You are on page 1of 574

Richard Fitzpatrick

Theoretical Fluid Mechanics


Contents

1 Mathematical Models of Fluid Motion 1


1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 What is a Fluid? . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Volume and Surface Forces . . . . . . . . . . . . . . . . . . . . . 2
1.4 General Properties of Stress Tensor . . . . . . . . . . . . . . . . . 4
1.5 Stress Tensor in a Static Fluid . . . . . . . . . . . . . . . . . . . . 6
1.6 Stress Tensor in a Moving Fluid . . . . . . . . . . . . . . . . . . 7
1.7 Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.8 Conservation Laws . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.9 Mass Conservation . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.10 Convective Time Derivative . . . . . . . . . . . . . . . . . . . . . 11
1.11 Momentum Conservation . . . . . . . . . . . . . . . . . . . . . . 12
1.12 Navier-Stokes Equation . . . . . . . . . . . . . . . . . . . . . . . 14
1.13 Energy Conservation . . . . . . . . . . . . . . . . . . . . . . . . 14
1.14 Equations of Incompressible Fluid Flow . . . . . . . . . . . . . . 17
1.15 Equations of Compressible Fluid Flow . . . . . . . . . . . . . . . 18
1.16 Dimensionless Numbers in Incompressible Flow . . . . . . . . . . 19
1.17 Dimensionless Numbers in Compressible Flow . . . . . . . . . . 20
1.18 Fluid Equations in Cartesian Coordinates . . . . . . . . . . . . . 24
1.19 Fluid Equations in Cylindrical Coordinates . . . . . . . . . . . . . 25
1.20 Fluid Equations in Spherical Coordinates . . . . . . . . . . . . . 27
1.21 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

2 Hydrostatics 33
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.2 Hydrostatic Pressure . . . . . . . . . . . . . . . . . . . . . . . . 33
2.3 Buoyancy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.4 Equilibrium of Floating Bodies . . . . . . . . . . . . . . . . . . . 35
2.5 Vertical Stability of Floating Bodies . . . . . . . . . . . . . . . . 37
2.6 Angular Stability of Floating Bodies . . . . . . . . . . . . . . . . 37
2.7 Determination of Metacentric Height . . . . . . . . . . . . . . . . 39
2.8 Energy of a Floating Body . . . . . . . . . . . . . . . . . . . . . 43
2.9 Curve of Buoyancy . . . . . . . . . . . . . . . . . . . . . . . . . 44
2.10 Rotational Hydrostatics . . . . . . . . . . . . . . . . . . . . . . . 49

2
3

2.11 Equilibrium of a Rotating Liquid Body . . . . . . . . . . . . . . . 51


2.12 Maclaurin Spheroids . . . . . . . . . . . . . . . . . . . . . . . . 54
2.13 Jacobi Ellipsoids . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2.14 Roche Ellipsoids . . . . . . . . . . . . . . . . . . . . . . . . . . 62
2.15 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

3 Surface Tension 73
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
3.2 Young-Laplace Equation . . . . . . . . . . . . . . . . . . . . . . 74
3.3 Spherical Interfaces . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.4 Capillary Length . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.5 Angle of Contact . . . . . . . . . . . . . . . . . . . . . . . . . . 78
3.6 Jurins Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.7 Capillary Curves . . . . . . . . . . . . . . . . . . . . . . . . . . 81
3.8 Axisymmetric Soap-Bubbles . . . . . . . . . . . . . . . . . . . . 86
3.9 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

4 Incompressible Inviscid Flow 95


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.2 Streamlines, Stream Tubes, and Stream Filaments . . . . . . . . . 95
4.3 Bernoullis Theorem . . . . . . . . . . . . . . . . . . . . . . . . 96
4.4 Euler Momentum Theorem . . . . . . . . . . . . . . . . . . . . . 98
4.5 dAlemberts Paradox . . . . . . . . . . . . . . . . . . . . . . . . 98
4.6 Flow Through an Orifice . . . . . . . . . . . . . . . . . . . . . . 100
4.7 Sub-Critical and Super-Critical Flow . . . . . . . . . . . . . . . . 102
4.8 Flow over Shallow Bump . . . . . . . . . . . . . . . . . . . . . . 103
4.9 Stationary Hydraulic Jumps . . . . . . . . . . . . . . . . . . . . . 105
4.10 Tidal Bores . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
4.11 Flow over a Broad-Crested Weir . . . . . . . . . . . . . . . . . . 110
4.12 Vortex Lines, Vortex Tubes, and Vortex Filaments . . . . . . . . . 111
4.13 Circulation and Vorticity . . . . . . . . . . . . . . . . . . . . . . 112
4.14 Kelvin Circulation Theorem . . . . . . . . . . . . . . . . . . . . 113
4.15 Irrotational Flow . . . . . . . . . . . . . . . . . . . . . . . . . . 114
4.16 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

5 Two-Dimensional Incompressible Inviscid Flow 119


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
5.2 Two-Dimensional Flow . . . . . . . . . . . . . . . . . . . . . . . 119
5.3 Velocity Potentials and Stream Functions . . . . . . . . . . . . . 122
5.4 Two-Dimensional Uniform Flow . . . . . . . . . . . . . . . . . . 123
5.5 Two-Dimensional Sources and Sinks . . . . . . . . . . . . . . . . 124
5.6 Two-Dimensional Vortex Filaments . . . . . . . . . . . . . . . . 126
5.7 Two-Dimensional Irrotational Flow in Cylindrical Coordinates . . 130
5.8 Flow Past a Cylindrical Obstacle . . . . . . . . . . . . . . . . . . 132
5.9 Motion of a Submerged Cylinder . . . . . . . . . . . . . . . . . . 136
4

5.10 Inviscid Flow Past a Semi-Infinite Wedge . . . . . . . . . . . . . 141


5.11 Inviscid Flow Over a Semi-Infinite Wedge . . . . . . . . . . . . . 143
5.12 Two-Dimensional Jets . . . . . . . . . . . . . . . . . . . . . . . . 144
5.13 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147

6 Two-Dimensional Potential Flow 149


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
6.2 Complex Functions . . . . . . . . . . . . . . . . . . . . . . . . . 149
6.3 Cauchy-Riemann Relations . . . . . . . . . . . . . . . . . . . . . 150
6.4 Complex Velocity Potential . . . . . . . . . . . . . . . . . . . . . 152
6.5 Complex Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . 153
6.6 Method of Images . . . . . . . . . . . . . . . . . . . . . . . . . . 154
6.7 Conformal Maps . . . . . . . . . . . . . . . . . . . . . . . . . . 160
6.8 SchwarzChristoel Theorem . . . . . . . . . . . . . . . . . . . 165
6.9 Free Streamline Theory . . . . . . . . . . . . . . . . . . . . . . . 168
6.10 Complex Line Integrals . . . . . . . . . . . . . . . . . . . . . . . 179
6.11 Blasius Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 181
6.12 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

7 Axisymmetric Incompressible Inviscid Flow 193


7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
7.2 Axisymmetric Flow . . . . . . . . . . . . . . . . . . . . . . . . . 193
7.3 Stokes Stream Function . . . . . . . . . . . . . . . . . . . . . . . 193
7.4 Axisymmetric Velocity Fields . . . . . . . . . . . . . . . . . . . 195
7.5 Axisymmetric Irrotational Flow in Spherical Coordinates . . . . . 196
7.6 Uniform Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
7.7 Point Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
7.8 Dipole Point Sources . . . . . . . . . . . . . . . . . . . . . . . . 199
7.9 Flow Past a Spherical Obstacle . . . . . . . . . . . . . . . . . . . 200
7.10 Motion of a Submerged Sphere . . . . . . . . . . . . . . . . . . . 203
7.11 Conformal Maps . . . . . . . . . . . . . . . . . . . . . . . . . . 205
7.12 Flow Around a Submerged Oblate Spheroid . . . . . . . . . . . . 210
7.13 Flow Around a Submerged Prolate Spheroid . . . . . . . . . . . . 213
7.14 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215

8 Incompressible Boundary Layers 219


8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
8.2 No Slip Condition . . . . . . . . . . . . . . . . . . . . . . . . . . 219
8.3 Boundary Layer Equations . . . . . . . . . . . . . . . . . . . . . 220
8.4 Self-Similar Boundary Layers . . . . . . . . . . . . . . . . . . . 224
8.5 Boundary Layer on a Flat Plate . . . . . . . . . . . . . . . . . . . 229
8.6 Wake Downstream of a Flat Plate . . . . . . . . . . . . . . . . . . 234
8.7 Von Karman Momentum Integral . . . . . . . . . . . . . . . . . . 239
8.8 Boundary Layer Separation . . . . . . . . . . . . . . . . . . . . . 240
8.9 Criterion for Boundary Layer Separation . . . . . . . . . . . . . . 244
5

8.10 Approximate Solutions of Boundary Layer Equations . . . . . . . 247


8.11 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253

9 Incompressible Aerodynamics 257


9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
9.2 Theorem of Kutta and Zhukovskii . . . . . . . . . . . . . . . . . 258
9.3 Cylindrical Airfoils . . . . . . . . . . . . . . . . . . . . . . . . . 260
9.4 Zhukovskiis Hypothesis . . . . . . . . . . . . . . . . . . . . . . 263
9.5 Vortex Sheets . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
9.6 Induced Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
9.7 Three-Dimensional Airfoils . . . . . . . . . . . . . . . . . . . . . 271
9.8 Aerodynamic Forces . . . . . . . . . . . . . . . . . . . . . . . . 275
9.9 Ellipsoidal Airfoils . . . . . . . . . . . . . . . . . . . . . . . . . 278
9.10 Simple Flight Problems . . . . . . . . . . . . . . . . . . . . . . . 281
9.11 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283

10 Incompressible Viscous Flow 287


10.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
10.2 Flow Between Parallel Plates . . . . . . . . . . . . . . . . . . . . 287
10.3 Flow Down an Inclined Plane . . . . . . . . . . . . . . . . . . . . 289
10.4 Poiseuille Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . 291
10.5 Taylor-Couette Flow . . . . . . . . . . . . . . . . . . . . . . . . 292
10.6 Flow in Slowly-Varying Channels . . . . . . . . . . . . . . . . . 293
10.7 Lubrication Theory . . . . . . . . . . . . . . . . . . . . . . . . . 296
10.8 Stokes Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
10.9 Axisymmetric Stokes Flow . . . . . . . . . . . . . . . . . . . . . 299
10.10 Axisymmetric Stokes Flow Around a Solid Sphere . . . . . . . . 300
10.11 Axisymmetric Stokes Flow In and Around a Fluid Sphere . . . . . 306
10.12 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309

11 Waves in Incompressible Fluids 313


11.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
11.2 Gravity Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
11.3 Gravity Waves in Deep Water . . . . . . . . . . . . . . . . . . . . 315
11.4 Gravity Waves in Shallow Water . . . . . . . . . . . . . . . . . . 317
11.5 Energy of Gravity Waves . . . . . . . . . . . . . . . . . . . . . . 318
11.6 Wave Drag on Ships . . . . . . . . . . . . . . . . . . . . . . . . . 320
11.7 Ship Wakes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
11.8 Gravity Waves in a Flowing Fluid . . . . . . . . . . . . . . . . . 328
11.9 Gravity Waves at an Interface . . . . . . . . . . . . . . . . . . . . 329
11.10 Steady Flow over a Corrugated Bottom . . . . . . . . . . . . . . . 331
11.11 Surface Tension . . . . . . . . . . . . . . . . . . . . . . . . . . . 332
11.12 Capillary Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . 334
11.13 Capillary Waves at an Interface . . . . . . . . . . . . . . . . . . . 335
11.14 Wind Driven Waves in Deep Water . . . . . . . . . . . . . . . . . 336
6

11.15 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338

12 Terrestrial Ocean Tides 341


12.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
12.2 Tide Generating Potential . . . . . . . . . . . . . . . . . . . . . . 341
12.3 Decomposition of Tide Generating Potential . . . . . . . . . . . . 343
12.4 Expansion of Tide Generating Potential . . . . . . . . . . . . . . 344
12.5 Surface Harmonics and Solid Harmonics . . . . . . . . . . . . . . 345
12.6 Planetary Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . 346
12.7 Total Gravitational Potential . . . . . . . . . . . . . . . . . . . . 347
12.8 Planetary Response . . . . . . . . . . . . . . . . . . . . . . . . . 348
12.9 Laplace Tidal Equations . . . . . . . . . . . . . . . . . . . . . . . 353
12.10 Harmonics of Forcing Term in Laplace Tidal Equations . . . . . . 357
12.11 Response to Equilibrium Harmonic . . . . . . . . . . . . . . . . . 359
12.12 Global Ocean Tides . . . . . . . . . . . . . . . . . . . . . . . . . 360
12.13 Non-Global Ocean Tides . . . . . . . . . . . . . . . . . . . . . . 366
12.14 Useful Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
12.15 Transformation of Laplace Tidal Equations . . . . . . . . . . . . 369
12.16 Another Useful Lemma . . . . . . . . . . . . . . . . . . . . . . . 371
12.17 Basis Eigenfunctions . . . . . . . . . . . . . . . . . . . . . . . . 371
12.18 Auxilliary Eigenfunctions . . . . . . . . . . . . . . . . . . . . . . 373
12.19 Gyroscopic Coecients . . . . . . . . . . . . . . . . . . . . . . . 374
12.20 Proudman Equations . . . . . . . . . . . . . . . . . . . . . . . . 376
12.21 Hemispherical Ocean Tides . . . . . . . . . . . . . . . . . . . . . 380
12.22 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 385

13 Equilibrium of Compressible Fluids 393


13.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393
13.2 Isothermal Atmosphere . . . . . . . . . . . . . . . . . . . . . . . 394
13.3 Adiabatic Atmosphere . . . . . . . . . . . . . . . . . . . . . . . 394
13.4 Atmospheric Stability . . . . . . . . . . . . . . . . . . . . . . . . 396
13.5 Eddington Solar Model . . . . . . . . . . . . . . . . . . . . . . . 397
13.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 404

14 One-Dimensional Compressible Inviscid Flow 407


14.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407
14.2 Thermodynamic Considerations . . . . . . . . . . . . . . . . . . 407
14.3 Isentropic Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . 411
14.4 Sound Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . 411
14.5 Bernoullis Theorem . . . . . . . . . . . . . . . . . . . . . . . . 414
14.6 Mach Number . . . . . . . . . . . . . . . . . . . . . . . . . . . . 415
14.7 Sonic Flow through a Nozzle . . . . . . . . . . . . . . . . . . . . 416
14.8 Normal Shocks . . . . . . . . . . . . . . . . . . . . . . . . . . . 420
14.9 Piston-Generated Shock Wave . . . . . . . . . . . . . . . . . . . 425
14.10 Piston-Generated Expansion Wave . . . . . . . . . . . . . . . . . 427
7

14.11 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430

15 Two-Dimensional Compressible Inviscid Flow 437


15.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437
15.2 Oblique Shocks . . . . . . . . . . . . . . . . . . . . . . . . . . . 437
15.3 Supersonic Flow in Corner or over Wedge . . . . . . . . . . . . . 441
15.4 Weak Oblique Shocks . . . . . . . . . . . . . . . . . . . . . . . . 443
15.5 Supersonic Compression by Turning . . . . . . . . . . . . . . . . 445
15.6 Supersonic Expansion by Turning . . . . . . . . . . . . . . . . . 447
15.7 Detached Shocks . . . . . . . . . . . . . . . . . . . . . . . . . . 450
15.8 Shock-Expansion Theory . . . . . . . . . . . . . . . . . . . . . . 451
15.9 Thin-Airfoil Theory . . . . . . . . . . . . . . . . . . . . . . . . . 453
15.10 Croccos Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 458
15.11 Homenergic Homentropic Flow . . . . . . . . . . . . . . . . . . . 459
15.12 Small-Perturbation Theory . . . . . . . . . . . . . . . . . . . . . 460
15.13 Subsonic Flow Past a Wave-Shaped Wall . . . . . . . . . . . . . . 464
15.14 Supersonic Flow Past a Wave-Shaped Wall . . . . . . . . . . . . . 466
15.15 Linearized Subsonic Flow . . . . . . . . . . . . . . . . . . . . . . 468
15.16 Linearized Supersonic Flow . . . . . . . . . . . . . . . . . . . . 470
15.17 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 472

A Vectors and Vector Fields 483


A.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 483
A.2 Scalars and Vectors . . . . . . . . . . . . . . . . . . . . . . . . . 483
A.3 Vector Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . 484
A.4 Cartesian Components of a Vector . . . . . . . . . . . . . . . . . 486
A.5 Coordinate Transformations . . . . . . . . . . . . . . . . . . . . 487
A.6 Scalar Product . . . . . . . . . . . . . . . . . . . . . . . . . . . . 489
A.7 Vector Area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 491
A.8 Vector Product . . . . . . . . . . . . . . . . . . . . . . . . . . . . 492
A.9 Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 495
A.10 Scalar Triple Product . . . . . . . . . . . . . . . . . . . . . . . . 496
A.11 Vector Triple Product . . . . . . . . . . . . . . . . . . . . . . . . 498
A.12 Vector Calculus . . . . . . . . . . . . . . . . . . . . . . . . . . . 498
A.13 Line Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . 499
A.14 Vector Line Integrals . . . . . . . . . . . . . . . . . . . . . . . . 502
A.15 Surface Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . 502
A.16 Vector Surface Integrals . . . . . . . . . . . . . . . . . . . . . . . 504
A.17 Volume Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . 505
A.18 Gradient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506
A.19 Grad Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . 510
A.20 Divergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 511
A.21 Laplacian Operator . . . . . . . . . . . . . . . . . . . . . . . . . 514
A.22 Curl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516
A.23 Useful Vector Identities . . . . . . . . . . . . . . . . . . . . . . . 519
8

A.24 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 520

B Cartesian Tensors 525


B.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 525
B.2 Tensors and Tensor Notation . . . . . . . . . . . . . . . . . . . . 525
B.3 Tensor Transformation . . . . . . . . . . . . . . . . . . . . . . . 528
B.4 Tensor Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . 531
B.5 Isotropic Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . 533
B.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 536

C Non-Cartesian Coordinates 539


C.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 539
C.2 Orthogonal Curvilinear Coordinates . . . . . . . . . . . . . . . . 539
C.3 Cylindrical Coordinates . . . . . . . . . . . . . . . . . . . . . . . 543
C.4 Spherical Coordinates . . . . . . . . . . . . . . . . . . . . . . . . 546
C.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 548

D Ellipsoidal Potential Theory 551


D.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 551
D.2 Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 551
D.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 555

E Calculus of Variations 557


E.1 Indroduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 557
E.2 Euler-Lagrange Equation . . . . . . . . . . . . . . . . . . . . . . 557
E.3 Conditional Variation . . . . . . . . . . . . . . . . . . . . . . . . 559
E.4 Multi-Function Variation . . . . . . . . . . . . . . . . . . . . . . 562
E.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 562

Bibliography 565
1
Mathematical Models of Fluid Motion

1.1 Introduction
This chapter sets 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 (which means that re-
peated roman subscripts are assumed to be summed from 1 to 3). (See Appendix B.)
More information on the fundamental assumptions that underlie conventional fluid
mechanics, as well as the basic equations that describe fluid motion, can be found in
Batchelor 2000.

1.2 What is a Fluid?


By definition, a solid material is rigid. 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 above 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. This does not preclude the possibility that such an element
may oer 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

1
2 Theoretical Fluid Mechanics

dierence between these two types of fluid lies in their relative compressibility. To be
more exact, 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.
A macroscopic fluid ultimately consists of a huge number of individual molecules.
However, most practical applications of fluid mechanics are concerned with behav-
ior on lengthscales that are far longer 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 ele-
ments, we really mean elements that are suciently small that the bulk fluid proper-
ties (such as mass density, pressure, and velocity) are approximately constant across
them, but are still suciently large that they contain a very great number of molecules
(which implies that we can safely neglect any statistical variations in the bulk prop-
erties). 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. In other words, there is no preferred direction in a
fluid.
2. Fluids are Newtonian. In other words, there is a linear relationship between the
local shear stress and the local rate of strain, as first postulated by Isaac Newton
(1642-1727). 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. In other words, the macroscopic motion of ordinary fluids is
well described by Newtonian dynamics, and both quantum and relativistic eects
can be safely ignored.
It should be noted that the previous 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 eects on macroscopic lengthscales). However, most practical
applications of fluid mechanics involve the equilibrium and motion of bodies of wa-
ter 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.

1.3 Volume and Surface Forces


Generally speaking, fluids are acted upon by two distinct types of force. The first type
is long range in naturethat is, such that it decreases relatively slowly with increas-
Mathematical Models of Fluid Motion 3

ing 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 suciently small volume
element. In this situation, the net force acting on the element becomes directly pro-
portional 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. (1.1)

The second type of force is short range in nature, and is most conveniently mod-
eled 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 den-
sity 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 previous 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). (1.2)
In tensor notation (see Appendix B), the previous equation can be written

fi = i j dS j , (1.3)

where 11 = ( x ) x , 12 = ( x )y , 21 = (y ) x , et cetera. (Note that, because the


subscript j is repeated in the previous equation, it is assumed to be 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
4 Theoretical Fluid Mechanics

first-order tensor (as all forces are proper vectors), and the dS i are the components
of an arbitrary first-order tensor (as surface elements are also proper vectorssee
Section A.7and Equation (1.3) holds for surface elements whose normals point in
any direction), so application of the quotient rule (see Section B.3) to Equation (1.3)
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,
because 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, because they drive shearing
motion in which parallel layers of fluid slide relative to one another.

1.4 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
 
fi = Fi dV + i j dS j , (1.4)
V 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 . Mak-
ing use of the tensor divergence theorem (see Section B.4), the previous expression
becomes  
i j
fi = Fi dV + dV. (1.5)
V V x j

In the limit V 0, it is reasonable to suppose that the Fi and i j /x j are ap-


proximately constant across the element. In this situation, both contributions on the
right-hand side of the previous equation scale as V. According to Newtonian dy-
namics, the i-component of the net force acting on the volume 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
Mathematical Models of Fluid Motion 5

surface S is written [see Equations (A.46) and (B.6)]


 
i = i jk x j Fk dV + i jk x j kl dS l , (1.6)
V S

where the first and second terms on the right-hand side are due to volume and sur-
face forces, respectively. [Here, i jk is the third-order permutation tensor. See Equa-
tion (B.7).] Making use of the tensor divergence theorem (see Section B.4), the
previous expression becomes
 
(x j kl )
i = i jk x j Fk dV + i jk dV, (1.7)
V V xl
which reduces to
  
kl
i = i jk x j Fk dV + i jk k j dV + i jk x j dV, (1.8)
V V V xl
because xi /x j = i j . [Here, i j is the second-order identity tensor. See Equa-
tion (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 previous equation scale as V 4/3 , V, and V 4/3 , respectively (because x V 1/3 ).
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 and density of the fluid
are approximately constant across the element, we deduce that the rate of change of
its angular momentum scales as V 4/3 (because the net rate of change of the linear
momentum scales as V, so the net rate of change of the 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 Equation (1.8) in the limit that the volume
of the element approaches zero (because 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 (1.9)
throughout the fluid. The previous relation shows that the stress tensor must be sym-
metric: that is,
ji = i j . (1.10)
It immediately follows that the stress tensor only has six independent components
(i.e., 11 , 22 , 33 , 12 , 13 , and 23 ).
It is always possible to choose the orientation of a set of Cartesian axes in such a
manner that the non-diagonal 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 refer-
ence to such principal axes, the diagonal components of the stress tensor i j become
6 Theoretical Fluid Mechanics

so-called principal stresses11 , 22 , 33 , say. 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 com-
pression 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: that is,

ii = 11 + 22 + 33 . (1.11)

1.5 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: that is,
1
3 ii , 0, 0

0, 3 ii ,
1
0 , (1.12)

0, 0, 1

3 ii

and 
11 13 ii , 0, 0


0, 22 3 ii ,
1
0 . (1.13)

0, 0, 33 13 ii
The tensor (1.12) is isotropic (see Section B.5), and corresponds to the same
normal force per unit area acting inward (because 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 tensor (1.13) represents the departure of the stress tensor from an isotropic
form. The diagonal components of this tensor have zero sum, in view of Equa-
tion (1.11), 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 act to change the shape of the vol-
ume element, either elongating or compressing it along one of its symmetry axes.
Moreover, this tendency cannot be oset by any volume force acting on the element,
Mathematical Models of Fluid Motion 7

because 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 1.4). We pre-
viously defined a fluid as a material that is incapable of withstanding any tendency
of applied forces to change its shape. (See Section 1.2.) It follows that if the diag-
onal components of the tensor (1.13) 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 (1.12) every-
where in a stationary fluid. Furthermore, this is true irrespective of the orientation of
the coordinate axes, because the components of an isotropic tensor are rotationally
invariant. (See Section B.5.)
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 , (1.14)

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 (1.3).] 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, after its discoverer Blaise Pascal (1623-
1662), 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.

1.6 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 , (1.15)

where p = ii /3 is the static pressure: that is, minus the normal stress acting in
any direction. The normal stress at a given point in a moving fluid generally varies
with direction. In other words, 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: that is,
1
p = ii . (1.16)
3
8 Theoretical Fluid Mechanics

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 , (1.17)

where
dii = 0. (1.18)
Moreover, because i j and i j are both symmetric tensors, it follows that di j is also
symmetric: that is,
d ji = di j . (1.19)
It is clear that the so-called deviatoric stress tensor, di j , is a consequence of fluid
motion, because 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. We would expect the force distribution within the
fluid to be the same in both frames of reference, because the fluid does not accelerate
in either. However, in the first frame, the fluid appears stationary and the deviatoric
stress tensor is therefore zero, while 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 previous discussion,
is that the components of the deviatoric stress tensor are linear functions of these
velocity gradients: that is,
vk
di j = Ai jkl . (1.20)
xl
Here, Ai jkl is a fourth-order tensor (this follows from the quotient rule because di j
and vi /x j are both proper second-order tensors). Any fluid in which the deviatoric
stress tensor takes the previous form is termed a Newtonian fluid, because Newton
was the first to postulate a linear relationship between shear stresses and velocity
gradients.
In an isotropic fluidthat is, a fluid in which there is no preferred directionwe
would expect the fourth-order tensor Ai jkl to be isotropicthat is, to have a form in
which all physical distinction between dierent 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 , (1.21)

where , , and are arbitrary scalars (which can be functions of position and time).
Mathematical Models of Fluid Motion 9

Thus, it follows from Equations (1.20) and (1.21) that


vk vi v j
di j = i j + + . (1.22)
xk x j xi
However, according to Equation (1.19), di j is a symmetric tensor, which implies that
= , and
di j = ekk i j + 2 ei j , (1.23)
where

1 vi v j
ei j = + (1.24)
2 x j xi
is called the rate of strain tensor. Finally, according to Equation (1.18), di j is a
traceless tensor, which yields 3 = 2 , and


1
di j = 2 ei j ekk i j , (1.25)
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 , (1.26)
3
where p(r, t) and (r, t) are arbitrary scalars.

1.7 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 (1.25) 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
v1
d12 = d21 = . (1.27)
x2
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 acting 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 (Batchelor 2000).
10 Theoretical Fluid Mechanics

1.8 Conservation Laws


Suppose that (r, t) is the density of some bulk fluid property (e.g., mass, momen-
tum, or 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

= dV, (1.28)
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 
= v dS, (1.29)
S
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 . (1.30)
dt
In other words, the rate of increase in the amount of the property contained within
V is the dierence 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 previous conservation
law can also be written
d
+ = S . (1.31)
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.

1.9 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 
M= dV, (1.32)
V
Mathematical Models of Fluid Motion 11

where dV is an element of V. Furthermore, the mass flux across S , and out of V, is


[see Equation (1.29)] 
M = v dS, (1.33)
S
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: that is,
dM
+ M = 0 (1.34)
dt
[cf., Equation (1.31)]. Here, we are assuming that there is no mass generation (or
destruction) within V (because individual molecules are eectively indestructible).
It follows that  

dV + v dS = 0, (1.35)
V t S
because V is non-time-varying. Making use of the divergence theorem (see Sec-
tion A.20), the previous equation becomes


+ ( v) dV = 0. (1.36)
V t

However, this result is true irrespective of the size, shape, or location of volume V,
which is only possible if

+ ( v) = 0 (1.37)
t
throughout the fluid. The previous expression is known as the equation of fluid con-
tinuity, and is a direct consequence of mass conservation.

1.10 Convective Time Derivative


The quantity (r, t)/t, appearing in Equation (1.37), 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
(r + v t, t + t) (r, t) D
lim = + v = , (1.38)
t0 t t Dt
where we have Taylor expanded (r + v t, t + t) up to first order in t, and where
D
= +v = + vi . (1.39)
Dt t t xi
Clearly, the so-called convective time derivative, D/Dt, represents the time derivative
seen in the local rest frame of the fluid.
12 Theoretical Fluid Mechanics

The continuity equation (1.37) can be rewritten in the form

1 D D ln
= = v, (1.40)
Dt Dt

because ( v) = v + v. [See Equation (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
   
DV vi
= v dS = vi dS i = dV = v dV, (1.41)
Dt S S V xi V

where S is the bounding surface of the element, and use has been made of the diver-
gence theorem. In the limit that V 0, and v is approximately constant across
the element, we obtain
1 DV D ln V
= = v. (1.42)
V Dt Dt
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.

1.11 Momentum Conservation


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

Pi = vi dV. (1.43)
V

Moreover, the flux of i-momentum across S , and out of V, is [see Equation (1.29)]

i = vi v j dS j . (1.44)
S

Finally, the i-component of the net force acting on the fluid within V is
 
fi = Fi dV + i j dS j , (1.45)
V S

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
Mathematical Models of Fluid Motion 13

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 (1.31)]

dPi
+ i = fi , (1.46)
dt

which can be written


   
( vi )
dV + vi v j dS j = Fi dV + i j dS j , (1.47)
V t S V S

because the volume V is non-time-varying. Making use of the tensor divergence


theorem, this becomes
 

( vi ) ( vi v j ) i j
+ dV = Fi + dV. (1.48)
V t x j V x j

However, the previous result is valid irrespective of the size, shape, or location of
volume V, which is only possible if

( vi ) ( vi v j ) i j
+ = Fi + (1.49)
t x j x j

everywhere inside the fluid. Expanding the derivatives, and rearranging, we obtain



v j vi vi i j
+ vj + vi + + vj = Fi + . (1.50)
t x j x j t x j x j

In tensor notation, the continuity equation (1.37) is written

v j
+ vj + = 0. (1.51)
t x j x j

So, combining Equations (1.50) and (1.51), we obtain the following fluid equation of
motion,


vi vi i j
+ vj = Fi + . (1.52)
t x j x j

An alternative form of this equation is

Dvi Fi 1 i j
= + . (1.53)
Dt x j

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

1.12 Navier-Stokes Equation


Equations (1.24), (1.26), and (1.53) can be combined to give the equation of motion
of an isotropic, Newtonian, classical fluid:



Dvi p vi v j 2 v j
= Fi + + . (1.54)
Dt xi x j x j xi xi 3 x j
This equation is generally known as the Navier-Stokes equation, and is named after
Claude-Louis Navier (1785-1836) and George Gabriel Stokes (1819-1903). In sit-
uations 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 p vi 1 2v j
= Fi + + . (1.55)
Dt xi x j x j 3 xi x j
When expressed in vector form, the previous expression becomes

Dv v 1
+ (v ) v = F p + v + ( v) ,
2
(1.56)
Dt t 3
where use has been made of Equation (1.39). Here,
bi
[(a )b]i = a j , (1.57)
x j
( 2 v)i = 2 vi . (1.58)
Note, however, that the previous identities are only valid in Cartesian coordinates.
(See Appendix C.)

1.13 Energy Conservation


Consider a fixed volume V surrounded by a surface S . The total energy content of
the fluid contained within V is
 
1
E= E dV + vi vi dV, (1.59)
V V 2

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 (1.29)]



1 1
E = E + vi vi v j dS j = E + vi vi v j dV, (1.60)
S 2 V x j 2
Mathematical Models of Fluid Motion 15

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: that is,
dE . .
+ E = W Q, (1.61)
dt
. . . .
where W is the net rate of work, and Q the net heat flux. It can be seen that W Q is
the eective energy generation rate within V [cf., Equation (1.31)].
The net rate at which volume and surface forces do work on the fluid within V is

.  
(vi i j )
W= vi Fi dV + vi i j dS j = vi F i + dV, (1.62)
V S V x j
where use has been made of the tensor divergence theorem.
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: that is,
T
q i = Ai j , (1.63)
x j
where the Ai j are the components of a second-rank tensor (which can be functions
of position and time). In an isotropic fluid we would expect Ai j to be an isotropic
tensor. (See Section B.5.) However, the most general second-order isotropic tensor
is simply a multiple of i j . Hence, we can write

Ai j = i j , (1.64)

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 = , (1.65)
xi
or, equivalently,
q = T. (1.66)
Moreover, it is a matter of experience that heat flows down temperature gradients:
that is, > 0. We conclude that the net heat flux out of volume V is
 

. T T
Q= dS i = dV, (1.67)
S xi V xi xi
where use has been made of the tensor divergence theorem.
16 Theoretical Fluid Mechanics

Equations (1.59)(1.62) and (1.67) can be combined to give the following energy
conservation equation:
 


1 1
E + vi vi + E + vi vi v j dV
V t 2 x j 2


T
= vi F i + vi i j + dV. (1.68)
V x j x j
However, this result is valid irrespective of the size, shape, or location of volume V,
which is only possible if




1 1 T
E + vi vi + E + vi vi v j = vi F i + vi i j + (1.69)
t 2 x j 2 x j x j
everywhere inside the fluid. Expanding some of the derivatives, and rearranging, we
obtain


D 1 T
E + vi vi = vi F i + vi i j + , (1.70)
Dt 2 x j x j
where use has been made of the continuity equation, (1.40). The scalar product of v
with the fluid equation of motion, (1.53), yields


Dvi D 1 i j
vi = vi vi = vi F i + vi . (1.71)
Dt Dt 2 x j
Combining the previous two equations, we get


DE vi T
= i j + . (1.72)
Dt x j x j x j
Finally, making use of Equation (1.26), we deduce that the energy conservation equa-
tion for an isotropic Newtonian fluid takes the general form


DE p vi 1 T
= + + . (1.73)
Dt xi x j x j
Here,



vi 1 vi vi vi v j 2 vi v j
= di j = 2 ei j ei j eii e j j = + (1.74)
x j 3 x j x j x j xi 3 xi x j
is the rate of heat generation per unit volume due to viscosity. When written in vector
form, Equation (1.73) becomes
DE p ( T )
= v+ + . (1.75)
Dt
According to the previous 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.
Mathematical Models of Fluid Motion 17

1.14 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. 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: that is,
D
= 0. (1.76)
Dt
In this case, the continuity equation (1.40) reduces to
v = 0. (1.77)
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 (1.42), 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 incom-
pressible fluid is uniform in space. Hence, it follows from Equation (1.76) 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): that is,
F = , (1.78)
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
Dv p
= + 2 v, (1.79)
Dt
where

= (1.80)

is termed the kinematic viscosity, and has units of meters
squared per second. Roughly
speaking, momentum diuses a distance of order t meters in t seconds as a conse-
quence of viscosity. The kinematic viscosity of water at 20 C is about 1.0106 m2 /s
(Batchelor 2000). It follows that viscous momentum diusion in water is a relatively
slow process.
The complete set of equations governing incompressible flow is
v = 0, (1.81)
Dv p
= + 2 v. (1.82)
Dt
18 Theoretical Fluid Mechanics

Here, and are regarded as known constants, and (r, t) as a known function.
Thus, we have four equationsnamely, Equation (1.81), plus the three components
of Equation (1.82)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.

1.15 Equations of Compressible Fluid Flow


In many situations of general interest, the flow of gases is compressible. In other
words, 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 (1.40), and the
Navier-Stokes equation (1.56), must be augmented by the energy conservation equa-
tion (1.75), 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 (Reif 1965)

cV
E= T, (1.83)
M
M p
T= , (1.84)
R

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 corre-
sponds to 6.0221 1024 molecules. Here, we have assumed, for the sake of simplic-
ity, 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, Equa-
tions (1.40), (1.75), (1.83), and (1.84) can be combined to give



1 Dp p D M 2 p
=+ , (1.85)
1 Dt Dt R

where
cp cV + R
= = (1.86)
cV cV

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 (Reif 1965).] (See Section 14.2.) The ratio of specific
heats of dry air at 20 C is 1.40 (Batchelor 2000).
Mathematical Models of Fluid Motion 19

The complete set of equations governing compressible ideal gas flow are
D
= v, (1.87)
Dt

Dv p 2 1
= + v + ( v) , (1.88)
Dt 3



1 Dp p D M 2 p
=+ , (1.89)
1 Dt Dt R
where the dissipation function is specified in terms of and v in Equation (1.74).
Here, , , , M, and R are regarded as known constants, and (r, t) as a known
function. Thus, we have five equationsnamely, Equations (1.87) and (1.89), plus
the three components of Equation (1.88)for five unknownsnamely, the density,
(r, t), the pressure, p(r, t), and the three components of the velocity, v(r, t).

1.16 Dimensionless Numbers in Incompressible Flow


It is helpful to normalize the equations of incompressible fluid flow, (1.81)(1.82),
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, (1.90)

2
Dv 1 1 v
= 1+ 2 + p 2 + , (1.91)
Dt Fr Re Fr Re

where D/Dt = /t + v , and


L V0
Re = , (1.92)

V0
Fr = . (1.93)
(g L)1/2
Here, the dimensionless quantities Re and Fr are known as the Reynolds number
and the Froude number, respectively. [After Osborne Reynolds (18421912) and
William Froude (18101879), 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
20 Theoretical Fluid Mechanics

relatively important compared to inertia when Fr  1, and vice versa. Note that, in
principal, Re and Fr are the only quantities in Equations (1.90) and (1.91) that can be
significantly greater or smaller than unity.
For the case of water at 20 C, located on the surface of the Earth (Batchelor
2000),

Re 1.0 106 L(m) V0 (m s1 ), (1.94)


Fr 3.2 101 V0 (m s1 )/[L(m)]1/2. (1.95)

Thus, if L 1 m and V0 1 m s1 , as is often the case for terrestrial water dynamics,


then the previous expressions suggest that Re
1 and Fr O(1). In this situa-
tion, the viscous term on the right-hand side of Equation (1.91) becomes negligible,
and the (unnormalized) incompressible fluid flow equations reduce to the following
inviscid, incompressible, fluid flow equations:

v = 0, (1.96)
Dv p
= . (1.97)
Dt

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 (Batchelor 2000)

Re 1.0 104 L(m) V0 (m s1 ), (1.98)


Fr 3.2 101 V0 (m s1 )]/[L(m)]1/2. (1.99)

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 previous expressions, that Re  1
and Fr  1. In this situation, the inertial term on the left-hand side of (1.91) becomes
negligible, and the (unnormalized) incompressible fluid flow equations reduce to the
following inertia-free, incompressible, fluid flow equations:

v = 0, (1.100)
p
0= + 2 v. (1.101)

1.17 Dimensionless Numbers in Compressible Flow


It is helpful to normalize the equations of compressible ideal gas flow, (1.87)(1.89),
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 spa-
tial variation lengthscale, V0 a typical fluid velocity, 0 a typical mass density, and g
Mathematical Models of Fluid Motion 21

a typical gravitational acceleration (assuming that represents a gravitational poten-


tial 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
= v, (1.102)
Dt


Dv 1 1 p 1 2 1
= 1+ 2 + 2 + v ( v) , (1.103)
Dt Fr Re Fr Re 3



1 Dp p0 + p D 1 2 p0 + p
= + , (1.104)
1 Dt Dt 1 + Re (1 + 1/Fr 2 ) Re Pr
1
p0 = , (1.105)
Ma (1 + 1/Fr 2 + 1/Re)
2

where D/Dt /t + v ,
L V0
Re = , (1.106)

V0
Fr = , (1.107)
(g L)1/2

Pr = , (1.108)
H
V0
Ma =  , (1.109)
p0 /0
and

= , (1.110)
0
M
H = . (1.111)
R 0
Here, the dimensionless numbers Re, Fr, Pr, and Ma are known as the Reynolds
number, Froude number, Prandtl number, and Mach number, respectively. [The
latter two numbers are named after Ludwig Prandtl (18751953) and Ernst Mach
(18381916), respectively.] The Reynolds number is the typical ratio of inertial to
viscous forces within the gas, the square of the Froude number is the typical ratio
of inertial to gravitational forces, the Prandtl number is the typical ratio of the mo-
mentum and thermal diusion rates, and the Mach number is the typical ratio of the
gas flow and sound propagation speeds. Thus, thermal diusion is far faster than
momentum diusion when Pr  1, and vice versa. Moreover, the gas flow is termed
 Ma  1, supersonic when Ma
1, and transonic when Ma O(1).
subsonic when
Note that p0 /0 is the speed of sound in the undisturbed gas (Reif 1965). The
22 Theoretical Fluid Mechanics

quantity H is called the thermal diusivity of the gas, and has units of
meters squared
per second. Thus, heat typically diuses through the gas a distance H t meters in
t seconds. The thermal diusivity of dry air at atmospheric pressure and 20 C is
about H = 2.1 105 m2 s1 (Batchelor 2000). It follows that heat diusion in air
is a relatively slow process. The kinematic viscosity of dry air at atmospheric pres-
sure and 20 C is about = 1.5 105 m2 s1 (Batchelor 2000). Hence, momentum
diusion in air is also a relatively slow process.
For the case of dry air at atmospheric pressure and 20 C (Batchelor 2000),

Re 6.7 104 L(m) V0 (m s1 ), (1.112)


1 1
Fr 3.2 10 V0 (m s )/[L(m)] 1/2
, (1.113)
Pr 7.2 101 , (1.114)
Ma 2.9 103 V0 (m s1 ). (1.115)

Thus, if L 1 m and V0 1 m s1 , as is often the case for subsonic air dynamics


close to the Earths surface, then the previous expressions suggest that Re
1,
Ma  1, and Fr, Pr O(1). It immediately follows from Equation (1.105) that
p0
1. However, in this situation, Equation (1.104) 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. (1.116)
Dt
Equation (1.102) then gives
v  1. (1.117)

It is evident that subsonic (i.e., Ma  1) gas flow is essentially incompressible. The


fact that Re
1 implies that such flow is also essentially inviscid. In the incompress-
ible 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:

v = 0, (1.118)
Dv p
= . (1.119)
Dt

It follows that the equations which govern subsonic gas dynamics close to the surface
of the Earth are essentially the same as those that 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, Equa-
tions (1.105) and (1.112)(1.115) yield Re, Fr
1 and Ma, Pr, p0 O(1). It follows
that the final two terms on the right-hand sides of Equations (1.103) and (1.104) can
be neglected. Thus, the (unnormalized) compressible ideal gas flow equations reduce
Mathematical Models of Fluid Motion 23

to the following set of inviscid, adiabatic, ideal gas, flow equations:

D
= v, (1.120)
Dt
Dv p
= , (1.121)
Dt


D p
= 0. (1.122)
Dt

In particular, if the initial distribution of p/ is uniform in space, as is often the


case, then Equation (1.122) 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
(Reif 1965)



cV p
S= ln . (1.123)
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 for which this
is the case is termed homentropic. Equation (1.122) ensures that the entropy of a co-
moving gas element is a constant of the motion in transonic flow. A gas for which this
is the case is termed isentropic. In the homentropic case, the previous compressible
gas flow equations simplify somewhat to give

D
= v, (1.124)
Dt
Dv p
= , (1.125)
Dt


p
= . (1.126)
p0 0

Here, p0 is atmospheric pressure, and 0 is the density of air at atmospheric pressure.


Equation (1.126) 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 eect on the temperature
distribution within the gas. Incidentally, a gas in which thermal diusion is negligible
is generally termed adiabatic.
24 Theoretical Fluid Mechanics

1.18 Fluid Equations in Cartesian Coordinates

Let us adopt the conventional Cartesian coordinate system, (x, y, z). According to
Equation (1.26), the various components of the stress tensor are

v x
xx = p + 2 , (1.127)
x
vy
yy = p + 2 , (1.128)
y
vz
zz = p + 2 , (1.129)
z


v x vy
xy = yx = + , (1.130)
y x


v x vz
xz = zx = + , (1.131)
z x


vy vz
yz = zy = + , (1.132)
z y

where v is the velocity, p the pressure, and the viscosity. The equations of com-
pressible fluid flow, (1.87)(1.89) (from which the equations of incompressible fluid
flow can easily be obtained by setting = 0), become

D
= , (1.133)
Dt


Dv x 1 p 2 1
= + vx + , (1.134)
Dt x x 3 x


Dvy 1 p 2 1
= + vy + , (1.135)
Dt y y 3 y


Dvz 1 p 2 1
= + vz + , (1.136)
Dt z z 3 z



1 D p D M 2 p
=+ , (1.137)
1 Dt Dt R
Mathematical Models of Fluid Motion 25

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
= + + , (1.138)
x y z
D
= + vx + vy + vz , (1.139)
Dt t x y z
2 2 2
2 = + + , (1.140)
x 2 y 2 z 2



2
2
v x 2 vy 2 vz 1 v x vy
= 2 + + + +
x y z 2 y x

2
2
1 v x vz 1 vy vz
+ + + + . (1.141)
2 z x 2 z y

Here, , , , and M are treated as uniform constants.

1.19 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

vr
rr = p + 2 , (1.142)
r


1 v vr
= p + 2 + , (1.143)
r r
vz
zz = p + 2 , (1.144)
z


1 vr v v
r = r = + , (1.145)
r r r


vr vz
rz = zr = + , (1.146)
z r


1 vz v
z = z = + , (1.147)
r z
26 Theoretical Fluid Mechanics

whereas the equations of compressible fluid flow become

D
= , (1.148)
Dt
Dvr v2 1 p
=
Dt r r r


2 vr 2 v 1
+ vr 2 2 + , (1.149)
r r 3 r
Dv vr v 1 p 1
+ =
Dt r r r


2 2 vr v 1
+ v + 2 + , (1.150)
r r 2 3 r


Dvz 1 p 2 1
= + vz + , (1.151)
Dt z z 3 z



1 D p D M 2 p
=+ , (1.152)
1 Dt Dt R

where

1 (r vr ) 1 v vz
= + + , (1.153)
r r r z
D v
= + vr + + vz , (1.154)
Dt t r r z


1 1 2 2
2 = r + 2 + 2, (1.155)
r r r r 2 z

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

2
2
1 vr vz 1 v 1 vz
+ + + + .
r
(1.156)
2 z r 2 z
Mathematical Models of Fluid Motion 27

1.20 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
rr = p + 2 , (1.157)
r


1 v vr
= p + 2 + , (1.158)
r r


1 v vr cot v
= p + 2 + + , (1.159)
r sin r r


1 vr v v
r = r = + , (1.160)
r r r


1 vr v v
r = r = + , (1.161)
r sin r r


1 v 1 v cot v
= = + , (1.162)
r sin r r

whereas the equations of compressible fluid flow become

D
= , (1.163)
Dt

Dvr v + v
2 2
1 p 2 2 vr 2 v
= + vr 2 2
Dt r r r r r

2 cot v 2 v 1
2 + , (1.164)
r2 r sin 3 r

Dv vr v cot v
2
1 p 1 2 2 vr
+ = + v + 2
Dt r r r r

v 2 cot v 1
+ , (1.165)
r 2 sin2 r 2 sin 3 r

Dv vr v + cot v v 1 p 1 2 v
+ = + v
Dt r r sin r sin r sin2
2

2 vr 2 cot v 1
+ + + , (1.166)
r 2 sin2 r 2 sin 3 r sin



1 D p D M 2 p
=+ , (1.167)
1 Dt Dt R
28 Theoretical Fluid Mechanics

where
1 (r 2 vr ) 1 (sin v ) 1 v
= + + , (1.168)
r 2 r r sin r sin
D v v
= + vr + + , (1.169)
Dt t r r r sin



1 2 1 1 2
2 = 2 r + 2 sin + , (1.170)
r r r r sin r 2 sin2 2

2
2
2
vr 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
+ + . (1.171)
2 r sin r r

1.21 Exercises
1.1 Equations (1.66), (1.75), and (1.87) 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
= v,
Dt
q = T,

where v is the fluid velocity, T the temperature, and the thermal conductiv-
ity. According to a standard theorem in thermodynamics (Reif 1965),
p
T dS = dE d,
2

where S is the entropy per unit mass. Moreover, the entropy flux density at a
given point in the fluid is (Hazeltine and Waelbroeck 2004)
q
s = Sv + ,
T
Mathematical Models of Fluid Motion 29

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 thermodynam-
ics.

1.2 The Navier-Stokes equation for an incompressible fluid of uniform mass den-
sity 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 con-
straint requires that
v = 0.
Finally, the quantity
v
is generally referred to as the fluid vorticity.
Derive the following vorticity evolution equation from the Navier-Stokes
equation:
D
= ( ) v + 2 .
Dt
1.3 Consider two-dimensional incompressible fluid flow. Let the velocity field
take the form
v = v x (x, y, t) e x + vy (x, y, t) ey.
Demonstrate that the equations of incompressible fluid flow (see Exercise 1.2)
can be satisfied by writing


vx = ,
y

vy = ,
x
30 Theoretical Fluid Mechanics

where

+ [, ] = 2 ,
t
and
= 2 .
Here, [A, B] ez A B, and 2 = 2 /x 2 + 2 /y 2 . Furthermore, the
quantity is termed a stream function, because v = 0. In other words,
the fluid flow is everywhere parallel to contours of .
1.4 Consider incompressible irrotational flow: that is, flow that satisfies

v = 0,

as well as
Dv p
= + 2 v,
Dt
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 previous 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,
because the velocity field is derived from a scalar potential.
1.5 The equations of inviscid adiabatic ideal gas flow are
D
= v,
Dt
Dv p
= ,
Dt


D p
= 0.
Dt
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: that is,
p/t = /t = 0.
Mathematical Models of Fluid Motion 31

Demonstrate that
1 2 p
H= v + +
2 1
is a constant of the motion. In other words, DH/Dt = 0. This result is known
as Bernoullis theorem.
1.6 The equations of inviscid adiabatic non-ideal gas flow are
D
= v,
Dt
Dv p
= ,
Dt
DE p D
= 0.
Dt 2 Dt
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: that is, p/t =
/t = 0. Demonstrate that
1 2 p
H= v +E+ +
2
is a constant of the motion. In other words, DH/Dt = 0. This result is a more
general form of Bernoullis theorem.
1.7 Demonstrate that Bernoullis theorem for incompressible, inviscid fluid flow
takes the form DH/Dt = 0, where
1 2 p
H= v + + .
2
32 Theoretical Fluid Mechanics
2
Hydrostatics

2.1 Introduction
This chapter discusses the mechanical equilibrium of incompressible fluids, a topic
that is conventionally termed hydrostatics. More information on the fundamentals of
hydrostatics can be found in Lamb 1928. More information on Maclaurin spheroids,
Jacobi ellipsoids, and Roche ellipsoids can be found in Chandrasekhar 1969, and
Lamb 1993.

2.2 Hydrostatic Pressure


Consider a body of water which 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 (1.79), the air/water
system remains in mechanical equilibrium (i.e., v = Dv/Dt = 0) provided

0 = p + , (2.1)

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) = , (2.2)
0 z0
where 0 is the (approximately uniform) mass density of water. Here, the compara-
tively small mass density of air has been neglected. Because = (z) and = (z),
it immediately follows, from Equation (2.1), that p = p(z), where

dp
= g. (2.3)
dz
We conclude that constant pressure surfaces in a stationary body of water take the
form of horizontal planes. Making use of Equation (2.2), the previous equation can

33
34 Theoretical Fluid Mechanics

be integrated to give

p0 z>0
p(z) = , (2.4)
p0 0 g z z0

where p0 105 N m2 is atmospheric pressure at ground level (Batchelor 2000).


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 (Batchelor 2000), we deduce that hydrostatic pressure in water
rises at the rate of 1 atmosphere (i.e., 105 N m2 ) every 10.2 m increase in depth
below the surface.

2.3 Buoyancy
Consider the air/water system described in the previous section. Let V be some vol-
ume, bounded by a closed surface S , that straddles the plane z = 0, and is thus par-
tially 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 1.3)
 
fi = i j dS j + Fi dV, (2.5)
S V

where
i j = p i j (2.6)
is the stress tensor for a static fluid (see Section 1.5), and
F = g ez (2.7)
the gravitational force density. (Recall that the indices 1, 2, and 3 refer to the x-, y-,
and z-axes, respectively. Thus, f3 fz , et cetera.) The first term on the right-hand
side of Equation (2.5) 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 (2.5)(2.7) yield the
following expression for the net force:
f = B + W, (2.8)
where 
p
Bi = dV, (2.9)
V xi
and
W x = Wy = 0, (2.10)

Wz = g dV. (2.11)
V
Hydrostatics 35

Here, B is the net surface force, and W the net volume force.
It follows from Equations (2.4) and (2.9) that

B = M0 g ez , (2.12)

where M0 = 0 V0 . Here, V0 is the volume of that part of V which lies below the
waterline, and M0 the total mass of water contained within V. Moreover, from Equa-
tions (2.2), (2.10), and (2.11),

W = M0 g ez. (2.13)

It can be seen that the net surface force, B, is directed vertically upward, and ex-
actly 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 wa-
ter 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.

2.4 Equilibrium of Floating Bodies


Consider the situation described in the previous section. Suppose that the fluid con-
tained 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
[because the force balance criterion (2.3) can be integrated to give the pressure dis-
tribution (2.4) 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 (because 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: that is, the water
that previously occupied V. In other words, from Equation (2.12),

B = W0 ez , (2.14)

where W0 = M0 g and M0 are the weight and mass of the displaced water, respec-
tively. 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.
36 Theoretical Fluid Mechanics

A floating body of weight W is acted upon by two forces: namely, its own weight,

W = W ez , (2.15)

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). So as 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, after Archimedes of Syracuse (c.
278212 BCE). The requirement of zero net torque implies that, in equilibrium, the
center of gravity, G, and center of buoyancy, H, of a floating body must 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, (2.16)
V

where

s= (2.17)
0

is termed the bodys specific gravity. (Recall, that V0 is the submerged volume, and
0 the mass density of water.) We conclude, from Equation (2.16), 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, because the submerged vol-
ume 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 (2.16), 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: that is, V0 V1 . According to Equation (2.16), the body can only remain
in equilibrium in this configuration if its specific gravity is

V1 V V0
s = = = 1 s. (2.18)
V V

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.
Hydrostatics 37

2.5 Vertical Stability of Floating Bodies


Consider a floating body of weight W that, in equilibrium, has a submerged volume
V0 . Thus, the bodys downward weight is balanced by the upward buoyancy force,
B = 0 V0 g: that is, 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 , (2.19)
g dt 2 0
which reduces to the simple harmonic equation

d 2 1 g
= 1 . (2.20)
dt 2 0
We conclude that if a floating body of mean draft 0 is subject to a small vertical
displacement then it oscillates about its equilibrium position at the characteristic fre-
quency 
g
= . (2.21)
0
It follows that such a body is unconditionally stable to small vertical displacements.
Incidentally, the previous calculation neglects the phenomenon of added mass, by
which some of the water surrounding the floating body oscillates in sympathy with
it, thereby increasing the bodys eective inertia. (See Sections 5.9 and 7.10.)

2.6 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 perpen-
dicular 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 passes through the origin. Let GH be that, originally
38 Theoretical Fluid Mechanics

M M

H H H H

STABLE UNSTABLE
Figure 2.1
Stable and unstable configurations for a floating body.

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 passing through H  meet GH at M. (See Figure 2.1.) 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 2.1, the weight and the buoyancy force
generate a net torque of magnitude = W sin . Here, is the length MG: that
is, 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 2.1,
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: that is, provided that the metacenter
lies above the center of gravity. Because we have already demonstrated that a floating
body is unconditionally stable to small vertical displacements (and as 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 sucient condition for the stability of a floating body to
a general small perturbation (made up of arbitrary linear and angular components) is
Hydrostatics 39

that its metacentric height be positive for angular displacements about any horizontal
axis.

2.7 Determination of Metacentric Height


Suppose that the floating body considered in the previous section is in an equilibrium
state. Let A0 be the cross-sectional area at the waterline: that is, in the plane z = 0.
Because the body is assumed to be symmetric with respect to the x = 0 and y = 0
planes, we have
  
x dx dy = y dx dy = x y dx dy = 0, (2.22)
A0 A0 A0

where the integrals are taken over the whole cross-section at z = 0. Let (x, y) be
the bodys draft: that is, 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
  
V0 = dx dy dz = (x, y) dx dy. (2.23)
A0 0 A0

It also follows from symmetry that


 
x (x, y) dx dy = y (x, y) dx dy = 0. (2.24)
A0 A0

The depth of the unperturbed center of buoyancy below the surface of the water is
 

A 0
z dx dy dz 1 2 A0
h= 0 = 2 (x, y) dx dy = 0 , (2.25)
V0 2 V0 A 0 2 V0
where  1/2
A 2 (x, y) dx dy
0 = 0 . (2.26)
A0

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
 

V0 = [(x, y) y] dx dy = V0 + y dx dy = V0 , (2.27)
A0 A0

where use has been made of Equations (2.22) and (2.23). Thus, the submerged vol-
ume is unchanged, as should be the case for a purely angular displacement. The new
40 Theoretical Fluid Mechanics

depth of the center of buoyancy is


  
z dx dy dz   
 A0 0 1
h = = 2 (x, y) 2 y (x, y) + O( 2 ) dx dy = h,
V0 2 V0 A0
(2.28)
where use has been made of Equations (2.24) and (2.25). 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
   
A0 0
y dx dy dz A0
y [(x, y) y] dx dy x2 A0
= = , (2.29)
V0 V0 V0

where use has been made of Equation (2.24). Here,


 2 1/2
A y dx dy
x = 0 ,
(2.30)
A0

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


It follows, from the previous analysis, that if the floating body under considera-
tion 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 2.1 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 buoy-
ancy. 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 metacen-
ter 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
zM = + zH , (2.31)
V0

and the metacentric height, = z M zG , becomes

x2 A0
= zGH , (2.32)
V0

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

y2 A0
= zGH , (2.33)
V0
Hydrostatics 41

where
 2 1/2
A x dx dy
y = 0 ,
(2.34)
A0

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 , (2.35)
V0
where
2 = x2 cos2 + y2 sin2 . (2.36)

Thus, the minimum value of 2 is the lesser of x2 and y2 . It follows that the equilib-
rium state in question is unconditionally stable provided it is stable to small ampli-
tude 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, respec-
tively. 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 (2.25).]
Moreover, by symmetry, the center of gravity is a height c/2 above the bottom sur-
face 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
 a/2  b/2
a/2 b/2
y 2 dx dy b2
x2 = = . (2.37)
ab 12
Hence, from Equation (2.32), the metacentric height is

b2 c
= (1 s). (2.38)
12 s c 2
The stability criterion > 0 yields

b2
s (1 s) > 0. (2.39)
6 c2
Because the maximum value that s (1 s) can take is 1/4, it follows that the block is
stable for all specific gravities when

2
c < c0 = b. (2.40)
3
42 Theoretical Fluid Mechanics

On the other hand, if c > c0 then the block is unstable for intermediate specific
gravities such that s < s < s+ , where

1 1 c02 /c 2
s = , (2.41)
2
and is stable otherwise. Assuming that the block is stable, its angular equation of
motion is written
d 2
I 2 = W sin W , (2.42)
dt
where
 a/2  b/2  cs c
W W  2 
I= (y 2 + z 2 ) dx dy dz = b + 4 [(1 s)3 + s 3 ] c 2
g V a/2 b/2 s c 12 g
(2.43)
is the moment of inertia of the block about the x-axis. Thus, we obtain the the simple
harmonic equation
d 2
= 2 , (2.44)
dt 2
where
W g c02 4 s (1 s) c 2
2 = = . (2.45)
I s c c02 + (8/3) [(1 s)3 + s 3 ] c 2
We conclude that the block executes small amplitude angular oscillations about the x-
axis at the angular frequency . For the case of rotation about the y-axis, the previous
analysis is unchanged except that a b. The previous analysis again neglects the
phenomenon of added mass, and, therefore, underestimates the eective inertia of
the block. (See Sections 5.9 and 7.10.)
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 per-
pendicular to the ships length), because the radius of gyration for pitching greatly
exceeds that for rolling. As is clear from Equation (2.45), a ship with a relatively
small metacentric height (for rolling) has a relatively long roll period, and vice versa.
An excessively low metacentric height increases the chances of a ship capsizing if
the weather is rough, if its cargo/ballast shifts, or if the ship is damaged and partially
flooded. For this reason, maritime regulatory agencies, such as the International Mar-
itime Organization, specify minimum metacentric heights for various dierent types
of sea-going vessel. A relatively large metacentric height, on the other hand, gener-
ally 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 2.4, that if a body of specific gravity s floats in verti-
cal 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 shall now
demonstrate that these positions are either both stable, or both unstable, provided the
Hydrostatics 43

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
V 1 H1 G = V 2 H2 G = H1 H2 , (2.46)
V1 + V2

where H1G is the distance between points H1 and G, et cetera. The metacentric
heights in the first and second positions are


2 A 1 V1 V2
1 = H1 G = A 2 H1 H2 , (2.47)
V1 V1 V1 + V2


2 A 1 V1 V2
2 = H2 G = A
2
H1 H2 , (2.48)
V2 V2 V1 + V2
respectively, where A and are the area and radius of gyration of the common water-
line section, respectively. Thus,

2
1 V1 V2
1 2 = A
2
H1 H2 0, (2.49)
V1 V2 V1 + V2

which implies that 1  0 as 2  0, and vice versa. It follows that the first and
second positions are either both stable, both marginally stable, or both unstable.

2.8 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 eectively 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. 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 (Fitzpatrick 2012). However, such
44 Theoretical Fluid Mechanics

r
p
H0
H
B

Figure 2.2
Curve of buoyancy for a floating body.

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.

2.9 Curve of Buoyancy


Consider a floating body in vertical force balance that is slowly rotated about a hor-
izontal 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 2.2, 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 buoy-
ancy to shift horizontally, rather than vertically, in the plane perpendicular to the axis
of rotation. Thus, the dierence in vertical height, zGH , between the center of gravity
Hydrostatics 45

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. If R is the
radius of curvature of the curve of buoyancy then, according to a standard result in
dierential calculus (Lamb 1928),

dr
R=r . (2.50)
dp

Writing this result in the form


R
r = p, (2.51)
r
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, coin-
cident 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: that is,
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 corre-
sponding values of r, p, and R. For neighboring points on the curve, we can write

dr 
r r0 =  (p p0 ), (2.52)
d p  H0

or
R0
r r0 = (p p0 ). (2.53)
r0
It follows that p p0 has the same sign as r r0 (because 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 simultane-
ously with those of p. Consequently, a stable equilibrium configuration corresponds
to a point of minimum r on the curve of buoyancy: that is, a minimum in the distance
GH between the center of gravity and the center of buoyancy.
We can use the previous 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, because, according to the analysis of
Section 2.6, 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
46 Theoretical Fluid Mechanics

O
x
Q

P Q
P

A B
y

Figure 2.3
Beam of square cross-section floating with two corners immersed.

case where the waterline PQ is parallel to a side. (See Figure 2.3.) If the length of a
side is 2 a then Equation (2.16) yields
AP = BQ = 2 a s. (2.54)

Suppose that the beam is turned through an angle > 0 such that the waterline
assumes the position P Q in Figure 2.3, but still intersects two opposite sides. The
lengths AP and BQ satisfy
BQ AP = 2 a tan . (2.55)

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): that is,

(AP + BQ ) a = 4 a 2 s. (2.56)


It follows that
AP = a (2 s tan ), (2.57)

BQ = a (2 s + tan ). (2.58)

The constraint that the waterline intersect two opposite sides of the beam implies that
AP > 0, and, hence, that
tan < 2 s. (2.59)
Hydrostatics 47

The coordinates of the center of buoyancy, H, which is the mean center of the trapez-
ium P ABQ , are
a a
a ah(x)
x dx dy (2/3) a 3 tan a
x =  a  a = 2
= tan , (2.60)
dx dy 4a s 6 s
a ah(x)
a a
a ah(x)
y dx dy
y =  a  a
a ah(x)
dx dy
4 a 3 s (1 s) (1/3) a 3 tan2 a
= = (1 s) a tan2 , (2.61)
4 a2 s 12 s
where
h(x) = 2 a s + x tan . (2.62)
Thus, if u = r 2 /a 2 = ( x 2 + y 2 )/a 2 then
2
t2 t2
u= + (1 s) , (2.63)
36 s 2 12 s

where t = tan . A stable equilibrium state corresponds to a minimum of r with


respect to , and, hence, of u with respect to t. However,

du t  
= t 2 12 s (1 s) + 2 , (2.64)
dt 36 s 2
2
d u 1  
2
= 3 t 2 12 s (1 s) + 2 . (2.65)
dt 36 s 2
The minima and maxima of u occur when du/dt = 0, d 2 u/dt 2 > 0 and du/dt = 0,
d 2 u/dt 2 < 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

1
s2 s + >0: (2.66)
6

that is, 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

t 2 = 12 s (1 s) 2. (2.67)

Such equilibria only exist for s > 0.2113, and are stable. Finally, in order to satisfy
the constraint (2.59), we must have t < 2 s, which, in combination with the previous
equation, implies that
8 s 2 6 s + 1 > 0, (2.68)
or s < 0.25.
48 Theoretical Fluid Mechanics

Q

O
x



A P B
y

Figure 2.4
Beam of square cross-section floating with one corner immersed.

Suppose that the constraint (2.59) is not satisfied, so that the immersed portion of
the beams cross-section is triangular. (See Figure 2.4.) It is clear that

BQ
= tan . (2.69)
BP
Moreover, the area of the triangle P BQ in Figure 2.4 must match that of the rect-
angle PABQ in Figure 2.3, in order to ensure that the submerged volume remain
invariant: that is,
1
BP BQ = 4 a 2 s. (2.70)
2
It follows that

BP = (8 s/ tan )1/2 a, (2.71)


BQ = (8 s tan )1/2 a, (2.72)

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

BP = 3 z 1 a, (2.73)
BQ = 3 z a. (2.74)

The coordinates of the center of buoyancy, H, which is the mean center of triangle
Hydrostatics 49

P BQ , are

x = a BP /3 = a (1 z 1 ), (2.75)

y = a BQ /3 = a (1 z), (2.76)

because 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 = r 2 /a 2 = ( x 2 + y 2 )/a 2 then

u = (1 z1 )2 + (1 z)2 , (2.77)
du 2 2 (z 2 1) (z 2 1 z + 1)
= , (2.78)
dz z3
d 2 u 2 2 (z 4 2 1 z + 3)
= . (2.79)
dz 2 z4
Moreover, the constraint (2.59) yields
3
z> . (2.80)
2
The stable and unstable equilibria correspond to du/dz = 0, d 2 u/dz 2 > 0 and du/dz,
d 2 u/dz 2 < 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 2 1 z + 1 = 0. (2.81)

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.

2.10 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
50 Theoretical Fluid Mechanics

appears stationary in a non-inertial co-rotating reference frame. Moreover, according


to standard Newtonian dynamics (Fitzpatrick 2012), the force balance equation for
the fluid in the co-rotating frame takes the form (cf., Section 2.2)

0 = p + + ( r), (2.82)

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 previous
equation represents the fictitious centrifugal force density. Without loss of generality,
we can assume that = ez . It follows that

0 = p + ( +  ), (2.83)

where
1
 = 2 (x 2 + y 2 ) (2.84)
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 r 2 . Assuming that the pressure distribution is
axisymmetric, so that p = p(r, z), the force balance equation, (2.83), reduces to

p ( +  )
+ = 0, (2.85)
r r
p ( +  )
+ = 0, (2.86)
z z
or
p
2 r = 0, (2.87)
r
p
+ g = 0. (2.88)
z
The previous two equations can be integrated to give


1 2 2
p(r, z) = p0 + r g z , (2.89)
2

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 : that is, it is the constant pressure surface whose
Hydrostatics 51

pressure matches that of the atmosphere. It follows that the surface of the water is
the paraboloid of revolution
2 2
z= r , (2.90)
2g
where r is the perpendicular distance from the axis of rotation, and z = 0 the on-axis
height of the surface.
According to the analysis of Section 2.3, 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
(Fitzpatrick 2012), the equation of horizontal motion of a general co-rotating body is
..
M (r 2 rG ) = M0 2 rH , (2.91)
.
where = d/dt. 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 (rH rG ) < 0. (2.92)

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 M0 ) 2 rG > 0. (2.93)

In other words, a fully submerged body drifts radially outward from the rotation axis.
The previous 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, while floating objects, such as sticks, tend to collect on the inner side.

2.11 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
52 Theoretical Fluid Mechanics

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 New-
ton, Maclaurin, Jacobi, Meyer, Liouville, Dirichlet, Dedekind, Riemann, and other
celebrated scientists, in the 17th, 18th, and 19th centuries (Chandrasekhar 1969).
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 sucient to allow the actual shape of the Earth to
deviate significantly from that of a hypothetical liquid Earth (Fitzpatrick 2012).
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 ficti-
tious 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. An ellipsoid is the three-
dimensional generalization of an ellipse. Let us adopt the right-handed Cartesian
coordinate system x1 , x2 , x3 . An ellipse whose principal axes are aligned along the
x1 - and x2 -axes satisfies

x12 x22
+ = 1, (2.94)
a12 a22

where a1 and a2 are the corresponding principal radii. Moreover, as is easily demon-
strated,

A= dA = a1 a2 , (2.95)

1 2
xi2 dA = a A, (2.96)
4 i

x1 x2 dA = 0, (2.97)

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 x22 x32


+ + = 1, (2.98)
a12 a22 a32

where a1 , a2 , and a3 are the corresponding principal radii. Moreover, as is easily


Hydrostatics 53

demonstrated,

4
V= dV = a1 a2 a3 , (2.99)
3

1 2
xi2 dV =a V, (2.100)
5 i
 
x1 x2 dV = x2 x3 dV = 0, (2.101)

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 (2.100)]
1
I33 = M (a12 + a22 ), (2.102)
5
where M is its mass. Thus, the planets angular momentum is
1
L = I33 = M (a12 + a22 ) , (2.103)
5
and its rotational kinetic energy becomes
1 1
K= I33 2 = M (a12 + a22 ) 2 . (2.104)
2 10
According to Equations (2.83) and (2.84), the fluid pressure distribution within
the planet takes the form

1
p = p0 2 (x12 + x22 ) , (2.105)
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 D that the gravitational po-
tential inside a homogeneous self-gravitating ellipsoidal body can be written (Chan-
drasekhar 1969; Lamb 1993)

3 
= G M 0 i xi ,
2
(2.106)
4 i=1,3

where G is the universal gravitational constant (Yoder 1995), and



du
0 = , (2.107)
0

du
i = , (2.108)
0 (ai + u)
2

= (a12 + u)1/2 (a22 + u)1/2 (a32 + u)1/2 . (2.109)


54 Theoretical Fluid Mechanics

Thus, we obtain



1 3 3 3
p = p0 G M 1 2 x12 + G M 2 2 x22 + G M 3 x32 , (2.110)
2 2 2 2
where p0 is the central fluid pressure. 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



1 3 3 3
G M 1 x1 + G M 2 x2 + G M 3 x3 = p0 , (2.111)
2 2 2 2 2
2 2 2 2
whenever
x12 x22 x32
+ + = 1. (2.112)
a12 a22 a32
The previous two equations can only be simultaneously satisfied if

2 2
1 a12 = 2 a 2 = 3 a32 . (2.113)
(3/2) G M (3/2) G M 2
Rearranging the previous expression, we obtain

2 a1 a3 2 u du
= (a2 a32 ) , (2.114)
2 G a2 0 (a2 + u) (a3 + u)
2 2

subject to the constraint





a12 a22 du a32
(a12 a22 ) 2 = 0, (2.115)
0 (a1 + u) (a2 + u) (a3 + u)
2 2

where use has been made of Equation (2.99).


Finally, according to Appendix D, the net gravitational potential energy of the
planet is
3
U = G M 2 0 . (2.116)
10
Hence, the bodys total mechanical energy becomes
1 3
E = K+U = M (a12 + a22 ) 2 G M 2 0 . (2.117)
10 10

2.12 Maclaurin Spheroids


One, fairly obvious, way in which the constraint (2.115) can be satisfied is if a2 = a1 .
In other words, if the planet is rotationally symmetric about its axis of rotation. An
ellipsoid that is rotationally symmetric about a principal axisor, equivalently, an
Hydrostatics 55

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 : that is, 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 . (2.118)
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.
Let u = a12 and = e132
/z 2 1. Setting a2 = a1 in Equation (2.114), we obtain

2 d
= (1 e13 ) e13
2 1/2 2
2 G 0 (1 + ) 2 (1 + e 2 )3/2
13
2 1/2  e13  e13
2 (1 e13 ) 2
z dz z 2 dz
= (1 e13 )
2
. (2.119)
3
e13 0 (1 z 2 )1/2 0 (1 z 2 )3/2
Performing the integrals, which are standard (Speigel, Liu, and Lipschutz 1999), we
find that
2
2 3 2 e13 3
=
(1 e 2 1/2
) sin 1
e 13 2 (1 e13
2
). (2.120)
2 G 3
e13 13
e13
This famous result was first obtained by Colin Maclaurin (1698-1746) in 1742. Fi-
nally, in order to calculate the potential energy, (2.116), we need to evaluate
  e13
1 d 2 dz
0 = =
a1 0 (1 + ) (1 + e13 )2 1/2 a1 e13 0 (1 z 2 )1/2
2 sin1 e13
= . (2.121)
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 to define =
/(2 G )1/2, L = L/(G M 3 a0 )1/2 , and E = E/(G M 2 /a0 ). The previous analysis
leads to the following set of equations that specify the properties of the so-called
Maclaurin spheroids:

cos
=
2
(1 + 2 cos )
2
3 cos , (2.122)
sin2 sin
1/2
6 (cos )1/6
L= (1 + 2 cos2 ) 3 cos , (2.123)
5 sin sin

3 (cos )1/3
E= (1 4 cos )
2
+ 3 cos . (2.124)
10 sin2 sin
56 Theoretical Fluid Mechanics

e13 L E e13 L E
0.00 0.00000 0.00000 0.60000 0.60 0.31729 0.18037 0.56233
0.05 0.02582 0.01266 0.59980 0.65 0.34484 0.20286 0.55320
0.10 0.05168 0.02540 0.59919 0.70 0.37239 0.22834 0.54200
0.15 0.07758 0.03830 0.59817 0.75 0.39967 0.25792 0.52800
0.20 0.10357 0.05144 0.59672 0.80 0.42612 0.29345 0.51001
0.25 0.12967 0.06491 0.59479 0.85 0.45046 0.33833 0.48587
0.30 0.15591 0.07882 0.59236 0.90 0.46932 0.39994 0.45107
0.35 0.18231 0.09329 0.58936 0.95 0.47045 0.50074 0.39272
0.40 0.20889 0.10846 0.58572 0.96 0.46472 0.53194 0.37485
0.45 0.23567 0.12450 0.58135 0.97 0.45418 0.57123 0.35273
0.50 0.26267 0.14163 0.57612 0.98 0.43475 0.62486 0.32351
0.55 0.28989 0.16013 0.56986 0.99 0.39389 0.71209 0.27916

Table 2.1
Properties of the Maclaurin spheroids.

These properties are set out in Table 2.1.


In the limit, 0, in which the planet is relatively slowly rotating (i.e.,  1),
and its degree of flattening consequently slight, Equations (2.122)(2.124) reduce to

15
e13 , (2.125)
2

6
L , (2.126)
5
3
E . (2.127)
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
m= = , (2.128)
g0 2
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
= . (2.129)
a0 2
Thus, it follows from (2.125) that
5
 m. (2.130)
4
Hydrostatics 57

0.2

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 2.5
Normalized angular velocity squared of a Maclaurin spheroid (solid) and a Jacobi
ellipsoid (dashed) versus the eccentricity e13 in the x1 -x3 plane.

For the case of the Earth ( = 7.27 105 rad. s1 , a0 = 6.37 106 m, g0 =
9.81 m s1Yoder 1995), we obtain
1
m . (2.131)
291
Thus, it follows that, were the Earth homogeneous, its figure would be a spheroid,
flattened at the poles, of ellipticity

5 1 1
 . (2.132)
4 291 233
This result was first obtained by Newton. The actual ellipticity of the Earth is about
1/294 (Yoder 1995), 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 2.5 and 2.6 illustrate the variation of the normalized angular velocity,
, and angular momentum, L, of a Maclaurin spheroid with its eccentricity, e13 ,
as predicted by Equations (2.122)(2.124). It can be seen, from Figure 2.5, that
there is a limit to how large the normalized angular velocity of such a spheroid can
become. The limiting value corresponds to = 0.47399, and occurs when e13 =
0.92995. For values of lying below 0.47399 there are two possible Maclaurin
spheroids, one with an eccentricity less than 0.92995, and one with an eccentricity
58 Theoretical Fluid Mechanics

1.2
1.1
1
0.9
0.8
0.7
L 0.6
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 2.6
Normalized angular momentum of a Maclaurin spheroid (solid) and a Jacobi ellip-
soid (dashed) versus the eccentricity e13 in the x1 -x3 plane.

greater than 0.92995. Note, however, from Figure 2.6, that despite the fact that the
angular velocity, , of a Maclaurin spheroid varies in a non-monotonic manner with
the eccentricity, e13 , the angular momentum, 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.

