2017F Phys363 Readings PDF
2017F Phys363 Readings PDF
AF
Intermediate Classical Mechanics Course Notes
Christopher O’Donovan
[email protected]
Version 0.3.9s
November 15, 2017
DR
ii
T
AF
DR
This work is licensed under the “Creative Commons Attribution Non-Commercial No Derivatives” license. http:
//creativecommons.org/.
T
Contents
1 Physics Review 1
1.1 Newton’s Second Law for Linear Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2
AF 1.1.1 Oscillating Systems . . . . . . . . .
1.2 Newton’s Second Law for Rotating Systems
1.2.1 Central Forces . . . . . . . . . . .
1.2.2 CoM Coordinates & Reduced Mass
1.2.3 Fixed Axis Rotations . . . . . . . .
1.2.4 Rigid, Planar Bodies . . . . . . . .
1.2.5 Parallel Axis Theorem . . . . . . .
1.2.6 Rotational Kinetic Energy . . . . .
1.3 Central Forces . . . . . . . . . . . . . . . .
1.3.1 Effective Potentials . . . . . . . . .
1.3.2 Rolling . . . . . . . . . . . . . . .
1.3.3 Further Problems . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
8
9
9
9
9
10
13
14
18
19
21
23
25
25
32
2.2.1 Generalized Eigenvectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.3 Further Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
DR
3 The Inertia Tensor 41
3.1 Vectors and their Transformation Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.2 Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.2.1 Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.2.2 Rotational Kinetic Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.2.3 Multilinearity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.3 Tensors and their Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.3.1 Transforming Tensors between Representations . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.4 Tensor Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.4.1 The Divergence Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.5 The Inertia Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.5.1 Properties of the Inertia Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.6 Useful Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.6.1 Parallel Axis Theorem (kAT) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.6.2 Principal Axes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.6.3 Perpendicular Axis Theorem (⊥AT) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.6.4 A Handy Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.7 Further Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
iii
iv CONTENTS
4 Rotational Kinematics 63
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.2 Accelerating Reference Frames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.2.1 Inertial Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.3 Rotating Reference Frames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
T
4.3.1 Rotation Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.3.2 Angular Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.3.3 General Vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.3.4 Newton’s Second Law in Non-Inertial Reference Frames . . . . . . . . . . . . . . . . . . . . 67
4.3.5 Effective Potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.4 General Motion of a Rigid Body . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.4.1 The Instantaneous Axis of Rotation (instantaneous axis of rotation (IAOR)) . . . . . . . . . . 69
4.4.2 Rolling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.5 Further Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
7
AF Midterm Review
5.1 Coupled Oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2 Inertia Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3 Rotational Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Virtual Work
~ . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
75
75
77
82
87
87
87
89
89
92
97
97
97
98
7.2 D’Alembert’s Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
7.2.1 Applied, Constraint, and Inertial Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
7.3 Constraint Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
DR
7.4 Further Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
T
10.3.1 Several Dependent Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
10.3.2 Variational Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
10.3.3 That Interesting Property of Lagrangians . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
AF
A Vectors
A.1 Cartesian Coordinates . . . . . .
A.2 Curvilinear Coordinates . . . . .
A.2.1 Polar Coordinates . . . .
A.2.2 Cylindrical Coordinates
A.2.3 Spherical Coordinates .
B Index Notation
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
165
B.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
B.2 Covariant & Contravariant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
B.3 Further Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
C Vector Multiplication
169
175
F Rotations 183
DR
F.0.1 3D Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
F.0.2 Active and Passive Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
M Formulas 217
M.1 Analytical Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
M.1.1 Lagrangian Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
M.1.2 Hamiltonian Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
M.1.3 Small Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
T
M.1.4 Non-inertial Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
M.1.5 Inertia Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
M.1.6 Rotational Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
M.2 Helpful Identities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
M.2.1 Cylindrical Coordinates
~r = ~r(s, φ, z) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
M.2.2 Spherical Coordinates
~r = ~r(r, θ, φ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
M.2.3 Vector Identities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
M.2.4 Tensor Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
AF M.2.5 Trigonometry . . . . . .
M.2.6 Series . . . . . . . . . .
Glossary
Index
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
219
219
219
222
DR
T
List of Examples
AF
1.4
1.5
1.6
1.7
1.8
1.9
1.10
1.11
2.1
2.2
2.3
2.4
2.5
3.1
Motion of a Charged Particle in a Magnetic Field
MoI of a Hoop . . . . . . . . . . . . . . . . . .
Thin Hemispherical Shell CoM . . . . . . . . .
Sphere on a Stick . . . . . . . . . . . . . . . . .
Conical Pendulum II . . . . . . . . . . . . . . .
Conical Pendulum - III . . . . . . . . . . . . . .
Rolling down an Incline I . . . . . . . . . . . .
Ferris Wheel . . . . . . . . . . . . . . . . . . .
Complex Oscillators . .
State Vector Oscillators .
Two Coupled Oscillators
Energetic Oscillators . .
Triple Pendulum . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
6
11
11
14
15
19
22
23
25
27
32
37
40
47
3.2 Cauchy Stress Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.3 Sheer Delight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.4 Moments of Inertia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
DR
3.5 Easy Cube . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.6 Inertia Tensor of a Cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.7 Inertia Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.8 Inertia Tensor of a Rectangular Plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.9 Spinning Plate Redux . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.10 The Maxwell Field Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.1 Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
6.2 Torque Free, Axial Symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
7.1 Leverage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
vii
viii LIST OF EXAMPLES
T
7.7 Ropes and Pulleys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
7.8 Inclined Pulleys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
AF 9.1
9.2
9.3
9.4
9.5
9.6
9.7
10.1
10.2
10.3
10.4
10.5
10.6
10.7
10.8
Bead on Wire I . . . . . . . . . . .
Bead on Wire II . . . . . . . . . . .
Bead on a Wire III . . . . . . . . .
Spherical Pendulum IV . . . . . . .
Electromagnetic Interaction I . . .
Bead on a Parabolic Wire . . . . .
Bead on a Rotating Parabolic Wire .
Shortest Distance . . . . . . . . . . . . .
Shortest Distance . . . . . . . . . . . . .
Pens and Pencils . . . . . . . . . . . . .
Surface Area . . . . . . . . . . . . . . .
The Brachistochrone Problem . . . . . .
The Brachistochrone Problem, Revisited
Falling . . . . . . . . . . . . . . . . . .
Lagrangian Optics . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
125
127
130
130
132
136
136
137
140
140
142
143
146
146
147
10.9 Isoperimetric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
12.14Multiplier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
T
AF
DR
x LIST OF EXAMPLES
T
AF
DR
T
List of Exercises
AF
1.4
1.5
1.6
1.7
1.8
2.1
2.2
2.3
2.4
3.1
3.2
3.3
3.4
3.5
Thick Hemispherical Shell CoM . . . . . . . . . . . . . .
Various MoI . . . . . . . . . . . . . . . . . . . . . . . .
Sliding Ball . . . . . . . . . . . . . . . . . . . . . . . . .
Bead on Wire . . . . . . . . . . . . . . . . . . . . . . . .
Another Oscillator . . . . . . . . . . . . . . . . . . . . .
Normal Modes . .
Double Pendulum
String Theory . . .
Easy Double . . .
.
.
.
.
.
.
.
.
Stressful Torque . . . . . .
Inertia Tensor of a Cube . .
Broken Lollypop . . . . . .
Spinning Rectangular Plate .
Two Rods . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
12
13
23
24
24
35
37
38
38
50
54
55
59
61
3.6 Triangular Plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.1 Carousel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
DR
4.2 Wheel Travelling in a Circle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.3 Bead on a Parabolic Wire . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
xi
xii LIST OF EXERCISES
T
11.3 Sliding Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
Physics Review
AF
1.1
Resources
Ch. 1,2,3,5,8,14 of Marion & Thornton
Appendix I: State Vectors
ẍ = −
k
m
x.
Week 1
The EOM, combined with the initial conditions, describes the motion of a system. It is usually in the form of a
set of coupled differential equations in the unknown positions of the parts of the system, ~ri . These equations can be
solved analytically or numerically.
DR
Exercise 1.1 Atwoods Machine.
An Atwood’s machine has masses m1 > m2 , which start at rest and level with each other. (a) Calculate the speed of
the masses after they have moved a distance h from their starting position. Assume the pulley is light. (b) If, instead,
the pulley is cylindrical and has a mass m3 and radius r calculate the speed of the masses after they have fallen a
distance h from their starting position.
Example 1.1 Trajectory Motion.
Consider two dimensional trajectory motion (see Fig. 1.1, left) with equations of motion given by the usual expres-
sions,
These equations are valid (assuming negligible air resistance, etc) but are only useful for this specific choice of coor-
dinate system.
Consider the similar problem with the projectile along an inclined plane (see Fig. 1.1, right). This problem may
be easier to analyze using the primed coordinates,
1
2 CHAPTER 1. PHYSICS REVIEW
T
AF Figure 1.1: Two dimensional trajectory motion using two different orientations of the coordinate axes.
Solution
(a) See the diagram on the left in Figure 1.3
1.1. NEWTON’S SECOND LAW FOR LINEAR SYSTEMS 3
ẑ
O
T
θ
`
ŷ
φ ~v
x̂
r̂
AF
Figure 1.2: The bob of a conical pendulum travels in an horizontal circle at a constant speed. The string traces out a cone.
m~ac
~FT
m~g
θ
Figure 1.3: On the left is the FBD for the bob of a conical pendulum has two forces and one acceleration. On the right is are
the three vectors arranged in a triangle – the sum of the two force vectors is equal to the mass times the acceleration vector.
DR
(b) I have
m~r¨ = F~T + m~g
where m~g = −mgẑ and, at the moment pictures, F~T = FT (x̂ sin θ + ẑ cos θ).
(c) I have ~r¨ = −x̂ω 2 a, where a = ` sin θ is the radius of the circular trajectory of the bob, so N 2 L gives
(d) Rearranging and dividing the equations from part (c) eliminates the tension, FT , and the mass, m, giving
(e) Rearranging and squaring both equations from part (c) gives
2 2
(FT sin θ) = −mω 2 ` sin θ
2
(FT cos θ) = (mg)2
T
and adding these allows me to isolate the tension, FT ,
g/` cos θ
FT2 ω ` sin θ)2 + (mg)2
= (m2
>
2
= (mg tan θ) + (mg)2
= (mg sec θ)2
⇒FT = mg sec θ
sin θ =
mω 2 a
FT
=
mg
FT
FT
⇒
mω 2 ` sin θ
FT = mg sec θ
⇒ ω2 =
m`
FT
A particle, mass m, is subjected to a one-dimensional force F (t) = kte−αt , where k and α are constants. Assume
X
that at t = 0 the particle is at rest at x = 0. (a) Use a computer to plot the acceleration of the particle, ẍ(t) = ẍ(t).
(b) Calculate the velocity, ẋ(t), of the particle by integrating the EOM (Hint: integrate by parts). (c) Calculate the
position, x(t), of the particle by integrating your solution to part (a). (d) Use a computer to calculate v(t) and x(t)
numerically (Hint: use the state-vector ψ(t)~ = [x(t), v(t)]). (e) Plot both the analytical, from parts (a) & (b), and
numerical results, from part (c).
DR
Solution
(a) See Listing 1.1 and Figure 1.4.
(b) I will need the integral
Z t
®
u=t
−αt ⇒ du = dt
I= dt te
0 dv = dt e−αt
⇒ v = −1
α e
−αt
òt Z t
t −1 −αt
ï
= − e−αt − dt e
α 0 0 α
òt
t 1 1 t
ï
= − e−αt − 2 e−αt = − 2 (1 + αt)e−αt 0
α α 0 α
1
= 2 1 − (1 + αt)e−αt .
α
For this problem I have mv̇ = kte−αt , from N2L, so
Z t Z v 0 k Z t
dv k
dt = dv = v −
v◦ = dt te−αt = I
0 dt v◦ m 0 m
k
⇒v= 1 − (1 + αt)e−αt .
mα2
1.1. NEWTON’S SECOND LAW FOR LINEAR SYSTEMS 5
1 ################################# constants
2 m = 1.0 # mass, m [kg]
3 k = 1.0 # drag force per unit mass, k [N/m]
4 alpha = 0.5 # time constant, α [Hz]
5 x0 = 0.0 # initial height, x1,0 [m]
6 v0 = 0.0 # initial speed, v0 [m/s]
7 N = 100 # number of time steps [dimensionless]
8 tN = 20.0 # maximum time, tN [s]
T
9 ################################# calculated constants
10 Dt = tN / N # time step size, ∆ t [s]
11 t = arange( N ) * Dt # length-N array, ti = i∆ t, i = 0, 1, 2, ..., (N − 1)
12 beta = ( m, k, alpha ) # parameter vector
13 a = k / m * t * exp( -alpha * t ) # analytic result
14 ################################# plot a(t) = ẍ(t)
15 plot( alpha*t, m*alpha*a/k )
16 xlabel( r'$\alpha t$', fontsize=14 )
17 ylabel( r'$m\alpha \ddot{x}/k$', fontsize=14 )
18 grid( 'on' )
Listing 1.1: This code plots the acceleration shown in Figure 1.4.
AF 0.40
0.35
0.30
0.25
mαẍ/k
0.20
0.15
0.10
0.05
0.00
0 2 4 6 8 10
αt
Figure 1.4: A plot of the acceleration, ẍ(t) = kt −αt
e for Example 1.3.
DR
m
k 1 −αt 1 k 1
Å ã
= 2
t + e − − 2
1 − (1 + αt)e−αt
mα α α mα α
k
= 3
αt − 2 + (2 + αt)e−αt
mα
(d) See Listing 1.2.
6 CHAPTER 1. PHYSICS REVIEW
1 ################################# variables
2 # Initialize an N × 2 array with zeros for the state vector, ψ(t) ~ = [ x(t), v(t) ],
3 # one row for each of the N times, ti = 0...tN , and one column for each xi and vi .
4 psi = zeros( [ N, 2 ] )
5 ~
# The zeroth row contains the initial condition, ψ(0) = ψ~o , the values at t=0.
6 psi[0,:] = x0, v0
7 ################################ state vector derivative function
8 def psiDot( phi, t, beta ) :
T
9 """
10 This function accepts the current state vector and returns its time
11 derivative. This abstraction encapsulates all of the physics (ie the
12 \gls{eom}) in this function. It also creates an obvious
13 extension to higher dimensional problems.
14
15 Further, it creates an hierarchy in which the solution method (which
16 makes use of this function) is independent of the particular problem.
17
18 These three principles, encapsulation, abstraction and hierarchy, are
19 the foundation for creating well-written computer programs.
20 """
21 x, v = phi # unpack the current state vector
22 m, k, alpha = beta # the parameters
23
24
25
26
27
28
29
1
2
3
4
5
6
7
8
9
AF a = k / m * t * exp( -alpha * t )
x, v = psi.T
clf()
# Newton's second law, a = kt
return array( [ v, a ] ) # time derivative of state vector, ψ~˙
me
~i+1 = ψ
psi[i+1,:] = psi[i,:] + psiDot( psi[i,:], t[i], beta ) * Dt # ψ
# extract the rows from the transpose of psi
−α t
~˙ i dt
~i + ψ
Listing 1.2: This code calculates the position and velocity using Euler’s algorithm and the concept of a state-vector,
Listing 1.3: This code plots the velocity, v(t) = ẋ(t), in Figure 1.5.
(e) See Listing 1.3 and Figure 1.5 for the velocity and See Listing 1.4 and Figure 1.6 for the displacement.
DR
X
Listing 1.4: This code plots the displacement, x(t), in Figure 1.6.
1.1. NEWTON’S SECOND LAW FOR LINEAR SYSTEMS 7
1.0
0.8
T
mα2 ẋ/k
0.6
0.4
0.2 numeric
analytic
0.0
AF
Solution
(a) For this problem I am given two vectors, B
0 2
k
mα2
αt
6 8
10
~ & ~v◦ and I need to select a coordinate system. I will first choose B
in the ẑ direction (almost always a convenient choice for a uniform magnetic field). Next, since two vectors (in
~ & ~v◦ to be in the x-z plane.
general) define a plane I will choose B
In Cartesian coordinates I have
~ = B ẑ
B 1
(b) Newton’s second law, combined with the Lorentz force equation, gives me
DR
m~v˙ = q~v × B
~ = qB(vx x̂ + vy ŷ + vz ẑ) × ẑ
m(v̇x x̂ + v̇y ŷ + v̇z ẑ) = qB(vy x̂ − vx ŷ)
⇒ v̇x = ωL vy , v̇y = −ωL vx , v̇z = 0, 2
qB
where ωL ≡ m is the the Larmor frequency.
(c) The equations for vx and vy are coupled. If I introduce the complex quantity
ṽ ≡ vx + ivy
then its time derivative is
ṽ˙ = v̇x + iv̇y
= ωL vy − iωL vx
= −iωL (vx + ivy )
˙
⇒ ṽ = −iωL ṽ 2
which has solutions ṽ = ṽ◦ e−iωL t . The velocity as a function of time is, therefore,
~v (t) = v◦ (x̂ cos α cos(ωL t) − ŷ cos α sin(ωL t) + ẑ sin α)
= v⊥ (x̂ cos(ωL t) − ŷ sin(ωL t)) + vk ẑ. 2
8 CHAPTER 1. PHYSICS REVIEW
0.5
0.4
T
mα3 x/k
0.3
0.2
0.1 numeric
analytic
0.0
AF
The motion of aÄcharged particle,
equation, F~ = q E
radius
where ω ≡ qB
~ + ~v × B
ä
0 2
(d) The motion is a helix with the axis parallel to the magnetic field.
k
mα3
αt
6
2
8 10
αt − 2 + (2 + αt)e−αt for Example 1.3.
r=
mv
qB
= ,
v
ω
~
X
m is the cyclotron frequency. (b) Choose the z-axis to lie in the direction of B and let the plane containing
~ ~
E and B be the y-z plane. What are the EOMs? (c) Show that the solution for the z component of the motion is given
by
DR
qEz 2
z(t) = z◦ + ż◦ t + t ,
2m
where z◦ ≡ z(0) and ż◦ ≡ ż(0). (d) Use the initial conditions x(0) = −v◦ /ω, ẋ(0) = Ey /B, y(0) = 0 & ẏ(0) = v◦
to show that the solutions for the x & y components of the motion are
v◦ Ey
x(t) = − cos(ωt) + t
ω B
v◦
y(t) = sin(ωt)
ω
Ey
(Hint: introduce u(t) = x(t) + iy(t) and use the ansatz u(t) = C1 e−iωt + B t + C2 . (e) Show that the time averages
of the x and y components of the velocity are
Ey
hẋi = x̂ & hẏi = 0
B
x3
F (x) = −kx + k
a2
T
where k and a are constants. (a) What is the potential energy, U (x)? Choose your zero of potential energy at the
origin, i.e. U (0) = 0. (Hint: You must have U (0) = 0 for later parts of this question.) (b) Identify the equilibrium
points? (c) Plot U (x). Label the equilibrium points. (d) What is the minimum speed, vo , which the particle must have
at the origin in order to never return? (e) What is the angular frequency, ω, of small oscillations about the point of
stable equilibrium? (f) Integrate the EOM, mẍ = F , to find v(x). Take v(0) = vo as your BC. (Hint: Recall that
dt = v dx .) (g) Plot the constant energy contours in the x-ẋ plane. Label any special contours that separate regions
dv dv
of qualitatively different motion. (h) What is x(t) for the special case of vo2 = ka /2m? Plot the result. (Hint: This
2
AF
1.2
1.2.1
1.2.2
1.2.3
Newton’s Second Law for Rotating Systems
Central Forces
CoM Coordinates & Reduced Mass
Fixed Axis Rotations
In this section we will consider planar rigid bodies which rotate about the ẑ axis perpendicular to its plane (i.e. the
rigid body is confined to the x-y plane). Because the body is rigid, if one part of the body is rotated by an angle
dφ about ẑ then all other parts of the body have the same rotation, dφ, about the ẑ axis. It is thus convenient to use
cylindrical coordinates with the ẑ axis along the rotation axis.1
The distance that a given part of the rigid body (with mass dm and position ~r) moves when it rotates by dφ about
the ẑ axis is
~˙ × ~r.
d~r = φ
The distance from dm to the origin, r = |~r|, does not change with the rotation dφ because this is the only motion of
the body.2
The velocity of the part dm is then
DR
d~r ~
dφ
~v = = ~r˙ = ~˙ × ~r,
× ~r = φ
dt dt
~˙ = φ̇ẑ is the angular velocity (because we are considering only rotational motion about ẑ we have ż = ṡ = 0).
and φ
The acceleration of the part of the rigid body, dm, (due to the rotation about the ẑ axis) is then
~v
~¨ × ~r + φ
˙~v = φ ~˙ × ~r˙ = φ
~¨ × ~r + φ
~˙ × φ~˙ × ~r ,
~¨ is the angular acceleration. We can use the identity from Example 12.6,
in which φ
to write
~¨ × ~r + (φ
~v˙ = φ ~˙ · ~r)φ
~˙ − (φ̇2 )~r.
1 Recall from subsection A.2.2 that cylindrical coordinates are the same as polar coordinates with the addition of the ẑ unit vector perpendicular
to both ŝ and φ̂. The position vector is ~ r˙ = ṡŝ + sŝ˙ + ż ẑ = ṡŝ + φ̇φ̂ + ż ẑ.
r = sŝ + z ẑ, and the velocity vector is ~
2 The general motion of a rigid body can be divided into a translational motion of the centre of mass ( COM ) of the rigid body and a rotational
motion about an axis. We are currently considering only the rotational motion about ẑ.
10 CHAPTER 1. PHYSICS REVIEW
~˙ = φ̇ẑ so φ
Since, for the moment, we are only considering rotations about ẑ I have φ ~˙ · ~r = 0, and the acceleration is
~¨ × ~r − (φ̇2 )~r.
~v˙ = φ
The first term is in the x-y plane and orthogonal to ~r – it is the tangential acceleration. The second term is the ω 2 r
T
centripetal acceleration which is in the radial direction.
In cylindrical coordinates I have (for a planar object in the x-y plane) ~r = sŝ so that the acceleration has radial
and tangential components:
~v˙ = ŝ φ̈r − r̂ φ̇2 r.
Torque
Recall that N 2 L for a particle, mass m, is X
m~v˙ = F~i
AF
so that the net torque on the particle is
~r × m~v˙ =
~ ≡ ~r × m~v ,
i
and that the torque due to the force F~i , relative to the origin O, is
~ i = ~r × F~i
N
~r × F~i =
~˙ = ~r × m~v˙ ,
L
X
i
~ i.
N
so that
X
~˙ =
L ~ i,
N
DR
i
Choice of Origin
The angular momentum clearly depends upon the choice of origin. However, this last equation is always valid so we
have the freedom to choose any origin which can often be exploited to simplify a problem. See Exercise 1.6 for a
particularly dramatic example of the importance of the choice of origin.
Angular Momentum
~˙ × ~r, and the angular momentum vector will
The velocity of an infinitesimal part, position ~r & mass dm, is then ~v = φ
be
~ = ~r × (dm ~v )
dL
1.2. NEWTON’S SECOND LAW FOR ROTATING SYSTEMS 11
T
Z Z
L = dm rv = dm r2 φ̇ = I φ̇
where Z Z
I≡ dm r2 = dm (x2 + y 2 )
Moments of Inertia
AF
The total mass, M , of an object is
and the COM of an object, total mass M , is the position ~rCoM such that
Z
~rCoM dm = M~rCoM = dm ~r.
I=
Z
Z
dm
dm s2
The MOI of an object depends upon the total mass and how that mass is distributed over the object, as well as the
axis about which the object is rotated. It is the rotational equivalent to mass.
Example 1.5 MoI of a Hoop.
Find the MOI of a thin uniform hoop, total mass m and radius a. (Hint: linear density λ = m/2πa)
Solution
R
DR
I have I = dm r2 with r = a, a constant, so that
Z Z 2π Z 2π 2π
2 2 2 >= ma2
I=a dm = a λa dθ = a (m/2πa)λa dθ
0 0
Solution
(a) This system has rotational symmetry about the x̂3 axis so I know that ~c, the position of the COM, is along the x̂3
axis and c1 = c2 = 0.
Now consider the ring of radius s = a sin θ at x3 = a cos θ, where θ is the usual polar angle. The hemisphere
has area 21 4πa2 so the areal density is σ = 2πa
m
2 and the mass of the thin ring (width a dθ) is
ê3
T
s a dθ
x3
a
θ
ê2
AF ê1
Figure 1.7: The mass of a ring, circumference 2πs and width a dθ, of a hemispherical shell (mass m, radius a) is dm =
σ(2πs)(a dθ), where θ is the usual polar angle, s = a sin θ is the radius of the ring, and σ is its areal density.
I therefore have
mc3 =
Z
= 2πσa3
dm x3 = 2πσa
1
Z
= 2πσa3 12 u2 0 = πa3
1
2
u = sin θ ⇒ du = dθ cos θ
du u,
m
2πa2
⇒ c3 = 21 a
Z
0
π/2
dθ sin θ(a cos θ)
(b) With s = a sin θ the distance of the mass dm from the x̂3 axis the MOI is
DR
Z Z π/2
2
I3 = dm s2 = 2πσa2 dθ sin θ(a sin θ)
0
Z π/2 2/3
= 2πσa 4
dθ *3 θ = 2 πa4 m
sin
⇒ I3 = 13 ma2
3 2πa2
0
(c) The speed of the ring at height x3 = a cos θ is x3 ω so the angular momentum of the ring is
dL3 = dm s(ωs)
where s = a sin θ is the distance of the ring from the x̂3 axis and the total angular momentum is
Z Z
L3 = dm ωs2 = ω dm s2 = ωI3 ⇒ L3 = 13 ma2 ω.
T
Z
Z = 12 ωi ωj dm (xk xk δij − xi xj )
T = 21 dm ~v · ~v ↔
Z ~TI ω
= 12 ωi Iij ωj = 12 ω ~,
= 12 dm (~
ω × ~r) · (~
ω × ~r) where Z
Z Iij = dm (xk xk δij − xi xj )
= 12 dm ω 2 r2 − (~
ω · ~r) 2
AF
in which I have used the identity from Example 12.5,
~ × B)
(A ~ · (C
~ × D)
~ = (A
~ · C)(
~ =
L
=
Z
Z
~˙ so that
will have the same angular velocity, θ,
˙
dm ~r × θ~ × ~r
˙
˙
dm (~r · ~r)θ~ − (θ~ · ~r)~r
Calculate the MOI for an axis through the COM and perpendicular to the plane of the object, of (a) a thin uniform rod,
total mass m and length `. (Hint: linear density λ = m/`), (b) a thin uniform disk, total mass m and radius a. (Hint:
areal density σ = m/πa2 ), (c) a thin uniform plate, total mass m and sides a & b. (Hint: areal density σ = m/ab),
(d) a (thin) uniform cylinder, total mass m, radius a and length `. (Hint: volumetric density ρ = m/πa2 `), and (e) a
uniform sphere, total mass m and radius a. (Hint: volumetric density ρ = 3m/4πa3 ).
~s 0 dm
O0
~c ~s
Figure
R 1.8: The MOI of a laminar object, mass M , about an axis perpendicular to the object through the origin at O is
I = dm s2 where s is the distance from the origin to the infinitesimal mass dm. However, since ~s = ~s 0 + ~c where ~c is
the position of the COM relative to O, we have I = ICoM + M c2 , where ICoM is the moment of inertia about a parallel axis
through the COM. This is the parallel axis theorem.
14 CHAPTER 1. PHYSICS REVIEW
T
Figure 1.9: Sphere on a Stick
so the same MOI relative to the origin O0 located at the COM with position ~c relative to O is
Z
I = dm (c1 + x01 )2 + (c2 + x02 )2
AF
1.2.6
2 2
=
Z
= ICoM + M c2
1
2
2
2
dm (x01 + x02 ) + 2c1
1
T = 2 dm v = 2 dm (rθ̇) = 2 θ̇
Z
2
dm
x0
*
1
0
+ 2c2
where c2 = x01 + x02 is the distance between O and O0 and ICoM is the moment
Z
dm
x0
*
2
0
+ (c21 + c22 )
R R
where I ≡ dm r2 = dm (x2 + y 2 ) is, again, the MOI of the body about the fixed axis. The quantity I is a property
of the body and the choice of origin, and is independent of the dynamics.
DR
Example 1.7 Sphere on a Stick.
A solid sphere of mass m1 = 5m2 and radius a is mounted on a the top of a rod of mass m2 and length 3a (see
Figure 1.9). The bottom of the rod is hinged and the rod starts from rest in the vertical position. (a) What is the
MOI , I, of the system about the pivot? (b) How far from the pivot, c, is the COM of the system? (c) Use N 2 L for
~ = I φ,
rotations, N ~¨ to find the instantaneous angular acceleration of this system. (d) Use conservation of energy to
find the angular velocity, φ̇, of the system as a function of the angle, φ, of the rod relative to the vertical. (Hint: The
kinetic energy of an object rotating with angular velocity φ̇ is T = 12 I φ̇2 , where I is the moment of inertia about the
rotation axis.) (e) What is the speed of the COM of the sphere when the rod is horizontal if ?
Solution
(a) The moments of inertia for a sphere and rod about their COM are 25 m1 a2 & 12 m2 ` ,
1 2
respectively. For the entire
system I use the PAT for each of the sphere and rod and then add the results,
Z Z Z
I = dm s2 = dm1 s2 + dm2 s2
sphere rod
2 2 2
= 2
5 m1 a + m1 (` + a) + 1
12 m2 ` + m2 (`/2)2 .
T
where I have used m1 = 5m2 and ` = 3a.
where ~c is the vector position of the COM, and φ̂ is the unit vector out of the page and φ is the angle the stick
makes with the vertical.
~ = Iφ
N 2 L for rotations is N ~¨ = I φ
~ φ̂ so
43g
AF (85m2 a2 )φ̈φ̂ =
(d) Taking the zero of potential energy when the system is horizontal (i.e. φ = π/2) I have
and
T = 12 I φ̇2
2
This is exactly the EOM for a compound pendulum (i.e. not a simple pendulum). The presence of the positive
sign is because I measured the angle, φ, from the upward direction (i.e. the stable equilibrium is at φ = π). This
ordinary differential equation (ODE) does not have a simple solution – neither the angular velocity, φ̇, nor the
43
2
43 g
⇒ φ̇2 = (1 − cos φ) 2
85 a
(e) The speed of the COM of the sphere when the rod is horizontal, φ = π/2, is then
… …
43 g 688 √
v = ω(` + a) = 4aω = 4a = ga. 2
85 a 85
X
~
ω
O0 ~L0
T
θ̂
φ̂
ŝ
θ
θ
ẑ r̂
AF
of the circular trajectory, L
~L
O
~r
~r 0
~v
Figure 1.10: The conical pendulum with the angular momentum vectors identified for the two cases: the origin at the centre
~ = ma2 ω ẑ, and the origin at the pivot, L ~ 0 = ma2 ω θ̂. In both cases the angular momentum is
orthogonal to the plane defined by the position and velocity vectors, however, for the second case the angular momentum is
not a constant vector – it traces out the cone as it rotates in the plane of the bob and string. In this diagram the velocity vector,
~v , is orthogonal to all the other vectors (which are therefore in the same plane).
On the right are the spherical and cylindrical basis vectors. Because the usual definition of the polar angle, θ, is from ẑ to
r̂ I have defined the ẑ axis as downwards. The relationship between the cylindrical and spherical basis vectors is therefore a
rotation of θ about the φ̂ axis. The velocity will therefore be in the −φ̂ direction and the angular velocity will be in the −ẑ
direction.
Solution
(a) With the origin at the centre of the circular trajectory I have, using cylindrical coordinates, ŝ, φ̂ & ẑ, with ẑ
pointing down (see Figure 1.10):
DR
~ = −ωẑ & ~r = aŝ
ω
so
~ × ~r = −ω ẑ × a ŝ = −aω φ̂
~v = ω
and the angular momentum is
~ = m~r × ~v = m(aŝ) × (−aω φ̂) = −ma2 ω ẑ,
L
which is a constant.
so
ẑ = r̂ cos θ − θ̂ sin θ,
and the velocity of the bob is
T
which is unchanged from (a) (as it must be). Alternatively, I could use the cylindrical basis,
This works because the cross product of two vectors is independent of the choice of basis, an important principle
in physics.
I can calculate the angular momentum using cylindrical coordinates using
AF
so that
L
r̂ = ẑ cos θ + ŝ sin θ
constant but is, instead, rotating in the same plane as the bob – see Figure 1.10.
~ 0 is no longer a
(c) The two angular momenta have the same magnitude but their directions are different. In fact L
X
DR
18 CHAPTER 1. PHYSICS REVIEW
~˙ = N
T
L ~ = ~r × F~ = 0,
r̂
ï ò ï òï ò
cos φ sin φ x̂
=
AF ï ò ñ ˙ô
d r̂
dt φ̂
r̂
= ˙ = φ̇
φ̂
− sin φ cos φ ŷ
Ä
ä
ä
= r̂ r̈ − rφ̇2 + φ̂
òï ò
− sin φ cos φ x̂
− cos φ − sin φ ŷ
Ä
= φ̇
r φ̇ ,
.
˙
~r¨ = r̈r̂ + ṙr̂˙ + ṙφ̇φ̂ + rφ̈φ̂ + rφ̇φ̂
= r̈r̂ + ṙφ̇φ̂ + ṙφ̇φ̂ + rφ̈φ̂ − rφ̇2 r̂
Ä
r dt
and N 2 L is
1 dÄ 2 ä
F~ = m~r¨ = r̂ m r̈ − rφ̇2 + φ̂
Ä ä
mr φ̇ .
DR
r dt
When the only force acting on a body is a central force I have F~ = F r̂ and the radial and angular components of
N 2 L are
Ä ä
m r̈ − rφ̇2 = F ⇒ mr̈ = F + mrφ̇2
m dÄ 2 ä
r φ̇ = 0 ⇒ mr2 φ̇ = L, a constant of the motion.
r dt
I can use the last equation to eliminate φ̇ = L/mr2 from the first equation, leaving
L2
mr̈ = F + mr(L/mr2 )2 = F + .
mr3
This equation has the form of a one-dimensional system with the additional force,
L2
Å 2 ã
d L
Fcf ≡ = − ,
mr3 dr 2mr2
called the centrifugal force. These so-called fictitious forces are known as inertial forces because they originate
L2
from the inertial term of N 2 L, m~r¨. It is often convenient to use the corresponding potential, Vcf (r) ≡ 2mr 2 so that
ẑ
O
T
θ
`
ŷ
φ ~v
x̂
r̂
AF
1.3.1
Figure 1.11: For the conical pendulum the position and velocity vectors are orthogonal, ~r · ~v = `r̂ · ~r˙ = 0.
Effective Potentials
For a conservative system the force can be written as the gradient of a scalar potential, F~ = −∇V
to introduce the effective potential,
Veff (r) ≡ V (r) + Vcf (r)
so that we can use our previous knowledge of potentials to examine the dynamics of a problem.
Example 1.9 Conical Pendulum - III.
~ , and it is convenient
Consider (once again) the spherical pendulum (length `, mass m). Put the origin, O, at the pivot and use spherical
coordinates, θ and φ, for this problem (see Figure 8.1). (a) Solve this problem using Newton’s second law in
spherical coordinates. [Hint: Use the expression for r̂¨ in subsection M.2.2 ] (b) The z components can be integrated
to give L̇ = 0 in which L is the z component of the angular momentum. What is L? (c) The magnitude of the angular
momentum, L, is a constant. Use this to eliminate φ̇ from the θ component, leaving an EOM in θ(t). (d) Re-write
your EOM in the form m`θ̈ = − dVdθeff , where Veff is an effective potential. (e) Plot the effective potential as a function
p
DR
of angle,
θ, for 0 ≤ L ≤ 2L◦ where L◦ ≡ m ` g is the unit of angular momentum. (f) Calculate θ◦ such that
2 3
dVeff
dθ θ=θ = 0. This is the equilibrium angle. (g) Write the effective potential as a Taylor series about the equilibrium
◦
angle, θ◦ (i.e. θ̈ θ = θ◦ = 0). (h) What is the frequency of small oscillations about the equilibrium?
Solution
(a) With ~r = `r̂ the acceleration is ~r¨ = `r̂¨ and Newton’s second law is
T
Ä ä
m` 2θ̇φ̇ cos θ + φ̈ sin θ =0.
Ä ä
(b) The z component can be integrated to d
dt m`2 φ̇ sin2 θ = 0) so
(c) I have φ̇ = L
m`2 sin2 θ
, so that
ã2
L g
Å
1
2
3
4
5
AF
(d) I have
so that
θ̈ =
################################# constants
m
ell
g
N
= 1.0
= 1.0
= 10.
= 100
2
m` sin θ
L2 cos θ
θ̈ = −
2
= 2 4 3 − sin θ
m ` sin θ
Veff (θ) =
# mass, m [kg]
# length, ` [m]
Å
sin θ cos θ − sin θ
g
`
L2
dθ 2m2 `4 sin2 θ
2
L2
4 2
2m ` sin θ
−
g
`
g
`
cos
− cos θ
`
θ
ã
(e) The code below creates Figure 1.12. I can see that the equilibrium point (i.e. the minimum of Veff (θ)) is at θ = 0
for L = 0, as expected, and that as L increases it moves towards θ = π/2, again as expected.
L2 cos θ g sin4 θ◦
0= 3 − sin θ ⇒ L2 = m2 `3 g .
2 4
m ` sin θ ` cos θ◦
sin4 θ◦
For L2 = m2 `3 g I need to find the root of 0 = 1 − cos θ◦ which I do using the following code:
1 from scipy.optimize import newton
2 def f( theta ) :
3 return 1-sin(theta)**4 / cos(theta)
4 theta_o = newton( f, pi/3 )
5 print( '{:.5f}'.format( theta_o/pi ) ) # report the root
1.3. CENTRAL FORCES 21
10
T
Veff /V◦
0
L = 0.0 L◦
L = 0.5 L◦
−5
L = 1.0 L◦
L = 1.5 L◦
L = 2.0 L◦
−10
3
4
5
6
7
8
9
10
AF 0.0 0.2
Figure 1.12: A plot of Veff (θ). The equilibrium point (i.e. the minimum of Veff (θ)) is at θ = 0 for L = 0, as expected, and
(h) I have
00 L2 (1 − 4 cos2 θ) g
Veff (θ) = + cos θ
m2 `4 sin4 θ `
so
00 sin4 θ◦ 1 − 4 cos2 θ◦ g
Veff (θ◦ ) = −m2 `3 g 4 + cos θ◦
cos θ◦ m ` sin θ◦
2 4 `
g 1 − 3 cos2 θ◦
=−
` cos θ◦
1.3.2 Rolling
An object is rolling when the point of contact of the rolling object with a surface, Q, has zero velocity – if it did have
a velocity then the object would be, instead, sliding. That the velocity of this point is zero is often imposed upon a
system as a constraint. This point is know as the instant centre of rotation (ICOR).
22 CHAPTER 1. PHYSICS REVIEW
10
T
6
Veff /V◦
4
2
V (θ)
U (θ)
0
ê02
O0
ê01 Q
~rC0
α
~rQC
p
C
0.0
0
~r˙ C
0.2
~g
α
0.4
θ/π
0.6 0.8
Ff
FN
1.0
θ
C
m~g
α
ṡ
θ̇
For example, a cylinder rolling upon a horizontal plane has a COM velocity ~vC and a rotational velocity ω ~ . These
DR
two quantities are related by the “rolling constraint” on the velocity of the instantaneous point of contact,
~ × ~a
0 = ~vC + ω
∆rC + a ∆θ = C, a constant.
Solution
(a) See Figure 1.14, right.
(b) Using the PAT the MOI is
I = I◦ + ma2 = 12 ma2 + ma2 = 32 ma2 .
1.3. CENTRAL FORCES 23
(c) Because I am summing the torques about the point Q the constraint force makes no contribution and I get
2g
( 23 ma2 )θ̈ = mga sin α ⇒ θ̈ = sin α.
3a
X
T
Exercise 1.6 Sliding Ball.
Consider a bowling ball (mass m, radius a) which has an initial velocity v◦ but no rotational velocity. The ball slides
along the floor (coefficient of kinetic friction µ) and the friction creates a torque on the ball. Eventually the ball will
be rolling (i.e. the speed of its COM, v, will be equal to its angular velocity, ω, multiplied by its radius, a, so that there
will be no relative motion between the point of contact of the ball and the ground). (a) Draw the FBD of the ball as it
slides. (b) What are N 2 L for the ball (both linear and rotational; use the COM for the origin)? (c) How long after the
ball starts sliding will it be rolling? (d) What is the speed, vroll , of the rolling ball. (e) How much work was done by
friction? (f) Place the origin at a point on the floor and calculate the angular momentum, L ~ ◦ , of the ball at the moment
AF
relationship L
Lucy, mass m, is riding a Ferris wheel of radius a. (a) Draw the FBD for Lucy. (b) Write Newton’s second law for
Lucy. Take the force F~ as the reaction force of the Ferris wheel on Lucy. (c) Find Lucy’s acceleration, ~r¨, in polar
coordinates and unit vectors. The vector ~r = r̂ a is Lucy’s position vector. (d) What is the force due to gravity, m~g ,
in terms of the same polar coordinates and unit vectors? (e) Using polar coordinates to resolve the components of the
force and acceleration acting on Lucy and so find the radial and tangential components, Fr & Fφ , of the reaction force,
F~ . (f) If the Ferris wheel is rotating with a constant angular velocity ω, what is the reaction force? (g) What angular
speed, ω◦ , is necessary so that Lucy just leaves her seat at one point and where is this point?
Solution
(a) I have
m~r¨ = m~g + F~ .
(b) If Lucy’s position vector is r̂ = x̂ cos φ + ŷ sin φ where φ is the angle ~r makes with the horizontal then ~r = r̂ a,
DR
~r˙ = r̂˙ a = φ̂ aφ̇, and
˙
~r¨ = φ̂ aφ̇ + φ̂ aφ̈ = −r̂ aφ̇2 + φ̂ aφ̈.
