Vous êtes sur la page 1sur 267

Physical Oceanography OCEA-4120 / OCEA-5120

c _1993-2011 Dan Kelley


5652 Oceanography LSC; 494-1694
mailto:Dan.Kelley@Dal.Ca?subject=OCEx120
http://myweb.dal.ca/kelley/OCE5120/index.php
ii
Preface
About this document. This is a supplement to OCEA-4120/5120 (and various cross
listings in other departments) at Dalhousie University, a course that introduces Physical
Oceanography at the senior undergraduate and graduate levels. Students come to this course
with a variety of backgrounds, and so not every reader will approach these notes in the
same way. For example, all students should read Chapter 1 (Introduction and Overview)
in Part I, but most students will have a sufciently deep background in Mathematics and
Physics to skip Chapters 2 and 3 in Part II. The meat of the course is in Parts III and IV,
and most of this material will be covered in class. Various ancillary topics (e.g. numerical
modelling, dealing with data) are covered in Part V. The lectures will provide a guide
through all of this, and much more. Stewart [2005] provides an online copy of course notes
that are a good supplement to this course. The following dead-tree texts may also prove
helpful: Knauss [1978], Pickard [1979], Pond and Pickard [1978] and Gill [1982].
About the cover. Winds and air temperature at Halifax are shown, along with the tem-
perature of water pumped from the Northwest Arm of Halifax Harbour into the Aquatron
facility of the Life Sciences Centre. As this course develops, you will learn how infer a
great deal of dynamical information from such diagrams, when considered in the context of
general theories.
Colophon. This document was prepared in L
A
T
E
X, using the hyperref package to
provide easy navigation in a PDF viewer. The page geometry is set up for two-sided printing
with a wide margin on the outer edge; to get the proper margins, be sure to print with page
scaling turned off.
iii
iv
Contents
I Course scope 1
1 Introduction & Overview 3
1.1 What is Physical Oceanography? . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 About the Course . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Satellite images . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 How the ocean is unlike a bathtub . . . . . . . . . . . . . . . . . . . . . . 5
1.4.1 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4.2 Stratication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4.3 Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4.4 Wind Forcing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.4.5 Thermohaline Forcing . . . . . . . . . . . . . . . . . . . . . . . . 10
1.4.6 Mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.4.7 Geometry (again): Continental shelves . . . . . . . . . . . . . . . 11
1.5 Course topics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.6 Table of symbols and keywords . . . . . . . . . . . . . . . . . . . . . . . . 12
II Review 17
2 Mathematical tools for Physical Oceanography 19
2.1 The beauty of proof . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.2 Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.3 Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3.1 Integers, rational numbers, and irrational numbers . . . . . . . . . 22
2.3.2 Imaginary Numbers . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3.3 Complex Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.4 Calculus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.4.1 Ordinary differentiation . . . . . . . . . . . . . . . . . . . . . . . 25
2.4.2 Partial differentiation . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.4.3 Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.5 Differential equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.5.1 Ordinary differential equations . . . . . . . . . . . . . . . . . . . . 27
2.5.2 Linear ODE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.5.3 Constant coefcient linear ODE . . . . . . . . . . . . . . . . . . . 30
2.5.4 Scaling Differential Equations . . . . . . . . . . . . . . . . . . . . 31
2.6 Dimensional Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.7 Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.8 Vector dot product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.9 Vector cross product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.10 Curl of a vector eld . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.11 Divergence of a vector eld . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.12 Gradient of a scalar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
v
vi CONTENTS
3 Review of Physics 35
3.1 Momentum: Newtons Second Law . . . . . . . . . . . . . . . . . . . . . 35
3.2 Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.2.1 Concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.2.2 Solid-body application . . . . . . . . . . . . . . . . . . . . . . . . 37
3.2.3 Fluid application . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
III Ocean Physics: General Issues 39
4 Seawater properties 41
4.1 Temperature, Salinity, Density . . . . . . . . . . . . . . . . . . . . . . . . 41
4.2 Hydrographic Sampling . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.3 Temperature distribution in the ocean . . . . . . . . . . . . . . . . . . . . 42
4.3.1 Explanation for depth variation T(z) in temperate seas . . . . . . . 42
4.3.2 Explanation for depth variation T(z) in polar seas . . . . . . . . . . 47
4.3.3 The Adiabatic Effect . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.4 Salinity distribution in the ocean . . . . . . . . . . . . . . . . . . . . . . . 48
4.5 Pressure distribution in the ocean . . . . . . . . . . . . . . . . . . . . . . . 50
4.6 How T, S, p and are measured . . . . . . . . . . . . . . . . . . . . . . . 51
4.6.1 Instruments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.6.2 How S is measured . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.6.3 How p is measured . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.7 Equation of state: = (S, T, p) . . . . . . . . . . . . . . . . . . . . . . . 57
4.8 Linear equation of state . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.9 Stratication and stability . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.10 Baroclinic and Barotropic States . . . . . . . . . . . . . . . . . . . . . . . 61
4.11 Fluid energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.11.1 Kinetic energy of a moving uid . . . . . . . . . . . . . . . . . . . 62
4.11.2 Gravitational potential energy . . . . . . . . . . . . . . . . . . . . 62
The General Case . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
Unstratied case . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
Two-layer case . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
Linearly-stratied case A . . . . . . . . . . . . . . . . . . . . . . . 64
Linearly-stratied case B . . . . . . . . . . . . . . . . . . . . . . . 64
Energy increase through mixing a water column . . . . . . . . . . . 64
Application: mixed-layer energetics . . . . . . . . . . . . . . . . . 65
4.12 Fluxes across the air-sea interface . . . . . . . . . . . . . . . . . . . . . . 65
4.12.1 Momentum ux Wind stress . . . . . . . . . . . . . . . . . . . . 65
4.12.2 Heat ux . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.13 Ways to look at T, S, p and data . . . . . . . . . . . . . . . . . . . . . . 67
4.13.1 The T-S diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
Cabbelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Salt fountain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Salt ngering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Diffusive instability . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.14 Other ocean tracers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
CONTENTS vii
5 Dynamics of Ocean Currents 77
5.1 Concepts of Fluid Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . 77
5.1.1 Continuum Hypothesis . . . . . . . . . . . . . . . . . . . . . . . . 77
5.1.2 Newtons Second Law (F = ma) . . . . . . . . . . . . . . . . . . . 77
Newtons Second Law for Solid-body mechanics . . . . . . . . . . 78
Newtons Second Law for Fluid mechanics . . . . . . . . . . . . . 78
5.1.3 Eulerian and Lagrangian Notation . . . . . . . . . . . . . . . . . . 79
Eulerian vs Lagrangian acceleration . . . . . . . . . . . . . . . . . 79
Extension to three dimensions . . . . . . . . . . . . . . . . . . . . 80
5.1.4 Fluxes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
5.1.5 Conservation laws . . . . . . . . . . . . . . . . . . . . . . . . . . 81
5.1.6 Conservation of mass . . . . . . . . . . . . . . . . . . . . . . . . . 81
5.1.7 Incompressible ow: continuity equation . . . . . . . . . . . . . . 82
Derivation of continuity equation . . . . . . . . . . . . . . . . . . 82
Example of use of continuity equation . . . . . . . . . . . . . . . . 82
5.1.8 Conservation of S, T, etc . . . . . . . . . . . . . . . . . . . . . . . 83
5.1.9 Eddy Fluxes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
5.2 Forces on a Fluid Element . . . . . . . . . . . . . . . . . . . . . . . . . . 83
5.2.1 Pressure gradients . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.2.2 Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.2.3 Hydrostatic equation . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.2.4 Coriolis Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.2.5 Molecular friction . . . . . . . . . . . . . . . . . . . . . . . . . . 88
5.2.6 Turbulent friction . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
5.3 Basic Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . 90
5.3.1 The Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . 90
5.3.2 Scaling the Equations of motion . . . . . . . . . . . . . . . . . . . 90
5.3.3 The Reynolds Number (advection vs friction) . . . . . . . . . . . . 91
5.3.4 The Rossby Number (advection vs Coriolis) . . . . . . . . . . . . . 93
5.4 Geostrophic Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
5.4.1 Thermal Wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
5.4.2 One-Layer Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
5.4.3 Two-Layer Flow (Margules Equations) . . . . . . . . . . . . . . . 97
5.4.4 Reference velocities (level of no motion) . . . . . . . . . . . . . 99
5.4.5 Dynamic Height . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
5.4.6 Constraint on Geostrophic Flow over Sloped Bottoms . . . . . . . 100
5.5 Wind forcing: slab response . . . . . . . . . . . . . . . . . . . . . . . . . 101
5.5.1 Case with constant forcing . . . . . . . . . . . . . . . . . . . . . . 101
Steady-state solution . . . . . . . . . . . . . . . . . . . . . . . . . 101
Time-varying solution . . . . . . . . . . . . . . . . . . . . . . . . 102
5.5.2 Case with time-varying forcing . . . . . . . . . . . . . . . . . . . 102
Analytical approach . . . . . . . . . . . . . . . . . . . . . . . . . 102
Numerical approach . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.6 Wind Forcing: depth-varying (Ekman) response . . . . . . . . . . . . . . . 104
viii CONTENTS
IV Ocean Physics: Limiting Cases 107
6 Wind-driven ocean circulation 109
6.1 Sverdrup transport: derivation 1 . . . . . . . . . . . . . . . . . . . . . . . 109
6.2 Sverdrup transport: derivation 2 . . . . . . . . . . . . . . . . . . . . . . . 113
6.2.1 Vorticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
6.2.2 Potential vorticity . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
6.2.3 Ekman pumping, potential vorticity, and Sverdrup ow . . . . . . . 114
6.3 The return ow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
6.4 Rossby waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
7 Thermohaline circulation 119
7.1 Movement of deep water . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
7.2 Thermocline theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
7.2.1 The Munk balance . . . . . . . . . . . . . . . . . . . . . . . . . . 120
7.2.2 The Ventilated Thermocline . . . . . . . . . . . . . . . . . . . . . 123
7.3 General circulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
8 Continental Shelf Circulation 125
8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
8.2 Processes and scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
8.3 Wind-driven coastal upwelling . . . . . . . . . . . . . . . . . . . . . . . . 126
8.4 Shelf-break upwelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
8.5 Shelf waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
8.6 Storm surges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
8.7 Thermohaline circulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
8.7.1 Buoyancy currents . . . . . . . . . . . . . . . . . . . . . . . . . . 129
8.7.2 Fronts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
8.7.3 Tides on Shelves and Slopes . . . . . . . . . . . . . . . . . . . . . 130
9 Estuarine Circulation 133
9.1 Classication of Estuaries . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
9.1.1 Topographic Classication . . . . . . . . . . . . . . . . . . . . . . 133
9.1.2 Hydrographic classication . . . . . . . . . . . . . . . . . . . . . 135
9.2 Mixing: the Richardson number . . . . . . . . . . . . . . . . . . . . . . . 135
9.3 Estuarine hydrography . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
9.3.1 Hansen-Rattray classication . . . . . . . . . . . . . . . . . . . . . 137
9.3.2 Fluxes in Type 3 estuaries . . . . . . . . . . . . . . . . . . . . . 138
9.3.3 Flushing time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
9.4 Knudsens Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
10 Ocean waves 141
10.1 Preview: what were trying to understand . . . . . . . . . . . . . . . . . . 141
10.2 Phase velocity and group velocity . . . . . . . . . . . . . . . . . . . . . . 141
10.3 Capillary Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
10.4 Surface gravity Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
10.4.1 Deep-water (or short) waves . . . . . . . . . . . . . . . . . . . . 145
10.4.2 Shallow-water (or long) waves . . . . . . . . . . . . . . . . . . . 146
10.4.3 Standing Waves, Seiches, Resonance . . . . . . . . . . . . . . . . 149
10.4.4 Energy in surface gravity waves . . . . . . . . . . . . . . . . . . . 150
10.4.5 Momentum in surface gravity waves . . . . . . . . . . . . . . . . . 151
10.5 Internal gravity waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
10.5.1 Two-layer internal gravity waves . . . . . . . . . . . . . . . . . . . 151
10.5.2 Continuously-stratied internal gravity waves . . . . . . . . . . . . 152
10.6 Effects of rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
CONTENTS ix
11 Turbulence and mixing 153
11.1 Characteristics of Turbulence . . . . . . . . . . . . . . . . . . . . . . . . . 153
11.2 Measuring turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
11.3 Turbulent scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
11.3.1 Ozmidov Scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
11.3.2 Batchelor Scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
11.4 Effect of turbulence on the mean ow . . . . . . . . . . . . . . . . . . . . 155
11.5 Bottom Boundary Layers in Rotating Fluid . . . . . . . . . . . . . . . . . 156
11.5.1 The Geostrophic Layer . . . . . . . . . . . . . . . . . . . . . . . . 156
11.5.2 The Ekman Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
11.5.3 Constant-stress Layer (aka Log Layer) . . . . . . . . . . . . . . . . 157
11.5.4 Laminar Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
11.6 Mixing, diffusion and dispersion . . . . . . . . . . . . . . . . . . . . . . . 157
12 Light in the sea 161
12.1 Electromagnetic radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
12.1.1 Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
12.1.2 Speed of Light . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
12.1.3 Index of Refraction . . . . . . . . . . . . . . . . . . . . . . . . . . 161
12.1.4 Net radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
12.1.5 Stefan-Boltzmann law for black-body radiation . . . . . . . . . . . 161
12.1.6 Weins Displacement Law for Spectral Peak . . . . . . . . . . . 163
12.1.7 Effect of radiation at the sea surface . . . . . . . . . . . . . . . . . 163
Refraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
Reection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
12.1.8 Effects on radiation passing through sea water . . . . . . . . . . . . 163
13 Tides 165
13.1 Measuring tides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
13.2 Tidal Forcing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
13.2.1 Nontechnical explanation of tidal forces . . . . . . . . . . . . . . . 167
13.2.2 Technical explanation of tidal forces . . . . . . . . . . . . . . . . . 168
13.3 Tidal analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
13.4 Empirical tidal prediction . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
13.5 Tidal prediction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
13.6 Tidal mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
13.7 Residual tidal effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
13.8 Tidal power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
13.8.1 General considerations . . . . . . . . . . . . . . . . . . . . . . . . 175
13.8.2 Tidal power in Bay of Fundy . . . . . . . . . . . . . . . . . . . . . 176
13.8.3 Political considerations . . . . . . . . . . . . . . . . . . . . . . . . 177
14 Ocean Acoustics 179
14.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
14.2 Ocean acoustic spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
14.3 Sound and compressibility . . . . . . . . . . . . . . . . . . . . . . . . . . 179
14.4 SOFAR (deep sound) channel 1 . . . . . . . . . . . . . . . . . . . . . 180
14.5 SOFAR (deep sound) channel 2 . . . . . . . . . . . . . . . . . . . . . 180
14.6 Refraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
14.7 AD(C)P acoustic doppler (current) proler . . . . . . . . . . . . . . . . . 182
14.8 RAFOS Instrument . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
14.9 Acoustic Tomography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
14.10ATOC experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
x CONTENTS
V Extras 187
15 Laboratory Experiments 189
15.1 Salt-ngers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
15.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
15.1.2 Method 1: heat/dye ngers . . . . . . . . . . . . . . . . . . . . . . 189
15.1.3 Method 2: sugar/salt ngers . . . . . . . . . . . . . . . . . . . . . 189
16 Dealing with Data 191
16.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
16.2 Signicant digits and uncertainties in calculation . . . . . . . . . . . . . . 191
16.3 Processing density proles . . . . . . . . . . . . . . . . . . . . . . . . . . 192
16.4 Calculating the Buoyancy Frequency . . . . . . . . . . . . . . . . . . . . . 194
16.5 Coordinate systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
16.5.1 Orthogonal coordinate systems . . . . . . . . . . . . . . . . . . . . 196
16.5.2 Right-handed coordinate systems . . . . . . . . . . . . . . . . . . 196
16.5.3 Angles in Geography, Meteorology, and Oceanography . . . . . . . 196
Angle conversions . . . . . . . . . . . . . . . . . . . . . . . . . . 197
Wind from, currents towards . . . . . . . . . . . . . . . . . . . . . 197
16.5.4 How to rotate cartesian coordinates . . . . . . . . . . . . . . . . . 197
16.5.5 Latitude-longitude versus x y coordinates . . . . . . . . . . . . . 198
16.6 Units: the good, the bad, and the ugly . . . . . . . . . . . . . . . . . . . . 198
16.7 Pitfalls in computing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
16.8 Dealing with mathematically-based text . . . . . . . . . . . . . . . . . . . 199
17 Important Numerical Values 201
17.1 General Properties of Matter . . . . . . . . . . . . . . . . . . . . . . . . . 201
17.2 Properties of the Earth . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
17.3 Properties of Seawater . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
17.4 Salt Water and Sea-ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
17.4.1 Sea-ice: Thermal Conductivity . . . . . . . . . . . . . . . . . . . . 201
17.4.2 Sea-ice: Specic Heat . . . . . . . . . . . . . . . . . . . . . . . . 202
17.5 Fresh Water and Pure Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
17.5.1 Latent Heat of Freezing . . . . . . . . . . . . . . . . . . . . . . . 202
17.5.2 Latent Heat of Vaporization . . . . . . . . . . . . . . . . . . . . . 202
17.5.3 Latent Heat of Sublimation . . . . . . . . . . . . . . . . . . . . . . 202
17.5.4 Pure Ice: Density . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
17.5.5 Pure Ice: Specic Heat Capacity . . . . . . . . . . . . . . . . . . . 203
17.5.6 Pure Ice: Thermal Diffusivity . . . . . . . . . . . . . . . . . . . . 203
17.6 Properties of Air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
17.6.1 Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
17.6.2 Specic Heat Capacity . . . . . . . . . . . . . . . . . . . . . . . . 203
17.6.3 Thermal Diffusivity . . . . . . . . . . . . . . . . . . . . . . . . . . 203
17.7 Properties of the Ocean . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
CONTENTS xi
18 Techniques 205
18.1 Symbolic Mathematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
18.2 The Matlab Language . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
18.3 The R Language . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
18.4 Numerical Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
18.4.1 Who should do modelling? . . . . . . . . . . . . . . . . . . . . . . 208
18.4.2 History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
18.4.3 Computational cost . . . . . . . . . . . . . . . . . . . . . . . . . . 210
Why compute the cost? . . . . . . . . . . . . . . . . . . . . . . . . 210
Online cost calculator . . . . . . . . . . . . . . . . . . . . . . . . 210
Cost in RAM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
Cost in time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
Range of processor capabilities . . . . . . . . . . . . . . . . . . . 212
18.4.4 Example: inertial motion . . . . . . . . . . . . . . . . . . . . . . . 212
18.4.5 Eulerian and Lagrangian models . . . . . . . . . . . . . . . . . . . 213
18.4.6 Vertical grid schemes . . . . . . . . . . . . . . . . . . . . . . . . . 213
18.4.7 Finite-difference versus nite-element models . . . . . . . . . . . . 216
18.4.8 Space-time applications: explicit schemes . . . . . . . . . . . . . 216
18.4.9 Space-time applications: implicit schemes . . . . . . . . . . . . 218
18.4.10 Special problems for ocean models . . . . . . . . . . . . . . . . . 220
18.4.11 Code example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
18.5 Data Assimilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
19 People in Oceanography 225
20 Physical Oceanography in Just Two Pages 229
20.1 Gravitational stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
20.2 Hydrostatic Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . 229
20.3 Geostrophy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
20.4 Ekman Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
20.5 Potential Vorticity, Sverdrup ow . . . . . . . . . . . . . . . . . . . . . . . 230
20.6 Rossby Number, Radius, . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
20.7 Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
21 Answers to Exercises 231
xii CONTENTS
List of Figures
1.1 Geometry of a bathtub . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2 Geometry of Pacic Ocean . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Topography of earth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Stratication in a bathtub . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.5 Stratication in the ocean . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.6 Coriolis effect on rotating planet . . . . . . . . . . . . . . . . . . . . . . . 7
1.7 Wind-driven upper-ocean circulation . . . . . . . . . . . . . . . . . . . . . 9
1.8 Thermohaline circulation in a bathtub . . . . . . . . . . . . . . . . . . . . 10
1.9 Thermohaline circulation in the world ocean . . . . . . . . . . . . . . . . . 10
1.10 Cross-section of continental shelf . . . . . . . . . . . . . . . . . . . . . . . 11
1.11 Side view of shelf-edge front . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.12 Locations of shing boats off Eastern coast of Canada . . . . . . . . . . . . 13
2.1 Denition sketch of a function . . . . . . . . . . . . . . . . . . . . . . . . 21
2.2 Denition sketch for Eulers Rule . . . . . . . . . . . . . . . . . . . . . . 24
2.3 Denition sketch of a derivative . . . . . . . . . . . . . . . . . . . . . . . 25
3.1 Diver jumping off a bridge . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.1 Typical WOCE sampling density . . . . . . . . . . . . . . . . . . . . . . . 42
4.2 Global surface temperature . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.3 Seasonal variation of upper-ocean temperature . . . . . . . . . . . . . . . . 44
4.4 Zonal temperature section . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.5 Temperature variation with latitude and depth . . . . . . . . . . . . . . . . 46
4.6 Schematic temperature prole evolution . . . . . . . . . . . . . . . . . . . 47
4.7 Schematic of depth variation in polar waters . . . . . . . . . . . . . . . . . 47
4.8 Variation of temperature T and potential temperature in deep waters . . . 48
4.9 Global surface temperature February) . . . . . . . . . . . . . . . . . . . . 49
4.10 Surface salinity in relation to evaporation minus precipitation . . . . . . . . 49
4.11 Prole showing Mediterranean water . . . . . . . . . . . . . . . . . . . . . 50
4.12 Niskin and Nansen bottles . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.13 CTD schematic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.14 Salinity in terms of conductivity ratio and temperature . . . . . . . . . . . 56
4.15 Equation of state in C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.16 Variation of
t
with temperature and salinity . . . . . . . . . . . . . . . . . 60
4.17 Denition sketch of stratication . . . . . . . . . . . . . . . . . . . . . . . 61
4.18 Stratication in St Lawrence Estuary . . . . . . . . . . . . . . . . . . . . . 62
4.19 Denition sketch for potential energy. . . . . . . . . . . . . . . . . . . . . 63
4.20 Denition sketch of wind prole . . . . . . . . . . . . . . . . . . . . . . . 66
4.21 Map of air-sea heat ux . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.22 Sample prole diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.23 Schematic temperature-salinity diagram . . . . . . . . . . . . . . . . . . . 70
xiii
xiv LIST OF FIGURES
4.24 T-S diagram on Scotian Shelf . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.25 Schematic of salt fountain thought experiment . . . . . . . . . . . . . . . 72
4.26 Molecular diffusion in seawater . . . . . . . . . . . . . . . . . . . . . . . . 73
4.27 Ocean regions that are susceptable to diffusive instability . . . . . . . . . . 74
4.28 Meridional oxygen section in Atlantic . . . . . . . . . . . . . . . . . . . . 75
5.1 Flow in a narrowing pipe . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5.2 Flux into one side of a box-shaped uid element . . . . . . . . . . . . . . . 81
5.3 Heuristic derivation of continuity equation . . . . . . . . . . . . . . . . . 81
5.4 Pressure forces exerted on opposite sides of a box-shaped uid element . . 84
5.5 Derivation of coriolis equation . . . . . . . . . . . . . . . . . . . . . . . . 86
5.6 Denition sketch of molecular friction . . . . . . . . . . . . . . . . . . . . 88
5.7 Denition sketch of turbulent friction . . . . . . . . . . . . . . . . . . . . 89
5.8 Flow at moderate Re . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
5.9 Denition sketch of isobar eld . . . . . . . . . . . . . . . . . . . . . . . . 94
5.10 Velocity in geostrophic balance . . . . . . . . . . . . . . . . . . . . . . . . 94
5.11 Flow on a weather map . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
5.12 Denition sketch for one-layer ow . . . . . . . . . . . . . . . . . . . . . 97
5.13 Denition sketch for two-layer ow . . . . . . . . . . . . . . . . . . . . . 98
5.14 Constraint on geostrophic ow with sloped bottom . . . . . . . . . . . . . 101
5.15 Slab model of wind response . . . . . . . . . . . . . . . . . . . . . . . . . 103
5.16 Ekman spiral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
6.1 Schematic of Hadleys model of atmospheric circulation . . . . . . . . . . 110
6.2 Schematic of 3-cell model of atmospheric circulation . . . . . . . . . . . . 110
6.3 Schematic of wind stress on North Atlantic Ocean . . . . . . . . . . . . . . 111
6.4 Actual wind stresses on North Atlantic . . . . . . . . . . . . . . . . . . . . 111
6.5 Conservation of potential vorticity . . . . . . . . . . . . . . . . . . . . . . 114
6.6 Ekman pumping and Sverdrup ow . . . . . . . . . . . . . . . . . . . . . 115
6.7 Rossby waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
7.1 Schematic thermocline balance . . . . . . . . . . . . . . . . . . . . . . . . 120
7.2 W usts north-south cross-sections in North Atlantic . . . . . . . . . . . . . 121
7.3 Stommels schematic diagram of the thermohaline circulation . . . . . . . . 122
7.4 Munks thermocline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
7.5 Meridional heat ux according to Ganachaud & Wunsch (2000) . . . . . . 124
8.1 Coastal upwelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
8.2 Coastal upwelling, with Ekman divergence and upwelling . . . . . . . . . . 127
8.3 Shelf-edge upwelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
8.4 Shelf waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
8.5 Buoyancy currents in Northwest Atlantic shelf . . . . . . . . . . . . . . . . 130
8.6 Shelf-edge front . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
8.7 Shelf-edge front: 3D view . . . . . . . . . . . . . . . . . . . . . . . . . . 131
8.8 Infrared image of California Current . . . . . . . . . . . . . . . . . . . . . 132
9.1 Estuarine section . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
9.2 Estuary classication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
9.3 Hansen-Rattray regime diagram . . . . . . . . . . . . . . . . . . . . . . . 137
LIST OF FIGURES xv
9.4 Side view of type 3 estuary . . . . . . . . . . . . . . . . . . . . . . . . . . 138
9.5 Estuarine circulation calculation . . . . . . . . . . . . . . . . . . . . . . . 139
9.6 Denition sketch for tidal ushing time . . . . . . . . . . . . . . . . . . . 139
9.7 Denition sketch for Knudsen theorem . . . . . . . . . . . . . . . . . . . 140
10.1 Large wave with surfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
10.2 Denition sketch of propogating wave . . . . . . . . . . . . . . . . . . . . 142
10.3 Group velocity of waves . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
10.4 The tanh() function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
10.5 Phase speeds of deep-water and shallow-water waves . . . . . . . . . . . . 147
10.6 Particle paths in deep-water and shallow-water waves . . . . . . . . . . . . 147
10.7 Refraction of shallow-water waves . . . . . . . . . . . . . . . . . . . . . . 148
10.8 Focussing of shallow-water waves on headlands . . . . . . . . . . . . . . . 148
10.9 Denition sketch of seiche . . . . . . . . . . . . . . . . . . . . . . . . . . 149
10.10Denition sketch for calculation of wave potential energy . . . . . . . . . . 150
10.11Side view of wave at interface in two-layer uid . . . . . . . . . . . . . . . 151
10.12Nonlinear (soliton) waves . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
12.1 Stefan-Boltzmann Law for Blackbody Radiation . . . . . . . . . . . . . . . 162
12.2 Weins law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
13.1 M
2
tides in the North Atlantic . . . . . . . . . . . . . . . . . . . . . . . . 166
13.2 Stilling well . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
13.3 Moon and Earth pair . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
13.4 Tidal forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
13.5 Tides in Halifax Harbour . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
13.6 Spring-neap tides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
13.7 Tide-prediction machine . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
13.8 Webtide domain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
13.9 Global M2 tides inferred by satellite . . . . . . . . . . . . . . . . . . . . . 174
13.10M
2
tides in Bay of Fundy and Gulf of Maine . . . . . . . . . . . . . . . . . 176
14.1 Acoustic power spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
14.2 Sound speed prole . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
14.3 Sound rays in SOFAR channel . . . . . . . . . . . . . . . . . . . . . . . . 182
14.4 ADP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
14.5 ADCP data display . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
14.6 ATOC experiment in Greenland Sea . . . . . . . . . . . . . . . . . . . . . 184
14.7 ATOC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
16.1 N
2
calculation in R . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
16.2 Latex example 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
16.3 Latex example 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
18.1 Symbolic-mathematics software demonstration. . . . . . . . . . . . . . . . 206
18.2 Plotting CTD data in R language . . . . . . . . . . . . . . . . . . . . . . . 207
18.3 Numerical simulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
18.4 POM users group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
18.5 Numerical integration in Perl . . . . . . . . . . . . . . . . . . . . . . . . . 214
18.6 Numerical grid: level . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
xvi LIST OF FIGURES
18.7 Numerical grid: reduced gravity . . . . . . . . . . . . . . . . . . . . . . . 215
18.8 Numerical grid: isopycnal . . . . . . . . . . . . . . . . . . . . . . . . . . 215
18.9 Numerical grid: sigma . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
18.10Finite-element mesh for tidal model of east-coast Canada . . . . . . . . . . 217
18.11Computational molecule . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
18.12CFL condition for diffusion equation . . . . . . . . . . . . . . . . . . . . . 218
18.13Explicit solution for diffusion equation . . . . . . . . . . . . . . . . . . . . 219
18.14Snippet of Princeton Ocean Model (matrices) . . . . . . . . . . . . . . . . 221
18.15Snippet of Princeton Ocean Model (viscosity) . . . . . . . . . . . . . . . . 222
18.16Numerical grid: 3D view . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
19.1 Gordon Riley . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
19.2 Henry Stommel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
19.3 Walter Munk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
19.4 Harald Sverdrup portrait . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
List of Tables
1.1 Keywords used in text . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.2 Symbols used in text . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
8.1 Scales of some phenomena . . . . . . . . . . . . . . . . . . . . . . . . . . 125
11.1 Values of mixing parameters . . . . . . . . . . . . . . . . . . . . . . . . . 156
13.1 Principal tidal periods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
xvii
xviii LIST OF TABLES
Part I. Course scope
1
CHAPTER 1
Introduction & Overview
And all this science,
I dont understand
Its just my job,
ve days a week.
Rocketman c _1972 Elton John and Bernie Taupin
F
IRST, AN INTRODUCTION. I am Dan Kelley, a Physical Oceanographer at Dalhousie
University. You can nd me at room 5652 in the Oceanography building, phone me
at 494-1694, or email at Dan.Kelley@Dal.Ca. Please feel free to contact me Contact information
any time to ask questions; if demand gets exhausting, Ill set up specic ofce hours, but
until then, lets play it casually.
1.1. What is Physical Oceanography?
When I tell people in bars that Im an Oceanographer, they make several assumptions
about my work:
1. that my work must be really interesting (exact quote)
2. that I dive
3. that I work on whales
4. that Ive met the famous Bob Ballard (of Titanic fame no, not James Im the King
of the World Cameron, rather the science guy who discovered the location of that
famous ship on the ocean bottom).
Item (2) is probably the most common assumption, and the most wrong. Oh, to be misun-
derstood by the masses.
Physical Oceanography is the study of how the water in the ocean moves, mixes, and
interacts with the atmosphere above and the land below. Its also the study of why the water
moves, mixes, and interacts. Physical Oceanography (almost always abbreviated PO in
class) is important for issues ranging from the sublime to the ridiculous: e.g., understanding
the massively complex and interconnected climate system of the earth, developing strategies
for disposing of nuclear wastes, and telling Halifax where to put a sewage outow.
Physical Oceanographers study currents, waves, tides, mixing, fronts, and so on. Much of
this is pure research, but a good deal of it is service-oriented. By that I mean that often the
reason we study particular systems is to help Biological Oceanographers understand patterns Topics in PO
of phytoplankton, zooplankton, and sh abundance, or to help Chemical Oceanographers
understand nutrient and tracer distributions, or even to help Geological Oceanographers
understand beach dynamics or sediment distributions.
Physical Oceanographers play an important role in these other disciplines because the
motion of the sea very often determines the biology, chemistry, etc, in the sea. For example, Multidisciplinary aspect
productivity is high on the continental shelves (and hence humanity can use sh for a large
part of its protein requirement) largely because the currents and mixing supply the biology
with warmth, light and nutrients, and protect it from dispersion; thus sheries focus on the
edge of the continental shelf because of the peculiar Physical Oceanography there.
The connection of Physical Oceanography to other aspects of Oceanography is a tradi-
tional motivation I would have told you about in similar terms 10 years ago. But now there is
a sweeping new reason for keen interest in Physical Oceanography. You are all aware of the
feared greenhouse effect. According to many indications, the climate depends partly on Climate
the ocean. This means that to understandand thus predictclimate, well have to understand
the ocean.
3
4 CHAPTER 1. INTRODUCTION & OVERVIEW
1.2. About the Course
When I was a postdoc, sitting on a beach in Cape Cod, I thought of how I might teach a
course like this. (Then, of course, I had no idea that I ever would actually do it!) I wanted to
think of a format that would accomplish three goals.
Hold the interest an introductory class made up of students, some of whom (gasp!)
were not planning further study in Physical Oceanography. (Many texts do this fairly
well.)
Ensure that the class gets a solid overview of the principles involved in Physical
Oceanography as well as being exposed to specic cases. That is, ensure that the
students see the forest as well as the trees. (Few courses or texts do this.)
Place the textbook theories in a realistic context with actual observations. (No course
or text I know of does this well.)
I think these goals are accomplished reasonably well by the following scheme, which is
used in this class.
1. Start with observations of the ocean, to establish phenomenology.
2. Use this as a springboard for explaining the fundamentals of wind-driven, thermoha-
line, and shelf circulation.
3. Reinforce this train of thought, and preview the topics to be covered in the course per
se, by contrasting the dynamics of the ocean with the dynamics of more familiar uid
systems.
4. Later on, in subsequent lectures, introduce each new major subject with the facts rst,
followed by theory later.
5. As the course continues, continually refer back to (1), (2) and (3).
In class Ill show you a gallery of satellite photographs, and provide a free-ranging
discussion of them. You may nd the next section helpful, in consolidating what Ive said in
class.
1.3. Satellite images
Satellite imagery is revolutionizing Oceanography
1
. The global viewpoint answers
questions about the ocean that we hadnt thought to ask previously. Before, all we had to go
on was ship observations. Heres why that is frustrating: Imagine a ship, steaming in a line
from Halifax to Bermuda
2
. Every so often, the ship makes measurements of temperature,
salinity, etc, developing a prole of water properties at an (x, y) point. String these points Prole
together along the ship track, and you have done a section. Great. Now you know what Section
the state of the ocean was along a single plane over a particular time period. But what was
the ocean like 100 km to starboard or port? You simply dont know and cant guess. Its a
lot like drawing a line along a complicated wallpaper pattern and reading off the colours
under the line: you cannot hope to gure out the pattern, let alone decide whether it contains
roses or red carnations. Of course, weve long tried to get around the problem by sending
out several ships at a time, and by deploying instruments in the ocean to record properties
over long periods of time. This helps a lot, and is still the only way to really know what goes
on below the surface, but its simply too expensive
3
to do much of. As you can imagine,
the results are often exceedingly confusing, since the ocean is not a piece of wallpaper; the
patterns dont repeat, they are not simple, and they change over a whole range of timescales.
1
The satellite people tell us that it will change the world very soon in a year or two but theyve been telling
us that for decades now so a grain of salt is needed.
2
Guess at what time of year this scientic cruise is overwhelmed by volunteers!
3
A ship costs something like $10 000 to $20 000 per day, so a day of ship time is roughly equivalent to an annual
graduate-student stipend . . . professors who have to pay for both need to weight the benets of these two things!
1.4. HOW THE OCEAN IS UNLIKE A BATHTUB 5
Figure 1.1: Geometry of a Bathtub, shown in side view. Note that the lengthscales H (water
depth) and L (width of basin) are similar that is, within a factor of 10 of each other.
Figure 1.2: Geometry of the Pacic Ocean, shown in side view. The water depth, at this
scale, is actually much thinner than the line!
Satellites, on the other hand, provide a global view of things. We can see meanders of
currents, the complicated interactions between colliding currents, etc. Satellites give us
only a surface view, usually of temperature (top few millimeters) or colour (top few meters)
or sealevel (top of the top!). This is a real limitation, but on balance satellites provide an
important addition to our toolbox. Ill tell you more about how they are used in the weeks
ahead.
To be honest, theres another reason why Physical Oceanographers are attracted to
satellite data. As a group, we are pasty-faced nerds who suffer severe sensory deprivation
between anxiety attacks. Some of us like to look at satellite images partly because they are
pretty. I hope you think they are pretty too. More importantly, I hope you read the captions
on the back of this set of NASA images, and reread them as the course goes on.
1.4. How the ocean is unlike a bathtub
Many of the essential elements of Physical Oceanography can be pointed out effectively
by contrast with something familiar to us all: a bathtub. And so we might ask How is the
ocean unlike a bathtub?
1.4.1. Geometry
A bathtub is almost as deep as it is long (Fig. 1.1). Now, consider the Pacic, which is
3000 km wide and 3 km deep. Obviously the aspect ratio is radically different from that of a
bathtub. To draw the Pacic with a 0.5 mm pencil, Ill have to draw a line 50 cm long. It
wont even t on a page! So in Figure 1.2 I show a scale drawing of just the Eastern half of
the Pacic.
The conclusion is that oceanic geometry is very different from bathtub geometry. You
can imagine the ocean as an exceedingly thin skin sitting atop the earth, much like a layer of
onionskin. This geometry is not just a curiosity; this fact has some crucial consequences for
the circulation patterns of ocean currents, as Ill explain in the weeks to come.
It is also worth pointing out that the ocean bottom is not at; Figure 1.3 illustrates this.
6 CHAPTER 1. INTRODUCTION & OVERVIEW
Figure 1.3: Topography of earth. Note mid-ocean ridges, e.g. in the Atlantic, and also the
difference between coastlines and continental shelves. Readers who are unfamiliar with
map projections should note that this representation (linear in latitude) exaggerates the area
of high-latitude features; for example, Greenland seems to have the same area as South
America in this diagram, but it is actually 8 times smaller.
Figure 1.4: Density Stratication in a Bathtub. The left panel is a side view, and the right
panel is a graph of the water density as a function of depth in the water column. The density
is constant, so the line is straight.
1.4.2. Stratication
The water in your bathtub is equally dense everywhere
4
. However, in the ocean, the deep
water is signicantly denser than the shallow water. We say the water is stratied; this
word calls to mind layering, and in fact we often think of the ocean as consisting of a light
upper layer atop a heavy bottom layer. Figures 1.4 and 1.5 show the contrast. Stratied
systems have dense water on the bottom and light water above. It takes mechanical work to
raise heavy water through light water, so the stratication inhibits mixing of deep waters and
surface water. An important biological consequence is the limitation of supply of essential
nutrients (stored in the deep water) to the well-lit surface waters where phytoplankton live.
An important physical consequence of the increase of density with depth is that it gives
rise to internal waves in the ocean. These are much like the waves you can make in a
bottle containing water and oil. They are of interest because they can sometimes break,
much like surface waves do, and the result is vertical mixing. (Also, the military is really
quite interested in internal waves, because submarines moving through the water create
internal waves. You can gure the rest out.)
4
Unless you add a lot of bath oil!
1.4. HOW THE OCEAN IS UNLIKE A BATHTUB 7
Figure 1.5: Density Stratication in the Ocean. The left panel is a side view, and the right
panel shows that water density increases with depth in the ocean.
Figure 1.6: The Coriolis Effect Due to the Earths Rotation. Left panel: a projectile is shot
from the North Pole towards the dot. The intended projectile path through space is shown in
the heavy arrow curve. Longitude and latitude lines are drawn on the earth for reference.
Right panel: by the time the projectile gets South to the latitude of the dot, the Earth has
rotated underneath it, so that the projectile passes to the West of the dot.
Horizontal stratication variations of density along horizontal surfaces is crucially
important to oceanic motion since it causes horizontal variations of pressure. Horizontal
pressure gradients can drive currents. (Youll discover, however, that the relationship
between pressure force and current is quite counterintuitive, due to the inuence of the
rotation of the earth.)
1.4.3. Rotation
The rotation of the earth about its axis leads to the Coriolis effect. This is the tendency
for objects moving in the Northern Hemisphere to drift off towards the right of their motion.
In the Southern Hemisphere, the drift is towards the left of motion. The effect is strongest at
the poles and decreases smoothly to zero at the Equator.
Figure 1.6 is a sketch of how the Coriolis effect works. It shows a gun being red from
the North Pole towards Halifax
5
. Assume there is no air friction. The path of the projectile,
viewed from a point in space, is a great circle. But the path as viewed by an observer
sitting on the Earth is quite different, since the earth rotates independently of the path of
the bullet. By the time the bullet reaches as far south as Halifax, the earth will have spun
enough that it falls to the West of its intended target. The projectile path deviates towards
the right of the direction of motion in the Northern Hemisphere, to the left in the Southern
Hemisphere.
Ill argue (and, hopefully demonstrate in the following weeks) that you already under-
stand the main laws of uid mechanics just from everyday experience. But youve never
experienced the Coriolis effect directly, since the Coriolis effect is weak by the standards of
everyday life. When you walk to work you certainly dont slide off the right hand side of
the sidewalk
6
. If the Earth rotated a lot faster, though, you might have to compensate for the
5
OK, Halifax Regional Municipality, for political correctness.
6
Pub crawls aside.
8 CHAPTER 1. INTRODUCTION & OVERVIEW
Coriolis effect. Heres a formula describing the Coriolis effect:
dv
dt
=f u (1.1)
where u is the velocity component to the East, t is time, and v is the velocity component to
the North
7
. Here the Coriolis parameter is
f = 2sin (1.2)
where is the rotation rate of the Earth and is the latitude. Numerically, the Coriolis
parameter is dened by
f = 1.4510
4
s
1
sin (1.3)
Note that = 0 at the Equator, so sin0 = 0 and f = 0 there; the Coriolis effect vanishes
there. Similarly, = 90

at the North Pole, so f = 1.4510


4
s
1
there, and =90

at
the South Pole, so f =1.4510
4
s
1
there.
Example. Im on a big at lake covered with perfectly slippery ice. Someone (a student?)
pushes me toward the South, giving me a speed of 1 m/s. (Question: how fast do you walk, in
m/s?) At Halifax the latitude is 45

N, so f = 110
4
s
1
. According to equation (1.1), my
acceleration to the West will be 110
4
m/s
2
. To get an idea how big that number is, lets
substitute t = 1h = 3600s. The rough result is that my westward speed will increase from
0 m/s to something like 0.4 m/s. Thats almost half my intended forward speed! In other
words, Ill be going signicantly off course in an hour. Then, why dont we continually walk
in circles? The answer is that this acceleration is very low compared to the accelerations
generated by everyday forces. To see this, lets set this Coriolis acceleration equal to the
sideways gravitational acceleration you feel when you walk along a sloping hill. Then well
calculate the slope of the hill. We want the sideways component of gravitational acceleration
to be 1.210
4
m/s
2
; therefore the slope of the hill must satisfy
1.210
4
m/s
2
= (9.8m/s
2
)sin (1.4)
(since the accelaration due to gravity is g =9.8m/s
2
). Solving, we get =1.210
5
radians,
or = 0.0007

. This is damn small; the slope across a 6 m wide road is just 0.5 mm, the
thickness of the mark made by a mechanical pencil! Ask yourself whether you could see
such a tiny slope, and youll realize that the Coriolis force is very weak compared to most
forces you feel during the day!
In summary, the ocean is strongly affected by the Coriolis force simply because the other
forces are very weak.
1.4.4. Wind Forcing
The familiar Gulf Stream is a return ow for a Southward midlatitude oceanic ow
driven by wind stress. We call this circulation pattern a wind-driven gyre; specically, the
system containing the Gulf Stream is called a sub-tropical gyre. A similar sub-tropical gyre
exists in the South Atlantic, the North and South Pacic, and the Indian Oceans. So-called
sub-polar gyres exist poleward of these. Figure 1.7 shows the general pattern.
Notice the similarity between the gyres in the various oceans. This is because the patterns
of wind forcing (i.e. wind-stress) are similar, mostly being a function of latitude
8
. In
Chapter 6 you will see more about how these gyres work, but for now just note that the
essential ingredients to the dynamics are (1) wind forcing and (2) latitudinal
9
variations of
the Coriolis parameter f .
7
If you dont understand that the left hand side of this equation refers to the rate at which v varies with time,
and therefore represents acceleration, youd better bring this up in class. There will be a lot more where that came
from. Youll need to understand simple calculus like this to understand a lot of what I tell you in the coming weeks.
8
As the course progresses, you will see that the laws governing oceanic motion are the same as those governing
atmospheric motion; you may also nd that the patterns in weather maps start to make sense.
9
Oceanographers often refer to latitudinal variations in things like wind-stress. This means variations in a
south-north direction. The adjective [meridional??] is also used for this direction. By contrast, longitudinal
variations, also called [zonal??] variations, are in the east-west direction. The terms meridional and zonal are
also used to refer to averages, and this can get confusing, e.g. a zonally-averaged model has variables that depend
meridionally.
1.4. HOW THE OCEAN IS UNLIKE A BATHTUB 9
Figure 1.7: Upper panel: Schematic of wind-driven upper-ocean circulation, showing
subtropical gyres in both Atlantic and Pacic Oceans, along with subpolar gyres in the
northern halves of these two oceans. The lack of a meriodonal coastal boundary prevents
such gyres to the south of South America, and instead we observe a circumpolar current in
the southern ocean, running eastward around Antarctic continent. Lower panel: circulation
streamlines calculated using Sverdrup theory (Chapter 6) based on wind patterns (Source:
Figure 11.3 of Stewart [2005])
10 CHAPTER 1. INTRODUCTION & OVERVIEW
Figure 1.8: Thermohaline Circulation in a Bathtub (view from the side). Imagine that there
water is cooled at the top left corner of the tub, and warmed at the bottom right corner. The
cooling causes a density increase, and the warming causes a decrease, so the result is that
water sinks at the left-hand end of the tub and rises at the right-hand end. This sets up a
circulation cell.
Figure 1.9: Schematized thermohaline circulation in the world ocean (view from above).
[Pond and Pickard, 1978, Figure 10.1, after Stommel] Water sinks at S1 and S2.
1.4.5. Thermohaline Forcing
If you were to put a heater at the surface of one end of your bathtub and a cooler at the
other end, youd nd that water under the cooler would sink (since cool water is heavier
than warm water) and that water under the heater would rise. Youd get a circulation cell
like that in Figure 1.8.
In the ocean, a somewhat similar thing goes on. At rst guess, youd think the cooler was
at the poles and the heater at the Equator. Thats almost correct, but the geographical rate of
cooling and the geometry of the ocean bottom (among other things) make the situation more
complicated. In fact, there is a circulation cell, but it looks more like the conceptual sketch
in Figure 1.9.
Strong surface cooling and evaporation in the Northern North Atlantic Ocean and in the
waters near Antarctica (S1 and S2 in Figure 1.9) causes surface waters to get heavy and sink
to the depths. This is like the cooled side of the bathtub. As well see in later chapters, the
way this dense water is distributed to the oceans is not intuitive. It surges down western
boundaries of ocean basins, and more slowly distributes into the basins themselves. And,
there is no direct analogue of the heater in the bathtub; rather, heating is distributed along
1.5. COURSE TOPICS 11
Figure 1.10: Continental Shelf, drawn to scale. The black region is the earth below, and the
very thin white region is the ocean.
the surface. The upwelling occurs throughout the ocean basins.
1.4.6. Mixing
If you put some red dye in the bottom of your tub and stir a bit, youll end up with pink
bathwater. (Hey, its your party.) Things are quite different in the ocean. The mixing is
so weak, at least compared to the vertical density gradient, that there isnt enough energy
to move the heavy bottom water up to the surface. The timescale to mix bottom water to
the surface is hundreds of thousands of years. This isolation of surface waters and deeper
waters, and some exceptions to it, is fundamental to how the oceans work, and to how the
atmosphere-ocean climate system works.
1.4.7. Geometry (again): Continental shelves
In many ways, the continental shelves are the parts of the ocean of greatest direct interest
to humanity. Biological oceanographers care about continental shelves because thats where
the sh are (see below).
Figure 1.10 shows a picture of the continental shelf South of Nova Scotia, more or less
to scale. The shelf is about 100 m deep, compared to a deep ocean of 4000 m, and the
shelf width is about 100 km. These numbers roughly apply along this coast. The shelf
off California is 10 times narrower. That is important to the Physical Oceanography and
Biological Oceanography.
(By the way, this word, scale, is an important one. We will often focus on the scales
of features in the ocean. Some of the words youll hear include, in order of increase
in size: [micro-scale??], [ne-scale??], [shelf-scale??], [meso-scale??], [basin-scale??],
and [planetary-scale??]. Some of these scales are dynamical in nature (meso-scale and, to
some extent, ne-scale and micro-scale), while others are geographical (e.g. basin-scale).
Both varieties of scale are relevant, but often in different ways.)
The water over the continental shelves is generally lighter than in the deep sea because
of a combination of solar input, river input, etc. There is often a front between shelf waters
and deep-sea waters (Figure 1.11), and this front is of major interest to all oceanographers.
Its highly productive, as Figure 1.12 shows. This is because the deep-sea waters contain
little life and therefore lots of nutrients. The shelf break front combines these nutrient-rich
waters with the warmer well-stratied shelf waters.
1.5. Course topics
The Table of Contents gives an idea of what Ill cover in this course. Essentially all of
this material must be covered at least supercially to give you the knowledge I think
you need. However, I can do some of the topics either quickly or in depth. Please feel free
to steer the course in the direction of your interest, through questions and suggestions. Ill
solicit this feedback often, but Id be a lot happier if you give it freely. (The only dumb
question is the one you dont ask.)
12 CHAPTER 1. INTRODUCTION & OVERVIEW
Figure 1.11: Side view of shelf-edge front. The blue curves are isopycnals, contours of
equal density. A tight packing of isopycnals indicates that density varies rapidly with depth
in the water column. Thus, the shallow waters at the left are homogeneous in density, in
contrast to those in deep water. The boundary between this well-mixed shallow water and
the stratied deeper water is a front, and its signature at the surface is a rapid change in
water properties, such as temperature.
1.6. Table of symbols and keywords
Table 1.1 lists keywords that have special meaning in Physical Oceanography, and table
1.2 lists some Greek and mathematical symbols Ill be using here. You may also nd it
helpful to consult the glossary, near the end of this document.
Whats Next. The next two chapters review some mathematical and physical topics. Most
students will be familiar with these things, and so may wish to skip to chapter 4.
1.6. TABLE OF SYMBOLS AND KEYWORDS 13
75 70 65 60 55 50 45 40
4
0
4
5
5
0
5
5
1400
1200
1000
800
600
400
200
50
100
150
200
250
300
350
400
450
500
500
1000
1500
2000
2000
3000
4000
5000
6000
7000
8000
library(oce)
pdf("eastern_canada_topo.pdf", width=10, height=7)
eastern.canada <- read.topo("eastern_canada_topo.dat")
par(cex=0.8)
plot(eastern.canada,
water.z=-c(seq(50,500,50), seq(500,2000,500),
seq(2000,8000,1000)))
Figure 1.12: Top: locations of shing vessels. The dataset covers 1989 to 2002, and
includes ships that are outbound from, or inbound to, Canadian ports; this explains the
absence of ships on the rich shing grounds of e.g. Georges Bank. The colour-coded dots
represent vessel counts in regions spanning 6-minute of latitude and longitude. (The diagram
was constructed from ECAREG data by Angelia Vanderlaan, a Dalhousie PhD student, who
owns the copyright.) Middle: bathymetry of region. Bottom: R code used to create middle
panel.
14 CHAPTER 1. INTRODUCTION & OVERVIEW
Table 1.1: Some keywords used in the text.
Keyword Meaning
to be advected to be carried along by a current
advection of . . . transport of . . . by a current
baroclinic state with constant-P surfaces and constant- surfaces at angle
barotropic state with constant-P surfaces and constant- surfaces parallel
closed boundary a boundary that reects waves (e.g. coastline)
diagnostic model model in which velocity is inferred from measured density eld
dynamic height horizontal gradient of this yields geostrophic velo. (5.4.5)
efold reduce by a factor of 1/e
front region in which gradient of water properties is anomalously large
geopotential surface a surface along which a ball wont roll (g =)
geostrophic state with pressure-gradient force balancing coriolis force(5.4)
gyre region with closed streamlines (or closed mean streamlines)
homogeneous same at all locations (e.g. homogeneous turbulence)
instability growing perturbation
inviscid frictionless
isobaric having same pressure (e.g. isobaric surface)
isopycnal having same uid density
isotropic same in all directions (e.g. isotropic turbulence)
M2 semidiurnal moon tidal constituent with period 12.42h
mean statistical average
open boundary a boundary that doesnt reect waves (e.g. opening of a cove)
perturbation signal - mean(signal)
prognostic model model in which variables (e.g. u, v) are time-stepped (predicted)
Rossby adjustment response to suddenly imposed pressure gradient, ending in geostropy
S2 semidiurnal sun tidal constituent with period 12.00h
statistically stationary same stats (e.g. variance) at all times or all locations
steady-state not varying with respect to time t
streamline line indicating path of a 2-D current
(ow cannot cross a streamline)
1.6. TABLE OF SYMBOLS AND KEYWORDS 15
Table 1.2: Some symbols used in the text, with units in square brackets and dening
references following the symbol..
Symbol Name Meaning
Greek Delta A change in something
partial or di Partial derivative
Greek beta d f /dy where y is north-south distance (5.2.4)
Greek eta Sea-surface elevation [m]
Greek zeta Interfacial elevation [m]
Greek nu Molecular kinematic viscosity 10
6
[m
2
/s]
Greek mu Dynamic viscosity, = [Pas = kg/(ms)]
Greek rho Mass density [kg/m
3
] (4.7)
Greek sigma (S, T, p) 1000 [kg/m
3
] (4.7)

t
(S, T, 0) 1000 [kg/m
3
] (4.7)
Greek Omega Rotation vector of earth [1/s] (5.2.4)
f Coriolis parameter 1.4610
4
sin(latitude) [1/s] (5.2.4)
g - Gravitational accel. [m/s
2
]
g
/
- Reduced grav. accel. = g/
0
[m/s
2
]
h or H Water depth [m]
N Buoyancy freq. Bobbing freq. of internal waves [1/s] (4.9)
S - Salinity [PPT or PSU] (4.1)
T - Temperature [

C] (4.1)
Greek theta Potential temperature [

C] (4.3.3)
A
V
- Vertical eddy viscosity [m
2
/s] (5.2.6)
A
H
- Horizontal eddy viscosity [m
2
/s] (5.2.6)
K
V
- Vertical eddy diffusivity [m
2
/s] (5.2.6)
K
H
- Horizontal eddy diffusivity [m
2
/s] (5.2.6)
L
E
Ekman length (2A
V
/f )
1/2
Thickness of veering Ekman layer (5.6)
Re Reynolds no. UL/ Low value laminar ow (5.3.3)
Ri Richardson no. N
2
/(u/z)
2
Low value mixing in stratied uid (9.2)
Ro Rossby no. U/( f L) Low value geostrophy (5.3.4)
L - Typical length scale of a system [m]
L
R
(gh)
1/2
/ f or (g
/
h)
1/2
/f Natural lengthscale of fronts [m] (8.3)
x - Vector coordinate [m] (2.7)
u - Vector velocity [m/s] (2.7)
u, v, w - Components of velocity vector [m/s]
x, y, z - Components of position vector [m]
t - Time [s]
p or P - Pressure [Pa=N/m
2
= kgm
1
s
2
] (5.2.1)
U - Typical speed scale of a system [m/s]
16 CHAPTER 1. INTRODUCTION & OVERVIEW
Part II. Review
17
CHAPTER 2
Mathematical tools for Physical
Oceanography
Ill stay home forever
Where two & two always
makes up ve
The lukewarm c _2003 Radiohead
Cher: If its a concusion, you have to keep her awake. Ask her questions.
Elton: Whats 7 times 7?
Cher: No, stuff she knows.
Clueless c _1995
T
WO PLUS TWO EQUALS FOUR. Its four whether you count on your ngers or on
your toes, whether you count with sticks or stones, whether with a slide-rule
1
or a
supercomputer. Its four whether youre a bleeding-heart liberal or a mean-spirited
right-winger, whether you be black, white, male, female, young, old . . . . The thing is that
the statement two plus two equals four doesnt depend on speaker at all, and that gives it
a universal truth.
Mathematical truth come in handy, sometimes. In particular, it is the strength of this
mathematical foundation that holds up Physics and other sciences. If you believe an equation
(perhaps f = ma) then you must believe, without further thought, the double of that equation
(2f = 2ma). And doubling isnt all we can do. If you believe an equation, oh lets just say
f v =
1

P
x
(2.1)
[which youll nd out later is a geostrophic equation, (5.76)], then you must believe that
the vertical derivative of the equation is also true, that is
2
f
v
z
=
1

x
_
P
z
_
(2.2)
Whats more, if you believe that A = B and that B =C, then you must believe that A =C.
This means that you can not just multipy and differentiate equations, but you can combine
them also. For example, knowing that P/z =g [which youll later come to know/love
as the hydrostatic equation, (5.34)] and using it in the above gives us what well be calling
the thermal wind equation (5.83), i.e.
v
z
=
g
f

x
(2.3)
Now, some folks can get all tied up in knots trying to gure out what this last equation
means. Thats ne, but it could take you a lot of time. And, in the meantime, if you know
what the geostrophic equation means, and what the hydrostatic equation means, then
why not save your thinking time for concepts that are actually new (e.g. angular-momentum
conservation, for example) . . . let the mathematics take care of the details. Relying on the
mathematical transformations is no more dangerous than relying on two plus two equals
four.
OK, now that Ive got that off my chest, I should say that this chapter contains a review
of the mathematics (calculus, mainly) that youll need for the course. It is what youd need
to do a rst year undergraduate Physics course or an advanced Physics class in high school.
1
Ask an old prof what that is, but rst get a friend to promise to phone you in a few minutes, so you can escape
gracefully.
2
I hope some of you are confused right now. What about the , you ask, cant it vary? Well, to those who are
confused, I say youre on the right track, and ask for patience until I discuss something called the Boussinesq
approximation. To the others, I suggest you stop believing the writer and pay more attention to the reader.
19
20 CHAPTER 2. MATHEMATICAL TOOLS FOR PHYSICAL OCEANOGRAPHY
2.1. The beauty of proof
In this course, little attention is paid to mathematical rigour, in the sense that results are
typically stated without a proof being offered. This is typical in science, and its appropriate
for this class, given the background and plans of the students, and the need to focus on
the matters at hand. However, there is something beautiful about mathematical proofs,
something I might be able to illustrate with a simple example.
First, a denition (for more, see Section 2.3.1). A rational number is one that can be
written as the ratio of two integers from which common factors have been removed. Any
number that is not rational is called irrational.
Theorem.

