR AND SECOND QUANTIZATION
Remark 2: For an n-oscillator system, the Hamiltonian may be diagonalized to a system of n
decoupled oscillators, as was done in Problem 1.8.10 for 2 oscillators. In such a system, the
Hamiltonian may be written as:

H =
N

i=1
N

j=1
P
i
P
j

ij
2m
+
1
2
N

i=1
N

j=1
x
i
V
ij
x
j
with
V
ij
=

2
V
x
i
x
j

0
=

2
V
x
j
x
i

0
= V
ji
The time dependent solution of the quantized HO can be captured by acting the propagator (3.4.1)
on the time-independent solution. So our goal here is to solve the time-independent Schrodinger
equation in the coordinate basis:

H[E
n
) = E
n
[E
n
)
where

H is given in (1). In position representation:
x[

H[E
n
)
_


2
2m
d
2
dx
2
+
1
2
m
2
x
2
_

E
(x)
x[ E
n
[E
n
) E
E
(x)
So the time independent eigenvalue equation is:
d
2
dx
2

E
+
2m

2
_
E
1
2
m
2
x
2
_

E
= 0 (2)
Following the procedure is Shankar:
1. Non-dimensionalize (2). Let
y =
_
m

x and =
E

(3)
Then
d
dx

_
m

d
dy
And thus, under the change of coordinates, (2) becomes
d
2
dy
2

E
+ (2 y
2
)
E
= 0 (4)
Remark 3: The scaling denes a natural length scale in the problem

m
6.1. HARMONIC OSCILLATOR IN POSITION REPRESENTATION 75
2. Investigate the asymptotic behavior.
Consider lim
y0
and lim
y
.
lim
y
(4)
d
2

dy
2
y
2
= 0 (5)
The solution as y then is:
= Ay
m
e
y
2
/2
(6)
lim
y0
(4)
d
2

dy
2
+ 2 = 0 (7)
This has a solution of the form:
= Acos(
_
2y) +Bsin(
_
2y) (8)
Remark 4: In Eq (7), we have already neglected terms of order y
2
, so we should
only keep terms with order < y
2
in (8).
A +B
_
2y +O(y
2
) (9)
3. Make an ansatz of a solution that satises the asymptotic behavior. Note that an
ansatz is like an educated guess as to the solution.
(y) = u(y)e
y
2
/2
(10)
Now, substitute (10) into (4) to derive a DE for u(y):
d
2

dy
2
=
d
dy
_
du
dy
e
y
2
/2
uye
y
2
/2
_
=
d
2
u
dy
2
e
y
2
/2
2
du
dy
ye
y
2
/2
+uy
2
e
y
2
/2
ue
y
2
/2
= (2 y
2
)ue
y
2
/2
by (4)

d
2
u
dy
2
2y
du
dy
+ (2 1)u = 0 (11)
4. Solve (11); Try a power series solution. Lets try:
u(y) =

n=1
C
n
y
2
(12)
76 CHAPTER 6. HARMONIC OSCILLATOR AND SECOND QUANTIZATION
Substituting (12) into (11) yields:
0 =

n=0
C
n
_
n(n 1)y
n2
2ny
n
+ (2 1)y
n

n=0
C
n
_
n(n 1)y
n2
+ (2n + 2 1)y
n

n=2
C
n
n(n 1)y
n2
+

n=0
C
n
(2 1 2n)y
n
=

n=0
C
n+2
(n + 2)(n + 1)y
n
+

n=0
C
n
(2 1 2n)y
n
=

n=0
[C
n+2
(n + 2)(n + 1) +C
n
(2 1 2n)] y
n
Since the ys are linearly independent, the coecients must vanish and this results
in a recursion relation for the coecients:
C
n+2
=
C
n
(2 1 2n)
(n + 2)(n + 1)
(13)
So given C
0
and C
1
, (13) gives all the rest of the coecients in (12), and thus the
solution to (10)...and thus the solution to our non-dimensionalized Schrodinger
equation (4).
Remark 5: The solution given by (3),(10),(12),and (13) doesnt seem to constrain the energy
eigenvalues as expected. However, recall the asymptotic solution as y (See (6)). For a
nite wavefunction, we take the negative in the exponential, and thus u(y)
y
y
m
. But
the solution (12) does not terminate at nite n unless C
n
= 0 for some n. The recursion
relation (13) has C
n+2
vanish for arbitrary C
n
when
= n +
1
2
(14)
So physically realizable solutions only occur at discretized energy eigenvalues. From (3)
and (14):
E = (n +
1
2
), n = 0, 1, 2, . . . (15)
Remark 6: Here we have only positive (or 0) values of n since we can easily show that the energy
eigenvalues can not be negative.
_

H
_
=
1
2m
[ P
2
[)
. .
positive
+
1
2
m
2
[ x
2
[)
. .
positive
0
6.2. SECOND QUANTIZATION: CREATION AND ANNIHILATION OPERATORS 77
Remark 7: For n odd, let C
0
= 0, and for n even, let C
1
= 0. Then the solution to (4) is
(y) = u(y)e
y
2
/2
= e
y
2
/2
_
C
0
+C
2
y
2
+C
4
y
4
+ +C
n
y
n
, n = even
C
1
y +C
3
y
3
+ +C
n
y
n
, n = odd
. .
These are Hermite Polynomials,Hn(y)
Remark 8: Normalizing (15) to get C
0
or C
1
gives the nal solution

E
(x)
(n+1/2)
(x)
n
(x)
=
_
m
2
2n
(n!)
2
_
1/2
exp
_
mx
2
2
_
H
n
_
_
m

_
1/2
x
_
(17)
6.2 Second Quantization: Creation and Annihilation Operators
6.2.1 Harmonic Oscillator in Energy Eigenbasis
Remark 1: Here we solve the HO using an entirely dierent approach using creation
and annihilation operators. This method is due to Dirac and is central to QED.
One advantage is that we do not have to solve any dierential equationswe
can extract the energy eigenvalues from the underlying operator algebra. In
this approach, we can also extract the physical content from the Hilbert space
and the operators that act on it, rather than from the wave functions obtained
from the solution of the Schrodinger equation.
Denition 1: The Creation and Annihilation operators are given by the following:
a

=
_
m
2
_
q
i
m
p
_
a =
_
m
2
_
q +
i
m
p
_
respectively, where q is the position operator. The reason for the names will
become obvious shortly.
Remark 2: The commutation relation for a and a

are:
[ a, a] = 0 = [ a

, a

]
78 CHAPTER 6. HARMONIC OSCILLATOR AND SECOND QUANTIZATION
[ a, a

] =
m
2
_
q +
i
m
p, q
i
m
p
_
=
m
2
_
[ q, q] +
1
m
2

2
[ p, p] +
i
m
[ p, q]
i
m
[ q, p]
_
=
m
2
_
_

2i
m
[ q, p]
..
i
_
_
= 1
Remark 3: we can obtain the operators p and q in terms of a and a

:
q =
_

2m
( a + a

)
p = i
_
m
2
( a a

)
Therefore, we can write the Hamiltonian as:

H =
1
2m
_

m
2
_
( a a a a

a a

+ a

)+
+
1
2
m
2
_

2m
_
( a a + a a

+ a

a + a

)
=
1
2
( a a

+ a

a)
=
_

N +
1
2
_
()
Where we dene the number operator

N a

a.
Remark 4: With the Hamiltonian given in the form (), the problem is reduced to
nding eigenvalues of

N.
To determine the eigenvalues of

N, we can denote the eigenstates of

N:

N [) = [)
and establish a relationship between the eigenstates.
6.2. SECOND QUANTIZATION: CREATION AND ANNIHILATION OPERATORS 79
Remark 5: If [) is an eigenstate of

N, so is a [) (an eigenstate).

N a [) = a

a a [)
= ( a a

1) a [)
= a( a

a
..

N
) [) a [)
= a [) a [)
= ( 1) a [)
with eigenvalue ( 1). Similarly, a

is an eigenstate of

N with eigenvalue
( + 1).

N a

[) = ( + 1) a

[)
Remark 6: The eigenstates [) , a [) , a

[) of

N are also eigenstates of

H with
energy eigenvalues:
_
+
1
2
_
for [)
_

1
2
_
for a [)
_
+
3
2
_
for a

[)
Repeated application of a or a

generates a whole series of energy eigenstates.


Eg:
E-States E-Values
.
.
.
.
.
.
a
2
[)
_

3
2
_

a [)
_

1
2
_

[)
_
+
1
2
_

[)
_
+
3
2
_

[)
_
+
5
2
_

.
.
.
.
.
.
Remark 7: Eigenstates a
n
[) will have negative energies if n . This would
correspond to energies below the minimum HO potential, which is physically
not allowed. To x this, we require to be an integer, say n. Then

N( a
n
[)) = ( n)( a
n
[)) = 0 for = n
80 CHAPTER 6. HARMONIC OSCILLATOR AND SECOND QUANTIZATION
and thus

H( a
n
[)) =
1
2
( a
n
[))
where
1
2
is the energy for the a
n
[) state. But that means that

H( a
n+1
[)) =
1
2
( a
n+1
[))
where we have negative energywhich we just argued was physically impossible.
How is this possible? Let us calculate the norm of a
n+1
[). To see how this
works, well go back to the wavefunction language. Thus, we have a [) a.
a[ a) =
m
2
_
(q +

m

q
)

(q +

m

q
)
_
=
m
2
_
dq
__
q +

m

q
_

___
q +

m

q
_

_
=
m
2
_
dq

_
q

m

q
__
q +

m

q
_
by int. by parts
=
_

a
_
=
_


N
_
=
_

N
_
Hence, the norm of the state a is the expectation of

N in the state .
Therefore, the norm of a( a
n
[ = n) equals the expectation value of

N in the state
a
n
[ = n). But this equals zero, since a
n
[ = n) is an eigenstate of

N with eigenvalue
equal to 0 (as shown above). Thus we conclude that the state a
n+1
[ = n) is identically
0. Therefore, as long as is an integer, the states with negative energy eigenvalues
are not generated (So our s must by integer to be physical).
Remark 8: To summarize: The eigenstates of

N are represented by [0) , [1) , [2) , . . . , [n) , . . .
with eigenvalues 0, 1, 2, . . . , n, . . .. These states are also energy eigenstates with
energy eigenvalues
1
2
,
3
2
, . . . , (n +
1
2
), . . ..
Remark 9: We can see why

N is the number operator: Its eigenvalue counts the
number of energy quanta in a given eigenstate. Similarly, we can see why
a and a

are the annihilation and creation operators. From Remark 7:


a [n) has norm = n[ a

a [n) = n[

N [n)
6.2. SECOND QUANTIZATION: CREATION AND ANNIHILATION OPERATORS 81
and
a
2
[n) has norm = n[ ( a

)
2
a
2
[n) = n[ a


N a [n)
..
E-state w/
E-value n1
= (n 1) n[ a

a [n)
= nn 1[

N [n 1)
Thus we deduce that
a [n) = C

[n 1)
a

[n) = C
+
[n + 1)
with C

and C
+
constants. We can see that the a removes (annihilates) one quanta of
energy while a

adds (creates) one quanta of energy-and hence their respective names.


To nd C
+
and C

, we take n[n) = 1 and use:


n[ a

a [n) = n[

N [n) = nn[n) = n
= n 1[ C

[n 1)
= [C

[
2
n 1[n 1) = [C

[
2
C
n
=

n
So that
a [n) =

n[n 1)
To nd C
+
:
a

[n) = C
+
[n + 1)
a a

[n) = aC
+
[n + 1)
a a

[n) = C
+

n + 1 [n)
( a

a + 1) [n) = C
+

n + 1 [n)
(n + 1) [n) = C
+

n + 1 [n)
C
+
=

n + 1
So that
a

[n) =

n + 1 [n + 1)
Remark 10: we can construct any eigenstate of [n) from the ground state.
[n) =
( a

)
n

n!
[0)
82 CHAPTER 6. HARMONIC OSCILLATOR AND SECOND QUANTIZATION
6.2.2 Eigenfunctions in the Position Representation
We can apply the above method to obtain the eigenfunctions in the position represen-
tation without solving the Schrodinger equation. We have
a [0) = 0
which, in position representation, corresponds to
a
0

_
q +

m
d
dq
_

0
= 0

d
0
dq
=
m

q
0
_
d
0

0
=
_

m

q dq
ln(
0
) =
m
2
q
2
+C

0
= Ae

m
2
q
2
where A is a normalization constant.
1 =
_

0
= A
2
_
e

q
2
dq = A
2
_
2
2m
A =
_
m

_
1/4
Giving the ground state wave-function:

0
(q) =
_
m

_
1/4
e

m
2
q
2
The excited states can be obtained by repeatedly applying the creation operator:
[n) =
1

n!
( a

)
n
[0)
which, in the position representation, corresponds to:

n
(q) =
_
m

_
1/4
1

n!
_
m
2
_
n/2
_
q

m
d
dq
_
n
e

m
2
q
2
This is equivalent to:

n
(q) =
_
m

_
1/4
1

n!
1
2
n/2
e

m
2
q
2
H
n
__
m

q
_
where H
n
is a Hermite polynomial.
6.3. COHERENT STATES: MINIMUM UNCERTAINTY WAVE PACKETS 83
6.3 Coherent States: Minimum Uncertainty Wave Packets
6.3.1 Uncertainty Relations in Energy Eigenstates
Lets calculate the Heisenberg uncertainty relations in the energy representation. From
6.2, Remark 3:
n[ q [n) =
_

2m
n[ ( a + a

) [n)
=
_

2m
n[ (

n[n 1) +

n + 1 [n + 1))
= 0
Similarly,
n[ p [n) = 0
Remark 1: This result is expected from the symmetry of the stationary HO wave-
function
n
(q):
0(q)
1(q)
2(q)
The uncertainty in position is given by:
n[ q
2
[n) =

2m
n[ ( a
2
+ a a

+ a

a + ( a

)
2
) [n)
=

2m
n[
_
_
n(n 1) [n 1) + (n + 1) [n) +n[n) +
_
(n + 1)(n + 2) [n + 2)
_
=

2m
(2n + 1)
84 CHAPTER 6. HARMONIC OSCILLATOR AND SECOND QUANTIZATION
The uncertainty in momentum is given by:
n[ p
2
[n) =
m
2
(2n + 1)
Remark 2: We can rewrite these as
n[ q
2
[n) =
E
n
m
2
, n[ p
2
[n) = mE
n
xp =
_
n[ q
2
[n)
_
n[ p
2
[n) =
_
E
2
n

2
=

2
(2n + 1)
So the minimum in uncertainty according to the Heisenberg relations is /2.
This is satised by the ground stateall other states are not minimum uncer-
tainty states.
6.3.2 Minimum Uncertainty States
It is possible to construct non-stationary minimum uncertainty states. It turns out
that these states are eigenstates of the annihilation operator.
a [) = [)
where is a complex number.
Remark 1: We must make sure these states exist. Expand the state [) in terms of
the stationary states [n).
[) =

n=0
C
n
[n)
a [) =

n=1
C
n

n[n 1) =

n=0
C
n+1

n + 1 [n) =

n=0
C
n
[n)
C
n+1

n + 1 = C
n
C
n
=

n

n!
C
0
which gives
[) = C
0

n=0

n!
[n)
and C
0
is obtained by normalizing:
1 = [) =

n
[C
n
[
2
= [C
0
[
2

n=0
[[
2n
n!
= [C
0
[
2
e
||
2
6.3. COHERENT STATES: MINIMUM UNCERTAINTY WAVE PACKETS 85
[C
0
[
2
= e
||
2
C
0
= e
||
2
/2
which has an arbitrary phase chosen to be 0. Thus we have
[) = e
||
2
/2

n=0

n!
[n)
which is known as the coherent state with complex amplitude .
Remark 2: These states were rst constructed by Schrodinger as minimum un-
certainty wavepackets. They are now used extensively in quantum optics
coherent states are important in describing optical coherence.
Remark 3: We can show that [) is indeed a minimum uncertainty state (See HW!)
6.3.3 Physical Meaning of the Coherent State
Let us relate the coherent state [) to the ground state:
[) = e
||
2
/2

n=0

n!
[n)
= e
||
2
/2

n=0

n!
( a

)
n

n!
[0)
= e
||
2
/2
e
a

[0)
Consider now [0) = e

a
[0) where [0) is the only constant term in the exponential
expansion that is not zero. Then
[) = e
||
2
/2
e
a

a
[0) ()
Now we want to make use of the following relationship. Let

A and

B be operators.
Then [

A,

B] is their commutation relation. If

A and

B commute with [

A,

B], eg
[

A, [

A,

B]] = [

B, [

A,

B]] = 0
Then
e

A+

B
= e

A
e

B
e
[

A,

B]/2
So let

A = a

and

B =

a. Thus we have:
a

a] = [[
2
[ a, a

] = [[
2
e
a

a
= e
a

a
e
||
2
/2
86 CHAPTER 6. HARMONIC OSCILLATOR AND SECOND QUANTIZATION
and thus
[) = e
a

a
[0) =

T() [0)
where

T() = e
a

a
is called the displacement operator.
To see what this means, assume = x (a real number) for simplicity. Then

D(x) = e
x( a

a)
= e
x
q
2
m
( pi)
= e
x
q
2
m
d
dq
= e
Q
d
dq
where Q = x
_
2
m
is a number with dimensions of length. (Which implies and x
are dimensionless. The displacement operator acting on an arbitrary function of q is:

DF(q) = e
Q
d
dq
F(q) =
_
1 Q
d
dq
+
1
2
Q
2
d
2
dq
2
+
_
F(q)
= F(q Q) for small Q
Remark 1: So the action of

D on an arbitrary function is to displace the function
by an amount Q (and hence the name).
Remark 2: Expressing [) =

D() [0) in the coordinate representation:

=x
(q)
. .
Wavefunction of
coherent state
where |=|x
= e
Q
d
dq
. .

D() in coord.
representation
with =x

n=0
(q)
. .
Wavefunction of
HO in state |0
= e
Q
d
dq
_
m

_
1/4
e

m
2
q
2
=
_
m

_
1/4
e

m
2
(qQ)
2
So the coherent state has the same gaussian wavefunction as the HO ground state, it
is just displaced a distance Q = x
_
2
m
.
6.3. COHERENT STATES: MINIMUM UNCERTAINTY WAVE PACKETS 87
Q
q
0(q)
0
Remark 3: The coherent state is not stationary (it evolves with time according to
the Schrodinger equation). The formal solution is given by
[(t)) = e

Ht
[)
..
Init. State
= exp
_

_
a

a +
1
2
_
t
_
[)
= exp
_
i
_
a

a +
1
2
_
t
_
exp
_

1
2
[[
2
_

n=0

n!
[n)
= exp
_

[[
2
2
_

n=0
exp
_
i
_
n +
1
2
_
t
_

n

n!
[n)
= e

1
2
(it+||
2
)

n=0
_
e
it
_
n

n!
[n)
= e
it/2
. .
overall arbitrary
phase can be
dropped

e
it

_
Which gives the evolution of q(t)) and p(t)):
q(t)) =

e
it

e
it

_
=
_

2m
_

e
it

e
it

_
+

e
it

e
it

_
_
=
_

2m
_
e
it
+e
it

_
=
_

2m
(cos(t)( +

) i sin(t)(

))
And so at t = 0
q(0)) =
_

2m
( +

)
88 CHAPTER 6. HARMONIC OSCILLATOR AND SECOND QUANTIZATION
And we can show that
p(0)) = i
_
m
2
(

)
Thus we have that
q(t)) = q(0)) cos(t) +
1
m
p(0)) sin(t)
And similarly
p(t)) = p(0)) cos(t) m q(0)) sin(t)
which are the solutions to the Heisenberg equations of motion for the operators q and
p.
Remark 4: The time dependence of the coherent state can be represented as a
trajectory in the complex plane.
And we see that the circular trajectory in the complex plane is related to the phase
space trajectory of classical Hamiltonian mechanics.
Remark 5: Consider the coherent state wave packet in coordinate representation.
From Remark 2, we can see q) = Q (the mean position), while all coherent
states have q =
_

2m
(Remark 3 - HW). And the wave packet moves in a
harmonic potential without changing its width. The coherent state of the HO
6.3. COHERENT STATES: MINIMUM UNCERTAINTY WAVE PACKETS 89
is a non-dispersing wavepacket. This is signicant since we can recall that free
particle wave-packets always disperse (Ch. 4, 4.1).
Remark 6: We can calculate the explicit time dependence of the wave packet. Re-
mark 3 implies:
[(t)) = e
i

Ht
[) =

e
it

_
=

D(e
it
) [0)
where

D(e
it
) = exp
_
e
it
a

e
it

a
_
In the coordinate representation, the time dependent coherent state is represented by
the wave function:

e
it

(q) =

D(e
it
)
. .
Displacement
Operator
_
m

_
1/4
e

m
2
q
2
To nd what the Displacement operator actually is:

D = e
it
a

e
it

= e
it

_
m
2
_
q
i
m
p
_
e
it

_
m
2
_
q +
i
m
p
_
= q
_
m
2
_
e
it
e
it

_
+ p
_
i

2m
_
_
e
it
+

e
it
_
= exp
_
i

( q p(t)) p q(t)))
_


D(e
it
) = e
i

( q p(t) p q(t))
= e
i

q p(t)
e

p q(t)
e

1
2
2
p(t) q(t)
i
..
[ q, p]
= e

i
2
p(t) q(t)
e
i

q p(t)
e

p q(t)
= e
i
2
q(t) p(t)
e
i

q p(t)
e
q(t)

q
The last term is the displacement operator that we had when we let =

. If we
have an arbitrary , then the displacement operator generalizes and acquires time
dependent phase factors.
The coordinate representation of the time dependent coherent state becomes:

e
it

(q) = e
i
2
q(t) p(t)
. .
overall phase
changes with time
and is indep of q
_
m

_
1/4
e
i

q p(t)
. .
this phase carries
info about mean
momentum
e

m
2
( q q(t))
2
. .
Gaussian envelope
carries info about
mean position and
q, p
90 CHAPTER 6. HARMONIC OSCILLATOR AND SECOND QUANTIZATION
In the last comment, we can nd p by knowing that [) is a minimum uncertainty
and knowing q. The related picture is a gaussian wavepacket that oscillates back
and forth along the classical HO trajectory with a xed width.
If we want to represent this by the coherent state trajectory in the complex plane, the
width of the gaussian can be related to the radius of a small circle.
p
x
Im((t))
Re((t))
t
Remark 7: There is another class of minimum uncertainty states which are closely
related to coherent states. These are known as Squeeze States.
As minimum uncertainty states, the squeeze states satisfy qp = /2, but they
dier from coherent states since they have either:
q <
_

2m
and p >
_
m
2
or
p <
_
m
2
and q >
_

2m
6.3. COHERENT STATES: MINIMUM UNCERTAINTY WAVE PACKETS 91
while preserving the minimum uncertainty. In this case, the uncertainty in p or the
uncertainty in q is less than it is for a coherent state.
To visualize the squeezed states, dene quadrature phase operators:
x
1
=
1
2
( a + a

) =
_
m
2
q
x
2
=
i
2
( a a

) =
_
1
2m
p
In a coherent state, these operators have uncertainties x
1
= x
2
=
1
2
. In a squeeze
state, x
1
<
1
2
or x
2
<
1
2
such that x
1
x
2
=
1
4
. Graphically, the minimum
uncertainty states lie along the line x
1
x
2
=
1
4
.
x2
x1
1
2
1
2
Coherent State
All other states along the line are squeeze states.
A complex plot, in this case, can be represented by a plot on the x
1
x
2
plane. In
this picture, the squeeze states are represented as:
x1 x1
x2 x2
x1 - Quadrature Squeezed x2 - Quadrature Squeezed
The time evolution of a squeeze state is analogous to a rotation of frequency in the
complex plane (in this case, the x
1
x
2
plane). This is seen in Fig. 6.1.
The wave packet in coordinate representation has a gaussian envelope whose width
changes as it oscillates in the harmonic potential, as can be seen in Fig. 6.2
More generally, the coordinate representation is squeezed twice during the cycle,
though not necessarily at the end or middle of an oscillation. Eg: See Fig. 6.3
92 CHAPTER 6. HARMONIC OSCILLATOR AND SECOND QUANTIZATION
Figure 6.1: The initial squeezing is in x
1
, which evolves in time to a squeezing in x
2
.
Figure 6.2: The outer gaussians are squeezed in position, while the center gaussian is squeezed in
momentum.
Figure 6.3: Example of squeeze cycles not fully squeezed on the main axes.
Chapter 7
Systems with N Degrees of Freedom
7.1 Tensor Products (Direct Products)
Denition 1: Let [
1
) be a basis vector of the Hilbert space V
1
so that the set of all [
1
) spans
V
1
. Similarly, let the set of all [
2
) be basis vectors that span V
2
. Then [
1
) [
2
) forms a
basis vector in the combined Hilbert space V
1
V
2
.
Remark 1: Physically, V
1
and V
2
may correspond to Hilbert spaces for 2 individual particles
(1 and 2), while V
1
V
2
represents the Hilbert space for the combined 2 particle system.
Alternatively, this may represent single particle system that spans 2 (or more) dimensions.
Remark 2: Consider an operator

1
that operates on the space V
1
with eigenstates [
1
) and
eigenvalues
1
. Eg:

1
[
1
) =
1
[
1
)
In the combined system, we write:

1
[
1
) [
2
)

(1)(2)
1
[
1
) [
2
)

(1)
1


I
(2)
[
1
) [
2
)

(1)
1
[
1
)

I
(2)
[
2
)

(1)
1

1
_
[
2
)
where the superscripts denote the target Hilbert space (either V
1
or V
2
) on which the
operator acts, and where

(1)
1

1
_
=

(1)
1
[
1
). Similarly, we can write

(2)
2

(1)(2)
2


I
(1)

(2)
2
So that

(2)
2
[
1
) [
2
) = [
1
)

(2)
2

2
_
93
94 CHAPTER 7. SYSTEMS WITH N DEGREES OF FREEDOM
The bottom line is that

(i)
i
only acts on the basis vectors [
i
) which span V
i
.
Remark 3: It is common to use a compact representation of the vectors which span the combined
space.
[
1
) [
2
) [
1
) [
2
) [
1

2
)
Since this also represents the simultaneous eigenkets of operators

(1)
1
and

(2)
2
.
Remark 4: The inner product of the direct products
_

_
([
1
) [
2
)) =

1
_

2
_
Similarly,
_

_
_

(1)
1

(2)
2
_
([
1
) [
2
)) =

(1)
1
[
1
)

(2)
2
[
2
)
Remark 5: Properties of the Tensor Product of Operators:
Denote

(1)
1


I
(2)

1
acts only on [
1
)

I
(2)

(2)
2

2
acts only on [
2
)

1
,

2
_
= 0

(1)
1

(2)
2
__

(1)
1

(2)
2
_
=
_

_
(1)
1

_
(2)
2
Squares of sums:
_

1
+

2
_
2
=
_

(1)(2)
1
+

(1)(2)
2
_
2
=
_

2
1
_
(1)


I
(2)
+

I
(1)

2
2
_
(2)
+ 2

(1)
1

(2)
2
Remark 6: The time evolution of the state vectors that are composed of direct products is still
governed by the Hamiltonian:

H =
p
2
1
2m
1
+
p
2
2
2m
2
+

V ( x
1
, x
2
)
for a two particle system in 1 dimension. If the Hamiltonian is separable:

H =

H
1
+

H
2


V ( x
1
, x
2
) =

V
1
( x
1
) +

V
2
( x
2
)
7.2. IDENTICAL PARTICLES 95
then the two particles will evolve independently of one another, giving the time evolution
as
[(t)) = [E
1
) e
iE
1
t/
[E
2
) e
iE
2
t/
which is obtained by solving the individual Schrodinger equations (with E = E
1
+ E
2
).
Separation of variables is often employed in solving the Schrodinger equations in position
representation. If the Hamiltonian is not separable (ie.