2.13 Jacobi Ellipsoids


If a2  a1 (i.e., if the outer boundary of the rotating body is ellipsoidal, rather than
spheroidal) then the constraint (2.115) can only be satisfied when



a12 a22 a32 du
= 0. (2.133)
0 (a12 + u) (a22 + u) (a3 + u)
2

Without loss of generality, we can assume that a1 a2 a3 . Let

a2 = a1 cos , (2.134)
a3 = a1 cos , (2.135)
Hydrostatics 59

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 = (1 a32 /a22 )1/2 = sin , (2.136)


e13 = (1 a32 /a12 )1/2 = sin , (2.137)

respectively. It is also helpful to define

= sin1 (sin / sin ). (2.138)

Let sin2 = [a12 /(a12 + u)] sin2 . Here, u = 0 corresponds to = , and u =


to = 0. Equations (2.115) and (2.114) transform to (Darwin 1886)

1 + (sin tan cos )2
E(, ) 2 F(, ) + E(, )
cos2
sin tan cos (1 + sin2 )
= 0, (2.139)
cos2
and

F(, ) E(, ) cos E(, ) cos2
=2
2
+ , (2.140)
tan sin tan tan3 cos2 tan2 cos2
respectively, where

E(, ) = (1 sin2 sin2 )1/2 d, (2.141)
0

d
F(, ) = , (2.142)
0 (1 sin2 sin2 )1/2
are special functions known as incomplete elliptic integrals (Abramowitz and Stegun
1965). The integral 0 , defined in Equation (2.107), transforms to

2 (cos cos )1/3


0 = F(, ). (2.143)
a0 sin
Finally, making use of some of the analysis in the previous two sections, the nor-
malized angular momentum, and normalized mechanical energy, of the planet can be
written

6 1 + cos2
L= , (2.144)
10 (cos cos )2/3
3 (cos cos )1/3 3 1 + cos2
E= F(, ) + 2, (2.145)
5 sin 20 (cos cos )2/3
respectively.
The constraint (2.139) is obviously satisfied in the limit 0, because this
60 Theoretical Fluid Mechanics

e12 e13 L E e12 e13 L E


0.00 0.81267 0.43257 0.30375 0.50452 0.60 0.85585 0.42827 0.30984 0.50138
0.05 0.81293 0.43257 0.30375 0.50459 0.65 0.86480 0.42609 0.31296 0.49975
0.10 0.81372 0.43257 0.30375 0.50459 0.70 0.87510 0.42288 0.31760 0.49734
0.15 0.81504 0.43256 0.30377 0.50458 0.75 0.88705 0.41807 0.32462 0.49372
0.20 0.81691 0.43253 0.30380 0.50457 0.80 0.90102 0.41069 0.33562 0.48814
0.25 0.81934 0.43248 0.30388 0.50453 0.85 0.91761 0.39879 0.35390 0.47908
0.30 0.82237 0.43237 0.30402 0.50445 0.90 0.93778 0.37787 0.38783 0.46295
0.35 0.82603 0.43220 0.30427 0.50432 0.95 0.96340 0.33353 0.46860 0.42782
0.40 0.83037 0.43191 0.30468 0.50410 0.96 0.96950 0.31776 0.50078 0.41499
0.45 0.83544 0.43146 0.30532 0.50376 0.97 0.97605 0.29691 0.54672 0.39771
0.50 0.84131 0.43078 0.30628 0.50326 0.98 0.98317 0.26722 0.62003 0.37241
0.55 0.84808 0.42976 0.30772 0.50250 0.99 0.99101 0.21809 0.76872 0.32842

Table 2.2
Properties of the Jacobi ellipsoids.

implies that 0 and E(, ), F(, ) . Of course, this limit corresponds to


the axisymmetric Maclaurin spheroids discussed in the previous section. Carl Jacobi
(1804-1851), in 1834, was the first researcher to obtain the very surprising result that
Equation (2.139) 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 (2.139), (2.140), (2.144), and
(2.145), are set out in Table 2.2, and illustrated in Figures 2.5 and 2.6. 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 2.5 and 2.6, 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., = 0.43257),
but no maximum angular momentum.
Figure 2.7 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 that the corre-
sponding 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, rotat-
Hydrostatics 61

0.6

0.5

E 0.4

0.3

0.2

0.1

0
0 0.2 0.4 0.6 0.8 1 1.2
L

Figure 2.7
Normalized mechanical energy of a Maclaurin spheroid (solid) and a Jacobi ellipsoid
(dashed) versus the normalized angular momentum.

ing, liquid planet, the relevant constraints are that the mass, volume, and net angu-
lar 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 2.7, that at rela-
tively high angular momentum (i.e., L > 0.30375, e13 > 0.81267), when the Jacobi
ellipsoid solutions exist, they are stable equilibrium states (because 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., 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 out by the results of direct stability
analysis performed on the Maclaurin spheroids and Jacobi ellipsoids (Chandrasekhar
1969). 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 uncondi-
tionally stable for e13 < 0.93858, but are unconditionally unstable for e13 > 0.93858,
evolving toward lower energy pear shaped equilibria (which are, themselves, un-
stable in the presence of weak dissipation).
62 Theoretical Fluid Mechanics

2.14 Roche Ellipsoids


Consider a homogeneous liquid moon of mass M that 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 (Fitz-
patrick 2012). 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 (Fitzpatrick 2012)
G (M + M  )
2 = . (2.146)
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). 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 (Murray and Dermott 1999).
Because a synchronously rotating moon is completely stationary in the afore-
mentioned non-inertial frame, its internal pressure, p, is governed by a force balance
equation of the form [cf., Equation (2.83)]

0 = p + ( +  +  ), (2.147)

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


= x1 R + x (2.148)
2 M + M 2

the centrifugal potential due to the fact that the non-inertial frame is rotating (about
Hydrostatics 63

an axis parallel to the x3 -axis and passing through point C  ) at the angular velocity
[cf., Equation (2.84)]. Suppose that the moon is much less massive that the planet
(i.e., M/M   1). In this limit, the centrifugal potential (2.148) reduces to


1 x1 (1/2) x12 + (1/2) x22
+ , (2.149)
2 R R2

where use has been made of Equation (2.146).


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
G M  x x 2 + x22 + x32
 = 1 2 1 + 1
R R R2

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

1 + + + , (2.150)
R R R2

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


The previous two equations can be combined to give


3 2 1 2
 +  x1 x3 , (2.151)
2 2

where
G M
= , (2.152)
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 ). (2.153)

The previous 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 dierent spatial
variation of the centrifugal force and the planets gravitational force of attraction.
This dierent variation causes these two forces, which balance one another at the
center of the moon, to not balance away from the center (Fitzpatrick 2012). 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).
64 Theoretical Fluid Mechanics

Suppose that the bounding surface of the moon is the ellipsoid

x12 x22 x32


+ + = 1, (2.154)
a12 a22 a32

where a1 a2 a3 . It follows, from Appendix D, that the gravitational potential of


the moon at an interior point can be written

3 

= G M 0 i xi ,
2
(2.155)
4 i=1,3

where the integrals i , for i = 0, 3, are defined in Equations (D.30) and (D.31).
Hence, from Equations (2.147) and (2.151), the pressure distribution within the moon
is given by



1 3 3 3
p = p0 G M 1 3 x12 + G M 2 x22 + G M 3 + x32 , (2.156)
2 2 2 2

where p0 is the central pressure. 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
G M 1 3 x12 + G M 2 x22 + G M 3 + x32 = p0 , (2.157)
2 2 2 2

whenever
x12 x22 x32
+ + = 1. (2.158)
a12 a22 a32
The previous two equations can only be simultaneously satisfied if

3
1 a12 = 2 a22 = 3 + a 2. (2.159)
(3/2) G M (3/2) G M 3

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

2 F(, ) E(, )
1 = , (2.160)
a13 sin3 sin2

2 E(, ) F(, ) cos sin
2 = 2 , (2.161)
a1 sin3 sin2 cos2 sin2 cos2 cos

2 E(, ) cos sin
3 = 2 + , (2.162)
a1 sin3 cos2 cos2 cos
Hydrostatics 65

0.05

0.04

0.03
e13 e12

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 2.8
Properties of the Roche ellipsoids.

where the incomplete elliptic integrals E(, ) and F(, ) are defined in Equa-
tions (2.141) and (2.142), respectively. Thus, Equation (2.159) yields


1 cos2
= F(, ) (1 + cos2 ) E(, ) 1 +
sin tan tan cos2

sin sin cos cos
+ , (2.163)
cos2

subject to the constraint



cos2 sin sin cos cos
0 = cos F(, ) 2 E(, ) + E(, )
2

cos2 cos2

cos2
+ (3 + cos2 ) E(, ) + [F(, ) cos2 2 E(, )]
cos2

2 sin sin cos cos
+ , (2.164)
cos2

where
M  a03
= , (2.165)
M R3
and a0 = (a1 a2 a3 )1/3 is the mean radius of the moon. The dimensionless parameter
66 Theoretical Fluid Mechanics

e12 e13 e12 e13


0.00 0.00000 0.00000 0.52 0.56740 0.33440
0.04 0.04613 0.00213 0.56 0.60632 0.38204
0.08 0.09223 0.00852 0.60 0.64445 0.43094
0.12 0.13809 0.01913 0.64 0.68182 0.48027
0.16 0.18364 0.03392 0.68 0.71848 0.52890
0.20 0.22879 0.05282 0.72 0.75446 0.57532
0.24 0.27346 0.07573 0.76 0.78984 0.61729
0.28 0.31756 0.10253 0.80 0.82472 0.65150
0.32 0.36104 0.13308 0.84 0.85923 0.67265
0.36 0.40383 0.16721 0.88 0.89353 0.67151
0.40 0.44588 0.20470 0.92 0.92793 0.62978
0.44 0.48718 0.24528 0.96 0.96294 0.50135
0.48 0.52769 0.28865 1.00 1.00000 0.00000

Table 2.3
Properties of the Roche ellipsoids.

measures the strength of the tidal distortion field, generated by the planet, that acts
on the moon. There is an analogous parameter,

M a0
3
 = , (2.166)
M R 3
where a0 is the mean radius of the planet, which measures the tidal distortion field,
generated by the moon, that acts on the planet. 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 
 , (2.167)
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
previous condition reduces to M  M  , or, equivalently, a0 < a0 . In other words,
neglecting the tidal distortion of the planet, while retaining that of the moon, is gener-
ally 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 (2.163) and (2.164), which describe the ellipsoidal equilibria of a syn-
chronously rotating, relatively low mass, liquid moon due to the tidal force field of
the planet about which it orbits, were first obtained by Edouard Roche (1820-1883)
in 1850. The properties of the so-called Roche ellipsoids are set out in Table 2.3, and
Figures 2.8 and 2.9.
Hydrostatics 67

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 2.9
Properties of the Roche ellipsoids.

It can be seen, from Table 2.3 and Figure 2.8, 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 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
15
2
e12 e13
2
. (2.168)
2
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 (Yoder 1995). It follows that e13 = 7.50 103 , and (a1 a3 )/a1
2
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 2.3 and Figure 2.9, 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 if
this maximum strength is exceeded then the moon is tidally disrupted by the planet.
68 Theoretical Fluid Mechanics

The equilibrium condition < 0.06757 is equivalent to



 1/3
R
> 2.455 , (2.169)
a0

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 previous 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 eects. 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 planets radius (assuming that
the moon and the planet have approximately the same mass density). Of course, rel-
atively 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.

2.15 Exercises
2.1 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 aected?
(Lamb 1928.)

2.2 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
b 1
> 1 .
a s

2.3 Show that the equilibrium of a solid of uniform density floating with an edge
or corner just emerging from the water is unstable. (Lamb 1928.)

2.4 Prove that if a solid of uniform density floats with a flat face just above the
waterline then the equilibrium is stable. (Lamb 1928.)

2.5 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 rota-
tions about its axis, which means that the corresponding metacentric height
is zero.] (Lamb 1928.)

2.6 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
Hydrostatics 69

value, prove that the equilibrium is only stable when the specific gravity of
the cylinder lies outside the range

1 a 2
1 1 2 2 .
2 h

2.7 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

1 1 a2
.
2 4 h2
(Lamb 1928.)
2.8 Show that the cylinder of the preceding exercise can float with its axis hori-
zontal provided
h
> 3 sin(s ),
2a
where s is the specific gravity of the cylinder. (Lamb 1928.)
2.9 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. (Lamb 1928.)
2.10 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 eect 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 eect diminished when a longitudinal partition bisects the tank?
(Lamb 1928.)
2.11 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.
2.12 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
c 2 < a 2.
2
Show that if c 2 > (1/2) a 2 then the upright position is stable when h lies in
the range 
c c 2 a 2 /2.
70 Theoretical Fluid Mechanics

2.13 A thin cylindrical vessel of cross-sectional area A floats upright, being im-
mersed 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. (Lamb 1928.)

2.14 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. (Lamb
1928.)

2.15 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

(Lamb 1928.)

2.16 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.

2.17 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 a 2 /g < 4 h and 2 a 2 /g < 4 (H h), the net radial thrust on the
vertical walls of the vessel is

2
2 a2
a h2 g 1 + .
4gh

2.18 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
a 4 2.
4
Hydrostatics 71

2.19 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
1 3
Mg M 2 a,
4 16
respectively, where M is the mass of the liquid.
2.20 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.
2.21 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. (Batchelor 2000.)
2.22 Verify Equations (2.95)(2.97) and (2.99)(2.101).
2.23 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.
(a) Demonstrate that 
5
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.
72 Theoretical Fluid Mechanics

(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 ho-
mogeneous body with the same mass and volume as the Earth, this max-
imum velocity corresponds to a minimum rotation period of 2 h 25 m.
3
Surface Tension

3.1 Introduction
As is well known, small drops of water in air, and small bubbles of gas in water, tend
to adopt spherical shapes. This observation, and a host of other natural phenomena,
can only be accounted for on the hypothesis that an interface between two dierent
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 (Reif 1965). 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. More information on surface tension can be found in Lamb 1928, and
Batchelor 2000.
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). Because this range is so small, the number of molecules in a
macroscopic system whose free energies are aected 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

73
74 Theoretical Fluid Mechanics

2
S
C t
1
tn

Figure 3.1
Interface between two immiscible fluids.

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
(Batchelor 2000). The surface tension at most oil/air interfaces is much lower
typically, 2102 N m1 (Batchelor 2000). 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 (Batchelor
2000).
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 af-
fected 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.

3.2 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 3.1.)
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.
Surface Tension 75

The net force acting on S is


 
f = (p1 p2 ) n dS + t n dr, (3.1)
S C

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 dierence
across the interface, whereas the second term is the net surface tension force. Note
that body forces play no role in Equation (3.1), because the interface has zero vol-
ume. Furthermore, viscous forces can be neglected, because both fluids are static. In
equilibrium, the net force acting on S must be zero: that is,
 
(p1 p2 ) n dS = t n dr. (3.2)
S C

(In fact, the net force would be zero even in the absence of equilibrium, because the
interface has zero mass.)
Applying the curl theorem (see Section A.22) to the curve C, we find that
 
F dr = F dS, (3.3)
C S

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


 
F t dr = F n dS . (3.4)
C S

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


 
(g b) t dr = (g b) n dS . (3.5)
C S

However, the vector identity (A.179) yields

(g b) = ( g) b + (b ) g, (3.6)

as b is a constant vector. Hence, we get


 
b t g dr = b [(g) n ( g) n] dS , (3.7)
C S

where b (g) n bi (g j /xi ) n j . Now, because b is also an arbitrary vector, the


previous equation gives
 
! "
t g dr = (g) n ( g) n dS . (3.8)
C S

Taking g = n, we find that


 
t n dr = [(n) n ( n) n] dS . (3.9)
C S
76 Theoretical Fluid Mechanics

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


 
t n dr = ( n) n dS , (3.10)
C S

which can be combined with Equation (3.2) to give



! "
(p1 p2 ) ( n) n dS = 0. (3.11)
S

Finally, given that S is arbitrary, the previous expression reduces to the pressure
balance constraint
p = n, (3.12)
where p = p1 p2 . The previous relation is generally known as the Young-Laplace
equation, and is named after Thomas Young (1773-1829), who developed the quali-
tative theory of surface tension in 1805, and Pierre-Simon Laplace (1749-1827) who
completed the mathematical description in the following year. The Young-Laplace
equation can also be derived by minimizing the free energy of the interface. (See
Section 3.8.) 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.

3.3 Spherical Interfaces


Generally speaking, the equilibrium shape of an interface between two immiscible
fluids is determined by solving the force balance equation (2.1) 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 fluidfor example,
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 dierence in
densities on the two sides of the interface, is suciently small, we can safely ignore
the eect 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
Equation (3.12), the mean curvature n of the interface is also constant. Because
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. Thus, the equilibrium
shape of the interface is that which has the least surface area for a given volume: in
other words, a sphere.
Surface Tension 77

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 Equation (C.65), 
1 r 2  2
n = 2  = . (3.13)
r r r=R R
The Young-Laplace equation, (3.12), then gives

2
p = . (3.14)
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: for instance, 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 dierence is twice that across
a single interface.

3.4 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 constant. On the
other hand, the pressure in the liquid varies as p = p0 g z (see Chapter 2), 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, (3.12), yields
g z = n, (3.15)
where n is the normal to the interface directed from liquid to air. If R represents
the typical radius of curvature of the interface then the left-hand side of the previous
equation dominates the right-hand side whenever R
l, and vice versa. Here,

1/2

l= (3.16)
g

is known as the capillary length, and takes the value 2.7 103 m for pure water
at 20 C (Batchelor 2000). We conclude that the eect of surface tension on the
shape of an liquid/air interface is likely to dominate the eect of gravity when the
interfaces radius of curvature is much less than the capillary length, and vice versa.
78 Theoretical Fluid Mechanics

2 1


3 r

Figure 3.2
Interface between a liquid (1), a gas (2), and a solid (3).

3.5 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 3.2 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 . (3.17)
However, an equilibrium state is one which minimizes the free energy, implying that
the previous expression is zero for arbitrary (small) r: that is,
23 13
cos = . (3.18)
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 previous 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.
Surface Tension 79

air
a

R

glass tube
liquid
h

air free surface z = 0


liquid
Figure 3.3
Elevation of liquid level in a capillary tube.

As explained in Section 3.1, we would generally expect 12 and 23 to be positive.


On the other hand, 13 could be either positive or negative. Now, because | cos | 1,
Equation (3.18) can only be solved when 13 lies in the range

23 + 12 > 13 > 23 12 . (3.19)

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 (Batchelor 2000).
80 Theoretical Fluid Mechanics

3.6 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 3.3.) 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 3.4, the shape of the internal liquid/air interface within a capillary tube is not
significantly aected by gravity. Thus, from Section 3.3, 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 3.3) reveals that
a
R= . (3.20)
cos
Hence, from Equation (3.13), the mean curvature of the interface is given by
2 2 cos
n= = , (3.21)
R a
where is the associated surface tension. [The minus sign in the previous expression
arises from the fact that n points towards the center of curvature of the interface,
whereas the opposite is true for Equation (3.13).] Finally, from Equation (3.15),
application of the Young-Laplace equation to the interface yields
2 cos
gh = , (3.22)
a
which can be rearranged to give
2 cos
h . (3.23)
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, and is named after
its discoverer, James Jurin (1684-1750). 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 Equations (3.16) and (3.23), because
h l2
= 2 cos 2 . (3.24)
a a
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) (Batchelor 2000). 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.
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).
Surface Tension 81

3.7 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 (x). The shape of the interface is assumed to be y-independent. The unit
normal to the interface (directed from liquid to air) is thus
(z f ) ez f x e x
n= = , (3.25)
|(z f )| (1 + f x2 )1/2
where f x = d f /dx. Hence, the mean curvature of the interface is (Riley 1974)
f xx
n = , (3.26)
(1 + f x2 )3/2

where f xx = d 2 f /dx 2 . According to Equations (3.15) and (3.16), the shape of the
interface is governed by the nonlinear dierential equation

l 2 f xx
f = . (3.27)
(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 previous equation by f x /l 2 ,
and integrating with respect to x, we obtain

1 f2
= C 2, (3.28)
(1 + f x )
2 1/2 2l
where C is a constant. It follows that
f2
C 1, (3.29)
2 l2
and
1 C f 2 /2 l 2
= . (3.30)
fx [1 (C f 2 /2 l 2 )2 ]1/2
Let
2
C= 1, (3.31)
k2
where 0 < k < 1, and
2l
f = (1 k 2 sin2 )1/2 . (3.32)
k
Thus, from Equations (3.31) and (3.32),

f2
C = cos(2 ), (3.33)
2 l2
82 Theoretical Fluid Mechanics

z/l

0
2 1 0 1 2
x/l

Figure 3.4
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 so the constraint (3.29) implies that /4 3/4. Moreover, Equations (3.30)
and (3.33) reduce to
1 dx 1
= = . (3.34)
fx d f tan(2 )
It follows from Equations (3.32) and (3.34) that
dx dx d f l k cos(2 )
= = , (3.35)
d d f d (1 k 2 sin2 )1/2
which can be integrated to give
 /2
x k cos(2 )
= d, (3.36)
l (1 k 2 sin2 )1/2
assuming that x = 0 when = /2. Thus, we get


x 2 2
= k F(, k) + E(, k), (3.37)
l k k
where
E(, k) = E(/2, k) E(, k), (3.38)
F(, k) = F(/2, k) F(, k), (3.39)
Surface Tension 83

0.8

0.7

0.6

0.5

z/l 0.4

0.3

0.2

0.1

0
1 0.5 0 0.5 1
x/l

Figure 3.5
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 .

and

E(, k) = (1 k 2 sin2 )1/2 , (3.40)
0

F(, k) = (1 k 2 sin2 )1/2 , (3.41)
0

are types of incomplete elliptic integral (Abramowitz and Stegun 1965). In conclu-
sion, the interface shape is determined parametrically by


x 2 2
= k F(, k) + E(, k), (3.42)
l k k
z 2
= (1 k 2 sin2 )1/2 , (3.43)
l k
where /4 3/4. Here, the parameter k is restricted to lie in the range 0 < k <
1.
Figure 3.4 shows the capillary curves predicted by Equations (3.42) and (3.43) for
various dierent values of k. Here, we have chosen the plus sign in Equation (3.43).
However, if the minus sign is chosen then the curves are simply inverted: that is,
x x and z z. In can be seen that all of the curves shown in the figure are
84 Theoretical Fluid Mechanics

symmetric about x = 0: that is, 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 that the plates in question lie at x = d. Furthermore, let
the angle of contact of the interface with the plates be , where < /2. Because 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 Equation (3.43). 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 (3.34) (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 (3.42), we can numerically
determine the value of k for which x = d. The interface shape is then given by
Equations (3.42) and (3.43), 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 3.5. 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
(Abramowitz and Stegun 1965)

k2
E(, k) ( + sin cos ), (3.44)
4
k2
F(, k) + ( + sin cos ), (3.45)
4
where = /2 . Thus, Equations (3.42) and (3.43) reduce to
x k
sin(2 ), (3.46)
l 2
z 2 k
[1 cos(2 )]. (3.47)
l k 2
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 (3.46), this constraint yields
d k
cos . (3.48)
l 2
Thus, the height that the liquid rises between the two platesthat is, h = z(x = 0) =
z( = /2) 2 l/kis given by
cos
h . (3.49)
gd
Surface Tension 85

1.1
1
0.9
0.8
0.7
0.6
z/l
0.5
0.4
0.3
0.2
0.1
0
0 1 2 3 4 5
x/l

Figure 3.6
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 .

This result is the form taken by Jurins law, (3.23), 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, because 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. Because C = 1, it follows from Equation (3.28) that

h2
= 1 sin , (3.50)
2 l2
or
h = 2 l sin(/4 /2). (3.51)
Furthermore, again recalling that C = 1, Equation (3.30) can be integrated to give
 
h
df h/2l
1 2 y2
x= =l dy, (3.52)
z fx z/2l y (1 y 2 )1/2

where we have chosen the minus sign, and y = f /(2 l). Making the substitution
86 Theoretical Fluid Mechanics

y = sin u, this becomes


 sin1 (h/2 l)

sin1 (h/2 l)
x 1 1 + cos u
= 2 sin u du = ln + 2 cos u , (3.53)
l sin1 (z/2 l) sin u sin u sin1 (z/2 l)

which reduces to



1/2
1/2
x 1 2 l 1 2 l h2 z2
= cosh cosh + 4 2 4 2 , (3.54)
l z h l l

because cosh1 (z) ln[z + (z 2 1)1/2 ]. Thus, Equations (3.51) and (3.54) specify
the shape of a liquid/air interface that meets an isolated vertical plate at x = 0. In
particular, Equation (3.51) 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 3.6 shows an example interface calculated for = 25 .

3.8 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

(r f ) er fz ez
n = , (3.55)
|(r f )| (1 + fz2 )1/2

where fz d f /dz. Hence, from Equation (C.39), the mean curvature of the surface
is given by
1 d f
n= . (3.56)
f fz dz (1 + fz2 )1/2
The Young-Laplace equation, (3.12), then yields

f fz d f
= , (3.57)
a dz (1 + fz2 )1/2

where

a= . (3.58)
p0
Here, is the net surface tension, including the contributions from the internal and
external soap/air interfaces. Moreover, p0 = p is the pressure dierence between
the interior and the exterior of the bubble. Equation (3.57) can be integrated to give

f f2
= + C, (3.59)
(1 + fz )
2 1/2 2a
Surface Tension 87

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, (3.60)

where S is area of the bubble surface, and V the enclosed volume. The first term on
the right-hand side of the previous expression represents the work needed to over-
come surface tension, whereas the second term represents the work required to over-
come the pressure dierence, p0 , between the exterior and the interior of the bub-
ble. 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 (Fitzpatrick 2012). It follows that the equilibrium shape of the
bubble is such as to minimize
 z2  z2
E= 2 f (1 + fz2 )1/2 dz p0 f 2 dz, (3.61)
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
 z2
L( f, fz ) dz, (3.62)
z1

where
L( f, fz ) = 2 f (1 + fz2 )1/2 p0 f 2 , (3.63)
subject to the constraint that f is fixed at the limits. This is a standard problem
in the calculus of variations. (See Appendix E.) In fact, because the functional
L( f, fz ) does not depend explicitly on z, the minimizing function is the solution of
[see Equation (E.14)]
L
L fz = C, (3.64)
fz
where C  is an arbitrary constant. Thus, we obtain

f f2
2 = C , (3.65)
(1 + fz2 )1/2 2 a

which can be rearranged to give Equation (3.59). Hence, we conclude that applica-
tion of the Young-Laplace equation does indeed lead to a bubble shape that mini-
mizes the net free energy of the soap/air interfaces.
Consider the case p0 = 0, in which there is no pressure dierence across the
surface of the bubble. In this situation, writing C = b > 0, Equation (3.59) reduces
to
f = b (1 + fz2 )1/2 . (3.66)
88 Theoretical 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 0.2 0.4 0.6
z/c

Figure 3.7
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 Equa-
tion (3.66) is such as to minimize the surface area of the bubble (because the only
contribution to the free energy of the soap/air interfaces is directly proportional to
the bubble area). The previous equation can be rearranged to give

2 1/2
f
fz = 2 1 , (3.67)
b
which leads to
 r  r
df df
z z0 = = 2 /b 2
= b cosh1 (r/b), (3.68)
b fz b (f 1)1/2
or
r = b cosh(|z z0 |/b), (3.69)
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 follows, from Equation (3.69), that
z0 = 0, and
r = b cosh(z/b). (3.70)
Surface Tension 89

Here, the parameter b must be chosen so as to satisfy

c = b cosh(d/b). (3.71)

For example, if d = 0.65 c then b = 0.6416 c, and the resulting bubble shape is
illustrated in Figure 3.7.
Let d/c = and d/b = u, in which case the previous equation becomes

G(u) = u cosh u = 0. (3.72)

Now, the function G(u) attains a maximum value


1
G(u0 ) = u0 , (3.73)
tanh u0

when u0 = sinh1 (1/). Moreover, if G(u0 ) > 0 then Equation (3.72) possesses
two roots. It turns out that the root associated with the smaller value of u mini-
mizes 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 (3.72) 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 dierence across the
surface of the bubble. In this situation, writing

2 a = + , (3.74)
2 a C = , (3.75)

Equation (3.59) becomes


( + ) f
= f 2 + , (3.76)
(1 + fz2 )1/2
which can be rearranged to give

( 2 f 2 )1/2 ( f 2 2 )1/2
fz = . (3.77)
f 2 +
We can assume, without loss of generality, that || > ||. It follows, from the previous
expression, that || f ||. Hence, we can write

f 2 = 2 cos2 + 2 sin2 , (3.78)


2 2
k2 = , (3.79)
2
90 Theoretical Fluid Mechanics

1.1
1
0.9
0.8
0.7
0.6
r/

0.5
0.4
0.3
0.2
0.1
0
2 1 0 1 2
z/

Figure 3.8
Radius versus axial distance for an unduloid soap bubble calculated with k = 0.95.

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

f = || (1 k 2 sin2 )1/2 , (3.80)


= sgn() || (1 k )2 1/2
, (3.81)

and

dz 1 df
= = f + , (3.82)
d fz d f
which can be integrated to give
 
|z| = || E(, k) + sgn( ) (1 k 2 )1/2 F(, k) , (3.83)

where E(, k) and F(, k) are incomplete elliptic integrals [see Equations (3.40) and
(3.41)]. 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 k 2 )1/2 r/ 1 for
/2 0, and 0.5 /(p0 ) < 1 for 1 k > 0, where

r = (1 k 2 sin2 ), (3.84)
 
|z| = E(, k) + (1 k 2 )1/2 F(, k) . (3.85)
Surface Tension 91

1.1
1
0.9
0.8
0.7
0.6
r/

0.5
0.4
0.3
0.2
0.1
0
0.5 0 0.5
z/

Figure 3.9
Radius versus axial distance for a positive pressure nodoid soap bubble calculated
with k = 0.95.

The axisymmetric curve parameterized by the previous pair of equations is known


as an unduloid. Note that an unduloid bubble always has positive internal pressure
(relative to the external pressure): that is, p0 > 0. An example unduloid soap bubble
is illustrated in Figure 3.8
In the second case, > 0 and < 0. It follows that (1 k 2 )1/4 r/ 1 for 0
0, and 0 < /(p0 ) 0.5 for 0 < k 1, where 0 = sin1 ([1 (1 k 2)1/2 ]1/2 /k),
and

r = (1 k 2 sin2 ), (3.86)
 
|z| = E(, k) (1 k 2 )1/2 F(, k) . (3.87)

The axisymmetric curve parameterized by the previous pair of equations is known


as an nodoid. This particular type of nodoid bubble has positive internal pressure:
that is, p0 > 0. An example positive pressure nodoid soap bubble is illustrated in
Figure 3.9.
In the third case, < 0 and > 0. It follows that (1 k 2 )1/2 | r/||
| (1 k 2 )1/4 for /2 0 (or /2 0 ), and 0 > /(p0 ||) 0.5 for
92 Theoretical Fluid Mechanics

0.6

0.5

0.4
r/||

0.3

0.2

0.1

0
0.1 0.2 0.3 0.4 0.5
z/||

Figure 3.10
Radius versus axial distance for a negative pressure nodoid soap bubble calculated
with k = 0.95.

0 < k 1, where

r = || (1 k 2 sin2 ), (3.88)
 
|z| = || E(, k) (1 k 2 )1/2 F(, k) . (3.89)

The axisymmetric curve parameterized by the previous pair of equations is again a


nodoid. However, this particular type of nodoid bubble has negative internal pres-
sure: that is, p0 < 0. An example negative pressure nodoid soap bubble is illustrated
in Figure 3.10.

3.9 Exercises
3.1 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 . (Lamb 1928.)

3.2 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 Tension 93

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. (Lamb 1928.)

3.3 A film of water is held between two parallel plates of glass a small distance
2 d 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. (Lamb 1928.)

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

2

dx
2 x

per unit breadth of the corrugations. Here, x is measured horizontally, per-


pendicular to the corrugations. Moreover, denotes the elevation of the sur-
face 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) a 2 k 2 .

(Lamb 1928.)

3.5 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+ .] (Lamb 1928.)

3.6 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
94 Theoretical Fluid Mechanics

the plane of the ring, and passing through its center. The mass is held to-
gether by surface tension alone. Show that the section of the ring must be
approximately circular. Demonstrate that

1/2
2
= ,
a c2

where is the density, the surface tension, and c the radius of the cross-
section. (Lamb 1928.)
3.7 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 a 3 + 4 a 2 = p0 (a13 + a23 ) + 4 (a12 + a22 ),

where p0 is the ambient pressure, and the surface tension of the soap/air
interfaces. (Batchelor 2000.)
3.8 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 di-
ameter 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.) (Batchelor 2000.)
3.9 Two small solid bodies are floating on the surface of a liquid. Show that the
eect 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. (Batchelor 2000.)
4
Incompressible Inviscid Flow

4.1 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.
Further information on these concepts can be found in Faber 1995, Batchelor 2000,
and Milne-Thompson 2011.

4.2 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.
If the flow is unsteady then the configuration of the stream tubes and filaments
changes from time to time. However, if 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, because the flow is, by definition, always tangential to
these walls. Moreover, the walls are fixed in space and time, because 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 suciently 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 where the cross-sectional
areas are S 1 and S 2 , respectively. Consider the section of the filament lying between
these points. Because 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

95
96 Theoretical Fluid Mechanics

D
D
C
C

B
B A
A
Figure 4.1
Bernoullis theorem.

implies that
v1 S 1 = v2 S 2 . (4.1)
This is the simplest manifestation of the equation of fluid continuity discussed in
Section 1.9. The previous result is equivalent to the statement that the product of
the flow 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 speed at that point becomes infinite. Leaving this case out
of consideration, it follows that stream filaments in steadily flowing incompressible
fluids either form closed loops, or terminate at the boundaries of the fluid. The same
is, of course, true of streamlines.

4.3 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 quan-
tity
p
+T (4.2)

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 4.1. Let us
Incompressible Inviscid Flow 97

denote the values of quantities at AB and CD by the suxes 1 and 2, respectively.


Thus, p1 , v1 , 1 , S 1 , T1 are the pressure, flow speed, mass density, cross-sectional
area, and total energy per unit mass, respectively, at AB, et cetera. 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. Because 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 . (4.3)
 
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 ). (4.4)
In the absence of viscous energy dissipation, this energy increase must equal the net
work done on the fluid by the pressures at AB and CD, which is


p1 p2
p1 S 1 v1 t p2 S 2 v2 t = m . (4.5)
1 2
Equating expressions (4.4) and (4.5), we find that
p1 p2
+ T1 = + T2 , (4.6)
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) v 2, and the potential energy per unit mass, , and Bernoullis theorem thus
becomes
p 1 2
+ v + = constant along a streamline. (4.7)
2
If we focus on a particular streamline, 1 (say), then Bernoullis theorem states that
p 1 2
+ v + = C1 , (4.8)
2
where C1 is a constant characterizing that streamline. If we consider a second stream-
line, 2 (say), then
p 1 2
+ v + = C2 , (4.9)
2
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 4.15), so that
p 1 2
+ v + =C (4.10)
2
throughout the fluid.
98 Theoretical Fluid Mechanics

4.4 Euler Momentum Theorem


Consider the stream filament shown in Figure 4.1. Let S 1 and S 2 be the cross-
sectional areas at AB and CD, respectively, and let v1 and v2 be the corresponding
flow velocities. Assuming the the flow is both steady and incompressible, Eulers
momentum theorem, which is named after Leonhard Euler (1707-1783), states that,
neglecting external forces, the resultant force due to the pressure of the surrounding
fluid on the walls and ends of the filament is equivalent to forces S 1 v12 and S 2 v22
acting normally outward at the ends AB and CD, respectively.
The proof is straightforward. According to Newtons second law of motion, the
resultant force must produce the change in the momentum of the fluid that occu-
pies the portion of the filament between AB and CD at any given instant of time, t.
Suppose that at time t + t the fluid in question occupies the portion of the filament
between A B and C  D . The momentum of the fluid in question has then increased
by the momentum of the fluid between CD and C  D , and decreased by the momen-
tum of the fluid between AB and A B . Hence, there has been a gain of momentum
S 2 v2 t v2 at CD, and a loss of momentum S 1 v1 t v1 at AB. Thus, the net
rate of charge of momentum consists of a gain S 2 v2 v2 at CD, and a loss S 1 v1 v1
at AB. This net rate of change is produced solely by the thrusts acting on the walls
and ends of the filaments. It follows that these thrusts are equivalent to the forces
S 1 v12 and S 2 v22 acting normally outward at AB and CD, respectively.
If p1 and p2 are the pressures at AB and CB, respectively, then the thrusts act-
ing normally inward on the ends of the filament are p1 S 1 at AB and p2 S 2 at CD.
According to Eulers theorem, the thrusts exerted on the walls plus the thrusts acting
on the ends are equivalent to the normal outward forces S 1 v12 at AB and S 2 v22 at
CD. It follows that the thrusts exerted by the walls on the fluid are equivalent to the
normal outward forces S 1 (p1 + v12 ) at AB and S 2 (p2 + v22 ) at CD. Conversely,
the thrusts exerted by the fluid on the walls are equivalent to normal inward forces
S 1 (p1 + v12 ) at AB and S 2 (p2 + v22 ) at CD.
Note, finally, that the Euler momentum theorem obviously also applies to a
stream tube, as long as the flow through the ends of the tube is uniform across the
cross-section.

4.5 dAlemberts Paradox


Consider a long straight tube through which an inviscid fluid flows at the constant
speed V. If we place a small obstacle, A, in the middle of the tube then the flow in the
immediate neighborhood of A will be modified, but that a great distance upstream
or downstream of A will presumably remain undisturbed. In general, in order to
maintain the obstacle at rest, we need to exert both a force and a couple on it. Let
F be the component of the force in the direction of the original flow. Neglecting
Incompressible Inviscid Flow 99

S1 S2
V V
A

Figure 4.2
The dAlembert paradox.

external forces, such as gravity, F is the resultant in the direction of the flow of the
pressure thrusts acting on the boundary of A.
Consider two cross-sections, S 1 and S 2 , a great distance upstream and down-
stream of A, respectively. The fluid between these sections can be split into a great
many stream filaments to each of which Eulers momentum theorem is applicable.
The outer filaments are bounded by the walls of the tube, and so the thrust compo-
nents acting on these are directed perpendicular to the flow. The walls of the filaments
in contact with A are acted on by the obstacle, which exerts on them a force whose
component in the direction of the flow is F. By Eulers theorem, the resultant of all
of the thrusts acting on the fluid in the tube is

S 1 V 2 + S 2 V 2 , (4.11)

which vanishes, because S 1 = S 2 .


By Bernoullis theorem, the pressure p1 acting across S 1 is the same as the pres-
sure p2 acting across S 2 . Thus, according to Eulers theorem,

p1 S 1 F p2 S 2 = 0, (4.12)

which implies that


F = 0. (4.13)
The surprising result that the net parallel force exerted by an inviscid fluid stream
on a stationary obstacle placed in its path is zero is known as dAlemberts paradox,
after the French scientist Jean-Baptiste dAlembert (17171783).
If we suppose the walls of the tube to recede to infinity then we obtain the case of
an obstacle immersed in a moving stream that is unbounded in every direction. The
previous proof still shows that F = 0.
Finally, if we impose on the whole system a uniform velocity V in the direction
opposite to that of the stream then the fluid at a great distance is reduced to rest, and
the obstacle A moves with the uniform speed V. However, superposing a uniform
100 Theoretical Fluid Mechanics

B
A v

Figure 4.3
Outflow through an orifice.

flow does not alter the dynamical conditions. Therefore, we conclude that the resis-
tance to a solid body moving with uniform velocity through an unbounded inviscid
fluid, otherwise at rest, is zero.

4.6 Flow Through an Orifice


Consider the situation, illustrated in Figure 4.3, in which a horizontal jet of fluid
emerges from an orifice in the side of a container. As shown in the figure, the jet nar-
rows over a short distance beyond the orifice that is comparable with the jet diameter
to form what is generally known as a vena contractathat is, a contracted vein.
The jet is bound to narrow in this manner because of the curvature of the lines of
flow as they pass through the orifice. The narrowing of the jet implies the existence
of a transverse pressure gradient. In other words, the pressure at A, on the axis of the
jet, is higher than the atmospheric pressure that acts at B. The pressure excess at A
suggests that the fluid on the axis is still accelerating longitudinally as it leaves the
orifice. Only in the vena contracta does the flow velocity becomes uniform, and the
pressure atmospheric, all the way across the jet.
Let S be the cross-sectional area of the orifice, and C S that of the vena contracta.
Here, C is known as the contraction coecient. Let us apply Bernoullis theorem to
a streamline that starts on the surface of the fluid within the container, and ends in the
vena contracta. Suppose that the surface of the fluid lies a height h above the orifice.
Let us assume that the fluid close to the surface is essentially at rest (which implies
Incompressible Inviscid Flow 101

that the outflow through the orifice is not suciently strong to cause the surface level
to drop at a significant rate.) Let v be the uniform fluid velocity in the vena contracta.
Of course, the pressure is atmospheric both at the surface of the fluid and in the vena
contracta. It follows that
1
g h = v 2, (4.14)
2
or
v = (2 g h)1/2. (4.15)

In other words, the eux velocity of the fluid from the orifice is the same as that
it would have acquired by falling a height h under gravity. This result is known as
Torricellis law, after Evangelista Torricelli (1608-1647). Finally, the discharge rate
of fluid flowing through the orifice is

Q = C S v = C S (2 g h)1/2. (4.16)

Let p = p0 + g h be the hydrostatic pressure at the level of the orifice when


the orifice is closed. Here, p0 is atmospheric pressure. The fluid experiences a
thrust S p from the section of the wall directly opposite the orifice, and a thrust
S p from the section of the wall closing the orifice. Let us suppose, as a first
approximation, that the hydrostatic pressure remains unaltered when the orifice is
opened. In this situation, the fluid experiences a thrust S p from the section of the
wall directly opposite the orifice, and a thrust S p0 from the orifice. The net thrust,
S (p p0 ) = S g h, is responsible for accelerating the jet. Now, the jets rate of
momentum outflow is v C S v. Momentum conservation yields

S g h = C S v 2 = 2 C S g h, (4.17)

where use has been made of Equation (4.15). Thus, we conclude that the contraction
coecient takes the value 1/2.
In reality, Bernoullis theorem suggests that when the orifice is opened the pres-
sure on the walls in the neighborhood of the orifice will fall below the hydrostatic
value, which implies that the accelerating thrust is actually greater than S (p p0 ).
Consequently, C > 1/2. Obviously, C cannot exceed unity, so we conclude that, in
general, 1/2 < C < 1. For instance, if the orifice is a circular hole punched in a thin
plate then the contraction coecient is observed to take the value 0.62 (Batchelor
2000).
Suppose, however, that we fit a small cylindrical nozzle projecting inward from
the orifice, as shown in Figure 4.4. In this case, the original assumption that the
pressure on the walls in the neighborhood of the orifice is hydrostatic is essentially
correct. This follows because the region where the lines of flow are converging on the
orifice is far removed from the walls, and the velocity of the fluid in contact with the
walls is negligible. Thus, the contraction coecient is exactly 1/2. This arrangement
is known as a Borda mouthpiece, after Jean-Charles Borda (17331799).
102 Theoretical Fluid Mechanics

Figure 4.4
A Borda mouthpiece.

4.7 Sub-Critical and Super-Critical Flow


Consider a shallow stream of depth h, uniform width, and uniform flow velocity v,
that is fed from a deep reservoir whose surface lies a height H above the (horizontal)
bed of the stream. Here, H is usually referred to as the head height. Assuming that
the water in the reservoir is eectively stationary, application of Bernoullis equation
to a streamline lying on the surface of the water (where the pressure is atmospheric)
yields
v2
H =h+ . (4.18)
2g
Let Q be the flow rate per unit width of the stream, which is assumed to be fixed. It
follows that
Q = h v. (4.19)
The previous two equations can be combined to give

H = F(v), (4.20)

where
Q v2
F(v) = + . (4.21)
v 2g
It is easily demonstrated that the function F(v) attains its minimum value,

1/3
3 Q2
Fc = , (4.22)
2 g
Incompressible Inviscid Flow 103

when v = vc , where
vc = (Q g)1/3 . (4.23)
We conclude that, as long as H > (3/2) (Q /g) , Equation (4.20) possesses two
2 1/3

possible solutions that are consistent with a given head height and flow rate. In one
solution, the stream flows at a relatively slow velocity, v , which is such that v < vc .
In the other, the stream flows at a relatively fast velocity, v+ , which is such that
v+ > vc . The corresponding depths are h = Q/v and h+ = Q/v+ , respectively.
It is helpful to introduce the dimensionless Froude number,
v
Fr =  . (4.24)
gh

(See Section 1.15.) Note that g h is the characteristic propagation velocity of a
gravity wave in shallow water of depth h. (See Section 11.4.) Hence, if Fr < 1
then the streams flow velocity falls below the wave speedsuch flow is termed
sub-critical. On the other hand, if Fr > 1 then the flow velocity exceeds the wave
speedsuch flow is termed super-critical.
We can combine Equations (4.18), (4.19), and (4.24) to give

H = G(Fr), (4.25)

where
1/2

Q2 1 Fr 4/3
G(Fr) = + . (4.26)
g Fr 2/3 2
It is easily demonstrated that G(Fr) attains its minimum value

1/3
3 Q2
Gc = , (4.27)
2 g

when Fr = Frc , where


Frc = 1. (4.28)
Hence, we again conclude that, as long as H > (3/2) (Q 2/g)1/3 , there are two pos-
sible flow velocities of the stream (parameterized by two dierent Froude numbers)
that are consistent with a given head height and flow rate. However, it is now clear
that the smaller velocity is sub-critical (i.e., Fr < 1), whereas the larger velocity is
super-critical (i.e., Fr > 1).

4.8 Flow over Shallow Bump


Consider a shallow stream of depth H, uniform width, and uniform flow velocity V.
Suppose that there is a very shallow bump of height d  H on the (horizontal) bed
of the stream, as shown in Figure 4.5. Suppose, further, that, at the point where the
104 Theoretical Fluid Mechanics

V v
H

Figure 4.5
Flow over a shallow bump.

stream passes over the top of the bump, its velocity is v, and its surface rises a height
h  H above the unperturbed surface.
Fluid continuity yields
H V = (H + h d) v. (4.29)
Furthermore, application of Bernoullis equation to a streamline lying on the surface
of the water (where the pressure is atmospheric) gives
1 2 1
gH+ V = g (H + h) + v 2 . (4.30)
2 2
The previous equation reduces to
v 2 = V 2 2 g h. (4.31)
Eliminating v between Equations (4.29) and (4.31), we obtain
H 2 V 2 = (H + h d) 2 (V 2 2 g h), (4.32)
which can be rearranged to give
 
V 2 (H + h d)2 H 2 = 2 g h (H + h d)2 , (4.33)
or
2
2
hd h hd
2
Fr 1 +
1 = 2 1+ . (4.34)
H H H
Here,
V
Fr = (4.35)
gH
is the Froude number of the unperturbed flow. Finally, given that h/H  1 and
d/H  1, Equation (4.34) reduces to
d
h . (4.36)
1 1/Fr 2
Incompressible Inviscid Flow 105

z
h2 v2

h1 v1

Figure 4.6
A hydraulic jump.

It follows, from the previous expression, that if the flow is super-critical, so that
Fr > 1, then h is positive. On the other hand, if the flow is sub-critical, so that
Fr < 1, then h is negative. Thus, if a super-critical shallow stream passes over a very
shallow bump on its bed then the surface of the stream becomes slightly elevated.
On the other hand, if a sub-critical stream passes over such a bump then the surface
of the stream becomes slightly depressed. A similar eect occurs when there is a
narrowing of the channel in the horizontal direction. A more sophisticated version
of the previous calculation, which does not necessarily assume that the stream is
shallow, can be found in Section 11.10.

4.9 Stationary Hydraulic Jumps


Under certain circumstances, water flowing within a horizontal open channel (i.e., a
stream) of constant width is found to have a depth that changes very rapidly over a
short section of the channel. This phenomenon is known as a hydraulic jump, and is
illustrated in Figure 4.6.
Suppose that the jump is stationary. Let h1 , v1 , and p1 be the depth, flow veloc-
ity, and pressure, of the water, respectively, upstream of the jump. Similarly let h2 ,
v2 , and p2 , be the depth, flow velocity and pressure, respectively, downstream of the
jump. As before, v1 and v2 are assumed to be uniform across the channel. We are also
assuming that the flow is laminar (i.e., smooth) both upstream and downstream of the
jump. Note that the region of the channel in which the jump occurs is generally asso-
ciated with violent mixing that gives rise to significant transfer of mechanical energy
from the laminar (i.e., smooth) to the turbulent component of the flow. (The latter
component is localized in the vicinity of the jump, and is continuously dissipated by
viscosity on small-scales.) Consequently, we cannot use Bernoullis equation to ana-
lyze the jump, because this equation assumes that the laminar component of the flow
conserves mechanical energy. However, we can still make use of fluid continuity, as
106 Theoretical Fluid Mechanics

well as the Euler momentum theorem, neither of which depend on the conservation
of mechanical energy.
Fluid continuity yields
Q = h 1 v1 = h 2 v2 , (4.37)
where Q is the fixed flow rate per unit width. Furthermore, the Euler momentum
theorem (see Section 4.4) implies that
 h2  h1
F21 = (p2 + v22 ) dz (p1 + v12 ) dz. (4.38)
0 0

Here, F21 is the horizontal thrust per unit width exerted by the channel bed on the
water lying between points 1 and 2. Moreover, z measured vertical height above the
bed. Assuming the usual linear pressure variation with depth, we can write

p1 (z) = p0 + g (h1 z), (4.39)


p2 (z) = p0 + g (h2 z), (4.40)

where p0 is atmospheric pressure. Hence, we deduce that


F21 1 2 1 2 Q
= h2 h1 + (v2 v1 ). (4.41)
g 2 2 g

The neglect of frictional drag at the channel bed implies that F21 = 0. Thus, we
obtain
1 2 1 2 Q
h h + (v2 v1 ) = 0. (4.42)
2 2 2 1 g
The previous equation possesses the trivial solution h1 = h2 and v1 = v2 , which
corresponds to the absence of a hydraulic jump. Eliminating v1 and v2 between Equa-
tions (4.37) and (4.42), and canceling a common factor h2 h1 , we obtain the non-
trivial solution
2 Q2
h2 h1 (h2 + h1 ) = . (4.43)
g
Now, the upstream Froude number is defined
v1 Q
Fr1 =  = . (4.44)
g h1 g h13/2
1/2

Eliminating Q between the previous two equations, we obtain



2
h2 h2
+ 2 Fr12 = 0, (4.45)
h1 h1

which can be solved to give




h2 1
= 1 + 1 + 8 Fr12 . (4.46)
h1 2
Incompressible Inviscid Flow 107

Here, we have neglected an unphysical solution in which h2 /h1 is negative. Note that
h2
1 as Fr1  1. (4.47)
h1
The downstream Froude number is defined
v2 Q
Fr2 =  = . (4.48)
g h2 g h23/2
1/2

Eliminating Q between Equations (4.43) and (4.48), we obtain



2
h1 h1
+ 2 Fr22 = 0, (4.49)
h2 h2
which can be solved to give


h1 1
= 1 + 1 + 8 Fr2 .
2
(4.50)
h2 2
Here, we have again neglected an unphysical solution in which h1 /h2 is negative. We
can combine Equations (4.46) and (4.50) to give

G(Fr1 ) G(Fr2 ) = 1, (4.51)

where
1 
G(x) 1 + 1 + 8 x 2 . (4.52)
2
Note that

G(x)  1 as x  1, (4.53)

which implies that

Fr1  1 as Fr2  1. (4.54)

In other words, if the upstream flow is super-critical then the downstream flow is
sub-critical, and h2 > h1 . On the other hand, if the upstream flow is sub-critical then
the downstream flow is super-critical, and h2 < h1 .
By analogy with Equation (4.18), we can define the head heights of the flows
upstream and downstream of the jump as

1 v12
H1 = h 1 + , (4.55)
2 g
and
1 v22
H2 = h 2 + , (4.56)
2 g
respectively. Of course, in the absence of any transfer of mechanical energy from the
108 Theoretical Fluid Mechanics

laminar to the turbulent component of the flow, within the jump, we would expect
H1 = H2 . (Because this is what Bernoullis equation predicts.) In the presence of
the transfer, we expect H1 > H2 . In other words, we expect there to be a head loss
across the jump. Note that it is impossible for H2 to exceed H1 , because this would
imply a transfer of mechanical energy from the turbulent to the laminar component
of the flow, within the jump, which violates the second law of thermodynamics. Let

v12 v2
H L = H1 H2 = h 1 h 2 + 2 (4.57)
2g 2g
be the positive definite head loss. Making use of some previous definitions, we can
write

HL h2 1 2 h12
= 1 + Fr1 1 2 .
(4.58)
h1 h1 2 h2
Now, according to Equation (4.45),


1 h2 h2
Fr12 = +1 . (4.59)
2 h1 h1
Hence, combining the previous two equations, we obtain


2
h2
+ 1 ,
1 h1 h2 h2
HL = 1 4 + (4.60)
4 h2 h1 h1 h1
or
(h2 /h1 1)3
HL = . (4.61)
4 h2 /h1
Thus, the thermodynamic constraint HL > 0 implies that h2 > h1 . However, as we
have already seen, h2 > h1 implies that Fr1 > 1. In other words, a stationary hydraulic
jump can only occur when the upstream flow is super-critical, and the downstream
flow sub-critical.
A hydraulic jump often occurs at the base of a spillway from a dam, where the
flow is accelerated to super-critical speeds. Because a hydraulic jump always results
in a loss of mechanical energy from the flow, spillways are sometimes designed to
promote jumps, so as to deliberately remove energy from the flow, thereby reducing
the danger from excessive currents in flood control.

4.10 Tidal Bores


A tidal bore is a sort of hydraulic jump that propagates up (i.e., upstream) a river
estuary. The upper part of Figure 4.7 shows such a bore in the local rest frame of
the Earth. The bore is propagating at the velocity V up a river of uniform width,
and depth h1 , that is flowing downstream at the velocity u1 . The flow behind the
Incompressible Inviscid Flow 109

Earth frame
V
u2
h2

h1 u1

Co-moving frame

V u2
h2

h1 V + u1

Figure 4.7
A tidal bore.

bore is of depth h2 , and is flowing upstream at the velocity u2 . The lower part of
the figure shows the same phenomenon in the rest frame of the bore. In this frame,
we observe a stationary hydraulic jump with an upstream depth and flow velocity
h1 and v1 = V + u1 , respectively, and a downstream depth and flow velocity h2 and
v2 = V u2 , respectively. Making use of Equations (4.44) and (4.45), we obtain

2
h2 h2 2 (u1 + V)2
+ = 0, (4.62)
h1 h1 g h1

which can be rearranged to give


1/2
g h2 (h1 + h2 )
V = u1 + . (4.63)
2 h1

Thus, we deduce that the speed of the bore, relative to the unperturbed river, is a
simple function of the upstream and downstream depths. Note that, in the limit
h1 h2 h, the previous equation reduces to V u1 + g h. In other words, a
weak bore degenerates intoan ordinary shallow water gravity wave propagating at
the characteristic velocity g h relative to the stream.
Tidal bores are found in river estuaries where a funneling eect causes the speed
of the incoming tide to increase to such a point that the flow becomes super-critical.
For example, bores can be observed daily on the River Severn in England.
110 Theoretical Fluid Mechanics

v(x)
h(x)

d(x)

Figure 4.8
Flow over a broad-crested weir.

4.11 Flow over a Broad-Crested Weir


Consider the situation, illustrated in Figure 4.8, in which a broad-crested weir is
placed in a shallow stream. The purpose of the weir is to impede the flow in such a
manner that there is a transition from sub-critical flow, upstream of the weir, to super-
critical flow, immediately downstream of the weir. (There is usually a transition back
to sub-critical flow, via a hydraulic jump, some way downstream of the weir.)
Let x measure horizontal distance, and let h(x), v(x), and d(x) be the stream depth,
stream velocity, and the height of the weir above the stream bed, respectively. (Here,
we are assuming that the velocity remains uniform across the stream at any point on
the weir.) A direct generalization of the analysis of Section 4.7 reveals that

v2
H =h+d+ , (4.64)
2g

and
Q = h v, (4.65)
where (via continuity) the flow rate, Q, is not a function of x. It follows that

H d(x) = F(v), (4.66)

where F(v) is defined in Equation (4.21). Suppose that the weir attains its maximum
height, d(0), at x = 0. It follows that H d(x) passes through a minimum value at
x = 0. Hence, we would expect F(v) to also pass through its minimum value at x = 0.
It follows that

1/3
3 Q2
H d(0) = Fc = , (4.67)
2 g
v(0) = vc = (Q g)1/3, (4.68)
Incompressible Inviscid Flow 111

where use has been made of Equations (4.22) and (4.23). However,

Q = h(0) v(0), (4.69)

where h(0) is the depth of the stream as its passes over the highest point of the weir.
Thus, we deduce that

3/2  1/2
2
Q = [g h (0)]
3 1/2
= g [H d(0)] 3 . (4.70)
3

In other words, the flow rate (per unit width) is very simply related to the depth of
the stream as it passes over the highest point of the weir, or, alternatively, the depth
of the highest point of the weir below the water surface in the reservoir feeding the
stream. For this reason, weirs are commonly used as devices to both measure and
control flow rates in streams.

4.12 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 , so that

= v. (4.71)

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 4.9.) 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. The divergence theorem (see Section A.20), applied to the section,
yields  
dS = dV, (4.72)

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


ever,
=v0 (4.73)
[see Equation (A.173)], implying that

dS = 0. (4.74)
112 Theoretical Fluid Mechanics

S2

A
S1

Figure 4.9
A vortex filament.

Now, dS = 0 on the curved surface of the filament, because 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

dS = S 2 2 S 1 1 = 0. (4.75)

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 to vanish, and, therefore, the vorticity, , would have to be-
come infinite. Thus, a vortex filament must either form a closed vortex ring, or must
terminate at the fluid boundary.
Because 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.

4.13 Circulation and Vorticity


Consider a closed curve C situated entirely within a moving fluid. The vector line
integral (see Section A.14) 
C = v dr, (4.76)
C
Incompressible Inviscid Flow 113

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 (i.e.,
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 the curl theorem (see Section A.22),
 
C = v dr = dS. (4.77)
C S

Thus, we conclude that circulation and vorticity are intimately related to one another.
In fact, according to the previous 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 Equation (4.77) is that
the surface S must lie entirely within the fluid.

4.14 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 in-
dependent of time. The proof is as follows. The circulation around a given loop C is
defined 
C = v dr. (4.78)
C
However, for a loop that is co-moving with the fluid, we have dv = d(dr/dt) =
d(dr)/dt. Thus,  
dC dv
= dr + v dv. (4.79)
dt C dt C
By definition, dv/dt = Dv/Dt for a co-moving loop. (See Section 1.10.) More-
over, the equation of motion of an incompressible inviscid fluid can be written [see
Equation (1.79)]

Dv p
= + , (4.80)
Dt
because is a constant. Hence,


dC p 1
= v 2 + dr = 0, (4.81)
dt C 2
because v dv = d(v 2 /2) = (v 2 /2) dr (see Section A.18), and p/ v 2 /2 + is
obviously a single-valued function.
114 Theoretical Fluid Mechanics

C

Figure 4.10
A vortex tube.

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 Figure 4.10.) The normal component of the vorticity over
the surface enclosed by C is zero, because all vorticity vectors are tangential to this
surface. Thus, from Equation (4.77), 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, because
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 4.10.

4.15 Irrotational Flow


Flow is said to be irrotational when the vorticity has the magnitude zero every-
where. It immediately follows, from Equation (4.77), 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 circu-
lation 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 (1.39) and (1.79),
Incompressible Inviscid Flow 115

takes the form



v p
+ (v ) v = + , (4.82)
t
because is a constant. However, from Equation (A.171),

(v ) v = (v 2 /2) v . (4.83)

Thus, we obtain

v p 1 2
= + v + + v . (4.84)
t 2
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. (4.85)
Dt
Thus, if = 0, initially, then D/Dt = 0, and, consequently, = 0 at all subsequent
times.
Suppose that O is a fixed point, and P an arbitrary movable point, in an irrota-
tional fluid. Let O and P be joined by two dierent paths, OAP and OBP (say). It
follows that OAPBO is a closed curve. Because the circulation around such a curve
in an irrotational fluid is zero, we can write
 
v dr + v dr = 0, (4.86)
OAP PBO

which implies that  


v dr = v dr = P (4.87)
OAP OBP

(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 position vector is r,
P
then we have eectively defined a scalar field (r) = O v dr.
Consider a point Q that is suciently 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)  Q
= Q + P = v dr v . (4.88)
P

The previous equation becomes exact in the limit that || 0. Because Q is arbi-
trary (provided that it is suciently close to P), the direction of the vector is also
arbitrary, which implies that
v = . (4.89)
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
116 Theoretical Fluid Mechanics

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. (4.90)

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 = (4.91)

and
v = 0. (4.92)
These two expressions can be combined to give (see Section A.21)

2 = 0. (4.93)

In other words, the velocity potential in an incompressible irrotational fluid satisfies


Laplaces equation.
According to Equation (4.84), if the flow pattern in an incompressible inviscid
fluid is also irrotational, so that = 0 and v = , then we can write


p 1 2
+ v + = 0, (4.94)
2 t
which implies that
p 1 2
+ v + = C(t), (4.95)
2 t
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 previous equation can be rewritten
p 1 2
+ v + = C, (4.96)
2 t
where C is constant in both space and time. Expression (4.96) is a generalization
of Bernoullis theorem (see Section 4.3) 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, (4.97)
2
which demonstrates that for steady irrotational flow the constant in Bernoullis theo-
rem is the same on all streamlines. (See Section 4.3.)
Incompressible Inviscid Flow 117

4.16 Exercises
4.1 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. (Milne-
Thomson 1958.)
4.2 Water of mass density and pressure p flows through a curved pipe of uni-
form cross-sectional area S , whose radius of curvature is R, at the uniform
speed v. Demonstrate that there is a net force per unit length S (pp0 + v 2 )/R
acting on the pipe, and that this force is everywhere directed away from the
pipes local center of curvature. Here, p0 is atmospheric pressure.
4.3 Water is held in a right circular conical tank whose apex lies vertically below
the center of its base. The water initially fills the tank to a height h above the
vertex. Let a be the initial radius of the surface of the water inside the tank.
A small hole of area S (that is much less than a 2) is made at the bottom of
the tank. Demonstrate that the time required to empty the tank is at least

1/2
5 a2 h
.
2 S 2g

4.4 Water is held in a spherical tank of radius a, and initially fills the tank to a
height h < 2 a above its lowest point. A small hole of area S (that is much
less than a 2 ) is made at the bottom of the tank. Demonstrate that the time
required to empty the tank is at least


2 2 1 5/2
a h 3/2
h .
S (2 g)1/2 3 5

4.5 Water is held in two contiguous tanks whose cross-sectional areas, A1 and
A2 , are independent of height. A small hole of area S (where S  A1 , A2 )
is made in the wall connecting the tanks. Assuming that the initial dierence
in water level between the two tanks is h, show that the time required for the
water levels to equilibrate is at least


1/2
2 A1 A2 h
.
S A1 + A2 2 g

4.6 For a channel of width W, having a discharge rate Q, show that there is a
critical depth hc , where

2 1/3
Q
hc = ,
gW2
which must be exceeded before a hydraulic jump is possible.
118 Theoretical Fluid Mechanics

4.7 Show that for a stationary hydraulic jump in a rectangular channel, the up-
stream Froude number Fr1 , and the downstream Froude number Fr2 , are re-
lated by
8 Fr12
Fr22 = .
[(1 + 8 Fr12 )1/2 1] 3

4.8 Consider a simply-connected 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.]
4.9 In Exercise 4.8, 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.
4.10 An incompressible fluid flows in a simply-connected volume V bounded by a
surface S . The normal flow at the boundary is prescribed. Show that the flow
pattern with the lowest kinetic energy is irrotational. This result is known
as the Kelvin minimum energy theorem. [Hint: Try writing v = + v,
where is the velocity potential of the irrotational flow pattern. Let v = 0
throughout V, and v dS = 0 on S . Show that the kinetic energy is lowest
when v = 0 throughout V.]
5
Two-Dimensional Incompressible Inviscid
Flow

5.1 Introduction
This chapter investigates two-dimensional, incompressible, inviscid flow. More in-
formation on this subject can be found in Batchelor 2000, and Milne-Thompson
2011.

5.2 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 . (5.1)

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 5.1.) Suppose
that fluid is neither created nor destroyed in the region, R (say), bounded by these
curves. Because 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 (in Figure 5.1) 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. 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. Because 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 point P has Cartesian coordinates (x, y), then

119
120 Theoretical Fluid Mechanics

B
A
Figure 5.1
Two-dimensional flow.

we can write
= (x, y, t). (5.2)
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 5.2.)
Let 1 and 2 be the fluxes from right to left across curves AP1 and AP2 . Using
similar arguments to those employed previously, 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 . If P1 and P2 both lie on the same streamline then the flux across P1 P2
is zero, because 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 suciently 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,
2 1
v = . (5.3)
s
In the limit s 0, the perpendicular velocity from right to left across ds becomes

d
v = . (5.4)
ds
Thus, in Cartesian coordinates, by considering infinitesimal arcs parallel to the x-
Two-Dimensional Incompressible Inviscid Flow 121

P2
P1

A
Figure 5.2
Two-dimensional flow.

and y-axes, we deduce that


vx = , (5.5)
y

vy = . (5.6)
x
These expressions can be combined to give

v = ez = z . (5.7)

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 [because
(AB) 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 , (5.8)

where
vy v x
z = . (5.9)
x y
Thus, it follows from Equations (5.5) and (5.6) that

2 2
z = + = 2 . (5.10)
x 2 y 2
122 Theoretical Fluid Mechanics

Hence, irrotational two-dimensional flow is characterized by

2 = 0. (5.11)

When expressed in terms of cylindrical coordinates (see Section C.3), Equa-


tion (5.7) yields
v = vr (r, , t) er + v (r, , t) e , (5.12)

where
1
vr = , (5.13)
r

v = . (5.14)
r
Moreover, the vorticity is = z ez , where


1 1 2
z = r + 2 . (5.15)
r r r r 2

5.3 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 . (5.16)

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 4.15.) Hence,


vx = , (5.17)
x

vy = . (5.18)
y

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 5.2.) Hence,


vx = , (5.19)
y

vy = . (5.20)
x
Two-Dimensional Incompressible Inviscid Flow 123

Finally, if the flow is both irrotational and incompressible then Equations (5.17)
(5.18) and (5.19)(5.20) hold simultaneously, which implies that

= , (5.21)
x y

= . (5.22)
x y
It immediately follows, from the previous two expressions, that

2 2 2 2
= = = 2, (5.23)
x 2 x y y x y
or
2 2
+ = 0. (5.24)
x 2 y 2
Likewise, it can also be shown that
2 2
+ = 0. (5.25)
x 2 y 2
We conclude that, for two-dimensional, irrotational, incompressible flow, the veloc-
ity potential and the stream function both satisfy Laplaces equation. Equations (5.21)
and (5.22) also imply that
= 0. (5.26)
In other words, the contours of the velocity potential and the stream function cross at
right-angles.

5.4 Two-Dimensional Uniform Flow


Consider a steady two-dimensional flow pattern that is uniform: in other words, a
pattern which is 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 , (5.27)

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 (5.5) and
(5.6), that the stream function for steady uniform flow takes the form

(x, y) = V0 (sin 0 x cos 0 y) . (5.28)

When written in terms of cylindrical coordinates, this becomes

(r, ) = V0 r sin( 0 ). (5.29)


124 Theoretical Fluid Mechanics

Note, from Equation (5.28), that 2 /x 2 = 2 /y 2 = 0. Thus, it follows from


Equation (5.10) that uniform flow is irrotational. Hence, according to Section 4.15,
such flow can also be derived from a velocity potential. In fact, it is easily demon-
strated that
(r, ) = V0 r cos( 0 ). (5.30)

5.5 Two-Dimensional Sources and Sinks


Consider a uniform line source, coincident with the z-axis, that emits fluid isotropi-
cally 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 5.3.) In other words, we expect

v = vr (r) er , (5.31)

where r = (x 2 + y 2 )1/2 . 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,

v dS = 2 r vr (r) = Q, (5.32)
S

which implies that


Q
vr (r) = . (5.33)
2 r
According to Equations (5.13) and (5.14), the stream function associated with a
line source of strength Q that is coincident with the z-axis is
Q
(r, ) = . (5.34)
2
Note that the streamlines, = c, are directed radially away from the z-axis, as illus-
trated in Figure 5.3. Note, also, that the stream function associated with a line source
is multivalued. However, this does not cause any particular diculty, because the
stream function is continuous, and its gradient is single valued.
It follows from Equation (5.34) that /r = 2 / 2 = 0. Hence, according to
Equation (5.15), 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
Q
(r, ) = ln r. (5.35)
2
A uniform line sink, coincident with the z-axis, which absorbs fluid isotropically
Two-Dimensional Incompressible Inviscid Flow 125

Figure 5.3
Streamlines of the flow generated by a line source coincident with the z-axis.

at the steady rate of Q unit volumes per unit length per unit time has an associated
steady flow pattern
Q
v= er , (5.36)
2 r
whose stream function is
Q
(r, ) = . (5.37)
2
This flow pattern is also irrotational, and can be derived from the velocity potential
Q
(r, ) = ln r. (5.38)
2
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 arrange-
ment is known as a doublet or dipole 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 Q
Q (r) = = tan1 (y/x) (5.39)
2 2
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
126 Theoretical Fluid Mechanics

located at r = r0 . Furthermore, the stream function associated with a line sink of


strength Q located at r = r0 is Q (r r0 ). 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), (5.40)
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
D sin( 0 )
(r, ) = , (5.41)
2 r
where D = Q d is termed the strength of the dipole source. The previous stream
function is antisymmetric across the line = 0 joining the source to the sink. It
follows that the associated dipole flow pattern,
D cos( 0 )
vr (r, ) = , (5.42)
2 r2
D sin( 0 )
v (r, ) = , (5.43)
2 r2
is symmetric across this line. Figure 5.4 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 o like 1/r 2 .
A dipole flow pattern is necessarily irrotational because it is a linear superposition
of two irrotational flow patterns. The associated velocity potential is
D cos( 0 )
(r, ) = . (5.44)
2 r

5.6 Two-Dimensional Vortex Filaments


Consider a vortex filament of intensity that is coincident with the z-axis. By sym-
metry, we expect the associated flow pattern to circulate isotropically around the
filament. (See Figure 5.5.) In other words, we expect
v = v (r) e . (5.45)
#
According to Section 4.13, the circulation, vdr, 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

r = v dr = 2 r v (r) = , (5.46)
Two-Dimensional Incompressible Inviscid Flow 127

1
0.8
0.6
0.4
0.2
y 0
0.2
0.4
0.6
0.8
1
1 0.5 0 0.5 1
x

Figure 5.4
Streamlines of the flow generated by a dipole line source (with D > 0) coincident
with the z-axis, and aligned along the x-axis. The flow is outward along the posi-
tive x-axis and inward along the negative x-axis. Positive and negative contours are
shown as solid and dashed lines, respectively.

or

v (r) = . (5.47)
2 r
According to Equations (5.13) and (5.14), the stream function associated with a
vortex filament of intensity that is coincident with the z-axis is

(r, ) = ln r. (5.48)
2
Note that the streamlines, = c, circulate around the z-axis, as illustrated in Fig-
ure 5.5.
It can be seen, from Equation (5.48), that (/r)(r /r) = / = 0. Hence,
it follows from Equation (5.15) that z = 2 = 0. In other words, the flow
pattern associated with a straight vortex filament is irrotational. This is a somewhat
surprising result, because there is a net circulation of the flow around the filament,
and, according to Section 4.13, 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: that is,

/( a 2) ra
z = . (5.49)
0 r>a
128 Theoretical Fluid Mechanics

Figure 5.5
Streamlines of the flow generated by a line vortex coincident with the z-axis.

Note that the intensity of the filament (i.e., the product of its vorticity and cross-
sectional area) is still . According to Equation (5.15), and assuming that = (r),


1 d d /( a 2) ra
r = . (5.50)
r r dr 0 r>a

The solution that is well behaved at r = 0, and continuous (up to its first derivative)
at r = a, is 
(/4) (r 2/a 2 1) ra
(r, ) = . (5.51)
(/2) ln(r/a) r>a
This expression is equivalent to Equation (5.48) (apart from an unimportant additive
constant) outside the filament, but diers inside. The associated circulation velocity,
v (r) = /r, is

(/2) (r/a 2) ra
v (r) = , (5.52)
(/2) (1/r) r>a

whereas the circulation, r (r) = 2 r v (r), is written



(r/a)2 ra
r (r) = . (5.53)
r>a

Thus, we conclude that the flow pattern associated with a straight vortex filament is
Two-Dimensional Incompressible Inviscid Flow 129

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 (1.149), yields

dp v2
= , (5.54)
dr r

which can be integrated to give


 v2
p = p dr, (5.55)
r r

where p is the pressure at infinity. Making use of expression (5.52), we obtain



p (/2) (/2 a)2 (2 r 2 /a 2 ) ra
p(r) = . (5.56)
p (/2) (/2 a)2 (a/r)2 r>a

It follows that the minimum pressure occurs at the center of the vortex (r = 0), and
takes the value
$ %2
p0 = p . (5.57)
2 a
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 . (5.58)
p 2

Finally, because the flow pattern outside a straight vortex filament is irrotational,
it can be derived from a velocity potential. It is easily demonstrated that the appro-
priate potential takes the form


(r, ) = . (5.59)
2

Note that the previous potential is multivalued. However, this does not cause any
particular diculty, because the potential is continuous, and its gradient is single
valued.
130 Theoretical Fluid Mechanics

5.7 Two-Dimensional Irrotational Flow in Cylindrical Coordi-


nates
In a two-dimensional flow pattern, we can automatically satisfy the incompressibil-
ity 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 (5.10) yields

2 = 0. (5.60)

In cylindrical coordinates, because = (r, , t), this expression implies that (see
Section C.3)

1 1 2
r + 2 = 0. (5.61)
r r r r 2
Let us search for a separable solution of Equation (5.61) of the form

(r, ) = R(r) (). (5.62)

It is easily seen that



r d dR 1 d 2
r = , (5.63)
R dr dr d 2
which can only be satisfied if


d dR
r r = m 2 R, (5.64)
dr dr
d 2
= m 2 , (5.65)
d 2
where m 2 is an arbitrary (positive) constant. The general solution of Equation (5.65)
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 unphysi-
cal). It follows that m can only take integer values (and that m 2 must be a positive,
rather than a negative, constant). The general solution of Equation (5.64) is a linear
combination of r m and rm factors, except for the special case m = 0, when it is a lin-
ear combination of r 0 and ln r factors. Thus, the general stream function for steady
two-dimensional irrotational flow (that extends over all values of ) takes the form

(r, ) = 0 + 0 ln r + (m r m + m rm ) sin[m ( m )], (5.66)
m>0

where m , m , and m are arbitrary constants. We can recognize the first few terms
on the right-hand side of the previous 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
Two-Dimensional Incompressible Inviscid Flow 131
5
4
3
2
1
y/a 0
1
2
3
4
5
5 4 3 2 1 0 1 2 3 4 5
x/a

Figure 5.6
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.

pattern generated by a vortex filament of intensity 2 0 , coincident with the z-axis.


(See Section 5.6.) 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 5.4.) Finally, the term 1 sin( 1 )/r corresponds to a dipole flow pattern.
(See Section 5.5.)
The velocity potential associated with the irrotational stream function (5.66) sat-
isfies [see Equations (4.89) and (5.7)]

1
= , (5.67)
r r
1
= . (5.68)
r r

It follows that

(r, ) = 0 0 + (m r m m rm ) cos[m ( 0 )]. (5.69)
m>0
132 Theoretical Fluid Mechanics

5.8 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. 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 5.4)
(r, ) V0 r sin as r/a . (5.70)
Given that the fluid velocity field a large distance upstream of the cylinder is irrota-
tional (because we have already seen that the flow pattern associated with uniform
flow is irrotationalsee Section 5.4), it follows from the Kelvin circulation theorem
(see Section 4.14) that the velocity field remains irrotational as it is convected past
the cylinder. Hence, according to Section 5.2, the stream function of the flow satisfies
Laplaces equation,
2 = 0. (5.71)
The appropriate boundary condition at the surface of the cylinder is simply that the
normal fluid velocity there be zero, because 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, (5.72)
because is undetermined to an arbitrary additive constant. It follows that we are
searching for the most general solution of Equation (5.71) that satisfies the bound-
ary conditions (5.70) and (5.72). Comparison with Equation (5.66) reveals that this
solution takes the form
& $r % $ r a% '
(r, ) = V0 a ln sin , (5.73)
a a r
where

= , (5.74)
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 the curl theorem.) Let us assume that 0, for the sake of
definiteness.
Figure 5.65.8 show streamlines of the flow calculated for various dierent val-
ues of the normalized circulation, . For < 2, there exist a pair of points on the
Two-Dimensional Incompressible Inviscid Flow 133
5
4
3
2
1
y/a 0
1
2
3
4
5
5 4 3 2 1 0 1 2 3 4 5
x/a

Figure 5.7
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.

surface of the cylinder at which the flow speed is zero. These are known as stagnation
points, and can be located in Figures 5.6 and 5.7 as the points at which streamlines
intersect the surface of the cylinder at right-angles. The tangential fluid velocity at
the surface of the cylinder is


vt () = v (a, ) =  = V0 ( + 2 sin ). (5.75)
r r=a
The stagnation points correspond to the points at which vt = 0 (because 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 o the surface of the cylinder, as illustrated in Figure 5.8 (the stagnation point
corresponds to the point at which two streamlines cross at right-angles).
The irrotational form of Bernoullis theorem, (4.97), 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
1  
p = p0 + V02 v 2 , (5.76)
2
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
1  
P() = p(a, ) = p0 + V02 vt2 = p1 + V02 (cos 2 2 sin ) , (5.77)
2
134 Theoretical Fluid Mechanics

where p1 = p0 (1/2) V02 (1 + 2 ). The net force per unit length exerted on the
cylinder by the fluid has the Cartesian coordinates

F x = P cos a d, (5.78)

Fy = P sin a d. (5.79)

Thus, it follows from Equation (5.77) that

F x = 0, (5.80)
Fy = 2 V02 a = V0 (). (5.81)

The component of the force that 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 that the fluid exerts in a direction per-
pendicular to that of the unperturbed flow is usually called lift. Hence, the previous
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. 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). Because the flow pattern
is initially irrotational, and because the flow pattern well upstream of the cylinder is
assumed to remain irrotational, the Kelvin circulation theorem indicates that 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
 

dC v p 1 2
= dr = + v + v dr = v dr, (5.82)
dt C t C 2 C

where use has been made of Equation (4.84) (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,

dC
= z v dS. (5.83)
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 Equation (4.85), the fact that = z ez , and the fact that
Two-Dimensional Incompressible Inviscid Flow 135
5
4
3
2
1
y/a 0
1
2
3
4
5
5 4 3 2 1 0 1 2 3 4 5
x/a

Figure 5.8
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.