T
v 2 = (~v · ẑ)2 + (~v · x̂)2 to find v(t).
2.0
1.5
1.0
0.5
U (x)/ka2
0.0
AF −0.5
−1.0
−1.5
−2.0
−2.0−1.5−1.0−0.5 0.0 0.5 1.0 1.5 2.0
U (x) = ka
Å 4
2 x
a4
− 2
x2
a
ã
where a is a constant with units of length and k is the spring constant. [2017S midterm] (a) What is the force on the
particle, F (x)? (b) Identify the equilibrium points and state which are stable and which are unstable. (c) Plot F (x).
Label the equilibrium points. (d) What is the kinetic energy, To , which the particle must have at x = a in order to
never return? (e) What is the angular frequency, ω, of small oscillations about the stable equilibrium points? (f) What
DR
is v(x). Take v(0) = vo as your BC. (g) Plot the constant energy contours in the x-ẋ plane. Label any special contours
that separate regions of qualitatively different motion.
T
Chapter 2
AF
2.1
Resources
Thornton & Marion – §3.4, Ch. 12
Boas – Ch. 2, §7.2
Appendix H: Linear Algebra
Appendix I: State Vectors
Single Oscillator
Example 2.1 Complex Oscillators.
Consider the simple harmonic oscillator (mass m, spring constant k; see Figure 2.1) which has an EOM
Week 2
mẍ = −kx
in which x is an equilibrium coordinate (i.e. it measures the displacement from the equilibrium position) in Figure 2.1.
DR
(a) Solve this ODE using the ansatz x̃(t) = ãe−iωt . The complex constants ã = |ã|eiδ will contain an amplitude and
phase, two constants determined by the initial conditions. The actual solutions will be the real parts of x̃(t),
(b) Find is ã given the initial conditions x(0) = x◦ and ẋ(0) = v◦ . (See Figure 2.2 for a discussion of the mapping
between the complex plane and the state vector, ψ ~ ≡ [x, v].) (c) What is the solution to the original EOM? (d) Plot ã
and x̃(t) in the complex plane. Discuss, briefly, the connection between this plot and the real solution for x(t). (e) Plot
ẋ(x) in the ẋ-x plane (aka phase space) for several different choices of initial conditions. (f) Plot the constant energy
contours ẋ-x plane. (g) Discuss, briefly, the connection between the last two plots.
k x
m
Figure 2.1: A simple harmonic oscillator (mass m, spring constant k) has an EOM mẍ = −kx in which x is an glsequilibrium
coordinate (i.e. it measures the displacement from the equilibrium position).
25
26 CHAPTER 2. COUPLED OSCILLATIONS & NORMAL MODES
Im v
iv◦ x
ã = x◦ + ψ(0) = ◦
~
ω v◦
v◦ /ω
|Ã|
δ Re x
T
x◦
−ωt
x̃(t) = ãe−iωt ~ x(t)
ψ(t) =
v(t)
Figure 2.2: On the left the constant complex number ã = x◦ + iv◦ /ω and the time-varying complex number x̃(t) =
x(t) + ωi v(t) = ãe−iωt can be plotted in the complex plane (aka Argand plane). On the right the clockwise trajectory
Solution
(a) Using the given ansatz, I have
x◦ = Re (x̃(0)) = Re ãe
Ä
˙
ä Å
v◦ = Re x̃(0) = Re −iωãe
*0
−i
ωt
ã
= Re (ã)
−i*0
ωt
ã
~
(b) I have
Å
= ω Re (−iã) = ω Im (ã)
~
and at an arbitrary time ψ(t) are
plotted. The one-to-one mapping between the phase space trajectory and the rotation of x̃(t) in the complex plane (i.e.
⇒ ã = x◦ + iv◦ /ω.
I have used the complex ansatz precisely to avoid these sorts of manipulations.
2.1. SINGLE OSCILLATOR 27
1 ################################# constants
2 k = 1.0 # N/m
3 m = 1.0 # kg
4 v0 = 0.0 # m/s
5 x0 = 1 # m
6 n = 2 # number of periods to plot
7 N = 100 # number of time steps per period
8 ################################# calculated constants
T
9 w2 = k / m # square of the angular frequency
10 w = sqrt( w2 ) # angular frequency
11 T = 2*pi/w # the period
12 tMax = n * T # n oscillations
13 Dt = T / N # time step -- the subinterval length
14 t = arange( n*N ) * Dt # time steps: t = 0, dt, 2dt, ... , nT − dt
15 ################################# calculate and plot the trajectories
16 clf()
17 for a in linspace( 0, x0, 10 ) :
18 x = a * cos( w * t )
19 v = - w * a * sin( w * t )
20 plot( x/x0, v/(w*x0) )
21 # make it pretty
22 xlabel( '$x/x_\circ$' )
23
24
AF ylabel( '$\dot{x}/x_\circ\omega$'
grid( 'on' )
)
(d) The point ã = x◦ + iv◦ /ω is a distance |ã|2 = x2◦ + (v◦ /ω)2 from the origin and makes an angle δ =
tan−1 (v◦ /x◦ ω) with the real axis.
The point x̃(t) = ãe−iωt = x(t) + iv(t)/ω is a distance |x̃|2 = x2 + (v/ω)2 from the origin and makes an
angle δ = tan−1 (v/xω) with the real axis.
The initial energy is
|x̃|2 = |ã|2 .
DR
This means that the point x̃ traces out a circle, radius |ã| & angular frequency ω, in the complex plane (see left
hand side (LHS) diagram in Figure 2.2).
(e) See Listing 2.1 and Figure 2.3.
(f) See Listing 2.2 and Figure 2.4.
(g) Since this is a conservative system the phase space trajectories are constant energy surfaces.
X
ẋ = v
v̇ = F/m
1.0
0.5
T
ẋ/x◦ ω
0.0
−0.5
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
AF E0 = 0.5 * k * x0**2
################################# constants
0.5
ẋ/x◦ ω
0.0
−0.5
−1.0
−1.0 −0.5 0.0 0.5 1.0
x/x◦
T
9 ################################# calculated constants
10 w = sqrt( k / m ) # unit of frequency, ω
11 T = 2*pi/w # the period, T = 2π/ω
12 v0 = a * w # largest initial velocity, v◦ = aω
13 tMax = n * T # n oscillations
14 Dt = T / N # time step -- the subinterval length
15 t = arange( n*N ) * Dt # time steps: t = 0, dt, 2dt, ... , nT − dt
16 ################################# time derivative of the state vector
17 def psiPrime( psi, t, w ) : # returns ψ ~˙
18 x, v = psi # unpack ψ ~ = [x, v]
19 a = - (w**2) * x # a = F/m = −ω 2 x
20 psiDot = array( [v, a] ) # pack ψ~˙ = [ẋ, v̇]
21 return psiDot
22
23
24
25
26
27
28
29
30
31
32
33
AF ################################# calculate and plot the state vectors
clf()
args
psi0
x, v
= ( w, )
grid( 'on' )
= x0, vi
d x
ï ò ï ò
dt v
~ =
=
ẋ
v̇
Z t
⇒
~˙ =
Z
~˙ =
ψ
t
ï
ï
v
F/m
v
ò
ò
,
ψ(t) dt ψ dt .
0 0 F/m
(a) Using F = −kx calcuate x(t) using Python’s scipy.integrate.odeint() function and plot the result. Take
DR
x(0) = x◦ and ẋ(0) = 0 as your initial conditions. On the same figure plot the analytic result, x(t) = x◦ cos(ωt), in
order to validate your computations. (b) Now plot the result for F = −mω 2 sin(x) with ω = π. Take x(0) = π/4
2
and v(0) = 0 as your initial conditions. On the same figure plot the harmonic oscillator result, x(t) = ωm cos(ωt), in
order to compare the solutions. (Hint: In the x(0) → 0 limit this solution should be the same as in part (a). You can
use this to check your computations.)
Solution
(a) See Listing 2.3 and Figure 2.5.
(b) See Listing 2.4 and Figure 2.6.
X
30 CHAPTER 2. COUPLED OSCILLATIONS & NORMAL MODES
1.0
T
0.5
ẋ/aω
0.0
−0.5
1
2
3
4
5
6
7
8
9
10
11
12
AF n
N
w
T
tMax
Dt
Figure 2.5: A plot of ψ(t)
The trajectories are circles.
= 1.0
= 5
= 40
= sqrt(g/b)
= 2*pi/w
= n * T
= T / N
−1.0
−1.0 −0.5
~ for Example 2.2(a), for the harmonic oscillator with initial states ψ(0)
~ = [0, paω], p = 1, ..., 10.
T
AF 1.0
0.5
q̇/πω
0.0
−0.5
−1.0
−1.0 −0.5 0.0 0.5 1.0
q/π
DR
~ for Example 2.2(b) for the simple pendulum. For small oscillations (near the origin) the trajectories
Figure 2.6: A plot of ψ(t)
are almost circular but become progressively less so for larger oscillations, eventually becoming non-oscillatory for large
enough initial velocities. Because this system is conservative, these phase space trajectories are equivalent to the constant
energy surfaces, E = T + U = 12 mẋ2 − g cos x. Compare this plot to the similar one in ??.
32 CHAPTER 2. COUPLED OSCILLATIONS & NORMAL MODES
x1 x2
k εk k
m1 m2
T
Figure 2.7: The two masses, mi , are each connected to a spring with spring constant k and coupled by a third spring
with spring constant k. The displacements, xi , are equilibrium coordinates – they are displacements measured from the
equilibrium positions of the system.
AF In order to describe the state of a system I will also need to specify the corresponding momenta, mi ẋi , and the
vector space consisting of these n coordinates and n momenta is known as phase space and the vectors in this space
are known as state vectors.
This distinction between the configuration space (dimension n) and the phase space (dimension 2n) is important.
Any given system will have n EOM and the solutions (for various boundary conditions (BCs)) will a paths in phase
space.
If I choose a different basis, âi , for the configuration space this will transform the EOM. In this section we will
examine such a change in basis for a system of coupled oscillators. In the next chapter we will extend these ideas to
more general systems.
Example 2.3 Two Coupled Oscillators.
Consider a pair of identical oscillators (mass m, spring constant k) that are weakly coupled (see Figure 2.7 with
m ≡ m1 = m2 ). The EOM for the oscillators (from N 2 L) are
where ε is a small quantity and the xi are measured from the equilibrium position (known as equilibrium coordinates).
In the ε → 0 limit we recover two uncoupled oscillators.
(a) Use the ansatz,
DR
˜(t) = ~a
~x ˜e−iωt ,
˜ is the amplitude vector (and the tilde indicates a complex quantity) to find two coupled algebraic equations
in which ~a
˜ = mω 2~a
in its components, ãi .2 (b) Re-write the equations in the form K · ~a ˜, where K is the 2 × 2 spring matrix.
(c) This equation K − ω M ·~a
2 ˜ (where M ≡ m1 is the mass matrix) is a set of linear equations that will have non-trivial
solutions only if3
det K − ω 2 M = 0.
The resulting polynomial in ω 2 is known as the secular equation with two roots, ω± . Find the values of ω±2
that satisfy
this criteria; the ω± are the eigenvalues and ω± are the characteristic frequencies. (Hint: Introduce the dimensionless
2
quantity λ ≡ mω 2 /k.) (d) Find the corresponding values of â ˜± that are solutions to K − ω 2 M · â
˜± = 0. These
±
eigenvectors are the normal modes. (e) The general motion of the system will be (the real part of) a superposition of
the normal modes,
˜(t) = ~a
~x ˜+ e−iω+ t + ~a
˜− e−iω− t ,
each of which oscillates with its own characteristic frequency. What is ~x(t) for the initial conditions ~x(0) = [ xx◦◦ 12 ],
~x˙ (0) = [ vv◦◦ 12 ]? (Hint: You will need to find the real part of ~x
˜(t)]
2 The complex constants ã = |ã |eiδi will contain an amplitude and phase, two constants determined by the initial conditions. The actual
i i
˜, ~
solutions will be the real parts of the ~
x x(t) = Re(~
x˜(t)).
3 The trivial solution is ~
˜
a = 0 which is the equilibrium solution with nothing moving.
2.2. COUPLED OSCILLATORS 33
Solution
˜(t) = ~a
(a) Using the ansatz, ~x ˜eiωt , I have
¨˜ ˜eiωt = −ω 2 ~x
˜
~x = −ω 2~a
so the equations of motion can be written as
T
−(1 + ε)kã1 + εkã2 + mω 2 ã1 e−iωt =0
εkã1 − (1 + ε)kã2 + mω 2 ã2 e−iωt = 0,
AF
(b) I can rearrange the above to
(c) I have
K − ω2 M =
ï
ï
(1 + ε)k ã1
1 + ε −ε
−εk ã1
1 + ε −ε
−ε 1 + ε
−ε 1 + ε
where λ ≡ mω 2 /k is a dimensionless quantity.
ò
ò ï ò
k − ω2
k
ï
m
0
ã1
ã2
−εk ã2
+(1 + ε)k ã2
˜ = mω 2~a
and these can be written as the eigenvalue equation, K · ~a ˜ ,4
= mω
ï ò
2 ã1
ã2
0
m
.
ò ï
=
= mω 2 ã1
= mω 2 ã2
1+ε−λ
−ε
−ε
1+ε−λ
ò
k
If this matrix is non-singular then it’s inverse exists and the only solution will be the trivial solution, ~x = 0.
However, this matrix is singular if its determinant is zero and this will yield two specific values of λ for which
there are non-trivial solutions.
The secular equation is found by setting the determinant of the above matrix equal to zero,
1 + ε − λ −ε
DR
0 = = (1 + ε − λ)2 − ε2
−ε 1 + ε − λ
⇒ 1 + ε − λ = ±ε ⇒ λ± ∈ {1, 1 + 2ε},
and the characteristic frequencies (i.e. those for which there is a non-trivial solution) of this system are
2 k
⇒ ω± = (1 + ε ∓ ε)
m
(d) For ω−
2
= (1 + 2ε)k/m I have λ− = 1 + 2ε and
ï ò
1+ε−λ −ε
0= â
−ε 1+ε−λ −
ï ò
1 + ε − (1 + 2ε) −ε
= â
−ε 1 + ε − (1 + 2ε) −
ï ò ï ò
1 1 1 1
= −ε â ⇒ â− = 2√
1 1 − −1
4 The process of going from the pair of coupled equations to the matrix-vector form is purely mechanical – going in the opposite direction is
trivial, I am just doing it backwards instead of forwards.
34 CHAPTER 2. COUPLED OSCILLATIONS & NORMAL MODES
This normal mode corresponds to the two masses moving in opposite directions and the presence of the central
spring (of strength εk) increases ω−
2
and decreases the period, T− = 2π/ω− .
For ω+
2
= k/m I have λ+ = 1 and
ï ò
1+ε−λ −ε
0= â
T
−ε 1+ε−λ +
ï ò
1 + ε − (1) −ε
= â
−ε 1 + ε − (1) +
ï ò ï ò
1 −1 1
=ε â+ ⇒ â+ = √12
−1 1 1
This normal mode corresponds to the two masses moving together and the central spring remains uncompressed.
It is identical to the uncoupled problem (i.e. ε = 0), hence the frequency is identical to that problem.
˜± = ã± â± , but not their (complex)
Note that we have determined the direction, â± , of the amplitude vectors, ~a
AF length, ã± .
(e) The general motion is a superposition (i.e. a linear combination) of the normal modes,
˜(t) = ~a
~x
~x
˜+ e−iω+ t + ~a
˜˙ (0) = −iω+~a
˜− e−iω− t = ã+ â+ e−iω+ t + ã− â− e−iω− t
with the ã± determined by the initial conditions.
At t = 0 I have
˜(0) = ~a
˜+ e−i
ï ò
x
~x(0) = ◦1 = Re ~x
x◦2
ω+
t
*
˜+ e−i
Ä
ω+
t
*
0
0
˜− e−i
+ ~a
˜− e−i
− iω−~a ω−
˜(0) = Re ~a
ï
ä
t
*
ò
ω−
t
* 0
˜+ + ~a
= ~a
0
˜− .
˜+ − iω−~a
= −iω+~a
so I need to find the amplitude vectors consistent with the initial conditions,
˜+ + ~a
= √12 Re ã+ [ 11 ] + ã− −11
˜−
ä
˜−
1 ã+ + ã−
= 2 Re
√
ã+ − ã−
ï ò
v ˜˙ (0) = Re −iω+~a
DR
~x˙ (0) = ◦1 = Re ~x ˜+ − iω−~a
˜−
Ä ä
v◦2
˜+ + ω−~a ˜−
Ä ä
= Im ω+~a
= √12 Im ω+ ã+ [ 11 ] + ω− ã− −11
ï ò
1 ω+ ã+ + ω − ã−
= √2 Im
ω+ ã+ − ω− ã−
1 i
⇒ ã± = √ (x◦1 ± x◦2 ) + √ (v◦1 ± v◦2 )
2 2ω±
ä
˜± = 1 (x◦1 ± x◦2 ) + i (v◦1 ± v◦2 )
Ä
1 −iω± t
& ~a 2 ω± ±1 e
Another Interpretation
For the previous problem I can consider my set of coupled EOM as
T
¨ = −K · ~x
m~x
where ï ò
1 + ε −ε
K≡ k
−ε 1 + ε
is the spring matrix, with eigenvectors, â± = ±11 .
If I next create a block matrix with the normalized eigenvectors, â± , as the column vectors,
A ≡ â+ â−
then I know that
AF
is diagonal with diagonal elements ω±
where
2
m
dt2
Ä ä
K0 ≡ AT · K · A
A · AT = 1,
¨ = −AT · K · (A · AT ) · ~x
AT · m~x
d2 T
~x0 ≡ AT · ~x
is the configuration space vector in the âi basis – i.e. the basis in which the spring matrix,
configuration space vectors are the linear combinations of the xi given by
mK ,
1 0
is diagonal. These
T
˜ = ~a
in which M is the mass matrix.5 With the usual ansatz, ~x ˜e−iωt , I have the matrix equations
K − ω 2 M · ~a˜ = 0,
AF
and similarly for M. Also, both K and M are positive definite,7
˜† · K · ~x
k = ~x
Ä
= ~x
Ä
= ~x
˜
&
˜ †
˜† · K† · ~x
ä
˜ † = k†
˜† · K · ~x
ä
KT = K
so, since both K and M are positive definite, the eigenvalues, ω 2 are real and positive.
DR
Degenerate Eigenvalues
It is possible for an eigenvalue to be degenerate. We can then find normal modes that span the degenerate subspace in
the usual way.
Orthogonal Eigenvectors
If ωi2 and ωj2 are distinct eigenvalues (i.e. if i 6= j then ωi2 6= ωj2 ) with eigenvectors ~ai and ~aj then the scalar quantity
~a†i · K · ~aj = ~a†i · ω22 M · ~aj = ωj2 ~a†i · M · ~aj
Ä ä
5
The usual eigenvalue equation has M = m1. Here I relax this restriction.
6
In quantum mechanics the matrices corresponding to observables are Hermitian, H† = H, instead.
7 For any vector ~˜ the quantity 1 ~
x ˜† · K · ~
x ˜ is the potential energy of the springs. Since I am using equilibrium coordinates this quantity is always
x
2
x = 0). Consequently, K is positive definite. Similarly, 12 ~
positive (except for the trivial solution, ~ ˜˙ · M · ~
x ˜˙ is the kinetic energy.
x
8 In index notation,
† †
x̃†i kij x̃j † †
= x̃i kij x̃j = x̃i kij x̃†j
= x̃j kji x̃†i = x̃†i kij
†
x̃j .
1 ˜† ˜.
This quantity is twice the potential energy of this system, U = 2
~
x ·K·~
x
2.2. COUPLED OSCILLATORS 37
ä†
= ~a†j · ω12 M · ~ai = ω12 ~a†i · M† · ~aj .
Ä Ä ä
T
= ω12 ~a†i · M · ~aj .
Ä ä
Subtracting, I have
Ä
0 = ωi2 − ωj2 ~a†j · M · ~ai
ä
AF
Example 2.4 Energetic Oscillators.
The kinetic and the potential energies for the coupled oscillators from Example 2.3 are
that Ė = 0.
Solution
(a) The time derivative of twice the total energy is
2Ė =
= ~x
T = 12 ~x˙ · M · ~x˙
(a) Find an expression for the total time derivative of the energy, Ė =
d Ä˙
dt
& U = 12 ~x · K · ~x.
~x · M · ~x˙ + ~x · K · ~x
¨ · M · ~x˙ + ~x˙ · M · ~x
= 2~x˙ · M · ~x
Ė = ~x˙ · M · ~x
Ä
ä
d
dt E.
¨ + ~x˙ · K · ~x + ~x · K · ~x˙
¨ + 2~x˙ · K · ~x, M & K are both symmetric
¨ + K · ~x .
ä
(b) This is a conservative system. Show
X
T = 1
2
θ̇i mij θ̇j = 1
m θ̇2
2 11 1
+ 12 m12 θ̇1 θ̇2 + 21 m21 θ̇2 θ̇1 + 21 m22 θ̇22 , = 1
m θ̇2
2 11 1
+ m12 θ̇1 θ̇2 + 12 m22 θ̇22
ij
where m12 = m21 (i.e. the mass matrix is symmetric), then the mass matrix is exactly M = [mij ].
38 CHAPTER 2. COUPLED OSCILLATIONS & NORMAL MODES
`1
θ1
~g
T
m1
θ̂1
r̂1
`2
ê2 θ2
ê1 m2
θ̂2
r̂2
AF Figure 2.8: The double pendulum with different possible bases, the Cartesian basis, êi , and two possible polar bases, r̂i &
θ̂i . For the polar bases the r̂i unit vectors are in the direction of increasing ri and the θ̂i unit vectors are in the direction of
increasing θi . Crucially, the polar bases are not constant unit vectors, rather they move with the position vectors, ~ri .
cos(θ2 − θ1 ).)
Exercise 2.3 String Theory.
î
θ1
θ2
ó
. What are the mass and spring matrices, M & K, for this problem? (f) This
is a conservative system so Ė = 0. Use this to find the EOMs for this system. (Hint: See Example 2.4.) (g) What
are the velocities of the two bobs, ~r˙ , in terms of the r̂i -θ̂i bases? (Hint: Recall from subsection A.2.1 that ṙ = θ̂θ̇
and θ̇ = −r̂θ̇.) (h) Use your answers to part (g) to find the potential and kinetic energies. (Hint: Note that r̂1 · r̂2 =
3`
m
m
DR
A light, extensible string is secured horizontally between two walls a distance 3` apart. Two masses, each of mass
m, are secured to a at points equidistant from each other and from the walls as shown. The tension in the string,
F , is much greater than mg so that during the small transverse oscillations of the two masses the tension can be
considered a constant (and, hence, the string is horizontal when the system is in equilibrium and the oscillations are
small and transverse). The oscillations are transverse. (a) What are the equations of motion for the two masses?
Use x1 and x2 as the p distances of the masses from their equilibrium positions (i.e. equilibrium coordinates). Please
use the constant Ω = F/m` to simplify your equations. The oscillations are transverse. (b) Make the usual ansatz,
xi (t) = ai cos(ωt), to find two linear equations in the ai . Write your result in matrix form, M · ~a = 0, where ~a = [ aa12 ]
and M is a 2 × 2 matrix. (Note that there will be no part marks for the subsequent parts of this question if this part
is wrong so please triple check your answer for M.) (c) What are the frequencies, ωi , of the system? (d) What is the
normalized eigenvector corresponding to the “fast mode”, âfast ? (e) Will the general motion be periodic? Justify your
answer. (f) Find the frequencies and normal modes for three equal masses spaced equidistantly along a similar string
of length 4`?
`
θ1
~g
T
3m
θ̂1
r̂1
`
ê2 θ2
ê1 m
θ̂2
r̂2
AF
You mayP
~r˙ = θ̇i ∂θ
∂~
r
i
Figure 2.9: The double pendulum with `1 = `2 ≡ ` and m1 = 3m2 ≡ 3m. cf. Figure 2.8 and Exercise 2.2.
use either the Cartesian basis, êi , of the polar bases, r̂i & θ̂i . [2017S midterm] (a)P
¨
˙
¨
Ė = 0 = θ~ · M · θ~ + K · θ~
˙
E = 12 θ~ · M · θ~ + 12 θ~ · K · θ~ + U◦ ,
˜ ˜ −iωt
and the equations ofmotion are M · θ~ + K · θ~ = 0. We can the ansatz θ~ = ~a e
Find the potential energy,
U = − i mi~g · ~ri . (Hint: Find ~r = ~r(θ~ ) first.) (b) Find the kinetic energy, T = 21 i mi~r˙i · ~r˙i . (Hint: Find
first.) (c) Taylor expand your potential and kinetic energies (about θi and θ̇i , respectively) and discard any
terms higher than quadratic.11 Then write the energy, E = T + U , in the form
˙
simplify the algebra.) (e) What are the eigenvectors, â± ? (f) Describe (in words or pictures) the two normal modes.
(g) Show that the eigenvectors are orthogonal. (h) Find M−1 , re-write the equations of motion as
DR
¨ ¨
M−1 · M · θ~ + M−1 · K · θ~ = θ~ + M−1 · K · θ~ = 0,
11 The multidimensional Taylor expansion of f (~
x) about ~a is
x) ≈ f
f (~ + (~ ~
x − ~a)T · ∇f + 21 (~ ~
~ ⊗ ∇f
x − ~a)T · ∇ · (~
x − ~a) + · · · ,
~
x=~
a ~
x=~
a ~
x=~
a
~ ⊗∇
in which ∇ ~ is the outer product of two gradient operators (aka the Hessian matrix),
∂2 ∂2 ∂2
···
∂x1 ∂x1 ∂x1 ∂x2 ∂x1 ∂xn
∂2 ∂2 ∂2
···
~ = ∂x2 ∂x1
~ ⊗∇
∇ ∂x2 ∂x2 ∂x2 ∂xn ,
.. .. .. ..
. . . .
∂2 ∂2 ∂2
···
∂xn ∂x1 ∂xn ∂x2 ∂xn ∂xn
∂2
~ ⊗∇
or, in index notation, ∇ ~ = ~ ⊗∇
, or even more concisely as ∇ ~ = ∂i ∂j . Clearly we would want to use this notation to find
ij ∂xi ∂xj ij
higher order terms,
x) ≈ f
f (~ + (xi − ai ) ∂i f + 21 (xi − ai )(xj − aj ) ∂i ∂j f + 1
3!
(xi − ai )(xj − aj )(xk − ak ) ∂i ∂j ∂k f + ··· .
~
x=~
x◦ ~
x=~
a ~
x=~
a ~
x=~
a
`1
40 θ1
CHAPTER 2. COUPLED OSCILLATIONS & NORMAL MODES
m1
use the ansatz, and find the eigenvalues of the resulting equation. θ2 `2
Example 2.5 Triple Pendulum. m2
Consider the triple pendulum as shown to the right with `
m1 = 6m, m2 = 2m, m3 = m and `1 = `2 = `3 = `. Use θ3 3
the angles θa , a = 1, 2, 3, as your generalized coordinates. m3
T
(a) What are the positions, ~ri , i = 1, 2, 3, of the three masses in terms of the lengths `i and the angles θa ? (b) What
are the velocities, ~r˙i , i P
= 1, 2, 3, of the three masses in terms of the lengths `i and the angles θa ? (c) What is the
kinetic energy, T = 21 i mi~r˙i · ~r˙iP ? Can you think of a clever short-cut to obtain the same result? (d) What is the
gravitational potential energy, U = i −mi~g · ~ri ? (e) What are the spring and mass matricies,
X
∂~ri ∂~ri ∂ 2 U
[M]a,b = mi · & [K]a,b =
i
∂θa ∂θb 0 ∂θa ∂θb 0
in the small angle approximation? (f) Use θa = Aa eiωt and the equation for small oscillations,
AF ¨
M · θ~ + K · θ~ = 0,
to find the frequencies of the normal modes. (Hint: If you introduce the dimensionless constant λ = `ω 2 /g then
λ = 3 is your friend.) (g) Calculate and then sketch each of the normal modes. (Hint: Check that they are orthogo-
~ (i) What is
nal.) (h) What are the normal coordinates for this system, ~q, in terms of the generalized coordinates, θ?
Lagrange’s equation in matrix form in terms of the normal coordinates? Write out the kinetic and potential energy
matrices explicitly.
DR
T
Chapter 3
AFResources
Thornton & Marion – §11.3-7
§10.1-7 of Boas
Appendix A: Vectors
Appendix B: Index Notation
Appendix C: Vector Multiplication
Appendix D: Coordinate Systems Transformations
Appendix E: Rotation Generator
Appendix F: Rotations
Appendix G: Vector Calculus
Appendix H: Linear Algebra
Week 3
3.2 Tensors
Vector quantities such as displacement, momentum, etc, are all rank-1 tensors. Scalar quantities such as mass or
distance are rank-0 tensors. Rank-2 tensors are also common in physics but, by appropriate choice of coordinate
41
42 CHAPTER 3. THE INERTIA TENSOR
T
time). tation of the vector (technically, xµ is a contravariant four-
It is convenient to introduce x0 ≡ ct so that the four- vector while xµ is a covariant four-vector).
vector in index notation can be written as ~τ = xµ êµ , µ = The rule, then, for doing dot products in index notation is
0, 1, 2, 3 (Greek rather than Roman indices are used as a that any sum over a repeated index must have one upper and
nemonic that this xµ represents an event rather than a po- one lower index.
sition, with µ = 0, 1, 2, 3. The reason for the upper index, In terms of matrices, I can write
xµ , will be explained shortly).
The “distance” of the event, ~τ , from the origin, known as [xµ ] = [x0 , x1 , x2 , x3 ]
the proper time interval, τ , is x0
−x1
AF
are non-zero).
τ 2 = ~τ · ~τ = (xν êν ) · (xµ êµ )
= xν xµ (êν · êµ ).
0
0
0
0
−1
.
↔
so that
µ
[∂ ] =
c
[j ] = cρ, J~ ,
µ
1
µ T
[xµ ] = g[x ] =
∂ ~
c ∂t
,∇
−x2
−x3
τ 2 = xν xν = [xµ ][xµ ].
The shorthand [xµ ] = [ct, ~r ] is often used to represent a
four-vector for position (aka event), so that τ 2 = (ct)2 −~r ·~r.
Other four-vectors of interest are
[pµ ] =
1
E, p
~ ,
system, we can often treat them in a mathematically less sophisticated manner (i.e. so that only the diagonal elements
four-momentum
four-current
four-gradient.
Examples of rank-2 tensors are the stress-strain tensor, σ , from which the net force on a surface S can be calcu-
lated, Z
↔
F~ = σ · d~a,
DR
S
↔
the inertia tensor, I , from which the angular momentum of an object can be calculated,
↔
~ = I ·ω
L ~
↔
and the susceptibility tensor, χ, from which the dielectric polarization density, P~ , can be calculated,
↔
P~ = ε0 χ · E.
~
It is important to note that ~v is a real vector representing a real physical quantity such as velocity while the triplets
[v1 , v2 , v3 ] and [v10 , v20 , v30 ] are representations of the vector in terms of specific set of basis vectors (i.e. {ê1 , ê2 , ê3 }
and {ê01 , ê02 , ê03 }).
The laws of physics do not depend upon the choice of basis vectors and we can express them in a basis-independent
manner. This key insight allows us exploit our freedom to choose a basis to simplify our calculations (usually by
T
identifying symmetries in the underlying system).
AF
in which the integral is over the object.
If the angular velocity of the object is ω
~r × ~v = ~r × (~
ω × ~r) = (~r · ~r)~
ω − (~r · ω
~◦ =
L
~ )~r.
Z
dm (xk xk δij − xi xj )ωj êi
Z
= ωj êi dm (xk xk δij − xi xj )
~ × ~r, so that
= êi Iij ωj
DR
where Z
Iij = dm (xk xk δij − xi xj )
~ × B)
(A ~ · (C
~ × D)
~ = (A
~ · C)(
~ B ~ · D)
~ − (A
~ · D)(
~ B ~ · C).
~
44 CHAPTER 3. THE INERTIA TENSOR
T
Z
= 21 ωi ωj dm (xk xk δij − xi xj )
↔
= 21 ωi Iij ωj = 12 ω
~TI ω
~,
where Z
Iij = dm (xk xk δij − xi xj )
AF
3.2.3 Multilinearity
A Linear Function
A linear function, f , of a vector is a function that accepts a single vector as its argument and returns a scalar, and that
has the property
f (a~u + b~v ) = a f (~u) + b f (~v ).
This means that if I write a vector in terms of its components, ~v = vi êi , then
f (~v ) = vi f (êi ),
so if I know how the basis vector are transformed by the function then I know how any vector is transformed.
Since I only need to know how the basis vectors transform it is convenient to introduce the notation
so that
fi ≡ f (êi )
f (~v ) = vi fi ,
and I can see that the requirement that f (êi ) be a linear function has resulted in the equivalent of a dot product,
DR
f (~v ) = f~ · ~v , where f~ = [fi ] is the matrix with elements fi . Note, in particular, that if I had used a different basis,
~v = vi0 ê0i , then I would have fi0 ≡ f (ê0i ) (with f (~v ) the same function) and
This is a long way of saying that linear algebra is an appropriate vehicle for linear functions.
A Bilinear Function
A bilinear function of two vectors is a function that accepts two vectors as its arguments and returns a scalar and that
has the property
f (a~u + b~v , cw
~ + d~x) = a f (~u, cw
~ + d~x) + b f (~v , cw
~ + d~x)
= ac f (~u, w)
~ + ad f (~u, ~x)
+ bc f (~v , w)
~ + bd f (~v , ~x).
This means that if I write two vector in terms of their components, ~u = ui êi & ~v = vj êj , then
so if I know how the basis vector are transformed by the function then I know how any vector is transformed.
3.3. TENSORS AND THEIR REPRESENTATION 45
This can be extended in a straightforward manner to a multilinear function of r vectors which returns a scalar. We
will largely stick with r = 2 but larger values are not uncommon in physics.
It is convenient to introduce the notation
fij = f (êi , êj )
so that we can write
T
f (~u, ~v ) = ui vj f (êi , êj ) = ui fij vj ,
which has the usual vector-matrix-vector interpretation,
f (~u, ~v ) = ~u · F · ~v ,
where F ≡ [fij ] is a matrix with elements fij . It is important to note that this last equation is independent of the choice
of basis.
AF
3.3 Tensors and their Representation
Consider first the tensor product of two vectors,
The quantity ~u ⊗ ~v (or êi êj ) is called a dyad (sometimes written as êij ) and a dyadic is a sum of dyads.1 Tensor
products are also called dyadic products in some contexts. Each dyad has a magnitude and two directions associated
with it (which may be the same or different). In the above example the dyad ~u ⊗ ~v has magnitude uv and directions
û ⊗ v̂. How we interpret the magnitude and directions depends upon the application.
Note that in these expressions the product of the two basis vectors, êi êj , is not a inner product but is, rather, the
tensor, outer or dyadic product; each pair of unit vectors is called a dyad.
If we think of ui vj as a square matrix then êi êj is the “basis” for that matrix.
↔
The identity rank-2 tensor, 1 is defined by
↔ ↔
1 · ~v = ~v · 1 = ~v ,
and for a êi basis I have ~v · êi = vi which is the ith component of ~v . I can then write
DR
~v = vi êi = (~v · êi )êi = ~v · (êi êi )
and
↔
1 = êi êi .
With this definition I can show that
↔
1 · ~v = (êi êi ) · (vj êj ) = vj êi (êi · êj ) = vj êi δ + ij = vi êi = ~v ,
↔
and 1 is the identity for both left and right multiplication.
However, consider the left and right inner product of the vector ~u with each dyad. First the left inner product:
~u · (êi êj ) = (uk êk ) · (êi êj ) = uk (êk · êi )êj = uk δki êj = ui êj
1 If we use êi as our basis then we can represent a vector as a column vector, e.g.
î v1 ó
vê1 → 0 .
0
Similarly, the dyads êi êj can be considered as the basis for a matrix, e.g.
0 00
î ó
v21 ê2 ê1 → v21 0 0 .
0 00
T
In general, the representation of a rank-2 tensor (with respect to a particular basis) will have two indices, aij :
↔
A = a11 ê1 ê1 + a12 ê1 ê2 + a21 ê2 ê1 + a22 ê2 ê2
= aij êi êj
a11
ï òï ò
a12 ê1
= ê1 , ê2 ,
a21 a22 ê2
in which the usual matrix-matrix multiplication rules are followed.
↔
The aij = [A]ij are the representation of A in the êi êj basis.
AF In cases in which the basis in unambiguous it is sometimes written as just the matrix representation,
A=
ï
a11 a12
a21 a22
ò
but this is a only a specific representation of the tensor rather than the actual tensor.
Because of the potential confusion over which type of vector-vector multiplication is involved the symbol ⊗ is
sometimes used,
inner product: ~u · ~v = ui vi
cross product: ~u × ~v = êi εijk uj vk
outer product: ~u ⊗ ~v = (ui êi ) ⊗ (vj êj ) = ui vj (êi êj ).
The inner product of a rank-2 tensor with a vector (i.e. a rank-1 tensor) is
↔ a a12 ê1 v
ï òï ò ï ò
A · ~v = ê1 , ê2 11 · ê1 , ê2 1
a21 a22 ê2 v2
a11 a12 ê1 · ê1 ê1 · ê2 v1
ï òï òï ò
DR
= ê1 , ê2
a21 a22 ê2 · ê1 ê2 · ê2 v2
a11 a12 v1
ï òï ò
= ê1 , ê2
a21 a22 v2
a v + a12 v2
ï ò
= ê1 , ê2 11 1
a21 v1 + a22 v2
a v
ï ò
= ê1 , ê2 1j j = ê1 a1j vj + ê2 a2j vj = êi aij vj
a2j vj
↔
which is another vector. It is convenient to remember that [A · ~v ]i = aij vj , rather than working through this matrix
↔
algebra each time (but to keep in mind that the aij and vj are representations of A and ~v using a specific basis, êi ).
↔
A very important caveat to the above derivation for A · ~v is that it relies upon the basis, êi , being orthonormal.
If the intermediate matrix, [êi · êj ], was not the identity matrix, 1, then this calculation would be complicated by the
presence of this matrix, called a metric3 , which would tell us how to do dot products in the non-orthonormal basis.
Examples of where this occurs are special and general relativity (see the sidebar on p. 42 for a brief discussion of the
Minkowski metric).
2
See Appendix C for a discussion on the different types of vector multiplication.
3
The metric tensor generalizes the dot product for non-orthonormal bases. It is called the metric because the dot product is what defines lengths
and angles.
3.3. TENSORS AND THEIR REPRESENTATION 47
Solution
T
(a) First the left inner product:
~ · (~u ⊗ ~v ) = (wk êk ) · ui vj (êi êj ) = ui vj wk (êk · êi )êj = ui vj wk δki êj = ui vj wi êj = (~u · w)~
w ~ v
(~u ⊗ ~v ) · w
~ = ui vj (êi êj ) · (wk êk ) = ui vj wk êi (êj · êk ) = ui vj wk êi δjk = ui vj wj êi = (~v · w)~
~ u
(c) These two expressions are clearly different. The inner product of a vector with a dyadic does not commute. X
AF
3.3.1
so that
where
Transforming Tensors between Representations
Recall that the relationship between two sets of basis vectors, ê0k & êi , is given by (see Appendix F)
λT
ê0k =λk` ê`
0
⇒ êi =λT
↔
T
ik êk =λik λk` ê` = δi` ê`
0
is the representation of the tensor [A0 ]ij = a0ij in the primed basis. We see that rank-2 tensors require two factors of
λij , one for each index. Such transformations are are termed similarity transformations and are also widely used in
DR
quantum mechanics (see, for instance, Example 12.10 and Exercise 12.10).
In matrix notation I can write the last equation above as
A0 = ΛAΛT
↔
and say that the matrices A and A0 are similar in the sense that they represent the same tensor quantity, A, in two
different bases. See subsection H.1.2 for a discussion of an intuitive way to understand similarity transformations.
T
↔
~ · A = (êi ∂i ) · (êj ajk êk ) = ∂i ajk êk δij = ∂j ajk êk ,
∇
which is a rank-3 tensor (and (êi êj êk ) is a triad). In general the divergence operation reduces the rank of the tensor by
one while the gradient operation increases it by one. It may be useful to remember the expressions
AF î
∇
∇
ó
~ ⊗ ~v = ∂i vj & ∇
ij
~ · ~v = ∂i vi & ∇
h
h
~ ⊗A
↔
↔i
i
ijk
= ∂i ajk .
~ · A = ∂i aij ,
j
in the first pair there are only free indices while in the second two there is one summation index.
~ · ~v ≡ lim 1
∇
∆τ →0 ∆τ ∂V
I
d~a · ~v .
This result is valid for arbitrary closed surface ∂V with volume ∆τ . So we can build a finite volume, V, (with surface
∂V) as the sum of N volumes, ∆τi , i = 1...N ,
X XI I
~
∆τi ∇ · ~v = d~a · ~v = d~a · ~v ,
i i ∂Vi ∂V
DR
in which we have canceled the fluxes through all of the internal surfaces.
We can then take the N → ∞ limit (or the ∆τi → dτ limit) and convert the sums into integrals to get,
Z I
dτ ∇~ · ~v = d~a · ~v ,
V ∂V
T
AF
Figure 3.1: The forces on each face of a cubical element within a solid can have components parallel and perpendicular to
the face. The perpendicular component is a tensile force (or compressive if it is negative) and the components parallel to each
face make up the shear force.
image by Sanpaz (Own work) [CC BY-SA 3.0 or GFDL], via Wikimedia Commons
is
↔
↔
In order to describe these forces we use the Cauchy stress tensor, σ , such that the force on the surface element d~a
dF~ = d~a · σ .