V ( x
1
, x
2
) ,=

V
1
( x
1
) +

V
2
( x
2
)),
then there is not a general method for solving the Schrodinger equations unless

V ( x
1
, x
2
) =

V ([ x
1
x
2
[), in which case the problem can be separated into center of mass plus the motion
about the center of mass (see Central Potential, coming soon!)
7.2 Identical Particles
Classically, two identical particles can be distinguished by fallowing their trajectories (ie. by looking
at their non-identical histories), which can be done without disturbing the system. Quantum me-
chanically, no such trajectory exists for particles, and thus there is no physical basis for distinguish-
ing 2 or more particles. This implies that the same state vector must describe two congurations
which dier only in particle exchange.
7.2.1 Symmetric and Anti-symmetric States: Bosons and Fermions
Consider the following state vector:
[) = [x
1
= a, x
2
= b) [ab) ()
which represents the result of a measurement that nds a particle at a, and one at
b. If the particles are distinguishable, then the state () is distinct from the state
[) = [x
1
= b, x
2
= a) [ba). However, if the particles are indistinguishable, then we
can not dierentiate one state from another. Instead, we write the state vector as a
superposition of these possible states. (These will dene a 2-D degenerate eigenspace.)
[(a, b)) = [ab) + [ba)
These two states are physically equivalent, so we can also write
[(a, b)) = [(b, a))
where is an arbitrary phase (complex number).
[ab) + [ba) = ([ba) + [ab))
= , =
=
2
= 1
Taking the +, we dene the symmetric wavefunction:
[(a, b))
S
= [ab) +[ba)
96 CHAPTER 7. SYSTEMS WITH N DEGREES OF FREEDOM
and the - gives the anti-symmetric wavefunction:
[(a, b))
A
= [ab) [ba)
Denition 1: Particles whose wavefunctions are symmetric are called Bosons, and
those with anti-symmetric wavefunctions are called Fermions.
Remark 1: One of the primary consequences of the anti-symmetric wavefunction is
the Pauli Exclusion Principle:
Suppose we have a 2 fermion system with the following state vector:
[(
1
,
2
))
A
= [
1

2
) [
2

1
)
so that one fermion is in state
1
, and the other is in state
2
. If
1
=
2

[(
1
,
2
))
A
= 0, which implies that no two fermions can be in the same
quantum state.
Consider the 2 particle Hilbert space V
12
made up of all vectors of the form [
1
)
[
2
). Since each pair of vectors [
1
= a,
2
= b) and [
1
= b,
2
= a) can form one
symmetric and one anti-symmetric state, if follows that V
12
can be decomposed into
symmetric and anti-symmetric parts:
V
12
= V
S
..
Bosonic
Hilbert
Space
V
A
..
Fermionic
Hilbert
Space
This leads to an ambiguity in determing whether the state of the system is a member
of the symmetric Hilbert space (V
S
) or the anti-symmetric Hilbert space (V
A
). (We
will see in 2.2 that the measurement made depends on which case applies.) To remove
the ambiguity, we introduce the Symmetrization Postulate, which states:
The state of a system containing N identical particles is either all symmetric
or all anti-symmetric with respect to permutation (particle exchange) of the
N particles
The Corollary to this is that the symmetry (or anti-symmetry) will be independent
of the number of particles.
Remark 2: The normalized states for a system of n identical bosons is given by:
[1, 2, . . . , n)
S
=
1

n!
([1, 2, . . . , n) +[2, 1, 3, . . . , n) + +n! permutations)
where 1, 2, . . . , n are generic state labels.
7.2. IDENTICAL PARTICLES 97
Remark 3: The normalized states for a system of n identical fermions is given by
the Slate Determinant:

1,2,...,n
(x
1
, x
2
, . . . , x
n
) =
1

n!

1
(x
1
)
2
(x
1
)
n
(x
1
)

1
(x
2
)
.
.
.
.
.
.
.
.
.

1
(x
n
)
n
(x
n
)

The determinant gives all the correct minus signs in the anti-symmetric wave-
function, and vanishes if any two js or x
j
s are the same.
7.2.2 Distinguishing Fermions and Bosons
The are two ways to identify whether a particle is a fermion or a boson:
1. Spin (intrinsic angular momentum)
2. Experiments
We will discuss spin in much more detail later, though we note here that bosons have
integral spin (0, , 2, . . .), and fermions have half-integral spin (

2
,
3
2
, . . .).
To determine the nature of the particle experimentally, place two identical particles
in a 1-D (for simplicity) box. Let

S
(x
1
, x
2
)

A
(x
1
, x
2
)
_
The 2 particle wave function for bosons and fermions
In general, each particle could have dierent energy levels, so

S,A
=
1

n
(x
1
)
m
(x
2
)
n
(x
2
)
m
(x
1
)
where n,m are energy levels and the + is for the symmetric cases and the - for
the anti-symmetric cases. Also, n ,= m (they dont have the same energy).
The probability distribution in x-space is determined by
P
S,A
(x
1
, x
2
) = 2 [
S,A
(x
1
, x
2
)[
2
()
Remark 1: The factor of 2 in () is a consequence of the normalization condition.
We require:
1 =

dx
1
dx
2
[
S,A
(x
1
, x
2
)[
2
and accounting for the double counting in the x
1
x
2
integral:
1 =

1
2
P
S,A
(x
1
, x
2
) dx
1
dx
2
98 CHAPTER 7. SYSTEMS WITH N DEGREES OF FREEDOM
where
P
S,A
(x
1
, x
2
) =

x
1
x
2
[
S,A
)
S,A

2
=

2
S,A
(x
1
, x
2
)
where the

2 in for convenience.
Taking ():
P
S,A
(x
1
, x
2
) = 2

2
(
n
(x
1
)
m
(x
2
)
n
(x
2
)
m
(x
1
)

2
= [
n
(x
1
)
m
(x
2
)[
2
+[
m
(x
1
)
n
(x
2
)[
2

[
n
(x
1
)

m
(x
1
)
m
(x
2
)

n
(x
2
) +

n
(x
1
)
m
(x
1
)

m
(x
2
)
n
(x
2
)[
Remark 2: So the dierenence between fermiens and bosons is detectable in the in-
terferecence in the probability distribution. In the extreme cases, lets consider
the limit that these particle have as x
1
x
2
x. Then
P
S
(x
1
x, x
2
x) 2
_
[
n
(x)[
2
[
m
(x)[
2
+[
n
(x)[
2
[
m
(x)[
2
_
P
A
(x
1
x, x
2
x) 0
Where the antisymmetric case is consistant with the Pauli exclusion principle.
Thus, two fermions will show probability 0 of being in the same state, while
two bosons double the probability density for two distinct particles to be in
the same state.
Chapter 8
Classical Limit and WKB
Approximation
Correspondence Principle: In the limit that 0, laws of quantum mechanics must reduce
to the laws of classical mechanics.
Remark 1: gives a measure of the energy spacing between discrete energy levels. Thus, in order
to justisfy the limit as 0, the system must have large quantum numbers.
Remark 2: The limit as 0 is called the classical approximation, and the conditions of validity
are the same as those of geometrical optics.
Remark 3: Note that in the classical limit, we have:
[q
i
, p
i
] = i
ij
0
q
i
p
i


2
0. This implies
q
i
=
_

q
2
i
_
q
i
)
2
which is the same as ignoring uxuations about the mean.
In order to ignore the eects of uxuations, we require that:
The mean values approximately follow classic laws of motion,
Dimensions of the wave packets are small (and remain small as the system
evolves) compared to the characteristic dimensions of the problem.
Remark 4: Since wavepackets disperse with time, the classical approximation is, in general, only
valid over a nite time interval.
Remark 5: The correspondence principle, as stated above, is most useful when applied to com-
mutators, uncertainty, and dierences is energy levels, angular momentum, etc... A second
99
100 CHAPTER 8. CLASSICAL LIMIT AND WKB APPROXIMATION
formulation of the classical approximation basically amounts to:
lim
Geo. Optics
Schrodinger Eqn. Eqns. of Classical Mechanics
In this view, the wavefunction represents a statistical mixture of classical systems. The
density of this mixture, at some point in conguration space, equals the probability density
of a quantum system at that point.
8.1 Ehrenfests Theorem
Consider a system which evolves according to the Schrodinger equation:
i

t
=

H i

t
= (

H)

And the mean value of an observable



A is:
_

A
_
= [

A[)
Since in general evolves in time,
_

A
_
will also evolve in time:
d
dt
_

A
_
=
d
dt
[

A[)
=
d
dt
_
dq
1
. . . dq
n


A
=
_
dq
1
. . . dq
n
_

A +

t
+


A
t

_
=
_
dq
1
. . . dq
n
_

1
i
(

H)


A +


A
_
1
i

H
_
+


A
t

_
=
_
dq
1
. . . dq
n
_

1
i

(

H

A) +
1
i

(

A

H) +


A
t

_
=
1
i
_
_
[

A,

H]
_
+
_


A
t
__
Thus, if

A has no explicit time dependence, then
d
dt
_

A
_
=
1
i
_
[

A,

H]
_
()
Remark 1: An immediate consequence of () is that if [

A,

H] = 0, then

A does not change with
time and is a constant of motion (a conserved quantity). This is analogous to classical
mechanics where we have

A,

H = 0 (The Poisson bracket from 2.1.3).
8.1. EHRENFESTS THEOREM 101
To connect with classical mechanics, apply () to q
i
and p
i
:
d
dt
q
i
) =
1
i
_
[ q
i
,

H]
_
d
dt
p
i
) =
1
i
_
[ p
i
,

H]
_
()
Now consider the q
i
and p
i
commutator with

A( q
1
, q
2
, . . . , q
n
, p
1
, p
2
, . . . , p
n
)
where

A is only well dened when the order in each term is specied. We note that:
_
q
i
,

F( q
1
, q
2
, . . . , q
n
)
_
= 0
_
p
i
,

F( p
1
, p
2
, . . . , p
n
)
_
= 0
_
q
i
,

G( p
1
, p
2
, . . . , p
n
)
_
= i


G
p
i
_
p
i
,

F( q
1
, q
2
, . . . , q
n
)
_
= i


F
q
i
Generalizing the above, one can show that
_
q
i
,

A
_
= i


A
p
i
_
p
i
,

A
_
= i


A
q
i
Using this result in () gives:
d
dt
q
i
) =
_


H
p
i
_
d
dt
p
i
) =
_


H
q
i
_ ()
which are the quantum averages of Hamiltons equations of motion.
Remark 2: Equation () represents Ehrenfests Theorem, which states that equations of motion
for the quantum mean values coordinates and momenta are equivalent to those of classical
mechanics.
Remark 3:
102 CHAPTER 8. CLASSICAL LIMIT AND WKB APPROXIMATION
WARNING! CUIDADO! ACHTUNG!
This is not exactly equivalent to saying that mean values of p and q follow classic
trajectories in phase space. For this, we require that


H
p
i
and


H
q
i
to be at most
linear functions of q and p since:
q
i
p
i
) , = q
i
) p
i
)
_
1
q
2
i
_
,=
1
q
i
)
2
, etc
unless uctuations about the averages can be neglected.
8.2 Classical Limit and Wavepacket Spreading
Though Ehrenfests theorem provides a formal connection between quantum mechanics and classical
mechanics, it doesnt allow us to think of the motion of a quantum wave function as we do the
trajectory of a classical particle.
To a good approximation, we can do this if
1. Quantum means follow the classical trajectories
2. q and p are small on a scale of dimensions of the trajectory.
Using point 2 above to justify the approximate validity of point 1 (they are related), let
q = q) + q
p = p) + p
Then
q p) = ( q) + q) ( p) + p))
= q) p) + q p)
Where the second term is approximately qp. If this term is small compared to q and p, we can
approximately absorb it into q) p). For example, consider:

H =
p
2
2m
+

V ( q)

d
dt
q) =
p)
m
and
d
dt
p) =
_

V ( q)
q
_
Then

V ( q) =

V ( q) + q)


V ( q)) +

V

( q)) q +
1
2

( q)) q
2


V ( q)

V ( q)) +
1
2

( q))

q
2
_
8.2. CLASSICAL LIMIT AND WAVEPACKET SPREADING 103

_
d
d q

V ( q)
_

( q)) +
1
2

( q))

q
2
_
If the

q
2
_
is small enough to neglect, then
d
dt
p)
d
dq

V ( q))
Remark 4: If

q
2
_
starts out small, does it remain small? To answer this, we use equation ()
from 8.1:
d
dt

q
2
_
=
1
i
__
q
2
,

H
__
=
1
i
__
q
2
,
p
2
2m
__
=
1
i
__
( q q))
2
,
p
2
2m
__
.
.
.
=
1
m
( q p + p q) 2 q) p))
Similarly,
d
dt

q
2
_
=
_
1
m
( q p + p q 2 q) p
_

d
2
dt
2

q
2
_
=
1
im
__
q p + p q 2 q) p,

H
__
.
.
.
=
4
m
_
_

H
_

p)
2
2m


V ( q))

V

( q))

q
2
_
_
Lets let
=
_

H
_

p)
2
2m


V ( q))
. .
Classical Energy

d
2
dt
2

q
2
_

4
m
_


V

( q))

q
2
_
_
()
One can use the above to determine the error introduced by replacing
_
d

V
d q
_
by
d
d q

V ( q)) in Ehrenfests Equation


104 CHAPTER 8. CLASSICAL LIMIT AND WKB APPROXIMATION
Remark 5: For a harmonic oscillator with potential

V ( q) =
1
2
m
2
q
2
,
()
d
2
dt
2

q
2
_
=
4
m
4
2

q
2
_
while the equation for the mean is:
d
2
dt
2
q) =
2
q)
which we get from solving Ehrenfests equations for the harmonic oscillator potential. The
two above equations imply that q) oscillates at frequency , while

q
2
_
oscillates at a
frequency of 2.
But this looks just like our squeeze states!
Remark 6: For a free particle,

V ( q) = 0:

d
2
dt
2

q
2
_
=
4
m

=
4
m
_

p
2
_
2m

p)
2
2m
_
=
2
m
2

p
2
_
. .
const.

q
2
_
(t) =

p
2
_
t
2
+
d
dt

q
2
_

t=0
+

q
2
_
(0)
This implies that the wavepacket disperses in time, and thus the classical approximation
remains valid only for a limited time.
8.3 Classical Limit of Schrodinger Equation and the WKB Ap-
proximation
The alternative formulation of the classical limit is to make a connection between the ow of
probability density by the Schrodinger equation and the ow of probability density in classical
statistical mechanics.
8.3. CLASSICAL LIMIT OF S-EQN AND WKB APPROXIMATION 105
8.3.1 Quantum Mechanical Probability Current
The quantum mechanical probability density in quantum mechanics is given by
P(r, t) =

(r, t)(r, t)

_
P(r, t) dr = 1
Now imagine that the probability density represents a particle or uid density. Then,
the evolution of the probability density (according to the Schrodinger equation) may
be thought of as a result of uid ow:
dP
dt
=
P
t
+

J = 0
where

J is the probability current and the whole expression equals 0 so that normal-
ization is conserved.

P
t
=

J
But
P
t
=

t
(

)
=

t
+

t
=
1
i
(

H)

+
1
i

(

H)
=
i

_
(

H)

(

H)
_
()
And so

J must be dened to agree with the right hand side of ().
Remark 1: There is some ambiguity in the denition of the probability current
since satisfying () does not uniquely determine

J. We choose a symmetrized
denition:

J(r) =
1
2
_

p
m
+
_
p
m

_
=
1
2
_
i
m

+
i
m

_
=
i
2m
(

) ()
Remark 2:
p
2m
is chosen so that () looks like velocity times density as in classical
current density.
106 CHAPTER 8. CLASSICAL LIMIT AND WKB APPROXIMATION
So

J, with denition (), gives:


J =
i
2m
_

_
=
i

p
2
2m

p
2
2m

_
=
i

_
p
2
2m
+V (r)
_

_
p
2
2m
+V (r)
_

_
=
i

(

H) (

H)

_
=
P
t
probability ow
8.3.2 Connect Probability Current in Classical Mechanics to Quantum
Mechanics
To formally connect the ow of probability in QM to that in CM, write the wave
function in terms of amplitude and phase variable.
(r, t) = A(r, t)
. .
amplitude
e
iS(r,t)/
. .
phase
Substituting this into the Schrodinger equation to obtain 2-coupled equations:
i

t
=

H =

2
2m

2
+V
i

t
_
Ae
S/
_
=
_

2
2m

2
+V
_
Ae
iS/
.
.
.
Equating real and imaginary parts:
.
.
.

A
t
=
1
m
A S
1
2m
A
2
S
S
t
=
1
2m
(S)
2
V +

2
2m

2
A
_

_
()
From this we can obtain an expression for the equation of motion for the probability
current.
P(r) =

= A(r)
2
8.3. CLASSICAL LIMIT OF S-EQN AND WKB APPROXIMATION 107

P
t
= 2A
A
t
=
2A
m
A S
A
2
m

2
S

t
A
2
=
1
m

_
A
2
S
_
And we can identify the probability current density:

J =
1
m
A
2
S
P
t
=

J
Remark 1: The classical limit here is taking 0 in equation (). This implies:
S
t
=
1
2m
(S)
2
V ()
This gives the ow of probability density in phase space. The S that satis-
es () is the principle function of Hamilton as seen in the Hamilton-Jacobi
formulation of classical mechanics. (See Phys 505!)
Remark 2: For the special case of stationary states (

H = E):
A
t
= 0 and
S
t
= E
and equation () becomes

_
A
2
S
_
= 0
(S)
2
+ 2m(V E) =

2

2
A
A
_
_
_
8.3.3 WKB Approximation
Here, the idea is similar to the classical approximation, but less extreme, since we
keep the dependence to the lowest order. Taking the wavefunction:
(r) = A(r)e
iS(r)/
and write this as
(r) = exp
_
i

W(r)
_
with
W(r) = S(r) +

i
ln (A(r))
Remark 1: A(r) and S(r) still evolve accordingly to () in 8.3.2, but they are no
longer restricted to be real functions.
108 CHAPTER 8. CLASSICAL LIMIT AND WKB APPROXIMATION
The idea now is to expand our W(r) to order
2
and drop higher level terms. Take,
for example, the case of stationary states in 1-D:
d
dx
_
A
2
dS
dx
_
= 0 (1)
_
dS
dx
_
2
+ 2m(V (x) E) =

2
A
d
2
A
dx
2
(2)
(1) 2A
dA
dx
dS
dx
+A
2
d
2
S
dx
2
= 0
2
dA
dx
dS
dx
+A
d
2
S
dx
2
= 0
_
1
A
dA
dx
dx =
1
2
_
d
dx
_
dS
dx
_
dS
dx
dx
ln(A) =
1
2
ln
_
dS
dx
_
+ constant
A = C
_
dS
dx
_
1/2
Substituting this into (2):
_
dS
dx
_
2
= 2m(E V ) +
2
_
dS
dx
_
1/2
d
dx
_

1
2
_
dS
dx
_
3/2
d
2
S
dx
2
_
= 2m(E V ) +
2
_
dS
dx
_
1/2
_
3
4
_
dS
dx
_
5/2
_
d
2
S
dx
2
_
2

1
2
_
dS
dx
_
5/2
d
3
S
dx
3
_
= 2m(E V ) +
2
_
_
3
4
_
d
2
S
dx
2
dS
dx
_
2

1
2
_
d
3
S
dx
3
dS
dx
_
_
_
(3)
Remark 2: Solving (3) exactly is equivalent to solving the original Schrodinger
equation. The WKB approximation enters here: we begin by expanding S in
powers of
2
. Let S = S
0
+ S
1

2
+ . . ., and substitute into (3). Doing this
will generate a system of equations for S
0
, S
1
, . . . by equating powers of
2
.
To Zeroeth Order:
_
dS
0
dx
_
2
= 2m(E V )
To order
2
:
2
dS
0
dx
dS
1
dx
=
3
4
_
d
2
S
0
dx
2
dS
0
dx
_
2

1
2
d
3
S
0
dx
3
dS
0
dx
8.3. CLASSICAL LIMIT OF S-EQN AND WKB APPROXIMATION 109
Remark 3: In general, the expansion of S in powers of
2
does not convergeit is
an asymptotic expansion (initially decrease, and then increase in error), so a
good approximation is obtained if a series is truncated at some nite values.
The simplest approximation is truncation at zero order:
(x) =
_

_
dx
_
2m(E V (x)) +

, E > V (x)
i
_
dx
_
2m(V (x) E) +

, E < V (x)
where
p
m and

are arbitrary constants. Then


A = const
_
dS
dx
_
1/2
A
0
(x) =
_

(2m(E V (x)))
1/4
, E > V (x)

(2m(V (x) E))


1/4
, E < V (x)
where

and

are arbitrary constants. So the solution will be a linear combination


of the plus and minus solutions:
(r) = e
i

W(r)
with W(r) = S(r) +

i
ln (A(r))
gives
For E > V (x):
(x) = (2m(E V (x))
1/4
_
C
+
e
i

R
dx

2m(EV (x))
+C

e
i

R
dx

2m(EV (x))
_
(4)
For E < V (x):
(x) = (2m(V (x) E)
1/4
_
D
+
e
1

R
dx

2m(V (x)E)
+D

e
1

R
dx

2m(V (x)E)
_
(5)
Remark 4: To determine the accuracy of the solutions (4) and (5), consider the term
O(
2
) in the expansion of S. Note that once we have S
0
, we can, in principle,
nd S
1
. In order for this to be a good approximation, we require that

2
S
1

[S
0
[
m
2
[dV/dx[
[2m(E V (x))[
3/2
1
(See Messiah Eqn VI.47 for details) Thus, the potential must change slowly
in space compared to the rate of change of zero approximation for the wave
function. This does not happen when E = V (turning points in the classical
trajectories).
110 CHAPTER 8. CLASSICAL LIMIT AND WKB APPROXIMATION
Remark 5: The solutions (4) and (5) represent oscillatory solutions in classically
allowed regions and exponential decay in classically forbidden regions. These
solutions must still match at the boundaries; however, the solutions are only
asymptotic and not necessarily well dened near the boundaries. If the form of
the wavefunction can be determined unambiguously near the boundary, then
it is possible to connect the two solutions. (See Messiah Ch. 6, 11 for HO for
how to do this in this case)
Chapter 9
Symmetries
In 2.1.4, we stated Nothers Theorem: symmetries in the system correspond to conservation laws.
This holds for whether we are speaking of classical or quantum systems. Here we investigate from
a quantum mechanical viewpoint.
9.1 Translations, Translational Invariance, and Conservation of
Momentum
Classically, if the Hamiltonian is invariant with respect to x x +a, then p
x
is conserved.
Remark 1: In quantum mechanics, we do not have a well dened position or momentum, so we
replace these quantities by their quantum averages:
x x)
p p)
Remark 2: In analogy with classical mechanics, we expect translations to give:
x) x) +
p) p)
_
Translation
_

H
_

_

H
__
Translational Invariance
_

p
_
= 0
_
Conservation Law
9.1.1 Active Translations
Recall 1.2.4: There were 2 types of transformations:
the Active: transformed vectors in Hilbert space
111
112 CHAPTER 9. SYMMETRIES
the Passive: transformed Hilbert space operators
Here, we dene the (innitesimal) translation operator

T() to have the following
action:

T () [) [

)
so that
x) x) +
or that

[ x [

) = [ x[) +
Consider the action of the translation operator on the position eigenstate:

T () [x) = [x +)
which shifts the position to the right by amount . Then, if x[) = (x), we have
x[

) = x[

T () [)
= x[

T ()
_

[y) y[ dy [)
= x[
_

dy [y +) y[)
= x[
_

dy [y) y [)
=
xy
y [)
= x [)
= (x )
Remark 1:

T () [x) = [x +) is not the most general result. In fact, we could write:

T () [x) = e
ig(x)/
. .
space dependent
phase factor
[x +)
This result still gives:
x) x) +
but gives
p) x[

T

() p

T () [x)
= x +[ e
ig(x)/
(i)
d
dx
e
ig(x)/
[x +)
= x +[ e
ig(x)/
_
(i)
i

(x)e
ig(x)/
[x +) + (i)e
ig(x)/

x
[x +)
_
= p) +
_
dg
dx
_
9.1. TRANSLATIONS, TRANSLATIONAL INVARIANCE, AND CONS. OF MOM. 113
For translations, we require that
p) p) g = constant
and we take this constant to be zero (for simplicity).
Denition 1: Translational Invariance is dened by
[

H[) =

[

H[

)
i.e., the expectation of the Hamiltonian is invariant under innitesimal trans-
lation.
Remark 2: According to Nothers Theorem, translational symmetry implies a con-
servation law. (Here, it will be conservation of momentum.)
Proposition 1: Momentum, p, is the generator of innitesimal translations.
Proof: Since is small, we expand

T () in a Taylor series about
to O():

T () =

I +

G
where

T (0) no translations. Then:
x[

T () [x) = x[

) (x )
Expanding both sides to O():
x[

I +

G[) = (x)
d
dx
(x) + x[

G[) = (x)
d
dx
x[

G[) =
d
dx
=
1
i
x[ p [)


G =
1
i
p


T () =

I +

i
p
Proposition 2: Momentum is conserved in a translationally invariant system.
Proof: We begin with the denition of translational invariance:
[

H[) =

[

H[

)
= [

T

()

H

T () [)
= [
_

I

i
p
_

H
_

I +

i
p
_
[)
= [

H[)

i
[ p

H

H p [) +O(
2
)
114 CHAPTER 9. SYMMETRIES
Since we are translationally invariant, O(
2
) = [ [ p,

H] [) =
0. And by Ehrenfests theorem:
_
[ p,

H]
_
= 0
_

p
_
= 0 momentum is conserved
Remark 3: So far, we have considered only innitesimal translations. If we need nite
translations, say x) x) + a, where a , we divide a into N intervals of
a/N and let N . Then we consider innitesimal translations:

T
_
a
N
_
=

I
ia
N
p
and dene our nite translation operator as

T (a) = lim
N
_

T
_
a
N
__
N
= e
ia p/
In coordinate basis:

T (a) e
a
d
dx
So
x[

T (a) [) = x[ e
a
d
dx
[)
= (x) a
d
dx
+
1
2
a
2
d
2

dx
2
+
= (x a)
Remark 4: For consecutive translations

T (a)

T (b) = e
ia p/
e
ib p/
= e
i(a+b) p/
=

T (a +b)
9.1.2 Passive Transformations
Passive translations leave quantum states invariant, whil shifting the coordinate sys-
tem to the left by the amount (assuming position is shifted to the right in the active
case). See Shankar for the complete derivation of the translational operator in passive
version. Briey, the requirement that the position and momentum operators must
obey is:

() x

T () x +

() p

T () p
Expanding

T () in powers of results in:

T () =

I +

i
p
9.1. TRANSLATIONS, TRANSLATIONAL INVARIANCE, AND CONS. OF MOM. 115
which we note is the same expression as in the active case. In the passive transforma-
tion picture, translational invariance requires

()

H

T () =

H
which again requires that the Hamiltonian is invariant under translation. Now we
want to show that this is equivalent to

H
_
x +

I, p
_
=

H ( x, p)
Proposition: For any

( x, p) that can be expanded in a power series, and for and
unitary operator

U:

U =

U,

U

U
_
Proof: Expand

in a power series and consider a typical term
(i.e. p x p). Then, for this term, we have

p x p

U =

U

U

U

U

U

U
Collapsing the expansion, we get:

U,

U

U
_
Corollary:

( x, p)

T

T =

T,

T

T
_
=

_
x +

I, p
_
Proof:

T is unitary. The rest follows from above.
Thus we see that for

=

H, translational invariance implies that

H( x, p) =

H( x +

I, p)
9.1.3 Translations For a Many Particle System
For a system of N particles, the translated many body wave-function becomes:
x
1
, . . . , x
N
[

T () [) = (x
1
, x
2
, . . . , x
N
)
and, to O(),
x
1
, . . . , x
N
[
_

I
i

p
_
[) = (x
1
, . . . , x
N
)
N

i=1

x
i
116 CHAPTER 9. SYMMETRIES


T () =

I
i

i=1
p
i
=

I
i

T
where

T is the total momentum operator.
Remark 1: For many particles, translations generalize to

() x
i

T () = x
i
+

() p
i

T () = p
i
and translational invariance becomes:

()

H( x
1
, x
2
, . . . , x
N
; p
1
, p
2
, . . . , p
N
)

T () =

H( x
1
+

I, x
2
+

I, . . . , x
N
+

I; p
1
, p
2
, . . . , p
N
)
=

H( x
1
, x
2
, . . . , x
N
; p
1
, p
2
, . . . , p
N
)
Remark 2: For a single particle system, translational invariance implies free particle
which implies that V = 0. For a system of particles, translational invari-
ance implies than V = V ( x
i
x
j
)so that the Hamiltonian depends only on
interactions between particles and not on external interactions.
Remark 3: In a translationally invariant system:

(a)

H

T (a) =

H and

T

=

T


H =

T (a)

H

T

(a)

T (a) ,

H
_
= 0

T (a) ,

U(t)
_
= 0
This tells us that a system that starts translationally invariant remains trans-
lationally invariant.
9.2 Time Translational Invariance and Energy Conservation
We expect that homogeneity of time will give the same result if an experiment is repeated at
dierent times.
A system is prepared in an initial state, [
0
) at time t
1
. The Hamiltonian at this time is H
1
, and
the system evolves for a short time .
[(t
1
+)) =

U() [
0
)
= e
i

H
1
/
[
0
)
=
_

I
i

H
1

_
[
0
)
9.3. DISCRETE SYMMETRIES: PARITY AND TIME REVERSAL 117
where

H
1
=

H(t
1
). The experiment is repeated, only system is prepared at time t
2
:
[(t
2
+)) =
_

I
i

H
2

_
[
0
)
where

H
2
=

H(t
2
). Since the outcome does not depend on when the experiment is done, we require
that
[(t
1
+)) = [(t
2
+))


H(t
1
) =

H(t
2
)
This holds for arbitrary t
1
and t
2
, so we can conclude that


H
t
= 0
which implies that

H has no explicit time dependence.
Remark 1: Time translational invariance mean

H is time-independent.
Remark 2: Remember Eherenfests Theorem:
d
dt
_

A
_
=
1
i
__

A,

H
__
+
_


A
t
_
Setting

A =

H and using that
_


H
t
_
= 0, we have that
d
dt
_

H
_
=
1
i
__

H,

H
__
= 0
which implies that we have conservation of energy.
9.3 Discrete Symmetries: Parity and Time Reversal
9.3.1 Parity Invariance
Denition 1: The Parity operator,

, is dened by its action on the position eigen-
state:

[x) = [x)
And it has the following properties:
1.