2 is an irrational number.
Proof. Assume, for the moment, that the theorem is false. Therefore, we can
write

2 =
a
b
where a and b are integers that have no common divisors. Squaring both sides
and rearranging yields
a
2
= 2b
2
.
The factor 2 in this equation dictates that a
2
must be an even number (i.e.
a multiple of 2). But since a factor of 2 cannot appear from nowhere upon
squaring an integer (a fact you should try to prove), it follows that a must also
be a multiple of 2. This means that a
2
must be a multiple of 4. Therefore, by
the above equation, b
2
must be a multiple of 2 . . . and that means that b must
be a multiple of 2.
We have shown that a and b share 2 as a common factor. And now we come
to the crux of the matter. If a and b share any common factor, then we have
contradicted the original assumption that

2 is rational.
The only way out of the contradiction is to admit that the original assumption
must have been false. In other words,

2 is an irrational number. This proves


the theorem.
You may be interested to learn that this proof is a hundred generations old, having been
constructed, or at least recorded, by Pythagoras
3
of Samos (569BC475BC).
I nd it remarkable that such an elegant thing as this proof was created so very long ago,
in an age that many might regard as crude. And we can only wonder whether other sentient
beings have pondered these same truths, for none of this is tied to our culture, our species,
even our world.
2.2. Functions
Functions are the building block of mathematics. When we state that f is a function of t,
we mean that at each value of t there is one and only one value of f . (This restriction that f
have only one value at any particular value of t is sometimes relaxed, the result being called
multivalued functions, but this is far beyond the present scope.). See gure 2.1.
We call t the independent variable (since it is supposed to take on any value it pleases)
and f the dependent variable (since its value is determined by the value of t). The notation
f (t) is typically spoken as f of t or f at t.
It may help to give a few examples of functions.
The [linear??] function includes such equations as
f (t) = 1+2t (2.4)
3
http://www-history.mcs.st-and.ac.uk/

history/Mathematicians/Pythagoras.
html.
2.2. FUNCTIONS 21
Figure 2.1: Denition sketch of a function. The graph in the left panel represents a function:
at any given x, there is only one value for f . On the other hand, the right panel shows
something that is not a function; note that f has two values at some values of x (e.g. x = x
0
),
for instance.
f (t) = a+bt (2.5)
If you draw a graph of f (t) on the vertical axis and t on the horizontal axis, youll
nd a line that intercepts the x = 0 axis at the value f (0) = a. The slope of that line
will be b.
Mathematically, a function f = f (x) is [linear??] if f (Ax) = Af (x) and if f (x
1
+
x
2
) = f (x
1
) + f (x
2
). As an exercise, you should reconcile that denition with the
graphical interpretation.
The square function:
f (t) =t
2
(2.6)
Sinusoid:
f (t) = sin(t) (2.7)
Sinusoid with xed phase lag:
f (t) = sin(t ) (2.8)
where is xed. (You can tell that is considered to be xed because it is not listed
as an independent variable in parentheses on the left-hand side of the equation.)
Sinusoid with variable phase lag, this lag being a second independent variable
f (t, ) = sin(t ) (2.9)
Here, appears in the list of independent variables, so it is considered to be variable.
Thus, the above equation illustrates a function of more than one variable. Another
good example of a function of several variables is T(x, y, z, t), the temperature eld as
a function of position (denoted by x, y, and z coordinates) and time t.
A note on units. Some functions, such sin(x) and exp(x), only make sense if the inde-
pendent variable, x, is [nondimensional??]. For example, consider a sinusiodal function
meant to represent tidal elevation, peaking twice per day. It might have the form sin(t),
where (the frequency) must have units of inverse-time in order to cancel the units of t (the
time of day). To see that this makes sense, imagine that the function were instead sin(t),
with t being the time (and no frequency in the formula) . . . imagine the confusion that would
arise because t could be measured in seconds or hours or millenia. (Hint: one of the rst
things you should do, when confronted with an unfamiliar equation, is to check whether the
units of all components of the equation make sense; see Section 2.6 below.)
22 CHAPTER 2. MATHEMATICAL TOOLS FOR PHYSICAL OCEANOGRAPHY
2.3. Numbers
2.3.1. Integers, rational numbers, and irrational numbers
It is common in mathematics to use words that have a different meaning in everyday
conversation. This section describes a prime example: complex numbers are not really all
that complex (in the sense of being complicated).
Before explaining complex numbers, it would make sense to explain other types of
numbers.
The simplest numbers are the numbers used to count things, e.g. 0, 1, 2, and so forth.
In some sense, these are the most basic numbers of all. But an abstraction makes these
numbers much more useful. That abstraction is negative numbers. Now, you can never have,
say, 5 pennies in your pocket. So in this sense, negative numbers are a bit odd, arent
they? But you could have zero dollars to your name, and owe somebody 5 pennies, couldnt
you? So, in that sense, your net worth would be 5 pennies. Thus, it is useful to add things
such as 1, 2, 3, . . . to the list of counting numbers 0, 1, 2, . . . . This whole list we call
the integers. There are an innite number of integers because if you list some, I can always
add 1 to the highest number and then put that in the list also, but then you can top me by
inserting another number in the same way.
The next most complicated type of number is the so-called rational number. These are
made of the ratio of integers. Examples are 2/3, 0.9 (i.e. 9/10), etc There are an innite
number of rational numbers
4
. Indeed, there are an innite number of rationals between
any two integers, n and n +1 say, and that might make you guess that there are more
rationals than integers . . . but your intuition would be wrong, since it is possible to count
the rationals
5
, so there cannot be more of one type than the other.
The next most complicated type of number is the so-called irrational number. An
irrational number cannot be expressed as a fraction of integers. An example is , which can
be approximated with various rational numbers (e.g. 31/10 is a reasonable approximation,
while 314/100 is a better approximation), but which cannot be exactly represented as a
rational number. Another commonly-used real number is e, the base of natural logarithms.
There are innitely many irrational numbers
6
.
Instead of referring to positive integers, and integers, and rational numbers, and real
numbers, lets just call this whole list as the real number system. If you can come to grips
with why these numbers are called real, youre a better person than I am. Its just a name
dont read too much into it. By the same token, lets now move on to so-called imaginary
numbers. They arent any more imaginary (in the common use of that word) than real
numbers are real, so dont get stuck on the word.
2.3.2. Imaginary Numbers
Consider the number 4. It has a familiar square-root, the value 2. But doesnt 2 also,
when multiplied by itself, yield 4? Yes, and therefore the two square roots of 4 are 2 and
2. Indeed, all numbers have two square roots, one positive and one negative.
Next consider the number 16. It can be written 44. From simple arithmetic, we can
write

16 =

(44) =

4. More generally,

(ab) =

b.
4
Consider any two distinct rational numbers a/b and c/d. Their average, (ad +bc)/(2(bd)), is also a rational
number, and it cannot equal either a/b or c/d. Thus we have demonstrated that at least one rational exists between
any given rationals, and that means there must be an innite number of rationals.
5
List all the rationals as
1/1 1/2 1/3 1/4 . . .
2/1 2/2 2/3 2/4 . . .
3/1 3/2 3/3 3/4 . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
and start counting, starting at the top-left corner and then sweeping along diagonals, e.g. the rst element in the
count is 1/1, the second is 2/1, the third is 1/2, and so on. No rationals are skipped, and the counting integers will
never run out, and so this process maps the rationals onto the integers in a one-to-one way.
6
Let a be any given irrational number. Consider its double, 2a. If this is rational, it may be written as 2a = b/c,
where b and c are integers. This yields a = b/(2c) and, but 2c is an integer, so this means that a must be rational,
in contradiction to the original assumption. Thus, 2a must be irrational. This reasoning applies with any integer
multiplier of a, so there must be an innite number of irrationals.
2.3. NUMBERS 23
Now, things get interesting. What is

4? Well, from the above rule, we can write this


as

4, which we may rewrite as 2

1. There is nothing special about 4, of


course. What we have discovered is that the square root of any negative number is equal to
the square root of its positive sister, times the square root of 1. And so we have reduced
the problem of understanding negative square roots to the simpler problem of understanding
the square root of 1, so, weve made good progress on this problem of roots of negative
numbers.
When in doubt, give something a name. We use the symbol i for this root, by the
denition
7
i

1 (2.10)
Note that we are taking just the positive root in the above denition. Obviously if i
2
=1
then (i)
2
=1 also, since the negative in the left-hand side cancels when the squaring is
done. Thus, i is also a square root of 1. However, by convention, we take the positive
root of 1 as our denition of i (in equation (2.10)).
Lets not worry for the moment about what i means; that will become clearer as it is
used
8
. In particular, some useful properties of i are as follows.
Straightforward application of the denition of i yields these rules:
i
2
=1
i
3
= i i
2
=i
i
4
= i i
3
= i (i) = 1
(2.11)
and so on
9
.
The exponential of i is very important, and it may be calculated as follows, from
power-series expansions
10
. We consider the exponent of i, where is some number
(which you may like to think of as an angle).
e
i
= 1+(i) +
(i)
2
2!
+
(i)
3
3!
+
(i)
4
4!
+
(i)
5
5!
+
(i)
6
6!
. . .
= 1+i
()
2
2!
i

3
3!
+

4
4!
+i

5
5!

6
6!
. . .
= [1

2
2!
+

4
4!

6
6!
. . . ] + i [

3
3!
+

5
5!
. . . ]
= cos +i sin
(2.12)
where we have used the above list of powers of i, done some grouping of terms, and
then recognized the power series for sine and cosine. This rule, that
e
i
= cos +i sin, (2.13)
is known as Eulers Rule and it is very important. Go ahead and forget the proof, Eulers rule
but remember this!
Time-dependent motions are often broken down into sinusoidal components (by so-
called Fourier Analysis), and many motions are sinusoidal to begin with, and for
that reason there is great interest in sinusoidal signals. Eulers Rule is very useful
for such signals. Imagine a cork bobbing on the ocean in a simple wave eld, such
that the corks vertical displacement is of the form =
0
cos(t), measured in
meters from mean sea-level, say. So, what is the velocity? Well, you may remember
7
Confusingly, some engineering texts use the symbol j for what were calling i here. The reason is that the
symbol i has long been taken to stand for electric current, so j is used instead.
8
This is like wondering what negative integers mean . . . they mean something because they are useful. You
cant have negative cash in your pocket, but you can owe somebody money and thats the same thing.
9
You might benet from the practice youd get by extending this list to a few more terms.
10
Here Im assuming that you know, or are willing to look up, the expansions for exp(x), sin(x), and cos(x).
Also, Im assuming you know that e.g. 4! is 1234 by denition.
24 CHAPTER 2. MATHEMATICAL TOOLS FOR PHYSICAL OCEANOGRAPHY
Figure 2.2: Denition sketch of Eulers Rule. Suppose the sloped line is of unit length and
that the angle of the line to the horizontal (x) axis is . Then the projection onto the x-axis is
x = cos, while the projection onto the y-axis is y = sin. We say that e
i
= cos +i sin,
so we have e
i
= x +iy.
that the derivative of cosine is the negative of sine, so that we have w = d/dt as
w =
0
sin(t). Thats ne. But you may also note that the real part of
0
e
it
is

0
cos(t), so that we can write the velocity as the real part of d(e
it
)/dt, i.e. the
real part of ie
it
, i.e., using Eulers rule, the sin term as written above. Thus, you
can forget your rules of differentiating sines and cosines; these rules are bound up in
the rule for differentiating exponentials. (You may complain that Ive not discussed
differentiation yet! Well, dont shoot me, just keep reading.)
2.3.3. Complex Numbers
Youve just seen so-called complex numbers; they are nothing more than the sum of a
real number and an imaginary number, as weve just used in Eulers Rule.
It is useful to catalogue some of the properties of complex numbers. First, lets consider
addition. What is the sum of one complex number, a+ib say, and another, c +id say? Well,
by straightforward rules of arithmetic we have:
(a+ib) +(c +id) = a+ib+c +id
= (a+c) +i (b+d)
(2.14)
which gives us the simple rule: just add the real and imaginary components separately.
Subtraction is similar.
Multiplication is a bit messier:
(a+ib) (c +id) = a(c +id) +ib(c +id)
= ac +iad +ibc +i
2
bd
= ac +iad +ibc bd
= (ac bd) +i(ad +bc)
(2.15)
with division being a bit messier still. However, many folks do multiplication and division
by rst converting to the exponential form with Eulers Rule. Thus, the rst step in multi-
plying is to write a+ib as re
i
where r =

a
2
+b
2
and = arctan(y/x) as determined by
straightforward inversion of the Euler Rule
11
. If we similarly convert c +id to, say, Re
i
,
then we can perform the multiplication trivially:
re
i
Re
i
= rRe
i(+)
(2.16)
by the simple rules of arithmetic. Thus, the length of the product of two complex numbers
is the product of their lengths; the angle is the sum of their angles
12
.
11
Exercise: verify this.
12
Draw a few more lines on Figure 2.2 if this isnt crystal clear.
2.4. CALCULUS 25
f(x)
x
x
0
x
1
f
1
f
0
Slope=
dx
d f
Figure 2.3: Denition sketch of derivative. As x(= x
1
x
0
) approaches 0, the ratio f /x
approaches the derivative d f /dx.
2.4. Calculus
The basic operations of calculus are differentiation and integration. Understanding the
meaning of differentiation is crucial both to understanding integration and to understanding
differential equations, so well start with differentiation.
2.4.1. Ordinary differentiation
Suppose that you are walking from home toward the liquor store
13
. Further suppose
that you continuously measure your distance from your starting location. Lets denote that
distance by x, and the time by t. Being mortal, you can be in only in one place at one time,
so clearly x = x(t) is a function of time t.
How could you determine the speed at which you are walking? Well, you could just
measure the time it takes to walk the whole way call that t
total
and divide this into the
total distance x
total
, giving the velocity estimate
v
x
total
t
total
(2.17)
but this is crude. Its a measure of your gross speed over the whole trip, but not of your
instantaneous speed.
A better measure of your speed at some particular time t would be calculated by measur-
ing the distance you move over some small time. Call the distance x and the time interval
t. (Note: here were introducing the notation that writing the Greek letter before a letter
representing some quantity stands for a change in that quantity.) Then an estimate of the
velocity is
v
x
t
(2.18)
As you make t smaller and smaller, this velocity estimate becomes a better approxima-
tion to the true velocity. Youll nd that your estimates of the velocity will approach a xed
value as you make t smaller and smaller. This limiting xed value is the true velocity.
This process of calculating ratios like that in (2.18), taking the limit for ever smaller t,
is so useful that it has a special name: differentiation. It holds a central place in mathematics
and Physics. We denote
14
the derivative of f with respect to t by the symbol d f /dt and
dene it by
d f
dt
= lim
t0
x
t
(2.19)
13
Or to school, even.
14
Note: The symbol d f /dt is not a ratio, but rather a shorthand way of writing the messy right hand side of
(2.19). Sometimes we write the derivative as f
/
, which is somewhat less confusing.
26 CHAPTER 2. MATHEMATICAL TOOLS FOR PHYSICAL OCEANOGRAPHY
where the lim part of the equation indicates that we are to take the value for the fraction
which you get in the limit of smaller and smaller values of t.
The differentiation principle, which Ive introduced as a way to work with a set of
numbers, is extended to work for idealized functions. Long lists of these have been tabulated,
and any rst mathematics or physics student can rattle them off endlessly. Lets work just
one example. Consider the linear function
f (t) = a+bt (2.20)
where a and b are constant numbers. And lets suppose we want to evaluate the derivative
d f /dt at some arbitrary value of which well just call t for now. Now, the value of f is
calculated as the difference between the value at t and the value at t +t:
f = f (t +t) f (t) (2.21)
which we can substitute into (2.20) to get
f = [a+b(t +t)] [a+bt] (2.22)
which can be simplied to
f = a+bt +bt abt
= bt +bt bt
= bt
(2.23)
Substituting this into (2.18) we get
d f
dt
=
f
t
(2.24)
=
bt
t
(2.25)
= b (2.26)
so that we see that the derivative of a linear function is just the slope of the line.
2.4.2. Partial differentiation
When something is a function of more than one independent variable, there is a problem
dening a derivative following the above reasoning. So we dene partial derivatives,
one for each independent variable. For example, if h = h(x, y) is a function of the two
independent variables x and y
15
, then the so-called partial derivatives are denoted by
h/x and h/y. Note the use of instead of d in this formula. Calculating partial
derivatives is easy. For example, to calculate h/x, simply consider the variation of h
with respect to x, holding y xed at some particular value of interest
16
. For example, if
h(x, y) = ax
2
+xy, with a being a constant, then h/x is equal to 2ax +y. Note that the
(xy)/x term is equal to y since y acts as a constant for differentiation with respect to x.
That is, more explicitly:
h
x
=

_
ax
2
+xy
_
x
(2.27)
which simplies to
h
x
=
(ax
2
)
x
+
(xy)
x
(2.28)
Since a is a constant for differentiation with respect to x, and since y is also a constant for
differentiation with respect to x, this becomes
h
x
= 2ax +y (2.29)
15
e.g., h(x, y) might be the depth of the water at some location(x,y)
16
Geometrical interpretation: suppose that h(x, y) is the water depth. Then h/x at say y = 0 is the slope of the
bottom of the ocean looking in the x direction, at the location y = 0.
2.5. DIFFERENTIAL EQUATIONS 27
2.4.3. Integration
Integration is the opposite of differentiation. If the derivative of some function f = f (x)
is say g = g(x) = d f /dx, then the integral of g with respect to x is f (x) plus an arbitrary
constant, i.e.
_
gdx = f +constant (2.30)
Here we have established the notation that
_
gdx means the integral of g = g(x) with respect
to x. The dx notation is an historical artifact; if you just read this as some item d
multiplied by another item x, then you need to take a quick peek at a calculus book.
Integration is more difcult than differentiation. While you can learn to differentiate
most functions of interest in a few weeks, you can keep learning integration tricks for the
rest of your life.
2.5. Differential equations
Dynamical systems are described by differential equationsthat is, equations with terms
involving derivatives. A good example is Newtons law, which relates the acceleration of an
object (which is the second derivative of the position of the object) to the force on object.
Example. Consider a mass on the end of a spring. Howmay we describe its dynamics?
Let us denote by x the distance the mass has moved from a location of rest (when
the spring is not stretched or compressed beyond its neutral, unloaded, state). Note
that in this problem x is the dependent variable (the solution), and t, time, is the
independent variable. Then the force on the spring is given by kx, where k, a constant
of proportionality with units mass/time
2
, depends on the stiffness of the spring. The
minus sign says that the force acts opposite to the direction in which the spring is pulled.
The acceleration of the mass is d
2
x/dt
2
, so that if the mass is m, then Newtons law
then says that
m
d
2
x
dt
2
=kx (2.31)
This is an ordinary differential equation that may be solved for x as a function of
time, x = x(t), provided that initial conditions are known. For example, if x = 0
at t = 0, and the velocity is dx/dt = 1 at t = 0, the solution valid for all time is
x =

(m/k)sin(

(k/m)t). You might want to substitute this back into the differential
equation to check if it is indeed a solution.
Formal solutions are known to only a handful of types of differential equations, but
approximate solutions can be found by computers running numerical models. Differential
equations are divided into two types: ODEs and PDEs. It is simpler to start with ODEs.
2.5.1. Ordinary differential equations
Denition An ordinary differential equation (ODE) is an equation for some function
u = u(x) which contains terms involving u and derivatives of u with respect to x, the
independent coordinate
17
.
ODEs appear in the study of dynamical systems because acceleration d
2
x/dt
2
is related
to force, and various forces are related to location x and velocity dx/dt. (Think of a damped
spring, such as those attached to doors: the restoring force is proportional to position, and
the damping is proportional to velocity.)
Perhaps the simplest non-trivial DE is
du
dt
= 1 (2.32)
17
The related eld of PDEs (partial differential equations, which contain partial derivatives, such as u/x and
u/y) is more complicated and extremely useful.
28 CHAPTER 2. MATHEMATICAL TOOLS FOR PHYSICAL OCEANOGRAPHY
with u(0) = 0 as initial condition. The initial condition (or IC, for short) refers to the
value of the function at time t equal to zero
18
. The initial condition is something akin to a
constant of integration in the topic of integral calculus it selects one particular solution,
from amongst an innite number of general solutions. The general solution of this DE, i.e.
the solution ignoring the initial condition, is u(t) =t +A, where A is a constant whose value
may be anything at all. For this to not contradict the IC, however, it is clear that A must be
zero
19
.
Notation. It is sometimes convenient to write du/dx as u
/
, with the prime denoting
derivative. Thus, the previous equation could also be written more compactly as u
/
= 1.
One problem with this notation, in Fluid Dynamics, is that the prime symbol is often
used to denote a deviation from a mean, e.g. in Section 4.12.1, and so Ill avoid the
prime notation for differentiation.
Finding analytical solutions of arbitrary ODEs is challenging, and approaching reasonable
skill for a limited class of problems requires years of study. For simple DEs, an appropriate
method is simply to guess a solution and check it by substitution into the DE
20
Some of the common trial solutions, say for f = f (t), are algebraic functions such as
t, t
2
, etc (more generally, t
n
), circular functions such as sint and cost, and exponential
functions such as logt and expt. In each case, somewhat more general trial solutions are
usual consideredthat is, solutions with extra constants, such as Asin(Bt) for examplebut
the above list is clearer without this extra clutter.
It may help you to think about some examples of differential equations for real-world
problems.
You are carrying coffee from the coffee joint in the Killam library, to your oh-so-fun
oceanography class in the LSC. Its beautiful outside one of those autumn mornings
when you can see for miles, when scattered morning clouds near the horizon have
been dipped in purple, reminding you of a gathering sunset. Its a crisp morning.
You realize that sweaters wont cut it for much longer, and that makes you enjoy the
morning even more, because you know whats coming. And whats coming is, in
a word, cold. In the winter you wont likely walk very far in the outdoors holding
coffee, because the rate of heat-ux to the atmosphere is proportional to the difference
in temperature of the coffee (T
c
) and the air (T
a
). Clearly, this heat loss makes the
coffee cool, since there is no source of heat
21
. Denoting the rate of change of the
coffee temperature dT
c
/dt, with t as time, and letting k measure the proportionality
constant in the heat-ux law, we see that the following must be true:
d(T
c
T
a
)
dt
=k(T
c
T
a
). (2.33)
This isnt a full statement of the problem, though, because we havent taken into
account the initial coffee temperature. Lets do that, by adding a statement of the
initial condition as
T
c
(0) = T
a
+T
0
. (2.34)
Thus T
0
is the initial heat difference between the coffee and the air.
To remove some of the clutter, we may create a new variable, called f , dened by
f = T
c
T
a
(2.35)
18
Actually, the phrase initial condition would be used even if the independent variable were x, or something
else, instead of time. Dont confuse yourself by thinking that the use of the word initial implies, necessarily, a
time-dependent problem.
19
Exercise: check this.
20
Youll recognize this trial solution technique if youve studied techniques of integration in your calculus
courses.
21
Radioactivity, inserted by the government into the coffee in an effort to cause mutants who will obey gun-
control laws, is an exception.
2.5. DIFFERENTIAL EQUATIONS 29
so that the DE becomes
d f
dt
=k f (2.36)
and the IC becomes
f (0) = T
0
(2.37)
A hockey puck is struck with initial velocity u
0
. It slides over a very large rink,
slowing due to friction. The friction force is proportional to the speed
22
. From
f = ma, we see that
du
dt
=ku (2.38)
where k is a friction coefcient with units of inverse-time.
A single cell of the species Cellus Exampilis is put onto a plate of nutrients. The
species reproduces asexually
23
, splitting once per hour. The number of new cells
produced in any hour is thus equal to the number of existing cells at the start of the
hour, so that
dC
dt
=C (2.39)
where C is the cell concentration.
Well, that was quite a lot of words in those examples. Now, I invite you to take a moment
and glance back at the equations in the above list, ignoring the words. You should notice that
each of these examples yields the same form of differential equation. That means that they
are all actually the same problem, in the Mathematical sense, despite the differing words
used to describe them. And that means that solving one problem solves all of the problems.
To a great extent, Mathematics and Physics are tied up in the search for analogous problems.
To avoid annoying coffee lovers (who want the word coffee) or hockey players (puck)
or janitors (cells), lets discuss all of these problems together, in stripped-down words.
The problem is to solve
d f
dt
+ f = 0 (2.40)
with initial condition f (0) = 10. To solve (2.40), lets try substituting some functions
f = f (t) whose derivatives d f /dt we know. If any particular trial function ts the equation,
then we have a solution to the DE.
First, try f (t) =t. Our test will be to calculate the derivative, substitute into our DE, and
see if the result can hold true for all time t. Note that d f /dt = 1 in this case, so that, upon
substitution in equation 2.40, we are left with
1+t
?
= 0 (2.41)
Well, thats no good it certainly doesnt hold for all time! But what about f (t) = t
n
?
Substitution into (2.40) yields
nt
n1
+t
n
?
= 0 (2.42)
which, again, is not true for all time.
Lets move on to the circular functions. First, consider the trial solution f (t) = sint. The
derivative is cost, so that substitution into (2.40) yields
cost +sint
?
= 0 (2.43)
and this is clearly not true, for all time
24
. The trial function f = cost has a similar problem.
Lets move on to the exponential functions, starting with f = lnt. This has derivative 1/t,
so that substitution into (2.40) yields
t +lnt
?
= 0 (2.44)
22
Does this make sense? Consider the case of zero speed, for example.
23
Oh what fun.
24
As an exercise, see if this is ever true!
30 CHAPTER 2. MATHEMATICAL TOOLS FOR PHYSICAL OCEANOGRAPHY
which is no good.
Next, try f = expt. It has derivative expt, so that substitution into (2.40) yields
expt +expt
?
= 0 (2.45)
which is no good. But, we seem to be getting pretty close, dont we? We just need to get a
negative sign in there, somehow! Lets try f =expt instead. This yields
expt expt
?
= 0 (2.46)
which is no good either. In fact, any multiplicative constant (instead of the 1 constant here)
will be of no help to us. However, lets try exp(t) instead. This has derivative exp(t),
so that our trial DE becomes
exp(t) +exp(t)
?
= 0 (2.47)
Well, now were onto something, since this ts! This is true, for all values of t. Therefore
we have a solution.
But is this the only solution, or are their others that would work also? To explore further,
lets generalize a bit, and consider the trial solution f = Aexp(Bt). This has derivative
ABexp(Bt), so that substitution into (2.40) yields
ABexp(Bt) +Aexp(Bt)
?
= 0 (2.48)
This is true for any value of A whatsoever, but its only true for one particular B value,
namely the value B =1 that weve already seen. From this, we conclude that the general
solution to (2.40) must be f (t) = Aexp(t), where A can take on any value. The next
step is to nd the particular value of A that makes our general solution satises the initial
condition. Substitution yields f (0) = A, and (2.40) requires that f (0) = 10, so clearly we
must have A = 10. And thus, at last, we have the full solution to our DE, equation (2.40):
f (t) = 10exp(t). Easy, eh wot?
2.5.2. Linear ODE
Denition A linear ODE is an ODE in which u(x) and its derivatives appear only once
per term.
Example u
//
+2x
2
u
/
+e
x
u = sin(x) is linear, but u
//
u
/
= e
x
u +3x
2
u
/
is not (since the
LHS violates the condition).
Solution In general, linear ODEs are easier to solve than nonlinear ODEs.
2.5.3. Constant coefcient linear ODE
Denition A constant coefcient linear ODE is a linear ODE in which the coefcients
multiplying u, u
/
, u
//
, etc, are constants that is, not functions of x.
Example au
//
+bu
/
+cu = 0 is a linear, constant coefcient, ODE.
Solution The trick to solving linear, constant coefcient ODEs is to assume the answer is
an exponential i.e., u =e
x
, where is a constant which we will subsequently determine
25
.
For exponential functions of this form, the derivative u
/
=
du
dx
is e
x
, which (and this is the
beauty of the trick) is just u. Similarly, d
n
u/dx
n
is
n
u. (Exercise: convince yourself this
is true, before proceeding further.) Thus, all derivate terms turn into terms like u times
raised to some power, so that the differential equation turns into an algebraic equation. This
is a remarkable simplication.
For example, substituting u = e
x
into
au
//
+bu
/
+cu = 0 (2.49)
25
Astute readers will realize that I chose the sample DE, equation (2.40), to lay the groundwork for the present
topic.
2.6. DIMENSIONAL ANALYSIS 31
yields
a
2
u+bu+cu = 0 (2.50)
which, cancelling the u term by dividing through both sides by u, yields
a
2
+b +c = 0, (2.51)
which is an algebraic equation for in terms of a, b, and c. Youll recall that there are
two solutions to a quadratic like this, which well denote

=(b
_
(b
2
4ac))/2a and

+
=(b+
_
(b
2
4ac))/2a.
Thus the solution is
u = Aexp(
+
x) +Bexp(

x) (2.52)
where A and B are constants whose values are determined by the initial conditions. For
example, if u were equal to 0 at x = 0, then the sum of A and B must equal 0 (since at x = 0
each of the two exponential functions equals 1. If in addition du/dt were equal to 1 at x = 0,
then we would know that A
+
+B

(the derivative of 2.52, evaluated at x = 0) must equal


1. These two constraints at x = 0 are enough to determine the values of the parameters A
and B that is, they provide two simultaneous equations in two unknowns.
Note: If the equation has multiple roots, that is, if b
2
= 4ac, then
+
=

; how then to
get two solutions? The trick is to form a second solution as x times the exponential term.
Thus, the missing term is xexp(
+
x), and u = (A+Bx)exp(
+
x) is the complete solution.
If the multiple root is 3rd order, then the three terms are e
x
, xe
x
, and x
2
e
x
.
The following formula, known as Eulers formula, is useful in translating exponential
functions of complex argument
26
into sines and cosines:
e
it
= cost +i sint (2.53)
This and other aspects of linear, constant-coefcient, differential equations is very ably
discussed in any number of texts, including Schaum outlines.
Exercise Solve u
//
+2u
/
+5u = 0, and rewrite the answer in terms of e
x
, sin2x, and
cos2x, with initial conditions: u(0) = 1, u
/
(0) = 0.
2.5.4. Scaling Differential Equations
The process of [scaling??] differential equations refers to a process of estimating the
size of the terms in a differential equation. If one term is clearly much smaller than another,
it can often be ignored. For example, if you could show that the Asin(x) term in the equation
(1+Asin(x))
dx
dt
= . . . (2.54)
is small compared to 1, then the simpler equation
dx
dt
= . . . (2.55)
results. In this case, this simplication would be reasonable so long as A is much less than 1,
since sin(x) is never greater than 1.
2.6. Dimensional Analysis
Every rods a gain
And theres victory in every quarter mile
The eld beyond the plow c _1980 Stan Rogers
God is a concept
by which we measure our pain
God c _1970 John Lennon
26
A complex number is one made up of the sum of a real part and an imaginary part. A real number is just a
normal number, and an imaginary number is a real number multiplied by the square root of 1.
32 CHAPTER 2. MATHEMATICAL TOOLS FOR PHYSICAL OCEANOGRAPHY
In the lectures for the class, you will notice that I pay close attention to the units of
physical quantities. This is intentional, since dimensional analysis is a very handy and
powerful tool. Use of this tool is an art it becomes easier with experience, and some people
do it more naturally than others.
There are 3 ways to use dimensions:
1. As a nal check of a formula. The formula must be dimensionally consistent, or
else a mistake has been made. For example, a diffusive time-scale may be given by
= 4d
2
/, with d a lengthscale (units m) and a diffusivity (units m
2
/s). Clearly,
from these units, has units of s. Now, ask what would happen if I wrote the formula
as = 4d/. In a ash, you can see that the RHS does not have units of time, and
that proves that the formula must be in error.
2. As an order-of-magnitude guess at a result. For example, how far will heat diffuse
in 1 second in a medium of diffusivity = 10
6
m
2
/s? We expect that the distance
increases with time, and note the units of t are m
2
, so we guess the distance as

(t) = 10
3
m.
3. As a signal about dimensionless ratios or numbers. These occur two ways
From nondimensionalizing or scale analysis of the relevant equations. (This
is a technique where we estimate the sizes of the various terms in the equations).
Well do this later, to come up with, for example, the Reynolds Number.
As a way to reduce the number of parameters we need to consider in working
with a situation. There is a formal theorem, called the Buckingham Pi-Theorem
(or theorem) , to help us. Given n variables describing a physical situation, and
among these variables there are k different physical dimensions (mass, length,
time, temperature, etc), then there are only nk dimensionless combinations to
consider. Thus, it tells us how many dimensionless parameters to look for in a
problem.
2.7. Vectors
A [vector??] in 3-space can be thought of as a description of a directed line segment
that is, a description of the orientation of the segment and of its length. In 3-space a
[vector??] has 3 components. Often vectors are described in Cartesian space, a space
whose axes are at right angles. Vectors are normally written with an arrow over them, or
in boldface (e.g., a velocity [vector??] might be denoted u or x), and the three Cartesian
components are normally written with the same symbol without the decoration, and with
subscripts 1,2,3 (e.g., u might be written (u
1
, u
2
, u
3
). Often these velocity components are
denoted ([u??], [v??], [w??]).
Two special cases of interest are
Position vectors, usually denotedx, with components written either as (x
1
, x
2
, x
3
) or
(x, y, z).
Velocity vectors, usually denotedu, with components written either as (u
1
, u
2
, u
3
) or
(u, v, w). (Note that u is the x-component, while u is the whole [vector??]. Confusing
eh?)
The length of a [vector??]

A, in analogy to the length of a directed line segment is
denoted A and equals (A
2
1
+A
2
2
+A
2
3
)
1/2
. (Length is sometimes called norm.)
A good example of [vector??] length is that the speed (independent of direction) is the
length of the velocity (dependent on direction). You can measure the speed of a car by how
fast the wheels turn, but you have to look at the car in 3-space to get the velocity.
2.8. VECTOR DOT PRODUCT 33
2.8. Vector dot product
The dot product, or inner between two vectors

A and

B is written

A

B and dened
as A
1
B
1
+A
2
B
2
+A
3
B
3
. It is equal to the product of the lengths of the two vectors, times
the cosine of the angle between them. Since the cosine of zero is 1, and the cosine of 90

is 0, the dot product between two vectors is zero if they are at right angles, and equal to
the product of the two lengths if they are parallel. Note that it is a number (properly, a
[scalar??]), not a vector.
Application. the dot product of the vector representing the force on an object, and the
velocity vector of the object, is proportional to the rate of energy being poured into the
object by the force. If they are at right angles (e.g., the gravity force acting on a skater
gliding over very smooth ice) then there is no energy exchange. If they are parallel (e.g.,
gravity force acting on a falling bungee jumper) then there is energy exchange.
2.9. Vector cross product
The cross product (

C say) of two 3-space vectors



A and

B is another vector, dened
by

C = (A
2
B
3
A
3
B
2
, A
1
B
3
A
3
B
2
, A
1
B
2
+A
2
B
1
). It is at right angles to both

A and

B,
obeying the [right-hand-rule??]. Its length is dened to be the product of the lengths of
the two vectors, times the sine of the angle between them. Thus, if

A and

B are parallel, the
cross product is a zero length vector, while if they are at right angles, the cross product is a
vector of length equal to the product of the lengths of

A and

B.
Application. the Coriolis term in the equations of motion is proportional to the cross
product of the vector representing the spin of the earth and the vector representing water
velocity.
2.10. Curl of a vector eld
The [curl??] of a vector eld
27

A, denoted curl

A or

A, is dened by [nabla-
times??]

A = (A
3
/y A
2
/z, A
1
/z A
3
/x, A
2
/x A
1
/y). Here [nabla??]
is a derivative operator.
Physical meaning: The curl of the velocity eld is the [vorticity??]. You must not think
of curl as meaning that motions are circular; it is possible to have circular ow elds with
zero curl. The diagnostic test for nonzero curl is whether a cork put into the water would
spin. Thus, for example, a [shear??] ow (as in a deck of cards sitting on a table, being
pushed over into a long line of cards) has curl, even though there is no curved motion.
2.11. Divergence of a vector eld
The [divergence??] of a [vector??], denoted [nabla-dot??]

A, is dened by

A =
A
1
x
+
A
2
y
+
A
3
z
(2.56)
It is a [scalar??], not a vector in fact, it is a [scalar??] function of a [vector??] variable.
The divergence operator is very important in uid mechanics. Of particular importance
is the fact that the divergence of a velocity eld measures the ow outward from a point
of interest. Since water is incompressible in many applications, its divergence is normally
identically zero.
The divergence operator is also used in biological oceanography, in which the velocity
eld is taken as the velocity of biological particles, e.g. larvae. It is of great interest to know
whether such particles are retained in various ow regimes.
Application: Imagine a room with windows and doors open. If wind enters one window
and leaves another, there will be zero divergence. But if wind leaves through one window
27
A vector eld is a function of space whose value is a vector. Example: the velocity eld as a function of x, y,
z, and t (here time t is being thought of as part of space).
34 CHAPTER 2. MATHEMATICAL TOOLS FOR PHYSICAL OCEANOGRAPHY
(because of a vacuum cleaner, perhaps), without having a compensating inow anywhere,
there will be positive divergence.
Most ocean ows are non-divergent (except acoustic motions). Many are also [2D??].
The combination permits the ow to be described by a [streamfunction??], whose (cross-
) derivatives give the velocity components. For example, consider ow in (x, y) with
components ([u??],[v??]). If such ow is non-divergent, then we have
u/x +v/y = 0
and so if we invent a function = (x, y) satisfying
u =

y
, v =

x
(2.57)
then we may describe this ow by means of the single function = (x, y), instead of with
the two functions u = u(x, y) and v = v(x, y). For example, we could contour , and that
would fully describe the ow in a single graph.
2.12. Gradient of a scalar
The [gradient??] of a [scalar??] eld a = a(x, y, z) is a [vector??] pointing in the
direction of greatest increase of a. It is dened by
a =
_
a
x
,
a
y
,
a
z
_
(2.58)
In the equations of motion, the gradient of the pressure eld, p, appears as a force (the
pressure-gradient force, not surprisingly).
Application: Imagine a topographic map. The gradient is a vector which is perpendicular
to the contour lines, and is largest where the slope is greatest.
CHAPTER 3
Review of Physics
O
NE OF THE THINGS YOU MAY FIND FRUSTRATING in the presentation of physics
is a tendency to concentrate on the general case and not particular cases. There
are two reasons for this.
The rst reason is that it is a very powerful approach in a mathematical science. The
mathematics, by itself, suggests cases that should be considered. For example, an equation
might have two terms (corresponding to two forces in a physical circumstance), one term
being proportional to some number , the other being proportional to 1/. (The symbol is
the lower-case Greek letter epsilon.) From this, and thinking entirely in the mathematical
realm, forgetting entirely about the physical realm, it is clear that two limits are of interest.
In the limit of large , the second term becomes neglible and we can therefore write a new
equation that simply skips the small term. In the limit of small , the other term can be
neglected
1
. Often the full equations are difcult to solve but these reduced equations are
easier to solve.
The second reason is that generalizing kills two, three, or a gazillion, birds with one
stone. Rather than consider the ow in a natural-gas pipeline (to take a local Nova Scotian
example), a physicist would prefer to think of ows in cylinders in general. That way, the
results might apply to not just pipelines, but also to the water in your pipes, the blood in
your veins, the sap in the tree in the back yard, and so on. Modied slightly, the theory
would also shed light on trafc ow.
3.1. Momentum: Newtons Second Law
Many people couldnt tell you what Newtons rst and third laws are, but almost anybody
can tell you the [newtons-second-law??] is is
f = ma (3.1)
This tells how a body of mass m will be accelerated (a) by a given force f . Actually, f
is the (vector) sum of all forces on the body. In (3.1), the acceleration a is dened as the
rate of change (time-derivative) of velocity u, i.e. a = du/dt. Since the velocity is the
time-derivative of position (x, say), we may rewrite (3.1) as
f = m
du
dt
(3.2)
or
f = m
d
2
x
dt
2
(3.3)
In the above, the x-component has been written, but the same thing follows for y- and
z-components also.
The unit of force is the [Newton??], named for you-know-who, and (3.1) provides a
handy way to gure out what this corresponds to, in meters and so forth. The right-hand
side has units kg times m/s
2
, so we infer that 1N = 1kgm/s
2
.
In solid-body mechanics, (3.1) is quite straightforward, since the mass m is quite easy to
state, and so is the acceleration a. For example, imagine a ball rolling down a hill . . . m is
the mass of the ball, and a (in vector form) is the acceleration in three dimensions. For a
1
Later on, youll see an example, in terms of the so-called Reynolds number. When this number is small, the
mathematical term representing friction is large compared to another term relating to turbulence, and so the ow
is smooth and predictable. When this number is large, the friction term becomes neglible and the ow becomes
turbulent and difcult to predict in detail.
35
36 CHAPTER 3. REVIEW OF PHYSICS
Figure 3.1: A diver jumps off a bridge in Rome.
uid ow, however, things are more complicated. One approach is to imagine the uid as
being made up of a great number of tiny elements, each one responding to a particular set of
forces, and each obeying (3.1) individually. We might, in fact, imagine innitesmal elements
of uid, each with mass m say. Thats the rst step, one which leads us in Chapter 5 to
discuss force per unit mass, instead of force. The next step is to realize that the acceleration
must be dened element-by-element, tracing the elements as they move through space. This
leads us to derive the so-called Lagrangian acceleration in Chapter 5. But, in saying this Im
running close to giving away the movies plot in the trailer the week before . . . so stay tuned.
Exercise 3.1 Fate of a Clumsy Olympic Diver
a
A diver jumps from a springboard that is 3m above the surface of a swimming
pool. Its an ungainly sort of dive more a stumble, really, with no upward
motion whatever. (a) How long will it take the diver to reach the water? (b) What
will be her vertical speed when she hits the water? (c) How would these things
be changed if the diver ran off the end of the board, instead of falling off?
a
Answer in chapter 21.
3.2. ENERGY 37
Exercise 3.2 Fate of a Bridge Diver
a
As a New Years celebration, Italians sprinkle wine onto the water and dive in
(see Figure 3.1). (a) How long will it take before the bridge diver hits the water?
(b) How fast will he be going when he hits the water?
a
Answer in chapter 21.
Exercise 3.3 Automobile (F1) Engines
a
New rules prevent Formula-one engines from being turbo-charged. So, to get the
required horsepower, the engines have been modied. Car and Driver magazine,
Oct 1999, gives some data on these engines. The engine rotates at 18,000 RPM
and the pistons have a stroke of 46mm. The connecting rod and piston head weigh
15 ounces together. What is the maximum velocity of the pistons? What is the
maximum acceleration of the pistons? What force is exerted on the connecting
rod? If youre really keen, you might want to nd out the kind of metal used in
these parts, and its breaking strength, and then work out the overdesign factor
that is employed in these parts.
a
Answer in chapter 21.
3.2. Energy
3.2.1. Concepts
The terms work and energy are used in everyday life, but they also have precise
meanings in Physics. In fact, they share the same unit, the [Watt??], and they may be
interchanged in various ways . . . but for more on this, youll need to consult a Physics
textbook; here I just want to sketch some uid-mechanical nuances to these terms.
Work is the result of a force acting through a distance. Both those words, force and
distance are crucial. You do work if you push a car up a hill at a constant speed (increasing
its potential energy) or if you accelerate a car (increasing its kinetic energy) on a level
surface. But you do no work at all (except for chemical work in your muscles) by pushing
on a stationary car. (Maybe a better example is that you do no work by standing on the
ground, even though there is a force continuously applied.)
The energy of an object increases if you do work on it. If you climb stairs, you increase
your potential energy. (Strictly speaking, this is gravitational potential energy Im talking
about here. There are other forms of potential energy as well, e.g. the energy in a coiled
spring.) This could be released, as kinetic energy, if you fell down the stairs. Kinetic
energy refers to the energy in a moving object. Add this information to what you learned in
grammar school, and you should have a picture in your head of PE being related to height
and KE being related to speed. Lets quantify that a bit.
3.2.2. Solid-body application
Consider an object of mass m that is raised a height z. If the gravitional acceleration is
denoted [g??], then the balls potential energy (PE, henceforth) is increased by the product Pot. Energy of a solid body
of the force mg times the distance z. Thus, the PE is increased by the amount mgz. The
unit is the [Joule??], abbreviated J. (Conversion: 1J = 1Nm, or equivalently 1J=1kgm
2
/s
2
.)
For example, quarter-pounder hamburger
2
, with m = 0.2kg raised z = 0.5m from a
table to a mouth has its PE increased by 1J. Well, what does that mean, really, 1 [Joule??]?
One way to come to grips with the number is to ask what kinetic energy (KE, henceforth) Kin. Energy of a solid body
would be recovered, by dropping the burger back onto the table. During this path, the KE
will increase by 1J, and the KE is mv
2
/2, so that the nal speed will be v = (2KE/m)
1/2
,
i.e. about 4m/s. That speed is comparable to jogging speed
3
2
OK, veggieburger. And 1/4 pound is both disgustingly big and non-metric. I hate what I write sometimes.
Sheesh.
3
If you wish to experiment along these lines, be sure to drop the hot burger on your lap so you can sue
MacDonalds for 8 million dollars.
38 CHAPTER 3. REVIEW OF PHYSICS
3.2.3. Fluid application
In studies of uid ows, energies are often expressed as energy densities per unit mass.
Nominally, then, KE becomes u
2
/2, and PE translates similarly. But it makes no sense to
get into details until the forces in uids have been discussed, so we will delay discussion of
energetic issues until section 4.11.
Part III. Ocean Physics:
General Issues
39
CHAPTER 4
Seawater properties
B
EFORE DESCRIBING THE MOTION OF THE SEA, we must be able to describe its
physical state, or hydrographic properties. The main hydrographic variables are
[temperature??] and [salinity??]. These affect [density??], which in turn affects
[pressure??]. Since [pressure??] is a key force driving ocean ows, temperature and
salinity are intimately connected to ocean dynamics.
4.1. Temperature, Salinity, Density
Temperature T: units

C
Salinity S: roughly speaking, [salinity??] is the number of gram of salt in a kilogram
of salt water
1
. Expressed in this way, S should be unitless, right? Well, its not that
simple. The matter of whether S has units is (remarkably) unsettled; as of 2003,
some journals accept (or demand) units, while others refuse units. History comes into
this. Before the 1980s the units for salinity were ppt [parts per thousand], sometimes
written as

. (Remember, it is grams salt per kilogram of seawater, so thats parts


per thousand.) In the 1980s, some changes were made to the salinity denition, and
it became popular to distinguish values in this newly-dened quantity with the unit
[PSU??], an abbreviation for [practical-salinity-unit??]. Some journals, however,
state that salinity is to have no unit. Thus, in reading the literature you might see
salinities reported as S = 35

, S = 35 PPT, S = 35 PSU, or S = 35. As of 2008,


a new wrinkle has arisen, with SCOR/IAPSO Working group 127 discussing a new
denition of salinity . . . but this has yet to make its way into common use. Students
who want to follow up on the continuing saga of salinity measurement might start
with Seitz et al. [2008], but there is no need to get into such details within the context
of this course.
Density, denoted [rho??], is measured in the unit kg/m
3
.
Pressure, denoted p or P, is measured
2
in the unit [Pascal??], abbreviated [Pa??].
Differences in [density??] provide one important force driving motion in the sea. For
a discrete object, motion depends on mass and shape. For a uid, these parameters are
replaced by [density??], denoted . Density (or factors determining density such as T or
S) can often be used to identify packets of water, or water masses. These can sometimes
be tracked over long distances, using the notion that (say) S is conserved following a uid
[parcel??]. Ability to measure and/or predict the distribution of T, S, is important for
biology, chemistry, meteorology, climatology.
4.2. Hydrographic Sampling
Figure 4.1 shows an indication of sampling in the [WOCE??] (WOCE) program. The
spacing along a given transect is normally about 30 km, but the spacing between transects is
vastly greater. The transects are widely separated in time. As this course progresses, you
should return to this gure, and ponder what sort of phenomena can be measured with such
ship-based sampling . . . and what cannot.
1
There are more precise ways of dening salinity, and a chemical oceanographer would be happy to tell you
about them. However, the denition given, which is the historical denition, is also the easiest to visualize.
2
You may also run across other units, e.g. [decibar??], abbreviated [dbar??], in oceanographic work and
[millibar??], abbreviated [mbar??], in atmospheric work. Neither is metric, neither is recommended, neither
works naturally in our equations. But both are appear in the literature so youve got to be aware of them. By
denition, 1[dbar??] = 10
4
[Pa??]. This is a convenient unit because 1 [dbar??] corresponds to 1m of water; thus
youll often see [dbar??] on the y axes of oceanographic [prole??]s, especially older ones.
41
42 CHAPTER 4. SEAWATER PROPERTIES
Figure 4.1: Map showing high-quality hydrographic sampling done in WOCE (world ocean
circulation experiment). Each line of dots results from a ship making a transect; station
spacing along the transect is tight, while spacing between transects is much larger. To
compare the sampling of atmospheric data (e.g. pressure, which is typically recorded in
any moderately sized city), compare the inter-transect spacing to what you know to be the
inter-city spacing in, say, North America.
4.3. Temperature distribution in the ocean
Some useful facts
3
:
Global mean ocean temperature 3.5

C
75% of the ocean is between 0

and 6

C
50% of the ocean is between 1.3

and 3.8

C
Sea-surface temperature (often written SST) varies from < 12

C near Antarctica
to > 28

C at the Equator
4
. While surface water has large temperature variations with
geography, the bulk of the water column is remarkably isothermal. Figure 4.2 shows the
annual average of sea-surface temperature
5
.
Figure 4.3 shows that the surface temperature really only applies to a thin skin in the
upper ocean. Note the rapid fall to very cold water below 100 m in this example, and recall
that the average temperature of the deep ocean is about 3

C. Another way to see this is in a


cross-section, e.g. Figure 4.4.
4.3.1. Explanation for depth variation T(z) in temperate seas
Figure 4.5 illustrates the variation of temperature over depth, across a range of latitudes.
Roughly speaking, the pattern is as you might expect: the ocean is generally warmer nearer
the equator, and generally warmer near the surface
6
.
Lets take rst the depth pattern. Apart from high latitudes (northerly and southerly
waters), the ocean cools with depth. That seems simple enough, since the ocean derives
its heat from the sun (whose radiation cannot penetrate far into the sea) and the air (which
interacts only with the ocean surface).
If the ocean behaved like a solid that was heated on a surface, we would see a simple
pattern of cooling with depth, perhaps something like the sketch in the left panel of Figure 4.6.
But that is nothing like what we see in the ocean, as youll recall from Figure 4.3.
3
Reference: Pickard p. 33-42; Duxbury and Duxbury p. 141-145.
4
The extreme cold near Antarctica is central to the thermohaline circulation in the ocean, and thus to climate.
5
Exercise: Download the data le from Collier and Durack [2006] and construct a similar diagram, but for
temperature at a depth of 500 m
6
There are exceptions, though, and these tell us a great deal about how the ocean functions. But we need to
begin with the coarse patterns; we saw the lumber before we sand it.
4.3. TEMPERATURE DISTRIBUTION IN THE OCEAN 43
library(ncdf)
jpeg("SST.jpg", width=1000, height=500)
#ftp://ftp.nodc.noaa.gov/pub/data.nodc/woa/WOA05nc/annual
file <- open.ncdf("/data/hydrographic/WOA05nc/t00an1.nc")
T <- get.var.ncdf(file,"t00an1")[,,1]
x <- get.var.ncdf(file, "lon")
y <- get.var.ncdf(file, "lat")
col <- hsv(h=seq(2/3, 0, length.out=16), 1, 1)
par(mar=rep(0.2, 4))
imagep(x, y, T, zlim=c(-2,32), ylim=c(-90, 90),
col=oceColorsJet, breaks=seq(-2, 30, 2), asp=1, axes=FALSE,
xlab="",ylab="")
Figure 4.2: Top: Annually-averaged surface temperature distribution, based on the 2005
version of the Levitus atlas [Collier and Durack, 2006]. Bottom: R code to create the
diagram, using a data le downloaded in 2007 from ftp://ftp.nodc.noaa.gov/
pub/data.nodc/woa/WOA05nc/annual/. Exercise: modify the R code to plot
temperature at other depths, to see how deep the spatial pattern extends.
44 CHAPTER 4. SEAWATER PROPERTIES
Figure 4.3: Variation of upper-ocean temperature through seasons. [Pickard Figure 9]
4.3. TEMPERATURE DISTRIBUTION IN THE OCEAN 45
library(ncdf)
jpeg("Tsection.jpg", width=1000, height=800)
file <- open.ncdf("/data/hydrographic/WOA05nc/T00an1.nc")
temperature <- get.var.ncdf(file,"t00an1")
depths <- get.var.ncdf(file,"depth")
lon.index <- 180
lon <- get.var.ncdf(file, "lon")[lon.index]
col <- hsv(h=seq(2/3, 0, length.out=17), 1, 1)
x <- seq(-89.5,89.5,by=1)
i <- which(depths < 2000)
y <- depths[i]
col <- hsv(h=seq(2/3, 0, length.out=17), 1, 1)
T.section <- temperature[lon.index,,i]
par(cex=1.5)
image(x, y, T.section, col=col,
xlim=c(-90, 90), ylim=rev(range(y)), zlim=c(-2,32),
xlab="Latitude", ylab="Depth [m]", axes=FALSE,
main=sprintf("Temperature section at %.1f E", lon))
contour(x, y, T.section, levels=seq(-2,32,2), add=TRUE, labcex=2)
box()
axis(2)
axis(1, at=seq(-90, 90, 45))
abline(h = seq(6000,0,by=-500), col="darkgray")
abline(v=seq(-90, 90, 45), col="darkgray")
Figure 4.4: Top: Temperature section based on Levitus (2005) atlas [Collier and Durack,
2006]. Bottom: R code to create diagram. Exercise: modify the code to explore other
longitudes, to see if the double-lobed pattern is the exception, or the rule.
46 CHAPTER 4. SEAWATER PROPERTIES
T
e
m
p
e
r
a
t
u
r
e
,

C
90S 45S Eq 45N 90N
0
5
1
0
1
5
2
0
2
5
0m
100m
500m
library(ncdf)
pdf("T-z-lat.pdf", width=8, height=6)
#ftp://ftp.nodc.noaa.gov/pub/data.nodc/woa/WOA05nc/annual
file <- open.ncdf("/data/hydrographic/WOA05nc/t00an1.nc")
temperature <- get.var.ncdf(file,"t00an1")
lon <- get.var.ncdf(file, "lon")
lat <- get.var.ncdf(file, "lat")
depth <- get.var.ncdf(file, "depth")
plot(lat, apply(temperature[,,1],2,mean,na.rm=TRUE),
xlab="",
ylab=expression(paste("Temperature, ", degree, "C")),
type=l, lwd=2,
xlim=c(-90,90),axes=FALSE,frame.plot=TRUE)
axis(1,labels=c("90S","45S","Eq","45N","90N"),
at=seq(-90,90,45))
axis(2, at=seq(0, 30, 5))
d <- which(depth == 100)
lines(lat, apply(temperature[,,d],2,mean,na.rm=TRUE),
lwd=2, lty="dashed")
d <- which(depth == 500)
lines(lat, apply(temperature[,,d],2,mean,na.rm=TRUE),
lwd=2, lty="dotted")
abline(v=seq(-90, 90, 45), lwd=0.5, col="darkgray")
abline(h=seq(0, 30, 5), lwd=0.5, col="darkgray")
legend("topright", c("0m","100m","500m"),
lwd=rep(2,3),
lty=c("solid","dashed","dotted"),bg="white")
Figure 4.5: Top: Variation of annually-averaged in-situ temperature with latitude and depth,
drawn from data in the 2005 version of the Levitus atlas [Collier and Durack, 2006]. Bottom:
R code to create the diagram.
4.3. TEMPERATURE DISTRIBUTION IN THE OCEAN 47
Figure 4.6: Left: Schematic [prole??]s of temperature through time, if there were no
wind-mixing. Solid line is early spring, dashed line is later in summer, after solar heating.
Right: the result of surface mixing.
Figure 4.7: Schematic of depth variation in polar waters. Solid line is winter, dashed
summer.
The explanation for the observed pattern is simply that the ocean is not a solid. Its a uid,
and uids can move. That movemement redistributes heat. The main factor in explaining
the observed pattern is wind-induced mixing, which tends to homogenize a layer near the
ocean surface. A competition between the two effects, heating and mixing, gives proles
which look schematically like the right panel of Figure 4.6.
The upper waters are called the wind-mixed layer, while the waters with large dT/dz
are called the [thermocline??].
Of course, solar input and wind-mixing are not constant over the seasons. Summer solar
input is greater than winter solar input, and winter mixing is greater than summer mixing.
The result is that in summer there are 2 thermoclines: one generated by this summers
activity, and a deeper permanent one. In winter, the summer thermocline is erased by heat
loss to the atmosphere.
These features can be seen in Figure 4.3.
4.3.2. Explanation for depth variation T(z) in polar seas
Figure 4.7 shows the seasonal variation of T(z) in polar upper waters. The main difference
from temperate waters, of course, is that its a lot colder there. But note that, while there
is some summer warming in the top 100m or so, the main annual-average effect of the
atmosphere is to cool rather than warm the ocean. Thus, surface waters in polar regions are
cooler than deeper waters
7
.
The fact that the permanent thermocline is warm on top in temperate seas, but cold on
top in polar seas suggests that, on average, heat is mixed down in temperate areas, but cold
is mixed down in polar regions.
4.3.3. The Adiabatic Effect
In very deep water, T increases with depth. Figure 4.8 shows an example.
7
You might then think that the situation is unstable, but it turns out the surface waters are so fresh that they are
lighter than the deeper waters even though they are colder.
48 CHAPTER 4. SEAWATER PROPERTIES
Figure 4.8: Variation of [temperature??] (normally denoted T) and [potential-
temperature??] (normally denoted [theta??]) in deep waters. The difference between the
two is more marked below about 1000m depth. Note that in very deep waters the temperature
T increases with depth (due to compression of the water), while potential temperature
remains constant. Source: [Knauss, 1978, Fig. 1.4].
This variation in temperature results merely from the compressibility of seawater (in
a way Ill explain in a moment), and it is not important dynamically. Exchanging water
between two different depths (say between 8000m and 10000m) would not affect the shape
of T(z) and would result in no forces or instabilities.
For this reason, we often describe temperature variations in deep water with (units

C),
the [potential-temperature??], rather than with T. is the temperature that the [parcel??]
of water would have if it were transported to the surface in an [adiabatic??] way, i.e.
without being allowed to mix with the surrounding waters, or exchange heat with them,
along its trip upwards Computer algorithms for potential temperature and other such things
can be found at ftp://nemo.ucsd.edu/pub/user_software.
Note that is approximately constant in the deep waters in Figure 4.8, while T isnt.
We calculate based upon local values of T, S, etc. The rate of change of the difference
between T and with depth
(T )
z
(4.1)
can be calculated theoretically. It depends upon the specic heat C
p
, the absolute (Kelvin)
temperature, the coefcient of thermal expansion a, and the density. The result is that
(T )
z
(4.2)
is > 0.2

C/km at 20

C, whereas at 0

C, it ranges from 0.03

C/km at the surface to


0.1

C/km at 4 km depth (as in Figure 4.8)


4.4. Salinity distribution in the ocean
Figure!4.9 shows the global pattern of surface salinity. Comparison with Figure 4.10
shows that the pattern of S(y) is largely determined by the difference between evaporation
and precipitation. The latter Figure shows that subtropical/tropical E P is strongly positive
(because of strong evaporation) so the surface waters are salty. In higher latitudes and near
the Equator, this pattern is reversed.
4.4. SALINITY DISTRIBUTION IN THE OCEAN 49
library(ncdf)
jpeg("SSS.jpg", width=1000, height=500)
#ftp://ftp.nodc.noaa.gov/pub/data.nodc/woa/WOA05nc/annual
#file <- nc_open("/data/hydrographic/WOA05nc/S00an1.nc")
file <- open.ncdf("/data/hydrographic/WOA05nc/S00an1.nc")
S <- get.var.ncdf(file,"s00an1")[,,1]
x <- get.var.ncdf(file, "lon")
y <- get.var.ncdf(file, "lat")
col <- hsv(h=seq(2/3, 0, length.out=16), 1, 1)
par(mar=rep(0.2, 4))
imagep(x, y, S, zlim=c(30,38), ylim=c(-90, 90),
col=oceColorsJet, breaks=seq(30, 38, 1/3), asp=1, axes=FALSE,
xlab="",ylab="")
Figure 4.9: Top: Annually-averaged surface salinity distribution, based on the 2005 version
of the Levitus atlas [Collier and Durack, 2006]. Bottom: R code to create the diagram.
Figure 4.10: Relationship between zonally-averaged surface salinity and E P, the net
difference between evaporation and precipitation.
50 CHAPTER 4. SEAWATER PROPERTIES
Figure 4.11: The Mediterranean water shows up as a middepth peak in S(z). Contrast the
[prole??] in the Northeast Atlantic with prole in Northwestern Atlantic. Solid line: NW
Atlantic; dashed line NE Atlantic.
The average oceanic salinities in the Pacic are 34.62. In the Atlantic the average is
34.90. This small but signicant contrast reects the difference in deep water circulation
patterns in the two oceans.
The depth variation of S(z) gives us clues about where the water in various depths came
from. Figure 4.11 shows an example. In the Eastern North Atlantic, there is a strong
mid-depth salinity peak because of outow from the Mediterranean Sea. This signal can be
traced over most of the North Atlantic, but it decreases to the zero near North America.
4.5. Pressure distribution in the ocean
The largest pressure variation in the ocean is with depth (although the small horizontal
variations are very important in ocean dynamics). For many applications, the ocean may be
assumed to be in a [hydrostatic??] state, yielding the [hydrostatic-pressure??] equation
P(0) P(z) =
_
0
z
g(z) dz (4.3)
so that, with P(0) written as P
atmospheric
P(z) = P
atmospheric
+
_
0
z
g(z) dz (4.4)
Note that the z in the limit of the integral is a negative number within the water column, e.g.
at a depth of 5m, z =5m. Therefore, since g is positive in the convention used in these
notes, the integral evaluates to a positive number. This makes sense: the ocean pressure
must be higher than the atmospheric pressure. The physical interpretation of (4.4) is that the
pressure at a given depth in the ocean is determined by the weight of the overlying water
plus atmospheric pressure.
For rough calculations some simplications can be made. Firstly, the gravitational
acceleration [g??] is essentially constant, in any circumstance. Also, varies by only few
percent; youll recall the range is roughly 1020 kg/m
3
to 1040 kg/m
3
. Therefore, the integral
can be written
P = P
atmosphere
+g
avg
D (4.5)
at a depth of D metres, where
avg
is the average density of the water column between the
surface and z. Since [g??] 9.8 m/s
2
and 1000 kg/m
3
, 1m of water corresponds to
10
4
Pa. Recall from the evening news that typical atmospheric pressures are about 10
5
Pa,
so that the weight of 10m of water gives the same pressure as the entire atmosphere.
4.6. HOW T, S, P AND ARE MEASURED 51
Nearly always in Oceanography, the atmospheric pressure is ignored, when discussing
pressure within the water column. If we say the pressure in the water is 100 Pa, we mean
that the pressure is 100 Pa in excess of the atmospheric pressure.
Example. In an ocean trench at 10, 000m, the pressure reaches 10
8
Pa = 10
5
kPa or 1,000
atmospheres
8
. At this depth, the compressibility of water leads to a reduction in volume of
410
10
10
8
or 4%. (See Knauss [1978] for tables.)
In more typical depths does not vary very greatly with pressure and P =gz provides
a good approximation. Indeed, Oceanographers often regard pressure as a measurement of
depth, because instruments measure P, not z. Youll often see pressure on the y-axis of an
ocean prole or an section.
4.6. How T, S, p and are measured
This section is out of date; youd be better off reading the Emery and Thomson [1998]
treatment.
4.6.1. Instruments
Figure 4.12 illustrate the sort of bottle used for hydrographic sampling. A reversing
thermometer (which stores temperature at depth by a clever scheme that occurs when the
thermometer is turned upside down) measures temperature and the bottle captures water
that can be analysed for salinity and other chemical and biological properties. Until recent
decades, the bulk of oceanographic hydrographic data came from such devices.
Deep-sea reversing thermometer protected - enclosed in glass case, so p constant
Small correction for temperature change coming to surface.
Accuracy: 0.01

C
Depth resolution: discrete intervals
An old (c. 1900) but well-tested technique
Cost: $100
CTD (conductivity, temperature, depth)
Basic temperature measurement of a platinum wire (electrical resistance is a
function of T)
Accuracy: 0.002

C on T, 0.005 PSU on salinity (; DFO report authored by


Ross Hendry, BIO, 1993)
Depth resolution: continuous (1 10 m resolution), via a digitquartz pressure
gauge.
Modern tool, data transmitted by conducting cable to logger or computer
Cost: $4000 to $100,000
BT (bathythermograph)
1930
Bourdon tube - gas in a spiral tube, temperature changes pressure changes
tube unwinds.
Accuracy: 0.1

C
Depth resolution: continuous trace to predetermined depth (100, 200, 800 m,
etc.)
Can be used while ship is underway
Obsolete; replaced by XBT
Cost: N/A
8
Go to the movie The Abyss for some technical details of life at 1,000 atmospheres.
52 CHAPTER 4. SEAWATER PROPERTIES
Figure 4.12: Left: A niskin bottle laid out on a table. The white attachment holds a reversing
thermometer. For scale note that the diameter of the cylinder is 0.2m or so. Right: Diagram
indicating how a Nansen bottle overturns, when a messenger is dropped along the cable
(I), triggering a release so that it pivots (II) to a nal position (III). Note that at step II a
messenger at the bottom of the bottle is released, so that a bottle below will be triggered
when the messenger has slid down the cable to the level of the next bottle.
4.6. HOW T, S, P AND ARE MEASURED 53
Figure 4.13: A CTD instrument, manufactured by Sea-Bird Electronics Inc. (model 25).
(Source: http://www.seabird.com/products/profilers.htm.)
54 CHAPTER 4. SEAWATER PROPERTIES
XBT (expendable BT)
Thermistor, trailing copper wire signal back to recorder.
Accuracy: 0.1

C (better resolution)
Depth resolution: continuous to pre-determined depth
Convenient: ships, helicopter, etc
$40 per throw
RS 5 manually balance bridges, portable shallow water applications
Accuracy: 0.5

C (resolution 0.05

C)
Depth resolution: discrete down to 100 m.
Cost: few $10
2
Research
Accuracy 0.001

C
Resolution < 1cm.
Remote sensing (satellite or aircraft)
Radiation from sea surface approximately black body at infra-red wavelengths.
Intensity depends strongly on temperature and not much on anything else.
Accuracy 12

C (resolution to 0.1

C)
Depth resolution: none - indicates very thin surface layer only
Provides a broad spatial coverage, pictures often fascinating
Cost: $10
6
for satellite, $10
3
/year to buy data
Moored and towed instruments
1. Thermistors plus data loggers moored in specic locations
2. Thermistor chain - moored or towed vertical array
Research instruments
Resolution 10
5
C
4.6.2. How S is measured
The sea contains every natural element for which adequate detection has been devised
plus innumerable complex compounds. The major ions in solution are:
Chloride ion Cl

55% of dissolved material


Sodium Na
+
30.6%
Sulphate SO

4
7.7%
Magnesium Mg
++
3.7%
Potassium K
+
1.1%
The total concentration of dissolved material varies but these proportions are approxi-
mately constant.
Denition Salinity is the total amount of dissolved material, formally dened as the
total amount of solid material in g/kg when all carbonate has been converted to oxide,
all bromine and iodine replaced by chlorine, and all organic matter completely oxidised.
Typical values are as follows
Typical value 35
Oceanic average 34.7302. (Hint: dont remember all these digits.)
We need to measure salinity to at least 0.01 in the open ocean.
4.6. HOW T, S, P AND ARE MEASURED 55
Exercise 4.1 Salinity
a
In terms of kitchen-based tools, what does a salinity of 35 PSU mean?
a
Answer in chapter 21.
If one follows the denition, salinity proves to be difcult to measure directly. However,
Dittimar (1884), using samples from the Challenger expedition, found that the ratios of the
ion constituents are very nearly constant.
Salinity estimated using most abundant ion, chlorine. Chlorinity measured by titration
with AgNO
3
using potassium chromate as the indicator, giving measure of total halogen
content (Cl

, Br

, I

)
Chemical method of measuring salinity. The following formulare used:
S = 0.03+1.8050 Cl(1902 Knudsen) (4.6)
S = 1.80655 Cl (1969) (4.7)
to an accuracy (using titration) of 0.02 using a calibration against standard sea water.
Physical method of measuring salinity. Modern measurements of salinity are made
by measuring the conductivity of the water. The difculty is that conductivity is strongly
temperature dependent (see Figure 4.14). To obtain a measurement of salinity an instrument
must either
Measure temperature and compensate internally to give a direct salinity reading (RS-5
for example)substantial loss in accuracy.
Measure temperature simultaneously, record temperature and conductivity and make
corrections during analysisaccurate.
Put the sample in a bath at a known T and compare it to standard seawatermost
accurate.
Conductivity can be measured by electrical inductance. The inductive coupling between
two coils (mutual inductance) depends on the resistance of the water - low conductance
resulting in resistive loss and lower coupling.
There are two types of instrument based on this technique: (a) laboratory salinometers,
with which the conductivity of sample of sea water is compared with that of standard sea
water, and (b) CTDs, the workhorse of Oceanographic measurement.
Most post-1974 CTDs and lab salinometers use platinum electrode cells to measure
seawater conductivity. A known A.C. (alternating current) current is passed through the
central pair of electrodes, and the voltage drop generated by this current is sensed at the outer
pair of electrodes. The ratio of current to voltage is proportional to the seawater conductance.
This technique is so accurate it has surpassed the ability to measure salinity by any other
means. From 1970-1978 CTDs were calibrated by noting that the TS curve in the deep
ocean was very well-dened, and the salinities were shifted so that different CTDs agreed.
In 1978 a new denition of salinity in terms of conductivity became the new standard
9
.
Note that, while CTDs read out almost continuously (typically at something like 24Hz),
their depth resolution is limited by sensor resolution (time constant, ushing) , and the need
to measure T and C of the same bit of water. Errors in such things lead to errors in computed
salinity (and density).
Other methods of measuring salinity:
Refractive index: The refractive index depends principally on Cl

and temperature,
relatively insensitive to other constituents.
f
25
1.33250 = 0.000328Cl (4.8)
9
Reference: [UNESCO??] Tech. papers in marine science: number 37 - background papers; number 38 -
tables; number 44 - computer programs. You can obtain these from Division of Marine Sciences, [UNESCO??],
Place de Fontenoy, 75700 Paris, FRANCE
56 CHAPTER 4. SEAWATER PROPERTIES
Conductivity Ratio [unitless]
T
e
m
p
e
r
a
t
u
r
e
,

C
0.4 0.5 0.6 0.7 0.8 0.9
0
5
1
0
1
5
2
0
2
5
3
0

1
0


1
5


2
0


2
5


3
0


3
5


4
0


4
5

library(oce)
pdf("S_from_C_T.pdf", width=5, height=4)
n <- 100
cc <- seq(0.4, 0.9, length.out=n)
cond <- rep(cc, n)
tt <- seq(-2, 30, length.out=n)
temperature <- rev(rep(tt, n))
pressure <- rep(0, n
*
n)
dim(cond) <- dim(temperature) <- dim(pressure) <- c(n, n)
temperature <- t(temperature)
S <- swSCTp(cond, temperature, pressure)
par(mar=c(4.3, 4.3, 1, 1))
contour(cc, tt, S,
levels=seq(2.5, 50, 5),
lty="dashed",
drawlabels=FALSE,
xlab="Conductivity Ratio [unitless]",
ylab=expression(paste("Temperature, ",degree,"C")))
contour(cc, tt, S,
levels=seq(0, 50, 5),
labcex=1,
add=TRUE)
Figure 4.14: Top: Salinity [PSU] in terms of conductivity ratio and temperature, according
to the [UNESCO??] formulation [Fofonoff and Millard, 1983]. Bottom: R code used to
create the diagram.
4.7. EQUATION OF STATE: = (S, T, P) 57
American Optical makes an easy to use refractometer which reads directly in salinity.
This is great for estuarine work.
Direct density measurements: Usually this means using hydrometer and referring to
Kundsen (1901) tables to convert at standard temperature and thence to Cl

or S.
Remote sensing from aircraft/satellite In certain frequency bands of the electromag-
netic spectrum, the emissivity of the sea surface does depend on the salinity (as well
as on the temperature). Using a 1 1.5 GHz SAR (synthetic aperature radar), one
gets t 2

C and S 5 (around 35), top 1 cm only.


