Вы находитесь на странице: 1из 282

Fluid Mechanics

Richard Fitzpatrick
Professor of Physics
The University of Texas at Austin

Contents
1

Overview
1.1
Intended Audience . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2
Major Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3
To Do List . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Mathematical Models of Fluid Motion


2.1
Introduction . . . . . . . . . . . . . . . . . . .
2.2
What is a Fluid? . . . . . . . . . . . . . . . . .
2.3
Volume and Surface Forces . . . . . . . . . . .
2.4
General Properties of Stress Tensor . . . . . . .
2.5
Stress Tensor in a Static Fluid . . . . . . . . . .
2.6
Stress Tensor in a Moving Fluid . . . . . . . . .
2.7
Viscosity . . . . . . . . . . . . . . . . . . . . .
2.8
Conservation Laws . . . . . . . . . . . . . . .
2.9
Mass Conservation . . . . . . . . . . . . . . .
2.10 Convective Time Derivative . . . . . . . . . . .
2.11 Momentum Conservation . . . . . . . . . . . .
2.12 Navier-Stokes Equation . . . . . . . . . . . . .
2.13 Energy Conservation . . . . . . . . . . . . . .
2.14 Equations of Incompressible Fluid Flow . . . .
2.15 Equations of Compressible Fluid Flow . . . . .
2.16 Dimensionless Numbers in Incompressible Flow
2.17 Dimensionless Numbers in Compressible Flow
2.18 Fluid Equations in Cartesian Coordinates . . . .
2.19 Fluid Equations in Cylindrical Coordinates . . .
2.20 Fluid Equations in Spherical Coordinates . . . .
2.21 Exercises . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

9
9
9
10
11
12
13
14
15
15
16
16
18
18
20
21
22
23
25
26
27
28

Hydrostatics
3.1
Introduction . . . . . . . . . . . . .
3.2
Hydrostatic Pressure . . . . . . . . .
3.3
Buoyancy . . . . . . . . . . . . . .
3.4
Equilibrium of Floating Bodies . . .
3.5
Vertical Stability of Floating Bodies
3.6
Angular Stability of Floating Bodies
3.7
Determination of Metacentric Height
3.8
Energy of a Floating Body . . . . .
3.9
Curve of Buoyancy . . . . . . . . .
3.10 Rotational Hydrostatics . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

31
31
31
31
32
33
34
35
38
38
42

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

7
7
7
7

FLUID MECHANICS
3.11
3.12
3.13
3.14
3.15

Equilibrium of a Rotating Liquid Body


Maclaurin Spheroids . . . . . . . . .
Jacobi Ellipsoids . . . . . . . . . . . .
Roche Ellipsoids . . . . . . . . . . . .
Exercises . . . . . . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

44
46
49
51
57

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

61
61
61
63
63
64
65
66
70
75

Incompressible Inviscid Fluid Dynamics


5.1
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2
Streamlines, Stream Tubes, and Stream Filaments . . . . . . .
5.3
Bernoullis Theorem . . . . . . . . . . . . . . . . . . . . . . .
5.4
Vortex Lines, Vortex Tubes, and Vortex Filaments . . . . . . .
5.5
Circulation and Vorticity . . . . . . . . . . . . . . . . . . . .
5.6
Kelvin Circulation Theorem . . . . . . . . . . . . . . . . . . .
5.7
Irrotational Flow . . . . . . . . . . . . . . . . . . . . . . . . .
5.8
Two-Dimensional Flow . . . . . . . . . . . . . . . . . . . . .
5.9
Two-Dimensional Uniform Flow . . . . . . . . . . . . . . . .
5.10 Two-Dimensional Sources and Sinks . . . . . . . . . . . . . .
5.11 Two-Dimensional Vortex Filaments . . . . . . . . . . . . . . .
5.12 Two-Dimensional Irrotational Flow in Cylindrical Coordinates
5.13 Inviscid Flow Past a Cylindrical Obstacle . . . . . . . . . . . .
5.14 Inviscid Flow Past a Semi-Infinite Wedge . . . . . . . . . . . .
5.15 Inviscid Flow Over a Semi-Infinite Wedge . . . . . . . . . . .
5.16 Velocity Potentials and Stream Functions . . . . . . . . . . . .
5.17 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

77
77
77
77
79
80
80
81
83
85
86
87
90
91
94
95
97
98

2D Potential Flow
6.1
Introduction . . . . . . . .
6.2
Complex Functions . . . .
6.3
Cauchy-Riemann Relations
6.4
Complex Velocity Potential
6.5
Complex Velocity . . . . .
6.6
Method of Images . . . . .
6.7
Conformal Maps . . . . . .
6.8
Complex Line Integrals . .
6.9
Theorem of Blasius . . . .
6.10 Exercises . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

101
101
101
102
102
103
104
109
113
114
118

Surface Tension
4.1
Introduction . . . . . . . . .
4.2
Young-Laplace Equation . .
4.3
Spherical Interfaces . . . . .
4.4
Capillary Length . . . . . . .
4.5
Angle of Contact . . . . . .
4.6
Jurins Law . . . . . . . . .
4.7
Capillary Curves . . . . . . .
4.8
Axisymmetric Soap-Bubbles
4.9
Exercises . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

Incompressible Boundary Layers


121
7.1
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
7.2
No Slip Condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
7.3
Boundary Layer Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

CONTENTS
7.4
7.5
7.6
7.7
7.8
7.9
7.10
7.11
8

Self-Similar Boundary Layers . . . . . . . . . . . . .


Boundary Layer on a Flat Plate . . . . . . . . . . . .
Wake Downstream of a Flat Plate . . . . . . . . . . .
Von Karman Momentum Integral . . . . . . . . . . .
Boundary Layer Separation . . . . . . . . . . . . . .
Criterion for Boundary Layer Separation . . . . . . .
Approximate Solutions of Boundary Layer Equations
Exercises . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

125
128
132
136
137
140
142
147

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

149
149
149
151
153
158
159
159
162
165
167
168

Incompressible Viscous Flow


9.1
Introduction . . . . . . . . . . . . . . . . . . . . . . . .
9.2
Flow Between Parallel Plates . . . . . . . . . . . . . . .
9.3
Flow Down an Inclined Plane . . . . . . . . . . . . . . .
9.4
Poiseuille Flow . . . . . . . . . . . . . . . . . . . . . .
9.5
Taylor-Couette Flow . . . . . . . . . . . . . . . . . . . .
9.6
Flow in Slowly-Varying Channels . . . . . . . . . . . . .
9.7
Lubrication Theory . . . . . . . . . . . . . . . . . . . .
9.8
Stokes Flow . . . . . . . . . . . . . . . . . . . . . . . .
9.9
Axisymmetric Stokes Flow . . . . . . . . . . . . . . . .
9.10 Axisymmetric Stokes Flow Around a Solid Sphere . . . .
9.11 Axisymmetric Stokes Flow In and Around a Fluid Sphere
9.12 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

171
171
171
172
174
174
175
177
179
180
181
185
188

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

191
191
191
193
194
195
196
198
202
203
205
205
206
207
208
209

Incompressible Aerodynamics
8.1
Introduction . . . . . . . . . . . .
8.2
Theorem of Kutta and Zhukovskii
8.3
Cylindrical Airfoils . . . . . . . .
8.4
Zhukovskiis Hypothesis . . . . .
8.5
Vortex Sheets . . . . . . . . . . .
8.6
Induced Flow . . . . . . . . . . .
8.7
Three-Dimensional Airfoils . . . .
8.8
Aerodynamic Forces . . . . . . . .
8.9
Ellipsoidal Airfoils . . . . . . . .
8.10 Simple Flight Problems . . . . . .
8.11 Exercises . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

10 Waves in Incompressible Fluids


10.1 Introduction . . . . . . . . . . . . . .
10.2 Gravity Waves . . . . . . . . . . . . .
10.3 Gravity Waves in Deep Water . . . . .
10.4 Gravity Waves in Shallow Water . . .
10.5 Energy of Gravity Waves . . . . . . .
10.6 Wave Drag on Ships . . . . . . . . . .
10.7 Ship Wakes . . . . . . . . . . . . . .
10.8 Gravity Waves in a Flowing Fluid . . .
10.9 Gravity Waves at an Interface . . . . .
10.10 Steady Flow over a Corrugated Bottom
10.11 Surface Tension . . . . . . . . . . . .
10.12 Capillary Waves . . . . . . . . . . . .
10.13 Capillary Waves at an Interface . . . .
10.14 Wind Driven Waves in Deep Water . .
10.15 Exercises . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

FLUID MECHANICS

11 Equilibrium of Compressible Fluids


11.1 Introduction . . . . . . . . . .
11.2 Isothermal Atmosphere . . . .
11.3 Adiabatic Atmosphere . . . . .
11.4 Atmospheric Stability . . . . .
11.5 Eddington Solar Model . . . .
11.6 Exercises . . . . . . . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

211
211
211
212
213
213
219

A Vectors and Vector Fields


A.1 Introduction . . . . . . . . . . . .
A.2 Scalars and Vectors . . . . . . . .
A.3 Vector Algebra . . . . . . . . . . .
A.4 Cartesian Components of a Vector
A.5 Coordinate Transformations . . . .
A.6 Scalar Product . . . . . . . . . . .
A.7 Vector Area . . . . . . . . . . . .
A.8 Vector Product . . . . . . . . . . .
A.9 Rotation . . . . . . . . . . . . . .
A.10 Scalar Triple Product . . . . . . .
A.11 Vector Triple Product . . . . . . .
A.12 Vector Calculus . . . . . . . . . .
A.13 Line Integrals . . . . . . . . . . .
A.14 Vector Line Integrals . . . . . . .
A.15 Surface Integrals . . . . . . . . . .
A.16 Vector Surface Integrals . . . . . .
A.17 Volume Integrals . . . . . . . . . .
A.18 Gradient . . . . . . . . . . . . . .
A.19 Grad Operator . . . . . . . . . . .
A.20 Divergence . . . . . . . . . . . . .
A.21 Laplacian Operator . . . . . . . .
A.22 Curl . . . . . . . . . . . . . . . .
A.23 Useful Vector Identities . . . . . .
A.24 Exercises . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

223
223
223
223
225
226
227
228
229
231
233
234
234
235
237
237
239
239
240
243
243
246
247
250
250

B Cartesian Tensors
B.1 Introduction . . . . . . . . .
B.2 Tensors and Tensor Notation
B.3 Tensor Transformation . . .
B.4 Tensor Fields . . . . . . . .
B.5 Isotropic Tensors . . . . . .
B.6 Exercises . . . . . . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

253
253
253
255
257
259
261

C Non-Cartesian Coordinates
C.1 Introduction . . . . . . . . . . . . .
C.2 Orthogonal Curvilinear Coordinates
C.3 Cylindrical Coordinates . . . . . . .
C.4 Spherical Coordinates . . . . . . . .
C.5 Exercises . . . . . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

265
265
265
268
270
272

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

D Calculus of Variations
273
D.1 Euler-Lagrange Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
D.2 Conditional Variation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275

CONTENTS
D.3
D.4

Multi-Function Variation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276


Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277

E Ellipsoidal Potential Theory

279

FLUID MECHANICS

Overview

1 Overview

1.1 Intended Audience


This book presents a single semester course on fluid mechanics that is intended primarily for advanced undergraduate
students majoring in physics. A thorough understanding of physics at the lower-division level, including a basic
working knowledge of the laws of mechanics, is assumed. It is also taken for granted that students are familiar
with the fundamentals of multi-variate integral and differential calculus, complex analysis, and ordinary differential
equations. On the other hand, vector analysis plays such a central role in the study of fluid mechanics that a brief,
but fairly comprehensive, review of this subject area is provided in Appendix A. Likewise, those aspects of cartesian
tensor theory, orthogonal curvilinear coordinate systems, and the calculus of variations, that are required in the study
of fluid mechanics are outlined in Appendices B, C, and D, respectively.

1.2 Major Sources


The material appearing in Appendix A is largely based on the authors recollections of a vector analysis course given
by Dr. Stephen Gull at the University of Cambridge. Major sources for the material appearing in other chapters and
appendices include:
Statics, Including Hydrostatics and the Elements of the Theory of Elasticity H. Lamb, 3rd Edition (Cambridge University Press, Cambridge UK, 1928).
Hydrodynamics H. Lamb, 6th Edition (Dover, New York NY, 1945).
Theoretical Aerodynamics L.M. Milne-Thomson, 4th Edition, Revised and enlarged (Dover, New York NY, 1958).
Ellipsoidal Figures of Equilibrium S. Chandrasekhar (Yale University Press, New Haven CT, 1969).
Boundary Layer Theory H. Schlichting, 7th Edition (McGraw-Hill, New York NY, 1970).
Mathematical Methods for the Physical Sciences K.F. Riley (Cambridge University Press, Cambridge UK, 1974).
Fluid Mechanics L.D. Landau, and E.M. Lifshitz, 2nd Edition (Butterworth-Heinemann, Oxford UK, 1987).
Physical Fluid Dynamics D.J. Tritton, 2nd Edition (Oxford University Press, Oxford UK, 1988).
Fluid Dynamics for Physicists T.E. Faber, 1st Edition (Cambridge University Press, Cambridge UK, 1995).
Schaums Outline of Fluid Dynamics W. Hughes, and J. Brighton, 3rd Edition (McGraw-Hill, New York NY, 1999).
An Introduction to Fluid Dynamics G.K. Batchelor (Cambridge University Press, Cambridge UK, 2000).
Theoretical Hydrodynamics L.M. Milne-Thomson, 5th Edition (Dover, New York NY, 2011).

1.3 To Do List
1. Add chapter on vortex dynamics.
2. Add chapter on 3D potential flow.
3. Add appendix on group velocity and Fourier transforms.
4. Add chapter on incompressible flow in rotating systems.
5. Add chapter on instabilities.

FLUID MECHANICS
6. Add chapter on turbulence.
7. Add chapter on 1D compressible flow.
8. Add chapter on sound waves.
9. Add chapter on compressible boundary layers.
10. Add chapter on supersonic aerodynamics.
11. Add chapter on convection.

Mathematical Models of Fluid Motion

2 Mathematical Models of Fluid Motion

2.1 Introduction
In this chapter, we set forth the mathematical models commonly used to describe the equilibrium and dynamics of
fluids. Unless stated otherwise, all of the analysis is performed using a standard right-handed Cartesian coordinate
system: x1 , x2 , x3 . Moreover, the Einstein summation convention is employed (so repeated roman subscripts are
assumed to be summed from 1 to 3see Appendix B).

2.2 What is a Fluid?


By definition, a solid material is rigid. Now, although a rigid material tends to shatter when subjected to very large
stresses, it can withstand a moderate shear stress (i.e., a stress that tends to deform the material by changing its shape,
without necessarily changing its volume) for an indefinite period. To be more exact, when a shear stress is first applied
to a rigid material it deforms slightly, but then springs back to its original shape when the stress is relieved.
A plastic material, such as clay, also possess some degree of rigidity. However, the critical shear stress at which it
yields is relatively small, and once this stress is exceeded the material deforms continuously and irreversibly, and does
not recover its original shape when the stress is relieved.
By definition, a fluid material possesses no rigidity at all. In other words, a small fluid element is unable to
withstand any tendency of an applied shear stress to change its shape. Incidentally, this does not preclude the possibility
that such an element may offer resistance to shear stress. However, any resistance must be incapable of preventing
the change in shape from eventually occurring, which implies that the force of resistance vanishes with the rate of
deformation. An obvious corollary is that the shear stress must be zero everywhere inside a fluid that is in mechanical
equilibrium.
Fluids are conventionally classified as either liquids or gases. The most important difference between these two
types of fluid lies in their relative compressibility: i.e., gases can be compressed much more easily than liquids. Consequently, any motion that involves significant pressure variations is generally accompanied by much larger changes
in mass density in the case of a gas than in the case of a liquid.
Of course, a macroscopic fluid ultimately consists of a huge number of individual molecules. However, most
practical applications of fluid mechanics are concerned with behavior on length-scales that are far larger than the
typical intermolecular spacing. Under these circumstances, it is reasonable to suppose that the bulk properties of a
given fluid are the same as if it were completely continuous in structure. A corollary of this assumption is that when,
in the following, we talk about infinitesimal volume elements, we really mean elements which are sufficiently small
that the bulk fluid properties (such as mass density, pressure, and velocity) are approximately constant across them,
but are still sufficiently large that they contain a very great number of molecules (which implies that we can safely
neglect any statistical variations in the bulk properties). The continuum hypothesis also requires infinitesimal volume
elements to be much larger than the molecular mean-free-path between collisions.
In addition to the continuum hypothesis, our study of fluid mechanics is premised on three major assumptions:
1. Fluids are isotropic media: i.e., there is no preferred direction in a fluid.
2. Fluids are Newtonian: i.e., there is a linear relationship between the local shear stress and the local rate of strain,
as first postulated by Newton. It is also assumed that there is a linear relationship between the local heat flux
density and the local temperature gradient.
3. Fluids are classical: i.e., the macroscopic motion of ordinary fluids is well-described by Newtonian dynamics,
and both quantum and relativistic effects can be safely ignored.
It should be noted that the above assumptions are not valid for all fluid types (e.g., certain liquid polymers, which
are non-isotropic; thixotropic fluids, such as jelly or paint, which are non-Newtonian; and quantum fluids, such as
liquid helium, which exhibit non-classical effects on macroscopic length-scales). However, most practical applications

10

FLUID MECHANICS

of fluid mechanics involve the equilibrium and motion of bodies of water or air, extending over macroscopic lengthscales, and situated relatively close to the Earths surface. Such bodies are very well-described as isotropic, Newtonian,
classical fluids.

2.3 Volume and Surface Forces


Generally speaking, fluids are acted upon by two distinct types of force. The first type is long-range in nature
i.e., such that it decreases relatively slowly with increasing distance between interacting elementsand is capable
of completely penetrating into the interior of a fluid. Gravity is an obvious example of a long-range force. One
consequence of the relatively slow variation of long-range forces with position is that they act equally on all of the
fluid contained within a sufficiently small volume element. In this situation, the net force acting on the element
becomes directly proportional to its volume. For this reason, long-range forces are often called volume forces. In the
following, we shall write the total volume force acting at time t on the fluid contained within a small volume element
of magnitude dV, centered on a fixed point whose position vector is r, as
F(r, t) dV.

(2.1)

The second type of force is short-range in nature, and is most conveniently modeled as momentum transport within
the fluid. Such transport is generally due to a combination of the mutual forces exerted by contiguous molecules, and
momentum fluxes caused by relative molecular motion. Suppose that x (r, t) is the net flux density of x-directed fluid
momentum due to short-range forces at position r and time t. In other words, suppose that, at position r and time t, as a
direct consequence of short-range forces, x-momentum is flowing at the rate of | x | newton-seconds per meter squared
per second in the direction of vector x . Consider an infinitesimal plane surface element, dS = n dS , located at point
r. Here, dS is the area of the element, and n its unit normal. (See Section A.7.) The fluid which lies on that side of the
element toward which n points is said to lie on its positive side, and vice versa. The net flux of x-momentum across the
element (in the direction of n) is x dS newtons, which implies (from Newtons second law of motion) that the fluid
on the positive side of the surface element experiences a force x dS in the x-direction due to short-range interaction
with the fluid on the negative side. According to Newtons third law of motion, the fluid on the negative side of the
surface experiences a force x dS in the x-direction due to interaction with the fluid on the positive side. Short-range
forces are often called surface forces because they are directly proportional to the area of the surface element across
which they act. Let y (r, t) and z (r, t) be the net flux density of y- and z- momentum, respectively, at position r and
time t. By a straightforward extension of above argument, the net surface force exerted by the fluid on the positive side
of some planar surface element, dS, on the fluid on its negative side is
f = (x dS, y dS, z dS).

(2.2)

In tensor notation (see Appendix B), the above equation can be written
fi = i j dS j ,

(2.3)

where 11 = ( x ) x , 12 = ( x )y , 21 = (y ) x , etc. (Note that, since the subscript j is repeated, it is assumed to be


P
summed from 1 to 3. Hence, i j dS j is shorthand for j=1,3 i j dS j .) Here, the i j (r, t) are termed the local stresses
in the fluid at position r and time t, and have units of force per unit area. Moreover, the i j are the components of a
second-order tensor (see Appendix B), known as the stress tensor. [This follows because the fi are the components
of a first-order tensor (since all forces are proper vectors), and the dS i are the components of an arbitrary first-order
tensor (since surface elements are also proper vectorssee Section A.7and (2.3] holds for surface elements whose
normals point in any direction), so application of the quotient rule (see Section B.3) to Equation (2.3) reveals that the
i j transform under rotation of the coordinate axes as the components of a second-order tensor.] We can interpret
i j (r, t) as the i-component of the force per unit area exerted, at position r and time t, across a plane surface element
normal to the j-direction. The three diagonal components of i j are termed normal stresses, since each of them gives
the normal component of the force per unit area acting across a plane surface element parallel to one of the Cartesian
coordinate planes. The six non-diagonal components are termed shear stresses, since they drive shearing motion in
which parallel layers of fluid slide relative to one another.

Mathematical Models of Fluid Motion

11

2.4 General Properties of Stress Tensor


The i-component of the total force acting on a fluid element consisting of a fixed volume V enclosed by a surface S is
written
I
Z
i j dS j ,
(2.4)
Fi dV +
fi =
S

where the first term on the right-hand side is the integrated volume force acting throughout V, whereas the second
term is the net surface force acting across S . Making use of the tensor divergence theorem (see Section B.4), the above
expression becomes
Z
Z
i j
Fi dV +
fi =
dV.
(2.5)
V x j
V
In the limit V 0, it is reasonable to suppose that the Fi and i j /x j are approximately constant across the element.
In this situation, both contributions on the right-hand side of the above equation scale as V. Now, according to
Newtonian dynamics, the i-component of the net force acting on the element is equal to the i-component of the rate
of change of its linear momentum. However, in the limit V 0, the linear acceleration and mass density of the fluid
are both approximately constant across the element. In this case, the rate of change of the elements linear momentum
also scales as V. In other words, the net volume force, surface force, and rate of change of linear momentum of an
infinitesimal fluid element all scale as the volume of the element, and consequently remain approximately the same
order of magnitude as the volume shrinks to zero. We conclude that the linear equation of motion of an infinitesimal
fluid element places no particular restrictions on the stress tensor.
The i-component of the total torque, taken about the origin O of the coordinate system, acting on a fluid element
that consists of a fixed volume V enclosed by a surface S is written [see Equations (A.46) and (B.6)]
I
Z
i jk x j kl dS l ,
(2.6)
i jk x j Fk dV +
i =
S

where the first and second terms on the right-hand side are due to volume and surface forces, respectively. [Here,
i jk is the third-order permutation tensor. See Equation (B.7).] Making use of the tensor divergence theorem (see
Section B.4), the above expression becomes
Z
Z
(x j kl )
i jk
i jk x j Fk dV +
i =
dV,
(2.7)
xl
V
V
which reduces to
i =

i jk x j Fk dV +

i jk k j dV +

i jk x j

kl
dV,
xl

(2.8)

since xi /x j = i j . [Here, i j is the second-order identity tensor. See Equation (B.9).] Assuming that point O lies
within the fluid element, and taking the limit V 0 in which the Fi , i j , and i j /x j are all approximately constant
across the element, we deduce that the first, second, and third terms on the right-hand side of the above equation scale
as V 4/3 , V, and V 4/3 , respectively (since x V 1/3 ). Now, according to Newtonian dynamics, the i-component of the
total torque acting on the fluid element is equal to the i-component of the rate of change of its net angular momentum
about O. Assuming that the linear acceleration of the fluid is approximately constant across the element, we deduce
that the rate of change of its angular momentum scales as V 4/3 (since the net linear acceleration scales as V, so the
net rate of change of angular momentum scales as x V, and x V 1/3 ). Hence, it is clear that the rotational equation of
motion of a fluid element, surrounding a general point O, becomes completely dominated by the second term on the
right-hand side of (2.8) in the limit that the volume of the element approaches zero (since this term is a factor V 1/3
larger than the other terms). It follows that the second term must be identically zero (otherwise an infinitesimal fluid
element would acquire an absurdly large angular velocity). This is only possible, for all choices of the position of
point O, and the shape of the element, if
i jk k j = 0
(2.9)
throughout the fluid. The above relation shows that the stress tensor must be symmetric: i.e.,
ji = i j .

(2.10)

12

FLUID MECHANICS

It immediately follows that the stress tensor only has six independent components (i.e., 11 , 22 , 33 , 12 , 13 , and
23 ).
Now, it is always possible to choose the orientation of a set of Cartesian axes in such a manner that the nondiagonal components of a given symmetric second-order tensor field are all set to zero at a given point in space. (See
Exercise B.6.) With reference to such principal axes, the diagonal components of the stress tensor i j become socalled principal stresses11 , 22 , 33 , say. Of course, in general, the orientation of the principal axes varies with
position. The normal stress 11 acting across a surface element perpendicular to the first principal axis corresponds
to a tension (or a compression if 11 is negative) in the direction of that axis. Likewise, for 22 and 33 . Thus, the
general state of the fluid, at a particular point in space, can be regarded as a superposition of tensions, or compressions,
in three orthogonal directions.
The trace of the stress tensor, ii = 11 + 22 + 33 , is a scalar, and, therefore, independent of the orientation of
the coordinate axes. (See Appendix B.) Thus, it follows that, irrespective of the orientation of the principal axes, the
trace of the stress tensor at a given point is always equal to the sum of the principal stresses: i.e.,
ii = 11 + 22 + 33 .

(2.11)

2.5 Stress Tensor in a Static Fluid


Consider the surface forces exerted on some infinitesimal cubic volume element of a static fluid. Suppose that the
components of the stress tensor are approximately constant across the element. Suppose, further, that the sides of the
cube are aligned parallel to the principal axes of the local stress tensor. This tensor, which now has zero non-diagonal
components, can be regarded as the sum of two tensors: i.e.,

and

1
3 ii


11 31 ii

0
0

0
1
3 ii
0

0
0
1

3 ii

22 13 ii
0

0
0
33 31 ii

(2.12)

(2.13)

The first of the above tensors is isotropic (see Section B.5), and corresponds to the same normal force per unit
area acting inward (since the sign of ii /3 is invariably negative) on each face of the volume element. This uniform
compression acts to change the elements volume, but not its shape, and can easily be withstood by the fluid within the
element.
The second of the above tensors represents the departure of the stress tensor from an isotropic form. The diagonal
components of this tensor have zero sum, in view of (2.11), and thus represent equal and opposite forces per unit
area, acting on opposing faces of the volume element, which are such that the forces on at least one pair of opposing
faces constitute a tension, and the forces on at least one pair constitute a compression. Such forces necessarily tend to
change the shape of the volume element, either elongating or compressing it along one of its symmetry axes. Moreover,
this tendency cannot be offset by any volume force acting on the element, since such forces become arbitrarily small
compared to surface forces in the limit that the elements volume tends to zero (because the ratio of the net volume
force to the net surface force scales as the volume to the surface area of the element, which tends to zero in the limit
that the volume tends to zerosee Section 2.4). Now, we have previously defined a fluid as a material that is incapable
of withstanding any tendency of applied forces to change its shape. (See Section 2.2.) It follows that if the diagonal
components of the tensor (2.13) are non-zero anywhere inside the fluid then it is impossible for the fluid at that point to
be at rest. Hence, we conclude that the principal stresses, 11 , 22 , and 33 , must be equal to one another at all points
in a static fluid. This implies that the stress tensor takes the isotropic form (2.12) everywhere in a stationary fluid.
Furthermore, this is true irrespective of the orientation of the coordinate axes, since the components of an isotropic
tensor are rotationally invariant. (See Section B.5.)

Mathematical Models of Fluid Motion

13

Fluids at rest are generally in a state of compression, so it is convenient to write the stress tensor of a static fluid in
the form
i j = p i j ,
(2.14)
where p = ii /3 is termed the static fluid pressure, and is generally a function of r and t. It follows that, in a stationary
fluid, the force per unit area exerted across a plane surface element with unit normal n is p n. [See Equation (2.3).]
Moreover, this normal force has the same value for all possible orientations of n. This well-known resultnamely,
that the pressure is the same in all directions at a given point in a static fluidis known as Pascals law, and is a direct
consequence of the fact that a fluid element cannot withstand shear stresses, or, alternatively, any tendency of applied
forces to change its shape.

2.6 Stress Tensor in a Moving Fluid


We have seen that in a static fluid the stress tensor takes the form
i j = p i j ,

(2.15)

where p = ii /3 is the static pressure: i.e., minus the normal stress acting in any direction. Now, the normal stress at
a given point in a moving fluid generally varies with direction: i.e., the principal stresses are not equal to one another.
However, we can still define the mean principal stress as (11 +22 +33 )/3 = ii /3. Moreover, given that the principal
stresses are actually normal stresses (in a coordinate frame aligned with the principal axes), we can also regard ii /3
as the mean normal stress. It is convenient to define pressure in a moving fluid as minus the mean normal stress: i.e.,
1
p = ii .
3

(2.16)

Thus, we can write the stress tensor in a moving fluid as the sum of an isotropic part, p i j , which has the same form
as the stress tensor in a static fluid, and a remaining non-isotropic part, di j , which includes any shear stresses, and also
has diagonal components whose sum is zero. In other words,
i j = p i j + di j ,

(2.17)

dii = 0.

(2.18)

where
Moreover, since i j and i j are both symmetric tensors, it follows that di j is also symmetric: i.e.,
d ji = di j .

(2.19)

It is clear that the so-called deviatoric stress tensor, di j , is a consequence of fluid motion, since it is zero in a static
fluid. Suppose, however, that we were to view a static fluid both in its rest frame and in a frame of reference moving
at some constant velocity relative to the rest frame. Now, we would expect the force distribution within the fluid to
be the same in both frames of reference, since the fluid does not accelerate in either. However, in the first frame,
the fluid appears stationary and the deviatoric stress tensor is therefore zero, whilst in the second it has a spatially
uniform velocity field and the deviatoric stress tensor is also zero (because it is the same as in the rest frame). We,
thus, conclude that the deviatoric stress tensor is zero both in a stationary fluid and in a moving fluid possessing no
spatial velocity gradients. This suggests that the deviatoric stress tensor is driven by velocity gradients within the fluid.
Moreover, the tensor must vanish as these gradients vanish.
Let the vi (r, t) be the Cartesian components of the fluid velocity at point r and time t. The various velocity
gradients within the fluid then take the form vi /x j . The simplest possible assumption, which is consistent with the
above discussion, is that the components of the deviatoric stress tensor are linear functions of these velocity gradients:
i.e.,
vk
.
(2.20)
di j = Ai jkl
xl

14

FLUID MECHANICS

Here, Ai jkl is a fourth-order tensor (this follows from the quotient rule because di j and vi /x j are both proper secondorder tensors). Any fluid in which the deviatoric stress tensor takes the above form is termed a Newtonian fluid, since
Newton was the first to postulate a linear relationship between shear stresses and velocity gradients.
Now, in an isotropic fluidthat is, a fluid in which there is no preferred directionwe would expect the fourthorder tensor Ai jkl to be isotropicthat is, to have a form in which all physical distinction between different directions
is absent. As demonstrated in Section B.5, the most general expression for an isotropic fourth-order tensor is
Ai jkl = i j kl + ik jl + il jk ,

(2.21)

where , , and are arbitrary scalars (which can be functions of position and time). Thus, it follows from (2.20) and
(2.21) that
v j
vi
vk
i j +
+
.
(2.22)
di j =
xk
x j
xi
However, according to Equation (2.19), di j is a symmetric tensor, which implies that = , and
di j = ekk i j + 2 ei j ,
where
1 vi v j
ei j =
+
2 x j xi

(2.23)

(2.24)

is called the rate of strain tensor. Finally, according to Equation (2.18), di j is a traceless tensor, which yields 3 =
2 , and
!
1
(2.25)
di j = 2 ei j ekk i j ,
3
where = . We, thus, conclude that the most general expression for the stress tensor in an isotropic Newtonian fluid
is
!
1
i j = p i j + 2 ei j ekk i j ,
(2.26)
3
where p(r, t) and (r, t) are arbitrary scalars.

2.7 Viscosity
The significance of the parameter , appearing in the previous expression for the stress tensor, can be seen from the
form taken by the relation (2.25) in the special case of simple shearing motion. With v1 /x2 as the only non-zero
velocity derivative, all of the components of di j are zero apart from the shear stresses
d12 = d21 =

v1
.
x2

(2.27)

Thus, is the constant of proportionality between the rate of shear and the tangential force per unit area when parallel
plane layers of fluid slide over one another. This constant of proportionality is generally referred to as viscosity. It is a
matter of experience that the force between layers of fluid undergoing relative sliding motion always tends to oppose
the motion, which implies that > 0.
The viscosities of dry air and pure water at 20 C and atmospheric pressure are about 1.8 105 kg/(m s) and
1.0 103 kg/(m s), respectively. In neither case does the viscosity exhibit much variation with pressure. However, the
viscosity of air increases by about 0.3 percent, and that of water decreases by about 3 percent, per degree Centigrade
rise in temperature.

Mathematical Models of Fluid Motion

15

2.8 Conservation Laws


Suppose that (r, t) is the density of some bulk fluid property (e.g., mass, momentum, energy) at position r and time t.
In other words, suppose that, at time t, an infinitesimal fluid element of volume dV, located at position r, contains an
amount (r, t) dV of the property in question. Note, incidentally, that can be either a scalar, a component of a vector,
or even a component of a tensor. The total amount of the property contained within some fixed volume V is
Z
dV,
(2.28)
=
V

where the integral is taken over all elements of V. Let dS be an outward directed element of the bounding surface of
V. Suppose that this element is located at point r. The volume of fluid that flows per second across the element, and so
out of V, is v(r, t)dS. Thus, the amount of the fluid property under consideration that is convected across the element
per second is (r, t) v(r, t) dS. It follows that the net amount of the property that is convected out of volume V by fluid
flow across its bounding surface S is
Z
v dS,

(2.29)

where the integral is taken over all outward directed elements of S . Suppose, finally, that the property in question is
created within the volume V at the rate S per second. The conservation equation for the fluid property takes the form
d
= S .
dt

(2.30)

In other words, the rate of increase in the amount of the property contained within V is the difference between the
creation rate of the property inside V, and the rate at which the property is convected out of V by fluid flow. The above
conservation law can also be written
d
+ = S .
(2.31)
dt
Here, is termed the flux of the property out of V, whereas S is called the net generation rate of the property within
V.

2.9 Mass Conservation


Let (r, t) and v(r, t) be the mass density and velocity of a given fluid at point r and time t. Consider a fixed volume
V, surrounded by a surface S . The net mass contained within V is
Z
dV,
(2.32)
M=
V

where dV is an element of V. Furthermore, the mass flux across S , and out of V, is [see Equation (2.29)]
Z
M =
v dS,

(2.33)

where dS is an outward directed element of S . Mass conservation requires that the rate of increase of the mass
contained within V, plus the net mass flux out of V, should equal zero: i.e.,
dM
+ M = 0
dt

(2.34)

[cf., Equation (2.31)]. Here, we are assuming that there is no mass generation (or destruction) within V (since individual molecules are effectively indestructible). It follows that
Z
Z

v dS = 0,
(2.35)
dV +
S
V t

16

FLUID MECHANICS

since V is non-time-varying. Making use of the divergence theorem (see Section A.20), the above equation becomes
#
Z "

+ ( v) dV = 0.
(2.36)
V t
However, this result is true irrespective of the size, shape, or location of volume V, which is only possible if

+ ( v) = 0
t

(2.37)

throughout the fluid. The above expression is known as the equation of fluid continuity, and is a direct consequence of
mass conservation.

2.10 Convective Time Derivative


The quantity (r, t)/t, appearing in Equation (2.37), represents the time derivative of the fluid mass density at the
fixed point r. Suppose that v(r, t) is the instantaneous fluid velocity at the same point. It follows that the time derivative
of the density, as seen in a frame of reference which is instantaneously co-moving with the fluid at point r, is
lim

t0

D
(r + v t, t + t) (r, t)
=
+ v =
,
t
t
Dt

(2.38)

where we have Taylor expanded (r + v t, t + t) up to first order in t, and where

D
.
=
+ v =
+ vi
Dt t
t
xi

(2.39)

Clearly, the so-called convective time derivative, D/Dt, represents the time derivative seen in the local rest frame of
the fluid.
The continuity equation (2.37) can be rewritten in the form
1 D D ln
=
= v,
Dt
Dt

(2.40)

since ( v) = v + v [see (A.174)]. Consider a volume element V that is co-moving with the fluid. In general,
as the element is convected by the fluid its volume changes. In fact, it is easily seen that
Z
Z
Z
Z
DV
vi
dV =
=
v dV,
(2.41)
v dS =
vi dS i =
Dt
V
S
S
V xi
where S is the bounding surface of the element, and use has been made of the divergence theorem. In the limit that
V 0, and v is approximately constant across the element, we obtain
1 DV
D ln V
=
= v.
V Dt
Dt

(2.42)

Hence, we conclude that the divergence of the fluid velocity at a given point in space specifies the fractional rate of
increase in the volume of an infinitesimal co-moving fluid element at that point.

2.11 Momentum Conservation


Consider a fixed volume V surrounded by a surface S . The i-component of the total linear momentum contained within
V is
Z
vi dV.

Pi =

(2.43)

Mathematical Models of Fluid Motion

17

Moreover, the flux of i-momentum across S , and out of V, is [see Equation (2.29)]
Z
vi v j dS j .
i =

(2.44)

Finally, the i-component of the net force acting on the fluid within V is
I
Z
i j dS j ,
Fi dV +
fi =
V

(2.45)

where the first and second terms on the right-hand side are the contributions from volume and surface forces, respectively.
Momentum conservation requires that the rate of increase of the net i-momentum of the fluid contained within
V, plus the flux of i-momentum out of V, is equal to the rate of i-momentum generation within V. Of course, from
Newtons second law of motion, the latter quantity is equal to the i-component of the net force acting on the fluid
contained within V. Thus, we obtain [cf., Equation (2.31)]
dPi
+ i = fi ,
dt
which can be written

( vi )
dV +
t

vi v j dS j =

(2.46)

Fi dV +

i j dS j ,

(2.47)

since the volume V is non-time-varying. Making use of the tensor divergence theorem, this becomes
Z "
V

#
!
Z
i j
( vi ) ( vi v j )
Fi +
dV =
dV.
+
t
x j
x j
V

(2.48)

However, the above result is valid irrespective of the size, shape, or location of volume V, which is only possible if
i j
( vi ) ( vi v j )
= Fi +
+
t
x j
x j

(2.49)

everywhere inside the fluid. Expanding the derivatives, and rearranging, we obtain
!
!
i j
v j

vi
vi
vi +
= Fi +
+
.
+ vj
+ vj
t
x j
x j
t
x j
x j

(2.50)

Now, in tensor notation, the continuity equation (2.37) is written


v j

+
= 0.
+ vj
t
x j
x j

(2.51)

So, combining Equations (2.50) and (2.51), we obtain the following fluid equation of motion,
!
i j
vi
vi

= Fi +
.
+ vj
t
x j
x j

(2.52)

Fi 1 i j
Dvi
.
=
+
Dt

x j

(2.53)

An alternative form of this equation is

The above equation describes how the net volume and surface forces per unit mass acting on a co-moving fluid element
determine its acceleration.

18

FLUID MECHANICS

2.12 Navier-Stokes Equation


Equations (2.24), (2.26), and (2.53) can be combined to give the equation of motion of an isotropic, Newtonian,
classical fluid: i.e.,
"
!#
!
2 v j

p
vi v j
Dvi

.
(2.54)
+
+
= Fi

Dt
xi x j
x j xi
xi 3 x j
This equation is generally known as the Navier-Stokes equation. Now, in situations in which there are no strong
temperature gradients in the fluid, it is a good approximation to treat viscosity as a spatially uniform quantity, in which
case the Navier-Stokes equation simplifies somewhat to give

#
" 2
Dvi
1 2 v j
p
vi
.
+
+
= Fi
Dt
xi
x j x j 3 xi x j

(2.55)

When expressed in vector form, the above expression becomes


"
#
"
#
1
v
Dv
=
+ (v) v = F p + 2 v + (v) ,

Dt
t
3

(2.56)

where use has been made of Equation (2.39). Here,


bi
,
x j

[(a)b]i

aj

(2 v)i

2 vi .

(2.57)
(2.58)

Note, however, that the above identities are only valid in Cartesian coordinates. (See Appendix C.)

2.13 Energy Conservation


Consider a fixed volume V surrounded by a surface S . The total energy content of the fluid contained within V is
Z
Z
1
E dV +
E=
vi vi dV,
(2.59)
V 2
V
where the first and second terms on the right-hand side are the net internal and kinetic energies, respectively. Here,
E(r, t) is the internal (i.e., thermal) energy per unit mass of the fluid. The energy flux across S , and out of V, is [cf.,
Equation (2.29)]
!
! #
"
Z
Z

1
1
(2.60)
E + vi vi v j dV,
E + vi vi v j dS j =
E =
2
2
V x j
S
where use has been made of the tensor divergence theorem. According to the first law of thermodynamics, the rate of
increase of the energy contained within V, plus the net energy flux out of V, is equal to the net rate of work done on
the fluid within V, minus the net heat flux out of V: i.e.,

. .
dE
+ E = W Q,
dt
.

(2.61)

where W is the net rate of work, and Q the net heat flux. It can be seen that W Q is the effective energy generation
rate within V [cf., Equation (2.31)].
Now, the net rate at which volume and surface forces do work on the fluid within V is
#
Z "
Z
. Z
(vi i j )
vi F i +
dV,
(2.62)
vi i j dS j =
vi Fi dV +
W=
x j
V
S
V
where use has been made of the tensor divergence theorem.

Mathematical Models of Fluid Motion

19

Generally speaking, heat flow in fluids is driven by temperature gradients. Let the qi (r, t) be the Cartesian components of the heat flux density at position r and time t. It follows that the heat flux across a surface element dS,
located at point r, is q dS = qi dS i . Let T (r, t) be the temperature of the fluid at position r and time t. Thus, a general
temperature gradient takes the form T/xi . Let us assume that there is a linear relationship between the components
of the local heat flux density and the local temperature gradient: i.e.,
q i = Ai j

T
,
x j

(2.63)

where the Ai j are the components of a second-rank tensor (which can be functions of position and time). Now, in an
isotropic fluid we would expect Ai j to be an isotropic tensor. (See Section B.5.) However, the most general secondorder isotropic tensor is simply a multiple of i j . Hence, we can write
Ai j = i j ,

(2.64)

where (r, t) is termed the thermal conductivity of the fluid. It follows that the most general expression for the heat
flux density in an isotropic fluid is
T
qi =
,
(2.65)
xi
or, equivalently,
q = T.

(2.66)

Moreover, it is a matter of experience that heat flows down temperature gradients: i.e., > 0. We conclude that the net
heat flux out of volume V is
!
Z
Z
.

T
T

dV,
(2.67)
dS i =
Q=
xi
xi
V xi
S
where use has been made of the tensor divergence theorem.
Equations (2.59)(2.62) and (2.67) can be combined to give the following energy conservation equation:
!#
! #)
"
Z ( "

1
E + vi vi +
E + vi vi v j dV
2
x j
2
V t
!#
Z "

T
=
vi F i +
vi i j +
dV.
x j
x j
V
However, this result is valid irrespective of the size, shape, or location of volume V, which is only possible if
"
!#
! #
"
!

1
T
E + vi vi +
E + vi vi v j = vi F i +
vi i j +
t
2
x j
2
x j
x j
everywhere inside the fluid. Expanding some of the derivatives, and rearranging, we obtain
!
!
1

T
D
E + vi vi = vi F i +
vi i j +
,

Dt
2
x j
x j

(2.68)

(2.69)

(2.70)

where use has been made of the continuity equation (2.40). Now, the scalar product of v with the fluid equation of
motion (2.53) yields
!
i j
Dvi
D 1
vi
.
(2.71)
=
vi vi = vi F i + vi
Dt
Dt 2
x j
Combining the previous two equations, we get

vi
T
DE
i j +
=

.
Dt
x j
x j x j

(2.72)

20

FLUID MECHANICS

Finally, making use of (2.26), we deduce that the energy conservation equation for an isotropic Newtonian fluid takes
the general form
"
!#

T
p vi 1
DE

.
(2.73)
+ +
=
Dt
xi
x j x j
Here,
=

!
!
vi
vi vi
1
vi v j 2 vi v j
di j = 2 ei j ei j eii e j j =
+

x j
3
x j x j x j xi 3 xi x j

(2.74)

is the rate of heat generation per unit volume due to viscosity. When written in vector form, Equation (2.73) becomes
DE
p
( T )
= v + +
.
Dt

(2.75)

According to the above equation, the internal energy per unit mass of a co-moving fluid element evolves in time as
a consequence of work done on the element by pressure as its volume changes, viscous heat generation due to flow
shear, and heat conduction.

2.14 Equations of Incompressible Fluid Flow


In most situations of general interest, the flow of a conventional liquid, such as water, is incompressible to a high degree
of accuracy. Now, a fluid is said to be incompressible when the mass density of a co-moving volume element does not
change appreciably as the element moves through regions of varying pressure. In other words, for an incompressible
fluid, the rate of change of following the motion is zero: i.e.,
D
= 0.
Dt

(2.76)

v = 0.

(2.77)

In this case, the continuity equation (2.40) reduces to

We conclude that, as a consequence of mass conservation, an incompressible fluid must have a divergence-free, or
solenoidal, velocity field. This immediately implies, from Equation (2.42), that the volume of a co-moving fluid
element is a constant of the motion. In most practical situations, the initial density distribution in an incompressible
fluid is uniform in space. Hence, it follows from (2.76) that the density distribution remains uniform in space and
constant in time. In other words, we can generally treat the density, , as a uniform constant in incompressible fluid
flow problems.
Suppose that the volume force acting on the fluid is conservative in nature (see Section A.18): i.e.,
F = ,

(2.78)

where (r, t) is the potential energy per unit mass, and the potential energy per unit volume. Assuming that
the fluid viscosity is a spatially uniform quantity, which is generally the case (unless there are strong temperature
variations within the fluid), the Navier-Stokes equation for an incompressible fluid reduces to
p
Dv
=
+ 2 v,
Dt

where
=

(2.79)

(2.80)

is termed the kinematic


viscosity, and has units of meters squared per second. Roughly speaking, momentum diffuses
a distance of order t meters in t seconds as a consequence of viscosity. The kinematic viscosity of water at 20 C
is about 1.0 106 m2 /s. It follows that viscous momentum diffusion in water is a relatively slow process.

Mathematical Models of Fluid Motion

21

The complete set of equations governing incompressible flow is


v =
Dv
=
Dt

0,

(2.81)

+ 2 v.

(2.82)

Here, and are regarded as known constants, and (r, t) as a known function. Thus, we have four equations
namely, Equation (2.81), plus the three components of Equation (2.82)for four unknownsnamely, the pressure,
p(r, t), plus the three components of the velocity, v(r, t). Note that an energy conservation equation is redundant in the
case of incompressible fluid flow.

2.15 Equations of Compressible Fluid Flow


In many situations of general interest, the flow of gases is compressible: i.e., there are significant changes in the
mass density as the gas flows from place to place. For the case of compressible flow, the continuity equation (2.40),
and the Navier-Stokes equation (2.56), must be augmented by the energy conservation equation (2.75), as well as
thermodynamic relations that specify the internal energy per unit mass, and the temperature in terms of the density
and pressure. For an ideal gas, these relations take the form
E

cV
T,
M
M p
,
R

(2.83)
(2.84)

where cV is the molar specific heat at constant volume, R = 8.3145 J K1 mol1 the molar ideal gas constant, M the
molar mass (i.e., the mass of 1 mole of gas molecules), and T the temperature in degrees Kelvin. Incidentally, 1 mole
corresponds to 6.0221 1024 molecules. Here, we have assumed, for the sake of simplicity, that cV is a uniform
constant. It is also convenient to assume that the thermal conductivity, , is a uniform constant. Making use of these
approximations, Equations (2.40), (2.75), (2.83), and (2.84) can be combined to give
!
!
M 2 p
Dp p D
1
=+
,
(2.85)

1 Dt
Dt
R

where
=

cp
cV + R
=
cV
cV

(2.86)

is the ratio of the molar specific heat at constant pressure, c p , to that at constant volume, cV . (Incidentally, the result
that c p = cV + R for an ideal gas is a standard theorem of thermodynamics.) The ratio of specific heats of dry air at
20 C is 1.40.
The complete set of equations governing compressible ideal gas flow are
D
=
Dt
Dv
=
Dt
!
Dp p D
1
=

1 Dt
Dt

v,

"
#
1
p
2
v + (v) ,

3
!
M 2 p
+
,

(2.87)
(2.88)
(2.89)

where the dissipation function is specified in terms of and v in Equation (2.74). Here, , , , M, and R are regarded
as known constants, and (r, t) as a known function. Thus, we have five equationsnamely, Equations (2.87) and
(2.89), plus the three components of Equation (2.88)for five unknownsnamely, the density, (r, t), the pressure,
p(r, t), and the three components of the velocity, v(r, t).

22

FLUID MECHANICS

2.16 Dimensionless Numbers in Incompressible Flow


It is helpful to normalize the equations of incompressible fluid flow, (2.81)(2.82), in the following manner: = L ,
v = v/V0 , t = (V0 /L) t, = /(g L), and p = p/( V02 + g L + V0 /L). Here, L is a typical spatial variation lengthscale, V0 a typical fluid velocity, and g a typical gravitational acceleration (assuming that represents a gravitational
potential energy per unit mass). All barred quantities are dimensionless, and are designed to be comparable with unity.
The normalized equations of incompressible fluid flow take the form
v = 0,
Dv
Dt

(2.90)

= 1+

1
1
v
p 2 +
+
,
2
Re
Re
Fr
Fr

(2.91)

where D/Dt = /t + v, and


Re

Fr

L V0
,

V0
.
(g L)1/2

(2.92)
(2.93)

Here, the dimensionless quantities Re and Fr are known as the Reynolds number and the Froude number, respectively.
The Reynolds number is the typical ratio of inertial to viscous forces within the fluid, whereas the square of the Froude
number is the typical ratio of inertial to gravitational forces. Thus, viscosity is relatively important compared to inertia
when Re 1, and vice versa. Likewise, gravity is relatively important compared to inertia when Fr 1, and vice
versa. Note that, in principal, Re and Fr are the only quantities in Equations (2.90) and (2.91) that can be significantly
greater or smaller than unity.
For the case of water at 20 C, located on the surface of the Earth,
Re
Fr

1.0 106 L(m) V0 (m s1 ),

3.2 10

V0 (m s )/[L(m)]

(2.94)
1/2

(2.95)

Thus, if L 1 m and V0 1 m s1 , as is often the case for terrestrial water dynamics, then the above expressions suggest that Re 1 and Fr O(1). In this situation, the viscous term on the right-hand side of (2.91) becomes negligible,
and the (unnormalized) incompressible fluid flow equations reduce to the following inviscid, incompressible, fluid flow
equations,
v =
Dv
=
Dt

0,

(2.96)
p
.

(2.97)

For the case of lubrication oil at 20 C, located on the surface of the Earth, 1.0 104 m2 s1 (i.e., oil is about
100 times more viscous than water), and so
Re

Fr

1.0 104 L(m) V0 (m s1 ),

3.2 101 V0 (m s1 )]/[L(m)]1/2.

(2.98)
(2.99)

Suppose that oil is slowly flowing down a narrow lubrication channel such that L 103 m and V0 101 m s1 . It
follows, from the above expressions, that Re 1 and Fr 1. In this situation, the inertial term on the left-hand
side of (2.91) becomes negligible, and the (unnormalized) incompressible fluid flow equations reduce to the following
inertia-free, incompressible, fluid flow equations,
v =
0

0,

(2.100)

p
+ 2 v.

(2.101)

Mathematical Models of Fluid Motion

23

2.17 Dimensionless Numbers in Compressible Flow


It is helpful to normalize the equations of compressible ideal gas flow, (2.87)(2.89), in the following manner: =
L , v = v/V0 , t = (V0 /L) t, = /0 , = /(g L), = (L/V0 )2 , and p = (p p0 )/(0 V02 + 0 g L + 0 V0 /L).
Here, L is a typical spatial variation length-scale, V0 a typical fluid velocity, 0 a typical mass density, and g a typical
gravitational acceleration (assuming that represents a gravitational potential energy per unit mass). Furthermore, p0
corresponds to atmospheric pressure at ground level, and is a uniform constant. It follows that p represents deviations
from atmospheric pressure. All barred quantities are dimensionless, and are designed to be comparable with unity.
The normalized equations of compressible ideal gas flow take the form
D
Dt

v,

(2.102)

!
"
#
2
1 p
1
1
1
2 +
= 1+ 2 +
v (v) ,
Re
Re
3
Fr
Fr
"
!
!
#
p0 + p D
Dp
1
1

2 p0 + p
+

,

=
1 Dt

Dt
1 + Re (1 + 1/Fr 2 ) Re Pr
Dv
Dt

p0

M2

1
,
(1 + 1/Fr 2 + 1/Re)

(2.103)
(2.104)
(2.105)

where D/Dt /t + v,
Re

Fr

Pr

L V0
,

V0
,
(g L)1/2

,
H
V0
,
p
p0 /0

and
=
H

(2.106)
(2.107)
(2.108)
(2.109)

,
0

(2.110)

M
.
R 0

(2.111)

Here, the dimensionless numbers Re, Fr, Pr, and M are known as the Reynolds number, Froude number, Prandtl
number, and Mach number, respectively. The Reynolds number is the typical ratio of inertial to viscous forces within
the gas, the square of the Froude number the typical ratio of inertial to gravitational forces, the Prandtl number the
typical ratio of the momentum and thermal diffusion rates, and the Mach number the typical ratio of gas flow and
sound propagation speeds. Thus, thermal diffusion is far faster than momentum diffusion when Pr 1, and vice
versa. Moreover, the p
gas flow is termed subsonic when M 1, supersonic when M 1, and transonic when
M O(1). Note that p0 /0 is the speed of sound in the undisturbed gas. The quantity H is called the thermal
diffusivityof the gas, and has units of meters squared per second. Thus, heat typically diffuses through the gas a
distance H t meters in t seconds. The thermal diffusivity of dry air at atmospheric pressure and 20 C is about
H = 2.1 105 m2 s1 . It follows that heat diffusion in air is a relatively slow process. The kinematic viscosity of
dry air at atmospheric pressure and 20 C is about = 1.5 105 m2 s1 . Hence, momentum diffusion in air is also a
relatively slow process.
For the case of dry air at atmospheric pressure and 20 C,
Re

6.7 104 L(m) V0 (m s1 ),

(2.112)

24

FLUID MECHANICS

3.2 101 V0 (m s1 )/[L(m)]1/2,

(2.113)

Pr

7.2 10 ,

(2.114)

Fr

2.9 103 V0 (m s1 ).

(2.115)

Thus, if L 1 m and V0 1 m s , as is often the case for subsonic air dynamics close to the Earths surface, the
above expressions suggest that Re 1, M 1, and Fr, Pr O(1). It immediately follows from Equation (2.105)
that p0 1. However, in this situation, Equation (2.104) is dominated by the second term in square brackets on its
left-hand side. Hence, this equation can only be satisfied if the term in question is small, which implies that
D
1.
Dt

(2.116)

v 1.

(2.117)

Equation (2.102) then gives


Thus, it is evident that subsonic (i.e., M 1) gas flow is essentially incompressible. The fact that Re 1 implies
that such flow is also essentially inviscid. In the incompressible inviscid limit (in which v = 0 and Re 1), the
(unnormalized) compressible ideal gas flow equations reduce to the previously derived inviscid, incompressible, fluid
flow equations: i.e.,
v =
Dv
=
Dt

0,

(2.118)
p
.

(2.119)

It follows that the equations which govern subsonic gas dynamics close to the surface of the Earth are essentially the
same as those which govern the flow of water.
Suppose that L 1 m and V0 300 m s1, as is typically the case for transonic air dynamics (e.g., air flow over the
wing of a fighter jet). In this situation, Equations (2.105) and (2.112)(2.115) yield Re, Fr 1 and M, Pr, p0 O(1).
It follows that the final two terms on the right-hand sides of Equations (2.103) and (2.104) can be neglected. Thus,
the (unnormalized) compressible ideal gas flow equations reduce to the following set of inviscid, adiabatic, ideal gas,
flow equations,
D
Dt
Dv
Dt
!
D p
Dt

v,

(2.120)

p
,

(2.121)

0.

(2.122)

In particular, if the initial distribution of p/ is uniform in space, as is often the case, then Equation (2.122) ensures
that the distribution remains uniform as time progresses. In fact, it can be shown that the entropy per unit mass of an
ideal gas is
!
cV
p
S=
(2.123)
ln .
M

Hence, the assumption that p/ is uniform in space is equivalent to the assumption that the entropy per unit mass of
the gas is a spatial constant. A gas in which this is the case is termed homentropic. Equation (2.122) ensures that the
entropy of a co-moving gas element is a constant of the motion in transonic flow. A gas in which this is the case is
termed isentropic. In the homentropic case, the above compressible gas flow equations simplify somewhat to give
D
Dt
Dv
Dt
p
p0

= v,
p
,

.
0

(2.124)

(2.125)

(2.126)

Mathematical Models of Fluid Motion

25

Here, p0 is atmospheric pressure, and 0 is the density of air at atmospheric pressure. Equation (2.126) is known as the
adiabatic gas law, and is a consequence of the fact that transonic gas dynamics takes place far too quickly for thermal
heat conduction (which is a relatively slow process) to have any appreciable effect on the temperature distribution
within the gas. Incidentally, a gas in which thermal diffusion is negligible is generally termed adiabatic.

2.18 Fluid Equations in Cartesian Coordinates


Let us adopt the conventional Cartesian coordinate system, x, y, z. According to Equation (2.26), the various components of the stress tensor are
v x
,
x
vy
p + 2
,
y

p + 2

xx

yy

zz

p + 2

(2.129)

xy = yx

v x
y

(2.130)

xz = zx

v x
z

yz = zy

vy
z

vz
,
z
!
vy
,
+
x
!
vz
,
+
x
!
vz
,
+
y

(2.127)
(2.128)

(2.131)
(2.132)

where v is the velocity, p the pressure, and the viscosity. The equations of compressible fluid flow, (2.87)(2.89)
(from which the equations of incompressible fluid flow can easily be obtained by setting = 0), become
D
Dt
Dv x
Dt
Dvy
Dt
Dvz
Dt
!
1
D p D

1 Dt
Dt

(2.133)

(2.134)

=
=
=

!
1
1 p 2
vx +
,

+
x
x
3 x
!
1 p 2
1

+
vy +
,
y
y
3 y
!
1
1 p 2
vz +
,

+
z
z

3 z
!
M 2 p
+
,

(2.135)
(2.136)
(2.137)

where is the mass density, the ratio of specific heats, the heat conductivity, M the molar mass, and R the molar
ideal gas constant. Furthermore,

v x vy vz
+
+
,
x
y
z

(2.138)

D
Dt

+ vx
+ vy
+ vz ,
t
x
y
z

(2.139)

2
2
2
+
+
,
x2 y2 z2

!
!
!2
!2
v x 2
vy 2
1 v x vy
vz

2
+
+
+
+
x
y
z
2 y
x

(2.140)

26

FLUID MECHANICS
1 v x vz
+
+
2 z
x

!2

!2
1 vy vz
+
+
.
2 z
y

(2.141)

In the above, , , , and M are treated as uniform constants.

2.19 Fluid Equations in Cylindrical Coordinates


Let us adopt the cylindrical coordinate system, r, , z. Making use of the results quoted in Section C.3, the components
of the stress tensor are
rr

zz

r = r

rz = zr

z = z

vr
,
r
!
1 v vr
p + 2
,
+
r
r
p + 2

vz
,
z
!
1 vr v v
,
+

r
r
r
!
vr vz

,
+
z
r
!
1 vz v
,
+

r
z

p + 2

(2.142)
(2.143)
(2.144)
(2.145)
(2.146)
(2.147)

whereas the equations of compressible fluid flow become


D
Dt
Dvr v2

Dt
r

Dv vr v
+
Dt
r

1 p

r
r

!
vr
2 v 1
2
vr 2 2
,
+

r
r 3 r

1 p 1

r r

!
2
2 vr v
1
v + 2
,
2+

3r
r r
!
1
1 p 2
vz +
,

+
z
z

3 z
!
M 2 p
+
,

Dvz
=
Dt
!
1
D p D
=

1 Dt
Dt

(2.148)

(2.149)

(2.150)
(2.151)
(2.152)

where
=
D
Dt

1 (r vr ) 1 v vz
+
+
,
r r
r
z

v
+ vr
+
+ vz ,
t
r
r
z
!
1

1 2
2
r
+ 2 2 + 2,
r r r
r
z

(2.153)
(2.154)
(2.155)

Mathematical Models of Fluid Motion

!
vr 2
1 v

+
+
= 2
r
r
!2
1 vr vz
1
+
+
+
2 z
r
2

27
!2

vz
+
z

!2

1 1 vr v v
+
+

2 r
r
r

!2
v 1 vz
+
.
z
r

vr
r

!2
(2.156)

2.20 Fluid Equations in Spherical Coordinates


Let us, finally, adopt the spherical coordinate system, r, , . Making use of the results quoted in Section C.4, the
components of the stress tensor are
vr
,
r
!
1 v vr
= p + 2
,
+
r
r
= p + 2

rr

!
1 v vr cot v
,
+ +
r sin
r
r
!
1 vr v v
,
+

=
r
r
r
!
1 vr v v
=
,
+

r sin
r
r
!
1 v 1 v cot v
,
+

=
r sin r
r

= p + 2

r = r
r = r
=

(2.157)
(2.158)
(2.159)
(2.160)
(2.161)
(2.162)

whereas the equations of compressible fluid flow become

Dvr

Dt

D
Dt
2
v + v2
r

2
Dv vr v cot v
+
Dt
r

Dv vr v + cot v v
+
Dt
r

1
D p D

1 Dt
Dt

2vr
1 p 2
2 v
vr 2 2

+
r
r

r
r
!
v 1
2 cot v
2

,
+

r2
r2 sin 3 r
2 vr
1 p 1 2
v + 2

+
r r

r
!
1
2 cot v
v
,
+

r2 sin2 r2 sin 3r

v
p
1
1 2
v

+
2
r sin r sin
r sin2
!

1
vr 2 cot v
2
,
+
+
+
r2 sin2 r2 sin 3r sin
!
M 2 p
+
,

(2.163)

(2.164)
(2.165)
(2.166)

(2.167)
(2.168)

where

1 (sin v )
1 v
1 (r2 vr )
+
+
,
2
r
r sin

r sin
r

(2.169)

28

FLUID MECHANICS
v

v
+ vr
+
+
,
t
r
r r sin
!
!
1 2
1

2
2
=
r
+
sin

+
,
r

r2 r
r2 sin
r2 sin2 2

!
!2
!2
vr 2
1 v vr
1 v vr cot v

+
+
+
+ +
= 2
r
r
r
r sin
r
r
!2
!2
1
1 vr v v
1 1 vr v v
+
+

+
2 r
r
r
2 r sin
r
r
!2
1
1 v 1 v cot v
+
+

.
2 r sin r
r

D
Dt

(2.170)
(2.171)

(2.172)

2.21 Exercises
2.1. Equations (2.66), (2.75), and (2.87) can be combined to give the following energy conservation equation for a non-ideal
compressible fluid:
DE p D

= q,

Dt
Dt
where is the mass density, p the pressure, E the internal energy per unit mass, the viscous energy dissipation rate per unit
volume, and q the heat flux density. We also have
D
Dt
q

v,

T,

where v is the fluid velocity, T the temperature, and the thermal conductivity. Now, according to a standard theorem in
thermodynamics,
p
T dS = dE 2 d,

where S is the entropy per unit mass. Moreover, the entropy flux density at a given point in the fluid is
s = Sv +

q
,
T

where the first term on the right-hand side is due to direct entropy convection by the fluid, and the second is the entropy flux
density associated with heat conduction.
Derive an entropy conservation equation of the form
dS
+ S = S ,
dt
where S is the net amount of entropy contained in some fixed volume V, S the entropy flux out of V, and S the net rate of
entropy creation within V. Give expressions for S , S , and S . Demonstrate that the entropy creation rate per unit volume
is
qq
.
= +
T
T2
Finally, show that 0, in accordance with the second law of thermodynamics.

2.2. The Navier-Stokes equation for an incompressible fluid of uniform mass density takes the form
Dv
p
=
+ 2 v,
Dt

where v is the fluid velocity, p the pressure, the potential energy per unit mass, and the (uniform) kinematic viscosity.
The incompressibility constraint requires that
v = 0.
Finally, the quantity

Mathematical Models of Fluid Motion

29

is generally referred to as the fluid vorticity.


Derive the following vorticity evolution equation from the Navier-Stokes equation:
D
= ( ) v + 2 .
Dt
2.3. Consider two-dimensional incompressible fluid flow. Let the velocity field take the form
v = vx (x, y, t) ex + vy (x, y, t) ey .
Demonstrate that the equations of incompressible fluid flow (see Exercise 2.2) can be satisfied by writing

where

,
y

vx

vy

,
x

= 2 ,
t
x y
x y

and
= 2 .

Here, 2 = 2 /x2 + 2 /y2 . Furthermore, the quantity is termed a stream function, since v = 0: i.e., the fluid flow is
everywhere parallel to contours of .
2.4. Consider incompressible irrotational flow: i.e., flow that satisfies
Dv
Dt

0,

p
+ 2 v,

as well as
v = 0.

Here, v is the fluid velocity, the uniform mass density, p the pressure, the potential energy per unit mass, and the
(uniform) kinematic viscosity.
Demonstrate that the above equations can be satisfied by writing
v = ,
where
2 = 0,

and

1 2 p
+ v + + = C(t).
t
2

Here, C(t) is a spatial constant. This type of flow is known as potential flow, since the velocity field is derived from a scalar
potential.

2.5. The equations of inviscid adiabatic ideal gas flow are


D
Dt
Dv
Dt
!
D p
Dt

v,

0.

p
,

Here, is the mass density, v the flow velocity, p the pressure, the potential energy per unit mass, and the (uniform)
ratio of specific heats. Suppose that the pressure and potential energy are both time independent: i.e., p/t = /t = 0.
Demonstrate that
1
p
H = v2 +
+
2
1
is a constant of the motion: i.e., DH/Dt = 0. This result is known as Bernoullis theorem.

30

FLUID MECHANICS

2.6. The equations of inviscid adiabatic non-ideal gas flow are


D
Dt
Dv
Dt
p D
DE
2
Dt
Dt

v,

0.

p
,

Here, is the mass density, v the flow velocity, p the pressure, the potential energy per unit mass, and E the internal
energy per unit mass. Suppose that the pressure and potential energy are both time independent: i.e., p/t = /t = 0.
Demonstrate that
1
p
H = v2 + E + +
2

is a constant of the motion: i.e., DH/Dt = 0. This result is a more general form of Bernoullis theorem.
2.7. Demonstrate that Bernoullis theorem for incompressible, inviscid fluid flow takes the form DH/Dt = 0, where
H=

1 2 p
v + + .
2

Hydrostatics

31

3 Hydrostatics

3.1 Introduction
This chapter discusses hydrostatics, which is the study of the mechanical equilibrium of incompressible fluids.

3.2 Hydrostatic Pressure


Consider a body of water that is stationary in a reference frame that is fixed with respect to the Earths surface. In this
chapter, such a frame is treated as approximately inertial. Let z measure vertical height, and suppose that the region
z 0 is occupied by water, and the region z > 0 by air. According to Equation (2.79), the air/water system remains in
mechanical equilibrium (i.e., v = Dv/Dt = 0) provided
0 = p + ,

(3.1)

where p is the static fluid pressure, the mass density, = g z the gravitational potential energy per unit mass, and g
the (approximately uniform) acceleration due to gravity. Now,
(
0
z>0
(z) =
,
(3.2)
0
z0
where 0 is the (approximately uniform) mass density of water. Here, the comparatively small mass density of air has
been neglected. Since = (z) and = (z), it immediately follows, from (3.1), that p = p(z), where
dp
= g.
dz

(3.3)

We conclude that constant pressure surfaces in a stationary body of water take the form of horizontal planes. Making
use of (3.2), the above equation can be integrated to give
(
p0
z>0
p(z) =
,
(3.4)
p0 0 g z
z0
where p0 105 N m2 is atmospheric pressure at ground level. According to this expression, pressure in stationary
water increases linearly with increasing depth (i.e., with decreasing z, for z < 0). In fact, given that g 9.8 m s2 and
0 103 kg m3 , we deduce that hydrostatic pressure rises at the rate of 1 atmosphere (i.e., 105 N m2 ) every 10.2 m
increase in depth below the surface.

3.3 Buoyancy
Consider the air/water system described in the previous section. Let V be some volume, bounded by a closed surface
S , that straddles the plane z = 0, and is thus partially occupied by water, and partially by air. The i-component of the
net force acting on the fluid (i.e., either water or air) contained within V is written (see Section 2.3)
Z
Z
Fi dV,
(3.5)
i j dS j +
fi =
V

where
i j = p i j

(3.6)

is the stress tensor for a static fluid (see Section 2.5), and
F = g ez

(3.7)

32

FLUID MECHANICS

the gravitational force density. (Recall that the indices 1, 2, and 3 refer to the x-, y-, and z-axes, respectively. Thus,
f3 fz , etc.) The first term on the right-hand side of (3.5) represents the net surface force acting across S , whereas the
second term represents the net volume force distributed throughout V. Making use of the tensor divergence theorem
(see Section B.4), Equations (3.5)(3.7) yield the following expression for the net force:
f = B + W,
where
Bi =

p
dV,
xi

(3.8)

(3.9)

and
Wx

Wz

Wy = 0,
Z
g dV.

(3.10)
(3.11)

Here, B is the net surface force, and W the net volume force.
It follows from Equations (3.4) and (3.9) that
B = M0 g ez ,

(3.12)

where M0 = 0 V0 . Here, V0 is volume of that part of V which lies below the waterline, and M0 the total mass of water
contained within V. Moreover, from Equations (3.2), (3.10), and (3.11),
W = M0 g ez .

(3.13)

It can be seen that the net surface force, B, is directed vertically upward, and exactly balances the net volume force, W,
which is directed vertically downward. Of course, W is the weight of the water contained within V. On the other hand,
B, which is generally known as the buoyancy force, is the resultant pressure of the water immediately surrounding
V. We conclude that, in equilibrium, the net buoyancy force acting across S exactly balances the weight of the water
inside V, so that the total force acting on the contents of V is zero, as must be the case for a system in mechanical
equilibrium. We can also deduce that the line of action of B (which is vertical) passes through the center of gravity of
the water inside V. Otherwise, a net torque would act on the contents of V, which would contradict our assumption
that the system is in mechanical equilibrium.