/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, because C = 0 originally, as 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 previous 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 a manifestation of dAlemberts
paradox, which was introduced in Section 4.5. dAlemberts result is paradoxical
because it would seem, at first sight, to be a reasonable approximation to neglect
viscosity altogether 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.
136 Theoretical Fluid Mechanics

V
x
a

Figure 5.9
Cylinder moving through an inviscid fluid

5.9 Motion of a Submerged Cylinder


Consider the situation, illustrated in Figure 5.9, in which an impenetrable rigid cylin-
der of radius a and infinite length, whose symmetry axis runs parallel to the z-
direction, is moving through an incompressible, inviscid fluid at the time-dependent
velocity V = V x (t) e x . Assuming that the fluid and cylinder were both initially sta-
tionary, it follows that the fluid velocity field was initially irrotational. Thus, ac-
cording to the Kelvin circulation theorem, the fluid velocity field remains irrotational
when the cylinder starts to move. Thus, we can write
v = , (5.84)
where v is the fluid velocity. Moreover, because the fluid is incompressible, we have
v = 2 = 0. (5.85)
Let x, y, z be Cartesian coordinates in the initial rest frame of the fluid, and let r,
be cylindrical coordinates in a frame of reference that co-moves with the cylinder,
as shown in Figure 5.9. In the following, all calculations are performed in the rest
frame. We expect the fluid a long way from the cylinder to remain stationary. In
other words,
(r, , t) 0 as r . (5.86)
Moreover, because the cylinder is impenetrable, we require that
V er = v er |r=a , (5.87)
Two-Dimensional Incompressible Inviscid Flow 137

or 

 = V x cos . (5.88)
r r=a
It is easily demonstrated that the solution to Equation (5.85), subject to the boundary
conditions (5.86) and (5.88), is

a2
(r, , t) = V x (t) cos . (5.89)
r
Hence,
a2
vr (r, , t) = V x (t) cos , (5.90)
r2
a2
v (r, , t) = V x (t) 2 sin . (5.91)
r
The general form of Bernoullis theorem, (4.96), which applies to an irrotational
flow field, yields
p 1 2 p0
+ v = , (5.92)
2 t
where is the uniform fluid mass density, and p0 the fluid pressure at infinity. Thus,
the pressure distribution at the surface of the cylinder can be written
 

p(a, , t) = p0 v 2  +
1
 V |r=a . (5.93)
2 r=a t r=a
The final term on the right-hand side of the previous equation arises because
 
 
 =  V . (5.94)
t  x,y t r,
Hence, we obtain
1 dV x
p(a, , t) = p0 V x2 + V x2 cos(2 ) + a cos . (5.95)
2 dt
The net force per unit length exerted on the cylinder by the fluid has the Cartesian
components

F x (t) = p(a, , t) cos a d, (5.96)

Fy (t) = p(a, , t) sin a d. (5.97)

It follows that
dV x
F x (t) = a 2 , (5.98)
dt
Fy (t) = 0, (5.99)
138 Theoretical Fluid Mechanics

or
dV
F = m , (5.100)
dt
where m = a 2 is the mass per unit length of the fluid displaced by the cylinder.
Suppose that the cylinder is subject to an external (i.e., not due to the fluid) force
per unit length Fext . The equation of motion of the cylinder is thus
dV
m = F + Fext , (5.101)
dt
where m is the cylinders mass per unit length. The previous two equations can be
combined to give
dV
(m + m ) = Fext . (5.102)
dt
In other words, the cylinder moves under the action of the external force, Fext , as
if its mass per unit length were m + m , rather than m. Here, m + m is commonly
referred to as the cylinders virtual mass (per unit length), whereas m is termed the
added mass (per unit length).
The origin of added mass is easily explained. According to Equations (5.90) and
(5.91), the total kinetic energy per unit length of the fluid surrounding the cylinder is
  $ a %4
1 2 1 1
Kfluid = v dV = 2 V x2 r dr = m V x2 . (5.103)
fluid 2 a 2 r 2
However, the kinetic energy per unit length of the cylinder is
1
Kcylinder = m V x2 . (5.104)
2
Thus, the total kinetic energy per unit length is
1
K = Kfluid + Kcylinder = (m + m ) V x2 . (5.105)
2
In other words, the kinetic energy of the fluid surrounding the cylinder can be ac-
counted for by supposing that a mass (per unit length) m of the fluid co-moves with
the cylinder, and that the remainder of the fluid remains stationary. This entrained
fluid mass accounts for the added mass of the cylinder. Note that the added mass
is independent of the speed of the cylinder (i.e., it is the same whether the cylinder
moves slowly or rapidly.) In the present case, the added mass is equal to the mass of
the displaced fluid. However, this is not a general rule. (In general, the added mass
of a object moving through an inviscid fluid is proportional to the displaced mass,
but the constant of proportionality is not necessarily unity, and depends on the shape
of the object.)
Let us generalize the previous calculation to allow the cylinder to move in any
direction in the x-y plane: that is,

V(t) = V x (t) e x + Vy (t) ey . (5.106)


Two-Dimensional Incompressible Inviscid Flow 139

Furthermore, let the fluid possess the initial circulation in the x-y plane. According
to the Kelvin circulation theorem, this circulation remains constant in time. Thus, we
must now solve Equation (5.85) subject to the boundary conditions

(r, , t) as r . (5.107)
2 r
and 

 = V x cos Vy sin . (5.108)
r r=a
It is easily demonstrated that the appropriate solution is
a2 a2
(r, , t) = + V x (t) cos + Vy (t) sin . (5.109)
2 r r r
Hence,
a2 a .2
vr (r, , t) = V x (t) cos + V y (t) sin , (5.110)
r2 r2
a2 a2
v (r, , t) = + V x (t) 2 sin Vy (t) 2 cos . (5.111)
2 r r r
Bernoullis theorem yields
p 1 2 p0
+ v + g r sin = , (5.112)
2 t
where we have assumed that the fluid and cylinder are both situated in a gravitational
field of uniform acceleration g = g ey . Thus, the pressure distribution at the surface
of the cylinder can be written
 

p(a, , t) = p0 v 2  g a sin
1
 V |r=a , (5.113)
2 r=a t r=a
which yields
$ %2
1
p(a, , t) = p0 V +
2
+ (V x2 Vy2 ) cos(2 ) + 2 V x Vy sin(2 )
2 2 a
dV x dVy
+ a cos + a sin g a sin
dr dr
 
V x sin Vy cos . (5.114)
a
It follows from Equations (5.96) and (5.97) that the force per unit length exerted on
the cylinder by the fluid has the Cartesian components
dV x
F x (t) = Vy m , (5.115)
dt
dVy
Fy (t) = V x m + m g. (5.116)
dt
140 Theoretical Fluid Mechanics

Here, the first terms on the right-hand sides of the previous two equations are the
components of the lift (per unit length) acting on the cylinder, due to the fluid cir-
culation, whereas the final term on the right-hand side of the second equation is the
buoyancy force (per unit length) acting on the cylinder. The cylinders equation of
motion,
dV x
m = Fx, (5.117)
dt
dVy
m = Fy m g, (5.118)
dt
leads to
d 2X
= 2 Y, (5.119)
dt 2


d 2Y m m
= 2
X + g, (5.120)
dt 2 m + m
where = /(m + m ), and (X, Y) are the Cartesian components of the cylinders
axis. Let us assume that X = Y = dX/dt = dY/dt = 0 at t = 0. It follows that


m m g
X(t) = [sin( t) t] , (5.121)
m + m 2


m m g
Y(t) = [1 cos( t)] . (5.122)
m + m 2
Consider, first, the case in which there is no circulation of the flow: that is, = 0.
In this case, the previous two equations reduce to

X(t) = 0, (5.123)



mm
Y(t) = g t 2. (5.124)
m + m
In other words, the cylinder moves vertically (i.e., in the y-direction) with the con-
stant acceleration

s1
ay = g, (5.125)
s+1
where s = m/m is the cylinders specific gravity. It follows that if the cylinder is
much denser than the fluid (i.e., s
1) then it accelerates downward at the acceler-
ation due to gravity: that is, ay = g. However, if the cylinder is much less dense
than the fluid (i.e., 0 < s  1) then it accelerates upward at the acceleration due to
gravity: that is ay = +g. Note that, in the latter case, the upward acceleration is lim-
ited by the cylinders added mass (i.e., in the absence of added mass, the acceleration
would be infinite.)
In the general case, in which the fluid circulation is non-zero, the trajectory of
the cylinder is a cycloid. In particular, assuming that m > m , the lift acting on
Two-Dimensional Incompressible Inviscid Flow 141

Figure 5.10
Inviscid flow past a wedge.

the cylinder prevents it from falling through the fluid a distance greater than 2 [(m
m )/(m + m )] (g/ 2). Once the cylinder has fallen through this distance, it starts
to rise again, until it attains its original height, and the motion then repeats itself
ad infinitum. Moreover, the cylinder simultaneously moves horizontally (i.e. in the
x-direction) at the mean velocity [(m m )/(m + m )] (g/).

5.10 Inviscid Flow Past a Semi-Infinite Wedge


Consider the situation, illustrated in Figure 5.10, 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.
Because the flow is two-dimensional, incompressible, and irrotational, it can be
represented in terms of a stream function that satisfies Laplaces equation. More-
over, in cylindrical coordinates, this equation takes the form (5.61). The boundary
conditions on the stream function are

(r, /2) = (r, 2 /2) = (r, ) = 0. (5.126)

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,
142 Theoretical Fluid Mechanics

y 0

3
3 2 1 0 1 2 3
x

Figure 5.11
Streamlines of inviscid incompressible irrotational flow past a 90 wedge.

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 the well-known result that
is constant on a streamline. It is easily demonstrated that

A
(r, ) = r 1+m sin [(1 + m) ( )] (5.127)
1+m
is a solution of Equation (5.61). Moreover, this solution satisfies the boundary con-
ditions provided (1 + m) (1 /2) = 1, or

m= . (5.128)
2
Because, as is well known, the solutions to Laplaces equation (for problems with
well-posed boundary conditions) are unique (Riley 1974), we can be sure that Equa-
tion (5.127) 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 r m , (5.129)

where m 0. Note that the tangential velocity is zero at the apex of the wedge.
Because the normal velocity is also zero at this point, we conclude that the apex is
a stagnation point of the flow. Figure 5.11 shows the streamlines of the flow for the
case = 1/2.
Two-Dimensional Incompressible Inviscid Flow 143

 x

Figure 5.12
Inviscid flow over a wedge.

5.11 Inviscid Flow Over a Semi-Infinite Wedge


Consider the situation illustrated in Figure 5.12 in which an incompressible irrota-
tional 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.
Because 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. (5.130)

These boundary conditions ensure that the normal velocity at the surface of the
wedge is zero. It is easily demonstrated that
A
(r, ) = r 1m cos [(1 m) ( /2)] (5.131)
1m
is a solution of Laplaces equation, (5.61). Moreover, this solution satisfies the
boundary conditions provided that (1 m) (1 /2) = 1/2, or

m= , (5.132)
1 + 
144 Theoretical Fluid Mechanics

y 0

3
3 2 1 0 1 2 3
x

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

where  = 1 . Because the solutions to Laplaces equation are unique, we


can again be sure that Equation (5.131) 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 , (5.133)
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 5.13 shows the streamlines of the flow for the case
= 1/2.

5.12 Two-Dimensional Jets


Consider a two-dimensional jet of width h and uniform speed v that impinges on a
rigid plate, at an angle to the normal to the plate, and divides into two streams,
running along the plate, whose widths are ultimately uniform, and equal to h1 and
h2 . (See Figure 5.14.) We expect the pressure well upstream of the plate to be atmo-
spheric all the way across the jet, because the jet is not yet decelerating. Likewise,
we expect the pressure well downstream of the strike point to be atmospheric all
the way across the two streams that run along the plate, because the streams are no
Two-Dimensional Incompressible Inviscid Flow 145

h1

h v

h2

Figure 5.14
A two-dimensional jet impinging on a rigid plate.

longer accelerating (and, hence, their widths are constant). It immediately follows
from Bernoullis theorem that the asymptotic flow speed in the two streams is v, as
indicated in Figure 5.14.
Fluid continuity demands that h v = h1 v + h2 v, or

h = h1 + h2 . (5.134)

Moreover, because the plate can only exert a normal force on the fluid, it follows
that the parallel (to the plate) momentum flux of the incident jet must equal the net
parallel momentum flux of the two streams. In other words,

h v 2 sin = h1 v 2 h2 v 2 , (5.135)

or
h sin = h1 h2 . (5.136)
Equations (5.134) and (5.136) can be combined to give
h
h1 = (1 + sin ), (5.137)
2
h
h2 = (1 sin ). (5.138)
2
Let us apply the Euler momentum theorem (see Section 4.4) to the flux tube
shown in Figure 5.14, treating the plate as one of the walls of the tube. By symmetry,
146 Theoretical Fluid Mechanics

h1 v 2

h1 /2

hv2
O
F

h2 /2

h2 v 2

Figure 5.15
Determination of the center of pressure.

the net force (per unit width perpendicular to the page) exerted across the walls and
ends of the tube is the resultant of the three forces shown in Figure 5.15. It is easily
demonstrated that this net force, F, is normal to the plate, as shown in the figure, and
of magnitude
F = h v 2 cos . (5.139)
Obviously, the force in question corresponds to the force exerted by the plate on the
fluid. An equal and opposite force is exerted by the fluid on the plate. (Note that there
is no net pressure force on the flux tube because it is entirely surrounded by fluid at
the same pressure: i.e., atmospheric pressure.) The line of action of the resultant
force passes through the so-called center of pressure, C. This is displaced slightly
from the point, O, where the middle of the jet strikes the plate. (See Figure 5.15.)
Let d be the distance OC. We can calculate d by demanding that the net moment of
the three forces shown in Figure 5.15 about point C be zero. In other words,

h1 h2
h v 2 cos d h1 v 2 + h2 v 2 = 0, (5.140)
2 2
which yields
h
d= tan . (5.141)
2
Two-Dimensional Incompressible Inviscid Flow 147

5.13 Exercises
5.1 For the case of the two-dimensional motion of an incompressible fluid, deter-
mine the condition that the velocity components

v x = a x + b y,
vy = c x + d y

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

v x = 2 c x y,
vy = c (a 2 + x 2 y 2 )

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.
5.3 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 dp 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
2 r2
p(r) = + p0 ,
2 a4
where p0 is the pressure at the center of the vortex.
5.4 Consider the cylindrical vortex discussed in Exercise 5.3. If p(r) is the pres-
sure at radius r external to the vortex, demonstrate that

2
p(r) = + p ,
2r2
where p is the pressure at infinity.
5.5 Show that the stream function for the cylindrical vortex discussed in Ex-
ercises 5.3 and 5.4 is (r) = (1/2) a 2 ln(r/a) for r > a, and (r) =
(1/4) (r 2 a 2 ) for r a.
148 Theoretical Fluid Mechanics

5.6 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 v x

x y

at the center of the circle. Neglect terms of order r 3 .


5.7 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 sat-
isfied when vr = a k r n exp[k (n + 1) ] and v = a r n exp[k (n + 1) ].
Determine the stream function, and show that the fluid speed at any point is

(n + 1) 1 + k 2 /r,

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


+ [, ] = 0,
t
2 = ,
2 = + 2 ,

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


5.10 For irrotational, incompressible, inviscid motion in two dimensions show that

q q = q 2 q,

where q = |v|.
6
Two-Dimensional Potential Flow

6.1 Introduction
This chapter describes the use of complex analysis to facilitate calculations in two-
dimensional, incompressible, irrotational fluid dynamics. Incidentally, incompress-
ible, irrotational flow is usually referred to as potential flow, because the associ-
ated velocity field can be represented in terms of a velocity potential that satisfies
Laplaces equation. (See Section 4.15.) In the following, all flow patterns are as-
sumed to be such that the z-coordinate is ignorable. In other words, the fluid ve-
locity 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 in the z-direction, and have uniform cross-sections. Hence, it
is only necessary to specify the locations of line sources, vortex filaments, and solid
surfaces in the x-y plane. More information on the use of complex analysis in two-
dimensional fluid mechanics can be found in Batchelor 2000, Milne-Thomson 1958,
Milne-Thomson 2011, and Lamb 1993.

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)

where r = x 2 + y 2 and = tan1 (y/x) are the modulus and argument of z, re-
spectively, but can also be identified with the corresponding plane polar coordinates.
Finally, Eulers theorem (Riley 1974),

e i = cos + i sin , (6.3)

149
150 Theoretical Fluid Mechanics

implies that

x = r cos , (6.4)
y = r sin . (6.5)

We can define functions of the complex variable, F(z), in the same way that we
define functions of a real variable. For instance,

F(z) = z 2 , (6.6)
1
F(z) = . (6.7)
z
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) = z 2 , (6.9)

then
F(x + i y) = (x + i y)2 = (x 2 y 2 ) + 2 i x y, (6.10)
giving

(x, y) = x 2 y 2 , (6.11)
(x, y) = 2 x y. (6.12)

6.3 Cauchy-Riemann Relations


We can define the derivative of a complex function in the same way that we define
the derivative of a real function. In other words,
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 dierentiable) 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 dierent possible approach directions, but if we look at a
regular complex function, F(z) = z 2 (say), then

dF
= 2z (6.14)
dz
Two-Dimensional Potential Flow 151

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 (x + x, y) + i (x + x, y) (x, y) i (x, y)
= lim |x|0
dz x

= +i . (6.15)
x x
Suppose that we now approach z along the imaginary axis, so that z = i y. We get

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


= lim |y|0
dz i y

= i + . (6.16)
y y

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


= , (6.17)
x y

= . (6.18)
x y

These expressions are called the Cauchy-Riemann relations, and are, in fact, su-
cient to ensure that all possible ways of taking the limit (6.13) give the same result
(Riley 1974).
The Cauchy-Riemann relations can be combined to give

2 2 2 2
+ = + = + = 0. (6.19)
x 2 y 2 x 2 y 2 x x y y

In other words,

2 = 0, (6.20)
= 0,
2
(6.21)
= 0. (6.22)

It follows that the real and imaginary parts of a well-behaved function of the complex
variable both satisfy Laplaces equation. Furthermore, the contours of these functions
cross at right-angles.
152 Theoretical Fluid Mechanics

6.4 Complex Velocity Potential


Equations (6.17)(6.18) are identical to Equations (5.21)(5.22). 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, (6.23)
where V0 is real. It follows that

(r, ) = V0 r cos , (6.24)


(r, ) = V0 r sin . (6.25)

It can be seen, by comparison with the analysis of Section 5.4, that the complex ve-
locity potential (6.23) corresponds to uniform flow of speed V0 directed along the
x-axis. Furthermore, as is easily demonstrated, the complex velocity potential asso-
ciated 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 .
Suppose that
Q
F(z) = ln z, (6.26)
2
where Q is real. Because ln z = ln r + i (Riley 1974), it follows that
Q
(r, ) = ln r, (6.27)
2
Q
(r, ) = . (6.28)
2
Thus, according to the analysis of Section 5.5, the complex velocity potential (6.26)
corresponds to the flow pattern of a line source, of strength Q, located at the origin.
(See Figure 5.3.) 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 . It can be seen, from Equation (6.26), that the complex velocity
potential of a line source is singular at the location of the source.
Suppose that

F(z) = i ln z, (6.29)
2
where is real. It follows that

(r, ) = , (6.30)
2

(r, ) = ln r. (6.31)
2
Thus, according to the analysis of Section 5.6, the complex velocity potential (6.29)
Two-Dimensional Potential Flow 153

corresponds to the flow pattern of a vortex filament of intensity located at the


origin. (See Figure 5.5.) As a simple generalization of this result, the complex
potential of a vortex filament, of intensity , located at the point (x0 , y0 ), is F(z) =
i (/2) ln(z z0 ), where z0 = x0 + i y0 . According to Equation (6.29), the complex
velocity potential of a vortex filament is singular at the location of the filament.
Suppose, finally, that

$z%
a2
F(z) = V0 z + +i ln , (6.32)
z 2 a
where V0 , a, and , are real. It follows that


a2
(r, ) = V0 r + cos , (6.33)
r 2

$r%
a2
(r, ) = V0 r sin + ln . (6.34)
r 2 a
Thus, according to the analysis of Section 5.8, the complex velocity potential (6.32)
corresponds to uniform flow of unperturbed speed V0 , running parallel to the x-axis,
around an impenetrable circular cylinder of radius a, centered on the origin. (See
Figures 5.6, 5.7, and 5.8.) Here, is the circulation of the flow about the cylinder.
It can be seen that = 0 on the surface of the cylinder (r = a), which ensures
that the normal velocity is zero on this surface (because the surface corresponds to a
streamline), as must be the case if the cylinder is impenetrable.

6.5 Complex Velocity


Equations (5.17), (5.20), and (6.15) imply that
dF
= +i = v x + i vy . (6.35)
dz x x
Consequently, dF/dz is termed the complex velocity. It follows that
  2
 dF  = v 2 + v 2 = v 2 , (6.36)
 dz  x y

where v is the flow speed.


A stagnation point is defined as a point in a flow pattern at which the flow speed,
v, falls to zero. (See Section 5.8.) According to the previous expression,
dF
=0 (6.37)
dz
at a stagnation point. For instance, the stagnation points of the flow pattern produced
154 Theoretical Fluid Mechanics

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 (6.32) to
zero. It follows that the stagnation points satisfy the quadratic equation


dF a2
= V0 1 2 + i = 0. (6.38)
dz z 2 z

The solutions are 


z
= i 1 2 , (6.39)
a
where = /(4 V0 a), with the proviso that |z|/a > 1, because the region |z|/a < 1
is occupied by the cylinder. Thus, if  1 then there are two stagnation points on
the surface of the cylinder at x/a = 1 2 and y/a = . 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.
According to Section 4.15, Bernoullis theorem in an steady, irrotational, incom-
pressible fluid takes the form
1 2
p+ v = p0 , (6.40)
2
where p0 is a uniform constant. Here, gravity (and any other body force) has been
neglected. Thus, the pressure distribution in such a fluid can be written
 
1  dF  2
p = p0   . (6.41)
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 fol-
lows that both stream functions satisfy Laplaces equation: that is, 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, cor-
responding 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 in the fol-
lowing, 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
Two-Dimensional Potential Flow 155

Figure 6.1
Two line sources.

the resulting flow pattern is the sum of the complex potentials of each source taken
in isolation. Hence, from Section 6.4,

Q Q Q
F(z) = ln(z i a) ln(z + i a) = ln(z 2 + a 2 ). (6.42)
2 2 2

Thus, the stream function of the flow pattern (which is the imaginary part of the
complex potential) is


Q 2 xy
(x, y) = tan1 2 . (6.43)
2 x y2 + a2

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 previous stream function as that
generated by a single line source of strength Q, located at the point (0, a), in the
presence of a planar rigid boundary at y = 0. This follows because the stream func-
tion 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 Laplaces 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.
156 Theoretical Fluid Mechanics
6

y/a 3

0
3 2 1 0 1 2 3
x/a

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

The complex velocity associated with the complex velocity potential (6.42) is
dF Q z
= . (6.44)
dz z + a2
2

Hence, the flow speed at the boundary is


 
|x|/a
v(x, 0) =   (x, 0) =
dF Q
. (6.45)
dz a 1 + x 2 /a 2
It follows from Equation (6.41) (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 $ Q %2 x 2 /a 2
p(x, 0) =   = . (6.46)
2 dz y=0 2 a (1 + x 2 /a 2 )2

Thus, the excess force per unit length (in the z-direction) acting on the boundary in
the y-direction is
 
Q2 2 Q2
Fy = p(x, 0) dx = 2 d = . (6.47)
2 a (1 + 2 )2 4 a
The fact that the force is positive implies that the boundary is attracted to the source,
and vice versa.
Two-Dimensional Potential Flow 157

Figure 6.3
Two vortex filaments.

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
F(z) = i ln(z i a) i ln(z + i a) = i ln . (6.48)
2 2 2 z+ ia
Thus, the stream function of the flow pattern is
2
x + (y a)2
(x, y) = ln 2 . (6.49)
2 x + (y + a)2
As before, (x, 0) = 0, which implies that there is zero flow normal to the plane y = 0
(because the plane corresponds to a streamline). Hence, in the region y > 0, we could
interpret the previous 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 circular 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
158 Theoretical Fluid Mechanics
6

y/a 3

0
3 2 1 0 1 2 3
x/a

Figure 6.4
Stream lines of the two-dimensional flow pattern due to a vortex filament at (0, a) in
the presence of a rigid boundary at y = 0.

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 Laplaces 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. The
complex velocity potential in the analog problem is simply

Q Q Q Q (z b) (z c)
F(z) = ln(z b) ln(z c) + ln z = ln . (6.50)
2 2 2 2 z

Hence, writing z = r e i , the corresponding stream function takes the form



Q 1 (r b c/r) sin
(r, ) = tan . (6.51)
2 (r + b c/r) cos (b + c)

We require the surface of the cylinder, r = a, to be a streamline: that is, (a, ) =


constant. This is easily achieved by setting c = a 2 /b. Thus, the stream function
becomes
Q 1 (r/a a/r) sin
(r, ) = tan . (6.52)
2 (r/a + a/r) cos (b/a + a/b)
The corresponding streamlines in the region external to the cylinder are illustrated in
Figure 6.6.
Two-Dimensional Potential Flow 159

y y

Q Q Q Q
x x
a b c

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

1
y/a

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

Figure 6.6
Stream lines of the two-dimensional flow pattern due to a line source at (2 a, 0) in
the presence of a rigid circular cylinder of radius a centered on the origin.
160 Theoretical Fluid Mechanics

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 dierentiable) func-
tion. 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 cor-
responding 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 ), (6.53)


0 + d  = f (z0 + dz ). (6.54)

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. If |dz |, |dz |  1
then

d  = f  (z0 ) dz , (6.55)


d  = f  (z0 ) dz , (6.56)

where f  (z) = d f /dz. Hence,


d  dz
= . (6.57)
d  dz
Thus, it follows that
|d  | |dz |
= , (6.58)
|d  | |dz |
and
arg(d  ) arg(d  ) = arg(dz ) arg(dz ). (6.59)
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.
Suppose that F() = (, ) + i (, ) is a well-behaved function of the complex
variable . It follows that 2 = 2 = = 0. Hence, the functions (, ) and
(, ) can be interpreted as the velocity potential and stream function, respectively,
of some two-dimensional, incompressible, irrotational flow pattern, where and
are Cartesian coordinates. However, if = f (z), where f (z) is well-behaved, then
F() = F[ f (z)] = G(z) = ( ((x, y), where G(z) is also well-behaved. It
(x, y) + i
follows that 2( ( = (
= 2 ( = 0. In other words, the functions (
(x, y) and
Two-Dimensional Potential Flow 161

y
d 
d 
0

dz 
z0
dz 

Figure 6.7
A conformal map.

((x, y) can be interpreted as the velocity potential and stream function, respectively,

of some new, two-dimensional, incompressible, irrotational flow pattern, where x and


y are Cartesian coordinates. In other words, we can use a conformal map to convert
a given two-dimensional, incompressible, irrotational flow pattern into another, quite
dierent, pattern. Incidentally, 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 . (6.60)

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
F() = i ln (6.61)
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,
$z%
G(z) = F(i e z/a ) = i ln tanh , (6.62)
2 2a
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, and the point = (0, 1) is mapped
162 Theoretical Fluid Mechanics

D
A C 1
a D x C A A B

A B
E E

Figure 6.8
The conformal map = i e z/a .

to the origin. Moreover, if the line = 0 is a streamline in the -plane then the lines
y = a/2 are also streamlines in the z-plane. Thus, these lines could all correspond to
rigid boundaries. The stream function associated with the previous complex velocity
potential,
cosh( x a) cos( y/a)
(x, y) = ln , (6.63)
cosh( x/a) + cos( y/a)
is shown in Figure 6.9.
As a second example, consider the map

= z 2. (6.64)

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 a 2) to the
point z = (a, a). As we saw in Section 6.6, in the region > 0, the velocity potential
Q
F() = ln( 2 + 4 a 4 ), (6.65)
2
corresponds to the flow pattern generated by a line source of strength Q, located at
the point = (0, 2 a 2 ), in the presence of a rigid plane at = 0. Thus, the complex
velocity potential
Q
G(z) = F(z 2 ) = ln(z 4 + 4 a 4 ), (6.66)
2
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 previous complex potential,

Q 4 x y (x 2 y 2 )
(x, y) = tan1 4 , (6.67)
2 x 6 x2 y2 + y4 + 4 a4
Two-Dimensional Potential Flow 163

0.5
y/a

0.5
1 0.5 0 0.5 1
x/a
Figure 6.9
Stream lines of the two-dimensional flow pattern due to a vortex filament at the origin
in the presence of two rigid planes at y = a/2.

y/a

0
0 1 2 3
x/a

Figure 6.10
Stream lines of the two-dimensional flow pattern due to a line source at (a, a) in the
presence of two rigid planes at x = 0 and y = 0.
164 Theoretical Fluid Mechanics

is shown in Figure 6.10.


As a final example, consider the map
l2
z=+ , (6.68)

where l is real and positive. Writing = r e i , we find that
x = 2 l cosh[ln(r/l)] cos = (r + l 2 /r) cos , (6.69)
y = 2 l sinh[ln(r/l)] sin = (r l 2 /r) sin . (6.70)
Thus, the map converts the circle + = c in the -plane, where c > l, into the
2 2 2

ellipse
$ x %2 $ y %2
+ =1 (6.71)
a b
in the z-plane, where
a = 2 l cosh[ln(c/l)] = c + l 2 /c, (6.72)
b = 2 l sinh[ln(c/l)] = c l /c. 2
(6.73)
Note that the center of the ellipse lies at the origin, and its major and minor axes
run parallel to the x- and the y-axes, respectively. As we saw in Section 6.4, in the
-plane, the complex velocity potential


c2
F = V0 + , (6.74)

represents uniform flow of unperturbed speed V0 , running parallel to the -axis,
around a circular cylinder of radius c, centered on the origin. Thus, assuming that
c > l, in the z-plane, the potential represents uniform flow of unperturbed speed
V0 , running parallel to the x-axis [which follows because at large |z| the map (6.68)
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 a,
aligned along the x-axis, and minor radius b, aligned along the y-axis. Note that
l = (a 2 b 2 )1/2 /2 and c = (a + b)/2. The corresponding stream function in the
z-plane is

c2
(x, y) = V0 r sin , (6.75)
r
where
r = l exp(cosh1 p), (6.76)


1 y p
= tan , (6.77)
x [p 2 1]1/2
  1/2
x 2 /l 2 + y 2 /l 2 + 4 + [x 2 /l 2 + y 2 /l 2 + 4]2 16 x 2 /l 2 1/2

p = . (6.78)
8

Figure 6.11 shows the streamlines of the flow pattern calculated for c = 1.5 l.
Two-Dimensional Potential Flow 165
4

y/l 0

4
4 3 2 1 0 1 2 3 4
x/l

Figure 6.11
Stream lines of the two-dimensional flow pattern due to uniform flow parallel to the
x-axis around an elliptical cylinder.

6.8 SchwarzChristoel Theorem


The SchwarzChristoel theorem is an important mathematical result that allows
a polygonal boundary in the z-plane to be mapped conformally onto the real axis,
= 0, in the -plane. It is conventional to map the region inside the polygon in the
z-plane onto the upper half, > 0, of the -plane. If the interior angles of the polygon
are , , , then the appropriate map is

d
= K ( a)1/ ( b)1/ ( c)1/ , (6.79)
dz

where K is a constant, and a, b, c, are the (real) values of that correspond to the
vertices of the polygon (Milne-Thompson 2011).
It is often convenient to take the point in the -plane that corresponds to one
of the vertices of the polygonsay, that given by = ato be at infinity. In this
case, the factor ( a) in Equation (6.79) becomes eectively constant, and can be
absorbed into a new constant of proportionality, K  .
Consider, for example, a semi-infinite strip in the z-plane, for which = 0,
= /2, and = /2. This is mapped onto the upper half of the -plane, with the
zero-angle vertex corresponding to a point at infinity in the -plane, by means of the
166 Theoretical Fluid Mechanics

y z-plane

z0 + i /K

x

z0

-plane
c b

Figure 6.12
Conformal transformation of a semi-infinite strip in the z-plane into the upper half of
the -plane.

transformation
d
= K  ( b)1/2 ( c)1/2 , (6.80)
dz
where K  is real and positive. We can integrate the previous expression to give
1 1
= (b + c) + (b c) cosh[K  (z z0 )]. (6.81)
2 2
Here, the points = b and = c in the -plane correspond to the vertices z = z0 and
z = z0 + i /K  , respectively, in the z-plane. (See Figure 6.12.)
Suppose that b = 0, c = w, z0 = 0, and /K  = w. In this case, the transforma-
tion (6.81) becomes
w& $ z% '
= cosh 1 , (6.82)
2 w
which implies that
w& $ x% $ y% '
= cosh cos 1 , (6.83)
2 w w
w $ x % $ y %
= sinh sin . (6.84)
2 w w
The transformation (6.82) maps the semi-infinite strip in the z-plane that is bounded
by the lines y = 0, x = 0, and y = w into the upper half of the -plane. The
transformation also maps the origin of the -plane to the origin of the z-plane, and
the -axis to the lines y = w, x = 0, and y = 0. Now, in the -plane,
Q
F= ln (6.85)

Two-Dimensional Potential Flow 167

y/w
0.5

0
0 1 2
x/w
Figure 6.13
Stream function due to a line source located in the left-hand corner of a semi-infinite
strip bounded by rigid planes at y = 0, x = 0, and y = w.

is the complex potential associated with a line source of strength 2 Q, located at the
origin. By symmetry, the -axis corresponds to a streamline, and can, therefore, be
replaced by a rigid boundary. It follows that, in the z-plane, the same potential is
that due to a line source, of strength Q, located in the lower left-hand corner of a
semi-infinite strip bounded by rigid planes at y = 0, x = 0, and y = w. (Incidentally,
the source strength in the z-plane is Q, rather than 2 Q, because, by symmetry, half
the output from the source in the -plane goes below the -axis and, therefore, does
not map to the semi-infinite strip in the z-plane.) The corresponding stream function
is

Q Q 1
= arg() = tan , (6.86)

and is illustrated in Figure 6.13.
An infinite strip in the z-plane is a polygon with two zero-angle vertices, both at
infinity. The required transformation follows from Equation (6.79) by setting =
= 0, ( a)/a 1, and K a K  . Thus, we obtain

d
= K  ( b), (6.87)
dz
which can be integrated to give

= b + eK (zz0 )
. (6.88)

This transformation maps the point = b + 1 to the point z = z0 , the -axis to the two
lines y = Im(z0 ) and y = Im(z0 )+/K  , and the upper half of the -plane to the region
between the lines y = Im(z0 ) and y = Im(z0 ) +/K  , as illustrated in Figure 6.14. It is
clear that the transformation (6.60), studied in the preceding section, is just a special
case of the transformation (6.88).
168 Theoretical Fluid Mechanics

y z-plane
A B

/K  x
z0
A B C

-plane
b b+1
 
B A A B C

Figure 6.14
Conformal transformation of an infinite strip in the z-plane into the upper half of the
-plane.

6.9 Free Streamline Theory


Consider a situation in which steady, incompressible, irrotational flow is partly bound-
ed by rigid walls, and partly by free streamlines of unknown shape on which the
pressure takes a known constant value. For instance, the free streamlines might cor-
respond to the interface of a liquid with the atmosphere, in which case the constant
pressure would correspond to that of the atmosphere. According to Bernoullis theo-
rem, (6.40), (neglecting gravity) the fluid speed is constant on a free streamline.
Let us define the new complex variable




dz 1 1
= ln = ln = ln + i , (6.89)
d F v x i vy v

where z = x + i y, F = + i , and v and are the magnitude and direction (relative to


the x-axis) of the velocity vector (v x , vy ). (See Section 6.5.) Here, F = F, = ,
and = . It follows that the real part of is constant on each free streamline,
whereas the imaginary part is constant on each straight segment of the rigid boundary.
Thus, the whole boundary of the fluid is represented in the -plane by a polygon.
We know, from the SchwarzChristoel theorem (see Section 6.8), that it is always
possible to map the interior (or exterior) of a polygon in one complex plane onto the
upper half of another complex plane. Hence, it is possible to find conformal relations
between and a new complex variable , and between F and , such that the flow
region is mapped onto the upper half of the -plane in both cases. In this way, we
can find a relation between and F, from which an expression for F in terms of z
follows via integration.
Two-Dimensional Potential Flow 169

F -plane
y A B = 1 C
A
= 0
z-plane
A B = 1 C 
B
-plane
= 1
B A
d C
/2
b C ln(v/V )
= 0 C
x A
B

C
-plane
= 1
B

= 0
A A B C C B A

Figure 6.15
Conformal transformations required for the determination of the flow from an orifice
in a plane wall in two dimensions.

As an example of the application of free streamline theory, let us calculate the


contraction ratio of a two-dimensional jet of liquid emerging from an orifice. Sup-
pose that the orifice in question is a long thin slot in a plane wall of small thickness,
and that the wall forms part of a large vessel containing liquid. The speed of the
liquid on the free streamlines that emerge from the edges of the orifice is uniform,
and equal to V, say. This is also the speed of the interior of the jet far downstream
of the orifice, where (neglecting the eects of gravity) the streamlines are straight
and parallel. (See Figure 6.16.) Let the two streamlines bounding the flow, on which
= +1 and = 1 , saybe ABC and A BC  , respectively, where A, A , C, and
C  all represent points at infinity. Figure 6.15 shows the corresponding straight-line
boundaries in the F- and -planes, where is now defined in a more convenient
manner as
$V %
dz
= ln V = ln + i . (6.90)
d F v
We, first, need to map the semi-infinite strip A BC  CBA in the -plane to the
upper half of the -plane. The conformal transformation of a semi-infinite strip onto
the upper half of another complex plane is achieved by the relation (6.81). Adapting
this relation to our current needs, we take K  = 1, z0 = i /2, and b = c = 1, so
that
= i sinh (6.91)
170 Theoretical Fluid Mechanics

In particular, this transformation maps the points B = i /2 and B = i /2 in the


-plane to the points B = 1 and B = 1, respectively, in the -plane.
Next, we need to map the infinite strip ABCA BC  in the F-plane to the upper
half of the -plane. The conformal transformation of a semi-infinite strip onto the
upper half of another complex plane is achieved by the relation (6.88). Adapting this
relation to our current needs, we take K  = /(2 1 ), b = 0, z0 = i 1 , so that


F
= i exp . (6.92)
2 1
The flow field has now been mapped onto the upper half of the -plane in two
coincident ways, giving



F i dz 1 d F
= i exp = i sinh = V . (6.93)
2 1 2 d F V dz
Hence,
dz
V = i + (1 2 )1/2 , (6.94)
d F
where, with cuts in the z-plane at AB and A B , the correct branch of (1 2)1/2 is that
which is real and positive when = 0. Integration, with the aid of Equation (6.92),
yields
V
(z z1 ) = i ( 1) (1 2 )1/2 + tanh1 [(1 2 )1/2 ], (6.95)
2 1
where z1 is a constant. However, = 1 at the point B, where z = i d (2 d being the
width of the orifice), so z1 = i d, and
V
(z i d) = i ( 1) (1 2 )1/2 + tanh1 [(1 2 )1/2 ], (6.96)
2 1
Finally, the required relationship between z and F is obtained by eliminating be-
tween Equations (6.92) and (6.96).
On the free streamline BC, we have
= 1 , = V s, = i , (6.97)
where s denotes distance measured along the streamline from B. It follows from
Equations (6.91) and (6.92) that


V
= sin = exp s . (6.98)
2 1
Thus, making use of Equation (6.96), the equation of streamline BC can be written,
in parametric form, as
2 1  
x= tanh1 (cos ) cos ) , (6.99)
V
2 1
y=d (1 + sin ) . (6.100)
V
Two-Dimensional Potential Flow 171

1
0.8
0.6
0.4
0.2
y/d

0
0.2
0.4
0.6
0.8
1
0 1 2
x/d

Figure 6.16
Free streamlines of a liquid jet emerging from an orifice in a plane wall in two di-
mensions.

Thus, the asymptotic semi-width of the jet is

2 1
b = lim y(s) = d . (6.101)
s V

Far from the orifice, which corresponds to 0, Equation (6.94) yields

d F 1 1
= . (6.102)
dz v x i vy V

In other words, as expected, the velocity profile becomes uniform across the jet far
downstream of the orifice, which implies that 1 = b V. It follows, from Equa-
tion (6.101), that the contraction ratio of a two-dimensional liquid jet emerging from
an orifice in a plane wall takes the value

b
= = 0.61. (6.103)
d +2

Somewhat surprisingly, this value is very close to the observed contraction ratio of
a three-dimensional jet emerging from a circular hole in a plane wall. (See Sec-
tion 4.6.) According to Equations (6.99), (6.100), and (6.103), the equation of the
172 Theoretical Fluid Mechanics

0.9

0.8
b/d

0.7

0.6

0.5
0 1 2
(2/)

Figure 6.17
Contraction ratio of a liquid jet emerging from a two-dimensional orifice formed by
a gap between two semi-infinite plane walls that subtend an angle 2 .

free streamline BC can be written


x 2  
= tanh1 (cos ) cos ) , (6.104)
d +2
y + 2 sin
= , (6.105)
d +2
for /2 0. By symmetry, the equation of the free streamline BC  is obtained
from the previous equation via the transformation x x, y y. The streamlines
BC and BC  are plotted in Figure 6.16.
We can easily extend the previous calculation to determine the contraction ratio
of a liquid jet emerging from a two-dimensional orifice formed by a gap between
two semi-infinite plane walls that subtend an angle 2 . In this case, the free stream-
line BC, on which = +1 , corresponds to v = V and = , whereas the free
streamline BC  , on which = 1 , corresponds to v = V and = +. Hence, the
transformation (6.93) generalizes to give

$ %
F
= i exp = i sinh , (6.106)
2 1 2

which implies that


dz   2 /
V = i + (1 2 )1/2 . (6.107)
d F
Two-Dimensional Potential Flow 173

Far from the orifice, which corresponds to 0, the previous expression yields
d F/dz 1/V. In other words, a long way downstream of the orifice, the velocity
profile is again uniform across the jet. Hence, we can write 1 = b V, where b is the
semi-width of the jet far from the orifice. Combining the previous two equations, we
obtain
dz d   2 /
= i + (1 2 )1/2 . (6.108)
2 b
Now, = 1 corresponds to the point B, at which z = i d, where d is the semi-width
of the orifice. Thus, integration of the previous expression yields

z id 2 1
dk   2 /
= i k + (1 k 2 )1/2 . (6.109)
b k

Making the transformation k = sin , we get


 /2

z id 2 2 d
= exp i . (6.110)
b sin1 () tan

As before, we have = exp[(/2) (s/b)] on the free streamline BC, where s denotes
distance measured along the streamline from B. Far downstream of the orifice, 0
and Im(z) b. Hence, we obtain the following expression for the contraction ratio
of the jet:

 1
b 2 /2 sin[(2 /) ]
= 1+ d . (6.111)
d 0 tan
This contraction ratio is plotted as a function of in Figure 6.17. It can be seen that
the ratio takes the value 1 when = 0, which corresponds to a two-dimensional jet
emerging from a gap between two parallel planes. Not surprisingly, there is no con-
traction in this case, because the streamlines of the flow are already parallel before
they emerge from the orifice. On the other hand, the ratio takes the value 1/2 when
= , which corresponds to the two-dimensional equivalent to the Borda mouth-
piece discussed in Section 4.6. In this case, the contraction ratio is exactly the same
as that of a three-dimensional Borda mouthpiece, which is not surprising, because
there is nothing in the discussion of Section 4.6 that depends crucially on the shape
of the mouthpiece cross-section.
As a second example of the use of free streamline theory, consider the flow past
a flat plate normal to a liquid stream of infinite extent. The pressure in the cavity
formed behind the plate is assumed to be equal to that of the undisturbed stream.
Thus, the speed of the liquid on the free streamlines bounding the cavity is equal to
V, which is the uniform flow speed far upstream of the plate. Let us take = 0 on
the central streamline that divides at the stagnation point O, and later becomes the
two free streamlines. (See Figure 6.18.)
The correspondence between the various points on the streamline = 0 in the
z-, F-, and -planes is specified in Figure 6.18. The flow occupies the whole of
the F-plane, apart from a thin slit running along the positive section of the real axis.
As previously, the procedure is to find transformations that map the flow regions in
174 Theoretical Fluid Mechanics

F -plane
y
C
A O B C

B C

z-plane

B F 1/2-plane
A

b C B O B C

O
A x
= 0
-plane O
B
/2
C A ln(V /v)
C O
B O
B
-plane

O B C, A, C  B O
C

Figure 6.18
Conformal transformations required for the determination of the flow past a flat plate
with a cavity at ambient pressure.

both the F- and -planes coincidentally onto the upper half of the -plane. The
semi-infinite strip in the -plane has the same width, position, and orientation as that
shown in Figure 6.15. Thus, the appropriate relationship between and is again

= i sinh , (6.112)

with the locations of the points B and B again being = 1 and = +1, respectively.
The appropriate relationship between F and can be recognized by noting, first, that
the flow occupies the upper half of the F 1/2 -plane, and, second, that an inversion and
change of sign are needed to bring corresponding points on the two real axes into
coincidence. Thus, we obtain

1/2
kV
= , (6.113)
F

where k is some positive constant.


The required relation between F and is given by

1/2

kV i dz 1 d F
= = i sinh = V . (6.114)
F 2 d F V dz
Two-Dimensional Potential Flow 175

Hence,

1/2
1/2
1 d F kV kV
= i + 1 , (6.115)
V dz F F
where the relevant branch of (1 k V/F)1/2 is that which is positive on AO. Integra-
tion, making use of the fact that F = 0 when z = 0, yields
 F/k V  F/k V  
z du
= 1/2
= i u 1/2 + (1 1/u)1/2 du, (6.116)
k 0 i u + (1 1/u)1/2 0

or

1/2
1/2

1/2
1/2
F 1/2
1 +2 i
z F F F F
= 1 ln + +i . (6.117)
k kV kV kV kV kV 2

We can now evaluate the constant k making use of the fact that at the point B, where
= 1,

F
z = i b, = 1. (6.118)
kV
Here, b is the semi-width of the plate. Thus, we obtain
2b
k= . (6.119)
+4
On the free streamline BC,

F = = V (k + s), = i , (6.120)

where s represents distance along the streamline measured from B. Thus, on this
streamline,

1/2
k
= sin = . (6.121)
k+s
Hence, according to Equation (6.117), the equation of the streamline BC can be
written, in parametric form, as
$
s %1/2 $ s %1/2
x = (k + s) s k ln 1 +
1/2 1/2
+ , (6.122)
k k

y = 2 (k + s)1/2 k1/2 + k. (6.123)
2
By symmetry, the equation of the free streamline BC  is obtained from the previ-
ous equation via the transformation x x, y y. We deduce that the cavity
downstream of the plate extends to infinity, and that its boundary asymptotes to the
parabola

8b
y2 = 4 k x = x. (6.124)
+4
176 Theoretical Fluid Mechanics

From Bernoullis theorem, the net force per unit length exerted by the fluid on
the plate, which is obviously normal to the plate, and, therefore, directed parallel to
the x-axis, is
 b  b

1 1
D= (p p0 ) x=0 dy = V 2 v 2 dy (6.125)
b b 2 2
 kV

= V2b d . (6.126)
0 y
In conventional parlance, this force constitutes a drag, because it is directed parallel
to the undisturbed flow. (See Section 5.8.) Now, F/(k V) = /(k V) < 1 on OB, so

1/2
1 d F 1
= i 1/2 + i 1 , (6.127)
V dz

where = /(k V). However, d F/dz = v x i vy = i /y. Hence,


 1
1/2

1
1 d,
1
D = V b V k
2 2
1/2 (6.128)
0

which yields

2
D= V 2 b. (6.129)
+4
Finally, if we define the drag coecient, in the conventional manner (see Section 8.8),
then we obtain
f 2
CD = = = 0.88. (6.130)
V b + 4
2

We can extend the previous calculation to determine the flow past a flat plate
inclined at an angle to a liquid stream of infinite extent. Let the plate lie in the
plane x = 0. The flow field then occupies a semi-infinite strip in the -plane that
has the same width, position, and orientation as that in the previous calculation.
Consequently, the appropriate relation between and is again


i dz 1 d F
= i sinh = V . (6.131)
2 d F V dz
It follows that
1 d F
= i + 1 2. (6.132)
V dz
As before, the two edges of the plate, B and B , correspond to = 1 and = +1,
respectively. Furthermore, the stagnation point, O, on the front surface of the plate
corresponds to || . (See Figure 6.18.) However, the latter point is no longer
necessarily located at the plates midpoint.
The transformation (6.113) generalizes to give
kV
F = . (6.133)
( cos )2
Two-Dimensional Potential Flow 177

At the points A, C, and C  , |F| and, hence, cos . (See Figure 6.18.) It
follows from Equation (6.132) that d F/dz V (sin + i cos ), which implies that
v x V sin and vy V cos . In other words, far upstream and downstream of
the plate, the fluid velocity vector subtends an angle with the plane x = 0.
On the front surface of the plate, B OB, || 1 and d F/dz = i V /y. Hence,
from Equation (6.132),

1 1
= 2 1 = , (6.134)
V y 2 1
where the upper/lower signs correspond to > 0 and < 0, respectively. (The signs
are chosen to ensure that /y 0 as || .) Moreover, F = on B OB, so
Equation (6.133) implies that

kV
= . (6.135)
( cos )2

It follows that
dy y d 2k  
= = 21 . (6.136)
d ( cos )3
Writing
1 cos cos
= , (6.137)
cos cos
the points B, O, and B correspond to = 0, = , and = , respectively.
Furthermore,


d cos cos
= sin d, (6.138)
( cos )3 sin4
sin sin
2 1 = . (6.139)
cos cos
Hence, Equations (6.136)(6.139) yield

2k
dy = (1 cos cos + sin sin ) sin d, (6.140)
sin4
which can be integrated to give

k & $ %'
y= 2 cos + cos sin2
+ sin sin cos + sin , (6.141)
sin4 2
where we have chosen the constant of integration so as to ensure that the midpoint
of the plate lies at y = 0. We expect y = b when = 0, where b is the semi-width of
the plate. We, thus, deduce, from the previous expression, that

2 b sin4
k= . (6.142)
sin + 4
178 Theoretical Fluid Mechanics

In particular, the stagnation point, O, at which = , lies a distance


2b & $ %'
y() = 2 cos (1 + sin2 ) + sin (6.143)
sin + 4 2
from the plates midpoint.
Suppose that
= cos + , (6.144)
where ||  1. It follows, from Equation (6.133), that
kV
F = , (6.145)
2
and, from Equation (6.132), that
1 d F i

1+ , (6.146)
V dz sin
where z = (sin + i cos ) z = x + i y . Here, x and y are Cartesian coordinates
oriented such that the unperturbed flow is parallel to the x -axis. The previous two
expressions can be combined to give
z 2i
2 . (6.147)
k sin
Now, is real and negative on the free streamline BC, and real and positive on the
free streamline BC  . (See Figure 6.18.) We conclude that, in the small-|| limit, the
equations of these streamlines can be written, in parametric form, as
k
x , (6.148)
2
2k
y . (6.149)
sin
In other words, far downstream of the plate, the free streamlines, which form the
boundaries of the cavity behind the plate, asymptote to the parabola



2 4k  8 b sin2 
y = x = x. (6.150)
sin2 sin + 4
From Bernoullis theorem, the net force per unit length exerted by the fluid on
the plate, which is obviously normal to the plate, and, therefore, directed parallel to
the x-axis, is
 b  b
1 vy2
f = (p p0 ) x=0 dy = V 2
1 2 dy
b 2 b V x=0
 1
vy2 y d
1 2
1
= V2 d
2 +1 V d
 1

V vy d
= k V 2 , (6.151)
+1 vy V ( cos )3
Two-Dimensional Potential Flow 179

where use has been made of Equation (6.135). However, according to Equations (6.134)
and (6.139),
V vy y 1 2 sin sin
=V = 2 2 1 = . (6.152)
vy V V y cos cos
Thus, making use of Equation (6.138), we obtain

2kV2 2 kV2
f = sin d = . (6.153)
sin3 0 sin3
Hence, it follows from Equation (6.142) that


2 sin
f = V 2 b. (6.154)
sin + 4
The drag, D, is the component of the force acting on the plate in the direction of the
unperturbed flow: that is,


2 sin2
D = f sin = V 2 b. (6.155)
sin + 4
On the other hand, the lift, L, is the component of the force acting perpendicular to
the direction of the unperturbed flow. (See Section 5.8.) Thus,


2 sin cos
L = f cos = V 2 b. (6.156)
sin + 4
Finally, it can be seen, from a comparison between Equations (6.124), (6.129),
(6.150), and (6.155), that there appears to be a connection between the drag force
exerted on the plate, and the shape of the parabola to which the cavity formed behind
the plate asymptotes. In fact,

2
D y
= lim , (6.157)
V2 x 4 x
cavity boundary

where the unperturbed flow runs parallel to the x-axis. It turns out that this relation
is a general one for two-dimensional flow past a body of arbitrary shape with an
attached cavity at ambient pressure (Batchelor 2000).

6.10 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:

J= F(z) dz. (6.158)
C
180 Theoretical Fluid Mechanics

Because 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

Jr = ( dx dy), (6.159)
C

Ji = ( dx + dy). (6.160)
C

However, we can also write the previous expressions in the two-dimensional vector
form

Jr = A dr, (6.161)
C

Ji = B dr, (6.162)
C

where dr = (dx, dy), A = (, ), and B = (, ). According to the curl theorem


(see Section A.22),
 
A dr = ( A)z dS , (6.163)
C S
 
B dr = ( B)z dS , (6.164)
C S

where S is the region of the x-y plane enclosed by C. Hence, we obtain





Jr = + dS , (6.165)
S x y



Ji = dS . (6.166)
S x y
Let

J= F(z) dz, (6.167)
C

J = F(z) dz, (6.168)
C

where C  is a closed curve in the complex plane that completely surrounds the smaller
curve C. Consider
J = J J  . (6.169)
Writing J = Jr + i Ji , a direct generalization of the previous analysis reveals that



Jr = + dS , (6.170)
S x y



Ji = dS , (6.171)
S x y
Two-Dimensional Potential Flow 181

where S is now the region of the x-y plane lying between the curves C and C  . Sup-
pose that F(z) is well-behaved (i.e., finite, single-valued, and dierentiable) through-
out S . It immediately follows that its real and imaginary components, and , re-
spectively, satisfy the Cauchy-Riemann relations, (6.17)(6.18), throughout S . How-
ever, 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
 
dF
= (v x dx + vy dy) = Re dz, (6.172)
C C dz

where F(z) is the complex velocity potential of the flow, and use has been made of
Equation (6.35). 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 while still remaining in the fluid, and not passing over a singularity of the
complex velocity, dF/dz.

6.11 Blasius Theorem


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 (6.41), which yields
 
1  dF  2
P = p0   . (6.173)
2 dz
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 suciently 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 6.19, 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, (6.174)
dY = P dx. (6.175)

The pressure force (per unit length) acting across the element also contributes to a
182 Theoretical Fluid Mechanics

P dx

dx

P dy
dy
y dl
P dl

Figure 6.19
Force acting across a short section of a curve.

moment (per unit length), M, acting about the z-axis, where

dM = x dY y dX = P (x dx + y dy). (6.176)

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


X= P dy, (6.177)
C

Y= P dx, (6.178)
C

M= P (x dx + y dy), (6.179)
C

respectively, where the integrals are taken (counter-clockwise) around the curve C.
Finally, given that the pressure distribution on the curve takes the form (6.173), and
Two-Dimensional Potential Flow 183

that a constant pressure obviously yields zero force and zero moment, we find that
   2
X=
1  dF  dy,
(6.180)
2  
C dz
   2
1
Y=  dF  dx, (6.181)
2  
C dz
   2
1
M=  dF  (x dx + y dy). (6.182)
2  dz 
C

Now, z = x + i y, and z = x i y, where indicates a complex conjugate. Hence,


d z = dx i dy, and i dz = dy + i dx. It follows that
   2
X iY =
1
i  dF  d z. (6.183)
2  
C dz

However,
  2
 dF  d z = dF d F d z = dF d F, (6.184)
 dz  dz d z dz
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
d F = dF = dz, (6.185)
dz dz dz

which implies that



2
1 dF
X iY = i dz. (6.186)
2 C dz
This result is known as the Blasius theorem, after Paul Blasius (18831970).
Now, x dx + y dy = Re(z d z). Hence,

   2
1
M = Re  dF  z d z , (6.187)
2  
C dz

or, making use of an analogous argument to that employed previously,




1  dF 2

M = Re z dz , (6.188)
2 C dz

In fact, Equations (6.186) and (6.188) 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
184 Theoretical Fluid Mechanics

C x

C
Q

Figure 6.20
Source in the presence of a rigid boundary.

there is a rigid boundary at y = 0. As we have seen, the complex velocity in the


region y > 0 takes the form
dF Q z
= . (6.189)
dz z2 + a2
Suppose that we evaluate the Blasius integral, (6.188), about the contour C shown in
Figure 6.20. 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 cor-
responds to a streamline (because 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. Observe, 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 6.20.) Hence, we can evaluate the Blasius integral around C  without
changing its value. Thus,

2 
1 dF 1 $ Q %2 z2
X iY = i dz = i dz, (6.190)
2 C dz 2 C (z + a )
2 2 2

or

1 $ Q %2 1 2 1
X iY = i + + dz. (6.191)
8 C (z i a)
2 (z + i a) (z i a) (z + i a)
Two-Dimensional Potential Flow 185

Writing z = i a +  e i , dz = i  e i d, and taking the limit  0, we find that

i Q2
X iY = . (6.192)
4 a

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 (6.47). Incidentally,
it is easily demonstrated from Equation (6.188) 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 flow-
ing fluid whose velocity is V = V0 (cos 0 , sin 0 ). From Section 6.4, the complex
velocity potential of the net flow is

Q
F(z) = ln z V0 z ei 0 . (6.193)
2

The net force (per unit length) acting on the source (which is calculated by per-
forming 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. (6.194)

This 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 previous 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 uni-
formly 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 . (6.195)
2

The net force (per unit length) acting on the filament (which is calculated by per-
forming the Blasius integral around a small circle centered on the filament) is (see
Exercise 6.2)
F = V ez . (6.196)

This 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 previous 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 filament (excluding
the velocity field of the filament itself).
186 Theoretical Fluid Mechanics

6.12 Exercises
6.1 Demonstrate that a line source of strength Q (running along the z-axis) sit-
uated 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 perpen-
dicular 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 per-
pendicular 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 ex-
periences 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 per-
pendicular 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


circular 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 b (b 2 a 2 )

directed toward the source.


Two-Dimensional Potential Flow 187

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
D ei
F(z) = .
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 inter-
sects 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 x 2
as well as a force (per unit length) whose x- and y-components are
D1 D2
X= 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.
6.11 A dipole line source of complex strength D e i runs parallel to, and is lo-
cated a perpendicular distance a from, a rigid planar boundary. Show that the
boundary experiences a force per unit length

D2
8 a 3
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),
188 Theoretical Fluid Mechanics

where z = x + i y, = + i , and c is real and positive. Show that


c sinh
x= ,
cosh cos
c sin
y= .
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
2 c i V0
F(z) = 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 . Demonstrate 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
Two-Dimensional Potential Flow 189

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
2 V0
v= [1 cos( n/2)] .
n2
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
F(z) =
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 dierence between the top and
bottom of the log is 4 V02 /32.
6.17 Show that the complex potential


c2 i
F(z) = V ei + e ,

where
l2
z=+ .

(l < c) represents uniform flow of unperturbed speed V, whose direction sub-
tends a (counter-clockwise) angle with the x-axis, around an impenetrable
elliptic cylinder of major radius a = c + l 2 /c, aligned along the x-axis, and
minor radius b = c l 2 /c, aligned along the y-axis. Demonstrate that the
moment per unit length (about the z-axis) exerted on the cylinder by the flow
is

M = V 2 (a 2 b 2 ) sin(2 ).
2
Hence, deduce that the moment acts to turn the cylinder broadside-on to the
flow (i.e., = /2 is a dynamically stable equilibrium state), and that the
equilibrium state in which the cylinder is aligned with the flow (i.e., = 0)
is dynamically unstable.
190 Theoretical Fluid Mechanics

6.18 Consider a simply-connected region of a two-dimensional flow pattern bounded


on the inside by the closed curve C1 (lying in the x-y plane), and on the out-
side by the closed curve C2 . Here, C1 and C2 do not necessarily correspond
to streamlines of the flow. Demonstrate that the kinetic energy per unit length
(in the z-direction) of the fluid lying between the two curves is
  
1
K= d d [] d ,
2 C2 C1 C3

where is the fluid mass density, the velocity potential, and the stream
function. Here, C3 is a curve that runs from C1 to C2 , and [] denotes the
amount by which the velocity potential increases as the argument of x + i y
increases by 2.
6.19 Show that the complex potential
 
F(z) = V c 2 e i l 2 ei 1

where
l2
z=+ .

(l < c) represents the flow pattern around an impenetrable elliptic cylinder
of major radius a = c + l 2 /c, aligned along the x-axis, and minor radius
b = c l 2 /c, aligned along the y-axis, moving with speed V, in a direction
that makes a counter-clockwise angle with the x-axis, through a fluid that
is at rest far from the cylinder. Demonstrate that the kinetic energy per unit
length of the flow pattern is
1  
K= V 2 a 2 sin2 + b 2 cos2 ,
2
where is the fluid mass density. Hence, show that the cylinders added mass
per unit length is
 
madded = a 2 sin2 + b 2 cos2 .

6.20 Demonstrate from Equation (6.110) that the equation of the free streamline
BC, in the case of a liquid jet emerging from a two-dimensional orifice of
semi-width d formed by a gap between two semi-infinite plane walls that
subtend an angle 2 , can be written parametrically as:
 /2

1
x 2 2 d 2 /2 2 d
= cos 1+ sin ,
d sin1 () tan 0 tan


1
y 2 /2 2 d 2 /2 2 d
=1 sin 1+ sin ,
d sin1 () tan 0 tan
where 0 1. Here, the orifice corresponds to the plane x = 0, and the
Two-Dimensional Potential Flow 191

flow a long way from the orifice is in the +y-direction. Show that for the
case of a two-dimensional Borda mouthpiece, = , the previous equations
reduce to
x 1 
= ln() 1 + 2 ,
d
y 1 1  1 
= + sin () + (1 2 )1/2 .
d 2
Finally, show that the previous equations predict that the free streamline is
re-entrant, with x/d attaining its minimum value (1 ln 2)/(2) = 0.153
when y/d = 3/4 + 1/(2) = 0.909.
192 Theoretical Fluid Mechanics
7
Axisymmetric Incompressible Inviscid Flow

7.1 Introduction
This chapter investigates axisymmetric, incompressible, inviscid flow. More infor-
mation on this subject can be found in Batchelor 2000, and Milne-Thompson 2011.

7.2 Axisymmetric Flow


A flow pattern is said to be axisymmetric when it is identical in every plane that
passes through a certain straight-line. The straight-line in question is referred to
as the symmetry axis. Let us set up a Cartesian coordinate system in which the
symmetry axis corresponds to the z-axis. The flow is most conveniently described in
terms of the cylindrical coordinates (, , z), or the spherical coordinates (r, , ).
Here,  = (x 2 + y 2 )1/2 , = tan1 (y/x), r = (x 2 + y 2 + z 2 )1/2 , and = cos1 (z/r).
(See Appendix C.) In particular,  = r sin and z = r cos .

7.3 Stokes Stream Function


Consider a fixed point A lying on the symmetry axis, and an arbitrary point P. Let
us join A to P via two dierent curves, AQ1 P and AQ2 P, that both lie in the same
plane. (See Figure 7.1.) We shall refer to this plane as the meridian plane. The
position of a given point in the meridian plane can be specified either in terms of
the cylindrical coordinates (, z), or the spherical coordinates (r, ). If the meridian
curves AQ1 P and AQ2 P rotate about the symmetry axis then closed surfaces are
formed. Assuming that the flow pattern is incompressible, the flux of fluid from right
to left (in Figure 7.1) across the surface generated by AQ2 P must match that in the
same direction across the surface generated by AQ1 P. Let us denote the flux across
either of these surfaces by 2 . Here, is known as the Stokes stream function. If
we keep AQ1 P fixed, and replace AQ2 P by any other meridian curve joining A to P,
then the value of is clearly unaltered. Thus, the stream function depends on the

193
194 Theoretical Fluid Mechanics

R v
 s
R

P
Q3
Q2
Q1

z
A B

Figure 7.1
Definition of the Stokes stream function.

position of the arbitrary point P, and, possibly, on that of the fixed point A. In fact,
if we take another fixed point B on the symmetry axis, and draw the meridian curve
BQ3 P, then the flux across the surface generated by BQ3 P will be the same as that
across the surface generated by AQ1 P, because, by symmetry, there is no flow across
AB. (See Figure 7.1.) It follows that the value of does not depend on the particular
fixed point that is used in its definition, as long as this point lies on the symmetry
axis. Hence, we conclude that the value of the stream function at P depends solely
on the position of P. Furthermore, if P lies on the axis then = 0.
Consider two neighboring points, R and R , lying in the meridian plane. (See
Figure 7.1.) The flux from right to left across the surface generated by revolving any
line joining R to R about the symmetry axis is 2 R 2 R . If the distance RR
takes the infinitesimal value s then we can write

2 (R R ) = 2  s v , (7.1)

where v is the normal flow velocity (from right to left) across the straight-line RR
in the meridian plane. (See Figure 7.1.) It follows that

1
v = . (7.2)
 s
In particular, if we suppose that ds is, in turn, equal to d, dz, dr, and r d then we
obtain the following expressions for the in-plane velocity components in cylindrical
Axisymmetric Incompressible Inviscid Flow 195

and spherical coordinates:

1 1
v = , vz = , (7.3)
 z  
1 1
vr = , v = . (7.4)
r sin r r sin r

7.4 Axisymmetric Velocity Fields


According to the analysis of Appendix C, Equations (7.3) and (7.4) imply that

v = . (7.5)

When the fluid velocity is written in this form it becomes obvious that the incom-
pressibility constraint v = 0 is satisfied [because (A B) 0see Equa-
tions (A.175) and (A.176)]. It is also clear that the Stokes stream function, , is
undefined to an arbitrary additive constant.
In fact, the most general expression for an axisymmetric incompressible flow
pattern is
v = + , (7.6)

where = (, z) is the angular velocity of flow circulating about the z-axis.
(This follows because ( ) = 0 when / = 0. See Appendix C.) The
previous expression implies that v = 0 (because = 0 when / =
0). In other words, when plotted in the meridian plane, streamlines in a general
axisymmetric flow pattern correspond to contours of .
Making use of the vector identities (A.176) and (A.178), we can also write Equa-
tion (7.6) in the form
v = ( ) + . (7.7)

It follows from the identity (A.177) that

v = [ ( )] + 2 ( ) +
= 2 ( ) + , (7.8)

because ( ) = 0, by symmetry. Hence, the vorticity of a general axisymmetric


flow pattern is written
= + e , (7.9)
196 Theoretical Fluid Mechanics

where = (, z), and [see Equations (C.52)(C.54)]


$%
= e 2 ( ) = 2 3
 


1 1 2
= +
    z 2


1 2 1 1
= + . (7.10)
r sin r 2 r 3 sin
In the following, we shall concentrate on axisymmetric flow patterns in which there
is no circulation about the z-axis (i.e., = v = 0).

7.5 Axisymmetric Irrotational Flow in Spherical Coordinates


In an irrotational flow pattern, we can automatically satisfy the constraint v = 0
by writing
v = . (7.11)
Suppose, however, that, in addition to being irrotational, the flow pattern is also
incompressible: that is, v = 0. In this case, Equation (7.11) yields

2 = 0. (7.12)

In spherical coordinates, assuming that the flow pattern is axisymmetric, so that =


(r, ), the previous equation leads to (see Section C.4)



1 2 1
r + sin = 0. (7.13)
r 2 r r r 2 sin
Let us search for a separable solution of Equation (7.13) of the form

(r, ) = R(r) (). (7.14)

It is easily seen that





1 d 2 dR 1 d d
r = sin , (7.15)
R dr dr sin d d
which can only be satisfied provided


d 2 dR
r l (l + 1) R = 0, (7.16)
dr dr

d d
(1 2 ) + l (l + 1) = 0, (7.17)
d d
Axisymmetric Incompressible Inviscid Flow 197

where = cos , and l (l + 1) is a constant. The solutions to Equation (7.17) that are
well behaved for in the range 1 to +1 are known as the Legendre polynomials,
and are denoted the Pl (), where l is a non-negative integer (Jackson 1962). (If l is
non-integer then the solutions are singular at = 1.) In fact,

(1) l d l  
2 l
Pl () = 1 . (7.18)
2 l l! d l
Hence,

P0 () = 1, (7.19)
P1 () = , (7.20)
1 2 
P2 () = 3 1 , (7.21)
2
1 3 
P3 () = 5 3 , (7.22)
2
et cetera. The general solution of Equation (7.16) is a linear combination of r l and
r(l+1) factors. Thus, the general axisymmetric solution of Equation (7.12) is written
 
(r, ) = l r l + l r(l+1) Pl (cos ), (7.23)
l=0,

where the l and l are arbitrary coecients. It follows from Equations (7.4) that the
corresponding expression for the Stokes stream function is
 $ l l %  dPl
(r, ) = 0 + r l+1 r l 1 2 , (7.24)
l=1,
l+1 l d

where = cos .

7.6 Uniform Flow


Consider a uniform steady stream of velocity v = V ez . Consider the flux (in the
minus z-direction) across a plane circle of radius  that lies in the x-y plane, and
whose center coincides with the z-axis. From the definition of the Stokes stream
function (see Section 7.3), we have 2 (, z) =  2 V, or
1
(, z) = V  2 . (7.25)
2
When expressed in terms of spherical coordinates, the previous expression yields
1
(r, ) = V r 2 sin2 . (7.26)
2
198 Theoretical Fluid Mechanics

Of course, uniform flow is irrotational [this is clear from a comparison of Equa-


tions (7.10) and (7.25)], so we can also represent the flow pattern in terms of a ve-
locity potential: that is (see Section 5.4),
(, z) = V z, (7.27)
or
(r, ) = V r cos . (7.28)
It follows, from the previous analysis, that the velocity field of a uniform stream,
running parallel to the z-axis, can either be written v = , with specified by
Equations (7.25)(7.26), or v = , with specified by Equations (7.27)(7.28).

7.7 Point Sources


Consider a point source, coincident with the origin, that emits fluid isotropically at
the steady rate of Q volumes per unit time. By symmetry, we expect the associated
steady flow pattern to be isotropic, and everywhere directed radially away from the
source. In other words,
v = v(r) er , (7.29)
where r is a spherical coordinate. Consider a spherical surface S of radius r whose
center coincides with the source. In a steady state, the rate at which fluid crosses this
surface must be equal to the rate at which the source emits fluid. Hence,

v dS = 4 r 2 vr (r) = Q, (7.30)
S

which implies that


Q
vr (r) = . (7.31)
4 r 2
Of course, v = 0.
According to Equations (7.4), the Stokes stream function associated with a point
source at the origin is such that = (), and is obtained by integrating
Q 1
vr = = 2 . (7.32)
4 r 2 r sin
It follows that
Q Q z
= cos = . (7.33)
4 4 ( 2 + z 2 )1/2
It is clear, from a comparison of Equations (7.10) and (7.33), that the previously
specified flow pattern is irrotational. Hence, this pattern can also be derived from a
velocity potential. In fact, by symmetry, we expect that = (r). The potential itself
is obtained by integrating
Q
vr = = . (7.34)
4 r 2 r
Axisymmetric Incompressible Inviscid Flow 199


P

r1
r r2

1 2
z
d/2 O d/2

Figure 7.2
A dipole source.

It follows that
Q Q
= = . (7.35)
4 r 4 ( 2 + z 2 )1/2

7.8 Dipole Point Sources


Consider the flow pattern generated by point source of strength Q located on the
symmetry axis at z = +d/2, and a point source of strength Q (i.e., a point sink)
located on the symmetry axis at z = d/2. It follows, by analogy with the analysis
of the previous section, that the stream function and velocity potential at a general
point, P, lying in the meridian plane, are

Q
= (cos 2 cos 1 ) , (7.36)
4
and

Q 1 1
= , (7.37)
4 r2 r1
respectively. Here, r1 , r2 , 1 , and 2 are defined in Figure 7.2.
In the limit that the product D = Q d remains constant, while d 0, we obtain a
so-called dipole point source. According to the sine rule of trigonometry,

r1 r2 d
= = . (7.38)
sin 2 sin 1 sin(2 1 )
200 Theoretical Fluid Mechanics

However, sin(2 1 ) = 2 sin[(2 1 )/2] cos[(2 1 )/2], so we obtain


d (sin 2 sin 1 )
r1 r2 = . (7.39)
2 sin[(2 1 )/2] cos[(2 1 )/2]
In fact, sin 2 sin 1 = 2 cos[(2 + 1 )/2] sin[(2 1 )/2], which leads to
d cos[(2 + 1 )/2]
r1 r2 = . (7.40)
cos[(2 1 )/2]
Thus, in the limit 2 1 and r2 r1 r, we get

r1 r2 d cos . (7.41)

Hence, according to Equation (7.37),




Q r1 r2 D cos
= . (7.42)
4 r1 r2 4 r 2
Equation (7.36) implies that




Q z d/2 z + d/2 Q 1 1 d 1 1
= = z + . (7.43)
4 r2 r1 4 r2 r1 2 r2 r1
Thus, in the limit 2 1 and r2 r1 r, we obtain


Q d cos2 d D sin2 D 2
= = , (7.44)
4 r r 4 r 4 ( + z 2 )3/2
2

where use has been made of Equation (7.41), as well as the fact that z = r cos .
Figure 7.3 shows the stream function of a dipole point source located at the origin.
Incidentally, Equations (7.26), (7.28), (7.33), (7.35), (7.42), and (7.44) imply that
the terms in the expansions (7.23) and (7.24) involving the constants 0 , 1 , and 1
correspond to a point source at the origin, uniform flow parallel to the z-axis, and a
dipole point source at the origin, respectively. Of course, the term involving 0 is
constant, and, therefore, gives rise to no flow.

7.9 Flow Past a Spherical Obstacle


Consider the steady flow pattern produced when an impenetrable rigid spherical ob-
stacle is placed in a uniformly flowing, incompressible, inviscid fluid. For instance,
suppose that the radius of the sphere is a, and that its center coincides with the origin.
Furthermore, let the unperturbed fluid velocity be of magnitude V, and be directed
parallel to the z-axis. We expect the flow pattern to remain unperturbed very far away
from the sphere. In other words, we expect v(r, , ) V ez as r/a . Given that
the fluid velocity field a large distance upstream of the sphere is irrotational (because
Axisymmetric Incompressible Inviscid Flow 201

1
0.8
0.6
0.4
0.2
x 0
0.2
0.4
0.6
0.8
1
1 0.5 0 0.5 1
z

Figure 7.3
Contours of the stream function of a dipole source located at the origin.

a uniform flow pattern is automatically irrotational), it follows from the Kelvin cir-
culation theorem that the velocity field remains irrotational as it is convected past the
sphere. (See Section 4.14.) Hence, we can write v = , where
2 = 0 (7.45)
(because the fluid is incompressible.) The boundary conditions are
(r, , ) = V r cos as r , (7.46)
and
(a, , )
= 0. (7.47)
r
The latter constraint arises because the surface of the sphere is impenetrable, which
implies that v er = 0 at r = a.
Let us search for an axisymmetric solution of Equation (7.45) of the form
(r, ) = V r P1 (cos ) + 1 r 2 P1 (cos ). (7.48)
It can be seen, by comparison with Equation (7.23), that the previous expression def-
initely solves Equation (7.45). Moreover, the expression also automatically satisfies
the boundary condition (7.46) [because P1 (cos ) = cos ]. The remaining boundary
condition, (7.47), yields 1 = V/(2 a 3). Hence, we obtain

r 1 $ a %2
(r, ) = V a + cos , (7.49)
a 2 r
202 Theoretical Fluid Mechanics

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

Figure 7.4
Contours of the stream function generated by a spherical obstacle of radius a placed
in the uniform flow field v = V ez .

or


a3
vr (r, ) = V 1 3 cos , (7.50)
r


1 a3
v (r, ) = V 1 + sin . (7.51)
2 r3
Because the solutions of Laplaces equation, subject to well-posed boundary condi-
tions, are unique (Riley 1974), we can be sure that the previous axisymmetric solu-
tion is the most general solution to the problem.
It is clear, by comparison with Equations (7.28) and (7.42), that the velocity po-
tential (7.49) is the superposition of that associated with uniform flow with velocity
V, parallel to the z-axis, and a dipole point source of strength D = 2 V a 3 , located
at the origin. Thus, making use of Equations (7.26) and (7.44), the associated stream
function takes the form


1 a3 1 a3
= V r 2 sin2 1 3 = V  2 1 . (7.52)
2 r 2 ( 2 + z 2 )3/2
Figure 7.4 show the contours of this stream function.
Bernoullis theorem yields (see Section 4.3)
p 1 2 p0 1
+ v = + V 2, (7.53)
2 2
Axisymmetric Incompressible Inviscid Flow 203

where is the uniform fluid mass density, and p0 the fluid pressure at infinity. Thus,
making use of Equations (7.50) and (7.51), the pressure distribution on the surface
of the sphere can be written
1 9
p(a, , ) = p0 V2 + V 2 cos(2 ). (7.54)
16 16
The net force exerted on the sphere by the fluid has the Cartesian components

F x = p(a, , ) a 2 sin cos d, (7.55)

Fy = p(a, , ) a 2 sin sin d, (7.56)

Fz = p(a, , ) a 2 cos d, (7.57)

where the integrals are over all solid angle. Thus, it follows that
F x = Fy = Fz = 0. (7.58)

In other words, the fluid exerts zero net force on the sphere, in accordance with
dAlemberts paradox. (See Section 4.5.)

7.10 Motion of a Submerged Sphere


Consider a situation in which an impenetrable rigid sphere of radius a is moving
through an incompressible, inviscid fluid at the time dependent velocity V = Vz (t) ez .
Assuming that the fluid and sphere were both initially stationary, it follows that the
fluid velocity field was initially irrotational. Thus, according to the Kelvin circulation
theorem, the fluid velocity field remains irrotational when the sphere starts to move.
Hence, we can write v = , and
2 = 0 (7.59)
(because the fluid is incompressible).
Let x, y, z be Cartesian coordinates in the initial rest frame of the fluid, and let r,
, be spherical coordinates in a frame of reference that co-moves with the sphere.
In the following, all calculations are performed in the rest frame. We expect the flow
pattern set up around the sphere to be axisymmetric (i.e., independent of .) We also
expect the fluid a long way from the sphere to remain stationary. In other words,

(r, , t) 0 as r . (7.60)
Moreover, because the sphere is impenetrable, we require that
V er = v er |r=a , (7.61)
204 Theoretical Fluid Mechanics

or 

 = Vz cos . (7.62)
r r=a
It is easily demonstrated that the solution to Equation (7.59), subject to the boundary
conditions (7.60) and (7.62), is

1 a3
(r, , t) = Vz (t) 2 cos . (7.63)
2 r
Hence,
a3
vr (r, , t) = Vz (t) cos , (7.64)
r3
1 a3
v (r, , t) = Vz (t) 3 sin . (7.65)
2 r
The general form of Bernoullis theorem, (4.96), which applies to an irrotational
flow field, yields
p 1 2 p0
+ v + g r cos = , (7.66)
2 t
where is the uniform fluid mass density, p0 the fluid pressure at infinity, and it is
assumed that fluid and cylinder are both situated in a gravitational field of uniform
acceleration g ez. Thus, the pressure distribution at the surface of the sphere can be
written
 
1 2  

p(a, , t) = p0 v  g a cos +  V |r=a . (7.67)
2 r=a t r=a
The final term on the right-hand side of the previous equation arises because
 
 
 =  V . (7.68)
t x,y,z t r,
Hence, we obtain
1 9 1 dVz
p(a, , , t) = p0 Vz2 + Vz2 cos(2 ) g a cos + a cos . (7.69)
16 16 2 dt
The net force exerted on the sphere by the fluid is specified by Equations (7.55)
(7.57). It follows that
F x = Fy = 0, (7.70)
and
1  dVz
F z = m g m , (7.71)
2 dr
where m = (4/3) a 3 is the mass of the fluid displaced by the sphere. The equation
of vertical (i.e., in the z-direction) motion of the sphere is thus,
dVz
m = Fz m g, (7.72)
dt
Axisymmetric Incompressible Inviscid Flow 205

where m is the mass of the sphere. Hence,




1  dVz
m+ m = m g + m g. (7.73)
2 dt
It follows that the sphere moves under the action of its weight, and the buoyancy
force exerted on it by the fluid (i.e., the first and second terms, respectively, on the
right-hand side of the previous equation), as if it had the virtual mass m + (1/2) m. In
other words, as it moves, the sphere eectively entrains an added mass of fluid equal
to half of the displaced fluid mass. The net vertical acceleration of the sphere can be
written

s1
az = g, (7.74)
s + 1/2
where s = m/m is the spheres specific gravity. For the case of a bubble of gas
in a liquid, we expect 0 < s  1, because the gas is inevitably much less dense
than the liquid. It follows that the bubble accelerates vertically upward at twice the
acceleration due to gravity.
The origin of the spheres added mass is easily explained. According to Equa-
tions (7.64) and (7.65), the total kinetic energy of the fluid surrounding the sphere
is

1 2
Kfluid = v dV
fluid 2
  $ %6

1 a 1
= V2 cos2 + sin2 2 r 2 sin dr d, (7.75)
2 a 0 r 4
which reduces to
1  2
Kfluid = m V . (7.76)
4
However, the kinetic energy of the sphere is
1
Ksphere = m V 2. (7.77)
2
Thus, the total kinetic energy is


1 1  2
K = Kfluid + Ksphere = m+ m V . (7.78)
2 2
In other words, the kinetic energy of the fluid surrounding the sphere can be ac-
counted for by supposing that a mass m /2 of the fluid co-moves with the sphere, and
that the remainder of the fluid remains stationary.