4.6.3. How p is measured
There are various techniques for measuring pressure:
Wire out on a bottle cast. (Method has difculty with wire angle.)
Protected and unprotected thermometers - one bulb of mercury exposed to ambient
pressure gives a slightly higher reading of temperature than a protected thermometer
(reversed at the same depth). Accuracy is 5 m for 100-1,000 m depths (unprotected
reads about 1

C higher per 100 m).


Mechanical motion - bellows or Bourdon tube. Not very accurate.
Vibrotron - vibrating wire held between diaphragms. Pressure on one diaphragm
changes vibration frequency of wire which is detected electromagnetically. Quartz
crystal oscillor is similar. The crystal can be cut so that the lattice vibrations are
pressure dependent not temperature dependent (cut for temperature sensitivity its
an accurate thermometer). Accuracy - high (better than 0.01 m of surface change -
generally limited by uncertainties in water density).
Strain gauge - deection of a membrane stretches a wire, causing a resistance change.
Accuracy 0.1% - used in CTDs.
Capacitance changes - or resistance changes. Uses carbon granules and semiconduc-
tors; same principle as per an audio microphone.
4.7. Equation of state: = (S, T, p)
The [in-situ??] density is a complicated function of salinity, temperature and pressure
= (S, T, p) (4.9)
varies between 1020 kg/m
3
and 1030 kg/m
3
in most of the ocean. The actual relationship
between and S, T, p is called the [equation-of-state??], and is given by sets of tables
of complex polynomials (for computer analysis) now standardized internationally
10
. The
following page has a graph of the dependence of on S and T at atmospheric pressure, the
next two pages copy the [UNESCO??] report, and the following page gives a table you can
use for hand computation.
Example: Seawater with salinity S=30 PSU and in-situ temperature T = 20

C has in-situ
density = 1020.954 kg/m
3
at atmospheric pressure.
Exercise 4.2 Freshwater density
a
Roughly what is the mass of the water in your bathtub?
a
Answer in chapter 21.
10
Reference: [UNESCO??] report no. 38; Gill Appendix A3.1.
58 CHAPTER 4. SEAWATER PROPERTIES
Exercise 4.2 Seawater density
a
What is the density of seawater at salinity S = 35 PSU, temperature T = 10

C
and pressure p = 100 dbar (i.e. at roughly 100 m depth)? If the water were to
be raised adiabatically to the surface, what would its temperature and density
become?
a
Answer in chapter 21.
As the rst two digits of [rho??] do not change over the oceanic extremes, we often use
[sigma??] instead, dened in various related forms (sometimes called anomaly forms) as
(S, T, p) = (S, T, p) 10
3
kg m
3
(4.10)
and

t
= (S, T, 0) 10
3
(4.11)

= (S, , 0) 10
3
(4.12)
The variants [sigma-t??] and [sigma-theta??] are used rather than [sigma??] to
partially sidestep the [adiabatic??] [lapse-rate??] problem
The quantity [sigma-theta??] is often called a potential density anomaly, with the
potential part referring to the use of potential temperature. This should properly be
called potential density anomaly referenced to surface pressure since =(S, T, p) is the
potential temperature referenced to the surface. (One could dene potential temperatures
with reference to any surface . . . e.g.
1
is a shorthand for (S, T, p, p
re f
) with p
re f
=
1000 dbar, and thus we might dene
1
= (S, (S, T, p, p
re f
), p
re f
) 1000 kg/m
3
with
p
re f
= 1000 dbar. Further subtleties exist, relating to what are called neutral surfaces, but
these are beyond the present scope.)
Since it is difcult to describe all of this in words, it is very common to speak the symbols
aloud; in seminars and in scientic conversation, one would say sigma tee, not potential
density anomaly referenced to surface pressure.
Early oceanographers introduced a large number of operational denitions to help them
with the analysis of bottle data. These included
specic volume 1/
specic volume anomaly =
1
(S, T, p)
1
(35, 0, p)
thermosteric anomaly
To some extent, these quantities have disappeared with computer data processing. Nowadays,
one computes using the UNESCO [equation-of-state??]; see Figure 4.15 for sample
code, and Figure 4.16 for the T-S relationship in terms of
t
.
4.8. Linear equation of state
A linearized equation of state may be written
=
0
[1(T T
0
) +(SS
0
)] (4.13)
where this is a linear expansion about some point (S
0
, T
0
), at which the density is
0
. In
the above equation, we have used the so-called thermal expansion coefcient [alpha??]
dened by
=
1

T
(4.14)
and the so-called haline contraction coefcient [beta??] dened by
=
1

S
(4.15)
Since this linearized equation of state is so simple, its used commonly for quick calculations
(e.g. at the blackboard) and in theories. The thermal expansion depends on temperature,
falling from 210
4
C
1
at room temperature to 210
5
C
1
below roughly
5

C. The haline contraction coefcient is roughly constant, over many applications, with
value = 810
4
PSU
1
.
4.8. LINEAR EQUATION OF STATE 59
/
*
@(#) rho.c 1.1 7/9/91 DEK
*
/
/
*
*
rho() -- density of seawater (UNESCO eqn of state)
*
*
SYNTAX double rho(double S,double T,double p)
*
*
UNITS S in psu; T in degC; p in dbar
*
*
RETURN VALUE density in kg/(m
**
3)
*
/
#include <math.h>
double
rho(double S, double T, double p)
{
double rhow, Kw, Aw, Bw, p1, S12, ro, xkst;
rhow = 999.842594 +
T
*
(6.793952e-2 +
T
*
(-9.095290e-3 +
T
*
(1.001685e-4 +
T
*
(-1.120083e-6 + T
*
6.536332e-9))));
Kw = 19652.21
+ T
*
(148.4206 +
T
*
(-2.327105 +
T
*
(1.360477e-2 - T
*
5.155288e-5)));
Aw = 3.239908 +
T
*
(1.43713e-3 +
T
*
(1.16092e-4 -
T
*
5.77905e-7));
Bw = 8.50935e-5 +
T
*
(-6.12293e-6 +
T
*
5.2787e-8);
p1 = 0.1
*
p;
S12 = sqrt(S);
ro = rhow +
S
*
(8.24493e-1 +
T
*
(-4.0899e-3 +
T
*
(7.6438e-5 +
T
*
(-8.2467e-7 + T
*
5.3875e-9))) +
S12
*
(-5.72466e-3 +
T
*
(1.0227e-4 -
T
*
1.6546e-6) +
S12
*
4.8314e-4));
xkst = Kw +
S
*
(54.6746 +
T
*
(-0.603459 +
T
*
(1.09987e-2 -
T
*
6.1670e-5)) +
S12
*
(7.944e-2 +
T
*
(1.6483e-2 +
T
*
(-5.3009e-4)))) +
p1
*
(Aw +
S
*
(2.2838e-3 +
T
*
(-1.0981e-5 +
T
*
(-1.6078e-6)) +
S12
*
(1.91075e-4)) +
p1
*
(Bw +
S
*
(-9.9348e-7 +
T
*
(2.0816e-8 +
T
*
(9.1697e-10)))));
return (ro / (1.0 - p1 / xkst));
}
Figure 4.15: Equation of state, in the C programming language.
60 CHAPTER 4. SEAWATER PROPERTIES
30 31 32 33 34 35
0
5
1
0
1
5
2
0
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
28
pdf("TS-blank.pdf", width=9, height=10)
library(oce)
fake.S <- c(30, 35)
fake.T <- c(-2, 20)
fake.p <- c(0, 0)
fake.data <- as.ctd(fake.S, fake.T, fake.p)
# use white ink for the fake data
plotTS(fake.data, col="white", rho.levels=20)
grid(col="red")
Figure 4.16: Top: Variation of [sigma-t??] with temperature and salinity, from the
UNESCO equation of state. Bottom: R code used to create the diagram. (The blue line is
the freezing-point curve, which sets a lower limit on observed seawater properties.)
4.9. STRATIFICATION AND STABILITY 61
Figure 4.17: Denition sketch for [stratication??]. The situation illustrated in the prole
on the left is gravitionally stable, since heavy water is underlying lighter water, while the
case on the right is unstable, and will lead to overturning.
4.9. Stratication and stability
Density usually increases with depth in the ocean. If not, the situation is unstable and
strong [convection??] should result, returning the system to stability. Figure 4.17 illustrates
stable and unstable [stratication??].
Stability is often expressed in terms of N, the so-called Brunt-Vaisala frequency, or, in
more modern terms, the [buoyancy-frequency??] :
N
2
=
g

z
(4.16)
where
0
is a reference density, e.g. the average density in the region. (The density in a
given region varies so little that for many calculations it doesnt much matter whether we
pick
0
as the average density, or the deep-water density, etc.)
N has units s
1
. This is the frequency at which a [parcel??] of uid would bob up
and down if displaced vertically. Computer algorithms for [buoyancy-frequency??]
and other such things will be provided in class/tutorials. Note that if the [in-situ??]
[density??] is used in the above equation, a correction term g
2
/C
2
(with C the speed
of sound, roughly 1500 m/s) must be subtracted to account for the compressibility
of seawater; this (small) correction neednt be used if the potential density is used
instead of the in-situ density.
Even when we have lighter water over heavier water the system is not necessarily
stable. Molecular scale diffusional processes may be important - double diffusion, of
which more will be said later.
Typical values of N follow. Ive given them in cycles per hour, since some folks nd
it useful to visualize the waves bobbing up and down in that way. But you should
convert these to radians per second, as Ive done for the rst element of the list, to get
familiar with the values of N as they would appear in a formula.
Seasonal thermocline: 10 cycle/h (0.017s
1
)
Main thermocline: 3 cycle/h
Deep water: 1/20 cycle/h
4.10. Baroclinic and Barotropic States
Two words youll hear often in PO seminars are [baroclinic??] and [barotropic??].
A uid is said to be in a [baroclinic??] state if the P and vectors are at an
angle. Another way to say this is that planes of constant density and pressure intersect.
Mathematically, a system is baroclinic if P ,= 0.
By contrast, a uid is said to be in a [barotropic??] state if these vectors are parallel.
62 CHAPTER 4. SEAWATER PROPERTIES
z

[

m

]

[ kg m
3
]

8
0

6
0

4
0

2
0
0
12 14 16 18 20 22 24
Data
Mean
Fit
Figure 4.18: Density proles in St Lawrence Estuary, taken over a 24-hour period, by a
ship transiting across the channel at the SLEIWEX-2008 study site. The darker lines are the
average and a t to a simple function = (z). Note that there are signicant deviations of
individual proles from the mean prole and the model prole. The deviations involve depth-
dependent variations in /z (well-mixed layers sandwiched between highly-stratied
regions) that vary from prole to prole, as well as some systematic deviations (e.g. the
mean and model proles overestimate /z near the surface).
4.11. Fluid energy
4.11.1. Kinetic energy of a moving uid
Consider a small cube of uid, of volume xyz. This will have mass xyz, and if
it is moving at a speed v, it will have [kinetic-energy??] xyzv
2
/2, from elementary
considerations (recall section 3.2). The unit for this kinetic energy is the [Joule??] (1J is
equivalent to 1Nm or 1kgm
2
s
2
).
From this we may derive per unit formulations Kinetic Energy of a uid
KE per unit volume =
1
2
u
2
(4.17)
(which has unit J/m
3
or N/m
2
or kgm
1
s
2
) as well as
KE per unit mass =
1
2
u
2
(4.18)
(which has unit J/kg or Nm/kg or m
2
/s
2
). Both of the per unit forms given above are
common in the literature, although sometimes youll also see KE expressed per unit surface
area, per metre across a wave, etc. (Hint: to keep track of the formulation being used in a
given instance, examine the units.)
4.11.2. Gravitational potential energy
The General Case
Consider an ocean column, with density = (z) at depth z above the ocean bottom.
Imagine constructing this water column by taking slabs of water and lifting them
11
from
the ocean bottom up to the depth at which they are observed. We may calculate the PE of
the water column by adding up the [potential-energy??] contributions that are associated
11
The same ideas hold if you do the thought experiment from the ocean surface downwards, and you should
try that, to see how the Mathematics works out.
4.11. FLUID ENERGY 63
Figure 4.19: Denition sketch for potential energy. Imagine raising the slab from the ocean
bottom at z = 0 to some height z above the bottom. Since the slab has horizontal-planar
area A and thickness z, it has a mass of (z)Az kilograms and it weighs g(z)Az
[Newton??]. The work done in raising it from the bottom, the product of this weight and the
distance of raising, is thus gz(z)Az [Joule??]. This, then, is PE associated with this slab,
at its nal resting position. The PE of the whole water column is calculated by summing the
contributions of an innite number of innitely thin slabs; see (4.19) in the text.
with lifting up each slab. Let the slab have horizontal-planar area A and an innitesmally
thin vertical thickness z. Consider the slab that ends up at a depth of z, i.e. a slab which
has density (z). This slab weighs g(z)Az [Newton??], and raising it through z meters
requires doing gz(z)Az [Joule??] of work. This work is stored in the PE. Thus, the PE
of the entire water column (relative to the bottom) may be calculated simply by adding up
the contributions of the slabs in the interval 0 < z < H, which is done in the integral
PE =
_
H
0
Agzdz (4.19)
(which has unit J). Dividing by the plan area A (Figure 4.19), we get
PE per unit surface area =
_
H
0
gzdz (4.20)
(which has unit J/m
2
or N/m or kgs
2
). Dividing by the water depth yields
PE per unit volume =
1
H
_
H
0
gzdz (4.21)
(which has unit J/m
3
or N/m
2
or kgm
1
s
2
). and this can be converted by dividing by the
mean density, = H
1
_
H
0
dz, yielding Potential Energy of a uid
PE per unit mass =
1
H
_
H
0
gzdz (4.22)
(which has unit J/kg or m
2
/s
2
).
All of these formulae nd their place in the literature, but youll most often nd the per
formulae used. All of this may make more sense to you if you think about some special
cases. Lets focus on the per unit mass formula, for concreteness.
Unstratied case
If (z) =
0
for all z, then (4.22) yields the PE per unit mass as
1
2
gH. The
1
2
factor may
make sense if you imagine that this water column has its centre of mass at a distance
1
2
H off
the bottom. In that sense, our formula is g height, which is just the formula for a solid-body
case (recall section 3.2). And what are we to make of the value obtained with this formula
what does that mean? Well, lets take an example. Suppose we have an above-ground
64 CHAPTER 4. SEAWATER PROPERTIES
swimming pool
12
with a depth of 100m. The PE per unit mass is
1
2
gH 500J/kg, or
500m
2
/s
2
. Notice that the unit is the square of a speed unit. To contextualize that, suppose
that the walls of the pool are suddenly removed, allowing the water to rush out onto the at
Texas landscape. If all of the PE is converted to KE (equal to
1
2
u
2
), the out-rushing water
will achieve a speed of u 30m/s. This is about 100 kilometers/hour, i.e. highway speed. In
some sense, then the potential of this PE is huge. But of course, we dont expect the walls
of the pool to disappear. And, if I had written this in terms of the ocean on a continental
shelf of depth 100m, we wouldnt expect the continental boundaries to disappear, either. So,
in some sense, this PE is bound up by geometry, and not really available to be converted
into KE. However, there is a store of PE that is available, under some scenarios of vertical
exchange, and this is the PE that is stored in the [stratication??]. The next few items
cover this, starting with the simplest case, of two-layer [stratication??].
Two-layer case
Again, consider a water column of height H, but now suppose there are two layers, of
differing densities: suppose that z =
1
2
H marks the boundary between a lower layer of density

0
+ and and upper layer of density
0
. Then (4.22) yields PE =
1
2
gH (1
1
2

0
).
Note that if
0
, this reduces to the unstratied case, as expected. It is common to
focus on the PE associated with [stratication??] alone, and to do this, we may subtract
the unstratied PE from this two-layer formulation, with the result being a stratication PE
per unit mass of
1
4

0
gH.
Linearly-stratied case A
Consider the density prole (z) =
0
(1N
2
z/g)where N is a constant. Using (4.22)
we thus may calculate the PE per unit density as
1
2
gH
_
1
2
3
N
2
H
g
_
. Notice that this reduces
to the unstratied case if H g/N
2
, which one might express as H/H
0
1, where H
0
is a [stratication??] lengthscale dened by H
0
= g/N
2
. You might nd it instructive to
calculate the stratication PE, as in the two-layer case, and to compare the two. (Note:
if you nd doing this integral scary, consult section 18.1 for an introduction to symbolic-
mathematics programs, using this integral as an example.)
Linearly-stratied case B
Now consider the density prole (z) =
0
[1N
2
(z
1
2
H)/g]where N (the [buoyancy-
frequency??] dened in (4.16)) is a constant. (Notice the subtly-different form than in
case A. You might wish to graph the two proles, to see how they differ.) This has PE per
unit density equal to
1
2
gH
_
1
1
6
N
2
H
g
_
. Notice how the factor on N
2
H/g has changed,
simply by changing the depth at which the density has the value
0
, without changing the
[stratication??], N. This sort of subtlety comes up often in these calculations; while you
might guess immediately, perhaps on dimensional grounds, that a term of the form N
2
H/g
must come into play, it would be difcult to guess the coefcient on this term.
Energy increase through mixing a water column
Since the mean density in linear case B is
0
, we may compare this PE with the PE
for an unstratied case of equal density. From the above, the difference between the two
is
1
12
N
2
H
2
, with the smaller energy being in the stratied case. Thus, we may conclude
that mixing up a water column of height H, starting with an initial [stratication??] of N,
increases
13
the [potential-energy??] per unit mass by an amount
1
12
N
2
H
2
. (Again, one
might guess the N
2
H
2
part of this result, but I dont know anybody who could guess the
1
12
-th part, except for folks who have worked through such examples mathematically.) This
formula for PE change during homogenization of a stratied uid comes up a lot in studies
of ocean mixing, where the breaking of internal waves may locally mix uid. If the waves
have a particular energy, E say, then they can mix up a region of height

2E /N. So, we
could try to infer mixing rates (which are difcult to observe directly, as Ill explain in class)
from the heights of mixed patches [Galbraith and Kelley, 1996].
12
This is a big pool. Kind of a Texas thing. Obey the lifeguard or get shot.
13
You may remember that the PE increases simply by realizing that mixing the water column involves pulling
heavy deep water up to the surface, which requires work, and pushing buoyant surface water downwards, which
also requires work.
4.12. FLUXES ACROSS THE AIR-SEA INTERFACE 65
Application: mixed-layer energetics
Consider a coastal region with a top/bottom density difference = 5kg/m
3
and a depth
H = 50m. (Look at some coastal data to see if these numbers are roughly sensible; Im
partly picking them because the numbers work out cleanly!) This gives N
2
= 10
3
s
2
, so
that the PE change in mixing this water column is 0.2J/kg. In class, Ill discuss how to put
this number in the context of mixed-layer deepening, location of fronts in the coastal ocean,
and other applications. But, in case you want something to puzzle over, read the next section
and think about the work done by the wind, and then return to the following formula, which
I will argue is the timescale for wind-stress to cause mixing of an initially stratied surface
layer of thickness H:
t

w

a
N
2
H
3
C
D
U
3
(4.23)
where U is the speed of the wind and [CD??] is a dimensionless [drag-coefcient??] to be
discussed in the next section. The proportionality factor takes into account the efciency
of the conversion of wind-stress work into mixing; typically E 0.2, as well as the factor
1/12 that weve been seeing in the formulae. If initial [stratication??] PE, proportional to
N
2
H
2
, is somehow set up as a constant in a given region (owing to summertime surface
heating, say), then the timescale to create a mixed layer of thickness H is proportional to
H/U
3
. This ratio has quite well-known, in a slightly different context
14
. as the Simpson-
Hunter criterion [Simpson, 1981].
4.12. Fluxes across the air-sea interface
The ocean is strongly inuenced by:
Wind stress - ux of momentum from atmosphere
Solar radiation - heat ux
Evaporation/precipitation - ux of water coastal runoff (and heat)
Various other uxes, of gases, particles etc. are signicant in chemical, geochemical,
biological, climate balances, i.e. CO
2
, O
2
.
4.12.1. Momentum ux Wind stress
(For details of calculation methods, see the TOGA-COARE bulk formulae, which in
1997 were on the web at http://www.coaps.fsu.edu/COARE/flux_algor; a
web-search for COARE or TOGA will help you nd this material if its moved by the time
you read this.)
Tangential wind stress at the ocean surface may be regarded as a vertical transfer of
horizontally-directed momentum. The horizontal momentum is
a
u, while of course w is
doing the transporting, so the stress is =
a
uw. This may vary over time on a variety of
timescales. Often we are interested in slow variations of (say, over hours or days), even
though we know that there are local bursts that vary over seconds. Let overbar denote the
slowly-varying component of some quantity, e.g. u might be velocity averaged over an
hour. Let prime, e.g. u
/
denote variations of velocity about this slowly-varying signal. Thus,
u = u +u
/
, and it follows that u
/
= 0 by denition. This same decomposition, called the
Reynolds decomposition, can be done for w as well. Thus, the slowly-varying stress may be
written
[tau??] [shear-stress??]
=
a
uw (4.24)
where we are assuming that
a
does not vary appreciably over timescales of interest,
and we have dropped the overline on because we are focussing only on its slowly varying
component.
14
Ill discuss this in class. In this criterion, the mixing comes from tidal ows over the ocean bottom, but
otherwise all the ideas are identical to those shown here.
66 CHAPTER 4. SEAWATER PROPERTIES
Figure 4.20: Denition sketch for U
10
, the wind velocity at 10 m above the sea surface. U
10
is conveniently measured, and is used in many air-sea ux parameterizations.
But, given the denition of u and u
/
, and the same for w, we may write the velocity
product in the stress as
uw = (u+u
/
) (w+w
/
) (4.25)
which can be simplied to
uw = u w+u
/
w
/
(4.26)
since uctuating quantities vanish upon averaging.
Now, the wind cant blow straight down into the sea, so that near the air-sea boundary,
we must have w = 0. This means that the rst term above must vanish, so that (4.24) can be
written
=
a
u
/
w
/
(4.27)
The overlined term is a covariance, and for this reason the above formula is often called an
eddy-covariance formulation of stress. It is the gold standard of measurement. Unfortunately,
it is really quite difcult to make measurements for use in this formula (and impossible to
mimic them in weather models). For this reason, uxes are commonly calculated with what
is called a bulk formula of the form

0
=
a
C
D
U
2
(4.28)
Here, U is the wind speed measured 10m above the water surface (or extrapolated to this
height; see Figure 4.20), and [CD??] is a dimensionless [drag-coefcient??] with a value
of order (1.50.5)10
3
. Although there are theoretical reasons to accept a formula of the
form of (4.28), it could also be viewed as essentially an empirical result. The value for C
D
must come from calibration against covariance-based stress estimates. If things were simple,
C
D
would be a constant, but in fact it seems to vary with wind speed, with the stability of
the lower atmosphere, with wave height, etc. Despite uncertainties as to the value of C
D
and its variation, the formula (4.28) is used routinely in calculations and in models. Really,
there is nothing else to do. Measuring uxes at a point with the covariance technique is very
difcult, and trying to make a weather model simulate the turbulent eddies that exchange
momentum near the sea surface is simply impossible at this time. Its a bulk formulation or
nothing.
4.13. WAYS TO LOOK AT T, S, P AND DATA 67
Exercise 4.4 Drag force from wind stress
a
A 10 m/s wind blows steadily over the water surface. (a) What is the wind-stress
that arises from this? (b) What if the wind is 5 m/s? (c) What if the wind is
80 km/hour?
a
Answer in chapter 21.
A note on units. Stresses are typically of the order 0.1Pa. A [Pascal??] is 1
[Newton??] per square metre. An old unit used before Pascals became popular is dynes per
square centimetre; one [Pascal??] is 10 dynes per square centimetre.
4.12.2. Heat ux
(For details of calculation methods, see the TOGA-COARE bulk formulae, which in
1997 were on the web at http://www.coaps.fsu.edu/COARE/flux_algor; a
web-search for COARE or TOGA will help you nd this material if its moved by the time
you read this.)
The net rate of heat gain by the ocean depends on a number of factors, none of which are
particularly easy to estimate reliably. In principle, the heat ux A is given by
A = Q
s
Q
b
Q
h
Q
e
(4.29)
where Q
s
is the net shortwave radiation absorbed (about half is direct sunlight, half scattered
light from sky and clouds). Q
b
is the long wave back radiation (Complex formulae for Q
s
and Q
b
which depend on cloud cover, ice, humidity, temperature, etc).
Sensible heat ux
Q
h
=
a
C
H
C
p
(T
S
T
10
)U
10
(4.30)
where C
H
is a coefcient of heat exchange C
p
the specic heat of air, T
S
the surface
temperature of the water T
10
air temperature at 10m. C
H
C
D
as a rst given (O.K. for
neutral stability.
Latent heat ux
Q
e
=
a
LC
E
(q
S
q
10
)U
10
(4.31)
where L is the latent heat of evaporation, C
E
a coefcient C
H
q
s
is the humidity at
surface (assuming saturation), q
10
the humidity at 10 m determined from wet and dry bulb
temperatures [Knauss, 1978, Chapter 2].
Note: Q
e
is typically much more important than Q
h
.
Figure 4.21 shows the annual balance for A in the Atlantic. There is clearly a large loss
of heat when a warm current such as the Gulf Stream is moving northward.
There is a very important question which concerns the global heat balance. More heat is
clearly absorbed by the earth at the equator than at the poles. The earth is in approximate
steady state so there must be a substantial heat ux towards the poles. Is this balance by
uxes in the ocean or is the atmosphere signicant?
Bryden and Hall (Science 22 Feb. 1980, 207, 884-886) estimating the oceanic component
using a section at 25

N in the Atlantic, come up with a value of 1.1 0.3 10


5
watts.
However satellite radiation measurements suggest the oceans should transport 2.510
15
watts
15
.
4.13. Ways to look at T, S, p and data
Large sets of observational data of S, T and as functions of horizontal space, depth and
time are not particularly easy to assimilate. A basic concept is to look at a data set every
possible way (4- or 5- dimensional plots being rather impractical).
There are several tried-and-true methods of looking at oceanographic data:
15
For more on global budgets, see [Knauss, 1978, Chapter 3].
68 CHAPTER 4. SEAWATER PROPERTIES
Figure 4.21: Map of air-sea heat ux in units of kcal/cm
2
/year. Negative values mean
the air is being heated by the water (e.g., -120 kcal/cm
2
/year over Gulf Stream). Source:
[Knauss, 1978, Fig. 3.6].
Proles The top two panels of Figure 4.22 illustrate proles, i.e. diagrams in which
some quantity of interest is plotted as a function of depth, drawn with depth on the
vertical axis, with the water surface at the top of the diagram.
Successive proles, taken in the same place at different times. The top panel of
Figure 4.3 is an example. It shows the temperature prole in the upper layers at
different times of the year, revealing systematic changes in both the temperature and
thickness of the upper mixed layer.
At short intervals or small distances, the proles may be very similar. A standard
technique is to spread them apart by offsetting the scale of F. This makes it easier to
follow the behaviour of any particular feature in the prole without getting a jumble
of overlapping lines.
Sections The same information can be plotted as contours in a diagram with vertical
position z running along the vertical axis and horizontal position x running along the
horizontal. Such a diagram is called a section (like a cross-section).
Horizontal maps The simplest maps are horizontal mappings of conditions at a given
depth; Figure 4.2 is an example. Sometimes, it proves useful to plot properties on
a particular density surface, as opposed to plotting at a xed depth. This is because
water tends to move and mix on surfaces of constant density.
In W usts atlas of the Atlantic based on the Meteor expeditions of 1925-7, S, T, oxygen
etc. are plotted using their values at the core of the water mass under consideration.
Characteristic diagrams, explained in the next section.
4.13.1. The T-S diagram
Any two properties can be plotted against each other but the T-S relation is the most
commonly used and generally the most useful. The data from a given station provides a
curve on the diagram. Practitioners have several choices when constructing such diagrams.
For some purposes, the data are joined with line segments, and for others the data are
presented just as dots. Sometimes depth is colour-coded, sometimes not. Sometimes a single
trace is shown, for a single cast of the CTD (or set of bottle observations), and sometimes
multiple traces are drawn. It all depends on the purpose that the practitioner has in mind.
4.13. WAYS TO LOOK AT T, S, P AND DATA 69
P
r
e
s
s
u
r
e

[
d
b
a
r
]
4 6 8 10 12 14
Temperature [C]
4
0
3
0
2
0
1
0
0
30.0 30.5 31.0 31.5
Salinity [PSU]
P
r
e
s
s
u
r
e

[
d
b
a
r
]
22.5 23.0 23.5 24.0 24.5 25.0

[ kg m
3
]
4
0
3
0
2
0
1
0
0
0.0000 0.0005 0.0010 0.0015 0.0020 0.0025
N
2
[ s
2
]
30.0 30.5 31.0 31.5
4
6
8
1
0
1
2
1
4
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
22.5 23
23.5
24
24.5
25
CTD Station
File: bedford_basin_0302.cnv
Ship: Divcom3
Cruise: Halifax Harbour
Station: 1
Depth: 44.141
Location: 44.6843N 63.6439W
pdf("profile.pdf", width=8, height=10)
library(oce)
data(ctd)
plot(ctd)
Figure 4.22: Top: Sample CTD summary diagram, with data from the St Lawrence Internal
Wave Experiment (SLEIWEX); the estuarine location explains the low salinities and linear
TS relationship. Bottom: R code to produce the diagram. Exercise: use the online R help
in this package, to learn how to (a) plot a TS diagram, (b) plot a T prole, and (c) extract a
portion of the data set, say from 10 m to 20 m, and repeat this overview plot with it.
70 CHAPTER 4. SEAWATER PROPERTIES
Figure 4.23: Schematic [TS-diagram??], showing relationship to proles. The curved
lines are lines of constant density, i.e. [isopycnal??] lines. As an aid to understanding the
gure, note on the proles at the top that points 3 and 4 have nearly equal temperature, but
varying salinities, so they are aligned almost horizontally on the graph. Similarly, points 1
and 2 are close in both T and S.
For example, Figure 4.24 shows observations made in the year 2007 at station 2, which
is in 150 m of water, a few kilometres south of Halifax on the Scotian Shelf
16
. The labels
beside the traces indicate day numbers, starting at 1 for 2008-Jan-01. A casual glance
indicates the expected warming in summer, but careful scrutiny and statistical analysis might
be required to learn much more; that is the nature of working with real data, and the point of
showing a messy graph here.
Uses of the [TS-diagram??]:
Identication of water masses. (see e.g. [Knauss, 1978, Fig. 9.7].) Either a point or
a section of the curve may serve to characterize the water observed. A comparison with
the empirical picture of the oceans built for the whole set of historical measurements.
Stability. Density values may be included on a T-S diagram. The angle between the
station curve and the lines of constant [sigma-t??] provides a measure of the local
stability, and the possibility of double-diffusive instability.
Mixing. If two water types mix, provided heat and salt are conserved the resulting
mixture will have T,S properties that plot on the line joining the parent water types.
Figure 4.23 guides an example. If we take a unit of the mixture (T
3
,S
3
) which consists
of (1a) of (T
1
, S
1
) and a of (T
2
, S
2
) then
T
3
= (1a)T
1
+aT
2
(4.32)
and
S
3
= (1a)S
1
+aS
2
(4.33)
So the new mixture lies on the line joining (T
1
, S
1
) (T
2
, S
2
) close to T
1
, S
1
if a is small,
close to T
2
, S
2
if a is close to one.
16
Source: http://www.meds-sdmm.dfo-mpo.gc.ca/zmp/hydro/station_year_e.html?
a=2&y=2008.
4.13. WAYS TO LOOK AT T, S, P AND DATA 71
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
16
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
51
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
63
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
95
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
98
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
112
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
128
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
172
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
189
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
200
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
214
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
215
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
221
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
242
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
272
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
284
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
284
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
292
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
311
30 31 32 33 34
0
5
1
0
1
5
Salinity [PSU]
T
e
m
p
e
r
a
t
u
r
e

[

C
]
21 22 23 24
25
26
27
346
pdf("station-2.pdf", width=7, height=7)
library(oce)
## http://www.meds-sdmm.dfo-mpo.gc.ca
file <- "CTD_SS22007_B.CSV"
d <- read.csv(file, skip=38) # header of 38 lines
t <- unique(d[,7]) # time, in days
Slim <- range(d[,17], na.rm=TRUE) # S in col 17
Tlim <- range(d[,15], na.rm=TRUE) # T in col 15
pmax <- max(d[,13], na.rm=TRUE) # p in col 13
nt <- length(t)
for (i in 1:nt) { # draw casts separately
dd <- d[d[,7] == t[i],]
ddd <- as.ctd(dd[,17], dd[,15], dd[,13])
if (i != 1) par(new=TRUE) # superimpose plots
plotTS(ddd, Slim=Slim, Tlim=Tlim, col="blue", type=l)
text(dd[1,17], dd[1,15], sprintf("%.0f",t[i]),
pos=3, cex=0.9, col="red")
}
Figure 4.24: Top: [TS-diagram??] showing measurements made during the year 2008 at
Station 2, south of Halifax. Each trace corresponds to a single prole, and the label indicated
the day on which the prole was done. Bottom: R code used to make the diagram. Advanced
exercise: download the data yourself, using the link mentioned on the previous page, and
make a similar plot, but showing monthly or seasonal averages, instead of individual casts.
72 CHAPTER 4. SEAWATER PROPERTIES
Figure 4.25: Schematic of salt fountain thought experiment. A copper pipe extends from
the warm, salty surface waters down to the cooler, fresher waters below.
In addition, the T-S diagram provides a useful way of explaining a variety of convective
phenemena that are particular to the heat/salt-stratied ocean.
Cabbelling
The lines of constant [sigma-t??] on the T-S diagram are slightly curved, mixing lines
are straight. Hence if two different water types are mixed the mixture is denser than the
parents and should sink. This is not thought to be generally important but may be locally
signicant in the Antarctic. (For an entry to the literature on this, see McDougall [1987].)
Salt fountain
Since near the surface particularly, salinity often decreases with depth (i.e., S is heavy
on top) it should be possible to sustain a ow if we put a long tube from deep water to the
surface. Temperature is exchanged through the tube walls but not salt. Hence after some
initial pumping a ow should become self sustaining. Figure 4.25 shows this.
Salt ngering
If we have an initially stable density interface with hot salty water on top of colder fresher
water, molecular processes may slowly cause instabilities. The molecular diffusion of heat
is 100 larger than the diffusivity of salt (Figure 4.26) so the temperature gradient diffuses
much quickly than the salinity gradient. The stability effect reduces, eventually one has
instability, a buckling of the interface which increases to form salt ngers. The net effect is
that the salt falls down, the top mass of water becomes even lighter, and the bottom one
more heavy
17
.
Diffusive instability
In regions with relatively colder and fresher water overlying warmer and saltier water
that is, the reverse of the salt-ngering case a second type of double-diffusive instability
may occur. This is sometimes called diffusive instability and sometimes the layering
case. Figure 4.27 indicates regions of the ocean that are susceptable to this form of double-
diffusion.
Further reading on double-diffusion. Reviews of research done on double-diffusive
phenomena appear about once per decade. Many of these have been co-written by Stewart
Turner, whose wonderful textbook [Turner, 1973] also touches on this topic. Since double-
diffusion may occur in many uids (stars, magma chambers, . . . ), some of the reviews may
be tangential to your interests. For oceanographic applications, the best review may be that
of Schmitt [1994]. This is slightly dated, and focusses on the salt-nger case. For a newer
treatment of the layering case, see [Kelley et al., 2003] and references contained therein.
4.14. OTHER OCEAN TRACERS 73
Figure 4.26: Properties of molecular diffusion in seawater.
74 CHAPTER 4. SEAWATER PROPERTIES
Figure 4.27: Ocean regions that are susceptable to diffusive instability, after Figure 1 of
Kelley et al. [2003].
4.14. Other ocean tracers
Much of the broad-scale ow in the world ocean was intuited by examination of tracers.
So far, we have focussed on the use of T and S. Another tracer that is often examined is
oxygen concentration. Figure 4.28 illustrates. Two patterns are quite clear in this panel.
First, there is a region of low oxygen concentration near the equator, centred about 500 m
below the surface. A similar pattern is seen in the Pacic as well, and in each case it results
from biological activity associated with equatorial upwelling. The second pattern is perhaps
more subtle, and to see it, look near the surface in the southern waters. The blue colour in
the diagram indicates high levels of oxygen. This increase is the result of a ux of oxygen
from the air to the sea, and so it may not be too surprising. But now for the intriguing thing.
Examine the top 1000 m near 45

S. Notice the contours that are bulging downward and


northward? These reveal that the surface waters of the southern ocean are subducting
and travelling northward. This is evidence of thermohaline overturning, caused by deep
convection in the waters near Antarctica. Now that you can see the pattern, look to the north
of the diagram, but this time look across the full depth. There, you can see that there is
high-oxygen water subducting from the waters near Greenland, and extending southward.
You can follow these contours all the way past the equator. This is a signal of thermohaline
circulation caused by deep convection in the Greenland Sea. As you can see, the Greenland
Sea convection creates deep watermasses, whereaas the Antarctic convection creates an
intermediate watermass.
This sort of inference about ows based on tracer patterns harkens back to W usts core
method. It is not terribly quantitative as just described, but the general idea can be extended
by inserting tracer elds into data-assimilative models (see Section 18.5).
17
For more on this, see section 15.1 for an experiment you can perform in your kitchen, to demonstrate salt
ngers.
4.14. OTHER OCEAN TRACERS 75
library(ncdf)
jpeg("osection.jpg", width=1000, height=750)
file <- open.ncdf("/data/hydrographic/WOA05nc/o00an1.nc")
#ftp://ftp.nodc.noaa.gov/pub/data.nodc/woa/WOA05nc/annual
oxygen <- get.var.ncdf(file,"o00an1")
depths <- get.var.ncdf(file,"depth")
lat <- get.var.ncdf(file, "lat")
lon <- get.var.ncdf(file, "lon")
lon.focus <- 30 # degrees west
lon.index <- 360 - lon.focus
col <- hsv(h=seq(0, 2/3, length.out=11), 0.9, 0.9)
oxygen.section <- oxygen[lon.index,,]
par(mar=c(4,4,3,1))
par(cex=1.5)
image(lat, depths, oxygen.section, col=col,
xlim=c(-90, 90), ylim=rev(range(depths)), zlim=c(0,8),
xlab="Latitude", ylab="Depth [m]", axes=FALSE,
main=sprintf("Oxygen [mL/L] at %.1f W",
360-lon[lon.index]))
contour(lat, depths, oxygen.section, levels=seq(0.5,10.5,1),
add=TRUE, drawlabels=FALSE)
contour(lat, depths, oxygen.section, levels=seq(0,10,1),
add=TRUE, labcex=2)
box()
axis(2)
axis(1, at=seq(-90, 90, 45))
abline(h = seq(6000,0,by=-1000), col="darkgray")
abline(v=seq(-90, 90, 45), col="darkgray")
Figure 4.28: Top: Annually-averaged variation of oxygen concentration in the Atlantic
ocean, in ml/l, at a latitude in which a north-south transect contains both the North and
South Atlantic basins. The dataset is from the 2005 version of the Levitus atlas [Collier
and Durack, 2006]. Bottom: R code to create the diagram. Exercise: plot the section at a
different location, or download another dataset (a nutrient, perhaps) and make a section plot
of its distribution.
76 CHAPTER 4. SEAWATER PROPERTIES
CHAPTER 5
Dynamics of Ocean Currents
T
HIS CHAPTER INTRODUCES THE BASIC PRINCIPLES OF FLUID MECHANICS. Much
of what youll read here is general in nature; it applies as much to the blood in your
veins as to ocean currents. Later chapters will build on this foundation, to discuss
the ocean more specically.
5.1. Concepts of Fluid Mechanics
5.1.1. Continuum Hypothesis
I have been drifting all this time
Living somewhere in my mind.
You and me c _2001 Jimmy Rankin
Fluid can be divided up conceptually into very small [parcel??]s without losing the
properties of the larger volume. Even though we know that uids actually consist of
molecules bouncing around, we blur a view a bit, ignoring these details, and instead continuum hypothesis
concentrate on bulk properties, e.g. the local [density??] will be dened as the local value
of mass per unit volume, calculated as the limiting value for a series of vanishingly small
volume elements centered at a point in the uid.
What is involved in our neglect of phenomena ocurring on atomic/molecular scales?
To answer this, note that 1mm
3
of water contains 10
19
molecules, which means that the
distance between molecules is less than 10
5
mm. Thats small, indeed
1
, smaller than any
ocean ows well consider in this course.
It is important to note that the continuum hypothesis doesnt say that the small scales
are unimportant. For example, the density of water at a given temperature and pressure
is controlled by the inter-molecular forces, occuring on these tiny scales. Similarly, the
viscosity of water is controlled by inter-molecular forces. The point is that we parameterize
these forces in a bulk way, by stating that the uid has a certain density or viscosity and
not trying to explain these values using rst principles Physics of molecular forces.
5.1.2. Newtons Second Law (F = ma)
Force = mass acceleration (5.1)
This equation, Newtons second law, says that you have to push things (apply a force)
to make them change the speed at which they move (accelerate). [acceleration??] is the
rate of change of velocity: a = du/dt, where u is the velocity. Think of push-starting a car
(changing the speed the speed fromu = 0 to some u > 0), or driving a car into a snowbank
(changing the speed fromu > 0 tou = 0); a force is required in each case.
Notation. The arrows drawn above the

F and a symbols indicate that these are
[vector??] quantities, i.e. quantities whose description involves both magnitude and
direction. For example, velocityu is a vector (involving speed and direction of travel),
but temperature T is not. (Sometimes, especially in British literature, vectors are
denoted with boldface notation, instead of with an arrow.)
You probably understand

F =ma intuitively, even if you dont understand it intellectually.
Knowing how much force it takes to accelerate a given object by a given amount is what
gives you the ability throw a baseball without looking immensely clutzy
2
. Unfortunately,
though, your intuition is probably based mainly on solid-body mechanics: the mechanics
describing the interactions between forces and solid objects like balls, chairs, etc. If you
were a whale or a canary, youd understand uid mechanics instead.
1
Think about that distance. Try comparing it to the width of something in everyday life, to get an idea of what
the distance means.
2
Come to think of it, by that criterion I dont understand

F = ma . . .
77
78 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
Newtons Second Law for Solid-body mechanics
The velocity of an object remains constant (at zero or any other value) unless an unbal-
anced force acts on the object. If there is an unbalanced force, then the acceleration a is
proportional to the ratio of the net force

F to the mass m. Note that the accelerationa is the
rate of change of the velocity with time. We denote velocity by u, so that acceleration is
dened bya = du/dt. Of course the velocity is just the derivative of the position: u = dx/dt.
If your position x is xed, then u = 0 (you have no velocity); by the same token if your
velocity is xed, thena = 0 (you have no acceleration).
We will write Newtons second law as:

F = ma (5.2)
This is the second best recognized equation in Physics, e = mc
2
being classier in the public
view, even though (5.2) plays a larger role in your everyday experience. Equation (5.2)
simply states that you have to apply forces to accelerate things, and that more force is
required for heavy things.
The only tricky thing is that its acceleration, not velocity, that matters. We all know
that things sit still (maintain zero velocity) unless pushed. Equation (5.2) states that moving
things keep going at the same rate unless forces are exerted: a moving hockey puck keeps
sliding down the ice at a constant speed in a constant direction unless some force (a stick, a
skate, or just the friction of the ice) is applied to it.
Physical Oceanographers seldom talk about acceleration explicitly. After all, who thinks
in terms of acceleration in their daily life? Its easier to think in terms of velocity. Since
acceleration is the time rate of change of velocity, we lose nothing by recasting Newtons
second law in terms of velocity:
du
dt
=

F
m
(5.3)
Newtons Second Law for Fluid mechanics
In uid mechanics, there is no direct analogue to the mass of the object. The mass of a
chosen chunk water depends on its size. Imagine writing (5.2) for the water in a bathtub.
Should we use the mass of all of the water in the tub (thus nding the gross response of the
whole tub to the oor collapsing), or should we instead focus on hand-sized chunks (thus
nding the response to your rubber ducky going all Moby Dick on your teeny weeny toy
boat)? Clearly the second view is better than the rst, if the goal is to understand the details
of the ow.
We take this process to the limit by thinking of the law being applied to innitesimally
small volumes of water. Imagine (5.3) applying to a small chunk of water. Denote its volume
V and its mass m. Then (5.3) is, in terms of force per unit volume,
du
dt
= Force per unit volume
V
m
(5.4)
Since V/m is 1/, where represents the density, this can be simplied to
du
dt
=
Force per unit volume

(5.5)
From now on, well usually work in terms of force per unit mass, so Newtons second
law will take the form
du
dt
=

F (5.6)
where well understand by convention that

F is the force per unit mass.
Equation (5.6) is called the [momentum-equation??], and it is about to become your
very good friend/enemy.
5.1. CONCEPTS OF FLUID MECHANICS 79
Figure 5.1: Flow in a narrowing pipe. Water entering at the left must speed up as it
encounters the constriction.
5.1.3. Eulerian and Lagrangian Notation
The acceleration dened by Newtons Second Law is the so-called Lagrangian accelera-
tion, the rate at which a given uid [parcel??] increases its velocity as it moves with the
uid ow. This is not the same acceleration you would calculate by time-differentiating the
velocity measured at a given xed point in space (the Eulerian acceleration)
3
Eulerian vs Lagrangian acceleration
All these places have their moments
In my life c _1996 John Lennon and Paul McCartney
I hate to question John and Paul, but how can a place have a moment? In your life, to
take up the phrase of the song, has a similar situation (place) seemed very different, years
later? The euler-lagrange discussion, the subject of this section, is about place and time.
Consider a quasi-one-dimensional situation - ow in a pipe (Figure 5.1). Focus on
a particle which is at some initial position x = x
0
at some inital time t = t
0
. Denote by
u = u(x, t) the current measured at an arbitrary location x at arbitrary time t. (This is the
current youd measure with a current meter placed at that location at that time.)
A short time t later the particle will have moved to the new location x
0
+ut. At this
new location, and at this new time, the velocity of the particle will be u(x
0
+ut, t
0
+t),
since this is the velocity of the uid ow there. From the perspective of the moving particle,
therefore, the change in the velocity du/dt is the limit as t 0 of the ratio
u(x
0
+ut, t
0
+t) u(x
0
, t
0
)
t
(5.7)
We call this acceleration of the particle the Lagrangian acceleration. By a Taylor expansion,
the limit yields
4
du
dt
=
u
t
+u
u
x
(5.8)
This equation states that: Lagrangian acceleration (the acceleration of a particle) is
the sum of local acceleration (rate of change of velocity at a xed point in the uid) plus
3
I recently heard a new story about Euler. Like hundreds of mathematicians over the last 300 years, Euler
had tried very hard to solve Fermats last theorem. Years before, Fermat had written about this seemingly simple
theorem, a sort of analogue to Pythagorass theorem but didnt prove it, claiming I have a truly marvelous
demonstration of this proposition, which this margin is too narrow to contain. Although one might verify in an
hour that the theorem seems to be true, it is not easy to prove. Euler was no slouch at mathematics (having set up
the background for calculus, among other accomplishments) but, try as he might, he could not crack this puzzle.
Finally he became so frustrated that he asked a friend to break into Fermats house, to see if Fermat might have left
behind a scrap of paper containing the unmarginal proof. This is but one of dozens of fascinating aspects to the
history of this theorem. For more on Fermats theorem, which was nally proved in 1994, see Singh, S. and K. A.
Ribet, Fermats last stand, Scientic American, November 1997. Although you wont be able to learn much about
the proof (if you had enough mathematics for that, you wouldnt be reading this page), youll learn a little about
the mathematician who nally proved it (in several hundreds of pages), how he worked in secret for a decade on
the problem, telling nobody except his wife. PS: he told his wife only on the honeymoon ...
4
It might be good for you to verify this for yourself. Basically, use the fact that the change in a function, u say,
is approximately given by t times u/t, for the variation over time (see section 2.4.1), plus a similar term for
variation in space. The spatial variation is xu/x, and since x is just t times the velocity u, the spatial term is
tuu/x. Now, the derivative is just the sum of these terms, divided by t, and taken in the limit of smaller and
smaller t. Doing this, noting that the t cancels, one is left with the indicated result.
80 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
advective acceleration (change of velocity achieved by being transported into a region with
a different ambient velocity).
Schematically, we may write
d
dt
=

t
+u

x
(5.9)
The Lagrangian d/dt term is the acceleration experienced by a given tagged water
[parcel??]
5
. The Eulerian /t term is the local acceleration at a point. Its what you get if
you put a current meter at a given point in space, measure the velocity of the uid streaming
past, and then differentiate with respect to time.
To understand the difference between the Lagrangian and Eulerian description, consider
the ow in a narrowing pipe. Suppose that the ow is steady (/t = 0) that is, a current
meter placed anywhere in the ow will measure a velocity u which doesnt change over
time. Think of a particular water [parcel??] entering the left-hand side of the pipe. As it
moves to the right, it enters a region where the pipe is narrower, so the ow is faster. Clearly,
then, the [parcel??] accelerates; du/dt ,= 0. Equation (5.8) tells us that the acceleration is
du/dt = uu/x.
Extension to three dimensions
Extending this idea to three dimensions we get
du
dt
=
u
t
+u
u
x
+v
u
y
+w
u
z
(5.10)
dv
dt
=
v
t
+u
v
x
+v
v
y
+w
v
z
(5.11)
dw
dt
=
w
t
+u
w
x
+v
w
y
+w
w
z
(5.12)
or, in [vector??] notation,
du
dt
=
u
t
+u u (5.13)
We call the form of the derivative above a [material-derivative??]. It is sometimes
denoted d/dt and sometimes D/Dt; the latter highlights that it is somewhat unusual in its
form.
We work with the Eulerian description of the oceanic ows when dealing with ob-
servations from current meters. Lagrangian descriptions of oceanic ows come from
measurements made by drift cards or bottles, bottom drifters, drogues, free drifting buoys,
etc. Analytical and numerical models are usually done in an Eulerian framework, but they
give us the elds u/x, etc, so we can easily convert to a Lagrangian framework.
5.1.4. Fluxes
One of the most practical issues in Oceanography is the ux of properties. For example,
in this town there is concern about the ux of sewage into the harbour. The ux of water
is the volume owing through unit area in unit time. Consider the schematic in Figure 5.2.
The volume transported through the area A in time t is uAt. The volume ux per unit
area, per unit time is therefore just the velocity u.
5
Another way to understand the [nonlinear??] acceleration terms like uu/x is to imagine a crowd walking
toward a slidewalk, such as youll nd in large airports. The crowd motion is like the uid motion; the motion of an
individual person is like the motion of an individual particle in the uid. Upstream of the slidewalk, individuals all
walk with a given (xed) speed, and on the slidewalk, individuals all walk with a (larger) xed speed. Measuring the
speed of walkers at any given location in space reveals that u/t = 0. This might seem to imply that individuals
do not accelerate, but obviously individuals are accelerating when they step onto the slidewalk. (The acceleration
is strong enough that some people are startled by it, especially when they step off after their ride.) How might we
represent this acceleration? Well, the essence is that the individual has moved from some region where the ow is
slow (the walking part) to another region where the ow is swift (the slidewalk part). We express a spatial variation
of velocity mathematically by partial derivatives such as u/x, u/y, etc, and we represent movement from
one region to another by the velocity compontents u, v, and w. Note that motion in the x direction will lead to a
dependence of acceleration on u/x, and that the change in velocity across some distance x will be xu/x.
Therefore the rate of change of velocity, found by dividing the above by the time t it takes to move that distance
x, is (x/t)u/x, which is just uu/x since the velocity u is by denition the ratio of distance x travelled in
some small time t. This reasoning applies to all directions equally, so we see that the full 3D acceleration of the
particle in the uid must be as in the text.
5.1. CONCEPTS OF FLUID MECHANICS 81
Figure 5.2: Flux into one side of a box-shaped uid element. Velocity of uid is u; area of
side of box is A.
Figure 5.3: Heuristic derivation of continuity equation, which states that water is neither
created nor destroyed. So, if the water is [incompressible??], the velocity gradients must
balance. The sides of the box have dimensions (x,y,z) as indicated. The mass ux in the
left side is u, while that out the right side is u+x
(u)
x
.
The mass ux is
uAt
At
= u (5.14)
Correspondingly, the salinity ux is Su and and the temperature ux is Tu.
5.1.5. Conservation laws
An important idea related to the uxes of properties is the concept of conservation. Many
of the properties of the uid are not involved in complex chemical or biological reactions,
so we should be able to account for the changes we observe in a volume of uid in terms of
the ux of the property into the volume.
We can have relatively simple conservation equations for salt, temperature, sand etc.
These just equate changes in a given volume to uxes into and out of the volume.
5.1.6. Conservation of mass
The basic conservation equation in uid mechanics is the conservation of mass. Consider
Figure 5.3, which shows a small volume xed in a particular location in space (ie, not
moving with the uid), with sides of length x, y, z. The rate of change of mass in this
volume is
dMass
dt
=xyz

t
(5.15)
Consider uxes in the x-direction. The mass ux into the left-hand side of the box is u.
The mass ux out of the right hand side of the box is u+x
(u)
x
. Therefore the difference,
the net gain of mass due to uxes in the x-direction, is
_
u
_
u+x
(u)
x
__
yz =xyz
(u)
x
(5.16)
Extending this using a similar argument for the other two directions and equating to the
left-hand-side of (5.15) yields

t
+
(u)
x
+
(v)
y
+
(w)
z
= 0 (5.17)
82 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
This is the equation for the conservation of mass.
In vector form this is

t
+ (u) = 0 (5.18)
This may also be written as

t
+u +u = 0 (5.19)
which lets us write
d
dt
=u, (5.20)
Note that these equations assume no sources or sinks of mass, and no diffusion of mass.
These effects are easily added to the equation, for applications in which they are relevant.
5.1.7. Incompressible ow: continuity equation
Derivation of continuity equation
For most ow problems, sea water can be regarded as being essentially [incompress-
ible??]
6
, so that [density??] is conserved for a given water [parcel??]. We express this
mathematically by writing
d
dt
= 0 (5.21)
Substituting this into (5.20) yields
u = 0 (5.22)
or, writing out the terms,
u
x
+
v
y
+
w
z
= 0 (5.23)
This equation is called the [continuity-equation??], and it is deeply important to an
understanding of Physical Oceanography. You need to nd a way to understand this equation
in your bones.
Example of use of continuity equation
Imagine a continental shelf 100 km wide and 100 m deep. Draw an x-axis at right angles
to the coast, with x = 0 at the coast. Draw a z-axis pointing upwards, with z = 0 at the
surface and z =100 m at the bottom of the ocean. Suppose that something (for example,
wind [stress??]) is causing an offshore ow of 0.1 ms
1
in the top 20 m at x = 100km. At
the coast there can be no ow directed onshore or offshore, since the coast is a solid wall, so
that u = 0 at x = 0. Similarly, there can be no upward ow at the ocean surface (trust me on
thisthe gravity force is so strong that it tends to inhibit surface movements on large scales
in the ocean), so that w = 0 at z = 0.
Youll guess immediately that the only way to conserve water is for there to be an upward
velocity at the bottom of this upper layer, that is, at z = 20 m. Our continuity equation
tells us what this velocity must be. We have
u
x
+
w
z
= 0 (5.24)
We can estimate u/x as simply the change in u over the given range of x. This is
0.1 ms
1
divided by 100 km, or u/x = 110
6
s
1
. Substituting this into the previous
equation, and solving for w/z = u/x, gives w/z = 110
6
s
1
. This is just
the variation in w over the range of variation, and we know that w = 0 at the surface, so this
implies that w at the depth 20 m must be 20m times w/z, or w = 210
5
m/s. Since a
day is approximately 10
5
s, this is a speed of about 1 m per day upward at z = 20 m. (This
value for w, and the other numbers in this sample calculation, are useful to have in mind,
when thinking of wind-induced upwelling on continental shelves, e.g. for rapid calculations
of nutrient uxes to the mixed layer.)
6
The exception is for sound waves. Strictly speaking, of course, no uid is incompressible. But your intuition
should tell you that water is very nearly incompressible, given the forces it is likely to feel in the ocean: you can
easily squeeze air in a bicycle pump, but the stongest among you could not measurably squeeze water in a pump.
5.2. FORCES ON A FLUID ELEMENT 83
5.1.8. Conservation of S, T, etc
We can create similar laws for conservation of S, T etc. For salinity we get
(S)
t
+
(uS)
x
+
(vS)
y
+
(wS)
z
= 0 (5.25)
or, using conservation of mass
S
t
+u
S
x
+v
S
y
+w
S
z
= 0 (5.26)
or, in vector notation,
S
t
+u S = 0 (5.27)
This states that S is constant following a uid [parcel??]: dS/dt = 0. This appears to
be a very simple result but in practice it is hard to measure u and S accurately enough to
estimate the size of these terms.
Note that this equation must be modied in circumstances in which salinity (or tempera-
ture, or other scalar) is added by sources or removed by sinks, or in which the scalar diffuses.
For example, heat is provided by solar forcing, and salinity is provided by evaporation
(which takes away water, leaving salt behind). In most cases of large-scale ow, molecular
diffusive uxes of scalars are unimportant, but an analogy, the so-called eddy diffusion
inserts a term into the time-averaged conservation equation that looks just like a diffusion
termsee next section.
5.1.9. Eddy Fluxes
Well I went to bed in Memphis
And I woke up in Hollywood.
Steve McQueen c _2002 Sheryl Crow
Commonly, we regard u, S, etc, as having a mean values, u, S plus uctuations u
/
, S
/
. (By
denition, u
/
= S
/
= 0.) If the ux of S is then averaged over some time interval
uS = (u+u
/
)(S+S
/
) (5.28)
then averages of terms like uS
/
vanish, leaving
uS = uS+u
/
S
/
(5.29)
Thus, the total [advection??] is the sum of the [advection??] of mean S by the mean
current and the correlation between perturbation S
/
and perturbation currents u
/
. The
correlation part is the eddy ux.
5.2. Forces on a Fluid Element
Recall that the basic [momentum-equation??] is
7
du
dt
=
1