=

1
Proof :

2
[x) =

[x) = [x)

2
=

I

1
=

I

1
118 CHAPTER 9. SYMMETRIES
2.

has e-values 1, which follows from

2
=

I and

acting on an
eigenstate.
Eigenvalue = 1 Even parity
Eigenvalue = 1 Odd parity
3.

=

(Hermitian), and

=

I (Unitary)
The action of the parity operator on an arbitrary ket is

[) =

[y) y[) dy
=
_

[y) y[) dy
=
_

[y) y[) dy
where the last line is by a simple change of variables. We know that x[) =
(x), and thus from the above we nd that
[x)

[) = (x)
and even or odd parity corresponds to the even or odd nature of the function
(x).
Remark 1:

[p) = [p) (which follows from the denition of p in position represen-
tation.)
Remark 2: The parity operator is the mirror image of a function about the origin.
Remark 3: For passive transformations

x = x

p = p
and parity invariance happens when


H( x, p)

=

H( x, p) =

H( x, p)
which means that

U(t) =

U(t)

and thus parity is preserved over time.


Remark 4: If weak interactions are present,
_

,

H
_
,= 0 and thus parity is not
preserved over time.
9.3. DISCRETE SYMMETRIES: PARITY AND TIME REVERSAL 119
Remark 5: Parity transformations are dierent from space and time translations
(and rotations) since this is a discrete transformation. It may be thought of
as a mirror image of some process.
9.3.2 Time-Reversal
Denition 1: The time-reversed state is dened by
x
R
(t) = x(t) and p
R
(t) = p(t)
(The system is moving in reverse.)
Time-Reversal Invariance (TRI): From an initial state of a system x(0) and p(0),
the system evolves and at time T is at x(T) and p(T). Then we reverse the
system and run for time T. We have TRI when
x(2T) = x(0) and p(2T) = p(0)
Remark 1: TRI exists when you can time evolve the system forward or backward in
time and there is no violation of physical laws.
Remark 2: Quantum mechanically, time reversal is consistent with our

since, in position representation,


x x and p p
under complex conjugation. For example, if we have an initial state (x, 0)
and we time evolve that state for time T:
(x, T) = e
i

HT/
(x, 0)
If we then time-reverse the system for time T:
(x, T)

(x, T) = e
i

H

T/

(x, 0)
We then evolve the system forward again (for time T)
e
i

HT/
e
i

H

T/

(x, 0)
Now, for TRI, we want (x, 2T) = (x, 0) and

H =

H

, the latter of which


happen automatically whenever the Hamiltonian in real (for ever powers of p).
Remark 3: This is another example of discrete symmetry.
120 CHAPTER 9. SYMMETRIES
9.4 Rotations, Rotational Invariance, and Conservation of Angu-
lar Momentum
9.4.1 Rotations
9.4.1.1 Rotations in 2D
Recall in classical mechanics, we dene a counterclockwise rotation of a vector
about the z-axis by:
x = xcos(
0
) y sin(
0
)
y = xsin(
0
) +y cos(
0
)

_
x
y
_
=
_
cos(
0
) sin(
0
)
sin(
0
) cos(
0
)
__
x
y
_
And we identify a rotation matrix
R() =
_
cos() sin()
sin() cos()
_
which also rotates momentum vectors.
In QM, dene the operator

U[R( z)] as the operator which rotates a vector in
Hilbert space:
[)

U[R]
[
R
) =

U[R] [)
Here we abbreviate

U[R( z)]

U[R] and the rotated vector [
R
) must satisfy
x)
R
=
R
[ x[
R
) = x) cos(
0
) y) sin(
0
)
y)
R
=
R
[ y[
R
) = x) sin(
0
) + y) cos(
0
)
where
x) = [ x[)
y) = [ y[)
9.4. ROTATIONS, ROTATIONAL INVARIANCE, AND CONS. OF ANG. MOM. 121
and similar expression hold for the x and y components of momentum.
Remark 1: Operating

U[R] on position eigenkets:

U[R] [x, y) = [xcos(


0
) y sin(
0
), xsin(
0
) +y cos(
0
))
Now consider innitesimal rotations about the z-axis:

U[R( z)] =

I +

z

L
z
i
where

L
z
is the generator of innitesimal rotations. Consider now this action on
a position eigenket:

U[R] [x, y) = [x y
z
, y +x
z
)
From this, we can show that
_
x, y

I +

z

L
z
i

_
= (x +y
z
, y x
z
)
Expanding this to O(
z
):
_
x, y

_
+

z
i
_
x, y

L
z

_
= (x, y) +

x
(
z
y) +

y
(
z
x)

_
x, y

L
z

_
=
_
iy

x
ix

y
_
(x, y)
So in position representation

L
z
x
_
i

y
_
y
_
i

x
_
or, more generally,

L
z
= x p
y
y p
x
()
Remark 2: Putting () into momentum representation and acting on (p
x
, p
y
)
rotates the momentum space wavefunctionso that momentum expecta-
tions are consistant with classical rotations.
Remark 3: The passive version of a rotational transformation has:

[R] x

U[R] = x y
z

[R] y

U[R] = x
z
+ y

[R] p
x

U[R] = p
x
p
y

[R] p
y

U[R] = p
y
+ p
x

z
_

_
()
122 CHAPTER 9. SYMMETRIES
Remark 4: Substitute the innitesimal version of

U[R] (

I +
z

Lz
i
) into () and
get, to O(
z
):
_
x,

L
z
_
= i y
_
y,

L
z
_
= i x
_
p
x
,

L
z
_
= i p
y
_
p
y
,

L
z
_
= i p
y
_


L
z
= x p
y
y p
x
Remark 5: For nite rotations, take N innitesimal rotations of size
0
/N as
N

U[R(
0
z)] = lim
N
_

I
i

0
N

L
z
_
N
= e
i
0

Lz/
Writing

L
z
x
_
i

y
_
y
_
i

x
_
in polar coordinates yields

L
z
i

So

U[R(
0
z)] = e

And
e

(, ) =
_
1
0

+
1
2!

2
0

2
+
_
(, )
= (,
0
)
Remark 6: Two consecutive rotations

U[R(

0
z)]

U[R
0
z)] =

U[R((

0
+
0
) z)]
Remark 7: Physically,

L
z
is the angular momentum operator, which is analo-
gous to the classical denition, and is the generator of innitesimal rota-
tions about the z axis.
Remark 8: If the system is invariant under rotations about the z axis, then

[R]

H( x, p
x
, y, p
y
)

U[R] =

H( x, p
x
, y, p
y
)
9.4. ROTATIONS, ROTATIONAL INVARIANCE, AND CONS. OF ANG. MOM. 123
Expanding this to order
z
for innitesimal rotations gives
_

L
z
,

H
_
= 0
which implies
1.
_

L
z
_
equals a constant, which then implies that angular mo-
mentum is conserved.
2. Outcomes of experiments on rotationally invariant systems will
be the same for diering orientations of the system.
3.

L
z
and

H can be simultaneously diagonalized, and therefore a
common eigenbasis exists
Remark 9: A transformation which consists of a product of translations and ro-
tations will, in general, depend on the order of individual transformations
e.g. translations and rotations do not commute.
9.4.1.2 Rotations in 3D
The results of the previous section may be generalized to include rotations about
all 3 coordinate axes. The corresponding rotation matrices are:
R( z) =
_
_
cos() sin() 0
sin() cos() 0
0 0 1
_
_
R( x) =
_
_
1 0 0
0 cos() sin()
0 sin() cos()
_
_
R( y) =
_
_
cos() 0 sin()
0 1 0
sin() 0 cos()
_
_
The generalization gives the components of angular momentum:

L
x
= y p
z
z p
y

L
y
= z p
x
x p
z

L
z
= x p
y
y p
x
_

_
()
Remark 1: Equatios () are related by cyclic permutations of indices. i.e.
xyz yzx zxy
124 CHAPTER 9. SYMMETRIES
Remark 2: One can consider products of innitesimal translations to derive
commutation relations between components of angular momentum. We
know that
_
_
x
y
z
_
_

U[R(x x)]

_
_
x
y
x
z
z +
x
y
_
_
,
_
_
x
y
z
_
_

U[R(y y)]

_
_
x +
y
z
y
z
y
x
_
_
,
_
_
x
y
z
_
_

U[R(z z)]

_
_
x
z
y
y +
z
x
z
_
_
So now if we consider the following sequence of rotations:
R(
x
x)R(
y
y)R(
x
x)R(
y
y)

_
_
x
y
z
_
_
R(x x)

_
_
x
y
x
z
z +
x
y
_
_
R(y y)

_
_
x +
y
(z +
x
y)
y
x
z
z +
x
y
y
x
_
_
R(x x)

_
_
x +
y
z +
y

x
y
y
x
z +
x
(z +
x
y
y
x)
z +
x
y
y
x
x
(y
x
z)
_
_
R(y y)

_
_
x +
y
z +
y

x
y
y
(z
y
x +
2
x
z)
y +
2
x
y
x

y
x
z
y
x +
2
x
z +
y
(x +
y
z +
y

x
y)
_
_
=
_
_
x +
y

x
y +
2
y
x
y

x
z
y
x

y
x +
2
x
y
z + (
2
x
+
2
y
)z +
x

2
y
y
_
_

_
_
x +
y

x
y
y
x

y
x
z
_
_
Which is equivalent to
_
_
x
y
z
_
_
R(xy z)

_
_
x +
y

x
y
y
x

y
x
z
_
_
From this, we require that the quantum rotations must satisfy

U[R(
y
y)]

U[R(
x
x)]

U[R(
y
y)]

U[R(
x
x)] =

U[R(
x

y
z)] ()
9.4. ROTATIONS, ROTATIONAL INVARIANCE, AND CONS. OF ANG. MOM. 125
Writing

U[R(
x
x)] =

I +

x

L
x
i

U[R(
y
y)] =

I +

y

L
y
i

U[R(
z
z)] =

I +

z

L
z
i
and expanding and matching coecients gives
_

L
x
,

L
y
_
= i

L
z
Similarly, one can nd
_

L
y
,

L
z
_
= i

L
x
_

L
z
,

L
x
_
= i

L
y
Remark 3: The commutation relations are also related by cyclic permutations
of the indices.
Remark 4: Alternative ways to express these commutation relations are

L

L = i

L
which is legitimate since the components dont commute.
Remark 5: Also, introducing the Levi-Civita fully antisymmetric tensor
ijk
where

ijk
=
_

_
0 if i = j, j = k, k = i
1 for even permutations
1 for odd permutations
and
123
= 1, we can write
_

L
i
,

L
j
_
= i
ijk

L
k
where summation over repeated indices is implied.
126 CHAPTER 9. SYMMETRIES
Remark 6: For nite rotations in 3Dlet

=

be an arbitrary axis of rotation.


Then

L

is the generator of innitesimal rotations about that axis and

U[R(

)] = lim
N
_

I +
1
i

L
_
N
= e
i

L/
= e
i

L/
9.4.2 Eigenvalues of Angular Momentum
9.4.2.1 Eigenvalues of

L
z
For a problem that is invariant with respect to rotations about the z-axis,
_

L
z
,

H
_
= 0 there exists a common eigenbasis for

L
z
and

H
To nd that basis, start with the eigenvalue problem for

L
z
:

L
z
[l
z
) = l
z
[l
z
)
In the coordinate basis,

L
z
i

:
i

lz
(, ) = l
z

lz
(, )

lz
(, ) = R()e
ilz/
()
where R() is an arbitrary function of .
Remark 1: For
lz
(, ) to be normalized, we require:
_

0
d
_
2
0
d

lz

lz
= 1
Remark 2: For the normalization requirement, l
z
seems arbitrary. However, it
is not since we require our angular momentum operator to be Hermitian:
_

L
z

2
_
=
_

L
z

1
_

In the coordinate basis:


_

0
_
2
0

1
_
i

2
d d =
__

0
_
2
0

2
_
i

1
d d
_

=
_

0
_
2
0

2
_
i

1
d d
= i
_

0
d
2

2
0

_

0
_
2
0

1
_
i

2
d d
9.4. ROTATIONS, ROTATIONAL INVARIANCE, AND CONS. OF ANG. MOM. 127
Thus, to be Hermitian we want the rst term to vanish, which happens
when (, 0) = (, 2). Combining this with () implies that
1 = e
ilz2/
l
z
= m where m = 0, 1, 2, . . .
Remark 3: Eigenvalues of

L
z
are discrete and mis called the magnetic quantum number.
Remark 4: At this point, R() is arbitrary. However, if we are looking for a
simultaneous eigenbasis for

L
z
and

H, then we use the eigenfunctions of

L
z
and () in the eigenvalue problem for

H. Thus the energy eigenvalues
and eigenfunctions determine R().
Remark 5: It is convenient to introduce the following functions:

m
() =
1

2
e
im
This is the normalized angular part of ():
_
2
0

n
()
m
() d =
nm
9.4.2.2 Solutions to Problems with Azimuthal Symmetry
Once we have determined the eigenfunctions of angular momentum, we can pro-
ceed to solve the eigenvalue equation for

H.
_

L
z
,

H
_
= 0

H is not an explicit function of


V (, ) =

V ()
In cylindrical coordinates, the Schrodinger equation is
_

2
2
_

2

2
+
1

+
1

2
_
+V ()
_

E
(, ) = E
E
(, )
Remark 1: here is equal to the mass, and is used so as to not be confused
with the magnetic quantum number.
Trying a solution of the form:

E,m
(, ) = R
Em
()
m
()
128 CHAPTER 9. SYMMETRIES
which is an eigenfunction of

L
z
and
m
=
1

2
e
im
. Substituting this into the
Schrodinger equation:
_

2
2
_

2

2
+
1


m
2

2
_
+V ()
_
R
Em
() = ER
Em
()
Remark 2: The azimuthally symmetric problem is reduced to solving a 1-D ra-
dial problem. The angular contribution is an eective repulsive potential
that corresponds to the centrifugal force.
9.4.2.3 Eigenvalues of

L
2
and

L
z
Dene

L
2
=

L
2
x
+

L
2
y
+

L
2
z
. One can show that
_

L
2
,

L
i
_
= 0 for all i = x, y, z
Remark 1: We already have shown that if the Hamiltonian is invariant under
rotations about the z-axis, then then
_

H,

L
z
_
= 0 and

L
z
is thus conserved. It
follows that if

H is invariant under arbitrary rotations, then
_

H,

L
i
_
= 0 for all i = x, y, z
and
_

H,

L
2
_
= 0
implies that

L
2
,

L
x
,

L
y
, and

L
z
are conserved.
Remark 2: Since the components of the angular momentum vector do not
commute, we can not construct a common eigenbasis between

H,

L
2
,

L
i
for all i = x, y, z. So we pick a single component, usually

L
z
, and form a
common basis with

H and

L
2
.
Now to nd a common eigenbasis between

L
z
and

L
2
. Let [) be a common
eigenvector so that

L
2
[) = [)

L
z
[) = [)
Remark 3: Recall the HO problem where we dened the annihilation and cre-
ation operators in terms of x, p. Here we dene

L
+
,

L

in terms of

L
x
and

L
y
.
9.4. ROTATIONS, ROTATIONAL INVARIANCE, AND CONS. OF ANG. MOM. 129
Denition 1: Let



L
x
i

L
y
be the raising and lowering operators. We note that
_

L
z
,

L

_
=

L
2
,

L

_
= 0
Remark 4:

L

raise/lower the eigenvalue of



L
z
while leaving the eigenvalues of

L
2
xed.

L
z
_

[)
_
=
_

L
z

_
[)
=

L

L
z

_
[)
=

L

( ) [)
= ( )

L

[)
and

L
2
_

[)
_
=

L

L
2
[)
_
=
_

[)
_
Remark 5: Since

L

raises/lowers the eigenvalue of the z-component of the


angular momentum, we deduce that

[) = C

(, ) [, )
However, given a state [), we can not raise or lower an arbitrary
amount as [l
z
[

l
2
. This implies
_

L
2


L
2
z

_
=
2
=
_

L
2
x
+

L
2
y

_
0

2
Remark 6: Since
2
is bounded by , there must exist a state [
max
) that can
not be raised any further.

L
+
[
max
) = 0
130 CHAPTER 9. SYMMETRIES
Similarly, there must be a [
min
) that can not be lowered.

[
min
) = 0
Starting with the top state and considering the above, we have that

L
+
[
max
) = 0
=
_

L
x
i

L
y
__

L
x
+i

L
y
_
[
max
)
=
_

L
2
x
+

L
2
y
+i
_

L
x

L
y


L
y

L
x
__
[
max
)
=
_

L
2


L
2
z

L
z
_
[
max
)
=
_

2
max

max
_
[
max
)
=
max
(
max
+)
Similarly, starting with the bottom state:

L
+

[
min
) = 0
=
_

L
2


L
2
z
+

L
z
_
[
min
)
=
_

2
min

min
_
[
min
)
=
min
(
min
)
Comparing values of implies that
min
=
max
.
Remark 7: One can also start at the top state and use

L

to get to the bottom


(or use

L
+
to go from [
min
) to [
max
)). In general, this will take k
steps, where k = 0, 1, 2, . . ., so that

max

min
= 2
max
= k

max
=
k
2
where k = 0, 1, 2, . . .
=
2
_
k
2
__
k
2
+ 1
_
Remark 8: Let
max

=
k
2
denote the angular momentum of the state.
Remark 9: Note that in 9.4.2.1, we found the eigenvalues of

L
z
to be m (where
m is an integer), while here we nd that the eigenvalues can be half-
integers. The reason is that here we used only commutation relations and
not the specic form of

L
z
(i

). So, in fact, the result here is valid


for any angular momentum with the same commutation relations without
9.4. ROTATIONS, ROTATIONAL INVARIANCE, AND CONS. OF ANG. MOM. 131
specifying the type of angular momentum. We will study two types of
angular momentum: orbital angular momentum (denoted by

L), which
has only integer eigenvalues, and spin angular momentum (denoted by

S), which can have integer or half-integer eigenvalues. The total angular
momentum,

J =

L+

S, obeys the same commutation relations that were


derived for

L, and hence the eigenvalues derived above hold for


J. For
total angular momentum then:

J
2
[jm) = j(j + 1)
2
[jm) , where j =
1
2
, 1,
3
2
, . . .

J
z
[jm) = m [jm) , where m = j, j 1, . . . , j + 1, j
For

J =

L:

L
2
[lm) = l(l + 1)
2
[lm) , where l = 0, 1, 2, . . .

L
z
[lm) = m [lm) , where m = l, l 1, . . . , l + 1, l
Remark 10: We will discuss rules for addition of angular momentum shortly
for now, we will just use the notion of total angular momentum and
determine general eigenfunctions.
9.4.3 Matrix Representation of Angular Momentum
Finding the matirx representation of

J
2
,

J
x
,

J
y
,

J
z
is sucient for determining the
eigenvectors of angular momentum. We will look at the coordinate representation of
these eigenvectors when we undertake central potential problems.
9.4.3.1 Matrix Elements of

J
2
and

J
z
These are both diagonal and therefore straight forward to compute. For

J
2
we
have:
_
jm

J
2

jm
_
= j(j + 1)
2
, where j = 0,
1
2
, 1,
3
2
, . . .


J
2
=
_
_
_
_
_
_
_
_
_
_
_
0 0
3
4

2
3
4

2
2
2
2
2
2
2
0
.
.
.
_
_
_
_
_
_
_
_
_
_
_
132 CHAPTER 9. SYMMETRIES
And for

J
z
we have:
_
jm

J
z

jm
_
= m


J
z
=
_
_
_
_
_
_
_
_
_
_
_
0 0
1
2

1
2

0
.
.
.
_
_
_
_
_
_
_
_
_
_
_
9.4.3.2 Matrix Elements of

J
x
and

J
y
These are non-diagonal. We will use

J
x
=
1
2
_

J
+
+

J

_
and

J
y
=
1
2i
_

J
+

_
and nd the action of

J

on the state [jm). We know:

[jm) = C

(j, m) [j, m1)


where we need to determine C

(j, m). In dual space, we have


jm[

J

= jm[

J

= C

j, m1[
Consider just C
+
for now. We know that
_
jm


J
+

jm
_
= [C
+
[
2
j, m + 1[j, m + 1) = [C
+
[
2
and
_
jm

J
2


J
2
z

J
z

jm
_
= [C
+
[
2
[C
+
[
2
= j(j + 1)
2
m
2

2
m
2
=
2
(j m)(j +m+ 1)
C
+
=
_
(j m)(j +m+ 1)
Remark 1: In general, C
+
can have an overall phase, which we choose to be
unity.
Similarly, considering C

:
_
jm

J
+

J

jm
_
= [C

[
2
9.4. ROTATIONS, ROTATIONAL INVARIANCE, AND CONS. OF ANG. MOM. 133
We nd that
C

=
_
(j +m)(j m+ 1)

[jm) =
_
(j m)(j m+ 1) [j, m1)
Using this to determine the matrix elements of

J
x
and

J
y
:
_
j

J
x

jm
_
=
1
2
_
j

J
+
+

J

jm
_
=
1
2
_
j

_
(j m)(j +m+ 1)

j, m + 1
_
+
1
2
_
j

_
(j +m)(j m+ 1)

jm
_
=
1
2

_
_
(j m)(j +m+ 1)
j,j

m+1,m
+
_
(j +m)(j m+ 1)
j,j

m1,m

_
Thus

J
x
=
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
0 0
0

2

2
0
0

2
0

2
0

2
0

2
0
0
.
.
.
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
Similarly, for

J
y
we have
_
j

J
y

jm
_
=
1
2i
_
j

J
+

jm
_
=

2i
_
_
(j m)(j +m+ 1)
j,j

m+1,m

_
(j +m)(j m+ 1)
j,j

m1,m

_
Thus

J
y
=
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
0 0
0

2i

2i
0
0

i

2
0

2
0

i

2
0

i

2
0
0
.
.
.
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
Remark 2: Though

J
x
and

J
y
are not diagonal, they are block diagonal (all
matrix elements have
j,j
). This implies that the blocks do not mix upon
multiplication, and we can conclude that:
_

J
(j)
x
,

J
(j)
y
_
= i

J
(j)
z
for j = 0,
1
2
, 1, . . .
134 CHAPTER 9. SYMMETRIES
9.4.4 Finite Rotations
Recall from 9.4.1.2 Remark 6 that nite rotations may be expressed as

U[R()] = e
i

J/
Remark 1: We have all components of

J; however,

J is block diagonal with only
blocks of the same j multiplying. So, schematically, we represent

U[R] by the
following:

U[R]
|jm basis
_
_
_
_
_
_
_
_
D
(0)
0
D
(1/2)
D
(1)
0
.
.
.
_
_
_
_
_
_
_
_
where D
(j)
is the (2j + 1) dimensional block for a given j.
Remark 2: Since a vector [
j
) need only be spanned by 2j+1 vectors [jj) , . . . , [j, j),
we only need D
(j)
[R] to rotate [
j
) (which stays within the subspace V
j
).
Consider
D
(j)
[R()] = e
i

j
(j)
/
=

n=0
1
n!
_
i

_
n _

j
(j)
_
n
()
Remark 3: This series expansion can be truncated at n = 2j.
D
(j)
[R] =
2j

n=0
f
n
()
_

j
(j)
_
n
The terms
_

j
(j)
_
n
for n > 2j can be written as linear combinations of the
rst 2j terms, and the f
n
() represents that combination.
Remark 4: Consider the subspace V
j
spanned by basis vectors [jm) where m =
j, . . . , j. The subspace is identied by the eigenvalue of

J
2
: j(j +1)
2
. Since
_

J
2
,

U[R]
_
= 0
then a rotation of any state spanned in this basis will not change the eigenvalues
and therefore this is called an invariant subspace (invariant under arbitrary
rotations).
9.4. ROTATIONS, ROTATIONAL INVARIANCE, AND CONS. OF ANG. MOM. 135
Remark 5: Furthermore, the invariant subspaces are irreducible, i.e., it does not
contain any other invariant subspace.
Remark 6: The block diagonal matrices, D
(j)
, which make up

U[R] are an irreducible
representation of rotation in the [jm) basis.
Remark 7: Also, since
_

H,

J

_
= 0, each block represented by D
(j)
will have a single
energy eigenvalue E
j
. This implies that all states with a given j are degenerate
in the rotationally invariant problem.
Remark 8: Classically, this degeneracy is because one state can be rotated to an-
other without changing the energy. Quantum mechanically, a rotation com-
bined with

J

is required to take one state to another, but these operators all


commute with

H, so the energy is unchanged under

U[R] and

J

.
136 CHAPTER 9. SYMMETRIES
Chapter 10
Central Potential
10.1 Hamiltonian in Spherical Coordinates
For a particle with mass and momentum p, the Hamiltonian in a spherically symmetric potential,
V (r), is

H =
p
2
2
+

V (r)
which corresponds to a time independent Schrodinger equation:

H(r)
_

2
2

2
+

V (r)
_
(r) = E(r) (1)
In spherical coordinates, this is:
_

2
2
_
1
r
2

r
r
2

r
+
1
r
2
sin()

sin()

+
1
r
2
sin
2
()

2
_
+

V (r)
_

E
(r, , ) = E(r, , )
(2)
Remark 1: Though we can tackle (2) as is, it is useful at this point to note the spherical symmetry
and thus that we have conservation of angular momentum. Consequently, we can express
(1) in terms of a conserved quantity.
Remark 2: Recall that

L = rp = i(r) = angular momentum, and that all three components
of angular momentum commute with the Hamiltonian.
Remark 3: For the radial momentum, we will not use i

r
=
r
r
p since it is not Hermitian.
Instead, we will use the symmetrized form:
p
r
=
1
2
_
r
r
p + p
r
r
_
Remark 4: p
r
commutes with any function of and , and also with l
x
, l
y
, l
z
, but
[ r, p
r
] = i
137
138 CHAPTER 10. CENTRAL POTENTIAL
Proposition:
p
2
= p
2
r
+

L
2
r
2
for r ,= 0
Proof:

L
2
= (r p) (r p)
= r
2
p
2
r
2
p
2
r
by vector identities
So an alternate form of the Hamiltonian for a spherically symmetric system is:

H =
p
2
r
2
+

L
2
2r
2
+

V (r) (3)
Remark 5: Comparing the Schrodinger equation obtained by (3)
_
p
2
r
2
+

L
2
2r
2
+

V (r)
_
(r, , ) = E(r, , ) (4)
with (2) gives us

L
2
=

2
sin
2
()
_
sin()

_
sin()

_
+

2

2
_
(5)
10.2 Solution of Spherically Symmetric Schrodinger Equation
Recall from Chapter 9 that spherical symmetry implies


H is invariant under transformation

H,

L
i
_
= 0 for i = x, y, z

H,

L
2
_
= 0
There exists an eigenbasis common to

H,

L
2
and one

L
i
, usually

L
z
.
The steps to nding a common eigenbasis are:
1. Separate the angular and radial parts of the Schrodinger equation.
2. Solve the eigenvalue problem of

L
z
and

L
2
for the angular part.
3. Solve the full problem to get the radial part.
Remark 1: The particular form of V (r) enters only in Step 3!
10.2. SOLUTION OF SPHERICALLY SYMMETRIC SCHRODINGER EQUATION 139
10.2.1 Solution of the Angular Part
Weve actually done this using the bra-ket notation in previous chapters using:

H[n, l, m) = E
lm
[n, l, m)
and starting by solving the eigenvalue equation for

L
2
and

L
z
. Again, we will solve
the eigenvalue equation for

L
2
and

L
z
, but this time well do it in coordinate repre-
sentation. Following the same strategy as before, well start with

L
+
[l, l) = 0
In coordinate representation:

= e
i
_

i cot()

_
We let [l, l)
l
l
(r, , ) so that we have:
_

+i cot()

l
l
(r, , ) = 0 ()
Remark 1:
l
l
is an eigenfunction of

L
z
with eigenvalue l:

L
z
= i

So let

l
l
(r, , ) = U
l
l
(r, )e
il
Substituting this into () gives:
_

+ cot()l
_
U
l
l
(r, ) = 0 ()
Remark 2: We can ignore the r dependence in (), and we can write
dU
l
l
U
l
l
=
l cos()
sin()
d =
l
sin()
d (sin())
which is satised if
U
l
l
(r, ) = R(r) (sin())
l
Remark 3: R(r) is normalizable, but otherwise arbitraryin actuality, it is deter-
mined by the radial part of the solution.
140 CHAPTER 10. CENTRAL POTENTIAL
Remark 4: Requiring R(r) to be normalizable with respect to r implies that the
angular part must be normalizable with respect to the angles. A function that
satises this requirement is:
Y
l
l
(, ) = (1)
l
_
2l + 1
2
_
1/2
1
2
l
l!
(sin())
l
e
il
()
Remark 5: Applying the lowering operator to this top state

[l, l) = [(l +l)(1)]


1/2
[l, l 1)
=

2l [l, l 1)
Y
l1
l
(, ) =
1

2l
()
_

i cot()

_
. .