7.11 Conformal Maps


As we saw in Section 6.7, conformal maps are extremely useful in the theory of
two-dimensional, irrotational, incompressible flows. It turns out that such maps also
206 Theoretical Fluid Mechanics

have applications to the theory of axisymmetric, irrotational, incompressible flows.


Consider the general coordinate transformation

z + i  = f ( + i ), (7.79)

where f is an analytic function. The Cauchy-Riemann relations (see Section 6.3)


yield

z 
= , (7.80)

 z
= . (7.81)

It follows, from the previous two expressions, that = 0. In other words, and
are orthogonal coordinates in the meridian plane. [Incidentally, we are assuming
that (, , ) are a right-handed set of coordinates.] Furthermore, it can also be
shown from the Cauchy-Riemann relations that

2
2 1/2
2
2 1/2
z  z 
h = h = + = + , (7.82)

where h = || 1 and h = || 1 . Writing the flow velocity in terms of a velocity


potential, so that v = , or, alternatively, in terms of a Stokes stream function, so
that v = , we get

1 1
v = = , (7.83)
h  h
1 1
v = = . (7.84)
h  h

Of course, writing the velocity field in terms of a Stokes stream function ensures
that the field is incompressible, which also implies that 2 = 0. The additional
requirement that the field be irrotational yields = 0. Making use of the analysis
of Appendix C, this requirement reduces to

(h v ) (h v )
= 0, (7.85)

or


1 1
+ = 0. (7.86)
 
Let = 0 represent the surface of an axisymmetric solid body moving with
velocity V = V ez through an incompressible irrotational fluid that is at rest a long
way from the body. Let the fluid occupy the region 0 < < , where far
from the body. (See Figure 7.5.) Let be an angular coordinate such that = 0 on
the positive z-axis, and = on the negative z-axis. The fact that the fluid is at rest
Axisymmetric Incompressible Inviscid Flow 207


E
=

B
= 0

V
z
F = A C =0 D

Figure 7.5
An axisymmetric solid body moving through an incompressible irrotational fluid.

at infinity implies that asymptotes to a constant a long way from the body. Without
loss of generality, we can chose this constant to be zero. Thus, one constraint on the
system is that
( , ) = 0. (7.87)
The appropriate constraint at the surface of the body is that

v = V e , (7.88)

where e = /||. However, we can write V = 0 , where 0 (, z) =


(1/2) V  2 . (See Section 7.6.) Hence, from Equation (7.83), the previous con-
straint becomes
0
= (7.89)

when = 0 . Integrating, making use of the constraint (7.87) (which implies that
= 0 on the z-axis, where is constant, by symmetry), we obtain

1
(0 , ) = V  2 . (7.90)
2
We can also set the velocity potential, , to zero at infinity, and on the z-axis.
The total kinetic energy of the fluid surrounding the moving body is
  
1 1 1
K = v 2 dV = ()2 dV = ( ) dV, (7.91)
2 2 2

where we have made use of the fact that 2 = 0. Here, is the fluid mass density,
208 Theoretical Fluid Mechanics

and dV is an element of the volume obtained by rotating the area ABCDEFA, shown
in Figure 7.5, about the z-axis. Making use of the divergence theorem, we obtain
 
1 1 1
K = n 2  ds = 2  ds. (7.92)
2 2 ABC h

where ds is an element of the curve ABCDEFA, and n is an outward pointing, unit,


normal vector to the area ABCDEFA. Here, we have made use of the fact that the
velocity potential is zero at infinity (i.e., on DEF), and also on the z-axis (i.e., on CD
and FA). On the curve ABC, we can write ds = h d. Furthermore, it follows from
Equations (7.82) and (7.83) that h = h , and / =  1 /. Thus,
 

K = d , (7.93)
0 =0

or  =


K = d . (7.94)
=0 
=0

As a simple example, consider the conformal map

z + i  = c exp ( + i ), (7.95)

where c is real and positive. It follows that

z = c e cos , (7.96)
 = c e sin , (7.97)

which implies that


z 2 +  2 = r 2, (7.98)
where
r() = c e . (7.99)
Thus, the constant- surfaces are concentric spheres of radius r(). If we set

a = c e 0 (7.100)

then the problem reduces to that of a sphere, of radius a, moving through a fluid that
is at rest at infinity. This problem was solved, via dierent methods, in Section 7.10.
The constraints (7.87) and (7.90) yield

( , ) = 0, (7.101)
1
(0 , ) = V c2 e 2 0 sin2 , (7.102)
2
where use has been made of Equation (7.97). This suggests that we can write

(, ) = F() sin2 . (7.103)


Axisymmetric Incompressible Inviscid Flow 209

Substitution into the governing equation, (7.86), gives




d dF
e 2 e F = 0, (7.104)
d d
whose most general solution is

F() = A e 2 + B e . (7.105)

The constraints (7.101) and (7.102) yield

A = 0, (7.106)
1
B = V c 2 e 3 0 , (7.107)
2
respectively. Thus, we obtain

1 sin2
(, ) = V a 3 . (7.108)
2 c e
Now, from Equations (7.84) and (7.97),
1 1 sin
= = V a3 , (7.109)
 2 (c e )2
which can be integrated to give
1 cos
(, ) = V a3 . (7.110)
2 (c e )2
Note that the previous expression is formally the same as expression (7.63), as long
as we make the identifications V Vz (t), c e r, and .
On the surface of the sphere, = 0 , we obtain
1
(0 , ) = V a 2 sin2 , (7.111)
2
1
(0 , ) = V a cos . (7.112)
2
Thus,
 =
 
 1 1
3
K = d = a3 V 2 2 d = a V 2. (7.113)
=0  2 1 3
=0

As is clear from the analysis of Section 7.10, the spheres added mass can be written
K 2 3
madded = 2
= a . (7.114)
(1/2) V 3
Hence, we arrive at the standard result that the added mass is half the displaced mass
[i.e., half of (4/3) a 3 ].
210 Theoretical Fluid Mechanics

7.12 Flow Around a Submerged Oblate Spheroid


Consider the conformal map

z + i  = c sinh( + i ), (7.115)

where c is real and positive. It follows that

z = c sinh cos , (7.116)


 = c cosh sin . (7.117)

Let

a = c cosh 0 , (7.118)
b = c sinh 0 . (7.119)

Thus, in the meridian plane, the curve = 0 corresponds to the ellipse


$  %2 $ z %2
+ = 1. (7.120)
a b
We conclude that the surface = 0 is an oblate spheroid (i.e., the three-dimensional
surface obtained by rotating an ellipse about a minor axis) of major radius a and
minor radius b. The constraints (7.87) and (7.90) yield

( , ) = 0, (7.121)
1
(0 , ) = V c 2 cosh2 0 sin2 , (7.122)
2
respectively. Setting (, ) = F() sin2 , and substituting into the governing equa-
tion, (7.86), we obtain


d 1 dF 2
F = 0, (7.123)
d cosh d cosh
which can be rearranged to give


d dF
cosh 2 sinh F = 0. (7.124)
d d
On integration, we get
dF
cosh 2 sinh F = B, (7.125)
d
which can be rearranged to give


d F B
= . (7.126)
d cosh
2
cosh3
Axisymmetric Incompressible Inviscid Flow 211

It follows that

d
F() = B cosh2
cosh3 


1 sinh  1 1
= B cosh
2
tan
2 cosh2  sinh 


1 1
= B sinh cosh2 tan1 , (7.127)
2 sinh

where use has been made of the constraint (7.121). Let e = (1 b 2 /a 2 )1/2 be the
eccentricity of the spheroid. Thus, cosh 0 = 1/e, sinh 0 = (1 e 2 )1/2 /e, c = e a,
and tan1 (1/ sinh 0 ) = sin1 e. The constraint (7.122) yields

1 1 (1 e 2 )1/2 sin1 e
V a = F(0 ) = B
2
, (7.128)
2 2 e e2
or
V a2 e2
B= . (7.129)
e (1 e 2 )1/2 sin1 e
Hence,
1

2 2 sinh cosh tan (1/ sinh )
2
1
(, ) = V a e sin2 . (7.130)
2 e (1 e 2 )1/2 sin1 e
Finally, from Equation (7.84),

1
= , (7.131)
c cosh sin

which can be integrated to give



1 sinh tan1 (1/ sinh )
(, ) = V a e cos . (7.132)
e (1 e 2 )1/2 sin1 e
It is easily demonstrated that

1
(0 , ) = V a 2 sin2 , (7.133)
2
and
e (1 e 2 )1/2 sin1 e
(0 , ) = V a cos . (7.134)
sin1 e e (1 e 2 )1/2
Thus,
 =

 2 3 2 e (1 e )
2 1/2
sin1 e
K = d = a V . (7.135)
=0  3 sin1 e e (1 e 2 )1/2
=0
212 Theoretical Fluid Mechanics

1
x/a

3
3 2 1 0 1 2 3
z/a

Figure 7.6
Contours of the Stokes stream function for the case of a disk of radius a, lying in the
x-y plane, and placed in a uniform, incompressible, irrotational flow directed parallel
to the z-axis.

It follows that the added mass of the spheroid is


4
madded = a 3 G(e), (7.136)
3
where
e (1 e 2 )1/2 sin1 e
G(e) = (7.137)
sin1 e e (1 e 2 )1/2
is a monotonic function that varies between 1/2 when e = 0 and 2/ when e = 1.
In the limit e 1, our spheroid asymptotes to a thin disk of radius a moving
through the fluid in the direction perpendicular to its plane. Expressions (7.130),
(7.132), and (7.136) yield
1
(, ) = V a 2 sin2 , (7.138)
2
2
(, ) = V a cos , (7.139)

and
8
a 3,
madded = (7.140)
3
respectively. Here, z = sinh cos and  = cosh sin . It follows that, in the
Axisymmetric Incompressible Inviscid Flow 213

instantaneous rest frame of the disk,


1    1
(, z) = V a 2 + z 2 +  2 (a 2 + z 2 +  2 )2 4 a 2  2 + V  2 . (7.141)
4 2
This flow pattern, which corresponds to that of a thin disk placed in a uniform flow
perpendicular to its plane, is visualized in Figure 7.6.

7.13 Flow Around a Submerged Prolate Spheroid


Consider the conformal map

z + i  = c cosh( + i ), (7.142)

where c is real and positive. It follows that

z = c cosh cos , (7.143)


 = c sinh sin . (7.144)

Let

a = c cosh 0 , (7.145)
b = c sinh 0 . (7.146)

Thus, in the meridian plane, the curve = 0 corresponds to the ellipse


$  %2 $ z %2
+ = 1. (7.147)
b a
We conclude that the surface = 0 is an prolate spheroid (i.e., the three-dimensional
surface obtained by rotating an ellipse about a major axis) of major radius a and
minor radius b. The constraints (7.87) and (7.90) yield

( , ) = 0, (7.148)
1
(0 , ) = V c 2 sinh2 0 sin2 , (7.149)
2
respectively. Setting (, ) = F() sin2 , and substituting into the governing equa-
tion, (7.86), we obtain


d 1 dF 2
F = 0. (7.150)
d sinh d sinh
The solution that satisfied the constraint (7.148) is
1  ! "
F() = B cosh + sinh2 ln tanh(/2) . (7.151)
2
214 Theoretical Fluid Mechanics

Let e = (1 b 2 /a 2 )1/2 be the eccentricity of the spheroid. Thus, cosh 0 = 1/e,


sinh 0 = (1 e 2 )1/2 /e, c = e a, and tanh(0 /2) = [(1 e)/(1 + e)]1/2 . The constraint
(7.149) yields
V a2 e2
B= 1
. (7.152)
e (1 e ) + (1/2) ln [(1 e)/(1 + e)]
2

Hence,
 ! " 
1 cosh + sinh2 ln tanh(/2)
(, ) = V b 2 1 sin2 . (7.153)
2 e + e 2 (1 e 2 ) (1/2) ln [(1 e)/(1 + e)]

Finally, from Equation (7.84),

1
= , (7.154)
c sinh sin

which can be integrated to give


 ! " 
1 + cosh ln tanh(/2)
(, ) = V a cos . (7.155)
(1 e 2 ) 1 + (1/2 e) ln [(1 e)/(1 + e)]

It is easily demonstrated that

1
(0 , ) = V b 2 sin2 , (7.156)
2
and
 
1 + (1/2 e) ln [(1 e)/(1 + e)]
(0 , ) = V a cos . (7.157)
(1 e 2 ) 1 + (1/2 e) ln [(1 e)/(1 + e)]

Thus,
 =


K = d
=0 
=0
 
2 1 + (1/2 e) ln [(1 e)/(1 + e)]
= a b2 V 2 . (7.158)
3 (1 e 2 ) 1 + (1/2 e) ln [(1 e)/(1 + e)]

It follows that the added mass of the spheroid is

4
madded = a b 2 G(e), (7.159)
3
where  
1 + (1/2 e) ln [(1 e)/(1 + e)]
G(e) = . (7.160)
(1 e 2 ) 1 + (1/2 e) ln [(1 e)/(1 + e)]
is a monotonic function that takes the value 1/2 when e = 0, and asymptotes to
(1 e) ln[1/(1 e)] as e 1.
Axisymmetric Incompressible Inviscid Flow 215

7.14 Exercises
7.1 Demonstrate that the streamlines that pass through a stagnation point in an
incompressible, irrotational, axisymmetric flow pattern cross at right angles.
7.2 A point source of strength Q is placed at the origin in a uniform stream of
incompressible fluid of velocity v = V ez . Show that the stream function of
the resultant flow pattern is
1 Q
(r, ) = V r 2 sin2 + cos , (7.161)
2 4
where r, , are spherical coordinates. Demonstrate that the flow pattern
possesses a single stagnation point at r = a, = , where
$ Q %1/2
a= ,
4 V
and that the streamline, other than  = 0, that passes through this stagnation
point satisfies
r 1
= .
a | sin(/2)|
The flow pattern (7.161) can be reinterpreted as that which results when
a blunt obstacle lying to the right of the previously specified streamline is
placed in a uniform stream of velocity v = V ez . Show that the obstacle in
question has the asymptotic (i.e., as z ) thickness 4 a. Demonstrate that
the pressure distribution over the surface of the obstacle is

3
p = p0 + V sin (/2)
2 2
sin (/2) 1 ,
2
2
where is the fluid mass density, and p0 the pressure at infinity. Show that the
maximum pressure, p = p0 + (1/2) V 2 , on the surface occurs at = ,and
that the minimum pressure, p = p0 (1/6) V 2 , occurs at = 2 sin1 (1/ 3).
Finally, demonstrate that these are, respectively, the maximum and minimum
pressures attained in the whole flow pattern.
7.3 A and B are point sources of strengths Q and Q , respectively, in an infi-
nite incompressible fluid. Here, Q > Q > 0. Show that the equation of a
streamline is
Q cos Q cos  = constant,
where ,  are the angles that AP, BP make with AB. P being a general point.
In addition, show that the cone defined by the equation
2 Q
cos = 1
Q
divides the streamlines issuing from A into two sets, one extending to infinity,
and the other terminating at B. (Milne-Thompson 2011.)
216 Theoretical Fluid Mechanics

7.4 A and B are point sources of equal strength Q (where Q > 0) located at
(, z) = (0, a) and (0, +a), respectively in a uniform stream of incom-
pressible fluid of velocity v = V ez . Show that the stream function of the
resultant flow pattern is
1 Q
= V 2 + (cos 1 cos 2 ), (7.162)
2 4
where ,  are the angles that AP, BP make with AB. P = (, z) being a
general point. Demonstrate that the flow pattern has two stagnation points,
located at C = (0, l) and D = (0, +1), where

(l 2 a 2 )2 = 2 a b 2 l,

and b 2 = Q/(2 V). Show that the streamline (other that  = 0) that passes
through these points satisfies

 2 = b 2 (cos 1 cos 2 ),

and also passes through the points E = (h, 0) and F = (h, 0), where

h2 2a
= 2 .
b 2 (h + a 2 )1/2
The flow pattern (7.162) can be reinterpreted as that which results when an
axisymmetric solid body of oval cross-section CEDF, lying inside the previ-
ously specified streamline, is placed in a uniform stream of velocity v = V ez .
Such an obstacle is known as a Rankine solid.
7.5 Verify that $A %
= cos + B r 2 sin2
r2
is a possible stream function for an axisymmetric, incompressible, irrota-
tional flow pattern, and find the corresponding velocity potential. (Milne-
Thompson 2011.)
7.6 A sphere, a great depth below the surface of an incompressible fluid, is pro-
jected with velocity V at an inclination of 45 to the horizontal. If the density
of the sphere is twice that of the fluid, prove that the greatest height above the
point of projection attained by the sphere is 5 V 2 /(8 g). (Milne-Thompson
2011.)
7.7 A sphere of radius a is placed in an incompressible fluid flowing with the
uniform velocity V = V ez . Show that the streamlines of the resultant flow
pattern satisfy
  sin2
a3 r3 = constant,
r
where r, are spherical coordinates whose origin coincides with the center
Axisymmetric Incompressible Inviscid Flow 217

of the sphere. If the sphere is divided into two halves by a diametral plane
lying in the x-y plane, show that the resultant force between the two parts is
less than it would have been if the fluid were at rest, the pressure at infinity
remaining the same, by an amount a 2 V 2 /16, where is the fluid density.
(Milne-Thompson 2011.)
7.8 A sphere of radius a is moving along the z-axis through an incompressible
fluid with the variable speed V. Show that the pressure on the surface of the
sphere is least at the intersection of the surface with the plane

2 a 2 dV
z= ,
9 V 2 dt
the center of the sphere being instantaneously at the origin. (Milne-Thompson
2011.)
7.9 Consider the conformal map

z + i  = c sinh( + i ).

where c is real and positive. Show that the stream function


1 2
(, ) = c V cos
2
can be interpreted as that of incompressible irrotational flow, with mean speed
V, through a circular hole of radius c in an infinite plane wall, corresponding
to z = 0.
218 Theoretical Fluid Mechanics
8
Incompressible Boundary Layers

8.1 Introduction
Previously, in Sections 4.5, 5.8 and 7.9, we saw that a uniformly flowing incompress-
ible fluid that is modeled as being completely inviscid is incapable of exerting a drag
force on a rigid stationary obstacle placed in its path. This resultwhich is known
as dAlemberts paradoxis surprising because, 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 fluid. Viscosity is as-
sumed to have a significant eect on the flow inside the layer, but a negligible eect
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. More information on such boundary layers can be found in Batchelor 2000
and Schlichting 1987.

8.2 No Slip Condition


We saw previously (for instance, in Section 5.8) 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 5.2.) 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 (Riley 1974). 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.

219
220 Theoretical Fluid Mechanics

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 com-
ponents of the fluid velocity are found to be zero at the obstacles surface. The addi-
tional constraint that the tangential fluid velocity be zero at a rigid stationary bound-
ary 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.

8.3 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 8.1.) Let be the typical normal thickness
of the boundary layer. The layer thus extends over the region 0 < y < . The fluid
that occupies the region < y < , and thus lies outside the layer, is assumed to
be both irrotational and (eectively) 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.
Suppose that the equations of irrotational flow have already been solved to de-
termine 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., at 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 irrota-
tional flow and the boundary layer are eectively two-dimensional. Likewise, we are
also assuming that all flows are steady, so that any time variation can be neglected.
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
Incompressible Boundary Layers 221

y irrotational fluid

U (x)
boundary layer

solid surface
Figure 8.1
A boundary layer.

(see Section 1.14)


v x vy
+ = 0, (8.1)
x y

2
v x v x 1 p vx 2vx
vx + vy = + + , (8.2)
x y x x 2 y 2

2
vy vy 1 p vy 2 vy
vx + vy = + + , (8.3)
x y y x 2 y 2
where is the (constant) density, and the kinematic viscosity. Here, Equation (8.1)
is the equation of continuity, whereas Equations (8.2) and (8.3) 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), (8.4)
p(x, y) P(x) (8.5)
as y/ . Here, P(x) is the fluid pressure at the outer edge of the layer, and
dU 1 dP
U = (8.6)
dx dx
(because 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, (8.7)
vy (x, 0) = 0. (8.8)
222 Theoretical Fluid Mechanics

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, because the layer is
assumed to be thin. It is helpful to define the normalized variables
x
X= , (8.9)
L
y
Y= , (8.10)

vx
V x (X, Y) = , (8.11)
U0
vy
Vy (X, Y) = , (8.12)
U1
p
P(X, Y) = , (8.13)
p0
where U1 and p0 are constants. All of these variables are designed to be O(1) inside
the layer. Equation (8.1) yields
U0 V x U1 Vy
+ = 0. (8.14)
L X Y
In order for the terms in this equation to balance one another, we need

U1 = U0 . (8.15)
L
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 (8.2) gives


U02 V x V x p 0 P $ U 0 % $ %2 2 V x 2 V x
Vx + Vy = + + . (8.16)
L X Y L X 2 L X 2 Y 2
In order for the pressure term on the right-hand side of the previous equation to be of
similar magnitude to the advective terms on the left-hand side, we require that

p0 = U02 . (8.17)

Furthermore, in order for the viscous term on the right-hand side to balance the other
terms, we need
U1 1
= = , (8.18)
L U0 Re1/2
where
U0 L
Re = (8.19)

Incompressible Boundary Layers 223

is the Reynolds number of the flow external to the layer. (See Section 1.16.) 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 (8.3) yields


1 Vy Vy P 1 1 2 Vy 2 Vy
Vx + Vy = + + . (8.20)
Re X Y Y Re Re X 2 Y 2
In the limit Re
1, this reduces to

P
= 0. (8.21)
Y
Hence, P = P(X), where
dP dU
= U , (8.22)
dX dX
U(X) = U/U0 , and use has been made of Equation (8.6). 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
V x Vy
+ = 0, (8.23)
X Y
V x Vy d U 2 Vy
Vx + Vy =U + , (8.24)
X Y X Y 2
subject to the boundary conditions

V x (X, ) = U(X), (8.25)

and

V x (X, 0) = 0, (8.26)
Vy (X, 0) = 0. (8.27)

In unnormalized form, the previous set of layer equations are written


v x vy
+ = 0, (8.28)
x y
v x v x dU 2vx
vx + vy =U + , (8.29)
x y dx y 2
subject to the boundary conditions

v x (x, ) = U(x) (8.30)


224 Theoretical Fluid Mechanics

(note that y = really means y/ ), and

v x (x, 0) = 0, (8.31)
vy (x, 0) = 0. (8.32)

Equation (8.28) can be automatically satisfied by expressing the flow velocity in


terms of a stream function: that is,

vx = , (8.33)
y

vy = . (8.34)
x
In this case, Equation (8.29) reduces to

3 2 2 dU
+ =U , (8.35)
y 3 x y 2 y x y dx
subject to the boundary conditions
(x, )
= U(x), (8.36)
y
and

(x, 0) = 0, (8.37)
(x, 0)
= 0. (8.38)
y
To lowest order, the vorticity internal to the layer, = ez , is given by

2
= , (8.39)
y 2
whereas the x-component of the viscous force per unit area acting on the surface of
the obstacle is written (see Section 1.18)
 
 v x  2 
xy y=0 =  =  . (8.40)
y y=0 y 2 y=0

8.4 Self-Similar Boundary Layers


The boundary layer equation, (8.35), takes the form of a nonlinear partial dier-
ential equation that is extremely dicult to solve exactly. However, considerable
progress can be made if this equation is converted into an ordinary dierential equa-
tion by demanding that its solutions be self-similar. Self-similar solutions are such
Incompressible Boundary Layers 225

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: that is, 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 (8.35) of the general form
1/2 1/2
2 U0 x m+1 2
(x, y) = f () = U0 x m f (), (8.41)
m+1 (m + 1) U0 x m1
where 1/2
(m + 1) U0 x m1
= y. (8.42)
2
This implies that (x) = [2 /(m + 1) U0 x m1 ]1/2 , and U(x) = U0 x m . Here, U0 and
m are constants. Moreover, U0 x m has dimensions of velocity, whereas m, , and f ,
are dimensionless. Transforming variables from x, y to x, , we find that
  
  m 1 
 =  +  , (8.43)
x y x 2 x  x
 1/2 
 (m + 1) U0 x m1 
 =  . (8.44)
y x 2  x
Hence,
1/2
U0 x m1
= [(m + 1) f + (m 1) f  ], (8.45)
x 2 (m + 1)

= U0 x m f  , (8.46)
y
1/2
2 (m + 1) U03 x 3 m1
=
f  , (8.47)
y 2 2
2 U0 x m1
= [2 m f  + (m 1) f  ], (8.48)
x y 2
3 (m + 1) U02 x 2 m1 
= f , (8.49)
y 3 2
where  = d/d. Thus, Equation (8.35) becomes
1 dU 2
(m + 1) f  + (m + 1) f f  2m f  2 = . (8.50)
U02 x 2 m1 dx
Because the left-hand side of the previous equation is a (non-constant) function of ,
while the right-hand side is a function of x (and as 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 (8.51)
U02 x 2 m1 dx
226 Theoretical Fluid Mechanics

then
U(x) = U0 x m (8.52)
(which is consistent with our initial guess), and

f  + f f  + (1 f  2 ) = 0, (8.53)

where
2m
= . (8.54)
m+1
Expression (8.53) is known as the Falkner-Skan equation. The solutions to this equa-
tion that satisfy the physical boundary conditions (8.36)(8.38) are such that

f (0) = f  (0) = 0, (8.55)

and

f  () = 1, (8.56)

f () = 0. (8.57)

(The final condition corresponds to the requirement that the vorticity tend to zero at
the edge of the layer.) Note, from Equations (8.39), (8.42), (8.47), (8.52), (8.55), and
(8.56), that the normally integrated vorticity within the boundary layer is

dy = U(x). (8.58)
0

Furthermore, from Equations (8.40), (8.47), and (8.52), the x-component of the vis-
cous force per unit area acting on the surface of the obstacle is
 $ %1/2
xy y=0 = U 2
1
(m + 1)1/2 2 f  (0). (8.59)
2 Ux
It is convenient to parameterize this quantity in terms of a skin friction coecient,

xy y=0
cf = . (8.60)
(1/2) U 2
It follows that
(m + 1)1/2 2 f  (0)
c f (x) = , (8.61)
[Re(x)]1/2
where
U(x) x
Re(x) = (8.62)

is the eective 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 (8.41), the width of the
boundary layer is approximately
(x) 1
, (8.63)
x [Re(x)] 1/2
Incompressible Boundary Layers 227

1.5

f (0) 1

0.5

0 1 2

Figure 8.2
f  (0) calculated as a function of for solutions of the Falkner-Skan equation.

which implies that (x) x (m1)/2 .


If m > 0 then the external tangential velocity profile, U(x) = U0 x m , corresponds
to that of irrotational inviscid flow incident, in a symmetric fashion, on a semi-infinite
wedge whose apex subtends an angle , where = 2 m/(m + 1). (See Section 5.10,
and Figure 5.10.) 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 8.5.) 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 5.11, and Figure 5.11.) 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, (8.53), possesses no general analytic
solutions. However, this equation is relatively straightforward to solve via numeri-
cal methods. Figure 8.2 shows f  (0), calculated numerically as a function of =
2 m/(m + 1), for the solutions of Equation (8.53) that satisfy the boundary conditions
(8.55)(8.57). In addition, Figure 8.3 shows f  () versus , calculated numerically
for various dierent values of m. Because 2 as m , solutions of the Falkner-
228 Theoretical 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 1 2 3 4 5 6 7 8 9 10

Figure 8.3
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.

Skan equation with > 2 have no physical significance. For 0 < < 2, it can be
seen, from Figures 8.2 and 8.3, that there is a single solution branch characterized by
f  () > 0 and f  (0) > 0. This branch is termed the forward flow branch, because 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 for-
ward flow branch is characterized by a positive skin friction coecient, 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, because 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 coecient. The
reversed flow solutions are probably unphysical, because reversed flow close to the
wall is generally associated with a phenomenon known as boundary layer separation
(see Section 8.10) that 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 equa-
tion 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.
Incompressible Boundary Layers 229

y
boundary layer
plate


U0
x
L

wake

Figure 8.4
Flow over a flat plate.

8.5 Boundary Layer on a Flat Plate


Consider a flat plate of length L, infinite width, and negligible thickness, that 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 8.4.) 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 be-
low the plate. The fluid flow outside the boundary layers remains eectively inviscid,
whereas that inside the layers is modified by viscosity. It follows that the flow ex-
ternal to the layers is unaected 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 section. [See Equation (8.52).] (Here,
we are assuming that the flow upstream of the trailing edge of the plate, x = L, is
unaected 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 conse-
quence 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
230 Theoretical Fluid Mechanics

1
0.9
0.8
0.7
0.6
vx / U0

0.5
0.4
0.3
0.2
0.1
0
5 4 3 2 1 0 1 2 3 4 5
y/

Figure 8.5
Tangential velocity profile across the boundary layers located above and below a flat
plate of negligible thickness located at y = 0.

the upper one, the tangential velocity profile across the both layers is written

v x (x, y) = U0 f  (), (8.64)

where $ U %1/2
0
= |y|. (8.65)
2x
Here, f () is the solution of
f  + f f  = 0 (8.66)
that satisfies the boundary conditions

f (0) = f  (0) = 0, (8.67)

and

f  () = 1, (8.68)

f () = 0. (8.69)

Equation (8.66) is known as the Blasius equation.


It is convenient to define the so-called displacement thickness of the upper bound-
ary layer, 
v x (x, y)
(x) = 1 dy, (8.70)
0 U0
Incompressible Boundary Layers 231

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
x 
(x) = 2 [1 f  ()] d. (8.71)
U0 0

In fact, the numerical solution of Equation (8.66), subject to the boundary condi-
tions (8.67)(8.69), yields

1/2
x
(x) = 1.72 . (8.72)
U0
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
= , (8.73)
L Re1/2
where
U0 L
Re = (8.74)

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.
The tangential velocity profile across the both boundary layers, which takes the
form
 |y|
v x (x, y) = U0 f 1.22 , (8.75)
(x)
is plotted in Figure 8.5. In addition, the vorticity profile across the layers, which is
written
U0  |y|
(x, y) = sgn(y) 1.22 f 1.22 , (8.76)
(x) (x)
is shown in Figure 8.6. 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
 L 
D=2 xy y=0 dx, (8.77)
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 (8.59) (with m = 0) that

1/2
1/2
  L
L
D = U02 2 f (0) x1/2 dx = U02 2 2 f  (0). (8.78)
U0 0 U0
232 Theoretical Fluid Mechanics

0.6
0.5
0.4
0.3
0.2
/ (U0/)
0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
5 4 3 2 1 0 1 2 3 4 5
y/

Figure 8.6
Vorticity profile across the boundary layers located above and below a flat plate of
negligible thickness located at y = 0.

In fact, the numerical solution of (8.66) yields


U02 L
D = 1.33 = 1.33 U0 ( U0 L)1/2 . (8.79)
Re1/2
The previous discussion is premised on the assumption that the flow in the up-
per (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 /, ex-
ceeds a critical value which is about 600 (Batchelor 2000). In this case, small-scale
z-dependent disturbances spontaneously grow to large amplitude, and the layer be-
comes turbulent. Because 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 Equation (8.73), 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 8.7 in which an initially ir-
rotational 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 aected by
Incompressible Boundary Layers 233

boundary layer plate

U0 d

potential flow l

Figure 8.7
Flow between two flat parallel plates.

viscosity. On the other hand, the flow outside the layers is irrotational and essen-
tially inviscidthis type of flow is usually termed potential flow (because it can be
derived from a velocity potential satisfying Laplaces equation). 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 8.7.)
Assuming that, prior to merging, the two boundary layers do not significantly aect
one another, their thickness, (x), is given by formula (8.72), 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

d
(l) = . (8.80)
2

It follows that
l
= 11.8 Re, (8.81)
d
where
U0 d
Re = . (8.82)

Thus, when an irrotational high Reynolds number fluid passes between two parallel
plates then the region of potential flow extends a comparatively long distance be-
tween the plates, relative to their spacing (i.e., l/d
1). By analogy, if an irrotational
high Reynolds number fluid passes into a pipe then the fluid remains essentially irro-
tational until it has travelled a considerable distance along the pipe, compared to its
diameter. Obviously, these conclusions are modified if the flow becomes turbulent.
234 Theoretical Fluid Mechanics

8.6 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 par-
allel 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 8.4.) Within the wake, the flow is modified by viscosity, and possesses finite
vorticity. Outside the wake, the downstream flow remains irrotational, and eectively
inviscid.
Because 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), (8.83)


vy (x, y) = v(x, y), (8.84)

where
|u|  U0 . (8.85)
Assuming that, within the wake,

1
, (8.86)
x x
1
, (8.87)
y

where  x is the wake thickness, fluid continuity requires that


v u. (8.88)
x
The flow external to the boundary layers, and the wake, is both uniform and essen-
tially inviscid. Hence, according to Bernoullis theorem, the pressure in this region
is also uniform. [See Equation (8.22).] However, as we saw in Section 8.3, 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, because the wake is formed via the convection of the
boundary layers downstream of the plate. We conclude that

p(x, y) p0 (8.89)

everywhere in the fluid, where p0 is a constant.


Incompressible Boundary Layers 235

The x-component of the fluid equation of motion is written



2
v x v x 1 p vx 2vx
vx + vy = + + . (8.90)
x y x x 2 y 2

Making use of Equations (8.83)(8.89), the previous expression reduces to

u 2u
U0 2. (8.91)
x y
The boundary condition
u(x, ) = 0 (8.92)
ensures that the flow outside the wake remains unperturbed. Note that Equation (8.91)
has the same mathematical form as a conventional diusion equation, with x play-
ing the role of time, and /U0 playing the role of the diusion coecient. Hence,
by analogy with the standard solution of the diusion equation, we would expect
( x/U0 )1/2 (Riley 1974).
As can easily be demonstrated, the self-similar solution to Equation (8.91), sub-
ject to the boundary condition (8.92), is

2
Q y
u(x, y) = exp 2 , (8.93)

where
1/2
x
(x) = 2 , (8.94)
U0
and Q is a constant. It follows that

u dy = Q, (8.95)


because, as is well-known, exp(t 2 ) dt = (Riley 1974). 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(y 2 / 2 ), (8.96)
U0 U0

is plotted in Figure 8.8. In addition, the vorticity profile across the wake, which is
written
(x, y) Q 2 y
= exp(y 2 / 2 ) (8.97)
U0 / U0
is shown in Figure 8.9. It can be seen that the profiles pictured in Figures 8.8 and
8.9 are essentially smoothed out versions of the boundary layer profiles shown in
Figures 8.5 and 8.6, respectively.
Suppose that the plate and a portion of its trailing wake are enclosed by a cuboid
236 Theoretical Fluid Mechanics

1
0.9
0.8
0.7
0.6
vx / U0

0.5
0.4
0.3
0.2
0.1
0
3 2 1 0 1 2 3
y/

Figure 8.8
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
0.2
/ (U0/)

0.1
0
0.1
0.2
0.3
0.4
0.5
3 2 1 0 1 2 3
y/

Figure 8.9
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 237

v(x) y

y=h
plate wake
U0 U0 u(y)

y = h
x = l x=l
v(x)

Figure 8.10
Control volume surrounding a flat plate and its trailing wake.

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 8.10.) 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 1.9)

v dS = 0, (8.98)
S

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, Equation (8.98) yields
 h  h  l
U0 dy + [U0 u(y)] dy + 2 v(x) dx = 0, (8.99)
h h l

or  
h l
u(y) dy = 2 v(x) dx. (8.100)
h l

However, given that u 0 for |y|


, and because h
, it is a good approximation
to replace the limits of integration on the left-hand side of the previous expression by
238 Theoretical Fluid Mechanics

. Thus, from Equation (8.95),


 h  l
u(y) dy = 2 v(x) dx Q, (8.101)
h l

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 1.11)
 
v x v dS = F x + x j dS j , (8.102)
S 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. In an incompressible fluid (see Section 1.6),


vi v j
i j = p i j + + . (8.103)
x j xi

Hence, we obtain
 h  h
! "
U0 dy +
2
U0 u(y) 2 dy
h h
 l  h
d
+2 U0 v(x) dx = D 2 u(y) dy, (8.104)
l dl h

because the pressure within the fluid is essentially uniform, the tangential fluid ve-
locity at y = h is U0 , and v is assumed to be negligible at x = l. Making use of
Equation (8.101), as well as the fact that Q is independent of l, we get

D = U0 Q. (8.105)

Here, we have neglected any terms that are second order in the small quantity u. A
comparison with Equation (8.79) reveals that

Q = 1.33 ( U0 L)1/2 , (8.106)

or $ L %1/2
Q
= 0.664 . (8.107)
U0 x
Hence, from Equations (8.96) and (8.97), the velocity and vorticity profiles across
the layer are
v x (x, y) $ L %1/2
= 1 0.375 exp(y 2 / 2 ), (8.108)
U0 x
Incompressible Boundary Layers 239

and
(x, y) $ L %1/2 y
= 0.749 exp(y 2 / 2 ), (8.109)
U0 / x

respectively, where (x) = 2 ( x/U0 )1/2 . Finally, because the previous 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 previous 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. Because 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 suciently far downstream of the plate. Our
analysis, which eectively 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 .

8.7 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 8.11.) Moreover, let y represent
distance across the boundary layer, measured normal to the surface. Suppose that the
boundary layer is suciently 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 (8.28) and (8.29)]

u v
+ = 0, (8.110)
x y
u u dU 2u
u +v U = 2, (8.111)
x y dx y

subject to the standard boundary conditions

u(x, ) = U(x), (8.112)


u(x, 0) = v(x, 0) = 0. (8.113)
240 Theoretical Fluid Mechanics

Here, U(x) is the external tangential fluid velocity at the edge of the layer. Integrating
(8.111) across the layer, making use of the boundary conditions (8.113), leads to
 

u  dU u u
 = U u v dy
y y=0 0 dx x y

dU (U u) (U u)
= (U u) +u +v dy
0 dx x y

dU (U u) v
= (U u) +u (U u) dy
0 dx x y

dU (U u) u
= (U u) +u + (U u) dy
0 dx x x
 
dU d
= (U u) dy + u (U u) dy. (8.114)
dx 0 dx 0
Here, we have integrated the final term on the right-hand side by parts, making use
of Equations (8.110), (8.112), and (8.113). Let us define the displacement thickness
of the layer [see Equation (8.70)]
 $
u%
1 (x) = 1 dy, (8.115)
0 U
as well as the so-called momentum thickness
 $
u u%
2 (x) = 1 dy. (8.116)
0 U U
It follows from Equation (8.114) that

u  d dU
 = U 2 2 + U (1 + 2 2 ). (8.117)
y y=0 dx dx

This important result is known as the von Karman momentum integral, and is fun-
damental to many of the approximation methods commonly employed to calcu-
late boundary layer thicknesses on the surfaces of general obstacles placed in high
Reynolds number flows. (See Section 8.10.)

8.8 Boundary Layer Separation


As we saw in Section 8.5, 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
Incompressible Boundary Layers 241

separation point
potential flow streamlines

boundary layer
obstacle wake
stagnation point

Figure 8.11
Boundary layer separation.

number of the external flow. Suppose, however, that the obstacle is not streamlined:
that is, 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 8.11, 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 un-
perturbed 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.
Outside the boundary layer, and the wake, the flow pattern is irrotational and
essentially inviscid. So, from Section 5.8, the tangential flow speed just outside the
boundary layer (neglecting any circulation of the external flow around the cylinder)
is
U() = 2 U0 sin , (8.118)
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 obsta-
cle, 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
242 Theoretical Fluid Mechanics

take place at a point on the surface of an obstacle where there is deceleration of the
external tangential flow. In addition, from Section 5.8, the pressure just outside the
boundary layer (and, hence, on the surface of the obstacle, because the pressure is
uniform across the layer) is

P() = p1 + U02 cos(2 ), (8.119)

where p1 is a constant. 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 compara-
tively 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 (8.119) 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: that is, P( ) = P(). (See Figure 8.12.) 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 8.12 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 signif-
icant 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 parameterize this
force in terms of a dimensionless drag coecient,
D
CD = , (8.120)
U02 a

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 coecient of order
unity. The exact value of the coecient 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, because 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
Incompressible Boundary Layers 243

0 /2 3/2
P () p1

stagnation point

separation points

Figure 8.12
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.

high Reynolds number flow. Such gradients can be significantly reduced by stream-
lining the obstacle: that is, 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 suciently streamlined obstacle,
thereby significantly decreasing, or even eliminating, the associated form drag (es-
sentially, 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 (8.79) and (8.120), the
drag coecient 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 5.8, 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 separation.
In this case, the region of the fluid in which viscosity plays a significant role is local-
ized to a thin boundary layer on the surface of the obstacle, and the resultant friction
244 Theoretical Fluid Mechanics

drag scales as Re1/2 , and, therefore, disappears in the inviscid limit (essentially, be-
cause 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 in-
viscid limit. The presence of significant fluid vorticity within the wake invalidates
irrotational fluid dynamics. Consequently, the pressure on the back side of the obsta-
cle 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 coecient associated with this
type of drag is generally of order unity, and does not tend to zero as the Reynolds
number tends to infinity.

8.9 Criterion for Boundary Layer Separation


As we have seen, the boundary layer equations (8.110)(8.113) 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 eectively
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 bound-
ary layer towards the interior of the fluid. In other words, we expect the normal ve-
locity to become comparable with the tangential velocity. However, we have seen
that the ratio v/u is of order Re1/2 . [See Equation (8.18).] Hence, an increase of
v to such a degree that v u implies an increase by a factor Re1/2 . For suciently
large Reynolds numbers, we may suppose that v eectively increases by an infinite
factor. Indeed, if we employ the dimensionless form of the boundary layer equations,
(8.23)(8.27), 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 previous discussion,

v(x0 , y) = (8.121)

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,
Incompressible Boundary Layers 245

we can expand x0 x in powers of u u0 (at fixed y). Because (x/u)u=u0 = 0, the


first term in this expansion vanishes identically, and we are left with
 
x0 x = f (y) (u u0 )2 + O (u u0 )3 , (8.122)

or

u(x, y) u0 (y) + (y) x0 x, (8.123)

where = 1/ f is some function of y. From the equation of continuity,

v u (y)
= . (8.124)
y x 2 x0 x

Upon integration, the previous expression yields

(y)
v(x, y) , (8.125)
x0 x

where  y
1
(y) = (y ) dy . (8.126)
2
The equation of tangential motion in the boundary layer, (8.111), is written

u u dU 2u
u +v =U + 2. (8.127)
x y dx y

As is clear from Equation (8.123), the derivative 2 u/y 2 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
previous expression become infinite as x x0 . Hence, in the immediate vicinity of
the separation point,
u u
u +v 0. (8.128)
x y
Because u/x = v/y, we can rewrite this equation in the form

v u $v%
u +v = u 2 0. (8.129)
y y y u

Because u does not, in general, vanish at x = x0 , we conclude that

$v%
0. (8.130)
y u

In other words, v/u is a function of x only. From Equations (8.123) and (8.125),

v (y)
= + O(1). (8.131)
u u0 (y) x0 x
246 Theoretical Fluid Mechanics

Hence, if this ratio is a function of x alone then (y) = (1/2) A u0(y), where A is a
constant: that is,
A u0 (y)
v(x, y) . (8.132)
2 x0 x
Finally, because Equation (8.126) yields = 2 d/dy = A du0 /dy, we obtain
du0
u(x, y) u0 (y) + A x0 x. (8.133)
dy
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, because the square roots become imaginary. This im-
plies that the solutions of the boundary layer equations cannot sensibly be continued
beyond the separation point.
The standard boundary conditions at the surface of the obstacle require that u =
v = 0 at y = 0. It, therefore, follows from Equations (8.132) and (8.133) that
u0 (0) = 0, (8.134)

du0 
 = 0. (8.135)
dy  y=0

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 (19081968) (Landau and Lifshitz 1987).
If the constant A in expressions (8.132) and (8.133) 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 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: that is,
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.
The dimensionless boundary layer equations, (8.23)(8.27), 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, (8.111), is
written 
2 u  1 dU 1 dP
 = = , (8.136)
y y=0
2 U dx dx
Incompressible Boundary Layers 247

where P(x) is the pressure just outside the layer, and use has been made of Equa-
tion (8.6). Because u is positive, and increases away from the surface (upstream of
the separation point), it follows that ( 2 u/y 2)y=0 > 0 at the separation point itself,
where (u/y)y=0 = 0. Hence, according to the previous equation,


dU
< 0, (8.137)
dx x=x0


dP
> 0. (8.138)
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.

8.10 Approximate Solutions of Boundary Layer Equations


The boundary layer equations, (8.110)(8.113), take the form

u v
+ = 0, (8.139)
x y
u u dU 2u
u +v U = 2, (8.140)
x y dx y
subject to the boundary conditions

u(x, ) = U(x), (8.141)


u(x, 0) = 0, (8.142)
v(x, 0) = 0. (8.143)

Furthermore, it follows from Equations (8.140), (8.142), and (8.143) that



2 u  dU
 = U . (8.144)
y 2 y=0 dx

The previous expression can be thought of as an alternative form of Equation (8.143).


As we saw in Section 8.4, the boundary layer equations can be solved exactly when
U(x) takes the special form U0 x m . However, in the general case, we must resort to
approximation methods.
Following Pohlhausen (Schlichting 1987), let us assume that
u(x, y)
= f (), (8.145)
U(x)
248 Theoretical Fluid Mechanics

where = y/(x), and /x  1/. In particular, suppose that



a + b + c 2 + d 3 + e 4 01
f () = , (8.146)
1 >1
where a, b, c, d, and e are constants. This expression automatically satisfies the
boundary condition (8.141). Moreover, the boundary conditions (8.142) and (8.144)
imply that a = 0, and
f  (0) = (x), (8.147)
where  = d/d, and
2 dU
= . (8.148)
dx
Finally, let us assume that f , f  , and f  are continuous at = 1: that is,
f (1) = 1, (8.149)

f (1) = 0, (8.150)

f (1) = 0. (8.151)
These constraints corresponds to the reasonable requirements that the velocity, vor-
ticity, and viscous stress tensor, respectively, be continuous across the layer. Given
that a = 0, Equations (8.146), (8.147), and (8.149)(8.151) yield
f () = F() + G() (8.152)
for 0 1, where
F() = 1 (1 ) 3 (1 + ), (8.153)
1
G() = (1 ) 3 . (8.154)
6
Thus, the tangential velocity profile across the layer is a function of a single param-
eter, , which is termed the Pohlhausen parameter. The behavior of this profile is
illustrated in Figure 8.13. 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. Expression (8.152) is only
an approximation, because it satisfies some, but not all, of the boundary conditions
satisfied by the true velocity profile. For instance, dierentiation of Equation (8.140)
with respect to y reveals that ( 3 u/y 3)y=0 f  (0) = 0, which is not the case for
expression (8.152).
It follows from Equations (8.115), (8.116), and (8.152)(8.154) that
 1

3
1 (x) = (1 f ) d = , (8.155)
0 10 120
 1

37 2
2 (x) = f (1 f ) d = . (8.156)
0 315 945 9072
Incompressible Boundary Layers 249

1
0.9
0.8
0.7
0.6
f 0.5

0.4
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8 1

Figure 8.13
Pohlhausen velocity profiles for = 12 (solid curve) and = 12 (dashed curve).

Furthermore,

u  U  U$ %
 = f (0) = 2+ . (8.157)
y y=0 6

The von Karman momentum integral, (8.117), can be rearranged to give



U d2 22 dU 1 2 u 
2 + 2+ =  . (8.158)
dx dx 2 U y y=0

Defining

22 dU
(x) = , (8.159)
dx

we obtain


d
U = 2 [F2 () {2 + F1 ()}] = F(), (8.160)
dx dU/dx
250 Theoretical Fluid Mechanics

1
F

0.15 0.1 0.05 0 0.05 0.1


Figure 8.14
The function F() (solid curve) and the linear function 0.47 6 (dashed line).

where

2
37 2
= , (8.161)
315 945 9072

)

1 3 37 2
F1 () = = , (8.162)
2 10 120 315 945 9072
 $

2 u  % 37 2
F2 () =  = 2 + , (8.163)
U y y=0 6 315 945 9072



37 2 116 2 1 2
F() = 2 2 + + 2 + 3 .
315 945 9072 315 945 120 9072
(8.164)

It is generally necessary to integrate Equation (8.158) 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. 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 Equation (8.161), that = 0.1567.
As was first pointed out by Walz (Schlichting 1987), and is illustrated in Fig-
ure 8.14, it is a fairly good approximation to replace F() by the linear function
Incompressible Boundary Layers 251

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 approxi-
mation, Equations (8.159) and (8.160) reduce to the linear dierential equation

d U 22 dU 22
= 0.47 5 , (8.165)
dx dx
which can be integrated to give

22 0.47 x
= U 5 (x ) dx , (8.166)
U6 0

assuming that the stagnation point corresponds to x = 0. It follows that



0.47 dU x 5 
= U (x ) dx . (8.167)
U 6 dx 0
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 8.5.) It follows from Equations (8.166) and (8.167) that
2 (x) 0.69
= , (8.168)
x Re1/2
where Re = U0 x/, and = 0. Moreover, according to Equations (8.148) and
(8.162), = 0 and 1 /2 = 2.55. Hence, the displacement width of the boundary
layer becomes
1 (x) 1.75
= . (8.169)
x Re1/2
This approximate result compares very favorably with the exact result, (8.73).
Suppose that x = a and U() = 2 U0 sin , which corresponds to uniform trans-
verse flow around a circular cylinder of radius a. (See Section 8.8.) Equation (8.167)
yields

cos
() = 0.47 6 sin5  d . (8.170)
sin 0
Figure 8.15 shows () determined from the previous 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 x m . If m is negative then, as illustrated in Fig-
ure 8.16, this corresponds to uniform flow over the back surface of a semi-infinite
wedge whose angle of dip is
m
= . (8.171)
1+m 2
(See Section 8.4.) It follows from Equation (8.167) that
0.47 m 0.47
= = . (8.172)
1+5m /2 4
252 Theoretical Fluid Mechanics

0.2
s

0.1

0 20 40 60 80 100 120
()

Figure 8.15
The function () for flow around a circular cylinder.

Figure 8.16
Flow over the back surface of a semi-infinite wedge.
Incompressible Boundary Layers 253

We expect boundary layer separation on the back surface of the wedge when <
s = 0.1567. This corresponds to > s , where
( s )
s = 13 . (8.173)
2 0.47 + 4 ( s)
Hence, boundary layer separation can be prevented by making the wedges angle of
dip suciently shallow: that is, by streamlining the wedge, which has the eect of
reducing the deceleration of the flow on the wedges back surface. 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 8.4 can no longer be found. This suggests
that the absence of self-similar solutions for m < m is related to boundary layer
separation.

8.11 Exercises
8.1 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 (8.35) can be found such
that is a function of y/x only. Demonstrate that this solution yields
$ %1/2
u(x, y) U0 y
=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 tanh + 2
2
2
is a suitable solution of the previous dierential equation, where tanh2 =
2/3.
8.2 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.
y 3 x y 2 y x y
Here, the symmetry axis of the jet is assumed to run along the x-direction,
254 Theoretical Fluid Mechanics

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 momen-


tum flux of the jet parallel to its symmetry axis,

M= u 2 dy,

to be independent of x.
Consider a self-similar stream function of the form

(x, y) = 0 x p F(y/x q ).

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 x 1/3 F(y/x 2/3).
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 .

8.3 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) = v s , where v s is the (constant) suction velocity.
Demonstrate that, in the presence of suction, the von Karman velocity integral
becomes 
u  d dU
 = U2 2 +U (1 + 2 2 ) + U v s .
y y=0 dx dx
Suppose that

sin( y) 0 y /(2 )
u(x, y) = U(x) ,
1 y > /(2 )
Incompressible Boundary Layers 255

where = (x). Demonstrate that the displacement and momentum widths


of the boundary layer are

1 = (/2 1) 1 ,
2 = (1 /4) 1,

respectively. Hence, deduce that

(/2 1)2 d1 dU
= U (1 /4) + 1 + (/2 1) v s.
1 dx dx

Consider a boundary layer on a flat plate, for which U(x) = U0 . Show that,
in the absence of suction,

1/2
1/2
8 x
1 = (/2 1) ,
4 U0

but that in the presence of suction


(/2 1)
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/.
256 Theoretical Fluid Mechanics
9
Incompressible Aerodynamics

9.1 Introduction
This chapter investigates the forces exerted on a stationary obstacle situated in a uni-
form, high Reynolds number, subsonic (and, therefore, eectively incompressible
see Section 1.17) wind, on the assumption that the obstacle is suciently 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 fun-
damental to the theory of flight. Further information on this subject can be found in
Milne-Thomson 1958.
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 8.5 and 8.6.) It follows that, for the flow external to the
boundary layer and wake, we can write

v = , (9.1)

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: that is,
2 = 0. (9.2)
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 8.2.)
However, as described in the previous chapter, this inconsistency is resolved by the
boundary layer, across which the tangential velocity is eectively discontinuous, be-
ing 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 disconti-
nuity in the tangential velocity across the layer implies the presence of bound vortices
covering the surface of the airfoil (see Section 9.7), and also gives rise to a friction
drag acting on the airfoil in the direction of the external flow. However, the mag-
nitude of this drag scales as Re 1/2 , where Re is the Reynolds number of the wind.
(See Section 8.5.) Hence, such drag becomes negligibly small in the high Reynolds

257
258 Theoretical Fluid Mechanics

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 trans-
versely around a stationary airfoil of infinite length (in the z-direction) and uniform
cross-section (parallel to the x-y plane).

9.2 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.) The air velocity a great distance from
the airfoil must tend toward this uniform velocity. Thus, for suciently large |z|, we
can write (see Section 6.4)

dF A B
= V ei + + 2 + . (9.3)
dz z z

According to Equation (6.172), the circulation, , of air about the airfoil is deter-
mined by performing the integral

dF
Re dz = (9.4)
C dz

around a loop C that lies just above the airfoil surface. However, as discussed in
Section 6.10, the value of this integral is unchanged if it is performed around any
loop that can be continuously deformed onto C, while 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 filament). Because (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,

  
1
= Re i V Re i (+)
+ A + O(R ) d = 2 Im(A), (9.5)

which implies that


dF B
= V ei + i + + (9.6)
dz 2 z z 2
at large |z|.
As discussed in Section 6.11, the net force (per unit length) acting on the airfoil,
Incompressible Aerodynamics 259

L = X e x + Y ey , is determined by performing the Blasius integral,



2
1 dF
i dz = X i Y, (9.7)
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 instead we perform it around a large circle of radius
R, centered on the origin. Far from the airfoil,

2
dF V e i 82 V B e i 2
= V 2 e2i + i + + O(z 3 ). (9.8)
dz z 42 z 2
So, we obtain

1 2 V ei
X iY = i V 2 R e i (+2 ) + i + O(R 1 ) d = i e i V , (9.9)
2
or
X + i Y = i ei V = e i (/2) V . (9.10)
In other words, the resultant force (per unit length) acting on the airfoil is of magni-
tude 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 (9.10) is known as the theorem of Kutta
and Zhukovskii, after the German scientist M. W. Kutta (18671944), and the Rus-
sian scientist N. E. Zhukovskii (18471921), who discovered it independently. Note
that (at fixed circulation) the lift is independent of the shape of the airfoil. Further-
more, according to the KuttaZhukovskii 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, that is discussed in Section 9.8.
As discussed in Section 6.11, the net moment per unit length (about the origin),
M, acting on the airfoil is determined by performing the integral

2
1
z dz = M
dF
Re (9.11)
2 C dz

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 V R e i (+)
M = Re i V 2 R 2 e 2 i (+) + i
2
i

8 V B e
2 2
+ + O(R 1 ) d , (9.12)
42
or  
M = Re 2 V B e i (/2) . (9.13)
260 Theoretical Fluid Mechanics

9.3 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

$ %
a 2 i
F() = V e i + e +i ln , (9.14)
2 a

whereas the associated stream function takes the form



$r%
a2
(r, ) = V r sin( + ) + ln , (9.15)
r 2 a

where = r e i . It follows that

dF V a 2 e i
= V ei + i . (9.16)
d 2 2

Comparison with Equation (9.6) (with z = ) reveals that

B = V a 2 ei . (9.17)

Hence, Equations (9.10) and (9.13) yield


 
V = V cos e + sin e , (9.18)
 
L = V sin e + cos e , (9.19)
M = 0, (9.20)

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 (because the lift generates zero moment
about the origin). (See Figure 9.1.)
Of course, a cylindrical airfoil of circular cross-section is completely unrealistic,
because its back side (i.e., the side opposite to that from which the wind is incident)
is not suciently streamlined to prevent boundary layer separation. (See Chapter 8.)
However, as described in Section 6.7, we can use the conformal transformation

l2
z=+ (9.21)

Incompressible Aerodynamics 261

Figure 9.1
A cylindrical airfoil of circular cross-section.

to transform a cylinder of circular cross-section in the -plane to a cylinder of ellipti-


cal cross-section in the z-plane. (Note that both cross-sections have centroids located
at the origin.) Moreover, a cylindrical airfoil of elliptical cross-section that is su-
ciently elongated, and whose major axis subtends a suciently small angle with the
incident wind direction, constitutes a realistic airfoil, because 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 9.2, is
obtained when the parameters a and l are given the following values:
1
a= (c + ), (9.22)
4
1
l = (c 2 2 )1/2 . (9.23)
4
In this case, the surface of the airfoil satisfies the parametric equations
c
x= cos , (9.24)
2

y = sin . (9.25)
2
In particular, the airfoils leading and trailing edges correspond to = 0 and = ,
respectively.
According to Equations (9.16) and (9.21), the complex velocity in the z-plane is
262 Theoretical Fluid Mechanics

d
x
C F
c

Figure 9.2
A cylindrical airfoil of elliptical cross-section.

given by


dF dF d i V a 2 e i 2
= = Ve +i . (9.26)
dz d dz 2 2 2 l2

Thus, on the airfoil surface, where = a e i , we obtain



dF (c + )
= i V sin( + ) + . (9.27)
dz (c + ) ( cos + i c sin )

A long way from the airfoil, z l 2 /z, so that Equation (9.26) reduces to

dF V (l 2 e i a 2 ei )
V ei + i + . (9.28)
dz 2 z z2

A comparison with Equation (9.6) 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 (9.21) does not modify the circulation. The transformation
also does not modify the external wind speed or direction [because, from (9.16) and
(9.28), 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 a 2 ei ) (9.29)
Incompressible Aerodynamics 263

in the complex z-plane. Hence, Equations (9.10) and (9.13) reveal that

V = V e , (9.30)
L = V e , (9.31)

M = V 2 (c 2 2 ) sin(2 ), (9.32)
8
where V is the wind vector, L the lift vector, and M the moment of the lift vector
about the origin. Here,
e = cos e x + sin ey (9.33)
is a unit vector parallel to the incident wind direction, and

e = e ez = sin e x + cos ey (9.34)

is a unit vector perpendicular to the wind direction. We conclude that, for a cylindri-
cal 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= = (c ) (9.35)
L cos 4

in front of the cross-sections centroid, C, where = /[ V (c+)]. (See Figure 9.2.)


Incidentally, the point, F, at which the line of action of the lift intersects the airfoils
major axis is conventionally termed the focus of the airfoil.

9.4 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 9.3 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 speed
immediately above the layer. [See Equation (9.64).] Moreover, the wake is emitted
by the airfoils trailing edge, and subsequently convected by the external air flow.
(See Section 8.6.) Note that the flow is irrotational everywhere apart from inside the
boundary layer and the wake. According to the analysis of Section 5.8, 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: that is,

d
= z v dS. (9.36)
dt
264 Theoretical Fluid Mechanics

dS
C

Wake
Boundary Layer
U

Trailing Edge Airfoil

Figure 9.3
The boundary layer and wake of a streamlined airfoil. Only shaded regions posses
non-zero vorticity.

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 , (9.37)
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.
Because 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 . (9.38)
dt

According to Equation (9.27), the tangential velocity just above the surface of the
Incompressible Aerodynamics 265

airfoil is

c sin i cos dF
U() = Re
(c 2 sin2 + 2 cos2 )1/2 dz ||=a

(c + )
= V sin( + ) + . (9.39)
(c + ) (c 2 sin2 + 2 cos2 )1/2

Hence, given that the airfoils trailing edge corresponds to = , we obtain


$
c + %
U0 = V sin . (9.40)
(c + )

Thus, Equation (9.38) yields

d
= + sin , (9.41)
dt

where = /[ V (c + )], t = t/t0 , and t0 = /(k V). Assuming that the circulation
of the flow about the airfoil is initially zero (i.e., = 0 at t = 0), the previous
equation can be solved to give
! "
= sin 1 exp(t) . (9.42)

Clearly, as t the normalized circulation asymptotes to the constant value

= sin . (9.43)

The corresponding constant value of the unnormalized circulation is

= V (c + ) sin . (9.44)

According to Equation (9.40), when the circulation, , takes the value the tangen-
tial 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 stagna-
tion point of the flow. This conclusion is known as Zhukovskiis hypothesis, after its
discoverer N. E. Zhukovskii (18471921).
Incidentally, it should be clear, from the previous discussion, that the air circula-
tion 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 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 (as-
suming that it was initially zero), in accordance with the Kelvin circulation theorem.
(See Section 5.8.) This implies, from Equation (9.31), that zero lift would act on
the airfoil, irrespective of its shape, and irrespective of the incident wind speed or
266 Theoretical Fluid Mechanics
2

y/c 0

2
2 1 0 1 2
x/c

Figure 9.4
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.

direction. In other words, flight would be impossible. Fortunately, as long as the tan-
gential 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: that is, such
that the trailing edge is a stagnation point. Thereafter, the circulation remains con-
stant (assuming that the wind speed and direction remain constant). Figures 9.4 and
9.5 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 (9.31) and (9.44), when the air circulation about the air-
foil has attained its steady-state value, , the lift acting on the airfoil becomes

L = V = V 2 (c + ) sin . (9.45)

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
Incompressible Aerodynamics 267
2

y/c 0

2
2 1 0 1 2
x/c

Figure 9.5
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 = .

termed the angle of attack. Finally, from Equations (9.35) and (9.43), the focus of
the airfoil is located a distance

1 sin 1
d= (c ) = (c ) (9.46)
4 4

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 previous 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 ne-
glect 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 8.10. 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 (9.49).] It follows that

1
ds = (dx 2 + dy 2 )1/2 = h() d, (9.47)
2
268 Theoretical Fluid Mechanics

180
()

0
0 10 20
()

Figure 9.6
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-dash
dotted, 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 .

where
h() = (c 2 sin2 + 2 cos2 )1/2 . (9.48)
Moreover, from Equations (9.39) and (9.44), the tangential air speed just above the
surface of the airfoil can be written
f ()
U() = V (c + ) , (9.49)
h()
with
f () = sin( + ) + sin . (9.50)
In addition, it can be shown that


d f
ln = g(), (9.51)
d h

where
cos( + ) (c 2 2 ) cos sin
g() = . (9.52)
f () h 2 ()
Incompressible Aerodynamics 269


y
vz (y = 0+)
z
vz (y = 0)
dz
Figure 9.7
Side view of a vortex sheet.

According to the analysis of Section 8.10, the separation points are located at = ,
where ( ) = 0.1567, and

h 4 () g() f 5 ( ) 
() = 0.47 4 
d . (9.53)
f 5 () 0 h ( )

Here, > + > 0 and 0 > > . Moreover, the x- and y-coordinates of the sep-
aration points are x = (c/2) cos and y = (/2) sin , respectively. Figure 9.6
shows the angular locations of the separation points, calculated as a function of the
angle of attack, for cylindrical airfoils of various dierent ellipticity, /c. (Note that
has been re-expressed as an angle in the range 0 to 2.) It can be seen that for a
blu 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 (because 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. However, in the latter case, as the angle of attack is grad-
ually 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).

9.5 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 9.7.) The vorticity within the sheet can be written
= x (y) e x, (9.54)
270 Theoretical Fluid Mechanics

where (y) is a Dirac delta function. 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 . (9.55)

In other words, a vortex sheet induces a discontinuity in the tangential flow across
the sheet. The previous expression can easily be generalized to give

= n v, (9.56)

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 previous relation
holds locally for non-planar and non-uniform vortex sheets.

9.6 Induced Flow


A vortex filament is necessarily associated with fluid flow circulating about the fila-
ment. 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 equa-
tion
0 j = B (9.57)
can be inverted to give the well-known Biot-Savart law (Fitzpatrick 2008)

1 j(r ) (r r ) 3 
B(r) = d r. (9.58)
4 |r r | 3
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, (9.59)

we can write 
1 (r ) (r r ) 3 
v(r) = d r. (9.60)
4 |r r | 3
This expression allows us to determine the flow field induced by a given vorticity
distribution. In particular, for a vortex filament of intensity the previous expression
reduces to 
1 (r ) (r r ) 
v(r) = dl , (9.61)
4 |r r | 3
Incompressible Aerodynamics 271

where dl is an element of length along the filament. Likewise, for a vortex sheet of
intensity per unit length , we obtain

1 (r ) (r r ) 
v(r) = dS , (9.62)
4 |r r | 3
where dS is an element of area of the sheet.

9.7 Three-Dimensional Airfoils


Let us now take into account the fact that realistic three-dimensional airfoils are of
finite size. Consider Figure 9.8, which shows a top view of a stationary airfoil of
finite size, situated in a (predominately) horizontal wind of velocity V = V e . 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 coor-
dinate 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 9.9.) 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: that is, 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 9.9, which shows a back view of the airfoil shown in Figure 9.8. 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. Because 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 9.8.)
The air that leaves the trailing edge of the wing at some point Q must have im-
pinged on the leading edge at the dierent points P and P , depending on whether
it travelled over the wings upper or lower surfaces, respectively. 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 ac-
quires 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
272 Theoretical Fluid Mechanics

V
x
airfoil

c(z)
z

z = b/2 z=0 z = b/2


Figure 9.8
Top view of a three-dimensional airfoil of finite size.


Y Y z
+
(z)

Figure 9.9
Back view of a three-dimensional airfoil of finite size, indication the pressure varia-
tion over its surface.
Incompressible Aerodynamics 273
x

P z P

Q Q

Figure 9.10
Top view of the airflow over the top (left) and bottom (right) surfaces of a three-
dimensional airfoil of finite size.

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 trail-
ing 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 9.5.) 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 inci-
dent wind. It follows that the vorticity per unit length in the wake can be written

= I(z) e , (9.63)

where I(z) = vz , and vz is tangential velocity discontinuity across the wake. [See
Equation (9.56).]
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 sur-
face. 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 con-
tains 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 (9.56)
yields
S = n U, (9.64)

where n is an outward unit normal to the airfoil surface.


We conclude that a stationary airfoil situated in a uniform wind of constant ve-
locity 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 [because n and U are both essen-
tially parallel to the x-y planesee Equation (9.64)], whereas that in is parallel
to the incident wind direction. (See Figure 9.11.) The vortex filaments within S are
generally termed bound filaments (because they cannot move o the airfoil surface).
Conversely, the vortex filaments within are generally termed free filaments. The
274 Theoretical Fluid Mechanics

S y

Figure 9.11
Vortex structure around a wing.

air velocity both inside and outside S can be written

v = V + v + vS , (9.65)

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+ ), (9.66)


v(P ) = V + v (P ) + vS (P ). (9.67)

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, as the air within S is stationary. Moreover,
v (P+ ) = v (P ) = v (P), because 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 )] . (9.68)
2
This quantity can be identified as the velocity induced at point P by the bound vor-
tices on S , excluding the contribution from the local bound vortex at P (because this
vortex induces equal and opposite velocities at P+ and P ). Finally, taking half the
sum of Equations (9.66) and (9.67), we obtain

1
U(P) = V + v (P) + vS (P). (9.69)
2
Incompressible Aerodynamics 275

9.8 Aerodynamic Forces


The net aerodynamic force acting on an three-dimensional airfoil of finite size can
be written 
A= p n dS , (9.70)
S
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) v 2, where
p0 is a constant pressure. Because a constant pressure exerts no net force on a closed
surface, we get 
1
A = U 2 n dS , (9.71)
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, (9.72)

because n U = 0 on the surface. Hence,



1
A = U (n U) dS . (9.73)
2 S

Making use of Equations (9.64) and (9.69), the previous expression can be written

A = (V + v + vS ) S dS = L + D + F, (9.74)
S

where

L = V S dS , (9.75)
S

D= v S dS , (9.76)
S

F= vS S dS . (9.77)
S

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 : (9.78)
that is, 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
276 Theoretical Fluid Mechanics

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 e . 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.
Making use of the curl theorem, we can easily show that

z dl = (z), (9.79)
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 (9.75) that
 b/2
L = V (z) dz e . (9.80)
b/2

This expression is the generalization of Equation (9.31) for a three-dimensional air-


foil 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 eectively concentrated onto a line)that

v = w(z) e (9.81)

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 (9.76) that
 b/2
D= w(z) (z) dz e . (9.82)
b/2

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 (9.62) and (9.77), the force F is written
 
[S (r ) (r r )] S (r)
F= dS dS  . (9.83)
4 S S  |r r | 3

We can interchange primed and unprimed variables without changing the value of
the integral. Hence,
 
[S (r) (r r)] S (r ) 
F= dS dS . (9.84)
4 S S  |r r| 3
Incompressible Aerodynamics 277

l O x

z
r r z

P
Figure 9.12
Semi-infinite vortex filament.

Taking the half the sum of the previous two equations, we obtain
 
[S (r ) (r r )] S (r) + [(r r ) S (r)] S (r )
F= dS dS  .
8 S S  |r r | 3
(9.85)
However, (a b) c + (b c) a + (c a) b = 0. Thus, the previous expression
yields  
[S (r ) S (r)] (r r )
F= dS dS  . (9.86)
8 S S  |r r | 3
But, the assumption (9.78) 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, because vorticity is divergence
free. Hence,
(z) = (z + dz) + I(z) dz, (9.87)
which implies that
d
I(z) = . (9.88)
dz
Finally, consider a semi-infinite straight vortex filament of vortex intensity =
e x that terminates at the origin, O, as shown in Figure 9.12. 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 (9.61), the induced velocity at P is v = vy ey , where
 /2

vy = cos d = . (9.89)
4 z 0 4 z
278 Theoretical Fluid Mechanics

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  b/2   
1 I(z ) dz 1 d(z )
w(z) = = , (9.90)
4 b/2 z z 4 z z
where use has been made of Equation (9.88).

9.9 Ellipsoidal Airfoils


Consider an ellipsoidal airfoil whose outer surface is specified by the parametric
equations
c0
x= sin cos , (9.91)
2
0
y= sin sin , (9.92)
2
b
z = cos , (9.93)
2
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 : that is,
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 (9.44) holds at fixed z, we deduce that
the air circulation about the wing satisfies

(z) = V c(z) sin = 0 sin , (9.94)

where
0 V c0 . (9.95)
Here, the angle of attack, , is assumed to be small. From Equations (9.90) and
(9.94), the downwash velocity in the region |z| < b/2 is given by


0 cos  d 0 cos d
w() = = 1+ . (9.96)
2 b 0 cos  cos 2 b 
0 cos cos

The integrand appearing in the integral



d
(9.97)
0 cos  cos
Incompressible Aerodynamics 279

is singular when  = . However, we can still obtain a finite value for the integral
by taking its Cauchy principal part: that is,

  
d d
lim + . (9.98)
0 0 cos  cos + cos  cos

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


d

1 sin (1/2) ( +  )
= lim ln

0 cos cos 0 sin sin (1/2) (  ) 0

+

1 sin (1/2) ( + )

+ ln
sin sin (1/2) ( )



1 sin( /2)
= lim ln = 0, (9.99)
0 sin sin( + /2)

which implies that


0
w() = . (9.100)
2b
In the region |z| > b/2, we can write = 2 z/b, so that


0 d 0 ||
w() = 1 = 1  .
2b
(9.101)
2b 0 cos  + 2 1

Hence, we conclude that the downwash velocity profile induced by an ellipsoidal


airfoil takes the form

0 1 |z| < b/2
w(z) = . (9.102)
2 b 1 |z|/(z b /4)
2 2 1/2
|z| > b/2

This profile is shown in Figure 9.13. 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 directly 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.
280 Theoretical Fluid Mechanics
1.2

0.8

0.4

w/(0/2 b) 0

0.4

0.8

1.2

1.6

2
2 1 0 1 2
z/b)

Figure 9.13
Downwash velocity profile induced at the trailing edge by an ellipsoidal airfoil.

It follows from Equation (9.93) and (9.94) that


 
b/2
0 b
(z) dz = sin2 d = 0 b. (9.103)
b/2 2 0 4

Hence, Equation (9.80), (9.82), and (9.100) yield the following expression for the lift
and induced drag acting on an ellipsoidal airfoil,


L= V b 0 , (9.104)
4

D = 02 . (9.105)
8

The surface area of the airfoil in the x-z plane is


S = b c0 . (9.106)
4

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: that is,

b2 4 b
A= = . (9.107)
S c0
Incompressible Aerodynamics 281

It thus follows from Equation (9.95) that

L = L0 , (9.108)
2
D= L0 2 , (9.109)
A
where
L0 = V 2 S . (9.110)

9.10 Simple Flight Problems


Figure 9.14 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 , (9.111)

whereas horizontal force balance gives

T cos( + ) = D cos + L sin . (9.112)

Let us assume that the angles and are both small. According to Equations (9.108)
and (9.109), L L0 and D (2 L0 /A) 2 . Thus, L O() and D O( 2 ).
Moreover, it is clear from Equations (9.111) and (9.112) that T O( 2 ) and W
O(). Thus, to lowest order in , Equation (9.111) yields
W
, (9.113)
L0
whereas Equation (9.112) gives
T 2 W
. (9.114)
W A L0
Expression (9.113) relates the angle of attack to the ratio of the aircrafts weight to
its (theoretical) maximum lift (at a given airspeed). Expression (9.114) relates the
aircrafts angle of controlled (i.e., at constant airspeed) ascent to the thrust developed
282 Theoretical Fluid Mechanics

airfoil T

V


x

D
W

Figure 9.14
Side view of a fixed wing aircraft in flight.

by its engine. An unpowered aircraft, such as a glider, has zero thrust. For such
an aircraft, Equation (9.114) reveals that the angle of controlled decentwhich is
usually termed the glide angletakes the value
2 W
g= . (9.115)
A L0
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. (9.116)
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 9.4, 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 (9.110) and (9.113), the requirement < c is equivalent to

1/2
W
V > Vs = . (9.117)
S c
Incompressible Aerodynamics 283

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.

9.11 Exercises
9.1 Consider the integral

cos(n  ) d
In () = ,
0 cos  cos
where n is a non-negative integer. This integral is defined by its Cauchy
principal value
  
cos(n  ) d cos(n  ) d
In () = lim + .
0 0 cos  cos 
+ cos cos

As was demonstrated in Section 9.9,

I0 = 0.

Show that
I1 = ,
and
In+1 + In1 = 2 cos In ,
and, hence, that
sin(n )
In = .
sin
9.2 Suppose that an airfoil of negligible thickness, and wingspan b, has a width
whose z variation is expressed parametrically as

c() = c sin( ),
=1,3,5,

for 0 , where
b
z= cos .
2
Show that the air circulation about the airfoil takes the form

() = sin( ),
=1,3,5,

where = V c . Here, is the angle of attack (which is assumed to be


284 Theoretical Fluid Mechanics

small). Demonstrate that the downwash velocity at the trailing edge of the
airfoil is  sin( )
w() = .
=1,3,5,
2 b sin

Hence, show that the lift and induced drag acting on the airfoil take the values

L= V b 1 ,
4

D= 2,
8 =1,3,5,

respectively. Demonstrate that the drag to lift ratio can be written



D 2  c 2

= 1 + ,
L A c2
=3,5,7, 1

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).
9.3 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 eectively subject
to two forces. First, its weight, W = W ey , and second its lift L = k v vy e x +
k v v x 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
dv x v vy
= ,
dt h
dvy v v x
= g,
dt h
where h = k g/W is a positive constant with the dimensions of length. Show
that
1 2
v + g y = E,
2
 
where E is a constant. Suppose that v x = g h (1 + u) and vy = g h w, where
|u|, |w|  1. Demonstrate that, to first order in perturbed quantities,

du g
w,
dt h

dw g
2 u.
dt h
Incompressible Aerodynamics 285

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.
286 Theoretical Fluid Mechanics
10
Incompressible Viscous Flow

10.1 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 1.14, the equations governing incompressible
viscous fluid motion can be written

v = 0, (10.1)
Dv
= P + 2 v, (10.2)
Dt
where the quantity
P = p + , (10.3)
which is a combination of the actual fluid pressure, p, and the gravitational potential
energy per unit volume, , is known as the eective pressure. Here, is the fluid
mass density, the fluid viscosity, and the gravitational potential. More informa-
tion on this topic can be found in Batchelor 2000.

10.2 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 measur-
ing 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 10.1.)
Suppose that there is a uniform eective pressure gradient in the x-direction, so
that
dP
= G, (10.4)
dx
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

287
288 Theoretical Fluid Mechanics

y=d

vx(y)
y P

x y=0

Figure 10.1
Viscous flow between parallel plates.

dierence to the final result. Suppose that the fluid velocity profile between the plates
takes the form
v = v x (y) e x . (10.5)
From Section 1.18, this profile automatically satisfies the incompressibility con-
straint v = 0, and is also such that Dv/Dt = 0. Hence, Equation (10.2) reduces
to
P
2v = , (10.6)

or. taking the x-component,
d 2vx G
= . (10.7)
dy 2
If the two plates are stationary then the solution that satisfies the no slip constraint
(see Section 8.2), v x (0) = v x (d) = 0, at each plate is
G
v x (y) = y (d y). (10.8)
2
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
 d
Gd3
Q= v x dy = . (10.9)
0 12
Note that this flux is directly proportional to the eective 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 (10.7)
is

G dy
v x (y) = y (d y) + U . (10.10)
2 d
Incompressible Viscous Flow 289

Figure 10.2
Viscous flow down an inclined plane.

Hence, the modified velocity profile is a combination of parabolic and linear profiles.
This type of flow is known as Couette flow, in honor of Maurice Couette (1858
1943). The net volume flux (per unit width) of fluid between the plates becomes

Gd3 1
Q= + U d. (10.11)
12 2

10.3 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 trans-
verse 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 10.2.)
The generalized pressure gradient within the fluid is written
dP
= G = g sin , (10.12)
dx
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 , (10.13)
290 Theoretical Fluid Mechanics

and Equation (10.2) again reduces to

d 2vx G
= , (10.14)
dy 2

where
G = g sin . (10.15)
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), because there is nothing above the interface that can exchange mo-
mentum with the fluid (assuming that the finite inertia and viscosity of air are both
negligible.) Hence, from Section 1.18, we get the boundary condition

dv x 
 = 0. (10.16)
dy y=h

The solution to Equation (10.14) that satisfies the boundary conditions is


G
v x (y) = y (2 h y). (10.17)
2
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
 h
G h 3 g sin h 3
Q= v x dy = = , (10.18)
0 3 3
where use has been made of Equation (10.15). 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
h= . (10.19)
g sin

Suppose that the rate at which fluid is poured down the plane is suddenly in-
creased 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. (10.20)
dQ

Let the interface between the layers of dierent 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
Incompressible Viscous Flow 291

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, (10.21)

or
1/3
dQ G h 2 9 Q 2 g sin
V= = = . (10.22)
dh
As can easily be verified, this velocity is twice the maximum fluid velocity in the
layer.