F (5.30)
where

F is the net force on the uid [parcel??].
The important forces for dynamical oceanography are pressure gradients, gravity, the
Coriolis force, and friction; these are discussed in the next several sections.
84 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
Figure 5.4: Pressure forces exerted on opposite sides of a box-shaped uid element. The
pressure force is inwards, so that the rightward force exerted on the left hand side is p
times the area of the left side, while the rightward force exerted on the right hand side is
PxP/x; thus the net rightward force is xP/x. Note very carefully that the
resultant force is related to the gradient of pressure, P/x, not to the pressure itself.
5.2.1. Pressure gradients
The pressure at a point in the uid acts equally in all directions. Pressure is a scalar. To
understand how pressure can exert a net, unbalanced, force, consider a small uid element
in the shape of a box with sides x, y and z (Figure 5.4)
The net force in the x direction is the difference between the rightward (positive) pressure
force P exerted on the left-hand side of the cube and the negative force P+x
P
x
exerted on
the right-hand side of the cube.
Acceleration in the x-direction is du/dt so
du
dt
=
1

P
x
(5.31)
The same procedure holds for the other two directions, so in vector form, we have
du
dt
=
1

P (5.32)
5.2.2. Gravity
The downward force due to gravity is mg for a solid body of mass m. For our little box
of uid, this is gxyz. So, writing force per unit volume, and noting that gravity affects
only the z-direction balance, we modify the z-component of the [momentum-equation??]
to
dw
dt
=
1

P
z
g (5.33)
5.2.3. Hydrostatic equation
A motionless uid has w = 0, so that clearly dw/dt = 0, so that in turn the last equation
becomes
P
z
=g (5.34)
the so-called hydrostatic equation. This also holds approximately, provided that
w
t
+u
w
x
+v
w
y
+w
w
z
<< g (5.35)
In this case, if the density is constant (or approximately constant) then pressure is given
by
P = P
atm
+g( z) (5.36)
7
I write the reciprocal of density as a multiplier in this way so that the similarity of this equations to those that
follow will be more evident.
5.2. FORCES ON A FLUID ELEMENT 85
where is the surface displacement (upwards). (Recall again that z < 0.)
The horizontal gradient of pressure at any given depth z is
P
x
= g

x
(5.37)
Well see later how this formula relates surface slopes to forces within the uid, with
implications for things ranging from surface waves with lengthscales O(10)m to currents
with lengthscales O(100)km.
Exercise 5.1 Flow on a tilted table
a
A large, at table is oriented to be perfectly level. A high edge runs along the
sides of the table, so that it can contain water. Water is poured onto the table and
allowed to come to rest. Suddenly, the table is tilted to some small angle from
level. (a) Discuss what happens, physically. (b) Quantify your answer.
a
Answer in chapter 21.
5.2.4. Coriolis Effect
Wheres the sense
in staying right?
Survive c _1999 David Bowie & Reeves Gabrels
Why does the wind blow roughly parallel to the isobars on a weather map instead of
directly from regions of high pressure to regions of low pressure? The reason is the apparent
force associated with motion in a rotating frame of reference. However our day to day
experience does not really give us any intuition about the Coriolis effect.
The basic reason for the existence of a Coriolis force is that Newtons laws are valid
for an inertial frame of reference (i.e. one in which the x, y, z coordinates do not accelerate).
However we view the motion of the ocean and atmosphere relative to a coordinate system
xed to the earth. Since the earth is rotating, the coordinate system is accelerating.
We take the centre of the Earth as the origin of a coordinate system rotating with the
Earth (see Figure 5.5). The Earth spins at a rate of radians per second; this is here denoted
by a rotation vector

, with magnitude , aligned due North.
The [parcel??], viewed from a non- rotating, non-accelerating frame (inertial, denoted
by subscript i), has position vector

R
i
= (x
i
, y
i
, z
i
) (5.38)
The position in the rotating frame is

R
r
= (x
r
, y
r
, z
r
) (5.39)
Visualizing rotating coordinates. imagine a merry-go-round with rotating coordi-
nates (x
r
, y
r
, z
r
) painted on it, with origin at the centre. The inertial coordinates (x
i
, y
i
, z
i
)
are painted on the ground adjacent. We want to write true inertial-frame acceleration
d
2

R
i
/dt
2
in terms of the coordinates as measured by d
2

R
r
/dt
2
in the rotating frame.
Consider rst the case where a [parcel??] of uid at point

P is xed to the earth, so
that an observer on earth measures no motion (d

R
r
/dt = 0). Someone in the inertial frame
of reference will see point

P describe a circle of radius r =[

R
r
[ cos in time period 2/.
Here, is latitude (i.e., the angle between the

R
r
vector and the equatorial plane) and r is
the distance of P from the axis of rotation. It is important to note that does not correspond
to a solar day (24 hours), but rather to a siderial day (23.93447 hours). The instantaneous
86 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
Figure 5.5: Derivation of Coriolis equation. Three vectors are indicated by the thick
arrowed lines: (1)

, the vector pointing up from the North Pole, representing the rotation of
the earth; (2)

R
r
, a coordinate pointing from the centre of the Earth to a point located on the
Earths surface, measured in a reference system which rotates with the earth (thus,

R
r
would
not vary with time, for a given geographical location, but would if the point were to move
relative to Earths surface); and (3) d

R
i
/dt, which is the rate of change of the vector pointing
to the location, measured in an inertial coordinate system not rotating with the Earth. Thus,
the rotation of the Earth alone gives d

R
i
/dt =

R
r
, and when the motion of the point
with respect to the Earths surface is added in, we get d

R
i
/dt = d

R
r
/dt +

R
r
; taking
derivatives, the acceleration is thus d
2

R
i
/dt
2
= d
2

R
r
/dt
2
+2

R
r
/dt +

R
r
. The
second term, proportional to the rate of rotation of the Earth times the velocity measured in
the coordinate system rotating with the Earth, is the Coriolis term.
velocity of the uid [parcel??] is therefore [R
r
[ cos, directed in the azimuthal (East)
direction, as shown in the view from above.
Now, given the denition of a vector cross products (see section 2.9), the velocity of
point P may be written
d

R
i
dt
=

R
r
(5.40)
If the uid [parcel??] is moving at velocity d

R
r
/dt relative to the Earth, we must add that
velocity to the apparent velocity due to the rotation:
d

R
i
dt
=
d

R
r
dt
+

R
r
(5.41)
We simply apply this rule twice to get the true acceleration of a uid [parcel??] as
measured from the rotating earth:
d
2

R
i
dt
2
=
_
d
dt
+

__
d
dt
+

R
r
(5.42)
=
d
2

R
r
dt
2
+2

R
r
dt
+

R
r
(5.43)
5.2. FORCES ON A FLUID ELEMENT 87
The terms on the right-hand-side are, in order:
(1) Acceleration as measured in the rotating frame.
(2) Coriolis acceleration 2

u.
(3) Centripetal acceleration

(

R
r
). This is a steady acceleration which causes
no motion. It makes the Earth bulge a bit, and affects the local value of gravity.
Test case 1: consider a point at the pole, where

and

R
r
are parallel, so that the
cross product is zero that there is no centripetal acceleration because the point isnt
circling. Test case 2: consider a point at the equator, where

and

R
r
are at right
angles, so that their cross product is maximal and of length equal to the product
of the length of

and

R
r
, and

(

R
r
) reduces to a vector of length equal to
the product of
2
and the earth radius. In each case (and generally) the length of

R
r
) equals the radius of the latidude-circle intersecting the point

R
r
, times
the square of

.
To simplify the Coriolis term further, lets calculate the acceleration in a rotating frame
which has z aligned to the local up:
d
2

R
r
dt
2
=

i
du
dt
+

j
dv
dt
+

k
dw
dt
(5.44)
where

i,

j and

k are unit vectors directed East, North and upwards, and u, v and w are the
velocity components in these directions. Let the latitude be . In this coordinate system, the
Earths rotation vector is

jcos +

ksin (5.45)
so that the Coriolis acceleration can be written
2

R
r
dt
=

i2vsin +

j2usin +

i2wcos

k2ucos (5.46)
The rst two terms are the usual Coriolis terms. The third term is much smaller as w is
usually much smaller than the horizontal velocities. The fourth term is not important directly
for oceanography, but has to be considered when making accurate gravity measurements
from a ship or aircraft.
The value of is
= 2/siderial day = 7.2910
5
radian/s (5.47)
To avoid writing the sin termrepeatedly, we dene the so-called [Coriolis-parameter??]
f = 2sin (5.48)
which takes on values
f =1.4610
4
s
1
sin in general
110
4
s
1
at midlatitudes
(5.49)
Looking at the horizontal (

i and

j) components in (5.46) we have accelerations


du
dt
f v =
F
x

(5.50)
dv
dt
+ f u =
F
y

(5.51)
A water [parcel??] moving north appears to be forced to the +u direction, eastwards. A
[parcel??] moving eastwards appears to be forced to the south. In each case the force tends
to make the particle veer to the right in the Northern hemisphere.
The rules of Coriolis forces are:
88 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
Figure 5.6: Denition sketch of mechanism of molecular friction.
In the Northern hemisphere, moving objects continually veer towards the right of their
path of motion. This veering is stronger near the poles, and diminishes to zero at the
Equator.
In the Southern hemisphere things are exactly reversed. The veering is to the left.
This is explained by the equations, since the latitude is negative so f is negative.
The Coriolis force is zero at the equator, since sin = 0 there.
Example of displacements due to the Coriolis force: An ocean current at latitude
= 45, moving to the North at v = 1 m/s, accelerates to the East at 10
4
m/s
2
, so that after
an hour it will achieve an Eastward speed of about 0.36 m/s. (In fact, by this time v will
have changed a little, because the [parcel??] will begin to describe an inertial circle.)
5.2.5. Molecular friction
In general, friction tends to oppose motion. As you slide on a patch of ice, the friction
always acts to slow you down, not to speed you up (lucky about that). For uid motion,
friction acts to make all gradients of velocity linear through space. Imagine a deck of playing
cards arranged in a slab, sitting on a table. If you put your hand on the top card and slide
it, it will tend to accelerate the card just below it through friction. Similarly, this second
card will exert a frictional force on the card below. The friction force between two cards is
related to the rate of sliding between the cards. Thus a way to have zero acceleration (ie,
to have the individual cards moving at constant speeds) is to have the friction force from
the card above equal and opposite to that from the card below. That is, steady state can be
achieved when the card velocity is a linear function of distance through the deck of cards.
We would write this linear function by saying that u/z = 0 (since the solution to this
differential equation is u = constant, as you can verify easily).
Now, for non-steady state, it is the change of frictional [stress??] across a given card that
matters. Since the stress is proportional to terms like u/z, this means that the acceleration
should be proportional to terms like
2
u/x
2
,
2
u/y
2
and
2
u/z
2
. The proportionality
constant is called the [kinematic-viscosity??], denoted (units m
2
s
1
). See Figure 5.6.
Consider Figure ??. Suppose that Brownian motion exhanges uid elements between
A (where ow is relatively fast) and B (where ow is relatively slow). The swiftly-moving
element from B will speed up the ow at A. Similarly, ow at B will be slowed down.
Thus, the swift ow slows down, and the slow ow speeds up. The net effect is that spatial
variations in speed are reduced. (You may nd it helpful to think of an analogy with
temperature, in which random movements might exchange parcels of warm and cool uid,
so that temperature becomes more uniform in space.)
Experiments on non-turbulent shear ows ows suggests that the [shear-stress??] is
proportional to the shear. For example, if the ow is u = u(z), this would be expressed as
=
u
z
(5.52)
5.2. FORCES ON A FLUID ELEMENT 89
Figure 5.7: Turbulent friction. Compare with the explanation of molecular friction; the
main difference is that the exchange is not through (weak) Brownian motion, but rather
through (strong) turbulent motion.
where is called the [dynamic-viscosity??] . Another variant form of viscosity is called
the kinematic viscosity , dened by
=

(5.53)
For water, is a function of temperature. For most calculations it is sufciently accurate
to use the constant value 1.410
6
m
2
/s.
5.2.6. Turbulent friction
In a turbulent ow [momentum??] is transferred by the random motion in the uid.
Figure (5.7) draws a parallel to the case of molecular friction.
The [stress??] is
=u
/
w
/
(5.54)
where u
/
and w
/
are the turbulent uctuations and the overbar represents a suitable time
average. We call u
/
w
/
the Reynolds Stress. For modelling studies, we often assume that
this stress is proportional to the [shear??], in direct analogue to molecular friction. This is
mostly a bad assumption, but it is nevertheless made commonly, simply because it permits
the construction of approximate theories for how uid ows work
8
.
This assumption leads us to dene an eddy viscosity
e
, analogous to the molecular
[kinematic-viscosity??], , via

e
u
z
=u
/
w
/
(5.55)
where u is the average ow pattern dened over the same timescales as the turbulent averages.
Equation (5.55) isnt very useful unless
e
is a reasonably simple function of the ow eld.
The turbulent viscosity
e
then enters the basic equations in just the same way as .
Typically
e
so that the turbulent effects far outweigh the molecular effects. This is
simply because turbulence exchanges uid elements much more vigorously than Brownian
motion does.
Ive used the notation
e
to highlight the analogy to but in the literature youll often
see this quantity denoted A. Also, since mixing is greatly inhibited in the vertical direction
and not in the horizontal directions, youll often see A
V
to indicate the eddy viscosity in the
vertical direction and A
H
for the eddy viscosity in the horizontal direction. Corresponding
to these eddy viscosities are eddy diffusivities, normally denoted K
V
and K
H
.
8
I hope the merits of using approximate theories have been emphasized in previous science courses. In particular,
you should be aware that the making of an assumption in the course of creating a theory does not mean that the
theoretician believes the assumption. If you wanted to nd out how long it takes a cat to fall 6 stories, you might
assume the cat is a sphere, and use a formula for a falling sphere. Then you might assume that it is in the shape
of a cylinder, and use that formula. Then you might reason that the true solution may lie between these two,
demonstrably untrue, solutions.
90 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
Horizontal eddy viscosities used in modelling studies, denoted A
H
, are in the range
10
2
m
2
/s to 10
4
m
2
/s, while vertical viscosities A
V
are in the range 10
5
m
2
/s to 10
2
m
2
/s.
These ranges are huge, and this is partly because the ocean is so very different from place
to place and time to time, and partly because A
V
and A
H
are difcult to measure, so we
are uncertain of their values. Indeed, the values of A
V
and A
H
used in modelling studies
often come from tuning to get the large-scale solutions to look like the observations, and
have nothing to do with actual measurements of the [mixing??] itself. This is a major, but
necessary, aw in all numerical models.
5.3. Basic Equations of Motion
5.3.1. The Equations of Motion
We can combine the various forcing terms weve just examined into a set of basic
equations, the [momentum-equation??] set, describing the large scale motion of a uid on
a rotating sphere.
In the x-direction (east)
u
t
+u
u
x
+v
u
y
+w
u
z
f v =
1

P
x
+friction (5.56)
similarly for the y direction (north)
v
t
+u
v
x
+v
v
y
+w
v
z
+ f u =
1

P
y
+friction (5.57)
and in the vertical
w
t
+u
w
x
+v
w
y
+w
w
z
=g
1

P
z
+friction (5.58)
For the friction term, we have to take into account the difference between horizontal
diffusion and vertical diffusion of [momentum??] and replace the general term
e
by A
h
for the horizontal eddy viscosity and A
V
for the vertical. The friction term is then, for u

x
_
A
H
u
x
_
+

y
_
A
H
u
y
_
+

z
_
A
V
u
z
_
+
2
u (5.59)
In addition to the [momentum-equation??]s we have the [continuity-equation??] for
[incompressible??] ow,
u
x
+
v
y
+
w
z
= 0 (5.60)
Exercise 5.2 Tidal velocities in harbour.
a
The tidal range in Halifax harbour is roughly 2 m (thats the difference in height
between high tide and low tide). What tidal velocities should this generate? (Hint:
you need to think about areas.)
a
Answer in chapter 21.
5.3.2. Scaling the Equations of motion
Not all of the terms in (5.56), (5.57), and(5.58), are equally important in all uid systems.
In some circumstances, for example, pressure forces might balance friction forces. An
example of this is ow in the capillaries of your body. The ow is very ordered, (laminar)
because friction tends to smooth ows out. In other circumstances pressure forces might
balance the inertial terms (that is, u/t and uu/x). This leads to more disordered ows,
or even to turbulent ows. An example of this is a white-water river.
In uid mechanics, we often try to classify ows roughly into such categories. This is
done with dimensionless numbers, which represent approximate ratios of terms in the
[momentum-equation??]. The process of determining relevant dimensionless numbers
typically involves doing something called [scaling??] the equations, which is illustrated in
the next few sections.
5.3. BASIC EQUATIONS OF MOTION 91
5.3.3. The Reynolds Number (advection vs friction)
For the x component of the [momentum-equation??] we have
u
t
+u
u
x
+v
u
y
+w
u
z
= f v
1

P
x
+
2
u (5.61)
where

2
u =

2
u
x
2
+

2
u
y
2
+

2
u
z
2
(5.62)
Note that were now looking at just molecular friction, so A
V
and A
H
dont enter into the
friction term. Lets scale these equations. Suppose that a typical value for the velocity,
to within a factor of 10 or so, is U. Well call this a velocity scale. Further, suppose that
velocity varies by this amount over a spatial length scale L. Then we can derive the timescale
via the ratio L/U. Lets dene dimensionless velocity, space and time variables by dividing
the numerical values by these scales:
(u

, v

, w

) = (u/U, v/U, w/U) (5.63)


(x

, y

, z

) = (x/L, y/L, z/L) (5.64)


t

=Ut/L (5.65)
Substituting these into Equation 5.61 yields
U
2
L
u

t
+
U
2
L
u

x
+
U
2
L
v

y
+
U
2
L
w

z
=
1
L
P
x

+
U
L
2

2
u

(5.66)
so that the ratio between the [advection??] terms and the viscous term is
uu/x

2
u

U U/L
U/L
2
=
UL

(5.67)
This last quantity is so important that we give it a name, the [Reynolds-number??]
Re =
UL

(5.68)
Without solving for the complete ow eld, we can use the Reynolds number to get a
qualitative picture of the ow. If Re1, then friction will be important to the ow, and one
may expect simple ows (sometimes, and probably unhelpfully, called laminar ows).
On the other hand, if Re1 then friction will not be important to the ow, and the motion
will be more complicated, e.g. unsteady or even turbulent. There are no cut-and-dried
rules on ow characterization by Re, however. This is partly because so many ows of
interest involve intermediate Re values, and partly because many situations involve turbulent
regions and nonturbulent regions, together
9
. Figure 5.8 illustrates some ows at a range of
intermediate values of Re.
Biological implications. You might enjoy reading Steven Vogels article in Physics
Today (Exposing lifes limits with dimensionless numbers, November 1998, pages
22-27), about applications of dimensionless numbers and uid mechanics to biology.
For a more specic treatment of life at low [Reynolds-number??], see Purcell [1977].
Non-biologists tend to enjoy these articles every bit as much as biologists enjoy them.
9
Stirring honey with a spoon is unlikely to cause turbulence. Stirring cream into coffee might . . . but not just
along the edge of the spoon, where the ow is nessarily limited in its direction
92 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
Figure 5.8: Streamlines of ow in the lee of a sphere, for various values of Re. Note that
for Re less than 20 or 30, the downstream motion is reminiscent of upstream motion, but at
greater Re values, eddies start to develop in the lee of the sphere. At still higher Re, waves
will form behind the sphere, and at still higher values, an unsteady pattern of waves and
eddies will form. At very high Re, the lee region experiences turbulence. The source for
this image is Batchelor [1967] Figure 4.12.8; see that book for additional illustrations, and
see various websites for movies of such ows.
5.4. GEOSTROPHIC FLOW 93
5.3.4. The Rossby Number (advection vs Coriolis)
Suppose u, v have typical magnitude U and vary over a characteristic distance L, while w
has magnitude W and varies over vertical distance H.
Substituting these into the continuity equation
u
x
+
v
y
+
w
z
= 0 (5.69)
yields the scale equation
U
L
+
U
L
+
W
H
0 (5.70)
in which represents equality to within an order of magnitude. This yields a scale relation-
ship
U
L

W
H
(5.71)
(after ignoring signs and a factor of 2 in the U/L terms).
Now, consider the v [momentum-equation??]:
v
t
+u
v
x
+v
v
y
+w
v
z
+ f u =
1

P
y
+friction (5.72)
which suggests a scale relationship
U
2
L
+
U
2
L
+
U
2
L
+
U
2
L
+ f U . . . (5.73)
(where the time scale for the rst term has been scaled with an advective timescale L/U, and
(5.71 has been used to eliminate W). Thus, the ratio of the advective terms to the Coriolis
term can be wrtiten
[uu/x +vu/y +wu/z[
[ f v[
=
U
f L
(5.74)
This is called the [Rossby-number??], dened by
Ro =
U
f L
(5.75)
Without solving for the details of the ow, we can use the Rossby number to qualitatively
describe the ow. For small values of Ro, ow will be affected greatly by the Coriolis effect.
(This may yield [geostrophic??] ow, if some further conditions about friction are met.)
By contrast, if Ro is large (say, greater than 0.5 or so), then the Coriolis effect will not be
too important, and ow will be non-[geostrophic??].
As an example, if a region of the ocean is described by values U 0.1m/s, f = 10
4
s
1
,
and L = 10
5
m, we calculate Ro10
2
which implies that Coriolis forces outweigh advec-
tive forces. In practice, this means that Coriolis accelerations balance pressure gradients.
This state is called [geostrophy??], and ow in such a state is said to be [geostrophic??].
5.4. Geostrophic Flow
For steady ow with small frictional effects
10
and large Coriolis effects (small Ro), the
(x,y) [momentum-equation??] set reduces to
f v =
1

P
x
f u =
1

P
y
(5.76)
To understand this, consider the case where the pressure gradient is in the +x direction
(Figure 5.9). That is, pressure is independent of y, but increases as x increases.
94 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
Figure 5.9: Isobars (contours of equal pressure) in a simple pressure eld in the northern
hemisphere. See corresponding geostropic velocity eld in Figure 5.10.
Figure 5.10: Velocity in geostrophic balance with pressure eld of Figure 5.9.
Figure 5.11: Flow on a weather map. Fluid circles around high and low pressure zones,
with the higher pressure to the right, in the Northern hemisphere. (Figure scanned from
[Pedlosky, 1979, gure 1.2.1a].)
5.4. GEOSTROPHIC FLOW 95
Then P/y =0 and P/x >0 so that (5.76) tells us that f u =0 and that f v >0. Thus,
the ow is to the North [in the Northern hemisphere where f > 0]
This expains the ow you see on weather maps, e.g. Figure 5.11.
Instead of owing from pressure highs to pressure lows, surface winds circle around
the highs and lows. In the Northern hemisphere, they circle clockwise around highs and
counterclockwise around lows. (An easy way to remember this is that ows trace along
lines of constant pressure, keeping the high pressure side to their right.)
The hydrostatic condition gives the pressure by
P(z) =
_

z
g dz (5.77)
where is the surface elevation about the plane it would occupy if there were no motion. In
most cases of interest the density variations over the depth range are minimal, in the sense
that (z) , overbar denoting depth average, so this integral can be written
P(z) g( z) (5.78)
Consider ow in a strait aligned East-West, so y = 0 is the South shore and y =W is the
North shore. At the surface (z = 0)
f u =
1

P
y
(5.79)
= g

y
(5.80)
so the difference between uid level at the North side of the strait and the South side is
=
f
g
_
W
0
udy (5.81)
where the integral is across the strait.
If we measure from tide gauges and u from current meters, we can calibrate the
measurement as a measure of ow through the strait. This is not as easy as it seems,
because its hard to measure small values of . For example, suppose the strait is 1 km
wide and the ow is 0.1 m/s to the East. Then North side will be just f uW/g 1 mm lower
than the South side.
Inferring a velocity eld from pressure/density elds is an example of a so-called [diag-
nostic??] calculation. By contrast, a [prognostic??] velocity calculation makes predictions
based on differential equations for velocity variations (i.e. retaining terms such as u/t
and uu/x).
Exercise 5.3 Geostrophic Flow
a
A geostrophic ow runs along Nova Scotia, from Cape Breton towards Yarmouth.
Suppose that it has velocity of roughly 10 cm/s and is 30 km wide. Consider the
deformation in sealevel associated with the ow. The sealevel at the coast will be
either higher or lower than the level that would exist without the current. (a) Is it
higher or lower? (b) Estimate the change of sealevel using the geostrophic and
hydrostatic approximations.
a
Answer in chapter 21.
10
We compare the scale of the friction and coriolis effects with the Ekman number, Ek = /( f L
2
), which is the
ratio of the scales of the friction and coriolis terms, with L being the lengthscale of velocity variation.
96 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
5.4.1. Thermal Wind
Reference: [Cushman-Roisin, 1994, section 13-1].
If we do not have good measurements of currents, elevation or pressure, but we do have
good measurements of density, may we still calculate currents? The answer is that we can
calculate current [shear??]
11
through the so-called thermal wind equations, and that if
we can infer (somehow) a constant of integration, then we can calculate velocities as well!
Differentiating (5.76) with respect to z, and exchanging the order of the x and z derivatives
on the right-hand side, we get


z
( f v) =

x
P
z

z
( f u) =

y
P
z
(5.82)
If the ow is hydrostatic, then P/z =g, and of course f is not a function of depth, so
we get
(v)
z
=
g
f

x
(u)
z
=
g
f

y
(5.83)
The above pair make up the [thermal-wind??] equations. They permit the calculation of
velocities in terms of density elds.
Another form, simpler but more restricted, may be derived if the so-called Boussinesq
approximation. This approximation states the conditions under which it is permissible to
neglect spatial derivatives of density. Consider
(v)
z
=
g
f

x
which yields
v

z
+
v
z
=
g
f

x
.
When may v/z be neglected, compared with v/z? Consider the ratio [v/z[/[ v/z[.
This may be rewritten as [g/ /z[/[g/v v/z[, which is of order N
2
H/g, where N is
the buoyancy frequency (section 4.9) and H v/(v/z) is the vertical scale over which
velocity v varies. Using typical values N 10
2
s
1
, H 1000 m yields N
2
H/g 10
2
,
i.e. the v/z term can be expected to be roughly 2 orders of magnitude smaller than
the v/z term. To neglect this small term is to make the boussinesq approximation is
employed
12
. Then we may write the thermal wind equations approximately as
v
z
=
g
f

x
u
z
=
g
f

y
(5.84)
Much of what we know of the currents in the ocean has come from the use of these
thermal-wind equations, in either the rst form above (5.83) or in the approximated form
(5.84). The procedure is to measure (z) at a sequence of stations, and then to map these
out, using some form of contouring (by computer now, by hand years ago) to yield (x, y, z),
from which the partial derivatives of may be inferred, and thus the values of u/z and
v/z. The nal step is to integrate these derivatives, making some assumption about the
constant of integration; see discussion in the next section.
5.4. GEOSTROPHIC FLOW 97
Figure 5.12: Denition sketch for one-layer ow, with the (level) ocean bottom at z = 0 and
free surface (interface to atmosphere) at z = (x, y). The density is uniform throughout
the layer, while the velocity (u, v) may vary with (x, y).
5.4.2. One-Layer Flow
Consider one-layer ow in the 2-D (x, z) plane (Figure 5.12). Let be the depth of the
free surface and the density, homogeneous everywhere, be . Within this layer, we have
f v =
1

P
x
f u =
1

P
y
(5.85)
under the geostrophic and boussinesq approximations, and with the hydrostatic approxima-
tion this becomes
v =
g
f

x
u =
g
f

y
(5.86)
Thus, all thats needed to calculate the velocity eld is to know the shape = (x, y) of the
ocean surface. Well, thats an all that we dont often have, since it is difcult to measure
the shape of the ocean surface. There are, however, some cases in which we can do this:
1. Time-varying ows in enclosed coastal regions, e.g. in straits that are wide enough to
be geostrophic, may be inferred from time-varying measurements of sealevel made
on the coast.
2. Satellite measurements of ocean-surface tilts may soon be useful in such calculations,
although at the present time this is in the testing stages. (Its been in that stage a long
time large signals, such as the tilt of the Gulf Stream, are easy to infer; smaller tilts
are harder. The difculty is partly one of calibration of the satellite signal, and of
removal of affect from fog, sea-state, etc.)
5.4.3. Two-Layer Flow (Margules Equations)
Consider two-layer geostrophic ow in the 2-D (x, z) plane (Figure 5.13). Let h be the
height of the interface between two layers of density
1
(upper layer) and
2
(lower layer),
respectively, and be the height of the sea surface.
Within the upper layer, the equations of the last section apply:
u
1
=
g
f

y
v
1
=
g
f

x
(5.87)
11
The shears in this case are vertical derivatives of horizontal currents.
12
Exercise: Think of a ow for which the boussinesq approximation would be poor.
98 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
Figure 5.13: Denition sketch for two-layer ow, with the (level) ocean bottom at z = 0, a
free surface at z = (x, y) and an interface, between the two layers at z = h(x, y). Subscripts
1 and 2 denote values of quantities in the upper and lower layer, respectively.
For the lower layer, we must start again, following the reasoning of the one-layer case.
The key point is that the pressure in the lower layer is given by
P
2
= g
2
h+g
1
( h) (5.88)
(see Figure 5.13) if the ow is hydrostatic, and therefore
f u
2
=
1

2
P
2
y
f v
2
=
1

2
P
2
x
(5.89)
can be written
u
2
=
g
f
_

2
h
y
+

1

y
_
v
2
=
g
f
_

2
h
x
+

1

x
_ (5.90)
Let =
2

1
(a positive number) represent the density contrast between the layers
and
0
represent the average of the two layer densities. In many cases we can make the
assumption
13
that the relative density difference between the layers is small ( <<
0
),
yielding
u
2
=
g
f
_

0
h
y
+

y
_
v
2
=
g
f
_

0
h
x
+

x
_ (5.91)
Thus, calculating either (u
1
, v
1
) or (u
2
, v
2
) requires knowledge of derivatives of =
(x, y). This is difcult to measure. However, if we are satised with just knowing the
difference in the velocity components between the two layers, we may subtract to get
u
1
u
2
=
g
f

0
h
y
v
1
v
2
=
g
f

0
h
x
(5.92)
This is [margules-equation??] for layered ow. It yields differences in current speeds in
the two layers, given the relative density difference in the layers and the slope of the interface.
13
This is a type of boussinesq approximation, in which terms of order /
0
are ignored when added to terms
of order 1. As an exercise, derive a version of (5.91) that does not use this approximation, and then compare
the difference between the two equations, in the context of the measurement uncertainties in other terms in the
equation.
5.4. GEOSTROPHIC FLOW 99
You may think of this as a discrete (nite-difference) analogue of the (continuous-derivative)
thermal wind equations.
Special case If the lower layer is motionless (or stagnant), so that u
2
= v
2
= 0, we get
u
1
=
g
f

0
h
y
v
1
=
g
f

0
h
x
(5.93)
and these equations are said to describe 1
1
2
layer ow.
Aside: the quantity g/
0
appears often in such equations, and it is given a special
name: the reduced gravity. It is sometimes also called [buoyancy??].
5.4.4. Reference velocities (level of no motion)
A thread running through the previous few sections is that we are able to diagnose
vertical changes in ows, but that we cannot diagnose ows themselves without making
some assumptions. The difculty is simply that measurements of the spatial gradients of
total pressure are difcult to make, whereas gradients in density are easy to measure.
Thus, in all calculations of the thermal wind ilk, one must either be satised with
knowledge of [shear??]s (e.g. u/z) or one must make an assumption about a constant of
integration, and infer velocities subject to that assumption.
For example, in the two-layer ow just discussed, we can infer upper-layer velocities from
easy-to-measure interface variations, provided that we assume the lower layer is motionless.
(More generally, we can use the equations of the last section to infer the difference between
upper- and lower-layer velocities.) Or, to take up the more general case, the thermal wind
equations give us shears, from which simple depth-integration will yield velocities, provided
that we know a constant of integration.
Historically, oceanographers assumed that there would be a [level-of-no-motion??]
somewhere deep in the ocean (where things seemed reasonably quiescent), a level where
there would be no horizontal pressure gradients. One could then calculate the velocity eld
and the slope (or height) of the sea surface relative to this particular depth.
It was clear that this approximation was not likely to be very satisfactory in shallow
water or in areas with strong deep currents. Starting in the 1970s, it has been shown that the
equations of motion plus the continuity conditions can be used in a more sophisticated way
and a depth of no motion can (at least in principle) be obtained from standard hydrographic
data (that is (x, y, z)). An entry to the literature (which is beyond the level of this course)
can be had by searching on the words beta spiral and inverse technique.
In principle, current meter data can be used to determine the ow at particular depth which
can be used to calibrate the standard geostrophic calculations. However the varibility of
ocean circulation is a problem here. One needs long term measurements to get a reasonable
estimate of the mean even in areas of reasonably strong currents.
5.4.5. Dynamic Height
Suppose we wish to get a general picture of the ow over a large area. We would be
interested in the height (D say) of an isobar relative to an assumed constant pressure surface
somewhat deep in the ow.
For example at the surface we know that the velocity is given by
v
s
=
g
f

x
(5.94)
u
s
=
g
f

y
(5.95)
So if we can map (x, y), we will have a picture of the ow pattern.
100 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
The pressure eld is given by
dp
dz
=g (5.96)
so that we can dene an elemental depth increment z in terms of an elemental pressure
increment p:
z =
p
g
(5.97)
The dynamic height relative to some reference level (on which the pressure is p
r
)
D
1
D
r
=
1
g
_
p
1
p
r
_
1
(S, T, p)

1
(35, 0, p)
_
dp (5.98)
so this is the height of the p
1
isobaric surface above p
r
. For the sea surface p
1
= p
a
, the
atmospheric pressure. As a note of caution regarding notation, Ive also seen the above
equation with a negative sign in it, and Ive seen it both with the 1/g multiplier ([Knauss,
1978, Ch 6]) and without it ([Gill, 1982, Section 7.7]).
Maps of dynamic height play the same role as the pressure contours you see on weather
maps do. They let us map out currents owing along contour lines of D.
5.4.6. Constraint on Geostrophic Flow over Sloped Bottoms
Recall (5.76), which reads
f v =
1

P
x
f u =
1

P
y
(5.99)
and suppose that f does not vary with space in the ow regime of interest
14
. In the case
with being constant in a horizontal plane, differentiation reveals that u/x +v/y is
identically zero. Therefore, if we substitute into the continuity equation (section 5.1.7), i.e.
u
x
+
v
y
+
w
z
= 0 (5.100)
we see that
w
z
= 0 (5.101)
Throughout our analysis, we have been assuming that w = 0 at the ocean surface. (You
might nd it useful to see how this assumption threads through the analysis.) Now, the
above equation links w values throughout the water column, in an intriguing way.
Suppose the bottom is at. If there is to be no ow into the bottom, then w = 0 there,
and (5.101) implies (by simple integration) that w = 0 at all depths.
If, on the other hand, the bottom is sloped, any non-zero horizontal ow must cause
vertical ow (since ow must run parallel to the bottom), and thus, by integrating (5.101),
the vertical velocity must be non-zero throughout the water column. But this contradicts the
assumption that w = 0 at the surface. The result is that [geostrophic??] ow cannot cross
lines of constant depth. This is one way of stating the very important [taylor-proudman-
theorem??]. It says that ows must run along lines of constant depth. (More generally, if f
varies, ow must run along lines of constant f /H, where H = H(x, y) is the bottom depth.)
Consider an arbitrary domain, say an ocean with land boundaries on the edges. There are
two types of isobaths in such a domain: (a) isobaths which, at some point, run into land and
(b) isobaths which form closed contours. Since ow must run along isobaths, and cannot run
into land, there must be no ow along isobaths of type (a). However, ow is permitted along
isobaths of type (b). You should imagine stagnant regions, painted by contours that at some
14
Of course, f varies with latitude, but this variation will be small if the scale of the region of study is small.
5.5. WIND FORCING: SLAB RESPONSE 101
Figure 5.14: Sketch of constraint on geostrophic motion in a domain with rigid walls and
non-at bottom [Cushman-Roisin, 1994, gure 4.4]. Since ow cannot enter a wall, there
may be no oceanic geostrophic ow along bathymetric contours that touch land boundaries.
point touch land, and (possibly) moving regions, with closed contours. Flow along closed
contours is sometimes called ow in a Taylor column. This is illustrated in Figure 5.14.
How may we break this constraint? By permitting non-geostrophic ow. This means,
simply, permitting the (time-dependent, advective, friction) terms that are ignored in
[geostrophic??] ows. The next section is an example of non-geostrophic ow, in which
both time-dependence and friction are permitted, in addition to the pressure and coriolis
terms weve used in this section on [geostrophy??].
5.5. Wind forcing: slab response
Our scale anlysis has suggested the importance of [geostrophy??], but how is the balance
achieved? Consider the effect of the wind on a surface mixed layer of depth H. Suppose the
wind blows from east to west, exerting a x-direction stress .
The governing equations, ignoring non-linear terms, and using linear friction (that is,
writing
2
u =u), are
u
t
f v = F u (5.102)
v
t
+ f u = v (5.103)
note that here the wind stress, , is expressed here via F = /(H), where H is the layer
depth.
5.5.1. Case with constant forcing
Steady-state solution
If F is a constant, then the nal steady-state solution for the velocity (u
0
, v
0
), is given
by
15
f v
0
= F u
0
f u
0
=v
0
(5.104)
15
To obtain this, simply cross out the terms with time derivatives, since they must be zero at steady state.
102 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
so
v
0
=
F f
f
2
+
2
(5.105)
u
0
=
F
f
2
+
2
(5.106)
Note that the wind is blowing eastward but the ow is not eastwards. In fact if f
(the usual situation) the ow is southerly at right angles to the wind.
Time-varying solution
The full solution, with time-dependence, is given by
16
u = u
0
(1e
t
cos f t) v
0
e
t
sin f t
v = v
0
(1e
t
cos f t) +u
0
e
t
sin f t
(5.107)
The solution contains two separate physical behavioural patterns:
The inertial oscillation is given by
u
t
f v = 0 (5.108)
v
t
+ f u = 0 (5.109)
so that

2
u
t
2
+ f
2
u = 0 (5.110)
This is harmonic motion at frequency f . If one arbitrarily disturbs the ocean (gives it
a kick) it will tend to oscillate at the local inertial period. Commonly, records from
moored current meters etc, show a noticeable peak at the inertial frequency f , and
also a strong tendency for the current vector at this frequency to rotate to the right.
The nal tendency towards a steady state, with ow almost at right angles to the wind
if the friction is small.
5.5.2. Case with time-varying forcing
The solutions given above were made simple by the assumption that F was a constant.
However, since the problem is linear, solutions for time-varying forcing are not too difcult
to obtain.
Analytical approach
One approach, for a repeating function
17
F = F(t) is to note that repeating functions can
be decomposed into a series of sine and cosine functions. By Eulers rule (Equation 2.13) we
may write each sine-cosine pair as e
it
, with being the frequency. Thus, the F in (5.102)
can be written as ae
it
, where a is a (constant) complex coefcient indicating the strength
of oscillation at frequency . Substitution into (5.102) reveals that the solution must be
proportional to e
it
. And thats the beauty of it if u = u
0
e
it
(where u
0
is a constant) then
u/t = iu
o
e
it
. In other words, the time derivative turns into a multiplying factor of i,
and something remarkable happens: we are left with a couple of algebraic equations to work
with, not differential equations. Yah! Were back at middle school, not university.
Numerical approach
Many software packages perform numerical integration of partial differential equations.
An example is provided in Figure 5.15. You may nd it instructive to run this example (in
the [R??] language) with differing values of , to get a physical feel for the meaning of this
parameter. Then you might like to try moving beyond this code, e.g. by setting it up for a
quadratic stress law, instead of the linear stress law.
16
Discussion in class will help you to see the meaning of the solution; it is not crucial that you can solve this
mathematically.
17
Or a function that can be thought to be repeating, e.g. learning about the response to a single storm by seeing
the response to a series of storms
5.5. WIND FORCING: SLAB RESPONSE 103
0.0 0.5 1.0 1.5 2.0 2.5 3.0

1
.
5

1
.
0

0
.
5
0
.
0
0
.
5
1
.
0
1
.
5
Time [d]
u

a
n
d

v

[
m
/
s
]
u
v
1.5 1.0 0.5 0.0 0.5 1.0 1.5

1
.
5

1
.
0

0
.
5
0
.
0
0
.
5
1
.
0
1
.
5
u [m/s]
v

[
m
/
s
]
pdf("slab_model.pdf", width=9, height=6)
library(odesolve)
mom.eqn <- function(t, x, parms = c(1.0e-4, 0, 0)) {
u <- x[1]
v <- x[2]
f <- parms[1]
lambda <- parms[2]
F <- parms[3]
dudt <- F + f
*
v - lambda
*
u
dvdt <- - f
*
u - lambda
*
v
list(c(dudt, dvdt))
}
slab.model <- function(F=1e-4, lambda=0.3e-4, f=1.0e-4) {
IC <- c(0, 0) # motionless at start
t <- 86400
*
3
*
(seq(0, 500) / 500) # 3 days
params <- c(f, lambda, F)
sol <- lsoda(IC, t, mom.eqn, params)
u <- sol[,2]
v <- sol[,3]
par(mfrow=c(1, 2))
par(mar=c(4,4,1,1))
day <- t / 86400
lim <- c(-1.5, 1.5)
plot(day, u, l,
xlab="Time [d]", ylab="u and v [m/s]", ylim=lim)
lines(day, v, l, lty="dashed")
abline(h=0)
legend("top", c("u","v"), lty=c("solid","dashed"))
par(pty="s")
plot(u, v, l, xlab="u [m/s]", ylab="v [m/s]",
xlim=lim, ylim=lim, asp=1)
grid()
}
slab.model()
Figure 5.15: Slab model of the response to a wind stress = 0.1 Pa in the x direction
(yielding F = / = 1 10
4
in the code). Top left: Time series of response of u and v
components. Top right: Phase diagram of response. Bottom: [R??] code to create the
graphs.
104 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
5.6. Wind Forcing: depth-varying (Ekman) response
Now we turn from slab models (in which no variation is permitted in the z direction) to
the next level of sophistication, in which u and v are permitted to vary with depth. This
problem was rst considered as a result of observations made by Nansen on the classic
voyage of the Fram in the North Polar Sea, 1893-1896. Nansen noted that ice drifted 20-40

to the right of the wind. This led Ekman [1905] to consider the balance of forces in the
upper layers of the ocean.
Ekman considered the steady response of a homogeneous, unbounded ocean to a steady
wind stress. Assuming no pressure gradient forces, and a depth-independent eddy viscosity
A
V
, the equations of motion are
f v = A
V

2
u
z
2
(5.111)
f u = A
V

2
v
z
2
(5.112)
with surface boundary conditions
18
at z = 0
A
V
u
z
=

0

(5.113)
A
V
v
z
= 0 (5.114)
and deep-water boundary conditions
u 0 and v 0 at z (5.115)
The solution is
u = U
0
e
z/D
e
cos
_
z
D
e

4
_
(5.116)
v = U
0
e
z/D
e
sin
_
z
D
e

4
_
(5.117)
where
U
0
=

0

(A
V
f )
1/2
(5.118)
and
D
e
=
_
2A
V
f
_
1/2
(5.119)
Figure 5.16 shows these elds.
Note that
The surface ow (z = 0) is 45

to the right of the wind in the northern hemisphere


(left in southern hemisphere)
The deection increases to the right in successively deeper layers.
The depth at which friction is important is limited to layer of scale given by the
[Ekman-layer??] scale, given by
D
e
=

2A
V
f
(5.120)
The net mass ux is equal to / f and directed at right angles to the wind (in N.
hemisphere)
18
Here, and arbitrarily, we take the wind stress to be in the x direction.
5.6. WIND FORCING: DEPTH-VARYING (EKMAN) RESPONSE 105
Figure 5.16: Ekman spiral. (a) Force balance. (b) Plan view at the surface; note the ow at
45

to the right of the wind. (c) Perspective view, showing the spiraling of current vectors.
(d) Plan view, showing the arrows for current vectors as in (c), but here without perspective;
note again the angle to the wind, and the spiraling to the right. [Figure 9.5, Pond and
Pickard].
We call /f the Ekman mass transport
19
. Ekman transports are crucial to understanding
many elements of ocean circulation:
Ekman transports (indirectly) drive the wind-driven ocean circulation.
Ekman transport causes upwelling near coasts and at shelf breaks, greatly enhancing
biological productivity there.
Exercise 5.4 Ekman layer thickness part 1
a
At 45 degrees North, what is the Ekman layer thickness if the vertical viscosity is
A
V
= 10
2
m
2
/s?
a
Answer in chapter 21.
Exercise 5.5 Ekman layer thickness part 2
a
If a current-meter string sited at 45 degrees North indicates that the Ekman-layer
thickness 30m, what is the A
V
in the region?
a
Answer in chapter 21.
Exercise 5.6 Ekman transport
a
A steady wind of U
a
= 10m/s blows on the ocean. (a) What is the wind-stress?
(b) What Ekman ux results in response to the wind?
a
Answer in chapter 21.
19
The Ekman volume transport is, of course, /( f ).
106 CHAPTER 5. DYNAMICS OF OCEAN CURRENTS
Further reading. Virtually all Oceanography textbooks touch on the issue of Ekman ow,
but, despite 100 years of work, they seldom go much beyond the work of Ekman [1905].
That is not because anyone thinks that Ekman said the nal word; it is mainly because more
sophisticated treatments are mathematically nontrivial; see Madsen [1977] for an example
of the complications that arise by relaxing the assumption about the vertical dependence of
K
V
, for example.
Part IV. Ocean Physics:
Limiting Cases
107
CHAPTER 6
Wind-driven ocean circulation
S
O FAR, WE HAVE CONSIDERED a uniform wind blowing over an unbounded ocean.
But real winds vary over space and time, and real oceans are bounded. So, you may
well ask, just what do real winds do when they blow over real oceans?
The rst part of the story the pattern of the winds is best told in different course. Such
a course will tell you, for example, about the Physics behind Hadleys idea of the winds on
the earth (Figure 6.1), and a more modern three-cell model of circulation (Figure 6.2). Well
have the latter model in mind in this course.
6.1. Sverdrup transport: derivation 1
The mean wind pattern is a series of large scale ows in alternating directions. Figure
6.3 shows this schematically, while Figure 6.4 shows the real wind patterns.
What is the response of the ocean to this wind? We know that the surface response takes
the form of Ekman uxes. But we want the total response, including that of the deep water
(from say 100 m to 4000 m depth) below the Ekman layer.
Lets start with the equation of motion including both pressure (which acts at all depths)
and friction (which acts only in the surface Ekman layer).
f v =
1

p
x
+
1

x
W
z
(6.1)
f u =
1

p
y
+
1

y
W
z
(6.2)
Integrating equations 6.1 and 6.2 down to some presumed level of no motion much deeper
than the thickness of the Ekman layer gives

_
0
h
f vdz =
_
0
h
p
x
dz +
x
W
(6.3)
_
0
h
f udz =
_
0
h
p
y
dz +
y
W
(6.4)
where
x
W
and
y
W
are the x and y components of the wind stress. Note that the shape of the
stress prole in the water column (z) does not enter into this depth-integrated equation;
only the surface stress does, provided that we integrate down to a level where = 0 (that is,
below the Ekman depth).
Dening depth-integrated [transport??] of mass in the x and y directions by
M
x
=
_
0
h
udz (6.5)
M
y
=
_
0
h
vdz (6.6)
we can write equations 6.3 and 6.4 as
109
110 CHAPTER 6. WIND-DRIVEN OCEAN CIRCULATION
Figure 6.1: Schematic of the Hadley [1735] model of atmospheric circulation. Air rises at
the equator (where the sun is directly overhead) and descends at the poles.
Figure 6.2: Schematic of three-cell model of atmospheric circulation. Air rises at the equator
and descends at mid-latitudes (roughly 30N and 30S), forming a subtropical high-pressure
zone.
6.1. SVERDRUP TRANSPORT: DERIVATION 1 111
Figure 6.3: Schematic wind stresses on the North Atlantic, here plotted as a function of
latitude, presuming no dependence on longitude.
Figure 6.4: Actual wind stresses on the North Atlantic. The contours show pressure
eld and the arrows show the wind stress imparted on the water, averaged over the years
1961-1970. [Figure 5, Thompson and Hazen, 1983, Canadian Technical Report of Fisheries
and Aquatic Sciences number 1214, Fisheries and Oceans]
112 CHAPTER 6. WIND-DRIVEN OCEAN CIRCULATION
f M
y
=
_
0
h
p
x
dz +
x
W
(6.7)
f M
x
=
_
0
h
p
y
dz +
y
W
(6.8)
Cross-differentiating to remove the pressure term
1
gives

( f M
y
)
y

( f M
x
)
x
=

x
W
y

y
W
x
(6.9)
or (noting that f /x = 0)
f
M
y
y
f
M
x
x

d f
dy
M
y
=

x
W
y

y
W
x
(6.10)
This can be further simplied by depth-integrating the continuity equation
(u)
x
+
(v)
y
+
(w)
z
= 0 (6.11)
using w = 0 at the surface and at z =h, to get
M
x
x
=
M
y
y
(6.12)
so equation 6.10 becomes

d f
dy
M
y
=

x
W
y

y
W
x
(6.13)
For convenience, we denote d f /dy by :
=
f
y
(6.14)
so we can write
M
y
=

y
W
x

x
W
y
(6.15)
This is the fundamental equation of large-scale wind-driven ow. It is called the Sverdrup
relation, and the South to North ow M
y
is called the Sverdrup transport. In words,
equation 6.15 says
(Sverdrup transport to the North) = curl of wind stress (6.16)
The Sverdrup transport is the total transport due to the wind, including both the surface-
trapped Ekman transport and the deep transport. The ratio of this total transport to the
surface Ekman component is
Ekman transport
Sverdrup transport


W
L

L
Rtan
< 1 (6.17)
where L is the lengthscale over which the wind eld varies (1000 km), R is the radius of
the Earth (6000 km) and is the latitude. The Ekman transport is small compared to the
Sverdrup transport, except near the equator.
1
Cross-differentiation of the x, y momentum equations amounts to taking a curl. Since the pressure force is the
gradient of a scalar, and since the curl of a gradient is always zero, this always removes the pressure terms from the
equations. This trick is used a lot in uid mechanics.
6.2. SVERDRUP TRANSPORT: DERIVATION 2 113
The [beta-plane??]. Note that = 2cos/R is a function of latitude. At the poles
= 0, and is a maximum at the equator, where f = 0. For some problems we can
ignore the variation of f with y, and set up a Cartesian (x, y, z) coordinate system tangent
to the earths surface. This is called the [f-plane??] approximation. For wind-driven
circulation we must take d f /dy into account, so we make the -plane approximation;
that is, we assume that is constant. The -plane approximation is valid over study
regions smaller than 10

, a distance of a few thousand kilometers: f f


0
+y, and
y f
0
.
6.2. Sverdrup transport: derivation 2
6.2.1. Vorticity
It is instructive to derive the formula 6.15 for the Sverdrup transport in more physical
terms. To do this, we will introduce the idea of [vorticity??].
The vorticity is dened as the curl of the velocity eld u(x, y, z). In most large-scale
ows, the horizontal components of the vorticity are not important. The vertical component,
which we will denote , is given by
=
v
x

u
y
(6.18)
We call the relative [vorticity??], f the planetary vorticity, and f + the total [vor-
ticity??].
The relative [vorticity??] is typically much smaller than the planetary [vorticity??].
Their ratio / f U/Lf is the Rossby Number, Ro, which is typically 1.
Exercise 6. Vorticity of simple ows
a
What is the [vorticity??] of (a) a shear ow and (b) a solid-body rotation?
a
Answer in chapter 21.
6.2.2. Potential vorticity
So far, weve been making extensive use of the momentum equations. These say that
momentum is conserved, and that any changes in momentum are caused by forces. In solid
body mechanics, there is a way to write these momentum equations in terms of the spin
of the object in question; this gives a conservation law for angular momentum. In uid
mechanics a similar conservation law applies: the conservation of [potential-vorticity??].
Well mostly be using a form of the [potential-vorticity??] dened for a layer of water of
thickness H:
=
+ f
H
(6.19)
The conservation equation, valid in frictionless ow, is:
d
dt
=
d
dt
_
+ f
H
_
= 0 (6.20)
Figure 6.5 illustrates the principle. If you stretch a column of water, its radius decreases.
Then, by analogy to a ballerina or skater who pulls her/his arms inwards, the water column
spins faster.
114 CHAPTER 6. WIND-DRIVEN OCEAN CIRCULATION
Figure 6.5: Principle of conservation of [potential-vorticity??]. On the left is a slowly
spinning column of uid (the short arrow extending from above showing the vector repre-
senting the rotation rate; note that the vorticity of this spinning body is 2 times the angular
rotation rate). On the right is the state achieved if this column were to be stretched along
the axis of rotation. As the radius is diminished, the spin rate is increased proportionally;
thus the ratio of total spin rate to thickness of the column is a xed quantity. On a rotating
planet, the total rate of spin includes not only the uid spin, as observed by an observer on
the surface of the planet, but also the rotation rate of the planet.
6.2.3. Ekman pumping, potential vorticity, and Sverdrup ow
A simple physical explanation of Sverdrup ow can now be put together (Figure 6.6).
Suppose that the wind stress has negative curl, as in the Subtropical gyre which contains
the Gulf Stream. Then the x-component of wind at point B is greater than that at point A to
the south. Therefore the southward Ekman ux is greater at B than at A, and, consequently,
surface water accumulates between A and B. Therefore water columns in the lower layer
get squashed. Since ( f +)/H is conserved for the lower layer, the reduction in H must
be matched by either a reduction in or in f . The ows which occur in the ocean are
so sluggish that is very much less than f . Therefore, changes in cannot alter f +
signicantly. The only way to reduce f + is to reduce f ; that is, for the uid to move
South to a region where f is smaller.
Negative wind-stress curl Ekman convergence squashing water columns in deep
water Southward ow
The Sverdrup ow occurs in the whole water depth, while Ekman transport is conned
to the surface Ekman layer.
Lets explore the division of ux between the surface Ekman layer and the deep water.
In practice
y
W
is small, so the wind-stress curl can be written

W

1

x
W
y
(6.21)
If is small, then Equation 6.20) gives
1
f
d f
dt

1
H
dH
dt
(6.22)
so that
1
f
d f
dt

w
H
(6.23)
6.2. SVERDRUP TRANSPORT: DERIVATION 2 115
Figure 6.6: Ekman pumping and Sverdrup ow. Top panel: overhead view of North
Atlantic, showing the increase in speed of the Eastward winds as one goes further to the
North. Associated with the Eastward wind is Southward Ekman transport in the upper
part of the water column; this Southward ux of mass increases as one goes further to the
North. Thus, water must accumulate at point B marked on the gure. Lower panel: side
view, with points A, B,C from upper panel indicated for orientation. The curved line denotes
the boundary between the upper layer and the deeper water. The convergence of Ekman
transport at B causes a thickening of the upper layer there, and thus a thinning of the deeper
layer. Imagine a cylinder of uid in the lower layer, of initial thickness H. The effect of
the Ekman convergence is to decrease H as time goes on, and because f /H must remain
constant, through conservation of [potential-vorticity??] given neglible local vorticity, the
uid column must move South to a region where f is lower.
116 CHAPTER 6. WIND-DRIVEN OCEAN CIRCULATION
where w is the vertical velocity imposed upon the top of the deep water by the convergence
of Ekman transport in the upper water. Now, by the Ekman relation,
w =
1

y
_

x
W
f
_
=
1
f

x
W
y
+

x
W
f
2
(6.24)
Scales. The velocity scale is w
W
/pf L. For winds of 10 m/s, w
0.2/(10
3
10
4
10
6
) or 210
6
m/s or w 60 m/year. Generally this is much more
important than rainfall or evaporation (which produce equivalent terms in these equa-
tions).
Since the changes in f are wholly advective, we can write
1
f
d f
dt
=
1
f
v
d f
dy
=
v
f
(6.25)
Using this together with Equation 6.24,
v
f
=
1
H
_
1
f
d
W
dy

f
2
_
(6.26)
M
y
D
= vH =
1

W
y
+

W
f
(6.27)
where the subscript D refers to the deep water. Consider the case when the Sverdrup
transport is zero, so the rst term on the right-hand side is zero. Then the deep water ux
M
y
D
is
W
/ f , or just the negative of the surface Ekman layer ux . . . the surface ow is
returned in the deep ow. In other words, the rst term in Equation 6.27 is the total Sverdrup
ow. When Equation 6.27 is added to the Ekman ow, we thus have a second derivation of
the Sverdrup relation (Equation 6.15) for the special case of East-West winds.
6.3. The return ow
To summarize: in the Subtropical North Atlantic (between 20

N and 40

N) there is a
slow southward drift of water. This cannot continue indenitely, because the North Atlantic
would dry up. How is this water returned to the North?
To see the answer, start with the constraint (Equation 6.20) that [potential-vorticity??]
is conserved:
d
dt
=
d
dt
_
+ f
H
_
= 0 (6.28)
That is, ( + f )/H is constant for a given [parcel??] of uid as it moves. For steady
ow the path a particle traces as it moves is called a streamline, so ( + f )/H should be
conserved along streamlines.
Imagine a water column starting at the Northward side of the gyre. It has = 0. It gets
squashed and moves South to maintain at its initial value. It arrives at the Southward
side of the gyre with = 0 and then enters the return ow. This return ow will be along
a boundary so that the addition of sidewall friction to the momentum balance can allow
an escape from the Sverdrup relation. But will it be the Eastern boundary or the Western
boundary? We know that it must reach the Northward end of its journey with = 0, so
that it will be at the top of the gyre with the same vorticity as it had before it began its
circuit. When it travels North, f increases. Assuming H is constant, this means that will
decrease (become more negative), which we know must not happen. There is one way to
keep from decreasing: if the current is along the Western boundary, then friction will
impart counterclockwise, or positive, vorticity. This can cancel the decrease in which
would otherwise occur. (Exercise: demonstrate that a return ow along the Eastern boundary
would have precisely the wrong effect.)
6.4. ROSSBY WAVES 117
Figure 6.7: Schematic diagram of [Rossby-wave??] propagation in a at-bottomed ocean.
An overhead view is shown, with North at the top of the gure. The thick sinudoidal line
denotes a locus of points originally lying along the thin horizontal line. Thus, the points
originally were on a line of constant latitude and constant value of f . The deformation
of the line, presumed to occur because of an agency of no concern here, induces negative
relative vorticity in the points to the North of the original line, and positive for points
to the South, in order that [potential-vorticity??] ( + f )/H be conserved. Points moved
further from the line receive larger increments of than those closer to the line. These
increments of are indicated by the little curved arrows, with arrows extending further
around the circle denoting larger spin, e.g., c spins faster than its neighbors b and d. The
spin of c pushes b to the North. Point c does not move, since it sees equal forces to the
North and South by its neighbors. Point d is moved to the South by c, while e is moved to
the South by the combined effect of d and f . Point f is moved to the South by g, while g
remains motionless and h is moved to the North by g. The net effect is shown in the lower
panel, where the original conguration is drawn dashed, and the original points in gray, and
the induced conguration is shown in the solid line. The net effect is a movement of the
sinusiodal pattern to the West. Thus is illustrated the mechanism of the Westward motion of
at-bottom [Rossby-wave??]s.
In summary:
Centre of the gyre: Wind eld with negative vorticity tendency causes the Ekman
ow to converge.
Ekman covergence causes downward Ekman pumping, squashing water columns
beneath the Ekman layer.
To conserve [potential-vorticity??], the deep water moves Southward.
The water return North in a Western boundary current. As it moves North, f increases.
This tends to reduce , but this reduction is cancelled by sidewall friction, which
tends to increase .
6.4. Rossby waves
The variation of f with y also leads to [Rossby-wave??]s. The essentials of the Physics
are shown in Figure 6.7. In mid-ocean a small northward displacement produces a decrease
in relative vorticity
118 CHAPTER 6. WIND-DRIVEN OCEAN CIRCULATION
f increases = decreases (6.29)
f decreases = increases (6.30)
As the Figure 6.7 shows, these displacements provide a restoring force which creates a wave,
a so-called planetary [Rossby-wave??]. These propagates westwards (in both hemispheres.)
In fact it is ( f +)/H which is conserved, so a sloping seabed (H varying) produces a
topographic beta effect that is analogous to that due to latitudinal changes in f . If the water
moves upslope H decreases, decreases. The resulting topographic [Rossby-wave??]s
move along the slope, with shallow water on their right.
Rossby waves may be signalled by a sequence of clockwise and anti-clockwise eddies
which move to the west (or along the slope with shallow water to the right) at the [Rossby-
wave??] phase speed
C
p
=
_
L
2
_
2
(6.31)
where L is the wavelength.
Further reading. The issue of western intensication, e.g., the reason for the Gulf Stream
being on the western side of the Atlantic, is covered well in Hughes and de Cuevas [2001].
Although the bulk of the paper is at a level that is a bit too high for this course, the
introduction might be useful to you.
CHAPTER 7
Thermohaline circulation
S
OLAR HEATING, COMBINED WITH WIND MIXING, will tend over time to gradually
warm the entire water column. But observations reveal that the deep waters are
very cold
1
so if there is to be a steady state, there must be a source of cold water.
This source is through the thermohaline circulation. In polar regions, strong atmospheric
cooling and evaporation increase the density of surface waters so much that they sink to
the bottom of the ocean. This deep cold water circulates throughout the world ocean. It
slowly upwells, supplying a source of cold to balance the source of heat from the sun. The
balance is between mixing of solar heat downward and advection of cold upward by weak
vertical currents (Figure 7.1).
This balance is addressed in detail in Section 7.2.1. But rst, we might ask how the deep
water gets from the poles to the central waters of the gyres.
Early work by W ust provided us with a picture of how the deep water circulates. Figure
7.2 shows the large-scale signature of the thermohaline circulation. In the middle panel,
the salinity signature of the North Atlantic Deep Water can be traced back to the source
near Greenland. One can also see in the O
2
panel that the deep waters are high in oxygen
because of the contact they had with the atmosphere in polar regions where the deep water
is formed.
7.1. Movement of deep water
The necessity to conserve potential vorticity governs the ow of the deep waters involved
in the thermohaline circulation (Figure 7.3).
Throughout the open gyre, the deep water upwells from the bottom towards the surface.
Since w = 0 at the ocean bottom, this means that w/z > 0; thus the upwelling stretches
the vortex lines, increasing H. To conserve ( f +)/H, the uid must move North, where
f is larger. ( cannot increase to compensate the increase in H, because f ; the
Rossby number is too low.) So, somewhat counterintuitively, the open ocean thermohaline
circulation is toward the North (thus towards the source). The deep water must ow from
North to South as a sidewall boundary current, for only with the addition of friction can the
force balance allow movement to the South. This boundary current must be on the Western
side of the basin, because only there do frictional torques supply negative vorticity. This
supply is needed to keep from increasing as the water ows South where f is less. (This
reasoning parallels that for the return current of the Sverdrup ow.)
1
Historical note: The fact that deep waters are cold has been known for a long time. Mariners commonly
lower lead lines over the sides to determine the water depth. (Water depth is an important variable since it is
no fun to run aground!). The weight at the bottom of the line was a mass of lead, hence the name. It has long
been observed that this lead weight would be very cold upon returning from the depths, even regions with warm
surface waters. The only explanation of this is that the deep waters are cold. This was reported by Robert Boyle in
1671, but we have no way of knowing how common the knowledge of cold deep water was at the time of reporing.
However, Boyle also reports that it was common for mariners in hot tropical regions to lower wine bottle on ropes,
letting them hang in the deep waters for a while, in order to get a cool beverage. (It was white wine, we can only
assume.) That this practice was apparently common implies that the knowledge of the coldness of deep water must
have been somewhat widespread in Boyles day. Systematic measurements of the cold temperatures of deep waters
were made in the early part of the next century. Ellis, on a slave-trading ship named Earl of Halifax (an ignoble
reference to the town in which we live, Id say), carried out measurements of the deep waters off Africa using a
lowered thermometer, reporting on his results in 1751. Within a few decades, a theory for the ocean circulation, in
which cold polar waters sink under warm tropical wates, was being put together. Richard Kirwin describes this
circulation in 1787, as does Delam etherie in 1797. But the most signicant work was by Count Rumford, who in
1798 laid out a fairly modern circulation scheme, driven by pressure differences associated with density variations
of the owing water. Humboldts writings in 1814 added to the picture. Thus, crude observations made at least as
early as the mid 1600s, followed up by better observations made in the middle 1700s led to a theory in the early
1800s. For more on the history, see Deacon [1971].
119
120 CHAPTER 7. THERMOHALINE CIRCULATION
Figure 7.1: Schematic thermocline balance: transport of solar heat downwards by mixing,
balanced by transport of cold upwards by weak vertical currents.
7.2. Thermocline theory
7.2.1. The Munk balance
In a seminal paper, Munk [1966] laid the foundation for thermocline theory, starting with
the conservation equation
T
t
+u
T
x
+v
T
y
+w
T
z
=

x
_
K
H
T
x
_
+

y
_
K
H
T
y
_
+

z
_
K
V
T
z
_
(7.1)
where K
H
and K
V
are the horizontal and vertical eddy diffusivities; that is, K
V
=w
/
T
/
/(T/z)+
molecular diffusivity.
The simplest steady balance is between downward mixing of heat and upward advection
of cold:
w
T
z
=

z
_
K
V
T
z
_
(7.2)
so, if K
V
is constant (and w is constant by continuity) the solution is of the form
T = T
deep
+(T
sur f ace
T
deep
)exp
_
wz
K
V
_
(7.3)
Munk used observed proles T(z) to t for the value of K
V
/w. Figure 7.4 shows Munks
Figure 3. It can be seen that the data can be tted using K
V
/w 1 km. To get the separate
values for K
V
and w, Munk looked at
14
C proles. This isotope decays at a known rate,
obeying the equation
w