Y
l
l
Repeating this gives the spherical harmonics. For m 0 they are
Y
m
l
(, ) = (1)
l
_
(2l + 1)!
4
_
1/2
1
2
l
l!
_
(l +m)!
(2l)!(l m)!
_
1/2
e
im
(sin )
m
d
lm
d(cos )
lm
(sin )
2l
And for m < 0:
Y
m
l
= (1)
m
(Y
m
l
)

These function satisfy the orthonormality condition:


_
Y
m
l

(, )Y
m

l
(, ) d =
l,l

m,m
with d = sin dd
Remark 6: The Spherical-Harmonics are closely related to the Associated Legendre
Polynomials. For P
m
l
with 0 m l, they are:
Y
m
l
(, ) =
_
(2l + 1)(l m)!
4(l +m)!
_
1/2
(1)
m
e
im
P
m
l
(cos )
10.2.2 Solution to the Radial Part
Since the solutions will be simultaneous eigenfunctions of

L
2
and

H, we assume a
solution of the form

nlm
(r, , ) = R
El
(r)Y
m
l
(, )
10.2. SOLUTION OF SPHERICALLY SYMMETRIC SCHRODINGER EQUATION 141
Substituting this into the spherically symmetric Schrodinger equation:
_
p
2
r
2
+

L
2
2r
2
+

V (r)
_
Y
m
l
(, )R(r) = E
lm
Y
m
l
(, )R(r)
_
p
2
r
2
+
l(l + 1)
2
2r
2
+

V (r)
_
Y
m
l
(, )R(r) = E
lm
Y
m
l
(, )R(r)
_
p
2
r
2
+
l(l + 1)
2
2r
2
+

V (r) E
_
R
El
(r) = 0
Since
p
r
= i
1
r

r
r = i
_

r
+
1
r
_
p
2
r
=
2
_
_
_
_
_
_

2
r
2
+
1
r

r
+

r
1
r
..

1
r
2
+
1
r

r
+
1
r
2
_
_
_
_
_
_
=
2
_

2
r
2
+
2
r

r
_
=

2
r
_
r

2
r
2
+

r
+

r
_
=

2
r
_

r
_
r

r
_
+

r
_
=

2
r

2
r
2
r
Thus we have that
_

2
2

2
r
2
r +
l(l + 1)
2
2r
2
+

V (r) E
_
R
El
(r) = 0
To solve this, let U
El
(r) = rR
El
(r):
_


2
2r

2
r
2
+
l(l + 1)
2
2r
3
+

V (r) E
r
_
U(r) = 0
_

2
2

2
r
2
+
l(l + 1)
2
2r
2
+

V (r) E
_
U(r) = 0 ()
Remark 1: () is called the radial equation and closely resembles the 1-D Schrodinger
equation. The dierences are:
0 < r < instead of < x <
142 CHAPTER 10. CENTRAL POTENTIAL
For l ,= 0, there exists a repulsive centrifugal barrier
l(l+1)
2
2r
2
. So the
total eective potential is
V
eff
= V (r) +
l(l + 1)
2
2r
2
Boundary conditions on U(r) are dierent.
Remark 2: As in the 1-D Schrodinger equation, the asymptotic solutions (ie, the
solutions as r ) are exponentials which decay to 0 (E < 0) for xed
discrete values of E (bound state solutions), or they are oscillatory solutions
(E > 0) which are the continuum or unbound states.
Remark 3: The solutions are generally degenerate with respect to the energy eigen-
value:
Energies with discrete spectrum are degenerate with respect to quan-
tum number m, and may be, but not necessarily, degenerate with re-
spect to the angular momentum quantum number l.
Continuous energy spectrum are always innitely degenerate, since for
each E > 0 there are eigenfunctions for all values of l = 0, 1, 2, . . .
and m = 0, 1, 2, . . . , l (assuming that V (r) 0 monotonically as
r ).
Remark 4: We must determine boundary conditions at r = 0 so that the solutions
are physically acceptable. First of all, we require that
D
l
(r) =

2
2
d
2
dr
2
+
l(l + 1)
2
2r
2
+V (r)
to be Hermitian with respect to functions U
El
. Here,
D
l
(r)U
El
= EU
El
is the eigenvalue equation representing the radial equation, ().
10.2. SOLUTION OF SPHERICALLY SYMMETRIC SCHRODINGER EQUATION 143
Remark 5: Requiring that D
l
(r) is Hermitian is equivalent to requiring that p
r
be
Hermitian (Convince yourself of this with Exercise 12.6.3 in Shankar).
[ p
r
[) =
_
d
_

0
r
2
dr

_
i
1
r

r
r
_

=
_
d
_

0
r dr

_
i

r
r
_

=
_
d
_

0
dr rY
m
l

El
_
i

r
rY
m
l
R
El
_
=
_
dY
m
l

Y
m
l
. .
1
_

0
dr (i)rR

El

r
(rR
El
)
= i [rR
El
[
2

0
+i
_

0
dr
_

r
(rR

El
)
_
rR
El
= i [rR
El
[
2

0
+i
_
dY
m
l

Y
m
l
_

0
dr
_

r
(rR

El
)
_
rR
El
= i [rR
El
[
2

0
+[ p
r
[)

So p
r
is Hermitian if
i [rR
El
[
2

0
[U
El
[
2

0
0
Since we require to be normalizable, this does vanish for sure at the upper
limit. For physical solutions then, we have the requirement that
U
El
(r = 0) = 0
Our goal now is to understand how solutions to the radial equation behave at the
origin. To investigate this, lets try an expansion of the form
U
El
(r 0) r
s
_
1 +a
1
r +a
2
r
2
+. . .
_
where s, a
1
, a
2
, . . . are constants and s > 0.
We will assume that if V (r) as r 0, then it does so no faster than
1
r
(the
Coulomb potential). Lets substitute this into the radial equation:
_

2
2
d
2
dr
2
+
l(l + 1)
2
2r
2
+V (r) E
_
_
r
s
+a
1
r
s+1
+a
2
r
s+2
+. . .
_
= 0

2
2
_
s(s 1)r
s2
+a
1
r
s1
(s + 1)s

+
l(l + 1)
2
2
_
r
s2
+a
1
r
s1
_
+V (r)r
s
Er
s
= 0
Keeping only the dominant terms leaves

2
2
_

_
s(s 1)
. .
radial
KE
l(l + 1)
. .
Centrifugal
KE
_

_
r
s2
= 0
144 CHAPTER 10. CENTRAL POTENTIAL
s(s 1) = l(l + 1)
s = l + 1 or s = l
So, near the origin:
U
El
r0
r
l+1
which is the regular solution.
or
U
El
r0
r
l
which is the irregular solution
Remark 6: The irregular solution is rejected since it is not normalizable, and it does
not satisfy the boundary conditions U
El
(0) = 0.
Remark 7: It is clear that the irregular solution must be discarded for l ,= 0, but
it also must be discarded for l = 0. If l = 0, then the corresponding wave
function:

0
= Y
0
0
R
E0
= Y
0
0
1
r
U
E0

1
r
which doesnt satisfy the Schrodinger equation:
_

H E
_

0
=
2k

(r) ,= 0
since
2 1
r
= 4(r).
10.2.3 Examples: Free Particles and Central Square Well Potentials
There are several classic problems that correspond to central potential problems:
Free particle in spherical coordinates: V (r) = V
0
= constant
Isotropic H.O: V (r)/
1
2

2
r
2
Hydrogen atom: V (r) =
e
2
r
3D versions of square wells and step potentials
10.2.3.1 Free Particles and Spherical Bessel Functions
Consider a free particle such that V (r) = V
0
for all r > 0. Then the solution to
the spherically symmetric Schrodinger equation is:

Elm
(r, , ) = R
El
(r)Y
m
l
(, )
And the radial equation is:
_

2
2
d
2
dr
2
+
l(l + 1)
2
2r
2
+V
0
E
_
U
El
(r) = 0
10.2. SOLUTION OF SPHERICALLY SYMMETRIC SCHRODINGER EQUATION 145
with U
El
(r) = rR
El
(r) (See 10.2.2). Consider now the case where E > V
0
. Let
k
2
=
2(EV
0
)

2
.

_
d
2
dr
2
+k
2

l(l + 1)
r
2
_
U
El
= 0
Or, in terms of R
El
:
_
d
2
dr
2
+k
2

l(l + 1)
r
2
_
rR
El
= 0
.
.
.
.
.
.
r
_
d
2
dr
2
+
2
r
d
dr
+k
2

l(l + 1)
r
2
_
R
El
= 0
_
d
2
d(kr)
2
+
2
(kr)
d
d(kr)
+ 1
l(l + 1)
(kr)
2
_
R
El
= 0 ()
Remark 1: () is a dierential equation whose solutions are spherical Bessel
functions.
j
l
(kr) The proper spherical Bessel functions
n
l
(kr) The Neumann functions
h
(+)
r
(kr) The Hankel functions of the 1st kind
h
()
r
(kr) The Hankel functions of the 2nd kind
Remark 2: The general solution is a linear combination of j
l
(kr) and n
l
(kr)
which are dened in terms of ordinary Bessel functions:
j
l
(kr) =
_

2kr
_
1/2
J
l+1/2
(kr)
n
l
(kr) = (1)
l
_

2kr
_
1/2
J
l1/2
(kr)
Remark 3: Near the origin:
j
l
(kr)
kr0
(kr)
l
regular
n
l
(kr)
kr0
(kr)
(l+1)
irregular
Remark 4: The Hankel functions are given by
h
()
l
= n
l
(kr) ij
l
(kr)
146 CHAPTER 10. CENTRAL POTENTIAL
which are both irregular at the origin.
Remark 5: For physical solutions, only regular solutions are allowed in regions
containing the origin. However, the other functions may be important
for matching discontinuities in the potential (Eg. a central square well
problem).
Remark 6: Asymptotically,
j
l
(kr)
kr

1
kr
sin
_
kr
l
2
_
n
l
(kr)
kr

1
kr
cos
_
kr
l
2
_
and
h
()
l
(kr)
kr

1
kr
e
i(krl/2)
Remark 7: Asymptotically, the factor of
1
kr
in j
l
(kr), n
l
(kr), and h
()
l
(kr) gives
the inverse square law behavior expected in a central potential problem.
Remark 8: To lowest order, the spherical Bessel functions are:
j
0
(kr) =
sin(kr)
kr
, j
1
(kr) =
sin(kr)
(kr)
2

cos(kr)
kr
n
0
(kr) =
cos(kr)
kr
, n
1
(kr) =
cos(kr)
(kr)
2
+
sin(kr)
kr
Remark 9: The solution to the spherical Schrodinger equation with V = V
0
and
E > V
0
is

Elm
(r, , ) = j
l
(kr)Y
m
l
(, ) ()
where k
2
=
2(EV
0

2
and the solutions satisfy

Elm

m
r
2
dr d =
2
k
2
(k k

)
l,l

m,m

where
_

0
j
l
(kr)j
l
(k

r)r
2
dr =

2k
2
(k k

)
10.2. SOLUTION OF SPHERICALLY SYMMETRIC SCHRODINGER EQUATION 147
Remark 10: For V (r) = V
0
and E < V
0
, we simply let
k =
_
2(E V
0
)

2
i = i
_
2(V
0
E)

2
so that the analogs in trigonometric functions become analogs of hyper-
bolic functions. In this case, it turns out that h
(+)
l
is the only solution
that doesnt diverge at innity.
h
(+)
l
(ir)
r

e
r
r
whereas all other functions diverge as
e
r
r
.
Remark 11: In Cartesian coordinates, the solution for the free particle is given
in terms of plane waves:

k
e
i

kr
which are eigenfunctions of the energy
p
2
2
and the components of linear
momentum p
x
, p
y
, p
z
. These plane waves form a complete set of functions
which may be used to expand any wavepacket. The spherical standing
wave () in Remark 9 is an alternate complete set.
We can expand the plane waves in terms of the spherical standing waves ():
e
i

kr
= 4

l=0
l

m=l
i
l
j
l
(kr)Y
m
l

k
k
_
Y
m
l
_
r
r
_
where the two terms in parenthesis denote angles and which dene the unit
vectors in the

k and r directions. If

k is along the z-axis, this simplies to
e
i

kr
= e
ikr cos

l=0
(2l + 1)i
l
j
l
(kr)P
l
(cos )
where P
l
(cos ) are the Legendre polynomials. In this case, a particle moving along
the z-axis has no angular momentum in this direction. Thus, Y
m
l
Y
0
l
and the m in
the sum in completed explicitly.
10.2.3.2 Central Square Well
Consider now the central potential dened by
V (r) =
_
V
0
, 0 < r < a
0, r > a
148 CHAPTER 10. CENTRAL POTENTIAL
For E < 0 (Bound case)::
R
El
(r) =
_
Aj
l
(kr), r < a regular at origin
Bh
(+)
l
(ir), r > a nite at innity
Solutions must match at r = a.
Aj
l
(ka) = Bh
(+)
l
(ia)
Akj

l
(ka) = Bh
(+)
l

(ia)i
Recursion relations relate derivatives of spherical Bessel functions to other
spherical Bessel functions. Let us consider the l = 0 case. Then:
Asin(ka)
ka
= B
_
cos(ia)
ia
+i
sin(ia)
ia
_
= B
e
i(ia)
ia
= B
e
a
a
(The B absorbed the i) (1)
and
Ak
_
cos(ka)
ka

sin(ka)
(ka)
2
_
= Bi
_
ie
a
a

e
a
i(a)
2
_
A
_
cos(ka)
sin(ka)
ka
_
= B
_
e
a

e
a
a
_
(2)
Adding equations (1) and (2) yields:
Acos(ka) = Be
a
= Asin(ka)
_

k
_
by (1)
tan(ka) =
k

Solving graphically: We found the bottom line by nding in terms of k


and plugging into
k

:
10.2. SOLUTION OF SPHERICALLY SYMMETRIC SCHRODINGER EQUATION 149
10.2.3.3 The Coulomb Potential: Hydrogen and H-like Atoms (Ions)
Consider the 2 particle coulomb interaction between an electron and a proton, as
in a Hydrogen atom. In the center of mass frame, this becomes a central potential
problem, with a Schrodinger equation:

H =
_


2
2m

2
+
z
1
z
2
e
2
r
_
= E
where m =
m
1
m
2
m
1
+m
2
is the reduced mass and z
1
= +1(protons) and z
2
= 1(electrons).
If one neglects nuclear structure, then this equation also describes other 1 electron
atoms such as He
+
, Li
2+
(z
1
= +2, +3, . . .); it also approximates atoms with a
single electron outside one or more complete shells.
Recall the Bohr model, in which we toake the electron to be distributed uniformly
over a spherical surface of radius r
0
. In this case, the mean potential energy is
given by

V (r) =
e
2
r
0
With this model, we can assume the particle is conned on the scale of r
0
, and
by the Heisenberg uncertainty principle, the uncertainty in momentum is at least
p

r
0
Letting this be the mean value of momentum, the total energy in the ground
state is:
E
g
=
1
2m
_

r
0
_
2

e
2
r
0
The stable conguration is where this is at a minimum:
dE
g
dr
0
=
2
2
2mr
3
0
+
e
2
r
2
0
= 0
150 CHAPTER 10. CENTRAL POTENTIAL
r
0
=

2
me
2
And plugging this into the ground state energy gives:
E
g
=

2
2m
_
me
2

2
_
2
e
2
_
me
2

2
_
=
1
2
me
4

2
Remark 1: r
0
is called the Bohr Radius.
Remark 2: The simple arguments actually give the exact results calculated
from the Schrodinger equation! However, in order to get the excited state
energies, we must actually solve the Schrodinger equation.
Thus, we have a classic central potential problem, so the angular and radial parts
separate and the main problem is solving the radial equation. The angular part
is the same as always, for the radial we have:
_
d
2
dr
2
+
2m

2
_
E +
e
2
r
_

l(l + 1)
r
2
_
U
El
(r) = 0
We are solving for the bound states, and hence are looking in the region where
E < 0. We begin by making the change of variables = r
_

2mE

2
(i.e., = kr
from the free particle solution).

_
_
d
2
d
2
1 +
e
2
E
_

2mE


l(l + 1)

2
_
_
U
El
(r) = 0
Let
=
1
E
_

2mE

2
=
_
2m

2
E
Now, we know that
U
El

l+1
as 0
U
El
e

as
10.2. SOLUTION OF SPHERICALLY SYMMETRIC SCHRODINGER EQUATION 151
Hence, we will let
U
El
=
l+1
e

v
El
()
to explicitly capture the asymptotic behavior. Substituting this into the radial
equation:

d
2
d
2
_

l+1
e

v
El
()
_
=
d
d
_
(l + 1)
l
e

v
El

l+1
e

V
El
+
l+1
e

El
_
= (l + 1)l
l1
e

v
El
(l + 1)
l
e

v
El
+
l
(l + 1)e

El

l
(l + 1)e

v
El
+
l+1
e

v
El

l+1
e

El
+
l
(l + 1)e

El

l+1
e

El
+
l+1
e

El
.
.
.
=
__
l(l + 1)
l1
2(l + 1)
l
+
l+1
_
v
El
+
_
(l + 1)
l

l+1
_
2v

El
+
l+1
v

El
_
e

And the radial equation becomes:


_

l+1
d
2
d
2
+
_
(l + 1)
l

l+1
_
2
d
d
+
_

l+1
2(l + 1) +l(l + 1)
l
_
+
_
1 +
e
2


l(l + 1)

2
_

l+1
_
v
El
= 0

d
2
d
2
+ 2 (l + 1 )
d
d
+
_
2(l + 1) +e
2
_
_
v
El
= 0 ()
We will try solutions of the form:
v
El
() =

k=0
C
k

k
()
Substituting this into ():

k=0
_
k(k 1)C
k
k 1 + 2 (l + 1 ) kC
k

k1
+
_
2(l + 1) +e
2
_
C
k

k
_
= 0

k=0
_
[2(l + 1) +e
2
2k]C
k

k
+ [(k(k 1) + 2k(l + 1)]C
k

k1
_
= 0

k=0
[2(l + 1) +e
2
2k]C
k

k
+

k=1
[k(k 1) + 2k(l + 1)]C
k

k1
= 0

k=0
[2(l + 1) +e
2
2k]C
k

k
+

k=0
[(k + 1)k + 2(k + 1)(l + 1)]C
k+1

k
= 0

k=0
_
[2(l + 1) +e
2
2k]C
k
+ [(k + 1)k + 2(k + 1)(l + 1)]C
k+1
_

k
= 0
152 CHAPTER 10. CENTRAL POTENTIAL

C
k+1
C
k
=
2(l + 1) +e
2
2k
(k + 1)k + 2(k + 1)(l + 1)
=
e
2
+ 2(l + 1 +k)
(k + 1)(2(l + 1) +k)
And hence we see that we get a recursion relation between the successive coef-
cients in the polynomial expansion (so long as we are assuming a power series
solution for R(r)). Consider now again the series solution (). To see how this
behaves for large , consider the form of the radial equation given earlier in this
section:
_
d
2
d
2
1 +
e
2


l(l + 1)

2
_
U
El
= 0 ()
For large , try a solution of the form:
U
El

a
e
b

d
2
d
2
U
El

d
d
_
a
a1
e
b
+b
a
e
b
_
a(a 1)
a2
e
b
+ 2ab
a1
e
b
+b
2

a
e
b

_
a(a 1)
a2
+ 2ab
a1
+b
2

e
b
Keeping the terms up to order
a1
in ():
2ab
a1
+b
2

a
+e
2

a1
= 0
b = 1 a =
e
2
2
So, U
El

e
2
/2
e

dominates in the limit. Now,


U
El
=
l+1
e

v
El
()
with
v
El
() =

k=0
C
k

k
.
But, we also have:
U
El
=
l+1
e

k=0
C
k

e
2
/2
e

k=0
C
k

l1e
2
/2
e
2
In order to avoid a divergent solution, we can force the series to terminate at
nite k by requiring the coecient to vanish at some k. This can happen if we
set the numerator in the recursion relation for successive coecients to 0.
e
2
+ 2(k +l + 1) = 0
10.2. SOLUTION OF SPHERICALLY SYMMETRIC SCHRODINGER EQUATION 153
Since we have that =
_
2m

2
E
(where m = mass), then we can conclude:
E =
2m

2
=
me
4
2
2
(k +l + 1)
2
for k, l = 0, 1, 2, . . .
Remark 3: We dene n = k +l + 1 to be the principle quantum number, and
E =
me
4
2
2
n
2
, for n = 1, 2, 3, . . .
the allowed energies.
Remark 4: Given n = 1, 2, 3, . . .:
l = n k 1 = n 1, n 2, . . . , 1, 0 (The ang. mom. Q #s)
are allowed values for l, and there in degeneracy for dierent l at each n:

l = 0
n1
(2l + 1) = 2
1
2
(n 1)(n) +n = n
2
(# degeneracy)
Remark 5: A Rydberg (R
y
) is a unit of energy that measures the energy levels
of Hydrogen:
R
y
=
me
4
2
2
E
n
=
1
n
2
R
y
Remark 6: Equation (),

d
2
d
2
+ 2 (l + 1 )
d
d
+
_
2(l + 1) +e
2
_
_
v
El
= 0
is the Laplace Equation, which has one solution that is regular at the
origin. That solution can be expressed as a conuent hypergeometric
series:
F
1
(2l + 2 e
2
; 2l + 2; ) =

k=0
(2l + 2 +k e
2
)
(2l + 2 e
2
)
(2l + 1)!
(2l + 1 +k)!

k
k!
where
(z) =
_

0
t
z1
e
t
dt
(n + 1) = n! for integer, positive n
154 CHAPTER 10. CENTRAL POTENTIAL
Remark 7: The nal wave functions are obtained from the recursion relation
for the coecients. Note that the series:
v
El
() =
nl1

k=0
C
k

k
terminates at n l 1. The recursion relation can be solved:
C
k+1
C
k
=
2(l + 1 +) e
2
(k + 1)(2l + 2 +k)
C
k+1
=
2(l + 1 +k n)C
k
(k + 1)(2l + 2 +k)
where we used that fact that e
2
= 2n since n = l +1+k. Letting C
0
= 1
implies:
C
1
2(l + 1 n
2l + 2
C
2
=
2(l + 2 n)
2(2l + 3)
C
1
C
k
=
2
k
k!
(l + 1 n)(l + 2 n) (l +k n)
(2l + 2)(2l + 3) (2l +k + 1)
=
2
k
k!
(1)
k
(n l k)!
(n 1)!
(2l + 1)!
(2l +k + 1)!
Thus,
v
El
() =
nl1

k=0
(1)
k
k!
(n l k)!
(n 1)!
(2l + 1)!
(2l +k + 1)!
(2)
k
which in the Laguerre Polynomial: L
2l+1
nl1
(2).
Remark 8: Finally, we write the normalized wavefunction for the Hydrogen
atom:

nlm
=
2
n
2

(n l 1)!
3(n + 1)!
r
3/2
0
_
2r
nr
0
_
(2)
l
e

L
2l+1
nl1
(2)Y
m
l
(, )
with
=
_
2mE

2
r =
r

_
2m
2
e
4
2
2
n
2
=
rme
2
n
2
=
1
n
r
r
0
where
r
0
=

2
me
2
is the Bohr radius.
Chapter 11
Angular Momentum: Spin
Spin in another concept that completes the angular momentum picture. Loosely, spin is imagined as
angular momentum associated with a particle rotating about its axis (like Earth). However, there
really is no such mechanical analogy. Nevertheless, spin represents an intrinsic angular momentum
of a particle that is not associated with the angular momentum operator

L.
Principle Evidence of Spin: The principle evidence of spin is in the behavior of electrons in
a magnetic eld, where the degeneracy is lifted (Zeemann eect). For example, for l = 1:
where H is the external magnetic eld. Spin is an angular momentum quantity and thus
its corresponding operator behaves much the same as

L (See 9.4).
11.1 Brief Review of Properties of Angular Momentum Operators
From 9.4.2.3, we know that eigenvalues of angular momentum as determined strictly by commu-
tator algebra are integer or
1
2
integer quantum numbers:

J
2
[jm) = j(j + 1)
2
[jm)

J
z
[jm) = m [jm)
155
156 CHAPTER 11. ANGULAR MOMENTUM: SPIN
where j = 0, 1, 2, . . . or j =
1
2
,
3
2
,
5
2
, . . . and m = j, . . . , 0 and

J is the total angular momentum.
The commutation relations hold for all types of angular momentum:
_

J
i
,

J
2
_
= 0
_

J
i
,

J
i
_
= i
ijk

J
k

,

J
z
_
=

J
+
,

J

_
= 2

J
z
where

J

=

J
x
i

J
y
are the raising and lowering operators.

[jm) =
_
(j m)(j m+ 1) [j, m1)
11.2 Spin: Evidence in Atomic Spectroscopy
Let us briey review the physical context in which electron spin was proposed. Consider a multi-
electron atom in a magnetic eld. If we did not know about spin, we would write the Hamiltonian
as:

H =

H
0


H

M
where

H is the externally applied eld and

M =
e
2c

L =
e
2c

z
j=1
( r
j
p
j
) is the magnetic moment
obtained by summing over the z magnetic moments of the individual electrons.
Remark 1: Lets recall that the concept of electron magnetic moment comes from the classical
denition for the magnetic moment in a current loop.
11.2. SPIN: EVIDENCE IN ATOMIC SPECTROSCOPY 157

l
=
1
c
e
2r/v
r
2
_
r p
[r p[
_
=
1
c
e
2r
1
m

r
[r[
p

. .

r
|r|
v

=v sin(90)=v
r
2
_
r p
[r p[
_
=
e
2mc
r p
=
e
2mc

L
Remark 2:

H
0
is invariant under rotations about the nucleus (coulomb interaction depends only
on distance between particles), and therefore commutes with all components of angular
momentum

H
0
,

L
2
,

L
z
form a mutually commuting set of operators. We choose H to be
in the z-direction and form a representation in terms of [nlm) where:

H
0
[nlm) = E
nl
0
[nlm)

L
2
[nlm) =
2
l(l + 1) [nlm)

L
z
[nlm) = m[nlm)
and E
nl
0
energies are 2l + 1 -fold degenerate since they are independent of m. The Hamil-
tonian is written

H =

H
0

e
2c
H

L
z
Since

H commutes with

H
0
,

L
2
, and

L
z
, the [nlm) states are also eigenstates of

H, with
eigenvalues:
E
nl
0

e
2c

. .