3.4 Equilibrium of Floating Bodies


Consider the situation described in the previous section. Suppose that the fluid contained within V is replaced by
a partially submerged solid body whose outer surface corresponds to S . Furthermore, suppose that this body is in
mechanical equilibrium with the surrounding fluid (i.e., it is stationary, and floating on the surface of the water). It
follows that the pressure distribution in the surrounding fluid is unchanged [since the force balance criterion (3.3) can
be integrated to give the pressure distribution (3.4) at all contiguous points in the fluid, provided that the fluid remains
in mechanical equilibrium]. We conclude that the net surface force acting across S is also unchanged (since this is
directly related to the pressure distribution in the fluid immediately surrounding V), which implies that the buoyancy
force acting on the floating body is the same as that acting on the displaced water: i.e., the water that previously
occupied V. In other words, from (3.12),
B = W0 ez ,
(3.14)
where W0 = M0 g and M0 are the weight and mass of the displaced water, respectively. The fact that the buoyancy
force is unchanged also implies that the vertical line of action of B passes through the center of gravity, H (say), of the
displaced water. Incidentally, H is generally known as the center of buoyancy.
A floating body of weight W is acted upon by two forces: namely, its own weight,
W = W ez ,

(3.15)

Hydrostatics

33

and the buoyancy force, B = W0 ez , due to the pressure of the surrounding water. Of course, the line of action of W
passes through the bodys center of gravity, G (say). Now, to remain in equilibrium, the body must be subject to zero
net force and zero net torque. The requirement of zero net force yields W0 = W. In other words, in equilibrium, the
weight of the water displaced by a floating body is equal to the weight of the body, or, alternatively, in equilibrium, the
magnitude of the buoyancy force acting on a floating body is equal to the weight of the displaced water. This famous
result is known as Archimedes principle. The requirement of zero net torque implies that, in equilibrium, the center
of gravity, G, and center of buoyancy, H, of a floating body lie on the same vertical straight-line.
Consider a floating body of mass M and volume V. Let = M/V be the bodys mean mass density. Archimedes
principle implies that, in equilibrium,
V0
= s,
(3.16)
V
where

s=
(3.17)
0
is termed the bodys specific gravity. (Recall, that V0 is the submerged volume, and 0 the mass density of water.)
We conclude, from (3.16), that the volume fraction of a floating body that is submerged is equal to the bodys specific
gravity. Obviously, the specific gravity must be less than unity, since the submerged volume fraction cannot exceed
unity. In fact, if the specific gravity exceeds unity then it is impossible for the buoyancy force to balance the bodys
weight, and the body consequently sinks.
Consider a body of volume V and specific gravity s that floats in equilibrium. It follows, from Equation (3.16), that
the submerged volume is V0 = s V. Hence, the volume above the waterline is V1 = V V0 = (1 s) V. Suppose that the
body is inverted such that its previously submerged part is raised above the waterline, and vice versa: i.e., V0 V1 .
According to (3.16), the body can only remain in equilibrium in this configuration if its specific gravity changes to
s =

V1 V V0
=
= 1 s.
V
V

(3.18)

We conclude that for every equilibrium configuration of a floating body of specific gravity s there exists an inverted
equilibrium configuration for a body of the same shape having the complementary specific gravity 1 s.

3.5 Vertical Stability of Floating Bodies


Consider a floating body of weight W which, in equilibrium, has a submerged volume V0 . Thus, the bodys downward
weight is balanced by the upward buoyancy force, B = 0 V0 g: i.e., 0 V0 g = W. Let A0 be the cross-sectional
area of the body at the waterline (i.e., in the plane z = 0). It is convenient to define the bodys mean draft (or mean
submerged depth) as 0 = V0 /A0 . Suppose that the body is displaced slightly downward, without rotation, such that
its mean draft becomes 0 + 1 , where |1 | 0 . Assuming that the cross-sectional area in the vicinity of the waterline
is constant, the new submerged volume is V0 = A0 (0 + 1 ) = V0 + A0 1 , and the new buoyancy force becomes
B = 0 V0 g = W + 0 A0 g 1 = (1 + 1 /0 ) W. However, the weight of the body is unchanged. Thus, the bodys
perturbed vertical equation of motion is written
W d 2 1
W
= W B = 1 ,
g dt2
0

(3.19)

which reduces to the simple harmonic equation


d 2 1
g
= 1 .
0
dt2

(3.20)

We conclude that when a floating body of mean draft 0 is subject to a small vertical displacement it oscillates about
its equilibrium position at the characteristic frequency
r
g
.
(3.21)
=
0

34

FLUID MECHANICS

STABLE

UNSTABLE

Figure 3.1: Stable and unstable configurations for a floating body.


It follows that such a body is unconditionally stable to small vertical displacements. Of course, the above analysis
presupposes that the oscillations take place sufficiently slowly that the water immediately surrounding the body always
remains in approximate hydrostatic equilibrium.

3.6 Angular Stability of Floating Bodies


Let us now investigate the stability of floating bodies to angular displacements. For the sake of simplicity, we shall
only consider bodies that have two mutually perpendicular planes of symmetry. Suppose that when such a body is in
an equilibrium state the two symmetry planes are vertical, and correspond to the x = 0 and y = 0 planes. As before, the
z = 0 plane coincides with the surface of the water. It follows, from symmetry, that when the body is in an equilibrium
state its center of gravity, G, and center of buoyancy, H, both lie on the z-axis.
Suppose that the body turns through a small angle about some horizontal axis, lying in the plane z = 0, that goes
through the origin. Let GH be that, originally vertical, straight-line that passes through the bodys center of gravity,
G, and original center of buoyancy, H. Owing to the altered shape of the volume of displaced water, the center of
buoyancy is shifted to some new position H . Let the vertical straight-line going through H meet GH at M. See
Figure 3.1. Point M (in the limit 0) is called the metacenter. In the disturbed state, the bodys weight W acts
downward through G, and the buoyancy 0 V0 g acts upward through M. Let us assume that the submerged volume,
V0 , is unchanged from the equilibrium state (which excludes vertical oscillations from consideration). It follows that
the weight and the buoyancy force are equal and opposite, so that there is no net force on the body. However, as can be
seen from Figure 3.1, the weight and the buoyancy force generate a net torque of magnitude = W sin . Here, is
the length MG: i.e., the distance between the metacenter and the center of gravity. This distance is generally known as
the metacentric height, and is defined such that it is positive when M lies above G, and vice versa. Moreover, as is also
clear from Figure 3.1, when M lies above G the torque acts to reduce , and vice versa. In the former case, the torque
is known as a righting torque. We conclude that a floating body is stable to small angular displacements about some
horizontal axis lying in the plane z = 0 provided that its metacentric height is positive: i.e., provided that the metacenter
lies above the center of gravity. Since we have already demonstrated that a floating body is unconditionally stable to
small vertical displacements (and since it is also fairly obvious that such a body is neutrally stable to both horizontal
displacements and angular displacements about a vertical axis passing through its center of gravity), it follows that
a necessary and sufficient condition for the stability of a floating body to a general small perturbation (made up of
arbitrary linear and angular components) is that its metacentric height be positive for angular displacements about any

Hydrostatics

35

horizontal axis.

3.7 Determination of Metacentric Height


Suppose that the floating body considered in the previous section is in an equilibrium state. Let A0 be the crosssectional area at the waterline: i.e., in the plane z = 0. Since the body is assumed to be symmetric with respect to the
x = 0 and y = 0 planes, we have
Z
Z
Z
x y dx dy = 0,
(3.22)
y dx dy =
x dx dy =
A0

A0

A0

where the integrals are taken over the whole cross-section at z = 0. Let (x, y) be the bodys draft: i.e., the vertical
distance between the surface of the water and the bodys lower boundary. It follows, from symmetry, that (x, y) =
(x, y) and (x, y) = (x, y). Moreover, the submerged volume is
Z
Z Z
(x, y) dx dy.
(3.23)
dx dy dz =
V0 =
A0

A0

It also follows from symmetry that


Z

x (x, y) dx dy =
A0

y (x, y) dx dy = 0.

Now, the depth of the unperturbed center of buoyancy below the surface of the water is
R R
Z
z dx dy dz
2 A0
1
A0 0
h=
=
2 (x, y) dx dy = 0 ,
V0
2 V0 A 0
2 V0
where

(3.24)

A0

R
1/2
A 2 (x, y) dx dy
0

.
0 =

A0

(3.25)

(3.26)

Finally, from symmetry, the unperturbed center of buoyancy lies at x = y = 0.


Suppose that the body now turns through a small angle about the x-axis. As is easily demonstrated, the bodys
new draft becomes (x, y) (x, y) y. Hence, the new submerged volume is
Z
Z

y dx dy = V0 ,
(3.27)
[(x, y) y] dx dy = V0 +
V0 =
A0

A0

where use has been made of Equations (3.22) and (3.23). Thus, the submerged volume is unchanged, as should be the
case for a purely angular displacement. The new depth of the center of buoyancy is
R R
Z
z dx dy dz
1
A0 0

h =
=
[ 2 (x, y) 2 y (x, y) + O( 2 )] dx dy = h,
(3.28)
V0
2 V0 A 0
where use has been made of (3.24) and (3.25). Thus, the depth of the center of buoyancy is also unchanged. Moreover,
from symmetry, it is clear that the center of buoyancy still lies at x = 0. Finally, the new y-coordinate of the center of
buoyancy is
R
R R
y [(x, y) y] dx dy
y dx dy dz
2 A0
A
A0 0
= 0
= x
,
(3.29)
V0
V0
V0
where use has been made of (3.24). Here,
1/2
R 2
A y dx dy
0
,
(3.30)
x =

A0

36

FLUID MECHANICS

is the radius of gyration of area A0 about the x-axis.


It follows, from the above analysis, that if the floating body under consideration turns through a small angle about
the x-axis then its center of buoyancy shifts horizontally a distance x2 A0 /V0 in the plane perpendicular to the axis of
rotation. In other words, the distance HH in Figure 3.1 is x2 A0 /V0 . Simple trigonometry reveals that HH /MH
(assuming that is small). Hence, MH = HH / = x2 A0 /V0 . Now, MH is the height of the metacenter relative to the
center of buoyancy. However, the center of buoyancy lies a depth h below the surface of the water (which corresponds
to the plane z = 0). Hence, the z-coordinate of the metacenter is z M = x2 A0 /V0 h. Finally, if zG and zH = h are the
z-coordinates of the unperturbed centers of gravity and buoyancy, respectively, then
x2 A0
+ zH ,
V0

(3.31)

x2 A0
zGH ,
V0

(3.32)

zM =
and the metacentric height, = z M zG , becomes
=

where zGH = zG zH . Note that, since x2 A0 /V0 > 0, the metacenter always lies above the center of buoyancy.
A simple extension of the above argument reveals that if the body turns through a small angle about the y-axis
then the metacentric height is
y2 A0
=
zGH ,
(3.33)
V0
where

1/2
R 2
A x dx dy
0
,
y =

A0

(3.34)

is radius of gyration of area A0 about the y-axis. Finally, as is easily demonstrated, if the body rotates about a horizontal
axis which subtends an angle with the x-axis then
=

2 A0
zGH ,
V0

(3.35)

where
2 = x2 cos2 + y2 sin2 .
2

x2

(3.36)

y2 .

Thus, the minimum value of is the lesser of and


It follows that the equilibrium state in question is unconditionally stable provided it is stable to small amplitude angular displacements about horizontal axes normal to its two
vertical symmetry planes (i.e., the x = 0 and y = 0 planes).
As an example, consider a uniform rectangular block of specific gravity s floating such that its sides of length a, b,
and c are parallel to the x-, y-, and z-axes, respectively. Such a block can be thought of as a very crude model of a ship.
The volume of the block is V = a b c. Hence, the submerged volume is V0 = s V = s a b c. The cross-sectional area of
the block at the waterline (z = 0) is A0 = a b. It is easily demonstrated that (x, y) = 0 = V0 /A0 = s c. Thus, the center
of buoyancy lies a depth h = 02 A0 /2 V0 = s c/2 below the surface of the water [see Equation (3.25)]. Moreover, by
symmetry, the center of gravity is a height c/2 above the bottom surface of the block, which is located a depth s c
below the surface of the water. Hence, zH = h = s c/2, zG = c/2 s c, and zGH = c (1 s)/2. Consider the stability
of the block to small amplitude angular displacements about the x-axis. We have
x2

R a/2 R b/2

a/2 b/2

y2 dx dy

ab

b2
.
12

(3.37)

Hence, from (3.32), the metacentric height is


=

b2
c
(1 s).
12 s c 2

(3.38)

Hydrostatics

37

The stability criterion > 0 yields


b2
s (1 s) > 0.
(3.39)
6 c2
Since the maximum value that s (1 s) can take is 1/4, it follows that the block is stable for all specific gravities when
r
2
c < c0 =
b.
(3.40)
3
On the other hand, if c > c0 then the block is unstable for intermediate specific gravities such that s < s < s+ , where
q
1 1 c02 /c 2
s =
,
(3.41)
2
and is stable otherwise. Assuming that the block is stable, its angular equation of motion is written
I
where
I=

W
gV

a/2

a/2

b/2

b/2

d2
= W sin W ,
dt2

cs c

(y2 + z2 ) dx dy dz =

s c


W  2
b + 4 [(1 s)3 + s3 ] c2
12 g

(3.42)

(3.43)

is the moment of inertia of the block about the x-axis. Thus, we obtain the the simple harmonic equation
d2
= 2 ,
dt2
where
2 =

c02 4 s (1 s) c2
W
g
.
=
2
I
s c c0 + (8/3) [(1 s)3 + s3 ] c2

(3.44)

(3.45)

We conclude that the block executes small amplitude angular oscillations about the x-axis at the angular frequency
. However, this result is only accurate in the limit in which the oscillations are sufficiently slow that the water
surrounding the block always remains in approximate hydrostatic equilibrium. For the case of rotation about the
y-axis, the above analysis is unchanged except that a b.
The metacentric height of a conventional ship whose length greatly exceeds its width is typically much less for
rolling (i.e., rotation about a horizontal axis running along the ships length) than for pitching (i.e., rotation about a
horizontal axis perpendicular to the ships length), since the radius of gyration for pitching greatly exceeds that for
rolling. As is clear from Equation (3.45), a ship with a relatively small metacentric height (for rolling) has a relatively
long roll period, and vice versa. Now, an excessively low metacentric height increases the chances of a ship capsizing
if the weather is rough, or if its cargo/ballast shifts, or if it is damaged and partially flooded. For this reason, maritime
regulatory agencies, such as the International Maritime Organization, specify minimum metacentric heights for various
different types of sea-going vessel. A relatively large metacentric height, on the other hand, generally renders a ship
uncomfortable for passengers and crew, because the ship executes short period rolls, resulting in large g-forces. Such
forces also increase the risk that cargo may break loose or shift.
We saw earlier, in Section 3.4, that if a body of specific gravity s floats in vertical equilibrium in a certain position
then a body of the same shape, but of specific gravity 1 s, can float in vertical equilibrium in the inverted position.
We can now demonstrate that these positions are either both stable, or both unstable, provided the body is of uniform
density. Let V1 and V2 be the volumes that are above and below the waterline, respectively, in the first position. Let H1
and H2 be the mean centers of these two volumes, and H that of the whole volume. It follows that H2 is the center of
buoyancy in the first position, H1 the center of buoyancy in the second (inverted) position, and H the center of gravity
in both positions. Moreover,
!
V1 V2
H1 H2 ,
(3.46)
V 1 H1 G = V 2 H2 G =
V1 + V2

38

FLUID MECHANICS

where H1 G is the distance between points H1 and G, etc. The metacentric heights in the first and second positions are
#
"
!
1
V1 V2
2 A
(3.47)
A 2
H1 H2 ,
H1 G =
1 =
V1
V1
V1 + V2
#
"
!
2 A
1
V1 V2
2 =
(3.48)
A 2
H1 H2 ,
H2 G =
V2
V2
V1 + V2
respectively, where A and are the area and radius of gyration of the common waterline section, respectively. Thus,
"
!
#2
1
V1 V2
2
1 2 =
A
H1 H2 0,
(3.49)
V1 V2
V1 + V2
which implies that 1 T 0 as 2 T 0, and vice versa. It follows that the first and second positions are either both stable,
both marginally stable, or both unstable.

3.8 Energy of a Floating Body


The conditions governing the equilibrium and stability of a floating body can also be deduced from the principle of
energy.
For the sake of simplicity, let us suppose that the water surface area is infinite, so that the immersion of the body
does not generate any change in the water level. The potential energy of the body itself is W zG , where W is the bodys
weight, and zG the height of its center of gravity, G, relative to the surface of the water. If the body displaces a volume
V0 of water then this effectively means that a weight 0 V0 of water, whose center of gravity is located at the center
of buoyancy, H, is removed, and then spread as an infinitely thin film over the surface of the water. This involves a
gain of potential energy of 0 V0 zH , where zH is the height of H relative to the surface of the water. Now, vertical
force balance requires that W = 0 V0 . Thus, the potential energy of the system is W zGH (modulo an arbitrary additive
constant), where zGH = zG zH is the height of the center of gravity relative to the center of buoyancy.
According to the principles of statics, an equilibrium state corresponds to either a minimum or a maximum of the
potential energy. However, such an equilibrium is only stable when the potential energy is minimized. Thus, it follows
that a stable equilibrium configuration of a floating body is such as to minimize the height of the bodys center of
gravity relative to its center of buoyancy.

3.9 Curve of Buoyancy


Consider a floating body in vertical force balance that is slowly rotated about a horizontal axis normal to one of its
vertical symmetry planes. Let us take the center of gravity, G, which necessarily lies in this plane, as the origin of a
coordinate system that is fixed with respect to the body. As illustrated in Figure 3.2, as the body rotates, the locus of its
center of buoyancy, H, as seen in the fixed reference frame, appears to traces out a curve, AB, in the plane of symmetry.
This curve is known as the curve of buoyancy. Let r represent the radial distance from the origin, G, to some point,
H, on the curve of buoyancy. Note that the tangent to the curve of buoyancy is always orientated horizontally. This
follows because, as was shown in the previous section, small rotations of a floating body in vertical force balance cause
its center of buoyancy to shift horizontally, rather than vertically, in the plane perpendicular to the axis of rotation.
Thus, the difference in vertical height, zGH , between the center of gravity and the center of buoyancy is equal to the
perpendicular distance, p, between G and the tangent to the curve of buoyancy at H. An equilibrium configuration
therefore corresponds to a maximum or a minimum of p as point H moves along the curve of buoyancy. However,
the equilibrium is only stable if p is minimized. Now, if R is the radius of curvature of the curve of buoyancy then,
according to a standard result in differential calculus,
dr
.
dp

(3.50)

R
p,
r

(3.51)

R=r
Writing this result in the form
r =

Hydrostatics

39

r
p
H0
H
B

Figure 3.2: Curve of buoyancy for a floating body.


it can be seen that maxima and minima of p, which are the points on the curve of buoyancy where p = 0, correspond
to the points where r = 0, and are, thus, coincident with maxima and minima of r. In other words, an equilibrium
configuration corresponds to a point of maximum or minimum r on the curve of buoyancy: i.e., a point at which GH
meets the curve at right-angles. At such a point, r = p, and the potential energy consequently takes the value W r.
Let H0 be a point on the curve of buoyancy, and let r0 , p0 , and R0 be the corresponding values of r, p, and R. For
neighboring points on the curve, we can write

dr
(p p0 ),
(3.52)
r r0 =
d p H0
or

r r0 =

R0
(p p0 ).
r0

(3.53)

It follows that p p0 has the same sign as r r0 (since R0 and r0 are both positive). [The fact that R0 is positive
(i.e., dr/d p > 0) follows from the previously established result that the metacenter, which is the center of curvature of
the curve of buoyancy, always lies above the center of buoyancy, implying that the curve of buoyancy is necessarily
concave upwards.] Hence, the minima and maxima of r occur simultaneously with those of p. Consequently, a stable
equilibrium configuration corresponds to a point of minimum r on the curve of buoyancy: i.e., a minimum in the
distance GH between the center of gravity and the center of buoyancy.
We can use the above result to determine the stable equilibrium configurations for a beam of square cross-section,
and uniform specific gravity s, that floats with its length horizontal. In order to achieve this goal, we must calculate
the distance GH for all possible configurations of the beam that are in vertical force balance. However, we need only
consider cases where s < 1/2, since, according to the analysis of Section 3.6, for every stable equilibrium configuration
with s = s0 < 1/2 there is a corresponding stable inverted configuration with s = 1 s0 > 1/2, and vice versa.
Let us define fixed rectangular axes, x and y, passing through the center of the middle section of the beam, and
running parallel to its sides. Let us start with the case where the waterline PQ is parallel to a side. See Figure 3.3. If
the length of a side is 2 a then (3.16) yields
AP = BQ = 2a s.
(3.54)

40

FLUID MECHANICS

x
Q

P
P

Figure 3.3: Beam of square cross-section floating with two corners immersed.
Suppose that the beam is turned through an angle > 0 such that the waterline assumes the position P Q , in
Figure 3.3, but still intersects two opposite sides. The lengths AP and BQ satisfy
BQ AP = 2 a tan .

(3.55)

Moreover, the area of the trapezium P ABQ must match that of the rectangle PABQ in order to ensure that the
submerged volume remains invariant (otherwise, the beam would not remain in vertical force balance): i.e.,
(AP + BQ ) a = 4 a2 s.

(3.56)

AP

a (2 s tan ),

(3.57)

a (2 s + tan ).

(3.58)

It follows that

BQ

The constraint that the waterline intersect two opposite sides of the beam implies that AP > 0, and, hence, that
tan < 2 s.
The coordinates of the center of buoyancy, H, which is the mean center of the trapezium P ABQ , are
Ra Ra
x dx dy (2/3) a3 tan
a
a ah(x)
tan ,
x = R a R a
=
=
2s
6
s
4
a
dx dy
a ah(x)
Ra Ra
y dx dy
a ah(x)
y = R a R a
dx dy
a ah(x)
=

a
4 a3 s (1 s) (1/3) a3 tan2
tan2 ,
= (1 s) a
12 s
4 a2 s

(3.59)

(3.60)

(3.61)

where
h(x) = 2 a s + x tan .

(3.62)

Hydrostatics

41

Q
O

Figure 3.4: Beam of square cross-section floating with one corner immersed.
Thus, if u = r2 /a2 = ( x 2 + y 2 )/a2 then
#2
"
t2
t2
,
u=
+ (1 s)
12 s
36 s2

(3.63)

where t = tan . Now, a stable equilibrium state corresponds to a minimum of r with respect to , and, hence, of u with
respect to t. However,
du
dt
d2 u
dt2

=
=

t
36 s2
1
36 s2

h
i
t2 12 s (1 s) + 2 ,

h
i
3 t2 12 s (1 s) + 2 .

(3.64)
(3.65)

The minima and maxima of u occur when du/dt = 0, d 2 u/dt2 > 0 and du/dt = 0, d 2 u/dt2 < 0, respectively. It follows
that the symmetrical position, t = 0, in which the sides of the beam are either parallel or perpendicular to the waterline,
is always an equilibrium, but is only stable when
s2 s +

1
>0:
6

(3.66)

i.e., when s < 1/2 1/ 12 = 0.2113. It is also possible to obtain equilibria in asymmetric positions such that t is the
root of
t2 = 12s (1 s) 2.
(3.67)
Such equilibria only exist for s > 0.2113, and are stable. Finally, in order to satisfy the constraint (3.59), we must have
t < 2 s, which, in combination with the above equation, implies that
8 s2 6 s + 1 > 0,

(3.68)

or s < 0.25.
Suppose that the constraint (3.59) is not satisfied, so that the immersed portion of the beams cross-section is
triangular. See Figure 3.4. It is clear that
BQ
= tan .
(3.69)
BP

42

FLUID MECHANICS

Moreover, the area of the triangle P BQ , in Figure 3.4, must match that of the rectangle PABQ, in Figure 3.3, in order
to ensure that the submerged volume remain invariant: i.e.,
1
BP BQ = 4 a2 s.
2

(3.70)

It follows that
BP
BQ

=
=

(8 s/ tan )1/2 a,
(8 s tan )

1/2

(3.71)

a,

(3.72)

or, writing z2 = tan and 2 = (8/9) s,


BP

3 z1 a,

(3.73)

3 z a.

(3.74)

BQ

The coordinates of the center of buoyancy, H, which is the mean center of triangle P BQ , are
x

y =

a BP /3 = a (1 z 1 ),

(3.75)

a BQ /3 = a (1 z),

(3.76)

since the perpendicular distance of the mean center of a triangle from one of its sides is one third of the perpendicular
distance from the side to the opposite vertex. Thus, if u = r2 /a2 = ( x 2 + y 2 )/a2 then
u =

(1 z1 )2 + (1 z)2 ,
2

du
dz

2 (z 1) (z
z3

d2 u
dz2

2 2 (z4 2 1 z + 3)
.
z4

z + 1)

(3.77)
,

(3.78)
(3.79)

Moreover, the constraint (3.59) yields


3
.
(3.80)
2
The stable and unstable equilibria correspond to du/dz = 0, d 2 u/dz2 > 0 and du/dz, d 2 u/dz2 < 0, respectively. It
follows that the symmetrical position, z = 1, in which the diagonals of the beam are either parallel or perpendicular to
the waterline is an equilibrium provided < 2/3, or s < 1/2, but is only stable when > 1/2, or s > 9/32 = 0.28125.
It is also possible to obtain equilibria in asymmetric positions such that z is the root of
z>

z2 1 z + 1 = 0.

(3.81)

Such equilibria only exist for 2/3 < < 1/2, or 1/4 < s < 9/32, and are stable.
In summary, the stable equilibrium configurations of a beam of square cross-section, floating with its length horizontal, are such that the sides are either parallel or perpendicular to the waterline for s < 0.2113, such that two corners
are immersed but the sides and diagonals are neither parallel nor perpendicular to the waterline for 0.2113 < s < 0.25,
such that only one corner is immersed but the sides and diagonals are neither parallel nor perpendicular to the waterline for 0.25 < s < 0.28125, and such that the diagonals are either parallel or perpendicular to waterline for
0.28125 < s < 0.5. For s > 0.5, the stable configurations are the same as those for a beam with the complimentary
specific gravity 1 s.

3.10 Rotational Hydrostatics


Consider the equilibrium of an incompressible fluid that is uniformly rotating at a fixed angular velocity in some
inertial frame of reference. Of course, such a fluid appears stationary in a non-inertial co-rotating reference frame.

Hydrostatics

43

Moreover, according to standard Newtonian dynamics, the force balance equation for the fluid in the co-rotating frame
takes the form (cf., Section 3.2)
0 = p + + ( r),
(3.82)
where p is the static fluid pressure, the mass density, the gravitational potential energy per unit mass, and r a
position vector (measured with respect to an origin that lies on the axis of rotation). The final term on the right-hand
side of the above equation represents the fictitious centrifugal force density. Without loss of generality, we can assume
that = ez . It follows that
0 = p + ( + ),
(3.83)
where

1
(3.84)
= 2 (x2 + y2 )
2
is the so-called centrifugal potential. Recall, incidentally, that is a uniform constant in an incompressible fluid.
As an example, consider the equilibrium of a body of water, located on the Earths surface, that is uniformly
rotating about a vertical axis at the fixed angular velocity . It is convenient to adopt cylindrical coordinates (see
Section C.3), r, , z, whose symmetry axis coincides with the axis of rotation. Let z increase upward. It follows that
= g z and = (1/2) 2 r2 . Assuming that the pressure distribution is axisymmetric, so that p = p(r, z), the force
balance equation, (3.83), reduces to
( + )
p
+
r
r
p
( + )
+
z
z

0,

(3.85)

0,

(3.86)

or
p
2 r
r
p
+ g
z

0,

(3.87)

0.

(3.88)

!
1 2 2
r gz ,
2

(3.89)

The previous two equations can be integrated to give


p(r, z) = p0 +

where p0 is a constant. Thus, constant pressure surfaces in a uniformly rotating body of water take the form of
paraboloids of revolution about the rotation axis. Suppose that p0 represents atmospheric pressure. In this case, the
surface of the water is the locus of p(r, z) = p0 : i.e., it is the constant pressure surface whose pressure matches that of
the atmosphere. It follows that the surface of the water is the paraboloid of revolution
z=

2 2
r ,
2g

(3.90)

where r is the perpendicular distance from the axis of rotation, and z = 0 the on-axis height of the surface.
Now, from Section 3.3, it is plain that the buoyancy force acting on any co-rotating solid body, which is wholly
or partially immersed in the water, is the same as that which would maintain the mass of water displaced by the body
in relative equilibrium. In the case of a floating body, this mass is limited by the continuation of the waters curved
surface through the body. Let points G and H represent the centers of gravity and buoyancy, respectively, of the body.
Of course, the latter point is simply the center of gravity of the displaced water. Suppose that G and H are located
perpendicular distances rG and rH from the axis of rotation, respectively. Finally, let M be the mass of the body,
and M0 the mass of the displaced water. It follows that the buoyancy force has an upward vertical component M0 g,
and an outward horizontal component M0 2 rH . Thus, according to standard Newtonian dynamics, the equation of
horizontal motion of a general co-rotating body is

..

M (r 2 rG ) = M0 2 rH ,

(3.91)

44

FLUID MECHANICS

where = d/dt. Now, from Archimedes principle, M0 = M for the case of a floating body that is less dense than
water. However, if the body is of uniform density then rH > rG , as a consequence of the curvature of the waters
surface. Hence, we obtain
..r = 2 (r r ) < 0.
(3.92)
H
G
In other words, a floating body drifts radially inward towards the rotation axis. On the other hand, M0 < M for a fully
submerged body that is more dense than water. However, if the body is of uniform density then its centers of gravity
and buoyancy coincide with one another, so that rH = rG . Hence, we obtain

..r = (M M ) 2 r > 0.
0
G

(3.93)

In other words, a fully submerged body drifts radially outward from the rotation axis. The above analysis accounts for
the common observation that objects heavier than water, such as grains of sand, tend to collect on the outer side of a
bend in a fast flowing river, whilst floating objects, such as sticks, tend to collect on the inner side.

3.11 Equilibrium of a Rotating Liquid Body


Consider a self-gravitating liquid body in outer space that is rotating uniformly about some fixed axis passing through
its center of mass. What is the shape of the bodys bounding surface? This famous theoretical problem had its origins
in investigations of the figure of a rotating planet, such as the Earth, that were undertaken by Newton, Maclaurin,
Jacobi, Meyer, Liouville, Dirichlet, Dedekind, Riemann, and other celebrated scientists, in the 17th, 18th, and 19th
centuries.1 Incidentally, it is reasonable to treat the Earth as a liquid, for the purpose of this calculation, because the
shear strength of the solid rock out of which the terrestrial crust is composed is nowhere near sufficient to allow the
actual shape of the Earth to deviate significantly from that of a hypothetical liquid Earth.
Now, in a co-rotating reference frame, the shape of a self-gravitating, rotating, liquid planet is determined by a
competition between fluid pressure, gravity, and the fictitious centrifugal force. The latter force opposes gravity in the
plane perpendicular to the axis of rotation. Of course, in the absence of rotation, the planet would be spherical. Thus,
we would expect rotation to cause the planet to expand in the plane perpendicular to the rotation axis, and to contract
along the rotation axis (in order to conserve volume).
For the sake of simplicity, we shall restrict our investigation to a rotating planet of uniform density whose outer
boundary is ellipsoidal. Now, an ellipsoid is the three-dimensional generalization of an ellipse. Let us adopt the righthanded Cartesian coordinate system x1 , x2 , x3 . An ellipse whose principal axes are aligned along the x1 - and x2 -axes
satisfies
x12 x22
+
= 1,
(3.94)
a12 a22
where a1 and a2 are the corresponding principal radii. Moreover, as is easily demonstrated,
Z
A =
dA = a1 a2 ,
Z
1 2
xi2 dA =
a A,
4 i
Z
x1 x2 dA = 0,

(3.95)
(3.96)
(3.97)

where A is the area, dA an element of A, and the integrals are taken over the whole interior of the ellipse. Likewise, an
ellipsoid whose principal axes are aligned along the x1 -, x2 -, and x3 -axes satisfies
x12
a12
1 See

x22
a22

x32
a32

= 1,

Ellipsoidal Figures of Equilibrium, S. Chandrasekhar (Yale University Press, New Haven CT, 1969).

(3.98)

Hydrostatics

45

where a1 , a2 , and a3 are the corresponding principal radii. Moreover, as is easily demonstrated,
Z
4
V =
dV = a1 a2 a3 ,
3
Z
1 2
a V,
xi2 dV =
5 i
Z
Z
x1 x2 dV =
x2 x3 dV = 0,

(3.99)
(3.100)
(3.101)

where V is the volume, dV an element of V, and the integrals are taken over the whole interior of the ellipsoid.
Suppose that the planet is rotating uniformly about the x3 -axis at the fixed angular velocity . The planets moment
of inertia about this axis is [cf., Equation (3.100)]
I33 =

1
M (a12 + a22 ),
5

(3.102)

where M is its mass. Thus, the planets angular momentum is


L = I33 =

1
M (a12 + a22 ) ,
5

(3.103)

and its rotational kinetic energy becomes


K=

1
1
I33 2 =
M (a12 + a22 ) 2 .
2
10

(3.104)

According to Equations (3.83) and (3.84), the fluid pressure distribution within the planet takes the form
#
"
1
(3.105)
p = p0 2 (x12 + x22 ) ,
2
where is the gravitational potential (i.e., the gravitational potential energy of a unit test mass) due to the planet,
= M/V the uniform planetary mass density, and p0 a constant. However, it is demonstrated in Appendix E that the
gravitational potential inside a homogeneous self-gravitating ellipsoidal body can be written

3
= G M 0
i xi2 ,
(3.106)
4
i=1,3

where G is the gravitational constant, and


0

du
,

(3.107)

du
,
(ai2 + u)

(3.108)

=
Thus, we obtain
1
p = p0
2

"

(a12 + u)1/2 (a22 + u)1/2 (a32 + u)1/2 .

!
!
#
3
3
3
2
2
2
2
2
G M 1 x1 + G M 2 x2 + G M 3 x3 ,
2
2
2

(3.109)

(3.110)

where p0 is the central fluid pressure. Now, the pressure at the planets outer boundary must be zero, otherwise there
would be a force imbalance across the boundary. In other words, we require
!
!
#
"
3
3
3
1
(3.111)

G M 1 2 x12 + G M 2 2 x22 + G M 3 x32 = p0 ,


2
2
2
2

46

FLUID MECHANICS

whenever

x12
a12

x22
a22

x32
a32

= 1.

(3.112)

The previous two equations can only be simultaneously satisfied if


!
!
2
2
2
a = 2
a 2 = 3 a32 .
1
(3/2) G M 1
(3/2) G M 2

(3.113)

Rearranging the above expression, we obtain


2
a1 a3 2
(a2 a32 )
=
2 G
a2
subject to the constraint
(a12

a22 )

(a22

u du
,
+ u) (a32 + u)

du
= 0,
2

2
2
(a1 + u) (a2 + u) (a3 + u)
a12 a22

a32

(3.114)

(3.115)

where use has been made of Equation (3.99).


Finally, according to Section E, the net gravitational potential energy of the planet is
U=

3
G M 2 0 .
10

(3.116)

Hence, the bodys total mechanical energy becomes


E = K+U =

3
1
M (a12 + a22 ) 2
G M 2 0 .
10
10

(3.117)

3.12 Maclaurin Spheroids


One, fairly obvious, way in which the constraint (3.115) can be satisfied is if a2 = a1 . In other words, if the planet
is rotationally symmetric about its axis of rotation. Now, an ellipsoid that is rotationally symmetric about a principal
axisor, equivalently, an ellipsoid with two equal principal radiiis known as a spheroid. In fact, if a2 = a1 then
the cross-section of the planets outer boundary in any plane passing though the x3 -axis is an ellipse of major radius
a1 , in the direction perpendicular to the x3 -axis, and minor radius a3 , in the direction parallel to the x3 -axis. Here,
we are assuming that a1 > a3 : i.e., that the planet is flattened along its axis of rotation. The degree of flattening is
conveniently measured by the eccentricity,
e13 (1 a32 /a12 )1/2 .
(3.118)
Thus, if e13 = 0 then there is no flattening, and the planet is consequently spherical, whereas if e13 1 then the
flattening is complete, and the planet consequently collapses to a disk in the x1 -x2 plane.
2
Let u = a12 and = e13
/z2 1. Setting a2 = a1 in Equation (3.114), we obtain
2
2 G

(1

2 1/2
2 (1 e13
)
3
e13

d
2 (1 + e 2 )3/2
(1
+
)
0
13
#
"Z e13
Z e13
2
z2 dz
z dz
2
.
(1 e13 )
(1 z2 )1/2
(1 z2 )3/2
0
0

2 1/2 2
e13
) e13

Performing the integrals, which are standard,2 we find that


2
3 2 e13

3
2
2
2 1/2
1

=
(1 e13 ) sin e13 2 (1 e13 ).
3
2 G
e13
e13
2 See

(3.119)

(3.120)

Schaums Mathematical Handbook of Formulas and Tables, 2nd Edition, Murray R. Spiegel, (Mc-Graw Hill, New York NY, 1998).

Hydrostatics

47
e13
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
0.45
0.50
0.55

0.00000
0.02582
0.05168
0.07758
0.10357
0.12967
0.15591
0.18231
0.20889
0.23567
0.26267
0.28989

b
L

0.00000
0.01266
0.02540
0.03830
0.05144
0.06491
0.07882
0.09329
0.10846
0.12450
0.14163
0.16013

b
E

0.60000
0.59980
0.59919
0.59817
0.59672
0.59479
0.59236
0.58936
0.58572
0.58135
0.57612
0.56986

e13
0.60
0.65
0.70
0.75
0.80
0.85
0.90
0.95
0.96
0.97
0.98
0.99

0.31729
0.34484
0.37239
0.39967
0.42612
0.45046
0.46932
0.47045
0.46472
0.45418
0.43475
0.39389

b
L

0.18037
0.20286
0.22834
0.25792
0.29345
0.33833
0.39994
0.50074
0.53194
0.57123
0.62486
0.71209

b
E

0.56233
0.55320
0.54200
0.52800
0.51001
0.48587
0.45107
0.39272
0.37485
0.35273
0.32351
0.27916

Table 3.1: Properties of the Maclaurin spheroids.


This famous result was first obtained by Colin Maclaurin in 1742. Finally, in order to calculate the potential energy,
(3.116), we need to evaluate
Z
Z e13
dz
d
2
2 sin1 e13
1
.
(3.121)
=
=
0 =
2 )1/2
a1 0 (1 + ) (1 + e13
a1 e13 0 (1 z2 )1/2 a1
e13
Let e13 = sin . Thus, = 0 corresponds to no rotational flattening, and = /2 to complete flattening. Moreover,
a1 = a0 (cos )1/3 and a3 = a0 (cos )2/3 , where a0 = (a1 a2 a3 )1/3 = (3 V/4)1/3 is the mean radius. It is also helpful
b = E/(G M 2 /a0 ). The above analysis leads to the following
to define b
= /(2 G )1/2, b
L = L/(G M 3 a0 )1/2 , and E
set of equations which specify the properties of the so-called Maclaurin spheroids:
#
"
cos

2
2 =
b

3
cos

,
(3.122)
(1
+
2
cos
)
sin
sin2

"
#1/2

6 (cos )1/6
b
(1 + 2 cos2 )
3 cos
L =
,
(3.123)
5
sin
sin
#
"
1/3

2
b = 3 (cos )
E
+
3
cos

.
(3.124)
(1

4
cos
)
10 sin2
sin

These properties are set out in Table 3.1.


In the limit, 0, in which the planet is relatively slowly rotating (i.e., b
1), and its degree of flattening
consequently slight, Equations (3.122)(3.124) reduce to

15
e13
,
b
(3.125)
2

6
b
L
,
b
(3.126)
5
b 3.
E
(3.127)
5
In other words, in the limit of relatively slow rotation, when the planet is almost spherical, its eccentricity becomes
directly proportional to its angular velocity. In this case, it is more conventional to parameterize angular velocity in
terms of
2 a0 3 2
= b
,
(3.128)
m=
g0
2

48

FLUID MECHANICS

0.2

2
0.1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


e13

Figure 3.5: Normalized angular velocity squared of a Maclaurin spheroid (solid) and a Jacobi ellipsoid (dashed)
versus the eccentricity e13 in the x1 -x3 plane.
where g0 = G M 2 /a02 is the mean surface gravitational acceleration. Furthermore, the degree of rotational flattening is
more conveniently expressed in terms of the ellipticity,
=

2
a1 a3 e13

.
a0
2

(3.129)

5
m.
4

(3.130)

Thus, it follows from (3.125) that

Now, for the case of the Earth (b


= 7.27 105 rad. s1 , a0 = 6.37 106 m, g0 = 9.81 m s1), we obtain
m

1
.
291

(3.131)

Thus, it follows that, were the Earth homogeneous, its figure would be a spheroid, flattened at the poles, of ellipticity

5 1
1

.
4 291 233

(3.132)

This result was first obtained by Newton. Now, the actual ellipticity of the Earth is about 1/294, which is substantially
smaller than Newtons prediction. The discrepancy is due to the fact that the Earth is strongly inhomogeneous, being
much denser at its core than in its outer regions.
Figures 3.5 and 3.6 illustrate the variation of the normalized angular velocity, b
, and angular momentum, b
L,
of a Maclaurin spheroid with its eccentricity, e13 , as predicted by Equations (3.122)(3.124). It can be seen, from
Figure 3.5, that there is a limit to how large the normalized angular velocity of such a spheroid can become. The

Hydrostatics

49

1.2
1.1
1
0.9
0.8
0.7
0.6
L
0.5
0.4
0.3
0.2
0.1
0

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


e13

Figure 3.6: Normalized angular momentum of a Maclaurin spheroid (solid) and a Jacobi ellipsoid (dashed) versus the
eccentricity e13 in the x1 -x3 plane.
limiting value corresponds to b
= 0.47399, and occurs when e13 = 0.92995. For values of b
lying below 0.47399 there
are two possible Maclaurin spheroids, one with an eccentricity less than 0.92995, and one with an eccentricity greater
than 0.92995. Note, however, from Figure 3.6, that despite the fact that the angular velocity, b
, of a Maclaurin spheroid
varies in a non-monotonic manner with the eccentricity, e13 , the angular momentum, b
L, increases monotonically with
e13 , becoming infinite in the limit e13 1. It follows that there is no upper limit to the angular momentum of a
Maclaurin spheroid.

3.13 Jacobi Ellipsoids


If a2 , a1 (i.e., if the outer boundary of the rotating body is ellipsoidal, rather than spheroidal) then the constraint
(3.115) can only be satisfied when

Z
a 2 du
a2 a2

2 1 2 2
= 0.
(3.133)
2 3
(a1 + u) (a2 + u) (a3 + u)
0
Without loss of generality, we can assume that a1 a2 a3 . Let
a2

a1 cos ,

(3.134)

a3

a1 cos ,

(3.135)

where . It follows that the cross-sections of the planets outer boundary in the x1 -x2 and x1 -x3 planes are ellipses
of eccentricities
e12
e13

=
=

(1 a32 /a22 )1/2 = sin ,

(3.136)

(1

(3.137)

a32 /a12 )1/2

= sin ,

50

FLUID MECHANICS

respectively. It is also helpful to define


= sin1 (sin / sin ).

(3.138)

Let sin2 = [a12 /(a12 + u)] sin2 . Here, u = 0 corresponds to = , and u = to = 0. Equations (3.115) and
(3.114) transform to 3
#
"
1 + (sin tan cos )2
E(, )
E(, ) 2 F(, ) +
cos2
sin tan cos (1 + sin2 )
cos2

and

0,

"

#
F(, ) E(, ) cos E(, )
cos2
b =2

,
+
tan sin tan
tan3 cos2 tan2 cos2
2

respectively, where

E(, ) =

F(, ) =

(3.140)

(3.139)

(1 sin2 sin2 )1/2 d,

(3.141)

d
,
(1 sin sin2 )1/2

(3.142)

are special functions known as incomplete elliptic integrals.4 The integral 0 , defined in (3.107), transforms to
0 =

2 (cos cos )1/3


F(, ).
a0 sin

(3.143)

Finally, making use of some of the analysis in the previous two sections, the normalized angular momentum, and
normalized mechanical energy, of the planet can be written

6 1 + cos2
b
,
b
(3.144)
L = =
10 (cos cos )2/3
b =
E

3 (cos cos )1/3


1 + cos2
3
b 2,

F(, ) +
5
sin
20 (cos cos )2/3

(3.145)

respectively.
Now, the constraint (3.139) is obviously satisfied in the limit 0, since this implies that 0 and E(, ),
F(, ) . Of course, this limit corresponds to the axisymmetric Maclaurin spheroids discussed in the previous section. Jacobi, in 1834, was the first researcher to obtain the very surprising result that (3.139) also has non-axisymmetric
ellipsoidal solutions characterized by > 0. These solutions are known as the Jacobi ellipsoids in his honor. The
properties of the Jacobi ellipsoids, as determined from Equations (3.139), (3.140), (3.144), and (3.145), are set out in
Table 3.2, and illustrated in Figures 3.5 and 3.6. It can be seen that the sequence of Jacobi ellipsoids bifurcates from the
sequence of Maclaurin spheroids when e13 = 0.81267. Moreover, there are no Jacobi ellipsoids with e13 < 0.81267.
However, as e13 increases above this critical value, the eccentricity, e12 , of the Jacobi ellipsoids in the x1 -x2 plane
grows rapidly, approaching unity as e13 approaches unity. Thus, in the limit e13 1, in which a Maclaurin spheroid
collapses to a disk in the x1 -x2 plane, a Jacobi ellipsoid collapses to a line running along the x1 -axis. Note, from
Figures 3.5 and 3.6, that, at fixed e13 , the Jacobi ellipsoids have lower angular velocity and angular momentum than
Maclaurin spheroids (with the same mass and volume). Furthermore, as is the case for a Maclaurin spheroid, there is a
maximum angular velocity that a Jacobi ellipsoid can have (i.e., b
= 0.43257), but no maximum angular momentum.
Figure 3.7 shows the mechanical energy of the Maclaurin spheroids and Jacobi ellipsoids plotted as a function of
their angular momentum. It can be seen that the Jacobi ellipsoid with a given angular momentum has a lower energy
3 See
4 See

On Jacobis Figure of Equilibrium for a Rotating Mass of Fluid, G.H. Darwin, Proc. Roy. Soc. London 41, 319 (1886).
Handbook of Mathematical Functions, M. Abramowitz, and I.A. Stegun (Dover, New York NY, 1965).

Hydrostatics

51

e12

e13

0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
0.45
0.50
0.55

0.81267
0.81293
0.81372
0.81504
0.81691
0.81934
0.82237
0.82603
0.83037
0.83544
0.84131
0.84808

0.43257
0.43257
0.43257
0.43256
0.43253
0.43248
0.43237
0.43220
0.43191
0.43146
0.43078
0.42976

b
L

0.30375
0.30375
0.30375
0.30377
0.30380
0.30388
0.30402
0.30427
0.30468
0.30532
0.30628
0.30772

b
E

0.50452
0.50459
0.50459
0.50458
0.50457
0.50453
0.50445
0.50432
0.50410
0.50376
0.50326
0.50250

e12

e13

0.60
0.65
0.70
0.75
0.80
0.85
0.90
0.95
0.96
0.97
0.98
0.99

0.85585
0.86480
0.87510
0.88705
0.90102
0.91761
0.93778
0.96340
0.96950
0.97605
0.98317
0.99101

0.42827
0.42609
0.42288
0.41807
0.41069
0.39879
0.37787
0.33353
0.31776
0.29691
0.26722
0.21809

b
L

0.30984
0.31296
0.31760
0.32462
0.33562
0.35390
0.38783
0.46860
0.50078
0.54672
0.62003
0.76872

b
E

0.50138
0.49975
0.49734
0.49372
0.48814
0.47908
0.46295
0.42782
0.41499
0.39771
0.37241
0.32842

Table 3.2: Properties of the Jacobi ellipsoids.


that the corresponding Maclaurin spheroid (i.e., the spheroid with the same angular momentum, mass, and volume).
This is significant because, in the presence of a small amount of dissipation (i.e., viscosity), we would generally expect
an isolated fluid system to slowly evolve toward the equilibrium state with the lowest energy, subject to any global
constraints on the system. For the case of a weakly viscous, isolated, rotating, liquid planet, the relevant constraints are
that the mass, volume, and net angular momentum of the system cannot spontaneously change. Thus, we expect such a
planet to evolve toward the equilibrium state with the lowest energy for a given mass, volume, and angular momentum.
This suggests, from Figure 3.7, that at relatively high angular momentum (i.e., b
L > 0.30375, e13 > 0.81267), when
the Jacobi ellipsoid solutions exist, they are stable equilibrium states (since there is no lower energy state to which
the system can evolve), whereas the Maclaurin spheroids are unstable. On the other hand, at relatively low angular
momentum (i.e., b
L < 0.30375, e13 < 0.81267), when there are no Jacobi ellipsoid solutions, the Maclaurin spheroids
are stable equilibrium states (again, because there is no lower energy state to which they can evolve). These predictions
are borne by the results of direct stability analysis performed on the Maclaurin spheroids and Jacobi ellipsoids.5 In
fact, such stability studies demonstrate that the Maclaurin spheroids are unstable in the presence of weak dissipation
for e13 > 0.81267, and unconditionally unstable for e13 > 0.95289. The Jacobi ellipsoids, on the other hand, are
unconditionally stable for e13 < 0.93858, but are unconditionally unstable for e13 > 0.93858, evolving toward lower
energy pear shaped equilibria (which are, themselves, unstable in the presence of weak dissipation).

3.14 Roche Ellipsoids


Consider a homogeneous liquid moon of mass M which is in a circular orbit of radius R about a planet of mass
M . Let C, C , and C be the center of the moon, the center of the planet, and the center of mass of the moon-planet
system, respectively. As is easily demonstrated, all three points lie on the same straight-line, and the distances between
them take the constant values CC = R and CC = [M /(M + M )] R. Moreover, according to standard Newtonian
dynamics, there exists an inertial frame of reference in which C is stationary, and the line CC rotates at the fixed
angular velocity , where
G (M + M )
.
(3.146)
2 =
R3
In other words, in the inertial frame, the moon and the planet orbit in a fixed plane about their common center of mass
at the angular velocity . It is convenient to transform to a non-inertial reference frame that rotates (with respect to the
inertial frame), about an axis passing through C , at the angular velocity . It follows that points C, C , and C appear
stationary in this frame. It is also convenient to adopt the standard right-handed Cartesian coordinates, x1 , x2 , x3 , and
to choose the coordinate axes such that = e3 , C = (0, 0, 0), C = (R, 0, 0), and C = ([M /(M + M )] R, 0, 0).
5 See

Ellipsoidal Figures of Equilibrium, S. Chandrasekhar (Yale University Press, New Haven CT, 1969).

52

FLUID MECHANICS

0.6
0.5

0.4
0.3
0.2
0.1
0

0.2

0.4

0.6

0.8

1.2

Figure 3.7: Normalized mechanical energy of a Maclaurin spheroid (solid) and a Jacobi ellipsoid (dashed) versus the
normalized angular momentum.
Thus, in the non-inertial reference frame, the orbital rotation axis runs parallel to the x3 -axis, and the centers of the
moon and the planet both lie on the x1 -axis.
Suppose that the moon does not rotate (about an axis passing through its center of mass) in the non-inertial reference frame. This implies that, in the inertial frame, the moon appears to rotate about an axis parallel to the x3 -axis, and
passing through C, at the same angular velocity as it orbits about C . This type of rotation is termed synchronous, and
ensures that the same hemisphere of the moon is always directed toward the planet. Such rotation is fairly common in
the Solar System. For instance, the Moon rotates synchronously in such a manner that the same hemisphere is always
visible from the Earth. Synchronous rotation in the Solar System is a consequence of process known as tidal locking.
Since a synchronously rotating moon is completely stationary in the aforementioned non-inertial frame, its internal
pressure, p, is governed by a force balance equation of the form [cf., Equation (3.83)]
0 = p + ( + + ),

(3.147)

where is the uniform internal mass density, the gravitational potential due to the moon, the gravitational
potential due to the planet, and

!2

1 2
M
2

R + x2
= x1
(3.148)

2
M+M
the centrifugal potential due to the fact that the non-inertial frame is rotating (about an axis parallel to the x3 -axis and
passing through point C ) at the angular velocity [cf., Equation (3.84)]. Suppose that the moon is much less massive
that the planet (i.e., M/M 1). In this limit, the centrifugal potential (3.148) reduces to

1 x1 (1/2) x12 + (1/2) x22


,
+

2 R
R2

(3.149)

Hydrostatics

53

where use has been made of (3.146).


Suppose that the planet is spherical. It follows that the potential is the same as that which would be generated
by a point mass M located at C . In other words,

1/2
x 2 + x22 + x32
G M
x
1 2 1 + 1

R
R
R2

x1 x12 (1/2) x22 (1/2) x32


G M

+ ,
+

1 +
R
R
R2

(3.150)

where we have expanded up to second order in x1 /R, etc.


The previous two equations can be combined to give

!
3 2 1 2
x x ,
+
2 1 2 3

(3.151)

where

G M
,
(3.152)
R3
and any constant terms have been neglected. Thus, the net force field experienced by the moon due to the combined
action of the fictitious centrifugal force and the gravitational force field of the planet is
=

( + ) = (3 x1 , 0, x3 ).

(3.153)

The above type of force field is known as a tidal force field, and clearly acts to elongate the moon along the axis
joining the centers of the moon and planet (i.e., the x1 -axis), and to compress it along the orbital rotation axis (i.e., the
x3 -axis). Moreover, the magnitude of the tidal force increases linearly with distance from the center of the moon. The
tidal force field is a consequence of the different spatial variation of the centrifugal force and the planets gravitational
force of attraction. This different variation causes these two forces, which balance one another at the center of the
moon, to not balance away from the center. As a result of the tidal force field, we expect the shape of the moon to
be distorted from a sphere. Of course, the moon also generates a tidal force field that acts to distort the shape of the
planet. However, we are assuming that the tidal distortion of the planet is much smaller than that of the moon (which
justifies our earlier statement that the planet is essentially spherical). As will be demonstrated later, this assumption
is reasonable provided the mass of the moon is much less than that of the planet (assuming that the planet and moon
have similar densities).
Suppose that the bounding surface of the moon is the ellipsoid
x12
a12

x22
a22

x32
a32

= 1,

(3.154)

where a1 a2 a3 . It follows, from Section E, that the gravitational potential of the moon at an interior point can be
written

3
2

= G M 0
i xi ,
(3.155)
4
i=1,3

where the integrals i , for i = 0, 3, are defined in Equations (E.30) and (E.31). Hence, from (3.147) and (3.151), the
pressure distribution within the moon is given by
"
!
! #
3
3
3
1
(3.156)
G M 1 3 x12 + G M 2 x22 + G M 3 + x32 ,
p = p0
2
2
2
2
where p0 is the central pressure. Now, the pressure must be zero on the moons bounding surface, otherwise this
surface would not be in equilibrium. Thus, in order to achieve equilibrium, we require
"
!
! #
1
3
3
3
(3.157)

G M 1 3 x12 + G M 2 x22 + G M 3 + x32 = p0 ,


2
2
2
2

54

FLUID MECHANICS

0.05

e13 e12

0.04

0.03

0.02

0.01

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


e13
Figure 3.8: Properties of the Roche ellipsoids.

whenever

x12

x22

x32

= 1.
a12 a22 a32
The previous two equations can only be simultaneously satisfied if
"
#
"
#
3

1
a12 = 2 a22 = 3 +
a 2.
(3/2) G M
(3/2) G M 3

(3.158)

(3.159)

Let a2 = a1 cos and a3 = a1 cos , where . It is also helpful to define = sin1 (sin / sin ). With the help
of some of the analysis presented in the previous section, the integrals i , for i = 1, 3, can be shown to take the form
1

F(, ) E(, )
2
,
3
sin
sin2
"
#
E(, )
2
F(, )
cos sin
,

cos2 cos
a12 sin3 sin2 cos2
sin2
"
#
2
E(, )
cos sin

,
+
cos2
cos2 cos
a12 sin3
a13

(3.160)
(3.161)
(3.162)

where the incomplete elliptic integrals E(, ) and F(, ) are defined in Equations (3.141) and (3.142), respectively.
Thus, (3.159) yields
"
!
1
cos2
2
=
F(, ) (1 + cos ) E(, ) 1 +
sin tan tan
cos2
#
sin sin cos cos
,
(3.163)
+
cos2

Hydrostatics

55
e12

e13

e12

e13

0.00
0.04
0.08
0.12
0.16
0.20
0.24
0.28
0.32
0.36
0.40
0.44
0.48

0.00000
0.04613
0.09223
0.13809
0.18364
0.22879
0.27346
0.31756
0.36104
0.40383
0.44588
0.48718
0.52769

0.00000
0.00213
0.00852
0.01913
0.03392
0.05282
0.07573
0.10253
0.13308
0.16721
0.20470
0.24528
0.28865

0.52
0.56
0.60
0.64
0.68
0.72
0.76
0.80
0.84
0.88
0.92
0.96
1.00

0.56740
0.60632
0.64445
0.68182
0.71848
0.75446
0.78984
0.82472
0.85923
0.89353
0.92793
0.96294
1.00000

0.33440
0.38204
0.43094
0.48027
0.52890
0.57532
0.61729
0.65150
0.67265
0.67151
0.62978
0.50135
0.00000

Table 3.3: Properties of the Roche ellipsoids.


subject to the constraint
0 =

#
"
cos2 sin sin cos cos

cos2 F(, ) 2 E(, ) + E(, )


cos2
cos2
#
"
cos2 2 sin sin cos cos
2
2
,
+
+(3 + cos ) E(, ) + [F(, ) cos 2 E(, )]
cos2
cos2

where
=

M a03
,
M R3

(3.164)

(3.165)

and a0 = (a1 a2 a3 )1/3 is the mean radius of the moon. The dimensionless parameter measures the strength of the
tidal distortion field, generated by the planet, that acts on the moon. There is an analogous parameter,
3
M a0
= 3,
M R

(3.166)

where a0 is the mean radius of the planet, which measures the tidal distortion field, generated by the moon, that acts
on the planet. Now, we previously assumed that the former distortion field is much stronger than the latter, allowing
us to neglect the tidal distortion of the planet altogether, and so to treat it as a sphere. This assumption is only justified
if , which implies that
M

,
(3.167)
M

where = M/[(4/3) a03] and = M /[(4/3) a0 3 ] are the mean densities of the moon and the planet, respectively.
Assuming that these densities are similar, the above condition reduces to M M , or, equivalently, a0 < a0 . In other
words, neglecting the tidal distortion of the planet, whilst retaining that of the moon, is generally only reasonable when
the mass of the moon is much less than that of the planet, as was previously assumed to be the case.
Equations (3.163) and (3.164), which describe the ellipsoidal equilibria of a synchronously rotating, relatively low
mass, liquid moon due to the tidal force field of the planet about which it orbits, were first obtained by Roche in 1850.
The properties of the so-called Roche ellipsoids are set out in Table 3.3, and Figures 3.8 and 3.9.
It can be seen, from Table 3.3 and Figure 3.8, that the eccentricity e12 = sin of a Roche ellipsoid in the x1 -x2
plane is almost equal to its eccentricity e13 = sin in the x1 -x3 plane. In other words, Roche ellipsoids are almost

56

FLUID MECHANICS

0.07
0.06
0.05
0.04
0.03
0.02
0.01
0

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


e13
Figure 3.9: Properties of the Roche ellipsoids.

spheroidal in shape, being elongated along the x1 -axis (i.e., the axis joining the centers of the moon and the planet),
and compressed by almost equal amounts along the x2 - and x3 -axes. In the limit 1, in which the tidal distortion
field due to the planet is weak, it is easily shown that
2
2
e12
e13

15
.
2

(3.168)

For the case of the tidal distortion field generated by the Earth, and acting on the Moon, which is characterized
by M/M = 0.01230 and R/a0 = 221.29, we obtain = 7.50 106 . It follows that e13 = 7.50 103 , and
2
(a1 a3 )/a1 e13
/2 = 2.81 105 . In other words, were the Moon a homogeneous liquid body, the elongation
generated by the tidal field of the Earth would be about 50 m.
It can be seen, from Table 3.3 and Figure 3.9, that the parameter attains a maximum value as the eccentricity
of a Roche ellipsoid varies from 0 to 1. In fact, this maximum value, = 0.06757, occurs when e12 = 0.8594 and
e13 = 0.8759. It follows that there is a maximum strength of the tidal distortion field, generated by a planet, that is
consistent with an ellipsoidal equilibrium of a synchronously rotating, homogeneous, liquid moon in a circular orbit
about the planet. It is plausible that when this maximum strength is exceeded the moon is tidally disrupted by the
planet. The equilibrium condition < 0.06757 is equivalent to
R

>
2.455
a0

!1/3

(3.169)

where = M/[(4/3) a03] and = M /[(4/3) a0 3 ] are the mean densities of the moon and the planet, respectively.
According to the above expression, there is a minimum orbital radius of a moon circling a planet. Below this radius,
which is called the Roche radius, the moon is presumably torn apart by tidal effects. The Roche radius for a synchronously rotating, self-gravitating, liquid moon in a circular orbit about a spherical planet is about 2.5 times the

Hydrostatics

57

planets radius (assuming that the moon and the planet are about the same density). Of course, relatively small objects,
such as artificial satellites, which are held together by internal tensile strength, rather than gravity, can orbit inside the
Roche radius without being disrupted.

3.15 Exercises
3.1. A hollow vessel floats in a basin. If, as a consequence of a leak, water flows slowly into the vessel, how will the level of the
water in the basin be affected?
3.2. A hollow spherical shell made up of material of specific gravity s > 1 has external and internal radii a and b, respectively.
Demonstrate that the sphere will only float in water if
!1/3
1
b
> 1
.
a
s
3.3. Show that the equilibrium of a solid of uniform density floating with an edge or corner just emerging from the water is
unstable.
3.4. Prove that if a solid of uniform density floats with a flat face just above the waterline then the equilibrium is stable.
3.5. Demonstrate that a uniform solid cylinder floating with its axis horizontal is in a stable equilibrium provided that its length
exceeds the breadth of the waterline section. [Hint: The cylinder is obviously neutrally stable to rotations about its axis,
which means that the corresponding metacentric height is zero.]

3.6. Show that a uniform solid cylinder of radius a and height h can float in stable equilibrium, with its axis vertical, if h/a < 2.
If the ratio h/a exceeds this value, prove that the equilibrium is only stable when the specific gravity of the cylinder lies
outside the range
r

1
a2
1 1 2 2 .
2
h

3.7. A uniform, thin, hollow cylinder of radius a and height h is open at both ends. Assuming that h > 2 a, prove that the cylinder
cannot float upright if its specific gravity lies in the range
r
1 a2
1

.
2
4 h2
3.8. Show that the cylinder of the preceding exercise can float with its axis horizontal provided

h
> 3 sin(s ),
2a
where s is the specific gravity of the cylinder.

3.9. Prove that any segment of a uniform sphere, made up of a substance lighter than water, can float in stable equilibrium with
its plane surface horizontal and immersed.
3.10. A vessel carries a tank of oil, of specific gravity s, running along its length. Assuming that the surface of the oil is at sea
level, show that the effect of the oils fluidity on the rolling of the vessel is equivalent to a reduction in the metacentric height
by A 2 s/V, where V is the displacement of the ship, A the surface-area of the tank, and the radius of gyration of this area.
In what ratio is the effect diminished when a longitudinal partition bisects the tank?
3.11. Find the stable equilibrium configurations of a cylinder of elliptic cross-section, with major and minor radii a and b < a,
respectively, made up of material of specific gravity s, which floats with its axis horizontal.
3.12. A cylindrical tank has a circular cross-section of radius a. Let the center of gravity of the tank be located a distance c above
its base. Suppose that the tank is pivoted about a horizontal axis passing through its center of gravity, and is then filled with
fluid up to a depth h above its base. Demonstrate that the position in which the tanks axis is upright is unstable for all filling
depths provided
1
c2 < a2 .
2
Show that if c2 > (1/2) a2 then the upright position is stable when h lies in the range
p
c c2 a2 /2.

58

FLUID MECHANICS

3.13. A thin cylindrical vessel of cross-sectional area A floats upright, being immersed to a depth h, and contains water to a depth
k. Show that the work required to pump out the water is 0 A k (h k) g.
3.14. A sphere of radius a is just immersed in water that is contained in a cylindrical vessel of radius R whose axis is vertical.
Prove that if the sphere is raised just clear of the water then the waters loss of potential energy is
!
2 a2
Wa 1
,
3 R2
where W is the weight of the water originally displaced by the sphere.
3.15. A sphere of radius a, weight W, and specific gravity s > 1, rests on the bottom of a cylindrical vessel of radius R whose axis
is vertical, and which contains water to a depth h > 2 a. Show that the work required to lift the sphere out of the vessel is
less than if the water had been absent by an amount
!
2 a3 W
ha
.
3 R2 s
3.16. A lead weight is immersed in water that is steadily rotating at an angular velocity about a vertical axis, the weight being
suspended from a fixed point on this axis by a string of length l. Prove that the position in which the weight hangs vertically
downward is stable or unstable depending on whether l < g/2 or l > g/2 , respectively. Also, show that if the vertical
position is unstable then there exists a stable inclined position in which the string is normal to the surface of equal pressure
passing though the weight.
3.17. A thin cylindrical vessel of radius a and height H is orientated such that its axis is vertical. Suppose that the vessel is filled
with liquid of density to some height h < H above the base, spun about its axis at a steady angular velocity , and the
liquid allowed to attain a steady state. Demonstrate that, provided 2 a2 /g < 4 h and 2 a2 /g < 4 (H h), the net radial
thrust on the vertical walls of the vessel is
!2
2 a2
a h2 g 1 +
.
4gh
3.18. A thin cylindrical vessel of radius a with a plane horizontal lid is just filled with liquid of density , and the whole rotated
about a vertical axis at a fixed angular velocity . Prove that the net upward thrust of the fluid on the lid is
1 4
a 2 .
4
3.19. A liquid-filled thin spherical vessel of radius a spins about a vertical diameter at the fixed angular velocity . Assuming that
the liquid co-rotates with the vessel, and that 2 > g/a, show that the pressure on the wall of the vessel is greatest a depth
g/2 below the center. Also prove that the net normal thrusts on the lower and upper hemispheres are
5
3
Mg+
M 2 a,
4
16
and

3
1
Mg
M 2 a,
4
16

respectively, where M is the mass of the liquid.


3.20. A closed cubic vessel filled with water is rotating about a vertical axis passing through the centers of two opposite sides.
Demonstrate that, as a consequence of the rotation, the net thrust on a side is increased by
1 4
a 2 ,
6
where a is the length of an edge of the cube, and the angular velocity of rotation.
3.21. A closed vessel filled with water is rotating at constant angular velocity about a horizontal axis. Show that, in the state of
relative equilibrium, the constant pressure surfaces in the water are circular cylinders whose common axis is a height g/2
above the axis of rotation.
3.22. Consider a homogeneous, rotating, liquid body of mass M, mean radius a0 , and angular velocity , whose outer boundary
is a Maclaurin spheroid of eccentricity e.

Hydrostatics

59

(a) Demonstrate that


e

2 (G M/a03 )1/2

in the low rotation limit, (G M/a03 )1/2 . Hence, show that e = 0.09262 for the case of a homogeneous body with
the same mass and volume as the Earth, which rotates once every 24 hours.
(b) Show that the critical angular velocity at which the bifurcation to the sequence of Jacobi ellipsoids takes place is
= 0.5298 (G M/a03 )1/2 ,
and occurs when e = 0.81267. Hence, show that, for the case of a homogeneous body with the same mass and volume
as the Earth, the bifurcation would take place at a critical rotation period of 2 h 39 m.
(c) Demonstrate that the maximum angular velocity consistent with a spheroidal shape is
= 0.5805 (G M/a03 )1/2 ,
and occurs when e = 0.92995. Hence, show that, for the case of a homogeneous body with the same mass and volume
as the Earth, this maximum velocity corresponds to a minimum rotation period of 2 h 25 m.

60

FLUID MECHANICS

Surface Tension

61

4 Surface Tension

4.1 Introduction
As is well-known, small drops of water in air, and small bubbles of gas in water, tend to adopt spherical shapes. This
phenomenon, and a host of other natural phenomena, can only be accounted for on the hypothesis that an interface
between two different media is associated with a particular form of energy whose magnitude is directly proportional to
the interfacial area. To be more exact, if S is the interfacial area then the contribution of the interface to the Helmholtz
free energy of the system takes the form S , where only depends on the temperature and chemical composition of
the two media on either side of the interface. It follows, from standard thermodynamics, that S is the work that must
be performed on the system in order to create the interface via an isothermal and reversible process. However, this
work is exactly the same as that which we would calculate on the assumption that the interface is in a state of uniform
constant tension per unit length . Thus, can be interpreted as both a free energy per unit area of the interface, and
a surface tension. This tension is such that a force of magnitude per unit length is exerted across any line drawn on
the interface, in a direction normal to the line, and tangential to the interface.
Surface tension originates from intermolecular cohesive forces. The average free energy of a molecule in a given
isotropic medium possessing an interface with a second medium is independent of its position, provided that the
molecule does not lie too close to the interface. However, the free energy is modified when the molecules distance
from the interface becomes less than the range of the cohesive forces (which is typically 109 m). Since this range is
so small, the number of molecules in a macroscopic system whose free energies are affected by the presence of an
interface is directly proportional to the interfacial area. Hence, the contribution of the interface to the total free energy
of the system is also proportional to the interfacial area. If only one of the two media in question is a condensed phase
then the parameter is invariably positive (i.e., such that a reduction in the surface area is energetically favorable).
This follows because the molecules of a liquid or a solid are subject to an attractive force from neighboring molecules.
However, molecules that are near to an interface with a gas lack neighbors on one side, and so experience an unbalanced
cohesive force directed toward the interior of the liquid/solid. The existence of this force makes it energetically
favorable for the interface to contract (i.e., > 0). On the other hand, if the interface separates a liquid and a solid,
or a liquid and another liquid, then the sign of cannot be predicted by this argument. In fact, it is possible for both
signs of to occur at liquid/solid and liquid/liquid interfaces.
The surface tension of a water/air interface at 20 C is = 7.28 102 N m1 . The surface tension at most oil/air
interfaces is much lowertypically, 2 102 N m1 . On the other hand, interfaces between liquid metals and
air generally have very large surface tensions. For instance, the surface tension of a mercury/air interface at 20 C is
4.87 101 N m1 .
For some pairs of liquids, such as water and alcohol, an interface cannot generally be observed because it is in
compression (i.e., < 0). Such an interface tends to become as large as possible, leading to complete mixing of the
two liquids. In other words, liquids for which > 0 are immiscible, whereas those for which < 0 are miscible.
Finally, the surface tension at a liquid/gas or a liquid/liquid interface can be affected by the presence of adsorbed
impurities at the interface. For instance, the surface tension at a water/air interface is significantly deceased in the
presence of adsorbed soap molecules. Impurities that tend to reduce surface tension at interfaces are termed surfactants.

4.2 Young-Laplace Equation


Consider an interface separating two immiscible fluids that are in equilibrium with one another. Let these two fluids
be denoted 1 and 2. Consider an arbitrary segment S of this interface that is enclosed by some closed curve C. Let
t denote a unit tangent to the curve, and let n denote a unit normal to the interface directed from fluid 1 to fluid 2.
(Note that C circulates around n in a right-handed manner.) See Figure 4.1. Suppose that p1 and p2 are the pressures
of fluids 1 and 2, respectively, on either side of S . Finally, let be the (uniform) surface tension at the interface.

62

FLUID MECHANICS

n
2

S
t

tn
Figure 4.1: Interface between two immiscible fluids.

The net force acting on S is


f=

(p1 p2 ) n dS +

t n dr,

(4.1)

where dS = n dS is an element of S , and dr = t dr an element of C. Here, the first term on the right-hand side is the
net normal force due to the pressure difference across the interface, whereas the second term is the net surface tension
force. Note that body forces play no role in (4.1), because the interface has zero volume. Furthermore, viscous forces
can be neglected, since both fluids are static. Now, in equilibrium, the net force acting on S must be zero: i.e.,
I
Z
t n dr.
(4.2)
(p1 p2 ) n dS =
C

(In fact, the net force would be zero even in the absence of equilibrium, because the interface has zero mass.)
Applying Stokes theorem (see Section A.22) to the curve C, we find that
Z
I
F dS,
F dr =

(4.3)

where F is a general vector field. This theorem can also be written


Z
I
F n dS .
F t dr =

(4.4)

Suppose that F = g b, where b is an arbitrary constant vector. We obtain


Z
I
(g b) n dS .
(g b) t dr =

(4.5)

However, the vector identity (A.179) yields


(g b) = ( g) b + (b ) g,
since b is a constant vector. Hence, we get
Z
I
t g dr = b [(g) n ( g) n] dS ,
b
where b (g) n bi (g j/xi ) n j . Now, since b is also an arbitrary vector, the above equation gives
Z
I


(g) n ( g) n dS .
t g dr =
I

(4.8)

(4.7)

Taking g = n, we find that

(4.6)

t n dr =

[(n) n ( n) n] dS .