14
C
z
=

z
_
K
V

14
C
z
_

14
C (7.4)
Fitting the solution of this equation to the
14
C data gives a second relationship between
K
V
and w. Therefore Munk could estimate K
V
and w separately, with the results being
w 10
7
m/s and K
V
10
4
m
2
/s.
7.2. THERMOCLINE THEORY 121
Figure 7.2: W usts South-North sections along the Western trough of the North Atlantic.
Note the North Atlantic Deep water, formed near Greenland, sinking and moving throughout
the basin. [Figure 23, Pickard]. Exercise: compare the lower panel with the oxygen section
shown in Figure 4.28.
122 CHAPTER 7. THERMOHALINE CIRCULATION
Figure 7.3: Stommels schematic diagram of the thermohaline circulation. S
1
is a source of
deep water (conceptually, in the Northern North Atlantic). It enters the basin in a Western
boundary current, and upwells throughout the gyre. The resultant vortex stretching causes
Northward movement in the interior of the gyre. [Figure 10.1, Pond and Pickard]
Figure 7.4: Figure 3 of Munk [1966], showing temperature (left) and salinity (right) proles
in the central Pacic waters, below 1 km depth. The dots are the data, and the lines are
curves with w/K
V
equal to 1.0 km
1
, 1.2 km
1
and 1.3 km
1
; the tightness of these lines
suggests the tightness of the K
V
inference from such data, if w could be considered to be
known.
7.3. GENERAL CIRCULATION 123
This value K
V
10
4
m
2
s
1
is often taken as a canonical value for use in calculations of
mixing in the deep ocean.
We now disbelieve this simple calculation. In practice, the uT/x, etc. terms cant be
neglected. A major part of the thermocline balance involves quasi-horizontal motion and
mixing along isopycnals. One element of this new view of the thermocline is the ventilated
thermocline theory.
7.2.2. The Ventilated Thermocline
The ventilated thermocline theory breaks away from Munks one dimensional view to
consider a three dimensional picture. To simplify matters, it assumes that vertical variation
is adequately regarded as a series of layers. This replaces the differential equation for
depth variation with a set of simpler equations for conservation within the layers. The
layers need not be at; in fact they slope so much that some deep layers contact the
atmosphere. The theory centres on this contact. Layers which contact the atmosphere are
said to be ventilated. The interaction with the atmosphere is through Ekman pumping and
precipitation. The simplest version of the model has 3 layers. The top layer is driven directly
by the wind stress curl and Ekman pumping. The middle layer is bowl-shaped and contacts
the atmosphere only in the North. The bottom layer is assumed to be motionless.
Where the middle layer is exposed to the wind, it is driven Southward by Ekman pumping.
When a water column in this middle layer slides under the top layer, it is isolated from direct
wind forcing. From then on, it must ow along geostrophic contours, lines of constant f /H,
H being the layers thickness. Since f decreases in a Southerly direction, the water column
follows a path which allows H to decrease just enough to compensate. This constrains the
motion.
Thinking along these lines, Jim Luyten, Joe Pedlosky and Henry Stommel of the Woods
Hole Oceanographic Institution Luyten et al. [1983] mapped out general properties of
solutions.
7.3. General circulation
The distinction between wind-driven and thermohaline circulation is not based upon
separation of the two processes in reality. Rather, it is a convenient artice that has allowed
us to approach the problem. The circulation of the oceans is a combination of the wind
driven and thermohaline circulation. Attempts have been made to model the combination
together, but it is difcult for several reasons Stommel [1984], including: (a) the geometry
of the real oceans is complex; (b) the equations are non-linear; (c) the forcing is variable; (d)
the eddy viscosities and diffusivities (A
H
, K
H
, A
V
, K
V
, etc) are unknown; furthermore, they
are probably functions of the ow itself (see 11.6.
Most existing models of the steady ow have been shown to be unrealistic. To understand
the ocean it is necessary to include the effects of eddies. Numerical models can include
eddies in a very limited way. Such models already mimic many of the observed features
of the ocean circulation. But the models involve some major assumptions (about eddy
coecients) that can only be sidestepped by greater computer power. It is estimated that
computer power will not be sufcient for adequate eddy resolution in these models for 1-2
decades.
A recent paper Ganauchaud and Wunsch [2000] casts a great deal of light on this topic in
general, and Ive presented their nal diagram, of [meridional??] (north-south) heat ux as
Figure 7.5.
Further reading on the thermohaline circulation. Wunsch [2002] discusses a confu-
sion in the literature about just what the term thermohaline circulation means, and he
highlights some key aspects of the dynamics. More details of heat uxes associated with the
thermohaline circulation, and the ocean ow/mixing in general, is provided by Ganauchaud
and Wunsch [2000]; this paper is also valuable because of its comments on how our un-
derstanding is limited by our observations. If you are a Physical Oceanography student,
you may also be interested in a paper written by a recent PhD student and his supervisor;
Greatbatch and Lu [2003] discuss two old models of the thermohaline circulation in the
context of numerical models.
124 CHAPTER 7. THERMOHALINE CIRCULATION
Figure 7.5: Meridional heat ux according to Figure 3 of Ganachaud & Wunsch (2000).
This is an important reference which I urge you to read!
CHAPTER 8
Continental Shelf Circulation
Reference Physical Oceanography of Coastal Waters, K.F. Bowden (1983). 302 pp. Ellis
Harwood Inc. (division of Wiley, New York).
T
HE CONTINENTAL SHELF REFERS TO THE SHALLOW, relatively at, zones near
the edges of continents. On this coast of Canada there are broad shelves, whereas
on the west coast the shelf is narrower; the reasons for the difference have to do
with the uid mechanics of [convection??] in the earth (and, yes, people who study this
mechanics are very familiar with the equations of chapter 5). The ocean circulation on
continental shelves is important, and this chapter will explain some of what we know about
this circulation.
8.1. Introduction
The dynamics of the continental margins is a vast, relatively new eld of physical
oceanography. Vast because most of the processes that occur in the deep ocean also
occur somewhere on the the continental shelf. Relatively new in the sense that relatively
little effort was applied to the study of this area before 1960. Deep-sea oceanography was
then the respectable eld; the exploration of the oceans and the understanding of general
circulation were the central priorities. Shelves were known to be complex both in terms of
the requirements for measurements (geostrophic estimates were known to be questionable
with doubts about the variability and the level of no motion) and the requirements for
theoretical understanding (the combination of density stratication with strong bottom
topography was known to be intractable). Consequently, no rm body of evidence was built
on continental shelf processes.
The increasing importance of the shelf, economically and politically, has provided a
strong impetus for detailed study of this messy area.
8.2. Processes and scales
As Table 8.1, reproduced from Table I of Mooers [1976], shows, there are important
processes on many time and space scales. Listing by timescale, we have:
Seconds to minutes Surface gravity waves; internal gravity waves.
Hours Tides (regular but with spring/neap cycle, internal tides).
Process Timescale U, m/s L, km W, km Vertical structure
Long-term mean decade 0.01 10
3
10
4
Margin width BT + perhaps 1 BC
Seasonal seasonal 0.1 10
3
Shelf width BT + at least 1 BC
Wind event 210 day 1 10
2
10
3
Shelf width Mainly BT
Tidal 1 day 0.1 10
3
10
4
Margin width BT + some BC
Inertial / f 0.1 10? 10? 10 BC modes
Edge wave minutes 0.1 1 0.110 Uniform
Surf. grav. wave seconds 0.1 0.1 0.0010.1 Uniform
Table 8.1: Scales for some phenomena. U is the velocity scale, L is the longshore scale
and W is the cross-shore scale. BT means barotropic and BC means baroclinic (1 BC
means 1 baroclinic mode). [after Table 1 of Mooers, 1976]
125
126 CHAPTER 8. CONTINENTAL SHELF CIRCULATION
Figure 8.1: Coastal upwelling. Arrows denote Ekman ow and return ow, M and M, in
units m
2
/s.
Days Low frequency shelf waves: periods days to weeks. Edge waves and shelf
resonances (barotropic): minutes to hours. Response to wind.
Weeks Timescale for individual lows, storms, etc., is 1-10 days on Scotian Shelf.
Seasonal Cycles in fresh water ow, tides, storm patterns, summer hurricanes, winds
and associated waves.
Decades Trends in river runoff (e.g., St. Lawrence), trends in wind patterns, ice cover,
sea-level, offshore inuences.
It is useful to dene regions where these various forces are inuential in different degrees.
Nearshore from the shoreline to 20m depth or so, the regime of strong surf zone
inuence, frequently vertically well mixed, vertically homogeneous in properties.
Persistent breaking of gravity waves near the shore drives longshore currents, rip
currents, nearshore circulation cells. River runoff provides thermocline-type forcing:
gradients of heat and salt, generally distributed at points along the coast. Strong
seasonal and geographic dependence.
Inner Shelf the true shelf regime, not dominated by nearshore or offshore effects.
Vertically stratied in summer, well mixed but horizontally stratied in winter.
Wind driven circulation, response of ow to wind events (esp. storms, hurricanes)
very noticeable. Balance between wind mixing, tidal mixing and buoyancy input
(solar heat, fresh water) important in determining the distribution of the density eld,
particularly the location of fronts.
Outer Shelf transition zone between shelf and ocean, generally strongly stratied, with
sloping [isopycnal??] surfaces often forming surface or bottom fronts. The outer
shelf region may be strongly inuenced by a major boundary current, intrusions of
eddies, etc. Internal tides may be generated at the shelf edge by the movement of the
barotropic tide on and off the shelf.
8.3. Wind-driven coastal upwelling
Consider a continental shelf in the Northern hemisphere. Suppose a longshore wind
blows with the coast on its left. Then the surface Ekman ux M
E
will be directed to the right
of the wind that is, away from the coast. Conservation of mass implies that a steady state
will require that onshore currents in the lower waters will return this ux. Figure 8.1 shows
this schematically. What are the details of ow, and what is the temporal development?
Clearly the ux must decay to zero at the coast, since there can be no ow into a
solid boundary. This leads to a divergence of ow in the upper waters, that can only be
compensated for by upward ow of deep water (Figure 8.2). This upward ow is coastal
8.3. WIND-DRIVEN COASTAL UPWELLING 127
Figure 8.2: Coastal upwelling, illustrating both Ekman divergence and coastal upwelling.
The divergence of ow in the upper layer warps the surface down (greatly exaggerated here)
and the interface up. As a result of the cross-shore oriented pressure gradient, a geostrophic
longshore jet develops near the coast.
upwelling. It is of great interest because the deep waters are rich in nutrients, so that
upwelling can lead to very productive biological blooms.
We can estimate the upwelling rate w as follows. Denote the lengthscale over which the
Ekman ux increases to the deep-water value by L. Then
w

W
f L
(8.1)
If L can be measured, then Equation 8.1 can be used to estimate w. But to predict L we must
look at the temporal development of upwelling.
The sequence of events envisioned is:
In the few hours after the wind starts, an Ekman layer forms. The offshore ux in this
layer is /( f ), except near the the coast.
At the coast, the outow of water causes the sea level to fall. The resulting pressure
gradient causes water to accelerate towards shore at all depths, in the return ow
sketched in Figures 8.1 and 8.2. The pressure gradient also accelerates a longshore
ow in the same direction as the wind.
The vertical velocity at the coast tilts the thermocline up at the coast.
This thermocline tilt counteracts the onshore pressure gradient in the lower layer, so
that the onshore pressure gradient is zero there, and there is no longshore ow in the
lower layer. But the upper layer still feels the pressure gradient from the surface tilt,
and the geostrophically-balanced longshore jet (the coastal jet) there continues.
With this dynamical picture in mind, we can write out the relevant equations and scale
them to get the width L of the coastal jet region. For simplicity, suppose the stratication
can be approximated by 2 layers, with densities
u
in the upper layer and
l
in the lower
layer. Denote the uplifting of the interface between the layers .
First, note that the continuity equation
u
x
+
w
z
= 0 (8.2)
implies the following scaling relationship
U
L

W
H
(8.3)
Next, note that conservation of onshore momentum
128 CHAPTER 8. CONTINENTAL SHELF CIRCULATION
f v =
1

p
x
= g
/

x
(8.4)
(where [g??]= g(
l

u
)/
u
is the [reduced-gravity??]) implies
f V g
/

L
(8.5)
Conservation of longshore momentum
v
t
+ f u = 0 (8.6)
implies
V f TU (8.7)
where T is the timescale for the response.
Combining Equations 8.3, 8.5 and 8.7 and noting that w /T yields
L

g
/
H
f
(8.8)
L is called the baroclinic [Rossby-radius??] of deformation
1
. It is the offshore length scale
over which the adjustment to the Ekman ux takes place. Many processes which rely upon
interfacial tilts in their dynamics have L as the lengthscale.
For a continuously stratied ow, as opposed to 2-layer ow, the baroclinic [Rossby-
radius??] is slightly different. We dene [g-prime??], the [reduced-gravity??], by
g
/
=H
_
g

z
_
= HN
2
(8.9)
so the [Rossby-radius??] becomes L = NH/f . For coastal upwelling, the scales f
10
4
s
1
, N 10
2
s
1
imply L 100H 10km. Then typical upwelling velocities are
(with
W
0.2Pa)
w

f L
210
4
m/s (8.10)
Note that w is very much larger than Ekman pumping upwelling velocities that drive the
Sverdrup ow in the interior of the gyre (see chapter 6).
8.4. Shelf-break upwelling
Upwelling similar to that at the coast is also seen at the shelf edge. Much of the physics
is the same. Figure 8.3 shows a denition sketch Huthnance [1981].
8.5. Shelf waves
As you might expect, the response to wind stress is not smooth. Many waves are
generated by changing wind stress and wind pressure.
One of the most important types of waves is the continental shelf wave. Such waves
are responsible for much of the variability in the 3-10 day band. In the Northern hemisphere,
such waves travel with the coast to their right (Figure 8.4, from Mooers [1976]).
1
The barotropic [Rossby-radius??] of deformation, in a sense the depth-integrated form of the baroclinic
radius, has g in place of the [reduced-gravity??] [g-prime??].
8.6. STORM SURGES 129
Figure 8.3: Side view of shelf-edge upwelling, after Huthnance 1981. Note that upwelling
occurs not just at the coast, but also at the shelf edge.
Figure 8.4: Shelf wave. (From C Mooers, 1976. Wind-driven currents on the Continental
Margin in Marine sediment transport and environmental management, ed Stanley and Swift.
Wiley)
8.6. Storm surges
Atmospheric storms can lead to temporary elevations or depressions of sealevel, due
to both the [inverted-barometer-effect??] and to the coastal convergence of stress-driven
ows. This so-called [storm-surge??] phenonomen is especially damaging in low-lying
coastal areas. For example, Bangladesh has lost thousands of lives to storm surges within
living memory. Financials costs are also high, e.g. a Florida surge in 1995 did $3 billion in
damage.
Physical Oceanographers at Dalhousie are spearheading real-time forecasts of storm
surges. See our websites.
Depending on the interests of the class, I may spend some time on storm surges during
lectures, perhaps using a presentation I prepared for high-school teachers
2
. For further
reading at layperson level, see my related essay for the Association of Science Teachers
journal
3
.
8.7. Thermohaline circulation
So far, we have only considered wind-driven motion. But in the shelves, as in the deep
sea, buoyancy input causes currents.
8.7.1. Buoyancy currents
The Southwestward drift of shelf waters off the Eastern coast of North America is caused
by buoyancy input from rivers. The relatively fresh, and therefore relatively light, water
exits from rivers and turns to the right that is, it turns in the Southward direction along the
coast (Figure 8.5). As it does so, it mixes with the saltier water on the shelf. So the details
2
http://www.phys.ocean.dal.ca/

kelley/publications/presentations_other/
2002/AST/
3
http://www.phys.ocean.dal.ca/

kelley/publications/publications_other/
Kelley_2003_AST/
130 CHAPTER 8. CONTINENTAL SHELF CIRCULATION
Figure 8.5: Buoyancy currents on the Northwest Atlantic shelf. (From W H Sutcliffe, R H
Loucks and K F Drinkwater, 1976. Journal of the Fisheries Research Board of Canada.)
of the process depend not only on the rate of freshwater input, but also upon mixing (and
thus upon solar input, which modies N, and also upon tidal currents, etc).
The combined action of thermohaline and wind-driven currents is not completely un-
derstood, although the generalities southward ow of the buoyancy current, etc are
rm.
8.7.2. Fronts
Generally, the waters of the shelf are isolated to some extent from those of the deeper
ocean by a shelf-edge front. Figure 8.6 shows a shelf-edge front. This Figure is taken from
a review by Csanady [1981, Circulation in the Coastal Ocean, part 2. EOS, 62(5)].
Csanady makes the point that the details of the frontal dynamics are not well understood.
We know in general why the front is there; in the shallow ocean the surface waters are
made lighter by both solar input and freshwater input, and strong mixing there creates a
light water mass that cannot mix with the offshore water. But the details of the frontal shape
are not well understood, partly because the front is not a static feature. The front is subject
to many instabilities, which exchange water across it and at times erase it more or less
completely. Figure 8.7 shows that the 3-dimensional structure of these fronts is complex;
we also know that the temporal evolution is complicated. Satellite techniques have proved
useful in monitoring the temporal evolution of frontal instabilities (Figure 8.8). Cross-front
exchanges are of vital interest to biological oceanography and sheries, because they can
bring nutrient-rich deep ocean water onto the shelf, and thus drive active biological systems.
Understanding these instabilities and parameterizing their uxes is at the cutting edge of
research.
8.7.3. Tides on Shelves and Slopes
In class, I will outline some effects of tidal [rectication??] and related phenomena.
8.7. THERMOHALINE CIRCULATION 131
Figure 8.6: Shelf-edge front. (Figure 7 in Csanady [1981, Circulation in the Coastal Ocean,
part 2. EOS, 62(5)].
Figure 8.7: Shelf-edge front: 3D view. (From Smith, 1978, Low-frequency uxes of
momentum, heat, salt, and nutrients at the edge of the Scotian shelf. J Geophys Res, 83,
4079-4096.)
132 CHAPTER 8. CONTINENTAL SHELF CIRCULATION
Figure 8.8: Left: Infrared image of California Current, with lighter colors indicating cooler
water. Upwelling near the coast brings cold water to the surface, which then extends offshore.
Tics on axes are 100 km apart. The dashed lines indicate 300 m and 3000 m isobathes.
(After Fig 1 of Flament et al. [1985].) Right: Satellite infrared image on 1994 sep 5 (source
http://www.oce.orst.edu/po/coastal.html).
CHAPTER 9
Estuarine Circulation
Im running home, babe
Like a river to the sea.
Your bright baby blues c _1976 Jackson Browne
M
UCH OF THE WORLD POPULATION lives near an estuary (a place where a river
meets the sea), and the waste products of this population often nd their way into
the water. Understanding estuarine circulation is, therefore, important to issues
of water quality, e.g. the issue of whether we should treat sewage before it enters Halifax
harbour (Figure ??).
Density effects are appreciable on scales of the shelf
1
. However, as the sources of fresh
water are normally localized, the density effects are even more noticeable close to the rivers.
Hence, we will concentrate on estuarine circulation.
An [estuary??] is a semi-enclosed coastal body of water which has a free connection
with the open sea, and within which sea water is measurably diluted with fresh water derived
from land drainage. Fischer [1976].
Estuaries are regions of great practical importance, so a substantial body of literature has
built up on the subject. While the circulation is driven by fresh water input or fresh water
evaporation (in the case of negative estuaries, such as the Mediterranean), the tidal action
and the topography are also extremely important in determining the ow structure.
9.1. Classication of Estuaries
9.1.1. Topographic Classication
Roughly speaking, topography denes 3 groups:
Drowned river valleys Sea level was 100m lower some 18,000 years ago. Rising sea
level since has ooded existing river valleys.
Features
Generally slow sedimentation rates.
Shallow (30m), large width/depth ratio, widening and deepening towards mouth.
Examples Chesapeake, Thames, St. Lawrence, St. Croix.
Fjords Gouged out by glacial action.
Features
U-shaped quasirectangular cross-section.
Deep, small width/depth ratio (commonly 1:10)
Rock bars or sills common at fjord mouths and intersections
River ow small compared to fjord volume
Examples Norway, B.C., Labrador, Bedford Basin.
Bar-built estuaries are similar to drowned valleys but with far more sediment which
forms a bar at the mouth.
Features
Shallow, usually only a few meters deep.
Large sediment carrying, seasonal river ow.
1
See for example Bumpus [1973].
133
134 CHAPTER 9. ESTUARINE CIRCULATION
Distance [ km ]
D
e
p
t
h

[

m

]
5
0
4
0
3
0
2
0
1
0
0
8
9
1
0


1
2

1
2

12
1
3


1
4

0 2 4 6 8 10 12
T
Distance [ km ]
5
0
4
0
3
0
2
0
1
0
0
3
0


3
0
.
4

30.4
3
0
.8


3
0
.
8

31
31.2
31.4
0 2 4 6 8 10 12
S
Distance [ km ]
D
e
p
t
h

[

m

]
5
0
4
0
3
0
2
0
1
0
0
22.5

2
3

2
3

23.5
23.5
24
24.5
0 2 4 6 8 10 12

296.35 296.40 296.45


4
4
.
6
2
4
4
.
6
4
4
4
.
6
6
4
4
.
6
8
4
4
.
7
0
4
4
.
7
2
Longitude
L
a
t
i
t
u
d
e
pdf("bedford-basin.pdf", width=8, height=7)
library(oce)
data(section)
data(coastlineHalifax)
plot(section, coastline=coastlineHalifax)
Figure 9.1: Top four panels: observations made in Bedford Basin by previous members
of this class, as part of a now-defunct eld component. The bottom-right of these panels
shows the station locations, with a square box surrounding the station designated to have
distance 0 km in the other panels. Note the pronounced reduction in S at this station, which
is near the outow of the Sackville River. Lower panel: R code used to create the diagram.
Advanced exercise: learn enough about the object hiearchy in the Oce package to modify
the R code to plot a TS diagram.
9.2. MIXING: THE RICHARDSON NUMBER 135
9.1.2. Hydrographic classication
While the topographic classication scheme clearly denes fjords and lagoons as being
different, it is clear that the drowned river valleys could differ considerably from one another.
Another useful classication is one based on salinity structure. There are 2 types (Figure
9.2).
Vertically homogeneous estuary For narrow estuaries with strong tidal currents, the
vertical mixing is essentially complete. There are two subcases:
Laterally homogeneous Salinity increases in the direction of the mouth. There
is seaward mean ow at all depths as shown in Figure 9.2. The salt inux up the
estuary arises from the assymetry of tidal transport of salt plus turbulent mixing.
Laterally variable If the estuary is wide compared to the internal Rossby radius,
Coriolis effects separate inow from outowhorizontal circulation pattern.
(e.g., Delaware Estuary)
Partially mixed estuary Tidal mixing not entirely complete. The ow is seaward at
the surface and landward at the bottom, with entrainment between these counterows.
9.2. Mixing: the Richardson number
A measure of the potential for mixing by currents is Ri, the Richardson number. There
are many formulations of Ri. Which formulation is used in analyzing a particular system
depends on the general properties of the ow and on the circumstances of measurement.
The most basic denition is the gradient Richardson Number , dened as
Ri =
N
2
_
u
z
_
2
(9.1)
which is a ratio of the stabilizing effect of the stratication to the destabilizating effect of
the shear. When Ri < 0.25, the uid is turbulent and mixing rates are high.
The problem is that gradient Richardson number is difcult to measure, requiring accurate
measurements of both (z) and u(z). In estuarine work, another formulation is often used:
the estuarine Richardson number
Ri
E
= g

F
WU
3
(9.2)
where F is the freshwater discharge in m
3
s
1
, W is the width of the estuary, U is the typical
r.m.s. tidal current, and is the density difference between river water and sea water.
2
An empirical mixing classication scheme based upon estuarine Richardson number is
2
Physical Meaning of Ri
E
The addition of fresh (therefore low-density) water at the surface decreases the potential energy of the system.
The rate of change of potential energy (per unit area) is
dPE
dt
= g

F
W
(9.3)
while the rate of supply of turbulent kinetic energy via bottom drag (equal to C
D
U
2
) acting against tidal velocity U
is
dKE
dt
=C
D
U
3
(9.4)
so that the ratio (dPE/dt)/(dKE/dt) is Ri
E
/C
D
. Thus Ri
E
measures the competitition between the stabilizing
effect of the input of buoyancy ux and the destabilizing effect of the mixing currents.
136 CHAPTER 9. ESTUARINE CIRCULATION
Figure 9.2: . Estuary classication by S structure [Figure 31, Pickard].
9.3. ESTUARINE HYDROGRAPHY 137
Figure 9.3: Left: Hansen-Rattray classication plot, after Dyers gure 8. (The axes
are difcult to read in this scan; the labels are u
s
/u
f
and s/S
0
.) Right: Hansen-Rattray
classication diagram, after Dyers gure 8, with classication into categories 1 (A and B),
2 (A and B) and 3 (A and B), as indicated in gure 2 of Hansen and Rattrays original paper.
Ri
E
< 0.08 mixed (9.5)
Ri
E
0.3 partially mixed (9.6)
Ri
E
> 0.8 highly stratied (9.7)
Example: with currents of 1ms
1
, river input of 40m
3
s
1
, width 1km and density dif-
ference of 20kgm
3
, we get Ri
E
= 0.01, which implies well-mixed conditions. But if the
current were changed to 0.2ms
1
, Ri
E
would be 1.2, indicating stratied conditions. Thus
the system is very sensitive to U.
9.3. Estuarine hydrography
Hydrographic characteristics can tell a lot about the dynamics of estuaries.
9.3.1. Hansen-Rattray classication
Hansen and Rattray devised a classication scheme based on two parameters:
The top to bottom salinity difference S divided by the mean over the cross-section S.
The ratio U
S
/U of the surface current to the mean over the cross-section.
The Hansen-Rattray scheme is to plot S/S versus U
S
/U as in Figure 9.3, and then to
read off the type of estuary from the plot. The identied types are
Type 1 Well mixed; essentially unstratied.
Type 2 Intermediate; partially stratied.
Type 3 Highly stratied; two-layer structure.
138 CHAPTER 9. ESTUARINE CIRCULATION
Figure 9.4: Side-view of type 3 estuary, with input and output salinities S
i
and S
o
, and
volume uxes (units m
3
/s) via input and output horizontal uxes M
i
and M
o
, riverine input
F, and precipitation P and evaporation P. The sloped thick line is the interface between the
upper and lower layers.
9.3.2. Fluxes in Type 3 estuaries
Type 3 estuaries are very strongly stratied, so much so that they can be regarded as
two-layer systems. This means that we can infer the surface velocity U
S
directly from the
hydrography and known inputs and outputs.
Figure 9.4 shows the situation. Denote the evaporative loss of water E, the precipitative
gain of water P, the input of freshwater from the river F, the ux of water into the estuary
from the ocean M
i
and the ux out of the estuary M
o
. (All of these have units m
3
s
1
.) Then
conservation of mass requires
M
o
+E = P+F +M
i
(9.8)
Conservation of salt requires
M
o

o
S
o
= M
i

i
S
i
(9.9)
(where S
i
is the salinity of the (deep) water entering the estuary and S
o
is the salinity of
the (shallow) water leaving the estuary). To a good degree of approximation this can be
rewritten
M
o
S
o
= M
i
S
i
(9.10)
Hence
M
o
=
S
i
S
i
S
o
(P+F E) (9.11)
M
i
=
S
o
S
i
S
o
(P+F E) (9.12)
For those of you who are interested in learning about the symbolic-mathematics language
called Sage, see Figure 9.5.
In cases where P and E are typically very small compared to the river input, we can
rewrite these as
M
o
=
S
i
F
S
i
S
o
(9.13)
and
M
i
=
S
o
F
S
i
S
o
(9.14)
S
i
S
o
is typically very much less than S
i
, so M
o
is very much greater than F: there is a
strong amplication of the river transport.
9.4. KNUDSENS THEOREM 139
Figure 9.5: Estuarine circulation, in the Sage language. The rst line denes F, E, etc as
symbolic variables. The next two lines set up the conservation equations. Finally, the fourth
line solves for M
i
and M
o
. The solution is in written in blue; compare this with (9.12) and
(9.11), noting that Sage has written the denominator differently than has been done here.
Figure 9.6: Denition sketch for tidal ushing time, showing high and low tide levels, the
volume V below the low-tide level, and the inter-tidal volume P.
9.3.3. Flushing time
Flushing by overturning motion
The ushing time of an estuary is not the volume of the estuary divided by F. Rather, it
is given by

O
=
Vol. of upper layer
M
o
(9.15)
Flushing by tides
If V is the low tide volume of the estuary and P the intertidal volume (see Figure 9.6),
then tidal ushing gives
3

T

V +P
P
(tidal period) (9.16)
Note: this formula tends to under-estimate the ushing time, especially in cases where there
is a sill, since it assumes that the water owing in over the sill is fully mixed with all the
water in the basin, even down to the very bottom. In practice, mixing is not this complete,
and therefore the ushing time tends to exceed
T
as calculated above.
9.4. Knudsens Theorem
We can formalize the simple arguments about salt conservation provided that the hori-
zontal mixing is negligible. Clearly this approach is most useful in cases with two distinct
layers.
Consider two sections, (1) and (2) with cross-sectional areas A
1
+A
/
1
and A
2
+A
/
2
(Figure
9.7). The continuity equation gives
3
Exercise: derive this, by considering a tracer initially at concentration C
0
, when the water volume is V, and
assuming complete mixing as the volume changes to V +P, etc.
140 CHAPTER 9. ESTUARINE CIRCULATION
Figure 9.7: Denition sketch for Knudsen theorem, showing a side view. Primed variables
are in the lower, unprimed in the upper. Subscripts 1 and 2 denote two vertical planes.
F = A
1
u
1
+A
/
1
u
/
1
(9.17)
and also
F = A
2
u
2
+A
/
2
u
/
2
(9.18)
where is the net input due to precipitation or runoff at the surface or sides.
The conservation of salt equation gives
A
1
u
1
S
1
+A
/
1
u
/
1
S
/
1
= 0 (9.19)
(i.e. the sum of rightward ow of salt in the upper layer and that in the lower must vanish)
and
A
2
u
2
S
2
+A
/
2
u
/
2
S
/
2
= 0 (9.20)
These four equations can then be solved for the four unknown velocities. We can also
compute the vertical velocity w as for water continuity in the bottom layer.
A
/
1
u
/
1
= wWL+A
/
2
u
/
2
(9.21)
(where W is the width), thus telling us about the rate of dilution of the upper layer.
Further reading . Good discussions are provided by Bowden [1983], Knauss [1978],
and Pickard [1979]. The book by Dyer [1972] is especially useful. Also, see Kramer et al.
[1994] for a description of methods and techniques of physical, biological, chemical and
geological sampling of estuaries. Articles on estuarine phenomena appear in many journals,
but there is also a semi-specialized journal, Estuarine, Coastal and Shelf Science, which
you may nd helpful.
CHAPTER 10
Ocean waves
10.1. Preview: what were trying to understand
Watching where the edge of water
meets and lies upon
the unmoving shore
An elegy for D. H. Lawrence c _1935 William Carlos Williams
Theyve got waves out on the ocean
theyre gonna wear away the land
Line em up c _1997 James Taylor
W
E ARE ALL FAMILIAR with surface waves on the ocean. This is but one type of
ocean wave. There are also waves in the interior of the ocean internalwaves,
in which the the [isopycnal??] surfaces bob up and down much like the surface
does. In both surface and internal waves the restoring force is gravitational. When the period
of surface or internal waves becomes long enough to be comparable to 2/ f , the Coriolis
force starts to modify the dynamics. The Coriolis force also plays a vital role in Rossby
waves, as we have seen already.
10.2. Phase velocity and group velocity
The sea at rest is at. If we introduce a bump, the sea level will oscillate about the mean
position, the wave propagating as it does so. The motion can be complex in both horizontal
dimensions, e.g., throwing a stone in a pond, getting out of your bath, etc. Although such
motions are complicated, we can think of them as being made up of the sum of simpler wave
motions, provided that the dynamical system is [linear??]. This linear approximation is
valid provided that the wave height is much less than the water depth. It fails near beaches
(see Figure 10.1), but it often holds well in deeper water.
The simpler waves we shall consider here are plane waves, that is, long-crested waves
with sinusoidal surface shape (Figure 10.2). Denote the direction in which the wave elevation
varies x and the orthogonal direction y. Let the wave have frequency and wavenumber k.
So the elevation of the surface about a state of rest is
= acos(kx t) (10.1)
where a is the wave amplitude (so 2a is the wave height H), k = 2/L is the wavenumber
(L being the wavelength), is the angular frequency = 2/T (T being the wave period).
Equation 10.1 represents a wave moving in the direction of increasing x. The wave shape
propagates at phase velocity C given by
C
P
=

k
(10.2)
In cases where /k is a constant, waves of different wavelengths will travel at the same
speed. But if /k is a function of k say, then waves of different wavelength will travel at
different speeds: they will disperse. Thus it is important to know how and k are related.
We call this relationship the dispersion relationship, and write it in its general form
= (k) (10.3)
The phase velocity C
P
is the velocity at which the wave shape travels. There are two
other velocities of interest:
141
142 CHAPTER 10. OCEAN WAVES
Figure 10.1: A wave arriving at Maui. Q: based on the size of the surfer, how tall is this
wave?
Figure 10.2: Wave propagating to the right, = acos(kx t). Dashed line is at some
time t
0
; solid line is at time t
0
+/(2).
10.2. PHASE VELOCITY AND GROUP VELOCITY 143
0 50 100 150

1
.
0
0
.
0
1
.
0
t

1

a
n
d

2
0 50 100 150

1
0
1
2
t

1
+

2
pdf("group_velo.pdf", width=6, height=5)
t <- seq(0, 150, length.out=500)
faster <- 9.5
slower <- 10.5
par(mfrow=c(2,1))
par(mar=c(4,4,1,1))
zeta.faster <- sin(2
*
pi
*
t / faster)
zeta.slower <- sin(2
*
pi
*
t / slower)
plot(t, zeta.faster, type=l,
ylab=expression(paste(zeta[1]," and ",zeta[2])))
lines(t, zeta.slower, lty="dashed")
plot(t, zeta.faster + zeta.slower, type=l,
ylab=expression(zeta[1]+zeta[2]))
Figure 10.3: Group velocity. Top: two waves, with periods 9.5 and 10.5. Middle: sum of
these two waves. Bottom: R code used to create the diagram. (As an exercise, you might
like to try this with different frequencies, e.g. try putting the frequencies closer together, to
whether the beating frequency increases or decreases.)
144 CHAPTER 10. OCEAN WAVES
The orbital velocity: the velocity at which water particles move.
The group velocity: the direction and speed at which the wave energy moves.
We shall have little more to say about the orbital velocity.
The group velocity can be understood in the following way (Figure 10.3). Consider two
waves of equal amplitude a but slightly different frequencies and wavenumbers
k k. The equation for the total surface elevation is
= acos(kx +kx t t) +acos(kx kx t +t) (10.4)
which can be rewritten
= 2a cos(kx t)cos(kx t) (10.5)
This is a wave of frequency , wavenumber k, like the average of the two original waves,
but modulated by a long-period, long-wavelength wave with frequency and wavelength
k. This modulation wave travels at speed /k. You can imagine the signal as being
a series of groups of waves. For this reason, the speed at which the modulation travels is
called the group velocity C
G
. Its general denition is in terms of the limit as 0 and
k 0:
C
G
=

k
(10.6)
which is in general not equal to the phase speed C
P
. For some types of waves, internal waves
for example, C
P
and C
G
may even be in different directions. For surface gravity waves C
P
,
C
G
are in the same direction.
10.3. Capillary Waves
Imagine a glassy calm sea that is suddenly hit with a wind. Almost immediately, tiny
ripples will form. These are capillary waves. The restoring force for these waves is surface
tension. Capillary waves are shorter than about 3cm. Until recently the only interest in
capillary waves was as the rst stage in wave generation, but interest has been rekindled,
since it was realized that if satellites could measure capillary waves then they could estimate
the wind stress on the ocean.
10.4. Surface gravity Waves
Surface gravity waves are the waves you see at the beach or at sea.
The equations describing the dynamics of surface gravity waves are the momentum
equations
du
dt
=
1

p
x
(10.7)
and
dw
dt
=g
1

p
z
(10.8)
(with g taken as a positive number) together with the continuity equation
u
x
+
w
x
= 0 (10.9)
subject to boundary conditions
w = 0 (10.10)
at the seabed z =h and
p = 0 (10.11)
d
dt
= 0 (10.12)
10.4. SURFACE GRAVITY WAVES 145
2 1 0 1 2

1
.
0

0
.
5
0
.
0
0
.
5
1
.
0
x
t
a
n
h
(
x
)
Figure 10.4: Illustration of the hyperbolic tangent function tanh(x) (smooth curve),
illustrating that tanh(x) 1 for x 0 and illustrating that tanh(x) 1 for x 0, while
tanh(x) x for [x[ 1 (straight line through origin).
at the sea surface z = . (The atmospheric pressure has been subtracted from all equations
because it drives no motion.)
Equations 10.7, 10.8 and 10.9 can be solved fairly easily if a [linear??] approximation
is made. The details are of little interest and will not be presented here. Instead, we will
focus on the dispersion relationship that results:

2
= gktanhkh (10.13)
where h is the depth of the water. As Figure 10.4 illustrates, the hyperbolic tangent
function, tanhx, is approximately x for [x[ 1 and approximately 1 for [x[ 1. Therefore
this dispersion equation has two limits, one for shallow water (h k
1
or h L/(2)) and
one for deep water (h k
1
or h L/(2)).
10.4.1. Deep-water (or short) waves
In deep water, where kh 1, the tanh term approaches 1, so that

2
= gk (10.14)
Therefore the phase velocity is
C
P
=

k
=
_
g
k
(10.15)
which can be rewritten in terms of the wavelength L as
C
P
=
_
gL
2
(10.16)
or in terms of the period T as
C
P
=
gT
2
= 1.56T (10.17)
NOTE: this last equation is an unusual one, because the units do not match; to use this
equation, you must express the time in seconds, and then youll get the phase speed in
146 CHAPTER 10. OCEAN WAVES
metres per second. (The equation is possibly the only one in these course notes that is
dimensionally-challenged in this way.)
The group velocity is
C
G
=

k
=
1
2
_
g
k
(10.18)
so that
C
G
=
C
P
2
(10.19)
Several properties of deep-water waves follow immediately:
Long waves (with large values of L and small values of k) travel faster than short
waves. A wave of length 10m travels at 8ms
1
while a wave of length 100m travels at
25ms
1
.
Neither the phase velocity nor the group velocity depends on the depth h. This means
the waves can propagate through regions of varying water depths without changing
speed (in contrast to shallow water waves, which, as we shall see, change speed as h
changes, and therefore are refracted.)
Deep-water waves are dispersive; C
P
depends upon k. Imagine waves arriving at a
beach, having travelled from a distant storm. The long waves travel the fastest (since
C
P
is proportional to

L), so the beach will rst be pounded by long swells suitable


for surng. (Convenient.)
The group velocity is exactly half the phase velocity.
Figure 10.5 shows the speed of deep-water waves, compared to shallow-water waves.
Figure 10.6 shows the orbital trajectories for deep-water and shallow-water waves. Note
that particles in deep-water waves go around in circles, and that the circles get smaller and
smaller in deeper and deeper water.
10.4.2. Shallow-water (or long) waves
In shallow water, where kh 1, the tanh term approaches kh, so that the dispersion
relationship becomes
= k
_
gh (10.20)
Therefore the phase speed is
C
P
=

k
=
_
gh (10.21)
and the group speed is
C
G
=
_
gh (10.22)
Shallow water waves are thus very different from deep water waves.
Shallow water waves are not dispersive; the phase velocity does not depend upon k or
.
The group velocity is the same as the phase velocity.
Neither the phase velocity nor the group velocity depends upon the wavelength of the
wave.
Both the phase velocity and the group velocity decrease as the water depth h decreases.
This leads to wave refraction (Figures 10.7, 10.8).
Figure 10.5 shows the speed of shallow-water waves.
Figure 10.6 shows that the orbital trajectories for shallow-water waves are ellipses.
10.4. SURFACE GRAVITY WAVES 147
Figure 10.5: Phase speeds of deep-water waves and shallow-water waves [Pond and Pickard,
1978, Fig12.3].
Figure 10.6: Particle orbits in deep-water waves and shallow-water waves. (Pond and
Pickard Figure 12.4).
148 CHAPTER 10. OCEAN WAVES
Figure 10.7: Refraction of shallow-water waves. As waves move from deeper water to
shallower water, they slow down. Therefore a wave travelling obliquely to isobaths tends
to turn into the shallow water. The result is that waves turn to approach the shore dead on.
(Pond and Pickard Figure 12.5)
Figure 10.8: Focussing of shallow-water waves on headlands. (Pond and Pickard Figure
12.6)
10.4. SURFACE GRAVITY WAVES 149
Figure 10.9: Denition sketches of various resonant standing waves (i.e. seiches), from
Pond and Pickard Figure 13.4. Top left: closed-region mode (half-wave resonance). Top
right: open-ended mode (quarter-wave resonance). Points A, B, etc are at different
phases of the wave; point C is called the node.
10.4.3. Standing Waves, Seiches, Resonance
Standing waves may result from oscillations of the water surface (so-called external
modes) or from the [isopycnal??] surfaces within the water column (so-called internal
modes). The next two subsections deal with these in turn.
External modes
Imagine water sloshing back and forth in your bathtub. You can make it slosh by jumping
into the tub
1
or by waving your hands back and forth underwater. In fact, if you wave your
hands at the right frequency, you can get some pretty dramatic sloshing. This is called
resonant forcing. In fact, there are several oscillations possible in a tub, e.g. from end to
end, from side to side, etc. There are resonances in which at a given instant of time the
water is sloshing high at one end and low at the other. There are also resonances in which
the water is high at the ends while it is low at the center, and then high at center and low at
the ends.
These different pattersn of resonance are called resonant modes. Your bathtub has
many such modes. So does a guitar string; musicians (and physicists) talk of the so-called
fundamental mode and of harmonic modes. The fundamental mode is the one with the
lowest frequency. This is also called the gravest mode.
Let us focus on the gravest mode for the time being that is, we are considering the
resonant mode with the longest wavelength. (It turns out that this longest-possible mode has
the lowest-possible resonant period.)
Figure 10.9 illustrates. The rst thing to notice is the form of the deformations in the sea
surface. In a closed domain, such as a lake, one-half of a full wavelength can t within the
domain. This is illustrated in the left panel. In a semi-closed domain, such as a harbour or
a fjord, the gravest mode has one-quarter of a wavelength. If we let the horizontal scale
be denoted L, then, and the wavelength , we have L = /2 for the left-hand panel and
L = /4 for the right-hand panel.
Now, how may we determine the periods of these oscillations? The answer stems from
the realization that a standing wave can be envisioned as the superimposition (addition) of
two identical propagating waves, travelling in opposite directions. (In the gure, imagine
one wave travelling to the right, the other to the left. Can you see how a standing wave
might result? If not, try drawing some pictures or doing the mathematics.)
In the gure, and in applications of interest, we denitely have the shallow-water variety
of waves, since the wavelength (roughly comparable to the basin length) is much greater
than the water depth. (Example: a harbour might be 10km long, and just 100m deep.) Thus,
the speed of the waves is given by (gh)
1/2
. Since the speed is the ratio of wavelength to
period, we have a way to predict the period (, say) of resonance.
For half-wave resonance (left panel in Figure 10.9) we have = 2L/(gh)
1/2
. For quarter-
wave resonance (right panel), we have = 4L/(gh)
1/2
.
1
Messy, but fun.
150 CHAPTER 10. OCEAN WAVES
Figure 10.10: Denition sketch for calculation of wave potential energy.
For example, the Bay of Fundy is roughly 200 km long, and roughly 50 m deep, and
therefore resonates at a period of roughly 10 hours. This is close to the period of the
semi-diurnal tides (chapter 13), which gets close to explaining why this region has famously
high tides. (In fact, a more complicated model is required to the domain adequately, and this
involves considering water in the Gulf of Maine as well as in the Bay of Fundy. If youd like
to learn more, look for papers by Chris Garrett on this, written in late 1970s or early 1980s.)
Internal modes
All of the reasoning in the previous section carries over to internal modes. The only
difference is that the speed of the waves is different. But youll have to skip ahead a bit to
see the formula for internal-wave speed. For a two-layer approximation to the stratication,
youll nd the speeds in section 10.5.1, for example.
10.4.4. Energy in surface gravity waves
The potential energy associated with the wave motion can be found by taking a suitable
average over a wavelength or wave period. We shall write the waves energy E as energy per
metre along the crest.
Figure 10.10 shows a small volume of uid which has thickness dx measured in the
direction of wave travel. The mass of this volume (per unit metre along the crest) is dx,
and its centre of mass is raised /2 above the base state, so its potential energy is
dPE =

2
gx
2
(10.23)
so that, summing over all such elements in one wavelength, we get
PE =
_
L
o
g
2
2
dx (10.24)
Since the wave is of form = acos(kx t), its average over the wavelength is a/2.
Thus the potential energy per unit metre is
PE =
1
4
ga
2
(10.25)
Similarly, the kinetic energy is given by
KE =
_
o
h
_
L
o
g
_
u
2
2
+
w
2
2
_
dx dz (10.26)
which, after some algebra, yields
KE =
1
4
ga
2
(10.27)
The total energy, the sum of potential plus kinetic, is thus
E =
1
2
ga
2
(10.28)
The energy travels at the group velocity, so the energy ux is EC
G
. When waves are not
breaking the energy ux is conserved.
10.5. INTERNAL GRAVITY WAVES 151
Figure 10.11: Side view of wave at interface in two-layer uid. Upper layer density is
1
,
while the lower layer density is
2
.
10.4.5. Momentum in surface gravity waves
Waves may lose energy by breaking, but momentum is conserved, because momentum
is only changed when forces act on the uid. (Actually, there is some loss of momentum
to friction when waves break, but the change in momentum is very much smaller than the
change in energy.)
Exercise 10.1 Bow-waves in Harbour
a
My apartment overlooks Halifax Harbour. Often I see large boats (e.g. QE2)
cruising along the edge of the harbour. If such a boat runs 50m from the harbour
edge, how long will it take the bow waves from the boat to reach the wharf?
a
Answer in chapter 21.
Exercise 10.2 Tsunami waves
a
Its a hot summer day in Halifax, and Im sitting by the shore, sipping a mint
julip
b
. My nerd oceanographer drinking buddy is an expert in Tsunami waves
(created by earthquakes at sea), and she has a beeper that goes off, indicating that
her seismometer has just detected an earthquake 100km offshore. How long do
we have to nish our drink?
a
Answer in chapter 21.
b
Hey, I said it was hot, so better drink a southern drink!
10.5. Internal gravity waves
10.5.1. Two-layer internal gravity waves
In surface gravity waves, the pressure forces which drive the motion result from tilts in
the surface. In internal gravity waves, the restoring forces result from tilts of [isopycnal??]
surfaces.
The simplest case to imagine is a two-layer uid (Figure 10.11). To see this, mark two
points along the bottom edge of this gure. Put one, named A, below the upward crest of
the wave, and put the second, named B, below the trough. Now, consider the pressure at
these two points. You should be able to see that the pressure at point A must be greater than
that at B. (There is more of the (heavier) lower layer uid above A.) The resulting pressure
gradient tends to force uid from A to B. There is a direct analogue to surface waves, and in
fact for two-layer internal gravity waves the central result, that is, the phase speed, is very
similar to that for surface gravity waves:
152 CHAPTER 10. OCEAN WAVES
Figure 10.12: Nonlinear internal wave at interface. [Figure 5, Sandstrom and Elliot, 1984]
C
Pi
=
_
gh
1

2
(10.29)
where
2
and
1
are the densities of the lower and upper layers respectively, and h
1
is the
thickness of the upper layer (presumed to be much thinner than the lower layer). Note that
this is the same as the formula for shallow water waves for a water column of the same
thickness as the upper layer, except that the acceleration g of gravity is reduced markedly by
the small term (
2

1
)/
2
.
This wave speed is for [linear??] waves. But quite a lot of the current interest in internal
waves centres on [nonlinear??] waves. These appear as isolated bumps on the interface,
rather than as simple sinusoidal wave trains. The bumps are very prominent, that is, the
elevation of the [isopycnal??] surfaces is not small compared to the mixed layer depth; that
is why the waves are nonlinear. Figure 10.12 shows an example from a paper by Sandstrom
and Elliot of the Bedford Institute of Oceanography (Sandstrom and Elliot [1984]). The
isolated waves, or solitons in current parlance, are denoted S1 and S2. It is thought that
such solitons, which are created as the tides sweep over changing bottom slopes at the shelf
edge, might play a signicant role in transporting nutrients to the continental shelf.
10.5.2. Continuously-stratied internal gravity waves
There are also waves within the thermocline. A large body of literature is concerned
with these waves, but Ill say nothing about them here in the interests of brevity. One point,
though: the whole eld of internal waves has been re-visited in a very big way in recent
years . . . see the assigned readings for more on that!
10.6. Effects of rotation
When the wave period becomes long enough to be comparable to 2/ f , waves feel the
inuence of rotation. This is the case, for example, with tides, for the tides are nothing more
than waves with slow period.
CHAPTER 11
Turbulence and mixing
Big whorls have little whorls,
Which feed on their velocity;
And little whorls have lesser whorls,
And so on to viscosity.
L. F. Richardson
T
URBULENCE AND MIXING ARE RELATED TOPICS partly because our interest in tur-
bulence arises from its ability to mix momentum and uid properties. We care about
the mixing of momentum because it indicates all manner of things, e.g. the thickness
of the Ekman layer. We care about the mixing of uid properties for reasons relating to
applications ranging from primary production (e.g. mixing of nutrients from deep waters up
toward the surface, where light permits photosynthesis) to climate (e.g. mixing of warm
water down from the surface inuences near-surface temperatures and thus inuences the
air-sea exchange of heat; and the simplest climate models view the ocean as a simple slab of
motionless water, with heat capacity corresponding to the mixed-layer thickness.)
11.1. Characteristics of Turbulence
Roughly speaking, turbulent ow is disordered ow ow that is difcult to predict from
moment to moment. But precisely speaking, turbulence is not easy to dene. In fact it is
usually dened in a negative sense, i.e. if a uid ow has such-and-such a property, then
it is not turbulent. For example, if ow is repeatable (like a wave) then it cannot be called
turbulent.
It might help to list a few properties of turbulence.
Irregularity of motion uctuating randomly about the mean u so
u = u+u
/
(11.1)
These uctuations produce perturbations of all the water properties, , S, T, nitrate
etc.
Note that the statistical description of this irregularity is often difcult as the turbulence
appears to intensify in bursts or events distributed in time at much lower frequencies
than the individual uctuations.
Efcient mixing. Under ocean waves one would also see an irregular pattern in
velocity and water properties. However the properties advect with the water, returning
to mean properties as the wave passes (to rst order)
A term such as u
/
T
/
/x where the overbar represents a suitable time average will be
small for wave motions
- not small for turbulent
- This means that a [parcel??] of uid cycling back and forth may change its T
through a cycle.
The turbulent velocity eld has vorticity
- in all three directions. The scales are not necessarily the same for vertical and hori-
zontal directions at large scales but tend to a condition of symmetry in all directions
(isotropy) at small scales.
153
154 CHAPTER 11. TURBULENCE AND MIXING
The measure of the tendency for the ow to be turbulent is the Reynolds Number,
Re =UL/ the ratio of the interial term in the acceleration to the molecular viscous
term, where U is a typical scale for speed, L is a typical length-scale over which
velocity varies, and is the molecular [kinematic-viscosity??] (see section 5.3.3 on
page 91).
Oceanic motion is usually in the high Re range but, as we will see, molecular processes
can be locally very signicant
Dissipation of turbulence - at small scales a cascade of eddy sizes occurs.
Large eddies break down into even smaller eddies until the velocity gradients are so
large that molecular viscosity becomes a very effective dissipation mechanism and
energy is lost as heat.
[At larger scales, the energy is transferred to both smaller and larger scales, particularly
when the vertical motion is smaller 2-D turbulence)
11.2. Measuring turbulence
Turbulent intensity is often dened, or measured, in terms of the rate of viscous
dissipation of turbulent kinetic energy. This is sometimes loosely called the [dissipation-
rate??] and it is normally designated with the symbol . It has units of W/kg, or equivalently
m
2
/s
3
. Physically, is the rate at which the water
1
is heated by friction between turbulent
eddies rubbing against each other. The W/kg unit indicates this.
Values of in the ocean are 10
10
W/kg (roughly the noise oor of instruments) to
perhaps 10
5
W/kg or so. Actually, it is difcult to say what typical values are, since
varies very widely and has been measured so seldom
2
.
11.3. Turbulent scales
We often think of turbulence as being forced at a large scale, and creating smaller and
smaller scales (eddies, if you will) with a lower limit on scale being set by the viscosity of
the uid
3
. The large scale could be the shear of steady or tidal currents rubbing on the ocean
bottom, or the breaking of internal waves; the point is that these scales are set, or controlled,
by some external agency that is independent of the turbulence.
11.3.1. Ozmidov Scale
In a stratied case, there is natural scale called the [Ozmidov-scale??] that represents the
competition between the tendency of turbulence to mix uid and the tendency of stratication
to inhibit mixing. From the last sentence, you can see that the scale should include both
a mixing parameter and a stratication parameter. Well, from the last section we have
identied a key mixing parameter as , and you know by now that the key stratication
parameter is N. The question is, can these two parameters be combined in such a way as to
create a quantity with length units. Well, sure. The unit is m
2
/s
3
. The N unit is 1/s. Thus,
/N
3
has unit m
2
, i.e. the product gets rid of the time unit. So, we just take the square root
of this to get a scale, which we call the [Ozmidov-scale??]:
L
o
= (/N
3
)
1/2
(11.2)
Often, the Ozmidov scale is taken as an indicator of the eddy scale. For example, the
Ozmidov scale is typically comparable to the scale of overturning motions, measured by the
[Thorpe-scale??], a measure of overturning motions inferred from CTD proles Galbraith
and Kelley [1996].
Note that L
o
increases as increases, i.e. the Ozmidov scale gets larger (or, the eddies
get larger), as the turbulence gets more intense.
1
Or whatever uid were talking about; the analysis is quite general, and the same notation is used in atmospheric
research, for example.
2
There are only a few groups in the world capable of measuring ; we have such folks in this department and
across the water at BIO.
3
Thats what the Richardson quote, at the start of this chapter, is about.
11.4. EFFECT OF TURBULENCE ON THE MEAN FLOW 155
11.3.2. Batchelor Scale
As the Ozmidov scale measures the competition between mixing and stratication, the
so-called [Batchelor-scale??] measures the competition between mixing and friction. As
before, the mixing is measured by . We measure the friction by the viscosity [nu??], which
has unit m
2
/s. As an exercise, you should verify that the following quantity has the unit of
length:
L
b
=
_

3
/
_
1/4
(11.3)
this is called the [Batchelor-scale??].
In contrast to the Ozmidov scale, the Batchelor scale gets smaller as the turbulence gets
stronger. Typically, L
o
> L
b
, and the window of scales between L
b
and L
o
gets wider as
the mixing gets more intense.
11.4. Effect of turbulence on the mean ow
We start by decomposing the velocity into mean (overbar) and deviating (prime) compo-
nents, i.e.
u = u+u
/
(11.4)
where () represents a temporal average of some quantity over some timescale that exceeds
the timescale of turbulent motions. That is, () =
1
_
t+
t
()dt where is the averaging
timescale. Then, taking this time integral of (11.4), we see that u
/
= 0.
Consider the equation of motion, for the x-component
u
t
+u
u
x
+v
u
y
+w
u
z
f v =
1

x
+
2
u (11.5)
where

2
=
_

2
x
2
+

2
y
2
+

2
z
2
_
(11.6)
and continuity requires
u
x
+
v
y
+
w
z
= 0 (11.7)
Combining the above and then taking a time average yields
du
dt
f v =
1

p
x
+
2
u

x
(u
/
)
2


y
(u
/
v
/
)

z
(u
/
w
/
) (11.8)
The turbulent terms disappear from the equations for the mean ow (averaged over a suitable
time), long cf. the turbulent time scales) except in the non-linear advection terms. They now
appear as stresses forcing or opposing the mean ow. the Reynolds stresses (cf. Reynolds
No)
If assume an eddy viscosity A
V
for which
A
V
u
z
=u
/
w
/
(11.9)
then the turbulent term above appears as

z
_
A
V
u
z
_
(11.10)
which has the same form as the molecular viscosity term, with A
V
taking the place of .
However, generally the eddy viscosity A
V
is much larger than the molecular viscosity .
Remember that the above is a denition that may be useful if we cannot measure the
u
/
w
/
terms directly. However, if we can estimate or measure u
/
w
/
directly we should do so
instead.
156 CHAPTER 11. TURBULENCE AND MIXING
Parameter Deep Sea Shelf Atmosphere
Horizontal current U 0.03 m/s 0.30 m/s -
Friction velocity u

0.001 m/s 0.01 m/s 1 m/s


Eddy viscosity 210
5
m
2
s
1
210
2
m
2
s
1
-
Laminar layer thickness 0.02 m usually zero zero
Logarithmic layer thickness 1 m 10 m 100 m
Ekman layer thickness 5 m 50 m 500-1000 m
Table 11.1: Typical values Wimbush and Munk [1970]. Caution: some of these values are
very uncertain.
11.5. Bottom Boundary Layers in Rotating Fluid
A frictional boundary layer is a region in which the frictional stresses become signicant.
Since the water velocity at a solid surface falls to zero, a shear must occur between the solid
surface and any mean ow. The question is, how does the velocity vary with distance from
the surface?
Suppose that the basic balance away from the bed is geostrophic and the density variation
close to the bed is small so that the pressure gradients associated with the mean ow are the
same as those in the boundary layer.
The momentum equations for the layer, driven by a pressure that varies in the x
f v =
1

P
x
+

z
_
A
V
u
z
_
f u =

z
_
A
V
v
z
_ (11.11)
The boundary conditions are as follows. Far above the boundary layer we have u = 0
(since P = P(x) so that P/y vanishes) and v = v
0
where
f v
0
=
1

p
x
(11.12)
and down at the seabed we have u = 0 and
A
V
u
z
=
b
(11.13)
where
b
is the bottom stress, perhaps given by a quadratic stress law akin to that used at the
air-sea boundary.
The above yield different solutions in different domains. The next few sections detail
these domains, starting with the geostrophic layer far from the boundary and ending with the
viscous laminar layer just next to the boundary. Remember that throughout this discussion, z
measures distance above the boundary (that is z = 0 is at the ocean bottom, not at the ocean
surface as it is through most of the rest of these notes.)
11.5.1. The Geostrophic Layer
As stated above, far from the boundary layer we have (u, v) = (0, v
0
), where the axes
have been aligned so that pressure varies only in the x direction. Geostrophy thus yields
v
0
=
1
f
p
x
(11.14)
11.5.2. The Ekman Layer
If A
V
is constant with respect to z, then the A
V
terms in the above equations may be
brought outside the derivative, yielding the much simpler system
(v v
0
) =
A
V
f