B
Hm
where
B
is the Bohr magneton. Thus, in the presence of a magnetic eld, the degeneracy
is lifted:
E
n1
0
:
E
n2
0
:
158 CHAPTER 11. ANGULAR MOMENTUM: SPIN
Remark 3: The above energy splitting was not always observed in atomic spectroscopic exper-
iments. If the number of electrons was odd, the number of splitting was even. Also, the
spacing in energy levels was not always given by
B
H. Given the even number of levels,
it seems there must be
1
2
integer spin values of angular momentum (2j + 1). Since orbital
momentum can not give
1
2
integer angular momentum quantum numbers, it was hypothe-
sized that electrons contain some intrinsic angular momentum that is independent of orbital
angular momentum. This implies electron spin.
Remark 4: The magnetic moment associated with spin is:

s
= g
s
e
2c

S
where

s
= spin magnetic moment

S = spin angular momentum


and g
s
is the factor that produces the correct energy level splitting in spectroscopic exper-
iments; g
s
2.
Remark 5: Since

S is an angular momentum operator, we assume
2
has eigenvalues
2
_
1
2
_ _
1
2
+ 1
_
for S =
1
2
, and that the

S
z
eigenvalues are
1
2
.
Remark 6: Spectroscopy gave physical evidence of spin. Dirac demonstrated mathematically that
electron spin was required to make the wave function covariant.
11.3 Spin Kinematics
11.3.1 Spin Operators, Spinors, and Eigenvalues
Consider an innitesimal rotation of an n-component wave function about the z-axis:
_

2
.
.
.

n
_

_
=
_
_
_
_
_
_
_
_
_

I
(n)

_
i

.
.
.
i

_
. .
(1)

S
z
. .
(2)
_
_
_
_
_
_
_
_
_
_

2
.
.
.

n
_

_
where
1. Corresponds to a physical rotation which assigns points in space to rotated
points
11.3. SPIN KINEMATICS 159
2. Corresponds to taking components of wave functions and transforming
them into linear combinations of other components
Equivalently:

_
=
_

I
i

L
z
+

S
z
_
_
[) =
_

I
i

J
z
_
[)
where

J
z
=

L
z
+

S
z
is the generator of rotations about the z-axis. More generally,

J =

L+

S
is the total angular momentum.
Remark 1:

J and

L obey angular momentum commutation relations (1). Since
_

L,

J
_
= 0 (

L acts on the spacial part of wave functions x,y,z, and



S acts on
components 1, . . . , n), it follows that

S also obeys angular momentum com-
mutation relations.
_

S
i
,

S
j
_
= i
ijk

S
k
Remark 2: In Chapter 9, we found that rotation operators are block diagonal matri-
ces and each block represents a rotation matrix for a specic value of angular
momentum. Since we are assuming that the electron has spin eigenvalues
1
2
,
the relevant block in D
(1/2)
, which implies that the electron spin operators are
given by:

S
x
=
_
0 1
1 0
_
,

S
y
=

2
_
0 i
i 0
_
,

S
z
=

2
_
1 0
0 1
_
Remark 3: Given the spin operators, the electron can be described as a 2-component
wave function:
=
_

+
(x, y, z)

(x, y, z)
_
=
+
_
1
0
_
+

_
0
1
_
where is called a spinor.
Remark 4: This can be represented as
[+)

1
2
,
1
2
_
[)

1
2
,
1
2
_
160 CHAPTER 11. ANGULAR MOMENTUM: SPIN
In an

S
z
basis, [+) points up the z-axis and [) points down. Here,

S
z
in
matrix form is:
_
+[
[
_

S
z
([+) [)) =
_
+[
[
__

2
[+)

2
[)
_
=

2
_
1 0
0 1
_
which is what we said above.
In the spirit of raising and lowering operators:

S
x
=
1
2
_

S
+
+

S

S
y
=
1
2i
_

S
+

_
where

S
+
[+) = 0

S
+
[) = [+)

[+) = [)

[) = 0
These give

S
x
and

S
y
matrices that are consistant with Remark 2.
Remark 5: Note that

S
2
[) =

2
2
_
1 +
1
2
_
[) =
3
4

2
[)

S
z
[) =

2
[)
Remark 6: One important dierence between spin and orbital angular momentum is
that the magnitude of spin is xed whereas the magnitude of orbital angular
momentum can change (as in an external eld).
Remark 7: So far, we are only considering the case where
_

L,

S
_
= 0. In some
instances, this is spin-orbit coupling, where
_

L,

S
_
,= 0, but well worry about
that later (perturbation theory). For now, we assume:

H =

H
o
+

H
s
Since we are taking the spin to evolve separately, the wavefunction separates:
[(t)) = [
o
(t)) [
s
(t))
11.3. SPIN KINEMATICS 161
Remark 8: Generalizing the spin components to an arbitrary direction n, rather
than in the z-direction:
n, [

S [n, )
are the spin states that point up or down the n axis. So if n points in the
direction of (, ):
n
x
= sin cos
n
y
= sin sin
n
z
= cos
And
n

S = n
x

S
x
+n
y

S
y
+n
z

S
z
=

2
_
n
z
n
x
in
y
n
x
+in
y
n
z
_
using Remark 2
=

2
_
cos sin e
i
sine
i
cos
_
Solving the eigenvalue problem gives the following eigenvectors:
[n
+
) =
_
cos
_

2
_
e
i/2
sin
_

2
_
e
i/2
_
, [n

) =
_
sin
_

2
_
e
i/2
cos
_

2
_
e
i/2
_
11.3.2 Pauli Spin Matrices
It is conventional to drop the

2
factor in the spin matrices

S
x
,

S
y
,

S
z
to get the Pauli
Spin Matrices:

S =

2


x
=
_
0 1
1 0
_
,
y
=
_
0 i
i 0
_
,
z
=
_
1 0
0 1
_
Remark 1: The Pauli matrices anti-commute:
[
i
,
j
]
+
=
i

j
+
j

i
= 0 (i ,= j)
Remark 2:
x

y
= i
z
and other cyclic permutations (by the anti-commuting na-
ture).
Remark 3: Tr(
i
= 0
162 CHAPTER 11. ANGULAR MOMENTUM: SPIN
Remark 4:
2
i
=

I. Moreover, (n ) =

I.
Remark 5: [
i
,
j
]
+
= 2
ij

I
Remark 6: [
x
,
y
] = 2i
z
and cyclic permutations
Remark 7: Tr(
i

j
) = 2
ij
. If we choose
0
(
1 0
0 1
), then
Tr (

) = 2
,
; , = 0, x, y, z
This implies that the

s are linearly independent, and thus any matrix M


can be written
M =

where m

=
1
2
Tr (M

)
11.3.3 Rotations
The rotation operator for spin
1
2
is in a simple form. Consider rotations about the
z-axis:

U[R( z)] = e
i

Sz/
= e
i z/2
=

k=0
(i)
2k
(2k!)
_

2
z
2
2
_
k
. .
Even terms
+

k=0
(i)
2k+1
(2k + 1)!
_

2
z
2
2
_
k

z
2
. .
Odd terms
Since
2
z
=

I:
= cos
_

2
_

I i sin
_

2
_

z
For rotations about a general axis, this becomes:

U[R(

)] = cos
_

2
_
i sin
_

2
_

11.4 Spin Dynamics
Consider a stationary electron in a magnetic eld B. The Hamiltonian is given by

H = B = g
s
e
2mc

S B
e
mc

S B
11.4. SPIN DYNAMICS 163
The time evolution for a state is given by
[(t)) =

U(t) [(0))
where the propagator

U(t) = e
i

Ht/
= exp
_

ie
mc

S Bt
_
Remark 1: We see that the propagator closely resembles the rotation operator e
i

S
. This
implies that the propagator rotates the state by an angle =
e
mc
Bt.
Consider now the case where B = B z. Then

U(t) = exp
_
i

eB
mc

S
z
t
_
= e
i
0
zt/2
=
_
e
i
0
t/2
0
0 e
i
0
t/2
_
where
0
=
mB
mc
. Lets let
[(0)) = [n+)
=
_
cos
_

2
_
e
i/2
sin
_

2
_
e
i/2
_
where [n+) indicates a spin up along an arbitrary axis. Then:
[(t)) =

U(t) [(0))
=
_
e
i
0
t/2
0
0 e
i
0
t/2
__
cos
_

2
_
e
i/2
sin
_

2
_
e
i/2
_
=
_
cos
_

2
_
e
i(
0
t)/2
sin
_

2
_
e
i(
0
t)/2
_
implying that the rotation angle changes at a rate
0
.
11.4.1 Spin
1
2
and the Bloch Sphere
Let us consider a geometrical representation for the state of a spin-
1
2
system. This
representation is the work of Felix Bloch and is used extensively in optical physics (a
spin-
1
2
system is isomorphic to a 2-state system and this makes a nice way to study
transitions in atoms and molecules). The most general state of a spin-
1
2
system is a
superposition of up and down states:
[) = C
1
[+) +C
2
[)
where C
1
and C
2
are complex numbers.
Remark 1: C
1
and C
2
are complex, and thus there exist 4 parameters that specify
the state. We can reduce this to 2 by noting:
164 CHAPTER 11. ANGULAR MOMENTUM: SPIN
[) = 1 [C
1
[
2
+[C
2
[
2
= 1
Overall phase of a state does not matter, we only need the relative
phase of C
1
and C
2
.
Remark 2: Let the 2 parameters be the angles , so that we can write
[) = cos
_

2
_
e
i/2
[+) + sin
_

2
_
e
i/2
[)
Let the state [) be represented by a point of the surface of the unit sphere. This is
called the Bloch sphere.
Figure 11.1: The Bloch Sphere
Remark 3: The axes on the Bloch sphere have physical signicance. Consider:
= 0 [) = [+) =

1
2
,
1
2
_
= [) = [) =

1
2
,
1
2
_
Thus the points represent eigenstates of

S
z
with eigenvalues
1
2
. Expressing
these in terms of Pauli matrices, the north and south pole of the Bloch sphere
represent:

z
) = 1 ( = 0)

z
) = 1 ( = )
11.4. SPIN DYNAMICS 165
Remark 4: The expectation values for
x
,
y
at these locations are:
= 0 :

x
) =
1

_
+

S
+
+

S

+
_
= 0

y
) =
1

_
+

(i)(

S
+

+
_
= 0
= :

x
) =
1

S
+
+

S

_
= 0

y
) =
1

(i)(

S
+

_
= 0
which suggests that the projection of the state onto each of the Cartesian axes
of the Bloch sphere corresponds to
x
) ,
y
) and
z
):
Remark 5: Consider the general case
[) = cos
_

2
_
e
i/2
[+) + sin
_

2
_
e
i/2
[)
166 CHAPTER 11. ANGULAR MOMENTUM: SPIN

x
) =
_
cos
_

2
_
e
i/2
+[ + sin
_

2
_
e
i/2
[
_ _

S
x
+

S

__
cos
_

2
_
e
i/2
[+) + sin
_

2
_
e
i/2
[)
_
= cos
_

2
_
e
i
+ sin
_

2
_
e
i
= sin() cos()

y
) = sin() sin()

z
) =
_
cos
_

2
_
e
i/2
+[ + sin
_

2
_
e
i/2
[
_

z
_
cos
_

2
_
e
i/2
[+) + sin
_

2
_
e
i/2
[)
_
= cos
2
_

2
_
sin
2
_

2
_
= cos()
These correspond to the Cartesian coordinates for a point located at , on
the unit sphere.
11.4.2 Rotations on the Bloch Sphere
Consider a rotation about the z-axis. From 11.3.3:

U[R( z)] = cos


_

2
_

I i sin
_

2
_

z
where is the angle we have rotated through. Applying this to the general state:

U[R( z)] [) =
_
cos
_

2
_

I i sin
_

2
_

z
_ _
cos
_

2
_
e
i/2
[+) + sin
_

2
_
e
i/2
[)
_
=
_
cos
_

2
_
i sin
_

2
_
cos
_

2
_
e
i/2
[+) +
_
cos
_

2
_
+i sin
_

2
_
sin
_

2
_
e
i/2
[)
= cos
_

2
_
e
i(+)/2
[+) + sin
_

2
_
e
i(+)/2
[)
Which shows that a point on the Bloch sphere represented by state , gets rotated
about the z-axis by an angle .
Now, consider a rotation about the y-axis:

U[R( y)] [) =
_
cos
_

2
_

I i
y
sin
_

2
_
_ _
cos
_

2
_
e
i/2
[+) + sin
_

2
_
e
i/2
[)
_
=
_
cos
_

2
_
cos
_

2
_
e
i/2
sin
_

2
_
sin
_

2
_
e
i/2
_
[+) +
_
cos
_

2
_
sin
_

2
_
e
i/2
+ sin
_

2
_
cos
_

2
_
e
i/2
_
[)
11.5. PARTICLES WITH SPIN 1 167
This is more complicated, but looking at the case where = 0, (in the
x
)
z
)
plane), we have:

U[R( y)] [( = 0)) = cos


_
+
2
_
[+) + sin
_
+
2
_
[)
Remark 1: Since the evolution of a state can be considered a rotation (11.4), the
evolution can be described as a point moving with time on the surface of the
Bloch sphere.
11.5 Particles with Spin 1
Electrons gave the rst clues to the existence of an intrinsic angular momentum. However, this
properly applies to a wide assortment of particles.
Remark 1: Recall that particles with
1
2
integer spin are called fermions. Particles with integer
spin are bosons.
Consider the rotated state of 11.3.1:

_
=
_

I
i

L
z
+

S
z
_
_
[)
The rotation matrix for this rotation is:
R
z
() =
_
_
cos sin 0
sin cos 0
0 0 1
_
_
168 CHAPTER 11. ANGULAR MOMENTUM: SPIN
11.5.1 Rotations of a Vector Field
(Disclaimer: Sometimes = 1 in this section)
Now, if we consider an arbitrary vector eld A(r), we can imagine this as having
a vector associated with every point in space.
When we rotate the vector eld, we are rotating the vectors A as well as the positions
r. we want to relate this rotation to angular momentum.
A A

, r Rr
So the transformation:
A

(Rr) = RA(r)
-or-
A

(r) = RA(R
1
r) ()
where the two dierent forms correspond to active or passive transformations. The
rotation matrix given for rotations about the z-axis has
R
1
z
() = R
z
()
Thus, the vector eld A

(r), according to (), is given by


A

(r) =
_
_
cos sin 0
sin cos 0
0 0 1
_
_
_
_
A
x
(R
1
r)
A
y
(R
1
r)
A
z
(R
1
r)
_
_
A

x
= A
x
(R
1
r) cos A
y
(R
1
r) sin
A

y
= A
x
(R
1
r) sin +A
y
(R
1
r) cos
A

z
= A
z
(R
1
r)
where
R
1
r =
_
_
x

_
_
=
_
_
xcos +y sin
y cos xsin
z
_
_
11.5. PARTICLES WITH SPIN 1 169
Since is small, we can write this as:
A

x
= A
x
_
_
xcos +y sin
. .
x+y
, y cos xsin
. .
yx
, z
_
_
A
y
(x +y, y x, z)
= A
x
(x +y, y x, z) A
y
(x, y, z) to O()
= A
x
(x, y, z) +
_
A
x
x

x=x

+
A
x
y

y=y

_
A
y
(r)
= A
x
(x, y, z) +
_
y

x
x

y
_
. .

Lz
A
x
(r) A
y
(r)
=
_

I
i

L
z
_
A
x
(r) A
y
(r)
Similarly,
A

y
(r) = A
x
(r) +
_

I
i

L
z
_
A
y
(r)
A

z
(r) =
_

I
i

L
z
_
A
x
(r)
A

(r) =
_

I
i

L
z
_
A(r)
. .
Usual Rotation

(iA
y
(r) x +iA
x
(r) y)
. .
Spin part
Thus we have
A

(r) =
_

I
i

L
z
+

S
z
_
_
A(r)
where we dene

S
z
A(r) =
i

A
y
(r) x +
i

A
x
(r) y
since

S
z
_
_
A
x
(r)
A
y
(r)
A
z
(r)
_
_
=
_
_
iA
y
(r)
iA
x
(r)
0
_
_
Therefore, we can represent

S
z
as the matrix

S
z
=
_
_
0 i 0
i 0 0
0 0 0
_
_

170 CHAPTER 11. ANGULAR MOMENTUM: SPIN


Similarly, rotations about the x and y axis give

S
x
=
_
_
0 0 0
0 0 i
0 i 0
_
_
,

S
y
=
_
_
0 0 i
0 0 0
i 0 0
_
_
Remark 1:

S
x
,

S
y
,

S
z
are related by row-column interchanges.
Remark 2: if

S
x
,

S
y
,

S
z
are components of angular momentum, they must obey
commutation relations. This can be checked directly. For example:

S
x

S
y


S
y

S
x
=
2
_
_
0 0 0
0 0 i
0 i 0
_
_
_
_
0 0 i
0 0 0
i 0 0
_
_

2
_
_
0 0 i
0 0 0
i 0 0
_
_
_
_
0 0 0
0 0 i
0 i 0
_
_
=
2
_
_
0 0 0
1 0 0
0 0 0
_
_

2
_
_
0 1 0
0 0 0
0 0 0
_
_
=
2
_
_
0 i 0
i 0 0
0 0 0
_
_
i
= i

S
z
Remark 3: The magnitude of angular momentum is given by

S

S =
_
_
0 0 0
0 0 i
0 i 0
_
_
2
+
_
_
0 0 i
0 0 0
i 0 0
_
_
2
+
_
_
0 i 0
i 0 0
0 0 0
_
_
2
= 2
_
_
1 0 0
0 1 0
0 0 1
_
_
= s(s + 1)

I
s(s + 1) = 2 s = 1 Spin 1
Remark 4: Working in the standard representation for a s = 1 system, [1, m), with

S
z
[1, m) = m [1, m) ; m = 0, 1
11.5. PARTICLES WITH SPIN 1 171
We nd that, in this representation,

S
z
=
_
_
1, 1[
1, 0[
1, 1[
_
_
S
z
_
[1, 1) [1, 0) [1, 1)
_
=
_
_
1 0 0
0 0 0
0 0 1
_
_
To nd the other components, well use the relations from 11.1:

S
+
[1, 0) =
_
(1 0)(1 + 0 + 1) [1, 1) =

2 [1, 1)

S
+
[1, 1) =
_
(1 + 1)(1 1 + 1) [1, 0) =

2 [1, 0)

[1, 1) =
_
(1 + 1)(1 1 + 1) [1, 0) =

2 [1, 0)

S
+
[1, 0) =
_
(1 0)(1 + 0 + 1) [1, 1) =

2 [1, 1)
So that

S
x

_
_
1, 1[
1, 0[
1, 1[
_
_
S
x
_
[1, 1) [1, 0) [1, 1)
_
=
1

2
_
_
0 1 0
1 0 1
0 1 0
_
_
Sp
y

_
_
1, 1[
1, 0[
1, 1[
_
_
S
y
_
[1, 1) [1, 0) [1, 1)
_
=
i

2
_
_
0 1 0
1 0 1
0 1 0
_
_
Remark 5: Notice that the matrices in Remark 4 are dierent than those given at
the beginning of this section. This is because the rst set are identied with
vector cartesian components, and the matrices in Remark 4 form a basis that
is not identied as such. To understand, recall that we dened the spin part
of the rotation by

S
z
A(r) = iA
y
(r) x +iA
x
(r) y
Let
u

=
1

2
( x i y)
with u
0
= z. Then any vector can be expressed in terms of these as
A = A
x
x +A
y
y +A
z
z
=
1

2
A
x
(u
+
+u

) +
1
i

2
A
y
(u
+
u

) +u
0
A
z
= A
+
u
+
+A

+A
0
u
0
172 CHAPTER 11. ANGULAR MOMENTUM: SPIN
with
A

=
1

2
(A
x
iA
y
) , A
0
= A
z
So

S
z
A(r) =
i

2
(u
+
+u

)
i

2
(A
+
(r) A

(r)) +
1

2
(u
+
u

)
1

2
(A
+
(r) +A

(r))
= u
+
A
+
(r) u

(r)
And thus

S
z
_
_
A
+
0
A

_
_
=
_
_
A
+
0
A

_
_


S
z
=
_
_
1 0 0
0 0 0
0 0 1
_
_
. .
Matrix from Remark 4
Similar arguments work for

S
x
,

S
y
. Physically, this change of basis can be
related to the polarization of a photon, which is a spin 1 particle. These spin
states of a photon are associated with circular polarization in the x-y plane
where m = 1, and linear polarization in the z direction (m = 0). Hence, it
is not associated with the 3 linear polarization directions. Therefore, u
+
, u

and A
+
, A

describe circular polarization of an EM eld.


Chapter 12
Addition of Angular Momentum
Consider the situation in which we have many particles that make up a system, and this system
has rotational invariance, and therefore angular momentum commutes with the Hamiltonian. The
total angular momentum of the system is the sum of the spin and orbital angular momentum for
all the particles. Given that we know the eigenstates for the individual pieces, we want to nd the
eigenstates of the total angular momentum.
12.1 Addition of 2 Angular Momenta
Consider 2 angular momenta,

j
1
,

j
2
, and let

J =

j
1
+

j
2
be the total angular momentum. The eigenstates for the individual pieces we already know:

j
2
1
[j
1
, m
1
) = j
1
(j
1
+ 1) [j
1
, m
1
)

j
1z
[j
1
, m
1
) = m
1
[j
1
, m
1
)
We have similar results for

j
2
. Let
[j
1
, j
2
, m
1
, m
2
) [j
1
, m
1
) [j
2
, m
2
) [j
1
, m
1
) [j
2
, m
2
)
be a representation based on the common eigenstates of

j
2
1
,

j
2
2
,

j
1z
,

j
2z
.
Remark 1: Since

J
z
=

j
1z
+

j
2z
,

J
z
[j
1
, j
2
, m
1
, m
2
) = (m
1
+m
2
) [j
1
, j
2
, m
1
, m
2
)
So the state [j
1
, j
2
, m
1
, m
2
) is also an eigenstate of

J
z
.
Remark 2: [j
1
, j
2
, m
1
, m
2
) is not an eigenstate of

J
2
, so we need to nd the eigenstates of

J
2
.
173
174 CHAPTER 12. ADDITION OF ANGULAR MOMENTUM
Remark 3: Since

j
2
1
commutes with every component of

j
1
and

j
2
(and similarly for

j
2
2
), we have
that
_

j
2
1
,

J
_
=
_

j
2
2
,

J
_
= 0
So we take

j
2
1
,

j
2
2
,

J
2
,

J
z
as our complete set of commuting observables, and instead of
[j
1
, j
2
, m
1
, m
2
), we use [j
1
, j
2
, J, M) (so

j
1z
and

j
2z
are replaced by J and M). In this
representation, we have:

j
2
1
[j
1
, j
2
, J, M) =
2
j
1
(j
1
+ 1) [j
1
, j
2
, J, M)

j
2
2
[j
1
, j
2
, J, M) =
2
j
2
(j
2
+ 1) [j
1
, j
2
, J, M)

J
2
[j
1
, j
2
, J, M) = J(J + 1) [j
1
, j
2
, J, M)

J
z
[j
1
, j
2
, J, M) = M[j
1
, j
2
, J, M)
With this representation, we need to nd
1. The allowed values of J and M for a given j
1
, j
2
, and
2. How to express [j
1
, j
2
, J, M) as a linear combination of [j
1
, j
2
, m
1
, m
2
).
12.1.1 Finding Allowed Values of J and M
For individual particles, we have allowed m values given by
m
1
= j
1
, j
1
+ 1, . . . , +j
1
m
2
= j
2
, j
2
+ 1, . . . , +2
1
and

j
1z
[j
1
, j
2
, m
1
, m
2
) = m
1
[j
1
, j
2
, m
1
, m
2
)

j
2z
[j
1
, j
2
, m
1
, m
2
) = m
2
[j
1
, j
2
, m
1
, m
2
)
and

J
z
[j
1
, j
2
, m
1
, m
2
) = (m
1
+m
2
) [j
1
, j
2
, m
1
, m
2
)
So we expect the M eigenvalue to be given by
M = (j
1
+j
2
), (j
1
+j
2
) + 1, . . . , +(j
1
+j
2
)
Remark 1: In general, there will be more than 1 state with the same M eigenvalue,
but these states will have dierent J eigenvalues.
12.1. ADDITION OF 2 ANGULAR MOMENTA 175
Remark 2: In order for us to get the allowed M values as [M[ j
1
+j
2
, we need a
J = j
1
+j
2
eigenvalue. This would give:
J : j
1
+j
2
M = (j
1
+j
2
), . . . , +(j
1
+j
2
)
j
1
+j
2
1 M = (j
1
+j
2
) + 1, . . . , (j
1
+j
2
) 1
j
1
+j
2
2 M = (j
1
+j
2
) + 2, . . . , (j
1
+j
2
) 2
.
.
.
But where does the sequence terminate? We can determine this with vectors.
If

j
1
and

j
2
can be represented as vectors (and they can), then the maximum
angular momentum should be obtained when they are parallel:
The minimum occurs when they are anti-parallel:
Remark 3: We can now nd the total number of [j
1
, j
2
, J, M) states. Without loss
of generality, let j
1
> j
2
. Then
# states = [2(j
1
+j
2
) + 1]
. .
Allowed M if
J=j
1
+j
2
+[2(j
1
+j
2
1) + 1]+. . .+[2(j
1
j
2
+ 1) + 1]+[2(j
1
j
2
) + 1]
Many of these terms cancel, leaving only (2j
1
+ 1) in each term. Since there
are (2j
2
+ 1) terms, the number of states is
# states = (2j
1
+ 1)(2j
2
+ 1)
which is the same number of states as we had in the [j
1
, j
2
, m
1
, m
2
) basis. So
we have the correct number of J and M eigenvalues.
Remark 4: To summarize: for a xed j
1
and j
2
, the (2j
1
+ 1)(2j
2
+ 1)-dimensional
space spanned by [j
1
, j
2
, m
1
, m
2
) has possible values of J given by j
1
+j
2
, j
1
+
j
2
1, . . . , [j
1
j
2
[, and for each J, there are (2J + 1) states [J, M) with
dierent M values.
176 CHAPTER 12. ADDITION OF ANGULAR MOMENTUM
12.1.2 Example of Addition of Two S=
1
2
States
Consider 2 particles with spin
1
2
. The individual eigenstates are:

S
1
=

1
2
,
1
2
_
1
and

1
2
,
1
2
_
1
[+)
1
and [)
1

S
2
=

1
2
,
1
2
_
2
and

1
2
,
1
2
_
2
[+)
2
and [)
2
We dene

S =

S
1
+

S
2
and we denote the eigenstates of

S
2
and

S
z
by [SM).
Remark 1: For the 4 [s
1
, s
2
, m
1
, m
2
) states: [+)
1
[)
2
; [+)
1
[)
2
; [)
1
[+)
2
; [)
1
[)
2
,
there are 4 [SM) states:
S = 1 [1, 1) , [1, 0) , [1, 1)
S = 0 [0, 0)
The state with S = 1 is called a triplet and the state with S = 0 is called the
singlet.
We now want to express the [SM) states in terms of the [s
1
, s
2
, m
1
, m
2
) states. Note
that there is only 1 state in each set with M = m
1
+ m
2
= 1 and one state with
M = m
1
+m
2
= 1. So we can write:
[1, 1) = [+)
1
[+)
2
[1, 1) = [)
1
[)
2
We can nd the other S = 1 states by using the lowering operator:
[1, 0) =
1

[1, 1)
=
1

2
_

S
1
+

S
2
_
[+)
1
[+)
2
=
1

2
([)
1
[+)
2
+[+)
1
[)
2
)
For the state corresponding to S = 0, we must have a linear combination of [+)
1
[)
2
, [)
1
[+)
2
which is orthogonal to the other states:
[0, 0) =
1

2
([+)
1
[)
2
[)
1
[+)
2
)
12.2. TWO NUCLEON SYSTEM 177
Remark 2: In the [s
1
, s
2
, m
1
, m
2
) basis, we have a tensor product of
1
2

1
2
, which is
equivalent to 1 0 in the [SM) basis. In other words,
1
2

1
2
(2s
1
+ 1)(2s
2
+ 1) = (
2
2
+ 1)(
2
2
+ 1) = 4
1 0
1

S=0
(2S + 1) = 1 + 3 = 4
which must be equal.
12.2 Two Nucleon System
Lets consider a system of 2 nucleons each with spin
1
2
, but with the addition of orbital angular
momentum. Only the angular momentum of the relative motion is relevant, so the total angular
momentum is given by:

J =

L+

S
1
+

S
2
=

L+

S
where

S =

S
1
+

S
2
is the total spin of the system and

L is the total angular momentum of relative
motion. We then have
[L, S, M
L
, M
S
) = [L, M
L
) [S, M
S
) = [L, M
L
) [S, M
S
)
with
L = 0, 1, 2, . . . M
L
= L, . . . , L
S = 1 M
S
= 1, 0, 1
S = 0 M
S
= 0
Giving the allowed

J values of
For S = 1

J =
_
L + 1, L, L 1 L 1
1 L = 0
S = 0

J = L
The states of total angular momentum are given by [L, S, J, M).
Remark 1: The quantum numbers L, S, J are related to spectroscopic language:
178 CHAPTER 12. ADDITION OF ANGULAR MOMENTUM
(a) General Form (b) Specic Instance
12.3. CLEBSCH-GORDAN COEFFICIENTS 179
12.3 Clebsch-Gordan Coecients
In 12.1, we noted that for a given j
1
, j
2
we needed to nd the allowed values of J, M (done in
12.1.1) and an expression for [j
1
, j
2
, J, M) in terms of [j
1
, j
2
, m
1
, m
2
) (done as an example of 2
spin
1
2
particles). We now wish to come up with a general scheme for expressing the eigenstates of
j
1
, j
2
, J, M. Formally, we can write:
[j
1
, j
2
, J, M) =

m
1
,m
2
j
1
, j
2
, m
1
, m
2
[j
1
, j
2
, J, M)
. .
Clebsch-Gordan Coecients
[j
1
, j
2
, m
1
, m
2
)
Remark 1: The Clebsch-Gordan coecients depend only on the angular momentum quantum
numbers, not on the specic details of a physical system. Thus they can be calculated
from the addition of angular momentum in general (without repeating for each physical
application).
Remark 2: In practive, extensive tables of these coecients are available.
Remark 3: To simplify notation, let
j
1
, j
2
, m
1
, m
2
[j
1
, j
2
, J, M) = j
1
, j
2
, m
1
, m
2
[J, M)
Remark 4: We already have some of the Clebsch-Gordan coecients from previous examples:
From 1.2 [0, 0) =
1

2
[+)
1
[)
2

2
[)
1
[+)
2

1
2
1
2
m
1
m
2

00
_
=
_

_
1

2
m
1
=
1
2
, m
2
=
1
2

2
m
1
=
1
2
, m
2
=
1
2
0 otherwise
Similarly,

1
2
1
2
m
1
m
2

11
_
=
_
1 m
1
= m
2
=
1
2
0 otherwise

1
2
1
2
m
1
m
2

1, 1
_
=
_
1 m
1
= m
2
=
1
2
0 otherwise

1
2
1
2
m
1
m
2

10
_
=
_

_
1

2
m
1
=
1
2
, m
2
=
1
2
1

2
m
1
=
1
2
, m
2
=
1
2
0 otherwise
180 CHAPTER 12. ADDITION OF ANGULAR MOMENTUM
Remark 5: Note that there can be an ambiguity in the overall phase (minus sign) of the states
[JM). This is xed by convention.
Remark 6: General properties of Clebsch-Gordan coecients:
1. j
1
j
2
m
1
m
2
[JM) = 0 unless M = m
1
+m
2
and j
1
+j
2
J [j
1
j
2
[.
2. They are real.
3. The convention for sign is that
j
1
j
2
j
1
, J j
1
[JM) > 0
where m
1
= j
1
and m
2
= J j
1
.
4. The equation
j
1
j
2
m
1
m
2
[JM) = (1)
j
1
+j
2
J
j
1
j
2
, m
1
, m
2
[J, M)
gives the coecients of M in terms of M.
5. Starting at the bottom (top) state and applying

J
+
(

J

) gives recurrence
relations between successive coecients:
_
J(J + 1) M(M + 1) j
1
j
2
m
1
m
2
[J, M + 1) =
_
j
1
(j
1
+ 1) m
1
(m
1
1) j
1
j
2
, m
1
1, m
2
[JM)
+
_
j
2
(j
2
+ 1) m
2
(m
2
1) j
1
j
2
m
1
, m
2
1[JM)
and
_
J(J + 1) M(M 1) j
1
j
2
m
1
m
2
[J, M 1) =
_
j
1
(j
1
+ 1) m
1
(m
1
+ 1) j
1
j
2
, m
1
+ 1, m
2
[JM)
+
_
j
2
(j
2
+ 1) m
2
(m
2
+ 1) j
1
j
2
m
1
, m
2
+ 1[JM)
Remark 7: The transformation from the [j
1
j
2
m
1
m
2
) basis to the [j
1
j
2
JM) basis is an example
of a unitary transformation between 2 orthonormal basis. Thus, for j
1
j
2
:
U
..
matrix elem.