10.4 Poiseuille Flow


Steady viscous fluid flow driven by an eective pressure gradient established between
the two ends of a long straight pipe of uniform circular cross-section is generally
known as Poiseuille flow, because it was first studied experimentally by Jean Poiseuille
(17971869) 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 (10.23)

is the uniform eective pressure gradient along the pipe, and

v = vz (r) ez (10.24)

the time independent velocity profile driven by this gradient. It follows from Sec-
tion 1.19 that v = 0 and Dv/Dt = 0. Hence, Equation (10.2) reduces to
G
2v = ez . (10.25)

Taking the z-component of this equation, we obtain


1 d dvz G
r = , (10.26)
r dr dr

where use has been made of Equation (1.155). The most general solution of the
previous equation is
G 2
vz (r) = r + A ln r + B, (10.27)
4
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
292 Theoretical Fluid Mechanics

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 ). (10.28)
4
The volume flux of fluid down the pipe is
 a
G a4
Q= 2 r vz dr = . (10.29)
0 8
According to the previous analysis, the quantity Q/a 4 should be directly proportional
to the eective pressure gradient along the pipe. The accuracy with which experi-
mental 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 (10.29), 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.

10.5 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, after Maurice Couette and Georey
Taylor (18861975).
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 , (10.30)
where (r) = v (r)/r is the angular velocity profile. Application of the no slip con-
dition at the two shells leads to the boundary conditions
(r1 ) = 1 , (10.31)
(r2 ) = 2 . (10.32)
It again follows from Section 1.19 that v = 0 and Dv/Dt = 0. Hence, Equa-
tion (10.2) reduces to
P
2v = . (10.33)

Incompressible Viscous Flow 293

Assuming that P = 0 within the fluid, because 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 1.19, the previous expression yields


1 d v v
r 2 = 0, (10.34)
r dr dr r
or

1 d 3 d
r = 0. (10.35)
r 2 dr dr
The solution of Equation (10.35) that satisfies the boundary conditions is

1 1 2 2 r12 1 r22
(r) = 2 2
+ . (10.36)
r r1 r22 r12 r22

It can be seen that this angular velocity profile is a combination of the solid body
rotation profile = constant, and the irrotational rotation profile r 2 .
From Section 1.19, the only non-zero component of the viscous stress tensor
within the fluid is
d $ v % d
r = r = r . (10.37)
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 r1 r (r1 ) = 4 2
2 . (10.38)
r1 r22

Likewise, the torque per unit height exerted on the outer cylinder is

1 2

2 = 2 r2 r (r2 ) = 4 2
2 . (10.39)
r1 r22

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).

10.6 Flow in Slowly-Varying Channels


According to Section 10.1, the equations governing steady, incompressible, viscous
fluid flow are

v = 0, (10.40)
(v ) v = P + 2 v. (10.41)
294 Theoretical Fluid Mechanics

As we saw in Sections 10.2 and 10.4, 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
P
2v = . (10.42)

Suppose, however, that the cross-section of the channel varies along its length. As
we shall demonstrate, provided this variation is suciently slow, the flow is still
approximately described by the previous 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 , (10.43)

where G(x) also varies on the same lengthscale. Assuming that /x O(1/l) and
/y O(1/d), it follows from Equation (10.40) that


vy d
O . (10.44)
vx l
Hence,


v x2
[(v ) v] x O , (10.45)
l

2
2 v x d
( v) x =
2

1 + O . (10.46)
y 2 l

The x-component of Equation (10.41) reduces to




2
2 v x vx d 2 d G

+ O + O = .
l
1 (10.47)
y 2 l

Thus, if
d
 1, (10.48)
l
vx d 2
1 (10.49)
l
in other words, if the channel is suciently narrow, and its cross-section varies
suciently slowly along its lengththen Equation (10.47) can be approximated as
2vx G
. (10.50)
y 2
This, of course, is the same as the equation governing steady, two-dimensional, vis-
cous flow between exactly parallel plates. (See Section 10.2.) Assuming that the
Incompressible Viscous Flow 295

plates are located at y = 0 and y = d(x), and making use of the analysis of Sec-
tion 10.2, the appropriate solution to the previous equation is
G(x)
v x (x, y) = y [d(x) y]. (10.51)
2
The volume flux (per unit width) of fluid between the plates is thus
 d
G(x) d 3 (x)
Q= v x dy = . (10.52)
0 12
However, for steady incompressible flow, this flux must be independent of x, which
implies that
dP
G(x) = = 12 Q d 3(x). (10.53)
dx
Suppose that a constant dierence in eective pressure, P, is established between
the fixed points x = x1 and x = x2 , where x2 x1 = l. Integration of the previous
equation between these two points yields

P = 12 Q l d 3 , (10.54)
 x2
where   = x ( ) dx/l. Hence, the volume flux (per unit width) of fluid between
1
the plates that is driven by the eective pressure dierence becomes
P 1
Q= . (10.55)
l 12 d 3 
Moreover, the eective pressure gradient at a given point is
dP P 1
G(x) = = , (10.56)
dx l d 3 (x) d 3 
which allows us to determine the velocity profile at that point from Equation (10.51).
Thus, given the average eective 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 previously, but adapting the anal-
ysis of Section 10.4, rather than that of Section 10.1, 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
G(z)  2 
vz (r, z) = a (z) r 2 , (10.57)
4
P 1
G(z) = , (10.58)
l a (z) a 4
4

P
Q= . (10.59)
l 8 a 4 
Here, Q is the volume flux of fluid down the pipe, P = P(z2 ) P(z1 ), l = z2 z1 ,
296 Theoretical Fluid Mechanics

y
d1 d
U d2
x

Figure 10.3
The lubrication layer between two planes in relative motion.

 z2
and   = z ( ) dz/l. The approximations used to derive the previous results are
1
valid provided
a
 1, (10.60)
l
vz a 2
 1. (10.61)
l

10.7 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 layer. This 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 oers 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 10.3.) 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, (10.62)
Incompressible Viscous Flow 297

where
d1 d2
= . (10.63)
l
As discussed in the previous section, provided that

d
 1, (10.64)
l
U d2
 1, (10.65)
l
the cross-section of the channel between the two bodies is suciently slowly varying
in the x-direction that the channel can be treated as eectively uniform at each point
along its length. Thus, it follows from Equation (10.10) that the velocity profile
within the channel takes the form

G(x) ! " d(x) y
v x (x, y) = y d(x) y + U , (10.66)
2 d

where
dp
G(x) = (10.67)
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
 d
G(x) d 3 (x) 1
Q= v x dy = + U d(x). (10.68)
0 12 2

Of course, in a steady state, this flux must be independent of x. Hence,



dp U 2Q
= G(x) = 6 2 3 , (10.69)
dx d (x) d (x)

where d(x) = d1 x. Integration of the previous equation yields




6 1 1 1 1
p(x) p0 = U Q 2 2 , (10.70)
d d1 d d1

where p0 = p(0). 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
previous equation that

d1 d2
Q=U , (10.71)
d1 + d2
and
6 U [d1 d(x)] [d(x) d2 ]
p(x) p0 = . (10.72)
d 2 (x) (d1 + d2 )
298 Theoretical Fluid Mechanics

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  l


6U d1 d1 d2
fy = [p(x) p1 ] dx = 2 ln 2 . (10.73)
0 d2 d1 + d2
Moreover, the net tangential force (per unit width) acting on the lower plane is




l
v x 2U d1 d1 d2
fx = dx = 2 ln 3 . (10.74)
0 y y=0 d2 d1 + d2

Of course, equal and opposite forces,




U l2 3 1+k
f x
= fx = ln 2k , (10.75)
d02 2 k2 1k


 U l 1 1+k
fy = fy = 2 ln 3k , (10.76)
d0 k 1k

act on the upper plane. Here, d0 = (d1 + d2 )/2 is the mean channel width, and
k = (d1 d2 )/(d1 + d2 ). It can be seen that if 0 < d2 < d1 then 0 < k < 1. The
eective coecient 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,
fx f  4 d0
Cf = = x = H(k), (10.77)
fy fy 3 l
where
)

1+k 3k 1+k
H(k) = k ln ln 2k . (10.78)
1k 2 1k
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 eective coecient
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 coecient 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.
Incompressible Viscous Flow 299

10.8 Stokes Flow


Steady flow in which the viscous force density in the fluid greatly exceeds the ad-
vective inertia per unit volume is generally known as Stokes flow, in honor of George
Stokes (18191903). Because, 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 1.16), Stokes flow implies Reynolds numbers that are much less than unity.
In the time independent, low Reynolds number limit, Equations (10.1) and (10.2)
reduce to

v = 0, (10.79)
0 = P + 2 v. (10.80)

It follows from these equations that

P = 2 v = ( v) = , (10.81)

where = v, and use has been made of Equation (A.177). Taking the curl of
this expression, we obtain
2 = 0, (10.82)
which is the governing equation for Stokes flow. Here, use has been made of Equa-
tions (A.173), (A.176), and (A.177).

10.9 Axisymmetric Stokes Flow


Let r, , be standard spherical coordinates. Consider axisymmetric Stokes flow
such that
v(r) = vr (r, ) er + v (r, ) e . (10.83)
According to Equations (A.175) and (A.176), we can automatically satisfy the in-
compressibility constraint (10.79) by writing (see Section 7.4)

v = , (10.84)

where (r, ) is the Stokes stream function (i.e., v = 0). It follows that

1
vr (r, ) = , (10.85)
sin
r2
1
v (r, ) = . (10.86)
r sin r
300 Theoretical Fluid Mechanics

Moreover, according to Section C.4, r = = 0, and

1 (r v ) 1 vr L()
(r, ) = = , (10.87)
r r r r sin
where (see Section 7.4)

2 sin 1
L= + 2 . (10.88)
r 2 r sin
Hence, given that || = 1/(r sin ), we can write

= v = L() . (10.89)

It follows from Equations (A.176) and (A.178) that

= [L()]. (10.90)

Thus, by analogy with Equations (10.84) and (10.89), and making use of Equa-
tions (A.173) and (A.177), we obtain

( ) = 2 = L 2 () . (10.91)

Equation (10.82) implies that


L 2 () = 0, (10.92)
which is the governing equation for axisymmetric Stokes flow. In addition, Equa-
tions (10.81) and (10.90) yield

P = [L()]. (10.93)

10.10 Axisymmetric Stokes Flow Around a Solid Sphere


Consider a solid sphere of radius a that is moving under gravity at the constant ver-
tical velocity V ez through a stationary fluid of density and viscosity . Here,
gravitational acceleration is assumed to take the form g = g ez . Provided that the
typical Reynolds number,
2V a
Re = , (10.94)

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.,
Incompressible Viscous Flow 301

/ = 0), and has no toroidal component (i.e., v = 0). The physical boundary
conditions at the surface of the sphere are
vr (a, ) = 0, (10.95)
v (a, ) = 0 : (10.96)
that is, 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 , (10.97)
v (r , ) V sin . (10.98)
Let us write
v = , (10.99)
where (r, ) is the Stokes stream function. As we saw in the previous section, ax-
isymmetric Stokes flow is characterized by
L 2 () = 0. (10.100)
Here, the dierential operator L is specified in Equation (10.88). The boundary
conditions (10.95)(10.98) reduce to


 = 0, (10.101)
r r=a


 = 0, (10.102)
r=a
1
(r , ) V r 2 sin2 . (10.103)
2
Equation (10.103) suggests that (r, ) can be written in the separable form
(r, ) = sin2 f (r). (10.104)
In this case,
2 cos f (r)
vr (r, ) = , (10.105)
r2
sin d f
v (r, ) = , (10.106)
r dr
and Equations (10.100)(10.103) reduce to

2 2
d 2
f = 0, (10.107)
dr 2 r 2

d f 
f (a) =  = 0, (10.108)
dr r=a
1
f (r ) V r 2. (10.109)
2
302 Theoretical Fluid Mechanics
3

z/a 0

3
3 2 1 0 1 2 3
x/a

Figure 10.4
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 (10.107) of the form f (r) = r n . We find
that
[n (n 1) 2] [(n 2) (n 3) 2] = 0, (10.110)
which implies that n = 1, 1, 2, 4. Hence, the most general solution to Equa-
tion (10.107) is
A
f (r) = + B r + C r 2 + D r 4 , (10.111)
r
where A, B, C, D are arbitrary constants. However, the boundary condition (10.109)
yields C = (1/2) V and D = 0, whereas the boundary condition (10.108) gives
A = (1/4) V a 3 and B = (3/4) V a. Thus, we conclude that

V (r a)2 (2 r + a)
f (r) = , (10.112)
4r
and the stream function becomes
V (r a)2 (2 r + a)
(r, ) = sin2 . (10.113)
4r
(See Figure 10.4.) From Equation (10.87), the fluid vorticity is


L() sin d 2 2 3 V a sin
(r, ) = = f = . (10.114)
r sin r dr 2 r 2 2r2
Incompressible Viscous Flow 303
3

z/a 0

3
3 2 1 0 1 2 3
x/a

Figure 10.5
Contours of the vorticity, , in the x-z plane for Stokes flow around a solid sphere.
Solid/dashed lines correspond to opposite signs of .

(See Figure 10.5.) Moreover, from Equation (10.81),

P = = ( r sin ). (10.115)

Hence,

P 3 V a cos
= , (10.116)
r r3
P 3 V a sin
= , (10.117)
2r2
which implies that the eective pressure distribution within the fluid is

3 V a cos
P(r, ) = p0 + , (10.118)
2r2
where p0 is an arbitrary constant. (See Figure 10.6.) However, P = p + , where
= g z = g r cos . Thus, the actual pressure distribution is

3V a
p(r, ) = p0 g r cos + cos . (10.119)
2r2
From Section 1.20, the radial and tangential components of the force per unit
304 Theoretical Fluid Mechanics
3

z/a 0

3
3 2 1 0 1 2 3
x/a

Figure 10.6
Contours of the eective 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 .

area exerted on the sphere by the fluid are




vr
fr () = rr (a, ) = p + 2 , (10.120)
r r=a


1 vr v v
f () = r (a, ) = + . (10.121)
r r r r=a

Now, vr (a, ) = v (a, ) = 0. Moreover, because v = 0, it follows from Equa-


tion (1.168) that (vr /r)r=a = 0. Finally, Equation (10.87) yields (v /r)r=a =
(a, ). Hence,

3V
fr () = p(a, ) = p0 + g a cos cos , (10.122)
2a
3V
f () = (a, ) = sin . (10.123)
2a

Thus, the force density at the surface of the sphere is

3V
f() = ez + (p0 + g a cos ) er . (10.124)
2a
Incompressible Viscous Flow 305

It follows that the net vertical force exerted on the sphere by the fluid is

Fz = f ez dS
S

3V
= 4 a 2 + 2 a 2 (p0 + g a cos ) cos sin d, (10.125)
2a 0

which reduces to
4 3
Fz = 6 a V + a g. (10.126)
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 previous equation as
the buoyancy force due to the weight of the fluid displaced by the sphere. (See
Chapter 2.) 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, (10.127)

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
V= 1 , (10.128)
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 a 3 g  
Re = = 1 . (10.129)
9 2  
For the case of a grain of sand falling through water at 20 C, we have / 2
and = 1.0 106 m2 /s (Batchelor 2000). Hence, Re = (a/6 105 ) 3 , where a
is measured in meters. Thus, expression (10.128), 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 (Batchelor 2000). Hence, Re = (a/4
105 ) 3 , where a is measured in meters. Thus, expression (10.128) only holds for
306 Theoretical Fluid Mechanics

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 (10.105), (10.106), and (10.112) yield

3 a $ a %2
vr (r, ) = V cos + V cos + O , (10.130)
2 r r
3 a $ a %2
v (r, ) = V sin V sin + O . (10.131)
4 r r
It follows that

vr v vr v2 V 2 a

[ (v ) v]r = vr + , (10.132)
r r r r2

and
2 vr V a
( 2 v)r 3 . (10.133)
r 2 r
Hence,
[ (v ) v]r V r $r%
Re , (10.134)
( 2 v)r a
where Re is the Reynolds number of the flow in the immediate vicinity of the sphere.
[See Equation (10.94).] Our analysis is based on the assumption that advective in-
ertia is negligible with respect to viscosity. However, as is clear from the previous
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: that is, for r
a/Re. In other
words, inertia always dominates viscosity, and our Stokes flow solution therefore
breaks down, at suciently large r/a.

10.11 Axisymmetric Stokes Flow In and Around a Fluid Sphere


Suppose that the solid sphere discussed in the previous section is replaced by a spher-
ical 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
$A %
(r, ) = sin2 + Br + C r2 + Dr4 , (10.135)
r
Incompressible Viscous Flow 307

and
A
(r, ) = sin + B r + C r + D r ,
2 2 4
(10.136)
r
respectively. Here, A, B, C, et cetera, are arbitrary constants. Likewise, the previous
analysis also allows us to deduce that
$A B %
vr (r, ) = 2 cos 3 + + C + D r 2 , (10.137)
r r
$ A B %
v (r, ) = sin 3 + + 2 C + 4 D r 2 , (10.138)
r r


6A
r (r, ) = sin 4 + 6 D r , (10.139)
r


12 A 6 B
rr (r, ) = p0 + g r cos + cos + 2 + 12 D r (10.140)
r4 r
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, ) = 0
that is, the normal velocity at the drop boundary must be zerootherwise, 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 be-
tween the two fluids are, firstly, continuity of tangential velocitythat is, v (a , ) =
v (a+ , )and, secondly, continuity of tangential stressthat is, r (a , ) = r (a+ , ).
These constraints yield A/a 3 + B/a V = 2 C + 4 D a 2 and 6 A/a 4 = 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 a 2+ 3 r
(r, ) = V a2 sin2 +2 , (10.141)
4 + r + a a
and

$ r %2
1 $ r %2
(r, ) = V a2 sin2 1 , (10.142)
4 + a a
respectively. (See Figures 10.7 and 10.8.)
The discontinuity in the radial stress across the drop boundary is
rr (a+ , ) rr (a , ) = p0 p0 + ( ) g a cos

V + (3/2)
3 cos . (10.143)
a +
308 Theoretical Fluid Mechanics

z/a 0

2
2 1 0 1 2
x/a

Figure 10.7
Contours of the Stokes stream function in the x-z plane for Stokes flow in and around
a fluid sphere. Here, / = 10.

z/a 0

2
2 1 0 1 2
x/a

Figure 10.8
Contours of the Stokes stream function in the x-z plane for Stokes flow in and around
a fluid sphere. Here, / = 1/10.
Incompressible Viscous Flow 309

The final physical constraint that must be satisfied at r = a is


2
rr (a+ , ) rr (a , ) = , (10.144)
a
where is the surface tension of the interface between the two fluids. (See Sec-
tion 3.3.) Hence, we obtain
2
p0 p0 = , (10.145)
a
and

a2 g +
V= 1 , (10.146)
3 + (3/2)
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 (10.128): that is, 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= . (10.147)
3

10.12 Exercises
10.1 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)
g sin
v x (y) = y (2 h y),
2
p(y) = p0 + g cos (h y),

where is the kinematic viscosity, the density, and p0 is atmospheric pres-


sure.
10.2 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
Q= (G + g sin ) ,
8
310 Theoretical Fluid Mechanics

where a is the pipe radius, the fluid viscosity, the fluid density, and G the
pressure gradient.
10.3 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 eective pressure gradient, and the viscosity. Find the mean
flow speed.
10.4 Consider viscous flow along a cylindrical pipe of elliptic cross-section. Sup-
pose 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

2 2 G
+ vz = ,
x 2 y 2
where G is the eective pressure gradient, and the fluid viscosity. Show
that
G a2 b2 b2 x2 a2 y2
vz (x, y) =
2 a2 + b2
is a solution of this equation that satisfies the no slip condition at the bound-
ary. Demonstrate that the flow rate is

G a3 b3
Q= .
4 a2 + b2
Finally, show that a pipe with an elliptic cross-section has lower flow rate
than an otherwise similar pipe of circular cross-section that has the same
cross-sectional area.
10.5 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 advec-
tive inertia) with uniform pressure (neglecting gravity) provided that


d 4 d
r = 0.
dr dr
Incompressible Viscous Flow 311

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

a3
(r) = 0 ,
r3
for r a. Show that the torque that the sphere exerts on the fluid is

= 8 0 a 3 .

10.6 Consider a solid sphere of radius a moving through a viscous fluid of vis-
cosity at the fixed velocity V = V ez . Let r, , be spherical coordinates
whose origin coincides with the instantaneous location of the spheres cen-
ter. 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
3 a
vr V cos ,
2 r
3 a
v V sin ,
4 r
9 a
rr p0 V cos 2 ,
2 r
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.


312 Theoretical Fluid Mechanics
11
Waves in Incompressible Fluids

11.1 Introduction
This chapter investigates low amplitude waves propagating through incompressible
fluids. More information on this topic can be found in Lighthill 1978, and Milne-
Thomson 2011.

11.2 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 suciently 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.
Because water is essentially incompressible, its equations of motion are

v = 0, (11.1)
v
+ (v ) v = p g ez + 2 v, (11.2)
t
where is the (uniform) mass density, the (uniform) viscosity, and g the (uniform)
acceleration due to gravity. (See Section 1.14.) Let us write

p(r, t) = p0 g z + p1 (r, t), (11.3)

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 2.) Substitution into Equation (11.2) yields
v
p1 + 2 v, (11.4)
t
where we have neglected terms that are second order in small quantities (i.e., terms
of order v 2 ).

313
314 Theoretical Fluid Mechanics

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, vis-
cosity is negligible as long as
(2 /g)1/3 5 105 m.] It follows that

v
p1 . (11.5)
t
Taking the curl of this equation, we obtain


0, (11.6)
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 = , (11.7)
where (r, t) is the velocity potential. (See Section 4.15.) However, from Equa-
tion (11.1), the velocity field is also divergence free. It follows that the velocity
potential satisfies Laplaces equation,

2 = 0. (11.8)

Finally, Equations (11.5) and (11.7) yield


p1 = . (11.9)
t
We now need to derive the physical constraints that must be satisfied at the wa-
ters upper and lower boundaries. It is assumed that the water is bounded from below
by a solid surface located at z = d. Because the water must always remain in con-
tact 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. (11.10)
z z=d

The waters upper boundary is a little more complicated, because it is a free surface.
Let represent the vertical displacement of this surface due to the wave. It follows
that 

= vz |z=0 =  . (11.11)
t z z=0
The appropriate physical constraint at the upper boundary is that the water pressure
there must equal atmospheric pressure, because there cannot be a pressure discon-
tinuity across a free surface (in the absence of surface tensionsee Section 11.11).
Accordingly, from Equation (11.3), we obtain

p0 = p0 g + p1 |z=0 , (11.12)
Waves in Incompressible Fluids 315

or
g = p1 |z=0 , (11.13)
which implies that  
 p1 
g = g  =  , (11.14)
t z z=0 t z=0
where use has been made of Equation (11.11). The previous expression can be com-
bined with Equation (11.9) to give the boundary condition
 
 1 2 
 = g  . (11.15)
z z=0 t 2 z=0
Let us search for a wave-like solution of Equation (11.8) of the form
(r, t) = F(z) cos( t k x). (11.16)
This solution actually corresponds to a propagating plane wave of wave vector k =
k e x , angular frequency , and amplitude F(z) (Fitzpatrick 2013). Substitution into
Equation (11.8) yields
d 2F
k 2 F = 0, (11.17)
dz 2
whose independent solutions are exp(+k z) and exp(k z). Hence, a general solution
to Equation (11.8) takes the form
(x, z, t) = A e k z cos( t k x) + B ek z cos( t k x), (11.18)
where A and B are arbitrary constants. The boundary condition (11.10) is satisfied
provided that B = A exp(2 k d), giving
 
(x, z, t) = A e k z + ek (z+2 d) cos( t k x), (11.19)
The boundary condition (11.15) then yields
  2  
A k 1 e2 k d cos( t k x) = A 1 + e2 k d cos( t k x). (11.20)
g
which reduces to the dispersion relation
2 = g k tanh(k d). (11.21)
The type of wave described in this section is known as a gravity wave.

11.3 Gravity Waves in Deep Water


Consider the so-called deep water limit,
k d
1, (11.22)
316 Theoretical Fluid Mechanics

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 (11.21) reduces to

= (g k)1/2 , (11.23)

because tanh(x) 1 as x . It follows that the phase velocity of gravity waves


in deep water is
$ g %1/2
vp = = . (11.24)
k k
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. 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 (11.23)] (Fitzpatrick 2013). 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 dierent wave numbers centered on k. Such a construct is known as a wave
pulse (Fitzpatrick 2013). As is well known, wave pulses propagate at the group
velocity (Fitzpatrick 2013),
d
vg = . (11.25)
dk
For the case of gravity waves in deep water, the dispersion relation (11.23) yields
1 $ g %1/2 1
vg = = vp. (11.26)
2 k 2
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, (11.27)
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 (11.19) yields

(x, z, t) = A e k z cos( t k x). (11.28)

Hence, [cf., Equations (11.45)(11.48)]

x (x, z, t) = a e k z cos( t k x), (11.29)


z (x, z, t) = a e k z sin( t k x), (11.30)
v x (x, z, t) = a e k z sin( t k x), (11.31)
vz (x, z, t) = a e kz
cos( t k x), (11.32)

and
p1 = g z , (11.33)
Waves in Incompressible Fluids 317

surface
z

x z=0

Figure 11.1
Motion of water particles associated with a deep water gravity wave propagating in
the x-direction.

where use has been made of Equations (11.7), (11.9), and (11.27). Here, a is the
amplitude of the vertical oscillation at the waters surface. According to Equa-
tions (11.29)(11.32), 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. The radius of the orbit decreases exponentially with increasing depth. Fur-
thermore, 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 11.1.)
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 (11.3) and (11.33) yield
p(x, z, t) = p0 + g h(x, z, t). (11.34)
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.

11.4 Gravity Waves in Shallow Water


Consider the so-called shallow water limit,
k d  1, (11.35)
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 (11.21) reduces to
= (g d)1/2 k, (11.36)
318 Theoretical Fluid Mechanics

because 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 , (11.37)

irrespective of wave number. We conclude thatunlike deep water wavesshallow


water gravity waves are non-dispersive in nature (Fitzpatrick 2013). In other words,
in shallow water, waves pulses and plane waves all propagate at the same speed. It
can be seen that the velocity (11.37) increases with increasing water depth.
For a plane wave of wave number k = k e x , in the limit k d  1, Equation (11.19)
yields
(x, z, t) = A [1 + k 2 (z + d)2 /2] cos( t k x). (11.38)
Hence, Equations (11.7) and (11.27) give [cf., Equations (11.45)(11.48)]

x (x, z, t) = a (k d)1 cos( t k x), (11.39)


z (x, z, t) = a (1 + z/d) sin( t k x), (11.40)
v x (x, z, t) = a (k d)1 sin( t k x) (11.41)
vz (x, z, t) = a (1 + z/d) cos( t k x). (11.42)

Here, a is again the amplitude of the vertical oscillation at the waters surface. Ac-
cording to the previous expressions, the passage of a shallow water gravity wave
causes a water particle located a depth h below the surface to execute an elliptical or-
bit, of horizontal radius a/(k d), and vertical radius a (1 h/d), about its equilibrium
position. 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.

11.5 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
tanh(k d)
v p = (g d) 1/2
. (11.43)
kd
Moreover, the ratio of the group to the phase velocity is

vg 1 2kd
= 1+ . (11.44)
vp 2 sinh(2 k d)
Waves in Incompressible Fluids 319

It follows 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
cosh[k (z + d)]
x (x, z, t) = a cos( t k x), (11.45)
sinh(k d)
sinh[k (z + d)]
z (x, z, t) = a sin( t k x), (11.46)
sinh(k d)
cosh[k (z + d)]
v x (x, z, t) = a sin( t k x), (11.47)
sinh(k d)
sinh[k (z + d)]
vz (x, z, t) = a cos( t k x). (11.48)
sinh(k d)
The mean kinetic energy per unit surface area associated with a gravity wave is
defined 
1 2
K= v dz, (11.49)
d 2
where
(x, t) = a sin( t k x) (11.50)
is the vertical displacement at the surface, and
 2
d(k x)
  = ( ) (11.51)
0 2

is an average over a wavelength. Given that cos2 ( tk x) = sin2 ( tk x) = 1/2,
it follows from Equations (11.47) and (11.48) that, to second order in a,
 0
1 cosh[2 k (z + d)] 1 2
K = a2 2 dz = g a 2 . (11.52)
4 d
2
sinh (k d) 4 g k tanh(k d)
Making use of the general dispersion relation (11.21), we obtain
1
K= g a 2. (11.53)
4
The mean potential energy perturbation per unit surface area associated with a
gravity wave is defined

1
U= g z dz + g d 2 , (11.54)
d 2
which yields
1 1 1
U =  g ( 2 d 2 ) + g d 2 = g  2 , (11.55)
2 2 2
or
1
U= g a 2. (11.56)
4
320 Theoretical Fluid Mechanics

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
1
E = K + U = g a 2. (11.57)
2
This energy depends on the wave amplitude at the surface, but is independent of the
wavelength, or the water depth.

11.6 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. Because 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 (11.43), yields
tanh(k d) V 2
= . (11.58)
kd gd
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 perpendicular 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 (Fitzpatrick 2013), 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. (11.59)
2 2
However, because V = v p , we obtain


1 vg 1 2kd
D = g a2 w 1 = g a2 w 1 , (11.60)
2 vp 4 sinh(2 k d)
Waves in Incompressible Fluids 321

where use has been made of Equation (11.44). Here, k d is determined implicitly in
terms of the ship speed via expression (11.58). Note that this expression cannot be
satisfied when the speed exceeds the critical value (g d)1/2, because gravity waves
cannot propagate at speeds in excess of this value. In this situation, no transverse
wave train can keep up with the ship, and the drag associated with such waves con-
sequently disappears. In fact, we can see, from the previous formulae, that when
V (g d)1/2 then k d 0, and so D 0. However, 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 eectively falling to zero as this speed is exceeded. Conse-
quently, 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 (11.60) reduces to
1
D= g a 2 w. (11.61)
4
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 . (11.62)
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 such that the phase
velocity of the waves matches the ships speed, V) produced by the ship is
 $ x%
1 l sin( k l) k l
a cos cos(k x) dx = : (11.63)
l 0 l kl +kl
that is, the amplitude is proportional to the Fourier coecient of the ships vertical
displacement pattern evaluated for a wave number that matches that of the wave train
(Fitzpatrick 2013). Hence, (at fixed bow wave displacement) the drag produced by
the transverse waves is
2
sin( Fr 2 ) 1
D a2 , (11.64)
Fr 2 1 + Fr 2
322 Theoretical Fluid Mechanics

1.1
1
0.9
0.8
D(a.u.) 0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8 1 1.2
Fr

Figure 11.2
Variation of wave drag with Froude number for a ship traveling through deep water.

where the dimensionless parameter

V
Fr = (11.65)
(g l)1/2

is known as the Froude number. (See Section 1.16.)


Figure 11.2 illustrates the variation of the wave drag with Froude number pre-
dicted by Equation (11.64). 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 inecient at driving short wavelength gravity waves. It can
also be seen that the drag increases as the Froude
number increases, reaching a rela-
tively 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 in-
stance, 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
Waves in Incompressible Fluids 323

implies a maximum speed of

Vc = 0.56 (g l)1/2 = 1.75 [l(m)]1/2 m/s = 3.4 [l(m)]1/2 kts. (11.66)

This characteristic speed is sometimes called the hull speed. It can be seen that the
hull speed increases with the length of the ship. In other words, long ships tend to
have higher hull speeds than short ones.

11.7 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 11.3. 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. 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
vp
sin = . (11.67)
V
This, therefore, is the condition that must be satisfied in order for an obliquely prop-
agating 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: that is,

v p = (g d)1/2, (11.68)

where d is the water depth. Hence, Equation (11.67) yields


1/2

1 (g d)
= sin . (11.69)
V
This equation can only be satisfied when

V > (g d)1/2 . (11.70)


324 Theoretical Fluid Mechanics

y D D

E
vp t


x
A A
Vt

C C

Figure 11.3
An oblique plane wave generated on the surface of the water by a moving ship.

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 Equa-
tion (11.69). This implies the scenario illustrated in Figure 11.4. 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  , et cetera.
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 (11.69), the faster the ship travels over the water, the smaller the angle
becomes. Shallow water wakes are especially dangerous to other vessels, and par-
ticularly 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, because, 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 gravity waves is v p = (g/k)1/2. Thus, Equation (11.67)
yields
v p $ g %1/2
sin = = . (11.71)
V kV2
It follows that in deep water any obliquely propagating gravity wave whose wave
number exceeds the critical value
g
k0 = (11.72)
V2
Waves in Incompressible Fluids 325

y
D
B 

B


x
A A A

C

C 
E

Figure 11.4
A shallow water wake.

can keep up with the ship, as long as its direction of propagation is such that Equa-
tion (11.71) is satisfied. In other words, the ship continuously excites gravity waves
with a wide range of dierent wave numbers and propagation directions. The wake
is essentially the interference pattern generated by these waves. As is well known
(Fitzpatrick 2013), an interference maximum generated by the superposition of plane
waves with a range of dierent 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: that is, vg = v p /2.
Consider Figure 11.5. 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.
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. Because 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 , (11.73)
y = vg t cos . (11.74)

Moreover,
dy
= tan , (11.75)
dx
because BC is the tangent to the curve APDthat is, the curve y(x)at point P. It
326 Theoretical Fluid Mechanics

y D C

P
vg t

x
A B A
Vt

Figure 11.5
Formation of an interference maximum in a deep water wake.

follows from Equation (11.71), and the fact that vg = v p /2, that


1
x = X 1 sin2 , (11.76)
2
1
y= X sin cos , (11.77)
2
where X() = V t. The previous three equations can be combined to produce
dy dy/d (1/2) dX/d sin cos + (1/2) X (cos2 sin2 )
= =
dx dx/d dX/d [1 (1/2) sin2 ] X sin cos
sin
= , (11.78)
cos
which reduces to
dX X
= . (11.79)
d tan
This expression can be solved to give
X = X0 sin , (11.80)
where X0 is a constant. Hence, the locus of our interference maximum is determined
parametrically by


1
x = X0 sin 1 sin2 , (11.81)
2
1
y= X0 sin2 cos . (11.82)
2
Waves in Incompressible Fluids 327

0.3

0.2 D

0.1
y/X0
A
0 C

0.1

0.2 B

0.3

0 0.1 0.2 0.3 0.4 0.5 0.6


x/X0

Figure 11.6
Locus of an interference maximum in a deep water wake.

Here, the angle ranges from /2 to +/2. The curve specified by the previous
equations is plotted in Figure 11.6. As usual, A is the instantaneous position of the
ship. It can be seen that the interference maximum essentially consists of the trans-
verse maximum BCD, and the two radial maxima AB and AD. As is easily demon-
strated, 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 dierent wave maximum curves of the form shown in Figure 11.6, correspond-
ing to many dierent 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 11.7.
This pattern, which is made up of interlocking transverse and radial wave maxima,
fills a wedge-shaped
regionknown as a Kelvin wedgewhose semi-angle takes the
value tan1 (1/ 8) = 19.47. This angle is 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. (11.83)
328 Theoretical Fluid Mechanics
6
5
4
3
2
1
y/0

0
1
2
3
4
5
6
1 0 1 2 3 4 5 6 7 8 9 10 11
x/0

Figure 11.7
A deep water wake.

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.

11.8 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), (11.84)


p(r, t) = p0 g z + p1 (r, t), (11.85)

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, (11.1)
and (11.2), reduce to

v1 = 0, (11.86)


p1
+ V v1 = , (11.87)
t
Waves in Incompressible Fluids 329

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 . (11.88)
t

The curl of Equation (11.87) implies that v1 = 0. Hence, we can write


v1 = , and Equation (11.87) yields


p1
+V = . (11.89)
t

Finally, Equation (11.86) gives


2 = 0. (11.90)
The most general traveling wave solution to Equation (11.90), with wave vector
k = k e x , and angular frequency , is

(x, z, t) = [A cosh(k z) + B sinh(k z)] cos( t k x). (11.91)

It follows from Equation (11.89) that

p1 (x, z, t) = k (V c) [A cosh(k z) + B sinh(k z)] sin( t k x), (11.92)

and from Equation (11.88) that

z (x, z, t, ) = (V c)1 [A sinh(k z) + B cosh(k z)] sin( t k x). (11.93)

Here, c = /k is the phase velocity of the wave.

11.9 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 11.8.)
Consider a gravity wave of angular frequency , and wavenumber k, propagating
through both fluids in the x-direction. Let

(x, t) = 0 sin( t k x) (11.94)

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 (11.91), with the constants
330 Theoretical Fluid Mechanics

z = d

 V

z=0

z V

x z = d

Figure 11.8
Gravity waves at an interface between two immiscible fluids.

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)]
(x, z, t) = (V c) 0 cos( t k x). (11.95)
sinh(k d)
In the upper fluid, the perturbed velocity potential must again be of the form (11.91),
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 )]
(x, z, t) = (V  c) 0 cos( t k x). (11.96)
sinh(k d )
Here, c = /k is the phase velocity of the wave. From Equations (11.85) and (11.92),
the fluid pressure just below 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). (11.97)
tanh(k d)
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). (11.98)
tanh(k d  )
In the absence of surface tension at the interface, these two pressure must equal one
another: that is,
! "z=0+
p z=0 = 0. (11.99)
Waves in Incompressible Fluids 331

Hence, we obtain the dispersion relation


k (V c)2 k  (V  c)2
(  ) g = + , (11.100)
tanh(k d) tanh(k d  )
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 (11.43) (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 . (11.101)

iii. In general, there are two values of c that satisfy the quadratic equation (11.100).
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. Because the complex roots of a quadratic equation
occur in complex conjugate pairs, one of the roots always corresponds to an
exponentially growing mode. In other words, an instability.
v. If both fluids are at rest (i.e., V = V  = 0), and of infinite depth, then the disper-
sion reduces to
g (  )
c2 = . (11.102)
k ( +  )
It follows that the configuration is only stable when >  : that is, 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 (11.100) yields
1/2  
tanh(k d) 1
c (g d)1/2 1 s [1 + tanh(k d)] . (11.103)
kd 2
Comparing this with Equation (11.43), we can see that the presence of the atmo-
sphere tends to slightly diminish the phase velocities of gravity waves propagating
over the surface of a body of water.

11.10 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),
332 Theoretical Fluid Mechanics

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.
We expect the velocity potential, perturbed pressure, and vertical displacement of
the water to be of the form (11.91), (11.92), and (11.93), respectively, with = c = 0,
because 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), (11.104)

which yields
V 1 [A sinh(k d) + B cosh(k d)] = a. (11.105)
At the free surface, we have

z (x, 0) = b sin(k x), (11.106)

which gives
b = V 1 B. (11.107)
In addition, pressure balance across the free surface yields

g b sin(k x) = p1 (x, 0) = k V A sin(k x), (11.108)

which leads to
g b = k V A. (11.109)
Hence, from Equations (11.105), (11.107), and (11.109),
a
b= , (11.110)
cosh(k d) (g/k V 2 ) sinh(k d)
or
a
b= , (11.111)
cosh(k d) (1 c 2 /V 2 )
where c = [(g/k) tanh(k d)]1/2 is the phase velocity of a gravity wave of wave number
k. [See Equation (11.21).] 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.

11.11 Surface Tension


As described in Chapter 3, there is a positive excess energy per unit area, , as-
sociated with an interface between two immiscible fluids. The quantity can also
Waves in Incompressible Fluids 333

be interpreted as a surface tension. Let us now incorporate surface tension into our
analysis. Suppose that the interface lies at

z = (x, t), (11.112)

where || is small. Thus, the unperturbed interface corresponds to the plane z = 0.


The unit normal to the interface is
(z )
n= . (11.113)
|(z )|
It follows that

nx , (11.114)
x
nz 1. (11.115)

The Young-Laplace Equation (see Section 3.2) yields

p = n, (11.116)

where p is the jump in pressure seen crossing the interface in the opposite direction
to n. However, from Equations (11.114) and (11.115), we have

2
n . (11.117)
x 2
Hence, Equation (11.116) gives

2
z=0 =
[p]z=0 .
+
(11.118)
x 2
This expression is the generalization of Equation (11.99) 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
(11.118) yields
2
g p1 |z=0 = 2 , (11.119)
x
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 Equa-
tion (11.9), p1 = (/t). Hence, the previous expression gives
  
 2  3 
g  +  =  . (11.120)
z z=0 t 2 z=0 z 2 x z=0
334 Theoretical Fluid Mechanics

This relation, which is a generalization of Equation (11.15), is the condition satis-


fied at a free surface in the presence of non-negligible surface tension. Applying
this boundary condition to the general solution, (11.19) (which already satisfies the
boundary condition at the bottom), we obtain the dispersion relation


k3
2 = g k + tanh(k d), (11.121)

which is a generalization of Equation (11.21) that takes surface tension into account.

11.12 Capillary Waves


In the deep water limit k d
1, the dispersion relation (11.121) simplifies to

k3
2 = g k + . (11.122)

It is helpful to introduce the capillary length,

1/2

l= . (11.123)
g

(See Section 3.4.) The capillary length of an air/water interface at s.t.p. is 2.7103 m
(Batchelor 2000). 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 intro-
duce the critical phase velocity

vc = (2 g l)1/2. (11.124)

This critical velocity takes the value 0.23 m/s for an air/water interface at s.t.p.
(Batchelor 2000). It follows from Equation (11.122) that the phase velocity, v p =
/k, of a surface water wave can be written

1/2
vp 1 1
= kl+ . (11.125)
vc 2 kl
Moreover, the ratio of the phase velocity to the group velocity, vg = d/dk, becomes

vg 1 1 + 3 (k l)2
= . (11.126)
v p 2 1 + (k l)2
In the long wavelength limit
c (i.e., k l  1), we obtain
vp 1
, (11.127)
v0 (2 k l)1/2
Waves in Incompressible Fluids 335

and
vg 1
. (11.128)
vp 2
We can identify this type of wave as the deep water gravity wave discussed in Sec-
tion 11.3.
In the short wavelength limit  c (i.e., k l
1), we get

1/2
vp kl
, (11.129)
vc 2
and
vg 3
. (11.130)
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,
Equation (11.129) can be rewritten

1/2
k
vp , (11.131)

which demonstrates that gravity plays no role in the propagation of a capillary wave.
In fact, its place is taken by surface tension. Finally, it is easily seen that the phase
velocity (11.125) attains the minimum value v p = vc when = c (i.e., when k l = 1).
Moreover, from Equation (11.126), 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 .

11.13 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 (11.97), (11.98), and (11.118) yield the
dispersion relation
k (V c)2  k (V  c)2
(  ) g + k 2 = + , (11.132)
tanh(k d) tanh(k d )
which is a generalization of the dispersion relation (11.100) that takes surface tension
into account. Here, c = /k is the phase velocity of a wave propagating along the
interface.
336 Theoretical Fluid Mechanics

For the case in which both fluids are at rest, and of infinite depth, the previous
dispersion relation simplifies to give

(  ) g + k 2 = ( +  ) k c 2 . (11.133)

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= (11.134)
g (1 s)

1/2
1s
c0 = 2 g l . (11.135)
1+s
It follows that

c2 1 1
= + k l . (11.136)
c02 2 k l
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. Further-
more, waves of all wavelength are able to propagate along the interface (i.e., c 2 > 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= . (11.137)
g (s 1)
The dispersion relation (11.133) then becomes



s1 1
c = gl
2
kl . (11.138)
s+1 kl

It is apparent that c 2 < 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 atmo-
spheric 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.)

11.14 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
Waves in Incompressible Fluids 337

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 (11.132) reduces to

(1 s) g + k 2 = k c 2 + s k (V  c)2 . (11.139)

This expression can be rearranged to give




2V s V 2 s
c
2
+ = c12 , (11.140)
1+s 1+s

which is a quadratic equation for the phase velocity, c, of the wave. Here,


g 1s k
c1 =
2
+ , (11.141)
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

1 c 2
c12 = + c (11.142)
2 c 0
where c = 2 l is the capillary wavelength, and l and c0 are defined in Equa-
tions (11.134) and (11.135), respectively.
For a given wavelength, , the wave velocity, c, attains its maximum value, cm ,
when dc/dV  = 0. According to the dispersion relation (11.140), this occurs when

V  = cm = (1 + s)1/2 c1 . (11.143)

If the wind has any other velocity, greater or less than cm , then the wave velocity is
less than cm .
According to Equation (11.140), the wave velocity, c, becomes complex, indicat-
ing an instability, when


2 (1 + s)2 2 (1 + s)2 c 2
V > c1 = + c . (11.144)
s 2s c 0

We conclude that if the wind speed exceeds the critical value

(1 + s)
Vc = c0 = 6.6 m/s = 12.8 kts (11.145)
s1/2
then waves whose wavelengths fall within a certain range, centered around c , are
unstable and grow to large amplitude.
The two roots of Equation (11.140) are
1/2
V s s V 2
c= c12 . (11.146)
1+s (1 + s)2
338 Theoretical Fluid Mechanics

Moreover, if
V  < (1 + s 1 )1/2 c1 (11.147)
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 previously then the waves cannot travel against the wind.
Because c1 has the minimum value c0 , it follows that waves traveling against the
wind are completely ruled out when

V  > (1 + s 1 )1/2 c0 = 6.6 m/s = 12.8 kts. (11.148)

11.15 Exercises
11.1 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.

11.2 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 (n 2 l 2 + m 2 b 2 )1/4 ,

where n and m are integers. Neglect surface tension.

11.3 Demonstrate that a sinusoidal gravity wave on deep water with surface eleva-
tion
= a cos( t k x)
possesses a mean momentum per unit surface area

1
a 2.
2

11.4 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.


Waves in Incompressible Fluids 339

11.5 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
c 2 = k g,
k g ( )
c2 = ,
coth(k d) +
can propagate along the layer.
11.6 Show that, taking surface tension into account, a sinusoidal wave of wavenum-
ber k and surface amplitude a has a mean kinetic energy per unit surface area
1
K= ( g + k 2 ) a 2 ,
4
and a mean potential energy per unit surface area
1
U= ( g + k 2 ) a 2 .
4
11.7 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.
11.8 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 11.7 to investigate the gen-
eration 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
3
y= X1 sec tan .
2
340 Theoretical Fluid Mechanics

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 x 2/3 .
12
Terrestrial Ocean Tides

12.1 Introduction
This chapter outlines the classical theory of terrestrial ocean tides, the study of which
was pioneered by Pierre-Simon Laplace (1749-1827). Our treatment of ocean tides
takes all major harmonics of the tide generating potential into account, as well as
the finite elasticity of the Earth, and the self-gravity of the oceans. As will become
apparent, the theory of terrestrial ocean tides is extremely complicated. Hence, for
the sake of simplicity and brevity, we shall only consider two relatively idealized
scenarios. First, tides in a frictionless ocean of constant depth that covers the whole
surface of the Earth, and, second, tides in a frictionless ocean of constant depth
that only covers one hemisphere of the Earth (and lies between two meridians of
longitude). It turns out that these two scenarios (in particular, the latter) exhibit most
of the observed features of terrestrial ocean tides. More information on terrestrial
tides can be found in Lamb 1993 and Cartwright 1999.

12.2 Tide Generating Potential


Consider an (almost) spherical planet of mass m, and radius a, and a spherical moon
of mass m , that execute (almost) circular orbits about their mutual center of mass.
Let r (where r
a) represent the vector position of the center of the moon, relative
to the center of the planet. It follows that (Fitzpatrick 2012)


 m
r = r (12.1)
m + m

is the vector position of the center of mass, relative to the center of the planet. More-
over, the planet and moon both orbit the center of mass with angular velocity 1 ,
where (Fitzpatrick 2012)
G (m + m )
12 = . (12.2)
r 3
Here, G is the universal gravitational constant (Yoder 1995).
Neglecting the planets axial rotation (which is introduced into our analysis in

341
342 Theoretical Fluid Mechanics

Section 12.6), the planets orbital motion about the center of mass causes its con-
stituent points to execute circular orbits of radius r with angular velocity 1 . The
centers of these orbits have the same spatial relation to the center of mass that the
constituent points have to the center of the planet. Moreover, the orbits lie in par-
allel planes. Thus, the common centripetal acceleration of the constituent points is
(Fitzpatrick 2012)
g centripetal = 12 r = centripetal . (12.3)
It follows, from the previous three equations, that
G (m + m )  G m  G m
centripetal (r) = r r = r r = r cos
r 3 r 3 r 2
G m r
=   P1 (cos ), (12.4)
r r
where r is a position vector relative to the center of the planet, and
r r
cos = . (12.5)
r r
Here,
1 dn
Pn (x) = [(x 2 1)n ], (12.6)
2 n n! dx n
for n 0, denotes a Legendre polynomial (Abramowitz and Stegun 1965). In partic-
ular,

P0 (x) = 1, (12.7)
P1 (x) = x, (12.8)
1
P2 (x) = (3 x 2 1). (12.9)
2
The gravitational force per unit mass exerted by the moon on one of the planets
constituent points, located at position vector r, is written

g moon (r) = moon , (12.10)

where (Fitzpatrick 2012)


G m G m  $ r %n
moon (r) = = Pn (cos ), (12.11)
|r r | r n=0, r

assuming that r > r. The tidal force per unit mass acting on the point (i.e., the force
per unit mass that is not compensated by the centripetal acceleration) is

g tide (r) = g moon (r) g centripetal = tide , (12.12)

where
tide (r) = moon (r) centripetal (r) (12.13)
Terrestrial Ocean Tides 343

is termed the tide generating potential. Thus, according to Equations (12.4) and
(12.11),
G m $ r %2 $ r %3
tide (r) =  1 +  P2 (cos ) + O  . (12.14)
r r r
The truncation of the expansion is ultimately justified by the fact that a/r  1.

12.3 Decomposition of Tide Generating Potential


Let r, , be right-handed spherical coordinates in a non-rotating reference frame
whose origin lies at the center of the planet, and whose symmetry axis coincides with
the planetary rotation axis. Thus, the vector position of a general point is
r = r sin cos er + r sin sin e + r cos e , (12.15)
where er = r/|r|, et cetera. Let the coordinates of the moons center be r ,  ,  .


It follows that
r = r sin  cos  er + r sin  sin  e + r cos  e . (12.16)
Hence, from Equation (12.5),
cos = cos cos  + sin sin  cos(  ). (12.17)
Now, according to the spherical harmonic addition theorem (Arfken 1985),
Pn (cos ) = Pn (cos ) Pn (cos  ) (12.18)
 (n m)!
+2 Pnm (cos ) Pnm (cos  ) cos[m (  )], (12.19)
m=1,n
(n + m)!

which implies that


1 1
P2 (cos ) = P20 (cos ) P20 (cos  ) + P (cos ) P21 (cos  ) cos(  )
3 2
1 2 ! "
+ P2 (cos ) P22 (cos  ) cos 2 (  ) . (12.20)
12
Here,
d m [Pn (x)]
Pnm (x) = (1)m (1 x 2 )m/2 , (12.21)
dx m
for n 0 and m n, denotes an associated Legendre function (Abramowitz and
Stegun 1965). In particular,
1
P20 (x) = (3 x 2 1), (12.22)
2
P21 (x) = 3 x (1 x 2 )1/2 , (12.23)
P22 (x) = 3 (1 x 2 ). (12.24)
344 Theoretical Fluid Mechanics

Note that Pn0 (x) = Pn (x).


Let
$ r %n
Rn(m) (r, , ) = Pnm (cos ) cos(m ), (12.25)
a
$ r %n
Rn(0) (r, ) = Pn0 (cos ), (12.26)
a
$ r %n
Rn(+m) (r, , ) = Pnm (cos ) sin(m ), (12.27)
a
where n m > 0. According to Equations (12.14), (12.20), and (12.25)(12.27),

G m a 2  0
tide (r, , ) = P2 (cos  ) R2(0) (r, )
r 3
1
+ P21 (cos  ) cos  R2(1) (r, , )
3
1
+ P21 (cos  ) sin  R2(+1) (r, , )
3
1 2
+ P (cos  ) cos(2  ) R2(2) (r, , )
12 2

1
+ P22 (cos  ) sin(2  ) R2(+2) (r, , ) . (12.28)
12

Here, we have neglected the unimportant constant term in Equation (12.14).

12.4 Expansion of Tide Generating Potential


Suppose that, in the r, , frame, the moons orbit is a Keplerian ellipse of major
radius R, and eccentricity 0 < e  1, lying in a fixed plane that is inclined at an
angle > 0 to the plane = /2. Thus, the closest and furthest distance between the
centers of the moon and the planet are R p = (1 e) R and Ra = (1 + e) R, respectively.
It follows that (Fitzpatrick 2013)

cos  = cos sin , (12.29)


tan
tan  = , (12.30)
cos
where

= 1 t + 2 e sin[(1 2 ) t] + O(e 2 ), (12.31)


1 2
r = R 1 e cos[(1 2 ) t] + O(e 2 ) . (12.32)
Terrestrial Ocean Tides 345

Here, it is assumed that the closest point on the moons orbit to the center of the planet
corresponds to = 2 t, where 2 is the uniform precession rate of this point. [It is
necessary to include such precession in our analysis because the Moons perigee pre-
cesses steadily in such a manner that it completes an orbit about the Earth once every
8.85 years. This eect is caused by the perturbing influence of the Sun (Fitzpatrick
2013).] Suppose that the inclination of the moons orbit to the planets equatorial
plane, = /2, is relatively small, so that sin  1. It follows that

cos  = sin cos(1 t) + O(e sin ), (12.33)


 = 1 t + 2 e sin[(1 2 ) t] + O(e 2 ) + O(sin2 ). (12.34)

Thus, Equations (12.22)(12.24), (12.28), and (12.32)(12.34) can be combined to


give the following expression for the tide generating potential due to a moon in a
low-eccentricity, low-inclination orbit:


G m a 2 1 3
tide (r, , ) = 1 + 3 e cos [(1 2 ) t] sin cos[2 1 t] R2(0) (r, )
2
R3 2 2
1
+ sin R2(1) (r, , )
2
1
+ sin R2(1) (r, , 2 1 t)
2
7 e (2)
R (r, , [(3/2) 1 (1/2) 2] t)
8 2
1
R2(2) (r, , 1 t)
4
e
+ R2(2) (r, , [(1/2) 1 + (1/2) 2] t)
8

+O(e 2 ) + O(e sin ) + O(sin2 ) . (12.35)

Here, we have retained a term proportional to sin2 in the previous expression, de-
spite the fact that we are formally neglecting O(sin2 ) terms, because the term in
question gives rise to important fortnightly tides on the Earth.

12.5 Surface Harmonics and Solid Harmonics


A surface harmonic of degree n (where n is a non-negative integer), denoted Sn (, ),
is defined as a well-behaved solution to

r 2 2 Sn + n (n + 1) Sn = 0 (12.36)
346 Theoretical Fluid Mechanics

on the surface of a sphere (i.e., r = constant). Here, r, , are standard spherical


coordinates, and 2 is the Laplacian operator. It follows that (Love 1927)
 ! "
Sn (, ) = cnm Pnm (cos ) cos(m ) + dnm Pnm (cos ) sin(m ) , (12.37)
m=0,n

where the cnm


and dnm are arbitrary coecients, and the Pnm (x) associated Legendre
functions.
A solid harmonic of degree n (where n is a non-negative integer), denoted Rn (r, , ),
is defined as a well-behaved solution to
2 Rn = 0 (12.38)
in the interior of a sphere (i.e., the region r < constant). It follows that (Love 1927)
Rn (r, , ) r n Sn (, ). (12.39)
In particular, the functions Rn(m) (r, , ) and Rn(0) (r, ), introduced in Section 12.3,
are solid harmonics of degree n. Note that the Cartesian coordinates xi (where i runs
from 1 to 3) are solid harmonics of degree 1. Moreover, Rn /xi is a solid harmonic
of degree n 1. Finally, tide (r), specified in Equation (12.35), is a solid harmonic
of degree 2.
The following results regarding solid harmonics are useful (Love 1927):
Rn Rn
xi =r = n Rn , (12.40)
xi r
Rn
2 (xi Rn ) = (Rn xi + xi Rn ) = 2 xi Rn = 2 , (12.41)
xi

1 d 2 d m+n
2 (r m Rn ) = 2 (r m+n Sn ) = r (r Sn n (n + 1) r
) m+n2
Sn
r 2 dr dr
= m (m + 2 n + 1) r m2 Rn . (12.42)
Here, use has been made of the Einstein summation convention (Riley 1974).

12.6 Planetary Rotation


Suppose that the planet is rotating rigidly about the axis = 0 at the angular velocity
(where
1 ). The planets rotational angular velocity vector is thus
= cos er sin e . (12.43)
Let
= t. (12.44)
It follows that r, , are spherical coordinates in a frame of reference that co-rotates
with the planet.
Terrestrial Ocean Tides 347

12.7 Total Gravitational Potential


The planet is modeled as a solid body of uniform mass density whose surface lies
at
r = a + b (, ), (12.45)
where 
b (, ) = b n (, ), (12.46)
n=1,

and b n is a surface harmonic of degree n. Suppose that the planetary surface is


covered by an ocean of uniform mass density whose surface lies at

r = a + d + t (, ), (12.47)

where 
t (, ) = t n (, ), (12.48)
n=1,

and t n is a surface harmonic of degree n. Here, d is the constant unperturbed depth


of the ocean, whereas b and t are the radial displacements of the oceans bottom
and top surfaces, respectively, generated by planetary rotation and tidal eects. It is
assumed that |b |, |t |  d  a.
The net gravitational acceleration in the vicinity of the planet takes the form

g(r) = , (12.49)

where (r) is the total gravitational potential. In the limit d/a  1, we can write

(r, , ) = 0 (r) + n (r, , ) + tide (r, , ), (12.50)
n=1,

where 
4 G + 4 G d (r a) ra
0 =
2
, (12.51)
0 r>a
and
2 n>0 = 4 G ( b n + n ) (r a). (12.52)
Here,
(, ) = t (, ) b (, ), (12.53)
is the net change in ocean depth due to planetary rotation and tidal eects, and n =
t n b n . Moreover, (x) is a Dirac delta function (Riley 1974). The boundary
conditions are that the n be well behaved as r 0, and

n (r) 0 (12.54)
348 Theoretical Fluid Mechanics

as r . It follows that
g
0 (r a) = (3 a 2 r 2 ) 3 g d , (12.55)
2a


g a2 d
0 (r > a) = 1+3 , (12.56)
r a

and


$ %n
3 r
n>0 (r a, , ) = g b n + n , (12.57)
2n+1 a


$ %(n+1)
3 r
n>0 (r > a, , ) = g b n + n , (12.58)
2n+1 a

where
G m 4
g= = Ga (12.59)
a2 3
is the mean gravitational acceleration at the planets surface. Note that, inside the
planet (i.e., r a), n (r) is a solid harmonic of degree n. We can identify the
three terms appearing on the right-hand side of Equation (12.50) as the equilibrium
gravitational potential generated by the planet and the ocean, the potential generated
by non-spherically-symmetric radial displacements of the planet and ocean surfaces,
and the tide generating potential, respectively.

12.8 Planetary Response


The interior of the planet is modeled as a uniform, incompressible, elastic, solid
possessing the stress-strain relation (Love 1927)


i j
i j = p i j + + , (12.60)
x j xi

and subject to the incompressibility constraint

= 0. (12.61)

Here, i j (r) is the stress tensor, i j the identity tensor, (r) the elastic displacement,
p(r) the pressure, and the (uniform) rigidity of the material making up the planet
(Riley 1974).
The planets equation of elastic motion in the co-rotating frame is (Fitzpatrick
2012)
2 i k 1 i j
+ 2 i jk j = , (12.62)
t 2 t x j xi xi
Terrestrial Ocean Tides 349

where i jk is the totally antisymmetric tensor (Riley 1974), and (Fitzpatrick 2012)
1
(r, ) = 2 r 2 sin2 = 0 (r) + 2 (r, ), (12.63)
2
2 a 2 $ r %2
0 (r) = , (12.64)
3 a
2 a 2 $ r %2 0
2 (r, ) = P2 (cos ). (12.65)
3 a
Note that 2 (r) is a solid harmonic of degree 2. The second term on the left-hand side
of Equation (12.62) is the Coriolis acceleration (due to planetary rotation), whereas
the final term on the right-hand side is the centrifugal acceleration (likewise, due to
planetary rotation). The first two terms on the right-hand side are the forces per unit
mass due to internal stresses and gravity, respectively.
Equations (12.60), (12.61), and (12.62) can be combined to give


2 2 p
+2 = ++ . (12.66)
t 2 t
Let us assume that 
/
 : (12.67)
a
that is, the typical timescale on which the tide generating potential varies (in the co-
rotating frame) is much longer than the transit time of an elastic shear wave through
the interior of the planet. In this case, we can neglect the left-hand side of Equa-
tion (12.66), and write

2 p
+ + 0, (12.68)

which is equivalent to saying that the interior of the planet always remains in an
equilibrium state.
Let
 ! "
p(r, , ) = p0 (r) + pn (r, , ) n (r, , ) ext (r, , ), (12.69)
n=1,

where
ext (r, , ) = tide (r, , ) + 2 (r, ), (12.70)
is the sum of the tide generating potential and the fictitious centrifugal potential due
to planetary rotation. Note that ext is a solid harmonic of degree 2. It follows from
Equations (12.50), (12.63), and (12.68) that


d p0
+ 0 + 0 = 0, (12.71)
dr
and 
2 pn = 0. (12.72)
n=1,
350 Theoretical Fluid Mechanics

Taking the divergence of the previous equation, and making use of Equations (12.61),
we find that 2 pn = 0, which implies that pn (r) is a solid harmonic (of degree n).
Equation (12.71) can be integrated to give


2
p0 (r) g 2 a (a 2 r 2 ) + g d + patm , (12.73)
2a 3
where use has been made of Equations (12.55) and (12.64), as well as

p0 (a) = g d + patm . (12.74)

Here, patm is the atmospheric pressure at sea level. The previous boundary condition
ensures that the mean pressure at the surface of the planet is able to support the
mean weight of the ocean, as well as the weight of the atmosphere. Incidentally, it is
assumed that g
2 a, and we have neglected terms of order 2 a d with respect
to terms of order g d.
The elastic displacement in the interior of the planet satisfies Equations (12.61)
and (12.72). It is helpful to define the radial component of the displacement
xi i
r = , (12.75)
r
as well as the stress acting (outward) across a constant r surface,

x j i j xi i (r r )
Xi = =p r i + , (12.76)
r r r r xi
where use has been made of Equation (12.60). The radial displacement at r = a is
equivalent to the displacement of the planets surface. Hence, according to Equa-
tion (12.45),
r (a, , ) = b (, ). (12.77)
The stress at any point on the surface r = a must be entirely radial (because a fluid
ocean cannot withstand a tangential stress), and such as to balance the weight of the
column of rock and ocean directly above the point in question. In other words,
! " $ xi %
Xi (a, , ) = g b + (d + ) . (12.78)
r r=a
It follows from Equation (12.57), (12.69), (12.70), (12.74), (12.76), and (12.78) that

 xi i (r r )
pn r i +
r r r xi
n=1, r=a


2 n 2


2 n 2
$ %
xi
= g 2 n 2 +
b n + n , (12.79)
n=1,
2n+1 2n+1 r r=a

where
2 (, ) = g 1 ext |r=a = g 1 (tide + 2 )r=a (12.80)
Terrestrial Ocean Tides 351

is a surface harmonic of degree 2. Here, 2 (, ), which has the dimensions of length,


parameterizes the perturbation due to tidal and rotational eects at the planets sur-
face. Moreover, i j is a Kronecker delta symbol (Riley 1974). Hence, we need to
solve Equations (12.61) and (12.72), subject to the boundary conditions (12.77) and
(12.79).
Let us try a solution to Equations (12.61) and (12.72) of the form

i = i n , (12.81)
n=1,

where (Love 1927)


pn n
i n = An r 2 + B n p n xi + . (12.82)
xi xi
Here, An and Bn are spatial constants, and n (r) is an arbitrary solid harmonic of
degree n. It follows that

r r n xi i n = (n An + Bn ) r 2 pn + n n , (12.83)

where use has been made of Equation (12.40). Moreover,

(r r n ) pn n
= (n An + Bn ) r 2 + 2 (n An + Bn ) pn xi + n , (12.84)
xi xi xi

and
i n pn n
r i n = n An r 2 + n Bn pn xi + (n 2) . (12.85)
r xi xi
Thus, the boundary conditions (12.77) and (12.79) become
 
(2 An + Bn ) r 2 pn + n n = a b n , (12.86)
r=a

and

pn n
(1 [2 n An + (2 + n) Bn] ) (pn xi )r=a (2 n An + Bn ) r + 2 (n 1)
2
xi xi r=a



2n2 2 n 2
= g 2 n 2 + b n + n (xi )r=a , (12.87)
2n+1 2n+1

respectively. Suppose that n (r) is such that

(2 n An + Bn ) a 2 pn (r) + 2 (n 1) n(r) = 0. (12.88)

In this case, the boundary conditions (12.86) and (12.87) reduce to



4 An + (n 2) Bn
a pn |r=a = b n , (12.89)
2 (n 1)
352 Theoretical Fluid Mechanics

and

(1 [2 n An + (2 + n) Bn] ) pn |r=a =



2n2 2 n 2
g 2 n 2 + b n + n , (12.90)
2n+1 2n+ 1

respectively.
The expression for i n given in Equation (12.82) satisfies Equations (12.61) and
(12.72) provided that

2 n An + (3 + n) Bn = 0, (12.91)
2 (2 n + 1) An + 2 Bn = 1, (12.92)

respectively, where use has been made of Equations (12.40)(12.42). It follows that

(3 + n)
An = , (12.93)
2 (2 n + 3) (n + 1)
2n
Bn = . (12.94)
2 (2 n + 3) (n + 1)

Hence, the boundary conditions (12.89) and (12.90) yield

b n = h2 2 n 2 + hn n n , (12.95)

where


3
n = , (12.96)
2n+1

)
2n+1 (2 n + 1) (2 n 2 + 4 n + 3)
hn = 1+ , (12.97)
2n2 (6 + n 2 ) ga

)
 2n+ 1 (2 n + 1) (2 n 2 + 4 n + 3)
hn = 1+ . (12.98)
3 (6 + n 2 ) ga

Here, the n parameterize the self-gravity of the ocean. Let



= ext + n . (12.99)
n=1,

This quantity can be recognized as the total perturbing potential at the planets sur-
face due to the combination of tidal, rotational, and ocean self-gravity, eects. It
follows from Equation (12.57), (12.80), and (12.95) that

  

= (1 + k2 ) 2 n 2 + (1 + kn ) n n , (12.100)
g r=a n=1,
Terrestrial Ocean Tides 353

where

)
3 (2 n + 1) (2 n 2 + 4 n + 3)
kn = 1+ , (12.101)
2n2 (6 + n 2 ) ga
1
(2 n + 1) (2 n 2 + 4 n + 3)
kn = 1 + . (12.102)
(6 + n 2 ) ga

Here, the dimensionless quantities hn , hn , kn , and kn are known as Love numbers
(Love 1911).

12.9 Laplace Tidal Equations


In the co-rotating frame, the linearized (in v) equations of motion of the ocean that
covers the surface of the planet are (Fitzpatrick 2012)

v = 0, (12.103)


v p
+ 2 v = ++ , (12.104)
t

where v(r) is the fluid velocity, and p(r) the fluid pressure. Here, (r) is the total
gravitational potential, (r) the centrifugal potential (due to planetary rotation), and
2 v the Coriolis acceleration (likewise, due to planetary rotation). Furthermore,
we have neglected the finite compressibility of the fluid that makes up the ocean. Let
r, , be spherical coordinates in the co-rotating frame, and let z = r a. It is helpful
to define u = v , v = v , and w = vr . Assuming that |z|  a, the previous equations of
motion reduce to
1 1 v w
(sin u) + + 0, (12.105)
a sin a sin z

and


u 1 p
2 cos v = ++ , (12.106)
t a


v 1 p
+ 2 cos u + 2 sin w = ++ , (12.107)
t a sin


w p
2 sin v = ++ . (12.108)
t z

In accordance with Equations (12.45), (12.47), and (12.53), the lower and upper
surfaces of the ocean lie at
z = b (, , t), (12.109)
354 Theoretical Fluid Mechanics

and
z = d + b (, , t) + (, , t), (12.110)
respectively, where
|b |, ||  d  a. (12.111)
It follows that
b
w|z=0 = , (12.112)
t
(b + )
w|z=d = . (12.113)
t
Let us assume that u = u(, , t) and v = v(, , t): that is, the horizontal components
of the fluid velocity are independent of z. Integration of Equation (12.105) from z = 0
to z = d, making use of the previous two boundary conditions, yields

d v
= (sin u) + . (12.114)
t a sin
Equations (12.112)(12.114) imply that



d d
w u, v  u, v. (12.115)
a a
In accordance with the analysis of Section 12.8, we can write


2 2
+ = g a (z d) + (, , t), (12.116)
3
where is the total perturbing potential due to tidal, rotation, and ocean self-
gravity, eects. We have neglected an unimportant constant term. We have also
neglected the z-dependence of the perturbing potential, because the variation length-
scale of this potential is a, rather than d, so that
d
() d  , (12.117)
z a
where () = |z=d |z=0 . Let us assume that the ocean is in approximate
vertical force balance: that is,


p
+ + 0. (12.118)
z
It follows that
p patm
+ + = P(, , t), (12.119)

where patm is the (uniform and constant) pressure of the atmosphere. However, pres-
sure balance at the oceans upper surface requires that

p|z=d+b + = patm . (12.120)


Terrestrial Ocean Tides 355

Hence, we deduce that

P(, , t) = ( + )z=d+b + g (b + ) + , (12.121)

to first order in small quantities [i.e., b /a, /a, /(g a), and 2 a/g]. Thus, Equa-
tions (12.106), (12.107), and (12.115) yield

u 1 ! "
2 cos v = g (b + ) + , (12.122)
t a
v 1 ! "
+ 2 cos u = g (b + ) + . (12.123)
t a sin

Let us justify our previous assumption that the ocean is in approximate vertical
force balance. Equations (12.108), (12.115), and (12.119) imply that

p patm
+ + = P(, , t) + O(d v), (12.124)

where we have assumed that /t O() and /z d 1 . However, Equa-


tions (12.121)(12.123) yield

P g a v
d v. (12.125)

Hence, Equation (12.124) becomes

p patm
+ + = P(, , t), (12.126)

which is equivalent to Equation (12.119).


Let us justify our previous assumption that u only depends weakly on z. It follows
from Equations (12.117) and (12.122) that


u , (12.127)
a
() d d
u u  u, (12.128)
a a a a

where u = u|z=d u|z=0 . In a similar manner, it can be shown that v  v.


According to Equations (12.95) and (12.100),
  
b = h2 2 n 2 + hn n n , (12.129)
n=1,
  
g 1 = (1 + k2 ) 2 n 2 + (1 + kn ) n n . (12.130)
n=1,
356 Theoretical Fluid Mechanics

Hence, Equations (12.114), (12.122), and (12.123) give



d v
= (sin u) + , (12.131)
t a sin

u g 
2 cos v = (1 + k2 h2 ) 2
(1 + kn hn ) n n ,
 
t a n=1,
(12.132)

v g 
+ 2 cos u = (1 + k2 h2 ) 2 (1 + kn hn ) n n .
 
t a sin n=1,
(12.133)

We can write 
(, ) = n (, ), (12.134)
n=1,

where [cf., Equation (12.37)]



n (, ) = cnm Pnm (cos ) e i m , (12.135)
m=0,n

and the cnm are arbitrary complex constants. Now, (Abramowitz and Stegun 1965)


2 (n + m)!
Pnm (cos ) Pnm (cos ) sin d = n n . (12.136)
0 2 n + 1 (n m)!
Hence,
  
(1 + kn hn ) n n = G F (, ;  ,  ) ( ,  ) sin  d d , (12.137)
n=1, 0

where
  (2 n + 1) (n m)!
G F (, ;  ,  ) = (1 + kn hn )
n=1, m=0,n
4 (n + m)!

n Pnm (cos ) Pnm (cos  ) e i m ( ) . (12.138)

is termed the Farrell Greens function (Farrell 1972). It follows that



d v
= (sin u) + , (12.139)
t a sin

u g
2 cos v = (1 + k2 h2 ) 2 G F d , (12.140)
t a

v g
+ 2 cos u = (1 + k2 h2 ) 2 G F d , (12.141)
t a sin
Terrestrial Ocean Tides 357

j j mj j Classification
0 (1 + 2 ) 0 0 Equilibrium
1 3 e 2 0 1 2 Long Period
2 +(3/2) sin2 2 0 2 1 Long Period
3 sin 2 1 2 1 Diurnal
4 sin 2 1 Diurnal
5 +(7 e/4) 2 2 2 3 1 + 2 Semi-diurnal
6 +(1/2) 2 2 2 2 1 Semi-diurnal

Table 12.1
The principal harmonics of the forcing term in the Laplace tidal equations.

where d sin  d d . Equations (12.139)(12.141) are known collectively as


the Laplace tidal equations, because they were first derived (in simplified form) by
Laplace (Lamb 1993).
The Laplace tidal equations are a closed set of equations for the perturbed ocean
depth, (, , t), and the polar and azimuthal components of the horizontal ocean
velocity, u(, , t), and v(, , t), respectively. Here, a is the planetary radius, g the
mean gravitational acceleration at the planetary surface, d the mean ocean depth,
the planetary angular rotation velocity, and hn , hn , kn , kn are Love numbers defined
in Equations (12.97), (12.98), (12.101), and (12.102), respectively. Amongst other
things, the Love numbers determine the elastic response of the planet to the forcing
term, 2 . Finally, the Farrell Greens function parameterizes the self-gravity of the
ocean.

12.10 Harmonics of Forcing Term in Laplace Tidal Equations


Making use of Equations (12.25)(12.27), (12.35), (12.44), (12.65), and (12.80), we
can write the forcing term in the Laplace tidal equations in the form


2 (, , t) = Re j (, , t), (12.142)
j

where

j (, , t) = j P2 j (cos ) e i (m j + j t) .
m
(12.143)
358 Theoretical Fluid Mechanics

j j mj j Classification Symbol
0 ( + M + E ) 0 0 Equilibrium S 0 , M0
1 +(3/2) sin2  E 0 2 E Long Period S sa
2 3 e M M 0 M M Long Period Mm
3 +(3/2) sin2  M 0 2 M Long Period Mf
4 sin  M 1 2 M Diurnal O1
5 sin  E 1 2 E Diurnal P1
6 sin  ( M + E ) 1 Diurnal K1
7 +(7 e M /4) M 2 2 3 M + M Semi-diurnal N2
8 +(1/2) M 2 2 2 M Semi-diurnal M2
9 +(1/2) E 2 2 2 E Semi-diurnal S2

Table 12.2
The principal harmonics of the forcing term in the Laplace tidal equations for the
Earth.

The amplitudes, j , azimuthal mode numbers, m j , and frequencies, j , of the princi-


pal harmonics of the forcing term are specified in Table 12.1, where

2 a2
1 = , (12.144)
3g
1 m a 4
2 = . (12.145)
2 m R3
Here, m is the planetary mass, m the mass of the moon, and R the moons orbital
major radius. As indicated in the table, the harmonics of the forcing term can be
divided into four classes: an equilibrium term that is time independent; two long
period terms that oscillate with periods much longer than the planets diurnal (i.e.,
rotational) period; two diurnal terms that oscillate at periods close to the planets
diurnal period; and two semi-diurnal terms that oscillate at periods close to half
the planets diurnal period. Here, we have neglected a semi-diurnal term that is
significantly smaller than the other two.
For the Earth-Moon system, m = 5.97 1024 kg, m = 7.35 1022 kg, a =
6.37 106 m, R = 3.84 108 m, = 7.29 105 rad. s1 (Yoder 1995). It follows
that g = 9.82 m s2 , as well as 1 = , 2 = M , where

= 7.32 103 m, (12.146)


M = 1.79 101 m. (12.147)

Moreover, = , e = e M , 1 = M , and 2 = M , where  = 23.44, e M = 0.05488,


360 / M = 27.322 (solar) days, 360/ M = 8.847 years (Yoder 1995). Here, m,
a, , and  are the Earths mass, mean radius, (sidereal) rotational angular velocity,
and inclination of the equatorial plane to the ecliptic, respectively. Furthermore, m ,
Terrestrial Ocean Tides 359

j j (m) m j f j = j /2 Period Symbol


2 3
1 +1.93 10 0 2.73 10 182.62 days S sa
2 2.95 102 0 1.81 102 27.555 days Mm
3 +4.25 102 0 3.65 102 13.661 days M f
4 7.12 102 1 4.63 101 25.819 hours O1
5 3.24 102 1 4.97 101 24.066 hours P1
6 10.36 102 1 5.00 101 23.934 hours K1
7 +1.72 102 2 9.45 101 12.658 hours N2
8 +8.95 102 2 9.63 101 12.421 hours M2
9 +4.07 102 2 9.97 101 12.000 hours S 2

Table 12.3
The principal harmonics of the forcing term in the Laplace tidal equations for the
Earth, excluding the equilibrium harmonic.

R, e M , M , and M are the Moons mass, orbital major radius, orbital eccentricity,
mean orbital angular velocity, and rate of perigee precession, respectively. Here, we
have neglected the small (i.e., about 5 ) inclination of the Moons orbital plane to the
ecliptic.
For the Earth-Sun system, m = 1.99 1030 kg and R = 1.50 1011 m (Yoder
1995). It follows that 1 = and = E , where

E = 8.14 102 m. (12.148)

Moreover, = , e = eE , 1 = E , and 2 = E , where eE = 0.0167, 360/1 =


365.24 days, and 360 /E = 20940 years (Yoder 1995). Here, m is the solar mass.
Furthermore, R, eE , E , and E are the Earths orbital major radius, orbital eccen-
tricity, mean orbital angular velocity, and rate of perihelion precession, respectively.
Combining lunar and solar eects, we can write the forcing term in the Laplace
tidal equations for the Earth in the form (12.142)(12.143). The properties of the
principal harmonics of the forcing term are specified in Tables 12.2 and 12.3. The
symbols are due to G.H. Darwin (1845-1912).