2
u
z
2
u =
A
V
f

2
v
z
2
(11.15)
11.6. MIXING, DIFFUSION AND DISPERSION 157
The solution to these equations, i.e. to the so-called bottom Ekman layer problem, is
straightford for these constant coefcients. If A
V
varies, the solution is more complicated of
course.
11.5.3. Constant-stress Layer (aka Log Layer)
Below the Ekman layer, we may expect velocity to decrease, i.e. u 0, so that (returning
to (11.11)) we see that

z
_
A
V
u
z
_
0 (11.16)
and thus

z
0 (11.17)
In other words the stress, , is constant, and equal to the boundary stress
b
, i.e.
(z) =
b
(11.18)
Such a layer is called a constant stress layer. It is also known as a log layer, since the
velocity varies as the logarithm of the distance above the bottom. Why? Because empirical
evidence (and simple theoretical considerations) suggest that that A
V
is linear in distance
from the wall, i.e. A
V
= z where 0.4 is a constant known as Von Karmans constant.
Thus, we have
z
u
z
=

= u

(11.19)
which has the solution
u =
u

ln
_
z
z
0
_
(11.20)
where z
0
is a constant of integration known as the roughness length, named this way because
studies have suggested that it is related to the roughness of the bottom
4
. The logarithm in
this formula yields the name log layer.
11.5.4. Laminar Layer
Below this constant-stress layer, we may have a layer in which turbulence vanishes and
we have laminar ow:

u
z
= u
2

(11.21)
where u

, dened as (
b
/)
1/2
, is known as the friction velocity. Solving yields
u =
u
2

(11.22)
in the laminar sub-layer.
11.6. Mixing, diffusion and dispersion
References Textbooks: [Knauss, 1978, chapter 4]; [Pond and Pickard, 1978, pages 136-
140]. Research papers: Okubo [1971] (general, summarizing many studies); Garrett [1984]
(icebergs).
The presence of turbulent eddies means not only transfer of momentum but also of the
other water properties, temperature, salinity, nutrients buoyancy (i.e. density) etc.,
Typical equation for a conservation property S is
S
t
+u
S
x
+v
S
y
+w
S
z
= K
s

2
S+I (11.23)
4
For example, empirical results of ow over sandy bottoms suggest that z
0
= D/30, where D is the size of the
grains of sand Nikuradse [1932].
158 CHAPTER 11. TURBULENCE AND MIXING
where I is a source or sink of S. Here, K
s
is the molecular diffusivity (basically a function of
the size of the molecule).
As usual we have the continuity conditon
u
x
+
v
y
+
w
z
= 0 (11.24)
so we may write
S
t
+
(uS)
x
+
(vS)
y
+
(wS)
z
= K
s

2
S+I (11.25)
and then letting S consist of a mean, S plus a uctuation S
/
.
S = S+S
/
(11.26)
where
S
/
= 0 (11.27)
we get the equation for the change in S on timescales long compared to the turbulent
uctuations.
S
t
+u
S
x
+v
S
y
+w
S
z
= K
s

2
S+I (11.28)
where
I =
u
/
S
/
x
+
v
/
S
/
y
+
w
/
S
/
z
(11.29)
where u
/
S
/
is the transport of S by the turbulent velocity eld.
Youll see that the reasoning here is just like that for Reynolds stresses. And, similarly,
the next step in the analysis is to dene eddy diffusivities K
H
and K
V
through equations of
the form
u
/
S
/
= K
H
S
x
(11.30)
v
/
S
/
= K
H
S
y
(11.31)
w
/
S
/
= K
V
S
x
(11.32)
Combining the above, we can express turbulent uxes in terms of large-scale (overbar)
quantities. Formerly, at least, this makes sense. But it is of no practical utility unless we
can then dene K
H
and K
V
in some reasonable terms; the quantities will not be of any use
if they need to be dened differently instant by instant, spot by spot, in the ocean. The
mathematical analysis only proves to be of use if we can further related K
H
and K
V
to some
large-scale properties of the ocean, or of the forcing. This is the parameterization problem
and it is at the heart of most studies of mixing. The research community is a very long way
from knowing how to write trustworthy formulas for K
H
and K
V
.
But lack of trust does not translate to lack of use! We need to make progress on a wide
variety of problems relating to mixing, and so it is very common to just formulate a theory
in terms of, say, xed K
V
, and then to try to infer its value from observations. One way to
do that is to inject a dye into the water and to track its rate of spread, comparing the latter to
the predictions of models with varying values of K
H
and K
V
. For example, if a thin patch
could be put into the water over a wide horizontal area, the value of K
H
would drop out of
the problem (at least, near the centre of the patch), so we would be left with
C
t
= K
V

2
C
z
2
(11.33)
11.6. MIXING, DIFFUSION AND DISPERSION 159
and so the analysis could consist of measuring C =C(z) at one time, and then a later time,
and computing the LHS of the equation by differencing over time, and computing the
RHS from differencing over space; plotting these quantities would let K
V
be estimated by
regression
5
. The point is that such an experiment could give us a value of K
V
. This would
be tabulated, along with other relevant quantities (N, vertical shear, Richardson number,
etc), and added to a database. The goal would then be to expand the database to include a
wide range of parameter space (e.g. covering typical values of N), in hopes of determining
an empirical relationship between what we know (can measure or can model), such as N,
and what we want (uxes).
Progress on such work is proving to be rather slow. The reason is partly that it is so
difcult to make measurements at sea, and partly that turbulence is simply a complicated
phenomenon. (Understanding turbulence is one of a handful of grand problems left in
Physics.)
Further reading about mixing and turbulence. The textbook of Tennekes and Lumley
[1972] provides a good introduction to turbulence, although at a level slightly higher than the
present course. Truthfully, the same can be said about the level of any relevant material, e.g.
review papers. Ive provided this chapter mainly just to familiarize you with terms. If you
need to work on mixing in your thesis topic, say, then youll want to do that by collaborating
with a specialist. As an example of the application of mixing analysis to estuarine situations,
and as a provider of recent references to the literature, you might like to consult Peters
[1999]. A major thrust of mixing work in recent years has focussed mixing induced by
internal waves. Since such waves may be expected to have a meridional dependence, we
might expect a similar result for mixing. This issue is taken up by Gregg et al. [2003], who
propose that deep-ocean mixing is a function of latitude. Much of my own research program
focuses on mixing driven by internal waves, except in coastal waters.
5
Actually, this is not the way this is done, but I dont want to get too distracted on the solution of the above-stated
DE. See Ledwell et al. [1986] for a classic introduction to the dye-release method. Or see Kelley and Van Scoy
[1999] for an application based on bomb-injected tracer.
160 CHAPTER 11. TURBULENCE AND MIXING
CHAPTER 12
Light in the sea
This chapter under construction; see online version for updates
May the light in the land of plenty
Shine on the truth, someday.
Land of plenty c _2001 Leonard Cohen
12.1. Electromagnetic radiation
12.1.1. Spectrum
A whole spectrum of EM (electro-magnetic) radiation is active on this planet, including
e.g. gamma rays, radio waves, etc., in addition to visible light. This EM radiation is visible
if it has wavelengths of roughly 300 to 800 nm (nanometers).
12.1.2. Speed of Light
There is no dispersion of light, in the sense that deep-water waves disperse that is, all
wavelengths travel at the same speed. In a vacuum, this speed is C = 310
8
m/s. In sea
water, the speed is reduced to 2.310
8
m/s, and varies with the chlorinity of the water.
12.1.3. Index of Refraction
If we denote the light speed in the medium by V, then the so-called index of refraction is
dened as n =C/V. Thus, from the above, seawater has n 1.33.
12.1.4. Net radiation
All bodies radiate a spectrum of electromagnetic radiation. We can dene I

() as the
spectral density of energy ux per unit increment in centred at wavelength .
Energy ux in wavelength band
1

2
is therefore
_

2

1
I

()d (12.1)
in units of W m
3
(power per area, per wavelength), and therefore the total energy radiated
is
I
0
=
_

0
I

()d (12.2)
in units of W m
2
.
For example, the intensity of light from the sun impinging on the earth is of order
700 W/m
2
at noon on a clear day.
12.1.5. Stefan-Boltzmann law for black-body radiation
All objects which are at temperatures above absolute zero emit radiation, and the intensity
of this radiation depends on the temperature. In fact, the dependence on temperature is very
strong, going as the fourth power; the so-called Stefan-Boltzmann law expresses this:
I
0
= T
4
(12.3)
where T is the absolute temperature measured in K, degrees Kelvin
1
and is a universal
constant known as the Stefan-Boltzmann constant, of value 5.67010
8
W/(m
2
K
4
).
For an example, the earth is at roughly T = 283

K, so that it emits I
0
= 365 W/m
2
. On
the other hand, the much hotter sun is at T = 6000

K, and it emits I
0
= 10
8
W/m
2
.
161
162 CHAPTER 12. LIGHT IN THE SEA
Figure 12.1: Blackbody radiation according to Stefan-Boltzmann law. Note points indi-
cating temperature of earth and sun, and the associated radiative uxes, in watts per unit
area.
12.1. ELECTROMAGNETIC RADIATION 163
Figure 12.2: Weins law for the color spectrum of blackbody radiation. The peak of the
spectrum is at wavelength
max
= /T, where = 2.9 10
3
m

K. Caution: this gure


uses wavelengths in nm, or 10
9
m, so the constant appears as 2.910
6
nm

K in the gure.
12.1.6. Weins Displacement Law for Spectral Peak
The radiation from a warm object is in many wavelengths, and the distribution amongst
these is given by Weins law. It states that higher temperatures yield smaller wavelengths,
and that the peak of the spectrum will be at the wavelength

max
= /T (12.4)
where = 2.910
3
m

K is a universal constant.
To visualize this, think of a piece of metal, heated at various temperatures. At relatively
low temperatures, it will glow red, and as it gets hotter, it will start to glow yellow. This is
consistent with Weins law, since red light has a longer wavelength (higher ) than yellow
light.
For example, the sun at T = 6000

K emits radiation with peak power at the wavelength


5 10
7
m, or 500 nm. This is viewed as a yellow hue. Not surprisingly, given how
evolution works this wavelength is in the middle of the visible range.
To be more specic, 90% of the EM radiation from the sun is in range 150400 nm, and
50% is in visible 360780 nm.
12.1.7. Effect of radiation at the sea surface
Refraction
Reection
12.1.8. Effects on radiation passing through sea water
Further reading . References: Knauss, p. 262-267; Jerlovs Optical Oceanography.
1
Recall that the absolute temperature is 273

C plus the Celcius temperature, and that the symbol



K is used for
the unit of absolute temperatures.
164 CHAPTER 12. LIGHT IN THE SEA
CHAPTER 13
Tides
There is a tide in the affairs of men,
Which, taken at the ood, leads on to fortune;
Omitted, all the voyage of their life
Is bound in shallows and in miseries.
Julius Csar, Act iv. Sc. 3, William Shakespeare
T
IDES MAKE THE SEA SURFACE GO UP AND DOWN, and the water slosh to and fro.
The vertical movement is important in various navigational applications (e.g. ships
have to wait for high tide to enter or leave certain harbours), for ooding applications
(a storm surge at high tide is more damaging than one at low tide), etc. The horizontal
movement is important for all the reasons that currents are important, including its effect on
navigation, pollutant dispersal and transport, and so forth. In some areas, tides may also
provide a worthwhile source of CO
2
-free power.
Tides have been part of human experience for a long time. For example, the early
travelers from the Mediterranean observed with surprise the large sea-level changes on the
Atlantic shores of Northern Europe. Every mariner pays attention to tides, as do those who
visit beaches with sand ats.
A common three-word description for Physical Oceanography is tides and currents.
The purpose of this chapter is to explain that rst word, if only a little. Lets start with a
diagram. Figure 13.1 shows the semidiurnal lunar tide (called M2 in tidal literature
1
in
the North Atlantic, computed using sparse measurements available in the 1970s. Later on,
you should compare this with Figure 13.9, constructed using satellite measurements which
cover the whole area continuously. But for now, study the diagram until you can see that the
tide (north of 20N, say) takes the form of a wave that circles the north Atlantic, centred at
an [amphidromic-point??] off Newfoundland
2
.
13.1. Measuring tides
A reasonable measurement of sea-level elevation through a tidal cycle can be obtained
simply by observing sea level on a graduated pole. In doing this, it is necessary to average
across high-frequency waves, except in cases with very large tides and small waves. In the
mid-1800s, a simple device called a stilling well was invented to get around this wave
problem. The device is illustrated in Figure 13.2. A cylinder (white, in this diagram) is
open at the top and the bottom. If the openings are large enough, then water in the cylinder
would track water outside. But, if the water opening is made small enough, then water
will enter the cylinder so slowly that water in the cylinder will not be able to keep up with
high-frequency waves. In this way, the waves are stilled. Normally the system is tuned to
smooth over waves of periods less than 5 minutes or so. The other key component, more
visible in the diagram, is a oat that sits on the surface of the water within the cylinder. It
is connected to a pulley with a wire, and tension is kept on this wire (e.g. by a weight), so
that the pulley turns as the oat rises or falls. Then, its a simple matter to hook up a pen
to drag across a slowly-rotating drum, to produce a graph, or to hook up electronics that
translate the pulley rotation into a signal that is recorded locally and/or transmitted to a data
repository.
Stilling wells are awkward to locate on an open coast and are almost always found in
harbours. Since the motivation for measuring the tides in such circumstances is navigation,
tide gauges are sometimes installed by the local harbour authorities.
1
The exact meaning of M2 will become clearer a little later on; the important thing is that it is a twice-daily
tidal constituent that is signicant in many regions.
2
Exercise: Determine which way the wave goes, and at what speed.
165
166 CHAPTER 13. TIDES
Figure 13.1: M2 tides in the North Atlantic. Dashed ([cotidal-line??]s) indicate tidal
amplitude, while solid ([cophase-line??]s) indicate phase. Note the sparseness of deep-
ocean data. (Source: Figure 1 of Brown [1984], based on measurements reported by other
authors in 1978.) Compare with Figure 13.9.
Figure 13.2: A stilling well for measuring tidal elevation even in the presence of
high-frequency waves. (Source: http://oceanservice.noaa.gov/education/
kits/tides/media/supp_tide10b.html) On the pier you can see a tide staff,
essentially a graduated pole, and a tide house containing a pulley connected to the actual
stilling well.
13.2. TIDAL FORCING 167
Figure 13.3: Viewof earth-moon pair, rotating about a common point named the barycenter.
Source: http://tidesandcurrents.noaa.gov/restles3.html.
Another approach is to measure the pressure at the sea bottom (or some other xed sub-
surface location). By the hydrostatic relationship, variations in pressure can be translated
into changes in sea level, provided that atmospheric pressure is also measured
3
.
13.2. Tidal Forcing
13.2.1. Nontechnical explanation of tidal forces
Most people know that tides are caused by the gravitational pull of the moon (and the
sun) on the rotating earth. Now, the earth rotates once per day, so how can it be that there
are two tides per day, in many locations? Consider just the moon, for example. Shouldnt its
gravity pull the water towards it, making the earth bulge in that direction . . . and so shouldnt
we get a high tide when the moon is directly overhead, just once per day?
The observed fact is that the water bulges out on the side closest to the moon, and the
opposite side, yielding two tides per day, roughly
4
. How can that be?
Well, there are simple answers and more complicated answers. Lets start with the basics.
The tidal force is gravitational, and such forces are proportional to the reciprocal of the
square of distance, Force 1/Distance
2
. Take a glance at Figure 13.4, and imagine that the
earth is completely covered by water. Focus on the points marked A and B in the diagram,
and think about what happens to a parcel of water at each point. The moons mass exerts
a (small) gravitational force at each point, but its stronger at B than at A, simply because
B is closer to the moon and so the reciprocal of the distance is larger there. This explains
why water would bulge out at B. Now, imagine that the earth is rotating about its centre,
along an axis oriented vertically on the diagram. That will mean that the water will always
be bulged up underneath the moon. The moon is almost xed (it rotates around the earth
once a month, and a month is a fair bit longer than a day), and so, to rst order, you can see
that there would be a bulge of water that, to a person on the earth, follows the moon through
the sky.
Thats one tide a day. But we have two tides, so now we need a new idea. This is to
recognize that the earth and moon are spinning around each other, which brings centrifugal
3
You should be able to nd a mathematical explanation for a crude rule used in the extraction of atmospheric
pressure signals, namely that 1 millibar in the air matches 1cm in the water.
4
Roughly because the various tidal components interact, etc. This will come up shortly.
168 CHAPTER 13. TIDES
A B


Moon
(mass )
Earth
(mass )
Figure 13.4: Denition sketch for explanation of tidal forces. (Not to scale.)
effects into play. Figure 13.3, from a useful NOAA website on tides, illustrates this idea
of the earth and moon rotating around a common centre. But, before we consider the
earth-moon system, lets back up a moment and consider a simpler case. Imagine that the
earth and moon had equal mass. In that case, one thing would not just spin around the other;
they would both spin around the centre of mass. And, for equal masses, that centre would
lie at the midpoint between the bodies. Now, imagine a coordinate system that is centred
at this centre of mass, and that rotates with the objects. In that system, you would see no
rotation. What youd see is two objects that are drawn together by gravitational forces, but
that are not moving toward one another. You would say that there must be a force repelling
the objects. In Physics, we call this the centrifugal force; it is a ctional force, which is a
phrase meaning that the force arises only by this choice of a rotating coordinate system.
It is this, the centrifugal force, that explains the second tide of the day. For, where gravity
makes a bulge at point B, the centrifugal force makes a bulge at A, which is on the outside
of the spinning earth-moon system
5
. So, thats it. Gravity makes one bulge, centrifugal
force makes another. Two bulges per day.
13.2.2. Technical explanation of tidal forces
Of course, there are some details to consider. You are probably aware that the moon is
much less massive than the earth, and so the centre of gravity is not halfway between the
two objects. Actually, it is inside the earth! But, as it turns out, this fact does not affect the
main concepts involved; we are still left with two bulges per day.
The actual tidal forces can be worked out from Newtons law of universal gravitation,
which states that two objects of masses m
1
and m
2
, separated by a distance R, experience an
attractive (gravitational) force given by
F = G
m
1
m
2
R
2
(13.1)
where G is a universal constant equal to 6.6710
11
m
3
kg
1
s
2
. (The unit of G could also
be Nm
2
/kg
2
, if you nd that simpler to understand.)
Now, return to our water parcels at A and B on Figure 13.4. Lets say the parcels have
equal mass, m
/
say. Then, the left-ward force at A on Figure 13.4 is
G
m
/
M
r
2
G
m
/
m
(D+r)
2
(13.2)
where the rst term comes from the earths gravity, and the second from the moons gravity.
Note that we are considering left-ward forces on the diagram, i.e. forces that make for a
tidal bulge at A. Similarly, the right-ward (high tide) gravity force at B is
G
m
/
M
r
2
+G
m
/
m
(Dr)
2
(13.3)
5
You might nd it helpful to think of the centrifugal force holding water at the bottom of a bucket being spun
on a rope.
13.2. TIDAL FORCING 169
Notice that equations 13.2 and 13.3 share a common term. This is inward force of earth
gravity, and since it is identical at A and B, it is not a tide-generating force. That leaves us
with one term in each equation, and those terms are identical except for the denominators. A
little manipulation lets us write these forces as
G
m
/
m
D
2
1
1+x
2
(13.4)
and
G
m
/
m
D
2
1
1x
2
(13.5)
respectively, where x = r/D. Now, the earth-moon distance D = 410
8
m is much smaller
than the earth radius r = 6 10
6
m, so x is a small number. Therefore, we can use the
approximate rule
6
that 1/(1x
2
) 1+2x and 1/(x+x
2
) 12x. With this approximation,
we have
G
m
/
m
D
2
(12
r
D
) (13.6)
as the outward force at A and
G
m
/
m
D
2
(1+2
r
D
) (13.7)
as the outward force at B. Expanding, we get
G
m
/
m
D
2
+2
r
D
G
m
/
m
D
2
(13.8)
as the outward force at A and
G
m
/
m
D
2
+2
r
D
G
m
/
m
D
2
(13.9)
as the outward force at B. The rst term in these last two equations is directed toward the
moon at each point, and is independent of the earth radius. These things suggest that this is
not a tide-generating force. (Indeed, is related to the rotation of the earth-moon pair.) As for
the second terms in these equations, these are tide-generating forces.
Note that the forces are outward at both A and B, verifying the physical argument of two
tides per day. But the mathematics has given us something new: the fact that the tidal forces
at the inside and outside points are equal.
The Mathematical treatment also lets us compare the tide force to the simple gravity
force. To do that, we examine the ratio of the tide-producing force (say, the second term in
equation 13.9) to the gravity force. This is
2
r
D
G
m
/
m
D
2
_
G
m
/
M
r
2
(13.10)
which simplies to
2
m
M
_
r
D
_
3
(13.11)
Using values m = 610
24
kg (earth mass), M = 710
22
kg (moon mass), r = 410
5
m
(earth radius) and D = 4 10
8
m (earth-moon distance), we see that the ratio of the tide-
generating force to the gravity force is 2 10
7
. (In words, the gravity force is twenty
million times stronger than the tide force.) Thats very small, indeed . . . but gravity cancels
out in this problem, so its intense effect is, well, ineffective!
Thats it for our sketch of the mathematics of tidal forces. For more, see various textbooks,
and especially original sources such as Doodson [1921]. Doodson, especially, is a good
source for insights on the remarkable suite of tidal forcing frequencies, a few of which are
given in Table 13.1.
6
If you remember Taylor expansions, you can see this quite quickly. If not, I suggest you just try this in a
calculator, for x = 0.01 as in the present case, and for a few other small values. The act of keying in the numbers
may help you to remember the rule, even if you cannot derive it from calculus.
170 CHAPTER 13. TIDES
Semidiurnal Diurnal Long
M
2
12.4206 h K
1
23.9344 h M
f
327.85 h
S
2
12.0000 h O
1
25.8194 h M
m
661.31 h
N
2
12.6584 h P
1
24.0659 h S
sa
4383.05 h
K
2
11.9673 h Q
1
26.8684 h
Table 13.1: Principal tidal periods, grouped into three categories: roughly twice per day,
roughly once per day and longer than a day. This list is not exhaustive (Doodson [1921]
lists 399 different constituents), nor is it universal (different sites have different principal
constituents). Note that subscript 2 is used for semidiurnal components, and subscript 1 for
diurnal components.
13.3. Tidal analysis
We model sea level variation = (t) as being of the form
= +
/
+
0s
sin
0
t +
0c
cos
0
t +. . . (13.12)
Here is mean sea level,
/
represents non-tidal variation caused by waves, by wind, etc,
and the sine and cosine terms represent a tide of frequency
0
. (The use of both sin and
cosine terms permits the specication of tidal phase, i.e. the timing of high tide with respect
to some reference time.) To this, we must add terms with other tidal frequencies
1
,
2
, and
so on.
In the above equation, we know the frequencies of the terms, but not the amplitudes
0s
and so on. Expressed in this way, what we have is a regression problem. Given a long time
series of sea level, we must use statistics to nd appropriate values for
0s
,
0c
, etc. This
procedure is called tidal analysis, or tidal decomposition.
One practical issue in tidal analysis has to do with the length of the sea-level records. As
you can imagine, an hour of data will not let you distinguish between an M
2
tide (period
12.42 h) and an S
2
tide (period 12 h). Generations ago, Lord Rayleigh proposed a criterion
for the time interval needed to distinguish between sinusoids of nearby frequencies, and this
criterion is a part of tidal analysis.
Nowadays, tidal analysis is often done with Mike Foremans Fortran package [Foreman,
1977, revised 2004] or with Rich Pawlowiczs t-tide Matlab package [Pawlowicz et al.,
2002], which is a recoding (and slight reinterpretation) written in the Matlab language.
These packages examine the length of the record, and then select a subset of the known tidal
frequencies that can be resolved. They can also, especially in Foremans code, do some
sophisticated estimation of non-resolved nearby constituents, through something called
satellite correction (unrelated to satellites in space).
My own oce package, in the R language
7
, also does tidal prediction, albeit (as of
summer 2008) without satellite/nodal correction. Figure 13.5 shows the results of analyzing
a year-long sea level record in Halifax Harbour.
13.4. Empirical tidal prediction
Once the tidal amplitudes are known for a given location, equation 13.12 can be used
to predict tides for long intervals into the future. By the early years of the 1900s, this
was done routinely for important harbours. Lacking electronic computers, the calculations
were carried out by ingenious mechanical computers consisting of a clockwork-style
arrangement of pulleys of various sizes. Figure 13.7 illustrates.
In recent years, of course, electronic computers have taken over, and hydrographic
agencies use them to produce printed and web tidal predictions of harbours and other sites
of interest, based on sea-level data.
For research work, many people now use the predictions of numerical models that had
been set up to incorporate local data. The idea is to assemble local sea-level data in a region
7
R is an open-source statistical language, available at http://www.r-project.org (August 2009).
13.4. EMPIRICAL TIDAL PREDICTION 171
0.00 0.05 0.10 0.15 0.20 0.25 0.30
0
.
0
0
.
2
0
.
4
0
.
6
0
.
8
1
.
0
1
.
2
1
.
4
Frequency [ cph ]
A
m
p
l
i
t
u
d
e

[

m

]
SA O1
K1
M2
S2
M4
pdf("tide-halifax.pdf",width=7,height=7)
library(oce)
data(sealevelHalifax)
tidalAnalysis <- tidem(sealevelHalifax)
plot(tidalAnalysis)
Figure 13.5: Top: Tides in Halifax Harbour. Note the strong semidiurnal components
(e.g. M
2
and S
2
, but also including nearby satellites), indicating two tides per day. This
is an unusual representation, showing the cumulative amplitude over frequency, instead
of spectral spikes. The advantage of this format is that it shows the relative strength of
constituent groups at a glance, e.g. the semidiurnal tides are predominant in this harbour (as
they are along the Nova Scotia coast), but the diurnal tidal components are also signicant,
which leads to a noticeable spring-neap cycle (see Figure 13.6). Bottom: R code to create
the diagram.
172 CHAPTER 13. TIDES
0 5 10 15

1
.
0

0
.
5
0
.
0
0
.
5
1
.
0
S
u
r
f
a
c
e

e
l
e
v
a
t
i
o
n

[
m
]
0 5 10 15

1
.
0

0
.
5
0
.
0
0
.
5
1
.
0
Time [day]
S
u
r
f
a
c
e

e
l
e
v
a
t
i
o
n

[
m
]
pdf("spring-neap.pdf", width=8, height=8)
t <- seq(0, 15, 1/24)
par(mfrow=c(2,1))
par(mar=c(4,4,1,1))
hour <- t
*
24
M2 <- 1.00
*
sin(2
*
pi
*
hour / 12.4206)
S2 <- 0.25
*
sin(2
*
pi
*
hour / 12)
plot(t, M2, type=l, col="blue",
xlab="", ylab="Surface elevation [m]")
lines(t, S2, col="red")
plot(t, M2 + S2, type=l, col="darkgreen",
xlab="Time [day]", ylab="Surface elevation [m]")
Figure 13.6: Top: Surface elevation timeseries for an M2 tide (blue) plus a weaker S2 tide
(red). The tidal amplitudes roughly correspond to those in Halifax Harbour (Figure 13.5).
Middle: sum of M2 and S2, displaying a noticeable spring-neap cycle. Bottom: R code
Bottom: R code to create the diagram. Exercises: (a) adjust the tidal amplitudes to see
effect on the spring-neap amplitude; (b) adjust the phase of one of the components, to see
the effect on spring-neap timing.
13.5. TIDAL PREDICTION 173
Figure 13.7: Tide-prediction machine once used at the National Observatory, Rio de Janeiro,
Brazil. At the bottom right are a series of pulleys of various sizes, representing various tidal
periods. (Downloaded from http://tidesandcurrents.noaa.gov/predma3.
html on 2008-07)
of interest, and then to set up a tidal model that tries to t these data, based on dynamical
(not statistical) techniques. Various models have been set up in various places. For Canadian
waters, a popular tool is WebTide
8
. This is a java-based program that works on MSwindows
and Linux (and on OS X, in my own experience). It produces graphs of sea level height and
of horizontal currents, based on the predictions of (ofine) numerical models that were set
up to use observations local to the prediction region. Figure 13.8 illustrates the domains,
as of the summer of 2008. The software is very easy to use, and any student doing work
in a region where tides are important really should take a half hour and set it up. A sister
package called WebDrogue
9
makes it easy to do predictions of where tides advect water
parcels. It is also very easy to use, but, as of summer 2008, its domain was more limited
(Figure 13.8).
All of the above has been aimed at producing precise tide predictions at a point, e.g. a
harbour, based on local data and either statistical inference or a regional tidal model. A
broader view (albeit one lacking this precise geographical detail) can be constructed using
satellite measurements of the changing sea-surface shape. Figure 13.9 illustrates.
13.5. Tidal prediction
It is important to note that the tidal forcing does not provide a simple means to predict
tidal response. For example, the theory would predict roughly equal tides at Halifax and
the Bay of Fundy, because the two locations are very close together, compared with the
radius of the earth, and so they experience similar gravitational forces from the moon and
sun. However, Halifax has ho-hum (1 m) tides, whereas the Bay of Fundy has exciting
(10 m) tides [Garrett, 1972]. If the forcing is that the water has dynamical features that
give it a life of its own. You should think of the tides as gravity waves. The wavelengths
are typically 100 km (in weird areas like the Bay of Fundy) to 1000 km or larger. The
water depth is only 0.2 km on shelves and 40 times that in the ocean generally, and so you
should see immediately (section 10.4.2) that tides are shallow-water waves. (Its intriguing
8
http://www.mar.dfo-mpo.gc.ca/science/ocean/coastal_hydrodynamics/
WebTide/webtide.html
9
http://www.mar.dfo-mpo.gc.ca/science/ocean/coastal_hydrodynamics/
WebDrogue/webdrogue.html
174 CHAPTER 13. TIDES
Figure 13.8: Geographical domains of the WebTide tidal elevation/ow model (top) and
WebDrogue tidal-displacement model (bottom) as of summer 2008.
Figure 13.9: M
2
tide inferred from Topex/Posieden satellite. (Image courtesy of Richard Ray,
Space Geodesy branch, NASA/GSFC http://svs.gsfc.nasa.gov/stories/
topex/images/TidalPatterns_hires.tif.) Compare with Figure 13.1.
13.6. TIDAL MIXING 175
that the global-scale tide waves act just like tiny waves on beaches.) You will recall that
shallow-water waves naturally travel at a speed controlled by the local water depth h. Now
it gets tricky, for the tidal wave should travel around the earth once per day, but it would
like to move at the speed (gh)
1/2
, which is much smaller
10
. Worse yet, h varies all over
the planet. And of course we have these funny beach-like things called continents that get
in the way. Taking all of this together, and throwing in friction just for fun, and you get a
somewhat messy picture.
By messy I partly mean computationally demanding, since it is necessary for the
domain to span global scales but also resolve the details of shelf and even harbour scales.
Furthermore, there is a need to consider baroclinic responses, and that means that we have
to also predict the ocean stratication, which is in itself a very great challenge. Indeed,
buried in the problem of predicting tides are so many other problems that the whole thing is
a nightmare . . . or a beautiful puzzle, depending on your view!
13.6. Tidal mixing
Tides create turbulence that contributes to vertical mixing of stratied water. Since tides
can vary greatly over small spatial ranges, this turbulence can vary also, and so it is possible
to for tides to create distinct variations in the vertical stratication.
A rough measure of the competition between the decrease of stratication by tides and
the increase of stratication by surface buoyancy ux (e.g. surface heating, or river outow)
is the Simpson-Hunter parameter
QH/U
3
(13.13)
in which Q is a surface buoyancy ux (in units m
2
/s
3
), U is a tidal speed, and H is water
depth [Simpson et al., 1981] (see Garrett et al. [1978] for an application to the Bay of
Fundy and Gulf of Maine). This is a measure of the work required to mix the water column,
divided by the source of energy that the tide provides for turbulent mixing. Contours of this
parameter (or variants developed subsequently) provide a hint as to the possibility of mixing
the water column. Indeed, there is good evidence that fronts between mixed and stratied
waters occur for certain values of this parameter. This is of interest because these fronts tend
to be very active biologically, and so we can predict these areas from physical principals
alone.
Tidal mixing is important in many estuaries, and since the mixing and the estuarine
overturning are linked (see chapter 9), we see that tides play a role in long-period ow, as
well as ow on time scales of hours.
13.7. Residual tidal effects
A tidal ow turning a corner can cause centrifugal upwelling, and this is possibly
signicant near Yarmouth. In addition, various nonlinear effects can rectify tidal currents, as
will be illustrated in class for jets near headlands, ow around Georges Bank, etc. Examples
will be discussed in class.
13.8. Tidal power
13.8.1. General considerations
The ideas for tidal energy are similar to those of wave energy (section 10.4.4). There are
many details relating to the generation system barrage or turbine, single-ow turbine or
double-ow, etc but some general calculations can be made in any case.
Consider an ocean region of plan (surface) area A and of some very small height dz. The
mass of that region is Adz, and so its weight is gAdz. Imagine that this region is raised
by some height z. The energy needed to do the lifting, given by the product of the force and
10
Exercise: Do some calculations of these two speeds, in both the deep ocean and on continental shelves.
176 CHAPTER 13. TIDES
Figure 13.10: M
2
tides in the Gulf of Maine and Bay of Fundy. Cotidal lines are in
centimetres. This is based on Figure 2 of Brown [1984], in turn based a diagram (published
in the grey literature) by Moody and colleagues in the early 1980s. For reference, the red
line pointing to Minas basin is 185 km long, at this scale.
the distance, is gAzdz. From this, you can see that step-by-step raising of a whole column
by a height takes an energy input of
_

0
gAzdz =
1
2
gA
2
(13.14)
This provides an upper bound on the energy that could be extracted from this raised water
mass, whether by turbines or some other scheme. (We can never extract more energy than
that was put in by the tidal forces, so this is an upper bound. Please refer back to the
discussion of waves in section 10.4.4, if youre confused as to why we are looking only at
potential energy here.) It is useful to express this as energy per unit area, as
E =
1
2
g
2
(13.15)
in units of J/m
2
. One may divide this by the time it takes to build up this sea level
11
,
yielding
P =
1
2
g
2

(13.16)
as an estimate of time-averaged tidal power per unit area, in units of W/m
2
.
13.8.2. Tidal power in Bay of Fundy
The Bay of Fundy has famously high tides, and their energy has been exploited for
centuries in one form or another. Over the past 50 years or so, it has been suggested that
this energy be extracted to make electricity. There are many aspects to this issue, including
environmental and economic impacts as well as practical issues of siting the generating
stations and getting the power to market. In this section, I provide a rough sketch of some of
the physical issues involved in extracting tidal power from the Minas Basin.
Consider Figure 13.10, which indicates that the entrance point to the Minas Basin of the
Bay of Fundy has a tidal amplitude of 4 m. Given the spatial trend, it may be reasonable to
11
Think of this as the time interval during which barrages are open, or as the time during which turbines work
efciently in high ow . . . it all depends on the scheme used to extract the power.
13.8. TIDAL POWER 177
estimate
12
the amplitude within the basin as 4.50.5 m. Using 13.16 with timescale
= 2 10
4
s (6 hours, say) yields P = 5 1 W/m
2
as the averaged tidal power per plan
area
13
. The area of the Basin is roughly 10 km by 50 km, or A 510
8
m
2
. Thus, a rough
estimate of the tidal power in the basin is AP 4110
9
W, or 41 GW.
Not all of this power is available, however. For one thing, the turbines that are envisioned
to extract energy from the ow will tend to block the ow. Theoretical considerations (which
are beyond the level of this course) suggest that only about 1/5th of the power in a tidal
stream can be extracted by a turbine eld [Garrett and Cummins, 2005, 2007]. Accordingly,
the estimate of available power is 0.80.2 GW. In addition, there are also losses involved
in the mechanics of the turbines, and in power generation, which yield efciency factor
14
of
0.4. This implies that Minas Basin might supply 0.40.1 GW to the power grid.
This power source can be put into context by comparing it with the electrical use in North
American homes, usually taken to be 1.5 kW. In these terms, the anticipated Minas Basin
power could supply a quarter of a million homes. This is a large number in practical terms,
since Nova Scotia has a population of about a million. Assuming 4 persons per residence.
Of course, not all electricity use is residential, and not all residential power is electrical
15
13.8.3. Political considerations
The calculation provided above is crude, requiring no mathematical or physical ideas
invented since the 1700s. The characteristics of the tides have been known for 300 years in
this region (and tidal power was rst extracted there at about the same time). And essentially
modern-day technologies for generating power from tidal ows have existed for a generation.
So everything is in place, today.
Actually, everything was in place in the 1970s, the last time that oil prices spiked and
North Americans became worried about energy. A test project was set up in a sub-region of
the Bay of Fundy near Annapolis Royal. According to Nova Scotia Power
16
, this station
produces 50 GWh annually, which works out to 6 MW in more sensible units
17
Again
assuming an electrical draw of 1.5 kW per house, this station can supply electricity for 4 000
houses. So this is, really, just a pilot project, and not a signicant solution in and of itself.
After a spike in prices in the 1970s, oil became again a cheap commodity, and interest
in so-called alternative energy sources dwindled. Flash forward to the present day and
were back where we were, all over again. There is now serious talk of putting in serious
(non-pilot) power-extraction facilities in the Bay of Fundy, with capacity of 300 MW or
so
18
.
The are several practical issues involved. One is this power is episodic, meaning alterna-
tive sources would be required in slack times, storage would have to be set up, or the power
would have to be dumped into a larger market that could withstand the uctuations. The
favoured scheme is to sell to the larger market, but rst it will be necessary to construct new
high-voltage lines. Of course, it will also be necessary to run power lines along the ocean
bottom. With such strong currents, there may be difculties in doing that wiring, so there
has been some talk of putting the wires in conduits drilled through the rocks that underly the
movable sediments. There is a hint of irony in this idea, since lateral horizontal drilling is a
technology perfected by the oil industry!
12
This value is roughly consistent with other maps of tides in the region, and the uncertainty is consistent with
that inferred by the recent modelling efforts of Dupont et al. [2005].
13
Exercise: Compare this with the solar power, and then use Figure 13.9 to estimate of the signicance of tidal
power globally.
14
This efciency factor is from a report to the Nova Scotia Power Corporations, designated EPRI-TP-003 NS
Rev2, and available at http://oceanenergy.epri.com/attachments/streamenergy/reports/
Tidal_003_NS_Site_Survey_Report_REV_2.pdf as of 2008-07.
15
Exercise: Do some research on these larger carbon footprint issues, for Nova Scotia in particular.
16
See http://www.electricalline.com/images/mag_archive/18.pdf
17
Caution: the cited document states the capacity as 20 MW, but that seems to be the maximal power, not the
power averaged over a tide.
18
See http://www.cbc.ca/news/background/energy/tidal-power.html
178 CHAPTER 13. TIDES
CHAPTER 14
Ocean Acoustics
This chapter is under construction. At the moment it consists mainly of a repeat of
a set of overheads on the topic, and for that reason it is mostly a set of lists. So, please
ignore formatting, but if you nd errors, or if there is a topic you think I should add,
please let me know. Thanks! Dan.
14.1. Introduction
Physics of sound:
Water is [compressible??] on small timescales (we lied to you before).
Water is springy a restoring force competes with compressive forces. The spring
constant sets sound speed C =

(P/)
,S
1500m/s. NB: property of uid, not
source.
Sound penetrates water easily, EM does not.
Applications:
Animals communication, hunting, bottom-type detection, . . .
Ship sound military, timid instruments, . . .
Measure sh, bugs, sediment, . . . by backscatter or beam attenuation
Uses: communications, SONAR, AD(C)P, RAFOS, tomography
Layperson ref: Oceanus, vol 20(2), Spring 1971
14.2. Ocean acoustic spectrum
Unit dB (decibel) is logarithmic, with dB=10log
10
(p
2
/p
2
re f
), with p
re f
=1 dyne/cm
2
in water.
Acoustic doppler prolers use frequencies of order 1 MHz.
14.3. Sound and compressibility
Mass ux has components u, v, and w.
Net mass ux into a volume V (xed in space) changes the mass inside, changing the
density this conservation of mass (continuity) is:

t
+(u)
x
+(v)
y
+(w)
z
= 0 (14.1)
Let =
0
+
/
(x, y, z, t) (NB
/
<<
0
), so that
0 =
0
(u
x
+v
y
+w
z
) large continuity
+
/
(u
x
+v
y
+w
z
) + small since
/

0
+(
/
t
+u
/
x
+v
/
y
+w
/
z
) = 0
Third RHS term second RHS term if O(
/
t
) O(
/
u
x
). Let be timescale of sound,
U be scale velocity and L be scale length. Can ignore third term if 1/ U/L, i.e. if
L/U.
Swimmer: L 1 m, U 1 m/s so can ignore sound if its timescale is 1 s.
Q: What is in your stereo?
179
180 CHAPTER 14. OCEAN ACOUSTICS
Figure 14.1: Source: [Knauss, 1978, Figure 12.2].
14.4. SOFAR (deep sound) channel 1
Sound speed C =C(S, T, p), or, nominally, C =C(S, T, p). The speed in m/s is
C =1449+4.6T 0.055T
2
+0.0003T
3
+(1.390.012T)(S25) +0.017p
where T is in

C, S is in PSU, and p is in db (1db, here 1 decibar, not 1 decibel, corresponds
to 1 m).
In mid-latitudes (why do I say this??) . . .
Near surface, C/T is largest term therefore C decreases with depth as cooler
water is encountered.
Deeper, T becomes more constant and C/z becomes largest term therefore C
increases with depth.
Therefore C reaches a minimum at some intermediate depth (nominally 1km.)
14.5. SOFAR (deep sound) channel 2
For example, for the canonical prole T(p) = 3+20exp(p/300), with p in dbar, one
has a result as shown in Figure 14.2.
14.6. Refraction
May think of propagating sound waves (or EM waves!) using ray-paths or modes.
Sound ray-paths curve in towards region of slow speed, as waves curve in towards the
shore.
See http://oalib/njit.edu/rays.html, which has code to trace paths.
14.6. REFRACTION 181
5 10 15 20

3
0
0
0

2
5
0
0

2
0
0
0

1
5
0
0

1
0
0
0

5
0
0
0
Temperature, C
z
,

m
1480 1490 1500 1510 1520 1530

3
0
0
0

2
5
0
0

2
0
0
0

1
5
0
0

1
0
0
0

5
0
0
0
Sound speed, m/s
z
,

m
pdf("sound_speed.pdf", width=8, height=8)
library(oce)
n <- 100
z <- seq(-3000, 0, length.out=n)
T <- 3 + 20
*
exp(z / 300)
S <- rep(35, n)
par(mfrow=c(1,2))
plot(T, z,
xlab=expression(paste("Temperature, ",degree,"C")),
ylab="z, m",
type=l, col="darkblue", lwd=3)
grid(col="lightgray", lty="solid")
## assume p (dbar) is same as z (m)
soundSpeed <- swSoundSpeed(S, T, -z)
plot(soundSpeed, z,
xlab="Sound speed, m/s",
ylab="z, m",
type="l", col="darkblue", lwd=3)
grid(col="lightgray", lty="solid")
Figure 14.2: Top left: Presumed variation of temperature in an ocean with xed salinity
S = 35 PSU. Top right: Variation of sound speed. Bottom: R code to make diagram.
182 CHAPTER 14. OCEAN ACOUSTICS
Figure 14.3: Sound paths in sofar channel. Left: vertical prole of mode-3 wave. Right: ray
paths resulting from a source at 1000 m depth at a range of 0 km.
Figure 14.4: Left: Acoustic-doppler current proler, showing beams (three red disks). Right:
geometry of beams. See e.g. http://www.sontek.com for more on such devices,
including principles of operation.
14.7. AD(C)P acoustic doppler (current) proler
The ADCP (or sometimes ADP) device uses the concept of doppler shift to measure
water velocities in a prole (i.e. as a function of depth).
Idea of doppler shift:
Imagine sound of train whistle as it passes by.
Doppler-shifted frequency F
doppler
= F
source
(1+V/C) where V is source approach
speed and C is sound speed.
The ADP (Figure 14.4)
Combines water speed (not velocity) estimates along several (3 or more) directions,
to infer velocity components.
Data are gated in time (thus in depth).
Typical freq. 1MHz (wavelength 0.001m).
14.8. RAFOS INSTRUMENT 183
library(oce)
data(adp)
jpeg("adcp_data.jpg", width=800, height=400, pointsize=22)
plot(adp, which=c("u1", "u2"), mar=c(1.5,2.5,1,1.5))
Figure 14.5: Plot of the adp dataset from the oce R library, showing one day of variation
in two horizontal components of velocity in a tidal estuary. The white trace near the top of
each panel indicates the surface, where data quality is low.
Workhorse in PO very useful, but costs 100K$.
Issues: ship navigation; bottom tracking; tide removal; biology . . .
14.8. RAFOS Instrument
SOFAR oats the instrument has a sound source; location is inferred from travel
times to receivers at xed locations.
RAFOS is SOFAR spelled backwards the oat is the receiver.
See http://www.oc.nps.navy.mil/

garfield and http://www.taygeta.


com/rafos.html.
14.9. Acoustic Tomography
Large-scale (up to basin-scale) technique averages over turbulence and internal
waves.
Data: travel time of sound pulses between source/receiver pairs.
Coded (30s) signals measured to 10
3
s. Inaudible over most of path.
184 CHAPTER 14. OCEAN ACOUSTICS
Figure 14.6: ATOC experiment in Greenland Sea
Infer average temperature along a sound ray-path (i.e. characteristic path) from
sound-speed along path.
Since sound speed C =C(S, T, p), must know (S, T, p) along path requires inversion
technique.
Proposed to measure whole-ocean temperature.
Technique very costly.
Political problems since worries about damage to marine mammals
Greenland Sea experiment see http://www.oal.whoi.edu/tomo2.html.
14.10. ATOC experiment
Climate change expected to heat ocean. ATOC (Acoustic Thermometry of Ocean
Climate) experiment is planned to measure T = T(t) over basin scales.
Sound speed C =C(S, T, p), so travel times along raypaths yields average temperature
along raypaths, if T = T(S, p) known. Measure spatial variation tomographically,
with crossed raypaths.
ATOC is 4-year, $40 M feasibility study, including funds to study damage to marine
mammals from 200 dB sound-sources.
See Scripps (UCSD) site http://atocdb.ucsd.edu
14.10. ATOC EXPERIMENT 185
Figure 14.7: ATOC experiment. Left: the idea of source-sink pairs. Right: photo of a
bottom-mounted source.
186 CHAPTER 14. OCEAN ACOUSTICS
Part V. Extras
187
CHAPTER 15
Laboratory Experiments
L
ABORATORY EXPERIMENTS PLAY A KEY ROLE in uid mechanics. In this chapter
you will nd descriptions of some simple experiments that you can do in your
kitchen, if youre so inclined.
15.1. Salt-ngers
15.1.1. Introduction
A laboratory model of salt-nger may be set up with any two substances that diffuse at
different rates. In the ocean, the substances are heat and salt. Unfortunately, heat is difcult
to control in the laboratory, and spurious convective motion tends to occur, confusing the
uid ows under study
1
. For this reason, it is easiest, for classroom demonstration purposes,
to use substances other than heat and salt.
The sections below discuss two methods. The rst is probably easiest.
15.1.2. Method 1: heat/dye ngers
The following instructions are courtesy of Barry Ruddick, who credits the advice of
George Veronis.
Float a layer of hot water over a deeper layer of cold water. The hot layer shoulld be
about 5 cm thick. The hot water should have a FEW grains of sodium ourescein dye
dissolved into it. This has a trivial effect on the density, so its easy to oat the layer on top
of the cold by using, say a scrap of cardboard.
Then you just leave it sit. The surface heat loss decreases the temperature difference,
and when the density ratio R becomes less than 100, ngers can form. This works so well
because the rst ngers to form at large density ratio are also very large. Also, the long
time (10-50 minutes) for ngers to form allows time for viscous dissipation of all the initial
disturbances that induced by lling, and this yields ngers that are quite regular.
You start the class off by lling the tank, then go on with the lecture. Five minutes before
the end o the class, turn off the lights, and use a ashlight to show the ngers.
This works best in a tank with square side-walls. To construct it, use 1/4-inch plate glass
in the correct size pieces, with the edges smoothed with emery cloth to prevent cuts. Then
clean all edges with alcohol, and then glue them together with clear silicone seal. Duct tape
provides a convenient clamp while this glue sets.
15.1.3. Method 2: sugar/salt ngers
Every kitchen has the materials needed to set up a salt-nger experiment. Sugar and
salt diffuse at different rates. Sugar diffuses more slowly than salt diffuses, so sugar in our
experiment plays the role of salt in the ocean. Just to confuse things, salt in the experiment
takes the role of heat in the ocean, since salt diffuses faster than sugar.
1
For example, experiments set up in a beaker suffer from the fact that the beaker itself transmits heat, leading
to so-called side-wall ows. Imagine a setup with a layer of cold water sitting below a layer of hot water ...
heat transmitted along the walls of the beaker will heat up the cold water near the edges of the container, causing
[convection??] as uid near the walls warms and grows less dense. This effect can be reduced, in relative terms,
by using a very thin-walled beaker or beaker that is so large that the side-wall ows are a minor part of the overall
ow. But were still left with the problem of heat-ow through the beaker, e.g. from a warm uid to a cool room.
This too will set up convective patterns inside the beaker, and for this reason it is common to insulate beakers with
thick styrofoam. Uh oh, now we cant see the ow! Fine, you say, well have to cut a hole in the styrofoam for a
viewing port. But wont that let heat escape or enter the domain? Sheesh!! You get the picture.
189
190 CHAPTER 15. LABORATORY EXPERIMENTS
Youll need a beaker with transparent sides, two mixing containers, some sugar and salt.
Youll also need some food coloring, dye, or ink ... something dark enough to see when
diluted but that can be cleaned up and. If youre doing this in a kitchen, dont use a poison
dye or your roommates and family will sue me. Youll also need a piece of paper to which
youve glued a string that is twice as long as your beaker is high.
In one container, mix the top uid, with 1 part salt and 2 parts sugar for 48 parts fresh
water. Add a few drops of food coloring, or ink
In the second container, mix the bottom uid, with 2.5 parts salt and 1 part sugar for
48 parts water.
Half-ll the clear-sided beaker with the bottom uid. Then cover the surface with the
paper, and dangle the string over the top of the beaker. You might want to tape the string on
the outside of the beaker, or at least keep a nger on it. Then, carefully pour the top uid
on top of the paper. The paper acts as a momentum buffer, to prevent the top uid from
mixing into the bottom uid. Youll be able to pour faster as time goes along.
Then, carefully pull out the paper with the string. This is a step that is easy to mess up,
but if you read these directions fully before starting, I bet youll have cut the paper to a
shape, and attached the string in such a way, that you can pull the paper out without mixing
the whole beaker up!
If youve succeeded in leaving a sharp interface between the two layers, sit back and wait.
After about a half hour, you should notice salt-ngers forming at the interface, forming a
very ne comb feature.
For more on this sort of lay-persons experiment in salt-ngers, see Strong [1971].
CHAPTER 16
Dealing with Data
T
EXTBOOKS TALK ABOUT THE THEORY of Physical Oceanography, but they seldom
tell you how to actually deal with physical-oceanographic data. Ill try to ll in this
gap here.
16.1. Overview
There are no how-to guides telling how to do even the most basic things with oceano-
graphic data, such as calculating the buoyancy frequency, N. (If you dont remember what
N is and why its important, have a look back at section 4.9 on page 61.)
The lack of guidelines on doing day-to-day physical oceanography is an anomaly. For
example, in Biological or Chemical Oceanography, such guides exist and are often referred
to daily
1
. In addition to how-to guides, research papers in other elds often have detailed
instructions on the procedure followed in the laboratory. You could ip the pages in a whole
issue of the Journal of Physical Oceanography and nd not one section named procedure.
So why are the elds different? Im not sure. Maybe its because research in Biology and
Chemistry often involves technical work done by support staff who lack a deep understanding
of the science behind the work. If you walk into a typical Biological Oceanography
laboratory, you may see a half-dozen undergraduates peering through microscopes, counting
things and measuring things. These folks need to be told how to recognize organisms, and
how to prepare samples, etc. It makes sense to pass the instructions on the work along
in printed form, saving time spent by high-level staff and ensuring uniformity between
laboratories. The key is that a lot of the work is done by low-level staff, who wont be the
folks writing the papers. In Physical Oceaongraphy, the day-to-day work is more often
done by the folks who wrote the proposals and will write the papers. These folks are highly
trained and are in it for the long term; they are already spun up on techniques. They dont
need guidebooks as much as the undergraduate who, a week before, had never seen the
organism she looks forward to counting through the months of her summer.
But I think we are entering a new era, one in which at least some processing of Physical-
Oceanographic data will be done by non-experts. This is largely because of new software
tools. Simply stated, it is easier to work in matlab than in fortran
2
. Some of you
reading this will want to work with Physical-Oceanographic data in your research, and for
this reason Ive started a little guide on some of the issues. My including this material is
unusual, and Id like your opinion as to whether it is worthwhile. Also, if there is a topic Ive
not covered, please let me know. I started sketching this out for the autumn-2000 course,
and Ill leave the chapter alone, to concentrate on other parts of the notes, unless I hear that
you want more work done along these lines.
16.2. Signicant digits and uncertainties in calculation
Suppose youre pulled over for speeding, and the peace ofcer asks how fast you were
going. One hundred kilometers per hour, you proclaim indignantly. And what of it?
Whats your frickin problem, you putrid nazi pig, you decide to add, before kneeing him in
the groin.
1
Hm, is that a silver hake larva or a silversh that crawled from the corner of the lab onto the microscope slide?
Lets just (1) check in this handy-dandy book of pictures and (2) send an email to somebody complaining that this
building is disgusting.
2
By the way, I highly recommend you learn matlab. Youll have available specialized things such as the
equation of state (which is a formula that runs over about 20 lines), in addition to a whole host of important tools,
such as fourier transforms and wavelet analysis. You just cant do in excel what you can do in matlab, and the
modest learning curve is well worth the effort.
191
192 CHAPTER 16. DEALING WITH DATA
This should be enough to land you in court, where youll have the pleasure of explaining
error bars to the judge
3
.
The question is: what does 100 kph mean? That depends on many signicant digits we
are proposing. It could mean that the speed is accurate to the three digits indicated, i.e.
between 90.5 kph and 100.5 kph. The ofcer would have you dead to rights on this one,
because after all, you were doing 140 kph. But maybe there is only 1 signicant digit; you
mean that it is between 50 kph and 150 kph, i.e. you were rounding to 1 digit.
The problem is that in normal notation, we have no real way to indicate that we trust only
the rst digit (so the speed is 1xx in this case, where x is an unknown digit), or the rst two
digs (10x), etc.
It was for this reason that scientic notation was devised. If we right a speed as 1.00
10
2
kph, then we mean that the speed is > 0.99510
2
kph and < 1.00510
2
kph, since
rounding the last digit in either instance yields 1.0010
2
kph. We say that 1.0010
2
kph
has three signicant digits.
The use of the scientic notation, with the symbol is quite important, since it makes
things very clear. Suppose Im giving you directions to my house. I tell you that its 1 km
from a curve in the road until my house. By that I mean that the distance is between 0.5 km
and 1.5 km. However, you dont like km, so you convert to m. Does it make sense to
write the distance as 1000 m? Well, not if you think it means the real distance is between
999.5 m and 1000.5 m! Writing the distance in scientic notation, I might say 110
0
km
and you might say 110
3
m, and there is clearly no confusing in the whether we agree on
the preciseness.
Now, I come to the most important point. Suppose that youve done a calculation using
some data. You must be sure not to let the number of digits in the nal result of the
calculation become much larger than the initial number of digits. Suppose that we are
interested in the product of two quantities. One is 1.1 and the other is 2.222. The rst is, as
written, known to 1 part in 10 parts, i.e. to 10 percent. The second, however, is evidently
known very precisely. And what of the answer how precise can it be? Your calculator
will indicate the product as 2.4442, a value with 5 digits. Surely this is nonsensical, and
your calculator can indicate why! We infer that the rst number, 1.1, probably lies between
1.05 and 1.15. Typing these values into the calculator yields the limiting results 2.3331 and
2.5551; these differ in the second digit, so wed be best off writing 2.4 for the result, since
that means the value is between 2.35 and 2.45.
The pattern is clearly illustrated by the example just given: the least-accurate value sets a
limit on the result. The rule is to retain only as many digits as you know to be signicant.
By now youve realized that the signicant-digit notation is a sort of poor cousin to error
bars. It isnt as powerful a notation as error bars are, since the implied uncertainty may only
be to a single digit. With error bars, we might say our speed was 10028 kph say. There
is no way to express this uncertainty with signicant digits alone. However, for quick and
dirty applications, signicant digits are often enough.
16.3. Processing density proles
This section will, for concreteness, focus on density proles. Nearly everything that is
said could also be said about any other hydrographic quantity, however, so when you read
density you might think temperature, or optical density, etc. But Ill focus on density,
, since it enters into most dynamical aspects of the ocean.
Density is, of course, is a function of time and of space. Mathematically, we would write
= (x, y, z, t). But of course we cant measure such a eld in practice; we have know way
of knowing the value through all space and time, or even through all the space of some
limited domain of interest. Instead, the vast majority of density observations are made with
a single type of instrument: the CTD proler (see section 4.6.1 on page 51).
3
No, Im not talking from experience.
16.3. PROCESSING DENSITY PROFILES 193
Lets suppose for the time being that CTD-style measurements are appropriate for the
task at hand. That is, we are interested in only a particular geographical location and in a
particular time, i.e. the time at which the CTD instrument was in use
4
.
Mathematically, we might write = (z) for the dependence of density on depth z. But
the data measured by a CTD are not of this functional form!
To begin with, the CTD does not measure either density or depth directly. Density must
be inferred from temperature and salinity (normally, through the UNESCO equation of state
(see section 4.7 on page 57). Indeed, the salinity is not measured either, and it must be
inferred from the measured conductivity by using another UNESCO-supplied empirical
equation (see Figure 4.14). Depth must be inferred from the observed pressure, using the
hydrostatic relationship (see section 4.5 on page 50).
But lets suppose that this initial processing has been done, as it often is by the low-level
CTD software, and that we have been presented density measurements. We may still not be
ready to calculate derivatives!
As the boat rocks and heaves in waves, the CTD may go up and down in the water
column, so that the a given depth may be sampled repeatedly, perhaps yielding different
densities each time. This means that, mathematically, (z) is not a single-valued function,
and that it is not possible to even dene a derivative! For this reason, a common rst step in
processing CTD data is to make z monotonic by some means. CTD instruments are normally
meant to be used only while they are sinking, since the sensors are at the bottom of the
package and it is essential that they run into water that has not been mixed up by the CTD
package itself. Therfore, it is common to discard any data collected while the instrument is
rising. (This is easy to do, since the instrument records pressure and the pressure drops as
the instrument rises.)
At this stage we will have available a series in which each sample represents water that
is deeper than the previous sample. (Since we measure z positive upwards, normally with
z = 0 at the ocean surface, all the z numbers will be negative, and thus each sample will
have a smaller z value than the preceding one.)
Let i be an index to number the data. That is, our data set consists of two sets of numbers,
z
i
and
i
, where i ranges over the integers from 1 to n. (Here n is the length of the data
series.)
We know that z
i
is monotonic, but that doesnt mean that the samples are equispaced in
depth. However, for many purposes not just for computing derivatives it is important that
data be equispaced. Therefore, the next step is often to cast the data onto a uniform depth
grid. This casting means to interpolate the observations onto a series of standard depths.
For example, we might have observations that are spaced at something like 5 cm to 10 cm
and wish to cast these onto a 1-meter grid. How should this be done? Lets suppose that
we have data at depths of 9.6 m, 9.9 m, 10.3 m, and 10.5 m. Should we just pick the density
at the closest depth (i.e. at 9.9 m)? That scheme is problematic since density is probably
varying with depth and sometimes well be close to the target depth and other times we
wont, so there is this introduces uncertainty even if the data are without error. Maybe we
linearly interpolate between the points just above and below the target depth. That makes
a bit more sense, but it is pretty sensitive to errors in the data
5
since it is using just two
datapoints and discarding the information carried by nearby points. A third approach might
be to average all the observations that fall between 9.5 m and 10.5 m. In the language of
time-series analysis
6
this amounts to a very crude lter, a so-called boxcar lter.
A difculty arises with the boxcar method if the data are distributed oddly in depth. For
example, there may be many more points between 9.5 m and 10 m than there are between
10 m and 10.5 m, and if density is varying systematically, then this will introduce an error
even if the data themselves are error-free.
4
Note that a CTD is normally dropped and then raised again at about 1 m/s, so it will take an hour or two to
make a single observation in the deep ocean. (Exercise: verify this statement, noting that the mean ocean depth is
roughly 4km.) As an exercise, you might want to think of ways in which the ocean state might change in such a
timescale.
5
And there are errors in data!
6
And time-series analysis is what were talking about, with depth taking the place of time.
194 CHAPTER 16. DEALING WITH DATA
This suggests that a better approach might be to use regression regression to t a line
through the data that are between 9.5 m and 10.5 m, and to extract the value of this line
at z=10 m. This seems to make sense, but it also presents new choices well have to
make. Should the line be tted by least-squares or by another means? (For example, some
researchers in Statistics argue that we should minimize not the sum of squares of model-
to-data mist, but rather the sum of absolute-values.) And should the data near 10 m be
weighted higher in the linear t than data far from this depth? And, furthermore, whats
special about a 1-meter range over which to t the line? It might make sense to use data
over a wider range (or a narrower range).
Implicit in what Ive written above is that there is no theoretical or rst-principles reason
to do one thing instead of another. Often an investigator will try a range of techniques that
seem reasonable (such as those listed above, and others you might imagine), checking to see
whether the results of the nal analysis depend greatly on the technique. This is an example
of a so-called sensitivity analysis. The idea is that the arbitrary choice of one particular
technique is made more palatable if it turns out that many similar techniques yield the same
results.
16.4. Calculating the Buoyancy Frequency
The buoyancy frequency (also known as Brunt-Vaisala frequency; see section 4.9 on
page 61), N, is dened by the equation
N
2
=
g

z
(16.1)
where g = 9.8 m/s
2
is the acceleration due to gravity,
0
is a constant reference density,
which can be approximated as the depth-averaged density
7
and the derivative is the thing
that is problematic, so Ill discuss this in some detail.
Lets suppose to begin with that the density data have been processed as in the previous
section, and that they are cast on a uniform grid.
Your rst thought might be estimate the derivative at a given depth by computing a
so-called rst difference of the data. One way to do that would be to take (
i

i1
)/(z
i

z
i1
) as an estimate of /z at depth (z
i
+z
i1
)/2. (Here i may take on the values 2, 3, ...
up to n.)
However, a two-point derivative is very sensitive to noise. In practice, with density
proles containing small-scale wiggles, it is prudent to smooth the data before calculating
derivatives. There are many ways in which this may be done. One method is illustrated in
Figure 16.1, which shows the computation of N
2
using the oce package of the R language
8
,
based on an articial dataset in which noise is present. This package computes N
2
after rst
approximating (z) with a smoothing spline.
Smoothing splines are not the only method used, however. Others include the following.
Smooth (z) by applying a so-called boxcar lter. For example, with data spaced
at increments of 1m, one might average the values 1m above and below. This is a
boxcar lter of width 2m. If such a lter failed to yield a smooth (z) curve and
mainly positive values of N
2
, a larger lter would be used. Normally, one guesses on
a reasonable lter by looking at a graph, and then tries a few lters with widths that
are in the same ballpark, to see if N
2
is remains unchanged by a range of seemingly
reasonable choices. Exercise: what do you think would happen in this case, if the
lter width were 50m?
Apply a more sophisticated lter (e.g. the so-called butterworth lters are popular)
and then calculate the derivative by rst difference. The advantage of this is that the
boxcar lter is suffers from so-called spectral leakage, which is beyond the present
discussion.
7
As an example, try some calculations with your own data, in which you use either this density or the bottom
density.
8
R is an open-source statistical language, available at http://www.r-project.org (August 2009).
16.4. CALCULATING THE BUOYANCY FREQUENCY 195
P
r
e
s
s
u
r
e