_
j
1
j
2
m
1
m
2
. .
col. index

JM
..
row index
_
such that
UU

= U

U = I

JM
j
1
j
2
m
1
m
2
[JM)

JM

1
j

2
m

1
m

2
_
=
m
1
,m

m
2
,m

2
and

m
1
m
2
JM[j
1
j
2
m
1
m
2
)

j
1
j
2
m
1
m
2

_
=
J,J

M,M

12.4. IRREDUCABLE TENSOR OPERATORS 181


12.4 Irreducable Tensor Operators
12.4.1 Irreducable Invariant Subspaces
Consider a state [, j, m) such that

J
2
[, j, j) = j(j + 1) [, j, j)

J
z
[, j, j) = j [, j, j)
and the (2j + 1) orthonormal states obtained by 11.1:
[, j, m) =

(j +m)!
(2j)!(j m)!

J
jm

[, j, j)
which span the subspace of Hilbert space having eigenvalues of

J
2
and degeneracy
index . Let
(J)
denote this subspace. Rotations involve only operators

J
x
,

J
y
,

J
z
(or, alternatively,

J
+
,

J

,

J
z
), so that a state that is initially in
(J)
is rotated into
another state in
(J)
. We say that
(J)
is invariant under rotations.
Remark 1: Consider a state [) in
(J)
. All possible rotations of [) will span the
entire
(J)
. Because of this property, we say that
(J)
is irreducible under
rotations.
Remark 2: These rotations can be represented in standard representation by a ma-
trix:
R
(J)
MM

(, , )
. .
(2J+1)(2J+1)
matrix
=
_
, J, M

R(, , )

, J, M

_
The R
(J)
MM

are called rotation matrices and represent rotations in Hilbert


space. The rotated basis vectors

R[, J, M) can be expressed in terms of
the non-rotated basis using these rotation matrices:

R[, J, M) =

, J, M

_
_
, J, M

, J, M
_
=

R
(J)
MM

, J, M

_
182 CHAPTER 12. ADDITION OF ANGULAR MOMENTUM
12.4.2 Irreducable Tensors
Scalars, vectors, and tensors are distinguished by their transformation properties. For
instance, consider transformations under rotations:
Scalar: S S

= U

SU = sU

U = S
So scalars are invariant under rotations
Vector: v
i
v

i
= U

v
i
U = R
ij
v
j
and, equivalently,
v
i
v

i
= Uv
i
U

= R
ij
v
j
since

U[R]

U[R
1
] =

U

[R]
Denition 1: A rank k tensor operator is a collection of (2k+1) operators denoted
T
(k)
q
, q = k, k + 1, . . . , k that transform according to:

T
(k)
q


U

T
(k)
q

U

=

T
q
R
q

q
()
Remark 1: These tensor operators are irreducible under rotations.
Remark 2: Equation () represents a transformation of a group (basis) of opera-
tors. The equivalent and more familiar case is a transformation of states:

R[k, q) =

k, q

_
R
(k)
q

q
..
kq

R[kq)
12.4.3 Wigner-Eckart Theorem
This theorem is widely used in atomi and nuclear physics and is used to calculate
matrix elements for evaluating transition rates between states. It says that the matrix
elements in a standard representation of the q
th
component in a k
th
order irreducible
tensor operator is given by:
_
JM

T
(k)
q

_
=
1

2J + 1
. .
unimportant
conversion
factor
J[

T
(k)

_
. .
reduced matrix element

kM
q

JM
_
. .
CG-coecient
Remark 1: The Wigner-Eckart Theorem does not tell us what the reduced matrix
element is this depends on the particular tensor operator under consideration.
12.4. IRREDUCABLE TENSOR OPERATORS 183
The value lies in the way the q, M, M

quantum numbers are extracted in


the Clebsch-Gordan coecients. For example, in atomic physics, the M, M

quantum numbers for the initial and nal states in a radiative transition are
connected with the polarization of the emitted photon. In this example, the
Wigner-Eckart theorem tells us the ratios of radiative decay rates for dierent
polarizations. These ratios are given by ratios of the squares of the Clebsch-
Gordan coecients.
Remark 2: The Clebsch-Gordan coecients J

kM

q[JM) give the angular momen-


tum selection rules:
q = M M

J J

k J +J

184 CHAPTER 12. ADDITION OF ANGULAR MOMENTUM


Chapter 13
Perturbation Theory
Most QM problem can not be solved exactly, thus approximations are an important part of the
actual applications. Perturbation theory is one of the main approximation techniques used. The
idea is to approximate eigenvalues and eigenstates of a Hamiltonian which diers slightly from a
Hamiltonian whose eigenvalues and eigenstates are known exactly.
13.1 Perturbation Theory for a Non-Degenerate, Discrete, Energy
Level
Consider the Hamiltonian:

H =

H
0
+

V
where

H
0
is a Hamiltonian whose eigenvalues and eigenstates E
(0)
n
,

n
(0)
_
are known. So

H
0

n
(0)
_
= E
(0)
n

n
(0)
_
For now we consider only the case where the states are discrete and non-degenerate.

V is a
perturbation with the size of the perturbation and

V has the same size as

H
0
. The eigenvalues
and eigenstates of the full Hamiltonian are given by:

H
n
[n) = E
n
[n) ()
where we write
E
n
= E
(0)
n
+E
(1)
n
+
2
E
(2)
n
+
[n) = [n) +

n
(1)
_
+
2

n
(2)
_
+
_
_
_
()
So that
lim
0
E
n
= E
(0)
n
and lim
0
[n) =

n
(0)
_
Substituting () into () gives:
_

H
0
+

V
__

n
(0)
_
+

n
(1)
_

n
(2)
_
+
_
=
=
_
E
(0)
n
+E
(1)
n
+
2
E
(2)
n
+
__

n
(0)
_
+

n
(1)
_

n
(2)
_
+
_
185
186 CHAPTER 13. PERTURBATION THEORY
Now, equating power of on both sides:

H
0

n
(k)
_
+

V

n
(k1)
_
= E
(0)
n

n
(k)
_
+E
(1)
n

n
(k1)
_
+ +E
(k)
n

n
(0)
_
or _
_
_
_

H
0
E
(0)
n
_

n
(1)
_
=
_

V E
(1)
n
_

n
(0)
_
, k = 1
_

H
0
E
(0)
n
_

n
(k)
_
=
_

V E
(1)
n
_

n
(k1)
_
+

k
j=2
E
(j)
n

n
(kj)
_
, k 2
where terms with superscripts (k) refer to the kth order term in the perturbative expansion.
13.1.1 First Order Approximation
Consider k = 1. Then we can apply

n
(0)

to
_

H
0
E
(0)
n
_

n
(1)
_
=
_

V E
(1)
n
_

n
(0)
_
()

_
n
(0)

H
0
. .
E
(0)
n
E
(0)
n
_

n
(1)
_
=
_
n
(0)

V E
(1)
n
_

n
(0)
_
0 =
_
n
(0)

n
(0)
_
+E
(1)
n
E
(1)
n
=
_
n
(0)

n
(0)
_
Thus the rst order correction to the energy is just the expectation of the perturbation
taken with respect to the unperturbed state. To get the 1st order correction to the
eigenstate, we must expand

n
(1)
_
in the basis of the unperturbed state:
n
(1)
=

m
_
m
(0)

n
(1)
_

m
(0)
_
where m is the unperturbed state. To nd the coecients, we will apply

m
(0)

to
():
_
m
(0)

H
0
. .
E
(0)
m
E
(0)
n
__
n
(1)

=
_
m
(0)

V E
(1)
n
..
0
_

n
(0)
_

_
m
(0)

n
(1)
_
=
_
m
(0)

n
(0)
_
E
(0)
m
E
(0)
n
where the E
(1)
n
term goes to zero since we are considering when m ,= n and thus

m
(0)

n
(0)
_
= 0.
13.1. PERTURBATION THEORYFOR A NON-DEGENERATE, DISCRETE, ENERGY LEVEL187
Remark 1: At this point, it becomes important that we are considering a non-
degenerate state since this trick will not work if E
(0)
m
= E
(0)
n
for m ,= n.
Remark 2: In the expansion of

n
(1)
_
in terms of

m
(0)
_
, there will be a coecient
corresponding to n = m which appears to yield a diverging coecient. To x
this, we set
_
n
(1)

n
(0)
_
= 0
We can justify this by considering
[n) =

n
(0)
_
+

n
(1)
_
+
2

n
(2)
_
+
If

n
(k)
_
has an

n
(0)
_
term in its expansion, then

n
(k)
_
=
_
n
(0)

n
(k)
_

n
(0)
_
+

n
(k)
_

where the last term is an expansion of

n
(k)
_
in terms of

m
(0)
_
with m ,= n.
Thus we have
[n) =

n
(0)
_
_
1 +

k=1

k
_
n
(0)

n
(k)
_
_
. .
all n=m terms here
+

n
(1)
_

+
2

n
(2)
_

+
. .
expansions with n = m
=

n
(0)
_
+

n
(1)
_

+
2

n
(2)
_

+
where
= 1 +

k=1

k
_
n
(0)

n
(k)
_
is a scaling factor that can be absorbed into the length of [n). Dividing by
and letting
1

[n) [n) and

n
(k)
_

=
1

n
(k)
_

gives:
[n) =

n
(0)
_
+

n
(1)
_

+
2

n
(2)
_

+
which is the expansion of [n) with states

n
(k)
_

having no projection onto the


state

n
(0)
_
. This is equivalent to writing
[n) =

n
(0)
_
+

n
(1)
_
+
2

n
(2)
_
+
and setting
_
n
(0)

n
(k)
_
= 0, k = 1, 2, 3, . . .
188 CHAPTER 13. PERTURBATION THEORY
Remark 3: The procedure used to obtain the rst order correction in eigenvalues
and eigenstates can be applied successively to get higher order corrections in
terms of previous determined corrections of lower orders. Taking terms
k
with k 2 implies:
_

H
0
E
(0)
n
_

n
(k)
_
=
_

V E
(1)
n
_

n
(k1)
_
+
k

j=2
E
(j)
n

n
(kj)
_
on which we apply

n
(0)

to get
0 =
_
n
(0)

_
_
V E
(1)
n
..
0
_
_

n
(k1)
_
+
k

j=2
E
(j)
n
_
n
(0)

n
(kj)
_
=
_
n
(0)

n
(k1)
_
+
_
_
_E
(k)
n
+
k2

i=1
E
(ki)
n
_
n
(0)

n
(1)
_
. .
0
_
_
_
E
(k)
n
=
_
n
(0)

n
(k1)
_
Now, applying

m
(0)

to terms O(
k
) implies
_
m
(0)

H
0
E
(0
n
_

n
(k)
_
=
_
m
(0)

V E
(1)
n
_

n
(k1)
_
+
k

j=2
E
(j)
n
_
m
(0)

n
(kj)
_
_
m
(0)

n
(k)
_
=
1
_
E
(0)
m
E
(0)
n
_
_
_
_
_
_
_
_
_

_
m
(0)

V E
(1)
n
_

n
(k1)
_
+
k1

j=2
E
(j)
n
_
m
(0)

n
(kj)
_
. .
to k 1 since

m
(0)

n
(0)

=0
_
_
_
_
_
_
_
_
which are coecients for the expansion

n
(k)
_
=

m
_
m
(0)

n
(k)
_

m
(0)
_
13.1.2 Example: Harmonic Oscillator in an Electric Field
Consider a particle of charge q and mass m in a harmonic oscillator potential and
corresponding to an unperturbed Hamiltonian:

H
0
=
p
2
2m
+
1
2
m
2
x
2
13.1. PERTURBATION THEORYFOR A NON-DEGENERATE, DISCRETE, ENERGY LEVEL189
Applying an electric eld with magnitude f along the +x direction corresponds to an
electrostatic potential = fx and potential energy

V = qf x
The full Hamiltonian then is

H =

H
0
+

V
where is the strength of the perturbation. The rst order correction to the energy
is given by
E
(1)
n
=
_
n
(0)

n
(0)
_
= qf
_
n
(0)

n
(0)
_
= qf
_

2m
_
1/2 _
n
(0)

+ a

n
(0)
_
Remark 1: Physically, the average interaction of the particle in the electric eld is
0 since the particles unperturbed wavefunction is a H.O.
E
(1)
n
= qf
_
_

(0)
n
_

x
_

(0)
n
_
dx = qf
_

(0)
n

2
. .
even
x
..
odd
dx = 0
The rst order correction to the eigenstate is:
_
m
(0)

n
(1)
_
=

m
(0)

(qf x)

n
(0)
_
E
(0)
m
E
(0)
n
= qf
_

2m
_
1/2

m
(0)

a + a

n
(0)
_
E
(0)
m
E
(0)
n
=
qf
_

2m
E
(0)
m
E
(0)
n
_
n
m,n1
+

n + 1
m,n+1
_

n
(1)
_
= qf
_

2m

m
1
E
(0)
m
E
(0)
n
_
n
m,n1
+

n + 1
m,n+1
_

m
(0)
_
= qf
_

2m
_
_
_
_
_
_

(n 1)
(0)
_
E
(0)
n1
E
(0)
n
. .
h
+

n+

(n + 1)
(0)
_
E
(0)
n+1
E
(0)
n
. .
h
_
_
_
_
_
_
= qf
_
1
2m
3
_

n + 1

(n + 1)
(0)
_

(n 1)
(0)
__
190 CHAPTER 13. PERTURBATION THEORY
So, to rst order, our eigenstate is:
[n) =

n
(0)
_
+

n
(1)
_
=

n
(0)
_
+qf
_
1
2m
3
_

n + 1

(n + 1)
(0)
_

(n 1)
(0)
__
Since the rst order energy correction was 0, lets calculate the 2nd order correction:
E
(2)
n
=
_
n
(0)

n
(1)
_
= qf
_
n
(0)

n
(1)
_
= qf
_

2m
_
n
(0)

a + a

n
(1)
_
= (qf)
2
1
2m
2
_
n
(0)

a + a

n + 1

(n + 1)
(0)
_

(n 1)
(0)
___
=
(qf)
2
2m
2
_
n
(0)

_
(n + 1)

n
(0)
_

n 1

(n 2)
(0)
_
+

n + 1

n + 1

(n + 2)
(0)
_
n

n
(0)
__
=
(qf)
2
2m
2
(n + 1 n)
=
(qf)
2
2m
2
Remark 2: This is a problem that can be solved exactly, which makes it a nice test
for perturbation theory. Consider:

H =
p
2
2m
+
1
2
m
2
x
2
qf x
=
p
2
2m
+
1
2
m
2
_
x
qf
m
2
_
2
. .
shifted H.O.

1
2
q
2
f
2

2
m
2
. .
const. potential that
we dont care about
resulting in energy eigenvalues given by
E
n
= E
(0)
n

q
2
f
2

2
2m
2
. .
energy shift
= E
(0)
n
+E
(1)
n
. .
0
+
2
E
(2)
n
and eigenstates of the H.O. with oscillator displaced by:
Q =
qf
m
2
= x
_
2
m
x =
qf
m
2
_
m
2
=
qf

2mm
2
[n) +D(x)

n
(0)
_
13.1. PERTURBATION THEORYFOR A NON-DEGENERATE, DISCRETE, ENERGY LEVEL191
where D(x) is the displacement operator of 6.3.3.
D(x) = e
x( a

+ a)
1 +x
_
a

+ a
_
. .
to 1st order
[n) =

n
(0)
_
+
qf

2m
3
_

n + 1

(n + 1)
(0)
_

(n 1)
(0)
__
Here we note that computing higher order terms in the perturbation expansion
is equivalent to taking higher order terms in the D(x) expansion. Also, note
that the energy eigenvalue is exact at 2nd ordercomputing higher order terms
in E
n
will give 0 contribution.
13.1.3 Example: Stark Eect
See Homework!
The picture is: Here we have that L | and the rotation occurs in the plane dened
by the angles and relative to the Cartesian axis.
E =
1
2
I
2
=
1
2

L
2
I
=
1
2
_

L
2
x
+

L
2
y
+

L
2
z
_
I
with
I =
1
2
mD
2
Hint:
cos Y
m
l
=
_
(l + 1 +m)(l + 1 m)
(2l + 1)(2l + 3)
_
1/2
Y
m
l+1
+
_
(l +m)(l m)
(2l + 1)(2l 1)
_
1/2
Y
m
l1
We will see that 2nd order perturbation removes the degeneracy in the unperturbed
eigenvalue.
Note: a degeneracy remains with respect to the sign of m which physically corresponds
to the conguration of a rigid rotor by invariance with respect to .
192 CHAPTER 13. PERTURBATION THEORY
13.2 Degenerate Perturbation Theory
Now, consider the case where the unperturbed eigenstates are degenerate. Start with the simplest
case of 2-fold degeneracy (higher order degeneracy is straight forward).

H
0

n
(0)
, 1
_
= E
(0)
n

n
(0)
, 1
_

H
0

n
(0)
, 2
_
= E
(0)
n

n
(0)
, 2
_
_
_
_
(1)
where the 1s and 2s and the degeneracy index. So eigenstates

n
(0)
, 1
_
and

n
(0)
, 2
_
(which are
assumed to be orthonormal), give the same eigenvalue E
(0)
n
. We expect the full solution to look
like:

H[n, 1) = E
n
1
[n, 1)

H[n, 2) = E
n
2
[n, 2)
_
(2)
to give E
n
E
(0)
n
as 0. Also, for now, assume that one eigenstate approaches

n
(0)
, 1
_
and
the other approaches

n
(0)
_
. To simplify notation, lets let:

n
(k)
, j
_
[k)
j
with j = 1, 2 and k = the order of perturbation. Then,
E
n
1
= E
(0)
n
+E
(1)
n
1
+ +
k
E
(k)
n
1
+
E
n
2
= E
(0)
n
+E
(1)
n
2
+ +
k
E
(k)
n
2
+
_
(3)
and
[n, 1) = [0)
1
+[1)
1
+ +
k
[k)
1
+
[n, 2) = [0)
2
+[1)
2
+ +
k
[k)
2
+
_
(4)
Now, substitute (3) and (4) into (2) to get the expansion in terms of powers of . As outlined in
13.1, the terms to 1st order in give:
_

H
0
E
(0)
n
_
[1)
1
=
_

V E
(0)
n
1
_
[0)
1
_

H
0
E
(0)
n
_
[1)
2
=
_

V E
(0)
n
2
_
[0)
2
_
_
_
(5)
Multiplying on the left by 0[
1
gives:
0 =
_
0

0
_
1 1
+E
(1)
n
1
(6)
0 =
_
0

0
_
1 2
(7)
13.2. DEGENERATE PERTURBATION THEORY 193
While multiplying on the left by 0[
2
gives:
0 =
_
0

0
_
2 1
(8)
0 =
_
0

0
_
2 2
+E
(1)
n
2
(9)
Remark 1: This analysis is sucient if equations (7) and (8) hold. In reality, however, they may
not. In fact, if they do hold, this situation is identical to the non-degenerate results (as in the
rigid rotor problem) and weve gained nothing. If (7) and (8) do not hold, then we have to
do something new. In this case, what weve done must be incorrect since (7) and (8) would
be contradictory. What went wrong above was the assumption that [E
n
, 1) and [E
n
, 2)
approached specic [0)
1
and [0)
2
respectively as 0. In fact, the full eigenstates can
hypothetically approach any mutually orthogonal linear combination of those states (since
any linear combination of these states is also an eigenstate of

H
0
). We must determine the
linear combinations for the degenerate perturbed states.
Let us proceed without making any assumptions on the actual form of the unperturbed degenerate
eigenstates, (let those be mutually orthogonal, normalized, linear combinations of the degenerate
states), such that the expansions given by (4) still hold.
[n, 1) = [0)
1
+[1)
1
+ +
k
[k)
1
+
[n, 2) = [0)
2
+[1)
2
+ +
k
[k)
2
+
_
(4)
where [0)
1
and [0)
2
are now unknown states. Then the 1st order in the perturbation expansion still
gives
_

H
0
E
(0)
n
_
[1)
1
=
_

V E
(0)
n
1
_
[0)
1
_

H
0
E
(0)
n
_
[1)
2
=
_

V E
(0)
n
2
_
[0)
2
_
_
_
(5)
where we multiply (5) by any

n
(0)

for which (1) holds to get 0 on the left hand side. Formally,
we can express this as:

P
_

V E
(1)
n
1
_
[0)
1
= 0 (6)

P
_

V E
(1)
n
2
_
[0)
2
= 0 (7)
where

P =

n
(0)
, 1
_
n
(0)
, 1

n
(0)
, 2
_
n
(0)
, 2

is the projection onto the subspace of degenerate


states.

n
(0)
, 1
_
and

n
(0)
, 2
_
form an orthonormal basis over this subspace. In practice, this
calculation reduces to the diagonalization of a 2 2 matrix. To see this, lets let
[0)
1
=

n
(0)
, 1
_
+

n
(0)
, 2
_
(8)
where , are unknown. Then (6) gives
_

n
(0)
, 1
__
n
(0)
, 1

n
(0)
, 2
__
n
(0)
, 2

__

V E
(1)
n
1
__

n
(0)
, 1
_
+

n
(0)
, 2
__
= 0
194 CHAPTER 13. PERTURBATION THEORY
0 =

n
(0)
, 1
____
n
(0)
, 1

n
(0)
, 1
_
E
(1)
n
1
_
+
_
n
(0)
, 1

n
(0)
, 2
_

_
+
+

n
(0)
, 2
____
n
(0)
, 2

n
(0)
, 2
_
E
(1)
n
1
_
+
_
n
(0)
, 2

n
(0)
, 1
_

__
n
(0)
, 1

n
(0)
, 1
_
E
(1)
n
1
_
+
_
n
(0)
, 1

n
(0)
, 2
_
= 0
__
n
(0)
, 2

n
(0)
, 2
_
E
(1)
n
1
_
+
_
n
(0)
, 2

n
(0)
, 1
_
= 0
_
_
_
_
_
n
(0)
, 1

n
(0)
, 1
_ _
n
(0)
, 1

n
(0)
, 2
_
_
n
(0)
, 2

n
(0)
, 1
_ _
n
(0)
, 2

n
(0)
, 2
_
_
_
E
(1)
n
_
1 0
0 1
_
_
_
_

_
=
_
0
0
_
(9)
which is the standard 2 2 matrix eigenvalue equation. Solving the eigenvalues, eigenvectors, and
normalizing gives rst order energy corrections, and the right linear combinations of the unper-
turbed degenerate ground states.
Remark 2: Solving the characteristic equation given by (9) can result in 1st order energy correc-
tions (E
(1)
n
1,2
) that are the same or distinct, depending on the roots of the quadratic equation.
So, to rst order, the degeneracy may or may not be lifted.
13.2.1 Example: Coupled Harmonic Oscillator
Consider 2 identical H.O.s with frequency that are coupled by an interaction en-
ergy proportional to the product of oscillator displacements from equilibrium. The
Hamiltonian is:

H =

H
0
+

V
where

H
0
=
_
a

a +

b
_

V = g
_
a + a

__

b +

_
and where a, a

and

b,

are annihilation and creation operators, and we drop 2 factors


1
2
in

H
0
.
The unperturbed energy is
E
N
= E
n
+E
m
= N = (n +m)
And the degeneracy arises because there is more than 1 way to break up the energy.
All we require is that N = n+m and the degeneracy increases with increasing energy.
n and m are numbers of quanta in independent, uncoupled oscillators.
States Degeneracy
[0, 0) none
[1, 0) , [0, 1) 2-fold
[0, 2) , [1, 1) , [2, 0) 3-fold
13.2. DEGENERATE PERTURBATION THEORY 195
Consider 1st order perturbation theory for the 1st excited state (2-fold degeneracy).
In the limit g 0, the exact eigenstate goes to

E
(0)
N
1,2
_
= [0, 1) + [1, 0)
The matrix elements required in (9) of the previous section are
_
0, 1

g
_
a + a

__

b +

0, 1
_
= 0
_
1, 0

g
_
a + a

__

b +

1, 0
_
= 0
_
0, 1

g
_
a + a

__

b +

1, 0
_
= g
_
1, 0

g
_
a + a

__

b +

0, 1
_
= g
And (9) becomes
_
E
(1)
N
1
g
g E
(1)
N
2
_
_

_
=
_
0
0
_
Solving the characteristic equation gives
E
(1)
N
1
= g
E
(1)
N
2
= g
_
_
_
(10)
And the ground state, unperturbed eigenstates follow:

E
(1)
N
1
_
=
1

2
([0, 1) +[1, 0))

E
(1)
N
2
_
=
1

2
([0, 1) [1, 0))
_

_
(11)
Now we need 1st order corrections to the eigenstates

E
(1)
N
, 1
_
=

M=N
_
E
(0)
M

E
(0)
N
, 1
_

E
(0)
m
_
=

M=N
_
E
(0)
M

E
(0)
N
, 1
_
E
(0)
M
E
(0)
N
1

E
(0)
M
_
(12)
Remark 1: This is the same procedure used in non-degenerate perturbation theory,
but the summation is more restrictive since E
(0)
N
1
and E
(0)
N
2
have no projection
on the degenerate subspace.
Remark 2: An expression similar to (12) gives

E
(0)
N
, 2
_
.
196 CHAPTER 13. PERTURBATION THEORY
To get the 1st order correction to the eigenstates, calculate the matrix elements
_
E
(0)
M

E
(0)
N
, (1 or 2
_
and specically for the H.O.:
[n, m)
. .
all unperturbed
states other than
|1,0 & |0,1
g
_
a + a

__

b +

_
([0, 1) [1, 0)) = n, m[
g

2
_

2 [1, 2)

2 [2, 1)
_

1, 2[

2
([0, 1) [1, 0)) = g
2, 1[

2
([0, 1) [1, 0)) = g
_

_
(13)
Remark 3: The states [1, 2) and [2, 1) both have energy E
(0)
M
= 3.
Plugging (13) into (12) gives:

E
(1)
N
1
_
=
g
3
[1, 2)
g
3
=
g
2
([1, 2) +[2, 1)) (14)

E
(1)
N
2
_
=
g
2
([1, 2) [2, 1)) (15)
Remark 4: The interacting Hamiltonian in this example is often approximated by

V
RWA
= g
_
a

b + a

_
This gives the 1st order corrections that are exact (no higher order approxima-
tions from a

b, a

). This interaction is referred to as the rotating wave approximation.