(a) Using the orthonormal êi basis, what are the forces, dFi , on the infinitesimal area d~a? (b) What is the in-
finitesimal force, dF~ on the infinitesimal area, d~a = da ê1 (i.e. perpendicular to ê1 )? (c) Briefly describe each term
DR
in your answer to part (b). (Hint: See Figure 3.1) (d) What is the net force on the material enclosed by the surface S?
(e) Briefly describe the result if the Cauchy stress tensor is diagonal in a certain representation and an applied force
is parallel to the ê1 basis vector. (f) Briefly describe the result if the Cauchy stress tensor is not diagonal in a certain
representation and an applied force is parallel to the ê1 basis vector.
Solution
(a) I have
↔
dF~ = d~a · σ = (dai êi ) · (êj σjk êk ) = dai σjk êk δij = daj σjk êk .
dF~ = da σjk êk δj1 = da σ1k êk = da (σ11 ê1 + σ12 ê2 + σ13 ê3 ).
(c) The first component, da σ11 , is perpendicular to the area d~a and is a tension or compression (depending upon its
sign). The other two components are parallel to the area d~a and are the two components of the shear force (see
Figure 3.1).
(d) The net force on the surface, S, is Z
↔
F~ = d~a · σ .
S
50 CHAPTER 3. THE INERTIA TENSOR
↔
(e) If the tensor σ is diagonal and a force is applied along one of the basis directions then the substance will deform
in a uniform manner along that direction (for the case of an isotropic material the stress tensor then behaves as a
scalar, the hydrostatic pressure).
↔
(f) If σ is not diagonal and a force is applied along one of the basis directions then there will be a shear force acting
T
orthogonal to the applied force and the substance will deform in direction that is not parallel to the applied force.
X
AF ↔
where σ is the Cauchy stress tensor. Introduce the force per unit volume, f~ ≡
Solution
I have
so that
I
Z
d~a · σ +
V
↔
↔
d~a · σ =
Z
↔
dτ f~ = 0.
Apply the divergence theorem to the first term and so find an expression for the force density.
~ · σ + f~) = 0.
dτ (∇
~ ·σ
dτ ∇
↔
~
dF
dτ , so that for equilibrium
Since the volume V is arbitrary this expression must hold for any choice of volume and thus the integrand must be
zero,
~ ·↔
f~ + ∇ σ = 0 or fj + ∂i σij = 0.
DR
This is known as Cauchy’s first law. X
(a) Write the above equation using index notation. (b) Apply the divergence theorem,
I Z I Z
d~a · ~v = ~ · ~v
dτ ∇ → dai vi = dτ ∂i vi .
∂V V ∂V V
(c) Identify those parts of the integrand that are zero by Cauchy’s first law. (d) Show that the stress tensor must be
symmetric,
↔T ↔
σ =σ or σij = σji .
∂xi
(Hint: ∂xj = δij .)
3.5. THE INERTIA TENSOR 51
T
Consider once again now the rotational kinetic energy of a rigid body,
Z
T = 21 dm ~v · ~v
Z
˙ ˙
= 12 dm (θ~ × ~r) · (θ~ × ~r)
Z
˙
= 21 dm θ̇2 r2 − (θ~ · ~r)2
AF
in which I have used the identity from Example 12.5,
where
~ × B)
(A ~ · (C
~ × D)
~ = (A
~ · C)(
~ B ~ · D)
1
Z Ä
↔
Z
⇒
~ B
ä
~ · C).
↔
ä
~
˙ ↔ ~˙
T = 12 θ~ · I · θ,
I = dm (~r · ~r) 1 − ~r ⊗ ~r
↔ ↔
DR
is the inertia tensor (and 1 ≡ êi êi . is the identity dyadic). In index notation I have I = êi Iij êj so that
Z
[I]ij = Iij = dm (xk xk δij − xi xj )
~ × (B
A ~ × C)
~ = (A
~ · C)
~ B~ − (A
~ · B)
~ C.
~
T
= dm (δij θ̇j )xk xk − θ̇j (xi xj )
Z
=θ̇j dm (xk xk δij − xi xj )
↔
=Iij θ̇j ⇒ ~ = I · θ~˙
L
↔
where I is, again, the inertia tensor.
AF
3.5.1 Properties of the Inertia Tensor
Z
Ñ
⇒ I = dm 0
xk xk
0
x2 x2 + x3 x3
= dm −x2 x1
−x3 x1
0
xk xk
0
0
xk xk
−x1 x2
x1 x1 + x3 x3
−x3 x2
I11 =
Z
↔
In a specific orthonormal basis, êi , the inertia tensor, I , is represented as I, a 3 × 3 matrix,
Z
dm (xk xk δij − xi xj )
[I]ij = Iij =
x1 x1 x1 x2 x1 x3
0 − x2 x1 x2 x2 x3 x3
x3 x1 x3 x2 x3 x3
−x1 x3
−x2 x3
x1 x1 + x2 x2
dm (xk xk − x1 x1 )
é
Z
= dm (x1 x1 + x2 x2 + x3 x3 − x1 x1 )
DR
Z
= dm (x2 x2 + x3 x3 )
and similarly for I22 and I33 . The quantity (x2 x2 + x3 x3 ) is the perpendicular distance of dm from the x1 axis and is
what we were taught in first year is the the moment of inertia about the x1 axis – this is only the case if Iij is diagonal.
The off-diagonal elements, called products of inertia, have the form
Z !
0
Iij = − xi xj ,
dm xk xk δij i 6= j
Z
=− dm xi xj .
↔ ↔T
The inertia tensor is symmetric (i.e. I = I ) and so there is a basis in which its representation, I, is diagonal.
This choice of basis will, as we shall see, greatly simplify many calculations.
Example 3.4 Moments of Inertia.
Calculate the moments of inertia of the following objects about their COM. (a) Rod, length `, mass M , about perpen-
dicular axis, (b) Disk, radius a, mass M , about perpendicular axis (c) Sphere, radius a, mass M , about perpendicular
axis (d) Cone, radius a, height h mass M , about axis
3.5. THE INERTIA TENSOR 53
Eigenvalues & Eigenvectors case the eigenvectors of the degenerate subspace must be or-
thonormalized) .
The mathematical process used to find the eigenvalues and
eigenvectors is very important in physics (and engineering).
It is supposed to be taught in your first year algebra course Diagonalization
and it will certainly be used extensively in any quantum me-
T
We want to find the matrix U such that the quantity UKU−1
chanics course you take. is diagonal. If the columns of U−1 are the orthonormal eigen-
The process is straightforward. In order to find the eigen- vectors of K,
values and eigenvectors of a square n × n matrix K we write U = [v̂1 v̂2 ... v̂n ]
K~v = λ~v ⇒ (K − λ1) ~v = 0 then U−1 = UT , the transpose of U since the eigenvectors
form an orthonormal set (i.e. v̂i · v̂j = δij ).
where 1 is the unit matrix. For a non-trivial solution (i.e.
Now we have
~v 6= 0) we must have
AF
This yields a polynomial in λ of degree n, where n × n is
the size of the matrix K. The roots of this polynomial are the
eigenvalues.
For each eigenvalue, λi , we go back to the defining equa-
tion, (K − λi 1) ~vi = 0, to find the corresponding eigenvec-
tors, ~vi , to within an overall multiplicative constant (which
we choose so that the eigenvectors have unit length, v̂i ).
Things can become slightly more complicated if some of
the eigenvalues are complex (in which case we need the Her-
mitian conjugate for inner products) or repeated (in which
Solution
and since , the eigenvectors form an orthonormal set
UT KU =
λ1
0
0
0
..
.
0
λ2
0
0
..
.
(a) With the origin at the centre of the rod, and with λ = M/` the linear density so that dm = λ dx, I have
Irod =
Z
dm x21 = λ
Z `/2
dx x2
0
0
λ2
0
..
.
···
···
···
..
···
.
0
0
0
..
λn
−`/2
1
3 `/2 M2 3 1 2
=λ 3 x −`/2 = ` = 12 M a
3` 8
DR
(b) The moment of inertia about the cylinder axis, ê3 , is easy to calculate with cylindrical coordinates. With ρ =
M/πa2 h for the volume density so that dm = ρ dτ , and with dτ = dz dφ s, ds I have
Z 2
:s
Idisk = dm (x 1 x
1 + x2 x2 )
Z h/2 Z 2π Z a
=ρ dz dφ s ds s2
−h/2 0 0
a
= ρ(h)(2π) 41 s4 0
= 12 πρha4
M
Å ã
= 12 π ha4
πa2 h
= 12 M a2
2
(c) Isphere = M a2
5
3
(d) Icone = M a2 . X
10
54 CHAPTER 3. THE INERTIA TENSOR
T
eigenvectors, v̂i , of I? (e) Sketch the cube, the basis vectors, and the eigenvectors. (f) What is the orthogonal matrix,
A, such that I0 = AT IA is diagonal? (g) Starting with the inertia tensor, I◦ , from part (a), use the parallel axis theorem
to find the inertia tensor about one corner of the cube. (h) Identify the principle axes for a cube with the origin at one
corner using symmetry arguments. (i) Using only the moment of inertia of a rod of mass m and length a about an axis
through its centre of mass and perpendicular to the rod (ie, ma /12), calculate the principle moments of inertia for a
2
AF
If an object is not fixed at one point then we can break the motion into two components, translation of the COM and
rotation about the (CoM) (i.e. we can perform a Galilean transformation to the COM coordinate system).
M~c =
◦
Iij = Iij◦
then, relative to a point displaced ~c from the CoM the inertia tensor
+ M ((ck ck )δij − ci cj )
Proof: The CoM of a rigid body of total mass M is located at the point ~c = ci êi defined by
Z
dm ~r or ci =
1
M
Z
dm xi
and the tensor matrix for an origin at −~c relative to the COM is
◦
Iij = Iij + M (ck ck δij − ci cj )
3.6. USEFUL THEOREMS 55
T
Example 3.5 Easy Cube.
Use the PAT to calculate the inertia matrix I for the cube in Exercise 3.2 by starting from I◦ at the COM.
Solution
(a) I will need to add the quantity M (ck ck δij − ci cj ) to I◦ , the inertia matrix of the CoM system. Here I have ~c = ci êi
with ci = 12 a so that
Ñ é
2 1 0 0 1 1 1
Ma
M (ck ck δij − ci cj ) = 30 1 0 − 1 1 1
AF ⇒I=
=
M a2
12
M a2
12
=
0 2 0 +
0 0 2
8 −3 −3
−3 8 −3 2
−3 −3 8
M a2
4
+ M (ck ck δij − ci cj )
2 0 0
M
a
12
0 0 1
2 −1 −1
−1 2 −1 2
−1 −1 2
2
6
−3
−3
1 1 1
−3
6
−3
−3
−3
6
in which I have put the summations in explicitly (because I have violated the summation convention).
56 CHAPTER 3. THE INERTIA TENSOR
T
ê1 axis is orthogonal to the mirror plane). Then any off-diagonal elements of the inertia matrix containing x1
(i.e. the products of inertia I12 and I13 ) will be zero (because for every element of mass, dm, with position +x1
there is a corresponding element, dm, with position −x1 ). This will be one of the principle axes.
2. If the rigid body has a rotational symmetry then the rotation axis will be a principal axis.
3. If the order, n, of the rotational symmetry (i.e. if a rotation of 2π/n is a symmetry operation then the order of
the rotational symmetry is n) is greater than two then any two axes perpendicular to the rotational symmetry
axis will also be principles axes and the moments of inertia about these axes will be equal. This is known as
dynamical axial symmetry.
AF 4. If the object has two different axes with dynamical axial symmetry then any three mutually perpendicular axes
can be principle axes and the moment of inertia for any axis will be the same. This is known as dynamical
spherical symmetry and in this case the inertia tensor can be treated as a scalar (i.e. Iij = Iδij ).
Most objects of interest will have some symmetry and we can then exploit these rules to find the principle axes and,
along with the parallel and perpendicular axis theorems (and other tricks), simplify the calculation of the inertia tensor.
Example 3.6 Inertia Tensor of a Cylinder.
Consider a uniform cylinder of mass m, radius a and height h. (a) What are the principal axes of the cylinder?
Explain your reasoning. (b) What are the moments of inertial about the principal axes? (c) What is the inertia tensor
about the centre of the end of the cylinder? (d) What is the inertia tensor about an edge of the cylinder?
Solution
(a) The cylinder has a rotational symmetry about its axis so this is a principal axis. Any plane containing this axis is
a mirror plane so the other two principal axes are any pair of axes perpendicular to the cylinder axis (and to each
other).
(b) I will use cartesian coordinates with the origin at the centre of the cylinder and the ê3 axis along the cylinder axis.
Since these are the principal axes for the cylinder the inertia tensor will be diagonal.
The moment of inertia about the cylinder axis, ê3 , is easy to calculate with cylindrical coordinates. With
DR
dτ = dz dφ s, ds and V = πa2 h I have ρ = m/πa2 h for the density and
Z 2 Z h/2 Z 2π Z a
◦
:s
I33 = dm (x 1 x
1 + x2 x2 ) = ρ dz dφ s ds s2
−h/2
0 0
a m
=ρ(h)(2π) 14 s4 0 = 12 πρha4 = 12 π ha4 = 21 ma2
πa2 h
The moment of inertia about the orthogonal axes, ê1 & ê2 , are equal but trickier to calculate. With x2 = s sin φ
I have
Z Z h/2 Z 2π Z a
◦ ◦ s2 sin2 φz 2
I11 = I22 = dm ( x2 x2 +
*
x3 x3 ) = ρ
*
dz dφ s ds (s2 sin2 φ + z 2 )
−h/2 0 0
Ç Z a Z 2π Z h/2
å
=ρ h ds s3 dφ sin2 φ + πa2 dz z 2
0 0 −h/2
m
1 4 2 1 3
= h 4a (π) + πa 12 h = 14 ma2 + 1
12 mh
2
πa2 h
The inertia tensor for this choice of axes is then
2
3a + h2 0 0
m
◦
I = 0 3a2 + h2 0
12
0 0 6a2
3.6. USEFUL THEOREMS 57
T
1 0 0 0 0 0 2 1 0 0
h
(ck ck )δij − ci cj = c2 0 1 0 − c2 0 0 0 = 0 1 0
4
0 0 1 0 0 1 0 0 0
and
2 2
3a + h2 0 0 h 0 0
m m
Iij = 0 3a2 + h2 0 + 0 h2 0
12 4
0 0 6a2 0 0 0
2
AF and
=
Iij =
m
12
m
12
3a + 4h
2
0
0
3a + h2
2
0
0
3a + h2
2
1 0 0
0
3a2 + 4h2
0
(ck ck )δij − ci cj = a2 0 1 0 − a2 0
0 0 1
0
0
0
6a2
The inertia tensor remains diagonal because the x0 and y 0 axes are orthogonal to mirror planes.
(d) I will again use the parallel axis theorem with ~c = aê1 I have ck ck
3a2 + h2
0
0
1
0
= a2 so
0 0
0 0
0
0 + m0
6a2
0
0
0
0
0 0 = a2 0
0
a2
0
0
0
a2
0
1
0
0
1
0
m
= 0 15a2 + h2 0
12
0 0 18a2
DR
The inertia tensor remains diagonal because the y 0 and z 0 axes are orthogonal to mirror planes. X
This is the perpendicular axis theorem. It applies to a thin plate of any shape, known as a laminar, meaning “flat.”
Example 3.7 Inertia Tensors.
Find the inertia tensors of a rod, a disk and a cone about their COM. Use the results from Ex. 3.4 and the perpendicular
axis theorem. Take ê3 along the axis used in Ex. 3.4. Explain your reasoning.
Solution
(a)
1 0 0
1
Irod = M `2 0 0 0
12
0 0 1
58 CHAPTER 3. THE INERTIA TENSOR
(b) 1
0 0
1 2
Idisk = M a2 0 1
2 0
2
0 0 1
(c)
T
1
0 0
3 2
Icone = M a2 0 1
2 0
10
0 0 1
X
Example 3.8 Inertia Tensor of a Rectangular Plate.
Consider a rectangular plate of with dimensions `1 × `2 × `3 and mass M . (a) What are the principal axes? Justify
your answer. (b) Consider the inertia tensor of the plate about its COM with respect to the principal axes. Write
out the three integrals for the diagonal components (but don’t perform the integrals). Introduce the volume density,
AF
ρ ≡ M/`1 `2 `3 . (c) Now consider the case in which
2
R
Ii = dm si = ρ
`3 `1 , `2 (i.e. a thin plate). Write out the three integrals for the
diagonal components and do the x3 integral, ρ dx3 = ρ`3 = σ, where σ ≡ M/`1 `2 is the surface density. (d) What
is the relationship between I1 , I2 and I3 ?
Solution
(a) There are three mirror planes so if I place the origin at the COM and the êi axis parallel to the dimensions `i then
these axes will define mirror planes and so will be the principal axes. 1
(b) The inertia tensor,
=ρ
Z
dx1
Z
Iij =
`i /2
dxi
Z
Z
dm (xk xk δij − xi xj ),
−`1 /2
2π
dφ
Z
dsi s3i
`2 /2
−`2 /2
dx2
Z `3 /2
−`3 /2
dx3 s2i
−`i /2 0
where
s2i = xk xk − x2i
DR
is the square of the distance of dm from the êi axis. I therefore have
Z `2 /2 Z `3 /2 Z `1 /2
2 2
I1 =ρ dx2 dx3 (x2 + x3 ) dx1 1
−`2 /2 −`3 /2 −`1 /2
Z `3 /2 Z `1 /2 Z `2 /2
I2 =ρ dx3 dx1 (x23 + x21 ) dx2 1
−`3 /2 −`1 /2 −`2 /2
Z `1 /2 Z `2 /2 Z `3 /2
I3 =ρ dx1 dx2 (x21 + x22 ) dx3 1
−`1 /2 −`2 /2 −`3 /2
(d) I have I1 + I2 = I3 . 3
This is the perpendicular axis theorem. This theorem applies to a thin plate of any shape, known as a
laminar, meaning “flat.”
X
T
Exercise 3.4 Spinning Rectangular Plate.
Consider a retangular plate (size `1 × `2 , mass M ) which is rotating about its diagonal with angular speed ω. Consider
the principle axis system as the unprimed coordinates and the primed coordinate system having ê01 along the diagonal.
(a) What is the inertia tensor in the unprimed, principle axis system? (b) What is the inertia tensor in the primed
coordinate system? (Hint: Check your answer for the case `1 = `2 .) (c) Calculate the kinetic energy in the unprimed
system. (d) Calculate the kinetic energy in the primed system.
AF
3.6.4 A Handy Theorem
The moment of inertia of a body with inertia tensor
Iij =
Z
In̂ = ni Iij nj .
Proof: Consider the mass element dm at position ~r. If the distance of dm from the n̂ axis is s then the moment of
inertia about n̂ is Z
In̂ = dm s2 .
so
Z
In̂ = dm (xk xk δij − xi xj )ni nj
=ni nj Iij .
This result isn’t that surprising when you consider the expression for the rotational kinetic energy with ω
~ = ωn̂,
↔
T = 21 ω ~ = 12 ωi Iij ωj
~TI ω
= 12 (ωni )Iij (ωnj ) = 21 ω 2 (ni Iij nj )
= 21 ω 2 In̂
Solution
For a rectangular plate in the principle axis system the inertia tensor is
2
`2 0 0
M 0 `21
I◦ = 0
T
3
0 0 `1 + `222
and
ê1 `1 + ê2 `2
n̂ = ê1 cos α + ê2 sin α = p 2
`1 + `22
so
2
M `2 0 0 `1
I 0 =ni Iij nj = `1 `2 0 0 `21 0 `2
3(`21 + `22 )
0 0 `21 + `22 0
=
M
3(`21 + `22 )
2M `21 `22
3(`21 + `22 )
2
`1 `2
T = 21 ω 2 I 0 =
`1
`21 `2 0 `2 =
0
`21 `22
3(`21 + `22 )
M ω2
M
X
(`2 `2 + `21 `22 )
3(`21 + `22 ) 1 2
X
DR
3.7. FURTHER PROBLEMS 61
T
0 Ex /c Ey /c Ez /c
↔ −Ex /c 0 Bz −By
F =
−Ey /c −Bz
0 Bx
−Ez /c By −Bx 0
Exercise 3.5 Two Rods.
Consider a thin rod, linear density η and length `1 (i.e. the other dimensions of the rod are much less than `1 and so can
be neglected.). Place the ê3 axis along the rod and the origin at one end of the rod. (Hint: The moment of inertial of a
thin rod, mass M length `, about its COM is 12 1
M `2 . Also, the mass of the rod is η`1 .) (a) Calculate the inertia tensor,
AF
I(1) , of rod 1 at the origin. (Hint: Is this a principle axis system?) (b) A second thin rod (also linear density η but of
length `2 ) is fixed to the end of the first rod such that the angle between the two rods is α. Calculate the inertia tensor,
I(2) , of the second rod at the origin using the same coordinate system as in (a). Place the two rods in
(i.e. so that ê2 is orthogonal to the plane of the rods). (Hint: A rotation about the ê2 axis is Λ(α) =
result.)
M
0
x
ê1 -ê3 plane
cos α 0 sin α
representation then you can just add them together, I = I(1) + I(2) .) (d) For the case `1 = `2 , what are the principal
↔0
axes and what is the inertia tensor, I , in the principal axis system? (e) Now consider a rod (linear density η, length
2 =
1 0
− sin α 0 cos α
(c) What is the inertia tensor, I, for the two rod system? (Hint: If the inertia tensors of the two rods use the same
`1 + `2 ). Calculate the inertia tensor with the origin at a point a distance `1 from one end. Take ê3 along the rod. (Hint:
By “calculate” I mean that you will only get both bonus marks if you start from scratch rather than using an earlier
a −
A ê
x
i
1
.)
T
AF
DR
T
Chapter 4
Rotational Kinematics
AF
4.1
Resources
Thornton & Marion – Ch. 10
Wikipedia - rotations in 3D
Wikipedia - rigid body dynamics
Introduction
Week 4
In the next chapter we will be examining the rotation of rigid bodies. If we use an inertial reference frame (IRF) to
describe such a system (the fixed frame) then the representation of the inertia tensor in the fixed frame will be given
by
in which Iij0
0
Ik` = λki λ`j Iij 0 T
= λki Iij λj`
is its representation in the rigid body’s non-inertial reference frame (NIRF) (the body frame), and λki ,
is the rotation matrix between the body frame and the fixed frame. If the body is rotating then this matrix will be a
function of time, λki (t). As a consequence, any expression involving the derivative of the inertia tensor of the rigid
body will be a very complicated entity,
DR
Ä ä
I˙k` = λ̇ki λ`j + λki λ̇`j Iij
0
.
and, if I use the Euler angles (see section F.0.1) to define the orientation of the rigid body,1
cos ψ sin ψ 0 1 0 0 cos φ sin φ 0
Λ = sin ψ cos ψ 00 cos θ sin θ sin φ cos φ 0
0 0 1 0 sin θ cos θ 0 0 1
cos ψ cos φ − cos θ sin ψ sin φ − cos ψ sin φ − cos θ cos φ sin ψ sin ψ sin θ
= cos φ sin ψ + cos ψ cos θ sin φ cos ψ cos θ cos φ − sin ψ sin φ − cos ψ sin θ
sin θ sin φ cos φ sin θ cos θ
I will have
d ∂λki ∂λki ∂λki
λ̇ki = λki = φ̇ + θ̇ + ψ̇,
dt ∂φ ∂θ ∂ψ
so there will be a sum of six products of three matrices!
As an alternative we can reformulate kinematics in a NIRF so that the inertia tensor is not a function of time. This
has the added advantage that we can choose the orientation of the NIRF to coincide with the principal axes of the rigid
body, thereby simplifying things even further. I will refer to this NIRF as the body coordinates and the IRF as the
fixed coordinates.
1 I introduced these in §F.0.1 (p. 184) to describe the relative orientation of two coordinate systems.
63
64 CHAPTER 4. ROTATIONAL KINEMATICS
T
Figure 4.1: There are twelve possible ways to specify Euler angles. This one, labeled z-x-z, is most common in physics. It
AF consists of a rotation of φ about the ê03 -axis followed by a rotation of θ about the ê1 -axis, followed by a rotation of ψ about
the ê3 -axis.
fixed frame
ê01
ê03
0
O0 ê2
~rO0
body frame
ê1
ê3
O ê2
Figure 4.2: The position of the body coordinates origin, O, relative to the fixed coordinates is ~rO0 , and the acceleration of the
body coordinate system is ~r¨O0 .
and an observer in the NIRF will experience an additional force, −m~r¨O0 , which we refer as an inertial force – the
apparent force an object in the non-inertial reference frame will experience due to the acceleration of the reference
frame.
Example 4.1 Ascending Elevator.
Alice is in an elevator ascending with an acceleration ~a. She holds a spring scale attached to a bob, mass m. What is
the magnitude and direction of the force measured by the scale?
4.3. ROTATING REFERENCE FRAMES 65
Solution
I have ~r¨ 0 = ~r¨ + ~r¨O0 so the scale measures the apparent force, F~ , on the stationary bob (i.e. ~r¨ = 0) in the NIRF,
m~r¨ 0 = F~ + m~g
T
m~r¨ = 0 = F~ + m~g − (m~r¨ 0 ) O
This is the force on Alice’s hand in which holds the scale stationary. X
Equivalence Principle
Einstein’s equivalence principle – that gravitational and inertial mass are the same thing – is the basis of his general
theory of relativity which replaces the gravitational force with the curvature of spacetime. In other words, a gravita-
tional force is actually an inertial force (similar to the centrifugal and Coriolis inertial forces) and that we observe it
AF
because we are not in an inertial reference frame, i.e. we in the presence of a mass (the Earth) which itself curves
spacetime.
4.3
4.3.1
Rotating Reference Frames
Rotation Vectors
In terms of the Euler angles the rotation matrix, Λ, is complicated but it is still an orthogonal rotation matrix which
rotates the body coordinate system, êi , into the fixed coordinates system, ê0i . As we saw in Exercise 12.7 in section F.0.1
on Euler angles, a rotation matrix has three eigenvalues, {1, e±iα }, and the eigenvector corresponding to the real
eigenvalue, α̂,
↔
Λ · α̂ = α̂ or λij αj = αi
(with αi αi = 1) is unchanged by the rotation. This unit vector defines the rotation axis. The complex eigenvalues,
e±iα , give the angle of rotation; it is convenient to let the magnitude of the rotation vector equal to the angle of rotation
with the sense of the rotation given by the right hand rule (RHR),
α
~ = αα̂.
rotation about the symmetry axis – the inertia tensor is a constant with respect to rotations about this axis – so that the Lagrangian will be invariant
∂L
with respect to ψ, ∂ψ = 0.
3 show this
4 Actually, as axial vectors; see section F.0.1.
66 CHAPTER 4. ROTATIONAL KINEMATICS
α̂
dα
T
P d⃗
rP′
r⃗P , r⃗P (t)
′
r⃗P′ (t + dt)
O, O′
´¹¹ ¹ ¹ ¹ ¹ ¹ ¹ ¹ ¶
AF Figure 4.3: Consider a rigid body rotating about the axis α̂ with the origins of the NIRF and the IRF coincident and on the axis
of rotation (so that ~rO0 = 0 and ~rP = ~rP0 (t)). The position vector of the point P, ~rP , on the rigid body sweeps out a cone in the
rP0 (t). The rotation vector, d~
IRF, ~
(i.e. since α
4.3.3
α = α̂ dα, tells us how the vector ~rP0 (t) evolves to ~rP0 (t + dt) in time dt: d~rP0 = d~
The velocity of the point relative to the fixed system will then be5
where ω~ ≡α
~ 0 = Λ·ω
ω
~r˙P0 = α
~ =ω
~ is an eigenvector of the rotation matrix so are d~
General Vector
The expression, ~r˙P0 = ω
~,
~˙ × ~r = ω
~˙ ).
α and α
↔
~ × ~r
~˙ is the angular velocity. Note in particular that this vector, because it is parallel to the rotation vector, is
~ × ~r, is true, in particular, for the basis vectors in the body frame,
α × ~rP .
ê˙ i = ω
~ × êi ,
so that with the two origins coincident and on the axis of rotation, the position vector of a point, P, is,
DR
~rP0 = ~rP
ri0 ê0i = ri êi
and the time derivatives are
d 0 0 d
(ri êi ) = (vi êi )
dt dt
ṙi0 ê0i = v̇i êi + vi ê˙ i
~ × êi
= v̇i êi + vi ω
~ × ~r
= ~v + ω
~˙ 0 , given by
~ = Q0 ê0 = Qi êi , has an inertial time derivative, Q
A general vector, Q i i
d 0 0 d
(Q ê ) = (Qi êi )
dt i i dt
Q̇0i ê0i = Q̇i êi + Qi ê˙ i = Q̇i êi + Qi ω
~ × êi = (Q̇i êi ) + ω
~ × (Qi êi )
~˙ 0 = Q
⇒ Q ~˙ + ω ~
~ ×Q
5 See Appendix E for a more rigorous derivation of this expression.
4.3. ROTATING REFERENCE FRAMES 67
where Q~˙ ≡ Q̇i êi is the body time derivative, how Q~ changes from the perspective of the body frame.
This is an important result; it relates the rate of change of the representation of a vector in the two coordinate
systems. We can write this relationship as the operators
d0 d
T
= ~×
+ω
dt dt
in order to describe the relationship between time derivatives of vectors represented in the two coordinate systems.
The second term takes care of how the body basis vectors change.
d0 d
AF
(where ω
~r 0 − ~rO0 =~r
~˙ 0 = ω
dt
=(~r¨ + ω
=
dt
~ × ~r)
~r¨ 0 − ~r¨ 0 =(~r¨ + ω
~˙ × ~r + 2~
~˙ + ω~ ×ω
~ ×,
+ω
~ defines the axis and angular speed of the body frame’s rotation) to find expressions relating the velocity and
acceleration of a particle in the two reference frames. I have
~ × ~r˙ ) + ω
~ × ~r˙ ) + ω
~˙ × ~r + ω
~˙ × ~r + ω
ω × ~r˙ + ω
~ × ~r˙
~ × (~r˙ + ω
~ × (~
~ × ~r)
ω × ~r).
is Newton’s second law for a rotating coordinate system. The two extra terms on the right are inertial forces, the first
is called the Coriolis force and the second is called the centrifugal force.6
Example 4.2 Ferris Wheel.
Alice is now a passenger on a Ferris wheel which rotates at a constant angular frequency ω. She still has her spring
scale attached to the bob, mass m. What is the magnitude and direction of the force measured by the scale? Assume
that Alice measures t = 0 at the bottom of the Ferris wheel.
Solution
With the origin at the hub of the Ferris wheel The bob’s position in polar coordinates is ~r = rr̂ with r the distance
from the hub. The velocity and acceleration are then7
and
˙
~r¨ = r̈r̂ + ṙr̂˙ + ṙθ̇θ̂ + rθ̈θ̂ + rθ̇θ̂
= r̈r̂ + ṙθ̇θ̂ + ṙθ̇θ̂ + rθ̈θ̂ − rθ̇2 r̂
T
Ä ä Ä ä
= r̈ − rθ̇2 r̂ + 2ṙθ̇ + rθ̈ θ̂
The apparent force, F~ , on the stationary bob (i.e. ṙ = 0) in the NIRF, therefore includes the centrifugal force,
Fcf = mω 2~r, but not the Coriolis force, F~cf = −2mṙω θ̂,
~
AF
4.3.5 Effective Potentials
m~r¨ = 0 = F~ + m~g + F~cf
F~ = −m(−g ŷ) − mω 2~r
= mg ŷ − mω 2 rr̂,
r̂ = x̂ cos φ + ŷ sin φ.
I find φ using the BC, φ(t) = ωt + φ◦ so that φ(0) = −π/2 = φ◦ , and the scale measures the force
This is the force on Alice’s hand in which holds the scale stationary.
φ = ωt − π/2.
It is often the case that we can express inertial forces in terms of potentials. In this way we can study a complicated
X
and in Example 4.2 concerning a Ferris wheel, I have for an object moving under the force F~ = −∇V
~ , N 2 L in the
NIRF is
m~r¨ = F~ − mg ŷ + mω 2 rr̂ = −∇
~ V (~r ) + mgy − 1 m(ωr)2 .
2
The interpretation of the effective potential can often simplify the interpretation of a complicated system – the various
minima, maxima, saddle points, etc assist in identifying equilibria, small oscillations, and other motions of interest.
Exercise 4.1 Carousel.
Consider the case of a hockey puck (mass m) on an icy merry-go-round rotating with constant angular velocity ω.
Assume that the merry-go-round surface is horizontal and frictionless. Use Cartesian coordinates fixed to the merry-
go-round (i.e. use the NIRF body coordinates) with the origin at the centre of the merry-go-round and ẑ vertical (so
~ = ωẑ and ~r = xx̂ + y ŷ). (a) What is the centrifugal force, F~cf , for this system? (b) What is the Coriolis
that ω
force, F~co , for this system? (c) Use the appropriate version of Newton’s second law in the body coordinates to find
the equation of motion for the hockey puck . (d) Use a computer to calculate and plot the apparent motion of the puck
in the body coordinate system using the initial conditions ~r◦ = ax̂ and ṙ◦ = −v◦ ŷ for several values of v◦ near aω.
(e) Use a computer to calculate and plot the√apparent motion of the puck in the body coordinate system using the initial
conditions ~r◦ = ax̂ and ṙ◦ = −v◦ (x̂ + ŷ)/ 2 for several values of v◦ near aω. (f) Solve this problem analytically for
the two initial conditions given in parts (d) and (e). (Hint: Introduce ũ ≡ x + iy. You may leave your answers as a
complex function, ũ(t).)
4.4. GENERAL MOTION OF A RIGID BODY 69
P
~rP0 ê2
T
~rP
ê03 ê1
ê3 O
O0 ~rO0
ê02
ê01
Figure 4.4: The un-primed coordinate system, with origin at O, is attached to the rigid body; these are the body coordinates.
The primed coordinate system, with origin at O0 , is an inertial coordinate system; these are the fixed coordinates. The
position of the point P is either ~rP0 or ~r 0 and these are related by ~rP0 = ~rO0 + ~rP .
AF
4.4 General Motion of a Rigid Body
I will now examine the general motion of a rigid body that is both accelerating and rotating. But first I must show that
the angular velocity vector is common for all points in the rigid body.
~r˙Q0 − ~r˙O0 = ω
~ O × ~rQ .
I could, however, use the point Q as my body origin, so that the velocity of the point P would be
~r˙P0 − ~r˙Q0 = ω
~ Q × ~rQP ,
DR
where ω
~ Q is the angular velocity for rotations when Q is the origin.
But in the body frame I have ~rP = ~rQ + ~rQP so
~r˙P0 − (~r˙O0 + ω
~ O × ~rQ ) = ω
~ Q × (~rP − ~rQ )
~r˙ − ~r˙ = (~
P
0 0
O ωO − ω ~ Q ) × ~rQ + ω
~ O × ~rP .
But, from above I have
~r˙P0 − ~r˙O0 = ω
~ O × ~rP .
so I see that I must have ω ~ Q and there is only one angular velocity of the rigid body. This is an important result.
~O = ω
T
A
Q
AF ~ · ~r˙O0 = 0
Case 1: ω
If the velocity of the body origin, ~r˙O0 , and the angular velocity, ω
~r˙ 0 = ~r˙ 0 + ω × ~r 0 , is also orthogonal to the angular velocity:
A
Figure 4.5: Consider a body at two instants in time separated by ∆t. We can use this construction to identify a new origin, Q,
for which the motion consists of a pure rotation. The instantaneous axis of rotation passes through this origin and is orthogonal
to the figure.
The velocity of the body origin, ~r˙O0 , and the angular velocity, ω
~ , are either orthogonal or they are not.
~ · ~r˙Q0 =
ω ~r˙O0 + ω
~ ·
ω ω × ~rQ0 ) = 0.
~ · (~
DR
I can then write
~r˙Q0 = ~r˙O0 + ω
~ × ~rQ0
1 Ä ˙
~ × (~r˙O0 × ω
ä
= 2 ( ~ · ~rO0 )~
ω ω+ω ~ × ~rQ0
~) + ω
ω Å
1 ˙0
ã
0
=ω ~× ~r × ω
~ + ~
rQ ,
ω2 O
and then choose as a new origin for the body frame (see Fig. 4.5)
1 ˙0 1
~rQ0 = − ~r × ω
~ = 2ω~ × ~r˙O0
ω2 O ω
so that ~r˙Q0 = ω
~ × ~rQ0 = 0 and the motion of the body consists of a pure rotation (i.e. the motion has no translational
component) about the axis parallel to ω ~ which passes through the point ~rQ0 . This point with ~r˙Q0 = 0 is known as the
ICOR and the axis parallel to ω~ passing through the ICOR is known as the instantaneous axis of rotation (IAOR).
Note that all points on the IAOR have zero velocity. This can be used to identify the IAOR prior to commencing the
solution to a problem.
4.4. GENERAL MOTION OF A RIGID BODY 71
0 0
0
~r˙ P = ω
~ × ~rP ~ × ~rP
ω ~r˙ P = ~r˙ C + ω
~ × ~rP
P P
~rP0 ~rP0 0
ω
~ ~r˙ C ω
~
~rQP 0 ~rCP 0
~r˙ C ~r˙ C
T
C C
~rQC ~rC0
O0 O0 ~rCQ
~rQ0
Q Q
AF
Figure 4.6: A rolling rigid body, radius a, rolling on a surface has ~r˙Q0 = 0, where Q is the ICOR; this is the rolling constraint.
On the left the body origin is at Q, the ICOR, and on the right the body origin is at C, the centre of the disk. Rolling is a
rotation about the ICOR, ~r˙P0 = ω
rotation about the centre, ω
~ · ~r˙O0 6= 0
Case 2: ω
If, instead, the velocity of the body origin, ~r˙O0 , and the angular velocity, ω
origin, ~rQ0 = ω̂ × ~rO0 , so that
~r˙Q0 = ~r˙O0 + ω
~ × ~rQ0
= ((ω̂ · ~rO0 )ω̂ + (
= (ω̂ · ~rO0 )ω̂
ω̂ ×
~ , are not orthogonal then I can choose a new
rO0(×(ω̂)) + (
((~
( (
ω̂ ×
((ω̂
(( ~rO0 )
×(
(
~ – the rotation and translation are along the same axis, known as a screw axis (examples are
a bullet from a rifled gun or a “spiral” throw in American football).
In either case, identifying the ICOR or the screw axis will simplify the description of the motion.
4.4.2 Rolling
DR
The term “rolling” has a specific meaning in the context of rigid body motion. It means that there is a point of contact
between two bodies at which the relative motion of the two bodies is zero. If this were not the case the two bodies
would be “sliding” instead of rolling. Usually we speak of one body rolling upon the other, which is usually an inertial
system so that we can attach the body frame to the rolling body and the inertial frame to the inertial body.
The point of contact is, of course, the ICOR and the IAOR is tangent to the contacting surfaces at this point. The
angular velocity vector of the rolling body, ω ~ , is along the ICOR and the velocity of any point, ~rP0 , on the rolling body
is, as usual,
~r˙P0 = ~r˙O0 + ω
~ × ~rP
where ~r˙O is the velocity of the body origin.
0
If we label the ICOR as point Q, then the rolling constraint, ~r˙Q0 = 0, results in ~r˙Q0 = 0 = ~r˙O0 + ω ~ × ~rQ . or
~r˙O = ~rQ × ω
0
~ , in which ~rQ is the vector from the body origin to the ICOR.
There are two special points on a rolling object: The ICOR and the centre of the object. It is usually convenient to
choose one of these as the body origin.
T
Case 2: Body Origin at the Centre
If, instead, I choose the body origin at the centre of the rolling body then the velocity of any point P in the rigid body
is (see Figure 4.6, right)
~r˙P0 = ~r˙C0 + ω
~ × ~rCP = ω
~ × (~rCP − ~rCQ ),
and, in particular, for the ICOR, I have
~r˙Q0 = ~r˙C0 + ω
~ × ~rCQ
AF
exactly as above (i.e. ~rCQ = −~rQC ).
It is interesting to examine the velocity of any point P in the rigid body is (see Figure 4.6, right)
~r˙P0 = ~r˙C0 + ω
~ × ~rCP
which is the velocity of the body origin at the centre, plus the rotational velocity relative to the centre. One can think
of rolling as a superposition of a linear translation at velocity ~r˙C0 and a rotational motion with angular velocity, ω
about the centre, along with the rolling constraint ~r˙C0 = ~rCP × ω
Exercise 4.2 Wheel Travelling in a Circle.
~.
Consider a circular wheel (radius a) which rolls without slipping, at a constant angular speed ω = |~
path (radius b) on a horizontal plane with the plane of the wheel vertical.
~,
ω |, in a circular
For this problem it is convenient to think of the motion of the wheel as consisting of an “orbital” part about the
~ and a “spin” part about the axis of the wheel (with angular velocity ~n) so that
circular path (with angular velocity Ω)
the total angular velocity is ω
~ =Ω ~ + ~n. Consequently, choose a rotating reference frame with the origin, O, at the
centre of the circular path, with ê3 vertical, and with the ê1 axis passing through the centre of the wheel. This frame
DR
rotates with angular velocity Ω~ = Ωê3 .
Choose the inertial frame to have the same origin and for the two coordinate systems to be coincident at t = 0. At
other times we then have
ê1 cos(Ωt) sin(Ωt) 0 ê01
ê2 = − sin(Ωt) cos(Ωt) 0ê02 .
ê3 0 0 1 ê03
(a) Calculate the orbital velocity, ~r˙C0 (i.e. the velocity of the centre of the wheel). (b) Calculate the spin velocity, ~r˙B ,
of the bottom of the wheel in the orbiting reference frame. (c) What is the velocity, ~r˙B0 , of the bottom of the wheel?