[
d
b
a
r
]
26.0 26.5 27.0 27.5

[ kg m
3
]
1
0
0
8
0
6
0
4
0
2
0
0
0e+00 1e04 2e04 3e04 4e04
N
2
[ s
2
]
P
r
e
s
s
u
r
e

[
d
b
a
r
]
26.0 26.5 27.0 27.5

[ kg m
3
]
1
0
0
8
0
6
0
4
0
2
0
0
0e+00 1e04 2e04 3e04 4e04
N
2
[ s
2
]
pdf("N2-example.pdf", width=8, height=8)
library(oce)
par(mfrow=c(1,2))
n <- 100 # number of fake data points
p <- seq(0, 100, length.out=n)
S <- rep(35, length.out=n)
t.error <- 0.050
t <- 10 - 5
*
tanh((p-50)/20) + rnorm(n, 0, t.error)
ctd <- as.ctd(S, t, p)
plotProfile(ctd, "density+N2")
t <- 10 - 5
*
tanh((p-50)/20) + rnorm(n, 0, 5
*
t.error)
ctd <- as.ctd(S, t, p)
plotProfile(ctd, "density+N2")
Figure 16.1: N
2
calculation in R. Top left: density and N
2
prole with fake data formed by a
smooth curve with added noise. Top right: the same, but with more noise. Bottom: R code
to make the diagram. Exercises: (a) examine the documentation for the N
2
calculation in
this package, and try alternative methods and parameters; (b) examine real proles to get a
better idea of noise, and do more experiments along these lines; (c) continue, with real data.
196 CHAPTER 16. DEALING WITH DATA
Find /z not by a [rst-order-difference??] method, but rather by determining
the slope of a local linear approximation, using least-squares or other techniques to t
the line. Of course, as in the previous section, this raises the question of how local is
appropriate. (For example, do you think it would be useful to use the whole dataset?)
It also raises the question of whether to weigh nearby data more heavily than data
further away.
In the context of this course, it is not reasonable to make a rm recommendation on
the best technique. There is simply no theory on the best method. Statisticians arent of
too much help, because the rst question they will ask is what is the error model and
there is no general answer to that; sometimes, the error is caused by noise in an instrument,
other times by heaving of the ship, other times by internal wave straining, other times by
interleaving, etc. Furthermore, it is difcult to x on a method for determining N
2
without
knowing why it is being computed, e.g. one persons signal is another persons noise.
16.5. Coordinate systems
16.5.1. Orthogonal coordinate systems
In work at the level of this course, coordinate systems will always be orthogonal or
cartesian
9
. Also, the x and y coordinates will always measure distances along a horizontal
plane. The orientation of x and y may vary, however. Usually x will increase towards the
east, while y increases to the north. However, for local coastal work, its also common to
align either the x axis or the y axis in an along-shore direction. Unfortunately, there is very
little agreement on whether the x or the y axis should run parallel to the coast
10
.
When it comes to the z axis, things get even more confusing. Everybody agrees that
the z axis should be aligned vertically, but some measure z to increase upwards and others
measure it to increase downwards. Worse yet, some put z = 0 at the ocean surface, while
others put it at the ocean bottom.
16.5.2. Right-handed coordinate systems
At least there is one saving grace. All but the idiotic
11
use a so-called right-handed
coordinate, i.e. one obeying the [right-hand-rule??]. To understand this, hold out your
right hand and consider your thumb, your index nger, and your second nger. Point these
such that they are at right angles
12
. Then if your index nger is aligned to point towards
increasing x values and your second nger similarly along the y axis, your thumb will point
towards positive z values.
Use of the right-hand coordinate system ensures that the equations of motion given in
these notes will be true. So, not using a right-hand coordinate system is about as bright as
deciding that youd prefer, for reasons of personal choice, to drive on the wrong side of the
road.
16.5.3. Angles in Geography, Meteorology, and Oceanography
Geographers and hikers mark an angle of 0 degrees as pointing north, with angles
increasing clockwise. However, Mathematicians, Physicists, Engineers, and their friends say
that 0

lies along east, with angles increasing counter-clockwise. Two zeros, two orientations.
But wait, theres more. Some people measure angles in degrees, others in radians, and others
in grads. Its enough to make your head spin
13
.
9
That is, the x, y and z axes will be at right angles to one another.
10
This is, I admit, stupid. But its what youll have to deal with in the literature.
11
And there are some idiots in physical oceanography, Ill concede. Lets just not name names until the course is
over in case its moi.
12
Worth practicing, in case the cops get you to touch your nose this way.
13
By the way, did the head of the girl in The Exorcist spin 180 degrees or /2 radians? Or was it some other
angle? I was too freaked-out to take notes.
16.5. COORDINATE SYSTEMS 197
Angle conversions
Running around a circle, youll cover 360 degrees or 2 radians. Therefore, to convert
from radians to degrees, multiply by 57.296.
Mathematics is always written in radians, not degrees. Computer programs nearly
always work in radians, not degrees. Get used to this or youll get into trouble
14
.
I wont dene a grad since I think youll never run across them in your work.
Given a geographical-type angle
g
(where the subscript g stands for geographical)
the math-type (subscript m) is given by

m
= (90
g
) (16.2)
Naturally, this works only if the angles are in degrees; if they are in radians, substitute
/2 for 90 in the above.
Wind from, currents towards
For reasons not worth knowing, meteorologists often
15
follow the weather-based conven-
tion that a North Wind blows from the north, not towards the north. Thus a ow towards
negative x may be written as a positive number in a datale obtained from a meteorologist.
Or not. Go gure.
This is a problem because pretty much everybody working in uid mechanics would take
a positive x velocity component as indicating ow towards increasing x coordinates.
But it can be a lot worse that this! Some meteorological data are given not in xy-
coordinate-component form, but in strength-and-angle form. But some folks say an angle
of 0 means ow from the north (or to the north!), and others take an angle of 0 to be to
(or from) the east. But wait, theres more! Some folks measure angles about this zero in a
clockwise way, others in a counterclockwise way.
The moral of this sad story? Be careful when working with meteorological data. If
data les have headers, read them. Read all the documents. Ask questions. Be worried.
Check with colleagues. Plot graphs (especially so-called rose diagrams, which are like
histograms in a polar coordinates system) and see if they make sense, e.g. are there quadrants
from which the wind never comes (a bad sign since thats unlikely) or does the wind come
from the opposite direction to the one you know, from other data?
16.5.4. How to rotate cartesian coordinates
Lots of times youll want to rotate coordinate systems. You may have model results that
give currents in along-shore versus cross-shore directions, but want an east versus north
view.
This is easy. Let be the angle, measured clockwise, from the positive x-axis in the rst
coordinate system to that in the new system. If the original x and y velocity components are
denoted u andv, then the new coordinates (marked with a prime, say) will be
u
/
= ucos vsin (16.3)
v
/
= usin +vcos (16.4)
(Reminder: virtually all computer languages do sine and cosine in terms of radians, not
degrees.)
14
You may be wondering why the layperson uses degrees while scientists use radians. The answer is that radians
are more natural. Consider the Taylor series approximation to sin(x), which is x +. . . where . . . represents
terms proportional to x raised to a power higher than unity, and which therefore become negligible compared
to the x term as x 0. Thus, as x 0, sin(x) x, but only if x is measured in radians. If its in degrees,
sin(x) x

(/180), which you will have to admit is a lot messier. Wanting not to have silly terms like /180 in
formulis why mathematicians prefer radians. Mathematicians are a reasonable lot and we should try to make
them happy.
15
But not always . . . gotta keep outsiders confused!
198 CHAPTER 16. DEALING WITH DATA
16.5.5. Latitude-longitude versus x y coordinates
The earth has a radius of r = 6371 km, so the circumference is 2r or 40030 km (this is
by design, the pole-to-equator distance was to be 10, 000 km by denition). Now consider
latitude on the earth. It measures the angle to the north of the equator, peaking at 90

at the
north pole and troughing at 90

at the south pole. The distance along the surface of the


earth from equator to pole is 10000 km, which is 90 degrees of latitude
16
In other words, 1
degree of latitude is 111.2 km.
Not to leave the English system out of the discussion, note that 1 degree of latitude (but
not longitude) is 60 nautical miles; since charts and maps show minutes as well as degrees.
There are 60 minutes in a degree, so a minute of latitude is a nautical mile of distance. This
is the rst thing you should look at when you see a chart; only wimps use distance bars.
While were at it, note that a knot (used to measure the speed of ships and airplanes and
wind and whatnot) is to 1 nautical mile per hour.
Now we come to the fun part. This correspondence of 1 degree of latitude and 111 km
holds everywhere. But what about longitude? It measures distance in the east-west direc-
tion
17
. At the equator, 1 degree of longitude measures a distance that matches a degree of
latitude (roughly 111 km). But longitude lines converge as the poles are approached. Indeed,
if you sat down at the north pole, your bum would span 360 degrees of longitude
18
.
Now, just as its impossible to wrap a globe without crinkling the wrapping paper, so it
is impossible to warp a latitude-longitude coordinate system into a cartesian system. But
you can make a reasonable approximation over limited areas. Lets say we are interested in
the region near a latitude
0
and longitude
0
and y = 0 at
0
, and lets measure x (aligned
positive east) and y (aligned positive north) in km. Then we can convert (latitude, longitude)
pairs, or ( ) pairs, to (x, y) using
x =F (
0
) cos (16.5)
y =F (
0
) (16.6)
where F =111 km/degree. Indeed, if the range of latitudes is small, we could write cos(
0
),
a constant value, instead of cos(), in the x equation.
Conversion from (x, y) to latitude-longitude is done by a simple inversion of the preceding
equations
19
.
16.6. Units: the good, the bad, and the ugly
Units can mess you up worse than cheap liquor or bad shoes. But there is a solution:
convert everything to S.I. units. If you get a datale with speeds in knots, convert to m/s.
Dont use Fahrenheit; use Celsius. Density is in kg/m
3
, not g/cm
3
. Sure, your nger is about
1cm wide, but for goodness sake write that as 0.01m if you need to for a calculation.
That last word of that last paragraph is all-important. You can use any crazy unit for
describing something and it makes sense to say an island is 100 km wide instead of
100000 m wide, to be sure but just be careful to use S.I. (meters, in the island case) if
youre doing a calculation of, say, T/x.
For example, the Reynolds number of an island, UL/, will have a meaningless value if
u is in knots, L in kilometers, and in m
2
/s.
The solution is simple. Just convert to S.I. base units before doing any calculations.
Distances are in meters (m). Times are in seconds (s). Masses are in kilograms (kg).
Densities are in kg/m
3
. Forces are in newtons (1N=1kg m/s
2
). Pressures are in Pascals
16
If youre a history buff, you may know why this distance, 10000 km, is so simple.
17
Wearisomely, there are a lot of conventions here also. Some say longitude is 0 at the so-called Greenwich
meridian, in England, with a positive value a bit to the east and a negative value a bit to the west. This would be
east longitude. But others, notably older folks, would insist that negative values never be used, that positive west
longitude values be used to measure the distance of a moron walking to the west out of Greenwich.
18
A feat not to be sniffed at ;-)
19
Here is an extra note of caution: virtually all computer languages report arc-sine and arc-cosine values in
radians, not degrees.
16.7. PITFALLS IN COMPUTING 199
\section{Overview}
There are no how-to guides telling how to do even the most
basic things with oceanographic data, such as calculating the
buoyancy frequency, $N$. (If you dont remember what $N$ is
and why its important, have a look back at
section\ref{s:stratification} on
page\pageref{s:stratification}.)
Figure 16.2: Snippet of LaT
E
X code, used near the start of this chapter. Note the use of ref
to refer to a section number, and of pageref, to refer to a page.
(1Pa=1N/m
2
). Oceanographers also often measure pressure in so-called decibars; 1 decibar
(or 1 db, to introduce a non-metric abbreviation) is 10
4
Pa and it is conveniently close to
the pressure change over 1 meters worth of water depth. Work is measured in the unit
[Joule??] (1 J=1 Nm). Energy has the same unit. Power, or the time rate of energy change,
is measured in watts (1 W=1 J/s). Temperatures are in Celsius or Kelvin degrees.
16.7. Pitfalls in computing
1. Ive said this above, but Ill say it again. You probably have your calculator set up to
work in degrees, not radians. Please dont be fooled into thinking that your computer
is set up the same way, cuz it aint! Most computer programs measure angles in
radians, not in degrees. See section 16.5.3 on page 197 for the conversion method.
2. The arc functions (arcsine, etc) return radians, not degrees. Also, they arent smart
enough to gure out quadrants for you. In particular, the oft-used atan(x) function
returns an angle between and radians. When working with wind components
and similar data, this is problematic. You need to use atan2(y,x) instead of atan.
Note that atan2(y,x) takes arguments y, x instead of x, y.
(Exercise: try this out, with a real or articial dataset in which wind components
(u, v) come from directions in all the quadrants. First, compute the strength and angle
of the wind, as W =

(u
2
+v
2
) and = atan(v/u), and then convert back using
u
/
=W cos() and v
/
=W sin(). Youll nd that the computed u
/
and v
/
values dont
agree with the input u and v values. Now try it again using = atan2(y, x) and see
if things work out better. Check by plotting or by looking at the numbers by eye;
you cannot check by using the are they equal operator in your computer language
because of roundoff errors.
3. As a general rule, try not to be too fancy with software. For example, if you dont
really know what fourier transforms are, you should do some reading before using
them in a research context. This is because the defaults that software provides are
often not what any reasonable person would use in practice.
16.8. Dealing with mathematically-based text
The standard method of doing this is to use either the T
E
X or the LaT
E
X text markup
language. The keyword here is markup. It is not a clicky-pointy sort of thing. It has
a somewhat steep learning curve, compared to msoft/word, but it rewards the effort of
climbing this curve with a delightful power over the details, and with elegant-looking
mathematics and fonts. Ive seen no program that makes mathematics look anywhere near
as good, and Ive seen no program that lls text so well. These things result, I think, from
a lot of care spent by the author (Donald Knuth). He is a mathematician who spent years
designing this program. I believe he spent months on making the letter s look good. He
scoured mathematics journals and texts, decade by decade, deciding what sort of spacing,
font density, etc, looks good.
To give an idea of what LaT
E
X looks like, Figure 16.2 shows the rst few lines of the le
for this chapter. So, as you can see, this isnt very difcult at all. It looks like normal text,
200 CHAPTER 16. DEALING WITH DATA
Extending this using a similar argument for the
other two directions and equating to the left-hand-side of
(\ref{eq:dyn.dmass.dt}) yields
\begin{equation}
\frac{\partial\rho}{\partial t}
+ \frac{\partial (\rho u)}{\partial x}
+ \frac{\partial (\rho v)}{\partial y}
+ \frac{\partial (\rho w)}{\partial z}
=
0
\end{equation}
This is the equation for the \emph{conservation of mass}.
Figure 16.3: Snippet of LaT
E
X code, to typeset an important equation of these course notes.
with some backslashes and curly braces to designate the meaning of the words (e.g. this
thing is a chapter title; this thing is a section title). The aesthetics of the entire document can
be changed by simply using a different style sheet. Thus, I can print these course notes in
a large font or a small one, with indented section headings or with lined-up ones, on large
pages or on small pages, in a single column or in three columns, all by changing one word
in the preamble.
But it gets better. Turn to page 81, where the important equation for conservation of mass
is discussed. Now examine Figure 16.3 to see how it was typeset. Let me draw your eye
to a few things. First, referring to equations is pretty easy (with the ref command), and
labelling them is easy to (the label command); writing the equation itself is also not very
difcult. You must begin it and end it. If you want a fraction, you use frac. To get the
Greek letter for density, use rho.
If you spend 5 minutes studying the above, and comparing to the material as it appears in
its formatted form, you will already know enough to read LaT
E
X, and youll be well along
the way to writing it.
Why bother? Because the results are beautiful. Because other folks (including virtually
all physical scientists and mathematicians) use this format. Because many journals demand
it and most accept it. Because most publishers of mathematically-complex material accept
it. Because it will put hair on your chest . . . hang on, I may be going a bit overboard here!
Further reading on working with Oceanographic data. Emery and Thomson [1998] is
worth consulting for the details of techniques. However, in Physical Oceanography we lack
the sort of handbook of techniques that is useful in other branches of science, and so many
of the important techniques are scattered throughout the oceanographic literature.
CHAPTER 17
Important Numerical Values
T
HIS CHAPTER LISTS NUMERICAL VALUES FOR SOME IMPORTANT QUANTITIES. If
you really care about a calculation, however, youd be well advised to check out
the references Ive cited, to make sure that the values I give here are correct. Also,
please ignore the uneven writing and formatting here.
17.1. General Properties of Matter
Reference: Gill [1982]
Coefcient in Stefan-Boltzmann constant and black-body radiation law = 5.670
10
8
W/(m
2
K
4
) (see chapter 12).
Coefcient in Weins law = 2.910
3
m

K is a universal constant (see chapter 12).


17.2. Properties of the Earth
Earth radius = 6.410
6
m
Earth mass = 6.010
24
kg
Gravitational acceleration 9.8ms
2
(a slightly-varying function of latitude and height).
17.3. Properties of Seawater
Reference: Pickard [1979]
The mean ocean temperature is 3.5

C.
The mean ocean salinity is 34.7PSU.
75% of the ocean water has temperature between 0 and 6

C and salinity between 34 and


35PSU.
17.4. Salt Water and Sea-ice
Latent Heat of Fusion
Best ref prob Ono [1967]. His table 3, following his equation (16), gives in cgs units;
multiply by 4184 to get to J/kg. Example: for =2, 2 < S < 10, he gives L
f
= (2.5
3.2)10
5
J/kg. For 8 < <2, 2 <S <10, he gives L
f
= (2.53.5)10
5
J/kg. For =
1.8

C (freezing) and S = (7, 12) (perhaps reasonable values), L


f
= (2.66, 2.15)10
5
J/kg.
Note: Moore and Wallace [1988] quote Ono [1967] as giving, for sea ice of S = 7 and at
T =1.8

C (quoting Ono [1967]), L


f
= 2.6710
5
J/kg.
17.4.1. Sea-ice: Thermal Conductivity
Weeks and Ackley [1986] p127 (eqn 1.75) and p128 (g 84) give
k = 9.828exp(0.0057T) (17.1)
where T is in Kelvin and k is the conductivity inW/m/K. EG: at -2C, k =2.10 W/m/K;
at -20C, k = 2.32W/m/K.
Pounder [1965] (p118) gives
k = 2.24(14.810
3
T)W

C
1
m
1
(17.2)
EG: at -2C k = 2.26W/m/K; at -20C k = 2.46.
This is for pure ice. Sea-ice seems to be much the same except for air-fraction (p128).
201
202 CHAPTER 17. IMPORTANT NUMERICAL VALUES
Conversion to diffusivity requires using C
Pi
, which varies a lot.
Conductivity has been inferred by modelling timeseries of temperature within sea-ice,
whilst monitoring air temperature and ice growth rates Lewis [1967]. Lewis states his
accuracy as 0.1 W/degC/m (get this by multiplying his cgs units by 4.184 J/cal and
100 cm/m), and has k ranging from 2.0W/degC/m at the ice bottom to 2.6 W/degC/m
at the top. His table 3 compares his measurements to a theory by Anderson [1960].
Anyway, the agreement is good; at a glance, within < 5%.
Anderson [1960] gives in g7 k as a function of T and S. For T ranging from -2 to
-30

C, k ranges from 0.6 to 2.5 W/m/degC, depending a little on S.


To get from k to , use = k/C
p
/, where of course now we need to know C
p
=
C
p
(S, T); see next section. For an example, using T =1

C and 2 < S < 10, so that


C
p
= (1.14.7) 10
4
J/kg/degC, and = 1000 kg/m
2
, and using a pure-ice density
of 917, the total range on given the Lewis [1967] range of k is 0.52.610
7
m
2
/s.
17.4.2. Sea-ice: Specic Heat
This quantity is Difcult to determine, because its difcult to say just what ice really
is (slush? newly-formed sea ice? multiyear sea ice?) Ono [1967] may be the best reference.
His table 2 gives (cgs units; multiply by 4184 to get J/kg/degC) tabulations of his equation
11. For example, for =2, S =7, C
p
=3.3710
4
J/kg/degC. For =2 and 2 <S <10,
C
p
= (1.14.7) 10
4
J/kg/degC.
Check: Pounder [1965]s table 5 gives, for =2 and S = 8, C
p
= 3.810
4
J/kg/degC,
same as Ono [1967] . . . Pounder doesnt say where he got his gures in the table, so Ono
[1967] is probably the best reference.
17.5. Fresh Water and Pure Ice
17.5.1. Latent Heat of Freezing
For water at 0

C, Iribarne and Godson [1981] Appendix 1 give: L


f
= 0.33410
6
Jkg
1
17.5.2. Latent Heat of Vaporization
For water at 0 Celcius, Iribarne and Godson [1981] Appendix 1 give: L
v
= 2.501
10
6
Jkg
1
As a function of temperature, Gill [1982] gives L
v
= 2.500810
6
2.310
3
t Jkg
1
17.5.3. Latent Heat of Sublimation
For water at 0 Celcius, Iribarne and Godson [1981] Appendix 1 give: L
s
= 2.834
10
6
Jkg
1
As a function of temperature, Gill [1982] gives L
s
= 2.83910
6
3.6(t +35)
2
Jkg
1
17.5.4. Pure Ice: Density
For freshwater ice (I guess), Weast [1985] page F-1 gives

i
= 917kgm
3
(17.3)
while Pease [1987] table 1 gives 950 (with no ref though). Probably most in-depth ref is
Pounder [1965], who gives value of 910 for winter ice at -15C, and a range from 850 for
old surface ice from which considerable brine drainage has occurred to 940 for very cold
winter ice.
17.6. PROPERTIES OF AIR 203
17.5.5. Pure Ice: Specic Heat Capacity
NB: 1 cal=4.184 J
The factor of 2 disagreement below is bothersome! Clearly something is wrong
here, so better look up if I need this.
For freshwater ice (I guess; temperature unknown), Weast [1985] page D172 gives
C
p
= 1004Jkg
1
C
1
(17.4)
Again, for freshwater ice at 0

C, Iribarne and Godson [1981] appendix 1 give


C
p
= 2106Jkg
1
C
1
(17.5)
17.5.6. Pure Ice: Thermal Diffusivity
Smith [1974] Table 4.5-9 p421 gives heat ux F = (0.005, 0.0039, 0.0022) cal cm
2
,
for temperature gradient dT/dx = 0.01

Cm
1
. The numbers are for various investigators
(pretty old stuff, from 1800s). The average of these is F = 0.0160.006 Jm
2
. The heat
equation yields
=
F
C
p
dT/dx
(17.6)
which, using
i
= 917 kg m
3
(see above) and C
p
= 4223Jkg
1
C
1
, one gets
= (4.11.5) 10
7
m
2
s
1
(17.7)
17.6. Properties of Air
17.6.1. Density
Air density is roughly

a
= 1.2kg/m
3
(17.8)
17.6.2. Specic Heat Capacity
Weast [1985] page F11-12 gives
C
p,a
= 1004Jkg
1
C
1
(17.9)
17.6.3. Thermal Diffusivity
17.7. Properties of the Ocean
Typical ocean depth 4000 m.
204 CHAPTER 17. IMPORTANT NUMERICAL VALUES
CHAPTER 18
Techniques
F
EW GENERAL-LEVEL TEXTBOOKS TALK ABOUT THE TECHNIQUES that youll hear
about in seminars, and that are used in todays literature. This chapter tries to ll that
gap by sketching some important techniques, in isolation of the problems to which
they are applied.
18.1. Symbolic Mathematics
Calculations of the potential energy of a uid, as discussed in section 4.11.2, require the
evaluation of integrals. There, it was stated that the PE per unit mass, for a linearly-stratied
uid with density prole (z) =
0
(1N
2
z/g), where N is a constant, is given by this can
was calculated using (4.22), i.e.
PE per unit mass =
1
H
_
H
0
g
_
1
N
2
z
g
_
zdz (18.1)
which yields the value PE =
1
2
gH
_
1
2
3
N
2
H
g
_
. Even though this is an easy integral to do,
it might be a good example for learning how to use symbolic-mathematics software.
Two popular symbolic-mathematics packages are Maple and Mathematica. Up until
2004, I favoured Maple (and its cheap at Dalhousie), so thats what Ive chosen to illustrate
in Figure 18.1. Note that its very easy to do the integral, but that it can be difcult to express
the results into a form that is similar to what you might write with a pencil. But the real
advantage of Maple (and similar software) is that it can handle much more complicated
integrals than this one!
Figure 18.1 also shows the Maple help facility. I got the window on the right by clicking
on the collect word in my worksheet, on the left.
Further reading about Maple. Youll nd the Maple help facility to be very handy,
especially as you grow more familiar with Maple. But I think you would be well-served to
read the deeper treatments that texts provide. On my bookshelf are a couple of books about
Maple; Corless [1995] is the better of these. But, since Maple is in continual development,
you might be better off buying a newer book.
18.2. The Matlab Language
The Matlab language is increasingly becoming the language of choice for data processing
in Oceanography. The mat in the name stands for matrix, and the language started its life
with a focus on matrix mathematics. But it also has reasonably good graphics capabilities,
and it has toolboxes to handle a growing number of specialized tasks, e.g. in spectral
analysis, etc.
One problem with Matlab, though, is its cost: our research group pays thousands of
dollars per year in licensing fees. Therefore, you may want to explore the Octave language,
which is open-source and free of cost (see http://www.octave.org/ for more).
205
206 CHAPTER 18. TECHNIQUES
Figure 18.1: Illustration of the use of the Maple symbolic-mathematics software to perform
a simple integral.
18.3. The R Language
The R language
1
specializes in statistical work. It is open-source and actively developed.
The project website holds good tutorials, and there are several books that provide more. If
youve used SPlus before, youll see that R is nearly identical (except that it has smarter
scoping rules). If not, youll have to be patient for a while, to learn how to work with it. I
have set up an Oceanographic library for R called oce, which you may nd helpful in your
work. (It is part of the standard R package list, and it is installed in the normal way, so I
wont detail it here.) For example, it knows about Seabird CTD format, so you can slurp in
CTD les and plot them in just two commands (Figure 18.2). It also knows all about the
equation of state of seawater (see 4.7). If you are doing complicated statistical work, you
really should spend the time to learn R; it is the language used by cutting-edge statistics
researchers.
Oh, and just so you dont think R is the only solution, you should see Chapter 21 for
examples in Python and Matlab.
18.4. Numerical Models
No oceanographer should be ignorant of the basic ideas of [numerical-modelling??]
techniques. Most Physical Oceanographers end up using numerical models in their work,
either directly or by collaboration. This statement is increasingly applicable to other types
of Oceanographers as well.
The basic ideas of [numerical-modelling??] are simple enough if you understand
Calculus, and especially the concepts of a Taylor expansion. However, the devil is in the
details. Youll have to read a lot more than this section to get all those details! For a general
introduction to the numerical solution of differential equations, you might consult Zwillinger
[1989], which is also a good general reference on differential equations. To learn more about
1
R is an open-source statistical language, available at http://www.r-project.org (August 2009).
18.4. NUMERICAL MODELS 207
Figure 18.2: Using the oce library in the R language to read and plot data from a Seabird
CTD instrument. Note that the format of the data le is somewhat complex, with a header
portion that must be analysed to determine the actual data that are stored.
208 CHAPTER 18. TECHNIQUES
Figure 18.3: Snapshot of model ocean temperature at 15m in the western north Atlantic.
(After Fig. 7 of Maltrud and McClean [2004]).
ocean-circulation modelling in particular, see for example the recent text by Haidvogel and
Beckmann [1999].
This chapter will sketch enough of the topic of [numerical-modelling??] to let you
interact with a modellor, but not enough to let you be a modellor. We have several courses
that will get you further along the road to modelling, but I realize that many of you reading
these notes will never take those courses. Thats why Ive written these words.
Let me start with some terms: OS means Operating System; RAM means Random
Access Memory (the fast memory in your computer, as opposed to the much-slower disk),
CPU means Central Processing Unit (the brains in your computer). Some other important
terms that you are bound to hear in a modelling seminar, or during collaborating with a
modeller, are grid, explicit, implicit, CFL, etc. But youll have to read other sources
to learn what these mean . . .
18.4.1. Who should do modelling?
About putting [numerical-modelling??] programs into the hands of non-experts, a
Dalhousie Physical Oceanographer once said isnt that like giving a handgun to a 5 year
old? (See also Figure 18.4.) We may soon enter an era in which some useful models can
be used reliably by non-experts, but we just arent there as of the time of writing. For now,
the nuts and bolts are best left to mechanics. However, everybody needs to know enough to
interact with the mechanic . . . its no fun leaving the car for an oil change and coming back
to nd thousands of dollars have been spent installing a new engine. Communication, thats
the trick! If youre working with a modeller, you should learn enough about [numerical-
modelling??] to know the scale of the work (are you asking for an afternoons work, or a
years work?), the possibilities and the limitations.
18.4.2. History
It is worth pointing out that [numerical-modelling??] is a new thing. The core math-
ematical technique used in this course is Calculus. Hints of calculus go back a long way
indeed; in trying to work out the area of a circle, Archimedes pondered innite series
18.4. NUMERICAL MODELS 209
To: pom-users@splash.Princeton.EDU
Subject: model request
I request ocean model.I work in university.
README
pom.changes
pom.f
pom.out
USERGUIDE.7-98.ps
curvigrid
Please let me know some information.
Thanks.
To: pom-users@splash.Princeton.EDU
Subject: OBCs
Dear pom user and owner:
I would like to know, what the best open boundary
conditions for a chain sea mountain?
thank you for help.
To: pom-users@splash.Princeton.EDU
Subject: I dont know how to couple MM5 with POM.
Dear pom user and owner:
I am very interested in Air-Ocean interaction,Now I
have a Meteorological Model(MM5) . However, I dont
know how to couple it with POM.It seems to have rather
simply content in the USERGUIDER for the atmospheric
researhers.My god, how can I do?
Figure 18.4: Samples of introductory questions posed on the [POM??] (Princeton Ocean
Model) email group. The aim of the present chapter is to indicate why such questions go
unanswered. I request is asking for the [POM??] code, but this correspondant already
has code, since the email contains a listing of the [POM??] les. Working with [POM??]
requires programming that is far beyond the ability of someone doesnt realize that les
named README and USERGUIDE should be consulted. I would like to know is asking
a much harder question, and evidently knows something about modelling already, but a
question of this type is unlikely to be answered because it is very far from being a simple
matter; things this complicated are normally only discussed usefully in the peer-reviewed
literature. Similarly, I am very interested seeks guidance on a task that is much too
difcult for this discussion group, since the envisioned work could take months or years to
accomplish.
210 CHAPTER 18. TECHNIQUES
comprised of areas of tiny triangles. This visionary thinking was done in about 225 BC.
Deeply signicant progress didnt see the light of day for nearly two millenia after that
time. In the 1500s, the ancient puzzle of determining the area of simple forms (e.g. the
ellipse) were again taken up, by men whose names you should recognize: Descartes, Fermat,
Huygens, etc. But the door to the modern mathematical techniques of calculus was really
opened by no less than Newton. This was in 1666. To place that in history, its a century
after Shakespeares birth and a century before the the birth of Davy Crockett, of the wild
west fame.
In other words, calculus goes back a long way, ten generations by the usual account-
ing of roughly 35y per generation. By contrast, computers are a new thing, developed
within living memory. For example, U. Pennsylvania celebrated the 50th anniversary
of ENIAC in 1996 (see http://www.seas.upenn.edu/

museum for more on the


party). FORTRAN, the programming language that unleashed these early computers (and
whose daughters are still used today) is traced back to the mid 1950s [Backus, 1979,
http://www.computer.org/annals/an1979/a1021abs.htm]. Ocean modelling began not very
long after, so that by the late 1960s the framework of modern numerical models had been
rmly established. Indeed, much of the code in some of todays models is barely modied
from those years. This sort of code, old and trustworthy, is called legacy code.
18.4.3. Computational cost
Why compute the cost?
Before embarking on a modelling project, one should always estimate the computational
cost of the work. Otherwise you may be setting out to dig a foundation with a teaspoon
. . . sure, you can do it, but it will take a lifetime.
The two key aspects of computational cost are CPU time and RAM usage. In theory,
memory and time are linked, in the sense that you could run a model that was larger than
your RAM, but the OS would have to page out to the disk, and that would slow the work so
dramatically (perhaps by a factor of 1000 or even more) as to make it impractical.
The rst thing you need to know is that simulations can be very slow. When you type
a number in a spreadsheet, the screen will normally update in the blink of an eye. This is
because the calculation involved in updating all the visible cells is trivial. However, the
calculations involved in ocean model simulations often take days or months of CPU time,
and such time commitments are signicant with respect to the career of a researcher.
For nite-difference models, it is not difcult to develop crude estimates of time and
memory demands. Since different models employ different algorithms (e.g. some might have
more scratch arrays than others), calibration is required to make the estimates accurate to
better than a factor of 10 or so. Luckily, the calibration is easy to do; just put in a pared-down
model run and compare its memory and time requirements to those calculated below, and
adjust my formulae appropriately.
Online cost calculator
At http://www.phys.ocean.dal.ca/

kelley/PO/modelling/model_
cost_POM.html Ive set up an online calculator to estimate the cost of using a 3D
ocean model, given a domain geometry of your own choosing. (Full documentation of the
calculation is provided on the website.)
Cost in RAM
RAM is often measured in bytes (b), e.g. a typical PC might have 0.1Gb, and a typical
workstation might have 1Gb of memory. It is crucial that the sum of the storage of all the
arrays in the model code to be less than M, the RAM capacity. The rst thing you need to
know about the model is the word size, w i.e. the number of bytes consumed by a single
number (at the precision of the model). Typically w is 4 or 8. Thus, RAM can hold M/w
numbers in total.
It remains only to determine how many numbers the model requires. Ocean models often
use the same grid resolution in x and y, with a different resolution in z, the vertical. Denoting
these resolutions x, y, and z, and letting L
x
, L
y
, and L
z
be the span of the domain of
18.4. NUMERICAL MODELS 211
interest, a 3D model thus has L
x
L
y
L
z
/(xyz) elements to consider. Models normally
have arrays containing the values of S, T and at each of the grid points, and perhaps also
arrays for P, and temporary arrays for each of these, used in updating. Thus, the number of
3D arrays, v, is perhaps as high as 15 or so
2
. To this we would have to add several [2D??]
arrays, e.g. for surface elevation, for bottom depth, etc., but so long as L
z
/z is considerably
in excess of 10 or so, which it typically is, we can ignore a handfull of [2D??] arrays in
comparison to the much-larger [3D??] arrays.
Thus the memory requirement is that
M > wvL
x
L
y
L
z
/(xyz) (18.2)
For example, we might model a cove or a harbour, with L
x
=L
y
=10, 000mand L
z
=30m,
with resolution x = y = 50m and z = 1m. For w = 4 and v = 15, then, the memory
must exceed M = 0.2Gb or so. This is within the realm of a desktop computer. However,
youll note that the resolutions are somewhat crude. For example, if you are interested in
ow around coastline features that are smaller than the chosen x, the model will be of no
help. Put another way, suppose we need to resolve features down to x =y = 5m. That
increases the M by a factor of 100, putting it beyond the range of typical PCs, and into the
range of more expensive workstations.
Cost in time
The main elements in computing time costs are the number of oating-point operations
the CPU can do in a second, F say, the number of operations that are required per time-step
of the model, n
s
, and of course the total simulation interval of the model run, .
The model typically computes derivatives by doing 1 or 2 calculations, e.g. T/x
involves subtracting two T values and dividing by a distance, and so terms like uT/x
involve roughly 3 calculations, plus additional calculations to look up the values of T at the
appropriate locations. Depending on the computer architecture, these look-up calculations
may be trivial or they may be as demanding as the actual derivative calculation. It might
be reasonable to estimate that terms like uT/x take up to 10 calculations, and thus the
three nonlinear terms take up 30 calculations. The mixing terms may impose an additional
load, and so, taking all things together, updating T may take up to 100 calculations at each
grid-point. If the model also uses the salinity, S, then this should be doubled. In addition, it
takes about 50 numerical operations to calculate from S and T. Working with the 2 or 3
components of velocity elds may add 200 calculations to the sum, and more calculations
are of course required to for the pressure equation. Taking these together, the number of
operations is likely to be of order 500 per grid point, per time step. This number may be
accurate to a factor of 3 or so, and calibration tests can easily establish it more precisely for
a given model type.
Since a 3D model has of order L
x
L
y
L
z
/(xyz) grid points, the computational load
should be roughly n
s
= 500L
x
L
y
L
z
/(xyz) operations per time-step. Thus, a simulation
designed to cover units of simulated time, at a time-step of t, will require n
s
/t op-
erations. Clearly, the computation time is made smaller as t is made larger. However,
t cannot be made so large as to smooth over a phenomenon of interest. Also, large
time-steps can cause spurious results in some integration schemes, e.g. explicit schemes
(see section 18.4.8 below) become unstable if t exceed x
2
/(2K), where K is the hori-
zontal diffusivity. Thus, in such schemes, the model run would be expected to take up to
500L
x
L
y
L
z
/(xyz)K/x
2
/F seconds of clock time.
For example, consider a computer capable of F 10
6
operations per second (typical
of machines you might have access to as I write this, in 2001) with L
x
= L
y
= 10, 000m
and x =y = 50m, with L
z
= 50m and z = 1m, and with K = 100m
2
/s as the horizontal
diffusion coefcient, and suppose we wish to run this model for 10 days, perhaps to
determine the response to a storm passing through. This simulation would take roughly
4 days of CPU time on a fast CPU. Thus, this simulation would run in about real time,
taking a clock second to simulate a second in the ocean. Note that decreasing the spatial
2
This variable, v, is of course easily calculated for a given computer code, simply by compiling it with the
desired resolution and domain.
212 CHAPTER 18. TECHNIQUES
resolution x by a factor of 5, perhaps to account for headlands that are 10m wide, increases
the cost by a factor of 5
3
, or roughly a hundred, so that it would take not a few days, but
a full year to run 10 days worth of simulation time. This would be an impractical model
simulation for a graduate student trying to write a thesis quickly!
Range of processor capabilities
Moving from a fast desktop PC to larger computers typically available in regional centers
may improve processing speed by a factor of 10, but rarely by a factor of 100. This factor
of 10 in speed usually comes with a cost factor of 100 or so. This performance/price
relationship is relevant in planning model runs. Your supervisor may buy you a $5000
linux box to run your model, but there is little chance youll get a quarter-million dollars
for a multi-processor box to speed up your work by a factor of 10. And, indeed, the range
between the fastest machines on the planet and the machine on your desk is no more than
about 1000. Since computation time goes as the 5th power of the spatial resolution, that
buys you only 1000
1/5
, or a factor of 4, improvement in the model resolution that can be
handled in a given clock time.
Another factor should be mentioned here, since youll hear it in talking with modellors.
This is that the fast computers typically gain speed by virtue of having several processors.
The problem is that exploiting these processors together requires special coding techniques,
and not all models are formulated to do this. So, if youre working with a modellor who
uses single-processor code, and they say it will take them a month or a year to recode the
model to work on a multi-processor box, they arent lying just to keep you out of their hair.
18.4.4. Example: inertial motion
Consider the differential equations
du
dt
= f v
dv
dt
=f u
(18.3)
for inertial motion (see 5). This is a trivial coupled set of linear DEs, whose solution is
the so-called inertial circle. As a concrete example, lets take the initial condition at t = 0
to be u = 0m/s v = 0.1m/s, and suppose that the location is in the Northern hemisphere at
mid-latitudes, where f = 110
4
s
1
. Then the solution is
u =U sin( f t)
v =U cos( f t)
(18.4)
where U = 0.1m/s.
We might try to solve this numerically by writing the derivatives, using a Taylor expansion,
i.e.
u(t) = u(0) +t
du
dt
[
0
+
1
2
t
2
d
2
u
dt
2
[
0
+. . . (18.5)
where ()[
0
means to evaluate the derivative () at t = 0.
In theory, if t is made arbitrarily small, we can approximate this as
u(t) = u(0) +t
du
dt
[0 (18.6)
to an arbitrary precision
3
.
3
In practice, however, we cannot make t arbitrarily small since computers cannot represent arbitrarily small
numbers. In a computer using real arithmetic, numbers below a certain threshold become indistinguishable from
zero. This is called the underow limitation. Your PC is a 32-bit machine, and it cannot distinguish numbers
that are < 10
38
from zero. Therefore, if the product of t and du/dt falls below this, u(t) will not be different
from u(0). Another limitation is that computers cannot add two numbers that range greatly in size if t is made
too small, then when the computer adds tdu/dt to u(0), it will just get u(0) back again. A way to think of this
is to imagine that the computer is working with decimal numbers, but that it can only handle a certain number
of signicant digits. For example, if it could handle only 2 signicant digits, then adding 0.0001 to 0.1 would
result in 0.1 again, since the small number falls off the end of the calculation. Your PC is running a 32-bit chip,
and it can handle roughly 10 to 20 signicant (decimal) digits, depending on the software youre using. Scientic
workstations typically run a 64-bit chip and they can handle 20 to 40 signicant (decimal) digits. Deeper insights
into numerical limitations are best sought after learning more about the techniques themselves.
18.4. NUMERICAL MODELS 213
With a [rst-order-difference??] approximation to du/dt, then, (18.3) would be rewrit-
ten
u = f vt
v =f ut
(18.7)
and these equations let us step forward from the initial conditions. For example, we know
that u(0) = 0, and that v(0) = 0.1, so that we can estimate u(t) as f vt =0.1f t. For an
example, if t is taken to be 1000s (roughly 15 minutes), we would estimate v at t =t as
0.1m/s, the initial value, plus 0.1m/s 10
4
s
1
1000s =0.01m/s. That is, we would
estimate that v should drop to 0.09m/s in that time. The true solution provided by (18.4)
differs from this estimate by about 0.5%.
You may well ask whether this error accumulates. After 100 time-steps of 1000s each
(to get us up to a day of integration), will we have an error of 50%, which would surely
be unacceptable? The way to answer the question is to perform the 1000 steps, and to
check the results. Many computing languages provide simple ways to perform a repeated
calculation 1000 times. In the Perl language, the method is as illustrated in Figure 18.5. The
answer is that at the end of the day, the error is about 0.003m/s, i.e. roughly 3% of the true
solution. This may well be acceptable for the problem at hand. However, a simple test
4
reveals that decreasing t by a factor of 10, reduces the nal error to 0.3%. In this case, then
the error scales as the inverse of the timestep. You may not nd this surprising in view of the
Taylor approximation. Of course, the computational cost (time on the computer) increases
by a factor of 10 as well. This touches on a central facet of [numerical-modelling??]: the
competition between accuracy and cost.
To be honest, there is really no need to hand-code the rst-difference solutions in Perl, or
any other language. Many software packages will integrate the differential equations for
you directly. Matlab and Octave are good for this, as is Maple. Another good choice is R;
see Figure 5.15 on page 103 for an example (and suggestions for exercises).
18.4.5. Eulerian and Lagrangian models
Eulerian models, such as those sketched above, have the advantage of enormous sim-
plicity. Just write a matrix, whose elements correspond to spatial locations, and adjust it
through multiple timesteps.
By contrast, Lagrangian models dene their grids to be attached to the moving water.
For some purposes that is exactly what is wanted
5
, but unfortunately lagrangian models can
be quite complicated to code.
Most ocean models are eulerian in nature, even if the questions being attacked are
fundamentally lagrangian.
18.4.6. Vertical grid schemes
The vertical coordinate is special in ocean models, since gravity works vertically but not
horizontally, and since the ocean is much wider than it is deep. Therefore, its common
to treat the vertical coordinate differently from the horizontal coordinates. There are three
classes of vertical grids
Level models (Figure 18.6) dene the elds (S, T, etc.) at particular depth (or pressure)
levels, the same levels at each horizontal location. Sometimes the levels are equispaced in
the vertical, sometimes not.
Some models employ so-called reduced gravity grids, and that means that just 2 layers
are present, the lower one being thought of as innitely deep (Figure 18.7).
The class of models known as [isopycnal??] models (Figure 18.8) dene the elds at par-
ticular [density??] values. This has some strong advantages but they come at considerable
book-keeping cost in the coding, and some problems with unique density coordinates.
Sigma models (Figure 18.9) dene elds at particular fractions of local water depth.
For example, one level might be dened to be everywhere at the exact mid-depth in the
4
Try it!
5
Exercise: name some such purposes.
214 CHAPTER 18. TECHNIQUES
#!/usr/bin/perl -w
# Set Coriolis parameter for mid-latitude calculation.
$f = 1e-4;
# Will take 100 steps of 1000s each.
$n = 100;
$dt = 1000;
# Set the initial conditions ...
$U = 0.1;
$u = 0;
$v = $U;
# ... then step forward in time using the first term
# of a Taylor expansion, at each time-step
# printing (time, v, v_error)
print " t v Error\n";
for ($i = 0; $i < $n; $i++) {
$u += $dt
*
($f
*
$v); # line A
$v += $dt
*
(-$f
*
$u); # line B
$t += $dt;
$v_true = $U
*
cos($f
*
$t);
printf "%.0f %.6f %.4f\n", $t, $v, ($v - $v_true);
}
Figure 18.5: Perl code to integrate the equations of inertial motion. If youre not familiar
with the Perl language, note that Perl ignores anything after a hash symbol (#) on a given
line, that Perl designates variable names with a dollar sign, and that variables named to
the left of += are incremented by the amount resulting from the expression to the right of
the +=. The guts of the calculation are at the lines labelled line A and line B. Note
that Perl is not a language that would be commonly used for real numerical work. A better
choice might be the R language; see the text.
Figure 18.6: A level grid. Source: Ocean Models (c) 1995 by the Computational Science
Education Project, http://csep1.phy.ornl.gov/CSEP/OM/OM.html
18.4. NUMERICAL MODELS 215
Figure 18.7: A reduced-gravity grid. Source: Ocean Models (c) 1995 by the Compu-
tational Science Education Project, http://csep1.phy.ornl.gov/CSEP/OM/OM.
html.
Figure 18.8: An isopycnal grid. Source: Ocean Models (c) 1995 by the Computational
Science Education Project, http://csep1.phy.ornl.gov/CSEP/OM/OM.html.
Figure 18.9: A sigma grid. Source: Ocean Models (c) 1995 by the Computational Science
Education Project, http://csep1.phy.ornl.gov/CSEP/OM/OM.html.
216 CHAPTER 18. TECHNIQUES
water column. Some sigma models equispace the levels through depth, while others (like
the S models) concentrate levels near the top and bottom.
As you might expect, there are other types of vertical grid, and of course there are
strengths and weaknesses to each approach. Exercise: list what might be weaknesses in each
approach. For example, noting the fact that the ocean depth varies from location to location,
do some of the models throw away storage on grid points that are below the bottom? In
your thinking, also consider the fact that the stratication varies greatly from location to
location. Another issue, for the [isopycnal??] and sigma models, is that interpolation will
be required to calculate anything along a level surface . . . can you think of anything that
depends on whether a surface (e.g. an isopycnal) is level or tilted?
18.4.7. Finite-difference versus nite-element models
The example in the last section is a nite-difference model, since it is based on assuming
that a derivative may be replaced by a [rst-order-difference??] (or higher-order) approxi-
mation. The wonderful thing about nite-difference models is that they easily let you cast a
grid over a domain, e.g. approximating spatial derivatives on a space step of x, etc., as
the temporal derivative was approximated on a time step of t. A prediction results at
each of many rectangles of the solution space.
Another class of models is the so-called nite-element model. Such models approximate
a change in variables in a somewhat different, and more exible, way. Typically, the
derivatives are calculated in triangles, not rectangles. Furthermore, the triangles need not
be of equal size. Therefore, nite-element grids are much more capable of modelling
ow in complex domains than are nite-difference grids
6
. Unfortunately, the exibility
comes at a considerable cost in programming and algorithmic complexity. For off-the-shelf
models, this is not an issue. However, for research models, requiring great congurability,
e.g. in data-assimilative applications, this can be very problematic indeed. This is because
nite-difference code is sometimes regarded as being simpler to read and modify than
nite-element code.
The ocean modelling community is roughly evenly divided on whether to use nite-
element or nite-difference models. Some models are best suited to some applications, of
course. Also, some models will be faster in some circumstances than others. But there is
also a great deal of inertia involved in the decision of which model to use. Simply stated,
it is hard to switch a research team from one type of model to another. It is even harder
to justify discarding expertise gained through years of labour with one type of model. In
an ideal world, selecting a model type for a new application would be merely a technical
issue. In the real world, experience, familiarity and the potential for collaboration are central
concerns that must be weighed into the decision.
18.4.8. Space-time applications: explicit schemes
Consider the diffusion equation for a tracer in the vertical:
T
t
= K

2
T
z
2
(18.8)
where K is the diffusivity (perhaps molecular, perhaps turbulent).
We might try to solve this be writing the spatial derivatives as rst differences, and also
doing the same for the time derivative. Doing this for the time derivative permits us to write
the numerical solution at the next time in an explicit formula that contains quantities from
previous times. Because there is an explicit formula, this is called an [explicit??] scheme
(section 18.4.9 outlines the implicit scheme, which is another popular approach).
This may be cast in a nite-difference form by writing the derivatives in terms of [rst-
order-difference??] approximations. You should do this as an exercise. For notation, use
T
i, j
to stand for T(it, jz). That is, let superscript i stand for time-step and j stand for
space-step. Your task is to write T
i+1, j
, in terms of items such as T
i, j1
, T
i, j+1
, etc. Connect
6
Just imagine drawing a coastline shape on 1/4-inch graph paper, and approximating the shape of the ocean
using the little squares that line up on the graph paper. Now let yourself use triangles, of any size.
18.4. NUMERICAL MODELS 217
Figure 3: The nite element mesh with 17055 nodes and 30839 elements.
7
Figure 18.10: Mesh used by the WebTide model of tides of eastern Canada and northern
U.S.
Figure 18.11: Illustration of the connectedness of terms in a rst-order explicit scheme
for the diffusion equation F/t = K
2
F/z
2
. The white boxes hold terms used in the
computation of the gray box. Here i refers to space and j to time. The advantage of such
diagrams, sometimes called computational molecules, is that they can indicate the gist
of a scheme at a glance. For example, this is an explicit scheme since the i, j +1 term is
computed solely on the basis of j terms.
218 CHAPTER 18. TECHNIQUES
Figure 18.12: Demonstration of CFL stability condition for explicit schemes for solving
the diffusion equation. Here, a tracer was initially inserted at concentration T(z) = 1 in the
2.5-cm tall patch centered at z = 0. Lower panel: simulation with t = 0.2K/z
2
, i.e. in
the stable range for the CFL condition. This graph is visually identical to the results with
any time-step t that satises tK/z
2
<< 1/2. By contrast, the upper panel shows the
results just on the border of instability. The wiggled contours indicate spurious variation.
This erroneous behaviour intensies if tK/z
2
is increased beyond 1/2; for example, a
graph with a CFL parameter of 0.6 has so many contours that it simply blackens the axis
area. (Data created by Perl script shown in Figure 18.13.)
your formulae to the initial conditions, i.e. T
0, j
, and to boundary conditions. Figure 18.11
illustrates the connectedness of various terms in a rst-order difference scheme. Youll see
gures like this in explanations of numerical schemes.
The update equation for T has a term that contains the expression tK/z
2
. This is a
nondimensional number expressing the size of the time step in terms of the time it takes to
diffuse across one spatial step. Your intuition may suggest that this is an important parameter
in the numerical technique, and youd be right! This parameter is called the CFL parameter
(for Courant, Friedrichs and Levy, three folks who thought about this sort of problem),
and it must be small or else all heck breaks out. The form of the error that arises if the
CFL stability criterion is violated is an alternating two gridpoint oscillation . . . but the
best way for you to learn about that is just to try coding this up for yourself. Figure 18.13
shows sample code in the Perl language. Figure 18.12 illustrates for a case that is just barely
unstable . . . if the CFL condition were violated much more than this, the numerical estimate
of the solution would have exponentially-growing errors. For computational efciency,
tK/z
2
should be as large as possible (i.e. time-steps should be as large as possible) but
the CFL stability requirement demands that tK/z
2
be no larger than 1/2.
18.4.9. Space-time applications: implicit schemes
In addition to explicit techniques, as sketched in the last section, ocean modellors also
have available so-called [implicit??] schemes. In these, the solution at some time-step is
computed based not just on properties of the previous time-step but also on (some) properties
18.4. NUMERICAL MODELS 219
#!/usr/bin/perl -w
# Solve diffusion equation, demonstrating issue with CFL condition.
# Next line sets the CFL parameter. Do not exceed roughly 0.4
# or instability ensues. The value 0.49 gives enough wiggles
# to demonstrate, without filling the whole domain with crazy
# contours. The value 0.20 yields smooth contours.
$CFL = 0.20; # CFL parameter
$nz = 21; # number of spatial grid steps
$dz = 0.025; # size of spatial grid
$K = 1e-5; # vertical diffusivity
$dt = $CFL
*
$dz
*
$dz / $K;
$final_time = 1000; # in seconds
$nt = $final_time / $dt; # number of steps
print $dt
*
$K / $dz / $dz, "\n";
print $nz, " ", $nz
*
$dz, "\n";
print $nt, " ", $nt
*
$dt, "\n";
# Set initial conditions
for ($i = 0; $i < $nz; $i++) {
$T[$i] = 0;
}
$T[$nz / 2] = 1;
# Saving this constant factor speeds execution
$factor = $K
*
$dt / $dz / $dz;
for ($t = 0; $t < $nt; $t += 1) {
for ($i = 1; $i < $nz - 1; $i++) {
$Tnew[$i] = $T[$i]
+ $factor
*
($T[$i-1] + $T[$i+1] - 2
*
$T[$i]);
}
for ($i = 1; $i < $nz - 1; $i++) {
$T[$i] = $Tnew[$i];
}
for ($i = 0; $i < $nz; $i++) {
printf "%f ", $T[$i];
}
print "\n";
}
Figure 18.13: Perl script for explicit solution to diffusion equation, used for Figure 18.12.
220 CHAPTER 18. TECHNIQUES
of the present time-step. The computation has no explicit formula (of the type that I hope
you wrote down, in response to my suggestion in a footnote of the last section). Rather, it
requires the inversion of a matrix.
The advantage of implicit schemes is that they are not bound by the CFL condition as
explicit schemes are. That means an implicit scheme can take large time steps. However,
there are also disadvantages. For one thing, the coding is less straightforward since it
involves matrices that may be very large and therefore best dealt with using special-case
matrix inverters (which rely on the sparseness of the matrices). For another, implicit schemes
may suffer in accuracy compared with explicit schemes. And, of course, the use of a large
time-step means that rapidly-varying phenomena cannot be modelled.
18.4.10. Special problems for ocean models
Time-scales of variation
Surface modes of variation have very short timescales, compared with interior modes. To
see this, note that surface waves move at speed (gH)
1/2
where H is the water depth, yielding
an inherent timescale of L/(gH)
1/2
in a domain of horizontal scale L, whereas interior waves
move at speed (gH/)
1/2
), yielding a timescale of that is (/)
1/2
longer. Typically
/ is of order 10
3
or so, which means that the interior modes evolve 30 times more
slowly than exterior modes. The exterior modes thus set a bound on the speed of numerical
integration. For this reason, many numerical models separate the interior and exterior modes,
using a large time-step for the former and small one for the latter. This increases efciency,
at the cost of requiring code to link the two modes together from time to time.
The problem of open boundaries
If you are modelling a local bay, there is little difculty in providing boundary conditions
at the sea-land interfaces: you may say (simply and conventionally) that water cannot
penetrate the coastline, or you may say (more rarely, and mainly in high-tide areas) that
it can swell up over the land, but only a certain distance. However, the open boundary,
i.e. the region where the bay meets the sea, poses special difculties. Just consider the
sealevel-provided pressure gradient at that place, for example. It cannot be known unless
the sealevel is known (predicted) both inside the bay and outside it. But that means that the
model has to extend outside the bay, ultimately through the whole world ocean. Well, we
cant model the whole world ocean because our computer is small, so we must do something
reasonable at this open boundary, which means making some assumptions about the open
ocean beyond the bay. This is the problem of open boundaries. Many solutions have been
attempted to resolve this problem, including most prominently radiation conditions. These
are conditions that couple time-varying sealevel tilts and velocities together, in accordance
with the tilt/velocity relationship that holds for gravity waves.
18.4.11. Code example
Most models are written in the FORTRAN programming language. Coding models is
very difcult (e.g. often running to several person-years or even person-decades, adding up
all the pieces of the puzzle). Even reading a pre-coded model is arduous, but it is worth
trying, if only to gain some appreciation for the scale of the work of model development
7
.
Figure 18.14 shows a snippet from the popular Princeton Ocean model. To give an idea
of the scale of the model, it comprises over 2900 lines of fortran code, so the snippet is just
1% of the whole code. One of the rst things youll want to look at in a code is the start of
the le, since that is where storage is set aside. (In other words, this is where you can look
to learn how much storage the model requires.) The DIMENSION statements are declaring
space for various arrays, e.g. ADVX(IM,JM,KB) is a 3D array that holds a term relating to
advection in the x direction, TCLIM(IM,JM,KB) holds a climatological temperature eld,
etc. The manual will tell you more on each of these things. But youd do well to note that
almost nothing about this code snippet is immediately understandable. For an example, can
you tell which (if any) of these arrays stores the temperature eld? About all you can be
reasonably certain of is that GRAV probably holds the gravitational acceleration, given the
variable name and the value used, and that PI is .
7
The mental anguish in getting the algorithm working probably outweighs that of coding by a factor of 10 or so.
Blood, sweat, and tears may hide behind the most innocent-looking lines of code!
18.5. DATA ASSIMILATION 221
INCLUDE comblk97.h
LOGICAL SEAMT
DIMENSION UTB(IM,JM),VTB(IM,JM),UTF(IM,JM),VTF(IM,JM),
1 ADVX(IM,JM,KB),ADVY(IM,JM,KB),ADVUA(IM,JM),ADVVA(IM,JM),
2 TSURF(IM,JM),SSURF(IM,JM),
2 DRHOX(IM,JM,KB),DRHOY(IM,JM,KB),DRX2D(IM,JM),DRY2D(IM,JM),
3 TCLIM(IM,JM,KB),SCLIM(IM,JM,KB),ADX2D(IM,JM),ADY2D(IM,JM),
4 SWRAD(IM,JM),
5 X(IM),Y(JM)
C--------------------------------------------------------------------
REAL ISPI,ISP2I
DATA PI/3.141592654/,SMALL/1.E-10/
C OPEN(40,FILE=INCOND ,FORM=UNFORMATTED)
GRAV=9.806
TBIAS=0.
SBIAS=0.
IINT=0
TIME=0.
NREAD=0
Figure 18.14: Snippet from the start of the Princeton Ocean Model. Note the denitions
of several matrices, e.g. UTB(IM,JM); a good way to learn about a model is to examine
the dimensions of such matrices. As this code illustrates, it is not uncommon for in-code
documentation to be sparse.
Further down in the model, youll nd code that works with lateral mixing parameter-
izations. This is shown in Figure 18.15. A glance is enough to reveal that this is a triple
loop, evidently running across x, y, and z directions. Also, you can see that something is
being done with velocity derivatives here, since the U(I+1,J,K)-U(I,J,K)/DX(I,J) term has
the form of a rst-order approximation to a derivative. But as to the exact details, youd be
well advised to start by translating the fortran to a cleaner mathematical notation, and also
reading the literature on the Smagorinsky scheme employed here.
The point of showing these code snippets here is to let you know that modelling work is
non-trivial. Beyond the coding work, there is the work of the algorithms and parameteriza-
tions within the code. For example, the second snippet shows the Smagorinsky scheme, and
before using such a scheme, youd be very well advised to read several papers about its pros
and cons compared with competitive schemes.
18.5. Data Assimilation
Introduction
Data are from Mars and theories are from Venus.
From the preceeding pages, you may have the idea that we know a lot about the theory
of the ocean, and that is true. In fact, the so-called Navier-Stokes equations
8
are thought to
be a very accurate representation of ocean ows. The problem is that these equations cannot
be solved in closed form by writing down a solution in terms of simple functions. No,
the equations admit so many solutions all the types of ows that youve ever seen, and an
innite number more that the solutions cannot be written down, except in mathematically-
degenerate cases. (For example, the constant-viscosity equations for steady Ekman dynamics
are one such mathematically-degenerate case.) But, this is no academic exercise in writing
formulae; we need to predict ocean ows, to understand the climate, to predict where
pollution will go, to predict storm surges, etc.
The only path is to work for numerical, not analytical, solutions to the equations, i.e. to
do [numerical-modelling??]. This is dealt with in section 18.4, but you neednt understand
all of that to get an idea of what data assimilation is all about.
8
Which you can learn about more properly in my Fluid Dynamics course.
222 CHAPTER 18. TECHNIQUES
C
**********************************************************************
C HOR VISC = HORCON
*
DX
*
DY
*
SQRT((DU/DX)
**
2+(DV/DY)
**
2
C +.5
*
(DU/DY+DV/DX)
**
2)
C
**********************************************************************
C If MODE.EQ.2 then initial values of AAM2D are used. If one wishes
C Smagorinsky lateral viscosity and diffusion for an external mode
C calculation, then appropiate code can be adapted from that below
C and installed after s.n 102 and before s.n. 5000 in subroutine ADVAVE.
DO 95 K=1,KBM1
DO 95 J=2,JMM1
DO 95 I=2,IMM1
AAM(I,J,K)=HORCON
*
DX(I,J)
*
DY(I,J)
1
*
SQRT( ((U(I+1,J,K)-U(I,J,K))/DX(I,J))
**
2
2 +((V(I,J+1,K)-V(I,J,K))/DY(I,J))
**
2
3 +.5E0
*
(.25E0
*
(U(I,J+1,K)+U(I+1,J+1,K)-U(I,J-1,K)-U(I+1,J-1,K))
4 /DY(I,J)
5 +.25E0
*
(V(I+1,J,K)+V(I+1,J+1,K)-V(I-1,J,K)-V(I-1,J+1,K))
6 /DX(I,J))
**
2)
95 CONTINUE
Figure 18.15: Snippet from the Princeton Ocean Model, in a section dealing with horizontal
viscosity. Exercise: write the equation in the loop in standard mathematical notation, to try
to learn what this Smagorinsky mixing scheme is all about.
Figure 18.16: A 3D view of a computational grid. Source: Ocean Models (c) 1995 by the
Computational Science Education Project, http://csep1.phy.ornl.gov/CSEP/
OM/OM.html.
18.5. DATA ASSIMILATION 223
The problem with the models is that they contain unknowns. Some things are understood
very well, e.g. the Coriolis parameter, the gravitational forces, and so on. But other things
are poorly known, even though they may be important to the ocean dynamics. Examples
might be the strength of the wind blowing on the ocean at a given place and time or the
strength of bottom-drag friction that is occuring someplace. Sometimes these things take
the form of boundary conditions (e.g. the wind), and sometimes they take the form of
parameters (e.g. the C
D
for bottom drag). If these things are being observed, well thats
ne, and we may proceed more or less directly to a numerical prediction of the ocean. But
sometimes these particular things are not observed, but other dynamically-relevant things
are observed. For example, we might be trying to predict storm surges at Halifax, and we
might have good sea-level information about the observed surge at some other location.
This sea-level information, or data, does not appear as a forcing term in our equations,
so its measurement doesnt help us directly. However, it is information that is dynamically
relevant. The goal of data assimilation is to assimilate such data into models to use
whatever information we have, even if it is slightly indirect, to constrain the solution
provided by the model.
An analogy might help here. The models that are used for weather prediction are
prognostic models, just like the ocean models we are familiar with. That is, they make
predictions about the future, based on knowledge from the past, by integrating equations
of the form (. . . )/t = . . . . But would it make sense to use a weather prediction from
last month that says today will be rainy, when a glance out the window indicates that it is
perfectly ne, and your memory tells you it has been ne for days? No, of course not. Your
own internal model will tell you to ignore the old predictions, and to instead make a new
prediction based on the data that you have on hand. But it might not make sense to entirely
ignore the old predictions, since they were, in themselves, based on perfectly useful data. It
seems that the best plan would be to try to use whatever data are on hand, to try to pull an
errant model back to reality. That is the essence of data assimilation the incorporation of
observations into models. As you might imagine, there is no one technique of doing this.
Several approaches are being explored in the research community.
Returning to the weather example, one approach would be to discard yesterdays pre-
dictions as soon as todays observations come in. But that has a problem, because the
observations may contain errors, and they may also be inconsistent with the dynamics of the
model. For example, the wind you measure in your backyard may be biased because of the
house nearby, so that the data from your wind gauge, inserted in a model, would bias the
ow direction, perhaps in contradiction to other data. Also, the wind may be quite gusty,
but you may be writing down the observations for today at some particular time, and so
they may have a large random component that will mess up a model. Indeed, when new
daily data are inserted into weather models, the model solutions get perturbed at rst, giving
inaccurate results, until the model sorts out the contradictory data. This might argue, then,
for a sort of continuous blending-in of data, rather than doing so at particular times. Whether
to insert or to blend, and how to blend these are the sorts of issues that are being examined
at present. Different approaches may be better for different models, and even for the same
models being used for different purposes.
One popular application of data assimilation is operational oceanography, which
combines realtime ocean measurements to make timely predictions of the ocean state, much
like weather prediction.
Example (nudging)
A simple example may clarify this. Suppose we have a model for ocean temperature, of
the form
T
t
= . . . (18.9)
where . . . stands for various effects, e.g. advection, solar heating, vertical mixing, etc. Now,
if we (somehow) knew the initial state of the ocean, T = T(x, y, z), then the (18.9) could, in
principle, be used to make predictions for any future time. But the initial measurements
are necessarily awed, and the . . . terms will also be imperfectly known, so you might
224 CHAPTER 18. TECHNIQUES
not be surprised to nd that the model makes predictions that drift away from the initial
observations.
Imagine that the model makes predictions that seem reasonable except that the inter-
season variation in temperature is too small. And further, imagine that have access to
historical ocean-temperature data in the form of seasonal T(x, y, z) values which we may
interpolate into a continues time axis and denote T
o
(x, y, z, t).
Now consider a new model identical to the rst except for the addition of a single term,
T
t
= (T T
0
)/ (18.10)
where is a time-scale. This last term will be zero if the predicted ocean state matches the
observations from previous seasons. But if the T is too warm, this term will be a negative
number; that means it will have the effect of cooling the ocean. Similarly, if T < T
O
, this
term will warm the ocean.
We call this term a nudging term, because it nudges the ocean back into a desired
state. The term can be made weaker or stronger, by making larger or smaller. To
understand the meaning of , imagine that the observations indicate that the ocean suddenly
warms at some particular time; is the time it would take for the ocean to adjust to this
observed temperature.
Nudging is a form of data assimilation that is not just simple to understand but also
reasonably effective
9
. Sometimes the nudging term is applied only at large scales, so that the
small-scale predictions of the model are not affected. It may be applied to various aspects
of the ocean state, not just to temperature; e.g. nudging velocity based on current-meter
observations. It may also be done using historical data or continuously incoming (realtime)
data.
Further reading: Thacker [1988] provides an extensive, gentle introduction to data
assimilation, suitable for many students in this class. Students with strong mathematics
skills could also tackle the more technical treatment provided by Thacker and Long [1988].
9
Im being vague here. To learn more, come to our seminar series or read some of the papers were publishing
lately.
CHAPTER 19
People in Oceanography
T
HIS CHAPTER PRESENTS SOME PHOTOS of oceanographers whose names youll
hear in the course. The photo sources are listed in the captions. Ill leave it to
other people to write biographies of these scientists, but where Ive seen a good
quote, Ive put it below.
Figure 19.1: Gordon Riley is shown working in a laboratory on R/V Atlantis, ca. 1940.
Riley set up the Department of Oceanography at Dalhousie University. (See http://
oceanography.dal.ca/ocean_1094.html for a biography, including a note on
the source of this image.)
225
226 CHAPTER 19. PEOPLE IN OCEANOGRAPHY
Figure 19.2: Henry Stommel is shown in his mature years. (Image source: http://
www.whoi.edu/media/stommel.html.) Upon receiving WHOIs Bigelow Medal
in 1974, Stommel said Most human history has not afforded men much chance to pursue
their curiosity, except as hobby of the rich or within the refuge of a monastery. We can count
ourselves fortunate to live in a society and at a time when we are actually paid to explore
the universe.
227
Figure 19.3: Walter Munk is shown in his mature years. (Image source: http:
//scrippsnews.ucsd.edu/Releases/?releaseID=654.) You may enjoy the
viewing Munks answers to biographical questions, at http://igpp.ucsd.edu/
aboutigpp/history/munk.html.
228 CHAPTER 19. PEOPLE IN OCEANOGRAPHY
Figure 19.4: Harald Sverdrup shown on the Scripps pier, with a current meter (1940),
and in a portrait by Paul Williams (1946). (Image source http://scilib.ucsd.edu/
sio/archives.)
CHAPTER 20
Physical Oceanography in Just
Two Pages
Everything is everything.
Youre missing c _2002 Bruce Springsteen
ive got a new song to sing
its longer than all the others combined
and doesnt mean a thing
grandpas interview c _2003 Neil Young
I
THINK THE WAY TO STUDY FOR AN EXAMINATION is to write a three-page summary
of the course material. In case it helps, I provide here such a summary. It centres on
concepts and formulae, rather than geography (e.g. what we mean by Gulf Stream)
or techniques (e.g. the TS diagram), since Ive found that most students can remember the
latter without difculty. But of course your summary may look nothing like mine.
20.1. Gravitational stability
Stability is measured by
N
2
=
g