Remark 5: In this case, the 1st order corrections lifted the degeneracy, and higher
order corrections may proceed with non-degenerate perturbation theory. If
degeneracy was not lifted at rst order, we must diagonalize another matrix
(use non-degenerate perturbation theory) again.
The second order corrections in energy for coupled H.O. are given by
E
(2)
N
1
=
_
E
(0)
N
, 1

E
(1)
N
, 1
_
=
1

2
([0, 1) +1, 0[) g
_
a + a

__

b +

_
_
g
2
_
([1, 2) +[2, 1))
=
g
2

13.3. TIME DEPENDENT PERTURBATION THEORY 197


and
E
(2)
N
2
=
_
E
(0)
N
, 2

E
(1)
n
, 2
_
=
g
2

So, to second order, we have


E
N
1
=
_
+g
g
2

_
E
N
2
=
_
g
g
2

_
Remark 6: Note that it was the non-rotating wave approximation part of the inter-
action that gives the second order corrections ( a

b term). This is an example


of a non-energy conserving, or virtual transition.
or
Only the intermediate step is non-conserving, overall, energy is conserved.
13.3 Time Dependent Perturbation Theory
For this analysis, we are going to consider the unitary propagator (aka time-evolution operator):

U(t, t
0
) = e
i

H(tt
0
)/
(1)
which gives the solution to the Schrodinger equation:
[(t)) =

U(t, t
0
) [(t
0
)) (2)
The propagator itself obeys the equation of motion.

U(t, t
0
) =
1
i

H

U(t, t
0
) (3)
198 CHAPTER 13. PERTURBATION THEORY
with the initial condition that

U(t
0
, t
0
) = 1. Formally integrating (3) gives

U(t, t
0
) = 1 +
1
i
_
t
t
0

H

U(, t
0
) d (4)
Time dependent perturbation theory is based on (4). Our 1st goal is to rewrite (4) so that the
dynamics generated by the unperturbed interacting part of the Hamiltonian are separate.

H(t) =

H
0
(t) +

V (t) (5)
Here we are allowing

H
0
and

V to explicitly depend on time.
Remark 1: We will use the interaction picture to separate the perturbed and unperturbed dy-
namics. Let

U
0
(t, t
0
) represent the propagator which governs the unperturbed motion such
that

U
0
(t, t
0
) satises

U(t, t
0
) =
1
i

H
0
(t)

U
0
(t, t
0
) (6)
with initial condition

U
0
(t
0
, t
0
) = 1. Now,

U
0
(t, t
0
) can be thought of as a rotation in
Hilbert space. If we go into a reference frame that rotates at a rate determined by

U
0
(t, t
0
),
then the state only evolves in that frame if

V (t) ,= 0. In the case that

V (t) ,= 0, we are in
the interaction picture.
Remark 2: Alternatively, a non-rotating frame in which the state evolves according to both

H
0
and

V
0
gives the usual Schrodinger picture, while a frame that rotates at a rote determined
by both

H
0
and

V
0
(the state is stationary in this frame) gives the Heisenberg picture.
Remark 3: The states in the Schrodinger and interaction pictures are related by
[
S
(t))
. .
Schrodinger
picture
=

U
0
(t, t
0
) [
I
(t))
. .
Interaction
picture
(7)
so that if

V (t) = 0, [
I
(t)) is stationary, and [
I
(t)) = [
I
(t
0
)) = [
S
(t
0
)). If

V (t) ,= 0,
then the state in the interaction picture evolves according to
[
I
(t)) =

U
I
(t, t
0
) [
I
(t
0
)) (8)
where

U
I
(t, t
0
) is the interaction picture evolution operator.
Remark 4: The evolution equation for

U
I
(t, t
0
) can be obtained via (7) and (8):
[
S
(t)) =

U
0
(t, t
0
)

U
I
(t, t
0
) [
I
(t
0
))
=

U
0
(t, t
0
)

U
I
(t, t
0
) [
S
(t
0
)) (9)
Comparing to (2):


U
S
(t, t
0
) =

U
0
(t, t
0
)

U
I
(t, t
0
) (10)
13.3. TIME DEPENDENT PERTURBATION THEORY 199
Substituting(10) and (5) into (3):

t
_

U
0

U
I
_
=
1
i
_

H
0
+

V
_

U
0

U
I
_


U
0
t
_

U
I
+

U
0
_


U
I
t
_
=
_
1
i

H
0

U
0
_

U
I
+
1
i

V

U
0

U
I



U
I
t
=
1
i
_

U
0

V

U
0
_

U
I
(11)
since

U
0
is unitary (and thus

U

U
0
= I).
Dening

V
I
(t) =

U

0
(t, t
0
)

V (t)

U
0
(t, t
0
) (12)
we have that

U
I
(t, t
0
) =
1
i

V
I

U
I
(t, t
0
) (13)
Formally integrating gives

U
I
(t, t
0
) = 1 +
1
i
_
t
t
0

V
I

U
I
(, t
0
) d (14)
Time dependent perturbation corrections are obtained by iterating (14):

U
I
(t, t
0
) = 1 +
1
i
_
t
t
0

V
I
() d
_
1 +
1
i
_

t
0

V
I
(

U
I
(

, t
0
) d

_
= 1 +
1
i
_
t
t
0

V
I
() d
1

2
_
t
t
0
d
_

t
0
d


V
I
()

V (

U
I
(

, t
0
) (15)
Higher order corrections can be obtained via an innite number of iterations. In powers of

V (t),
this becomes:

U
I
(t, t
0
) = 1 +

n=1
_
1
i
_
n
_
t
t
0
d
n
_
n
t
0
d
n1

_

2
t
0
d
1

V
I
(
n
)

V
I
(
n1
)

V
I
(
1
) (16)
where the order of integration is such that
t
0
<
1
<
2
< <
n1
<
n
< t
Remark 5: We now transform (16) back into the Schrodinger picture using

U
S
(t, t
0
) =

U
0
(t, t
0
)

U
I
(t, t
0
)
and

V
I
(
j
) =

U

0
(
j
, t
0
)

V (
j
)

U
0
(
j
, t
0
)
Also, note that

U
0
(
j
, t
0
)

0
(
j1
, t
0
) =

U
0
(
j
,
j1
)
200 CHAPTER 13. PERTURBATION THEORY
Thus we have that

U
S
(t, t
0
) =

U
0
(t, t
0
) +

n=1
_
1
i
_
n
_
t
t
0
d
n
_
n
t
0
d
n1

_

2
t
0
d
1


U
0
(t,
n
)

V (
n
)

U
0
(
n
,
n1
)
. .

U
0
(t,t
0
)

U

0
(
n
, t
0
)

V (
n
)

U
0
(
n
, t
0
)
. .

V
I

0
(
n1
,t
0
)

V (
n1
)

U
0
(
2
,
1
)

V (
1
)

U
0
(
1
, t
0
) (17)
Remark 6: Starting at the earliest times (right) and moving to interacting (left) in the integrand
represents
1. Free evolution from t
0
to
1
2. Interaction at
1
3. Free evolution from
1
to
2
.
.
.
4. Interaction at
n
5. Free evolution from
n
to t
So that the full evolution represents a sum over all elementary time evolutions, taking into account
all possible numbers of interactions between t
0
and t.
13.3.1 Transition Properties Between Energy States
For this analysis, consider the case where the free Hamiltonian

H
0
is time independent.
Then

U
0
(
j
,
j1
) = e
i

H
0
(
j

j1
)/
(18)
and we consider a system with discrete energies E
1
, E
2
, . . .. We develop the perturba-
tion theory in terms of jumps between the levels.
The transition probability between some initial state [E
a
) and some nal state [E
b
) is
dened as
W
ab
=

_
E
b

U
S
(t, t
0
)

E
a
_

2
(19)
13.3. TIME DEPENDENT PERTURBATION THEORY 201
Remark 1: The transition probability depends on the interval t
0
t during which
the transition takes place. We expect non-zero transition probabilities if there
is an interaction

V () which couples the initial and nal states [E
a
) , [E
b
). The
interaction is specied in terms of matrix elements in the unperturbed energy
basis:

V () =

k,l
V
kl
() [E
k
) E
l
[ (20)
where V
kl
() is:
V
kl
() =
_
E
a

V ()

E
l
_
(21)
Substituting (17), (18) and (20) into (19) gives:
W
ab
=

n=1
_
1
i
__
t
t
0
d
n
_
n
t
0
d
n1

_

2
t
0
d
1

kn,ln

k
n1
,l
n1

k
1
,l
1
_
E
b

e
i

H
0
(tn)/

E
kn
_

V
kn,ln
(
n
)
_
E
ln

e
i

H
0
(n
n1
)/

E
k
n1
_

V
k
n1
,l
n1
(
n1
)

E
l
n1

e
i

H
0
(
2

1
)/
[E
k
1
) V
k
1
,l
1
(
1
)
_
E
ln

e
i

H
0
(
1
t
0
)/

E
a
_

2
=

n=1
_
1
i
_
n
_
t
t
0
d
n
_
n
t
0
d
n1

_

2
t
0
d
1

ln

l
n1

l
2
e
iE
b
(tn)/
V
b,ln
(
n
)
e
iE
ln
(n
n1
)/
V
ln,l
n1
(
n1
) e
iE
l
2
(
2

1
)/
V
l
2
,a
e
iEa(
1
t
0
)/

2
(22)
Remark 2:

n=1
sum over the number of intermediate state interactions between
t
0
and t

j
t
0
integration over all possible times for intermediate state interactions
V
l
j
,l
j1
strength of interaction connecting intermediate states

E
l
j1
_
and

E
l
j
_
( ) transition amplitude is sum over amplitudes for transitions tak-
ing place via a set of intermediate or virtual states [E
l
2
) , [E
l
3
) , . . . , [E
ln
).
There are n 1 intermediate states (stepping stones) so that there are n
steps between [E
a
) and [E
b
).
Remark 3: Diagrammatically, the transitions can be represented as:
where the vertices are where the state of the system changes.
202 CHAPTER 13. PERTURBATION THEORY
Remark 4: To rst order in perturbation theory, (22) looks like
W
ab
=
1

_
t
t
0
e
iE
b
(t)/
V
ba
()e
iEa(t
0
)/
d

2
=
1

_
t
t
0
e
i
ba

V
ba
() d

2
(23)
where
ba
=
E
b
Ea

.
13.3.2 Time-Independent Interactions
The time integral in (23) is easy if the interaction is independent of time:
W
ab
=
1

2
[V
ba
[
2

e
i
ba
t
e
i
ba
t
0
i
ba

2
=
1

2
[V
ba
[
2
1

2
ba
_
1 + 1 e
i
ba
(tt
0
)
e
i
ba
(tt
0
)
_
=
2

2
[V
ba
[
2
1

2
ba
(1 cos (
ba
(t t
0
))) (24)
Let t = t t
0
and dene
f(, t)
2(1 cos(t))

2
(25)
Remark 1: f(, t) contains information about energy conservation during the tran-
sition. It is peaked at = 0, with a maximum f(0, t) = (t)
2
. This is seen
in Fig 13.1.
In the nal limit t big, this function becomes highly peaked about = 0 with
half-width

t
and
f(, t)
t
A() (26)
13.3. TIME DEPENDENT PERTURBATION THEORY 203
Figure 13.1: Plot of f(, t).
where A is found by integrating (25) over .
A = 2
_

1 cos(t)

2
d
= 2t (27)
where the integral is done by contour integration. (Let f(z) =
1e
izt
z
2
and the
integral is the real part of this along the x-axis.) So, in the limit of innite interaction
time, the transition probability is 0 unless

ba
=
E
b
E
a

= 0 (28)
This is a statement of conservation of energy: A mismatch of energies is tolerated for
nite time intervals as long as
Et , with E =

t
(29)
Transitions occurring in a nite time t involve an uncertainty about the energies.
These transitions can occur if E
a
and E
b
are close enough together so that the total
system energy uncertainty can bring them together during a nite interaction. Note
that E in (29) is the dierence in unperturbed energies, not the energy shift due to
the perturbation.
Remark 2: For transitions between states of same energy:
W
ab
=
1

2
[V
ba
[
2
(t)
2
(30)
which has limited validity since W
ab
must not exceed 1, even as t .
Hence (30) is valid only if
t

[V
ba
[
204 CHAPTER 13. PERTURBATION THEORY
13.3.3 Transition to a Continuum of States
Often, [E
b
) is not a discrete state, but is part of a continuous spectrum:
In these cases, we must nd the transition probability to the state whose energy falls
in the range of energies centered at E
b
. Let B denote the subspace of nal states with
energies in the range E
b
E to E
b
+E and dene
W
aB
=

all states
in B
W
ab
(31)
where W
ab
is given by (24). For a continuum of states, we replace the sum in (31)
with an integral which includes the density of states:
E
b
dE
b
= Number of states with energy in the range E
b
to E
b
E (32)
So (31) becomes
W
aB
=
_
E
0
+E
E
0
E
dE (E)
2

_
E

E
a
_

2
_
1 cos
_
EEa

t
_
_
EEa

_
2
_
(33)
Remark 1: If the range of nal energies is suciently small, then
(E) (E
b
)

_
E

E
a
_

_
E
b

E
a
_

2
= [V
ba
[
2
in which case
W
aB
= 2(E) [V
ba
[
2
_
E
0
+E
E
0
E
dE
_
1 cos
_
EEa

t
_
(E E
a
)
2
_
(34)
Remark 2: From 13.3.2.R1, we know that the function in (34) has a max value
of
_
t

_
2
at E = E
a
, then secondary maxima get smaller as deviations from
the resonant energies (E = E
a
) decrease. The width of the central maximum
13.3. TIME DEPENDENT PERTURBATION THEORY 205
is

t
, and we assume that the integration range is much larger than this
width.
E

t
(35)
With this approximation, we consider (34) for 2 distinct cases:
1. [E
b
E
a
[ > E implies that the nal states are non-conserving by an
amount larger than the width of the transition probability maximum
In this case,
1
(E E
a
)
2

1
E
b
E
a
)
2
So that
_
E
0
+E
E
0
E
dE
_
1 cos
_
E E
a

t
__
. .
2Et
And so
W
aB
=
4(E
b
)E [V
ba
[
2
(E
b
E
a
)
2
(36)
2. [E
b
E
a
[ < E implies that the transition probability peak falls within
the range of nal energies. This means that the transition is energy
conserving. Now,
_
E
0
+E
E
0
E
in (34)
_

and (34) in given by (27). Thus


W
aB
=
2t

(E
b
) [V
ba
[
2
(37)
206 CHAPTER 13. PERTURBATION THEORY
Remark 3: Eqn (37) is an important result since it represents the dynamical behavior
in many quantum systems. I.e., imagine N
a
systems in state [E
a
) at time t.
A short time t later, the number of systems in state [E
a
) is
N
a
(t + t) = N
a
(t)
_
1
2

(E
b
) [V
ba
[
2
t
_
So
N
a
(t + t) N
a
(t)
t
=
2

(E
b
) [V
ba
[
2
N
a
(t)


N
a
= N
a
(38)
where
=
W
aB
t
=
2

(E
b
) [V
ba
[
2
(39)
and thus is a transition rate (transition probability per unit time), and (38)
describes an exponential decay at this rate.
N
a
(t) = N
a
(0)e
t
(40)
Examples of where this holds: -decay, spontaneous emission from an excited
state of an atom or molecule.
13.3.4 Periodic Perturbations
Consider a harmonic interaction of the form:

V (t) =

A

e
it
+

Ae
it
with

A an unspecied time-independent operator. Then (23) becomes
W
ab
=
1

_
t
0
+t
t
0
d
__
E
b

E
a
_
e
i(
ba
)
+
_
E
b

E
a
_
e
i(
ba
+)
_

2
(41)
And the integral is largest when
ba
= 0 or
ba
+ = 0. These are resonance
conditions corresponding to energy conservation.
E
b
= E
a
+
E
b
= E
a

_
(42)
Physically, these represent an energy quanta absorbed or emitted (respectively) from
the system providing the periodic potential, as seen in Fig. 13.2.
Remark 1: We require that

t
so that the frequency is much greater than the
width of the transition resonance peak. This establishes that the system has enough
time to feel the periodicity of the perturbation.
13.3. TIME DEPENDENT PERTURBATION THEORY 207
(a) Resonance and
energy conservation
with absorption of a
quanta
(b) Resonance and
energy conservation
with emission of a
quanta
Figure 13.2: Resonances of Periodic Perturbations
Following through with the integration in (41) gives us:
W
ab
=
2

_
E
b

E
a
_

2
1 cos ((
ba
)t)
(
ba
)
2
+
2

_
E
b

E
a
_

2
1 cos ((
ba
)t)
(
ba
)
2
Considering transitions to a continuous band of nal states gives energy conserving
transition rates:
1. For the case where E
b
+E > E
a
+ > E
b
E:
W
ab
t
=
2

(E
a
+)

_
E
a
+

E
a
_

2
(43)
2. For the case where E
b
+E > E
a
> E
b
E:
W
ab
t
=
2

(E
a
)

_
E
a

E
a
_

2
(44)
Remark 2: Equations (43) and (44) give transition rates that are often referred to
as Fermis Golden Rules. The supplied by the perturbation provides the
energy to make the transition energy conserving.
13.3.5 Sudden and Adiabatic Perturbations
Two situations in which the response of the quantum systems changes is easy to
calculate is:
1. Sudden Approximation: Here we assume the system can not change its
state arbitrarily fast, so a sudden change in external eld will not change
the state of the quantum system during the eld change.
2. Adiabatic Perturbation: When perturbation is slow enough that the state
of the system changes adiabatically with applied perturbations.
208 CHAPTER 13. PERTURBATION THEORY
Chapter 14
Scattering
14.1 Scattering Cross Sections
Consider a general scattering experiment:
Where the particle ux J is the number of particles crossing unit area per unit time. Also, the
beam is weak enough so that the particles scatter independently (no coherence).
Remark 1: We neglect coherence of scattered waves between scattererseach scatterer acts as if
it were alone. This assumption fails for x-ray diraction.
We dene:
^ = # of particles recorded at the detector per unit time
^ = JN()d (1)
where () is the dierential scattering cross section, and is a parameter that characterizes the
collision between the incident particles and scatterers. The dierential refers to the detector only
monitoring a small solid angle d.
Remark 2: The total scattering cross section is

tot
=
_
() d (2)
209
210 CHAPTER 14. SCATTERING
Remark 3: Though the target can be quite complicated (an atom made up of electrons and
protons, etc...), it is often sucient to describe the scatterer by a central potential

V (r).
(if we consider elastic collions only!) This is what we will do in this section.
Remark 4: The goal is to relate what is measured in experiments (the dierential cross section),
to a solution to the Schrodinger equation.
Assume incoming particles have energy E, and momentum k. Then the stationary solution must
satisfy
_


2
2m

2
+

V (r)
_

k
= E
k
(3)
where

V (r) denes the interaction between particles and scatterers, and E is the energy of the
incident and outgoing particles. We want a solution that describes the scattering process, so we
look for a solution of the form:

k
(r, t) = e
iEt/

k
(r) (4)
with asymptotic form

k
(r) e
ikr
..
incoming
particles
+ f()
e
ikr
r
. .
outgoing spherical
wave of scattered
particles
(5)
as r . This means that
k
(r, t) gives an expanding spherical wave.
Remark 5: The e
ikr
represents a beam of particles with speed
k
m
and a density of 1 particle per
unit volume. The corresponding particle ux is
J = 1
k
m
The ux of particles entering the detector at angle (, ) is:
^ =
_
number of outgoing
particles per unit
volume
_

k
m
(
Detector
area
)
For large distances r, the rst term is
|f()|
2
r
2
and the detector area is r
2
d.
^ = [f()[
2
k
m
d (6)
Comparing (6) with (1) gives:
()
. .
measured
= [f()[
2
. .
calculated
Remark 6: Now the main task is to calculate what is called the scattering amplitude f().
14.1. SCATTERING CROSS SECTIONS 211
Remark 7: Often we are considering the scattering of individual particles o one another. In
this case, an incoming singe particle would be described by a time independent wave-packet
rather than a plane wave, and the picture would look like: where the big particle is being
scattered by the smaller. The incident particle (wavepacket) has velocity v =
k
m
=
2
m
.
The longitudial and transverse dimensions of the wavepacket are determined by the spread
in momentum.
In the above picture weve made the approximations that:
l, d a - otherwise the scattering depends on the details of the shape of the incoming
wavepacket.
l, d - otherwise the wavepacket description of the incoming particle will not give
well dened energy and momentum to be small compared to the scale on which the
scattering properties change with energy and momentum.
d b - for non-trivial results (if d b, then there is no signicant perturbation,
assuming short range potentials that fall o faster than
1
r
).
D a, - so that the inuence of the potential on the detector has dropped to 0
(detector is asymptotic regime).
Dsin d - so as to separate the transmitted and scattered parts of the incident
wave packet:
Mathematically, for the description we would multiply the incoming plane wave by a real envelope
function which depends on the distance from the scatterer (more precisely, on the impact parame-
ter). We would eventually derive a description of the scattered wave packet that depends on the
solution to the Schrodinger equation in a central potential. Given this description and the above
212 CHAPTER 14. SCATTERING
constraints, we can show that the dierential cross section (which depends on the probability for a
wavepacket to scatter within d of (location of the detector)) is still given by
() = [f
k
()[
2
(7)
where the asymptotic solution to the Schrodinger equation is:

k
(r) e
ikr
+f
k
()
e
ikr
r
(5)
14.2 Waves and Phase Shifts
Our goal now is to nd f() for a given central potential. Recall that in 10.2.2 and 8.2.3 we
found that the solution to the spherically symmetric Schrodinger equation could be written
(r, , ) =

l,m
A
lm
U
El
(r)
r
Y
lm
(, )
where U
El
is a solution to the radial equation (10.2.2.E2) and Y
lm
(, ) are the spherical harmonics.
Remark 1: The scattering problem has azimuthal symmetry. For initial propagation direction,
aligned with the z-axis, the solution is independent, and only m = 0 terms are involved.
Thus we have
(r, ) =

l=0
a
l
U
El
r
P
l
(cos ) (8)
where P
l
(cos ) is the Legendre polynomial and a
l
=
_
2l+1
4
A
l,0
.
Remark 2: The goal is to match (8) to the asymptotic form (5)
(r, )
r
e
ikr
+f()
e
ikr
r
where we are using arrows to represent asymptotically similar.
If we take k in the z-direction, the incident plane wave can also be expressed as:
e
ikr
=

l=0
(2l + 1)(i)
l
j
l
(kr)P
l
(cos ) (9)
(See 10.2.3.R11). Expanding f() in Legendre polynomials:
f() =

l=0
f
l
P
l
(cos ) (10)
14.2. WAVES AND PHASE SHIFTS 213
We require U
El
(r) to be a regular solution to the radial equation, and asymptotically, we can write
it as
U
El
(r)
r
sin
_
kr
l
2
+
l
_
(11)
where
l
is called a phase shift. The asymptotic form of (8) is:
(r, )
r

l=0
a
l
_
(i)
l
2i
e
i
l
e
ikr
r

(i)
l
2i
e
i
l
e
ikr
r
_
P
l
(cos ) (12)
In order to match this to the asymptotic form (5), we write the asymptotic form of j
l
(kr) in (9):
(r, )
r

l=0
_
(2l + 1)
(i)
l
kr
_
(i)
l
2i
e
ikr

(i)
l
2i
e
ikr
_
+f
l
e
ikr
r
_
P
l
(cos ) (13)
Now, (12) and (13) must match as there are 2 dierent forms of the same incoming and outgoing
spherical waves. This requirement gives an expression for a
l
:
Incoming Waves:
a
l
_
(i)
l
2i
e
i
l
_
= (2l + 1)
i
l
k
_
i
l
2i
_
a
l
=
(2l + 1)
k
i
l
e
i
l
(14)
Outgoing Waves: (matching coecients)
a
l
_
(i)
l
2i
e
i
l
_
=
(2l + 1)i
l
k
(i)
l
2i
+f
l

(2l + 1)e
i
l
k
e
i
l
2i
=
(2l + 1)
2ik
+f
l
f
l
=
(2l + 1)
2ik
_
e
2i
l
1
_
(15)
And the expression for the scattering amplitude in terms of the phase shifts is:
f() =
1
k

l=0
(2l + 1)e
i
l
sin(
l
)P
l
(cos ) (16)
Remark 3: Solving the dierential scattering cross-section is reduced to nding the phase shift
for the dierent angular momentum components of the scattered wave.
Remark 4: Note that we can write the asymptotic solution (13) with a
l
from (14).
(r, )
r

l=0
(2l + 1)
2i
_

(1)
l
e
ikr
kr
+e
2i
l
e
ikr
kr
_
P
l
(cos )
214 CHAPTER 14. SCATTERING
where this phase factor shifts the phase of the outgoing partial wave relative to the incident
plane wavethis is where the entire scattering process is captured. (Partial waves refer to
individual terms in the spherical wave expansion.)
Remark 5: The dierential scattering cross-section is terms of phase shifts is:
() = [f()[
2
=
1
k
2

l,l

(2l + 1)(2l

+ 1)e
i(
l

l
)
sin(
l
) sin(
l
)P
l
(cos )P
l
(cos ) (17)
Integrating this over a unit sphere gives the total scattering cross-section. Noting the
orthogonality of the Legendre polynomials:

tot
=
4
k
2

l=0
(2l + 1) sin
2
(
l
) (18)
We can write

tot
=

l=0

l
with

l
=
4
k
2
(2l + 1) sin
2
(
l
)
_

_
(19)
and
l
gives the contribution to scattering cross-section from partial waves with angular momentum
l.
Remark 6: The partial wave expansion is most useful when only a few terms in (18) contribute.
To get an idea of which terms dominate, consider a scattering (classical) problem. Here
V (r) = 0 for r > r
0
. Now the angular momentum of the incoming particle is L = bp
and is a constant of motion. If b > r
0
(L > r
0
p), then there is no scattering. Quantum
mechanically, L l (more precisely, L
2
l(l + 1)) and p = k. Thus, we expect the
strongest scattering when
l < kr
0
(20)
and the weakest scattering when
l > kr
0
(21)
Generally speaking, we expect
l
to become unimportant for l kr
0
. This holds if V (r) = 0
for r > r
0
, but turns out to be true also if V (r) is very small (though not strictly 0) at some
range r
0
.
14.3. EXAMPLE: SCATTERING FROM A HARD SPHERE 215
Remark 7: Equation (8) (weak scattering) holds when there is either low energy (k 0), or high
angular momentum (l ). Recall that high angular momentum particles will miss the
target if L > r
0
p. Physically, lower energy corresponds to large wavelength. In the limit
r
0

=
kr
2
1, a small scatterer will have little eect on a large incident wave. (Imagine a
small obstacle in sound and water waves.) Any scattering that does occur will be isotropic
since large wavelength waves cannot probe the spacial structure of the scatterer.
14.3 Example: Scattering from a Hard Sphere
Consider the potential:
V (r) =
_
, 0 < r < r
0
0, r
0
< r <
The general solution to the radial equation (10.2.3) is
R
l
(r) = A
l
j
l
(kr) +B
l

l
(kr)
where A
l
, B
l
are constants and the BCs are
A
l
j
l
(r
0
k) +B
l

l
(kr
0
) = 0
B
l
= A
l
j
l
(kr
0

l
(kr
0
)
R
l
(r) = A

l
(
l
(kr
0
)j
l
(kr) j
l
(kr
0
)
l
(kr))
where A

l
=
A
l

l
(kr
0
)
. The asymptotic form of the solution (10.3.1.R6) is
R
l
(r)
r

l
(kr
0
)
kr
sin
_
kr
l
2
_
+
j
l
(kr
0
)
kr
cos
_
kr
l
2
_
Comparing this with the general form (14.2.E11):
R
l
(r)
r
1
kr
sin
_
kr
l
2
+
l
_

r
1
kr
sin
_
kr
l
2
_
cos
l
+
1
kr
cos
_
kr
l
2
_
sin
l
216 CHAPTER 14. SCATTERING

l
(kr
0
) = cos
l
and j
l
(kr
0
) = sin
l
tan
l
=
j
l
(kr
0
)

l
(kr
0
)

l
= arctan
_
j
l
(kr
0
)

l
(kr
0
)
_
sin
2
(
l
) =
j
2
l
(kr
0
)

2
l
(kr
0
) +j
2
l
(kr
0
)
Remark 1: At low energies, the spherical Bessel functions become
j
l
(kr)
r0
(kr)
l
(2l + 1)!!

l
(kr)
r0
(2l 1)!!
(kr)
l+1
So that the partial cross section for scattering o a hard sphere is

l

k0
4
k
2
(2l + 1)(kr
0
)
4l+2
[(2l + 1)!!(2l 1)!!]
2
The dominant cross section is s-wave scattering (l=0):

l

k0
4
k
2
(kr
0
)
2
= 4r
2
0
where r
0
is the length scale characterizing the range of the scatterer, and is sometime called
the scattering length.
Remark 2: At high energies, well use the asymptotic forms of the Bessel functions (see Messiah)
to get:
()
k
1
4
r
2
0
_
1 + cot
2
(