(4.9)

Surface Tension

63

But, (n) n (1/2) (n2) = 0, because n is a unit vector. Thus, we obtain


Z
I
t n dr = ( n) n dS ,

which can be combined with (4.2) to give


Z

(4.10)


(p1 p2 ) ( n) n dS = 0.

(4.11)

Finally, given that S is arbitrary, the above expression reduces to the pressure balance constraint
p = n,

(4.12)

where p = p1 p2 . The above relation is generally known as the Young-Laplace equation, and can also be derived
by minimizing the free energy of the interface. (See Section 4.8.) Note that p is the jump in pressure seen when
crossing the interface in the opposite direction to n. Of course, a plane interface is characterized by n = 0. On
the other hand, a curved interface generally has n , 0. In fact, n measures the local mean curvature of the
interface. Thus, according to the Young-Laplace equation, there is a pressure jump across a curved interface between
two immiscible fluids, the magnitude of the jump being proportional to the surface tension.

4.3 Spherical Interfaces


Generally speaking, the equilibrium shape of an interface between two immiscible fluids is determined by solving the
force balance equation (3.1) in each fluid, and then applying the Young-Laplace equation to the interface. However,
in situations in which a mass of one fluid is completely immersed in a second fluide.g., a mist droplet in air, or a
gas bubble in waterthe shape of the interface is fairly obvious. Provided that either the size of the droplet or bubble,
or the difference in densities on the two sides of the interface, is sufficiently small, we can safely ignore the effect of
gravity. This implies that the pressure is uniform in each fluid, and consequently that the pressure jump p is constant
over the interface. Hence, from (4.12), the mean curvature n of the interface is also constant. Since a sphere is the
only closed surface which possesses a constant mean curvature, we conclude that the interface is spherical. This result
also follows from the argument that a stable equilibrium state is one which minimizes the free energy of the interface,
subject to the constraint that the enclosed volume be constant. In other words, the equilibrium shape of the interface
is that which has the least surface area for a given volume: i.e., a sphere.
Suppose that the interface corresponds to the spherical surface r = R, where r is a spherical coordinate. (See
Section C.4.) It follows that n = er |r=R . (Note, for future reference, that n points away from the center of curvature of
the interface.) Hence, from (C.65),

1 r2
2
n= 2
(4.13)
= .
r r r=R R
The Young-Laplace equation, (4.12), then gives

2
.
(4.14)
R
Thus, given that p is the pressure jump seen crossing the interface in the opposite direction to n, we conclude that the
pressure inside a droplet or bubble exceeds that outside by an amount proportional to the surface tension, and inversely
proportional to the droplet or bubble radius. This explains why small bubbles are louder that large ones when they
burst at a free surface: e.g., champagne fizzes louder than beer. Note that soap bubbles in air have two interfaces
defining the inner and outer extents of the soap film. Consequently, the net pressure difference is twice that across a
single interface.
p =

4.4 Capillary Length


Consider an interface separating the atmosphere from a liquid of uniform density that is at rest on the surface of the
Earth. Neglecting the density of air compared to that of the liquid, the pressure in the atmosphere can be regarded as

64

FLUID MECHANICS

Figure 4.2: Interface between a liquid (1), a gas (2), and a solid (3).
constant. On the other hand, the pressure in the liquid varies as p = p0 g z (see Chapter 3), where p0 is the pressure
of the atmosphere, g the acceleration due to gravity, and z measures vertical height (relative to the equilibrium height
of the interface in the absence of surface tension). Note that z increases upward. In this situation, the Young-Laplace
equation (4.12) yields
g z = n,
(4.15)
where n is the normal to the interface directed from liquid to air. Now, if R represents the typical radius of curvature
of the interface then the left-hand side of the above equation dominates the right-hand side whenever R l, and vice
versa. Here,
!1/2

(4.16)
l=
g
is known as the capillary length, and takes the value 2.7 103 m for pure water at 20 C. We conclude that the effect
of surface tension on the shape of an liquid/air interface is likely to dominate the effect of gravity when the interfaces
radius of curvature is much less than the capillary length, and vice versa.

4.5 Angle of Contact


Suppose that a liquid/air interface is in contact with a solid, as would be the case for water in a glass tube, or a drop of
mercury resting on a table. Figure 4.2 shows a section perpendicular to the edge at which the liquid, 1, the air, 2, and
the solid, 3, meet. Suppose that the free energies per unit area at the liquid/air, liquid/solid, and air/solid interfaces are
12 , 13 , and 23 , respectively. If the boundary between the three media is slightly modified in the neighborhood of
the edge, as indicated by the dotted line in the figure, then the area of contact of the air with the solid is increased by a
small amount r per unit breadth (perpendicular to the figure), whereas that of the liquid with the solid is decreased by
r per unit breadth, and that of the liquid with the air is decreased by r cos per unit breadth. Thus, the net change
in free energy per unit breadth is
23 r 13 r 12 r cos .
(4.17)
However, an equilibrium state is one which minimizes the free energy, implying that the above expression is zero for
arbitrary (small) r: i.e.,
23 13
cos =
.
(4.18)
12
We conclude that, in equilibrium, the angle of contact, , between the liquid and the solid takes a fixed value that
depends on the free energies per unit area at the liquid/air, liquid/solid, and air/solid interfaces. Note that the above

Surface Tension

65

air
a

liquid

glass tube

air
liquid

free surface

z=0

Figure 4.3: Elevation of liquid level in a capillary tube.


formula could also be obtained from the requirement that the various surface tension forces acting at the edge balance
one another, assuming that it is really appropriate to interpret 13 and 23 as surface tensions when one of the media
making up the interface is a solid.
As explained in Section 4.1, we would generally expect 12 and 23 to be positive. On the other hand, 13 could be
either positive or negative. Now, since | cos | 1, Equation (4.18) can only be solved when 13 lies in the range
23 + 12 > 13 > 23 12 .

(4.19)

If 13 > 23 + 12 then the angle of contact is 180, which corresponds to the case where the free energy at the
liquid/solid interface is so large that the liquid does not wet the solid at all, but instead breaks up into beads on its
surface. On the other hand, if 13 < 23 12 then the angle of contact is 0 , which corresponds to the case where the
free energy at the liquid/solid interface is so small that the liquid completely wets the solid, spreading out indefinitely
until it either covers the whole surface, or its thickness reaches molecular dimensions.
The angle of contact between water and glass typically lies in the range 25 to 29 , whereas that between mercury
and glass is about 127.

4.6 Jurins Law


Consider a situation in which a narrow, cylindrical, glass tube of radius a is dipped vertically into a liquid of density
, and the liquid level within the tube rises a height h above the free surface as a consequence of surface tension.
See Figure 4.3. Suppose that the radius of the tube is much less than the capillary length. A tube for which this is
the case is generally known as a capillary tube. According to the discussion in Section 4.4, the shape of the internal
liquid/air interface within a capillary tube is not significantly affected by gravity. Thus, from Section 4.3, the interface
is a segment of a sphere of radius R (say). If is the angle of contact of interface with the glass then simple geometry
(see Figure 4.3) reveals that
a
.
(4.20)
R=
cos

66

FLUID MECHANICS

Hence, from Equation (4.13), the mean curvature of the interface is given by
n=

2 cos
2
=
,
R
a

(4.21)

where is the associated surface tension. [The minus sign in the above expression arises from the fact that n points
towards the center of curvature of the interface, whereas the opposite is true for Equation (4.13).] Finally, from (4.15),
application of the Young-Laplace equation to the interface yields
gh =

2 cos
,
a

(4.22)

which can be rearranged to give


2 cos
.
(4.23)
ga
This result, which relates the height, h, to which a liquid rises in a capillary tube of radius a to the liquids surface
tension, , is known as Jurins law. Note that the assumption that the radius of the tube is much less than the capillary
length is equivalent to the assumption that the height of the interface above the free surface of the liquid is much
greater than the radius of the tube. This follows, from (4.16) and (4.23), because
h

h
l2
= 2 cos 2 .
a
a

(4.24)

Thus, the ordering a l implies that h a.


For the case of water at 20 , assuming a contact angle of 25 , Jurins law yields h(mm) = 13.5/a(mm). Thus,
water rises a height 13.5 mm in a capillary tube of radius 1 mm, but rises 13.5 cm in a capillary tube of radius 0.1 mm.
Note that in the case of a liquid, such a mercury, that has an oblique angle of contact with glass, so that cos < 0, the
liquid level in a capillary tube is depressed below that of the free surface (i.e., h < 0).

4.7 Capillary Curves


Let adopt Cartesian coordinates on the Earths surface such that z increases vertically upward. Suppose that the
interface of a liquid of density and surface tension with the atmosphere corresponds to the surface z = f (x), where
the liquid occupies the region z < f (z). Note that the shape of the interface is y-independent. The unit normal to the
interface (directed from liquid to air) is thus
n=

(z f )
ez f x e x
,
=
|(z f )| (1 + f x2 )1/2

(4.25)

where f x = d f /dx. Hence, the mean curvature of the interface is


n=

f xx
,
(1 + f x2 )3/2

(4.26)

where f xx = d2 f /dx2 . According to (4.15) and (4.18), the shape of the interface is governed by the nonlinear differential equation
l2 f xx
f =
.
(4.27)
(1 + f x2 )3/2
where the vertical height, f , of the interface is measured relative to its equilibrium height in the absence of surface
tension. Multiplying the above equation by f x /l2 , and integrating with respect to x, we obtain
1
f2
= C 2,
2
1/2
(1 + f x )
2l
where C is a constant. It follows that
C

f2
1,
2 l2

(4.28)

(4.29)

Surface Tension

67

2
z/l
1

0
2

0
x/l

Figure 4.4: Capillary curves for /4 3/4 and (in order from the top to the bottom) k = 0.6, 0.7, 0.8, 0.9, and
0.99.
and

Let

1
C f 2 /2 l 2
=
.
fx
[1 (C f 2 /2 l 2 )2 ]1/2

(4.30)

2
1,
k2

(4.31)

2l
(1 k2 sin2 )1/2 .
k

(4.32)

C=
where 0 < k < 1, and
f =
Thus, from (4.31) and (4.32),

f2
= cos(2 ),
2 l2
and so the constraint (4.29) implies that /4 3/4. Moreover, Equations (4.30) and (4.33) reduce to
C

(4.33)

1
dx
1
=
=
.
fx d f
tan(2 )

(4.34)

dx dx d f
l k cos(2 )
,
=
=
d d f d
(1 k2 sin2 )1/2

(4.35)

It follows from (4.32) and (4.34) that

which can be integrated to give

/2

k cos(2 )
d,
(1 k2 sin2 )1/2

(4.36)

!
x
2
2
F(, k) + E(,
k),
= k
l
k
k

(4.37)

x
=
l

assuming that x = 0 when = /2. Thus, we get

68

FLUID MECHANICS

0.8
0.7
0.6
0.5
z/l 0.4
0.3
0.2
0.1
0
1

0.5

0
x/l

0.5

Figure 4.5: Liquid/air interface for a liquid trapped between two vertical parallel plates located at x = l. The contact
angle of the interface with the plates is = 30 .
where
k) =
E(,
k) =
F(,

E(/2, k) E(, k),

(4.38)

F(/2, k) F(, k),

(4.39)

and
E(, k) =

F(, k) =

(1 k2 sin2 )1/2 ,

(4.40)

(1 k2 sin2 )1/2 ,

(4.41)

are types of incomplete elliptic integral.1 In conclusion, the interface shape is determined parametrically by
!
2
2
x
F(, k) + E(,
k),
= k
l
k
k
z
l

2
(1 k2 sin2 )1/2 ,
k

(4.42)
(4.43)

where /4 3/4. Here, the parameter k is restricted to lie in the range 0 < k < 1.
Figure 4.4 shows the capillary curves predicted by (4.42) and (4.43) for various different values of k. Here, we
have chosen the plus sign in (4.43). However, if the minus sign is chosen then the curves are simply inverted: i.e.,
x x and z z. In can be seen that all of the curves shown in the figure are symmetric about x = 0: i.e., z z as
x x. Consequently, we can use these curves to determine the shape of the liquid/air interface which arises when a
liquid is trapped between two flat vertical plates (made of the same material) that are parallel to one another. Suppose
1 See

Handbook of Mathematical Functions, M. Abramowitz, and I.A. Stegun (Dover, New York NY, 1965).

Surface Tension

69

that the plates in question lie at x = d. Furthermore, let the angle of contact of the interface with the plates be ,
where < /2. Since the angle of contact is acute, we expect the liquid to be drawn upward between the plates, and
the interface to be concave (from above). This corresponds to the positive sign in (4.43). In order for the interface to
meet the plates at the correct angle, we require f x = 1/ tan at x = d and f x = 1/ tan at x = +d. However, if one
of these boundary conditions is satisfied then, by symmetry, the other is automatically satisfied. From Equation (4.34)
(choosing the positive sign), the latter boundary condition yields tan(2 ) = 1/ tan at x = +d, which is equivalent to
x = +d when = 3/4/2. Substituting this value of into Equation (4.42), we can numerically determine the value
of k for which x = d. The interface shape is then given by Equations (4.42) and (4.43), using the aforementioned value
of k, and in the range /4 + /2 to 3/4 /2. For instance, if d = l and = 30 then k = 0.9406, and the associated
interface is shown in Figure 4.5. Furthermore, if we invert this interface (i.e., x x and z z) then we obtain the
interface which corresponds to the same plate spacing, but an obtuse contact angle of = 180 30 = 150.
Consider the limit k 1, which is such that the distance between the two plates is much less than the capillary
length. It is easily demonstrated that, at small k, 2
k)
E(,
k)
F(,

k2

( + sin cos ),
4
k2

+ ( + sin cos ),
4

(4.44)
(4.45)

where = /2 . Thus, Equations (4.42) and (4.43) reduce to


x
l
z
l

sin(2 ),
2
2 k

[1 cos(2 )].
k 2

(4.46)
(4.47)

It follows that the interface is a segment of the curved surface of a cylinder whose axis runs parallel to the y-axis. If
the distance between the plates is 2 d, and the contact angle is , then we require x = d when = 3/4 /2 (which
corresponds to = /4 + /2). From Equation (4.46), this constraint yields
d k
cos .
l
2

(4.48)

Thus, the height that the liquid rises between the two platesi.e., h = z(x = 0) = z( = /2) 2 l/kis given by
cos
.
gd

(4.49)

This result is the form taken by Jurins law, (4.23), for a liquid drawn up between two parallel plates of spacing 2 d.
Consider the case k = C = 1, which is such that the distance between the two plates is infinite. Let the leftmost
plate lie at x = 0, and let us completely neglect the rightmost plate, since it lies at infinity. Suppose that h = z(x = 0)
is the height of the interface above the free surface of the liquid at the point where the interface meets the leftmost
plate. If is the angle of contact of the interface with the plate then we require f x = 1/ tan at x = 0. Since C = 1, it
follows from (4.28) that
h2
= 1 sin ,
(4.50)
2 l2
or
h = 2 l sin(/4 /2).
(4.51)
Furthermore, again recalling that C = 1, Equation (4.30) can be integrated to give
x=

2 See

df
=l
fx

h/2l

z/2l

1 2 y2
dy,
y (1 y2 )1/2

Handbook of Mathematical Functions, M. Abramowitz, and I.A. Stegun (Dover, New York NY, 1965).

(4.52)

70

FLUID MECHANICS

1.1
1
0.9
0.8
0.7
z/l

0.6
0.5
0.4
0.3
0.2
0.1
0

x/l
Figure 4.6: Liquid/air interface for a liquid in contact with a vertical plate located at x = 0. The contact angle of the
interface with the plate is = 25 .
where we have chosen the minus sign, and y = f /2l. Making the substitution y = sin u, this becomes
x
=
l
which reduces to

sin1 (h/2l)

sin1 (z/2l)

!
#sin1 (h/2l)
!
"
1
1 + cos u
,
+ 2 cos u
2 sin u du = ln
sin u
sin u
sin1 (z/2l)

!
!
!1/2
!1/2
2l
h2
x
2l
z2
cosh1
+ 4 2
= cosh1
4 2
,
l
z
h
l
l

(4.53)

(4.54)

since cosh1 (z) ln[z + (z2 1)1/2 ]. Thus, Equations (4.51) and (4.54) specify the shape of a liquid/air interface that
meets an isolated vertical plate at x = 0. In particular, (4.51) gives the height that the interface climbs up the plate
(relative to the free surface) due to the action of surface tension. Note that this height is restricted to lie in the range
2 l h 2 l, irrespective of the angle of contact. Figure 4.6 shows an example interface calculated for = 25 .

4.8 Axisymmetric Soap-Bubbles


Consider an axisymmetric soap-bubble whose surface takes the form r = f (z) in cylindrical coordinates. See Section C.3. The unit normal to the surface is
n

er fz ez
(r f )
=
,
|(r f )| (1 + fz2 )1/2

(4.55)

where fz d f /dz. Hence, from (C.39), the mean curvature of the surface is given by
n=

"
#
1 d
f
.
f fz dz (1 + fz2 )1/2

(4.56)

Surface Tension

71

The Young-Laplace equation, (4.12), then yields


"
#
f
f fz
d
=
,
a
dz (1 + fz2 )1/2
where
a=

(4.57)

.
p0

(4.58)

Here, is the net surface tension, including the contributions from the internal and external soap/air interfaces. Moreover, p0 = p is the pressure difference between the interior and the exterior of the bubble. Equation (4.57) can be
integrated to give
f2
f
+ C,
(4.59)
=
2
1/2
2a
(1 + fz )
where C is a constant.
Suppose that the bubble occupies the region z1 z z2 , where z1 < z2 , and has a fixed radius at its two end-points,
z = z1 and z = z2 . This could most easily be achieved by supporting the bubble on two rigid parallel co-axial rings
located at z = z1 and z = z2 . The net free energy required to create the bubble can be written
E = S p0 V,

(4.60)

where S is area of the bubble surface, and V the enclosed volume. The first term on the right-hand side of the
above expression represents the work needed to overcome surface tension, whilst the second term represents the work
required to overcome the pressure difference, p0 , between the exterior and the interior of the bubble. Now, from the
general principles of statics, we expect a stable equilibrium state of a mechanical system to be such as to minimize the
net free energy, subject to any dynamical constraints. It follows that the equilibrium shape of the bubble is such as to
minimize
Z z2
Z z2
E=
2 f (1 + fz2 )1/2 dz p0
f 2 dz,
(4.61)
z1

z1

subject to the constraint that the bubble radius, f , be fixed at z = z1 and z = z2 . Hence, we need to find the function
f (z) that minimizes the integral
Z z2
L( f, fz ) dz,
(4.62)
z1

where
L( f, fz ) = 2 f (1 + fz2 )1/2 p0 f 2 ,

(4.63)

subject to the constraint that f is fixed at the limits. This is a standard problem in the calculus of variations. (See
Appendix D.) In fact, since the functional L( f, fz ) does not depend explicitly on z, the minimizing function is the
solution of [see Equation (D.14)]
L
= C,
(4.64)
L fz
fz
where C is an arbitrary constant. Thus, we obtain
#
"
f2
f
= C,

2
(1 + fz2 )1/2 2 a

(4.65)

which can be rearranged to give Equation (4.59). Hence, we conclude that application of the Young-Laplace equation
does indeed lead to a bubble shape that minimizes the net free energy of the soap/air interfaces.
Consider the case p0 = 0, in which there is no pressure difference across the surface of the bubble. In this situation,
writing C = b > 0, Equation (4.59) reduces to
f = b (1 + fz2 )1/2 .

(4.66)

72

FLUID MECHANICS

1
0.9
0.8
0.7
0.6
r/c 0.5
0.4
0.3
0.2
0.1
0

0.6 0.4 0.2

0
z/c

0.2

0.4

0.6

Figure 4.7: Radius versus axial distance for a catenoid soap bubble supported by two parallel co-axial rings of radius
c located at z = 0.65 c.
Moreover, according to the previous discussion, the bubble shape specified by (4.66) is such as to minimize the surface
area of the bubble (since the only contribution to the free energy of the soap/air interfaces is directly proportional to
the bubble area). The above equation can be rearranged to give
f2
fz = 2 1
b
which leads to
z z0 =

df
=
fz

or

(f

!1/2

df
= b cosh1 (r/b),
1)1/2

2 /b 2

r = b cosh(|z z0 |/b),

(4.67)

(4.68)

(4.69)

where z0 is a constant. This expression describes an axisymmetric surface known as a catenoid.


Suppose, for instance, that the soap bubble is supported by identical rings of radius c that are located a perpendicular distance 2 d apart. Without loss of generality, we can specify that the rings lie at z = d. It thus follows, from
(4.69), that z0 = 0, and
r = b cosh(z/b).
(4.70)
Here, the parameter b must be chosen so as to satisfy
c = b cosh(d/b).

(4.71)

For example, if d = 0.65 c then b = 0.6416 c, and the resulting bubble shape is illustrated in Figure 4.7.
Let d/c = and d/b = u, in which case the above equation becomes
G(u) = u cosh u = 0.

(4.72)

Surface Tension

73

1.1
1
0.9
0.8

r/

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
2

0
z/

Figure 4.8: Radius versus axial distance for an unduloid soap bubble calculated with k = 0.95.
Now, the function G(u) attains a maximum value
G(u0 ) = u0

1
,
tanh u0

(4.73)

when u0 = sinh1 (1/). Moreover, if G(u0 ) > 0 then Equation (4.72) possesses two roots. It turns out that the root
associated with the smaller value of u minimizes the interface system energy, whereas the other root maximizes the
free energy. Hence, the former root corresponds to a stable equilibrium state, whereas the latter corresponds to an
unstable equilibrium state. On the other hand, if G(u0 ) < 0 then Equation (4.72) possesses no roots, implying the
absence of any equilibrium state. The critical case G(u0 ) = 0 corresponds to u = uc and = c , where uc tanh uc = 1
and c = 1/ sinh uc . It is easily demonstrated that uc = 1.1997 and c = 0.6627. We conclude that a stable equilibrium
state of a catenoid bubble only exists when c , which corresponds to d 0.6627 c. If the relative ring spacing d
exceeds the critical value 0.6627 c then the bubble presumably bursts.
Consider the case p0 , 0, in which there is a pressure difference across the surface of the bubble. In this situation,
writing
2a =
2aC

+ ,

(4.74)

(4.75)

Equation (4.59) becomes


( + ) f
= f 2 + ,
(1 + fz2 )1/2

(4.76)

which can be rearranged to give


(2 f 2 )1/2 ( f 2 2 )1/2
.
(4.77)
f 2 +
We can assume, without loss of generality, that || > ||. It follows, from the above expression, that || f ||.
Hence, we can write
fz =

f2

2 cos2 + 2 sin2 ,

(4.78)

74

FLUID MECHANICS

1.1
1
0.9
0.8

r/

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0.5

0
z/

0.5

Figure 4.9: Radius versus axial distance for a positive pressure nodoid soap bubble calculated with k = 0.95.
k2

where 0 /2 and 0 < k 1. It follows that

(4.79)

|| (1 k2 sin2 )1/2 ,

(4.80)

sgn() || (1 k )

(4.81)

!
dz
1 df

,
=
= f +
d
fz d
f

(4.82)

h
i
|z| = || E(, k) + sgn( ) (1 k2 )1/2 F(, k) ,

(4.83)

=
and

which can be integrated to give

2 2
,
2

2 1/2

where E(, k) and F(, k) are incomplete elliptic integrals [see Equations (4.40) and (4.41)]. Here, we have assumed
that = 0 when z = 0. There are three cases of interest.
In the first case, > 0 and > 0. It follows that (1 k2 )1/2 r/ 1 for /2 0, and 0.5 /(p0 ) < 1 for
1 k > 0, where
r

|z| =

(1 k2 sin2 ),
h
i
E(, k) + (1 k2 )1/2 F(, k) .

(4.84)

(1 k2 sin2 ),
h
i
E(, k) (1 k2 )1/2 F(, k) .

(4.86)

(4.85)

The axisymmetric curve parameterized by the above pair of equations is known as an unduloid. Note that an unduloid
bubble always has positive internal pressure (relative to the external pressure): i.e., p0 > 0. An example unduloid soap
bubble is illustrated in Figure 4.8
In the second case, > 0 and < 0. It follows that (1 k2 )1/4 r/ 1 for 0 0, and 0 < /(p0 ) 0.5
for 0 < k 1, where 0 = sin1 ([1 (1 k2 )1/2 ]1/2 /k), and
r

|z| =

(4.87)

Surface Tension

75

0.6
0.5

r/||

0.4
0.3
0.2
0.1
0

0.1

0.2

0.3
z/||

0.4

0.5

Figure 4.10: Radius versus axial distance for a negative pressure nodoid soap bubble calculated with k = 0.95.
The axisymmetric curve parameterized by the above pair of equations is known as an nodoid. This particular type
of nodoid bubble has positive internal pressure: i.e., p0 > 0. An example positive pressure nodoid soap bubble is
illustrated in Figure 4.9.
In the third case, < 0 and > 0. It follows that (1 k2 )1/2 | r/|| | (1 k2 )1/4 for /2 0 (or
/2 0 ), and 0 > /(p0 ||) 0.5 for 0 < k 1, where
r
|z|

= || (1 k2 sin2 ),
h
i
= || E(, k) (1 k2 )1/2 F(, k) .

(4.88)
(4.89)

The axisymmetric curve parameterized by the above pair of equations is again a nodoid. However, this particular type
of nodoid bubble has negative internal pressure: i.e., p0 < 0. An example negative pressure nodoid soap bubble is
illustrated in Figure 4.10.

4.9 Exercises
4.1. Show that if N equal spheres of water coalesce so as to form a single spherical drop then the surface energy is decreased by
a factor 1/N 1/3 .
4.2. A circular cylinder of radius a, height h, and specific gravity s floats upright in water. Show that the depth of the base below
the general level of the water surface is
2
sh +
cos ,
a
where is the surface tension at the air/water interface, and the contact angle of the interface with the cylinder.
4.3. A film of water is held between two parallel plates of glass a small distance 2d apart. Prove that the apparent attraction
between the plates is
2 A cos
+ L sin ,
d
where is the surface tension at the air/water interface, the angle of contact of the interface with glass, A the area of the
film, and L the circumference of the film.

76

FLUID MECHANICS

4.4. Show that if the surface of a sheet of water is slightly corrugated then the surface energy is increased by

!2

dx

per unit breadth of the corrugations. Here, x is measured horizontally, perpendicular to the corrugations. Moreover, denotes
the elevation of the surface above the mean level. Finally, is the surface tension at an air/water interface. If the corrugations
are sinusoidal, such that
= a sin(k x),
show that the average increment of the surface energy per unit area is (1/4) a2 k2 .
4.5. A mass of liquid, which is held together by surface tension alone, revolves about a fixed axis at a small angular velocity ,
so as to assume a slightly spheroidal shape of mean radius a. Prove that the ellipticity of the spheroid is
=

2 a3
,
8

where is the uniform mass density, and the surface tension. [If r+ is the maximum radius, and r the minimum radius,
then a = (r+2 r )1/3 , and the ellipticity is defined = (r+ r )/r+ .]
4.6. A liquid mass rotates, in the form of a circular ring of radius a and small cross-section, with a constant angular velocity ,
about an axis normal to the plane of the ring, and passing through its center. The mass is held together by surface tension
alone. Show that the section of the ring must be approximately circular. Demonstrate that
=

2
a c2

!1/2

where is the density, the surface tension, and c the radius of the cross-section.
4.7. Two spherical soap bubbles of radii a1 and a2 are made to coalesce. Show that when the temperature of the gas in the
resulting bubble has returned to its initial value the radius a of the bubble satisfies
p0 a3 + 4 a2 = p0 (a13 + a23 ) + 4 (a12 + a22 ),
where p0 is the ambient pressure, and the surface tension of the soap/air interfaces.
4.8. A rigid sphere of radius a rests on a flat rigid surface, and a small amount of liquid surrounds the contact point, making a
concave-planar lens whose diameter is small compared to a. The angle of contact of the liquid/air interface with each of the
solid surfaces is zero, and the surface tension of the interface is . Show that there is an adhesive force of magnitude 4 a
acting on the sphere. (It is interesting to note that the force is independent of the volume of liquid.)
4.9. Two small solid bodies are floating on the surface of a liquid. Show that the effect of surface tension is to make the objects
approach one another if the liquid/air interface has either an acute or an obtuse angle of contact with both bodies, and to
make them move away from one another if the interface has an acute angle of contact with one body, and an obtuse angle of
contact with the other.

Incompressible Inviscid Fluid Dynamics

77

5 Incompressible Inviscid Fluid Dynamics

5.1 Introduction
This chapter introduces some of the fundamental concepts that arise in the theory of incompressible, inviscid (or, to
be more exact, high Reynolds number) fluid motion.

5.2 Streamlines, Stream Tubes, and Stream Filaments


A line drawn in a fluid such that its tangent at each point is parallel to the local fluid velocity is called a streamline. The
aggregate of all the streamlines at a given instance in time constitutes the instantaneous flow pattern. The streamlines
drawn through each point of a closed curve constitute a stream tube. Finally, a stream filament is defined as a stream
tube whose cross-section is a curve of infinitesimal dimensions.
When the flow is unsteady then the configuration of the stream tubes and filaments changes from time to time.
However, when the flow is steady then the stream tubes and filaments are stationary. In the latter case, a stream tube
acts like an actual tube through which the fluid is flowing. This follows because there can be no flow across the walls,
and into the tube, since the flow is, by definition, always tangential to these walls. Moreover, the walls are fixed in
space and time, since the motion is steady. Thus, the motion of the fluid within the tube would be unchanged were the
walls replaced by a rigid frictionless boundary.
Consider a stream filament of an incompressible fluid whose motion is steady. Suppose that the cross-sectional
area of the filament is sufficiently small that the fluid velocity is the same at each point on the cross-section. Moreover,
let the cross-section be everywhere normal to the direction of this common velocity. Suppose that v1 and v2 are the
flow speeds at two points on the filament at which the cross-sectional areas are S 1 and S 2 , respectively. Consider the
section of the filament lying between these points. Since the fluid is incompressible, the same volume of fluid must
flow into one end of the section, in a given time interval, as flows out of the other, which implies that
v1 S 1 = v2 S 2 .

(5.1)

This is the simplest manifestation of the equation of fluid continuity discussed in Section 2.9. The above result is
equivalent to the statement that the product of the speed and cross-sectional area is constant along any stream filament
of an incompressible fluid in steady motion. Thus, a stream filament within such a fluid cannot terminate unless the
velocity at that point becomes infinite. Leaving this case out of consideration, it follows that stream filaments in
steadily flowing incompressible fluids are either closed loops, or terminate at the boundaries of the fluid. The same is,
of course, true of streamlines.

5.3 Bernoullis Theorem


In its most general form, Bernoullis theoremwhich was discovered by Daniel Bernoulli (17001783)states that,
in the steady flow of an inviscid fluid, the quantity
p
+T
(5.2)

is constant along a streamline, where p is the pressure, the density, and T the total energy per unit mass.
The proof is straightforward. Consider the body of fluid bounded by the cross-sectional areas AB and CD of the
stream filament pictured in Figure 5.1. Let us denote the values of quantities at AB and CD by the suffixes 1 and 2,
respectively. Thus, p1 , v1 , 1 , S 1 , T1 are the pressure, fluid speed, density, cross-sectional area, and total energy per
unit mass, respectively, at AB, etc. Suppose that, after a short time interval t, the body of fluid has moved such that it
occupies the section of the filament bounded by the cross-sections A B and C D , where AA = v1 t and CC = v2 t.
Since the motion is steady, the mass m of the fluid between AB and A B is the same as that between CD and C D , so
that
m = S 1 v1 t 1 = S 2 v2 t 2 .
(5.3)

78

FLUID MECHANICS

D
D
C

B
A

B
A

Figure 5.1: Bernoullis theorem.


Let T denote the total energy of the section of the fluid lying between A B and CD. Thus, the increase in energy of
the fluid body in the time interval t is
(m T2 + T ) (m T1 + T ) = m (T2 T1 ).

(5.4)

In the absence of viscous energy dissipation, this energy increase must equal the net work done by the fluid pressures
at AB and CD, which is
!
p1 p2
p1 S 1 v1 t p2 S 2 v2 t = m

.
(5.5)
1 2
Equating expressions (5.4) and (5.5), we find that
p2
p1
+ T1 =
+ T2 ,
(5.6)
1
2
which demonstrates that p/ + T has the same value at any two points on a given stream filament, and is therefore
constant along the filament. Note that Bernoullis theorem has only been proved for the case of the steady motion of
an inviscid fluid. However, the fluid in question may either be compressible or incompressible.
For the particular case of an incompressible fluid, moving in a conservative force-field, the total energy per unit
mass is the sum of the kinetic energy per unit mass, (1/2) v2, and the potential energy per unit mass, , and Bernoullis
theorem thus becomes
p 1 2
+ v + = constant along a streamline.
(5.7)
2
If we focus on a particular streamline, 1 (say), then Bernoullis theorem states that
p 1 2
+ v + = C1 ,
2

(5.8)

where C1 is a constant characterizing that streamline. If we consider a second streamline, 2 (say), then
p 1 2
+ v + = C2 ,
2

(5.9)

where C2 is another constant. It is not generally the case that C1 = C2 . If, however, the fluid motion is irrotational
then the constant in Bernoullis theorem is the same for all streamlines (see Section 5.7), so that
p 1 2
+ v + =C
2

(5.10)

Incompressible Inviscid Fluid Dynamics

79

S2
B

A
S1
Figure 5.2: A vortex filament.
throughout the fluid.

5.4 Vortex Lines, Vortex Tubes, and Vortex Filaments


The curl of the velocity field of a fluid, which is generally termed vorticity, is usually represented by the symbol :
i.e.,
= v.
(5.11)
A vortex line is a line whose tangent is everywhere parallel to the local vorticity vector. The vortex lines drawn
through each point of a closed curve constitute the surface of a vortex tube. Finally, a vortex filament is a vortex tube
whose cross-section is of infinitesimal dimensions.
Consider a section AB of a vortex filament. The filament is bounded by the curved surface that forms the filament
wall, as well as two plane surfaces, whose vector areas are S1 and S2 (say), which form the ends of the section at
points A and B, respectively. See Figure 5.2. Let the plane surfaces have outward pointing normals that are parallel (or
anti-parallel) to the vorticity vectors, 1 and 2 , at points A and B, respectively. Gausss theorem (see Section A.20),
applied to the section, yields
I
Z
dS =

dV,

(5.12)

where dS is an outward directed surface element, and dV a volume element. However,


=v0
[see Equation (A.173)], implying that

dS = 0.

(5.13)

(5.14)

Now, dS = 0 on the curved surface of the filament, since is, by definition, tangential to this surface. Thus, the
only contributions to the surface integral come from the plane areas S1 and S2 . It follows that
I
dS = S 2 2 S 1 1 = 0.
(5.15)
This result is essentially an equation of continuity for vortex filaments. It implies that the product of the magnitude of
the vorticity and the cross-sectional area, which is termed the vortex intensity, is constant along the filament. It follows
that a vortex filament cannot terminate in the interior of the fluid. For, if it did, the cross-sectional area, S , would have

80

FLUID MECHANICS

to vanish, and, therefore, the vorticity, , would have to become infinite. Thus, a vortex filament must either form a
closed vortex ring, or must terminate at the fluid boundary.
Since a vortex tube can be regarded as a bundle of vortex filaments whose net intensity is the sum of the intensities
of the constituent filaments, we conclude that the intensity of a vortex tube remains constant along the tube.

5.5 Circulation and Vorticity


Consider a closed curve C situated entirely within a moving fluid. The vector line integral (see Section A.14)
I
v dr,
(5.16)
C =
C

where dr is an element of C, and the integral is taken around the whole curve, is termed the circulation of the flow
around the curve. The sense of circulation (e.g., either clockwise or counter-clockwise) is arbitrary.
Let S be a surface having the closed curve C for a boundary, and let dS be an element of this surface (see Section A.7) with that direction of the normal which is related to the chosen sense of circulation around C by the right-hand
circulation rule (see Section A.8). According to Stokes theorem (see Section A.22),
Z
I
dS.
(5.17)
v dr =
C =
S

Thus, we conclude that circulation and vorticity are intimately related to one another. In fact, according to the above
expression, the circulation of the fluid around loop C is equal to the net sum of the intensities of the vortex filaments
passing through the loop and piercing the surface S (with a filament making a positive, or negative, contribution to the
sum depending on whether it pierces the surface in the direction determined by the chosen sense of circulation around
C and the right-hand circulation rule, or in the opposite direction). One important proviso to (5.17) is that the surface
S must lie entirely within the fluid.

5.6 Kelvin Circulation Theorem


According to the Kelvin circulation theorem, which is named after Lord Kelvin (18241907), the circulation around
any co-moving loop in an inviscid fluid is independent of time. The proof is as follows. The circulation around a given
loop C is defined
I
C =

v dr.

However, for a loop that is co-moving with the fluid, we have dv = d(dr/dt) = d(dr)/dt. Thus,
I
I
dC
dv
v dv.
=
dr +
dt
C dt
C

(5.18)

(5.19)

However, by definition, dv/dt = Dv/Dt for a co-moving loop (see Section 2.10). Moreover, the equation of motion of
an incompressible inviscid fluid can be written [see Equation (2.79)]
!
Dv
p
=
+ ,
(5.20)
Dt

since is a constant. Hence,


dC
=
dt

!
p 1 2
v + dr = 0,
2

(5.21)

since v dv = d(v2 /2) = (v2 /2) dr (see Section A.18), and p/ v2 /2 + is obviously a single-valued function.
One corollary of the Kelvin circulation theorem is that the fluid particles that form the walls of a vortex tube at a
given instance in time continue to form the walls of a vortex tube at all subsequent times. To prove this, imagine a
closed loop C that is embedded in the wall of a vortex tube but does not circulate around the interior of the tube. See

Incompressible Inviscid Fluid Dynamics

81

Figure 5.3: A vortex tube.


Figure 5.3. The normal component of the vorticity over the surface enclosed by C is zero, since all vorticity vectors are
tangential to this surface. Thus, from (5.17), the circulation around the loop is zero. By Kelvins circulation theorem,
the circulation around the loop remains zero as the tube is convected by the fluid. In other words, although the surface
enclosed by C deforms, as it is convected by the fluid, it always remains on the tube wall, since no vortex filaments
can pass through it.
Another corollary of the circulation theorem is that the intensity of a vortex tube remains constant as it is convected
by the fluid. This can be proved by considering the circulation around the loop C pictured in Figure 5.3.

5.7 Irrotational Flow


Flow is said to be irrotational when the vorticity has the magnitude zero everywhere. It immediately follows, from
Equation (5.17), that the circulation around any arbitrary loop in an irrotational flow pattern is zero (provided that the
loop can be spanned by a surface that lies entirely within the fluid). Hence, from Kelvins circulation theorem, if an
inviscid fluid is initially irrotational then it remains irrotational at all subsequent times. This can be seen more directly
from the equation of motion of an inviscid incompressible fluid which, according to Equations (2.39) and (2.79), takes
the form
!
v
p
+ (v ) v =
+ ,
(5.22)
t

since is a constant. However, from Equation (A.171),

Thus, we obtain

(v ) v = (v2 /2) v .

(5.23)

!
v
p 1 2
=
+ v + + v .
t
2

(5.24)

Taking the curl of this equation, and making use of the vector identities 0 [see Equation (A.176)], A 0
[see Equation (A.173)], as well as the identity (A.179), and the fact that v = 0 in an incompressible fluid, we obtain
the vorticity evolution equation
D
= ( ) v.
(5.25)
Dt
Thus, if = 0, initially, then D/Dt = 0, and, consequently, = 0 at all subsequent times.

82

FLUID MECHANICS

Suppose that O is a fixed point, and P an arbitrary movable point, in an irrotational fluid. Let O and P be joined by
two different paths, OAP and OBP (say). It follows that OAPBO is a closed curve. Now, since the circulation around
such a curve in an irrotational fluid is zero, we can write
Z
Z
v dr = 0,
(5.26)
v dr +
PBO

OAP

which implies that

OAP

v dr =

OBP

v dr = P

(5.27)

(say). It is clear that P is a scalar function whose value depends on the position of P (and the fixed point O), but not
on the path taken between O and P. Thus, if O is the origin of our coordinate system, and P an arbitrary point whose
RP
position vector is r, then we have effectively defined a scalar field (r) = O v dr.
Consider a point Q that is sufficiently close to P that the velocity v is constant along PQ. Let be the position
vector of Q relative to P. It then follows that (see Section A.18)
Z Q
= Q + P =
v dr v .
(5.28)
P

The above equation becomes exact in the limit that || 0. Since Q is arbitrary (provided that it is sufficiently close
to P), the direction of the vector is also arbitrary, which implies that
v = .

(5.29)

We, thus, conclude that if the motion of a fluid is irrotational then the associated velocity field can always be expressed
as minus the gradient of a scalar function of position, (r). This scalar function is called the velocity potential, and
flow which is derived from such a potential is known as potential flow. Note that the velocity potential is undefined to
an arbitrary additive constant.
We have demonstrated that a velocity potential necessarily exists in a fluid whose velocity field is irrotational.
Conversely, when a velocity potential exists the flow is necessarily irrotational. This follows because [see Equation (A.176)]
= v = = 0.
(5.30)
Incidentally, the fluid velocity at any given point in an irrotational fluid is normal to the constant- surface that passes
through that point.
If a flow pattern is both irrotational and incompressible then we have
v =

(5.31)

v = 0.

(5.32)

and
These two expressions can be combined to give (see Section A.21)
2 = 0.

(5.33)

In other words, the velocity potential in an incompressible irrotational fluid satisfies Laplaces equation.
According to Equation (5.24), if the flow pattern in an incompressible inviscid fluid is also irrotational, so that
= 0 and v = , then we can write
!
p 1 2

= 0,
(5.34)
+ v +
2
t
which implies that

p 1 2
+ v +
= C(t),
2
t

(5.35)

Incompressible Inviscid Fluid Dynamics

83

P
C

B
A
Figure 5.4: Two-dimensional flow.
where C(t) is uniform in space, but can vary in time. In fact, the time variation of C(t) can be eliminated by adding the
appropriate function of time (but not of space) to the velocity potential, . Note that such a procedure does not modify
the instantaneous velocity field v derived from . Thus, the above equation can be rewritten

p 1 2
+ v +
= C,
2
t

(5.36)

where C is constant in both space and time. Expression (5.36) is a generalization of Bernoullis theorem (see Section 5.3) that takes non-steady flow into account. However, this generalization is only valid for irrotational flow. For
the special case of steady flow, we get
p 1 2
+ v + = C,
(5.37)
2
which demonstrates that for steady irrotational flow the constant in Bernoullis theorem is the same on all streamlines.
(See Section 5.3.)

5.8 Two-Dimensional Flow


Fluid motion is said to be two-dimensional when the velocity at every point is parallel to a fixed plane, and is the same
everywhere on a given normal to that plane. Thus, in Cartesian coordinates, if the fixed plane is the x-y plane then we
can express a general two-dimensional flow pattern in the form
v = v x (x, y, t) e x + vy (x, y, t) ey .

(5.38)

Let A be a fixed point in the x-y plane, and let ABP and ACP be two curves, also in the x-y plane, that join A to an
arbitrary point P. See Figure 5.4. Suppose that fluid is neither created nor destroyed in the region, R (say), bounded by
these curves. Since the fluid is incompressible, which essentially means that its density is uniform and constant, fluid
continuity requires that the rate at which the fluid flows into the region R, from right to left across the curve ABP, is
equal to the rate at which it flows out the of the region, from right to left across the curve ACP. Now, the rate of fluid
flow across a surface is generally termed the flux. Thus, the flux (per unit length parallel to the z-axis) from right to
left across ABP is equal to the flux from right to left across ACP. Since ACP is arbitrary, it follows that the flux from
right to left across any curve joining points A and P is equal to the flux from right to left across ABP. In fact, once the
base point A has been chosen, this flux only depends on the position of point P, and the time t. In other words, if we
denote the flux by then it is solely a function of the location of P and the time. Thus, if point A lies at the origin, and

84

FLUID MECHANICS

P2
P1

A
Figure 5.5: Two-dimensional flow.
point P has Cartesian coordinates x, y, then we can write
= (x, y, t).

(5.39)

The function is known as the stream function. Moreover, the existence of a stream function is a direct consequence
of the assumed incompressible nature of the flow.
Consider two points, P1 and P2 , in addition to the fixed point A. See Figure 5.5. Let 1 and 2 be the fluxes from
right to left across curves AP1 and AP2 . Now, using similar arguments to those employed above, the flux across AP2 is
equal to the flux across AP1 plus the flux across P1 P2 . Thus, the flux across P1 P2 , from right to left, is 2 1 . Now,
if P1 and P2 both lie on the same streamline then the flux across P1 P2 is zero, since the local fluid velocity is directed
everywhere parallel to P1 P2 . It follows that 1 = 2 . Hence, we conclude that the stream function is constant along a
streamline. The equation of a streamline is thus = c, where c is an arbitrary constant.
Let P1 P2 = s be an infinitesimal arc of a curve that is sufficiently short that it can be regarded as a straight-line.
The fluid velocity in the vicinity of this arc can be resolved into components parallel and perpendicular to the arc. The
component parallel to s contributes nothing to the flux across the arc from right to left. The component perpendicular
to s contributes v s to the flux. However, the flux is equal to 2 1 . Hence,
v =

2 1
.
s

(5.40)

In the limit s 0, the perpendicular velocity from right to left across ds becomes
v =

d
.
ds

(5.41)

Thus, in Cartesian coordinates, by considering infinitesimal arcs parallel to the x- and y-axes, we deduce that

,
y

vx

vy

.
x

(5.42)
(5.43)

These expressions can be combined to give


v = ez = z .

(5.44)

Incompressible Inviscid Fluid Dynamics

85

Note that when the fluid velocity is written in this form then it immediately becomes clear that the incompressibility
constraint v = 0 is automatically satisfied [since (A B) = 0see Equations (A.175) and (A.176)]. It is also
clear that the stream function is undefined to an arbitrary additive constant.
The vorticity in two-dimensional flow takes the form
= z ez ,
where

(5.45)

vy v x

.
x
y

(5.46)

2 2
+ 2 = 2 .
x2
y

(5.47)

z =
Thus, it follows from Equations (5.42) and (5.43) that
z =

Moreover, irrotational two-dimensional flow is characterized by


2 = 0.

(5.48)

When expressed in terms of cylindrical coordinates (see Section C.3), Equation (5.44) yields
v = vr (r, , t) er + v (r, , t) e ,

(5.49)

where
vr

1
,
r

.
r

(5.50)
(5.51)

Moreover, the vorticity is = z ez , where


z =

!
1
1 2
r
+ 2
.
r r r
r 2

(5.52)

5.9 Two-Dimensional Uniform Flow


Consider a steady two-dimensional flow pattern that is uniform: i.e., such that the fluid velocity is the same everywhere
in the x-y plane. For instance, suppose that the common fluid velocity is
v = V0 cos 0 e x + V0 sin 0 ey ,

(5.53)

which corresponds to flow at the uniform speed V0 in a fixed direction that subtends a (counter-clockwise) angle 0
with the x-axis. It follows, from Equations (5.42) and (5.43), that the stream function for steady uniform flow takes
the form
(5.54)
(x, y) = V0 (sin 0 x cos 0 y) .
When written in terms of cylindrical coordinates, this becomes
(r, ) = V0 r sin( 0 ).

(5.55)

Note, from (5.54), that 2 /x2 = 2 /y2 = 0. Thus, it follows from Equation (5.47) that uniform flow is
irrotational. Hence, according to Section 5.7, such flow can also be derived from a velocity potential. In fact, it is
easily demonstrated that
(r, ) = V0 r cos( 0 ).
(5.56)

86

FLUID MECHANICS
y

Figure 5.6: Streamlines of the flow generated by a line source coincident with the z-axis.

5.10 Two-Dimensional Sources and Sinks


Consider a uniform line source, coincident with the z-axis, that emits fluid isotropically at the steady rate of Q unit
volumes per unit length per unit time. By symmetry, we expect the associated steady flow pattern to be isotropic, and
everywhere directed radially away from the source. See Figure 5.6. In other words, we expect
v = vr (r) er .

(5.57)

Consider a cylindrical surface S of unit height (in the z-direction) and radius r that is co-axial with the source. In a
steady state, the rate at which fluid crosses this surface must be equal to the rate at which the section of the source
enclosed by the surface emits fluid. Hence,
Z
v dS = 2 r vr (r) = Q,
(5.58)
S

which implies that


Q
.
(5.59)
2 r
According to Equations (5.50) and (5.51), the stream function associated with a line source of strength Q that is
coincident with the z-axis is
Q
(5.60)
(r, ) = .
2
Note that the streamlines, = c, are directed radially away from the z-axis, as illustrated in Figure 5.6. Note, also, that
the stream function associated with a line source is multivalued. However, this does not cause any particular difficulty,
since the stream function is continuous, and its gradient single-valued.
Note, from Equation (5.60), that /r = 2 / 2 = 0. Hence, according to (5.52), z = 2 = 0. In other
words, the steady flow pattern associated with a uniform line source is irrotational, and can, thus, be derived from a
velocity potential. In fact, it is easily demonstrated that this potential takes the form
vr (r) =

(r, ) =

Q
ln r.
2

(5.61)

Incompressible Inviscid Fluid Dynamics

87

A uniform line sink, coincident with the z-axis, which absorbs fluid isotropically at the steady rate of Q unit
volumes per unit length per unit time has an associated steady flow pattern
v=

Q
er ,
2 r

(5.62)

whose stream function is

Q
.
2
This flow pattern is also irrotational, and can be derived from the velocity potential
(r, ) =

(r, ) =

Q
ln r.
2

(5.63)

(5.64)

Consider a line source and a line sink of equal strength, which both run parallel to the z-axis, and are located a
small distance apart in the x-y plane. Such an arrangement is known as a dipole or doublet line source. Suppose that
the line source, which is of strength Q, is located at r = d/2 (where r is a position vector in the x-y plane), and that
the line sink, which is also of strength Q, is located at r = d/2. Let the function
Q (r) =

Q
Q
=
tan1 (y/x)
2
2

(5.65)

be the stream function associated with a line source of strength Q located at r = 0. Thus, Q (r r0 ) is the stream
function associated with a line source of strength Q located at r = r0 . Furthermore, the stream function associated
with a line sink of strength Q located at r = r0 is Q (r r0 ). Now, we expect the flow pattern associated with the
combination of a source and a sink to be the vector sum of the flow patterns generated by the source and sink taken in
isolation. It follows that the overall stream function is the sum of the stream functions generated by the source and the
sink taken in isolation. In other words,
(r) = Q (r d/2) Q (r + d/2) d Q (r),

(5.66)

to first order in d/r. Hence, if d = d (cos 0 e x + sin 0 ey ) = d [cos( 0 ) er sin( 0 ) e ], so that the line joining
the sink to the source subtends a (counter-clockwise) angle 0 with the x-axis, then
(r, ) =

D sin( 0 )
,
2
r

(5.67)

where D = Q d is termed the strength of the dipole source. Note that the above stream function is antisymmetric across
the line = 0 joining the source to the sink. It follows that the associated dipole flow pattern,
vr (r, )

v (r, )

D cos( 0 )
,
2
r2
D sin( 0 )
,
2
r2

(5.68)
(5.69)

is symmetric across this line. Figure 5.7 shows the streamlines associated with a dipole flow pattern characterized by
D > 0 and 0 = 0. Note that the flow speed in a dipole pattern falls off like 1/r 2 .
A dipole flow pattern is necessarily irrotational since it is a linear superposition of two irrotational flow patterns.
The associated velocity potential is
D cos( 0 )
.
(5.70)
(r, ) =
2
r

5.11 Two-Dimensional Vortex Filaments


Consider a vortex filament of intensity that is coincident with the z-axis. By symmetry, we expect the associated
flow pattern to circulate isotropically around the filament. See Figure 5.8. In other words, we expect
v = v (r) e .

(5.71)

88

FLUID MECHANICS

1
0.8
0.6
0.4
0.2
y

0
0.2
0.4
0.6
0.8
1
1

0.5

0
x

0.5

Figure 5.7: Streamlines of the flow generated by a dipole line source coincident with the z-axis and aligned along the
x-axis. The flow is outward along the positive x-axis and inward along the negative x-axis.
H
Now, according to Section 5.5, the circulation, v dr, around any closed curve in the x-y plane is equal to the net
intensity of the vortex filaments that pass through the curve. Consider a circular curve of radius r that is concentric
with the origin. It follows that
I
r =
v dr = 2 r v (r) = ,
(5.72)
or
v (r) =

.
2 r

(5.73)

According to Equations (5.50) and (5.51), the stream function associated with a vortex filament of intensity that
is coincident with the z-axis is

ln r.
(5.74)
(r, ) =
2
Note that the streamlines, = c, circulate around the z-axis, as illustrated in Figure 5.8.
It can be seen, from Equation (5.74), that (/r)(r /r) = / = 0. Hence, it follows from (5.52) that
z = 2 = 0. In other words, the flow pattern associated with a straight vortex filament is irrotational. This
is a somewhat surprising result, since there is a net circulation of the flow around the filament, and, according to
Section 5.5, non-zero circulation implies non-zero vorticity. The paradox can be resolved by supposing that the
filament has a small, but finite, radius. In fact, let the filament have the finite radius a, and be such that the vorticity is
uniform inside this radius, and zero outside: i.e.,
z =

/ a2
0

ra
.
r>a

(5.75)

Note that the intensity of the filament (i.e., the product of its vorticity and cross-sectional area) is still . According to
Equation (5.52), and assuming that = (r),
! (
/ a2
1 d d
r
=
r r dr
0

ra
.
r>a

(5.76)

Incompressible Inviscid Fluid Dynamics

89

Figure 5.8: Streamlines of the flow generated by a line vortex coincident with the z-axis.
The solution that is well-behaved at r = 0, and continuous (up to its first derivative) at r = a, is
(
(/4) (r2/a2 1)
ra
(r, ) =
.
(/2) ln(r/a)
r>a

(5.77)

Note that this expression is equivalent to (5.74) (apart from an unimportant additive constant) outside the filament, but
differs inside. The associated circulation velocity, v (r) = /r, is
(
(/2) (r/a2)
ra
v (r) =
,
(5.78)
(/2) (1/r)
r>a
whereas the circulation, r (r) = 2 r v (r), is written
(
(r/a)2
r (r) =

ra
.
r>a

(5.79)

Thus, we conclude that the flow pattern associated with a straight vortex filament is irrotational outside the filament,
but has finite vorticity inside the filament. Moreover, the non-zero internal vorticity generates a constant net circulation
of the flow outside the filament. In the limit in which the radius of the filament tends to zero, the vorticity within the
filament tends to infinity (in such a way that the product of the vorticity and the cross-sectional area of the filament
remains constant), and the region of the fluid in which the vorticity is non-zero becomes infinitesimal in extent.
Let us determined the pressure profile in the vicinity of a vortex filament of finite radius. Assuming, from symmetry, that p = p(r), Equation (2.149), yields
v2
dp
= ,
(5.80)
dr
r
which can be integrated to give
Z 2
v
p = p
dr,
(5.81)
r
r
where p is the pressure at infinity. Making use of expression (5.78), we obtain
(
p (/2) (/2 a)2 (2 r2 /a2 )
ra
p(r) =
.
(5.82)
2
2
p (/2) (/2 a) (a/r)
r>a

90

FLUID MECHANICS

It follows that the minimum pressure occurs at the center of the vortex (r = 0), and takes the value
 2
.
p0 = p
2 a

(5.83)

Now, under normal circumstances, the pressure in a fluid must remain positive, which implies that a vortex filament
of intensity , embedded in a fluid of density and background pressure p , has a minimum radius of order
!1/2  

amin
.
(5.84)
p
2
Finally, since the flow pattern outside a straight vortex filament is irrotational, it can be derived from a velocity
potential. In fact, it is easily demonstrated that the appropriate potential takes the form
(r, ) =

.
2

(5.85)

Note that the above potential is multivalued. However, this does not cause any particular difficulty, since the potential
is continuous, and its gradient single-valued.

5.12 Two-Dimensional Irrotational Flow in Cylindrical Coordinates


As we have seen, in a two-dimensional flow pattern, we can automatically satisfy the incompressibility constraint,
v = 0, by expressing the pattern in terms of a stream function. Suppose, however, that, in addition to being
incompressible, the flow pattern is also irrotational. In this case, Equation (5.47) yields
2 = 0.

(5.86)

In cylindrical coordinates, since = (r, , t), this expression implies that (see Section C.3)
!
1 2
1
r
+ 2
= 0.
r r r
r 2

(5.87)

Let us search for a separable steady-state solution of Equation (5.87) of the form

It is easily seen that

(r, ) = R(r) ().

(5.88)

!
r d dR
1 d2
,
r
=
R dr dr
d 2

(5.89)

which can only be satisfied if


r

m2 R,

(5.90)

d2
d 2

m2 ,

(5.91)

d dR
r
dr dr

where m2 is an arbitrary (positive) constant. The general solution of Equation (5.91) is a linear combination of
exp( i m ) and exp(i m ) factors. However, assuming that the flow extends over all values, the function ()
must be single-valued in , otherwise and, hence, vwould not be be single-valued (which is unphysical). It
follows that m can only take integer values (and that m2 must be a positive, rather than a negative, constant). Now, the
general solution of Equation (5.90) is a linear combination of rm and rm factors, except for the special case m = 0,
when it is a linear combination of r0 and ln r factors. Thus, the general stream function for steady two-dimensional
irrotational flow (that extends over all values of ) takes the form
X
(r, ) = 0 + 0 ln r +
(m rm + m rm ) sin[m ( m )],
(5.92)
m>0

Incompressible Inviscid Fluid Dynamics

91

where m , m , and m are arbitrary constants. We can recognize the first few terms on the right-hand side of the above
expression. The constant term 0 has zero gradient, and, therefore, does not give rise to any flow. The term 0 ln r is
the flow pattern generated by a vortex filament of intensity 2 0 , coincident with the z-axis. (See Section 5.11.) The
term 1 r sin( 1 ) corresponds to uniform flow of speed 1 whose direction subtends a (counter-clockwise) angle 1
with the minus x-axis. (See Section 5.9.) Finally, the term 1 sin( 1 )/r corresponds to a dipole flow pattern. (See
Section 5.10.)
The velocity potential associated with the irrotational stream function (5.92) satisfies [see Equations (5.29) and
(5.44)]

r
1
r
It follows that
(r, ) = 0 0 +

X
m>0

1
,
r

= .
r

(m rm m rm ) cos[m ( 0 )].

(5.93)
(5.94)

(5.95)

5.13 Inviscid Flow Past a Cylindrical Obstacle


Consider the steady flow pattern produced when an impenetrable rigid cylindrical obstacle is placed in a uniformly
flowing, incompressible, inviscid fluid, with the cylinder orientated such that its axis is normal to the flow. For instance,
suppose that the radius of the cylinder is a, and that its axis corresponds to the line x = y = 0. Furthermore, let the
unperturbed fluid velocity be of magnitude V0 , and be directed parallel to the x-axis. Now, we expect the flow pattern
to remain unperturbed very far away from the cylinder. In other words, we expect v(r, ) V0 e x as r/a , which
corresponds to a boundary condition on the stream function of the form (see Section 5.9)
(r, ) V0 r sin

as r/a .

(5.96)

Given that the fluid velocity field a large distance upstream of the cylinder is irrotational (since we have already
seen that the flow pattern associated with uniform flow is irrotationalsee Section 5.9), it follows from the Kelvin
circulation theorem (see Section 5.6) that the velocity field remains irrotational as it is convected past the cylinder.
Hence, according to Section 5.8, the stream function of the flow satisfies Laplaces equation,
2 = 0.

(5.97)

The appropriate boundary condition at the surface of the cylinder is simply that the normal fluid velocity there be
zero, since the fluid must stay in contact with the cylinder, but cannot penetrate its surface. Hence, vr (a, )
(1/a) /|r=a = 0, which implies that
(a, ) = 0,
(5.98)
since is undetermined to an arbitrary additive constant. It follows that we are searching for the most general solution
of (5.97) that satisfies the boundary conditions (5.96) and (5.98). Comparison with Equation (5.92) reveals that this
solution takes the form


r  r a

(5.99)
sin ,
(r, ) = V0 a ln
a
a r
where

,
(5.100)
=
2 a V0
and is the circulation of the flow around the cylinder. (Note that the velocity field can be irrotational, but still possess
nonzero circulation around the cylinder, because a loop that encloses the cylinder cannot be spanned by a surface lying
entirely within the fluid. Thus, zero fluid vorticity does not necessarily imply zero circulation around such a loop from
Stokes theorem.) Let us assume that 0, for the sake of definiteness.

92

FLUID MECHANICS

5
4
3
2
1
y/a

0
1
2
3
4
5
5 4 3 2 1 0 1
x/a

Figure 5.9: Streamlines of the flow generated by a cylindrical obstacle of radius a, whose axis runs along the z-axis,
placed in the uniform flow field v = V0 e x . The normalized circulation is = 0.
Figure 5.95.11 show streamlines of the flow calculated for various different values of the normalized circulation,
. For < 2 there exist a pair of points on the surface of the cylinder at which the flow speed is zero. These are known
as stagnation points, and can be located in Figures 5.9 and 5.10 as the points at which streamlines intersect the surface
of the cylinder at right-angles. Now, the tangential fluid velocity at the surface of the cylinder is


= V0 ( + 2 sin ).
vt () = v (a, ) =
(5.101)
r r=a
The stagnation points correspond to the points at which vt = 0 (since the normal velocity is automatically zero at the
surface of the cylinder). Thus, the stagnation points lie at = sin1 (/2). When > 2 the stagnation points coalesce
and move off the surface of the cylinder, as illustrated in Figure 5.11 (the stagnation point corresponds to the point at
which two streamlines cross at right-angles).
The irrotational form of Bernoullis theorem, (5.37), can be combined with the boundary condition v V0 as
r/a , as well as the fact that is constant in the present case, to give
p = p0 +


1  2
V0 v 2 ,
2

(5.102)

where p0 is the constant static fluid pressure a large distance from the cylinder. In particular, the fluid pressure on the
surface of the cylinder is
P() = p(a, ) = p0 +


1  2
V0 vt2 = p1 + V02 (cos 2 2 sin ) ,
2

(5.103)

where p1 = p0 (1/2) V02 (1 + 2 ). The net force per unit length exerted on the cylinder by the fluid has the Cartesian
components
I
F x = P cos a d,
(5.104)
I
Fy = P sin a d.
(5.105)

Incompressible Inviscid Fluid Dynamics

93

5
4
3
2
1
y/a

0
1
2
3
4
5
5 4 3 2 1 0 1
x/a

Figure 5.10: Streamlines of the flow generated by a cylindrical obstacle of radius a, whose axis runs along the z-axis,
placed in the uniform flow field v = V0 e x . The normalized circulation is = 1.
Thus, it follows from (5.103) that
Fx
Fy

0,

2 V02

(5.106)
a = V0 ().

(5.107)

Now, the component of the force which a moving fluid exerts on an obstacle, placed in its path, in a direction parallel
to that of the unperturbed flow is usually called drag. On the other hand, the component of the force which the fluid
exerts in a direction perpendicular to that of the unperturbed flow is usually called lift. Hence, the above equations
imply that if a cylindrical obstacle is placed in a uniformly flowing inviscid fluid then there is zero drag. On the other
hand, as long as there is net circulation of the flow around the cylinder, the lift is non-zero. Now, lift is generated
because (negative) circulation tends to increase the fluid speed directly above, and to decrease it directly below, the
cylinder. Thus, from Bernoullis theorem, the fluid pressure is decreased above, and increased below, the cylinder,
giving rise to a net upward force (i.e., a force in the +y-direction).
Suppose that the cylinder is placed in a fluid which is initially at rest, and that the fluids uniform flow velocity, V0 ,
is then very slowly ramped up (in such a manner that no vorticity is induced in the upstream flow at infinity). Since
the flow pattern is initially irrotational, and since the flow pattern well upstream of the cylinder is assumed to remain
irrotational, the Kelvin circulation theorem indicates Hthat the flow pattern around the cylinder also remains irrotational.
Consider the time evolution of the circulation, = C v dr, around some fixed curve C that lies entirely within the
fluid, and encloses the cylinder. We have
!
#
I
I
I "
dC
v
p 1 2
v dr,
(5.108)

=
dr =
+ v + v dr =
dt
2
C
C t
C
where use has been made of (5.24) (with assumed constant). However, = z ez in two-dimensional flow, and
dr ez = dS, where dS is an outward surface element of a unit depth (in the z-direction) surface whose normal lies in
the x-y plane, and that cuts the x-y plane at C. In other words,
I
dC
(5.109)
= z v dS.
dt
S
We, thus, conclude that the rate of change of the circulation around C is equal to minus the flux of the vorticity across
S [assuming that vorticity is convected by the flow, which follows from (5.25), the fact that = z ez , and the fact

94

FLUID MECHANICS

5
4
3
2
1
y/a

0
1
2
3
4
5
5 4 3 2 1 0 1
x/a

Figure 5.11: Streamlines of the flow generated by a cylindrical obstacle of radius a, whose axis runs along the z-axis,
placed in the uniform flow field v = V0 e x . The normalized circulation is = 2.5.
that /z = 0 in two-dimensional flow]. However, we have already seen that the flow field surrounding the cylinder
is irrotational (i.e., such that z = 0). It follows that C is constant in time. Moreover, since C = 0 originally,
because the fluid surrounding the cylinder was initially at rest, we deduce that C = 0 at all subsequent times. Hence,
we conclude that, in an inviscid fluid, if the circulation of the flow around the cylinder is initially zero then it remains
zero. It follows, from the above analysis, that, in such a fluid, zero drag force and zero lift force are exerted on the
cylinder as a consequence of the fluid flow. This result is known as dAlemberts paradox, after the French scientist
Jean-Baptiste le Rond dAlembert (17171783). DAlemberts result is paradoxical because it would seem, at first
sight, to be a reasonable approximation to neglect viscosity alltogether in high Reynolds number flow. However, if we
do this then we end up with the nonsensical prediction that a high Reynolds number fluid is incapable of exerting any
force on an obstacle placed in its path.

5.14 Inviscid Flow Past a Semi-Infinite Wedge


Consider the situation, illustrated in Figure 5.12, in which incompressible irrotational flow is incident on a impenetrable rigid wedge whose apex subtends an angle . Let the cross-section of the wedge in the x-y plane be both
z-independent and symmetric about the x-axis. Furthermore, let the apex of the wedge lie at x = y = 0. Finally, let the
upstream flow a large distance from the wedge be parallel to the x-axis.
Since the flow is two-dimensional, incompressible, and irrotational, it can be represented in terms of a stream
function that satisfies Laplaces equation. Moreover, in cylindrical coordinates, this equation takes the form (5.87).
The boundary conditions on the stream function are
(r, /2) = (r, 2 /2) = (r, ) = 0.

(5.110)

The first two boundary conditions ensure that the normal velocity at the surface of the wedge is zero. The third
boundary condition follows from the observation that, by symmetry, the streamline that meets the apex of the wedge
splits in two, and then flows along its top and bottom boundaries, combined with well-known result that is constant
on a streamline. It is easily demonstrated that
(r, ) =

A
r1+m sin [(1 + m) ( )]
1+m

(5.111)

Incompressible Inviscid Fluid Dynamics

95

Figure 5.12: Inviscid flow past a wedge.


is a solution of (5.87). Moreover, this solution satisfies the boundary conditions provided (1 + m) (1 /2) = 1, or

.
(5.112)
m=
2

Since, as is well-known, the solutions to Laplaces equation (for problems with well-posed boundary conditions) are
unique, we can be sure that (5.111) is the correct solution to the problem under investigation. According to this
solution, the tangential velocity on the surface of the wedge is given by
vt (r) = A rm ,

(5.113)

where m 0. Note that the tangential velocity is zero at the apex of the wedge. Since the normal velocity is also zero
at this point, we conclude that the apex is a stagnation point of the flow. Figure 5.13 shows the streamlines of the flow
for the case = 1/2.

5.15 Inviscid Flow Over a Semi-Infinite Wedge


Consider the situation illustrated in Figure 5.14 in which an incompressible irrotational fluid flows over an impenetrable rigid wedge whose apex subtends an angle . Let the cross-section of the wedge in the x-y plane be both
z-independent and symmetric about the y-axis. Furthermore, let the apex of the wedge lie at x = y = 0. Finally, let the
upstream flow a large distance from the wedge be parallel to the x-axis.
Since the flow is two-dimensional, incompressible, and irrotational, it can be represented in terms of a stream
function that satisfies Laplaces equation. The boundary conditions on the stream function are
(r, [3 ] /2) = (r, [1 ] /2) = 0.

(5.114)

These boundary conditions ensure that the normal velocity at the surface of the wedge is zero. It is easily demonstrated
that
A
r1m cos [(1 m) ( /2)]
(5.115)
(r, ) =
1m
is a solution of Laplaces equation, (5.87). Moreover, this solution satisfies the boundary conditions provided that
(1 m) (1 /2) = 1/2, or

m=
,
(5.116)
1 +

96

FLUID MECHANICS

3
2
1
y

0
1
2
3
3

0
x

Figure 5.13: Streamlines of inviscid incompressible irrotational flow past a 90 wedge.

Figure 5.14: Inviscid flow over a wedge.

Incompressible Inviscid Fluid Dynamics

97

3
2
1
y

0
1
2
3
3

0
x

Figure 5.15: Streamlines of inviscid incompressible irrotational flow over a 90 wedge.


where = 1 . Since the solutions to Laplaces equation are unique, we can again be sure that (5.115) is the correct
solution to the problem under investigation. According to this solution, the tangential velocity on the surface of the
wedge is given by
vt (r) = A rm ,
(5.117)
where m 0. Note that the tangential velocity, and hence the flow speed, is infinite at the apex of the wedge. However,
this singularity in the flow can be eliminated by slightly rounding the apex. Figure 5.15 shows the streamlines of the
flow for the case = 1/2.

5.16 Velocity Potentials and Stream Functions


As we have seen, a two-dimensional velocity field in which the flow is everywhere parallel to the x-y plane, and there
is no variation along the z-direction, takes the form
v = v x (x, y, t) e x + vy (x, y, t) ey .

(5.118)

Moreover, if the flow is irrotational then v = 0 is automatically satisfied by writing v = , where (x, y, t) is
termed the velocity potential. (See Section 5.7.) Hence,
vx

vy

,
x

.
y

(5.119)
(5.120)

On the other hand, if the flow is incompressible then v = 0 is automatically satisfied by writing v = z , where
(x, y, t) is termed the stream function. (See Section 5.8.) Hence,

,
y

vx

vy

.
x

(5.121)
(5.122)

98

FLUID MECHANICS

Finally, if the flow is both irrotational and incompressible then Equations (5.119)(5.120) and (5.121)(5.122) hold
simultaneously, which implies that

,
y

.
y

(5.123)
(5.124)

It immediately follows, from the previous two expressions, that

or

2
2
2
2
=
=

=
,
x y y x
x2
y2

(5.125)

2 2
+
= 0.
x2 y2

(5.126)

2 2
+
= 0.
x2 y2

(5.127)

Likewise, it can also be shown that

We conclude that, for two-dimensional, irrotational, and incompressible flow, the velocity potential and the stream
function both satisfy Laplaces equation. Equations (5.123) and (5.124) also imply that
= 0 :

(5.128)

i.e., the contours of the velocity potential and the stream function cross at right-angles.