(d) Apply the rolling constraint, ~r˙B0 = 0, to find a relationship between Ω and ω. (e) What is the angular velocity,
ω
~ =Ω ~ + ~n, of the wheel? (f) What is the acceleration, ~a, of the ICoR?
Ld ? Again, use x as your generalized coordinate. Use this system for the remainder of this question. (e) What is the
conjugate momentum, p, for the bead? (f) What is the equation of motion, ẍ = f (ẋ, x), for the bead? (g) What is the
Hamiltonian, H(x, p), for the bead? (h) What is the energy, E = T + V , in terms of x and p? (i) What is E − H?
(j) Give a brief physical interpretation as to what, if anything, is conserved in light of your answers to parts (g) through
(h). (k) What are Hamiltonian’s equations, ṗ = − ∂H ∂x & ẋ = ∂p , for the bead? (l) Equilibrium points occur when
∂H
T
ṗ = ẋ = 0. Identify any equilibrium points and state under what conditions they will be stable? (Hint: Examine ṗ near
the equilibrium – if the force is a restoring force the equilibrium is stable.) (m) What are the equations of motion for
parts (c). Give a brief physical interpretation of each equation. (n) Redo part (c) in cylindrical coordinates, (s, φ, z),
and then use the constraints η1 = z − 12 αx2 = 0 and η2 = φ − ωt − φo to find the constraint forces (φo is a constant).
Give a physical interpretation for each force.
AF
DR
74 CHAPTER 4. ROTATIONAL KINEMATICS
T
AF
DR
T
Chapter 5
Midterm Review
AF
5.1 Coupled Oscillators
Example 5.1 Three Blocks.
Week 5
Consider the three blocks (each of mass m) which rest on a frictionless surface and are free to slide in the ê1 direction
only. The two outside blocks are attached to vertical walls by identical springs (spring constant k) and to the centre
block by two weak springs (each of spring constant εk with ε a small, dimensionless constant). Newton’s second law
tells us
where each xi is measured from the block’s equilibrium position. This can be written in matrix form as
x 1 + ε −ε 0 x1
d2 1
m1 2 x2 = −k −ε 2ε −ε x2
dt
x3 0 −ε 1 + ε x3
DR
h 1+ε−ε 0
i
¨ = −K~x where K ≡ k
or M~x −ε −2ε −ε is the “spring matrix.” This is a set of three coupled, second order ODEs
0 −ε 1+ε
which can be solved with the ansatz ~x = ~ae iωt
. This gives
−Mω 2 ~x = −K~x ⇒ K − mω 2 1 ~x = 0
where λ ≡ mω 2 /k. This, of course, is an eigenvalue problem. (a) What are the eigenvalues, λi ? (b) What are
the orthonormal eigenvectors,
v̂i ? (c) We can now create the matrix S, whose column vectors are the orthonormal
eigenvectors, S ≡ v̂1 v̂2 v̂3 . We can then multiply the equations of motion, m~x ¨ = −K~x, by the matrix S T and
insert the identity SS = 1 between K and ~x,
T
d2 T
m S ~x = −ST K SST ~x = − ST KS ST ~x
dt2
d2 0
⇒m (~x ) = −K0 ~x 0
dt2
75
76 CHAPTER 5. MIDTERM REVIEW
T
Solution
(a) The eigenvalues are
λi ∈ {??, ??, ??}. 2
v̂1 = ??
v̂2 = ??
v̂3 = ??
AF
1+ε−λ −ε
0
−ε −4ε −ε
0 −ε 1−ε
−ε −2ε−λ −ε = (1 + ε − λ)2 (2ε − λ)
−ε 1+ε−λ
h 1−ε−ε 0
iî α1 ó
α2
α3
=0⇒
3
εα2 =
4α2 =
εα2 =
⇒
(1 − ε)α1
−ε(α1 + α3 ) ⇒
(1 − ε)α3
λ ∈ {1 + ε, 2ε, 1 + ε}
~v2 =
î ε ó
1−ε .
ε
The other two eigenvectors are degenerate (with eigenvalue 1 + ε) and I can choose them as I wish so long as
and
1 h 1 i
v̂1 = √ 0 1
2 −1
DR
1 î ε ó
v̂2 = √ 1−ε 1
3ε − 2ε + 1 ε
2
1 h 1−ε i
v̂3 = p −2ε 1
2(3ε2 − 2ε + 1) 1−ε
(c) The similarity transformation, K 0 ≡ S T KS, will diagonalize the spring matrix and the diagonal elements will be
the eigenvalues, ï ò
λ1 0 0
K0 = 0 λ2 0 k 2
0 0 λ3
and the X
x0i = T
Sij xj = v̂i · ~x. 2
j
(d) In the primed system the spring matrix is diagonal and the equations of motion are decoupled into three indepen-
dent equations, ẍ0i = −ωi2 x0i , with solutions x0i (t) = a0i cos(ωi t + δi0 ), in which the a0i and δi0 are constants given
by the boundary conditions and ωi2 = λi k/M are the frequencies. +2
The components of ~x 0 , x0i = v̂i · ~x, are known as the “normal modes,” which are independent periodic oscilla-
tions with frequency ωi . The general solution for this problem will be a linear combination of these modes.
X
5.2. INERTIA TENSOR 77
T
centre of mass? (b) What is the inertia tensor, I, at the origin? (c) Now consider a primed coordinate system which is
rotated an angle α = tan−1 (a/h) about the ê2 axis (i.e. ê01 is along the edge of the cone). What is the inertia tensor
for this primed coordinate system? h cos α 0 sin α i
(Hint: A rotation about the ê2 axis is Λ(α) = 0 1 0 .) (d) If the disk rotates about the ê03 axis with angular
− sin α 0 cos α
↔
speed ω, what will be its kinetic energy, T = 12 ω ~ = 12 ωi Iij ωj ? (Hint: You may use your inertia tensor from
~T · I · ω
either (b) or (c), or some other method if you wish.)
Solution
AF
(a) The moment of inertia about the disk axis, ê3 , is easy to calculate with polar coordinates. With σ = M/πa2 for
the surface density so that dm = σ da, and with da = dφ s ds I have
= 12 πσa4 = 12 π
1 , so
Å
M
:s
πa2
1 + x2 x2 ) = σ
0
dφ
Z a
a4 = 21 M a2 1
I◦ =
1
a
s ds s2 = σ(2π) 41 s4 0
M a2 î 1 0 0 ó
4
î1 0 0ó
010 .
002
1
î1 0 0ó î0 0 0ó
~c · ~c 1 − ~c ⊗ ~c = a2 010 − a2 000 = a2 010 1
001 000 001
and î0 0 0ó î1 0 0ó
I = I ◦ + M a2 010 = 41 M a2 050 2
001 006
T
= 21 ω 2 (I11 sin2 α + I33 cos2 α)
tensors of the two rods use the same representation then you can just add them together, I = I
↔0
↔
the case `1 = `2 , what are the principal axes and what is the inertia tensor, I , in the principal axis system? (e) Now
consider a rod (linear density η, length `1 + `2 ). Calculate the inertia tensor with the origin at a point a distance `1
from one end. Take ê3 along the rod. (Hint: By “calculate” I mean that you will only get both bonus marks if you start
from scratch rather than using an earlier result.)
Solution
↔(1) ↔(2)
+I .) (d) For
(a) For a “thin” rod I33 = 0 1 and rotation symmetry about the axis of the rod tells me that I11 = I22 . The parallel
axis theorem gives me
DR
◦
I11 = I11 + (η`1 )(`1 /2)2 = 1 3
12 η`1 + 14 η`31 ⇒ I11 = I22 = 31 η`31 . 1
Because of the rotation symmetry about the ê3 axis this is a principal axis system and
so
↔(1) î1 0 0ó
⇒ I = 31 η`31 010 .
000
The rod can be considered a laminar in either the ê1 or ê2 directions so that I22 +I33 = I11 and I11 +I33 = I22 .
5.2. INERTIA TENSOR 79
(b) The principal axis inertia tensor for the second rod will be
the same as in (a) with `1 replaced with `2 . I need to rotate
the basis by −α about the ê2 axis:
↔(2) cos α 0 sin α 1 0 0 cos α 0 − sin α
T
I = 31 η`32 0 1 0 0 1 0 0 1 0
− sin α 0 cos α 0 0 0 sin α 0 cos α
cos2 α 0 − sin α cos α
= 13 η`32 0 1 0 . 2
− sin α cos α 0 sin2 α
AFwhere dm = η d` =
↔
I = 1
η dx1
sin α ,
3 100
î
3 η`1 0 0 0
010
ó
+ 1
I11 =
3
3 η`2
↔
0
0 1
I = 31 η`31
dm (
0
sin α cos α 0 sin2 α
x20
2 +
* x2 x1 cot
1 α3
(c) I have
The rods can be considered a laminar in the ê2 direction so that I11 + I33 = I22 .
1
3η
ò
.
0
`31 +`32
0
`32 sin α cos α
0
`32 sin2 α
ô
2
I could find the eigenvalues and eigenvectors but there is a reflection symmetry in the plane bisecting the angle
between the two rods so the principal axes must be at
DR
ê03 = ê1 sin(α/2) + ê3 cos(α/2)
ê01 = ê1 cos(α/2) − ê3 sin(α/2). 1
⇒ λ ∈ {2, 1 ± cos α} 1
↔(2) ↔(1)
Alternatively, I can use my result from (b) with α → α/2 for I and with α → −α/2 for I , and then
80 CHAPTER 5. MIDTERM REVIEW
T
cos2 (α/2) 0 sin(α/2) cos(α/2)
+ 31 η` 3
0 1 0
2
sin(α/2) cos(α/2) 0 sin (α/2)
2
cos (α/2) 0 0
= 23 η`3 0 1 0 .
2
0 0 sin (α/2)
The rods can be considered a laminar in the ê2 direction so that I11 + I33 = I22 .
(e) I have a rod of length ` ≡ `1 + `2 and linear density η. Its moments of inertial about an its centre of mass (at 21 `)
1
= 12
1
= 12
↔(◦)
I
1 0 0
0 0 0
1 0 0
=
1
12 η`
3 100
î
010
000
η`3 0 1 0 + 14 η` ∆`2 0 1
η 0 1 0` `2 + 3∆`2 +1
ó
so I need to move the axis by a displacement of ~c = 21 |`2 − `1 |ê3 from the CoM, so:
î1 0 0ó
~c · ~c 1 − ~c ⊗ ~c = 14 (`2 − `1 )2 0 1 0
↔(◦)
+ η`(~c · ~c 1 − ~c ⊗ ~c)
1 0
0 0
000
0
0
0 0 0
↔ 1 0 0
⇒ I = 13 η(`1 + `2 )0 1 0, +1
3 3
DR
0 0 0
X
Example 5.4 Thin Plates.
ê2
Consider a thin plate, mass
√ M , in the shape of a right trian- B
x2 = b(1 − x1 /a)
gle (sides a, b and c = a2 + b2 ). Place the origin at the
right angle with the ê1 axis along a side of length a, the ê2 b
M
axis along a side of length b and the ê3 axis perpendicular A ê
1
to the plate. O a
(a) Calculate the position, ~c, of the centre of mass of the plate. (Hint: Recall that the area of this triangle is 12 ab.)
(Hint 2: For parts (a) and (b) it is critical to have the limits on your integral correct. Draw a picture.) (b) Calculate the
5.2. INERTIA TENSOR 81
inertia tensor at the origin. (Hint: Is this a principle axis system?) (c) What are the principle axes about the origin?
Solution
R R Ra R b(1−x1 /a)
T
(a) I have M~c = dm ~r so M (c1 ê1 + c2 ê2 ) = dm (ê1 x1 + ê2 x2 ) and M c1 = σ 0 dx1 0 dx2 x1 =
Ra b(1−x1 /a)
σ 0 dx1 x1 [x2 ]0 = 3 M a. Similarly, M c2 = 3 M b so that ~c = 3 (aê1 + bê2 ) 2 .
1 1 1
Ra R b(1−x1 /a)
(b) I11 = σ 0
dx1 0
dx2 x21 ⇒ I11 = 16 M a2 1 , similarly, I22 = 16 M b2 1 , and by the perpendicular
2
0
0
2
2a
−ab 2b2
2
0
2(a + b ) − λ
−ab
=
0
0
0
2(a2 + b2 )
.
Consider a right circular cone (mass M , height h and radius a at base). Place the origin, O, at the apex and the ê3
axis along the axis of the cone as shown to the right. (a) Calculate the position, ~c, of the centre of mass of the cone.
(Hint: Recall that the volume of a cone is 31 πha2 .) (b) Calculate the inertia tensor at the origin. (Hint: What are
the principal
X
h cosaxes?) (c)i What is the inertia tensor along an edge of the cone? (Hint: A rotation about the ê2 axis is
α 0 sin α
Λ(α) = 0 1 0 .) (d) For what value of ha will the cone have dynamical spherical symmetry? (Hint: A guess
− sin α 0 cos α
will cost you nothing.)
DR
Solution
(a) The CoM will clearly be along the axis so, with tan α = a/h, I have
Z Z 2π Z h Z z tan α
M~c = M cê3 = dm z = ρê3 dφ dz z s ds
0 0 0
Z h Z h
1 2
z tan α
= ρê3 (2π) dz z 2s 0
= πρê3 tan2 α dz z 3
0 0
2
1
4 h
1
3 πha ρ cê3 = πρê3 (a/h)2 4z 0
= 14 πρê3 (a/h)2 h4 = 41 πρê3 a2 h2 ⇒ c = 43 h 2
(b) The principal axes are the as given in the figure so the inertia tensor is diagonal. The cone has a continuous
rotational symmetry about its axis so I11 = I22 and I only need to calculate I11 and I33 .
Z Z 2π Z h Z z tan α
I33 = dm s2 = ρ dφ dz s ds s2
0 0 0
Z h Z h
1 z tan α
= 2πρ dz 4 s4 0 = 12 πρ tan4 α dz z 4
0 0
h
= 2 π(3M/πha2 )(a/h)4 51 z 5 0
1
⇒ I33 = 3
10 M a
2
2
82 CHAPTER 5. MIDTERM REVIEW
Z Z 2π Z h Z z tan α
I11 = dm (x22 + x23 ) =ρ dφ dz s ds (s2 cos2 φ + z 2 )
0 0 0
Z Z Z Z Z2πh Z
2π
* πh
2
z tan α
2
2π
2
z tan α
=ρ dφ
cos φ
dz s ds s + ρ dφ
> dz z s ds
0 0 0 0 0 0
Z Z
T
h 1 z tan α h 1
4 2 z tan α
=πρ dz 4s 0
+ 2πρ dz z 2 2s 0
0 0
Z h Z h
π
= ρ tan4 α dz z 4 + πρ tan2 α dz z 4
4
ãÅ0 4 a 2 ã
0
π 3M a
Å
1 5 h
= 2
+4 5 z 0 ⇒ I11 = I22 = 203
M (a2 + 4h2 ) 2
4 πha h h
2
1 0 0 1 0 0 a + 4h2 0 0
3M 2 3M 3M
I+ h 0 1 0 = a2 0 1 0 = 0 a2 + 4h2 0 1
AFI have
5
=
0 0 0
001
I0 = ΛT IΛ = 0
2
0
0
î1 0 0ó
3M
20
2
1
cos α + 6 sin α 0 0
2
0 0 2
5 0
0 6
cos α 0 − sin α I11
0 0
sin α 0 cos α
0
20
0 0 2a2
0
I22
0
0
cos α
0 0
I33 − sin α
(d) I will first need the intertia tensor, I◦ , about the centre of mass. With ~c = 34 hê3 I will use the parallel axis theorem
î0 0 0ó
001
a + 4h2
î1 0 0ó
with ck ck δij − ci cj = c23 0 1 0 − c23 0 0 0 = c23 0 1 0 = 16
000
0
9 2 100
0
î
000
9M
ó
h 0 1 0 so, with Iij = Iij
2
h 0 0
◦
+ M (~c · ~c 1 − ~c ⊗ ~c),
0 sin α
1 0
0 cos α
I◦ = 0 a2 + 4h2 0 − 0 h2 0
20 2 16
0 0 2a 0 0 0
2 2
12a + 3h 0 0
DR
M
= 0 12a2 + 3h2 0 +1
80
0 0 24a2
h
12a2 + 3h2 = 24a2 ⇒ =2 +1
a
ê3 , ê03
A
n̂
α
T
C
ê2
O, O0
Ωt Q ê1
ê01
ê02
AF
Figure 5.1: A coin rolls on a plane such that its centre remains fixed and the point of contact with the horizontal plane traces
out a circle. The angle between ê01 and ê1 is Ωt.
The disk will have a spin angular velocity ~n such that its total angular velocity is ω ~ =Ω ~ + ~n.
(a) What is the relationship between the primed and unprimed basis vectors? (b) What are the point of contact,
~rQ , and the spin axis, n̂, in terms of both the primed and unprimed basis vectors? (c) Calculate ~r˙Q0 , the velocity of the
point of contact, Q, in the primed, inertial system? Express your answer in terms of both the primed and unprimed
basis vectors. (Hint: How are the êi rotating relative to the ê0i ?) (d) Calculate ~r˙CQ , the velocity of the lowest point on
the disk relative to the rotating coordinate basis. Express your answer in terms of the unprimed basis vectors. (Hint:
~rCQ = ~rQ − ~rC is a vector from the spin axis to the point Q.) (e) Use the rolling constraint to find the relationship
between n and Ω where ~n = nn̂ and Ω ~ = Ωê3 are the spin and precession angular velocities, respectively. (f) What is
the instantaneous axis of rotation, ω̂, where ω
(g) What is ω
~ =Ω~ + ~n is the total angular velocity? Express your answer in terms of
both the primed and unprimed basis vectors. (Hint: You should be able to answer this question without any equations.)
~ · ~n? Sketch the instantaneous axis of rotation, ω̂, on the figure. (h) What is the velocity of the highest
point on the disk, ~r˙A0 ? (i) (BONUS) Calculate the relationship between the magnitude of the spin angular velocity, n,
and that of the orbital angular velocity, Ω, using a non-inertial, double-primed coordinate basis fixed to the disk with
its origin at the centre of the disk with the ê003 perpendicular to the disk (i.e. the non-inertial basis is rotating with
angular velocity ω ~ =Ω ~ + ~n relative to the primed, inertial basis and the spin angular momentum is ~n = nê00 ).
3
DR
Solution
ê1 = ê01 cos φ + ê02 sin φ & ê2 = −ê01 sin φ + ê02 cos φ, φ = Ωt
(b)
~rQ0 = ê1 a cos α = a cos α(ê01 cos φ + ê02 sin φ) 2
n̂ = ê3 cos α + ê1 sin α = ê03 cos α + sin α(ê01 cos φ + ê02 sin φ) 2
with φ = Ωt.
84 CHAPTER 5. MIDTERM REVIEW
~r˙Q0 = ~r˙Q + Ω
~ × ~rQ = (Ωê3 ) × (ê1 a cos α)
T
the point B is travelling in a circle, radius a cos α, with frequency Ω.
(d) I have ~rCQ = a(ê1 cos α − ê3 sin α) so
~r˙CQ = ~n × ~rCQ = n(ê3 cos α + ê1 sin α) × a(ê1 cos α − ê3 sin α) = naê2 2
(e)
0 = ~r˙CQ + ~r˙Q0 = naê2 + ê2 aΩ cos α ⇒ n = −Ω cos α 2
AF
(f)
This vector is a constant in the rotating frame (and is why we chose this frame for the rotating basis). This vector
has length Ω sin α so
ω̂ = sin αê03 − cos α(ê01 cos φ + ê02 sin φ) = sin αê03 − cos αê1 +1
Alternatively: Since the two points Q and C are stationary the rotation axis must pass through these points,
parallel to QC,
ω̂ = sin αê03 − cos α(ê01 cos φ + ê02 sin φ) = sin αê03 − cos αê1 2
DR
(g) These are orthogonal, ω
~ · ~n = Ω sin α(ê3 sin α − ê1 cos α) · (−Ω cos α)(ê3 cos α + ê1 sin α) = 0. 2
~ , so ~r˙A0 = 0.
(h) This point is on the instantaneous axis of rotation, ω 2
Alternatively,
~rA0 = ~rA = ê3 2a sin α + a(ê3 cos α − ê1 sin α) = ê3 2a sin α − ê1 a cos α
~r˙CA = ~n × ~rCA = n(ê3 cos α + ê1 sin α) × a(−ê1 cos α + ê3 sin α)
= −naê2 = −~r˙CQ , as expected
so
~r˙A0 = ~r˙A + Ω
~ × ~rA = ~n × ~rCA + Ω
~ × (~rC + ~rCA ) = ω
~ × ~rCA , as expected
= Ω sin α(ê3 sin α − ê1 cos α) × a(−ê1 cos α + ê3 sin α) = 0
5.3. ROTATIONAL KINEMATICS 85
(i) With φ̇ = Ω, ψ̇ = n and θ = α these are Euler’s angles. I am given φ◦ = ψ◦ = 0 so φ = Ωt and ψ = nt and the
formula sheet gives me
T
Since the axis of rotation is in the plane of the disk, ê003 · ω
~ = 0 and I have n = −Ω cos α +1
AF
DR
86 CHAPTER 5. MIDTERM REVIEW
T
AF
DR
T
Chapter 6
AF
6.1
Resources
Thornton & Marion – Ch. 11
Introduction
The motion of rigid bodies is significantly more complicated than that of particles. However, there are several ways
in which we can make the description of rigid body dynamics less cumbersome. In particular, using the principle axes
for the body frame basis and using Euler angles (see section F.0.1) for the generalized coordinates (with ê3 along any
cylindrical axis) will greatly simplify the problem.
˙0
↔ ↔0
~˙ 0 = I · ω
L ~˙ = N
~ + I ·ω ~ 0,
the rate of change of the angular momentum is equal to the net torque, N ~ 0.
Because the representation of the inertia tensor in a fixed frame changes during the motion of the body (i.e. İ 6= 0)
it is convenient to use the body frame in which the representation of the inertia tensor is constant. The angular
momentum vector can be represented in either basis, L ~ = L0 ê0 = Li êi and the time derivative is then1
i i
↔ ↔
~˙ 0 = L
L ~˙ + ω
~ ×L ~˙ + ω
~ = I ·ω ~ × (I · ω ~ 0,
~) = N
in which N ~ 0 = P ~r 0 × F~ .
~ = P ~r × F~ is the net torque in the body frame and is related to that in the fixed frame, N
The advantage of this expression over the previous one is that it contains no time derivatives of the inertia tensor
(because it is constant in the body frame).
1 Recall that the time-derivative operator in the IRF is related to that in the NIRF by d 0
= d ~˙
~ ×. In this expression the prime on the L
+ω
dt dt
indicates that the time-derivative is calculated in the primed reference frame.
87
88 CHAPTER 6. RIGID BODY DYNAMICS
T
N1 −ω3 L2 + ω2 L3
= N2 − ω3 L1 − ω1 L3
N3 −ω2 L1 + ω1 L2
N1 + ω3 L2 − ω2 L3
= N2 − ω3 L1 + ω1 L3
N3 + ω2 L1 − ω1 L2
which is a set of three coupled first order ODEs in the Li .
If I make the obvious choice to use the principal axes of the object as the body frame basis then
AF
and I have
I22 ω̇3
N1 + ω3 I22 ω2 − ω2 I33 ω3
N1 + ω2 ω3 (I22 − I33 )
N3 + ω1 ω2 (I11 − I22 )
I22 ω̇2 = N2 − ω3 I11 ω1 + ω1 I33 ω3 = N2 + ω3 ω1 (I33 − I11 ).
N3 + ω2 I11 ω1 − ω1 I22 ω2
Consider the motion of a body in the absence of a net torque. (a) Differentiate Euler’s equations and use the original
equations to eliminate the first derivatives. (b) Assume that the angular velocity is almost parallel to the ê1 axis,
~ = ω1 ê1 + 2 ê2 + 3 ê3 . What are the second order equations for the three components of ω
ω ~ ? Keep only first order
terms in the i . (c) What can you conclude if I11 is the largest or smallest of the three moments of inertia? (d) What if
I22 < I11 < I33 ?
Solution
(a) Differentiate the first of Euler’s equations for a rigid body I get
I11 ω̈1 = (ω̇2 ω3 + ω2 ω̇3 )(I22 − I33 )
DR
I can now substitute in the other two equations,
I33 − I11 I11 − I22
Å ã
I11 ω̈1 = ω3 ω1 ω3 + ω2 ω1 ω2 (I22 − I33 )
I22 I33
I11 I22 I33 ω̈1 = − ω32 (I33 − I11 )I33 + ω22 (I11 − I22 )I22 (I33 − I22 )ω1 .
I can find the other two equations by cyclically permuting the indices,
I11 I22 I33 ω̈2 = − ω12 (I11 − I22 )I11 + ω32 (I22 − I33 )I33 (I11 − I33 )ω2
I11 I22 I33 ω̈3 = − ω22 (I22 − I33 )I22 + ω12 (I33 − I11 )I11 (I22 − I11 )ω3
(b) If the motion initially has an angular velocity almost parallel to the ê1 axis, ω ~ = ω1 ê1 + 2 ê2 + 3 ê3 , then
I11 I22 I33 ω̈1 = − 23 (I33 − I11 )I33 + 22 (I11 − I22 )I22 (I33 − I22 )ω1
I11 I22 I33 ¨2 = − ω12 (I11 − I22 )I11 + 23 (I22 − I33 )I33 (I11 − I33 )2
I11 I22 I33 ¨3 = − 22 (I22 − I33 )I22 + ω12 (I33 − I11 )I11 (I22 − I11 )3
2 I have used the cross product matrix,
0 −ω3
ñ ô
ω2
[~
ω ]× ≡ [Ωij ] = ω3 0 −ω1
−ω2 ω1 0
because I felt like it. See Appendix E for a derivation.
6.3. AXIALLY SYMMETRIC SYSTEMS (AKA TOPS) 89
T
I33 − I11 I22 − I11
Å ãÅ ã
¨3 = −ω12 3
I33 I22
Ä äÄ ä
I33 −I11 I22 −I11
where Ω̃2 ≡ ω12 I33 I22 is the frequency of oscillation about the equilibrium ω
~ ◦ = ω1 ê1 .
(c) If I11 is either the largest or smallest of the three moments of inertia then I can write
¨i = −Ω̃2 i
Ä äÄ ä
where Ω̃2 ≡ ω12 I33I−I33
11 I22 −I11
I22 is the frequency of oscillation about the equilibrium ω ~ ◦ = ω1 ê1 . Any
~ from the ê1 axis will be oscillatory3 and a rotation about the ê1 axis will be stable. Examples of
AF
6.3
deviations of ω
such stable motions are frisbees and arrows.
(d) If I22 < I11 < I33 then
In general, the angular velocity vector and the body axis will not be parallel – if they are parallel then the problem
reduces to rotations in a plane perpendicular to the spin axis (i.e. all axial vectors will be parallel and perpendicular to
all of the velocity vectors).
For a system with dynamical axial symmetry, with ê3 along the symmetry axis, the inertia tensor in the principle
axis basis of the body frame is
I11 0 0
X
I = 0 I11 0
0 0 I33
The angular momentum is then
DR
↔
~ = I ·ω
L ~
= I11 (ω1 ê1 + ω2 ê2 ) + I33 ω3 ê3
= I11 (ω1 ê1 + ω2 ê2 + ω3 ê3 ) + (I33 − I11 )ω3 ê3
~ + (I33 − I11 )ω3 ê3 .
= I11 ω
From this I can conclude that the angular momentum is in the ω
~ -ê3 plane (see Figure 6.1).
are constants. The angle α between ω ~ and L ~ must therefore be fixed and I can write the angular velocity vector as
~ = nê3 + Ωê3 (see Figure 6.1; the parallelogram rule is used to add the components). I do so because the ω
ω 0
~ -ê3 plane
~ = Lê0 with angular velocity Ω
is therefore rotating about the L ~ – a type of motion called precession.
3
3 Note that the frequency of such oscillations will be the same in the two directions.
90 CHAPTER 6. RIGID BODY DYNAMICS
⃗
ω ê′3
ω3′
ê3 ⃗
L
ω3
T
α
⃗
n
⃗
Ω
θ ê2
ω2
θ ê′2
ω2′ O, O′
´¹¹ ¹ ¹ ¹ ¹ ¹ ¹ ¹ ¶
momentum, Ω 6= ω30 .
I11 0
I = 0 I11 0 .
0 0 I33
0
~ are co-planar, the ω
~ , ê3 , and L ~ -ê3 plane. For a torque-free
~ = Lê03 . Because the kinetic energy, T = 1 ω
~ is fixed and I can write the angular velocity vector as ω
~
2
~
~ · L,
~ = nê3 + Ωê03
~ -ê3 plane rotates at angular velocity Ω about the ê03 axis (i.e.
the system precesses). It is important to note that the spin, ~n, is not component of the angular momentum along the body
symmetry axis, n 6= ω3 , and that the precession is not the component of the angular momentum parallel to the angular
Consider first an object with dynamic axial symmetry which is torque free. In the principle axes (PA) system I have
I11 = I22 , and the inertia tensor is
(a) What are Euler’s equations for this system? Write your coupled ODEs in terms of the variables, ωi and a constant,
Ω. (b) Solve these coupled ODEs by introducing the complex quantity,
ω̃ ≡ ω1 + iω2 .
DR
(c) Describe ω
~ (t) in the body frame. (d) In the absence of any torque both the angular momentum vector, L, ~ and the
kinetic energy, T = 2 ω1 ~
~ · L, will be a constants. Use this to Describe ω ~
~ (t) fixed frame. (Hint: Choose L = Lê03 .)
(e) Calculate the quantity L~ · (~
ω × ê3 ) and draw any conclusions you can from this.
Solution
(a) Euler’s equations simplify to
I11 − I33
ω̇1 = ω2 ω3 = Ωω2
I11
I33 − I11
ω̇2 = ω3 ω1 = −Ωω1
I11
ω̇3 = 0
I11 −I33
where Ω ≡ I11 ω3 is a constant.
ω̃ ≡ ω1 + iω2 ,
6.3. AXIALLY SYMMETRIC SYSTEMS (AKA TOPS) 91
⃗
ω
ê′3
ê3
T
O, O′
´¹¹ ¹ ¹ ¹ ¹ ¹ ¹ ¹ ¶
Figure 6.2: In the body frame the angular momentum vector traces out the body cone while in the fixed frame it traces out
the space cone. These two cones touch along the angular momentum vector and one can visualize the body cone rolling over
AF
the space cone in a manner similar to Ex. 4.2.
so that
(c) In the ê1 -ê2 plane ω1 ê1 + ω2 ê2 will trace out a circle and the angular velocity vector, ω
in the body frame with ê3 as its axis and with angular frequency Ω ≡ I11I−I
Figure 6.2).
11
33
ω3 .
~ , will thus trace out a cone
This is the body cone (see
(d) Since the projection of the angular velocity vector onto the angular momentum vector is constant and I can con-
clude that in the fixed frame the cone traced out by the angular velocity vector will have the angular momentum
vector as its axis. And that it will precess around it at angular frequency Ω. This is the space cone (see Figure 6.2).
DR
(e) Further, consider the “volume” given by,
~ · (~
L ω × ê3 ) = (I11 ω1 ê1 + I22 ω2 ê2 + I33 ω3 ê3 ) · (ω2 ê1 − ω1 ê2 )
= I11 ω1 ω2 − I22 ω2 ω1 = 0,
so the three vectors are co-planar. It is convenient to choose the fixed ê03 axis as parallel to the constant angular
momentum,
L~ = Lê0 ,
3
In the previous example the angles α and θ were constant so the angular velocity vector, ω
~ , traces out what are known
as the body cone in the body frame and the space cone in the fixed frame. In this way the body cone rolls around the
space cone similar to Exercise 4.2.
~˙ and ~n˙ will be opposite and the body cone rolls inside the space
Note that for the case I33 > I11 the signs of Ω
cone (cf. ??).
92 CHAPTER 6. RIGID BODY DYNAMICS
x 3′ x 3′
L L x3
ω ω
x3
θ θ ω3
T
Ω
Space Space α
fixed α ω3 fixed
Ω
cone cone
Body
cone Body cone
(a) (b)
Figure 6.3: The space and body cones. [figure from Thornton & Marion, Figure 10.1]
AF Figure 6.4: Recall that, while the Euler vectors (aka orientation vectors) do not add as vectors, their derivates do add as
~˙ = φ̇ ê03 + θ̇ ê1 + ψ̇ ê3 . Note that the basis vectors
~˙ + θ~˙ + ψ
vectors. In terms of the Euler angles the angular velocity is ω
~ =φ
used in this expression do not form an orthonormal basis so care must be exercised when contemplating or calculating the
components of the angular velocity.
DR
6.3.2 The Components of ω
~
For an object with dynamic axial symmetry the inertia tensor is unchanged by rotations about its symmetry axis. It
is then very convenient to divide its angular velocity into a spin part along the axis, ~n = nê3 and an orbital part,
~ = Ωê0 , describing the motion of the body frame as it precesses about the (constant) angular momentum vector,
Ω 3
~ = Lê0 – i.e. the object’s spin motion about its axis can be considered independently of the precession of the body
L 3
frame about the angular momentum vector (because the inertia tensor does not change for rotations about the spin
axis). The total angular momentum is then
ω ~ = nê3 + Ωê0 .
~ = ~n + Ω 3
It is important to note that the spin is not equal to the component of the angular momentum along the body symmetry
axis, ~n 6= ω
~ · ê3 = ω3 , and that the precession is not the component of the angular momentum parallel to the angular
momentum, Ω ~ 6= ω~ · ê03 = ω30 (see Figure 6.1).
In the body frame the angular momentum vector traces out the body cone while in the fixed frame it traces out the
space cone. These two cones touch along the angular momentum vector and one can visualize the body cone rolling
over the space cone in a manner similar to Ex. 4.2.
6.3. AXIALLY SYMMETRIC SYSTEMS (AKA TOPS) 93
Euler Angles
It is now convenient to re-introduce the Euler angles (see Figure F.3). We can see that the spin angular momentum,
~n, is exactly ψ ~˙ and the orbital (precession) angular momentum, Ω, ~˙ This leaves the obvious question:
~ is exactly φ.
What about θ? ~˙ This last part is the “nutation” angular velocity – which we view as an oscillation of the spin axis as it
T
precesses about the angular momentum, L. ~
In the presence of nutation oscillation the total angular momentum will be
ω ~˙ + θ~˙ + ψ.
~ =φ ~˙
AF Ä
~˙ = ψ̇ ê3 = ψ̇(sin θ(ê1 sin φ + ê2 cos φ) + ê0 cos θ),
~n = ψ
~ = ê01 ψ̇ sin θ sin φ + θ̇ cos φ + ê02 −ψ̇ sin θ cos φ + θ̇ sin φ + ê03 ψ̇ cos θ + φ̇
ω
Ä ä Ä ä Ä
= ê1 φ̇ sin θ sin ψ + θ̇ cos ψ + ê2 φ̇ sin θ cos ψ − θ̇ sin ψ + ê3 φ̇ cos θ + ψ̇
ä
ä
DR
4 Or I can recall from section F.0.1 that I have for a zxz Euler rotation,
ôñ 0 ô
cos ψ sin ψ 0 1 0 0 cos φ sin φ 0
ñ ô ñ ôñ ôñ
ê1 ê1
ê2 = sin ψ cos ψ 0 0 cos θ sin θ sin φ cos φ 0 ê02
ê3 0 0 1 0 sin θ cos θ 0 0 1 ê03
ôñ 0 ô
cos ψ cos φ − cos θ sin ψ sin φ − cos ψ sin φ − cos θ cos φ sin ψ sin ψ sin θ
ñ
ê1
= cos φ sin ψ + cos ψ cos θ sin φ cos ψ cos θ cos φ − sin ψ sin φ − cos ψ sin θ ê02
sin θ sin φ cos φ sin θ cos θ ê03
94 CHAPTER 6. RIGID BODY DYNAMICS
P N
T
R
AF Figure 6.5: The angular orientation of an object can be represented by Euler angles, φ, θ, & ψ, and the angular velocity by
the sum of their time derivatives, ω
~ =φ
ê3
n
⃗
Q
~˙ Here ψ
~˙ + θ~˙ + ψ.
a
˙
C
θ
b
θ
Ω
⃗
⊙
~˙ is the precession about the angular
~˙ is the spin, about the body axis, φ
momentum vector (depicted vertical here), and θ~ is the nutation, an oscillation of the body axis as it precesses. [image By
User Herbye (German Wikipedia). Designed by Dr. H. Sulzer - Original, CC BY-SA 3.0]
ê′3
ê2
ê′2
O, O′ ê1 , ê′1
´¹¹ ¹ ¹ ¹ ¹ ¹ ¹ ¹ ¶ ´¹¹ ¹ ¹ ¹ ¹ ¹ ¹ ¹ ¹ ¹ ¶
Figure 6.6: A lollipop (mass m, radius a) is resting on the ground with its stick pivoted at point O. As the lollipop rolls the
DR
point of contact with the ground traces out a circle of radius b. For this problem the NIRF, êi , is rotating at angular velocity
~ = Ωê03 relative to the IRF, ê0i .
Ω
page)? (b) What is the relationship between êi and ê0i as a function of time? (Hint: Take φ = Ωt as the rotation about
ê03 .) (c) What are the principle axes and what is the inertia tensor about the pivot point? (d) What is the magnitude
and direction of the angular velocity vector, ω ~ ? (e) What are the components of ω ~ (i.e. write ω
~ in terms of the two
bases, ωi êi and ωi0 ê0i as well as the sum of the spin and precession angular velocities, ~n + Ω)? ~ (f) What is the angular
momentum, L, ~ in the principle axis system? (g) What is the time rate of change of the angular momentum vector in the
inertial reference frame, L ~˙ 0 ? (h) The torque on the lollipop is caused by the gravitational force, m~g , acting through the
centre of mass and the normal force, FN , acting through the contact point (the force at the pivot produces no torque).
What is the normal force? (Hint: What do you get in the Ω → 0 limit?)
Exercise 6.2 Tip Top.
6.3. AXIALLY SYMMETRIC SYSTEMS (AKA TOPS) 95
Consider a top made from a disk of mass m and radius a and a light spindle of length 2` perpendicular to the disk
and passing through its centre with the centres of the disk and spindle coinciding (i.e. the disk is halfway down the
spindle). The base of the top is fixed at the origin of a fixed coordinate system, {ê01 , ê02 , ê03 }, with ê03 vertical. A rotating
coordinate system has ê3 along the axis of the top. Gravity is acting upon the top. (a) What are the principle axes
and principle moments of inertia at the origin. Justify your answers. (b) What is the net torque, N ~ , about the origin?
T
(c) What is the angular velocity, ω ~ , of the top in terms of ê3 and its derivatives? (Hint: Start with ê˙ 3 = ω ~ × ê3 and
take the cross product with ê3 .) (d) What is the angular momentum, L, ~ of the top in terms of ê3 and its derivatives?
(e) What is the equation of motion for the top? (Hint: We know that L ~˙ = N~ g where N ~ g is the torque caused by gravity,
−mgê3 , acting at the centre of mass of the top, `ê3 .) (f) Write ê3 = ê3 + ~ and so find the equation of motion for ~.
0
Is this equation valid only for small = |~ |? (g) Assume is small and linearize your equation of motion (Hint: this
means drop any terms proportional to n for n > 1 and , ˙ etc, is proportional to ). (h) Take ~ = 1 ê01 + 2 ê02 + 3 ê03 .
For small you can take 3 ≈ 0. What are the equations of motion for 1 and 2 ? (i) These equations of motion should
(again) be familiar; introduce a complex variable, q̃ = 1 + i2 , and solve the resulting equation for q̃(t). (j) Under
what conditions will the top be stable in the upright position? (k) For a = ` = 1 cm and m = 1 g what values of ω, in
AF
terms of rpm, are required for stability?
DR
96 CHAPTER 6. RIGID BODY DYNAMICS
T
AF
DR
T
Chapter 7
Virtual Work
AF
7.1
Resources
Thornton & Marion – §7.3
7.1.1
Static Equilibrium
Constraints & Generalized Coordinates
For a system of N particles there will be N position vectors, ~ri , i = 1, 2..., N , with 3N components. The number of
quantities, n, that must be specified in order to completely describe the configuration of the system, called the degrees
of freedom, will be less than or equal to the 3N components of the N , ~ri .
Week 7
If n < 3N then there are 3N −n constraints on the system. The n quantities, qa , a = 1, 2, ..., n, are the generalized
coordinates, and their time derivatives, dt d
qa ≡ q̇a , are the generalized velocities.
The positions of the N particles, ~ri , will be functions of the n generalized coordinates, qa , and time, t:
Note in particular that the motions described by the qa are consistent with the constraints – ie the constraints remain
satisfied
Specifying the qa is not enough to describe the state of the system; the q̇a are also needed, a total of 2n quantities.
Because the qa and q̇a are specified
î óindependently, we consider them as independent variables. The state of a system
is a vector in phase space, ψ ~ = q~˙ , and the state of the system as it evolved in time traces out a trajectory in phase
q
~
~
space, ψ(t).
It is convenient to collect the n qa in an ordered n-tuple, ~q ≡ (q1 , q2 , ..., qn ), so that ~q˙ = (q̇1 , q̇2 , ..., q̇n ) and
~ri = ~ri (~q, t).
The relationships between the ~q, ~q˙, and ~q¨ are called the equations of motion and are second order ODEs in ~q. And
the solutions of these equations, ~q(t), describe a path in the n-dimensional configuration space.
The total differentials of the position vectors are,
∂~ri ∂~ri
d~ri = dqa + dt.
∂qa ∂t
97
98 CHAPTER 7. VIRTUAL WORK
T
because a displacement can not actually happen instantaneously.