2
)J
2
1
(kr
0
sin )
_
()
and

tot

k
2r
2
0
This limit corresponds to very short wavelengths, from which we might expect the classical
results:
() =
1
4
r
2
0

tot
= r
2
0
_
_
_
Classical results
We do not get these results since the wave nature of the scattering process does not com-
pletely vanish. This is because the hard sphere has sharp edges, so that any wavelength,
no matter how short, is aected by the discontinuity and there are diraction eects. The
Bessel function in () is the diraction term.
14.4. SCATTERING FROM A SQUARE WELL RESONANCES 217
14.4 Scattering from a Square Well Resonances
Consider the potential
V (r) =
_
V
0
, r < r
0
0, r > r
0
Remark 1: From before we know that a 1D square well exhibits resonances which are very sharp
when the well is very deep and incident particles have low energies. Similar resonances
occur when scattering from a 3D well.
Remark 2: Recall from 10.3.2 that regular solutions inside the square well are
R
l
(r) = j
l
(r) (r < r
0
)
with
=
_
2m

2
(E +V
0
)
and we have chosen to normalize R
l
(r) for convenience. Outside of the well, the solution is
a superposition of the Hankel functions.
R
l
(r) = A
l
h
(+)
l
(kr) +B
l
h
()
l
(kr) (r > r
0
)
with
k =
_
2mE

2
These solutions correspond to E > 0 since we are interested in the unbound states outside
the well for scattering phenomena. The boundary conditions require continuity at r = r
0
:
j
l
(r
0
) = A
l
h
(+)
l
(kr
0
) +B
l
h
()
l
(kr
0
) (1)
and continuity of the derivatives at r = r
0
:
j

l
(r
0
) = k
_
A
l
h
(+)
l
(kr
0
) +B
l
h
()
l
(kr
0
)
_
(2)
218 CHAPTER 14. SCATTERING
Remark 3: We also want to match the asymptotic form of the solution outside the well to the
scattering solution:
R
l
(r)
r
1
kr
sin
_
kr
l
2
+
l
_

r
1
kr
1
2i
_
e
i(krl/2)
e
i
l
e
i(krl/2)
e
i
l
_
Inserting the asymptotic form of the Hankel functions:
R
l
(r)
r
1
kr
_
A
l
e
i(krl/2)
+B
l
e
i(krl/2)
_
Comparing the forms gives
A
l
= e
i
l
, B
l
= e
i
l
e
2i
l
=
A
l
B
l
So we need only the ratio of the unknown coecients. From (1) then,
A
l
B
l
=
1
B
l
j
l
(r
0
h
(+)
l
(kr
0
)

h
()
l
(kr
0
)
h
(+)
l
(kr
0
)
Now, taking (1) kh
(+)
l
(kr
0
) (2) h
(+)
l
(kr
0
):
kj
l
(r
0
)h
(+)
l
(kr
0
) j

l
(r
0
)h
(+)
l
(kr
0
) = A
l
k
0
..
_
h
(+)
l
(kr
0
)h
(+)
l
(kr
0
) h
(+)
l
(kr
0
)h
(+)
l
(kr
0
)
_
+
+B
l
k
_
h
()
l
(kr
0
)h
(+)
l
(kr
0
) h
()
l
(kr
0
)h
(+)
l
(kr
0
)
_
= B
l
k
_
2i
kr
2
0
_
where the last term is from the Wronskian relation of Hankel functions (Messiah X.38).
e
2i
l
=
h
()
l
(kr
0
)
h
(+)
l
(kr
0
)

j
l
(r
0
)
h
(+)
l
(kr
0
)

k2i
(kr
0
)
2
1
kj
l
(r
0
)h
(+)
l
(kr
0
) j

l
(r
0
)h
(+)
l
(kr
0
)
=
h
()
l
(kr
0
)
h
(+)
l
(kr
0
)
_
_
1 2i
k

1
(kr
0
)
2
1
h
()
l
(kr
0
)
_
k

h
(+)
l
(kr
0
)
j

l
(r
0
)
j
l
(r
0
)
h
(+)
l
(kr
0
)
_
_
_
Remark 4: For a deep potential and moderately low energies:
r
0
l and k
14.4. SCATTERING FROM A SQUARE WELL RESONANCES 219
So
e
2i
l

h
()
l
(kr
0
)
h
(+)
l
(kr
0
)
=
1 ij
l
(kr
0
)/
l
(kr
0
)
1 +ij
l
(kr
0
)/
l
(kr
0
)
=
1
_
j
l
(kr
0
)
n
l
(kr
0
)
_
2

2ij
l
(kr
0
)

l
(kr
0
)
1 +
_
j
l
(kr
0
)

l
(kr
0
)
_
2
(3)
= cos(2
l
) +i sin(2
l
)
tan
l

j
l
(kr
0
)

l
(kr
0
)
which is the same result as for the hard sphere.
Remark 5: For certain energies, the denominator in the second term of (3) vanishes. In this case,
note that
j

l
(r
0
)
j
l
(r
0
)

r
0
l
cot (r
0
l/2)
For simplicities sake, let kr
0
l so that
h
()
l
(kr
0
)
h
(+)
l
(kr
0
)

kr
0
l
e
2i(kr
0
l/2)
(kr
0
)
2
h
()
l
(kr
0
)h
(+)
l
(kr
0
)
kr
0
l
1
(kr
0
)
2
h
()
l
(kr
0
)h
(+)
l
(kr
0
)
kr
0
l
i
Thus (3) becomes
e
2i
l
e
2i(kr
0
l/2)
_
1
2i
k

i
k

cot
_
r
0

l
2
_
_
e
2i(kr
0
l/2)
cot(r
0
l/2) +i
k

cot(r
0
l/2) i
k

l
= (kr
0
l/2) +
where
tan =
k

cot(r
0
l/2)
Now is generally negligible when
k

1, unless r
0
l/2 = (2m+1)/2, where m is an
integer. When we have this secondary case, we have resonances.
220 CHAPTER 14. SCATTERING
The distance between resonances is r
0
= .

_
_
2m

2
(E + E +V
0
)
_
2m

2
(E +V
0
)
_
r
0
=

_
1 +
m

2
E

2
1
_
r
0
=
E =

2

mr
0
Assuming k V
0
E
2

2m

2
V
0

2
0
:
E
2V
0
mr
0
Since
Which implies
where in the range (), tan goes to 0 and 0 which corresponds to a change of in
. So
l
= (kr
0
l/2) + increases by each resonance. In between resonances, background
scattering is determined by
l
= (kr
0
l/2) +.
Chapter 15
Relativistic Quantum Mechanics: The
Dirac Equation
The Dirac equation is a relativistic description of a quantum system that includes the relativistic
nature of some systems (like high-Z atoms) as well as spin and spin-orbit coupling.
15.1 Spinless Free Particles - The Klein-Gordan Equation
Recall that in deriving the Schrodinger equation, we started with the kinetic energy:
H =
p
2
2m
and let
p p
H i

t
_
_
_
i

t
[) =
p
2
2m
[)
In deriving a relativistic description, let
H
2
= c
2
p
2
+m
2
c
4
So that the substitution of operators gives

2
t
2
[) =
1

2
_
c
2
p
2
+m
2
c
4
_
[) ()
This is the Klein-Gordan equation.
Remark 1: Note that we considered H
2
rather than H =
_
c
2
p
2
+m
2
c
4
. This is because the
former gives a symmetric description of space and time as desired by relativity. Otherwise,
expanding H in a momentum basis:
H = mc
2
_
1 +
1
2
p
2
m
2
c
2

1
8
p
4
m
4
c
4
+
_
221
222 CHAPTER 15. RELATIVISTIC QUANTUM MECHANICS: THE DIRAC EQUATION
gives even and odd derivatives not on an equal footing.
Remark 2: () is a good description for spinless particles, but is insucient for particles with
S ,= 0.
15.2 Dirac Equation for Free Particle
The Dirac equation for a free particle is given by
i

t
[) =
_
c p +

mc
2
_
[) (1)
where
=
_
0
0
_
,

=
_

I 0
0

I
_
(2)
where
is the Pauli spin matrices


I is the 2x2 identity
c is the speed of light
Remark 1: In the derivation of the Klein-Gordan equation, we took H H
2
to get the space-time
symmetry. Here, Dirac assumed that
_
c
2
p
2
+m
2
c
4
p+ something without momentum
to get space-time symmetry in an equation of 1st order in space and time.
c
2
p
2
+m
2
c
4
= c p +

mc
2
where ,

are determined by matching coecients:
c
2
_
p
2
x
+p
2
y
+p
2
z
_
+m
2
c
4
= c
2
( p)
2
+mc
3
_
p

+

p
_
+

2
m
2
c
4
Matching coecients gives the following relations:
1.
2
i
=

2
= I where i = x, y, z
2.
i

j
+
j

i
= [
i
,
j
]
+
= 0 for i ,= j
3.


i
+
i

=
_

,
i
_
+
= 0
Remark 2: From these relationships, we conclude that ,

are traceless Hermitian matrices with
eigenvalues = 1.
Remark 3: Traceless with eigenvalues=1 even dimensional. But they cannot be the 2x2 Pauli
matrices (since there are 4 matrices here, and there are only 3 Pauli matrices with these
properties) try 4x4 matrices
15.3. ELECTROMAGNETIC INTERACTIONS 223
Remark 4: The choice of and

given in (2) satisfy these requirements, though it is not unique
(since the properties of Remark 2 are preserved under a unitary transformation).
Remark 5: [) in (1) is called the Lorentz Spinor and has 4 components (to be compatible with
4x4 s and

s).
Remark 6: [) = a constant since the Hamiltonian is Hermitian.
15.3 Electromagnetic Interactions
The classical Hamiltonian for a particle with charge q is:
H =
__
p
q
c
A
_
c
2
+m
2
c
4

1/2
+q
where and Aare the electrodynamic scalar and vector potentials. Generalizing the Dirac equation
so that p p
q
c
A and adding the electromagnetic potential energy:
i

t
=
_
c
_
p
q
c
A
_
+

mc
2
+q
_
(1)
15.3.1 Electron Spin and Magnetic Moment
To see how spin and magnetic moment drop out of the Dirac equation, it is sucient to
let = 0 and keep only O(
v
c
)
2
terms. Let (t) = e
iEt/
to nd energy eigenstates.
Plugging this into (1):
i
_

iE

_
e
iEt/
=
_
c
_
p
q
c
A
_
+

mc
2
_
e
iEt/

E =
_
c

+

mc
2
_
(2)
where

= p
q
c
A (3)
Remark 1:

is called the kinetic momentum operator.
Remark 2: Recall that is a 4 component spinor and that ,

are given by
=
x
and

=
z


I
which suggests that we can write
=
_

_
(4)
224 CHAPTER 15. RELATIVISTIC QUANTUM MECHANICS: THE DIRAC EQUATION
where and are 2 component spinors.
Plugging (4) into (2):
E
_

_
=
_
mc
2

I c

c

mc
2

I
_ _

_
_
(E mc
2
)

I c

c

(E +mc
2
)

I
_ _

_
=
_
0
0
_
which gives a set of coupled equations for and :
(E mc
2
) = c

(5)
(E +mc
2
) = c

(6)
From (6):
=
c

E +mc
2
(7)
Remark 3: In the non-relativistic limit,

c
_
p
q
c
A
_
E
s
+mc
2
+mc
2

c(mv)
2mc
2

1
2
v
c
1
where
E
s
is the Schrodinger energy mc
2
mv is the typical momentum
So in the non-relativistic limit, is the small component and is the big component.


2mc
(8)
In the non-relativistic limit, (5) becomes:
E
s
= c

=
1
2m
_

__

_
(9)
=
1
2m
_


+i

and


=
iq
c
B. So now (9) becomes
_
1
2m
_
p
q
c
A
_
2

q
2mc
B
_
= E
s
(10)
which describes a spin
1
2
particle in a magnetic eld. The energy splitting (Zeemann
eect) gives
E =
eB
mc
_

2
_
and the picture in Fig. 15.1:
Here 2
B
B = 2E = g
B
B g = 2s + 1, and we see that electron spin and magnetic
moment are accounted for in the Dirac equation.
15.4. THE DIRAC EQUATION FOR THE HYDROGEN ATOM 225
Figure 15.1: Zeemann splitting
15.4 The Dirac Equation for the Hydrogen Atom
For the hydrogen atom:
V = e =
e
2
r
and the Dirac equation is
i

t
=
_
c
_
p
q
c
A
_
+

mc
2

e
2
r
_

We still decompose into the 2 component spinors and obtain equations for and (15.3 (5)
and (6) with E E V ).
(E V mc
2
) = c p (1)
(E V +mc
2
) = c p (2)
Remark 1: In absence of a B eld, we can let A = 0 so

p, and hence the above results in
(1) and (2).
From (2), we have
= (E V +mc
2
)
1
c p (3)
which is substituted into (1), giving:
(E

V +mc
2
) = c p(E

V +mc
2
)
1
c p (4)
Remark 2: Note that we were careful to keep the order correct, as p can operate of

V , so order
does matter.
Remark 3: Solving (4) and plugging into (3) gives the solution to the full Lorentz spinor .
Remark 4: Keep O(
v
c
)
2
, (4) becomes:
(E
s


V ) = c p(E


V + 2mc
2
)
1
c p

c
2
2mc
2
( p) ( p)
=
1
2m
p
2

which is just the non-relativistic Schrodinger equation.


226 CHAPTER 15. RELATIVISTIC QUANTUM MECHANICS: THE DIRAC EQUATION
15.4.1 Fine Structure
Recall that in solving the Schrodinger equation for the hydrogen atom, we resorted to
perturbation theory to get the ne structure of the energy eigenspectrum (previous
HW). The time structure was due to relativistic eec8ts which emerge naturally from
the Dirac equation. Expanding (4) to O(
v
c
)
4
(one order higher than that required
to recover the Schrodinger equation), and letting E = E
s
+ mc
2
in (4) with E
s
the
solution for the Schrodinger equation gives:
(E
s
V + 2mc
2
)
1
=
1
2mc
2
_
1 +
E
s
V
2mc
2
_
1
(15.1)

1
2mc
2
_
1
E
s
V
2mc
2
_
(15.2)
=
1
2mc
2

E
s
V
4m
2
c
4
(15.3)
And (4) becomes
(E
s
V ) = c p
_
1
2mc
2

E
s
V
4m
2
c
4
_
c p
E
s
=
_

_
1
2m
( p)
2
. .
p
2
+V
p(E
s
V )( p)
4m
2
c
4
_

_
E
s
=
_

_
p
2
2m
+V
p(E
s
V )( p)
4m
2
c
4
. .
extra piece giving ne structure
_

_
(5)
Remark 1: Note that (5) is not simply a more complicated Schrodinger equation as
E
s
is on both sides of the equation.
Remark 2: From 15.4.R4, the Schrodinger equation is recovered from O(
v
c
)
2
, so:
E
s
=
_
p
2
2m
+V
_
are O(
v
c
)
2
(E
s
V ) =
p
2
2m
+O(
v
c
)
4
since
p(E
s
V )( p)
4m
2
c
4

v
2
c
2
(E
s
V )

v
2
c
2
_
p
2
2m
+O(
v
c
)
4
_
15.4. THE DIRAC EQUATION FOR THE HYDROGEN ATOM 227
So in the third term of the rhs of (5), it will be sucient to let (E
s
V )
p
2
2m
. But we have
(E
s
V ) p
..
awkward

So
(E
s
V ) p = [(E
s
V ) p]
= pE
s
+
_

V p + p

V p

V
_
= p(E
s
V ) +
_
p,

V
_

which we substitute into (5):


E
s
=
_
_
p
2
2m
+

V
( p)
2 p
2
2m
+ ( p)
_

_
p,

V
__
4m
2
c
4
_
_

=
_
_
p
2
2m
+

V
p
4
8m
3
c
2

p
_
p,

V
_
4m
2
c
2

i p
_
p,

V
_
4m
2
c
2
_
_

=

H (6)
Remark 3: (6) now looks like the Schrodinger equation!
Remark 4: On the rhs of (6), the 3rd term is the relativistic correction to the kinetic
energy.
Remark 5: The 5th term can be manipulated:
[ p, f(x)] = i

x
f +if

x
= i
_
f
x
_
if

x
+if

x
= i
_
f
x
_
228 CHAPTER 15. RELATIVISTIC QUANTUM MECHANICS: THE DIRAC EQUATION

i p
_
p,

V
_
4m
2
c
2
=
i
4m
2
c
2
p
_

_
i
_

e
2
r
_
. .
+e
2
r
r
3
_

_
=
e
2
4m
2
c
2
1
r
3
p r
=
e
2
4m
2
c
2
1
r
3
r p
where it is ok to switch r and p since the p operator never operates on the conjugate
coordinate.
=
e
2
2m
2
c
2
1
r
3

S

L
=

H
S.O.
the spin-orbit interaction
where it is ok to switch r and p since the p operator never operates on the
conjugate coordinate.
Remark 6: The 4rth term in (6) is not Hermitian.

_
[[
2
d
3
r ,= constant
is not a good Schrodinger wave function
15.4. THE DIRAC EQUATION FOR THE HYDROGEN ATOM 229
But what is required is
_

d
3
r =
_
_
[[
2
+[[
2

d
3
r = constant
=
_
_
[[
2
+
( p)

( p)
4m
2
c
2
_
d
3
r
=
_
_
[[
2
+

( p)

( p)
4m
2
c
2
_
d
3
r
=
_ _
[[
2
+

p
2

4m
2
c
2
_
d
3
r
=
_

_
1 +
p
2
2m
2
c
2
_
d
3
r
=
_

__
1 +
p
2
8m
2
c
2
__
1 +
p
2
8m
2
c
2
__
d
3
r
=
_

_
_
1 +
p
2
8m
2
c
2
_

_
1 +
p
2
8m
2
c
2
_
_
d
3
r
=
_ _
1 +
p
2
8m
2
c
2

_
1 +
p
2
8m
2
c
2

_
d
3
r
=
_
[
s
[
2
d
3
r
So to O(
v
c
)
4
,
s
has a constant norm. But now we must write (6) in terms of

s
.

s
=
_
1 +
p
2
8m
2
c
2
_
(7)
Since E
s
=

H by (6), we have that
E
s
_
1 +
p
2
8m
2
c
2
_
1

s
=

H
_
1 +
p
2
8m
2
c
2
_
1

s
E
s

s
=
_
1 +
p
2
8m
2
c
2
_

H
_
1 +
p
2
8m
2
c
2
_
1

_
1 +
p
2
8m
2
c
2
_

H
_
1
p
2
8m
2
c
2
_

s
=
_

H +
1
8m
2
c
2
_
p
2

H

H p
2
_
_

s
+O(
v
c
)
6

H +
1
8m
2
c
2
_
p
2
,

H
_
_

s
=

H
s

s
(8)
230 CHAPTER 15. RELATIVISTIC QUANTUM MECHANICS: THE DIRAC EQUATION


H
s
=

H +
1
8m
2
c
2
_
p
2
,

H
_


H +
1
8m
2
c
2
_
p
2
,

V
_
+O(
v
c
)
6


H
s
=

H +
1
8m
2
c
2
_
p
2
,

V
_
(9)
Remark 7: Equation (9) gives an extra term that comes from creating a suitable
wavefunction with constant norm, so that (8) is a suitable Schrodinger equation
that includes relativistic eects in the Dirac equation. Combining this extra
term in (9) with the non-Hermitian 4rth term in (6) gives:
1
8m
2
c
2
_
p p,

V
_

1
4m
2
c
2
p
_
p,

V
_
=
1
8m
2
c
2
_
p
_
p,

V
_
+
_
p,

V
_
p 2 p
_
p,

V
__
where we used [AB, C] = A[B, C] + [A, C]B.
=
1
8m
2
c
2
_
p
_
p,

V
_
+
_
p,

V
_
p
_
=

2
8m
2
c
2

V
=
e
2

2
8m
2
c
2

2
_
1
r
_
=
e
2

2
2m
2
c
2

3
(r)
=

H
D
which is the Darwin term!
Remark 8: Putting these together (Relativistic Kinetic Energy, Spin-Orbit Interac-
tion, and Darwin) gives the potentials that, in perturbation theory, resulted
in the ne structure for the Hydrogen atom.
Remark 9: Recall that the Darwin term contributed when l = 0 and the S.O.
coupling contributed only when l ,= 0 such that the contribution gave the
same result independent of l. This is in spite of the fact that

H
D
has nothing
to do with angular momentum. Physically, this term comes because a particle
can not be localized more that

mc
(the compton wavelength).
V (r)

V (r) = some smeared average about r
Remark 10: Solving the full Dirac equation (rather than just to O(
v
c
)
4
), gives the
15.5. NEGATIVE ENERGY SOLUTIONS 231
energy spectrum:
E
nj
= mc
2
_
_
1 +
_

n (j +
1
2
) +
_
(j +
1
2
)
2

1/2
_
2
_
_
1/2
(11)
Expanding in powers of :
O(
0
) rest energy
O() Schrodinger energy
O(
2
) ne structure (compare with perturbation results)
Remark 11: Equation (11) is still not 100% correct when compared to experiment
(One can not account for Lamb shift for example). To get this correct, the
E + M eld itself must also be quantized (i.e. the eld must be treated as a
bunch of harmonic oscillators).
15.5 Negative Energy Solutions
In general
Relativity
. .
particle production
with enough energy
+Quantum Mechanics
. .
arbitrarily enough
energy for short time
=
the possible outcome of a system
of 1 or several particles
Question: How does Dirac theory account for this?
Answer: With Negative Energy Solutions!
For simplicity, let = c = 1, then the free particle Dirac equation is:
i

t
=
_
p +

m
_

=
_

_
E m p
p E +m
_ _

_
=
_
0
0
_
or
(E m) = p
(E +m) = p
=
p
2
E
2
m
2

p
2
E
2
m
2
= 1
E
2
= p
2
+m
2
E =
_
p
2
+m
2
()
232 CHAPTER 15. RELATIVISTIC QUANTUM MECHANICS: THE DIRAC EQUATION
Remark 1: We can not neglect the negative energy solutions, though note that if p = 0 E =
mc
2
.
But what do we do with the negative energy solutions?
Even if +mc
2
states are stable, quantum uctuations should be enough to allow transitions to
negative energy states. The problem is that this is not observed.
15.5.1 Dirac Solution
The vaccuum is really occupied by fermions, which dont allow positive energy particles
to occupy negative energy states (Pauli-exclusion principle). Consider
Remark 1: The lled Dirac sea is unobservable, but the hole is observableit has
change +e and [E[ mc
2
, and is called a positron.
Remark 2: When an electron jumps back into the hole left in the Dirac sea, 2
particles are destroyed (an electron and a positron), but an energy 2mc
2
is
liberated in the form of a photon.
Remark 3: Note that this explanation only works if the particles are fermions. But
bosons can have negative energy solutions (and the negative energy solutions
of the spinless Klein-Gordan equation are what motivated Dirac to look for a
rst order theory in the rst place).
15.5.2 Feynmanns Solution
Negative energy solutions only travel backwards in time.
15.5. NEGATIVE ENERGY SOLUTIONS 233
Remark 1: We still have the notion of an anti-particle (positron in this case). This
is how things progress:
What we observe is the positron (particles with positive energy) created at t
d
and destroyed
at t
c
. This implies that the creation of the positive energy positron coincides with the
destruction of the negative energy electron (and vice-versa).
234 CHAPTER 15. RELATIVISTIC QUANTUM MECHANICS: THE DIRAC EQUATION
Chapter 16
Feynmann Path Integral Formulation
of Quantum Mechanics
In the Schrodinger approach, we rst nd eigenvalues and eigenvectors of the Hamiltonian, then
determine the propagator,

U(t) (the full solution). In the path integral approach,

U(t) is found
directly.
16.1 Path Integrals - Concepts and the Classical Action
Here we consider only a single particle in 1-D (the generalization to higher dimensions is straight
forward).
where the initial and nal points are xed.
Recipe:
1. Draw all the paths between (x
0
, t
0
) and (x, t).
2. Calculate the action S[x, t] for each path.
235
236 CHAPTER 16. FEYNMANN PATH INTEGRAL
3. U(x, t; x
0
, t
0
) = A

all paths
e
iS[x,t]/
, where A is a normalization factor.
Remark 1: Recall that the action is a functional dened by
S =
_
t
t
0
L(t

) dt

(1)
where L(t) is the Lagrangian (usually L = T U). Also, a functional eats a function and
excretes a number.
Remark 2: Note that the propagator is not a weighted sum, thus, the classical action does not
seem to be more likely (as it should in the classical limit). However, the paths are all treated
as phases, which add together constructively near the path that corresponds to the classical
action, and destructively for paths away from the classical action path by more than

.
Example: For a free particle with x = t:
Here the alternate path is x = t
2
. Letting (x
0
, t
0
) = (0, 0) and (x, t) = (1 cm, 1 s):
S
cl
=
_
1
0
_
1
2
m x
2

dt =
..
x=t x=1
1
2
m
_
1
0
dt =
1
2
m
S
alt
=
_
1
0
_
1
2
m x
2

dt = 2m
_
1
0
t
2
dt =
2
3
m
For a classical particle, m 1 gram, which implies that the phase changes by
S
alt
S
cl

=
_
2
3

1
2
_
m

=
m
6
1.6 10
26

So we can ignore non-classical paths. For an electron, m 10


27
grams, implying that
S
alt
S
cl

=
m
6
.16
and so we can not ignore non-classical paths.
16.2.

U(T) FOR A FREE PARTICLE 237
16.2

U(t) for a free particle
Recall that propagating

U(t) as the solution to the time-dependent Schrodinger equation gave
U(x, t; x
0
, t
0
) U(x, t; x
0
) =
_
m
2it
_
1/2
e
im(xx
0
)
2
/2t
(1)
It turns out that for potentials of the form
V = a +bx +cx
2
+d x +ex x
U(t) = A(t)e
iS
cl
/
(2)
with A(t) a normalization (HW). A free particle is of this form, so show that (2) is sucient to
obtain the exact answer (1). For a free particle: a = b = c = d = e = 0. So,
S =
_
t
N
t
0
L(t) dt =
_
t
N
t
0
1
2
m x
2
dt (3)
Imagine that the path is broken up into n-discrete steps so that as N , the path is continuous:
Remark 1: In this picture, draw all the paths from (x
0
, t
0
) to (x
N
, t
N
) corresponds to using all
possible values of x at a given t (like moving the beads up and down on an abacus).
238 CHAPTER 16. FEYNMANN PATH INTEGRAL
Remark 2: Calculate the action for all possible paths is represented in the discretized form by
S =
N1

t=0
m
2
_
x
i+1
x
i

_
2
with = t (4)
Remark 3: Finally, calculate the propagator as a sum of all paths. Symbolically, this is written
as
_
x
N
x
0
e
iS[x(t)]/
T[x, t] (5)
where
_
x
N
x
0
T[x, t] Integrate between all paths between x
0
, x
N
In this case, we write
U(x
N
, t
N
; x
0
, t
0
) =
_
x
N
x
0
e
iS[x,t]/
T[x(t)]
= lim
N
0
A
_

dx
1
_

dx
2

_

dx
N1
exp
_
im
2
N1

i=0
(x
i+1
x
i
)
2

_
Scaling x by
_
m
2
:
y
i
=
_
m
2
_
1/2
x
i
= lim
N
0
_
2
m
_
1
2
(N1)
_

dy
1
_

dy
2

_

dy
N1
e

P
N1
i=0
(y
i+1
y
i
)
2
i
Taking terms with y
1
implies
_

dy
1
e
(y
1
y
0
)
2
i
e
(y
2
y
1
)
2
i
=
_
i
2
e
(y
2
y
0
)
2
2i
Inserting this result into the next integral...
_
i
2
_

dy
2
e
(y
3
y
2
)
2
/i
e
(y
2
y
0
)
2
/2i
=
_
(i)
2
3
_
1/2
e
(y
3
y
0
)
2
/3i
Doing all the N 1 integrations gives:
U(x
N
, t
N
; x
0
, t
0
) = lim
N
0
A
_
2
m
_
1
2
(N1)
(i)
1
2
(N1)
N
1/2
e
(y
N
y
0
)
2
/Ni
= lim
N
0
A
_
2i
m
_
(N1)/2
N
1/2
e
(x
N
x
0
)
2
m/2iN
= lim
N
0
A
_
2i
m
_
N/2 _
m
2iN
_
1/2
e
(x
N
x
0
)
2
m/2iN
16.2.

U(T) FOR A FREE PARTICLE 239
Remark 4: Taking the limit N and 0 gives N = t, which is the exact answer in (1) if
A =
_
2i
m
_
N/2
This means we can write
_
T[x, t] lim
N
0
_
m
2i
_
1/2
_

dx
1
_
m
2i
_
1/2
_

_

dx
N1
_
m
2i
_
1/2
(6)
Remark 5: The propagator, U(x
f
, t
f
; x
i
, t
i
) is sometimes called the Feynmann amplitude.

You might also like