12.11 Response to Equilibrium Harmonic


The j = 0 harmonic of the forcing term, which is associated with the Earths ax-
ial rotation, is special, because the associated oscillation frequency is zero. In this
case, Equations (12.139)(12.141) yield u = v = 0. Hence, it follows from Equa-
360 Theoretical Fluid Mechanics

tions (12.129), (12.142) and (12.143), as well as Table 12.2, that



h2 (1 + k2 h2 ) 2
b () = h2 + ( + M + E ) P2 (cos ), (12.149)
1 (1 + k2 h2 ) 2

1 + k2 h 2
() = ( + M + E ) P2 (cos ). (12.150)
1 (1 + k2 h2 ) 2
For the Earth-Moon-Sun system, + M + E = 7.32 103 m. Given the relatively
large size of + M + E , we expect the steady-state response to the equilibrium
harmonic to be fluid-like (otherwise, the elastic stress within the Earth would exceed
the yield stress) (Fitzpatrick 2012). In other words, /( g a) = 0 for the j = 0
harmonic, which implies from Equations (12.97), (12.98), (12.101), and (12.102)
that h2 = 5/2, h2 = 5/3, k2 = 3/2, and k2 = 1. Thus, it follows from the previous
two equations that
5
b () = ( + M + E ) P2 (cos ) = 18.3 P2(cos ) km, (12.151)
2
() = 0. (12.152)
We deduce that the Earths rotation causes a planetary equatorial (i.e., = /2) bulge
of about 9.2 km, and a polar (i.e., = 0) flattening of 18.3 km, but does not give rise
to any spatial variation in ocean depth. The observed equatorial bulge and polar flat-
tening of the Earth are 7.1 km and 14.3 km, respectively (Yoder 1995). Our estimates
for these values are too large because, for the sake of simplicity, we are treating the
Earth as a uniform body. In reality, the Earth possesses a mass distribution that is
strongly concentrated in its core.

12.12 Global Ocean Tides


Suppose that the surface of the Earth is completely covered by an ocean of uniform
mean depth d. The motion of this ocean under the action of the tide generating
potential is governed by the Laplace tidal equations, (12.139)(12.141), which can
be written

1
(sin ) + + = 0, (12.153)
sin
.. . g d ( F)
2 cos = , (12.154)
a2
.. . g d ( F )
+ 2 cos = 2 , (12.155)
a sin
where

F (, , t) = (1 + k2 h2 ) 2 (, , t) + G F (, ;  ,  ) ( ,  , t) d , (12.156)
Terrestrial Ocean Tides 361

and
a
u= , (12.157)
d t
a
v= . (12.158)
d t
.
Here, /t.
It is convenient to write (Love 1913)
1
= , (12.159)
sin
1
= + . (12.160)
sin
It follows from Equation (12.153) that

D = , (12.161)

where

1 1 2
D sin + . (12.162)
sin sin 2
Forming (sin )1 (12.154)/ (sin )1 [sin (12.155)/], we obtain



. .
D +2 + 2 cos D sin = 0. (12.163)
t

Similarly, forming (sin )1 (/) [sin (12.154)] (sin )1 (12.155)/, we get





. . gd
D+2 + 2 cos D sin = 2 D ( F ). (12.164)
t a
Let = cos . Consider the response, j (, , t), of the ocean to a particular
harmonic of the forcing term that takes the form [see Equation (12.143)]

2 (, , t) = j P2 j () e i (m j + j t) ,
m
(12.165)

where j is a constant. We can write


 zn
Pn j () e i (m j + j t) ,
m
(, , t) = (12.166)
n=1,
n (n + 1)

n Pn j () e i (m j + j t) ,
m
(, , t) = i (12.167)
n=1,

where the Pnm (x) are associated Legendre functions, and the zn and n are constants.
Now, by definition (Abramowitz and Stegun 1965),

D Pnm () e i m = n (n + 1) Pnm() e i m . (12.168)


362 Theoretical Fluid Mechanics

Hence, Equations (12.161) and (12.166) give



zn Pn j () e i (m j + j t) .
m
j (, , t) = (12.169)
n=1,

Equations (12.136) and (12.138) imply that


 
G F (, ;  ,  ) j ( ,  , t) d = (1 + kn hn ) n zn Pn j () e i (m j + j t) .
m

n=1,
(12.170)
It follows from Equations (12.163) and (12.164) that
    m
(1 2 ) d

m

n f j n (n + 1) + m j Pn j () + zn Pn j () = 0,
n=1,
n (n + 1) d
(12.171)
and
  n (n + 1) fj mj

  m
zn (1 [1 + kn hn ] n ) + f j Pn j ()
2

n=1,
n (n + 1)

d m 6 m
+ f j n n (n + 1) (1 2 ) Pn j () = (1 + k2 h2 ) j P2 j (), (12.172)
d
where (Lamb 1993)
j
fj = , (12.173)
2
4 a2 2
= . (12.174)
gd
However, as is well known (Abramowitz and Stegun 1965),

$ n+m %
nm+1 m
Pn () =
m
Pn+1 () + P m (), (12.175)
2n+1 2 n + 1 n1

2 dPn
m
n (n m + 1) m (n + 1) (n + m) m
(1 ) = Pn+1 () + Pn1 (). (12.176)
d 2n+1 2n+1
Equations (12.171), (12.172), (12.175), and (12.176) can be combined to give
 


n + 1 n mj n n + mj + 1
m j n (n + 1) f j n + zn1 + zn+1 = 0, (12.177)
n 2n 1 n+ 1 2n+3
and

n (n + 1)   mj fj
(1 [1 + kn hn ] n ) + f j zn
2
n (n + 1)

(n 1) (n + 1) (n m j )
+ fj n1
2n1

n (n + 2) (n + m j + 1) 6
+ fj n+1 = (1 + k2 h2 ) j n 2 .
2n+3
(12.178)
Terrestrial Ocean Tides 363

Finally, Equations (12.177) and (12.178) yield the tridiagonal matrix equation (Love
1913)
6
a(n) zn2 + b(n) zn + c(n) zn+2 = (1 + k2 h2 ) j n 2 (12.179)

for n = 2, 4, 6, , where
f j n (n + 1) (n m j 1) (n m j )
a(n) = , (12.180)
(2 n 3) (2 n 1) [m j n (n 1) f j ]
n (n + 1) 3 4 mj fj
b(n) = 1 [1 + kn hn ] n + f j2
n (n + 1)
f j (n 1)2 (n + 1) (n m j ) (n + m j )

n (2 n 1) (2 n + 1) [m j n (n 1) f j ]
f j (n + 2)2 n (n m j + 1) (n + m j + 1)
, (12.181)
(n + 1) (2 n + 1) (2 n + 3) [m j (n + 1) (n + 2) f j ]
f j n (n + 1) (n + m j + 1) (n + m j + 2)
c(n) = . (12.182)
(2 n + 3) (2 n + 5) [m j (n + 1) (n + 2) f j ]

Note that z0 = 0. The solution of a tridiagonal matrix equation is, of course, a


relatively trivial numerical exercise (Press, et al. 2007).
We can most conveniently characterize the jth tidal harmonic in terms of the
function j (), which is defined such that

j (, , t) = j () cos(m j + j t). (12.183)

Here, = /2 = sin1 () represents planetary latitude. It follows that


 m
j () = zn Pn j (sin ). (12.184)
n=2,

The j , m j , and f j values for the various harmonics of the forcing term in the Laplace
tidal equations for the Earth are specified in Table 12.3. Note that, on the longitudinal
meridian = 0, the Moon and Sun culminate simultaneously at t = 0. Moreover, at
this time, the Sun and Moon are both passing though the summer solstice, and the
Moon is passing though its perigee. The mean density of the Earth is such that / =
5.5 (Yoder 1995). Moreover, the mean depth of the Earths oceans is d = 3.8 km
(Yoder 1995), which implies that = 23.1. Finally, the Earth responds elastically to
the relatively low-amplitude lunar and solar tidal gravitational fields in such a manner
that /( g a) 0.35 (Fitzpatrick 2012).
Figures 12.1, 12.2, and 12.3 show the long-period, diurnal, and semi-diurnal
solutions of the Laplace tidal equations together with the corresponding equilibrium
tides (i.e., the tides calculated in the absence of ocean inertia),

1 + k2 h 2 m
j () = j P2 j (sin ). (12.185)
1 (1 + k2 h2 ) 2
364 Theoretical Fluid Mechanics

0.01 0.04

0.03
0.005
0.02

0.01
j 0 j
0

0.005 0.01

0.02
0.01
0.03
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
sin() sin()

Figure 12.1
Long-period solutions to the Laplace tidal equations (left panel) together with the
corresponding equilibrium tides (right panel). The solid, dashed, and dash-dotted
lines correspond to j = 1, j = 2, and j = 3, respectively. Here, j is measured in
meters.

It can be seen from Figure 12.1 that the long-period solutions are direct (i.e.,
they have the same signs as the corresponding equilibrium tides). However, the am-
plitudes of these solutions are about 4 times lower than those of the equilibrium tides.
The j = 1 (S sa ) and j = 3 (M f ) solutions cause a slight increase in the equatorial
bulge of the Earths oceans when the Sun and Moon, respectively, pass through the
equinoxes, and a slight reduction when they pass through the solstices. The j = 2
(Mm ) solution causes a slight increase in the equatorial bulge of the Earths oceans
when the Moon passes through its perigee, and a slight reduction when it passes
through its apogee.
It can be seen from Figure 12.2 that the diurnal solutions are inverted (i.e., they
have the opposite signs to the corresponding equilibrium tides). The amplitude of
the j = 4 solution is about 6 times lower than that of the j = 4 equilibrium tide.
The amplitudes of the j = 5 and j = 6 solutions are negligibly small. (In fact, the
amplitude of the j = 6 solution is identically zero.) The j = 4 (O1 ) solution causes
the Earths oceans to displace slightly toward the Moon when it passes through the
equinoxes, and away from the Moon when it passes through the solstices.
Finally, it can be seen from Figure 12.3 that the semi-diurnal solutions are in-
verted in equatorial regions, and direct in polar regions. Moreover, the amplitudes of
these solutions are similar to those of the corresponding equilibrium tides. The dom-
inant j = 8 (M2 ) solution generates a high tide at a particular point on the Earths
surface every 12.421 hours. The j = 9 (S 2 ) solution causes an enhancement of the
high tidea so-called spring tideat the new and full moons (i.e., every half a syn-
odic month, or 14.77 days). Finally, the j = 7 (N2 ) solution causes an enhancement
Terrestrial Ocean Tides 365

0
0.12

0.005 0.09

j j
0.06
0.01

0.03

0.015
0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
sin() sin()

Figure 12.2
Diurnal solutions to the Laplace tidal equations (left panel) together with the corre-
sponding equilibrium tides (right panel). The solid, dashed, and dash-dotted lines
correspond to j = 4, j = 5, and j = 6, respectively. Here, j is measured in meters.

0.1
0.05
0.2
0
0.05
0.1
j j
0.15
0.1
0.2
0.25
0.3
0.35 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
sin() sin()

Figure 12.3
Semi-diurnal solutions to the Laplace tidal equations (left panel) together with the
corresponding equilibrium tides (right panel). The solid, dashed, and dash-dotted
lines correspond to j = 7, j = 8, and j = 9, respectively. Here, j is measured in
meters.
366 Theoretical Fluid Mechanics

of the spring tidea so-called perigean spring tidewhenever the new or full moon
coincides with the passage of the moon through its perigee.

The previously described global solutions to the Laplace tidal equations were
first obtained (in simplified form) by Laplace (Lamb 1993). Let us consider how
these solutions compare with the observed terrestrial tides. One obvious prediction
of the global solutions is that the time variation in ocean depth at a given point on
the Earths surface should be expressible as a superposition of terms that oscillate at
constant amplitude (but not necessarily in phase with one another) with the periods
specified in Table 12.3. This is indeed found to be the case (Cartwright 1999). An-
other prediction is that the amplitudes of the long-period solutions should be small
compared to the amplitudes of the semi-diurnal solutions. This is also found to be
the case. Yet another prediction is that amplitudes of the diurnal solutions should
be small compared to the amplitudes of the semi-diurnal solutions. This turns out
not to be the case (Cartwright 1999). In fact, this predication is an artifact of the
fact that we are assuming that the ocean covering the Earths surface is of constant
mean depth. If we relax this assumption, and allow the ocean depth to vary in a
realistic manner, then the amplitudes of the semi-diurnal solutions become compa-
rable to those of the diurnal solutions (Lamb 1993). The final prediction is that a
given solution is such that the tide at a given point on the Earth either oscillates in
phase or in anti-phase with the corresponding equilibrium solution (i.e., the tide is
either direct or inverted). In particular, for the dominant M2 tide, this implies that,
at a given point on the Earths surface, there should be either a tidal maximum or a
tidal minimum whenever the Moon culminates. It turns out that this is not the case
(Cartwright 1999). Indeed, the observed phase dierence between the M2 tide and
the corresponding equilibrium tide varies from place to place on the Earths surface,
and can take any value. In other words, high tide at a given point on the Earth can
occur three hours before the Moon culminates, or two hours after, et cetera, and the
time lag between high tide and the culmination of the Moon varies from place to
place. We conjecture that this eect is caused by the continents impeding the flow of
tidal waves around the Earth. Hence, we now need to consider tides in oceans that
do not cover the whole surface of the Earth.

12.13 Non-Global Ocean Tides

Suppose that the surface of the Earth is covered by an ocean of uniform depth d that
extends from = 0 to = , and from = to = + , where |+ | < 2. For
consistency with the previous analysis, we must imagine that the remaining surface
is covered by a frozen (i.e., = 0) ocean of uniform depth d. The motion of the
ocean under the action of the tide generating potential is governed by the Laplace
Terrestrial Ocean Tides 367

tidal equations, (12.153)(12.155), which are written



1
(sin ) + + = 0, (12.186)
sin
.. .gd
2 cos = ( F ), (12.187)
a 2
.. . gd
+ 2 cos = 2 ( F ), (12.188)
a sin
where

F (, , t) = (1 + k2 h2 ) 2 (, , t) + G F (, ;  ,  ) ( ,  , t) d . (12.189)

The finite extent of the ocean in the direction introduces the additional constraint

(, , t) = 0 (12.190)

into the problem.

12.14 Useful Lemma


Let X(, , t), Y(, , t), P(, , t), and Q(, , t) be well-behaved functions of and
. Suppose that
1 Y
DP = (sin X) + , (12.191)
sin
where
1 P(, , t)
= Y(, , t), (12.192)
sin
Suppose, further, that

1 X
DQ = (sin Y) , (12.193)
sin

where
Q(, , t) = 0. (12.194)
We wish to demonstrate that
P 1 Q
X= , (12.195)
sin
1 P Q
Y= + . (12.196)
sin
368 Theoretical Fluid Mechanics

Let

P 1 Q
X = X + + , (12.197)
sin
1 P Q
Y = Y + . (12.198)
sin

Note, from Equations (12.192) and (12.194), that

Y  (, , t) = 0. (12.199)

It follows from Equations (12.191), (12.193), (12.197), and (12.198) that

Y 
(sin X  ) + = 0, (12.200)

X 
(sin Y  ) = 0. (12.201)

Equations (12.199) and (12.201) imply that

P
X = , (12.202)

1 P
Y = , (12.203)
sin

and
1 P (, , t)
= 0. (12.204)
sin

Furthermore, Equation (12.200) yields

D P = 0. (12.205)

Multiplying by P , integrating over the ocean, and making use of the boundary con-
dition (12.204), we obtain


 2
2
P 1 P
+ d = 0. (12.206)
sin

Here, is the surface of the Earth that is covered by ocean, and d = sin d d.
It follows that P is a constant. Thus, Equations (12.202) and (12.203) imply that
X  = Y  = 0, and, hence, that Equations (12.195) and (12.196) are valid.
Terrestrial Ocean Tides 369

12.15 Transformation of Laplace Tidal Equations


Let X = , Y = , P = , and Q = in Equations (12.191)(12.196). It follows that
[cf., Equation (12.161)]

D = , (12.207)

1
D = (sin ) , (12.208)
sin

where
1 (, , t)
= 0, (12.209)
sin
(, , t) = 0. (12.210)

Here, use has been made of Equations (12.186) and (12.190). Furthermore [cf.,
Equations (12.159) and (12.160)],

1
= , (12.211)
sin
1
= + . (12.212)
sin

Let X = cos , Y = cos , P =  , and Q =  in Equations (12.191)


(12.196). It follows that

 1
D = (sin cos ) (cos ) , (12.213)
sin

1
D  = (sin cos ) + (cos ) , (12.214)
sin

where
1  (, , t)
= cos (, , t), (12.215)
sin
 (, , t) = 0. (12.216)

Furthermore,
 1 
cos = , (12.217)
sin
1  
cos = + . (12.218)
sin
370 Theoretical Fluid Mechanics

Substitution of Equations (12.211), (12.212), (12.217), and (12.218) into Equa-


tions (12.187) and (12.188) yields

A 1 B
+ = 0, (12.219)
sin
B 1 A
= 0, (12.220)
sin

where

.. . gd
A = 2  ( F), (12.221)
a2
.. .
B = 2 . (12.222)

Equations (12.219) and (12.220) can be combined to give

D A = 0, (12.223)
D B = 0. (12.224)

Moreover, it follows from Equations (12.188), (12.190), (12.209), (12.210), (12.215),


and (12.216) that

1 A(, , t)
= 0, (12.225)
sin
B(, , t) = 0. (12.226)

We have already seen that the solution of Equation (12.205), subject to the bound-
ary condition (12.204), is P = constant. It follows that the solution of Equa-
tion (12.223), subject to the boundary condition (12.225), is A = constant. Anal-
ogous arguments reveal that the solution of Equation (12.224), subject to the bound-
ary condition (12.226), is B = 0. Hence, we deduce that the Laplace tidal equations,
(12.186)(12.188), are equivalent to the following set of equations:

D = , (12.227)
.. . gd
2  ( F ) = constant, (12.228)
a2
.. .
2 = 0, (12.229)

where , ,  ,  , and F are defined in Equations (12.211), (12.212), (12.217),


(12.218), and (12.189), respectively.
Terrestrial Ocean Tides 371

12.16 Another Useful Lemma


Suppose that g(, ) and f (, ) are two well-behaved functions that satisfy either
1 f (, )
= 0, (12.230)
sin
1 g(, )
= 0, (12.231)
sin
or

f (, ) = 0, (12.232)
g(, ) = 0. (12.233)

It is easily demonstrated that


 

g f 1 g f
g D f d = + d = f D g d. (12.234)
sin2

12.17 Basis Eigenfunctions


Suppose that the r (, ) are well-behaved solutions of the eigenvalue equation

D r + r r = 0, (12.235)

subject to the boundary conditions


1 r (, )
= 0. (12.236)
sin
It immediately follows from Equations (12.234) and (12.235) that
 
(r D r r D r ) d = (r r ) |r | 2 d = 0. (12.237)

Hence, we deduce that the r are real. It follows that we can choose the r to be real
functions. Equations (12.234) and (12.235) also yield
  
2
2
r 1 r
r D r d = r r d =
2
+ sin d, (12.238)

which implies that the r are positive. Integration of Equation (12.235), subject to
the boundary condition (12.236), gives

r r d = 0. (12.239)

372 Theoretical Fluid Mechanics

Because r is positive, this implies that



r d = 0. (12.240)

Finally, Equations (12.234) and (12.235) yield


 
( s D r r D s ) d = ( s r ) s r d = 0. (12.241)

It follows that 
s r d = 0 (12.242)

if s  r . As is well known (Riley 1974), if r and r are linearly independent
solutions of (12.235) corresponding to the same eigenvalue, r , then it is always
possible to choose linear combinations of them that satisfy

r r d = 0. (12.243)

This argument can be extended to multiple linearly independent solutions corre-


sponding to the same eigenvalue. Hence, we conclude that it is possible to choose
the r such that they satisfy the orthonormality condition
 
r r s d = s r s d = r s . (12.244)

Let F(, ) be a well-behaved function. Suppose that


1 F(, )
= 0. (12.245)
sin
We can automatically satisfy the previous boundary condition by writing

F= ar r . (12.246)
r=1,

(Note that F is undetermined to an arbitrary additive constant which is chosen so as


to ensure that F d = 0.) Here, 1 is the smallest eigenvalue of Equation (12.235),
2 the next smallest eigenvalue, and so on. It follows from Equation (12.244) that

a r = r F r d. (12.247)

Suppose that the r (, ) are well-behaved solutions of the eigenvalue equation

D r + r r = 0, (12.248)

subject to the boundary conditions

r (, ) = 0. (12.249)
Terrestrial Ocean Tides 373

Using analogous arguments to those employed previously, we can show that the r
are real and positive, and that the r can be chosen so as to satisfy the orthonormality
constraint  
r r s d = s r s d = r s . (12.250)

Let F(, ) be a well-behaved function. Suppose that

F(, ) = 0. (12.251)

We can automatically satisfy the previous boundary condition by writing



F= ar r . (12.252)
r=1,

It follows from Equation (12.250) that



ar = r F r d. (12.253)

12.18 Auxilliary Eigenfunctions


Let
cos r
X= , (12.254)
sin
r
Y = cos , (12.255)

P = r , (12.256)
Q= r (12.257)

in Equations (12.191)(12.196). It follows that





 1 r r
D r = cos cos , (12.258)
sin



1 r 1 r
D r = sin cos + cos , (12.259)
sin sin

and
1 r (, ) r
= cos , (12.260)
sin
r (, ) = 0, (12.261)
374 Theoretical Fluid Mechanics

as well as
cos r  1 r
= r , (12.262)
sin sin
r 1 r r
cos = + . (12.263)
sin
Let
r
X = cos , (12.264)

cos r
Y= , (12.265)
sin
P = r , (12.266)
Q= r (12.267)
in Equations (12.191)(12.196). It follows that



 1 r 1 r
D r = sin cos + cos , (12.268)
sin sin



1 r r
D r = cos cos , (12.269)
sin
and
1 r (, ) cos r (, )
= , (12.270)
sin sin
r (, ) = 0, (12.271)
as well as
r  1 r
cos = r , (12.272)
sin
cos r 1 r r
= + . (12.273)
sin sin

12.19 Gyroscopic Coecients


Let
 
r, s = s r s d = r D s d



r s 1 r s
= + d, (12.274)
sin2
Terrestrial Ocean Tides 375

where use has been made of Equations (12.235) and (12.236). It follows from Equa-
tions (12.262) and (12.263) that


cos r s r s
r, s = d
sin


1 r s r s
d. (12.275)
sin
However, the second term on the right-hand side of the previous equation integrates
to zero with the aid of Equation (12.261). Hence, we are left with


cos r s r s
r, s = d. (12.276)
sin
Let
 
r, s = s r s d = r D s d



r s 1 r s
= + d, (12.277)
sin2
where use has been made of Equations (12.235) and (12.236). It follows from Equa-
tions (12.272) and (12.273) that


r s 1 r s
r, s = cos + d
sin2


1 r s r s
d. (12.278)
sin
However, the second term on the right-hand side of the previous equation integrates
to zero with the aid of Equation (12.271). Hence, we are left with


r s 1 r s
r, s = cos + d. (12.279)
sin2
Let
 
r, s = s r s d = r D s d



r s 1 r s
= + d, (12.280)
sin2
where use has been made of Equations (12.248) and (12.261). It follows from Equa-
tions (12.262) and (12.263) that


r s 1 r s
r, s = cos + d
sin2


1 r s r s
+ d. (12.281)
sin
376 Theoretical Fluid Mechanics

However, the second term on the right-hand side of the previous equation integrates
to zero with the aid of Equation (12.249). Hence, we are left with


r s 1 r s
r, s = cos + d. (12.282)
sin2

Finally, let
 
r, s = s r s d = r D s d



r s 1 r s
= + d, (12.283)
sin2

where use has been made of Equations (12.248) and (12.271). It follows from Equa-
tions (12.272) and (12.273) that


cos r s r s
r, s = d
sin


1 r s r s
+ d. (12.284)
sin

However, the second term on the right-hand side of the previous equation integrates
to zero with the aid of Equation (12.249). Hence, we are left with


cos r s r s
r, s = d. (12.285)
sin

Incidentally, the r,s , r,s , r,s , and r,s are known collectively as gyroscopic
coecients (Proudman 1916).

12.20 Proudman Equations


We can automatically satisfy the boundary condition (12.209) by writing

(, , t) = pr (t) r (, ). (12.286)
r=1,

Likewise, we can automatically satisfy the boundary condition (12.210) by writing



(, , t) = pr (t) r (, ). (12.287)
r=1,
Terrestrial Ocean Tides 377

It follows from Equations (12.211), (12.212), (12.227), and (12.235) that



r 1 r

= pr + pr , (12.288)
r=1,
sin

r 1 r

= pr pr , (12.289)
r=1,
sin

= r pr r . (12.290)
r=1,

Let
 
A r = r  r d =  D r d (12.291)



 r 1  r
= + d, (12.292)
sin2
where use has been made of Equations (12.235) and (12.236). It follows from Equa-
tions (12.217) and (12.218) that


r r
Ar = cos + d
sin


1  r  r
d. (12.293)
sin

However, the second term on the right-hand side of the previous equation integrates
to zero with the aid of Equation (12.216). Hence, we are left with


r r
Ar = cos + d. (12.294)
sin

Finally, Equations (12.276), (12.282), (12.288), and (12.289) can be combined with
the previous equation to give
  3 4
r  r d = r, s p s + r, s ps . (12.295)
s=1,

Let
 
Ar = r  r d =  D r d



 r 1  r
= + d, (12.296)
sin2
where use has been made of Equations (12.216) and (12.248). It follows from Equa-
378 Theoretical Fluid Mechanics

tions (12.217) and (12.218) that




r r
Ar = cos + d
sin


1  r  r
+ d. (12.297)
sin

However, the second term on the right-hand side of the previous equation integrates
to zero with the aid of Equation (12.249). Hence, we are left with


r r
Ar = cos + d. (12.298)
sin

Finally, Equations (12.279), (12.285), (12.288), and (12.289) can be combined with
the previous equation to give
  3 4
r  r d = r, s p s + r, s ps . (12.299)
s=1,

Operating on Equations (12.228) and (12.229) with r ( ) r d and with

r ( ) r d, respectively, yields the so-called Proudman equations (Proudman
1916):

..  . . gd gd
pr + 2 (r, s p s + r, s ps ) + r pr = 2 Fr , (12.300)
s=1,
a2 a
..  . .
pr + 2 (r, s p s + r, s ps ) = 0, (12.301)
s=1,

where 
F r = r F r d. (12.302)

Here, use has been made of Equations (12.240), (12.244), (12.250), (12.286), (12.287),
(12.290), (12.295), and (12.299). Finally, it follows from Equations (12.138), (12.189),
and (12.290) that 
F r = Pr r, s p s , (12.303)
s=1,

where 
P r = r (1 + k2 h2 ) 2 r d, (12.304)

and  
r, s = (1 + kn hn ) n n,mr n,ms , (12.305)
n=1, m=0,n

with 
n,mr = r Pnm (cos ) e i m r (, ) d. (12.306)

Terrestrial Ocean Tides 379

Here,
Pnm (x) = bnm Pnm (x), (12.307)
where
1/2
2 n + 1 (n m)!
bnm = . (12.308)
4 (n + m)!
It follows from Equation (12.136) that

n n
Pnm (cos ) Pnm (cos ) sin d = . (12.309)
0 2

Consider the response of the ocean to a particular harmonic of the tide generating
potential for which

2 (, , t) = j P2 j (cos ) e i (m j + j t) .
m
(12.310)

Assuming a common exp( i j t) time dependence of the pr and the pr , Equations


(12.300) and (12.301) yield
  
( 1 r f j2 ) pr + (i f j r, s 1 r, s ) p s + i f j r, s ps
s=1,
m
1 (1 + k2 h2 ) j 2, rj
= mj , (12.311)
b2

and 
f j2 pr + i ( f j r, s p s + f j r, s ps ) = 0, (12.312)
s=1,

where f j and are defined in Equations (12.173) and (12.174), respectively.


The task in hand (see Section 12.13) is to solve the Laplace tidal equations,
(12.186)(12.189), subject to the constraint (12.190). Our basic approach is to ex-
pand the fields appearing in the Laplace tidal equationsnamely, , , and in
terms of a set of basis eigenfunctionsthe r and the r (see Section 12.17)
that are defined in such a manner as to automatically satisfy the boundary condi-
tions. The expansion is specified in Equations (12.288)(12.290). The Laplace tidal
equations then reduce to the Proudman equations, (12.300)(12.301), which are a
set of coupled ordinary dierential equations for the weights in the expansionthe
pr (t) and the pr (t). For a particular harmonic of the tide generating potential, the
weights all oscillate at the same frequency, and the Proudman equations reduce fur-
ther to give two coupled matrix equations for the amplitudes of the weightssee
Equations (12.311) and (12.312). In order to solve the matrix equations, we need
to calculate the gyroscopic coecients, r,s , r,s , r,s , r,s , as well as the n,r
m
.
These quantities are integrals involving the eigenfunctions and the associated Leg-
endre functions, and are defined in Equations (12.276), (12.279), (12.282), (12.285),
and (12.306), respectively.
380 Theoretical Fluid Mechanics

12.21 Hemispherical Ocean Tides


Consider a hemispherical ocean for which = 0 and + = . In general, there
are many linearly independent solutions to the eigenvalue problem (12.235) and
(12.236), subject to the orthonormality constraint (12.244), that correspond to a given
value of r . It is convenient to dierentiate these solutions by means of two indices,
n and m. Thus, r nm and r nm , where the index n ranges from 1 to .
Moreover, for a given value of n, the index m ranges from 0 to n. Now,



m = m = 0




cos(m ) cos(m ) d =
/2 m = m  0 . (12.313)


0 m  m
0

Hence, it follows that

nm (, ) = anm Pnm (cos ) cos(m ), (12.314)

and
nm = n (n + 1), (12.315)
where 
(4/nm )1/2 m>0
anm = . (12.316)
(2/nm )1/2 m=0
Likewise, the solutions to the eigenvalue problem (12.248) and (12.249), subject to
the orthonormality constraint (12.250), is such that r nm and r nm , where

nm (, ) = anm Pnm (cos ) sin(m ), (12.317)

and
nm = n (n + 1). (12.318)

We also have (see Section 12.19)


  
 cos nm nm nm nm
r, s n,m,nm
=
d, (12.319)
sin
  
m, m nm nm 1 nm nm
r, s n, n = cos + d, (12.320)
sin2
  
m, m nm nm 1 nm nm
r, s n, n = cos + d, (12.321)
sin2
  
m, m cos nm nm nm nm
r, s n, n = d, (12.322)
sin
Terrestrial Ocean Tides 381

which yields


  
m, m nm nm
1
nm nm
n, n = d d, (12.323)
0 1
  1  
m, m  nm nm 1 nm nm
n, = (1 ) + d d,
2
n
0 1 1 2
(12.324)
  1  
m  nm nm
1 nm n
m
n,m,n = (1 2 ) + d d, (12.325)
0 1 1 2
  1  
m, m  nm nm nm nm
n, n = d d, (12.326)
0 1

where = cos . However,


 
 2 m/(m 2 m 2 ) m + m odd
sin(m ) cos(m ) d = . (12.327)
0 0 m + m even

m, m
Hence, if m + m is even then the n, n are zero. Otherwise, we obtain
  
n,m,nm 2 2 m dPnm
1
dPnm m
 = m Pn + m 2 Pn d, (12.328)
cnm cnm m 2 m 2
1 d d
m, m   1 
n, 2m dPnm dPnm m 2 

P P  d, (12.329)
n
= 2 (1 ) +
2 m m

cnm cnm m m 2 1 d d 1 2 n n
m, m
n,
  1 
n 2 m m m dPnm dPnm m
 = 2 P + P  d, (12.330)
cnm cnm m m 2 1 n d d n

as well as  
m m , m
n,m,n = n, n , (12.331)
where
cnm = anm bnm . (12.332)
Let  1
 
In,m,nm = Pnm () Pnm () d. (12.333)
1

It follows, from symmetry, that In,m,nm = 0 when n + n m m is odd. When
n + n m m is even (Wong 1998),

  
In,m,nm = n,p
m
nm ,p
p=0,pmax p =0,pmax

[(n + n m m 2p 2p + 1)/2] [(m + m + 2p + 2p + 2)/2]


, (12.334)
[(n + n + 3)/2]
382 Theoretical Fluid Mechanics

where (x) denotes a gamma function (Abramowitz and Stegun 1965),

(1)m+p (n + m)!
n,p
m
= , (12.335)
2 m+2p (m + p)! p! (n m 2p)!

and 
(n m)/2 n m even
pmax = , (12.336)
(n m 1)/2 n m odd
with an analogous definition for pmax .
It can be shown that (Longuet-Higgins and Pond 1970)

n,m,nm n (n + 1) + m (m + 1)  
 n (n + 1) n (n + 1) + m
 
 = 2m In,m,nm
cnm cnm m + 1 m 2 + m 2


1 
+  I m, m +2 , (12.337)
m + 1 n, n

m
n,m,n 2m  
 = (n 1) (n + 1) (n + m ) In,m,nm 1
cnm cnm (m 2 
m ) (2 n + 1)
2

 
+n (n + 2) (n m + 1) In,m,nm +1 , (12.338)

m, m
n, n 2 m m m, m
 = I  . (12.339)
cnm cnm m 2 m 2 n, n

According to Equation (12.306), we can also write n,r
m
n,n
m,m
 , where

m,m
 m,m m,m

n,n = (anm )1 n,n + i n,n , (12.340)

and

n,m,nm = n n m m . (12.341)

Here, n,m,nm = 0 if m + m is even, and


 2m   
n,m,nm = 
nm cnm cnm In,m,nm , (12.342)
m m
2 2


otherwise. It follows from Equation (12.305) that r, s n,m,nm , where
   
n,m,nm = (1 + kn hn ) n (anm )2 n,m,nm i n,m,nm + n,m,nm . (12.343)

Here, n,m,nm = 0 if m + m is even, and
   
n,m,nm = (1 + kn hn ) n (anm )2 n,n
m,m
 (1 + kn  hn ) n (anm )2 nm ,n,m (12.344)
Terrestrial Ocean Tides 383

otherwise. Now, if m and m are both even then


 m

odd
   
n,m,nm = (1 + kn  hn ) n (anm )2 nm ,n,m nm ,n,m ; (12.345)
n =1, m =0,n

if m and m are both odd then


 m

even
   
n,m,nm = (1 + kn  hn ) n (anm )2 nm ,n,m nm ,n,m ; (12.346)
n =1, m =0,n


and if m + m is odd then n,m,nm = 0.
m
If we let pr i n pnm and pr i n+1 pn then Equations (12.311) and (12.312)
become
     
1 [nm (1 + kn hn ) n (anm )2 ] f j2 n,m,nm + f j n,m,nm
n =1, m =0,n
   
 m m
+ 1 (n,m,nm n,m,nm ) pnm f j n,m,n pn
1 (1 + k2 h2 ) j  m ,m m ,m

= m 2, nj + 2, nj , (12.347)
c2 j

and
     
m, m m m, m m
f j2 n,m,nm + f j n, n pn + f j
n, n pn = 0, (12.348)
n =1, m =0,n

respectively. Here,
  
n,m,nm = i n+n n,m,nm , (12.349)
  
n,m,nm = i n+n +1 n,m,nm , (12.350)
  
m m
n,m,n = i n+n n,m,n , (12.351)
  
m, m m, m
n, n = i
n+n
n, n , (12.352)
m, m  m, m
n, n = i n+n +1 n, n , (12.353)
 
n+n +1
n,m,nm =i n,m,nm , (12.354)
  
n,m,nm = i n+n n,m,nm , (12.355)
  
n,m,nm = i n+n +1 n,m,nm . (12.356)
 
By symmetry, n,m,nm and n,m,nm are only non-zero when m + m is even, and n + n is

m, m m, m m, m 
even; n,m,nm , n, n , n, n , and n, n are only non-zero when m + m is odd, and
 
m m,
n + n is odd; and n, n and n, n are only non-zero when m + m is odd, and n + n
m, m

is even. It follows that all quantities appearing in Equations (12.347) and (12.348)
384 Theoretical Fluid Mechanics

are real. Once we have solved these equations to obtain the pnm (which is a relatively
straightforward numerical task), we can reconstruct the tidal elevation as follows:

(, , t) = c (, ) cos( j t) + s (, ) sin( j t), (12.357)

where
n
even 
c (, ) = nm m
n (, ) cos(n /2), (12.358)
n=1, m=0,n
n
odd 
s (, ) = nm m
n (, ) sin(n /2), (12.359)
n=1, m=0,n

and
nm = nm pnm . (12.360)
Thus, the tidal amplitude at a given point on the ocean is
 1/2
||(, ) = c2 (, ) + s2 (, ) . (12.361)

It is easily demonstrated that


t(, ) (, )
= , (12.362)
T 2
where
1 s (, )
(, ) = m j + tan . (12.363)
2 c (, )
Here, T = 2/ j is the oscillation period of the harmonic of the tide generating
potential under consideration, t is the time-lag between the peak tide at a given
point on the ocean and the maximal tide generating potential at = /2, and is
the corresponding phase-lag.
Figures 12.4 and 12.5 show the amplitude and phase-lag of the ( j = 3) M f long-
period tide in a hemispherical ocean of mean depth d = 3.8 km (which corresponds to
= 23.1), calculated assuming that / = 5.5 and /( g a) = 0.35. The calculation
includes all azimuthal harmonics up to nmax = 20. Note that only a quarter of the
ocean is shown, because the amplitude is symmetric about the meridians = 0
and = /2, whereas the phase-lag is symmetric about the meridian = 0, and
antisymmetric about the meridian = /2. Here, /2 . Given that 3 =
4.25 102 m (see Table 12.3), the maximum amplitude of the tide is about 3.5 cm,
and occurs at the poles. Moreover, it is clear from a comparison with Figure 12.1
that the tide is direct (i.e., it is in phase with the equilibrium tide). In fact, the M f
tide in a hemispherical ocean is about four times larger in amplitude than that in a
global ocean (i.e., an ocean that covers the whole surface of the Earth) of the same
depth. Otherwise, the two tides have fairly similar properties.
Figures 12.6 and 12.7 show the amplitude and phase-lag of the ( j = 6) K1 diur-
nal tide in a hemispherical ocean of mean depth d = 3.8 km (which corresponds to
Terrestrial Ocean Tides 385

= 23.1), calculated assuming that / = 5.5 and /( g a) = 0.35. The calculation


includes all azimuthal harmonics up to nmax = 20. Given that 6 = 10.36 102 m
(see Table 12.3), the maximum amplitude of the tide is about 15.8 cm, and occurs at
mid-latitudes. This is very dierent to the case of a global ocean, where the ampli-
tude of the K1 tide is identically zero everywhere. (See Figure 12.2.) Note that the
tidal phase-lag only exhibits a fairly weak dependence on the azimuthal angle, . In
fact, the K1 tide in a hemispherical ocean essentially oscillates in anti-phase with the
equilibrium tide at the oceans central longitudinal meridian ( = /2), except close
to the poles, where it oscillates in phase. Again, this is very dierent to the case of a
global ocean, where the tidal maximum lies on a meridian of longitude, and rotates
steadily around the Earth from east to west.

Figures 12.8 and 12.9 show the amplitude and phase-lag of the ( j = 8) M2 semi-
diurnal tide in a hemispherical ocean of mean depth d = 3.8 km (which corresponds
to = 23.1), calculated assuming that / = 5.5 and /( g a) = 0.35. The calcula-
tion includes all azimuthal harmonics up to nmax = 20. Given that 8 = 8.95 102 m
(see Table 12.3), the maximum amplitude of the tide is about 60.7 cm, and occurs
at the poles. This is very dierent to the case of a global ocean, where the ampli-
tude of the M f tide is zero at the poles. (See Figure 12.3.) Another major dierence
is that, in a hemispherical ocean, the tidal maxima circulate around points of zero
tidal amplitudesuch points are known as amphidromic points. Of course, in the
case of a global ocean, the tidal maxima lie on opposite meridians of longitude, and
rotate steadily around the Earth from east to west. The systems of tidal waves, cir-
culating around amphidromic points, that is evident in Figure 12.9, are known as
amphidromic systems, and are one of the the major features of tides in real oceans
(Cartwright 1999). Generally speaking, the sense of circulation of amphidromic sys-
tems in real oceans is counter-clockwise (seen from above) in the Earths northern
hemisphere, and clockwise in the southern hemisphere.

In conclusion, our investigation of tides in a hemispherical ocean suggests that


the impedance of the flow of tidal waves around the Earth, due to the presence of
the continents, is likely to have a comparatively little eect on long-period tides, but
a very significant eect on diurnal and semi-diurnal tides. In particular, for semi-
diurnal tides, the impedance gives rise to the formation of amphidromic systems.

12.22 Exercises

12.1 Consider a tidal wave whose wavelength is very much less than the radius of
the Earth. This corresponds to the limit n
1, where n is an azimuthal mode
number. Suppose, however, that the ocean depth, d, is allowed to vary with
386 Theoretical Fluid Mechanics

Figure 12.4
Relative amplitude, ||/3 , of the ( j = 3) M f long-period tide in a hemispherical
ocean.

Figure 12.5
Phase-lag, , of the ( j = 3) M f long-period tide in a hemispherical ocean.
Terrestrial Ocean Tides 387

Figure 12.6
Relative amplitude, ||/|6 |, of the ( j = 6) K1 diurnal tide in a hemispherical ocean.

Figure 12.7
Phase-lag, , of the ( j = 6) K1 diurnal tide in a hemispherical ocean.
388 Theoretical Fluid Mechanics

Figure 12.8
Relative amplitude, ||/8 , of the ( j = 8) M2 semi-diurnal tide in a hemispherical
ocean.

Figure 12.9
Phase-lag, , of the ( j = 8) M2 semi-diurnal tide in a hemispherical ocean.
Terrestrial Ocean Tides 389

position. Show that, in this limit, the Laplace tidal equations are written

1 (v d)
= (sin u d) + ,
t a sin
u g
2 cos v = (  ),
t a
v g
+ 2 cos u = (  ),
t a sin

where
 = (1 + k2 h2 ) 2 .

12.2 Consider short-wavelength tidal waves in a region of the ocean that is su-
ciently localized that it is a good approximation to treat cos and sin as con-
stants. We can define local Cartesian coordinates, x, y, z, such that dx = a d,
dy = a sin d, and dz = r a. It follows that the x-axis is directed south-
ward, the y-axis is directed eastward, and the z-axis is directed vertically
upward. Show that, when expressed in terms of this local coordinate system,
the Laplace tidal equations derived in the previous exercise reduce to

(u d) (v d)
= ,
t x y
u (  )
f v = g ,
t x
v (  )
+ f u = g ,
t y

where
f = 2 cos
is a constant. In fact, the previous equations are the linearized equations of
motion of a body of shallow water confined to a tangent plane that touches
the Earth at the angular coordinates , . This plane is known as the f -plane
because, as a consequence of the Earths diurnal rotation, it rotates about ez
at the angular velocity f .

12.3 Demonstrate that the set of equations derived in the previous exercise can be
written in the coordinate-free (in the x-y plane) form


= (d v),
t
v
+ f ez v = g (  ),
t
where v = u e x + v ey . Let = ez v and = v. Assuming that d and
390 Theoretical Fluid Mechanics

f are constants, demonstrate that the previous equations are equivalent to


2
+ f 2 g d 2 = f g 2  ,
t 2


2 2

+ f 2
g d 2
= g ,
t 2 t

where 2 denotes a two-dimensional Laplacian (in the x-y plane).


12.4 Consider free (i.e.,  = 0) plane-wave solutions to the f -plane equations,
derived in Exercise 12.2, of the form

(r, t) = 0 e i ( tkr) ,
u(r, t) = u0 e i ( tkr) ,
v(r, t) = v0 e i ( tkr) .

Here, r = (x, y), and 0 , u0 , v0 are constants. Assuming that d is constant,


show that


k x i f ky
u0 = g 0 ,
2 f 2


ky + i f k x
v0 = g 0 , ,
2 f 2
and
2 = f 2 + c 2 k 2,

where c = d g. This type of wave is known as a Poincare wave.
12.5 Suppose that the region y 0 corresponds to an ocean of constant depth d,
whereas the region y > 0 corresponds to land. Consider free solutions to the
f -plane equations in the region y < 0. We can trivially satisfy the constraint
v(x, 0, t) = 0 by searching for solutions which are such that v = 0 for all y 0.
Show that the most general such solution takes the form

(x, y, t) = Z1 (x + c t, y) + Z2 (x c t, y),

where c = d g, and

Z1 (x + c t, y) = Z1 (x + c t, 0) e+s y/l ,
Z2 (x c t, y) = Z2 (x c t, 0) es y/l .

Here, Z1 (x, 0) and Z2 (x, 0) are arbitrary functions, l = d g/| f |, and s =
sgn( f ). These solutions are known as Kelvin waves. Deduce that Kelvin
waves propagate along coastlines, at the speed c, in such a manner as to keep
the coastline to the right of the direction of propagation in the Earths northern
hemisphere, and to the left in the southern hemisphere.
Terrestrial Ocean Tides 391

12.6 We can take into account the latitude dependence of the parameter f by writ-
ing f = f0 x, where f0 = 2 cos and = 2 sin /a. Let us assume that
the ocean is of constant depth, d. Furthermore, let us search for an almost in-
compressible, free solution of the Laplace tidal equations which is such that
u /y and v /x. By eliminating from the final two f -plane
equations, show that


2 + 0.
t y
By searching for a wavelike solution of the previous equation of the form
= 0 e i ( tkr) , deduce that

k 2 + ky = 0.

This is the dispersion relation of a so-called Rossby wave. Demonstrate that


Rossby waves always travel with a westward component of phase velocity.
Finally, show that it is reasonable to neglect compression provided that 
| f0 |.
392 Theoretical Fluid Mechanics
13
Equilibrium of Compressible Fluids

13.1 Introduction
In this chapter, we investigate the equilibria of compressible fluids, such as gases.
As is the case for an incompressible fluid (see Chapter 2), a compressible fluid in
mechanical equilibrium must satisfy the force balance equation

0 = p + , (13.1)

where p is the static fluid pressure, the mass density, and the gravitational po-
tential energy per unit mass. In an ideal gas, the relationship between p and is
determined by the energy conservation equation, (1.89), which can be written



D p M 2 p
= . (13.2)
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 previous equation (because it is zero in a stationary gas). The
limits in which the left- and right-hand sides of Equation (13.2) are dominant are
termed the adiabatic and isothermal limits, respectively. In the isothermal limit, in
which thermal transport is comparatively large, so that Equation (13.2) 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 conse-
quently related according to the isothermal gas law,

p
= constant. (13.3)

On the other hand, in the adiabatic limit, in which thermal transport is negligible, so
that Equation (13.2) can only be satisfied when D/Dt(p/ ) 0, the pressure and
density are related according to the adiabatic gas law,

p
= constant. (13.4)

393
394 Theoretical Fluid Mechanics

13.2 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. Conse-
quently, 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 func-
tions 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 (13.1) that
dp
= g. (13.5)
dz
Now, in an isothermal atmosphere, in which the temperature, T , is assumed not to
vary with height, the ideal gas equation of state (1.84) yields [cf., Equation (13.3)]
p RT
= . (13.6)
M
The previous two equations can be combined to give
dp gM
= p. (13.7)
dz RT
Hence, we obtain
p(z) = p0 exp(z/H), (13.8)
2
where p0 10 N m
5
is atmospheric pressure at ground level, and
RT
H= (13.9)
gM
is known as the isothermal scale height of the atmosphere. Using the values T =
273 K (0 C), M = 29 103 kg, 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 (13.6) and (13.8) yield
(z) = 0 exp(z/H), (13.10)
where 0 = p0 /(g H) is the atmospheric mass density at z = 0. According to Equa-
tions (13.8) and (13.10), 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.

13.3 Adiabatic Atmosphere


In fact, the temperature of the Earths atmosphere is not uniform, but instead de-
creases steadily with increasing altitude. This eect is largely due to the action of
Equilibrium of Compressible Fluids 395

convection currents. When a packet of air ascends, under the influence of such cur-
rents, the diminished pressure at higher altitudes causes it to expand. Because this
expansion generally takes place far more rapidly than heat can diuse 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, while remaining in approximate
vertical force balance (such a state is known as a convective equilibrium), and that
the eect of heat conduction is negligible (because the mixing takes place too rapidly
for thermal diusion to aect the temperature), we would expect the adiabatic gas
law, (13.4), to oer a much more accurate description of the relationship between
atmospheric pressure and density than the isothermal gas law, (13.3).
Let p = p/p0 , = /0 , and T = T/T 0 , where p0 , 0 , and T 0 are the pressure,
mass density, and temperature of the atmosphere, respectively, at ground level. The
adiabatic gas law, (13.4), can be combined with the ideal gas equation of state, (13.6),
to give
p = = T /(1) . (13.11)
The isothermal scale height of the atmosphere is conveniently redefined as [cf., Equa-
tion (13.9)]
R T0 p0
H= = . (13.12)
gM g 0
Equations (13.5), (13.11), and (13.12) yield
d p
= p 1/ , (13.13)
dz
where z = z/H, or, from Equation (13.11),
dT 1
= . (13.14)
dz
The previous equation can be integrated to give


1 z
T (z) = T 0 1 . (13.15)
H
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.
Equations (13.11) and (13.15) yield

/(1)
1 z
p(z) = p0 1 , (13.16)
H

1/(1)
1 z
(z) = 0 1 . (13.17)
H
396 Theoretical Fluid Mechanics

Because /( 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 (because they cannot take negative or imaginary values). In contrast,
an isothermal atmosphere has a diuse 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, because the adiabatic
gas law does not hold at very high altitudes.

13.4 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 required for
heat to diuse across it (but still a suciently 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, (13.4). 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 packets 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 packets
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 :
that is, when
 
p  p 
 <  , (13.18)
z2 z1

for any z2 > z1 . It follows that the atmosphere is only stable to vertical convection
when p/ is a monotonically increasing function of altitude. As is easily demon-
strated, this stability criterion can also be written

d ln p
< , (13.19)
d ln
Equilibrium of Compressible Fluids 397

or, making use of the ideal gas equation of state,


d ln T
< 1. (13.20)
d ln
Convection is triggered in regions of the atmosphere where the previous stability
criterion is violated. However, such convection acts to relax these regions back to
a marginally-stable state in which p/ is uniform: in other words, an adiabatic
equilibrium.

13.5 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 coincides with
the Suns geometric center, and making the simplifying (and highly accurate) as-
sumption that the mass distribution within the Sun is spherically symmetric, we find
that
dm
= 4 r 2 , (13.21)
dr
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 (Fitzpatrick 2012). In other words,
d Gm
= 2 , (13.22)
dr r
where (r) is the gravitational potential energy per unit mass, and d/dr the radial
gravitational acceleration. The force balance criterion (13.1) yields
dp d
+ = 0, (13.23)
dr dr
where p(r) is the pressure. The previous three equations can be combined to give


1 d r 2 dp
= 4 G . (13.24)
r 2 dr dr

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 complicated. This energy is
398 Theoretical Fluid Mechanics

then transported from the core to the outer boundary via a combination of convec-
tion 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 dicult subject. Finally,
once the energy reaches the boundary of the star, it is radiated away. The following
ingenious model, due to Eddington (Eddington 1926), 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 suciently high 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 , (13.25)
where
kT
pg = (13.26)
mp
is the gas pressure (modeling the plasma within the Sun as an ideal gas of free elec-
trons and ions), and (Chandrasekhar 1967)
1
pr = T 4 (13.27)
3
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
Equation (13.26), we have expressed M/R in the equivalent form m p /k.
Let
pg = (1 ) p, (13.28)
pr = p, (13.29)
where the parameter is assumed to be uniform. In other words, the ratio of the ra-
diation pressure to the gas pressure is assumed to be the same everywhere inside the
Sun. This fairly drastic assumption turns outperhaps, somewhat fortuitously (Mes-
tel 1999)to lead to approximately the correct internal pressure-density relation for
the Sun. In fact, Equations (13.26)(13.29) can be combined to give
p = K 4/3 , (13.30)
where
4 1/3
k 3
K = .
m p (1 )
4 4
(13.31)
Equilibrium of Compressible Fluids 399
2

, y 1

0
0 1 2 3 4 5 6 7

Figure 13.1
The functions () (solid) and y() (dashed).

It can be seen, by comparison with Equation (13.4), that the previous pressure-
density relation takes the form of an adiabatic gas law with an eective 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 Equation (13.30), 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 Equation (13.30) that

pc = K 4/3
c , (13.32)

and from Equations (13.26) and (13.28) that

pc m p
Tc = (1 ) . (13.33)
c k

Suppose that
T
= , (13.34)
Tc

where (r) is a dimensionless function. According to Equations (13.26), (13.28), and


400 Theoretical 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 13.2
Solar temperature versus mass fraction obtained from the Eddington Solar Model
(solid) and the Standard Solar Model (dashed).

(13.30),

= 3, (13.35)
c
p
= 4. (13.36)
pc

Moreover, it is clear, from the previous 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 , (13.37)

where is a dimensionless radial coordinate, and



1/2
K
a= . (13.38)
G c2/3

Thus, the center of the Sun corresponds to = 0, and the edge to = 1 (say), where
(1 ) = 0, and
R = 1 a. (13.39)
Equations (13.35)(13.38) can be used to transform the equilibrium relation (13.24)
Equilibrium of Compressible Fluids 401

4
log10 ( [kg m3])

1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
m/M

Figure 13.3
Solar mass density versus mass fraction obtained from the Eddington Solar Model
(solid) and the Standard Solar Model (dashed).

into the non-dimensional form




1 d 2 d
= 3 . (13.40)
2 d d
Moreover, Equation (13.21) can be integrated, with the aid of Equations (13.35),
(13.37), and (13.40), and the physical boundary condition m(0) = 0, to give
m = 4 c a 3 y, (13.41)
where
d
y() = 2
. (13.42)
d
Equation (13.40) is known as the Lane-Emden equation (of degree 3), and can, unfor-
tunately, only be solved numerically (Chandrasekhar 1967). The appropriate solution
takes the form = 1 2 /6 + O( 4 ) when  1, and must be integrated to = 1 ,
where (1 ) = 0. Figure 13.1 shows (), and the related function y(), obtained via
numerical methods. Note that 1 = 6.897, and y1 = y(1 ) = 2.018.
According to Equation (13.41), the solar mass, M = m(R), can be written
M = 4 c a 3 y1 , (13.43)
which reduces, with the aid of Equations (13.31) and (13.38), to

= 2, (13.44)
(1 )4
402 Theoretical Fluid Mechanics

where
M
= 2 , (13.45)
M0
and
4 1/2
3
M0 =
1 k
4 y1 = 3.586 1031 kg.
( G)3 m p
(13.46)

Moreover, it is easily demonstrated that

G M2/3
K= . (13.47)
(4 y1 )2/3

According to Equations (13.44) and (13.45), 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 (13.44) can be inverted to give the approximate solution

1 4
= 2 + O( 6 ). (13.48)
4
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 in the following), 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 ecient than convective energy transport. Indeed, in
regions of the Sun in which convection occurs, we would expect the convective trans-
port 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 (13.32), (13.33), (13.39), (13.43), and (13.47) yield

1 G M m p
Tc = (1 ) = 1.34 107 K, (13.49)
4 y1 Rk
13 M
c = = 7.63 104 kg m3, (13.50)
4 y1 R 3
14 G M2
pc = = 1.24 1016 N m2 , (13.51)
16 y12 R4

where the solar radius R has been given the observed value 6.960 108 m. The ac-
tual values of the Suns central temperature, density, and pressure, as determined by
the so-called Standard Solar Model (SSM),1 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 ,
1 http://en.wikipedia.org/wiki/Standard\_Solar\_Model
Equilibrium of Compressible Fluids 403
17

16

15
log10 (p [N m2])
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 13.4
Solar pressure versus mass fraction obtained from the Eddington Solar Model (solid)
and the Standard Solar Model (dashed).

and pc obtained from the Eddington model lie within a factor of two of those ob-
tained from the much more accurate SSM. Figures 13.2, 13.3, and 13.4 show the
temperature, density, and pressure profiles, respectively, obtained from the SSM 2
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 13.5 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.

2 The SSM data is obtained from http://www.ap.stmarys.ca/guenther/evolution/ssm1998.

html
404 Theoretical Fluid Mechanics
2.5

3
log10[/(1 )]

3.5

4
0 0.2 0.4 0.6 0.8 1
m/M

Figure 13.5
Ratio of radiation pressure to gas pressure calculated from the Eddington Solar
Model (solid) and the Standard Solar Model (dashed).

13.6 Exercises
13.1 Prove that the fraction of the whole mass of an isothermal atmosphere that
lies between the ground and a horizontal plane of height z is

1 ez/H .

Evaluate this fraction for z = H, 2 H, 3 H, respectively.


13.2 If the absolute temperature in the atmosphere diminishes upwards according
to the law
T z
=1 ,
T0 c
where c is a constant, show that the pressure varies as
p $ z %c/H
= 1 .
p0 c

13.3 If the absolute temperature in the atmosphere diminishes upward according


to the law
T 1
= ,
T0 1 + z
Equilibrium of Compressible Fluids 405

where is a constant, show that the pressure varies as




p z 1 z2
= exp .
p0 H 2 H

13.4 Show that if the absolute temperature, T , in the atmosphere is any given func-
tion of the altitude, z, then the vertical distribution of pressure in the atmo-
sphere is given by 
p T 0 z dz
ln = .
p0 H 0 T

13.5 Show that if the Earth were surrounded by an atmosphere of uniform temper-
ature 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.

13.6 Show that if the whole of space were occupied by air at the uniform tem-
perature T then the densities at the surfaces of the various planets would be
proportional to the corresponding values of
$g M a%
exp ,
RT
where a is the radius of the planet, and g its surface gravitational acceleration.

13.7 Prove that in an atmosphere arranged in horizontal strata the work (per unit
mass) required to interchange two thin strata of equal mass without distur-
bance of the remaining strata is

1  1 
p1 p2

2 11 ,
1 1 2

where the suxes refer to the initial states of the two strata. Hence, show that
for stability the ratio p/ must increase upwards.

13.8 A spherically symmetric star is such that m(r) is the mass contained within ra-
dius r. Show that the stars total gravitational potential energy can be written
in the following three alternative forms:
 M  M  R
Gm 1
U = dm = dm = 3 p dV
0 r 2 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 r 2 dr.
406 Theoretical Fluid Mechanics

13.9 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
1/2
(n + 1) K (n1)/n
= c .
4 G
Show that the physical solution to the Lane-Emden equation, which is such
that (0) = 1 and (1 ) = 0, for some 1 > 0, is
2
=1
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
3 G M2
U= ,
5n R
where M is the total mass, and R = a 1 the radius.
13.10 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. Demon-
strate that the central pressure is
5 G M2
pc = ,
4 R
where M is the mass of the star. Finally, show that the total gravitational
potential energy of the star can be written
26 G M2
U= .
35 R
14
One-Dimensional Compressible Inviscid Flow

14.1 Introduction
This chapter investigates one-dimensional, compressible, inviscid flow. Flow is said
to be one-dimensional when the fluid properties only depend on a single Cartesian
coordinate. Flow is said to be compressible when there is a significant variation in
the mass density along a given streamline. Generally speaking, compressible flow is
much more common in gases than in liquids. As described in Section 1.17, subsonic
gas dynamics (i.e., dynamics in which the typical flow speed is much less than the
propagation speed of sound waves in the gas) is essentially incompressible, and is
governed by the same equations that govern the incompressible flow of liquids. On
the other hand, sonic gas dynamics (i.e., dynamics in which the typical flow speed is
comparable with the sound speed) exhibits significant dierences to subsonic dynam-
ics. Hence, this chapter will concentrate on sonic gas dynamics. More information
on this subject can be found in Liepmann & Roshko 1957, Hughes & Brighton 1999,
and Milne-Thomson 2011.

14.2 Thermodynamic Considerations


Consider a quantity of gas contained in some form of enclosure. If the system is
left alone for a suciently long time then it will attain an equilibrium state in which
all macroscopically measurable quantities become independent of time. Examples
of such quantities are the mass, m, the pressure, p, the volume, V, and the absolute
temperature, T . It is helpful to distinguish between so-called extensive and intensive
equilibrium quantities. A quantity is said to be extensive if it is proportional to the
mass of the gas present in the enclosure. On the other hand, a quantity is said to
be intensive if it is independent of the mass. Thus, the mass, m, and volume, V,
are extensive quantities, whereas the pressure, p, and the absolute temperature, T ,
are intensive. For every extensive quantity (except the mass), we can introduce a
corresponding intensive quantity by dividing by the mass. For instance, the intensive
quantity that corresponds to the volume, V, is the specific volume, v = V/m, which is
simply the volume per unit mass (i.e., the volume of a unit mass of gas). Of course,
v = 1/, where is the mass density.

407
408 Theoretical Fluid Mechanics

A gas is said to be ideal if it obeys the law (Reif 1965),

p v = R T, (14.1)

where the constant R is termed the specific gas constant. This quantity is related to
the molar gas constant, R, introduced in Section 1.15, according to R = R/M, where
M is the molar massthat is, the mass of NA = 6.022 1023 gas molecules. Here,
NA is Avogadros number (Reif 1965).
For all gases, whether in mean motion, or not, there exists an internal energy
function, E, that is independent of the mean motion, and dependent only on the
variables of state, p, , and T . Moreover, this function is such that when a small
quantity of heat, q, is added to the system (Reif 1965),

q = dE + p dV. (14.2)

Thus, the quantity dE is the excess of the energy supplied over the mechanical work
done by the pressure. Of course, internal energy is an extensive quantity. The cor-
responding intensive quantity is the specific internal energy, E = E/m. In an ideal
gas, the specific internal energy, E, is a function of the absolute temperature, T , alone
(Reif 1965). It follows that
dE = k(T ) dT, (14.3)
and Equation (14.2) becomes

q = m (k dT + p dv). (14.4)

The heat capacity at constant volume of the gas is defined



q 
C V =  . (14.5)
dT v
Thus, Equation (14.4) implies that

C V = m k. (14.6)

It is clear that C V is an extensive quantity. The corresponding intensive quantity is


the specific heat capacity at constant volume,

C V
CV = = k. (14.7)
m
Finally, the molar specific heat capacity at constant volume, which is the heat capac-
ity of Avogadros number of gas molecules, takes the form

cV = M C V . (14.8)

It follows, from the preceding equations, that


cV
dE = CV dT = dT. (14.9)
M
One-Dimensional Compressible Inviscid Flow 409

In an ideal gas, cV is a constant, independent of the temperature (Reif 1965), in which


case the previous equation can be integrated to give [cf., Equation (1.83)]
cV
E = CV T = T. (14.10)
M
Equation (14.1) yields
v d p + p dv = R dT. (14.11)
Combining the previous equation with Equations (14.4) and (14.7), we obtain
! "
q = m (CV + R) T v d p . (14.12)

Now, the heat capacity at constant pressure of the gas is defined



q 
C p =  (14.13)
dT  p

Thus, Equation (14.12) implies that

C p = m (CV + R). (14.14)

It is clear that C p is an extensive quantity. The corresponding intensive quantity is


the specific heat capacity at constant pressure,

C p
Cp = = CV + R. (14.15)
m
Finally, the molar specific heat capacity at constant pressure, which is the heat ca-
pacity of Avogadros number of gas molecules, takes the form [cf., Equation (1.86)]

c p = M C p = cV + R. (14.16)

The molar specific heat capacity at constant pressure of an ideal gas is a constant,
independent of the temperature.
The enthalpy of a gas is defined (Reif 1965)

H = E + p V. (14.17)

This quantity is obviously extensive. The corresponding intensive quantity is the


specific enthalpy, H = H/m. It follows that, for an ideal gas,
cp
H = E + p v = E + R T = Cp T = T, (14.18)
M
where use has been made of Equations (14.1), (14.10), (14.15), and (14.16).
In the preceding analysis, we denoted a small quantity of heat by q, rather than
dQ, because there is, in general, no function Q of which q is an exact dierential.
However, we can write (Reif 1965)

q = T dS , (14.19)
410 Theoretical Fluid Mechanics

where dS is the dierential of an extensive function, S , called the entropy. To justify


the previous expression, note from Equations (14.1), (14.4) and (14.7) that



dT p dT dv
dS = m CV + dv = m CV +R . (14.20)
T T T v

However, Equation (14.1) yields

d p dv dT
+ = . (14.21)
p v T

Hence, we obtain


dv dp
dS = m C p + CV , (14.22)
v p

where use has been made of Equation (14.15). In an ideal gas, CV and C p are con-
stants, independent of T , so the previous expression can be integrated to give


p
S = m CV ln(p v ) = m CV ln , (14.23)

where
Cp
= (14.24)
CV

is the constant ratio of specific heats. Obviously, the specific entropy is defined



S p
S= = CV ln . (14.25)
m

Furthermore, it follows from Equations (14.2), (14.17), and (14.19) that

dE = p dV + T dS , (14.26)
dH = V d p + T dS , (14.27)

or, equivalently,

dE = p dv + T dS, (14.28)
dH = v d p + T dS. (14.29)
One-Dimensional Compressible Inviscid Flow 411

14.3 Isentropic Flow


In the limit of vanishing viscosity and heat conduction, the equations of compressible
ideal gas flow, introduced in Section 1.15, can be written

D
= v, (14.30)
Dt
Dv p
= , (14.31)
Dt


D p
= 0, (14.32)
Dt

where v is the flow velocity, and the potential energy per unit mass. It is clear from
Equations (14.25) and (14.32) that the specific entropy is constant along a streamline,
but not necessarily the same constant on dierent streamlines. Such flow is said to
be isentropic. From Equation (14.29), isentropic flow is characterized by

dp
dH = (14.33)

along a streamline. More generally, isentropic flow is characterized by p/ , /T 1 ,


and T /p 1 , constant along streamlines.

14.4 Sound Waves


Suppose that v = u(x, t) e x , = (x, t), p = p(x, t), and = 0 in Equations (14.30)
and (14.31). We obtain

u
+u + = 0, (14.34)
t x x
u u 1 p
+u + = 0. (14.35)
t x x

Equation (14.32) implies that the flow is isentropic. In other words, p/ = constant.
Thus, it follows that
p
= c2 , (14.36)
x x
where
dp p
c2 = . (14.37)
d
412 Theoretical Fluid Mechanics

Hence, Equations (14.34) and (14.35) become


u
+u + = 0, (14.38)
t x x
u u c 2
+u + = 0. (14.39)
t x x
Unfortunately, these equations are dicult to solve exactly, because of the presence
of nonlinear terms such as u u/x. (See Example 14.9.)
Let us write u(x, t) = u1 (x, t), (x, t) = 0 + 1 (x, t), and p(x, t) = p0 + p1 (x, t),
where quantities with the subscript 1 are assumed to be much smaller that corre-
sponding quantities with the subscript 0. Thus, we are now considering the evolution
of small-amplitude pressure and density perturbations in a stationary gas of uniform
density and pressure 0 and p0 , respectively. To first order in small quantities, the
previous two equations yield
1 1 u1
+ = 0, (14.40)
0 t x
u1 c02 1
+ = 0, (14.41)
t 0 x
where 
p0
c0 = . (14.42)
0
The general solution to Equations (14.40) and (14.41) is well known (Fitzpatrick
2013):
1 (x, t)
= F(x c0 t), (14.43)
0
u1 (x, t) = c0 F(x c0 t), (14.44)

where F(x) is an arbitrary function. According to this solution, a small-amplitude


density perturbation of arbitrary shape propagates through the gas, either in the pos-
itive or negative x-directions (corresponding to the upper and lower signs, respec-
tively), without changing shape, at the constant speed c0 . This type of perturbation
is known as a sound wave, and c0 is consequently referred to as the sound speed.
It is clear, from the previous analysis, that the general expression for the local
sound speed in an isentropic ideal gas of pressure p, density , and temperature T , is

p 
c= = R T, (14.45)

where use has been made of Equation (14.1). It follows that the speed of sound is
solely determined by the temperature of an ideal gas.
Let us now consider the eect of finite wave amplitude on the propagation of a
sound wave through an isentropic ideal gas. The previous analysis suggest that, in a
One-Dimensional Compressible Inviscid Flow 413

frame of reference
 moving with the local flow velocity, u, a sound wave propagates
at the speed c = p/. Thus, the net wave propagation speed is

c = u + c. (14.46)

Using the isentropic law, p/ = constant, to eliminate p from c = p/, we
deduce that
(1)/2

c = c0 . (14.47)
0
Now, Equations (14.43) and (14.44) suggest that, in the presence of a sound wave
propagating in the +x-direction, the general dierential relationship between the lo-
cal flow velocity and the density is
d
du = c . (14.48)

Making use of Equation (14.47), we can integrate the previous expression to give

(1)/2
2 c0
1 =
d 2
u= c = (c c0 ), (14.49)
0 1 0 1
or

1
c = c0 + u. (14.50)
2
It follows that

+1
c = c + u = c0 + u, (14.51)
2
or

(1)/2


+ 1



c = c0
1+ 1 . (14.52)
1 0

Thus, we conclude that the net wave propagation speed is higher than c0 in regions
of compression (i.e., > 0 ), and lower in regions of rarefaction (i.e., < 0 ). This
implies that a finite-amplitude sound wave distorts as it propagates in such a man-
ner that the regions of compression tend to catch up with the regions of rarefaction,
leading to a monotonic increase with time of the velocity and temperature gradients
in the former regions, and a monotonic decrease in the latter regions. Eventually, the
velocity and temperature gradients in the regions of compression become so large
that it is no longer valid to neglect the eects of viscosity and heat conduction, lead-
ing to a local breakdown of the isentropic gas law. Viscosity and heat conduction
prevent any further steepening of the velocity and temperature gradients in the re-
gions of compression, and the wave subsequently propagates without additional dis-
tortion. However, it has now eectively been transformed into a shock wave. (See
Section 14.8.)
It is clear, from the previous discussion, that, at finite amplitude, there is an im-
portant dierence in the behavior of compression and rarefaction waves propagating
414 Theoretical Fluid Mechanics

through a gas. A compression wave tends to steepen (i.e., the temperature and ve-
locity gradients tend to increase), and eventually becomes a shock wave, in which
case it is no longer isentropic. A rarefaction wave, on the other hand, always re-
mains isentropic, because it tends to flatten, thereby, further reducing the velocity
and temperature gradients. Hence, there are no rarefaction, or expansion, shocks.

14.5 Bernoullis Theorem


According to Bernoullis theorem, which was introduced in Section 4.3, the quantity
p/ + T is constant along a streamline in steady inviscid flow, where T is the total
energy per unit mass. For the case of a compressible fluid, T = E + (1/2) v 2 + .
Hence, we deduce that
p 1
E + + v2 + (14.53)
2

is constant along a streamline. In particular, for an ideal gas, we find that

1 2 1
H+ v + = Cp T + v 2 + (14.54)
2 2

is constant along a streamline, where use has been made of Equation (14.18).
Consider a solid object moving through an ideal gas. Generally speaking, there is
at least one stagnation point in front of the object, where the gas comes to rest relative
to it. At this point, the gas is adiabatically compressed, and there is an associated rise
in temperature. In the frame of reference in which the object is at rest, and the
gas a long way from it moves with constant speed and temperature, application of
Bernoullis theorem yields

1 2
Cp T + v = Cp T0, (14.55)
2

where T 0 is the stagnation point temperature, and T the temperature at some general
point where the flow speed is v. Thus, the total temperature rise due to adiabatic
compression is
v2
T = T 0 T = , (14.56)
2 Cp

where v and T are the asymptotic flow speed and temperature, respectively. It can
be seen that the stagnation temperature rise only depends on the velocity dierence
between the object and the gas, and is independent of the gass density, temperature,
or pressure. Note, however, that a lower molar mass implies a higher specific heat,
and, hence, a smaller stagnation temperature rise.
One-Dimensional Compressible Inviscid Flow 415

14.6 Mach Number


In an ideal gas, the local Mach number of the flow is defined (see Section 1.17)

v
Ma = , (14.57)
c

where c = R T is the local sound speed. [See Equation (14.45).] Setting v 2 =
Ma 2 c 2 in Equation (14.55), we obtain

1
T 1
= 1 + ( 1) Ma 2
, (14.58)
T0 2

where use has been made of Equations (14.15) and (14.24), which imply that



1
CV = R, (14.59)
1



CP = R. (14.60)
1

Incidentally, the relation (14.58) is valid for any streamline, because the stagna-
tion temperature, T 0 , can be defined, by means of Equation (14.55), even when the
streamline in question does not pass through a stagnation point. We can combine
Equation (14.58) with the isentropic relation, p / = constant along a streamline
(see Section 14.3), as well as the ideal gas law, (14.1), to give


/(1) /(1)
p T 1
= = 1+ ( 1) Ma 2 , (14.61)
p0 T0 2

1/(1) 1/(1)
T 1
= = 1+ ( 1) Ma 2 . (14.62)
0 T0 2

Here, p0 and 0 are the pressure and density, respectively, at the stagnation point. In
principle, the stagnation values, T 0 , p0 , and 0 , can be dierent on dierent stream-
lines. However, if a solid object moves through a homogeneous ideal gas that is
asymptotically at rest then the stagnation parameters become true constants, inde-
pendent of the streamline. Such flow is said to be homentropic.
A point where the speed of a steadily flowing ideal gas equals the local speed of
sound, v = c, is termed a sonic point. The sonic temperature, T 1 , pressure, p1 , and
density, 1 , are simply related to the stagnation values. In fact, setting Ma = 1 in
416 Theoretical Fluid Mechanics

Equations (14.58), (14.61), and (14.62), we obtain




T1 2
= , (14.63)
T0 +1

/(1)
p1 2
= , (14.64)
p0 +1

1/(1)
1 2
= . (14.65)
0 +1
Finally, if we combine Equations (14.58), and (14.61)(14.65), then we get

 1
T 1  2
= 1+ Ma 1 , (14.66)
T1 +1

 /(1)
p 1  2
= 1+ Ma 1 , (14.67)
p1 +1

 1/(1)
1  2
= 1+ Ma 1 . (14.68)
1 +1

14.7 Sonic Flow through a Nozzle


Consider an ideal gas flowing steadily through a straight nozzle with a slowly-varying
cross-sectional area, A = A(x), where x measures distance along the nozzle. The tem-
perature, T , density, , pressure, p, and normal velocity, u = v x , are all assumed to
be slowly-varying functions of x that are constant across any cross section. Thus,
in this quasi-one-dimensional problem, we are eectively disregarding the compara-
tively small components of the flow velocity orthogonal to the x-axis. Furthermore,
because all streamlines have the same parameter values in any cross section, the flow
is homentropic.
Of course, the net mass flow rate, Q, must be constant along the nozzle, so

Q = (x) A(x) u(x). (14.69)

In particular, A u = 1 A1 u1 , where the subscript 1 refers to the sonic point (assum-


ing that such a point exists). Using u/c = Ma and u1 /c1 = 1, we find that


A 1 u1 1 c1 1 1 $ T 1 %1/2 1
= = = , (14.70)
A1 u Ma c Ma T
where use has been made of Equation (14.45). It follows from Equations (14.66) and
(14.68) that

 (+1)/[2 (1)]
A 1 1  2
= 1+ Ma 1 . (14.71)
A1 Ma +1
One-Dimensional Compressible Inviscid Flow 417

0
0 1 2 3
Ma

Figure 14.1
Simple model of a de Laval nozzle for = 7/5. The solid, dashed, long-dashed,
dash-dotted, and dotted curves show A/A1 , p/p1 , /1 , T/T 1 , and u/u1 , respectively.