Holonomic Constraints
Note that we have specified the constraint equation in a limiting manner. For example, they don’t depend upon any of
the velocities, ~r˙i , or higher derivative. Such constraints are called holonomic.
Note that the forces that keep a rigid body rigid are also constraints,
AF
but that we don’t usually include the internal forces of a rigid body in our calculations (generally by specifying the
COM coordinates and the Euler angles, a total of six generalized coordinates).
then of the 3N coordinates only 3N − C = n will be independent. We call n the degrees of freedom of the problem
at hand.
When we write Newton’s second law for our system of N particles we can classify the forces on each particle
as either applied, F~i(A) , or constraint, F~i(C) forces, where we define the constraint forces as those orthogonal to those
consistent with the constraints (e.g. normal forces). For a system in equilibrium we have
N
X
0= F~i(A) + F~i(C) , i = 1...N,
i=1
in which the internal constraint forces (i.e. the action-reaction pairs that make a rigid body rigid) are not included
because we are summing over all of the particles.
Now consider the virtual work done on the system in the virtual displacements δ~ri ,
T
Example 7.1 Leverage.
Consider the rigid beam, length `1 + `2 , with a fulcrum as depicted. The two forces, F~i are applied at each end.
Use the principle of virtual work to calculate the ration F1 /F2 so that the system remains in balance. Use θ as your
generalized coordinate and put the origin at the fulcrum.
Solution
The points of application of the two forces are
AFso
F1 `1 = F2 `2 ⇒
F1 `2
= .
F2 `1
X
DR
Example 7.2 Virtual Ladder.
A ladder of length ` and mass m is leaning against a vertical wall and makes an angle α with the horizontal floor. The
wall is smooth but the floor is rough. Use the principle of virtual work to find the friction, FF , that prevents the ladder
from slipping. Choose the angle α as your generalized coordinate.
Solution
Here the constraint forces are the normal reactions at the wall and floor. They do
no work in the virtual displacement δα and the applied forces are the force due to
gravity, FG , and the friction, FF .
Putting the origin at where the wall meets the floor, the position of the centre of
mass (where FG acts) is
Since δα is a motion consistent with the constraints (i.e. it is an “allowed” motion) I must have
T
FF = 12 mg cot α.
Note that, conventionally, we would sum the torques about the the point where the two constraint forces meet,
giving
X
~ = 0 = (mg)( 1 ` cos α)(+ẑ) + (FF )(` sin α)(−ẑ) ⇒ FF = 1 mg cot α.
N 2 2
AF
Example 7.3 Simple Machine.
Consider a simple machine consisting of two rods (each
of length ` and negligible mass) connected as shown. The
pins at A, B and C are frictionless and A is fixed while C
can move horizontally. The pins A and C are at the same
height above the horizontal surface. The force F~1 is applied
vertically at B and the force F~2 is applied horizontally at C.
The system is in equilibrium. Use the principle of virtual
work to find the force F~2 in terms of the force F~1 . (Hint:
Use the angle θ for your virtual displacement.)
Solution
I have F~1 = −F1 ŷ and F~2 = −F2 x̂ so the virtual work done is
1
2 F1 = NC = F2 cot θ ⇒ F2 = 21 F1 tan θ,
as before. X
7.2. D’ALEMBERT’S PRINCIPLE 101
θ̂
θ
r̂
T
Figure 7.1: A simple pendulum using polar basis, ~r = `r̂.
AF
so that Ä
i
so that Newton’s second law for each particle takes the form
and we can perform the same procedure we did for virtual work,
δW = 0 = δW (I) + δW (A) +
δW
ä
*(C)0
i = 1...N
This is d’Alembert’s principle: The net work done by the applied forces and the inertial forces in a virtual displacement
is zero. (Note that the introduction of inertial forces reduces dynamics to statics.)
With d’Alembert’s principle we considered virtual displacements about the instantaneous state of the system.
Such virtual displacements were introduced by Johann Bernoulli and D’Alembert. By introducing inertial forces they
reduced dynamics to statics. This differential form was originally proposed by d’Alembert but Lagrange recast the
ideas into a form that is very easy to use for dynamical problems.
Whenever there is a differential form of a physical principle there is inevitably an integral form. The most important
DR
integral form in classical mechanics is Hamilton’s principle.1
In quantum mechanics most systems are holonomic and it is d’Alembert’s principle and its variations that are most
useful rather than Newton’s second law.
Example 7.4 Constrained Bob.
Consider the simple pendulum which traces out a circular arc. We have N = 1. (a) What are the constraints on this
system? (b) How many degrees of freedom does it have? (c) Discuss how this problem is first approached and how
we will now describe it using the language of constraints and degrees of freedom.
Solution
(a) We have the three components of the position vector of the bob, ~r = ~r(x, y, z). The bob is constrained to move
in a single vertical plane, say y = 0. And the distance of the bob from the pivot is constrained by the inextensible
string of length `, x2 + z 2 − `2 = 0.
(c) We usually use polar instead of Cartesian coordinates because we are able to eliminate one of the equations
resulting from Newton’s second law (i.e. the radial part), thereby reducing the problem to a set of f equations
where f is the number of degrees of freedom of the problem. (i.e. For this problem there is only one independent
variable which we have identified as θ(t).) X
102 CHAPTER 7. VIRTUAL WORK
y
x
`
θ
m
T
Figure 7.2: A simple pendulum using Cartesian basis, ~r = êi xi = `(ê1 sin θ − ê2 cos θ). Note that (ê1 sin θ − ê2 cos θ) = r̂,
the unit vector used in Figure 7.1.
AF
Solution
We have tan θ = −x/y and x2 + y 2 = `2 so the position of the bob, ~r,
(a) in terms of the generalized coordinates are
Ä p
~r(x) =x̂x + ŷ − `2 − x2
Ä p ä
ä
(b) Since I know that I will be taking derivatives, etc, I will use ~r(θ),
θ `δθ sin θ
~r =`r̂
~r˙ =`r̂˙ = `θ̇θ̂
DR
˙
~r¨ =`θ̂θ̂ + `θ̈θ̂
= − `θ̇2 r̂ + `θ̈θ̂
The applied force is F~ (a) = −mg ŷ so the virtual work done by the applied force is
δW (a) = F~ (a) · δ~r = −mg ŷ · `(x̂ cos θ + ŷ sin θ) δθ = −mg` sin θ δθ.
1 The
Rt
action integral, I = dt L, is stationary under arbitrary variations of the coordinates; see section 10.1 for more details.
t◦
7.2. D’ALEMBERT’S PRINCIPLE 103
T
which, as we already knew, is the equation of motion of a simple pendulum.
(c) Because virtual displacements occur at a moment frozen in time the work done by gravity in a virtual displacement,
δθ, is the same as in part (b), δW (a) = −mg` sin θ δθ, However, the acceleration of the bob is more complicated
because ` = `(t),
~r =`r̂
~r˙ =`r̂
˙ + `r̂˙ = `r̂
˙ + `θ̇θ̂
¨ + `˙r̂˙ + `˙θ̇θ̂ + `θ̈θ̂ + `θ̂θ̂˙
~r¨ =`r̂
~ ~ (a)
d Ä 2 ä
dt
=(`¨ − `θ̇2 )r̂ + (2`˙θ̇ + `θ̈)θ̂
m` θ̇ = −mg` sin θ,
ä
δW = −m` g sin θ + 2`˙θ̇ + θ̈` δθ = 0
Solution
If I put the origin on the ceiling then the positions of the two masses are ~ri = −xi x̂ for i ∈ {1, 2}. The virtual
displacements are then, δ~ri = −δxi x̂ With δx1 + δx2 = 0 so that
δ~r2 = +δx1 x̂
.
The applied forces are mi gx̂ and the inertial forces are −mi ẍi . The work done by the applied forces is
δW (i) = (−mi ẍi x̂) · (δxi x̂) = −m1 ẍ1 δx1 − m2 ẍ2 δx2 .
There is a constraint, x1 + x2 = c (c some constant) which gives me δx1 + δx2 = 0 and ẍ1 + ẍ2 = 0 so that
giving
m1 − m2
T
ẍ1 = g
m1 + m2
X
h
AF
mass m2 slides on the inclined surface without friction. Use d’Alembert’s princi-
ple to find the acceleration of m1 .
7.3
y
Constraint Equations
s
m
α
g
x(t) M
m
α
g
(a) For a block of mass m on a frictionless inclined plane (angle α to the horizontal, as shown to the left) what is the
position, ~r = (x, y), in terms of the generalized coordinate s? How about in terms of x or y? (b) Use d’Alembert’s
principle to find the equation of motion. State clearly your virtual displacement. (c) A block of mass m slides on a
frictionless surface which moves horizontally according to x(t). Use d’Alembert’s principle to find the equation of
motion. (Hint: The acceleration of the block is not parallel to the incline.) For what values of x(t) will the block
accelerate up the incline? (Hint: Does your answer make sense in the α → 0 limit?) (d) A block of mass m slides on
DR
a frictionless inclined surface of mass M which is free to slide on a frictionless horizontal surface. Use d’Alembert’s
principle to find the equations of motion of both the block and the inclined plane. Does your answer make sense in
the α → 0 and M → ∞ limits? (e) One of your equations of motion in (d) can be easily integrated. Give a physical
interpretation of this integral.
θ
̂
ı̂ m
x M
O
Xf
∂~r
are the virtual displacements, δ~r1 and δ~r2 , of the masses M and m, respectively? Recall that δ~r = δqa .
a=1
∂qa
(b) What is the work done in a virtual displacement by the inertial forces and applied forces, δW (i) and δW (a) , re-
spectively? (c) What are the equations of motion corresponding to the virtual displacements δx and δθ? Simplify your
answers but don’t bother isolating ẍ or θ̈.
T
Exercise 7.3 Spherical Pendulum.
(a) Write each of the Cartesian unit vectors, x̂, ŷ & k̂, in terms of the spherical unit vectors, r̂, θ̂ & φ̂. (Hint: Check
that your unit vectors are mutually orthogonal.) (b) Write each of the spherical unit vectors, r̂, θ̂ & φ̂, in terms of
the Cartesian unit vectors, x̂, ŷ & k̂, and then find the time derivative of each. Express your answers in terms of the
spherical unit vectors. (c) Calculate the second time derivative of each of the spherical unit vectors, r̂, θ̂ & φ̂. You may
use your results from (b). (d) Consider the spherical pendulum consisting of an inextensible string of length ` and a
bob of mass m. The position of the bob in spherical coordinates is ~r = `r̂ and our generalized coordinates are θ and
φ. Calculate ∂~r/∂θ and ∂~r/∂φ and so find an expression for δ~r. (e) The applied force is m~g and the inertial force is
−m~r¨. Calculate the virtual work,
AF ,
δW = m~g − m~r¨ · δ~r.
,
∂θ ∂ θ̇ dt ∂ θ̇
(i) Put your results from (h) into Lagrange’s equation
d ∂L
Å
dt ∂ q̇a
Ä
∂L
∂qa
=0
,
ä
(Hint: If all of the vectors in the above are in spherical coordinates the virtual work is much easier to calculate.) (f) As
θ and φ are independent each of the coefficients of δθ and δφ in your virtual work expression can be set equal to
zero. Use this to identify two equations of motion. (g) Show that your solutions to part (f) give the expected results
in both the θ̇ = 0 and φ = 0 limits. (Hint: We are measuring θ from the positive z-axis so θ̈ is positive in the
direction towards the equilibrium position. This may change what your “expected” results are.) (h) Now calculate the
Lagrangian, L = T − V , where T = 21 m~r˙ · ~r˙ is the kinetic energy and V = −m~g · ~r is the potential energy, as well
as the derivatives,
∂L ∂L d ∂L
Å ã
∂L ∂L d ∂L
,
Å
∂φ ∂ φ̇ dt ∂ φ̇
ã
DR
106 CHAPTER 7. VIRTUAL WORK
T
pended from the mass M by a light, inextensible string of
θ
length a and is free to swing in the vertical plane containing
the horizontal bar (i.e. the plane of this page). ŷ m
This problem has two degrees of freedom. Use x and θ, as
indicated on the diagram, for your generalized coordinates. x̂
(a) Using the point O as your origin, what are the position vectors, ~rM & ~rm , of the two masses as functions of the
generalized coordinates in terms of the Cartesian unit vectors, x̂ & ŷ? What are the corresponding first and second
time derivatives? (b) Using the point o as your origin, what is the position vector, ~rm , of the mass m as a function
of the generalized coordinates in terms of the polar unit vectors, r̂ & θ̂? What are the corresponding first and second
time derivatives? (c) What are the virtual displacements δ~rM and δ~rm in terms of the Cartesian unit vectors, x̂ & ŷ?
AF
(d) What is the virtual displacement δ~rm in terms of the polar unit vectors, r̂ & θ̂? (e) What is the virtual work, δW (A) ,
done by the applied force(s)? (f) What is the virtual work, δW (I) , done by the inertial force(s)? (g) Use d’Alembert’s
Principle to find the equations of motion for this system.
DR
T
Chapter 8
Lagrangian Mechanics
AF
8.1
Resources
Thornton & Marion – Ch. 7
Boas – §5.5
Newton’s second law for the N positions at which the forces act, ~ri , is mi~r¨i = F~i . This can be written as
But we already know that the virtual work done by the individual constraint forces is zero, δWi(C) = F~i(C) · δ~ri = 0, so
N
X N
X
δW = F~i(A) · δ~ri + F~i(I) · δ~ri = 0,
i=1 i=1
which states that the net virtual work done, δW , by the applied and inertial forces is zero. This is d’Alembert’s
principle (described in more detail in chapter 7).
1 In this chapter there is no implied sum over roman indices in the middle of the alphabet (e.g. i, j, k, etc) which enumerate the N positions at
which the forces act, ~ri . However, for repeated indices roman indices at the beginning of the alphabet (e.g. a, b, c, etc) the sum is over the degrees
of freedom (DOF), f , is implied.
107
108 CHAPTER 8. LAGRANGIAN MECHANICS
If the system has f DOF then these can be described by f generalized coordinates, {qa | a = 1...f }, and this allows
me to write the virtual displacement of the ith particle as2
∂~ri
~ri = ~ri ({qa }, t) ⇒ δ~ri = δqa . (implied sum over repeated index a = 1...f )
∂qa
T
An important aspect of this is that the qa are independent – they each represent a different virtual displacement which
can be examined in isolation.
The virtual work done by the applied forces is
N
X N
X ∂~ri
δW (A) = F~i(A) · δ~r = F~i(A) · δqa = Qa δqa
i=1 i=1
∂qa
AF
where the
N
X
i=1
¨
mi~ri ·
(I)
δW = −
i=1
∂~ri
∂qa
=
XN
N
X
i=1
d
dt
Å
Qa ≡
mi~r¨i · δ~r = −
˙
mi~ri ·
i=1
N
X
i=1
∂~ri
∂qa
F~i(A) ·
mi~r¨i ·
ã X
−
N
i=1
∂~
r
N
X
i
∂qa
˙
∂~ri
∂qa
are the generalized forces. Note that the generalized displacements and generalized forces are both scalar quantities.
!
δqa .
mi~ri ·
d ∂~ri
Å
dt ∂qa
ã
(8.1)
Now, I have ~ri = ~ri ({qa }, t) so the velocities and their partial derivatives are
DR
∂~ri ∂~ri
~r˙i = q̇b +
∂qb ∂t
∂~r˙i ∂ ∂~ri ∂ ∂~ri ∂ 2~ri ∂ 2~ri
Å ã Å ã
= q̇b + = q̇b + (8.2)
∂qa ∂qa ∂qb ∂qa ∂t ∂qa ∂qb ∂qa ∂t
˙
∂~ri ∂~ri ∂ q̇b ∂~ri ∂~ri
= = δa,b = . (8.3)
∂ q̇a ∂qb ∂ q̇a ∂qb ∂qa
However,
d ∂~ri ∂~r˙i
Å ã
= (8.4)
dt ∂qa ∂qa
Substituting Equation 8.3 and Equation 8.4 into Equation 8.1, I get
N
X XN Å
∂~ri d ∂~ri d ∂~ri
Å ã Å ãã
mi~r¨i · = mi~r˙i · − mi~r˙i ·
i=1
∂qa i=1
dt ∂qa dt ∂qa
XN
∂~r˙i ∂~r˙i
T
Ç Ç å å
d ˙ ˙
= mi~ri · − mi~ri ·
i=1
dt ∂ q̇a ∂qa
N
X d ∂ Ä1 ˙ ˙ ä ∂ Ä1 ˙ ˙ ä
Å Å ã ã
= mi~ri · ~ri − mi~ri · ~ri
i=1
dt ∂ q̇a 2 ∂qa 2
N
!! N
!
d ∂ X ∂ X
= 1
mi~r˙i · ~r˙i − 1
mi~r˙i · ~r˙i
dt ∂ q̇a i=1 2 ∂qa i=1 2
d ∂T ∂T
Å ã
= − .
AF
where
δW =
=−
N Ä
X
d
dt
i=1
Å
∂qa
d
dt
T ≡
Å
ä
N
X
i=1
∂T
∂ q̇a
∂T
∂ q̇a
ã
ã
1
−
∂T
r˙i
2 mi~
∂qa
∂T
∂qa
· ~r˙i
ã
− Qa δqa = 0
= Qa
Solution
(a) The positions and velocities of the two mass are
~rM = x̂ x ⇒ ~r˙M = x̂ ẋ
~rm = x̂ x + ŷ b + a(x̂ cos θ − ŷ sin θ) ⇒ ~r˙m = x̂ ẋ − aθ̇(x̂ sin θ + ŷ cos θ).
110 CHAPTER 8. LAGRANGIAN MECHANICS
T
= 12 M ẋ2 + 12 mẋ2 − maẋθ̇ sin θ + 12 m(aθ̇)2 .
(c) The applied forces are those due to gravity so the generalized forces are
∂~rM ∂~rm
Qx = M~g · + m~g · = −M g ŷ · x̂ − mg ŷ · x̂ = 0
∂x ∂x
and
∂~rM ∂~rm
Qθ = M~g · + m~g · = −M g ŷ · 0̂ − mg ŷ · a(−x̂ sin θ − ŷ cos θ) = mga cos θ.
∂θ ∂θ
AFThe generalized force Qx is the x̂ component of the total force while Qθ is the ẑ component (i.e. orthogonal to
the page) of the total torque.
dt
Ä
d ∂T
ä
∂ q̇a − ∂qa = Qa , are for q1 ≡ x,
dÄ
dt
∂T
ä
ä
M ẋ + mẋ − maθ̇ sin θ − 0 = 0
⇒
M ẋ + m(ẋ − aθ̇ sin θ) = Px ,
which tells us that the horizontal linear momentum is a constant, Px , and for q2 ≡ θ,
dÄ
−ẍ sin θ −
cos θ + aθ̈ = g cos θ −
ẋθ̇
ẋθ̇
cos θ,
which tells us that the time-rate-of-change of the total angular momentum, Lz = ma(aθ̇ − ẋ sin θ), is equal to the
net torque, mga cos θ.
Conservative Forces
If the applied forces are conservative they can be written as the divergence of a potential
X
F~i(A) = −∇
~ i Vi
DR
then
N
X XN
∂~ri ~ i Vi · ∂~ri
Qa = F~i(A) · =− ∇
i=1
∂q a i=1
∂qa
XN Å
∂Vi ∂xi ∂Vi ∂yi ∂Vi ∂zi
ã
=− + +
i=1
∂xi ∂qa ∂yi ∂qa ∂zi ∂qa
N
∂ X ∂V
=− Vi , = − ,
∂qa i=1 ∂qa
P
where V ({~ri }) ≡ i Vi (~ri ) is the total potential.
The work done by the applied forces is
∂V
δW (A) = Qa δqa = − δqa = −δV,
∂qa
as expected.
If the potential is a function of the generalized coordinates but not the generalized velocities (or higher derivatives)
then we can introduce the Lagrangian,
L=T −V
8.1. THE EULER-LAGRANGE EQUATION 111
T
8.1.1 Lagrangian Method
The ELE make finding the equations of motion of a suitable system algorithmic.
3. write the kinetic and potential energies, T & V , in terms of these coordinates, their time derivatives, {q̇a }, a =
1 . . . f , and time, t, and form the Lagrangian, L = T − V
AF
4. substitute L in the ELE and perform the derivatives
Solution
(a) The potential is
V = −F~ (A) · ~r = −(−mg ŷ) · ~rm = mg ŷ · (x̂ x + ŷ b + a(x̂ cos θ − ŷ sin θ)) = mgb − mga sin θ
so the Lagrangian is
DR
L = 12 M ẋ2 + 12 mẋ2 − maẋθ̇ sin θ + 12 m(aθ̇)2 − mgb + mga sin θ
ẑ
O
T
θ
`
ŷ
φ ~v
x̂
r̂
AF
Solution
Figure 8.1: For the conical pendulum the position and velocity vectors are orthogonal, ~r · ~v = `r̂ · ~r˙ = 0.
(a) I have ~r = `r̂ for the position fo the bob, and r̂˙ = θ̇θ̂ + φ̇ sin θφ̂ so the kinetic and potential energies are
dÄ 2 ä
m` φ̇ sin2 θ = 0
dt
dÄ 2 ä
m` θ̇ = m`2 φ̇2 sin θ cos θ + mg` sin θ.
dt
DR
(c) I have Lz ≡ m`2 φ̇ sin2 θ a constant. This is the ẑ component of the angular momentum. It is constant because
there is no torque about the ẑ axis.
dÄ 2 ä g
m` θ̇ = mg` sin θ ⇒ θ̈ = sin θ.
dt `
This is a simple plane pendulum (it lacks a minus sign because the angle θ is measured from the upwards vertical
rather than the downwards vertical as is more usual).
(f) With ~r = `r̂ the acceleration is ~r¨ = `r̂¨ and Newton’s second law gives
T
Ä ä
m` −θ̇2 − φ̇2 sin2 θ = − FT + mg cos θ
Ä ä
m` θ̈ − φ̇2 sin θ cos θ = − mg sin θ
Ä ä
m` 2θ̇φ̇ cos θ + φ̈ sin θ =0.
Ä ä
The second and third equations are the same as I found in (b) (the third can be integrated to d
dt m`2 φ̇ sin2 θ = 0).
The first equation is new and it gives F~T , the constraint force.
X
AF
8.2
8.2.1
Other things to do with Lagrangians
Semi-Holonomic Constraints
Note that the rolling constraint is a semi-holonomic constraint if it is integrable, usually to give a relationship between
s, the distance travelled by the centre of the rolling rigid body, and θ, the angle through which it rotates. For example,
a rolling cylinder, radius a, and angular velocity, θ̇, has ṡ = aθ̇ so ds = a dθ and ∆s = a ∆θ.
dL =
∂L
· d~r +
dV
d~r
∂L
· d~r +
· d~r˙ +
∂L
∂t
∂L
dt.
dt.
∂~r ˙
∂~r ∂t
I can therefore write
∂L dV ∂L
= & = m~r˙,
DR
∂~r d~r ∂~r˙
and I only need to write the total derivative of L to find its partial derivatives.
So for Example 8.3 I had the Lagrangian
L = 12 m`2 θ̇2 + 21 m`2 φ̇2 sin2 θ − mga cos θ.
so
∂L ∂L ∂L
dL = · d~r + · d~r˙ + dt = (mg` sin θ)dθ + (m`2 φ̇ sin2 θ)dθ̇ + (0)dt
∂~r ˙
∂~r ∂t
but
∂f ∂f
f˙(~q, t) = q̇b +
∂qb ∂t
so
∂ f˙ ∂f ∂f
T
= δa,b =
∂ q̇a ∂qb ∂qa
(which is “cancelling the dots”) and
∂ f˙ ∂2f ∂2f
= q̇b + .
∂qa ∂qa ∂qb ∂qa ∂t
I therefore have
d ∂ f˙ d ∂f ∂2f ∂2f ∂ f˙
= = q̇b + =
dt ∂ q̇a dt ∂qa ∂qa ∂qb ∂qa ∂t ∂qa
AF
so that the ELE gives the same solution for L0 as for L – adding a complete time-derivative, f˙, of a function of the
generalized coordinates and time, f = f (~q, t), does not change the resulting EOM.
8.3
In subsection 10.3.3 I will show another much more elegant proof of this statement.
d ∂L
Å
dt ∂ φ̇
ã
=
dÄ 2
dt
ä
m` φ̇ sin2 θ = 0.
The quantity m`2 φ̇ sin2 θ is therefore a constant of the motion (i.e. it doesn’t change during the motion). This
quantity, it turns out, is the ẑ component of the angular momentum, Lz = m`2 φ̇ sin2 θ, which, as there is no torque
P~ ~˙
about this axis, is a constant, N = L.
DR
We can see, however, that whenever one of the generalized coordinates, qa , is missing from the Lagrangian,
∂qa = 0, then corresponding quantity,
∂L
∂L
pa ≡ ,
∂ q̇a
known as the conjugate momentum, will be constant. The coordinate qa is said to be cyclic.
Often the conjugate momenta are components of the linear or angular momentum but they need not be.
Example – Translation
We can also turn this argument around: If, for example, q1 does not appear in the Lagrangian, ∂q ∂L
1
= 0, and if a change
in q1 , δq1 , corresponds to a translation of the system in, say, the x̂ direction, δx x̂, then, since V = V (qa ) we have
0
N
∂L ∂T ∂V 7 ∂ X1
= − = mi (ẋ2i + ẏi2 + żi2 )
∂ q̇1 ∂ q̇1 ∂ q̇1 ∂ q̇1 i=1 2
Ö 1 0 0 è
XN N
∂ ẋ
i
7 ∂ ẏi
7 7 X
∂ żi
= mi ẋi + ẏi + żi = mi ẋi
i=1 ∂ q̇1 ∂ q̇1 ∂ q̇1 i=1
8.3. CONJUGATE MOMENTA, SYMMETRY & CONSERVATION LAWS 115
so, since ∂L
∂q1 = 0, Lagrange’s equation gives
N
! N
d ∂L d X X
Å ã
0= = mi ẋi ⇒ mi ẋi = constant.
dt ∂ q̇1 dt i=1 i=1
T
From this we can conclude that the if the Lagrangian is invariant under a translation in the x̂ direction then the x̂
component of the linear momentum is conserved.
This is a profound result: The symmetry of the system under a translation (the system over here has the same
dynamics as the system over there) leads to an invariance in the Lagrangian ( ∂q
∂L
1
= 0) which, in turn, leads to a
conservation law. The generalization of this result is known as Nöther’s Theorem.
Example – Rotation
AF
Let’s look at another example: If q1 again doesn’t appear in the Lagrangian, ∂q
8.3.1 Symmetry
∂L
1
= 0, and if a change in q1 , δq1 ,
corresponds to a rotation, δφ, about a certain azis, say the ẑ axis, then, using cylindrical coordinates we have
Symmetry, or self-similarity, means that certain transformations leave a system unchanged. Common geometric ex-
amples of symmetries are translations, rotations, reflections, scaling and combinations thereof. In quantum mechanics
internal symmetries, such as exchange, spin, charge (and many more) are also possible. An important distinction is
that in classical mechanics the symmetries are almost always approximate while in quantum mechanics the symme-
tries are exact – a fact that can have important consequences on, for example, statistics. The mathematical study of
symmetry is group theory.
If a system under study has a particular symmetry then the corresponding Lagrangian will be invariant with respect
to that transformation. Translation and rotation symmetries will manifest as the partial derivative of the Lagrangian
with respect to the direction or rotation being zero. We say that the Lagrangian is invariant to those transformations
and, if the generalized coordinates are chosen appropriately, then that coordinate will not appear in the Lagrangian
(although its time derivative will likely appear). Such coordinates are termed cyclic coordinates.
The conjugate momentum is defined as a partial derivative of the Lagrangian,
DR
∂L
pa ≡ ,
∂ q̇a
and is “conjugate” to the corresponding coordinate, qα . The ELE can then be written as
∂L
ṗα = .
∂qα
ṗa = 0
and so pa is a constant of the motion – the EOM corresponding to qa is ṗa = 0 or, equivalently, pa is constant.
This is a special case of Nöther’s Theorem:
If a system has a continuous symmetry then there is a choice of generalized coordinate, qa for which the
Lagrangian is invariant, ∂q
∂L
a q̇a , the corresponding conjugate momentum is conserved.
= 0, and pa = ∂∂L
T
the horizontal bar (i.e. the plane of this page). θ
This problem has two degrees of freedom. Use x and θ, as
indicated on the diagram, for your generalized coordinates. ŷ m
x̂
(a) Using the point O as your origin, what are the position vectors, ~rM & ~rm , of the two masses as functions of the
generalized coordinates in terms of the Cartesian unit vectors, x̂ & ŷ? What are the corresponding first and second
time derivatives? (b) Using the point O as your origin, what is the position vector, ~rm , of the mass m as a function
of the generalized coordinates in terms of the polar unit vectors, r̂ & θ̂? What are the corresponding first and second
time derivatives? (c) What are the kinetic and potential energies, T & V , for this system? (d) Find the equations of
AF
motion for this system. (e) What are the equations of motion for this system in the M m limit? (f) Identify any
symmetries of this system which lead to an invariance in the Lagrangian (i.e. any cyclic coordinates) along with the
corresponding conserved quantities (i.e. the corresponding conjugate momentum).
Solution
(a) For the mass M I have
Using r̂ = sin θx̂ − cos θŷ I have for the mass m: ~rm = xx̂ + ar̂ so
T
into my solutions from part (a).
(c) For the mass M I have TM = 21 M ẋ2 and VM = 0. For the mass m I can use the velocity vectors from either part
(a) or part (b) to find Tm and Vm (I get the same result in either case as I must). I have
and
Vm = −m~g · ~rm = +mg ŷ · ((x + a sin θ)x̂ − (a cos θ)ŷ) = −mga cos θ.
So
AF
(d) I have a cyclic coordinate, ∂L
∂x = 0 so
d
dt
T = 21 M ẋ2 + 21 m(ẋ2 + a2 θ̇2 + 2aẋθ̇ cos θ)
V = −mga cos θ.
Å
∂L
∂ ẋ
so
d
dt
∂L
Å
∂ θ̇
ã
=
d Ä 2
dt
ã
=
∂L
∂θ
d Ä
dt
ä
ä
M ẋ + mẋ + maθ̇ cos θ = 0
(e) I have → 0 so I can divide the first EOM by M and take this limit, leaving ẍ = 0 , and so the second is
m
M
g
θ̈ = − sin θ . This, of course, is the equation of a simple pendulum with a fixed support.
a
(f) The system is unchanged by a displacement in the horizontal direction so x is a cyclic coordinate and the
Lagrangian is invariant under changes in x, ∂L ∂x = 0. Consequently, dt ∂ ẋ = 0 and the conjugate momentum,
d ∂L
Solution
T
where r2 ≡ (x2 − x1 )2 + (y2 − y1 )2 + (z2 − z1 )2 is the distance between the atoms.
The velocities are ~r˙i = x̂ẋi + ŷ ẏi + ẑ żi so the kinetic energy is
AF so
L = 21 m1 (ẋ21 + ẏ12 + ż12 ) + 12 m2 (ẋ22 + ẏ22 + ż22 )
~+
~r1 = R
µ
m1
~˙ + µ ~r˙
~r˙1 = R
m1
~r ~−
& ~r2 = R
~˙ − µ ~r˙
& ~r˙ = R
m2
µ
m2
2
~r
~˙ = x̂Ẋ + ŷ Ẏ + ẑ Ż and ~r˙ = ṙr̂ + rθ̇θ̂ + rφ̇ sin θφ̂. The kinetic energy is then
with R
m1 m2
where M ≡ m1 + m2 is the total mass and µ ≡ m1 +m2 is the reduced mass.
4
This is awesome! I have ∂L
∂X = ∂L
∂Y = ∂L
∂Z = ∂L
∂φ = 0, and of the six degrees of freedom, four are conserved.
X
4 Thismethod pays even more dividends for a relativistic system when a centre of momentum system is used, M R~˙ = m1 ~
r˙1 + m2 ~
r˙2 = 0, in
˙~
which a boost of −R is applied to the system – this makes the problem one dimensional with equal and opposite momenta).
8.4. FURTHER PROBLEMS 119
T
or radius 2a and a small mass m is attached to the other
end, as shown below. The tension in the string increases ~r m
linearly with its extension from its unstretched length with y
a spring constant k. Assume that gravity acts vertically, that
the string is taught along the surface of the hemisphere, and
that the mass m remains in contact with the hemisphere at
all times. x
(a) How many degrees of freedom, f , are there for this problem? What will you use for your generalized coordi-
nates? (b) Write the position, ~r, and velocity, ~r˙ , in terms of your generalized coordinates. Don’t forget that the radius
of the hemisphere is 2a and not a. (c) Write the kinetic and potential energies, T & V and so form the Lagrangian,
AF
L = T − V . Don’t forget that the radius of the hemisphere is 2a and not a. (d) Use the ELE to find the equations of
motion for this system. (e) Identify any symmetries of this system which lead to an invariance in the Lagrangian (i.e.
a cyclic coordinate) along with the corresponding conserved quantities (i.e. the corresponding conjugate momentum).
Give a physical interpretation for any conserved quantities. (f) With the right initial conditions it is possible for the
mass m to travel in an horizontal circle just barely in contact with the sphere (ie the contact force is zero). What is
the angular momentum, L,
equations of motion.
~ for this situation? BONUS: Can you express your result, L
in terms of both cartesian and polar unit vectors. (h) Calculate the virtual work done by the applied forces, δW (a) .
(i) Calculate the virtual work done by the inertial force, δW (i) = −m~r¨ · δ~r. (j) Use d’Alembert’s principle to find the
Consider a bead of mass m that is free to slide along a circular hoop of radius a which is oriented in a vertical
plane. (a) Using the y-coordinate as your generalized coordinate write the position, ~r, and velocity, ~r˙ , in terms of the
generalized coordinate. (b) Write the kinetic and potential energies, T & V and so form the Lagrangian, L = T − V .
(c) Find the EOM for this system. (d) Show that you get the expected results for your EOM for the case y = 0 and
give a brief explanation as to why this is the expected result.
(a) Using cylindrical coordinates with the x-z plane vertical as shown, write the position, ~r, and velocity, ~r˙ , in terms
of both cartesian and cylindrical unit vectors. (Hint: The formula sheet may be helpful.) (b) Write the kinetic and
potential energies, T & V and so form the Lagrangian, L = T − V . (c) Find the equations of motion for this system.
(d) Identify any symmetries of this system which lead to an invariance in the Lagrangian along with the corresponding
conserved quantities. Give a physical interpretation for any conserved quantities. (e) Verify that you get the expected
result for φ̈ in the small angle approximation for angles in which φ is close to π, and give a brief explanation as to
Xn
∂~r
why it is the expected result. (f) Calculate the virtual displacement, δ~r = δqa , for this system. (g) Calculate
a=1
∂qa
the virtual work done by the applied force, δW (a) = m~g · δ~r. (h) Calculate the virtual work done by the inertial force,
δW (i) = −m~r¨ · δ~r. (i) Use d’Alembert’s principle to find the equations of motion. (j) The system is now tilted so that
the axis of the cylinder makes an angle α with the horizontal such that the x-z plane remains vertical and the y-axis
remains horizontal (i.e. it is rotated by α about the y-axis). Find the new equations of motion. You may use either
d’Alembert’s or Lagrange’s method. Verify that your result reduces to your earlier result in the α → 0 limit. (k) Verify
that you get the expected result for z̈ in the φ = π, φ̇ = 0 case and give a brief explanation as to why each is the
expected result.
120 CHAPTER 8. LAGRANGIAN MECHANICS
x
z
~g
α
T
m
y ~r
Solution
AF There are two degrees of freedom and I choose the lengths of the springs, x1 & x2 , as the generalized coordinates.
The positions of the two masses are then
V = 21 k1 (x1 − `1 )2 + 21 k2 (x2 − `2 )2 .
1
ï ò
m2 −m2
M−1 = ,
m1 m2 −m2 m1 + m2
so
ï ò
¨ + M−1 K~x + M−1 U◦ = 0 ¨ = − k1 /m1 −k2 /m2
M−1 M~x ⇒ ~x ~x − M−1 U◦ ,
−k1 /m1 k2 /µ
m1 m2
where µ ≡ m1 +m2 is the reduced mass.
and I can solve for the accelerations,
k1 k2
ẍ1 = − (x1 − `1 ) + (x2 − `2 )
m1 m2
m1 − m2 k1
ẍ2 = − k2 (x2 − `2 ) + (x1 − `1 )
m1 m2 m1
X
8.4. FURTHER PROBLEMS 121
T
m1
two generalized coordinates – state explicitly what you choose as coordinates and
then find the EOM using the ELE. (c) Now consider the double pendulum as shown. θ2 `2
Using θ1 & θ2 as generalized coordinates find the EOM using the ELE. (d) For your m2
answer to (c) use the small angle approximation, cos θ ≈ 1 & sin θ ≈ θ, to find the
equations of motion in this limit.
AF
DR
122 CHAPTER 8. LAGRANGIAN MECHANICS
T
AF
DR
T
Chapter 9
AF
9.1
9.1.1
Effective Potentials
The Lagrangian in an Accelerating Frame
With the expression for the displacements in the IRF,
ä2
T = 21 m~r˙ 0 · ~r˙ 0 = 21 m ~r˙O0 + ~r˙
Ä
02
= 12 mṙO + m~r˙O0 · ~r˙ + 21 mṙ2
DR
The first term is a function of time only and so it will not contribute to the EOM (i.e. its partial derivatives with respect
to ~r and ~r˙ are both zero). The second term is
d Ä˙0 ä
m~r˙O0 · ~r˙ = m ~r · ~r − m~r · ~r¨O0
dt O
and the complete time-derivative term won’t contribute to the EOM (see ??). The Lagrangian is then
where
Veff (~r ) ≡ m~r · ~r¨O0 + V (~r )
is the effective potential.
The EOM in the NIRF is then
d ∂L d
= (mṙ) = mr̈
˙
dt ∂~r dt
∂L ~ eff
= −∇V
∂~r
⇒ mr̈ = −∇V ~ eff = m~r¨O0 − ∇V.
~
123
124 CHAPTER 9. LAGRANGIAN MECHANICS OF ROTATING & OSCILLATING SYSTEMS
T
9.1.2 The Lagrangian in a Rotating Frame
Consider an IRF and a NIRF whose origins coincide (i.e. ~rO0 = 0) so that the relation between the velocities in the IRF
and NIRF is
~r˙ 0 = ~r˙ + ω
~ × ~r.
The kinetic energy is then
T = 21 m~r˙ 0 · ~r˙ 0 = 21 m(~r˙ + ω
~ × ~r )2
= 1 mṙ2 + 1 m(~
2 2 ω × ~r )2 + m~r˙ · (~
ω × ~r ),
AF
so the general Lagrangian is
m~r¨ + mω
~˙ × ~r + m~
ω × ~r˙ = −
⇒
p~ =
m~r¨ = −
∂L
∂~r˙
p~˙ = ∇L
~
∂V
∂~r
∂V
∂~r
ω × ~r )2 + m~r · (~r˙ × ω
= 21 mṙ2 + 12 m(~
ω × ~r )2 + m~r˙ · (~
L = 21 mṙ2 + 12 m(~
= m~r˙ + m~
ω × ~r ) − V (~r )
~ ) + m~r˙ × ω
ω × (~r × ω
+ m~
ω × (~r × ω
+ m~
~ and F~cf = m~
~
~ ) + 2m~r˙ × ω ~˙ × ~r.
~ − mω
∂H
p~˙ = − = p~ × ω ~
~ − ∇V
∂~r
∂H p~
~r˙ = = −ω~ × ~r.
T
∂~p m
AF
angle α to the vertical. The wire is rotating about its vertical axis with
angular frequency ω and gravity acts downwards. (a) Using spherical
coordinates (with constraints1 θ = α, a constant, and φ̇ = ω), find the
EOM using the ELE. (b) Express your result in terms of an effective po-
tential in the rotating system, mr̈ = − dr
d
Veff . (c) Is energy conserved
for this system and is L = 2 mṙ − Veff ? (d) Plot the effective potential
Solution
1 2
(a) There is one degrees of freedom for this problem. In spherical coordinates r, the distance of the bead from the
pivot, is my generalized coordinate. The position and velocity of the bead are
~r =rr̂
0 ω α
α
m
∂L
ṗ = ⇒ mr̈ = mrω 2 sin2 α − mg cos α
∂r
(b) I have
d
mr̈ = − − 12 mω 2 r2 sin2 α + mgr cos α
dr
126 CHAPTER 9. LAGRANGIAN MECHANICS OF ROTATING & OSCILLATING SYSTEMS
1.0
0.5
T
Veff /mg`
0.0
−0.5
−1.0
AF T = 21 ω~ · L, 2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
r/`
Figure 9.1: A plot of the effective potential, Veff (r), with α = π/4, for Example
ω × ~r)2 ,
Vcf = − 12 m(~
√ 9.1. The unit length for this problem is
` ≡ g/ω 2 and the unit energy is mg`. There is an unstable equilibrium at r◦ = 2g/ω 2 , and a stable equilibrium at r = 0.
or, alternatively, I = m(r sin α)2 is the moment of inertia of the bead about the ẑ axis so, with Lz = Iω and
~ this term is − 1 Iω 2 , the negative of the rotational kinetic energy.
Note that energy is not conserved for this system – there is an external agency that is making the wire rotate at
angular frequency ω(t).
(c) Energy is not conserved for this system – there is an external agency that is making the wire rotate at angular
frequency ω(t) – and L = 21 mṙ − Veff .
DR
(d) The code below produces the plot in Figure 9.1.