5.17 Exercises
5.1. Liquid is led steadily through a pipeline that passes over a hill of height h into the valley below, the speed at the crest being
v. Show that, by properly adjusting the ratio of the cross-sectional areas of the pipe at the crest and in the valley, the pressure
may be equalized at these two places.
5.2. For the case of the two-dimensional motion of an incompressible fluid, determine the condition that the velocity components
vx

a x + b y,

vy

cx+dy

satisfy the equation of continuity. Show that the magnitude of the vorticity is c b.
5.3. For the case of the two-dimensional motion of an incompressible fluid, show that
vx

2 c x y,

vy

c (a2 + x2 y2 )

are the velocity components of a possible flow pattern. Determine the stream function and sketch the streamlines. Prove that
the motion is irrotational, and find the velocity potential.
5.4. A cylindrical vortex in an incompressible fluid is co-axial with the z-axis, and such that z takes the constant value for
r a, and is zero for r > a, where r is a cylindrical coordinate. Show that
1 d p 2 r
= 4 ,
dr
a
where p(r) is the pressure at radius r inside the vortex, and the circulation of the fluid outside the vortex is 2 . Deduce that
p(r) =
where p0 is the pressure at the center of the vortex.

2 r 2
+ p0 ,
2 a4

Incompressible Inviscid Fluid Dynamics

99

5.5. Consider the cylindrical vortex discussed in Exercise 5.4. If p(r) is the pressure at radius r external to the vortex, demonstrate
that
2
p(r) = 2 + p ,
2r
where p is the pressure at infinity.
5.6. Show that the stream function for the cylindrical vortex discussed in Exercises 5.4 and 5.5 is (r) = (1/2) a2 ln(r/a) for
r > a, and (r) = (1/4) (r2 a2 ) for r a.
5.7. Consider a volume V whose boundary is the surface S . Suppose that V contains an incompressible fluid whose motion is
irrotational. Let the velocity potential be constant over S . Prove that has the same constant value throughout V. [Hint:
Consider the identity (A A) A A + A 2 A.]
5.8. In Exercise 5.7, suppose that, instead of taking a constant value on the boundary, the normal velocity is everywhere zero
on the boundary. Show that is constant throughout V.
5.9. Prove that in the two-dimensional motion of a liquid the mean tangential fluid velocity around any small circle of radius r is
r, where 2 is the value of
vy vx

x
y
at the center of the circle. Neglect terms of order r3 .
5.10. Show that the equation of continuity for the two-dimensional motion of an incompressible fluid can be written
(r vr ) v
+
= 0,
r

where r, are cylindrical coordinates. Demonstrate that this equation is satisfied when vr = a k rn exp[k (n + 1) ] and
v = a rn exp[k (n + 1) ]. Determine the stream function, and show that the fluid speed at any point is

(n + 1) 1 + k2 /r,
where is the stream function at that point (defined such that = 0 at r = 0).
5.11. Demonstrate that streamlines cross at right-angles at a stagnation point in two-dimensional, incompressible, irrotational flow.
5.12. Consider two-dimensional, incompressible, inviscid flow. Demonstrate that the fluid motion is governed by the following
equations:

+ [, ]
t

0,

+ 2 ,

where v = ez , [A, B] = ez A B, and = p/ + (1/2) v2 + .


5.13. For irrotational, incompressible, inviscid motion in two-dimensions show that
q q = q 2 q,
where q = |v|.

100

FLUID MECHANICS

2D Potential Flow

101

6 2D Potential Flow

6.1 Introduction
This chapter discusses the use of complex analysis to simplify calculations in two-dimensional, incompressible, inviscid, irrotational, fluid dynamics. Incidentally, incompressible, inviscid, irrotational flow is usually referred to a
potential flow, since the velocity field can be represented in terms of a velocity potential that satisfies Laplaces equation. In the following, all flow patterns are assumed to be such that the z-coordinate is ignorable. In other words, the
fluid velocity is everywhere parallel to the x-y plane, and /z = 0. It follows that all line sources and vortex filaments
run parallel to the z-axis. Moreover, all solid surfaces are of infinite extent along the z-axis, and have uniform crosssections. Hence, it is only necessary to specify the locations of line sources, vortex filaments, and solid surfaces in the
x-y plane.

6.2 Complex Functions


The complex variable is conventionally written
z = x + i y,

(6.1)

where i represents the square root of minus one. Here, x and y are both real, and are identified with the corresponding
Cartesian coordinates. (Incidentally, z should not be confused with a z-coordinate: this is a strictly two-dimensional
discussion.) We can also write
z = r e i ,
(6.2)
p
1
2
2
where r = x + y and = tan (y/x) are termed the modulus and argument of z, respectively, but can also be
identified with the corresponding plane polar coordinates. Finally, De Moivres theorem,
e i = cos + i sin ,

(6.3)

implies that
x

r cos ,

(6.4)

r sin .

(6.5)

We can define functions of the complex variable, F(z), just like we would define functions of a real variable. For
instance,
F(z) =
F(z) =

z2 ,
1
.
z

(6.6)
(6.7)

For a given function, F(z), we can substitute z = x + i y and write


F(z) = (x, y) + i (x, y),

(6.8)

where (x, y) and (x, y) are real two-dimensional functions. Thus, if


F(z) = z2 ,

(6.9)

F(x + i y) = (x + i y)2 = (x2 y2 ) + 2 i x y,

(6.10)

then
giving

(x, y) =

x2 y 2 ,

(6.11)

(x, y) =

2 x y.

(6.12)

102

FLUID MECHANICS

6.3 Cauchy-Riemann Relations


We can define the derivative of a complex function in just the same manner that we would define the derivative of a
real function: i.e.,
dF
F(z + z) F(z)
= lim |z|
.
(6.13)
dz
z
However, we now have a problem. If F(z) is a well-behaved function (i.e., finite, single-valued, and differentiable)
then it should not matter from which direction in the complex plane we approach the point z when taking the limit in
Equation (6.13). There are, of course, many different possible approach directions, but if we look at a regular complex
function, F(z) = z2 (say), then
dF
= 2z
(6.14)
dz
is perfectly well-defined, and is, therefore, completely independent of the details of how the limit is taken in Equation (6.13).
The fact that Equation (6.13) has to give the same result, no matter from which direction we approach z, means
that there are some restrictions on the forms of the functions (x, y) and (x, y) in Equation (6.8). Suppose that we
approach z along the real axis, so that z = x. We obtain
dF
dz

lim |x|0

(x + x, y) + i (x + x, y) (x, y) i (x, y)

=
+i
.
x
x
x

(6.15)

Suppose that we now approach z along the imaginary axis, so that z = i y. We get
dF
dz

lim |y|0

(x, y + y) + i (x, y + y) (x, y) i (x, y)



= i
+
.
i y
y y

(6.16)

But, if F(z) is a well-behaved function then its derivative must be well-defined, which implies that the above two
expressions are equivalent. This requires that

,
y

.
y

(6.17)
(6.18)

These expressions are called the Cauchy-Riemann relations, and are, in fact, sufficient to ensure that all possible ways
of taking the limit (6.13) give the same result.

6.4 Complex Velocity Potential


Note that Equations (6.17)(6.18) are identical to Equations (5.123)(5.124). This suggests that the real and imaginary parts of a well-behaved function of the complex variable can be interpreted as the velocity potential and stream
function, respectively, of some two-dimensional, irrotational, incompressible flow pattern. For instance, suppose that
F(z) = V0 z,

(6.19)

where V0 is real. It follows that


(r, )

V0 r cos ,

(6.20)

(r, )

V0 r sin .

(6.21)

It can be seen, by comparison with the analysis of Section 5.9, that the complex velocity potential (6.19) corresponds
to uniform flow of speed V0 directed along the x-axis. Furthermore, as is easily demonstrated, the complex velocity
potential associated with uniform flow of speed V0 whose direction subtends a (counter-clockwise) angle 0 with the
x-axis is F(z) = V0 z ei 0 .

2D Potential Flow

103

Suppose that
F(z) =

Q
ln z,
2

(6.22)

where Q is real. Since ln z = ln r + i , it follows that


(r, )

(r, )

Q
ln r,
2
Q
.
2

(6.23)
(6.24)

Thus, according to the analysis of Section 5.10, the complex velocity potential (6.22) corresponds to the flow pattern
of a line source, of strength Q, located at the origin. (See Figure 5.6.) As a simple generalization of this result, the
complex potential of a line source, of strength Q, located at the point (x0 , y0 ), is F(z) = (Q/2) ln(z z0 ), where
z0 = x0 + i y0 . Note, from (6.22), that the complex velocity potential of a line source is singular at the location of the
source.
Suppose that

F(z) = i
ln z,
(6.25)
2
where is real. It follows that
(r, )

(r, )

,
2

ln r.
2

(6.26)
(6.27)

Thus, according to the analysis of Section 5.11, the complex velocity potential (6.25) corresponds to the flow pattern
of a vortex filament of intensity located at the origin. (See Figure 5.8.) Note, from (6.25), that the complex velocity
potential of a vortex filament is singular at the location of the filament.
Suppose, finally, that
!
z

a2
+i
,
(6.28)
ln
F(z) = V0 z +
z
2
a
where V0 , a, and , are real. It follows that
(r, ) =
(r, )

!
a2

V0 r +
cos
,
r
2
!
r

a2
sin +
.
ln
V0 r
r
2
a

(6.29)
(6.30)

Thus, according to the analysis of Section 5.13, the complex velocity potential (6.28) corresponds to uniform inviscid
flow of unperturbed speed V0 , running parallel to the x-axis, around an impenetrable cylinder of radius a, centered on
the origin. (See Figures 5.9, 5.10, and 5.11.) Here, is the circulation of the flow about the cylinder. Note that = 0
on the surface of the cylinder (r = a), which ensures that the normal velocity is zero on this surface, as must be the
case if the cylinder is impenetrable.

6.5 Complex Velocity


It follows from Equations (5.119), (5.122), and (6.15) that
dF

=
+i
= v x + i vy .
dz
x
x
Consequently, dF/dz is termed the complex velocity. Note that
2
dF = v 2 + v 2 = v2 ,
x
y
dz

(6.31)

(6.32)

104

FLUID MECHANICS

where v is the flow speed.


A stagnation point is defined as a point in a flow pattern where the flow speed, v, falls to zero. It follows, from the
previous expression, that
dF
=0
(6.33)
dz
at a stagnation point. For instance, the stagnation points of the flow pattern produced when a cylindrical obstacle
of radius a, centered on the origin, is placed in a uniform flow of speed V0 , directed parallel to the x-axis, and the
circulation of the flow around is cylinder is , are found by setting the derivative of the complex potential (6.28) to
zero. It follows that the stagnation points satisfy the quadratic equation
!
dF

a2
= V0 1 2 + i
= 0.
(6.34)
dz
2 z
z
The solutions are

q
z
= i 1 2 ,
(6.35)
a
where = /(4 V0 a), with the proviso that |z|/a > 1, since the region |z|/a < 1 is occupied
by the cylinder. Thus,
p
2
if 1 then there are two stagnation points on the surface of the cylinder at x/a = 1 and y/a =p. On the
other hand, if > 1 then there is a single stagnation point below the cylinder at x/a = 0 and y/a = 2 1.
Now, according to Section 5.7, Bernoullis theorem in an steady, irrotational, incompressible fluid takes the form
p+

1 2
v = p0 ,
2

(6.36)

where p0 is a uniform constant, and where gravity (or any other body force) has been neglected. Thus, the pressure
distribution in such a fluid can be written

1 dF 2
p = p0 .
(6.37)
2 dz

6.6 Method of Images

Let F1 (z) = 1 (x, y) + i 1 (x, y) and F2 (z) = 2 (x, y) + i 2 (x, y) be complex velocity potentials corresponding to
distinct, two-dimensional, irrotational, incompressible flow patterns whose stream functions are 1 (x, y) and 2 (x, y),
respectively. It follows that both stream functions satisfy Laplaces equation: i.e., 2 1 = 2 2 = 0. Suppose that
F3 (z) = F1 (z) + F2 (z). Writing F3 (z) = 3 (x, y) + i 3 (x, y), it is clear that 3 (x, y) = 1 (x, y) + 2 (x, y). Moreover,
2 3 = 2 1 + 2 2 = 0, so 3 also satisfies Laplaces equation. We deduce that two complex velocity potentials,
corresponding to distinct, two-dimensional, irrotational, incompressible flow patterns, can be superposed to produce
a third velocity potential that corresponds to another two-dimensional, irrotational, incompressible flow pattern. As
described below, this idea can be exploited to determine the flow patterns produced by line sources and vortex filaments
in the vicinity of rigid boundaries.
As an example, consider a situation in which there are two line sources of strength Q located at the points (0, a).
See Figure 6.1. The complex velocity potential of the resulting flow pattern is the sum of the complex potentials of
each source taken in isolation. Hence, from Section 6.4,
F(z) =

Q
Q
Q
ln(z i a)
ln(z + i a) =
ln(z2 + a2 ).
2
2
2

Thus, the stream function of the flow pattern (which is the imaginary part of the complex potential) is
!
2 xy
Q
(x, y) =
tan1 2
.
2
x y2 + a 2

(6.38)

(6.39)

Note that (x, 0) = 0, which implies that there is zero flow normal to the plane y = 0. Hence, in the region y > 0, we
could interpret the above stream function as that generated by a single line source of strength Q, located at the point

2D Potential Flow

105

Q
a
x
a
Q

Figure 6.1: Two line sources.


(0, a), in the presence of a planar rigid boundary at y = 0. This follows because the stream function satisfies 2 = 0
everywhere in the region y > 0, has the requisite singularity (corresponding to a line source of strength Q) at (0, a),
and satisfies the physical boundary condition that the normal velocity be zero at the rigid boundary. Moreover, as is
well-known, the solutions of Poissons equation are unique. The streamlines of the resulting flow pattern are shown in
Figure 6.2. Incidentally, we can think of the two sources in Figure 6.1 as the images of one another in the boundary
plane. Hence, this method of calculation is usually referred to as the method of images.
Now, the complex velocity associated with the complex velocity potential (6.38) is
dF
z
Q
.
=
2
dz
z + a2

(6.40)

Hence, the flow speed at the boundary is



dF
|x|/a
Q
v(x, 0) = (x, 0) =
.
dz
a 1 + x2 /a2

(6.41)

It follows from (6.37) (and the fact that the flow speed at infinity is zero) that the excess pressure on the boundary, due
to the presence of the source, is

dF 2
x2 /a2
 Q 2
.
(6.42)
p(x, 0) = =
2 dz y=0
2 a (1 + x2 /a2 )2

Thus, the excess force per unit length (in the z-direction) acting on the boundary in the y-direction is
Z
Z
Q2
Q2
2
Fy =
p(x, 0) dx = 2
.
d =
2
2
4 a
2 a (1 + )

(6.43)

The fact that the force is positive implies that the boundary is attracted to the source, and vice versa.
As a second example, consider the situation, illustrated in Figure 6.3, in which there are two vortex filaments of
intensities and situated at (0, a). As before, the complex velocity potential of the resulting flow pattern is the
sum of the complex potentials of each filament taken in isolation. Hence, from Section 6.4,
!

z ia
.
(6.44)
ln(z i a) i
ln(z + i a) = i
ln
F(z) = i
2
2
2
z+ ia

106

FLUID MECHANICS

6
5
4
y/a 3
2
1
0
3

0
x/a

Figure 6.2: Stream lines of the 2D flow pattern due to a line source at (0, a) in the presence of a rigid boundary at
y = 0.

a
x
a

Figure 6.3: Two vortex filaments.

2D Potential Flow

107

6
5
4
y/a 3
2
1
0
3

0
x/a

Figure 6.4: Stream lines of the 2D flow pattern due to a vortex filament at (0, a) in the presence of a rigid boundary at
y = 0.
Thus, the stream function of the flow pattern is
#
" 2
x + (y a)2

.
ln 2
(x, y) =
2
x + (y + a)2

(6.45)

As before, (x, 0) = 0, which implies that there is zero flow normal to the plane y = 0. Hence, in the region y > 0,
we could interpret the above stream function as that generated by a single vortex filament of intensity , located at
the point (0, a), in the presence of a planar rigid boundary at y = 0. The streamlines of the resulting flow pattern are
shown in Figure 6.4. We conclude that a vortex filament reverses its sense of rotation (i.e., ) when reflected
in a boundary plane.
As a final example, consider the situation, illustrated in Figure 6.5, in which there is an impenetrable cylinder of
radius a, centered on the origin, and a line source of strength Q located at (b, 0), where b > a. Consider the so-called
analog problem, also illustrated in Figure 6.5, in which the cylinder is replaced by a source of strength Q, located at (c,
0), where c < a, and a source of strength Q, located at the origin. We can think of these two sources as the images
of the external source in the cylinder. Moreover, given that the solutions of Poissons equation are unique, if the analog
problem can be adjusted in such a manner that r = a is a streamline then the flow in the region r > a will become
identical to that in the actual problem. Now, the complex velocity potential in the analog problem is simply
#
"
Q
Q
Q
Q
(z b) (z c)
.
(6.46)
F(z) =
ln(z b)
ln(z c) +
ln z = ln
2
2
2
2
z
Hence, writing z = r e i , the corresponding stream function takes the form
#
"
Q
(r b c/r) sin
1
(r, ) =
.
tan
2
(r + b c/r) cos (b + c)

(6.47)

Now, we require the surface of the cylinder, r = a, to be a streamline: i.e., (a, ) = constant. This is easily achieved
by setting c = a2 /b. Thus, the stream function becomes
#
"
(r/a a/r) sin
Q
.
(6.48)
tan1
(r, ) =
2
(r/a + a/r) cos (b/a + a/b)
The corresponding streamlines in the region external to the cylinder are shown in Figure 6.6.

108

FLUID MECHANICS

Q
a

Q
c

Figure 6.5: A line source in the presence of an impenetrable cylinder.

4
3
2
1
y/a

0
1
2
3
4
4 3 2 1

0
x/a

Figure 6.6: Stream lines of the 2D flow pattern due to a line source at (2a, 0) in the presence of a rigid cylinder of
radius a centered on the origin.

2D Potential Flow

109

d
d
0
x

dz
z0

dz

Figure 6.7: A conformal map.

6.7 Conformal Maps


Let = + i and z = x + i y, where , , x, and y are real. Suppose that = f (z), where f is a well-behaved (i.e.,
single-valued, non-singular, and differentiable) function. We can think of = f (z) as a map from the complex z-plane
to the complex -plane. In other words, every point (x, y) in the complex z-plane maps to a corresponding point (, )
in the complex -plane. Moreover, if f (z) is indeed a well-behaved function then this mapping is unique, and also has
a unique inverse. Suppose that the point z = z0 in the z-plane maps to the point = 0 in the -plane. Let us investigate
how neighboring points map. We have
0 + d

f (z0 + dz ),

(6.49)

0 + d

f (z0 + dz ).

(6.50)

In other words, the points z0 + dz and z0 + dz in the complex z-plane map to the points 0 + d and 0 + d in the
complex -plane, respectively. Now, if |dz |, |dz | 1 then
d
d
where f (z) = d f /dz. Hence,

Thus, it follows that

=
=

f (z0 ) dz ,

f (z0 ) dz ,

(6.51)
(6.52)

d dz
= .
d
dz

(6.53)

|d | |dz |
=
,
|d |
|dz |

(6.54)

arg(d ) arg(d ) = arg(dz ) arg(dz ).

(6.55)

and

Now, we can think of dz and dz as infinitesimal vectors connecting neighboring points in the complex z-plane to
the point z = z0 . Likewise, d and d are infinitesimal vectors connecting the corresponding points in the complex
-plane. It is clear, from the previous two equations, that, in the vicinity of z = z0 , the mapping from the complex
z-plane to the complex -plane is such that the lengths of dz and dz expand or contract by the same factor, and
the angle subtended between these two vectors remains the same. See Figure 6.7. This type of mapping is termed
conformal.

110

FLUID MECHANICS

Suppose that F() = (, ) + i (, ) is a well-behaved function of the complex variable . It follows that 2 =
= = 0. Hence, the functions (, ) and (, ) can be interpreted as the velocity potential and stream
function, respectively, of some two-dimensional, inviscid, incompressible flow pattern, where and are Cartesian
coordinates. However, if = f (z), where f (z) is well-behaved, then F() = F[( f (z)] = G(z) = e
(x, y) + i e
(x, y), where
e = e
e = 0. In other words, the functions e
e(x, y)
G(z) is also well-behaved. It follows that 2e
= 2

(x, y) and
can be interpreted as the velocity potential and stream function, respectively, of some new, two-dimensional, inviscid,
incompressible flow pattern, where x and y are Cartesian coordinates. In other words, we can use a conformal map
to convert a given two-dimensional, inviscid, incompressible flow pattern into another, quite different, pattern. Note,
incidentally, that a conformal map converts a line source into a line source of the same strength, and a vortex filament
into a vortex filament of the same intensity. (See Exercise 6.12.)
As an example, consider the conformal map
= i e z/a .
(6.56)
2

Writing = r e i , it is easily demonstrated that x = a ln r/ and y = a (/ 1/2). Hence, the positive -axis ( = 0)
maps to the line y = a/2, the negative -axis ( = ) maps to the line y = a/2, and the region > 0 (0 )
maps to the region a/2 < y < a/2. Moreover, the points = (0, 1) map to the points z = a (0,1/2 1/2). See
Figure 6.8. As we saw in Section 6.6, in the region > 0, the velocity potential
!
i

(6.57)
ln
F() = i
2
+i
corresponds to the flow pattern generated by a vortex filament of intensity , located at the point = (0, 1), in the
presence of a rigid plane at = 0. Hence,
G(z) = F(i e z/a ) = i

z

,
ln tanh
2
2a

(6.58)

corresponds to the flow pattern generated by a vortex filament of intensity , located at the origin, in the presence of
two rigid planes at y = a/2. This follows because the line = 0 is mapped to the lines y = a/2, the point = (0, 1)
is mapped to the origin, and if the line = 0 corresponds to a streamline then the lines y = a/2 also correspond to
streamlines. The stream function associated with the above complex velocity potential,
"
#

cosh( x a) cos( y/a)


(x, y) = ln
,
(6.59)

cosh( x/a) + cos( y/a)


is shown in Figure 6.9.
As a second example, consider the map
= z2 .

(6.60)

This maps the positive -axis to the positive x-axis, the negative -axis to the positive y-axis, the region > 0 to the
region x > 0, y > 0, and the point = (0, 2 a2 ) to the point z = (a, a). As we saw in Section 6.6, in the region > 0,
the velocity potential
Q
ln( 2 + 4 a4 ),
(6.61)
F() =
2
corresponds to the flow pattern generated by a line source of strength Q, located at the point = (0, 2 a2 ), in the
presence of a rigid plane at = 0. Thus, the complex velocity potential
G(z) = F(z2 ) =

Q
ln(z4 + 4a4 ),
2

(6.62)

corresponds to the flow pattern generated by a line source of strength Q, located at the point z = (a, a), in the presence
of two orthogonal rigid planes at y = 0 and x = 0. The stream function associated with the above complex potential,
"
#
4 x y (x2 y2 )
Q
,
(6.63)
tan1 4
(x, y) =
2
x 6 x2 y 2 + y 4 + 4 a 4

2D Potential Flow

111

C
a

1
A A

B
E

Figure 6.8: The conformal map = i e z/a .

1
0.8
0.6
0.4
0.2
y/a

0
0.2
0.4
0.6
0.8
1
1

0.5

0
x/a

0.5

Figure 6.9: Stream lines of the 2D flow pattern due to a vortex filament at the origin in the presence of two rigid planes
at y = a/2.

112

FLUID MECHANICS

2
y/a
1

x/a
Figure 6.10: Stream lines of the 2D flow pattern due to a line source at (a, a) in the presence of two rigid planes at
x = 0 and y = 0.
is shown in Figure 6.10.
As a final example, consider the map
l2
,

(6.64)

2 l cosh[ln(r/l)] cos ,

(6.65)

2 l sinh[ln(r/l)] sin .

(6.66)

z=+
where l is real and positive. Writing = r e i , we find that
x

y =

Thus, the map converts the circle 2 + 2 = a2 in the -plane, where a > l, into the ellipse
!2
!2
x
y
+
=1
2 l cosh[ln(a/l)]
2 l sinh[ln(a/l)]

(6.67)

in the z-plane. Note that the center of the ellipse lies at the origin, and its major and minor axes run parallel to the xand the y-axes, respectively. As we saw in Section 6.4, in the -plane, the complex velocity potential
!
a2
F = V0 +
,
(6.68)

corresponds to uniform flow of unperturbed speed V0 , running parallel to the -axis, around a circular cylinder of
radius a, centered on the origin. Thus, assuming that a > l, in the z-plane the potential corresponds to uniform flow
of unperturbed speed V0 , running parallel to the x-axis (which follows because at large |z| the map (6.64) reduces to
z = , and so the flow at large distances from the origin is the same in the complex z- and -planes), around an elliptical
cylinder of major radius 2 l cosh[ln(a/l)] = a + l 2 /a, aligned along the x-axis, and minor radius 2 l sinh[ln(a/l)] =
a l 2 /a, aligned along the y-axis. The corresponding stream function in the z-plane is
!
a2
sin ,
(6.69)
(x, y) = V0 r
r

2D Potential Flow

113

4
3
2
1
y/l

0
1
2
3
4
4 3 2 1

0
x/l

Figure 6.11: Stream lines of the 2D flow pattern due to uniform flow parallel to the x-axis around an elliptical cylinder.
where
r

l exp(cosh1 p),

(6.70)

p
y
,
x [p2 1]1/2


 1/2
x2 /l 2 + y2 /l 2 + 4 + [x2 /l 2 + y2 /l 2 + 4]2 16 x2 /l 2 1/2

8
tan1

(6.71)

(6.72)

Figure 6.11 shows the streamlines of the flow pattern calculated for a = 1.5 l.

6.8 Complex Line Integrals


Consider the line integral of some function F(z) of the complex variable taken (counter-clockwise) around a closed
curve C in the complex plane:
I
F(z) dz.
(6.73)
J=
C

Since dz = dx + i dy, and writing F(z) = (x, y) + i (x, y), where (x, y) and (x, y) are real functions, it follows that
J = Jr + i Ji , where
I
( dx dy),
(6.74)
Jr =
C

Ji

( dx + dy).

(6.75)

However, we can also write the above expressions in the two-dimensional vector form
I
A dr,
Jr =
C

(6.76)

114

FLUID MECHANICS
Ji

B dr,

where dr = (dx, dy), A = (, ), and B = (, ). Now, according to Stokes theorem (see Section A.22),
Z
I
( A)z dS ,
A dr =
C

(6.78)

(6.77)

B dr

( B)z dS ,

where S is the region of the x-y plane enclosed by C. Hence, we obtain


!
Z

Jr =
dS ,
+
y
S x
!
Z

dS .

Ji =
y
S x
Let
J

(6.79)

(6.80)
(6.81)

F(z) dz,

(6.82)

F(z) dz,

(6.83)

where C is a closed curve in the complex plane that completely surrounds the smaller curve C. Consider
J = J J .
Writing J = Jr + i Ji , a direct generalization of the previous analysis reveals that
!
Z

dS ,
+
Jr =
y
S x
!
Z

Ji =
dS ,

y
S x

(6.84)

(6.85)
(6.86)

where S is now the region of the x-y plane lying between the curves C and C . Suppose that F(z) is well-behaved (i.e.,
finite, single-valued, and differentiable) throughout S . It immediately follows that its real and imaginary components,
and , respectively, satisfy the Cauchy-Riemann relations, (6.17)(6.18), throughout S . However, if this is the case
then it is apparent, from the previous two expressions, that Jr = Ji = 0. In other words, if F(z) is well-behaved
throughout S then J = J .
The circulation of the flow about some closed curve C in the x-y plane is defined
I
I
dF
dz,
(6.87)
= (v x dx + vy dy) = Re
C dz
C
where F(z) is the complex velocity potential of the flow, and use has been made of Equation (6.31). Thus, the
circulation can be evaluated by performing a line integral in the complex z-plane. Moreover, as is clear from the
previous discussion, this integral can be performed around any loop that can be continuously deformed into the loop
C whilst still remaining in the fluid, and not passing over a singularity of the complex velocity, dF/dz.

6.9 Theorem of Blasius


Consider some flow pattern in the complex z-plane that is specified by the complex velocity potential F(z). Let C be
some closed curve in the complex z-plane. The fluid pressure on this curve is determined from Equation (6.37), which
yields

1 dF 2
P = p0 .
(6.88)
2 dz

2D Potential Flow

115

P dx

dx

P dy

dy

dl

P dl

x
Figure 6.12: Force acting across a short section of a curve.
Let us evaluate the resultant force (per unit length), and the resultant moment (per unit length), acting on the fluid
within the curve as a consequence of this pressure distribution.
Consider a small element of the curve C, lying between x, y and x + dx, y + dy, which is sufficiently short that it
can be approximated as a straight line. Let P be the local fluid pressure on the outer (i.e., exterior to the curve) side
of the element. As illustrated in Figure 6.12, the pressure force (per unit length) acting inward (i.e., toward the inside
of the curve) across the element has a component P dy in the minus x-direction, and a component P dx in the plus
y-direction. Thus, if X and Y are the components of the resultant force (per unit length) in the x- and y-directions,
respectively, then
dX

P dy,

(6.89)

dY

P dx.

(6.90)

The pressure force (per unit length) acting across the element also contributes to a moment (per unit length), M, acting
about the z-axis, where
dM = x dY y dX = P (x dx + y dy).
(6.91)
Thus, the x- and y-components of the resultant force (per unit length) acting on the of the fluid within the curve, as
well as the resultant moment (per unit length) about the z-axis, are given by
I
P dy,
(6.92)
X =
Y

P dx,

(6.93)

P (x dx + y dy),

(6.94)

respectively, where the integrals are taken (counter-clockwise) around the curve C. Finally, given that the pressure
distribution on the curve takes the form (6.88), and that a constant pressure obviously yields zero force and zero
moment, we find that
I 2
1
dF dy,
X =

(6.95)

2
C dz

116

FLUID MECHANICS
I 2
dF dx,

C dz
I 2
1
dF (x dx + y dy).
=
dz
2
1
=
2

Y
M

(6.96)
(6.97)

Now, z = x + i y, and z = x i y, where indicates a complex conjugate. Hence, dz = dx i dy, and i dz = dy + i dx.
It follows that
I 2
1
dF dz.
X iY = i
(6.98)
dz
2
C

However,

2
dF dz = dF d F dz = dF d F,

dz
dz dz
dz

(6.99)

where dF = d + i d and d F = d i d. Suppose that the curve C corresponds to a streamline of the flow, in which
case = constant on C. Thus, d = 0 on C, and so d F = dF. Hence, on C,
!2
dF
dF dF
dF =
dF =
dz,
(6.100)
dz
dz
dz
which implies that
1
X iY = i
2

dF
dz

!2

dz.

(6.101)

This result is known as the Blasius theorem.


Now, x dx + y dy = Re(z dz). Hence,
1
M = Re
2

!
I 2
dF z dz ,

C dz

or, making use of an analogous argument to that employed above,

1 I dF 2
z dz ,
M = Re
2
C dz

(6.102)

(6.103)

Note, finally, that Equations (6.101) and (6.103) hold even when is not constant on the curve C, as long as C can
be continuously deformed into a constant- curve without leaving the fluid or crossing over a singularity of (dF/dz) 2.
As an example of the use of the Blasius theorem, consider again the situation, discussed in Section 6.6, in which
a line source of strength Q is located at (0, a), and there is a rigid boundary at y = 0. As we have seen, the complex
velocity in the region y > 0 takes the form
dF
z
Q
.
(6.104)
=
dz
z2 + a2
Suppose that we evaluate the Blasius integral, (6.103), about the contour C shown in Figure 6.13. This contour runs
along the boundary, and is completed by a semi-circle in the upper half of the z-plane. As is easily demonstrated, in
the limit in which the radius of the semi-circle tends to infinity, the contribution of the curved section of the contour to
the overall integral becomes negligible. In this case, only the straight section of the contour contributes to the integral.
Note that the straight section corresponds to a streamline (since it is coincident with a rigid boundary). In other words,
the contour C corresponds to a streamline at all constituent points that make a finite contribution to the Blasius integral,
which ensures that C is a valid contour for the application of the Blasius theorem. In fact, the Blasius integral specifies
the net force (per unit length) exerted on the whole fluid by the boundary. Note, however, that the contour C can be
deformed into the contour C , which takes the form of a small circle surrounding the source, without passing over a
singularity of (dF/dz) 2 . See Figure 6.13. Hence, we can evaluate the Blasius integral around C without changing its
value. Thus,
!2
I
I
dF
1  Q 2
1
z2
dz = i
dz,
(6.105)
X iY = i
2
2 2
2
2

C dz
C (z + a )

2D Potential Flow

117

y
C

x
C

Figure 6.13: Source in the presence of a rigid boundary.


or

1  Q 2
X iY = i
8

I "
C

#
1
1
2
dz.
+
+
(z i a)2 (z + i a) (z i a) (z + i a)

(6.106)

Writing z = i a + e i , dz = i e i d, and taking the limit 0, we find that


X iY =

i Q2
.
4 a

(6.107)

In other words, the boundary exerts a force (per unit length) F = ( Q 2 /4 a) ey on the fluid. Hence, the fluid
exerts an equal and opposite force F = ( Q 2 /4 a) ey on the boundary. Of course, this result is consistent with
Equation (6.43). Incidentally, it is easily demonstrated from (6.103) that there is zero moment (about the z-axis)
exerted on the boundary by the fluid, and vice versa.
Consider a line source of strength Q placed (at the origin) in a uniformly flowing fluid whose velocity is V =
V0 (cos 0 , sin 0 ). From Section 6.4, the complex velocity potential of the net flow is
F(z) =

Q
ln z V0 z ei 0 .
2

(6.108)

The net force (per unit length) acting on the source (which is calculated by performing the Blasius integral around
a large loop that follows streamlines, and then shrinking the loop to a small circle centered on the source) is (see
Exercise 6.1)
F = Q V.
(6.109)
Note that the force acts in the opposite direction to the flow. Thus, an external force F, acting in the same direction
as the flow, must be applied to the source in order for it to remain stationary. In fact, the above result is valid even in
a non-uniformly flowing fluid, as long as V is interpreted as the fluid velocity at the location of the source (excluding
the velocity field of the source itself).
Finally, consider a vortex filament of intensity placed at the origin in a uniformly flowing fluid whose velocity
is V = V0 (cos 0 , sin 0 ). From Section 6.4, the complex velocity potential of the net flow is
F(z) = i

ln z V0 z ei 0 .
2

(6.110)

The net force (per unit length) acting on the filament (which is calculated by performing the Blasius integral around a
small circle centered on the filament) is (see Exercise 6.2)
F = V ez .

(6.111)

118

FLUID MECHANICS

Note that the force is directed at right-angles to the direction of the flow (in the sense obtained by rotating V through
90 in the opposite direction to the filaments direction of rotation). Again, the above result is valid even in a nonuniformly flowing fluid, as long as V is interpreted as the fluid velocity at the location of the filament (excluding the
velocity field of the filament itself).

6.10 Exercises
6.1. Demonstrate that a line source of strength Q (running along the z-axis) situated in a uniform flow of (unperturbed) velocity
V (lying in the x-y plane) and density experiences a force per unit length
F = Q V.
6.2. Demonstrate that a vortex filament of intensity (running along the z-axis) situated in a uniform flow of (unperturbed)
velocity V (lying in the x-y plane) and density experiences a force per unit length
F = V ez .
6.3. Show that two parallel line sources of strengths Q and Q , located a perpendicular distance r apart, exert a radial force per
unit length Q Q /(2 r) on one another, the force being attractive if Q Q > 0, and repulsive if Q Q < 0.
6.4. Show that two parallel vortex filaments of intensities and , located a perpendicular distance r apart, exert a radial force
per unit length /(2 r) on one another, the force being repulsive if > 0, and attractive if < 0.
6.5. A vortex filament of intensity runs parallel to, and lies a perpendicular distance a from, a rigid planar boundary. Demonstrate that the boundary experiences a net force per unit length 2 /(4 a) directed toward the filament.
6.6. Two rigid planar boundaries meet at right-angles. A line source of strength Q runs parallel to the line of intersection of the
planes, and is situated a perpendicular distance a from each. Demonstrate that the source is subject to a force per unit length

3 2 Q2
8 a
directed towards the line of intersection of the planes.
6.7. A line source of strength Q is located a distance b from an impenetrable cylinder of radius a < b (the axis of the cylinder
being parallel to the source). Demonstrate that the cylinder experiences a net force per unit length
Q2
a2
2
2 b (b a2 )

directed toward the source.


6.8. A dipole line source consists of a line source of strength Q, running parallel to the z-axis, and intersecting the x-y plane at
(d/2) (cos , sin ), and a parallel source of strength Q that intersects the x-y plane at (d/2)( cos , sin ). Show that, in
the limit d 0, and Q d D, the complex velocity potential of the source is
F(z) =

D ei
.
2 z

Here, D e i is termed the complex dipole strength.


6.9. A dipole line source of complex strength D e i is placed in a uniformly flowing fluid of speed V0 whose direction of motion
subtends a (counter-clockwise) angle 0 with the x-axis. Show that, while no net force acts on the source, it is subject to a
moment (per unit length) M = D V0 sin( 0 ) about the z-axis.

6.10. Consider a dipole line source of complex strength D1 e i 1 running along the z-axis, and a second parallel source of complex
strength D2 e i 2 that intersects the x-y plane at (x, 0). Demonstrate that the first source is subject to a moment (per unit
length) about the z-axis of
D1 D2
M=
sin(1 + 2 ),
2 x2
as well as a force (per unit length) whose x- and y-components are
D1 D2
cos(1 + 2 ),
x3
D1 D2
Y =
sin(1 + 2 ),
x3
respectively. Show that the second source is subject to the same moment, but an equal and opposite force.
X

2D Potential Flow

119

6.11. A dipole line source of complex strength D e i runs parallel to, and is located a perpendicular distance a from, a rigid planar
boundary. Show that the boundary experiences a force per unit length
D2
8 a3
acting toward the source.
6.12. Demonstrate that a conformal map converts a line source into a line source of the same strength, and a vortex filament into
a vortex filament of the same intensity.
6.13. Consider the conformal map
z = i c cot(/2),
where z = x + i y, = + i , and c is real and positive. Show that
x

c sinh
,
cosh cos
c sin
.
cosh cos

Demonstrate that = 0 , where 0 0 , maps to a circular arc of center (0, c cot 0 ), and radius c |cosec 0 |, that connects
the points (c, 0), and lies in the region y > 0. Demonstrate that = 0 + maps to the continuation of this arc in the region
y < 0. In particular, show that = 0 maps to the region |x| > c on the x-axis, whereas = maps to the region |x| < c.
Finally, show that = 0 maps to a circle of center (c coth 0 , 0), and radius c |cosech| 0 .
6.14. Consider the complex velocity potential
F(z) =

2 c i V0
cot(/n),
n

where
z = i c cot(/2).
Here, V0 , n, and c are real and positive. Show that

dF
4V0 sin2 (/2)
.
= 2
dz
n sin2 (/n)

Hence, deduce that the flow at |z| is uniform, parallel to the x-axis, and of speed V0 . Show that
(, ) =

2 V0 c
sin(2 /n)
.
n cosh(2 /n) cos(2 /n)

Hence, deduce that the streamline = 0 runs along the x-axis for |x| > c, but along a circular arc connecting the points (c,
0) for |x| < c. Furthermore, show that if 1 < n < 2 then this arc lies above the x-axis, and is of maximum height
"
#
cos( n/2) + 1
h=c
,
sin( n/2)
but if 2 < n < 3 then the arc lies below the x-axis, and is of maximum depth
"
#
cos( n/2) + 1
d=c
.
| sin( n/2)|
Hence, deduce that if 1 < n < 2 then the complex velocity potential under investigation corresponds to uniform flow of
speed V0 , parallel to a planar boundary that possesses a cylindrical bump (whose axis is normal to the flow) of height h and
width 2 c, but if 2 < n < 3 then the potential corresponds to flow parallel to a planar boundary that possesses a cylindrical
depression of depth d and width 2 c. Show, in particular, that if n = 1 then the bump is a half-cylinder, and if n = 3 then the
depression is a half-cylinder. Finally, demonstrate that the flow speed at the top of the bump (in the case 1 < n < 2), or the
bottom of the depression (in the case 2 < n < 3) is
v=

2 V0
[1 cos( n/2)] .
n2

120

FLUID MECHANICS

6.15. Show that z = cosh( /a) maps the semi-infinite strip 0 a, 0 in the -plane onto the upper half (y 0) of the
z-plane. Hence, show that the stream function due to a line source of strength Q placed at = (0, a/2), in the rectangular
region 0 a, 0 bounded by the rigid planes = 0, = 0, and = a, is
(, ) =

Q sinh( /a) sin( /a)


.
2 [sinh2 ( /a) + cos2 ( /a)]

6.16. Show that the complex velocity potential


a V0
tanh(a /z)
can be interpreted as that due to uniform flow of speed V0 over a cylindrical log of radius a lying on the flat bed of a deep
stream (the axis of the log being normal to the flow). Demonstrate that the flow speed at the top of the log is (2 /4) V0 .
Finally, show that the pressure difference between the top and bottom of the log is 4 V02 /32.
F(z) =

Incompressible Boundary Layers

121

7 Incompressible Boundary Layers

7.1 Introduction
Previously, in Section 5.13, we saw that a uniformly flowing incompressible fluid that is modeled as inviscid is incapable of exerting a drag force on a rigid stationary obstacle placed in its path. This result is surprising since, in practice,
a stationary obstacle experiences a significant drag when situated in such a fluid, even in the limit that the Reynolds
number tends to infinity (which corresponds to the inviscid limit). In this chapter, we shall attempt to reconcile these
two results by introducing the concept of a boundary layer. This is a comparatively thin layer that covers the surface of
an obstacle placed in a high Reynolds number incompressible fluidviscosity is assumed to have a significant effect
on the flow inside the layer, but a negligible effect on the flow outside. For the sake of simplicity, we shall restrict our
discussion to the two-dimensional boundary layers that form when a high Reynolds number fluid flows transversely
around a stationary obstacle of infinite length and uniform cross-section.

7.2 No Slip Condition


We saw, in Section 5.13, that when an inviscid fluid flows around a rigid stationary obstacle then the normal fluid
velocity at the surface of the obstacle is required to be zero. However, in general, the tangential velocity is non-zero.
In fact, if the fluid velocity field is both incompressible and irrotational then it is derivable from a stream function that
satisfies Laplaces equation. (See Section 5.8.) It is a well-known property of Laplaces equation that we can either
specify the solution itself, or its normal derivative, on a bounding surface, but we cannot specify both these quantities
simultaneously. Now, the constraint of zero normal velocity is equivalent to the requirement that the stream function
take the constant value zero (say) on the surface of the obstacle. Hence, the normal derivative of the stream function,
which determines the tangential velocity, cannot also be specified at this surface, and is, in general, non-zero.
In reality, all physical fluids possess finite viscosity. Moreover, when a viscous fluid flows around a rigid stationary
obstacle both the normal and the tangential components of the fluid velocity are found to be zero at the obstacles
surface. The additional constraint that the tangential fluid velocity be zero at a rigid stationary boundary is known as
the no slip condition, and is ultimately justified via experimental observations.
The concept of a boundary layer was first introduced into fluid mechanics by Ludwig Prandtl (18751953) in
order to account for the modification to the flow pattern of a high Reynolds number irrotational fluid necessitated by
the imposition of the no slip condition on the surface of an impenetrable stationary obstacle. According to Prandtl, the
boundary layer covers the surface of the obstacle, but is relatively thin in the direction normal to this surface. Outside
the layer, the flow pattern is the same as that of an idealized inviscid fluid, and is thus generally irrotational. This
implies that the normal fluid velocity is zero on the outer edge of the layer, where it interfaces with the irrotational
flow, but, in general, the tangential velocity is non-zero. However, the no slip condition requires the tangential velocity
to be zero on the inner edge of the layer, where it interfaces with the rigid surface. It follows that there is a very
large normal gradient of the tangential velocity across the layer, which implies the presence of intense internal vortex
filaments trapped within the layer. Consequently, the flow within the layer is not irrotational. In the following, we
shall attempt to make the concept of a boundary layer more precise.

7.3 Boundary Layer Equations


Consider a rigid stationary obstacle whose surface is (locally) flat, and corresponds to the x-z plane. Let this surface
be in contact with a high Reynolds number fluid that occupies the region y > 0. See Figure 7.1. Let be the typical
normal thickness of the boundary layer. The layer thus extends over the region 0 < y < . Now, the fluid that occupies
the region < y < , and thus lies outside the layer, is assumed to be both irrotational and (effectively) inviscid. On
the other hand, viscosity must be included in the equation of motion of the fluid within the layer. The fluid both inside
and outside the layer is assumed to be incompressible.

122

FLUID MECHANICS

irrotational fluid

U (x)

boundary layer

x
solid surface
Figure 7.1: A boundary layer.
Suppose that the equations of irrotational flow have already been solved to determine the fluid velocity outside the
boundary layer. This velocity must be such that its normal component is zero at the outer edge of the layer (i.e., y ).
On the other hand, the tangential component of the fluid velocity at the outer edge of the layer, U(x) (say), is generally
non-zero. Here, we are assuming, for the sake of simplicity, that there is no spatial variation in the z-direction, so that
both the irrotational flow and the boundary layer are effectively two-dimensional. Likewise, we are also assuming that
all flows are steady, so that any time variation can be neglected. Now, the motion of the fluid within the boundary layer
is governed by the equations of steady-state, incompressible, two-dimensional, viscous flow, which take the form (see
Section 2.14)
v x vy
+
x
y

0,

vx

v x
v x
+ vy
x
y

vx

vy
vy
+ vy
x
y

(7.1)

!
1 p
2 v x 2 v x
+
,
+
x
x2
y2
!
2 vy 2 vy
1 p

+
,
+
y
x2
y2

(7.2)
(7.3)

where is the (constant) density, and the kinematic viscosity. Here, Equation (7.1) is the equation of continuity,
whereas Equations (7.2) and (7.3) are the x- and y-components of the fluid equation of motion, respectively. The
boundary conditions at the outer edge of the layer, where it interfaces with the irrotational fluid, are
v x (x, y) U(x),

(7.4)

p(x, y)

(7.5)

P(x)

as y/ . Here, P(x) is the fluid pressure at the outer edge of the layer, and
U

dU
1 dP
=
dx
dx

(7.6)

(since vy = 0, and viscosity is negligible, just outside the layer). The boundary conditions at the inner edge of the
layer, where it interfaces with the impenetrable surface, are
v x (x, 0) =

0,

(7.7)

vy (x, 0) =

0.

(7.8)

Incompressible Boundary Layers

123

Of course, the first of these constraints corresponds to the no slip condition.


Let U0 be a typical value of the external tangential velocity, U(x), and let L be the typical variation length-scale of
this quantity. It is reasonable to suppose that U0 and L are also the characteristic tangential flow velocity and variation
length-scale in the x-direction, respectively, of the boundary layer. Of course, is the typical variation length-scale
of the layer in the y-direction. Moreover, /L 1, since the layer is assumed to be thin. It is helpful to define the
normalized variables
x
,
(7.9)
X =
L
y
Y =
,
(7.10)

vx
,
(7.11)
V x (X, Y) =
U0
vy
Vy (X, Y) =
,
(7.12)
U1
p
b
,
(7.13)
P(X, Y) =
p0
where U1 and p0 are constants. All of these variables are designed to be O(1) inside the layer. Equation (7.1) yields
U0 V x U1 Vy
+
= 0.
L X
Y

(7.14)

In order for the terms in this equation to balance one another, we need
U1 =

U0 .
L

(7.15)

In other words, within the layer, continuity requires the typical flow velocity in the y-direction, U1 , to be much smaller
than that in the x-direction, U0 .
Equation (7.2) gives
!
"
#
U02
P  U0   2 2 V x 2 V x
V x
V x
p0 b
+
+ Vy
+
Vx
=
.
(7.16)
L
X
Y
L X
2
L
X 2
Y 2
In order for the pressure term on the right-hand side of the above equation to be of similar magnitude to the advective
terms on the left-hand side, we require that
(7.17)
p0 = U02 .
Furthermore, in order for the viscous term on the right-hand side to balance the other terms, we need
U1
1
=
,
=
L U0 Re1/2

(7.18)

where

U0 L
(7.19)

is the Reynolds number of the flow external to the layer. The assumption that /L 1 can be seen to imply that
Re 1. In other words, the normal thickness of the boundary layer separating an irrotational flow pattern from a rigid
surface is only much less than the typical variation length-scale of the pattern when the Reynolds number of the flow
is much greater than unity.
Equation (7.3) yields
#
!
"
Vy
Vy
1
b
P
1 1 2 Vy 2 Vy
.
(7.20)
+
Vx
=
+ Vy
+
Re
X
Y
Y Re Re X 2
Y 2
Re =

In the limit Re 1, this reduces to

b
P
= 0.
Y

(7.21)

124

FLUID MECHANICS

Hence, b
P=b
P(X), where

b
db
P
b dU ,
= U
dX
dX

(7.22)

b
U(X)
= U/U0 , and use has been made of (7.6). In other words, the pressure is uniform across the layer, in the direction
normal to the surface of the obstacle, and is thus the same as that on the outer edge of the layer.
Retaining only O(1) terms, our final set of normalized layer equations becomes

Vx

V x Vy
+
X
Y

Vy
V x
+ Vy
X
Y

subject to the boundary conditions


and

0,

(7.23)

b 2 Vy
b dU +
U
,
X
Y 2

(7.24)

b
V x (X, ) = U(X),

(7.25)

V x (X, 0) =

0,

(7.26)

Vy (X, 0) =

0.

(7.27)

In unnormalized form, the above set of layer equations are written

vx

v x vy
+
x
y

0,

v x
v x
+ vy
x
y

(7.28)
dU
2 v x
+ 2 ,
dx
y

(7.29)

subject to the boundary conditions


v x (x, ) = U(x)

(7.30)

v x (x, 0) =

0,

(7.31)

vy (x, 0) =

0.

(7.32)

(note that y = really means y/ ), and

Now, Equation (7.28) can be automatically satisfied by expressing the flow velocity in terms of a stream function: i.e.,

,
y

vx

vy

.
x

(7.33)
(7.34)

In this case, Equation (7.29) reduces to

dU
3 2 2

+
=U
,
3
2
y
x y
y x y
dx

(7.35)

subject to the boundary conditions


(x, )
= U(x),
y

(7.36)

and
(x, 0) =
(x, 0)
y

0,

(7.37)

0.

(7.38)

Incompressible Boundary Layers

125

To lowest order, the vorticity internal to the layer, = ez , is given by


=

2
,
y2

(7.39)

whereas the x-component of the viscous force per unit area acting on the surface of the obstacle is written (see Section 2.18)



2
v x
= 2 .
(7.40)
xy y=0 =
y y=0
y y=0

7.4 Self-Similar Boundary Layers

The boundary layer equation, (7.35), takes the form of a nonlinear partial differential equation that is extremely
difficult to solve exactly. However, considerable progress can be made if this equation is converted into an ordinary differential equation by demanding that its solutions be self-similar. Self-similar solutions are such that, at
a given distance, x, along the layer, the tangential flow profile, v x (x, y), is a scaled version of some common profile: i.e., v x (x, y) = U(x) F[y/(x)], where (x) is a scale-factor, and F(z) a dimensionless function. It follows that
(x, y) = U(x) (x) f [y/(x)], where f (z) = F(z).
Let us search for a self-similar solution to Equation (7.35) of the general form
"
#1/2
"
#1/2
2 U0 xm+1
2
(x, y) =
f () = U0 xm
f (),
(7.41)
m+1
(m + 1) U0 xm1
where

"

(m + 1) U0 xm1
=
2

#1/2

y.

(7.42)

This implies that (x) = [2 /(m + 1) U0 xm1 ]1/2 , and U(x) = U0 xm . Here, U0 and m are constants. Moreover, U0 xm
has dimensions of velocity, whereas m, , and f , are dimensionless. Transforming variables from x, y to x, , we find
that





m 1
=
+
,
(7.43)
x y
x
2 x
x




y x

Hence,

"

"

(m + 1) U0 x
2

U0 xm1
2 (m + 1)

#1/2

#
m1 1/2



.
x

(7.44)

U0 xm f ,

(7.46)

2
y2

(7.47)

2
x y

1/2

(m + 1) U03 x3m1
f ,

2

U0 xm1
[2 m f + (m 1) f ],
2

(7.48)

3
y3

(m + 1) U02 x2m1
f ,
2

(7.49)

[(m + 1) f + (m 1) f ],

(7.45)

where = d/d. Thus, Equation (7.35) becomes


(m + 1) f + (m + 1) f f 2m f 2 =

dU 2
1
.
U02 x2m1 dx

(7.50)

126

FLUID MECHANICS

Since the left-hand side of the above equation is a (non-constant) function of , whilst the right-hand side is a function
of x (and since and x are independent variables), the equation can only be satisfied if its right-hand side takes a
constant value. In fact, if
1
dU 2
= 2m
(7.51)
U02 x2m1 dx
then
U(x) = U0 xm

(7.52)

f + f f + (1 f 2 ) = 0,

(7.53)

(which is consistent with our initial guess), and

where

2m
.
(7.54)
m+1
Expression (7.53) is known as the Falkner-Skan equation. The solutions to this equation that satisfy the physical
boundary conditions (7.36)(7.38) are such that
=

f (0) = f (0) = 0,

(7.55)

f () = 1,

(7.56)

(7.57)

and

f () = 0.

(The final condition corresponds to the requirement that the vorticity tend to zero at the edge of the layer.) Note, from
(7.39), (7.42), (7.47), (7.52), (7.55), and (7.56), that the normally integrated vorticity within the boundary layer is
Z
dy = U(x).
(7.58)
0

Furthermore, from (7.40), (7.47), and (7.52), the x-component of the viscous force per unit area acting on the surface
of the obstacle is
 1/2

1
(7.59)
xy y=0 = U 2
(m + 1)1/2 2 f (0).
2
Ux
It is convenient to parameterize this quantity in terms of a skin friction coefficient,

xy y=0
cf =
.
(7.60)
(1/2) U 2
It follows that

(m + 1)1/2 2 f (0)
,
c f (x) =
[Re(x)]1/2

where
Re(x) =

U(x) x

(7.61)

(7.62)

is the effective Reynolds number of the flow on the outer edge of the layer at position x. Hence, c f (x) x(m+1)/2 .
Finally, according to Equation (7.41), the width of the boundary layer is approximately
1
(x)
,

x
[Re(x)] 1/2
which implies that (x) x(m1)/2 .

(7.63)

Incompressible Boundary Layers

127

1.5

f (0)

0.5

0
0

Figure 7.2: f (0) calculated as a function of for solutions of the Falkner-Skan equation.
Note that if m > 0 then the external tangential velocity profile, U(x) = U0 xm , corresponds to that of irrotational
inviscid flow incident, in a symmetric fashion, on a semi-infinite wedge whose apex subtends an angle , where
= 2m/(m + 1). (See Section 5.14, and Figure 5.12.) In this case, U(x) can be interpreted as the tangential velocity
a distance x along the surface of the wedge from its apex (in the direction of the flow). By analogy, if m = 0 then
the external velocity profile corresponds to that of irrotational inviscid flow parallel to a semi-infinite flat plate (which
can be thought of as a wedge whose apex subtends zero angle). In this case, U(x) can be interpreted as the tangential
velocity a distance x along the surface of the plate from its leading edge (in the direction of the flow). (See Section 7.5.)
Finally, if m < 0 then the external velocity profile is that of symmetric irrotational inviscid flow over the back surface
of a semi-infinite wedge whose apex subtends an angle (1 ) , where = m/(1 + m). (See Section 5.15, and
Figure 5.13.) In this case, U(x) can be interpreted as the tangential velocity a distance x along the surface of the wedge
from its apex (in the direction of the flow).
Unfortunately, the Falkner-Skan equation, (7.53), possesses no general analytic solutions. However, this equation
is relatively straightforward to solve via numerical methods. Figure 7.2 shows f (0), calculated numerically as a
function of = 2 m/(m + 1), for the solutions of (7.53) that satisfy the boundary conditions (7.55)(7.57). In addition,
Figure 7.3 shows f () versus , calculated numerically for various different values of m. Note that, since 2 as
m , solutions of the Falkner-Skan equation with > 2 have no physical significance. For 0 < < 2 it can be
seen, from Figures 7.2 and 7.3, that there is a single solution branch characterized by f () > 0 and f (0) > 0. This
branch is termed the forward flow branch, since it is such that the tangential velocity, v x () f (), is in the same
direction as the external tangential velocity [i.e., v x ()] across the whole layer (i.e., 0 < < ). The forward flow
branch is characterized by a positive skin friction coefficient, c f f (0). It can also be seen that for < 0 there
exists a second solution branch, which is termed the reversed flow branch, since it is such that the tangential velocity
is in the opposite direction to the external tangential velocity in the region of the layer immediately adjacent to the
surface of the obstacle (which corresponds to = 0). The reversed flow branch is characterized by a negative skin
friction coefficient. Note that the reversed flow solutions are probably unphysical, since reversed flow close to the wall
is generally associated with a phenomenon known as boundary layer separation (see Section 7.10) which invalidates
the boundary layer orderings. It can be seen that the two solution branches merge together at = = 0.1989, which
corresponds to m = m = 0.0905. Moreover, there are no solutions to the Falkner-Skan equation with < or
m < m . The disappearance of solutions when m becomes too negative (i.e., when the deceleration of the external flow
becomes too large) is also related to boundary layer separation.

128

FLUID MECHANICS

1
0.9
0.8
0.7
0.6
f

0.5
0.4
0.3
0.2
0.1
0
0.1
0

9 10

Figure 7.3: Solutions of the Falkner-Skan equation. In order from the left to the right, the various solid curves
correspond to forward flow solutions calculated with m = 4, 1, 1/3, 1/9, 0, 0.05, and 0.0904, respectively. The
dashed curve shows a reversed flow solution calculated with m = 0.05.

7.5 Boundary Layer on a Flat Plate


Consider a flat plate of length L, infinite width, and negligible thickness, which lies in the x-z plane, and whose two
edges correspond to x = 0 and x = L. Suppose that the plate is immersed in a low viscosity fluid whose unperturbed
velocity field is v = U0 e x . See Figure 7.4. In the inviscid limit, the appropriate boundary condition at the surface of
the plate, vy = 0corresponding to the requirement of zero normal velocityis already satisfied by the unperturbed
flow. Hence, the original flow is not modified by the presence of the plate. However, when we take the finite viscosity
of the fluid into account, an additional boundary condition, v x = 0corresponding to the no slip conditionmust be
satisfied at the plate. The imposition of this additional constraint causes thin boundary layers, of thickness (x) L, to
form above and below the plate. The fluid flow outside the boundary layers remains effectively inviscid, whereas that
inside the layers is modified by viscosity. It follows that the flow external to the layers is unaffected by the presence
of the plate. Hence, the tangential velocity at the outer edge of the boundary layers is U(x) = U0 . This corresponds to
the case m = 0 discussed in the previous sectionsee Equation (7.52). (Here, we are assuming that the flow upstream
of the trailing edge of the plate, x = L, is unaffected by the edges presence, and, is, therefore, the same as if the plate
were of infinite length. Of course, the flow downstream of the edge is modified as a consequence of the finite length
of the plate.)
Making use of the analysis contained in the previous section (with m = 0), as well as the fact that, by symmetry,
the lower boundary layer is the mirror image of the upper one, the tangential velocity profile across the both layers is
written
v x (x, y) = U0 f (),
(7.64)
where
=

 U 1/2
0
|y|.
2x

(7.65)

Here, f () is the solution of


f + f f = 0

(7.66)

Incompressible Boundary Layers

129

y
boundary layer
plate

U0

x
L

wake
Figure 7.4: Flow over a flat plate.
that satisfies the boundary conditions
f (0) = f (0) = 0,

(7.67)

f () = 1,

(7.68)

(7.69)

and

f () = 0.
Equation (7.66) is known as the Blasius equation.
It is convenient to define the so-called displacement thickness of the upper boundary layer,
#
Z "
v x (x, y)
dy,
1
(x) =
U0
0

(7.70)

which can be interpreted as the distance through which streamlines just outside the layer are displaced laterally due to
the retardation of the flow within the layer. (Of course, the thickness of the lower boundary layer is the same as that of
the upper layer.) It follows that
!1/2
Z
x
(x) =
[1 f ()] d.
(7.71)
2
U0
0
In fact, the numerical solution of (7.66), subject to the boundary conditions (7.67)(7.69), yields
x
(x) = 1.72
U0

!1/2

(7.72)

Hence, the thickness of the boundary layer increases as the square root of the distance from the leading edge of the
plate. In particular, the thickness at the trailing edge of the plate is
(L)
1.72
,
=
L
Re1/2
where

(7.73)

U0 L
(7.74)

is the appropriate Reynolds number for the interaction of the flow with the plate. Note that if Re 1 then the thickness
of the boundary layer is much less than its length, as was previously assumed.
Re =

130

FLUID MECHANICS

1
0.9
0.8

vx / U0

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
5 4 3 2 1 0 1
y/

Figure 7.5: Tangential velocity profile across the boundary layers located above and below a flat plate of negligible
thickness located at y = 0.
The tangential velocity profile across the both boundary layers, which takes the form
#
"
|y|
,
v x (x, y) = U0 f 1.22
(x)
is plotted in Figure 7.5. In addition, the vorticity profile across the layers, which is written
#
"
U0
|y|
(x, y) = sgn(y) 1.22
,
f 1.22
(x)
(x)

(7.75)

(7.76)

is shown in Figure 7.6. Note that the vorticity is negative in the upper boundary layer (i.e., y > 0), positive in the lower
boundary layer (i.e., y < 0), and discontinuous across the plate (which is located at y = 0). Finally, the net viscous
drag force per unit width (along the z-axis) acting on the plate in the x-direction is
Z L

(7.77)
xy y=0 dx,
D=2
0

where the factor of 2 is needed to take into account the presence of boundary layers both above and below the plate. It
follows from Equation (7.59) (with m = 0) that
D=

U02

U0

!1/2

Z
2 f (0)

1/2

dx =

U02

L
U0

!1/2

2 2 f (0).

(7.78)

In fact, the numerical solution of (7.66) yields


D = 1.33

U02 L
Re1/2

= 1.33 U0 ( U0 L)1/2 .

(7.79)

The above discussion is premised on the assumption that the flow in the upper (or lower) boundary layer is both
steady and z-independent. It turns out that this assumption becomes invalid when the Reynolds number of the layer,

/ (U0/)

Incompressible Boundary Layers

131

0.6
0.5
0.4
0.3
0.2
0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
5 4 3 2 1 0 1
y/

Figure 7.6: Vorticity profile across the boundary layers located above and below a flat plate of negligible thickness
located at y = 0.
U0 /, exceeds a critical value which is about 600. In this case, small-scale z-dependent disturbances spontaneously
grow to large amplitude, and the layer becomes turbulent. Since x1/2 , if the criterion for boundary layer turbulence
is not satisfied at the trailing edge of the plate, x = L, then it is not satisfied anywhere else in the layer. Thus, the
previous analysis, which neglects turbulence, remains valid provided U0 (L)/ < 600. According to (7.73), this
implies that the analysis is valid when 1 Re < 1.2 105, where Re = U0 L/ is the Reynolds number of the external
flow.
Consider, finally, the situation illustrated in Figure 7.7 in which an initially irrotational fluid passes between two
flat parallel plates. Let d be the perpendicular distance between the plates. As we have seen, the finite viscosity of the
fluid causes boundary layers to form on the inner surfaces of the upper and lower plates. The flow within these layers
possesses non-zero vorticity, and is significantly affected by viscosity. On the other hand, the flow outside the layers
is irrotational and essentially inviscidthis type of flow is usually termed potential flow (since it can be derived from
a velocity potential satisfying Poissons equation). Now, the thickness of the two boundary layers increases like x1/2 ,
where x represents distance, parallel to the flow, measured from the leading edges of the plates. It follows that, as x
increases, the region of potential flow shrinks in size, and eventually disappears. See Figure 7.7. Assuming that, prior
to merging, the two boundary layers do not significantly affect one another, their thickness, (x), is given by formula
(7.72), where U0 is the speed of the incident fluid. The region of potential flow thus extends from x = 0 (which
corresponds to the leading edge of the plates) to x = l, where
(l) =
It follows that

d
.
2

l
= 11.8 Re,
d

where

(7.80)

(7.81)

U0 d
.
(7.82)

Thus, when an irrotational high Reynolds number fluid passes between two parallel plates then the region of potential
flow extends a comparatively long distance between the plates, relative to their spacing (i.e., l/d 1). By analogy, if
Re =

132

FLUID MECHANICS

boundary layer

plate

U0

potential flow

Figure 7.7: Flow between two flat parallel plates.


an irrotational high Reynolds number fluid passes into a pipe then the fluid remains essentially irrotational until it has
travelled a considerable distance along the pipe, compared to its diameter. Obviously, these conclusions are modified
if the flow becomes turbulent.

7.6 Wake Downstream of a Flat Plate


As we saw in the previous section, if a flat plate of negligible thickness, and finite length, is placed in the path of a
uniform high Reynolds number flow, directed parallel to the plate, then thin boundary layers form above and below
the plate. Outside the layers, the flow is irrotational, and essentially inviscid. Inside the layers, the flow is modified
by viscosity, and has non-zero vorticity. Downstream of the plate, the boundary layers are convected by the flow, and
merge to form a thin wake. See Figure 7.4. Within the wake, the flow is modified by viscosity, and possesses finite
vorticity. Outside the wake, the downstream flow remains irrotational, and effectively inviscid.
Since there is no solid surface embedded in the wake, acting to retard the flow, we would expect the action of
viscosity to cause the velocity within the wake, a long distance downstream of the plate, to closely match that of the
unperturbed flow. In other words, we expect the fluid velocity within the wake to take the form
v x (x, y) =

U0 u(x, y),

(7.83)

vy (x, y) =

v(x, y),

(7.84)

where
|u| U0 .

(7.85)

Assuming that, within the wake,

1
,
x
1
,

(7.86)
(7.87)

where x is the wake thickness, fluid continuity requires that


v

u.
x

(7.88)

Now, the flow external to the boundary layers, and the wake, is both uniform and essentially inviscid. Hence, according
to Bernoullis theorem, the pressure in this region is also uniformsee Equation (7.22). However, as we saw in

Incompressible Boundary Layers

133

Section 7.3, there is no y-variation of the pressure across the boundary layers. It follows that the pressure is uniform
within the layers. Thus, it is reasonable to assume that the pressure is also uniform within the wake, since the wake is
formed via the convection of the boundary layers downstream of the plate. We conclude that
p(x, y) p0

(7.89)

everywhere in the fluid, where p0 is a constant.


The x-component of the fluid equation of motion is written
!
v x
v x
1 p
2 v x 2 v x
vx
+ 2 .
+ vy
=
+
x
y
x
x2
y

(7.90)

Making use of (7.83)(7.89), the above expression reduces to


U0

u
2 u
2.
x
y

(7.91)

The boundary condition


u(x, ) = 0

(7.92)

ensures that the flow outside the wake remains unperturbed. Note that Equation (7.91) has the same mathematical
form as a conventional diffusion equation, with x playing the role of time, and /U0 playing the role of the diffusion
coefficient. Hence, by analogy with the standard solution of the diffusion equation, we would expect ( x/U0 )1/2 .
As can easily be demonstrated, the self-similar solution to (7.91), subject to the boundary condition (7.92), is
!
y2
Q
(7.93)
exp 2 ,
u(x, y) =

where
x
(x) = 2
U0
and Q is a constant. It follows that

!1/2

u dy = Q,

(7.94)

(7.95)

since, as is well-known, exp(t2 ) dt = . As expected, the width of the wake scales as x1/2 .
The tangential velocity profile across the wake, which takes the form
v x (x, y)
Q 1
=1
exp(y2 / 2 ),
U0
U0

(7.96)

is plotted in Figure 7.8. In addition, the vorticity profile across the wake, which is written
(x, y)
Q 2 y
=
exp(y2 / 2 )

U0 /
U0

(7.97)

is shown in Figure 7.9. It can be seen that the profiles pictured in Figures 7.8 and 7.9 are essentially smoothed out
versions of the boundary layer profiles shown in Figures 7.5 and 7.6, respectively.
Suppose that the plate and a portion of its trailing wake are enclosed by a cuboid control volume of unit depth (in
the z-direction) that extends from x = l to x = +l and from y = h to y = h. See Figure 7.10. Here, l L and
h (l), where L is the length of the plate, and (x) the width of the wake. Hence, the control volume extends well
upstream and downstream of the plate. Moreover, the volume is much wider than the wake.
Let us apply the integral form of the fluid equation of continuity to the control volume. For a steady-state, this
reduces to (see Section 2.9)
I
v dS = 0,
(7.98)
S

134

FLUID MECHANICS

1
0.9
0.8

vx / U0

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
3

0
y/

Figure 7.8: Tangential velocity profile across the wake of a flat plate of negligible thickness located at y = 0. The
profile is calculated for Q/(U0 ) = 0.5.

0.5
0.4
0.3
/ (U0/)

0.2
0.1
0
0.1
0.2
0.3
0.4
0.5
3

0
y/

Figure 7.9: Vorticity profile across the boundary layers above and below a flat plate of negligible thickness located at
y = 0. The profile is calculated for Q/(U0 ) = 0.5.

Incompressible Boundary Layers

135

v(x)

y
y=h

plate

wake
U0 u(y)

U0

y = h
x=l

x = l
v(x)

Figure 7.10: Control volume surrounding a flat plate and its trailing wake.
where S is the bounding surface of the control volume. The normal fluid velocity is U0 at x = l, U0 u(y) at x = l,
and v(x) at y = h, as indicated in the figure. Hence, (7.98) yields

U0 dy +

or

[U0 u(y)] dy + 2

u(y) dy = 2

v(x) dx = 0,

(7.99)

v(x) dx.

(7.100)

However, given that u 0 for |y| , and since h , it is a good approximation to replace the limits of integration
on the left-hand side of the above expression by . Thus, from Equation (7.95),
Z

u(y) dy = 2
h

l
l

v(x) dx Q,

(7.101)

where Q is independent of x. Note that the slight retardation of the flow inside the wake, due to the presence of the
plate, which is parameterized by Q, necessitates a small lateral outflow, v(x), in the region of the fluid external to the
wake.
Let us now apply the integral form of the x-component of the fluid equation of motion to the control volume. For
a steady-state, this reduces to (see Section 2.11)
Z
Z
x j dS j ,
(7.102)
v x v dS = F x +
S

where F x is the net x-directed force exerted on the fluid within the control volume by the plate. It follows, from
Newtons third law of motion, that F x = D, where D is the viscous drag force per unit width (in the z-direction)
acting on the plate in the x-direction. Now, in an incompressible fluid (see Section 2.6),
!
vi v j
i j = p i j +
+
.
(7.103)
x j xi

136

FLUID MECHANICS

Hence, we obtain

U02 dy +



U0 u(y) 2 dy

+2

U0 v(x) dx =

d
D 2
dl

u(y) dy,

(7.104)

since the pressure within the fluid is essentially uniform, the tangential fluid velocity at y = h is U0 , and v is assumed
to be negligible at x = l. Making use of Equation (7.101), as well as the fact that Q is independent of l, we get
D = U0 Q.