Moreover, the previous two equations can be combined with Equation (14.68) to give

 1/2
u 1  2
= Ma 1 + Ma 1 . (14.72)
u1 +1
Figure 14.1 shows A/A1 , p/p1 , /1 , T/T 1 , and u/u1 , plotted as functions of the
local Mach number, Ma, for an ideal gas with = 7/5. Here, use has been made
of Equations (14.71), (14.67), (14.68), (14.66), and (14.72), respectively. Inspecting
the curves, we can see, somewhat surprisingly, that the cross-sectional area function,
A/A1 , attains a minimum value when Ma = 1. In fact, the figure indicates that,
for subsonic flow (Ma < 1), a decreasing cross-sectional area of the nozzle in the
direction of the gas flow leads to an increasing flow speed, and decreasing pressure,
density, and temperature. However, for supersonic flow (Ma > 1), this behavior is
reversed, and an increasing cross-sectional area of the nozzle leads to an increasing
flow speed, and decreasing pressure, density, and temperature. We conclude that
local Mach number of gas flowing through a converging nozzle (i.e., a nozzle whose
cross-sectional area decreases monotonically in the direction of the gas flow) can
never exceed unity. Moreover, the maximum Mach number (i.e., unity) is achieved
on exit from the nozzle, where the cross-sectional area is smallest. On the other hand,
the local Mach number of gas flowing through a converging-diverging nozzle (i.e., a
nozzle whose cross-sectional area initially decreases in the direction of the gas flow,
attains a minimum value, and then increases) can exceed unity. For this to happen, the
418 Theoretical Fluid Mechanics

flow conditions must be arranged such that the sonic point corresponds precisely to
the narrowest point of the nozzle, which is generally known as the throat. In this case,
as the gas flows through the converging part of the nozzle, the local cross-sectional
area, A, travels down the left-hand, subsonic branch of the A/A1 curve shown in
Figure 14.1, while the flow speed, u, simultaneously increases. After passing through
the throat at the sonic speed, the gas flows through the diverging part of the nozzle,
and the cross-sectional area travels up the right-hand, supersonic branch of the A/A1
curve, while the flow speed continues to increase. The type of converging-diverging
nozzle just described is known as a de Laval nozzle, after its inventor, Gustaf de Laval
(18451913).
Consider a de Laval nozzle whose gas supply is derived from a large reservoir.
Assuming that the gas in the reservoir is essentially at rest, it follows that the temper-
ature, pressure, and density of the gas in the reservoir correspond to the stagnation
temperature, pressure, and densityT 0 , p0 , and 0 , respectively. Equation (14.63)
(14.65) then specify the temperature, pressure, and density of the gas at the throat of
the nozzleT 1, p1 , and 1 , respectivelyin terms of the temperature, pressure, and
density of the gas in the reservoir.
Suppose that a de Laval nozzle exhausts gas into the atmosphere, whose pressure
is patm . Now, for the case of incompressible flow, the pressure of the gas exhausted
from a nozzle, pe (say), must match the ambient pressure, patm . The reason for this is
that any mismatch between the exhaust and ambient pressures is instantly communi-
cated to the whole fluid by means of sound waves that travel infinitely fast (because
an incompressible fluid corresponds to the limit ). Of course, the sound speed
is finite in a compressible gas. However, in subsonic compressible flow, upstream
communication is still possible, because the local sound speed exceeds the local flow
speed. However, in supersonic compressible flow, upstream communication is im-
possible, because sound waves cannot catch up with the flow. Consequently, in the
case of a nozzle with a subsonic exhaust speed, we would generally expect the ex-
haust pressure to match the ambient pressure. However, in the case of a nozzle with
a supersonic exhaust speed, the exhaust pressure can be significantly dierent to the
ambient pressure.
Equation (14.61) yields
5

(1)/
p0
1,
2
Ma e = (14.73)
1 pe

where Ma e is the Mach number of the gas exhausted by the nozzle, pe the exhaust
pressure, and p0 the reservoir pressure. Suppose that M e < 1that is, the exhaust
speed of the gas is subsonic. In this case, we would expect pe = patm , so that
5


p0 (1)/
1.
2
Ma e = (14.74)
1 patm

It follows that the critical value of p0 /patm at which the exhaust speed becomes sonic
One-Dimensional Compressible Inviscid Flow 419

(i.e. Ma e = 1) is


/(1)
p0 +1
= . (14.75)
patm crit 2
Thus, in order for a de Laval nozzle to achieve supersonic exhaust speeds, p0 /patm
must exceed this critical value. For the case of a gas with = 7/5, we find that
(p0 /patm )crit = 1.89. Note that if p0 /patm does not exceed the critical value then,
as the gas flows through the converging part of the nozzle, its local cross-sectional
area, A, travels down the left-hand, subsonic branch of the A/A1 curve shown in
Figure 14.1. However, when the gas passes through the throat of the nozzle, the
area turns around, and then backtracks up the left-hand branch while the gas passes
through the diverging part of the nozzle. In this case, the Mach number never reaches
unity. In fact, in the converging part of the nozzle, the flow speed increases, while
the pressure, density, and temperature decrease. The flow speed attains its maximum,
subsonic value at the throat of the nozzle, while the pressure, density, and pressure
simultaneously attain minimum values. Finally, in the diverging part of the nozzle,
the flow speed decreases, while the pressure, density, and temperature increase.
Equations (14.64) and (14.67) yield



(1)/
1 2 p0
1+ (Ma e 1) =
2
. (14.76)
+1 + 1 pe
Moreover, Equations (14.45) and (14.63) imply that
5

2
u1 = R T 0 . (14.77)
+1
Hence, Equations (14.72), (14.73), (14.76), and (14.77) give the following expression
for the exhaust speed from a de Laval nozzle,
5

(1)/
2 pe
.
ue = R T 0 1 (14.78)
1 p0
Because pe cannot be negative, we deduce that there is an upper limit to the exhaust
speed: namely,

1/2 
2
ue max = c0 = 2 C p T 0 , (14.79)
1
where use has been made of Equation (14.60). Here, c0 is the reservoir sound speed.
If the exhaust pressure of a de Laval nozzle is higher than the ambient pressure,
pe > patm , then the gas is said to be under-expanded. In this case, a pattern of
standing shock waves, called shock diamonds, forms in the exhaust plume external
to the nozzle. On the other hand, if the exhaust pressure is lower than the ambient
pressure, pe < patm , then the gas is said to be over-expanded. In this case, a static
shock front forms inside the diverging part of the nozzle. (See Section 14.8.) As the
gas passes through the front, its speed drops abruptly from a supersonic to a subsonic
value, whereas the pressure, density, and temperature all increase abruptly. As the
subsonic gas flows through the remainder of the nozzle, its velocity decreases further.
420 Theoretical Fluid Mechanics

14.8 Normal Shocks


As previously described, there is an eective discontinuity in the flow speed, pres-
sure, density, and temperature, of the gas flowing through the diverging part of an
over-expanded Laval nozzle. This type of discontinuity is known as a normal shock.
Let us investigate the properties of such shocks.
Our fundamental equations are the mass conservation equation [see Equation
(14.30)],
D
= v, (14.80)
Dt
the momentum conservation equation [see Equation (14.31)],
Dv p
= , (14.81)
Dt
and the energy conservation equation [see Equation (1.75)],
DE p
= v. (14.82)
Dt
In writing the previous equations, we have neglected viscosity, heat conduction, and
potential energy. Note that we have used the more fundamental energy conservation
equation, (1.75), rather than Equation (14.32), because the latter equation incorpo-
rates the ideal gas law, and this law is not valid inside the shock, because the gas
there is not in thermodynamic equilibrium.
Consider a compressible gas flowing steadily down a duct of constant cross-
sectional area A. Let x measure distance along the duct. The gass temperature,
T (x), density, (x), pressure p(x), specific internal energy E(x), and normal velocity
u(x) = v x , are all assumed to be constant across any cross-section, and independent
of time. Let the shock be situated at x = 0. Suppose that in the region upstream of the
shock, x < 0, the temperature, density, pressure, specific internal energy, and flow
speed of the gas take the constant values T 1 , 1 , p1 , E1 , and u1 , respectively. Like-
wise, suppose that in the region downstream of the shock, x > 0, the temperature,
density, pressure, specific internal energy, and flow speed of the gas take the constant
values T 2 , 2 , p2 , E2 , and u2 , respectively. Of course, in the immediate vicinity of
the shock, T (x), (x), p(x), E(x), and u(x) are all rapidly-varying functions of x. We
wish to find the relationship between the upstream and downstream gas parameters.
We can achieve this goal by invoking conservation of mass, momentum, and energy
across the shockin other words, by making use of Equations (14.80)(14.82).
The mass continuity equation (14.71) yields
d du
u = , (14.83)
dx dx
which can be rearranged to give
d
( u) = 0. (14.84)
dx
One-Dimensional Compressible Inviscid Flow 421

The momentum conservation equation (14.72) yields

du 1 dp
u = , (14.85)
dx dx

which can be rearranged to give

du d p
u + = 0, (14.86)
dx dx
or
d  2 
u + p = 0, (14.87)
dx
where use has been made of Equation (14.84). Finally, the energy conservation equa-
tion (14.73) yields
dE p du
u = , (14.88)
dx dx
which can be rearranged to give

dE d dp
u + (p u) = u , (14.89)
dx dx dx
or

d 2 du 1 du 2 d 1
( u E + p u) = u = u = u ,
3
(14.90)
dx dx 2 dx dx 2
where use has been made of Equations (14.84) and (14.86). Thus, we obtain


d 1
u E + p u + u 3 = 0, (14.91)
dx 2

which reduces to

d p 1
E + + u2 = 0 (14.92)
dx 2
with the aid of Equation (14.84). The previous equation can also be written


d 1
H + u 2 = 0, (14.93)
dx 2

where H = E + p/ is the specific enthalpy of the gas. [See Equation (14.18).]


Integrating Equations (14.84), (14.87), and (14.93) across the shock, we obtain

1 u1 = 2 u2 , (14.94)
1 u12 + p1 = 2 u22 + p2 , (14.95)
1 2 1
H1 + u1 = H2 + u22 , (14.96)
2 2
where H1 = E1 + p1 /1 , et cetera. We shall assume that the gas upstream and
422 Theoretical Fluid Mechanics

downstream of the shock obeys the ideal gas law, and has the same ratio of specific
heats, . It follows that

p1,2 = R 1,2 T 1,2 , (14.97)





H1,2 = R T 1,2 . (14.98)
1

where use has been made of Equations (14.1), (14.18), and (14.60). Let Ma1 and Ma2
be the Mach numbers upstream and downstream of the shock, respectively. Thus,
u1,2
Ma1,2 =  , (14.99)
R T 1,2

where use has been made of Equations (14.45) and (14.57). Equations (14.96),
(14.98), and (14.99) can be combined to give

T 2 1 + (1/2) ( 1) Ma12
= . (14.100)
T 1 1 + (1/2) ( 1) Ma22

Equations (14.94) and (14.99) yield



2 u1 T 1 Ma1
= = , (14.101)
1 u2 T 2 Ma2
or 
p2 T 2 Ma1
= , (14.102)
p1 T 1 Ma2
where use has been made of Equation (14.97). Finally, Equations (14.95), (14.97),
and (14.99) give
p2 1 + Ma12
= . (14.103)
p1 1 + Ma22
Eliminating T 2 , 2 , and p2 between Equations (14.100), (14.101), and (14.103),
we obtain 1/2
2 + ( 1) Ma12
Ma2 = . (14.104)
1 + 2 Ma12
Note that 2
dMa22 1+
= 2
< 0. (14.105)
dMa12 1 + 2 Ma1
Moreover, it is easily verified that Ma2 = 1 when Ma1 = 1. Hence, we deduce that

Ma2  1 as Ma1  1. (14.106)

In other words, if the downstream flow is subsonic then the upstream flow is super-
sonic, and vice versa. Equations (14.1), (14.100), (14.101), (14.103), and (14.104)
One-Dimensional Compressible Inviscid Flow 423

can be combined to give

p2 1 + 2 M12
= , (14.107)
p1 +1
2 u1 ( + 1) Ma12
= = , (14.108)
1 u2 2 + ( 1) Ma12
T2 p2 1
= . (14.109)
T1 p1 2
The previous three equations completely describe the conditions downstream of the
shock in terms of the upstream conditions. We can rearrange these equations to give

1 + ( + 1) (p2 /p1 )
Ma12 = , (14.110)
2
2 u1 1 + ( + 1) (p2 /p1 )
= = , (14.111)
1 u2 + 1 + ( 1) (p2 /p1 )


T2 p2 + 1 + ( 1) (p2/p1 )
= . (14.112)
T1 p1 1 + ( + 1) (p2/p1 )

The latter two equations are known as the Rankine-Hugoniot relations.


According to Equation (14.25), the jump in specific entropy across the shock is


p2 1
S2 S1 = CV ln . (14.113)
p1 2

It follows from Equations (14.107) and (14.108) that


 
S2 S1 = CV ln G(Ma12 ) , (14.114)

where

1 + 2x 2 + ( 1) x
G(x) = . (14.115)
+1 ( + 1) x
It is easily demonstrated that

dG [2 + ( 1) x] 1
= 2 ( 1) (1 x) 2 0. (14.116)
dx [( + 1) x] +1
Moreover, G(1) = 1. Hence, we deduce that

S2 S1  0 as Ma1  1. (14.117)

Now, the second law of thermodynamics forbids a spontaneous decrease in the spe-
cific entropy of the gas as it passes through the shock (Reif 1965). In other words, the
second law of thermodynamics demands that S2 S1 0. It follows that Ma1 0.
However, it is easily seen from Equations (14.107)(14.109) that p2 /p1 = 2 /1 =
424 Theoretical Fluid Mechanics

u2 /u1 = T 2 /T 1 = 1 when Ma1 = 1. In other words, there is no shock (i.e., no dis-


continuity in the properties of the gas at x = 0) when the upstream Mach number
is exactly unity. Thus, we conclude that the only type of normal shock that is con-
sistent with the second law of thermodynamics is one in which the upstream flow
is supersonic, and the downstream flow subsonicthat is, Ma1 > 1 and Ma2 < 1.
It is apparent from Equations (14.107)(14.109) that if Ma1 > 1 then p2 /p1 > 1,
2 /1 > 1, and T 2 /T 1 > 1. In other words, the passage of the gas through the shock
front leads to both compression and heating.
The dimensionless parameter
p p2 p1
= (14.118)
p1 p1
is a convenient measure of shock strength. Thus, a shock is said to be weak if
p/p1  1, and strong if p/p1
1.
Consider a weak shock. According to Equations (14.111), (14.112), (14.114),
and (14.116),
1 p
, (14.119)
1 p1


T 1 p
, (14.120)
T1 p1

3
S ( + 1) p
, (14.121)
R 12 2 p1
where = 2 1 , et cetera, and use has been made of Equation (14.59). It can be
seen that the flow across the shock is isentropic (i.e., p/ and T /p 1 are constant)
to first order in the shock strength. This is the case because the increase in specific
entropy across the shock is only third order in the shock strength.
Consider a strong shock. It can be seen from Equations (14.111) and (14.112)
that
2 +1
, (14.122)
1 1


T2 1 p
, (14.123)
T1 + 1 p1
as p/p1 . Clearly, there is a finite limit to the degree of compression of gas
passing through a strong shock. Thus, in the case of a strong shock, the large increase
in the pressure across the shock front is predominately caused by a large increase in
the temperature. In other words, the gas flowing through a strong shock is subject
moderate compression, but intense heating.
Finally, let us transform to a frame of reference that is co-moving with the gas up-
stream of the shock. In this reference frame, the shock appears to propagate through
a stationary gas of pressure, density, and temperature, p1 , 1 , and T 1 , respectively,
at the speed u1 . In other words, in the new reference frame, our stationary shock is
One-Dimensional Compressible Inviscid Flow 425

transformed into a shock wave. (See Section 14.4.) The pressure, density, and tem-
perature behind the wave are p2 , 2 , and T 2 , respectively. Also, the gas behind the
wave follows the shock front at the speed u1 u2 . We can regard Ma1 as the Mach
number of the shock wave in the unperturbed gas. It follows from Equation (14.110)
that

+ 1 p
Ma1 = 1 +
2
. (14.124)
2 p1
Thus, for a weak shock wave (i.e., p/p1  1),

Ma1 1. (14.125)

In other words, a weak shock wave propagates through the unperturbed gas at the
local sound speed. Indeed, such a wave is essentially indistinguishable from a con-
ventional sound wave. On the other hand, for a strong shock wave (i.e., p/p1
1),
5

+ 1 p
Ma1 . (14.126)
2 p1

We conclude that a strong shock wave propagates through the unperturbed gas at a
speed that greatly exceeds the local sound speed.

14.9 Piston-Generated Shock Wave


Consider the situation illustrated in Figure 14.2 in which, at t = 0, a tight-fitting
piston is suddenly pushed into a stationary gas, contained in a uniform tube, at the
steady speed V p , generating a shock front that propagates away from the piston, and
into the gas, at the speed V s . Suppose that the piston is located at x = 0 at t = 0,
and moves in the +x-direction. The gas can be divided into two regions. In region 1,
which lies to the right of the shock wave (i.e., x > V s t), the gas remains undisturbed.
Hence, its velocity is u1 = 0. Let its pressure, density, and temperature be p1 , 1 ,
and T 1 , respectively. In region 2, which lies between the piston and the shock wave
(i.e., V p t < x < V s t), the gas is co-moving with the piston, so its velocity is u2 = V p .
Let its pressure, density, and temperature, be p2 , 2 , and T 2 , respectively. Thus, the
system is formally the same as that discussed in the final paragraph of the previous
section, provided that we make the identifications V s = u1 and V p = u1 u2 .
Let = (p2 p1 )/p1 be our conventional measure of shock strength. It follows
from Equations (14.111) and (14.112) that
2 2 + ( + 1)
= , (14.127)
1 2 + ( 1)

T2 2 + ( 1)
= (1 + ) , (14.128)
T1 2 + ( + 1)
426 Theoretical Fluid Mechanics

t
x = Vp t x = Vs t

ath
th
pa
on p
k
oc
sh
pist

Vp Vs
region 2 region 1

Figure 14.2
A piston-generated shock wave.

where is the gass ratio of specific heats. We can define the Mach number of the
shock wave as Ma s = V s /c1 = u1 /c1 , where c1 is the sound speed in the unperturbed
gas. According to Equation (14.110),


1/2
+1
Ma s = 1 + . (14.129)
2

Finally, we can define the Mach number of the piston as Ma p = V p /c1 . Thus,



u1 u2 u1 u2 1
Ma p = = 1 = Ma s 1 , (14.130)
c1 c1 u1 2

where use has been made of Equation (14.111). Equations (14.127) and (14.129) can
be combined with the previous equation to give

1 1 +1
= 2+ . (14.131)
2 2
Ma p 2

Thus, Equations (14.127)(14.129), and (14.131), fully determine the parameters of


the shocked gas, as well as the speed of the shock front, in terms of the parameters
of the unperturbed gas and the piston speed. In particular, for a weak shock (i.e.,
One-Dimensional Compressible Inviscid Flow 427

 1),

Ma p , (14.132)


+1
Ma s 1 + Ma p , (14.133)
4

whereas for a strong shock (i.e.,


1),

( + 1) Ma p2
, (14.134)
2


+1
Ma s Ma p . (14.135)
2

Thus, a weak shock is associated with a subsonic piston (i.e., Ma p  1), whereas a
strong shock is associated with a supersonic piston (i.e., Ma p
1).

14.10 Piston-Generated Expansion Wave


Consider the situation illustrated in Figure 14.3 in which, at t = 0, a tight-fitting
piston is suddenly withdrawn from a stationary gas, contained in a uniform tube,
at the steady speed V p , generating an expansion wave that propagates away from
the piston, and into the gas. Suppose that the piston is located at x = 0 at t = 0,
and moves in the x-direction. As described in Section 14.4, the expansion wave is
isentropic.
The sudden withdrawal of the piston creates a step-function change in the flow
speed, pressure, density, and temperature, of the gas inside the tube, such that the
immediately adjacent gas moves with the piston. As described in Section 14.4, this
step function flattens as the expansion wave begins to propagate. Locally, within
the wave, the disturbance travels at the sound speed. Because the temperature varies
across the wave (which starts out as a step in this variable), the local speed of distur-
bance propagation also varies across the wave. On the edge of the wave adjacent to
the region of undisturbed gas, which we shall denote region 1, the temperature and
propagation speed are greatest. On the opposite edge, the temperature and propaga-
tion speed are least. Let the region lying between the piston and the left edge of the
expansion wave be denoted region 2. In region 1, the flow speed, pressure, density,
and temperature, of the gas are u1 = 0, p1 , 1 , and T 1 , respectively. On the other
hand, in region 2, the flow speed, pressure, density, and temperature, of the gas are
u2 = V p (i.e., the gas in region 2 is co-moving with the piston), p2 , 2 , and T 2 ,
respectively. As shown in Figure 14.3, the expansion wave expands, or fans out, as
it propagates. In fact, the x-t plot is a fan of constant sonic speed lines that show the
development of the wave. These lines are called characteristics, and follow the path
of local isentropic disturbances. The absolute speed of the disturbance is the sum of
428 Theoretical Fluid Mechanics

t x = [c1 (1/2)( + 1) Vp] t

x = c1 t
x = Vp t
expansion front

pist
on p
ath

Vp c1
region 2 region 1

Figure 14.3
A piston-generated expansion wave.

the local sonic speed and the local gas flow speed. The terminating characteristic on
the right has slope dx/dt = c1 , where c1 is the region-1 sound speed. The terminating
characteristic on the left has the slope dx/dt = c2 V p , where c2 is the region-2 sound
speed. The latter slope may be either positive or negative, depending on whether c2
is greater than, or less than, V p .
Making use of isentropic relationships (see Section 14.3), we can explicitly eval-
uate the structure of the fan as follows. In terms of c1 , the local sonic speed in the
fan can be written [cf., Equation (14.47)]

(1)/2

c(x) = c1 . (14.136)
1
Moreover, the relationship between the gas speed, u(x), and the density in the wave
is [cf., Equation (14.48)]
d
du = c . (14.137)

The previous two expressions yield [cf., Equation (14.50)]
1
c(x) = c1 + ( 1) u(x). (14.138)
2
Hence, the absolute disturbance speed in the fan is [cf., Equation (14.51)]
1
c (x) = c(x) + u(x) = c1 + ( + 1) u(x). (14.139)
2
One-Dimensional Compressible Inviscid Flow 429

In particular, the terminating characteristic on the left (for which u = V p ) has the
slope
dx 1
= c2 V p = c1 ( + 1) V p . (14.140)
dt 2
It follows that
1
c2 = c1 ( 1) V p . (14.141)
2
Standard isentropic relations (see Section 14.3) then yield the density and pressure
changes across the fan:

2 1 V p 2/(1)
= 1 ( 1) , (14.142)
1 2 c1

p2 1 V p 2 /(1)
= 1 ( 1) . (14.143)
p1 2 c1
Here, use has been made of the fact that c T 1/2 . At the critical piston withdrawal
speed V p = [2/( 1)] c1, the pressure and density in region 2 are both reduced to
zero, and the terminating characteristic on the left co-moves with the piston. Any
further increase in the withdrawal speed makes no dierence to the flow.
The equation of motion of the right edge of the expansion wave is
x1 = c1 t. (14.144)

Likewise, the equation of motion of the left edge is



1
x2 = c1 ( + 1) V p t. (14.145)
2
Finally, the equation of motion of a general point in the expansion wave is

1
x = c1 + ( + 1) u t. (14.146)
2
The previous three equations can be combined to give


x x2
u(x) = 1 + Vp (14.147)
x1 x2
for x2 x x1 . In other words, the flow speed varies linearly with x inside the
expansion wave. It follows from Equations (14.138) and (14.141) that
c(x) c2 x x2
= (14.148)
c1 c2 x1 x2
for x2 x x1 . Thus, the sound speed also varies linearly with x inside the expan-
sion wave. However, according to Equation (14.136),

(1)/2 2(1)/2 x x2
= (14.149)
1(1)/2 2(1)/2 x1 x2
430 Theoretical Fluid Mechanics

for x2 x x1 . Furthermore, standard isentropic relations yield


(1)/2
p (1)/2 p2 x x2
= (14.150)
p1(1)/2 p2(1)/2 x1 x2

for x2 x x1 . Thus, neither the density nor the pressure vary linearly with x inside
the expansion wave.

14.11 Exercises
14.1 Prove the following useful theorems regarding partial derivatives:

)

X Y
=1 ,
Y Z X Z


)

X X Y
= ,
Y Z W Z W Z


)

X Z Z
=
Y Z Y X X Y

14.2 (a) The specific internal energy of a (not necessarily ideal) gas is defined by

dE = T dS p dv.

Demonstrate that



E E
T= , p= ,
S v v S

and


T p
= .
v S S v
(b) The specific enthalpy of a gas is defined by

H = E + p v.

Demonstrate that



H H
T= , v= ,
S p p S

and


T v
= .
p S S p
One-Dimensional Compressible Inviscid Flow 431

(c) The specific Helmholtz free energy of a gas is defined by

F = E T S.

Demonstrate that



F F
S= , p= ,
T v v T

and


S p
= .
v T T v

(d) The specific Gibbs free energy of a gas is defined by

G = H T S.

Demonstrate that



G G
S= , v= ,
T p p T

and


S v
= .
p T T p

14.3 Demonstrate that the specific heat at constant volume of a (not necessarily
ideal) gas can be written



E S
cv = =T .
T v T v

Likewise, show that the specific heat at constant pressure takes the form



H S
cp = =T .
T v T p

14.4 The quantities




1 v
= ,
v T p


1 v
T = ,
v p T


1 v
S = ,
v p S
432 Theoretical Fluid Mechanics

are known as the coecient of thermal expansion, the isothermal compress-


ibility, and the adiabatic compressibility, respectively. Demonstrate that for a
(not necessarily ideal) gas,

T v 2
c p cv = ,
T
T v 2
T S = . (14.151)
cp

Hence, deduce that


c p T
= .
cv S
Show that for the special case of an ideal gas, = 1/T , T = 1/p, and S =
1/( p). Hence, obtain the following standard results for an ideal gas:
cp
= ,
cv
c p cv = R,



cp = R,
1

14.5 Show that for an ideal gas



2 1/2
v
Ma = 1 1 ( 1) v ,
c0 2 c0

2
T 1 v
= 1 ( 1) ,
T0 2 c0

2 /(1)
p 1 v
= 1 ( 1) ,
p0 2 c0

2 1/(1)
1 v
= 1 ( 1) ,
0 2 c0

where Ma is the Mach number, v the flow speed, and T 0 , p0 , 0 , and c0 , are the
temperature, pressure, density, and sound speed, respectively, at the stagnation
point.
14.6 Consider the flow of an isentropic ideal gas down a straight nozzle with a
slowly-varying cross-sectional area, A(x), where x measures distance along
the nozzle. Let u(x), (x), c(x), and Ma(x) be the flow speed, density, sonic
speed, and Mach number, respectively. Demonstrate that
d
c2 + u du = 0,

One-Dimensional Compressible Inviscid Flow 433

diaphragm

p1 p4 p4 > p1

shock contact surface expansion front

|Vs| |u2| c4

region 1 region 2 region 3 region 4

Figure 14.4
A shock tube.

and
d dA du
+ + = 0.
A u
Hence, show that
dA du
= (Ma 2 1).
A u
Deduce that the throat of the nozzle (where A attains its minimum value) either
corresponds to a sonic point (i.e, Ma = 1), or a point of maximum or minimum
flow speed. Finally, demonstrate that

dMa 1 + (1/2) ( 1) Ma 2 dA
= .
Ma 1 Ma 2 A

14.7 As indicated in Figure 14.4, a shock tube is a tube of uniform cross section that
is divided by a diaphragm into two chambers that contain dierent gases at dif-
ferent pressures. Let x measure distance along the tube, and let the diaphragm
be located at x = 0. Suppose that the chamber to the left of the diaphragm
(which lies at x < 0) is filled with stationary gas of pressure, density, temper-
ature, and ratio of specific heats, p1 , 1 , T 1 , and 1 , respectively. Likewise,
suppose that the chamber to the right of the diaphragm (which lies at x > 0) is
filled with stationary gas of pressure, density, temperature, and ratio of specific
heats, p4 , 4 , T 4 , and 4 , respectively. It is assumed that p4 > p1 . At t = 0,
the diaphragm is ruptured. As indicated in the figure, a shock wave subse-
quently travels to the left with speed |V s |, and an expansion wave to the right
with speed c4 . The so-called contact surface marks the boundary between
the two dierent gases that were originally on either side of the diaphragm.
Neglecting diusion, the gases do not mix, but are permanently separated by
the contact surface. On either side of the contact surface, which moves to the
left with speed |u2 |, the temperatures and densities can be dierent, but the
pressures and flow velocities must be the same. We can divide the gas in the
434 Theoretical Fluid Mechanics

tube into four regions. Regions 1 lies to the left of the shock wave. Region 2
lies between the shock wave and the contact surface. Region 3 lies between
the contact surface and the expansion wave. Region 4 lies to the right of the
expansion wave. Thus, we expect the flow velocity, pressure, density, tem-
perature, and ratio of heats to be u1 = 0, p1 , 1 , T 1 , and 1 , respectively, in
Region 1; u2 , p2 , 2 , T 2 , and 1 , respectively in Region 2; u2 , p2 , 3 , T 3 , and
4 , respectively in Region 3; and u4 = 0, p4 , 4 , T 4 , and 4 , respectively, in
Region 1. Note that u2 and V s are negative. Demonstrate that

1/2
p2 2/1
|u2 | = c1 1 ,
p1 2 1 + (1 + 1) (p2 /p1 1)

and
(4 1)/4
2 c4 p2
,
|u2 | = 1
4 1 p4
where c1 and c4 are the sound speeds in Regions 1 and 4, respectively. Hence,
obtain
2 4 /(4 1)
p4 p2 (4 1) (c1 /c4 ) (p2 /p1 1)
= 1   .
p1 p1 2 1 2 1 + (1 + 1) (p2/p1 1)

This expression give the shock strength, p2 /p1 , implicitly as a function of


the diaphragm pressure ratio, p4 /p1 . All other quantities of interest can be
expressed in terms of the shock strength. Show that


1/2
1 + 1 p2
|V s | = c1 1 + 1 ,
2 1 p1
2 2 1 + (1 + 1) (p2 /p1 1)
= ,
1 2 1 + (1 1) (p2 /p1 1)

T2 p2 2 1 + (1 1) (p2 /p1 1)
= ,
T1 p1 2 1 + (1 + 1) (p2 /p1 1)

1/4
3 p2 /p1
= ,
4 p4 /p1

and

( 1)/4
T3 p2 /p1 4
= .
T4 p4 /p1

14.8 Show that the maximum shock strength and shock speed attainable in a (uni-
form) shock tube, in the limit p4 /p1 , are
 
p 2 c4 2 1 (1 + 1)
= ,
p 1 c1 (4 1)
One-Dimensional Compressible Inviscid Flow 435

and

1 + 1
|V s | = c4 ,
4 1
respectively. Show, further, that the contact surface moves at the speed


2
|u2 | = c4 .
4 2

14.9 Show that Equations (14.38) and (14.39) can be written in the form


2 c c u
+u +c = 0,
1 t x x
u u 2 a c
+u + = 0,
t x 1 x
where c is the sound speed. By adding and subtracting the previous equations,
obtain


2c
+ (u + c) u+ = 0,
t x 1


2c
+ (u c) u = 0.
t x 1

These equations indicate that the quantities P = u + 2 c/( 1) and Q =


u 2 c/( 1) are constant on curves that have the slope dx/dt = u + c and
dx/dt = u c, respectively. These curves are called characteristics, and P
and Q are known as Riemann invariants. In situations in which all the Q
characteristics originate from regions where the gas is at rest, we expect Q to
be constant throughout the gas. Deduce that [cf., Equation (14.50)]


1
c = c0 + u,
2

where c0 is the stagnation sound speed. Show that




+ (u + c) (u + c) = 0.
t x

Hence, conclude that the P characteristics are straight-lines.


14.10 Consider a one-dimensional sound wave propagating through an ideal gas
whose unperturbed sound speed is c0 . At time t = 0, the velocity perturba-
tion, u(x, t), due to the wave, has the form u0 sin(k x), where u0 and k are both
positive. Demonstrate that shocks form after a time
2
t
( + 1) k u0
436 Theoretical Fluid Mechanics

has elapsed. [Hint: Shocks are associated with the crossing of dierent P
characteristics.] Show that in the time interval t = 0 to t = t a local maximum
of u(x, t) travels a distance

2 c0 1
1+ ,
( + 1) u0 k

14.11 An ideal gas is initially at rest in a uniform tube, and occupies the region to
the right of a tight-fitting piston whose position is X(t) = (1/2) a t 2 for t > 0.
Here, a > 0. Show that a shock first forms at t = t and x = x , where


2 c0
t = ,
+1 a

2
2 c0
x = .
+1 a

Here, c0 is the stagnation sound speed, and the ratio of specific heats.
15
Two-Dimensional Compressible Inviscid Flow

15.1 Introduction
This chapter investigates two-dimensional, supersonic, compressible, inviscid flow.
Flow is said to be two-dimensional when the fluid properties only depend on two
Cartesian coordinates. More information on this subject can be found in Liepmann
& Roshko 1957, Hughes & Brighton 1999, Emanuel 2000, and Anderson 2003.

15.2 Oblique Shocks


Consider a normal shock in which the flow upstream and downstream of the shock
front is parallel to the x-axis, and front itself lies in the y-z plane. (See Section 14.8.)
Suppose that u1 and u2 are the upstream and downstream flow speeds, respectively.
Let us now view this shock in a frame of reference that moves with respect to our
original frame at the constant velocity v ey . As illustrated in Figure 15.1, viewing
the shock in the new reference frame has the eect of adding a component of velocity
of magnitude v, directed parallel to the y-axis, to both the upstream and the down-
stream flow. The resultant upstream velocity is now of magnitude w1 = (u12 + v 2 )1/2 ,
and subtends an angle = tan1 (u1 /v), known as the wave angle, with the shock
front. It is evident that may be adjusted to any value via a suitable choice of v.
Now, because u2 is not the same as u1 , the inclination of the downstream flow to the
shock front is dierent to that of the upstream flow. In other words, the direction
of the flow turns abruptly as it passes through the shock. Because u2 is always less
than u1 (see Section 14.8), the deviation is always towards the shock front. In other
words, the deflection angle, , defined in Figure 15.1, is positive.
The relationship between the conditions upstream and downstream of the shock is
easily obtained from the analysis of Section 14.8, because viewing a normal shock in
a moving frame of reference does not change the relationship between the upstream
and downstream conditions in the original reference frame. The only modification is
that the upstream Mach number is now defined as Ma1 = w1 /c1 , rather than u1 /c1 .
Here, c1 is the upstream sound speed. Thus, given that u1 = w1 sin , we simply
need to make the transformation Ma1 Ma1 sin in the previous analysis. Hence,

437
438 Theoretical Fluid Mechanics

2
w
u1 x
u2
v w1

Figure 15.1
An oblique shock.

Equations (14.107), (14.108), (14.109), and (14.113), yield

p2 1 + 2 Ma12 sin2
= , (15.1)
p1 +1
2 u1 ( + 1) Ma12 sin2
= = , (15.2)
1 u2 2 + ( 1) Ma12 sin2
  
T2 1 + 2 Ma12 sin2 2 + ( 1) Ma12 sin2
= , (15.3)
T1 ( + 1) 2 Ma12 sin2


2 1 1

1
S2 S1
1 + 2 Ma1 sin ( + 1) Ma1 sin
2 2 2

= ln





, (15.4)
R
+ 1 2 + ( 1) Ma1 sin
2 2

respectively. Here, p1 , 1 , T 1 , and S1 are the upstream pressure, density, temper-


ature, and specific entropy, respectively, p2 , 2 , T 2 , and S2 are the corresponding
downstream quantities, is the ratio of specific heats, R the specific gas constant,
and use has been made of Equation (14.59). As before (see Section 14.8), the second
law of thermodynamics demands that S2 S1 0, which implies that Ma1 sin 1.
This sets a minimum inclination of the upstream flow to the shock front for a given
upstream Mach number. The maximum inclination is, of course, = /2. Thus,

, (15.5)
2
Two-Dimensional Compressible Inviscid Flow 439

where

1 1
= sin (15.6)
Ma1
is termed the Mach angle. The downstream Mach number can be calculated by not-
ing that Ma2 = w2 /c2 , rather that u2 /c2 , where c2 is the downstream sound speed,
and u2 = w2 sin( ). In other words, we simply need to make the transforma-
tion Ma2 Ma2 sin( ) in the previous analysis. Thus, it follows from Equa-
tion (14.104) that

2 + ( 1) Ma12 sin2
Ma22 sin2 ( ) = . (15.7)
1 + 2 Ma12 sin2

According to Figure 15.1,


u1
tan = , (15.8)
v
and
u2
tan( ) = . (15.9)
v
Eliminating v, and then making use of Equation (15.2), we obtain

tan( ) u2 1 2 + ( 1) Ma12 sin2


= = = . (15.10)
tan u1 2 ( + 1) Ma12 sin2

Now,
tan tan
tan( ) , (15.11)
1 + tan tan
so Equation (15.10) can be rearranged to give
 
2 cot Ma12 sin2 1
tan = ! ". (15.12)
2 + Ma12 + cos(2 )

Here, use has been made of the identity cos(2 ) cos2 sin2 . The previous
expression implies that = 0 at = and = /2, which are the limits of the range
of allowed values for defined in Equation (15.5). Within this range, is positive,
and, must, therefore, have a maximum value, max . This is illustrated in Figure 15.2,
where the relationship between and , for an ideal gas with = 1.4, is plotted for
various values of Ma1 .
If < max then, for each value of and Ma1 , there are two possible solutions,
corresponding to two dierent values of . The larger value of corresponds to the
stronger shock [because, according to Equation (15.1), the shock strength, p2 /p1 1,
is a monotonically increasing function of ].
Also shown in Figure 15.2 is the locus of solutions for which Ma2 = 1. In the
solution with the stronger shock, the downstream flow always becomes subsonic. On
the other hand, in the solution with the weaker shock, the downstream flow remains
supersonic, except for a small range of values of that are slightly smaller that max .
440 Theoretical Fluid Mechanics

40

30
( )

20

10

0
0 10 20 30 40 50 60 70 80 90
( )

Figure 15.2
Oblique shock solutions for = 1.4. The solid curves show solutions for Ma1 = 1.2,
1.4, 1.6, 2, 3, 4, 5, 10, and , in order from the innermost to the outermost curve.
The dashed curve shows the locus of solutions for which = max . The dash-dotted
curve shows the locus of solutions for which Ma2 = 1. Solutions to the left and right
of the dash-dotted curve correspond to Ma2 > 1 and Ma2 < 1, respectively.

Equation (15.10) can be rearranged to give




1 + 1 tan( ) 1
= , (15.13)
Ma12 sin2 2 tan 2

which can be further reduced to




+1 sin sin
Ma12 sin 1 =
2
Ma12 . (15.14)
2 cos( )

For small deflection angles, , the previous expression can be approximated by



+1
Ma12 sin 1
2
Ma1 tan .
2
(15.15)
2
Two-Dimensional Compressible Inviscid Flow 441

shock front

Ma2
Ma1


wall

wall

Figure 15.3
Supersonic flow in concave corner.

15.3 Supersonic Flow in Corner or over Wedge


In practice, what determines and ? In other words, how are oblique shocks gen-
erated? In non-viscous flow, any streamline can be replaced by a solid boundary.
Thus, the oblique shock flow, described in the previous section, provides the solution
to the problem of supersonic flow in a concave corner, as illustrated in Figure 15.3.
For a given values of Ma1 and , the values of and Ma2 are determined. [See Exer-
cise 15.1 and Equation (15.7).] For the present, we shall only consider cases in which
Ma2 > 1. This restricts the oblique shock solution to the weaker of the two possible
solutions. We also require that be less than max . The stronger shock solution, as
well as cases in which > max , will be discussed later. (See Section 15.7.)
By symmetry, supersonic flow in a concave corner that turns though an angle
is equivalent to one half of the flow pattern that results when a supersonic fluid
is normally incident on a symmetric wedge of nose angle 2 . This is illustrated in
Figure 15.4. It is clear, from the figure, that the presence of the wedge only aects
the flow in the region that lies to the right of the two shock fronts that are attached to
the wedge apex, P. In particular, the presence of the apex at P only aects the flow
in the region that lies to the right of the shock fronts.
The part of Figure 15.2 that we are presently considering (i.e., Ma2 > 1) is such
that a decrease in the wedge angle, , corresponds to a decrease in the wave angle,
. When decreases to zero (causing the wedge to disappear), decreases to the
limiting value . Moreover, the shock strength (which parameterizes the jump in
quantities across the shock front) becomes zero. (See Section 15.4.) In fact, there
is no disturbance in the flow in the limit 0. Thus, in Figure 15.4, there is no
longer anything unique about the point P; indeed, this point might correspond to any
point in the flow. The angle is simply a characteristic angle associated with the
442 Theoretical Fluid Mechanics

shock front

Ma2
Ma1

wedge

P
Ma1

Ma1
Ma2

shock front

Figure 15.4
Supersonic flow over symmetric wedge.

local Mach number of the flow, according to the relation (15.6). As we have already
mentioned, is known as the Mach angle. Note that 0 /2 for Ma1 1.

The lines of inclination (with respect to the upstream directed streamline) that
may be drawn at any point in the flow are known as Mach lines. At a general point,
P, there are always two lines that intersect the local streamline at the angle . (In
three-dimensional flow, the Mach lines define a conical surface with apex P, known
as a Mach cone.) Thus, a two-dimensional supersonic flow pattern is associated with
two families of Mach lines. These are conventionally distinguished by the labels
(+) and (). Those lying in the (+) set run to the right of the (upstream directed)
streamline, whereas those in the () set run to the left. (See Figure 15.5.) Mach lines
are also sometimes called characteristics. It is clear, from our previous discussion,
that the flow conditions at point P can only aect those at some other point Q if the
latter point lies between the (+) and () Mach lines passing through point P. (See
Figure 15.5, as well as Exercise 15.2.) Hence, we deduce that, unlike subsonic flow,
there is no upstream influence in supersonic flow.

Oblique shocks can also be distinguished by the labels (+) and (), according to
which set of characteristics they asymptote to in the limit of zero shock strength.
Two-Dimensional Compressible Inviscid Flow 443

()

streamline

Q

P
(+)
Figure 15.5
Right- and left-running Mach lines [labeled (+) and (), respectively] emanating
from an arbitrary point, P, in a supersonic flow field.

15.4 Weak Oblique Shocks


For small deflection angles, , the oblique shock equations reduce to relatively simple
expressions. In fact, we already saw in Equation (15.15) that if  1 then


+1
Ma1 sin 1
2 2
Ma1 tan .
2
(15.16)
2
It follows that is either close to or /2, depending on whether Ma2 > 1 or
Ma2 < 1. (See Figure 15.2.) For the present, we are only considering the former case
(Ma2 > 1), for which we may use the approximation [see Equation (15.6)]
1
tan tan =  . (15.17)
Ma12 1

Equation (15.16) then reduces to





+ 1 Ma1 2

Ma1 sin 1
2 2
 . (15.18)
2
Ma1 1
2

This is the basic relation that is needed to obtain all other approximate expressions,
because these expressions only depend on the normal component of the upstream
flow, Ma1 sin . Thus, Equation (15.1) yields


p Ma12
 , (15.19)
p1
Ma12 1
444 Theoretical Fluid Mechanics

where p = p2 p1 . It is apparent that the shock strength, p/p1 , is directly propor-


tional to the deflection angle, .
The changes in the other flow quantities, except the specific entropy, across the
shock front are also directly proportional to . (See Exercise 15.3.) The change in the
entropy, on the other hand, is proportional to the third power of the shock strength
(see Section 14.8), and, hence, to the third power of the deflection angle.
To explicitly find the deviation of the wave angle, , from the Mach angle, , we
write
= + , (15.20)
where   . Thus, we have

sin = sin( + ) sin +  cos . (15.21)

However, by definition, sin = 1/Ma1 , and cos = (Ma12 1)1/2 /Ma1 [see Equa-
tion (15.6)], so we obtain

Ma1 sin 1 +  Ma12 1, (15.22)

or 
Ma12 sin 1 + 2  Ma12 1.
2
(15.23)
Finally, comparison with Equation (15.18) reveals that


+ 1 Ma12
 . (15.24)
4 Ma12 1

In other words, for a finite deflection angle, , the wave angle, , diers from the
Mach angle, , by an amount, , that is of the same order of magnitude as .
The change in flow speed across the shock front is obtained from the ratio

w22 u22 + v 2 (u2 /v)2 + 1 tan2 ( ) + 1 cos2


= = = = , (15.25)
w12 u12 + v2 (u1 /v) + 1
2 tan + 1
2 cos2 ( )

where use has been made of Equations (15.8) and (15.9). However,


Ma 2
1
2 
cos = 1 sin
2 2 1
1  , (15.26)
Ma1
Ma12 1
2

where use has been made of Equation (15.23). Similarly,




Ma 2
1
2 ( )
cos2 ( ) 1
1  . (15.27)
Ma1
Ma12 1
2
Two-Dimensional Compressible Inviscid Flow 445

shock weak shocks


(a) (b)

Ma1 Ma1


wall

Mach lines
(c)

Ma1

Figure 15.6
Compression of supersonic flow that is turned through an angle . Case (a) shows a
single turn through an angle . Case (b) shows the turn subdivided into a number of
smaller turns of magnitude . Case (c) shows a smooth continuous turn through an
angle .

Hence, to first order in ,


w2
1  , (15.28)
w1
Ma12 1
w
 , (15.29)
w1
Ma12 1

where w = w2 w1 .

15.5 Supersonic Compression by Turning


A shock front increases the pressure and density of the fluid that passes through it.
In other words, the front compresses the flow. A simple method for compressing a
supersonic flow is to turn it through an oblique shock, by deflecting the wall through
an angle , as shown in Figure 15.6(a).
The wall deflection can be subdivided into several segments, which make smaller
446 Theoretical Fluid Mechanics

corners of angle , as shown in Figure 15.6(b). Compression then occurs through


successive oblique shocks. These shocks divide the flow field close to the wall into
segments of uniform flow. Further out, the shocks must intersect one another, be-
cause they are convergent. However, for the present, we shall only consider the flow
close to the wall. In this region, each segment of the flow is independent of the fol-
lowing segment. In other words, the flow pattern can be constructed step by step,
proceeding downstream. This property of limited upstream influence exists as long
as the deflection does not become so large that the flow is rendered subsonic.
To compare the compression in the two cases, (a) and (b), of Figure 15.6, we shall
make use of the approximate expressions for weak shocks obtained in the preceding
section. For each shock in case (b), we have

p , (15.30)
S ()3 . (15.31)

Thus, if the complete turn is subdivided into n equal segments, so that

= n , (15.32)

then the overall pressure and specific entropy changes are

pn p1 n , (15.33)
3
Sn S1 n ()3 ()2 . (15.34)
n2
Hence, if the compression is achieved by means of n
1 weak shocks then the
net specific entropy increase can be reduced very significantly, compared to that
generated by a single shock that produces the same deflection. In fact, the reduction
is by a factor of order 1/n 2.
In the limit that n , the smooth turn of Figure 15.6(c) is obtained. In this
case, the net increase in specific entropy becomes vanishingly small. In other words,
the compression becomes isentropic. We can also deduce the following results. First,
the shocks become vanishingly weak, their limiting positions being straight Mach
lines emanating from the wall. Second, each segment of uniform flow becomes van-
ishingly narrow, and eventually coincides with a Mach line. Thus, on each Mach
line, both the flow inclination and the Mach number are constant. Third, the limited
upstream influence is preserved. That is, the flow upstream of a given Mach line
is not aected by any downstream modifications in the wall. Finally, the approxi-
mate expression for the change in flow speed across a weak shock, Equation (15.29),
becomes the dierential expression
dw d
= . (15.35)
w Ma 2 1
Let us now consider what happens to the flow further away from the wall, where
the weak shocks and Mach lines of Figure 15.6(b) and Figure 15.6(c), respectively,
Two-Dimensional Compressible Inviscid Flow 447

shock front

Ma2

b Ma1

envelope of Ma2
Mach lines

a Ma1

wall

Figure 15.7
Convergence of Mach lines in continuous compression by turning.

converge. Consider Figure 15.7. Because of the convergence of the Mach lines,
the change from the initial Mach number, Ma1 , to the final Mach number, Ma2 , is
achieved in a much shorter distance on streamline b than on streamline a. Thus,
the velocity gradients on streamline b are higher than those on streamline a. Now,
an intersection of Mach lines would imply infinitely high velocity gradients, because
there would be two dierent values of the Mach number at a single point. In practice,
this does not occur, because, in the region where the Mach lines converge, and before
they cross, the velocity gradients become high enough that the conditions are no
longer isentropic. In fact, a shock wave develops, as illustrated in the figure.

15.6 Supersonic Expansion by Turning


Up to now, we have only considered turns that are concave. That is, turns in which
the wall is deflected into the oncoming flow. Let us now consider what happens in a
convex turn, where the wall is deflected away from the oncoming flow. In particular,
let us investigate supersonic flow over a convex corner, such as that illustrated in
Figure 15.8(a).
A convex turn through a single oblique shock, like that illustrated in Figure 15.8(a),
is not possible. In such a turn, the deviation of the flow, as it passes through shock, is
inevitably away from the plane of the shock front. Such a deviation would require the
downstream normal flow speed to exceed the upstream normal flow speed (because
the upstream and downstream tangential flow speeds are equal). (See Section 15.2.)
448 Theoretical Fluid Mechanics
Mach lines
(a) shock (b)
Ma1
Ma1 2
1
Ma2
2 Ma2
1
wall

(c)
Ma1

Ma2

Figure 15.8
Expansion of supersonic flow via turning. Case (a) shows a single turn through an
oblique shock, which is forbidden on thermodynamic grounds. Case (b) shows a
Prandtl-Mayer expansion fan. Case (c) shows smooth continuous expansion.

However, although this would satisfy the equations of fluid motion, it would lead
to a spontaneous decrease in specific entropy across the shock front, and is, thus,
forbidden by the second law of thermodynamics. (See Section 14.8.)
What actually happens is as follows. The nonlinear mechanism that tends to
steepen a compression (see Sections 14.4 and 15.5) produces the opposite eect in
an expansion. In particular, instead of being convergent, the Mach lines are divergent,
as shown in Figure 15.8(b) and (c). Consequently, there is a tendency for velocity
gradients to decrease with increasing distance from the wall. Thus, an expansion is
isentropic throughout the fluid.
The expansion at a corner takes place via a fan of straight Mach lines that is
generally known as a Prandtl-Mayer expansion fan. [See Figure 15.8(b).] The ar-
guments that lead to this conclusion are as follows. First, the flow up to the corner
is uniform, at Mach number Ma1 (say), and, thus, the leading Mach line must be
straight, and inclined to the flow at the Mach angle 1 = sin1 (1/Ma1 ). Because of
the limited upstream influence in supersonic flow, the same argument may be applied
to each succeeding portion of the flow. The terminating Mach line stands at the angle
2 = sin1 (1/Ma2 ) to the wall. Here, Ma2 is the Mach number of the downstream
flow. Second, because there is no characteristic length to define a scale in this prob-
lem, any variation of flow parameters can only be with respect to angular position,
Two-Dimensional Compressible Inviscid Flow 449

measured relative to the corner. That is, flow conditions must be constant along rays
emanating from the corner.
Figure 15.8(c) shows a typical expansion over a continuous convex turn. Because
the flow is isentropic, it is also reversible (i.e., if the direction of the flow is reversed
then the fluid is isentropicaly compressed).
Equation (15.35), which specifies the relationship between and Ma in an isen-
tropic compression or expansion by turning, can be written
 dw
d = Ma 2 1 , (15.36)
w
or  Ma 
dw
+ const. = Ma 2 1 (Ma). (15.37)
w
In order to evaluate the integral, and, thus, derive an explicit form for the function
(Ma), we need to rewrite w in terms of Ma using the definition
w = c Ma. (15.38)

Here, c is the local sound speed. Given that c T [see Equation (14.45)], it follows
from Equation (14.58) that


c02 1
=1+ Ma 2 , (15.39)
c2 2
where c0 is the stagnation sound speed. Thus,


dw dMa dc 1 dMa
= + = . (15.40)
w Ma c 1 + [( 1)/2] Ma 2 Ma
The function (Ma), which is known as the Prandtl-Mayer function, is then given by
(see Exercise 15.10)
 Ma
M2 1 dM
(Ma) (15.41)
1 1 + [( 1)/2] M M 2

1/2
1/2 $
+1 1 1/2 %
1
= tan (Ma 1) tan1 Ma 2 1
2
.
1 +1

Here, the constant of integration has been arbitrarily chosen such that = 0 cor-
responds to Ma = 1. A supersonic Mach number, Ma, is thus associated with a
definite value of the function . In fact, as Ma increases from 1 to , increases
monotonically from 0 to max , where

1/2
+ 1
max = 1 . (15.42)
2 1

Equation (15.37) yields


= (Ma) (Ma1 ), (15.43)
450 Theoretical Fluid Mechanics

d
(a) (b)
b
c
c Ma = Ma1
Ma > 1
b = max

Ma < 1 afterbody
Ma1 a Ma = 1
d a

/2
shoulder
Ma > 1

Figure 15.9
Detached shock waves. (a) Detached shock on wedge with afterbody. [At lettered
points on the shock, flow deflection is given by corresponding points in (b).] (b)
Deflection angle versus wave angle for fixed Ma1 .

as the relationship between the deflection angle, , and the Mach number, Ma, in an
isentropic compression or expansion by turning. In a compressive turn, decreases,
whereas, in an expansive turn, it increases: in each case, by an amount equal to the
change in . Here, we have set 1 = 0, because only the change in matters. Note
that is positive in compressive turns, and negative in expansive turns.

15.7 Detached Shocks


Let us return to the problem of supersonic flow, of Mach number Ma1 , over a sym-
metric wedge of nose angle 2 , that was previously discussed in Section 15.3. What
happens when the wedge angle, , is greater than max ?
In fact, there is no rigorous analytical treatment for cases where exceeds max .
Experimentally, it is observed that the flow configurations are like those sketched in
Figure 15.9(a). The flow is compressed through a curved shock front, known as a
bow shock, that is detached from, and stands some distance upstream of, the wedge
apex. The shape of the bow shock, as well as its detachment distance, depend on the
geometry of the wedge, as well as the upstream Mach number, Ma1 .
On the central streamline, where the shock is normal, as well as on the nearby
ones, where it is nearly normal, the flow is compressed to subsonic speeds. Farther
out, as the shock becomes weaker, its inclination becomes less steep, approaching
the upstream Mach angle asymptotically. Thus, conditions along the bow shock run
the whole range of oblique shock solutions for the given Mach number. This is
Two-Dimensional Compressible Inviscid Flow 451

illustrated in Figure 15.9(b), which shows one curve from Figure 15.2. The complete
description of the bow shock is complicated by the fact that its location cannot be
found explicitly, because of the subsonic influence upstream, which implies that the
entire flow field must be solved as one. There is an absolute limit of , just greater
than 45 , beyond which the shock will always detach, no matter how large the value
of the upstream Mach number, Ma1 . Hence, if the wedge is replaced by a blunt-nosed
obstacle then the shock will always detach (because = 90 at the nose of such an
obstacle).
For a given wedge angle, , the sequence of events with decreasing upstream
Mach number, Ma1 , is as follows. When Ma1 is suciently high, the shock front
is attached to the wedge apex, and its straight portion is independent of both the
shoulder and the afterbody. [See Figure 15.9(a).] The straight portion of the shock
front lies between the wedge apex and the point where the first Mach line, emanating
from the shoulder, intersects it. As Ma1 decreases, the wave angle, , increases. With
a further decrease in the Mach number, a value is reached at which the fluid behind
the shock becomes subsonic. The shoulder now has an eect on the whole shock
front, which may become curved, although still remaining attached to the wedge
apex. These conditions correspond to the region between the lines Ma2 = 1 and
= max in Figure 15.2. At the Mach number corresponding to = max , the shock
front starts to detach from the apex. This Mach number is called the detachment
Mach number. With a further decrease of Ma1 , the detached shock moves upstream
of the wedge apex.

15.8 Shock-Expansion Theory


It is possible to solve many problems in two-dimensional supersonic flow by patch-
ing together appropriate combinations of the oblique shock wave, described in Sec-
tion 15.2, and the Prandtl-Mayer expansion fan, described in Section 15.6. For ex-
ample, let us consider the flow over a simple two-dimensional airfoil section.
Figure 15.10 shows a flat plate inclined at an angle of attack 0 to the oncoming
supersonic flow. The streamline ahead of the leading edge is non-inclined, because
there is no upstream influence. Moreover, the flow streams over the upper and lower
surfaces are completely independent of one another. Thus, the flow on the upper
surface is turned through an expansion angle 0 by means of a Prandtl-Mayer expan-
sion fan attached to the leading edge of the airfoil, whereas the flow on the lower side
is turned through a compression angle 0 by means of an oblique shock. The flow
on the upper surface is recompressed to the upstream pressure p1 by means of an
oblique shock wave attached to the trailing edge of the airfoil. Likewise, the flow on
the lower surface is re-expanded to the upstream pressure by means of an expansion
fan. The uniform pressures, p2 and p2 , respectively, on the upper and lower surfaces
of the airfoil can easily be calculated by means of oblique shock theory and Prandtl-
Mayer expansion theory. Given the pressures, the lift and drag per unit transverse
452 Theoretical Fluid Mechanics

expansion fan shock

p1 p2
plate
Ma1 p2 0 p1
slipstream
c

shock expansion fan

Figure 15.10
A flat lifting plate. Ma1 is the upstream Mach number, p1 , p2 , et cetera, denote
pressures, and 0 is the angle of attack.

length of the airfoil are simply

L = (p2 p2 ) c cos 0 , (15.44)


D= (p2 p2 ) c sin 0 , (15.45)

respectively, where c is the chord-length (i.e., width). (See Figure 15.10.) The in-
crease in entropy of the flow along the upper surface of the airfoil is not the same as
that for the flow along the lower surface, because the upper and lower shock waves
occur at dierent Mach numbers. Consequently, the streamline attached to the trail-
ing edge of the airfoil is a slipstreamthat is, it separates flows with the same pres-
sures, but slightly dierent speeds, temperatures, and densitiesinclined at a small
angle relative to the free stream. (The angle of inclination is determined by the re-
quirement that the pressures on both sides of the slipstream be equal to one another.)
Note that the drag that develops on the airfoil is of a completely dierent nature
to the previously discussed (see Chapter 9) drags that develop on subsonic airfoils,
such as friction drag, form drag, and induced drag. This new type of drag is termed
supersonic wave drag, and exists even in an idealized, inviscid fluid. It is ultimately
due to the trailing shock waves attached to the airfoil.
Comparatively far from the airfoil, the attached shock waves and expansion fans
intersect one another. The expansion fans then attenuate the oblique shocks, mak-
ing them weak and curved. At very large distances, the shock waves asymptote to
free-stream Mach lines. However, this phenomenon does not aect the previous cal-
culation of the lift and drag on a flat-plate airfoil.
Two-Dimensional Compressible Inviscid Flow 453

15.9 Thin-Airfoil Theory


The shock-expansion theory of the previous section provides a simple and general
method for computing the lift and drag on a supersonic airfoil, and is applicable
as long as the flow is not compressed to subsonic speeds, and the shock waves re-
main attached to the airfoil. However, the results of this theory cannot generally be
expressed in concise analytic form. In fact, the theory is mostly used to obtain nu-
merical solutions. However, if the airfoil is thin, and the angle of attach small, then
the shocks and expansion fans attached to the airfoil become weak. In this situa-
tion, shock-expansion theory can be considerably simplified by using approximate
expressions for weak shocks and expansion fans.
The basic approximate expression [cf. Equation (15.19)]

p Ma 2
 , (15.46)
p Ma 2 1
specifies the relative change in pressure across either a weak oblique shock (see Sec-
tion 15.4) or a weak expansion fan (see Exercise 15.6) that deflects flow of Mach
number Ma through an angle . Because, in the weak wave approximation, the
pressure, p, never greatly diers from the upstream pressure, p1 , and the Mach num-
ber, Ma, never diers appreciably from the upstream Mach number, Ma1 , we can
write

p p1 Ma12
 , (15.47)
p1 Ma 2 1
1

which is correct to first order in . Here, is the deflection angle relative to the
upstream flow.
It is convenient to define a dimensionless quantity known as the pressure coe-
cient:
p p1
Cp = , (15.48)
q1
where q1 = (1/2) 1 w12 , and 1 and w1 are the upstream density and flow speed,

respectively. Given that the upstream sound speed is c1 = p1 /1 , and Ma1 =
w1 /c1 , we obtain
2 p p1
Cp = , (15.49)
Ma12 p1
which yields
2
Cp =  . (15.50)
Ma12 1
This is the fundamental formula of thin-airfoil theory. It states that the pressure coef-
ficient is proportional to the local deflection of the flow from the upstream direction.
Consider the flat-plate airfoil shown in Figure 15.10. The deflection angle is 0
454 Theoretical Fluid Mechanics

on the upper surface of the airfoil, and +0 on the lower surface. (A positive deflec-
tion angle corresponds to compression, and a negative deflection angle to expansion.)
Thus, the pressure coecients on the upper and lower surfaces are

2 0
C pU =  (15.51)
Ma12 1

and
2 0
C pL =  , (15.52)
Ma12 1

respectively. It is convenient to define the dimensionless coecient of lift,

L
CL = , (15.53)
q1 c

and the dimensionless coecient of drag,

D
CD = . (15.54)
q1 c

It follows that
(pL pU ) c cos 0
CL = = (C pL C pU ) cos 0 , (15.55)
q1 c
(pL pU ) c sin 0
CD = = (C pL C pU ) sin 0 . (15.56)
q1 c

Making use of Equations (15.51) and (15.52), as well as the conventional small angle
approximations cos 0 1 and sin 0 0 , we obtain

4 0
CL =  , (15.57)
Ma12 1

4 02
CD =  . (15.58)
Ma12 1

The focus (see Section 9.3) of the airfoil is at the midchord. Moreover, the ratio
D/L 2 = (Ma12 1)1/2 /4 is independent of 0 .
As a second example, consider the diamond-section airfoil pictured in Figure 15.11.
This airfoil has a nose angle 2 , and zero angle of attack. The pressure coecient
on the front face of the airfoil is
2
C pF =  , (15.59)
Ma12 1
Two-Dimensional Compressible Inviscid Flow 455
shock expansion fan

p2 p3
Ma1
2 2 t
p2 p3

Figure 15.11
A diamond-section airfoil. Ma1 is the upstream Mach number. p2 and p3 denote
pressures.

whereas that on the rear face is


2
C pR =  . (15.60)
Ma1 1
2

It follows that the pressure dierence is

4
p2 p3 =  q1 , (15.61)
Ma12 1

giving a drag

42
D = (p2 p3 ) t = (p2 p3 )  c =  q1 c, (15.62)
Ma12 1

where t and c are the thickness and chord-length of the airfoil, respectively. (See
Figure 15.11.) Thus,

D 42 4 $ t %2
CD = =  =  . (15.63)
q1 c c
Ma1 1
2
Ma1 1
2

Figure 15.12 shows the cross-section of an arbitrary airfoil. The cross-section


456 Theoretical Fluid Mechanics

yU (x)

yC (x)

yL(x)

xL xR
x

Figure 15.12
An arbitrary airfoil.

is assumed to be uniform in the z-direction, with the upstream flow parallel to the
x-axis. The upper surface of the airfoil corresponds to the curve y = yU (x), the lower
surface to the curve y = yL (x), and the camber line (i.e., the centerline) to the curve
y = yC (x). Furthermore, the leading and trailing edges of the airfoil lie at x = xL
and x = xR , respectively. Hence, yU (xL ) = yC (xL ) = yL (xL ) and yU (xR ) = yC (xR ) =
yL (xR ). By definition,
yU (x) + yL (x)
yC (x) = , (15.64)
2
for xL x xR . It is helpful to define the half-width of the airfoil,
yU (x) yL (x)
h(x) = , (15.65)
2
for xL x xR . Note that h(xL ) = h(xR ) = 0. We can also define the mean angle of
attack of the airfoil:
yU (xL ) yU (xR ) yL (xL ) yL (xR ) yC (xL ) yC (xR )
0 = = = . (15.66)
xR x L xR x L xR x L
Thus, we can write

yC (x) = yC (xL ) 0 (x xL ) C (x), (15.67)

where C (xL ) = C (xR ) = 0. Here,

(x) = 0 + C (x) (15.68)


Two-Dimensional Compressible Inviscid Flow 457

is the local angle of attack of the camber line. Thus, the camber function, C (x),
parameterizes the deviations of (x) from 0 across the width of the airfoil. Equa-
tions (15.64), (15.65), and (15.67) yield

yU (x) = yC (x) + h(x) = yC (xL ) 0 (x xL ) C (x) + h(x), (15.69)


yL (x) = yC (x) h(x) = yC (xL ) 0 (x xL ) C (x) h(x). (15.70)

Hence, the airfoil shape is completely specified by the thickness function, h(x), the
camber function, C (x), and the mean angle of attack, 0 .
The pressure coecients on the upper and lower surfaces of the airfoil are [see
Equations (15.50)]

2 dyU
C pU =  , (15.71)
dx
Ma12 1


2 dyL
C pL =  , (15.72)
dx
Ma12 1

respectively. It follows from Equations (15.68)(15.70) that

dyU dh
= (x) + , (15.73)
dx dx
dyL dh
= (x) . (15.74)
dx dx
Now, the lift and drag per unit transverse length acting on the airfoil are given by
[cf., Equations (15.53)(15.56)]
 xR  
L = q1 C pL C pU dx, (15.75)
xL
 xR


dyL dyU
D = q1 C pL + C pU dx. (15.76)
xL dx dy

Thus, it follows from Equations (15.71)(15.74) that


 xR
4 q1
D=  (x) dx, (15.77)
Ma12 1 xL
 xR
2
2
2 q1 dyL dyU
L=  dx + dx dx
Ma12 1 xL
 xR
2

dh 2
+ dx.
4 q1
=  (15.78)
dx
Ma12 1 xL
458 Theoretical Fluid Mechanics

It is helpful to define the chord-average operator:



1 xR
y y(x) dx, (15.79)
c xL
where c = xR xL is the chord-length. Taking the average of Equation (15.73),
making use of Equation (15.66), as well as the fact that h(xL ) = h(xR ) = 0, we obtain
= 0 . (15.80)
However, the average of Equation (15.68) yields
= 0 + C , (15.81)
which implies that C = 0. We can also write

2 = (0 + C )2 = 02 + 2 0 C + C2 = 02 + C2 . (15.82)
Hence, the coecients of lift and drag, L/(q1 c) and D/(q1 c), are written
4 4 0
CL =  =  , (15.83)
Ma12 1 Ma12 1


dh 2
CD = 
4 + 2
dx
Ma1 1
2



dh 2
= 
4 + + 2 ,
C
2
dx 0 (15.84)
Ma1 1
2

respectively. Thus, in thin-airfoil theory, the lift only depends on the mean angle
of attack, whereas the drag splits into three components. Namely, a drag due to
thickness, a drag due to lift, and a drag due to camber.

15.10 Croccos Theorem


Bernoullis theorem for the steady, inviscid flow of an ideal gas, in the absence of
body forces, implies that, on a given streamline,
1 2
H+ v = H0 , (15.85)
2
where H is the specific enthalpy, H0 the stagnation enthalpy, and v the flow velocity.
(See Section 14.5.) The fluid equation of motion, (14.31), reduces to
p
(v ) v = . (15.86)

Two-Dimensional Compressible Inviscid Flow 459

However, according to Equation (A.171),


(v ) v (v 2 /2) v , (15.87)
where = v. Hence, we obtain
p
v = (v 2 /2) + . (15.88)

Now, Equation (14.29) implies that
p
H = + T S, (15.89)

where T is the temperature, and S the specific entropy. Moreover, Equation (15.85)
yields
H + (v 2 /2) = H0 . (15.90)
Thus, we arrive at Croccos theorem:
v = H0 T S. (15.91)
In most aerodynamic flows, the fluid originates from a common reservoir, which
implies that the stagnation enthalpy, H0 , is the same on all streamlines. Such flow is
termed homenergic. It follows from Equation (14.18) that the stagnation temperature,
T 0 , is the same on all streamlines in homenergic flow. According to Croccos theo-
rem, an irrotational (i.e., = 0 everywhere) homenergic (i.e., H0 = 0 everywhere)
flow pattern is also homentropic (i.e., S = 0 everywhere). Conversely, a homener-
gic, homentropic flow pattern is also irrotational (at least, in two dimensions, where
v and cannot be parallel to one another).

15.11 Homenergic Homentropic Flow


Consider a steady, two-dimensional, homenergic, homentropic flow pattern, in the
absence of body forces. Suppose that all quantities are independent of the Cartesian
coordinate z, and that the flow velocity, q, is confined to the x-y plane. Let q =
u e x + v ey . Equations (14.25), (14.30), and (14.31) reduce to

( u) + ( v) = 0, (15.92)
x y
u u 1 p
u +v = , (15.93)
x y x
v v 1 p
u +v = , (15.94)
x y y


p
= , (15.95)
p0 0
460 Theoretical Fluid Mechanics

where p0 and 0 are the uniform stagnation pressure and density, respectively. (Note

that p0 and 0 must be uniform because the stagnation specific entropy, S0 p0 /0 ,
and the stagnation temperature, T 0 p0 /0 , are both uniform.) Equation (15.95)
implies that
p = c 2 , (15.96)

where c = p/ is the sound speed. Hence, Equations (15.92)(15.94) yield

u v u v
+ = , (15.97)
x y x y
u u c 2
u2 +uv = u , (15.98)
x y x
v v c 2
uv + v2 = v . (15.99)
x y y

Summing the previous two equations, and then making use of Equation (15.97), we
obtain

   
2 u 2 v u v
u c
2
+ v c
2
+uv + = 0. (15.100)
x y y x

Finally, given that c T , Equation (14.58) implies that

1  
c 2 = c02 ( 1) u 2 + v 2 , (15.101)
2
where c0 is the stagnation sound speed.

15.12 Small-Perturbation Theory


A great number of problems of interest in compressible fluid mechanics are con-
cerned with the perturbation of a known flow pattern. The most common case is that
of uniform, steady flow. Let U denote the uniform flow velocity, which is directed
parallel to the x-axis. The density, pressure, and temperature are also assumed to be
uniform, and are denoted , p , and T , respectively. The corresponding sound
speed is c , and the Mach number is Ma = U/c . Finally. the velocity field of the
unperturbed flow pattern is

u = U, (15.102)
v = 0. (15.103)

Suppose that a solid body, such as an airfoil, is placed in the aforementioned flow
pattern. The cross-section of the body is assumed to be independent of the Cartesian
Two-Dimensional Compressible Inviscid Flow 461

coordinate z. The body disturbs the flow pattern, and changes its velocity field, which
is now written

u = U + u , (15.104)

v=v, (15.105)

where u and v are known as induced velocity components. We are interested in


situations in which u /U  1 and v /U  1.
Equation (15.101) can be combined with the previous two equations to give
1  
c 2 = c2 ( 1) 2 U u + u 2 + v 2 . (15.106)
2
It then follows from Equation (15.100) that


 2

 2 
2 u

v 2
u 
+ 1 u 1 v u
(1 Ma ) + = Ma ( + 1) + +
x y U 2 U 2 U x


2

2
2
u + 1 v 1 u v

+ Ma ( 1) + +
U 2 U 2 U y




2 v u u v
+ Ma 1+ + . (15.107)
U U y x
The previous equation is exact. However, if u /U and v /V are small then it becomes
possible to neglect many of the terms on the right-hand side. For instance, neglecting
terms that are third-order in small quantities, we obtain
u v u u u v
(1 Ma2 ) + Ma2 ( + 1) + Ma2 ( 1)
x y U x U y


v u v
+ Ma2 + . (15.108)
U y x
Furthermore, if we neglect terms that are second-order in small quantities then we
get the linear equation
u v
(1 Ma2 ) + 0. (15.109)
x y
Note, however, than in so-called transonic flow, where Ma 1, the coecient of
u /x on the left-hand side of Equation (15.108) becomes very small. In this situ-
ation, it is not possible to neglect the first term on the right-hand side. However, the
condition Ma 1 does not aect the term v /y on the left-hand side of Equa-
tion (15.108), and so the other terms on the right-hand side can still be neglected.
Thus, transonic flow is governed by the non-linear equation
u v u u
(1 Ma2 ) + Ma2 ( + 1) . (15.110)
x y U x
462 Theoretical Fluid Mechanics

On the other hand, subsonic (i.e., Ma < 1) and supersonic flow (i.e., Ma > 1)
are both governed by Equation (15.109). Another situation in which certain terms
on the right-hand side of Equation (15.108) must be retained is hypersonic flow (i.e.,
Ma
1). This follows because, although u /U and v /V are small, their products
with Ma2 can still be non-negligible. Roughly speaking, Equation (15.109) is valid
for 0 Ma 0.8 and 1.2 Ma 5. In other words, transonic flow corresponds
to 0.8 < Ma < 1.2, and hypersonic flow to Ma > 5 (Anderson 2003).
The pressure coecient is defined
p p 2 p p
Cp = = . (15.111)
(1/2) U 2 Ma2 p

(See Section 15.9.) Equation (15.101) implies that


1 1
c2 + ( 1) q 2 = c2 + ( 1) U 2 , (15.112)
2 2
where q 2 = (U + u ) 2 + v 2 . Moreover, Equation (14.61) yields
/(1)
p 2 + ( 1) Ma2
= , (15.113)
p 2 + ( 1) Ma 2
where Ma = q/c. The previous two equations can be combined to give

/(1)
p 1 q2
= 1 + ( 1) Ma2 1 2 . (15.114)
p 2 U

Hence, we obtain

/(1)
2
1 ,
2 1
1 + ( 1) Ma 1 q
Cp = 2
(15.115)
Ma2 2 U 2

which reduces to


2  2 /(1)
Cp =
2 1 1 ( 1) Ma2 2 u + u + v 1 . (15.116)
Ma2 2 U U2

Using the binomial expansion on the expression in square brackets, and neglecting
terms that are third-order, or higher, in small quantities, we obtain

 2
 2
2 u u v
C p + (1 Ma )
2
+ .
U
(15.117)
U U

For two-dimensional flows, in the limit in which Equation (15.109) is valid, it is


consistent to retain only first-order terms in the previous equation, so that
2 u
Cp . (15.118)
U
Two-Dimensional Compressible Inviscid Flow 463

Let
f (x, y) = 0 (15.119)
be the equation of the surface of the solid body that perturbs the flow. At the surface,
the velocity vector of the flow must be perpendicular to the local normal: that is, the
flow must be tangential to the surface. In other words,
q f = 0, (15.120)
which reduces to
f f
(U + u ) + v = 0. (15.121)
x y
Neglecting u with respect to U, we obtain
v f /x dy
= , (15.122)
U f /y dx
where dy/dx is the slope of the surface, and v /U the approximate slope of a stream-
line.
Now, the body has to be thin in order to satisfy our assumption that the induced
velocities are relatively small. This implies that the coordinate y diers little from
zero (say) on the surface of the body. Hence, we can write


  v
v (x, y) = v (x, 0) + y + (15.123)
y y=0
in the immediate vicinity of the surface. Within the framework of small-perturbation
theory, it is consistent to neglect all terms on the right-hand side of the previous
equation after the first. Hence, the boundary condition (15.122) reduces to


dy
v (x, 0) = U , (15.124)
dx body
respectively.
Because a homenergic, homentropic flow pattern is necessarily irrotational (see
Section 15.10), we can write
q = U e x , (15.125)
where is the perturbed velocity potential. (See Section 4.15.) It follows that

u = , v = . (15.126)
x y
Hence, Equations (15.109), (15.118), and (15.124) become
2 2
(1 Ma2 ) + 0, (15.127)
x 2 y 2
2
Cp , (15.128)
U x


(x, 0) dy
U , (15.129)
y dx body
respectively.
464 Theoretical Fluid Mechanics

15.13 Subsonic Flow Past a Wave-Shaped Wall


The following simple example serves to clarify many of the concepts introduced
in the previous section. Consider flow past a straight wall with a small-amplitude
sinusoidal modulation whose surface is specified by

y =  sin( x). (15.130)

Here,  is the amplitude of the modulation, whereas l = 2/ is the wavelength. In


the limit  0, the unperturbed flow is of uniform speed U, directed parallel to
the x-axis. The corresponding Mach number is Ma . The fluid occupies the region
y >  sin( x).
The perturbed flow is governed by Equation (15.127),

2 2
(1 Ma2 ) + 0, (15.131)
x 2 y 2
and is subject to the boundary condition (15.129), which can be written
(x, 0) dy
= U = U  cos( x). (15.132)
y dx
We also expect the perturbed flow to become vanishingly small far from the wall,
which implies that
(x, ) (x, )
= = 0. (15.133)
x y
Consider subsonic flow, for which 1 Ma2 m 2 > 0. Equation (15.131) is of
the elliptic type
2 1 2
+ 2 =0 (15.134)
x 2 m y 2
(Arfken 1985). Let us search for a separable solution of the form

(x, y) = F(x) G(y). (15.135)

We obtain
F  1 G
= 2 , (15.136)
F m G
where  denotes derivative with respect to argument. Given that the left-hand side of
the previous equation is a function of x only, whereas the right-hand side is a function
of y only, the two sides must equal the same constant. In other words,
F 
= k 2 , (15.137)
F
G
= +k 2 . (15.138)
G
Two-Dimensional Compressible Inviscid Flow 465

Here, the sign of the constant has been chosen so as to allow the boundary condition
(15.133) to be satisfied. The most general solution of Equation (15.137) is

F(x) = A1 sin(k x) + A2 cos(k x), (15.139)

where A1 and A2 are arbitrary constants. Likewise, the most general solution of
Equation (15.138) is
G(y) = B1 e m k y + B2 e+m k y , (15.140)
where B1 and B2 are arbitrary constants.
The boundary condition (15.133) implies that B2 = 0. The boundary condi-
tion (15.132) then yields

m k B1 [A1 sin(k x) + A2 cos(k x)] = U  cos( x). (15.141)

This condition can only be satisfied at all x if A1 = 0, k = , and m B1 A2 = U .


Thus, we obtain
U
(x, y) = cos( x) em y , (15.142)
m
U 
u = = sin( x) em y , (15.143)
x m

v = = U  cos( x) em y . (15.144)
y

According to Equation (15.128), the pressure coecient at the wall can be written

2  2
Cp  = sin( x). (15.145)
U x y=0
1 Ma 2

It immediately follows that there is zero net drag force acting on the wall, because the
pressure variations are in phase with the walls sinusoidal modulations, and, hence,
symmetrical about the crests and troughs.
Let us now consider under what circumstances the approximations made in the
small-perturbation theory developed in the previous section are valid for the problem
under investigation. In fact, small-perturbation theory is premised on three assump-
tions. First, that u /U  1 and v /V  1. It is evident from Equations (15.143) and
(15.144) that this assumption is valid provided

  1. (15.146)
1 Ma2

Second, in Equation (15.110) it is assumed that

u u u
(1 Ma2 )
Ma2 ( + 1) , (15.147)
x U x
466 Theoretical Fluid Mechanics

which implies that


Ma2 ( + 1) 
 1. (15.148)
(1 Ma2 ) 3/2
Finally, in Equation (15.123), it is assumed that the second term on the right-hand
side is negligible compared to the first, when evaluated at the surface of the wall,
which implies that

 1 Ma2  1. (15.149)
Hence, we deduce that the analysis described in this section is valid provided

(1 Ma2 ) 3/2 1
 1 Ma2 , ,  . (15.150)
Ma ( + 1)
2
1 Ma
2

15.14 Supersonic Flow Past a Wave-Shaped Wall


Suppose that the unperturbed flow in the problem considered in the previous sec-
tion is supersonic, so that Ma2 1 2 > 0. The perturbed flow is governed by
Equation (15.127), which is now of the hyperbolic type

2 1 2
0 (15.151)
x 2 2 y 2

(Arfken 1985). As is easily verified, the previous equation has the general solution

(x, y) = f (x y) + g(x + y), (15.152)

where f and g are arbitrary functions (Fitzpatrick 2013). It is clear that f is constant
along lines x y = const., whereas g is constant along lines x + y = const.. These
lines are inclined at the Mach angle, = cot1 [(Ma2 1) 1/2 ], to the undisturbed
flow. They are, in fact, the Mach lines, or characteristics, of the unperturbed flow.
(See Section 15.3.) The former characteristics are inclined downstream. In other
words, they originate at the wall. The latter characteristics are inclined upstream. In
other words, they originate at infinity. Because there are no sources at infinity, the
latter characteristics carry no perturbation, which implies that g = 0.
The boundary condition at the wall, (15.129), yields

f  (x) = U  cos( x), (15.153)

where  denotes derivative with respect to argument. It follows that

U
f (x) = sin( x), (15.154)

Two-Dimensional Compressible Inviscid Flow 467

and, hence, that


U
(x, y) = f (x y) = sin[ (x y)], (15.155)

U
u = = cos[ (x y)], (15.156)
x

v = = U  cos[ (x y)]. (15.157)
y
Note that, unlike the case of subsonic flow, the amplitude of the perturbed velocity
does not decrease with increasing distance from the wall.
The pressure coecient at the wall is

2  2
Cp  =  cos( x). (15.158)
U x y=0
Ma2 1

It can be seen, by comparison with Equation (15.145), that the maxima and minima
of the pressure are now phase-shifted by /2 radians with respect to the correspond-
ing maxima and minima in the subsonic case. Hence, the pressure distribution at
the wall is anti-symmetric with respect to the crests and troughs of the wall. Con-
sequently, a net drag force is exerted on the wall. The mean coecient of drag per
wavelength is [see Equation (15.56)]

1 l dy
CD = Cp dx, (15.159)
l 0 dx
which is obtained by replacing the sine of the slope of the wall by the tangent, dy/dx.
This approximation is valid within the context of the small-perturbation theory. How-
ever, according to Equations (15.130) and (15.158), C p can be written
2 dy
Cp =  , (15.160)
dx
Ma2 1

which implies that



2
2 dy
CD =  . (15.161)
dx
Ma2 1
Here, the bar denotes a period average of the form

2  l
2
dy 1 dy
= dx. (15.162)
dx l 0 dx
Equations (15.160) and (15.161) are valid for any small-amplitude, periodic modula-
tion of the wall. For the particular sinusoidal modulation under consideration in this
section, we find that
( ) 2
CD =  . (15.163)
Ma2 1
468 Theoretical Fluid Mechanics

The discussion of the range of validity of the approximations used in deriving the
previous results follows the same lines as in the subsonic case.