1 ##################### constants
2 m = 1 # mass, m
3 g = 10 # free-fall acceleration due to gravity, g
4 w = 1 # rotational frequency, ω
5 a = pi / 4 # angle, α = π/4
6 ell = g / w**2 # unit of length, ` = g/ω 2
7 Vo = m*g*ell # unit of energy, V◦ = mg`
8 ##################### the potential
9 r = linspace( 0, 4, 100 ) * ell # domain, 0 ≤ r ≤ 4`
10 V = -0.5*m*( w*r*sin(a) )**2 + m*g*r*cos(a) # Veff (r) = − 12 m(ω r sin α)2 + mgr cos α
11 plot( r/ell, V/Vo )
There is also a stable equilibrium at r = 0 (i.e. r ≥ 0 for this problem so this is a local minimum in the
effective potential).
If r < r◦ then the bead will slide down the wire towards the pivot (i.e. towards the stable equilibrium and
away from the unstable equilibrium). If r > r◦ then the bead will slide up the wire (i.e. away from the unstable
equilibrium). X
T
Example 9.2 Bead on Wire II.
Consider a bead of mass m that is free to slide on a wire that makes an angle α to the vertical. The wire is free to
rotate about its vertical axis and gravity acts downwards. (a) Using spherical coordinates (with constraint θ = α, a
constant), find the EOM using the ELE. (b) Your Lagrangian should be invariant with respect to φ so that you have
a constant of the motion Lz ≡ ∂L ∂ φ̇
. Use this EOM to eliminate φ̇ from the other EOM so that it contains only the
single generalized coordinate r, and its time derivatives. (c) Express your result in terms of an effective potential,
mr̈ = − dr d
Veff (i.e. as a function of r and its derivatives only). (d) Is energy conserved for this system and is
L = 2 mṙ − Veff ? (e) Calculate the position, r◦ , of the stable equilibrium as well as U◦ ≡ Veff (r◦ ) and the effective
1 2
AF
spring constant, keff ≡ drd2
2 Veff (r◦ ), in terms of r◦ (i.e. eliminate Lz in favour of r◦ ). (f) Plot the effective potential as
a function of the generalized coordinate, r, as well as the parabola U (r) = Veff (r◦ ) + 21 keff (r − r◦ )2 . (g) In terms of
r◦ and other constants, what is the frequency of small oscillations?
Solution
(a) There are two DOF for this problem. In spherical coordinates r, the distance of the bead from the pivot, and φ, the
azimuthal angle, are my generalized coordinates. The position and velocity of the bead are
The first term is the kinetic energy for the motion along the wire and the second term is the kinetic energy for the
rotational motion perpendicular to the wire.
The gravitational potential energy is
DR
V = −m~g · ~r = −m(−gẑ) · rr̂ = mgr cos α
∂L
ṗr = ⇒ mr̈ = mrφ̇2 sin2 α − mg cos α
∂r
∂L
ṗφ = ⇒ mr2 φ̇ sin2 α = Lz , a constant
∂φ
Lz
(b) I have φ̇ = m(r sin α)2 so
ã2
Lz L2z
Å
mr̈ = mr sin2 α − mg cos α = − mg cos α
m(r sin α)2 mr3 sin2 α
128 CHAPTER 9. LAGRANGIAN MECHANICS OF ROTATING & OSCILLATING SYSTEMS
(c) I have
d L2z
Å ã
mr̈ = − + mgr cos α
dr 2mr2 sin2 α
L2z
⇒ Veff (r) = + mgr cos α.
T
2m(r sin α)2
The first term can be viewed as the centrifugal potential for this problem; I = m(r sin α)2 is the moment of inertia
of the bead about the ẑ axis so, with Lz = I φ̇, this term is 21 I φ̇2 , the rotational kinetic energy).
(d) Energy is conserved for this system and L 6= 12 mṙ − Veff .
(e) I have a stable, dynamic equilibrium at
L2z L2z
mr̈ =0= − mg cos α ⇒ r◦3 = ,
r=r◦ mr◦3 sin2 α m2 g cos α sin2 α
AF so
keff =
dr2 r=r◦
=
r◦3 m2 g cos α sin2 α
=
2m(r sin α)2
and the effective spring constant at the equilibrium point, r◦ , is
d2 Veff
−
d
dr
mr̈
mr◦4 sin2 α
=
=
+ mgr cos α = 23 mgr◦ cos α.
−
−3L2z
mr◦4 sin2 α
3mg cos α
r◦
.
U (r) = 23 mgr◦ cos α 1 + (1 − r/r◦ )2 .
(g) Near r = r◦ the EOM can be approximated using the first few terms in the Taylor expansion of the effective
potential,2
!
d d U◦ d 0 d 2 keff
*
mr̈ ≈ − U (r) = − Veff
*
+ Veff
(r − r◦ ) + 21 Veff
*
(r − r◦ )2 + · · ·
dr dr
dr r=r◦ dr2
r=r◦ r=r◦
r − r◦
= −keff (r − r◦ ) = −3mg cos α
r◦
2 The term proportional to (r − r ) is missing because its coefficient, d V
◦ dr eff r=r
= − mr̈ = 0, is the equilibrium condition.
◦ r=r◦
9.1. EFFECTIVE POTENTIALS 129
4.0 1.510
3.5
3.0 1.505
T
2.5
Veff /mg`
Veff /mg`
2.0 1.500
1.5
1.0 1.495
0.5
0.0 1.490
AF1
2
k eff (r −
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
r ◦ )2
, for
r/`
Figure 9.2: A plot of the effective potential, Veff (r), with α = π/4, as well
Example 9.2. There is a stable
ω=
equilibrium
…
g
r◦
at
3 cos α.
r ◦ =
1.30 1.35
p
1.40 1.45
r/`
1.50
which is a simple harmonic oscillator about the equilibrium point r = r◦ with frequency ω 2 = keff /m so
Solution
T
(a) For Example 9.1 I have
L = 21 mṙ2 + 21 mr2 ω 2 sin2 α − mgr cos α,
so ∂L
∂t = (mr2 ω sin2 α)ω̇ is zero if ω is a constant.
The conjugate momentum is p = ∂L
∂ ṙ = mṙ so
AF
(b) For Example 9.2 I have
so ∂L
∂t = 0.
The conjugate momenta are pr =
= T + V − m(rω sin α)2 ,
= Tr + Veff
∂L
∂ ṙ = mṙ and pφ = ∂L
∂ φ̇
= mr2 φ̇ sin2 α ≡ Lz (a constant) so
h = mṙ2 + mr2 φ̇2 sin2 α − 12 mṙ2 − 12 mr2 φ̇2 sin2 α + mgr cos α
= 21 mṙ2 + 12 mr2 φ̇2 sin2 α + mgr cos α
= T + V.
ã2
Lz
Å
= 21 mṙ2 + 12 m(r sin α)2 + mgr cos α
m(r sin α)2
DR
L2z
= 21 mṙ2 + + mgr cos α
2m(r sin α)2
= Tr + Veff .
Solution
(a) I have two degrees of freedom and I choose to use spherical coordinates, {r, θ, φ}, with the constraint r − ` = 0.
(b) I have
L = 21 m(`θ̇)2 + 21 m(`φ̇ sin θ)2 − mg` cos θ.
T
The first term is the kinetic energy in the θ̂ direction, the second is the kinetic energy in the φ̂ direction, and the
third is the negative of the gravitational potential energy.
(c) The conjugate momenta are
pφ
pφ = m(` sin θ)2 φ̇ ⇒ φ̇ =
m(` sin θ)2
pθ
pθ = m`2 θ̇ ⇒ θ̇ =
m`2
AF
(d)
(e) I have
(f)
(g)
so that the remaining EOM is
d
dt
Lz
m(` sin θ)2
⇒
ã2
φ̇ =
+ mg` sin θ =
Lz
m(` sin θ)2
L2z cos θ
m`2 sin3 θ
+ mg` sin θ
(h)
p 2 Å
pφ
ã2
θ
h(θ, φ, pθ , pφ ) = 12 m`2 + 1
2 m(` sin θ)2
+ mg` cos θ
m`2 m(` sin θ)2
p2θ p2φ
= + + mg` cos θ
2m`2 2m(` sin θ)2
(i) I have
∂h pθ
= = θ̇
∂pθ m`2
∂h pφ
= = φ̇
∂pφ m(` sin θ)2
∂h p2φ cos θ
− = + mg` sin θ = ṗθ
∂θ m`2 sin3 θ
∂h
− =0 = ṗφ
∂φ
132 CHAPTER 9. LAGRANGIAN MECHANICS OF ROTATING & OSCILLATING SYSTEMS
(j) I have
h(θ, φ, θ̇, φ̇) = 12 m(`θ̇)2 + 12 m(`φ̇ sin θ)2 + mg` cos θ
so
∂h
− = 0, unchanged
T
∂φ
∂h
− = −m(`φ̇)2 sin θ cos θ + mg` sin θ
∂θ
p2φ cos θ
= − 2 3 + mg` sin θ 6= ṗθ .
m` sin θ
The difference is due to this ∂h
∂θ partial derivative having {φ, θ̇, φ̇} constant while in the previous part I had
{φ, pθ , pφ } constant.
X
AF
Exercise 9.3 Cubical Snoozing.
(a) Consider a top of mass m made from a uniform cube of side a with a light spindle of length 2` which passes through
the body diagonal such that the centres of the cube and spindle coincide (ie, the spindle passes from one corner of the
cube, through the centre and out the farthest corner). What are the principle axes and principle moments of inertia for
this top about an origin at the end of the spindle? (b) Starting from a fixed coordinate system, the IRF, with ê03 vertical
draw a diagram showing the transformation to the primed principle axis coordinate system using Euler’s angles, φ, θ
& ψ. (c) Write each of φ, ~˙ θ~˙ & ψ
~˙ in terms of {ê1 , ê2 , ê3 }, the NIRF. (Note that due to the axial symmetry we do not
require the principle axes to rotate with the top but rather that ê3 remain along the axis of the top and ê2 remain in
~ 3 -~
the L-ê ω plane. Hint #1: What is ψ for this choice of of primed coordinates? Hint #2: Your diagram above may be
helpful.) (d) What is the kinetic energy, T , of the top using Euler’s angles as generalized coordinates? (e) What is the
potential energy, V , of the top using Euler’s angles as generalized coordinates? (f) Form the Lagrangian and identify
the cyclic coordinates. (g) What are the corresponding conserved conjugate momenta? Give a physical interpretation
of each and introduce suitable constants. (h) What is the equation of motion for the remaining non-cyclic coordinate?
(i) Use your constants from (g) to eliminate the other coordinates from your equation of motion. (j) Now assume that
the remaining angle is small and use the small angle approximation (ie, sin θ ∼ θ, 1 − cos θ ∼ 21 θ2 & cos θ ∼ 1)
to obtain a linearized equation. (Hint: There should be a simple relationship between the two constants introduced
above.) (k) What condition is necessary for the top to be stable in the upright orientation?
DR
9.2 Velocity Dependent Potentials
Example 9.5 Electromagnetic Interaction I.
Usually we can only form a Lagrangian when the forces are conservative, F~i = −∇ ~ i V . However, for the important
~
case of the electromagnetic interaction, for which we have the Lorentz force, F = eE ~ + e ~r˙ × B,
~ we can still form a
c
Lagrangian if we write the electric and magnetic fields in terms of the scalar and vector potentials,
~
E ~ − 1 ∂A
~ = −∇φ & ~ =∇
B ~ × A.
~
c ∂t
The Lagrangian for a particle, mass m and charge e, moving in electric and magnetic fields is
T
matrix calculus notation. Replace the total time derivative of the vector potential with4
~ ~ ~
~˙ r, t) = ∂ A · d~r + ∂ A = ~r˙ · ∇ ~ + ∂ A = ẋi ∂i Aj êj + ∂t Aj êj .
Ä ä
A(~ ~ A
∂~r dt ∂t ∂t
(b) Calculate the gradient of the Lagrangian, ∇ ~ ~r L using matrix calculus notation. (c) Use the ELE in matrix calculus
notation to find the EOM. (d) Show that your result in (c) is the same as the Lorentz force (you may require a vector
identity to do so). (e) Repeats parts (a)-(d) using index notation.
AF
Solution
(a) Starting with the Lagrangian,
(b) I have
~ + e∇
~ = −e∇φ
∇L
c
c
Ä
~ ~r˙ · A
Ä ä
~ ,
e
L = 12 m~r˙ · ~r˙ − eφ + ~r˙ · A,
∂L
c
e~
= m~r˙ + A,
c
ä
~ A
~
~
∂t
å
~ + ∂A .
d ∂L ∂L
= ,
˙
dt ∂~r ∂~r
or
DR
p~˙ = ∇L
~
~
Ç å
e Ä ˙ ~ ä ~ ∂A ~ + e∇
m~r¨ + ~ ~r˙ · A
Ä ä
~r · ∇ A + = −e∇φ ~
c ∂t c
~
Ç å
¨ ~ 1 ∂A e Ä ~ Ä ˙ ~ä Ä ˙ ~ ä ~ä
m~r = e −∇φ − + ∇ ~r · A − ~r · ∇ A
c ∂t c
and
~
Ç å
¨ 1 ∂A e ~ + e ~r˙ × B
~
m~r = e −∇φ − + ~r˙ × (∇
~ × A)
~ = eE ~
c ∂t c c
the Lorentz force.
4 ~
∂A
The quantity ∂~
r
= [∂i Aj ] is the Jacobian matrix, the matrix of partial derivatives of a vector function.
134 CHAPTER 9. LAGRANGIAN MECHANICS OF ROTATING & OSCILLATING SYSTEMS
O0
~rC0 θ
C 0
~r˙ C
α
ê02 ~rQC
T
ê01 Q ~g
α
mẍi + e
c ẋj ∂j Ai + e
pi =
ṗi
c ∂ t Ai
= ∂i L
e
= mẋi + Ai ,
c
∂i L = −e∂i φ + ec ẋj ∂i Aj ,
= −e∂i φ + ec ẋj ∂i Aj
mẍi = e −∂i φ − 1c ∂t Ai + ec ẋj (∂i Aj − ∂j Ai ).
axis of the cylinder is perpendicular to the steepest decent (Figure 1.14).6 (a) Describe the IAOR. (b) Use the rolling
constraint, ~r˙Q0 = 0, to find the relationship between the the rotation of the cylinder, θ, and the displacement of the
centre of the cylinder, rC 0
. Assume that θ = 0 when rC 0
= 0. (c) Put the body origin at Q, use θ for the generalized
coordinate, and find the Lagrangian, L1 (θ, θ̇). (d) Put the body origin at C, use rC 0
for the generalized coordinate, and
find the Lagrangian, L2 (rC , ṙC ). (e) Use the Lagrange multiplier, F , and the constraint equation from part (a) to find
0 0
T
the Lagrangian,7 L3 (~q, ~q˙ where ~q = (rC0
, θ, F ). (f) Use L3 to find the EOMs. (g) Solve the EOMs for F and give a
physical interpretation of this Lagrange multiplier. (h) Solve the EOMs for rC 0
(t). Take rC 0 0
= ṙC (0) = 0 as your BC.
AF
DR
6
P
In Example 1.10 I did this problem using the rotational version of N 2 L, I θ̈ = N , and summed the torques about the point Q so that the
constraint force made no contribution,
2g
( 21 ma2 + ma2 )θ̈ = mga sin α ⇒ θ̈ = sin α.
3a
T
A bead of mass m slides along a smooth wire in the shape of a parabola, z = Ax2 . The wire is rotating about the ẑ axis
with angular frequency ω and gravity acts downwards, ~g = −gẑ. (a) Using cylindrical coordinates (with constraints8
z = Ax2 , and φ̇ = ω), find the EOM using the ELE. (b) Express your result in terms of an effective potential
in the rotating system, mr̈ = − dr d
Veff . (c) Plot the effective potential as a function of the generalized coordinate.
(d) Describe the resulting motion.
An analogy between classical mechanics and electrical circuits can be made using the substitutions:
2
1
2 mv → 12 Lq̇ 2
q2
AF
where q is the charge and q̇ is the current.
−m~g · ~r → 12 CV 2 =
1
2 βv
2
→ 12 Rq̇ 2
2C
DR
8
R
Note that the second constraint is semi-holonomic because it can be integrated, φ = dt ω.
T
Chapter 10
Calculus of Variations
AF
10.1
Resources
Thornton & Marion – Ch. 6
Boas – §9.1-7
Hamilton’s Principle
A more general formulation of the Euler-Lagrange equation is that the “true” path, q(t), is “stationary with respect
to variations in the path”, and is a special case of Hamilton’s Principle, which states that the path taken between two
points, ~rA and ~rB , is such that the action,
S[qi (t)] ≡
Z tB
dt L(qi (t), q̇i (t), t)
Week 10
tA
is stationary,
DR
δS[qi (t)] = 0,
where L(qi (t), q̇i (t), t) is the Lagrangian function, L = T − V , expressed in terms of the generalized coordinates,
qi (t). The action, S[qi (t)], is a functional (i.e. its argument is the set of functions, qi (t)) which returns a scalar.
Example 10.1 Shortest Distance.
Consider the problem of finding the shortest distance between two points in a plane. If we use a Cartesian coordinate
system the the distance between two adjacent points, ~r(A) = x(A) êx + y (A) êy & ~r(B) = x(B) êx + y (B) êy is
Z r (B)
~ Z r (B)
~ p Z x(B)
ã2
dy
Å
I= ds = dx2 + dy 2 = dx 1+ .
r (A)
~ r (A)
~ x(A) dx
I know wish to find the path, y = f (x), that minimizes the quantity I. Consider the path y = F (x) = f (x)+η(x)
which deviates by a small amount, η(x), from the shortest path, f (x), with η(x(A) ) = η(x(B) ) = 0 so that this path
also passes through the two end points.
My integral for the path length is now a function of . I have
Z x(B) »
2
I() = dx 1 + (F 0 ) ,
x(A)
where F 0 (x) = d
dx F (x).
137
138 CHAPTER 10. CALCULUS OF VARIATIONS
Solution
I wish to minimize the path length with respect to so I have dI
d = 0 when = 0. Differentiating,
Z Z
dI x(B)
d » 2
x(B)
F0 dF 0
d
= dx 1 + (F 0 ) = dx » .
T
d x(A) d x(A) 1 + (F 0 )
2
dF 0
Since F 0 (x) = f 0 (x) + η 0 (x) and d = η 0 (x) I can see that putting = 0 results in F 0 (x) = f 0 (x), yielding
Z x(B)
dI f 0 (x) η 0 (x)
=0 = dx .
d =0
»
2
x(A) 1 + (f 0 )
AF so that
giving
dI
d =0
=0 = »
d
dx1
Ñ
du =
f 0 (x)
1 + (f )
»
0
f 0 (x)
u= »
1 + (f 0 )
1 + (f 0 )
d
dx
η(x)
2
é
Ñ
x(B)
(A)
x
=0 ⇒ »
»
−
2
0
f (x)
1 + (f 0 )
Z x(B)
x(A)
dx η(x)
f 0 (x)
&
d
dx
1 + (f 0 )
2
é
≡C
dx
»
dv = η 0 (x) dx
1
f
+
0
&
(x)
(f 0 )2
v = η(x)
Since η(x(A) ) = η(x(B) ) = 0 the boundary term is zero and, since η(x) is an arbitrary function, I must have1
a
θ1
T
land C
water
b
θ2
x B
c
AF
Figure 10.1: Jill wishes to travel from point A to point B as quickly as possible. She can run at speed v1 and swim at speed
v2 . At what point, C, should she enter the water in order to minimize her travel time?
S=
Z
x(A)
x(B)
dx L(x, y, y 0 ).
in which L(x, y, y 0 ) is some given function. Let y = F (x) = f (x) + η(x) be a function which deviates by a small
amount, η(x), from the function which minimizes S, f (x), with η(x(A) ) = η(x(B) ) = 0 so that this path also passes
through the two end points.
My integral is now a function of . I have
Z x(B)
S() = dx L(x, F, F 0 ),
x(A)
where F 0 (x) = dx
d
F (x) = f 0 (x) + η 0 (x).
DR
I wish to minimize the integral with respect to so I have dS
d = 0 when = 0. Differentiating,
Z x(B) Z x(B)
dS d ∂L dF ∂L dF 0
Å ã
= (L(x, F, F 0 )) =
dx dx +
d x(A) d x(A) ∂F d ∂F 0 d
Z x(B)
∂L ∂L 0
Å ã
= dx η(x) + η (x) .
x(A) ∂F ∂F 0
∂L
u= & dv = η 0 (x) dx
∂F 0
so that
d ∂L
Å ã
du = dx & v = η(x)
dx ∂F 0
giving
Z x(B) x(B) Z x(B)
∂L 0 ∂L d ∂L
Å ã
dx η (x) = η(x) − dx η(x) .
∂F 0 ∂F 0 (A) dx ∂F 0
x(A) x x(A)
140 CHAPTER 10. CALCULUS OF VARIATIONS
Since η(x(A) ) = η(x(B) ) = 0 the boundary term is zero and I am left with
Z x(B)
dS ∂L d ∂L
Å Å ãã
=0 = dx − η(x).
d =0 x(A) ∂y dx ∂y 0
Since η(x) is an arbitrary function (and F |=0 = f = y the unknown function), I must have2
T
d ∂L ∂L
Å ã
− = 0.
dx ∂y 0 ∂y
This is Euler’s equation. The result of calculating the various derivatives in this equation is a second-order ODE. When
calculating the derivatives care must be taken to distinguish between the full and partial derivatives.
The function L(x, y, y 0 ) which minimizes the integral
Z x(B)
S= dx L(x, y, y 0 ).
AF
is the solution of the second-order ODE
Solution
I have
dy
where y 0 = dx
∂L
∂y = 0 and
S=
r (B)
~
r (A)
~
ds =
~
»
~
r (A)
d ∂L
Å
dx ∂y
r (B)
0
x(A)
p
−
∂L
∂y
= 0.
Consider the problem of finding the shortest distance between two points in a plane. If we use a Cartesian coordinate
system the the distance between two adjacent points, ~r(A) = x(A) êx + y (A) êy & ~r(B) = x(B) êx + y (B) êy is
Z Z
dx2 + dy 2 =
2
Z
∂L y0
=
∂y 0
»
2
1 + (y 0 )
DR
so the Euler equation gives
Ñ é
d ∂L d y0 y0
Å ã
0
=0= » ⇒» = C,
dx ∂y dx 1 + (y 0 )
2
1 + (y 0 )
2
Solution
p dy
p
(a) I have ds = dx 1 + (y 0 )2 with y 0 ≡ dx so L = 1
x 1 + (y 0 )2 , the partial derivatives are
∂L y0 ∂L
= p & =0
∂y 0 x 1 + (y 0 )2 ∂y
T
and Euler’s equation gives
Ç å
d y0
p =0
dx x 1 + (y 0 )2
y0
⇒ p =a, a constant
x 1 + (y 0 )2
⇒ (y 0 )2 =a2 x2 (1 + (y 0 )2 )
d
dx
Ç
p
dy
dx
∂L
∂y 0
xy 0
⇒ y0 = ± p
=
⇒y =±
1 + (y 0 )2
xy 0
p
p
1 + (y 0 )2
å
=0
»
ax
1 − (ax)2
1 − (ax)2 + b,
&
∂L
∂y
=0
b a constant
xy 0
⇒p =a, a constant
1 + (y 0 )2
⇒ (xy 0 )2 =a2 (1 + (y 0 )2 )
DR
⇒ (1 − (a/x)2 )(y 0 )2 =(a/x)2
a/x a
⇒ y0 = ± p = ±√
1 − (a/x)2 x2 − a2
d ∂L
Å ã
=0
dx ∂y 0
and it was particularly easy to integrate this second order ODE into the first order ODE
∂L
= a,
∂y 0
142 CHAPTER 10. CALCULUS OF VARIATIONS
y
yf
y(x)
yi
T
x
xi xf
Figure 10.2: Find the curve, y(x), that connects two points such that the surface of revolution has minimum area.
AF
in which a is a constant. This equation is called the first integral
If, instead, I have ∂L
where x0 ≡ dy .
dx
∂x = 0 then,
I then have
S=
instead
d
dy
Z
of
Ç
writing
xf
xi
ds
ds =
=
p of Euler’s p
p
dx
dx L(y, y 0 ) =
∂ L̃
∂x0
å
2 + dy 2 =
∂ L̃
− =0
∂x
0
Z
yi
»
yf
equation.
1 + (y 0 )2 dx I can use
dx2 + dy 2 = 1 + (x0 )2 dy
dy
y0
⇒
L(y, y 0 ) =
∂ L̃
∂x0
= a,
Z
yi
yf
dy L̃(y, x0 )
a constant
Solution
(a) I have for the area of a ring of width ds and radius y,
p »
dA = (2πy) ds = (2πy) dx2 + dy 2 = (2πy) 1 + (y 0 )2 dx
p
so L(x, y, y 0 ) = 2πy 1 + (y 0 )2 .
T
d
Ç
2πyy 0
å »
p = 2π 1 + (y 0 )2
dx 1 + (y 0 )2
(y 0 )2 + yy 00 (yy 0 )(y 0 y 00 ) »
p − = 1 + (y 0 )2
1 + (y 0 )2 (1 + (y 0 )2 )3/2
(1 + (y 0 )2 )((y 0 )2 + yy 00 ) − (yy 0 )(y 0 y 00 ) =(1 + (y 0 )2 )2
(1 + (y 0 )2 )y − y(y 0 )2 y 00 =(1 + (y 0 )2 )2 − (1 + (y 0 )2 )(y 0 )2
yy 00 =(1 + 2(y 0 )2 + (y 0 )4 ) − ((y 0 )2 + (y 0 )4 )
so
p
dy ∂x0
ã
∂x
7
− =0
∂L
∂x0
⇒ y 00 =
(c) I have, instead, for the area of a ring of width ds and radius y,
p
=p
2πyx0
1 + (x0 )2
»
1 + (y 0 )2
= a,
y
d ∂L
Å
∂L
0
∂L
∂x0
= a,
,
a constant
a constant
0.0
0.5
T
1.0
y
1.5
2.0
AF
Solution
(b)
T =
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Z
ds
=
Z
x/π
dx
v=
p
2gy.
1 + (y 0 )2
√ ⇒T =
1
2ga2
(u
− sin u) and y =
(a) I have E = 12 mv 2 − mgy and the initial energy is Ei = 0. Since energy conserved (the wire is smooth) I have
p
1 + (y 0 )2
1
4ga2
(1 − cos u).
v 2gy 2gy
DR
(c)
Z Z p
ds 1 + (x0 )2 1 + (x0 )2
T = = dy √ ⇒ T̃ =
v 2gy 2gy
(d) Since ∂ L̃
∂x = 0 I will use the expression from (c). Euler’s equation is
0
Ç å
d ∂ L̃ ∂ L̃
∂ L̃
− =0 ⇒ = a, a constant
dy ∂x0 ∂x ∂x0
so
∂ L̃ 1 x0
= √ p = a, a constant
∂x0 2gy 1 + (x0 )2
⇒ (x0 )2 =2ga2 y(1 + (x0 )2 )
0 2ga2 y
⇒x =± .
1 − 2ga2 y
10.2. EULER’S EQUATION 145
This can be integrated using the substitution 2ga2 y = sin2 (u/2) so that 2ga2 dy = du sin(u/2) cos(u/2) and
s
2ga2 y du sin(u/2) cos(u/2) sin2 (u/2)
dx = ± dy = ±
1 − 2ga2 y 4ga2 1 − sin2 (u/2)
T
du sin(u/2) cos(u/2) sin(u/2) du
=± =± sin2 (u/2)
4ga2 cos(u/2) 4ga2
du
=± (1 − cos u)
2ga2
1
⇒x=± (u − sin u) + b, b another constant
2ga2
AF b = 0, leaving
Alternative Derivation
y=
2ga2
dL
=
∂L ∂L dy
+
∂L dy 0
+ 0
sin2 (u/2) =
=
x=±
y=
∂L
1
4ga2
+ y0
1
2ga2
∂L
4ga2
(1 − cos u),
are the parametric equations for a cycloid. Because (0, 0) is the initial point I have cos u0 = 1 so sin u0 = 0 and
(u − sin u)
(1 − cos u)
∂L
+ y 00 0
X
dx ∂x ∂y dx ∂y dx ∂x ∂y ∂y
∂L dL ∂L ∂L
⇒ y 00 = − − y0
DR
∂y 0 dx ∂x ∂y
so that
d 0 ∂L 00 ∂L 0 d ∂L
Å ã Å ã
y =y +y
dx ∂y 0 ∂y 0 dx ∂y 0
dL ∂L ∂L d ∂L
Å ã Å ã
= − − y0 + y0
dx ∂x ∂y dx ∂y 0
!
d ∂L ∂L
Å ã : 0 dL ∂L
=y 0 − + −
dx ∂y 0 ∂y dx ∂x
∂L d ∂L
Å ã
⇒ = L − y0 0 .
∂x dx ∂y
This equation will be revisited in later weeks.3
3 ∂L d
We will learn that this alternative formulation, with L = L(t, x, ẋ), ∂t
= dt
L − ẋ ∂L
∂ ẋ
, is widely used in classical mechanics. The
quantity
∂L
H(t, x, p) ≡ p ẋ(t, x, p) − L(t, x, ẋ(t, x, p)), p≡
∂ ẋ
is the Hamiltonian and the relationship between H and L given here is a Legendre transform. The quantity ẋ(t, x, p) is usually found by inverting
the definition of p (i.e. H can’t be an explicit function of ẋ).
146 CHAPTER 10. CALCULUS OF VARIATIONS
If ∂L
∂x = 0 then
∂L
L − y0 = a, a constant.
∂y 0
I can see that this is equivalent to
∂ L̃
T
=0
∂x0
by recalling that x0 = (y 0 )−1 and L̃ = x0 L,
−(x0 )−2 = −(y 00 )2
0
∂ ∂L 1 ∂L ∂y
7 ∂L
0
(x0 L) = L + x0 0 = L + 0 0 0 = L − y 0 0 .
∂x ∂x y ∂y ∂x ∂y
Example 10.6 The Brachistochrone Problem, Revisited. Ä ä
Use the expression for T found in part (b) of Ex. 10.5 and ∂T
∂x =
d
dx
∂T
T − y 0 ∂y 0 . to find the solution for this problem.
AF
Solution
10.3
From Ex. 10.5.b I have T =
»
q
1+(y 0 )2
2gy so
(y 0 )2 =
∂T
∂x= 0 and T − y 0 ∂y
a=
1
2ga2 y
1 + (y 0 )2
2gy
−1
∂T
0 = a, a constant. Therefore
Ç
− y0 p
2ga2 y(1 + (y 0 )2 ) =1 + (y 0 )2 − (y 0 )2
y0
2gy(1 + (y 0 )2 )
which, since x0 = 1/y 0 , is the same ODE found in Ex. 10.5.d. The solution is the cycloid.
Lagrangians
å
Solution
(a) I have ~r˙ = ẏ ŷ and V = −m~g · ~r = −m(−g ŷ) · (y ŷ) = +mgy, so
L = 21 mẏ 2 − mgy.
T
(b) I have for the partial derivatives
∂L ∂L
= mẏ & = −mg.
∂ ẏ ∂y
so
d
(mẏ) − (−mg) = 0 ⇒ mÿ = −mg
dt
which, of course, is Newton’s second law for this problem.
It is particularly important to note that once I had expressions for the position vector, ~r = y ŷ, and the potential
AF energy, V = −m~g · ~r, the remainder of the solution consisted of calculating the inner products and various
derivatives present in the Euler equation with
In fact, once the Lagrangian, L, is calculated the remainder of the problem is purely an exercise in mathematics
(i.e. L contains all of the physics of this problem).
For each of the following problems find an expression for the Lagrangian (in terms of the given variables and constants)
and use Euler’s equation with L = T − V to find the equations of motion. In each case give a brief description of the
solution (i.e. describe the physics in words). (a) One dimensional spring with ~r = xx̂ and V = 12 k~r · ~r. (b) Circular
motion with ~r = a(x̂ cos θ + ŷ sin θ) and V = −m~g · ~r. (c) Motion of a small bead, mass m, along a frictionless
wire shaped as y = a cosh(x/a) in the presence of a uniform gravitational field ~g = −g ŷ (a is constant). (In order to
understand the physics of this problem, consider the solution in the limit of small x.)
Example 10.8 Lagrangian Optics.
Fermat’s principle, aka the principle of least time, is the idea that the path taken by light between two points in space is
that which requires the least time. Fermat’s principle can be used to derive the laws of reflection and refraction (which
we did in Assn. 10.1). Rt
The time taken to travel between two points is (a) Fermat’s principle states that T = tif dt is stationary with
DR
respect to variations in the path. Show that
Z ~rf
cT = ds n(~r )
~
ri
in which v(~r ) = ṡ = is the speed of light as a function of position, and n(~r ) = c/v(~r ) is the refractive index as a
ds
dt
function of position (and c is the speed of light in vacuum). The integral of the quantity ds n is known as the optical
path length. (b) Take y(x) as the stationary path and find the function L(x, y, y 0 ) such that
Z xf
cT = dx L(x, y, y 0 ).
xi
(c) Use Euler’s equation to find the second order ODE, the solution of which is the stationary path through a medium
with refractive index n(~r ). This ODE forms the basis of Lagrangian optics. (d) Show that in the case of constant
refractive index the solution from Ex. 10.2 is recovered.
Solution
(a)
Z tf Z tf Z ~
rf
1 c 1
T = dt = dt ṡ = ds n(~r )
ti c ti v(~r ) c ~
ri
148 CHAPTER 10. CALCULUS OF VARIATIONS
so
T
»
L(x, y, y 0 ) = 1 + (y 0 )2 n(x, y(x))
AF
10.3.1
so Euler’s equation gives
d
dx
Ç
n
p
å
y0
−
1 + (y 0 )2
»
1 + (y 0 )2
å
=0
∂n
∂y
= 0.
This is a second order differential equation, the solutions of which, y(x), are the paths taken by light through a
Euler’s equation can be extended to the case of n dependent variables, yi (x), and their derivatives, yi0 (x), for i = 1...n.
X
I now have the problem of finding the functions yi (x) which make the functional
Z x(B)
S[yi (x)] = dx L(x, yi , yi0 ), i = 1...n
DR
x(A)
stationary.
Let yi = Fi (x) = fi (x) + ηi (x) be a set of functions which deviate by the small amount, ηi (x), from the
functions fi (x) which make S stationary, with ηi (x(A) ) = ηi (x(B) ) = 0.
I therefore have
Z x(B) Z x(B) ã Z x(B)
dS d ∂L dFi ∂L dFi0 ∂L ∂L 0
Å Å ã
= dx L(x, Fi , Fi0 ) = dx + = dx η i (x) + η (x) ,
d x(A) d x(A) ∂Fi d ∂Fi0 d x(A) ∂Fi ∂Fi0 i
in which the sums over the repeated indices is implied (recall the discussion of total derivatives in §G.0.3).
I can now integrate the second term by parts with
∂L
ui = & dvi = dx ηi0 (x)
∂Fi0
so that
d ∂L
Å ã
dui = & vi = ηi (x)
dx ∂Fi0
giving
Z x(B) x(B) Z x(B)
∂L 0 ∂L d ∂L
Å ã
dx η (x) = η (x) − dx η (x) .
∂Fi0 i ∂Fi0
i (A) i
dx ∂Fi0
x(A) x x(A)
10.3. LAGRANGIANS 149
Since ηi (x(A) ) = ηi (x(B) ) = 0 the boundary term is zero and I am left with
Z x(B)
dS ∂L d ∂L
Å Å ãã
= dx − ηi (x).
d =0 x(A) ∂yi dx ∂yi0
T
Since the ηi (x) are arbitrary I must have5
d ∂L ∂L
Å ã
− = 0, i = 1...n
dx ∂yi0 ∂yi
AF
10.3.2 Variational Notation
We have been examining Euler’s equation as a method of finding stationary solutions of
S[xi (t)] =
Z tB
dt L(t, xi , ẋi ).
tA
In this expression I have written S[xi (t)] to indicate that S is a functional – it is a function of the functions xi (t).
Recall that we assumed the stationary solutions were fi (t) and wrote xi = fi (t) + ηi (t) (with the ηi the varia-
tions). We then took the derivative of the functional S with respect to the parameter and, in the end, took = 0 so
that fi (t) = xi (t). This entire process is referred to as “the first variation of S with respect to ” and is written as
δS[xi ] ≡d
=
Z tB
dS
d =0
dt δL(t, xi , ẋi )
tA
ZtB
∂L ∂L
Å ã
= dt δxi + δ ẋi
tA ∂xi ∂ ẋi
DR
Z tB Å
∂L ∂L
ã
=d dt ηi + η̇i
tA ∂xi ∂ ẋi
The last term is then integrated by parts, giving
Z tB
d ∂L ∂L
Å Å ãã
δS[xi ] = 0 = d dt ηi −
tA dt ∂xi ∂ ẋi
d ∂L ∂L
Å ã
⇒ − = 0,
dt ∂ ẋi ∂xi
Euler’s equation.
Because this is a useful concept it is worth formalizing it. The variation of xi is
d
δxi = d (fi (t) + ηi (t)) = d ηi (t).
d =0
Similarly,
dÄ˙ ä
δ ẋi = d fi (t) + η̇i (t) = d η̇i (t),
d =0
Ä Ä ää
5 I could, for example, choose each ηi (x) to be positive wherever ∂L
− d ∂L
is positive and negative where it is negative.
∂yi dx ∂yi0
150 CHAPTER 10. CALCULUS OF VARIATIONS
T
10.3.3 That Interesting Property of Lagrangians
In subsection 8.2.3 I showed that adding a complete time-derivative, f˙, of a function of the generalized coordinates
and time, f = f (~q, t), had no effect on the resulting EOM. This is now obvious,
Z tB Z tB Z tB
df
S 0 [~q(t)] = dt L0 = dt L + dt
tA tA tA dt
= S[~q(t)] + f (~q(tA ), tA ) − f (~q(tB ), tB )
AF
so that
U = dm gy = µg ds y = µg
δS 0 [~q(t)] = δS[~q(t)]
(i.e. the function f is already evaluated at tA or tB and so is a constant) and the EOM from the action S 0 are the same
as those from the action S.
Consider the case of two pegs at the same height a distance w apart. A uniform cable of length ` hangs between the
pegs. Find the shape of the curve defined by the cable by considering that the potential energy of the cable,
Z Z Z w
(Hint: The constraint equation is ` = ds = −w/2 dx 1 + (y 0 )2 , so since U is stationary and ` is constant you can
find the stationary solution of F = U + λ( ds − `), in which λ is the Lagrange multiplier.)
DR
T
Chapter 11
Hamiltonian Dynamics
AF
11.1
Resources
Thornton & Marion – Ch. 7
Wikipedia - Hamiltonian Mechanics
Wikipedia - William Rowan Hamilton
Introduction
We have a system of n second order differential equations given by the ELE,
d ∂L
Å
dt ∂ q̇a
ã
−
∂L
∂qa
= 0, a = 1...n.
Week 11
It is possible to make this into a system of 2n first order differential equations. The easiest way is to introduce
va ≡ q̇a
DR
to give
∂L d ∂L ∂2L ∂2L ∂2L
Å ã
= = v̇b + vb +
∂qa dt ∂va ∂vb ∂va ∂qb ∂va ∂t ∂va
so the 2n first order differential equations are
∂L ∂2L ∂2L
∂qa − ∂qb ∂va vb − ∂t ∂va
v̇b = ∂2L
∂vb ∂va
q̇a = va ,
which are, indeed, 2n first order differential equations in the qa and va . But they are certainly inelegant.
If, instead, we use the equations for the conjugate momenta, pa , we can make the equations beautiful.
I will have
∂L
pa = pa (~q, ~q˙, t). ≡
∂ q̇a
So, if the determinant
∂pa ∂ 2 L
=
∂ q̇ ∂ q̇ ∂ q̇ 6= 0
b a b
151
152 CHAPTER 11. HAMILTONIAN DYNAMICS
is non-zero then the expression for pa can be inverted to write the generalized velocities as functions of the generalized
coordinates and their conjugate momenta (and time) and I can write the ELE as 2n first order differential equations,
T
ṗa = ṗa (~q, p~, t) =
∂qa q̇a =q̇a (~q,~p,t)
in which I have used q̇a = q̇a (~q, p~, t) to eliminate the generalized velocities in favour of the conjugate momenta after
performing the partial derivatives.
These 2n equations are certainly more elegant than those resulting from the substitution va = q̇a . This process is
easier to think of as a Legendre transformation which I will examine in the next section.
AF
The Hamiltonian is defined as a function of the generalized coordinates and their conjugate momenta (and time)
so I must have
H(~q, p~, t) = q̇a pa − L(~q, ~q˙, t),
dH =
∂H
∂qa
∂qa
∂L
∂t
dqa +
dqa −
dt.
∂H
∂pa
∂ q̇a
dpa +
∂H
∂t
dt,
q̇a = q̇a (~q, p~, t).
∂H
q̇a =
DR
∂pa
∂H
ṗa = −
∂qa
∂H ∂L
=− .
∂t ∂t
These are Hamilton’s equations (aka canonical equations).
Note, however, that ∂q∂a means different things when applied to the Lagrangian and the Hamiltonian – for the
Lagrangian the generalized velocities are kept constant while for the Hamiltonian the conjugate momenta are kept
constant.
An Interesting Result
The total derivative of the Hamiltonian with respect to time is
∂H ∂H ∂H
Ḣ = q̇a + ṗa +
∂qa ∂pa ∂t
∂H ∂H ∂H ∂H ∂H ∂H
= − + = ,
∂qa ∂pa ∂pa ∂qa ∂t ∂t
so if the Hamiltonian does not explicitly depend upon time then it is a constant of the motion.
11.2. THE HAMILTONIAN 153
T
of motion. Use the same generalized coordinates. (c) Plot both the potential energy, V (r) = 12 k(r − a)2 , and the
effective potential, Veff (r). (d) What angular velocity ω◦ , is required for a circular orbit with radius r◦ ? (e) Derive the
frequency of small oscillations, ω, about the circular orbit with radius r◦ . Write your solutions in terms of k, m, r◦
and a.