(7.105)

Here, we have neglected any terms that are second-order in the small quantity u. A comparison with Equation (7.79)
reveals that
Q = 1.33 ( U0 L)1/2 ,
(7.106)
or

 L 1/2
Q
= 0.664
.
U0
x

(7.107)

Hence, from (7.96) and (7.97), the velocity and vorticity profiles across the layer are

and

 L 1/2
v x (x, y)
= 1 0.375
exp(y2 / 2 ),
U0
x

(7.108)

 L 1/2 y
(x, y)
= 0.749
exp(y2 / 2 ),
U0 /
x

(7.109)

where (x) = 2 ( x/U0 )1/2 . Finally, since the above analysis is premised on the assumption that |1 v x/U0 | = |u|/U0
1, it is clear that the previous three expressions are only valid when x L (i.e., well downstream of the plate).
The above analysis only holds when the flow within the wake is non-turbulent. Let us assume, by analogy with
the discussion in the previous section, that this is the case as long as the Reynolds number of the wake, U0 (x)/,
remains less than some critical value that is approximately 600. Since the Reynolds number of the wake can be written
2 Re1/2 (x/L)1/2 , where Re = U0 L/ is the Reynolds number of the external flow, we deduce that the wake becomes
turbulent when x/L > 9 104 /Re. Hence, the wake is always turbulent sufficiently far downstream of the plate. Our
analysis, which effectively assumes that the wake is non-turbulent in some region, immediately downstream of the
plate, whose extent (in x) is large compared with L, is thus only valid when 1 Re 9 104 .

7.7 Von Karman Momentum Integral


Consider a boundary layer that forms on the surface of a rigid stationary obstacle of arbitrary shape (but infinite length
and uniform cross-section) placed in a steady, uniform, transverse, high Reynolds number flow. Let x represent arc
length along the surface, measured (in the direction of the external flow) from the stagnation point that forms at the
front of the obstacle. (See Figure 7.11.) Moreover, let y represent distance across the boundary layer, measured
normal to the surface. Suppose that the boundary layer is sufficiently thin that it is well approximated as a plane slab
in the immediate vicinity of a general point on the surface. In this case, writing the velocity field within the layer in
the form v = u(x, y) e x + v(x, y) ey, it is reasonable to model this flow using the slab boundary layer equations [see
Equations (7.28) and (7.29)]

u v
+
x y

0,

u
dU
u
+v
U
x
y
dx

2 u
,
y2

(7.110)
(7.111)

Incompressible Boundary Layers

137

subject to the standard boundary conditions


u(x, ) =
u(x, 0) = v(x, 0) =

U(x),

(7.112)

0.

(7.113)

Here, U(x) is the external tangential fluid velocity at the edge of the layer. Integrating (7.111) across the layer, making
use of the boundary conditions (7.113), leads to
!
Z

dU
u
u
u

dy
U
u
v
=

y y=0
dx
x
y
0
#
Z "
dU
(U u)
(U u)
=
(U u)
+u
+v
dy
dx
x
y
0
#
Z "
dU
(U u)
v
(U u)
=
dy
+u
(U u)
dx
x
y
0
#
Z "
(U u)
u
dU
dy
+u
+ (U u)
(U u)
=
dx
x
x
0
Z
Z
d
dU
(U u) dy +
u (U u) dy.
(7.114)
=
dx 0
dx 0
Here, we have integrated the final term on the right-hand side by parts, making use of Equations (7.110), (7.112), and
(7.113). Let us define the displacement thickness of the layer [see Equation (7.70)]
Z 
u
1 (x) =
dy,
(7.115)
1
U
0
as well as the so-called momentum thickness
2 (x) =

It follows from (7.114) that

u
u 
1
dy.
U
U


u
dU
d
= U2 2 +U

(1 + 2 2 ).
y y=0
dx
dx

(7.116)

(7.117)

This important result is known as the von Karman momentum integral, and is fundamental to many of the approximation methods commonly employed to calculate boundary layer thicknesses on the surfaces of general obstacles placed
in high Reynolds number flows. (See Section 7.10.)

7.8 Boundary Layer Separation


As we saw in Section 7.5, when a high Reynolds number fluid passes around a streamlined obstacle, such as a slender
plate that is aligned with the flow, a relatively thin boundary layer form on the obstacles surface. Here, by relatively
thin, we mean that the typical transverse (to the flow) thickness of the layer is L/Re1/2 , where L is the length
of the obstacle (in the direction of the flow), and Re the Reynolds number of the external flow. Suppose, however,
that the obstacle is not streamlined: i.e., the surface of the obstacle is not closely aligned with the streamlines of the
unperturbed flow pattern. In this case, the typically observed behavior is illustrated in Figure 7.11, which shows the
flow pattern of a high Reynolds number irrotational fluid around a cylindrical obstacle (whose axis is normal to the
direction of the unperturbed flow). It can be seen that a stagnation point, at which the flow velocity is locally zero,
forms in front of the obstacle. Moreover, a thin boundary layer covers the front side of the obstacle. The thickness
of this layer is smallest at the stagnation point, and increases towards the back side of the obstacle. However, at
some point on the back side, the boundary layer separates from the obstacles surface to form a vortex-filled wake
whose transverse dimensions are similar to those of the obstacle itself. This phenomenon is known as boundary layer
separation.

138

FLUID MECHANICS

separation point
potential flow streamlines

boundary layer
obstacle

wake

stagnation point

Figure 7.11: Boundary layer separation.


Outside the boundary layer, and the wake, the flow pattern is irrotational and essentially inviscid. So, from Section 5.13, the tangential flow speed just outside the boundary layer (neglecting any circulation of the external flow
around the cylinder) is
U() = 2 U0 sin ,
(7.118)
where U0 is the unperturbed flow speed, and is a cylindrical coordinate defined such that the stagnation point
corresponds to = 0. Note that the tangential flow accelerates (i.e., increases with increasing arc-length, along the
surface of the obstacle, in the direction of the flow) on the front side of the obstacle (i.e., 0 /2), and decelerates
on the back side. Boundary layer separation is always observed to take place at a point on the surface of an obstacle
where there is deceleration of the external tangential flow. In addition, from Section 5.13, the pressure just outside the
boundary layer (and, hence, on the surface of the obstacle, since the pressure is uniform across the layer) is
P() = p1 + U02 cos 2,

(7.119)

where p1 is a constant. Note that the tangential pressure gradient is such as to accelerate the tangential flow on the front
side of the obstaclethis is known as a favorable pressure gradient. On the other hand, the pressure gradient is such
as to decelerate the flow on the back sidethis is known as an adverse pressure gradient. Boundary layer separation
is always observed to take place at a point on the surface of an obstacle where the pressure gradient is adverse.
Boundary layer separation is an important physical phenomenon because it gives rise to a greatly enhanced drag
force acting on a non-streamlined obstacle placed in a high Reynolds number flow. This is the case because the pressure
in the comparatively wide wake that forms behind a non-streamlined obstacle, as a consequence of separation, is
relatively low. To be more exact, in the case of a cylindrical obstacle, Equation (7.119) specifies the expected pressure
variation over the obstacles surface in the absence of separation. It can be seen that the variation on the front side of
the obstacle mirrors that on the back side: i.e., P() = P(). (See Figure 7.12.) In other words, the resultant pressure
force on the front side of the obstacle is equal and opposite to that on the back side, so that the pressure distribution
gives rise to zero net drag acting on the obstacle. Figure 7.12 illustrates how the pressure distribution is modified as
a consequence of boundary layer separation. In this case, the pressure between the separation points is significantly
less than that on the front side of the obstacle. Consequently, the resultant pressure force on the front side is greater in
magnitude than the oppositely directed force on the back side, giving rise to a significant drag acting on the obstacle.
Let D be the drag force per unit width (parallel to the axis of the cylinder) exerted on the obstacle. It is convenient to

Incompressible Boundary Layers

139

/2

3/2

P () p1

stagnation point

separation points
Figure 7.12: Pressure variation over surface of a cylindrical obstacle in a high Reynolds number flow both with
(dashed curve) and without (solid curve) boundary layer separation.
parameterize this force in terms of a dimensionless drag coefficient,
CD =

D
,
U02 a

(7.120)

where is the fluid density, and a the typical transverse size of the obstacle (in the present example, the radius of
the cylinder). The drag force that acts on a non-streamlined obstacle placed in a high Reynolds number flow, as a
consequence of boundary layer separation, is generally characterized by a drag coefficient of order unity. The exact
value of the coefficient depends strongly on the shape of the obstacle, but only relatively weakly on the Reynolds
number of the flow. Consequently, this type of drag is termed form drag, since it depends primarily on the external
shape, or form, of the obstacle. Form drag scales roughly as the cross-sectional area (per unit width) of the vortex-filled
wake that forms behind the obstacle.
Boundary layer separation is associated with strong adverse pressure gradients, or, equivalently, strong flow deceleration, on the back side of an obstacle placed in a high Reynolds number flow. Such gradients can be significantly
reduced by streamlining the obstacle: i.e., by closely aligning its back surface with the unperturbed streamlines of
the external flow. Indeed, boundary layer separation can be delayed, or even completely prevented, on the surface
of a sufficiently streamlined obstacle, thereby significantly decreasing, or even eliminating, the associated form drag
(essentially, by reducing the cross-sectional area of the wake). However, even in the limit that the form drag is reduced
to a negligible level, there is still a residual drag acting on the obstacle due to boundary layer viscosity. This type
of drag is called friction drag. As is clear from a comparison of Equations (7.79) and (7.120), the drag coefficient
associated with friction drag is O(Re1/2 ), where Re is the Reynolds number of the flow. Friction drag thus tends to
zero as the Reynolds number tends to infinity.
The phenomenon of boundary layer separation allows us to resolve dAlemberts paradox. Recall, from Section 5.13, that an idealized fluid that is modeled as inviscid and irrotational is incapable of exerting a drag force on
a stationary obstacle, despite the fact that very high Reynolds number, ostensibly irrotational, fluids are observed to
exert significant drag forces on stationary obstacles. The resolution of the paradox lies in the realization that, in such
fluids, viscosity can only be neglected (and the flow is consequently only irrotational) in the absence of boundary layer

140

FLUID MECHANICS

separation. In this case, the region of the fluid in which viscosity plays a significant role is localized to a thin boundary
layer on the surface of the obstacle, and the resultant friction drag scales as Re1/2 , and, therefore, disappears in the
inviscid limit (essentially, because the boundary layer shrinks to zero thickness in this limit). On the other hand, if the
boundary layer separates then viscosity is important both in a thin boundary layer on the front of the obstacle, and in a
wide, low-pressure, vortex-filled, wake that forms behind the obstacle. Moreover, the wake does not disappear in the
inviscid limit. The presence of significant fluid vorticity within the wake invalidates irrotational fluid dynamics. Consequently, the pressure on the back side of the obstacle is significantly smaller than that predicted by irrotational fluid
dynamics. Hence, the resultant pressure force on the front side is larger than that on the back side, and a significant
drag is exerted on the obstacle. The drag coefficient associated with this type of drag is generally of order unity, and
does not tend to zero as the Reynolds number tends to infinity.

7.9 Criterion for Boundary Layer Separation


As we have seen, the boundary layer equations (7.110)(7.113) generally lead to the conclusion that the tangential
velocity in a thin boundary layer, u, is large compared with the normal velocity, v. Mathematically speaking, this
result holds everywhere except in the immediate vicinity of singular points. But, if v u then it follows that the
fluid moves predominately parallel to the surface of the obstacle, and can only move away from this surface to a very
limited extent. This restriction effectively precludes separation of the flow from the surface. Hence, we conclude that
separation can only occur at a point at which the solution of the boundary layer equations is singular.
As we approach a separation point, we expect the flow to deviate from the boundary layer towards the interior of
the fluid. In other words, we expect the normal velocity to become comparable with the tangential velocity. However,
we have seen that the ratio v/u is of order Re1/2 [see Equation (7.18)]. Hence, an increase of v to such a degree that
v u implies an increase by a factor Re1/2 . For sufficiently large Reynolds numbers, we may suppose that v effectively
increases by an infinite factor. Indeed, if we employ the dimensionless form of the boundary layer equations, (7.23)
(7.27), the situation just described is formally equivalent to an infinite value of the dimensionless normal velocity, Vy ,
at the separation point.
Let the separation point lie at x = x0 , and let x < x0 correspond to the region of the boundary layer upstream of
this point. According to the above discussion,
v(x0 , y) =
(7.121)
at all y (except, of course, y = 0, where the boundary conditions at the surface of the obstacle require that v = 0). It
follows that the deriviative v/y is also infinite at x = x0 . Hence, the equation of continuity, u/x + v/y = 0,
implies that (u/x) x=x0 = , or x/u = 0, if x is regarded as a function of u and y. Let u(x0 , y) = u0 (y). Close to the
point of separation, x0 x and u u0 are small. Thus, we can expand x0 x in powers of u u0 (at fixed y). Since
(x/u)u=u0 = 0, the first term in this expansion vanishes identically, and we are left with
i
h
(7.122)
x0 x = f (y) (u u0 )2 + O (u u0 )3 ,
or
where = 1/

u(x, y) u0 (y) + (y)

x0 x,

(7.123)

f is some function of y. From the equation of continuity,


v
u
(y)
.
=

y
x 2 x0 x

(7.124)

(y)
,
v(x, y)
x0 x

(7.125)

Upon integration, the above expression yields

where
(y) =

1
2

(y ) dy .

(7.126)

Incompressible Boundary Layers

141

The equation of tangential motion in the boundary layer, (7.111), is written


u

u
dU
2 u
u
+v
=U
+ 2.
x
y
dx
y

(7.127)

As is clear from Equation (7.123), the derivative 2 u/y2 does not become infinite as x x0 . The same is true of
the function U dU/dx, which is determined from the flow outside the boundary layer. However, both terms on the
left-hand side of the above expression become infinite as x x0 . Hence, in the immediate vicinity of the separation
point,
u
u
u
+v
0.
(7.128)
x
y
Since u/x = v/y, we can rewrite this equation in the form
u

v
u
v
0.
+v
= u2
y
y
y u

Since u does not, in general, vanish at x = x0 , we conclude that


v
0.
y u

(7.129)

(7.130)

In other words, v/u is a function of x only. Now, from (7.123) and (7.125),
(y)
v
=
+ O(1).

u u0 (y) x0 x

(7.131)

Hence, if this ratio is a function of x alone then (y) = (1/2) A u0(y), where A is a constant: i.e.,
v(x, y)

A u0 (y)
.

x0 x

(7.132)

Finally, since (7.126) yields = 2 d/dy = A du0 /dy, we obtain


u(x, y) u0 (y) + A

du0
x0 x.
dy

(7.133)

The previous two expressions specify u and v as functions of x and y near the point of separation. Beyond the point
of separation, that is for x > x0 , the expressions are physically meaningless, since the square roots become imaginary.
This implies that the solutions of the boundary layer equations cannot sensibly be continued beyond the separation
point.
Now, the standard boundary conditions at the surface of the obstacle require that u = v = 0 at y = 0. It, therefore,
follows from Equations (7.132) and (7.133) that
u0 (0) =

du0

=
dy y=0

0,

(7.134)

0.

(7.135)

Thus, we obtain the important prediction that both the tangential velocity, u, and its first derivative, u/y, are zero at
the separation point (i.e., x = x0 and y = 0). This result was originally obtained by Prandtl, although the argument we
have used to derive it is due to L.D. Landau.
Note that if the constant A in expressions (7.132) and (7.133) happens to be zero then the point x = x0 and y = 0,
at which the derivative u/y vanishes, has no particular properties, and is not a point of separation. However, there is
no reason, in general, why A should take the special value zero. Thus, in practice, a point on the surface of an obstacle
at which u/y = 0 is always a point of separation.
Incidentally, if there were no separation at the point x = x0 (i.e., if A = 0) then we would have u/y < 0 for
x > x0 . In other words, u would become negative as we move away from the surface, y being still small. That is, the

142

FLUID MECHANICS

fluid beyond the point x = x0 would move tangentially, in the region of the boundary layer immediately adjacent to
the surface, in the direction opposite to that of the external flow: i.e., there would be back-flow in this region. In
practice, the flow separates from the surface at x = x0 , and the back-flow migrates into the wake.
Note that the dimensionless boundary layer equations, (7.23)(7.27), are independent of the Reynolds number of
the external flow (assuming that this number is much greater than unity). Thus, it follows that the point on the surface
of the obstacle at which u/y = 0 is also independent of the Reynolds number. In other words, the location of the
separation point is independent of the Reynolds number (as long as this number is large, and the flow in the boundary
layer is non-turbulent).
At y = 0, the equation of tangential motion in the boundary layer, (7.111), is written

2 u
1 dP
1 dU
2 =
=
,
(7.136)

y y=0
U dx
dx
where P(x) is the pressure just outside the layer, and use has been made of (7.6). Now, since u is positive, and increases
away from the surface (upstream of the separation point), it follows that (2 u/y2)y=0 > 0 at the separation point itself,
where (u/y)y=0 = 0. Hence, according to the above equation,
!
dU
< 0,
(7.137)
dx x=x0
!
dP
> 0.
(7.138)
dx x=x0

In other words, we predict that the external tangential flow is always decelerating at the separation point, whereas the
pressure gradient is always adverse (i.e., such as to decelerate the tangential flow), in agreement with experimental
observations.

7.10 Approximate Solutions of Boundary Layer Equations


The boundary layer equations, (7.110)(7.113), take the form

u v
+
x y

0,

u
u
dU
+v
U
x
y
dx

2 u
,
y2

(7.139)
(7.140)

subject to the boundary conditions


u(x, ) =

U(x),

(7.141)

u(x, 0) =

0,

(7.142)

v(x, 0) =

0.

(7.143)

Furthermore, it follows from (7.140), (7.142), and (7.143) that



2 u
dU
.
2 = U

dx
y y=0

(7.144)

The above expression can be thought of as an alternative form of (7.143). As we saw in Section 7.4, the boundary
layer equations can be solved exactly when U(x) takes the special form U0 xm . However, in the general case, we must
resort to approximation methods.
Following Pohlhausen, let us assume that
u(x, y)
= f (),
(7.145)
U(x)

Incompressible Boundary Layers

143

where = y/(x), and /x 1/. In particular, suppose that


(
a + b + c 2 + d 3 + e 4
f () =
1

01
,
>1

(7.146)

where a, b, c, d, and e are constants. This expression automatically satisfies the boundary condition (7.141). Moreover,
the boundary conditions (7.142) and (7.144) imply that a = 0, and
f (0) = (x),

(7.147)

where = d/d, and

2 dU
.
dx
Finally, let us assume that f , f , and f are continuous at = 1: i.e.,
=

(7.148)

f (1) =

1,

(7.149)

f (1) =

0,

(7.150)

0.

(7.151)

f (1) =

These constraints corresponds to the reasonable requirements that the velocity, vorticity, and viscous stress tensor,
respectively, be continuous across the layer. Given that a = 0, Equations (7.146), (7.147), and (7.149)(7.151) yield
f () = F() + G()

(7.152)

for 0 1, where
F() =
G() =

1 (1 ) 3 (1 + ),
1
(1 ) 3 .
6

(7.153)
(7.154)

Thus, the tangential velocity profile across the layer is a function of a single parameter, , which is termed the
Pohlhausen parameter. The behavior of this profile is illustrated in Figure 7.13. Note that, under normal circumstances, the Pohlhausen parameter must lie in the range 12 12. For > 12, the profile is such that f () > 1
for some < 1, which is not possible in a steady-state solution. On the other hand, for < 12, the profile is such
that f (0) < 0, which implies flow reversal close to the wall. As we have seen, flow reversal is indicative of separation.
Indeed, the separation point, f (0) = 0, corresponds to = 12. Note that expression (7.152) is only an approximation, since it satisfies some, but not all, of the boundary conditions satisfied by the true velocity profile. For instance,
differentiation of (7.140) with respect to y reveals that (3 u/y3 )y=0 f (0) = 0, which is not the case for expression
(7.152).
It follows from Equations (7.115), (7.116), and (7.152)(7.154) that
!
Z 1
3

(1 f ) d =
1 (x) =
,
(7.155)

10 120
0
!
Z 1
37

2
f (1 f ) d =
2 (x) =
.
(7.156)

315 945 9072


0
Furthermore,



u
U
U
=
2+
.
f (0) =
y y=0

Now, the von Karman momentum integral, (7.117), can be rearranged to give
!

1
U
2 u
d2 22 dU
.
2+
=
2
+

dx
dx
2
U y y=0

(7.157)

(7.158)

144

FLUID MECHANICS

1
0.9
0.8
0.7
0.6
f 0.5
0.4
0.3
0.2
0.1
0

0.2

0.4

0.6

0.8

Figure 7.13: Pohlhausen velocity profiles for = 12 (solid curve) and = 12 (dashed curve).
Defining
(x) =
we obtain

22 dU
,
dx

d
= 2 [F2 () {2 + F1 ()}] = F(),
U
dx dU/dx

(7.159)

(7.160)

where
=
F1 () =
F2 () =
F() =

!2

2
37
,

315 945 9072


!,
!
1
3
37

2
=

,
2
10 120
315 945 9072
!


2 u

2
 37
= 2 +
,

U y y=0
6 315 945 9072
!"
!
#
116
2

2
1
2
37
2
2 +

+
+
3 .
2
315 945 9072
315
945 120
9072

(7.161)
(7.162)
(7.163)
(7.164)

It is generally necessary to integrate Equation (7.158) from the stagnation point at the front of the obstacle, through the
point of maximum tangential velocity, to the separation point on the back side of the obstacle. Now, at the stagnation
point we have U = 0 and dU/dx , 0, which implies that F() = 0. Furthermore, at the point of maximum tangential
velocity we have dU/dx = 0 and U , 0, which implies that = = 0. Finally, as we have already seen, = 12 at
the separation point, which implies, from (7.161), that = 0.1567.
As was first pointed out by Walz, and is illustrated in Figure 7.14, it is a fairly good approximation to replace F()
by the linear function 0.47 6 for in the physically relevant range. The approximation is particularly accurate on
the front side of the obstacle (where > 0). Making use of this approximation, Equations (7.159) and (7.160) reduce

Incompressible Boundary Layers

145

1
F

0
0.15

0.1

0.05

0.05

0.1

Figure 7.14: The function F() (solid curve) and the linear function 0.47 6 (dashed line).
to the linear differential equation

which can be integrated to give

d U 22
dU 22

= 0.47 5
,
dx
dx
22 0.47
=

U6

(7.165)

U 5 (x ) dx ,

(7.166)

assuming that the stagnation point corresponds to x = 0. It follows that


Z
0.47 dU x 5
=
U (x ) dx .
U 6 dx 0

(7.167)

Recall that the separation point corresponds to x = x s , where (x s ) s = 0.1567.


Suppose that U(x) = U0 , which corresponds to uniform flow over a flat plate. (See Section 7.5.) It follows from
Equations (7.166) and (7.167) that
0.69
2 (x)
,
(7.168)
=
x
Re1/2
where Re = U0 x/, and = 0. Moreover, according to Equations (7.148) and (7.162), = 0 and 1 /2 = 2.55.
Hence, the displacement width of the boundary layer becomes
1 (x)
1.75
=
.
x
Re1/2

(7.169)

This approximate result compares very favorably with the exact result, (7.73).
Suppose that x = a and U() = 2 U0 sin , which corresponds to uniform transverse flow around a circular
cylinder of radius a. (See Section 7.8.) Equation (7.167) yields
Z
cos
sin5 d .
(7.170)
() = 0.47 6
sin 0

146

FLUID MECHANICS

0.2

0.1

20

40

60
()

80

100

120

Figure 7.15: The function () for flow around a circular cylinder.

Figure 7.16: Flow over the back surface of a semi-infinite wedge.


Figure 7.15 shows () determined from the above formula. It can be seen that = s = 0.1567 when = s 108.
In other words, the separation point is located 108 from the stagnation point at the front of the cylinder. This suggests
that the low pressure wake behind the cylinder is almost as wide as the cylinder itself, and that the associated form
drag is comparatively large.
Suppose, finally, that U = U0 xm . If m is negative then, as illustrated in Figure 7.16, this corresponds to uniform
flow over the back surface of a semi-infinite wedge whose angle of dip is
m
=
.
(7.171)
1+m 2
(See Section 7.4.) It follows from (7.167) that
=

0.47 m
0.47
=
.
1+5m
/2 4

(7.172)

Now, we expect boundary layer separation on the back surface of the wedge when < s = 0.1567. This corresponds
to > s , where
( s )

13 .
(7.173)
s =
2 0.47 + 4 ( s )

Incompressible Boundary Layers

147

Hence, boundary layer separation can be prevented by making the wedges angle of dip sufficiently shallow: i.e., by
streamlining the wedge, which has the effect of reducing the deceleration of the flow on the wedges back surface.
Note that the critical value of m (i.e., m s = 0.0125) at which separation occurs in our approximate solution is very
similar to the critical value of m (i.e., m = 0.0905) at which the exact self-similar solutions described in Section 7.4
can no longer be found. This suggests that the absence of self-similar solutions for m < m is related to boundary layer
separation.

7.11 Exercises
7.1. Fluid flows between two non-parallel plane walls, towards the intersection of the planes, in such a manner that if x is
measured along a wall from the intersection of the planes then U(x) = U0 /x, where U0 is a positive constant. Verify that
a solution of the boundary layer equation (7.35) can be found such that is a function of y/x only. Demonstrate that this
solution yields
" 1/2 #
u(x, y)
y
U0
=F
,
U(x)

x
where u = /y, and
F F 2 = 1,
subject to the boundary conditions F(0) = 0 and F() = 1. Verify that

!
z
F(z) = 3 tanh2 + 2
2
is a suitable solution of the above differential equation, where tanh2 = 2/3.
7.2. A jet of water issues from a straight narrow slit in a wall, and mixes with the surrounding water, which is at rest. On the
assumption that the motion is non-turbulent and two-dimensional, and that the approximations of boundary layer theory
apply, the stream function satisfies the boundary layer equation

3 2 2

+
= 0.
y3
x y2
y x y

Here, the symmetry axis of the jet is assumed to run along the x-direction, whereas the y-direction is perpendicular to this
axis. The velocity of the jet parallel to the symmetry axis is
u(x, y) =

,
y

where u(x, y) = u(x, y), and u(x, y) 0 as y . We expect the momentum flux of the jet parallel to its symmetry axis,
Z
M=
u2 dy,

to be independent of x.
Consider a self-similar stream function of the form
(x, y) = 0 x p F(y/xq ).
Demonstrate that the boundary layer equation requires that p + q = 1, and that M is only independent of x when 2 p q = 0.
Hence, deduce that p = 1/3 and q = 2/3.
Suppose that
(x, y) = 6 x1/3 F(y/x2/3 ).
Demonstrate that F(z) satisfies

F + 2 F F + 2 F 2 = 0,
subject to the constraints that F (z) = F (z), and F (z) 0 as z . Show that
F(z) = tanh( z)
is a suitable solution, and that
M = 48 2 3 .

148

FLUID MECHANICS

7.3. The growth of a boundary layer can be inhibited by sucking some of the fluid through a porous wall. Consider conventional
boundary layer theory. As a consequence of suction, the boundary condition on the normal velocity at the wall is modified
to v(x, 0) = vs , where vs is the (constant) suction velocity. Demonstrate that, in the presence of suction, the von Karman
velocity integral becomes

d
dU
u
= U2 2 +U
(1 + 2 2 ) + U vs .

y y=0
dx
dx
Suppose that

u(x, y) = U(x)

0 y /(2 )
,
y > /(2 )

sin( y)
1

where = (x). Demonstrate that the displacement and momentum widths of the boundary layer are
1

(/2 1) 1 ,
(1 /4) 1 ,

respectively. Hence, deduce that


(/2 1)2
d1 dU
= U (1 /4)
+
1 + (/2 1) vs .
1
dx
dx
Consider a boundary layer on a flat plate, for which U(x) = U0 . Show that, in the absence of suction,
8
1 = (/2 1)
4

!1/2

x
U0

!1/2

but that in the presence of suction


(/2 1)
.
vs
Hence, deduce that, for a plate of length L, suction is capable of significantly reducing the thickness of the boundary layer
when
vs
1
,

U0
Re1/2
where Re = U0 L/.
1 =

Incompressible Aerodynamics

149

8 Incompressible Aerodynamics

8.1 Introduction
This chapter investigates the forces exerted on a stationary obstacle situated in a uniform, high Reynolds number wind,
on the assumption that the obstacle is sufficiently streamlined that there is no appreciable separation of the boundary
layer from its back surface. Such an obstacle is termed an airfoil (or aerofoil). Obviously, airfoil theory is fundamental
to the theory of flight. The flow around an airfoil is essentially irrotational and inviscid everywhere apart from a thin
boundary layer localized to its surface, and a thin wake emitted by its trailing edge. (See Sections 7.5 and 7.6.) It
follows that, for the flow external to the boundary layer and wake, we can write
v = ,

(8.1)

which automatically ensures that the flow is irrotational. Assuming that the flow is also incompressible, so that v = 0,
the velocity potential, , satisfies Laplaces equation: i.e.,
2 = 0.

(8.2)

The appropriate boundary condition at the surface of the airfoil is that the normal velocity be zero. In other words,
n = 0, where n is a unit vector normal to the surface. In general, the tangential velocity at the airfoil surface,
obtained by solving 2 = 0 in the external region, subject to the boundary condition n = 0 on the surface, is
non-zero. Of course, this is inconsistent with the no slip condition, which demands that the tangential velocity be zero
at the surface. (See Section 7.2.) However, as described in the previous chapter, this inconsistency is resolved by the
boundary layer, across which the tangential velocity is effectively discontinuous, being non-zero on the outer edge
of the layer (where it interfaces with the irrotational flow), and zero on the inner edge (where it interfaces with the
airfoil). The discontinuity in the tangential velocity across the layer implies the presence of bound vortices covering
the surface of the airfoil (see Section 8.7), and also gives rise to a friction drag acting on the airfoil in the direction of
the external flow. However, the magnitude of this drag scales as Re1/2 , where Re is the Reynolds number of the wind.
(See Section 7.5.) Hence, such drag becomes negligibly small in the high Reynolds number limit. In the following,
we shall assume that any form drag, due to the residual separation of the boundary layer at the back of the airfoil, is
also negligibly small. Moreover, for the sake of simplicity, we shall initially restrict our discussion to two-dimensional
situations in which a high Reynolds number wind flows transversely around a stationary airfoil of infinite length (in
the z-direction) and uniform cross-section (parallel to the x-y plane).

8.2 Theorem of Kutta and Zhukovskii


Consider a two-dimensional airfoil that is at rest in a uniform wind of speed V whose direction subtends a (clockwise)
angle with the negative x-axis. It follows that the wind velocity is V = V cos e x + V sin ey , and the corresponding complex velocity is dF/dz = V e i . (See Section 6.4.) Now, the air velocity a great distance from the airfoil must
tend toward this uniform velocity. Thus, for sufficiently large |z|, we can write (see Section 6.4)
A B
dF
= V ei + + 2 + .
dz
z z

(8.3)

According to Equation (6.87), the circulation, , of air about the airfoil is determined by performing the integral
I
dF
dz =
(8.4)
Re
C dz
around a loop C that lies just above the airfoil surface. However, as discussed in Section 6.8, the value of this integral
is unchanged if it is performed around any loop that can be continuously deformed onto C, whilst not passing through
the airfoil surface, or crossing a singularity of the complex velocity, dF/dz (i.e., a line source or a z-directed vortex

150

FLUID MECHANICS

filament). Since (in the high Reynolds number limit in which the boundary layer and the wake are infinitely thin) there
are no line sources or z-directed vortex filaments external to the airfoil, we can evaluate the integral around a large
circle of radius R, centered on the origin. It follows that z = R e i and dz = i R e i d = i z d. Hence,
= Re i

I h

i !
V R e i (+) + A + O(R 1 ) d = 2 Im(A),

(8.5)

which implies that


dF

B
= V ei + i
+
+
dz
2 z z 2

(8.6)

at large |z|.
As discussed in Section 6.9, the net force (per unit length) acting on the airfoil, L = X e x + Y ey , is determined by
performing the Blasius integral,
!2
I
dF
1
dz = X i Y,
(8.7)
i
2
C dz
around a loop C that lies just above the airfoil surface. However, as before, the value of the integral is unchanged if
we perform it instead around a large circle of radius R, centered on the origin. Now, far from the airfoil,
dF
dz

!2

= V 2 e2i + i

So, we obtain
X iY =

1 2
i
2

V e i 82 V B e i 2
+
+ O(z 3 ).
z
42 z 2

#
I "
V ei
V 2 R e i (+2 ) + i
+ O(R1 ) d = i e i V ,

(8.8)

(8.9)

or
X + i Y = i ei V = e i (/2) V .

(8.10)

In other words, the resultant force (per unit length) acting on the airfoil is of magnitude V , and has the direction
obtained by rotating the wind vector through a right-angle in the sense opposite to that of the circulation. This type of
force is known as lift, and is responsible for flight. The result (8.10) is known as the theorem of Kutta and Zhukovskii,
after the German scientist M.W. Kutta, and the Russian scientist N.E. Zhukovskii, who discovered it independently.
Note that (at fixed circulation) the lift is independent of the shape of the airfoil. Furthermore, according to the Kutta
Zhukovskii theorem, there is zero drag acting on the airfoil (i.e., zero force acting in the direction of the wind). In
reality, there is always a small friction drag due to air viscosity, as well as a (hopefully) small form drag due to residual
separation of the boundary layer from the back of the airfoil. There is actually a third type of drag, known as induced
drag, which is discussed in Section 8.8.
As discussed in Section 6.9, the net moment per unit length (about the origin), M, acting on the airfoil is determined
by performing the integral

1 I dF 2
z dz = M
Re
(8.11)
2
dz
C

around a loop C that lies just above the airfoil surface. As before, we can deform C into a circle of radius R, centered
on the origin, without changing the value of the integral. Hence, we obtain
1
M = Re i
2

or

I "

V R e

2 i (+)

# !
V R e i (+) 82 V B e i 2
1
+ O(R ) d ,
+i
+

42

i
h
M = Re 2 V B e i (/2) .

(8.12)

(8.13)

Incompressible Aerodynamics

151

8.3 Cylindrical Airfoils


For the moment, let us work in the complex -plane, where = + i . Consider a cylindrical airfoil with a circular
cross-section of radius a, centered on the origin, that is situated in a uniform, high Reynolds number wind of speed
V whose direction subtends a (clockwise) angle with the negative -axis. Let be the circulation of air around the
airfoil. A slight generalization of the analysis of Section 6.4 reveals that the appropriate complex velocity potential is
!
 
a2 i
i
,
(8.14)
+i
e
ln
F() = V e +

2
a
whereas the associated stream function takes the form

!
a2
r
(r, ) = V r
ln
sin( + ) +
,
r
2
a

(8.15)

where = r e i . It follows that


dF

V a2 e i
= V ei + i

.
d
2
2
Comparison with Equation (8.6) (with z = ) reveals that
B = V a2 ei .

(8.16)

(8.17)

Hence, Equations (8.10) and (8.13) yield


V
L
M



= V cos e + sin e ,


= V sin e + cos e ,

= 0,

(8.18)
(8.19)
(8.20)

where V is the wind vector, L the lift vector, and M the moment of the lift vector about the origin. We conclude that,
for a cylindrical airfoil of circular cross-section, the lift vector is normal to the wind vector, and the line of action of
the lift passes through the centroid of the cross-section (since the lift generates zero moment about the origin). See
Figure 8.1.
Of course, a cylindrical airfoil of circular cross-section is completely unrealistic, since its back side (i.e., the side
opposite to that from which the wind is incident) is not sufficiently streamlined to prevent boundary layer separation.
(See Chapter 7.) However, as described in Section 6.7, we can use the conformal transformation
z=+

l2

(8.21)

to transform a cylinder of circular cross-section in the -plane to a cylinder of elliptical cross-section in the z-plane.
(Note that both cross-sections have centroids located at the origin.) Moreover, a cylindrical airfoil of elliptical crosssection that is sufficiently elongated, and whose major axis subtends a sufficiently small angle with the incident wind
direction, constitutes a realistic airfoil, since its back side is, for the most part, closely aligned with the external flow.
An elliptical airfoil of width c and thickness < c, as shown in Figure 8.2, is obtained when the parameters a and
l are given the following values:
1
(c + ),
4
1 2
l =
(c 2 )1/2 .
4
In this case, the surface of the airfoil satisfies the parametric equations
c
cos ,
x =
2

y =
sin .
2
a =

(8.22)
(8.23)

(8.24)
(8.25)

152

FLUID MECHANICS

Figure 8.1: A cylindrical airfoil of circular cross-section.

x
C

M
Figure 8.2: A cylindrical airfoil of elliptical cross-section.

Incompressible Aerodynamics

153

In particular, the airfoils leading and trailing edges correspond to = 0 and = , respectively.
According to Equations (8.16) and (8.21), the complex velocity in the z-plane is given by
"
#
!
dF dF d

V a2 e i
2
=
= V ei + i

.
dz
d dz
2
2
2 l2
Thus, on the airfoil surface, where = a e i , we obtain
#
"
(c + )

dF
.
= i V sin( + ) +
dz
(c + ) ( cos + i c sin )

(8.26)

(8.27)

A long way from the airfoil, z l 2 /z, so that Equation (8.26) reduces to
dF

V (l 2 e i a2 ei )
V ei + i
+
.
dz
2 z
z2

(8.28)

A comparison with Equation (8.6) reveals that the circulation of air around the airfoil takes the same value, , in both
the complex - and z-planes. In other words, the conformal transformation (8.21) does not modify the circulation.
Note that the transformation also does not modify the external wind speed or direction [since, from (8.16) and (8.28),
dF/d = dF/dz = V e i at very large || and |z|]. On the other hand, it is clear that the constant B, which takes the
value zero in the complex -plane, takes the value
B = V (l 2 e i a2 ei )

(8.29)

in the complex z-plane. Hence, Equations (8.10) and (8.13) reveal that
V =

V ek ,

(8.30)

L =

V e ,

V 2 (c2 2 ) sin(2 ),
8

(8.31)

(8.32)

where V is the wind vector, L the lift vector, and M the moment of the lift vector about the origin. Here,
ek = cos e x + sin ey

(8.33)

is a unit vector parallel to the incident wind direction, and


e = ek ez = sin e x + cos ey

(8.34)

is a unit vector perpendicular to the wind direction. We conclude that, for a cylindrical airfoil of elliptic cross-section,
the lift vector is normal to the wind vector, but the line of action of the lift intersects the major axis of the airfoil a
distance
M
1
sin
d=
(8.35)
= (c )
L cos 4
b

b = /[ V (c + )]. See Figure 8.2. Incidentally, the point, F, at


in front of the cross-sections centroid, C, where
which the line of action of the lift intersects the airfoils major axis is conventionally termed the focus of the airfoil.

8.4 Zhukovskiis Hypothesis


According to the previous analysis, the lift acting on a cylindrical airfoil of elliptic cross-section, situated in a uniform,
high Reynolds number wind, depends on the circulation, , of air about the airfoil. But, how can we determine the
value of this circulation?
Figure 8.3 shows the boundary layer and wake of a streamlined airfoil. The boundary layer, which is localized on
the surface of the airfoil, has a vortex intensity per unit length in the z-direction equal to U, where U is the tangential air

154

FLUID MECHANICS

dS
C

V
Wake
Boundary Layer

U
Trailing Edge

Airfoil

Figure 8.3: The boundary layer and wake of a streamlined airfoil. Only shaded regions posses non-zero vorticity.
speed immediately above the layer. [See Equation (8.64).] Moreover, the wake is emitted by the airfoils trailing edge,
and subsequently convected by the external air flow. (See Section 7.6.) Note that the flow is irrotational everywhere
apart from inside the boundary layer and the wake. Now, according to the analysis of Section 5.13, the rate of change
of the circulation, , around some curve C that encloses the airfoil is equal to minus the flux of z-directed vorticity
across this curve: i.e.,
I
d
= z v dS.
(8.36)
dt
Here, v is the wind velocity, the wind vorticity, and dS an outward surface element (of unit depth in the z-direction)
lying on C. We expect the vorticity flux to be independent of the size and shape of C, otherwise the circulation of the
flow, , about the airfoil would not have a unique value. In the limit that C becomes very large, v V, where V is
the incident wind velocity. Thus,
d
= V z ,
(8.37)
dt
where V is the wind speed, and z the vortex intensity per unit length in the wake (at the point where it crosses the
curve C). Here, we are assuming that the vorticity within the wake is convected by the flow, giving rise to a net flux
of vorticity across C. Since the wake is essentially an extension of the boundary layer, it is reasonable to assume that
its vortex intensity per unit length is proportional to that in the boundary layer at the airfoils trailing edge, where the
wake and boundary layer intersect. In other words, z = k U0 , where U0 is the tangential velocity immediately above
the trailing edge of the airfoil, and k is a constant. It follows that
d
= k V U0 .
dt
According to Equation (8.27), the tangential velocity just above the surface of the airfoil is
#
#
"
"
dF
(c + )

c sin i cos
.
= V sin( + ) +
U() = Re
2
2
2
2
2
1/2
2
dz

(c
+
)
(c sin + cos )
(c sin + 2 cos2 )1/2
||=a

(8.38)

(8.39)

Incompressible Aerodynamics

155

y/c

2
2

0
x/c

Figure 8.4: Streamlines around a slender cylindrical airfoil of elliptic cross-section situated in a uniform, high
Reynolds number wind. The parameters for this calculation are /c = 0.1, = /12, and = 0.
Hence, given that the airfoils trailing edge corresponds to = , we obtain
#
"
c + 

.
V sin
U0 =
(c + )

(8.40)

Thus, Equation (8.38) yields


db

= b
+ sin ,
dt

(8.41)

where b
= /[ V (c + )], t = t/t0 , and t0 = /(k V). Assuming that the circulation of the flow about the airfoil is
initially zero (i.e., b
= 0 at t = 0), the above equation can be solved to give


b
= sin 1 exp(t) .

Clearly, as t the normalized circulation b


asymptotes to the constant value

(8.42)

b
= sin .

(8.43)

= V (c + ) sin .

(8.44)

The corresponding constant value of the unnormalized circulation is

Note that, according to Equation (8.40), when the circulation, , takes the value the tangential velocity at the
airfoils training edge, U0 , is zero. In other words, the steady-state circulation set up around the airfoil is such as to
render its trailing edge a stagnation point of the flow. This conclusion is known as Zhukovskiis hypothesis, after its
discoverer N.E. Zhukovskii.
Incidentally, it should be clear, from the above discussion, that the air circulation about the airfoil is only able to
change its value because of the presence of the boundary layer, and the associated wake that trails from the airfoils

156

FLUID MECHANICS

y/c

2
2

0
x/c

Figure 8.5: Streamlines around a slender cylindrical airfoil of elliptic cross-section situated in a uniform, high
Reynolds number wind. The parameters for this calculation are /c = 0.1, = /12, and = .
trailing edge. This follows because the flow is irrotational everywhere except within the boundary layer and the wake.
Moreover, as we have seen, a change in circulation is necessarily associated with a net vorticity flux away from the
airfoil, and such a flux cannot be carried by an irrotational wind. Thus, in the absence of the boundary layer and the
wake, the air circulation about the airfoil would be constrained to remain zero (assuming that it was initially zero),
in accordance with the Kelvin circulation theorem. (See Section 5.13.) This implies, from Equation (8.31), that zero
lift would act on the airfoil, irrespective of its shape, and irrespective of the incident wind speed or direction. In
other words, flight would be impossible. Fortunately, as long as the tangential air velocity at the trailing edge of
the airfoil is non-zero, the wake that trails behind the airfoil carries a net flux of z-directed vorticity, which causes
the airfoil circulation to evolve in time. This process continues until the circulation becomes such that the tangential
velocity at the airfoils trailing edge is zero: i.e., such that the trailing edge is a stagnation point. Thereafter, the
circulation remains constant (assuming that the wind speed and direction remain constant). Figures 8.4 and 8.5 show
the streamlines of the flow around a slender cylindrical airfoil of elliptic cross-section situated in a uniform, high
Reynolds number wind whose direction of incidence is slightly inclined to the airfoils major axis. In the first figure,
the air circulation about the airfoil is zero. In the second figure, the circulation is such as to make the trailing edge of
the airfoil a stagnation point.
According to Equations (8.31) and (8.44), when the air circulation about the airfoil has attained its steady-state
value, , the lift acting on the airfoil becomes
L = V = V 2 (c + ) sin .

(8.45)

Note that the lift is positive (i.e., upward) when > 0 (i.e., when the wind is incident on the airfoils bottom surface),
negative (i.e., downward) when < 0 (i.e., when the wind is incident on the airfoils top surface), and zero when = 0
(i.e., when the wind is incident parallel to airfoils major axis). Incidentally, the angle is conventionally termed the
angle of attack. Finally, from Equations (8.35) and (8.43), the focus of the airfoil is located a distance
d=

sin 1
1
(c )
= (c )
4
b
4

(8.46)

Incompressible Aerodynamics

157

()

180

10
()

20

Figure 8.6: The angular locations of the boundary layer separation points, , calculated as a function of the angle of
attack, , for a cylindrical airfoil of elliptic cross-section situated in a uniform, high Reynolds number wind. The solid,
dashed, short-dashdotted, and long-dashdotted curves correspond to airfoils of ellipticity /c = 1.0, 0.5, 0.25, and
0.125, respectively. The trailing edge of the airfoil is located at = 180 .
in front of the centroid of its cross-section. In the limit that the airfoil becomes very thin (i.e., c), this distance
asymptotes to c/4. Thus, we conclude that the focus of a thin airfoil, which is defined as the point of action of the lift,
is located one quarter of the way along the airfoil from its leading edge.
The above analysis is premised on the assumption that there is no appreciable separation of the boundary layer
from the back of the airfoil, which implies the neglect of form drag. We can check that this assumption is reasonable
by calculating the approximate locations of the boundary layer separation points using the analysis of Section 7.10.
Let s represent arc-length along the surface of the airfoil, measured from the front stagnation point. Assuming that,
in accordance with Zhukovskiis hypothesis, the circulation is such that = , this stagnation point is located at
= 0 , where 0 = 2 . [See Equation (8.49).] It follows that
1
h() d,
2

(8.47)

h() = (c2 sin2 + 2 cos2 )1/2 .

(8.48)

ds = (dx 2 + dy 2 )1/2 =
where

Moreover, from (8.39) and (8.44), the tangential air speed just above the surface of the airfoil can be written
U() = V (c + )

f ()
,
h()

(8.49)

with
In addition, it can be shown that

f () = sin( + ) + sin .

(8.50)

!
f
d
= g(),
ln
d
h

(8.51)

158

FLUID MECHANICS

vz (y = 0+)

z
vz (y = 0)

dz

Figure 8.7: Side view of a vortex sheet.


where

cos( + ) (c2 2 ) cos sin

.
(8.52)
f ()
h2 ()
According to the analysis of Section 7.10, the separation points are located at = , where ( ) = 0.1567, and
Z
h4 () g() f 5 ( )
() = 0.47
d .
(8.53)
4
f 5 ()
0 h ( )
g() =

Here, > + > 0 and 0 > > . Moreover, the x- and y-coordinates of the separation points are x = (c/2) cos
and y = (/2) sin , respectively. Figure 8.6 shows the angular locations of the separation points, calculated as
a function of the angle of attack, for cylindrical airfoils of various different ellipticity, /c. (Note that has been
re-expressed as an angle in the range 0 to 2.) It can be seen that for a bluff airfoil (e.g., /c = 1) the angular distance
between the separation points is large, indicating the presence of a wide wake, and a high associated form drag (since
the magnitude of form drag is roughly proportional to the width of the wake). On the other hand, for a slender airfoil
(e.g., /c = 0.125) the angular distance between the separation points is much smaller, indicating the presence of a
narrow wake, and a low associated form drag. Note, however, that, in the latter case, as the angle of attack is gradually
increased from zero, there is an initial gradual increase in the angular distance between the separation points, followed
by an abrupt, and very large, increase. We would expect there to be a similar gradual increase in the form drag,
followed by an abrupt, and very large, increase. The value of the angle of attack at which this abrupt increase occurs
is termed the critical angle of attack. We conclude that the previous analysis, which neglects form drag, is valid only
for slender airfoils whose angles of attack do not exceed the critical value (which is generally only a few degrees).

8.5 Vortex Sheets


A vortex sheet is defined as a planar array of parallel vortex filaments. Consider a uniform vortex sheet, lying in the
x-z plane, in which the vortex filaments run parallel to the x-axis. See Figure 8.7. The vorticity within the sheet can
be written
= x (y) e x .
(8.54)
Here, = x e x is the sheets vortex intensity per unit length. Let vz (y = 0+ ) and vz (y = 0 ) be the z-component of
the fluid velocity immediately above and below the sheet, respectively. Consider a small rectangular loop in the y-z
plane that straddles the sheet, as shown in the figure. Integration of = v around the loop (making use of the curl
theorem) yields
vz vz (y = 0+ ) vz (y = 0 ) = x .
(8.55)

In other words, a vortex sheet induces a discontinuity in the tangential flow across the sheet. The above expression can
easily be generalized to give
= n v,
(8.56)
where is the sheets vortex intensity per unit length, n is a unit vector normal to the sheet, and v is the jump in
tangential velocity across the sheet (traveling in the direction of n). Furthermore, it is reasonable to assume that the
above relation holds locally for non-planar and non-uniform vortex sheets.

Incompressible Aerodynamics

159

8.6 Induced Flow


A vortex filament is necessarily associated with fluid flow circulating about the filament. Let us determine the relationship between the filament vorticity and the flow field that it induces. This problem is mathematically identical to
determining the magnetic field generated by a current filament. In the latter case, the Maxwell equation
0 j = B
can be inverted to give the well-known Biot-Savart law
Z
j(r ) (r r ) 3
1
d r.
B(r) =
4
|r r | 3

(8.57)

(8.58)

Here, j is the current density, and B the magnetic field-strength. By analogy, given that vorticity is related to fluid
velocity via the familiar relation
= v,
(8.59)
we can write

1
v(r) =
4

(r ) (r r ) 3
d r.
|r r | 3

(8.60)

This expression allows us to determine the flow field induced by a given vorticity distribution. In particular, for a
vortex filament of intensity the above expression reduces to
Z
(r ) (r r )
1
v(r) =
dl ,
(8.61)
4
|r r | 3
where dl is an element of length along the filament. Likewise, for a vortex sheet of intensity per unit length , we
obtain
Z
(r ) (r r )
1
v(r) =
dS ,
(8.62)
4
|r r | 3
where dS is an element of area of the sheet.

8.7 Three-Dimensional Airfoils


Let us now take into account the fact that realistic three-dimensional airfoils are of finite size. Consider Figure 8.8,
which shows a top view of a stationary airfoil of finite size, situated in a (predominately) horizontal wind of velocity
V = V ek . In the following, we shall sometimes refer to such an airfoil as a wing (although it actually represents a
pair of wings on a standard fixed wing aircraft). Let us adopt the coordinate system shown in the figure, which is
such that the x-z plane is horizontal, the wind is incident predominately from the x-direction, and the y-axis points
vertically upward. The wing is assumed to lie in the x-z plane. Let b be the wingspan, and let c(z) and (z) be the
width and thickness, respectively, of the wing cross-section (parallel to the x-y plane). See Figure 8.9. Suppose
that the wing is symmetric about the median plane, z = 0, so that c(z) = c(z) and (z) = (z). It follows that
c(z > b/2) = (z > b/2) = 0: i.e., the wing extends from z = b/2 to z = b/2.
Suppose that air circulation is set up around the wing parallel to the x-y plane in such a manner as to produce an
upward lift. It follows that the average pressure on the lower surface of the wing must exceed that on its upper surface.
Consider Figure 8.9, which shows a back view of the airfoil shown in Figure 8.8. As we go from the median plane
(z = 0) to a wing tip, Y, whether along the upper or the lower surface of the wing, we must arrive at the same pressure
at Y. It follows that there is a drop in pressure as we move outward, away from the median plane, along the wings
bottom surface, and a further drop in pressure as we move inward, toward the median plane, along the upper surface.
Since air is pushed in the direction of decreasing pressure, it follows that the air that impinges on the wings leading
edge, and then passes over its upper surface, deviates sideways toward the median plane. Likewise, the air that passes
over the wings lower surface deviates sideways away from the median plane. See Figure 8.8.
Now, the air that leaves the trailing edge of the wing at some point Q must have impinged on the leading edge at
the different points P and P , depending on whether it travelled over the wings upper or lower surfaces, respectively.

160

FLUID MECHANICS

V
airfoil

c(z)

z = b/2

z=0

z = b/2

Figure 8.8: Top view of a three-dimensional airfoil of finite size.

y
b
Y

(z)

Figure 8.9: Back view of a three-dimensional airfoil of finite size, indication the pressure variation over its surface.

Incompressible Aerodynamics

161
x

P
Q

Figure 8.10: Top view of the airflow over the top (left) and bottom (right) surfaces of a three-dimensional airfoil of
finite size.
Moreover, air that travels to Q via the wings upper surface acquires a small sideways velocity directed towards the
median plane, whereas that which travels to Q via the lower surface acquires a small sideways velocity directed away
from the median plane. On the other hand, the air speed at Q must be the same, irrespective of whether the air arrives
from the wings upper or lower surface, because the pressure (which, according to Bernoullis theorem, depends on the
air speed) must be continuous at Q. Thus, we conclude that there is a discontinuity in the direction of the air emitted
by the trailing edge of a wing. This implies that the interface, , between the two streams of air that travel over the
upper and lower surfaces of the wing is a vortex sheet. (See Section 8.5.) Of course, this vortex sheet constitutes the
wake that trails behind the airfoil. Moreover, we would generally expect the wake to be convected by the incident
wind. It follows that the vorticity per unit length in the wake can be written
= I(z) ek ,

(8.63)

where I(z) = vz , and vz is tangential velocity discontinuity across the wake. [See Equation (8.56).]
As we saw previously, the boundary layer that covers the airfoil is such that the tangential velocity U just outside
the layer is sharply reduced to zero at the airfoil surface. Actually, the nature of the substance enclosed by the surface
is irrelevant to our argument, and nothing is changed in our analysis if we suppose that this region contains air at rest.
Thus, we can replace the airfoil by air at rest, and the boundary layer by a vortex sheet, S , with a vortex intensity per
unit length S that is determined by the velocity discontinuity U between the air just outside the boundary layer and
that at rest in the region where the airfoil was previously located. In fact, Equation (8.56) yields
S = n U,

(8.64)

where n is an outward unit normal to the airfoil surface.


We conclude that a stationary airfoil situated in a uniform wind of constant velocity is equivalent to a vortex sheet
S , located at the airfoil surface, and a wake that trails behind the airfoil, the airfoil itself being replaced by air at rest.
The vorticity within S is largely parallel to the z-axis [since n and U are both essentially parallel to the x-y planesee
Equation (8.64)], whereas that in is parallel to the incident wind direction. See Figure 8.11. The vortex filaments
within S are generally termed bound filaments (since they cannot move off the airfoil surface). Conversely, the vortex
filaments within are generally termed free filaments. The air velocity both inside and outside S can be written
v = V + v + vS ,

(8.65)

where V is the external wind velocity, v the velocity field induced by the free vortex filaments that constitute , and
vS the velocity field induced by the bound filaments that constitute S .
Consider some point P that lies on S . Let P+ and P be two neighboring points that are equidistant from P, where
P+ lies just outside S , and P lies just inside S , and the line P P+ is normal to S . We can write
v(P+ )

V + v (P+ ) + vS (P+ ),

(8.66)

v(P )

V + v (P ) + vS (P ).

(8.67)

162

FLUID MECHANICS

x
z
V
Figure 8.11: Vortex structure around a wing.
However, v(P+ ) = U(P), where U(P) is the tangential air velocity just above point P on the airfoil surface, and
v(P ) = 0, since the air within S is stationary. Moreover, v (P+ ) = v (P ) = v (P), since we expect v to be
continuous across S . On the other hand, we expect the tangential component of vS to be discontinuous across S . Let
us define
1
vS (P) = [vS (P+ ) + vS (P )] .
(8.68)
2
This quantity can be identified as the velocity induced at point P by the bound vortices on S , excluding the contribution
from the local bound vortex at P (since this vortex induces equal and opposite velocities at P+ and P ). Finally, taking
half the sum of Equations (8.66) and (8.67), we obtain
1
U(P) = V + v (P) + vS (P).
2

(8.69)

8.8 Aerodynamic Forces


The net aerodynamic force acting on an three-dimensional airfoil of finite size can be written
Z
p n dS ,
A=

(8.70)

where the integral is taken over the surface of the airfoil, S . Here, n is an outward unit normal vector on S , dS is
an element of S , and p is the air pressure. From Bernoullis theorem (in an irrotational fluid), we can write p =
p0 (1/2) v2, where p0 is a constant pressure. Since a constant pressure exerts no net force on a closed surface, we
get
Z
1
A = U 2 n dS ,
(8.71)
2
S
where U is the tangential air velocity just above the surface of the airfoil. Now,
U (n U) = U 2 n (n U) U = U 2 n,
since n U = 0 on the surface. Hence,

1
A=
2

U (n U) dS .

Making use of Equations (8.64) and (8.69), the above expression can be written
Z
A = (V + v + vS ) S dS = L + D + F,
S

(8.72)

(8.73)

(8.74)

Incompressible Aerodynamics

163

where
L
D
F

= V
=
=

S dS ,

(8.75)

v S dS ,

(8.76)

vS S dS .

(8.77)

Here, V, v , and vS are the incident wind velocity, the velocity induced by the free vortices in the wake, and the
velocity induced by the bound vortices covering the surface of the airfoil, respectively. The forces L and D are called
the lift and the induced drag, respectively. (Note, that L now represents a net force, rather than a force per unit length.)
We shall presently demonstrate that the force F is negligible.
Let us assume that
S z ez :
(8.78)
i.e., that the bound vortices covering the surface of the airfoil run parallel to the z-axis. This assumption is exactly
correct for an airfoil of infinite wingspan and constant cross-section. Moreover, it is a good approximation for an
airfoil of finite wingspan, provided the airfoils length greatly exceeds its width (i.e., b c). Now, the incident wind
velocity is written V = V ek . Moreover, dS = dl dz, where dl is an element of length that runs parallel to the x-y plane
whilst lying on the airfoil surface. Now, making use of the curl theorem, we can easily show that
I
z dl = (z),
(8.79)
C

where the closed curve C is the intersection of the airfoil surface with the plane z = z, and (z) is the air circulation
about the airfoil in this plane. Thus, it follows from Equation (8.75) that
L = V

b/2

b/2

(z) dz e .

(8.80)

This expression is the generalization of Equation (8.31) for a three-dimensional airfoil of finite size. As before, the lift
is at right-angles to the incident wind direction.
Let us make the further assumptionknown as the lifting line approximation (because the lifting action of the
wing is effectively concentrated onto a line)that
v = w(z) e

(8.81)

throughout S , where w(z) e is the induced velocity due to the free vortices in , evaluated at the trailing edge of the
airfoil. Here, the velocity w(z) is called the downwash velocity. It follows from Equation (8.76) that
D=

b/2

b/2

w(z) (z) dz ek .

(8.82)

Note that the induced drag is parallel to the incident wind direction. The origin of induced drag is as follows. It
takes energy to constantly resupply free vortices to the wake, as they are swept downstream by the wind (note that a
vortex filament possesses energy by virtue of the kinetic energy of its induced flow pattern), and this energy is supplied
by the work done in opposing the induced drag. The drag acting on a well-designed airfoil (i.e., an airfoil with an
aerodynamic shape that minimizes form drag) situated in a high Reynolds number wind (which implies that the friction
drag is negligible) is generally dominated by induced drag.
According to Equations (8.62) and (8.77), the force F is written
Z Z

[S (r ) (r r )] S (r)
F=
dS dS .
(8.83)
4 S S
|r r | 3

164

FLUID MECHANICS

rr

Figure 8.12: Semi-infinite vortex filament.


We can interchange primed and unprimed variables without changing the value of the integral. Hence,
Z Z
[S (r) (r r)] S (r )

dS dS .
F=
4 S S
|r r| 3

(8.84)

Taking the half the sum of the previous two equations, we obtain
Z Z

[S (r ) (r r )] S (r) + [(r r ) S (r)] S (r )


F=
dS dS .
8 S S
|r r | 3

(8.85)

However, (a b) c + (b c) a + (c a) b = 0. Thus, the above expression yields


Z Z
[S (r ) S (r)] (r r )

dS dS .
F=
8 S S
|r r | 3

(8.86)

But, the assumption (8.78) implies that S (r ) S (r) 0. Hence, F is negligible, as was previously stated.
Consider a closed surface covering the small section of the airfoil lying between the parallel planes z = z and
z = z + dz. The flux of vorticity into the surface due to bound vortices at z is (z). The flux of vorticity out of the
surface due to bound vortices at z + dz is (z + dz). Finally, the flux of vorticity out of the surface due to the free
vortices in the part of the wake lying between z and z + dz is I(z) dz. However, the net flux of vorticity out of a closed
surface is zero, since vorticity is divergence free. Hence,
(z) = (z + dz) + I(z) dz,

(8.87)

which implies that


d
.
(8.88)
dz
Finally, consider a semi-infinite straight vortex filament of vortex intensity = e x that terminates at the origin,
O, as shown in Figure 8.12. Let us calculate the flow velocity induced by this filament at the point P = (0, 0, z). From
the diagram l = z tan , dl = z sec2 d, |r r | = z sec , and |r r | = z ey . Hence, from Equation (8.61), the
induced velocity at P is v = vy ey , where
Z /2

vy =
.
(8.89)
cos d =
4 z 0
4 z
I(z) =

This result allows us to calculate the downwash velocity, w(z) = vy (z), induced at the trailing edge of the airfoil by
the semi-infinite free vortices in the wake. The vortex intensity in the small section of the wake lying between z and
z + dz is I(z) dz, so we obtain
Z b/2
Z
I(z ) dz
1
d(z )
1
=
,
(8.90)
w(z) =
4 b/2 z z
4
z z
where use has been made of (8.88).

Incompressible Aerodynamics

165

8.9 Ellipsoidal Airfoils


Consider an ellipsoidal airfoil whose outer surface is specified by the parametric equations
x

c0
sin cos ,
2
0
sin sin ,
2
b
cos ,
2

z =

(8.91)
(8.92)
(8.93)

where 0 and 0 2. Here, b is the wingspan, c0 the maximum wing width, and 0 the maximum
wing thickness. Note that the wings cross-section is elliptical both in the x-y and the x-z planes. It is assumed that
b > c0 0 : i.e., the wingspan is greater than the wing width, which in turn is much greater than the wing thickness.
At fixed (i.e., fixed z), the width and thickness of the airfoil are c() = c0 sin and () = 0 sin , respectively.
Assuming that the two-dimensional result (8.44) holds at fixed z, we deduce that the air circulation about the wing
satisfies
(z) = V c(z) sin = 0 sin ,
(8.94)
where
0 V c0 .

(8.95)

Here, the angle of attack, , is assumed to be small. From Equations (8.90) and (8.94), the downwash velocity in the
region |z| < b/2 is given by
!
Z
Z
cos
0
cos d
d
0
1
+
.
(8.96)
=
w() =

2 b 0 cos cos 2 b

0 cos cos
Now, the integrand in the integral

d
cos cos

(8.97)

is singular when = . However, we can still obtain a finite value for the integral by taking its principal part: i.e.,
lim

d
+
cos cos

!
d
.
cos cos

(8.98)

Physically, this is equivalent to omitting the contribution of the local free vortex at a given point on the airfoils trailing
edge to the downwash velocity induced at that point, which is reasonable because a vortex induces zero velocity at its
center. Hence, we obtain

#!
#!+
"
"
Z

d
1
sin (1/2) ( + )
sin (1/2) ( + )

1
+
=
lim
ln
ln

0
sin
sin (1/2) ( ) 0
sin
sin (1/2) ( )
0 cos cos
"
#!
sin( /2)
1
ln
= 0,
(8.99)
= lim
0 sin
sin( + /2)
which implies that
w() =

0
.
2b

(8.100)

In the region |z| > b/2, we can write = 2 z/b, so that

0
1
w() =
2b

!
||
0
d
=
1 p
.
cos +
2b
2 1

(8.101)

166

FLUID MECHANICS

1.2
0.8

w/(0/2 b)

0.4
0
0.4
0.8
1.2
1.6
2
2

0
z/b)

Figure 8.13: Downwash velocity profile induced at the trailing edge by an ellipsoidal airfoil.
Hence, we conclude that the downwash velocity profile induced by an ellipsoidal airfoil takes the form
(
|z| < b/2
0 1
.
w(z) =
2 b 1 |z|/(z2 b2 /4)1/2
|z| > b/2

(8.102)

This profile is shown in Figure 8.13. It can be seen that the downwash velocity is uniform and positive in the region
between the wingtips (i.e., b/a < z < b/2), but negative and decaying in the region outside the wingtips. Hence, we
conclude that as air passes over an airfoil subject to an upward lift it acquires a net downward velocity component,
which, of course, is a consequence of the reaction to the lift. On the other hand, the air immediately behind and to the
sides of the airfoil acquires a net upward velocity component. In other words, the lift acting on the airfoil is associated
with a downwash of air behind, and an upwash behind and to either side of, the airfoil. The existence of upwash
slightly behind and to the side of a flying object allows us to explain the V-formation adopted by wild geesea bird
flying in the upwash of another bird needs to generate less lift in order to stay in the air, and, consequently, experiences
less induced drag.
It follows from Equation (8.93) and (8.94) that
Z b/2
Z
0 b 2

(z) dz =
sin d = 0 b.
(8.103)
2
4
b/2
0
Hence, Equation (8.80), (8.82), and (8.100) yield the following expression for the lift and induced drag acting on an
ellipsoidal airfoil,
L

V b 0 ,
4

02 .
8

(8.104)
(8.105)

Now, the surface area of the airfoil in the x-z plane is


S =

b c0 .
4

(8.106)

Incompressible Aerodynamics

167

L
y
airfoil

V
x

D
W

Figure 8.14: Side view of a fixed wing aircraft in flight.


Moreover, the airfoils aspect-ratio is conventionally defined as the length to width ratio for a rectangle of length b
that has the same area as the airfoil: i.e.,
b2 4 b
A=
.
(8.107)
=
S
c0
It thus follows from Equation (8.95) that
L =
D

L0 ,
2
L0 2 ,
A

(8.108)
(8.109)

where
L0 = V 2 S .

(8.110)

8.10 Simple Flight Problems


Figure 8.14 shows a side-view schematic of a fixed wing aircraft flying in a straight-line at constant speed through
stationary air. Here, x is a horizontal coordinate, and y a vertical coordinate. The center of mass of the aircraft
is assumed to be moving with some fixed velocity V that subtends an angle with the horizontal. Thus, the wind
velocity in the aircrafts rest frame is V. Let the aircrafts wings, which are assumed to be parallel to its fuselage,
be inclined at an angle + to the horizontal. It follows that is the angle of attack. The aircraft is subject to four
forces: the thrust, T, developed by its engine, which is assumed to act parallel to its fuselage; the lift L, which acts
at right-angles to V; the induced drag, D, which acts in the opposite direction to V; and the weight, W, which acts
vertically downward.
Vertical force balance yields
T sin( + ) + L cos = W + D sin ,
(8.111)
whereas horizontal force balance gives
T cos( + ) = D cos + L sin .

(8.112)

168

FLUID MECHANICS

Let us assume that the angles and are both small. According to Equations (8.108) and (8.109), L L0 and
D (2 L0 /A) 2 . Thus, L O() and D O(2 ). Moreover, it is clear from (8.111) and (8.112) that T O(2 ) and
W O(). Thus, to lowest order in , Equation (8.111) yields

W
,
L0

(8.113)

whereas Equation (8.112) gives


2 W
T
.
(8.114)

W A L0
Expression (8.113) relates the angle of attack to the ratio of the aircrafts weight to its (theoretical) maximum lift (at
a given airspeed). Expression (8.114) relates the aircrafts angle of controlled (i.e., at constant airspeed) ascent to the
thrust developed by its engine. Now an unpowered aircraft, such as a glider, has zero thrust. For such an aircraft,
(8.114) reveals that the angle of controlled decentwhich is usually termed the glide angletakes the value

g=

2 W
.
A L0

(8.115)

At fixed airspeed, V, and wing surface area, S , (which implies that L0 is fixed) this angle can be minimized by making
the wing aspect-ratio, A, as large as possible. This result explains accounts for the fact that gliders (and albatrosses)
have long thin wings, rather than short stubby ones. For a powered aircraft, the critical thrust to weight ratio required
to maintain level flight (i.e., = 0) is
T
= g.
(8.116)
W
Hence, this ratio is minimized by minimizing the glide angle, which explains why long-haul aircraft, which generally
need to minimize fuel consumption, tend to have long thin wings. Finally, as we saw in Section 8.4, if the angle of
attack exceeds some (generally small) critical value c then boundary layer separation occurs on the back sides of
the wings, giving rise to a greatly increased level of drag acting on the aircraft. In aerodynamics, this phenomenon is
called a stall. As is clear from Equations (8.110) and (8.113), the requirement < c is equivalent to
!1/2
W
.
(8.117)
V > Vs =
S c
In other words, a stall can be avoided by keeping the airspeed above the critical value V s , which is known as the stall
speed. Note that the stall speed decreases with decreasing altitude, as the air becomes denser.

8.11 Exercises
8.1. Consider the integral

cos(n ) d
,

0 cos cos
where n is a non-negative integer. This integral is defined by its principal value
"Z
#
Z
cos(n ) d
cos(n ) d
In () = lim
+
.

0
cos cos
0
+ cos cos
In () =

As was demonstrated in Section 8.9,


I0 = 0.
Show that
I1 = ,
and
In+1 + In1 = 2 cos In ,
and hence that
In =

sin(n )
.
sin

Incompressible Aerodynamics

169

8.2. Suppose that an airfoil of negligible thickness, and wingspan b, has a width whose z variation is expressed parametrically as
X
c sin( ),
c() =
=1,3,5,

for 0 , where
z=

b
cos .
2

Show that the air circulation about the airfoil takes the form
X
sin( ),
() =
=1,3,5,

where = V c . Here, is the angle of attack (which is assumed to be small). Demonstrate that the downwash velocity
at the trailing edge of the airfoil is
X sin( )
w() =
.
2 b sin
=1,3,5,
Hence, show that the lift and induced drag acting on the airfoil take the values
L

V b 1 ,
4
X

2,

8 =1,3,5,

respectively. Demonstrate that the drag to lift ratio can be written

X
D
2
= 1 +
L
A

=3,5,7,

c2
,
c12

where A is the aspect ratio. Hence, deduce that the airfoil shape (in the x-y) plane that minimizes this ratio (at fixed aspect
ratio) is an ellipse (i.e., such that c = 0 for > 1).

8.3. Consider a plane that flies with a constant angle of attack, and whose thrust is adjusted such that it cancels the induced drag.
The plane is effectively subject to two forces. First, its weight, W = W ey , and second its lift L = k v vy ex + k v vx ey .
Here, x and y are horizontal and vertical coordinates, respectively, v is the planes instantaneous velocity, and k is a positive
constant. Note that the lift is directed at right angles to the planes instantaneous direction of motion, and has a magnitude
proportional to the square of its airspeed. Demonstrate that the planes equations of motion can be written
dvx
dt
dvy
dt

=
=

v vy
,
h
v vx
g,
h

where h = k g/W is a positive constant with the dimensions of length. Show that
1 2
v + g y = E,
2
p
p
where E is a constant. Suppose that vx = g h (1 + u) and vy = g h w, where |u|, |w| 1. Demonstrate that, to first order in
perturbed quantities,
r
g
du

w,
dt
h
r
g
dw
2
u.
dt
h
Hence, deduce that if the plane is flying horizontally
at some speed v0 , and is subject to a small perturbation, then its altitude
oscillates sinusoidally at the angular frequency = 2 g/v0 . This type of oscillation is known as a phugoid oscillation.

170

FLUID MECHANICS

Incompressible Viscous Flow

171

9 Incompressible Viscous Flow

9.1 Introduction
This chapter investigates incompressible flow in which viscosity plays a significant role throughout the bulk of the
fluid. Such flow generally takes place at relatively low Reynolds number. From Section 2.14, the equations governing
incompressible viscous fluid motion can be written
v
Dv

Dt

0,

(9.1)

P + 2 v,

(9.2)

where the quantity


P = p + ,

(9.3)

which is a combination of the actual fluid pressure, p, and the gravitational potential energy per unit volume, , is
known as the effective pressure. Here, is the fluid density, the fluid viscosity, and the gravitational potential.