15.15 Linearized Subsonic Flow


The aim of this section is to modify the two-dimensional, incompressible, subsonic
aerodynamic theory discussed in Chapter 9 so as to take the finite compressibility of
air into account.
Consider compressible, subsonic flow over a thin airfoil at a small angle of attack.
Such a situation can be analyzed using the small-perturbation theory introduced in
Section 15.12. As before, the unperturbed flow is of uniform speed U, directed
parallel to the x-axis, and the associated Mach number is Ma . The perturbed flow,
due to the presence of the airfoil, is governed by Equation (15.127), which can be
written in the elliptic form (see Section 15.13)
2 1 2
+ 0, (15.164)
x 2 m 2 y 2
where 
m= 1 Ma2 . (15.165)
Equation (15.164) can be transformed into a familiar incompressible form by
introducing a transformed coordinate system (, ), such that
= x, (15.166)
= m y. (15.167)
In this transformed space, we define a transformed perturbed velocity potential,
(, ), such that
(, ) = m (x, y). (15.168)
Now,

= 1, = 0, = 0, = m, (15.169)
x y x y
so

= + = , (15.170)
x x x

= + =m . (15.171)
y y y
It follows that
1 2 1 2 2 2
= , = , = , = m 2. (15.172)
x m y x 2 m 2 y 2
Two-Dimensional Compressible Inviscid Flow 469

Thus, Equation (15.164) transforms to give

2 2
+ 0. (15.173)
x 2 y 2
However, this is simply Laplaces equation, which governs incompressible flow. (See
Section 4.15.) Hence, (, ) represents an incompressible flow in (, ) space that
is related to a compressible flow, represented by (x, y), in (x, y) space.
Suppose that the surface of the airfoil is given by y = f (x) in (x, y) space, and
= q() in (, ) space. According to Equation (15.129), the boundary condition on
the flow at the airfoil surface is
df (x, 0)
U (15.174)
dx y
in (x, y) space. The analogous boundary condition in (, ) space is

dq (, 0)
U . (15.175)
d
However,

= , (15.176)
y
which implies that
df dq
= . (15.177)
dx d
In other words, the shape of the airfoil in (x, y) space is the same as that in (, )
space. Hence, the coordinate transformation (x, y) (, ) transforms compressible
flow over an airfoil in (x, y) space to incompressible flow over the same airfoil in
(, ) space.
According to Equation (15.128), the pressure coecient in (x, y) space is written

2 2
Cp = . (15.178)
U x U m
The corresponding incompressible pressure coecient in (, ) space is

2
C p0 . (15.179)
U
A comparison of the previous two equations reveals that
C p0
Cp =  . (15.180)
1 Ma2

This result is known as the Prandtl-Glauert rule: it is a similarity rule that relates
incompressible flow over a given two-dimensional airfoil to subsonic compressible
470 Theoretical Fluid Mechanics

flow over the same airfoil. Given that the coecient of lift is directly related to the
pressure coecient (see Section 15.9), we can also deduce that

C L0
CL =  , (15.181)
1 Ma2

where C L is the coecient of lift for compressible flow, and C L0 the corresponding
coecient for incompressible flow. It is clear that compressibility has the eect of
increasing the lift on a given airfoil.
Equation (15.181) indicates that the coecient of lift goes to infinity as Ma ap-
proaches unity, which is an impossible result. The quandary is resolved by recalling
that linearized theory breaks down in the transonic regime (where Ma 1). In fact,
the Prandlt-Glauert rule is only accurate up to Mach numbers of approximately 0.8.
The eect of compressibility on subsonic flow-fields can be seen by noting that

1 u
u = = =  , (15.182)
x m
1 Ma2

where u = / is the incompressible velocity perturbation parallel to the unper-


turbed flow. It can be seen that as Ma increases, the parallel velocity perturbation,
u , also increases relative to u. In other words, compressibility strengthens the dis-
turbance of the flow introduced by the airfoil.
In conventional inviscid, incompressible aerodynamic theory, a two-dimensional
airfoil experiences zero aerodynamic drag (assuming that there is no separation of
the boundary layer). (See Chapters 8 and 9.) This is due to the fact that, in the
absence of friction, the pressure distributions over the forward and rearward portions
of the airfoil exactly cancel in the direction of the unperturbed flow. Note that the
compressible pressure coecient, C p , only diers from the incompressible pressure
coecient, C p0 , by a constant scale-factor. Thus, if the distribution of C p0 over the
surface of the airfoil results in zero drag then the distribution of C p will also result
in zero drag. In other words, in subsonic flow, compressibility does not give rise to
additional aerodynamic drag.

15.16 Linearized Supersonic Flow


The aim of this section is to re-examine the problem of supersonic flow past a
thin, two-dimensional airfoil using the small-perturbation theory developed in Sec-
tion 15.12.
As before, the unperturbed flow is of uniform speed U, directed parallel to the
x-axis, and the associated Mach number is Ma . The perturbed flow, due to the
presence of the airfoil, is governed by Equation (15.127), which can be written in the
Two-Dimensional Compressible Inviscid Flow 471

hyperbolic form (see Section 15.14)


2 1 2
0, (15.183)
x 2 2 y 2
where 
= Ma2 1. (15.184)
As we saw in Section 15.14, the general solution of Equation (15.183) is
(x.y) = f (x y) + g(x + x), (15.185)
where f and g are arbitrary functions. Because disturbances only propagate along
downstream-running Mach lines (i.e., Mach lines that originate at the airfoil), we
only need the function f for the upper surface, and the function g for the lower
surface. Thus,
(x, y) = f (x y) y > 0, (15.186)
(x, t) = g(x + y) y < 0. (15.187)
On the upper surface, the boundary condition (15.129) yields


dy (x, 0+)
U = = f  (x), (15.188)
dx U x
which implies that

 U dy
f (x) = . (15.189)
dx U
Similarly,

 U dy
g (x) = . (15.190)
dx L
According to Equation (15.128), the pressure coecient on the surface of the airfoil
is given by

2
Cp . (15.191)
U x y=0
Hence, we obtain
2 
Cp = f (x) (15.192)
U
on the upper surface, and
2 
Cp = g (x) (15.193)
U
on the lower surface. Thus,


2 dyU
C pU =  , (15.194)
dx
Ma2 1


2 dyL
C pL =  . (15.195)
dx
Ma2 1
472 Theoretical Fluid Mechanics

This is the same result as that obtained in Section 15.9, where the pressure on thin
airfoils was obtained by an approximation to the shock-expansion method. In fact,
for the purposes of calculating the velocity and pressure perturbations on the surface
of the airfoil, the linearized theory discussed in this section is equivalent to the weak
wave approximations of Section 15.9.

15.17 Exercises
15.1 Show that Equation (15.12) can be written in the form

X 3 + b X 2 + c X + d = 0, (15.196)

where

X = sin2 ,

Ma12 + 2 2
b = 2
+ sin ,
Ma1


2 Ma12 + 1 + 1 2 1 2
c= + + sin ,
Ma14 2 Ma12
cos2
d= .
Ma14

Let

1 1
Q= c b 2,
3 9
1 1 3
R = (b c 3 d) b ,
6 27
D = Q 3 + R 2,

Demonstrate that the oblique shock solution only exists for D < 0 [i.e., when
Equation (15.196) possesses three real roots.] Show that the strong shock so-
lution, s , and the weak shock solution, w , are given by


s
s = tan1 ,
1 s


1 w
w = tan ,
1 w
Two-Dimensional Compressible Inviscid Flow 473

where
1 
s = b + 2 Q cos ,
3
1   
w = b Q cos 3 sin ,
3

1 1 D

= tan + .

3 R

Here, = 0 if R 0, and = if R < 0.


15.2 Assuming that information propagates with respect to a two-dimensional su-
personic flow pattern at the local sound speed, show that, in order for the flow
at some point P to aect the flow at some other point Q, the latter point must
lie between the (+) and () characteristics that pass through P.
15.3 Show that for a weak oblique shock with ,
1 p
,
1 p1


T 1 p
,
T1 p1

3
S ( + 1) p
,
R 12 2 p1
where = 2 1 , et cetera, and


p Ma12
 .
p1
Ma 2 1
1

Here, is the wave angle, the Mach angle,  1 the deflection angle, the
ratio of specific heats, R the specific gas constant, and Ma1 the upstream Mach
number. Furthermore, p1 , 1 , T 1 , and S1 are the upstream pressure, density,
temperature, and specific entropy, respectively, whereas p2 , 2 , T 2 , and S2 are
the corresponding downstream quantities. Show, also, that

2

Ma
1 + [( 1)/2] Ma1
 .
Ma1 Ma1 1
2

15.4 Show that for a weak oblique shock

1 + ,
2 ,
474 Theoretical Fluid Mechanics

Mach line

Ma1


wall
|2|

wall

Figure 15.13
Flow of Mach number Ma1 over convex corner of deflection angle |2 |.

where   1. Here, 1 and 2 are the Mach angles upstream and downstream
of the shock front, respectively. Moreover, is the wave angle, and  1
the deflection angle. Hence, deduce that the shock front subtends the same
angle, , with the Mach lines upstream and downstream of it. In other words,
the shock position is the average of the Mach line positions on either side
of it. Consider supersonic flow incident on a wedge of small nose angle, with
an afterbody, as illustrated in Figure 15.9(a). Assume that the shock front is
attached to the apex of the wedge, and that the flow downstream of the shock
is supersonic. Use the result just proved to show that the shape of the shock
front in the region of attenuation by expansion (i.e., in the region in which
the shock front is intersected by Mach lines emanating from the shoulder) is
parabolic. [Hint: Use the well-known optical result that a parabolic mirror
perfectly focuses a parallel beam of light rays.] (Leipmann & Roshko 1957.)

15.5 Show that, to second order in the deflection angle, , the relative change in
pressure across a weak oblique shock is written


p Ma1
2
Ma12  
 + ( + 1) Ma14 4 (Ma12 1) 2 ,
p1 4 (Ma1 1) 2
Ma12 1
2

where p = p2 p1 . Here, p1 is the upstream pressure, p2 the downstream


pressure, Ma1 the upstream Mach number, and the ratio of specific heats.
(Leipmann & Roshko 1957.)

15.6 An ideal gas of pressure, density, temperature, and Mach number p1 , 1 , T 1 ,


and Ma1 , respectively, flows over a convex corner that turns through an angle
Two-Dimensional Compressible Inviscid Flow 475

|2 |, as shown in Figure 15.13. Consider a particular Mach line in the Prandtl-


Mayer expansion fan that subtends an angle with the continuation of the
upstream wall, as shown in the figure. Let p, , T , Ma, ||, , and be the
pressure, density, temperature, Mach number, magnitude of the deflection an-
gle, Mach angle, and Prandtl-Mayer function, respectively, on the Mach line
in question. Furthermore, let

1/2
+1
= (Ma1 ) + z,
2 1


1 1/2
1

z1 = tan (Ma1 1) ,
2 1/2
+1
where is the ratio of specific heats. Show that, inside the fan,

1 1/2 tan z
= cot (Ma1 1)
2
,
tan z1
 
1/2
1 1 1/2 tan z +1
|| = tan (Ma1 1)
2 1/2
tan (Ma1 1)
2
+ (z z1 ),
tan z1 1
tan 2 z
Ma 2 = 1 + (Ma12 1) ,
tan 2 z1
 
1 1 1/2 tan z
= (Ma1 ) + tan (Ma1 1) 2 1/2
tan (Ma1 1)
2
tan z1

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

2 /(1)
cos z
p = p1 ,
cos z1

2/(1)
cos z
= 1 ,
cos z1

2
cos z
T = T1 ,
cos z1
where z1 z z2 . Here, z2 is defined implicitly by ||(z2 ) = |2 |. Demonstrate
that the fan extends over the range of angles 2 1 , where
 
1 = cot1 (Ma12 1)1/2 ,

1/2
+1
2 = 1 (z2 z1 ).
1

Show that , p, 0, Ma ,

1/2
+ 1
max 1 ,
2 1
476 Theoretical Fluid Mechanics

incident shock
reected shock

Ma1
Ma2
Ma3

1 1
wall
Figure 15.14
Reflection of oblique shock by wall. Here, Ma1 , Ma2 , et cetera, are Mach numbers.

and
|| |max | max (Ma1 ),
in the limit that z /2. Hence, deduce that if |2 | > |max | then the the fan
only extends over the region |max | 1 , and the region |2 | < <
|max | is occupied by a vacuum (i.e., a gas with zero pressure and density).
Assuming that ||  1, show that

1/2
2 1 Ma12 1
|| (z z1 ),
1 +1 Ma12

1 Ma12
1 ||,
2 Ma12 1
and


p
Ma 2

 1
||,
p1 Ma 2 1
1

1 p
,
1 p1
T 1 p
,
T1 p1

Ma 1 + [( 1)/2] Ma12 p
,
Ma1 Ma12 p1

where p = p p1 , et cetera. Of course, the previous four relations are the


Two-Dimensional Compressible Inviscid Flow 477

same as those for a weak shock. (See Exercise 15.3.) Why is this not surpris-
ing? (Hint: The jump in specific entropy across a weak shock is third order in
the deflection angle.) Deduce that to second order in the deflection angle,


p Ma1 Ma12  
2

|| + ( + 1) Ma 4
4 (Ma 2
1) || 2
p1 4 (Ma 2
1) 2 1 1
Ma1 1
2 1

for a weak Prandtl-Mayer fan. (See Exercise 15.5.)

15.7 If an oblique shock is intercepted by a wall then it is reflected, as illustrated in


Figure 15.14. Calculate 1 1 , assuming that the shocks are suciently weak
that the approximate expressions of Section 15.4 can be used. Demonstrate
that, in this limit,


+ 1 Ma12
1 1 ,

1 1 2
4 Ma1 1
2

where 1  1 is the deflection angle of the incident shock, and the ratio
of specific heats. Show that 1 1 > 0 if 1 < Ma1 < [4/(3 )]1/2 and
1 1 < 0 if Ma1 > [4/(3 )]1/2 . Demonstrate that Ma2 Ma1 (1 ) and
Ma3 Ma1 (1 2 ), where


1 + [( 1)/2] Ma12
=  1 .
Ma12 1

15.8 Figure 15.15 shows a situation in which two oblique shocks of the same family
[in this case, the () family], produced by successive concave corners in a wall,
merge together to form a single stronger shock [of the () family]. Assuming
that the shocks are suciently weak that the approximate expressions of Sec-
tion 15.4 can be used (which implies that 1  1 and 2  1), demonstrate
that
1 + 2 ,
where


+1 Ma12
1 = 1 ,
4 Ma12 1


+ 1 Ma12
2 =
2 ,
4 Ma12 1

and is the ratio of specific heats. Show that

3 1 + 2 ,
478 Theoretical Fluid Mechanics

shock Ma4

Ma1
slipstream
3

Ma3
shock shock

Ma1
Ma2
1
2
wall 1

Figure 15.15
Merging of two oblique shocks of the same family produced by successive concave
corners of deflection angles 1 and 2 . Here, Ma1 , Ma2 , et cetera, are Mach numbers.

and also that Ma2 Ma1 (1 1 ) and Ma3 Ma4 Ma1 (1 1 2 ), where

2
1 + [( 1)/2] Ma1
1 =  1 ,

Ma12 1


1 + [( 1)/2] Ma12
2 =  2 .
Ma12 1

Demonstrate that the strength of the merged shock is approximately the sum
of the strengths of the two component shocks, and, hence, that the pressures
on either side of the slipstream shown in the figure are equal (at least, to first
order in 1 and 2 ). Finally, show that

S4 S3 ( + 1) Ma16
1 2 (1 + 2 ),
R 4 (Ma12 1)3/2

where S is specific entropy, and R the specific gas constant. It is, thus, clear
that the specific entropy is not quite the same on either side of the slipstream.

15.9 If two shocks of opposite families intersect then they pass through one another,
but are slightly bent in the process, as illustrated in Figure 15.16. Assuming
that the shocks are suciently weak that the approximate expressions of Sec-
tion 15.4 can be used (which implies that 1  1, 2  1, et cetera), show
Two-Dimensional Compressible Inviscid Flow 479

(+) ()

1 Ma2 Ma3
2

Ma1

2
Ma3
Ma2
1

() (+)

Figure 15.16
Crossing of two oblique shocks of dierent families. Here, Ma1 , Ma2 , et cetera, are
Mach numbers, and 1 , 2 , et cetera, are deflection angles.

that


p2 p1 Ma1 2
 1 ,

Ma1 1
p1 2



p3 p1 Ma12
 (1 + 2 ),
p1 Ma 2 1
1


p2 p1 Ma1 
2
 1 ,

Ma12 1
p1


p3 p1 Ma12 
 (1 + 2 ),
p1 Ma 2 1
1

where p denotes pressure. Hence, deduce that


2 1 ,
2 1 ,
1 1 .
Show, that the respective strengths of the two shocks are unaected by the
intersection (at least, to first order in the deflection angles).
480 Theoretical Fluid Mechanics

15.10 Show that Equation (14.66) can be written in the form




1 c2
sin +
2
cos2 = 12 ,
+1 w
where is the Mach angle, the ratio of specific heats, c1 the sound speed at
the sonic point, and w the flow speed. Deduce that


dw b2 1
Ma 2 1 = 2 d,
w b + tan2
where
1
b2 = ,
+1
and, hence, that
 

dw 1 1
Ma 2 1 = tan1 tan + const.
w b b
Finally, demonstrate that

1/2
1/2
 +1 1
1
Ma
(Ma 1)
dw
(Ma) Ma 1
2
= tan 2
1 w 1 +1
$ 1/2 %
tan1 Ma 2 1 .

(Leipmann & Roshko 1957.)


15.11 Show that for a thin, symmetrical airfoil, with zero angle of attack, whose
profile is a lens defined by two circular arcs, the drag coecient is
16 $ t %2
CD =  ,
c
3 Ma12 1

where Ma1 is the upstream Mach number, t the maximum thickness, and c
the chord-length. Demonstrate that for a given thickness ratio, t/c, the airfoil
with the minimum drag is a symmetric diamond profile. (Leipmann & Roshko
1957.)
15.12 Prove that on a supersonic swept-back wing of infinite span the thin-airfoil
pressure coecient, (15.50), is multiplied by the sweepback factor,
1
,
1 n2
where
tan
n=  .
Ma12 1
Two-Dimensional Compressible Inviscid Flow 481

Here, is the sweepback angle, and Ma1 the upstream Mach number. [Hint:
Resolve Ma1 into components normal and parallel to the leading edge. The
flow may then be studied in the plane normal to the leading edge using standard
thin-airfoil theory.] (Leipmann & Roshko 1957.)
15.13 Consider the problem of subsonic flow past a wave-shaped wall that was dis-
cussed in Section 15.13. Show that if the flow is bounded by a second wall
that lies at y = b (where b > 0) then
U cosh[m (b y)]
(x, y) = cos( x) .
m sinh(m b)
Show, on the other hand, that if the flow is bounded by a free surface at y = b
(where p = p ) then
U sinh[m (b y)]
(x, y) = cos( x) .
m cosh(m b)
(Leipmann & Roshko 1957.)
482 Theoretical Fluid Mechanics
A
Vectors and Vector Fields

A.1 Introduction

This appendix outlines those aspects of vector algebra, vector calculus, and vector
field theory that are useful in the study of fluid dynamics. Most of the material ap-
pearing in this appendix is reproduced from Fitzpatrick 2008. For more information
on vector algebra, vector calculus, and vector field theory see Schey 1992 and Milne-
Thomson 2011.

A.2 Scalars and Vectors

Many physical entities (e.g., mass and energy) are entirely defined by a numerical
magnitude (expressed in appropriate units). Such entities, which have no directional
element, are known as scalars. Moreover, because 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, and 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.

483
484 Theoretical Fluid Mechanics

A.3 Vector Algebra



Suppose that the displacements PQ and QR, shown in Figure A.2, represent the
vectors a and b, respectively. 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 previous
equation is equivalent to
c = a + b = b + a. (A.3)
We conclude that the addition of vectors is commutative. It can also be shown that
the associative law holds: that is,

a + (b + c) = (a + b) + c. (A.4)

The null vector, 0, is represented by a displacement of zero length and arbitrary


direction. Because 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: that is,

a + (a) = 0. (A.6)

Figure A.1
A vector.
Vectors and Vector Fields 485

b
a
c=a+b
Q
S

a
b

Figure A.2
Vector addition.

c=ab

b a

c
b

Figure A.3
Vector subtraction.

We can also define the dierence 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, because n a = |n| (a), it follows
that n a is a vector whose magnitude is |n| times that of a, and whose direction is
486 Theoretical Fluid Mechanics

y P

O x

Figure A.4
A right-handed Cartesian coordinate system.

opposite to a. These 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 conventional 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).
Vectors and Vector Fields 487

According to the three-dimensional generalization of the Pythagorean theorem,


the distance OP |r| = r is given by

r = x 2 + y 2 + z 2. (A.13)

By analogy, the magnitude of a general vector a takes the form



a = a x2 + ay2 + az2 . (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)

A.5 Coordinate Transformations


A Cartesian coordinate system allows position and direction in space to be repre-
sented 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 dierent origins, and dierent
orientations of the coordinate axes. In general, a given vector a will have dierent
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, because vectors are

represented by moveable line elements in space (i.e., in Figure A.2, PQ and S R rep-
resent the same vector), it follows that the components of a general vector are not
aected 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)



y = x sin + y cos , (A.18)
z = z. (A.19)

Consider the vector displacement r OP. Note that this displacement is represented
488 Theoretical Fluid Mechanics

y
y z

x

x
O

Figure A.5
Rotation of the coordinate axes about Oz.

by the same symbol, r, in both coordinate systems, because the magnitude and di-
rection of r are manifestly independent of the orientation of the coordinate axes. The
coordinates of r do depend on the orientation of the axes: that is, 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 Equa-
tions (A.17)(A.19): that is,

a x = a x cos + ay sin , (A.20)


ay = a x sin + ay cos , (A.21)
az = az . (A.22)

Moreover, there are similar transformation rules for rotation about Ox and Oy. Equa-
tions (A.20)(A.22) eectively constitute the definition of a vector: in other words,
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 Equa-
tions (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 trans-
form in accordance with Equations (A.20)(A.22). Of course, scalar quantities are
invariant under 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 and acceleration), automat-
ically satisfy Equations (A.20)(A.22). There are, however, other physical quantities
that have both magnitude and direction, but are not obviously related to displace-
ments. We need to check carefully to see whether these quantities are really vectors.
(See Sections A.7 and A.9.)
Vectors and Vector Fields 489

A.6 Scalar Product


A scalar quantity is invariant under all possible rotational transformations. The in-
dividual components of a vector are not scalars because they change under transfor-
mation. 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
the
coordinate

45 about Oz. In the new coordinate system, a (1/ 2, 1/ 2, 0) and
axes through
b (1/ 2, 1/ 2, 0), giving a % b = 1/2. Clearly, a % b is not invariant under
rotational transformation, so the previous 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 Equa-
tions (A.20)(A.22), a b takes the form

a b = (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
is easily demonstrated that it is also invariant under rotation about Ox and Oy. We
conclude that a b is a true scalar, and that the definition (A.24) is a good one.
Incidentally, a b is the only simple combination of the components of two vectors
that transforms like a scalar. It is readily shown that the dot product is commutative
and distributive: that is,

a b = b a,
a (b + c) = a b + a c. (A.26)

The associative property is meaningless for the dot product, because we cannot have
(a b) c, as a b is scalar.
We have shown that the dot product a b is coordinate independent. But what is
the geometric significance of this property? In the special case where a = b, we get

a b = a x2 + ay2 + az2 = |a| 2 = a 2 . (A.27)

So, the invariance of a a is equivalent to the invariance of the magnitude of vector a


under transformation.
490 Theoretical Fluid Mechanics

b
ba


O a A

Figure A.6
A vector triangle.

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)

where (AB) denotes the length of side AB. It follows that

a b = |a| |b| cos . (A.30)

In this case, the invariance of ab 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:

ab
cos = . (A.31)
|a| |b|

The work W performed by a constant force F that 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 that 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.
Vectors and Vector Fields 491

Figure A.7
A vector area.

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
(because 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 sur-
faces, 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 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: that is, 
S= Si . (A.34)
i
492 Theoretical Fluid Mechanics

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

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 dierent 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, because 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 that 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), b = (1, 0, 0). In this case,
a b = 0. However,
if we rotate
the 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, previous 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)

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
Vectors and Vector Fields 493

thumb

ab

middle finger
b

a index finger

Figure A.8
The right-hand rule for cross products. Here, is less that 180.

correctly under rotation about Ox and Oy. Thus, a b is a proper vector. Inciden-
tally, 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
anti-commutative,
a b = b a, (A.39)
distributive,
a (b + c) = a b + a c, (A.40)
but is not associative,
a (b c)  (a b) c. (A.41)
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,

a a b = 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)
494 Theoretical Fluid Mechanics

Thus,

|a b| = |a| |b| sin , (A.44)

where is the angle subtended between a and b. Clearly, a a = 0 for any vector,
because 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). In other words, 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, as illustrated in 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: that is,

l = r p. (A.47)

a
S
b b

Figure A.9
A vector parallelogram.
Vectors and Vector Fields 495



P

O Q
r sin

Figure A.10
A torque.

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 prob-
lem is that the addition of rotations is not commutative, whereas vector addition
is commuative. Figure A.11 shows the eect of applying two successive 90 rota-
tions, 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 dierent 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 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 previous 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)
where
= x e x + y ey + z ez . (A.50)
496 Theoretical Fluid Mechanics

x
y
z -axis x-axis

x-axis z -axis

Figure A.11
Eect of successive rotations about perpendicular axes on a six-sided die.

Clearly, we can define a rotation vector, , but it only works for small angle rotations
(i.e., suciently small that the small angle approximations of sine and cosine are
good). According to the previous equation, a small z-rotation plus a 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,

= lim , (A.51)
t0 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 that precesses about the origin with
some angular velocity is
da
= a. (A.52)
dt

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 and Vector Fields 497

c
b

Figure A.12
A vector parallelepiped.

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:
that is,
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, (A.58)
so
rbc
= . (A.59)
abc
Analogous expressions can be written for and . The parameters , , and are
uniquely determined provided a b c  0: that is, provided the three vectors are
non-coplanar.
498 Theoretical 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(bc)  (ab)c. In fact, it can be demonstrated
that
a (b c) (a c) b (a b) c (A.60)
and
(a b) c (a c) b (b c) a. (A.61)
Let us try to prove the first of the previous 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: in fact, b c (0, 0, b x cy ). Hence, a (b c) lies in the x-y plane: in fact,
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
da a(t + t) a(t)
= lim . (A.63)
dt t0 t
When written out in component form this becomes


da da x day daz
, , . (A.64)
dt dt dt dt
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
da d db
= b+ . (A.66)
dt dt dt
Vectors and Vector Fields 499

Q
y f

P
O x P Q l
.

Figure A.13
A line integral.

Moreover, it is easily demonstrated that


d da db
(a b) = b+a , (A.67)
dt dt dt
and
d da db
(a b) = b+a . (A.68)
dt dt dt
Hence, it can be seen that the laws of vector dierentiation are analogous to those in
conventional calculus.

A.13 Line Integrals


Consider a two-dimensional function f (x, y) that 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
 Q
f (x, y) dl = Area under the curve, (A.69)
P

where dl = dx 2 + dy 2 .
For example, consider the integral of f (x, y) = x y 2 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,
 Q  1 2
x y 2 dl = x 3 2 dx = . (A.70)
P 0 4
500 Theoretical Fluid Mechanics

y
Q = (1, 1)

1 2

2
P = (0, 0) x

Figure A.14
An example line integral.

The integration along route 2 gives


    1 
Q 1
 
x y 2 dl = x y 2 dx + x y 2 dy
P 0 y=0 0  x=1
 1
1
=0+ y 2 dy = . (A.71)
0 3

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: that
is,  Q
! "
f (x, y) dx + g(x, y) dy . (A.72)
P

For example, consider the integral


 Q  
y dx + x 3 dy (A.73)
P

along the two routes indicated in Figure A.15. Along route 1, we have x = y + 1 and
dx = dy, so
 Q    1   17
y dx + x 3 dy = y dy + (y + 1)3 dy = . (A.74)
P 0 4

Along route 2,
    2 
Q   1
  7
y dx + x dy =3
x dy +
3
y dx = . (A.75)
P 0  x=1 1 y=1 4

Again, the integral depends on the path of integration.


Vectors and Vector Fields 501

y
2 Q = (2, 1)

2 1

O P = (1, 0) x

Figure A.15
An example line integral.

Suppose that we have a line integral that does not depend on the path of integra-
tion. It follows that
 Q
( f dx + g dy) = F(Q) F(P) (A.76)
P

for some function F. Given F(P) for some point P in the x-y plane,
 Q
F(Q) = F(P) + ( f dx + g dy) (A.77)
P

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:
 Q  Q
( f dx + g dy) = dF(x, y) = F(Q) F(P). (A.78)
P P

Thus,
dF(x, y) = f (x, y) dx + g(x, y) dy. (A.79)

For instance, if F = x 3 y then dF = 3 x 2 y dx + x 3 dy and


 Q    Q
3 x 2 y dx + x 3 dy = x 3 y P (A.80)
P

is independent of the path of integration.


It is clear that there are two distinct types of line integralthose that depend only
on their endpoints and not on the path of integration, and those that 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.)
502 Theoretical Fluid Mechanics

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
 Q  Q
A dr = (A x dx + Ay dy + Az dz). (A.81)
P P

For instance, if A is a force-field then the line integral is the work done in going from
P to Q.
For example, consider the work done by a repulsive inverse-square central field,
F = r/|r 3 |. 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
a
1 1 1
W= 2 dx = = . (A.82)
x x a

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,
because F is perpendicular to dr. In the second part,
 0
0
y dy 1 1
W= = = . (A.83)
(a 2 + y 2 )3/2 (y 2 + a 2 )1/2 a

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
  
f (x, y, z) dS = lim f (x, y, z) S i (A.84)
S S i 0
i

is a surface integral. For instance, the volume of water in a lake of depth D(x, y) is
 
V= D(x, y) dS . (A.85)
Vectors and Vector Fields 503

Q P
x
O a 1

Figure A.16
An example vector line integral.

To evaluate this integral, we must split the calculation into two ordinary integrals.
The volume in the strip shown in Figure A.17 is
 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: that is,
 y2  x2 (y)  
V= dy D(x, y) dx D(x, y) dx dy. (A.87)
y1 x1 (y) S

Of course, the integral can be evaluated by taking the strips the other way around:
that is,  x2  y2 (x)
V= dx D(x, y) dy. (A.88)
x1 y1 (x)
Interchanging the order of integration is a very powerful and useful trick. But great
care must be taken when evaluating the limits.
For example, consider  
x y 2 dx dy, (A.89)
S
where S is shown in Figure A.18. Suppose that we evaluate the x integral first:

 1y 2 1y
x y2
dy x y dx = y dy
2 2
= (1 y)2 dy. (A.90)
0 2 0 2
Let us now evaluate the y integral:
 1
2
y y4 1
y3 + dy = . (A.91)
0 2 2 60
504 Theoretical Fluid Mechanics

y
y2

dy

y1

x1 x2 x

Figure A.17
Decomposition of a surface integral.

We can also evaluate the integral by interchanging the order of integration:


 1  1x  1
x 1
x dx y dy =
2
(1 x)3 dx = . (A.92)
0 0 0 3 60
In some cases, a surface integral is just the product of two separate integrals. For
instance,  
x 2 y dx dy (A.93)
S
where S is a unit square. This integral can be written
 1  1
 1
 1
1 1 1
dx x y dy =
2 2
x dx y dy = = , (A.94)
0 0 0 0 3 2 6

because 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 very many infinitesimal surfaces is
   
v dS = lim v cos dS , (A.95)
S dS 0

where is the angle subtended between the normal to the surface and the flow veloc-
ity.
Vectors and Vector Fields 505

(0, 1)

1y =x

(0, 0) (1, 0) x

Figure A.18
An example surface integral.

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 incom-
pressible fluid flow between two surfaces S 1 and S 2 that end on the same rim. (See
Figure A.23.) The volume between the surfaces is constant, so what goes in must
come out, and    
v dS = v dS. (A.96)
S1 S2

It follows that  
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
  
f (x, y, z) dV, (A.98)
V

where V is some volume, and dV = dx dy dz is a small volume element. The volume


element is sometimes written d 3 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
506 Theoretical Fluid Mechanics

line joining the centroid to the apex. In fact, the height of the center of mass is given
by    )  
z= z dV dV. (A.99)

The bottom integral is just the volume of the pyramid, and can be written
    a  (az)/2  (az)/2  a
dV = dz dy dx = (a z)2 dz
0 (az)/2 (az)/2 0
 a  a 1
= (a 2 2 a z + z 2 ) dz = a 2 z a z 2 + z 3 /3 = a 3 . (A.100)
0
0 3
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
    a  (az)/2  (az)/2  a
z dV = z dz dy dx = z (a z)2 dz
0 (az)/2 (az)/2 0
 a  a
= (z a 2 2 a z 2 + z 3 ) dz = a 2 z 2 /2 2 a z 3 /3 + z4 /4
0
0
1 4
= a . (A.101)
12
Thus, )
1 4 1 3 1
z = a a = a. (A.102)
12 3 4
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) that represents (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 dis-
tance 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 vicin-
ity 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
Vectors and Vector Fields 507

y contours of h(x, y)

high

dr
direction of steepest ascent
P

low
O x

Figure A.19
A two-dimensional gradient.

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 direc-
tion is the direction of steepest ascent. The cos variation exhibited in the previous
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
grad h , . (A.104)
x y

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 h
= , = . (A.106)
x y

For small dx = x x0 and dy = y y0 , the function h is coincident with the tangent


plane. It follows that
h h
dh = dx + dy. (A.107)
x y
508 Theoretical Fluid Mechanics

But, grad h (h/x, h/y) and dr (dx, dy), so


dh = grad h dr. (A.108)
Incidentally, the previous equation demonstrates that grad h is a proper vector, be-
cause 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
grad T , , . (A.109)
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 neighboring point
oset by dr (dx, dy, dz) is
T T T
dT = dx + dy + dz. (A.110)
x y z
In vector form, this becomes
dT = grad T dr. (A.111)
Suppose that dT = 0 for some dr. It follows that
dT = grad T dr = 0. (A.112)
So, dr is perpendicular to grad T . Because 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
 Q  Q
dT = grad T dr = T (Q) T (P). (A.113)
P P
Q
This integral is clearly independent of the path taken between P and Q, so P grad T
dr must be path independent.
Q
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
 Q
A dr = V(Q), (A.114)
P
Vectors and Vector Fields 509

T = constant grad T
dr

isotherms

Figure A.20
Isotherms.

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
 Q+dx
V(Q + dx) = V(Q) + A dr = V(Q) + A x dx. (A.115)
Q

Hence,
V
= Ax, (A.116)
x

with analogous relations for the other components of A. It follows that

A = grad V. (A.117)

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

A dr = 0, (A.118)

#
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 going around a closed cycle, so A dr  0.
510 Theoretical Fluid Mechanics

A.19 Grad Operator


It is useful to define the vector operator



, , , (A.119)
x y z

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
grad f = f , , . (A.120)
x y z

For two scalar fields and ,

grad ( ) = grad + grad (A.121)

can be written more succinctly as

( ) = + . (A.122)

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 = x cos y sin , (A.123)


 
y = x sin + y cos , (A.124)

z=z. (A.125)

Now,




x y z

= 
+ 
+ 
, (A.126)
x x y ,z x x y ,z y x y ,z z

giving

= cos + sin , (A.127)
x x y
and
x = cos x + sin y . (A.128)

It can be seen, from Equations (A.20)(A.22), that the dierential operator trans-
forms in an analogous manner to a vector. This is another proof that f is a good
vector.
Vectors and Vector Fields 511
y + dy

y
z + dz

z
y

z x
x x + dx

Figure A.21
Flux of a vector field out of a small box.

A.20 Divergence
#
Let us start with a vector field A(r). Consider S A dS over some closed surface
S , where dS denotes an outward pointing surface element. 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. In fact,  
A dS = A dS = A S = 0, (A.129)

because the vector area S of a closed surface is zero.


Suppose, now, that A is 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 A x
A x (x + dx) dy dz A x (x) dy dz = dx dy dz = dV, (A.130)
x x
where dV = dx dy dz is the volume element. (See Figure A.21.) There are analo-
gous contributions from the sides normal to the y- and z-axes, so the total of all the
contributions is 

A x Ay Az
A dS = + + dV. (A.131)
x y z
The divergence of a vector field is defined

A x Ay Az
div A = A = + + . (A.132)
x y z
512 Theoretical Fluid Mechanics

interior contributions cancel


.

exterior contributions survive

Figure A.22
The divergence theorem.

Divergence is a good scalar (i.e., it is coordinate independent), because it is the dot


product of the vector operator with A. The formal definition of A is
#
A dS
A = lim . (A.133)
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. This states that for any volume V surrounded by a closed surface S ,
 
A dS = A dV, (A.134)
S V

where dS is an outward pointing volume element. The proof is very straightforward.



We divide up the volume into very many infinitesimal 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 Equa-
tion
 (A.131) for each cube individually. This tells us that the summation is equivalent
to A dV over the whole volume. Thus, the integral of AdS 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,

A dS = 0 (A.135)
S

for any closed surface S . So, for two surfaces, S 1 and S 2 , on the same rim,
 
A dS = A dS, (A.136)
S1 S2
Vectors and Vector Fields 513

S S2

S1
rim

Figure A.23
Two surfaces spanning the same rim (right), and the equivalent closed surface (left).

as illustrated in 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
integral depends on the rim, but not on the nature of the surface that spans it. On the
other hand, if A  0 then the integral depends on both the rim and the surface.
#
Consider an incompressible fluid whose velocity field is v. It is clear that v
dS = 0 for any closed surface, because 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 dierential relation:

v x vy vz
+ + = 0. (A.137)
x y z

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

 
A dS = A dV. (A.138)
S V

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.
514 Theoretical Fluid Mechanics

A.21 Laplacian Operator


So far we have encountered



= , , , (A.139)
x y z

which is a vector field formed from a scalar field, and


A x Ay Az
A= + + , (A.140)
x y z
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)
#
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



h dS = c T dV , (A.142)
S t

where c is the specific heat. It follows from the divergence theorem that
T
h = c . (A.143)
t

1 2

Figure A.24
Divergent lines of force.
Vectors and Vector Fields 515

Taking the divergence of both sides of Equation (A.141), and making use of
Equation (A.143), we obtain
T
( T ) = c . (A.144)
t
If is constant then the previous equation can be written
c T
(T ) = . (A.145)
t
The scalar field (T ) takes the form




T T T
(T ) = + +
x x y y z z
2 T 2 T 2T
= + + 2 T. (A.146)
x 2 y 2 z 2
Here, the scalar dierential operator
2 2 2
2 + 2+ 2 (A.147)
x 2 y z
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 re-
duces to 2 T/x 2 . Now, 2 T/x 2 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,
2T 2T
2T = + . (A.148)
x 2 y 2
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/y 2 is large and
positive, whereas 2 T/x 2 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:
c T
2T = . (A.149)
t
It is clear that if 2 T is positive in some small region then the value of T there is less
than the local average value, so T/t > 0: that is, the region heats up. Likewise,
if 2 T is negative then the value of T is greater than the local average value, and
heat flows out of the region: that is, T/t < 0. Thus, the previous heat conduction
equation makes physical sense.
516 Theoretical Fluid Mechanics

A.22 Curl
Consider a vector field A(r),
# and a loop that lies in one plane. The integral of A
around this loop is written A dr, where# dr is a line element of the loop. 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.
For a small loop, we expect A dr to # be proportional to the area of the loop.
Moreover, for a fixed-area loop, we expect Adr to depend on#the orientation of the
loop. One particular orientation will give the maximum value: 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
#
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.
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, as shown in Figure A.25.
The contribution from sides 1 and 3 is
Az
Az (y + dy) dz Az (y) dz = dy dz. (A.151)
y
The contribution from sides 2 and 4 is
Ay
Ay (z + dz) dy + Ay (z) dy = dy dz. (A.152)
y
So, the total of all contributions gives


Az Ay
A dr = dS , (A.153)
y z
where dS = dy dz is the area of the loop.
# small) loop in the y-z plane. We can divide it
Consider a non-rectangular (but still
into rectangular elements, and 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 

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


A x Az
A dr = dS y . (A.155)
z x
Vectors and Vector Fields 517

z + dz 4

1 3

z
y 2 y + dy

Figure A.25
A vector line integral around a small rectangular loop in the y-z plane.

Finally, if the loop is in the x-y plane then dS = (0, 0, dS z ) and




Ay A x
A dr = dS z . (A.156)
x y
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,
   
A dr = A dr1 + A dr2 + A dr3 , (A.157)

giving 
A dr = curl A dS = |curl A| |dS| cos , (A.158)

where

Az Ay A x Az Ay A x
curl A = , , , (A.159)
y z z x x y
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, because 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)
518 Theoretical Fluid Mechanics
z

dS
2 1

x y

Figure A.26
Decomposition of a vector area into its Cartesian components.

[See Equation (A.52).] Let us evaluate


# v on the axis of rotation. The x-component
is proportional to the integral v dr around a loop in the y-z plane.
# This is plainly
zero. Likewise, the y-component 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 ).
O 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, because 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: that is, 2 . Hence, a vector field with A = 0
everywhere is said to be irrotational.
Another important result of vector field theory is the curl theorem:
 
A dr = A dS, (A.163)
C S

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 Equa-
tion (A.158) for each of the small loops, we can see that the contribution from all of
Vectors and Vector Fields 519

the loops is also equal to the integral of A dS across the whole surface. This
proves the theorem.

One immediate consequence of the curl theorem is that A is incompressible.


Consider any two surfaces, S 1 and S 2 , that share the same rim. (See Figure A.23.) It
# curl theorem that A dS is the same for both surfaces. Thus,
is clear from the
it follows that A #dS = 0 for anyclosed surface. However, we have from the
divergence theorem that A dS = ( A) dV = 0 for any volume. Hence,

( A) 0. (A.164)

So, A is a solenoidal field.

#
We have seen that for a conservative field 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 )
520 Theoretical Fluid Mechanics

(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) = A ( B) + B ( A) + (A ) B + (B ) A, (A.171)
= 2 , (A.172)
A = 0, (A.173)
( A) = A + A , (A.174)
(A B) = B A A B, (A.175)
= 0, (A.176)
( A) = ( A) A, 2
(A.177)
( A) = A + A, (A.178)
(A B) = ( B) A ( A) B + (B ) A (A ) B. (A.179)

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.
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|.
Vectors and Vector Fields 521

A.5 Find the scalar product a b and the vector product a b when

(a) a = e x + 3 ey ez , b = 3 e x + 2 ey + ez ,
(b) a = e x 2 ey + ez , b = 2 e x + 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.
522 Theoretical Fluid Mechanics

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
bc
A= ,
abc
plus cyclic permutations, are said to be reciprocal vectors. Show that

a = (B C)/(A B C),

plus cyclic permutations.


A.12 In the notation of the previous exercise, 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 .
#
A.13 Evaluate A dr for
x e x + y ey
A= 
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 x 3 + 3 x 2 y 2 , 2 x 3 y + 6 y, 6).

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 y 2 z 2 y 2 , y 3 z 2 + x 2 y, 3 x 2 x 2 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).
Vectors and Vector Fields 523

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.
524 Theoretical Fluid Mechanics
B
Cartesian Tensors

B.1 Introduction
As we saw in Appendix A, many physical entities can be represented mathemati-
cally 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 par-
ticular types of a more general class of mathematical constructs known as 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 nec-
essary 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. For more information on Cartesian tensors see Jeries
1961, Riley 1974, and Temple 2004.

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:
for instance, 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: for instance,
ai j , i j . Moreover, an nth-order tensor is represented as a variable with n subscripts:
for instance, 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.
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

525
526 Theoretical Fluid Mechanics

tensor is not simply another name for an n-dimensional matrix. A matrix is merely 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 previous 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: that is, 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: for instance, the second-order tensor ai b j
is the outer product of the two first-order tensors ai and bi . Likewise, when two ten-
sors are multiplied together in a manner that involves contraction then the resulting
tensor is called an inner product: for instance, 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 . It can be seen
from Equation (B.2) that the scalar product of two vectors is equivalent to the inner
product of the corresponding first-order tensors.
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 previous expression can be written more compactly as
(a b)i = i jk a j bk . (B.6)
Here,


+1 if i, j, k is an even permutation of 1, 2, 3


i jk =
1 if i, j, k is an odd permutation of 1, 2, 3 (B.7)


0 otherwise
Cartesian Tensors 527

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: for
instance, 113 = 212 = 0. Furthermore, it follows from Equation (B.7) that

i jk =  jki = ki j = k ji =  jik = ik j . (B.8)

It is helpful to define the second-order identity tensor (also known as the Kroe-
necker delta tensor), 
1 if i = j
i j = . (B.9)
0 otherwise
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)

et cetera.
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 previous expression, let us consider the various
cases that arise. As is easily seen, the right-hand side of Equation (B.16) takes the
values

+1 if j = l and k = m  j, (B.17)
1 if j = m and k = l  j, (B.18)
0 otherwise. (B.19)

Moreover, in each product on the left-hand side of Equation (B.16), 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). Because 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 (as each  factor would then have repeated
indices, and so be zero). Thus, the left-hand side of Equation (B.16) reproduces
Equation (B.19), as well as the conditions on the indices in Equations (B.17) and
(B.18). The left-hand side also reproduces the values in Equations (B.17) and (B.18)
because 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
528 Theoretical Fluid Mechanics

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 Equation (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 ), (B.21)

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.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 trans-
formation properties under rotation of the coordinate axes. In fact, a scalar is in-
variant under rotation of the coordinate axes. On the other hand, according to Equa-
tions (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)
Cartesian Tensors 529

where
Ri j = i j k ki j (B.26)
is a rotation matrix (which is not a tensor, because 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 Equa-
tions (B.8), (B.10), and (B.12). Likewise, it is easily demonstrated that

Rki Rk j = i j . (B.29)

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 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)
whereas a third-order tensor transforms as

ai jk = Ril R jm Rkn almn . (B.34)

The generalization to higher-order tensors is straightforward. For the case of a scalar,


which is a zeroth-order tensor, the transformation rule is particularly simple: that is,

a = a. (B.35)
530 Theoretical Fluid Mechanics

By analogy with Equation (B.27), the inverse transform is exemplified by

ai jk = Rli Rm j Rnk almn . (B.36)

Incidentally, because 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 dierent
order, because such an equation would only be valid in one particular coordinate
system, and so could not have any physical significance (because 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)

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 previous proof is more rig-
orous than that given 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
Cartesian Tensors 531

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

(aik Ril Rkm alm ) bjk = 0. (B.46)

However, the bi j are arbitrary, so the previous equation can only be satisfied, in gen-
eral, 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: that is, 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: for instance, (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 space:
532 Theoretical Fluid Mechanics

for instance, ai (x). It stands to reason, then, that a tensor field is a set of tensors
associated with every point in space: for instance, 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 fields encountered in physics are smoothly varying and dieren-
tiable. 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 ten-
sor. In order to demonstrate this, we need the transformation rule for the xi , which is
the same as that for a first-order tensor: that is,
xi = Ri j x j . (B.51)
Thus,
xi
= Ri j . (B.52)
x j
It is also easily shown that
xi
= R ji . (B.53)
xj
Now,
ai ai xk (Ril al ) al
 =  = R jk = Ril R jk , (B.54)
x j xk x j xk xk
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 previous argument,
that dierentiating a tensor field increases its order by one: for instance, ai j /xk
is a third-order tensor. The only exception to this rule occurs when dierentiation
and contraction are combined. Thus, ai j /x j is a first-order tensor, because 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 = . (B.55)
xi
The divergence (see Section A.20) of a vector field is a contracted second-order ten-
sor field that transforms as a scalar:
ai
a = . (B.56)
xi
Finally, the curl (see Section A.22) of a vector field is a contracted fifth-order tensor
that transforms as a vector
ak
( a)i = i jk . (B.57)
x j
Cartesian Tensors 533

The previous definitions can be used to prove a number of useful results. For
instance,

2
( )i = i jk = i jk = 0, (B.58)
x j xk x j xk
which follows from symmetry because ik j = i jk whereas 2 /xk x j = 2 /x j xk .
Likewise,

ak ak
( a) = i jk = i jk = 0, (B.59)
xi x j xi x j
which again follows from symmetry. As a final example,

  a j bk
(a b) = i jk a j bk = i jk bk + i jk a j
xi xi xi
ak bk
= bi i jk ai i jk = b ( a) a ( b). (B.60)
x j x j

According to the divergence theorem (see Section A.20),


 
ai
ai dS i = dV, (B.61)
S V xi

where S is a closed surface surrounding the volume V. The previous theorem is


easily generalized to give, for example,
 
ai j
ai j dS i = dV, (B.62)
S V xi

or  
ai j
ai j dS j = dV, (B.63)
S V x j
or even  
a
a dS i = dV. (B.64)
S V xi

B.5 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 (Hodge 1961).
534 Theoretical Fluid Mechanics

The most general second-order isotropic tensor, ai j , is such that

ai j = Rip R jq a pq = ai j (B.65)

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.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)

which reduces to
2 ai j + a ji = a ss i j . (B.69)
Interchanging the labels i and j, and then taking the dierence between the two
equations thus obtained, we deduce that

ai j = a ji . (B.70)

Hence,
a ss
ai j = i j , (B.71)
3
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 previous 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)
Cartesian Tensors 535

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)

The solution to the previous equation must satisfy

aik j = a jik = ak ji = ai jk . (B.84)

This implies, from Equation (B.8), that

ai jk = i jk . (B.85)

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 previous 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)

Let us assume, for the moment, that

ai jss = a ssi j , (B.94)


aisks = a sisk , (B.95)
aissl = a sils . (B.96)
536 Theoretical Fluid Mechanics

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


1
ai jkl + aikl j + ail jk = ai jlk + aik jl + ailk j = ( + + ) (i j kl + ik jl + il jk ). (B.106)
5
This can be seen by swapping the indices k and l in the previous expression. Finally,
substitution into Equation (B.102) yields
ai jkl = i j kl + ik jl + il jk , (B.107)

where
= (4 )/10, (B.108)
= (4 )/10, (B.109)
= (4 )/10. (B.110)

B.6 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 ,
Cartesian Tensors 537

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:

(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 coecients in the previous expression transform under rota-
tion 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).


538 Theoretical Fluid Mechanics

B.5 If
A i 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 previous
form only has a non-trivial solution when (Riley 1974)
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: that is, A ji = Ai j .
(a) Demonstrate that the eigenvalues of Ai j are all real, and that the eigen-
vectors can be chosen to be real.
(b) Show that eigenvectors of Ai j corresponding to dierent eigenvalues
are orthogonal to one another. Hence, deduce that the three eigenvec-
tors 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
i j 2 ui
= 2 ,
x j t


1 uk ul
i j = ci jkl + ,
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 stiness 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.
C
Non-Cartesian Coordinates

C.1 Introduction
Non-Cartesian coordinates are often employed in fluid mechanics to exploit the sym-
metry 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 orthog-
onal curvilinear coordinates. The two most commonly occurring examples of this
class in fluid mechanics are cylindrical and spherical coordinates. (Incidentally, the
Einstein summation convention is not used in this appendix.) For more information
on orthogonal curvilinear coordinates see Milne-Thomson 2011.

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 dierent points in space. Because 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 that 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: that is,

ei e j = i j . (C.1)

In this case, u1 , u2 , u3 are said to constitute an orthogonal coordinate system. Sup-


pose, further, that
e1 e2 e3 = 1 (C.2)
at all points in space, so that u1 , u2 , u3 also constitute a right-handed coordinate

539
540 Theoretical Fluid Mechanics

system. It follows that


ei e j ek = i jk . (C.3)
Finally, a general vector A, associated with a particular point in space, can be written

A= A i ei , (C.4)
i=1,3

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
du1 du2 du3 
dx = e1 + e2 + e3 = hi dui ei . (C.5)
|u1 | |u2 | |u3 | i=1,3

Hence, from (C.1), an element of length (squared) in the u1 , u2 , u3 coordinate system


is written 
dx dx = hi2 dui2 . (C.6)
i=1,3

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 , 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)

It can be seen that [see Equation (A.176)]

ui = 0, (C.8)

and 2
h
i ui = 0. (C.9)
h
The latter result follows from Equations (A.175) and (A.176) because (h12 /h) u1 =
u2 u3 , et cetera. Finally, it is easily demonstrated from Equations (C.1) and (C.3)
that

ui u j = hi2 i j , (C.10)
1
ui u j uk = h i jk . (C.11)

Consider a scalar field (u1 , u2 , u3 ). It follows from the chain rule, and the rela-
tion ei = hi ui , that
  1
= ui = ei . (C.12)
i=1,3
ui h ui
i=1,3 i
Non-Cartesian Coordinates 541

Hence, the components of in the u1 , u2 , u3 coordinate system are


1
()i = . (C.13)
hi ui
Consider a vector field A(u1 , u2 , u3 ). We can write
  
h hi2
A= (Ai ei ) = (hi Ai ui ) = Ai ui
i=1,3 i=1,3 i=1,3
hi h
 h2
1 h

h
= i
ui Ai = Ai , (C.14)
i=1,3
h hi i=1,3
h ui hi

where use has been made of Equations (A.174), (C.9), and (C.10). Thus, the diver-
gence of A in the u1 , u2 , u3 coordinate system takes the form
 1
h
A= Ai . (C.15)
i=1,3
h ui hi

We can write
  
A= (Ak ek ) = (hk Ak uk ) = (hk uk ) uk
k=1,3 k=1,3 k=1,3
 (hk Ak )
= u j uk , (C.16)
j,k=1,3
u j

where use has been made of Equations (A.178), (C.8), and (C.12). It follows from
Equation (C.11) that
 (hk Ak )  hi (hk Ak )
( A)i = ei A = hi ui u j uk = i jk .
j,k=1,3
u j j,k=1,3
h u j
(C.17)
Hence, the components of A in the u1 , u2 , u3 coordinate system are
 hi (hk Ak )
( A)i = i jk . (C.18)
j,k=1,3
h u j

Now, 2 = [see Equation (A.172)], so Equations (C.12) and (C.15) yield


the following expression for 2 in the u1 , u2 , u3 coordinate system:
 1 h
2 = . (C.19)
i=1,3
h ui hi2 ui

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:
1
(A ) B = [(A B) (A B) ( A) B + ( B) A
2
A ( B) B ( A)] . (C.20)
542 Theoretical Fluid Mechanics

Making use of Equations (C.13), (C.15), and (C.18), as well as the easily demon-
strated results

AB= Ai Bi , (C.21)
i=1,3

AB= i jk A j Bk , (C.22)
j,k=1,3

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:


A j Bi A j B j h j Ai B j hi
[(A ) B]i = + . (C.23)
j=1,3
h j u j hi h j ui hi h j u 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
(A)i j = . (C.24)
x j

In an orthogonal curvilinear coordinate system, the previous expression generalizes


to

(A)i j = [(e j ) A]i . (C.25)

It thus follows from Equation (C.23), and the relation (ei ) j = ei e j = i j , that

1 Ai A j h j  Ak hi
(A)i j = + i j . (C.26)
h j u j hi h j ui h h uk
k=1,3 i k

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 previous equation reduces (after a great deal of
tedious algebra) to the following expression for the components of 2 A in the u1 , u2 ,
Non-Cartesian Coordinates 543

u3 coordinate system:
  2
1 hi 1 h j

( A)i = Ai +
2 2
Aj
j=1,3
hi h j hi u j ui h j ui u j

h A j h j hi2 Ai hi h j2
+ 2
hi h j2 hi ui u j h h j2 u j u j h
2 2 2
A j h 1 h j h j h h j h j 2 h 2
j
+ +
hi h j3 h j ui u j h h j2 ui h u j h ui u j h

2
Ai 2 hi 2 hi

. (C.28)
hi h hi u j
2
j u
j
2

Note, again, that the commonly quoted result ( 2 A)i = 2 Ai is only valid in Carte-
sian coordinate systems (for which h1 = h2 = h3 = h = 1).

C.3 Cylindrical Coordinates



In the cylindrical coordinate system, u1 = r, u2 = , and u3 = z, where r = x 2 + y 2 ,
= tan1 (y/x), and x, y, z are standard Cartesian coordinates. Thus, r is the perpen-
dicular 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|. Of course, the unit basis vectors
er , e , and ez are mutually orthogonal, so Ar = A er , et cetera.
As is easily demonstrated, an element of length (squared) in the cylindrical coor-
dinate system takes the form

dx dx = dr 2 + r 2 d 2 + dz 2 . (C.30)

Hence, comparison with Equation (C.6) reveals that the scale factors for this system
are

hr = 1, (C.31)
h = r, (C.32)
hz = 1. (C.33)
544 Theoretical Fluid Mechanics

y e er

r z


x
O

Figure C.1
Cylindrical coordinates.

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)

respectively, whereas a volume element takes the form

dV = r dr d dz. (C.37)

According to Equations (C.13), (C.15), and (C.18), gradient, divergence, and curl
in the cylindrical coordinate system are written

1
= er + e + ez , (C.38)
r r z
1 1 A Az
A= (r Ar ) + + , (C.39)
r r r z




1 Az A Ar Az 1 1 Ar
A= er + e + (r A ) ez ,
r z z r r r r
(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
Non-Cartesian Coordinates 545

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

A2
[(A ) A]r = A Ar , (C.42)
r
Ar A
[(A ) A] = A A + , (C.43)
r
[(A ) A]z = A Az . (C.44)

Let us define the symmetric gradient tensor


 
6 = 1 A + (A)T .
A (C.45)
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 corresponding element of ST is S ji . According to
Equation (C.26), the components of A 6 in the cylindrical coordinate system are

6 rr = Ar
(A) , (C.46)
r
6 = 1 A Ar
(A) + , (C.47)
r r
6 zz = Az
(A) , (C.48)
z


6 r = (A)
6 r = 1 1 Ar A A
(A) + , (C.49)
2 r r r


6 zr =
6 rz = (A) 1 Ar Az
(A) + , (C.50)
2 z r


6 z =
6 z = (A) 1 A 1 Az
(A) + . (C.51)
2 z r

Finally, from Equation (C.28), the components of 2 A in the cylindrical coordi-


nate system are

Ar 2 A
( 2 A)r = 2 Ar , (C.52)
r 2 r 2
2 Ar A
( 2 A) = 2 A + 2 2, (C.53)
r r
( 2 A)z = 2 Az . (C.54)
546 Theoretical Fluid Mechanics

z
r y

x
O

Figure C.2
Spherical coordinates.

C.4 Spherical Coordinates



In the spherical coordinate system, u1 = r, u2 = , and u3 = , where r = x 2 + y 2 + z 2 ,
= 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 = /|| . Of course, the unit vectors er ,
e , and e are mutually orthogonal, so Ar = A er , et cetera.
As is easily demonstrated, an element of length (squared) in the spherical coor-
dinate system takes the form
dx dx = dr 2 + r 2 d 2 + r 2 sin2 d2 . (C.56)
Hence, comparison with Equation (C.6) reveals that the scale factors for this system
are
hr = 1, (C.57)
h = r, (C.58)
h = r sin . (C.59)
Non-Cartesian Coordinates 547

Thus, surface elements normal to er , e , and e are written

dS r = r 2 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 = r 2 sin dr d d. (C.63)

According to Equations (C.13), (C.15), and (C.18), gradient, divergence, and curl
in the spherical coordinate system are written

1 1
= er + e + e , (C.64)
r r r sin
1 2 1 1 A
A= (r Ar ) + (sin A ) + , (C.65)
r 2 r r sin r sin

1 1 A
A= (sin A ) er
r sin r sin

1 Ar 1
+ (r A ) e
r sin r r

1 1 Ar
+ (r A ) e , (C.66)
r r r

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 Lapla-
cian of a scalar field becomes



1 2 1 1 2
2 = 2 r + 2 sin + . (C.67)
r r r r sin r 2 sin 2
2

Moreover, from Equation (C.23), the components of (A )A in the spherical


coordinate system are

A2 + A2
[(A ) A]r = A Ar , (C.68)
r
Ar A cot A2
[(A ) A] = A A + , (C.69)
r
Ar A + cot A A
[(A ) A] = A A + . (C.70)
r

6 in the spherical coor-


Now, according to Equation (C.26), the components of A
548 Theoretical Fluid Mechanics

dinate system are

6 rr = Ar
(A) , (C.71)
r
6 = 1 A Ar
(A) + , (C.72)
r r
6 = 1 A Ar cot A
(A) + + , (C.73)
r sin r r


6 r =
6 r = (A) 1 1 Ar A A
(A) + , (C.74)
2 r r r


6 r =
6 r = (A) 1 1 Ar A A
(A) + , (C.75)
2 r sin r r


6 =
6 = (A) 1 1 A 1 A cot A
(A) + . (C.76)
2 r sin r r

Finally, from Equation (C.28), the components of 2 A in the spherical coordinate


system are

2Ar 2 A 2 cot A 2 A
( 2 A)r = 2 Ar 2 2 , (C.77)
r 2 r r 2 r sin
2 Ar A 2 A
( 2 A) = 2 A + 2 , (C.78)
r 2 r sin
2 2 r sin
A 2 Ar 2 cot A
( 2 A) = 2 A + + 2 . (C.79)
r sin r sin
2 2 2 2 r sin

C.5 Exercises
C.1 Find the Cartesian components of the basis vectors er , e , and ez of the cylin-
drical 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 ra-
dius a and length l is a 2 l. Demonstrate that the moment of inertia of a uni-
form cylinder of mass M and radius a about its symmetry axis is (1/2) M a 2 .

C.3 Use spherical coordinates to prove that the volume of a sphere of radius a is
(4/3) a 3. 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 a 2 .

C.4 For what value(s) of n is (r n er ) = 0, where r is a spherical coordinate?


Non-Cartesian Coordinates 549

C.5 For what value(s) of n is (r n er ) = 0, where r is a spherical coordinate?


C.6 (a) Find a vector field F = Fr (r) er satisfying F = r m for m 0. Here,
r is a spherical coordinate.
(b) Use the divergence theorem to show that
 
1
r m dV = r m+1 er dS,
V m + 3 S

where V is a volume enclosed by a surface S .


(c) Use the previous 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) = ,
4 0 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 equa-
tions x = (u 2 v 2 )/2, y = u v, z = z, where x, y, z are Cartesian coordi-
nates, 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 equa-
tions x = cosh cos , y = sinh sin , z = z, where x, y, z are Cartesian
coordinates, and 0 , < , are orthogonal. Find the scale fac-
tors h , h , hz . What shapes are the = constant and = constant surfaces?
Write an expression for f in elliptical cylindrical coordinates.
550 Theoretical Fluid Mechanics
D
Ellipsoidal Potential Theory

D.1 Introduction
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, (D.1)
a12 a22 a32
where a1 , a2 , and a3 are the principal radii along the x1 -, x2 -, and x3 -axes, respec-
tively. 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. More information
on ellipsoidal potential theory can be found in Chandrasekhar 1969.

D.2 Analysis
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 D.1.) If the cone subtends a solid angle d
then a volume element is written dV = r 2 dr d, where r measures displacement
from P along the axis of the cone. Thus, from standard classical gravitational theory
(Fitzpatrick 2012), the contribution to the potential takes the form
 r  0
G G
d = dV dV, (D.2)
0 r r (r)
where r = |PQ|, r = |PR|, and is the constant mass density of the ellipsoid.
Hence, we obtain

 r  r
1
d = G r dr + r dr d = G (r 2 + r 2 ) d. (D.3)
0 0 2
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

1
= G (r 2 + r 2 ) d. (D.4)
4

551
552 Theoretical Fluid Mechanics

d
r
O x P

R
Figure D.1
Calculation of ellipsoidal gravitational potential.

From Figure D.1, the position vector of point Q, relative to the origin, O, is

x = x + r n, (D.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. (D.6)

However, Q and R both lie on the bodys outer boundary. It follows, from Equa-
tion (D.1), that r and r are the two roots of

xi + r n i 2
= 1, (D.7)
i=1,3
ai

which reduces to the quadratic

A r 2 + B r + C = 0, (D.8)

where
 n2
A= i
, (D.9)
ai2
i=1,3
 xi n i
B=2 , (D.10)
i=1,3
ai2
 x2
C= i
1. (D.11)
i=1,3
ai2
Ellipsoidal Potential Theory 553

According to standard polynomial equation theory (Riley 1974), r +r = B/A, and
r r = C/A. Thus,
B2 C
r 2 + r 2 = (r + r )2 2 r r = 2
2 , (D.12)
A A
and Equation (D.4) becomes

 2   
1 i=1,3 xi2 /ai2
2 2
1 i=1,3 xi ni /ai
= G  2 +  2
d. (D.13)
2 i=1,3 ni /ai
2
i=1,3 ni /ai
2 2

The previous expression can also be written



 2  
i, j=1,3 xi x j ni n j /(ai a j ) 1 i=1,3 xi2 /ai2
2 2
1
= G  2 +  2
d. (D.14)
2 i=1,3 ni /ai
2
n 2 /a 2
i=1,3 i i

However, the cross terms (i.e., i  j) integrate to zero by symmetry, and we are left
with

   2

1
2 x 2 2
n /a 4
1 x 2
/a i
= G  d.
i=1,3 i i i=1,3 i

i
+  (D.15)
2 2 n 2 /a 2
n
i=1,3 i
2 /a i
2 i=1,3 i i

Let 
d
J=  . (D.16)
i=1,3 ni /ai
2 2

It follows that 
1 J 2 ni2/ai4
=  2 d. (D.17)
ai ai ni2 /ai2
i=1,3

Thus, Equation (D.15) can be written



1 
= G J Ai xi ,
2
(D.18)
2 i=1,3

where
J 1 J
Ai = . (D.19)
ai2 ai ai
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 ), (D.20)

and d = sin d d. We find, from Equation (D.16), that


 /2  /2 2 1
sin cos2 sin2 sin2 cos2
J=8 sin d + + d. (D.21)
0 0 a12 a22 a32
554 Theoretical Fluid Mechanics

Let t = tan . It follows that


 /2   /2
dt sin d
J=8 sin d = 4 , (D.22)
0 0 a + b t2 0 (a b)1/2
where
sin2 cos2
a= + , (D.23)
a12 a32
sin2 cos2
b= + . (D.24)
a22 a32

Hence, we obtain
 /2
sin sec2 d
J= 4 a1 a2 a32 . (D.25)
0 (a12 + a32 tan2 )1/2 (a22 + a32 tan2 )1/2

Let u = a32 tan2 . It follows that



du
J = 2 a1 a2 a3 , (D.26)
0
where
= (a12 + u)1/2 (a22 + u)1/2 (a32 + u)1/2 . (D.27)
Now, from Equations (D.19), (D.26), and (D.27),


2 a1 a2 a3 du 1 du
Ai = 2 a a a
1 2 3
ai2 0 ai ai 0


2 a1 a2 a3 1 du
= du = 2 a1 a2 a3 . (D.28)
0 ai 0 (ai + u)
ai 2

Thus, Equations (D.18), (D.26), and (D.28) yield



3 

= G M 0 i xi ,
2
(D.29)
4 i=1,3

where

du
0 = , (D.30)
0

du
i = . (D.31)
0 (ai2 + u)

Here, M = V and V = (4/3) a1 a2 a3 are the bodys mass and volume, respec-
tively.
Ellipsoidal Potential Theory 555

The total gravitational potential energy of the body is written (Fitzpatrick 2012)

1
U= dV, (D.32)
2

where the integral is taken over all interior points. It follows from Equation (D.29)
that
3 2

1 
U = G M 0 i ai .
2
(D.33)
8 5 i=1,3

In writing
 the previous expression, use has been made of the easily demonstrated
result xi2 dV = (1/5) ai2 V. Now,

d $u% 1  ai2
2 = + , (D.34)
du i=1,3 (ai2 + u)

so
 
  ai2 du d $u% 1
i ai2 = = 2 + du = 0 . (D.35)
i=1,3 0 i=1,3
(ai2 + u) 0 du

Hence, we obtain
3
U= G M 2 0 . (D.36)
10

D.3 Exercises
D.1 Demonstrate that the volume of an ellipsoid whose bounding surface satis-
fies
x12 x22 x32
+ + = 1,
a12 a22 a32
is V = (4/3) a1 a2 a3 .

D.2 Demonstrate that the moments of inertia about the three Cartesian axes of a
homogeneous ellipsoidal body of mass M, whose bounding surface satisfies
(x1 /a1 )2 + (x2 /a2 )2 + (x3 /a3 )2 = 1, are

M 2
I1 = (a + a32 ),
5 2
M 2
I2 = (a + a32 ),
5 1
M 2
I3 = (a + a22 ).
5 1
556 Theoretical Fluid Mechanics

D.3 According to MacCullaghs formula (Fitzpatrick 2012), the gravitational po-


tential a relatively long way from a body of mass M whose center of mass
coincides with the origin, and whose principal moments of inertial are I1 , I2 ,
and I3 (assuming that the principal axes coincide with the Cartesian axes),
takes the form

G M G (I1 + I2 + I3 ) 3 G (I1 x12 + I2 x22 + I3 x32 )


(x1 , x2 , x3 ) + ,
r 2r3 2r5
where r = (x12 + x22 + x32 )1/2 . Demonstrate that if the body in question is a ho-
mogeneous ellipsoid whose bounding surface satisfies (x1 /a1 )2 + (x2 /a2 )2 +
(x3 /a3 )2 = 1 then

GM GM 2 2 1 2
(x1 , x2 , x3 ) a x (x + x3 ) 2
(D.37)
r 5r5 1 1 2 2

GM 2 2 1 2 GM 2 2 1 2
a x (x + x 2
) a x (x + x 2 .
2
)
5r5 2 2 2 1 3
5r5 3 3 2 1

D.4 Show that the gravitational potential external to a homogeneous ellipsoidal


body of mass M, whose outer boundary satisfies (x1 /a1 )2 +(x2 /a2 )2 +(x3 /a3 )2 =
1, takes the form

3 
(x1 , x2 , x3 ) = G M 0 i xi2 , (D.38)
4 i=1,3

where

du
0 = ,


du
i = ,
(ai2 + u)

and = (a12 + u)1/2 (a22 + u)1/2 (a32 + u)1/2 . Here, is the positive root of

 xi2
= 1.
i=1,3
ai2 +

Demonstrate that, at large , Equation (D.38) reduces to Equation (D.37).


E
Calculus of Variations

E.1 Indroduction
This appendix gives a brief overview of the calculus of variations. More information
on this topic can be found in Riley 1974.

E.2 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 E.1. Now, l takes the form
 B  b
l= (dx + dy )
2 2 1/2
= [1 + y 2 (x)]1/2 dx, (E.1)
A 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.
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. (E.2)
The meaning of the previous 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
 b
I= F(y, y, x) dx, (E.3)
a

557
558 Theoretical Fluid Mechanics

B
y

a b x
Figure E.1
Dierent paths between points A and B.

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
 b

F F
I = y +  y dx, (E.4)
a y y
where y = d(y)/dx. Setting I to zero, we obtain
 b

F F
y +  y dx = 0. (E.5)
a y y
This equation must be satisfied for all possible small perturbations y(x).
Integrating the second term in the integrand of the previous equation by parts, we
get
 b
b
F d F F
y dx + y = 0. (E.6)
a y dx y y a
However, 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 previous equation is zero. Thus, we obtain
 b

F d F
y dx = 0. (E.7)
a y dx y
The previous 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


d F F
= 0. (E.8)
dx y y
Calculus of Variations 559

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 (E.8) simplifies to
F
= const. (E.9)
y
Next, suppose that F does not depend explicitly on x. Multiplying Equation (E.8) by
y , we obtain

d F F
y 
y = 0. (E.10)
dx y y
However,


d  F  d F F
y =y + y  . (E.11)
dx y dx y y
Thus, we get

d  F F F
y = y + y  . (E.12)
dx y y y
Now, if F is not an explicit function of x then the right-hand side of the previous
equation is the total derivative of F, namely dF/dx. Hence, we obtain


d  F dF
y 
= , (E.13)
dx y dx
which yields
F
y F = const. (E.14)
y

Returning to the case under consideration, we have F = 1 + y 2 , according to
Equation (E.1) and (E.3). Hence, F is not an explicit function of y, so Equation (E.9)
yields
F y
=  = c, (E.15)
y 1 + y 2
where c is a constant. So,
c
y = = const. (E.16)
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.

E.3 Conditional Variation


Suppose that we wish to find the function y(x) which maximizes or minimizes the
functional  b
I= F(y, y, x) dx, (E.17)
a
560 Theoretical Fluid Mechanics

subject to the constraint that the value of


 b
J= G(y, y , x) dx (E.18)
a

remains constant. We can achieve our goal by finding an extremum of the new func-
tional K = I + J, where (x) is an undetermined function. We know that J = 0,
because 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 d [ G] [ G]
+ = 0. (E.19)
dx y y dx y y

In principle, the previous equation, together with the constraint (E.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
that 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 that minimizes the systems potential energy
(Fitzpatrick 2012). Now, the potential energy of the chain is written
  a/2
U = g y ds = g y (1 + y 2 )1/2 dx, (E.20)
a/2

where ds = dx 2 + dy 2 is an element of length along the chain, and g is the acceler-
ation 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
  a/2
l= ds = (1 + y 2 )1/2 dx (E.21)
a/2

remains constant.
It follows, from the previous discussion, that we need to minimize the functional
 a/2
K = U + l = ( g y + ) (1 + y 2 )1/2 dx, (E.22)
a/2
Calculus of Variations 561

where is an, as yet, undetermined constant. Because the integrand in the functional
does not depend explicitly on x, we have from Equation (E.14) that

y 2 ( g y + ) (1 + y 2 )1/2 ( g y + ) (1 + y 2 )1/2 = k, (E.23)

where k is a constant. This expression reduces to


$ y %2
y 2 =  + 1, (E.24)
h

where  = /k, and h = k/( g).


Let
y
 + = cosh z. (E.25)
h
Making this substitution, Equation (E.24) yields

dz
= h 1 . (E.26)
dx
Hence,
x
z = + c, (E.27)
h
where c is a constant. It follows from Equation (E.25) that

y(x) = h [ + cosh(x/h + c)]. (E.28)

The previous 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. (E.29)

Hence, c = 0, and  = cosh(a/2 h). It follows that

y(x) = h [cosh(a/2 h) cosh(x/h)]. (E.30)

The final unknown constant, h, is determined via the application of the constraint
(E.21). Thus,
 a/2  a/2
l= (1 + y 2 )1/2 dx = cosh(x/h) dx = 2 h sinh(a/2 h). (E.31)
a/2 a/2

Hence, the equilibrium configuration of the chain is given by the curve (E.30), which
is known as a catenary (from the Latin for chain), where the parameter h satisfies

l $ a %
= sinh . (E.32)
2h 2h
562 Theoretical Fluid Mechanics

E.4 Multi-Function Variation


Suppose that we wish to maximize or minimize the functional
 b
I= F(y1 , y2 , , yF , y1 , y2 , , yF , x) dx. (E.33)
a

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 E.2 yields F
separate Euler-Lagrange equations,


d F F
 = 0, (E.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. (E.35)
yk

Likewise, if F does not explicitly depend on x then all F Euler-Lagrange equations


simplify to
F
yi  F = const, (E.36)
yi
for i = 1, F .

E.5 Exercises
E.1 Find the extremal curves y = y(x) of the following constrained optimization
problems, using the method of Lagrange multipliers:
1  1
(a) 0
y 2 + x2 dx, such that 0 y 2 dx = 2.
 
(b) 0 y 2 dx, such that y(0) = y() = 0, and 0 y 2 dx = 2.
1  
(c) 0 y dx, such that y(0) = y(1) = 1, and 1 + y 2 dx = 2/3.

E.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
Calculus of Variations 563

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.
E.3 Find the curve y(x), in the interval 0 x p, which is of length , and
maximizes  p
y dx.
0
564 Theoretical Fluid Mechanics
565

Bibliography
Abramowitz, M. (ed.), and Stegun, I. (ed.) 1965. Handbook of Mathematical Func-
tions: with Formulas, Graphs, and Mathematical Tables. Dover.

Anderson, J.D. Jr. 2003. Modern Compressible Flow, with Historical Perspective,
3rd Edition. McGraw-Hill.

Arfken, G. 1985. Mathematical Methods for Physicists, 3rd Edition. Academic


Press.

Batchelor, G.K. 2000. An Introduction to Fluid Dynamics. Cambridge.

Cartwright, D.E. 1999. Tides. A Scientific History. Cambridge.

Chandrasekhar, S. 1967. An Introduction to the Study of Stellar Structure. Dover.

Chandrasekhar, S. 1969. Ellipsoidal Figures of Equilibrium. Yale.

Darwin, G.H. 1886. Proceedings of the Royal Society of London 41, 319.

Emanuel, G. 2000. Analytical Fluid Dynamics. CRC Press.

Eddington, A.S. 1926. The Internal Constitution of the Stars. Cambridge.

Faber, T.E. 1995. Fluid Dynamics for Physicists, 1st Edition. Cambridge.

Farrell, W.E. 1972. Reviews of Geophysics 10, 761.

Fitzpatrick, R. 2008. Maxwells Equations and the Principles of Electromagnetism.


Jones & Bartlett.

Fitzpatrick, R. 2012. An Introduction to Celestial Mechanics. Cambridge.

Fitzpatrick, R. 2013. Oscillations and Waves: An Introduction. CRC.

Hazeltine, R.D., and Waelbroeck, F.L. 2004. The Framework of Plasma Physics.
Westview.

Hodge, P.G., Jr. 1961. American Mathematical Monthly 68, 793.

Hughes, W., and Brighton, J. 1999. Schaums Outline of Fluid Dynamics, 3rd Edi-
tion. McGraw-Hill.

Jackson, J.D. 1962. Classical Electrodynamics, 2nd Edition. Wiley.

Jereys, H. 1961. Cartesian Tensors. Cambridge.

Lamb, H. 1928. Statics, Including Hydrostatics and the Elements of the Theory of
Elasticity, 3rd Edition. Cambridge.
566

Lamb, H. 1993 Hydrodynamics, 6th Edition. Cambridge.


Landau, L.D., and Lifshitz, E.M. 1987. Fluid Mechanics, 2nd Edition. Butterworth-
Heinemann.
Lighthill, J, 1978. Waves in Fluids. Cambridge.
Liepmann, H.W., and Roshko, A. 1957. Elements of Gasdynamics. Wiley.
Longuet-Higgins, M.S., and Pond, G.S. 1970. Philosophical Transactions of the
Royal Society of London A 266, 193.
Love, A.E.H. 1911. Some Problems of Geodynamics: Being an Essay to Which the
Adams Prize in the University of Cambridge was Adjudged in 1911. Cambridge.
Love, A.E.H. 1913. Proceedings of the London Mathematical Society 12, 309.
Love, A.E.H. 1927. A Treatise on the Mathematical Theory of Elasticity, 4th Edi-
tion. Cambridge.
Mestel, L. 1999. Phys. Reports 311, 295.
Milne-Thomson, L.M. 1958. Theoretical Aerodynamics, 4th Edition, Revised and
enlarged. Dover.
Milne-Thomson, L.M. 2011. Theoretical Hydrodynamics, 5th Edition. Dover.
Murray, C.D., and Dermott, S.F. 1999. Solar System Dynamics. Cambridge.
Press, W.H., Teukolsky, S.A., Vetterling, W.T., and Flannery, B.P. 2007. Numerical
Recipes: The Art of Scientific Computing, 3rd Edition. Cambridge.
Proudman, J. 1916. Proceedings of the London Mathematical Society 18, 51.
Reif, F. 1965. Fundamentals of Statistical and Thermal Physics. McGraw-Hill.
Riley, K.F. 1974. Mathematical Methods for the Physical Sciences. Cambridge.
Schey, H.M. 1992. Div, Grad, Curl, and All That, 2nd Edition. Norton.
Schlichting, H. 1987. Boundary Layer Theory, 7th Edition. McGraw-Hill.
Spiegel, M.R., Liu, J., and Lipschutz, S. 1999. Mathematical Handbook of Formu-
las and Tables, 2nd Edition. McGraw-Hill.
Temple, G. 2004. Cartesian Tensors. Dover.
Wong, B.R. 1998. Journal of Physics A: Mathematical and General 31, 1101.
Yoder, C.F., 1995. Astrometric and Geodetic Properties of Earth and the Solar
System, in Global Earth Physics: A Handbook of Physical Constants, Ahrens, T.
(ed.). American Geophysical Union.

You might also like