Solution
(a) The net force on the mass is the central force F~ = −k(r − a)r̂ and in polar coordinates the acceleration is
2mṙφ̇ + mrφ̈ =0
d
dt
Ä ä
mr̈ = −k(r − a) + mr
⇒ mr̈ = −
=
dt
⇒ mr̈ = − k(r − a) + mrφ̇2
dÄ 2 ä
−k(r
mr φ̇ =0 1
− a) +
mr3
mr2 φ̇ = 0, the angular momentum is a constant, pφ ≡ mr2 φ̇. And the first as
p 2 p2φ
p2r p2φ
H = 21 m(pr /m)2 + 21 mr2 (pφ /mr2 )2 + 12 k(r − a)2 = + + 21 k(r − a)2
2m 2mr2
and the equations of motion are
∂H p2φ ∂H
ṗr = − = − k(r − a) & ṗφ = − =0 2
∂r mr3 ∂φ
and
∂H ∂H
ṙ = + = pr /m & φ̇ = + = pφ /mr2
∂ ṗr ∂ p˙φ
(as above).
(c) Dynamic equilibrium occurs at r = r◦ and pφ = p◦ such that
p2◦
mr̈ = 0 = −k(r◦ − a) − ⇒ p2◦ = mkr◦3 (r◦ − a).
r=r◦ 2mr◦2
154 CHAPTER 11. HAMILTONIAN DYNAMICS
1 ################################# constants
2 k = 1.0
3 a = 1.0
4 m = 1.0
5 N = 1000
6 ################################# calculated constants
7 w = sqrt(k/m) # natural spring frequency
8 r0 = 6.0/5.0 * a # (conveniently chosen) radius of circular orbit
T
9 w0 = w * sqrt( 1 - a / r0 ) # orbital frequency
10 W = w * sqrt( 4 - 3*a / r0 ) # natural spring frequency in rotating system
11 V0 = 0.5 * k * ( r0 - a ) * ( 2*r0 - a ) # natural unit for energy, Veff (ro )
12 ################################# plot V (r), the effective potential
13 r = linspace( 0, 2, N ) * r0 + 0.01 # offset from zero to avoid NaN at origin
14 p0 = m * w0 * r0 * r0 # angular moment for circular orbit r = 2a
15 Vs = 0.5 * k * ( r - a )**2 # spring potential
16 Vc = p0 * p0 / ( 2 * m * r * r ) # centrifugal potential
17 V = Vs+Vc # effective potential
18 figure(0)
19 plot( r/r0, V/V0, lw=2 )
20 plot( r/r0, Vs/V0, 'k--' )
21 plot( r/r0, Vc/V0, 'k.', ms=2 )
22 xlabel( r'$r/r_o$' )
23
24
25
AF ylabel( r'$V/V_0$' )
axis( [0,2, 0, 2] )
grid('on')
Listing 11.1: The effective potential for Example 11.1, plotted using the code in Example 11.1.
It is convenient to define
V◦ ≡ Veff (r◦ ) =
=
p2◦
2mr2
1
2
+ 2 k(r − a)
mkro3 (ro − a) 1
2mro2
r=r◦
(d) Circular orbits, radius r◦ , are at the minimum of the effective potential,
dVeff p2φ
= 0 = − + k(r − a) ⇒ p2◦ ≡ p2φ |r=r◦ = mkr◦3 (r◦ − a)
dr r=ro mr3
r=ro
DR
(note that ṗr = −Veff
0
) so with p◦ = mω◦ r◦2 I get
k
ω◦2 = (r◦ − a)/r◦ 2
m
which is the orbital frequency for circular orbits of radius r◦ .
d2 Veff
(e) For small oscillations I need to find the effective spring constant, Keff ≡ dr 2 r=r , so
o
Keff k 4r◦ − 3a
ω2 = = · . 2
m m r◦
Note that this is the frequency of oscillations about r = r◦ and is distinct from the orbital frequency, ω◦ . X
2.0
1.5
T
V /V0
1.0
0.5
0.0
AF p
Vo ≡ Veff (ro ) =
p
0.0
p2θ
2mr2
1
0.5
2
+ 2 k(r − a)
1.0
r/ro
r=ro
1.5 2.0
= 21 k(2ro − a)(ro − a)
3
unit of distance and k/m as your unit of angular frequency. Take vx = 0 and vy = r◦ ω◦ and x = 1.1r◦ , y = 0 as
your initial conditions. Use
as your unit of energy and T ≡ 2π m/k as your unit of time. Choose the frequency of small oscillations, ω, as three
times the orbital frequency, ω◦ .(b) Plot the numeric results for r − r◦ as a function of time for 0 ≤ t ≤ 3T . (c) Plot
the position of the mass on an x-y plot; include a plot of the circle r = r◦ for comparison.
Consider a bead of mass m constrained to move along a wire z = (a) What are
3 α x in a uniform gravitational field, ~
g = −g k̂. We now rotate the
1 2 3
Hamiltonian’s equations for the bead? (b) Plot the effective potential, Veff /V1 , versus x/x1 where x1 ≡ ω 2 /gα2 and
√
V1 ≡ Veff (x1 ) = mω 6 /6g 2 α4 . (c) Plot several contours of constant H/V1 . Use ω = gα. (d) Calculate (and then
plot) the trajectory in phase space, (x/x1 , p/po ), where po ≡ mω/α, using the initial conditions (1, 0) and (1.1, 0).
Exercise 11.3 Sliding Mass.
PHYS363 2014S
Test #2 Student Number: -
156 CHAPTER 11. HAMILTONIAN DYNAMICS
2. Consider a small mass m which slides without friction in a slot along a diameter a
of a horizontal disk of mass 2m and radius a and which is connected by a light
Consider spring (spring
a small mass constant
m which k, zero equilibrium
slides length) to
without friction in the
a centre of the disk (see
figure
slot along a diameter Thea horizontal
at right). of horizontaldisk diskis of
freemass
to rotate and its vertical axis.
2m about
radius a Consider
and which carefully how many
is connected bydegrees
a lightofspring
freedom there are and what generalized
(spring
coordinates will you use. Reading the entire question before beginning may be
m
constant k, zero equilibrium length) to the centre of the
T
helpful.
disk (see figure at right). Use x, the position along the slot,
(1) (a) i. How many degrees of freedom, f , there are?
and θ, the angle the positive x direction makes with a fixed
ii. What are the best generalized coordinates, {q↵ }, for this problem
direction in the horizontal plane, as your generalized coor-
(i.e. those that give as many conserved quantities as possible)?
dinates. R
iii. What is the moment of inertia, I = dm r2 , of the disk? 2m
(Assume the slot is negligible.)
(a) What. . .are Hamiltonian’s equations for this problem? (b)
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .√
Plot the effective potential, V . . . ./ka
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .eff
2
, versus x/a, for
............................................
several values of pθ /pc between zero and two, where pc ≡ a km is the critical angular momentum for this problem. 2
(c) Plot several contours of constant H/ka for several values of pθ /pc between zero and two (bonus: make a movie
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
in
AF
of these contours
(2)phase (b)
(2)
. . . . . . . . . . . with
space,
What (x/a,
. . . . . . . . . . .p
is thepLagrangian,
. .θ. . . .changing
. . . . . . . . . . . . . . . . . . . from
x /pc ), using
. . . . . . . . . . . zero
. . . . . . . . . . to
. . . . . two
. . . . . . . . . as
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
DR
.....................................................................................................................................................................................................................
⇣ ⌘
(2) (d) Use the Euler-Lagrange equation, dt d @L
@ q̇a @qa = 0, to find the equations of motion. (One of the generalized
@L
coordinates should be cyclic; if not, go back to part (a) and start again.)
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
.....................................................................................................................................................................................................................
Page 5 of 7
11.3. FURTHER PROBLEMS 157
T
potential. (a) Calculate, again, the conjugate momentum, p~ = ∂L r˙
∂~
. (b) What is the Hamiltonian, H(~r, p~) = p~ · ~r˙ − L?
(c) It turns out that the choice of the potentials, φ & A,~ is not unique. If we perform a “gauge transformation”
1 ∂λ ~0 = A
~ + ∇λ
~
φ0 = φ − , A , λ = λ(~r, t)
c ∂t
we can write the new Lagrangian as, L0 = L + ec g(~r, t), where L is the original Lagrangian. Find this new function,
g(~r, t). (d) Show that the addition of such function to the Lagrangian does not alter the resulting equation of motion.
AF
Solution
(a) The conjugate momentum is
(b)
(c)
(d)
so
p~ =
∂L
∂~r˙
e~
= m~r˙ + A,
c
X
DR
158 CHAPTER 11. HAMILTONIAN DYNAMICS
Week 12
T
AF
DR
T
Appendix A
Vectors
AF
A vector is a quantity with both magnitude and direction. Examples that frequently occur in classical mechanics are
displacement, velocity, acceleration, force, and momentum. It is important to recognize that these vectors are physical
quantities that exist independently of any choice of coordinate system. Since the laws of physics are independent of the
choice of coordinate system it is valuable to mathematically represent the laws of physics in terms of vector quantities
because this makes their coordinate independence manifest.
For example, Newton’s second law is written as F~ = m~a and we can use this to write the equations of motion for
a system (i.e. the differential equation which has solutions that describes the motion of a system).
In order to represent vectors mathematically it is convenient to introduce a set of basis vectors (i.e. a set of
orthonormal vectors that span the vector space). There are a number of different possible choices, the choice of which
depends upon the problem at hand.
Solution
(a) The gravitational force is F~g = m~g so Newton gives me
m~a = m~g .
159
160 APPENDIX A. VECTORS
~v˙ = ~g
T
Z t Z t
dt ~v˙ =
0
dt0 ~g
t◦ t◦
Z t Z t
d~v
dt0 = ~g dt0
t◦ dt0 t◦
[~v ]tt◦ = ~g [t0 ]tt◦
~v − ~v◦ = ~g (t − t◦ ).
AF
A.2
With ~r˙ = ~v I can integrate a second time to give
t◦
t
d~r
Z
dt0 0 − ~v◦
dt
t◦
t
dt (~r˙ − ~v◦ ) = ~g
0
Curvilinear Coordinates
Z
t◦
t
dt0 = ~g
Z
Z
t◦
t
t◦
t
dt0 (t0 − t◦ )
dt0 t0 − ~g t◦
2
[~r ]tt◦ − ~v◦ (t0 − t◦ ) = ~g [ 21 t0 ]tt◦ − ~g t◦ (t0 − t◦ )
~r − ~r◦ − ~v◦ (t − t◦ ) = 12 ~g (t2 − t2◦ ) − ~g t◦ t0 + ~g t2◦
= 12 ~g (t2 − t2◦ − 2t◦ t0 + 2t2◦ )
⇒ ~r = ~r◦ + ~v◦ (t − t◦ ) + 21 ~g (t − t◦ )2 .
Z
t◦
t
dt0
It is often convenient to use non-Cartesian coordinate systems. The three most common such systems are polar,
cylindrical, and spherical.
DR
A.2.1 Polar Coordinates
An alternative to Cartesian coordinates is polar coordinates. In this system the position vector is
~r = rŝ
so ŝ is the unit vector in the ~r direction. We measure φ as the angle the unit vector ŝ makes with the positive x̂ axis (in
the counter-clockwise direction) and the vector orthogonal to ŝ is φ̂ which points in the direction of increasing φ (see
Figure A.1).
If the position vector, ~r, is a function of time then so are the basis vectors are ŝ and φ̂,
ŝ
ï ò ï ò
x̂
=A
φ̂ ŷ
A.2. CURVILINEAR COORDINATES 161
φ̂
T
r̂
ŷ
~r
AF
Solution
(a) I have
ï ò
x̂
ŷ
= A−1
x̂
Figure A.1: In polar coordinates the orthogonal basis vectors are r̂ and φ̂ and the position vector is always ~r = rr̂. The φ̂
basis vector is in the direction of increasing φ. Crucially, if the position vector, ~r, is a function of time then so are the basis
vectors are r̂ and φ̂.
in which A is a 2 × 2 matrix. (b) What is the matrix for the inverse relationship,
ŝ
ï ò
φ̂
ŝ
ï ò ï
φ̂
=
cos φ
− sin φ
sin φ
cos φ
òï ò
x̂
ŷ
(b) and, since this matrix is orthogonal,1 the inverse is the transpose,
DR
ï ò
cos φ − sin φ
A−1 = AT =
sin φ cos φ
1 An orthogonal matrix is one whose rows and columns are orthogonal unit vectors. If we represent such a matrix as a block matrix,
A = ê1 ê2
(i.e. the columns are the unit vectors êi ) then
h i h i
ê1 ê1 · ê1 ê1 · ê2
AT A = ê1 ê2 = =1 ⇒ AT = A−1
ê2 ê2 · ê1 ê2 · ê2
162 APPENDIX A. VECTORS
ẑ
z φ̂
~
r
ŝ
T
z
φ s
Figure A.2: Cylindrical coordinates are the same as polar coordinates with the addition of the ẑ unit vector perpendicular to
both ŝ and φ̂.
AF x
z
φ
θ
~
r
y
r̂
φ̂
θ̂
ŝ cos φ
φ̂ = − sin φ
0
sin φ 0 x̂
cos φ 0ŷ
0 1 ẑ
ẑ
ŝ˙ = φ̇φ̂
DR
˙
φ̂ = −φ̇ŝ
ẑ˙ = 0
~r = sŝ + z ẑ
~r˙ = ṡŝ + sφ̇φ̂ + ż ẑ
~r¨ = (s̈ − sφ̇2 )ŝ + (2ṡφ̇ + sφ̈)φ̂ + z̈ ẑ
r̂ sin θ cos φ sin θ sin φ cos θ x̂
θ̂ = cos θ cos φ cos θ sin φ − sin θŷ
φ̂ − sin φ cos φ 0 ẑ
A.2. CURVILINEAR COORDINATES 163
T
˙
φ̂ = −φ̇(sin θr̂ + cos θθ̂)
Ä ä Ä ä
r̂¨ = −θ̇2 − φ̇2 sin2 θ r̂ + θ̈ − φ̇2 sin θ cos θ θ̂
Ä ä
+ 2θ̇φ̇ cos θ + φ̈ sin θ φ̂
~r = rr̂
~r˙ = ṙr̂
Ä + rθ̇θ̂ + rφ̇ sin θφ̂ ä
~r¨ = r̈ − rθ̇2 − rφ̇2 sin2 θ r̂
Ä ä
+ 2ṙθ̇ + rθ̈ − rφ̇2 sin θ cos θ) θ̂
AFÄ
~r = Ä
r sin θ(cos φx̂ + sinä φŷ) + r cos θẑ
~r˙ = ṙ sin θ + rθ̇ cos θ (cos φx̂ + sin φŷ)
+Ä ä cos φŷ)
+ ṙ cos θ − rθ̇ sin θ ẑ
DR
164 APPENDIX A. VECTORS
T
AF
DR
T
Appendix B
Index Notation
AF
B.1
Resources
Thornton & Marion §1.6-17
Boas §4.1-12
Boas §10.1-5
Vectors
Introduction
For tensors with rank greater than one it is more convenient to work in index notation. In this notation the basis vectors
are êi , i = 1, 2, 3, and a general vector, ~v , can be written as
It is important to remember that in the above two equations the vector ~v is the same vector and only its reprepresen-
tation in the two bases, vi & vi0 , are different.
~v = vi êi
= vi0 ê0i
which is a very compact notation. Clearly this notation is trivial to extend to three or more dimensions.
Note in particular that a summation index can be replaced by any other index,
165
166 APPENDIX B. INDEX NOTATION
without affecting the meaning or validity of the equation. Note also that an index may only appear either once or twice
in each expression. An index that appears only once in an expression, called a free index, must appear on each side of
the equation exactly once. An index may only appear either once or twice in each expression.
Matrices
T
Matrices can also be represented using index notation. For example, if I have the matrix
a11 a12 a13
A = a21 a22 a23
a31 a32 a33
then the elements of the matrix are aij and [aij ] is the matrix made up on the matrix elements aij ,
A = [aij ],
AF
and [A]ij is the ij th element of the matrix A,
x1
x3
[A]ij = aij .
The brackets, [ and ], are used to change between the two different notations.
Note in particular that the indices of a matrix are always row number first then column number. This is an arbitrary
convention that is best adhered to.
Matrix Multiplication
The matrix, A, multiplied by a vector, ~v , is
~x = A · ~v
a11 a12 a13 v1
in which I have used Einstein summation convention (i.e. there is an implied sum of the index j).
Switching to index notation, I have
xi = aij vj .
in which i and k are free indices while j is a dummy or summation index. This mapping between the two notations
requires that the summation indices are adjacent. For example, if I have
C = ABT ,
then
cik = aij bT
jk = aij bkj ,
T
B.3 Further Problems
Exercise 12.1 Maxwell’s Equations.
Maxwell’s equations,
~ ·E
ε◦ ∇ ~ =ρ ~ ·B
∇ ~ =0
~ ~
∇~ ×E~ = − ∂B ~ ×B
∇ ~ = µ◦ J~ + µ◦ ε◦ ∂ E ,
∂t ∂t
AF
Introduce
~ ×B
∇
qV ≡
IS ≡
ΦS ≡
Z
Z
V
S
dτ ρ
d~a · J~
d~a · E
S
~
~
R your equations.
H (a) Write Gauss’ law, ε◦ ∇~ ·E
~ r, t).
relate the electric and magnetic fields, E(~r, t) & B(~r, t), to the charge and current distributions ρ(~r, t) &, J(~
~ = ρ, using index
notation. (b) Use Gauss’ theorem in index form, V dτ ∂i vi = ∂V dai vi where ∂V is the boundary of the volume V,
to re-write Gauss’ law as an integral equation. (c) Change your result back to vector notation. (d) Write Ampére’s law,
R
~ = µ◦ J~ + µ◦ ε◦ ∂ E~ , using index notation. (e) Use Stokes’ theorem in index form, dai εijk ∂j vk =
∂t S
H
∂S
where ∂S is the boundary of the surface S, to re-write Ampére’s law as an integral equation. (f) Change your result
dxi vi ,
back to vector notation. (g) Take the divergence, êi ∂i , of your answer to part (d), substitute in your answer to part (a),
and simplify your result. (h) Give a physical interpretation to your result in (g). [2017S midterm]
Exercise 12.2 Cartesian.
In cartesian coordinates the position vectors is ~r = xi êi (assume Einstein summation convention for this question). In
√
DR
polar coordinates the position vectors is ~r = sŝ where s = xi xi is the length of the vector ~r, and ŝ is the unit vector
in the r̂ direction. (Hint: ŝ is a function of φ, the angle between ~r and the positive ê1 axis.) (a) Use the fact that ŝ
is in the ∂s
∂
~r direction and φ̂ is in the ∂φ
∂
~r direction to find expressions for ŝ and φ̂ in terms of the êi basis. (Hint:
~r = s(ê1 cos φ + ê2 sin φ)) (Hint 2: Check that ŝ · φ̂ = 0.) (b) The total derivative of ψ(~r) in cartesian coordinates is
∂ψ ~
dψ = dxi = d~r · ∇ψ
∂xi
while in polar coordinates it is
∂ψ ∂ψ
dψ = ds + dφ.
∂s ∂φ
Use the expression for the total derivative of d~r in polar coordinates, d~r = ∂~
r ∂~
r ~
to find an expression for ∇ψ
∂s ds + ∂φ dφ,
~
in polar coordinates with the ŝ-φ̂ basis. (Hint: Use your results from (d) for d~r.) (Hint 2: The scalar quantity d~r · ∇ψ
is independent of the choice of coordinate system.)
Exercise 12.3 In Summation.
Consider the case in which the position of a particle is a function of time, ~r(t) = xi (t) êi . The êi are the Cartesian
orthonormal basis vectors; assume Einstein summation convention for this question. Express your answers in index
notation. (a) Using index notation, calculate rṙ and ~r · ~r˙ , where ~r˙ ≡ d~ dt , r = |~
r
r|, etc. (b) Using index notation,
d ˙
calculate dt (~r · ~r). (c) (d) Plot ~r(t) for |~r| = r◦ , a constant. (Hint: What is ~r · ~r for this case?)
168 APPENDIX B. INDEX NOTATION
(Hint: Use the properties of the Levi-Civita symbol, εijk .) (b) Give a physical interpretation of the result from (a).
T
(c) Use index notation to show that
Ä ä Ä ä ÄÄ ä ä ÄÄ ä ä
~a × ~b × ~c × d~ = ~a × ~b · d~ ~c − ~a × ~b · ~c d~
AF
DR
T
Appendix C
Vector Multiplication
AF
The Kronecker Delta
Inner Product
An important example of its use is the inner product of two orthonormal basis vectors,
The quantity (êi ⊗ êj ), sometimes referred to as a dyad, can be thought of as the basis for a rank-2 tensor and (ai bj )
as the representation of the tensor in that basis.
Cross Product
169
170 APPENDIX C. VECTOR MULTIPLICATION
so that
0 −u3 u2 ê1 v1
[~u ]× · ~v = ê1 ê2 ê3 u3 0 −u1 ê2 · ê1 ê2 ê3 v2
−u2 u1 0 ê3 v3
T
0 −u3 u2 v1
= ê1 ê2 ê3 u3 0 −u1 v2
−u2 u1 0 v3
−u3 v1 + u2 v3
= ê1 ê2 ê3 u3 v1 − u1 v3
−u2 v1 + u1 v2
= ~u × ~v
In the êi basis the representation of the cross product tensor is the cross product matrix,
AF
Levi-Civita Symbol
It is now convenient to introduce the Levi-Civita aymbol,
εijk
+1
= −1
0
0
[~u ]× → u3
−u2 u1
−u3 u2
0
−u1 .
0
In words, εijk is 1 if (i, j, k) is an even permutation of (1, 2, 3), −1 if it is an odd permutation, and 0 otherwise (i.e.
if one or more of the indices is repeated). The Levi-Civita symbol is sometimes referred to as “the anti-symmetric
tensor” because exchanging any two indices changes the sign of the Levi-Civita symbol,
(this is akin to taking the transpose of a 2D matrix and it is the changing sign that makes the Levi-Civita symbol
DR
anti-symmetric).
T
and so the i-th component of the cross product is
[~u × ~v ]i = εijk uj vk .
AF
Proof: Consider the triple vector product identity1
~ × C)
in index notation the kth component of (B
~ × (B
A ~ × C)
~ = (A
~ · C)
~ B~ − (A
~ · B)
~ C~
The identity can also be demonstrated by considering the fact that both ijk and k`m must both consist of a
permutation of the three integers 123 in order for their product to be non-zero. If we first permute the indices,
then we can see that since k is the same for both Levi-Civita symbols there are only two cases for which the sum of
products is non-zero:
i = ` and j = m ⇒ εijk ε`mk = +1
®
Solution
(a) I have
~a · ~b = ~aT~b = ai êi · bj êj = ai bj (êi · êj ) = ai bj δij = ai bi .
(b) I have
~a · ~b = a0i ê0i · b0j ê0j = a0i b0j ê0i · ê0j = a0i b0j δij = a0i b0i .
1 The proof for this identity is rather dull. The x component is
~ × (B
[A ~ × C)]
~ x = Ay (Bx Cy − By Cx ) − Az (Bz Cx − Bx Cz )
= Bx (Ay Cy + Az Cz ) − Cx (Ay By + Az Bz )
= Bx (Ax Cx + Ay Cy + Az Cz ) − Cx (Ax Bx + Ay By + Az Bz )
~ · C)B
= (A ~ x − (A
~ · B)C
~ x.
This is an important result. If the vectors ~a & ~b represent physical quantities then their inner product must be
T
independent of the choice of coordinate system (i.e. it will be a rank-0 tensor). This exactly what I have shown in
part (c). X
AF
Solution
In index notation I have
so
(in which I have changed the dummy indices in the central expression, jki → ijk) so that
X
DR
Example 12.5 Cross-Dot.
Show that
~ × B)
(A ~ · (C
~ × D)
~ = (A
~ · C)(
~ B ~ · D)
~ − (A
~ · D)(
~ B ~ · C).
~
Solution
I have (Â × B̂)k = εijk ai bj and (Ĉ × D̂)k = εlmk cl dm so
~ × B)
(A ~ · (C
~ × D)
~ = εijk εlmk ai bj cl dm .
Using the identity εijk εlmk = δil δjm − δim δjl gives
~ × B)
(A ~ · (C
~ × D)
~ = (δil δjm − δim δjl )ai bj cl dm
= (ai ci )(bj dj ) − (ai di )(bj cj )
~ · C)(
= (A ~ B ~ · D)~ − (A ~ · D)(
~ B ~ · C)
~
X
2 An outer product is also known as a tensor product.
173
where Ä ä
A~B
~C ~ · (B
~ =A ~ × C)
~ =B
~ · (C
~ × A)
~ =C
~ · (A
~ × B)
~
T
is the triple scalar product from Example 12.4.
Example 12.6 Triple Vector Product.
Calculate the vector triple products (a) ~a × (~b × ~c) and (b) (~a × ~b) × ~c, using index notation. Put your final results
back into vector notation. (Hint: The identity εijk εk`m = δi` δjm − δim δj` may be helpful.)
Solution
AF
(a) In index notation I have
so
= am b` c` − a` bm c` 1
= (~b · ~c)am − (~a · ~c)bm 1
so
(~a × ~b) × ~c = (~b · ~c)~a − (~a · ~c)~b. +1
X
174 APPENDIX C. VECTOR MULTIPLICATION
T
AF
DR
T
Appendix D
AF
If we start with the vector ~v in two different coordinate systems,
and take the inner product with ê0j (since the index i is already repeated we can not re-use it) we get
~v · ê0j = vi êi · ê0j = vi0 ê0i · ê0j = vi0 δij ⇒ vj0 = vi êi · ê0j
where δij is the Kronecker Delta symbol (note that the orthonormality of the êi is essential for this identity – we
wouldn’t have ê0i · ê0j = δij otherwise).
Similarly, if we take the inner product with êj we get
so that
vj = vi0 λij .
DR
Next, consider the transpose of Λ is
T
λ11 λ12 λ13 λ11 λ21 λ31
ΛT = [λij ]T = λ21 λ22 λ23 = λ12 λ22 λ32 = [λji ] ⇒ λTij = λji
λ31 λ32 λ33 λ13 λ23 λ33
I therefore have
vj0 = vi λT
ij = λji vi .
1 Here I use the notation that if Λ is a matrix and λij is its ijth element, then [Λ]ij = λij and [λij ] = Λ.
175
176 APPENDIX D. COORDINATE SYSTEMS TRANSFORMATIONS
T
Example 12.7 Rotations in 2D.
Consider a rotation of the basis {ê1 , ê2 } by a counter-clockwise angle α into the primed basis {ê01 , ê02 } so that
ï 0ò ï ò
ê1 ê
= Λ(α) 1 .
ê02 ê2
(a) Re-write this expression in index notation (Hint: the ijth element of Λ is λij = [Λ]ij ). (b) Use λij = ê0i · êj to
find the rotation matrix Λ. (c) Use your result to write expressions for the vi0 in terms of the vj . (d) Find the inverse
rotation matrix, Λ−1 , such that
AF
Solution
(a) The expression can be written as
Λ(α) =
ï ò
=
v1
v2
ï
ï
ï
ï 0ò
v
= Λ 10 .
v2
ê01 · ê1
ê02 · ê1
cos α
cos(π/2 + α)
ê01 · ê2
ê02 · ê2
cos(π/2 − α)
ò
cos α
ò
ò
cos α sin α
= .
− sin α cos α
DR
(c) I have ï 0ò ï ò ï òï ò
v1 v1 cos α sin α v1
= Λ =
v20 v2 − sin α cos α v2
so that
(e) With ï ò ï ò ï òï 0 ò
v1 v cos α − sin α v1
= ΛT 1 =
v2 v2 sin α cos α v20
177
I get
T
X
Orthogonal Transformations
Consider the determinant of this last expression,
det(δjk ) = δ1i δ2j δ3k ijk = 1 = det λT
ji λik
= det λT
ji det(λik )
= det(λij ) det(λik )
AF
so that
det(λij ) = ±1.
2
The case for which det(λij ) = +1 are called “rotations” and those for which det(λij ) = −1 are called “improper
rotations” (i.e. a combination of a reflection and a rotation). The distinction between these two types of rotations is
important when dealing with pseudo-vectors, i.e. those defined by the right-hand-rule such as angular momentum, L,
or angular velocity, ω
~.
~
DR
178 APPENDIX D. COORDINATE SYSTEMS TRANSFORMATIONS
T
AF
DR
T
Appendix E
Rotation Generator
AF
Consider the point P in two different reference frames (see Figure E.1). Its position is
where ~rP0 is the position of P relative to the IRF, ~rP is the position of P relative to the NIRF, and ~rO0 is the vector
connecting the origins of the two reference frames.
Recall that we can transform from one basis to another using the rotation matrix,
so that
0
ê0i = λij êj
~r = ri êi = ri λT 0
T 0
0
0
ki êi = λki λij êj = δjk êj ⇒ êk = λki êi = êi λik
and we can represent the position of P, ~r, relative to the NIRF in terms of either basis,
P
~rP0 ê2
~rP
ê03 ê1
ê3 O
O0 ~rO0
ê02
ê01
Figure E.1: The un-primed, NIRF has its origin at O. The primed, IRF, with origin at O0 . The position of the point P is either
~rP0 or ~r 0 and these are related by ~rP0 = ~rO0 + ~rP .
179
180 APPENDIX E. ROTATION GENERATOR
λT T
ij (t + ∆t) − λij (t)
λ̇T
ij λjk êk = lim λjk êk
∆t→0 ∆t
!
T
δ
1 * ik
= lim λT
ij (t + ∆t)λjk (t) λT
− (t)λjk (t) êk
ij
∆t→0 ∆t
so the first term maps the NIRF basis, êk , to the IRF basis, λjk (t)êk = ê0j , at time t and then back to the NIRF basis,
ij (t + ∆t)êj , at time t + ∆t.
0
êi = λT
As ∆t → 0 this must be close to an identity operator so I can write, to first order,
2
λT
ij (t + ∆t)λjk (t) = δik − Ωik ∆t + O(∆t )
AF
so, putting the pieces together, I get
ΩT
Ä
T
ik = Ωki = − λ̇ij λjk
which is an extremely interesting result: The time derivative is replaced with a matrix multiplication, λ̇ij = −Ωik λT
In the ∆t → 0 limit this is an exact result. The matrix Ωij is called the generator of the rotation and is an
important quantity.
The velocity of the point P is then
= −λ̇T
Ä ä
ij λjk êk
T T
kj λji = −λ̇jk λij = −λij λ̇jk
kj .
d T
λij λjk = 0 = λ̇T T
ij λjk + λij λ̇jk = −Ωik − Ωki ⇒ Ωik = −ΩT
ik
dt
DR
and so the generator is an anti-symmetric 3 × 3 matrix. It therefore has three independent elements, and I can write its
elements as
T
⇒ ~r˙P0 = ~r˙O0 + ~r˙ + ω
~ × ~r.
We can interpret this as stating that the velocity of the point P in the IRF, ~r˙P0 , is equal to the velocity of the NIRF origin,
~r˙O0 , plus the velocity of the point P in the NIRF, ~r˙ , plus the velocity the point P due to the rotation of the NIRF, ω
~ × ~r.
AF
DR
182 APPENDIX E. ROTATION GENERATOR
T
AF
DR
T
Appendix F
Rotations
AF
Consider the vector ~v , and its representation in terms of two different bases (see Figure F.1). In terms of {ê1 , ê2 } I
have
= v cos θ, sin θ
ê0
ï ò
ê2
òï ò
.
ê0
ï ò
= v cos θ cos α + sin θ sin α, sin θ cos α − cos θ sin α 10
ê2
Since the these two quantities are the same I must have
ï ò ï 0ò
ê1 ê
DR
= Λ(−α) 10
ê2 ê2
where ï ò
cos α − sin α
Λ(−α) ≡
sin α cos α
Figure F.1: The vector ~v can be represented using two different coordinate systems.
183
184 APPENDIX F. ROTATIONS
is the rotation matrix which rotates the primed basis, {ê01 , ê02 }, by an angle −α, into the unprimed basis, {ê1 , ê2 },
(the minus is so that counter-clockwise rotations will be positive and is only a convention).
Rotation matrices are orthogonal and have a determinant of one and an inverse given by its transpose,
det(Λ(−α)) = 1
T
Λ−1 (α) = Λ(−α) = ΛT (α).
I therefore define the rotation matrix, Λ, as that which rotates the basis vectors by a counter-clockwise angle α by
ï ò
cos α sin α
Λ(α) ≡
− sin α cos α
and I note carefully that the 21 element carries the minus sign.1
AF
F.0.1 3D Rotations
There are many ways that we can describe the relative orientation of two coordinate systems. In two dimensions
rotations are within the plane (i.e. about a single axis perpendicular to the plane) and so there is only one angle needed
to specify the rotation. In three dimensions rotations can be about an arbitrary axis and so require three angles. For
example, two to define the direction of the axis and a third to specify how much to rotate about that axis.
Direction Cosines
Perhaps the simplest method to define a rotation in three dimensions is to specify the axis as a unit vector, v̂, and to
give the angles, αi , this vector makes with the basis vectors, êi ,
X
vi2 = vi vi = 1 =
X
v̂ · êi = vi = cos αi .
and so only two of the three angles are independent. The third independent angle is, of course, the rotation about the
axis defined by v̂; the sense of the rotation is given by the RHR for the axis.
DR
I can define the rotation by a vector ~v = vv̂ with the rotation about the axis v̂ being given by the magnitude of the
vector |~v | = v; such a vector is known as the rotation vector or the Euler vector. However, rotations in 3D do not, in
general, commute and so such a quantity will not behave as a vector (although its derivatives will – see §4.3 on p. 65).
Euler Angles
One very useful system to describe the relative orientation of two coordinate systems is attributed to Euler and, like all
such systems, consists of three angles. The Euler angles are {φ, θ, ψ} and consist of
Note that there are 12 possible ways to define Euler angles; this one is called 313. The Wikipedia article on Euler
angles may also be also helpful.
Here the three parameters required to describe the orientation of a body (as mentioned at the start of this chapter)
consist of two to define the direction of the axis (φ and θ as used in spherical coordinates) and a third (ψ) to specify
how much to rotate about that axis.
1 In physics we specify the elements of a matrix by the row and column numbers, always in that order.
185
ê2
ê3 ê3
ê2
T
ê2 ê1 ê1
ê1 ê3
AF
ê1
cos ψ
ê2
sin ψ 0 1
ê3
ê1
0 0 cos φ sin φ 0
Λ = − sin ψ cos ψ 00 cos θ sin θ − sin φ cos φ 0
0 0 1 0 − sin θ cos θ 0 0 1
ê3
Figure F.2: A rotation by π/2 about the ê3 axis followed by a rotation by π/2 about the ê1 axis (top row) is not equivalent to
a rotation by π/2 about the ê1 axis followed by a rotation by π/2 about the ê3 axis (bottom row). In these diagrams the faces
orthogonal to the ê1 axes are green, those orthogonal to the ê2 axes are blue, and those orthogonal to the ê3 axes are red
cos ψ cos φ − cos θ sin ψ sin φ cos ψ sin φ + cos θ cos φ sin ψ sin ψ sin θ
= − cos φ sin ψ − cos ψ cos θ sin φ cos ψ cos θ cos φ − sin ψ sin φ cos ψ sin θ
DR
sin θ sin φ − cos φ sin θ cos θ
Rotation matrices have three eigenvalues, {1, e±iα }, and the eigenvector corresponding to the real eigenvalue, α̂,
Λα̂ = α̂ or λij αj = αi
(with αi αi = 1) is unchanged by the rotation. This unit vector defines the rotation axis. The complex eigenvalues,
e±iα , give the angle of rotation; it is convenient to introduce the rotation vector (aka the Euler vector), parallel to α̂
and with magnitude equal to the angle of rotation,
α
~ = αα̂.
Angular Velocity
Consider the path of a particle moving in an arbitrary path in space with position vector given by ~r(t). During any
infinitesimal time interval dt its path may be considered as an infinitesimal arc of a circle. The axis through the centre
186 APPENDIX F. ROTATIONS
T
Figure F.3: There are twelve possible ways to specify Euler angles. This one, labeled z-x-z, is most common in physics. It
AF
consists of a rotation of φ about the ê03 -axis followed by a rotation of θ about the ê1 -axis, followed by a rotation of ψ about
the ê3 -axis. Pictured here is a snapshot at ψ = 0.
DR
Figure F.4: If we choose the angular velocity vector, ω ~ , as parallel to the IAOR then the velocity of the particle is given by
~ ×~r, where ~r is the position vector of the particle and the origin, O, is any point on the IAOR. [Figure 1-18 of Thornton & Marion]
~v = ω
of this circle is the IAOR. As a particle moves along the arc of a circle its angular position changes by dθ and its
angular speed is
dθ
ω= = θ̇.
dt
If the radius of the circular arc is R = r sin α then the particle’s linear speed is
v = Rω = rω sin θ.
The direction of the velocity vector, ~v , is orthogonal to the radius of the circular arc, R, and in the plane of the circular
arc.
It is convenient (Figure F.4) to define the angular velocity vector, ω~ , as along the instantaneous axis of rotation and
to have a magnitude given by θ̇ so that
~ × ~r.
~v = ω
187
T
and it rotates the unprimed basis {ê1 , ê2 } by a counter-clockwise angle α into the primed basis {ê01 , ê02 }. The vector ~v
is unchanged by this rotation of the basis vectors, only its representation has changed. This is known as an passive or
alias transformation (i.e. the vector itself is unchanged).
A passive transformation describes the same vector in two different coordinate systems.
There also exist active transformations in which the vector itself is rotated and the basis remains unchanged. The
two types of transformations are closely related mathematically but different from a physical point of view.
In particular, an active rotation would rotate the vector ~v by a counter-clockwise angle α into the vector ~v 0 ,
cos θ
AF
where
~v = v(ê1 cos θ + ê2 sin θ) = ê1 , eˆ2 v
cos θ cos α − sin θ sin α
= ê1 , eˆ2 v
Λ(−α) = ΛT (α) ≡
ï
sin θ cos α + cos θ sin α
cos α − sin α cos θ
= ê1 , eˆ2 v
ï òï
sin α cos α sin θ
ï
cos θ
sin θ
ò
cos(−α)
− sin(−α)
ò
ï
ò
sin θ
sin(−α)
cos(−α)
ò
ò ï
=
ï
sin(θ + α)
cos α
sin α
ò
− sin α
cos α
ò
is again a rotation matrix (as defined earlier) but it is here rotating the vector by +α instead of the basis by −α.
When the vector ~v is rotated by a counter-clockwise angle α into the vector ~v 0 its components in the {ê1 , ê2 } basis
transform as
~v 0 = ΛT (α)~v ,
DR
which is an active transformation. We should compare to the result of a passive transformation,
ï 0ò ï ò
ê1 ê1
= Λ(α) ,
ê02 ê2
2 Thisis one of those cases in which the minus sign is incredibly important. Consequently, a mistake will be a serious mistake rather than the
usual minor mistake that most sign errors are.
188 APPENDIX F. ROTATIONS
T
AF
DR
T
Appendix G
Vector Calculus
AF
Vector calculus is a subset of tensor analysis for the special case of rank-1 tensors. It is worthwhile reviewing vector
calculus using index notation prior to examining tensor calculus.1
where ∂i ≡ ∂x
∂
i
is the partial derivative with respect to xi .
The gradient of the scalar field f (~r ) is then
~ (~r ) = êi ∂i f.
∇f
df ∂f dxi ∂f
= = ẋi = ẋi ∂i f,
dt ∂xi dt ∂xi
1 Another notation that may be encountered is matrix calculus. With this notation the gradient of a scalar function f (~
r ) is represented as
h i
~ = ∂f ∂f ∂f
∇f ··· →
∂x1 ∂xn ∂~
r
and the Jacobian matrix of the vector function ~v (~
r ) is represented by
∂v ∂v1
1
···
∂x1 ∂xn
↔
J =
.. .. .. d~v
. . . →
d~r
.
∂vm ∂vm
···
∂x1 ∂xn
This notation is convenient as the product and chain rules come out similar to those for scalar derivatives.
189
190 APPENDIX G. VECTOR CALCULUS
where ∂i ≡ ∂x ∂
i
is again the partial derivative with respect to xi . This is the chain rule.
In vector notation I have
df
= ~r˙ · ∇f
~
dt
where ~r = êi xi .
T
G.0.3 Partial & Total Derivatives
Recall that the distinction between a partial derivative and a total derivative is very important. If I have a function
y(x, t) = y(x(t), t)
∂y ∂y(x, t)
AF dy
dt
=
∂y dx ∂y dt
∂x dt
dy
dt
+
∂t dt
=
∂t
∂y
∂x
~ · d~r = ∇y
= ∇y
dt
=
ẋ +
∂t
in which x is considered a constant, while the total derivative is
∂y
∂t
= ∂t y,
or ẏ = ẋ ∂x y + ∂t y.
Clearly this differs from the partial derivative due to the inclusion of the ẋ ∂x y term.
If I have, instead, a function of ~r,
~ · ~r˙ = ∂y ẋi
∂xi
Relativity
It is easy to see the convenience of replacing t with x0 = ct and have the indices run from 0 to 3 instead of 1 to 3. The
above would become
y = y(xi ), i = 0...3
and the total derivative with respect to x0 would be
ẏ = ẋi ∂i y, i = 0...3
in which ẋ0 = c dt
dt
= c.
Jacobians
If, instead, I have a vector quantity, vj = vj (xi ), then the gradient of this would be ∂i vj which is exactly the Jacobian
matrix. The gradient is the generalization of the concept of a derivative to higher dimensions.
191
The divergence is the contraction of the gradient, ∂i vi , which can be used with the divergence theorem (here dτ is the
infinitesimal volume element),
Z I Z I
T
dτ ∇~ · ~v = dA~ · ~v → dτ ∂i vi = dai vi ,
V ∂V V ∂V
AF
Example 12.8 The Radius Vector.