9.2 Flow Between Parallel Plates


Consider steady, two-dimensional, viscous flow between two parallel plates that are situated a perpendicular distance
d apart. Let x be a longitudinal coordinate measuring distance along the plates, and let y be a transverse coordinate
such that the plates are located at y = 0 and y = d. See Figure 9.1.
Suppose that there is a uniform effective pressure gradient in the x-direction, so that
dP
= G,
dx

(9.4)

where G is a constant. Here, the quantity G could represent a gradient in actual fluid pressure, a gradient in gravitational
potential energy (due to an inclination of the plates to the horizontal), or some combination of the twoit actually
makes no difference to the final result. Suppose that the fluid velocity profile between the plates takes the form
v = v x (y) e x .

(9.5)

From Section 2.18, this profile automatically satisfies the incompressibility constraint v = 0, and is also such that
Dv/Dt = 0. Hence, Equation (9.2) reduces to
P
,
(9.6)
2 v =

y=d
P

vx(y)

y=0
Figure 9.1: Viscous flow between parallel plates.

172

FLUID MECHANICS

Figure 9.2: Viscous flow down an inclined plane.


or. taking the x-component,
G
d2v x
= .

dy2

(9.7)

If the two plates are stationary then the solution that satisfies the no slip constraint, v x (0) = v x (d) = 0, at each plate is
v x (y) =

G
y (d y).
2

(9.8)

Thus, steady, two-dimensional, viscous flow between two stationary parallel plates is associated with a parabolic
velocity profile that is symmetric about the midplane, y = d/2. The net volume flux (per unit width in the z-direction)
of fluid between the plates is
Z d
Gd3
Q=
v x dy =
.
(9.9)
12
0
Note that this flux is directly proportional to the effective pressure gradient, inversely proportional to the fluid viscosity,
and increases as the cube of the distance between the plates.
Suppose that the upper plate is stationary, but that the lower plate is moving in the x-direction at the constant speed
U. In this case, the no slip boundary condition at the lower plate becomes v x (0) = U, and the modified solution to
Equation (9.7) is
!
G
dy
.
(9.10)
v x (y) =
y (d y) + U
2
d
Hence, the modified velocity profile is a combination of parabolic and linear profiles. This type of flow is known as
Couette flow. The net volume flux (per unit width) of fluid between the plates becomes
Q=

Gd3 1
+ U d.
12 2

(9.11)

9.3 Flow Down an Inclined Plane


Consider steady, two-dimensional, viscous flow down a plane that is inclined at an angle to the horizontal. Let x
measure distance along the plane, and let y be a transverse coordinate such that the surface of the plane corresponds
to y = 0. Suppose that the fluid forms a uniform layer of depth h covering this surface. See Figure 9.2.

Incompressible Viscous Flow

173

The generalized pressure gradient within the fluid is written


dP
= G = g sin ,
dx

(9.12)

where g is the acceleration due to gravity. In this case, there is no gradient in the actual pressure in the x-direction,
and the flow down the plane is driven entirely by gravity. As before, we can write
v = v x (y) e x ,

(9.13)

d2v x
G
= ,

dy2

(9.14)

G = g sin .

(9.15)

and Equation (9.2) again reduces to

where
Application of the no slip condition at the surface of the plane, y = 0, yields the standard boundary condition v x (0) = 0.
However, the appropriate
physical constraint at the fluid/air interface, y = h, is that the normal viscous stress there

be zero (i.e., xy y=h = 0), since there is nothing above the interface that can exchange momentum with the fluid
(assuming that the finite inertia and viscosity of air are both negligible.) Hence, from Section 2.18, we get the boundary
condition

dv x
= 0.
(9.16)
dy y=h
The solution to Equation (9.14) that satisfies the boundary conditions is
v x (y) =

G
y (2 h y).
2

(9.17)

Thus, the profile is again parabolic. In fact, it is the same as the lower half of the profile obtained when fluid flows
between two (stationary) parallel plates situated a perpendicular distance 2 h apart.
The net volume flux (per unit width in the z-direction) of fluid down the plane is
Z h
G h 3 g sin h 3
v x dy =
Q=
=
,
(9.18)
3
3
0
where use has been made of Equation (9.15). Here, = / is the kinematic viscosity of the fluid. Thus, given the
rate Q that fluid is poured down the plane, the depth of the layer covering the plane becomes
!1/3
3Q
.
(9.19)
h=
g sin
Suppose that the rate at which fluid is poured down the plane is suddenly increased slightly from Q to Q + Q. We
would expect an associated change in depth of the layer covering the plane from h to h + h, where
!
dh
h =
Q.
(9.20)
dQ
Let the interface between the layers of different depth propagate in the x-direction at the constant velocity V. In a
frame of reference that co-moves with this interface, the volume fluxes (per unit width) immediately to the right and
to the left of the interface are Q V h and Q + Q V (h + h), respectively. However, in a steady state, these fluxes
must equal one another. Hence,
Q = V h,
(9.21)
or
dQ G h 2
9 Q 2 g sin
V=
=
=
dh

!1/3

As can easily be verified, this velocity is twice the maximum fluid velocity in the layer.

(9.22)

174

FLUID MECHANICS

9.4 Poiseuille Flow


Steady viscous fluid flow driven by an effective pressure gradient established between the two ends of a long straight
pipe of uniform circular cross-section is generally known as Poiseuille flow, since it was first studied experimentally by
J. L. M. Poiseuille in 1838. Suppose that the pipe is of radius a. Let us adopt cylindrical coordinates whose symmetry
axis coincides with that of the pipe. Thus, z measures distance along the pipe, r = 0 corresponds to the center of the
pipe, and r = a corresponds to the pipe wall. Suppose that
P = G ez

(9.23)

is the uniform effective pressure gradient along the pipe, and


v = vz (r) ez

(9.24)

the time independent velocity profile driven by this gradient. It follows from Section 2.19 that v = 0 and Dv/Dt = 0.
Hence, (9.2) reduces to
G
(9.25)
2 v = ez .

Taking the z-component of this equation, we obtain


!
G
1 d dvz
r
= ,
r dr dr

(9.26)

where use has been made of (2.155). The most general solution of the above equation is
vz (r) =

G 2
r + A ln r + B,
4

(9.27)

where A and B are arbitrary constants. The physical constraints are that the flow velocity is non-singular at the center
of the pipe (which implies that A = 0), and is zero at the edge of the pipe [i.e., vz (a) = 0], in accordance with the no
slip condition. Thus, we obtain
G
vz (r) =
(a 2 r 2 ).
(9.28)
4
The volume flux of fluid down the pipe is
Q=

2 r vz dr =

G a4
.
8

(9.29)

According to the above analysis, the quantity Q/a 4 should be directly proportional to the effective pressure gradient
along the pipe. The accuracy with which experimental observations show that this is indeed the case (at relatively low
Reynolds number) is strong evidence in favor of the assumptions that there is no slip at the pipe walls, and that the
flow is non-turbulent. In fact, the result (9.29), which is known as Poiseuilles law, is valid experimentally provided
the Reynolds number of the flow, Re = U a/, remains less than about 6.5 103 . Here, U = Q/ a 2 is the mean
flow speed. On the other hand, if the Reynolds number exceeds the critical value 6.5 103 then the flow in the pipe
becomes turbulent, and Poiseuilles law breaks down.

9.5 Taylor-Couette Flow


Consider two thin cylindrical shells with the same vertical axis. Let the inner and outer shells be of radius r1 and r2 ,
respectively. Suppose that the annular region r1 r r2 is filled with fluid of density and viscosity . Let the inner
and outer cylinders rotate at the constant angular velocities 1 and 2 , respectively. We wish to determine the steady
flow pattern set up within the fluid. Incidentally, this type of flow is generally known as Taylor-Couette flow.

Incompressible Viscous Flow

175

It is convenient to adopt cylindrical coordinates, r, , z, whose symmetry axis coincides with the common axis of
the two shells. Thus, the inner and outer shells correspond to r = r1 and r = r2 , respectively. Suppose that the flow
velocity within the fluid is written
v = v (r) e = r (r) e ,
(9.30)
where (r) = v (r)/r is the angular velocity profile. Application of the no slip condition at the two shells leads to the
boundary conditions
(r1 ) =

1 ,

(9.31)

(r2 ) =

2 .

(9.32)

It again follows from Section 2.19 that v = 0 and Dv/Dt = 0. Hence, (9.2) reduces to
2 v =

P
.

(9.33)

Assuming that P = 0 within the fluid, since any flow is driven by the angular rotation of the two shells, rather than by
pressure gradients or gravity, and again making use of the results quoted in Section 2.19, the above expression yields
!
v
1 d v
r
2 = 0,
(9.34)
r dr dr
r
or

!
1 d 3 d
r
= 0.
dr
r 2 dr

(9.35)

The solution of (9.35) that satisfies the boundary conditions is

1 1 2 2 r12 1 r22

.
(r) = 2 2
+
r r1 r22
r12 r22

(9.36)

Note that this angular velocity profile is a combination of the solid body rotation profile = constant, and the
irrotational rotation profile r2 .
From Section 2.19, the only non-zero component of the viscous stress tensor within the fluid is
d  v 
d
r = r
= r
.
(9.37)
dr r
dr
Thus, the viscous torque (acting in the -direction) per unit height (in the z-direction) exerted on the inner cylinder is

1 2
.
1 = 2 r12 r (r1 ) = 4 2
(9.38)
r1 r22

Likewise, the torque per unit height exerted on the outer cylinder is

1 2
2
.
2 = 2 r2 r (r2 ) = 4 2
r1 r22

(9.39)

As expected, these two torques are equal and opposite, and act to make the two cylinders rotate at the same angular
velocity (in which case, the fluid between them rotates as a solid body).

9.6 Flow in Slowly-Varying Channels


According to Section 9.1, the equations governing steady, incompressible, viscous fluid flow are
v =
(v ) v =

0,

(9.40)
2

P + v.

(9.41)

176

FLUID MECHANICS

As we saw in Sections 9.2 and 9.4, for the case of flow along a straight channel of uniform cross-section, v and
(v ) v are both identically zero, and the governing equations consequently reduce to the simple relation
2 v =

P
.

(9.42)

Suppose, however, that the cross-section of the channel varies along its length. As we shall demonstrate, provided this
variation is sufficiently slow, the flow is still approximately described by the above relation.
Consider steady, two-dimensional, viscous flow, that is predominately in the x-direction, between two plates that
are predominately parallel to the y-z plane. Let the spacing between the plates, d(x), vary on some length scale l d.
Suppose that
P = G(x) e x ,
(9.43)
where G(x) also varies on the same length scale. Assuming that /x O(1/l) and /y O(1/d), it follows from
Equation (9.40) that
!
vy
d
.
(9.44)
O
vx
l
Hence,
[(v ) v] x

(2 v) x

!
v x2
,
l

!2
2 v x
d
.
1 + O
l
y2

The x-component of Equation (9.41) reduces to

!
!2
d
2 v x
G
vx d 2
+O
= .
1 + O
2
l
l

(9.45)
(9.46)

(9.47)

Thus, if

d
l

1,

(9.48)

vx d 2
l

(9.49)

i.e., if the channel is sufficiently narrow, and its cross-section varies sufficiently slowly along its lengththen
Equation (9.47) can be approximated as
G
2 v x
.
(9.50)

y2
This, of course, is the same as the equation governing steady, two-dimensional, viscous flow between exactly parallel
plates. See Section 9.2. Assuming that the plates are located at y = 0 and y = d(x), and making use of the analysis of
Section 9.2, the appropriate solution to the above equation is
v x (x, y) =

G(x)
y [d(x) y].
2

The volume flux (per unit width) of fluid between the plates is thus
Z d
G(x) d 3 (x)
Q=
v x dy =
.
12
0

(9.51)

(9.52)

However, for steady incompressible flow, this flux must be independent of x, which implies that
G(x) =

dP
= 12 Q d 3(x).
dx

(9.53)

Incompressible Viscous Flow

177

Suppose that a constant difference in effective pressure, P, is established between the fixed points x = x1 and x = x2 ,
where x2 x1 = l. Integration of the above equation between these two points yields
R x2

P = 12 Q l hd 3 i,

(9.54)

where h i = x ( ) dx/l. Hence, the volume flux (per unit width) of fluid between the plates that is driven by the
1
effective pressure difference becomes
1
P
.
(9.55)
Q=
l 12 hd 3 i
Moreover, the effective pressure gradient at a given point is
G(x) =

dP P
1
,
=
3
dx
l d (x) hd 3 i

(9.56)

which allows us to determine the velocity profile at that point from Equation (9.51). Thus, given the average effective
pressure gradient, P/l, and the variable separation, d(x), we can fully specify the flow between the plates.
Using analogous arguments to those employed above, but adapting the analysis of Section 9.4, rather than that of
Section 9.1, we can easily show that steady viscous flow down a straight pipe of circular cross-section, whose radius
a varies slowly with distance, z, along the pipe, is characterized by
vz (r, z) =
G(z) =
Q

i
G(z) h 2
a (z) r 2 ,
4
1
P
,
l a 4 (z) ha 4 i
P

l 8 ha 4 i

(9.57)

Here, Q is the volume flux of fluid down the pipe, P = P(z2 ) P(z1 ), l = z2 z1 , and h i =
approximations used to derive the above results are valid provided

(9.58)
(9.59)
R z2
z1

( ) dz/l. The

a
l

1,

(9.60)

vz a 2
l

1.

(9.61)

9.7 Lubrication Theory


It is well-known that two solid bodies can slide over one another particularly easily when there is a thin layer of
fluid sandwiched between them. Moreover, under certain circumstances, a large positive pressure develops within the
layerthis phenomenon is exploited in hydraulic bearings, whose aim is to substitute fluid-solid friction for the much
larger friction that acts between solid bodies that are in direct contact with one another. Once set up, the fluid layer in
hydraulic bearings offers great resistance to being squeezed out, and is often capable of supporting a useful load.
Consider the simple two-dimensional case of a solid body with a plane surface (that is almost parallel to the x-z
plane) gliding steadily over another such body, the surface of the gliding body being of finite length l in the direction
of the motion (the x-direction), and of infinite width (in the z-direction). See Figure 9.3. Experience shows that the
plane surfaces need to be slightly inclined to one another. Suppose that 1 is the angle of inclination. Let us
transform to a frame of reference in which the upper body is stationary. In this frame, the lower body moves in the
x-direction at some fixed speed U. Suppose that the upper body extends from x = 0 to x = l, and that the surface of
the lower body corresponds to y = 0. Let d(x) be the thickness (in the y-direction) of the fluid layer trapped between
the bodies, where d(0) = d1 and d(l) = d2 . It follows that
d(x) = d1 x,

(9.62)

178

FLUID MECHANICS

y
d1

d2

l
Figure 9.3: The lubrication layer between two planes in relative motion.
where
=

d1 d2
.
l

(9.63)

As discussed in the previous section, provided that


d
l

1,

(9.64)

U d2
l

1,

(9.65)

the cross-section of the channel between the two bodies is sufficiently slowly varying in the x-direction that the channel
can be treated as effectively uniform at each point along its length. Thus, it follows from Equation (9.10) that the
velocity profile within the channel takes the form
#
"

d(x) y
G(x) 
,
(9.66)
y d(x) y + U
v x (x, y) =
2
d
where

dp
(9.67)
dx
is the pressure gradient. Here, we are neglecting gravitational forces with respect to both pressure and viscous forces.
The volume flux per unit width (in the z-direction) of fluid along the channel is thus
Z d
G(x) d 3 (x) 1
v x dy =
Q=
+ U d(x).
(9.68)
12
2
0
G(x) =

Of course, in a steady state, this flux must be independent of x. Hence,


"
#
dp
U
2Q
= G(x) = 6 2
,

dx
d (x) d 3 (x)
where d(x) = d1 x. Integration of the above equation yields

!
1
1
1
6 1

p(x) p0 =
Q 2 2 ,
U

d d1
d
d1

(9.69)

(9.70)

where p0 = p(0). Now, assuming that the sliding block is completely immersed in fluid of uniform ambient pressure
p0 , we would expect the pressures at the two ends of the lubricating layer to both equal p0 , which implies that p(l) = p0 .
It follows from the above equation that
!
d1 d2
,
(9.71)
Q=U
d1 + d2

Incompressible Viscous Flow

179

and

6 U [d1 d(x)] [d(x) d2 ]


.
(9.72)

d 2 (x) (d1 + d2 )
Note that if d1 > d2 then the pressure increment p(x) p0 is positive throughout the layer, and vice versa. In other
words, a lubricating layer sandwiched between two solid bodies in relative motion only generates a positive pressure,
that is capable of supporting a normal load, when the motion is such as to drag (by means of viscous stresses) fluid
from the wider to the narrower end of the layer. The pressure increment has a single maximum in the layer, and its
value at this point is of order l U/d 2 , assuming that (d1 d2 )/d1 is of order unity. This suggests that very large
pressures can be set up inside a thin lubricating layer.
The net normal force (per unit width in the z-direction) acting on the lower plane is
"
!
!#
Z l
6U
d1
d1 d2
fy = [p(x) p1 ] dx = 2 ln
2
.
(9.73)

d2
d1 + d2
0
p(x) p0 =

Moreover, the net tangential force (per unit width) acting on the lower plane is
!
"
!
!#
Z l
d1
2U
d1 d2
v x
2 ln
3
.
dx =
fx =

y y=0

d2
d1 + d2
0

(9.74)

Of course, equal and opposite forces,


f x

fy

!
#
"
U l2 3
1+k
2k ,
ln
fx =
1k
d02 2 k2
"
!
#
U l 1
1+k
fy =
2 ln
3k ,
d0 k
1k

(9.75)
(9.76)

act on the upper plane. Here, d0 = (d1 + d2 )/2 is the mean channel width, and k = (d1 d2 )/(d1 + d2 ). Note that if
0 < d2 < d1 then 0 < k < 1. The effective coefficient of friction, C f , between the two sliding bodies is conventionally
defined as the ratio of the tangential to the normal force that they exert on one another. Hence,
Cf =
where

f 4 d0
fx
= x =
H(k),
fy
fy
3 l

"
!
#, "
!
#
1+k
1+k
3k
H(k) = k ln

ln
2k .
1k
2
1k

(9.77)

(9.78)

The function H(k) is a monotonically decreasing function of k in the range 0 < k < 1. In fact, H(k 0) 3/(4 k),
whereas H(k 1) 1. Thus, if k O(1) [i.e., if (d1 d2 )/d1 O(1)] then C f O(d0 /l) 1. In other words, the
effective coefficient of friction between two solid bodies in relative motion that are separated by a thin fluid layer is
independent of the fluid viscosity, and much less than unity. This result is significant because the coefficient of friction
between two solid bodies in relative motion that are in direct contact with one another is typical of order unity. Hence,
the presence of a thin lubricating layer does indeed lead to a large reduction in the frictional drag acting between the
bodies.

9.8 Stokes Flow


Steady flow in which the viscous force density in the fluid greatly exceeds the advective inertia per unit volume is
generally known as Stokes flow. Since, by definition, the Reynolds number of a fluid is the typical ratio of the advective
inertia per unit volume to the viscous force density (see Section 2.16), Stokes flow implies Reynolds numbers that are
much less than unity. Now, in the time independent, low Reynolds number limit, Equations (9.1) and (9.2) reduce to
v =
0 =

0,

(9.79)
2

P + v.

(9.80)

180

FLUID MECHANICS

It follows from these equations that


P = 2 v = ( v) = ,

(9.81)

where = v, and use has been made of Equation (A.177). Taking the curl of this expression, we obtain
2 = 0,

(9.82)

which is the governing equation for Stokes flow. Here, use has been made of Equations (A.173), (A.176), and (A.177).

9.9 Axisymmetric Stokes Flow


Let r, , be standard spherical coordinates. Consider axisymmetric Stokes flow such that
v(r) = vr (r, ) er + v (r, ) e .

(9.83)

According to Equations (A.175) and (A.176), we can automatically satisfy the incompressibility constraint (9.79) by
writing
v = ,
(9.84)
where (r, ) is a stream function (i.e., v = 0). It follows that
vr (r, )

v (r, )

1
,
sin
1
.
r sin r

r2

(9.85)
(9.86)

Moreover, according to Section C.4, r = = 0, and


1 (r v ) 1 vr
L()

=
,
r r
r
r sin

(r, ) =

(9.87)

where
L=

2
sin 1
+ 2
.
r2
r sin

(9.88)

Hence, given that || = 1/(r sin ), we can write


= v = L() .

(9.89)

It follows from Equations (A.176) and (A.178) that


= [L()].

(9.90)

Hence, by analogy with Equations (9.84) and (9.89), and making use of (A.173) and (A.177), we obtain
( ) = 2 = L 2 () .

(9.91)

L 2 () = 0,

(9.92)

Equation (9.82) implies that


which is the governing equation for axisymmetric Stokes flow. In addition, Equations (9.81) and (9.90) yield
P = [L()].

(9.93)

Incompressible Viscous Flow

181

9.10 Axisymmetric Stokes Flow Around a Solid Sphere


Consider a solid sphere of radius a that is moving under gravity at the constant vertical velocity V ez through a stationary fluid of density and viscosity . Here, gravitational acceleration is assumed to take the form g = g ez . Now,
provided the typical Reynolds number,
2V a
,
(9.94)
Re =

is much less than unity, the flow around the sphere is an example of axisymmetric Stokes flow. Let us transform to
a frame of reference in which the sphere is stationary, and centered at the origin. Adopting the standard spherical
coordinates r, , , the surface of the sphere corresponds to r = a, and the surrounding fluid occupies the region r > a.
By symmetry, the flow field outside the sphere is axisymmetric (i.e., / = 0), and has no toroidal component (i.e.,
v = 0). The physical boundary conditions at the surface of the sphere are
vr (a, ) =

0,

(9.95)

v (a, ) =

0:

(9.96)

i.e., the normal and tangential fluid velocities are both zero at the surface. A long way from the sphere, we expect the
fluid velocity to asymptote to v = V ez . In other words,
vr (r , )

V cos ,

(9.97)

v (r , )

V sin .

(9.98)

Let us write
v = ,

(9.99)

L 2 () = 0.

(9.100)

where (r, ) is the stream function. As we saw in the previous section, axisymmetric Stokes flow is characterized by

Here, the differential operator L is specified in Equation (9.88). The boundary conditions (9.95)(9.98) reduce to



= 0,
(9.101)
r r=a



= 0,
(9.102)
r=a
1
V r 2 sin2 .
(9.103)
(r , )
2
Equation (9.103) suggests that (r, ) can be written in the separable form
(r, ) = sin2 f (r).

(9.104)

In this case,
vr (r, )

v (r, )

2 cos f (r)
,
r2
sin d f
,
r dr

(9.105)
(9.106)

and Equations (9.100)(9.103) reduce to


d2
2

dr 2 r 2
f (a) =

!2


d f

dr r=a

f (r )

0,

(9.107)

0,

(9.108)

1
V r 2.
2

(9.109)

182

FLUID MECHANICS

3
2
1
z/a

0
1
2
3
3

0
x/a

Figure 9.4: Contours of the stream function in the x-z plane for Stokes flow around a solid sphere.
Let us try a test solution to Equation (9.107) of the form f (r) = r n . We find that
[n (n 1) 2] [(n 2) (n 3) 2] = 0,

(9.110)

which implies that n = 1, 1, 2, 4. Hence, the most general solution to Equation (9.107) is
f (r) =

A
+ B r + C r 2 + D r 4,
r

(9.111)

where A, B, C, D are arbitrary constants. However, the boundary condition (9.109) yields C = (1/2) V and D = 0,
whereas the boundary condition (9.108) gives A = (1/4) V a 3 and B = (3/4) V a. Thus, we conclude that
f (r) =

V (r a)2 (2 r + a)
,
4r

and the stream function becomes


(r, ) = sin2

V (r a)2 (2 r + a)
.
4r

(9.112)

(9.113)

See Figure 9.4. From (9.87), the fluid vorticity is


!
L()
sin d2
3 V a sin
2
(r, ) =
=
f =

.
r sin
r
dr 2 r 2
2r2

(9.114)

See Figure 9.5. Moreover, from (9.81),


P = = ( r sin ).

(9.115)

Hence,
P
r
P

3 V a cos
,
r3
3 V a sin
,
=
2r2
=

(9.116)
(9.117)

Incompressible Viscous Flow

183

3
2
1
z/a

0
1
2
3
3

0
x/a

Figure 9.5: Contours of the vorticity, , in the x-z plane for Stokes flow around a solid sphere. Solid/dashed lines
correspond to opposite signs of .
which implies that the effective pressure distribution within the fluid is
P(r, ) = p0 +

3 V a cos
,
2r2

(9.118)

where p0 is an arbitrary constant. See Figure 9.6. However, P = p + , where = g z = g r cos . Thus, the actual
pressure distribution is
3V a
p(r, ) = p0 g r cos +
cos .
(9.119)
2r2
From Section 2.20, the radial and tangential components of the force per unit area exerted on the sphere by the
fluid are
!
vr
fr () = rr (a, ) = p + 2
,
(9.120)
r r=a
!
1 vr v v
f () = r (a, ) =
+

.
(9.121)
r
r
r r=a
Now, vr (a, ) = v (a, ) = 0. Moreover, since v = 0, it follows from (2.169) that (vr /r)r=a = 0. Finally,
Equation (9.87) yields (v /r)r=a = (a, ). Hence,
fr ()

p(a, ) = p0 + g a cos

f ()

(a, ) =

3V
cos ,
2a

3V
sin .
2a

(9.122)
(9.123)

Thus, the force density at the surface of the sphere is


f() =

3V
ez + (p0 + g a cos ) er .
2a

(9.124)

184

FLUID MECHANICS

3
2
1
z/a

0
1
2
3
3

0
x/a

Figure 9.6: Contours of the effective pressure, P p0 , in the x-z plane for Stokes flow around a solid sphere.
Solid/dashed lines correspond to opposite signs of P p0 .
It follows that the net vertical force exerted on the sphere by the fluid is
I
Z
3V
(p0 + g a cos ) cos sin d,
f ez dS =
4 a2 + 2 a2
Fz =
2a
S
0

(9.125)

which reduces to

4 3
a g.
(9.126)
3
By symmetry, the horizontal components of the net force both average to zero. We can recognize the second term on
the right-hand side of the above equation as the buoyancy force due to the weight of the fluid displaced by the sphere.
(See Chapter 3.) Moreover, the first term can be interpreted as the viscous drag acting on the sphere. Note that this
drag acts in the opposite direction to the relative motion of the sphere with respect to the fluid, and its magnitude is
directly proportional to the relative velocity.
Vertical force balance requires that
Fz = M g,
(9.127)
Fz = 6 a V +

where M is the spheres mass. In other words, in a steady state, the weight of the sphere balances the vertical force
exerted by the surrounding fluid. If the sphere is composed of material of mean density then M = (4/3) a 3 .
Hence, in the frame in which the fluid a large distance from the sphere is stationary, the steady vertical velocity with
which the sphere moves through the fluid is
!

2 a2 g
1
,
(9.128)
V=
9

where = / is the fluids kinematic viscosity. Obviously, if the sphere is more dense than the fluid (i.e., if / > 1)
then it moves downward (i.e., V < 0), and vice versa. Finally, the typical Reynolds number of the fluid flow in the
vicinity of the sphere is


2 V a 4 a3 g


.
Re =
1

=
(9.129)

9 2

Incompressible Viscous Flow

185

For the case of a grain of sand falling through water at 20 C, we have / 2 and = 1.0 106 m2 /s. Hence,
Re = (a/6 105 ) 3 , where a is measured in meters. Thus, expression (9.128), which is strictly speaking only valid
when Re 1, but which turns out to be approximately valid for all Reynolds numbers less than unity, only holds for
sand grains whose radii are less than about 60 microns. Such grains fall through water at approximately 8 103 m/s.
For the case of a droplet of water falling through air at 20 C and atmospheric pressure, we have / = 780 and
= 1.5 105 m2 /s. Hence, Re = (a/4 105 ) 3 , where a is measured in meters. Thus, expression (9.128) only holds
for water droplets whose radii are less than about 40 microns. Such droplets fall through air at approximately 0.2 m/s.
At large values of r/a, Equations (9.105), (9.106), and (9.112) yield

It follows that

and

vr (r, )

v (r, )

 a 2
3
a
,
V cos + O
2
r
r
 a 2
a
3
.
V sin V sin + O
4
r
r

V cos +

vr v vr v2 V 2 a
+

[ (v ) v]r = vr
,
r
r
r
r2
( 2 v)r

2 vr V a
3 .
r 2
r

(9.130)
(9.131)

(9.132)

(9.133)

Hence,
[ (v ) v]r V r
r
Re ,

a
( v)r

(9.134)

where Re is the Reynolds number of the flow in the immediate vicinity of the sphere. [See Equation (9.94).] Now, our
analysis is based on the assumption that advective inertia is negligible with respect to viscosity. However, as is clear
from the above expression for the ratio of inertia to viscosity within the fluid, even if this ratio is much less than unity
close to the spherein other words, if Re 1it inevitably becomes much greater than unity far from the sphere:
i.e., for r a/Re. In other words, inertia always dominates viscosity, and our Stokes flow solution therefore breaks
down, at sufficiently large r/a.

9.11 Axisymmetric Stokes Flow In and Around a Fluid Sphere


Suppose that the solid sphere discussed in the previous section is replaced by a spherical fluid drop of radius a. Let
the drop move through the surrounding fluid at the constant velocity V ez . Obviously, the fluid from which the drop is
composed must be immiscible with the surrounding fluid. Let us transform to a frame of reference in which the drop
is stationary, and centered at the origin. Assuming that the Reynolds numbers immediately outside and inside the drop
are both much less than unity, and making use of the previous analysis, the most general expressions for the stream
function outside and inside the drop are

+ Br + C r2 + Dr4 ,

(9.135)

(r, ) = sin2 + B r + C r 2 + D r 4 ,
r

(9.136)

(r, ) = sin2
and

A
r

respectively. Here, A, B, C, etc. are arbitrary constants. Likewise, the previous analysis also allows us to deduce that
vr (r, )
v (r, )


B
2
,
+
C
+
D
r
r3 r
 A

B
= sin 3 + + 2 C + 4 D r 2 ,
r
r

= 2 cos

A

(9.137)
(9.138)

186

FLUID MECHANICS

z/a

2
2

0
x/a

Figure 9.7: Contours of the stream function in the x-z plane for Stokes flow in and around a fluid sphere. Here,
/ = 10.
!
6A
+
6
D
r
,
r4

r (r, )

= sin

rr (r, )

12 A 6 B
+ 2 + 12 D r
= p0 + g r cos + cos
r4
r

(9.139)
!

(9.140)

in the region r > a, with analogous expressions in the region r < a. Here, , , p0 are the viscosity, density, and
ambient pressure of the fluid surrounding the drop. Let , , and p0 be the corresponding quantities for the fluid that
makes up the drop.
In the region outside the drop, the fluid velocity must asymptote to v = V ez at large r/a. This implies that
C = (1/2) V and D = 0. Furthermore, vr (a, ) = 0i.e., the normal velocity at the drop boundary must be zero
otherwise, the drop would change shape. This constraint yields A/a 3 + B/a + (1/2) V = 0.
Inside the drop, the fluid velocity must remain finite as r 0. This implies that A = B = 0. Furthermore, we
again require that vr (a, ) = 0, which yields C + D a 2 = 0.
Two additional physical constraints that must be satisfied at the interface between the two fluids are, firstly,
continuity of tangential velocityi.e., v (a , ) = v (a+ , )and, secondly, continuity of tangential stressi.e.,
r (a , ) = r (a+ , ). These constraints yield A/a 3 + B/a V = 2 C + 4 D a 2 and 6 A/a4 = 6 D a, respectively.
At this stage, we have enough information to determine the values of A, B, C, and D. In fact, the stream functions
outside and inside the drop can be shown to take the form
"
!
!
 r 2 #
1
2+4 r
a

2
2
(r, ) = V a sin

+2
,
(9.141)
4
+ r
+ a
a
and
(r, ) =

!
"
 r 2 #
 r 2
1
1
,
V a2 sin2
4
+ a
a

(9.142)

Incompressible Viscous Flow

187

z/a

2
2

0
x/a

Figure 9.8: Contours of the stream function in the x-z plane for Stokes flow in and around a fluid sphere. Here,
/ = 1/10.
respectively. See Figures 9.7 and 9.8.
The discontinuity in the radial stress across the drop boundary is
rr (a+ , ) rr (a , ) = p0 p0 + ( ) g a cos 3

"
#
V + (3/2)
cos .
a
+

(9.143)

The final physical constraint that must be satisfied at r = a is


rr (a+ , ) rr (a , ) =

2
,
a

(9.144)

where is the surface tension of the interface between the two fluids. (See Section 4.3.) Hence, we obtain
p0 p0 =
and

a2 g

V=
1
3

!"

2
,
a

(9.145)

#
+
,
+ (3/2)

(9.146)

where = / is the kinematic viscosity of the surrounding fluid. The fact that we have been able to completely satisfy
all of the physical constraints at the interface between the two fluids, as long as the drop moves at the constant vertical
velocity V, proves that our previous assumptions that the interface is spherical, and that the drop moves vertically
through the surrounding fluid at a constant speed without changing shape, were correct. In the limit, , in which
the drop is much more viscous than the surrounding fluid, we recover Equation (9.128): i.e., the drop acts like a solid
sphere. On the other hand, in the limit , and , which is appropriate to an air bubble rising through a liquid,
we obtain
a2 g
V=
.
(9.147)
3

188

FLUID MECHANICS

9.12 Exercises
9.1. Consider viscous fluid flow down a plane that is inclined at an angle to the horizontal. Let x measure distance along
the plane (i.e., along the path of steepest decent), and let y be a transverse coordinate such that the surface of the plane
corresponds to y = 0, and the free surface of the fluid to y = h. Show that within the fluid (i.e., 0 y h)
vx (y)

p(y)

g sin
y (2 h y),
2
p0 + g cos (h y),

where is the kinematic viscosity, the density, and p0 is atmospheric pressure.


9.2. If a viscous fluid flows along a cylindrical pipe of circular cross-section that is inclined at an angle to the horizontal show
that the flow rate is
a4
(G + g sin ) ,
Q=
8
where a is the pipe radius, the fluid viscosity, the fluid density, and G the pressure gradient.
9.3. Viscous fluid flows steadily, parallel to the axis, in the annular region between two coaxial cylinders of radii a and n a, where
n > 1. Show that the volume flux of fluid flow is
"
#
G a4 4
(n2 1)2
Q=
n 1
,
8
ln n
where G is the effective pressure gradient, and the viscosity. Find the mean flow speed.
9.4. Consider viscous flow along a cylindrical pipe of elliptic cross-section. Suppose that the pipe runs parallel to the z-axis, and
that its boundary satisfies
x2 y2
+
= 1.
a2 b2
Let
v = vz (x, y) ez .
Demonstrate that

!
G
2
2
vz = ,
+
x2 y2

where G is the effective pressure gradient, and the fluid viscosity. Show that
vz (x, y) =

G a2 b2 b2 x2 a2 y2
2
a2 + b2

is a solution of this equation that satisfies the no slip condition at the boundary. Demonstrate that the flow rate is
Q=

G a3 b3
.
4 a2 + b2

Finally, show that a pipe with an elliptic cross-section has lower flow rate than an otherwise similar pipe of circular crosssection that has the same cross-sectional area.
9.5. Consider a velocity field of the form
v(r) = r 2 (r) sin2 ,

where r, , are spherical coordinates. Demonstrate that this field satisfies the equations of steady, incompressible, viscous
fluid flow (neglecting advective inertia) with uniform pressure (neglecting gravity) provided that
!
d 4 d
r
= 0.
dr
dr
Suppose that a solid sphere of radius a, centered at the origin, is rotating about the z-axis, at the uniform angular velocity
0 , in a viscous fluid, of viscosity , that is stationary at infinity. Demonstrate that
(r) = 0

a3
,
r3

for r a. Show that the torque that the sphere exerts on the fluid is
= 8 0 a 3 .

Incompressible Viscous Flow

189

9.6. Consider a solid sphere of radius a moving through a viscous fluid of viscosity at the fixed velocity V = V ez . Let r, ,
be spherical coordinates whose origin coincides with the instantaneous location of the spheres center. Show that, if inertia
and gravity are negligible, the fluid velocity, and the radial components of the stress tensor, a long way from the sphere, are
vr

rr

3
a
V cos ,
2
r
a
3
V sin ,
4
r
9
a
p0 V cos 2 ,
2
r
0,

respectively. Hence, deduce that the net force exerted on the fluid lying inside a large spherical surface of radius r, by the
fluid external to the surface, is
F = 6 a V,
independent of the surface radius.

190

FLUID MECHANICS

Waves in Incompressible Fluids

191

10 Waves in Incompressible Fluids

10.1 Introduction
This chapter investigates low amplitude waves propagating through incompressible fluids.

10.2 Gravity Waves


Consider a stationary body of water, of uniform depth d, located on the surface of the Earth. This body is assumed
to be sufficiently small compared to the Earth that its unperturbed surface is approximately planar. Let the Cartesian
coordinate z measure vertical height, with z = 0 corresponding to the aforementioned surface. Suppose that a small
amplitude wave propagates horizontally through the water, and let v(r, t) be the associated velocity field.
Since water is essentially incompressible, its equations of motion are
v =

v
+ (v ) v =
t

0,

(10.1)

p g ez + 2 v,

(10.2)

where is the (uniform) density, the (uniform) viscosity, and g the (uniform) acceleration due to gravity. (See
Section 2.14.) Let us write
p(r, t) = p0 g z + p1 (r, t),
(10.3)
where p0 is atmospheric pressure, and p1 the pressure perturbation due to the wave. Of course, in the absence of the
wave, the water pressure a depth h below the surface is p0 + g h. (See Chapter 3.) Substitution into (10.2) yields

v
p1 + 2 v,
t

(10.4)

where we have neglected terms that are second-order in small quantities (i.e., terms of order v 2 ).
Let us also neglect viscosity, which is a good approximation provided that the wavelength is not ridiculously small.
[For instance, for gravity waves in water, viscosity is negligible as long as (2 /g)1/3 5 105 m.] It follows that

v
p1 .
t

(10.5)

Taking the curl of this equation, we obtain

0,
(10.6)
t
where = v is the vorticity. We conclude that the velocity field associated with the wave is irrotational. Consequently, the previous equation is automatically satisfied by writing

v = ,

(10.7)

where (r, t) is a velocity potential. (See Section 5.7.) However, from Equation (10.1), the velocity field is also
divergence-free. It follows that the velocity potential satisfies Laplaces equation,
2 = 0.

(10.8)

Finally, Equations (10.5) and (10.7) yield

.
(10.9)
t
We now need to derive the physical constraints that must be satisfied at the waters upper and lower boundaries. It
is assumed that the water is bounded from below by a solid surface located at z = d. Since the water must always
p1 =

192

FLUID MECHANICS

remain in contact with this surface, the appropriate physical constraint at the lower boundary is vz |z=d = 0 (i.e., the
normal velocity is zero at the lower boundary), or



= 0.
(10.10)
z z=d

The waters upper boundary is a little more complicated, since it is a free surface. Let represent the vertical displacement of this surface due to the wave. It follows that

.
= vz |z=0 =
(10.11)
t
z z=0
The appropriate physical constraint at the upper boundary is that the water pressure there must equal atmospheric
pressure, since there cannot be a pressure discontinuity across a free surface (in the absence of surface tensionsee
Section 10.11). Accordingly, from (10.3), we obtain
p0 = p0 g + p1 |z=0 ,

(10.12)

g = p1 |z=0 ,

(10.13)

or
which implies that
g




p1

=
,
= g
t
z z=0
t z=0

(10.14)

where use has been made of (10.11). The above expression can be combined with Equation (10.9) to give the boundary
condition



2
1
= g
(10.15)
.
z z=0
t2 z=0
Let us search for a wave-like solution of Equation (10.8) of the form

(r, t) = F(z) cos( t k x).

(10.16)

This solution actually corresponds to a propagating plane wave of wave vector k = k e x , angular frequency , and
amplitude F(z). Substitution into Equation (10.8) yields
d2 F
k2 F = 0,
dz2

(10.17)

whose independent solutions are exp(+k z) and exp(k z). Hence, a general solution to (10.8) takes the form
(x, z, t) = A e k z cos( t k x) + B ek z cos( t k x),

(10.18)

where A and B are arbitrary constants. The boundary condition (10.10) is satisfied provided that B = A exp(2 k d),
giving
i
h
(10.19)
(x, z, t) = A e k z + ek (z+2 d) cos( t k x),

The boundary condition (10.15) then yields




2 
1 + e2 k d cos( t k x).
A k 1 e2 k d cos( t k x) = A
g

(10.20)

which reduces to the dispersion relation


2 = g k tanh(k d).
The type of wave described in this section is known as a gravity wave.

(10.21)

Waves in Incompressible Fluids

193

10.3 Gravity Waves in Deep Water


Consider the so-called deep water limit,
k d 1,

(10.22)

in which the depth, d, of the water greatly exceeds the wavelength, = 2/k, of the wave. In this limit, the gravity
wave dispersion relation (10.21) reduces to
= (g k)1/2 ,
(10.23)
since tanh(x) 1 as x . It follows that the phase velocity of gravity waves in deep water is
vp =

 g 1/2
.
=
k
k

(10.24)

Note that this velocity is proportional to the square root of the wavelength. Hence, deep water gravity waves with
long wavelengths propagate faster than those with short wavelengths. Now, the phase velocity, v p = /k, is defined
as the propagation velocity of a plane wave with the definite wave number, k [and a frequency given by the dispersion
relation (10.23)]. Such a wave has an infinite spatial extent. A more realistic wave of finite spatial extent, with an
approximate wave number k, can be formed as a linear superposition of plane waves having a range of different wave
numbers centered on k. Such a construct is known as a wave pulse. As is well-known, wave pulses propagate at the
group velocity,
d
vg =
.
(10.25)
dk
For the case of gravity waves in deep water, the dispersion relation (10.23) yields
vg =

1  g 1/2 1
= vp.
2 k
2

(10.26)

In other words, the group velocity of such waves is half their phase velocity.
Let (r, t) be the displacement of a particle of water, found at position r and time t, due to the passage of a deep
water gravity wave. It follows that

= v,
(10.27)
t
where v(r, t) is the perturbed velocity. For a plane wave of wave number k = k e x , in the limit k d 1, Equation (10.19)
yields
(x, z, t) = A e k z cos( t k x).
(10.28)
Hence, [cf., Equations (10.45)(10.48)]
x (x, z, t)
z (x, z, t)

=
=

a e k z cos( t k x),
ae

kz

sin( t k x),

kz

sin( t k x),

v x (x, z, t)

ae

vz (x, z, t)

a e k z cos( t k x),

(10.29)
(10.30)
(10.31)
(10.32)

and
p1 = g z ,

(10.33)

where use has been made of Equations (10.7), (10.9), and (10.27). Here, a is the amplitude of the vertical oscillation at
the waters surface. According to Equations (10.29)(10.32), the passage of the wave causes a water particle located a
depth h below the surface to execute a circular orbit of radius a e k h about its equilibrium position. Note that the radius
of the orbit decreases exponentially with increasing depth. Furthermore, whenever the particles vertical displacement
attains a maximum value the particle is moving horizontally in the same direction as the wave, and vice versa. See
Figure 10.1.

194

FLUID MECHANICS

surface
z
x

z=0

Figure 10.1: Motion of water particles associated with a deep water gravity wave propagating in the x-direction.
Finally, if we define h(x, z, t) = z (x, z, t) z as the equilibrium depth of the water particle found at a given point
and time then Equations (10.3) and (10.33) yield
p(x, z, t) = p0 + g h(x, z, t).

(10.34)

In other words, the pressure at this point and time is the same as the unperturbed pressure calculated at the equilibrium
depth of the water particle.

10.4 Gravity Waves in Shallow Water


Consider the so-called shallow water limit,
k d 1,

(10.35)

in which the depth, d, of the water is much less than the wavelength, = 2/k, of the wave. In this limit, the gravity
wave dispersion relation (10.21) reduces to
= (g d)1/2 k,
(10.36)
since tanh(x) x as x 0. It follows that the phase velocities and group velocities of gravity waves in shallow water
all take the fixed value
v p = vg = (g d)1/2 ,
(10.37)
irrespective of wave number. We conclude thatunlike deep water wavesshallow water gravity waves are nondispersive in nature: i.e., waves pulses and plane waves all propagate at the same speed. Note, also, that the velocity
(10.37) increases with increasing water depth.
For a plane wave of wave number k = k e x , in the limit k d 1, Equation (10.19) yields
(x, z, t) = A [1 + k2 (z + d)2 /2] cos( t k x).

(10.38)

Hence, Equations (10.7) and (10.27) give [cf., Equations (10.45)(10.48)]


x (x, z, t)

a (k d)1 cos( t k x),

(10.39)

z (x, z, t)

a (1 + z/d) sin( t k x),

(10.40)

sin( t k x)

v x (x, z, t)

a (k d)

vz (x, z, t)

a (1 + z/d) cos( t k x).

(10.41)
(10.42)

Waves in Incompressible Fluids

195

Here, a is again the amplitude of the vertical oscillation at the waters surface. According to the above expressions,
the passage of a shallow water gravity wave causes a water particle located a depth h below the surface to execute an
elliptical orbit, of horizontal radius a/(k d), and vertical radius a (1 h/d), about its equilibrium position. Note that the
orbit is greatly elongated in the horizontal direction. Furthermore, its vertical radius decreases linearly with increasing
depth such that it becomes zero at the bottom (i.e., at h = d). As before, whenever the particles vertical displacement
attains a maximum value the particle is moving horizontally in the same direction as the wave, and vice versa.

10.5 Energy of Gravity Waves


It is easily demonstrated, from the analysis contained in the previous sections, that a gravity wave of arbitrary
wavenumber k, propagating horizontally through water of depth d, has a phase velocity
#1/2
"
1/2 tanh(k d)
.
(10.43)
v p = (g d)
kd
Moreover, the ratio of the group to the phase velocity is
"
#
vg 1
2kd
1+
.
=
vp 2
sinh(2 k d)

(10.44)

Note, that neither the phase velocity nor the group velocity of a gravity wave can ever exceed the critical value
(g d)1/2. It is also easily demonstrated that the displacement and velocity fields associated with a plane gravity wave
of wavenumber k e x , angular frequency , and surface amplitude a, are
x (x, z, t) =
z (x, z, t) =
v x (x, z, t) =
vz (x, z, t) =

cosh[k (z + d)]
cos( t k x),
sinh(k d)
sinh[k (z + d)]
sin( t k x),
a
sinh(k d)
cosh[k (z + d)]
a
sin( t k x),
sinh(k d)
sinh[k (z + d)]
a
cos( t k x).
sinh(k d)

Now, the mean kinetic energy per unit surface area associated with a gravity wave is defined
Z
1 2
v dzi,
K=h
2
d

(10.45)
(10.46)
(10.47)
(10.48)

(10.49)

where
(x, t) = a sin( t k x)

(10.50)

is the vertical displacement at the surface, and


h i =

( )

d(k x)
2

(10.51)

is an average over a wavelength. Given that hcos2 ( t k x)i = hsin2 ( t k x)i = 1/2, it follows from (10.47) and
(10.48) that, to second order in a,
Z 0
cosh[2 k (z + d)]
1
2
1
K = a2 2
.
(10.52)
dz = g a2
2
4
4
g k tanh(k d)
sinh (k d)
d
Making use of the general dispersion relation (10.21), we obtain
K=

1
g a2 .
4

(10.53)

196

FLUID MECHANICS

The mean potential energy perturbation per unit surface area associated with a gravity wave is defined
Z
1
g z dzi + g d 2 ,
U=h
2
d

(10.54)

which yields
1
1
1
U = h g ( 2 d 2 )i + g d 2 = g h 2 i,
2
2
2

(10.55)

or

1
g a 2.
(10.56)
4
In other words, the mean potential energy per unit surface area of a gravity wave is equal to its mean kinetic energy
per unit surface area.
Finally, the mean total energy per unit surface area associated with a gravity wave is
U=

E = K+U =

1
g a 2.
2

(10.57)

Note that this energy depends on the wave amplitude at the surface, but is independent of the wavelength, or the water
depth.

10.6 Wave Drag on Ships


Under certain circumstances (see the following section), a ship traveling over a body of water leaves behind it a train of
gravity waves whose wavefronts are transverse to the ships direction of motion. Since these waves possess energy that
is carried away from the ship, and eventually dissipated, this energy must have been produced at the ships expense.
The ship consequently experiences a drag force, D. Suppose that the ship is moving at the constant velocity V. We
would expect the transverse waves making up the train to have a matching phase velocity, so that they maintain a
constant phase relation with respect to the ship. To be more exact, we would generally expect the ships bow to always
correspond to a wave maximum (because of the pile up of water in front of the bow produced by the ships forward
motion). The condition v p = V, combined with expression (10.43), yields
tanh(k d) V 2
=
.
kd
gd

(10.58)

Suppose, for the sake of argument, that the wave train is of uniform transverse width w. Consider a fixed line drawn
downstream of the ship at right-angles to its path. The rate at which the length of the train is increasing ahead of this
line is V. Therefore, the rate at which the energy of the train is increasing ahead of the line is (1/2) g a 2 w V, where
a is the typical amplitude of the transverse waves in the train. As is well-known, wave energy travels at the group
velocity, rather than the phase velocity. Thus, the energy flux per unit width of a propagating gravity wave is simply
E vg . Wave energy consequently crosses our fixed line in the direction of the ships motion at the rate (1/2) g a 2 w vg .
Finally, the ship does work against the drag force, which goes to increase the energy of the train in the region ahead of
our line, at the rate D V. Energy conservation thus yields
1
1
g a 2 w V = g a 2 w vg + D V.
2
2

(10.59)

#
!
"
vg
1
1
2kd
2
2
,
= ga w 1
D = ga w 1
2
vp
4
sinh(2 k d)

(10.60)

However, since V = v p , we obtain

where use has been made of Equation (10.44). Here, k d is determined implicitly in terms of the ship speed via
expression (10.58). Note that this expression cannot be satisfied when the speed exceeds the critical value (g d)1/2,
since gravity waves cannot propagate at speeds in excess of this value. In this situation, no transverse wave train can

Waves in Incompressible Fluids

197

keep up with the ship, and the drag associated with such waves consequently disappears. In fact, we can see, from
the above formulae, that when V (g d)1/2 then k d 0, and so D 0. Note, however, that the transverse wave
amplitude, a, generally increases significantly as the ship speed approaches the critical value. Hence, the drag due to
transverse waves actually peaks strongly at speeds just below the critical speed, before effectively falling to zero as
this speed is exceeded. Consequently, it usually requires a great deal of propulsion power to force a ship to travel at
speeds faster than (g d)1/2.
In the deep water limit k d 1, Equation (10.60) reduces to

1
g a 2 w.
(10.61)
4
Note that (at fixed wave amplitude) this expression is independent of the wavelength of the wave train, and, hence,
independent of the ships speed. This result is actually rather misleading. In fact, (at fixed wave amplitude) the drag
acting on a ship traveling through deep water varies significantly with the ships speed. We can account for this
variation by incorporating the finite length of the ship into our analysis. A real ship moving through water generates
a bow wave from its bow, and a stern wave from its stern. Moreover, the bow wave tends to have a positive vertical
displacement, because water naturally piles up in front of the bow due to the forward motion of the ship, whereas the
stern wave tends tends to have a negative vertical displacement, because water rushes into the void left by the stern.
Very roughly speaking, suppose that the vertical displacement of the water surface caused by the ship is of the form
 x
(x) cos .
(10.62)
l
Here, l is the length of the ship. Moreover, the bow lies (instantaneously) at x = 0 [hence, (0) > 0], and the stern
at x = l [hence, (l) < 0]. For the sake of simplicity, the upward water displacement due to the bow is assumed
to equal the downward displacement due to the stern. At fixed bow wave displacement, the amplitude of transverse
gravity waves of wave number k = g/V 2 (chosen so that the phase velocity of the waves matches the ships speed, V)
produced by the ship is
Z
 x
sin( k l) k l
1 l
:
(10.63)
cos cos(k x) dx =
a
l 0
l
kl +kl
i.e., the amplitude is proportional to the Fourier coefficient of the ships vertical displacement pattern evaluated for a
wave number that matches that of the wave train. Hence, (at fixed bow wave displacement) the drag produced by the
transverse waves is
"
#2
sin( F 2 )
1
2
Da
,
(10.64)
F 2 1 + F 2
where the dimensionless parameter
V
(10.65)
F=
(g l)1/2
is known as the Froude number. See Section 2.16.
Figure 10.2 illustrates the variation of the wave drag with Froude number predicted by Equation (10.64). As we
can see, if the Froude number is much less than unity, which implies that the wavelength of the wave train is much
smaller than the length of the ship, then the drag is comparatively small. This is the case because the ship is extremely
inefficient at driving short wavelength gravity waves. It can also be seen
that the drag increases as the Froude number
increases, reaching a relatively sharp maximum when Fr = Frc = 1/ = 0.56, and then falls rapidly. Now, Fr = Frc
corresponds to the case in which the length of the ship is equal to half the wavelength of the wave train. In this
situation, the bow and stern waves interfere constructively, leading to a particularly large amplitude wave train, and,
hence, to a particularly large wave drag. The smaller peaks visible in the figure correspond to other situations in which
the bow and stern waves interfere constructively. (For instance, when the length of the ship corresponds to one and a
half wavelengths of the wave train.) A heavy ship with large a displacement, and limited propulsion power, generally
cannot overcome the peak in the wave drag that occurs when Fr = Frc . Such a ship is, therefore, limited to Froude
numbers in the range 0 < Fr < Frc , which implies a maximum speed of
D=

Vc = 0.56 (g l)1/2 = 1.75 [l(m)]1/2 m/s = 3.4 [l(m)]1/2 kts.

(10.66)

This characteristic speed is sometimes called the hull speed. Note that the hull speed increases with the length of the
ship: i.e., long ships have higher hull speeds than short ones.

198

FLUID MECHANICS

0.3

D(a.u.)

0.2

0.1

0.2

0.4

0.6
Fr

0.8

1.2

Figure 10.2: Variation of wave drag with Froude number for a ship traveling through deep water.

10.7 Ship Wakes


Let us now make a detailed investigation of the wake pattern generated behind a ship as it travels over a body of water,
taking into account obliquely propagating gravity waves, in addition to transverse waves. For the sake of simplicity,
the finite length of the ship is neglected in the following analysis. In other words, the ship is treated as a point source
of gravity waves. Consider Figure 10.3. This shows a plane gravity wave generated on the surface of the water by a
moving ship. The water surface corresponds to the x-y plane. The ship is traveling along the x-axis, in the negative
x-direction, at the constant speed V. Suppose that the ships bow is initially at point A , and has moved to point A after
a time interval t. Now, the only type of gravity wave that is continuously excited by the passage of the ship is one that
maintains a constant phase relation with respect to its bow. In fact, as we have already mentioned, the bow should
always correspond to a wave maximum. An oblique wavefront associated with such a wave is shown in the figure.
Here, the wavefront C D , which initially passes through the bow at point A , has moved to CD after a time interval t,
such that it again passes through the bow at point A. Of course, the wavefront propagates at the phase velocity, v p . It
follows that, in the right-angled triangle AA E, the sides AA and A E are of lengths V t and v p t, respectively, so that
sin =

vp
.
V

(10.67)

This, therefore, is the condition that must be satisfied in order for an obliquely propagating gravity wave to maintain a
constant phase relation with respect to the ship.
In shallow water, all gravity waves propagate at the same phase velocity: i.e.,
v p = (g d)1/2,
where d is the water depth. Hence, Equation (10.67) yields
"
#
1/2
1 (g d)
= sin
.
V

(10.68)

(10.69)

Note that this equation can only be satisfied when


V > (g d)1/2 .

(10.70)

Waves in Incompressible Fluids

199

D
E
vp t

A
Vt

Figure 10.3: An oblique plane wave generated on the surface of the water by a moving ship.

D
B
B

C
C
E
Figure 10.4: A shallow water wake.

In other words, the ship must be traveling faster than the critical speed (g d)1/2 . Moreover, if this is the case then
there is only one value of that satisfies Equation (10.69). This implies the scenario illustrated in Figure 10.4. Here,
the ship is instantaneously at A, and the wave maxima that it previously generatedwhich all propagate obliquely,
subtending a fixed angle with the x-axishave interfered constructively to produce a single strong wave maximum
DAE. In fact, the wave maxima generated when the ship was at A have travelled to B and C , the wave maxima
generated when the ship was at A have travelled to B and C , etc. We conclude that a ship traveling over shallow
water produces a V-shaped wake whose semi-angle, , is determined by the ships speed. Indeed, as is apparent from
Equation (10.69), the faster the ship travels over the water, the smaller the angle becomes. Shallow water wakes
are especially dangerous to other vessels, and particularly destructive of the coastline, because all of the wave energy
produced by the ship is concentrated into a single large wave maximum. Note, finally, that the wake contains no
transverse waves, since, as we have already mentioned, such waves cannot keep up with a ship traveling faster than
the critical speed ( g)1/2 .
Let us now discuss the wake generated by a ship traveling over deep water. In this case, the phase velocity of

200

FLUID MECHANICS

D C

P
vg t

B
Vt

Figure 10.5: Formation of an interference maximum in a deep water wake.


gravity waves is v p = (g/k)1/2 . Thus, Equation (10.67) yields
v p  g 1/2
=
sin =
.
V
kV2

(10.71)

It follows that in deep water any obliquely propagating gravity wave whose wave number exceeds the critical value
k0 =

g
V2

(10.72)

can keep up with the ship, as long as its direction of propagation is such that Equation (10.71) is satisfied. In other
words, the ship continuously excites gravity waves with a wide range of different wave numbers and propagation
directions. The wake is essentially the interference pattern generated by these waves. Now, as is well-known, an interference maximum generated by the superposition of plane waves with a range of different wave numbers propagates
at the group velocity, vg . Furthermore, as we have already seen, the group velocity of deep water gravity waves is half
their phase velocity: i.e., vg = v p /2.
Consider Figure 10.5. The curve APD corresponds to a particular interference maximum in the wake. Here, A
is the ships instantaneous position. Consider a point P on this curve. Let x and y be the coordinates of this point,
relative to the ship. Now, the interference maximum at P is part of the plane wavefront BC emitted some time t earlier,
when the ship was at point A . Let be the angle subtended between this wavefront and the x-axis. Since interference
maxima propagate at the group velocity, the distance A P is equal to vg t. Of course, the distance AA is equal to V t.
Simple trigonometry reveals that
x

V t vg t sin ,

(10.73)

vg t cos .

(10.74)

Moreover,
dy
= tan ,
(10.75)
dx
since BC is the tangent to the curve APDi.e., the curve y(x)at point P. It follows from Equation (10.71), and the
fact that vg = v p /2, that
!
1
2
(10.76)
x = X 1 sin ,
2
y

1
X sin cos ,
2

(10.77)

Waves in Incompressible Fluids

201

0.3
0.2

y/X0

0.1
0

0.1
B

0.2
0.3
0

0.1

0.2

0.3 0.4
x/X0

0.5

0.6

Figure 10.6: Locus of an interference maximum in a deep water wake.


where X() = V t. The previous three equations can be combined to produce
sin
dy dy/d (1/2) dX/d sin cos + (1/2) X (cos2 sin2 )
=
=
=
,
2
dx dx/d
cos
dX/d [1 (1/2) sin ] X sin cos

(10.78)

X
dX
=
.
d
tan

(10.79)

X = X0 sin ,

(10.80)

which reduces to

This expression can be solved to give

where X0 is a constant. Hence, the locus of our interference maximum is determined parametrically by
!
1
sin2 ,
2

X0 sin 1

1
X0 sin2 cos .
2

(10.81)
(10.82)

Here, the angle ranges from /2 to +/2. The curve specified by the above equations is plotted in Figure 10.6. As
usual, A is the instantaneous position of the ship. It can be seen that the interference maximum essentially consists
of the transverse maximum BCD, and the two radial maxima AB and AD. As is easily demonstrated, point C, which
corresponds
to = 0, lies at x = X0 /2, y = 0. Moreover, the two cusps, B and D, which correspond to =
tan1 (1/ 8) = 19.47, lie at x = (8/27)1/2 X0 , y = (1/27)1/2 X0 .
The complete interference pattern that constitutes the wake is constructed out of many different wave maximum
curves of the form shown in Figure 10.6, corresponding to many different values of the parameter X0 . However, these
X0 values must be chosen such that the wavelength of the pattern along the x-axis corresponds to the wavelength
0 = 2/k0 = 2 V 2 /g of transverse (i.e., = 0) gravity waves whose phase velocity matches the speed of the
ship. This implies that X0 = 2 j 0 , where j is a positive integer. A complete deep water wake pattern is shown in
Figure 10.7. Note that this pattern, which is made up of interlocking transverse and radialwave maxima, fills a wedgeshaped regionknown as a Kelvin wedgewhose semi-angle takes the value tan1 (1/ 8) = 19.47. This angle is

FLUID MECHANICS

y/0

202

6
5
4
3
2
1
0
1
2
3
4
5
6
1 0 1 2 3 4 5 6 7 8 9 10 11
x/0
Figure 10.7: A deep water wake.

independent of the ships speed. Finally, our initial assumption that the gravity waves that form the wake are all deep
water waves is valid provided k0 d 1, which implies that
V (g d)1/2.

(10.83)

In other words, the ship must travel at a speed that is much less than the critical speed (g d)1/2. This explains why the
wake contains transverse wave maxima.

10.8 Gravity Waves in a Flowing Fluid


Consider a gravity wave traveling through a fluid that is flowing horizontally at the uniform velocity V = V e x . Let us
write
v(r, t) = V + v1 (r, t),
p(r, t) =

(10.84)

p0 g z + p1 (r, t),

(10.85)

where v1 and p1 are the small velocity and pressure perturbations, respectively, due to the wave. To first order in small
quantities, the fluid equations of motion, (10.1) and (10.2), reduce to
v1
!

+ V v1
t

= 0,
=

(10.86)
p1
,

(10.87)

respectively. We can also define the displacement, (r, t), of a fluid particle due to the passage of the wave, as seen in
a frame co-moving with the fluid, as
!

+ V = v1 .
(10.88)
t
The curl of Equation (10.87) implies that v1 = 0. Hence, we can write v1 = , and (10.87) yields
!
p1

+V = .
t

(10.89)

Waves in Incompressible Fluids

203

z = d
V

z=0
z

z = d

x
Figure 10.8: Gravity waves at an interface between two immiscible fluids.
Finally, Equation (10.86) gives
2 = 0.

(10.90)

The most general traveling wave solution to (10.90), with wavevector k = k e x , and angular frequency , is
(x, z, t) = [A cosh(k z) + B sinh(k z)] cos( t k x).

(10.91)

It follows from Equation (10.89) that


p1 (x, z, t) = k (V c) [A cosh(k z) + B sinh(k z)] sin( t k x),

(10.92)

and from Equation (10.88) that


z (x, z, t, ) = (V c)1 [A sinh(k z) + B cosh(k z)] sin( t k x).

(10.93)

Here, c = /k is the phase velocity of the wave.

10.9 Gravity Waves at an Interface


Consider a layer of fluid of density , depth d , and uniform horizontal velocity V , situated on top of a layer of
another fluid of density , depth d, and uniform horizontal velocity V. Suppose that the fluids are bounded from above
and below by rigid horizontal planes. Let these planes lie at z = d and z = d , and let the unperturbed interface
between the two fluids lie at z = 0. See Figure 10.8.
Consider a gravity wave of angular frequency , and wavenumber k, propagating through both fluids in the xdirection. Let
(x, t) = 0 sin( t k x)
(10.94)
be the small vertical displacement of the interface due to the wave. In the lower fluid, the perturbed velocity potential
must be of the form (10.91), with the constants A and B chosen such that vz |z=d = 0 and z (x, 0, t) = (x, t). It follows
that A = (V c) 0 / tanh(k d) and B = (V c) 0 , so that
(x, z, t) = (V c) 0

cosh[k (z + d)]
cos( t k x).
sinh(k d)

(10.95)

In the upper fluid, the perturbed velocity potential must again be of the form (10.91), with the constants A and B
chosen such that vz |z=d = 0 and z (x, 0, t) = (x, t). It follows that A = (V c) 0 / tanh(k d ) and B = (V c) 0 , so
that
cosh[k (z d )]
cos( t k x).
(10.96)
(x, z, t) = (V c) 0
sinh(k d )

204

FLUID MECHANICS

Here, c = /k is the phase velocity of the wave. From Equations (10.85) and (10.92), the fluid pressure just below the
interface is
p0 g + p1 (x, 0 , t)

p(x, 0 , t) =

p0 g 0 sin( t k x) k (V c) A sin( t k x)
#
"
k (V c)2
0 sin( t k x).
p0 g
tanh(k d)

=
=

(10.97)

Likewise, the fluid pressure just above the interface is


p(x, 0+ , t)

=
=
=

p0 g + p1 (x, 0+ , t)

p0 g 0 sin( t k x) k (V c) A sin( t k x)
"
#
k (V c)2
p0 g +
0 sin( t k x).
tanh(k d )

Now, in the absence of surface tension at the interface, these two pressure must equal one another: i.e.,
 z=0+
p z=0 = 0.

(10.98)

(10.99)

Hence, we obtain the dispersion relation

( ) g =

k (V c)2 k (V c)2
+
,
tanh(k d)
tanh(k d )

(10.100)

which takes the form of a quadratic equation for the phase velocity, c, of the wave. We can see that:
i. If = 0 and V = 0 then the dispersion relation reduces to (10.43) (with v p = c).
ii. If the two fluids are of infinite depth then the dispersion relation simplifies to
( ) g = k (V c)2 + k (V c)2 .

(10.101)

iii. In general, there are two values of c that satisfy the quadratic equation (10.100). These are either both real, or
form a complex conjugate pair.
iv. The condition for stability is that c is real. The alternative is that c is complex, which implies that is also
complex, and, hence, that the perturbation grows or decays exponentially in time. Since the complex roots of
a quadratic equation occur in complex conjugate pairs, one of the roots always corresponds to an exponentially
growing mode: i.e., an instability.
v. If both fluids are at rest (i.e., V = V = 0), and of infinite depth, then the dispersion reduces to
c2 =

g ( )
.
k ( + )

(10.102)

It follows that the configuration is only stable when > : i.e., when the heavier fluid is underneath.
As a particular example, suppose that the lower fluid is water, and the upper fluid is the atmosphere. Let s = / =
1.225 103 be the specific density of air at s.t.p. (relative to water). Putting V = V = 0, d , and making use of
the fact that s is small, the dispersion relation (10.100) yields
#1/2 (
"
)
1
1/2 tanh(k d)
1 s [1 + tanh(k d)] .
c (g d)
(10.103)
kd
2
Comparing this with (10.43), we can see that the presence of the atmosphere tends to slightly diminish the phase
velocities of gravity waves propagating over the surface of a body of water.

Waves in Incompressible Fluids

205

10.10 Steady Flow over a Corrugated Bottom


Consider a stream of water of mean depth d, and uniform horizontal velocity V = V e x , that flows over a corrugated
bottom whose elevation is z = d + a sin(k x), where a is much smaller than d. Let the elevation of the free surface of
the water be z = b sin(k x). We wish to determine the relationship between a and b.
Now, we expect the velocity potential, perturbed pressure, and vertical displacement of the water to be of the form
(10.91), (10.92), and (10.93), respectively, with = c = 0, since we are looking for a stationary (i.e., non-propagating)
perturbation driven by the static corrugations in the bottom. The boundary condition at the bottom is
z (x, d) = a sin(k x),

(10.104)

V 1 [A sinh(k d) + B cosh(k d)] = a.

(10.105)

z (x, 0) = b sin(k x),

(10.106)

b = V 1 B.

(10.107)

which yields
At the free surface, we have
which gives
In addition, pressure balance across the free surface yields
g b sin(k x) = p1 (x, 0) = k V A sin(k x),

(10.108)

g b = k V A.

(10.109)

which leads to
Hence, from (10.105), (10.107), and (10.109),
b=
or

a
,
cosh(k d) (g/k V 2 ) sinh(k d)

(10.110)

a
,
cosh(k d) (1 c2 /V 2 )

(10.111)

b=

where c = [(g/k) tanh(k d)]1/2 is the phase velocity of a gravity wave of wave number k. See Equation (10.21). It
follows that the peaks and troughs of the free surface coincide with those of the bottom when |V| > |c|, and the troughs
coincide with the peaks, and vice versa, when |V| < |c|. If |V| = |c| then the ratio b/a becomes infinite, implying that
the oscillations driven by the corrugations are not of small amplitude, and, therefore, cannot be described by linear
theory.

10.11 Surface Tension


As described in Chapter 4, there is a positive excess energy per unit area, , associated with an interface between two
immiscible fluids. The quantity can also be interpreted as a surface tension. Let us now incorporate surface tension
into our analysis. Suppose that the interface lies at
z = (x, t),

(10.112)

where || is small. Thus, the unperturbed interface corresponds to the plane z = 0. The unit normal to the interface is
n=

(z )
.
|(z )|

(10.113)

nx

,
x

(10.114)

nz

1.

It follows that

(10.115)

206

FLUID MECHANICS

Now, the Young-Laplace Equation yields


p = n,

(10.116)

where p is the jump in pressure seen crossing the interface in the opposite direction to n. See Section 4.2. However,
from (10.114) and (10.115), we have
2
(10.117)
n 2.
x
Hence, Equation (10.116) gives
2
z=0+
[p]z=0
=

.
(10.118)

x2
This expression is the generalization of (10.99) that takes surface tension into account.
Suppose that the interface in question is that between a body of water, of density and depth d, and the atmosphere.
Let the unperturbed water lie between z = d and z = 0, and let the unperturbed atmosphere occupy the region z > 0.
In the limit in which the density of the atmosphere is neglected, the pressure in the atmosphere takes the fixed value
p0 , whereas the pressure just below the surface of the water is p0 g + p1 |z=0 . Here, p1 is the pressure perturbation
due to the wave. The relation (10.118) yields
g p1 |z=0 =

2
,
x2

(10.119)

where is the surface tension at an air/water interface. However, /t = (/z)z=0, where is the perturbed velocity
potential of the water. Moreover, from (10.9), p1 = (/t). Hence, the above expression gives




2
3
+
g
(10.120)
.
=
z z=0 t2 z=0 z 2 x z=0

This relation, which is a generalization of Equation (10.15), is the condition satisfied at a free surface in the presence
of non-negligible surface tension. Applying this boundary condition to the general solution, (10.19) (which already
satisfies the boundary condition at the bottom), we obtain the dispersion relation
!
k3
tanh(k d),
(10.121)
2 = g k +

which is a generalization of (10.21) that takes surface tension into account.

10.12 Capillary Waves


In the deep water limit k d 1, the dispersion relation (10.121) simplifies to
2 = g k +
It is helpful to introduce the capillary length,

l=
g

k3
.

!1/2

(10.122)

(10.123)

(See Section 4.4.) The capillary length of an air/water interface at s.t.p. is 2.7 103 m. The associated capillary
wavelength is c = 2 l = 1.7 102 m. Roughly speaking, surface tension is negligible for waves whose wavelengths
are much larger than the capillary wavelength, and vice versa. It is also helpful to introduce the critical phase velocity
vc = (2 g l)1/2.