z
where N is called the buoyancy frequency (it is the frequency of internal waves) or, less
meaningfully, the Brunt-Vaisaila frequency. The ratio
Ri =
N
2
(u/z)
2
,
called the Richardson number, measures the competition between the tendency of vertical
shears of horizontal currents to cause overturning and the tendency of gravitational stability
to inhibit such overturning motions. (Here u stands for the magnitude of the horizontal
velocity components, not for the x-component as it normally does.)
Q: should mixing increase at high Ri or low Ri?
20.2. Hydrostatic Approximation
The z-component of the momentum equations yields the hydrostatic approximation
p
z
=g since all other terms in the equation (e.g. w/t) are small compared to g. Q: of
what use is this equation?
A special case for layered ow is P(z) = P
atm
+g( z) where z is the a depth of
interest and measures the deviation of the water surface from the level z = 0. Q: how does
this relate to gravity waves? . . . to geostrophic currents?
20.3. Geostrophy
A central formula is
f v =
1

p
x
Q: what is formula for u?
Exercise: write an essay, with mathematics and calculations, about geostrophy in wide,
frictionless rivers, e.g. considering
=
f
g
_
W
0
udy
229
230 CHAPTER 20. PHYSICAL OCEANOGRAPHY IN JUST TWO PAGES
20.4. Ekman Flow
Understand basic elements of Ekman ow (wind, transport to the right, direction of spiral,
etc). Know the formula for the Ekman ux
ux in units of m
2
s
1
=

wind
f
20.5. Potential Vorticity, Sverdrup ow
You should understand conservation of potential vorticity, and be able to draw pictures
of the physics. The important application of Sverdrup ow should be clear to you.
20.6. Rossby Number, Radius, . . .
You should know the form of the [Rossby-number??] and [Rossby-radius??] and
know these mean physically. You should know (or be able to derive) typical values in
various situations.
20.7. Waves
Shallow-water waves have phase speed C
P
=

gh where water-depth is h; deep-water


waves have phase speed C
P
=
_
g/k where k is the wavenumber. Q: (1) What are the group
speeds in these two cases? (2) What are the dispersion relationships in these two cases? (3)
What is the general signicance of the dispersion relationship?
CHAPTER 21
Answers to Exercises
Well I tell them theres no problem, only solutions
Watching The Wheels c _1980 John Lennon
A3.1 Fate of a Clumsy Olympic Diver (page 36)
Question. A diver jumps from a springboard that is 3m above the surface of a swimming
pool. Its an ungainly sort of dive more a stumble, really, with no upward motion whatever.
(a) How long will it take her to reach the water? (b) What will be her vertical speed when
she hits the water? (c) How would these things be changed if the diver ran off the end of the
board, instead of falling off?
Answer. Well start with a few simplications. Lets suppose that the diver holds her
body rigid during descent. (Otherwise well have to worry that she could delay the time of
impact by pulling her knees up to her chest, for example.) Lets also assume no air friction.
(Although adding air friction isnt terribly hard in principle, it does require calibration for
body shape, etc. You may know, for example, that male swimmers often shave their chests,
to reduce friction in the water.) Finally, lets assume that the intitial fall speed is zero that
the swimmer does not jump up before leaving the diving board.
Now lets move on to notation. Let z be the coordinate measuring distance above the
water surface, and let = (t) be the height of the swimmer as a function of time t. While
were at it, lets let t = 0 denote the time when the diver leaves the diving board. Thus, we
have initial condition that (t) = h.
For later times, Newtons second law gives
m
d
2

dt
2
=mg, (21.1)
where g is the acceleration due to gravity. Notice that the mass m cancels in this expression
1
,
yielding
d
2

dt
2
=g. (21.2)
Integrating once with respect to time, and dening w = d/dt as the fall-speed, yields
w =gt +A, (21.3)
The constant of integration A is determined by the initial conditions. Since the diver starts
from rest (i.e. w = 0 at t = 0) A must be zero, so that the fall-speed is
w =gt. (21.4)
A second integration gives
=
1
2
gt
2
+B, (21.5)
where B is another constant of integration
2
. From the initial condition that (0) = h, we see
that B = h, so we are left to conclude that the diver has trajectory
= (t) =
1
2
gt
2
+h. (21.6)
1
There is something deep and important about this, but Ill leave that to another Physics course!
2
You will have expected, I presume, to encounter two constants of integration for a second-order differential
equation.
231
232 CHAPTER 21. ANSWERS TO EXERCISES
From this, one may calculate the time to impact, simply by substituting = 0 and solving
for time, which we may denote as t
i
, say. The result is t
i
=
_
2h/g. For our particular case,
with h = 3 m and with g = 9.8m/s
2
as usual, we obtain t
i
= 0.8 s, a time that is, in a word,
a heart-beat. Its short enough that its a challenge to see what the heck is going on during
the fall . . . thats why non-experts like me can tell little about the skill of divers on television,
without a commentator to say ah, good one or too bad, her career is over now.
Now we move on to the issue of the impact speed. This is easily determined by substitut-
ing t
i
into (21.4), yielding impact velocity
w
i
=gt =g
_
2h/g =

2
_
gh. (21.7)
Ive written it out in this way, with a

gh term, so that you can see a parallel with the
velocity of shallow-water gravity waves see section 10.4.2. Substituting yields w 8m/s,
a velocity approaching that of the swiftest 100 m runners. (As an extra example to do
yourself, calculate how high the board would have to be to achieve 10 m/s, which is more
typical of world-class sprinter speeds.)
A3.2 Fate of a Bridge Diver (page 37)
Question. As a New Years celebration, Italians sprinkle wine onto the water and dive
in. (a) How long will it take before the bridge diver hits the water? (b) How fast will he be
going when he hits the water?
Answer. First we must estimate how high the bridge is above the water surface. (Youll
need to refer back to Figure 3.1 on page 36 to follow this.) Let me start by assuming that the
lighted triangle on the bottom-right is the bridge. Then Ill assume that this bridge is level
with the road that can be seen in the background to the right of the divers legs. Given this,
the height of the bridge above water can be estimated by inspecting the staircase leading
from the road to the water. The dark fuzz along the stairway is a mass of people. It is unsafe
(and illegal) for the railing of a staircase to be below the center of mass of average humans,
because otherwise folks could topple over the railings easily. (Until recently, the central
staircase in Oceanography at Dalhousie was unsafe in this way.) However, if the railing is
too high, people whine because they have no view. From this we can infer that the railing
is probably 1/2 the average height of humans, or about 1 m. Given this scale, the staircase
height can be gauged to be 10 m to 20 m. Designating the height as h, we have h = 15 m as
a rough estimate
3
.
(a) Neglecting air friction, the diver will accelerate at the rate g =9.8m/s
2
. Therefore
Newtons second law gives us
d
2
z/dt
2
= g (21.8)
Here z is the divers vertical coordinate and t is time. According to the principle that the
laws of physics are independent of a translation of a coordinate system, we may put z = 0
anywhere we like. Lets choose z = 0 to be at the bridge surface, so that the water surface is
at z =h. Correspondingly, lets choose t = 0 to represent the time when the diver jumps
from the bridge. Integrating the above differential equation twice yields z as a function of
time t. The result is z = gt
2
/2. From this we get t =

2z/g for t as a function of z


(the time at any given position). The time to travel from z = 0 to z = h is thus

2h/g.
Substituting, we get a time of

220m/(9.8m/s
2
), or roughly 3 s.
(b) Integrating d
2
z/dt
2
=g once, and dening w = dz/dt as the vertical velocity, we
get w =gt. Substituting t = 3s, we thus get that w =30 m/s. (The negative sign means
the diver is falling, of course.) How fast is this? Well, consider a 100 m sprint. The best
runners in the world can cover that distance in about 10 s, so 10 m/s is the fastest a human
can run; from this we see that the diver is going three times as fast as the swiftest runner.
Put another way, 30 m/s is 303600 m/hour, or 108, 000 m/h, or 108 km/h. Thus, the diver
will hit the water at highway speed. Does this sound like fun to you?
3
If you think this scale-estimation analysis is silly, think again estimate physical scales rapidly, for use in
quick calculations, is a big part of physical science. Its the surest way to impress physics nerds!
233
A3.3 Automobile (F1) Engines (page 37)
Question. New rules prevent F1 engines from being turbo-charged. So, to get the
required horsepower, the engines have been modied. Car and Driver magazine, Oct 1999,
gives some data on these engines. The engine revs at 18,000 RPM and the pistons have a
stroke of 46mm. The connecting rod and piston head weigh 15 ounces together. What is the
maximum velocity of the pistons? What is the maximum acceleration of the pistons? What
force is exerted on the connecting rod?
Answer. Let y measure the coordinate of the piston, with y = 0 midway through the
cylinder, y = L at the top of the stroke and y =L at the bottom of the stroke.
We will describe the cyclic motion by y = Lsin(2t/, where is the period. From the
problem statement, l =46mm/2, i.e. l =0.023 m. At 18000 RPM, or 300 cycles per second,
we thus have = 1/300 s or = 0.0033 s.
The velocity is v = dy/dt, or v = 2L/ cos(2t/), so that the maximum velocity is
v

= 2L/, or 43 m/s. This converts to 150 kph, or 80 mph. Not too shabby.
Similarly, the maximum acceleration is a = (2/)
2
L, or 81, 000m/s
2
(which is 8000
times the acceleration of gravity).
To get the force, well use f = ma, with m = 0.4 kg for 15 ounces. (One ounce is 30
grams. Ask your dealer.) Therefore the force is 32, 000 Newtons. Recalling that 1 kg weighs
10 Newtons, this corresponds to the weight of 3200 kg, or 7000 pounds, or about 3 tons. Any
way you convert it, its a lot!
How close this engine is to coming apart at the seams? To answer this youd have to
know the type of metal and the cross-sectional area of whatever is being tugged at. A rough
calculation may be done by noting that the tensile strength of steel is in the range of 30,000
to 100,000 pounds per square inch. My reading of the photo in the Car and Driver article
suggests that parts of the apparatus are about 1 inch wide, or less. Id really doubt that its
more than 0.5 inch thick, so the cross-sectional area is perhaps a less than 0.2 square inches.
Since 7000 pounds is tugging on this, we have at least 24 000 pounds of force per square
inch. Luckily for the car driver, this tension is under the strength limit for the metal.
Note that my calculation is pretty rough; if I had guessed that the metal piece is 1/8-th
inch thick, the force per unit area would be doubled.
Kinesiology students might like to gure out, for comparison, how much axial stress a
bone (say, a femur) can take before it breaks . . . Im guessing its less than 3 tons.
A4.1 Salinity (page 55)
Question. In terms of kitchen-based tools, what does a salinity of 35 PSU mean?
Answer. When youre done with a 1-liter container of milk, thrown in as much table
salt as you can put in the palm of your hand (roughly 35 g). Then ll the container with
water from the tap. Since water density is = 1000 kg/m
3
, and since a litre is 1/1000th of a
cubic metre, you will have added 1kg of water. Thus, the salinity is 35 g per kilogram, i.e.
S = 35PSU.
You might like to taste the water, but please do not drink it, because salt is not good for
your blood pressure.
A4.2 Freshwater density (page 57)
Question. Roughly what is the mass of the water in your bathtub?
Answer. A bathtub is roughly 0.5 m wide and 2 m long, and often lled to about 0.3 m
depth. Thus, the volume is V = 0.5m2m0.3m or V = 0.3m
3
. The density of fresh
water is = 1000kg/m
3
(with a slight dependence on temperature, but one that is certainly
too small to consider in this rough calculation), and so we see that the mass is M = V, or
M = 300kg.
234 CHAPTER 21. ANSWERS TO EXERCISES
You might nd it instructive to compare this result to your body weight.
A4.3 Seawater density (page 58)
Question. What is the density of seawater at salinity S = 35 PSU, temperature T = 10

C
and pressure p = 100 dbar (i.e. at roughly 100 m depth)? If the water were to be raised
adiabatically to the surface, what would its temperature and density become?
Code to handle the density of seawater is available in many computing languages.
Answer in the R language. Using the oce package of the R language
4
, First, we must
load the package (you may need to install it rst)
> library(oce)
(the > is the R system prompt, i.e. material on lines starting with that symbol is a command
given to R) and then we compute the in-situ seawater density = (S, T, p) as
> sw.rho(35, 10, 100)
[1] 1027.404
(The [1] indicates that only a single number results from the command, in this case the
density = 1027.404 kg/m
3
.)
Next, we compute the temperature of the water parcel if it is raised adiabatically to the
surface (i.e., without exchanging heat with the surrounding water)
> sw.theta(35, 10, 100)
[1] 9.988452
(This is just the potential temperature, =(S, T, p), referenced to the surface; the package
permits reference to other pressures, of course.)
Finally, the density at this the surface
5
is given by
> sw.rho(35, sw.theta(35, 10, 100), 0)
[1] 1026.954
where weve kept S the same, and changed the local pressure to p = 0, which indicates
atmospheric pressure.
Before moving on to other computing languages, it is useful to note that these calculations
reveal some rough rules that are worth remembering. A well-mixed water column is not
uniform in in-situ temperature; rather, the temperature will increase by roughly 0.01

C per
100 m of depth, while the density will increase by about 0.05 kg/m
3
.
Answer in the Python language. First, we load the seawater package
6
>>> from seawater import
*
(note that the prompt symbol in Python is >>>), and then we compute the density
>>> dens(35, 10, 100)
1027.4044374075813
Good. That matches the R result to within the number of digits printed by R. Next, we
calculate the potential temperature
4
R is an open-source statistical language, available at http://www.r-project.org (August 2009).
5
Since this is a calculation that is done often, it is provided by a one-step subroutine in this package, and in
those of described next. But its more useful to write things out explicitly, for the present purpose.
6
At http://www.imr.no/

bjorn/python/seawater/index.html (August 2008).


235
>>> temppot0(35, 10, 100)
9.988470406375999
and it agrees with the R result to the number of digits normally used in such calculations.
(Small differences in algorithms explain the differences; there are multiple methods of doing
these calculations, based on formulae developed from regressing different data sets.) Finally,
we calculate the surface density
>>> dens(35, temppot(35, 10, 100), 0)
1026.9543890212447
in good agreement with the R results.
Answer in the Matlab language. The CSIRO seawater library
7
produces similar results
(commands are followed here by the results, or answers):
We begin by telling matlab where to nd the seawater library
>> addpath /matlab/seawater
(where, again, the prompt is >>, and where of course you may store the library in another
place, at your whim.) Then, we proceed as in the previous examples:
>> sw_dens(35,10,100)
ans =
1027.40443740758
which matches the previous calculations, as does the potential temperature
>> sw_ptmp(35,10,100,0)
ans =
9.9884517161377
and the density at the surface
>> sw_dens(35,sw_ptmp(35,10,100,0),0)
ans =
1026.95438902125
which the impatient may also compute with
>> sw_pden(35,10,100,0)
ans =
1026.95438902125
A4.4 Drag force from wind stress (page 67)
Question. A 10 m/s wind blows steadily over the water surface. (a) What is the wind-
stress that arises from this? (b) What if the wind is 5 m/s? (c) What if the wind is 80 km/hour?
Answer. Air density is roughly 1.2 kg/m
3
, and using a C
D
value of 1.5 10
3
in the
drag-force law (4.28), we get the following.
(a) For a 10 m/s wind, the stress is = 1.5 kg/m
3
1.210
3
(10m/s)
2
= 0.2Pa. (Note
that 10m/s is a reasonably brisk wind, but one that you can walk in easily.)
(b) For a 5 m/s wind, the stress is = 0.05 Pa. (Note that halving the wind-speed reduces
the drag by a factor of 4.)
7
At http://www.cmar.csiro.au/datacentre/ext_docs/seawater.htm (August 2008).
236 CHAPTER 21. ANSWERS TO EXERCISES
(c) An 80 km/hour wind is 80 10
3
m/3600s = 22m/s. This yields a stress of =
1.5kg/m
3
1.510
3
(22m/s)
2
= 1.0Pa. (Note that 80 km/hour is a strong wind, one that
makes it very difcult to walk.)
These examples may seem to be just so many numbers to you, but I do encourage you to
go through the motions of performing similar calculations for examples that you make up,
so that you will become familiar with the values that are relevant to the ocean.
A5.1 Flow on a tilted table (page 85)
Question. A large, at table is oriented to be perfectly level. A high edge runs along the
sides of the table, so that it can contain water. Water is poured onto the table and allowed
to come to rest. Suddenly, the table is tilted to some small angle from level. (a) Discuss
what happens, physically. (b) Quantify your answer.
Answer. (a) Flow runs down the table. Waves may form, especially near the outow
edge, where overow may occur depending on the edge height and the angle of tilt.
(b) Anticipating that our mathematical solution will match the intuitive solution, lets
start by dening a coordinate system with x running down the slope. That way, well expect
both the forcing and the response to be in the x-direction. Let the slope of the tilt table be
denoted by .
We shall consider only the initial response that is, at times that are small compared the
time it will take uid to pile up on one end of the table. (Well come back to estimating this
time in a little while.)
Finally, we shall ignore frictional drag on the bottom, reasoning that it is proportional to
the square of velocity, which will be small during the time when the water is just beginning
to ow.
Because the x-axis is tilted downward at the angle , the weight of the water exerts a
force in the x-direction, yielding a force of g in the positive x-direction.
The equation of motion, based on these assumptions, is
u
t
+u
u
x
= g (21.9)
Now consider the ratio of the rst two terms. If a velocity scale is U, a length-scale is
L (say, the width of the table) and a time-scale is , then the ratio of the space-derivative
to the time-derivative is U/L. If this [nondimensional??] number is small, then the
nonlinear (advective) term may be ignored from the equation of motion. As it turns out,
this [nondimensional??] number may be made arbitrarily small (i.e., we can satisfy the
inequality U/L 1) by choosing L large enough (for given U and ), or by considering
only a small interval of time . We shall make this small-time assumption. The physics
of this is that the water at the edge of the tilt-table runs into the wall, and therefore cannot
move, so it has u = 0 there, whereas at the center of the table, u is free to take any value.
Thus the scale for u/x is U/L, for the case of a table with retaining walls.
With the small-time approximation, then, the nonlinear (advective) term vanishes, and
the equation of motion becomes
u
t
= g (21.10)
from which, with initial condition u(t) = 0 at t = 0, yields, after a simple integration,
u = gt. This is the desired solution. For example, for a table tilted by 5 degrees (which
is a slope of = 0.1), we get u = 9.8m/s
2
0.1 t, which yields a speed of 1m/s after 1
second, and 2 m/s after 2 s. Thats quite fast, in oceanographic terms! You might like to try
an experiment to see if this makes sense. Dont use water though its too messy. Instead,
put a ball-bearing (or tennis ball, or whatever) on one edge of a 2m-long table. Then lift one
edge of the desk by 5 degrees. (For a 2m-wide desk, just raise the edge by about 20cm, or
about a one-octave span of your ngers and thumb on a piano.) Out loud, count seconds
by saying one thousand and one, one thousand two to measure 2s. Ask your lovely and
237
talented assistant to put a chalk mark on the table, at each of these well-calibrated seconds.
If our prediction is reasonable, the speed between 1s and 2s should be of order 1m/s, so
the chalk lines should be separated by about about 1m, or a reasonable fraction of the table
width.
We may now return to the issue of the period over which (21.10), sans the nonlinear term,
is valid. We have an estimate u = gt, so that at time t = , U = g. Thus, our criterion
for neglecting the uu/x term, i.e. U/L 1, can be written g
2
/L 1. So, for our
L = 2m example, we require that for times less than about
_
L/(g), or times less
than
_
2m/(10m/s
2
0.1) 1s to 2s. So, seems that our 2m-wide table would be just about
wide enough for the uid-style version of our experiment to be valid
8
.
A5.2 Tidal velocities in harbour (page 90)
Question. The tidal range in Halifax harbour is roughly 2 m (thats the difference in
height between high tide and low tide). What tidal velocities should this generate? (Hint:
you need to think about areas.)
Answer. If you live in Nova Scotia, and have gone to the beach as often as I did when I
was a kid, youll know that there are two tides per day in these parts. Therefore, the period
is roughly 12 hours, so that the tide ows in over 6 hours, and then ows out over 6 hours.
Well start by calculating the volume ux of the tide. We know the elevation difference
is 2 m, so if we can estimate the harbor surface-area (the area as seen from above), then we
can calculate the amount of water that enters each tide. Ill guess that the harbor, together
with Bedford Basin, has area corresponding roughly to a square 5 km on a side. Thus the
area is 310
7
m
2
. (Youll note that Im using just 1 signicant digit here maybe even
thats too much given how Im slinging estimates around without even consulting a map!)
Therefore the volume entering in each tide is V 2m 310
7
m
2
, or V 610
7
m
3
. Now,
the period through which this ows, 6 hours, is 210
4
s, so that the volume ux is the ratio
(V 610
7
m
3
)/(210
4
s), or 310
3
m
3
/s.
Now, computing the tidal velocities is a simple matter of dividing this volume ux by
the area through which the water ows. This area is given by the product of the width
of the harbor, at the inlet point (by Point Pleasant Park). This is, perhaps, 2 km width by
30 m depth, or 610
4
m
2
. Dividing this into the volume ux gives us the velocity estimate
(310
3
m
3
/s)/(610
4
m
2
), or roughly 0.05 m/s.
This result, 5 cm/s, can be visualized
9
by rubbing your right index nger across your left
palm (a distance of roughly 5 cm), counting one thousand and one.
Youve certainly seen junk oating in the harbor. If not, wait for a sunny day and get a
beer at one of the harbor-side pubs, and watch the water move. Question: is the estimate
weve derived out by a factor of 10, either way?
A5.3 Geostrophic Flow (page 95)
Question. A depth-independent geostrophic ow runs along Nova Scotia, from Cape
Breton towards Yarmouth. It has velocity of roughly 10 cm/s and is about 30 km wide. (Each
of these numbers is probably good to a factor of 2.) Consider the deformation in sealevel
associated with the ow. The sealevel at the coast will be either higher or lower than the
level that would exist without the current. (a) Is it higher or lower? (b) Estimate the change
of sealevel using the geostrophic and hydrostatic approximations.
Answer. (a) The level at the coast is higher than it otherwise would be. To see this, note
that the pressure-gradient and coriolis forces must balance if the ow is to be steady. The
coriolis force is towards the shore (to the right of the ow), so the pressure-gradient force
8
But if you try this experiment, Im not responsible for cleaning up the kitchen oor!
9
Its worth doing the exercise it helps to attach a visual or tactile measure to numerical quantities like this.
238 CHAPTER 21. ANSWERS TO EXERCISES
must be away from the shore. This pressure-gradient force is achieved by having sealevel
being high at the coast.
Consider an x-axis that runs along the coast, with x increasing in the Cape Breton
direction, and a y-axis that runs at right angles to the coast, with y increasing seaward. Then
the current is described by the statement that u =0.1m/s.
Let denote the sealevel deformation, measured positive for sealevel higher than in the
no-ow situation. With hydrostatic pressure (see (5.34), the sealevel tilt d/dy controls the
pressure-gradient term through the rule

P
y
=g

y
(21.11)
so that the equation of motion becomes (see section 5.4)
f u =g

y
(21.12)
At the latitude of Nova Scotia, f = 110
4
s
1
.
We may estimate the spatial derivative of sealevel by d/dy = /y. Here, y =
310
4
m (30 km) as the width of the current, and is the sealevel deformation we are
seeking. So, substituting, we get = f uy/g, or = [1 10
4
s
1
(0.1m/s)
310
4
m]/[10m/s
2
] or = 0.03 m. Given our stated uncertainty in the current speed and
width, we might reasonably state that the sealevel increase at the cost is roughly in the range
of 1 cm to 5 cm.
A5.4 Ekman layer thickness part 1 (page 105)
Question. At 45 degrees North, what is the Ekman layer thickness if the vertical viscosity
is A
V
= 10
2
m
2
/s?
Answer. The Ekman scale, given by (5.120), is D
e
=
_
2A
V
/f . For the latitude of
45N, f = 110
4
s
1
, and with A
V
= 10
2
m
2
/s, we obtain an Ekman scale thickness of
D
e
= 14 m. Given that A
V
has been specied only to within an order of magnitude, however,
the result is likely to be uncertain to a factor of about 3 (roughly, the square-root of 10),
so its best to report it as D
e
10 m, where the symbol indicates an order-of-magnitude
uncertainty.
A5.5 Ekman layer thickness part 2 (page 105)
Question.f a current-meter string sited at 45 degrees North indicates that the Ekman-layer
thickness 30 m, what is the A
V
in the region?
Answer: A scale of D
e
= 30 m, at latitude 45

N (where f = 110
4
s
1
) yields A
V
=
D
2
e
f /2, or A
V
= (30m)
2
110
4
s
1
/2, or A
V
= 510
2
m
2
/s.
A5.6 Ekman transport (page 105)
Question. A steady wind of U
a
=10 m/s blows on the ocean. (a) What is the wind-stress?
(b) What Ekman ux results in response to the wind?
Answer. (a) The wind-stress is given by (4.28), i.e.
=
a
C
D
U
2
a
, (21.13)
With atmospheric density
a
= 1kg/m
2
and with C
D
= 110
3
this yields = 0.1 Pa.
(b) The Ekman volume ux per metre of cross-section is given by M
e
= /( f ), in
units of m
2
/s. (To understand this unit, visualize an imaginary vertical plane aligned
239
with the direction of the wind, with Ekman ow passing through the plane. Our unit
is the volume ux per meter along the length of this plane.) In our case, we get M
e
=
0.1Pa/(1000kg/m
3
110
4
s
1
) (Note that we must use here the water density, not the
air density as in the stress calculation!) Thus M
e
= 1m
2
/s. If the Ekman layer thickness
is roughly D
e
= 10m (as in examples above), this corresponds to an Ekman velocity of
U
e
= M
e
/D
e
= (1m
2
/s)/(10m), or about 0.1m/s.
To help visualize the speed, trace your right nger across your left palm, counting one
thousand and one to mark a second of motion across this roughly 10 cm distance. Note
that the wind-speed is much swifter than the ocean Ekman speed that is created by the wind
stress
10
.
A6.1 Vorticity of simple ows (page 113)
Question. What is the vorticity of (a) a shear ow and (b) a solid-body rotation?
Answer. (a) Consider a ow in the x direction, sheared in the y direction. (Any directions
could be chosen for this.) Let the shear be A, i.e. u = u(y) is given by u = Ay. Thus, the
vorticity, dened by =v/xu/y is equal to A. (b) It is easiest to think of solid-body
rotation in a polar coordinate system, i.e. with coordinates (r, ), where r measures the
distance from the centre of the spinning body and (in radians, positive counterclockwise)
measures the angle with respect to some reference point. For solid-body rotation at angular
frequency (radians/second, positive for anticlockwise motion), the tangential velocity is
V = r, which can be translated into cartesian (x, y) velocity components u =r sin()
and v = r cos() from simple geometry. But the coordinate x is just r sin() and y is
r cos(), so we have u =y and v = x. Thus, substituting into the equation above, we
get = (), i.e. the vorticity is = 2. In words, the vorticity of a spinning body is
2 times the angular rotation rate.
A10.1 Bow-waves in Harbour (page 151)
Question. My apartment overlooks Halifax Harbour. Often I see large boats (e.g. QM2)
cruising along the edge of the harbour. If such a boat runs 50 m from the harbour edge, how
long will it take the bow waves from the boat to reach the wharf?
Answer. First, well have to make some guesses about the wave characteristics, to see
if they are deep-water or shallow-water waves. The criterion involves the wavenumber
k = 2/, where is the wavelength. I think, based on observation, that is of order a few
meters, say 3m. (Do you agree, to within a factor of 3 or 10?) Thus, the wavenumber
k = 2/ is of 2m
1
. The water, deep enough to hold a cruise ship, is certainly in excess
of h = 10m. Thus kh must be at least of value 20. This means that the inequality kh 1
certainly applies, so that the waves fall solidly into the deep-water category.
Therefore the phase speed of the waves is C
P
=
_
g/k and the group speed is C
G
=
(
_
g/k)/2. The set of waves from the ship travel at the group speed. Substituting, we get
C
G
=
_
(10m/s
2
)/(10m
1
)/2, or C
G
= 0.7m/s. Thus it will take roughly 50m/(0.7m/s),
or about 60s, a minute, for the wave to reach the wharf.
If you think about it, this timescale is pretty large in human terms. If you sit on the edge
of the wharf on a warm day, concentrating on keeping your ice-cream from dripping on your
clothes, you probably wont be paying enough attention to gure out, when you hear waves
lapping the wharf, exactly which ship it was caused by!
A10.2 Tsunami waves (page 151)
10
For reference, you could bet your tuition that nobody in your class can sustain a 10 m/s running speed for more
than a few seconds. It is difcult to run like the wind!
240 CHAPTER 21. ANSWERS TO EXERCISES
Question. Its a hot summer day in Halifax, and Im sitting by the shore, sipping a mint
julip
11
. My nerd oceanographer drinking buddy is an expert in Tsunami waves (created by
earthquakes at sea), and she has a beeper that goes off, indicating that her seismometer has
just detected an earthquake 100km offshore. How long do we have to nish our drink?
Answer. Tsunami waves have very long wavelength, being comparable to the width of
the ocean bottom that is bumping up and down in the earthquake. Therefore the wavelength
greatly exceeds the water depth h, so that kh = 2h/ will be 1, revealing that the
tsunami wave will be of the shallow-water variety.
The phase speed of shallow-water waves equals the group speed, each being given by
C =

gh. We know that the earthquake is 100 km offshore, and a glance at a high-school
atlas conrms that the water depth over that offshore domain is roughly 100 m (say, to within
a factor of 2). Thus the wave speed is C =
_
(10m/s
2
) 100m, or 30 m/s. (Note that this is
a large speed, compared with the speed of the waves youve seen running up onto a beach.)
Given a distance of 100 km, the transit time is 10010
3
m/(30m/s) or roughly 3000 s. So it
will take under an hour for the wave to arrive, drowning those who didnt heed my warning
and catch the #1 bus back to the high point in Halifax (Dalhousie).
. . . oh yeah, about that nal drink. You have time to gulp and run to the bus-stop
12
, but
not enough to order another round.
11
Hey, I said it was hot, so better drink a southern drink!
12
If you drink and drive, you fail this course.
Bibliography
D. L. Anderson. The physical constants of sea ice. (journal not known), 13:310318, 1960.
John Backus. The history of FORTRAN I, II, and III. IEEE Annals of the history of
computing, 1(1), 1979. URL http://www.computer.org/annals/an1979/
a1021abs.htm.
George K. Batchelor. An Introduction to Fluid Dynamics. Cambridge University Press,
Cambridge, 1967.
K. F. Bowden. Physical Oceanography of Coastal Waters. Harwood/Wiley, 1983.
Wendell S. Brown. A comparison of Georges Bank, Gulf of Maine and New Eng-
land Shelf tidal dynamics. Journal of Physical Oceanography, 14(1):145167,
1984. URL http://dx.doi.org/10.1175%2F1520-0485%281984%29014%
3C0145%3AACOGBG%3E2.0.CO%3B2.
D. F. Bumpus. A description of the circulation on the continetal shelf of the east coast of the
United States. Prog. Oceanogr., 6:111158, 1973.
Mark A. Collier and Paul J. Durack. CSIRO netCDF version of the NODC World Ocean
Atlas 2005. Technical report, CSIRO, 2006. URL http://apdrc.soest.hawaii.
edu/datadoc/woa05.htm.
Robert M. Corless. Essential Maple: an introduction for scientic programmers. Springer-
Verlag, 1995.
Benoit Cushman-Roisin. Introduction to Geophysical Fluid Dynamics. Prentice-Hall,
Englewood Cliffs, New Jersey, 1994.
Margaret. Deacon. Scientists and the sea, 1650-1900 : a study of marine science / Margaret
Deacon. Academic Press, New York, 1971. ISBN 0122078500.
A. T. Doodson. The harmonic development of the tide-generating potential. Proc Roy
Soc London A, 100(704):305329, 1921. URL http://www.jstor.org/stable/
93989.
Fr ed eric Dupont, Charles G. Hannah, and David Greenberg. Modelling the sea level of the
Upper Bay of Fundy. Atmosphere-Ocean, 43(1):3347, 2005.
K. R. Dyer. Estuaries: A Physical Introduction. John Wiley, 1972.
V. W. Ekman. On the inuence of the earths rotation on ocean currents. Ark. Math. Astron.
Fys., 2(1-53), 1905.
William J. Emery and R. E. Thomson. Data Analysis Methods in Physical Oceanography.
Pergamon, 1998.
Hugo B. Fischer. Mixing and dispersal in estuaries. Annu. Rev. Fluid Mech., 8:107133,
1976.
Pierre Flament, L. Armi, and L. Washburn. The evolving structure of an upwelling lament.
Geophys. Res., 90:1176511778 and 1183511836, 1985.
N. P. Fofonoff and R. C. Millard. Algorithms for computation of fundamental properties of
seawater. Technical Report 44, UNESCO Division of Marine Science, Paris, 1983.
241
242 BIBLIOGRAPHY
M. G. G. Foreman. Manual for tidal heights analysis and prediction. Pacic Marine Science
Report, 77(10):158, 1977, revised 2004. URL http://www.pac.dfo-mpo.gc.
ca/sci/osap/projects/tidpack/tidpack_e.htm.
Peter S. Galbraith and Dan E. Kelley. Identifying overturns in CTD proles. Journal of
Atmospheric and Ocean Technology, 13:688702, 1996.
Alexandre Ganauchaud and Carl Wunsch. Improved estimates of global ocean circulation,
heat transport and mixing from hydrographic data. Nature, 408:453456, 2000.
C. J. R. Garrett, J. R. Keeley, and D. A. Greenberg. Tidal mixing versus thermal stratication
in the Bay of Fundy and Gulf of Maine. Atmosphere-Ocean, 16(4):403423, 1978.
Chris Garrett and Patrick Cummins. The power potential of tidal currents in channels.
Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences,
461(2060):25632572, 2005. URL http://dx.doi.org/10.1098/rspa.2005.
1494.
Chris Garrett and Patrick Cummins. The efciency of a turbine in a tidal channel. J
Fluid Mech, 588:243251, 2007. URL http://journals.cambridge.org/
production/action/cjoGetFulltext?fulltextid=1346064.
Christopher Garrett. Tidal currents and eddy statistics from iceberg trajectories off Labrador.
Science, 227(4692):13331335, 1984.
Christopher Garrett. Tidal resonance in the Bay of Fundy and Gulf of Maine. Nature, 238
(5365):441443, 1972. URL http://dx.doi.org/10.1038/238441a0.
Adrian E. Gill. Atmosphere-Ocean Dynamics. Academic Press, New York, 1982.
Richard J. Greatbatch and Jian Lu. Reconciling the Stommel box model with the Stommel-
Aron s model: a possible role for southern hemisphere wind forcing? J. Phys. Oceanogr.,
33:16181632, 2003.
Michael C. Gregg, Thomas B. Sanford, and David P. Winkel. Reduced mixing from the
breaking of internal waves in equatorial waters. Nature, 422(3 April):513515, 2003.
Geo. Hadley. Concerning the cause of the general trade-winds: By Geo. Hadley, Esq; F.
R. S. Philosophical Transactions (1683-1775), 39:5862, 1735. ISSN 02607085. URL
http://www.jstor.org/stable/103976.
Dale B. Haidvogel and Ajke Beckmann. Numerical ocean circulation modelling. Imperial
College Press, River Edge, NJ, 1999.
Chris W. Hughes and Beverly A. de Cuevas. Why western boundary currents in realistic
oceans are inviscid: a link between form stress and bottom pressure torques. J. Phys.
Oceanogr., 31:28712885, 2001.
John M. Huthnance. Waves and currents near the continental shelf edge. Prog. Oceanogr.,
10:193226, 1981.
J. V. Iribarne and W. L. Godson. Atmospheric Thermodynamics. Reidel, Dordrecht, Holland,
1981.
D. E. Kelley and K. A. Van Scoy. A basin-wide estimate of vertical mixing in the upper
pycnocline: spreading of bomb tritium in the north pacic ocean. J. Phys. Oceanogr., 29:
17591771, 1999.
Dan E. Kelley, H. J. S. Fernando, A. E. Gargett, J. Tanny, and E.

Ozsoy. The diffusive
regime of double-diffusive convection. Prog. Oceanogr., 56:461481, 2003.
John A. Knauss. Introduction to Physical Oceanography. Prentice-Hall, Englewood Cliffs,
NJ, 1978.
BIBLIOGRAPHY 243
K. J. M. Kramer, U. H. Brockmann, and R. M. Warwick. Tidal estuaries: manual of
sampling and analytical procedures. A. A. Balkema, Brookeld, VT, USA, 1994.
J. R. Ledwell, A. J. Watson, and W. S. Broecker. A deliberate tracer experiment in santa
monica basin. Nature, 323:322324, 1986. dye release in Santa Monica Basin. Trace for
50 days. Get diffusivity = (3.3+/-0.8)e-5 m*m/s at 800m, where N=(2.0+/-0.2)e-3rad/s.
E. L. Lewis. Physics of snow and ice. In H. Oura, editor, Heat ow through winter ice. Inst
of Low Temp. Sci., Hoddaido University, Sapporo, Japan, 1967.
Jim R. Luyten, Joseph Pedlosky, and Henry Stommel. The ventilated thermocline. J. Phys.
Oceanogr., 13:292309, 1983.
Ole Secher Madsen. A realistic model of the wind-induced Ekman boundary layer. Journal
of Physical Oceanography, 7(2):248255, 1977. URL http://dx.doi.org/10.
1175%2F1520-0485%281977%29007%3C0248%3AARMOTW%3E2.0.CO%3B2.
Mathew E. Maltrud and Julie L. McClean. An eddy resolving global 1/10

ocean simulation.
Ocean Modelling, in press, 2004.
Trevor J. McDougall. Thermobaricity, cabbeling, and water-mass conversion. J. Geophys.
Res., 92:54485464, 1987.
Chris Mooers. Wind-driven currents on the Continental Margin. In D. J. Stanley and D. J. P.
Swift, editors, Marine sediment transport and environmental management, chapter 2.
Wiley, 1976.
Robert M. Moore and Doug W. R. Wallace. A relationship between heat transfer to sea ice
and temperature-salinity properties of arctic ocean waters. J. Geophys. Res., 93:565571,
1988.
Walter H. Munk. Abyssal recipes. Deep-Sea Res., 13:707730, 1966.
J. Nikuradse. Gesetzm assigkeiten der turbulenten Str omung in glatten Rohren. VDI-
Forschungsheft, Berlin, 1932.
A. Okubo. Oceanic diffusion diagrams. Deep-Sea Research, 18:789802, 1971.
N. Ono. Specic heat and heat of fusion of sea ice. In H. Oura, editor, Physics of snow and
ice. Institute of Low Temperature Science, Hokkaido University, Sapporo, Japan, 1967.
Rich Pawlowicz, Bob Beardsley, and Steve Lentz. Classical tidal harmonic analysis in-
cluding error estimates in matlab using t tide. Computers & Geosciences, 28(8):929
937, 2002. URL http://www.sciencedirect.com/science/article/
B6V7D-46DP7HS-1/1/09396567bec54b42f742f47eafd31455.
C. H. Pease. The size of wind-driven coastal polynyas. J. Geophys. Res., 92:70497059,
1987.
Joseph Pedlosky. Geophysical Fluid Dynamics. Springer-Verlag, New York, 1979.
Hartmut Peters. Spatial and temporal variability of turbulent mixing in an estuary. J. Mar.
Res., 1999.
George L. Pickard. Descriptive Physical Oceanography. Pergamon, New York, 1979.
Stephen Pond and George L. Pickard. Introductory Dynamic Oceanography. Pergamon,
New York, 1978.
E. R. Pounder. The physics of ice. Pergamon, New York, 1965.
E. M. Purcell. Life at low Reynolds number. Am. J. Physics, 45:311, 1977. URL
http://scitation.aip.org/getabs/servlet/GetabsServlet?
prog=normal&id=AJPIAS000045000001000003000001&idtype=
cvips&gifs=yes.
244 BIBLIOGRAPHY
H. Sandstrom and Jim A. Elliot. Internal tide and solitions on the Scotian Shelf: a nutrient
pump at work. J. Geophys. Res., 89:64156426, 1984.
Raymond W. Schmitt. Double diffusion in oceanography. Annu. Rev. Fluid Mech., 26:
255285, 1994.
Steffen Seitz, Petra Spitzer, and Richard Brown. Consistency of practical salinity measure-
ments traceable to primary conductivity standards: Euromet project 918. Accreditation
and Quality Assurance: Journal for Quality, Comparability and Reliability in Chemical
Measurement, 13(10):601605, 10 2008. URL http://dx.doi.org/10.1007/
s00769-008-0444-0.
J. H. Simpson, D. J. Crisp, and C. Hearn. The shelf-sea fronts: Implications of their
existence and behaviour [and discussion]. Philosophical Transactions of the Royal
Society of London. Series A, Mathematical and Physical Sciences, 302(1472):531546,
1981. ISSN 00804614. URL http://www.jstor.org/stable/37036.
John H. Simpson. The shelf-sea fronts: implications of their existence and behaviour. Phil.
Trans. R. Soc. Lond. A, 302:531546, 1981.
F. G. W. Smith. Handbook of Marine Science, vol. 1. CRC Press, Cleveland Ohio, 1974.
Robert H. Stewart. Introduction to Physical Oceanography. Texas A&M University,
2005. URL http://oceanworld.tamu.edu/resources/ocng_textbook/
contents.html.
Henry Stommel. The delicate interplay between wind-stress and buoyancy input in ocean
circulation: the Goldbrough variations. Tellus, 36:111119, 1984.
C. L. Strong. The amateur scientist: Experiments with salt fountains and related instabilities
in water. Scientic American, pages 124127, 1971.
H. Tennekes and J. L. Lumley. A First Course in Turbulence. MIT Press, Cambridge, MA,
1972.
W. C. Thacker. Three lectures on tting numerical models to observations. Lectures to the
Physics Institute of GKSS, pages 165, 1988.
W. C. Thacker and R. B. Long. Fitting dynamics to data. J. Geophys. Res., 93:12271240,
1988.
J. Stewart Turner. Buoyancy Effects in Fluids. Cambridge University Press, Cambridge,
1973.
R. C. Weast. Handbook of Chemistry and Physics. CRC Press, Boca Raton, Fla, 1985.
W. F. Weeks and S. F. Ackley. The growth, structure, and properties of sea ice. In
Untersteiner, editor, The geophysics of sea ice, volume 146 of NATO ASI series, chapter 1,
pages 9164. Plenum, 1986.
Mark Wimbush and Walter Munk. The benthic boundary layer. In The Sea, volume 4,
chapter 19. Elsevier, 1970.
Carl Wunsch. What drives the thermohaline circulation? Science, 298:11791181, 2002.
Daniel Zwillinger. Handbook of Differential Equations. Academic Press, Boston, 1989.
Index
D
e
, Ekman-layer thickness, 104
G, gravitational constant, 168
N, buoyancy frequency, 61
TS diagram, 68
, the von Karman constant, 157
beta, 112
u velocity component, 8
u

, friction velocity, 157


v velocity component, 8
absolute temperature, 48
abyssal recipe of Munk, 120
acceleration
centripetal, 87
eulerian vs lagrangian, 79
air-sea uxes, see ux
approximation
Boussinesq, 96
continuum hypothesis, 77
incompressible ow, 82
aspect ratio of ocean, 5
baroclinic, 61
barotropic, 61
barycenter, 167
Batchelor scale, 155
bathythermograph, 51
beta, 112
black-body radiation, 161
bottom ekman layer, 156
Boussinesq approximation, 96
Brunt-vaisala frequency, 61
Buckingham theorm, 32
bulk ux formulae, 65, 67
buoyancy
currents on shelf, 129
frequency, 61
cabbelling, 72
calculations, error propagation in, 191
calculations, uncertainties in, 191
cartesian vector space, 32
centripetal acceleration, 87
CFL condition, in numerical modelling, 216
circulation, wind-driven, 112
classication of estuaries, 133
COARE, 65, 67
coastal jet, in upwelling, 127
coastal upwelling velocity, 127
complex numbers, 31
computational cost, in numerical modelling,
210
computational molecule, 216
computer requirements for modelling, 210
computers, range in capabilities, 212
conservation
of S, T, etc, 83
of mass, 81
constant-stress layer, 157
continental shelves, 11, 125
continuity equation, 82
and upwelling velocity, 127
depth-integrated, 112
continuum hypothesis, 77
Coriolis
and waves, 152
denition of parameter f , 87
effect, 7, 8, 85
force, 85
formula for the parameter, 8
CPU time requirements, for numerical mod-
elling, 211
cross product between vectors, 33
CTD probe, 51
curl of a vector eld, 33
data assimilation, 221
day length, for Coriolis parameter, 85
day, sidereal (for Coriolis), 85
decibar, pressure unit, 41
deep-water waves, 146
density
air, 203
seawater, 41, 57, 58, 233
destratication energetics, 64
differential equations, 27
differentiation, 25
diffusive instability, 72
diffusivity ofseawater, 72
digits, signicant, 191
dimensionless numbers, 32
dispersion relationship, 141
dissipation rate, 154
divergence of a vector eld, 33
dot product between vectors, 33
double-diffusion, 72
dynamic height, 99
earth mass, 169
eddy
ux, 83
Ekman
convergence, 115
245
246 INDEX
layer, 104
spiral, 104
transport, 105
upwelling, 116
Ekman number, 93
electromagnetic radiation, 161
energy
in a uid, 62
solid-body, 37
equation
of continuity, see continuity equation
of motion
dimensional, 90
nondimensional, 90
of state, 57, 58
equatorial upwelling, 74
error propagation in calculations, 191
estuary
circulation, 133
circulation ushing time, 139
classication, 133
denition, 133
sampling techniques, 133
tidal ushing time, 139
variant Richardson number, 135
Eulers rule, 23, 24, 31
eulerian vs lagrangian, in numerical mod-
elling, 213
evaporation, 48, 65
example of code, in numerical modelling,
220
experiment
SLEIWEX, 69
WOCE, 42
Fermats last theorem, 79
nite-difference numerical modelling, 216
nite-element numerical modelling, 216
shing at shelf-edge fronts, 11
fjord, 135
ux
across air-sea interface, 65
bulk formulae for computing, 65, 67
concept of, 80
eddy, 83
force
balance, 83
Coriolis, 85
friction, 88, 89
gravity, 84
pressure, 84
summary, 90
fortran computer language, 208
friction
force, 88
velocity, 157
viscosity coefcients, 89
fronts
illustration, 12
importance to shing, 11
function
denition, 21
general circulation, 123
geometry of ocean, 5
geostrophy
and bottom slope, 100
equations for, 93
gradient of a scalar, 34
graphical methods
prole, 68
section, 68
graphical methods
TS diagram, 68
introduction, 67
map, 68
gravity
force, 84
value of G, 168
waves, internal, 151
waves, surface, 144
group velocity, 144
Gulf Stream, 116
gyre circulation, 8, 112
haline contraction coefcient, 58
harbour modes, 149
heat
conservation, 83
ux, 67
height, dynamic, 99
history
of numerical modelling, 208
of oceanographers, 225
horizontal grid schemes, in numerical mod-
elling, 216
hydrostatic equation, 84, 95
imaginary number, 31
implicit vs explicit schemes, in numerical
modelling, 218
incompressible ow, 82
index of refraction, 161
inertial circle, 88
numerical modelling of, 212
infra-red radiation, 54
initial condition, 27
inner product between vectors, 33
inner shelf environment, 126
integration, 27
internal waves, 6, 141, 151
Kelvin temperature, 48
keywords list of, 12
kinetic energy
in waves, 150
INDEX 247
solid-body, 37
Lagrangian acceleration, 79
laminar layer, 157
latex, 199
latitude
effect on Coriolis parameter, 87
layered ow
one layer, 97
two-layer, 97
length of a vector, 32
level of no motion, 99
light, 161
linear equation of state, 58
list
keywords used in text, 12
symbols used in text, 12
local acceleration, 79
log layer, 157
longshore jet, in upwelling, 127
Luyten, Pedlosky and Stommel, 1983, 123
maps, 68
Margules equations, 97
mass
earth, 169
moon, 169
mass conservation, 81
Matlab language
code to work with equation of state, 235
matlab language, 205
mixed-layer dynamics, 65
mixing, 11
modelling, numerical see numerical, 206
moon mass, 169
Munks abyssal recipes, 120
nearshore environment, 126
Newtons second law, 35, 77
Newton, denition of unit, 35
nondimensionalizing differential equations,
32
norm of a vector, 32
North Atlantic Deep Water, 119
North Atlantic Wind stress, 109
nudging, in data assimilation, 223
number
Ekman, 93
Reynolds, 91
Richardson, 15, 135
Rossby, 93
numerical modelling, 206
CFL condition, 216
computational cost, 210
CPU time requirements, 211
eulerian vs lagrangian, 213
example of code, 220
nite-difference method in, 216
nite-element method in, 216
horizontal grid schemes, 216
implicit vs explicit schemes, 218
inertial circle example, 212
open boundaries, 220
range in computer capabilities, 212
storage requirements, 210
surface and interior modes, 220
the Princeton Ocean Model (POM), 208
vertical grid schemes, 213
oce, software, 170
ocean waves, 141
open boundaries in numerical modelling, 220
orbital velocity, 144
outer shelf environment, 126
Ozmidov scale, 154
partial differentiation, 26
phase velocity, 141
planetary rotation
coriolis effect, see Coriolis
POM, the Princeton Ocean Model, 208
potential energy
in waves, 150
solid-body, 37
stratied uid, 64
potential temperature, 48
potential vorticity
and thermohaline overturning, 119
conservation properties, 113
denition, 113
PPT, salinity unit, 41
precipitation, 48, 65
pressure, 41
distribution, 50
force, 84
measurement, 57
unit, 41
prime notation
deviation from mean, 65
differentiation, 28
prole, 4, 68
propagating errors in calculations, 191
PSU, salinity unit, 41
Python language
code to work with equation of state, 234
R language, see software
radiation
black-body, 161
infra-red, 54
RAM, needed for numerical modelling, 210
Re, see Reynolds number
reference velocities in sheared ow, 99
refractive index, 55
resonance, 149
Reynolds number, 91
248 INDEX
Ri, see Richardson number
Richardson number, 15, 135
Estuarine, 135
Rossby
number, Ro, 93
radius of deformation, 128
Rossby waves, 117
rotating coordinate system, 85
roughness length, 157
salinity
conservation, 83
denition, 41
distribution, 48
measured by CTD, 51
measured by refractive index, 55
measurement, 54
salt fountain, 72
salt-ngering, 72
sampling techniques, in estuaries, 133
satellite imagery, 4
scalar, 33
scale
analysis for equation of motion, 90
analysis of differential equations, 31
Ekman-layer thickness, 104
seawater
density, 41, 234
diffusivity, 72
viscosity, 72
section, 4, 68
seiches, 149
sensible heat ux, 67
shallow-water waves, 146
sheared ow, reference velocities, 99
shelf
front, 11
fronts, 130
thermohaline circulation, 129
upwelling at shelf break, 128
waves, 128
sidereal day, 85
sigma-t, density unit, 58
signicant digits, 191
Simpson-Hunter criterion, 65
SLEIWEX experiment, 69
sloped bottom, affect on geostrophic ow,
100
software
C examples
seawater density, 59
Matlab examples
calculate seawater properties, 235
oce, R oceanography package, 170, 206
Perl examples
solve diffusion equation, 219
Python examples
calculate seawater properties, 234
R examples
N
2
calculation, 195
calculate seawater software, 234
draw a bathymetric chart, 13
drawconductivity-temperature-salinity
contours, 56
draw ctd prole, 69
draw CTD summary plot, 69
draw isopycnal diagram, 60, 181
draw oxygen section, 75
draw section diagram, 134
draw SSS map, 49
draw SST map, 43
draw T section, 45
draw T(lat), 46
draw tide signals, 172
draw tide spectrum, 171
draw TS diagram of real-world data,
71
model wind-driven ow, 103
read and plot ADCP data, 183
read and plot CTD data, 207
R oce package, 170, 206
Sage examples
calculate estuarine circulation, 139
t-tide, for tidal analysis, 170
webdrogue, for tidal prediction, 173
webtide, for tidal prediction, 173
solar heating, 47, 119
spring, differential equation for, 27
spring-neap cycle, 172
stability, 61
standing waves, 149
steady response to wind forcing, 104
Stefan-Boltzmann law, 161
storm surges, 129
stratication, 6, 61
streamfunction, 34
subtropical gyre, 112
surface & interior modes in models, 220
surface gravity waves, 144
surges, storm, 129
Sverdrup
relation, 112
transport, 109, 112
symbols, list of, 12
t-tide, see software
Taylor column, 100
Taylor-Proudman theorem, 100
temperature
absolute, 48
conservation, 83
Kelvin, 48
measured by CTD, 51
potential, 48
thermal expansion coefcient, 58
thermal wind, 96
INDEX 249
thermocline, 47
balance in Munks theory, 120
thermohaline circulation, 10, 119
on shelves, 129
thickness of Ekman layer, 104
tidal ushing time in estuaries, 139
tidal power
Bay of Fundy, 176
general, 175
time dependent response to wind forcing,
101
timescale
for mixed-layer deepening, 65
for tidal ushing of an estuary, 139
TOGA, 65, 67
trial solution, for differential equations, 28
turbulent friction, 89
uncertainties, in calculations, 191
UNESCO equation of state, 58, 234
units, use of, 32
upwelling
coastal jet, 127
longshore jet, 127
velocity, 127
vector
cross product, 33
curl of, 33
denition, 32
divergence of, 33
dot product, 33
formed from gradient of a scalar, 34
inner product, 33
length, 32
norm, 32
velocity
friction, 157
waves
group speed, 144
phase, 141
ventilated thermocline, 123
vertical grid schemes, in numerical modelling,
213
viscosity, 89
viscosityof seawater, 72
von Karman constant, 157
vorticity, potential, see potential vorticity
wavelength of light, 161
waves, 141
deep water, 145, 146
dispersion relationship, 141
energy, 150
group velocity, 144
internal, 6, 141, 151
orbital velocity, 144
phase velocity, 141
role of Coriolis effect, 152
shallow water, 146
shelf, 128
standing, 149
web
program for seawater density, 58
webdrogue, see software
webtide, see software
wind stress, 65
causing surface mixed layer, 47
curl of, 112
driving coastal upwelling, 126
driving ocean circulation, 8, 112
North Atlantic, 109
time-dependent response to, 101
WOCE experiment, 42
Wust, 119

Vous aimerez peut-être aussi