Some identities involving the radius vector, ~r = xi êi , are particularly easy to prove using index notation. Write each
of the following in index notation and then prove the relationship. (a) ∇r
Solution
and therefore
∂ 2
∂xi
êi
r = 2r
∂r
∂
∂xi
∂xi
=
(r) = êi
∂xj
∂xi
∂ √
∂xi
xj + xj
∂xj
∂xi
~ = ~r , (b) ∇
( xk xk ) = êi √
xi
r
xk xk
∂r xi
~ · ~r = 3,
= êi
xi
r
=
∂xi r
Solution
(a) First, consider the ith component
~ u · ~v )]i = ∂i (uj vj ) = uj (∂i vj ) + vj (∂i uj )
[∇(~
T
~ × ~v ). The k th component of the quantity in parenthesis is [∇
Next, consider ~u × (∇ ~ × ~v ]k = εk`m ∂` vm , so that
Similarly,
~ × ~u)]i = vj (∂i uj ) − (~v · ∇)u
~ i
AF
(b) I have
(c) I have
~ × ~v ) + ~v × (∇
[~u × (∇
[~v × (∇
so the sum of these last two expressions is
~ × ~u)]i = uj ∂i vj + vj ∂i uj − (~u · ∇)v
⇒ ∇(~
~ u · ~v )]i − [(~u · ∇)~
= [∇(~
~ u · ~v ) = (~u · ∇)~
~ v + (~v · ∇)~
~ i − (~v · ∇)u
~ v + (~v · ∇)~
~ u + ~u × (∇
~ u ]i
~ i
~ × ~v ) + ~v × (∇
~ × ~u)
Linear Algebra
AF
H.1
Resources
Thornton & Marion §1.1-17
Boas §3.1-12
Boas §4.1-12
where 1 is the unit matrix. For a non-trivial solution (i.e. ~v 6= 0) the matrix K − λ1 must be singular (because if it
DR
wasn’t then the inverse would exist and multiplying both sides by this inverse would leave me with ~v = 0). The matrix
is singular only if
det (K − λ1) = 0.
This yields a polynomial in λ of degree n, where n × n is the size of the matrix K. The roots of this polynomial are
the eigenvalues.
For each eigenvalue, λi , we go back to the defining equation, (K − λi 1)~vi = 0, to find the corresponding eigen-
vectors, ~vi , to within an overall multiplicative constant (which we choose so that the eigenvectors have unit length).
Things can become slightly more complicated if some of the eigenvalues are complex and/or repeated.
H.1.1 Diagonalization
We want to find the matrix U such that the quantity U−1 KU is diagonal. If the columns of U are the eigenvectors of K,
then U−1 = UT , the transpose of U since the eigenvectors form an orthonormal set (i.e. ~vi · ~vj = δij ).
Now we have
KU = [λ1~v1 λ2~v2 ... λn~vn ]
193
194 APPENDIX H. LINEAR ALGEBRA
T
.. ..
. . . .
0 0 0 ··· λn
K~v = λ~v
AF
which I wish to solve. I first multiply both sides of the equation by UT , the matrix whose rows are the orthonormal
eigenvectors of K,
where
UT K~v = λUT~v ,
and then insert the identity, 1 = UUT , between K & ~v ,
0
K0 ≡ UT KU =
0
0
λ2
UT KUUT~v = UT KU UT~v = λ UT~v .
K0~v 0 = λ~v 0
λ1 0
0
0
0
0
..
0
.
0
0
0
λn
is now diagonal (with the eigenvalues as diagonal elements), and ~v 0 ≡ UT~v , is the vector ~v in the basis defined by the
orthogonal eigenvectors. In this new basis the resulting equation is trivially solvable.
Example 12.10 Similarity Transformations .
DR
(a) Find the eigenvalues and eigenvectors of the matrix
ï √ ò
2 3
A= √
3 4
and plot the normalized eigenvectors in the plane. (b) Construct the matrix U whose columns are the normalized
eigenvectors of A and calculate the product U † AU where U † is the Hermitian conjugate of U (ie, the transpose of
the complex conjugate). Such a transform is called a “similarity transform”. (c) Calculate U † U and U U√ †
and discuss
~ ~
what the result tells you about the the eigenvectors of A. (d) Solve the problem A~x = b where b = (1, 3) by direct
†
multiplication and then solving the resulting system of equations. Take ~x† = (x1 , x2 ). (e) Instead now solve A~x = ~b
by multiplying each side of the equation by U † and inserting U U † between A and ~x and then calculating U † AU . The
vectors ~x0 = U † ~x and ~b0 = U †~b represent the vectors in the basis defined by the eigenvectors of A. Solve the problem
for ~x0 in this basis and plot the results in the same diagram used for part (a). Now transform back to the unprimed
system, ~x = U~x0 , to find the solution to the original problem. (f) Discuss the relative difficulty of the two methods of
solving A~x = ~b if instead of A being a 2 × 2 matrix it was very large.
Solution
(a) The eigenvalues are ?? 1 and the eigenvectors are ?? 1 and
??. 1
H.1. EIGENVALUES & EIGENVECTORS 195
T
1
(d) I have
?? = ??
with solution
?? = ??. 1
AF where
and we have
and, finally,
A0 = U † AU = ??
~x0 = ???? = ??
~x = U~x0 = ??.
2
(f) If the matrix A was very large and we had to find the solutions ~x for many different vectors ~b then finding
X
T
AF
DR
T
Appendix I
State Vectors
AF
Consider two dimensional trajectory motion with a drag force proportional to the square of the velocity and anti-
parallel to it. We then have for Newton’s second law,
or, as components,
m~r¨ = F~g + F~D = m~g + mkv 2 (−v̂)
where m~g is the downwards force due to a uniform gravitational field. We will take initial conditions ~r(0) = ~ro and
~r˙ = ~vo where ~vo is a constant that makes an angle θ to the horizontal.
Because there is no reason to do otherwise, we will put our origin at ~ro and orient our coordinate system so that
~g = −gê2 and ~vo = vo (cos θê1 + sin θê2 ), giving for Newton’s second law,
ẍ1 = −k ẋ1
which are a pair of coupled second order ODEs.
»
ẋ21 + ẋ22 & ẍ2 = −g − k ẋ2
ẋ1 + ẋ22
»
ẋ21 + ẋ22
In the usual way we introduce v1 = ẋ1 and v2 = ẋ2 so that we have a system of four coupled first order ODEs,
ẋ1 = v1
DR
»
v̇1 = −k ẋ1 ẋ21 + ẋ22
ẋ2 = v2
»
v̇2 = −g − k ẋ2 ẋ21 + ẋ22 .
We can see how this could be extended to three dimensional problems or even to several interacting particles, but
that it would become quite cumbersome. We can also see that the first three lines contain the physics and the last four
lines contain four repetitions of the Euler algorithm.
197
198 APPENDIX I. STATE VECTORS
A core principle of good software design is “don’t repeat yourself” (DRY). We can introduce an abstraction that
adheres to this principle by introducing a state vector,
~i = (x1,i , x2,i , v1,i , v2,i )
ψ
T
~i+1 = ψ
ψ ~˙ i
~i + ∆tψ
where
~˙ i = (ẋ1,i , ẋ2,i , v̇1,i , v̇2,i ) = (v1,i , v2,i , a1,i , a2,i )
ψ
is the time derivative of the state vector. Two elements are, of course, trivial (i.e. ~r˙ = ~v ) but the other two contain (or
1 P ~
encapsulate) all of the physics (i.e. Newton’s second law, ~a = m F ).
This abstraction, then, both reduces repetition in the Euler algorithm and encapsulates the physics in a function,
ψ
AF
~˙ i (ψ
~i , t). This has an even further benefit in that we can now use algorithms that are more sophisticated than Euler’s
without delving into their details. (In fact, Euler’s algorithm performs particularly poorly for oscillatory motion so we
will need to examine alternative algorithms shortly.)
Computationally, we will use and N × 2n array to store the state vectors where N is the number of times steps
and 2n is the number of first order ODEs (and n is the number of degrees of freedom). There will be one row for each
time, ti , i = 0...N − 1 and one column for each first order ODE.
Consider the problem of trajectory motion in the presence of a drag force proportional to the square of the velocity,
Fdrag = −kv 2 , as well as gravity, m~g . The direction of the drag force is opposite to that of the velocity, F~ = −kv 2 v̂ =
−kv~v . Newton’s second law gives:
m~x¨ = −kv~x˙ + m~g ,
where ~x˙ = ~v is the velocity. I choose x̂1 in the horizontal direction and x̂2 in the vertical direction so that In terms of
the Cartesian components I have
DR
ẍ1 = −kvv1
ẍ2 = −kvv2 − mg.
The following pseudo-code implements the algorithm for calculating the derivative of the state-vector,
~˙ ψ,
function ψ( ~ t, β)~
~
# accepts ψ = (x1 , x2 , v1 , v2 ), time t,
# and parameters β ~ = (g, k)
~
(x1 , x2 , v1 , v2 ) = ψ # the state vector
(g, k)p= β ~ # the parameters
v = v12 + v22 # the speed
a1 = −kvv1 # Newton’s second law
a2 = −g − kvv2 # note that v 2 v̂ = v~v
~˙ = (v1 , v2 , a1 , a2 )
ψ # time derivative of the state vector
~˙
return ψ
~i = ψ(t
(a) Use Python to implement the algorithm to find ψ ~ i ) where ti = i ∆t, i = 0...N − 1. (b) Plot ẋ(t) and
the trajectory, x2 (x1 ).
199
1 #================================== constants
2 m = 1.0 # mass, m [kg]
3 g = 10.0 # gravity, g [m/s2 ]
4 k = 0.1 # drag force per unit mass, k [N/m]
5 x_10 = 0.0 # initial height, x1,0 [m]
6 x_20 = 0.0 # initial height, x2,0 [m]
7 v_0 = 100 # initial speed, v0 [m/s]
8 theta = pi / 6 # initial angle, θ = π/6 [radians]
T
9 N = 100 # number of time steps [dimensionless]
10 tN = 3.0 # maximum time, tN [s]
11 #================================== calculated constants
12 Dt = tN / N # time step size, ∆ t [s]
13 v_10 = v_0*cos(theta) # initial horizontal velocity, v1,0 = v0 cos θ [m/s]
14 v_20 = v_0*sin(theta) # initial vertical velocity, v2,0 = v0 sin θ [m/s]
15 t = arange( N ) * Dt # length-N array, ti = i∆ t, i = 0, 1, 2, ..., (N − 1)
p
16 vT = sqrt( g / k ) # the terminal velocity, v terminal = g/k
17 beta = ( g, k ) # parameter vector
18 #================================== variables
19 ~
# Initialize an N × 4 array with zeros for the state vector, ψ(t) = [ x1 (t), v1 (t), x2 (t), v2 (t) ],
20 # one row for each of the N times, ti , and one column for each xi and vi .
21 psi = zeros( [ N, 4 ] )
~ ~o , the values at t=0.
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
AF # The zeroth row contains the initial condition, ψ(0)
psi[0,:] = x_10, x_20, v_10, v_20
"""
x1, x2, v1, v2 = phi
g, k = beta
v = sqrt( v1*v1 + v2*v2 )
a1 = - k * v * v1
a2 = - g - k * v * v2
= ψ
~˙ ψ,
function ψ( ~ t, β)
~
~ ~ = (M♁ , G)
# accepts ψ = (x1 , x2 , v1 , v2 ), time t, and parameters β
200 APPENDIX I. STATE VECTORS
100 10
8
80
6
T
60 4
v [m/s]
x2 [m]
2
40 0
−2
20
−4
0 −6
AF 0.0 0.5
~
(x1 , x2 , v1 , v2 ) = ψ
(M♁ ,p G) = β
r = x21 + x22
a = GM♁ /r2
a1 = −a x1 /r
a2 = −a x2 /r
~
1.0
~˙ = (v1 , v2 , a1 , a2 )
ψ
~˙
return ψ
1.5
t [s]
2.0 2.5
# the parameters
# the radius
3.0
trajectory, x2 (x1 ), for a circular orbit and (c) for an elliptical orbit.
DR
T
Appendix J
Lagrange Multipliers
AFResources
§6.6 of Thornton & Marion
§4.9 of Boas
Wikipedia : Lagrange multiplier
Consider the problem of minimizing a function, f (xi ), i = 1...n, subject to the constraint
df =dxi ∂i f
dG =dxi ∂i G.
which, along with the constraint equation, G = 0, forms a set of n + 1 equations in n + 1 unknowns, xi & λ, which
can then be solved (although we usually don’t care about the value of λ).
This problem is identical to that of find the minimum of the auxiliary function
in which λ is the Lagrange multiplier (i.e. the gradients are parallel but not necessarily equal).
Now consider the new auxiliary function,
201
202 APPENDIX J. LAGRANGE MULTIPLIERS
g(x,y) = c
T
f(x,y) = d1
f(x,y) = d2
f(x,y) = d3
AF
which the two gradients, ∇f
Nexcis from wikipedia]
To find the extrema of a function f (x, y) subject to the constraint G(x, y) = 0, first form the function
These will be a set of three equations in three unknowns, x, y, and λ, which can then be solved. Note that the third equation,
∂F
∂λ
= 0, always results in the constraint equation.
The quantity λ is known as the Lagrange multiplier and it often has a physical significance.
DR
and its gradient,
~ x ,x ,λ F = ∇
∇ ~ x ,x f + λ∇
~ x ,x G + Gλ̂
1 2 1 2 1 2
~ x ,x ,λ F = 0 then
so if ∇ 1 2
~ x ,x f = −λ∇
∇ ~ x ,x G & G = 0
1 2 1 2
4.0
3.5
3.0
T
2.5
y/a
2.0
1.5
1.0
0.5
0.0
Figure J.2: In Example 12.12 we wish to find the shortest distance from the origin to the parabola y = a(1 − (x/a)2 )
Solution
(a) Using y = a(1 − 2(x/a)2 ) to eliminate y from f I get
f0 ≡
df
dx
√
= 2x(2(x/a)2 − 1) = 0 ⇒ x ∈ {0, ±a/ 2}.
It is not immediately obvious which of these points is a maximum and which is a minimum, so in this simple
problem it is worth while to find the second derivative:
DR
−2 at x = 0 (relative maximum)
®
d2 f
00
f ≡ = 12(x/a)2 − 2 = √
dx2 6a2 at x = ±a/ 2 (minimum).
√
The minimum I want then occurs at x = ±a/ 2, y = a/2.
dy
f 0 = 2x + 2y ,
dx
dy
and then, from y = a(1 − (x/a)2 ), I find the derivative y 0 ≡ dx ,
2x
y0 = − .
a
Substituting the second into the first, I get
4yx
f 0 = 2x − .
a
df
To minimize f , I set dx = 0 (or in the differential notation I set df = 0 for arbitrary dx). This gives
2x − 4xy/a = 0.
204 APPENDIX J. LAGRANGE MULTIPLIERS
This equation must now be solved simultaneously with the equation of the curve y = a(1 − (x/a)2 ). I get
√
3x − 4ax(1 − (x/a)2 ) = 0 ⇒ x ∈ {0, ±a/ 2},
as before.
T
To test for maxima or minima I need f 00 . Differentiating f 0 = 2x + 2yy 0 I get
2
f 00 = 2 + 2(y 0 ) + 2yy 00 .
f 00 (0) = −2a;
this point is the required minimum. Notice in particular here that we could do every step of (b) even if the equation
of the curve could not be solved for y.
Solution
(a) I have
f = x2 + (a2 /x)2 = x2 + a4 /x2
which has a minimum when
df 2
= 0 = 2x − 2a4 /x3 ⇒ 3 (x4 − a4 ) = 0 ⇒ x = ±a
dx x
(b) I have
F (x, y, λ) = x2 + y 2 + λ y − a2 /x
205
so
∂F
= 0 =2x + λa2 /x2
∂x
∂F
= 0 =2y + λ ⇒ λ = −2y
T
∂y
∂F
= 0 =y − a2 /x.
∂λ
The second tells me that λ = −2y so that the first gives
1/3
2x3 = −λa2 = 2a2 y ⇒ x = a2 y
and so
a2 a6 a4
y = a2 /x = ⇒ y 3
= = ⇒ y = ±a
AF
Example 12.14 Multiplier.
(a2 y)1/3 a2 y y
(Hint: Form the function F (x, y, λ) = xy 2 + λ (x2 + y 2 − a2 ) and calculate its gradient, ∇
Solution
(a) I have
so
F (x, y, λ) = xy 2 + λ (x2 + y 2 − a2 )
∂F
∂x
∂F
= 0 =y 2 + 2λx
X
Find the extrema on the surface f (x, y) = xy 2 subject to the constraint x2 + y 2 = a2 using (a) the method of
substitution, (b) the method of implicit differentiation, and (c) the method of Lagrange multipliers.
~ x,y,λ F = 0.)
= 0 =2xy + 2λy
∂y
∂F
= 0 =x2 + y 2 − a2
DR
∂λ
and there are six critical points:
√ √ √ √ √ √
( 2, 1); (− 2, 1); ( 2, −1); (− 2, −1); (0, 3); (0, − 3).
T
AF
DR
T
Appendix K
Legendre Transformations
AFResources
Thornton & Marion §7.10
Boas §11.4
Wikipedia : Legendre Transformation
arXiv.org : Making Sense of the Legendre Transform
A Legendre transformation converts a function of one variable into a function of another variable in a manner
similar to a Fourier or Laplace transformation. Each pair of variables, said to be conjugate to each other, contain the
same information but it is often easier to manipulate or control one variable than the other.
The Legendre transformation of the function f (x) is defined as1
df
s(x) ≡
df
dx
where x(s) is the inverse of s(x) (note that s(x) = dx must be a monotonically increasing function of x so that the
inverse, x(s), is a well defined function – see Figure K.1).
DR
1 Definitions with differing signs are possible and common. This version is particularly elegant as the inverse Legendre transformation is is then
y
y = f (x)
s = f 0 (x)
xs
f (x)
−g(s) x x
Figure K.1: The Legendre transformation of the function f (x) is g(s) = s x(s) − f (x(s)), where x(s) is the inverse of the
function s(x) ≡ f 0 (x). This can be visualized graphically by considering the y intercept of the tangent to y = f (x) – each
value of f (x) is mapped to −g(s) which is this intercept. This, of course, is only possible if f 0 (x), the slope of f (x), is a
steadily increasing function of x.xx
207
208 APPENDIX K. LEGENDRE TRANSFORMATIONS
T
ds ds dx ds ds
One of the most interesting properties of the Legendre transformation is that its inverse,
dfÊ
f (x) ≡ x s(x) − fÊ(s(x)), x(s) ≡
ds
is also a Legendre transformation.
To show this I write the total derivatives
AF df
dx
Ä
df = d x s(x) − fÊ(s(x))
ds
dx = s dx + x dx −
dx ⇒ s(x) =
df
dx
.
Find, fÊ(s) = xs − f , the Legendre transformation, and f (x) = xs − fÊ, the inverse Legendre transformation, of each
of the following: (a) f (x) = x4 (b) f (x) = ex (c) f (x) = 12 x2 .
Solution
1 4/3
fÊ(s) = s x(s) − f (x(s)) = √
1/3
I now have x = fÊ0 = 3(4/3)(s/4) (1/4) = (s/4)
1/3
3
4
4/3
s − (s/4)4/3 = 3(s/4) .
4/3
f (x) = x s(x) − fÊ(s(x)) = 4x4 − 3 4x3 /4 = x4 . X
DR
(b) I have s = f 0 = ex so x(s) = ln s and
fÊ(s) = s ln s − eln s = s ln s − s.
f (x) = x2 − 12 (x)2 = 21 x2 . X
So this function is special for the Legendre transformation, playing the same role as the Gaussian for the Fourier
transformation.
X
209
Alternate Form
An alternate form of the Legendre transformation is
fÊ = xs − f xf 0 − f
T
fÊ f0 f d f
Å ã
= − 2 =
x2 x x dx x
d f
Å ã
2
⇒ f (s) = x
Ê ,
dx x
in which it is understood that the substitution x = x(s) is performed after the differentiation.
Example 12.16 Alternate Legendre Transformations. Ä ä
d f
Recalculate the Legendre transformations from Example 12.15 using the alternate form, fÊ(s) = x2 dx x .
AF
Solution
(a) With x(s) =
p
3
s/4, I have
f (s) = x
Ê 2 d
dx x
d x4
Å ã
dx x
x 2 e
x
fÊ(s) = x2
x
d
dx
ã
» 4
= 3x4 = 3 3 s/4 = 3(s/4)
Å xã Å x
e ex
4/3
− 2 = xex − ex = s ln s − s X
Ç1
2x
x
2
å
= x2 ( 21 ) = 12 s2 X
X
2n Variables
DR
I start with a function f (~x, ~v ) in which ~x and ~v are n dimensional vectors with components xi and vi , respectively. If
the vi are, for some reason, inconvenient variables I introduce
∂f
pi (~x, ~v ) ≡ .
∂vi
∂pi ∂2f
in which I have eliminated the ~v on the right hand side, which is possible if the determinant of the matrix ∂vj = ∂vi ∂vj
is non-zero.
The total derivative of the above is
so
∂ fÊ
vi =
∂pi
∂f ∂ fÊ
T
=−
∂xi ∂xi
∂f ∂ fÊ
This last expression, ∂xi= − ∂x i
, is more subtle that it first appears as the variables held constant are different for the
two partial derivatives (i.e. f and fÊ are functions of different variables).
Example 12.17 More Legendre Transformations.
Find, L(~
Ê x, p~) = p~ · ~v − L(~x, ~v ), the Legendre transformation, and L(~x, ~v ) = p~ · ~v − L(~
Ê x, p~), the inverse Legendre
transformation, of each of the following: (a) L(~x, ~v ) = 2 m~v · ~v − 2 k(|~r| − a) , where ~r = êi ri & ~v = êi vi .
1 1 2
~ r)
(b) L(~x, ~v ) = 12 m~v · ~v − eφ(~r) + ec ~v · A(~
AF
Solution
(a) I have pi =
(b) I have pi =
∂L
∂vi
∂L
∂vi
= mvi so vi = pi /m and
= mvi + ec Ai so ~v = 1
m
= 2m
1
1
Ä
1
p~ − ec A
Ê x, p~) = ~v · m~v + e A
L(~ c
ä
mm
Ê x, p~) = pi pi − 1 m pi pi + 1 k(|~r| − a)2
L(~
m 2
2
= 2m p~ · p~ + 2 k(|~r| − a) .
Ä ä
~ and
2
~ − 1 m~v · ~v + eφ − e ~v · A
2
= 12 m~v · ~v + eφ
ä2
~ + eφ
p~ − ec A
c
~
dU = dQ + dW.
The additional heat changes the temperature, T , and entropy, S, of the system dQ = T dS, and the work done on the
system changes the pressure, P , and volume, V , of the system dW = −P dV (negative so that pressure is a positive
definite quantity).
K.1. THERMODYNAMIC POTENTIALS 211
I therefore have
dU = T dS − P dV.
However, the entropy of a system is often difficult to measure experimentally so the Helmholtz free energy,
H(S, P ) ≡ U + P V , is often used instead. This is a Legendre transformation,
T
dH = d(U + P V )
∂H ∂H
dS + dP = T dS − P dV + P dV + V dP
∂S ∂P
= T dS + V dP
so that
∂H ∂H
=T & = V,
∂S ∂P
both of which are eminently accessible by experiment. For a system at constant pressure, dP = 0, I have
AF dW
dH = dQ.
For example, the transformation from internal energy, U , to Helmholtz free energy, H, is found using the first and
second laws of thermodynamics, dU = dQ − dW , (where Q is heat energy and W is the work done) and dQ = T dS
(where T is the absolute temperature and S is the entropy). I have for the enthalpy, H(S, P, N ) = U + P V , so
dH = dU + P dV + V dP
= (dQ − ) +
P
= T dS + V dP
dV
+ V dP
One of the variables in each conjugate pair is intrinsic (i.e. proportional to the size of the system e.g. V , S, N ), and
the other is extrinsic (i.e. is independent of the size of the system e.g. P , T , or µ).
DR
212 APPENDIX K. LEGENDRE TRANSFORMATIONS
T
AF
DR
T
Appendix L
Table of Integrals
AF
Trigonometry
cos x = 21 (eix + e−ix )
sin x = 1
2i (e
ix
− e−ix )
ex = cosh x + sinh x
(L.1)
(L.2)
(L.3)
(L.4)
(L.5)
(L.6)
2 tan θ
tan 2θ = (L.13)
1 − tan2 θ
213
214 APPENDIX L. TABLE OF INTEGRALS
T
2 sin A sin B = cos(A − B) − cos(A + B) (L.18)
Series
∞ Å ãn
1 1 X r0
= Pn (cos α), cos α = r̂ · r̂0 (L.19)
r r n=0 r
X∞ n
(n)
AF f (x◦ + ) =
n=0
n!
f (x◦ ) = f (x◦ ) + f 0 (x◦ ) + 12 2 f 00 (x◦ ) + · · ·
(1 + x)n = 1 + nx +
ex = 1 + x +
sin x = x −
cos x = 1 −
1 2
tan x = x + x3 + x5 +
3!
1 2
2!
17 5
n(n − 1) 2
1 2
2!
1
2!
x + ···
x + ···
ln(1 ± x) = ±x − 21 x2 ± 31 x3 − 14 x4 + · · ·
1 3
x + x5 − · · ·
5!
1
x + x4 − · · ·
4!
x + ···
(L.20)
(L.21)
(L.22)
(L.23)
(L.24)
(L.25)
(L.26)
3 15 315
1 1 1 2 5
cot x = − x − x3 − x − ··· (L.27)
x 3 45 945
DR
Integrals
Indefinite
Z
1
dx xn = xn+1 , (n 6= −1) (L.28)
n+1
Z
dx
= ln x (L.29)
x
Z
1
dx sin(ax) = − cos(ax) (L.30)
a
Z
1
dx cos(ax) = sin(ax) (L.31)
a
Z
x 1
dx cos2 (ax) = + sin(2ax) (L.32)
2 4a
Z
x 1
dx sin2 (ax) = − sin(2ax) (L.33)
2 4a
215
Z
dx 1 x
= tan−1 (L.34)
x2 + a 2 a a
Z
dx 1 a+x
Å ã
= ln (L.35)
a2 − x2 a−x
T
2a
Z
dx x 1
2 2
= ± ln a2 ± x2 (L.36)
a ±x 2
Z
dx x ∓1
2 2 n+1
= (L.37)
(a ± x ) 2(a ± x2 )n
2
Z p
dx Ä ä
= ln x + x2 ± a2
√ (L.38)
x2 ± a2
Z
AF Z
√
p
dx x2
x2 + a2
Z
Z
√
Z
Z
2
Ä p
1 p
dx x3
√
x2 + a 2
√
dx
2
x a −x
dx
Z
− a2
dx x
x2 + a2
2
1
=
1
Ä
|a|
Ç
p
dx x2 ± a2 = 12 x x2 ± a2 ± a2 ln x + x2 ± a2
= sin−1 (x/a)
√
x2
= x2 + a2
p
p
= x x2 + a2 − a2 ln 2 x2 + a2 + 2x
= (x2 − 2a2 ) x2 + a2
3
sec−1 (x/a)
√
ää
å
(L.39)
(L.40)
(L.41)
(L.42)
(L.43)
(L.44)
dx 1 a + a2 ± x2
√ = − ln (L.45)
x a2 ± x2 a x
Z
DR
dx x
= √ (L.46)
(x2 + a2 )3/2 a x2 + a2
2
Z
dx x −1
2 2 3/2
=√ (L.47)
(x + a ) x2 + a 2
Definite
Z π Z π
π
dx sin(nx) sin(mx) = dx cos(nx) cos(mx) = δn,m (L.48)
0 0 2
Z 2π Z 2π
dx cos2 x = dx sin2 x = π (L.49)
0 0
Z ∞
b
dx e−ax sin bx = (L.50)
0 a2 + b2
Z ∞
a
dx e−ax cos bx = (L.51)
0 a2 + b2
Z ∞
dx 2
= 2 (L.52)
−∞ (x2 + a2 )3/2 a
216 APPENDIX L. TABLE OF INTEGRALS
Z ∞
dx x 1
= 4 (L.53)
0 (x2+a )2 3 4a
Z ∞
dx x3 1
= 2 (L.54)
T
0 (x2+ a2 )3 4a
AF
DR
T
Appendix M
Formulas
AF
M.1
M.1.1
Analytical Mechanics
d
dt
d
dt
Lagrangian Mechanics
∂T
∂ q̇a
∂L
∂ q̇a
−
−
∂T
∂qa
∂L
T =
Qa =
= Qa
∂qa
L=T −V
=0
X
N
r¨ · δ~
~ (a) − m~
δW = F
i=1
X
N
i=1
1
r=0
r˙
m~
2 i i
~i · ∂~
F
ri
∂qa
r˙i
·~
M.1.4
M.1.5
Non-inertial Systems
~centrifugal = −mΩ
F
~Coriolis = −2mΩ
F
Inertia Tensor
Z
00
keff = −Veff
~ × Ω
(x)
~ × ~
r˙
~ ×~
~
r+R
00
+ 12 (x − x◦ )2 Veff
0
Veff (x) = Veff (x◦ ) + (x − x◦ ) Veff (x◦ )
(x◦ ) + · · ·
M~c = dm ~
r
Z
M.1.2 Hamiltonian Mechanics Iij = dm(xk xk δij − xi xj )
DR
∂L Iij = I ◦ + M (c2 1 − ~c ⊗ ~c)
pa = ◦
= Iij + M (ck ck δij − ci cj )
∂ q̇a
X
f
H= pa q̇a − L
a=1
ṗa = −
∂H
& q̇a =
∂H M.1.6 Rotational Dynamics
∂qa ∂pa
M.1.3 Small Oscillations ~ = φ̇ sin ψ sin θ + θ̇ cos ψ x̂
ω
+ φ̇ cos ψ sin θ − θ̇ sin ψ ŷ
X
2
∂~ ri
ri ∂~ + φ̇ cos θ + ψ̇ ẑ
Ta,b = mi ·
∂qa ∂qb 0 ñ
I11 ω̇1
ô
N1 + ω2 ω3 (I22 − I33 )
ñ ô
i=1
I22 ω̇2 = N2 + ω3 ω1 (I33 − I11 )
∂ 2 V I22 ω̇3 N3 + ω1 ω2 (I11 − I22 )
Va,b =
∂qa ∂qb
0 r0 = R
~ ~0 + ~
r
X
f
~˙ 0 + (~
Ta,b q̈b + Va,b qb = 0 r˙ 0 = R
~ r˙ + ω
~ ×~
r)
b=1 ~ ~¨ 0 + ~
r¨ 0 = R r¨ + ω
~˙ × ~
r + 2~ r˙ + ω
ω ×~ ~ × (~
ω ×~
r)
217
218 APPENDIX M. FORMULAS
Euler Angles
For a zxz Euler rotation,
ê1 cos ψ sin ψ 0 1 0 0 cos φ sin φ 0 ê01
ê2 = − sin ψ cos ψ 00 cos θ sin θ − sin φ cos φ 0ê02
T
ê3 0 0 1 0 − sin θ cos θ 0 0 1 ê03
cos ψ cos φ − cos θ sin ψ sin φ cos ψ sin φ + cos θ cos φ sin ψ sin ψ sin θ ê01
= − cos φ sin ψ − cos ψ cos θ sin φ cos ψ cos θ cos φ − sin ψ sin φ cos ψ sin θê02
sin θ sin φ − cos φ sin θ cos θ ê03
Ω~ =φ ~˙ = φ̇ ê0 = φ̇(ê3 cos θ + sin θ(ê1 sin ψ + ê2 cos ψ))
3
~˙
θ = θ̇ (ê1 cos ψ − ê2 sin ψ) = θ̇(ê01 cos φ + ê02 sin φ)
~n = ψ ~˙ = ψ̇ ê3 = ψ̇(ê0 cos θ + sin θ(ê0 sin φ − ê0 cos φ)),
(snapshot at ψ = 0) Ä
3
ä
1
Ä
2
ä Ä ä
~ = ê01 ψ̇ sin θ sin φ + θ̇ cos φ + ê02 −ψ̇ sin θ cos φ + θ̇ sin φ + ê03 ψ̇ cos θ + φ̇
ω
AF
M.2
M.2.1
z
Helpful Identities
Cylindrical Coordinates
~r = ~r(s, φ, z)
~
r
ẑ
φ̂
ŝ
Ä
M.2.2
x
z
φ
ä
θ
Ä
Spherical Coordinates
~r = ~r(r, θ, φ)
~
r
y
r̂
φ̂
θ̂
ä Ä
= ê1 φ̇ sin θ sin ψ + θ̇ cos ψ + ê2 φ̇ sin θ cos ψ − θ̇ sin ψ + ê3 φ̇ cos θ + ψ̇
ä
" #
cos θ cos φ cos θ sin φ − sin θ
z
ñ ôñ ô
θ̂ x̂
φ̂ = − sin φ cos φ 0 ŷ
y
r̂ sin θ cos φ sin θ sin φ cos θ ẑ
DR
φ s
r̂˙ = θ̇θ̂ + φ̇ sin θφ̂
x
˙
θ̂ = −θ̇r̂ + φ̇ cos θφ̂
˙
φ̂ = −φ̇(sin θr̂ + cos θθ̂)
r̂¨ = −θ̇2 − φ̇2 sin2 θ r̂ + θ̈ − φ̇2 sin θ cos θ θ̂
ñ ŝ ô
cos φ sin φ 0
ñ ô ñ ôñ ô
x̂ x̂
φ̂ = Λz (φ) ŷ = − sin φ cos φ 0 ŷ
ẑ ẑ 0 0 1 ẑ + 2θ̇φ̇ cos θ + φ̈ sin θ φ̂
r = rr̂
~
ŝ˙ = φ̇φ̂ r˙ = ṙr̂ + rθ̇θ̂ + rφ̇ sin θφ̂
~
˙
φ̂ = −φ̇ŝ
˙ẑ = 0 r¨ = r̈ − rθ̇2 − rφ̇2 sin2 θ r̂
~
+ 2ṙθ̇ + rθ̈ − rφ̇2 sin θ cos θ) θ̂
r = sŝ + z ẑ
~
r˙ = ṡŝ + sφ̇φ̂ + ż ẑ
~ + 2ṙφ̇ sin θ + 2rθ̇φ̇ cos θ + rφ̈ sin θ φ̂
r¨ = (s̈ − sφ̇2 )ŝ + (2ṡφ̇ + sφ̈)φ̂ + z̈ ẑ
~ r = r sin θ(cos φx̂ + sin φŷ) + r cos θẑ
~
r = s cos φx̂ + s sin φŷ + z ẑ
~ r˙ = ṙ sin θ + rθ̇ cos θ (cos φx̂ + sin φŷ)
~
r˙ = (ṡ cos φ − sφ̇ sin φ)x̂ + (ṡ sin φ + sφ̇ cos φ)ŷ + ż ẑ
~
+ rφ̇ sin θ (− sin φx̂ + cos φŷ)
r¨ = (s̈ cos φ − 2ṡφ̇ sin φ − sφ̈ sin φ − sφ̇2 cos φ)x̂
~
+ (s̈ sin φ + 2ṡφ̇ cos φ + sφ̈ cos φ − sφ̇2 sin φ)ŷ + z̈ ẑ + ṙ cos θ − rθ̇ sin θ ẑ
M.2. HELPFUL IDENTITIES 219
~ · (B
~ × C)
~ =B
~ · (A
~ × C)
~ =C
~ · (A
~ × B)
~ sin x = 1
2i
(eix − e−ix )
A
eix = cos x + i sin x
T
~ × (B
A ~ × C)
~ = (A
~ · C)
~ B~ − (A
~ · B)
~ C~
cosh x = 1 x
(e + e−x )
~v = (n̂ · ~v )n̂ + n̂ × (~v × n̂) = ~vk + ~v⊥ 2
~ · (∇
~ × A)
~ =0 sinh x = 1 x
(e − e−x )
∇ 2
x
~ × (∇f
~ )=0 e = cosh x + sinh x
∇
sin A sin B sin C
~ × (∇
∇ ~ × A)
~ = ∇(
~ ∇~ · A)
~ − ∇2 A
~ = =
a b c
a2 = b2 + c2 − 2bc cos A
sin2 θ + cos2 θ = 1
sec2 θ = 1 + tan2 θ
M.2.4 Tensor Analysis
AF εijk =
+1 if (i, j, k) is (1, 2, 3), (2, 3, 1), or (3, 1, 2),
−1 if (i, j, k) is (3, 2, 1), (1, 3, 2), or (2, 1, 3),
0 if i = j, or j = k, or k = i.
Gauss’ theorem,
Z
εijk εk`m = δi` δjm − δim δj` .
I
det(A) = a1i a2j a3k εijk
u × ~v = êi εijk uj vk
~
~ = êi ∂ = êi ∂i
∇
∂xi
Z I
M.2.6
f (x◦ + ) =
Series
X
∞
n
n=0
(1 + x)n = 1 + nx +
sin 2θ = 2 sin θ cos θ
cos 2θ = cos2 θ − sin2 θ
tan 2θ =
n!
1
2 tan θ
1 − tan2 θ
sin(A ± B) = sin A cos B ± cos A sin B
cos(A ± B) = cos A cos B ∓ sin A sin B
2 sin A cos B = sin(A + B) + sin(A − B)
2 cos A cos B = cos(A + B) + cos(A − B)
2 sin A sin B = cos(A − B) − cos(A + B)
n(n − 1) 2
2!
x + ···
~ · ~v =
dτ ∇ d~a · ~v → dτ ∂i vi = dai vi ex = 1 + x + x2 + · · ·
V ∂V V ∂V
2!
ln(1 ± x) = ±x − 12 x2 ± 13 x3 − 14 x4 + · · ·
DR
where ∂V is the boundary of the volume V. 1 3 1
sin x = x − x + x5 − · · ·
3! 5!
Stokes’ theorem, 1 1
cos x = 1 − x2 + x4 − · · ·
2! 4!
Z I Z I 1 2 5 17 5
~ ×~v = tan x = x + x3 + x + x + ···
d~a · ∇ r ·~v
d~ → dai εijk ∂j vk = dxi vi , 3 15 315
S ∂S S ∂S 1 1 1 3 2 5
cot x = − x − x − x − ···
x 3 45 945
where ∂S is the boundary of the surface S.
220 APPENDIX M. FORMULAS
T
AF
DR
T
Glossary
AMPLITUDE VECTOR The complex amplitudes, ãb , which determine the linear combination of the normal modes,
η̂b (t), that will satisfy the BCs. The motion of the system is then described by the real part of
n
X
BC
AF
COM
DOF
ELE
BC s.. 32
A system with n DOF will have 2n BCs; because the amplitudes are complex they are each able to satisfy two
centre of mass. 9, 11, 12, 13, 14, 15, 21, 22, 23, 52, 54, 55, 57, 58, 61, 89, 98
CONFIGURATION SPACE The variables that describe the configuration of a system define a vector space known as
configuration space.. 32
Euler-Lagrange equation. 109, 110, 111, 113, 114, 115, 118, 120, 123, 124, 125, 127, 132, 133, 134, 136, 151
EOM equation of motion. 1, 4, 6, 8, 9, 15, 19, 25, 26, 27, 32, 35, 37, 112, 114, 115, 117, 119, 120, 123, 125, 127,
128, 130, 131, 133, 134, 136, 150
DR
EQUILIBRIUM COORDINATE A set of generalized coordinates which are displacements measured from the equilib-
rium positions of the system.. 25, 32, 36, 38
IAOR instantaneous axis of rotation. iv, 69, 70, 71, 134, 186, 185
IRF inertial reference frame. 63, 64, 65, 87, 89, 94, 123, 124, 132, 179, 180
N2L Newton’s second law. 1, 2, 3, 10, 14, 15, 18, 22, 23, 32, 64, 67, 68, 125, 134
NIRF non-inertial reference frame. 63, 64, 65, 67, 68, 87, 94, 123, 124, 125, 132, 179, 180
221
222 Glossary
NORMAL MODE A solution for an oscillatory system in which a single frequency oscillation is present,
where ωb are the eigen-frequencies and âb the corresponding eigen-vectors that satisfy the small-oscillation
equation,
T
(K − ω 2 M) · â = 0,
in which K is the spring matrix and M is the mass matrix.. 32, 33, 34, 35, 38, 40, 221
PA principle axes. 89
PAT parallel axis theorem. 14, 22, 54, 55
AF
PHASE SPACE
RHR
An n dimensional space (with n equal to the degrees of freedom of the system) in which the axes are
the generalized coordinates and the generalized velocities (or the generalized momenta). A point in phase space,
the state vector, uniquely describes the state of the system.. 25, 27, 32, 222
STATE VECTOR A vector that describes the state of a system as a point in phase space.. 25, 27, 32, 35, 222
DR
Index
T
action, 146 parallel axis theorem, 54
amplitude vector, 32 phase portrait, 155
ansatz, 25 phase space, 155
precession, 89
cancelling the dots, 114 products of inertia, 51, 52
canonical equations, 152
AF
Cauchy stress tensor, 49
centrifugal force, 18
centripetal acceleration, 10
characteristic frequencies, 32
constraint, 21
coordinate systems transformations, 175
cross-product matrix, 180
secular equation, 32
semi-holonomic constraint, 113
similarity transformation, 47
similarity transformations, 194
special relativity, 42
spring matrix, 32
state vector, 26
strain, 49
stress, 49
summation index, 165
torque, 10
Euler vector, 65, 185
vector calculus, 189
free index, 166 vector multiplication, 169
DR
vectors, 159
Hamilton’s equations, 152
Hessian matrix, 39
Lagrangian, 146
Legendre Transformations, 207
Levi-Civita symbol, 170, 171
Levi-Civita tensor, 180
metric, 46
moment of inertia, 14
223