(10.124)

This critical velocity takes the value 0.23 m/s for an air/water interface at s.t.p. It follows from (10.122) that the phase
velocity, v p = /k, of a surface water wave can be written
!#1/2
"
vp
1
1
kl+
.
(10.125)
=
vc
2
kl

Waves in Incompressible Fluids

207

Moreover, the ratio of the phase velocity to the group velocity, vg = d/dk, becomes
#
"
vg
1 1 + 3 (k l)2
.
=
v p 2 1 + (k l)2

(10.126)

In the long wavelength limit c (i.e., k l 1), we obtain

vp
1
,

v0
(2 k l)1/2

(10.127)

and

vg 1
.
vp 2
We can identify this type of wave as the deep water gravity wave discussed in Section 10.3.
In the short wavelength limit c (i.e., k l 1), we get
!1/2
vp
kl

,
vc
2

(10.128)

(10.129)

and

vg 3
.
(10.130)
vp 2
This corresponds to a completely new type of wave known as a capillary wave. Such waves have wavelengths that are
much less than the capillary wavelength. Moreover, (10.129) can be rewritten
!1/2
k
,
(10.131)
vp

which demonstrates that gravity plays no role in the propagation of a capillary waveits place is taken by surface
tension. Finally, it is easily seen that the phase velocity (10.125) attains the minimum value v p = vc when = c (i.e.,
when k l = 1). Moreover, from (10.126), vg = v p at this wavelength. It follows that the phase velocity of a surface
wave propagating over a body of water can never be less than the critical value, vc .

10.13 Capillary Waves at an Interface


Consider a layer of fluid of density , depth d , and uniform horizontal velocity V , situated on top of a layer of another
fluid of density , depth d, and uniform horizontal velocity V. Suppose that the fluids are bounded from above and
below by rigid horizontal planes. Let these planes be at z = d and z = d , and let the unperturbed interface between
the two fluids be at z = 0. Suppose that the elevation of the perturbed interface is z = , where = 0 sin( t k x).
Finally, let be the surface tension of the interface. Equations (10.97), (10.98), and (10.118) yield the dispersion
relation
k (V c)2 k (V c)2
+
,
(10.132)
( ) g + k2 =
tanh(k d)
tanh(k d )
which is a generalization of the dispersion relation (10.100) that takes surface tension into account. Here, c = /k is
the phase velocity of a wave propagating along the interface.
For the case in which both fluids are at rest, and of infinite depth, the above dispersion relation simplifies to give
( ) g + k2 = ( + ) k c2 .

(10.133)

Suppose that s = / is the specific gravity of the upper fluid with respect to the lower. In the case in which s < 1
(i.e., the upper fluid is lighter than the lower one), it is helpful to define
"
#1/2

l =
(10.134)
g (1 s)
!#1/2
"
1s
.
(10.135)
c0 = 2 g l
1+s

208

FLUID MECHANICS

It follows that

!
1 1
c2
+
k
l
.
=
c02 2 k l

(10.136)

Thus, we conclude that the phase velocity of a wave propagating along the interface between the two fluids achieves its
minimum value, c = c0 , when k l = 1. Furthermore, waves of all wavelength are able to propagate along the interface
(i.e., c2 > 0 for all k). In the opposite case, in which s > 1 (i.e., the upper fluid is heavier than the lower one), we can
redefine the capillary length as
"
#1/2

l=
.
(10.137)
g (s 1)
The dispersion relation (10.133) then becomes

c2 = g l

s1
s+1

kl

!
1
.
kl

(10.138)

It is apparent that c2 < 0 for k l < 1, indicating instability of the interface for waves whose wavelengths exceed the
critical value c = 2 l. On the other hand, waves whose wavelengths are less than the critical value are stabilized by
surface tension. This result is exemplified by the experiment in which water is retained by atmospheric pressure in an
inverted glass whose mouth is closed by a gauze of fine mesh (the purpose of which is to put an upper limit on the
wavelengths of waves that can exist at the interface.)

10.14 Wind Driven Waves in Deep Water


Consider the scenario described in the previous section. Suppose that the lower fluid is a body of deep water at rest,
and the upper fluid is the atmosphere. Let the air above the surface of the water move horizontally at the constant
velocity V . Suppose that is the density of water, s = / the specific gravity of air with respect to water, and the
surface tension at an air/water interface. With V = 0, k d , k d , the dispersion relation (10.132) reduces to
(1 s) g + k2 = k c2 + s k (V c)2 .

(10.139)

This expression can be rearranged to give


!
V 2 s
2 V s
+
= c12 ,
c
1+ s
1+s
2

which is a quadratic equation for the phase velocity, c, of the wave. Here,
!
k
g 1s
2
+
c1 =
,
k 1+s
(1 + s)
where c1 is the phase velocity that the wave would have in the absence of the wind. In fact, we can write
!
c 2
1
2
c
+
c1 =
2 c
0

(10.140)

(10.141)

(10.142)

where c = 2 l is the capillary wavelength, and l and c0 are defined in Equations (10.134) and (10.135), respectively.
For a given wavelength, , the wave velocity, c, attains its maximum value, cm , when dc/dV = 0. According to
the dispersion relation (10.140), this occurs when
V = cm = (1 + s)1/2 c1 .
If the wind has any other velocity, greater or less than cm , then the wave velocity is less than cm .
According to (10.140), the wave velocity, c, becomes complex, indicating an instability, when
!
(1 + s)2 2 (1 + s)2 c 2
V 2 >
c .
+
c1 =
s
2s
c
0

(10.143)

(10.144)

Waves in Incompressible Fluids

209

We conclude that if the wind speed exceeds the critical value


Vc =

(1 + s)
c0 = 6.6 m/s = 12.8 kts
s1/2

(10.145)

then waves whose wavelengths fall within a certain range, centered around c , are unstable and grow to large amplitude.
The two roots of Equation (10.140) are
"
#1/2
V s
s V 2
2
c=
c1
.
1+s
(1 + s)2

(10.146)

V < (1 + s1 )1/2 c1

(10.147)

Moreover, if
then these roots have opposite signs. Hence, the waves can either travel with the wind, or against it, but travel faster
when they are moving with the wind. If V exceeds the value given above then the waves cannot travel against the
wind. Since c1 has the minimum value c0 , it follows that waves traveling against the wind are completely ruled out
when
V > (1 + s1 )1/2 c0 = 6.6 m/s = 12.8 kts.
(10.148)

10.15 Exercises
10.1. Find the velocity potential of a standing gravity wave in deep water for which the associated elevation of the free surface is
z = a cos( t) cos(k x).
Determine the paths of water particles perturbed by the wave.
10.2. Deep water fills a rectangular tank of length l and breadth b. Show that the resonant frequencies of the water in the tank are
(g )1/2 (n2 l 2 + m2 b 2 )1/4 ,
where n and m are integers. You may neglect surface tension.
10.3. Demonstrate that a sinusoidal gravity wave on deep water with surface elevation
= a cos( t k x)
possesses a mean momentum per unit surface area
1
a 2.
2
10.4. A seismic wave passes along the bed of an ocean of uniform depth d such that the vertical perturbation of the bed is
a cos[k (x V t)]. Show that the amplitude of the consequent gravity waves at the surface is
!
#1
"
c2
a 1 2 cosh(k d) ,
V

where c is the phase velocity of waves of wavenumber k.


10.5. A layer of liquid of density and depth d has a free upper surface, and lies over liquid of infinite depth and density > .
Neglecting surface tension, show that two possible types of wave of wavenumber k, with phase velocities

can propagate along the layer.

c2

k g,

c2

k g ( )
,
coth(k d) +

210

FLUID MECHANICS

10.6. Show that, taking surface tension into account, a sinusoidal wave of wavenumber k and surface amplitude a has a mean
kinetic energy per unit surface area
1
K = ( g + k2 ) a2 ,
4
and a mean potential energy per unit surface area
U=

1
( g + k2 ) a2 .
4

10.7. Show that in water of uniform depth d the phase velocity of surface waves can only attain a stationary (i.e., maximum or
minimum) value as a function of wavenumber, k, when
"
#1/2
sinh(2 k d) 2 k d
k=
kc ,
sinh(2 k d) + 2 k d
where kc = ( g/)1/2 . Hence, deduce that the phase velocity has just one stationary value (a minimum) for any depth greater
than 31/2 kc1 4.8 mm, but no stationary values for depths less than that.
10.8. Unlike gravity waves in deep water, whose group velocities are half their phase velocities, the group velocities of capillary
waves are 3/2 times their phase velocities. Adapt the analysis of Section 10.7 to investigate the generation of capillary waves
by a very small object traveling across the surface of the water at the constant speed V. Suppose that the unperturbed surface
corresponds to the x-y plane. Let the object travel in the minus x-direction, such that it is instantaneously found at the origin.
Find the present position of waves that were emitted, traveling at an angle to the objects direction of motion, when it was
located at (X, 0). Show that along a given interference maximum the quantities X and vary in such a manner that X cos3
takes a constant value, X1 (say). Deduce that the interference maximum is given parametrically by the equations
!
1
x = X1 sec tan2
,
2
y

3
X1 sec tan .
2

Sketch this curve, noting that it goes through the points (0.5 X1 , 0) and (0, 1.3 X1 ), and asymptotes to y = 1.5 X11/3 x2/3 .

Equilibrium of Compressible Fluids

211

11 Equilibrium of Compressible Fluids

11.1 Introduction
In this chapter, we investigate the equilibria of compressible fluids such as gases. As is the case for an incompressible
fluid (see Chapter 3), a compressible fluid in mechanical equilibrium must satisfy the force balance equation
0 = p + ,

(11.1)

where p is the static fluid pressure, the mass density, and the gravitational potential energy per unit mass. In an
ideal gas, the relationship between p and is determined by the energy conservation equation, (2.89), which can be
written
!
!
D p
M 2 p
.
(11.2)
=

1 Dt
R

Here, is the ratio of specific heats, the thermal conductivity, M the molar mass, and R the molar ideal gas constant.
Note that the viscous heat generation term has been omitted from the above equation (since it is zero in a stationary
gas). The limits in which the left- and right-hand sides of Equation (11.2) are dominant are termed the adiabatic and
isothermal limits, respectively. In the isothermal limit, in which thermal transport is comparatively large, so that (11.2)
can only be satisfied when 2 (p/) 0, the temperature (recall that T p/ in an ideal gas) distribution in the gas
becomes uniform, and the pressure and density are consequently related according to the isothermal gas law,
p
= constant.

(11.3)

On the other hand, in the adiabatic limit, in which thermal transport is negligible, so that (11.2) can only be satisfied
when D/Dt(p/ ) 0, the pressure and density are related according to the adiabatic gas law,
p
= constant.

(11.4)

11.2 Isothermal Atmosphere


The vertical thickness of the atmosphere is only a few tens of kilometers, and is, therefore, much less than the radius
of the Earth, which is about 6000 km. Consequently, it is a good approximation to treat the atmosphere as a relatively
thin layer, covering the surface of the Earth, in which the pressure and density are only functions of altitude above
ground level, z, and the gravitational potential energy per unit mass takes the form = g z, where g is the acceleration
due to gravity at z = 0. It follows from Equation (11.1) that
dp
= g.
dz

(11.5)

Now, in an isothermal atmosphere, in which the temperature, T , is assumed not to vary with height, the ideal gas
equation of state (2.84) yields [cf., Equation (11.3)]
p RT
=
.

(11.6)

gM
dp
=
p.
dz
RT

(11.7)

p(z) = p0 exp(z/H),

(11.8)

The previous two equations can be combined to give

Hence, we obtain

212

FLUID MECHANICS

where p0 105 N m2 is atmospheric pressure at ground level, and


H=

RT
gM

(11.9)

is known as the isothermal scale height of the atmosphere. Using the values T = 273 K (0 C), M = 29, and
g = 9.8 m s2 , which are typical of the Earths atmosphere (at ground level), as well as R = 8.315 J mol1 K1 , we find
that H = 7.99 km. Equations (11.6) and (11.8) yield
(z) = 0 exp(z/H),

(11.10)

where 0 = p0 /(g H) is the atmospheric mass density at z = 0. According to Equations (11.8) and (11.10), in an
isothermal atmosphere, the pressure and density both decrease exponentially with increasing altitude, falling to 37%
of their values at ground level when z = H, and to only 5% of these values when z = 3 H.

11.3 Adiabatic Atmosphere


In fact, the temperature of the Earths atmosphere is not uniform, but instead decreases steadily with increasing altitude. This effect is largely due to the action of convection currents. When a packet of air ascends, under the influence
of such currents, the diminished pressure at higher altitudes causes it to expand. Since this expansion generally takes
place far more rapidly than heat can diffuse into the packet, the work done against the pressure of the surrounding
gas, as the packet expands, leads to a reduction in its internal energy, and, hence, in its temperature. Assuming that
the atmosphere is in a continually mixed state, whilst remaining in approximate vertical force balance (such a state is
known as a convective equilibrium), and that the effect of heat conduction is negligible (because the mixing takes place
too rapidly for thermal diffusion to affect the temperature), we would expect the adiabatic gas law, (11.4), to offer a
much more accurate description of the relationship between atmospheric pressure and density than the isothermal gas
law, (11.3).
b = T/T 0 , where p0 , 0 , and T 0 are the pressure, mass density, and temperature of the
Let b
p = p/p0 , b
= /0 , and T
atmosphere, respectively, at ground level. The adiabatic gas law, (11.4), can be combined with the ideal gas equation
of state, (11.6), to give
b /(1) .
b
p=b
= T
(11.11)
The isothermal scale height of the atmosphere is conveniently redefined as [cf., (11.9)]
H=

R T0
p0
.
=
g M g 0

(11.12)

Equations (11.5), (11.11), and (11.12) yield

where b
z = z/H, or, from (11.11),

db
p
= b
p 1/ ,
db
z

(11.13)

b
dT
1
=
.
db
z

(11.14)

The above equation can be integrated to give


T (z) = T 0 1

!
1 z
.
H

(11.15)

It follows that the temperature in an adiabatic atmosphere decreases linearly with increasing altitude at the rate of
[/( 1)] (T 0/H) degrees per meter. This rate is known as the adiabatic lapse rate of the atmosphere. Using the
values = 1.4, T 0 = 273 K, and H = 7.99 km, which are typical of the Earths atmosphere, we estimate the lapse rate
to be 9.8 K km1 . In reality, the lapse rate only takes this value in dry air. In moist air, the lapse rate is considerably
reduced because of the latent heat released when water vapor condenses.

Equilibrium of Compressible Fluids

213

Equations (11.11) and (11.15) yield


p(z) =

p0 1

1 z
H

!/(1)

(11.16)

(z)

0 1

1 z
H

!1/(1)

(11.17)

Since /( 1) 3.5 and 1/( 1) 2.5, it follows that pressure decreases more rapidly than density in an adiabatic
atmosphere. Moreover, the previous three equations imply that an adiabatic atmosphere has a sharp upper boundary
at z = [/( 1)] H 28 km. At this altitude, the temperature, pressure, and density all fall to zero. Of course, above
this altitude, the temperature, pressure, and density remain zero (since they cannot take negative or imaginary values).
In contrast, an isothermal atmosphere has a diffuse upper boundary in which the pressure and density never fall to
zero, even at extreme altitudes. It should be noted that, in reality, the Earths atmosphere does not have a sharp upper
boundary, since the adiabatic gas law does not hold at very high altitudes.

11.4 Atmospheric Stability


Suppose that the atmosphere is static (i.e., non-convecting). Moreover, let p(z) and (z) be the pressure and density,
respectively, as functions of altitude. Consider a packet of air that is in equilibrium with the surrounding air at some
initial altitude z1 , but subsequently moves to a higher altitude z2 . Thus, the packets initial pressure and density are
p1 = p(z1 ) and 1 = (z1 ), respectively. Now, at the higher altitude, the packet must adjust its volume in such a manner
that its pressure matches that of the surrounding air, otherwise there would be a force imbalance across the packet
boundary. It follows that the packet pressure at altitude z2 is p2 = p(z2 ). Assuming that the packet moves upward
on a much faster time-scale than that needed for heat to diffuse across it (but still a sufficiently slow time-scale that it
remains in approximate pressure balance with the surrounding air), we would expect its internal pressure and density
to be related according to the adiabatic gas law, (11.4). Thus, the packets density at altitude z2 is 2 = (p2 /p1 )1/ 1 .
Now, if 2 > (z2 ) then the packet is denser than the surrounding air. It follows that the packet weight exceeds the
buoyancy due to the atmosphere, causing the packet to sink back to its original altitude. On the other hand, if 2 < (z2 )
then the packet is less dense than the surrounding air. It follows that the buoyancy force exceeds the packet weight,
causing it to rise to an even higher altitude. In other words, the atmosphere is unstable to vertical convection when
[p(z2 )/p(z1 )]1/ (z1 ) < (z2 ) for any z2 > z1 : i.e., when


p
p

,
<
(11.18)
z2
z1

for any z2 > z1 . It follows that the atmosphere is only stable to vertical convection when p/ is a monotonically
decreasing function of altitude. As is easily demonstrated, this stability criterion can also be written
d ln p
< ,
d ln

(11.19)

d ln T
< 1.
d ln

(11.20)

or, making use of the ideal gas equation of state,

Convection is triggered in regions of the atmosphere where the above stability criterion is violated. However, such
convection acts to relax these regions back to a marginally stable state in which p/ is uniform: i.e., an adiabatic
equilibrium.

11.5 Eddington Solar Model


Let us investigate the internal structure of the Sun, which is basically a self-gravitating sphere of incandescent ionized
gas (consisting mostly of hydrogen). Adopting a spherical coordinate system (see Section C.4), r, , , whose origin

214

FLUID MECHANICS

coincides with the Suns geometric center, and making the simplifying (and highly accurate) assumption that the mass
distribution within the Sun is spherically symmetric, we find that
dm
= 4 r 2 ,
dr

(11.21)

where m(r) is the total mass contained within a sphere of radius r, centered on the origin, and (r) the mass density
at radius r. Now, as is well-known, the gravitational acceleration at some radius r in a spherically symmetric mass
distribution is the same as would be obtained were all the mass located within this radius concentrated at the center,
and the remainder of the mass neglected. In other words,
Gm
d
= 2 ,
dr
r

(11.22)

where (r) is the gravitational potential energy per unit mass, and d/dr the radial gravitational acceleration. The
force balance criterion (11.1) yields
d
dp
+
= 0,
(11.23)
dr
dr
where p(r) is the pressure. The previous three equations can be combined to give
!
1 d r 2 dp
= 4 G .
r 2 dr dr

(11.24)

In order to make any further progress, we need to determine the relationship between the Suns internal pressure
and density. Unfortunately, this relationship is ultimately controlled by energy transport, which is a very complicated
process in a star. In fact, a stars energy is ultimately derived from nuclear reactions occurring deep within its core,
the details of which are extremely involved. This energy is then transported from the core to the outer boundary via a
combination of convection and radiation. (Conduction plays a much less important role in this process.) Unfortunately,
an exact calculation of radiative transport requires an understanding of the opacity of stellar material, which is an
exceptionally difficult subject. Finally, once the energy reaches the boundary of the star it is radiated away. The
following ingenious model, due to Eddington,1 is appropriate to a star whose internal energy transport is dominated
by radiation. This turns out to be a fairly good approximation for the Sun. The main advantage of Eddingtons model
is that it does not require us to know anything about stellar nuclear reactions or opacity.
Now, the temperature inside the Sun is sufficiently large that radiation pressure cannot be completely neglected
with respect to conventional gas pressure. In other words, we must write the solar equation of state in the form
p = pg + pr ,
where
pg =

kT
mp

(11.25)

(11.26)

is the gas pressure (modeling the plasma within the Sun as an ideal gas of free electrons and ions), and
pr =

1
T 4
3

(11.27)

the radiation pressure (assuming that the radiation within the Sun is everywhere in local thermodynamic equilibrium
with the plasma). Here, T (r) is the Suns internal temperature, k the Boltzmann constant, m p the mass of a proton, and
the relative molecular mass (i.e., the ratio of the mean mass of the free particles making up the solar plasma to that
of a proton). Note that the electron mass has been neglected with respect to that of a proton. Furthermore, = 4/c,
where is the Stefan-Boltzmann constant, and c the velocity of light in a vacuum. Incidentally, in writing (11.26), we
have expressed M/R in the equivalent form m p /k.
1 A.S.

Eddington, The Internal Constitution of the Stars (Cambridge University Press, Cambridge UK, 1926).

Equilibrium of Compressible Fluids

215

, y 1

Figure 11.1: The functions () (solid) and y() (dashed).


Let
pg

= (1 ) p,

(11.28)

pr

= p,

(11.29)

where the parameter is assumed to be uniform. In other words, the ratio of the radiation pressure to the gas pressure
is assumed to be the same everywhere inside the Sun. This fairly drastic assumption turns outperhaps, somewhat
fortuitously 2 to lead to approximately the correct internal pressure-density relation for the Sun. In fact, Equations (11.26)(11.29) can be combined to give
p = K 4/3 ,
(11.30)
where

1/3
!
k 4 3

.
K =
m p 4 (1 )4

(11.31)

It can be seen, by comparison with Equation (11.4), that the above pressure-density relation takes the form of an
adiabatic gas law with an effective ratio of specific heats = 4/3. Note, however, that the actual ratio of specific heats
for a fully ionized hydrogen plasma, in the absence of radiation, is = 5/3. Hence, the 4/3 exponent, appearing in
(11.30), is entirely due to the non-negligible radiation pressure within the Sun.
Let T c = T (0), c = (0), and pc = p(0), be the Suns central temperature, density, and pressure, respectively. It
follows from (11.30) that
pc = K 4/3
(11.32)
c ,
and from (11.26) and (11.28) that
Tc =
2 L.

Mestel, Phys. Reports 311, 295 (1999).

pc m p
(1 ) .
c k

(11.33)

216

FLUID MECHANICS

log10 (T [K])

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


m/M
Figure 11.2: Solar temperature versus mass fraction obtained from the Eddington Solar Model (solid) and the Standard
Solar Model (dashed).
Suppose that
T
= ,
Tc

(11.34)

where is a dimensionless function. According to Equations (11.26), (11.28), and (11.30),

c
p
pc

3,

(11.35)

4.

(11.36)

Moreover, it is clear, from the above expressions, that = 1 at the center of the Sun, r = 0, and = 0 at the edge,
r = R (say), where the temperature, density, and pressure are all assumed to fall to zero. Suppose, finally, that
r = a ,

(11.37)

where is a dimensionless radial coordinate, and


a=

K
G c2/3

!1/2

(11.38)

Thus, the center of the Sun corresponds to = 0, and the edge to = 1 (say), where (1 ) = 0, and
R = 1 a.

(11.39)

Equations (11.35)(11.38) can be used to transform the equilibrium relation (11.24) into the non-dimensional form
!
1 d 2 d

= 3 .
(11.40)
d
2 d

Equilibrium of Compressible Fluids

217

log10 ( [kg m3])

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


m/M

Figure 11.3: Solar mass density versus mass fraction obtained from the Eddington Solar Model (solid) and the Standard Solar Model (dashed).
Moreover, Equation (11.21) can be integrated, with the aid of Equations (11.35), (11.37), and (11.40), and the physical
boundary condition m(0) = 0, to give
m = 4 c a3 y,
(11.41)
where
y() = 2

d
.
d

(11.42)

Equation (11.40) is known as the Lane-Emden equation (of degree 3), and can, unfortunately, only be solved numerically. The appropriate solution takes the form = 1 2 /6 + O(4 ) when 1, and must be integrated to = 1 ,
where (1 ) = 0. Figure 11.1 shows (), and the related function y(), obtained via numerical methods. Note that
1 = 6.897, and y1 = y(1 ) = 2.018.
According to Equation (11.41), the solar mass, M = m(R), can be written
M = 4 c a3 y1 ,

(11.43)

which reduces, with the aid of Equations (11.31) and (11.38), to

= 2,
(1 )4

(11.44)

where
= 2
and

M0 =

k
1
3
m
( G)
p

!4

M
,
M0

1/2
3
4 y1 = 3.586 1031 kg.

(11.45)

(11.46)

218

FLUID MECHANICS

17

log10 (p [N m2])

16
15
14
13
12
11
10

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


m/M

Figure 11.4: Solar pressure versus mass fraction obtained from the Eddington Solar Model (solid) and the Standard
Solar Model (dashed).
Moreover, it is easily demonstrated that
K=

G M 2/3
.
(4 y1)2/3

(11.47)

According to Equations (11.44) and (11.45), the ratio, /(1 ), of the radiation pressure to the gas pressure in a
radiative star is a strongly increasing function of the stellar mass, M, and mean molecular weight, . In the case of the
Sun, for which 1, Equation (11.44) can be inverted to give the approximate solution
= 2

1 4
+ O( 6 ).
4

(11.48)

Using the observed solar mass M = 1.989 1030 kg, and the value = 0.68 (which represents the best fit to the
Standard Solar Model mentioned below), we find that = 6.58 104 . In other words, the radiation pressure inside
the Sun is only a very small fraction of the gas pressure. This immediately implies that radiative energy transport is
much less efficient than convective energy transport. Indeed, in regions of the Sun in which convection occurs we
would expect the convective transport to overwhelm the radiative transport, and so to drive the local pressure-density
relation toward an adiabatic law with an exponent 5/3. Fortunately, convection only takes place in the Suns outer
regions, which contain a minuscule fraction of its mass.
Equations (11.31), (11.32), (11.39), (11.43), and (11.47) yield
Tc

G M mp
1
(1 )
= 1.34 107 K,
4 y1
Rk

(11.49)

13 M
= 7.63 104 kg m3,
4 y1 R 3

(11.50)

pc

14

G M2
= 1.24 1016 N m2 ,
16 y12 R4

(11.51)

Equilibrium of Compressible Fluids

219

log10[/(1 )]

2.5

3.5

0.2

0.4
0.6
m/M

0.8

Figure 11.5: Ratio of radiation pressure to gas pressure calculated from the Eddington Solar Model (solid) and the
Standard Solar Model (dashed).
where the solar radius R has been given the observed value 6.960 108 m. The actual values of the Suns central
temperature, density, and pressure, as determined by the so-called Standard Solar Model (SSM),3 which incorporates
detailed treatments of nuclear reactions and opacity, are T c = 1.58 107 K, c = 15.6 104 kg m3 , and pc =
2.38 1016 N m2 , respectively. It can be seen that the values of T c , c , and pc obtained from the Eddington model
lie within a factor of two of those obtained from the much more accurate SSM. Figures 11.2, 11.3, and 11.4 show
the temperature, density, and pressure profiles, respectively, obtained from the SSM 4 and the Eddington model. The
profiles are plotted as functions of the mass fraction, m(r)/M = y()/y1 , where = r/a. It can be seen that there
is fairly good agreement between the profiles calculated by the two models. Finally, Figure 11.5 compares the ratio,
/(1 ), of the radiation pressure to the gas pressure obtained from the SSM and the Eddington model. Recall, that
it is a fundamental assumption of the Eddington model that this pressure ratio is uniform throughout the Sun. In fact,
it can be seen that the pressure ratio calculated by the SSM is not spatially uniform. On the other hand, the spatial
variation of the ratio is fairly weak, except close to the edge of the Sun, where convection sets in, and the Eddington
model, thus, becomes invalid. We conclude that, despite its simplicity, the Eddington solar model does a remarkably
good job of accounting for the Suns internal structure.

11.6 Exercises
11.1. Prove that the fraction of the whole mass of an isothermal atmosphere which is included between the ground and a horizontal
plane of height z is
1 ez/H .
Evaluate this for z = H, 2 H, 3 H, respectively.

11.2. If the absolute temperature in the atmosphere diminishes upwards according to the law
z
T
=1 ,
T0
c
3 http://en.wikipedia.org/wiki/Standard\_Solar\_Model
4 The

SSM data is obtained from http://www.ap.stmarys.ca/guenther/evolution/ssm1998.html

220

FLUID MECHANICS
where c is a constant, show that the pressure varies as

z c/H
p
.
= 1
p0
c

11.3. If the absolute temperature in the atmosphere diminishes upward according to the law
1
T
=
,
T0
1 + z
where is a constant, show that the pressure varies as
!
p
z
1 z2
= exp
.
p0
H 2 H
11.4. Show that if the absolute temperature, T , in the atmosphere is any given function of the altitude, z, then the vertical distribution of pressure in the atmosphere is given by
Z
p
T 0 z dz
ln
=
.
p0
H 0 T
11.5. Show that if the Earth were surrounded by an atmosphere of uniform temperature then the pressure a distance r from the
Earths center would be
" 2
!#
p
a 1 1
= exp

,
p0
H r a
where a is the Earths radius.
11.6. Show that if the whole of space were occupied by air at the uniform temperature T then the densities at the surfaces of the
various planets would be proportional to the corresponding values of
!
gMa
,
exp
RT
where a is the radius of the planet, and g its surface gravitational acceleration.
11.7. Prove that in an atmosphere arranged in horizontal strata the work (per unit mass) required to interchange two thin strata of
equal mass without disturbance of the remaining strata is
!
 p1
1  1
p2
2 11
,

1
1 2
where the suffixes refer to the initial states of the two strata. Hence, show that for stability the ratio p/ must increase
upwards.
11.8. A spherically symmetric star is such that m(r) is the mass contained within radius r. Show that the stars total gravitational
potential energy can be written in the following three alternative forms:
Z
Z R
Z M
1 M
Gm
dm = 3
dm =
p dV
U=
r
2 0
0
0
Here, M is the total mass, R the radius, (r) the gravitational potential per unit mass (defined such that 0 as r ),
p(r) the pressure, and dV = 4 r2 dr.
11.9. Suppose that the pressure and density inside a spherically symmetric star are related according to the polytropic gas law,
p = K (1+n)/n ,
where n is termed the polytropic index. Let = c n , where c is the central mass density. Demonstrate that satisfies the
Lane-Emden equation
!
1 d 2 d

= n ,
2 d
d
where r = a , and
=

"

(n + 1) K (n1)/n
c
4 G

#1/2

Equilibrium of Compressible Fluids

221

Show that the physical solution to the Lane-Emden equation, which is such that (0) = 1 and (1 ) = 0, for some 1 > 0, is
=1

2
6

for n = 0,
=

sin

for n = 1, and
1
(1 + 2 /3)1/2
for n = 5. Determine the ratio of the central density to the mean density in all three cases. Finally, demonstrate that, in the
general case, the total gravitational potential energy can be written
=

U=

3 G M2
,
5n R

where M is the total mass, and R = a 1 the radius.


11.10. A spherically symmetric star of radius R has a mass density of the form
(r) = c (1 r/R).
Show that the central mass density is four times the mean density. Demonstrate that the central pressure is
pc =

5 G M2
,
4 R

where M is the mass of the star. Finally, show that the total gravitational potential energy of the star can be written
U=

26 G M 2
.
35 R

222

FLUID MECHANICS

Vectors and Vector Fields

223

A Vectors and Vector Fields

A.1 Introduction
This Appendix outlines those aspects of vector algebra, vector calculus, and vector field theory that are helpful in the
study of fluid dynamics.

A.2 Scalars and Vectors


Many physical entities (e.g., mass, energy) are entirely defined by a numerical magnitude (expressed in appropriate
units). Such entities, which have no directional element, are known as scalars. Moreover, since scalars can be
represented by real numbers it follows that they obey the laws of ordinary algebra. However, there exits a second
class of physical entities (e.g., velocity, acceleration, force) that are only completely defined when both a numerical
magnitude and a direction in space are specified. Such entities are known as vectors. By definition, a vector obeys

the same algebra as a displacement in space, and may thus be represented geometrically by a straight-line, PQ (say),
where the arrow indicates the direction of the displacement (i.e., from point P to point Q). See Figure A.1. The
magnitude of the vector is represented by the length of the straight-line.
It is conventional to denote vectors by bold-faced symbols (e.g., a, F) and scalars by non-bold-faced symbols (e.g.,
r, S ). The magnitude of a general vector, a, is denoted |a|, or just a, and is, by definition, always greater than or equal
to zero. It is convenient to define a vector with zero magnitudethis is denoted 0, and has no direction. Finally, two
vectors, a and b, are said to be equal when their magnitudes and directions are both identical.

A.3 Vector Algebra

Suppose that the displacements PQ and QR represent the vectors a and b, respectively. See Figure A.2. It can be seen

that the result of combining these two displacements is to give the net displacement PR. Hence, if PR represents the
vector c then we can write
c = a + b.
(A.1)
This defines vector addition. By completing the parallelogram PQRS , we can also see that

PR = PQ + QR = PS + S R .

(A.2)

However, PS has the same length and direction as QR, and, thus, represents the same vector, b. Likewise, PQ and S R
both represent the vector a. Thus, the above equation is equivalent to
c = a + b = b + a.

(A.3)

P
Figure A.1: A vector.

224

FLUID MECHANICS

R
b

c=a+b

S
a

b
P

Figure A.2: Vector addition.

c=ab
a

c
b

Figure A.3: Vector subtraction.


We conclude that the addition of vectors is commutative. It can also be shown that the associative law holds: i.e.,
a + (b + c) = (a + b) + c.

(A.4)

The null vector, 0, is represented by a displacement of zero length and arbitrary direction. Since the result of
combining such a displacement with a finite length displacement is the same as the latter displacement by itself, it
follows that
a + 0 = a,
(A.5)
where a is a general vector. The negative of a is defined as that vector which has the same magnitude, but acts in the
opposite direction, and is denoted a. The sum of a and a is thus the null vector: i.e.,
a + (a) = 0.

(A.6)

We can also define the difference of two vectors, a and b, as


c = a b = a + (b).

(A.7)

This definition of vector subtraction is illustrated in Figure A.3.


If n > 0 is a scalar then the expression n a denotes a vector whose direction is the same as a, and whose magnitude
is n times that of a. (This definition becomes obvious when n is an integer.) If n is negative then, since n a = |n| (a),
it follows that n a is a vector whose magnitude is |n| times that of a, and whose direction is opposite to a. These

Vectors and Vector Fields

225

z
y

Figure A.4: A right-handed Cartesian coordinate system.


definitions imply that if n and m are two scalars then
n (m a) =

n m a = m (n a),

(A.8)

(n + m) a =

n a + m a,

(A.9)

n (a + b) =

n a + n b.

(A.10)

A.4 Cartesian Components of a Vector


Consider a Cartesian coordinate system Oxyz consisting of an origin, O, and three mutually perpendicular coordinate
axes, Ox, Oy, and Oz. See Figure A.4. Such a system is said to be right-handed if, when looking along the Oz
direction, a 90 clockwise rotation about Oz is required to take Ox into Oy. Otherwise, it is said to be left-handed. It
is conventional to always use a right-handed coordinate system.
It is convenient to define unit vectors, e x , ey , and ez , parallel to Ox, Oy, and Oz, respectively. Incidentally, a
unit vector is a vector whose magnitude is unity. The position vector, r, of some general point P whose Cartesian
coordinates are (x, y, z) is then given by
r = x ez + y ey + z ez .
(A.11)
In other words, we can get from O to P by moving a distance x parallel to Ox, then a distance y parallel to Oy, and
then a distance z parallel to Oz. Similarly, if a is an arbitrary vector then
a = a x e x + a y ey + a z ez ,

(A.12)

where a x , ay , and az are termed the Cartesian components of a. It is coventional to write a (a x , ay , az ). It follows
that e x (1, 0, 0), ey (0, 1, 0), and ez (0, 0, 1). Of course, 0 (0, 0, 0).
According to the three-dimensional generalization of the Pythagorean theorem, the distance OP |r| = r is given
by
q
r=

x2 + y2 + z2 .

By analogy, the magnitude of a general vector a takes the form


q
a = a x2 + ay2 + az2 .

(A.13)

(A.14)

If a (a x , ay , az ) and b (b x , by , bz ) then it is easily demonstrated that


a + b (a x + b x , ay + by , az + bz ).

(A.15)

Furthermore, if n is a scalar then it is apparent that


n a (n a x , n ay , n az ).

(A.16)

226

FLUID MECHANICS

z
P

Figure A.5: Rotation of the coordinate axes about Oz.

A.5 Coordinate Transformations


A Cartesian coordinate system allows position and direction in space to be represented in a very convenient manner.
Unfortunately, such a coordinate system also introduces arbitrary elements into our analysis. After all, two independent
observers might well choose Cartesian coordinate systems with different origins, and different orientations of the
coordinate axes. In general, a given vector a will have different sets of components in these two coordinate systems.
However, the direction and magnitude of a are the same in both cases. Hence, the two sets of components must be
related to one another in a very particular fashion. Actually, since vectors are represented by moveable line elements

in space (i.e., in Figure A.2, PQ and S R represent the same vector), it follows that the components of a general vector
are not affected by a simple shift in the origin of a Cartesian coordinate system. On the other hand, the components
are modified when the coordinate axes are rotated.
Suppose that we transform to a new coordinate system, Ox y z , which has the same origin as Oxyz, and is obtained
by rotating the coordinate axes of Oxyz through an angle about Oz. See Figure A.5. Let the coordinates of a general
point P be (x, y, z) in Oxyz and (x , y , z ) in Ox y z . According to simple trigonometry, these two sets of coordinates
are related to one another via the transformation
x

x cos + y sin ,

(A.17)

(A.18)

x sin + y cos ,

z.

(A.19)

Consider the vector displacement r OP. Note that this displacement is represented by the same symbol, r, in
both coordinate systems, since the magnitude and direction of r are manifestly independent of the orientation of the
coordinate axes. The coordinates of r do depend on the orientation of the axes: i.e., r (x, y, z) in Oxyz, and
r (x , y , z ) in Ox y z . However, they must depend in a very specific manner [i.e., Equations (A.17)(A.19)] which
preserves the magnitude and direction of r.
The components of a general vector a transform in an analogous manner to Equations (A.17)(A.19): i.e.,
a x

= a x cos + ay sin ,

(A.20)

a y

= a x sin + ay cos ,

(A.21)

az

= az .

(A.22)

Moreover, there are similar transformation rules for rotation about Ox and Oy. Equations (A.20)(A.22) effectively
constitute the definition of a vector: i.e., the three quantities (a x , ay , az ) are the components of a vector provided that
they transform under rotation of the coordinate axes about Oz in accordance with Equations (A.20)(A.22). (And also
transform correctly under rotation about Ox and Oy). Conversely, (a x , ay , az ) cannot be the components of a vector
if they do not transform in accordance with Equations (A.20)(A.22). Of course, scalar quantities are invariant under

Vectors and Vector Fields

227

rotation of the coordinate axes. Thus, the individual components of a vector (a x , say) are real numbers, but they are not
scalars. Displacement vectors, and all vectors derived from displacements (e.g., velocity, acceleration), automatically
satisfy Equations (A.20)(A.22). There are, however, other physical quantities that have both magnitude and direction,
but which are not obviously related to displacements. We need to check carefully to see whether these quantities are
really vectors (see Sections A.7 and A.9).

A.6 Scalar Product


A scalar quantity is invariant under all possible rotational transformations. The individual components of a vector
are not scalars because they change under transformation. Can we form a scalar out of some combination of the
components of one, or more, vectors? Suppose that we were to define the percent product,
a % b a x bz + ay b x + az by = scalar number,

(A.23)

for general vectors a and b. Is a % b invariant under transformation, as must be the case if it is a scalar number?
Let us consider an example. Suppose that a (0, 1, 0) and b (1, 0, 0). It is easily seen that a % b= 1. Let
us now rotate
thecoordinate axes through 45 about Oz. In the new coordinate system, a (1/ 2, 1/ 2, 0) and

b (1/ 2, 1/ 2, 0), giving a % b = 1/2. Clearly, a % b is not invariant under rotational transformation, so the
above definition is a bad one.
Consider, now, the dot product or scalar product:
a b a x b x + ay by + az bz = scalar number.

(A.24)

Let us rotate the coordinate axes though degrees about Oz. According to Equations (A.20)(A.22), a b takes the
form
ab =

(a x cos + ay sin ) (b x cos + by sin )


+(a x sin + ay cos ) (b x sin + by cos ) + az bz

a x b x + ay by + az bz

(A.25)

in the new coordinate system. Thus, a b is invariant under rotation about Oz. It can easily be shown that it is also
invariant under rotation about Ox and Oy. We conclude that a b is a true scalar, and that the above definition is a
good one. Incidentally, a b is the only simple combination of the components of two vectors which transforms like a
scalar. It is readily shown that the dot product is commutative and distributive: i.e.,
ab =
a (b + c) =

b a,
a b + a c.

(A.26)

The associative property is meaningless for the dot product, because we cannot have (a b) c, since a b is scalar.
We have shown that the dot product a b is coordinate independent. But what is the geometric significance of this?
Well, in the special case where a = b, we get
a b = a x2 + ay2 + az2 = |a|2 = a2 .

(A.27)

So, the invariance of a a is equivalent to the invariance of the magnitude of vector a under transformation.
Let us now investigate the general case. The length squared of AB in the vector triangle shown in Figure A.6 is
(b a) (b a) = |a|2 + |b|2 2 a b.

(A.28)

However, according to the cosine rule of trigonometry,


(AB)2 = (OA)2 + (OB)2 2 (OA) (OB) cos ,

(A.29)

228

FLUID MECHANICS

ba

a
Figure A.6: A vector triangle.

where (AB) denotes the length of side AB. It follows that


a b = |a| |b| cos .

(A.30)

In this case, the invariance of a b under transformation is equivalent to the invariance of the angle subtended between
the two vectors. Note that if a b = 0 then either |a| = 0, |b| = 0, or the vectors a and b are mutually perpendicular.
The angle subtended between two vectors can easily be obtained from the dot product: i.e.,
cos =

ab
.
|a| |b|

(A.31)

The work W performed by a constant force F which moves an object through a displacement r is the product of
the magnitude of F times the displacement in the direction of F. If the angle subtended between F and r is then
W = |F| (|r| cos ) = F r.

(A.32)

The work dW performed by a non-constant force f which moves an object through an infinitesimal displacement
dr in a time interval dt is dW = f dr. Thus, the rate at which the force does work on the object, which is usually
referred to as the power, is P = dW/dt = f dr/dt, or P = f v, where v = dr/dt is the objects instantaneous velocity.

A.7 Vector Area


Suppose that we have planar surface of scalar area S . We can define a vector area S whose magnitude is S , and whose
direction is perpendicular to the plane, in the sense determined by a right-hand circulation rule (see Section A.8)
applied to the rim, assuming that a direction of circulation around the rim is specified. See Figure A.7. This quantity
clearly possesses both magnitude and direction. But is it a true vector? We know that if the normal to the surface makes
an angle x with the x-axis then the area seen looking along the x-direction is S cos x . This is the x-component of S
(since S x = e x S = e x n S = cos x S , where n is the unit normal to the surface). Similarly, if the normal makes an
angle y with the y-axis then the area seen looking along the y-direction is S cos y . This is the y-component of S. If
we limit ourselves to a surface whose normal is perpendicular to the z-direction then x = /2 y = . It follows that
S = S (cos , sin , 0). If we rotate the basis about the z-axis by degrees, which is equivalent to rotating the normal
to the surface about the z-axis by degrees, so that , then
S x = S cos ( ) = S cos cos + S sin sin = S x cos + S y sin ,

(A.33)

which is the correct transformation rule for the x-component of a vector. The other components transform correctly
as well. This proves both that a vector area is a true vector and that the components of a vector area are the projected
areas seen looking down the coordinate axes.
According to the vector addition theorem, the projected area of two plane surfaces, joined together at a line, looking
along the x-direction (say) is the x-component of the resultant of the vector areas of the two surfaces. Likewise, for

Vectors and Vector Fields

229

Figure A.7: A vector area.


many joined-up plane areas, the net area seen looking down the x-axis, which is the same as the area of the outer rim
seen looking down the x-axis, is the x-component of the resultant of all the vector areas: i.e.,
X
Si .
(A.34)
S=
i

If we approach a limit, by letting the number of plane facets increase, and their areas reduce, then we obtain a continuous surface denoted by the resultant vector area
X
S=
Si .
(A.35)
i

It is clear that the area of the rim seen looking down the x-axis is just S x . Similarly, for the areas of the rim seen
looking down the other coordinate axes. Note that it is the rim of the surface that determines the vector area rather
than the nature of the surface spanning the rim. So, two different surfaces sharing the same rim both possess the same
vector area.
In conclusion, a loop (not all in one plane) has a vector area S which is the resultant of the component vector areas
of any surface ending on the loop. The components of S are the areas of the loop seen looking down the coordinate
axes. As a corollary, a closed surface has S = 0, since it does not possess a rim.

A.8 Vector Product


We have discovered how to construct a scalar from the components of two general vectors, a and b. Can we also
construct a vector which is not just a linear combination of a and b? Consider the following definition:
a b (a x b x , ay by , az bz ).

(A.36)

Is a b a proper vector? Suppose that a = (0, 1, 0),


0, 0). In thiscase, a b = 0. However, if we rotate the
b = (1,
coordinate axes through 45 about Oz then a = (1/ 2, 1/ 2, 0), b = (1/ 2, 1/ 2, 0), and a b = (1/2, 1/2, 0).
Thus, a b does not transform like a vector, because its magnitude depends on the choice of axes. So, above definition
is a bad one.
Consider, now, the cross product or vector product:
a b (ay bz az by , az b x a x bz , a x by ay b x ) = c.

(A.37)

Does this rather unlikely combination transform like a vector? Let us try rotating the coordinate axes through an angle
about Oz using Equations (A.20)(A.22). In the new coordinate system,
c x

(a x sin + ay cos ) bz az (b x sin + by cos )

(ay bz az by ) cos + (az b x a x bz ) sin

c x cos + cy sin .

(A.38)

230

FLUID MECHANICS

thumb
ab

middle finger
b

index finger

Figure A.8: The right-hand rule for cross products. Here, is less that 180 .
Thus, the x-component of a b transforms correctly. It can easily be shown that the other components transform
correctly as well, and that all components also transform correctly under rotation about Ox and Oy. Thus, a b is a
proper vector. Incidentally, a b is the only simple combination of the components of two vectors that transforms like
a vector (which is non-coplanar with a and b). The cross product is anticommutative,
a b = b a,

(A.39)

a (b + c) = a b + a c,

(A.40)

a (b c) , (a b) c.

(A.41)

distributive,
but is not associative,
The cross product transforms like a vector, which means that it must have a well-defined direction and magnitude.
We can show that a b is perpendicular to both a and b. Consider a a b. If this is zero then the cross product must
be perpendicular to a. Now,
aab =
=

a x (ay bz az by ) + ay (az b x a x bz ) + az (a x by ay b x )
0.

(A.42)

Therefore, a b is perpendicular to a. Likewise, it can be demonstrated that a b is perpendicular to b. The vectors


a, b, and a b form a right-handed set, like the unit vectors e x , ey , and ez . In fact, e x ey = ez . This defines a unique
direction for a b, which is obtained from a right-hand rule. See Figure A.8.
Let us now evaluate the magnitude of a b. We have
(a b)2

(ay bz az by )2 + (az b x a x bz )2 + (a x by ay b x )2

(a x2 + ay2 + az2 ) (b x2 + by2 + bz2 ) (a x b x + ay by + az bz )2

|a|2 |b|2 (a b)2

|a|2 |b|2 |a|2 |b|2 cos2 = |a|2 |b|2 sin2 .

(A.43)

Thus,
|a b| = |a| |b| sin ,

(A.44)

where is the angle subtended between a and b. Clearly, a a = 0 for any vector, since is always zero in this case.
Also, if a b = 0 then either |a| = 0, |b| = 0, or b is parallel (or antiparallel) to a.
Consider the parallelogram defined by the vectors a and b. See Figure A.9. The scalar area of the parallelogram
is a b sin . By convention, the vector area has the magnitude of the scalar area, and is normal to the plane of the
parallelogram, in the sense obtained from a right-hand circulation rule by rotating a on to b (through an acute angle):

Vectors and Vector Fields

231

i.e., if the fingers of the right-hand circulate in the direction of rotation then the thumb of the right-hand indicates the
direction of the vector area. So, the vector area is coming out of the page in Figure A.9. It follows that
S = a b,

(A.45)

Suppose that a force F is applied at position r. See Figure A.10. The torque about the origin O is the product
of the magnitude of the force and the length of the lever arm OQ. Thus, the magnitude of the torque is |F| |r| sin .
The direction of the torque is conventionally defined as the direction of the axis through O about which the force tries
to rotate objects, in the sense determined by a right-hand circulation rule. Hence, the torque is out of the page in
Figure A.10. It follows that the vector torque is given by
= r F.

(A.46)

The angular momentum, l, of a particle of linear momentum p and position vector r is simply defined as the
moment of its momentum about the origin: i.e.,
l = r p.
(A.47)

A.9 Rotation
Let us try to define a rotation vector whose magnitude is the angle of the rotation, , and whose direction is parallel
to the axis of rotation, in the sense determined by a right-hand circulation rule. Unfortunately, this is not a good vector.
The problem is that the addition of rotations is not commutative, whereas vector addition is commuative. Figure A.11
shows the effect of applying two successive 90 rotations, one about Ox, and the other about the Oz, to a standard
six-sided die. In the left-hand case, the z-rotation is applied before the x-rotation, and vice versa in the right-hand case.
It can be seen that the die ends up in two completely different states. In other words, the z-rotation plus the x-rotation
does not equal the x-rotation plus the z-rotation. This non-commuting algebra cannot be represented by vectors. So,
although rotations have a well-defined magnitude and direction, they are not vector quantities.
But, this is not quite the end of the story. Suppose that we take a general vector a and rotate it about Oz by a small
angle z . This is equivalent to rotating the coordinate axes about the Oz by z . According to Equations (A.20)
(A.22), we have
a a + z ez a,
(A.48)
where use has been made of the small angle approximations sin and cos 1. The above equation can easily be
generalized to allow small rotations about Ox and Oy by x and y , respectively. We find that
a a + a,

(A.49)

= x e x + y ey + z ez .

(A.50)

where
Clearly, we can define a rotation vector, , but it only works for small angle rotations (i.e., sufficiently small that the
small angle approximations of sine and cosine are good). According to the above equation, a small z-rotation plus a

a
S
b

a
Figure A.9: A vector parallelogram.

232

FLUID MECHANICS

P
r

r sin

Figure A.10: A torque.

z
x
z -axis

x-axis

x-axis

z -axis

Figure A.11: Effect of successive rotations about perpendicular axes on a six-sided die.

Vectors and Vector Fields

233

c
b
Figure A.12: A vector parallelepiped.
small x-rotation is (approximately) equal to the two rotations applied in the opposite order. The fact that infinitesimal
rotation is a vector implies that angular velocity,

,
(A.51)
t
must be a vector as well. Also, if a is interpreted as a(t + t) in Equation (A.49) then it follows that the equation of
motion of a vector which precesses about the origin with some angular velocity is
= lim

t0

da
= a.
dt

(A.52)

A.10 Scalar Triple Product


Consider three vectors a, b, and c. The scalar triple product is defined a b c. Now, b c is the vector area of the
parallelogram defined by b and c. So, a b c is the scalar area of this parallelogram multiplied by the component of
a in the direction of its normal. It follows that a b c is the volume of the parallelepiped defined by vectors a, b, and
c. See Figure A.12. This volume is independent of how the triple product is formed from a, b, and c, except that
a b c = a c b.

(A.53)

So, the volume is positive if a, b, and c form a right-handed set (i.e., if a lies above the plane of b and c, in the sense
determined from a right-hand circulation rule by rotating b onto c) and negative if they form a left-handed set. The
triple product is unchanged if the dot and cross product operators are interchanged,
a b c = a b c.

(A.54)

The triple product is also invariant under any cyclic permutation of a, b, and c,
a b c = b c a = c a b,

(A.55)

but any anti-cyclic permutation causes it to change sign,


a b c = b a c.

(A.56)

The scalar triple product is zero if any two of a, b, and c are parallel, or if a, b, and c are coplanar.
If a, b, and c are non-coplanar then any vector r can be written in terms of them: i.e.,
r = a + b + c.

(A.57)

Forming the dot product of this equation with b c, we then obtain


r b c = a b c,
so

(A.58)

rbc
.
(A.59)
abc
Analogous expressions can be written for and . The parameters , , and are uniquely determined provided
a b c , 0: i.e., provided that the three vectors are non-coplanar.
=

234

FLUID MECHANICS

A.11 Vector Triple Product


For three vectors a, b, and c, the vector triple product is defined a (b c). The brackets are important because
a (b c) , (a b) c. In fact, it can be demonstrated that
a (b c) (a c) b (a b) c

(A.60)

(a b) c (a c) b (b c) a.

(A.61)

and

Let us try to prove the first of the above theorems. The left-hand side and the right-hand side are both proper
vectors, so if we can prove this result in one particular coordinate system then it must be true in general. Let us take
convenient axes such that Ox lies along b, and c lies in the x-y plane. It follows that b (b x , 0, 0), c (c x , cy , 0),
and a (a x , ay , az ). The vector b c is directed along Oz: i.e., b c (0, 0, b x cy ). Hence, a (b c) lies in the
x-y plane: i.e., a (b c) (ay b x cy , a x b x cy , 0). This is the left-hand side of Equation (A.60) in our convenient
coordinate system. To evaluate the right-hand side, we need a c = a x c x + ay cy and a b = a x b x . It follows that the
right-hand side is
RHS =
=

( [a x c x + ay cy ] b x , 0, 0) (a x b x c x , a x b x cy , 0)
(ay cy b x , a x b x cy , 0) = LHS,

(A.62)

which proves the theorem.

A.12 Vector Calculus


Suppose that vector a varies with time, so that a = a(t). The time derivative of the vector is defined
#
"
a(t + t) a(t)
da
.
= lim
dt t0
t

(A.63)

When written out in component form this becomes


!
da
da x day daz
.

,
,
dt
dt dt dt

(A.64)

Suppose that a is, in fact, the product of a scalar (t) and another vector b(t). What now is the time derivative of
a? We have
da x
d
d
db x
= ( b x ) =
bx +
,
(A.65)
dt
dt
dt
dt
which implies that
db
da d
=
b+ .
dt
dt
dt

(A.66)

da
db
d
(a b) =
b+a
,
dt
dt
dt

(A.67)

d
da
db
(a b) =
b+a
.
dt
dt
dt

(A.68)

Moreover, it is easily demonstrated that

and

Hence, it can be seen that the laws of vector differentiation are analogous to those in conventional calculus.

Vectors and Vector Fields

235

P
O

Q l

P
.

Figure A.13: A line integral.

A.13 Line Integrals


Consider a two-dimensional function f (x, y) which is defined for all x and y. What is meant by the integral of f along
a given curve joining the points P and Q in the x-y plane? Well, we first draw out f as a function of length l along the
path. See Figure A.13. The integral is then simply given by
Z

f (x, y) dl = Area under the curve,

(A.69)

p
where dl = dx2 + dy2 .
As an example of this, consider the integral of
f (x, y) = x y2 between P and Q along the two routes indicated in
Figure A.14. Along route 1 we have x = y, so dl = 2 dx. Thus,

Z 1
Z Q
2
x3 2 dx =
x y2 dl =
.
(A.70)
4
0
P
The integration along route 2 gives
Z

Q
2

x y dl =
P



Z 1


2
x y dx +
x y dy
y=0
x=1
0
0
Z 1
1
0+
y2 dy = .
3
0
Z

(A.71)

Note that the integral depends on the route taken between the initial and final points.
The most common type of line integral is that in which the contributions from dx and dy are evaluated separately,
rather that through the path-length element dl: i.e.,
Z Q


f (x, y) dx + g(x, y) dy .
(A.72)
P

As an example of this, consider the integral

Q
P

y dx + x3 dy

(A.73)

along the two routes indicated in Figure A.15. Along route 1 we have x = y + 1 and dx = dy, so
Z

Q
P

h
i Z
y dx + x3 dy =

h
i 17
y dy + (y + 1)3 dy =
.
4

(A.74)

236

FLUID MECHANICS

Q = (1, 1)

2
2
x

P = (0, 0)

Figure A.14: An example line integral.

y
2

Q = (2, 1)

P = (1, 0)
Figure A.15: An example line integral.

Along route 2,

Q
P

i Z
y dx + x3 dy =



Z 2


7
x3 dy +
y dx = .
x=1

4
1
y=1

Again, the integral depends on the path of integration.


Suppose that we have a line integral which does not depend on the path of integration. It follows that
Z Q
( f dx + g dy) = F(Q) F(P)

(A.75)

(A.76)

for some function F. Given F(P) for one point P in the x-y plane, then
Z Q
( f dx + g dy)
F(Q) = F(P) +

(A.77)

defines F(Q) for all other points in the plane. We can then draw a contour map of F(x, y). The line integral between
points P and Q is simply the change in height in the contour map between these two points:
Z Q
Z Q
( f dx + g dy) =
dF(x, y) = F(Q) F(P).
(A.78)
P

Thus,
dF(x, y) = f (x, y) dx + g(x, y) dy.
For instance, if F = x3 y then dF = 3 x2 y dx + x3 dy and
Z Q
 h
iQ
3 x2 y dx + x3 dy = x3 y
P

(A.79)

(A.80)

Vectors and Vector Fields

237

2
2

Q
O a

P
x

Figure A.16: An example vector line integral.


is independent of the path of integration.
It is clear that there are two distinct types of line integralthose which depend only on their endpoints and not on
the path of integration, and those which depend both on their endpoints and the integration path. Later on, we shall
learn how to distinguish between these two types (see Section A.18).

A.14 Vector Line Integrals


A vector field is defined as a set of vectors associated with each point in space. For instance, the velocity v(r) in a
moving liquid (e.g., a whirlpool) constitutes a vector field. By analogy, a scalar field is a set of scalars associated with
each point in space. An example of a scalar field is the temperature distribution T (r) in a furnace.
Consider a general vector field A(r). Let dr (dx, dy, dz) be the vector element of line length. Vector line
integrals often arise as
Z Q
Z Q
A dr =
(A x dx + Ay dy + Az dz).
(A.81)
P

For instance, if A is a force-field then the line integral is the work done in going from P to Q.
As an example, consider the work done by a repulsive inverse-square central field, F = r/|r3 |. The element of
work done is dW = F dr. Take P = (, 0, 0) and Q = (a, 0, 0). The first route considered is along the x-axis, so
!
" #a
Z a
1
1
1
= .
(A.82)
W=
2 dx =
x
a
x

The second route is, firstly, around a large circle (r = constant) to the point (a, , 0), and then parallel to the y-axis.
See Figure A.16. In the first part, no work is done, since F is perpendicular to dr. In the second part,
W=

#0
"
y dy
1
1
=
= .
(a2 + y2 )3/2
(y2 + a2 )1/2 a

(A.83)

In this case, the integral is independent of the path. However, not all vector line integrals are path independent.

A.15 Surface Integrals


Let us take a surface S , that is not necessarily co-planar, and divide it up into (scalar) elements S i . Then
Z Z
X
f (x, y, z) S i
f (x, y, z) dS = lim
S

S i 0

(A.84)

238

FLUID MECHANICS

y
y2

dy
y1
x1

x2

Figure A.17: Decomposition of a surface integral.


is a surface integral. For instance, the volume of water in a lake of depth D(x, y) is
Z Z
V=
D(x, y) dS .

(A.85)

To evaluate this integral we must split the calculation into two ordinary integrals. The volume in the strip shown in
Figure A.17 is
"Z x2
#
D(x, y) dx dy.
(A.86)
x1

Note that the limits x1 and x2 depend on y. The total volume is the sum over all strips: i.e.,
# Z Z
Z y2 "Z x2 (y)
D(x, y) dx dy.
dy
V=
D(x, y) dx
y1

(A.87)

x1 (y)

Of course, the integral can be evaluated by taking the strips the other way around: i.e.,
Z x2 Z y2 (x)
D(x, y) dy.
dx
V=

(A.88)

y1 (x)

x1

Interchanging the order of integration is a very powerful and useful trick. But great care must be taken when evaluating
the limits.
As an example, consider
Z Z
x y2 dx dy,

(A.89)

where S is shown in Figure A.18. Suppose that we evaluate the x integral first:
dy

Let us now evaluate the y integral:

1y

"

x2
x y dx = y dy
2
2

#1y
0

y2
(1 y)2 dy.
2

!
1
y4
y2
dy =
y3 +
.
2
2
60

We can also evaluate the integral by interchanging the order of integration:


Z 1
Z 1x
Z 1
x
1
x dx
y2 dy =
(1 x)3 dx =
.
60
0
0
0 3

(A.90)

(A.91)

(A.92)

Vectors and Vector Fields

239

y
(0, 1)
1y =x

(1, 0)

(0, 0)

Figure A.18: An example surface integral.


In some cases, a surface integral is just the product of two separate integrals. For instance,
Z Z
x2 y dx dy

(A.93)

where S is a unit square. This integral can be written


Z 1 Z 1
Z
2
dx
x y dy =
0

1
2

x dx
0

! Z

y dy =

1 1 1
= ,
3 2 6

(A.94)

since the limits are both independent of the other variable.

A.16 Vector Surface Integrals


Surface integrals often occur during vector analysis. For instance, the rate of flow of a liquid of velocity v through an
infinitesimal surface of vector area dS is v dS. The net rate of flow through a surface S made up of lots of infinitesimal
surfaces is
Z Z
hX
i
v dS = lim
v cos dS ,
(A.95)
dS 0

where is the angle subtended between the normal to the surface and the flow velocity.
Analogously to line integrals, most surface integrals depend both on the surface and the rim. But some (very
important) integrals depend only on the rim, and not on the nature of the surface which spans it. As an example of
this, consider incompressible fluid flow between two surfaces S 1 and S 2 which end on the same rim. See Figure A.23.
The volume between the surfaces is constant, so what goes in must come out, and
Z Z
Z Z
v dS.
(A.96)
v dS =
S2

S1

It follows that

Z Z

v dS

(A.97)

depends only on the rim, and not on the form of surfaces S 1 and S 2 .

A.17 Volume Integrals


A volume integral takes the form

Z Z Z

f (x, y, z) dV,

(A.98)

240

FLUID MECHANICS

where V is some volume, and dV = dx dy dz is a small volume element. The volume element is sometimes written
d3 r, or even d.
As an example of a volume integral, let us evaluate the center of gravity of a solid pyramid. Suppose that the
pyramid has a square base of side a, a height a, and is composed of material of uniform density. Let the centroid of
the base lie at the origin, and let the apex lie at (0, 0, a). By symmetry, the center of mass lies on the line joining the
centroid to the apex. In fact, the height of the center of mass is given by
,Z Z Z
Z Z Z
z=
z dV
dV.
(A.99)
The bottom integral is just the volume of the pyramid, and can be written
Z a
Z a
Z Z Z
Z a Z (az)/2 Z (az)/2
2
(a2 2 a z + z2 ) dz
(a z) dz =
dx =
dy
dz
dV =
0

(az)/2

(az)/2

h
ia 1
a2 z a z2 + z3 /3 = a3 .
0
3

(A.100)

Here, we have evaluated the z-integral last because the limits of the x- and y- integrals are z-dependent. The top integral
takes the form
Z a
Z a
Z (az)/2 Z (az)/2
Z Z Z
Z a
(z a2 2 a z2 + z3 ) dz
z (a z)2 dz =
dx =
dy
z dz
z dV =
0

=
Thus,

(az)/2

(az)/2

a2 z2 /2 2 a z3 /3 + z4 /4

ia
0

1 4
=
a.
12

,
1 4 1 3 1
a
a = a.
z =
12
3
4

(A.101)

(A.102)

In other words, the center of mass of a pyramid lies one quarter of the way between the centroid of the base and the
apex.

A.18 Gradient
A one-dimensional function f (x) has a gradient d f /dx which is defined as the slope of the tangent to the curve at x.
We wish to extend this idea to cover scalar fields in two and three dimensions.
Consider a two-dimensional scalar field h(x, y), which is (say) height above sea-level in a hilly region. Let dr
(dx, dy) be an element of horizontal distance. Consider dh/dr, where dh is the change in height after moving an
infinitesimal distance dr. This quantity is somewhat like the one-dimensional gradient, except that dh depends on the
direction of dr, as well as its magnitude. In the immediate vicinity of some point P, the slope reduces to an inclined
plane. See Figure A.19. The largest value of dh/dr is straight up the slope. It is easily shown that for any other
direction
!
dh
dh
=
cos ,
(A.103)
dr
dr max
where is the angle shown in Figure A.19. Let us define a two-dimensional vector, grad h, called the gradient of h,
whose magnitude is (dh/dr)max, and whose direction is the direction of steepest ascent. The cos variation exhibited
in the above expression ensures that the component of grad h in any direction is equal to dh/dr for that direction.
The component of dh/dr in the x-direction can be obtained by plotting out the profile of h at constant y, and then
finding the slope of the tangent to the curve at given x. This quantity is known as the partial derivative of h with respect
to x at constant y, and is denoted (h/x)y. Likewise, the gradient of the profile at constant x is written (h/y) x. Note
that the subscripts denoting constant x and constant y are usually omitted, unless there is any ambiguity. It follows
that in component form
!
h h
.
(A.104)
,
grad h
x y

Vectors and Vector Fields

241
contours of h(x, y)

high
dr

direction of steepest ascent

low
x

Figure A.19: A two-dimensional gradient.


Now, the equation of the tangent plane at P = (x0 , y0 ) is
hT (x, y) = h(x0 , y0 ) + (x x0 ) + (y y0 ).

(A.105)

This has the same local gradients as h(x, y), so


=

h
,
x

h
,
y

(A.106)

by differentiation of the above. For small dx = x x0 and dy = y y0 , the function h is coincident with the tangent
plane, so
h
h
dx +
dy.
(A.107)
dh =
x
y
But, grad h (h/x, h/y) and dr (dx, dy), so
dh = grad h dr.

(A.108)

Incidentally, the above equation demonstrates that grad h is a proper vector, since the left-hand side is a scalar, and,
according to the properties of the dot product, the right-hand side is also a scalar provided that dr and grad h are both
proper vectors (dr is an obvious vector, because it is directly derived from displacements).
Consider, now, a three-dimensional temperature distribution T (x, y, z) in (say) a reaction vessel. Let us define
grad T , as before, as a vector whose magnitude is (dT/dr)max , and whose direction is the direction of the maximum
gradient. This vector is written in component form
!
T T T
,
,
.
(A.109)
grad T
x y z
Here, T/x (T/x)y,z is the gradient of the one-dimensional temperature profile at constant y and z. The change
in T in going from point P to a neighbouring point offset by dr (dx, dy, dz) is
T
T
T
dx +
dy +
dz.
x
y
z

(A.110)

dT = grad T dr.

(A.111)

dT = grad T dr = 0.

(A.112)

dT =
In vector form, this becomes
Suppose that dT = 0 for some dr. It follows that

242

FLUID MECHANICS

T = constant

grad T
dr

isotherms
Figure A.20: Isotherms.
So, dr is perpendicular to grad T . Since dT = 0 along so-called isotherms (i.e., contours of the temperature), we
conclude that the isotherms (contours) are everywhere perpendicular to grad T . See Figure A.20.
It is, of course, possible to integrate dT . For instance, the line integral of dT between points P and Q is written
Z Q
Z Q
grad T dr = T (Q) T (P).
(A.113)
dT =
P

RQ
This integral is clearly independent of the path taken between P and Q, so P grad T dr must be path independent.
RQ
Consider a vector field A(r). In general, the line integral P A dr depends on the path taken between the end
points. However, for some special vector fields the integral is path independent. Such fields are called conservative
fields. It can be shown that if A is a conservative field then A = grad V for some scalar field V. The proof of this is
straightforward. Keeping P fixed, we have
Z Q
A dr = V(Q),
(A.114)
P

where V(Q) is a well-defined function, due to the path independent nature of the line integral. Consider moving the
position of the end point by an infinitesimal amount dx in the x-direction. We have
Z Q+dx
A dr = V(Q) + A x dx.
(A.115)
V(Q + dx) = V(Q) +
Q

Hence,
V
= Ax,
x
with analogous relations for the other components of A. It follows that

(A.116)

A = grad V.

(A.117)
R
The force field due to gravity is a good example of a conservative field. Now, if A(r) is a force-field then A dr
is the work done in traversing some path. If A is conservative then
I
A dr = 0,
(A.118)
H
where corresponds to the line integral around a closed loop. The fact that zero net work is done in going around a
closed loop is equivalent to the conservation of energy (which is why conservative fields are called conservative). A
good example of a non-conservative
field is the force field due to friction. Clearly, a frictional system loses energy in
H
going around a closed cycle, so A dr , 0.

Vectors and Vector Fields

243

A.19 Grad Operator


It is useful to define the vector operator

!

,
,
,
x y z

(A.119)

which is usually called the grad or del operator. This operator acts on everything to its right in a expression, until the
end of the expression or a closing bracket is reached. For instance,
!
f f f
.
(A.120)
,
,
grad f = f
x y z
For two scalar fields and ,
grad ( ) = grad + grad

(A.121)

( ) = + .

(A.122)

can be written more succinctly as


Suppose that we rotate the coordinate axes through an angle about Oz. By analogy with Equations (A.17)(A.19),
the old coordinates (x, y, z) are related to the new ones (x , y , z ) via
x

y =
z =
Now,

x
=
x
x

y ,z

x cos y sin ,

(A.123)

x sin + y cos ,

(A.124)

z.

+
x
x

(A.125)
!

y ,z

+
y
x

y ,z

,
z

(A.126)

giving

= cos
+ sin ,

x
x
y

(A.127)

x = cos x + sin y .

(A.128)

and
It can be seen, from Equations (A.20)(A.22), that the differential operator transforms in an analogous manner to a
vector. This is another proof that f is a good vector.

A.20 Divergence
H
Let us start with a vector field A(r). Consider S A dS over some closed surface S , where dS denotes an outward
pointing surfaceHelement. This surface integral is usually called the flux of A out of S . If A represents the velocity of
some fluid then S A dS is the rate of fluid flow out of S .
If A is constant in space then it is easily demonstrated that the net flux out of S is zero,
I
I
A dS = A dS = A S = 0,
(A.129)
since the vector area S of a closed surface is zero.
Suppose, now, that A isH not uniform in space. Consider a very small rectangular volume over which A hardly
varies. The contribution to A dS from the two faces normal to the x-axis is
A x (x + dx) dy dz A x (x) dy dz =

A x
A x
dx dy dz =
dV,
x
x

(A.130)

244

FLUID MECHANICS
y + dy

z + dz

z
y
z

x + dx

Figure A.21: Flux of a vector field out of a small box.


where dV = dx dy dz is the volume element. See Figure A.21. There are analogous contributions from the sides
normal to the y- and z-axes, so the total of all the contributions is
!
I
A x Ay Az
A dS =
dV.
(A.131)
+
+
x
y
z
The divergence of a vector field is defined
div A = A =

A x Ay Az
+
+
.
x
y
z

(A.132)

Divergence is a good scalar (i.e., it is coordinate independent), since it is the dot product of the vector operator with
A. The formal definition of A is
H
A dS
.
(A.133)
A = lim
dV0
dV
This definition is independent of the shape of the infinitesimal volume element.
One of the most important results in vector field theory is the so-called divergence theorem or Gauss theorem.
This states that for any volume V surrounded by a closed surface S ,
Z
I
A dV,
(A.134)
A dS =
V

where dS is an outward pointing volume


element. The proof is very straightforward. We divide up the volume into
R
lots of very small cubes, and sum A dS over all of the surfaces. The contributions from the interior surfaces cancel
out, leaving just the contribution from the outer surface. See Figure
A.22. We can use Equation (A.131) for each cube
R
individually. This tells us that the summation is equivalent to A dV over the whole volume. Thus, the integral
of A dS over the outer surface is equal to the integral of A over the whole volume, which proves the divergence
theorem.
Now, for a vector field with A = 0,
I
A dS = 0
(A.135)
S

for any closed surface S . So, for two surfaces, S 1 and S 2 , on the same rim,
Z
Z
A dS.
A dS =
S1

(A.136)

S2

See Figure A.23. (Note that the direction of the surface elements on S 1 has been reversed relative to those on the
closed surface. Hence, the sign of the associated surface integral is also reversed.) Thus, if A = 0 then the surface

Vectors and Vector Fields

245

S
interior contributions cancel
.

exterior contributions survive

Figure A.22: The divergence theorem.

S2
S1
rim

Figure A.23: Two surfaces spanning the same rim (right), and the equivalent closed surface (left).

246

FLUID MECHANICS

2
Figure A.24: Divergent lines of force.

integral depends on the rim but not on the nature of the surface which spans it. On the other hand, if A , 0 then
the integral depends on both the rim and the surface.
H
Consider an incompressible fluid whose velocity field is v. It is clear that v dS = 0 for any
R closed surface, since
what flows into the surface must flow out again. Thus, according to the divergence theorem, v dV = 0 for any
volume. The only way in which this is possible is if v is everywhere zero. Thus, the velocity components of an
incompressible fluid satisfy the following differential relation:
v x vy vz
+
+
= 0.
x
y
z

(A.137)

It is sometimes helpful to represent a vector field A by lines of force or field-lines. The direction of a line of force
at any point is the same as the local direction of A. The density of lines (i.e., the number of lines crossing a unit surface
perpendicular to A) is equal to |A|. For instance, in Figure A.24, |A| is larger at point 1 than at point 2. The number of
lines crossing a surface element dS is A dS. So, the net number of lines leaving a closed surface is
Z
I
A dV.
(A.138)
A dS =
S

If A = 0 then there is no net flux of lines out of any surface. Such a field is called a solenoidal vector field. The
simplest example of a solenoidal vector field is one in which the lines of force all form closed loops.

A.21 Laplacian Operator


So far we have encountered

!

=
,
,
,
x y z
which is a vector field formed from a scalar field, and
A=

A x Ay Az
+
+
,
x
y
z

(A.139)

(A.140)

which is a scalar field formed from a vector field. There are two ways in which we can combine gradient and divergence. We can either form the vector field (A) or the scalar field (). The former is not particularly interesting,
but the scalar field () turns up in a great many physical problems, and is, therefore, worthy of discussion.
Let us introduce the heat flow vector h, which is the rate of flow of heat energy per unit area across a surface
perpendicular to the direction of h. In many substances, heat flows directly down the temperature gradient, so that we
can write
h = T,
(A.141)
H
where is the thermal conductivity. The net rate of heat flow S h dS out of some closed surface S must be equal to
the rate of decrease of heat energy in the volume V enclosed by S . Thus, we have
!
Z
I

c T dV ,
(A.142)
h dS =
t
S

Vectors and Vector Fields

247

where c is the specific heat. It follows from the divergence theorem that
h = c

T
.
t

(A.143)

Taking the divergence of both sides of Equation (A.141), and making use of Equation (A.143), we obtain
T
.
t

(A.144)

c T
.
t

(A.145)

( T ) = c
If is constant then the above equation can be written
(T ) =
The scalar field (T ) takes the form
(T )

!
!
!
T
T
T
+
+
x x
y y
z z

2 T 2 T 2 T
+ 2 + 2 2 T.
x2
y
z

(A.146)

Here, the scalar differential operator


2

2
2
2
+ 2+ 2
2
x
y
z

(A.147)

is called the Laplacian. The Laplacian is a good scalar operator (i.e., it is coordinate independent) because it is formed
from a combination of divergence (another good scalar operator) and gradient (a good vector operator).
What is the physical significance of the Laplacian? In one dimension, 2 T reduces to 2 T/x2 . Now, 2 T/x2
is positive if T (x) is concave (from above) and negative if it is convex. So, if T is less than the average of T in its
surroundings then 2 T is positive, and vice versa.
In two dimensions,
2 T 2 T
2 T = 2 + 2 .
(A.148)
x
y
Consider a local minimum of the temperature. At the minimum, the slope of T increases in all directions, so 2 T is
positive. Likewise, 2 T is negative at a local maximum. Consider, now, a steep-sided valley in T . Suppose that the
bottom of the valley runs parallel to the x-axis. At the bottom of the valley 2 T/y2 is large and positive, whereas
2 T/x2 is small and may even be negative. Thus, 2 T is positive, and this is associated with T being less than the
average local value.
Let us now return to the heat conduction problem:
2 T =

c T
.
t

(A.149)

It is clear that if 2 T is positive then T is locally less than the average value, so T/t > 0: i.e., the region heats up.
Likewise, if 2 T is negative then T is locally greater than the average value, and heat flows out of the region: i.e.,
T/t < 0. Thus, the above heat conduction equation makes physical sense.

A.22 Curl
H
Consider a vector field A(r), and a loop which lies in one plane. The integral of A around
H this loop is written A dr,
where dr is a line element of the loop.
H If A is a conservative field then A = and A dr = 0 for all loops. In
general, for a non-conservative field,
A dr , 0.
H
For a small
loop
we
expect
A

dr
to be proportional to the area of the loop. Moreover, for a fixed-area loop
H
we expect A dr to depend on the orientation of the loop. One particular orientation will give the maximum value:

248

FLUID MECHANICS

z
4

z + dz
1
z

3
y

y + dy
y

Figure A.25: A vector line integral around a small rectangular loop in the y-z plane.
H

A dr = Imax . If the loop subtends an angle with this optimum orientation then we expect I = Imax cos . Let us
introduce the vector field curl A whose magnitude is
H
A dr
|curl A| = lim
(A.150)
dS 0
dS

for the orientation giving Imax . Here, dS is the area of the loop. The direction of curl A is perpendicular to the plane
of the loop, when it is in the orientation giving Imax , with the sense given by a right-hand circulation
rule.
H
Let us now express curl A in terms of the components of A. First, we shall evaluate A dr around a small
rectangle in the y-z plane. See Figure A.25. The contribution from sides 1 and 3 is
Az (y + dy) dz Az (y) dz =

Az
dy dz.
y

(A.151)

The contribution from sides 2 and 4 is


Ay (z + dz) dy + Ay (z) dy =
So, the total of all contributions gives

Ay
dy dz.
y

!
Az Ay

A dr =
dS ,
y
z

(A.152)

(A.153)

where dS = dy dz is the area of the loop.


Consider
a non-rectangular (but still small) loop in the y-z plane. We can divide it into rectangular elements, and
H
form A dr over all the resultant loops. The interior contributions cancel, so we are just left with the contribution
from the outer loop. Also, the area of the outer loop is the sum of all the areas of the inner loops. We conclude that
!
I
Az Ay
A dr =

dS x
(A.154)
y
z
is valid for a small loop dS = (dS x , 0, 0) of any shape in the y-z plane. Likewise, we can show that if the loop is in
the x-z plane then dS = (0, dS y , 0) and
!
I
A x Az
dS y .
(A.155)

A dr =
z
x
Finally, if the loop is in the x-y plane then dS = (0, 0, dS z ) and
!
I
Ay A x
dS z .

A dr =
x
y

(A.156)

Vectors and Vector Fields

249

dS

3
x

Figure A.26: Decomposition of a vector area into its Cartesian components.


Imagine an arbitrary loop of vector area dS = (dS x , dS y , dS z ). We can construct this out of three vector areas, 1,
2, and 3, directed in the x-, y-, and z-directions, respectively, as indicated in Figure A.26. If we form the line integral
around all three loops then the interior contributions cancel, and we are left with the line integral around the original
loop. Thus,
I
I
I
I
A dr =

giving

A dr1 +

A dr2 +

A dr3 ,

A dr = curl A dS = |curl A| |dS| cos ,

where
curl A =

!
Az Ay A x Az Ay A x
,

y
z z
x x
y

(A.157)

(A.158)

(A.159)

and is the angle subtended between the directions of curl A and dS. Note that
curl A = A.

(A.160)

This demonstrates that A is a good vector field, since it is the cross product of the operator (a good vector
operator) and the vector field A.
Consider a solid body rotating about the z-axis. The angular velocity is given by = (0, 0, ), so the rotation
velocity at position r is
v=r
(A.161)
[see
Equation (A.52)]. Let us evaluate v on the axis of rotation. The x-component is proportional to the integral
H
vH dr around a loop in the y-z plane. This is plainly zero. Likewise, the y-component
H is also zero. The z-component
is v dr/dS around some loop in the x-y plane. Consider a circular loop. We have v dr = 2 r r with dS = r2 .
Here, r is the perpendicular distance from the rotation axis. It follows that ( v)z = 2 , which is independent of r.
So, on the axis, v = (0 , 0 , 2 ). Off the axis, at position r0 , we can write
v = (r r0 ) + r0 .

(A.162)

The first part has the same curl as the velocity field on the axis, and the second part has zero curl, since it is constant.
Thus, v = (0, 0, 2 ) everywhere in the body. This allows us to form a physical picture of A. If we imagine
A(r) as the velocity field of some fluid then A at any given point is equal to twice the local angular rotation
velocity: i.e., 2 . Hence, a vector field with A = 0 everywhere is said to be irrotational.

250

FLUID MECHANICS

Another important result of vector field theory is the curl theorem or Stokes theorem:
Z
I
A dS,
A dr =

(A.163)

for some (non-planar) surface S bounded by a rim C. This theorem can easily be proved by splitting the loop up into
many small rectangular loops, and forming the integral around all of the resultant loops. All of the contributions from
the interior loops cancel, leaving just the contribution from the outer rim. Making use of Equation (A.158) for each of
the small loops, we can see that the contribution from all of the loops is also equal to the integral of A dS across
the whole surface. This proves the theorem.
One immediate consequence of Stokes theorem is that A is incompressible. Consider
any two surfaces, S 1
R
and S 2 , which share the same rim. See
Figure
A.23.
It
is
clear
from
Stokes
theorem
that

A
dS is the same for
H
both surfaces.H Thus, it follows
that

dS
=
0
for
any
closed
surface.
However,
we
have
from
the divergence
R
theorem that A dS = ( A) dV = 0 for any volume. Hence,
( A) 0.

(A.164)

So, A is a solenoidal field.


H
We have seen that for a conservative field
H A dr = 0 for any loop. This is entirely equivalent to A = .
However, the magnitude of A is lim dS 0 A dr/dS for some particular loop. It is clear then that A = 0 for
a conservative field. In other words,
() 0.
(A.165)
Thus, a conservative field is also an irrotational one.

A.23 Useful Vector Identities


Notation: a, b, c, d are general vectors; , are general scalar fields; A, B are general vector fields; (A )B
(A B x , A By , A Bz ) and 2 A = (2 A x , 2 Ay , 2 Az ) (but, only in Cartesian coordinatessee Appendix C).
a (b c) =

(a c) b (a b) c,

(A.166)

(a b) c =

(c a) b (c b) a,

(A.167)

(a b) (c d) =

(a c) (b d) (a d) (b c),

(A.168)

(a b) (c d) =

(a b d) c (a b c) d,

(A.169)

+ ,

(A.170)

A ( B) + B ( A) + (A )B + (B )A,

(A.171)

(A.172)

0,

(A.173)

A + A ,

(A.174)

B A A B,

(A.175)

0,

(A.176)

( A) 2 A,

(A.177)

A + A,

(A.178)

( B) A ( A) B + (B )A (A )B.

(A.179)

( ) =
(A B)

=
A

( A) =
(A B)

=
( A) =
( A) =
(A B)

A.24 Exercises

A.1. The position vectors of the four points A, B, C, and D are a, b, 3 a + 2 b, and a 3 b, respectively. Express AC, DB, BC, and

CD in terms of a and b.

Vectors and Vector Fields

251

A.2. Prove the trigonometric law of sines


sin a sin b sin c
=
=
A
B
C
using vector methods. Here, a, b, and c are the three angles of a plane triangle, and A, B, and C the lengths of the corresponding opposite sides.
A.3. Demonstrate using vectors that the diagonals of a parallelogram bisect one another. In addition, show that if the diagonals
of a quadrilateral bisect one another then it is a parallelogram.
A.4. From the inequality
a b = |a| |b| cos |a| |b|
deduce the triangle inequality
|a + b| |a| + |b|.
A.5. Find the scalar product a b and the vector product a b when
(a) a = ex + 3 ey ez ,

b = 3 ex + 2 ey + ez ,

(b) a = ex 2 ey + ez ,

b = 2 ex + ey + ez .

A.6. Which of the following statements regarding the three general vectors a, b, and c are true?
(a) c (a b) = (b a) c.

(b) a (b c) = (a b) c.
(c) a (b c) = (a c) b (a b) c.
(d) d = a + b implies that (a b) d = 0.

(e) a c = b c implies that c a c b = c |a b|.


(f) (a b) (c b) = [b (c a)] b.

A.7. Prove that the length of the shortest straight-line from point a to the straight-line joining points b and c is
|a b + b c + c a|
.
|b c|
A.8. Identify the following surfaces:
(a) |r| = a,
(b) r n = b,

(c) r n = c |r|,

(d) |r (r n) n| = d.
Here, r is the position vector, a, b, c, and d are positive constants, and n is a fixed unit vector.
A.9. Let a, b, and c be coplanar vectors related via
a + b + c = 0,
where , , and are not all zero. Show that the condition for the points with position vectors u a, v b, and w c to be colinear
is

+ + = 0.
u v w
A.10. If p, q, and r are any vectors, demonstrate that a = q + r, b = r + p, and c = p + q are coplanar provided that = 1,
where , , and are scalars. Show that this condition is satisfied when a is perpendicular to p, b to q, and c to r.
A.11. The vectors a, b, and c are non-coplanar, and form a non-orthogonal vector base. The vectors A, B, and C, defined by
A=

bc
,
abc

plus cyclic permutations, are said to be reciprocal vectors. Show that


a = (B C)/(A B C),
plus cyclic permutations.

252

FLUID MECHANICS

A.12. In the notation of the previous question, demonstrate that the plane passing through points a/, b/, and c/ is normal to
the direction of the vector
h = A + B + C.
In addition, show that the perpendicular distance of the plane from the origin is |h|1 .
H
A.13. Evaluate A dr for
x ex + y ey
A= p
x2 + y2
around the square whose sides are x = 0, x = a, y = 0, y = a.
A.14. Consider the following vector field:
A(r) = (8 x3 + 3 x2 y2 , 2 x3 y + 6 y, 6).

H
Is this field conservative? Is it solenoidal? Is it irrotational? Justify your answers. Calculate C A dr, where the curve C is
a unit circle in the x-y plane, centered on the origin, and the direction of integration is clockwise looking down the z-axis.
A.15. Consider the following vector field:
A(r) = (3 x y2 z2 y2 , y3 z2 + x2 y, 3 x2 x2 z).
Is this field conservative? Is it solenoidal? Is it irrotational? Justify your answers. Calculate the flux of A out of a unit sphere
centered on the origin.
A.16. Find the gradients of the following scalar functions of the position vector r = (x, y, z):
(a) k r,

(b) |r|n ,

(c) |r k|n ,

(d) cos(k r).


Here, k is a fixed vector.
A.17. Find the divergences and curls of the following vector fields:
(a) k r,

(b) |r|n r,

(c) |r k|n (r k),

(d) a cos(k r).


Here, k and a are fixed vectors.
A.18. Calculate 2 when = f (|r|). Find f if 2 = 0.

Cartesian Tensors

253

B Cartesian Tensors

B.1

Introduction

As we saw in Appendix A, many physical entities can be represented mathematically as either scalars or vectors,
depending on their transformation properties under rotation of the coordinate axes. However, it turns out that scalars
and vectors are particular types of a more general class of mathematical constructs called tensors. In fact, a scalar is a
tensor of order zero, and a vector is a tensor of order one. In fluid mechanics, certain important physical entities (i.e.,
stress and rate of strain) are represented mathematically by tensors of order greater than one. It is therefore necessary
to supplement our investigation of fluid mechanics with a brief discussion of the mathematics of tensors. For the sake
of simplicity, we shall limit this discussion to Cartesian coordinate systems. Tensors expressed in such coordinate
systems are known as Cartesian tensors.

B.2

Tensors and Tensor Notation

Let the Cartesian coordinates x, y, z be written as the xi , where i runs from 1 to 3. In other words, x = x1 , y = x2 , and
z = x3 . Incidentally, in the following, any lowercase roman subscript (e.g., i, j, k) is assumed to run from 1 to 3. We
can also write the Cartesian components of a general vector v as the vi . In other words, v x = v1 , vy = v2 , and vz = v3 . By
contrast, a scalar is represented as a variable without a subscript: e.g., a, . Thus, a scalarwhich is a tensor of order
zerois represented as a variable with zero subscripts, and a vectorwhich is a tensor of order oneis represented as
a variable with one subscript. It stands to reason, therefore, that a tensor of order two is represented as a variable with
two subscripts: e.g., ai j , i j . Moreover, an nth-order tensor is represented as a variable with n subscripts: e.g., ai jk is a
third-order tensor, and bi jkl a fourth-order tensor. Note that a general nth-order tensor has 3n independent components.
Now, the components of a second-order tensor are conveniently visualized as a two-dimensional matrix, just as the
components of a vector are sometimes visualized as a one-dimensional matrix. However, it is important to recognize
that an nth-order tensor is not simply another name for an n-dimensional matrix. A matrix is just an ordered set of
numbers. A tensor, on the other hand, is an ordered set of components that have specific transformation properties
under rotation of the coordinate axes. (See Section B.3.)
Consider two vectors a and b that are represented as ai and bi , respectively, in tensor notation. According to
Section A.6, the scalar product of these two vectors takes the form
a b = a1 b1 + a2 b2 + a3 b3 .

(B.1)

The above expression can be written more compactly as


a b = ai bi .

(B.2)

Here, we have made use of the Einstein summation convention, according to which, in an expression containing lower
case roman subscripts, any subscript that appears twice (and only twice) in any term of the expression is assumed to
be summed from 1 to 3 (unless stated otherwise). Thus, ai bi = a1 b1 + a2 b2 + a3 b3 , and ai j b j = ai1 b1 + ai2 b2 + ai3 b3 .
Note that when an index is summed it becomes a dummy index, and can be written as any (unique) symbol: i.e.,
ai j b j and aip b p are equivalent. Moreover, only non-summed, or free, indices count toward the order of a tensor
expression. Thus, aii is a zeroth-order tensor (because there are no free indices), and ai j b j is a first-order tensor
(because there is only one free index). The process of reducing the order of a tensor expression by summing indices
is known as contraction. For example, aii is a zeroth-order contraction of the second-order tensor ai j . Incidentally,
when two tensors are multiplied together without contraction the resulting tensor is called an outer product: e.g., the
second-order tensor ai b j is the outer product of the two first-order tensors ai and bi . Likewise, when two tensors are
multiplied together in a manner that involves contraction then the resulting tensor is called an inner product: e.g.,
the first-order tensor ai j b j is an inner product of the second-order tensor ai j and the first-order tensor bi . Note, from
Equation (B.2), that the scalar product of two vectors is equivalent to the inner product of the corresponding first-order
tensors.

254

FLUID MECHANICS

According to Section A.8, the vector product of two vectors a and b takes the form
(a b)1

a2 b3 a3 b2 ,

(B.3)

(a b)2

a3 b1 a1 b3 ,

(B.4)

(a b)3

a1 b2 a2 b1

(B.5)

in tensor notation. The above expression can be written more compactly as


(a b)i = i jk a j bk .
Here,
i jk

+1

1
=

if i, j, k is an even permutation of 1, 2, 3
if i, j, k is an odd permutation of 1, 2, 3
otherwise

(B.6)

(B.7)

is known as the third-order permutation tensor (or, sometimes, the third-order Levi-Civita tensor). Note, in particular,
that i jk is zero if one of its indices is repeated: e.g., 113 = 212 = 0. Furthermore, it follows from (B.7) that
i jk = jki = ki j = k ji = jik = ik j .
It is helpful to define the second-order identity tensor (also known as the Kroenecker delta tensor),
(
1
if i = j
i j =
.
0
otherwise

(B.8)

(B.9)

It is easily seen that


i j

ji ,

(B.10)

ii

3,

(B.11)

ik k j

i j ,

(B.12)

i j a j

ai ,

(B.13)

i j ai b j

ai bi ,

(B.14)

i j aki b j

aki bi ,

(B.15)

etc.
The following is a particularly important tensor identity:
i jk ilm = jl km jm kl .

(B.16)

In order to establish the validity of the above expression, let us consider the various cases that arise. As is easily seen,
the right-hand side of (B.16) takes the values
+1

if j = l and k = m , j,

(B.17)

if j = m and k = l , j,

(B.18)

otherwise.

(B.19)

Moreover, in each product on the left-hand side, i has the same value in both factors. Thus, for a non-zero contribution, none of j, k, l, and m can have the same value as i (because each factor is zero if any of its indices are repeated).
Since a given subscript can only take one of three values (1, 2, or 3), the only possibilities that generate non-zero
contributions are j = l and k = m, or j = m and k = l, excluding j = k = l = m (since each factor would then have
repeated indices, and so be zero). Thus, the left-hand side reproduces (B.19), as well as the conditions on the indices

Cartesian Tensors

255

in (B.17) and (B.18). The left-hand side also reproduces the values in (B.17) and (B.18) since if j = l and k = m then
i jk = ilm and the product i jk ilm (no summation) is equal to +1, whereas if j = m and k = l then i jk = iml = ilm
and the product i jk ilm (no summation) is equal to 1. Here, use has been made of Equation (B.8). Hence, the validity
of the identity (B.16) has been established.
In order to illustrate the use of (B.16), consider the vector triple product identity (see Section A.11)
a (b c) = (a c) b (a b) c.

(B.20)

In tensor notation, the left-hand side of this identity is written


[a (b c)]i = i jk a j (klm bl cm ),
where use has been made of Equation (B.6). Employing Equations (B.8) and (B.16), this becomes


[a (b c)]i = ki j klm a j bl cm = il jm im jl a j bl cm ,

(B.21)

(B.22)

which, with the aid of Equations (B.2) and (B.13), reduces to

[a (b c)]i = a j c j bi a j b j ci = [(a c) b (a b) c]i .

(B.23)

Thus, we have established the validity of the vector identity (B.20). Moreover, our proof is much more rigorous than
that given earlier (in Section A.11).

B.3

Tensor Transformation

As we saw in Appendix A, scalars and vectors are defined according to their transformation properties under rotation
of the coordinate axes. In fact, a scalar is invarient under rotation of the coordinate axes. On the other hand, according
to Equations (A.49) and (B.6), the components of a general vector a transform under an infinitesimal rotation of the
coordinate axes according to
ai = ai + i jk j ak .
(B.24)
Here, the ai are the components of the vector in the original coordinate system, the ai are the components in the rotated
coordinate system, and the latter system is obtained from the former via a combination of an infinitesimal rotation
through an angle 1 about coordinate axis 1, an infinitesimal rotation through an angle 2 about axis 2, and an
infinitesimal rotation through an angle 3 about axis 3. These three rotations can take place in any order. Incidentally,
a finite rotation can be built up out of a great many infinitesimal rotations, so if a vector transforms properly under an
infinitesimal rotation of the coordinate axes then it will also transform properly under a finite rotation.
Equation (B.24) can also be written
ai = Ri j a j ,
(B.25)
where
Ri j = i j k ki j

(B.26)

is a rotation matrix (which is not a tensor, since it is specific to the two coordinate systems it transforms between). To
first-order in the i , Equation (B.25) can be inverted to give
ai = R ji aj .

(B.27)

This follows because, to first-order in the i ,


Rik R jk

(ik l lik ) ( jk m m jk ) = ik jk l jk lik m ik m jk

i j l li j l l ji = i j ,

(B.28)

where the dummy index m has been relabeled l, and use has been made of Equations (B.8), (B.10), and (B.12).
Likewise, it is easily demonstrated that
Rki Rk j = i j .
(B.29)

256

FLUID MECHANICS

It can also be shown that, to first-order in the i ,


i jk Rli Rm j Rnk = lmn .

(B.30)

This follows because


i jk Rli Rm j Rnk

=
=

i jk (li a ali ) (m j b bm j ) (nk c cnk )


i

h
i jk li m j nk a ali m j nk + am j li nk + ank li m j

lmn a (imn ial + inl iam + ilm ian )

lmn a (ma nl ml na + na lm nm la + la mn ln ma )

lmn .

(B.31)

Here, there has been much relabeling of dummy indices, and use has been made of Equations (B.10) and (B.16). It
can similarly be shown that
i jk Ril R jm Rkn = lmn .
(B.32)
As a direct generalization of Equation (B.25), a second-order tensor transforms under rotation as
ai j = Rik R jl akl ,

(B.33)

ai jk = Ril R jm Rkn almn .

(B.34)

whereas a third-order tensor transforms as


The generalization to higher-order tensors is straight-forward. For the case of a scalar, which is a zeroth-order tensor,
the transformation rule is particularly simple: i.e.,
a = a.
(B.35)
By analogy with Equation (B.27), the inverse transform is exemplified by
ai jk = Rli Rm j Rnk almn .

(B.36)

Incidentally, since all tensors of the same order transform in the same manner, it immediately follows that two tensors
of the same order whose components are equal in one particular Cartesian coordinate system will have their components equal in all coordinate systems that can be obtained from the original system via rotation of the coordinate axes.
In other words, if
ai j = bi j
(B.37)
in one particular Cartesian coordinate system then
ai j = bi j

(B.38)

in all Cartesian coordinate systems (with the same origin and system of units as the original system). Conversely,
it does not make sense to equate tensors of different order, since such an equation would only be valid in one particular coordinate system, and so could not have any physical significance (since the laws of physics are coordinate
independent).
It can easily be shown that the outer product of two tensors transforms as a tensor of the appropriate order. Thus,
if
ci jk = ai b jk ,
(B.39)
and
ai

Ri j a j ,

(B.40)

bi j

Rik R jl bkl ,

(B.41)

Cartesian Tensors

257

then
ci jk

ai bjk = Ril al R jm Rkn bmn = Ril R jm Rkn al bmn

Ril R jm Rkn clmn ,

(B.42)

which is the correct transformation rule for a third-order tensor.


The tensor transformation rule can be combined with the identity (B.29) to show that the scalar product of two
vectors transforms as a scalar. Thus,
ai bi = Ri j a j Rik bk = Ri j Rik a j bk = jk a j bk = a j b j = ai bi ,

(B.43)

where use has been made of Equation (B.14). Again, the above proof is more rigorous than that given previously (in
Section A.6). The proof also indicates that the inner product of two tensors transforms as a tensor of the appropriate
order.
The result that both the inner and outer products of two tensors transform as tensors of the appropriate order is
known as the product rule. Closely related to this rule is the so-called quotient rule, according to which if (say)
ci j = aik b jk ,

(B.44)

where b jk is an arbitrary tensor, and ci j transforms as a tensor under all rotations of the coordinate axes, then aik
which can be thought of as the quotient of ci j and b jk also transforms as a tensor. The proof is as follows:
aik bjk

= ci j = Ril R jm clm = Ril R jm alk bmk = Ril R jm alk R pm Rqk bpq

= Ril Rqk alk b jq = Ril Rkm alm bjk ,

(B.45)

where use has been made of the fact that ci j and bi j transform as tensors, as well as Equation (B.28). Rearranging, we
obtain
(B.46)
(aik Ril Rkm alm ) bjk = 0.

However, the bi j are arbitrary, so the above equation can only be satisfied, in general, if
aik = Ril Rkm alm ,

(B.47)

which is the correct transformation rule for a tensor. Incidentally, the quotient rule applies to any type of valid tensor
product.
The components of the second-order identity tensor, i j , have the special property that they are invariant under
rotation of the coordinate axes. This follows because
i j = Rik R jl kl = Rik R jk = i j ,

(B.48)

where use has been made of Equation (B.28). The components of the third-order permutation tensor, i jk , also have
this special property. This follows because
ijk = Ril R jm Rln lmn = i jk ,

(B.49)

where use has been made of Equation (B.31). The fact that i jk transforms as a proper third-order tensor immediately
implies, from the product rule, that the vector product of two vectors transforms as a proper vector: i.e., i jk a j bk is a
first-order tensor provided that ai and bi are both first-order tensors. This proof is much more rigorous that that given
earlier (in Section A.8).

B.4

Tensor Fields

We saw in Appendix A that a scalar field is a set of scalars associated with every point in space: e.g., (x), where
x = (x1 , x2 , x3 ) is a position vector. We also saw that a vector field is a set of vectors associated with every point in

258

FLUID MECHANICS

space: e.g., ai (x). It stands to reason, then, that a tensor field is a set of tensors associated with every point in space:
e.g., ai j (x). It immediately follows that a scalar field is a zeroth-order tensor field, and a vector field is a first-order
tensor field.
Most tensor field encountered in physics are smoothly varying and differentiable. Consider the first-order tensor
field ai (x). The various partial derivatives of the components of this field with respect to the Cartesian coordinates xi
are written
ai
.
(B.50)
x j
Moreover, this set of derivatives transform as the components of a second-order tensor. In order to demonstrate this,
we need the transformation rule for the xi , which is the same as that for a first-order tensor: i.e.,
xi = Ri j x j .

(B.51)

xi
= Ri j .
x j

(B.52)

xi
= R ji .
xj

(B.53)

ai
a xk (Ril al )
al
= i
=
R jk = Ril R jk
,

x j xk xj
xk
xk

(B.54)

Thus,

It is also easily shown that

Now,

which is the correct transformation rule for a second-order tensor. Here, use has been made of the chain rule, as well
as Equation (B.53). [Note, from Equation (B.26), that the Ri j are not functions of position.] It follows, from the above
argument, that differentiating a tensor field increases its order by one: e.g., ai j /xk is a third-order tensor. The only
exception to this rule occurs when differentiation and contraction are combined. Thus, ai j /x j is a first-order tensor,
since it only contains a single free index.
The gradient (see Section A.18) of a scalar field is an example of a first-order tensor field (i.e., a vector field):
()i =

.
xi

(B.55)

The divergence (see Section A.20) of a vector field is a contracted second-order tensor field that transforms as a scalar:
a=

ai
.
xi

(B.56)

Finally, the curl (see Section A.22) of a vector field is a contracted fifth-order tensor that transforms as a vector
( a)i = i jk

ak
.
x j

(B.57)

The above definitions can be used to prove a number of useful results. For instance,
( )i = i jk

!
2

= i jk
= 0,
x j xk
x j xk

(B.58)

which follows from symmetry because ik j = i jk whereas 2 /xk x j = 2 /x j xk . Likewise,


( a) =

ak
ak
= 0,
i jk
= i jk
xi
x j
xi x j

(B.59)

Cartesian Tensors

259

which again follows from symmetry. As a final example,


(a b)


a j
bk

bk + i jk a j
i jk a j bk = i jk
xi
xi
xi
ak
bk
= bi i jk
ai i jk
= b ( a) a ( b).
x j
x j
=

According to the divergence theorem (see Section A.20),


Z
I
ai
dV,
ai dS i =
x
i
V
S

(B.60)

(B.61)

where S is a closed surface surrounding the volume V. The above theorem is easily generalized to give, for example,
Z
I
ai j
ai j dS i =
dV,
(B.62)
V xi
S
or

ai j
dV,
x j

(B.63)

(B.64)

a
dV.
xi

or even

ai j dS j =

B.5

a dS i =

Isotropic Tensors

A tensor which has the special property that its components take the same value in all Cartesian coordinate systems is
called an isotropic tensor. We have already encountered two such tensors: namely, the second-order identity tensor, i j ,
and the third-order permutation tensor, i jk . Of course, all scalars are isotropic. Moreover, as is easily demonstrated,
there are no isotropic vectors (other than the null vector). It turns out that the most general isotropic Cartesian tensors
of second-, third-, and fourth-order are i j , i jk , and i j kl + ik jl + il jk , respectively, where , , , , and
are scalars. Let us prove these important results.1
The most general second-order isotropic tensor, ai j , is such that
ai j = Rip R jq a pq = ai j
for arbitrary rotations of the coordinate axes. It follows from Equation (B.24) that, to first-order in the i ,


m mis a s j + m js ais = 0.

(B.65)

(B.66)

However, the i are arbitrary, so we can write


mis a s j + m js ais = 0.

(B.67)

Let us multiply by mik . With the aid of Equation (B.16), we obtain


(ii ks is ki ) a s j + (i j ks is k j ) ais = 0,

(B.68)

2 ai j + a ji = a ss i j .

(B.69)

which reduces to
Interchanging the labels i and j, and then taking the difference between the two equations thus obtained, we deduce
that
ai j = a ji .
(B.70)
1 This

proof is adapted from P.G. Hodge, Jr., American Mathematical Monthly 68, 793 (1961).

260

FLUID MECHANICS

Hence,

a ss
i j ,
3

ai j =

(B.71)

which implies that


ai j = i j .

(B.72)

For the case of an isotropic third-order tensor, Equation (B.67) generalizes to


mis a s jk + m js aisk + mks ai js = 0.

(B.73)

Multiplying by mit , m jt , and mkt , and then setting t = i, t = j, and t = k, respectively, we obtain
2 ai jk + a jik + ak ji

a ssk i j + a s js ik ,

(B.74)

2 ai jk + a jik + aik j

a ssk i j + aiss jk ,

(B.75)

2 ai jk + ak ji + aik j

a s js ik + aiss jk ,

(B.76)

respectively. However, multiplying the above equations by jk , ik , and i j , and then setting i = i, j = i, and k = i,
respectively, we obtain
2 aiss + a sis + a ssi

a ssi + a sis,

(B.77)

2 a sis + aiss + a ssi

a ssi + aiss,

(B.78)

2 a ssi + aiss + a sis

a sis + aiss,

(B.79)

respectively, which implies that


aiss = a sis = a ssi = 0.

(B.80)

Hence, we deduce that


2 ai jk + a jik + ak ji

0,

(B.81)

2 ai jk + a jik + aik j

0,

(B.82)

2 ai jk + ak ji + aik j

0.

(B.83)

aik j = a jik = ak ji = ai jk .

(B.84)

ai jk = i jk .

(B.85)

The solution to the above equation must satisfy

This implies, from Equation (B.8), that


For the case of an isotropic fourth-order tensor, Equation (B.73) generalizes to
mis a s jkl + m js aiskl + mks ai jsl + mls ai jks = 0.

(B.86)

Multiplying the above by mit , m jt , mkt , mlt , and then setting t = i, t = j, t = k, and t = l, respectively, we obtain
2 ai jkl + a jikl + ak jil + al jki

a sskl i j + a s jsl ik + a s jks il ,

(B.87)

2 ai jkl + a jikl + aik jl + ail jk

a sskl i j + aisks jl + aissl jk ,

(B.88)

2 ai jkl + ak jil + aik jl + ai jlk

ai jss kl + a s jsl ik + aissl jk ,

(B.89)

2 ai jkl + al jki + ailk j + ai jlk

ai jss kl + aisks jl + a s jks il ,

(B.90)

respectively. Now, if ai jkl is an isotropic fourth-order tensor then a sskl is clearly an isotropic second-order tensor, which
means that is a multiple of kl . This, and similar arguments, allows us to deduce that
a sskl

kl ,

(B.91)

a s jsl

jl ,

(B.92)

a s jks

jk .

(B.93)

Cartesian Tensors

261

Let us assume, for the moment, that


ai jss

a ssi j ,

(B.94)

aisks

a sisk ,

(B.95)

aissl

a sils .

(B.96)

Thus, we get
2 ai jkl + a jikl + ak jil + al jki

i j kl + ik jl + il jk ,

(B.97)

2 ai jkl + a jikl + aik jl + ailk j

i j kl + ik jl + il jk ,

(B.98)

2 ai jkl + ak jil + aik jl + ai jlk

i j kl + ik jl + il jk ,

(B.99)

2 ai jkl + al jki + ailk j + ai jlk

i j kl + ik jl + il jk .

(B.100)

Relations of the form


ai jkl = a jilk = akli j = alk ji

(B.101)

can be obtained by subtracting the sum of one pair of Equations (B.97)(B.100) from the sum of the other pair. These
relations justify Equations (B.94)(B.96). Equations (B.97) and (B.101) can be combined to give
2 ai jkl + (ai jlk + aik jl + ailk j )

= i j kl + ik jl + il jk ,

(B.102)

2 aikl j + (aik jl + ailk j + ai jlk )

= ik jl + il jk + i j kl ,

(B.103)

2 ail jk + (ailk j + ai jlk + aik jl )

= il jk + i j kl + ik jl .

(B.104)

The latter two equations are obtained from the first via cyclic permutation of j, k, and l, with i remaining unchanged.
Summing Equations (B.102)(B.104), we get
2 (ai jkl + aikl j + ail jk ) + 3 (ai jlk + aik jl + ailk j ) = ( + + ) (i j kl + ik jl + il jk ).

(B.105)

It follows from symmetry that


ai jkl + aikl j + ail jk = ai jlk + aik jl + ailk j =

1
( + + ) (i j kl + ik jl + il jk ).
5

(B.106)

This can be seen by swapping the indices k and l in the above expression. Finally, substitution into Equation (B.102)
yields
ai jkl = i j kl + ik jl + il jk ,
(B.107)
where

B.6

(4 )/10,

(B.108)

(4 )/10,

(B.109)

(4 )/10.

(B.110)

Exercises

B.1. Show that a general second-order tensor ai j can be decomposed into three tensors
ai j = ui j + vi j + si j ,
where ui j is symmetric (i.e., u ji = ui j ) and traceless (i.e., uii = 0), vi j is isotropic, and si j only has three independent
components.
B.2. Use tensor methods to establish the following vector identities:

262

FLUID MECHANICS
(a) a b c = a b c = b a c.
(b) (a b) c = (a c) b (c b) a.
(c) (a b) (c d) = (a c) (b d) (a d) (b c).
(d) (a b) (c d) = (a b d) c (a b c) d.
(e) ( a) = a + a .
(f) ( a) = a + a.
(g) (a b) = (b ) a (a ) b + ( b) a ( a) b.
(h) (a a) = 2 a ( a) + 2 (a ) a.
(i) ( a) = ( a) 2 a.

Here, [(b )a]i = b j ai /x j , and (2 a)i = 2 ai .


B.3. A quadric surface has an equation of the form
a x12 + b x22 + c x32 + 2 f x1 x2 + 2 g x1 x3 + 2 h x2 x3 = 1.
Show that the coefficients in the above expression transform under rotation of the coordinate axes like the components of a
symmetric second-order tensor. Hence, demonstrate that the equation for the surface can be written in the form
xi T i j x j = 1,
where the T i j are the components of the aforementioned tensor.
B.4. The determinant of a second-order tensor Ai j is defined
det(A) = i jk Ai1 A j2 Ak3 .
(a) Show that
det(A) = i jk A1i A2 j A3k
is an alternative, and entirely equivalent, definition.
(b) Demonstrate that det(A) is invariant under rotation of the coordinate axes.
(c) Suppose that Ci j = Aik Bk j . Show that
det(C) = det(A) det(B).
B.5. If
Ai j x j = xi
then and x j are said to be eigenvalues and eigenvectors of the second-order tensor Ai j , respectively. The eigenvalues of Ai j
are calculated by solving the related homogeneous matrix equation
(Ai j i j ) x j = 0.
Now, it is a standard result in linear algebra that an equation of the above form only has a non-trivial solution when
det(Ai j i j ) = 0.
Demonstrate that the eigenvalues of Ai j satisfy the cubic polynomial
3 tr(A) 2 + (A) det(A) = 0,
where tr(A) = Aii and (A) = (Aii A j j Ai j A ji )/2. Hence, deduce that Ai j possesses three eigenvalues1 , 2 , and 3 (say).
Moreover, show that
tr(A)

1 + 2 + 3 ,

det(A)

1 2 3 .

B.6. Suppose that Ai j is a (real) symmetric second-order tensor: i.e., A ji = Ai j .


(a) Demonstrate that the eigenvalues of Ai j are all real, and that the eigenvectors can be chosen to be real.

Cartesian Tensors

263

(b) Show that eigenvectors of Ai j corresponding to different eigenvalues are orthogonal to one another. Hence, deduce
that the three eigenvectors of Ai j are, or can be chosen to be, mutually orthogonal.
(c) Demonstrate that Ai j takes the diagonal form Ai j = i i j (no sum) in a Cartesian coordinate system in which the
coordinate axes are each parallel to one of the eigenvectors.
B.7. In an isotropic elastic medium under stress the displacement ui satisfies
2 ui
,
t2

i j
x j

i j

ci jkl

!
1 uk ul
,
+
2 xl xk

where i j is the stress tensor (note that ji = i j ), the mass density (which is a uniform constant), and
ci jkl = K i j kl + [ik jl + il jk (2/3) i j kl ].
the isotropic stiffness tensor. Here, K and are the bulk modulus and shear modulus of the medium, respectively. Show
that the divergence and the curl of u both satisfy wave equations. Furthermore, demonstrate that the characteristic wave
velocities of the divergence and curl waves are [(K + 4/3)/]1/2 and (/)1/2 , respectively.

264

FLUID MECHANICS

Non-Cartesian Coordinates

265

C Non-Cartesian Coordinates

C.1 Introduction
In fluid mechanics non-Cartesian coordinates are often used to exploit the symmetry of particular fluid systems. For
example, it is convenient to employ cylindrical coordinates to describe systems possessing axial symmetry. In this
Appendix, we investigate a particularly useful class of non-Cartesian coordinates known as orthogonal curvilinear
coordinates. The two most commonly occurring examples of this class in fluid mechanics are cylindrical and spherical
coordinates. (Note, incidentally, that the Einstein summation convention is not used in this Appendix.)

C.2 Orthogonal Curvilinear Coordinates


Let x1 , x2 , x3 be a set of standard right-handed Cartesian coordinates. Furthermore, let u1 (x1 , x2 , x3 ), u2 (x1 , x2 , x3 ),
u3 (x1 , x2 , x3 ) be three independent functions of these coordinates which are such that each unique triplet of x1 , x2 , x3
values is associated with a unique triplet of u1 , u2 , u3 values. It follows that u1 , u2 , u3 can be used as an alternative set
of coordinates to distinguish different points in space. Since the surfaces of constant u1 , u2 , and u3 are not generally
parallel planes, but rather curved surfaces, this type of coordinate system is termed curvilinear.
Let h1 = |u1 |1 , h2 = |u2 |1 , and h3 = |u3 |1 . It follows that e1 = h1 u1 , e2 = h2 u2 , and e3 = h3 u3 are a set
of unit basis vectors which are normal to surfaces of constant u1 , u2 , and u3 , respectively, at all points in space. Note,
however, that the direction of these basis vectors is generally a function of position. Suppose that the ei , where i runs
from 1 to 3, are mutually orthogonal at all points in space: i.e.,
ei e j = i j .

(C.1)

In this case, u1 , u2 , u3 are said to constitute an orthogonal coordinate system. Suppose, further, that
e1 e2 e3 = 1

(C.2)

at all points in space, so that u1 , u2 , u3 also constitute a right-handed coordinate system. It follows that
ei e j ek = i jk .
Finally, a general vector A, associated with a particular point in space, can be written
X
A=
A i ei ,

(C.3)

(C.4)

where the ei are the local basis vectors of the u1 , u2 , u3 system, and Ai = ei A is termed the ith component of A in this
system.
Consider two neighboring points in space whose coordinates in the u1 , u2 , u3 system are u1 , u2 , u3 and u1 + du1 ,
u2 + du2 , u3 + du3 . It is easily shown that the vector directed from the first to the second of these points takes the form
dx =

X
du1
du2
du3
e1 +
e2 +
e3 =
hi dui ei .
|u1 |
|u2 |
|u3 |
i

Hence, from (C.1), an element of length (squared) in the u1 , u2 , u3 coordinate system is written
X
dx dx =
hi2 dui2 .

(C.5)

(C.6)

Here, the hi , which are generally functions of position, are known as the scale factors of the system. Elements of area
that are normal to e1 , e2 , and e3 , at a given point in space, take the form dS 1 = h2 h3 du2 du3 , dS 2 = h1 h3 du1 du3 ,

266

FLUID MECHANICS

and dS 3 = h1 h2 du1 du2 , respectively. Finally, an element of volume, at a given point in space, is written dV =
h du1 du2 du3 , where
h = h1 h2 h3 .
(C.7)
Note that [see Equation (A.176)]
and

ui = 0,

(C.8)

hi
ui = 0.
h

(C.9)

The latter result follows from Equations (A.175) and (A.176) because (h12 /h) u1 = u2 u3 , etc. Finally, it is easily
demonstrated from (C.1) and (C.3) that
= hi2 i j ,

ui u j
ui u j uk

= h

(C.10)

i jk .

(C.11)

Consider a scalar field (u1 , u2 , u3 ). It follows from the chain rule, and the relation ei = hi ui , that
=

X 1
X
ui =
ei .
ui
hi ui
i
i

(C.12)

Hence, the components of in the u1 , u2 , u3 coordinate system are


()i =

1
.
hi ui

(C.13)

Consider a vector field A(u1 , u2 , u3 ). We can write


A

hi2
h

ui
(Ai ei ) =
(hi Ai ui ) =
Ai
h
h
i
i
i
i
!
!
2
Xh
X1 h
h
i
Ai =
Ai ,
ui
h
hi
h ui hi
i
i

(C.14)

where use has been made of Equations (A.174), (C.9), and (C.10). Thus, the divergence of A in the u1 , u2 , u3 coordinate
system takes the form
X1 h !
A=
(C.15)
Ai .
h ui hi
i
We can write
A

X
k

(Ak ek ) =

X (hk Ak )
j,k

u j

X
k

(hk Ak uk ) =

X
k

(hk uk ) uk

u j uk ,

(C.16)

where use has been made of Equations (A.178), (C.8), and (C.12). It follows from (C.11) that
( A)i = ei A =

X
j,k

hi

X
hi (hk Ak )
(hk Ak )
ui u j uk =
i jk
.
u j
h u j
j,k

(C.17)

Hence, the components of A in the u1 , u2 , u3 coordinate system are


( A)i =

X
j,k

i jk

hi (hk Ak )
.
h u j

(C.18)

Non-Cartesian Coordinates

267

Now, 2 = [see (A.172)], so Equations (C.12) and (C.15) yield the following expression for 2 in the u1 ,
u2 , u3 coordinate system:
X 1 h
2

.
=
(C.19)
h ui hi2 ui
i
The vector identities (A.171) and (A.179) can be combined to give the following expression for (A )B that is
valid in a general coordinate system:
(A )B =

1
[(A B) (A B) ( A) B + ( B) A
2
A ( B) B ( A)] .

Making use of Equations (C.13), (C.15), and (C.18), as well as the easily demonstrated results
X
AB =
Ai Bi ,

(C.20)

(C.21)

AB =

i jk A j Bk ,

(C.22)

j,k

and the tensor identity (B.16), Equation (C.20) reduces (after a great deal of tedious algebra) to the following expression for the components of (A )B in the u1 , u2 , u3 coordinate system:
X A j Bi A j B j h j Ai B j hi !
[(A ) B]i =
.
(C.23)

+
h j u j
hi h j ui
hi h j u j
j
Note, incidentally, that the commonly quoted result [(A )B]i = A Bi is only valid in Cartesian coordinate systems
(for which h1 = h2 = h3 = 1).
Let us define the gradient A of a vector field A as the tensor whose components in a Cartesian coordinate system
take the form
Ai
.
(C.24)
(A)i j =
x j
In an orthogonal curvilinear coordinate system, the above expression generalizes to
(A)i j = [(e j ) A]i .

(C.25)

It thus follows from (C.23), and the relation (ei ) j = ei e j = i j , that


(A)i j =

X Ak hi
A j h j
1 Ai

+ i j
.
h j u j hi h j ui
hi hk uk
k

(C.26)

The vector identity (A.177) yields the following expression for 2 A that is valid in a general coordinate system:
2 A = ( A) ( A).

(C.27)

Making use of Equations (C.15), (C.18), and (C.19), as well as (C.21) and (C.22), and the tensor identity (B.16), the
above equation reduces (after a great deal of tedious algebra) to the following expression for the components of 2 A
in the u1 , u2 , u3 coordinate system:
!
X( 2
1 h j
1 hi
(2 A)i = 2 Ai +

Aj
hi h j hi u j ui h j ui u j
j

2

h A j h j hi2 Ai hi h j


2
hi h j2 hi2 ui u j h
h j u j u j h

2
2
2
2
A j h 1 h j h j h h j h j
2 h j

+



hi h j3 h j ui u j h h j2 ui h u j h ui u j h

!2

2 hi
Ai 2 hi

(C.28)

.
2 h u
2

hi h
u
i
j
j

268

FLUID MECHANICS

e
r

er
z

Figure C.1: Cylindrical coordinates.


Note, again, that the commonly quoted result (2 A)i = 2 Ai is only valid in Cartesian coordinate systems (for which
h1 = h2 = h3 = h = 1).

C.3 Cylindrical Coordinates


p
In the cylindrical coordinate system, u1 = r, u2 = , and u3 = z, where r = x2 + y2 , = tan1 (y/x), and x, y, z
are standard Cartesian coordinates. Thus, r is the perpendicular distance from the z-axis, and the angle subtended
between the projection of the radius vector (i.e., the vector connecting the origin to a general point in space) onto the
x-y plane and the x-axis. See Figure C.1.
A general vector A is written
A = A r er + A e + A z ez ,
(C.29)
where er = r/|r|, e = /||, and ez = z/|z|. See Figure C.1. Of course, the unit basis vectors er , e , and ez are
mutually orthogonal, so Ar = A er , etc.
As is easily demonstrated, an element of length (squared) in the cylindrical coordinate system takes the form
dx dx = dr 2 + r2 d 2 + dz2 .

(C.30)

Hence, comparison with Equation (C.6) reveals that the scale factors for this system are
hr

= 1,

(C.31)

= r,

(C.32)

hz

= 1.

(C.33)

Thus, surface elements normal to er , e , and ez are written


dS r

r d dz,

(C.34)

dS

dr dz,

(C.35)

dS z

r dr d,

(C.36)

dV = r dr d dz.

(C.37)

respectively, whereas a volume element takes the form

Non-Cartesian Coordinates

269

According to Equations (C.13), (C.15), and (C.18), gradient, divergence, and curl in the cylindrical coordinate
system are written
=
A

er +
e +
ez ,
r
r
z
1
1 A Az
(r Ar ) +
+
,
r r
r
z
!
!
!
Ar Az
1
1 Ar
1 Az A

(r A )
er +
e +
ez ,
r
z
z
r
r r
r

(C.38)
(C.39)

(C.40)
respectively. Here, (r) is a general scalar field, and A(r) a general vector field.
According to Equation (C.19), when expressed in cylindrical coordinates, the Laplacian of a scalar field becomes
!
1

1 2 2
2
=
r
+ 2
+ 2.
(C.41)
r r r
r 2
z
Moreover, from Equation (C.23), the components of (A )A in the cylindrical coordinate system are
[(A )A]r

[(A )A]

[(A )A]z

A2
,
r
Ar A
,
A A +
r
A Az .

A Ar

(C.42)
(C.43)
(C.44)

Let us define the symmetric gradient tensor


h
i
g = 1 A + (A)T .
(C.45)
A
2
Here, the superscript T denotes a transpose. Thus, if the i j element of some second-order tensor S is S i j then the
g in the cylindrical coordinate
corresponding element of ST is S ji . According to Equation (C.26), the components of A
system are
g rr
(A)

g
(A)

g zz
(A)

g r = (A)
g r
(A)
g rz = (A)
g zr
(A)

g z = (A)
g z
(A)

=
=
=
=
=
=

Ar
,
r
1 A Ar
+ ,
r
r
Az
,
z
!
1 1 Ar A A
,
+

2 r
r
r
!
1 Ar Az
,
+
2 z
r
!
1 A 1 Az
.
+
2 z
r

(C.46)
(C.47)
(C.48)
(C.49)
(C.50)
(C.51)

Finally, from Equation (C.28), the components of 2 A in the cylindrical coordinate system are
(2 A)r

(2 A)

(2 A)z

Ar
2 A
,

r2 r2
2 Ar A
2,
2 A + 2
r
r

2 Ar

2 Az .

(C.52)
(C.53)
(C.54)

270

FLUID MECHANICS

z
r

Figure C.2: Spherical coordinates.

C.4 Spherical Coordinates


p
In the spherical coordinate system, u1 = r, u2 = , and u3 = , where r = x2 + y2 + z2 , = cos1 (z/r), =
tan1 (y/x), and x, y, z are standard Cartesian coordinates. Thus, r is the length of the radius vector, the angle
subtended between the radius vector and the z-axis, and the angle subtended between the projection of the radius
vector onto the x-y plane and the x-axis. See Figure C.2.
A general vector A is written
A = A r er + A e + A e ,
(C.55)
where er = r/|r|, e = /||, and e = /||. See Figure C.2. Of course, the unit vectors er , e , and e are
mutually orthogonal, so Ar = A er , etc.
As is easily demonstrated, an element of length (squared) in the spherical coordinate system takes the form
dx dx = dr 2 + r2 d 2 + r2 sin2 d2 .

(C.56)

Hence, comparison with Equation (C.6) reveals that the scale factors for this system are
hr

1,

(C.57)

r,

(C.58)

r sin .

(C.59)

Thus, surface elements normal to er , e , and e are written


dS r

r2 sin d d,

(C.60)

dS

r sin dr d,

(C.61)

dS

r dr d,

(C.62)

respectively, whereas a volume element takes the form


dV = r2 sin dr d d.

(C.63)

Non-Cartesian Coordinates

271

According to Equations (C.13), (C.15), and (C.18), gradient, divergence, and curl in the spherical coordinate
system are written
=
A =
A =

1
1

er +
e +
e ,
r
r
r sin

1 2
1
1 A
(r Ar ) +
(sin A ) +
,
r sin
r sin
r2 r
#
"

1 A
1
er
(sin A )
r sin
r sin
"
#
1 Ar 1
+

(r A ) e
r sin
r r
#
"
1 Ar
1
e ,
(r A )
+
r r
r

(C.64)
(C.65)

(C.66)

respectively. Here, (r) is a general scalar field, and A(r) a general vector field.
According to Equation (C.19), when expressed in spherical coordinates, the Laplacian of a scalar field becomes
!
!
1 2
1

1
2

2 = 2
r
+ 2
sin
+
.
(C.67)
2
r

r r
r sin
r2 sin 2
Moreover, from Equation (C.23), the components of (A )A in the spherical coordinate system are
[(A )A]r

A Ar

[(A )A]

A A +

[(A )A]

A2 + A2

,
r
Ar A cot A2

,
r
Ar A + cot A A
A A +
.
r

g in the spherical coordinate system are


Now, according to Equation (C.26), the components of A
g rr
(A)

g
(A)

g r = (A)
g r
(A)

g
(A)

g r = (A)
g r
(A)

g = (A)
g
(A)

=
=

Ar
,
r
1 A Ar
+ ,
r
r
A
1
Ar cot A

+
+
,
r sin
r
r
!
1 1 Ar A A
,
+

2 r
r
r
!
1
1 Ar A A
,
+

2 r sin
r
r

!
1 A 1 A cot A
1
.
+

2 r sin
r
r

(C.68)
(C.69)
(C.70)

(C.71)
(C.72)
(C.73)
(C.74)
(C.75)
(C.76)

Finally, from Equation (C.28), the components of 2 A in the spherical coordinate system are
(2 A)r

= 2 Ar

A
2Ar
2 A 2 cot A
2

,
2
2
2
2
r
r
r
r sin

(C.77)

(2 A)

= 2 A +

A
2 Ar
2
A
2

,
2
2
r
r2 sin r sin

(C.78)

272

FLUID MECHANICS
(2 A)

= 2 A

A
r2

sin

Ar
2 cot A
+ 2
.
r sin
sin
2

r2

(C.79)

C.5 Exercises
C.1. Find the Cartesian components of the basis vectors er , e , and ez of the cylindrical coordinate system. Verify that the vectors
are mutually orthogonal. Do the same for the basis vectors er , e , and e of the spherical coordinate system.
C.2. Use cylindrical coordinates to prove that the volume of a right cylinder of radius a and length l is a2 l. Demonstrate that
the moment of inertia of a uniform cylinder of mass M and radius a about its symmetry axis is (1/2) M a2 .
C.3. Use spherical coordinates to prove that the volume of a sphere of radius a is (4/3) a3 . Demonstrate that the moment of
inertia of a uniform sphere of mass M and radius a about an axis passing through its center is (2/5) M a2 .
C.4. For what value(s) of n is (rn er ) = 0, where r is a spherical coordinate?

C.5. For what value(s) of n is (rn er ) = 0, where r is a spherical coordinate?


C.6.

(a) Find a vector field F = Fr (r) er satisfying F = rm for m 0. Here, r is a spherical coordinate.

(b) Use the divergence theorem to show that


Z

rm dV =
V

1
m+3

rm+1 er dS,

where V is a volume enclosed by a surface S .


(c) Use the above result (for m = 0) to demonstrate that the volume of a right cone is one third the volume of the right
cylinder having the same base and height.
C.7. The electric field generated by a z-directed electric dipole of moment p, located at the origin, is
"
#
1 3 (er p) er p
,
E(r) =
40
r3
where p = p ez , and r is a spherical coordinate. Find the components of E(r) in the spherical coordinate system. Calculate
E and E.

C.8. Show that the parabolic cylindrical coordinates u, v, z, defined by the equations x = (u2 v2 )/2, y = u v, z = z, where x, y, z
are Cartesian coordinates, are orthogonal. Find the scale factors hu , hv , hz . What shapes are the u = constant and v = constant
surfaces? Write an expression for 2 f in parabolic cylindrical coordinates.
C.9. Show that the elliptic cylindrical coordinates , , z, defined by the equations x = cosh cos , y = sinh sin , z = z, where
x, y, z are Cartesian coordinates, and 0 , < , are orthogonal. Find the scale factors h , h , hz . What shapes
are the = constant and = constant surfaces? Write an expression for f in elliptical cylindrical coordinates.

Calculus of Variations

273

D Calculus of Variations

D.1 Euler-Lagrange Equation


It is a well-known fact, first enunciated by Archimedes, that the shortest distance between two points in a plane is
a straight-line. However, suppose that we wish to demonstrate this result from first principles. Let us consider the
length, l, of various curves, y(x), which run between two fixed points, A and B, in a plane, as illustrated in Figure D.1.
Now, l takes the form
Z B
Z b
2
2 1/2
l=
[dx + dy ] =
[1 + y 2 (x)]1/2 dx,
(D.1)
A

where y dy/dx. Note that l is a function of the function y(x). In mathematics, a function of a function is termed a
functional.

Figure D.1: Different paths between points A and B.


Now, in order to find the shortest path between points A and B, we need to minimize the functional l with respect
to small variations in the function y(x), subject to the constraint that the end points, A and B, remain fixed. In other
words, we need to solve
l = 0.
(D.2)
The meaning of the above equation is that if y(x) y(x) + y(x), where y(x) is small, then the first-order variation
in l, denoted l, vanishes. In other words, l l + O(y 2 ). The particular function y(x) for which l = 0 obviously
yields an extremum of l (i.e., either a maximum or a minimum). Hopefully, in the case under consideration, it yields a
minimum of l.
Consider a general functional of the form
I=

F(y, y, x) dx,

(D.3)

where the end points of the integration are fixed. Suppose that y(x) y(x) + y(x). The first-order variation in I is
written
!
Z b
F
F
I =
(D.4)
y + y dx,
y
y
a

274

FLUID MECHANICS

where y = d(y)/dx. Setting I to zero, we obtain


!
Z b
F
F
y + y dx = 0.
y
y
a
This equation must be satisfied for all possible small perturbations y(x).
Integrating the second term in the integrand of the above equation by parts, we get
!#
"
#b
Z b"
F
d F
F
y dx +
y = 0.

y dx y
y
a
a

(D.5)

(D.6)

Now, if the end points are fixed then y = 0 at x = a and x = b. Hence, the last term on the left-hand side of the above
equation is zero. Thus, we obtain
!#
Z b"
F
d F
y dx = 0.
(D.7)

y dx y
a
The above equation must be satisfied for all small perturbations y(x). The only way in which this is possible is for the
expression enclosed in square brackets in the integral to be zero. Hence, the functional I attains an extremum value
whenever
!
F
d F

= 0.
(D.8)
dx y
y
This condition is known as the Euler-Lagrange equation.
Let us consider some special cases. Suppose that F does not explicitly depend on y. It follows that F/y = 0.
Hence, the Euler-Lagrange equation (D.8) simplifies to
F
= const.
y
Next, suppose that F does not depend explicitly on x. Multiplying Equation (D.8) by y , we obtain
!
F
d F
y
= 0.
y

dx y
y
However,

!
!
F
d F
d F
y
=y
+ y .

dx
y
dx y
y

(D.9)

(D.10)

(D.11)

Thus, we get

!
d F
F
F
y
= y
(D.12)
+ y .
dx
y
y
y
Now, if F is not an explicit function of x then the right-hand side of the above equation is the total derivative of F,
namely dF/dx. Hence, we obtain
!
d F
dF
y
=
,
(D.13)
dx
y
dx
which yields
F
(D.14)
y F = const.
y
p
Returning to the case under consideration, we have F = 1 + y 2 , according to Equation (D.1) and (D.3). Hence,
F is not an explicit function of y, so Equation (D.9) yields

where c is a constant. So,

F
y
= c,
= p

y
1 + y 2

(D.15)

c
= const.
(D.16)
y =
1 c2
Of course, y = constant is the equation of a straight-line. Thus, the shortest distance between two fixed points in a
plane is indeed a straight-line.

Calculus of Variations

275

D.2 Conditional Variation


Suppose that we wish to find the function y(x) which maximizes or minimizes the functional
Z

I=
subject to the constraint that the value of
J=

F(y, y, x) dx,

(D.17)

G(y, y , x) dx

(D.18)

remains constant. We can achieve our goal by finding an extremum of the new functional K = I + J, where (x) is
an undetermined function. We know that J = 0, since the value of J is fixed, so if K = 0 then I = 0 as well. In
other words, finding an extremum of K is equivalent to finding an extremum of I. Application of the Euler-Lagrange
equation yields
#
!
!
"
d F
F
[ G]
d [ G]
= 0.
(D.19)

+
dx y
y
dx y
y
In principle, the above equation, together with the constraint (D.18), yields the functions (x) and y(x). Incidentally,
is generally termed a Lagrange multiplier. If F and G have no explicit x-dependence then is usually a constant.
As an example, consider the following famous problem. Suppose that a uniform chain of fixed length l is suspended
by its ends from two equal-height fixed points which are a distance a apart, where a < l. What is the equilibrium
configuration of the chain?
Suppose that the chain has the uniform density per unit length . Let the x- and y-axes be horizontal and vertical,
respectively, and let the two ends of the chain lie at (a/2, 0). The equilibrium configuration of the chain is specified
by the function y(x), for a/2 x +a/2, where y(x) is the vertical distance of the chain below its end points at
horizontal position x. Of course, y(a/2) = y(+a/2) = 0.
According to standard Newtonian dynamics, the stable equilibrium state of a conservative dynamical system is one
which minimizes the systems potential energy. Now, the potential energy of the chain is written
U = g

y ds = g

a/2

y [1 + y 2 ]1/2 dx,

(D.20)

a/2

p
where ds = dx2 + dy2 is an element of length along the chain, and g is the acceleration due to gravity. Hence,
we need to minimize U with respect to small variations in y(x). However, the variations in y(x) must be such as to
conserve the fixed length of the chain. Hence, our minimization procedure is subject to the constraint that
l=

ds =

a/2

[1 + y 2 ]1/2 dx

(D.21)

a/2

remains constant.
It follows, from the above discussion, that we need to minimize the functional
Z a/2
( g y + ) [1 + y 2 ]1/2 dx,
K = U +l =

(D.22)

a/2

where is an, as yet, undetermined constant. Since the integrand in the functional does not depend explicitly on x, we
have from Equation (D.14) that
y 2 ( g y + ) [1 + y 2 ]1/2 ( g y + ) [1 + y 2 ]1/2 = k,

(D.23)

where k is a constant. This expression reduces to



y 2
1,
y 2 = +
h

(D.24)

276

FLUID MECHANICS

where = /k, and h = k/ g.


Let
+

y
= cosh z.
h

(D.25)

Making this substitution, Equation (D.24) yields


dz
= h1 .
dx

(D.26)

Hence,

x
z = + c,
h
where c is a constant. It follows from Equation (D.25) that
y(x) = h [ + cosh(x/h + c)].

(D.27)

(D.28)

The above solution contains three undetermined constants, h, , and c. We can eliminate two of these constants by
application of the boundary conditions y(a/2) = 0. This yields
+ cosh(a/2 h + c) = 0.

(D.29)

Hence, c = 0, and = cosh(a/2 h). It follows that


y(x) = h [cosh(a/2 h) cosh(x/h)].
The final unknown constant, h, is determined via the application of the constraint (D.21). Thus,
Z a/2
Z a/2
l=
[1 + y 2 ]1/2 dx =
cosh(x/h) dx = 2 h sinh(a/2 h).
a/2

(D.30)

(D.31)

a/2

Hence, the equilibrium configuration of the chain is given by the curve (D.30), which is known as a catenary (from
the Latin for chain), where the parameter h satisfies
 a 
l
.
(D.32)
= sinh
2h
2h

D.3 Multi-Function Variation


Suppose that we wish to maximize or minimize the functional
Z b
F(y1 , y2 , , yF , y1 , y2 , , yF , x) dx.
I=

(D.33)

Here, the integrand F is now a functional of the F independent functions yi (x), for i = 1, F . A fairly straightforward
extension of the analysis in Section D.1 yields F separate Euler-Lagrange equations,
!
d F
F

= 0,
(D.34)
dx yi
yi
for i = 1, F , which determine the F functions yi (x). If F does not explicitly depend on the function yk then the kth
Euler-Lagrange equation simplifies to
F
= const.
(D.35)
yk
Likewise, if F does not explicitly depend on x then all F Euler-Lagrange equations simplify to
yi
for i = 1, F .

F
F = const,
yi

(D.36)

Calculus of Variations

277

D.4 Exercises
D.1. Find the extremal curves y = y(x) of the following constrained optimization problems, using the method of Lagrange
multipliers:
i
R1h
R1
(a) 0 y 2 + x2 dx, such that 0 y2 dx = 2.
R
R
(b) 0 y 2 dx, such that y(0) = y() = 0, and 0 y2 dx = 2.
R1
R p
1 + y 2 dx = 2/3.
(c) 0 y dx, such that y(0) = y(1) = 1, and

D.2. Suppose P and Q are two points lying in the x-y plane, which is orientated vertically such that P is above Q. Imagine there is
a thin, flexible wire connecting the two points and lying entirely in the x-y plane. A frictionless bead travels down the wire,
impelled by gravity alone. Show that the shape of the wire that results in the bead reaching the point Q in the least amount
of time is a cycloid, which takes the parametric form
x()

k ( sin ) ,

y()

k (1 cos ) ,

where k is a constant.
D.3. Find the curve y(x), in the interval 0 x p, which is of length and maximizes
Z p
y dx.
0

278

FLUID MECHANICS

Ellipsoidal Potential Theory

279

E Ellipsoidal Potential Theory

Let us adopt the right-handed Cartesian coordinate system x1 , x2 , x3 . Consider a homogeneous ellipsoidal body whose
outer boundary satisfies
x12 x22 x32
+
+
= 1.
(E.1)
a12 a22 a32
Let us calculate the gravitational potential (i.e., the potential energy of a unit test mass) at some point P (x1 , x2 , x3 )
lying within this body.
Consider the contribution to the potential at P from the mass contained within a double cone, whose apex is P,
and which is terminated in both directions at the bodys outer boundary. See Figure E.1. If the cone subtends a solid
angle d then a volume element is written dV = r2 dr d, where r measures displacement from P along the axis of
the cone. Thus, from standard classical gravitational theory, the contribution to the potential takes the form
d =

G
dV
r

G
dV,
(r)

(E.2)

where r = |PQ|, r = |PR|, and is the constant mass density of the ellipsoid. Hence, we obtain
d = G

r dr +

!
1
r dr d = G (r 2 + r 2 ) d.
2

(E.3)

The net potential at P is obtained by integrating over all solid angle, and dividing the result by two to adjust for double
counting. This yields
I
1
= G (r 2 + r 2 ) d.
(E.4)
4
From Figure E.1, the position vector of point Q, relative to the origin, O, is
x = x + r n,

(E.5)

where x = (x1 , x2 , x3 ) is the position vector of point P, and n a unit vector pointing from P to Q. Likewise, the
position vector of point R is
x = x + r n.
(E.6)
However, Q and R both lie on the bodys outer boundary. It follows, from (E.1), that r and r are the two roots of
X xi + r n i ! 2
= 1,
ai
i=1,3

(E.7)

A r2 + B r + C = 0,

(E.8)

which reduces to the quadratic


where
A

X n2
i

,
ai2
X xi n i

(E.9)

i=1,3

B =

i=1,3

X x2
i

i=1,3

(E.10)

1.

(E.11)

ai2

ai2

280

FLUID MECHANICS

Q
d
O

r
P

R
Figure E.1: Calculation of ellipsoidal gravitational potential.
Now, according to standard polynomial equation theory, r + r = B/A, and r r = C/A. Thus,
r 2 + r 2 = (r + r )2 2 r r =
and (E.4) becomes
1
= G
2

C
B2
2 ,
2
A
A


P
I 2 P
2 2
1 i=1,3 xi2 /ai2

i=1,3 xi ni /ai
d.

2 + P
2
2

P
2
2
i=1,3 ni /ai
i=1,3 ni /ai

The above expression can also be written

P
I 2 P
2 2
1 i=1,3 xi2 /ai2

i, j=1,3 xi x j ni n j /(ai a j )
1
d.
+ P
= G
2
P
2
2

2
2 /a 2
i=1,3 ni /ai
n
i=1,3 i
i

However, the cross terms (i.e., i , j) integrate to zero by symmetry, and we are left with

P
I P
2 i=1,3 xi2 ni2 /ai4 1 i=1,3 xi2 /ai2
1
d.
= G P
2 + P
2
2

2
i=1,3 ni /ai
n 2 /a 2
i=1,3

Let

J=

1 J
=
ai ai

It follows that

Thus, (E.15) can be written

where

(E.13)

(E.14)

(E.15)

d
.
2
2
i=1,3 ni /ai

P

2 ni2 /ai4

i=1,3

ni2 /ai2

2 d.

1
2
Ai xi ,
= G J
2
i=1,3
Ai =

(E.12)

J
1 J

.
2
ai ai
ai

(E.16)

(E.17)

(E.18)

(E.19)

Ellipsoidal Potential Theory

281

At this stage, it is convenient to adopt the spherical angular coordinates, and (see Section C.4), in terms of
which
n = (sin cos , sin sin , cos ),
(E.20)
and d = sin d d. We find, from (E.16), that
J=8

/2

sin d

/2

Let t = tan . It follows that


J=8

2
1
sin cos2 sin2 sin2 cos2
d.
+
+

a12
a22
a32

(E.21)

(E.22)

/2

sin d

dt
= 4
a + b t2

/2

sin d
,
(a b)1/2

where

Hence, we obtain
J = 4 a1 a2 a32

sin2 cos2
+
,
a12
a32

(E.23)

sin2 cos2
+
.
a22
a32

(E.24)

/2

sin sec2 d
.
(a12 + a32 tan2 )1/2 (a22 + a32 tan2 )1/2

Let u = a32 tan2 . It follows that


J = 2 a1 a2 a3

du
,

(E.25)

(E.26)

where
= (a12 + u)1/2 (a22 + u)1/2 (a32 + u)1/2 .

(E.27)

Now, from (E.19), (E.26), and (E.27),


Ai

Z !
du
du 1
=

2 a1 a2 a3

a
a

i
i
0
0
!
Z
Z
du
2 a1 a2 a3
1
=
du = 2 a1 a2 a3
.
2
ai
a

i
0 (ai + u)
0
2 a1 a2 a3
ai2

Thus, from (E.18), (E.26), and (E.28),

where

3
2

= G M 0
i xi ,
4
i=1,3
0

du
,

(ai2

(E.28)

(E.29)

(E.30)
du
.
+ u)

Here, M = V and V = (4/3) a1 a2 a3 are the bodys mass and volume, respectively.
The total gravitational potential energy of the body is written
Z
1
dV,
U=
2

(E.31)

(E.32)

282

FLUID MECHANICS

where the integral is taken over all interior points. It follows from (E.29) that

1
3
i ai2 .
U = G M 2 0
8
5

(E.33)

i=1,3

In writing the above, use has been made of the easily demonstrated result
2
so

i=1,3

i ai2 =

xi2 dV = (1/5) ai2 V. Now,

ai2
d u
1 X
= +
,
du
i=1,3 (ai2 + u)
X

i=1,3

ai2 du
(ai2 + u)

"

#
d u 1
+
du = 0 .
du

(E.34)

(E.35)

Hence, we obtain
U=

3
G M 2 0 .
10

(E.36)

Вам также может понравиться